id
stringlengths
27
33
source
stringclasses
1 value
format
stringclasses
1 value
text
stringlengths
13
1.81M
warning/0002/hep-th0002030.html
ar5iv
text
# 1 Introduction ## 1 Introduction Since late 1997 there has been a lot of interest in superstring theory in $`AdS_m\times S^n`$ space, as according to the Maldacena conjecture string theory on $`AdS_m\times S^n`$ can be compared to Conformal Field Theory in $`m1`$ dimensions. Initially, the greatest interest fell to IIB superstring theory on $`AdS_5\times S^5`$, due to $`AdS_5\times S^5`$ being a maximally supersymmetric vacuum of IIB supergravity and its CFT counterpart being $`𝒩=4`$, $`D=4`$ Super Yang Mills Theory. However the problems of quantising this theory so that it can be compared to it’s CFT counterpart and of our poor understanding of 4 dimensional CFT has led to much effort going into the study of string theory on $`AdS_3\times S^3\times ^4`$, which should be simpler to compare to CFT. The $`AdS_3\times S^3\times ^4`$ background is the near horizon geometry of a IIB D1 - D5-brane system. Finding the action of a D1 or D5-brane in $`AdS_3\times S^3\times ^4`$ is therefore like finding the action of one of the parallel and coincident D1 or D5-branes which cause the space to deform into $`AdS_3\times S^3\times ^4`$, if the relevant D1 or D5-brane has been slightly extracted from the rest of the D-branes. Here $`^4`$ is a Ricci-flat four dimensional compact manifold, such as $`K3`$ or $`T^4`$ . Despite this being a much studied system there has been a gap in the research so far, ie: finding the full $`\kappa `$-invariant and supersymmetric IIB effective actions of the D1 and D5-branes in $`AdS_3\times S^3`$ (although some papers have either: proposed a method for finding the D1-brane action ; given just the BI-action ; given a bosonic action and performed an energy analysis of the branes ; or used the BPS method to find a D5-brane BPS solution of supergravity in $`AdS_5\times S^5`$). In this paper we determine the action of the D1 and D5-branes in an $`AdS_3\times S^3`$ background, where both the D1 and D5-branes lie entirely in $`AdS_3\times S^3`$ and the $`^4`$ space has been compactified to zero volume. This D5-brane does not correspond to one of the D5-branes that are deforming the background to $`AdS_3\times S^3\times ^4`$, as such branes have 4 dimensions in $`^4`$. The method used was based on the method of Metsaev and Tseytlin in . The actions are described by constructions of Cartan forms on the coset superspace. The supersymmetric BI-action is found first, and then $`\kappa `$-symmetry and supersymmetry are used to determine what WZ term is needed to complete the action . The coset superspace of $`AdS_3\times S^3`$ is $`\frac{SU(1,1|2)^2}{SO(2,1)\times SO(3)}`$ which describes the spacetime supersymmetries the action should display. The D5-brane found above is of interest due to the fact that as the brane must lie in $`AdS_3\times S^3`$, instead of the larger space $`AdS_3\times S^3\times ^4`$, this can be interpreted as a constraint imposed on the fields of a D5 action in $`AdS_3\times S^3\times ^4`$, such that the solution must lie in $`AdS_3\times S^3`$ and with the $`^4`$ compactified away. This constraint on the solutions can be thought of as describing a brane that is lying in the background of another D5-brane (or branes), with the brane dimensions in $`^4`$ compactified away and with the branes intersecting in one spatial dimension. Such a configuration is S-dual to the NS5-NS5 configuration described in , which is BPS and thus such a D5-D5 system is also BPS. Next the action of the D5-brane in $`AdS_5\times S^5`$ was found using the background coset superspace of $`AdS_5\times S^5`$, $`\frac{SU(2,2|4)}{SO(4,1)\times SO(5)}`$. This was initially done as an exercise in preparation for handling a D5-brane in $`AdS_3\times S^3\times ^4`$, however it is of interest in its own right. This brane, described by the BPS method in , is related to the much studied D5-brane of (and others), which describes a D5-brane in the presence of N D3-branes, and connected to them by N fundamental strings. Such a brane describes a baryon vertex in the corresponding 4 dimensional gauge theory. To transform the D5-brane action found here to the D5 action of restrictions must be placed on which dimensions the D5-brane can lie in and some of the fermionic degree of freedom must be projected out to accomodate the presence of the fundamental strings as well as the D5-brane. Finally an attempt was made to find the most general D5-brane action in $`AdS_3\times S^3\times T^4`$. The action found using the method here is not fully $`\kappa `$-invariant and the reason why this is so is discussed. This paper has been structured as follows. In subsection 2.1 Cartan forms are introduced, and it is described how these objects can be used to describe D-brane actions. The $`\kappa `$-invariance constraint that ensures that the D-brane does not have too many fermionic degrees of freedom is also discussed. In subsection 2.2 the supersymmetric field strength $`F`$ is defined. In Sections 3 and 4 the actions of the D1-brane and D5-brane in $`AdS_3\times S^3`$ are found, respectively. In Section 5 the slightly more complicated case of a D5-brane action in $`AdS_5\times S^5`$ is found. Section 6 describes the attempt to find a fully supersymmetric and $`\kappa `$-invariant action in $`AdS_3\times S^3\times T^4`$. Section 7 is a summary of the results. Appendix A contains all the algebra and notation used in the three backgrounds, and Appendix B contains the Maurer-Cartan equations and other useful Cartan form relations. Appendix C contains the gauge fixed actions of the D1 and D5-branes, using the Killing gauge of . ## 2 Cartan 1 Forms and the Supercoset Method ### 2.1 Cartan Forms and the Action The notations and algebra used when finding actions on the three different backgrounds $`AdS_3\times S^3`$, $`AdS_5\times S^5`$ and $`AdS_3\times S^3\times T^4`$ can be found in Appendix A. They have been formulated in such a way that the resulting Maurer-Cartan Equations, the Cartan form variations and various other useful relations can be written out in the same way for the three different backgrounds. The notation used is based upon that in . The papers contain useful reviews of the embedding of brane coordinates in background coordinates for use in deriving the brane actions. As in , for a particular coset superspace the left invariant Cartan 1 forms are $`L^A=dX^ML_M^A`$, $`X^M=(x,\theta )`$ are the background super coordinates and $$G^1dG=L^{\widehat{a}}P_{\widehat{a}}+\frac{1}{2}L^{\widehat{a}\widehat{b}}J_{\widehat{a}\widehat{b}}+L^{\widehat{\alpha }}Q_{\widehat{\alpha }}$$ (2.1) where $`G=G(x,\theta )`$ is a coset representative of $`\frac{SU(1,1|2)^2}{SO(2,1)\times SO(3)}`$ in $`AdS_3\times S^3`$, $`\frac{SU(2,2|4)}{SO(4,1)\times SO(5)}`$ in $`AdS_5\times S^5`$ and $`\frac{SU(1,1|2)^2}{SO(2,1)\times SO(3)}\times U(1)^4`$ in $`AdS_3\times S^3\times T^4`$. The Cartan forms $`L^A`$ are inherently supersymmetric. They are also defined such that there is no mixing between the $`AdS_m`$, $`S^n`$ and $`^4`$ dimensions of space (ie: $`L^{aa^{}}=L^{aa^{\prime \prime }}=L^{a^{}a^{\prime \prime }}=0`$). In the rest of this paper, the spinor index $`\widehat{\alpha }`$ in $`L^{\widehat{\alpha }}`$ is not explicitly written. As in , the gauge choice $`G(x,\theta )=g(x)e^{\theta Q}`$ is chosen, needed to explicitly find the supersymmetric field strength $`F`$ and the R-R fields in $`S_{WZ}`$. $`g(x)`$ is a coset representative of $`\frac{SO(2,2)\times SO(4)}{SO(2,1)\times SO(3)}\frac{SO(2,1)^2\times SO(3)^2}{SO(2,1)\times SO(3)}`$ in $`AdS_3\times S^3`$, $`\frac{SO(4,2)\times SO(6)}{SO(4,1)\times SO(5)}`$ in $`AdS_5\times S^5`$ and $`\frac{SO(2,1)^2\times SO(3)^2}{SO(2,1)\times SO(3)}\times U(1)^4`$ in $`AdS_3\times S^3\times T^4`$. $`g(x)`$ satisfies $$g(x)^1dg(x)=e^{\widehat{a}}P_{\widehat{a}}+\frac{1}{2}\omega ^{\widehat{a}\widehat{b}}J_{\widehat{a}\widehat{b}}.$$ (2.2) The method used is to find the supersymmetric BI-action and to then use $`\kappa `$-symmetry to find the WZ-action. The total effective action of a Dp-brane can then be written as $`S=S_{BI}+S_{WZ}`$ (2.3) $`S_{BI}={\displaystyle _{M_{p+1}}}d^{p+1}\sigma \sqrt{det(G_{ij}+F_{ij})}`$ (2.4) $`G_{ij}=L_i^{\widehat{a}}L_j^{\widehat{b}}\eta _{\widehat{a}\widehat{b}}=_iX^ML_M^{\widehat{a}}_jX^NL_N^{\widehat{b}}\eta _{\widehat{a}\widehat{b}}L^{\widehat{a}}=d\sigma ^iL_i^{\widehat{a}}`$ (2.5) $`S_{WZ}={\displaystyle _{M_{p+1}}}e^FC,C={\displaystyle \underset{n=0}{\overset{p+1}{}}}C_n`$ (2.6) where $`C_n`$ are the R-R super-$`n`$-forms and the brane tension has been set to 1 . The general form of $`d(e^FC)`$ can be deduced from which endeavours to find the D-brane actions on a general background. However extra terms are needed for the WZ-actions here to make them fully $`\kappa `$-symmetric. The $`\kappa `$-transformation $$\delta _\kappa x^{\widehat{a}}=0,\delta _\kappa x^{\widehat{a}\widehat{b}}=0,\delta _\kappa \theta =\kappa ,\frac{1}{2}(1+\mathrm{\Gamma })\kappa =\kappa ,$$ (2.7) is applied to the BI-action and then the WZ-action is defined such that it cancels the $`\kappa `$ variation of the BI-action. $`\mathrm{\Gamma }`$ must have $`\pm 1`$ eigenvalues and $`\mathrm{\Gamma }^2=1`$. Thus $`\frac{1}{2}(1+\mathrm{\Gamma })`$ projects out half of the fermionic degrees of freedom, as needed for supersymmetry. Also, $`d^{p+1}\sigma \mathrm{\Gamma }`$ must be a $`p+1`$ form in order to relate $`\delta _\kappa S_{BI}`$ and $`\delta _\kappa S_{WZ}`$. The $`\mathrm{\Gamma }`$ for a D$`p`$-brane is $$d^{p+1}\sigma \mathrm{\Gamma }=(1)^n\frac{e^F}{_{}}\underset{n}{}\mathrm{\Gamma }_{(2n)}𝒦^n|_{vol},\mathrm{\Gamma }_{(2n)}=L^{\widehat{a}_1}\mathrm{\Gamma }_{\widehat{a}_1}\mathrm{}L^{\widehat{a}_{2n}}\mathrm{\Gamma }_{\widehat{a}_{2n}},$$ (2.8) where $`𝒦`$ and $``$ are defined in (A.2). ### 2.2 Defining $`F`$ As in , $`F`$ is defined using the gauge choice $`G(x,\theta )=g(x)e^{\theta Q}`$. $$F=dA+2i_0^1𝑑t\overline{\theta }\widehat{L}_t𝒦L_t,F=\frac{1}{2}d\sigma ^id\sigma ^jF_{ij},\widehat{L}=L^{\widehat{a}}\mathrm{\Gamma }_{\widehat{a}},$$ (2.9) where $`L_t^{\widehat{a}}(x,\theta )=L^{\widehat{a}}(x,t\theta )`$, $`L_t(x,\theta )=L(x,t\theta )`$. This definition holds in all the three backgrounds used in this paper, due to the choice of the notation used (Appendix A). It is not obvious that $`F`$ is supersymmetric, as it is not defined entirely by Cartan forms. It, however, can be shown that $$dF=i\overline{L}\widehat{L}𝒦L,$$ (2.10) which is supersymmetric. This is shown using equations (B.1), (B.2), (B.11), (B.12) and the Fierz identity $$(\overline{A}\mathrm{\Gamma }_{\widehat{a}}B)(\overline{C}\mathrm{\Gamma }^{\widehat{a}}D)=\frac{1}{2}(\overline{A}\mathrm{\Gamma }_{\widehat{a}}e_lD)(\overline{B}\mathrm{\Gamma }^{\widehat{a}}e_lC)+\frac{1}{2}(\overline{A}\mathrm{\Gamma }_{\widehat{a}}e_lC)(\overline{B}\mathrm{\Gamma }^{\widehat{a}}e_lD),$$ (2.11) where $`A,B,C,D`$ are fermionic 0-forms, and $`e_l=\{\mathrm{𝟏},,𝒥,𝒦\}`$. Thus, $`F`$ is supersymmetric if the variation of $`A`$ is defined such that it cancels the variation of the second term in $`\delta F`$ . ## 3 D1-brane in $`AdS_3\times S^3`$ The D1-brane BI-action is $$S_{BI}=_{M_2}d^2\sigma \sqrt{det(G_{ij}+F_{ij})},$$ (3.1) where $`i,j\{0,1\}`$ are brane coordinates. This is supersymmetric but not $`\kappa `$-invariant. From (2.8) the operator $`\mathrm{\Gamma }`$ which defines the $`\kappa `$ variation for this brane is $$\mathrm{\Gamma }=\frac{\epsilon ^{i_1i_2}(\mathrm{\Gamma }_{i_1i_2}𝒥+F_{i_1i_2})}{2_{BI}}.$$ (3.2) Using $$\delta det(G+F)=det(G+F)Tr((G+F)^1(\delta G+\delta F)),$$ (3.3) one can show that $$\delta _\kappa S_{BI}=2i_{M_2}\overline{\kappa }\widehat{L}𝒥L.$$ (3.4) Next, one shows that the $`\kappa `$-variation of the following supersymmetric WZ-action $$S_{WZ}=2i_{M_2}_0^1𝑑t\overline{\theta }\widehat{L}_t𝒥L_t=i_{M_3}\overline{L}\widehat{L}𝒥L$$ (3.5) cancels this variation. Equivalence of the two different forms of $`S_{WZ}`$ can be seen via $`S_{WZ}`$ $`=`$ $`{\displaystyle _{M_3}}{\displaystyle _0^1}𝑑t_t(i\overline{L}_t\widehat{L}_t𝒥L_t)`$ $`=`$ $`{\displaystyle _{M_3}}{\displaystyle _0^1}𝑑td(2i\overline{\kappa }\widehat{L}𝒥L)`$ $`=`$ $`{\displaystyle _{M_2}}{\displaystyle _0^1}𝑑t2i\overline{\theta }\widehat{L}_t𝒥L_t,`$ where (B.1), (B.2), (B.11), (B.12), (B.9) and (2.11) were used. $`S_{WZ}`$ is invariant under supersymmetry trasformations as the second form of the equation in (3.5) is composed entirely of Cartan forms. The full supersymmetric, $`\kappa `$-invariant D1-brane action in $`AdS_3\times S^3`$ is thus given by $$S^{D1}=_{M_2}d^2\sigma \sqrt{det(G_{ij}+F_{ij})}2i_{M_2}_0^1𝑑t\overline{\theta }\widehat{L}_t𝒥L_t.$$ (3.7) The results of applying the Killing gauge of to this action are contained in Appendix C. This gauge was adapted to $`AdS_3\times S^3`$ in . The results are not contained in the main body of the paper as they do not simplify the action as greatly as for the fundamental string (especially for the D5-brane actions found next). ## 4 D5-brane in $`AdS_3\times S^3`$ As mentioned earlier, the action described here is for the D5-brane lying entirely in $`AdS_3\times S^3`$, which can be interpreted as a D5-brane in $`AdS_3\times S^3\times ^4`$ with constraints imposed on the fields such that the solutions of the action must lie in $`AdS_3\times S^3`$, with the $`^4`$ compactified away. Such constraints on a D5-brane solution in $`AdS_3\times S^3\times ^4`$ describe branes in a BPS configuration . This action does not have solutions that describe the branes whose gravity is warping spacetime to an $`AdS_3\times S^3\times ^4`$ background. Such a D5-brane, with 4 directions compactified in $`^4`$, would have an action like the one of the D1-brane in $`AdS_3\times S^3`$ just found (3.7). This case of a D5-brane lying entirely in $`AdS_3\times S^3`$, with $`^4`$ being compactified away, has similarities to the case studied in where a D9-brane filling 10-dimensional space is investigated. The D5-brane has a BI-action of $$S_{BI}=_{M_6}d^6\sigma \sqrt{det(G_{ij}+F_{ij})}i,j\{0,\mathrm{},5\}.$$ (4.1) This time the operator $`\mathrm{\Gamma }`$ that defines the $`\kappa `$-invariance is $$\mathrm{\Gamma }=\frac{\epsilon ^{i_1\mathrm{}i_6}}{_{BI}}\left(\frac{\mathrm{\Gamma }_{i_1\mathrm{}i_6}}{6!}𝒥+\frac{\mathrm{\Gamma }_{i_1\mathrm{}i_4}F_{i_5i_6}}{2.4!}+\frac{\mathrm{\Gamma }_{i_1i_2}F_{i_3i_4}F_{i_5i_6}}{2^4}𝒥+\frac{F_{i_1i_2}F_{i_3i_4}F_{i_5i_6}}{3!.2^3}\right).$$ (4.2) Using this the same procedure as for the D1-brane is followed. The variation $`\delta _\kappa S_{BI}`$ turns out to be given by $$\delta _\kappa S_{BI}=2i_{M_6}\left(\frac{(\overline{\kappa }(\widehat{L})^5𝒥L)}{5!}+\frac{(\overline{\kappa }(\widehat{L})^3L)F}{3!}+\frac{(\overline{\kappa }\widehat{L}𝒥L)FF}{2}\right).$$ (4.3) The WZ term whose $`\kappa `$-variation cancels this is $$S_{WZ}=2i_{M_6}_0^1𝑑t\left(\frac{(\overline{\theta }(\widehat{L})^5𝒥L)}{5!}+\frac{(\overline{\theta }(\widehat{L})^3L)F}{3!}+\frac{(\overline{\theta }\widehat{L}𝒥L)FF}{2}\right).$$ (4.4) The $`\kappa `$-variation of this is $$\delta _\kappa S_{WZ}=2i_{M_6}\left(\frac{(\overline{\kappa }(\widehat{L})^5𝒥L)}{5!}+\frac{(\overline{\kappa }(\widehat{L})^3L)F}{3!}+\frac{(\overline{\kappa }\widehat{L}𝒥L)FF}{2}\right)+Y,$$ (4.5) where $$Y=_{M_6}i_\kappa _0^1𝑑t\left(\frac{8}{5!}(\overline{\theta }\sigma _{}((L_t^a\mathrm{\Gamma }_a)^5L_t^b^{}\mathrm{\Gamma }_b^{}L_t^a\mathrm{\Gamma }_a(L_t^b^{}\mathrm{\Gamma }_b^{})^5)𝒦L_t)\right)$$ $$+_{M_6}i_\kappa _0^1𝑑t\left(\frac{1}{3}(\overline{\theta }\sigma _{}((L_t^a\mathrm{\Gamma }_a)^4(L_t^a^{}\mathrm{\Gamma }_a^{})^4)L_t)F_t\right),$$ and where $`i_\kappa `$ is from $`\delta _\kappa =\{d,i_\kappa \}`$ and $`i_\kappa L=\kappa `$, $`i_\kappa L^{\widehat{a}}=0`$. Due to the indices $`a\{0,1,2\}`$ and $`a^{}\{3,4,5\}`$ having only 3 possibilities each, $`Y=0`$. The total effective action of a D5-brane in $`AdS_3\times S^3`$ is thus given by $$S^{D5}=_{M_6}d^6\sigma \sqrt{det(G_{ij}+F_{ij})}$$ $$2i_{M_6}_0^1𝑑t\left(\frac{(\overline{\theta }(\widehat{L}_t)^5𝒥L_t)}{5!}+\frac{(\overline{\theta }(\widehat{L}_t)^3L_t)F_t}{3!}+\frac{(\overline{\theta }\widehat{L}_t𝒥L_t)F_tF_t}{2}\right).$$ (4.6) When calculating this the Fierz identity eqn (2.11) was required as well as identities $`(\mathrm{\Gamma }^{\widehat{a}\widehat{b}\widehat{c}})_{(\widehat{\alpha }\widehat{\beta }}(\mathrm{\Gamma }_{\widehat{c}})_{\widehat{\gamma }\widehat{\delta })}2(\mathrm{\Gamma }^{[\widehat{a}})_{(\widehat{\alpha }\widehat{\beta }}(\mathrm{\Gamma }^{\widehat{b}]}𝒦)_{\widehat{\gamma }\widehat{\delta })}=0,`$ (4.7) $`(\mathrm{\Gamma }^{\widehat{a}\widehat{b}\widehat{c}\widehat{d}\widehat{e}}𝒥)_{(\widehat{\alpha }\widehat{\beta }}(\mathrm{\Gamma }_{\widehat{e}})_{\widehat{\gamma }\widehat{\delta })}4(\mathrm{\Gamma }^{[\widehat{a}\widehat{b}\widehat{c}})_{(\widehat{\alpha }\widehat{\beta }}(\mathrm{\Gamma }^{\widehat{d}]}𝒦)_{\widehat{\gamma }\widehat{\delta })}=0.`$ (4.8) The Killing guage fixed D5-brane action in $`AdS_3\times S^3`$ is given in Appendix C. ## 5 D5-brane in $`AdS_5\times S^5`$ The mathematics in this case is almost identical to the previous case (this is due to the intentional similarities in the notation), however in this case $`Y`$ is not zero, but we can use $$ϵ^{a_1\mathrm{}a_5}=i\sigma _1\mathrm{\Gamma }^{a_1\mathrm{}a_5},ϵ^{a_1^{}\mathrm{}a_5^{}}=\sigma _2\mathrm{\Gamma }^{a_1^{}\mathrm{}a_5^{}},$$ (5.1) to show that $`S^{D5}`$ $`=`$ $`{\displaystyle _{M_6}}d^6\sigma \sqrt{det(G_{ij}+F_{ij})}`$ (5.2) $`i{\displaystyle _{M_7}}\left({\displaystyle \frac{(\overline{L}(\widehat{L})^5𝒥L)}{5!}}+{\displaystyle \frac{(\overline{L}(\widehat{L})^3L)F}{3!}}+{\displaystyle \frac{(\overline{L}\widehat{L}𝒥L)FF}{2}}\right)`$ $`+{\displaystyle _{M_7}}\left({\displaystyle \frac{ϵ_{a_1\mathrm{}a_5}}{30}}L^{a_1}\mathrm{}L^{a_5}F+{\displaystyle \frac{ϵ_{a_1^{}\mathrm{}a_5^{}}}{30}}L^{a_1^{}}\mathrm{}L^{a_5^{}}F\right),`$ is the full $`\kappa `$-invariant, supersymmetric action. Eqn (5.2) can be rewritten as $$S^{D5}=S_{BI}2i_{M_6}_0^1𝑑t\left(\frac{(\overline{\theta }(\widehat{L}_t)^5𝒥L_t)}{5!}+\frac{(\overline{\theta }(\widehat{L}_t)^3L_t)F_t}{3!}+\frac{(\overline{\theta }\widehat{L}_t𝒥L_t)F_tF_t}{2}\right)$$ $$+_{M_7}_{WZ}^{BOSE},$$ (5.3) where $$_{WZ}^{BOSE}=_{WZ}|_{\theta =0}=\frac{ϵ_{a_1\mathrm{}a_5}}{30}e^{a_1}\mathrm{}e^{a_5}dA+\frac{ϵ_{a_1^{}\mathrm{}a_5^{}}}{30}e^{a_1^{}}\mathrm{}e^{a_5^{}}dA.$$ (5.4) The Killing guage fixed D5-brane action in $`AdS_5\times S^5`$ is contained in Appendix C. ## 6 D5-brane in $`AdS_3\times S^3\times T^4`$ As mentioned in the introduction, an attempt was made to find the action of a general D5-brane free to move in all the dimensions of $`AdS_3\times S^3\times T^4`$. The same method as used to find the other actions in this paper is used, however as will be shown, this method does not succeed in producing a fully $`\kappa `$-invariant action in this case. The case of extending the D1-brane action to $`AdS_3\times S^3\times T^4`$ is trivial, and in the notation of Appendix A, appears unchanged to (3.7). As mentioned in Appendix A, the coset representation for $`T^4`$ that is used here describes only translations in $`T^4`$ (ie $`U(1)^4`$). The D5-brane BI-action is $`S_{BI}={\displaystyle _{M_6}}d^6\sigma \sqrt{det(G_{ij}+F_{ij})},i,j\{0,\mathrm{},5\},`$ $`G_{ij}=L_i^aL_{ja}+L_i^a^{}L_{ja^{}}+L_i^{a^{\prime \prime }}L_{ja^{\prime \prime }},`$ (6.1) $`F=dA+2i{\displaystyle _0^1}𝑑t\overline{\theta }(L_t^a\mathrm{\Gamma }_a+L_t^a^{}\mathrm{\Gamma }_a^{}+L_t^{^{\prime \prime }}\mathrm{\Gamma }_{a^{\prime \prime }})𝒦L_t,`$ where $`\mathrm{\Gamma }`$ in given by (4.2). The $`\kappa `$-variation of the BI-action appears the same as (4.3) in the $`AdS_3\times S^3\times T^4`$ notation of Appendix A. The WZ-action required to cancel this should be of the form $`S_{WZ}^{D5}`$ $`=`$ $`2i{\displaystyle _{M_6}}{\displaystyle _0^1}𝑑t\left({\displaystyle \frac{(\overline{\theta }(\widehat{L}_t)^5𝒥L_t)}{5!}}+{\displaystyle \frac{(\overline{\theta }(\widehat{L}_t)^3L_t)F_t}{3!}}+{\displaystyle \frac{(\overline{\theta }\widehat{L}_t𝒥L_t)F_tF_t}{2}}\right)`$ (6.3) $`+{\displaystyle _{M_7}}E_7^{BOSE}`$ $`=`$ $`i{\displaystyle _{M_7}}\left({\displaystyle \frac{(\overline{L}(\widehat{L})^5𝒥L)}{5!}}+{\displaystyle \frac{(\overline{L}(\widehat{L})^3L)F}{3!}}+{\displaystyle \frac{(\overline{L}\widehat{L}𝒥L)FF}{2}}\right)`$ $`+{\displaystyle _{M_7}}E_7,`$ where $`E_7`$ should have the following properties to make the action $`\kappa `$-invariant * $`E_7`$ should be supersymmetric. * $`i_\kappa E_7=0`$. * $$\delta _\kappa E_7=\frac{8}{5!}(\overline{\kappa }\sigma _{}((L^a\mathrm{\Gamma }_a)^5L^{\stackrel{~}{b}}\mathrm{\Gamma }_{\stackrel{~}{b}}L^a\mathrm{\Gamma }_a(L^{\stackrel{~}{b}}\mathrm{\Gamma }_{\stackrel{~}{b}})^5)𝒦L)$$ $$\frac{1}{3}(\overline{\kappa }\sigma _{}((L^a\mathrm{\Gamma }_a)^4(L^{\stackrel{~}{a}}\mathrm{\Gamma }_{\stackrel{~}{a}})^4)L)F$$ (6.4) where $`\stackrel{~}{a}\{a^{},a^{\prime \prime }\}`$. No supersymmetric 7 form can be constructed out of the Cartan forms and $`F`$ that satisfies these conditions. ## 7 Summary In this paper the fully $`\kappa `$-invariant and supersymmetric actions were found for the D1 and D5-branes in $`AdS_3\times S^3`$ and the D5 action in $`AdS_5\times S^5`$. The solutions of the D5 action in $`AdS_3\times S^3`$ are a BPS system as they are S-dual to the two intersecting NS-five branes of , which are BPS. This D5-brane has no dimensions compactified on $`^4`$. A D5-brane that is parallel to the branes that are the source of the background field would have four dimensions compactified on $`^4`$ and the action would be the same as the D1-brane action. The D5-brane action in $`AdS_5\times S^5`$ that was found is of interest as it can be adjusted to describe the D5, D3, F1 configuration of which exhibits a baryon vertex, by placing some restrictions on the degrees of freedom. A fully $`\kappa `$-symmetric D5 action in $`AdS_3\times S^3\times T^4`$ failed to be found in this case, however I do not believe that this is a reason to believe that it does not exist. The WZ-action may take a form very different to more standard D-brane actions. It should be said though that most of the interesting dynamics occurs in the $`AdS_3\times S^3`$, not in $`T^4`$, so the action found in (3.7) is sufficient for most purposes. In Appendix C the Killing gauge was applied to the actions, however even though the D1-brane and D5-brane actions in $`AdS_3\times S^3`$ are reduced to being eigth order in fermions, they are still somewhat complicated. The D5-brane in $`AdS_5\times S^5`$ remains very complicated even in this gauge. It remains to be seen if some method, such as the T-duality transformation of can be used to simplify the actions. Acknowledgements I would like to thank my supervisor Peter Bouwknegt for providing useful guidance and R.R. Metsaev and A.A. Tseytlin for several invaluable correspondences. My research has been supported by an Australian Postgraduate Award scholarship. ## Appendix A Notation and Algebra for Various $`AdS\times S`$ Spaces The algebra follows along the lines of . ### A.1 $`AdS_3\times S^3\times T^4`$ The coset representation for $`T^4`$ that is used here describes only translations. That is, $`J_{a^{\prime \prime }b^{\prime \prime }}`$ are absent from the algebra. It is not certain if this situation was described in an ideal fashion. The notation and algebra used (with the radii of the spaces $`AdS_3`$, $`S^3`$ and $`T^4`$ set to 1) is as follows * $`a,b,c\{0,1,2\}`$ are indices of $`SO(2,1)`$ ($`AdS_3`$). * $`a^{},b^{},c^{}\{3,4,5\}`$ are indices of $`SO(3)`$ ($`S^3`$). * $`a^{\prime \prime },b^{\prime \prime },c^{\prime \prime }\{6,7,8,9\}`$ are indices of $`U(1)^4`$ ($`T^4`$). * $`\widehat{a},\widehat{b},\widehat{c}\{0,\mathrm{},9\}`$ are a combination of $`(a,b,c)`$, $`(a^{},b^{},c^{})`$ and $`(a^{\prime \prime },b^{\prime \prime },c^{\prime \prime })`$. * $`\alpha ,\beta ,\gamma \{1,2\}`$ are $`SO(2,1)`$ spinor indices. * $`\alpha ^{},\beta ^{},\gamma ^{}\{1,2\}`$ are $`SO(3)`$ spinor indices. * $`\alpha ^{\prime \prime },\beta ^{\prime \prime },\gamma ^{\prime \prime }\{1,\mathrm{},4\}`$ are $`SO(4)`$ spinor indices. * $`\widehat{\alpha },\widehat{\beta },\widehat{\gamma }\{1,\mathrm{},32\}`$ are $`D=10`$ Majorana-Weyl spinor indices. * $`I,J,K\{1,2\}`$ are $`𝒩=2`$ supersymmetry labels. * $`(P_a,J_{ab})`$, $`(P_a^{},J_{a^{}b^{}})`$ and $`(P_{a^{\prime \prime }},J_{a^{\prime \prime }b^{\prime \prime }})`$ are the $`SO(2,1)`$, $`SO(3)`$ and $`SO(4)`$ translations and rotations respectively. They are defined to be antihermitian. * $`32\times 32`$ $`\mathrm{\Gamma }`$ matrices of $`AdS_3\times S^3\times T^4`$ are $`\mathrm{\Gamma }^a=\gamma ^a1_21_4\sigma ^1,\gamma ^a:\gamma ^0=i\sigma ^3,\gamma ^1=\sigma ^1,\gamma ^2=\sigma ^2,`$ (A.1) $`\mathrm{\Gamma }^a^{}=1_2\gamma ^a^{}\overline{\gamma }_5\sigma ^2,\gamma ^a^{}:\gamma ^3^{}=\sigma ^1,\gamma ^4^{}=\sigma ^2,\gamma ^5^{}=\sigma ^3,\overline{\gamma }_5=\gamma _6\gamma _7\gamma _8\gamma _9,`$ $`\mathrm{\Gamma }^{a^{\prime \prime }}=1_21_2\gamma ^{a^{\prime \prime }}\sigma ^2.`$ * $`\pi _+=1_21_21_4\frac{1}{2}(1_2+\sigma _3),\sigma _+=1_21_21_4\frac{1}{2}(\sigma _1+i\sigma _2)`$ $`\sigma _{}=1_21_21_4\frac{1}{2}(\sigma _1i\sigma _2)`$ * $`\widehat{𝐂}=CC^{}C^{\prime \prime }i\sigma ^2`$ is the charge conjugation matrix, where $`C`$, $`C^{}`$ and $`C^{\prime \prime }`$ are the charge conjugation matrices of $`SO(2,1)`$, $`SO(3)`$ and $`SO(4)`$. * $`\{\mathrm{\Gamma }^{\widehat{a}},\mathrm{\Gamma }^{\widehat{b}}\}=2\eta ^{\widehat{a}\widehat{b}}`$ where $`\eta ^{\widehat{a}\widehat{b}}=(,+,\mathrm{},+)`$. * Supersymmetry generators $`Q^{I\widehat{\alpha }}=\left(\begin{array}{c}0\\ Q^{I\alpha \alpha ^{}\alpha ^{\prime \prime }}\end{array}\right)`$. In this notation $`\theta _{I\widehat{\alpha }}`$ and $`L_{I\widehat{\alpha }}`$ have the opposite chirality: $`L^{I\widehat{\alpha }}=\left(\begin{array}{c}L^{I\alpha \alpha ^{}\alpha ^{\prime \prime }}\\ 0\end{array}\right)`$. * $`Q^{1\widehat{\alpha }}`$ and $`Q^{2\widehat{\alpha }}`$ can be combined as $`Q=\left(\begin{array}{c}Q^1\\ Q^2\end{array}\right)`$. * $$=\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right)𝒥=\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right)𝒦=\left(\begin{array}{cc}1& 0\\ 0& 1\end{array}\right)\mathrm{act}\mathrm{on}Q=\left(\begin{array}{c}Q^1\\ Q^2\end{array}\right).$$ (A.2) The algebra $`[P_a,P_b]=J_{ab},[P_a^{},P_b^{}]=J_{a^{}b^{}},[P_{a^{\prime \prime }},P_{b^{\prime \prime }}]=0,`$ $`[P_a,J_{bc}]=\eta _{ab}P_c\eta _{ac}P_b,[P_a^{},J_{b^{}c^{}}]=\eta _{a^{}b^{}}P_c^{}\eta _{a^{}c^{}}P_b^{},`$ $`[J_{ab},J_{cd}]=\eta _{bc}J_{ad}+\eta _{ad}J_{bc}\eta _{ac}J_{bd}\eta _{bd}J_{ac},`$ (A.3) $`[J_{a^{}b^{}},J_{c^{}d^{}}]=\eta _{b^{}c^{}}J_{a^{}d^{}}+\eta _{a^{}d^{}}J_{b^{}c^{}}\eta _{a^{}c^{}}J_{b^{}d^{}}\eta _{b^{}d^{}}J_{a^{}c^{}},`$ $`[Q,P_{\widehat{a}}]={\displaystyle \frac{i}{2}}Q\sigma _+\mathrm{\Gamma }_{\widehat{a}},[Q,J_{\widehat{a}\widehat{b}}]={\displaystyle \frac{1}{2}}Q\mathrm{\Gamma }_{\widehat{a}\widehat{b}}\mathrm{if}\widehat{a},\widehat{b}5,`$ $`\{Q_{\widehat{\alpha }},Q_{\widehat{\beta }}\}=2i(\widehat{C}\mathrm{\Gamma }^{\widehat{a}}\pi _+)_{\widehat{\alpha }\widehat{\beta }}P_{\widehat{a}}+[(\widehat{C}\mathrm{\Gamma }^{ab}\sigma _{})_{\widehat{\alpha }\widehat{\beta }}J_{ab}(\widehat{C}\mathrm{\Gamma }^{a^{}b^{}}\sigma _{})_{\widehat{\alpha }\widehat{\beta }}J_{a^{}b^{}}].`$ ### A.2 $`AdS_3\times S^3`$ The notation and algebra for just $`AdS_3\times S^3`$ space (ie, assuming the $`T^4`$ is compactified to zero volume limit and ignoring it) is almost identical to the case of the more general $`AdS_3\times S^3\times T^4`$, except any terms with double primed indices are absent, thus any difficulties in treating The notable changes besides this are * $`\widehat{\alpha },\widehat{\beta },\widehat{\gamma }\{1,\mathrm{},8\}`$ are $`D=6`$ complex chiral spinor indices. * $`8\times 8`$ $`\mathrm{\Gamma }`$ matrices of $`AdS_3\times S^3`$ are $`\mathrm{\Gamma }^a`$ $`=`$ $`\gamma ^a1\sigma ^1,\gamma ^a:\gamma ^0=i\sigma ^3,\gamma ^1=\sigma ^1,\gamma ^2=\sigma ^2,`$ (A.4) $`\mathrm{\Gamma }^a^{}`$ $`=`$ $`1\gamma ^a^{}\sigma ^2,\gamma ^a^{}:\gamma ^3^{}=\sigma ^1,\gamma ^4^{}=\sigma ^2,\gamma ^5^{}=\sigma ^3.`$ ### A.3 $`AdS_5\times S^5`$ The third set of notation and algebra we use is for $`AdS_5\times S^5`$. This is exactly the same notation as used in . * $`a,b,c\{0,\mathrm{},4\}`$ are indices of $`SO(4,1)`$. * $`a^{},b^{},c^{}\{5,\mathrm{},9\}`$ are indices of $`SO(5)`$. * $`\widehat{a},\widehat{b},\widehat{c}\{0,\mathrm{},9\}`$ are a combination of $`(a,b,c)`$ and $`(a^{},b^{},c^{})`$. * $`\alpha ,\beta ,\gamma \{1,\mathrm{},4\}`$ are $`SO(4,1)`$ spinor indices. * $`\alpha ^{},\beta ^{},\gamma ^{}\{1,\mathrm{},4\}`$ are $`SO(5)`$ spinor indices. * $`\widehat{\alpha },\widehat{\beta },\widehat{\gamma }\{1,\mathrm{},32\}`$ are $`D=10`$ Majorana-Weyl spinor indices. * $`I,J,K\{1,2\},𝒩=2`$ supersymmetry labels. * $`(P_a,J_{ab})`$, $`(P_a^{},J_{a^{}b^{}})`$ and are the $`SO(4,1)`$ and $`SO(5)`$ translations and rotations respectively. They are defined to be antihermitian. * $`32\times 32`$ $`\mathrm{\Gamma }`$ matrices of $`AdS_5\times S^5`$ are $`\mathrm{\Gamma }^a`$ $`=`$ $`\gamma ^a1_4\sigma ^1,\gamma ^a:\gamma ^0=i\sigma ^31_2,\gamma ^{1,\mathrm{},3}=\sigma ^2\sigma ^{1,\mathrm{},3},\gamma ^4=\sigma ^11_2,`$ $`\mathrm{\Gamma }^a^{}`$ $`=`$ $`1_4\gamma ^a^{}\sigma ^2,\gamma ^a^{}:\gamma ^{5,\mathrm{},7}=\sigma ^2\sigma ^{1,\mathrm{},3},\gamma ^8=\sigma ^31_2,\gamma ^9=\sigma ^11_2.`$ * $`\widehat{𝐂}=CC^{}i\sigma ^2`$ is the charge conjugation matrix, where $`C`$ and $`C^{}`$ are the charge conjugation matrices of $`SO(4,1)`$ and $`SO(5)`$. * $`\{\mathrm{\Gamma }^{\widehat{a}},\mathrm{\Gamma }^{\widehat{b}}\}=2\eta ^{\widehat{a}\widehat{b}}`$ where $`\eta ^{\widehat{a}\widehat{b}}=(,+,\mathrm{},+)`$. * Supersymmetry generators $`Q^{I\widehat{\alpha }}=\left(\begin{array}{c}0\\ Q^{I\alpha \alpha ^{}}\end{array}\right)`$. In this notation $`\theta _{I\widehat{\alpha }}`$ and $`L_{I\widehat{\alpha }}`$ have the opposite chirality: $`L^{I\widehat{\alpha }}=\left(\begin{array}{c}L^{I\alpha \alpha ^{}}\\ 0\end{array}\right)`$. * $`Q^{1\widehat{\alpha }}`$ and $`Q^{2\widehat{\alpha }}`$ can be combined as $`Q=\left(\begin{array}{c}Q^1\\ Q^2\end{array}\right)`$. ## Appendix B Maurer-Cartan Equations and other Cartan form relations The Maurer-Cartan Equations, the variations for the Cartan forms and their $`_tL`$ equations (which will be explained below) for $`AdS_3\times S^3\times T^4`$, $`AdS_3\times S^3`$ and $`AdS_5\times S^5`$ can all be written out in the same form, as long as it is understood that the indices run over different ranges for each space, and that $`L^{a^{\prime \prime }b^{\prime \prime }}=0`$ for $`AdS_3\times S^3\times T^4`$. The equations of course appear exactly the same as the corresponding equations in . $`dL^{\widehat{a}}=L^{\widehat{a}\widehat{b}}L_{\widehat{b}}i\overline{L}\mathrm{\Gamma }^{\widehat{a}}L`$ (B.1) $`dL={\displaystyle \frac{i}{2}}\sigma _+\widehat{L}L{\displaystyle \frac{1}{4}}L^{\widehat{a}\widehat{b}}\mathrm{\Gamma }_{\widehat{a}\widehat{b}}L`$ (B.2) $`dL^{ab}=L^aL^bL^{ac}L_c^b+\overline{L}\mathrm{\Gamma }^{ab}\sigma _{}L`$ (B.3) $`dL^{a^{}b^{}}=Ł^a^{}L^b^{}L^{a^{}c^{}}L_c^{}^b^{}\overline{L}\mathrm{\Gamma }^{a^{}b^{}}\sigma _{}L`$ (B.4) The variations of the Cartan forms are $`\delta L^{\widehat{a}}=d\delta x^{\widehat{a}}+L^{\widehat{a}\widehat{b}}\delta x_{\widehat{b}}\delta x^{\widehat{a}\widehat{b}}L_{\widehat{b}}+2i\overline{L}\mathrm{\Gamma }^{\widehat{a}}\delta \theta ,`$ (B.5) $`\delta L=d\delta \theta {\displaystyle \frac{i}{2}}\sigma _+\widehat{L}\delta \theta +{\displaystyle \frac{1}{4}}L^{\widehat{a}\widehat{b}}\mathrm{\Gamma }_{\widehat{a}\widehat{b}}\delta \theta +{\displaystyle \frac{i}{2}}\sigma _+\delta x^{\widehat{a}}\mathrm{\Gamma }_{\widehat{a}}L{\displaystyle \frac{1}{4}}\delta x^{\widehat{a}\widehat{b}}\mathrm{\Gamma }_{\widehat{a}\widehat{b}}L,`$ (B.6) $`\delta x^{\widehat{a}}\delta X^ML_M^{\widehat{a}},\delta x^{\widehat{a}\widehat{b}}\delta X^ML_M^{\widehat{a}\widehat{b}},\delta \theta \delta X^ML_M^{\widehat{a}\widehat{b}}.`$ (B.7) Rescale $`\theta t\theta `$ to get the $`L_t`$ Cartan forms. $$L_t^{\widehat{a}}(x,\theta )L^{\widehat{a}}(x,t\theta ),L_t^{\widehat{a}\widehat{b}}(x,\theta )L^{\widehat{a}\widehat{b}}(x,t\theta ),L_t(x,\theta )L(x,t\theta ).$$ (B.8) The initial conditions of these are $$L_{t=0}^{\widehat{a}}=e^{\widehat{a}},L_{t=0}^{\widehat{a}\widehat{b}}=\omega ^{\widehat{a}\widehat{b}},L_{t=0}=0.$$ (B.9) It can be shown that $$_tF_t=2i\overline{\theta }\widehat{L}_t𝒦L_t.$$ (B.10) The defining equations for the Cartan forms can be shown to be $`_tL_t^{\widehat{a}}=2i\overline{\theta }\mathrm{\Gamma }^{\widehat{a}}L_t,`$ (B.11) $`_tL_t=d\theta {\displaystyle \frac{i}{2}}\sigma _+\widehat{L}\theta +{\displaystyle \frac{1}{4}}\mathrm{\Gamma }_{\widehat{a}\widehat{b}}L_t^{\widehat{a}\widehat{b}},`$ (B.12) $`_tL_t^{ab}=2\overline{\theta }\mathrm{\Gamma }^{ab}\sigma _{}L_t,_tL_t^{a^{}b^{}}=2\overline{\theta }\mathrm{\Gamma }^{a^{}b^{}}\sigma _{}L_t.`$ (B.13) The solutions to these equations can be found in equations B.14 - B.18 of and using the Killing gauge they are symplified in . ## Appendix C Killing Gauge Fixed Actions In this appendix the Killing gauge of is applied to the D1-brane in $`AdS_3\times S^3`$, and the D5-brane actions in both $`AdS_3\times S^3`$ and $`AdS_5\times S^5`$. ### C.1 D1-brane in $`AdS_3\times S^3`$ The Killing gauge is defined by $`𝒫_\pm ^{IJ}={\displaystyle \frac{1}{2}}(\delta ^{IJ}\pm \mathrm{\Gamma }_{}ϵ^{IJ}),`$ (C.1) $`𝒫_{}\theta =0,`$ (C.2) where $`\mathrm{\Gamma }_{}=i\mathrm{\Gamma }_{01}`$ in $`AdS_3\times S^3`$. The basis used to represent the spinors in the simplified action is $`\theta _\pm =𝒫_\pm \theta ,`$ (C.3) $`\theta _\pm ^1={\displaystyle \frac{1}{2}}(\theta ^1\pm \mathrm{\Gamma }_{}\theta ^2),`$ (C.4) $`\theta _\pm ^2={\displaystyle \frac{1}{2}}(\theta ^2\mathrm{\Gamma }_{}\theta ^1)=\mathrm{\Gamma }_{}\theta _\pm ^1.`$ (C.5) Letting $`p0,1`$ be the coordinates along the background D1-branes and $`q\{2,\mathrm{},5\}`$ be the transverse coordinates, $`\theta `$ and the Cartan forms can be shown to simplify to $`\theta _+^I=\sqrt{|y|}\eta _+^I,`$ (C.6) $`L_{t+}^I=t\sqrt{|y|}d\eta _+^I,`$ (C.7) $`L_{}^I=0,`$ (C.8) $`L_t^p=|y|(dx^pit^2\overline{\eta }_+^I\mathrm{\Gamma }^pd\eta _+^I),`$ (C.9) $`L_t^q={\displaystyle \frac{1}{|y|}}dy^q.`$ (C.10) Using this, the BI-action is $`S_{BI}={\displaystyle _{M_2}}d^2\sigma \sqrt{det(G_{ij}+F_{ij})},`$ (C.11) $`G_{ij}=|y|^2(_ix_p2i(\overline{\eta }_+^1\mathrm{\Gamma }_p_i\eta _+^1))(_jx^p2i(\overline{\eta }_+^1\mathrm{\Gamma }^p_j\eta _+^1))+{\displaystyle \frac{1}{|y|^2}}dy^qdy^q,`$ (C.12) $`ϵ^{ij}F_{ij}=ϵ^{ij}(2_iA_j+4i(\overline{\eta }_+^1\mathrm{\Gamma }_q_iy^q_j\eta _+^1)),`$ (C.13) while the WZ-action is found to be $$S_{WZ}=2_{M_2}(\overline{\eta }_+^1(dy^q\mathrm{\Gamma }_q)\mathrm{\Gamma }_{01}d\eta _+^1).$$ (C.14) ### C.2 D5-brane in $`AdS_3\times S^3`$ The BI-action of the D5-brane looks identical to that of the D1-brane, except of course that $`i,j\{0,\mathrm{},5\}`$ instead of $`\{0,1\}`$. The WZ-action is quite complicated in this gauge, however some terms disappear as the Killing gauge leaves only 8 fermionic degrees of freedom, so no term can be higher than eighth order in $`\eta _+^1`$. $`S_{WZ}=2i{\displaystyle _{M_6}}[{\displaystyle \frac{2}{5!}}\overline{\eta }_+^1({\displaystyle \frac{1}{2|y|^4}}(dy^q\mathrm{\Gamma }_q)^5+10(dy^q\mathrm{\Gamma }_q)^3({\displaystyle \frac{1}{2}}(dx^p\mathrm{\Gamma }_p)^2`$ $`i(dx^p\mathrm{\Gamma }_p)((\overline{\eta }_+^1\mathrm{\Gamma }^pd\eta _+^1)\mathrm{\Gamma }_p){\displaystyle \frac{2}{3}}((\overline{\eta }_+^1\mathrm{\Gamma }^pd\eta _+^1)\mathrm{\Gamma }_p)^2)+5|y|^5(dy^q\mathrm{\Gamma }_q)({\displaystyle \frac{1}{2}}(dx^p\mathrm{\Gamma }_p)^4`$ $`2i(dx^p\mathrm{\Gamma }_p)^3((\overline{\eta }_+^1\mathrm{\Gamma }^pd\eta _+^1)\mathrm{\Gamma }_p)4(dx^p\mathrm{\Gamma }_p)^2((\overline{\eta }_+^1\mathrm{\Gamma }^pd\eta _+^1)\mathrm{\Gamma }_p)^2+4i(dx^p\mathrm{\Gamma }_p)`$ $`((\overline{\eta }_+^1\mathrm{\Gamma }^pd\eta _+^1)\mathrm{\Gamma }_p)^3))\mathrm{\Gamma }_{}d\eta _+^1`$ $`+{\displaystyle \frac{1}{3!}}\overline{\eta }_+^1(2|y|^4({\displaystyle \frac{1}{2}}(dx^p\mathrm{\Gamma }_p)^3{\displaystyle \frac{3i}{2}}(dx^p\mathrm{\Gamma }_p)^2((\overline{\eta }_+^1\mathrm{\Gamma }^pd\eta _+^1)\mathrm{\Gamma }_p)2(dx^p\mathrm{\Gamma }_p)((\overline{\eta }_+^1\mathrm{\Gamma }^pd\eta _+^1)\mathrm{\Gamma }_p)^2`$ $`+i((\overline{\eta }_+^1\mathrm{\Gamma }^pd\eta _+^1)\mathrm{\Gamma }_p)^3)3(dx^p\mathrm{\Gamma }_p)(dy^q\mathrm{\Gamma }_q)^2+3i(dy^q\mathrm{\Gamma }_q)^2((\overline{\eta }_+^1\mathrm{\Gamma }^pd\eta _+^1)\mathrm{\Gamma }_p))\mathrm{\Gamma }_{}d\eta _+^1dA`$ $`+{\displaystyle \frac{i}{3}}\overline{\eta }_+^1(2|y|^4({\displaystyle \frac{1}{4}}(dx^p\mathrm{\Gamma }_p)^3i(dx^p\mathrm{\Gamma }_p)^2((\overline{\eta }_+^1\mathrm{\Gamma }^pd\eta _+^1)\mathrm{\Gamma }_p){\displaystyle \frac{3}{2}}(dx^p\mathrm{\Gamma }_p)((\overline{\eta }_+^1\mathrm{\Gamma }^pd\eta _+^1)\mathrm{\Gamma }_p)^2)`$ $`{\displaystyle \frac{3}{2}}(dx^p\mathrm{\Gamma }_p)(dy^q\mathrm{\Gamma }_q)^2+2i((\overline{\eta }_+^1\mathrm{\Gamma }^pd\eta _+^1)\mathrm{\Gamma }_p)(dy^q\mathrm{\Gamma }_q)^2)\mathrm{\Gamma }_{}d\overline{\eta }_+^1`$ $`(\overline{\eta }_+^1(dy^q\mathrm{\Gamma }_q)d\eta _+^1)(\overline{\eta }_+^1(dy^q\mathrm{\Gamma }_q)\mathrm{\Gamma }_{}d\eta _+^1)({\displaystyle \frac{1}{2}}dAdA+idA(\overline{\eta }_+^1(dy^q\mathrm{\Gamma }_q)d\eta _+^1)`$ $`{\displaystyle \frac{2}{3}}(\overline{\eta }_+^1(dy^q\mathrm{\Gamma }_q)d\eta _+^1)^2)].`$ (C.15) ### C.3 Killing Gauge Fixed D5-brane Action in $`AdS_5\times S^5`$ Again the BI-action is written as eqn (C.11) with $`i,j\{0,\mathrm{},5\}`$, and the WZ-action is found to be very similar to the $`AdS_3\times S^3`$ case, except that now $`\mathrm{\Gamma }_{}=\mathrm{\Gamma }_{0123}`$ and the WZ-action contains some extra terms due to the greater number of fermionic degrees of freedom, and also due to eqn (5.4). $`S_{WZ}=2i{\displaystyle _{M_6}}[{\displaystyle \frac{2}{5!}}\overline{\eta }_+^1({\displaystyle \frac{1}{2|y|^4}}(dy^q\mathrm{\Gamma }_q)^5+10(dy^q\mathrm{\Gamma }_q)^3({\displaystyle \frac{1}{2}}(dx^p\mathrm{\Gamma }_p)^2`$ $`i(dx^p\mathrm{\Gamma }_p)((\overline{\eta }_+^1\mathrm{\Gamma }^pd\eta _+^1)\mathrm{\Gamma }_p){\displaystyle \frac{2}{3}}((\overline{\eta }_+^1\mathrm{\Gamma }^pd\eta _+^1)\mathrm{\Gamma }_p)^2)+5|y|^5(dy^q\mathrm{\Gamma }_q)({\displaystyle \frac{1}{2}}(dx^p\mathrm{\Gamma }_p)^4`$ $`2i(dx^p\mathrm{\Gamma }_p)^3((\overline{\eta }_+^1\mathrm{\Gamma }^pd\eta _+^1)\mathrm{\Gamma }_p)4(dx^p\mathrm{\Gamma }_p)^2((\overline{\eta }_+^1\mathrm{\Gamma }^pd\eta _+^1)\mathrm{\Gamma }_p)^2+4i(dx^p\mathrm{\Gamma }_p)`$ $`((\overline{\eta }_+^1\mathrm{\Gamma }^pd\eta _+^1)\mathrm{\Gamma }_p)^3+{\displaystyle \frac{8}{5}}((\overline{\eta }_+^1\mathrm{\Gamma }^pd\eta _+^1)\mathrm{\Gamma }_p)^4))\mathrm{\Gamma }_{}d\eta _+^1`$ $`+{\displaystyle \frac{1}{3!}}\overline{\eta }_+^1(2|y|^4({\displaystyle \frac{1}{2}}(dx^p\mathrm{\Gamma }_p)^3{\displaystyle \frac{3i}{2}}(dx^p\mathrm{\Gamma }_p)^2((\overline{\eta }_+^1\mathrm{\Gamma }^pd\eta _+^1)\mathrm{\Gamma }_p)2(dx^p\mathrm{\Gamma }_p)((\overline{\eta }_+^1\mathrm{\Gamma }^pd\eta _+^1)\mathrm{\Gamma }_p)^2`$ $`+i((\overline{\eta }_+^1\mathrm{\Gamma }^pd\eta _+^1)\mathrm{\Gamma }_p)^3)3(dx^p\mathrm{\Gamma }_p)(dy^q\mathrm{\Gamma }_q)^2+3i(dy^q\mathrm{\Gamma }_q)^2((\overline{\eta }_+^1\mathrm{\Gamma }^pd\eta _+^1)\mathrm{\Gamma }_p))\mathrm{\Gamma }_{}d\eta _+^1dA`$ $`+{\displaystyle \frac{i}{3}}\overline{\eta }_+^1(2|y|^4({\displaystyle \frac{1}{4}}(dx^p\mathrm{\Gamma }_p)^3i(dx^p\mathrm{\Gamma }_p)^2((\overline{\eta }_+^1\mathrm{\Gamma }^pd\eta _+^1)\mathrm{\Gamma }_p){\displaystyle \frac{3}{2}}(dx^p\mathrm{\Gamma }_p)((\overline{\eta }_+^1\mathrm{\Gamma }^pd\eta _+^1)\mathrm{\Gamma }_p)^2`$ $`+{\displaystyle \frac{4i}{5}}((\overline{\eta }_+^1\mathrm{\Gamma }^pd\eta _+^1)\mathrm{\Gamma }_p)^3){\displaystyle \frac{3}{2}}(dx^p\mathrm{\Gamma }_p)(dy^q\mathrm{\Gamma }_q)^2+2i((\overline{\eta }_+^1\mathrm{\Gamma }^pd\eta _+^1)\mathrm{\Gamma }_p)(dy^q\mathrm{\Gamma }_q)^2)\mathrm{\Gamma }_{}d\overline{\eta }_+^1`$ $`(\overline{\eta }_+^1(dy^q\mathrm{\Gamma }_q)d\eta _+^1)(\overline{\eta }_+^1(dy^q\mathrm{\Gamma }_q)\mathrm{\Gamma }_{}d\eta _+^1)({\displaystyle \frac{1}{2}}dAdA+idA(\overline{\eta }_+^1(dy^q\mathrm{\Gamma }_q)d\eta _+^1)`$ $`{\displaystyle \frac{2}{3}}(\overline{\eta }_+^1(dy^q\mathrm{\Gamma }_q)d\eta _+^1)^2)]+{\displaystyle }_{M_7}_{WZ}^{BOSE},`$ (C.16) $`_{WZ}^{BOSE}=4|y|^3dx^0\mathrm{}dx^3dy^4dA+{\displaystyle \frac{4}{|y|^5}}dy^5\mathrm{}dy^9dA.`$ (C.17)
warning/0002/math0002190.html
ar5iv
text
# Existence of close pseudoholomorphic disks for almost complex manifolds and an application to Kobayashi-Royden pseudonorm ## Introduction Let $`(M^{2n},J)`$ be an almost complex manifold, i.e. $`J^2=\mathrm{𝟏}T^{}MTM`$. A mapping $`\mathrm{\Phi }(M_1,J_1)(M_2,J_2)`$ is called pseudoholomorphic if its differential preserves the complex multiplication in the tangent bundles: $`\mathrm{\Phi }_{}J_1=J_2\mathrm{\Phi }_{}`$. Denote by $`e=1T_0\text{C}^{_{}}`$ the unit vector. Let us also denote by $`D_R`$ the disk in $`\text{C}^{_{}}`$ of radius $`R`$, which is equipped with the standard complex structure $`J_0`$. Let $`vT_pM`$, $`p=\tau _Mv`$, with $`\tau _M:TMM`$ being used for the canonical projection. We say that a disk $`f:D_RM`$ passes in the direction $`v`$ if $`f_{}e=v`$. Due to theorem III from there exists a small pseudoholomorphic disk in the direction of an arbitrary vector $`v`$. We study non-small pseudoholomorphic disks which lie in a neighborhood of a given pseudoholomorphic disk. The main result is ###### Theorem 1 $`.`$ Let a nonconstant pseudoholomorphic disk of radius $`R`$ pass through a point $`p`$ at an almost complex manifold $`(M^{2n},J)`$: $$f_0:(D_R,J_0)(M,J),(f_0)_{}(0)e=v_00.$$ Then for every $`\epsilon >0`$ there exists a neighborhood $`𝒱=𝒱_\epsilon (v_0)`$ of the vector $`v_0TM`$ such that in the direction of each vector $`v𝒱`$ a pseudoholomorphic disk of radius $`R\epsilon `$ passes: $$f:(D_{R\epsilon },J_0)(M,J),f_{}(0)e=v.$$ This theorem has an important application in the theory of invariant metrics. In 1967 Kobayashi introduced a pseudodistance on complex manifolds, which is invariant under biholomorphisms. This gave rise to hyperbolic spaces theory \[3–5\]. Kobayashi pseudodistance is the maximal pseudodistance among all pseudodistances non-increasing under holomorphic mappings, which on the unit disk $`D_1\text{C}^{_{}}`$ coincides with the distance $`d_D`$, induced by infinitesimal Poincaré metric in Lobachevskii model $$dl^2=\frac{dzd\overline{z}}{(1|z|^2)^2}.$$ On a complex manifold $`M`$ the pseudodistance is defined by the formula $$d_M(p,q)=inf\underset{k=1}{\overset{m}{}}d_D(z_k,w_k),$$ where the infimum is taken over all holomorphic mappings $`f_k:D_1M`$, $`k=1,\mathrm{},m`$, such that $`f_1(z_1)=p`$, $`f_k(w_k)=f_{k+1}(z_{k+1})`$ and $`f_m(w_m)=q`$. In the paper the Kobayashi pseudodistance was extended to the case of arbitrary almost complex manifolds and it was shown that the basic properties of this pseudodistance are preserved. In 1970 Royden found an infinitesimal analog of the Kobayashi pseudodistance for complex manifolds. We define the corresponding notion in the category of almost complex manifolds and we prove, using theorem 1, the coincidence theorem (theorem 3). We obtain a hyperbolicity criterion (theorem 4). We also consider the reduction procedure, which allows to define geometric invariants of the moduli space for pseudoholomorphic curves. ## 1 . Existence of close pseudoholomorphic disks ### 1.1 . Reformulation of the main result Let us reformulate theorem 1 using the differential equation language in appropriate coordinates. To begin with choose these coordinates along the disk $`f_0(D_R)M`$. Due to existence of isothermal coordinates on surfaces the disk $`f_0(D_R)`$ can be defined in local complex coordinate system $`(z^1,\mathrm{},z^n)`$, which is specified in some neighborhood of the disk, via the formulae: $`|z^1|R`$, $`z^2=\mathrm{}=z^n=0`$. Moreover the disk will be pseudoholomorphic, $`J|_{Imf_0}=J_0`$, and $`v_0=(1,0,\mathrm{},0)T_0\text{C}{}_{}{}^{_{}}\text{C}^{_{}}`$. ###### Proposition 1 $`.`$ In an appropriate coordinate system the vector fields $`_k=/z^k`$ and $`\overline{}_k=/\overline{z}^k`$ at points of the disk $`f_0(D_R)`$ satisfy the conditions $$J_k=i_k,J\overline{}_k=i\overline{}_k.$$ (1) * Given equations are already satisfied on the disk for the vector fields $`_1`$, $`\overline{}_1`$. Further at the points of the disk we define transversal to this disk vector fields $`_k`$, $`\overline{}_k`$, $`k2`$, in such a way that all the union of $`2n`$ vectors forms a basis at each point and also that condition (1) is satisfied. Upon constructing the needed vector fields at the points of the disk we extend them to a neighborhood with the help of lemma 1. Obtained structure $`J`$ coincides with the structure $`J_0`$ on the disk and does not necessarily do so outside. y $`\mathrm{}`$ ###### Lemma 1 $`.`$ Let we be given $`k`$ standard commuting vector fields $`v_i=_i`$, $`i=1,\mathrm{},k`$, and also $`nk`$ transversal fields $`v_j`$, $`j=k+1,\mathrm{},n`$, along the disk $`D^k\mathrm{IR}^k\times \{0\}^{nk}\mathrm{IR}^n`$; at each point $`xD^k`$ all the vectors $`v_1,\mathrm{},v_n`$ forming a basis. Then there exist coordinates $`x^i`$ in a small neighborhood of the disk $`D^k`$ such that $`v_i(x)=_i=/x^i`$, $`i=1,\mathrm{},n`$, for all $`xD^k`$. * Since the commutators of vector fields along $`D^k`$ are determined by their $`1`$-prolongations outside $`D^k`$, we write the general form for a $`1`$-prolongation of the vector field $`v_1`$: $$v_1=\underset{r=1}{\overset{n}{}}\left(\delta _1^r+\underset{s=k+1}{\overset{n}{}}x^s\varphi _s^r(x_1,\mathrm{},x_k)\right)_r\text{mod}\mu ^2𝒟,$$ (2) where $`\mu ^2𝒟`$ is the submodule of the module of vector fields, consisting of the vector fields vanishing on the submanifold $`D^k\mathrm{IR}^n`$ to the second order. If on the disk $`D^k`$ the decomposition of the additional vector fields is written as $$v_j=\underset{s=1}{\overset{n}{}}a_j^s(x_1,\mathrm{},x_k)_s,j=k+1,\mathrm{},n,$$ (3) then the equations $`[v_1,v_j]=0`$ with $`x^{k+1}=\mathrm{}=x^n`$ have the following form $$\underset{r=1}{\overset{n}{}}(_1a_j^r(x^1,\mathrm{},x^k))_r=\underset{s=k+1}{\overset{n}{}}a_j^s(x^1,\mathrm{},x^k)\underset{r=1}{\overset{n}{}}\varphi _s^r(x^1,\mathrm{},x^k)_r.$$ This system decomposes (by $`r`$) on $`n`$ determinate systems of $`nk`$ linear equations with $`nk`$ unknowns. The matrix $`(a_j^s)_{k+1j,sn}`$ of each system is nondegenerate, hence the system possesses a solution. Thus the field $`v_1`$ is constructed. Let us rectify it: $`v_1=/x^1`$. We can assume that on the disk $`D^k\{x^{k+1}=\mathrm{}=x^n=0\}`$ the tangent vector fields have the original form $`v_i=_i`$. In new coordinates the coefficients of the decomposition (3) do not depend on $`x^1`$. So one can search for the prolongation of the field $`v_2`$ in the form similar to (2), but with no dependence on $`x^1`$. Continuing the process we get some coordinates $`x^1,\mathrm{},x^n`$, in which the disk $`D^k`$ belongs to the subspace $`\{x^{k+1}=\mathrm{}=x^n=0\}`$ and such that on this disk $$v_i=\frac{}{x^i},i=1,\mathrm{},k,v_j=\underset{s=1}{\overset{n}{}}a_j^s_s,a_j^s=\text{const},j=k+1,\mathrm{},n.$$ (4) Now we prolong the vector fields to a neighborhood via the formula (4). y $`\mathrm{}`$ Thus we introduce complex coordinates $`z^k=x^k+iy^k`$ in a neighborhood of the disk $`f_0(D^k)`$. Now our manifold, being contracted, has the form $$M_0=\{|z_1|R,|z_k|R_1,k2\}D_R\times (D_{R_1})^{n1}\text{C}{}_{}{}^{_{}}{}_{}{}^{n},R_1R,$$ (5) and the structure $`J`$ at points of the disk $`D_R=\{(x_1,y_1,0,\mathrm{},0,0)\}`$ has the form $$J\frac{}{x^k}=\frac{}{y^k},J\frac{}{y^k}=\frac{}{x^k}.$$ (6) Writing down Cauchy-Riemann equations $`f_{}J_0=Jf_{}`$ on the mapping of the disk $`f:(D_{R\epsilon },J_0)(M,J)`$ (similar to sec. 3.3 from ) and using the rectifying conditions (6), we get an equivalent formulation of the main statement ($`,\overline{}`$ are considered in the coordinates $`z^k`$): ###### Theorem 1 $`^{}.`$ Let $`n^2`$ functions $`a_{\overline{m}}^i`$ ($`i,m=1,\mathrm{},n`$) of the class $`C^{k+\lambda }`$, $`k\mathrm{Z}\mathrm{Z}_+`$, $`\lambda (0,1)`$ be given on a manifold $`M_0`$ of the form (5). Let $`\epsilon (0,R)`$ be an arbitrary small real number. If $`a_{\overline{m}}^i(z)=0`$ for all points $`zD_R\times \{0\}^{n1}M_0`$, then the equation $$\overline{}z^i+\underset{m=1}{\overset{n}{}}a_{\overline{m}}^i(z)\overline{}\overline{z}^m=0,z^i(0)=p^i,z^i(0)=u^i,i=1,\mathrm{},n,$$ has a solution $`z^i=z^i(\zeta )C^{k+1+\lambda }(D_{R\epsilon };M_0)`$ subject to the restriction that the neighborhood $`𝒱=𝒱(v_0)v`$ of the vector $`v_0=(1,0,\mathrm{},0)T_0M_0`$ is chosen sufficiently small. Here $`v=(p,u)`$, $`p=\tau _MvM_0`$, $`uT_pM_0`$. ### 1.2 . Covering of the neighborhood by disks and another reformulation ###### Theorem 1 $`^{\prime \prime }.`$ One can set $`p=0`$ in the formulation of theorem $`1^{}`$. Thus in the chosen coordinates the equation on the pseudoholomorphic disk sought for takes the following form: $$\{\begin{array}{ccc}\hfill \overline{}z^1& =& _{m=1}^na_{\overline{m}}^1(z)\overline{}\overline{z}^m,\hfill \\ \hfill \overline{}z^I& =& _{m=1}^na_{\overline{m}}^I(z)\overline{}\overline{z}^m.\hfill \end{array}$$ (7) $$z^1(0)=0,z^I(0)=0,(z^1(0),z^I(0))=(u^1,\mathrm{},u^n),$$ where the multiindex $`I`$ stands for $`(2,\mathrm{},n)`$. Next statement shows equivalence of theorems $`1^{}`$ and $`1^{\prime \prime }`$. ###### Proposition 2 $`.`$ For every pseudoholomorphic disk $`f_0:D_RM_0`$ and every $`\epsilon >0`$ a small neighborhood of the image $`f_0(D_{R\epsilon })`$ can be covered by the images of close pseudoholomorphic disks $`f`$ of radii $`R\delta `$, where $`\delta <\epsilon `$: $`𝒪(Imf_0(D_{R\epsilon }))_fImf(D_{R\delta })`$. * Perturb the almost complex structure $`J`$ in a neighborhood of the disk $`f_0(D_R)`$ so that it coincides with the standard integrable structure $`J_0`$ near the boundary of this neighborhood: for every $`\epsilon >0`$ there exists such an almost complex structure $`\stackrel{~}{J}`$ that $`\stackrel{~}{J}=J`$ in a small neighborhood of the disk $`f_0(D_{R\epsilon /2})=D_{R\epsilon /2}\times \{0\}^{n1}M_0`$ and $`\stackrel{~}{J}=J_0`$ in a neighborhood of the boundary of the manifold $`\stackrel{~}{M}=D_R\times (D_\delta )^{n1}M_0\text{C}_{}^{n}{}_{}{}^{_{}}`$. Further one can suppose that $`\stackrel{~}{M}\widehat{M}=S_R^2\times (S_{2R}^2)^{n1}(S^2)^n`$, where $`\widehat{M}`$ can be equipped with an almost complex structure $`\widehat{J}`$, which coincides with $`\stackrel{~}{J}`$ in $`\stackrel{~}{M}`$ and which equals the standard integrable structure $`J_0`$ in the complement. Let us supply the manifold $`\widehat{M}`$ with the symplectic structure $`\omega =\omega _0^{(1)}\omega _0^{(2)}\mathrm{}\omega _0^{(2)}`$, where $`\omega _0`$ is the standard volume form, and also $`\omega _0^{(1)}(S_R^2)=\pi R^2`$ and $`\omega _0^{(2)}(S_{2R}^2)=4\pi R^2`$. Decreasing if necessary the size of the neighborhood of the disk $`f_0(D_{R\epsilon /2})`$ we can suppose that the almost complex structure $`\widehat{J}`$ is tamed by the symplectic structure $`\omega `$, i.e. $`\omega (\xi ,\widehat{J}\xi )>0`$ for $`\xi 0`$. Denote by $`AH_2(\widehat{M};\mathrm{Z}\mathrm{Z})`$ the homology class of the sphere $`S_R^2\times \{\}^{n1}\widehat{M}`$. The disk $`f_0(D_{R\epsilon /2})`$ can be extended to the entire rational pseudoholomorphic curve $`u_0:S^2\widehat{M}`$, which lies in the class $`A`$. Let us consider the space $`(A,\widehat{J})`$ of entire pseudoholomorphic curves $`u:S^2\widehat{M}`$ of the class $`A`$. Since the class $`A`$ cannot be decomposed into a sum of homology classes $`_{i=1}^nA_i`$, $`n2`$, with $`\omega (A_i)>0`$, then Gromov compactness theorem ( sec. 1.5.B or Sec. 4.3.2) implies the compactness of the space $`(A,\widehat{J})/G`$, where $`GPSL_2`$ is the complex automorphisms group of the sphere $`(S^2,J_0)`$, $`dimG=6`$. Moreover for almost complex structure $`\widehat{J}`$ of general position the space $`(A,\widehat{J})`$ is a smooth manifold of dimension $`2n+4`$ sec. 2.1–2.2; sec. 3.1.2. Consider the space of nonparametrized pseudoholomorphic curves $`𝒲(A,\widehat{J})=(A,\widehat{J})\times _GS^2`$. This space is a compact manifold of dimension $`2n`$. Let us consider the evaluation map $`e:𝒲(A,\widehat{J})\widehat{M}`$, which is defined by the formula $`e(u,z)=u(z)`$ for $`zS^2`$, $`u(A,\widehat{J})`$. We suppose the group $`G`$ acts on $`\times S^2`$ by conjugation, $`\varphi (u,z)=(u\varphi ^1,\varphi (z))`$, $`\varphi G`$, whence the correctness of the definition for $`e`$. Since $`A`$-curves foliate the manifold $`\widehat{M}`$ outside a small neighborhood of the image $`u_0(S^2)`$ (because there $`\widehat{J}=J_0`$), the map $`e`$ has degree $`1`$, $`dege=1`$. Therefore through every point, close to the curve $`u_0(S^2)`$, some pseudoholomorphic curve $`u(S^2)`$ passes, which is homologous to the curve $`u_0(S^2)`$. To eliminate the general position condition for $`\widehat{J}`$ we take a sequence $`\widehat{J}_k`$ of the general position almost complex structures, which tends to $`\widehat{J}`$ in $`C^{\mathrm{}}`$-topology, and use the compactness theorem B.4.2. Intersecting the obtained set of pseudoholomorphic spheres $`u(S^2)`$ with a small neighborhood $`𝒪`$ of the disk $`f(D_{R\epsilon /2})`$, we get the desired set of the disks $`f(D)`$ in a neighborhood $`(𝒪,J)`$. Smoothness of these disks follows from the standard elliptic regularity sec. 5.4, 4.3; sec. B.4.1. y $`\mathrm{}`$ ### 1.3 . Spaces, norms and estimates Define the $`\lambda `$-Hölder norm of complex-valued functions on the disk $`D_R`$ of radius $`R`$ by the formula $`f=|f|+(2R)^\lambda H_\lambda [f]`$, $`\lambda (0,1)`$, where $$H_\lambda [f]=\underset{w0}{sup}\left|\frac{f(z+w)f(z)}{w^\lambda }\right|,|f|=sup|f(z)|.$$ The space $`C^\lambda (D_R,M_0)`$ of $`\lambda `$-Hölder maps consists of all maps $`f:D_RM_0`$, the components of which have finite $`\lambda `$-norms, $`f^i<\mathrm{}`$. The space $`C^{k+\lambda }(D_R,M_0)`$, $`k\mathrm{Z}\mathrm{Z}_+`$, of $`(k+\lambda )`$-Hölder maps consists of all maps, the partial derivatives of which up to the $`k`$-th order inclusive belong to $`C^\lambda `$. Let us also introduce the space $`B=C_{}^{k+\lambda }(D_R,M_0)`$ consisting of all maps $`fC^{k+\lambda }`$, $`f(0)=0`$, with the norm $`f=_1^nf^i`$. Note that the space $`B`$ can be also supplied with the norm $$f^{}=\mathrm{max}\{f,\overline{}f\}.$$ ###### Proposition 3 $`.`$ The spaces $`(C^\lambda ,)`$ and $`(C_{}^{1+\lambda },^{})`$ are Banach. The second statement follows from the first and the estimate 7.1.c–7.1.e $$f6Rf^{}.$$ (8) Consider the Cauchy operators $$Sf(w)=\frac{1}{2\pi i}_{D_R}\frac{f(\zeta )}{\zeta w}𝑑\zeta ,Tf(w)=\frac{1}{2\pi i}_{D_R}\frac{f(\zeta )}{\zeta w}𝑑\zeta d\overline{\zeta }.$$ Recall the basic properties of these operators 6.1–6.2: $$fC^\lambda (D)TfC^{1+\lambda }(D),\overline{}Tf=f,$$ (9) $$fC^\lambda (D)SfC^\lambda (IntD),\overline{}Sf=0,STf=0,$$ (10) $$f=Sf+T\overline{}f\text{ (Cauchy-Green-Pompéiu formula) },$$ (11) $$Tf^{}c_1f,Sfc_2f,$$ (12) Consider also the operators $`T_kf(w)=Tf(w)_{s=0}^k\frac{1}{s!}^sTf(0)w^s`$. ###### Lemma 2 $`.`$ For the points $`wIntD`$ the following formula holds $$T_kf(w)=\frac{w^{k+1}}{2\pi i}_{D_R}\frac{f(\zeta )}{(\zeta w)\zeta ^{k+1}}𝑑\zeta d\overline{\zeta }.$$ ###### Lemma 3 $`.`$ The operator $`T_{\mathrm{}}=lim_k\mathrm{}T_k`$ is defined for functions $`fC^\lambda (D_R)`$, and moreover $`T_{\mathrm{}}fC^{1+\lambda }(D_{R\epsilon })`$. ###### Lemma 4 $`.`$ $`T_k(w^l\overline{w}^m)=[\begin{array}{c}\frac{w^l\overline{w}^{m+1}}{m+1},l<k+m+2,\hfill \\ \frac{w^l\overline{w}^{m+1}}{m+1}\frac{R^{2\left(m+1\right)}}{m+1}w^{lm1},lk+m+2.\hfill \end{array}`$ * $`T_{\mathrm{}}(w^l\overline{w}^m)=w^l\overline{w}^{m+1}/(m+1)`$. Thus the operator $`T_{\mathrm{}}`$ represent the integration by $`\overline{\zeta }`$ of the polynomials on $`D_R`$. Let us also besides the space $`B`$ consider its closed subset $`B_\delta =\{f=(f_1,\mathrm{},f_n)B,|f_1\zeta |\delta ,|f_k|\delta ,k2\}`$. We will seek a solution $`f`$ of the Cauchy-Riemann equation (7) in the space $`B_\delta `$ for a small neighborhood $`𝒱`$ of the vector $`v_0`$. ### 1.4 . Proof of theorem 1<sup>′′</sup> Idea of the proof. Equation (7) was solved in the paper , theorem III, where the velocity vector $`v`$ was fixed and the radius $`R1`$ of the disk was supposed small. For this the linearization of almost complex structure at the point was considered. Because of the proximity of equations on pseudoholomorphic curves for the given almost complex structure $`J`$ and for the linearized one $`J_0`$ the following map was contractible: $$\mathrm{\Phi }:BB,(\mathrm{\Phi }f)^i(\zeta )=v^i\zeta +T_1\left(\underset{m}{}a_{\overline{m}}^i(f)\overline{}\overline{f}^m\right)(\zeta ).$$ (13) In our situation radius of the disk is not small, therefore the word for word carrying over the arguments from is possible only if the structure $`J`$ differs from $`J_0`$ on the disk $`f_0`$ by a second order smallness quantity, i.e. if the functions $`a_{\overline{1}}^i`$ on the disk $`f_0`$ as well as their derivatives vanish. In general situation it is not the case, so we linearize the almost complex structure $`J`$ along the disk $`f_0`$. Here the linearization is parametrized by the coordinate $`z^1=\zeta `$ along this disk. Solutions of complex linear equation behave similarly to solutions of the real equation $`\dot{x}=Ax+B`$: for non-small values of the parameter it is false that $`e^{At}1`$, so the terms of the series $`e^{At}=_{s=0}^{\mathrm{}}(At)^s/s!`$ are not absolute decreasing, but this property becomes true beginning with some number $`ss_0`$. Thus finite sums of the series for the exponent does not form a contracting sequence, yet to achieve this one should consider the sums beginning with some big number. Let us turn to the proof. Similarly to , starting with formulae (11), (10), we seek a solution of equation (7) in the form (13), but we replace the space $`B`$ by $`B_\delta `$. In fact, as noted above, we should change the definition of the operator $`\mathrm{\Phi }`$ to improve the convergence. Let us consider the automorphism of the space $`\text{C}_{}^{n}{}_{}{}^{_{}}`$, which comes from the contraction of the space $`M_0`$, $$z^1z^1,z^I\frac{z^I}{N},N1.$$ (14) Since $`a_{\overline{m}}^i=0`$ along the disk $`D_R\times \{0\}^{n1}M_0`$, the function $`a_{\overline{1}}^1`$ becomes small and the functions $`a_{\overline{1}}^I`$ become very close to their linearizations by the variables $`z^I`$ in the norm $`^{}`$ for large $`N`$ in equation (7). Consequently the first equation of (7), considered as one $`z^I`$-parametric equation, can be solved by the iteration method, when we use formula (13) for complex dimension $`1`$ and change $`B`$ to $`B_\delta `$. For small $`\delta `$ and big $`N`$ in (14) the estimates from sec. 5.2 yield the contractibility of the iteration procedure in the norm $`^{}`$. This iteration procedure will be denoted by $`z^1\mathrm{\Psi }^1(z^1,z^I)`$. To consider the second equation of (7) let us linearize the functions used in it by $`z^I`$: $$a_{\overline{1}}^I(z)=\underset{m2}{}\left(a_{\overline{1};m}^I(z^1)z^m+a_{\overline{1};\overline{m}}^I(z^1)\overline{z}^m\right)+\widehat{a}_{\overline{1}}^I(z).$$ (15) In this formula the functions $`\widehat{a}_{\overline{1}}^I(z)`$ have the second order of smallness along the disk $`D_RM_0`$. Let us also set $`\widehat{a}_{\overline{m}}^I(z)=a_{\overline{m}}^I(z)`$ when $`m1`$. According to Weierstrass theorem the coefficients at linear by $`z^I`$ terms in (15) are approximated by polynomials depending on $`z^1`$, $`\overline{z}^1`$ in the norm $`||`$ on $`D_R`$: $$a_{\overline{1};m}^I(z^1)=p_m^I(z^1,\overline{z}^1)+\alpha _m^I(z^1),a_{\overline{1};\overline{m}}^I(z^1)=p_{\overline{m}}^I(z^1,\overline{z}^1)+\alpha _{\overline{m}}^I(z^1),|\alpha _m^I|,|\alpha _{\overline{m}}^I|<\epsilon .$$ Let $`A^I(\zeta ,z^I)=_{m2}(p_m^I(\zeta ,\overline{\zeta })z^m+p_{\overline{m}}^I(\zeta ,\overline{\zeta })\overline{z}^m)`$, $`A_\delta ^I(\zeta ,z)=A^I(\zeta ,z^I)A^I(z^1,z^I)`$. Then the second equation of (7) can be written in the form $$\overline{}z^I(\zeta )=A^I(\zeta ,z^I)+U^I(z(\zeta )),$$ (16) where the summands of the remainder $`U^I=A_\delta ^I+U_1^I+U_2^I+U_3^I`$ have the form $$U_1^I(z)=A^I(z^1,z^I)(1\overline{}\overline{z}^1),U_2^I(z)=\underset{m}{}\widehat{a}_{\overline{m}}^I\overline{}\overline{z}^m,$$ $$U_3^I(z)=\underset{m2}{}(\alpha _m^I(z^1)z^m+\alpha _{\overline{m}}^I(z^1)\overline{z}^m)\overline{}\overline{z}^1.$$ We approximate equation (16) by the following equation with linear by $`z^I`$ right hand size and polynomial by $`\zeta `$ coefficients: $$\overline{}z^I(\zeta )=A^I(\zeta ,z^I).$$ (17) A solution of this equation can be constructed as the limit of the iteration procedure $$z_{(k+1)}^I=v^I\zeta T_{\mathrm{}}[A^I(\zeta ,z_{(k)}^I)].$$ (18) By the corollary of lemma 4 the iteration of integration by means of the operator $`T_{\mathrm{}}`$ has the form $`T_{\mathrm{}}^k(w^l\overline{w}^m)=w^lm!\overline{w}^{m+k}/(m+k)!`$, which implies that the iteration process (18) converges under any initial condition $`z_{(0)}^I`$ to a solution of equation (17), and moreover the convergence is exponential. In particular, beginning with some number $`k`$, the sequence $`z_{(k)}^I`$ is contractible. And what is more there exist constants $`C`$ and $`\mu `$, depending only on almost complex structure $`J`$ (i.e. on coefficients $`a_{\overline{m}}^i`$), such that for every $`k1`$ and polynomial $`p(\zeta ,\overline{\zeta })`$ the following inequality holds: $$T_{\mathrm{}}^k[A^I(\zeta ,p)]^{}Ce^{\mu R}p.$$ (19) We now define the iteration procedure to compute $`z^I(\zeta )`$. Let the iterative term $`z_{[r]}^I`$ be already constructed. Additionally in virtue of the previous step the given term is equal to the sum of a polynomial $`P_{[r]}^I(\zeta ,\overline{\zeta })`$ and a function $`\theta _{[r]}^I(\zeta )C^{1+\lambda }`$. Represent the last function by Weierstrass theorem as the sum of a polynomial (by $`\zeta `$, $`\overline{\zeta }`$) and an error: $`\theta _{[r]}^I(\zeta )=Q_{[r]}^I(\zeta )+q_{[r]}^I(\zeta )`$, $`|q_{[r]}^I|\nu |\theta _{[r]}^I|`$. Define the next term by the formula $$z_{[r+1]}^I(\zeta )=v^I\zeta T_{\mathrm{}}[A^I(P_{[r]}^I)]T_{\mathrm{}}^{k_r}[A^I(Q_{[r]}^I)]T_1[A^I(q_{[r]}^I)]+T_1[U^I(z_{[r]})].$$ Here $`A^I=A^I(\zeta ,)`$ and $`k_r`$ is such a number that beginning with number $`k_r`$ the sequence $`T_{\mathrm{}}^k[A^I(Q_{[r]}^I)]`$ contracts with the coefficient $`\epsilon _r`$. In addition (cf. (8)) the following estimates for the additional terms take place: $$A_\delta ^I(\zeta ,z^{})A_\delta ^I(\zeta ,z^{\prime \prime })c_3\delta z^{}z^{\prime \prime }^{},U_1^I(z^{})U_1^I(z^{\prime \prime })c_3\delta z^{}z^{\prime \prime }^{},$$ $$U_2^I(z^{})U_2^I(z^{\prime \prime })c_3\delta z^{}z^{\prime \prime }^{},U_3^I(z^{})U_3^I(z^{\prime \prime })c_4\epsilon z^{}z^{\prime \prime }^{}.$$ Taking inequality (12) into account we conclude that for small $`\delta `$, $`\epsilon `$, $`\epsilon _r`$ and $`\nu `$ the sequence $`z_{[r]}^I`$ is contractible: $`z_{[r+1]}^Iz_{[r]}^I^{}(1\kappa )z_{[r]}^Iz_{[r1]}^I^{}`$ for some $`\kappa <1`$ independent of $`r`$. Therefore, taking into consideration the iteration by $`\mathrm{\Psi }^1`$ for the variable $`z^1`$, we get a convergent in $`C^{1+\lambda }`$ sequence, the limit of which has to be the desired solution. Actually, set $`z_{[r+1]}^1=\mathrm{\Psi }^1(z_{[r]}^1,z_{[r]}^I)`$, taking as parameter $`z^I`$ the iterative term $`z_{[r]}^I`$. In what follows in determination of the term $`z_{[r+1]}^I`$ we assume $`z^1=z_{[r]}^1`$. Thus we obtain the sequence $`z_{[r]}`$. Due to the estimates considered and inequality (19) the terms and the limit of the sequence $`z_{[r]}`$ differ from its initial term $`z_{[0]}=v\zeta `$ less than exponentially by $`R`$ with respect to $`|vv_0|`$ in the norm $`^{}`$. Therefore for small $`|vv_0|1`$ all the terms and the limit of the iterative sequence lie in $`B_\delta `$. Hence the sequence converges in $`B_\delta `$. Now it is easily seen that the limit of the sequence $`z_{[r]}`$ is a solution of the equation (7). When the coefficients have smoothness $`a_{\overline{m}}^iC^{k+\lambda }(M_0)`$, then the obtained solution, which is of smoothness $`C^{1+\lambda }`$, will be actually of higher smoothness class $`C^{k+1+\lambda }`$. This follows from the standard elliptic regularity methods for our equation 5.4, 4.3, B.4.1. When $`a_{\overline{m}}^iC^{\mathrm{}}(M_0)`$ we get a smooth solution of the Cauchy-Riemann equation $`z(\zeta )C^{\mathrm{}}(D_{R\epsilon };M_0)`$. y $`\mathrm{}`$ ### 1.5 . Jet spaces and connection with $`h`$-principle Let us call the foliation by pseudoholomorphic disks any embedding (immersion) $`\mathrm{\Phi }:D_R\times N^{2n2}M`$ such that all the mappings $`\mathrm{\Phi }|_{D_R\times \{x\}}`$ are pseudoholomorphic and the image of the map $`\mathrm{\Phi }`$ covers the entire manifold $`M`$. The construction of proposition 2 together with the positivity of intersections in dimension $`4`$ ( 2.1.C<sub>2</sub>; 1.1) imply ###### Proposition 4 $`.`$ Let $`(M,J)`$ be a four-dimensional almost complex manifold. For every embedded (immersed) pseudoholomorphic disk $`f:D_RM`$ and every $`\epsilon >0`$ small neighborhood of the image $`f(D_{R\epsilon })`$ allows the foliation by pseudoholomorphic disks. Let us consider the manifold of pseudoholomorphic jets $`𝒥_{PH}^1(D_R;M)`$ of the mappings $`u:D_RM`$. Its points are triples $`(\zeta ,z,\mathrm{\Phi })`$, where $`\zeta D_R`$, $`zM`$, and $`\mathrm{\Phi }:(T_\zeta D_R,J_0)(T_zM,J(z))`$ is a complex linear mapping. It was shown in the paper that the manifold $`𝒥_{PH}^1`$ possesses a canonical almost complex structure $`J_{[1]}`$, which is equal to $`J_0JJ`$ regarding the induced by some minimal connection decomposition $`T_p𝒥_{PH}^1=T_\zeta D_RT_zMT_p`$, where $``$ is the fiber of the natural projection $`\tau :𝒥_{PH}^1(D_R,M)D_R\times M`$. The canonical projection $`\pi :𝒥_{PH}^1(D_R;M)M`$ is pseudoholomorphic and any pseudoholomorphic mapping $`f:D_RM`$ lifts canonically to the pseudoholomorphic mapping $`j^1f:D_R𝒥_{PH}^1(D_R;M)`$, $`j^1f(\zeta )=(\zeta ,f(\zeta ),d_\zeta f)`$. We define the structure $`J_{[1]}`$ in a different way (cf. , remark 1). If $`p=(\zeta ,z,\mathrm{\Phi })𝒥_{PH}^1`$, we can assume that the mapping $`\mathrm{\Phi }`$ is the differential at the point $`\zeta `$ of some small pseudoholomorphic disk $`u:D_\epsilon M`$. Denote by $`p^{(2)}`$ the 2-jet of the disk $`u`$ at the point $`\zeta D_\epsilon D_R`$. Consider the map $`j^1u:D_\epsilon 𝒥_{PH}^1`$. The tangent space at the point $`p`$ depends only on the value $`p^{(2)}`$. Denote this tangent space by $`L_{p^{(2)}}`$. Consider the natural projection $`\rho :𝒥_{PH}^1D_R`$ with the fiber $``$. We have $`T_p𝒥_{PH}^1=L_{p^{(2)}}T_p`$, both summand being naturally equipped with complex structures. Set $`J_{[1]}=J_0J`$. This structure does not depend on the choice of $`p^{(2)}`$, i.e. it is defined canonically. * Let us call a pseudoholomorphic disk $`g:D_R𝒥_{PH}^1(D_R,M)`$ holonomic, if the mapping $`g`$ is the 1-jet lifting of some pseudoholomorphic disk from $`D_R`$ to $`M`$: $`g=j^1f`$. Proposition 2 applied to a holonomic disk $`g=j^1f:D_R𝒥_{PH}^1`$ yields existence of a pseudoholomorphic disk $`g^{}`$ through each point arbitrary close to the image of the disk $`g`$, which however needs not be a holonomic disk, $`g^{}j^1(\pi g^{})`$. In this sense theorem 1 provides a more strong statement. Actually, closeness of initial points of the disks $`g=j^1f`$ and $`g^{}=j^1f^{}`$ in $`𝒥_{PH}^1(D_{R\epsilon };M)`$ means closeness of initial points and initial directions of the maps $`f`$ and $`f^{}`$ in $`TM`$. Thus theorem 1 implies existence of $`C^1`$-close disk $`f^{}`$, and we can set $`g^{}=j^1f^{}`$. Thus we proved ###### Theorem 2 $`.`$ Through every point, which is close to the image of embedded (immersed) holonomic pseudoholomorphic disk $`g:D_R𝒥_{PH}^1(D_R;M)`$, an embedded (immersed) holonomic pseudoholomorphic disk $`g^{}:D_{R\epsilon }𝒥_{PH}^1(D_{R\epsilon };M)`$ passes. y $`\mathrm{}`$ In other words, proposition 2 remains also valid in the holonomic situation. The statement just proved is a particular case of the so-called $`h`$-principle . It is also interesting to get the holonomic version of proposition 4. ## 2 . Kobayashi-Royden pseudonorm ### 2.1 . Definition of the pseudonorm and its main properties Let us consider the set $`(v)=_{r>0}_r(v)`$, where $`_r(v)`$ for $`r\mathrm{IR}_+`$ consists of pseudoholomorphic mappings $`f:D_1M`$, such that $`f_{}(0)e=rv`$. * Let us call the Kobayashi-Royden pseudonorm on an almost complex manifold $`M`$ the function on the tangent bundle $`TM`$, which is defined by the formula $$F_M(v)=\underset{(v)}{inf}\frac{1}{r}.$$ According to theorem III from the set $`_r(v)`$ is nonempty for small $`r`$, so the definition is correct. We call the function $`F_M`$ pseudonorm since it is nonnegative and homogeneous of degree one: $`F_M(tv)=|t|F_M(v)`$. However $`F_M`$ can vanish in some directions and the triangle inequality does not hold. The next statement follows from the very definition. ###### Proposition 5 $`.`$ Given any vector $`vTM_1`$ and any pseudoholomorphic mapping $`f:(M_1,J_1)(M_2,J_2)`$ we have $$F_{M_2}(f_{}v)F_{M_1}(v).$$ Let us fix some norm $`||`$ on $`TM`$. ###### Proposition 6 $`.`$ (i) There exists a constant $`C_K`$ for every compact $`KM`$ such that each vector $`vTM`$ with $`\tau _MvK`$ satisfies $$F_M(v)C_K|v|.$$ (ii) Let $`M`$ be a compact manifold (with possible boundary) equipped with an almost complex structure $`J`$, which is tamed by an exact symplectic form $`\omega =d\alpha `$, $`\omega (\xi ,J\xi )>0`$ for $`\xi 0`$. Then there exists such a constant $`c_M>0`$, that for all $`vTM`$ $$F_M(v)c_M|v|.$$ * For a small neighborhood $`U`$ of the point $`pM`$ the estimates of sec. 5.2a of the paper imply existence of a number $`\epsilon >0`$, dependent only on the almost complex structure $`J`$ and the neighborhood $`U`$, such that for every $`qU`$, $`vT_qM`$, $`|v|=1`$, and $`r(0,\epsilon )`$ there exists a pseudoholomorphic disk $`f:D_1M`$ such that $`f(0)=q`$, $`f_{}(0)e=rv`$. Setting $`C_U=1/\epsilon `$ we have $`F_M(v)C_U|v|`$ for all (now not necessarily unit) vectors $`v`$ for which $`\tau _MvU`$. Since a compact set can be covered by a finite number of neighborhoods $`U`$, the first statement of the proposition is proved. The second part is a reformulation of the nonlinear Schwarz lemma 1.3.A: if an almost complex structure $`J`$ on a compact manifold is tamed by an exact symplectic structure $`\omega `$, then the derivative at zero of any pseudoholomorphic disk $`f:D_1M`$, passing through a fixed point at the manifold, is bounded by a non-depending on the disk constant: $`|f_{}(0)e|<C`$. y $`\mathrm{}`$ ###### Proposition 7 $`.`$ The function $`F_M`$ is upper semicontinuous. * The inequality $`\overline{\underset{vv_0}{lim}}F_M(v)F_M(v_0)`$ is equivalent to the statement of theorem 1 because $`F_M(v)=inf(1/R)`$, where the lower bound is considered over all mappings $`f:D_RM`$, such that $`f_{}(0)e=v`$.y $`\mathrm{}`$ ### 2.2 . Coincidence theorem Define a function $`\overline{d}_M:M\times M\mathrm{IR}`$ by the formula $$\overline{d}_M(p,q)=\underset{\gamma }{inf}_0^1F_M(\dot{\gamma }(t))𝑑t,$$ where the lower bound is taken over all piecewise smooth paths $`\gamma `$ from the point $`p`$ to $`q`$. Propositions 6(i) and 7 imply correctness of the definition and ###### Proposition 8 $`.`$ The function $`\overline{d}_M`$ is pseudodistance. y $`\mathrm{}`$ ###### Theorem 3 $`.`$ Introduced pseudodistance coincides with the Kobayashi pseudodistance, $`d_M=\overline{d}_M`$. * The inequality $`\overline{d}_Md_M`$ is evident because $`F_M(v)=inf|\xi |`$, where the lower bound is taken over all pseudoholomorphic mappings $`f:D_1M`$, $`f_{}\xi =v`$, and the norm is count with respect to the Poincaré metric. Let us prove the reverse. We follow the Royden’s proof . Let $`\gamma `$ be a smooth curve from a point $`p`$ to a point $`q`$ such that $`_\gamma F_M<\overline{d}_M(p,q)+\epsilon `$. Due to upper semicontinuity there exists a continuous on $`[0,1]`$ function $`h`$, such that $`h(t)>F_M(\dot{\gamma }(t))`$ and $$_0^1h(t)𝑑t<\overline{d}_M(p,q)+\epsilon ,$$ i.e. for sufficiently dense partition $`0=t_0<t_1<\mathrm{}<t_k=1`$ we have $$\underset{i=1}{\overset{k}{}}h(t_{i1})(t_it_{i1})<\overline{d}_M(p,q)+\epsilon .$$ Consider arbitrary pseudoholomorphic curve $`u_t^\gamma :D_\delta M`$, which satisfies the conditions $`u_t^\gamma (0)=\gamma (t)`$ and $`(u_t^\gamma )_{}e=\dot{\gamma }(t)`$. Define for small $`\mathrm{\Delta }t\mathrm{IR}_+\text{C}^{_{}}`$ the curve $`\widehat{\gamma }(t;\mathrm{\Delta }t)=u_t^\gamma (\mathrm{\Delta }t)`$. Since $`\widehat{\gamma }(t;\mathrm{\Delta }t)=\gamma (t+\mathrm{\Delta }t)+O(|\mathrm{\Delta }t|^2)`$, propositions 8 and 6 imply that for small $`\mathrm{\Delta }t`$ it holds: $`d_M(\gamma (t),\gamma (t+\mathrm{\Delta }t))`$ $``$ $`d_M(\gamma (t),\widehat{\gamma }(t;\mathrm{\Delta }t))+d_M(\widehat{\gamma }(t;\mathrm{\Delta }t),\gamma (t+\mathrm{\Delta }t))`$ $``$ $`F_M(\dot{\gamma }(t))\mathrm{\Delta }t+O(|\mathrm{\Delta }t|^2)(1+\epsilon )h(t)\mathrm{\Delta }t.`$ Thus for sufficiently dense partition $`d_M(p,q){\displaystyle \underset{i=1}{\overset{k}{}}}d_M(\gamma (t_{i1}),\gamma (t_i))<(1+\epsilon )(\overline{d}_M(p,q)+\epsilon ).`$ Since $`\epsilon >0`$ is arbitrary constant, the theorem is proved. y $`\mathrm{}`$ ### 2.3 . Hyperbolicity and nonhyperbolicity * Almost complex manifold $`(M,J)`$ is called hyperbolic if the pseudodistance $`d_M`$ is a distance. Let us consider the unit tangent vectors bundle $`\tau _M^{(1)}:T_1MM`$ for some norm $`||`$, and let $`F_M^{(1)}:T_1M\mathrm{IR}`$ be the restriction of the Kobayashi-Royden pseudonorm to it. Proposition 6(i) and theorem 3 imply ###### Theorem 4 $`.`$ The function $`F_M^{(1)}`$ is bounded on compact subsets in $`M`$. Manifold $`M`$ is hyperbolic iff $`F_M^{(1)}`$ is bounded away from zero on compact subsets. Now let us consider the case of nonhyperbolic manifold $`M`$, for example let it possess pseudoholomorphic spheres. In the case of general position for the almost complex structure $`J`$, which is tamed by some symplectic form $`\omega `$ on $`M`$, the set of all pseudoholomorphic spheres in a fixed homology class $`AH_2(M;\mathrm{Z}\mathrm{Z})`$ (completed for compactness by the set of decomposable rational curves) is a finite-dimensional manifold $`(A;J)`$ . We define by the reduction procedure some pseudodistance on this manifold. Namely for any two pseudoholomorphic spheres $`f_i:S^2M`$, defined up to holomorphical reparametrization of $`S^2`$ let $$d_{}([f_1],[f_2])=d_M(p_1,p_2),$$ where $`p_iIm(f_i)`$ are arbitrary points on the images. It is easily seen that $`d_{}`$ is correctly defined pseudodistance on the manifold $``$. As an example note that the defined pseudodistance $`d_{}`$ is a distance for almost complex manifold $`M^4=\mathrm{\Sigma }_g^2\times S^2`$ with $`g>1`$, where the structure $`J`$ is tamed by the standard product symplectic form: as in proposition 4 one proves that $`M^4`$ is fibered by pseudoholomorphic spheres and there is an isomorphism $`\mathrm{\Sigma }_g^2`$. However in the case of four-dimensional manifolds this definition is of importance only in the case of zero self-intersection. Actually if $`AA>0`$ (for nonexceptional case $`AA0`$ ), then two spheres $`Im(f_1)`$ and $`Im(f_2)`$ of the given homology class do intersect. Thus $`d_{}([f_1],[f_2])=0`$. It was shown in the paper that for $`N`$ large enough the manifold $`\times \mathrm{IR}^{2N}`$ possesses a homotopically canonical almost complex structure $`\stackrel{~}{J}`$. Kobayashi pseudodistance $`d_{\times \mathrm{IR}^{2N}}`$ induces a pseudodistance $`\widehat{d}_{}`$ on $``$ via reduction over $`\mathrm{IR}^{2N}`$. In this connection there arises a natural question of existence of almost complex structures $`\stackrel{~}{J}`$ such that the pseudodistances $`\widehat{d}_{}`$ and $`d_{}`$ coincide. MSTU n.a. Baumann, Moscow; kruglikov`@`math.uit.no
warning/0002/hep-ph0002193.html
ar5iv
text
# FZJ-IKP(TH)-1999-37, FTUV-IFIC-99-1215 Chiral Unitary approach to meson-meson and meson-baryon interactions and nuclear applications ## Chapter 1 Introduction Nowadays it is believed that QCD is the theory for the strong interactions. However, while in the high energy regime, due to the asymptotic freedom, the theory has been successfully tested by the experiment, this is not the case for low energies. In this case, one is in the confinement regime of QCD and perturbative methods cannot be applied. On the other hand, in the low energy region the spectrum of QCD presents an interesting fact which is the appearance of the isospin pion triplet with a mass much smaller than the rest of the QCD states. This can be extended to SU(3) by considering the lowest octet of pseudoscalar states $`(\pi ,K,\eta )`$. This fact can be understood by the presence in the QCD Lagrangian of a chiral symmetry for the light quark sector $`(u,d,s)`$ which spontaneously breaks down giving rise, by the Goldstone theorem, to the aforementioned states. The presence of this symmetry pattern constrains also tremendously the interactions between these Goldstone bosons. As a consequence, a successful theory has emerged, Chiral Perturbation Theory ($`\chi PT`$), which exploits these facts giving rise to a series of Lagrangians in a power momentum expansion treating the quark masses in a perturbative way. This power series is valid up to some high energy scale, $`\mathrm{\Lambda }_{\chi PT}`$, which is of the order of 1 GeV. Hence, the $`\chi PT`$ expansion is valid for momenta $`p<<\mathrm{\Lambda }_{\chi PT}1`$ GeV. Although $`\chi PT`$ is a powerful tool for the low energy region, its convergence is limited to a narrow interval (for example, for energies below 0.5 GeV in meson-meson scattering or close to the threshold region in meson-baryon scattering). As a consequence, one of the most representative and interesting facts of strong interacting phenomena, resonances and their properties, cannot be studied. Furthermore, when one tries to increase the energy region of applicability of $`\chi PT`$ just by including higher orders, the predictive power of the theory is rapidly lost since the number of free parameters increases tremendously with the order. For instance, the leading order $`\chi PT`$ Lagrangian has essentially no free parameters, in the next-to-leading one there are 12 and in the next-next-to-leading order there are more than 100. Hence, the development of nonperturbative techniques which can extend the energy region of applicability of the theory, without loosing predictive power, is an important issue. This is the topic to which this report is devoted. We shall indeed see that this can be done successfully and that a good description of the meson-meson and meson-baryon interactions, even for energies above 1 GeV, arises from $`\chi PT`$, supplied with exact unitarity with or without explicit resonance exchanges. The manuscript is organized as follows. In chapter 2 an overall introduction to the emerging chiral Lagrangians is given. Chapter 3 is devoted to deducing the various nonperturbative methods reported here, as well as to discussing their applications to strong meson-meson and meson-baryon scattering processes. In chapters 4 and 5 it is discussed how the previous strong amplitudes are implemented to account for final and initial state interactions, which turn out to be crucial to understand the physics of many reactions. Applications of the former nonperturbative methods in nuclear physics are reported in chapter 6. Finally, conclusions and remarks are collected in the last section. Every chapter has its own introduction where one can find a more detailed, although still general, overview of its contents. ## Chapter 2 Effective chiral Lagrangians In this chapter we want to review the chiral Lagrangians that are going to be used in the following. After giving a brief account of chiral symmetry from the QCD Lagrangian we will report on the lowest and next to leading order $`\chi PT`$ Lagrangians without baryons. We will also discuss the inclusion of explicit meson resonance fields. Then, we will consider the meson-baryon system where the lowest order $`\chi PT`$ Lagrangian will be given. There are many good reviews about Chiral Perturbation Theory to which the interested reader is referred for further details. ### 2.1 Chiral symmetry The QCD Lagrangian with massless $`u`$, $`d`$ and $`s`$ quarks coupled to several external sources reads: $$_{QCD}=_{QCD}^0+i\overline{q}D^\mu \gamma _\mu q+\overline{q}\gamma ^\mu (v_\mu +\gamma _5a_\mu )q\overline{q}(si\gamma _5p)q$$ (2.1) where $`_{QCD}^0`$ is the part of the QCD Lagrangian for the heavier quarks $`c`$, $`b`$ and $`t`$ and gluons, $`v_\mu `$, $`a_\mu `$, $`s`$ and $`p`$ are the vector, axial, scalar and pseudoscalar external sources, $`D_\mu `$ is the covariant derivative for the $`SU(3)`$-colour gauge symmetry and $$q=\left(\begin{array}{c}u\\ d\\ s\end{array}\right)$$ (2.2) is a vector in the three dimensional flavour space. The Lagrangian eq. (2.1) exhibits a local $`SU(3)_LSU(3)_R`$ flavour symmetry under the following transformation rules $`q`$ $``$ $`g_R{\displaystyle \frac{1}{2}}(1+\gamma _5)q+g_L{\displaystyle \frac{1}{2}}(1\gamma _5)q`$ (2.3) $`v_\mu \pm a_\mu `$ $``$ $`g_{R,L}(v_\mu \pm a_\mu )g_{R,L}^{}+ig_{R,L}_\mu g_{R,L}^{}`$ $`s+ip`$ $``$ $`g_R(s+ip)g_L^{}`$ $`g_{R,L}`$ $``$ $`SU(3)_{R,L}`$ This chiral symmetry, which should be rather good in the light quark sector, is not seen in the hadronic spectrum. Although hadrons can be classified in $`SU(3)_VSU(3)_{R+L}`$ representations, degenerate multiplets with opposite parity are not observed. Moreover, the octet of the lightest pseudoscalar mesons ($`\pi `$, $`K`$, $`\eta `$) can be understood if the chiral $`SU(3)_LSU(3)_R`$ symmetry spontaneously breaks down to $`SU(3)_V`$. Then, according to the Goldstone theorem , an octet of pseudoscalar massless bosons appears in the theory, and these mesons will have the same quantum numbers as the broken generators of $`SU(3)_ASU(3)_{RL}`$. One thus expects that ($`\pi `$, K, $`\eta `$) are to be identified with this octet of Goldstone bosons. Furthermore, Chiral Symmetry is also explicitly broken by a mass term in eq. (2.1) $$\overline{q}M_qq$$ (2.4) when fixing the scalar source $`s(x)=M=diag(m_u,m_d,m_s)`$, where $`m_u`$, $`m_d`$ and $`m_s`$ are the masses of the respective quarks. Because of this term, the Goldstone bosons acquire a small mass giving rise to the masses of the ($`\pi `$, K, $`\eta `$). On the other hand $`SU(3)_V`$ in the hadronic spectrum is only an approximate symmetry because $`m_s`$ is much larger than $`m_u`$ or $`m_d`$. ### 2.2 Chiral Perturbation Theory Taking advantage of the mass gap separating the lightest pseudoscalar octet from the rest of the hadronic spectrum one can build an effective field theory containing only the Goldstone modes. These Goldstone fields, $`\stackrel{}{\varphi }`$, can be collected in a traceless $`SU(3)`$ matrix $$\mathrm{\Phi }=\frac{\stackrel{}{\lambda }}{\sqrt{2}}\stackrel{}{\varphi }=\left(\begin{array}{ccc}\frac{1}{\sqrt{2}}\pi ^0+\frac{1}{\sqrt{6}}\eta _8& \pi ^+& K^+\\ \pi ^{}& \frac{1}{\sqrt{2}}\pi ^0+\frac{1}{\sqrt{6}}\eta _8& K^0\\ K^{}& \overline{K}^0& \frac{2}{\sqrt{6}}\eta _8\end{array}\right)$$ (2.5) where $`\lambda _i`$ are the Gell-Mann’s matrices with Tr$`(\lambda _i\lambda _j)=2\delta _{ij}`$. From the field $`\mathrm{\Phi }`$ one builds the matrix $`U(\varphi )=e^{i\sqrt{2}\mathrm{\Phi }/f}`$, with $`f`$ a constant. This matrix transforms linearly under $`SU(3)_LSU(3)_R`$ as: $$U(\varphi )g_RU(\varphi )g_L^{}$$ (2.6) and, hence, the Goldstone boson fields $`\stackrel{}{\varphi }`$ transform in a non-linear form. The effective Lagrangian will be constructed as a power expansion series in terms of the external Goldstone momenta and the quark mass matrix. ### 2.3 Lowest and next to leading order $`𝝌𝑷𝑻`$ Lagrangian The lowest order chiral Lagrangian invariant under Lorentz transformations, parity and charge conjugation with only two derivatives and linear in the quark masses is $$_2=\frac{f^2}{4}<D_\mu U^{}D^\mu U+U^{}+^{}U>,$$ (2.7) where $`<>`$ means $`SU(3)`$-flavour trace and $$=2B_0(s+ip),$$ (2.8) with $`B_0`$ a constant and the covariant derivative is defined as $$D_\mu U=_\mu Uir_\mu U+iU\mathrm{}_\mu ,$$ (2.9) where $`r_\mu (\mathrm{}_\mu )=v_\mu +()a_\mu `$. The external sources can also be used to incorporate the electromagnetic and semileptonic weak interactions through the following relations: $`r_\mu `$ $`=`$ $`e𝒬A_\mu +\mathrm{}`$ $`\mathrm{}_\mu `$ $`=`$ $`e𝒬A_\mu +{\displaystyle \frac{e}{\sqrt{2}\mathrm{sin}\theta _W}}(W_\mu ^{}T_++HC)+\mathrm{}`$ (2.10) where $`𝒬={\displaystyle \frac{1}{3}}diag(2,1,1)`$ is the quark-charge matrix and $`T_+`$ is a $`3\times 3`$ matrix containg the relevant Cabibbo-Kobayashi-Maskawa factors $$T_+=\left(\begin{array}{ccc}0& V_{ud}& V_{us}\\ 0& 0& 0\\ 0& 0& 0\end{array}\right).$$ (2.11) On the other hand, the former Lagrangian given in eq. (2.7) is quoted as $`𝒪(p^2)`$ because it contains at most masses squared, see eq. (2.12), and two derivatives. In general, when no baryons are present, the Lagrangian will have an even number of power of masses and derivatives. In this way, we will have $`=_2+_4+_6+\mathrm{}`$, where the subindex indicates the power in the momenta. Fixing $`s(x)=M`$ and $`p(x)=0`$, the $``$ term in eq. (2.7) gives rise to a quadratic pseudoscalar mass term plus additional interactions proportional to the quark masses. This is the reason why in $`\chi PT`$ the quark masses are considered as $`𝒪(p^2)`$. In the isospin limit ($`m_u=m_d`$) with $`\widehat{m}={\displaystyle \frac{m_u+m_d}{2}}`$, the following relations arise from the lowest order $`\chi PT`$ Lagrangian, $`_2`$: $`m_\pi ^2`$ $`=`$ $`2\widehat{m}B_0`$ (2.12) $`m_K^2`$ $`=`$ $`(\widehat{m}+m_s)B_0`$ $`m_{\eta _8}^2`$ $`=`$ $`{\displaystyle \frac{2}{3}}(\widehat{m}+2m_s)B_0`$ satisfying the Gell-Mann-Okubo mass relation. $$3m_{\eta _8}^2=4m_K^2m_\pi ^2.$$ (2.13) From eqs. (2.12), valid in the isospin limit, we can write the mass matrix $``$ ($`s(x)=M`$ and $`p=0`$) present in the lowest order Lagrangian, $`_2`$, as: $$\left(\begin{array}{ccc}m_\pi ^2& 0& 0\\ 0& m_\pi ^2& 0\\ 0& 0& 2m_K^2m_\pi ^2\end{array}\right).$$ (2.14) The meaning of the constant $`f`$ can be appreciated when calculating from the lowest order Lagrangian, eq. (2.7), the axial current. Then $`f`$ becomes the pion decay constant in the chiral limit, that is, $$f=f_\pi +𝒪(m_q).$$ (2.15) The next to leading order Lagrangian, $`_4`$, is constructed with the same building blocks than $`_2`$, namely, eqs. (2.3), (2.5), (2.6) and (2.9). Preserving Lorentz invariance, parity and charge conjugation one has : $`_4`$ $`=`$ $`L_1D_\mu U^{}D^\mu U^2+L_2D_\mu U^{}D_\nu UD^\mu U^{}D^\nu U`$ (2.16) $`+`$ $`L_3D_\mu U^{}D^\mu UD_\nu U^{}D^\nu U+L_4D_\mu U^{}D^\mu UU^{}+^{}U`$ $`+`$ $`L_5D_\mu U^{}D^\mu U\left(U^{}+^{}U\right)+L_6U^{}+^{}U^2`$ $`+`$ $`L_7U^{}^{}U^2+L_8^{}U^{}U+U^{}U^{}`$ $``$ $`iL_9<F_R^{\mu \nu }D_\mu UD_\nu U^{}+F_L^{\mu \nu }D_\mu U^{}D_\nu U>+L_{10}<U^{}F_R^{\mu \nu }UF_{L,\mu \nu }>`$ $`+`$ $`H_1F_{R\mu \nu }F_R^{\mu \nu }+F_{L\mu \nu }F_L^{\mu \nu }+H_2^{}`$ In $`_4`$ there is also the anomalous term although we will not consider it. In the former equation we have also included the strength tensor: $`F_L^{\mu \nu }`$ $`=`$ $`^\mu \mathrm{}^\nu ^\nu \mathrm{}^\mu i[\mathrm{}^\mu ,\mathrm{}^\nu ]`$ (2.17) $`F_R^{\mu \nu }`$ $`=`$ $`^\mu r^\nu ^\nu r^\mu i[r^\mu ,r^\nu ]`$ In the $`_4`$ Lagrangian the terms proportional to $`H_1`$ and $`H_2`$ do not contain the pseudoscalar fields and are therefore not directly measurable. Thus, at $`𝒪(p^4)`$ we need ten additional coupling constants $`L_i`$ to determine the low-energy behaviour of the Green functions. These couplings have an infinite plus a finite part. The infinite part cancels with the infinites from loops, so that at the end only the finite parts, $`L_i^r`$, remain. In $`\chi PT`$ the $`\overline{MS}`$-1 scheme is the usual renormalization scheme. At the present time these $`L_i^r`$ constants have to be fitted to the phenomenology. In general, the number of free parameters increases drastically with the order of the chiral expansion so that for $`_6`$ there are more than one hundred free couplings. This implies that the predictive power of the theory is rapidly lost with higher orders. On the other hand, the convergence of the $`\chi PT`$ series is restricted to low energies, typically for $`\sqrt{s}<500`$ MeV, although this upper limit depends strongly on the process. Note that the lightest well established resonance, the $`\rho `$(770) has a mass of 770 MeV. This resonance introduces a pole in the $`T`$-matrix which cannot be reproduced by a power expansion. Thus, the masses of the heavier states not included in eq. (2.5), put a clear upper limit to the $`\chi PT`$ series and also give us an upper limit of the scale $`\mathrm{\Lambda }_{\chi PT}`$ over which $`\chi PT`$ is constructed $$\frac{𝒪(p^4)}{𝒪(p^2)}\frac{p^2}{\mathrm{\Lambda }_{\chi PT}^2}$$ (2.18) with $`\mathrm{\Lambda }_{\chi PT}M_\rho 1`$ GeV. One can also obtain an estimation of $`\mathrm{\Lambda }_{\chi PT}`$ by taking into account those contributions coming from loops when allowing a change in the regularization scale by a factor of $`𝒪(1)`$ . The result is that $$\mathrm{\Lambda }_{\chi PT}4\pi f_\pi 1.2\text{ GeV}$$ (2.19) which is of the same order of magnitude than $`M_\rho `$. ### 2.4 Chiral Lagrangians with meson resonances Following ref. we include hadron states heavier than the lightest pseudoscalar mesons ($`\pi `$, $`K`$, $`\eta `$). The former states will include vector ($`V`$), axial ($`A`$), scalar ($`S`$) and pseudoscalar ($`P`$) octets and scalar ($`S_1`$) and pseudoscalar ($`P_1`$) singlets. The exchange of these resonances between the Goldstone bosons contains the resonance propagators which for $`p^2<<M_R^2`$ can be expanded as $$\frac{1}{p^2M_R^2}=\frac{1}{M_R^2}\left(1+\frac{p^2}{M_R^2}+\left(\frac{p^2}{M_R^2}\right)^2+\mathrm{}\right)$$ (2.20) giving rise to contributions which should be embodied in the $`\chi PT`$ counterterms. However, from the equation above it is obvious that a resummation to all orders of such local contributions is obtained when including explicit resonance fields. The lowest order Lagrangian with resonance fields, conserving parity and charge conjugation is given in ref. . Its kinetic part, expressing the vector and axial vector octets in terms of antisymmetric tensor fields $`V_{\mu \nu }`$ and $`A_{\mu \nu }`$ (see below), is: $`_{Kin}(R=V,A)`$ $`=`$ $`{\displaystyle \frac{1}{2}}<^\lambda R_{\lambda \mu }_\nu R^{\nu \mu }{\displaystyle \frac{1}{2}}M_R^2R_{\mu \nu }R^{\mu \nu }>`$ $``$ $`{\displaystyle \frac{1}{2}}^\lambda R_{1,\lambda \mu }_\nu R_1^{\nu \mu }+{\displaystyle \frac{1}{4}}M_{R_1}^2R_{1,\mu \nu }R_1^{\mu \nu }`$ $`_{Kin}(R=S,P)`$ $`=`$ $`{\displaystyle \frac{1}{2}}<^\mu R_\mu RM_R^2R^2>+{\displaystyle \frac{1}{2}}^\mu R_1_\mu R_1{\displaystyle \frac{1}{2}}M_{R_1}^2R_1^2`$ (2.21) where the covariant derivative $`_\mu R`$ is defined as $$_\mu R=_\mu R+[\mathrm{\Gamma }_\mu ,R]$$ (2.22) with $$\mathrm{\Gamma }_\mu =\frac{1}{2}\left\{u^{}(_\mu ir_\mu )u+u(_\mu il_\mu )u^{}\right\}$$ (2.23) with $`u`$ such that $`u^2=U`$. The interaction Lagrangians of the octets and singlets of resonances with spin$``$1 to lowest order in the chiral expansion are given in ref. : Vector Octet, J$`{}_{}{}^{\text{PC}}=1^{}`$ $$\frac{F_V}{2\sqrt{2}}<V_{\mu \nu }f_+^{\mu \nu }>+\frac{iG_V}{\sqrt{2}}<V_{\mu \nu }u^\mu u^\nu >$$ (2.24) Axial Octet, J$`{}_{}{}^{\text{PC}}=1^{++}`$ $$\frac{F_A}{2\sqrt{2}}<A_{\mu \nu }f_{}^{\mu \nu }>$$ (2.25) Scalar Octet, J$`{}_{}{}^{\text{PC}}=0^{++}`$ $$c_d<Su_\mu u^\mu >+c_m<S\chi _+>$$ (2.26) Scalar Singlet, J$`{}_{}{}^{\text{PC}}=0^{++}`$ $$\stackrel{~}{c}_dS_1<u_\mu u^\mu >+\stackrel{~}{c}_mS_1<\chi _+>$$ (2.27) Pseudosalar Octet, J$`{}_{}{}^{\text{PC}}=0^+`$ $$id_m<P\chi _{}>$$ (2.28) Pseudoscalar Singlet, J$`{}_{}{}^{\text{PC}}=0^+`$ $$i\stackrel{~}{d}_mP_1<\chi _{}>$$ (2.29) where $`V_{\mu \nu }`$ is $$V_{\mu \nu }=\left(\begin{array}{ccc}\frac{1}{\sqrt{2}}\rho ^0+\frac{1}{\sqrt{6}}w_8& \rho ^+& K^+\\ \rho ^{}& \frac{1}{\sqrt{2}}\rho ^0+\frac{1}{\sqrt{6}}w_8& K^{0}\\ K^{}& \overline{K}^{\mathrm{\hspace{0.17em}0}}& \frac{2}{\sqrt{6}}w_8\end{array}\right)_{\mu \nu }$$ (2.30) and similarly for the rest of the octets. From eq. (2.24) and (2.25) the $`V`$ and $`A`$ resonances only couple at lowest order as octets. In the former equations the vector and axial octets are included as antisymmetric tensor fields, such that if $`|W,p>`$ represents a vector or axial resonance with momentum $`p`$ and mass $`M`$, then $$<0|W_{\mu \nu }|W,p>=iM^1\left[p_\mu ϵ_\nu (p)p_\nu ϵ_\mu (p)\right]$$ (2.31) with $`ϵ_\mu (p)`$ the polarization (axial)vector of the resonance state. The propagator is given by: $$\frac{M^2}{M^2p^2iϵ}[g_{\mu \rho }g_{\nu \sigma }(M^2p^2)+g_{\mu \rho }p_\nu p_\sigma g_{\mu \sigma }p_\nu p_\rho (\mu \nu )]$$ (2.32) In the Lagrangians given above we have also used the following objects which transform as $`SU(3)_V`$ octets: $`u_\mu `$ $`=`$ $`iu^{}D_\mu Uu^{}=u_\mu ^{}`$ (2.33) $`\chi _\pm `$ $`=`$ $`u^{}u^{}\pm u^{}u`$ $`f_\pm ^{\mu \nu }`$ $`=`$ $`uF_L^{\mu \nu }u^{}\pm u^{}F_R^{\mu \nu }u`$ In ref. the $`𝒪(p^4)`$ contributions resulting from the exchange of the above resonances are also studied. Note that this is the first order to which resonance exchange contributes to the $`\chi PT`$ series since their couplings to the Goldstone fields are $`𝒪(p^2)`$. At $`𝒪(p^4)`$ the resonance exchange gives contribution to all the terms of $`_4`$ in eq. (2.16). In fact, it is seen in that reference that, under certain assumptions for the mass and couplings of the scalar resonances, the numerical values obtained for the couplings $`L_i`$ present in $`_4`$ are saturated by the resonance contributions to them. ### 2.5 Chiral Lagrangians with baryons The inclusion of baryons in the chiral formalism is done in a similar way than the one used for the meson resonances, that is, exploiting their well defined linear transformation laws under $`SU(3)_V`$. We consider here the octet of baryons $$B=\left(\begin{array}{ccc}\frac{1}{\sqrt{2}}\mathrm{\Sigma }^0+\frac{1}{\sqrt{6}}\mathrm{\Lambda }^0& \mathrm{\Sigma }^+& p\\ \mathrm{\Sigma }^{}& \frac{1}{\sqrt{2}}\mathrm{\Sigma }^0+\frac{1}{\sqrt{6}}\mathrm{\Lambda }^0& n\\ \mathrm{\Xi }^{}& \mathrm{\Xi }^0& \frac{2}{\sqrt{6}}\mathrm{\Lambda }^0\end{array}\right)$$ (2.34) The lowest order baryon-meson Lagrangian with at most two baryons can be written as: $`_1=<\overline{B}i\gamma ^\mu _\mu B>M_B<\overline{B}B>`$ $`+`$ $`{\displaystyle \frac{1}{2}}D<\overline{B}\gamma ^\mu \gamma _5\{u_\mu ,B\}>`$ $`+`$ $`{\displaystyle \frac{1}{2}}F<\overline{B}\gamma ^\mu \gamma _5[u_\mu ,B]>`$ where $$_\mu B=_\mu B+[\mathrm{\Gamma }_\mu ,B],$$ (2.36) $`\mathrm{\Gamma }_\mu `$ is already defined in Eq. (2.23) and $`D+F=g_A=1.257`$ and $`DF=0.33`$ . Note that, while in the Lagrangians without baryons the number of derivatives is always even, in the case with baryons an odd number of derivatives also appear, and in fact, the former Lagrangian is $`𝒪(p)`$. On the other hand in eq. (2.5) $`M_B`$ is the baryon octet mass in the chiral limit. The baryon mass splitting begins to appear at $`𝒪(p^2)`$ satisfying the Gell-Mann-Okubo mass relation for baryons . Finally, from eq. (2.5) one can easily derive the well known Goldberger-Treiman relation and the Kroll-Ruderman term . ## Chapter 3 Nonperturbative models from chiral symmetry for meson-meson and meson-baryon interactions The effective chiral Lagrangian techniques have become a widespread tool to address the problem of the low energy interactions of Goldstone bosons . We have presented in chapter 2, the $`\chi PT`$ formalism which is the low energy effective theory of the strong interactions (QCD). Another example is the standard model strongly interacting symmetry breaking sector (SISBS) or the effective chiral Lagrangians in solid-state physics for high-$`T_c`$ superconductors . In all the cases, the chiral symmetry constraints are a powerful tool to determine the low energy matrix elements in a systematic way. These Lagrangians consist of an expansion on the powers of the external momenta of the Goldstone bosons over some typical scale $`\mathrm{\Lambda }`$, which is smaller than the masses of the heavier particles. For instance in QCD, resonances typically appear for $`\sqrt{s}0.8`$ GeV, so that $`\mathrm{\Lambda }_{\chi PT}1`$ GeV. Of course, when a resonance appears, there is no way to reproduce it from the perturbative expansion since it is associated to a pole in the scattering amplitude. Furthermore, as explained in section 2.3, there are also higher order corrections, as chiral loops, which make that $`\mathrm{\Lambda }_{\chi PT}<4\pi f_\pi 1.2`$ GeV . As a result, the $`\chi PT`$ expansion is typically valid up to energies around 500 MeV with $`\mathrm{\Lambda }_{\chi PT}1`$ GeV. Nevertheless, the constraints imposed by chiral symmetry breaking are rather powerful and not restricted to the region where $`\chi PT`$ is meant to converge . Another drawback of the effective chiral theories, is the appearance of a fast increasing number of free parameters (not fixed by the symmetry) as one increases the order of the calculation. At $`𝒪(p^2)`$ the $`\chi PT`$ Lagrangian without baryons only contains the masses of pions, kaons and etas and $`f_\pi `$. At $`𝒪(p^4)`$ several new free parameters appear: for instance in $`\chi PT`$ there are 12 parameters and in the SISBS one needs 13. At $`𝒪(p^6)`$ in $`\chi PT`$ there are more than 100 new parameters. That is, the predictive power of the theory is lost as we go higher in order. Because of the former reasons, nonperturbative schemes become necessary in order to go to higher energies and maintain the predictive power of the theory. ### 3.1 The inverse amplitude method with coupled channels An attempt to extend the constraints of chiral symmetry to higher energies, constructing a unitary $`T`$-matrix, is the Inverse Amplitude Method (IAM) . This approach proved efficient in reproducing low energy data and produced poles in the amplitudes associated to the $`\rho `$ and $`K^{}`$ in the vector channel as well as the $`\sigma `$ in the scalar one. It has also been applied to study the SISBS resonances that could appear at LHC . Since only elastic unitarity was imposed in the IAM, multichannel problems could not be addressed. In fact, the treatment of coupled channels has proved to be crucial in order to reproduce the basic features of the $`f_0`$ and $`a_0`$ resonances and in general for all the scalar sector with $`I=0,1,1/2`$ . As a consequence, neither the $`f_0`$ nor $`a_0`$ resonances could be obtained by the elastic IAM . We now proceed to the extension of the IAM with coupled channels which was given for first time in ref. . Let $`T_L^I`$ be a meson-meson partial wave amplitude with definite isospin $`I`$ and angular momentum $`L`$. If $`T`$ is the scattering matrix, $`S=IiT`$ with $`S`$ the $`S`$-matrix, then $`{\displaystyle \frac{T^I}{\sqrt{2}^\alpha }}`$ $`=`$ $`{\displaystyle \underset{L=0}{\overset{\mathrm{}}{}}}(2L+1)T_L^I(s)P_L(\mathrm{cos}\theta )`$ $`T_L^I(s)`$ $`=`$ $`{\displaystyle \frac{1}{2(\sqrt{2})^\alpha }}{\displaystyle _1^1}d\mathrm{cos}\theta P_L(\mathrm{cos}\theta )T^I(s,\mathrm{cos}\theta )`$ (3.1) where $`(\sqrt{2})^\alpha `$ is a symmetry factor to take care of the presence of identical particle states as $`\eta \eta `$ or $`\pi \pi `$ in the isospin limit. The index $`\alpha `$ can be 0,1 or 2 depending on the number of times these identical particle states appear in the corresponding partial wave amplitude. For instance, $`\alpha =2`$ for $`\pi \pi \pi \pi `$, $`\alpha =1`$ for $`\eta \eta K\overline{K}`$, $`\alpha =0`$ for $`K\pi K\pi `$ and so on. $`P_L(\mathrm{cos}\theta )`$ is the Legendre polynomial of $`L_{th}`$ degree. In the following we will omit the indexes $`L`$ and $`I`$ in a partial wave, although it should be kept in mind that we are considering partial wave amplitudes with definite $`L`$ and $`I`$, unless the contrary is said. In our normalization unitarity in coupled channels reads: $$\mathrm{Im}T_{if}=T_{in}\rho _{nn}T_{nf}^{}$$ (3.2) where $`\rho `$ is a real diagonal matrix whose elements account for the phase space of the two meson intermediate states $`n`$ which are physically accessible. With our normalization, $`\rho `$ is given by $$\rho _{nn}(s)=\frac{k_n}{8\pi \sqrt{s}}\theta (s(m_{1n}+m_{2n})^2)$$ (3.3) where $`k_n`$ is the on shell center mass (CM) momentum of the meson in the intermediate state $`n`$ and $`m_{1n},m_{2n}`$ are the masses of the two mesons in this state. Isolating $`\rho `$ from eq. (3.2) one has: $`\rho `$ $`=`$ $`T^1\mathrm{Im}TT^1`$ (3.4) $`=`$ $`{\displaystyle \frac{1}{2i}}T^1(TT^{})T^1`$ $`=`$ $`{\displaystyle \frac{1}{2i}}(T^1T^1)=\mathrm{Im}T^1`$ The former result is in fact the basis for the $`K`$-matrix formalism . From eq. (3.4) we can write: $$T^1=\mathrm{Re}T^1+i\rho K^1+i\rho $$ (3.5) where $`K`$ is the $`K`$-matrix which from the former equation is given by $$K^1=\mathrm{Re}T^1$$ (3.6) Once the $`K`$-matrix ($`\mathrm{Re}T^1`$) is given, the $`T`$-matrix follows by inverting eq. (3.5): $$T=[K^1+i\rho ]^1=[\mathrm{Re}T^1+i\rho ]^1$$ (3.7) We will approach the $`K`$-matrix by expanding $`\mathrm{Re}T^1`$ from the $`𝒪(p^4)`$ $`\chi PT`$ expansion of the $`T`$-matrix. In this way, unitarity will be fulfilled to all orders since we know exactly the imaginary part of $`T^1`$ from eq. (3.4). Another advantage of considering the expansion of $`T^1`$ is clear when $`T`$ has a pole. In this case, $`T^1`$ will have just a zero and hence its expansion will not be affected by this pole of $`T`$. Thus, expanding $`T^1`$ in powers of $`p^2`$ from the expansion of $`T`$, one has: $$TT_2+T_4+\mathrm{}$$ where $`T_2`$ is the lowest order $`\chi PT`$ amplitude and $`T_4`$ the $`𝒪(p^4)`$ contribution, $$T^1=T_2^1[1+T_4T_2^1+\mathrm{}]^1=T_2^1[1T_4T_2^1+\mathrm{}]=T_2^1\left[T_2T_4+\mathrm{}\right]T_2^1$$ Inverting the former equation we finally have: $$T=T_2\left[T_2T_4\right]^1T_2$$ (3.8) Note that eq. (3.1) fulfills the unitarity requirements given in eq. (3.4) because $`\mathrm{Im}T_4=T_2\rho T_2`$ above the physical thresholds. Taking into account eq. (3.6), the $`K`$-matrix resulting from eq. (3.1) is given by: $$K=T_2\left[T_2\mathrm{Re}T_4\right]^1T_2$$ (3.9) Eq. (3.8) is the extension of the IAM to coupled channels. In the next two sections we describe the application of the former formalism to the study of the meson-meson interactions up to $`\sqrt{s}1.2`$ GeV. For higher energies more than two meson states are needed since multipion states become increasingly important. #### 3.1.1 $`𝝅𝝅`$ and $`𝑲\overline{𝑲}`$ amplitudes In this section we are going to present the results of applying the IAM with coupled channels to the study of the $`\pi \pi `$ partial wave amplitudes with $`(I,L)`$=(0,0), (1,1) and (2,0). The pions couple with the $`K\overline{K}`$ channel in the waves (0,0) and (1,1), although, as we will see below, this coupling is negligible in the (1,1) case. This study was developed in ref. . In order to apply eq. (3.8) we need the $`\chi PT`$ amplitudes up to $`𝒪(p^4)`$. For $`\pi \pi \pi \pi `$ this calculation was done in ref. for the $`SU(2)`$ case and extended to $`SU(3)`$ in ref. . The $`\pi \pi K\overline{K}`$ amplitude can be obtained by crossing from the $`K\pi K\pi `$ one calculated in ref. . The $`K\overline{K}K\overline{K}`$ was first calculated in ref. . The amplitudes we are considering in this section depend on the $`L_i`$ couplings of the $`𝒪(p^4)`$ $`\chi PT`$ Lagrangian, eq. (2.16), that enter in their calculations. These constants are: $`L_1`$, $`L_2`$, $`L_3`$, $`L_4`$, $`L_5`$ and $`2L_6+L_8`$. They are fitted to the elastic $`\pi \pi `$ phase shifts in the partial waves $`(I,L)=(0,0)`$ and (1,1) as shown in Fig. 3.1 and 3.2. The fit was done using MINUIT. In the energy region $`\sqrt{s}=`$ 500–950 MeV the data from different experiments for S-wave $`\pi \pi `$ phase shifts are incompatible. Given that situation, the central value for each energy is taken as the mean value between the different experimental results . For $`\sqrt{s}=`$ 0.95–1 GeV, the mean value comes from . In both cases the error is the maximum between the experimental errors and the largest distance between the experimental points and the average value. The fit is good with a $`\chi ^2=1.3`$ per degree of freedom. The values of the $`L_i`$ at the $`M_\rho `$ scale are shown in Table 3.1. In the second column the values obtained from $`\chi PT`$ fits at $`𝒪(p^4)`$ to the low energy data are also shown. We can see that the values obtained by applying the IAM , when the errors are taken into account, are compatible with those from $`\chi PT`$. With the former values for the $`L_i`$ coefficients, the inelastic S-wave phase shifts for $`K\overline{K}\pi \pi `$, Fig. 3.3, $`{\displaystyle \frac{1\eta _{00}^2}{4}}`$ where $`\eta _{00}`$ is the inelasticity in the $`I=L=0`$ channel, Fig. 3.4, and the elastic $`\pi \pi `$ S-wave phase shifts with $`I=2`$, Fig. 3.5, are also calculated. In Fig. 3.3 one sees clearly the $`\eta \eta `$ threshold. Although this channel is not included in the unitarization process, it appears as an intermediate state in the loops of $`T_4`$. If we remove from $`T_4`$ the imaginary part coming from the intermediate $`\eta \eta `$ state above its threshold the dashed line results. It is clear then that the inclusion of the $`\eta \eta `$ channel in the unitarization procedure for the (0,0) channels is important. This point will be further considered in section 3.2.3 and at the end of section 3.3.1. The scattering lengths of the channels $`(I,L)`$=(0,0), (1,1) and (2,0) were also calculated in ref. . These scattering lengths are denoted by $`a_L^I`$. In Table 3.2 we show the values obtained for $`a_L^I`$ in ref. together with the experimental ones and the $`\chi PT`$ values up to $`𝒪(p^4)`$. We see in this table that a good agreement with experiment is accomplished. The values of ref. are also close to the ones from $`\chi PT`$ as one should expect because at low energies the IAM recovers the chiral expansion up to $`𝒪(p^4)`$. #### 3.1.2 Two-meson scattering below 1.2 GeV In the previous section the IAM with coupled channels was applied with the full $`𝒪(p^4)`$ $`\chi PT`$ amplitudes. The calculation of the latter amplitudes is rather involved and it is not done in all two-meson channels, for instance, in those channels with the $`\eta `$ meson. In this section, following ref. , an approximation to calculate the $`𝒪(p^4)`$ amplitude, which turns out to be technically much simpler and rather accurate at the phenomenological level, is presented. Then, eq. (3.8) will be applied to study the partial waves $`(I,L)`$=(0,0), (1,0), (2,0), (1/2,0), (3/2,0), (1,1), (1/2,1) and (0,1). One considers the channels shown in Table 3.3, which are supposed to be the dominant ones up to $`\sqrt{s}1.2`$ GeV. The $`T_4`$ amplitudes are then approximated in ref. by $$T_4T_4^P+T_2g(s)T_2$$ (3.10) where $`T_4^P`$ represents the polynomial tree level amplitudes from the $`𝒪(p^4)`$ Lagrangian, eq. (2.16), and are given in the Appendix A of ref. . On the other hand, $`g(s)`$ is a diagonal matrix corresponding to the loop integral with two meson propagators, given by: $$g_{nn}(s)=i\frac{d^4q}{(2\pi )^4}\frac{1}{q^2m_{1n}^2+iϵ}\frac{1}{(Pq)^2m_{2n}^2+iϵ}$$ (3.11) where $`P`$ is the total initial four-momentum of the two meson system. This $`g`$ matrix has the property $$\mathrm{Im}g_{nn}(s)=\rho _{nn}(s)$$ (3.12) as can be easily checked. The real part of $`g(s)`$ is divergent and requires regularization. In ref. it is evaluated by a cut off regularization with a maximum value, $`q_{max}`$, for the modulus of the three-momentum in the integral. Because the divergence in eq. (3.11) is only logarithmic its calculation making use of dimensional regularization is numerically equivalent when, depending on the considered renormalization scheme, the dimensional regularization scale $`\mu `$ is properly chosen as a function of $`q_{max}`$. For instance, in the $`\overline{MS}1`$ renormalization scheme, the one used in $`\chi PT`$, $`\mu 1.2q_{max}`$ . In section 3.2.2 the expression of $`g(s)`$ in dimensional regularization is given. When approximating $`T_4`$ by eq. (3.10) one is taking into account the close relationship between $`T_4^P`$ and the vector mesons and the dominant role that the unitarization with coupled channels of the lowest order $`\chi PT`$ amplitudes has in the scalar sector . In fact, the approach of ref. follows in the limit $`T_4^P=0`$. With respect to a full $`𝒪(p^4)`$ $`\chi PT`$ calculation, neither tadpoles nor loops in crossed channels are included. These are soft contributions which will be reabsorbed in the $`L_i`$ coefficients, which are now denoted by $`\widehat{L}_i`$ since differences begin to rise even at $`𝒪(p^4)`$ with respect to the $`L_i`$ of $`\chi PT`$. From Fig. 3.6 to 3.13 we show the fit of ref. to the meson-meson S and P-wave experimental data, phase shifts and inelasticities. In this reference an error was detected in the $`T_4^P`$ amplitude $`K^+K^{}K^0\overline{K}^0`$ calculated in ref. (the corrected expression is given in the Appendix of ref. ). As a result, the fit presented in ref. was redone making use of MINUIT and the results are the ones displayed in the former figures. In ref. several sets of values of the $`\widehat{L}_i`$ coefficients were found giving rise to fits of similar quality. The main difference between the different sets of values is in the value of $`\widehat{L}_7`$ which can even change sign. We report here the fit with the $`\widehat{L}_i`$ coefficients closer to the $`L_i`$ of $`𝒪(p^4)`$ $`\chi PT`$. In Table 3.4 the corresponding values of the couplings are given and compared with the ones of $`𝒪(p^4)`$ $`\chi PT`$. As we see from the figures, the results are in rather good agreement with a vast amount of experimental data. The resonances that appear are indicated with their corresponding names. Some comment is needed with respect to the (0,1) channel, Fig. 3.13. In this channel a pole appears with a mass around 910 MeV. Below 1.2 GeV there are two resonances with such quantum numbers. They are the $`\omega `$ and the $`\varphi `$, which fit well within the $`q\overline{q}`$ scheme, with practically ideal mixing, as $`\frac{1}{\sqrt{2}}(u\overline{u}+d\overline{d})`$ and $`s\overline{s}`$, respectively. In the limit of exact SU(3) symmetry these resonances manifest as one antisymmetric octet state and a symmetric singlet state. Since the spatial function of the $`K\overline{K}`$ state is antisymmetric its SU(3) wave function has to be also antisymmetric and therefore it only couples to the antisymmetric octet resonance. Of course, the amplitudes given by eq. (3.10), do contain some SU(3) breaking, but, in this channel only the $`K\overline{K}`$ state is considered, neglecting states with other mesons (like the three pion channel) and, hence, the formulae for this process do not contain any SU(3) symmetry breaking term. Thus, one just sees one pole, $`\omega _8`$, corresponding to the antisymmetric octet state of the exact SU(3) limit. The $`\omega _8`$ meson will be studied in further detail in section 4.2, in connection with the $`\varphi `$ resonance and its decays. ##### Pole positions, widths and partial decay widths In ref. the poles of the $`T`$-matrix in the complex plane, which appear in the unphysical Riemann sheets, are also studied and we refer to this reference for further details. These poles correspond to the $`f_0(980)`$, $`a_0(980)`$, $`\sigma `$, $`\rho (770)`$, $`K^{}(890)`$, $`\kappa `$ and the $`\omega _8`$ resonance. The changes in the values of the masses, partial and total decay widths of the different resonances , due to the error detected in ref. , are rather small except in the case of the $`a_0(980)`$ resonance. For the values of the $`\widehat{L}_i`$ coefficients given in Table 3.4, the pole of the $`a_0(980)`$ in the second sheet ($`k_{\pi ^0\eta }<0`$, $`k_{K\overline{K}}>0`$) appears at $`(1195i\mathrm{\hspace{0.17em}350})`$ MeV. However, a new one appears in the fourth sheet ($`k_{\pi ^0\eta }>0`$, $`k_{K\overline{K}}<0`$) at: $`(1111i\mathrm{\hspace{0.17em}118})`$ MeV. When doing a fit, the $`I=1`$, $`L=0`$ channel has very little statistical weight due to the lack of experimental data so that an improvement of this experimental situation would be very welcome. For instance, we will see in the next section a good reproduction of the experimental scalar data showing an $`a_0(980)`$ resonance with just a pole also in the second sheet. ### 3.2 Inclusion of Explicit Resonance Fields $`\chi PT`$ can be supplied with the exchange of explicit resonance fields . In doing this, a resummation up to an infinite order in the chiral expansion can be achieved from the expansion of the bare propagator of a resonance, as we have already seen in eq. (2.20). However, the amplitudes that can be built directly from $`\chi PT`$ at $`𝒪(p^4)`$ plus resonance exchanges as in ref. need a unitarization procedure, in order to compare directly with experimental data (phase shifts, inelasticities…) for the different energy regions, in particular, around the resonance masses. This is one of the aims of the present section. On the other hand, it is well known that the scalar sector is much more controversial than the vector one. In the latter case, the properties of the associated spectroscopy can be understood in terms of first principles coming directly from QCD as Chiral Symmetry and Large $`N_c`$ plus unitarity, once we admit VMD as dictated by phenomenology . In ref. the same basic principles than before, that is, Chiral Symmetry, Large $`N_c`$ and unitarity in coupled channels, were applied in order to study the scalar resonant channels. We will also pay special attention to the issue of the nature of the scalar resonances that we will find in the amplitudes. As it is well known, the low energy scalar resonances have been ascribed to different models as: conventional $`q\overline{q}`$ mesons , $`q^2\overline{q}^2`$ states , $`K\overline{K}`$ molecules , glueballs and/or hybrids . The question about the nature of the resonances is specially important in order to determine their contributions to the $`L_i`$ counterterms. For instance, the resonances considered in ref. are supposed to be preexisting ones (their masses are $`𝒪(1)`$ in Large $`N_c`$ counting rules) and their contributions to the previous couplings arise dominantly from their bare propagators. However, as it is shown below, there are also resonances, like the $`\sigma `$ or the $`a_0(980)`$, that are originated from the interactions between the pseudoscalars and hence their contributions to the $`L_i`$ come just from the loops of the pseudoscalars. #### 3.2.1 Formalism We present here a formalism based on the N/D method in order to provide physical amplitudes from $`\chi PT`$ supplied with the exchange of explicit resonance fields as given in ref. . This formalism was derived in ref. . A $`T(s)`$ partial wave amplitude has two kinds of cuts. The right hand cut required by unitarity and the unphysical cuts from crossing symmetry. In our chosen normalization, the right hand cut leads to eq. (3.4): $`\mathrm{Im}T^1`$ $`=`$ $`\rho (s)`$ (3.13) for $`s>s_{threshold}s_{th}`$. In the case of two particle scattering, the one we are concerned about, $`s_{th}=(m_1+m_2)^2`$ and $`\rho (s)`$ is given in eq. (3.3). The unphysical cuts comprise two types of cuts in the complex s-plane. For processes of the type $`a+aa+a`$ with $`m_1=m_2=m_a`$, there is only a left hand cut for $`s<s_{Left}`$. However for those of the type $`a+ba+b`$ with $`m_1=m_a`$ and $`m_2=m_b`$, apart from a left hand cut there is also a circular cut in the complex s-plane for $`|s|=m_2^2m_1^2`$, where we have taken $`m_2>m_1`$. In the rest of this section, for simplicity in the formalism, we will just refer to the left hand cut as if it were the full set of unphysical cuts. This will be enough for our purposes in this section. In any case, if one works in the complex $`p^2`$-plane all the cuts will be linear cuts and then only the left hand cut will appear in this variable. The left hand cut, for $`s<s_{Left}`$, reads: $$T(s+iϵ)T(siϵ)=2i\mathrm{Im}T_{Left}(s)$$ (3.14) The standard way of taking into account eqs. (3.13) and (3.14) is the N/D method . In this method a $`T(s)`$ partial wave is expressed as a ratio of two functions, $$T(s)=\frac{N(s)}{D(s)}$$ (3.15) with the denominator function, $`D(s)`$, bearing the right hand cut and the numerator function, $`N(s)`$, the unphysical cuts. In order to take explicitly into account the behavior of a partial wave amplitude near threshold, which vanishes as $`p^{2L}\nu ^L`$, we consider the new quantity, $`T^{}`$, given by: $$T^{}(s)=\frac{T(s)}{\nu ^L}$$ (3.16) which also satisfies relations of the type of eqs. (3.13) and (3.14). Hence, we can write: $$T^{}(s)=\frac{N^{}(s)}{D^{}(s)}$$ (3.17) From eqs. (3.13), (3.14) and (3.16), $`\text{N}^{}(s)`$ and $`\text{D}^{}(s)`$ will obey the following equations: $$\begin{array}{cc}\mathrm{Im}D^{}=\mathrm{Im}T^1N^{}=\rho (s)N^{}\nu ^L,\hfill & s>s_{th}\hfill \\ \mathrm{Im}D^{}=0,\hfill & s<s_{th}\hfill \end{array}$$ (3.18) $$\begin{array}{cc}\mathrm{Im}N^{}=\mathrm{Im}T_{Left}^{}D^{},\hfill & s<s_{Left}\hfill \\ \mathrm{Im}N^{}=0,\hfill & s>s_{Left}\hfill \end{array}$$ (3.19) where $`\mathrm{Im}T_{Left}^{}={\displaystyle \frac{\mathrm{Im}T_{left}}{\nu ^L}}`$. It is important to notice that $`N^{}`$ and $`D^{}`$ can be simultaneously multiplied by any arbitrary real analytic function without changing its ratio, $`\text{T}^{}`$, nor eqs. (3.18) and (3.19). In this way, we can always consider $`N^{}`$ free of poles. Thus, using dispersion relations for $`N_L^{}(s)`$, we write from eq. (3.19): $$N^{}(s)=\frac{(ss_0)^{n+1}}{\pi }_{\mathrm{}}^{s_{Left}}𝑑s^{}\frac{\mathrm{Im}T_{Left}^{}(s^{})D^{}(s^{})}{\nu (s^{})^L(s^{}s_0)^{n+1}(s^{}s)}+\underset{m=0}{\overset{n}{}}\overline{a}_m^{}s^m$$ (3.20) with $`n`$ such that $$\underset{s\mathrm{}}{lim}\frac{N^{}(s)}{s^{n+1}}=0$$ (3.21) In the following, the unphysical cuts will be considered in a perturbative way. In fact, this was the case in the former section where we reported the IAM results from refs. . In this method the unphysical cuts are only considered up to $`𝒪(p^4)`$ as in the $`\chi PT`$ calculations. That is, from the resummation done by the IAM one obtains fully unitarized partial wave amplitudes but satisfying crossing symmetry perturbatively up to $`𝒪(p^4)`$. As a matter of fact, we have seen that one can reproduce rather accurately the meson-meson interactions. Hence, the approach of considering the unphysical cuts in a perturbative way seems to be a realistic one for the physical region. First, we study the zeroth order approach, that is, no unphysical cuts at all, obtaining the most general structure that a partial wave amplitude has in such case. Note that any unitarization method without unphysical cuts must then implement this structure (as an example see the amplitudes of refs. ). Later on, we will also consider the inclusion of the unphysical cuts up to one loop calculated at $`𝒪(p^4)`$. In this case, the $`\chi PT`$ expansion up to $`𝒪(p^4)`$ will be recovered for the low energy region and the IAM, eq. (3.8), will also appear as a special case. ##### No unphysical cuts In this case, we have $`\mathrm{Im}T_{Left}=0`$ and then from eq. (3.20): $$N^{}(s)=\underset{m=0}{\overset{n}{}}a_m^{}s^m$$ (3.22) that is, $`N^{}`$ is just a polynomial. Then, after dividing $`N^{}`$ and $`D^{}`$ by this polynomial, one has: $$N^{}=1$$ (3.23) and the dispersion relation for $`D^{}`$ will read from eq. (3.18) $$D^{}(s)=\frac{(ss_0)^{L+1}}{\pi }_{s_{th}}^{\mathrm{}}𝑑s^{}\frac{\nu (s^{})^L\rho (s^{})}{(s^{}s)(s^{}s_0)^{L+1}}+\underset{m=0}{\overset{L}{}}a_ms^m+\underset{i}{\overset{M_L}{}}\frac{R_i}{ss_i}$$ (3.24) where the last sum takes into account the possible presence of poles of $`D^{}`$ (zeros of $`T^{}`$) inside and along the integration contour, which is given by a circle in the infinity deformed to engulf the real axis along the right hand cut. Each term of this sum is referred to as a Castillejo-Dalitz-Dyson (CDD) pole after ref. . Note that since $`N^{}=1`$ from eq. (3.17) $`T^{}={\displaystyle \frac{1}{D^{}}}`$. Let us come back to QCD and split the subtraction constants $`a_m`$ of eq. (3.24) in two pieces $$a_m=a_m^L+a_m^{SL}(s_0)$$ (3.25) The term $`a_m^L`$ will go as $`N_c`$, because in the $`N_c\mathrm{}`$ limit, the meson-meson amplitudes go as $`N_c^1`$ . Since the integral in eq. (3.24) is $`𝒪(1)`$ in this counting, the subleading term $`a_m^{SL}(s_0)`$ is of the same order and depends on the subtraction point $`s_0`$. This implies that eq. (3.24), when $`N_c\mathrm{}`$, will become $$D^{}(s)D^{\mathrm{}}(s)=\underset{m=0}{\overset{L}{}}a_m^Ls^m+\underset{i}{\overset{M_L^{\mathrm{}}}{}}\frac{R_i^{\mathrm{}}}{ss_i}$$ (3.26) where $`R_i^{\mathrm{}}`$ is the leading part of $`R_i`$ and $`M_L^{\mathrm{}}`$ counts the number of leading CDD poles. Clearly eq. (3.26) represents tree level structures, contact and pole terms, which have nothing to do with any kind of potential scattering, which in large $`N_c`$ QCD is suppressed. In order to determine eq. (3.26) one can make use of $`\chi PT`$ and of the paper . In this latter reference the way to include resonances with spin $`1`$, consistently with chiral symmetry at lowest order in the chiral power counting, is shown. It is also seen that, when integrating out the resonance fields, the contributions of the exchange of these resonances essentially saturate the next to leading $`\chi PT`$ Lagrangian. We will make use of this result in order to state that in the inverse of eq. (3.26) the contact terms come just from the lowest order $`\chi PT`$ Lagrangian and the pole terms from the exchange of resonances in the s-channel in the way given by ref. (consistently with our approximation of neglecting the left hand cut the exchange of resonances in crossed channels is not considered). In the latter statement it is assumed that the result of ref. at $`𝒪(p^4)`$ is also applicable to higher orders. That is, local terms appearing in $`\chi PT`$ and from eq. (3.26) of order higher than $`𝒪(p^4)`$ are also saturated from the exchange of resonances for $`N_c>>1`$, where loops are suppressed. In ref. it is proved that eq. (3.26) can accommodate the tree level amplitudes coming from lowest order $`\chi PT`$ and the Lagrangian given in ref. for the coupling of resonances (with spin$``$1) with the lightest pseudoscalars ($`\pi `$, $`K`$, $`\eta `$). Thus, if we denote by $`T_2`$ the $`𝒪(p^2)`$ $`\chi PT`$ amplitudes and by $`T^R`$ the contribution from the s-channel exchange of resonances according to ref. , we can write: $$T^{\mathrm{}}T_2+T^R=\nu ^L\left[D^{\mathrm{}}\right]^1$$ (3.27) On the other hand, we define the function $`g_L(s)`$ by $$g_L(s)\nu ^L=\underset{m=0}{\overset{L}{}}a_m^{SL}(s_0)s^m\frac{(ss_0)^{L+1}}{\pi }_{s_{th}}^{\mathrm{}}𝑑s^{}\frac{\nu (s^{})^L\rho (s^{})}{(s^{}s)(s^{}s_0)^{L+1}}$$ (3.28) After these definitions, we can write our final formula for $`T(s)`$, in the case that unphysical cuts are not considered, as: $$T(s)=\left[(T^{\mathrm{}})^1g_L(s)\right]^1$$ (3.29) The physical meaning of eq. (3.29) is clear. The $`T^{\mathrm{}}`$ amplitudes correspond to the tree level structures present before unitarization. The unitarization is then accomplished through the function $`g_L(s)`$. From the former comments the generalization of eq. (3.29) to coupled channels should be obvious. In this case, $`T^{\mathrm{}}(s)`$ is a matrix determined by the tree level partial wave amplitudes given by the lowest order $`\chi PT`$ Lagrangian and the exchange of resonances . For instance, $`[T^{\mathrm{}}(s)]_{11}=(T_2)_{11}+T_{11}^R`$, $`[T_L^{\mathrm{}}(s)]_{12}=(T_2)_{12}+T_{12}^R`$ and so on. Once we have $`T^{\mathrm{}}(s)`$ its inverse is the one which enters in eq. (3.29). Because $`N^{}(s)`$ is proportional to the identity, $`g_L(s)`$ will be a diagonal matrix, accounting for the right hand cut, as in the elastic case. In this way, unitarity, which in coupled channels reads (above the thresholds of the channels $`i`$ and $`j`$) $$[\mathrm{Im}T^1]_{ij}=\rho _{ii}(s)\delta _{ij}=\mathrm{Im}g_L(s)_{ii}\delta _{ij}$$ (3.30) is fulfilled. The matrix element $`g_L(s)_{ii}`$ obeys eq. (3.28) with the right masses corresponding to the channel $`i`$ and its own subtraction constants $`a_i^{SL}(s_0)`$. In ref. the coupling constants and resonance masses contained in $`\text{T}_L^{\mathrm{}}(s)`$ are fitted to the experiment. The same happens with the $`a_i^{SL}`$ although, as we will discuss below, they are related by $`SU(3)`$ considerations. In Appendix A of ref. the already mentioned coupled channel version of eq. (3.29) is deduced directly from the N/D method in coupled channels . ##### The unphysical cuts at one loop at $`𝒪(p^4)`$ In a full calculation for a meson-meson partial wave amplitude combining $`\chi PT`$ at $`𝒪(p^4)`$ and the exchange of resonances as done in ref. one has to include the diagrams depicted in Fig. 3.14. The lowest order $`\chi PT`$ amplitudes, $`T_2`$, plus the exchange of resonances in the s-channel, $`T^R`$, were already taken into account in the previous section. The sum of both contributions was denoted by $`T^{\mathrm{}}`$. One also generates through the $`g_L(s)`$ function the loops in the s-channel responsible for unitarity. Thus, in matrix notation at $`𝒪(p^4)`$ we have from eq. (3.29): $`T_2g_L(s)T_2`$. Note that the difference between the loops in the s-channel calculated in $`\chi PT`$ at $`𝒪(p^4)`$ and the ones we have obtained can be at most of a polynomial of second order in $`s`$. We denote the rest of contributions coming from the exchange of resonances and loops in the crossed channels, tadpole-like contributions and the previous difference between our loops in the s-channel and the ones from $`\chi PT`$ by $`T_{Left}`$, since they only have unphysical cuts. Then, in the present notation, a partial wave amplitude calculated as in ref. can be written as: $$T^{\mathrm{}}+T_{Left}+T_2g_L(s)T_2$$ (3.31) An N/D representation of the former amplitude can be done as follows. This representation contains the unphysical cuts up to the order considered in eq. (3.31), that is, up to one loop calculated at $`𝒪(p^4)`$. In matrix formalism: $`N`$ $`=`$ $`T^{\mathrm{}}+T_{Left}`$ (3.32) $`D`$ $`=`$ $`INg_L`$ $`T`$ $`=`$ $`D^1N`$ Making a chiral expansion of the former result, one has: $$T=N+NgN+𝒪(p^6\mathrm{},\mathrm{}^2)=T_0^{\mathrm{}}+T_{Left}+T_2g_LT_2+𝒪(p^6\mathrm{},\mathrm{}^2)$$ (3.33) reproducing eq. (3.31). In the former equation, $`𝒪(p^6\mathrm{},\mathrm{}^2)`$ indicates that the result is valid up to one loop calculated at $`𝒪(p^4)`$. One can also check that, up to the same order in the unphysical cuts, the $`N`$ and $`D`$ functions satisfy eqs. (3.18) and (3.19): $`\mathrm{Im}D`$ $`=`$ $`N\mathrm{Im}g_L=N\rho (s)s>s_{th}`$ (3.34) $`\mathrm{Im}D`$ $`=`$ $`𝒪(p^6\mathrm{},\mathrm{}^2)s\text{ in unphysical cuts}`$ $`\mathrm{Im}N`$ $`=`$ $`D\mathrm{Im}T=\mathrm{Im}T_{Left}+𝒪(p^6\mathrm{},\mathrm{}^2)s\text{ in unphysical cuts}`$ One can reabsorb $`T_0^{\mathrm{}}`$ in $`D`$ just by multiplying $`N`$ and $`D`$ at the same time by $`(T_0^{\mathrm{}})^1`$. In this way, neither their ratio nor their cut structure are modified since $`T_0^{\mathrm{}}`$ is just a matrix of real rational functions. Then one has $`N`$ $`=`$ $`I+(T^{\mathrm{}})^1T_{Left}`$ (3.35) $`D`$ $`=`$ $`(T^{\mathrm{}})^1(I+(T^{\mathrm{}})^1T_{Left})g_L(s)`$ In any case eq. (3.32) can also be written as: $$T=\left[(T^{\mathrm{}}+T_{Left})^1g_L(s)\right]^1$$ (3.36) setting $`T_{Left}`$ to zero we recover once again the limit case of eq. (3.29), where no unphysical cuts are included. The former equation has being used in ref. to describe the coupled channel scattering of $`K\pi `$ and $`K\eta ^{}`$ in order to obtain the scalar $`K\pi `$ form factor. It has also being used in ref. to describe the strong $`WW`$ scattering for a heavy Higgs boson. The main conclusion of this work is that for a general scenario with heavy particles with a mass much larger than $`4\pi v3`$ TeV, a isoscalar-scalar WW resonance would appear with a mass $``$1 TeV. As a consequence, this resonance, which would not be responsible for the spontaneous breaking of the electroweak symmetry $`SU(2)_L\times U(1)_Y`$, could be confused with a true Higgs particle with a mass around 1 TeV at LHC. From eq. (3.36) one can also reobtain the IAM with coupled channels, eq. (3.8). In order to do this, let us expand the inverted matrix in eq. (3.36) analogously as in eq. (3.1): $`(T^{\mathrm{}}+T_{Left})^1g_L(s)`$ $`=`$ $`(T_2+𝒯_4+\mathrm{})^1g_L(s)`$ (3.37) $`=`$ $`T_2^1(1+𝒯_4T_2^1+\mathrm{})^1g_L(s)`$ $`=`$ $`T_2^1(1𝒯_4T_2^1+\mathrm{})g_L(s)`$ $`=`$ $`T_2^1(T_2𝒯_4T_2g_L(s)T_2+\mathrm{})T_2^1`$ $``$ $`T_2^1(T_2T_4)T_2^1`$ In the previous equation $`𝒯_4`$ is the $`𝒪(p^4)`$ contribution of $`T^{\mathrm{}}+T_{Left}`$. Inverting the former result and assuming the saturation of the $`L_i`$ coefficients by the exchange of resonances , one recovers eq. (3.8). #### 3.2.2 Results I: The vector sector In ref. the $`\pi \pi `$ and $`K\pi `$ scattering with $`I`$=$`L`$=1 and $`I`$=1/2, $`L`$=1, respectively, were studied. As we will see below, one reproduces the well known features associated with the vectors: VMD and the KSFR relation for the couplings of these resonances to the pseudoscalars. For the special case of the P-waves, since the zero at threshold is a simple one, instead of eq. (3.24), we consider the slightly modified formula: $$\text{D}(s)=\underset{i}{}\frac{\gamma _i}{ss_i}+a\frac{ss_0}{\pi }_{s_{th}}^{\mathrm{}}𝑑s^{}\frac{\rho (s^{})}{(s^{}s)(s^{}s_0)}$$ (3.38) where the threshold zero has passed to poles in the denominator function, $`D`$. This equation is derived analogously to eq. (3.24) but working directly with $`T`$ rather than with $`T^{}`$. Another advantage of using eq. (3.38) instead of eq. (3.24) is that the comparison with the scalar sector will be more straightforward, because the dispersive integral will be the same. The integral in eq. (3.38) will be evaluated making use of dimensional regularization. It can be identified up to a constant with eq. (3.11). This identification is a consequence of the fact that both the integral in eq. (3.38) and the one in eq. (3.11) have the same cut and the same imaginary part along this cut, as it can be easily checked. Thus, one has: $`g_0(s)`$ $`=`$ $`a^{SL}(s_0){\displaystyle \frac{ss_0}{\pi }}{\displaystyle _{s_{th}}^{\mathrm{}}}𝑑s^{}{\displaystyle \frac{\rho (s^{})}{(s^{}s)(s^{}s_0)}}`$ (3.39) $`=`$ $`{\displaystyle \frac{1}{(4\pi )^2}}[\stackrel{~}{a}^{SL}(\mu )+\mathrm{ln}{\displaystyle \frac{m_2^2}{\mu ^2}}{\displaystyle \frac{m_1^2m_2^2+s}{2s}}\mathrm{ln}{\displaystyle \frac{m_2^2}{m_1^2}}{\displaystyle \frac{\lambda ^{1/2}(s,m_1^2,m_2^2)}{2s}}`$ $`\mathrm{ln}\left({\displaystyle \frac{m_1^2+m_2^2s+\lambda ^{1/2}(s,m_1^2,m_2^2)}{m_1^2+m_2^2s\lambda ^{1/2}(s,m_1^2,m_2^2)}}\right)]`$ for $`ss_{th}`$. For $`s<s_{th}`$ or $`s`$ complex one has the analytic continuation of eq. (3.39). The function $`\lambda ^{1/2}(s,m_1^2,m_2^2)`$ is given by $`\sqrt{(s(m_1+m_2)^2)(s(m_1m_2)^2)}`$. The regularization scale $`\mu `$, appearing in the last equality of eq. (3.39), plays a similar role than the arbitrary subtraction point $`s_0`$ present in the first one. In fact, both of them are arbitrary although the resulting $`g_0(s)`$ function is well defined because any change in $`\mu `$ or $`s_0`$ is reabsorbed in $`\stackrel{~}{a}^{SL}(\mu )`$ or $`a^{SL}(s_0)`$, respectively. The $`\stackrel{~}{a}^{SL}(\mu )`$ ‘constant’ will change under a variation of the scale $`\mu `$ to another one $`\mu ^{}`$ as $$\stackrel{~}{a}^{SL}(\mu ^{})=\stackrel{~}{a}^{SL}(\mu )+\mathrm{ln}\frac{\mu ^2}{\mu ^2}$$ (3.40) in order to have $`g_0(s)`$ invariant under changes of the regularization scale. We will take $`\mu =M_\rho =770`$ MeV . The function $`g_0(s)`$ is also symmetric under the exchange $`m_1m_2`$ and for the equal mass limit it reduces to $$g_0(s)=\frac{1}{(4\pi )^2}\left[\stackrel{~}{a}^{SL}(\mu )+\mathrm{ln}\frac{m_1^2}{\mu ^2}+\sigma (s)\mathrm{ln}\frac{\sigma (s)+1}{\sigma (s)1}\right]$$ (3.41) with $$\sigma (s)=\sqrt{1\frac{4m_1^2}{s}}$$ (3.42) Let us consider first the case of the P-wave $`\pi \pi `$ scattering . Taking into account the zero at threshold, from eq. (3.38) we have: $$\text{T}^{\pi \pi }(s)=\left[\frac{\gamma _1^{\pi \pi }}{s4m_\pi ^2}+\stackrel{~}{a}_{\pi \pi }^Lg_0^{\pi \pi }(s)+\underset{i=2}{}\frac{\gamma _i}{ss_i}\right]^1$$ (3.43) On the other hand, from $`\chi PT`$ and the exchange of the $`\rho `$ , one has: $$\text{T}^{\pi \pi \mathrm{}}(s)=\frac{2}{3}\frac{p_{\pi \pi }^2}{f^2}+g_v^2\frac{2}{3}\frac{p_{\pi \pi }^2}{f^2}\frac{s}{sM_\rho ^2}$$ (3.44) with $`p_{\pi \pi }^2`$ the three-momentum squared of the pions in the c.m., $`f=87.3`$ MeV the pion decay constant in the chiral limit . The deviation of $`g_v^2`$ with respect to unity measures the variation of the value of the $`\rho `$ coupling to two pions with respect to the KSFR relation , $`g_v^2=1`$. In ref. this KSFR relation is justified making use of large $`N_c`$ QCD (neglecting loop contributions) and an unsubtracted dispersion relation for the pion electromagnetic form factor (a QCD inspired high-energy behavior). Comparing eqs. (3.43) and (3.44), one needs only one additional CDD pole apart from the one at threshold and we obtain $`\stackrel{~}{a}^L`$ $`=`$ $`0`$ $`\gamma _1^{\pi \pi }`$ $`=`$ $`{\displaystyle \frac{6f^2(4m_\pi ^2M_\rho ^2)}{(M_\rho ^24m_\pi ^2(1g_v^2))}}`$ $`\gamma _2^{\pi \pi }`$ $`=`$ $`{\displaystyle \frac{6f^2}{g_v^21}}{\displaystyle \frac{g_v^2M_\rho ^2}{M_\rho ^2(1g_v^2)4m_\pi ^2}}`$ $`s_2`$ $`=`$ $`{\displaystyle \frac{M_\rho ^2}{g_v^21}}`$ (3.45) The former equation is an example of the matching between both representations: the one given by the N/D method , eq. (3.38), and the one derived from chiral symmetry . In the following, we will not consider more this matching and we will take directly $`T^{\mathrm{}}`$ as given by the lowest order $`\chi PT`$ amplitudes and the exchange of resonances in the s-channel , as discussed above. For the P-wave $`I=1/2`$ $`K\pi `$ elastic amplitude the tree level amplitude, $`T^{K\pi \mathrm{}}`$, is the same than for pions but multiplying it by 3/4 and substituting $`p_{\pi \pi }`$ by $`p_{K\pi }`$ and $`M_\rho `$ by $`M_K^{}`$, the mass of the $`K^{}(890)`$ resonance. The subleading constant $`\stackrel{~}{a}^{SL}`$ present in $`g_0(s)`$, eq. (3.39), should be the same for the $`\pi \pi `$ and $`K\pi `$ states because the dependence of the loop in eq. (3.11) on the masses of the intermediate particles is given by eq. (3.39). This point can be used in the opposite sense. That is, if it is not possible to obtain a reasonable good fit after setting $`\stackrel{~}{a}^{SL}`$ to be the same in both channels, some kind of SU(3) breaking is missing. Making use of the minimization program MINUIT, in ref. a simultaneous fit to the elastic $`\pi \pi `$ and $`K\overline{K}`$ phase shifts from threshold up to $`\sqrt{s}1.2`$ GeV was given. As a result one has: $`g_v^2`$ $`=`$ $`0.879\pm 0.016`$ $`\stackrel{~}{a}^{SL}`$ $`=`$ $`0.341\pm 0.042`$ (3.46) the errors are just statistical and are obtained by increasing in one unit the $`\chi ^2`$ per degree of freedom, $`\chi _{d.o.f.}^2`$. The $`\chi _{d.o.f.}^2`$ obtained is around 0.8. The fact that $`g_v`$ deviates from 1 just by a 6$`\%`$ states clearly that the KSFR result is phenomenologically successful. #### 3.2.3 Results II: The scalar sector We know consider the S-wave $`I`$=0,1 and 1/2 amplitudes . For the partial wave amplitudes with $`L`$=0 and $`I`$=0 and 1, coupled channels are fundamental in order to get an appropriate description of the physics involved up to $`\sqrt{s}1.3`$ GeV. This is an important difference with respect to the former vector channels, essentially elastic in the considered energy region. Up to $`\sqrt{s}=1.3`$ GeV the most important channels are: $$\begin{array}{cc}\text{I=0}\hfill & \pi \pi (1),K\overline{K}(2),\eta \eta (3)\hfill \\ \text{I=1}\hfill & \pi \eta (1),K\overline{K}(2)\hfill \\ \text{I=1/2}\hfill & K\pi (1),K\eta (2)\hfill \end{array}$$ (3.47) where the number between brackets indicates the index associated to the corresponding channel when using a matrix notation. For the $`I`$=0 S-wave, the $`4\pi `$ state becomes increasingly important at energies above $`1.21.3`$ GeV, so that, in this channel, one is at the limit of applicability of only two meson states when $`\sqrt{s}`$ is close to 1.4 GeV. In the $`I`$=1/2 channel, the threshold of the important $`K\eta ^{}`$ state is also close to 1.4 GeV. Thus, one cannot go higher in energies in a realistic description of the scalar sector without including the $`K\eta ^{}`$ and $`4\pi `$ states. In order to fix $`T^{\mathrm{}}`$ one needs to include explicit resonance fields. From ref. , two sets of resonances appear in the $`L=0`$ partial wave amplitudes. A first one, with a mass around 1 GeV, contains the $`I`$=0 $`f_0(4001200)`$ and $`f_0(980)`$ and the $`I`$=1 $`a_0(980)`$. A second set appears with a mass around $`1.4`$ GeV as the $`I`$=0 $`f_0(1370)`$ and the $`f_0(1500)`$, the $`I`$=1 $`a_0(1450)`$ or the $`I`$=1/2 $`K_0^{}(1430)`$. As a consequence, we first include the exchange of two scalar nonets, with masses around 1 and 1.4 GeV. In ref. the expressions for the $`T_{ij}^{\mathrm{}}`$ partial waves are collected. Once again, the minimization program MINUIT was used in order to fit the SU(3) related experimental data represented in Figs. 3.15–3.19. It happens that the couplings of the octet around 1 GeV and the singlet around 1.4 GeV are compatible with zero, giving rise to very narrow peaks that are not seen in experiment. In fact, one obtains an equally good fit by including one singlet around 1 GeV and an octet around 1.4 GeV. The values of the parameters given by the fit are : $$\begin{array}{cc}\text{Nonet (MeV)}\hfill & \\ c_d=19.1_{2.1}^{+2.4}\hfill & a^{SL}=.75\pm 0.2\hfill \\ c_m=15\pm 30\hfill & 𝒩=(9.4\pm 4.5)\mathrm{\hspace{0.17em}10}^5\text{MeV}^2\hfill \\ M_8=1390\pm 20\hfill & \\ \stackrel{~}{c}_d=20.9_{1.0}^{+1.6}\hfill & \chi _{d.o.f.}^2=1.07\hfill \\ \stackrel{~}{c}_m=10.6_{3.5}^{+4.5}\hfill & 188\text{points}\hfill \\ M_1=1021_{20}^{+40}\hfill & \end{array}$$ (3.48) ##### Resonances We consider now the resonance content of the former fit. The octet around 1.4 GeV gives rise to eight resonances which appear with masses very close to the physical ones, $`f_0(1500)`$, $`a_0(1450)`$ and $`K_0^{}(1430)`$ . However, a detailed study of the former resonances is not given because one has not included channels which become increasingly important for energies above $``$ 1.3 GeV as $`4\pi `$ in $`I`$=0 or $`K\eta ^{}`$ for $`I`$=1/2. This makes that the widths obtained from the pole position of the former resonances are systematically smaller than the experimental ones . Thus, a more detailed study, which included all the relevant channels for energies above 1.3 GeV, should be done in order to obtain a better determination of the parameters for this octet around 1.4 GeV. On the other hand, from Figs. 3.15 and 3.18, one can easily see two resonances with masses around 1 GeV, the well known $`f_0(980)`$ and $`a_0(980)`$ resonances. The first one is related to the singlet bare state with $`M_1=1020`$ MeV, but for the second there are no bare resonances to associate with, because the tree level resonance was included with a mass around 1.4 GeV and has evolved to the physical $`a_0(1450)`$. The situation is even more complex, because we also find in the amplitudes other poles corresponding to the $`f_0(4001200)\sigma `$ and to the $`K_0^{}\kappa `$. In Table 3.5 the pole positions of the resonances in the second sheet<sup>1</sup><sup>1</sup>1I sheet: Im $`p_1>`$0, Im $`p_2>`$0, Im $`p_3>`$0; II sheet: Im $`p_1<`$0, Im $`p_2>`$0, Im $`p_3>`$0 are given and also the modulus of the residues corresponding to the resonance $`R`$ and channel $`i`$, $`\zeta _i^R`$, given by $$|\zeta _i^R\zeta _j^R|=\underset{ss_R}{lim}|(ss_R)\text{T}_{ij}|$$ (3.49) where $`s_R`$ is the complex pole for the resonance $`R`$. While for the $`f_0(980)`$ one has a preexisting tree level resonance with a mass of 1020 MeV, for the other resonances present in Table 3.5 the situation is rather different. In fact, if one removes the tree level nonet contribution from $`T^{\mathrm{}}`$, the $`a_0(980)`$, $`\sigma `$ and $`\kappa `$ poles still appear as can be seen in Table 3.6. For the $`f_0(980)`$, in such a situation, one has not a pole but a very strong cusp effect in the opening of the $`K\overline{K}`$ threshold. In fact, by varying a little the value of $`a^{SL}`$ one can regenerate also a pole for the $`f_0(980)`$ from this strong cusp effect. In Table 3.6 we have not given an absolute value for the coupling of the $`f_0(980)`$ to the $`K\overline{K}`$ channel because one has not a pole for the given value of $`a^{SL}`$. However, the ratios between the different amplitudes are stable around the cusp position. As a result, the physical $`f_0(980)`$ will have two contributions: one from the bare singlet state with $`M_1=1020`$ MeV and the other one coming from meson-meson scattering, particularly $`K\overline{K}`$ scattering, generated by the lowest order $`\chi PT`$ Lagrangian. When the resonant tree level contributions are removed from $`T^{\mathrm{}}`$, only the lowest order, $`𝒪(p^2)`$, $`\chi PT`$ contributions remain. Thus, except for the contribution to the $`f_0(980)`$ coming from the bare singlet at 1 GeV, the poles present in Table 3.6 originate from a ‘pure potential’ scattering, following the nomenclature given in ref. . In this way, the source of the dynamics is the lowest order $`\chi PT`$ amplitudes. The constant $`a^{SL}`$ can be interpreted from the need to give a ‘range’ to this potential so that the loop integrals converge. In ref. it is also shown that these meson-meson states in the SU(3) limit, equaling the masses of the pseudoscalars, appear as a degenerate octet plus a singlet. Finally, in ref. an estimation of the unphysical cut contributions was done from ref. up to $`\sqrt{s}800`$ MeV for the resonance $`L=0`$ partial waves. As shown in Table 3.7 the influence is rather small in the physical region. ### 3.3 Bethe-Salpeter equation for S-wave meson-meson and meson-baryon In this sections we report about the use of the Bethe-Salpeter equation for the study of the S-wave meson-meson scattering and for the S-wave meson-baryon system with strangeness($`S`$)=$``$1 . In both cases the potential is the lowest order $`\chi PT`$ amplitude. The use of the Bethe-Salpeter equation together with $`\chi PT`$ was first considered in ref. in the meson-baryon sector. One of the advantages of the approach in refs. is that the Bethe-Salpeter equation, which is an integral equation, is reduced to an algebraic one. This is accomplished through an analysis of the renormalization process embodied in a Bethe-Salpeter scattering equation. #### 3.3.1 Bethe-Salpeter equation for S-wave meson-meson scattering To see how the former simplification occurs let us consider eq. (3.29). In that section, $`T^{\mathrm{}}`$ was defined to be the sum of the lowest order $`\chi PT`$ amplitudes plus the s-channel resonance exchanges. If we consider only the contribution from the lowest order $`\chi PT`$ we will have from that equation: $$T=\left[T_2^1g_0(s)\right]^1=\left[1T_2g_0(s)\right]^1T_2$$ (3.50) We can rewrite the previous equation in a form that will remind us of the Bethe-Salpeter equation: $$T=T_2+T_2g_0(s)T$$ (3.51) with $`g_0(s)`$ given in eq. (3.11). The above equation would correspond to a Bethe-Salpeter equation with a potential given by the corresponding lowest order $`\chi PT`$ partial wave, $`T_2`$. However, while a true Bethe-Salpeter equation is an integral equation the former one is algebraic. Note that one should have instead of $`T_2g_0(s)T`$ in eq. (3.51) the integral: $$(T_2g_0T)_{\alpha \beta }=\underset{j}{}\frac{d^4q}{(2\pi )^4}T_2(k,p;q)_{\alpha j}\frac{i}{\left(q^2m_{1j}^2\right)\left((Pq)^2m_{2j}^2\right)}T(q;k^{},p^{})_{j\beta }$$ (3.52) where P is the total momentum, $`k`$ and $`p`$ represent the initial momenta of the ingoing mesons and $`k^{}`$ and $`p^{}`$ the final momenta of the outgoing ones. Only when $`T_2`$ and $`T`$ are factorized on shell outside the integral in eq. (3.52) one recovers eq. (3.51). This is exactly what happens as stated by eq. (3.50). However, when considering this equation one has to recall also eq. (3.24) in order to realize that all the parameters that appear in $`T_2`$ have to be the physical or renormalized ones corresponding to the final positions and residues of the CDD poles or to the substraction constant that result from eq. (3.50). The final result present in ref. is obtained when approximating the ’physical values’ of the parameters in $`T_2`$ by the ones dictated by the lowest order $`\chi PT`$ results. That is, the constant $`f`$ is translated to $`f_\pi `$ and for the bare masses one takes the physical ones. Both in the IAM or in the N/D method with unphysical cuts, the parameters appearing in $`T_2`$ are renormalized according to $`\chi PT`$ at $`𝒪(p^4)`$. The argumentation given in ref. for factorizing on shell the potential and the physical amplitude in eq. (3.52) is discussed below and will be considered in further detail when discussing the Bethe-Salpeter equation for the meson-baryon scattering in the next subsection . Briefly, in the former reference the potential was splitted in two parts: $$T_2=V_{onshell}+V_{offshell}$$ (3.53) The V$`{}_{onshell}{}^{}(p_i)`$ part, with $`p_i^2=m_i^2`$, factorizes out of the integral in eq. (3.52) since it only depends of the Mandelstam variable $`s`$. For V<sub>off-shell</sub> it was realized that since it only involves terms proportional to $`p_i^2m_i^2`$, they cancel one of the mesons propagators in the loop and give rise to tadpole-like contributions which can be reabsorbed in the final values of the parameters present in $`T_2`$. So that, at the end, one only needs V<sub>on-shell</sub> and hence, eq. (3.51) follows. This result has been recently derived from an alternative point of view in ref. . In ref. the $`g_0(s)`$ function was calculated making use of a cut off regularization. The cut off<sup>2</sup><sup>2</sup>2The relation between $`\mathrm{\Lambda }`$ and the three momentum cut off $`q_{max}`$, introduced in section 3.1.2, is such that $`\mathrm{\Lambda }=\sqrt{m_K^2+q_{max}^2}`$., $`\mathrm{\Lambda }`$, was fixed to reproduce the experimental points, or in other words, to give the right value for the substraction constant $`a_0^{SL}`$ in eq. (3.28). On the other hand, making use of the IAM one also recovers eq. (3.50) when putting $`T_4^P=0`$ in eq. (3.10). In Fig. 3.20 we show the results of ref. compared with data. The agreement is rather good and surprising. On the other hand, poles corresponding to the resonances $`f_0(980)`$, $`a_0(980)`$ and $`\sigma (500)`$ were found and their masses, partial and total decay widths were also analyzed. Hence, we see that for the scalar sector with $`I`$=0,1 the unitarization of the lowest order $`\chi PT`$ amplitudes plays a very important role and also that the resonances which appear there with masses $``$1 GeV will have a large meson-meson component in their nature. The same results are expected to hold in the $`I`$=1/2 S-wave amplitude by $`SU(3)`$ symmetry as we have seen in section 3.2.3. In the P-waves, where the resonances that appear there are of preexisting nature, one cannot reproduce the substraction constant $`a^L`$ present in eq. (3.26) by a reasonable cut off and the method fails. In ref. it is argued that a cut off around 300 GeV would be needed to reproduce the $`\rho `$ which is a senseless result. On the other hand, whereas in section 3.2.3, ref. , one needs seven free parameters, a nonet of resonances with 6 parameters and a subleading constant, for describing the scalar sector, in this section, ref. , there is only one free constant, a cut off, for the $`I=0`$ and 1 S-waves. This is due to: 1) in ref. the fit is pushed up to $`\sqrt{s}1.4`$ GeV, while in ref. the fit is up to $`\sqrt{s}=1.2`$ GeV. In fact, the effect of this octet around 1.4 GeV is soft enough below 1.2 GeV to be reabsorbed in the cut off (or subleading constant). In this way, below 1.2 GeV, one would have needed only 4 parameters in the approach of ref. . 2) In this latter reference the $`\eta \eta `$ channel is included and in order to reproduce the $`{\displaystyle \frac{1\eta _{00}^2}{4}}`$ data one has had to include the singlet resonance around 1 GeV. The $`\eta \eta `$ channel was not considered, however, in ref. . Should one have taken the available data for $`\eta _{00}`$, which are measured with much worse precision than $`{\displaystyle \frac{1\eta _{00}^2}{4}}`$, the effect of the $`\eta \eta `$ channel would have been masked by the large errors in $`\eta _{00}`$. #### 3.3.2 Bethe-Salpeter equation for S-wave meson-baryon scattering The effective chiral Lagrangian techniques, which successfully describe meson-meson scattering at low energies , have also proved to be an excellent tool to study low energy properties of meson-baryon systems when the interaction is weak, as is the case of the S-wave $`\pi N`$ and $`K^+N`$ interactions , where the leading term in the chiral expansion $`𝒪(p)`$ is the dominant one close to threshold. The perturbative scheme breaks down –even at low energies– in the vicinity of a resonance. This is the case of the $`\overline{K}N`$ system in the $`S=1`$ sector with an isospin zero S-wave resonance, the $`\mathrm{\Lambda }(1405)`$. Originally treated as a bound $`\overline{K}N`$ state , this resonance was later interpreted as a conventional three-quark system . Analysis in terms of the cloudy-bag model reinforced the idea of the $`\mathrm{\Lambda }(1405)`$ being a $`\overline{K}N`$ bound state and, in the framework of the bound-state soliton model, it corresponds to a bound state of a kaon in the background potential of the soliton . The fact that the $`\mathrm{\Lambda }(1405)`$ resonance is located 27 MeV below the $`K^{}p`$ threshold makes it difficult to reproduce the scattering observables whithin the standard chiral Lagrangian techniques, unless the resonance is explicitly introduced as an elementary field or one resorts to nonperturbative techniques similar to those reviewed in previous sections for meson-meson scattering. A nonperturbative scheme to study the $`S=1`$ meson-baryon sector, yet using the input of the chiral Lagrangians, was employed in . A potential model was constructed such that, in Born approximation, it had the same S-wave scattering length as the chiral Lagrangian up to order $`p^2`$. This potential, which includes also finite range factors to regularize the integrals, was inserted in a set of coupled-channel Lippmann Schwinger integral equations. The channels included were those opened around the $`K^{}p`$ threshold, namely $`K^{}p,\overline{K}^0n,\pi ^0\mathrm{\Lambda },\pi ^+\mathrm{\Sigma }^{},\pi ^0\mathrm{\Sigma }^0`$ and $`\pi ^{}\mathrm{\Sigma }^+`$. By fiting five parameters, corresponding to, so far, unknown parameters of the second order chiral Lagrangian plus the range parameters of the potential, the $`\mathrm{\Lambda }(1405)`$ resonance was generated as a quasibound meson-baryon state and the cross sections of the $`K^{}pK^{}p,\overline{K}^0n,\pi ^0\mathrm{\Lambda },\pi ^+\mathrm{\Sigma }^{},\pi ^0\mathrm{\Sigma }^0,\pi ^{}\mathrm{\Sigma }^+`$ reactions at low energies, plus the threshold branching ratios $`\gamma `$ $`=`$ $`{\displaystyle \frac{\mathrm{\Gamma }(K^{}p\pi ^+\mathrm{\Sigma }^{})}{\mathrm{\Gamma }(K^{}p\pi ^{}\mathrm{\Sigma }^+)}}`$ $`R_c`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Gamma }(K^{}p\mathrm{charged}\mathrm{channels})}{\mathrm{\Gamma }(K^{}p\mathrm{all}\mathrm{channels})}}`$ (3.54) $`R_n`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Gamma }(K^{}p\pi ^0\mathrm{\Lambda })}{\mathrm{\Gamma }(K^{}p\mathrm{neutral}\mathrm{channels})}}`$ were well reproduced. Although the cross sections and position of the resonance could also be reproduced with only the lowest order Lagrangian and one potential range parameter, the impossibility of obtaining a good result for the double charge exchange ratio $`\gamma `$ was the reason for the need of including the S-wave terms of the next-to-leading order Lagrangian. However, a recent work , which shares many points with , showed that all the strangeness $`S=1`$ meson-baryon scattering observables near threshold were reproduced with the lowest order Lagrangian and one cut off. The main reason was the inclusion of the $`\eta \mathrm{\Lambda }`$ channel, neglected in ref. . The chiral scheme employed in ref. , and summarized below, includes all meson-baryon states that can be generated from the octet of pseudoscalar mesons and the octet of ground-state baryons, thus including in addition the $`\eta \mathrm{\Lambda },\eta \mathrm{\Sigma }^0`$ and the $`K^+\mathrm{\Xi }^{},K^0\mathrm{\Xi }^0`$ channels, adding up to a total of 10. It should be noted that if one sets equal baryon masses and equal meson masses in the scheme, one should get degenerate SU(3) multiplet states. This is only possible if all states of the $`0^{}`$ meson and $`\frac{1}{2}^+`$ baryon octets are included in the scheme and one should start from such a situation to have control on the SU(3) breaking due to unequal masses. At lowest order in momentum the interaction Lagrangian comes from the $`\mathrm{\Gamma }_\mu `$ term in the covariant derivative of eqs. (2.5) and (2.36) $$_1^{(B)}=\overline{B}i\gamma ^\mu \frac{1}{4f^2}[(\mathrm{\Phi }_\mu \mathrm{\Phi }_\mu \mathrm{\Phi }\mathrm{\Phi })BB(\mathrm{\Phi }_\mu \mathrm{\Phi }_\mu \mathrm{\Phi }\mathrm{\Phi })]$$ (3.55) which leads to a common structure for the meson-baryon amplitudes of the type $$V_{ij}=\frac{C_{ij}}{4f^2}\overline{u}(\stackrel{}{k^{}})\gamma ^\mu (k_\mu +k_\mu ^{})u(\stackrel{}{k})$$ (3.56) for the different channels, where $`u,\overline{u}`$ are the Dirac spinors and $`k,k^{}`$ the momenta of the incoming and outgoing mesons in the center of mass of the meson-baryon system. The particular values for the $`\overline{K}N`$ system of the SU(3) coefficients $`C_{ij}`$, which connect the meson-baryon channels $`j`$ and $`i`$, can be found in ref. . At low energies the spatial components can be neglected and the amplitudes reduce to $$V_{ij}=C_{ij}\frac{1}{4f^2}(k^0+k^0).$$ (3.57) This amplitude is inserted in a coupled channel Bethe-Salpeter equation $$T_{ij}=V_{ij}+\overline{V_{il}G_lT_{lj}}$$ (3.58) with $$\overline{V_{il}G_lT_{lj}}=i\frac{d^4q}{(2\pi )^4}\frac{M_l}{E_l(\stackrel{}{q})}\frac{V_{il}(k,q)T_{lj}(q,k^{})}{k^0+p^0q^0E_l(\stackrel{}{q})+iϵ}\frac{1}{q^2m_l^2+iϵ},$$ (3.59) where only the positive energy component of the fermion propagator has been kept. The Bethe-Salpeter equation sums up automatically the series of diagrams of Fig. 3.21. The loop integral in eq. (3.59) is logarithmically divergent and can be regularized by a cut off $`q_{\mathrm{max}}`$. The quantities $`M_l`$ and $`E_l`$ in eq. (3.59) correspond to the mass and energy of the intermediate baryon, while $`m_l`$ is the mass of the intermediate meson and $`k^0+p^0\sqrt{s}`$ is the total energy in the center of mass frame. The same arguments given in the previous section for the meson-meson sector allow one to retain only the on shell part of the amplitudes appearing in eq. (3.59), while the rest goes into renormalization of couplings and masses. Take, as an example, the one loop diagram of Fig. 3.21 with equal masses in the external and intermediate states for simplicity. We have $`V_{\mathrm{off}}^2`$ $`=`$ $`C(k^0+q^0)^2=C(2k^0+q^0k^0)^2`$ (3.60) $`=`$ $`C(2k^0)^2+2C(2k^0)(q^0k^0)+C(q^0k^0)^2`$ with $`C`$ a proportionality constant. The first term in the last expression is the on shell contribution $`V_{\mathrm{on}}^2`$, with $`V_{\mathrm{on}}C2k^0`$. Neglecting $`p^0E(\stackrel{}{q})`$ in eq. (3.59), a typical approximation in the heavy baryon formalism, the one loop integral for the second term of eq. (3.60) becomes $`2iV_{\mathrm{on}}{\displaystyle \frac{d^3q}{(2\pi )^3}\frac{dq^0}{2\pi }\frac{M}{E(\stackrel{}{q})}\frac{q^0k^0}{k^0q^0}\frac{1}{q^{02}\omega (\stackrel{}{q})^2+iϵ}}`$ (3.61) $`=`$ $`2V_{\mathrm{on}}{\displaystyle \frac{d^3q}{(2\pi )^3}\frac{M}{E(\stackrel{}{q})}\frac{1}{2\omega (\stackrel{}{q})}}V_{\mathrm{on}}q_{\mathrm{max}}^2`$ with $`\omega (\stackrel{}{q})^2=\stackrel{}{q}^2+m^2`$. This term, proportional to $`V_{\mathrm{on}}`$, has the same structure as the tree level term in the Bethe-Salpeter series and it can be reabsorbed in the lowest order Lagrangian by a suitable renormalization of the parameter $`f`$. Similarly, one of the two $`(q^0k^0)`$ factors in the last term of eq. (3.60) cancels the baryon propagator in eq. (3.59) while the remaining factor gives rise to another term proportional to $`k^0`$ (and hence $`V_{\mathrm{on}}`$) and a term proportional to $`q^0`$, which vanishes for parity reasons. These arguments can be extended to coupled channels and higher order loops with the conclusion that $`V_{il}`$ and $`T_{lj}`$ factorize with their on shell values out of the integral in eq. (3.59), reducing the problem to one of solving a set of algebraic equations, written in matrix form as $$T=V+VGT$$ (3.62) with $`G`$ a diagonal matrix given by $`G_l(\sqrt{s})`$ $`=`$ $`i{\displaystyle \frac{d^4q}{(2\pi )^4}\frac{M_l}{E_l(\stackrel{}{q})}\frac{1}{\sqrt{s}q^0E_l(\stackrel{}{q})+iϵ}\frac{1}{q^2m_l^2+iϵ}}`$ (3.63) $`=`$ $`{\displaystyle _{\stackrel{}{q}<q_{\mathrm{max}}}}{\displaystyle \frac{d^3q}{(2\pi )^3}}{\displaystyle \frac{1}{2\omega _l(\stackrel{}{q})}}{\displaystyle \frac{M_l}{E_l(\stackrel{}{q})}}{\displaystyle \frac{1}{\sqrt{s}\omega _l(\stackrel{}{q})E_l(\stackrel{}{q})+iϵ}}.`$ The value of the cut off in ref. , $`q_{\mathrm{max}}=630`$ MeV, was chosen to reproduce the $`K^{}p`$ threshold branching ratios , while the weak decay constant, $`f=1.15f_\pi `$, was taken in between the pion and kaon ones to optimize the position of the $`\mathrm{\Lambda }(1405)`$ resonance . The branching ratios as well as the predictions of the scattering lengths for $`K^{}p`$ and $`K^{}n`$ scattering are summarized in Table 3.8, where results omitting the $`\eta `$ channels are also shown. While the $`\eta `$ channels have a moderate effect on the isospin $`I=1`$ $`K^{}n`$ scattering length, $`a_{K^{}n}`$, they have a tremendous influence on the $`K^{}p`$ scattering observables, especially on the ratio $`\gamma `$ which changes by a factor of about 2. As was shown in ref. , it is the $`I=0`$ $`\eta \mathrm{\Lambda }`$ channel the one that was providing most of the changes. The $`K^{}p`$ scattering length is essentially in agreement with the most recent results from Kaonic hydrogen $`X`$ rays , in qualitative agreement with the scattering lengths determined from scattering data in , with an estimated error of 15%, and in remarkable agreement with the result from a combined dispersion relation and $`M`$-matrix analysis . Finally, the $`K^{}p`$ cross sections for some selected channels ($`K^{}pK^{}\pi `$, $`\overline{K}^0n`$, $`\pi ^+\mathrm{\Sigma }^{}`$, $`\pi ^{}\mathrm{\Sigma }^+`$) are compared with the low-energy scattering data in Fig. 3.22. The results using the isospin basis (short-dashed line) are close to those using the basis of physical states (solid line) but the cusp associated to the opening of the $`\overline{K}^0n`$ channel appears in the wrong place due to the use of an average mass for all the members of an isospin multiplet. The effects of neglecting the $`\eta `$ channels (long-dashed line) are much more significant. Close to threshold, the $`\pi ^{}\mathrm{\Sigma }^+`$ cross section is reduced by almost a factor of 3 and the $`\pi ^+\mathrm{\Sigma }^{}`$ cross section is reduced by a factor 1.3 when the $`\eta `$ channels are included. This enhances the ratio $`\gamma `$ by a factor 2.2 and makes the agreement with the experimental value possible using only the lowest order chiral Lagrangian. It is also remarkable that the $`\pi ^+\mathrm{\Sigma }^{}`$ cross section, which is zero at lowest order with the chiral Lagrangians, turns out to be about three times bigger that the analogous, allowed one, $`\pi ^{}\mathrm{\Sigma }^+`$. The multiple scattering with coupled channels is responsible for this. The predictions of this model for $`KN`$ scattering in the strangeness $`S=1`$ sector at low energies are quite satisfactory. As shown in ref. , the phase shifts in the isospin channel $`I=1`$ are about 15% smaller than experiment. This result is qualitatively similar to the one obtained in , where it was also shown that allowing for a $`K^+p`$ shorter range parameter (larger cut off) the agreement with data improves. The predicted $`KN`$ scattering lengths for isospin $`I=0`$ and $`I=1`$ are $`a(S=1,I=0)=2.4\times 10^7`$ fm and $`a(S=1,I=1)=0.26`$ fm, which compare favorably with present experimental data, $`0.02\pm 0.04`$ fm $`(I=0)`$ and $`0.32\pm 0.02`$ fm $`(I=1)`$ . Note that the scattering length in $`I=0`$ is zero at lowest order ($`T=V`$) and becomes finite, although negligibly small, as a consequence of the coupling to other channels when working in the particle basis. We have seen that the use of only one cut off parameter and the input of the lowest-order Lagrangian reproduces the low energy data in the $`S=1`$ sector as satisfactorily as the model of ref. , where the $`\eta `$ channels were omitted and the next-to-leading order terms of the chiral Lagrangian included. This is due to the fact that at low energies the $`\eta `$ meson loops only contribute to the real part of the amplitudes and this can effectively be taken into account by means of parameters of the second-order Lagrangian. Nevertheless, the values of the $`𝒪(p^2)`$ countertems of the meson-baryon Lagrangian given in ref. are affected by the resummation of the important contributions coming from the SU(3) channels with the $`\eta `$ meson and, hence, their actual values can be very different. The success in reproducing $`\overline{K}N`$ and $`KN`$ low energy scattering observables with the lowest-order Lagrangian and one cut off does not mean that this procedure can be generalized to all meson-baryon sectors. The richness of information available for meson-nucleon scattering requires the use of higher order Lagrangians, as it was the case in meson-meson scattering when including all the different channels . In fact, it has turned out to be impossible to dynamically reproduce the S-wave $`N^{}(1535)`$ resonance and the low energy scattering data in the $`S=0`$ sector with the lowest order Lagrangian and only one cut off. However, the extension of the model of ref. to pion induced reactions in the $`S=0`$ sector ($`\pi ^{}p\eta \mathrm{\Lambda },K^0\mathrm{\Lambda },K^+\mathrm{\Sigma }^{},\pi ^+pK^+\mathrm{\Sigma }^+`$) produces the $`N^{}(1535)`$ resonance as a quasibound $`K\mathrm{\Sigma }K\mathrm{\Lambda }`$ state , using the same parameters of their next-to-leading order Lagrangian fitted to the low energy $`\overline{K}N`$ ($`S=1`$) data. Simultaneously, the $`\eta `$ and $`K`$ photoproduction processes in the $`S=0`$ channels were also studied ($`\gamma p\eta p,K^+\mathrm{\Lambda },K^+\mathrm{\Sigma }^0,K^0\mathrm{\Sigma }^+`$) and with a few more parameters a global reproduction of the strong and electromagnetic cross section was obtained. The method is being extended to higher partial waves, to gain access to higher energies, other resonances and polarization observables . Recently, the unitarization of the Heavy Baryon $`\chi PT`$ amplitudes at $`𝒪(p^3)`$ has regained interest. In ref. , this is done making use of the IAM method, giving rise to a reasonable account of the scattering data up to around 1.2 GeV, including the region of the $`\mathrm{\Delta }(1232)`$ resonance in the $`P_{33}`$ partial wave. In ref. it is argued that a new rearrangement of the Heavy Baryon $`\chi PT`$ series, in terms of which the IAM is once again applied, leads to much better results than making use of the more straightforward version of the IAM used in . This seems to be the case for the $`P_{33}`$ partial wave, although a convincing argumentation for this rearragement is lacking in ref. , particularly when considering other partial waves <sup>3</sup><sup>3</sup>3With respect to this point, J.A.O. acknowleges very fruitful and enlightening discussions with José Ramón Peláez.. On the other hand, in ref. the unitarization of the elastic $`\pi N`$ scattering is accomplished using an adapted version of the method described in section 3.2.1 for the meson-baryon sector, in a fully relativistic way. Explicit resonance fields are included in ref. and a matching with the $`𝒪(p^3)`$ Heavy Baryon $`\chi PT`$ $`\pi N`$ amplitudes is given. The data are reproduced up to 1.3 GeV, where new channels would have to be introduced. The field is at a stage where rapid progress is being done and a clearer and broader picture of the role of chiral dynamics in meson-baryon scattering can be expected in the near future. ## Chapter 4 Final state interactions in meson pairs In this chapter we present how to use the previous meson-meson strong amplitudes to calculate processes with mesons in the final state. In many of them, the final state interactions between these mesons are crucial and give rise to corrections of even orders of magnitude in describing the physics involved. ### 4.1 The $`𝜸𝜸\mathbf{}`$ meson-meson reaction The $`\gamma \gamma `$ meson-meson reaction provides interesting information concerning the structure of hadrons, their spectroscopy and the meson-meson interactions, given the sensitivity of the reaction to the hadronic final state interactions (FSI) . In this sense, the study of these processes constitutes a very interesting test of consistency of the approaches for the scalar sector. In ref. , a unified theoretical description of the reactions $`\gamma \gamma \pi ^+\pi ^{}`$, $`\pi ^0\pi ^0`$, $`K^+K^{}`$, $`K^0\overline{K}^0`$, $`\pi ^0\eta `$ up to about $`\sqrt{s}=1.4`$ GeV was presented for the first time. The agreement with the experimental data was very good as can be seen in Fig. 4.4. For calculating the above processes one needs to correct for FSI the tree level amplitudes coming from Born terms, Fig. 4.1, in the case of the charged channels, and also from the exchange of vector and axial resonances in the crossed channels , Fig. 4.2. #### 4.1.1 FSI: S-wave In ref. the one loop corrections of the tree level amplitudes is first considered and then this result is extended to the string of loops represented in Fig. 4.3. The one loop contribution generated from the Born terms with intermediate charged mesons can be directly taken from the $`\chi PT`$ calculations of the $`\gamma \gamma \pi ^0\pi ^0`$ amplitude at $`𝒪(p^4)`$. The important point is that the $`𝒪(p^2)`$ $`\chi PT`$ amplitude connecting the charge particles with the $`\pi ^0\pi ^0`$ factorizes on shell outside the loop. One can schematically represent this situation by: $$\underset{a}{}L(s)_aT_{ab}^{(2)}(s)$$ (4.1) where the subindex $`a`$ represents the pair of intermediate charged mesons, $`b`$ the final ones and $`T^{(2)}`$ the on shell $`𝒪(p^2)`$ amplitude. The contribution of ref. beyond this first loop is to include all meson loops (Fig. 4.3) generated by the coupled channel Bethe-Salpeter equations of section 3.3.1, ref. . We also saw there that the on shell $`𝒪(p^2)`$ $`\chi PT`$ amplitudes factorize outside the loop integrals. Thus, the immediate consequence of introducing these loops is to substitute the on shell $`𝒪(p^2)`$ $`\pi \pi `$ amplitude in eq. (4.1), by the on shell meson-meson amplitude, $`T_{ab}(s)`$, evaluated in ref. . A similar procedure can be done to account for the FSI in the case of the tree level diagrams with the exchange of a resonance (vector or axial). As explained in ref. one can justify the accuracy of factorizing the strong amplitude for the loops with crossed exchange of resonances, since this result is correct for $`M_R^2\mathrm{}`$. Because we are dealing with real photons the intermediate axial or vector mesons are always off shell and the large mass limit is a sensible approximation. The errors were estimated to be below the level of $`5\%`$ for $`M_R`$ about 800 MeV. #### 4.1.2 D-wave contribution For the $`(2,2)`$ component we take the results of ref. , obtained using dispersion relations $$t_{BC}^{(2,2)}=\left[\frac{2}{3}\chi _{22}^{T=0}e^{i\delta _{20}}+\frac{1}{3}\chi _{22}^{T=2}e^{i\delta _{22}}\right]t_B^{(2,2)}$$ (4.2) where the functions $`\chi _{ij}(s)`$ are just first order polynomials in the $`s`$ variable. For the $`\gamma \gamma K^+K^{}`$ reaction the non resonant D-wave contribution is not needed because one is close to the $`K\overline{K}`$ threshold and furthermore the functions $`\chi _{ij}`$ are nearly zero close to the mass of the $`f_2`$ and $`a_2`$ resonances, which are also in the energy region we are considering. The resonance contribution in the $`D`$wave coming from the $`f_2(1270)`$ and $`a_2(1320)`$ resonances is parametrized in the standard way of a Breit-Wigner as done in ref. . The parameters of these resonances are completely compatible with the ones coming from the Particle Data Group . Once the FSI for the S- and D-waves have been taken into account, which completely dominate the $`\gamma \gamma `$meson-meson reactions up to the energies considered , one can compare with several experimental data. #### 4.1.3 Total and differential cross sections The experimental data correspond to total and differential cross sections. As can be seen in Fig. 4.4, the agreement is very good in all the channels considered. It is worth mentioning that the results presented are not a fit, since the parameters of the axial, vector and tensor resonances were taken from the literature. It is also worth remarking that in the figure corresponding to the $`\gamma \gamma K^+K^{}`$ reaction, the Born term, indicated by the long-dashed line, reduces to the short-dashed line when taking into account the FSI. This implies a large reduction of this Born term thanks to which a good reproduction of the data is obtained, hence solving a long standing problem . #### 4.1.4 Partial decay widths to two photons of the $`𝒇_\mathrm{𝟎}\mathbf{(}\mathrm{𝟗𝟖𝟎}\mathbf{)}`$ and $`𝒂_\mathrm{𝟎}\mathbf{(}\mathrm{𝟗𝟖𝟎}\mathbf{)}`$ The same procedure as in section 3.1.2 is followed in ref. in order to calculate the partial decay widths of the $`f_0(980)`$ and $`a_0(980)`$ in terms of the strong and photo-production amplitudes. From the amplitudes with isospin $`I=1`$ and 0 , one considers the terms which involve the strong $`M\overline{M}M\overline{M}`$ amplitude. Then, one isolates the part of the $`\gamma \gamma M\overline{M}`$ process which proceeds via the resonances $`a_0`$ and $`f_0`$ respectively. In the vicinity of the resonance the amplitude proceeds as $`M\overline{M}RM\overline{M}`$. Hence, eliminating the $`RM\overline{M}`$ part of the amplitude plus the $`R`$ propagator and removing the proper isospin Clebsch Gordan coefficients for the final states (1 for $`\pi ^0\eta `$ and $`1/\sqrt{2}`$ for $`K^+K^{}`$), one obtains the coupling of the previous resonances to the $`\gamma \gamma `$ channel. The results are: $$\begin{array}{ccc}\mathrm{\Gamma }_{a_0}^{\gamma \gamma }=0.78\mathrm{KeV};\hfill & \mathrm{\Gamma }_{a_0}^{\gamma \gamma }\frac{\mathrm{\Gamma }_{a_0}^{\eta \pi }}{\mathrm{\Gamma }_{a_0}^{tot}}=0.49\mathrm{KeV};\hfill & \mathrm{\Gamma }_{f_0}^{\gamma \gamma }=0.20\mathrm{KeV}\hfill \end{array}$$ (4.3) The calculated width for the $`f_0(980)`$ is smaller than the average value of ($`0.56\pm 0.11`$) KeV reported in the PDG . In doing this average the PDG refers to the work by Morgan and Pennington where they quote a width of $`(0.63\pm 0.14)`$ KeV. However, in a recent work by Boglione and Pennington they quote the much smaller width ($`0.28_{0.13}^{+0.09}`$) KeV . When taking into account the errors, the former result and the one from ref. are compatible. The value given above for the second magnitude in eq. (4.3) is larger than the value of ($`0.28\pm 0.04\pm 0.1)`$ KeV given in PDG. However, this value comes from references where a background is introduced in order to fit the data. In the analysis we have discussed in this section , no background is included and hence, in a natural way, the strength of the $`a_0(980)`$ to two photons is increased. ##### Conclusions As important features of the approach of ref. , we can remark: 1) The resonance $`f_0(980)`$ shows up weakly in $`\gamma \gamma \pi ^0\pi ^0`$ and barely in $`\gamma \gamma \pi ^+\pi ^{}`$. 2) In order to explain the angular distributions of the $`\gamma \gamma \pi ^+\pi ^{}`$ reaction there is not need of the hypothetical $`f_0(1100)`$ broad resonance suggested in other works . This also solves the puzzle of why it did not show up in the $`\gamma \gamma \pi ^0\pi ^0`$ channel. Furthermore, such resonance does not appear in the theoretical work of ref. , while the $`f_0(980)`$ showed up clearly as a pole of the $`T`$ matrix in $`I=0`$. 3) The resonance $`a_0`$ shows up clearly in the $`\gamma \gamma \pi ^0\eta `$ channel and the experimental results are well reproduced without the need of an extra background from a hypothetical $`a_0(11001300)`$ resonance suggested in ref. . 4) One can explain the drastic reduction of the Born term in the $`\gamma \gamma K^+K^{}`$ reaction in terms of final state interaction of the $`K^+K^{}`$ system. ### 4.2 The $`\mathit{\varphi }\mathbf{}𝜸𝑲^\mathrm{𝟎}\overline{𝑲}^\mathrm{𝟎}`$, $`𝜸𝝅^\mathrm{𝟎}𝝅^\mathrm{𝟎}`$ and $`𝜸𝝅^\mathrm{𝟎}𝜼`$ decays We first discuss the decay of the $`\varphi `$ meson to $`\gamma K^0\overline{K}^0`$ following ref. . With the previous formalism fixed, we will consider the decay of the $`\varphi `$ to $`\gamma \pi ^0\pi ^0`$ and $`\gamma \pi ^0\eta `$ . #### 4.2.1 The $`\mathit{\varphi }\mathbf{}𝜸𝑲^\mathrm{𝟎}\overline{𝑲}^\mathrm{𝟎}`$ decay The study of the process $`\varphi \gamma K^0\overline{K^0}`$ is an interesting subject since it provides a background to the reaction $`\varphi K^0\overline{K^0}`$. This latter process has been proposed as a way to study CP violating decays to measure the small ratio $`ϵ^{}/ϵ`$ , but, since this implies seeking for very small effects, a BR($`\varphi \gamma K^0\overline{K^0}`$ )$`10^6`$ will limit the scope of these perspectives. There are several calculations of this quantity . In ref. it is estimated for a non resonant decay process without including the $`f_0`$ and $`a_0`$ resonances. The issue was revisited in ref. . The approach introduced in ref. to treat the $`I=0,1`$ scalar meson-meson sector was the one used in ref. . The formalism, reviewed in section 3.3.1, will allow us to consider simultaneously the influence of the $`f_0(980)`$ and the $`a_0(980)`$ resonances, as well as their mutual interference, in a way that takes into account the energy dependence of their widths and coupling constants to the $`K\overline{K}`$ system. Furthermore, other possible contributions, non resonant, are also considered. The final state interactions will be taken into account following the way of ref. and discussed in the former section. As in previous works , in ref. the process $`\varphi \gamma K^0\overline{K^0}`$ is calculated through an intermediate $`K^+K^{}`$ loop which couples strongly to the $`\varphi `$ and the scalar resonances, see Fig. 4.5. For calculating the contribution of these loop diagrams one uses the minimal coupling to make the interaction between the $`\varphi `$ and the $`K^+K^{}`$ mesons gauge invariant, then we have $$H_{int}=(eA_\mu +g_\varphi \varphi _\mu )i(^\mu K^+K^{}K^+^\mu K^{})2eg_\varphi A^\mu \varphi _\mu K^+K^{},$$ (4.4) where $`g_\varphi `$ is the coupling constant between the $`\varphi `$ and the $`K^+K^{}`$ system <sup>1</sup><sup>1</sup>1There is an extra contribution that does not come from minimal coupling and is given by the term proportional to $`F_V`$ in eq. (2.24). However, this term only originates a correction to the coupling constant proportional to the momentum of the photon (see section 4.2.2) which, because we are very close to the threshold of the $`\varphi `$, is very tiny.. An essential ingredient to evaluate the loop in Fig. 4.5 is the strong amplitude connecting $`K^+K^{}`$ with $`K^0\overline{K^0}`$. As we said before, the amplitude calculated in ref. is the one used in ref. . This implies the sum of an infinite series of diagrams which is represented in Fig. 4.6 for the diagram of Fig. 4.5a, and the analogue sums corresponding to Figs. 4.5b,c,d. This way of taking into account the S-wave final state interactions is the same as shown above in Fig. 4.3 for the $`\gamma \gamma `$meson-meson reactions. This series gives rise to the needed corrections due to final state interactions and in fact, from the vertex connecting the $`K^+K^{}`$ with the $`K^0\overline{K^0}`$, this series is the same one as that in ref. which gives rise to the S-wave strong amplitude $`K^+K^{}K^0\overline{K^0}`$, see eq. (3.51). In this approach the vertex between the loops correspond to the on shell lowest order $`\chi PT`$ amplitude . Note that an analogous series before the loop with the emission of the photon is absorbed in the infinite series of diagrams contained in the $`\varphi `$ resonance propagator and its effective coupling. First of all, let us see that the strong amplitude connecting $`K^+K^{}`$ with $`K^0\overline{K^0}`$ calculated in the way shown in Fig. 4.6, ref. , must factorize out of the integral. In order to see this, following ref. , consider the diagrams in Fig. 4.5 but with the $`𝒪(p^2)`$ $`\chi `$PT amplitude connecting the kaons. This amplitude is given by $$<K^0\overline{K^0}|t|K^+K^{}>=\frac{1}{2}\left[t_{I=0}t_{I=1}\right]=\frac{1}{4f^2}\left[s+\frac{4m_K^2_ip_i^2}{3}\right],$$ (4.5) where $`f`$ is the pion decay constant, $`f93`$ MeV, $`I`$ refers to the isospin channel of the amplitude and the subindex $`i`$ runs from $`1`$ to $`4`$ and refers to any of the four kaons involved in the strong interaction. If the particle is on shell then $`p_i^2=m_K^2`$. In the present case $`p_{K^0}^2=p_{\overline{K^0}}^2=m_K^2`$ so one has $$\frac{1}{4f^2}\left[s+\frac{(m_K^2p_{K^+}^2)+(m_K^2p_K^{}^2)}{3}\right].$$ (4.6) The important point for the sequel is that the off shell part, which should be kept inside the loop integration, will not contribute. First of all let us note that, due to gauge invariance, the physical amplitude for $`\varphi \gamma K^0\overline{K^0}`$ has the form $$M(\varphi (p)\gamma (q)K^0\overline{K^0})=[g^{\mu \nu }(pq)p^\mu q^\nu ]ϵ_\mu ^\gamma ϵ_\nu ^\varphi H(pq,Q^2,qQ),$$ (4.7) where $`ϵ_\mu ^\gamma `$ and $`ϵ_\nu ^\varphi `$ are the polarization vectors of the photon and the $`\varphi `$ meson, $`Q=p_{K^0}+p_{\overline{K^0}}`$ and $`H`$ is an arbitrary scalar function. In the calculation of this loop contribution the problem is the presence of divergences in the loops represented in Fig. 4.5. Following refs. we will take into account the contribution of $`p^\mu q^\nu `$ of Figs. 4.5a,b, since Figs. 4.5c,d do not give such type of terms. Then, by gauge invariance, see formula (4.7), the coefficient for $`(pq)g^{\mu \nu }`$ is also fixed. In fact, as shown in ref , the $`p^\mu q^\nu `$ contribution will be finite since the off shell part of the strong amplitudes does not contribute, as we argue below, and then one is in the same situation than in the latter references. On the other hand, depending on the renormalization scheme chosen, additional tadpole like terms can appear . However, they do not contribute to the $`p^\mu q^\nu `$ structure and hence can be ignored. If we take the diagrams of Figs. 4.5a,b, which give the same contribution, we find the following amplitude for the sum of both diagrams: $`M^{}`$ $`=`$ $`ϵ_\mu ^\gamma ϵ_\nu ^\varphi {\displaystyle \frac{2eg_\varphi }{i}}{\displaystyle \frac{d^4k}{(2\pi )^4}\frac{(2k_\nu p_\nu )(2k_\mu q_\mu )}{(k^2m_K^2+iϵ)((kq)^2m_K^2+iϵ)((kp)^2m_K^2+iϵ)}}`$ (4.8) $`\times `$ $`{\displaystyle \frac{(1)}{4f^2}}\left[Q^2+{\displaystyle \frac{(m_K^2p_{K^+}^2)+(m_K^2p_K^{}^2)}{3}}\right].`$ The momentum for each particle in the loop is indicated in Fig. 4.5a and so one has $`p_{K^+}=kq`$, $`p_K^{}=kp`$. Concentrating in the off shell part of the strong amplitude, one has the following integral $`{\displaystyle \frac{d^4k}{(2\pi )^4}\frac{(2k_\nu p_\nu )(2k_\mu q_\mu )}{(k^2m_K^2+iϵ)((kq)^2m_K^2+iϵ)((kp)^2m_K^2+iϵ)}}`$ $`\times [(kq)^2m_K^2+(kp)^2m_K^2]=`$ (4.9) $`{\displaystyle \frac{d^4k}{(2\pi )^4}\frac{(2k_\nu p_\nu )(2k_\mu q_\mu )}{(k^2m_K^2+iϵ)((kp)^2m_K^2+iϵ)}}+{\displaystyle \frac{d^4k}{(2\pi )^4}\frac{(2k_\nu p_\nu )(2k_\mu q_\mu )}{(k^2m_K^2+iϵ)((kq)^2m_K^2+iϵ)}}`$ Taking into account that $$ϵ_\mu ^\varphi p^\mu =0;ϵ_\nu ^\gamma q^\nu =0(\text{Feynman gauge})$$ (4.10) then one only has $$\frac{d^4k}{(2\pi )^4}\frac{4k_\mu k_\nu }{(k^2m_K^2+iϵ)((kp)^2m_K^2+iϵ)}+\frac{d^4k}{(2\pi )^4}\frac{4k_\mu k_\nu }{(k^2m_K^2+iϵ)((kq)^2m_K^2+iϵ)}$$ (4.11) The above integrals do not give contribution to $`q^\mu p^\nu `$ since in each integral there is only one of the two vectors $`q`$ or $`p`$. In this way we see that the strong amplitude $`𝒪(p^2)`$ factorizes out on shell in (4.7). Note that the important point in the former argumentation is the form of the off shell part of the S-wave strong amplitude at $`𝒪(p^2)`$ and this is common to any other S-wave meson-meson amplitude at this order, as one can see in ref. . Next we consider the sum of all the infinite series represented in Fig. 4.6. In ref. this was accomplished by noting that at the one loop level the strong $`𝒪(p^2)`$ $`\chi PT`$ amplitude factorizes on shell, as we have seen. Then, one can apply the same technique as for the $`\gamma \gamma `$meson-meson and substitute the $`𝒪(p^2)`$ amplitude by the full one calculated in ref. . Then to all orders in the approach of ref. one has the amplitude $$t_S=\frac{1}{2}\left[t_{I=0}t_{I=1}\right].$$ (4.12) Note that the amplitude obtained in ref. contains also the resonances $`f_0(980)`$ and $`a_0(980)`$ which are generated dynamically. The final expression for the amplitude $`\varphi (p)\gamma (q)K^0\overline{K^0}`$, as given in ref. , is then $$M=ϵ_\mu ^\gamma ϵ_\nu ^\varphi \frac{2eg_\varphi }{i}t_S\frac{d^4k}{(2\pi )^4}\frac{(2k_\nu p_\nu )(2k_\mu q_\mu )}{(k^2m_K^2+iϵ)((kq)^2m_K^2+iϵ)((kp)^2m_K^2+iϵ)}$$ (4.13) This integral has been evaluated in ref. using dimensional regularization and confirmed in ref. , with the result $$M=\frac{eg_\varphi }{2\pi ^2im_K^2}I(a,b)[(pq)(ϵ_\gamma ϵ_\varphi )(pϵ_\gamma )(qϵ_\varphi )]t_S,$$ (4.14) with $`a=M_\varphi ^2/m_K^2`$, $`b=Q^2/m_K^2`$ and $$I(a,b)=\frac{1}{2(ab)}\frac{2}{(ab)^2}\left[f\left(\frac{1}{b}\right)f\left(\frac{1}{a}\right)\right]+\frac{a}{(ab)^2}\left[g\left(\frac{1}{b}\right)g\left(\frac{1}{a}\right)\right],$$ (4.15) where $`f(x)`$ $`=`$ $`\{\begin{array}{cc}\mathrm{arcsin}^2\left(\frac{1}{2\sqrt{x}}\right)& x>\frac{1}{4}\hfill \\ \frac{1}{4}\left[\mathrm{ln}\left(\frac{\eta _+}{\eta _{}}\right)i\pi \right]^2& x<\frac{1}{4}\hfill \end{array}`$ (4.18) $`g(x)`$ $`=`$ $`\{\begin{array}{cc}(4x1)^{\frac{1}{2}}\mathrm{arcsin}\left(\frac{1}{2\sqrt{x}}\right)& x>\frac{1}{4}\hfill \\ \frac{1}{2}(14x)^{\frac{1}{2}}\left[\mathrm{ln}\left(\frac{\eta _+}{\eta _{}}\right)i\pi \right]& x<\frac{1}{4}\hfill \end{array}`$ (4.21) $`\eta _\pm `$ $`=`$ $`{\displaystyle \frac{1}{2x}}\left(1\pm (14x)^{\frac{1}{2}}\right)`$ After summing over the final polarizations of the photon, averaging over the ones of the $`\varphi `$ and taking into account the phase space for three particles one obtains $$\mathrm{\Gamma }(\varphi \gamma K^0\overline{K^0})=\frac{dm_{12}^2dQ^2}{(2\pi )^3192M_\varphi ^3}\left|eg_\varphi \frac{I(a,b)}{2\pi ^2m_K^2}\right|^2(M_\varphi ^2Q^2)^2|t_S|^2$$ (4.22) where $`m_{12}^2=(q+p_{K^0})^2`$. Taking $`{\displaystyle \frac{g_\varphi ^2}{4\pi }}=1.66`$ from its width to $`K^+K^{}`$, $`M_\varphi =1019.41`$ MeV, $`\mathrm{\Gamma }(\varphi )=4.43`$ MeV, BR( $`\varphi K^0\overline{K^0})=0.34`$ and using the mass of the $`K^0`$ for the phase space considerations, ref. , one gets $$\begin{array}{cc}\hfill \mathrm{\Gamma }(\varphi \gamma K^0\overline{K^0})=& 2.22\times 10^7\mathrm{MeV}\hfill \\ \hfill BR(\varphi \gamma K^0\overline{K^0})=& 0.50\times 10^7\hfill \\ \hfill \frac{\mathrm{\Gamma }(\varphi \gamma K^0\overline{K^0})}{\mathrm{\Gamma }(\varphi K^0\overline{K^0})}=& 1.47\times 10^7\hfill \end{array}$$ (4.23) The uncertainties coming from the range of the possible values for the cut off give a relative error around $`20\%`$. Taking only into account the $`I=0`$ contribution $$\begin{array}{cc}\hfill \mathrm{\Gamma }(\varphi \gamma K^0\overline{K^0})=& 8.43\times 10^7\mathrm{MeV}\hfill \\ \hfill BR(\varphi \gamma K^0\overline{K^0})=& 1.90\times 10^7\hfill \\ \hfill \frac{\mathrm{\Gamma }(\varphi \gamma K^0\overline{K^0})}{\mathrm{\Gamma }(\varphi K^0\overline{K^0})}=& 5.58\times 10^7\hfill \end{array}$$ (4.24) and with only the $`I=1`$ $$\begin{array}{cc}\hfill \mathrm{\Gamma }(\varphi \gamma K^0\overline{K^0})=& 2.03\times 10^7\mathrm{MeV}\hfill \\ \hfill BR(\varphi \gamma K^0\overline{K^0})=& 4.58\times 10^8\hfill \\ \hfill \frac{\mathrm{\Gamma }(\varphi \gamma K^0\overline{K^0})}{\mathrm{\Gamma }(\varphi K^0\overline{K^0})}=& 1.35\times 10^7\hfill \end{array}$$ (4.25) We see that the process is dominated by the $`I=0`$ contribution and that the interference between both isospin channels is destructive. From the former results one concludes that the $`\varphi \gamma K^0\overline{K^0}`$ background will not be too significant for the purpose of testing CP violating decays from the $`\varphi K^0\overline{K^0}`$ process at DA$`\mathrm{\Phi }`$NE in the lines of what was expected in ref. . All these calculations have been done in a way that both the resonant and non-resonant contributions are considered at the same time and taking into account also the different isospin channels. #### 4.2.2 The $`\mathit{\varphi }\mathbf{}𝝅^\mathrm{𝟎}𝝅^\mathrm{𝟎}𝜸`$ and $`𝝅^\mathrm{𝟎}𝜼𝜸`$ decays Radiative $`\varphi `$ decay into neutral mesons has been a subject of interest often advocated as a source of information on the nature of the scalar meson resonances. Calculations in the line of the former section have been done using the amplitude for the $`K\overline{K}\pi ^0\pi ^0`$ amplitude from $`\chi PT`$. Other calculations have concentrated on the possibility of using the reactions to decide the nature of the $`f_0`$ resonance between several models like a $`K\overline{K}`$ molecule, a $`q\overline{q}`$ state or a $`q\overline{q}q\overline{q}`$ structure . The laboratories of Frascati and Novosibirsk have been actively pursuing research on this topic and very recently some novel results have been reported by two Novosibirsk groups . In ref. calculations along the lines reported in the previous section have been conducted following however the tensor approach to the vector mesons of ref. reported in section 2.4. This approach had been previously used for the $`\rho \pi ^+\pi ^{}\gamma `$ decay in the absence of final state interaction in ref. . The novelties with respect to the approach of the former section can be summarized in the basic couplings involved in Fig. 4.5 given by $`t_{\varphi K^+K^{}}`$ $`=`$ $`{\displaystyle \frac{G_VM_\varphi }{\sqrt{2}f^2}}(p_\mu p_\mu ^{})ϵ^\mu (\varphi )`$ $`t_{\varphi \gamma K^+K^{}}`$ $`=`$ $`\sqrt{2}e{\displaystyle \frac{G_VM_\varphi }{f^2}}ϵ_\nu (\varphi )ϵ^\nu (\gamma )`$ (4.26) $`{\displaystyle \frac{\sqrt{2}e}{M_\varphi f^2}}\left({\displaystyle \frac{F_V}{2}}G_V\right)P_\mu ϵ_\nu (\varphi )[k^\mu ϵ^\nu (\gamma )k^\nu ϵ^\mu (\gamma )]`$ with $`p_\mu `$, $`p_\mu ^{}`$ the $`K^+,K^{}`$ momenta, $`P_\mu `$, $`k_\mu `$ the $`\varphi `$ and photon momenta and $`f`$ the pion decay constant. The couplings of eq. (4.26) are easily induced from the Lagrangian of eq. (2.24). The $`\varphi `$ meson is introduced in the scheme by means of a singlet, $`\omega _1`$, going from SU(3) to U(3) through the substitution $`V_{\mu \nu }V_{\mu \nu }+I_3\frac{\omega _{1,\mu \nu }}{\sqrt{3}}`$, with $`I_3`$ the $`3\times 3`$ diagonal matrix. Then, assuming ideal mixing for the $`\varphi `$ and $`\omega `$ mesons $`\sqrt{{\displaystyle \frac{2}{3}}}\omega _1+{\displaystyle \frac{1}{\sqrt{3}}}\omega _8`$ $``$ $`\omega `$ $`{\displaystyle \frac{1}{\sqrt{3}}}\omega _1{\displaystyle \frac{2}{\sqrt{6}}}\omega _8`$ $``$ $`\varphi `$ (4.27) one obtains the required vertices after substituting in eq. (2.24) $`V_{\mu \nu }`$ by $`\stackrel{~}{V}_{\mu \nu }`$, given by $$\stackrel{~}{V}_{\mu \nu }\left(\begin{array}{ccc}\frac{1}{\sqrt{2}}\rho _{\mu \nu }^0+\frac{1}{\sqrt{2}}\omega _{\mu \nu }& \rho _{\mu \nu }^+& K_{\mu \nu }^+\\ \rho _{\mu \nu }^{}& \frac{1}{\sqrt{2}}\rho _{\mu \nu }^0+\frac{1}{\sqrt{2}}\omega _{\mu \nu }& K_{\mu \nu }^0\\ K_{\mu \nu }^{}& \overline{K}_{\mu \nu }^0& \varphi _{\mu \nu }\end{array}\right)$$ (4.28) As we can see, with respect to the approach of the former section the contact term contains an extra part, the term proportional to $`{\displaystyle \frac{F_V}{2}}G_V`$, which is gauge invariant by itself. The first part of the contact term proportional to $`G_V`$ is not gauge invariant and this requires the addition of the diagrams (a) and (b) of Fig. 4.5 to have a gauge invariant set, as discussed in the former section. This means that in addition to the terms discussed there one gets now an additional term proportional to $`({\displaystyle \frac{F_V}{2}}G_V)k`$, where $`k`$ is the photon momentum in the $`\varphi `$ rest frame. This term appears only with the structure of diagram (c). This has as a consequence that in the treatment of the final state interaction the first loop contains only the two meson propagator and hence is the same function $`g(M_I)`$ defined in eq. (3.11). This technical detail plus the use of the $`K^+K^{}\pi ^0\pi ^0`$ and $`\pi ^0\eta `$ amplitudes instead of the $`K^+K^{}K^0\overline{K^0}`$ in the former section are the basic modifications needed in this work. The values $`G_V=55`$ MeV and $`F_V=165`$ MeV which are suited to the $`\varphi K^+K^{}`$ and $`\varphi e^+e^{}`$ decay widths respectively have been used in ref. . In Fig. 4.7 we show the distribution $`dB/dM_I`$ for $`\varphi \pi ^0\pi ^0\gamma `$ which allows one to see the $`\varphi f_0\gamma `$ contribution since the $`f_0`$ is the important scalar resonance appearing in the $`K^+K^{}\pi ^+\pi ^{}`$ amplitude . The solid curve shows the results with $`F_VG_V>0`$, the sign predicted by vector meson dominance . The intermediate dotted line corresponds to taking $`G_V=67`$ MeV and $`F_V=154`$ MeV , which are the values of the parameters for $`\rho `$ decay, usually assumed as standard values. The two curves give us an idea of the theoretical uncertainties. The upper dashed curve is obtained considering $`F_VG_V<0`$. These results are compared in the figure with the recent ones of the Novosibirsk experiment . We can see that the shape of the spectrum is relatively well reproduced considering statistical and systematic errors (the latter ones not shown in the figure). The results considering $`F_VG_V<0`$ are in complete disagreement with the data. The finite total branching ratio for the $`\varphi \pi ^0\pi ^0\gamma `$ is $`0.8\times 10^4`$ . This latter number is slightly smaller than the result given in ref. , $`(1.14\pm 0.10\pm 0.12)\times 10^4`$, where the first error is statistical and the second one systematic. The result given in ref. is $`(1.08\pm 0.17\pm 0.09)\times 10^4`$, compatible with the prediction. If the values for $`F_V`$, $`G_V`$ of the $`\rho `$ decay are used one obtains $`1.7\times 10^4`$ . The branching ratio obtained for the case $`\varphi \pi ^0\eta \gamma `$ is $`0.87\times 10^4`$. The results obtained at Novosibirsk are $`(0.83\pm 0.23)\times 10^4`$ and $`(0.90\pm 0.24\pm 0.10)\times 10^4`$. If the values of the $`\rho `$ decay are used one obtains $`1.6\times 10^4`$ . The spectrum, not shown, is dominated by the $`a_0`$ contribution. The results of this section for these radiative decays are a striking success of the chiral unitary approach reported here . The evidence given by Fig. 4.7 in favour of the sign $`F_VG_V>0`$ is much stronger than the one given in ref. from the tail of the $`\rho \pi ^+\pi ^{}\gamma `$ distribution. ### 4.3 Vector and scalar pion form factors The scalar and vector form factors of the pion are defined respectively as $$\pi ^a(p^{})\pi ^b(p)\text{out}\left|\widehat{m}(\overline{u}u+\overline{d}d)\right|0=\delta ^{ab}m_\pi ^2\mathrm{\Gamma }(s)$$ (4.29) and $$\pi ^i(p^{})\pi ^l(p)\text{out}\left|\overline{q}\gamma _\mu \left(\frac{\tau ^k}{2}\right)q\right|0=iϵ^{ikl}(p^{}p)_\mu F_V(s)$$ (4.30) with $`\widehat{m}=(m_u+m_d)/2`$ and $`ϵ^{ijk}`$ the total antisymmetric tensor with three indices. Assuming elastic unitarity (valid up to the $`K\overline{K}`$ threshold and neglecting multipion states) and making use of the Watson final state theorem the phase of $`\mathrm{\Gamma }(s)`$ and $`F_V(s)`$ is fixed to be the one of the corresponding partial wave strong amplitude: $`\text{Im}\mathrm{\Gamma }(s+iϵ)`$ $`=`$ $`\mathrm{tan}\delta _0^0\text{Re}\mathrm{\Gamma }(s)`$ $`\text{Im}F_V(s+iϵ)`$ $`=`$ $`\mathrm{tan}\delta _1^1\text{Re}F_V(s)`$ (4.31) The solution of (4.31) is well known and corresponds to the Omnès type : $`\mathrm{\Gamma }(s)`$ $`=`$ $`P_0(s)\mathrm{\Omega }_0(s)`$ $`F_V(s)`$ $`=`$ $`P_1(s)\mathrm{\Omega }_1(s)`$ (4.32) with $$\mathrm{\Omega }_i(s)=\mathrm{exp}\left\{\frac{s^n}{\pi }_{4m_\pi ^2}^{\mathrm{}}\frac{ds^{}}{s_{}^{}{}_{}{}^{n}}\frac{\delta _i^i(s^{})}{s^{}siϵ}\right\}$$ (4.33) In (4.32) $`P_0(s)`$ and $`P_1(s)`$ are polynomials of degree fixed by the number of subtractions done in $`\mathrm{ln}\{\mathrm{\Omega }_0(s)\}`$ and $`\mathrm{ln}\{\mathrm{\Omega }_1(s)\}`$ minus one, and the zeros of $`F_V`$ and $`\mathrm{\Gamma }`$. For $`n=1`$, $`P_i(s)=1`$. This follows from the normalization requirement that $`\mathrm{\Gamma }(0)=F_V(0)=1`$ and the absence of zeros for those quantities.<sup>2</sup><sup>2</sup>2In principle $`\mathrm{\Gamma }(0)1`$. However, since this is the leading $`\chi PT`$ result and we are at low energies, the difference with respect one is very small . In ref. the previous dispersion integrals, eq. (4.33), are evaluated making use of the phase shifts calculated in the same reference, see section 3.1.1. The resulting vector and scalar form factors are shown in Figs. 4.8 and 4.9 respectively. The Omnès solution assumes the phase of the form factor to be that of the scattering amplitude, and that is true exactly only until the first inelastic threshold. The first inelastic threshold is the $`4\pi `$ one. However, as it was already said, its influence, in a first approach, is negligible. The first important inelastic threshold is the $`K\overline{K}`$ one around 1 GeV. This is essential in $`I=L=0`$ but negligible in $`I=L=1`$. This inelastic threshold, as discussed above, has been included in the approach and it is mostly responsible for the appearance of the $`f_0(980)`$ resonance. In the case of the $`F_V(s)`$ the agreement with existing data is quite satisfactory, with the dominant role played by the $`\rho (770)`$ resonance. On the other hand, taking into account possible uncertainties coming from orders higher than $`p^4`$ in $`\chi PT`$, the result obtained for the vector form factor is similar to the one obtained in ref. using another phase shift expression. For the $`I=L=0`$ channel the most dramatic effect is the openning of the $`K\overline{K}`$ channel. In Fig. 4.9 the continuous line corresponds to the use of eq. (4.32) integrating up to infinity in the Omnès formula (4.33). On the other hand, the dashed-dotted line corresponds to integrating only up to the openning of the $`K\overline{K}`$ threshold. The differences between both options are tremendous making evident that a coupled channel approach is necessary in order to describe properly the scalar form factor for energies above 400–500 MeV. Finally, the dashed line represents the scalar form factor unitarizing only with pions to obtain the $`\delta _{00,\pi \pi }`$ phase shift, in the line of the works . ### 4.4 $`𝜸𝒑\mathbf{}𝒑`$ meson-meson An interesting example of final state interaction of two mesons appears in the photoproduction of pairs of mesons in the $`\gamma p`$ reaction when the pair is produced with an invariant mass close to the mass of one resonance. An example of it is given in ref. where the photoproduction of scalar meson is studied. The process is depicted in Fig. 4.10 where the leading tree level process appears in diagram (a) and is a contact term induced from minimal coupling from the meson-baryon Lagrangian of eq. (3.55). This term is given by $$V_{\pi (K)}^\gamma =C_{\pi (K)}\frac{e}{2f^2}\overline{u}(p^{})\gamma ^\mu u(p)ϵ_\mu ,$$ (4.34) where $`C_\pi `$ =1 and $`C_K`$=2 stand for the case of the production of a $`\pi ^+\pi ^{}`$ or a $`K^+K^{}`$ pair respectively. The final state interaction of the mesons is accounted for by means of the rest of the diagrams in the figure and, similarly to the case of the $`\varphi `$ radiative decay, the photon lines must also be coupled to the lines in the loops to guarantee gauge invariance. Missing in the figure are the tree level diagrams of Bremmsstrahlung where the photon is coupled to any of the external meson lines in diagram (a). This is justified in the case of meson production close to threshold (for this purpose a photon beam of energy 1.7 GeV is suggested in ref. ) or in the case of production of neutral meson pairs like $`\pi ^0\pi ^0`$ or $`\pi ^0\eta `$. At higher energies there can be more involved production mechanisms and also at lower energies there are background terms which are important particularly for the case of $`\pi ^+\pi ^{}`$ production . Yet the resonant production should show up as a bump on top of the background and allow one to study the resonances in a different setup, and in addition study the production of the resonances in nuclei which we will address below. The technique to include final state interaction here follows closely the steps described in the study of the $`\gamma \gamma `$ meson-meson reaction in ref. and of the $`\varphi `$ radiative decay in ref. , which have been reported above. There is only one difference since the square of the momentum transferred by the nucleon (which plays here the role of the $`\varphi `$ mass squared in the $`\varphi `$ decay) can be here negative. In this case one has to extrapolate analytically the $`I(a,b)`$ function of eq. (4.15) and detailed expressions are given in ref. . The results obtained for the invariant mass distribution of the two mesons are given in Fig. 4.11 for different pairs of mesons in the final state. The figure shows clear peaks for the production of the $`f_0`$ and $`a_0`$ resonances in the $`\pi ^+\pi ^{}`$, $`\pi ^0\pi ^0`$ and $`\pi ^0\eta `$ production respectively. In these cases the ratio of the resonance signal to background is found to be optimal for the $`\pi ^0\pi ^0`$ and $`\pi ^0\eta `$ cases. The figure also shows the cross section for $`K^+K^{}`$ production close to threshold which is appreciably renormalized by final state interaction with respect to the Born contribution. The $`K^0\overline{K^0}`$ production is found to be very small. It is interesting to note that the signal for the $`\pi \pi `$ production around the $`f_0`$ resonance shows up with a peak. This is a novelty with respect to the case in $`\pi \pi `$ scattering where the $`f_0`$ shows up with a minimum in the cross section. The different structure of the loop function in the first loop in the figures with respect to the $`g(s)`$ function of the plain two meson propagator is responsible for this different interference pattern with the background from the $`\sigma `$ pole contribution. This was also the case in the radiative $`\varphi `$ decay where the $`f_0`$ resonance also appeared as a peak in the invariant mass distribution. #### 4.4.1 Scalar meson production in nuclei It is also relevant to see that the previous reaction can be suited for the observation of the resonance modification in nuclei. The $`f_0`$ and $`a_0`$ couple strongly to $`K\overline{K}`$ but they are narrow because there is little phase space for the decay into that state. However, the kaons are apprecibly renormalized in nuclei and hence one expects that these resonances will also be appreciably modified inside a nuclear medium. The interaction of $`K`$, $`\overline{K}`$ with nuclei is a subject that has attracted much attention . Interesting developments have been done recently in $`K^{}N`$ scattering using the chiral Lagrangians , which have allowed to tackle the problem of the $`K`$, $`\overline{K}`$ nucleus interaction with some novel results . The issue is not yet settled since there are still important discrepancies between the different results. The first thing which we observe is that if one looks for a proton in the final state, one can also have the $`\gamma np\pi ^{}\eta (K^{}K^0)`$ reactions and approximately one would expect a cross section $$\frac{d\sigma }{dM_I}|_AZ\frac{d\sigma }{dM_I}(p)+N\frac{d\sigma }{dM_I}(n).$$ (4.35) The latter cross section can proceed through the meson channels $`K^{}K^0`$ and $`\pi ^{}\eta `$, both in $`I=1`$. These cross sections are found in ref. to be one order of magnitude smaller than those on the proton target. Hence, in nuclei we should expect a cross section roughly $`Z`$ times the one of the proton, unless the properties of the resonances $`a_0`$ and $`f_0`$ are drastically modified in the medium, which is, however, what one expects. As noted above, the relatively small width of the $`f_0`$ resonance is due to the small coupling to the $`\pi \pi `$ channel. The coupling of $`a_0`$ to $`\pi ^0\eta `$ is comparatively much larger. These resonances, however, couple very strongly to the $`K\overline{K}`$ system but the decay is largely inhibited because the $`K\overline{K}`$ threshold is above the resonance mass. Only the fact that the $`f_0`$ and $`a_0`$ resonances have already a width for $`\pi \pi `$ and $`\pi \eta `$ decay, respectively, allows the $`K\overline{K}`$ decay through the tail of the resonance distribution. If the $`K^{}`$ develops a large width on its own this enlarges considerably the phase space for $`K\overline{K}`$ decay and the $`a_0`$, $`f_0`$ width should become considerably larger. Given the interest that the modifications of meson resonances in nuclei, like the $`\sigma `$ , $`\rho `$ , etc., is raising, the study of the modifications of the $`f_0`$ and $`a_0`$ is bound to offer us some insight into the nature of these resonances, that has been so much debated, and into the chiral unitary approach to these resonances which we are discussing in this work. Preliminary results using the $`K,\overline{K}`$ self-energies in the medium discussed in section 6.2 are already available and indicate a large increase of the $`f_0`$ width in the medium. ## Chapter 5 Initial and final state interaction in meson-baryon reactions In section 3.3.2 we studied the meson-baryon interaction around the region of the $`\mathrm{\Lambda }(1405)`$ and $`N^{}(1535)`$ resonances. We saw there how the unitarization in coupled channels was essential to reproduce the scattering data. In this chapter we show some examples of physical reactions where the meson-baryon interaction appears in the initial or final states. We shall show how the proper consideration of the initial state interactions along the lines discussed in section 3.3.2 brings a natural solution to one problematic reaction, the $`K^{}p`$ radiative capture, where the ratio of $`\mathrm{\Lambda }`$ to $`\mathrm{\Sigma }^0`$ production is abnormally low. The coupled channel unitary techniques will be also applied to study reactions in which the resonances appearing in the meson-baryon interactions are now generated in the final state. In addition we shall also devise how the techniques can be used to evaluate static properties of those resonances. ### 5.1 $`𝑲^{\mathbf{}}`$ proton radiative capture: $`𝑲^{\mathbf{}}𝒑\mathbf{}𝜸𝚲\mathbf{,}𝜸𝚺^\mathrm{𝟎}`$ The near threshold $`K^{}p\gamma Y`$ reaction with $`Y=\mathrm{\Lambda },\mathrm{\Sigma }^0`$ has long attracted a lot of interest, mainly because of the possibility of using this reaction to resolve the debates over the structure of the $`\mathrm{\Lambda }(1405)`$ resonance. Most of the earlier theoretical investigations neglected the initial strong $`K^{}p`$ interactions. It was first demonstrated by Siegel and Saghai that the initial $`K^{}p`$ interactions can drastically change the predicted capture rates and thus can significantly alter the interpretation of the data. With the phenomenological separable potentials, they, however, needed an about $`3050\%`$ deviation of the coupling constants from the SU(3) values to obtain an accurate description of the data. This unsatisfactory situation was revised in ref. , where it was investigated whether the data could be well described by treating the initial $`K^{}p`$ interactions within the unitary coupled-channel chiral approach of refs. , which has been discussed in detail in section 3.3.2. In the present section we report on the results obtained in ref. and we will show that, indeed, a satisfactory agreement with the experiment can be obtained without the need of SU(3) breaking. A detailed derivation of the $`K^{}p\gamma Y`$ reaction can be found in ref. , where standard electromagnetig vertices are used and the initial state strong $`K^{}p`$ interaction is described through the coupled-channel Bethe-Salpeter equation. The resulting amplitude in the center of mass frame $`P=(\sqrt{s},\stackrel{}{0})`$ reads $`T_{\gamma Y,K^{}p}(q,k^{})`$ $`=`$ $`Q_{\gamma Y,K^{}p}(q,k^{})+[QGT]_{\gamma Y,K^{}p}(q,k^{})+\mathrm{\Delta }_{\gamma Y,K^{}p}(q,k^{}),`$ (5.1) where the first term $`Q_{\gamma Y,MB}(q,k^{})=ie_M[\stackrel{}{\sigma }\stackrel{}{ϵ}]{\displaystyle \frac{C_{Y,MB}}{2f}}\left(1{\displaystyle \frac{\omega _M(k^{})}{2q}}+{\displaystyle \frac{\mu _M^2}{4qk^{}}}\mathrm{ln}{\displaystyle \frac{\omega _M(k^{})+k^{}}{\omega _M(k^{})k^{}}}\right),`$ (5.2) with $`MBK^{}p`$, collects the Born terms for the elementary $`K^{}p\gamma Y`$ reaction displayed in Fig. 5.1. Note that the above expression contains only the contact (Fig. 5.1d) and the meson exchange (Fig. 5.1c) terms. In the heavy-baryon approximation one can show that the baryon pole term (Fig. 5.1a) contributes only to the meson-baryon P-wave states, while some S-wave contributions from the baryon-exchange term (Fig. 5.1b) also vanish at threshold ($`K^{}`$ capture at rest). The charge and mass of the meson $`M`$ are denoted, respectively, by $`e_M`$ and $`\mu _M`$. The SU(3) coupling constants $`C_{Y,MB}`$ for the $`MBY`$ transition are given by $`C_{Y,MB}=X_{Y,MB}(D+F)+Z_{Y,MB}(DF),`$ (5.3) where the values of the $`X`$ and $`Z`$ coefficients are easily evaluated from the chiral Lagrangians and can be found in ref. . The initial state strong interactions are present in the second term of Eq. (5.1) $`[QGT]_{\gamma Y,K^{}p}(q,k^{})`$ $`=`$ $`{\displaystyle \underset{MB}{}}{\displaystyle \frac{d\stackrel{}{k}}{(2\pi )^3}\frac{M_B}{E_B(\stackrel{}{k})}\frac{1}{2\omega _M(\stackrel{}{k})}\frac{Q_{\gamma Y,MB}(q,k)}{\sqrt{s}E_B(\stackrel{}{k})\omega _M(\stackrel{}{k})+iϵ}}`$ (5.4) $`\times `$ $`T_{MB,K^{}p}(\stackrel{}{k}_{MB},\stackrel{}{k}^{},\sqrt{s}),`$ where $`T_{MB,K^{}p}`$ is the amplitude for the strong $`K^{}pMB`$ transition, as well as in the third term $`\mathrm{\Delta }_{\gamma Y,K^{}p}(q,k^{})`$ $`=`$ $`{\displaystyle \underset{MB}{}}{\displaystyle \frac{d\stackrel{}{k}}{(2\pi )^3}\frac{M_B}{E_B(\stackrel{}{k})}\frac{1}{2\omega _M(\stackrel{}{k}\stackrel{}{q})}\frac{1}{\sqrt{s}E_B(\stackrel{}{k})q^0\omega _M(\stackrel{}{k}\stackrel{}{q})+iϵ}}`$ (5.5) $`\times `$ $`ie_M{\displaystyle \frac{C_{Y,MB}}{2f}}{\displaystyle \frac{2[\stackrel{}{k}\stackrel{}{ϵ}][\stackrel{}{\sigma }(\stackrel{}{k}\stackrel{}{q})]}{(q^0+\omega _M(\stackrel{}{q}\stackrel{}{k}))^2\omega _M^2(\stackrel{}{k})}}T_{MB,K^{}p}(\stackrel{}{k}_{MB},\stackrel{}{k}^{},\sqrt{s}),`$ which is related to the exact treatment of the meson propagator. It corresponds to the contribution of the pole of the meson propagating between the emitted photon and the final hyperon $`Y`$ in Fig. 5.1c and in the loop diagrams generated by the initial state interactions. Note that, in the above expressions, the allowed intermediate states are the charged particle channels ($`MB=K^{}p,\pi ^+\mathrm{\Sigma }^{},\pi ^{}\mathrm{\Sigma }^+`$ and $`K^+\mathrm{\Xi }`$). Moreover, the strong amplitude, $`T_{MB,K^{}p}`$, appears with the on shell momentum $`k_{MB}`$ and factors out of the integral because, as shown in the appendix of ref. , the off shell piece can be absorbed in the renormalization of the charge. The results for the branching ratios, defined by $`B_{K^{}p\gamma Y}={\displaystyle \frac{\sigma _{K^{}p\gamma Y}(\sqrt{s}_{\mathrm{th}})}{\sigma _{K^{}p\mathrm{all}}(\sqrt{s}_{\mathrm{th}})}},`$ (5.6) where $`Y=\mathrm{\Lambda },\mathrm{\Sigma }^0`$ and $`\sqrt{s}_{\mathrm{th}}\mu _K^{}+M_p`$, are compared to the experimental data in Table 5.1. Neglecting initial meson-baryon interactions (first row in Table 5.1) gives a very weak branching ratio for $`\gamma \mathrm{\Sigma }^0`$ production and the predicted ratio between the two production rates is an order of magnitude larger than the data. This is in agreement with the findings of Siegel and Saghai. When the strong coupled-channel effects are included (third row in Table 5.1) the ratio is close to the experimental value. The predicted branching ratio for the $`\gamma \mathrm{\Lambda }`$ production is about 50% larger than the experimental value, but it is within the experimental uncertainty for the $`\gamma \mathrm{\Sigma }^0`$ production. As shown in ref. , the enhancement of the $`\gamma \mathrm{\Sigma }^0`$ production is essentially coming from the coupling of the photon to intermediate $`\pi ^+\mathrm{\Sigma }^{}`$ and $`\pi ^{}\mathrm{\Sigma }^+`$ states. The exact treatment of the meson propagator in Fig. 5.1c leads to a contribution from a second meson pole, $`\mathrm{\Delta }_{\gamma Y,K^{}p}`$, which can change the $`\gamma \mathrm{\Sigma }^0`$ branching ratio by about 40% and brings the predicted ratio closer to the experimental value. The influence of the $`\eta `$ channels on the strong $`T_{MB,K^{}p}`$ amplitudes has, as for the case of the low energy $`K^{}p`$ scattering data , a significant effect here. Comparing the third and fourth rows in Table 5.1 one sees that the predicted branching ratio for $`\gamma \mathrm{\Lambda }`$ production is increased by about 60% if the $`\eta `$ channels are omitted in the calculation of the strong amplitudes. It is thus clear that including the $`\eta `$ channels is also crucial in using this reaction to test the chiral SU(3) symmetry. We note that the $`\eta `$ channels were omitted in the model of ref. and, at the same time, the couplings had to be substantially changed with respect to their SU(3) values in order to obtain a good fit to the data. In retrospective one can say that the deviation of the coupling constants from their SU(3) value is in fact trying to restore the breaking that was induced by the omission of the $`\eta `$ channels. Finally, we note that the strong meson-baryon-baryon vertex in each of the photoproduction amplitudes should in principle have a form factor because hadrons are composite particles. The results obtained with a monopole form factor with a cut off of $`\mathrm{\Lambda }_\pi =1`$ GeV, a value which is commonly accepted, are shown in the fifth row of Table 5.1 and agree roughly with the data within experimental errors, which are of the order of 20%. If one compares with the central values of the experimental branching ratios, the results are on the upper edge of the $`B_{K^{}p\gamma \mathrm{\Lambda }}`$ ratio while those for $`B_{K^{}p\gamma \mathrm{\Sigma }^0}`$ are on the lower edge. Looked at it in the context that the coupled channels and unitarization have reduced the ratio $`R`$ by a factor 14, differences of the order of 10–20% are not so significative. Note that all coupling constants are consistent with the chiral SU(3) symmetry and the model depends on only the cut off parameter, which was fixed in the study of $`S=1`$ meson-baryon reactions . In this approach, neither the meson-baryon nor the photoproduction mechanisms involve the explicit consideration of excited hyperon states since the $`\mathrm{\Lambda }(1405)`$ resonance, which plays a key role in these reactions, is generated dynamically, hence strengthening the interpretation of the $`\mathrm{\Lambda }(1405)`$ as a quasi-bound meson-baryon system with $`S=1`$, as already supported by the study of the strong interactions in . ### 5.2 Photoproduction of the $`𝚲\mathbf{(}\mathrm{𝟏𝟒𝟎𝟓}\mathbf{)}`$ on protons and nuclei As we saw in section 3.3.2, the $`\mathrm{\Lambda }(1405)`$ resonance is produced dynamically by using the Bethe-Salpeter equation and the lowest order Lagrangian for meson-baryon interaction in S-wave suggesting that this resonance is like a quasibound meson-baryon state rather than a genuine $`3q`$ state. Further tests on the nature of the resonance can be done by studying different production processes. One of them was studied in ref. by means of photoproduction on the proton, i.e. $$\gamma pK^+\mathrm{\Lambda }(1405).$$ (5.7) The study of the reaction in nuclei is also of much interest since different studies predict sizeable changes of the resonance in nuclei which would have also repercussions on the properties of $`\overline{K}`$ inside a nuclear medium. This is a hot topic since it relates to the possibility of having kaon condensation in stars and to the puzzle of the strong $`\overline{K}`$ attraction needed to explain the $`K^{}`$ atoms which seems to violate the low density theorem. In ref. a study was done along lines similar to those exposed in section 4.3.1, however, in this case a meson and a baryon combine through final state interaction to give the $`\mathrm{\Lambda }(1405)`$. Diagrammatically the mechanism for the production is depicted in Fig. 5.2. As we can see there, a $`K^+`$ is produced together with another meson and a baryon which combine through final state interaction with the coupled channels in the $`S=1`$ sector and give rise to the $`\mathrm{\Lambda }(1405)`$ resonance. Once again one needs the vertex with two mesons a photon and a baryon line which is obtained via minimal coupling from eq. (3.56) and is given by $$V_{ij}^{(\gamma )}=C_{ij}\frac{e}{4f^2}(Q_i+Q_j)\overline{u}(p^{})\gamma ^\mu u(p)ϵ_\mu ,$$ (5.8) where $`Q_i`$, $`Q_j`$ are the initial and final meson charges and $`ϵ_\mu `$ the photon polarization vector and $`C_{ij}`$ the coefficients of eq. (3.56). Once again Bremsstrahlung terms on the meson lines of the tree diagram are negligible if the reaction is done close to threshold. In ref. the photon energy was chosen 1.7 GeV in the lab frame. In Fig. 5.3 we show $`d\sigma /dM_I`$ for the different channels. While all coupled channels collaborate to the building up of the $`\mathrm{\Lambda }`$(1405) resonance, most of them open up at higher energies and the resonance shape is only visible in the $`\pi ^+\mathrm{\Sigma }^{}`$, $`\pi ^{}\mathrm{\Sigma }^+`$, $`\pi ^0\mathrm{\Sigma }^0`$ channels. The $`\overline{K}N`$ production occurs at energies slightly above the resonace and the $`\pi ^0\mathrm{\Lambda }`$, with isospin one, only provides a small background below the resonance. It is interesting to see the different shapes of the three $`\pi \mathrm{\Sigma }`$ channels. This can be understood in terms of the isospin decomposition of the states $$|\pi ^+\mathrm{\Sigma }^{}=\frac{1}{\sqrt{6}}|2,0\frac{1}{\sqrt{2}}|1,0\frac{1}{\sqrt{3}}|0,0$$ (5.9) $$|\pi ^{}\mathrm{\Sigma }^+=\frac{1}{\sqrt{6}}|2,0+\frac{1}{\sqrt{2}}|1,0\frac{1}{\sqrt{3}}|0,0$$ (5.10) $$|\pi ^0\mathrm{\Sigma }^0=\sqrt{\frac{2}{3}}|2,0\frac{1}{\sqrt{3}}|0,0$$ (5.11) Disregarding the $`I=2`$ contribution which is negligible, the cross sections for the three channels are proportional to the modulus squared of the amplitude and hence they go as: $$\frac{1}{2}|T^{(1)}|^2+\frac{1}{3}|T^{(0)}|^2+\frac{2}{\sqrt{6}}\mathrm{Re}(T^{(0)}T^{(1)});\pi ^+\mathrm{\Sigma }^{}$$ (5.12) $$\frac{1}{2}|T^{(1)}|^2+\frac{1}{3}|T^{(0)}|^2\frac{2}{\sqrt{6}}\mathrm{Re}(T^{(0)}T^{(1)});\pi ^{}\mathrm{\Sigma }^+$$ (5.13) $$\frac{1}{3}|T^{(0)}|^2;\pi ^0\mathrm{\Sigma }^0$$ (5.14) The crossed term $`T^{(0)}T^{(1)}`$ is what makes these cross sections different. We can also see that $$3\frac{d\sigma }{dM_I}(\pi ^0\mathrm{\Sigma }^0)\frac{d\sigma }{dM_I}(I=0)$$ (5.15) $$\frac{d\sigma }{dM_I}(\pi ^0\mathrm{\Sigma }^0)+\frac{d\sigma }{dM_I}(\pi ^+\mathrm{\Sigma }^{})+\frac{d\sigma }{dM_I}(\pi ^{}\mathrm{\Sigma }^+)\frac{d\sigma }{dM_I}(I=0)+\frac{d\sigma }{dM_I}(I=1)$$ (5.16) This means that the real shape of the resonance must be seen in either the $`\pi ^0\mathrm{\Sigma }^0`$ channel or in the sum of the three $`\pi \mathrm{\Sigma }`$ channels, provided the I = 1 cross section (not the crossed terms which are relatively large) is small as it is the case. Incidentally, eqs. (5.12),(5.13) also show that the difference between the $`\pi ^+\mathrm{\Sigma }^{}`$ and $`\pi ^{}\mathrm{\Sigma }^+`$ cross sections gives the crossed term and hence provides some information on the $`I=1`$ amplitude. In Fig. 5.4 the results are recombined in a practical way from the experimental point of view. They show the $`I=0`$ contribution, the $`\mathrm{\Sigma }^0\pi ^0`$ contribution and the sum of all channels including the $`\pi ^0\mathrm{\Lambda }`$, and we see that they are all very similar and the total contribution is just the $`\mathrm{\Lambda }`$(1405) contribution plus a small background. In practical terms this result means that the detection of the $`K^+`$ alone (which sums the contribution of all channels) is sufficient to determine the shape and the strength of the $`\mathrm{\Lambda }`$(1405) resonance in this reaction. The study of the reaction in nuclei requires special care. Indeed, assume one uses the same set up as before with a nuclear target and measures the outgoing $`K^+`$. There the invariant mass will be given by $`M_I^2(p)=(q+pk)^2`$ $`=`$ $`M^2+m_K^22q^0k^0+2\stackrel{}{q}\stackrel{}{k}+`$ $`2p^0(q^0k^0)2\stackrel{}{p}(\stackrel{}{q}\stackrel{}{k})`$ with $`q,k,p`$ the momenta of the photon, $`K^+`$ and initial proton respectively. Since $`\stackrel{}{q}\stackrel{}{k}`$ has a large size, there will be a large spreading of invariant masses due to Fermi motion for a given set up of photon and $`K^+`$ momenta, unlike in the free proton case where $`M_I^2`$ is well determined<sup>1</sup><sup>1</sup>1 We are endebted to T. Nakano and J. K. Ahn for calling us the attention on this point. The nuclear cross section normalized to the number of protons (the neutrons through $`K^{}n`$ and coupled channels only contribute to $`I=1`$ with a small background) would be given by the convolution formula $$\frac{1}{Z}\frac{d\sigma }{dM_I}|_A\frac{2}{\rho _p}\frac{d^3p}{(2\pi )^3}\frac{d\sigma }{dM_I(\stackrel{}{p})};\rho _p=\frac{k_F^3}{3\pi ^2}$$ (5.18) where the integral over $`\stackrel{}{p}`$ ranges up to the Fermi momentum $`k_F`$. In order to show the effects of the Fermi motion, values of $`\stackrel{}{k}`$ corresponding to forward $`K^+`$ in the CM (and hence largest value of $`\stackrel{}{k}`$ in the lab frame) are chosen in ref. . These components would minimize the spreading of the $`M_I^2(\stackrel{}{p})`$ in eq. (5.2). Even then, the spreading of the invariant masses is so large that one looses any trace of the original resonance, as one can see in Fig. 5.4. The $`M_I`$ in the x axis of the figure in this case is taken for reference from eq. (5.2) for a nucleon at rest. This result simply means that in order to see genuine dynamical effects one would have to look at the invariant mass of the resonance from its decay product, $`\pi \mathrm{\Sigma }`$, tracing back this original invariant mass with appropiate final state interaction corrections. One interesting thing here is that the $`\mathrm{\Lambda }(1405)`$ resonance is produced with a large momentum in the nuclear lab frame. Because of that, Pauli blocking effects in the resonance decay, which are so important for the resonance at rest in the nucleus, become now irrelevant. Hence medium modifications of the resonance in the present situation should be attributed to other dynamical effects . Experiments on this reaction are now done at TJNAF and are being analysed. They are also scheduled to run with priority at LEPS of SPring8/RCNP. These experiments will allow to test current ideas on chiral symmetry for the elementary reaction. When used with nuclear targets they should provide us with much needed information on the in medium properties of the $`\mathrm{\Lambda }`$(1405) resonance and the $`K^{}`$ meson. This should help resolve questions like $`K^{}`$ condensation and the origin of the attraction seen in $`K^{}`$ atoms. ### 5.3 Radiative production of the $`𝚲\mathbf{(}\mathrm{𝟏𝟒𝟎𝟓}\mathbf{)}`$ resonance in $`𝑲^{\mathbf{}}`$ collisions on protons and nuclei One of the problems which we encountered in the former section regarding the study of the properties of the $`\mathrm{\Lambda }(1405)`$ resonance in nuclei is that the detection of the $`K^+`$ alone did not allow one to observe the shape of the resonance because the effect of Fermi motion of the nucleons in the nucleus produced a large spread of the invariant mass of the meson-baryon system building up the resonance. One had to reconstruct the invariant mass from the $`\pi \mathrm{\Sigma }`$ decay products. In this section we report on an alternative reaction to produce the $`\mathrm{\Lambda }(1405)`$ resonance which is dynamically quite different from the photoproduction process, hence offering extra tests of the chiral symmetry ideas in the baryon sector. Furthermore, it has an attractive feature since in this case an easy experimental set up is still sufficient to investigate the resonance properties in nuclei. The reaction reported here is the $`K^{}p\mathrm{\Lambda }(1405)\gamma `$ at low $`K^{}`$ energies, which was studied in ref. . Although the present reaction corresponds to a crossed channel of the $`\gamma pK^+\mathrm{\Lambda }(1405)`$ reaction studied in ref. and reported above, the two processes are rather different dynamically in their respective physical channels, with the dominant mechanisms in the photoproduction reaction being negligible in the present one, and others which could be proved negligible in the photoproduction one becoming now dominant. The set of diagrams considered in ref. is depicted in Fig. 5.5. The first line simply shows the diagrams of the Bethe-Salpeter equation which are utilized to generate the meson-baryon scattering $`T`$-matrix with coupled channels. The channels considered here are the same 10 channels considered in section 3.3.2. The rest of the diagrams in Fig. 5.5 stand for the radiative production of the $`\mathrm{\Lambda }(1405)`$. Apart from the strong $`MBM^{}B^{}`$ vertices of eq. (3.56) we also need the coupling of the photon to the baryons, the mesons, plus the contact term of diagram (2.a) of Fig. 5.5 required by gauge invariance. These vertices are standard and after the nonrelativistic reduction of the $`\gamma `$ matrices, are given in the Coulomb gauge, $`ϵ^0=0`$, $`\stackrel{}{ϵ}\stackrel{}{q}=0`$, with $`\stackrel{}{q}`$ the photon momentum, by $$a)it_{M^{}M\gamma }=2ieQ_M\stackrel{}{k}^{}\stackrel{}{ϵ}$$ (5.19) for the coupling of the photon to the mesons, with $`e`$ electron charge, $`Q_M`$ the charge of the meson, $`k^{}`$ the momentum of the outgoing meson and $`ϵ_\mu `$ the photon polarization vector, $$b)it_{B^{}B\gamma }=ie(Q_B\frac{\stackrel{}{p}+\stackrel{}{p}^{}}{2M_B}i\frac{\stackrel{}{\sigma }\times \stackrel{}{q}}{2M_B}\mu _B)\stackrel{}{ϵ}$$ (5.20) for the coupling of the photon to the baryons, with $`Q_B`$ the charge of the baryon, $`\stackrel{}{p}`$, $`\stackrel{}{p}^{}`$ the incoming, outgoing baryon momenta and $`M_B`$, $`\mu _B`$ the mass and magnetic moment of the baryon, and $$c)it_{B^{}M^{}BM\gamma }=iC_{ij}(Q_i+Q_j)\{\frac{\stackrel{}{p}+\stackrel{}{p}^{}}{2\overline{M}}i\frac{\stackrel{}{\sigma }\times (\stackrel{}{p}\stackrel{}{p}^{})}{2\overline{M}}\}\stackrel{}{ϵ}$$ (5.21) for the contact term of diagram (2.a) of Fig. 5.5, with $`C_{ij}`$ the coefficients of eq. (3.56), $`i,j`$ standing for a $`MB`$ state, $`Q_i`$, $`Q_j`$ the charges of the mesons, $`\overline{M}`$ an average mass of the baryons and $`\stackrel{}{p}`$, $`\stackrel{}{p}^{}`$ the momenta of the incoming, outgoing baryons. In ref. one is concerned with $`K^{}`$ with momenta below 500 MeV/c in the lab frame. In this energy domain it is easy to see that the Bremsstrahlung diagrams from mesons and baryons (diagrams (3.a), (4.a), (5.a), (6.a)) are of the same order of magnitude and that the contact term (diagram (2.a)) is of order $`q/2M`$ of the corresponding meson Bremsstrahlung diagrams ((3.a) and (5.a)). With CM photon momenta $`q`$ of the order of 150 MeV or below, the terms of row 2 represent corrections below the 8$`\%`$ level and are neglected. In addition, terms like in diagram (2.d), where the photon couples to internal vertices of the loops, vanish for parity reasons. The diagrams in row 7 of Fig. 5.5 where the photon couples to mesons in the loops vanish due to the gauge condition $`\stackrel{}{ϵ}\stackrel{}{q}=0`$ and the same happens to the diagrams in row 8, where the photon couples with the dielectric part to the baryons inside the loops ($`(\stackrel{}{p}+\stackrel{}{p}^{})`$ term of eq. (5.21)). The magnetic coupling of the photons in row 8 survives. Hence, the process is given, within the approximations mentioned, by the diagrams in rows 3, 4, 5, 6 plus the magnetic part in row 8. This situation is opposite to the one found in ref. for $`\mathrm{\Lambda }(1405)`$ photoproduction close to threshold, where the dominant terms came from the contact term and the Bremsstrahlung diagrams were negligible. If one inspects the series of terms in rows 3 and 4 of Fig. 5.5 one can see that the strong part of the interaction to the right of the electromagnetic vertex involves the series of terms of the Bethe-Salpeter equation and generates the $`T`$-matrix from the initial $`MB`$ state to the final $`M^{}B^{}`$ state after losing the energy of the photon, this is, with an argument $`M_I`$, where $`M_I`$ is the invariant mass of the $`M^{}B^{}`$ state. Similarly, in the rows (5) (6) the strong $`T`$-matrix factorizes before the electromagnetic vertex with an argument $`\sqrt{s}`$, with $`s`$ the Mandelstam variable for the initial $`K^{}p`$ system. In the diagrams of row (8) we have a loop with one meson and two baryons. The strong interaction to the left originates $`T(\sqrt{s})`$ and the one to the right $`T(M_I)`$. The loop function of row (8) contains two baryon propagators and for the small energies involved here can be obtained by differentiating the $`G(\sqrt{s})`$ function of a meson-baryon loop, eq. (3.63), with respect to $`\sqrt{s}`$, which duplicates the baryon propagator. By choosing an appropiate pair ($`ϵ_1,ϵ_2`$) of orthogonal photon polarization vectors, also orthogonal to $`\stackrel{}{q}`$, and summing over final photon and baryon polarizations plus averaging over the initial proton polarizations, one obtains the cross section for the process given by ($`\sigma `$ the cross section for each $`i,j`$ transition) $$\frac{d\sigma }{dM_Id\phi }=\frac{1}{2\pi }\frac{d\sigma }{dM_I}+\frac{d\sigma _I}{dM_Id\phi }\mathrm{cos}\phi ,$$ (5.22) with $`\phi `$ the azimutal angle formed by the plane containing the $`\stackrel{}{k}^{}`$ and $`\stackrel{}{q}`$ vectors and the one containing the $`\stackrel{}{k}`$ and $`\stackrel{}{q}`$ vectors. The only dependence on the azimutal angle $`\phi `$ comes in the $`\mathrm{cos}\phi `$ dependence which accompanies the interference cross section, $`\sigma _I`$, in eq. (5.22), which means that both $`d\sigma /dM_I`$ and $`d\sigma _I/dM_Id\phi `$ do not depend on the angle $`\phi `$. Explicit expressions for $`d\sigma /dM_I`$ and $`d\sigma _I/dM_Id\phi `$ are given in ref. . The results for $`d\sigma /dM_I`$ are shown in Fig. 5.6. There one can see the results for the cross sections in the $`K^{}p\pi ^{}\mathrm{\Sigma }^+\gamma ,\pi ^+\mathrm{\Sigma }^{}\gamma ,\pi ^0\mathrm{\Sigma }^0\gamma ,\pi ^0\mathrm{\Lambda }\gamma `$, $`K^{}p\gamma `$ channels. The cross section for $`K^{}p\overline{K}^0n\gamma `$ is very small, around 0.1 mb GeV<sup>-1</sup> in the range $`1.441.52`$ GeV, and is not plotted in the figure. The $`\mathrm{\Lambda }(1405)`$ peak appears clearly in the $`\pi \mathrm{\Sigma }`$ spectrum. It is interesting to notice the difference between the cross sections for the different $`\pi \mathrm{\Sigma }`$ channels. The origin of this is the same one discussed in ref. and reported in the former section due to the different isospin combinations of the three charged states and the crossed products of the $`I=1`$, $`I=0`$ amplitudes which appear in the cross section. The $`\pi ^0\mathrm{\Sigma }^0`$ has no $`I=1`$ component and since the $`I=2`$ component is negligible, the $`\pi ^0\mathrm{\Sigma }^0`$ distribution is very similar to the $`I=0`$ $`\mathrm{\Lambda }(1405)`$ distribution. The $`I=0`$ contribution alone, coming from the excitation of the $`\mathrm{\Lambda }(1405)`$, can be obtained using a combination of the three $`\pi \mathrm{\Sigma }`$ amplitudes $$(t_{K^{}p\pi ^{}\mathrm{\Sigma }^+}+t_{K^{}p\pi ^+\mathrm{\Sigma }^{}}+t_{K^{}p\pi ^0\mathrm{\Sigma }^0})/\sqrt{3}.$$ (5.23) The results for the pure $`I=0`$ excitation (dotted line) shown in Fig. 5.6 look very similar to the total strength around the $`\mathrm{\Lambda }(1405)`$ peak. Below the $`K^{}p`$ threshold there is some strength for $`I=1`$, $`\pi ^0\mathrm{\Lambda }`$ excitation, which is also very small. As a consequence of that, the sum of all channels in the $`\mathrm{\Lambda }(1405)`$ region, which requires exclusively the detection of the photon, has approximately the $`\mathrm{\Lambda }(1405)`$ shape and strength. It is interesting to observe the fast rise of the cross section in the $`K^{}pK^{}p\gamma `$ channel, showing the Bremsstrahlung infrarred divergence at large $`M_I`$ (small photon momentum). The other channels also would show the infrarred divergence at higher energies, when the photon momentum goes to zero. The relative larger weight of the $`K^{}pK^{}p\gamma `$ reaction at these energies, with respect to the other ones, is a reflection of the fact that the $`K^{}pK^{}p`$ cross section at values of $`M_I`$ or $`\sqrt{s}`$ of the order of 1500 MeV is much bigger than the other $`K^{}pM^{}B^{}`$ cross sections. There is a lower limit, which happens around 200 MeV/c for the $`K^{}`$ lab momentum, where the tails of the distribution, reflecting the Bremsstrahlung properties of the reaction, overlap with the peak of the resonance and hence the information on the $`\mathrm{\Lambda }(1405)`$ is lost. Let us now turn the attention to nuclei where we would consider the reaction $`K^{}A\mathrm{\Lambda }(1405)\gamma (A1)`$. In this case if one detects only the photon one has a distribution of invariant masses due to Fermi motion since now $`M_I^2=(k+p_Nq)^2`$ and $`p_N`$ runs over all nucleon momenta of the occupied states. One can fold the results of $`d\sigma /dM_I`$ with the distribution of $`M_I`$ coming from a Fermi sea of nucleons. The results in this case are shown in Fig. 5.6 for $`\rho =\rho _0/4`$, a likely effective density for this reaction, taking into account the distortion of the initial $`K^{}`$ through the nucleus. We can see a widening of the $`\mathrm{\Lambda }(1405)`$ distribution, with the shape only moderately changed, such that other effects from genuine changes of the $`\mathrm{\Lambda }(1405)`$ properties in the medium, predicted to be quite drastic , could in principle be visible. Certainly, the detailed measurement of the final meson-baryon in coincidence with the photon would allow a much better determination of the $`\mathrm{\Lambda }(1405)`$ properties than just the photon detection, and ultimately these exclusive measurements should also be performed. But the fact that the simple detection of the photon can provide interesting information is a welcome feature from the experimental point of view. The reactions discussed can be easily implemented at present facilities like KEK or Brookhaven. In Brookhaven some data from recent $`K^{}p`$ experiment with detection of photons in the final state are in the process of analysis . The present results should encourage the detailed analysis of the particular channels discussed here. ## Chapter 6 Further nuclear applications In sections 5.2 and 5.3 we already discussed how the $`\mathrm{\Lambda }(1405)`$ resonance could be generated in a nuclear environment and the additional information that this would bring on the nature of the resonance, plus the repercussions of these findings on the interaction of $`\overline{K}`$ with nuclei. In this chapter we will show some examples where the combination of the chiral unitary approach with the many body techniques proves to be a rather powerful tool to clarify issues which have remained so far controversial, particularly the problems of the $`\pi \pi `$ interaction in the nuclear medium and the interaction of $`K^{}`$ with nuclei. ### 6.1 The isoscalar $`𝝅𝝅`$ interaction in a nuclear medium The $`\pi \pi `$ interaction in a nuclear medium in the $`J=I=0`$ channel ($`\sigma `$ channel) has stimulated much theoretical work lately. It was realized that the attractive P-wave interaction of the pions with the nucleus led to a shift of strength of the $`\pi \pi `$ system to low energies and eventually produced a bound state of the two pions around $`2m_\pi 10`$ MeV . This state would behave like a $`\pi \pi `$ Cooper pair in the medium, with repercussions in several observable magnitudes in nuclear reactions . The possibility that such effects could have already been observed in some unexpected enhancement in the ($`\pi ,2\pi `$) reaction in nuclei was also noticed there. More recent experiments where the enhancement is seen in the $`\pi ^+\pi ^{}`$ channel but not in the $`\pi ^+\pi ^+`$ channel have added more attraction to that conjecture. Yet, it was early realized that constraints of chiral symmetry in the amplitude at low energies might affect those conclusions . In order to investigate the influence of chiral constraints in $`\pi \pi `$ scattering in the nuclear medium two different models for the $`\pi \pi `$ interaction were used in ref. . One of them did not satisfy the chiral constraints, while another one produced an amplitude behaving like $`m_\pi `$ in the limit of small pion masses. The conclusion of ref. was that, although in the chirally constrained model the building up of $`\pi \pi `$ strength at low energies was attenuated, it was still important within the approximations done in their calculations. Among these approximations there is the use of only $`\mathrm{\Delta }h`$ excitation with zero $`\mathrm{\Delta }`$ width to build up the $`\pi `$ nuclear interaction. Warnings were also given that results might depend on the off shell extrapolation of the $`\pi \pi `$ scattering matrix. Further refinements were done in ref. , where the width of the $`\mathrm{\Delta }`$ and coupling to $`1p\mathrm{\hspace{0.17em}1}h`$ and $`2p\mathrm{\hspace{0.17em}2}h`$ components were considered. The coupling of pions to the $`ph`$ continuum led to a dramatic re-shaping of the $`\pi \pi `$ strength distribution, but the qualitative conclusions about the accumulated strength at low energies remained. In ref. the importance of the coupling to the $`ph`$ components was reconfirmed and the use of more accurate models for the $`\pi \pi `$ interaction, as the Jülich model based on meson exchange , did not change the conclusions on the enhanced $`\pi \pi `$ strength at low energies. However, the use of a linear and nonlinear models for the $`\pi \pi `$ interaction, satisfying the chiral constraints at small energies, led to quite different conclusions and showed practically no enhancement of the $`\pi \pi `$ strength at low energies. The same conclusions were reached using the Jülich model with a subtracted dispersion relation so as to satisfy the chiral constraints. The latter model employed the Blakenbecler-Sugar equation in which the $`2\pi `$ intermediate states were placed on shell. The conclusion of this paper was that the imposition of chiral constraints in the $`\pi \pi `$ amplitude prevented the pairing instabilities shown by the other models not satisfying those constraints. In a further paper the authors showed, however, that the imposition of the chiral constraints by themselves did not prevent the pairing instabilities and uncertainties remained related to the off shell extrapolation of the $`\pi \pi `$ amplitude and the possible ways to implement the minimal chiral constraints. The situation, as noted in ref. , is rather ambiguous, but the studies done have certainly put the finger in the questions that should be properly addressed: chiral symmetry, off shell extrapolations, unitarity, etc. The chiral unitary methods discussed in section 3 address automatically all these questions and seem most appropriate to tackle the problem discussed above. That task was undertaken in ref. and we report here on their results. Since one is concerned about the S-wave $`\pi \pi `$ interaction, the Bethe-Salpeter approach is the most economical one to follow and this is what is done in ref. . In the nuclear medium the pions will get renormalized and hence their propagators will differ from the free ones. In addition the point four meson vertices also get renormalized as we see below. In order to illustrate the medium modifications in the $`\pi \pi `$ amplitude we show in Figs. 6.1, 6.2 and 6.3 the diagrams which are now involved at the level of just one loop. The pion is dressed by allowing it to excite $`ph`$ and $`\mathrm{\Delta }h`$ components, as is usually done in pion nuclear physics . The diagrams in Fig. 6.1 involve the usual $`ph`$ and $`\mathrm{\Delta }h`$ excitation of the pion, which acquires a self-energy leading to a modification of the pion propagator, and the two pion loop function. In ref. only the pions were renormalized but the kaons were also taken into account in the coupled channel Bethe-Salpeter equation. In Fig. 6.2 another sort of diagrams appears. These diagrams, which qualify as vertex corrections of the four meson vertex, come from the Lagrangian of Eq. (2.5), and involve a baryon line and three mesons. The need to consider this contact three meson term in connection with diagrams involving a pion pole, like one has in Fig. 6.1, was already known prior to the developments of $`\chi PT`$ and has been systematically used in studies of the $`\pi N\pi \pi N`$ reaction in refs. . The advent of $`\chi PT`$ has made it easier to extend these ideas to the strangeness sector where the effective Lagrangians were not available. The presence of the three meson contact term leads also to the term in Fig. 6.3 in a natural way. The interesting feature of these diagrams is that when one separates the four pion vertex appearing in Figs. 6.1, 6.2 into an on shell and an off shell part, as done in section 3.3.1, eq. (3.53), there is an exact cancellation between the off shell parts and the three meson contact terms of Figs. 6.2, 6.3, such that at the end only diagrams of the type of Fig. 6.1 must be evaluated and with all the four pion vertices evaluated on shell (i.e. taking $`p^2`$=$`m^2`$ for all the meson lines in the expression of the amplitude). This subtle cancellation, which also makes the work simpler technically, was first observed in ref. when checking that the results cannot depend on the arbitrary coefficient that one has in chiral theories at the level of three pion fields in the expansion of the $`U`$ function defined after eq. (2.5). The interplay found here for the pion pole term and the three meson contact term has been investigated in other nuclear problems before, where also interesting cancellations were found. In ref. it was shown that the real part of the kaon self-energy in the nuclear medium tied to the scattering of the kaons with the virtual pion nuclear cloud was zero, because of exact cancellations between the four meson terms and the contact three meson baryon terms. This solved a puzzle at the time where large uncertainties were tied to the off shell part of the $`K\pi `$ amplitude . The phenomenology also demanded that this real part be close to zero . Similarly, in ref. the pion self-energy tied to the interaction of the pions with the virtual pion cloud was found to be very small, with partial cancellations which became complete in the chiral limit. The results obtained for the imaginary part of the $`\pi \pi `$ amplitude are shown in Fig. 6.4 for different values of the Fermi momentum. One can observe a depletion of the strength in the region of 600-900 MeV, but more interesting, in connection with the experiments reporting an enhancement of the invariant mass of the two pions close to threshold, is the strength found in the figure in that region. The results shown in Fig. 6.4 are very similar to those of ref. where minimal chiral constraints were imposed in their models which eliminated the peaks below threshold found in earlier works. The situation is nevertheless still puzzling. Indeed, by using a model of ref. for the $`\pi \pi `$ interaction in the medium and a model for the $`(\pi ,\pi \pi )`$ reaction from ref. , a spectrum of invariant masses similar to the experimental one could be reproduced in ref. . A more detailed work was carried in ref. improving on the approximations done in ref. and it was found that the enhancement of the invariant mass found was narrowly tied to the approximations done, and when more accurate calculations were done there was no much sign of an enhanced invariant mass in the two pion mass distribution. One of the reasons for the apparent lack of enhancement is that there are large cancellations of pieces of the amplitude for the case of $`\pi ^+\pi ^{}`$ production which is not the case for $`\pi ^+\pi ^+`$ production. This latter process exhibits a peak at low invariant masses mostly due to phase space reasons. Should there be a renormalization of some terms in the nucleus which would alter the cancellation found in free space, then the results would look more like those of the $`\pi ^+\pi ^+`$ production and the experimental observation could be understood. Further work is necessary to understand better the process before we can see in this reaction a precursor of the chiral symmetry restoration, which is one of the appealing possibilities suggested so far . ### 6.2 The $`𝑲^{\mathbf{}}`$ nucleus interaction The properties of the kaons and antikaons in the nuclear medium have been the object of numerous investigations since the possibility of the existence of a kaon condensed phase in dense nuclear matter was pointed out . If the $`K^{}`$ meson develops sufficient attraction in dense matter it could be energetically more favorable, after a certain critical density, to neutralize the positive charge with antikaons rather than with electrons. A condensed kaon phase would then start to develop, changing drastically the properties of dense neutron star matter . In fact, the enhancement of the $`K^{}`$ yield in Ni+Ni collisions measured recently by the KaoS collaboration at GSI can be explained by assuming the $`K^{}`$ meson to feel a strong attraction in the medium , although alternative mechanisms, such as the production of antikaons via $`\mathrm{\Sigma }`$ hyperons, have also been suggested . Kaonic atom data, a compilation of which is given in ref. , also favor an attractive $`K^{}`$ nucleus interaction. The theoretical investigations that go beyond pure phenomenology have mainly followed two different strategies. One line of approach is that of the mean field models, built within the framework of chiral Lagrangians , based on the relativistic Walecka model extended to incorporate strangeness in the form of hyperons or kaons or using explicitly quark degrees of freedom . The other type of approach aims at obtaining the in-medium $`\overline{K}N`$ interaction microscopically by incorporating the medium modifications in a $`\overline{K}N`$ amplitude that reproduces the low energy scattering data and generates the $`\mathrm{\Lambda }(1405)`$ resonance dynamically . For instance, Pauli blocking on the intermediate nucleon states of the Lippmann-Schwinger or, alternatively, the Bethe-Salpeter equation makes the $`\overline{K}N`$ interaction density dependent and this, in turn, modifies the $`K^{}`$ properties from those in free space. These medium modifications were already included long time ago in the context of Brueckner-type many body theory using a separable $`\overline{K}N`$ interaction to obtain the kaon-nucleus optical potential for kaonic atoms. The more recent theoretical works take the $`\overline{K}N`$ interaction from the chiral Lagrangian. The blocking of intermediate states shifts the resonance to higher energy and this changes the $`\overline{K}N`$ interaction at threshold from being repulsive in free space to being attractive in the medium. A recent self-consistent calculation of the $`K^{}`$ self-energy has shown that the position of the resonance remains unchanged, due to a compensation of the repulsive Pauli blocking effects with the attraction felt by the $`K^{}`$ meson, in qualitatively agreement with was was noted in ref. using a constant mean field potential for the $`\overline{K}`$. Additional medium effects have been considered in a recent work , including the self-energy of the pions in the $`\pi \mathrm{\Lambda }`$, $`\pi \mathrm{\Sigma }`$ intermediate states, which couple strongly to the $`\overline{K}N`$ state, as well as the dressing of the baryons ($`N,\mathrm{\Lambda },\mathrm{\Sigma }`$) through density-dependent mean-field binding potentials. The starting point is the chiral model of ref. , described in section 3.3.2, which reproduces the $`\overline{K}N`$ low energy scattering observables. The medium effects on the $`\overline{K}N`$ interaction are incorporated replacing the free meson and baryon propagators in the meson-baryon loop of Eq. (3.63) by in-medium ones. For the nucleon, a mean-field propagator $$A(\sqrt{s}q^0,\stackrel{}{q},\rho )=\frac{1n(\stackrel{}{q}_{\mathrm{lab}})}{\sqrt{s}q^0E_l(\stackrel{}{q})+iϵ}+\frac{n(\stackrel{}{q}_{\mathrm{lab}})}{\sqrt{s}q^0E_l(\stackrel{}{q})iϵ}$$ (6.1) is taken, where $`n(\stackrel{}{q}_{\mathrm{lab}})`$ is the occupation probability of a nucleon of momentum $`\stackrel{}{q}_{\mathrm{lab}}`$ in the lab frame. For the hyperons ($`\mathrm{\Lambda }`$ and $`\mathrm{\Sigma }`$), the occupation probability is simply zero. The single particle energy, $`E_l(\stackrel{}{q})`$, now contains a mean-field potential of the type $`U_0\rho /\rho _0`$, with $`\rho _0=0.17`$ fm<sup>-3</sup> being the normal nuclear matter density. For the nucleon, a reasonable depth value is $`U_0^N=70`$ MeV, as suggested by numerous calculations of the nucleon potential in nuclear matter. For the $`\mathrm{\Lambda }`$ hyperon, it is reasonable to take $`U_0^\mathrm{\Lambda }=30`$ MeV, as implied by the extrapolation to very heavy systems of the experimental $`\mathrm{\Lambda }`$ single particle energies in $`\mathrm{\Lambda }`$ hypernuclei . For the $`\mathrm{\Sigma }`$ hyperon, there is no conclusive information on the potential. Early phenomenological analyses and calculations found the $`\mathrm{\Sigma }`$ atom data to be compatible with $`U_0^\mathrm{\Sigma }30`$ MeV, but more recent analysis do not exclude a repulsive potential in the nuclear interior . In the work of ref. the potential used is $`U^\mathrm{\Sigma }=30\rho /\rho _0`$ MeV, as commonly accepted for low densities, but the effects of using a repulsive depth of 30 MeV are also explored. The meson propagator for the $`\overline{K}`$ and $`\pi `$ mesons is replaced by the dressed one $$D_l(q^0,\stackrel{}{q},\rho )=\frac{1}{(q^0)^2\stackrel{}{q}^2m_l^2\mathrm{\Pi }_l(q^0,\stackrel{}{q},\rho )}=_0^{\mathrm{}}𝑑\omega \mathrm{\hspace{0.17em}2}\omega \frac{S_l(\omega ,\stackrel{}{q},\rho )}{(q^0)^2\omega ^2+iϵ},$$ (6.2) where $`\mathrm{\Pi }_l(q^0,\stackrel{}{q},\rho )`$ is the meson self-energy. The second equality in Eq. (6.2) is the Lehmann representation of the meson propagator and $`S_l(\omega ,\stackrel{}{q},\rho )=\mathrm{Im}D_l(\omega ,\stackrel{}{q},\rho )/\pi `$ is the meson spectral density which, in the case on undressed mesons, reduces to $`\delta (\omega \omega _l(\stackrel{}{q}))/2\omega _l(\stackrel{}{q})`$. With these modifications the loop integral becomes $`G_l(P^0,\stackrel{}{P},\rho )`$ $`=`$ $`{\displaystyle _{\stackrel{}{q}<q_{\mathrm{max}}}}{\displaystyle \frac{d^3q}{(2\pi )^3}}{\displaystyle \frac{M_l}{E_l(\stackrel{}{q})}}{\displaystyle _0^{\mathrm{}}}𝑑\omega S_l(\omega ,\stackrel{}{q},\rho )`$ (6.3) $`\times `$ $`\left\{{\displaystyle \frac{1n(\stackrel{}{q}_{\mathrm{lab}})}{\sqrt{s}\omega E_l(\stackrel{}{q})+iϵ}}+{\displaystyle \frac{n(\stackrel{}{q}_{\mathrm{lab}})}{\sqrt{s}+\omega E_l(\stackrel{}{q})iϵ}}\right\},`$ where $`(P^0,\stackrel{}{P})`$ is the total four-momentum in the lab frame and $`s=(P^0)^2\stackrel{}{P}^2`$. The in-medium $`\overline{K}N`$ interaction, $`T_{\mathrm{eff}}(P^0,\stackrel{}{P},\rho )`$, is then obtained by solving the coupled-channel Bethe-Salpeter equation using the dressed meson-baryon loop of Eq. (6.3). The S-wave $`\overline{K}`$ self-energy ($`\overline{K}=K^{}`$ or $`\overline{K}^0`$) is determined by summing the in-medium $`\overline{K}N`$ interaction over the nucleons in the Fermi sea $$\mathrm{\Pi }_{\overline{K}}^s(q^0,\stackrel{}{q},\rho )=2\underset{N=n,p}{}\frac{d^3p}{(2\pi )^3}n(\stackrel{}{p})T_{\mathrm{eff}}^{\overline{K}N}(q^0+E(\stackrel{}{p}),\stackrel{}{q}+\stackrel{}{p},\rho ).$$ (6.4) Note that a self-consistent approach is required since one calculates the $`\overline{K}`$ self-energy from the effective interaction $`T_{\mathrm{eff}}`$ which uses $`\overline{K}`$ propagators which themselves include the self-energy being calculated. A P-wave contribution to the $`\overline{K}`$ self-energy coming from the coupling of the $`\overline{K}`$ meson to hyperon-hole excitations is also included and the expression can be found in ref. . The pion self-energy is built from a model that contains the effect of one- and two-nucleon absorption and is conveniently modified to include the effect of nuclear short-range correlations (see ref. for details). To assess the importance of dressing the pions we show in Fig. 6.5 the spectral density of the $`\pi `$ meson in nuclear matter at density $`\rho =\rho _0`$ for several momenta. The strength is distributed over a wide range of energies and, as the pion momentum increases, the position of the peak is increasingly lowered from the corresponding one in free space as a consequence of the attractive pion-nuclear potential. Note that, to the left of the peaks, there appears the typical structure of the $`1p1h`$ excitations which give rise to $`1p1h\mathrm{\Lambda }`$ and $`1p1h\mathrm{\Sigma }`$ components in the effective $`\overline{K}N`$ interacion. The spectral function of a $`K^{}`$ meson of zero momentum is shown in Fig. 6.7 for various densities: $`\rho _0`$, $`\rho _0/2`$ and $`\rho _0/4`$. The results in the upper panel include only Pauli blocking effects, i.e. the nucleons propagate as in Eq. (6.1) but the mesons behave as in free space. At $`\rho _0/4`$ one clearly sees two excitation modes. The left one corresponds to the $`K^{}`$ pole branch, appearing at an energy smaller than the kaon mass, $`m_K`$, due to the attractive medium effects. The peak on the right corresponds to the $`\mathrm{\Lambda }(1405)`$-hole excitation mode, located above $`m_K`$ because of the shifting of the $`\mathrm{\Lambda }(1405)`$ resonance to energies above the $`K^{}p`$ threshold. As density increases, the $`K^{}`$ feels an enhanced attraction while the $`\mathrm{\Lambda }(1405)`$-hole peak moves to higher energies and loses strength, a reflection of the tendency of the $`\mathrm{\Lambda }(1405)`$ to dissolve in the dense nuclear medium. These features were already observed in ref. . The (self-consistent) incorporation of the $`\overline{K}`$ propagator in the Bethe-Salpeter equation softens the effective interaction, $`T_{\mathrm{eff}}`$, which becomes more spread out in energies. The resulting $`K^{}`$ spectral function (middle panel in Fig. 6.7) shows the displacement of the resonance to lower energies because, as noted by Lutz , the attraction felt by the $`\overline{K}`$ meson lowers the threshold for the $`\overline{K}N`$ states that had been increased by the Pauli blocking on the nucleons. This has a compensatory effect and the resonance moves backwards, slightly below its free space value. The $`K^{}`$ pole peak appears at similar or slightly smaller energies, but its width is larger, due to the strength of the intermediate $`\overline{K}N`$ states being distributed over a wider region of energies. Therefore the $`K^{}`$ pole and the $`\mathrm{\Lambda }(1405)`$-hole branches merge into one another and can hardly be distinguished. Finally, when the pion is dressed according to the spectral function shown in Fig. 6.5 the effective interaction $`T_{\mathrm{eff}}`$ becomes even smoother. The resulting $`K^{}`$ spectral function is shown in the bottom panel in Fig. 6.7. As seen by the long-dashed line, even at very small densities one no longer distinguishes the $`\mathrm{\Lambda }(1405)`$-hole peak from the $`K^{}`$ pole one. As density increases the attraction felt by the $`K^{}`$ is more moderate and the $`K^{}`$ pole peak appears at higher energies than in the other two approaches. However, more strength is found at very low energies, especially at $`\rho _0`$, due to the coupling of the $`K^{}`$ to the $`1p1h`$ and $`2p2h`$ components of the pionic strength. It is precisely the opening of the $`\pi \mathrm{\Sigma }`$ channel, on top of the already opened $`(1p1h)\mathrm{\Sigma }`$ and $`(2p2h)\mathrm{\Sigma }`$ ones, the reason for the cusp structure which appears slightly above 400 MeV. The isospin averaged in-medium scattering length, defined as $$a_{\mathrm{eff}}(\rho )=\frac{1}{4\pi }\frac{M}{m_K+M}\frac{\mathrm{\Pi }_{\overline{K}}(m_K,\stackrel{}{q}=0,\rho )}{\rho },$$ (6.5) is shown in Fig. 6.7 as a function of the nuclear density $`\rho `$. The change of $`\mathrm{Re}a_{\mathrm{eff}}`$ from negative to positive values indicates the transition from a repulsive interaction in free space to an attractive one in the medium. As shown by the dotted line, this transition happens at a density of about $`\rho 0.1\rho _0`$ when only Pauli effects are considered, in agreement with what was found in ref. . However, this transition occurs at even lower densities ($`\rho 0.04\rho _0`$) when one considers the self-energy of the mesons in the description, whether one dresses only the $`\overline{K}`$ meson (dashed line) or both the $`\overline{K}`$ and $`\pi `$ mesons (solid line). The deviations from the approach including only Pauli blocking or those dressing the mesons are quite appreciable over a wide range of densities. The thin solid lines show the results obtained with a repulsive $`\mathrm{\Sigma }`$ potential of the type $`U^\mathrm{\Sigma }=U_0^\mathrm{\Sigma }\rho /\rho _0`$, with $`U_0^\mathrm{\Sigma }=30`$ MeV. The deviations from the thick solid line, obtained for an attractive potential depth of $`U_0^\mathrm{\Sigma }=30`$ MeV, are smaller than 10% and only show up at the higher densities. The implications on kaonic atoms of the scattering length displayed in Fig. 6.7 or, equivalently, the $`K^{}`$ optical potential, $`V_{\mathrm{opt}}(\rho )=\mathrm{\Pi }_{\overline{K}}(m_K,\stackrel{}{q}=0,\rho )/2m_K`$, have been analyzed in the framework of a local density approximation, where the nuclear matter density $`\rho `$ is replaced by the density profile $`\rho (r)`$ of the particular nucleus . As can be seen from Fig. 6.8, both the energy shifts and widths of kaonic atom states agree well with the bulk of experimental data. The model reported here gives a $`K^{}`$ nuclear potential depth of $`44`$ MeV at $`\rho =\rho _0`$. This is about half the attraction of that obtained with other recent theories and approximation schemes , which give rise to potential depths at the center of the nucleus in between $`140`$ and $`75`$ MeV, and also lies very far from the depth of around $`200`$ MeV obtained from a best fit to $`K^{}`$ atomic data with a phenomenological potential that includes an additional non-linear density dependent term . On the other hand, the early Brueckner-type calculations of ref. also obtained a shallow $`K^{}`$-nucleus potential, of the order of $`40`$ MeV at the center of <sup>12</sup>C, and predicted reasonably well the $`K^{}`$ atomic data available at that time and recent self-consistent calculations find a moderate attraction of $`32`$ MeV. Acceptable fits to kaonic atom data have also been obtained using charge densities and a phenomenological $`T_{\mathrm{eff}}\rho `$ type potential with a depth of the order of $`50`$ MeV in the nuclear interior. But, when matter densities are used instead, the fit gives a potential depth of $`80`$ MeV . A comparison of kaonic atom results obtained with various $`K^{}`$-nucleus potentials can be found in ref. . A hybrid model, combining a relativistic mean field approach in the nuclear interior and a phenomenological density dependent potential at the surface that is fitted to $`K^{}`$ atomic data, also favors a strongly attractive $`K^{}`$ potential of depth $`180`$ MeV . In summary, although all models predict attraction for the $`K^{}`$-nucleus potential, there are still large discrepancies for the precise value of its depth, which has important implications for the occurrence of kaon condensation. It is then necessary to gather more data that could help in disentangling the properties of the $`\overline{K}`$ in the medium. Apart from the valuable information that can be extracted from the production of $`K^{}`$ in heavy-ion collisions, one could also measure deeply bound kaonic states, which have been predicted to be narrow and could be measured in $`(K^{},\gamma )`$ or $`(K^{},p)`$ reactions . ## Chapter 7 Conclusions After a short review of the basic concepts of chiral symmetry and of several chiral Lagrangians we have discussed various nonperturbative methods to deal with the meson-meson and meson-baryon interactions which allow one to extend the region of applicability of the theory to higher energies than in $`\chi PT`$, where the low lying mesonic and baryonic resonances appear. The common ground of all these methods was the exact implementation of unitarity in coupled channels. The constraints imposed by unitarity allow one to extract information contained in the chiral Lagrangians which is not accessible with the standard $`\chi PT`$ expansion. One of the procedures followed was the Inverse Amplitude Method, which relies upon the expansion of the inverse of the scattering matrix and gives rise to an expansion in powers of $`p^2`$ with a larger convergence radius than $`\chi PT`$. In that method one could extend the predictions for meson-meson interactions up to about 1.2 GeV, and all mesonic resonances up to this energy were well reproduced, as well as phase shifts and inelasticities. A second method relied upon the use of the N/D method and the hypothesis of resonance saturation. In this case, the use of the information contained in the lowest order chiral Lagrangian, together with chiral loops and the explicit exchange of some resonances, which are genuine QCD states in the sense that they would remain in the large $`N_c`$ limit, allow also a good description of the meson-meson data up to about 1.5 GeV. This second method is particularly rewarding for it allows one to dig into the nature of the mesonic resonances and separate those which are preexisting QCD resonances, in the limit of large $`N_c`$, from others which qualify as dynamical meson-meson resonances coming from the multiple scattering of the mesons. In this way, it was stablished that the low lying scalar resonances, the $`\sigma `$, $`\kappa `$, $`a_0(980)`$ and to large extent the $`f_0(980)`$, are generated dynamically from multiple scattering from the lowest order chiral Lagrangian. On the contrary, a singlet contribution to the $`f_0(980)`$ and a scalar octet around 1.35 GeV would be the lightest preexisting scalar states. This latter method also allows one to understand why in the case of the scalar sector a succesful reproduction of the data can be obtained simply by means of the lowest order chiral Lagrangian and the Bethe-Salpeter equation, together with a suitable cut off, or regularizing scale. In the meson-baryon problem, applications were only done in the scalar sector taking advantage of the simplification of the Bethe-Salpeter equation, which was found to be a suitable approach much as in the case of the meson-meson scalar sector. In this case, low lying resonances like the $`\mathrm{\Lambda }(1405)`$ or the $`N(1535)`$ were generated within that approach, and a good reproduction of the low energy scattering data was found, particularly in the case of the $`K^{}N`$ interaction and coupled channels. Applications to problems of initial and final state interaction have also been shown. Since the energy region of applicability of the reported methods is much larger than the one of $`\chi PT`$, one could tackle many new problems formerly inaccessible with plain $`\chi PT`$ theory. We have also shown how these chiral approaches to the meson-baryon and meson-meson interactions have repercussions in nuclear physics and provide a new perspective into problems which have been rather controversial up to now, like the $`\pi \pi `$ scattering in a nuclear medium or the $`K^{}`$ nucleus interaction. Several reactions which can bring new light into these problems have also been reviewed and the implementation of the experiments is already planned in some laboratories. The methods exposed here open new possibilities to face a large number of problems, of which we have only given a few examples. Extension of the methods to higher energies by incorporating channels with more than two particles and application to other physical domains remain as challenges for the future. Acknowledgements This work has been partially supported by CICYT contracts PB95-1249, PB96-0753 (Spain) and by the Eurodaphne project, EEC-TMR contract FMRX-CT98-0169. We would like to thank our colleagues H.C. Chiang, F. Guerrero, S. Hirenzaki, A. Hosaka, T.S.H. Lee, E. Marco, Ulf-G. Meissner, J.C. Nacher, M. Oka, Y. Okumura, A. Parreño, J.R. Peláez, A. Pich, H. Toki and M.J. Vicente Vacas for many stimulating discussions and for their collaboration in some of the works reported here.
warning/0002/hep-ex0002046.html
ar5iv
text
# Measurements of Cross Sections and Forward-Backward Asymmetries at the 𝐙 Resonance and Determination of Electroweak Parameters ## 1 Introduction The Standard Model (SM) of electroweak interactions is tested with great precision by the experiments performed at the LEP and SLC $`\mathrm{e}^+\mathrm{e}^{}`$ colliders running at centre-of-mass energies, $`\sqrt{s}`$, close to the $`\mathrm{Z}`$ mass. From measurements of the total cross sections and forward-backward asymmetries in the reactions $`\mathrm{e}^+\mathrm{e}^{}\mathrm{hadrons}(\gamma ),`$ $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}(\gamma ),`$ $`\mathrm{e}^+\mathrm{e}^{}\mu ^+\mu ^{}(\gamma ),`$ $`\mathrm{e}^+\mathrm{e}^{}\tau ^+\tau ^{}(\gamma ),`$ (1) the mass, total and partial widths of the $`\mathrm{Z}`$ and other electroweak parameters are obtained by L3 and other experiments . The $`(\gamma )`$ indicates the presence of radiative photons. The large luminosity collected in the years $`199395`$ enables a significant improvement on our previous measurements of $`\mathrm{Z}`$ parameters. An integrated luminosity of $`103\mathrm{pb}^1`$ was collected, corresponding to the selection of $`2.510^6`$ hadronic and $`2.510^5`$ leptonic events. Most of the data were collected at a centre-of-mass energy corresponding to the maximum annihilation cross section. In 1993 and 1995 scans, of the $`\mathrm{Z}`$ resonance were performed where runs at the $`\mathrm{Z}`$ pole alternated with runs at about $`1.8\mathrm{GeV}`$ on either side of the peak. Compared to previous measurements, our event samples on the wings of the $`\mathrm{Z}`$ resonance are increased by more than a factor of five. The LEP beam energies were precisely calibrated at the three energy points in $`199395`$ using the method of resonant depolarisation . As a result, the contributions to the errors on the $`\mathrm{Z}`$ mass and total width from the uncertainty on the centre-of-mass energy are reduced by factors of about five and three, respectively, as compared to the data collected before. The installation of silicon strip detectors in front of the small angle electromagnetic calorimeters allows a much more precise determination of the fiducial volume used for the luminosity measurement . This improvement, together with the reduced theoretical uncertainty on the small angle Bhabha cross section , allows more precise measurements of the cross sections, in particular that for $`\mathrm{e}^+\mathrm{e}^{}\mathrm{hadrons}(\gamma )`$. This results in a better determination of the invisible $`\mathrm{Z}`$ width, from which the number of light neutrino generations is deduced. In this article measurements of hadronic and leptonic cross sections and leptonic forward-backward asymmetries, obtained from the data collected between 1993 and 1995, are presented. These measurements are combined with our published results from the data collected in $`199092`$ . The complete integrated luminosity collected by L3 at the $`\mathrm{Z}`$ resonance is $`143\mathrm{pb}^1`$, consisting of about $`3.510^6`$ hadronic and $`3.510^5`$ leptonic events. The results on the properties of the $`\mathrm{Z}`$ boson and on other electroweak observables presented here are based on the final analyses of the complete data set collected at the $`\mathrm{Z}`$ resonance. This article is organised as follows: After a brief description of the L3 detector in Section 2, we summarise in Section 3 features of the $`199395`$ data analysis common to all final states investigated. Section 4 addresses issues related to the LEP centre-of-mass energy. The measurement of luminosity is described in Section 5. The event selection and the analysis of the reactions in (1) are discussed in Sections 6 to 9 and the results on the measurements of total cross sections and forward-backward asymmetries are presented in Section 10. A general description of the fits performed to our data is given in Section 11. Various fits for $`\mathrm{Z}`$ parameters are performed in Section 12 and the results of the fits in the framework of the SM are given in Section 13. We summarise and conclude in Section 14. The Appendices A and B give details on the treatment of the $`t`$-channel contributions in $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}(\gamma )`$ and on technicalities of the fit procedures, respectively. ## 2 The L3 Detector The L3 detector consists of a silicon microvertex detector , a central tracking chamber, a high resolution electromagnetic calorimeter composed of BGO crystals, a lead-scintillator ring calorimeter at low polar angles , a scintillation counter system, a uranium hadron calorimeter with proportional wire chamber readout and an accurate muon spectrometer. Forward-backward muon chambers, completed for the 1995 data taking, extend the polar angle coverage of the muon system down to 24 degrees with respect to the beam line. All detectors are installed in a 12 m diameter magnet which provides a solenoidal field of $`0.5\mathrm{T}`$ in the central region and a toroidal field of $`1.2\mathrm{T}`$ in the forward-backward region. The luminosity is measured using BGO calorimeters preceded by silicon trackers situated on each side of the detector. In the L3 coordinate system the direction of the $`\mathrm{e}^{}`$ beam defines the $`z`$ direction. The $`xy`$, or $`r\varphi `$ plane, is the bending plane of the magnetic field, with the $`x`$ direction pointing to the centre of the LEP ring. The coordinates $`\varphi `$ and $`\theta `$ denote the azimuthal and polar angles. ## 3 Data Analysis The data collected between 1993 and 1995 are split into nine samples according to the year and the centre-of-mass energy. Data samples at $`\sqrt{s}m_\mathrm{Z}`$ are referred to as peak, those at off-peak energies are referred to as peak$`2`$ and peak$`+2`$. The peak samples in 1993 and 1995 are further split into data taken early in the year (pre-scan) and those peak runs interspersed with off-peak data taking (scan) which coincide with the precise LEP energy calibration (see Section 4). Cross sections and leptonic forward-backward asymmetries are determined for each data sample. Acceptances, background contaminations and trigger efficiencies are studied for all nine data samples separately to take into account their possible dependence on the centre-of-mass energy and the time dependence of the detector status. Systematic errors are determined for the data samples individually. Average values for uncertainties are used if no dependence on the centre-of-mass energy or the data taking period is observed. Correlations of the systematic errors among the data sets are estimated and are taken into account in the analyses to determine electroweak parameters. Acceptances and background contaminations from $`\mathrm{e}^+\mathrm{e}^{}`$-interactions are determined by Monte Carlo simulations. The following event generator programs are used for the various signal and background processes: JETSET and HERWIG for $`\mathrm{e}^+\mathrm{e}^{}\mathrm{hadrons}(\gamma )`$; KORALZ for $`\mathrm{e}^+\mathrm{e}^{}\mu ^+\mu ^{}(\gamma )`$ and $`\mathrm{e}^+\mathrm{e}^{}\tau ^+\tau ^{}(\gamma )`$; BHAGENE , BHWIDE and BABAMC for large angle $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}(\gamma )`$; BHLUMI for small angle $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}(\gamma )`$; GGG for $`\mathrm{e}^+\mathrm{e}^{}\gamma \gamma (\gamma )`$; DIAG36 for $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}\mathrm{}^+\mathrm{}^{}`$; DIAG36, PHOJET and PYTHIA for $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}\mathrm{hadrons}`$. For the simulation of hadronic final states the fragmentation parameters of JETSET and HERWIG are tuned to describe our data as discussed in Reference . The generated events are passed through a complete detector simulation. The response of the L3 detector is modelled with the GEANT detector simulation program which includes the effects of energy loss, multiple scattering and showering in the detector materials. Hadronic showers are simulated with the GHEISHA program. The performance of the detector, including inefficiencies and their time dependence as observed during data taking, is taken into account in the simulation. With this procedure, experimental systematic errors on cross sections and forward-backward asymmetries are minimized. ## 4 LEP Energy Calibration The average centre-of-mass energy of the colliding particles at the L3 interaction point is calculated using the results provided by the Working Group on LEP Energy . Every 15 minutes the average centre-of-mass energy is determined from measured LEP machine parameters, applying the energy model which is based on calibration by resonant depolarisation . This model traces the time variation of the centre-of-mass energy of typically $`1\mathrm{MeV}`$ per hour. The average centre-of-mass energies are calculated for each data sample individually as luminosity weighted averages. Slightly different values are obtained for different reactions because of small differences in the usable luminosity. The errors on the centre-of-mass energies and their correlations for the 1994 data and for the two scans performed in 1993 and 1995 are given in form of a $`7\times 7`$ covariance matrix in Table 1. The uncertainties on the centre-of-mass energy for the data samples not included in this matrix, i.e. the 1993 and 1995 pre-scans, are $`18\mathrm{MeV}`$ and $`10\mathrm{MeV}`$, respectively. Details of the treatment of these errors in the fits can be found in Appendix B. The energy distribution of the particles circulating in an $`\mathrm{e}^+\mathrm{e}^{}`$-storage ring has a finite width due to synchrotron oscillations. An experimentally observed cross section is therefore a convolution of cross sections at energies which are distributed around the average value in a gaussian form. The spread of the centre-of-mass energy for the L3 interaction point as obtained from the observed longitudinal length of the particle bunches in LEP is listed in Table 2 . The time variation of the average energy causes a similar, but smaller, effect which is included in these numbers. All cross sections and forward-backward asymmetries quoted below are corrected for the energy spread to the average value of the centre-of-mass energy. The relative corrections on the measured hadronic cross sections amount to $`+1.7`$ per mill () at the $`\mathrm{Z}`$ pole and to $`1.1`$ and $`0.6`$ at the peak$`2`$ and peak$`+2`$ energy, respectively. The absolute corrections on the forward-backward asymmetries are very small. The largest correction is $`0.0002`$ for the muon and tau peak$`2`$ data sets. The error on the energy spread is propagated into the fits, resulting in very small contributions to the errors of the fitted parameters (see Appendix B). The largest effect is on the total width of the $`\mathrm{Z}`$, contributing approximately $`0.3\mathrm{MeV}`$ to its error. During the operation of LEP, no evidence for an average longitudinal polarisation of the electrons or positrons has been observed. Stringent limits on residual polarisation during luminosity runs are set such that the uncertainties on the determination of electroweak observables are negligible compared to their experimental errors . The determination of the LEP centre-of-mass energy in $`199092`$ is described in References . From these results the LEP energy error matrix given in Table 3 is derived. ## 5 Luminosity Measurement The integrated luminosity $``$ is determined by measuring the number of small-angle Bhabha interactions $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}(\gamma )`$. For this purpose two cylindrical calorimeters consisting of arrays of BGO crystals are located on either side of the interaction point. Both detectors are divided into two half-rings in the vertical plane to allow the opening of the detectors during filling of LEP. A silicon strip detector, consisting of two layers measuring the polar angle, $`\theta `$, and one layer measuring the azimuthal angle, $`\varphi `$, is situated in front of each calorimeter to precisely define the fiducial volume. A detailed description of the luminosity monitor and the luminosity determination can be found in Reference . The selection of small-angle Bhabha events is based on the energy depositions in adjacent crystals of the BGO calorimeters which are grouped to form clusters. The highest-energy cluster on each side is considered for the luminosity analysis. For about 98% of the cases a hit in the silicon detectors is matched with a cluster and its coordinate is used; otherwise the BGO coordinate is retained. The event selection criteria are: 1. The energy of the most energetic cluster is required to exceed $`0.8E_\mathrm{b}`$ and the energy on the opposite side must be greater than $`0.4E_\mathrm{b}`$, where $`E_\mathrm{b}`$ is the beam energy. If the energy of the most energetic cluster is within $`\pm 5\%`$ of $`E_\mathrm{b}`$ the minimum energy requirement on the opposite side is reduced to $`0.2E_\mathrm{b}`$ in order to recover events with energy lost in the gaps between crystals. The distributions of the energy of the most energetic cluster and the cluster on the opposite side as measured in the luminosity monitors are shown in Figure 1 for the 1993 data. All selection cuts except the one under study are applied. 2. The cluster on one side must be confined to a tight fiducial volume: * 32 mrad $`<\theta <`$ 54 mrad; $`|\varphi 90^{}|>11.25^{}`$ and $`|\varphi 270^{}|>11.25^{}`$. The requirements on the azimuthal angle remove the regions where the half-rings of the detector meet. The cluster on the opposite side is required to be within a larger fiducial volume: * 27 mrad $`<\pi \theta <`$ 65 mrad; $`|\varphi 90^{}|>3.75^{}`$ and $`|\varphi 270^{}|>3.75^{}`$. This ensures that the event is fully contained in the detectors and edge effects in the reconstruction are avoided. 3. The coplanarity angle $`\mathrm{\Delta }\varphi =\varphi (z<0)\varphi (z>0)`$ between the two clusters must satisfy $`|\mathrm{\Delta }\varphi 180^{}|<10^{}`$. The distribution of the coplanarity angle is shown in Figure 2. Very good agreement with the Monte Carlo simulation is observed. Four samples of Bhabha events are defined by applying the tight fiducial volume cut to one of the $`\theta `$-measuring silicon layers. Taking the average of the luminosities obtained from these samples minimizes the effects of relative offsets between the interaction point and the detectors. The energy and coplanarity cuts reduce the background from random beam-gas coincidences. The remaining contamination is very small: $`(3.4\pm 2.2)10^5`$. This number is estimated using the sidebands of the coplanarity distribution, $`10^{}<|\mathrm{\Delta }\varphi 180^{}|<30^{}`$, after requiring that neither of the two clusters have an energy within $`\pm 5\%`$ of $`E_\mathrm{b}`$. The accepted cross section is determined from Monte Carlo $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}(\gamma )`$ samples generated with the BHLUMI event generator at a fixed centre-of-mass energy of $`\sqrt{s}=91.25\mathrm{GeV}`$. The dependence on the centre-of-mass energy, as well as the contributions of $`\mathrm{Z}`$-exchange and $`\gamma \mathrm{Z}`$ interference, are calculated with the BHLUMI program. At $`\sqrt{s}=91.25\mathrm{GeV}`$ the accepted cross section is determined to be $`69.62\mathrm{nb}`$. The statistical error on the Monte Carlo sample contributes $`0.35`$ to the uncertainty of the luminosity measurement. The theoretical uncertainty on the Bhabha cross section in our fiducial volume is estimated to be $`0.61`$ . The experimental errors of the luminosity measurement are small. Important sources of systematic errors are: geometrical uncertainties due to the internal alignment of the silicon detectors ($`0.15`$ to $`0.27`$), temperature expansion effects ($`0.14`$) and the knowledge on the longitudinal position of the silicon detectors ($`0.16`$ to $`0.60`$). The precision depends on the accuracy of the detector surveys and on the stability of the detector and wafer positions during the different years. The polar angle distribution of Bhabha scattering events used for the luminosity measurement is shown in Figure 3. The structure seen in the central part of the $`+z`$ side is due to the flare in the beam pipe on this side. The imperfect description in the Monte Carlo does not pose any problem as it is far away from the edges of the fiducial volume. The overall agreement between the data and Monte Carlo distributions of the selection quantities is good. Small discrepancies in the energy distributions at high energies are due to contamination of Bhabha events with beam-gas interactions and, at low energies, due to an imperfect description of the cracks between crystals. The selection uncertainty is estimated by varying the selection criteria over reasonable ranges and summing in quadrature the resulting contributions. This procedure yields errors between $`0.42`$ and $`0.48`$ for different years. The luminosities determined from the four samples described above agree within these errors. The trigger inefficiency is measured using a sample of events triggered by only requiring an energy deposit exceeding $`30\mathrm{GeV}`$ on one side. It is found to be negligible. The various sources of uncertainties are summarized in Table 4. Combining them in quadrature yields total experimental errors on the luminosity of $`0.86`$, $`0.64`$ and $`0.68`$ in 1993, 1994 and 1995. Correlations of the total experimental systematic errors between different years are studied and the correlation matrix is given in Table 5. The error from the theory is fully correlated. Because of the $`1/s`$ dependence of the small angle Bhabha cross section, the uncertainty on the centre-of-mass energies causes a small additional uncertainty on the luminosity measurement. For instance, this amounts to $`0.1`$ for the high statistics data sample of 1994. This effect is included in the fits performed in Section 12 and 13, see Appendix B. The statistical error on the luminosity measurement from the number of observed small angle Bhabha events is also included in those fits. Table 6 lists the number of observed Bhabha events for the nine data samples and the corresponding errors on cross section measurements. Combining all data sets taken in $`199395`$ at $`\sqrt{s}m_\mathrm{Z}`$ the statistical error on the luminosity contributes 0.45 to the uncertainty on the pole cross section measurements. Higher order corrections from photon radiation to the small angle Bhabha cross section are studied with the photon spectrum of luminosity events. For this analysis events with two distinct energy clusters exceeding $`0.1E_\mathrm{b}`$ in one of the calorimeters are selected. The photon is identified as the lower energy cluster. The fraction of radiative events with $`E_\gamma >0.1E_\mathrm{b}`$ in the total low-angle Bhabha sample is 2% and the measured cross section, normalised to the expectation, is found to be $`0.993\pm 0.16`$. The observed spectrum from 1993 is shown in Figure 4 and good agreement is found with the Monte Carlo expectation. ## 6 $`𝐞^\mathbf{+}𝐞^{\mathbf{}}\mathbf{}\mathrm{𝐡𝐚𝐝𝐫𝐨𝐧𝐬}\mathbf{(}𝜸\mathbf{)}`$ #### Event Selection Hadronic $`\mathrm{Z}`$ decays are identified by their large energy deposition and high multiplicity in the electromagnetic and hadron calorimeters. The selection criteria are similar to those applied in our previous analysis : 1. The total energy observed in the detector, $`E_{\mathrm{vis}}`$, normalised to the centre-of-mass energy must satisfy $`0.5<E_{\mathrm{vis}}/\sqrt{s}<2.0`$; 2. The energy imbalance along the beam direction, $`E_{}`$, must satisfy $`|E_{}|/E_{\mathrm{vis}}<0.6`$; 3. The transverse energy imbalance, $`E_{}`$, must satisfy $`E_{}/E_{\mathrm{vis}}<0.6`$; 4. The number of clusters, $`N_{\mathrm{cl}}`$, formed from energy depositions in the calorimeters is required to be: a) $`N_{\mathrm{cl}}13`$ for $`|\mathrm{cos}\theta _t|0.74`$ (barrel region), b) $`N_{\mathrm{cl}}17`$ for $`|\mathrm{cos}\theta _t|>0.74`$ (end-cap region), where $`\theta _t`$ is the polar angle of the event thrust axis. Detailed analyses of the large data samples collected have been used to improve the Monte Carlo simulation of the detector response. Figures 5 to 9 show the distributions of the quantities used to select hadronic $`\mathrm{Z}`$ decays and the comparisons to the Monte Carlo predictions. In these plots all selection cuts are applied, except the one under study. Good agreement is observed between our data and the Monte Carlo simulations. #### Total Cross Section The acceptance for $`\mathrm{e}^+\mathrm{e}^{}\mathrm{hadrons}(\gamma )`$ events is determined from large samples of Monte Carlo events generated with the JETSET program. Applying the selection cuts, between $`99.30\%`$ and $`99.42\%`$ of the events are accepted depending on the year of the data taking and on differences in initial-state photon radiation at the various centre-of-mass energies. Monte Carlo events are generated with $`\sqrt{s^{}}>0.1\sqrt{s}`$ where $`\sqrt{s^{}}`$ is the effective centre-of-mass energy after initial state photon radiation. The acceptance for events in the data with $`\sqrt{s^{}}0.1\sqrt{s}`$ is estimated to be negligible. They are not considered as part of the signal and hence not corrected for. The interference between initial and final state photon radiation is not accounted for in the event generator. This effect modifies the angular distribution of the events in particular at very low polar angles where the detector inefficiencies are largest. However, the error from the imperfect simulation on the measured cross section, which includes initial-final state interference as part of the signal, is estimated to be very small ($`0.1\mathrm{pb}`$) in the centre-of-mass energy range considered here. Quark pairs originating from pair production from initial state radiation are considered as part of the signal if their invariant mass exceeds 50% of $`\sqrt{s}`$. To estimate the uncertainty on the acceptance on the modelling of the quark fragmentation, the determination of the acceptance is repeated using the HERWIG program. The detector simulations of both Monte Carlo programs are tuned in the same way to describe as closely as possible our data, e.g. in terms of energy resolution and cluster multiplicity. The remaining difference in acceptance is $`0.42`$ and we assign half of it as an estimate of the uncertainty on the acceptance of $`\mathrm{e}^+\mathrm{e}^{}\mathrm{hadrons}(\gamma )`$ events due to the modelling of quark fragmentation. Differences of the implementation of QED effects in both programs are studied and found to have negligible impact on the acceptance. Hadronic $`\mathrm{Z}`$ decays are triggered by the energy, central track, muon or scintillation counter multiplicity triggers. The combined trigger efficiency is obtained from the fraction of events with one of these triggers missing as a function of the polar angle of the event thrust axis. This takes into account most of the correlations among triggers. A sizeable inefficiency is only observed for events in the very forward region of the detector, where hadrons can escape through the beam pipe. Trigger efficiencies, including all steps of the trigger system, between $`99.829\%`$ and $`99.918\%`$ are obtained for the various data sets. Trigger inefficiencies determined for data sets taken in the same year are statistically compatible. Combining those data sets results in statistical errors of at most $`0.12`$ which is assigned as systematic error to all data sets. The background from other $`\mathrm{Z}`$ decays is found to be small: $`2.9`$ essentially only from $`\mathrm{e}^+\mathrm{e}^{}\tau ^+\tau ^{}(\gamma )`$. The uncertainty on this number is negligible compared to the total systematic error. The determination of the non-resonant background, mainly $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}\mathrm{hadrons}`$, is based on the measured distribution of the visible energy shown in Figure 5. The Monte Carlo program PHOJET is used to simulate two-photon collision processes. The absolute cross section is derived by scaling the Monte Carlo to obtain the best agreement with our data in the low end of the $`E_{\mathrm{vis}}`$ spectrum: $`0.32E_{\mathrm{vis}}/\sqrt{s}0.44`$. Consistently for all data sets, scale factors of $`1.1`$ are necessary. In the signal region contaminations from $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}\mathrm{hadrons}`$ between $`11.6\mathrm{pb}`$ and $`13.0\mathrm{pb}`$ are obtained for the different data sets. No dependence on $`\sqrt{s}`$ is observed. This is in agreement with results of a similar calculation performed with the DIAG36 program. Beam related background (beam-gas and beam-wall interactions) is small. To the extent that the $`E_{\mathrm{vis}}`$ spectrum is similar to that of $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}\mathrm{hadrons}`$, it is accounted for by determining the absolute normalisation from the data. As a check, the non-resonant background is estimated by extrapolating an exponential dependence of the $`E_{\mathrm{vis}}`$ spectrum from the low energy part into the signal region. This method yields consistent results. Based on these studies we assign an error on the measured hadron cross section of $`3\mathrm{pb}`$ due to the understanding of the non-resonant background. This error assignment is supported by our measurements of the hadronic cross section at high energies ($`130\mathrm{GeV}\sqrt{s}172\mathrm{GeV}`$) where the relative contribution of two-photon processes is much larger . The extrapolation of these studies back to the $`\mathrm{Z}`$ peak yields a similar result for the uncertainty. The contribution of random uranium noise and electronic noise in the detector faking a signal event is determined from a subsample of the event candidates. This subsample is obtained requiring that most of the observed energy stems either from the electromagnetic or the hadron calorimeter and that there be little matching between individual energy deposits and tracks. The $`E_{}/E_{\mathrm{vis}}`$ distribution of this subsample shows an $`\mathrm{e}^+\mathrm{e}^{}\mathrm{hadrons}(\gamma )`$ signal over a flat background (see Figure 10 for the 1994 data). This background is consistent with a constant noise rate, from which a background correction of $`7.4\mathrm{pb}`$ is derived. An uncertainty of $`1\mathrm{pb}`$ on the hadron cross section is assigned to all data sets from this correction. The absolute normalisation of the $`\mathrm{e}^+\mathrm{e}^{}\mathrm{hadrons}(\gamma )`$ signal in Figure 10 is not expected to be perfectly reproduced by the Monte Carlo simulation. However, this does not pose a serious problem as the noise rate is determined from the tail of the spectrum. The systematic error from event selection on the measured cross sections is estimated by varying the selection cuts. All cross section results are stable within $`\pm 0.3`$. The systematic errors to the cross section measurements $`\mathrm{e}^+\mathrm{e}^{}\mathrm{hadrons}(\gamma )`$ are summarised in Table 7. Uncertainties which scale with the cross section and absolute uncertainties are separated because they translate in a different way into errors on $`\mathrm{Z}`$ parameters, in particular on the total width. The scale error is further split into a part uncorrelated among the data samples, in this case consisting of the contribution of Monte Carlo statistics, and the rest which is taken to be fully correlated and amounts to $`0.39`$. The results of the $`\mathrm{e}^+\mathrm{e}^{}\mathrm{hadrons}(\gamma )`$ cross section measurements are discussed in Section 10. ## 7 $`𝐞^\mathbf{+}𝐞^{\mathbf{}}\mathbf{}𝝁^\mathbf{+}𝝁^{\mathbf{}}\mathbf{(}𝜸\mathbf{)}`$ #### Event Selection The selection of $`\mathrm{e}^+\mathrm{e}^{}\mu ^+\mu ^{}(\gamma )`$ in the 1993 and 1994 data is similar to the selection applied in previous years described in Reference . Two muons in the polar angular region $`|\mathrm{cos}\theta |<0.8`$ are required. Most of the muons, 88%, are identified by a reconstructed track in the muon spectrometer. Muons are also identified by their minimum ionising particle (MIP) signature in the inner sub-detectors, if less than two muon chamber layers are hit. A muon candidate is denoted as a MIP, if at least one of the following conditions is fulfilled: 1. A track in the central tracking chamber must point within $`5^{}`$ in azimuth to a cluster in the electromagnetic calorimeter with an energy less than 2 $`\mathrm{GeV}`$. 2. On a road from the vertex through the barrel hadron calorimeter, at least five out of a maximum of 32 cells must be hit, with an average energy of less than 0.4 $`\mathrm{GeV}`$ per cell. 3. A track in the central chamber or a low energy electromagnetic cluster must point within $`10^{}`$ in azimuth to a muon chamber hit. In addition, both the electromagnetic and the hadronic energy in a cone of $`12^{}`$ half-opening angle around the MIP candidate, corrected for the energy loss of the particle, must be less than 5 $`\mathrm{GeV}`$. Events of the reaction $`\mathrm{e}^+\mathrm{e}^{}\mu ^+\mu ^{}(\gamma )`$ are selected by the following criteria: 1. The event must have a low multiplicity in the calorimeters $`N_{\mathrm{cl}}15`$. 2. If at least one muon is reconstructed in the muon chambers, the maximum muon momentum must satisfy $`p_{\mathrm{max}}>0.6E_\mathrm{b}`$. If both muons are identified by their MIP signature there must be two tracks in the central tracking chamber with at least one with a transverse momentum larger than $`3\mathrm{GeV}`$. 3. The acollinearity angle $`\xi `$ must be less than $`90^{}`$, $`40^{}`$ or $`5^{}`$ if two, one or no muons are reconstructed in the muon chambers. 4. The event must be consistent with an origin of an $`\mathrm{e}^+\mathrm{e}^{}`$-interaction requiring at least one time measurement of a scintillation counter, associated to a muon candidate, to coincide within $`\pm 3\mathrm{ns}`$ with the beam crossing. Also, there must be a track in the central tracking chamber with a distance of closest approach to the beam axis of less than $`5\mathrm{mm}`$. As an example, Figure 11 shows the distribution of the maximum measured muon momentum for candidates in the $`199394`$ data compared to the expectation for signal and background processes. The acollinearity angle distribution of the selected muon pairs is shown in Figure 12. The experimental angular resolution and radiation effects are well reproduced by the Monte Carlo simulation. The analysis of the 1995 data in addition uses the newly installed forward-backward muon chambers. The fiducial volume is extended to $`|\mathrm{cos}\theta |<`$ 0.9. Each event must have at least one track in the central tracking chamber with a distance of closest approach in the transverse plane of less than $`1\mathrm{mm}`$ and a scintillation counter time coinciding within $`\pm 5\mathrm{ns}`$ with the beam crossing. The rejection of cosmic ray muons in the 1995 data is illustrated in Figure 13. For events with muons reconstructed in the muon chambers the maximum muon momentum must be larger than $`\frac{2}{3}E_\mathrm{b}`$. Every muon without a reconstructed track in the muon chambers must have a transverse momentum larger than $`3\mathrm{GeV}`$ as measured in the central tracking chamber. The polar angle distribution of muon pairs collected in 1995 is shown in Figure 14. #### Total Cross Section The acceptance for the process $`\mathrm{e}^+\mathrm{e}^{}\mu ^+\mu ^{}(\gamma )`$ in the fiducial volume $`|\mathrm{cos}\theta |<0.8`$ (0.9 for 1995 data) and for $`\xi <90^{}`$ is determined with events generated with the KORALZ program. We obtain acceptances between $`92.25\%`$ and $`93.04\%`$, mainly depending on the centre-of-mass energy. The systematic error on the cross section from imperfect description of detector inefficiencies is estimated to be $`2.7`$ (3.2 for the 1995 data). This number is calculated from a comparison with results obtained by removing events at the detector edges from the analysis and using different descriptions of time dependent detector inefficiencies. Smaller contributions to the systematic error arise from the statistical precision of the Monte Carlo simulations performed for the different data samples. Muon pairs are mainly triggered by the muon and the central track trigger. The trigger efficiencies are studied as a function of the azimuthal angle as inefficiencies are expected close to chamber boundaries. For the 1995 data also the polar angular dependence of the trigger efficiency is determined to account for effects in the forward region. Events with both muons reconstructed in the muon chambers are triggered with full efficiency. The efficiency of the central track trigger is independently determined using Bhabha events. The overall trigger efficiency varies between $`99.62\%`$ and $`99.90\%`$ for the different years of data taking. Systematic errors on the measured cross sections of less than $`1`$ are estimated from comparing a simulation of the central track trigger efficiency and its measurement with Bhabha events. A background of $`(1.35\pm 0.03)\%`$ remains in the sample arising from $`\mathrm{e}^+\mathrm{e}^{}\tau ^+\tau ^{}(\gamma )`$ events with both tau leptons decaying into muons. The error reflects Monte Carlo statistics and the uncertainty of the branching ratio $`\tau ^{}\mu ^{}\overline{\nu }_\mu \nu _\tau `$ . Other backgrounds from $`\mathrm{Z}`$ decays are smaller than $`0.1`$. The contamination from the non-resonant two-photon process $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}\mu ^+\mu ^{}`$ is $`0.11\mathrm{pb}`$, i.e. between $`0.1`$ and $`0.3`$ of the signal cross section, as determined using the DIAG36 Monte Carlo program. The residual contamination from cosmic ray muons in the event sample is determined from the sideband in the distribution of distance of closest approach to the beam axis after all other selection cuts are applied (Figure 13). Cosmic ray muons enter into the event sample at a rate of $`(9.7\pm 0.8)10^4`$ per minute of data taking which translates to background contaminations between $`1.9`$ and $`6.8`$ for the different data sets depending on their average instantaneous luminosity and the signal cross section. The statistical precision of the determination of the cosmic contamination causes a systematic error of $`0.3\mathrm{pb}`$ on the total muon pair cross section. By varying the selection cuts we determine systematic errors on the total cross section between $`1.3`$ and $`2.2`$. The systematic errors on the cross section measurements $`\mathrm{e}^+\mathrm{e}^{}\mu ^+\mu ^{}(\gamma )`$ are summarised in Table 8. Resonant four-fermion final states with a high-mass muon pair and a low-mass fermion pair are accepted. These events are considered as part of the signal if the invariant mass of the muon pair exceeds $`0.5\sqrt{s}`$. This inclusive selection minimizes errors due to higher order radiative corrections. Especially no cut is applied on additional tracks from low-mass fermion pairs in the final state . #### Forward-Backward Asymmetry The forward-backward asymmetry, $`A_{\mathrm{FB}}`$, is defined as: $$A_{\mathrm{FB}}=\frac{\sigma _\mathrm{F}\sigma _\mathrm{B}}{\sigma _\mathrm{F}+\sigma _\mathrm{B}},$$ (2) where $`\sigma _\mathrm{F}`$ is the cross section for events with the fermion scattered into the hemisphere which is forward with respect to the $`\mathrm{e}^{}`$ beam direction. The cross section in the backward hemisphere is denoted by $`\sigma _\mathrm{B}`$. Events with hard photon bremsstrahlung are removed from the sample by requiring that the acollinearity angle of the event be less than $`15^{}`$. The differential cross section in the angular region $`|\mathrm{cos}\theta |<0.9`$ can then be approximated by the lowest order angular dependence to sufficient precision: $$\frac{\mathrm{d}\sigma }{\mathrm{d}\mathrm{cos}\theta }\frac{3}{8}\left(1+\mathrm{cos}^2\theta \right)+A_{\mathrm{FB}}\mathrm{cos}\theta ,$$ (3) with $`\theta `$ being the polar angle of the final state fermion with respect to the $`\mathrm{e}^{}`$ beam direction. For each data set the forward-backward asymmetry is determined from a maximum likelihood fit to our data where the likelihood function is defined as the product over the selected events labelled $`i`$ of the differential cross section evaluated at their respective scattering angle $`\theta _i`$: $$L=\underset{i}{}\left(\frac{3}{8}\left(1+\mathrm{cos}^2\theta _i\right)+\left(12\kappa _i\right)A_{\mathrm{FB}}\mathrm{cos}\theta _i\right).$$ (4) The probability of charge confusion for a specific event, $`\kappa _i`$, is included in the fit. Only events with opposite charge assignment to the two muons are used for this measurement. The bias on the asymmetry measurement introduced by the use of the lowest order angular dependence (Equation 3) does not exceed 0.0003. This method does not require an exact knowledge of the acceptance as a function of the polar angle provided that the acceptance is independent of the muon charge. Events without a reconstructed muon in the muon chambers are included with the charge assignment obtained from the central tracking chamber in a similar way as for $`\mathrm{e}^+\mathrm{e}^{}`$ final states . This largely reduces effects of charge dependent acceptance in the muon chambers. The remaining asymmetry is estimated by artificially symmetrising the detector. For each known, inefficient detector element, the element opposite with respect to the centre of the detector is removed from the data reconstruction. The event selection is applied again and, for the large 1994 data set, the measured forward-backward asymmetry changes by $`0.0011\pm 0.0006`$. Half of this difference, $`0.0006`$, is assigned to all data sets as a systematic error on $`A_{\mathrm{FB}}`$ from a possible detector asymmetry. In 1995 the forward-backward muon chambers did not contribute significantly to the detector asymmetry. The values of $`\kappa _i`$ are obtained from the fraction of events with identical charges assigned to both muons. Besides its dependence on the transverse momentum, the charge measurement strongly depends on the number of muon chamber layers used in the reconstruction. The charge confusion is determined for each event class individually. The average charge confusion probability, almost entirely caused by muons only measured in the central tracking chamber, is $`(3.2\pm 0.3)`$, $`(0.8\pm 0.1)`$ and $`(1.0\pm 0.3)`$ for the years 1993, 1994 and 1995, respectively, where the errors are statistical. The improvement in the charge determination for 1994 and 1995 reflects the use of the silicon microvertex detector. The correction for charge confusion is proportional to the forward-backward asymmetry and it is less than $`0.001`$ for all data sets. To estimate a possible bias from a preferred orientation of events with the two muons measured to have the same charge we determine the forward-backward asymmetry of these events using the track with a measured momentum closer to the beam energy. The asymmetry of this subsample is statistically consistent with the standard measurement. Including these like-sign events in the 1994 sample would change the measured asymmetry by $`0.0008`$. Half of this number is taken as an estimate of a possible bias of the asymmetry measurement from charge confusion in the $`199394`$ data. The same procedure is applied to the 1995 data and the statistical precision limits a possible bias to $`0.0010`$. Differences of the momentum reconstruction in forward and backward events would cause a bias of the asymmetry measurement because of the requirement on the maximum measured muon momentum. We determine the loss of efficiency due to this cut separately for forward and backward events by selecting muon pairs without cuts on the reconstructed momentum. No significant difference is observed and the statistical error of this comparison limits the possible effect on the forward-backward asymmetry to be less than $`0.0004`$ and $`0.0009`$ for the $`199394`$ and 1995 data, respectively. Other possible biases from the selection cuts on the measurement of the forward-backward asymmetry are negligible. This is verified by a Monte Carlo study which shows that events not selected for the asymmetry measurement, but inside the fiducial volume and with $`\xi <15^{}`$, do not have a different $`A_{\mathrm{FB}}`$ value. The background from $`\mathrm{e}^+\mathrm{e}^{}\tau ^+\tau ^{}(\gamma )`$ events is found to have the same asymmetry as the signal and thus neither necessitates a correction nor causes a systematic uncertainty. The effect of the contribution from the two-photon process $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}\mu ^+\mu ^{}`$, further reduced by the tighter acollinearity cut on the measured muon pair asymmetry, can be neglected. The forward-backward asymmetry of the cosmic ray muon background is measured to be $`0.02\pm 0.13`$ using the events in the sideband of the distribution of closest approach to the interaction point. Weighted by the relative contribution to the data set this leads to corrections of $`0.0007`$ and $`+0.0003`$ to the peak$`2`$ and peak$`+2`$ asymmetries, respectively. On the peak this correction is negligible. The statistical uncertainty of the measurement of the cosmic ray asymmetry causes a systematic error of $`0.0001`$ on the peak and between $`0.0003`$ and $`0.0005`$ for the peak$`2`$ and peak$`+2`$ data sets. The systematic uncertainties on the measurement of the muon forward-backward asymmetry are summarised in Table 9. In $`199394`$ the total systematic error amounts to $`0.0008`$ at the peak points and to $`0.0009`$ at the off-peak points due to the larger contamination of cosmic ray muons. For the 1995 data the determination of systematic errors is limited by the number of events taken with the new detector configuration and the total error is estimated to be $`0.0015`$. In Figure 15 the differential cross sections $`\mathrm{d}\sigma /\mathrm{d}\mathrm{cos}\theta `$ measured from the $`199395`$ data sets are shown for three different centre-of-mass energies. The data are corrected for detector acceptance and charge confusion. Data sets with a centre-of-mass energy close to $`m_\mathrm{Z}`$, as well as the data at peak$`2`$ and the data at peak$`+2`$, are combined. The data are compared to the differential cross section shape given in Equation 3. The results of the total cross section and forward-backward asymmetry measurements in $`\mathrm{e}^+\mathrm{e}^{}\mu ^+\mu ^{}(\gamma )`$ are presented in Section 10. ## 8 $`𝐞^\mathbf{+}𝐞^{\mathbf{}}\mathbf{}𝝉^\mathbf{+}𝝉^{\mathbf{}}\mathbf{(}𝜸\mathbf{)}`$ #### Event Selection The selection of $`\mathrm{e}^+\mathrm{e}^{}\tau ^+\tau ^{}(\gamma )`$ events aims to select all hadronic and leptonic decay modes of the tau. $`\mathrm{Z}`$ decays into tau leptons are distinguished from other $`\mathrm{Z}`$ decays by the lower visible energy due to the presence of neutrinos and the lower particle multiplicity as compared to hadronic $`\mathrm{Z}`$ decays. Compared to our previous analysis the selection of $`\mathrm{e}^+\mathrm{e}^{}\tau ^+\tau ^{}(\gamma )`$ events is extended to a larger polar angular range, $`|\mathrm{cos}\theta _t|0.92`$, where $`\theta _t`$ is defined by the thrust axis of the event. Event candidates are required to have a jet, constructed from calorimetric energy deposits and muon tracks, with an energy of at least $`8\mathrm{GeV}`$. Energy deposits in the hemisphere opposite to the direction of this most energetic jet are combined to form a second jet. The two jets must have an acollinearity angle $`\xi <10^{}`$. There is no energy requirement on the second jet. High multiplicity hadronic $`\mathrm{Z}`$ decays are rejected by allowing at most three tracks matched to any of the two jets. In each of the two event hemispheres there should be no track with an angle larger than $`18^{}`$ with respect to the jet axis. Resonant four-fermion final states with a high mass tau pair and a low mass fermion pair are mostly kept in the sample. The multiplicity cut affects only tau decays into three charged particles with the soft fermion close in space leading to corrections of less than $`1`$. If the energy in the electromagnetic calorimeter of the first jet exceeds $`85\%`$, or the energy of the second jet exceeds $`80\%`$, of the beam energy with a shape compatible with an electromagnetic shower the event is classified as $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}(\gamma )`$ background and hence rejected. Background from $`\mathrm{e}^+\mathrm{e}^{}\mu ^+\mu ^{}(\gamma )`$ is removed by requiring that there be no isolated muon with a momentum larger than 80% of the beam energy and that the sum of all muon momenta does not exceed $`1.5E_\mathrm{b}`$. Events are rejected if they are consistent with the signature of two MIPs. To suppress background from cosmic ray events the time of scintillation counter hits associated to muon candidates must be within $`\pm 5\mathrm{ns}`$ of the beam crossing. In addition, the track in the muon chambers must be consistent with originating from the interaction point. In Figures 16 to 19 the energy in the most energetic jet, the number of tracks associated to both jets, the acollinearity between the two jets and the distribution of $`|\mathrm{cos}\theta _t|`$ are shown for the 1994 data. Data and Monte Carlo expectations are compared after all cuts are applied, except the one under study. Good agreement between data and Monte Carlo is observed. Small discrepancies seen in Figure 17 are due to the imperfect description of the track reconstruction efficiency in the central chamber. Their impact on the total cross section measurement is small and is included in the systematic error given below. Tighter selection cuts must be applied in the region between barrel and end-cap part of the BGO calorimeter and in the end-cap itself, reducing the selection efficiency (see Figure 19). This is due to the increasing background from Bhabha scattering. Most importantly the shower shape in the hadron calorimeter is also used to identify candidate electrons and the cuts on the energy of the first and second jet in the electromagnetic end-cap calorimeter are tightened to 75% of the beam energy. #### Total Cross Section Between $`70.21\%`$ and $`70.91\%`$ of the signal events are accepted inside the fiducial volume defined by $`|\mathrm{cos}\theta _t|0.92`$. The acceptance for $`\mathrm{e}^+\mathrm{e}^{}\tau ^+\tau ^{}(\gamma )`$ events depends on the tau decay products. The experimental knowledge of tau branching fractions translates to an uncertainty on the average acceptance of $`\mathrm{e}^+\mathrm{e}^{}\tau ^+\tau ^{}(\gamma )`$ events which contributes with $`2`$ to the systematic error on the cross section measurement. From the data the efficiency of the trigger system for selected $`\mathrm{e}^+\mathrm{e}^{}\tau ^+\tau ^{}(\gamma )`$ events is determined to be $`(99.71\pm 0.02)\%`$. The largest remaining background consists of Bhabha events, $`1.3\%`$ to $`3.7\%`$, depending on the centre-of-mass energy, entering into the sample predominantly at low polar angles. Background from $`\mathrm{Z}`$ decays into hadrons is determined to be between $`1.3`$ and $`2.7`$, depending on the data taking period, and $`7.5`$ from $`\mathrm{Z}`$ decays into muons. The statistical precision of the background determination by Monte Carlo simulations causes systematic errors between $`1.0`$ and $`3.3`$. Contaminations from non-resonant background are small: $`1`$ to $`2`$ from two-photon collisions and $`2`$ to $`3`$ from cosmic ray muons, depending on the centre-of-mass energy. The systematic error from the subtraction of non-resonant background is estimated to be $`1.2\mathrm{pb}`$. From variations of the above selection cuts contributions to the systematic error on the total cross section between $`5.3`$ and $`8.0`$ are estimated for different years, largely independent of the centre-of-mass energy. The main contribution arises from the definition of the fiducial volume by $`|\mathrm{cos}\theta _t|0.92`$, see Figure 19. The systematic errors on the $`\mathrm{e}^+\mathrm{e}^{}\tau ^+\tau ^{}(\gamma )`$ cross section measurements are summarised in Table 10. #### Forward-Backward Asymmetry The forward-backward asymmetry of $`\mathrm{e}^+\mathrm{e}^{}\tau ^+\tau ^{}(\gamma )`$ events is determined in the same way as described for muon pairs (Equation 4). The charge of a tau is derived from the sum of the charges of its decay products as measured in the central tracking and the muon chambers. The event sample selected for the cross section measurement is used requiring opposite and unit charge for the two tau jets. The average probability for a mis-assignment of both charges as determined from the ratio of like and unlike sign events is $`(7.4\pm 0.4)`$ in 1993. The use of the silicon microvertex detector reduced this mis-assignment to $`(2.5\pm 0.1)`$ and $`(1.3\pm 0.1)`$ in 1994 and 1995. Because the charge confusion probability is approximately independent of the polar angle this average value is used in the fit for $`A_{\mathrm{FB}}`$. The systematic error on the forward-backward asymmetry from the uncertainty in the determination and the treatment of the charge confusion probability is estimated to be less than $`0.0001`$ for all data sets. The effect of a possible detector asymmetry, in particular at the edges of the fiducial volume, is estimated from variation of the $`\mathrm{cos}\theta _t`$ cut. The statistical accuracy of this test limits this uncertainty to $`0.003`$ which is taken as a systematic error. The measured asymmetries are corrected for background contributions. The uncertainty on the background contamination, in particular from $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}(\gamma )`$, translates into an error of $`0.001`$ on the tau pair asymmetry. Large Monte Carlo samples are used to study a possible bias on the measured asymmetry from the fit method and from the selection cuts. In particular, energy and momentum requirements might preferentially select certain helicity configurations leading to a bias in the determination of $`A_{\mathrm{FB}}`$. The Monte Carlo simulation does not show evidence for such a bias and its statistical precision, $`0.0004`$, is taken as the systematic error. During the 1995 data taking, large shifts of the longitudinal position of the $`\mathrm{e}^+\mathrm{e}^{}`$-interaction point were observed caused by the reconfiguration of the LEP radio frequency system . However, they are found to have no sizeable effect on the measurement of the forward-backward asymmetry. The total systematic error assigned to the forward-backward asymmetry measurement of tau pairs is $`0.0032`$ (Table 11). It is fully correlated between the data sets. The measured differential cross sections, combining the data into three centre-of-mass energy points, are shown in Figure 20. The lines show the results of fits to the data using the functional form of Equation 3. Section 10 presents the measurements of the total cross section and the forward-backward asymmetry in $`\mathrm{e}^+\mathrm{e}^{}\tau ^+\tau ^{}(\gamma )`$. ## 9 $`𝐞^\mathbf{+}𝐞^{\mathbf{}}\mathbf{}𝐞^\mathbf{+}𝐞^{\mathbf{}}\mathbf{(}𝜸\mathbf{)}`$ #### Event Selection The analysis of the reaction $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}(\gamma )`$ is restricted to the polar angular range $`44^{}<\theta <136^{}`$ to increase the relative contribution of $`\mathrm{Z}`$ exchange to the measured cross section. The signature of $`\mathrm{e}^+\mathrm{e}^{}`$ final states is the low multiplicity high energy deposition in the electromagnetic calorimeter with associated tracks in the central tracking chamber. Most of the events are selected by requiring at least two clusters in the fiducial volume of the electromagnetic calorimeter, one with an energy greater than $`0.9E_\mathrm{b}`$ and the other with more than $`2\mathrm{GeV}`$. The polar angles are determined form the centre-of-gravity of the clusters in the calorimeter and the interaction point. Figure 21 shows the distribution of the highest energy cluster, $`E_1`$, normalised to the beam energy for events which pass all cuts except the requirement on the most energetic cluster. Electrons are discriminated from photons by requiring five out of 62 anodes of the central tracking chamber with a hit matching in azimuthal angle within $`\pm 3^{}`$ with the cluster in the calorimeter. Two electron candidates are required inside the fiducial volume and with an acollinearity angle $`\xi <25^{}`$. Figure 22 shows the distribution of the acollinearity angle. All other cuts except the one under study are applied. The event selection depends on the exact knowledge of imperfections of the electromagnetic calorimeter. The impact of the discrepancies seen in Figure 21 around the cut value is significantly reduced by accepting also events without a second cluster in the fiducial volume of the electromagnetic calorimeter. In this case a cluster in the hadron calorimeter is required consistent with an electromagnetic shower shape and at least $`7.5\mathrm{GeV}`$ opposite to the leading BGO cluster. This recovers events, up to $`4`$ of the total sample, with electrons leaking through the BGO support structure. Events failing the requirement on the most energetic cluster in the electromagnetic calorimeter are accepted if the sum of the energies of the four highest energy clusters anywhere in the electromagnetic calorimeter is larger than $`70\%`$ of the centre-of-mass energy. In addition this partially recovers radiative events. For all event candidates the total number of energy deposits, $`N_{\mathrm{cl}}`$, must be less than $`15`$ ($`12`$ for 1995 data). #### Total Cross Section The selection efficiency is determined using Monte Carlo events generated with the program BHAGENE, which generates up to three photons. Efficiencies between 97.37% and 98.53% are obtained for the different samples, where the differences originate from time dependent detector inefficiencies. The use of high multiplicity hadron events allows to monitor the status of each individual BGO crystal in short time intervals. Inefficient crystals, typically 100 out of 8000 in the barrel part, are identified and taken into account in the Monte Carlo simulation. This method, together with the redundancy of the selection cuts, reduces the systematic error on the selection efficiency. Limited Monte Carlo statistics causes systematic errors between $`0.4`$ and $`1.0`$. The calculation of the selection efficiency is checked using events generated with the programs BABAMC and BHWIDE. The efficiencies calculated with the different event generators agree within $`\pm 1`$ which is taken as an estimate of the systematic error. The efficiency of the electron and photon discrimination in the central tracking chamber is determined using a subsample of data events selected by a tight acollinearity cut ($`\xi <1^{}`$) and requiring two high energy clusters in the electromagnetic calorimeter ($`E>30\mathrm{GeV}`$). Here the contamination of $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}\gamma `$ events with one electron and the photon inside, and the other electron outside the fiducial volume is expected to be very small. In this sample, events with only one identified electron originate from mis-identified Bhabha events or from photon conversion of $`\mathrm{e}^+\mathrm{e}^{}\gamma \gamma `$ events. The contamination of the latter in this sample is $`0.4`$ to $`0.9`$ as calculated from Monte Carlo. After correction for this contamination, the probability that one of the electrons in $`\mathrm{e}^+\mathrm{e}^{}`$ final states fails the electron-photon discrimination is measured to be $`(0.7\pm 0.3)`$ and $`(1.0\pm 0.1)`$ for the 1993 and 1994 data, respectively. We correct for this effect. The method to determine this probability from the data is checked on fully simulated $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}(\gamma )`$ Monte Carlo events. Firstly by not applying the electron-photon discrimination, the contamination of events in the data used for the cross section measurement with one photon and only one electron in the fiducial volume is determined to be $`(2.6\pm 0.5)`$. This is in reasonable agreement with the Monte Carlo prediction of $`1.4`$. Then we apply the above method to determine the probability that an electron fails the electron-photon discrimination on the fully simulated events and compare it to the value obtained using the generator information. The result is consistent within $`0.6`$ which is assigned as a systematic error to the total cross section due to the simulation and determination of the electron-photon discrimination. In 1995 the quality criteria on the status of the central tracking chamber are relaxed to increase the data sample at the expense of a smaller efficiency on the electron identification and a larger systematic error. Between 1.9 and 2.8 of the electrons fail the electron-photon discrimination cuts as determined from Monte Carlo simulation. We correct for this effect and a systematic error of 1.5 is assigned to the total cross section measurement. Large angle Bhabha scattering events are triggered by the energy and the central track triggers. The overall trigger inefficiency is found to be $`0.1`$ and has a negligible effect on the cross section measurement. In the 1993 and 1994 data the longitudinal position of the $`\mathrm{e}^+\mathrm{e}^{}`$ interaction point is stable within $`\pm 2\mathrm{mm}`$. The corresponding uncertainty on the definition of the fiducial volume translates to a systematic error of $`0.5`$ on the cross section measurement. Imperfections of the description of the BGO geometry and the shower shape of electrons lead to a possible difference of the definition of the polar angle between data and Monte Carlo simulation. This difference is found to be less than $`0.1^{}`$, translating to a systematic error of $`0.5`$ on the cross section measurement. The large movements of the interaction point in 1995 are determined from our $`\mathrm{e}^+\mathrm{e}^{}\mathrm{hadrons}(\gamma )`$ data and the positions are used to calculate the scattering angle in $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}(\gamma )`$ events. The remaining systematic uncertainty on the definition of the fiducial volume, including the description of the BGO geometry, is estimated from a variation of the cut on the polar angle to be $`1.5`$. The selected sample contains about $`1\%`$ background from the process $`\mathrm{e}^+\mathrm{e}^{}\tau ^+\tau ^{}(\gamma )`$, only slightly depending on the centre-of-mass energy. Contaminations from hadronic Z decays and the process $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}`$ are below $`1`$ and the remaining background from $`\mathrm{e}^+\mathrm{e}^{}\gamma \gamma `$ is negligible. The error on the total cross section from background subtraction is $`0.4`$ to $`1.0`$ originating from limited Monte Carlo statistics. The systematic uncertainty of the event selection, estimated from variations of the selection cuts around their nominal values, varies between $`0.8`$ and $`2.7`$ for the various data sets. The systematic uncertainties contributing to the measurement of the cross section $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}(\gamma )`$ are summarised in Table 12. #### Forward-Backward Asymmetry The data sample for the forward-backward asymmetry measurement is obtained from the sample used for the measurement of the total cross section requiring in addition that each of the two electron candidates match with a track within 25 mrad in azimuth. The charge determination of the electrons is described in detail in reference . The charge confusion is measured with the $`\mathrm{e}^+\mathrm{e}^{}\mu ^+\mu ^{}(\gamma )`$ data sample which has an independent charge measurement from the muon spectrometer. We obtain for the probability of a wrong event orientation values between 0.5% and 4.6%. Lower values are due to the exploitation of the silicon microvertex detector in 1994 and 1995. We determine the asymmetry of a subsample with much lower charge confusion by excluding events with tracks close to the cathode and anode planes of the central tracking chamber. Comparing these results to those obtained from the full sample we derive a systematic error on $`A_{\mathrm{FB}}`$ of 0.002 from the uncertainty of the charge determination. In the event sample used for the asymmetry measurement the main background from $`\mathrm{e}^+\mathrm{e}^{}\tau ^+\tau ^{}(\gamma )`$ is reduced to about $`4`$ because the tight requirement on the matching between tracks and clusters in the electromagnetic calorimeter removes $`\tau ^{}\rho ^{}\nu _\tau `$ decays present in the cross section sample. It induces a correction of $`0.002`$ on the asymmetry for the peak$`2`$ and of less than $`0.0005`$ for the other data sets. The effect is largest at peak$`2`$ because of the difference of the $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}(\gamma )`$ and $`\mathrm{e}^+\mathrm{e}^{}\tau ^+\tau ^{}(\gamma )`$ asymmetries. The uncertainty on the asymmetry measurement from background subtraction is estimated to be $`0.0005`$. The asymmetry is determined from the number of events observed in the forward and backward hemispheres, correcting for polar angle dependent efficiencies and background. The scattering angle is defined by the polar angle of the electron, $`\theta _\mathrm{e}^{}`$. The determination of the asymmetry is repeated defining the angle by the positron, $`\theta _{\mathrm{e}^+}`$, and taking the average of the two $`A_{\mathrm{FB}}`$ values. This reduces the sensitivity of the result to the size of the interaction region and its longitudinal offset. Alternatively, we determine the forward-backward asymmetry using the scattering angle in the rest system of the final state electron and positron: $$\mathrm{cos}\theta ^{}=\frac{\mathrm{sin}(\theta _{\mathrm{e}^+}\theta _\mathrm{e}^{})}{\mathrm{sin}\theta _\mathrm{e}^{}+\mathrm{sin}\theta _{\mathrm{e}^+}}.$$ (5) This definition minimises the sensitivity to photon emission. A Monte Carlo study shows that it differs by less than $`0.0005`$ from the above definition of $`A_{\mathrm{FB}}`$ due to different radiative corrections. After correcting for this difference in the data the two approaches yield forward-backward asymmetries consistent within $`0.0015`$ which is taken as an estimate of the remaining uncertainty of the scattering angle from the knowledge of the interaction point. The contributions of the systematic error on the asymmetry measurement are summarised in Table 13. The differential cross sections of the process $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}(\gamma )`$ at three different centre-of-mass energy points are shown in Figure 23 together with the prediction of the ALIBABA program. ## 10 Results on Total Cross Sections and Forward-Backward Asymmetries The results of the measurements of the total cross section performed between 1993 and 1995 in the four reactions $`\mathrm{e}^+\mathrm{e}^{}\mathrm{hadrons}(\gamma )`$, $`\mathrm{e}^+\mathrm{e}^{}\mu ^+\mu ^{}(\gamma )`$, $`\mathrm{e}^+\mathrm{e}^{}\tau ^+\tau ^{}(\gamma )`$ and $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}(\gamma )`$ are listed in Tables 14 to 17. The measured cross sections for $`\mathrm{e}^+\mathrm{e}^{}\mathrm{hadrons}(\gamma )`$ are corrected to the full solid angle for acceptance and efficiencies, keeping a lower cut on the effective centre-of-mass energy of $`\sqrt{s^{}}>0.1\sqrt{s}`$. The measured cross sections for muon and tau pairs are extrapolated to the full solid angle and the full phase space using ZFITTER. The quoted Bhabha cross sections are for both final state leptons inside the polar angular range $`44^{}<\theta <136^{}`$, with an acollinearity angle $`\xi <25^{}`$ and for a minimum energy of $`1\mathrm{GeV}`$ of the final state fermions. In Table 17 the $`s`$-channel contributions to the cross section extrapolated to the full phase space are also given. Their calculation is described in Appendix A and they can be compared to the measurements of the other leptonic final states (Tables 15 and 16). Results of the measurements performed between 1990 and 1992 are presented in Reference . Figures 24 to 27 compare the measurements of the total cross sections performed in $`199095`$ at the $`\mathrm{Z}`$ pole to the result of the fit to all cross section measurements imposing lepton universality described in section 12.1. For Bhabha scattering the contributions from the $`s`$\- and $`t`$-channels and their interference are displayed separately. Good agreement between measurements in different years is observed. The measurements of the forward-backward asymmetry performed between 1993 and 1995 in the leptonic reactions $`\mathrm{e}^+\mathrm{e}^{}\mu ^+\mu ^{}(\gamma )`$, $`\mathrm{e}^+\mathrm{e}^{}\tau ^+\tau ^{}(\gamma )`$ and $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}(\gamma )`$ are listed in Tables 18 to 20. For muon and tau pairs the results are extrapolated to the full solid angle keeping a cut on the acollinearity of $`\xi <15^{}`$ and $`\xi <10^{}`$, respectively. The measurements for the process $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}(\gamma )`$ apply to the same polar angular range and cuts as the total cross section. Table 20 contains also the $`s`$-channel contributions to the asymmetry (see Appendix A) to be compared to the measurements for muon and tau pairs. Figures 28 to 30 compare these measurements to the results of the fit to all hadronic and leptonic cross section and forward-backward asymmetry measurements imposing lepton universality. For the Bhabha scattering the difference of the forward and backward cross sections in the $`s`$\- and $`t`$-channels and in the interference, all normalised to the total cross section, are displayed separately. Good agreement between measurements in different years is observed. For the fits presented in the following sections we include the cross section and forward-backward asymmetry measurements from $`199092`$ . All our measurements at the Z resonance performed in the period $`199095`$ are self-consistent. Qualitatively this can be seen from Figure 31 where for all 175 measurements the absolute difference between the measurements and the expectations, divided by the statistical error of the measurements, is shown. The expected cross sections and forward-backward asymmetries are calculated from the result of the five parameter fit presented in Section 12.3. The scattering of our measurements is compared with the one expected from a perfect Gaussian distribution. The agreement is satisfactory considering that due to their complicated correlations, systematic errors cannot be taken into account in this comparison. ## 11 Fits for Electroweak Parameters Different analyses are used to extract electroweak parameters from the measured total cross sections and forward-backward asymmetries. Firstly, we determine the electroweak parameters making a minimum of assumptions about any underlying theory, for example the SM. The first analysis uses only the total cross section data to determine the parameters of the $`\mathrm{Z}`$ boson, its mass, the total and partial decay widths to fermion pairs. The second analysis also includes the asymmetry data, which allows the determination of the coupling constants of the neutral weak current. In a third analysis we fit the cross section and forward-backward asymmetry measurements in the S-Matrix ansatz where all contributions from $`\gamma /\mathrm{Z}`$-interference are determined from the data. Finally, all our measurements on electroweak observables are interpreted in the framework of the SM in order to determine its free parameters. #### Lowest Order Formulae In all analyses, a Breit-Wigner ansatz is used to describe the $`\mathrm{Z}`$ boson. The mass, $`m_\mathrm{Z}`$, and the total width, $`\mathrm{\Gamma }_\mathrm{Z}`$, of the Z boson are defined by the functional form of the Breit-Wigner denominator, which explicitly takes into account the energy dependence of the total width . The total $`s`$-channel cross section to lowest order, $`\sigma ^{}`$, for the process $`\mathrm{e}^+\mathrm{e}^{}\mathrm{f}\overline{\mathrm{f}}`$, is given by the sum of three terms, the $`\mathrm{Z}`$ exchange, $`\sigma _\mathrm{Z}^{}`$, the photon exchange, $`\sigma _\gamma ^{}`$, and the $`\gamma /\mathrm{Z}`$-interference, $`\sigma _{\mathrm{int}}^{}`$: $`\sigma ^{}`$ $`=`$ $`\sigma _\mathrm{Z}^{}+\sigma _\gamma ^{}+\sigma _{\mathrm{int}}^{}`$ $`\sigma _\mathrm{Z}^{}`$ $`=`$ $`{\displaystyle \frac{12\pi }{m_\mathrm{Z}^2}}{\displaystyle \frac{\mathrm{\Gamma }_\mathrm{e}\mathrm{\Gamma }_\mathrm{f}}{\mathrm{\Gamma }_\mathrm{Z}^2}}{\displaystyle \frac{s\mathrm{\Gamma }_\mathrm{Z}^2}{(sm_\mathrm{Z}^2)^2+s^2\mathrm{\Gamma }_\mathrm{Z}^2/m_\mathrm{Z}^2}}`$ $`\sigma _\gamma ^{}`$ $`=`$ $`{\displaystyle \frac{4\pi \alpha ^2}{3s}}q_\mathrm{e}^2q_\mathrm{f}^2N_\mathrm{C}^\mathrm{f}`$ $`\sigma _{\mathrm{int}}^{}`$ $`=`$ $`{\displaystyle \frac{4\pi \alpha ^2}{3}}J_\mathrm{f}{\displaystyle \frac{sm_\mathrm{Z}^2}{(sm_\mathrm{Z}^2)^2+s^2\mathrm{\Gamma }_\mathrm{Z}^2/m_\mathrm{Z}^2}},\mathrm{f}=\mathrm{e},\mu ,\tau ,\mathrm{q}`$ (6) where $`q_\mathrm{f}`$ is the electric charge of the final-state fermion, $`N_\mathrm{C}^\mathrm{f}`$ its colour factor, and $`\alpha `$ the electromagnetic coupling constant. The pure photon exchange is determined by QED. The first analysis treats the mass and the total and partial widths of the $`\mathrm{Z}`$ boson as free and independent parameters. The interference of the Z exchange with the photon exchange adds another parameter, the $`\gamma /\mathrm{Z}`$-interference term, $`J_\mathrm{f}`$, besides those corresponding to mass and widths of the $`\mathrm{Z}`$. Since in the SM $`|\sigma _{\mathrm{int}}^{}(s)|\sigma ^{}(s)`$ for centre-of-mass energies close to $`m_\mathrm{Z}`$, it is difficult to measure $`J_\mathrm{f}`$ accurately using data at the $`\mathrm{Z}`$ only. The $`\gamma /\mathrm{Z}`$-interference term is usually taken from the SM , thus making assumptions about the form of the electroweak unification. The second analysis determines the vector and axial-vector coupling constants of the neutral weak current to charged leptons, $`g_\mathrm{V}^{\mathrm{}}`$ and $`g_\mathrm{A}^{\mathrm{}}`$, by using the forward-backward asymmetries in addition to the total cross sections. In lowest order, for $`\sqrt{s}=m_\mathrm{Z}`$ and neglecting the photon exchange, the $`s`$-channel forward-backward asymmetry for the process $`\mathrm{e}^+\mathrm{e}^{}\mathrm{}^+\mathrm{}^{}(\gamma )`$ is given by: $`A_{\mathrm{FB}}^{0,\mathrm{}}`$ $`=`$ $`{\displaystyle \frac{3}{4}}A_\mathrm{e}A_{\mathrm{}}\mathrm{with}`$ $`A_{\mathrm{}}`$ $`=`$ $`{\displaystyle \frac{2g_\mathrm{V}^{\mathrm{}}g_\mathrm{A}^{\mathrm{}}}{(g_\mathrm{V}^{\mathrm{}})^2+(g_\mathrm{A}^{\mathrm{}})^2}}.`$ (7) The energy dependence of the asymmetry distinguishes $`g_\mathrm{V}^{\mathrm{}}`$ and $`g_\mathrm{A}^{\mathrm{}}`$ . The experimental precision on the coupling constants is improved by also including information from tau-polarisation measurements which determine $`A_\mathrm{e}`$ and $`A_\tau `$ independently. In Equation 6, the leptonic partial width, $`\mathrm{\Gamma }_{\mathrm{}}`$, and the leptonic $`\gamma /\mathrm{Z}`$-interference term, $`J_{\mathrm{}}`$, are now expressed in terms of $`g_\mathrm{V}^{\mathrm{}}`$ and $`g_\mathrm{A}^{\mathrm{}}`$: $`\mathrm{\Gamma }_{\mathrm{}}`$ $`=`$ $`{\displaystyle \frac{G_\mathrm{F}m_\mathrm{Z}^3}{6\sqrt{2}\pi }}\left[(g_\mathrm{V}^{\mathrm{}})^2+(g_\mathrm{A}^{\mathrm{}})^2\right]`$ $`J_{\mathrm{}}`$ $`=`$ $`{\displaystyle \frac{G_\mathrm{F}m_\mathrm{Z}^2}{\sqrt{2}\pi \alpha }}q_\mathrm{e}q_{\mathrm{}}g_\mathrm{V}^\mathrm{e}g_\mathrm{V}^{\mathrm{}},`$ (8) where $`G_\mathrm{F}`$ is the Fermi coupling constant. The hadronic cross section is given by the sum over the five kinematically allowed flavours and their colour states. Because no separation of quark flavours is attempted, this approach cannot be applied to the hadronic final state. Therefore, the parameterisation of the first analysis is used to express the hadronic cross section in terms of $`\mathrm{\Gamma }_{\mathrm{had}}`$ and $`J_{\mathrm{had}}`$. Our data are also interpreted in the framework of the S-Matrix ansatz , which makes a minimum of theoretical assumptions. This ansatz describes the hard scattering process of fermion-pair production in $`\mathrm{e}^+\mathrm{e}^{}`$-annihilations by the $`s`$-channel exchange of two spin-1 bosons, a massless photon and a massive $`\mathrm{Z}`$ boson. The lowest-order total cross section, $`\sigma _{\mathrm{tot}}^0`$, and forward-backward asymmetry, $`A_{\mathrm{FB}}^0`$, for $`\mathrm{e}^+\mathrm{e}^{}\mathrm{f}\overline{\mathrm{f}}`$ are given as: $`\sigma _\mathrm{a}^0(s)`$ $`=`$ $`{\displaystyle \frac{4}{3}}\pi \alpha ^2\left[{\displaystyle \frac{g_\mathrm{f}^\mathrm{a}}{s}}+{\displaystyle \frac{sr_\mathrm{f}^\mathrm{a}+\left(s\overline{m}_\mathrm{Z}^2\right)j_\mathrm{f}^\mathrm{a}}{\left(s\overline{m}_\mathrm{Z}^2\right)^2+\overline{m}_\mathrm{Z}^2\overline{\mathrm{\Gamma }}_\mathrm{Z}^2}}\right]\mathrm{for}\mathrm{a}=\mathrm{tot},\mathrm{FB}`$ $`A_{\mathrm{FB}}^0(s)`$ $`=`$ $`{\displaystyle \frac{3}{4}}{\displaystyle \frac{\sigma _{\mathrm{FB}}^0(s)}{\sigma _{\mathrm{tot}}^0(s)}}.`$ (9) The S-Matrix parameters $`r_\mathrm{f}^\mathrm{a}`$, $`j_\mathrm{f}^\mathrm{a}`$ and $`g_\mathrm{f}^\mathrm{a}`$ are real numbers which express the size of the $`\mathrm{Z}`$ exchange, $`\gamma /\mathrm{Z}`$-interference and photon exchange contributions. Here, $`r_\mathrm{f}^\mathrm{a}`$ and $`j_\mathrm{f}^\mathrm{a}`$ are treated as free parameters while the photon exchange contribution, $`g_\mathrm{f}^\mathrm{a}`$, is fixed to its QED prediction. Each final state is thus described by four free parameters: two for cross sections, $`r_\mathrm{f}^{\mathrm{tot}}`$ and $`j_\mathrm{f}^{\mathrm{tot}}`$, and two for forward-backward asymmetries, $`r_\mathrm{f}^{\mathrm{FB}}`$ and $`j_\mathrm{f}^{\mathrm{FB}}`$. In models with only vector and axial-vector couplings of the $`\mathrm{Z}`$ boson, these four S-Matrix parameters are not independent of each other: $`r_\mathrm{f}^{\mathrm{tot}}`$ $``$ $`\left[(g_\mathrm{V}^\mathrm{e})^2+(g_\mathrm{A}^\mathrm{e})^2\right]\left[(g_\mathrm{V}^\mathrm{f})^2+(g_\mathrm{A}^\mathrm{f})^2\right],`$ $`j_\mathrm{f}^{\mathrm{tot}}`$ $``$ $`g_\mathrm{V}^\mathrm{e}g_\mathrm{V}^\mathrm{f},`$ $`r_\mathrm{f}^{\mathrm{FB}}`$ $``$ $`g_\mathrm{A}^\mathrm{e}g_\mathrm{V}^\mathrm{e}g_\mathrm{A}^\mathrm{f}g_\mathrm{V}^\mathrm{f},`$ $`j_\mathrm{f}^{\mathrm{FB}}`$ $``$ $`g_\mathrm{A}^\mathrm{e}g_\mathrm{A}^\mathrm{f}.`$ (10) Under the assumption that only vector- and axial-vector couplings exist, the S-Matrix ansatz corresponds to the second analysis discussed above without fixing $`J_{\mathrm{had}}`$ to the SM. The S-Matrix ansatz is defined using a Breit-Wigner denominator with $`s`$-independent width for the $`\mathrm{Z}`$ resonance. To derive the mass and width of the $`\mathrm{Z}`$ boson for a Breit-Wigner with $`s`$-dependent width, the following transformations are applied : $`m_\mathrm{Z}=\overline{m}_\mathrm{Z}+34.1\mathrm{MeV}`$ and $`\mathrm{\Gamma }_\mathrm{Z}=\overline{\mathrm{\Gamma }}_\mathrm{Z}+0.9\mathrm{MeV}`$. #### Radiative Corrections The QED radiative corrections to the total cross sections and forward-backward asymmetries are included by convolution and by the replacement $`\alpha \alpha (s)=\alpha /(1\mathrm{\Delta }\alpha )`$ to account for the running of the electromagnetic coupling constant . Weak radiative corrections are calculated assuming the validity of the SM and as a function of the unknown mass of the Higgs boson. The coupling constants which are real to lowest order are modified by absorbing weak corrections and become complex quantities . Effective couplings, $`\overline{g}_\mathrm{V}^{\mathrm{}}`$ and $`\overline{g}_\mathrm{A}^{\mathrm{}}`$, are defined which correspond to the real parts. When extracting $`\overline{g}_\mathrm{V}^{\mathrm{}}`$ and $`\overline{g}_\mathrm{A}^{\mathrm{}}`$ from the measurements, the small imaginary parts are taken from the SM. Observables such as the leptonic partial widths (Equation 8) and the leptonic pole asymmetry (Equation 7) are redefined by replacing the vector and axial-vector coupling constants by these effective couplings. The effective couplings of fermions are expressed in terms of the effective electroweak mixing angle, $`\mathrm{sin}^2\overline{\theta }_\mathrm{W}`$, and the effective ratio of the neutral to charged weak current couplings, $`\overline{\rho }=1/(1\mathrm{\Delta }\overline{\rho })`$ : $`\overline{g}_\mathrm{V}^\mathrm{f}`$ $`=`$ $`\sqrt{\overline{\rho }}(I_3^\mathrm{f}2q_\mathrm{f}\mathrm{sin}^2\overline{\theta }_\mathrm{W})`$ $`\overline{g}_\mathrm{A}^\mathrm{f}`$ $`=`$ $`\sqrt{\overline{\rho }}I_3^\mathrm{f},`$ (11) where $`I_3^\mathrm{f}`$ is the third component of the weak isospin of the fermion f. Due to weak vertex corrections, the definitions of $`\overline{\rho }`$ and $`\mathrm{sin}^2\overline{\theta }_\mathrm{W}`$ depend on the fermion. However, except for the b-quark, these differences are small compared to the experimental precision. Therefore, we define $`\mathrm{sin}^2\overline{\theta }_\mathrm{W}`$ as the effective weak mixing angle for a massless charged lepton. It is related to the on-shell definition of the weak mixing angle, $`\mathrm{sin}^2\theta _\mathrm{W}`$ by the factor $`\kappa `$: $$\kappa =\frac{\mathrm{sin}^2\overline{\theta }_\mathrm{W}}{\mathrm{sin}^2\theta _\mathrm{W}}\mathrm{with}\mathrm{sin}^2\theta _\mathrm{W}=1\frac{m_\mathrm{W}^2}{m_\mathrm{Z}^2}.$$ (12) #### Fits in the SM The fourth analysis to determine electroweak parameters uses the framework of the SM. By comparing its predictions with the set of experimental measurements, it is possible to test the consistency of the SM and to constrain the mass of the Higgs boson. The input parameters of the SM are $`\alpha `$, the fermion masses, $`m_\mathrm{H}`$, $`m_\mathrm{Z}`$ and the mass of the W boson, $`m_\mathrm{W}`$. QCD adds one more parameter, the strong coupling constant, $`\alpha _\mathrm{s}`$, which is relevant for hadronic final states. The Cabbibo-Kobayashi-Maskawa matrix, relating electroweak and mass eigenstates of quarks, is not important for total hadronic cross sections in neutral current interactions considered here. Concerning the fermion masses, only the mass of the top quark is important for SM calculations performed below. All other masses are too small to play a significant role or are known to sufficient precision. Generally, in SM calculations for observables at the $`\mathrm{Z}`$ resonance, the mass of the W is replaced by the Fermi coupling constant, $`G_\mathrm{F}`$, which is measured precisely in muon decay . These two parameters are related by $$\frac{G_\mathrm{F}}{\sqrt{2}}=\frac{\pi \alpha }{2}\frac{1}{m_\mathrm{Z}^2\mathrm{sin}^2\theta _\mathrm{W}\mathrm{cos}^2\theta _\mathrm{W}}\frac{1}{1\mathrm{\Delta }r},$$ where $`\mathrm{\Delta }r`$ takes into account the electroweak radiative corrections. These corrections can be split into QED corrections due to the running of the QED coupling constant, $`\mathrm{\Delta }\alpha `$, and pure weak corrections, $`\mathrm{\Delta }r_\mathrm{w}`$ : $`\mathrm{\Delta }r`$ $`=`$ $`\mathrm{\Delta }\alpha +\mathrm{\Delta }r_\mathrm{w}`$ $`\mathrm{\Delta }r_\mathrm{w}`$ $`=`$ $`\mathrm{cot}^2\theta _\mathrm{W}\mathrm{\Delta }\overline{\rho }+\mathrm{\Delta }r_{\mathrm{rem}}.`$ (13) The corrections $`\mathrm{\Delta }r_{\mathrm{rem}}`$, not absorbed in the $`\rho `$-parameter, are smaller than the main contributions discussed below but are nevertheless numerically important and included in the calculations. Weak radiative corrections originate mainly from loop corrections to the W propagator due to the large mass splitting in the top-bottom iso-spin doublet and Higgs boson loop corrections to the propagators of the heavy gauge bosons . To leading order they depend quadratically on $`m_\mathrm{t}`$ and logarithmically on $`m_\mathrm{H}`$. The $`\mathrm{Zb}\overline{\mathrm{b}}`$ vertex receives additional weak radiative corrections which depend on the top mass. Through the measurements of weak radiative corrections our results at the $`\mathrm{Z}`$ are sensitive to the mass of the top quark and the Higgs boson. This allows to test the SM at the one-loop level by comparing the top mass derived from our data with the direct measurement and to estimate the mass of the yet undiscovered Higgs boson which is one of the fundamental parameters of the SM. With this procedure the relevant parameters in SM fits are $`m_\mathrm{Z}`$, $`m_\mathrm{t}`$, $`m_\mathrm{H}`$, $`\alpha `$($`m_\mathrm{Z}`$) and $`\alpha _\mathrm{s}`$. #### Fitting Programs and Methods The programs ZFITTER and TOPAZ0 are used to calculate radiative corrections and SM predictions. For computational reasons the fits are performed using ZFITTER. Both programs include complete $`𝒪(\alpha ^2)`$ and leading $`𝒪(\alpha ^3\mathrm{ln}^3(s/m_\mathrm{e}^2))`$ QED calculations of initial state radiation . Final state corrections are calculated in $`𝒪(\alpha )`$ for QED and $`𝒪(\alpha _\mathrm{s}^3)`$ for QCD including also mixed terms $`𝒪(\alpha \alpha _\mathrm{s})`$. Interference of initial and final state radiation is included up to $`𝒪(\alpha )`$ corrections. Pair production by initial state radiation is implemented . Electroweak radiative corrections are complete at the one-loop level and are supplemented by leading $`𝒪(G_\mathrm{F}^2m_\mathrm{t}^4)`$ and sub-leading $`𝒪(G_\mathrm{F}^2m_\mathrm{t}^2m_\mathrm{Z}^2)`$ two-loop corrections. Complete mixed QCD-electroweak corrections of $`𝒪(\alpha \alpha _\mathrm{s})`$ with leading $`𝒪(G_\mathrm{F}m_\mathrm{t}^2\alpha _s^2)`$ terms are included together with a non-factorizable part . For the reaction $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}(\gamma )`$ the contributions from the $`t`$-channel photon and Z boson exchange and the $`s/t`$-interference are calculated with the programs ALIBABA and TOPAZ0 (see Appendix A). Electroweak parameters are determined in $`\chi ^2`$ fits using the MINUIT program. The $`\chi ^2`$ is constructed from the theoretical expectations, our measurements and their errors, including the correlations. Apart from experimental statistical and systematic errors, and theoretical errors, we take into account uncertainties on the LEP centre-of-mass energy. Technical details of the fit procedure are described in Appendix B. Theoretical uncertainties on SM predictions of cross sections, asymmetries, Z decay widths and effective coupling constants are studied in detail in Reference . Errors on the theoretical calculations of cross section and forward-backward asymmetries based on total and partial $`\mathrm{Z}`$ widths, effective couplings or S-Matrix parameters, as used in the fits of Section 12, arise mainly from the finite precision of the QED convolution. They are found to be small compared to the experimental precision and do not introduce sizeable uncertainties in the fit for $`\mathrm{Z}`$ parameters. Residual SM uncertainties in the imaginary part of the effective couplings are even smaller. The only exceptions are the theoretical uncertainty on the luminosity determination, discussed in Section 5, and the treatment of $`t`$-channel and $`s/t`$-interference contributions to the $`\mathrm{e}^+\mathrm{e}^{}`$ final state due to missing higher order terms and the precision of the ALIBABA program . Uncertainties on the Bhabha cross section and forward-backward asymmetry from calculations of the $`t`$-channel and $`s/t`$-interference contributions are discussed in Appendix A. Additional uncertainties arise in the calculation of SM parameters from the application of different re-normalisation schemes, momentum transfer scales for vertex corrections and factorisations schemes . By comparing different calculations as implemented in ZFITTER and TOPAZ0, we find that the impact of these theoretical uncertainties on the fit results for SM parameters presented in Section 13 is negligible compared to the experimental errors. ZFITTER and TOPAZ0 calculations in the SM framework are performed based on five input parameters: the masses of the Z and Higgs bosons, the top quark mass, the strong coupling constant $`\alpha _\mathrm{s}`$ at $`m_\mathrm{Z}`$ and the contribution of the five light quark flavours, $`\mathrm{\Delta }\alpha _{\mathrm{had}}^{(5)}`$, to the running of the QED coupling constant to $`m_\mathrm{Z}`$. For comparison to the SM we use the following set of values and uncertainties : $$\begin{array}{ccccc}\hfill m_\mathrm{Z}\text{ = }& \mathrm{91\hspace{0.17em}189.8}\pm 3.1\mathrm{MeV},\hfill & \hfill m_\mathrm{t}\text{ = }& 173.8\pm 5.2\mathrm{GeV},\hfill & 95.3\mathrm{GeV}m_\mathrm{H}1\mathrm{TeV},\hfill \\ \hfill \mathrm{\Delta }\alpha _{\mathrm{had}}^{(5)}\text{ = }& 0.02804\pm 0.00065,\hfill & \hfill \alpha _\mathrm{s}\text{ = }& 0.119\pm 0.002.\hfill & \end{array}$$ (14) The central values are calculated for $`m_\mathrm{H}=300\mathrm{GeV}`$. This arbitrary choice is motivated by the logarithmic dependence of electroweak observables on $`m_\mathrm{H}`$ and it leads to approximately symmetric theoretical errors. We use the default settings of ZFITTER which provide the most accurate calculations. Exceptions are that in all calculations we allow for the variation of the contribution of the five light quarks to the running of the QED coupling constant, $`\mathrm{\Delta }\alpha _{\mathrm{had}}^{(5)}`$. Secondly, as recommended by the authors of ZFITTER, the corrections of Reference are explicitly calculated for SM expectations and in fits in the SM framework (Section 13). In all other cases they are absorbed in the definitions of the parameters. ## 12 Determination of $`𝐙`$ Parameters ### 12.1 Mass, Total and Partial Widths of the $`𝐙`$ We determine the mass, the total width and the partial decay widths of the $`\mathrm{Z}`$ into hadrons, electrons, muons and taus in a fit to the measured total cross sections. These parameters describe the contribution of the $`\mathrm{Z}`$ exchange to the total cross section. The photon exchange and $`\gamma /\mathrm{Z}`$-interference contributions are fixed to their SM expectations. Two fits are performed: one assuming and one not assuming lepton universality, where in the first one a common leptonic width is defined as the decay width of the $`\mathrm{Z}`$ into a pair of massless charged leptons. The results of both fits are summarised in Table 21 and the correlation coefficients for the parameters determined in the two fits are given in Tables 22 and 23, respectively. The partial decay widths into the three charged lepton species are found to be consistent within errors. It should be noted that due to the mass of the tau lepton, $`\mathrm{\Gamma }_\tau `$ is expected to be $`0.19\mathrm{MeV}`$ smaller than $`\mathrm{\Gamma }_{\mathrm{}}`$. Our new results with significantly reduced errors are in agreement with the SM expectations and our previous measurements . For the mass $`m_\mathrm{Z}`$ and the total width $`\mathrm{\Gamma }_\mathrm{Z}`$ we obtain: $`m_\mathrm{Z}`$ $`=`$ $`\mathrm{91\hspace{0.17em}189.8}\pm 3.1\mathrm{MeV},`$ $`\mathrm{\Gamma }_\mathrm{Z}`$ $`=`$ $`\mathrm{2\hspace{0.17em}502.4}\pm 4.2\mathrm{MeV}.`$ (15) These are measurements of the mass of the $`\mathrm{Z}`$ boson with an accuracy of $`3.410^5`$ and of its total decay width of $`1.710^3`$. The contribution to the total errors on $`m_\mathrm{Z}`$ and $`\mathrm{\Gamma }_\mathrm{Z}`$ from the LEP energy is estimated by performing fits to the $`199395`$ data with and without taking into account LEP energy errors. From a quadratic subtraction of the errors of the fitted parameters we find $`\mathrm{\Delta }m_\mathrm{Z}(\mathrm{LEP})=1.8\mathrm{MeV}`$ and $`\mathrm{\Delta }\mathrm{\Gamma }_\mathrm{Z}(\mathrm{LEP})=1.3\mathrm{MeV}`$, in agreement with the estimates given in Reference . The impact of the uncertainties on SM parameters on the fit results is negligible. The largest effect is an uncertainty on the $`\mathrm{Z}`$ mass of $`\pm 0.2\mathrm{MeV}`$ caused by the calculation of the $`\gamma /\mathrm{Z}`$-interference contribution when varying the Higgs and top masses and $`\mathrm{\Delta }\alpha _{\mathrm{had}}^{(5)}`$ in the ranges given in Equation 14. Motivated by the different methods used to obtain the absolute scale of the LEP energy in the years $`199092`$, $`199394`$ and 1995, resulting in different uncertainties, we determine the mass of the $`\mathrm{Z}`$ for these three periods independently. The mass values obtained are consistent within their statistical errors. To check our results on $`m_\mathrm{Z}`$ and $`\mathrm{\Gamma }_\mathrm{Z}`$ the fit assuming lepton universality is repeated twice: i) using only the leptonic cross sections and ii) using only the $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}(\gamma )`$ data. The results for the mass and total width obtained this way are $`m_\mathrm{Z}=\mathrm{91\hspace{0.17em}198.7}\pm 8.2\mathrm{MeV}`$, $`\mathrm{\Gamma }_\mathrm{Z}=\mathrm{2\hspace{0.17em}508.1}\pm 13.4\mathrm{MeV}`$ using all three lepton species and $`m_\mathrm{Z}=\mathrm{91\hspace{0.17em}177}\pm 16\mathrm{MeV}`$, $`\mathrm{\Gamma }_\mathrm{Z}=\mathrm{2\hspace{0.17em}497}\pm 26\mathrm{MeV}`$ when using only Bhabha scattering data. Within the errors, dominated by the statistical errors of the measurements, these values are in agreement with those given in Table 21 where the $`\mathrm{e}^+\mathrm{e}^{}\mathrm{hadrons}(\gamma )`$ cross section measurements contribute most. Also, we conclude that there is no significant bias introduced in the determination of the mass and the total width of the Z boson by the treatment of the $`t`$-channel in Bhabha scattering. From the difference of the total width and the partial widths into hadrons and charged leptons, including their correlations, the decay width of the $`\mathrm{Z}`$ into invisible particles is derived to be $$\mathrm{\Gamma }_{\mathrm{inv}}=499.1\pm 2.9\mathrm{MeV}.$$ (16) This number is determined in the fit assuming lepton universality and it is in agreement with our direct determination of $`\mathrm{\Gamma }_{\mathrm{inv}}`$ from cross section measurements of the reaction $`\mathrm{e}^+\mathrm{e}^{}\nu \overline{\nu }\gamma (\gamma )`$ which yields $`\mathrm{\Gamma }_{\mathrm{inv}}=498\pm 12(\mathrm{stat})\pm 12(\mathrm{sys})\mathrm{MeV}`$. In the SM, the invisible width is exclusively given by the $`\mathrm{Z}`$ decays into neutrinos and the result can be interpreted as the number of neutrino generations $`N_\nu `$. Using the SM prediction $`\mathrm{\Gamma }_{\mathrm{}}/\mathrm{\Gamma }_\nu =0.5021\pm 0.0002`$ for the ratio of the $`\mathrm{Z}`$ decay width into charged leptons and neutrinos we obtain: $$N_\nu =\frac{\mathrm{\Gamma }_{\mathrm{inv}}}{\mathrm{\Gamma }_{\mathrm{}}}\left(\frac{\mathrm{\Gamma }_{\mathrm{}}}{\mathrm{\Gamma }_\nu }\right)^{\mathrm{SM}}=2.978\pm 0.014.$$ (17) This formula is used because the experimental precision on the ratio $`\mathrm{\Gamma }_{\mathrm{inv}}/\mathrm{\Gamma }_{\mathrm{}}`$ is better than that on $`\mathrm{\Gamma }_{\mathrm{inv}}`$. ### 12.2 Limits on Non-Standard Decays of the $`𝐙`$ From the measurements of total and partial $`\mathrm{Z}`$ decay widths presented in the previous section we derive experimental limits on additional $`\mathrm{Z}`$ decay widths not accounted for in the SM. These limits take into account experimental and theoretical errors added in quadrature. The latter are derived from adding in quadrature the changes in the theoretical predictions when varying the SM input parameters by their errors as given in Equation 14. This is motivated by the fact that these parameters are determined in independent experiments with the exception of the mass of the Higgs boson. A value of $`m_\mathrm{H}=1\mathrm{TeV}`$ is used here to calculate the Z widths which results in the lowest SM predictions and therefore in conservative limits. The 95% confidence level (C.L.) limits on non-standard decay widths, $`\mathrm{\Gamma }_{95}^{\mathrm{NP}}`$, are calculated using the formula : $`10.95`$ $`=`$ $`{\displaystyle \frac{_{\mathrm{}}^{\mathrm{\Gamma }^{\mathrm{exp}}}d\mathrm{\Gamma }G(\mathrm{\Gamma };\mathrm{\Gamma }^{\mathrm{SM}}+\mathrm{\Gamma }_{95}^{\mathrm{NP}},\mathrm{\Delta })}{_{\mathrm{}}^{\mathrm{\Gamma }^{\mathrm{exp}}}d\mathrm{\Gamma }G(\mathrm{\Gamma };\mathrm{\Gamma }^{\mathrm{SM}},\mathrm{\Delta })}}`$ $`\mathrm{with}G(\mathrm{\Gamma };\mu ,\mathrm{\Delta })`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2\pi }\mathrm{\Delta }}}\mathrm{exp}\left[{\displaystyle \frac{(\mathrm{\Gamma }\mu )^2}{2\mathrm{\Delta }^2}}\right],`$ (18) where $`\mathrm{\Gamma }^{\mathrm{exp}}`$ is our experimental result, $`\mathrm{\Gamma }^{\mathrm{SM}}`$ the SM expectation for $`m_\mathrm{H}=1\mathrm{TeV}`$ and $`\mathrm{\Delta }`$ the combined experimental and theoretical error. The limits obtained for the total, hadronic, leptonic and invisible widths, as well as for the three lepton species, are summarised in Table 24. Also listed are the differences of our measurements and the SM expectations together with their experimental and theoretical 68% C.L. errors. It should be noted that the results on the total and partial widths are correlated; hence the limits derived in this section cannot be applied simultaneously. ### 12.3 Fits to Total Cross Sections and Forward-Backward Asymmetries The measured leptonic forward-backward asymmetries are included in the fits. Besides $`m_\mathrm{Z}`$ and $`\mathrm{\Gamma }_\mathrm{Z}`$ the measurements are fitted to the hadronic pole cross section, $`\sigma _{\mathrm{had}}^0`$, the ratios of hadronic to leptonic widths, $`R_{\mathrm{}}`$, and the leptonic pole asymmetries, $`A_{\mathrm{FB}}^{0,\mathrm{}}`$, which are defined as: $$\sigma _{\mathrm{had}}^0=\frac{12\pi }{m_\mathrm{Z}^2}\frac{\mathrm{\Gamma }_\mathrm{e}\mathrm{\Gamma }_{\mathrm{had}}}{\mathrm{\Gamma }_\mathrm{Z}^2},R_{\mathrm{}}=\frac{\mathrm{\Gamma }_{\mathrm{had}}}{\mathrm{\Gamma }_{\mathrm{}}},A_{\mathrm{FB}}^{0,\mathrm{}}=\frac{3}{4}A_\mathrm{e}A_{\mathrm{}}(\mathrm{}=\mathrm{e},\mu ,\tau ).$$ (19) The advantage of this parameter set is that the parameters are less correlated than the partial widths. Two fits are performed, one with and one without assuming lepton universality. The results are listed in Table 25 and the correlation matrices are given in Tables 26 and 27. The 68% C.L. contours in the $`A_{\mathrm{FB}}^{0,\mathrm{}}R_{\mathrm{}}`$ plane are derived from these fits for the three lepton species separately and for all leptons combined (Figure 32). In this plot the contour of $`A_{\mathrm{FB}}^{0,\tau }R_\tau `$ is shifted by the difference in expectation for $`R_\tau `$ due to the tau mass to facilitate the comparison with the other leptons. Also for the forward-backward asymmetries good agreement among the lepton species is observed. Our results are in agreement with the SM expectations. From the measurements of the forward-backward pole asymmetries the polarisation parameter, $`A_{\mathrm{}}`$, can be derived for the three individual lepton types as well as the average value. The results are listed in Table 28. Because of their relation to the measured pole asymmetry (Equation 7) the results for $`A_\mathrm{e}`$, $`A_\mu `$ and $`A_\tau `$ are highly correlated. They are compared to $`A_\mathrm{e}`$ and $`A_\tau `$ derived from our measurements of the average and the forward-backward tau-polarisation . All measurements are in good agreement and yield an average value of $$A_{\mathrm{}}=0.1575\pm 0.0067.$$ (20) ### 12.4 Vector- and Axial-Vector Coupling Constants of Charged Leptons The effective coupling constants, $`\overline{g}_\mathrm{V}^{\mathrm{}}`$ and $`\overline{g}_\mathrm{A}^{\mathrm{}}`$, are obtained from a fit to cross section and forward-backward asymmetry measurements, and including our results from tau-polarisation. We use the results from tau-polarisation on $`A_\mathrm{e}`$ and $`A_\tau `$ as given in Table 28 together with a 8% correlation of the errors. The inclusion of tau-polarisation results significantly improves the determination of the effective coupling constants. Fits with and without assuming lepton universality are performed and the vector and axial-vector coupling constants so obtained are listed Table 29. The axial-vector coupling constant of the electron is taken to be negative, in agreement with the combination of results from neutrino-electron scattering and low energy $`A_{\mathrm{FB}}`$ measurements . All other signs are unambiguously determined by our measurements. The 68% C.L. contours in the $`\overline{g}_\mathrm{V}`$-$`\overline{g}_\mathrm{A}`$ plane are shown in Figure 33, revealing good agreement among the three lepton species and thus supporting lepton universality in neutral currents. This is quantified by calculating the ratio of muon and tau to electron coupling constants, taking into account their correlations (see Table 30). The average vector and axial-vector coupling constants of charged leptons are found to be $$\overline{g}_\mathrm{V}^{\mathrm{}}=0.0397\pm 0.0017,\overline{g}_\mathrm{A}^{\mathrm{}}=0.50153\pm 0.00053.$$ (21) The resulting axial-vector coupling constant $`\overline{g}_\mathrm{A}^{\mathrm{}}`$ is significantly different from its lowest order SM expectation $`1/2`$. This is interpreted as proof for the existence of weak radiative corrections from higher order processes and corresponds to a measurement of the $`\rho `$-parameter of $$\overline{\rho }=1.0061\pm 0.0021.$$ (22) The evidence for the existence of weak radiative corrections is illustrated in Figure 33. Measurements of forward-backward asymmetries and tau-polarisation can be compared in terms of the effective weak mixing angle, $`\mathrm{sin}^2\overline{\theta }_\mathrm{W}`$, defined by the ratio of the coupling constants (Equation 11). From the average leptonic pole asymmetry and the tau polarisation, the values listed in Table 31 are obtained. Our results obtained from the measurement of the forward-backward asymmetry of b-quarks and the measurement of the quark charge asymmetry, $`Q_{\mathrm{FB}}`$ are also shown. All four measurements of the weak mixing angle are in agreement with each other and the average yields: $$\mathrm{sin}^2\overline{\theta }_\mathrm{W}=0.23093\pm 0.00066.$$ (23) ### 12.5 Fits in the S-Matrix Framework The programs SMATASY together with ZFITTER, ALIBABA and TOPAZ0 are used for the calculation of the theoretical predictions, including QED radiative corrections, of total cross sections and forward-backward asymmetries. Further details can be found in Reference . The results of the fits in the S-Matrix framework with and without imposing lepton universality to the cross sections and forward-backward asymmetries measured at the Z resonance are shown in Table 32. The fitted parameters for electrons, muons, taus and hadrons are in agreement with each other and with the expectations from the SM. The correlations of the parameters as obtained in the two fits are shown in Tables 33 and 34, respectively. Large correlations among the parameters are observed. Of particular importance is the correlation of $`0.95`$ between the mass of the $`\mathrm{Z}`$ boson and the hadronic interference term, $`j_{\mathrm{had}}^{\mathrm{tot}}`$. It causes an increase of the error on $`m_\mathrm{Z}`$ with respect to the fits performed in Sections 12.1 and 12.3 where the $`\gamma /\mathrm{Z}`$-interference terms are fixed to their SM expectations. The correlation between $`m_\mathrm{Z}`$ and $`j_{\mathrm{had}}^{\mathrm{tot}}`$ is illustrated in Figure 34. Comparing the results on the $`\mathrm{Z}`$ boson mass obtained with the two analyses (Table 21 or 25 and Table 32) good agreement is found. From a quadratic subtraction we estimate the additional error on the $`\mathrm{Z}`$ mass arising from the experimental uncertainty on the hadronic interference term to be: $$m_\mathrm{Z}=\mathrm{91\hspace{0.17em}185.2}\pm 3.1\pm 9.8(j_{\mathrm{had}}^{\mathrm{tot}})\mathrm{MeV}.$$ (24) The interference between the photon and the $`\mathrm{Z}`$ is measured to much better precision at centre-of-mass energies below or above the $`\mathrm{Z}`$ resonance. By adding our measurements of hadronic and leptonic cross sections and forward-backward asymmetries above the $`\mathrm{Z}`$ resonance this contribution to the error on $`m_\mathrm{Z}`$ is significantly reduced. This will be reported in a forthcoming publication. ## 13 Results on SM Parameters We interpret our measurements in the framework of the SM to check its consistency by comparing our results to other measurements. The strategy will be to test at first QCD radiative corrections in terms of $`\alpha _\mathrm{s}`$ before we verify weak radiative corrections by comparing the top mass derived from our data at the $`\mathrm{Z}`$ resonance to the direct measurement. From our measurements at the $`\mathrm{Z}`$, the W mass is determined and compared to our result obtained above the W-pair threshold. Finally we use all our measurements of electroweak parameters and include the direct measurement of $`m_\mathrm{t}`$ to estimate the mass of the SM Higgs boson. Fits are performed to our data to determine the set of SM parameters given in Equation 14. The program ZFITTER is used for SM calculations. In all fits the QED coupling constant at the mass of the $`\mathrm{Z}`$ is calculated using the constraint on the contribution from the five light quark flavours to the running as obtained in Reference . The input data for the SM fits are our measurements of total cross sections and forward-backward asymmetries at the $`\mathrm{Z}`$ resonance performed between 1990 and 1995. In addition, our results from tau polarisation (Table 28), the effective weak mixing angles from b-quark forward-backward and form quark charge asymmetry (Table 31), as well as the partial decay width into b-quarks $`R_\mathrm{b}=\mathrm{\Gamma }_\mathrm{b}/\mathrm{\Gamma }_{\mathrm{had}}=0.2174\pm 0.0032`$ , are included. Firstly, the sensitivity of our data to QCD radiative corrections is exploited to determine the strong coupling constant at the mass of the Z boson: $`\alpha _\mathrm{s}`$ $`=`$ $`0.1226_{0.0060}^{+0.0066}.`$ (25) In this fit the result of Section 12.1 is obtained again for $`m_\mathrm{Z}`$ and $`\mathrm{\Delta }\alpha _{\mathrm{had}}^{(5)}`$ remains within the imposed constraint. The masses of the top quark and the Higgs boson are free parameters and their uncertainties are included in the error on $`\alpha _\mathrm{s}`$. The value for $`\alpha _\mathrm{s}`$ is in good agreement with our determination of the strong coupling constant from hadronic event topologies $`\alpha _\mathrm{s}=0.1216\pm 0.0017\pm 0.0058`$ . We use this measurement as an additional constraint and obtain for the top mass: $$m_\mathrm{t}=197_{16}^{+30}\mathrm{GeV}.$$ (26) This result for the top quark mass is based on our measurements of weak radiative corrections and their interpretation in the SM framework. The agreement with the direct $`m_\mathrm{t}`$ measurements by the CDF and D0 experiments, $`m_\mathrm{t}=173.8\pm 5.2\mathrm{GeV}`$ , means that the bulk of weak radiative corrections indeed originates from the large mass of the top quark. The result for the Higgs boson mass obtained in this fit, $`\mathrm{log}_{10}m_\mathrm{H}/\mathrm{GeV}=1.99_{0.66}^{+0.98}`$, is in agreement with the range allowed by the direct search and the SM (Equation 14). From the result of this fit which is based on measurements at the Z resonance a value for the mass of the W boson is derived: $$m_\mathrm{W}=80.523\pm 0.079\mathrm{GeV}.$$ (27) The 68% C.L. contour in the $`m_\mathrm{t}`$-$`m_\mathrm{W}`$ plane obtained in this fit is shown in Figure 35. This result for the W mass agrees well with our direct measurements of $`m_\mathrm{W}`$ performed at centre-of-mass energies between $`161\mathrm{GeV}`$ and $`183\mathrm{GeV}`$ which yield a combined value of $`m_\mathrm{W}=80.61\pm 0.15\mathrm{GeV}`$. We include the direct measurement of $`m_\mathrm{W}`$ to determine the electroweak radiative corrections $`\mathrm{\Delta }r`$, the parameters $`\rho `$ and $`\kappa `$, the effective weak mixing angle $`\mathrm{sin}^2\overline{\theta }_\mathrm{W}`$ and the on-shell definition $`\mathrm{sin}^2\theta _\mathrm{W}`$ as well as a combined result for $`m_\mathrm{W}`$: $$\begin{array}{cccccc}\hfill \mathrm{\Delta }r\text{ = }& 0.0257\pm 0.0043,\hfill & \hfill \rho \text{ = }& 1.0078\pm 0.0017,\hfill & \hfill \kappa \text{ = }& 1.0493\pm 0.0053,\hfill \\ \hfill \mathrm{sin}^2\overline{\theta }_\mathrm{W}\text{ = }& 0.23075\pm 0.00054,\hfill & \hfill \mathrm{sin}^2\theta _\mathrm{W}\text{ = }& 0.2199\pm 0.0013,\hfill & \hfill m_\mathrm{W}\text{ = }& 80.541\pm 0.069\mathrm{GeV}.\hfill \end{array}$$ (28) This value for the effective weak mixing angle $`\mathrm{sin}^2\overline{\theta }_\mathrm{W}`$ derived in a SM fit is in good agreement with the result obtained, in a less model-dependent way, from measurements of asymmetries (Table 31). Finally, we constrain the mass of the top quark to the combined value from the direct measurement of D0 and CDF. The five SM parameters and their correlations obtained in this fit are summarised in Tables 35 and 36, respectively. In particular, for the mass of the yet undiscovered SM Higgs boson, we obtain a value and an upper limit: $`m_\mathrm{H}`$ $`=`$ $`36_{19}^{+43}\mathrm{GeV},`$ (29) $`<`$ $`133\mathrm{GeV}95\%\mathrm{C}.\mathrm{L}.`$ Figure 36 shows the 68% and 95% C.L. contours in the $`m_\mathrm{t}`$-$`m_\mathrm{H}`$ plane and Figure 37 the dependence of the $`\chi ^2`$ of the fit on the Higgs mass from which the upper mass limit is derived. The result is compatible with the result of our direct search for the SM Higgs boson $`m_\mathrm{H}>95.3\mathrm{GeV}`$ . ## 14 Summary and Conclusion We report on the precise measurements of total cross sections and forward-backward asymmetries of the reactions $`\mathrm{e}^+\mathrm{e}^{}\mathrm{hadrons}(\gamma )`$, $`\mathrm{e}^+\mathrm{e}^{}\mu ^+\mu ^{}(\gamma )`$, $`\mathrm{e}^+\mathrm{e}^{}\tau ^+\tau ^{}(\gamma )`$ and $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}(\gamma )`$ at centre-of-mass energies at the peak and the wings of the $`\mathrm{Z}`$ resonance performed in the years $`199395`$. A total luminosity of $`103\mathrm{pb}^1`$ corresponding to 2.5 million hadronic and 250 thousand leptonic decays of the $`\mathrm{Z}`$ was collected which significantly improve our measurements of the resonance curve. Including the data samples collected in previous years, the total number of $`\mathrm{Z}`$ decays observed by the L3 detector during the first phase of LEP amounts to 4 million which are used to determine the properties of the $`\mathrm{Z}`$ and other SM parameters. All our measurements are consistent with the hypothesis of lepton universality. From the measured total hadronic and leptonic cross sections we obtain: $`m_\mathrm{Z}=\mathrm{91\hspace{0.17em}189.8}\pm 3.1\mathrm{MeV},`$ $`\mathrm{\Gamma }_\mathrm{Z}=\mathrm{2\hspace{0.17em}502.4}\pm 4.2\mathrm{MeV},`$ $`\mathrm{\Gamma }_{\mathrm{had}}=\mathrm{1\hspace{0.17em}751.1}\pm 3.8\mathrm{MeV},`$ $`\mathrm{\Gamma }_{\mathrm{}}=84.14\pm 0.17\mathrm{MeV}.`$ From these results, the decay width of the $`\mathrm{Z}`$ into invisible particles is derived to be $`\mathrm{\Gamma }_{\mathrm{inv}}=499.1\pm 2.9\mathrm{MeV}`$, which in the SM corresponds to a number of light neutrino species of: $$N_\nu =2.978\pm 0.014.$$ (30) Including our measurements of leptonic forward-backward asymmetries and tau polarisation the effective vector and axial-vector coupling constants of charged leptons to the $`\mathrm{Z}`$ are determined to be: $$\overline{g}_\mathrm{V}^{\mathrm{}}=0.0397\pm 0.0017,\overline{g}_\mathrm{A}^{\mathrm{}}=0.50153\pm 0.00053.$$ (31) For the effective weak mixing angle we obtain: $$\mathrm{sin}^2\overline{\theta }_\mathrm{W}=0.23093\pm 0.00066,$$ (32) including our measurements of the b-quark forward-backward and quark charge asymmetries. Our measurements are sensitive to higher order weak radiative corrections which depend on the masses of the top quark and the Higgs boson. Using in addition our measurements of the partial width $`\mathrm{Z}\mathrm{b}\overline{\mathrm{b}}`$ and $`\alpha _\mathrm{s}`$, we derive in the SM framework a top quark mass $$m_\mathrm{t}=197_{16}^{+30}\mathrm{GeV},$$ (33) which is in agreement with the direct measurements of $`m_\mathrm{t}`$. Using our direct measurement of $`m_\mathrm{W}`$ and the knowledge of $`m_\mathrm{t}`$ our data constrain the mass of the Higgs boson to $$m_\mathrm{H}<133\mathrm{GeV}95\%\mathrm{C}.\mathrm{L}.$$ (34) ## Acknowledgments We congratulate the CERN accelerator divisions for their achievements in the precise calibration of the beam energy and we express our gratitude for the excellent performance of the LEP machine. We like to thank W. Beenakker, D. Bardin, G. Passarino and their collaborators for many fruitful discussions and their help in theory questions. We acknowledge the contributions of the engineers and technicians who have participated in the construction and maintenance of this experiment. Appendix ## Appendix A Treatment of Contributions related to the $`𝒕`$-channel In the case of the process $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}(\gamma )`$, the existence of the $`t`$-channel exchange of photons and $`\mathrm{Z}`$ bosons and its interference with the $`s`$-channel exchange lead to additional complications. Analytical programs to calculate this process, such as ALIBABA and TOPAZ0, are not directly suited for fitting purposes, as computationally they are very time consuming. Thus, the following procedure is adopted. During the initialisation of a fit, ALIBABA is used once to calculate the predictions of the $`t`$-channel and $`s/t`$-interference contributions to the measured $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}(\gamma )`$ cross sections and forward-backward asymmetries. Calculations are performed at several centre-of-mass energy values in the vicinity of the data points allowing for a reduction of the statistical error of the Monte Carlo integration used by ALIBABA and the calculation of derivatives needed to construct the covariance matrix (see Appendix B). ZFITTER is employed during the fits to calculate the corresponding s-channel contributions as a function of the varying electroweak parameters. The contribution from $`s/t`$-interference depends on the fit parameter, most importantly on the Z boson mass. This dependence is taken into account by converting the difference between the Z mass used in the initialisation and the current fit value into an equivalent shift in the centre-of-mass energy at which the $`t`$-channel and $`s/t`$-interference contributions are calculated. The dependence of the $`s/t`$-interference on $`m_\mathrm{Z}`$ is responsible for a correlation of the results for $`m_\mathrm{Z}`$ and the electron $`s`$-channel cross section which amounts to $`+11\%`$ between $`m_\mathrm{Z}`$ and $`R_\mathrm{e}`$ in the nine parameter fit (Table 26). In the analytical program ZFITTER polar angular cuts can only be applied on the positron while in the experimental cross section measurements both electron and positron are required to lie within the fiducial volume. Correction factors are calculated with TOPAZ0 which allows for both types of cuts. Finally, cross sections of $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}(\gamma )`$ to be compared to the experimental measurement are calculated as: $`\sigma ^{\mathrm{th}}`$ $`=`$ $`\sigma ^{t+s/t,\mathrm{AL}}(44^{}<\theta _\mathrm{e}^{},\theta _{\mathrm{e}^+}<136^{})+\sigma ^{s,\mathrm{ZF}}(44^{}<\theta _{\mathrm{e}^+}<136^{})R`$ $`\mathrm{with}R`$ $`=`$ $`{\displaystyle \frac{\sigma ^{s,\mathrm{T0}}(44^{}<\theta _\mathrm{e}^{},\theta _{\mathrm{e}^+}<136^{})}{\sigma ^{s,\mathrm{T0}}(44^{}<\theta _{\mathrm{e}^+}<136^{})}}`$ (35) All cross sections are defined for an acollinearity angle cut of $`\xi <25^{}`$. The indices AL, T0 and ZF label cross sections calculated with the ALIBABA, TOPAZ0 and ZFITTER programs, respectively. Cuts on the polar angle of the electron and positron are given in parentheses. This procedure is applied because it combines the most accurate treatment of electroweak radiative corrections as available in ZFITTER and TOPAZ0 with the most complete calculations of $`t`$-channel contributions in ALIBABA. In the case of the forward-backward asymmetry, $`A_{\mathrm{FB}}=(\sigma _\mathrm{F}\sigma _\mathrm{B})/(\sigma _\mathrm{F}+\sigma _\mathrm{B})`$, the analogous calculations for the forward, $`\sigma _\mathrm{F}`$, and backward cross sections, $`\sigma _\mathrm{B}`$, are performed. The cross sections calculated with the ALIBABA and TOPAZ0 programs for all $`199095`$ data sets are listed in in Table 37. The theoretical uncertainties on the calculation of the $`t`$-channel contributions are listed in Table 38. These errors are applied to the measured cross sections and forward-backward asymmetries in the fits (see Appendix B). Due to the contribution of the $`t`$-channel the results from the Bhabha channel cannot be directly compared to the measurements of the other leptonic final states. To permit such a comparison, Equation 35 is used to calculate the $`s`$-channel contributions, replacing $`\sigma ^{\mathrm{th}}`$ by the measurements and using ZFITTER for the extrapolation to the full solid angle. The $`s`$-channel cross sections obtained this way, $`\sigma _\mathrm{e}^\mathrm{s}`$, without any cuts, and $`s`$-channel asymmetries, $`A_{\mathrm{FB}}^s`$, with an acollinearity angle cut of $`\xi <25^{}`$, are given in Tables 17 and 20. ## Appendix B Construction of the Covariance Matrix All fits for electroweak parameters described in this article are performed in the error matrix analysis. They consist of minimising a $`\chi ^2`$ function defined as $$\chi ^2=D^TV^1D,$$ (36) where $`D`$ is a column vector with elements defined by the difference between measurements $`\mathrm{\Omega }_i^{\mathrm{exp}}`$ and theoretical expectations $`\mathrm{\Omega }_i^{\mathrm{th}}`$ which are calculated during the fit as a function of the fit parameters: $$D_i=\mathrm{\Omega }_i^{\mathrm{exp}}\mathrm{\Omega }_i^{\mathrm{th}}.$$ (37) The index $`i`$ runs over all cross section ($`\mathrm{\Omega }_i=\sigma _i`$) and forward-backward asymmetry ($`\mathrm{\Omega }_i=A_{\mathrm{FB}}^{}{}_{,i}{}^{}`$) measurements considered in the fit. There are 100 cross section and 75 forward-backward asymmetry measurements at the $`\mathrm{Z}`$ resonance taken in $`199095`$. The covariance matrix $`V`$ is constructed in the following way from all experimental and theoretical errors affecting the measurements. The diagonal elements, $`V_{ii}`$, are obtained adding all individual errors of measurement $`i`$ in quadrature $`V_{ii}`$ $`=`$ $`\left(f_i\mathrm{\Delta }_i^{\mathrm{stat}}\right)^2+\left(f_i\mathrm{\Delta }_i^{\mathrm{unc}}\right)^2+\left(\mathrm{\Delta }_i^{\mathrm{cor}}\right)^2+\left(\mathrm{\Delta }_i^{\mathrm{abs}}\right)^2`$ (38) $`+V_{ii}^{\mathrm{lum}}+V_{ii}^{\mathrm{LEP}}+V_{ii}^{ϵ_{\mathrm{cms}}}+V_{ii}^t,`$ where $`\mathrm{\Delta }_i^{\mathrm{stat}}`$ is the statistical error of the measurement, $`\mathrm{\Delta }_i^{\mathrm{unc}}`$ the uncorrelated part of the systematic error, $`\mathrm{\Delta }_i^{\mathrm{cor}}=\delta _i^{\mathrm{cor}}\sigma _i^{\mathrm{th}}`$ the correlated systematic error which scales with the expected value and $`\mathrm{\Delta }_i^{\mathrm{abs}}`$ the correlated part of the systematic error which does not scale. In case of the forward-backward asymmetry the systematic error can simply be expressed in terms of $`\mathrm{\Delta }_i^{\mathrm{abs}}`$ only. The statistical and uncorrelated systematic errors which are derived from the measurements (Tables 14 to 20) are scaled during the fit with factors $`f_i`$ to the expected errors using the theoretical expectations: $$f_i=\sqrt{\frac{\sigma _i^{\mathrm{th}}}{\sigma _i^{\mathrm{exp}}}}(\mathrm{for}\sigma _i),f_i=\sqrt{\frac{1\left(A_{\mathrm{FB}}^{}{}_{,i}{}^{\mathrm{th}}\right)^2}{1\left(A_{\mathrm{FB}}^{}{}_{,i}{}^{\mathrm{exp}}\right)^2}\frac{\sigma _i^{\mathrm{exp}}}{\sigma _i^{\mathrm{th}}}}(\mathrm{for}A_{\mathrm{FB}}^{}{}_{,i}{}^{}).$$ (39) For cross section measurements there is an additional contribution of the luminosity measurement, $`V_{ii}^{\mathrm{lum}}`$. Its calculation as well as the contribution from the uncertainty on the LEP centre-of-mass energy, $`V_{ii}^{\mathrm{LEP}}`$, and its spread, $`V_{ii}^{ϵ_{\mathrm{cms}}}`$, both applied to cross sections and $`A_{\mathrm{FB}}`$ measurements, and the theoretical uncertainty on the subtraction of the $`t`$-channel and $`s/t`$-interference contribution to Bhabha scattering, $`V_{ii}^t`$, are described below. The off-diagonal elements are constructed from correlated error sources: $`V_{ij}`$ $`=`$ $`\mathrm{\Delta }_i^{\mathrm{cor}}\mathrm{\Delta }_j^{\mathrm{cor}}+\mathrm{\Delta }_i^{\mathrm{abs}}\mathrm{\Delta }_j^{\mathrm{abs}}`$ (40) $`+V_{ij}^{\mathrm{lum}}+V_{ij}^{\mathrm{LEP}}+V_{ij}^{ϵ_{\mathrm{cms}}}+V_{ij}^t(ij).`$ Experimental systematic errors, $`\mathrm{\Delta }^{\mathrm{cor}}`$ and $`\mathrm{\Delta }^{\mathrm{abs}}`$, are only applied to elements connecting the same observable, either cross section or asymmetry, of the same reaction. As above, for forward-backward asymmetry measurements contributions from $`\mathrm{\Delta }^{\mathrm{cor}}`$ and $`V^{\mathrm{lum}}`$ are not applicable. Contributions from $`V_{ij}^{\mathrm{lum}}`$, $`V_{ij}^{\mathrm{LEP}}`$, $`V_{ij}^{ϵ_{\mathrm{cms}}}`$ and $`V_{ij}^t`$ enter also into off-diagonal elements connecting measurements of different observables or reactions. All statistical errors and the individual contributions to systematic errors for the measurements performed in $`199395`$ are listed in Tables 14 to 20. For the measurements performed in $`199092`$ systematic errors are quoted in Reference as relative errors, $`\delta _i`$, for the cross section and absolute errors, $`\mathrm{\Delta }_i`$, for the forward-backward asymmetries. The correlation among these systematic errors is treated by using $`V_{ij}=(\mathrm{min}(\delta _i,\delta _j))^2\sigma _i^{\mathrm{th}}\sigma _j^{\mathrm{th}}`$ for the cross sections and $`V_{ij}=(\mathrm{min}(\mathrm{\Delta }_i,\mathrm{\Delta }_j))^2`$ for the asymmetries. Contributions from uncertainties on the luminosity and the LEP energy are added, where applicable. Correlations between experimental systematic errors in the $`199092`$ and $`199395`$ data sets are estimated in the same way by using the smaller values of $`199395`$. Exceptions are the measurements of the $`\mathrm{e}^+\mathrm{e}^{}\tau ^+\tau ^{}(\gamma )`$ cross section where due to the completely revised analysis in $`199395`$ an additional factor of $`0.72`$ is applied to $`\delta _i^{\mathrm{cor}}`$. Other contributions to elements connecting measurements of $`199092`$ and $`199395`$ are from $`V_{ij}^{ϵ_{\mathrm{cms}}}`$, $`V_{ij}^t`$ and the theoretical uncertainty in $`V_{ij}^{\mathrm{lum}}`$. #### Uncertainty on the Luminosity For the $`199395`$ cross sections the contributions to the covariance matrix from errors on the luminosity measurement are obtained from the sum of the total experimental errors, including their correlations, and the theoretical uncertainty: $$V_{ij}^{\mathrm{lum}}=\left(\delta _k^{\mathrm{lum},\mathrm{exp}}\sigma _i^{\mathrm{th}}\right)\left(\delta _l^{\mathrm{lum},\mathrm{exp}}\sigma _j^{\mathrm{th}}\right)\rho _{kl}^{\mathrm{lum},\mathrm{exp}}+\left(\delta _m^{\mathrm{lum},\mathrm{stat}}\right)^2\sigma _i^{\mathrm{th}}\sigma _j^{\mathrm{th}}+\left(\delta ^{\mathrm{lum},\mathrm{th}}\right)^2\sigma _i^{\mathrm{th}}\sigma _j^{\mathrm{th}}.$$ (41) The indices $`k`$ and $`l`$ label the years of the data sets $`i`$ and $`j`$. Total experimental errors on the luminosity, $`\delta ^{\mathrm{lum},\mathrm{exp}}`$, are given in Table 4 and their correlations, $`\rho _{kl}^{\mathrm{lum},\mathrm{exp}}`$, in Table 5. The statistical error on the luminosity measurement, $`\delta ^{\mathrm{lum},\mathrm{stat}}`$, is given in Table 6 for the nine data taking periods. It applies only to cross section measurements $`i`$ and $`j`$ performed in the same period $`m`$. The combined experimental and theoretical error on the luminosity determination in $`199092`$ is $`\delta ^{\mathrm{lum}}=6`$. It is treated as fully correlated and the corresponding terms in the covariance matrix are calculated as $`V_{ij}^{\mathrm{lum}}=(\delta ^{\mathrm{lum}})^2\sigma _i^{\mathrm{th}}\sigma _j^{\mathrm{th}}`$. For the $`199092`$ data the statistical error of the luminosity measurement is included in the quoted statistical errors of the hadron cross section measurements and it is neglected for the leptonic cross sections. The theoretical uncertainty, $`\delta ^{\mathrm{lum},\mathrm{th}}=0.61`$, is fully correlated among all cross section measurements in $`199095`$. The measurement of the luminosity, and hence of $`\sigma _i^{\mathrm{exp}}`$, depends on the centre-of-mass energy $`E`$ due to the approximate $`1/E^2`$ dependence of the Bhabha cross section: $$\frac{\mathrm{d}\sigma _i^{\mathrm{exp}}}{\mathrm{d}E}=\kappa _m\frac{\sigma _i^{\mathrm{exp}}}{E_i}.$$ (42) Because of the $`\gamma /\mathrm{Z}`$-interference and higher order contributions the exponent $`\kappa _m`$ differs from 2 and it is calculated with BHLUMI to be $`\kappa _m=1.95,2.27`$ and $`1.97`$ for the peak$`2`$, peak and peak$`+2`$ data sets, respectively. This dependence causes a small contribution to the uncertainty on the cross section measurements and it is taken into account for the $`199395`$ data together with the error on the LEP energy. #### Uncertainty on the Centre-of-Mass Energy The errors on the LEP centre-of-mass energy are transformed into equivalent errors of cross section and asymmetry measurements using the partial derivatives of the theoretical cross sections and forward-backward asymmetries with respect to the centre-of-mass energy, $`\mathrm{\Omega }_i^{\mathrm{th}}/E`$. The dependence of the measured cross section, via the luminosity, on the centre-of-mass energy is included where applicable. For the $`199395`$ data sets the terms in the covariance matrix are determined as $`V_{ij}^{\mathrm{LEP}}`$ $`=`$ $`v_{kl}^{\mathrm{LEP}}{\displaystyle \frac{\mathrm{d}\mathrm{\Omega }_i}{\mathrm{d}E}}{\displaystyle \frac{\mathrm{d}\mathrm{\Omega }_j}{\mathrm{d}E}},`$ $`\mathrm{with}{\displaystyle \frac{\mathrm{d}\mathrm{\Omega }_i}{\mathrm{d}E}}`$ $`=`$ $`{\displaystyle \frac{\sigma _i^{\mathrm{th}}}{E}}\kappa _m{\displaystyle \frac{\sigma _i^{\mathrm{exp}}}{E_i}}\mathrm{for}199395\mathrm{cross}\mathrm{sections},`$ (43) $`=`$ $`{\displaystyle \frac{\mathrm{\Omega }_i^{\mathrm{th}}}{E}}\mathrm{otherwise}.`$ where $`k`$ and $`l`$ are the indices of the LEP energy calibration periods corresponding to the data sets $`i`$ and $`j`$, with $`\kappa _m`$ the appropriate factor defined in Eq. 42. For a pair of data sets taken in any of the seven periods with precise LEP energy calibration, $`v_{kl}^{\mathrm{LEP}}`$ is the corresponding element of the LEP energy error matrix as given in Table 1. For data sets taken during the 1993 or 1995 pre-scans the LEP energy error is treated as uncorrelated with any other period. Hence, only elements connecting data sets taken during the same pre-scan receive contributions $`v_{kk}^{\mathrm{LEP}}=(\delta _k^{\mathrm{LEP}})^2`$, where $`\delta _k^{\mathrm{LEP}}`$ are the uncertainties on the centre-of-mass energy of $`18\mathrm{MeV}`$ and $`10\mathrm{MeV}`$ in 1993 and 1995, respectively. The uncertainties on the LEP centre-of-mass energy in $`199092`$ are listed in Table 3. They are uncorrelated to the energy errors in $`199395`$. The correction factors for cross sections, $`f_{ϵ_{\mathrm{cms}}}^i`$, and the absolute corrections for forward-backward asymmetries, $`\alpha _{ϵ_{\mathrm{cms}}}^i`$, applied to the measurements to account for the spread of the centre-of-mass energy, $`ϵ_{\mathrm{cms}}`$, can be calculated to sufficient precision from Taylor expansions: $$f_{ϵ_{\mathrm{cms}}}^i=1\frac{1}{2}\frac{1}{\sigma _i^{\mathrm{th}}}\frac{^2\sigma _i^{\mathrm{th}}}{E^2}ϵ_{\mathrm{cms}}^2,\alpha _{ϵ_{\mathrm{cms}}}^i=\frac{1}{2}\left[\frac{^2A_{\mathrm{FB}}^{}{}_{,i}{}^{\mathrm{th}}}{E^2}+2\frac{1}{\sigma _i^{\mathrm{th}}}\frac{\sigma _i^{\mathrm{th}}}{E}\frac{A_{\mathrm{FB}}^{}{}_{,i}{}^{\mathrm{th}}}{E}\right]ϵ_{\mathrm{cms}}^2.$$ (44) The contributions to the covariance matrix from the uncertainty on the centre-of-mass energy spread, $`\mathrm{\Delta }ϵ_{\mathrm{cms}}`$, are then given by $`V_{ij}^{ϵ_{\mathrm{cms}}}`$ $`=`$ $`\mathrm{\Delta }_i^{ϵ_{\mathrm{cms}}}\mathrm{\Delta }_j^{ϵ_{\mathrm{cms}}}`$ $`\mathrm{with}\mathrm{\Delta }_i^{ϵ_{\mathrm{cms}}}`$ $`=`$ $`2\sigma _i^{\mathrm{th}}\left(f_{ϵ_{\mathrm{cms}}}^i1\right){\displaystyle \frac{\mathrm{\Delta }ϵ_{\mathrm{cms}}}{ϵ_{\mathrm{cms}}}}\mathrm{for}\mathrm{cross}\mathrm{sections},`$ (45) $`=`$ $`2\alpha _{ϵ_{\mathrm{cms}}}^i{\displaystyle \frac{\mathrm{\Delta }ϵ_{\mathrm{cms}}}{ϵ_{\mathrm{cms}}}}\mathrm{for}\mathrm{A}_{\mathrm{FB}}.`$ The spread of the centre-of-mass energy and its error for the $`199395`$ data sets are given in Table 2. The centre-of-mass energy spread used in Reference for the $`199092`$ data has been revised and the values and errors given in Reference are used. The error on the centre-of-mass energy spread is fully correlated between all data sets of $`199095`$, hence all elements of the covariance matrix receive a contribution. #### Error on $`𝒕`$-channel Contributions The last term in Equations 38 and 40 applies to $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}(\gamma )`$ cross section and $`A_{\mathrm{FB}}`$ measurements only. It accounts for the uncertainty on the calculation of $`t`$-channel and $`s/t`$-interference contributions: $$V_{ij}^t=\mathrm{\Delta }_k^t\mathrm{\Delta }_l^t.$$ (46) The indices $`k`$ and $`l`$ refer to the centre-of-mass energies of the data set $`i`$ and $`j`$ and the corresponding errors on cross sections and asymmetries, $`\mathrm{\Delta }^\mathrm{t}`$, are listed in Table 38 . Off-diagonal elements receive a contribution only if the two data sets are both below, on or above the peak. Cross section and asymmetry measurements are also connected this way. #### Constraints In fits using other results, $`\mathrm{\Omega }_m^\mathrm{c}`$, than cross section and $`A_{\mathrm{FB}}`$ measurements these additional measurements are added with their errors, $`\mathrm{\Delta }_m^\mathrm{c}`$, to the $`\chi ^2`$ function. This applies to our measurements of tau-polarisation, b-quark and quark charge asymmetries, $`R_\mathrm{b}`$ and $`\alpha _\mathrm{s}`$, as well as to the value of $`\mathrm{\Delta }\alpha _{\mathrm{had}}^{(5)}`$ and the measurement of $`m_\mathrm{t}`$ used in SM fits. Statistical and systematic errors are added in quadrature to obtain $`\mathrm{\Delta }_m^\mathrm{c}`$: $$\chi ^2=D^TV^1D+\underset{m}{}\left(\frac{\mathrm{\Omega }_m^\mathrm{c}\mathrm{\Omega }_m^{\mathrm{th}}}{\mathrm{\Delta }_m^\mathrm{c}}\right)^2.$$ (47) The small correlation of 8% between $`A_\mathrm{e}`$ and $`A_\tau `$ measured from tau-polarisation is included by means of a covariance matrix. The L3 Collaboration: M.Acciarri,<sup>26</sup> P.Achard,<sup>19</sup> O.Adriani,<sup>16</sup> M.Aguilar-Benitez,<sup>25</sup> J.Alcaraz,<sup>25</sup> G.Alemanni,<sup>22</sup> J.Allaby,<sup>17</sup> A.Aloisio,<sup>28</sup> M.G.Alviggi,<sup>28</sup> G.Ambrosi,<sup>19</sup> H.Anderhub,<sup>48</sup> V.P.Andreev,<sup>6,36</sup> T.Angelescu,<sup>12</sup> F.Anselmo,<sup>9</sup> A.Arefiev,<sup>27</sup> T.Azemoon,<sup>3</sup> T.Aziz,<sup>10</sup> P.Bagnaia,<sup>35</sup> L.Baksay,<sup>43</sup> A.Balandras,<sup>4</sup> R.C.Ball,<sup>3</sup> S.Banerjee,<sup>10</sup> Sw.Banerjee,<sup>10</sup> A.Barczyk,<sup>48,46</sup> R.Barillère,<sup>17</sup> L.Barone,<sup>35</sup> P.Bartalini,<sup>22</sup> M.Basile,<sup>9</sup> R.Battiston,<sup>32</sup> A.Bay,<sup>22</sup> F.Becattini,<sup>16</sup> U.Becker,<sup>14</sup> F.Behner,<sup>48</sup> L.Bellucci,<sup>16</sup> J.Berdugo,<sup>25</sup> P.Berges,<sup>14</sup> B.Bertucci,<sup>32</sup> B.L.Betev,<sup>48</sup> S.Bhattacharya,<sup>10</sup> M.Biasini,<sup>32</sup> A.Biland,<sup>48</sup> J.J.Blaising,<sup>4</sup> S.C.Blyth,<sup>33</sup> G.J.Bobbink,<sup>2</sup> A.Böhm,<sup>1</sup> L.Boldizsar,<sup>13</sup> B.Borgia,<sup>35</sup> D.Bourilkov,<sup>48</sup> M.Bourquin,<sup>19</sup> S.Braccini,<sup>19</sup> J.G.Branson,<sup>39</sup> V.Brigljevic,<sup>48</sup> F.Brochu,<sup>4</sup> I.C.Brock,<sup>33</sup> A.Buffini,<sup>16</sup> A.Buijs,<sup>44</sup> J.D.Burger,<sup>14</sup> W.J.Burger,<sup>32</sup> A.Button,<sup>3</sup> X.D.Cai,<sup>14</sup> M.Campanelli,<sup>48</sup> M.Capell,<sup>14</sup> G.Cara Romeo,<sup>9</sup> G.Carlino,<sup>28</sup> A.M.Cartacci,<sup>16</sup> J.Casaus,<sup>25</sup> G.Castellini,<sup>16</sup> F.Cavallari,<sup>35</sup> N.Cavallo,<sup>37</sup> C.Cecchi,<sup>32</sup> M.Cerrada,<sup>25</sup> F.Cesaroni,<sup>23</sup> M.Chamizo,<sup>19</sup> Y.H.Chang,<sup>50</sup> U.K.Chaturvedi,<sup>18</sup> M.Chemarin,<sup>24</sup> A.Chen,<sup>50</sup> G.Chen,<sup>7</sup> G.M.Chen,<sup>7</sup> H.F.Chen,<sup>20</sup> H.S.Chen,<sup>7</sup> G.Chiefari,<sup>28</sup> L.Cifarelli,<sup>38</sup> F.Cindolo,<sup>9</sup> C.Civinini,<sup>16</sup> I.Clare,<sup>14</sup> R.Clare,<sup>14</sup> G.Coignet,<sup>4</sup> A.P.Colijn,<sup>2</sup> N.Colino,<sup>25</sup> S.Costantini,<sup>5</sup> F.Cotorobai,<sup>12</sup> B.Cozzoni,<sup>9</sup> B.de la Cruz,<sup>25</sup> A.Csilling,<sup>13</sup> S.Cucciarelli,<sup>32</sup> T.S.Dai,<sup>14</sup> J.A.van Dalen,<sup>30</sup> R.D’Alessandro,<sup>16</sup> R.de Asmundis,<sup>28</sup> P.Déglon,<sup>19</sup> A.Degré,<sup>4</sup> K.Deiters,<sup>46</sup> D.della Volpe,<sup>28</sup> P.Denes,<sup>34</sup> F.DeNotaristefani,<sup>35</sup> A.De Salvo,<sup>48</sup> M.Diemoz,<sup>35</sup> D.van Dierendonck,<sup>2</sup> F.Di Lodovico,<sup>48</sup> C.Dionisi,<sup>35</sup> M.Dittmar,<sup>48</sup> A.Dominguez,<sup>39</sup> A.Doria,<sup>28</sup> M.T.Dova,<sup>18,♯</sup> D.Duchesneau,<sup>4</sup> D.Dufournaud,<sup>4</sup> P.Duinker,<sup>2</sup> I.Duran,<sup>40</sup> S.Dutta,<sup>10</sup> H.El Mamouni,<sup>24</sup> A.Engler,<sup>33</sup> F.J.Eppling,<sup>14</sup> F.C.Erné,<sup>2</sup> P.Extermann,<sup>19</sup> M.Fabre,<sup>46</sup> R.Faccini,<sup>35</sup> M.A.Falagan,<sup>25</sup> S.Falciano,<sup>35,17</sup> A.Favara,<sup>17</sup> J.Fay,<sup>24</sup> O.Fedin,<sup>36</sup> M.Felcini,<sup>48</sup> T.Ferguson,<sup>33</sup> F.Ferroni,<sup>35</sup> H.Fesefeldt,<sup>1</sup> E.Fiandrini,<sup>32</sup> J.H.Field,<sup>19</sup> F.Filthaut,<sup>17</sup> P.H.Fisher,<sup>14</sup> I.Fisk,<sup>39</sup> G.Forconi,<sup>14</sup> L.Fredj,<sup>19</sup> K.Freudenreich,<sup>48</sup> C.Furetta,<sup>26</sup> Yu.Galaktionov,<sup>27,14</sup> S.N.Ganguli,<sup>10</sup> P.Garcia-Abia,<sup>5</sup> M.Gataullin,<sup>31</sup> S.S.Gau,<sup>11</sup> S.Gentile,<sup>35,17</sup> N.Gheordanescu,<sup>12</sup> S.Giagu,<sup>35</sup> Z.F.Gong,<sup>20</sup> G.Grenier,<sup>24</sup> O.Grimm,<sup>48</sup> M.W.Gruenewald,<sup>8</sup> M.Guida,<sup>38</sup> R.van Gulik,<sup>2</sup> V.K.Gupta,<sup>34</sup> A.Gurtu,<sup>10</sup> L.J.Gutay,<sup>45</sup> D.Haas,<sup>5</sup> A.Hasan,<sup>29</sup> D.Hatzifotiadou,<sup>9</sup> T.Hebbeker,<sup>8</sup> A.Hervé,<sup>17</sup> P.Hidas,<sup>13</sup> J.Hirschfelder,<sup>33</sup> H.Hofer,<sup>48</sup> G. Holzner,<sup>48</sup> H.Hoorani,<sup>33</sup> S.R.Hou,<sup>50</sup> I.Iashvili,<sup>47</sup> V.Innocente,<sup>17</sup> B.N.Jin,<sup>7</sup> L.W.Jones,<sup>3</sup> P.de Jong,<sup>2</sup> I.Josa-Mutuberría,<sup>25</sup> R.A.Khan,<sup>18</sup> M.Kaur,<sup>18,♢</sup> M.N.Kienzle-Focacci,<sup>19</sup> D.Kim,<sup>35</sup> J.K.Kim,<sup>42</sup> J.Kirkby,<sup>17</sup> D.Kiss,<sup>13</sup> W.Kittel,<sup>30</sup> A.Klimentov,<sup>14,27</sup> A.C.König,<sup>30</sup> E.Koffeman,<sup>2</sup> A.Kopp,<sup>47</sup> V.Koutsenko,<sup>14,27</sup> M.Kräber,<sup>48</sup> R.W.Kraemer,<sup>33</sup> W.Krenz,<sup>1</sup> A.Krüger,<sup>47</sup> H.Kuijten,<sup>30</sup> A.Kunin,<sup>14,27</sup> P.Ladron de Guevara,<sup>25</sup> I.Laktineh,<sup>24</sup> G.Landi,<sup>16</sup> K.Lassila-Perini,<sup>48</sup> M.Lebeau,<sup>17</sup> A.Lebedev,<sup>14</sup> P.Lebrun,<sup>24</sup> P.Lecomte,<sup>48</sup> P.Lecoq,<sup>17</sup> P.Le Coultre,<sup>48</sup> H.J.Lee,<sup>8</sup> J.M.Le Goff,<sup>17</sup> R.Leiste,<sup>47</sup> E.Leonardi,<sup>35</sup> P.Levtchenko,<sup>36</sup> C.Li,<sup>20</sup> S.Likhoded,<sup>47</sup> C.H.Lin,<sup>50</sup> W.T.Lin,<sup>50</sup> F.L.Linde,<sup>2</sup> L.Lista,<sup>28</sup> Z.A.Liu,<sup>7</sup> W.Lohmann,<sup>47</sup> E.Longo,<sup>35</sup> Y.S.Lu,<sup>7</sup> W.Lu,<sup>31</sup> K.Lübelsmeyer,<sup>1</sup> C.Luci,<sup>17,35</sup> D.Luckey,<sup>14</sup> L.Lugnier,<sup>24</sup> L.Luminari,<sup>35</sup> W.Lustermann,<sup>48</sup> W.G.Ma,<sup>20</sup> M.Maity,<sup>10</sup> L.Malgeri,<sup>17</sup> A.Malinin,<sup>17</sup> C.Maña,<sup>25</sup> D.Mangeol,<sup>30</sup> P.Marchesini,<sup>48</sup> G.Marian,<sup>15</sup> J.P.Martin,<sup>24</sup> F.Marzano,<sup>35</sup> G.G.G.Massaro,<sup>2</sup> K.Mazumdar,<sup>10</sup> R.R.McNeil,<sup>6</sup> S.Mele,<sup>17</sup> L.Merola,<sup>28</sup> M.Merk,<sup>33</sup> M.Meschini,<sup>16</sup> W.J.Metzger,<sup>30</sup> M.von der Mey,<sup>1</sup> A.Mihul,<sup>12</sup> H.Milcent,<sup>17</sup> G.Mirabelli,<sup>35</sup> J.Mnich,<sup>17</sup> G.B.Mohanty,<sup>10</sup> P.Molnar,<sup>8</sup> B.Monteleoni,<sup>16,†</sup> T.Moulik,<sup>10</sup> G.S.Muanza,<sup>24</sup> F.Muheim,<sup>19</sup> A.J.M.Muijs,<sup>2</sup> M.Musy,<sup>35</sup> M.Napolitano,<sup>28</sup> F.Nessi-Tedaldi,<sup>48</sup> H.Newman,<sup>31</sup> T.Niessen,<sup>1</sup> A.Nisati,<sup>35</sup> H.Nowak,<sup>47</sup> G.Organtini,<sup>35</sup> A.Oulianov,<sup>27</sup> C.Palomares,<sup>25</sup> D.Pandoulas,<sup>1</sup> S.Paoletti,<sup>35,17</sup> A.Paoloni,<sup>35</sup> P.Paolucci,<sup>28</sup> R.Paramatti,<sup>35</sup> H.K.Park,<sup>33</sup> I.H.Park,<sup>42</sup> G.Pascale,<sup>35</sup> G.Passaleva,<sup>17</sup> S.Patricelli,<sup>28</sup> T.Paul,<sup>11</sup> M.Pauluzzi,<sup>32</sup> C.Paus,<sup>17</sup> F.Pauss,<sup>48</sup> D.Peach,<sup>17</sup> M.Pedace,<sup>35</sup> S.Pensotti,<sup>26</sup> D.Perret-Gallix,<sup>4</sup> B.Petersen,<sup>30</sup> D.Piccolo,<sup>28</sup> F.Pierella,<sup>9</sup> M.Pieri,<sup>16</sup> P.A.Piroué,<sup>34</sup> E.Pistolesi,<sup>26</sup> V.Plyaskin,<sup>27</sup> M.Pohl,<sup>19</sup> V.Pojidaev,<sup>27,16</sup> H.Postema,<sup>14</sup> J.Pothier,<sup>17</sup> N.Produit,<sup>19</sup> D.O.Prokofiev,<sup>45</sup> D.Prokofiev,<sup>36</sup> J.Quartieri,<sup>38</sup> G.Rahal-Callot,<sup>48,17</sup> M.A.Rahaman,<sup>10</sup> P.Raics,<sup>15</sup> N.Raja,<sup>10</sup> R.Ramelli,<sup>48</sup> P.G.Rancoita,<sup>26</sup> A.Raspereza,<sup>47</sup> G.Raven,<sup>39</sup> P.Razis,<sup>29</sup>D.Ren,<sup>48</sup> M.Rescigno,<sup>35</sup> S.Reucroft,<sup>11</sup> T.van Rhee,<sup>44</sup> S.Riemann,<sup>47</sup> K.Riles,<sup>3</sup> A.Robohm,<sup>48</sup> J.Rodin,<sup>43</sup> B.P.Roe,<sup>3</sup> L.Romero,<sup>25</sup> A.Rosca,<sup>8</sup> S.Rosier-Lees,<sup>4</sup> S.Roth,<sup>1</sup> J.A.Rubio,<sup>17</sup> D.Ruschmeier,<sup>8</sup> H.Rykaczewski,<sup>48</sup> S.Saremi,<sup>6</sup> S.Sarkar,<sup>35</sup> J.Salicio,<sup>17</sup> E.Sanchez,<sup>17</sup> M.P.Sanders,<sup>30</sup> M.E.Sarakinos,<sup>21</sup> C.Schäfer,<sup>17</sup> V.Schegelsky,<sup>36</sup> S.Schmidt-Kaerst,<sup>1</sup> D.Schmitz,<sup>1</sup> H.Schopper,<sup>49</sup> D.J.Schotanus,<sup>30</sup> G.Schwering,<sup>1</sup> C.Sciacca,<sup>28</sup> D.Sciarrino,<sup>19</sup> A.Seganti,<sup>9</sup> L.Servoli,<sup>32</sup> S.Shevchenko,<sup>31</sup> N.Shivarov,<sup>41</sup> V.Shoutko,<sup>27</sup> E.Shumilov,<sup>27</sup> A.Shvorob,<sup>31</sup> T.Siedenburg,<sup>1</sup> D.Son,<sup>42</sup> B.Smith,<sup>33</sup> P.Spillantini,<sup>16</sup> M.Steuer,<sup>14</sup> D.P.Stickland,<sup>34</sup> A.Stone,<sup>6</sup> H.Stone,<sup>34,†</sup> B.Stoyanov,<sup>41</sup> A.Straessner,<sup>1</sup> K.Sudhakar,<sup>10</sup> G.Sultanov,<sup>18</sup> L.Z.Sun,<sup>20</sup> H.Suter,<sup>48</sup> J.D.Swain,<sup>18</sup> Z.Szillasi,<sup>43,¶</sup> T.Sztaricskai,<sup>43,¶</sup> X.W.Tang,<sup>7</sup> L.Tauscher,<sup>5</sup> L.Taylor,<sup>11</sup> B.Tellili,<sup>24</sup> C.Timmermans,<sup>30</sup> Samuel C.C.Ting,<sup>14</sup> S.M.Ting,<sup>14</sup> S.C.Tonwar,<sup>10</sup> J.Tóth,<sup>13</sup> C.Tully,<sup>17</sup> K.L.Tung,<sup>7</sup>Y.Uchida,<sup>14</sup> J.Ulbricht,<sup>48</sup> U.Uwer,<sup>17</sup> E.Valente,<sup>35</sup> G.Vesztergombi,<sup>13</sup> I.Vetlitsky,<sup>27</sup> D.Vicinanza,<sup>38</sup> G.Viertel,<sup>48</sup> S.Villa,<sup>11</sup> M.Vivargent,<sup>4</sup> S.Vlachos,<sup>5</sup> I.Vodopianov,<sup>36</sup> H.Vogel,<sup>33</sup> H.Vogt,<sup>47</sup> I.Vorobiev,<sup>27</sup> A.A.Vorobyov,<sup>36</sup> A.Vorvolakos,<sup>29</sup> M.Wadhwa,<sup>5</sup> W.Wallraff,<sup>1</sup> M.Wang,<sup>14</sup> X.L.Wang,<sup>20</sup> Z.M.Wang,<sup>20</sup> A.Weber,<sup>1</sup> M.Weber,<sup>1</sup> P.Wienemann,<sup>1</sup> H.Wilkens,<sup>30</sup> S.X.Wu,<sup>14</sup> S.Wynhoff,<sup>17</sup> L.Xia,<sup>31</sup> Z.Z.Xu,<sup>20</sup> B.Z.Yang,<sup>20</sup> C.G.Yang,<sup>7</sup> H.J.Yang,<sup>7</sup> M.Yang,<sup>7</sup> J.B.Ye,<sup>20</sup> S.C.Yeh,<sup>51</sup> J.M.You,<sup>33</sup> An.Zalite,<sup>36</sup> Yu.Zalite,<sup>36</sup> Z.P.Zhang,<sup>20</sup> G.Y.Zhu,<sup>7</sup> R.Y.Zhu,<sup>31</sup> A.Zichichi,<sup>9,17,18</sup> G.Zilizi,<sup>43,¶</sup> M.Zöller.<sup>1</sup> * I. Physikalisches Institut, RWTH, D-52056 Aachen, FRG<sup>§</sup> III. Physikalisches Institut, RWTH, D-52056 Aachen, FRG<sup>§</sup> * National Institute for High Energy Physics, NIKHEF, and University of Amsterdam, NL-1009 DB Amsterdam, The Netherlands * University of Michigan, Ann Arbor, MI 48109, USA * Laboratoire d’Annecy-le-Vieux de Physique des Particules, LAPP,IN2P3-CNRS, BP 110, F-74941 Annecy-le-Vieux CEDEX, France * Institute of Physics, University of Basel, CH-4056 Basel, Switzerland * Louisiana State University, Baton Rouge, LA 70803, USA * Institute of High Energy Physics, IHEP, 100039 Beijing, China * Humboldt University, D-10099 Berlin, FRG<sup>§</sup> * University of Bologna and INFN-Sezione di Bologna, I-40126 Bologna, Italy * Tata Institute of Fundamental Research, Bombay 400 005, India * Northeastern University, Boston, MA 02115, USA * Institute of Atomic Physics and University of Bucharest, R-76900 Bucharest, Romania * Central Research Institute for Physics of the Hungarian Academy of Sciences, H-1525 Budapest 114, Hungary * Massachusetts Institute of Technology, Cambridge, MA 02139, USA * KLTE-ATOMKI, H-4010 Debrecen, Hungary * INFN Sezione di Firenze and University of Florence, I-50125 Florence, Italy * European Laboratory for Particle Physics, CERN, CH-1211 Geneva 23, Switzerland * World Laboratory, FBLJA Project, CH-1211 Geneva 23, Switzerland * University of Geneva, CH-1211 Geneva 4, Switzerland * Chinese University of Science and Technology, USTC, Hefei, Anhui 230 029, China * SEFT, Research Institute for High Energy Physics, P.O. Box 9, SF-00014 Helsinki, Finland * University of Lausanne, CH-1015 Lausanne, Switzerland * INFN-Sezione di Lecce and Universitá Degli Studi di Lecce, I-73100 Lecce, Italy * Institut de Physique Nucléaire de Lyon, IN2P3-CNRS,Université Claude Bernard, F-69622 Villeurbanne, France * Centro de Investigaciones Energéticas, Medioambientales y Tecnologícas, CIEMAT, E-28040 Madrid, Spain$`\mathrm{}`$ * INFN-Sezione di Milano, I-20133 Milan, Italy * Institute of Theoretical and Experimental Physics, ITEP, Moscow, Russia * INFN-Sezione di Napoli and University of Naples, I-80125 Naples, Italy * Department of Natural Sciences, University of Cyprus, Nicosia, Cyprus * University of Nijmegen and NIKHEF, NL-6525 ED Nijmegen, The Netherlands * California Institute of Technology, Pasadena, CA 91125, USA * INFN-Sezione di Perugia and Universitá Degli Studi di Perugia, I-06100 Perugia, Italy * Carnegie Mellon University, Pittsburgh, PA 15213, USA * Princeton University, Princeton, NJ 08544, USA * INFN-Sezione di Roma and University of Rome, “La Sapienza”, I-00185 Rome, Italy * Nuclear Physics Institute, St. Petersburg, Russia * INFN-Sezione di Napoli and University of Potenza, I-85100 Potenza, Italy * University and INFN, Salerno, I-84100 Salerno, Italy * University of California, San Diego, CA 92093, USA * Dept. de Fisica de Particulas Elementales, Univ. de Santiago, E-15706 Santiago de Compostela, Spain * Bulgarian Academy of Sciences, Central Lab. of Mechatronics and Instrumentation, BU-1113 Sofia, Bulgaria * Laboratory of High Energy Physics, Kyungpook National University, 702-701 Taegu, Republic of Korea * University of Alabama, Tuscaloosa, AL 35486, USA * Utrecht University and NIKHEF, NL-3584 CB Utrecht, The Netherlands * Purdue University, West Lafayette, IN 47907, USA * Paul Scherrer Institut, PSI, CH-5232 Villigen, Switzerland * DESY, D-15738 Zeuthen, FRG * Eidgenössische Technische Hochschule, ETH Zürich, CH-8093 Zürich, Switzerland * University of Hamburg, D-22761 Hamburg, FRG * National Central University, Chung-Li, Taiwan, China * Department of Physics, National Tsing Hua University, Taiwan, China * Supported by the German Bundesministerium für Bildung, Wissenschaft, Forschung und Technologie * Supported by the Hungarian OTKA fund under contract numbers T019181, F023259 and T024011. * Also supported by the Hungarian OTKA fund under contract numbers T22238 and T026178. * Supported also by the Comisión Interministerial de Ciencia y Tecnología. * Also supported by CONICET and Universidad Nacional de La Plata, CC 67, 1900 La Plata, Argentina. * Also supported by Panjab University, Chandigarh-160014, India. * Supported by the National Natural Science Foundation of China. * Deceased.
warning/0002/quant-ph0002057.html
ar5iv
text
# On the Complexity of Quantum ACC ## 1 Introduction Advances in quantum computation in the last decade have been among the most notable in theoretical computer science. This is due to the surprising improvements in the efficiency of solving several fundamental combinatorial problems using quantum mechanical methods in place of their classical counterparts. These advances led to considerable efforts in finding new efficient quantum algorithms for classical problems and in developing a complexity theory of quantum computation. While most of the original results in quantum computation were developed using quantum Turing machines, they can also be formulated in terms of quantum circuits, which yield a more natural model of quantum computation. For example, Shor has shown that quantum circuits can factor integers more efficiently than any known classical algorithm for factoring. And quantum circuits have been shown (see Yao ) to provide a universal model for quantum computation. In the classical setting, small depth circuits are considered a good model for parallel computing. Constant-depth circuits, corresponding to constant parallel time, are of central importance. For example, constant-depth circuits of AND, OR and NOT gates of polynomial size (called AC<sup>(0)</sup> circuits) can add and subtract binary numbers. The class ACC extends AC<sup>(0)</sup> by allowing modular counting gates. The class TC<sup>(0)</sup>, consisting of constant-depth threshold circuits, can compute iterated multiplication. In studying quantum circuits, it is natural to consider the power of small depth circuit families. Quantum circuit models analogous to the central classical circuit classes have recently been studied by Moore and Nilsson and Moore . They investigated the properties of classes of quantum operators QAC$`{}_{wf}{}^{}{}_{}{}^{(0)}`$, QACC$`[q]`$, and QNC defined to be analogous to and to contain their classical counterparts. This paper is a contribution to this line of research. For example, a quantum analog of AC<sup>(0)</sup>, defined by Moore and denoted QAC$`{}_{wf}{}^{}{}_{}{}^{(0)}`$, is the class of families of operators which can be built out of products of constantly many layers consisting of polynomial-sized tensor products of one-qubit gates (analogous to NOT’s), Toffoli gates (analogous to AND’s and OR’s) and fan-out gates<sup>1</sup><sup>1</sup>1The subscript “$`wf`$” in the notation denotes “with fan-out.” The idea of fan-out in the quantum setting is subtle, as will be made clearer later in this paper. See Moore for a more in-depth discussion.. An analog of ACC$`[q]`$ (i.e., ACC circuit families only allowing Mod<sub>q</sub> gates) is QACC$`[q]`$, defined similarly to QAC$`{}_{wf}{}^{}{}_{}{}^{(0)}`$, but replacing the fan-out gates with quantum $`\mathrm{Mod}_q`$ gates (which we denote as $`\mathrm{MOD}_q`$). QACC is the same class but we allow $`\mathrm{MOD}_q`$ gates for every $`q`$. Moore proves the surprising result QAC$`{}_{wf}{}^{}{}_{}{}^{(0)}`$$`=`$ QACC$`[2]`$ $`=`$ QACC. This is in sharp contrast to the classical result of Smolensky that says ACC$`{}_{}{}^{(0)}[q]`$ ACC$`{}_{}{}^{(0)}[p]`$ for any pair of distinct primes $`q,p`$, which implies that for any prime $`p`$, AC$`{}_{}{}^{(0)}`$ ACC$`{}_{}{}^{(0)}[p]`$ ACC. This result showed that parity gates are as powerful as any other mod gates in QACC, but left open the complexity of $`\mathrm{MOD}_q`$ gates for $`q>2`$. In , Moore conjectured that QACC $``$ QACC$`[q]`$ for odd $`q`$. In this paper, we provide the missing ingredients to show that in fact QACC$`=`$ QACC$`[q]`$ for any $`q2`$. Moore’s result showed that parity is as good as any other $`\mathrm{MOD}_q`$ gate; our result further shows that any $`\mathrm{MOD}_q`$ gate is as good as any other. The main technical contribution is the application of the Quantum Fourier Transform (using complex $`q^{th}`$ roots of unity), and encodings of base $`q`$ digits using qubits. We also develop methods for proving upper bounds for language classes related to QACC. Our methods result in upper bounds for restricted QACC circuits. Roughly speaking, we show that QACC is no more powerful than P/Poly provided that a layer of “wire-crossings” in the QACC operator can be written as log many compositions of Kronecker products of controlled-not gates. We call this class QACC$`{}_{}{}^{\mathrm{log}}{}_{pl}{}^{}`$, where the “pl” is for this planarity condition. We show if one further restricts attention to the case where the number of multi-line gates (gates whose input is more than 1 qubit) is log-bounded then the circuits are no more powerful than TC<sup>(0)</sup>. We call this class QACC$`{}_{}{}^{\mathrm{log}}{}_{gates}{}^{}`$. These results hold for arbitrary complex amplitudes in the QACC circuits. To be more precise, it is necessary to show how a class of operators in QACC can define a language, as usually considered in complexity theory. In this paper, we define classes of languages EQACC, NQACC, and BQACC based on the expectation of observing a certain state after applying the QACC operator to the input state. For example, the class NQACC corresponds to the case where $`x`$ is in the language if the expectation of the observed state after applying the QACC operator is non-zero. This is analogous to the definition of the class NQP in Fenner et al. . In this paper, we show that NQACC$`{}_{}{}^{\mathrm{log}}{}_{gates}{}^{}`$is in TC<sup>(0)</sup> and NQACC$`{}_{}{}^{\mathrm{log}}{}_{pl}{}^{}`$is in P/poly. Although the proof uses some of the techniques developed by Yamakami and Yao to show that NQP$`{}_{C}{}^{}=`$ co-C<sub>=</sub>P, the small depth circuit case presents technical challenges not present in their setting. In particular, given a QACC operator built out of layers $`M_1,\mathrm{},M_t`$ and an input state $`|x,0^{p(n)}`$, we must show that a TC<sup>(0)</sup> circuit can keep track of the amplitudes of each possible resulting state as each layer is applied. After all layers have been applied, the TC<sup>(0)</sup> circuit then needs to be able to check that the amplitude of one possible state is non-zero. Unfortunately, there could be exponentially many states with non-zero amplitudes after applying a layer. To handle this problem we introduce the idea of a “tensor-graph,” a new way to represent a collection of states. We can extract from these graphs (via TC<sup>(0)</sup> or P/poly computations) whether the amplitude of any particular vector is non-zero. The exponential growth in the number of states is one of the primary obstacles to proving that all of NQACC is in TC<sup>(0)</sup> (or even P/Poly), and thus the tensor graph formalism represents a significant step towards such an upper bound. The reason the bounds apply only in the restricted cases is that although tensor graphs can represent any QACC operator, in the case of operators with layers that might do arbitrary permutations, the top-down approach we use to compute a desired amplitude from the graph no longer seems to work. We feel that it is likely that the amplitude of any vector in a tensor graph can be written as a polynomial product of a polynomial sum in some extension algebra of the ones we work with in this paper, in which case it is quite likely it can be evaluated in TC<sup>(0)</sup>. Another important obstacle to obtaining a TC<sup>(0)</sup> upper bound is that one needs to be able to add and multiply a polynomial number of complex amplitudes that may appear in a QACC computation. We solve this problem. It reduces to adding and multiplying polynomially many elements of a certain transcendental extension of the rational numbers. We show that in fact TC<sup>(0)</sup> is closed under iterated addition and multiplication of such numbers (Lemma 4.1 below). This result is of independent interest, and our application of tensor-graphs and these closure properties of TC<sup>(0)</sup> may prove useful in further investigations of small-depth quantum circuits. We now discuss the organization of the rest of this paper. In the next section we introduce the definitions and notations we use in this paper. Then in the following section we prove QACC$`[q]`$ $`=`$ QACC. Finally, in the last section, we prove the TC<sup>(0)</sup> and P/poly upper bounds for the restricted classes discussed above. ## 2 Preliminaries In this section we define the gates used as building blocks for our quantum circuits. Classes of operators built out of these gates are then defined. We define language classes that can be determined by these operators and give a couple definitions from algebra. Lastly, some closure properties of TC<sup>(0)</sup> are described. ###### Definition 2.1 By a one-qubit gate we mean an operator from the group $`U(2)`$. Let $`U=\left(\begin{array}{cc}u_{00}& u_{01}\\ u_{10}& u_{11}\end{array}\right)U(2)`$. $`_m(U)`$ is defined as: $`_0(U)=U`$ and for $`m>0`$, $`_m(U)`$ is $$_m(U)(|\stackrel{}{x},y)=\{\begin{array}{cc}u_{y0}|\stackrel{}{x},0+u_{y1}|\stackrel{}{x},1\hfill & \text{if }_{k=1}^mx_k=1\hfill \\ |\stackrel{}{x},y\hfill & \text{otherwise}\hfill \end{array}$$ Let $`X=\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right)`$. A Tofolli gate is a $`_m(X)`$ gate for some $`m0`$. A controlled-not gate is a $`_1(X)`$ gate. An (m-)spaced controlled-not gate is an operator that maps $`|y_1,\mathrm{},y_m,x`$ to $`|xy_1,y_2\mathrm{},y_m,x`$ or $`|y_1,\mathrm{},y_m,x`$ to $`|x,y_1\mathrm{},y_{m1},y_mx`$ An (m-ary) fan out gate $`F`$ is an operator that maps from $`|y_1,\mathrm{},y_m,x`$ to $`|xy_1,\mathrm{},xy_m,x`$. A $`\mathrm{MOD}_{q,r}`$ gate is an operator that maps $`|y_1,\mathrm{},y_m,x`$ to $`|y_1,\mathrm{},y_m,x(y_imodqr)`$. We use the following graphical notation for parity (i.e., $`\mathrm{MOD}_2`$) or, in the case of $`n=1`$, for controlled-not: and for $`\mathrm{MOD}_q`$: As discussed in , the no-cloning theorem of quantum mechanics makes it difficult to directly fan out qubits in constant depth (although constant fan-out is no problem). Thus it is necessary to define the operator $`F`$ as in the above definition; refer to for further details. Also, in the literature it is frequently the case that one says a given operator $`M`$ on $`|y_1,\mathrm{},y_m`$ can be written as a tensor product of certain gates. What is meant is that there is an permutation operator $`\mathrm{\Pi }`$ ( a map $`|y_1,\mathrm{},y_m`$ to $`|y_{\pi (1)},\mathrm{},y_{\pi (m)}`$ for some permutation $`\pi `$) such that $$M|y_1,\mathrm{}y_m=\mathrm{\Pi }_j^nM_j\mathrm{\Pi }^1|y_1,\mathrm{}y_m$$ where the $`M_i`$’s are our base gates, i.e., those gates for which no inherent ordering on the $`y_i`$ is assumed a priori. Since it is important to keep track of such details in our upper bounds proofs, we will always use Kronecker products of the form $`_j^nM_j`$ without unspoken permutations. Nevertheless, being able to do permutation operators (not conjugation by a permutation) intuitively allows our circuits to simulate classical wire crossings. To handle permutations, we allow our circuits to have controlled-not layers. A controlled-not layer is a gate which performs, in one step, controlled-not’s between an arbitrary collection of disjoint pairs of lines in its domain. That is, it performs $`\mathrm{\Pi }_j^n_1(X)\mathrm{\Pi }^1`$ for some permutation operator $`\mathrm{\Pi }`$. Moore Nilsson show that any permutation can be written as a finite product of controlled-not layers. We say a controlled-not layer is log-depth if it can be written as the composition of log many matrices each of which is the Kronecker product of identities and spaced controlled-not gates. $`M^n`$ is the $`n`$-fold Kronecker product of $`M`$ with itself. The next definitions are based on Moore . ###### Definition 2.2 QAC<sup>(k)</sup> is the class of families $`\{F_n\}`$, where $`F_n`$ is in $`U(2^{n+p(n)})`$, $`p`$ a polynomial, and each $`F_n`$ is writable as a product of $`O(\mathrm{log}^kn)`$ layers, where a layer is a Kronecker product of one-qubit gates and Toffoli gates or is a controlled-not layer. Also for all $`n`$ the number of distinct types of one qubit gates used must be fixed. QACC$`{}_{}{}^{(k)}[q]`$ is the same as QAC<sup>(k)</sup> except we also allow $`\mathrm{MOD}_q`$ gates. QACC$`{}_{}{}^{(k)}=_q`$QACC$`{}_{}{}^{(k)}[q]`$. QAC$`{}_{wf}{}^{}{}_{}{}^{(k)}`$ is the same as QAC<sup>(k)</sup> but we also allow fan-out gates. QACC is defined as QACC<sup>(0)</sup> and QACC$`[q]`$ is defined as QACC$`{}_{}{}^{(0)}[q]`$. QACC$`{}_{}{}^{\mathrm{log}}{}_{pl}{}^{}`$is QACC restricted to log-depth controlled not layers. QACC$`{}_{}{}^{\mathrm{log}}{}_{gates}{}^{}`$is QACC restricted so that the total number of multi-line gates in all layers is log-bounded. If $`𝒞`$ is one of the above classes, then $`𝒞_K`$ are the families in $`𝒞`$ with coefficients restricted to $`K`$. Let $`\{F_n\}`$ and $`\{G_n\}`$, $`G_n,F_nU(2^n)`$ be families of operators. We say $`\{F_n\}`$ is QAC<sup>(0)</sup> reducible to $`\{G_n\}`$ if there is a family $`\{R_n\}`$, $`R_nU(2^{n+p(n)})`$ of QAC<sup>(0)</sup> operators augmented with operators from $`\{G_n\}`$ such that for all $`n`$, $`𝐱,𝐲\{0,1\}^n`$, there is a setting of $`z_1,\mathrm{},z_{p(n)}\{0,1\}`$ for which $`𝐲|F_n|𝐱=𝐲,𝐳|R_n|𝐱,𝐳`$. Operator families are QAC<sup>(0)</sup> equivalent if they are QAC<sup>(0)</sup> reducible to each other. If $`𝒞_1`$ and $`𝒞_2`$ are families of QAC<sup>(0)</sup> equivalent operators, we write $`𝒞_1=𝒞_2`$. We refer to the $`z_i`$’s above as “auxiliary bits” (called “ancillae” in ). Note that in proving QAC<sup>(0)</sup> equivalence, the auxiliary bits must be returned to their original values in a computation. It follows for any $`\{F_n\}`$ QAC<sup>(0)</sup> that $`F_n`$ is writable as a product of finite number of layers. Moore shows QAC$`{}_{wf}{}^{(0)}=`$ QACC$`[2]`$ $`=`$ QACC. Moore places no restriction on the number of distinct types of one-qubit gates used in a given family of operators. We do this so that the number of distinct amplitudes which appear in matrices in a layer is fixed with respect to $`n`$. This restriction arises implicitly in the quantum Turing machine case of the upper bounds proofs in Fenner, et al. and Yamakami and Yao . Also, it seems fairly natural since in the classical case one builds circuits using a fixed number of distinct gate types. Our classes are, thus, more “uniform” than Moore’s. We now define language classes based on our classes of operator families. ###### Definition 2.3 Let $`𝒞`$ be a class of families of $`U(2^{n+p(n)})`$ operators where $`p`$ is a polynomial and $`n=|x|`$. 1. E$`𝒞`$ is the class of languages $`L`$ such that for some $`\{F_n\}𝒞`$ and $`\{\stackrel{}{z}_n|\}=\{z_{n,1},\mathrm{},z_{n,n+p(n)}|\}`$ a family of states, $`m:=|\stackrel{}{z}_n|F_n|x,0^{p(n)}|^2`$ is $`1`$ or $`0`$ and $`xL`$ iff $`m=1`$. 2. N$`𝒞`$ is the class of languages $`L`$ such that for some $`\{F_n\}𝒞`$ and $`\{\stackrel{}{z}_n|\}`$ a family of states, $`xL`$ iff $`|\stackrel{}{z}_n|F_n|x,0^{p(n)}|^2>0`$. 3. B$`𝒞`$ is the class of languages $`L`$ so that for some $`\{F_n\}𝒞`$ and $`\{\stackrel{}{z}|\}`$, $`xL`$ if $`|\stackrel{}{z}_n|F_n|x,0^{p(n)}|^2>3/4`$ and $`xL`$ if $`|\stackrel{}{z}_n|F_n|x,0^{p(n)}|^2<1/4`$ . It follows E$`𝒞`$ N$`𝒞`$ and E$`𝒞`$ B$`𝒞`$. We frequently will omit the ‘$``$’ when writing a class, so E$``$QACC is written as EQACC. Let $`|\mathrm{\Psi }:=F_n|x,0^{p(n)}`$. Notice that $`|\stackrel{}{z}_n|F_n|x,0^{p(n)}|^2=\mathrm{\Psi }|P_{|\stackrel{}{z}_n}|\mathrm{\Psi }`$, where $`P_{|\stackrel{}{z}_n}`$ is the projection matrix onto $`|\stackrel{}{z}_n`$. We could allow in our definitions measurements of up to polynomially many such projection observables and not affect our results below. However, this would shift the burden of the computation in some sense away from the QACC operator and instead onto preparation of the observable. Next are some variations on familiar definitions from algebra. ###### Definition 2.4 Let $`k>0`$. A subset $`\{\beta _i\}_{1ik}`$ of $`𝐂`$ is linearly independent if $`_{i=1}^ka_i\beta _i0`$ for any $`(a_1,\mathrm{},a_k)𝐐^k\{\stackrel{}{0}^k\}`$. A set $`\{\beta _i\}_{1ik}`$ is algebraically independent if the only $`p𝐐[x_1,\mathrm{},x_k]`$ with $`p(\beta _1,\mathrm{},\beta _k)=0`$ is the zero polynomial. We now briefly mention some closure properties of TC<sup>(0)</sup> computable functions that are useful in proving NQACC$`{}_{}{}^{\mathrm{log}}{}_{gates}{}^{}`$$``$ TC<sup>(0)</sup>. For proofs of the statements in the next lemma see . ###### Lemma 2.5 (1) TC<sup>(0)</sup> functions are closed under composition. (2) The following are TC<sup>(0)</sup> computable: $`x+y`$, $`x\stackrel{.\text{.}}{}y:=xy`$ if $`xy>0`$ and $`0`$ otherwise, $`|x|:=\mathrm{log}_2(x+1)`$, $`xy`$, $`x/y`$, $`2^{\mathrm{min}(i,p(|x|)}`$, and $`cond(x,y,z):=y`$ if $`x>0`$ and $`z`$ otherwise. (3) If $`f(i,x)`$ is a TC<sup>(0)</sup> computable then $`_{k=0}^{p(|x|)}f(k,x)`$, $`_{k=0}^{p(|x|)}f(k,x)`$, $`ip(|x|)(f(i,x)=0)`$, $`ip(|x|)(f(i,x)=0)`$, and $`\mu ip(|x|)(f(i,x)=0):=`$ the least $`i`$ such that $`f(i,x)=0`$ or $`p(x)+1`$ otherwise, are TC<sup>(0)</sup> computable. We drop the $`\mathrm{min}`$ from the $`2^{\mathrm{min}(i,p(|x|))}`$ when it is obvious a suitably large $`p(|x|)`$ can be found. We define $`max(x,y):=cond(1\stackrel{.\text{.}}{}(y\stackrel{.\text{.}}{}x)),x,y)`$ and define $`max_{ip(|x|)}(f(i))`$ $`:=`$ $`(\mu i`$ $`p(|x|)`$ $`)(jp(|x|)(f(j)\stackrel{.\text{.}}{}f(i)=0)`$ Using the above functions we describe a way to do sequence coding in TC<sup>(0)</sup>. Let $`\beta _{|t|}(x,w):=(w\stackrel{.\text{.}}{}w/2^{(x+1)|t|}2^{(x+1)|t|})/2^{x|t|}.`$ The function $`\beta _{|t|}`$ is useful for block coding. Roughly, $`\beta _{|t|}`$ first gets rid of the bits after the $`(x+1)|t|`$th bit then chops off the low order $`x|t|`$ bits. Let $`B=2^{|\mathrm{max}(x,y)|}`$, so that $`B`$ is longer than either $`x`$ or $`y`$. Hence, we code pairs as $`x,y:=(B+y)2B+B+x`$, and projections as $`(w)_1:=\beta _{\frac{1}{2}|w|\stackrel{.\text{.}}{}1}(0,\beta _{\frac{1}{2}|w|}(0,w))`$ and $`(w)_2:=\beta _{\frac{1}{2}|w|\stackrel{.\text{.}}{}1}(0,\beta _{\frac{1}{2}|w|}(1,w))`$. We can encode a poly-length, TC<sup>(0)</sup> computable sequence of numbers $`f(1),\mathrm{},f(k)`$ as the pair $`_i^k(f(i)2^{im}),m`$ where $`m:=|f(\mathrm{max}_i(f(i)))|+1`$. We then define the function which projects out the $`i`$th member of a sequence as $`\beta (i,w):=\beta _{(w)_2}(i,w)`$. We can code integers using the positive natural numbers by letting the negative integers be the odd natural numbers and the positive integers be the even natural numbers. TC<sup>(0)</sup> can use the TC<sup>(0)</sup> circuits for natural numbers to compute both the polynomial sum and polynomial product of a sequence of TC<sup>(0)</sup> definable integers. It can also compute the rounded quotient of two such integers. For instance, to do a polynomial sum of integers, compute the natural number which is the sum of the positive numbers in the sum using $`cond`$ and our natural number iterated addition circuit. Then compute the natural number which is the sum of the negative numbers in the sum. Use the subtraction circuit to subtract the smaller from the larger number and multiply by two. One is then added if the number should be negative. For products, we compute the product of the natural numbers which results by dividing each integer code by two and rounding down. We multiply the result by two. We then sum the number of terms in our product which were negative integers. If this number is odd we add one to the product we just calculated. Finally, division can be computed using the Taylor expansion of $`1/x`$. ## 3 QACC\[$`q`$\] In this section, we show QACC\[$`q`$\]$`=`$QACC for any $`q2`$. Let $`q𝐍`$, $`q2`$ be fixed throughout this discussion. Consider quantum states labelled by digits in $`D=\{0,\mathrm{},q1\}`$. By analogy with “qubit,” we refer to a state of the form, $$\underset{k=0}{\overset{q1}{}}c_k|k$$ with $`_k|c_k|^2=1`$ as a “qudigit.” Direct products of the basis states will be labelled by lists of eigenvalues, e.g., $`|x|y`$ is denoted as $`|x,y`$. We define three important operations on qudigits. The $`n`$-ary modular addition operator $`M_q`$ acts as follows: $$M_q|x_1,\mathrm{},x_n,b=|x_1,\mathrm{}x_n,(b+x_1+\mathrm{}+x_n)modq$$ The gate is represented graphically as in the following figure: Since $`M_q`$ merely permutes the states, it is clear that it is unitary. Similarly, the $`n`$-ary unitary base $`q`$ fanout operator $`F_q`$ acts as, $$F_q|x_1,\mathrm{}x_n,b=|(x_1+b)modq,\mathrm{}(x_n+b)modq,b.$$ We write $`F`$ for $`F_2`$, since it is the “standard” fan-out gate introduced by Moore (see Definition 2.1). Note that $`M_q^1=M_q^{q1}`$ and $`F_q^1=F_q^{q1}`$. Finally, the Quantum Fourier Transform $`H_q`$ (which generalizes the Hadamard transform $`H`$ on qubits) acts on a single qudigit as, $$H_q|a=\frac{1}{\sqrt{q}}\underset{b=0}{\overset{q1}{}}\zeta ^{ab}|b,$$ where $`\zeta =e^{\frac{2\pi i}{q}}`$ is a primitive complex $`q^{th}`$ root of unity. It is easy to see that $`H_q`$ is unitary, via the fact that $`_{\mathrm{}=0}^{q1}\zeta ^a\mathrm{}=0`$ iff $`a0modq`$. The first observation is that, analogous to parity and fanout for Boolean inputs, the operators $`M_q`$ and $`F_q`$ are “conjugates” in the following sense. ###### Proposition 3.1 $`M_q=(H_q^{(n+1)})^1F_q^1H_q^{(n+1)}.`$ Proof. We apply the operators $`H_q^{(n+1)}`$, $`F_q^1`$, and $`(H_q^{(n+1)})^1`$ in that order to the state $`|x_1,\mathrm{},x_n,b`$, and check that the result has the same effect as $`M_q`$. The operator $`H_q^{(n+1)}`$ simply applies $`H_q`$ to each of the $`n+1`$ qudigits of $`|x_1,\mathrm{},x_n,b`$, which yields, $`{\displaystyle \frac{1}{q^{\frac{(n+1)}{2}}}}{\displaystyle \underset{𝐲D^n}{}}{\displaystyle \underset{a=0}{\overset{q1}{}}}\zeta ^{𝐱𝐲+ab}|y_1,\mathrm{},y_n,a,`$ where $`𝐲`$ is a compact notation for $`y_1,\mathrm{},y_n`$, and $`𝐱𝐲`$ denotes $`_{i=1}^nx_iy_i`$. Then applying $`F_q^1`$ to the above state yields, $`{\displaystyle \frac{1}{q^{\frac{(n+1)}{2}}}}{\displaystyle \underset{𝐲D^n}{}}{\displaystyle \underset{a=0}{\overset{q1}{}}}`$ $`\zeta `$ <sup>x⋅y+ab</sup> $`|(`$ $`y_1`$ $`a)modq,\mathrm{},(y_na)modq,a.`$ By a change of variable, the above can be re-written as, $`{\displaystyle \frac{1}{q^{\frac{(n+1)}{2}}}}{\displaystyle \underset{𝐲D^n}{}}{\displaystyle \underset{a=0}{\overset{q1}{}}}\zeta ^{_{i=1}^nx_i(y_i+a)+ab}|y_1,\mathrm{},y_n,a`$ Finally, applying $`(H_q^{(n+1)})^1`$ to the above undoes the Fourier transform and puts the coefficient of $`a`$ in the exponent into the last slot of the state. The result is, $`(H_q^{(n+1)})^1F_q^1H_q^{(n+1)}|x_1,\mathrm{},x_n,b=`$ $`|x_1,\mathrm{},x_n,(b+x_1+\mathrm{}+x_n)modq,`$ which is exactly what $`M_q`$ would yield. . We now describe how the operators $`M_q`$, $`F_q`$ and $`H_q`$ can be modified to operate on registers consisting of qubits rather than qudigits. Firstly, we encode each digit using $`\mathrm{log}q`$ bits. Thus, for example, when $`q=3`$, the basis states $`|0,|1`$ and $`|2`$ are represented by the two-qubit registers $`|00,|01`$ and $`|10`$, respectively. Note that there remains one state (in the example, $`|11`$) which does not correspond to any of the qudigits. In general, there will be $`2^{\mathrm{log}q}q`$ such “non-qudigit” states. $`M_q`$, $`F_q`$ and $`H_q`$ can now be defined to act on qubit registers, as follows. Consider a state $`|x`$ where $`x`$ is a number represented as $`m`$ bits (i.e., an $`m`$-qubit register). If $`m<\mathrm{log}q`$, then $`H_q`$ leaves $`|x`$ unaffected. If $`0xq1`$ (where here we are identifying $`x`$ with the number it represents), then $`H_q`$ acts exactly as one expects, namely, $`H_q|x=(1/\sqrt{q})_{y=0}^{q1}\zeta ^{xy}|y.`$ If $`xq`$, again $`H_q`$ leaves $`|x`$ unchanged. Since the resulting transformation is a direct sum of unit matrices and matrices of the form of $`H_q`$ as it was originally set down, the result is a unitary transformation. $`M_q`$ and $`F_q`$ can be defined to operate similarly on $`m`$-qubit registers for any $`m`$: Break up the $`m`$ bits into blocks of $`\mathrm{log}q`$ bits. If $`m`$ is not divisible by $`\mathrm{log}q`$, then $`M_q`$ and $`F_q`$ do not affect the “remainder” block that contains fewer than $`\mathrm{log}q`$ bits. Likewise, in a quantum register $`|x_1,\mathrm{},x_n`$ where each of the $`x_i`$’s (with the possible exception of $`x_n`$) are $`\mathrm{log}q`$-bit numbers, $`M_q`$ and $`F_q`$ operate on the blocks of bits $`x_1,\mathrm{},x_n`$ exactly as expected, except that there is no affect on the “non-qudigit” blocks (in which $`x_iq`$), or on the (possibly) one remainder block for which $`|x_n|<\mathrm{log}q`$. Since $`M_q`$ and $`F_q`$ operate exactly as they did originally on blocks representing qudigits, and like unity for non-qudigit or remainder blocks, it is clear that they remain unitary. Henceforth, $`M_q`$, $`F_q`$, and $`H_q`$ should be understood to act on qubit registers as described above. Nevertheless, it will usually be convenient to think of them as acting on qudigit registers consisting of $`\mathrm{log}q`$ qubits in each. ###### Lemma 3.2 $`F_q`$ and $`M_q`$ are QAC<sup>(0)</sup>-equivalent. Proof. By Barenco et al. , any fixed dimension unitary matrix can be computed in fixed depth using one-qubit gates and controlled nots. Hence $`H_q`$ can be computed in QAC<sup>(0)</sup>, as can $`H_q^{(n+1)}`$. The result now follows immediately from Proposition 3.1. . The classical Boolean $`\mathrm{Mod}_q`$-function on $`n`$ bits is defined so that $`\mathrm{Mod}_q(x_1,\mathrm{},x_n)=1`$ $`\mathrm{iff}`$ $`_{i=1}^nx_i0(modq).`$ (The more common definition sets it to 1 if $`_{i=1}^nx_i`$ is not divisible by $`q`$, but this convention is less convenient in this setting, and is not important technically either.) We also define $`\mathrm{Mod}_{q,r}(x_1,\mathrm{},x_n)`$ to output 1 iff $`_{i=1}^nx_ir(modq)`$. Note that $`\mathrm{Mod}_q=\mathrm{Mod}_{q,0}`$. Reversible, quantum versions of these functions can also be defined. The operator $`\mathrm{MOD}_{q,r}`$ on $`n+1`$ qubits has the following effect: $$|x_1,\mathrm{},x_n,b|x_1,\mathrm{},x_n,b\mathrm{Mod}_{q,r}(x_1,\mathrm{},x_n).$$ We write $`\mathrm{MOD}_{q,0}`$ as $`\mathrm{MOD}_q`$. Since negation is built into the output (via the exclusive OR), it is easy to simulate negations using $`\mathrm{MOD}_{q,r}`$ gates. For example, by setting $`b=1`$, we can compute $`\neg \mathrm{Mod}_{q,r}`$. More generally, using one auxiliary bit, it is possible to simulate “$`\neg \mathrm{MOD}_{q,r}`$,” defined so that, $$|x_1,\mathrm{},x_n,b|x_1,\mathrm{},x_n,b(\neg \mathrm{Mod}_{q,r}(x_1,\mathrm{},x_n)),$$ using just $`\mathrm{MOD}_{q,r}`$ and a controlled-NOT gate. Thus $`\mathrm{MOD}_{q,r}`$ and $`\neg \mathrm{MOD}_{q,r}`$ are QAC<sup>(0)</sup>-equivalent. Moore’s version of $`\mathrm{MOD}_q`$ is our $`\neg \mathrm{MOD}_q`$. Observe that $`\mathrm{MOD}_{q,r}^1=\mathrm{MOD}_{q,r}`$. ###### Lemma 3.3 $`\mathrm{MOD}_q`$ and $`M_q`$ are QAC<sup>(0)</sup>-equivalent. Proof. First note that $`\mathrm{MOD}_q`$ and $`\mathrm{MOD}_{q,r}`$ are equivalent, since a $`\mathrm{MOD}_{q,r}`$ gate can be simulated by a $`\mathrm{MOD}_q`$ gate with $`qr`$ extra inputs set to the constant 1. Hence we can freely use $`\mathrm{MOD}_{q,r}`$ gates in place of $`\mathrm{MOD}_q`$ gates. It is easy to see that, given an $`M_q`$ gate, we can simulate a $`\mathrm{MOD}_q`$ gate. Applying $`M_q`$ to $`n+1`$ digits (represented as bits, but each digit only taking on the values 0 or 1) transforms, $$|x_1,\mathrm{},x_n,0|x_1,\mathrm{},x_n,(\underset{i}{}x_i)modq.$$ Now send the bits of the last block ($`_ix_imodq`$) to a Toffoli gate with all inputs negated and control bit $`b`$. The resulting output is exactly $`b\mathrm{Mod}_q(x_1,\mathrm{},x_n)`$. The bits in the last block can be erased by re-negating them and reversing the $`M_q`$ gate. This leaves only $`x_1,\mathrm{},x_n`$, $`O(n)`$ auxiliary bits, and the output $`b\mathrm{Mod}_q(x_1,\mathrm{},x_n)`$. The converse (simulating $`M_q`$ given $`\mathrm{MOD}_q`$) requires some more work. The first step is to show that $`\mathrm{MOD}_q`$ can also determine if a sum of digits is divisible by $`q`$. Let $`x_1,\mathrm{},x_nD`$ be a set of digits represented as $`\mathrm{log}q`$ bits each. For each $`i`$, let $`x_i^{(k)}`$ ($`0k\mathrm{log}q1`$) denote the bits of $`x_i`$. Since the numerical value of $`x_i`$ is $`_{k=0}^{\mathrm{log}q1}x_i^{(k)}2^k`$, it follows that $`{\displaystyle \underset{i=1}{\overset{n}{}}}x_i={\displaystyle \underset{k=0}{\overset{\mathrm{log}q1}{}}}{\displaystyle \underset{i=1}{\overset{n}{}}}x_i^{(k)}2^k.`$ The idea is to express this last sum in terms of a set of Boolean inputs that are fed into a $`\mathrm{MOD}_q`$ gate. To account for the factors $`2^k`$, each $`x_i^{(k)}`$ is fanned out $`2^k`$ times before plugging it into the $`\mathrm{MOD}_q`$ gate. Since $`k<\mathrm{log}q`$, this requires only constant depth and $`O(n)`$ auxiliary bits (which of course are set back to 0 in the end by reversing the fanout). Thus, just using $`\mathrm{MOD}_q`$ and constant fanout, we can determine if $`_{i=1}^nx_i0(modq)`$. More generally, we can determine if $`_{i=1}^nx_ir(modq)`$ using just a $`\mathrm{MOD}_{q,r}`$ gate and constant fanout. Let $`\widehat{\mathrm{MOD}}_{q,r}(x_1,\mathrm{},x_n)`$ denote the resulting circuit, that determines if a sum of digits is congruent to $`r`$ mod $`q`$. The construction of $`\widehat{\mathrm{MOD}}_{q,r}(x_1,\mathrm{},x_n)`$ is illustrated in the figure below for the case of $`q=3`$. In the figure, $`\mathrm{mod}(x)`$ denotes $`\mathrm{Mod}_{3,r}(x_1,\mathrm{},x_n)`$. The notation on the right will be used as a shorthand for this circuit: We can get the bits in the value of the sum $`_{i=1}^nx_imodq`$ using $`\widehat{\mathrm{MOD}}_{q,r}`$ circuits. This is done, essentially, by implementing the relation $`xmodq=_{r=0}^{q1}r\mathrm{Mod}_{q,r}(x)`$. For each $`r`$, $`0rq1`$, we compute $`\mathrm{Mod}_{q,r}(x_1,\mathrm{},x_n)`$ (where now the $`x_i`$’s are digits). This can be done by applying the $`\widehat{\mathrm{MOD}}_{q,r}`$ circuits in series (for each $`r`$) to the same inputs, introducing an auxiliary 0-bit for each application, as illustrated here. Let $`r_k`$ denote the $`k^{th}`$ bit of $`r`$. For each $`r`$ and for each $`k`$, we take the AND of the output of the $`\widehat{\mathrm{MOD}}_{q,r}`$ with $`r_k`$ (again by applying the AND’s in series, which is still constant depth, but introduces $`q`$ extra auxiliary inputs). Let $`a_{k,r}`$ denote the output of one of these AND’s. For each $`k`$, we OR together all the $`a_{k,r}`$’s, that is, compute $`_{r=0}^{q1}a_{k,r}`$, again introducing a constant number of auxiliary bits. Since only one of the $`r`$’s will give a non-zero output from $`\widehat{\mathrm{MOD}}_{q,r}`$, this collection of OR gates outputs exactly the bits in the value of $`_{i=1}^nx_imodq`$. Call the resulting circuit $`C`$, and the sum it outputs $`S`$. Finally, to simulate $`M_q`$, we need to include the input digit $`bD`$. To do this, we apply a unitary transformation $`T`$ to $`|S,b`$ that transforms it to $`|S,(b+S)modq`$. By Barenco, et al. (as in the proof of Lemma 3.2), $`T`$ can be computed in fixed depth using one-qubit gates and controlled NOT gates. Now using $`S`$ and all the other auxiliary inputs, we reverse the computation of the circuit $`C`$, thus clearing the auxiliary inputs. This is illustrated in this figure: The result is an output consisting of $`x_1,\mathrm{},x_n`$, $`O(n)`$ auxiliary bits, and $`(b+_{i=1}^nx_i)modq`$, which is the output of an $`M_q`$ gate. . It is clear that we can fan out digits, and therefore bits, using an $`F_q`$ gate (setting $`x_i=0`$ for $`1in`$ fans out $`n`$ copies of $`b`$). It is slightly less obvious (but still straightforward) that, given an $`F_q`$ gate, we can fully simulate an $`F`$ gate. ###### Lemma 3.4 For any $`q>2`$, $`F`$ and $`F_q`$ are QAC<sup>(0)</sup>-equivalent. Proof. By the preceeding lemmas, $`F_q`$ and $`\mathrm{MOD}_q`$ are QAC<sup>(0)</sup>-equivalent. By Moore’s result, $`\mathrm{MOD}_q`$ is QAC<sup>(0)</sup>-reducible to $`F`$. Hence $`F_q`$ is QAC<sup>(0)</sup>-reducible to $`F`$. Conversely, arrange each block of $`\mathrm{log}q`$ input bits to an $`F_q`$ gate as follows. For the control-bit block (which contains the bit we want to fan out), set all but the last bit to zero, and call the last bit $`b`$. Set all bits in the $`i^{th}`$ input-bit block to 0. Now the $`i^{th}`$ output of the $`F_q`$ circuit is $`b`$, represented as $`\mathrm{log}q`$ bits with only one possibly nonzero bit. Send this last output bit $`b`$ and the input bit $`x_i`$ to a controlled-NOT gate. The outputs of that gate are $`b`$ and $`bx_i`$. Now apply $`F_q^1`$ to the bits that were the outputs of the $`F_q`$ gate (which are all left unchanged by the controlled-not’s). This returns all the $`b`$’s to 0 except for the control bit which is always unchanged. The outputs of the controlled-not’s give the desired $`bx_i`$. Thus the resulting circuit simulates $`F`$, with $`O(n)`$ auxiliary bits. . ###### Theorem 3.5 For any $`q𝐍`$, $`q1`$, QACC $`=`$ QACC\[$`q`$\]. Proof. By the preceeding lemmas, fanout of bits is equivalent to the $`\mathrm{MOD}_q`$ function. By Moore’s result, we can do $`\mathrm{MOD}_q`$ if we can do fanout in constant depth. By our result, we can do fanout, and hence $`\mathrm{MOD}_2`$, if we can do $`\mathrm{MOD}_q`$. Hence QACC $`=`$ QACC\[$`2`$\] $``$ QACC\[$`q`$\]. . ## 4 Upper Bounds In this section, we prove the following upper bounds results NQACC$`{}_{}{}^{\mathrm{log}}{}_{gates}{}^{}`$$``$ TC<sup>(0)</sup>, BQACC$`{}_{}{}^{\mathrm{log}}{}_{𝐐,gates}{}^{}`$$``$ TC<sup>(0)</sup>, NQACC$`{}_{}{}^{\mathrm{log}}{}_{pl}{}^{}`$$``$ P/poly, and BQACC$`{}_{}{}^{\mathrm{log}}{}_{𝐐,pl}{}^{}`$$``$ P/poly. Suppose $`\{F_n\}`$ and $`\{z_n\}`$ determine a language $`L`$ in NQACC. Let $`F_n`$ be the product of the layers $`U_1,\mathrm{},U_t`$ and $`E`$ be the distinct entries of the matrices used in the $`U_j`$’s. By our definition of QACC, the size of $`E`$ is fixed with respect to $`n`$. We need a canonical way to write sums and products of elements in $`E`$ to be able to check $`|\stackrel{}{z}|U_1\mathrm{}U_t|x,0^{p(n)}|^2>0`$ with a TC<sup>(0)</sup> function. To do this let $`A=\{\alpha _i\}_{1im}`$ be a maximal algebraically independent subset of $`E`$. Let $`F=𝐐(A)`$ and let $`B=\{\beta _i\}_{0i<d}`$ be a basis for the field $`G`$ generated by the elements in $`(EA)\{1\}`$ over $`F`$. Since the size of the bases of $`F`$ and $`G`$ are less than the cardinality of $`E`$ the size of these bases is also fixed with respect to $`n`$. As any sum or product of elements in $`E`$ is in $`G`$, it suffices to come up with a canonical form for elements in $`G`$. Our representation is based on Yamakami and Yao . Let $`\alpha G`$. Since $`B`$ is a basis, $`\alpha =_{j=0}^{d1}\lambda _j\beta _j`$ for some $`\lambda _jF`$. We encode an $`\alpha `$ as a $`d`$-tuple (we iterate the pairing function from the preliminaries to make $`d`$-tuples) $`^{}\lambda _0^{},\mathrm{},^{}\lambda _{d1}^{}`$ where $`{}_{}{}^{}\lambda _{j}^{}`$ encodes $`\lambda _j`$. As the elements of $`A`$ are algebraically independent, each $`\lambda _j=s_j/u_j`$ where $`s_j`$ and $`u_j`$ are of the form $$\underset{\stackrel{}{k}_j,|\stackrel{}{k}_j|e}{}a_{\stackrel{}{k}_j}(\underset{i=1}{\overset{m}{}}\alpha _i^{k_{ij}}).$$ Here $`\stackrel{}{k}_j=(k_{1j},\mathrm{},k_{mj})𝐙^m`$, $`|\stackrel{}{k}_j|`$ is $`_ik_{ij}`$, $`a_{\stackrel{}{k}_j}𝐙`$, and $`e𝐍`$. In particular, any product $`\beta _m\beta _l=_{j=0}^{d1}\lambda _j\beta _j`$ with $`\lambda _j=s_j/u_j`$ and $`s_j`$ and $`u_j`$ in this form. We take a common denominator $`u`$ for elements of $`E\{\beta _m\beta _l\}`$ and not just $`E`$ since the $`\lambda _j`$’s associated with the $`\beta _m\beta _l`$ might have additional factors in their denominators not in $`E`$. Also fix an $`e`$ large enough to bound the $`|\stackrel{}{k_j}|`$’s which might appear in any element of $`E`$ or a product $`\beta _m\beta _l`$. This $`e`$ will be constant with respect to $`n`$. In multiplying $`t`$ layers of QACC circuit against an input, the entries in the result will be polynomial sums and products of elements in $`E\{\beta _m\beta _l\}`$, so we can bound $`|\stackrel{}{k}_j|`$ for $`\stackrel{}{k_j}`$’s which appear in the $`\lambda _j`$’s of such an entry by $`ep(n)`$. To complete our representation of $`\alpha G`$ we encode $`\lambda _j`$ as the sequence $`r,a_{\stackrel{}{k_j}},k_{1j},\mathrm{},k_{mj}`$ where $`r`$ is the power to which $`u`$ is raised and $`a_{\stackrel{}{k_j}},k_{1j},\mathrm{},k_{mj}`$ is the sequence of $`a_{\stackrel{}{k_j}},k_{1j},\mathrm{},k_{mj}`$’s that appear in $`s_j`$. By our discussion, the encoding of an $`\alpha `$ that appears as an entry in the output after applying a QACC operator to the input is of polynomial length and so can be manipulated in TC<sup>(0)</sup>. We have need of the following lemma: ###### Lemma 4.1 Let $`p`$ be a polynomial. (1) Let $`f(i,x)`$ TC<sup>(0)</sup> output encodings of $`a_{i,x}𝐙[A]`$. Then $`𝐙[A]`$ encodings of $`_{i=1}^{p(|x|)}a_{i,x}`$ and $`_{i=1}^{p(|x|)}a_{i,x}`$ are TC<sup>(0)</sup> computable. (2) Let $`f(i,x)`$ TC<sup>(0)</sup> output encodings of $`a_{i,x}G`$. Then $`G`$ encodings of $`_{i=1}^{p(|x|)}a_{i,x}`$ and $`_{i=1}^{p(|x|)}a_{i,x}`$ are TC<sup>(0)</sup> computable. Proof. We will abuse notation in this proof and identify the encoding $`f(i,x)`$ with its value $`a_{i,x}`$. So $`_if(i,x)`$ and $`_if(i,x)`$ will mean the encoding of $`_ia_{i,x}`$ and $`_ia_{i,x}`$ respectively. (1) To do sums, the first thing we do is form the list $`L1=f(0,x),\mathrm{},f(p(|x|),x)`$. Then we create a flattened list $`L2`$ from this with elements which are the $`a_{\stackrel{}{k_j}},k_{1j},\mathrm{},k_{mj}`$’s from the $`f(i,x)`$’s. $`L1`$ is in TC<sup>(0)</sup> using our definition of sequence from the preliminaries, and closure under sums and $`max_i`$ to find the length of the longest $`f(i,x)`$. To flatten $`L1`$ we use $`max_i`$ to find the length $`d`$ of the longest $`f(i,x)`$ for $`ip(|x|)`$. Then using max twice we can find the length of the longest $`a_{\stackrel{}{k_j}},k_{1j},\mathrm{},k_{mj}`$. This will be the second coordinate in the pair used to define sequence $`L2`$. We then do a sum of size $`dp(|x|)`$ over the subentries of $`L1`$ to get the first coordinate of the pair used to define $`L2`$. Given $`L2`$, we make a list $`L3`$ of the distinct $`\stackrel{}{k_j}`$’s that appear as $`a_{\stackrel{}{k_j}},k_{1j},\mathrm{},k_{mj}`$ in some $`f(i,x)`$ for some $`ip(|x|)`$. This list can be made from $`L2`$ using sums, $`cond`$ and $`\mu `$. We sum over the $`tlength(L2)`$ and check if there is some $`t^{}<t`$ such that the $`t^{}`$th element of $`L2`$ has same $`\stackrel{}{k}_j`$ as $`t`$ and if not add the $`t`$th elements $`\stackrel{}{k_j}`$ times 2 raised to the appropriate power. We know what power by computing the sum of the number of smaller $`t^{}`$ that passed this test. Using $`cond`$ and closure under sums we can compute in TC<sup>(0)</sup> a function which takes a list like $`L2`$ and a $`\stackrel{}{k_j}`$ and returns the sum of all the $`a_{\stackrel{}{k_j}}`$’s in this list. So using this function and the lists $`L2`$ and $`L3`$ we can compute the desired encoding. For products, since the $`\alpha _i`$’s of $`A`$ are algebraically independent, $`𝐙[A]`$ is isomorphic to the polynomial ring $`𝐙[y_1,\mathrm{},y_m]`$ under the natural map which takes $`\alpha _j`$ to $`y_j`$. We view our encodings $`f(i,x)`$ as $`m`$-variate polynomials in $`𝐙[y_1,\mathrm{},y_m]`$. We describe for any $`p^{}`$ a circuit that works for any TC<sup>(0)</sup> computable $`f(i,x)`$ such that $`_if(i,x)`$ is of degree less than $`p^{}`$ viewed as an $`m`$-variate polynomial. In $`TC^{(0)}`$ we define $`g(i,x)`$ to consist of the sequence of polynomially many integer values which result from evaluating the polynomial encoded by $`f(i,x)`$ at the points $`(i_1,\mathrm{},i_m)𝐍^m`$ where $`0i_s`$ and $`_si_sp^{}`$. To compute $`f(i,x)`$ at a point involves computing a polynomial sum of a polynomial product of integers, and so will be in $`TC^{(0)}`$. Using closure under polynomial integer products we compute $`k(j,x):=_i\beta (j,g(i,x))`$ where $`\beta `$ is the sequence projection function from the preliminaries. Our choice of points is what is called by Chung and Yao the $`p^{}`$-th order principal lattice of the $`m`$-simplex given by the origin and the points $`p^{}`$ from the origin in each coordinate axis. By Theorems 1 and 4 of that paper (proved earlier by a harder argument in Nicolaides ) the multivariate Lagrange Interpolant of degree $`p^{}`$ through the points $`k(j,x)`$ is unique. This interpolant is of the form $`P(y_1,\mathrm{},y_m)=_jp_j(y_1,\mathrm{},y_m)k(j,x)`$ where the $`p_j`$’s are polynomials which do not depend on the function $`f`$. An explicit formula for these $`p_j`$’s is given in Corollary 2 of Chung and Yao as a polynomial product of linear factors. Since these polynomials are all of degree less than $`p^{}`$, they have only polynomial in $`p^{}`$ many coefficients and in PTIME these coefficients can be computed by iteratively multiplying the linear factors together. We can then hard code these $`p_j`$’s (since they don’t depend on $`f`$) into our circuit and with these $`p_j`$’s, $`k(j,x)`$, and closure under sums we can compute the polynomial of the desired product in TC<sup>(0)</sup>. (2) We do sums first. Assume $`f(i,x):=_{j=0}^{d1}\lambda _{ij}\beta _j`$. One immediate problem is that the $`\lambda _{ij}`$ and $`\lambda _{i^{}j}`$ might use different $`u^r`$’s for their denominators. Since TC<sup>(0)</sup> is closed under poly-sized maximum, it can find the maximum value $`r_0`$ to which $`u`$ is raised. Then it can define a function $`g(i,x)=_{j=0}^{d1}\gamma _{ij}\beta _j`$ which encodes the same element of $`G`$ as $`f(i,x)`$ but where the denominators of the $`\gamma _{ij}`$’s are now $`u^{r_0}`$. If $`\lambda _j`$ was $`s_j/u^r`$ we need to compute the encoding $`s_ju^{r_0r}/u^{r_0}`$. This is straightforward from (1). Now $$\underset{i=1}{\overset{p(|x|)}{}}f(i,x)=\underset{i=1}{\overset{p(|x|)}{}}g(i,x)=\underset{j=0}{\overset{d1}{}}[(\underset{i=1}{\overset{p(|x|)}{}}s_{ij})/u^{r_0}]\beta _j,$$ where $`s_{ij}`$’s are the numerators of the $`\gamma _{ij}`$’s in $`g(i,x)`$. From part (1) we can compute the encoding $`e_j`$ of $`(_{i=1}^{p(|x|)}s_{ij})`$ in TC<sup>(0)</sup>. So the desired answer $`r_0,e_0,\mathrm{},r_0,e_{d1}`$ is in TC<sup>(0)</sup>. For products $`_{i=1}^{p(|x|)}f(i,x)`$, we play the same trick as the in the $`𝐙[A]`$ product case. We view our encodings of elements of $`G`$ as d-variate polynomials in $`F(y_0,\mathrm{},y_{d1})`$ under the map $`\beta _k`$ goes to $`y_k`$. (Note that this map is not necessarily an isomorphism.) We then create a function $`g(i,x)`$ which consists of the sequence of values obtained by evaluating $`f(i,x)`$ at polynomially many points in a lattice as in the first part of this lemma. Evaluating $`f(i,x)`$ at a point can easily be done using the first part of this lemma. We then use part (1) of this lemma to compute the products $`k(j,x)=\beta (j,g(i,x))`$. We then get the interpolant $`P(y_0,\mathrm{},y_{d1})=_jp_j(y_0,\mathrm{},y_m)k(j,x)`$. We non-uniformly obtain the encoding of $`p_j(\beta _0,\mathrm{},\beta _{d1})`$ expressed as an element of $`G`$. i.e., in the form $`_{w=0}^{d1}\lambda _{jw}\beta _w`$. Thus, the product $`_{i=1}^{p(|x|)}f(i,x)`$ is $$\underset{w=0}{\overset{d1}{}}(\underset{j}{}\lambda _{jw}k(j,w))\beta _w$$ The encoding of the products is the d-tuple given by $`_j\lambda _{j0}k(j,0),\mathrm{},_j\lambda _{jd1}k(j,d1)`$ Each of its components is a polynomial sum of a product of two things in $`F`$ and can be computed using the first part of the lemma. . For $`\{F_n\}`$ QAC$`{}_{wf}{}^{(0)}=`$ QACC, the vectors that $`F_n`$ act on are elements of a $`2^{n+p(n)}`$ dimensional space $``$<sub>1,n+p(n)</sub> space which is a tensor product of the 2-dimensional spaces $``$$`{}_{1}{}^{},\mathrm{}`$<sub>n+p(n)</sub>, which in turn are each spanned by $`|0,|1`$. We write $``$<sub>j,k</sub> for the subspace $`_{i=j}^k`$<sub>i</sub> of $``$<sub>1,n+p(n)</sub>. We now define a succinct way to represent a set of vectors in $``$<sub>1,n+p(n)</sub> which is useful in our argument below. A tensor graph is a directed acyclic graph with one source node of indegree zero, one terminal node of outdegree zero, and two kinds of edges: horizontal edges, which are unlabeled, and vertical edges, which are labeled with a pair of amplitudes and a product of colors and anticolors. (The color product may be the number 1.) We require that all paths from the source to the terminal traverse the same number of vertical edges and that no vertex can have vertical edge indegree greater than one or outdegree greater than one. For a color $`c`$ we write $`\stackrel{~}{c}`$ for its corresponding anticolor. The height of a node in a tensor graph is the number of vertical edges traversed to get to it on any path from the source; the height of an edge is the height of its end node. The width of a tensor graph is maximum number of nodes of the same height. As an example of a tensor graph where our color product is the number 1, consider the following figure: The rough idea of tensor graphs is that paths through the graph correspond to collections of vector in $`_{1,n}`$. For this particular figure the left path from the source node (s) to the terminal node (t) corresponds to the vectors given by $$|1(\frac{1}{\sqrt{2}}|0+\frac{1}{\sqrt{2}}|1)\frac{1}{2}|0$$ and the right hand path corresponds to $$|0(\frac{1}{\sqrt{2}}|0+\frac{1}{\sqrt{2}}|1)\frac{1}{2}|0.$$ A $``$<sub>j,k</sub>-term in a tensor graph is a maximal induced tensor subgraph between a node of height $`j1`$ and a node of height $`k`$. We also require that the horizontal indegree of the node at height $`j1`$ be zero and that the horizontal outdegree of the node at height $`k`$ be zero. For the graph we considered above there are two $``$<sub>1,2</sub>-terms and two $``$<sub>2,3</sub>-terms but only one $``$<sub>1,3</sub>-term corresponding to the whole figure. Colors are used to handle controlled-not layers. A color $`c`$ and its anticolor $`\stackrel{~}{c}`$ satisfy the following multiplicative properties: $`cc=\stackrel{~}{c}\stackrel{~}{c}=1`$ and $`c\stackrel{~}{c}=0`$. Given two distinct colors $`b`$ and $`c`$ we have $`bc=cb`$ and $`\stackrel{~}{b}c=c\stackrel{~}{b}`$. If $`a`$ is a product of colors and anticolors not involving the color $`b`$ or $`\stackrel{~}{b}`$ and $`c`$ is another product of colors we have $`a(bc)=(ab)c`$. We consider formal sums of products of complex numbers times colors. We require complex numbers to commute with colors and require colors and anticolors to distribute, i.e., if $`a`$, $`b`$, $`c`$ are colors or anticolors then $`a(b+c)=ab+ac`$ and $`(b+c)a=ba+ca`$. Finally, we require addition to work so that the above structure satisfies the axioms of an $`𝐂`$-algebra. Given a tensor graph $`G`$ denote by $`𝒜_G`$ the $`𝐂`$-algebra above. Since $$(aa)\stackrel{~}{a}=\stackrel{~}{a}0=a(a\stackrel{~}{a})$$ this algebra is not associative. However, in the sums we will consider the terms will never have more than two positions where a color or its anticolor can occur, so the products we will consider are associative. Using our our earlier encoding for the elements of $`𝐂`$ which could appear in a $`QACC`$ computation, it is straightforward to use sequence coding to get a TC<sup>(0)</sup> encodings of the relevant elements of $`𝒜_G`$. As an example of how colors affect amplitudes, consider the following picture: The amplitude of $`|1,0,0`$ in the left hand dotted path is $`b\frac{1}{\sqrt{2}}1\frac{1}{\sqrt{2}}b1=1/2`$ using commutativity and $`b^2=1`$. Its amplitude in the right hand dotted path would be zero because of the last vertical edge. However, vectors such as $`|0,0,1`$ would have nonzero amplitude in the right hand dotted path. Nevertheless, the amplitude of any vector $`|\stackrel{}{x}`$ in any path other than the dotted ones from $`s`$ to $`t`$ will be $`0`$ as $`b\stackrel{~}{b}=0`$. More formally, we define the amplitude of an $`|\stackrel{}{x}`$ in a vertical edge as equal to the left amplitude times the color product in the edge if $`\stackrel{}{x}`$ is $`|0`$ and equal to the right amplitude times the color product in the edge if $`\stackrel{}{x}`$ is $`\stackrel{}{1}`$. The amplitude of a vector $`|x_1,\mathrm{},x_j`$ in a path in a tensor graph is the product over $`k`$ from 1 to $`j`$ of the amplitude of the vectors $`|x_k`$ in the vertical edge of height $`k`$. The amplitude of a vector $`|x_j,\mathrm{},x_k`$ in an $`_{j,k}`$-term is the sum of its amplitude in its paths. The amplitude of a vector $`|x_1,\mathrm{},x_{p(n)}`$ in a tensor graph $`G`$ is defined to be the sum of its amplitudes in $`G`$’s $`_{1,p(n)}`$-terms. As we will be interested in families of tensor graphs $`\{G_n\}`$, corresponding to our circuit families we want to look at those families with a certain degree of uniformity. We say a family of tensor graphs $`\{G_n\}`$ is color consistent if: (1) the number of colors for edges of the same height is bounded by a constant $`k`$ with respect to $`n`$, (2) the number of heights in which a given color/anticolor can appear is exactly two (colors and their anticolors must appear on the same heights), (3) each color product at the same height is of the form $`_{i=0}^kl_i`$ where $`l_i`$ must be either a color $`c_i`$ or $`\stackrel{~}{c_i}`$ (it follows there are $`2^k`$ possible color products for edges at a given height). We say that a color/anticolor is active at a given height if the height is at or after the first height at which the color/anticolor occurs and is below the height of its second occurrence. The family is further said to be log-color depth if the number of active colors/anticolors of a given height is log-bounded. ###### Theorem 4.1 Let $`\{F_n\}`$ be a family of QACC operators and let $`\{\stackrel{}{z}_n|\}`$ a family of observables. (1) There is a color-consistent family of tensor graphs of width $`2^{2^{2t}}`$ and polynomial size representing the output amplitudes of $`U_1\mathrm{}U_t|\stackrel{}{z}_n`$ where $`U_i`$ are the layers of $`F_n`$. (2) If $`\{F_n\}`$ is in QACC$`{}_{}{}^{\mathrm{log}}{}_{pl}{}^{}`$then the family of tensor graphs will be of log-color depth. (3) If $`\{F_n\}`$ is in QACC$`{}_{}{}^{\mathrm{log}}{}_{gates}{}^{}`$then the number of paths from the source to the terminal node is polynomially bounded. Proof. The proof is by induction on $`t`$. In the base case, $`t=0`$, we do not multiply any layers, and we can easily represent this as a tensor graph of width 1. Assume for $`j<t`$ that $`U_j\mathrm{}U_1|\stackrel{}{x},0^{p(n)}`$ can be written as color consistent tensor graph of width $`2^{2^{2t}}`$ and polynomial size. There are two cases to consider: In the first case the layer is a tensor product of matrices $`M_1\mathrm{}M_\nu `$ where the $`M_k`$’s are Toffoli gates, one qubit gates, or fan-out gates (since QAC$`{}_{wf}{}^{(0)}=`$QACC); in the second case the layer is a controlled-not layer. For the first case we “multiply” $`U_t`$ against our current graph by “multiplying” each $`M_j`$ in parallel against the terms in our sum corresponding to $`M_j`$’s domain, say $``$$`_{j^{},k^{}}`$. If $`M_j=\left(\begin{array}{cc}u_{00}& u_{01}\\ u_{10}& u_{11}\end{array}\right)`$ with domain $``$$`_j^{}`$ is a one-qubit gate, then we multiply the two amplitudes in each vertical edge of height $`j^{}`$ in our tensor graph by $`M_j`$. This does not effect the width, size, or number of paths through the graph. If $`M_j`$ is a Toffoli gate, then for each term $`S`$ in $``$$`_{j^{},k^{}}`$ in our tensor graph we add one new term to the resulting graph. This term is added by adding a horizontal edge going out from the source node of $`S`$ followed by the new $``$$`_{j^{},k^{}}`$-term followed by a horizontal edge into the terminal node of $`S`$. The new term is obtained from the old one by setting to $`0`$ the left hand amplitudes of all edges in $`S`$ of height between $`j^{}`$ and $`k^{}1`$ and then if $`\alpha ,\gamma `$ is the amplitude of an edge of height $`k^{}`$ in the new term we change it to $`\gamma \alpha ,\alpha \gamma `$. This new term adjusts the amplitude for the case of a $`|1^{(k^{}j^{}1)}`$ vector in $``$$`_{j^{},k^{}1}`$ tensored with either a $`|0`$ or $`|1`$. This operation increases the width of the new tensor graph by the width of the $`_{j^{},k^{}}`$-term for each $``$$`_{j^{},k^{}}`$-term in the graph. Since the original graph has width $`2^{2^{2(t1)}}`$ there are at most this many starting and ending vertices for such terms. So there at most $`(2^{2^{2(t1)}})^2`$ such terms. Each of these terms has width at most $`2^{2^{2(t1)}}`$. Thus, the new width is at most $$2^{2^{2(t1)}}+(2^{2^{2(t1)}})^22^{2^{2(t1)}}<2^{2^{2t}}.$$ Notice this action adds one new path through the $`_{j^{},k^{}}`$ part of the graph for every existing one. Now suppose $`M_j`$ is a fan-out gate, let $`S`$ be a $``$$`_{j^{},k^{}}`$-term in our tensor graph and let $`e`$ be any vertical edge in $`S`$ in $``$$`_k^{}`$. Suppose $`e`$ has amplitude $`\alpha `$ for $`|0`$ and amplitude $`\gamma `$ for $`|1`$. In the new graph we change the amplitude of $`e`$ to $`\alpha ,0`$. We then add a horizontal edge out of the source node of $`S`$ followed by a new $``$$`_{j^{},k^{}}`$-term followed by a horizontal edge into the terminal node of $`S`$. The new term is obtained from $`S`$ by changing the amplitude for edges in $`_k^{}`$ with amplitudes $`\alpha ,\gamma `$ in $`S`$ to $`0,\gamma `$. The amplitudes of the non-$`_k^{}`$ edges in this term are the reverse of the corresponding edge in $`S`$, i.e., if the edge in $`S`$ had amplitude $`\delta ,\zeta `$ then the new term edge would have amplitude $`\zeta ,\delta `$. The same argument as in the Toffoli case shows the new width is bounded by $`2^{2^{2t}}`$ and that this action adds one new path through the $`_{j^{},k^{}}`$ part of the graph for every existing one. For the case of a controlled-not layer, suppose we have a controlled-not going from line $`i`$ onto line $`j`$. Let $`c,\overline{c}`$ be a new color, anti-color pair not yet appearing in the graph. Let $`e_i`$ be a vertical edge of height $`i`$ in the graph and let $`C_i,\alpha _i,\gamma _i`$ be respectively its color product and two amplitudes. Similarly, let $`e_j`$ be a vertical edge of height $`j`$ in the graph and $`C_j,\alpha _j,\gamma _j`$ be its color product and two amplitudes. In the new graph we multiply $`c`$ times the color product of $`e_i`$ and $`e_j`$ and change the amplitude of $`e_i`$ to $`\alpha _i,0`$. We then add a horizontal edge going out from the starting node of $`e_i`$, followed by a vertical edge with values $`C_i\stackrel{~}{c},0,\gamma _i`$ followed by a horizontal edge into the terminal node of $`e_i`$. In turn, we add a horizontal edge going out of the starting node of $`e_j`$, followed by a vertical edge with values $`C_j\stackrel{~}{c},\gamma _i,\alpha _j`$ followed by a horizontal edge into the terminal node of $`e_j`$. We handle all other controlled gates in this layer in a similar fashion (recall they must go to disjoint lines). We add at most a new vertex of a given height for every existing vertex of a given height. So the total width is at most doubled by this operation and $`22^{2^{2(t1)}}<2^{2^{2t}}`$. In the QACC$`{}_{}{}^{\mathrm{log}}{}_{pl}{}^{}`$case, simulating a layer which is a Kronecker product of spaced controlled-not gates and identity matrices, notice we would at most add one to the color depth at any place. So if a controlled-not layer is a composition of $`O(\mathrm{log})`$ many such layers it will increase the color depth by $`O(\mathrm{log})`$. In the QACC$`{}_{}{}^{\mathrm{log}}{}_{gates}{}^{}`$ case, notice that simulating a single controlled-not we add one new path for each existing path through the graph at each of the two heights affected. This gives three new paths on the whole subspace for each old one. Since we have handled the two possible layer cases and the changes we needed to make only increase the resulting tensor graph polynomially, we thus have established the induction step and (1) and (2) of the theorem. For (3), observe for each multi-line gate we handle in adding a layer we at most quadruple the number of paths through the subspace where that gate applies. Since there are at most logarithmically many such gates, the number of paths through the graph increases polynomially. . ###### Theorem 4.2 Let $`\{G_n\}`$ be a family of constant width color-consistent tensor graphs of vectors in $``$<sub>1,p(n)</sub>. Assume the coefficients of amplitudes in the $`\{G_n\}`$ can be encoded in $`TC^{(0)}`$ using our encoding scheme described earlier and that $`\{G_n\}`$ has log-color depth. Then the amplitude of any basis vector of $``$<sub>1,p(n)</sub> in $`G_n`$ is P/poly computable. If the number of paths through the graph from the source to the terminal node is polynomially bounded then the amplitude of any basis vector is TC<sup>(0)</sup> computable. Proof. Let $`G_n`$ be a particular graph in the family and let $`|\stackrel{}{x_n}`$ be the vector whose amplitude we want to compute. Assume that all graphs in our family have fewer than $`k`$ colors in any color product and have a width bounded by $`w`$. We will proceed from the source to the terminal node one height at a time to compute the amplitude. Since the width is $`w`$ the number of $`_1`$-terms is at most $`w`$ and each of these must have width at most $`w`$. Let $`\alpha _{1,1},\mathrm{},\alpha _{1,w}`$ (some of which may be zero) denote the amplitudes in $`𝒜_{G_n}`$ of $`|x_{n,1}`$ in each of these terms. The $`\alpha _{1,i}`$ are each sums of at most $`w`$ amplitudes times the color products of at most $`k`$ colors and anticolors, so the encoding of these $`w`$ amplitudes is TC<sup>(0)</sup> computable. Because of the restriction on the width of $`G_n`$ there are at most $`w`$ many $``$<sub>1,j</sub>-terms, $`w^2`$ many $``$<sub>j,j+1</sub>-terms, and $`w`$ many $``$<sub>1,j+1</sub>-terms. Fixing some ordering on the nodes of height $`j`$ and $`j+1`$ let $`\gamma _{j,i,k}`$ be the amplitude of $`|x_{n,j+1}`$ in the $``$<sub>j,j+1</sub>-term with source the $`i`$th node of height $`j`$ and with terminal node the $`k`$th node of height $`j+1`$. The amplitude is zero if there is no such $`_{j,j+1}`$-term. Then the amplitudes $`\alpha _{j+1,1},\mathrm{},\alpha _{j+1,w}`$ of the $``$<sub>1,j+1</sub>-terms can be computed from the amplitudes $`\alpha _{j,1},\mathrm{},\alpha _{j,w}`$ of the $``$<sub>1,j</sub>-terms using the formula $$\alpha _{j+1,k}=\underset{i=1}{\overset{w}{}}\alpha _{j,i}\gamma _{j,i,k}.$$ Thus $`\alpha _{j+1,k}`$ can be computed from the $`\alpha _{j,i}`$ using a polynomial sized circuit to do these adds and multiplies. Similarly, each $`\alpha _{j,k}`$ can be computed by polynomial sized circuits from the $`\alpha _{j1,k}`$’s and so on. Since we have log-color depth the number of terms consisting of elements in our field times color products in a $`\alpha _{j,k}`$ will be polynomial. So the size of the $`\alpha _{j,k}`$’s $`jp(n)`$, $`kw`$ will be polynomial in the input $`\stackrel{}{x}_n`$. So the size of the circuits for each $`\alpha _{j,k}`$ where $`jp(n)`$ and $`kw`$ will be polynomial size. There is only one $``$<sub>1,p(n)</sub>-term in $`G_n`$ and its amplitude is that of $`|\stackrel{}{x}_n`$, so this shows it has polynomial sized circuits. For the TC<sup>(0)</sup> result, if the number of paths is polynomially bounded, then the amplitude can be written as the polynomial sum of the amplitudes in each path. The amplitude in a path can in turn be calculated as a polynomial product of the amplitudes times the colors on the vertical edges in the path. Our condition on every color appearing at exactly two heights guarantees the color product along the whole path will be 1 or 0, and will be zero iff we get a color and its anticolor on the path. This is straightforward to check in TC<sup>(0)</sup>, so this sum of products can thus be computed in TC<sup>(0)</sup> using Lemma 4.1. . ###### Corollary 4.3 EQACC$`{}_{}{}^{\mathrm{log}}{}_{pl}{}^{}`$$``$NQACC$`{}_{}{}^{\mathrm{log}}{}_{pl}{}^{}`$$``$P/Poly, and BQACC$`{}_{}{}^{\mathrm{log}}{}_{𝐐,pl}{}^{}`$ $``$P/poly. EQACC$`{}_{}{}^{\mathrm{log}}{}_{gates}{}^{}`$$``$NQACC$`{}_{}{}^{\mathrm{log}}{}_{gates}{}^{}`$$``$TC<sup>(0)</sup>, and BQACC$`{}_{}{}^{\mathrm{log}}{}_{𝐐,gates}{}^{}`$$``$TC<sup>(0)</sup>. Proof. Given a a family $`\{F_n\}`$ of QACC$`{}_{}{}^{\mathrm{log}}{}_{pl}{}^{}`$operators and a family $`\{\stackrel{}{z}_n|\}`$ of states we can use Theorem 4.1 to get a family $`\{G_n\}`$ of log color depth, color-consistent tensor graphs representing the amplitudes of $`F_n^1|\stackrel{}{z}_n`$. Note $`\{F_n^1\}`$ is also a family of QACC$`{}_{}{}^{\mathrm{log}}{}_{pl}{}^{}`$operators since Toffoli and fan-out gates are their own inverses, the inverse of any one qubit gate is also a one qubit gate (albeit usually a different one), and finally a controlled-not layer is its own inverse. Theorem 4.2 shows there is a P/poly circuit computing the amplitude of any vector $`|\stackrel{}{x}_n`$ in this graph. This amounts to calculating $$\stackrel{}{x}_n|F_n^1|\stackrel{}{z}_n=\stackrel{}{z}_n|F_n|\stackrel{}{x}_n$$ If this is nonzero, then $`|\stackrel{}{z}_n|F_n|\stackrel{}{x}_n|^2>0`$, and we know $`\stackrel{}{x}`$ is in the language. In the BQACC<sub>Q</sub> case everything is a rational so P/poly can explicitly compute the magnitude of the amplitude and check if it is greater than $`3/4`$. The TC<sup>(0)</sup> result follows similarly from the TC<sup>(0)</sup> part of Theorem 4.1. . ## 5 Discussion and Open Problems A number of questions are suggested by our work. * Is all of NQACC in TC<sup>(0)</sup> or even P/Poly? We conjecture that NQACC is in TC<sup>(0)</sup>. As mentioned in the introduction, we have developed techniques that remove some of the important obstacles to proving this. * Are there any natural problems in NQACC that are not known to be in ACC? * What exactly is the complexity of the languages in EQACC, NQACC and BQACC<sub>Q</sub>? We entertain two extreme possibilities. Recall that the class ACC can be computed by quasipolynomial size depth 3 threshold circuits . It would be quite remarkable if EQACC could also be simulated in that manner. However, it is far from clear if any of the techniques used in the simulations of ACC (the Valiant-Vazirani lemma, composition of low-degree polynomials, modulus amplification via the Toda polynomials, etc.), which seem to be inherently irreversible, can be applied in the quantum setting. At the other extreme, it would be equally remarkable if NQACC and NQTC<sup>(0)</sup> (or BQACC<sub>Q</sub> and NQTC<sup>(0)</sup>) coincide. Unfortunately, an optimal characterization of QACC language classes anywhere between those two extremes would probably require new (and probably difficult) proof techniques. * How hard are the fixed levels of QACC? While lower bounds for QACC itself seem impossible at present, it might be fruitful to study the limitations of small depth QACC circuits (depth 2, for example). Acknowledgments: We thank Cris Moore for pointing out an error in an earlier version of Theorem 4.1, and Bill Gasarch for helpful comments and suggestions.
warning/0002/cond-mat0002418.html
ar5iv
text
# 1 Introduction ## 1 Introduction In this note, we are going to discuss the algebraic structure of a certain infinite-dimensional Lie algebra which appeared in the seminal work of Onsager of 1944 on the solution of 2D Ising model . Due to some other simpler and powerful methods introduced later in the study of Ising model, the algebra in the Onsager’s original work, now called the Onsager algebra, had not received enough attention in the substantial years until the 1980-s when its new-found role appeared in the superintegrable chiral Potts model, \[2-12\] (for a review of the subject, see and references therein ). Since the work of Onsager it was known that there exists an intimate relationship between the Onsager algebra and $`sl_2`$, a connection now clarified in . Namely, the Onsager algebra is isomorphic to the fixed subalgebra of $`sl_2`$-loop algebra (or alternatively, its central extension $`A_1^{(1)}`$, ) by a certain involution. A generalization of the Onsager algebra to other Kac-Moody algebras was later introduced in , with some interesting relations with integrable motions. Due to the close relationship of the Onsager algebra with 2D integrable models in statistical mechanics, especially in the chiral Potts model, a thorough mathematical understanding of this infinite-dimensional algebra is desirable to warrant the further investigation. In this note, we summarize our recent results on the algebraic study of the structure of the Onsager algebra. Detailed derivations, as well as extended references to the literature, may be found in . We have established the structure theorem of a certain class of (Lie)-ideals of the Onsager algebra, originated from the study of (reducible or irreducible) finite dimensional representations of the Onsager algebra, through the theory of the reciprocal polynomials. In the process, we have found a profound structure of the Onsager algebra with some interesting applications to solvable and nilpotent Lie algebras. It suggests that the further study of the ideals in the Onsager algebra and its possible generalization should provide certain useful information on the theory of solvable Lie algebras. The mathematical results obtained would be expected to have some feedback to the original physical theory. Such a program is now under progress and partial results are promising. ## 2 Hamiltonian of Superintegrable Chiral Potts Model The superintegrable chiral Potts $`N`$-state spin chain Hamiltonian has the following form of a parameter $`k^{}`$, $$H(k^{})=H_0+k^{}H_1,$$ with $`H_0,H_1`$ the Hermitian operators acting on the vector space of $`L`$-tensor of $`\text{IC}^N`$, defined by $$H_0=2\underset{l=1}{\overset{L}{}}\underset{n=1}{\overset{N1}{}}(1\omega ^n)^1X_l^n,H_1=2\underset{l=1}{\overset{L}{}}\underset{n=1}{\overset{N1}{}}(1\omega ^n)^1Z_l^nZ_{l+1}^{Nn},$$ where $`\omega =e^{\frac{2\pi \mathrm{i}}{N}}`$, $`X_l=I\mathrm{}\stackrel{l\mathrm{th}}{X}\mathrm{}I`$, $`Z_l=I\mathrm{}\stackrel{l\mathrm{th}}{Z}\mathrm{}I,(Z_{L+1}=Z_1)`$. Here $`I`$ is the identity operator, and $`X,Z`$ are the operators of $`\text{IC}^N`$ with the Weyl commutation relation, $`ZX=\omega XZ`$, which are defined by $`X|m>=|m+1>,Z|m>=\omega ^m|m>`$, $`m\text{ZZ}_N`$. The operator $`H(k^{})`$ is Hermitian for real $`k^{}`$, hence with the real eigenvalues. When $`N=2`$, $`X,Z`$, become the Pauli matrices, $`\sigma ^1,\sigma ^3`$, then it is the Ising quantum chain : $$\begin{array}{c}H(k^{})=_{l=1}^L\sigma _l^1+k^{}_{l=1}^L\sigma _j^3\sigma _{l+1}^3.\hfill \end{array}$$ For $`N=3`$, one obtains the $`\text{ZZ}_3`$-symmetrical self-dual chiral clock model with the chiral angles $`\phi =\varphi =\frac{\pi }{2}`$, $$\frac{\sqrt{3}}{2}H(k^{})=\underset{l=1}{\overset{L}{}}(e^{\frac{\pi \mathrm{i}}{6}}X_l+e^{\frac{\pi \mathrm{i}}{6}}X_l^2)+k^{}\underset{l=1}{\overset{L}{}}(e^{\frac{\pi \mathrm{i}}{6}}Z_lZ_{l+1}^2+e^{\frac{\pi \mathrm{i}}{6}}Z_l^2Z_{l+1}),$$ which was studied by Howes, Kadanoff and M. den Nijs . Note that the Potts model is given by the chiral angles, $`\phi =\varphi =0`$. For a general $`N`$, $`H(k^{})`$ was constructed in a paper of G. von Gehlen and R. Rittenberg , in which the models were shown to be ”superintegrable” in the sense that the Dolan-Grady (DG) condition is satisfied for $`A_0=2N^1H_0,A_1=2N^1H_0`$, $$[A_1,[A_1,[A_1,A_0]]]=16[A_1,A_0],[A_0,[A_0,[A_0,A_1]]]=16[A_0,A_1].$$ For a pair of operators, $`A_0,A_1`$, we denote, $`4G_1=[A_1,A_0]`$, and define an infinite sequence of operators, $`A_m,G_m,(m\text{ZZ})`$, by relations, $$A_{m1}A_{m+1}=\frac{1}{2}[A_m,G_1],G_m=\frac{1}{4}[A_m,A_0].$$ The DG-condition on $`A_1,A_0,`$ are equivalent to the statement that the collection of $`A_m,G_m`$ forms the Onsager algebra, which means they satisfy the Onsager relations, $$\begin{array}{ccc}[A_m,A_l]=4G_{ml},\hfill & [A_m,G_l]=2(A_{ml}A_{m+l}),\hfill & [G_m,G_l]=0.\hfill \end{array}$$ (1) The above relations ensure that the $`k^{}`$-dependence eigenvalues of $`H(k^{})`$ have the following special form as in the Ising model, $$a+bk^{}+2N\underset{j=1}{\overset{n}{}}m_j\sqrt{1+k^22k^{}\mathrm{cos}(\theta _j)},$$ where $`a,b,\theta _j,`$ are reals, and $`m_j`$’s take the values, $`m_j=s_j,(s_j+1),\mathrm{},s_j`$ with $`s_j`$ a positive half-integer . Mathematically, this result follows from the classification of the finite-dimensional unitary representations of the Onsager algebra, by the relationship of the algebra with the loop algebra of $`sl_2`$, $`L(sl_2):=\text{IC}[t,t^1]sl_2`$, which was identified in as follows. Definition. The Onsager algebra, denoted by $`\mathrm{𝖮𝖠}`$, is defined as one of the following equivalent conditions: (i) $`\mathrm{𝖮𝖠}`$ = the universal Lie algebra generated by two elements, $`A_0,A_1`$, with the DG-condition. (ii) $`\mathrm{𝖮𝖠}`$ = the Lie-subalgebra of $`L(sl_2)`$ fixed by the involution $`\widehat{\theta }`$, $$\widehat{\theta }:p(t)p(t^1),ef,fe,hh,$$ where $`p(t)\text{IC}[t,t^1]`$, $`e,f,h`$ are the standard basis of $`sl_2`$ with $`[e,f]=h,[h,e]=2e,[h,f]=2f`$. $`\mathrm{}`$ The sequence, $`A_m,G_m`$, associated to the universal elements, $`A_0,A_1`$ of (i) have the following expression as elements in (ii), $$A_m=2(t^me+t^mf),G_m=(t^mt^m)h,\mathrm{for}m\text{ZZ}.$$ We shall always make the above identification in what follows. For an element $`X`$ in $`L(sl_2)`$, the criterion of $`X`$ in $`\mathrm{𝖮𝖠}`$ is now given by $$X\mathrm{𝖮𝖠}X=p(t)e+p(t^1)f+q(t)h,\mathrm{with}q(t)+q(t^1)=0,$$ where $`p(t),q(t)\text{IC}[t,t^1]`$. In fact, $`q(t)`$ can always be written in the form, $`q(t)=q_+(t)q_+(t^1)`$ with $`q_+(t)\text{IC}[t]`$. ## 3 Closed Ideals of the Onsager Algebra A (non-trivial Lie) ideal $`𝖨`$ of a Lie algebra $`𝖫`$ (over IC) is called a closed ideal if it satisfies the following condition, $$𝖨=\{x𝖫|[x,𝖫]𝖨\},$$ or equivalently, $`𝖫/𝖨`$ has the trivial center. By Schur’s lemma, the kernel ideal of an irreducible representation of $`𝖫`$ in $`sl_n(\text{IC})`$ is always closed, which constitutes an important class of closed ideals. In this section, we shall describe the classification of closed ideals of the Onsager algebra, OA. Definition. (i) Let $`P(t)`$ be a non-trivial monic polynomial in $`\text{IC}[t]`$. We call $`P(t)`$ a reciprocal polynomial if $`P(t)=\pm t^dP(t^1)`$, where $`d`$ is the degree of $`P(x)`$. (ii) For a reciprocal polynomial $`P(t)`$, $`𝖨_{P(t)}`$ is the ideal of $`\mathrm{𝖮𝖠}`$ defined by $$𝖨_{P(t)}:=\{X=p(t)e+p(t^1)f+q(t)h\mathrm{𝖮𝖠}|p(t),q(t)P(t)\text{IC}[t,t^1]\}.$$ We shall call $`P(t)`$ the generating polynomial of the ideal $`𝖨_{P(t)}`$. $`\mathrm{}`$ It is easy to see that zeros of a reciprocal polynomial $`P(t)`$ not equal to $`\pm 1`$ must occur in the reciprocal pairs, and $`𝖨_{P(t)}`$ is invariant under the involution of $`\mathrm{𝖮𝖠}`$ , $`A_mA_m,G_mG_m`$. Through the Chinese remainder theorem, one can establish the a canonical (Lie-)isomorphism of the quotient algebras, $$\mathrm{𝖮𝖠}/𝖨_{P(t)}\stackrel{}{}\underset{j=1}{\overset{J}{}}\mathrm{𝖮𝖠}/𝖨_{P_j(t)},P(t):=\underset{j=1}{\overset{J}{}}P_j(t)$$ where $`P_j(t)`$ are pairwise relatively prime reciprocal polynomials. The role of reciprocal polynomials in the study of the Onsager algebra is given by the following structure theorem of closed ideals of $`\mathrm{𝖮𝖠}`$. Theorem 1. An ideal $`𝖨`$ of $`\mathrm{𝖮𝖠}`$ is closed if and only if $`𝖨=𝖨_{P(t)}`$ for a reciprocal polynomial $`P(t)`$ whose zero at $`t=\pm 1`$ are of the even multiplicity. The ideal $`𝖨_{P(t)}`$ is characterized as the minimal closed ideal of $`\mathrm{𝖮𝖠}`$ containing $`P(t)e+P(t^1)f`$. $`\mathrm{}`$ For an irreducible special representation of $`\mathrm{𝖮𝖠}`$ on a finite dimensional vector space $`V`$, $`\rho :\mathrm{𝖮𝖠}sl(V)`$, the generating polynomial of the kernel $`\mathrm{Ker}(\rho )`$ is given by $`P(t)=_{j=1}^nU_{a_j}(t)`$, where $`a_j\text{IC}^{}\{\pm 1\},a_ja_i^{\pm 1}`$ for $`ji`$, and $`U_a(t):=(ta)(ta^1)`$. In this situation, the evaluation morphism of $`\mathrm{𝖮𝖠}`$, induced from $`L(sl_2)`$, into the sum of $`n`$ copies of $`sl_2`$, $$e_{a_1,\mathrm{},a_n}:\mathrm{𝖮𝖠}\stackrel{n}{}sl_2,X(e_{a_1}(X),\mathrm{},e_{a_n}(X)),$$ gives rise the isomorphism between $`\mathrm{𝖮𝖠}/\mathrm{Ker}(\rho )`$ and $`\stackrel{n}{}sl_2`$. Hence irreducible representations of the $`sl_2`$-factors determine an irreducible representation of $`\mathrm{𝖮𝖠}`$. ## 4 Completion of the Onsager Algebra and Solvable Lie Algebras By the previous discussion, the study of closed ideals in $`\mathrm{𝖮𝖠}`$ can be reduced to the case, $`𝖨=𝖨_{P(t)}`$ with $`P(t)=(t\pm 1)^L`$ or $`U_a(t)^L`$ for $`L\text{ZZ}_0,a\text{IC}^{}\{\pm 1\}`$. In this report, we shall only discuss the case, $`P(t)=(t\pm 1)^L`$. As the map, $`tt`$, gives rise an involution of $`\mathrm{𝖮𝖠}`$, one has the isomorphism, $$\mathrm{𝖮𝖠}/𝖨_{(t1)^L}\mathrm{𝖮𝖠}/𝖨_{(t+1)^L}.$$ We may assume, $`P(t)=(t1)^L,(L\text{ZZ}_0)`$. Denote $`\pi _L,\pi _{KL}`$ the canonical projections, $$\begin{array}{cc}\pi _L:\mathrm{𝖮𝖠}\mathrm{𝖮𝖠}/𝖨_{(t1)^L},\hfill & \pi _{KL}:\mathrm{𝖮𝖠}/𝖨_{(t1)^L}\mathrm{𝖮𝖠}/𝖨_{(t1)^K},LK0,\hfill \end{array}$$ and $`\widehat{\mathrm{𝖮𝖠}}`$ the projective limit of the projective system, $`(\mathrm{𝖮𝖠}/𝖨_{(t1)^L},\pi _{KL})`$. Then $`\widehat{\mathrm{𝖮𝖠}}`$ is a Lie algebra and denote the canonical morphism, $`\psi _L:\widehat{\mathrm{𝖮𝖠}}\mathrm{𝖮𝖠}/𝖨_{(t1)^L},L\text{ZZ}_0`$, with the kernel, $`\widehat{\mathrm{𝖮𝖠}}^L:=\mathrm{Ker}(\psi _L)`$. We have, $`\widehat{\mathrm{𝖮𝖠}}/\widehat{\mathrm{𝖮𝖠}}^L\mathrm{𝖮𝖠}/𝖨_{(t1)^L}`$. We have a filtration of ideals in $`\widehat{\mathrm{𝖮𝖠}}`$, $$\widehat{\mathrm{𝖮𝖠}}=\widehat{\mathrm{𝖮𝖠}}^0\widehat{\mathrm{𝖮𝖠}}^1\mathrm{}\widehat{\mathrm{𝖮𝖠}}^L\mathrm{}.$$ There exists a morphism, $`\pi :\mathrm{𝖮𝖠}\widehat{\mathrm{𝖮𝖠}}`$, with $`\psi _L\pi =\pi _L`$. The Lie algebra $`\widehat{\mathrm{𝖮𝖠}}`$ is regarded as a completion of $`\mathrm{𝖮𝖠}`$. For the convenience, we shall write the element $`\pi (X)`$ of $`\widehat{\mathrm{𝖮𝖠}}`$ again by $`X`$ for $`X\mathrm{𝖮𝖠}`$ if no confusion could arise. There exists the unique sequence of elements, $`X_k,Y_k,(k\text{ZZ}_0)`$, in $`\widehat{\mathrm{𝖮𝖠}}`$ such that the following identities hold in $`\widehat{\mathrm{𝖮𝖠}}`$, $$A_m(=\pi (A_m))=\underset{k0}{}\frac{m^{(k)}}{k!}X_k,G_m(=\pi (G_m))=\underset{k0}{}(1)^k\frac{m^{(k)}}{k!}Y_k,$$ where $`x^{(n)}`$ is the shifted factorial defined by $`x^{(0)}:=1,x^{(n)}:=x(x1)\mathrm{}(xn+1),n\text{ZZ}_{>0}`$. In fact, with the infinite formal sum $`\widehat{\mathrm{𝖮𝖠}}`$ has $`X_k,Y_k`$ as the generators of the formal Lie algebra, while $`\widehat{\mathrm{𝖮𝖠}}^L`$ is the ideal generated by $`X_k,Y_k,(kL)`$. In $`\mathrm{𝖮𝖠}/𝖨_{(t1)^L}`$, one has, $$\psi _L(A_m)=\underset{k=0}{\overset{L1}{}}\frac{m^{(k)}}{k!}\psi _L(X_k),\psi _L(G_m)=\underset{k=0}{\overset{L1}{}}(1)^k\frac{m^{(k)}}{k!}\psi _L(Y_k),m\text{ZZ}.$$ The relations (1) for $`\mathrm{𝖮𝖠}`$ now become the following relations in $`\widehat{\mathrm{𝖮𝖠}}`$, $`{\displaystyle \underset{n,k0}{}}{\displaystyle \frac{a!b!}{n!k!}}s_a^ns_b^k[X_n,X_k]=4{\displaystyle \underset{k0}{}}(1)^{b+k}{\displaystyle \frac{(a+b)!}{k!}}s_{a+b}^kY_k,`$ $`{\displaystyle \underset{n,k}{}}(1)^n{\displaystyle \frac{a!b!}{n!k!}}s_a^ns_b^k[Y_n,X_k]=2(1(1)^a){\displaystyle \underset{k}{}}{\displaystyle \frac{(a+b)!}{k!}}s_{a+b}^kX_k,`$ $`[Y_n,Y_k]=0,`$ where $`s_k^n,S_k^n,(n,k\text{ZZ}_0),`$ are the Stirling numbers, i.e., the integers with the relations, $`x^{(n)}=_{k0}x^ks_k^n,x^n=_{k0}x^{(k)}S_k^n`$. Note that $`s_n^n=S_n^n=1`$, and $`s_k^n=S_k^n=0`$ for $`k>n`$. By the identities among Stirling’s numbers , $$\underset{k}{}s_k^aS_b^k=\delta _b^a,\frac{b!}{a!}\underset{k}{}(1)^ks_k^aS_b^k=(1)^a(\begin{array}{c}a1\\ b1\end{array}),\frac{a!}{j!}S_a^j=\underset{l}{}\frac{k!(a+l)!}{l!(j+k)!}s_k^lS_{a+l}^{j+k},$$ one obtains the commutation relations of $`X_k,Y_k`$, $$\begin{array}{cc}[X_n,X_k]\hfill & =4(1)^n(Y_{n+k}+_{a>0}(\begin{array}{c}a+k1\\ a\end{array})Y_{n+k+a}),\hfill \\ [Y_n,X_k]\hfill & =2(((1)^n1)X_{n+k}_{a>0}(1)^a(\begin{array}{c}a+n1\\ a\end{array})X_{n+k+a}),\hfill \\ [Y_n,Y_k]\hfill & =0.\hfill \end{array}$$ By the above relations, $`\widehat{\mathrm{𝖮𝖠}}^L/\widehat{\mathrm{𝖮𝖠}}^{L+1}`$ is abelian for a positive integer $`L`$, hence $`\widehat{\mathrm{𝖮𝖠}}/\widehat{\mathrm{𝖮𝖠}}^L`$ is a finite-dimensional solvable Lie-algebra. However, there exist certain non-trivial relations among $`Y_k`$. In fact, one can express $`Y_{2n}`$ in terms of $`Y_{2k+1}`$ for $`kn`$ in the algebra $`\widehat{\mathrm{𝖮𝖠}}`$; the explicit formula is given by $$Y_{2n}=\underset{kn}{}(1)^{kn+1}\frac{(4^{kn+1}1)B_{kn+1}}{kn+1}(\begin{array}{c}2k\\ 2n1\end{array})Y_{2k+1},$$ where $`B_j(j1)`$ are the Bernoulli numbers, $`B_j=(1)^{j1}b_{2j}`$, with $`\frac{x}{e^x1}=_{j=0}^{\mathrm{}}\frac{b_j}{j!}x^j`$. For the study of the Lie algebra structure of $`\widehat{\mathrm{𝖮𝖠}}`$, it is more natural to use a local coordinate system near $`t=1`$ for expressing elements in $`L(sl_2)`$. A convenient variable is, $`\lambda :=\frac{tt^1}{2}`$, near the origin, $`\lambda =0`$. The structure of $`\widehat{\mathrm{𝖮𝖠}}`$ can be visualized as a formal subalgebra of $`sl_2[[\lambda ]](=\text{IC}[[\lambda ]]sl_2)`$ as follows. Theorem 2 . Define $$\begin{array}{cc}sl_2\lambda \hfill & :=\text{IC}[[\lambda ^2]]h+\lambda \text{IC}[[\lambda ^2]]e+\lambda \text{IC}[[\lambda ^2]]fsl_2[[\lambda ]],\hfill \\ sl_2\lambda ^L\hfill & :=sl_2\lambda \lambda ^Lsl_2[[\lambda ]],(L\text{ZZ}_0).\hfill \end{array}$$ There is a formal Lie-algebra isomorphism, $$\widehat{\mathrm{𝖮𝖠}}sl_2\lambda ,$$ under which $`\widehat{\mathrm{𝖮𝖠}}^L`$ is corresponding to $`sl_2\lambda ^L`$. As a consequence, $$\mathrm{𝖮𝖠}/𝖨_{(t1)^L}sl_2\lambda /sl_2\lambda ^L.$$ $`\mathrm{}`$ By the above Theorem, $`\mathrm{𝖮𝖠}/𝖨_{(t1)^L}`$ is a solvable Lie algebra of dimension $`L+[\frac{L}{2}]`$, which has a basis consisting of $`\psi _L(X_k),\psi _L(Y_l),0k,l<L,l1(mod2)`$. By the structure of $`\mathrm{𝖮𝖠}/𝖨_{(t1)^L}`$, one can easily see that the criterion of trivial center for $`\mathrm{𝖮𝖠}/𝖨_{(t1)^L}`$ , (or equivalently, the closed ideal for $`𝖨_{(t1)^L}`$), is given by the integer $`L`$ to be even. For odd $`L`$, the center of $`\mathrm{𝖮𝖠}/𝖨_{(t1)^L}`$ is 1-dimensional. Using the relations, $$t1=\lambda 1+\sqrt{1+\lambda ^2},t^11=\lambda 1+\sqrt{1+\lambda ^2},$$ one obtains an explicit expression of $`A_m,G_m`$, in terms of $`\lambda `$ in the above theorem, in particular, $$A_0=2(e+f),A_1=2(ef)\lambda +2(e+f)\underset{j0}{}(\begin{array}{c}\frac{1}{2}\\ j\end{array})\lambda ^{2j}.$$ By taking the expression of of $`A_0,A_1`$ modular $`\lambda ^Lsl_2[[\lambda ]]`$, one can construct reducible representations of $`\mathrm{𝖮𝖠}`$ with the kernel ideal generated by $`(t1)^L`$. ## 5 Further Remarks For the understanding of representations of the Onsager algebra, both the reducible and irreducible ones, the structure of $`\mathrm{𝖮𝖠}/𝖨_{P(t)}`$ with a reciprocal polynomial $`P(t)`$ warrants the mathematical investigation. But, just to keep things simple, we restrict our attention in this present report only to the case, $`P(t)=(t\pm 1)^L,L1`$, or $`U_a(t),a\pm 1`$. With a similar argument, one can also obtain the structure of quotients for closed ideals generated by $`P(t)=U_a(t)^L,L\text{ZZ}_{>0}`$, ( for the details, see ). The results, together with the mixed types, should have some interesting applications and implications in solvable or nilpotent algebras, which we shall leave to future work. Here is the one example which appeared in : Example. $`\mathrm{𝖮𝖠}/𝖨_{(t^2+1)^2}sl_2[\epsilon ]/\epsilon ^2sl_2[\epsilon ]`$. Denote $`e_k,f_k,h_k,(k=0,1)`$ the basis elements of $`sl_2[\epsilon ]/\epsilon ^2sl_2[\epsilon ]`$ corresponding to the standard basis of $`sl_2`$ with the index $`k`$ indicating the grade of $`\epsilon `$. Define the linear map, $`\phi :\mathrm{𝖮𝖠}sl_2[\epsilon ]/\epsilon ^2sl_2[\epsilon ]`$, by the expression: $$\begin{array}{c}\phi (A_m)=2(\mathrm{i}^me_0+\mathrm{i}^mf_0+m\mathrm{i}^{m1}e_1+m\mathrm{i}^{m+1}f_1),\hfill \\ \phi (G_m)=(\mathrm{i}^m\mathrm{i}^m)h_0+m(\mathrm{i}^{m1}\mathrm{i}^{m1})h_1.\hfill \end{array}$$ Then $`\phi `$ is a surjective morphism. Since $`sl_2[\epsilon ]/\epsilon ^2sl_2[\epsilon ]`$ has only trivial center, $`\mathrm{Ker}(\phi )`$ is a closed ideal with $`(t^2+1)^2`$ as the generating polynomial. $`\mathrm{}`$ Generalizations of the Onsager algebra to other loop algebras or Kac-Moody algebras as in should provide ample examples of solvable algebras of the above type. We hope that the further development of the study of the subject will eventually lead to some interesting results in Lie-theory with possible applications in quantum integrable models.
warning/0002/astro-ph0002439.html
ar5iv
text
# C32, A Young Star Cluster in IC 1613 ## 1 Introduction Baade (1963) remarked on the fact, which he obviously considered remarkable, that the irregular galaxy IC 1613, a member of the Local Group that he had studied quite intensively, appeared to be devoid of star clusters. This fact was again discussed by van den Bergh (1979), who compared IC 1613 with the SMC, pointing out that the SMC’s many rich clusters are so conspicuous that, if IC 1613 has any, they should show up clearly on available plate material. Hodge (1978) had searched photographic plates taken with the Palomar Observatory’s 5m and the Lick Observatory’s 3m telescopes and published a list of 43 possible candidates for clusters, all of which were very inconspicuous, none having more than 6 resolved stars on the best images. Clearly, these objects were not comparable to the rich star clusters of the Magellanic Clouds, but he thought that at least some of them might be similar to small Galactic open clusters like the Pleiades. More recently Freedman (1988), in her CCD-based study of the color-magnitude diagram (CMD) of a portion of IC 1613, noted that she could not identify those candidate clusters that should have been visible on her images. On the other hand, Georgiev et al. (1999) observed a field centered on the most active area of star formation (east of our present field), identified 12 of the earlier cluster candidates and measured integrated UBV magnitudes for eight, six of which they found to have colors indicating ages of less than 10 million years. They also identified two new small, young clusters. They searched for massive clusters, either young or old globular clusters, but found none. For a related discsussion of the young massive cluster populations in galaxies and their correlations with integrated galaxy properties see Larsen & Richtler (2000). Thus it has been clear that IC 1613 must be poor in rich star clusters compared with the Magellanic Clouds, a fact which van den Bergh attributed to the lack of a history of strong shocks. But it has not been clear whether or not IC 1613 is lacking in a significant number of normal open clusters, as most of the published candidate clusters have not been examined on anything but the original photographic plates. This paper reports the results of a study of some of these candidates using HST and WIYN<sup>1</sup><sup>1</sup>1The WIYN Observatory is a joint facility of the University of Wisconsin-Madison, Indiana University, Yale University, and the National Optical Astronomy Observatories. telescope images. ## 2 Observational Material For our cluster search we have used multi-color images of IC 1613 that were obtained to study the stellar populations in IC 1613 , which have been fully discussed elsewhere (Cole et al., 1999). One set of images is a single pointing of HST centered near the geometrical center of IC 1613 at RA = $`1^h`$ $`04^m`$ $`48.7^s`$, declination = $`+02\mathrm{°}`$ 07$`\mathrm{}`$ 06.2$`\mathrm{}`$ (J2000). WFPC2 images were taken with the F439W, F555W and F814W filters. All are available from the HST Data Archive (PID 6865). The other set was obtained by the WIYN telescope at Kitt Peak towards the same position using a CCD detector and B, V and I filters, taken under 0.6 arcsec seeing conditions. Both sets of images are centered on the HI ”hole” (Skillman, 1987; Lake & Skillman, 1989; Westpfahl et al., 1996) at the center of the galaxy and do not include the most active star-forming regions, which are concentrated primarily to the east of these images (Hodge, 1978; Hodge et al., 1990; Price et al., 1990). The WIYN images are 6.8 arc minutes on a side, providing an area of 46.24 square arc minutes, while the entire area of IC 1613, to the V = 25 magnitudes per square arcsec isophote, is 283 square arc minutes. Thus the region surveyed here is about 16% of the main body of the galaxy. ## 3 Survey Results The HST images included only two of the previously-identified cluster candidates, nos. 22 and 33, both of which had been considered faint and doubtful. The HST images revealed both to be small asterisms, with no indication of the presence of any physical clustering of stars (Fig. 1). The WIYN images covered an additional 21 candidate clusters. One of these, cluster C32, has the appearance of a true, small but bright cluster (Fig. 2). Five other candidates are possible examples of small, very sparse clusters, for which the WIYN images are not definitive. The others are either small asterisms or background galaxies with foreground stars superimposed, making them appear to be star clusters when imaged with insufficient resolution. Table 1 lists the candidates examined and their disposition. ## 4 Cluster C 32 Cluster C32 (Fig. 2) is a small cluster located near the center of OB association No. 5 (Hodge, 1978). The OB association is involved with HII region S2 (Sandage, 1971), which is cataloged as HLG 13 (Hodge et al., 1990). The cluster is thus located in a small area of current star formation. The entire WIYN image is shown in Figure 3 with the position of C32 indicated. Figure 4 shows the radial distribution of stars in the cluster and its surroundings. The half-light diameter of the cluster is only of the order of 2.0 arcsecs. One arcsec at IC 1613’s distance (Lee et al., 1993) corresponds to 3.5 parsecs, implying that cluster C32 is approximately 7 parsecs across, making it similar in size to many small young open clusters in the Milky Way Galaxy. Photometry of the stars in the WIYN B and V images was extracted using the PSF-fitting software DAOPHOTII and ALLSTAR (Stetson, 1994). The magnitude zeropoints were determined via comparison with the B and V photometry of Freedman (1988). We first determined an astrometric solution for the WIYN data using the Digitized Sky Survey images available from the Space Telescope Science Institute. We then matched together the positions of relatively bright stars measured by Freedman in her Field 1 that appear in our WIYN images. The average offset between our instrumental psf-fit magnitudes and Freedman’s photometry was determined using a total of 24 and 25 stars in B and V, respectively. The rms scatter for the sample of comparison stars was 0.05 magnitudes in both filters. Figure 5 is a color-magnitude diagram that shows the positions of the few well-separated stars of C32 detected in both B and V superimposed on the CMD of the entire field of the WIYN images. Clearly the cluster is very young, with a main sequence that extends to M$`{}_{V}{}^{}=5`$. There are too few stars to establish a definitive age from the CMD, but the cluster is clearly younger than about 10 million years, a conclusion that agrees with its position in a young OB association that is enveloped in an HII region. To measure the integrated magnitude and color of C32, we first calculated the background as the mean flux in an annular region between $`6\mathrm{}`$ and $`8\mathrm{}`$ from the center of C32. This value was then subtracted from the flux within a circular aperture of radius 13 pc. The error in the flux was determined by randomly choosing 100 aperture centers within a $`1.6\mathrm{}\times 1.6\mathrm{}`$ area surrounding C32. At each of these positions, we calculated the flux and background using the same aperture and sky annulus as used for the cluster. The resulting flux values had an average of zero and a standard deviation that we assume as the probable error in the aperture magnitude of C32. Assuming a foreground reddening of E(B$``$V)$`=0.025`$ (Schlegel et al., 1998), the integrated color and magnitude of C32 within a radius of 13 pc are M$`{}_{V}{}^{}=5.78\pm 0.16`$ and (B$``$V)$`{}_{0}{}^{}=0.19\pm 0.08`$. These values of color and absolute magnitude are fully consistent with those of small young star clusters in other galaxies, such as M31 (Hodge et al., 1987). It is interesting to make a quantitative comparison of the cluster density in IC 1613 with that in other galaxies. We have counted the number of star clusters found in the LMC on small-scale plate surveys (approximately equivalent in physical scale and limiting absolute magnitude to the WYIN survey of IC 1613) in three areas each equal in physical size to the WYIN IC 1613 field and located at a corresponding optical surface brightness. While we have found only one certain star cluster in the IC 1613 field, the LMC fields averaged $`81\pm 9`$ clusters. Details of this comparison, as well as comparisons with other galaxies, are given in Hodge (2000). ## 5 Conclusions We have surveyed the central area of IC 1613 for star clusters. There are no globular clusters or massive young clusters present, but there are six very sparse groupings that may be open clusters like some of the smaller ones in our Galaxy. One of them, previously cataloged as C32, is a very young cluster embedded in an HII region and surrounded by an OB association. When the uncertain nature of many of the cluster candidates, their very small sizes and their implied small stellar populations are taken into account, it is clear that IC 1613 is a cluster-poor galaxy. We are grateful to NASA’s Space Telescope Science Institute for partial support of this research through grant AR-08362 to the University of Washington. The original observations were made with the NASA/ESA Hubble Space Telescope, obtained at the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy. Inc., under NASA contract NAS 5-26555, and with the WIYN Telescope at the Kitt Peak National Observatory. Figure Captions
warning/0002/astro-ph0002311.html
ar5iv
text
# On the origin of the color Tully-Fisher and color-magnitude relations of disk galaxies Instituto de Astronomía-UNAM, A.P. 70-264, 04510 México, D. F. <sup>1</sup><sup>1</sup>footnotetext: Also Osservatorio Astronomico di Brera, via E.Bianchi 46, I-23807 Merate, Italy $``$ The increment of the slope of the T-F relation (TFR) with the passband wavelength has been called the color TFR. For example, from the observed $`K`$ and $`B`$band TFRs one obtains that $`L_K/L_BV_m^{0.6}`$ (Pierce & Tully1999, ApJ, 387,47). Since $`L_K`$ traces the stellar disk mass $`M_s`$, then $`M_s/L_BV_m^{0.6}`$. The empirical fact that more luminous galaxies tend to be redder have been called the color-magnitude relation: $`(BH)\gamma `$log$`L_B`$, $`\gamma 0.41.2`$. $``$ Why the mass-to-luminosity ratio and the color index of disk galaxies do depend on $`V_m`$ (or luminosity)? At least there are three alternatives: the star formation (SF) efficiency, the gas infall efficiency and/or the internal face-on dust extinction depend on the galaxy mass. We explore the last alternative since self-consistent models of disk galaxy evolution within the hierarchical formation scenario show that the $`M_s/L_B`$ ratio and $`BH`$ do not significantly depend on mass or luminosity (Avila-Reese & Firmani 2000, RevMexA&A, v. 36, in press; Avila-Reese et al., this volume). $``$ Observations indeed show that dust and metallicity increase with $`L_B`$. In a more quantitative fashion, Wang & Heckman (1996: ApJ, 457, 645, WH) have established that the UV(young massive stars)-to-FIR (the same young stars +dust absorption) ratio decreases rapidly with $`L_B`$ $``$ the dust opacity increases with $`L_B`$: $`\tau _B=\tau _{B,o}(L_B/L_{B,o})^\beta `$ (eq. 1), with $`L_{B,o}=1.3\times 10^{10}L_B`$, $`\tau _{B,o}=0.8\pm 0.3`$, and $`\beta =0.5\pm 0.2`$ (WH). $``$ Applying the uniform slab model, and using eq. (1) with the central values for the constants, a good approximation for the extinction in the range of $`10^810^{11}L_B`$ is: $`A_B0.38+0.42`$log$`L_{B,10}+0.14(`$log$`L_{B,10})^2`$ (eq. 2). Assuming that in the $`H`$ band dust absorption is negligible, this result shows that $`(BH)`$ will redden due to dust extinction roughly as $`0.42`$log$`L_B`$, which reasonable agrees with that is observed. On the other hand, in the understanding that the origin of the TFR in the different bands is common, the fact that $`L_K/L_BV_m^{0.6}`$ is easily accounted for the $`B`$band extinction given by eq. (2). In the hierarchical scenario of galaxy formation indeed the TFR for any band is a common imprint of the mass-velocity relation of the CDM halos. $``$ In conclusion, the luminosity-dependent dust extinction reported by WH easily explains the color TF and color-magnitude relations of disk galaxies; there is not necessity to evoke mass dependent SF and gas infall efficiencies. The extinction could depend on mass because the efficiency of metal ejection out of the disk might be larger for smaller galaxies and/or because more massive galaxies have higher surface densities than the smaller ones.
warning/0002/math-ph0002038.html
ar5iv
text
# Riemannian manifolds with uniformly bounded eigenfunctions ## 0. Introduction This paper is concerned with the relation between the dynamics of the geodesic flow $`G^t`$ on the unit sphere bundle $`S^{}M`$ of a compact Riemannian manifold $`(M,g)`$ and the growth rate of the $`L^{\mathrm{}}`$-norms of its $`L^2`$-normalized $`\mathrm{\Delta }`$-eigenfunctions (or ‘modes’) $`\{\varphi _\lambda \}`$. Let $`V_\lambda :=\{\varphi :\mathrm{\Delta }\varphi _\lambda =\lambda \varphi _\lambda \}`$ denote the $`\lambda `$-eigenspace for $`\lambda Sp(\mathrm{\Delta })`$ and define (1) $$L^{\mathrm{}}(\lambda ,g)=\underset{\genfrac{}{}{0pt}{}{\varphi V_\lambda }{\varphi _{L^2}=1}}{sup}\varphi _L^{\mathrm{}},\mathrm{}^{\mathrm{}}(\lambda ,g)=\underset{ONB\{\varphi _j\}V_\lambda }{inf}(\underset{j=1,\mathrm{},dimV_\lambda }{sup}\varphi _j_L^{\mathrm{}}).$$ The universal bound $$L^{\mathrm{}}(\lambda ,g)=0(\lambda ^{\frac{n1}{4}})$$ holds for any $`(M,g)`$ in consequence of the local Weyl law \[Ho IV\] $$N(T,x):=\underset{j:\lambda _jT}{}|\varphi _{\lambda _j}(x)|^2=\frac{1}{(2\pi )^n}vol(M,g)T^{\frac{n}{2}}+R(T,x),R(T,x)=O(T^{\frac{n1}{2}}).$$ It is attained in the case of the standard $`(S^n,can)`$ (by the zonal spherical harmonics) but is far off in the case of irrational flat tori $`(T^n,ds^2)`$ where $`L^{\mathrm{}}(\lambda ,g)=O(1).`$ These cases represent the extremes, and the problem arises of characterizing the manifolds with extremal growth rates of $`L^{\mathrm{}}`$-norms of eigenfunctions. In this article, we are interested in the case of minimal growth: * Problem: Determine the $`(M,g)`$ for which $`\mathrm{}^{\mathrm{}}(\lambda ,g)=O(1)`$ and those for which $`L^{\mathrm{}}(\lambda ,g)=O(1).`$ The same kind of problem may be posed in the more general setting of semi-classical Schroedinger operators $`\mathrm{}^2\mathrm{\Delta }+V`$. The eigenvalue problem $`(\mathrm{}^2\mathrm{\Delta }+V)\varphi _j=E_j(\mathrm{})\varphi _j`$ now depends on $`\mathrm{}`$, and we are interested in the behaviour of eigenfunctions $`\varphi _j`$ in the semiclassical limit $`\mathrm{}0`$. The spectrum becomes dense around each regular value $`E`$ of the classical Hamiltonian $`H(x,\xi )=|\xi |_g^2+V(x)`$ on $`T^{}M`$, and for any $`0<\delta <1`$, the asymptotics of spectral data from an interval $`[Ec\mathrm{}^{1\delta },E+c\mathrm{}^{1+\delta }]`$ around $`E`$ will reflect the dynamics of the classical Hamiltonian flow $`\mathrm{\Phi }_t^E`$ on the energy surface $`X_E=\{H(x,\xi )=E\}.`$ We fix $`E`$ and $`0<\delta <1`$, and consider the eigenvalues $`E_j(h)[Ec\mathrm{}^{1\delta },E+c\mathrm{}^{1\delta }].`$ Denote by $`V_{E_j(\mathrm{})}`$ the eigenspace of eigenvalue $`E_j(h)`$ and put: (2) $$L^{\mathrm{}}(\mathrm{},E_j(\mathrm{});g,V)=\underset{\genfrac{}{}{0pt}{}{\varphi V_{E_j(\mathrm{})}}{\varphi _{L^2}=1}}{sup}\varphi _L^{\mathrm{}},\mathrm{}^{\mathrm{}}(h,E_j(\mathrm{});g,V)=\underset{ONB\{\varphi _j\}V_{E_j(\mathrm{})}}{inf}(\underset{j=1,\mathrm{},dimV_{E_j(\mathrm{})}}{sup}\varphi _L^{\mathrm{}}),$$ and pose the analogous questions: * Problem: Determine the $`(M,g,V)`$ for which there exists a regular energy level $`E`$ such that $`\mathrm{}^{\mathrm{}}(\mathrm{},E_j(\mathrm{});g,V)=O(1)`$ and the $`(M,g,V)`$ for which $`L^{\mathrm{}}(\mathrm{},E_j(\mathrm{});g,V)=O(1)`$ as $`\mathrm{}0`$ with $`E_j(\mathrm{})[Ec\mathrm{}^{1\delta },E+c\mathrm{}^{1\delta }]`$ for some $`c>0`$. The problem on Laplace operators is the same as the problem on Schroedinger operators in the case $`V=0`$, for any value of $`E>0.`$ The problems about $`\mathrm{}^{\mathrm{}}`$ asks which Laplacians or Schroedinger operators possess an ONBE (orthonormal basis of eigenfunctions) of minimal growth. The problems about $`L^{\mathrm{}}`$ ask which ones have the property that every ONBE has minimal growth. Obviously, the distinction between $`\mathrm{}^{\mathrm{}}`$ and $`L^{\mathrm{}}`$ only arises when the spectrum of $`\mathrm{\Delta }`$ is multiple. At the opposite extreme, one may ask which $`(M,g)`$ possess eigenfunctions which achieve the maximal rate of growth, but we will not discuss that problem here. One may also pose quantitative problems of giving upper and lower bounds on $`\mathrm{}^{\mathrm{}}(\lambda ,g),L^{\mathrm{}}(\lambda ,g)`$, and their $`L^p`$-analgoues, under various dynamical hypotheses. Some results on such quantitative problems will be given in a subsequent article \[TZ\]. The known connections between $`\mathrm{}^2\mathrm{\Delta }+V`$\- eigenfunctions and the dynamics of $`\mathrm{\Phi }_t^E`$ are not strong enough at present to answer these questions in the general setting of compact Riemannian manifolds. If, however, the systems are assumed to be completely integrable geodesic flows then much more can be said. Let us assume in fact that $`\mathrm{\Delta }`$ is quantum completely integrable in the (well-known) sense that there exist $`P_1,\mathrm{},P_n\mathrm{\Psi }^1(M)`$ ($`n`$ = dim $`M`$) satisfying * $`[P_i,P_j]=0`$; * $`dp_1dp_2\mathrm{}dp_n0`$ on a dense open set $`\mathrm{\Omega }T^{}M0`$ of ‘finite complexity’ (see below); * $`\sqrt{\mathrm{\Delta }}=\widehat{K}(P_1,\mathrm{},P_n)`$ for some polyhomogeneous function $`\widehat{K}`$ on $`^n0.`$ Here, $`\mathrm{\Psi }^m(M)`$ is the space of mth order pseudodifferential operators over $`M`$, and $`p_k=\sigma _{P_k}`$ is the principal symbol of $`P_k`$. Since $`\sigma _{[P_i,P_j]}=\{p_i,p_j\}`$ (the Poisson bracket), it follows that the $`p_j`$’s generate a homogeneous Hamiltonian action $`\mathrm{\Phi }_t`$ of $`t^n`$ on $`T^{}M0`$ with moment map $$𝒫:T^{}M0^n,𝒫=(p_1,\mathrm{},p_n).$$ We denote the image $`𝒫(T^{}M0)`$ by $`B`$, by $`B_{reg}`$ (resp. $`B_{sing}`$) the regular values (resp. singular values) of the moment map. By ‘finite complexity’ we mean the following: for each $`b=(b^{(1)},\mathrm{},b^{(n)})B`$, let $`m_{cl}(b)`$ denote the number of $`^n`$-orbits of the joint flow $`\mathrm{\Phi }_t`$ on the level set $`𝒫^1(b)`$. Then (3) $$\text{Finite complexity condition}:M:m_{cl}(b)<M(bB).$$ When $`bB_{reg}`$, then $`𝒫^1(b)`$ is the union of $`m_{cl}(b)`$ isolated Lagrangean tori. If $`bB_{sing}`$, then $`𝒫^1(b)`$ consists of a finite number of connected components, each of which is a finite union of orbits. These orbits may be Lagrangean tori, singular compact tori (ie. compact tori of dimension $`<n`$) , or non-compact orbits consisting of cylinders or planes. We will also make the following assumption on the quantum level: (4) $$\begin{array}{ccccc}\text{Bounded eigenvalue multiplicity }:\hfill & M^{}:m(\lambda )M^{}\hfill & & (\lambda ;m(\lambda )=dimV_\lambda ).\hfill & \end{array}$$ With this assumption, $`L^{\mathrm{}}`$ is bounded by a constant times $`\mathrm{}^{\mathrm{}}`$, so all ONBE’s are uniformly bounded if and only if one is. Without assumption (4), it is simple to construct an ONBE which is not uniformly bounded. We will recall the construction in §4, and discuss some open problems in which the bounded eigenvalue multiplicity is dropped. The Hamiltonian $`|\xi |_g=\sqrt{_{i,j=1}^ng^{ij}(x)\xi _i\xi _j}`$ is then given by $`|\xi |_g=K(p_1,\mathrm{},p_n)`$ where $`K`$ is the homogeneous term of order 1 of $`\widehat{K}.`$ Hence the geodesic flow commutes with a Hamiltonian $`^n`$-action, i.e. it is completely integrable. We will assume throughout the following properness assumption: Our main result is the following rigidity theorem: ###### Theorem 0.1. Suppose that $`\mathrm{\Delta }`$ is a quantum completely integrable Laplacian on a compact Riemannian manifold $`(M,g)`$, and suppose that the corresponding moment map satisfies (3). Then: (a) If $`L^{\mathrm{}}(\lambda ,g)=O(1)`$ then $`(M,g)`$ is flat. (b) If $`\mathrm{}^{\mathrm{}}(\lambda ,g)=O(1)`$ and if (4) holds, then $`(M,g)`$ is flat. More generally, suppose that $`\mathrm{}^2\mathrm{\Delta }+V`$ is a quantum completely integrable Schroedinger operator, and that the corresponding moment map $`𝒫`$ is proper and satisfies (3). Assume there exists an energy level $`E`$ such that: (a) $`L^{\mathrm{}}(\mathrm{},E_j(\mathrm{}));g,V;)=O(1)`$ as $`\mathrm{}0`$; (b) $`\mathrm{}^{\mathrm{}}(h,E_j(\mathrm{});g,V)=O(1)`$ as $`\mathrm{}0`$, and (4) holds. Then: $`E>\mathrm{max}V,`$ and $`(M,(EV)g)`$ is flat. If (a) (or (b)) holds for all energy levels $`E`$ in an interval $`E_1<E<E_2`$, then $`(M,g)`$ is flat and $`V`$ is constant. As mentioned above, (a)-(b) are equivalent so we only consider (a) henceforth. We recall that flat manifolds are manifolds carrying a flat metric. By the Bieberbach theorems (\[W\], Theorems 3.3.1 - 3.3.2), a flat manifold $`(M,g)`$ may be expressed as the quotient $`M=^n/\mathrm{\Gamma }`$ of $`^n`$ by a discrete (crystallographic) subgroup of Euclidean motions $`\mathrm{\Gamma }E(n).`$ The subgroup $`\mathrm{\Gamma }^{}:=\mathrm{\Gamma }^n`$ is normal and of finite index in $`\mathrm{\Gamma }`$ so there exists a flat torus $`T^n=^n/\mathrm{\Gamma }^{}`$ and a finite normal Riemannian cover $`\pi :T^nM`$ with deck transformation group $`G=\mathrm{\Gamma }/\mathrm{\Gamma }^{}.`$ For each $`n>0`$, there are only finitely many affine equivalence classes of flat compact connected $`(M,g)`$ of dimension $`n`$ (affinely equivalent= same fundamental group), and in low dimensions they have been classified (cf. \[W\]). The eigenfunctions $`\varphi _\lambda `$ of $`\mathrm{\Delta }_g`$ on $`(M,g)`$ may be lifted to $`G`$-invariant eigenfunctions $`\pi ^{}\varphi _\lambda `$ on $`T^n`$ and hence the eigenspace $`E_\lambda (M,g)`$ may be identified with the $`G`$-invariant eigenspace $`E_\lambda (T^n,g_T)^G`$. The latter eigenfunctions may be written as sums of exponential functions. Let us outline the proof of the Theorem (0.1) in the simplest case of toric integrable systems (see §1 for background), and then explain what more is involved in the case of general integrable systems. By definition, the geodesic flow $`G_g^t:T^{}MT^{}M`$ of a compact Riemannian manifold $`(M,g)`$ is toric integrable if it commutes with a Hamiltonian action of the n-torus $`^n/^n`$. Equivalently, if there exist global action variables $`\{(I_j,\theta _j):j=1,\mathrm{},n\}`$ for the geodesic flow, i.e. functions of $`(p_1,\mathrm{},p_n)`$ whose Hamilton flows are $`2\pi `$-periodic. The level sets $$T_I:=^1(I)$$ of the moment map $$=(I_1,\mathrm{},I_n):T^{}M0^n$$ are then orbits $`^n/^n(x_o,\xi _o)`$ of the torus action and hence are tori. The image $`B`$ of $`T^{}M0`$ under $``$ is a convex polyhedral cone and $``$ is a Lagrangean torus bundle over its interior. Such moment maps $``$ are the cotangent bundle analogues of toric varieties in algebraic geometry. In the toric case, it is always possible to quantize the action variables as first order pseudodifferential action operators $`\widehat{I}_j`$ which commute with $`\mathrm{\Delta }`$. The actions define a (projective) action of $`^n/^n`$ by Fourier integral operators, or equivalently, the joint spectrum $`Sp(\widehat{I}_1,\mathrm{},\widehat{I}_1)`$ is contained in an (off-centered) lattice $`^n+\mu `$. The joint eigenfunctions $$(\widehat{I}_1,\mathrm{},\widehat{I}_n)\varphi _\lambda =\lambda \varphi _\lambda \lambda ^n$$ are therefore quantizations of the invariant Lagrangean torii $`𝒯_\lambda `$ with integral actions $`\lambda ^n+\mu `$. In particular, eigenfunctions $`\{\varphi _\lambda \}`$ localize on the invariant tori in the semiclassical limit in the sense that for any zeroth order pseudodifferential operator $`A`$ (with symbol $`\sigma _A)`$, (5) $$(A\varphi _{k\lambda },\varphi _{k\lambda })=_{T_\lambda }\sigma _A𝑑\mu _\lambda +O(k^1),$$ where $`d\mu _\lambda `$ is the normalized Lebesgue (probability) measure on $`T_\lambda .`$ Hence, $`|\varphi _\lambda (x)|^2`$ measures the density of the natural projection $`\pi _\lambda :𝒯_\lambda M`$ at $`x`$. The proof of Theorem (0.1) in the toric case is based on the following simple Lemmas. First we have: Suppose that $`G^t`$ is toric integrable and that $`L^{\mathrm{}}(M,g)=0(1)`$. Then every invariant torus $`T_\lambda `$ has a non-singular projection to $`M`$. The proof uses the fact that for any invariant torus $`T_I`$, there exists a sequence of joint eigenfunctions $`\{\varphi _\lambda \}`$ of the quantum torus action which localizes on $`T_I`$. Uniform boundedness of the eigenfunctions then implies regular projection of the tori. The second ingredient in the proof of the main theorem in the case of toric integrable systems is the following purely geometric statement which follows from the recently proved Hopf conjecture (cf. \[BI\] \[CK\]). Suppose that $`(M,g)`$ is a compact Riemannian manifold with toric integrable geodesic flow, and suppose that all the invariant torii project regularly to $`M`$. Then $`(M,g)`$ is a flat manifold. By ‘projecting regularly’ we mean that the projection has no singular values, hence (in view of the dimensions) is a covering map. The proof of Theorem (0.1) in the case of general Hamiltonian $`^n`$ actions is basically similar, but there are some new complications to handle. Geometrically, the new features are that the fibers $`𝒫^1(b)`$ may have several components (‘geometric multiplicity’), that there may exist non-compact orbits (e.g. embedded cylinders), and that there may exist singular orbits lying over the interior of the image of $`T^{}M0`$ under $`𝒫.`$ Analytically, the main new feature is that modes need not localize on individual components of $`𝒫^1(b)`$. What does localize on individual tori are quasimodes, i.e. semiclassical Lagrangean distributions which approximately solve the eigenvalue problem. In the toric case, modes and quasimodes are the same but this is not the case in general. As originally stressed by Arnold \[A\], and as is evident from simple examples such as the symmetric double well potential, eigenfunctions may be linear combinations of quasi-modes with very close quasi-eigenvalues and in the classical limit their mass concentrates in some way on the union of the components. How the mass is distributed involves the question whether the tori are resonant or not, and whether or not there is tunnelling between tori. We will discuss such relations between modes and quasimodes in detail in \[TZ\], where we prove (among other things) that quasimodes have uniformly bounded sup norms when modes do and where we determine precisely how modes blow up around singular orbits. In this paper, we take a softer approach via quantum limits of eigenfunctions and semiclassical trace formulae. We close with some acknowledgements. We thank Bruce Kleiner for pointing out the paper \[M\], Leonid Polterovich for helpful comments on \[BP\], and Francois Lalonde for helpful comments on an earlier version of the paper. We would especially like to thank the referee of this paper for pointing out that one of our original (non-degeneracy) hypotheses could be removed from the proof of Theorem (0.1), and for several other corrections and improvements. To clarify the ingredients in the proof, we cut the original manuscript (which appeared on the lanl archive as math-ph/0002038) into two parts, the present qualitative one and the subsequent quantitative one (\[TZ\]). ## 1. Background ### 1.1. Completely integrable systems By a completely integrable system on $`T^{}M`$ we mean a set of $`n`$ independent, $`C^{\mathrm{}}`$ functions $`p_1,\mathrm{},p_n`$, on $`T^{}M`$ satisfying: | $`\{p_i,p_j\}=0`$ for all $`1i,jn`$; | | --- | | $`dp_1dp_2\mathrm{}dp_n0`$ on an open dense subset of $`T^{}M.`$ | The associated moment map is defined by (6) $$𝒫=(p_1,\mathrm{},p_n):T^{}MB^n.$$ We refer to to the set $`B`$ as the ‘image of the moment map.’ The Hamiltonians generate an action of $`^n`$ defined by $$\mathrm{\Phi }_t=\mathrm{exp}t_1\mathrm{\Xi }_{p_1}\mathrm{exp}t_2\mathrm{\Xi }_{p_2}\mathrm{}\mathrm{exp}t_n\mathrm{\Xi }_{p_n}.$$ We often denote $`\mathrm{\Phi }_t`$-orbits by $`^n(x,\xi )`$. The isotropy group of $`(x,\xi )`$ will be denoted by $`_{(x,\xi )}.`$ When $`^n(x,\xi )`$ is a compact Lagrangean orbit, then $`_{(x,\xi )}`$ is a lattice of full rank in $`^n`$, and is known as the ‘period lattice’, since it consists of the ‘times’ $`T^n`$ such that $`\mathrm{\Phi }_T|_{\mathrm{\Lambda }^{(j)}(b)}=Id.`$ We will need the following: ###### Definition 1.1. We say that: * $`bB_{sing}`$ if $`𝒫^1(b)`$ is a singular level of the moment map, i.e. if there exists a point $`(x,\xi )𝒫^1(b)`$ with $`dp_1\mathrm{}dp_n(x,\xi )=0`$. Such a point $`(x,\xi )`$ is called a singular point of $`𝒫`$. * a connected component of $`𝒫^1(b)`$ ($`bB_{sing}`$)is a singular component if it contains a singular point ; * an orbit $`^n(x,\xi )`$ of $`\mathrm{\Phi }_t`$ is singular if it is non-Lagrangean, i.e. has dimension $`<n`$; * $`bB_{reg}`$ and that $`𝒫^1(b)`$ is a regular level if all points $`(x,\xi )𝒫^1(b)`$ are regular, i.e. if $`dp_1\mathrm{}dp_n(x,\xi )0`$. * a component of $`𝒫^1(b)`$ ( $`bB_{sing}B_{reg}`$) is regular if it contains no singular points. By the Liouville-Arnold theorem \[AM\], the orbits of the joint flow $`\mathrm{\Phi }_t`$ are diffeomorphic to $`^k\times T^m`$ for some $`(k,m),k+mn.`$ By the properness assumption on $`𝒫`$, a regular level has the form (7) $$𝒫^1(b)=\mathrm{\Lambda }^{(1)}(b)\mathrm{}\mathrm{\Lambda }^{(m_{cl})}(b),(bB_{reg})$$ where each $`\mathrm{\Lambda }^{(l)}(b)T^n`$ is an $`n`$-dimensional Lagrangian torus. The classical (or geometric) multiplicity function $`m_{cl}(b)=\mathrm{\#}𝒫^1(b)`$, i.e. the number of orbits on the level set $`𝒫^1(b)`$, is constant on connected components of $`B_{reg}`$ and the moment map (6) is a fibration over each component with fiber (7). In sufficiently small neighbourhoods $`\mathrm{\Omega }^{(l)}(b)`$ of each component torus, $`\mathrm{\Lambda }^{(l)}(b)`$, the Liouville-Arnold theorem also gives the existence of local action-angle variables $`(I_1^{(l)},\mathrm{},I_n^{(l)},\theta _1^{(l)},\mathrm{},\theta _n^{(l)})`$ in terms of which the joint flow of $`\mathrm{\Xi }_{p_1},\mathrm{},\mathrm{\Xi }_{p_n}`$ is linearized \[AM\]. For convenience, we henceforth normalize the action variables $`I_1^{(l)},\mathrm{},I_n^{(l)}`$ so that $`I_j^{(l)}=0;j=1,\mathrm{},n`$ on the torus $`\mathrm{\Lambda }^{(l)}(b)`$. When $`bB_{reg}`$, the Lagrangean tori $`\mathrm{\Lambda }^{(j)}(b)`$ of $`𝒫^1(b)`$ carry two natural measures, which we take some care to distinguish. ###### Definition 1.2. We define: * Lebesgue measure $`d\mu _b^{(j)}=(2\pi )^nd\theta _1\mathrm{}d\theta _n`$ on $`\mathrm{\Lambda }^{(j)}(b)`$, as the normalized (mass one) $`\mathrm{\Phi }_t`$-invariant measure on this orbit; * The Liouville measure $`d\omega _b^{(j)}`$ on $`\mathrm{\Lambda }^{(j)}(b)`$, as the surface measure induced by the moment map $`𝒫`$, i.e. $$d\omega _b^{(j)}=\frac{dV}{dp_1\mathrm{}dp_n}$$ where $`dV`$ is the symplectic volume measure on $`T^{}M.`$ By the Liouville mass of $`\mathrm{\Lambda }^{(j)}(b)`$ we mean the integral $$\omega ^{(j)}(b):=_{\mathrm{\Lambda }^{(j)}(b)}𝑑\omega _b^{(j)}.$$ The Liouville mass of a compact Lagrangean orbit $`\mathrm{\Lambda }^{(j)}(b)`$ has a simple dynamical interpretation: it is the Euclidean volume of the fundamental domain of the common period lattice $`_b^{(j)}=_{(x,\xi )}^{(j)}`$ of points $`(x,\xi )\mathrm{\Lambda }^{(j)}(b)`$ , i.e. (8) $$\omega ^{(j)}(b)=Vol(^n/_b^{(j)}).$$ Indeed, by writing Liouville measure in local action-angle variables, we see that (9) $$d\omega _b^{(j)}=det(T_{\mathrm{}}^k(b))d\mu _b^{(j)},\text{where}T_{\mathrm{}}^k=\frac{I_k}{p_{\mathrm{}}}.$$ It is clear from the definition of the action-angle variables that $`_b^{(j)}`$ is generated by the rows $`(T_1^k,\mathrm{},T_n^k)`$, hence the determinant is the co-volume of the period lattice. We now turn to singular levels. When $`bB_{sing}`$ we first decompose (10) $$𝒫^1(b)=_{j=1}^r\mathrm{\Gamma }_{sing}^{(j)}(b)$$ the singular level into connected components $`\mathrm{\Gamma }_{sing}^{(j)}(b)`$ and then decompose (11) $$\mathrm{\Gamma }_{sing}^{(j)}(b)=_{k=1}^p^n(x_k,\xi _k)$$ each component into orbits. Both decompositions can take a variety of forms. The regular components $`\mathrm{\Gamma }_{sing}^{(j)}(b)`$ must be Lagrangean tori by the properness assumption. A singular components consists of finitely many orbits by the finite complexity assumption. The orbit $`^n(x,\xi )`$ of a singular point is necessarily singular, hence has the form $`^k\times T^m`$ for some $`(k,m)`$ with $`k+m<n.`$ Regular points may also occur on a singular component, whose orbits are Lagrangean and can take any one of the forms $`^k\times T^m`$ for some $`(k,m)`$ with $`k+m=n.`$ We will need the following result in the proof of Theorem (0.1): ###### Proposition 1.3. A singular component $`\mathrm{\Gamma }_{sing}^{(j)}(b)𝒫^1(b)`$ (with $`bB_{sing}`$) must contain a compact singular orbit $`^n(x,\xi )T^k,k<n.`$ Proof: It follows by a standard averaging argument \[M2\] that the set $`_{\mathrm{\Gamma }_{sing}^{(j)}}^I`$ of invariant probability measures supported on $`\mathrm{\Gamma }_{sing}^{(j)}`$ is non-empty: for any probability measure $`\mu _0`$ supported on $`\mathrm{\Gamma }_{sing}^{(j)}`$, the set of weak\* limit points of the set of finite time averages $`\mu _T=\frac{1}{vol\{|t|T\}}_{|t|T}(\mathrm{\Phi }_t)_{}\mu _0𝑑t`$ gives at least one non-trivial element of $`_{\mathrm{\Gamma }_{sing}^{(j)}}^I`$. Since $`\mathrm{\Gamma }_{sing}^{(j)}`$ consists of only finitely many orbits, any invariant measure in $`_{\mathrm{\Gamma }_{sing}^{(j)}}^I`$ is a finite sum of (ergodic) measures, each supported on just one orbit. The non-compact orbits $`^k\times T^m`$ obviously cannot carry invariant probability measures; hence, at least one orbit must be compact. ∎ We will need a further result on Hamiltonian $`^n`$-actions $`\mathrm{\Phi }_t`$. We define a non-zero period of $`\mathrm{\Phi }_t`$ to be a time $`T_b^{(j)}\{0\}`$ for some $`(b,j),`$ and denote the set of periods by $`𝒯`$. ###### Proposition 1.4. There exists a constant $`C>0,`$ which depends on the Riemannian manifold $`(M,g)`$, such that $`inf_{\{T𝒯\}}|T|C.`$ Proof: In the case of a Hamiltonian flow with Hamilton vector field $`\mathrm{\Xi }`$, this is a case of Yorke’s theorem \[Y\]. In fact, $`C=\frac{2\pi }{L}`$ where $`L=d\mathrm{\Xi }_{\mathrm{}}`$. In the case of $`^n`$ actions, we can apply Yorke’s theorem to any one parameter subgroup. ∎ ### 1.2. Hamiltonian torus actions In special cases (see \[D\] for the geometric conditions), the Hamiltonian $`^n`$ action descends to the Hamiltonian action of the torus $`^n/^n`$ on $`T^{}M`$. Such Hamiltonian torus actions are the cotangent space analogues of toric varieties in algebraic geometry. In this case, there exist generators $$:=(I_1,\mathrm{},I_n):T^{}MB^n$$ of of the Hamiltonian $`^n`$ action so that each $`I_j`$ generates a $`2\pi `$-periodic Hamiltonian flow. The components $`I_j`$ are called global action variables and $``$ is called a toric moment map. In the toric case, $`B`$ is a convex polyhedral cone, $`B_{reg}`$ is simply the interior of $`B`$, $`B_{sing}=B`$ (its boundary) and $`m_{cl}(b)1.`$ Since tori are now labelled by actions, we write $`T_I:=^1(I)`$. Singular orbits $`^n(x,\xi )`$ are obviously compact non-Lagrangean tori, and singular levels consist of just one singular orbit. Examples: (i) $`M=^n/^n,I_j=\xi _j`$, the usual linear coordinates on $`T^{}(^n/^n)`$. (ii) $`M=𝕊^2,I_1=p_\theta ,I_2=|\xi |_0`$, where $`p_\theta (x,\xi )=\xi (\frac{}{\theta })`$ (the infinitesimal generator of rotations around the $`z`$-axis), and where $`|\xi |_0`$ is the length function of the standard metric. ### 1.3. Riemannian manifolds with completely integrable geodesic flow Now suppose that $`g`$ is a Riemannian metric on $`M`$ and let $`H(x,\xi )=|\xi |_g`$ denote the associated length function on covectors. The Hamilton flow $`G_t`$ of $`H`$ on $`T^{}M0`$ is homogeneous of degree 1 with respect to the natural $`^+`$ action, and will be referred to as the geodesic flow. It leaves invariant the cosphere bundles $`S^{}M_E=\{H=E\}`$ and the flows $`G_t^E`$ on $`S^{}M_E`$ are all equivalent under dilation $`(x,\xi )E(x,\xi )`$ to $`G_1^t`$. The geodesic flow $`G_t`$ will be called integrable if it commutes with a homogeneous Hamiltonian action of $`^n`$. We may then put $`H=p_1`$. It is called toric integrable if it commutes with a homogeneous Hamiltonian action of $`^n/^n`$. Because $`m_{cl}(b)1`$ in this case, there exists a homogeneous function $`K`$ on $`B`$ such that $`H=K().`$ Examples: The following is a short list of examples: (i) $`M=^n/^n`$ and $`g`$ is flat. Then $`(M,g)`$ is toric integrable. (ii) $`M=𝕊^2`$ and $`g`$ is a rotationally invariant metric. If $`g`$ is of ’simple type’ (e.g. convex), then $`(M,g)`$ is toric integrable \[CV1\]. (iii) $`M=𝕊^2`$ and $`g`$ is the metric for which $`(𝕊^2,g)`$ is an ellipsoid. (iv) $`M=^2/^2`$ and $`g`$ is a Liouville metric (cf. \[B.K.S, KMS\]). (v) Bi-invariant metrics on compact Lie groups. Geodesic flow on $`SO(3)`$ is known as the Euler top. ### 1.4. Manifolds without conjugate points A Riemannian manifold $`(M,g)`$ is said to be without conjugate points if there exists a unique geodesic between each two points of its universal Riemannian cover $`(\stackrel{~}{M},\stackrel{~}{g})`$, or equivalently if every exponential map $`\mathrm{exp}_x:T_xMM`$ is non-singular. We will need the following geometric theorems on manifolds without conjugate points. ###### Theorem 1.5. \[M\] Let $`(M,g)`$ be a compact Riemannian manifold with (co)-geodesic flow $`G^t:T^{}M0T^{}M0`$. Suppose that $`G^t`$ preserves a (non-singular) Lagrangean foliation $``$ of $`T^{}M0`$, i.e. suppose that $`G^tL=L`$ for all leaves $`L`$ of $``$. Then $`(M,g)`$ has no conjugate points. The Hopf conjecture on tori without conjugate points was proved by Burago-Ivanov: ###### Theorem 1.6. \[BI\] Suppose that $`g`$ is a metric on the n-torus $`T^n`$ without conjugate points. Then $`g`$ is flat. ### 1.5. Integrable Newtonian flows on cotangent bundles We will also consider Newtonian flows, i.e. flows of classical Hamiltonians $`H(x,\xi )=\frac{1}{2}|\xi |^2+V(x)`$ on cotangent bundles $`T^{}M`$. Such Hamiltonians and their flows $`G_t`$ are no longer homogeneous. The invariant energy surfaces $`X_E=\{H=E\}`$ and the restricted flows $`G_t^E`$ of $`G_t`$ to $`X_E`$ may change drastically with $`E`$. In particular, it may be completely integrable for some values of $`E`$ and not others. Examples: (i) The spherical pendulum: $`M=𝕊^2`$, $`H=|\xi |^2+\mathrm{cos}\varphi `$; $`|\xi |^2`$ corresponds to the round metric and $`\varphi `$ is the azimuthal angle. (ii) The C. Neumann oscillator on $`T^{}𝕊^n.H=|\xi |^2+_{j=1}^n\alpha _jx_j^2`$ on $`T^{}𝕊^n`$. Here $`0<\alpha _1<\mathrm{}<\alpha _n`$ are constants, $`(x_1,\mathrm{},x_n)`$ are Cartesian coordinates on $`^{n+1}`$ and $`|\xi |^2`$ corresponds to the usual round metric. (iii) The Kowalevsky and Chaplygin tops \[He\]. We note that in the non-homogeneous case, the joint flows $`\mathrm{\Phi }_t^E`$ on each energy level are distinct systems, and may be integrable for only some values of $`E`$. An interesting case is the Chaplygin top \[He\], which is integrable only when the angular momentum integral is put equal to zero. ### 1.6. Rigidity theorems for Newtonian flows We will need a generalization of Mane’s rigidity theorem to Newtonian flows on tori. The following combines some ideas of Bialy-Polterovich \[BP\] and Knauf \[K\] to give a rigidity result when $`M`$ is a torus and $`H`$ is completely integrable with only compact regular orbits. In fact, it is more general: ###### Proposition 1.7. Suppose that $`g`$ is a metric and $`V(x)`$ is a potential on the $`n`$-torus $`𝕋^n`$ such that the Hamiltonian flow $`G_t^E`$ of $`H(x,\xi )`$ on $`X_E`$ preserves a $`C^1`$ Lagrangean foliation by tori which project regularly to $`𝕋^n`$. Then $`E>\mathrm{max}V`$ and $`(EV)g`$ is a flat metric. Proof: By ( \[K\], Theorem 2) no such invariant foliation exists unless $`E>\mathrm{max}V`$, so we may assume this is the case. The Jacobi metric $`(EV)g`$ is then a well-defined metric on $`𝕋^n`$. We denote by $`|\xi |_{J,E}^2`$ the associated homogeneous Hamiltonian (length squared of a covector). Since the sets $`\{H=E\}`$ and $`\{|\xi |_{J,E}^2=1\}`$ are the same, the latter carries a Lagrangean foliation by tori which project regularly to $`𝕋^n.`$ Since the geodesic flow $`G_{J,E}^t`$ of $`(EV)g`$ on $`\{|\xi |_{J,E}^2=1\}`$ coincides (up to a time re-parametrization) with $`G_t^E,`$ this foliation is invariant under $`G_{J,E}^t`$. Now let $`D_r:T^{}M0T^{}M0`$ be the dilation $`D_r(x,\xi )=(x,r\xi ).`$ Then $`D_r:\{|\xi |_{J,E}^2=1\}\{|\xi |_{J,E}^2=r^2\}`$ intertwines the geodesic flows on these sphere bundles (up to constant time reparametrization). Since $`D_r`$ is conformally symplectic it also carries the invariant Lagrangean torus foliation of $`\{|\xi |_{J,E}^2=1\}`$ to an invariant Lagrangean torus foliation of $`\{|\xi |_{J,E}^2=r^2\}`$. It follows that $`T^{}M0`$ carries a Lagrangean torus foliation invariant under the geodesic flow of the Jacobi metric. By Mane’s theorem, the geodesic flow has no conjugate points and so by Burago-Ivanov’s theorem, (E - V)g must be flat. ∎ ###### Corollary 1.8. With the same notation as above, suppose that there exists an interval $`[E_0ϵ,E_0+ϵ]`$ such that, for all $`E[E_0ϵ,E_0+ϵ]`$, $`G_t^E`$ preserves a Lagrangean foliation of by tori which project regularly to $`𝕋^n`$. Then: g is flat and $`V`$ is constant. Proof: The assumption implies that $`(EV)g`$ is flat for all $`E`$ in the interval. Let $`R_E`$ denote the curvature tensor of $`(EV)g.`$ It is clearly a real analytic function of $`E`$. Since $`R_E0`$ in $`[E_0ϵ,E_0+ϵ]`$, it must vanish identically. Therefore the Newton’s flow $`\mathrm{\Phi }_t`$ on $`T^{}T^n`$ has no conjugate points. By Remark 1.C and Theorem 1.B of \[BP\], it follows that $`g`$ is flat and $`V`$ is constant. ∎ ### 1.7. Semiclassical quantum integrable systems: semiclassical calculus We now provide the necessary background on quantum integrable systems. Since we wish to include quantizations of possibly inhomogeneous Hamiltonians, the proper framework is that of semiclassical pseudodifferential operators. First, we introduce symbols. On a given an open $`U^n`$, we say that $`a(x,\xi ;\mathrm{})C^{\mathrm{}}(U\times ^n)`$ is in the symbol class $`S^{m,k}(U\times ^n)`$, provided $$|_x^\alpha _\xi ^\beta a(x,\xi ;\mathrm{})|C_{\alpha \beta }\mathrm{}^m(1+|\xi |)^{k|\beta |}.$$ We say that $`aS_{cl}^{m,k}(U\times ^n)`$ provided there exists an asymptotic expansion: $$a(x,\xi ;\mathrm{})\mathrm{}^m\underset{j=0}{\overset{\mathrm{}}{}}a_j(x,\xi )\mathrm{}^j,$$ with $`a_j(x,\xi )S^{0,kj}(U\times ^n)`$. The associated $`\mathrm{}`$-quantization by $`Op_{\mathrm{}}(a)`$ is defined locally by the standard formula: $$Op_{\mathrm{}}(a)(x,y)=(2\pi \mathrm{})^n_^ne^{i(xy)\xi /\mathrm{}}a(x,\xi ;\mathrm{})𝑑\xi .$$ By using a partition of unity, one constructs a corresponding class, $`Op_{\mathrm{}}(S^{m,k})`$, of properly-supported $`\mathrm{}`$-pseudodifferential operators acting globally on $`C^{\mathrm{}}(M)`$; as is well known, it is independent of the choice of partition of unity. Given $`aS^{m_1,k_1}`$ and $`bS^{m_2,k_2}`$, the composition is given by $`Op_{\mathrm{}}(a)Op_{\mathrm{}}(b)=Op_{\mathrm{}}(c)+𝒪(\mathrm{}^{\mathrm{}})`$ in $`L^2(M)`$ where locally, $$c(x,\xi ;\mathrm{})\mathrm{}^{(m_1+m_2)}\underset{|\alpha |=0}{\overset{\mathrm{}}{}}\frac{(i\mathrm{})^{|\alpha |}}{\alpha !}(_\xi ^\alpha a)(_x^\alpha b).$$ ###### Definition 1.9. We say that the operators $`P_j^{\mathrm{}}Op_{\mathrm{}}(S_{cl}^{m,k});j=1,\mathrm{},n`$, generate a semiclassical quantum completely integrable system on $`M`$ if for each $`\mathrm{}`$, $$\underset{j=1}{\overset{n}{}}P_j^{\mathrm{}}P_j^{\mathrm{}}\text{is jointly elliptic on }T^{}M,$$ $$[P_i^{\mathrm{}},P_j^{\mathrm{}}]=0;1i,jn,$$ and the respective semiclassical principal symbols $`p_1,\mathrm{},p_n`$ generate a classical integrable system on $`T^{}M`$ with $`dp_1dp_2\mathrm{}dp_n0`$ on a dense open subset of $`T^{}M`$. We also assume that the finiteness condition (3) is satisfied. #### 1.7.1. Examples The basic examples we have in mind are where $`P_1^{\mathrm{}}=\mathrm{}^2\mathrm{\Delta }+VOp_{\mathrm{}}(S_{cl}^{0,2})`$ is a Schroedinger operator over a compact manifold $`M`$. Examples include: * Quantum integrable Laplacians $`\mathrm{\Delta }`$ such as Laplacians of Liouville metrics on the sphere or torus \[B.K.S\] \[KMS\], or of the ellipsoid \[T3\]. * Toric integrable Laplacians such as the flat Laplacian on $`𝐓^n`$, or Laplacians for surfaces of revolution of ‘simple type’ (see below and \[CV1\]). * The quantum spherical pendulum $`\mathrm{}^2\mathrm{\Delta }+\mathrm{cos}\varphi `$: $`M=S^2`$, $`\mathrm{\Delta }`$ is the standard Laplacian, $`V=\mathrm{cos}\varphi `$ where $`\varphi `$ is the azimuthal angle. The commuting operator is $`\mathrm{}\frac{}{\theta }`$, the generator of rotations around the $`z`$-axis. * The C. Neumann oscillator on $`𝕊^n`$. Here the quantum Hamiltonian is the Schroedinger operator $`\mathrm{}^2\mathrm{\Delta }+_{j=1}^n\alpha _jx_j^2`$ acting on $`C^{\mathrm{}}(𝕊^n)`$. Here, $`\mathrm{\Delta }`$ is the spherical, constant curvature Laplacian and the potential is the one described above. For the quantized C. Neumann system, one can construct quantum integrals that are all second-order, real-analytic, semiclassical partial differential operators on the sphere \[T3\]. * The quantized Euler, Lagrange and Kowalevsky tops. The Euler and Lagrange cases are classical \[He\], while the quantum Kowalevsky top was shown to be QCI recently by Heckman \[He\]. Here, the integrals are semiclassical differential operators in the enveloping algebra of $`so(3)^3`$ defined as follows: Let $`E_1,E_2,E_3`$ be the standard Pauli basis of $`so(3,)`$ and $`L_1,L_2,L_3`$ be the corresponding left-invariant vector fields defined by: $$L_i(f)(x):=\frac{d}{dt}\{f(x\mathrm{exp}tE_i)\}_{t=0}.$$ Fix a unit vector $`e^3`$ and define the $`C^{\mathrm{}}`$ functions on $`SO(3)`$ by $$Q_i(x):=xe_i,e.$$ Then, the space of operators generated by $`Q_1,Q_2,Q_3,L_1,L_2,L_3`$ can be identified with $`so(3)^3`$. Two of the quantum integrals are the quantized energy Schroedinger operator, $`P_1:=\frac{1}{4}\mathrm{}^2(L_1^2+L_2^2+2L_3^2)Q_1`$ and the quantized momentum operator, $`P_2=\mathrm{}_{j=1}^3Q_jL_j`$. In analogy with the classical case, the third quantum integral is a fourth-order partial differential operator defined as follows: Put $`K:=\mathrm{}^2(L_1+iL_2)^2+4(Q_1+iQ_2)`$. Then, in terms of $`K`$, $`P_3=KK^{}+K^{}K8\mathrm{}^4(L_1^2+L_2^2).`$ Homogeneous quantum completely integrable systems are the special case where $`\mathrm{}`$ occurs with the same power in each term and where the usual homogeneous symbols of the operators are all of order one, e.g. ns $`\mathrm{}\sqrt{\mathrm{\Delta }}`$ or $`\mathrm{}\sqrt{\mathrm{\Delta }+V}`$. In this case, one could remove $`\mathrm{}`$ and use the homogeneous symbolic calculus. However, it is often more convenient to convert homogeneous systems $`P_1,\mathrm{},P_n`$ into semiclassical ones by introducing a semiclassical parameter $`\mathrm{}`$ (with values in some sequence $`\{\mathrm{}_k;k=1,2,3,\mathrm{}\}`$ with $`\mathrm{}_k0`$) and semiclassically scaling the $`P_j`$’s: (12) $$P_j^{\mathrm{}}:=\mathrm{}P_j;j=1,2,\mathrm{},n.$$ When $`P_1=\sqrt{\mathrm{\Delta }},P_2,\mathrm{},P_n`$ are classical pseudodifferential operators of order one, then $`P_j^{\mathrm{}}:=\mathrm{}P_jOp(S_{cl}^{0,1})`$ generate the semiclassical quantum integrable system in the sense of Definition 1.9. ### 1.8. Quantum torus actions (see \[GS\] for many details on this case). Classical torus actions can always be quantized and produce the simplest examples of toric quantum integrable systems. The classical actions $`\{I_j\}`$ can be quantized as commuting pseudodifferential operators $`\widehat{I}_1,\mathrm{},\widehat{I}_n`$ whose joint spectrum $$Sp(\widehat{I}_1,\mathrm{},\widehat{I}_n)=\mathrm{\Lambda }(^n+\nu )B$$ is a lattice (translated by a Maslov index). The simplest case is that of the torus, where $`\widehat{I}_j=\frac{}{\theta _j}`$ (with $`\theta _j`$ denoting the usual angular coordinates). The operators $`\sqrt{\mathrm{\Delta }+1/4},\frac{}{\theta }`$ on $`S^2`$ provide another example. Less obviously, any convex surface of revolution has a toric integrable Laplacian (cf. \[CV1\]). Just as the classical multiplicity $`m_{cl}(b)1`$ in the toric case, so also the multiplicity $`m(\lambda )`$ of the joint eigenvalues is 1 for $`|\lambda |`$ sufficiently large \[CV.1\]. Hence up to a finite dimensional subspace, there is a unique (up to unit scalars) orthonormal basis of joint eigenfunctions $$\widehat{I}_j\varphi _\lambda =\lambda _j\varphi _\lambda ,\lambda =(\lambda _1,\mathrm{},\lambda _n)\mathrm{\Lambda }.$$ ### 1.9. Joint eigenvalue ladders In the next section, we will study the localization of sequences of eigenfunctions on level sets of the moment map. To obtain sequences which localize on a given level $`𝒫^1(b)`$ it is necessary to choose the corresponding joint eigenvalues to tend in an appropriate sense to $`b`$. Roughly speaking, such joint eigenvalues form an ‘eigenvalue ladder’. The term comes from the toric case, where the joint spectrum $`\mathrm{\Lambda }`$ of the action operators is a semi-lattice (i.e. the set of lattice points in a cone) We define ladders (or rays) in a direction $`\lambda `$ by: (13) $$𝐍_\lambda =\{k\lambda +\nu ,k=0,1,2,\mathrm{}\}\mathrm{\Lambda }.$$ In the case of quantizations of torus and other Hamiltonian compact group actions, semiclassical limits are essentially the same as limits along ladders (cf. \[GS\]\[CV1\]). In the $`^n`$ case, there is usually no optimal choice of the generators $`P_j`$, and their joint spectrum is quite far from a lattice. We therefore define a homogeneous ladder of eigenvalues in the direction $`b=(b^{(1)},b^{(2)},\mathrm{},b^{(n)})^n`$ to be a sequence satisfying (14) $$\{\lambda _k:=(\lambda _k^{(1)},\mathrm{},\lambda _k^{(n)})Spec(P_1,\mathrm{},P_n);j=1,..,n,\underset{k\mathrm{}}{lim}\frac{\lambda _k}{|\lambda _k|}=b\},$$ where $`|\lambda _k|:=\sqrt{|\lambda _k^{(1)}|^{\mathrm{\hspace{0.17em}2}}+\mathrm{}+|\lambda _k^{(n)}|^{\mathrm{\hspace{0.17em}2}}}`$. Finally, we introduce a notion of semiclassical ladders: We fix $`0<\delta <1`$, $`b=(b^{(1)},b^{(2)},\mathrm{},b^{(n)})^n`$, and define the set (15) $$𝐋_{b;\delta }(\mathrm{}):=\{b_j(\mathrm{}):=(b_j^{(1)}(\mathrm{}),b_j^{(2)}(\mathrm{}),\mathrm{},b_j^{(n)}(\mathrm{}))\text{Spec}(P_1,\mathrm{},P_n);|b_j(\mathrm{})b|C\mathrm{}^{1\delta }\}.$$ Here, $`b_j^{(1)}(\mathrm{})=E_j(\mathrm{})`$. Taking a sequence $`\mathrm{}0`$, the joint eigenvalues in $`𝐋_{b;\delta }(\mathrm{})`$ form a sequence tending to $`b`$ which is the analogue of a homogeneous ladder. ## 2. Localization on tori One of the main inputs in the proof of the Theorem is the localization of a ladder of joint eigenfunctions of a quantum completely integrable system in a regular direction $`bB_{reg}`$ on the level set $`𝒫^1(b)`$ of the moment map. In this section, we prove the relevant localization results. We first consider toric systems, where level sets are regular and connected and eigenfunctions necessarily localize on individual tori. In the general $`^n`$ case, ladders of eigenfunctions localize on the possibly disconnected level set $`𝒫^1(b)`$, and it is a complicated problem to determine how the limit eigenfunction mass (or ‘charge’) is distributed among the components. To deal with this problem, we define a notion of the charge of a component, and prove that every compact component of $`𝒫^1(b)`$ is charged by some sequence of eigenfunctions. This result will play an important role in the proof of the Theorem. ### 2.1. Toric integrable systems Let $`A\mathrm{\Psi }^o(M)`$ denote any zeroth order pseudodifferential operator and $`d\mu _\lambda `$ denote Lebesgue measure on the Lagrangian torus $`T_\lambda `$. In the toric case we have the following localization theorem: ###### Proposition 2.1. \[Z1\] For any ladder $`\{k\lambda +\nu :k=0,1,2,\mathrm{}\}`$ of joint eigenvalues, we have: $$(A\varphi _{k\lambda },\varphi _{k\lambda })=_{T_\lambda }\sigma _A𝑑\mu _\lambda +O(k^1).$$ We thus have: ###### Corollary 2.2. For any invariant torus $`T_\lambda S^{}M`$, there exists a ladder $`\{\varphi _{k\lambda },k=0,1,2,\mathrm{}\}`$ of eigenfunctions localizing on $`T_\lambda .`$ ### 2.2. $`^n`$-integrable systems The proper generalization of the toric localization result Proposition (2.1) to $`^n`$ actions says that ladders of joint eigenfunctions localize on level sets of the moment map rather than on individual tori. This result is more or less a folk theorem in the physics literature (see \[E, Be, Be2\]), and the rigorous result is in principle known to experts. However, we were unable to find the result in the literature, so we sketch the proof here. It uses some material on quantum Birkhoff normal forms from \[CV2\]. Let $`b`$ be a regular value of the moment map $`𝒫`$, let $$𝒫^1(b)=\mathrm{\Lambda }^{(1)}(b)\mathrm{}\mathrm{\Lambda }^{(m_{cl})}(b),$$ where the $`\mathrm{\Lambda }^{(l)}(b);l=1,\mathrm{},m`$ are $`n`$-dimensional Lagrangian tori, and $`d\mu _{\mathrm{\Lambda }^{(j)}(b)}`$ denote the normalize Lesbegue measure on the torus $`\mathrm{\Lambda }^{(j)}(b)`$. Let $`b_j(\mathrm{})𝐋_{b,\delta }(\mathrm{})`$ and define (16) $$c_l(\mathrm{};b_j(\mathrm{})):=Op_{\mathrm{}}(\chi _l)\varphi _{b_j(\mathrm{})},\varphi _{b_j(\mathrm{})};l=1,\mathrm{},m_{cl}(b).$$ We recall that $`\chi _l`$ is cutoff function which is equal to 1 in the neighbourhood $`\mathrm{\Omega }^{(l)}(b)`$ of the torus $`\mathrm{\Lambda }^{(l)}(b)`$ and vanishes on $`_{kl}\mathrm{\Omega }^{(k)}(b)`$. ###### Proposition 2.3. Let $`bB_{reg}`$, and let $`\{\varphi _{b_j(\mathrm{})}\}`$ be a sequence of $`L^2`$-normalizeed joint eigenfunctions of $`P_1,\mathrm{},P_n`$ with joint eigenvalues in the ladder $`𝐋_{b,\delta }(\mathrm{})`$ of (15). Then, for any $`aS^{0,\mathrm{}}`$, we have that as $`\mathrm{}0`$: $$Op_{\mathrm{}}(a)\varphi _{b_j(\mathrm{})},\varphi _{b_j(\mathrm{})}=\underset{l=1}{\overset{m}{}}c_l(\mathrm{};b_j(\mathrm{}))_{\mathrm{\Lambda }^{(j)}(b)}a𝑑\mu _{\mathrm{\Lambda }^{(j)}(b)}+𝒪(\mathrm{}^{1\delta }).$$ Here, $`d\mu _{\mathrm{\Lambda }^{(j)}(b)}`$ denotes Lebesgue measure on $`\mathrm{\Lambda }^{(j)}(b)`$. Proof: Let $`^{(l)}`$ be the pullback of the Maslov line bundle over $`\mathrm{\Lambda }^{(l)}`$ to the affine torus given by $`I_1^{(l)}=\mathrm{}=I_n^{(l)}=0`$ and $`\mathrm{\Omega }^{(l)}`$ be a sufficiently small neighbourhoodof $`\mathrm{\Lambda }^{(l)}`$ on which there exist action-angle variables $`(\theta ^{(l)},I^{(l)})`$. According to the quantum Birkhoff normal form (QBNF) construction \[CV2\], for $`l=1,\mathrm{},k`$ and $`j=1,\mathrm{},n`$, there exist $`\mathrm{}`$-Fourier integral operators, $`U_{b,\mathrm{}}^{(l)}:C^{\mathrm{}}(M)C^{\mathrm{}}(𝕋^n;^{(l)})`$, microlocally elliptic on $`\mathrm{\Omega }^{(l)}`$, together with $`C^{\mathrm{}}`$ symbols, $`f_j^{(l)}(x;\mathrm{})_{k=0}^{\mathrm{}}f_{jk}^{(l)}(x)\mathrm{}^k`$, with $`f_{j0}(0)=0`$ such that: (17) $$U_{b,\mathrm{}}^{(l)}f_j^{(l)}(P_1b^{(1)},\mathrm{},P_nb^{(n)};\mathrm{})U_{b,\mathrm{}}^{(l)}=_{\mathrm{\Omega }_0^{(l)}}\frac{\mathrm{}}{i}\frac{}{\theta _j}.$$ Moreover, when $`P_1,\mathrm{},P_n`$ are self-adjoint, the operator $`U_b^{(l)}`$ can be taken to be microlocally unitary. We now observe that the space of admissible \[CP\] solutions of the microlocal eigenfunction equation (18) $$P_k\varphi _{b_j(\mathrm{})}=_{\mathrm{\Omega }^{(l)}(b)}b_j^{(k)}(\mathrm{})\varphi _{b_j(\mathrm{})}$$ is one-dimensional. Indeed, such solutions are the same as solutions of $$f_k^{(l)}(P_1b^{(1)},\mathrm{},P_nb^{(n)};\mathrm{})\varphi _j=_{\mathrm{\Omega }^{(l)}(b)}f_k^{(l)}(b_j^{(1)}b^{(1)},\mathrm{},b_j^{(n)}b^{(n)};\mathrm{})\varphi _j.$$ We conjugate this equation to Birkhoff normal form (17) and use the fact that the microlocal solutions of the model equation $$\frac{\mathrm{}}{i}\frac{}{\theta }u_j=m_ju_j$$ are just multiples of $`\mathrm{exp}[i(n+\pi \gamma /4)\theta ]`$, where $`\gamma `$ is the Maslov index and $`n`$. Thus, the joint eigenfunctions $`\varphi _{b_j(\mathrm{})}`$ are given microlocally by (19) $$\varphi _{b_j(\mathrm{})}=_{\mathrm{\Omega }^{(l)}(b)}\sqrt{c_l(\mathrm{};b_j(\mathrm{}))}U_{b;\mathrm{}}^{(l)}(e^{i(n_j+\pi \gamma /4)\theta }).$$ The right sides of (19) are the usual quasimodes or semiclassical Lagrangian distributions \[CV2\] Now let $`\chi _l(x,\xi )C_0^{\mathrm{}}(T^{}M);l=1,\mathrm{}.,m_{cl}(b)`$ be a cutoff function which is identically equal to one on the neighbourhood $`\mathrm{\Omega }^{(l)}(b)`$ and vanishes on $`\mathrm{\Omega }^{(k)}(b)`$ for $`kl`$. For $`\mathrm{}`$ sufficiently small, we then have $$Op_{\mathrm{}}(a)\varphi _{b_j(\mathrm{})},\varphi _{b_j(\mathrm{})}=\underset{l=1}{\overset{m_{cl}(b)}{}}Op_{\mathrm{}}(a)Op_{\mathrm{}}(\chi _l)\varphi _{b_j(\mathrm{})},\varphi _{b_j(\mathrm{})}+𝒪(\mathrm{}^{\mathrm{}}).$$ It follows by (19), the semiclassical Egorov theorem and a Taylor expansion about the Lagrangian torus $`I^{(l)}=0`$ that: (20) $$\begin{array}{ccc}Op_{\mathrm{}}(a)Op_{\mathrm{}}(\chi _l)\varphi _{b_j(\mathrm{})},\varphi _{b_j(\mathrm{})}\hfill & =\hfill & c_l(\mathrm{};b_j(\mathrm{}))Op_{\mathrm{}}(a)Op_{\mathrm{}}(\chi _l)U_{b;\mathrm{}}^{(l)}(e^{i(n_j+\pi \gamma /4)\theta }),U_{b;\mathrm{}}^{(l)}(e^{i(n_j+\pi \gamma /4)\theta })\hfill \\ & & \\ & =\hfill & c_l(\mathrm{};b_j(\mathrm{}))U_{b;\mathrm{}}^{(l)}Op_{\mathrm{}}(a)Op_{\mathrm{}}(\chi _l)U_{b;\mathrm{}}^{(l)}e^{i(n_j+\pi \gamma /4)\theta },e^{i(n_j+\pi \gamma /4)\theta }\hfill \\ & & \\ & =\hfill & (2\pi )^nc_l(\mathrm{};b_j(\mathrm{}))\left(_{\mathrm{\Lambda }^{(l)}}a𝑑\mu _l+e(\mathrm{})\right)+𝒪(\mathrm{}),\hfill \end{array}$$ where $`e(\mathrm{})=Op_{\mathrm{}}(r)u_{\mathrm{}},u_{\mathrm{}}`$ for some function $`rC_0^{\mathrm{}}(𝕋^n\times D_1)`$ satisfying $`r(\theta ,I)=𝒪(|I|)`$ (recall, we have normalized the action variables so that $`I^{(l)}=0`$ on the torus $`\mathrm{\Lambda }^{(l)}(b)`$). Here, $`u_{\mathrm{}}(\theta )=\mathrm{exp}[i(m_1\theta _1+\mathrm{}+m_n\theta _n)]`$ with $`m_j(\mathrm{})=𝒪(\mathrm{}^{1\delta })`$. An integration by parts in the $`I_1,\mathrm{},I_n`$ variables shows that: $$(Op_{\mathrm{}}(r)u_{\mathrm{}},u_{\mathrm{}})=𝒪(\mathrm{}^{1\delta }),$$ and the proposition follows.∎ ### 2.3. Charge of compact Lagrangean orbits We now investigate the coefficients $`c_j(\mathrm{})`$ in Proposition (2.3) for ‘ladders’ of eigenfunctions. Our purpose is to show that there exist ladders for which the limit as $`\mathrm{}0`$ of $`c_j(\mathrm{})`$ is bounded below by a positive geometric constant. It is convenient at this point to introduce the language of quantum limits. #### 2.3.1. Quantum limits Let $`(P_1,\mathrm{},P_n)`$ denote a quantum integrable system, with classical integrable flow $`\mathrm{\Phi }_t`$. Fix $`E`$ and let $`M_I^E`$ denote the set of invariant probability measures for $`\mathrm{\Phi }_t^E`$ on $`X_E`$. For instance, $`M_I^E`$ includes the orbital averaging measures $`\mu _z`$, defined by $$_{X_E}f𝑑\mu _z=\underset{T\mathrm{}}{lim}\frac{1}{T^n}_{\mathrm{max}|t_j|T}f(\mathrm{\Phi }_t(z))𝑑t.$$ In the case of compact (torus) orbits, $`\mu _z`$ is the Lebesgue probability measure on the orbit of $`z`$. By the set $`𝒬_E`$ of ‘quantum limit’ measures of the quantum integrable system at energy level $`E`$, we mean the set of weak\* limits (as $`\mathrm{}0`$) of the measures $`d\mathrm{\Phi }_{b_j(\mathrm{})}`$ defined by (21) $$Op_{\mathrm{}}(a)\varphi _{b_j(\mathrm{})},\varphi _{b_j(\mathrm{})}=_{X_E}a𝑑\mathrm{\Phi }_{b_j(\mathrm{})},(b_j^{(1)}E).$$ We write $`d\mathrm{\Phi }_{b_j(\mathrm{})}d\mu 𝒬_E`$ for weak\* convergence to the limit as $`\mathrm{}0`$. It is an easy consequence of the semiclassical Egorov theorem that $`𝒬_EM_I^E.`$ When $`d\mu `$ equals Lebesgue probability measure on an orbit, we say that the sequence $`\{\varphi _{b_j(\mathrm{})}\}`$ localizes on the orbit. For background, terminology and references in a closely related context, we refer to \[JZ\]. We now consider quantum limits of eigenfunctions corresponding to a ladder of joint eigenvalues. Put: (22) $$𝐕_{b,\delta }(\mathrm{})=\{\varphi _{b_j(\mathrm{})}:b_j(\mathrm{})𝐋_{b;\delta }(\mathrm{})\}$$ There are many possible weak\* limit points of the set $`_{\mathrm{}[0,\mathrm{}_0]}𝐕_{b,\delta }(\mathrm{}).`$ We say: ###### Definition 2.4. For $`bB_{reg}`$, a ladder of eigenfunctions is a sequence $`_b:=\{\varphi _{b_j}(\mathrm{})\}`$ of joint eigenfunctions with the following properties: * $`b_j(\mathrm{})𝐋_{b,\delta }(\mathrm{})`$ as $`\mathrm{}0`$ forms an eigenvalue ladder; * $`d\mathrm{\Phi }_{b_j}(\mathrm{})`$ has a unique weak limit $`d\mathrm{\Phi }__b`$ as $`\mathrm{}0.`$ For a ladder of eigenfunctions, $`lim_\mathrm{}0c_{\mathrm{}}(\mathrm{};b_j(\mathrm{}))`$ exists for each $`\mathrm{}`$ in Proposition (2.3). ###### Definition 2.5. Given $`bB_{reg}`$, we say that the ladder $`_b=\{\varphi _{b_j(\mathrm{})}\}`$ gives charge $`c_l(_b):=lim_\mathrm{}0c_{\mathrm{}}(\mathrm{};b_j(\mathrm{}))`$ to the component torus $`\mathrm{\Lambda }^{(l)}(b)`$, and that it charges $`\mathrm{\Lambda }^{(l)}(b)`$ if $`c_l(_b)>0.`$ The limit in Definition (2.5) above clearly depends on the ladder $`_b`$. For instance, there could be sequences of joint eigenfunctions localizing on each single component of $`𝒫^1(b).`$ To obtain an invariant of the Lagrangean orbits which is independent of the ladder, we say: ###### Definition 2.6. The charge $`c(\mathrm{\Lambda }^{(l)}(b))`$ of a component torus $`\mathrm{\Lambda }^{(l)}(b)𝒫^1(b)`$ is defined by by the formula: $$\begin{array}{ccc}c(\mathrm{\Lambda }^{(l)}(b))\hfill & =\hfill & sup__bc_l(_b)\hfill \end{array}$$ where $`c_l`$ is the coefficient in the sum of Proposition (2.3). A useful formula for the charge is: ###### Proposition 2.7. $`c(\mathrm{\Lambda }^{(l)}(b))=lim\; sup_\mathrm{}0\mathrm{max}_{\varphi _{b_j(\mathrm{})}V_\delta (\mathrm{})}Op_{\mathrm{}}(\chi _l)\varphi _{b_j(\mathrm{})},\varphi _{b_j(\mathrm{})}.`$ Proof: (i) $``$: By definition, $`c_l(\mathrm{};b_j(\mathrm{}))=Op_{\mathrm{}}(\chi _l)\varphi _{b_j(\mathrm{})},\varphi _{b_j(\mathrm{})}`$ where $`\chi _l`$ is a cutoff to $`\mathrm{\Omega }_l.`$ Since $`V_{b,\delta }(\mathrm{})`$ is a finite set for each $`\mathrm{}`$, there exists $`\varphi _{b_j(\mathrm{})}^{\mathrm{max}}V_{b,\delta }(\mathrm{})`$ such that $`Op_{\mathrm{}}(\chi _l)\varphi _{b_j(\mathrm{})}^{\mathrm{max}},\varphi _{b_j(\mathrm{})}^{\mathrm{max}}=\mathrm{max}_{\varphi _{b_j(\mathrm{})}V_\delta (\mathrm{})}Op_{\mathrm{}}(\chi _l)\varphi _{b_j(\mathrm{})},\varphi _{b_j(\mathrm{})}.`$ We form the sequence $`\{\varphi _{b_j(\mathrm{})}^{\mathrm{max}}\}_{\mathrm{}\{\mathrm{}_k\}}`$ and then choose a sub-ladder $`_b^{\mathrm{max}}`$ with a unique quantum limit. Then $$\begin{array}{c}c(\mathrm{\Lambda }^{(l)}(b))lim_\mathrm{}0Op_{\mathrm{}}(\chi _l)\varphi _{b_j(\mathrm{})}^{max},\varphi _{b_j(\mathrm{})}^{max}\hfill \\ \\ lim\; sup_\mathrm{}0\mathrm{max}_{\varphi _{b_j(\mathrm{})}V_\delta (\mathrm{})}Op_{\mathrm{}}(\chi _l)\varphi _{b_j(\mathrm{})},\varphi _{b_j(\mathrm{})}.\hfill \end{array}$$ (ii) $``$: It is clear that for each ladder $`_b`$ we have $$c_l(_b)\underset{\mathrm{}0}{lim\; sup}\underset{\varphi _{b_j(\mathrm{})}V_\delta (\mathrm{})}{\mathrm{max}}Op_{\mathrm{}}(\chi _l)\varphi _{b_j(\mathrm{})},\varphi _{b_j(\mathrm{})}.$$ Therefore the same holds after taking the supremum over $`_b.`$ The following lemma is the main result of this section: ###### Lemma 2.8. Let $`\omega ^{(l)}(b)`$ denote Liouville measure of the Lagrangian torus $`\mathrm{\Lambda }^{(l)}(b);l=1,\mathrm{},m_{cl}(b)`$. Then, for all $`(b,l)B_{reg}\times \{1,\mathrm{},m_{cl}(b)\}`$ we have that $$c(\mathrm{\Lambda }^{(l)}(b))\frac{\omega ^{(l)}(b)}{_{j=1}^{m_{cl}(b)}\omega ^{(j)}(b)}.$$ Proof: Fix $`\zeta 𝒮(^n)`$ with $`\zeta 0`$, $`\stackrel{ˇ}{\zeta }C_0^{\mathrm{}}(^n)`$ and $`\stackrel{ˇ}{\zeta }(0)=1`$. Assume moreover that $`0^n`$ is the only point of intersection of supp $`\zeta `$ with the joint periods of the joint flow $`\mathrm{\Phi }_t`$. Let $`K`$ be a fixed compact neighbourhood of $`b=(b^{(1)},\mathrm{},b^{(n)})`$ and $`aS^{0,\mathrm{}}`$. Consider the localized semiclassical trace: (23) $$Tr_a(\zeta ):=\underset{b_j(\mathrm{})K}{}Op_{\mathrm{}}(a)\varphi _{b_j(\mathrm{})},\varphi _{b_j(\mathrm{})}\zeta \left(\frac{b_j(\mathrm{})b}{\mathrm{}}\right).$$ The localized semiclassical trace formula for commuting operators \[Ch\] implies that for any $`aS^{0,\mathrm{}}`$ and $`\zeta 𝒮(^n)`$ as above, (24) $$Tr_a(\zeta )=(2\pi )^n_{𝒫^1(b)}a𝑑\omega ^{(l)}(b)+𝒪(\mathrm{}).$$ So, in particular putting $`a(x,\xi )=\chi _l(x,\xi )`$, we have that: (25) $$Tr_{\chi _l}(\zeta )=(2\pi )^n_{\mathrm{\Lambda }^{(l)}(b)}\chi _l𝑑\omega ^{(l)}(b)+𝒪(\mathrm{})=(2\pi )^n\omega ^{(l)}(b)+𝒪(\mathrm{}),$$ since $`\chi _l=1`$ on the torus, $`\mathrm{\Lambda }^{(l)}(b)`$. On the other hand, since $`\zeta 𝒮(^n)`$, it follows that (26) $$Tr_{\chi _l}(\zeta )=\underset{\{b_j(\mathrm{})𝐋_{b,\delta }(\mathrm{})\}}{}Op_{\mathrm{}}(\chi _l)\varphi _{b_j(\mathrm{})},\varphi _{b_j(\mathrm{})}\zeta \left(\frac{b_j(\mathrm{})b}{\mathrm{}}\right)+𝒪(\mathrm{}^{\mathrm{}}).$$ Thus, by the definition (2.6) of the charge $`c(\mathrm{\Lambda }^{(l)}(b)`$ and the fact that $`\zeta 0`$, we have that: (27) $$|Tr_{\chi _l}(\zeta )|(2\pi )^n\left(\underset{\{b_j(\mathrm{})𝐋_{b,\delta }(\mathrm{})\}}{\mathrm{max}}Op_{\mathrm{}}(\chi _l)\varphi _{b_j(\mathrm{})},\varphi _{b_j(\mathrm{})}\right)\underset{\{b_j(\mathrm{})𝐋_{b,\delta }(\mathrm{})\}}{}\zeta \left(\frac{b_j(\mathrm{})b}{\mathrm{}}\right)+𝒪(\mathrm{}^{\mathrm{}}).$$ Next, by applying the trace formula once again we get that: (28) $$\underset{\{b_j(\mathrm{})𝐋_{b,\delta }(\mathrm{})\}}{}\zeta \left(\frac{b_j(\mathrm{})b}{\mathrm{}}\right)=(2\pi )^n\underset{j=1}{\overset{m_{cl}(b)}{}}\omega ^{(j)}(b)+𝒪(\mathrm{}).$$ Substituting (28) in (27) yields the estimate (29) $$|Tr_{\chi _l}(\zeta )|(2\pi )^n\underset{\{b_j(\mathrm{})𝐋_{b,\delta }(\mathrm{})\}}{\mathrm{max}}Op_{\mathrm{}}(\chi _l)\varphi _{b_j(\mathrm{})},\varphi _{b_j(\mathrm{})}\left(\underset{j=1}{\overset{m_{cl}(b)}{}}\omega ^{(j)}(b)\right)+𝒪(\mathrm{}).$$ The lemma then follows by combining (29) and (25) and letting $`\mathrm{}0`$. ∎ This yields a generalization of Corollary (2.2): ###### Corollary 2.9. For any $`bB_{reg}`$, and for any $`1\mathrm{}m_{cl}(b)`$, there exists a ladder $`_b^{(\mathrm{})}=\{\varphi _{b_j(\mathrm{})}\}`$ such that $`c_{\mathrm{}}(_b^{(\mathrm{})})\frac{\omega ^{(\mathrm{})}(b)}{_{j=1}^{m_{cl}(b)}\omega ^{(j)}(b)};`$ Thus, every regular torus orbit is charged by some ladder. This follows from Lemma ( 2.8), Proposition ( 2.7) and Proposition ( 2.3). #### 2.3.2. Charge of compact singular orbits Our next step is to prove that some compact singular orbits are also charged. To be precise, we have so far only defined the notion of charge for regular levels of the moment map (Definition (2.5). The analogous defintion in the case of a singular value $`b_sB_s`$ is as follows. Let $`𝒫^1(b_s)=_{j=1}^r\mathrm{\Gamma }_{sing}^{(j)}(b_s)`$ be the decomposition (10) into connected components. ###### Definition 2.10. When $`b_sB_{sing}`$, we define an eigenfunction ladder $`_{b_s}`$ to be a sequence of joint eigenfunctions with joint eigenvalues satisfying $`b_j(\mathrm{})b_s=o(1)`$ as $`\mathrm{}0`$ and with unique limit measure $`d\mathrm{\Phi }_{_{b_s}}`$. We say that $`_{b_s}`$ gives charge $`_{\mathrm{\Gamma }_{sing}^{(j)}(b_s)}𝑑\mathrm{\Phi }_{_{b_s}}`$ to the component $`\mathrm{\Gamma }_{sing}^{(j)}`$. Similarly, we say that it gives charge $`_{\mathrm{\Lambda }_{sing}^{(j)}(b_s)}𝑑\mathrm{\Phi }_{_{b_s}}`$ to any orbit $`\mathrm{\Lambda }_{sing}^{(j)}(b_s)`$ on $`\mathrm{\Gamma }_{sing}^{(j)}(b_s)`$ (see (11)). Finally, the charge $`c(\mathrm{\Gamma }^{(j)}(b_s)),`$ resp. $`c(\mathrm{\Lambda }^{(j)}(b_s))`$ of a component, resp. an orbit on the component, is the supremum of the same over all ladders $`_{b_s}`$ We then have: ###### Lemma 2.11. Let $`b_sB_{sing}`$, and let $`\{\mathrm{\Gamma }_{sing}^{(j)}(b)\}`$ denote the singular components of $`𝒫^1(b_s)`$. Then, there exists $`j`$ such that $`c(\mathrm{\Gamma }_{sing}^{(j)}(b))>0`$. Further, there exists a compact singular orbit $`\mathrm{\Lambda }^{(j)}(b_s)\mathrm{\Gamma }_{sing}^{(j)}(b)`$ such that $`c(\mathrm{\Lambda }^{(j)}(b_s))>0.`$ Proof: Let $`U_{sing}`$ be a $`\mathrm{\Phi }_t`$-invariant neighbourhood of $`_{j=1}^r\mathrm{\Gamma }_{sing}^{(j)}(b)`$. Let $`\{b_n\}B_{reg}`$ be a sequence of regular points such that $`b_nb_s`$. For each $`j`$ and sufficiently large $`n`$, there exists at least one component $`\mathrm{\Lambda }^{(\mathrm{})}(b_n)`$ of $`𝒫^1(b_n)`$ such that $`\mathrm{\Lambda }^{(\mathrm{})}(b_n)U_{sing}`$. By Lemma (2.8), $`\mathrm{\Lambda }^{(\mathrm{})}(b_n)`$ is charged by an amount $`\frac{\omega ^{(\mathrm{})}(b_n)}{_{j=1}^{m_{cl}(b_n)}\omega ^{(j)}(b_n)}`$. We now break up the discussion into two cases: Case 1: All $`^n`$-orbits of $`_{j=1}^r\mathrm{\Gamma }_{sing}^{(j)}(b)`$ are compact In this case, we just need a positive lower bound for the quotient $`\frac{\omega ^{(\mathrm{})}(b_n)}{_{j=1}^{m_{cl}(b_n)}\omega ^{(j)}(b_n)}`$ as $`n\mathrm{}.`$ A lower bound for the numerator is given by the minimal period of Yorke’s theorem (Proposition 1.4). Since all orbits (including the limit) are compact, the masses in the denominator have uniform upper bounds. Indeed, by (8) the masses are the co-volumes of the period lattices of $`\mathrm{\Lambda }^{(\mathrm{})}(b).`$ Since the period vectors generating the lattices are uniformly bounded as $`n\mathrm{}`$, the volumes are also uniformly bounded above. Hence the denominator is bounded above, and therefore the quotient is bounded below by a positive constant. ∎ Case 2: There exists a non-compact orbit in $`_{j=1}^r\mathrm{\Gamma }_{sing}^{(j)}(b)`$ In this case, the denominator will tend to infinity, so we need a better lower bound on the numerator. We claim that there exists $`\mathrm{}`$ such that $`\mathrm{\Lambda }^{(\mathrm{})}(b_n)U_{sing}`$ and $`c(\mathrm{\Lambda }^{(\mathrm{})}(b_n))\frac{1}{m_{cl}(b_n)}`$. To prove this, it suffices to find $`\mathrm{}`$ such that (30) $$\frac{\omega ^{(\mathrm{})}(b_n)}{_{j:\mathrm{\Lambda }^{(j)}(b_n)U_{sing}}\omega ^{(j)}(b_n)}\frac{1}{\mathrm{\#}\{j:\mathrm{\Lambda }^{(j)}(b_n)U_{sing}\}}.$$ The natural candidate is to choose $`\mathrm{}`$ such that (31) $$\omega ^{(\mathrm{})}(b_n)=\underset{\{j:\mathrm{\Lambda }^{(j)}(b_n)U_{sing}\}}{\mathrm{max}}\omega ^{(j)}(b_n).$$ We now prove that this choice of $`\mathrm{}`$ satisfies (30). We write $$\underset{j=1}{\overset{m_{cl}(b_n)}{}}\omega ^{(j)}(b_n)=\underset{j:\mathrm{\Lambda }^{(j)}(b_n)U_{sing}}{}\omega ^{(j)}(b_n)+\underset{j:\mathrm{\Lambda }^{(j)}(b_n)U_{sing}=\mathrm{}}{}\omega ^{(j)}(b_n).$$ The second term is bounded above by a constant $`C`$ independent of $`n`$. The first term tends to infinity since $`_{j=1}^r\mathrm{\Gamma }_{sing}^{(j)}(b)`$ contains a non-compact orbit. Indeed, at least one vector of the period lattice of $`\mathrm{\Lambda }^{(j)}(b_n)`$ must tend to infinity as $`n\mathrm{}`$ since the limit orbit is non-compact. It follows that the set of period lattices $`_b^{(j)}`$ is non-compact in the manifold of lattices of full rank of $`^n`$. Now according to Mahler’s theorem, any set $$\{\mathrm{\Gamma }^n|\gamma C,(\gamma \mathrm{\Gamma }\{0\}),\text{and}Vol(^n/\mathrm{\Gamma })K\}$$ is compact. By Yorke’s theorem (loc. cit.), the minimal period stays bounded below, so non-compactness of the lattices forces some volume $`\omega ^{(\mathrm{})}(b_n)\mathrm{}`$ as $`n\mathrm{}.`$ It follows that when a non-compact orbit exists in $`𝒫^1(b_s)`$, then for each $`\mathrm{}`$, $$\frac{\omega ^{(\mathrm{})}(b_n)}{_{j=1}^{m_{cl}(b_n)}\omega ^{(j)}(b_n)}=\frac{\omega ^{(\mathrm{})}(b_n)}{_{j:\mathrm{\Lambda }^{(j)}(b_n)U_{sing}}\omega ^{(j)}(b_n)}+o(1)\text{as}n\mathrm{}.$$ Then (30) follows if we select $`\mathrm{}`$ as in (31). ∎ We now complete the proof of Lemma (2.11). By the finite complexity condition, we have found $`\mathrm{\Lambda }^{(j_n)}(b_n)U_{sing}`$ such that $`c(\mathrm{\Lambda }^{(j_n)}(b_n))c:=\frac{1}{M}>0`$. Further, for each $`n`$, there exists a ladder $`_{b_n}`$ which gives charge $`c`$ to $`\mathrm{\Lambda }^{(j_n)}(b_n)𝒫^1U_{sing}`$. Let $`d\mathrm{\Phi }_{_{b_n}}`$ denote the unique weak limit measure of the ladder. Then let $`\nu `$ denote any weak\* limit of the sequence $`\{d\mathrm{\Phi }_{_{b_n}}\}`$. It follows that $`\nu `$ is an invariant probability measure supported on $`_{j=1}^r\mathrm{\Gamma }_{b_s}^{(j)}.`$ Indeed, its support must be contained in the set of limit points of the sequence of orbits $`\{\mathrm{\Lambda }^{(j_n)}(b_n)\}`$, hence in $`𝒫^1(b_s)U_{sing}.`$ Since $`𝒬`$ is closed in the weak\* topology (since it is a set of limit points), it follows further that $`\nu 𝒬.`$ Hence there exists a ladder $`_{b_s}`$ such that $`\mathrm{\Phi }_{_{b_s}}\nu ,`$ and which charges $`_{j=1}^r\mathrm{\Gamma }_{b_s}^{(j)}`$ by an amount $`c>0.`$ This proves the first part of the lemma. The second statement is an immediate consequence of Proposition(1.3): There must exist at least one compact singular orbit $`\mathrm{\Lambda }_{b_s}^{(j)}_{j=1}^r\mathrm{\Gamma }_{b_s}^{(j)}.`$ Since $`\nu `$ is an invariant probabililty measure, it must be supported on union of the compact singular orbits, hence must charge at least one such orbit. ∎ ## 3. Proof of the Theorem We break up the proofs into three steps. Step 1 is to show that the uniform boundedness assumption implies that all regular tori project without singularities to the base. Step 2 is to show that there are no singular tori. Step 3 is a geometric argument showing that any completely integrable system with no singular tori and with all tori projecting regularly to the base is flat. ### 3.1. Step 1: regular tori project regularly We first consider the simplest case of toric systems: #### 3.1.1. Toric integrable systems ###### Proposition 3.1. Suppose that $`(M,g)`$ is toric integrable and that $`L^{\mathrm{}}(E,g)=O(1).`$ Then every orbit of the torus action has a non-singular projection to $`M`$. In particular, the orbit foliation is a non-singular Lagrangean foliation. Proof: The assumption implies that the joint eigenfunctions $`\{\varphi _\lambda \}`$ of the quantum torus action have uniformly bounded sup-norms. By Proposition (2.1), for every invariant torus $`T_\lambda `$, there exists a ladder $`\{k\lambda ,k=1,2,\mathrm{}\}`$ of joint eigenvalues such that for all $`VC^{\mathrm{}}(M)`$ we have $$\underset{k\mathrm{}}{lim}_MV(x)|\varphi _{k\lambda }(x)|^2𝑑vol=_MV\pi _\lambda 𝑑\mu _\lambda .$$ If we have $`\varphi _{k\lambda }_{\mathrm{}}C`$ for all $`(k,\lambda )`$, then $$|_MV(x)|\varphi _{k\lambda }(x)|^2𝑑vol|CV_{L^1}(k)$$ and hence $$\underset{k\mathrm{}}{lim}|_MV(x)|\varphi _{k\lambda }(x)|^2𝑑vol|CV_{L^1}.$$ Therefore (32) $$|_MV\pi _\lambda 𝑑\mu _\lambda |CV_{L^1}$$ which implies that $`\pi _\lambda d\mu _\lambda `$ is a continuous linear functional on $`L^1(M)`$, hence belongs to $`L^{\mathrm{}}(M)`$. That is, we may write $`\pi _\lambda d\mu _\lambda =f_\lambda dvol`$, with $`f_\lambda _{\mathrm{}}C.`$ If $`\pi _\lambda `$ had a singular value, it is easy to check that $`\pi _\lambda d\mu _\lambda `$ would blow up there . Hence, $`\pi _\lambda `$ is a non-singular projection. ∎ Now we turn to the general case: #### 3.1.2. $`^n`$ actions ###### Proposition 3.2. All regular tori project diffeomorphically to the base. Proof: Since by Lemma (2.8) a regular torus $`\mathrm{\Lambda }^{(l)}(b)`$ has charge $`c(\mathrm{\Lambda }^{(l)}(b))\frac{\omega ^{(l)}(b)}{_{j=1}^{m_{cl}(b)}\omega ^{(j)}(b)}>0`$, it follows by Corollary (2.9) that there exists a ladder of joint eigenfunctions $`\{\varphi _{b_j(\mathrm{})}\}_b`$ with the property that: $$V\varphi _{b_j(\mathrm{})},\varphi _{b_j(\mathrm{})}=\underset{l=1}{\overset{m_{cl}(b)}{}}c_l(_b)_{\mathrm{\Lambda }^{(j)}(b)}V𝑑\mu _{\mathrm{\Lambda }^{(j)}(b)}+o(1),$$ where $`c_l(_b)\frac{\omega ^{(l)}(b)}{_{j=1}^{m_{cl}(b)}\omega ^{(j)}(b)}>0`$ and $`c_k(_b)0`$ for $`kl`$. Thus, we have (as in the toric case) that $$c_l(_b)_MV\pi _{}𝑑\mu _{\mathrm{\Lambda }^{(j)}(b)}CV_{L^1},$$ where we can take $`C=c_l(_b)L^{\mathrm{}}(\mathrm{},b_j(\mathrm{});g,V).`$ Since $`c_l(_b)>0`$ we can cancel it to find that the torus projects regularly. ∎ As an immediate of Proposition (3.2) we have: ###### Corollary 3.3. Let $`\{\pi _{}d\mu _\mathrm{\Lambda }\}`$ denote the set of projections to $`M`$ of normalized Lebesgue measures on compact Lagrangean tori $`\mathrm{\Lambda }X_E.`$ Then, under the assumptions of Theorem (0.1), the family is uniformly bounded as linear functionals on $`L^1(M).`$ ### 3.2. Non-existence of singular levels We have: ###### Lemma 3.4. Under the assumptions of Theorem (0.1), $`𝒫`$ has no singular levels; all orbits are Lagrangean. Proof: Existence of a compact singular orbit contradicts the the uniform boundedness of eigenfunctions assumption. Indeed, it follows from Lemma (2.11) that, for any $`VC^{\mathrm{}}(M)`$, there exist a compact, singular orbit $`\mathrm{\Lambda }_{sing}^{(l)}`$ and $`L^2`$-normalized joint eigenfunctions $`\{\varphi _{b_j(\mathrm{})}\}`$ such that for some $`c(\mathrm{\Lambda }_{sing}^{(l)})>0,`$ (33) $$c(\mathrm{\Lambda }_{sing}^{(l)})_{\mathrm{\Lambda }_{sing}^{(j)}}V\pi _{}𝑑\nu _lCV_{L^1(M)}.$$ However, the estimate in (33) cannot hold since by definition, compact singular orbits have dimension $`dim\mathrm{\Lambda }_{sing}^{(l)}<n`$. Therefore, there cannot exist singular levels of the moment map $`𝒫`$. ∎ ### 3.3. Completion of proof of Theorem We first complete the proof of Theorem (0.1) for general metrics with quantum completely integrable Laplacians. Subsequently we take up the case of Schroedinger operators. The first step is to consider projections of regular Lagrangean tori. By Proposition(3.2), the assumption of uniformly bounded eigenfunctions then applies to show that all Lagrangean torus orbits must project regularly to $`M`$. Furthermore, by Lemma (3.4) we know that the under the finite geometric multiplicity condition (3) and uniform boundedness condition on the eigenfunctions, there do not exist any singular leaves of the moment map. Consequently, the proof of Theorem (0.1) in the case of Laplacians is a direct consequence of the following: ###### Lemma 3.5. Suppose that the geodesic flow $`G^t`$ of $`(M,g)`$ commutes with a Hamiltonian $`^n`$ action. Suppose that there are no singular levels of the moment map, and suppose that each regular Lagrangean orbit $`^n(x,\xi )`$ has a non-singular projection to $`M`$. Then $`(M,g)`$ is a flat manifold. Proof: We will give two proofs of the lemma. First Proof: The first proof uses Mane’s theorem (1.5): Since the foliation by orbits has no singular leaves, Mane’s theorem implies that $`(M,g)`$ has no conjugate points. Since each leaf is compact, it must be a torus which covers $`M`$. Thus, there exists a cover $`p:T^nM`$. Lift the metric to $`p^{}g`$ on $`T^n`$. The lifted metric must have no conjugate points since the universal covering metric is the same. By the Burago-Ivanov theorem (1.6), the metric is flat. ∎ In the second proof, we do not use Mane’s theorem, and directly relate the condition on torus projections to non-existence of conjugate points. Second Proof: As above, let $`\pi :T^{}M0M`$ denote the natural projection and let $`\pi _I=\pi |_{T_I}.`$ Since each $`\pi _I:T_IM`$ is non-singular, and $`dimT_I=dimM`$, $`\pi _I`$ must be a covering map. #### 3.3.1. Case 1: $`M`$ is a torus Let us first assume that $`M`$ is a torus, i.e. diffeomorphic to $`^n/^n`$; we make no assumptions on the metric. From the fact that $`p_I`$ is a covering map, it follows by a result of Lalonde-Sikorav (\[LS\]) that the degree of $`\pi _I:T_IM`$ equals 1 for all $`I`$. Since $`\pi _I`$ is a diffeomorphism, there are well-defined inverse maps $$\pi _I^1:MT_I$$ with $`K(I)=1.`$ They define sections of $`\pi :S^{}MM`$ and hence are given by graphs of 1-forms $`\alpha _I:MS^{}M.`$ Thus, $`|\alpha _I(x)|1`$ where $`||`$ is the co-metric. We have $`\pi _I^1\alpha =\alpha _I^{}\alpha =\alpha _I`$ where $`\alpha `$ is the canonical 1-form. Since the tori $`T_I`$ are Lagrangean, and since $`d\alpha =\omega `$, the 1-forms are closed, i.e. $`d\alpha _I=0.`$ Now let $`p:\stackrel{~}{M}M`$ denote the universal cover of $`M`$ and let $`^n`$ denote the deck transformation group, with generators $`\alpha _1,\mathrm{},\alpha _n.`$ The metric $`g`$ lifts to a $`^n`$-periodic metric $`\stackrel{~}{g}`$ on $`\stackrel{~}{M}`$. We note that the corresponding geodesic flow $`\stackrel{~}{G}^t`$ is also completely integrable. Indeed, the cover $`p`$ induces the universal cover $`p_1:T^{}\stackrel{~}{M}T^{}M`$ whose deck transformation group we continue to denote by $`^n`$. Then $`\stackrel{~}{G}^t`$ commutes with the $`T^n`$-action on $`T^{}M0`$ generated by the lifted action integrals $`\stackrel{~}{I}_j=p_1^{}I_j.`$ The invariant tori $`T_I`$ therefore lift to $`\stackrel{~}{G}^t`$-invariant level sets $`\stackrel{~}{T}_I`$ of $`(\stackrel{~}{I}_1,\mathrm{},\stackrel{~}{I}_n).`$ Furthermore, the 1-forms $`\alpha _I`$ lift to $`^n`$-invariant closed 1-forms $`\stackrel{~}{\alpha _I}`$ on $`\stackrel{~}{M}.`$ They are exact $`\stackrel{~}{M}`$ and hence have the form $`dB_I`$ for some ‘potential’ $`B_IC^{\mathrm{}}(\stackrel{~}{M})`$. The gradient $`B_I`$ is then a $`^n`$-invariant vector field on $`\stackrel{~}{M}`$. Since $`|dB_I|1`$ we have $`|B_I|1.`$ We now claim that the integral curves of $`B_I`$ are lifts of geodesics on $`T_I.`$ To see this, we recall that the generator $`\mathrm{\Xi }_H`$ of the geodesic flow lies tangent to each torus $`T_I`$. Hence for each $`I`$ it projects from $`T_I`$ to a non-singular vector field $`\pi _I\mathrm{\Xi }_H=\mathrm{\Xi }_I`$ on $`M`$. We have $$B_I,\mathrm{\Xi }_I=dB_I(\mathrm{\Xi }_I)=\alpha _I(\mathrm{\Xi }_I)=\alpha |_{T_I},\mathrm{\Xi }_H|_{T_I}=1$$ since $`\mathrm{\Xi }_H`$ is a contact vector field for $`(S^{}M,\alpha ).`$ Since $`|B_I|=1`$ it follows that $`B_I=\mathrm{\Xi }_I`$. This relation holds for the lifts to $`\stackrel{~}{M}`$ and hence the integral curves of $`B_I`$ are the lifts of the geodesics on $`T_I.`$ We now claim that $`g`$ has no conjugate points, i.e. that each geodesic of $`\stackrel{~}{g}`$ on $`\stackrel{~}{M}`$ is length minimizing between each two points on it. This follows by a well-known argument: Let $`\stackrel{~}{x}`$ be any point of $`\stackrel{~}{M}`$, let $`\stackrel{~}{v}S_{\stackrel{~}{x}}\stackrel{~}{M}.`$ and let $`\gamma _{\stackrel{~}{v}}`$ be the geodesic of $`\stackrel{~}{g}`$ in the direction $`\stackrel{~}{v}.`$ To see that $`\gamma _{\stackrel{~}{v}}`$ is length minimizing between $`\stackrel{~}{x}`$ and any other point $`\gamma _{\stackrel{~}{v}}(t_o)`$, we project it to $`S^{}M`$. The image lies in one of the (possibly singular) invariant tori $`T_I`$ and by the above, $`\gamma _{\stackrel{~}{v}}`$ is an integral curve of $`B_I.`$ If it is not length minimizing to $`\gamma _{\stackrel{~}{v}}(t_o)`$, then there exists $`s_o<t_o`$ and a second geodesic $`\alpha `$ with $`\alpha (0)=\stackrel{~}{x},\alpha (s_o)=\gamma _{\stackrel{~}{v}}(t_o).`$ This leads to a contradiction since $$B_I(\alpha (s_o))=_0^{s_o}B_I,\alpha ^{}(s)𝑑s=_0^{t_o}B_I,\gamma _{\stackrel{~}{v}}^{}(s)𝑑s=t_o>s_o$$ but $$t_o=|_0^{s_o}B_I,\alpha ^{}(s)𝑑s|s_o$$ as $`|B_I|=1`$. Therefore, $`(T^n,g)`$ is a torus without conjugate points. Theorem A then follows in this case from the recent proof by Burago-Ivanov \[BI\] of the Hopf conjecture that a metric on $`T^n`$ with no conjugate points is flat.∎ #### 3.3.2. The general case We now consider the general case where $`M`$ is only covered by a torus $`T^n`$ (namely $`T_I`$ for each $`I`$). We denote by $`p:T^nM`$ a fixed d-fold covering map. For notational clarity we denote the metric on $`M`$ by $`g_M`$. By Lemma (3.1), there is a Hamiltonian torus action on $`T^{}M0`$ with the property that every orbit projects non-singularly to $`M`$. Let $`g_T=p^{}g_M`$ be the metric induced on $`T^n`$ by the cover. We claim that $`g_T`$ is a flat metric. Since $`p:(T^n,g_T)(M,g)`$ is a Riemannian cover, this will imply that $`g_M`$ is a flat metric and conclude the proof of (a). To prove $`g_T`$ is flat, we lift the torus foliation of $`T^{}M0`$ to $`T^{}T^n0`$. Given a metric $`g`$ on a manifold $`X`$ we denote by $`\stackrel{~}{g}:TXT^{}X`$ the induced bundle map $`\stackrel{~}{g}(X)=g(X,).`$ We also consider the bundle map: $`dp:TT^nTM`$. Since $`dp_x`$ is a fiber-isomorphism for each $`xT^n`$, $`p`$ is a d-fold covering map. It follows that $$F:T^{}(T^n)T^{}M,F:=\stackrel{~}{g_M}d\rho \stackrel{~}{g}_T^1$$ is also a d-fold covering map. Let $`𝒯`$ denote the foliation of $`T^{}M0`$ by orbits of the torus action. We define $`F^1𝒯`$ to be the foliation of $`T^{}T^n0`$ whose leaves are given by $`\stackrel{~}{T}_I:=F^1T_I`$ where $`\{T_I\}`$ are the leaves of $`𝒯.`$ (The associated involutive distribution of the $`n`$-planes $`\stackrel{~}{T}_{x,\nu }T_{x,\nu }T^{}T^n0`$ is defined by $`dF(\stackrel{~}{T}_{x,\nu })=T_{F(x,\nu )}T_{I(F(x,\nu )}`$.) This foliation could also defined as orbits of the commuting Hamiltonians $`F^{}I_j`$ on $`T^{}T^n0`$. Each of the leaves is compact, hence a torus. We note that $`F:\stackrel{~}{T}_IT_I`$ is always a smooth covering map. We then have the commutative diagrams: (34) $$\begin{array}{ccc}\stackrel{~}{T}_I\hfill & \stackrel{F}{}& \hfill T_I\\ \pi \hfill & & \hfill \pi \\ T^n\hfill & \stackrel{p}{}& \hfill M\end{array}$$ We claim that the map $`\pi :\stackrel{~}{T}_IT^n`$ is non-singular. If not, the map $`\pi F:\stackrel{~}{T}_IM`$ would be singular. But as observed above, it is a covering map. It further follows by the result of \[LS\] that $`\pi :\stackrel{~}{T}_IT^n`$ has degree one, hence is a diffeomorphism. We have now reduced to the previous case of the torus: the metric $`g_T`$ must be a flat metric, hence $`g_M`$ must be flat. This completes the second proof of Theorem (0.1) in the case of torus actions. ∎ ### 3.4. Proof of Theorem (0.1) for Schroedinger operators We now consider the case of semiclassical Schroedinger operators $`\mathrm{}^2\mathrm{\Delta }+V.`$ Our proof in the homogeneous case (i.e. $`V=0`$) was based on the use of semiclassical pseudodifferential operators, so it generalizes with little change. Proof: We fix an energy level $`E`$ and consider eigenvalues of $`\mathrm{}^2\mathrm{\Delta }+V`$ lying in $`[EC\mathrm{}^{1\delta },E+C\mathrm{}^{1\delta }]`$ for some fixed $`C>0.`$ The eigenfunctions we consider are the joint eigenfunctions of $`P_1,\mathrm{},P_n`$ with joint eigenvalues $`(E_j(\mathrm{})=b_j^{(1)}(\mathrm{}),\mathrm{},b_j^{(n)}(\mathrm{}))`$ respectively, satisfying $`b_j^{(1)}(\mathrm{})[EC\mathrm{}^{1\delta },E+C\mathrm{}^{1\delta }]`$ for some $`0<\delta <1.`$ We recall that $`b=(b^{(1)}=E,b^{(2)},\mathrm{},b^{(n)})`$ and $`E`$ corresponds to the energy shell $`X_E`$ of the classical Hamiltonian $`1/2|\xi |_g^2+V`$ corresponding to the quantum Hamiltonian $`P_1=\mathrm{}^2\mathrm{\Delta }+V`$. By assumption, the eigenfunctions corresponding to these joint eigenvalues are uniformly bounded independently of $`\mathrm{}\mathrm{}_0.`$ By Proposition (3.2), it follows that all Lagrangean torus orbits of $`\mathrm{\Phi }_t^E`$ on $`X_E`$ project regularly to the base. Indeed, the proof that the torus $`\mathrm{\Lambda }^{(j)}(b)`$ projects regularly only involves trace formula and quantum limits over joint eigenvalues in the set $`\{(E_j(\mathrm{})=b_j^{(1)}(\mathrm{}),b_j^{(2)}(\mathrm{}),\mathrm{},b_j^{(n)}(\mathrm{})):|b_j(\mathrm{})b|\mathrm{}^{1\delta }\}.`$ . Hence our assumption on uniform boundedness of the eigenfunctions of $`P_1=\mathrm{}^2\mathrm{\Delta }+V`$ with eigenvalues in the interval $`[Ec\mathrm{}^{1\delta },E+c\mathrm{}^{1\delta }]`$ is sufficient to obtain the result of Proposition (3.2) for the tori on the energy shell $`X_E.`$ Hence, by a simple covering space argument, we can without loss of generality assume that the base manifold is a torus. By Lemma (3.4), there are no singular levels of the moment map $`𝒫|_{X_E}.`$ Hence $`X_E`$ has a smooth Lagrangrean foliation invariant under $`\mathrm{\Phi }_t^E.`$ By Proposition (1.7), we must have that $`E>V_{max}`$ and the Jacobi metric $`(EV)g`$ is flat. If we additionally assume that the sup norms are bounded indepedently of $`\mathrm{}`$ and $`E`$ in some interval $`[E_0ϵ,E_0+ϵ]`$, then the Jacobi metrics $`(EV)g`$ are flat for all $`E`$ in this interval, and it follows by Corollary (1.8) that $`g`$ is flat and $`V`$ is constant. ∎ ## 4. Problems and Conjectures We conclude with some problems conjectures on integrable systems and their eigenfunctions. ### 4.1. Symplectic geometry of toric integrable systems Some of the ideas of this paper are relevant to purely geometric problems. ###### Conjecture 4.1. Suppose that $`g`$ is a metric on $`^n/^n`$ which is toric integrable. Then $`g`$ is flat. This would follow from the solution of the Hopf conjecture and from ###### Conjecture 4.2. Up to symplectic equivalence, the only homogeneous Hamiltonian torus action on $`T^{}(^n/^n)`$ is the standard one ($`\mathrm{\Phi }_t(x,\xi )=(x+t\xi ,\xi ).`$) Indeed, the geodesic flow of $`(^n/^n,g)`$ would preserve the Lagrangean foliation defined by orbits of $`\mathrm{\Phi }_t`$ and hence by Mane’s theorem $`g`$ would have no conjugate points. Since the time of the original submission of this article, these conjectures have been proved by E. Lerman and N. Shirokova \[LS\]. ### 4.2. Eigenfunctions We assume throughout that the Laplacian or Schroedinger operator was quantum completely integrable. It is natural to ask if the hypothesis can be weakened to classical integrability. ###### Conjecture 4.3. Suppose that $`(M,g)`$ is a compact Riemannian manifold with completely integrable geodesic flow. Suppose that $`\mathrm{\Delta }_g+V`$ is a Schroedinger operator on $`(M,g)`$ all of whose ONBE’s have uniformly bounded sup norms. Then $`(M,g)`$ is flat. Without the assumption of quantum complete integrability, it is not even known whether eigenfunctions localize on level sets of the classical moment map. There are also interesting problems in the converse direction. We will explain the difficulty of the next conjecture when we come to multiplicities. ###### Conjecture 4.4. Suppose that $`\mathrm{\Delta }_g+V`$ is a Schroedinger operator on a flat manifold $`(M,g)`$. Then for generic $`V`$, are the eigenfunctions uniformly bounded? We further note that all the questions about sup norms are equally reasonable in the non-compact case. ### 4.3. KAM and classically non-integrable systems We now consider the extent to which even classical complete integrability can be dropped. It is plausible that sup norm blow-up occurs whenever there exists a stable elliptic orbit of the geodesic flow. In that case one can construct quasimodes associated to the orbit which do blow up. The relation between modes and quasimodes can be quite complicated in general, but it is plausible that there should exist a sequence of modes which also blows up. KAM systems always contain such stable elliptic orbits. We plan to consider these issues in a future article. ### 4.4. Multiplicities and sup norms There are (well-known) relations between eigenvalue multiplicities and sup-norm blow up of eigenfunction. If there exists a sequence of eigenvalues of unbounded multiplicity, then there exists an ONBE with unbounded sup norms. Indeed, for each $`x`$, consider the eigenfunction $`\mathrm{\Pi }_E(x,)`$ where $`\mathrm{\Pi }_E`$ is the orthogonal projection onto the eigenspace $`V_E`$. Then $`\mathrm{\Pi }_E(x,\dot{)}`$ has $`L^2`$-norm equal to $`\sqrt{\mathrm{\Pi }_E(x,x)}`$. So the normalized eigenfunction is $`\varphi _E^x():=\mathrm{\Pi }_N(x,)/\sqrt{\mathrm{\Pi }_N(x,x)}`$. It is well-known and easy to see (by the Schwartz inequality) that $`\varphi _E^x()`$ has its maximum at $`x`$, where it equals $`\sqrt{\mathrm{\Pi }_N(x,x)}`$. Since $`_M\mathrm{\Pi }_N(x,x)𝑑vol(x)=m(E)`$ (with $`dvol(x)`$ the volume form), there must exist $`x`$ so that $`\mathrm{\Pi }_N(x,x)m(E).`$ Hence $`\varphi _E^x()_{\mathrm{}}\sqrt{m(E)}.`$ When $`(M,g)`$ is a rational torus, $`L^{\mathrm{}}(\lambda ,M,g)`$ therefore grows at a polynomial rate while $`\mathrm{}^{\mathrm{}}(\lambda ,M,g)`$ stays bounded. For instance, on a flat torus $`^n/L`$, an ONBE of the standard Laplacian $`\mathrm{\Delta }_0`$ is given by the exponentials $`e^{i\lambda ,x},`$ with $`\lambda \mathrm{\Lambda }:=L^{}`$, the dual lattice to $`L`$. The associated eigenvalue is $`E=|\lambda |^2`$ and its multiplicity $`m(E)`$ is the number of lattice points of $`\mathrm{\Lambda }`$ on the sphere of radius $`\sqrt{E}`$. Counting this number is a well-known problem in number theory when the lattice is rational. When $`L=^n`$, for instance, the multiplicity function $`m(E|)`$ has logarithmic growth for $`n=2`$, and polynomial growth in higher dimensions. Under a perturbation by a potential $`ϵV`$, there exists a smoothly varying orthonormal basis of eigenfunctions (sometimes called the Kato-Rellich basis). It is possible that for some potential $`V`$ on $`^n/^n`$ , the Kato-Rellich basis for the perturbation $`\mathrm{\Delta }_0+ϵV`$ may be a smooth deformation of the eigenfunctions just described with high sup norms. If so, it is then possible that even if the multiplicity is broken and all eigenvalues become simple, the eigenfunctions can still have unbounded sup norms. Conjecture (4.4) states that such potentials should be sparse. It would be of some interest to understand if there exist any potentials for which sup norm blow-up occurs. The most extreme case of multiplicity is of course that on the standard sphere $`(S^2,g_0)`$. At this time of writing, it remains an open problem whether $`\mathrm{}^{\mathrm{}}(\lambda ,g)=O(1)`$ on the standard sphere. The best result to date is the upper bound of VanderKam \[V\], that for a ‘random’ ONB of eigenfunctions $`\{\varphi _\lambda \}`$ the sup-norms satisfy $`\varphi _\lambda _{\mathrm{}}/\varphi _\lambda _{L^2}=O(\sqrt{\mathrm{log}\lambda }),`$ i.e. $`\mathrm{}^{\mathrm{}}(\lambda ,S^2,g_0)=𝒪(\sqrt{\mathrm{log}\lambda }).`$ Our methods do not apply to this problem. ### 4.5. Quantitative problems Can one weaken the hypothesis of uniform boundedness of eigenfunctions in $`L^{\mathrm{}}`$ in the rigidity results? It is plausible that our rigidity results holds as long as $`L^{\mathrm{}}(\lambda ,M,g)`$ lies below some threshold. One may ask the same question for the analogous $`L^p`$ quantities $`L^p(\lambda ,M,g)`$. In \[TZ\] (see also \[T1\]\[T2\]), we analyse sup norm blow-up of eigenfunctions near singular levels (among other things). We also study some cases of sup norm blow up near singular projections of regular levels. To obtain a threshfold of some generality one needs to estimate the minimal blow up corresponding to the possible types of singular behaviour.
warning/0002/astro-ph0002372.html
ar5iv
text
# The Luminosity Function of MS2255.7+2039 at 𝑧=0.288Based on observations made with the Nordic Optical Telescope, operated on the island of La Palma jointly by Denmark, Finland, Iceland, Norway, and Sweden, in the Spanish Observatorio del Roque de los Muchachos of the Instituto de Astrofisica de Canarias. ## 1 Introduction The faint end of the luminosity function (LF) is of great interest both in connection to the excess found in number counts of faint galaxies and for theories of cluster formation. The latter is demonstrated especially in a series of papers by Kaufmann et al. (1993), Kaufmann (1995a, b), Heyl et al. (1995) and Baugh et al. (1996), which show how the LF is related to the hierarchical clustering in e.g. the Cold Dark Matter (CDM) model. A steep faint end of the LF is unique not only to the CDM scenario, but is a generic prediction of hierarchical models of cluster formation (White & Frenk 1991). The difference of the LF in the field and in clusters of galaxies also gives important information about environmental effects related to the galaxy formation process. In particular, ram pressure stripping and effects of interaction are likely to be more important for clusters than in the field (e.g., Moore et al. 1996). There is in this respect a clear connection to the Butcher-Oemler effect (Butcher & Oemler 1984), seen for clusters, but not apparent in the field. The fraction of blue galaxies in clusters is larger at higher redshifts, which is often interpreted as an evolutionary effect. The evolution of the cluster LF is therefore of obvious interest. The field LF has been investigated locally by several groups. Some of these report local LFs with flat faint-end slopes of $`1.1\begin{array}{c}<\\ \end{array}\alpha \begin{array}{c}<\\ \end{array}1.0`$ (Loveday et al. 1992; Ellis et al. 1996), while others have found a significantly steeper LF with $`\alpha \begin{array}{c}<\\ \end{array}1.5`$ (Marzke et al. 1994; Lilly et al. 1995). At higher redshifts especially the CFRS survey (Lilly et al. 1995) and the Autofib survey (Ellis et al. 1996) give information about the LF in the field up to $`z1`$. Lilly et al. find that, while strong evolution is seen for the blue sample at $`z\stackrel{}{\stackrel{>}{}}\mathrm{\hspace{0.17em}\hspace{0.17em}0.5}`$, the red LF is changing little. At $`z0.5`$ the LF has brightened by $`1`$ magnitude in B. At higher redshifts the bright end stays constant, while the faint continues to increase, leading to a steepening of the faint end of the LF, from $`\alpha =1.1`$ locally to $`\alpha =1.5`$ at $`z0.5`$, in broad agreement with the Autofib survey. A steepening of the LF can explain the excess counts found in deep surveys, as discussed by Gronwall & Koo (1995). While the field LF has been studied in fair detail, there have been only a few CCD investigations of the LF of clusters of galaxies until recently. Relatively nearby clusters, like Virgo, Fornax and Coma, have been studied by a number of authors (e.g., Bernstein et al. 1995, hereafter BNTUW; Lobo et al. 1997; Biviano et al. 1995; De Propris et al. 1995; Trentham 1998a). While some earlier investigations yield faint-end slopes of $`\alpha 1.3`$ (e.g., Ferguson & Sandage 1988), more recent investigations, taking low surface brightness galaxies into account, point to steeper slopes, $`\alpha \begin{array}{c}<\\ \end{array}1.5`$ (e.g., Bothun et al. 1991). At higher redshifts the information about the cluster LF is scarce. The first detailed study was that by Driver et al. (1994b, henceforth DPDMD), who studied the R-band LF to $`R=24`$ for the cluster Abell 963 at $`z=0.206`$. While the high luminosity end can be well fitted with a Schechter function, they found an increase of faint galaxies between $`19\stackrel{}{\stackrel{<}{}}M_\text{R}\stackrel{}{\stackrel{<}{}}16.5`$, with a slope of $`\alpha 1.8`$. Further investigations at $`0.1\begin{array}{c}<\\ \end{array}z\begin{array}{c}<\\ \end{array}0.2`$ have strengthened the case for steep slopes, $`\alpha \begin{array}{c}<\\ \end{array}1.7`$, at $`19\begin{array}{c}<\\ \end{array}M_\text{R}`$ (Smith et al. 1997; Wilson et al. 1997, hereafter WSEC), although examples of flatter LFs also exist (Trentham 1998b). It is obvious that these results need confirmation for more clusters, especially at higher redshifts. Because of the strong evolution of the Butcher-Oemler effect, as well as from numerical simulations of cluster evolution (e.g., Kauffmann 1995a,b), one expects substantial evolution even from $`z=0.2`$ to $`z=0.4`$. The detection of the faint end of the LF then becomes more difficult, both because the galaxies are fainter, and because of increasing contamination from the background. As a first step we here report observations of the cluster MS2255.7+2039 at $`z=0.288`$. We will in this paper use $`\mathrm{H}_\text{0}=50\mathrm{km}\mathrm{s}^1\mathrm{Mpc}^1`$ and $`\mathrm{\Omega }_\text{M}=1`$. ## 2 The data The coordinates of the center of MS2255.7+2039 = Zw 8795, hereafter MS2255, are $`\alpha =22^\mathrm{h}55^\mathrm{m}40.^\mathrm{s}6`$, $`\delta =20{}_{}{}^{}30{}_{}{}^{}04.^{\prime \prime }2`$ (1950.0), and the redshift is $`z=0.288`$ (Stocke et al. 1991). MS2255 was detected as an X-ray cluster in the Einstein Observatory Extended Medium Sensitive Survey (EMSS), with an X-ray luminosity of $`2.0\times 10^{44}\mathrm{erg}\mathrm{s}^1`$ in the $`0.33.5`$ keV band (Gioia & Luppino 1994), fairly typical for the EMSS selected sample. The galactic extinction can be estimated in several ways. Based on the H i column density, the absorption in the direction of MS2255 ($`l=90.^{}32,b=34.^{}67`$) is $`A_\text{B}=0.18`$ (Burstein & Heiles 1982). With $`A_\text{R}=0.61A_\text{B}`$ we get $`A_\text{R}=0.11`$. On the other hand, the estimated X-ray column density is $`5.0\times 10^{20}\mathrm{cm}^2`$ (Gioia et al. 1990). With $`N_\text{H}=4.8\times 10^{21}E_{\text{B-V}}`$ (Bohlin et al. 1978) and $`A_\text{R}=2.4E_{\text{B-V}}`$ this gives $`A_\text{R}=0.25`$, considerably higher than the value deduced from the H i column density. Finally, the recently presented COBE/DIRBE - IRAS/ISSA dust maps (Schlegel et al. 1998) yield $`E_{\text{B-V}}=0.06`$, which leads to $`A_\text{R}=0.14`$. We will in this paper use $`A_\text{R}=0.14`$, but note that this may be in error by $`0.1`$ magnitude. ### 2.1 Observations The data were obtained with the 2.56 m Nordic Optical Telescope and the Andalucia Faint Object Spectrograph (ALFOSC) in June 1997. The ALFOSC contained a thinned, back-side illuminated Ford-Loral 2K<sup>2</sup> chip that yielded a field of $`6.^{}5\times 6.^{}5`$ and an image scale of 0$`.^{\prime \prime }`$189/pixel. Every data set consists of the usual bias, dark, twilight-flatfield, and standard-star images, beside the science frames. To determine the LF of the cluster it is crucial to correct for the contribution from the field. A nearby field at $`\alpha =22^\mathrm{h}54^\mathrm{m}55.^\mathrm{s}04,\delta =21{}_{}{}^{}32{}_{}{}^{}19.^{\prime \prime }00(1950.0)`$ was chosen in order to have similar galactic latitude and extinction properties as the cluster. Furthermore, the seeing conditions during the observations turned out to be similar. All these factors are crucial to ensure that the foreground and background will be as similar to the cluster field as possible. Systematic errors may otherwise easily enter into the subtraction of the background. The total exposure time was 5400 seconds for both the cluster and the background field, respectively, divided into exposures of 900 seconds. During the observations we rotated the instrument by 90 and/or made offsets by 10<sup>′′</sup> between different frames. This was done in order to suppress the influence of bad regions on the chip and to make it possible to create a night-sky flatfield from the object frames (e.g., Tyson 1988). A pointing error of the telescope at the time of the observations may have caused the centre of the cluster and the image field centre to differ slightly. ### 2.2 Reductions The bias level was determined from the overscan of each frame. These values were used to scale a master bias that was subtracted from the frames. We then used the shifted, bias-subtracted science frames to create a night-sky flatfield by removing objects with a ’smooth-and-clip’ (Näslund 1995), followed by an averaging of the frames. We corrected the large-scale gradients of the twilight flatfield by the night-sky master flat, and used the corrected twilight flatfield to flat the science frames. A narrow strip along the edges had to be excluded in the frames due to the structure of the thinned CCD. The flat-fielded frames were sky-subtracted, corrected for atmospheric extinction, aligned and finally combined (see Näslund 1995). The rotation and shifting of the frames decreased the effective area of the combined image, so that the area of the background and cluster images became $`5.^{}6\times 5.^{}5=31.0\mathrm{}`$ and $`5.^{}4\times 5.^{}6=30.2\mathrm{}`$, respectively (the final detection areas were reduced somewhat in order to avoid edge effects; see below). The images were calibrated using standard stars observed at intervals during the night. The seeing in the combined images was in both cases $`0.^{\prime \prime }85`$ (FWHM). The background field is shown in Fig. 1, while Fig. 2 displays the cluster. The individual frames of the cluster and background fields were checked for ’internal consistency’. We selected a few common objects in the magnitude range 17-21 and determined their brightness both in FOCAS (Jarvis & Tyson 1981; Valdes 1982, 1993) and DAOPHOT (Stetson 1987). All objects were consistent within a few hundredths of a magnitude or better. The weighted average of the standard stars observations yielded an statistical uncertainty of $`\pm 0.01`$ magnitude. ## 3 Analysis of the fields ### 3.1 Object catalogue The reduced fields were analyzed using the FOCAS package. As a limit to our detections we used a value of $`2.5\sigma `$ of the sky noise and a detection area $`A`$ of 20 pixels, which corresponds to the seeing. We then utilized the ordinary FOCAS procedure, including sky correction and splitting of multiple objects. The splitting procedure was checked by simulations and worked satisfactory in most cases. A region of 50 pixels along the edges was excluded when counting galaxies, which decreased the effective area of the background field to $`27.65\mathrm{}`$ and of the cluster field to $`27.02\mathrm{}`$. This procedure resulted in a catalogue of objects with a number of parameters such as isophotal R magnitudes $`(m_\text{R}^{\text{iph}})`$ and intensity weighted first-moment radii $`(r)`$. A plot of the detected galaxies is shown in (Fig. 3). ### 3.2 Simulations A substantial part of the analysis consists of simulations in order to determine completeness levels, fraction of noise detections, and magnitude corrections. When we approach the level of the sky noise we will obviously lose some galaxies in the noise, as well as pick up false detections. This problem is for a given total magnitude most severe for galaxies with large scale lengths and therefore low surface brightness. As discussed by, e.g., DPDMD and WSEC, the determination of this factor is far from trivial, and unfortunately, model-dependent assumptions about the galaxies are necessary. In one approach, discussed by DPDMD, one generates artificial galaxies with fixed parameters shifted to different redshifts, to describe the surface brightness distribution. An alternative method, used by WSEC, is to use real galaxies at brighter magnitudes, typical for the population in the field, as templates, and then rescale these to fainter magnitudes. This approach has the advantage of using realistic brightness profiles. It, however, implicitly assumes that the relative fractions of the different morphological types are the same for faint magnitudes as for bright, and also that the brightness profiles of a given type is independent of luminosity, except for a normalization factor. Neither of these assumptions are obvious. We used a generalized version of the DPDMD method in this investigation; the objects were not only characterized by their brightness, but also by their scale size. Simulated exponential disks were added to the background-field image, and these artificial galaxies were then detected with the same criteria as were used for the real data. After detection we could determine the magnitude correction depending on the objects position in the magnitude-scale length diagram (see Section 3.2.3). #### 3.2.1 Noise simulations First, a number of Poisson-noise frames were created, with a noise level corresponding to that of the data. We then used FOCAS to detect spurious features with different combinations of the upper sigma limit and detection area. The parameters we settled for, $`\sigma =2.5`$ and $`A=20`$ pixels ($``$ seeing), yielded six noise detections in an area of $`27\mathrm{}`$. This amounts to about 0.2% of the actual detections in the background field. #### 3.2.2 Completeness To estimate the completeness we simulated exponential disks of different magnitudes, scale lengths and inclinations. For each set of parameters, 51 exponential disks were generated in empty regions of the background-field image. The reason for positioning the simulated galaxies in empty regions was to isolate the detection completeness due to surface brightness, and treat other effects, like overlapping (cf. Section 3.2.4), separately. We used FOCAS with the same detection criteria as for the real data. The parameters of the detected artificial objects could then be extracted from the resulting catalogue. As mentioned below, and discussed by other authors (DPDMD, Trentham 1997), the possibility of detection, as well as the fraction of the total light recovered, varies with surface brightness, which in turn depends on scale length and inclination for a given magnitude. At fainter magnitudes, galaxies with short scale length are, as expected, more easily detected than those with long scale lengths. The simulations indicate that we detect all dwarf galaxies, modelled as exponential disks ($`0.5\begin{array}{c}<\\ \end{array}r_\text{d}\begin{array}{c}<\\ \end{array}2`$ kpc, where $`r_\text{d}`$ is the disk scale length), down to $`R25`$. An important complication is the possible presence of low surface brightness galaxies (LSBGs). These come in different flavours, such as the sample of blue Low-Surface-Brightness Galaxies of McGaugh & Bothun (1994) and the Giant Low-Surface-Brightness Galaxies of Sprayberry et al. (1995). The central surface brightness distribution of galaxies is not well known, although research during the last decade has shed more light on this particular type of object (Impey & Bothun 1997). If LSBGs are peaked around $`\mu _\text{B}^\text{0}23.2\mathrm{mag}/\mathrm{}`$<sup>′′</sup>(corresponding to $`\mu _\text{R}^\text{0}22.0\mathrm{mag}/\mathrm{}`$<sup>′′</sup>), as those observed by Sprayberry et al. and McGaugh & Bothun, we would detect LSBGs with small and intermediate scale lengths ($`r_\text{d}3`$ kpc), given the colours of the Sprayberry et al. sample. Even objects with lower surface brightness, like UGC 9024 ($`\mu _\text{B}^\text{0}=24.5,r_\text{d}=5.6/h`$ kpc), should be possible to detect at $`z=0.288`$, although such galaxies are close to our detection limit. Galaxies with lower surface brightness will accordingly escape detection. The most extreme cases known (e.g. Malin 1) are far beyond detection, but galaxies of this type are likely to be an order of magnitude fewer than objects of higher surface brightness (Davies et al. 1994). Furthermore, a study of Hubble Deep Field data by Driver (1999) shows that luminous low surface brightness galaxies are rare compared to their high surface brightness counterparts. One should, in any case, bear in mind that LSBGs below the surface-brightness detection limits may influence the faint-end slope by an unknown amount. #### 3.2.3 Magnitude correction From the simulations above, used to determine the completeness, we also estimated the total magnitude for each parameter set. During the analysis we developed a technique for magnitude corrections (Näslund 1998), which turned out to be similar to that of Trentham (1997). A large number of simulated galaxies, with different values of scale length ($`r_\text{d}`$), axial ratio ($`b/a`$) and total magnitude ($`m_\text{R}^{\text{tot}}`$), were generated in the background-field image. These objects were detected with the same setup as for the real data. We could at this stage calculate the magnitude correction as a function of isophotal magnitude ($`m_\text{R}^{\text{iph}}`$) and intensity weighted first-moment radius ($`r`$). We, finally, corrected the magnitudes of the detected galaxies from their position in the $`(m_\text{R}^{\text{iph}},r)`$ plane (Fig. 3). This method will obviously not be fully correct for elliptical galaxies that are better described by de Vaucouleur profiles than by exponential disks. Most of these are, however, of comparatively bright magnitudes, where the correction is small. If we assume that the faintest cluster members, for which the corrections are largest, have exponential profiles the application of the method is justified. This is plausible if the faintest galaxies are late-type spirals or dwarf spheroidals and/or dwarf irregulars. However, some of the galaxies close to our magnitude limit ($`R24`$) may be luminous dwarf ellipticals ($`16<M_\text{B}`$) that are better fitted by de Vaucouleurs profiles (Ferguson & Binggeli 1994). We performed a few simulations of de Vaucouleurs profiles with short scale lengths to mimic luminous dwarf ellipticals, and compared them to exponential disks of the same brightness. We find that the total magnitudes for the $`r^{1/4}`$ profile galaxies are underestimated by 0.15 magnitudes, similar to the findings of Trentham (1997). The magnitude correction can be applied in two ways. One can either correct for ’light loss’ without any assumptions about the cluster population, or one may use a priori information to constrain the distribution of points in the ($`m_\text{R}^{\text{iph}},r)`$ plane. If the faintest cluster members that we detect are exclusively dwarfs, this has two implications. Firstly, for objects with $`r_\text{d}\begin{array}{c}<\\ \end{array}2`$ kpc the catalogue should be more than 50% complete for $`m_\text{R}^{\text{iph}}<26`$. Secondly, the magnitude correction would in this case be fairly small, and the ambiguity at the faintest magnitudes is reduced, compared to a more complex population. This is because faint galaxies are found in a comparatively small region of the $`(m_\text{R}^{\text{iph}},r)`$ plane, and the reason is simply that faint, intrinsically large galaxies (i.e. LSBGs), will have their apparent scale lengths substantially reduced, and thereby be closer to the true dwarf region of the $`(m_\text{R}^{\text{iph}},r)`$ plane. As a result, the uncertainties in the magnitude correction increase at the faint end. If one for good reasons could justify the exclusion of non-dwarfs from this region of the $`(m_\text{R}^{\text{iph}},r)`$ plane, the corresponding magnitude correction would be more accurate. In addition, if all faint galaxies are dwarfs and have $`r^{1/4}`$ profiles, the luminosity after correction would be systematically underestimated (see above). On the other hand, if they all are luminous dwarf ellipticals the magnitude correction would not be increasing with magnitude as steeply as the one applied here, and the LF would therefore be less steep at faint magnitudes. However, there is no strong motivation for such an exclusion of intrinsically larger objects, and for the data presented here we used the first, more general, approach. The possible presence of larger, faint objects also calls for the use of exponential profiles. The magnitude-correction procedure was tested by generating a number of exponential disks distributed in magnitude according to a power law of slope 0.4. The galaxies were positioned along a grid in order to avoid crowding effects in this particular test. The simulations showed that the procedure managed to correct for the light loss well down to $`R24`$, close to the actual completeness limit (for $`r_\text{d}\begin{array}{c}<\\ \end{array}6`$ kpc) of the data (Näslund 1998). #### 3.2.4 Crowding Overlapping objects is another important factor, especially for faint objects. One way to see this is by noting that fainter objects have smaller effective field areas at their disposal. Stars and galaxies that are brighter in general occupy a larger apparent area in the image and fainter objects are shielded by them. We have tested different methods for this correction. In the first approach we added simulated compact objects of different magnitudes to the cluster and background images. To a first approximation, the fraction of recovered objects gives the detection probability as a function of magnitude. However, non-linear effects, connected with obscuration between especially the faint galaxies themselves, are likely to be important, and full simulations of the cluster and background would be needed to address these aspects. Instead of this method we adopted a more conservative correction procedure for the obscuration of faint objects. The cumulative area covered by objects up to a certain magnitude was calculated in half-magnitude intervals for the background and cluster field, respectively. It was then found that there is a clear difference between the area covered in the cluster image and the background - field image when it comes to bright objects. However, for fainter objects there is no substantial difference; the covered area increases similarly in both fields. In practice, this means that for objects brighter than $`R22`$ the obscured area is 8% in the cluster image, while the corresponding number for the background image is only 3.4%. We accordingly used these numbers to correct for the obscuration for objects fainter than $`R=22`$. We do therefore not include any magnitude dependence of the correction factor below $`R=22`$, as Smith et al. (1997) do, which implies that apparently small objects are not an important source of obscuration in our fields. Furthermore, these effects do not add linearly with brightness for faint objects. In the case that faint objects contribute somewhat themselves to the obscuration, we would have underestimated the faint - end slope of the LF slightly. We hope to perform a more thorough study of crowding effects elsewhere. ### 3.3 Foreground and background corrections #### 3.3.1 Stars MS2255 is at comparatively low galactic latitude ($`b=34.^{}67`$), and a substantial contamination by stars is expected. Because our comparison field is close to the cluster, this contamination is to a large extent reduced when we subtract the background counts from the cluster counts. With FOCAS, we can nevertheless separate stars from galaxies, based on the brightness profile, for $`R\begin{array}{c}<\\ \end{array}20`$. This eliminates the statistical errors in the cluster counts from this source. The saturated stars were removed interactively from the galaxy list, while remaining stars brighter than $`R=20`$ were detected and classified by FOCAS. We found in this way 25 stars in the cluster field, and 17 in the background field. This can be compared to the number expected from the galactic model by Bahcall & Soneira (1980) and Bahcall (1995), which for a field of $`27\mathrm{}^{}`$ gives 30 stars brighter than $`R=20`$. Within the statistical errors we consider the number of stars in the cluster field to be consistent with that expected from the model. The lower number for the background field is simply explained by the fact that the field was selected in a region void of bright objects. Besides bright stars, we also excluded small spurious objects around bright stars or galaxies, which can be a result of false detections by FOCAS in the cluster and background field (see Trentham 1997 for a discussion). This effect does not affect the faint-end slope significantly. #### 3.3.2 Background counts Guided by the completeness simulations, we decided to set the isophote limit for inclusion in the catalogue at $`m_\text{R}^{\text{iph}}=26`$, approximately corresponding to the magnitude limit after correction given above. According to the simulations we detect all pointlike sources at this isophotal magnitude. Inspection of the ($`m_\text{R}^{\text{iph}},r`$) plane shows that some of the detected objects fall below the simulated point sources, i.e. they have scale lengths smaller than the seeing. The apparent small scale length may be a result of the small number of counts (ADUs) for the faint objects. There is then a substantial probability that even for a point-like object the scale length will be smaller than the width of the PSF. Moreover, the simulations showed that disks with large scale lengths that are below the completeness limit could be detected as such compact objects, but almost all of them have $`m_\text{R}^{\text{iph}}`$ beyond our catalogue limit anyway, and hence cause no problem. The ambiguity in the interpretation of the nature of these objects made us test whether they influenced the shape of the cluster LF. We generated one list that included these objects, and one in which they were removed for both the cluster image and the background-field image. The resulting faint-end slope is in the latter case somewhat flatter compared to the slope including these objects, which is the one presented in this paper. In Fig. 4 we show the corrected R - band background counts, together with those obtained by the Hitchhiker team at WHT (Driver et al. 1994a) and the counts from BNTUW. As seen in the figure, the corrected background counts agree well with each other down to $`R24`$, within the limited statistics. While we have a total background area of $`27.65\mathrm{}^{}`$, Driver et al. had a total area of $`15.9\mathrm{}^{}`$. Their exposures with the WHT were, however, deeper in these fields. We also note that our counts are within the variations of other recent investigations such as those by Arnouts el al. (1999) and Fontana et al. (1999). ## 4 The cluster luminosity function ### 4.1 Error estimation of cluster counts Because of both statistical and systematic errors a careful error analysis has to be made. The standard method has been to assume Poisson statistics for the galaxies in the background field and cluster field, respectively, in some cases supplemented by an error for the field - to - field variation caused by large scale structures. Because we have only one background field, we cannot determine the field-to-field variation from our material. We therefore use the field - to - field statistics of BNTUW, with the characteristics of our data, to estimate the variation in the background counts as a function of magnitude (see also Trentham 1997). This value was added in quadrature to the Poisson variation in the cluster counts and the error due to the magnitude correction. The latter was estimated as the $`1\sigma `$ dispersion in a number of simulations (see Section 3.2.3). The two error sources are generally of comparable magnitude, but the field-to-field variation is systematically larger for fainter magnitudes ($`R>22.5`$). ### 4.2 The raw LF By subtracting the background counts from the cluster counts, we obtain the LF in terms of the apparent, isophotal magnitude. This ’raw’ LF is shown in Fig. 5. Then we applied the same distance modulus, including $`K`$-correction, as for the corrected distribution (see below). The resulting distribution can be fitted by a single Schechter function with $`\alpha =1.4`$ and $`M_\text{R}^{}=24`$. It is important to note that these numbers are arrived at partly because of the two ’low’ bins at $`M_\text{R}=20.75`$ and $`M_\text{R}=20.25`$. If we exclude these points, we obtain $`\alpha =1.3`$ and $`M_\text{R}^{}=22.8`$. Although the shape of the LF suggests that the bright end would be better fitted by a Gaussian, these low values can be explained as a statistical effect. Both points are fitted within the $`2\sigma `$ level by a Schechter function with parameters $`M_{\text{R,giants}}^{}=22.5`$, $`\alpha =1.0`$ (see Fig. 7). Monte Carlo simulations of this Schechter function with the same total number of galaxies as observed, show that a dip similar to that in Fig. 5 appears in $`(1020)\%`$ of the simulations. We therefore conclude that this dip is of marginal significance. The Schechter parameters can be compared to those of Valotto et al. (1997), who found $`\alpha =1.4`$ for rich clusters and $`\alpha =1.2`$ for poor. A linear fit for the range $`19<M_\text{R}<17`$ also yielded $`\alpha =1.3`$, which is the least model dependent estimate of the uncorrected faint-end slope. A comparison of ’raw’ LFs at different redshifts is, however, misleading. ### 4.3 The corrected LF For an unbiased comparison with samples at other redshifts, the isophotal magnitudes have to be related to the total magnitude, and then translated into absolute magnitudes. For the latter we included an R-band $`K`$-correction given by Metcalfe et al. (1991), who found a $`K`$-correction of 0.41 magnitudes for E/S0 galaxies at $`z=0.288`$, while the average for spirals was 0.09. We used a straight mean value $`K=0.15`$, but realize that the true range may be $`0.1\begin{array}{c}<\\ \end{array}K\begin{array}{c}<\\ \end{array}0.41`$. Because of the type dependence of the $`K`$-correction the intrinsic cluster LF is expected to be slightly redistributed, compared to the one observed. This redistribution should, however, be within the limits just mentioned. Thus, the distance modulus became $`(mM)=41.47K`$. In order to estimate the influence of the $`K`$-correction on the measured slope we made two tests. In both cases we kept the correction of the bright population fixed at $`K=0.15`$, but applied values of $`K=0.1`$ and $`K=0.4`$, respectively, to the dwarfs. This was done by simply assuming that all bins in the LF at $`R>23`$ represent the dwarf population, since this is the region where the dwarfs dominate in the corrected LF (see Fig. 7). In neither case is the faint-end slope for the the fitting method described below affected. The uncorrected LF presented in Section 4.2 represents a strict lower limit to the true LF. By applying the magnitude and crowding corrections, described in previous sections, to the cluster field and background field, respectively, we arrive at the LF shown in Fig. 6. While providing a useful description of the data, a Schechter fit is not unique, and we have tested different parametrizations of the corrected data. A single Schechter function with $`M_\text{R}^{}=23.9`$ and $`\alpha 1.43`$ gives a decent fit, although there is an indication of a break at $`M_\text{R}20`$. Binggeli et al. (1988) discussed the separation of the total LF into components for each Hubble type for the Virgo cluster. This is not possible for our data, because of the distance to the cluster, the lack of colour information, etc. We can, however, use the a priori information that the transition region between giants and dwarfs is at $`18\begin{array}{c}<\\ \end{array}M_\text{B}\begin{array}{c}<\\ \end{array}16`$. For comparison reasons, we add two separate Schechter functions, one representing giant galaxies and the other representing dwarfs. If the slope of the giants is fixed at $`\alpha =1.0`$, the procedure yields $`M_{\text{R,giant}}^{}=22.8`$ and $`M_{\text{R,dwarf}}^{}=18.9`$ for the two populations. This is close to the corresponding values found by DPDMD ($`22.5,19.0`$). The main discrepancy is in the slope of the dwarf population. For MS2255 we find $`\alpha 1.5`$. This can be compared to A963 for which DPDMD find $`\alpha 1.8`$. It is, however, important to note the coupling between $`M^{}`$ and $`\alpha `$. For $`M_{\text{R,dwarf}}^{}=19.5`$ our faint-end slope yields $`\alpha 1.8`$, while $`M_{\text{R,dwarf}}^{}=18.5`$ results in $`\alpha 1.2`$ (the dwarf/giant ratio was, however, not constrained in these tests, as in the case of DPDMD’s fit). The strong coupling between $`\alpha `$ and $`M_\text{R}^{}`$ makes it dangerous to draw any firm conclusions based on Schechter-function fits only. Instead, one should directly compare any two LFs, magnitude by magnitude, as shown below. In order to avoid the coupling of the Schechter-function parameters for the faint end of the LF, we also used a simple straight-line fit to the last five data bins, $`19.5M_\text{R}17`$ (as proposed by Trentham 1998b), which yielded $`1.6\begin{array}{c}<\\ \end{array}\alpha \begin{array}{c}<\\ \end{array}1.5`$. This is somewhat steeper than the ’raw’ LF, and is caused by the magnitude and obscuration corrections. This steepening of the LF is consistent with the bright end of the dwarf population detected in more nearby clusters (see discussion below). Although the procedures discussed above yield a formal value of the faint-end slope of $`\alpha 1.5`$, the uncertainties involved in these kind of studies make us emphasize that one should not focus on the exact value of the slope, but rather on the qualitative appearance of the LF. ### 4.4 Comparison with other clusters A major reason for the interest in the cluster LF is to study the evolutionary effects with redshift. As we have just discussed, this is done best by a direct comparison of the different LFs, i.e. by plotting them together, including all corrections. A problem is that the observations are in different filters and/or that the $`K`$-corrections are uncertain. A possibility to avoid some of these uncertainties is to compare the LFs in filters with central wavelengths adjusted to the redshift. In our case we note that at $`z=0.288`$, the R band corresponds to a wavelength between B and V at $`z=0`$. The other, more model dependent, method is to use a $`K`$-correction for a given $`z`$ and a given filter. This depends, however, on the population (section 4.3), as well as on evolution. Unfortunately, we are in most cases forced to use this alternative. In Fig. 7 we display the LF for MS2255 together with Trentham’s (1998a) LF of Coma ($`z=0.023`$) and the LF of A963 ($`z=0.206`$) by DPDMD, all adjusted to $`\mathrm{H}_\text{0}=50\mathrm{km}\mathrm{s}^1\mathrm{Mpc}^1`$. Both LFs were normalized to the same level as MS2255 for galaxies brighter than $`M_\text{R}=21`$. Although a fainter limit for this normalization would have been preferable, this value was chosen to avoid influence from the low points in the LF of MS2255 at $`M_\text{R}21`$. It is evident that all three clusters exhibit steep slopes at the faint end, and there is accordingly no qualitative difference between nearby and distant clusters in that respect. The slope of Coma is actually as steep as that of A963, while MS2255 displays a somewhat flatter faint end of the LF. The steep slope of the Coma LF was also noted by Smith et al. (1997), who used Coma data from Thompson & Gregory (1993). The steepening occurs at slightly different magnitudes. A963 has its break point around $`M_\text{R}=19.5`$. The steepening in MS2255 starts at approximately the same magnitude, while that in Coma occurs at a somewhat fainter magnitude ($`M_\text{R}18.5`$). The shape of the faint end of the cluster LF is a matter of controversy. While both nearby and more distant clusters show a steepening of the LF, the magnitude where this occurs differs substantially. To some extent this may be caused by a simple zero-point shift between filters. Most studies at intermediate redshifts ($`0.1z0.2`$) yield steep slopes ($`2\begin{array}{c}<\\ \end{array}\alpha \begin{array}{c}<\\ \end{array}1.7`$) for $`19\begin{array}{c}<\\ \end{array}M_\text{R}`$ (DPDMD; Smith et al. 1997; WSEC). However, the studies by Trentham (1998b, c), which in some cases are based on local clusters, give a more shallow slope ($`\alpha 1.4`$). In some local clusters, the steep slope starts at $`M_\text{R}14`$, which is a region beyond the limits of present studies for clusters around $`0.2\begin{array}{c}<\\ \end{array}z\begin{array}{c}<\\ \end{array}0.3`$. One reason for the discrepancies can be the different correction methods applied. It should be noted that earlier photographic investigations of nearby clusters also suffered from severe selections effects that work against low surface brightness objects. When such effects have been corrected for, the resulting slope is in the range $`1.5\begin{array}{c}<\\ \end{array}\alpha \begin{array}{c}<\\ \end{array}1.8`$ (Impey et al. 1988; Bothun et al. 1991). Another subject of interest is the possible existence of a universal cluster LF. In view of the discussion above, this is highly controversial and depends e.g. on the correction procedures applied. Composite LFs of clusters at $`0.02z0.2`$ have been presented both in B and in R, based on more than 15 (in B) and six clusters and groups (in R), respectively (Trentham 1998c, d). The composite R-band LF has a slope of $`\alpha 1.5`$ at $`M_\text{R}17`$, which is similar to what has been found here for MS2255. The steep slope of Coma (Trentham 1998a) seems to have been averaged out by the weighting procedure in the composite LF. Trentham noted, however, that his composite function may be valid only for the centres of rich clusters, i.e. regions dominated by ellipticals. WSEC studied two clusters, A665 and A1689, both at $`z=0.18`$, in V and I. This study is especially interesting because it gives some information about the colour dependence of the LF. WSEC found rising LFs with breaks around $`M_\text{V}=19`$ and $`M_\text{I}=21`$. Their V-band slope, after correction for incompleteness and obscuration, is very steep ($`\alpha 2`$), while that in the I band is significantly flatter ($`\alpha 1.1`$). This difference is interesting, since it indicates that the faintest detected galaxies indeed are blue. If this effect is real, one would expect an even steeper slope in B. Somewhat surprisingly, Trentham (1998b) did not find a comparatively steep B-band LF in his study of A665. Furthermore, while displaying very steep slopes ($`\alpha <2`$), the four Abell clusters investigated by De Propris et al. (1995) did not show any differences between B and I in this respect. The question of colour dependent slopes is therefore still unanswered. Few numerical simulations of cluster LFs exist. White & Springel (1999) discuss in a recent paper a combination of $`N`$-body and semianalytical modelling of the cluster population. Unfortunately, they only present the B-band LF, for which they find a faint end slope of $`\alpha =1.2`$. Although a direct comparison is difficult (R at $`z=0.3`$ corresponds approximately to V at $`z=0`$), this is considerably flatter than the observed slope presented here. ## 5 Summary and conclusions We have observed the galaxy cluster MS2255.7+2039 ($`z=0.288`$) and a background field at similar galactic latitude with the aim of determining the cluster LF. The isophotal magnitudes have been corrected for light loss according to results obtained from simulations. We have also compensated for obscuration due to bright, apparently large, objects in the images. The resulting cluster LF has a fairly steep faint-end slope ($`\alpha 1.5`$) faintward of the break in the profile around $`M_\text{R}=19`$. This slope is more shallow than some LFs found in clusters both locally and at $`z0.2`$, but similar to the slope of the composite LF derived by Trentham (1998c). Without focusing too strongly on the precise value of the slope, we conclude that MS2255.7+2039 exhibits a steepening LF at faint magnitudes. The evidence for steep faint-end slopes of cluster LFs is accumulating. There are now a number of fairly deep CCD investigations of nearby, as well as a few medium distant ($`z\begin{array}{c}<\\ \end{array}0.3`$) clusters, that all point to rising LFs at faint magnitudes. It therefore seems clear that a flat LF (i.e., $`\alpha =1`$) can be ruled out even at intermediate magnitudes ($`20\begin{array}{c}<\\ \end{array}M_\text{R}\begin{array}{c}<\\ \end{array}17`$). However, several questions remain unanswered. The uncertainties in the measured slopes are probably considerable, since different correction methods seem to yield deviating results, which probably explains the discrepancies between the LFs found for the same clusters as determined by different investigators (see section 4.4). Because of these uncertainties, it is too early to discuss any variation of the faint-end slope with $`z`$. The accuracy of the present study only allows us to claim that the cluster LF is non-flat at faint magnitudes ($`19\begin{array}{c}<\\ \end{array}M_\text{R}`$). The exact values of the slope and the magnitude where the steepening sets in are uncertain, and any trend with $`z`$ that may be present is dominated by these uncertainties. In addition, environmental differences, like richness or density, between clusters at the same $`z`$ could affect individual LFs, making distinctions in $`z`$ even more difficult to isolate. The uncertainties in the background subtraction is also a source of error, although the simulations by Driver et al. (1998) show that the faint-end slope of the LF can be reliably determined out to $`z0.3`$ with seeing and depth similar to those of the present data. Nevertheless, the errors due to the statistical background subtraction can probably be substantially reduced by using photometric redshifts. Work based on this approach is in progress. There are in the context of cluster LFs several important questions to answer in the future: Is there a universal LF for galaxy clusters at low redshifts, or is the steepness of the dwarf population different between clusters? Is there a colour dependence of the steepness of the dwarf population, as may be indicated in the study by WSEC? The K-band observations of five clusters by Trentham & Mobasher (1998) are especially interesting in this context. These data were, however, not deep enough to draw any conclusions about the faint end of the luminosity function. From the clear signs of evolution between $`z=0.5`$ and the present epoch for the field (e.g., Lilly et al. 1995), one would expect a corresponding evolution in clusters. From the galaxy harassment scenario for the Butcher-Oemler effect (Moore et al. 1996) one may expect a larger fraction of low luminosity galaxies in the past, and therefore a steeper LF. We hope to address some of these issues in the future. ###### Acknowledgements. We are grateful to Helmuth Kristen for obtaining a few images of the cluster in September 1995 and to Steven Jörsäter for some initial observations in 1994. We also thank Leif Festin for supplying some ALFOSC images that we could use to test the image quality prior to our observations. We are grateful to Margrethe Wold for discussions about completeness and to Tomas Dahlén for discussions and assistance with an additional consistency check. We also thank the referee, C. Gronwall, who provided several important suggestions that improved the presentation of this work. Last but not least, M. N. is very grateful to Stefan Larsson for discussions about statistics and related topics. The data presented here have been taken using ALFOSC, which is owned by the Instituto de Astrofisica de Andalucia (IAA) and operated at the Nordic Optical Telescope under agreement between IAA and the NBIfA of the Astronomical Observatory of Copenhagen. This research was supported by the Swedish Natural Sciences Research Council, and the Göran Gustafsson Foundation for Research in Natural Sciences and Medicine.
warning/0002/hep-th0002238.html
ar5iv
text
# A Geometrical Description of the Consistent and Covariant Chiral Anomaly ## 1 Introduction A mathematical rigorous definition of the consistent anomaly in the space-time formalism is in terms of a connection on a line bundle over $`𝒜`$, the space of gauge potentials. Similarly, the consistent Schwinger term is defined by a curvature on a line bundle over a dense subset of $`𝒜`$. Despite this, no geometrical description has been given for the covariant counterparts. In this paper we will obtain a geometrical interpretation of the covariant anomaly and Schwinger term in a similar way to what has been done in the consistent framework. For completeness, we will also review the consistent case. The result is that the consistent and covariant anomaly are given by connections on line bundles. The consistent Schwinger term is given by the curvature of a line bundle while for the covariant Schwinger term this is only true up to a (canonical) form defined by data coming from a part of space-time which are far from the point of interest. Explicit expressions for the anomalies and Schwinger terms have been given in , where the correct expression for the covariant Schwinger term was determined (there are two results in the literature, differing by a sign). The consistent and covariant descent equations will also be compared. The connecting term between consistent and covariant cochains is shown to be a local form on the space of gauge potentials. The paper is organized in the following way: In section 2 and 3 we consider the anomaly in the space-time respective Hamiltonian framework. In section 4 we review the descent equations and compare the consistent and covariant cochains. ## 2 The chiral anomaly in Euclidean space-time We will consider Weyl fermions coupled to an external gauge field $`A𝒜`$ in a $`2n`$-dimensional space-time $`M`$, a smooth, flat, compact and oriented Riemannian spin manifold without boundary. The group $`𝒢`$ of gauge transformations consists of diffeomorphisms $`\phi `$ of a principal bundle $`\pi :P\stackrel{G}{}M`$ such that the base remains unchanged. It acts on the affine space $`𝒜`$ of connections on $`P`$ by pull-back: $`A\phi :=\phi ^{}A`$. To make the action free (so $`𝒜/𝒢`$ will be a smooth manifold) we will assume that $`𝒢`$ only consists of diffeomorphisms that leaves a fixed point $`p_0P`$ unchanged. The generating functional is defined as $$\mathrm{exp}(W(A))=𝑑\psi 𝑑\overline{\psi }\mathrm{exp}(_M\overline{\psi }/_A^+\psi d^{2n}x),$$ where $`W`$ is the effective action and $`/_A^+=/_A(1+\gamma _5)/2=\gamma ^\mu (_\mu +A_\mu )(1+\gamma _5)/2`$. We will use conventions such that $`\gamma ^\mu `$ is hermitian and $`A_\mu `$ is anti-hermitian. The consistent chiral anomaly measures the lack of gauge invariance of the generating functional: $$\delta _X\mathrm{exp}(W(A))=\mathrm{exp}(W(A))\omega (A;X),$$ (1) $`X\text{Lie}𝒢`$. The reason for the minus sign is that the consistent anomaly is defined by $`\delta _XW(A)=\omega (A;X)`$. The effective action is however only defined up to terms $`c`$ which are local functionals in $`𝒜`$. This makes the generating functional defined up to $`\mathrm{exp}(c(A))`$ and the anomaly up to terms $`(\delta _Xc)(A)`$. To make mathematical sense out of the effective action we will first assume that $`/_A^+`$ is a Fredholm operator with zero index. We can then construct the canonical section $`\text{det}i/_A`$ of the determinant line bundle $`\text{DET}i/_A=\text{det ker}i/_A^+(\text{det coker}i/_A^+)^{}`$, see . To identify it with a functional, a reference section $`s_0`$ is chosen. Then we define $`\mathrm{exp}(W(A))`$ as $`\text{det}i/_A/s_0`$. Since all line bundles over an affine space are equivalent, it might appear meaningless to consider the quotient of two sections of a line bundle over $`𝒜`$. However, the determinant line bundle is not just a line bundle, but it is also equipped with a certain structure. For instance, below we will review the fact that the restriction to gauge directions (vectors tangent to the fibres in $`𝒜𝒜/𝒢`$) of its connection (covariant derivative) $``$ can be determined canonically. The variation $`\delta _X`$ will be defined with respect to a fixed $`s_0`$: $$\delta _X\mathrm{exp}(W(A)):=(_X\text{det}i/_A)/s_0.$$ Notice that it in general is not possible to find a section such that $`_Xs_0(A)=0`$ for all $`A𝒜`$. Thus, if we consider a variation at a point $`A^{}`$ which is not related to $`A`$ by a gauge transformation, then a different section $`s_0^{}`$ has to be chosen. Relation (1) is now equivalent with $$_X\text{det}i/_A=\text{det}i/_A\omega (A;X).$$ (2) Observe that this definition of the anomaly does not depend on the choice of $`s_0`$. Recall that a connection on a line bundle satisfies the Leibniz rule $`(s\lambda )=(s)\lambda +sd\lambda `$ for any section $`s`$ and function $`\lambda `$. The connection can be described by a 1-form $`\omega `$ on $`𝒜`$ by $`s=s\omega `$, where $`s`$ is a nowhere vanishing section. We will refer to $`\omega `$ as the pull-back of $``$ with respect to $`s`$. Observe that $`\omega `$ depends on the choice of pull-back: if $`\omega ^{}`$ is defined with respect to $`s^{}=s\lambda `$, then $`\omega ^{}=\omega +\lambda ^1d\lambda `$. For $`\lambda =e^c`$ it gives: $`\omega ^{}(A;X)=\omega (A;X)+(\delta _Xc)(A)`$, where $`X`$ in this case is any vector field on $`𝒜`$. From this abstract discussion it follows that if locality is disregarded, then the anomaly can be identified with minus the restriction to gauge directions of the connection on the determinant line bundle. When locality is taken into account, we obtain the following identification: The anomaly is minus the restriction to gauge directions of the pull-back of the connection with respect to sections given by $`\text{det}i/_A`$ multiplied with the exponential of local terms. The curvature (pulled-back to $`𝒜`$) corresponding to $``$ has been computed in (we will suppress the representation of the gauge group): $$F=c_n_M\text{tr}\left(^{n+1}\right),c_n=2\pi i\frac{1}{(n+1)!}\left(\frac{i}{2\pi }\right)^{n+1}.$$ Here, $`=(d+\delta )\alpha +\alpha ^2`$ is the curvature corresponding to a connection on the universal bundle $`P\times 𝒜M\times 𝒜`$. We thus see that the curvature (and connection) of the determinant line bundle depends on the choice of $`\alpha `$. We will soon see how it is possible to partly determine $`\alpha `$ by physical arguments. This will lead us to an expression for the anomaly. Notice that in order to compute the anomaly from $`F`$ we need to use the fact that the anomaly (the consistent and the covariant) is local. This since there certainly exist forms with $`\delta \omega =F`$ which are non-local even in gauge directions. The Dirac operator acts on sections of an associated bundle to $`PM`$. In the family index theorem one then consider the universal bundle as bundles $`PM`$ parameterized by $`𝒜`$. The first reasonable choice of $`\alpha `$ is then $`\alpha =A`$. In general, the bundle $`P𝒜`$ may twist over $`𝒜`$. This means that $`\alpha `$ is of the form $`A+a`$, where $`a`$ is a 1-form on $`𝒜`$ taking values in $`\text{Lie}G`$. Such a twist occurs for example when demanding gauge invariance, i.e. that $`A+a`$ should descend to a connection on $`(P\times 𝒜)/𝒢M\times 𝒜/𝒢`$, where $`𝒢`$ acts on $`P\times 𝒜`$ by $`(p,A)\phi :=(\phi ^1(p),A\phi )`$, . This is equivalent with that the determinant line bundle can be pushed forward to a line bundle on $`𝒜/𝒢`$ (where it is equipped with a curvature). This does not determine $`a`$ uniquely, but it implies that it is equal to the Faddeev-Popov ghost $`v`$ in gauge directions (one example is the non-local form $`a=(d_A^{}d_A)^1d_A^{}`$). This gives the consistent anomaly. To compute the anomaly, recall that $`\text{tr}\left(^{n+1}\right)\text{tr}\left(^{n+1}\right)`$ $`=`$ $`(d+\delta )\omega _{2n+1}(\alpha ,\alpha ^{})`$ $`\omega _{2n+1}(\alpha ,\alpha ^{})`$ $`=`$ $`(n+1){\displaystyle _0^1}𝑑t\text{tr}\left((\alpha \alpha ^{})_t^n\right)`$ (3) and $`_t`$ the curvature corresponding to $`(1t)\alpha ^{}+t\alpha `$. We will first assume that $`P`$ is trivial so that $`\alpha ^{}=0`$ is a possible choice. With $`\omega `$ the pull-back of $``$, $`F=\delta \omega `$, we get $$\omega =c_n_M\omega _{2n+1}(A+a,0)+\delta \chi ,$$ where $`\chi `$ is an arbitrary functional. Let us now restrict to gauge directions. Recall that the exterior differential $`\delta `$ then is equal to the BRST operator. The first term on the right hand side is now local. By restricting to local $`\chi `$’s, the right hand side becomes minus the consistent anomaly. The Russian formula is the fact that $`=dA+A^2`$ (in gauge directions). It implies that $`F=\delta \omega `$ is zero in gauge directions. This is the Wess-Zumino consistency condition. It implies that it is possible to regard the consistent anomaly as an element in a cohomology group. We will now get rid of two unwanted assumptions, namely that $`\text{ind}/_A^+=0`$ and that $`P`$ is trivial. For this reason we consider the ‘difference’ between two Dirac operators: $`i/_A^++(i/_A^{}^+)^{}=i(/_A^++/_A^{}^{})`$. This is a Dirac operator with zero index (since $`A^{}`$ is a connection on the same bundle $`P`$) and we can thus study the section $`\text{det}i(/_A^++/_A^{}^{})`$ of $`\text{DET}i(/_A^++/_A^{}^{})𝒜\times 𝒜`$, where $`A`$ is in the first factor of $`𝒜\times 𝒜`$ and $`A^{}`$ is in the second. The curvature on $`\text{DET}i(/_A^++/_A^{}^{})`$ is $$FF^{}=c_n_M\text{tr}\left(^{n+1}^{n+1}\right)=(\delta +\delta ^{})c_n_M\omega _{2n+1}(A+a,A^{}+a^{}).$$ (4) Since the consistent anomaly is defined with respect to variations of $`A`$, it is obtained by only considering form parts with respect to the first factor of $`𝒜\times 𝒜`$. With $`F=\delta \omega `$ we then get: $$\omega =c_n_M\omega _{2n+1}(A+a,A^{})+\delta \chi ,$$ where $`\chi `$ is local when restricted to gauge directions. The second determinant bundle has thus only been used as a reference. The gauge field $`A^{}`$ is unaffected by the variations we are interested in. The anomaly above is referred to as the consistent anomaly in the background connection $`A^{}`$, . Expressions obtained from two different background connections differ by a BRST operator acting on a local term and represents therefore the same anomaly. Notice that since we let everything be fixed in the second factor, we do not need to demand gauge invariance of $`\alpha ^{}`$. We may thus set $`a^{}=0`$. The corresponding canonical section and line bundle are then no longer symmetric in $`A`$ and $`A^{}`$ and we will for this reason denote them by $`\widehat{\text{det}}i(/_A^++/_A^{}^{})`$ respective $`\widehat{\text{DET}}i(/_A^++/_A^{}^{})`$. Notice that since $`A^{}`$ is regarded as fixed, $`\widehat{\text{DET}}i(/_A^++/_A^{}^{})`$ can be considered as a line bundle over $`𝒜`$ (the first factor of $`𝒜\times 𝒜`$) with section $`\widehat{\text{det}}i(/_A^++/_A^{}^{})`$. Let us make a comment on the line bundle $`\text{DET}i/_A^+(\text{DET}i/_A^{}^+)^{}`$. Despite the fact that it has the same curvature as $`\text{DET}i(/_A^++/_A^{}^{})`$, it is equipped with a different structure. For instance, it is easy to see that its connection evaluated on the canonical section $`\text{det}i/_A^+(\text{det}i/_A^{}^+)^{}`$ gives a different result than for $`\text{det}i(/_A^++/_A^{}^{})`$. In the first case one obtains a form which consist of two separate pieces on the first respective second factor of $`𝒜\times 𝒜`$. However, in the second case we saw in the expression for the consistent anomaly that the background connection also plays a role in the first factor. This is an example of the multiplicative anomaly (the determinant of a product is not necessary equal to the product of the determinants for the factors) in the family case, i.e. $`\text{DET}i/_A^+\text{DET}(i/_A^{}^+)^{}`$ $`=`$ $`\text{DET}\left(\begin{array}{cc}0& 1\\ i/_A^+& 0\end{array}\right)\text{DET}\left(\begin{array}{cc}0& i/_A^{}^{}\\ 1& 0\end{array}\right)`$ $``$ $`\text{DET}\left(\begin{array}{cc}0& i/_A^{}^{}\\ i/_A^+& 0\end{array}\right)=\text{DET}i(/_A^++/_A^{}^{}),`$ where we with the (non-)equality mean with respect to line bundles with structures. Consider now $`\widehat{\text{DET}}i(/_A^++/_A^{}^{})`$ and $`\widehat{\text{det}}i(/_A^++/_A^{}^{})`$ over the diagonal in $`𝒜\times 𝒜`$, i.e. we put $`A^{}=A`$. The curvature then looks as in eq. (4) with $`a^{}=0`$ and $`A^{}=A`$. The operator $`\delta ^{}`$ acts as $`\delta `$ on $`A^{}=A`$ in the second argument of $`\omega _{2n+1}`$ while it leaves $`A`$ and $`a`$ in the first argument unchanged. We then see that if we identify the diagonal with $`𝒜`$, then the bundle has the curvature $$F=\delta c_n_M\omega _{2n+1}(A+a,A),$$ (7) where $`\delta `$ now acts on the second argument as well. The corresponding connection $$\omega =c_n_M\omega _{2n+1}(A+a,A)+\delta \chi $$ is recognized as (minus) the covariant anomaly if restricted to gauge directions (where $`\chi `$ is local). It is now clear that also in the covariant formalism can the anomaly be defined by eq. (2) and the discussion that followed this equation, however, the connection and canonical section are defined with respect to $`\widehat{\text{DET}}i(/_A^++/_A^{})`$. The fact that the curvature is non-zero in gauge directions means that the Wess-Zumino consistency condition is not fulfilled. It also implies that the covariant anomaly can not be obtained from a variation of a functional. For instance, since the curvature is non-zero also in gauge directions it is impossible to find a reference section satisfying $`_Xs_0=0`$ at a point $`A𝒜`$. ## 3 The chiral anomaly in the Hamiltonian formalism We will here let $`M`$ be $`(2n1)`$-dimensional and interpreted as the physical space at a fixed time. $`𝒢`$ and $`𝒜`$ will be the corresponding group of gauge transformations respective space of potentials. The chiral anomaly manifest itself as the Schwinger term in the Hamiltonian formalism. We will take the starting point that the consistent Schwinger term is minus the curvature of the vacuum bundle restricted to gauge directions. To compute it we must therefore understand the vacuum first. The Hamiltonian $`H_A`$ decomposes the 1-particle Hilbert space into two pieces, depending on if the eigenvalues are bigger or less than some number $`\lambda `$: $$H=H_+(A,\lambda )H_{}(A,\lambda ).$$ The vacuum bundle $`\mathrm{\Omega }_\lambda (A)`$ is given by the filled up Dirac sea to a certain level $`\lambda `$. It is thus the complex span of the wedge product of the eigenvectors in $`H_{}(A,\lambda )`$. Naively, this is a line bundle over $`U_\lambda =\{A𝒜|\lambda |\text{spec}(H_A)\}`$. However, the vacuum bundle is not well-defined since it involves an infinite wedge product. Instead, we will consider the ‘quotient’ of two vacua. For this, consider a decomposition of the 1-particle space with respect to different $`A`$ and $`\lambda `$: $`H=H_+(A^{},\lambda ^{})H_{}(A^{},\lambda ^{})`$. The intuitive definition of the quotient of $`\mathrm{\Omega }_\lambda (A)`$ and $`\mathrm{\Omega }_\lambda (A^{})`$ is then $`\text{det}H_{}(A,\lambda )(\text{det}H_{}(A^{},\lambda ^{}))^{}`$, where det is the top wedge product. This ill-defined expression becomes simpler if we quote out a common infinite part from the first and the second factor. The quotient of the two vacuum bundles can then be defined as $`(\text{det}H_{}(A,\lambda )H_+(A^{},\lambda ^{}))(\text{det}H_+(A,\lambda )H_{}(A^{},\lambda ^{}))^{}`$. However, this is still not well-defined since the intersections are infinite dimensional spaces. Fortunately there exists renormalization methods to handle this problem. In 1 space dimension it is particular simple since the identity operator on $`H`$ becomes a Hilbert-Schmidt operator when restricted to the domain $`H_{}(A,\lambda )`$ ($`H_+(A,\lambda )`$) and the range $`H_+(A^{},\lambda ^{})`$ ($`H_{}(A^{},\lambda ^{})`$). Although the above definition of the quotient of two vacua is the most common used, we will use an alternative definition from . Instead of using the identity operator on $`H`$ to compare $`H_{}(A,\lambda )`$ and $`H_{}(A^{},\lambda ^{})`$, the Dirac operator will be used. For this, let $`A(t)`$, $`tI=[0,1]`$ be a path in $`𝒜`$ from $`A^{}`$ to $`A`$. It is often useful to consider $`t`$ as the time. By letting the wave functions $`\psi `$ depend on the extra parameter $`t`$, we see that the Dirac equation $`i/_{A(t)}^+\psi =0`$, $`i/_{A(t)}^+=i_tH_{A(t)}`$, can be used to identify the two differently composed Hilbert spaces. Then $`\text{DET}_{\lambda \lambda ^{}}(A(t)):=\text{det ker}i/_{A(t)}^+(\text{det coker}i/_{A(t)}^+)^{}`$ is a reasonable definition of $`\mathrm{\Omega }_\lambda (A)\mathrm{\Omega }_\lambda (A^{})^{}`$ if the boundary conditions $`\psi |_{t=0}H_+(A^{},\lambda ^{})`$ and $`\psi |_{t=1}H_{}(A,\lambda )`$ are used. This gives a line bundle over $`U_\lambda \times U_\lambda ^{}𝒜\times 𝒜`$ in a similar way as for the anomaly in the space-time formalism. Its curvature is $$F=c_n\left(_{M\times I}\text{tr}\left(^{n+1}\right)\frac{1}{2}\left(\widehat{\eta }_\lambda \right)_{[2]}+\frac{1}{2}\left(\widehat{\eta }_\lambda ^{}\right)_{[2]}\right),$$ (8) where $``$ is the curvature corresponding to a connection $`\alpha `$ on the universal bundle $`P\times I\times (U_\lambda \times U_\lambda ^{})M\times I\times (U_\lambda \times U_\lambda ^{})`$. The fact that $`M\times I`$ is a manifold with boundary gives rise to additional contributions, the 2-form parts of the $`\widehat{\eta }`$-forms, . Let us now compute the consistent Schwinger term from the above results. We thus want to compute $`F`$ when gauge variations of $`A=A(t=1)`$ is made. For this reason, we chose the simplest possible path, namely $`A(t)=(1t)A^{}+tA`$ (actually, in order to use theorems about determinant line bundles, $`t`$ should be replaced with a function $`f(t)`$ which is 0 in a neighbourhood of $`t=0`$ and 1 in a neighbourhood of $`t=1`$, ). Let us now regard $`A^{}`$ as a background connection. Then $`\text{DET}_{\lambda \lambda ^{}}(A(t))`$ is a line bundle over $`U_\lambda 𝒜`$. Clearly, the choice $`\alpha =(1t)A^{}+t(A+a)`$ should be made, where $`a`$ is equal to the ghost in gauge directions. It implies that $`=(d+d_t+\delta )((1t)A^{}+t(A+a))+((1t)A^{}+t(A+a))^2`$. By dimensional reasons, only one $`d_t`$ term can appear in $`_{M\times I}\text{tr}\left(^{n+1}\right)`$. It gives $$_{M\times I}\text{tr}\left(^{n+1}\right)=_M\omega _{2n+1}(A+a,A^{}).$$ (9) To compute the consistent Schwinger term, we now restrict eq. (8) to gauge directions. By construction, the $`\widehat{\eta }`$-forms vanishes then, . Thus, the consistent Schwinger term in a background connection $`A^{}`$ is given by minus the restriction of eq. (9) to gauge directions. We have shown that the vacuum bundle can be defined as a determinant bundle. As in the space-time formalism it is not the bundles themselves that are interesting on the level of anomalies, but the fact that the determinant line bundle is equipped with a natural connection in gauge directions. In fact, regarding Schwinger terms, the connection is only defined up to local 1-forms $`\chi `$ on $`𝒜`$. This makes the Schwinger term defined up to terms $`\delta \chi `$. The consistent Schwinger term is thus the curvature corresponding to the canonical connection, up to local forms, on the vacuum bundle. The relation $`\delta F=0`$, due to the Bianchi identity, gives the consistency condition for the Schwinger term. This can be seen directly from the fact that $`\delta {\displaystyle _M}\omega _{2n+1}(A+v,A^{})`$ $`=`$ $`{\displaystyle _M}(d+\delta )\omega _{2n+1}(A+v,A^{})`$ $`=`$ $`{\displaystyle _M}\left(\text{tr}\left(dA+A^2\right)^{n+1}\text{tr}\left(dA^{}+A^2\right)^{n+1}\right)=0`$ holds in gauge directions. As for the consistent anomaly, it implies that the consistent Schwinger term can be regarded as an element in a cohomology group. If $`t`$ is interpreted as time, then the consistent Schwinger term has been computed by comparing the vacuum bundle at the time of interest ($`t=1`$) with a reference vacuum bundle at another time ($`t=0`$). The reference vacuum is given by the Hamiltonian $`H_A^{}`$. Certainly, it is possible to use the field $`A`$ itself as a background, i.e. as for the anomaly we now put $`A^{}=A`$. In this case eq. (9) gives the covariant Schwinger term. Thus, the figure in the previous section illustrates as well the case of Schwinger terms (except for the fact that the line bundle over (a subset of) $`𝒜\times 𝒜`$ is different). That both the consistent and covariant anomaly are given by connections suggests that both the consistent and covariant Schwinger term are given by curvatures. For the covariant Schwinger term this is however not true in general. This is easy to understand since all curvatures of line bundles are closed by the Bianchi identity. Thus, if the covariant Schwinger term could have been obtained from a curvature, then it would also have to satisfy the consistency condition. This is however not true. The reason why the covariant Schwinger term in general is not obtained from a curvature is that the third term on the right hand side of eq. (8) is not necessary zero in gauge directions for $`A^{}=A`$. This term is determined by the boundary conditions at the manifold $`M\times \{t=0\}`$. Since we are interested in the other part $`M\times \{t=1\}`$ of the boundary when computing the Schwinger term, it is reasonable to not take into account the contribution from $`\widehat{\eta }_\lambda ^{}`$. We thus only pay attention to how the connection of the universal bundle at $`M\times \{t=1\}`$ is extended to the bulk and not to terms coming from data on the other part of the boundary. Let us give an alternative description, intended for readers unfamiliar with the $`\widehat{\eta }`$-forms. Extend $`M`$ to a manifold which have $`M`$ as its boundary and looks like a cylinder close to $`M`$, i.e. the extension is $`M\times I\stackrel{~}{M}`$, where $`\stackrel{~}{M}=M\times \{t=0\}`$. Then extend the gauge connections and gauge transformations in a smooth way to the bulk. The curvature of the corresponding bundle is $`F+\stackrel{~}{F}`$, where $`F`$ is given by eq. (8) and $`\stackrel{~}{F}`$ is defined with respect to $`\stackrel{~}{M}`$. If we consider the gauge connection $`A(t=0)`$ as a fixed background field $`A^{}`$ which doesn’t depend on $`A=A(t=1)`$, then $`\stackrel{~}{F}`$ is zero and the consistent anomaly is obtained. If $`A(t=0)=A(t=1)`$, then $$\stackrel{~}{F}=c_n\left(_{\stackrel{~}{M}}\text{tr}\left(\stackrel{~}{}^{n+1}\right)\frac{1}{2}(\widehat{\eta }_\lambda ^{})_{[2]}\right)$$ where $`\stackrel{~}{}=(d+\delta )A+A^2`$. It gives $$F+\stackrel{~}{F}=c_n\left(_{M\times I}\text{tr}\left(^{n+1}\right)+_{\stackrel{~}{M}}\text{tr}\left(\stackrel{~}{}^{n+1}\right)\frac{1}{2}(\widehat{\eta }_\lambda )_{[2]}\right).$$ As in the consistent case, $`\widehat{\eta }_\lambda `$ gives zero in gauge directions. The contribution from the second term comes from a manifold which is far from the manifold $`M\times \{t=1\}`$ of interest. For this reason, it can be disregarded. The same is true for the second term on the left hand side. The literature about covariant Schwinger terms can be divided into two parts depending on if they define it by an algebraic argument or by a physical argument, see for instance . The first discipline have a lack of physical understanding while the mathematical structure is unclear in the second. In this paper, on the other hand, a mathematical rigorous construction of an underlying geometrical and physical idea has been made. It is therefore not sound to use earlier results as a judge if the non-local $`\widehat{\eta }`$-form will give a contribution or not. Indeed, all the previous computations of the covariant Schwinger term seems already from the beginning to neglect the possibility of contributions from terms as the non-local $`\widehat{\eta }`$-form. From this point of view it seems more natural to define the covariant Schwinger term as the curvature above. We then obtain a non-local covariant Schwinger term which differs from earlier results by $`\frac{1}{2}(\widehat{\eta }_\lambda ^{})_{[2]}`$. However, since everything comes down to the definition we will for the rest of the paper disregard the $`\widehat{\eta }`$-form so that our result agrees with the previous ones. Notice that the discussion above also makes sense for the anomaly in the space-time formalism: With the mathematical methods of section 2 we obtain a contribution from an $`\widehat{\eta }`$-form when the space-time has a boundary. ## 4 Descent equations and connecting terms Let $`\omega _{2n+1k}^k(A+v,A^{})`$ denote the part of $`\omega _{2n+1}(A+v,A^{})`$ which have ghost degree $`k`$. Expansion in ghost degrees of the first expression in (2) with $`\alpha =A+v`$ and $`\alpha ^{}=A^{}`$ gives the consistent descent equations: $`\text{tr}\left(dA+A^2\right)^{n+1}\text{tr}\left(dA^{}+A^2\right)^{n+1}`$ $`=`$ $`d\omega _{2n+1}^0(A+v,A^{})`$ $`\delta \omega _{2n+1k}^k(A+v,A^{})`$ $`=`$ $`d\omega _{2nk}^{k+1}(A+v,A^{}),`$ $`k=0,1,\mathrm{},2n`$ $`\omega _0^{2n+1}(A+v,A^{})=0,`$ where it has been used that $`=dA+A^2`$ in gauge directions. From the previous sections it follows that the (non-integrated) anomaly and Schwinger term in a background $`A^{}`$ is $`c_n\omega _{2n}^1(A+v,A^{})`$ respective $`c_n\omega _{2n1}^2(A+v,A^{})`$. We thus see that the descent equations can be used for their computation. Further, it gives a relation between them. By integrating over $`M`$ it is also seen that the consistency condition is fulfilled. When $`A^{}=0`$ we obtain the ‘ordinary’ descent equations. It is not so well-known that there exist descent equations for the covariant anomaly and Schwinger term as well. The first relation in (2) with $`\alpha =A+v`$ and $`\alpha ^{}=A`$ gives the covariant descent equations : $`0`$ $`=`$ $`\omega _{2n+1}^0(A+v,A)`$ $`\delta \omega _{2n+1k}^k(A+v,A)`$ $`=`$ $`d\omega _{2nk}^{k+1}(A+v,A)`$ $`\left(\begin{array}{c}n+1\\ k+1\end{array}\right)\text{str}\left((\delta A)^{k+1}(dA+A^2)^{nk}\right),`$ $`k=0,1,\mathrm{},n`$ $`\delta \omega _{2n+1k}^k(A+v,A)`$ $`=`$ $`d\omega _{2nk}^{k+1}(A+v,A),k=n+1,\mathrm{},2n`$ $`\delta \omega _0^{2n+1}(A+v,A)`$ $`=`$ $`0,`$ where str is the symmetrized trace. This gives a computational scheme for the covariant anomaly and Schwinger term. However, as for the consistent case, the direct formula, eq. (2), is to prefer. Let us give the explicit expressions in the descent equations for the simplest case $`n=1`$: $`\delta 2\text{tr}\left(v(dA+A^2)\right)`$ $`=`$ $`d\text{tr}(v\delta A)\text{tr}(\delta A)^2`$ $`\delta \text{tr}(v\delta A)`$ $`=`$ $`d\left({\displaystyle \frac{1}{3}}\text{tr}v^3\right)`$ $`\delta \left({\displaystyle \frac{1}{3}}\text{tr}v^3\right)`$ $`=`$ $`0,`$ The triangle formula implies that $`\omega _{2n+1k}^k(A+v,A)`$ is up to trivial terms (which in the non-integrated setting takes the form $`(d+\delta )\chi `$) equal to $`\omega _{2n+1k}^k(A+v,A^{})\omega _{2n+1k}^k(A,A^{})`$. The difference between the consistent and covariant cochains is thus given by $`\omega _{2n+1k}^k(A,A^{})`$. When $`kn+1`$ this term is zero and the consistent and covariant cochains are equal. Certainly, this is in agreement with the fact that the consistent and covariant descent equations are equal in this case. Thus, half of the cochains are always equal. It explains the fact that the consistent Schwinger term is covariant when $`n=1`$, while this is not true for $`n2`$. The connecting term $`\omega _{2n+1k}^k(A,A^{})`$ has a very interesting property. It contains only the ghost in the combination $`\delta A`$. This implies that it can be extended to a local form on all of $`𝒜`$, not necessary in gauge directions. (With local we mean that it can be expressed as a trace of a polynomial in $`A`$, $`dA`$ and $`\delta A`$.) This is in contrary to $`\omega _{2n+1k}^k(A+v,A^{})`$ and $`\omega _{2n+1k}^k(A+v,A)`$ since if the ghost is extended to all of $`𝒜`$, for instance to $`(d_A^{}d_A)^1d_A^{}`$, then a non-local expression is obtained. That the covariant cochains are not in any cohomology group has sometimes been regarded as a ‘drawback’ of the covariant formalism. This is not really true. In fact, the covariant cochains can be regarded as elements in a cohomology group due to their bijective correspondence with such elements, namely the consistent cochains. In general, since they are in a one-to-one relation, the consistent and covariant terms are always on equal footing. Another example where this can be used is in the fact that the consistent anomaly and Schwinger term can be pushed forward to $`𝒜/𝒢`$. These geometrical objects on line bundles over (subsets of) $`𝒜`$ becomes topological on $`𝒜/𝒢`$. That this is not possible for the corresponding covariant terms follows from the fact that $`\widehat{\text{DET}}i(/_A^++/_A^{}^{})`$ can be pushed forward to $`𝒜/𝒢`$ while this is not true for $`\widehat{\text{DET}}i(/_A^++/_A^{})`$. We would also like to point out that the chiral anomaly and the Schwinger term (and the other cochains) can appear in different forms than just the consistent and covariant. In general, the connection $`A^{}`$ can be separated into a piece $`f(A)`$ which is independent of $`A`$ and a piece $`h(A)`$ which depends on $`A`$. The extreme case $`h(A)=A^{}`$, $`f(A)=0`$ gives the consistent cochains while $`h(A)=0`$, $`f(A)=A`$ gives the covariant cochains. Acknowledgments: I thank Dr C. Adam for an interesting discussion.
warning/0002/cond-mat0002056.html
ar5iv
text
# Fermi systems with long scattering lengths ## I Introduction Dilute degenerate Fermi systems with long scattering lengths are of interest for nuclear and neutron star matter (see, e.g., ). Recently, also dilute systems of cold Fermionic atoms have been trapped . The number density is sufficient for degeneracy to be observed and superfluidity is expected at critical temperatures similar to the onset of Bose-Einstein condensates, $`10100`$nK. S-wave scattering lengths can be very long, e.g.<sup>*</sup><sup>*</sup>*The sign convention of negative scattering length for attractive potentials is used. Also $`\mathrm{}=c=1`$. $`a=2160`$ Bohr radii for triplet $`{}_{}{}^{6}Li`$ and $`a18.8`$ fm for neutron-neutron interactions. These scattering lengths $`|a|`$ are much longer than typical range of the potentials, respectively $`R1`$ fm for strong interactions and $`R10100`$ Å for van der Waals forces. Generally, when $`|a|R`$ three density regimes naturally emerges: the low density (or dilute, $`k_F^1\stackrel{>}{}|a|`$), the high density ($`k_F^1\stackrel{<}{}R`$), and the intermediate density region ($`R\stackrel{<}{}k_F^1\stackrel{<}{}|a|`$). The latter “novel” region of densities is the object of study here. It will be shown that previous conjectures based on extrapolations from the dilute limit fail. Instead it is found that both the energy per particle and superfluid gaps scale with the Fermi energy. They depend only on statistical factors but not on the scattering length, range or other details of the interaction. Consequently, phase diagrams are dramatically altered and the stability criteria differ so that two spin systems are actually stable. The manuscript is organized as follows. In Sec. II the scaling and poles of the effective scattering amplitude are studied in a homogeneous many-body system going from dilute to intermediate densities. In Sec. III the ground state energies are calculated at intermediate densities and compared to well known results from the dilute and high density limits. In Sec. IV extensions to finite temperatures are discussed and a phase diagram is constructed displaying regions of superfluidity and spinodal instabilities. In Sec. V the properties of finite systems of Fermions are investigated as they are relevant for current experiments with magnetically trapped cold atoms. Finally, a summary and conclusion is given. ## II The effective scattering amplitude and superfluidity Consider a homogeneous many-body system of fermions of mass $`m`$ at density: $`\rho =\nu k_F^3/6\pi ^2`$, where $`\nu `$ is the statistical factor, e.g. $`\nu =4`$ for symmetric nuclear matter and $`\nu =2`$ for neutron matter as well as for the $`{}_{}{}^{40}K`$ atomic gas with two hyperspin states currently studied at JILA . The scattering lengths and Fermi momentum $`k_F`$ are assumed the same for all spin-isospin/hyperspin components in the system but interesting effects of varying the relative densities of the various components will be discussed at the end. Particles are assumed non-relativistic and to interact through attractive two-body contact interactions. The details of the potential is not important, only its range $`R`$ and scattering length $`a`$. We shall be particular interested in cases where $`R|a|`$, which occur when, for example, the two-body potential can almost support a bound state or resonance. At dilute or intermediate densities the particles interact via short range interactions that appear singular on length scales of order the interparticle distance $`k_F^1`$. Such systems are best described by resumming the multiple interactions in terms of the scattering amplitude. The Galitskii’s integral equations for the effective two-particle interaction or scattering amplitude in the medium is given by the ladder resummation $`\mathrm{\Gamma }(𝐩,𝐩^{},𝐏)`$ $`=`$ $`\mathrm{\Gamma }_0(𝐩,𝐩^{},𝐏)+m{\displaystyle \underset{𝐤}{}}\mathrm{\Gamma }_0(𝐩,𝐤,𝐏)`$ (1) $`\times `$ $`\left[{\displaystyle \frac{N(𝐏,𝐤)}{\kappa ^2k^2}}{\displaystyle \frac{1}{\kappa ^2k^2}}\right]\mathrm{\Gamma }(𝐤,𝐩^{},𝐏).`$ (2) Here, $`\mathrm{\Gamma }_0=4\pi a/m`$ is the s-wave scattering amplitude in vacuum; the total energy of the pair in the center of mass is $`\kappa ^2/m`$; $`𝐩,𝐤,𝐩^{}`$ are the relative momentum of the pair of interacting particles in the initial, intermediate and final states respectively, and $`𝐏`$ the total momentum; $`N(𝐏,𝐤)=+1`$ for particle-particle propagation ($`|𝐏\pm 𝐤|k_F`$), $`N(𝐏,𝐤)=1`$ for hole-hole propagation ($`|𝐏\pm 𝐤|k_F`$), and zero otherwise. For spin independent interactions the amplitudes contain a factor $`(1\delta _{\nu _1\nu _2})`$ that takes exchange into account between identical spins $`\nu _1=\nu _2`$. For obtaining BCS gaps it is sufficient to study pairs with $`𝐏=0`$ where Eq. (2) is simply $`\mathrm{\Gamma }=\mathrm{\Gamma }_0+2m{\displaystyle \underset{kk_F}{}}\mathrm{\Gamma }_0{\displaystyle \frac{1}{\kappa ^2k^2}}\mathrm{\Gamma }.`$ (3) Note the factor of 2 due to particle-particle and hole-hole propagation each contributing by the same amount in non-dense systems. The ladder resummation implicit in Eq. (2) insures that only momenta smaller than Fermi momenta enter. $`\mathrm{\Gamma }_0`$ vary on momentum scales $`1/Rk_F`$ only and can therefore be considered constant at low and intermediate densities. Eq. (3) is then easily solved for momenta near the Fermi surface $`\mathrm{\Gamma }`$ $`=`$ $`\mathrm{\Gamma }_0\left[1{\displaystyle \frac{2}{\pi }}k_Fa(2+{\displaystyle \frac{\kappa }{k_F}}\mathrm{ln}{\displaystyle \frac{k_F\kappa }{k_F+\kappa }})\right]^1.`$ (4) The in-medium scattering amplitude has a pole due to Cooper pairing when $`\mathrm{\Delta }`$ $``$ $`{\displaystyle \frac{k_F^2\kappa ^2}{m}}={\displaystyle \frac{(k_F+\kappa )^2}{m}}\mathrm{exp}({\displaystyle \frac{\pi }{2\kappa a}}2{\displaystyle \frac{k_F}{\kappa }})`$ (5) $``$ $`E_F{\displaystyle \frac{8}{e^2}}\mathrm{exp}({\displaystyle \frac{\pi }{2k_Fa}}),k_F|a|1,`$ (6) where $`E_F=k_F^2/2m`$ is the Fermi energy. The critical temperature is $`T_c=(\gamma /\pi )\mathrm{\Delta }`$, where $`\gamma =e^C`$ and $`C=0.577`$ is Euler’s constant. Eq. (6) is the BCS gap in the dilute limit which agrees with gaps calculated in . However, Gorkov et al. showed that spin fluctuations lead to a higher order correction $`(k_Fa)^2`$ in the denominator of Eq. (4) that is amplified by logarithmic terms $`\mathrm{ln}(\mathrm{\Delta })1/k_Fa`$. It contributes by a (negative) constant in the exponent and leads to a reduction of the gap in the dilute limit by a factor $`(4e)^{1/3}=2.215\mathrm{}`$ as compared to Eq. (6) for two spins. Generally for $`\nu `$ spins, isospins or hyperspins the gap is $`\mathrm{\Delta }=E_F{\displaystyle \frac{8}{e^2}}(4e)^{\nu /31}\mathrm{exp}\left[{\displaystyle \frac{\pi }{2ak_F}}\right].`$ (7) In the intermediate density region pairing must still occur since interaction are attractive. The validity of the expressions of Eqs. (6,7) in this density regime will be discussed further below. They predict that in the limits $`a\mathrm{}`$ and $`R0`$ the gap cannot depend on either $`a`$, $`R`$ or other details of the potential. For dimensional reasons the gap can therefore only be proportional to the Fermi energy. ## III Ground state energies The ground state energy is another crucial property of the system. In terms of the on-shell effective scattering amplitude it is $`E`$ $`=`$ $`{\displaystyle \underset{k_1\nu _1}{}}{\displaystyle \frac{k_1^2}{2m}}+{\displaystyle \frac{1}{2}}{\displaystyle \underset{k_1k_2\nu _1\nu _2}{}}\mathrm{\Gamma }(𝐩,𝐩,𝐏)(1\delta _{\nu _1\nu _2})`$ (8) $`=`$ $`{\displaystyle \frac{3}{5}}{\displaystyle \frac{k_F^2}{2m}}N+{\displaystyle \frac{\nu (\nu 1)}{2}}{\displaystyle \underset{k_1k_2}{}}\mathrm{\Gamma }(𝐩,𝐩,𝐏).`$ (9) Here, $`N=V\rho `$ is the number of particles and the summations $`\nu _1,\nu _2`$ include spin and isospin or hyperspin states. The factor $`(1\delta _{\nu _1\nu _2})`$ in the amplitude due to exchange has now been written explicitly. Antisymmetrization of the wave-function prevents identical particles to be in relative s-states. At low and intermediate densities, $`k_FR1`$, the exchange term is $`1/\nu `$ of the direct one for spin independent interactions. Before investigating the novel intermediate density region, a brief review of results at low and high densities is given. ### A Low densities (dilute): $`k_F|a|1`$ At low densities, $`k_F|a|1`$, gaps are small and have little effect on the total energy of the system. Expanding the effective scattering amplitude of Eq. (2) in the small quantity $`k_Fa`$, the energy per particle is obtained from Eq. (9) by summing over momenta of the two interacting particles $`{\displaystyle \frac{E}{N}}`$ $`=`$ $`E_F[{\displaystyle \frac{3}{5}}+(\nu 1){\displaystyle \frac{2}{3\pi }}k_Fa`$ (10) $`+`$ $`(\nu 1){\displaystyle \frac{4(112\mathrm{ln}2)}{35\pi ^2}}(k_Fa)^2+𝒪((k_Fa)^3)].`$ (11) It consists of respectively the Fermi kinetic energy, the standard dilute pseudo-potential proportional to the scattering length and density, and orders $`(k_Fa)^2`$ and higher . The zero temperature hydrodynamic sound speed squared can at low temperatures be expressed as $`s^2={\displaystyle \frac{1}{m}}\left({\displaystyle \frac{P}{\rho }}\right)={\displaystyle \frac{1}{m}}{\displaystyle \frac{}{\rho }}\left(\rho ^2{\displaystyle \frac{E/N}{\rho }}\right),`$ (12) With the energy per particle of Eq. (11) at low densities, the sound speed can be expanded as $`s^2={\displaystyle \frac{1}{3}}v_F^2\left(1+{\displaystyle \frac{2}{\pi }}(\nu 1)k_Fa+\mathrm{}\right).`$ (13) where $`v_F=k_F/m`$ is the Fermi velocity. It is commonly conjectured from the first two orders that the Fermion (and Bose) gases undergo spinodal instability when the sound speed squared becomes negative, which occurs when $`k_Fa\stackrel{<}{}\pi /2(\nu 1)`$. However, at the same densities the dilute approximation leading to Eq. (11) fails and so does the conjecture as will be shown below. ### B High densities, $`k_FR1`$ At high densities, $`k_FR\stackrel{>}{}1`$, the particle potentials overlap and each particle experience on average the volume integral of the potentials. The energy per nucleon consists of the Fermi kinetic energy and the Hartree-Fock potential, (see, e.g. Eq. 40.14): $`{\displaystyle \frac{E_{HF}}{N}}`$ $`=`$ $`{\displaystyle \frac{3}{5}}E_F+{\displaystyle \frac{\rho }{2}}{\displaystyle d^3rV(r)\left[1\frac{1}{\nu }\left(\frac{3j_1(rk_F)}{rk_F}\right)^2\right]}.`$ (14) The latter term is the exchange energy which vanishes at very high densities, $`k_FRR/r_01`$, leaving the Hartree potential term only. At lower densities, $`k_FRR/r_01`$, it is identical to the first integral, i.e., the Hartree direct term is $`\nu `$ times the Fock exchange term as also found in the dilute limit, Eqs. (15-11). As shown in , the Hartree potential is considerably less attractive than the dilute potential. In fact it vanishes when the range of the interaction goes to zero and the scattering length to infinity. Take for example a square well potential of range $`R`$ and depth $`V_0`$. Long scattering lengths requires $`V_0R^2\pi ^2/4m`$, and therefore the Hartree potential is $`V_0R^3R0`$. Only in the Born approximation do the Hartree (14) and dilute potentials (11) coincide since the Born scattering length is $`a_{Born}=(m/4\pi )d^3rV(r)`$. Short range repulsion complicate the high density limit. In nuclear and atomic system the repulsive core is only of slightly shorter range than the attractive force. It makes the liquid strongly correlated and the Hartree-Fock approximation fails . How the short range repulsion turn the attraction to repulsion at these even higher densities will, however, not affect the intermediate density region. ### C Intermediate densities, $`|a|k_F^1R`$ At intermediate densities, $`|a|k_F^1R`$, the scattering length expansion in Eq. (11) breaks down. Brueckner, Bethe and Goldstone pioneered such studies for nuclear matter and $`{}_{}{}^{3}He`$ where the range of interactions, scattering lengths and repulsive cores all are comparable in magnitude. In our case the range of interaction is small, $`k_FR1`$, and therefore all particle-hole diagrams are negligible. Higher order particle-particle and hole-hole diagrams do contribute by orders of $`\mathrm{\Gamma }_0(mk_F\mathrm{\Gamma })^n`$, It is evident from Eq. (6) that at intermediate densities $`\mathrm{\Gamma }`$ no longer is proportional to $`\mathrm{\Gamma }_0`$ or the scattering length but instead $`\mathrm{\Gamma }(mk_F)^1`$. Due to the very restricted phase space such higher order terms are usually neglected as in standard Brueckner theory. $`\mathrm{\Gamma }`$ of Eq. (2) can therefore be considered as a resummation of an important class of diagrams. The Cooper instability complicates the calculation of $`\mathrm{\Gamma }`$. If the gap is small the instability occurs only for pairs near the Fermi system with opposite momenta and spin and the effect on the total energy is small. The momentum dependence of the effective scattering amplitude also complicates a self-consistent calculation. These complications can be dealt with by approximating $`\mathrm{\Gamma }`$ by its momentum average value in Eq. (2). The momentum integrals are then analogous to those in the dilute limit (11), and one obtains from Eq. (9) $`{\displaystyle \frac{E}{N}}E_F\left[{\displaystyle \frac{3}{5}}+{\displaystyle \frac{(\nu 1)\frac{2}{3\pi }k_Fa}{1\frac{6}{35\pi }(112\mathrm{ln}2)k_Fa}}\right].`$ (15) This expression is valid for dilute systems, where it reproduces Eq. (11), and approximately valid within the Galitskii ladder resummation at intermediate densities, $`Rk_F^1|a|`$, where it reduces to $`{\displaystyle \frac{E}{N}}=E_F\left[{\displaystyle \frac{3}{5}}(\nu 1)c_1\right]=E_Fc_1(\nu _c\nu ),`$ (16) with $`c_1=35/9(112\mathrm{ln}2)0.40`$ and $`\nu _c=1+3/5c_12.5`$. Both the attractive and the kinetic part of the energy per particle are proportional to the Fermi energy at these intermediate energies as found for the gaps above. The other remarkable feature of Eq. (16) is that the energy per particle changes sign for a critical number of degrees of freedom, $`\nu _c2.5`$. Fermi systems with more degrees of freedom such as symmetric nuclear matter have negative energy per particle and will therefore implode, undergo spinodal decomposition and fragment . Contrarily, systems with $`\nu \stackrel{<}{}\nu _c`$ such as neutron matter and atomic systems with only two hyperspins have positive energy per particle and will therefore explode, if not contained. This is also evident from the sound speed squared which from Eqs. (12) and (16) becomes $`s^2=(5/9)c_1(\nu _c\nu )v_F^2.`$ (17) Calculations for pure neutron matter and symmetric nuclear matter at low densities by variational Monte Carlo and in neutron matter by Pade’ approximants to R-matrix calculations independently confirm the above results approximately in a limited range of intermediate densities. In the density range $`\rho _0\stackrel{>}{}\rho \stackrel{>}{}|a|^310^4\rho _0`$ the energy per particle is positive for neutron matter and negative for symmetric nuclear matter . They scale approximately with $`\rho ^{2/3}`$ with coefficients compatible with Eq. (16). In symmetric nuclear matter the intermediate density regime is, however, limited since protons also interact through the triplet channel, which has a shorter repulsive scattering length $`a_t5.4`$ fm, besides the singlet one, $`a_s=18.8`$ fm, relevant for neutron matter. Never-the-less, the ladder resummation in the Galitskii integral equation Eqs. (15,16) are supported by dimensional arguments and quantitatively it successfully predicts $`\nu _c`$ between that of neutron and symmetric nuclear matter. The ladder resummation therefore seems to include the most important class of diagrams. However, even small corrections can be important for the magnitude of the gap because they appear in the exponent as, e.g., found for induced interaction (compare Eq. (7) with (6)). In addition, the superfluidity decrease the energy of the system by $`\mathrm{\Delta }^2/2E_F`$, which can be significant at intermediate densities if gaps really are as large as the Fermi energy. The interesting feature of the intermediate density region, that energies and gaps are independent of the scattering length, leads to the remarkable fact that the system is also insensitive to whether the scattering length goes to plus or minus infinity. In other words, a many particle system is insensitive to whether the two-body system has a marginally bound state just above or below threshold; the two-body bound state or resonance is dissolved in matter at sufficiently high density, $`k_F|a|\stackrel{>}{}1`$. For positive scattering lengths a pair condensate of molecules may form at low densities but they dissolve at intermediate densities when the range of the two-body wavefunction exceeds the interparticle distance. In Fig. 1 the energy per particle is shown as function of density extending from the dilute and intermediate densities, Eqs. (11,15), to high densities, Eq. (14). At low densities the Fermi kinetic energy dominates but at intermediate densities, $`R\stackrel{<}{}k_F^1\stackrel{<}{}|a|`$, the attractive potential lower the energy by an amount proportional to the statistical factor. The two cases $`\nu =2`$ and $`\nu =4`$ are seen to saturate at positive and negative energies respectively. In the high density limit the attractive (Hartree) potential of Eq. (14) dominates and will lead to collapse of all Fermi systems in the absence of repulsive cores. Fig. 1 also shows the superfluid gaps of Eqs. (6,7) for dilute and intermediate density Fermi systems. At low densities they decrease exponentially as $`\mathrm{\Delta }E_F\mathrm{exp}(2/\pi k_F|a|)`$ whereas at intermediate densities the gaps are a finite fraction of the Fermi energy. At high densities the gap generally decreases rapidly with density . For example, for an attractive square well potential of range $`R`$ and depth $`V_0`$ with long scattering length (or marginally bound state), i.e. $`V_0R^2\pi ^2/4m`$, the gap decrease exponentially as $`\mathrm{\Delta }\mathrm{exp}(4k_FR/\pi )`$. When $`|a|R`$ plateaus appear for $`(E/N)/E_F`$ and $`\mathrm{\Delta }/E_F`$. Since $`E_F`$ also decrease with decreasing density, the gap itself is narrowly peaked near $`k_F1/R`$ as found in nuclear and neutron matter . Information on the density dependence can be obtained independently from calculations within the Wigner-Seitz cell approximation that has recently been employed for the strongly correlated nuclear liquid . The periodic boundary condition is a computational convenience which contains the important scale for nucleon-nucleon correlations given by the interparticle spacing. It naturally gives the correct low density Eq. (11) and high density Eq. (14) limits. At intermediate densities one obtains $`{\displaystyle \frac{E}{N}}`$ $`=`$ $`E_F\left[{\displaystyle \frac{3}{5}}c_2{\displaystyle \frac{\nu 1}{\nu ^{1/3}}}\right].`$ (18) Finite crystal momenta complicates the calculation of $`c_2`$. A lower (but reasonable) estimate $`c_20.25`$ can be calculated. The potential energy in Eq. (18) is also proportional to the kinetic one as found in Eq. (16) and of similar magnitude. The scaling with $`\nu ^{1/3}`$ arise because energies scale with the square of the inverse particle spacing, $`r_0^2`$, in the Wigner-Seitz cell approximation, and $`\rho =\nu k_F^3/6\pi ^2=(4\pi r_0^3/3)^1`$. The energy per particle can be calculated at all densities and finite values of $`a`$ and $`R`$ and the cross over between the three density regimes generally confirm the energy per particle shown in Fig. 1. ## IV Phase diagram Constructing a phase diagram from the low temperature degenerate regime to the high temperature classical regime requires a finite temperature generalization. For illustration we shall follow the procedure as in Ref. and employ the high temperature approximation for the additional thermal pressure. At high temperatures quantal effects are negligible and the energy per particle is simply given by the classical value $`E/N3T/2`$. The isothermal sound speed is within this approximation $`s_T^2={\displaystyle \frac{T}{m}}+s_{T=0}^2,`$ (19) where the zero temperature sound speed is given by Eq. (12) with energy per particle from Eqs. (15,11). The spinodal instability condition, $`s_T=0`$, determines the line of collapse $`T(\rho )`$ for long wavelength density fluctuations. The resulting phase diagram is shown in Fig. 2 for $`\nu =2,3,4,7,10`$ spin states. The lower ($`T\stackrel{<}{}T_F`$) and right ($`k_F|a|\stackrel{>}{}1`$) corner of the phase diagram is the spinodally unstable region where the system collapses. The region decreases for fewer spin states and is absent for $`\nu =2`$. For comparison the spinodal lines for $`\nu =2`$ and $`\nu =3`$ are shown when the dilute approximation of Eq. (11) is extrapolated into intermediate densities. Generally, the spinodally unstable regions based on the dilute approximation are substantially overestimated. The regions of superfluidity given by $`T_c=(\gamma /\pi )\mathrm{\Delta }`$ and Eq. (7) are also shown in Fig. 2. As for the spinodally unstable region it is the lower right corner that is superfluid. However, superfluidity extends to lower densities and therefore mechanical instability does not prevent the BCS-type pairing in the case of fermions. The opposite conclusion was reached for the pairing transition in Bose-Einstein condensates . As cooling becomes increasingly difficult at temperatures below the Fermi temperature we observe that the superfluid transition is readily obtained by increasing the density above $`k_F|a|\stackrel{>}{}1`$. The phase diagram is quantitatively correct at low as well as high temperatures. Around the Fermi energy it gives a qualitative description only due to the approximate thermal pressure employed. Furthermore, at intermediate densities the superfluid gaps become large exceeding the Fermi energy for large spins, and the corrections to the ground state energies can therefore not be ignored. ## V Finite systems The degenerate Fermi gases and Bose-Einstein condensates (BEC) produced so far contain $`n10^310^6`$ magnetically trapped alkali atoms. Some of them interact via long scattering lengths such as the triplet $`{}_{}{}^{6}Li`$ fermions with $`a=2160`$ Bohr radii and singlet $`{}_{}{}^{85}Rb_2`$ bosons with $`|a|\stackrel{>}{}10^3`$ Bohr radii. Large scattering lengths $`a\pm \mathrm{}`$ can be taylored by using Feshback resonances, i.e. hyperfine states close to threshold further tuned by magnetic fields. Fermi gases differ from BEC’s in several respects. Most importantly, whereas bosons sit at zero momentum states, fermions have considerable kinetic energy. Therefore, when interactions are small, a BEC has energy per particle $`\mathrm{}\omega `$ and size $`a_{osc}`$, where $`\omega =(\omega _{}\omega _z)^{1/3}`$ is the geometric average of the magnetic trap frequencies and $`a_{osc}=\mathrm{}/\sqrt{m\omega }`$ is the oscillator length. A degenerate gas of $`N`$ Fermionic atoms has a larger energy per particle $`N^{1/3}\mathrm{}\omega `$ and size $`LN^{1/6}a_{osc}`$. In current experiments with degenerate Fermi gases and BEC’s the densities are low so that the dilute potential applies and the energy per particle is approximately $`{\displaystyle \frac{E}{N}}{\displaystyle \frac{3}{4}}\left({\displaystyle \frac{6}{\nu }}\right)^{1/3}N^{1/3}\mathrm{}\omega +{\displaystyle \frac{\nu 1}{\nu }}{\displaystyle \frac{2\pi a}{m}}{\displaystyle \frac{N}{L^3}},`$ (20) where the average density in the trap has been approximated by $`\rho N/L^3`$ . For a small number of trapped atoms with attractive scattering lengths the system is meta-stable but for a large number of trapped atoms, $`N\stackrel{>}{}\nu (a_{osc}/(\nu 1)a)^6`$, the attractive potential overcomes the Fermi kinetic energy and the degenerate Fermi gas becomes unstable and implodes. However, around the same density $`k_F|a|\stackrel{>}{}1`$ and we enter the intermediate density region, where the potential of (15) should be applied instead of the dilute potential. The gas is therefore mechanically stable for two hyperspins only contrary to conclusions based on the dilute potential . Recent experiment on cold magnetically trapped Fermionic atoms observed egeneracy for $`{}_{}{}^{40}K`$ atoms in the two hyperfine states $`m_F=9/2,7/2`$. Current experimental oscillator lengths $`a_{osc}\mu m`$ are less than one order of magnitude longer than the atomic scattering length $`|a|`$ of $`{}_{}{}^{6}Li`$. It should be possible to reach intermediate densities, $`k_F|a|\stackrel{>}{}1`$, by trapping $`N\stackrel{>}{}10^6`$ $`{}_{}{}^{6}Li`$ atoms . The atomic gases offer the unique opportunity to vary the densities as well as the relative amount of the hyperfine states. Varying the composition is a convenient way to vary the gaps and attractive potential of Eqs. (11,15,14) through $`\nu `$ for given density and scattering length. In the limit where most atoms are in one of the states, the Fock and Hartree terms almost cancels and effectively $`\nu 1_+`$. More intricate systems of mixtures of fermions and bosons, e.g. $`{}_{}{}^{39,40,41}K`$ isotopes can also be studied. If the interaction is attractive it will contract the atomic cloud towards higher densities. Irrespective of whether the bosons or fermions attract or repel the induced interactions, which are of second order in the fermion-boson coupling, enhance the gap . An artificial “gravitational” or “Coulomb” force can be exerted on the atoms by shining laser light on the trapped cloud from many directions . It would add an energy per particle of order $`GNm^2/L`$ to Eq. (20) where $`G`$ is proportional to the laser field intensity. Such an interaction has several interesting consequences. If $`G`$ is attractive, it would contract the cloud towards higher densities which would increase gaps (see also ) and for sufficiently large $`G`$ the intermediate density region is entered. Depending on the strengths and sign of the scattering amplitude and gravitational interactions the kinetic energy of the atoms will be balanced by the magnetic trap and/or the scattering or gravitational interactions. The resulting phase diagram is much more complex. If such a strong attractive laser field is suddenly applied to the gas, the Jeans instability sets in and the gas collapses until balanced again by the kinetic energy. Subsequently, the system will “bounce” analogous to the initial stages of a supernova explosion. If, however, intermediate energies are reached and the number of spins exceed $`\nu \stackrel{>}{}2.5`$, then the collapse will be further accelerated by the attraction between atoms. The corresponding critical particle number is $`N_c(Gm^3a)^{3/2},`$ (21) at zero temperature. It differs from the standard Chandrasekhar mass by a factor $`(m|a|)^{3/2}`$ because the instability condition is $`k_F|a|1`$ whereas stars go unstable when the particles become relativistic $`k_Fm`$. ## VI Summary The energy per particle and superfluid gaps have been calculated for an homogeneous system of fermions interacting via a long attractive s-wave scattering length. In the intermediate region of densities, where the interparticle spacing $`(1/k_F)`$ is much longer than the range of the interaction but much shorter than the scattering length or $`|a|`$, the energy per particle and superfluid gaps are proportional to the Fermi energy. The energy per particle increases linearly with the spin-isospin or hyperspin statistical factor such that, e.g., symmetric nuclear matter is unstable in the intermediate density regions and undergoes spinodal decomposition whereas neutron matter and Fermionic atomic gases with few hyperspin states are mechanically stable. A phase diagram of Fermi gases at low and intermediate densities was constructed by including thermal pressures in the high temperature classical approximation. With the proper energy per particle at intermediate densities the spinodal region in the phase diagram was reduced substantially as compared to conjectures based on extrapolations from the dilute limit. Generally, mechanical instability does not prevent a superfluid transition for a wide range of densities. This is contrary to Bose gases, where spinodal instabilities exclude pairing transitions . The interaction energies of the many-body system were discussed for magnetically trapped cold degenerate gases of Fermi atoms. In such systems both superfluidity and the intermediate density region should be attainable. In these “novel” density regions the superfluid gaps can be large and the stability and sensitivity to the statistical factor $`\nu `$ can be studied. Adding a gravitationally like force by shining laser light on the atomic cloud further increase densities whereby collapse and bounce analogous to the early stages of supernova explosions may be studied.
warning/0002/astro-ph0002322.html
ar5iv
text
# Testing the connection between the X-ray and submillimetre source populations using Chandra ## 1 Introduction The spectrum of the X-ray Background (XRB) over the 1-7 keV band is a power-law of energy index 0.4 (Gendreau et al. 1995) which is flatter than that of any known class of extragalactic source. Following the original suggestion of Setti & Woltjer (1989), it is commonly assumed that this is due to the XRB being dominated by many absorbed sources. These sources, with different absorbing column densities and redshifts, combine to give the observed XRB spectrum (Madau, Ghisellini & Fabian 1994; Matt & Fabian 1994; Comastri et al. 1995; Wilman & Fabian 1999). The absorbed energy is presumably reradiated in the Far InfraRed (FIR) band, as observed in nearby heavily absorbed Active Galactic Nuclei (AGN) like that in the luminous IRAS galaxy NGC 6240 (Vignati et al. 1999). The residual harder X-ray emission which penetrates the absorbing medium, and the tail of the reradiated emission in the submillimetre band from such an source both exhibit a negative K-correction. This means that obscured AGN should be detectable to large redshifts in both bands. Observations of the XRB with Chandra show that much (more than 80 per cent) of it is resolved into point sources in the 2–8 keV band (Brandt et al. 2000; Mushotzky et al. 2000). More than half the intensity is due to a combination of hard sources identified with either otherwise normal bright galaxies or in optically faint or even invisible galaxies. Deep 850-$`\mu `$m submillimetre observations with SCUBA show that much (more than 80 per cent) of the submillimetre background seen by COBE-FIRAS (Fixsen et al. 1998) at this wavelength is resolved into discrete sources (Blain et al. 1999). A key question is what fraction of the deep X-ray and submillimetre source populations are related. The far-infrared and submillimetre background represents a significant part of the energy output of objects in the Universe. If the contribution of starbursts and AGN to the background can be separated using deep X-ray images, then the relative importance of high-mass star formation and AGN in heating the dust responsible for this emission can be determined. Recent modelling suggests that the AGN fraction in the submillimetre background is about 20 per cent (Almaini, Lawrence & Boyle 1999; Fabian & Iwasawa 1999; Gunn & Shanks 1999). The uncertainty is such that AGN may contribute in total between 10–50 per cent of the total energy output of stars (Fabian 1999). Smail et al. (1997, 1998) have used the SCUBA instrument at the JCMT (Holland et al 1999) to study the submillimetre sources in seven massive clusters of galaxies at redshifts between 0.2 and 0.4, exploiting the gravitational lensing magnification from the clusters to make an exceptionally deep 850-$`\mu `$m survey for background sources. Chandra images of two of the clusters, A 2390 and A 1835, have recently been obtained which, owing to the superb angular resolution, probe much deeper than any previous X-ray images of these regions, for example the ROSAT-HRI limit of $`8\times 10^{14}`$ erg s<sup>-1</sup> cm<sup>-2</sup> to the 0.1–2.0 keV X-ray flux of the SCUBA source SMMJ 14011+0252 in A 1835 (Ivison et al. 2000). Here we study the Chandra X-ray limits to the flux of the SCUBA sources and conversely SCUBA limits on the Chandra sources found in the images. Only one source is marginally detected in both wavebands. We then discuss the implications for obscured AGN models. ## 2 Observations and Results Chandra observed A 2390 on 1999 November 5 for a livetime of 9,126 s and A 1835 on 1999 December 11 for a total of 19,626 s. For each observation the cluster lies close to the aimpoint of the telescope on the ACIS-S back-illuminated CCD chip. Only mild temporal variations in count rate are seen through the observations so we use the whole exposure in this work. A 7-arcsec pointing offset evident in the A 1835 field from the position of the cluster X-ray peak and several other source identified in the field has been removed. We estimate that the source positions from the Chandra images are accurate to 1<sup>′′</sup> rms. The analysis of the cluster emission will be presented elsewhere; here we restrict ourselves to the positions of the SCUBA sources and of three X-ray sources in the Chandra fields which lie within the SCUBA field (i.e. within 1.5 arcmin of the cluster centre; Table 1). A further possible ($`2.8\sigma `$) X-ray source in the A 2390 field is also discussed since it is spatially coincident with a tentative ($`2\sigma `$) SCUBA source. No point-like X-ray sources are seen in the SCUBA region of the Chandra field of A 1835. The X-ray data were analysed using images in the 0.5–2 keV and 2–7 keV bands with one-arcsec pixels. The sources detected in the SCUBA field of A 2390 appear principally in just 4 neighbouring Chandra pixels (a few counts from the brightest sources occupy 2 further pixels). We therefore adopt a region of 4 sq. arcsec when estimating source fluxes and limits. The cluster emission means that the background is not flat, and so the background has been determined by averaging the 8 neighbouring pixels in images formed with 10 by 10 arcsec pixels, excluding the source pixel. Upper limits in the X-ray band have been obtained by using the Bayesian method of Kraft, Burrows & Nousek (1990) and are quoted at the 99 per cent confidence level (Table 2). We assume Galactic columns of $`6.8\times 10^{20}\mathrm{cm}^2`$ for A 2390 and $`2.3\times 10^{20}\mathrm{cm}^2`$ for A 1835. The submillimetre, radio and optical imaging of A 1835 used here is discussed in detail in Ivison et al. (2000), while the submillimetre observations of A 2390 are detailed in Smail et al. (1998; a new sources is reported here), the VLA 1.4-GHz radio map in Edge et al. (1999) and the optical imaging of this cluster with Hubble Space Telescope (HST) is described in Pelló et al. (1999). In addition, there is deep 6.7$`\mu `$m and 15-$`\mu `$m ISO imaging of A 2390 (Altieri et al. 1999; Lemonon et al. 1999), which is sensitive to hot dust emission from background galaxies out to $`z1`$. In Fig. 1 we show overlays of the Chandra and SCUBA images on optical images of A 1835 and A 2390. We note that the brightest Chandra source, CXOUJ215334.0+174240, is associated with a relatively large (4<sup>′′</sup> total extent) mid-type spiral with a blue nucleus and a faint companion. This galaxy has a colour and apparent magnitude similar to those expected for a slightly reddened $`L^{}`$ mid-type spiral at $`z=0.85\pm 0.15`$. While the next brightest X-ray source, CXOUJ215333.2+174209 has a more amorphous counterpart, with a bright compact nucleus and an asymmetric envelope and has a much bluer colour, its redshift is not strongly constrained by the current data. Counterparts to the remaining two Chandra sources are not seen in the optical imaging (although there is a very faint object in the vicinity of CXOUJ215334.4+174205); if they are background galaxies then, after correction for lensing amplification, they must be fainter than $`I27`$. CXOUJ215334.0+174240 is the only one of the four Chandra sources detected at 1.4 GHz with the VLA, at a flux density of $`2.6\pm 0.1`$ mJy, the 3-$`\sigma `$ limits on the remainder being $`<0.2`$ mJy. For the two X-ray sources which are detected in both X-ray bands we have determined an X-ray spectral index, $`\alpha _\mathrm{x}`$ from the flux ratios. These are both unphysically negative ($`0.8`$ and $`0.3`$; we define all spectral indices according to $`F\nu ^\alpha `$) and thus indicate intrinsic absorption. The brightest source, CXOUJ215334.0+174240, has $`90`$ counts, which enables a crude spectral analysis to be performed. A straight power-law spectrum with Galactic absorption is a poor fit and yields an energy index of $`0.3`$. Including additional soft X-ray absorption allows a better fit, although with no firm constraints on the index. The absorption however must exceed $`N=6\times 10^{21}\mathrm{cm}^2`$ at the 90 per cent confidence level. Note that the intrinsic absorption will be approximately $`(1+z)^3N`$, where $`z`$ is the redshift of the source. For an assumed energy index of 1, typical of quasars, and the redshift indicated by the optical colours, we find $`N=(6\pm 2)`$$`(9\pm 3)\times 10^{22}\mathrm{cm}^2`$, and an unabsorbed intrinsic 2–10 keV luminosity of 2.7–$`6.8\times 10^{44}\mathrm{erg}\mathrm{s}^1`$, for $`z=0.7`$–1, respectively. After correction for the lensing amplification factor of about 2, the source has an intrinsic $`L(2`$$`10\mathrm{keV})2`$$`3\times 10^{44}\mathrm{erg}\mathrm{s}^1`$. This makes it a strong contender to be one of the first genuine Type-II quasars (see discussions in Halpern et al. 1999; Vignati et al. 1999; Franceschini et al 1999). The X-ray limits on the SCUBA sources in the A 1835 field are about 100 times deeper than previously achieved using ROSAT data (Ivison et al. 2000). We have produced two submillimetre-to-X-ray spectral indices, $`\alpha _{SX}`$, using the measured values at 850$`\mu `$m and 2 keV in our rest frame. Rather than obtain one X-ray estimate for the spectral flux at 2 keV from a whole band measurement, which would be biased by the most sensitive lower energy end of the band, we have estimated it by converting the observed counts to fluxes in the 0.5–2 and 2–7 keV bands (listed in Table 2), using the response matrix of the ACIS-S chip and assuming an intrinsic energy index of 1 (which is roughly appropriate for the scattered flux from an absorbed AGN). ## 3 Discussion We compare expected values of $`\alpha _{\mathrm{SX}}`$ as a function of redshift in Fig. 2 for various classes of observed objects, both starbursts and AGN. We select 3C 273 as an example of a powerful quasar, and use the submillimetre data of Neugebauer, Soifer & Miley (1985) in our model; Arp 220 as a starburst using the submillimetre data of Sanders et al. (1999) and the X-ray data of Iwasawa (1999); and NGC 6240 as a powerful obscured (Compton-thick) AGN using an estimate of the submillimetre spectrum based on that from Arp 220 normalised by the FIR luminosity and the BeppoSAX spectrum from Vignati et al. (1999). We also show how $`\alpha _{\mathrm{SX}}`$ changes as the absorption in NGC 6240 decreases from the measured value of $`2\times 10^{24}`$ to $`5\times 10^{23}\mathrm{cm}^2`$ and also if the scattered fraction drops to 1 per cent. Data from several quasars at $`z>4`$ are shown for comparison with the 3C 273 prediction. The typical limits on $`\alpha _{\mathrm{SX}}`$ for the SCUBA galaxies are $`\alpha _{\mathrm{SX}}<1.2`$ (at 99 per cent confidence). Combining the limits for the 6 galaxies we obtain a 99 per cent confidence limit on a typical SCUBA galaxy of $`\alpha _{\mathrm{SX}}>1.34`$, with the strongest constraints for individual galaxies being $`\alpha _{\mathrm{SX}}>1.32`$ for SMMJ 14011+0252 ($`z=2.55`$) and SMMJ 14009+0252 in the A 1835 field. Thus the indices for this population are straightforwardly consistent with starbursts. They do not resemble Compton-thin AGN at any redshift and are inconsistent with the spectrum of NGC 6240 at any redshift. If these galaxies host a powerful AGN then either a) reprocessed radiation from the AGN provides only a minor fraction of the submillimetre luminosity, or b) the source must be Compton thick and in addition the fraction of any scattered X-ray emission must be less than one per cent. We note that X-ray limits on the hyperluminous IRAS galaxy F15307+3252 at $`z=0.92`$, which does contain an AGN since broad scattered emission lines are seen (Hines et al. 1995), show that it must have a very low X-ray scattered fraction ($`<1`$ per cent), possibly due to intrinsic absorption of the scattered emission (Fabian et al. 1996). If this is typical of obscured AGN more powerful than NGC 6240, then the present results would be consistent with obscured AGN provided that they are at relatively high redshift ($`z>2`$). Of the four Chandra sources we identified in our fields, two have probable optical counterparts suggesting that they are distant disk galaxies. The colours and luminosity of the brightest of these suggest it is an $`L^{}`$ mid-type galaxy at $`z1`$. If the two optically-unidentified Chandra sources are AGN, and the host galaxies have luminosities of $`L^{}`$ or greater, then they must be at $`z>2`$, or be intrinsically reddened. Obscuration of both AGN and surrounding spheroid is required by some models for the XRB (Fabian 1999). It is possible for these dust-enshrouded strongly-obscured AGN to be undetected in our 850-$`\mu `$m SCUBA maps if their dust is typically much hotter than 40 K. Comparing the radio flux and SCUBA limit for the brightest Chandra galaxy CXOUJ215334.0+174240 (§2), the redshift $`z`$ and dust temperature $`T_d`$ of the source must satisfy the relationship $`T_d>27(1+z)`$ K. Taking the redshift constraint from the optical, $`z0.85\pm 0.15`$, we derive $`T50`$ K. Such sources would not contribute substantially to the submillimetre background at $`1`$ mm, but might at shorter wavelengths. Eventual comparison with ISO imaging may shed further light on the nature of the dust emission in these sources. In summary, Chandra and SCUBA observations have been combined to probe the relation between the faint X-ray and submillimetre sky. Only one marginal source is seen in both datasets, and so in general we find deep submillimetre limits from SCUBA on the X-ray sources found using Chandra and deep X-ray limits from Chandra on the SCUBA-selected sources. The limits on background sources in these fields are particularly strong due to amplification by gravitational lensing. For the SCUBA galaxies, we cannot completely rule out the presence of either an obscured, weak AGN or a more powerful Compton-thick AGN in which any scattered flux is also very weak or absorbed. However, the simplest explanation of our results on the SCUBA sources is that they are predominantly powered by starbursts. Clearly the current sample is small and so we cannot make any definitive statement about the whole SCUBA population. We note that at the moment our conclusions are not inconsistent with suggestions that AGN powering 20 per cent of the SCUBA population, although a substantially higher fraction would be difficult to accommodate. For the Chandra sources in the A2390 field we find that they have optically faint counterparts, $`I>23`$–27. We identify one source as a probable Type-II obscured quasar at $`z1`$. We suggest that the remaining, typically fainter sources are either more distant, $`z>2`$, or intrinsically reddened. Observations of these fields in the near-infrared are urgently needed to test the nature of these sources. ## 4 Acknowledgements We thank L. van Speybroeck and his colleagues for the superb X-ray telescope and M. Weisskopf and the project team for the Chandra mission. ACF is grateful to NASA for the opportunity to participate in Chandra. The Royal Society is thanked for support by ACF, IRS, SWA, CSC and SE and PPARC by RJI. AWB and JPK thank the Raymond & Beverly Sackler Foundation and the LENSNET European Network, respectively.
warning/0002/hep-th0002248.html
ar5iv
text
# Effective Dual Higgs Mechanism with Confining Forces ## Abstract We consider the dual Yang-Mills theory which shows some kind of confinement at large distances. In the static system of the test color charges an analytic expression for the string tension is derived. 1. In this paper, we consider the model (in four-dimensional space-time (4d)) based on the dual description of a long-distance Yang-Mills (LDY-M) theory which can provide a quark confinement in a system of static test charges. This work follows the idea that the vacuum of quantum Yang-Mills (Y-M) theory is realized by a condensate of monopole-antimonopole pairs \[1-4\]. Since there are no monopoles as classical solutions with finite energy in a pure Y-M theory, it has been suggested by ’t Hooft going into the Abelian projection where the gauge group SU(2) is broken by a suitable gauge condition to its Abelian subgroup U(1). Now there is the well-known statement that the interplay between a quark and antiquark is analagous to the interaction between a monopole and an antimonopole in a superconductor. The topology of the Y-M SU(N) manifold and that of its Abelian subgroup $`[U(1)]^{N1}`$ are different, and new topological objects can appear in case of introducing the local gauge transformation of some gauge function in our model, eg., the field strength tensor for the gauge field $`A_\mu (x)`$ in quantum chromodynamics (QCD) with $`D_\mu (x)=_\mu +ieA_\mu (x)`$ $$F_{\mu \nu }(x)=\frac{1}{ie}\left([D_\mu (x),D_\nu (x)][_\mu ,_\nu ]\right),$$ which transforms with the gauge function $`\mathrm{\Omega }(x)`$ as $`F_{\mu \nu }(x)F_{\mu \nu }^\mathrm{\Omega }(x)=\mathrm{\Omega }(x)F_{\mu \nu }(x)\mathrm{\Omega }^1(x)=_\mu A_\nu ^\mathrm{\Omega }(x)_\nu A_\mu ^\mathrm{\Omega }(x)+`$ (1) $`+ie[A_\mu ^\mathrm{\Omega }(x),A_\nu ^\mathrm{\Omega }(x)]{\displaystyle \frac{1}{ie}}\mathrm{\Omega }(x)[_\mu ,_\nu ]\mathrm{\Omega }^1(x).`$ (2) The last term in (1) reflects the singular character of the above-mentioned gauge transformation. One can identify the Abelian projection by the replacement $$F_{\mu \nu }^\mathrm{\Omega }(x)F_{\mu \nu }^\alpha (x)=_\mu A_\nu ^\alpha (x)_\nu A_\mu ^\alpha (x)\frac{1}{ie}\mathrm{\Omega }(x)[_\mu ,_\nu ]\mathrm{\Omega }^1(x),$$ where $`A_\mu ^\mathrm{\Omega }A_\mu ^\alpha `$, while the label $`\alpha `$ reflects the Abelian world. This leads to the Dirac string and magnetic current $$J_\mu ^m(x)=\frac{1}{2ie}ϵ_{\mu \nu \rho \sigma }_\nu \mathrm{\Omega }(x)[_\rho ,_\sigma ]\mathrm{\Omega }^1(x)$$ in the Abelian gauge sector. Formally, a gauge group element, which transforms a generic SU(3) connection onto the gauge fixing surface in the space of connections, is not regular everywhere in spacetime. The projected (or transformed) connections contain topological singularities (or defects). Such a singularity may form the worldline(s) of magnetic monopoles. Hence, this singularity leads to the monopole current $`J_\mu ^{mon}`$. This is a natural way of the transformation from the Y-M theory to a model dealing with Abelian fields. Analytical models of the dual QCD with monopoles were intensively investigated \[6-9\]. We study the Lagrangian model where the fundamental variables are an octet of dual potentials coupled minimally to three octets of monopole (Higgs) fields. The dual gauge model is studied at the lowest order of the perturbative series using the canonical quantization. The basic manifestation of the model is that it generates the equations of motion where one of them for the scalar (Higgs) field looks like as a dipole-like field equation. The monopole fields obeying such an equation are classified by their two-point Wightman functions (TPWF). In the scheme presented in this work, the flux distribution in the tubes formed between two heavy color charges is understood via the following statement: the Abelian monopoles are excluded from the string region while the Abelian electric flux is squeezed into the string region. In our model, we use the dual gauge field $`\widehat{C}_\mu ^a(x)`$ and the monopole field $`\widehat{B}_i^a(x)`$ ($`i=1,\mathrm{},N_c(N_c1)/2`$ and $`a`$=1,…,8 is a color index) which are relevant modes for infrared behaviour. The local coupling of the $`\widehat{B}_i^a`$-field to the $`\widehat{C}_\mu ^a`$-field provides the mass of the dual field and, hence, a dual Meissner effect. The commutation relations, TPWF and Green’s functions as well-defined distributions in the space $`S(\mathrm{}^d)`$ of complex Schwartz test functions on $`\mathrm{}^d`$, will be defined in the text. 2. Let us consider the Lagrangian density (LD) $`L`$ of the $`U(1)\times U(1)`$ dual Higgs model corresponding to the LDY-M theory $`L=2Tr\left[{\displaystyle \frac{1}{4}}\widehat{F}^{\mu \nu }\widehat{F}_{\mu \nu }+{\displaystyle \frac{1}{2}}\left(D_\mu \widehat{B}_i\right)^2\right]W\left(\widehat{B}_i\right),`$ (3) where $$\widehat{F}_{\mu \nu }=_\mu \widehat{C}_\nu _\nu \widehat{C}_\mu ig[\widehat{C}_\mu ,\widehat{C}_\nu ],$$ $$D_\mu \widehat{B}_i=_\mu \widehat{B}_iig[\widehat{C}_\mu ,\widehat{B}_i],$$ $`\widehat{C}_\mu `$ and $`\widehat{B}_i`$ are the SU(3) matrices, g is the gauge coupling constant; $`\widehat{C}_\mu =_aC_\mu ^a\frac{1}{2}\lambda _a`$, $`\lambda ^a`$ are generators of SU(3). The Higgs fields develop their vacuum expectation values (v.e.v.) $`\widehat{B}_{0_i}`$ and the Higgs potential $`W(\widehat{B}_i)`$ has a minimum at $`\widehat{B}_{0_i}`$. The v.e.v. $`\widehat{B}_{0_i}`$ produce a color monopole generating current confining the electric color flux. Representing the quark sources by the Dirac string tensor $`\stackrel{~}{G}_{\mu \nu }(x)`$, we read Eq. (3) as $`L(\stackrel{~}{G}_{\mu \nu })={\displaystyle \frac{1}{3}}G_{\mu \nu }^2+4|(_\mu igC_\mu )\varphi |^2+2(_\mu \varphi _3)^2W(\varphi ,\varphi _3),`$ (4) where $`G_{\mu \nu }=_\mu C_\nu _\nu C_\mu +\stackrel{~}{G}_{\mu \nu }`$; $`\varphi (x)`$ and $`\varphi _3(x)`$ denote the complex scalar monopole fields. The effective potential becomes (see ) $$W(\varphi ,\varphi _3)=\frac{2}{3}\lambda \{11[2(B^2+\overline{B}^2B_0^2)^2+(B_3^2B_0^2)^2]$$ $$+7[2(B^2+\overline{B}^2)+B_3^23B_0^2]^2\},$$ where $`\varphi \varphi _1=\varphi _2=B_{1,2}i\overline{B}_{1,2},\varphi _3=B_3;`$ $`\lambda `$ is dimensionless. The invariance of the LD ( 4 ) under the local gauge transformation $`C_\mu (x)C_\mu (x)+(1/g)_\mu \theta _c(x)`$ and the phase transformation $`\varphi _i(x)\mathrm{exp}[iϵ_i\theta _c(x)]\varphi _i(x)`$ is assumed. Here, $`\theta _c(x)S(\mathrm{}^4)`$ is the real function, $`ϵ_1=(1,0),ϵ_2=(\frac{1}{2},\frac{1}{2}\sqrt{3}),ϵ_3=(\frac{1}{2},\frac{1}{2}\sqrt{3})`$ . The generating current of (4) is nothing but the monopole current confining the electric color flux $`J_\mu ^{mon}=(2/3)^\nu G_{\mu \nu }(x)`$. The formal consequence of the $`J_\mu ^{mon}`$ conservation, $`^\mu J_\mu ^{mon}=0`$, means that monopole currents form closed loops. Since $`\varphi _1`$ and $`\varphi _2`$ couple to $`C_\mu `$ in the same way, we choose $`B(x)=b(x)+B_0,\overline{B}(x)=\overline{b}(x),B_3(x)=b_3(x)+B_0`$ with $`B(x)_00`$, $`\overline{B}(x)_00`$, $`B_3(x)_00`$. In terms of the new fields $`b,\overline{b},b_3`$ the LD (4) is divided into two parts $`L=L_1+L_2`$ where $`L_1`$ in the lowest order of $`g`$ and $`\lambda `$ and with the minimal weak interaction looks like $$L_1=\frac{1}{3}G_{\mu \nu }^2+\mathrm{\hspace{0.17em}4}\left[(_\mu b)^2+(_\mu \overline{b})^2+\frac{1}{2}(_\mu b_3)^2\right]$$ $$+m^2C_\mu ^2\frac{4}{3}\mu ^2(50b^2+18b_3^2)+8m_\mu \overline{b}C_\mu .$$ Here, $`mgB_0`$ and $`\mu \sqrt{2\lambda }B_0`$. The equations of motion for the fields $`b,\overline{b},b_3`$ and $`C_\mu `$ are $$(\mathrm{\Delta }^2+\mu _1^2)b(x)=0;$$ $$\mathrm{\Delta }^2\overline{b}(x)+m(C)=0;$$ $$(\mathrm{\Delta }^2+\mu _2^2)b_3(x)=0;$$ $`(\mathrm{\Delta }^2+m_1^2)C_\mu (x)_\mu (C)+12m_\mu \overline{b}(x)^\nu \stackrel{~}{G}_{\mu \nu }(x)=0,`$ (5) where $`\mu _1^2=(50/3)\mu ^2,\mu _2^2=12\mu ^2,m_1^2=3m^2`$. The formal solution of equation (5) looks like $$C_\mu (x)=\alpha ^\nu \stackrel{~}{G}_{\mu \nu }(x)\beta _\mu \overline{b}(x),$$ with $`\alpha (3m^2)^1`$, $`\beta 4/m`$. We obtain that the dual gauge field is defined via the divergence of the Dirac string tensor $`\stackrel{~}{G}_{\mu \nu }(x)`$ shifted by the divergence of the scalar field $`\overline{b}(x)`$. For large enough $`\stackrel{}{x}`$, the monopole field is going to its v.e.v. while $`C_\mu (\stackrel{}{x}\mathrm{})0`$ and $`J_\mu ^{mon}(\stackrel{}{x}\mathrm{})8m^2C_\mu `$. It implies that in the $`d=2h`$ dimensions the $`\overline{b}(x)`$-field obeys the equation $`\mathrm{\Delta }^{2h}\overline{b}(x)0,h=2,3,\mathrm{},`$ (6) for a very weak $`C_\mu `$-field, but $`\mathrm{\Delta }^2\overline{b}(x)0`$. Here, the solutions of equation (6) obey locality, Poincare covariance and spectral conditions, and look like the dipole ”ghosts” at h=2. We define TPWF $`W_h(x)`$ in the d=2 h-dimensions $`W_h(x)=\overline{b}(x)\overline{b}(0)_0`$ as the distribution in the Schwartz space $`S^{}(\mathrm{}^{2h})`$ of temperate distributions on $`\mathrm{}^{2h}`$ which obeys the equation $`\mathrm{\Delta }^{2h}W_h(x)=0.`$ (7) The general solution of (7) should be Lorentz invariant and is given in the form at h=2 $`W_2(x)=a_1\mathrm{ln}{\displaystyle \frac{l^2}{x_\mu ^2+iϵx^0}}+{\displaystyle \frac{a_2}{x_\mu ^2iϵx^0}}+a_3,`$ (8) where $`a_i(i=1,2,3)`$ are the coefficients , $`l`$ is an arbitrary length scale. The coefficients $`a_1`$ and $`a_2`$ in (8) can be fixed using the canonical commutation relations (CCR) $`[C_\mu (x),\pi _{C_\nu }(0)]_{|_{x^0=0}}=ig_{\mu \nu }\delta ^3(\stackrel{}{x})`$ and $`[\overline{b}(x),\pi _{\overline{b}}(0)]_{|_{x^0=0}}=i\delta ^3(\stackrel{}{x})`$, respectively, with $`\pi _{C_\mu }(x)=\frac{4}{3}G_{0\mu }(x)`$ and $`\pi _{\overline{b}}(x)=8\left[^0\overline{b}(x)+mC^0(x)\right]`$. The standard commutator for the scalar field $`\overline{b}`$ is $$[\overline{b}(x),b(0)]=(2\pi )^2i\left[4a_1E_2(x)+a_2D_2(x)\right],$$ where $`E_2(x)=(8\pi )^1sgn(x^0)\theta (x^2)`$ and $`D_2(x)=(2\pi )^1sgn(x^0)\delta (x^2)`$ are taken into account. The direct calculation leads to $`a_1=(m^2/48\pi ^2)`$ and $`a_2=(1/24\pi ^2)`$. The propagator of the $`\overline{b}`$-field in $`S^{}(\mathrm{}_4)`$ is $`\widehat{\tau }_2(p)=weak\underset{\stackrel{~}{\kappa }^2<<1}{lim}{\displaystyle \frac{i}{3(2\pi )^4}}\{m^2[{\displaystyle \frac{1}{(p^2\kappa ^2+iϵ)^2}}+i\pi ^2\mathrm{ln}{\displaystyle \frac{\kappa ^2}{\stackrel{~}{\mu }^2}}\delta _4(p)]`$ (9) $`{\displaystyle \frac{1}{2}}{\displaystyle \frac{1}{p^2\kappa ^2+iϵ}}\}.`$ (10) Here, $`\kappa `$ is a parameter of representation and not an analogue of the infrared mass, $`\stackrel{~}{\kappa }^2\kappa ^2/p^2`$. To define the commutation relation $`[C_\mu (x),C_\nu (y)]`$, let us consider the canonical conjugate pair $`\{C_\mu ,\pi _{C_\nu }\}`$ $`[{\displaystyle \frac{4}{3}}C_\mu (x),_\nu C_0(0)_0C_\nu (0)g_{0\nu }(C(0))+\mathrm{\Delta }_{0\nu }(0)]_{|_{x^0=0}}=ig_{\mu \nu }\delta ^3(\stackrel{}{x}),`$ (11) where $`\mathrm{\Delta }_{\mu \nu }(x)=g_{\mu \nu }(C(x))\stackrel{~}{G}_{\mu \nu }(x)`$ tends to zero as $`x0`$ and the Dirac string tensor $`\stackrel{~}{G}_{\mu \nu }(x)`$ obeys the equation $`\left[\mathrm{\Delta }^2+(3m)^2\right]\stackrel{~}{G}_{\mu \nu }(x)=0`$. Obviously, the following form of the free $`C_\mu `$-field commutator: $`[C_\mu (x),C_\nu (0)]=ig_{\mu \nu }\left[\xi m_1^2E_2(x)+cD_2(x)\right],`$ (12) ensures the CCR (11) at large $`x_\mu ^2`$ with both $`\xi `$ and $`c`$ (in (12)) being real arbitrary numbers but $`\xi =\frac{3}{4}4c`$. The free dual gauge field propagator in $`S^{}(\mathrm{}_4)`$ in any local covariant gauge is given by $`\widehat{\tau }_{\mu \nu }(p)={\displaystyle d^4x\mathrm{exp}(ipx)\tau _{\mu \nu }(x)}`$ (13) $`=i\left[g_{\mu \nu }\left(1{\displaystyle \frac{1}{\zeta }}\right){\displaystyle \frac{p_\mu p_\nu }{p^2+iϵ}}\right]\left[\xi m_1^2\widehat{t}_1(p)+c\widehat{t}_2(p)\right],`$ (14) where $$\tau _{\mu \nu }(x)=\frac{ig_{\mu \nu }}{(4\pi )^2}\left[\xi m_1^2\mathrm{ln}(\stackrel{~}{\mu }^2x_\mu ^2+iϵ)+\frac{c}{x_\mu ^2+iϵ}\right];$$ $$\widehat{t}_1(p)=weak\underset{\stackrel{~}{\kappa }^2<<1}{lim}\left[\frac{1}{(p^2\kappa ^2+iϵ)^2}+i\pi ^2\mathrm{ln}\left(\frac{\kappa ^2}{\stackrel{~}{\mu }^2}\right)\delta _4(p)\right];$$ $$\widehat{t}_2(p)=weak\underset{\stackrel{~}{\kappa }^2<<1}{lim}\frac{1}{2}\frac{1}{(p^2\kappa {}_{}{}^{2}+iϵ)}.$$ The gauge parameter $`\zeta `$ in (13) is a real number. The following requirement $`(\mathrm{\Delta }^2)^2\stackrel{~}{\tau }_{\mu \nu }(x)=i\delta _4(x)`$ on Green’s function $`\tau _{\mu \nu }(x)`$ leads to that a constant $`c`$ has to be equal to zero and $`\stackrel{~}{\tau }_{\mu \nu }(x)=\tau _{\mu \nu }(x)/(\xi m_1^2)`$. 3. As for an approximate topological solution for this dual model, we fix the equations of motion $`^\nu G_{\mu \nu }=6ig[\varphi ^{}(_\mu igC_\mu )\varphi \varphi (_\mu +igC_\mu )\varphi ^{}],`$ (15) $`(_\mu igC_\mu )^2\varphi ={\displaystyle \frac{2}{3}}\lambda (32B_0^225|\varphi |^27\varphi _3^2)\varphi ,`$ (16) where $`\varphi (x)`$ is decomposed like $`\varphi (x)=\frac{1}{\sqrt{2}}\mathrm{exp}(if(x))[\chi (x)+B_0]`$ using the new scalar variables $`\chi (x)`$ and $`f(x)`$. The equation of motion (15) transforms into the following one: $$^\nu G_{\mu \nu }=6g(\chi +B_0)^2(gC_\mu _\mu f),$$ that means that the $`\overline{b}(x)`$-field is nothing but a mathematical realization of the ”massive” phase $`B_0f(x)`$ at large enough $`\stackrel{}{x}`$, i.e., $`\overline{b}(x)(B_0/2)S(x)f(x)`$ and $`S(x)(1+\chi (x)/B_0)^2`$. Integrating out $`G_{\mu \nu }`$ over the 2d surface element $`\sigma ^{\mu \nu }`$ in the flux $`\mathrm{\Pi }=G_{\mu \nu }(x)𝑑\sigma ^{\mu \nu }`$, we conclude that the phase $`f(x)`$ is varied by $`2\pi n`$ for any integer number $`n`$ associated with the topological charge inside the flux tube. Using the cylindrical symmetry we arrive at the field equation$`(\stackrel{}{C}(\stackrel{~}{C}(r)/r)\stackrel{}{e_\theta })`$ $$\frac{d^2\stackrel{~}{C}(r)}{dr^2}\frac{1}{r}\frac{d\stackrel{~}{C}(r)}{dr}3m^2[\mathrm{\hspace{0.17em}3}+2S(r)]\stackrel{~}{C}(r)+6nmB_0S(r)=0$$ with the asymptotic transverse behaviour of its solution $$\stackrel{~}{C}(r)\frac{4n}{7g}\sqrt{\frac{\pi mr}{2k}}e^{kmr}\left(\mathrm{\hspace{0.17em}1}+\frac{3}{8kmr}\right),k\sqrt{21}.$$ The field equation (16) is given by ($`\chi =\chi (r),S(r)=(1+\chi (r)/B_0)^2`$) $$\frac{d^2\chi (r)}{dr^2}+\frac{1}{r}\frac{d\chi (r)}{dr}\left\{\left[\frac{ng\stackrel{~}{C}(r)}{r}\right]^2+\frac{50}{3}\lambda B_0^2\left[1\frac{1}{2}S(r)\right]\right\}(\chi +B_0)0.$$ The profile of the color electric field in the flux tube at large $`r`$ looks like $$E_z(r)=\sqrt{\frac{\pi m}{2kr}}\left(km\frac{1}{2r}\right)e^{kmr}.$$ 4. Now, our aim is to obtain the confinement potential in an analytic form for the system of interacting static test charges of a quark and an antiquark. According to the distribution (9), the static potential in $`\mathrm{}^3`$ is a rising function with $`r=|\stackrel{}{x}|`$ $$P_{stat}(r)\frac{1}{2^{2h}\pi ^{3/2}}\frac{1}{(h1)!}\mathrm{\Gamma }(3/2h)r^{2h3},$$ and the Fourier transformation of the analytic function $`r^\sigma `$ at $`\sigma d,d2,\mathrm{}`$ is $$F\{r^\sigma \}=\left(\frac{4\pi }{p^2}\right)^{(\sigma +d)/2}\pi ^{\sigma /2}\frac{\mathrm{\Gamma }[(\sigma +d)/2]}{\mathrm{\Gamma }(\sigma /2)}.$$ In this paper, we define the static potential like $`P_{stat}(r)=lim_T\mathrm{}[T^1A(r)]`$ and the action $`A(r)`$ is given by the colour source-current part of LD $`L(p)=\stackrel{}{j}_\alpha ^\mu (p)\widehat{\tau }_{\mu \nu }(p)\stackrel{}{j}_\alpha ^\nu (p)`$ with the quark current $`\stackrel{}{j}_\alpha ^\mu (x)=\stackrel{}{Q}_\alpha g^{\mu 0}[\delta _3(\stackrel{}{x}\stackrel{}{x_1})\delta _3(\stackrel{}{x}\stackrel{}{x_2})]`$. Here, $`\stackrel{}{Q}_\alpha =e\stackrel{}{\rho }_\alpha `$ is the Abelian color-electric charge of a quark while $`\stackrel{}{\rho }_\alpha `$ is the weight vector of the SU(3) algebra: $`\rho _1=(1/2,\sqrt{3}/6)`$, $`\rho _2=(1/2,\sqrt{3}/6)`$, $`\rho _3=(0,1/\sqrt{3})`$ ; $`\stackrel{}{x}_1`$ and $`\stackrel{}{x}_2`$ are the position vectors of a quark and an antiquark, respectively. As a consequence of the dual field propagator (13), and using the following representation in the sense of generalized functions $$weak\underset{\stackrel{~}{\kappa }^2<<1}{lim}\left[\frac{1}{(p^2\kappa ^2+iϵ)^2}+i\pi ^2\mathrm{ln}\frac{\kappa ^2}{\stackrel{~}{\mu }^2}\delta _4(p)\right]=$$ $$=\frac{1}{4}\frac{^2}{p^2}\frac{1}{p^2iϵ}\mathrm{ln}\frac{p^2iϵ}{\stackrel{~}{\mu }^2}=\frac{1}{2}\frac{1}{(p^2+iϵ)^2}\left(53\mathrm{ln}\frac{p^2iϵ}{\stackrel{~}{\mu }^2}\right),$$ we get $`P_{stat}(r)={\displaystyle \frac{3\stackrel{}{Q}^2}{16\pi }}\left[\xi m^2r(12.4+6\mathrm{ln}\stackrel{~}{\mu }r)+O\left({\displaystyle \frac{c}{r}}\right)\right].`$ (17) Hence, the string tension $`a`$ in $`P_{stat}(r)=ar`$ emerges as $`a{\displaystyle \frac{9\stackrel{}{Q}^2}{64\pi }}m^2\left(12.4+3\mathrm{ln}{\displaystyle \frac{\stackrel{~}{\mu }^2}{m^2}}\right),\stackrel{~}{\mu }>9m,`$ (18) where $`r`$ in the logarithmic function in (17) has been changed by the characteristic length $`r_c1/m`$ which determines the transverse dimension of the dual field concentration, while $`\stackrel{~}{\mu }`$ is associated with the ”coherent length” inverse, and the dual field mass $`m`$ defines the ”penetration depth” in the type II superconductor. For a typical value of the electroweak scale $`\stackrel{~}{\mu }250GeV`$, we get $`a0.10GeV^2`$ for the mass of the dual $`C_\mu `$-field $`m=0.5GeV`$ and $`a0.31GeV^2`$ if $`m=1.0GeV`$. The experimental string tension $`a_{exp}`$ then determines the fixed value of the dual mass $`m_{fix}`$ (eg., $`m_{fix}0.78GeV`$ at $`a_{exp}0.2GeV^2`$). Doing the formal comparison, let us note that the string tension in paper is given by $`ϵ={\displaystyle \frac{g^2m_v^2}{8\pi }}\mathrm{ln}\left(1+{\displaystyle \frac{m_{s}^{}{}_{}{}^{2}}{m_{v}^{}{}_{}{}^{2}}}\right),`$ (19) with $`m_s`$ and $`m_v`$ being the masses of scalar and vector fields. We found that for a sufficiently long string $`r>>m^1`$ the $`r`$-behaviour of the static potential is dominant; for a short string $`r<<m^1`$ the singular interaction provided by the second term in (17) becomes important if the average size of the monopole is even smaller. 5. Finally, some conclusion is in order a). We have actually derived the analytic expressions of both the $`\overline{b}`$-field (9) and the dual gauge boson field (13) propagators in $`S^{}(\mathrm{}_4)`$. Our result should be regarded as the distributions (9) and (13) in a weak sense. The scheme is based on the flux-tube approach of Abelian dominance and monopole condensation. b). In this work, we have obtained that dual gauge bosons become massive due to their interaction with scalar field(s). But not every scalar species becomes massive since the symmetry breaking pattern is $`SU(3)U(1)\times U(1)`$ and one scalar field remains massless. We see that the fields $`b(x)`$ and $`b_3(x)`$ receive their masses and the $`\overline{b}(x)`$ field in combination with $`^\nu G_{\mu \nu }(x)`$ form the vector field $`C_\mu (x)`$ obeying the equation of motion for the massive vector field with the mass $`m=gB_0`$. The solution of the $`\overline{b}(x)`$-field can be identified as a ”ghost”-like particle in the substitute manner. Thus, we imply that two species of Abelian scalars (magnetic monopoles) are responsible for quark confinement. c). There is the first analytic result of having derived the potential (17) of static test charges at large distances in this paper. The form of this potential grows linearly with the distance $`r`$ apart from a logarithmic correction. The analytic comparison of $`ϵ`$ (19) with $`a`$ in (18) leads to the conclusion that we have obtained a similar behaviour of the string tension $`a`$ to those in the magnetic flux picture of the vortex and in the Nambu scheme , respectively, as well as in the dual Ginzburg-Landau model . d). Since no real physics can depend on the choice of the gauge group (where the Abelian group appears as a subgroup) it seems to be a new mechanism of confinement . The author is grateful to G.M. Prosperi for useful discussions and the kind hospitality at the University of Milan where this work has partly been done.
warning/0002/astro-ph0002373.html
ar5iv
text
# Properties of Proto–Planetary Nebulae ## 1. What is a Proto-Planetary Nebula? The proto-planetary nebula (PPN;a.k.a. post-AGB or pre-PN) stage of evolution immediately precedes the planetary nebula (PN) stage. The lifetime for this phase is $`\stackrel{<}{}`$1000 years and marks the time from when the star was forced off the asymptotic giant branch (AGB) by intensive mass loss to when the central star becomes hot enough (T$`{}_{\mathrm{eff}}{}^{}3\times 10^4`$ K) to photoionize the neutral circumstellar shell (Kwok 1993). We refer the reader to Kwok (1993) for a comprehensive review of PPN. For short recent reviews, see Hrivnak (1997) and van Winckel (1999). In this short review, I summarize the basic properties of PPN but focus primarily on the morphological studies because there have been numerous morphological studies in the past few years and because this particular conference is focused on morphology. ## 2. General Characteristics of PPN PPN are like PN in that the central star illuminates a detached circumstellar shell, however, we observe PPN using quite different techniques. Because PPN do not have ionized gas, we can not use the optically bright emission lines or radio free-free continuum commonly used for studies of PN. Instead tracers of dust and neutral gas are employed. In fact, PPN are identified as stars of spectral type B-K, luminosity class I with infrared excesses. These infrared excesses arise from the circumstellar dust which was originally created in the AGB wind. They emit broad ($``$20 km s<sup>-1</sup>), parabolic or double–horned lines of CO rotational lines and OH maser lines indicative of a remnant AGB circumstellar shell. These broad lines distinguish these PPN from pre-main sequence and Vega-excess stars which also have infrared excesses but typically narrower or non-existent molecular lines. Candidates for PPN are discovered using infrared sky surveys. One of the most famous PPN, AFGL 2688 (a.k.a the Egg Nebula), was discovered by Ney et al. (1975) in an infrared sky survey done by the airforce. Studies using the IRAS all sky survey have used two approaches to identify candidates. One way is to use IRAS colors (- vs. -), mark the locations of known AGB stars and PN and choose PPN candidates from the regions in between (van der Veen, Habing & Geballe 1989; Hrivnak, Kwok & Volk 1989; Hu et al. 1993). The second way is to cross correlate the IRAS catalog with optical star catalogs and choose objects in common (Oudmaijer et al. 1992). Several initial followup studies focused on ground based photometry observations and models of the spectral energy distributions (SEDs) of these PPN candidates. Van der Veen et al. (1989) identified four types of SEDs which they attributed to optical depth differences in the dust shells. Type I has a flat spectrum from 4 to 25 $`\mu `$m with a steep fall off at short wavelengths. Type II have a maximum near 25 $`\mu `$m and a gradual fall-off to shorter wavelengths. Type III have a maximum near 25 $`\mu `$m, a steep fall off at shorter wavelengths to a plateau between 1 and 4 $`\mu `$m. Type IV have two distinct maxima, one near 25 $`\mu `$m and the second at $`\lambda <2`$ $`\mu `$m. Hrivnak & Kwok (1991) suggested that the 21$`\mu `$m PPN were quite similar to objects like AFGL 2688 except for viewing angle based on the differences in their SEDs. The 21$`\mu `$m PPN have a Type IV SED while AFGL 2688 has a Type III SED and we could be viewing the 21$`\mu `$m PPN down the poles while we view AFGL 2688 edge-on. ## 3. Morphologies The study of PPN morphologies offers us insight on the axisymmetric PN issue because they are the missing link between two well studied groups: PN and AGB stars. Moreover, their morphologies are relatively pristine fossil records of the AGB mass loss process because shaping by the hot, fast stellar wind of a PN nucleus has probably not occured (see Schonberner in this volume). Hence we can determine “initial conditions” for the interacting winds models. Figure 1 shows our working model for a PPN circumtellar shell based on the observed evidence that most PN are axially symmetric (e.g. Balick 1987) while the outer shells of AGB circumstellar shells are spherically symmetric (e.g. Neri et al. 1998). In the PPN fossil record, radial distance directly corresponds to time. The maximum radius, $`\mathrm{R}_{\mathrm{max}}`$ marks when the mass loss began. As we move inwards, we see the spherically symmetric shell created by the AGB wind. The superwind radius, $`\mathrm{R}_\mathrm{S}`$, corresponds to when the mass loss began to increase in rate and to assume an axial as opposed to spherical symmetry. We note that our use of the term superwind is slightly modified from the intention of Iben & Renzini (1983) in that we include the symmetry change in addition to an increase in mass loss rate. The inner radius, $`\mathrm{R}_{\mathrm{in}}`$ marks when the mass loss stopped and the size of the inner radius reflects the dynamical age that has passed since the star left the AGB (Meixner et al. 1997). Studies of PPN morphologies have used tracers of dust (thermal infrared radition or optical/near-infared scattered starlight) or molecular gas (CO maps or maser maps). Since the molecular gas observations are covered by others in this volume (see Huggins; Alcolea; Fong et al.), I will focus on the tracers of dust. In PPN, the central star heats the dust and the dust both scatters the starlight and radiates in the thermal infrared ($`\lambda >5\mu `$m). In our working model, we expect that the scattered starlight will preferentially leak out of the lower density bipolar openings of the dustshell and will be the most intense in the inner regions where both the starlight and dust density are highest. In the thermal infrared, what we see depends on the wavelength we observe because dust radiates as a modified black body and hence depends sensitively on temperature. In these dust shells, the temperature decreases from at $``$200 K at the inner radius to $``$30 K at the outer radius. Mid-infrared (Mid-IR; 8-25$`\mu `$m) emission arises exclusively from the inner regions where the dust is warm. Far-IR and submillimeter radiation ($`>`$50 $`\mu `$m) arises from the outer regions as well as the inner regions, however, the angular resolution at these wavelengths is presently quite poor (10-40<sup>′′</sup>) which prevents investigation of the inner, axisymmetric regions. ### 3.1. Mid-IR Imaging Studies Ground based mid-IR imaging studies of PPN have had typical angular resolutions of about 1<sup>′′</sup> and enough spectral resolution to separate dust features and dust continuum. The larger telescopes coming on line, e.g. Keck, VLT, Gemini and MMT, promise diffraction limited performance as good as 0.<sup>′′</sup>2 (e.g. see Morris in this volume, and Jura & Werner 2000). A number of published mid-IR imaging studies of PPN have focused on one to five well resolved PPN and usually include radiative transfer modeling (Skinner et al. 1994; Hawkins et al. 1995; Hora et al. 1996; Dayal et al. 1998; Meixner et al. 1997; Skinner et al. 1997). A recent survey paper of 66 PPNe (Meixner et al. 1999) provides the largest data base of PPNe candidate mid-IR images to date. It also incorprates the results of previously published works. Considering all the published mid-IR images to date, there are three main points that can be made (Meixner et al. 1999). First, of the 73 PPNe candidates, 33% have been resolved with $``$1<sup>′′</sup> resolution. The cooler and brighter objects are easier to resolve probably because cooler shells are more distant from the central star and hence larger and brighter shells are either closer or more luminous which create larger mid-IR emission regions. Second, all of the well resolved PPNe are axisymmetric. Thirdly, there appear to be two morphological types in the well-resolved mid-IR PPNe candidates which we have called toroidal and core/elliptical. Toroidals, as exemplified by IRAS 07134+1005 (Fig. 2), have elliptical/round outer perimeters and two peaks which can be interpreted as limb brightened peaks of an equatorial density enhancement. The central star usually appears between the two peaks. Core/ellipticals, as exemplified by the Red Rectangle (Fig. 2), have tremendously bright, unresolved cores at their centers and low surface brightness elliptical shaped nebulae surrounding these cores. The extension of the low surface brightness emission is in the same direction as the optical reflection nebulosity found in these objects. ### 3.2. Optical Polarimetry and Imaging Studies The optical and near-IR polarimetry and imaging observations of the scattered starlight in PPN predate (1970’s) and far out number these mid-IR studies. Here I only have room to summarize some recent work. Two large survey polarimetry studies of PPN have revealed a large amount of polarization at the PPN stage which indicates that the PPN stage is axisymmetric. Using broad band polarimetry, Johnson & Jones (1991) investigated 38 objects ranging from the AGB to PPN to PN stages and found that the polarization increased from the AGB to PPN stage and then decreased in the PN stage. Using spectropolarimetry, Trammell, Dinerstein & Goodrich (1994), studied 31 PPN and found 80% had some intrinsic polarization. They also classified polarized PPN into Type 1 and Type 2. Both have large polarizations, but Type 1’s also have a large position angle rotation with wavelength which suggests that Type 1’s may be more bipolar in shape. See Gledhill et al. and Su et al. in this volume for recent near-IR polarimetry work on PPN. High angular resolution optical and imaging studies of PPN are numerous and have exploded with use of HST because the compact nature of PPNe is well suited for HST. A number of the studies focus on one or a few objects and range from phenomenological discussions to quantitative modelling in their interpretations (e.g. Sahai, Bujarrabal & Zijlstra 1999; Sahai et al. 1999; Kwok, Su & Hrivnak 1998; Su et al. 1998; Sahai et al. 1998; Kwok et al. 1996; Skinner et al. 1997; Trammell & Goodrich 1996; Bobrowsky et al. 1995; Latter et al. 1993; in this volume see Trammell et al., Hrivnak et al., Bobrowsky et al. and Bieging et al.). Two recent papers cover a significant number of PPN and strive for a more global picture of PPN. Hrivnak et al. (1999) pursued a ground based imaging study of 10 PPN with angular resolutions of $``$0.<sup>′′</sup>75. Ueta, Meixner & Bobrowsky (2000) imaged 27 PPN using the HST WFPC2 with angular resolutions of 0.<sup>′′</sup>046 (see also Ueta et al. in this volume). If we combine the imaging results from all these papers, we find that 80% of the 44 PPN have resolvable optical reflection nebulosities and all of the well resolved reflection nebulosities are axisymmetric. While many of these studies have focused on PPN with obscured central stars, the Ueta et al. (2000) survey included many PPN with optically visible, prominent stars and they discovered two types of optical morphology in the PPN. ## 4. Two Types of PPN The two types of PPN have been called DUst-Prominent Longitudinally-EXtended (DUPLEX) and Star-Obvious Low-level Elongated (SOLE) PPN (see Ueta et al. in this volume). These names describe their optical morphological appearance and their acronyms describe the two lobed structures seen in DUPLEX PPNe and the continuous structures seen in SOLE PPNe. Besides their optical appearances, DUPLEX and SOLE PPNe differ in their mid-IR morphologies: DUPLEX are core-ellipticals, while SOLE are toroidals. They also have distinctly different SEDs: DUPLEX have type II or III SEDs, while SOLE have type IV SEDs in the van der Veen et al. 1989 classification. Ueta et al. (2000) claim that the cause of these differences is the optical depth of the dust shell. SOLE nebulae have less dust optical depth than DUPLEX nebulae and, hence, the central star is visible no matter the inclination angle. They further suggest that DUPLEX PPNe may well be the precursors of bipolar PNe while SOLE PPNe may be the precursors to the elliptical PNe based on their differences in morphologies and galactic height distributions. This interpretation is controversial, judging by the avid discussion after my talk. The other point of view, presented by Hrivnak (in this volume) is that these optical morphological differences are due primarily to inclination angle differences. That is, all these PPNe are the same physical type of beast, just viewed from different angles on the sky. The best way to resolve such controversy is to make radiative transfer models using axially symmetric dust codes to derive optical depths, and structures for all of the PPNe and compare their derived properties. Here, we present model results of one DUPLEX PPN, IRAS 17150-3224, and one SOLE PPN, IRAS 17436+5003, to demonstrate that these two sources, which are among the best examples of their classes, are physically quite different. We use a radiative transfer code that we used in Meixner et al. (1997) and Skinner et al. (1997). Su et al. (1998 and in this volume) are using a different code with similar aims but have concentrated on primarily on DUPLEX PPN so far. We have constrained the model using our HST and mid-IR images and photometry from the literature. A full discussion of these models will appear in Meixner, Ueta & Bobrowsky (2000). Comparison of the model images and SED data demonstrates a reasonable match of the model with the data (Figs. 4 and 5). The derived parameters from the models appear in the Table and reveal the physical reasons for the apparent morphological differences. First, both objects are best fit by a 90 inclination angle; i.e. we are viewing both edge-on. Thus, we are not viewing the SOLE PPN, IRAS 17436+5003, near the pole which would be expected if our viewing angle were the main cause of the morphological differences. We note that other PPN in the Ueta et al. (2000) sample show qualitative evidence for inclination angles different than 90; e.g. unbalanced lobes for DUPLEX and less elliptical nebulae for SOLE. Second, the optical depth for the DUPLEX PPN, IRAS 17150-3224, is significantly higher than for the SOLE PPN, IRAS 17436+5003 and explains why we do not see the central star in the former but do in the latter. The cause for the difference in optical depth is the history of mass loss. IRAS 17150-3224 experienced a more intensive mass loss rate than IRAS 17436+5003 that resulted in a denser dust cocoon of significantly higher mass. Both this higher mass and the higher luminosity for this source suggest that IRAS 17150-3224 originated from a higher mass star. These results are, of course, distance dependent and the distance maybe uncertain by a factor of two. However, the optical depth is distance independent and even with the distance uncertainty it seems reasonable to conclude that IRAS 17150-3224 is more luminous and had a higher mass progenitor. ## 5. Summary points The observational evidence from a number of independent studies clearly shows that PPNe are intrinsically axisymmetric. Thus the axisymmetry that we observe in PNe must predate the PPNe stage. Most likely the axisymmetry originates at the end of the AGB phase, because observations of the outer regions of AGB envelopes show a spherical symmetry. With the variety of PNe morphologies discussed at this conference (e.g. round, elliptical, bipolar and point symmetric), we must now begin to ask: Do we see examples of PPNe with these corresponding subtle differences in morphologies? I think we are beginning to see these differences. The DUPLEX and SOLE PPNe may well be the precursors for bipolar and elliptical PNe. ## References Balick, B. 1987, AJ, 94, 671 Bobrowsky, M. et al. 1995; ApJ, 446, L89 Dayal, A., Hoffmann, W.F., Bieging, J.H., Hora, J.L, Deutsch, L.K., & Fazio, G.G. 1998, ApJ, 492, 603 Hawkins, G.W., Skinner, C.J., Meixner, M.M., Jernigan, J.G., Arens, J.F. Keto, E., and Graham, J. 1995, ApJ, 452, 314. Hora, J.L, Deutsch, L.K., Hoffmann, W.F., Fazio, & Giovanni, G. 1996, AJ, 112, 2064 Hrivnak, B.J. & Kwok, S. 1991, ApJ, 371, 631 Hrivnak, B.J., Kwok, S. & Volk, K. 1989, ApJ, 346, 265 Hrivnak, B.J. 1997, Proceedings of IAU Symposium No. 180, ed. Habing, H.J. & Lamers H.J.G.L.M. (Dordrecht: Kluwer Academic Publishers), p. 303 Hrivnak, B.J., Langhill, P.P., Su, K.Y.L. & Kwok, S. 1999, ApJ, 513, 421 Hu, J.Y., Slijkhuis, S., De Jong, T. & Jiang, B.W. 1993, A&AS, 100, 413 Iben, I. & Renzini, A. 1983, ARA&A, 21, 271 Jura, M. & Werner, M.W. 2000, ApJ, in press Kwok, S. 1993, ARA&A, 31, 63 Kwok, S., Hrivnak, B.J., Zhang, C.Y., & Langhill, P.L. 1996, ApJ, 472, 287 Kwok, S., Su, K.Y.L, Hrivnak, B.J. 1998, ApJ, 501, L117 Latter, W. B., Hora, J.L., Kelly, D. M., Deutsch, L.K., & Maloney, P.R. 1993, AJ, 106, 260 Meixner, M., Skinner, C.J., Graham, J.R., Keto, E., Jernigan, J.G., & Arens, J.F. 1997, ApJ, 482, 897 Meixner et al. 1999, ApJS, 122, 221 Meixner, M., Ueta, T., & Bobrowsky, M. 2000, ApJLet., submitted Neri, R., Kahane, C., Lucas, R., Bujarrabal, V. & Loup, C. 1998, A&AS, 122, 221 Ney, E. P., Merrill, K. M., Becklin, E. E., Neugebauer, G., Wynn-Williams, C. G. 1975, ApJ, 198, L129 Oudmaijer, R.D., van der Veen, W.E.C.J., Waters, L.B.F.M., Trams, N.R., Waelkens, C. & Engelsman, E. 1992, A&AS, 96, 625 Sahai et al. 1998, ApJ, 493, 301 Sahai, R., Bujarrabal, V. & Zijlstra, A. 1999, ApJ, 518, L115 Sahai, R., Zijlstra, A., Bujarrabal, V., Te Lintel Hekkert, P. 1999, AJ, 117, 1408 Skinner, C.J., Meixner, M.M., Hawkins, G., Keto, E., Jernigan, J.G., and Arens, J.F. 1994, ApJ, 423, L135 Skinner et al. 1997, A&A, 328, 290 Su, K.Y.L., Volk, K., Kwok, S., & Hirvnak, B.J. 1998, ApJ, 508, 744 Trammell, S.R. & Goodrich, R. W. 1996, ApJ, 468, L107 Ueta, T., Meixner, M. & Bobrowsky, M. 2000, ApJ, 528, in press van der Veen, W.E.C.J., Habing, H.J. & Geballe, T.R. 1989, A&A, 226, 108 Van Winckel, H. 1999, Proceedings of IAU Symposium No. 191, ed. Le Betre, T., Lebre, A. & Waelkens, C. (San Francisco: ASP), p. 465
warning/0002/cond-mat0002446.html
ar5iv
text
# Inter-Landau-level skyrmions versus quasielectrons in the 𝜈=2 quantum Hall effect ## I Introduction Ferromagnetic quantum Hall systems are known to have charged excitations, skyrmions, that involve texturing of spin. So far, these skyrmions have been considered only at odd integer filling factors and fractional filling factors smaller than 1. Skyrmions were predicted to be the lowest energy charged excitations in, for example, GaAs at filling factor $`\nu =1`$. This was later confirmed in several experiments. Under some conditions skyrmions are also predicted to form at higher odd filling factors and there are experiments claiming to have seen this. Since skyrmions involve moving particles to unoccupied spin-flipped states, they can be candidates for the lowest energy excitations only when there is a small single-particle gap for such spin flips. In most materials spin-orbit coupling decreases the effective Landé factor $`g`$ of the electrons—in for example GaAs $`g`$ is effectively reduced from 2 to $`0.44`$. At the same time the cyclotron gap $`\mathrm{}\omega _c`$ is normally increased due to the small effective mass of the conducting electrons in these materials. This means that the Zeeman energy $`g\mu _BB`$ will normally be small compared to the cyclotron gap $`\mathrm{}\omega _c`$. Hence small single-particle gaps will be found at odd filling factors. There are, however, materials where spin-orbit coupling is so strong that the magnitude of the Zeeman energy is instead strongly enhanced. In InSb the effective Landé factor $`g`$ is of the order of $`50`$. The ratio of the Zeeman energy to the cyclotron energy gap can also, as always, be increased by tilting the sample relative to the magnetic field. This makes it possible to reach a limit where the spin-split single-particle energy levels from the lowest and next lowest orbital Landau levels can come very close together or even cross (see Fig. 1). In this case there is a small single-particle gap at even filling factors and one can imagine the possibility of having “inter-Landau-level” skyrmions—these would be charged excitations that involve spins flipped from one orbital Landau level to another. In this paper we first review the paramagnetic to ferromagnetic phase transition that occurs at filling factor $`\nu =2`$ when single-particle levels from different orbital Landau levels come close. We present finite width calculations of how the Coulomb interaction affects this transition and comment on recent experiments at higher filling factors by Papadakis et al. on the spin splitting in AlAs quantum wells. We then investigate the charged excitations in the possible $`\nu =2`$ quantum Hall states and find that inter-Landau-level skyrmions are never the lowest energy charged excitations in this system, but do provide an effective driving mechanism for the phase transition and thereby limit the maximal hysteresis attainable in this case. ## II The Paramagnetic to Ferromagnetic Transition In the limit when the Coulomb energy is negligible compared to the gaps in the single-particle spectrum, we can specify the groundstate of a quantum Hall system at integer filling factors by giving the order in which the single-particle energy levels are filled. For finite but small Coulomb interaction, spin polarized states will be favored and the order in which levels are filled is determined by a competition between single-particle energies and the Coulomb energy. In the case when the cyclotron gap $`\mathrm{}\omega _c`$ is larger than both the Zeeman energy $`g\mu _BB`$ and the characteristic Coulomb energy $`e^2/ϵ\mathrm{}`$, the $`\nu =2`$ ground state is simply the lowest orbital Landau level with both spin states filled. This state is paramagnetic with no net polarization. Now if the Zeeman energy is increased the system will undergo a first order phase transition to the ferromagnetic state with the same spin filled in the two lowest Landau levels. The Coulomb interaction will cause this transition to occur before the single-particle energies cross. Evaluating the Coulomb energy in the paramagnetic state $`E_C^{\text{PM}}`$ and the ferromagnetic state $`E_C^{\text{FM}}`$ yields $`E_C^{\text{PM}}`$ $`=`$ $`2E_{00},`$ (1) $`E_C^{\text{FM}}`$ $`=`$ $`E_{00}+E_{11}+2E_{01}.`$ (2) Here $`E_{MN}`$ is the total exchange energy contribution of the filled level $`M`$ to level $`N`$, $`E_{MN}=`$ $``$ $`\frac{1}{2}{\displaystyle \underset{m,n}{}}{\displaystyle }d^2r{\displaystyle }d^2r^{}V(|𝐫𝐫^{}|)\times `$ (4) $`\varphi _{Mm}^{}(𝐫)\varphi _{Nn}(𝐫)\varphi _{Nn}^{}(𝐫^{})\varphi _{Mm}(𝐫^{}),`$ where $`\varphi _{Mm}`$ is the wave function with (angular) momentum $`m`$ in Landau level $`M`$. In the ideal (zero thickness) two-dimensional case $`E_{MN}`$ can be evaluated exactly, $$E_{MN}=N_\varphi \frac{1}{\sqrt{2}}_0^{\mathrm{}}𝑑te^{t^2}L_M(t^2)L_N(t^2),$$ (5) where $`L_M`$ are the Laguerre polynomials and $`N_\varphi `$ is the number of flux quanta in the system, i.e., the number of particles in each level. The difference in Coulomb energy between the two states, $`E_C^{\text{PM}}E_C^{\text{FM}},`$ in this ideal two-dimensional (2D) case is $`\frac{3}{8}\sqrt{\pi /2}N_\varphi `$; this means that the ferromagnetic state will have the lowest total energy when $$ϵ\mathrm{}\omega _cg\mu _BBϵ_c=\frac{3}{8}\sqrt{\frac{\pi }{2}}0.470,$$ (6) where energies here and onward are given in units of $`e^2/ϵ\mathrm{}`$. This instability may seem large but is drastically reduced when the finite width of the 2D electron gas is taken into account. Using a Gaussian subband approximation we calculate $`ϵ_c`$ numerically for different effective widths $`w`$. Results are given in Table I where $`w`$ is given in units of the magnetic length. The finite width dependence of $`ϵ_c`$ closely follows the form $$ϵ_c(w)=\frac{a}{w+b},$$ (7) where $`a0.170`$ and $`b0.359`$. It is worth noting that for small enough densities (and hence small magnetic fields) the ferromagnetic $`\nu =2`$ state has the lowest energy even for a vanishing Zeeman energy. This happens because the Coulomb energy—proportional to $`\sqrt{B}`$—is larger than the cyclotron energy—proportional to $`B`$—for small enough $`B`$. Again the critical density for this ferromagnetic ordering depends strongly on the width of the 2D gas. We can easily predict it for some common materials, now keeping the Zeeman energy at zero tilt angle at $`\nu =2`$; see Table II. In GaAs and InSb these critical densities are too low to be experimentally relevant. In AlAs we use an effective mass of $`m^{}=0.46m_e`$ corresponding to the 2D electron gas occupying ellipsoidic constant-energy surfaces with one major axis in the plane of the 2D interface and find a critical density that is larger than that used in experiments. This is compatible with experiments by Papadakis et al. who report on measurements where they see a fully polarized $`\nu =3`$ state at zero tilt angle in this system. We must note that, since the Coulomb energy normally is larger than the cyclotron gap in AlAs, the true ground state will contain a large mixture of different Landau levels and the simple analysis here is not expected to be precise. Nevertheless, the same method applied to the $`\nu =3`$ ground state with a physical width of 150 Å (as used in the setup of Papadakis et. al.) yields a critical density of $`3.8\times 10^{11}`$ cm<sup>-2</sup> below which the $`\nu =3`$ groundstate should be fully polarized. This agrees qualitatively with the range of densities used in this experiment \[(1.4–3.9)$`\times 10^{11}`$ cm<sup>-2</sup>\]. ## III The charged excitations We have used a time dependent Hartree-Fock method to look for spin-flip instabilities around a quasielectron or quasihole in both the paramagnetic and ferromagnetic $`\nu =2`$ ground states. This instability analysis is equivalent to taking the small spin limit of an inter-Landau-level skyrmion and will tell us whether or not the charged excitations involve extra flipped spins. Before describing the calculation we want to note that Kohn’s theorem, stating that electron-electron interactions cannot affect the cyclotron resonance in a translationally invariant system, does not imply anything for the excitations in this case since Kohn’s theorem does not apply when the excited state has a different spin from the ground state. In the paramagnetic case we write the ground state as $$\psi _0_{\text{PM}}=\underset{m=0}{}c_{0,m,}^{}c_{0,m,}^{}0,$$ (8) where $`c_{M,m,\sigma }^{}`$ creates an electron in Landau level $`M`$ with angular momentum $`m`$ and spin $`\sigma `$. We use the symmetric gauge where $`m`$ takes integer values from $`M`$ and upward. Next we create a charged excitation (a quasihole) at the origin by removing one electron from the upper level. To find out if further spin flips are favored we allow an inter Landau level spin wave around the hole, $$\mathrm{\Psi }^{}(q)c_{0,0,}\psi _0_{\text{PM}},$$ (9) where $`\mathrm{\Psi }^{}(q)`$ is a spin wave operator that flips a spin from the $`M=0`$ spin-down level to the $`M=1`$ spin-up level with change $`q`$ in momentum, $$\mathrm{\Psi }^{}\left(q\right)=\underset{k=1}{}\alpha _kc_{1,k+q,}^{}c_{0,k,}.$$ (10) We determine the $`\alpha _k`$’s and the corresponding energies by numerically diagonalizing the Coulomb Hamiltonian in this subspace for each $`q`$. Note that $`q`$ can take values from $`2`$ and upward here. For $`q=2`$ we find a negative eigenvalue of $`0.269`$ in the ideal 2D case. This means that the charged excitations will involve extra spin flips and form a charged spin-texture excitation—a skyrmion—if the single-particle gap to the next level is smaller than this instability, i.e., $`ϵ0.269e^2/ϵ\mathrm{}`$. However, this value is smaller than the instability to the ferromagnetic state described in the previous section ($`ϵ_c=0.470`$); thus the system will in this case undergo the transition to the fully polarized state before skyrmions become the preferred quasiparticles. Since this is a question of energetics it is not guaranteed that the same conclusion holds for finite widths of the 2D electron gas. Indeed, at filling factor $`\nu =3`$ skyrmions are predicted to form for finite widths but not in the ideal 2D case. In the present case, however, the two instabilities both decrease at roughly the same rate and the polarized electron/hole quasiparticle is favored up to a width of at least five magnetic lengths. In the ferromagnetic state, $$\psi _0_{\text{FM}}=\underset{m=0,n=1}{}c_{1,n,}^{}c_{0,m,}^{}0,$$ (11) the system can gain single-particle energy by flipping one spin from the filled $`M=1`$ spin-up level to the empty $`M=0`$ spin-down level below it. By first diagonalizing the Coulomb Hamiltonian of this spin wave with no hole present, we confirm that any such spin flip always costs more than the maximum possible single-particle gain—the lowest energy spinwave costs $`0.51e^2/ϵ\mathrm{}`$. Hence the ground state is stable and the phase transition is first order as promised. We then redo the diagonalization with a quasihole present and find that the spin wave is unaffected by the localized excitation, so in this case the spin wave again costs more Coulomb energy than the potential single-particle gain. Hence there is no situation where a skyrmion can have lower energy than an electron/hole quasiparticle in this case either. Since the transition between the paramagnetic and ferromagnetic phase is first order, it should be possible to “supercool” the system and stay in the metastable state. One could imagine, for example, starting in the stable paramagnetic phase and then slowly increase the Zeeman splitting while keeping the filling factor fixed close to $`\nu =2`$. This could be realized by rotating the sample in situ while increasing the total magnetic field. When the single-particle gap $`ϵ`$ falls below $`ϵ_c`$ ($`=0.47`$ in the 2D case) the paramagnetic state ceases to be the true ground state. However, there may still be no effective way for the system to overcome the barrier separating the two phases, in which case it will stay in the metastable paramagnetic phase for a very long time. In a region below $`ϵ_c`$ the barrier is large since a single spin flip still costs a large energy (even though we know that flipping all spins would take us to the true ground state). Naively, one could hope to form skyrmions by supercooling the paramagnetic state beyond the point where the skyrmion instability sets in (i.e., $`ϵ<0.269`$). But since forming a skyrmion does involve flipping spins it seems likely that instead of forming skyrmions in the metastable paramagnetic state this spin-flip instability rather provides an effective way for the system to undergo the phase transition and actually drives it to the true (ferromagnetic) ground state. Note that this instability will be present even at $`T=0`$ since there will always be a finite density of charged excitations in the system (except at the exact center of the $`\nu =2`$ plateaux). Once the instability sets in, these charged excitations will act as nucleation centers for the other phase and thus this instability sets an upper limit for how large a hysteresis one can obtain when approaching $`T=0`$. We can also note that the hysteresis is asymmetric in the sense that the paramagnetic state can survive as metastable further into the ferromagnetic region than the ferromagnetic state can into the paramagnetic region (the hysteresis is limited to $`0.269<ϵ<0.51`$, with $`ϵ_c=0.47`$ rather close to the upper of these limits). ## IV Conclusion We have investigated the nature of the charged excitations in a quantum Hall system with large Zeeman energy at filling factor $`\nu =2`$, and find that there is no instability to flip extra spins around the polarized quasiparticles either in the paramagnetic or in the ferromagnetic ground state. Hence no inter-Landau-level skyrmions can be the lowest energy charged excitations in this system. The skyrmion instability does, however, limit the possible hysteresis attainable in the paramagnetic to ferromagnetic phase transition. We also presented finite thickness calculations for the first order phase transition between the paramagnetic and the ferromagnetic ground state and predict some critical densities below which the ferromagnetic state has lowest energy even for zero tilt angle. ## V Acknowledgments The author wishes to thank Anders Karlhede, Kennet Lejnell, and Hans Hansson for numerous useful discussions. The author also thanks K. Mullen, N. Cooper, H. Fertig, G. Khodaparast, and S. Das Sarma.
warning/0002/astro-ph0002047.html
ar5iv
text
# THE PRIMORDIAL HELIUM ABUNDANCE: TOWARDS UNDERSTANDING AND REMOVING THE COSMIC SCATTER IN THE 𝑑⁢𝑌/𝑑⁢𝑍 RELATION ## 1 INTRODUCTION An accurate measurement of the primordial helium abundance would be an important test of standard big bang nucleosynthesis (Olive, Steigman & Skillman, 1997), and would also constrain the values of the photon-to-baryon ratio and $`\mathrm{\Omega }_b`$ (Olive, Steigman & Walker, 1999). The traditional procedure to measure the primordial helium abundance is to make use of the correlation between the helium mass fraction ($`Y`$) and metal abundance ($`Z`$). This correlation is then extrapolated to zero metallicity to estimate the primordial mass fraction of helium, $`Y_p`$. Spectroscopic observations of bright, low-metallicity extragalactic HII regions provide the data for these studies (e.g., Olive & Steigman, 1995; Olive, Steigman & Skillman, 1997; Izotov, Thuan & Lipovetsky, 1994, 1997; Izotov & Thuan, 1998; Torres-Peimbert, Peimbert & Fierro, 1989; Skillman, Terlevich & Terlevich, 1998). To be cosmologically useful the value of $`Y_p`$ has to be determined to better than 5%. Fortunately, abundance determinations from measurements of line ratios is relatively straightforward (Peimbert, 1975; Benjamin, Skillman & Smits, 1999), and can, in theory, give the desired accuracy. However, to reach the needed level of precision, any systematic errors involved with target selection, observations, and data analysis must be identified and corrected. Many such systematic errors have already been identified (Davidson & Kinman, 1985; Dinerstein & Shields, 1986; Pagel et al., 1992; Skillman et al., 1994; Peimbert, 1996; Izotov, Thuan & Lipovetsky, 1997; Steigman, Viegas & Gruenwald, 1997; Skillman, Terlevich & Terlevich, 1998), but any errors resulting from the so-called “ionization correction factor” (ICF) have so far been assumed to be small. The ICF corrects for the fact that some amount of atomic (i.e., unseen) helium might be present in ionized regions of hydrogen (Osterbrock, 1989; Peimbert, 1975). This correction traditionally has been assumed to be zero because measurements of the primordial helium abundance employ observations of bright extragalactic HII regions. These regions are excited by clusters of young stars with effective temperatures greater than 40,000 K. Calculations by Stasinska (1990) and Pagel et al. (1992) showed that the helium ICF should be negligibly small for these HII regions. As a result, recent determinations of $`Y_p`$ have assumed that the helium ICF is small. Very recently, Armour et al. (1999) presented calculations that showed that HII regions excited by stars with temperatures greater than 40,000 K can have non-negligible ICFs. Armour et al. (1999) found that the ICFs were often negative (i.e., the helium ionized zone is larger than the hydrogen one; Stasinska (1980, 1982), Peña (1986)) for the hardest stellar continua. These results were confirmed by Viegas, Gruenwald & Steigman (2000). In this paper, we follow up on the work of Armour et al. (1999), and develop observational diagnostics of when the He ICF is important and when it can be ignored. We then apply these diagnostics to the data of Izotov & Thuan (1998) to illustrate how our technique can improve the precision of the measurement of $`Y_p`$. We describe our calculations in § 2, and our results in § 3. The main results are summarized in § 4. ## 2 DESCRIPTION OF CALCULATIONS In order to investigate the effects of a non-negligible ICF on the determination of the primordial helium abundance, we ran photoionization models of HII regions and extracted the ICF for each nebula. These calculations are very similar to ones presented by Armour et al. (1999) and Bottorff et al. (1998), and were made with the development version of Cloudy, last described by Ferland et al. (1998). Since we are modeling HII regions, our models use the ISM abundances and grain model that were used and described by Armour et al. (1999). However, we scaled both the metal and grain abundances to lower values because, in this case, we are most interested in lower metallicity nebulae. The scaling was implemented so that all metals and grains were varied together relative to hydrogen and helium, but the He/H ratio was held constant. We modeled nebulae at three different metallicities: O/H=32, 64, and 128 parts per million \[ppm\] ($`Z=Z_{}/23,Z_{}/12`$, and $`Z_{}/6`$ respectively). For each metallicity, 1936 models were computed for each of the following spectral energy distributions: the LTE plane-parallel atmospheres of Kurucz (1991), the non-LTE, wind-blanketed, solar abundance CoStar atmospheres of Schaerer et al. (1996a, b), the earlier NLTE atmospheres of Mihalas (1972), and for completeness blackbodies. We also ran models using the subsolar-abundance CoStar atmospheres which are spectrally slightly harder than the solar abundance ones. These models resulted in slightly more negative ICFs, but the values of the line ratio cutoffs (§ 3.2 & § 3.3) were not changed from the ones calculated with a solar abundance atmosphere. For each atmosphere we computed models with 10 cm$`{}_{}{}^{3}n_H10^6`$ cm<sup>-3</sup>, 10$`{}_{}{}^{4}U10^{0.25}`$, and 40,000 K$`T_{eff}50,000`$ K, where $`U`$ is the ionization parameter defined as in Eq. 4 of Armour et al. (1999). Giant extragalactic HII regions that are observed are generally excited by large clusters, so this range of parameters should cover all such nebulae. We modeled the nebulae as plane-parallel constant density slabs, a simple way to characterize blister HII regions. Our proposed diagnostic indicators are the \[O III\] $`\lambda `$5007 and \[O I\] $`\lambda `$6300 lines (§ 3.2 & § 3.3), and these should be fairly independent of the assumed geometry (sphere, sheet, or evaporating blister). The \[O III\] $`\lambda `$5007/H$`\beta `$ ratio represents the cooling per recombination (the Stoy ratio) and so is primarily sensitive to the stellar temperature (Stoy, 1933; Kaler, 1978) rather than geometry. Similarly, the \[O I\] $`\lambda `$6300/H$`\beta `$ intensity ratio mostly measures the ”softness” of the hydrogen ionization front (where the line forms - Netzer & Davidson 1979). ## 3 RESULTS ### 3.1 Temperature Dependence Figure THE PRIMORDIAL HELIUM ABUNDANCE: TOWARDS UNDERSTANDING AND REMOVING THE COSMIC SCATTER IN THE $`dY/dZ`$ RELATION plots the ICF calculated from the Kurucz and CoStar models versus the stellar temperature. Note that we define the ICF such that an ICF of zero corresponds to zero correction: $$\mathrm{ICF}=\frac{\mathrm{H}^+/\mathrm{H}}{\mathrm{He}^+/\mathrm{He}}1,$$ (1) where the angle brackets denote the volume mean ionization fraction. This definition takes into account the presence of any He<sup>+2</sup> in the nebula. Figure THE PRIMORDIAL HELIUM ABUNDANCE: TOWARDS UNDERSTANDING AND REMOVING THE COSMIC SCATTER IN THE $`dY/dZ`$ RELATION clearly shows that one can obtain a non-negligible ICF for stars with temperatures greater than 40,000 K. The harder CoStar atmospheres give preferentially negative ICFs, which, if not taken into account, would result in a overestimate of the helium abundance. The same is true of the Mihalas atmospheres (not shown). These results agree with the calculations of Armour et al. (1999) and Viegas, Gruenwald & Steigman (2000). The softer Kurucz atmospheres result in preferentially positive ICFs (i.e., the helium ionized zone is smaller than the hydrogen one), although, at temperatures greater than 45,000 K, they can also give negative ICFs. The blackbody atmospheres, the least realistic, are softer still, but even they can produce negative ICFs in some models at the highest temperatures. Therefore, negative ICFs seem to be found at high stellar temperatures independent of the type of stellar atmosphere. ### 3.2 Metallicity Dependent Cutoff Criterion A negative ICF occurs as the results of penetrating high-energy photons preferentially ionizing helium, due to its large photoionization cross section. This tends to be important for lower ionization parameter models, since these have significant regions where H and He are partially ionized. We expect these nebula to be characterized by lower \[O III\] $`\lambda `$5007/H$`\beta `$ ratios (a measure of excitation) and larger \[O I\] $`\lambda `$6300/H$`\beta `$ ratios (since \[O I\] $`\lambda `$6300 is formed in warm atomic regions). The results presented in § 3.1 show that it is not appropriate to simply assume that the ICF is zero when a nebula is excited by a star with a temperature greater than 40,000 K. However, it would be important to develop observational diagnostics for when the ICF is important and when it is not. Figure THE PRIMORDIAL HELIUM ABUNDANCE: TOWARDS UNDERSTANDING AND REMOVING THE COSMIC SCATTER IN THE $`dY/dZ`$ RELATION shows such a diagnostic. A plot of ICF versus \[O III\] $`\lambda `$5007/H$`\beta `$ shows that beyond a line ratio of about 3–4 the ICF is negligible (for clarity results are shown for Kurucz and CoStar atmospheres only; a plot for blackbody and Mihalas atmospheres is very similar). However, the value of the cutoff will depend on metallicity. We found very small ICFs for line ratios greater than the following cutoff: $$\left([\mathrm{O}\mathrm{III}]\lambda 5007/\mathrm{H}\beta \right)_{\mathrm{Cutoff}}=(0.025\pm 0.004)\mathrm{O}/\mathrm{H}+(1.139\pm 0.306),$$ (2) where O/H is measured in parts per million. HII regions which have an \[O III\] $`\lambda `$5007/H$`\beta `$ ratio less than the cutoff for their metallicity might be subject to an ICF correction. Unless this correction can be made (and, in general, it cannot) these HII regions should be removed from the abundance analysis, as they will increase the scatter in the $`dY/dZ`$ relation used to determine the primordial helium abundance. ### 3.3 Metallicity Independent Cutoff Criterion One can improve the above result by finding emission line ratios that should be independent of metallicity. Figure THE PRIMORDIAL HELIUM ABUNDANCE: TOWARDS UNDERSTANDING AND REMOVING THE COSMIC SCATTER IN THE $`dY/dZ`$ RELATION plots the helium ICF versus the \[O III\] $`\lambda `$5007/\[O I\] $`\lambda `$6300 ratio. The figure shows that the ICF is negligible for a line ratio greater than about 300. Not surprisingly, this result is independent of metallicity. There are some Kurucz and blackbody models which result in a non-negligible ICF at line ratios larger than this cutoff. These models had low stellar temperatures (40,000–42,000 K), and were found over a narrow range in both $`\mathrm{log}U`$ ($`1.5`$ to $`2.25`$) and $`\mathrm{log}n_H`$ (1.0–3.0). Their positive ICF is a result of combining the low stellar temperatures and the softness of their atmospheres; their large \[O III\] $`\lambda `$5007/\[O I\] $`\lambda `$6300 ratio is a result of combining the fairly high ionization parameter with the low density. CoStar and Mihalas models, which are considered more “realistic”, result in only very small ICFs in this region. Therefore, we find that any HII region, regardless of its metallicity, that has an \[O III\] $`\lambda `$5007/\[O I\] $`\lambda `$6300 ratio less than about 300 might be subject to an ICF correction and should not be used to determine the primordial helium abundance. ### 3.4 Application to Real Data To see how these new results affect the determination of $`Y_p`$, we applied the metallicity independent rejection criteria to the data of Izotov & Thuan (1998), and the results are shown in Figure THE PRIMORDIAL HELIUM ABUNDANCE: TOWARDS UNDERSTANDING AND REMOVING THE COSMIC SCATTER IN THE $`dY/dZ`$ RELATION. There are a number of points to note from this Figure: 1. Our criterion rejects points over the entire range of metallicity, so there is no metallicity bias. The rejected points also fall evenly over the range of $`Y`$ values, which implies that there is no correlation between $`Y`$ and the \[O III\] $`\lambda `$5007/\[O I\] $`\lambda `$6300 ratio. 2. At an \[O III\] $`\lambda `$5007/\[O I\] $`\lambda `$6300 cutoff of 300 there is a 20% reduction in the weighted rms scatter about the best fit line. This is encouraging evidence that part of the scatter was due to the ICF, and the situation has indeed improved by the implementation of the cutoff. 3. The negative slopes predicted by using a cutoff$``$ 300 result from weighted fits to the small number of data that remain after applying the cutoff, and are probably not realistic. Given the current data, a cutoff value greater than 300 is too severe (not practical). 4. Despite the increased errors in the slope and the intercept, we find that implementing our cutoff will result in a larger value of $`Y_p`$. Specifically, we find $`Y_p=0.2489\pm 0.0030`$ when the cutoff is taken at a \[O III\] $`\lambda `$5007/\[O I\] $`\lambda `$6300 ratio of 300. Izotov & Thuan (1998) find $`0.2443\pm 0.0015`$ (our fit gives $`0.2443\pm 0.0013`$ with no cutoff), so the two results are barely consistent at the 1$`\sigma `$ level. This result is in the opposite direction of the shift predicted by the Monte Carlo simulations of Viegas, Gruenwald & Steigman (2000), but moves the value of $`Y_p`$ closer to the theoretically predicted values (Olive & Steigman, 1995). There is always the possibility that systematic errors are introduced whenever data are rejected by a certain criterion. The above selection criterion preferentially selects HII regions that have large \[O III\] $`\lambda `$5007/\[O I\] $`\lambda `$6300 ratios. This ratio is generally large whenever the \[O III\] $`\lambda `$5007 line is strong (this is consistent with our metallicity dependent criterion). The \[O III\] $`\lambda `$5007 line is a major source of cooling in a nebula, and so by selecting HII regions with strong \[O III\] $`\lambda `$5007 lines we are selecting regions ionized by hotter stars. Since these large extragalactic HII regions are generally ionized by clusters, ones with strong \[O III\] $`\lambda `$5007 lines are preferentially younger. However, this is unlikely to introduce a systematic error in the primordial helium abundance determination, as young clusters can form at any metallicity. Indeed, Figure THE PRIMORDIAL HELIUM ABUNDANCE: TOWARDS UNDERSTANDING AND REMOVING THE COSMIC SCATTER IN THE $`dY/dZ`$ RELATION shows that the points rejected by applying the cutoff span the entire range of metallicity. There is also the possibility that physical conditions within the HII regions may bias our results. For example, in some nebulae the intensity of the \[O I\] $`\lambda `$6300 line could be enhanced due to shock heating. This would lower the measured \[O III\] $`\lambda `$5007/\[O I\] $`\lambda `$6300 ratio, and could move it below our selection criterion. However, shock heating would not change the ICF of the nebula, so even though application of our criterion might reject such a HII region, it will not bias the determination of $`Y_p`$. Another potential situation in HII regions is that the nebula may be matter bounded (i.e., optically thin to the Lyman continuum) in certain solid angles or sectors. Because there will be no hydrogen ionization front in such sectors, and therefore no \[O I\] $`\lambda `$6300 line emission, the presence of such sectors will increase the measured \[O III\] $`\lambda `$5007/\[O I\] $`\lambda `$6300 ratio, possibly pushing it above our selection criterion. However, these sectors will also have little or no neutral helium within them, and so will have no ICF. Therefore, although the other, ionization bounded, sectors of the HII region could have a non-negligible ICF, this will be diluted by the matter bounded sectors. We anticipate that even if such an HII region were shifted into our selected data, there ought not be a large effect on determining $`Y_p`$. According to Fig. THE PRIMORDIAL HELIUM ABUNDANCE: TOWARDS UNDERSTANDING AND REMOVING THE COSMIC SCATTER IN THE $`dY/dZ`$ RELATION, to severely bias the results a number of points would have to shift rightward in \[O III\] $`\lambda `$5007/\[O I\] $`\lambda `$6300 by a factor larger than ten; but such a large shift in the line ratio would probably result in a large dilution in the ICF. More modeling would be needed to quantify how little an impact matter bounded sectors would have on the \[O III\] $`\lambda `$5007/\[O I\] $`\lambda `$6300 ratio and ICFs. ## 4 CONCLUSIONS In this paper we have shown the following: 1. There can be a non-negligible ICF correction for HII regions excited by stars with temperatures greater than 40,000 K. At temperatures higher than 45,000 K, the ICF is preferentially negative. This result is independent of the atmosphere of the O star. 2. There is a simple procedure to determine if an ICF correction needs to be made for a given HII region. If the \[O III\] $`\lambda `$5007/\[O I\] $`\lambda `$6300 ratio is greater than 300, then no correction is needed. This criterion is independent of metallicity. If the \[O I\] $`\lambda `$6300 line cannot be measured, then there is a metallicity dependent cutoff (Eq. 2) that can be used with the \[O III\] $`\lambda `$5007 line. 3. Applying the metallicity independent criterion to the data of Izotov & Thuan (1998) results in reducing the rms scatter about the best fit $`YZ`$ line by 20%. This will help remove systematic errors relating to unrecognized ICF effects, and ought to improve the reliability of the $`Y_p`$ determination. Furthermore, an analysis of the selected data gives a larger value of $`Y_p`$ than was originally measured, which is closer to the theoretically expected value. G.J.F. thanks CITA for its hospitality during a sabbatical year and acknowledges support from the Natural Science and Engineering Research Council of Canada through CITA. D.R.B. also acknowledges financial support from NSERC. We acknowledge useful comments from an anonymous referee.
warning/0002/hep-th0002106.html
ar5iv
text
# Abstract ## Abstract <br> We explore gauge fields – strings duality by means of the loop equations and the zigzag symmetry. The results are striking and incomplete. Striking—because we find that the string ansatz proposed in satisfies gauge theory Schwinger-Dyson equations precisely at the critical dimension $`D_{\text{cr}}=4`$. Incomplete—since we get these results only in the WKB approximation and only for a special class of contours. The ways to go beyond these limitations and in particular the OPE for operators defined on the loop are also discussed. February 2000 ## 1 Introduction Gauge fields – strings duality is an old and fundamental subject. It underwent a rapid and fascinating development in the last three years. In this duality color-electric flux lines of gauge theory are described as certain relativistic strings. It has been shown in that the “natural habitat” for these strings (and thus for the flux lines) is a five-dimensional curved space with the metric $$ds^2=d\varphi ^2+a^2(\varphi )d\stackrel{}{x}^2.$$ (1.1) The function $`a^2(\varphi )`$ (which represents the running string tension) must be of a special type. In order to properly describe the zigzag-invariant Wilson loop, it must have a horizon, where $`a^2(\varphi _H)=0`$, and an infinity where $`a^2(\varphi _I)=\mathrm{}`$. The precise form of $`a^2(\varphi )`$ is determined from the condition of conformal invariance on the world sheet. The gauge fields – strings duality was formulated in as an isomorphism between the closed string vertex operators and the gauge invariant operators of gauge theory. Another development was related to the $`𝒩=4`$ SYM theory and the 3-branes in the type IIB strings. It has been shown there that some of the SYM correlation functions (describing absorption of the external particles by the 3-brane) can be calculated by the use of the classical supergravity in which the 3-brane is replaced by the metric (1.1). It was further conjectured that the correspondence extends beyond the SUGRA approximation and the function $`a^2(\varphi )`$ in this case must correspond to the AdS space ($`a^2(\varphi )e^{\alpha \varphi }`$). In the subsequent papers , it was shown how to implement the above-mentioned isomorphism in the $`𝒩=4`$ SYM theory. After that the correspondence in this case has been confirmed in an almost infinite number of papers. But all is not well. There is no real understanding (beyond the heuristic arguments of -) why and, more importantly, when the correspondence works. Ideally, one would like to check that the Schwinger-Dyson equations of the Yang-Mills theory can be obtained from the string representation. This would clarify which string theories must be used for the various gauge theories. It is important to stress in this respect that the help of the D-branes in answering the above question is not always available. There are reasons to believe that in the most interesting non-supersymmetric cases the D-brane interpretation of the general $`\sigma `$-model metric (1.1) is not possible. In these cases the Schwinger-Dyson equations (formulated as loop equations) is our only tool. In the vast literature on the subject there have been occasional attempts to explore the loop equations, but no conclusive results have been reached so far. The reason, we believe, lies in some common misconceptions concerning loop equations. It is generally thought that the loop operator is singular and can be applied only to the regularized and non-universal Wilson loop. This is not necessarily so. In it has already been sketched how to implement the loop equation for the renormalized loops. In the present paper we shall apply the loop operator to the string functional integral. The results are incomplete but quite stunning. We will be dealing only with the very special contours—wavy lines (suggested in ) and only in the WKB approximation. This is obviously incomplete. We will find in this case that, first, the loop operator is well defined in any dimension (implying the zigzag symmetry of the string representation), and second, at the critical dimension $`D_{\text{cr}}=4`$ the loop equation is satisfied. This we find quite stunning. ## 2 The loop equation and the zigzag symmetry The loop equation has been introduced in and further explored in and (see also some related developments in , ). We will give its derivation now, which can be used for the renormalized Wilson loop. It will contain some additional elements to the old works. The Wilson loop is given by $$W[C]=\frac{1}{N}\text{Tr}P\mathrm{exp}_CA_\mu 𝑑x^\mu .$$ (2.1) The averaging in this formula is performed with the Yang-Mills action $$S=\frac{1}{4g_{YM}^2}\text{Tr}F_{\mu \nu }^2(dx)$$ ($`F_{\mu \nu }`$ is the Yang-Mills field strength). The idea of the loop equation is to find an operation in the loop space which, being applied to the LHS of (2.1), will give the Yang-Mills equations of motion at the RHS. To implement this idea, consider the second variational derivative of $`W[C]`$ $`{\displaystyle \frac{\delta ^2W}{\delta x_\mu (s)\delta x_\mu (s^{})}}=\text{Tr}P\left(_\mu F_{\mu \nu }(x(s))e^{{\scriptscriptstyle A_\mu 𝑑x_\mu }}\right)\dot{x}_\nu (s)\delta (ss^{})+`$ $`\text{Tr}P\left(F_{\mu \lambda }(x(s))F_{\mu \sigma }(x(s^{}))e^{{\scriptscriptstyle A_\mu 𝑑x_\mu }}\right)\dot{x}_\lambda (s)\dot{x}_\sigma (s^{})`$ (2.2) (where $`P`$ is the ordering along the contour). The main idea of the loop equation is to separate the first term, and to introduce the “loop Laplacian” which acting on $`W[C]`$ gives the Yang-Mills equation of motion. This can be achieved in several different ways. The problem to overcome is the singularity of the second term at $`s=s^{}`$, which makes it difficult to distinguish it from the first one. The “brute force” method would be to regularize the gauge theory with some cut-off $`\mathrm{\Lambda }`$ and the consider $`|ss^{}|1/\mathrm{\Lambda }`$. The second term is regular in this case, while the first one contains the $`\delta `$-function which is easy to pick up. The price to pay is the non-universal arbitrary regularization. We cannot be sure that on the string side the regularization is the same and thus the comparison of gauge fields and strings becomes strictly speaking impossible. It has already been noticed in that the way out of this difficult is to use the operator product expansion as $`ss^{}`$. Consider a set of operators $`\left\{𝒪_n(x)\right\}`$ which are *not* color singlets. We can in general examine a gauge invariant amplitude $$G_{n_1\mathrm{}n_N}(s_1,\mathrm{}s_N)=\text{Tr}P\left(𝒪_{n_1}(x(s_1))𝒪_{n_2}(x(s_2))\mathrm{}e^{_CA_\mu 𝑑x_\mu }\right).$$ Let us assume that the contour $`C`$ is smooth and non-selfintersecting. In this case we expect the OPE to have the form<sup>1</sup><sup>1</sup>1Such operator products were considered before in . $$𝒪_{n_1}(x(s_1))𝒪_{n_2}(x(s_2))=\frac{C_{n_1n_2}^m(x(s))}{\left|x(s_1)x(s_2)\right|^{\mathrm{\Delta }_{n_1}+\mathrm{\Delta }_{n_2}\mathrm{\Delta }_m}}𝒪_m(x(s))$$ (2.3) (where $`s=\frac{s_1+s_2}{2}`$ and $`\mathrm{\Delta }_n`$ are anomalous dimensions of the corresponding operators). Of course, in asymptotically free theories there are also powers of logarithms $`\mathrm{log}\left|x(s_1)x(s_2)\right|`$ in these formulas, which we do not display. In this way the OPE in the physical 4d space is transplanted to the one-dimensional contour $`C`$. The “open string” amplitudes $`G_{n_1\mathrm{}n_N}(s_1,\mathrm{}s_N)`$ have power singularities at the coinciding points. There is no room for the contact terms proportional to $`\delta (ss^{})`$ if we use renormalized correlators. Moreover, the possible contact terms in the $`x`$-space do not give contact terms in the $`s`$-space. Consider as an example a singularity $$\delta (x(s)x(s^{}))=\underset{\mathrm{\Delta }4}{lim}(4\mathrm{\Delta })\frac{1}{|x(s)x(s^{})|^\mathrm{\Delta }}.$$ As it was explained in , when one uses test functions, $$\delta (x(s)x(s^{}))f(s^{})𝑑s^{}=\underset{\mathrm{\Delta }4}{lim}(4\mathrm{\Delta })\frac{f(s^{})ds^{}}{|x(s)x(s^{})|^\mathrm{\Delta }}=0,$$ provided that the contour has no self-intersections, so that the only singularity in the integral comes from $`s=s^{}`$. It is also assumed that the integral is defined by analytic continuation. As a result of this discussion we conclude that the second variational derivative of the renormalized (and analytically regularized) Wilson loop has the form $$\frac{\delta ^2W}{\delta x_\mu (s)\delta x_\mu (s^{})}=(\widehat{L}(s)W)\delta (ss^{})+\underset{n}{}\frac{C_n(s,\{x(s)\})}{|x(s)x(s^{})|^{4\mathrm{\Delta }_n}},$$ (2.4) where $`C_n(s,\{x(s)\})\text{Tr}P\left(𝒪_n(x(s))e^{{\scriptscriptstyle A𝑑x}}\right)`$. Since $`|x(s)x(s^{})|\sqrt{\dot{x}^2(s)}|ss^{}|`$, the nonlocal power-like behavior of the second term for $`ss^{}`$ can be separated from the local $`s=s^{}`$ singularity of the first one. The loop Laplacian is defined as the coefficient $`\widehat{L}(s)W`$ in front of the $`\delta `$-function, and the loop equation for non-selfintersecting contours has the form $$\widehat{L}(s)W=0.$$ (2.5) In our applications it will be convenient to write (2.4) in the momentum space: $$\underset{p\mathrm{}}{lim}\frac{\delta ^2W}{\delta x_\mu \left(\frac{q}{2}+p\right)\delta x_\mu \left(\frac{q}{2}p\right)}=(\widehat{L}_qW)p^0+\underset{n}{}C_n(q)|p|^{\lambda _n}.$$ (2.6) This is an important formula. It shows that the Wilson loop belongs to a very special class of functionals, for which the asymptotics in (2.6) does not contain integer even powers of $`p`$, corresponding to the derivatives of the $`\delta `$-function. If it does, this means that the functional $`W[C]`$ *cannot* be represented as an ordered exponential (2.1). Some danger for this analysis would be presented by odd integers $`\mathrm{\Delta }_n`$ in (2.4), which would give a $`p^k(\mathrm{log}p+\text{const})`$ contribution. However, the lowest operators which can appear in (2.3), even when protected by non-renormalization theorem, have even dimensions. An exception from this is the operator $`_\alpha F_{\beta \gamma }`$ which can appear in the OPE: $$F_{\mu \lambda }(x(s))F_{\mu \sigma }(x(s^{}))\frac{(x(s)x(s^{}))_\mu _\mu F_{\lambda \sigma }}{\left|x(s)x(s^{})\right|^2}.$$ (We assume here the scenario in which the $`F_{\mu \nu }`$ operator has the normal dimension as a result of some non-renormalization theorem.) However, the singular contribution from this term to (2.2) vanishes after contraction with $`\dot{x}_\lambda (s)\dot{x}_\sigma (s^{})`$. The above analysis can be extended to the functionals represented by the iterated integrals $$W[C]=\underset{n}{}_{s_1<s_2\mathrm{}<s_n}\mathrm{\Gamma }_{\mu _1\mathrm{}\mu _n}^{(n)}(x(s_1),\mathrm{}x(s_n))\dot{x}_{\mu _1}(s_1)\mathrm{}\dot{x}_{\mu _n}(s_n)𝑑s_1\mathrm{}𝑑s_n$$ (2.7) (of which (2.1) is a special case). Such functionals behave well under the application of the loop Laplacian (see ). Hence, the second variational derivative for them has the structure (2.6). Another feature of the functionals (2.7) is the zigzag symmetry . When we change $`s\alpha (s)`$, they are invariant even if $`\alpha ^{}(s)`$ changes sign (and is not a diffeomorphism). It is likely that any zigzag-invariant functional can be represented in the form (2.7), if it satisfies some extra analyticity requirements. Without them many other forms are possible, e.g. $$I_{\alpha \beta }=𝑑s\left(\frac{\ddot{x}_\alpha \dot{x}_\beta \ddot{x}_\beta \dot{x}_\alpha }{\dot{x}^2}\right).$$ At present the precise form of these requirements is not known, although a possibility is that the functional must be nonsingular with respect to $`\dot{x}^2`$. In the following, we will use the term “zigzag symmetry” referring to the form (2.7). According to the above discussion, there is a *practical way* for testing the zigzag symmetry in this sense: take the asymptotics (2.6) and check for the terms *$`p^n`$* with $`n`$—positive even integer. If they are present, there is no place for the zigzag symmetry, and the functional does not represent any Yang-Mills theory. This conclusion, together with the formula (2.6), will be extensively used below. ## 3 The string representation: D-branes and/or $`\sigma `$-model. Let us summarize the facts about string representation of the Wilson loop. According to we must consider a 5d background with the metric (1.1). The origin of the fifth dimension is the Liouville degree of freedom which is unavoidable in the non-critical strings. The $`\sigma `$-model action of this string is given by $`S`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle d^2\xi \sqrt{g}g^{ab}(\xi )G_{MN}(z(\xi ))_az^M_bz^N}+\mathrm{\Phi }(z(\xi ))^{(2)}R(g)\sqrt{g}`$ (3.1) $`+ϵ^{ab}B_{MN}(z(\xi ))_az^M_bz^N+\text{other background fields}+\text{fermions}.`$ We introduced the 5d variable $`z^M=(x^\mu ,y)`$. We are also considering a fermionic non-critical string. The metric $`G_{MN}`$ has the form obtained by the obvious change of variables in (1.1): $$ds^2=\rho (y)(dy^2+d\stackrel{}{x}^2).$$ (3.2) As usual, the background fields are determined from the $`\beta `$-function equation , expressing the $`g^{ab}`$-independence of the theory. It is important to have the NSR fermions for the gauge fields – strings duality. It is also necessary to perform the non-chiral GSO-projection in the functional integral. The whole matter has been treated in the paper . Here we will just remind the reader the underlying logic with a few additional comments. There are two types of objects we would like to consider. First, there are the closed strings amplitudes. We introduce the closed string vertex operators, e.g. $$V_{\mu \nu }=d^2\xi \left(\psi _\stackrel{}{p}(y(\xi ))_ax^\mu _ax^\nu \right)e^{i\stackrel{}{p}\stackrel{}{x}(\xi )}$$ and claim that for the proper choice of the metric $`\rho (y)`$ the correlators of $`V`$’s on the string side are equal to the correlators of the gauge invariant operators on the Yang-Mills side. In other words, we conjecture an isomorphism $$\begin{array}{c}\text{Closed string}\\ \text{states}\end{array}\begin{array}{c}\text{Gauge invariant}\\ \text{operators}.\end{array}$$ (3.3) The intuitive reason for this conjecture is that gauge invariant operators like $`\text{Tr}F_{\mu \nu }^2`$ can be associated with tiny closed flux lines, represented by small Wilson loops. Indeed, we create a flux line by acting on the vacuum with the operator $$\text{Tr}P\mathrm{exp}\underset{C}{}A𝑑x\underset{C0}{}I+C_{\mu \nu \lambda \sigma }\text{Tr}(F_{\mu \nu }F_{\lambda \sigma })+\mathrm{}$$ The propagation of these loops forms the closed world surfaces with punctures represented by the vertex operators. This means that if we find the right action (3.1), the closed string $`S`$-matrix will be equal to the correlation functions of the local gauge invariant operators. But how to discover the right action? This question does not have a complete answer at present. There are two approaches the problem. One approach , , applicable to the supersymmetric Yang-Mills theories, is to start with the 3-branes of the Type IIB superstring. Then, on the one hand, it is known that the low energy excitation of the stack of $`N`$ three-branes are described by the supersymmetric Yang-Mills theory with the $`SU(N)`$ gauge group. On the other hand, 3-branes can be presumably replaced by the supergravity background the generate, and thus one expects and checks the isomorphism (3.3). The main problem with this approach is that it is far from being clear that non-supersymmetric theories can be described in the D-brane language. Also, even in the supersymmetric case, so far there is no real *derivation* of the isomorphism (3.3) from the first principles, although numerous checks have been performed. Another approach is to try to formulate the conditions under which the $`\sigma `$-model satisfies the Schwinger-Dyson equations of the Yang-Mills theory. The basic object in this approach is the Wilson loop, defined by $$W[\stackrel{}{x}(s)]=\underset{\{\begin{array}{c}\stackrel{}{x}|_D=\stackrel{}{x}(s)\\ y|_D=\mathrm{?}\end{array}}{}𝒟y(\xi )𝒟\stackrel{}{x}(\xi )e^{S[\stackrel{}{x}(\xi ),y(\xi )]}.$$ Here we take the world sheet with a disc topology (appropriate for large $`N`$). At the boundary of the disc we fixed $`\stackrel{}{x}`$ by the condition $`\stackrel{}{x}|_D=\stackrel{}{x}(s)`$. Now, we have to determine $`y|_D`$ and the background fields so that the loop equations are satisfied. The basic idea of was to notice that the metric $`\rho (y)`$ must have horizon and/or infinity points, so that $$\rho (y_H)=0\text{and}\rho (y_I)=\mathrm{}.$$ These are the only positions where we can place the contour. Indeed, if we choose some other $`y|_D`$, we will encounter the lack of the zigzag symmetry coming from the fact that the first term in (3.1) is (due to the presence of $`\sqrt{g}`$) invariant only under diffeomorphisms, but not under the zigzag transformations. Hence it must be either zero or infinite at the boundary. A more precise condition was given in . It was stated there that in order to describe gauge theory we must find the metric $`\rho (y)`$, the other background fields, and the boundary conditions $`y|_D`$ such that *the vector vertex operators, defined at the boundary, form a closed algebra*. Physically this means that open strings in this theory correspond to vector gluons. Also, the vector vertex operators are selected by the zigzag symmetry, since they are the only ones defined without the world sheet metric. When the D-brane description is applicable, the above condition is satisfied. But it is important that the condition be formulated without any reference to D-branes. In the case of the $`𝒩=4`$ SYM the D-branes and conformal symmetry considerations lead to the formula $$\rho (y)=\sqrt{g_{YM}^2N}\frac{1}{y^2}.$$ (3.4) It has been shown in , for the local operators and in , for the loops that the natural location for the gauge theory quantities is the infinity in this AdS space, that is $`y0`$. In this paper we shall adopt this prescription, although there is no real proof for it. In a sense, our check of the loop equations in this paper can be considered as such (incomplete) proof. One should keep in mind, however, that there is another zigzag-symmetric location, $`y=\mathrm{}`$. This point has no geometrical significance in the Euclidean signature (the Lobachevsky space) but is meaningful in the Lorentz signature (that is in the AdS space). In the latter case this point represents a horizon behind which the D-branes are hidden. This horizon is physically important because the absorption by the D-branes can be accounted for as the disappearance of the closed strings behind the horizon . It is therefore natural that the Wilson loop $`C`$ on the D-brane can be represented as a Wilson loop at the horizon . In this case, however, the loop generates the $`B_{\mu \nu }`$-field with $`B_{\mu \nu }=B_{\mu \nu }(x,\{C\})`$ . At present we do not know how to determine this field. We hope that the two types of boundary conditions yield the same result and will be studying the case $`y=0`$. The string representation involves various background fields. For example, since we need NSR fermions to exclude the boundary tachion (violating the zigzag symmetry), we have RR fields in our closed string spectrum, and they form a condensate in order to stabilize the AdS space. The concrete types and forms of background fields depend on the type of the gauge theory we are working with. The problem of finding the precise background in each particular case is not yet completely solved. Of course to check the complete loop equations we have to solve the above problem first. However, in the conformal cases it is possible to retreat to the quasiclassical domain by taking $`g_{YM}^2N1`$. In this case the string action takes the form: $$S[\stackrel{}{x}(\xi ),y(\xi )]=\frac{1}{2}\sqrt{g_{YM}^2N}\frac{d^2\xi }{y^2(\xi )}\left((_a\stackrel{}{x})^2+(_ay)^2\right)+O(1),$$ (3.5) where the terms $`O(1)`$ contain NSR fermions and RR fields. We arrive at the conclusion that in the WKB limit, the Wilson loop is given by , $$W[C]e^{\sqrt{g_{YM}^2N}A_{\mathrm{min}}[C]}.$$ (3.6) Here $$A_{\mathrm{min}}[C]=\mathrm{min}\frac{1}{2}\frac{d^2\xi }{y^2}\left((_a\stackrel{}{x})^2+(_ay)^2\right)$$ (3.7) is the minimal area of a surface bounded by the loop $`C`$. In the following sections we will analyze the action of the loop operator on this functional. ## 4 Minimal area in the Lobachevsky space Our contour is located at infinity (the absolute) of the Lobachevsky space. There are some general features of such minimal areas which we discuss in this section. The equations of motion for the action (3.7) have the form: $$\{\begin{array}{c}_a\left(\frac{1}{y^2}_a\stackrel{}{x}\right)=0,\hfill \\ ^2y=\frac{1}{y}\left((_ay)^2\left(_a\stackrel{}{x}\right)^2\right).\hfill \end{array}$$ (4.1) We also have to impose the Virasoro constraints $$\{\begin{array}{c}\left(_1\stackrel{}{x}\right)^2+\left(_1y\right)^2=\left(_2\stackrel{}{x}\right)^2+\left(_2y\right)^2,\hfill \\ _1\stackrel{}{x}_2\stackrel{}{x}+_1y_2y=0.\hfill \end{array}$$ (4.2) We are looking for the solutions satisfying the conditions $$\{\begin{array}{c}\stackrel{}{x}(\sigma ,0)=\stackrel{}{c}(\sigma ),\hfill \\ y(\sigma ,0)=0.\hfill \end{array}$$ (We renamed the variables: $`\xi ^1=\sigma `$, $`\xi ^2=\tau `$.) It is easy to see that the expansion in $`\tau `$ has the form: $$\{\begin{array}{c}\stackrel{}{x}=\stackrel{}{c}(\sigma )+\frac{1}{2}\stackrel{}{f}(\sigma )\tau ^2+\frac{1}{3}\stackrel{}{g}(\sigma )\tau ^3+\mathrm{}\hfill \\ y=a(\sigma )\tau +\frac{1}{3}b(\sigma )\tau ^3+\mathrm{}\hfill \end{array}$$ (4.3) After substituting this expansion into (4.1) and (4.2) we obtain $$\{\begin{array}{c}a^2(\sigma )=\left(\frac{d\stackrel{}{c}}{d\sigma }\right)^2,\hfill \\ \stackrel{}{f}(\sigma )=\left(\frac{d\stackrel{}{c}}{d\sigma }\right)^2\frac{d}{d\sigma }\left(\frac{_\sigma \stackrel{}{c}}{\left(_\sigma \stackrel{}{c}\right)^2}\right).\hfill \end{array}$$ (4.4) This guarantees that the leading term in the energy-momentum tensor (4.2) vanishes; we have $$\theta _{}=\frac{1}{y^2}\left(_\tau \stackrel{}{x}_\sigma \stackrel{}{x}+_\tau y_\sigma y\right)=\frac{1}{a^2\tau }\left[\stackrel{}{f}\stackrel{}{c}^{}+aa^{}\right]+\frac{1}{a^2}\left(\stackrel{}{g}\stackrel{}{c}^{}\right)+\mathrm{}$$ and due to (4.4) the first term vanishes, while the second one gives $$\theta _{}=\frac{1}{a^2}\left(\stackrel{}{g}\stackrel{}{c}^{}\right).$$ Analogously: $`\theta _{}`$ $`={\displaystyle \frac{1}{2y^2}}\left[\left(_\tau \stackrel{}{x}\right)^2+\left(_\tau y\right)^2\left(_\sigma \stackrel{}{x}\right)^2\left(_\sigma y\right)^2\right]`$ (4.5) $`={\displaystyle \frac{1}{a^2}}\left[\stackrel{}{f}^2+2ab\stackrel{}{c}^{}\stackrel{}{f}^{}a_{}^{}{}_{}{}^{2}\right].`$ The action (3.7) calculated on the classical solution (4.3) has the following structure (encountered previously in in a special case): $$A_{\mathrm{min}}[C]=\frac{L[C]}{ϵ}+𝒜[\stackrel{}{c}(\sigma )],$$ (4.6) where we introduced the cut-off $`y_{\mathrm{min}}=a(\sigma )\tau _{\mathrm{min}}=ϵ`$. In this formula $`L[C]`$ is the length of the contour $`C`$ and $`𝒜[\stackrel{}{c}(\sigma )]`$ is a finite functional. Our main interest is to derive variational equations for $`𝒜[C]`$. One might think that the usual Hamilton-Jacobi equations following from the conditions $`\theta _{}=\theta _{}=0`$ will give us a closed equation for $`𝒜`$. Unfortunately, life is not so simple. The above would be the case is we were able to find the coefficients $`\stackrel{}{g}`$ and $`b`$ in terms of $`\stackrel{}{c}`$. Indeed, it is easy to see that $$\frac{\delta 𝒜}{\delta \stackrel{}{c}(\sigma )}=\frac{\stackrel{}{g}(\sigma )}{a^2(\sigma )}$$ and if $`b`$ were known, the equation for $`𝒜`$ would follow from (4.5). There is an unpleasant surprise, however. Further iterations of (4.1) reveal that the functions $`\stackrel{}{g}(\sigma )`$ and $`b(\sigma )`$ in (4.3) are not determined by the small $`\tau `$-expansion and can be kept arbitrary. They are fixed by the global condition for the absence of singularities at finite $`\tau `$, and thus it is hard to find them explicitly. Because of this difficulty we will choose an alternative way of finding $`𝒜[C]`$. Namely, we will consider a special type of contours—wavy lines, and will develop a method of successive approximations for $`𝒜[C]`$. We will also see another derivation of (4.6). ## 5 The theory of wavy lines It was already suggested in that it is instructive to consider wavy lines instead of general contours, namely to look at a curve $$x^1(s)=s,x^i(s)=\varphi ^i(s),i=2,\mathrm{}D,$$ and to assume that $`\varphi ^i(s)`$ are small. Below we shall find the expansion of $`𝒜[\stackrel{}{\varphi }(s)]`$ up to the fourth order and then explore the action of the loop Laplacian. To perform this expansion it is convenient to consider the standard Hamilton-Jacobi equation for the minimal surface which has the well-known general form: $$G^{MN}(z)\frac{\delta A}{\delta z^M(s)}\frac{\delta A}{\delta z^N(s)}=G_{MN}\frac{dz^M}{ds}\frac{dz^N}{ds}.$$ For the Poincaré metric this equation becomes $$\left(\frac{\delta A}{\delta y(s)}\right)^2+\left(\frac{\delta A}{\delta \stackrel{}{x}(s)}\right)^2=\frac{1}{y^4(s)}\left\{\left(\frac{dy}{ds}\right)^2+\left(\frac{d\stackrel{}{x}}{ds}\right)^2\right\}.$$ (5.1) We want to explore the limit $`y(s)=y0.`$ We can do this by solving (5.1) with respect to $`\frac{\delta A}{\delta y}`$ and by noticing that $$\frac{A}{y}=ds\frac{\delta A}{\delta y(s)}|_{y(s)=y}.$$ (5.2) We have from (5.1) and (5.2) $$\frac{A}{y}=\frac{1}{y^2}𝑑s\sqrt{\left(\frac{d\stackrel{}{x}}{ds}\right)^2y^4\left(\frac{\delta A}{\delta \stackrel{}{x}(s)}\right)^2}.$$ (5.3) We see directly from (5.3) that $`A(y)`$ behaves like $$A(y)\underset{y0}{}\frac{L[C]}{y}+O(1).$$ To get further information we have to look at the following expansion of $`A`$: $$A=\underset{n}{}\frac{1}{n!}ds_1\mathrm{}ds_n\mathrm{\Gamma }_{i_1\mathrm{}i_n}(s_1,\mathrm{}s_n|y)(\varphi _{i_1}(s_1)\mathrm{}\varphi _{i_n}(s_n)).$$ Expansion of (5.3) together with the reparametrization invariance equation $$\frac{dx_\mu }{ds}\frac{\delta A}{\delta x_\mu (s)}=\frac{\delta A}{\delta x_1(s)}|_{x_1=s}+\frac{d\stackrel{}{\varphi }}{ds}\frac{\delta A}{\delta \stackrel{}{\varphi }}=0$$ (5.4) will give us equations for the $`\mathrm{\Gamma }`$’s. Let us expand (5.3): $`{\displaystyle \frac{A}{y}}`$ $`={\displaystyle \frac{L_0}{y^2}}+{\displaystyle \frac{1}{2}}{\displaystyle \left(y^2\stackrel{}{\pi }^2\frac{1}{y^2}\stackrel{}{\varphi }_{}^{}{}_{}{}^{2}\right)𝑑s}`$ (5.5) $`+{\displaystyle \frac{1}{8}}{\displaystyle \left(\frac{1}{y^2}\left(y^4\stackrel{}{\pi }^2\stackrel{}{\varphi }_{}^{}{}_{}{}^{2}\right)^2+4y^2\left(\stackrel{}{\varphi }^{}\stackrel{}{\pi }\right)^2\right)𝑑s}+\mathrm{}`$ where $`\stackrel{}{\pi }=\delta A/\delta \stackrel{}{\varphi }`$. If we perform the Fourier transform in $`s`$, we obtain the following equations for the coefficient functions: $$\{\begin{array}{c}\frac{d\mathrm{\Gamma }_2}{dy}=y^2\mathrm{\Gamma }_2^2\frac{p^2}{y^2},\hfill \\ \frac{d\mathrm{\Gamma }_4}{dy}=y^2\left(\underset{1}{\overset{4}{}}\mathrm{\Gamma }_2(p_i)\right)\mathrm{\Gamma }_4(p_1,\mathrm{}p_4)B_4(p_1,\mathrm{}p_4).\hfill \end{array}$$ (5.6) We assume here that $`A`$ $`={\displaystyle \frac{L_0}{y}}+{\displaystyle \frac{1}{2}}{\displaystyle \mathrm{\Gamma }_2(p)\left(\stackrel{}{\varphi }_p\stackrel{}{\varphi }_p\right)𝑑p}`$ $`{\displaystyle \frac{1}{8}}{\displaystyle \mathrm{\Gamma }_4(p_1,\mathrm{}p_4)\left(\stackrel{}{\varphi }_{p_1}\stackrel{}{\varphi }_{p_2}\right)\left(\stackrel{}{\varphi }_{p_3}\stackrel{}{\varphi }_{p_4}\right)\delta \left(p_i\right)𝑑p_1\mathrm{}𝑑p_4}+\mathrm{}`$ The only slightly complicated structure is the polynomial $`B_4`$ which is found from (5.5): $`(2\pi )B_4(p_1,\mathrm{}p_4)`$ $`=`$ $`C_4(p_1,\mathrm{}p_4)+D(p_1,p_2)+D(p_3,p_4)`$ (5.8) $`D(p_1,p_3)D(p_1,p_4)D(p_2,p_3)D(p_2,p_3).`$ In this formula $`C_4`$ $`=`$ $`\left(p_1p_2p_3p_4\right){\displaystyle \frac{1}{y^2}}+y^6\left(\omega _1\omega _2\omega _3\omega _4\right),`$ $`D(p_1,p_2)`$ $`=`$ $`(p_1p_2\omega _3\omega _4)y^2,`$ where we introduced the notation $`\omega _i=\mathrm{\Gamma }_2(p_i)`$. The equations (5.6) are easy to solve. From the first one we get $$\mathrm{\Gamma }_2(p,y)\omega (p,y)=\frac{p^2}{y(1+|p|y)}$$ (5.9) (this solution is the only one which is positive and regular for $`y\mathrm{}`$.) From the second one, $`\mathrm{\Gamma }_4`$ $`=`$ $`{\displaystyle _y^{\mathrm{}}}𝑑ye^{_0^y𝑑y_1y_1^2\left({\scriptscriptstyle \omega _i}\right)}B_4(p_1,\mathrm{}p_4|y)=`$ (5.10) $`=`$ $`{\displaystyle _y^{\mathrm{}}}𝑑y{\displaystyle \underset{i}{}}(1+|p_i|y)e^{{\scriptscriptstyle |p_i|y}}B_4(p_1,\mathrm{}p_4|y).`$ We see that $`\mathrm{\Gamma }_4`$ also has the structure of (5.8): $$(2\pi )\mathrm{\Gamma }_4=F(p_1,\mathrm{}p_4)+\mathrm{\Phi }_{12}+\mathrm{\Phi }_{34}\mathrm{\Phi }_{13}\mathrm{\Phi }_{14}\mathrm{\Phi }_{23}\mathrm{\Phi }_{24}.$$ (5.11) Taking the integral (5.10) and separating the finite part for $`y0`$ gives $`F`$ $`=`$ $`\left(2{\displaystyle \frac{ϵ_1ϵ_2ϵ_3ϵ_4+1}{\mathrm{\Delta }^3}}+{\displaystyle \frac{ϵ_1ϵ_2ϵ_3ϵ_4}{\mathrm{\Delta }^2}}\left({\displaystyle \frac{1}{|p_i|}}\right)+{\displaystyle \frac{_{i<j}|p_i||p_j|}{\mathrm{\Delta }p_1p_2p_3p_4}}{\displaystyle \frac{\mathrm{\Delta }}{p_1p_2p_3p_4}}\right)p_1^2p_2^2p_3^2p_4^2,`$ $`\mathrm{\Phi }_{12}`$ $`=`$ $`\left({\displaystyle \frac{2ϵ_1ϵ_2}{\mathrm{\Delta }^3}}+{\displaystyle \frac{ϵ_1ϵ_2}{\mathrm{\Delta }^2}}\left({\displaystyle \frac{1}{|p_1|}}+{\displaystyle \frac{1}{|p_2|}}\right)+{\displaystyle \frac{1}{\mathrm{\Delta }p_1p_2}}\right)p_1^2p_2^2p_3^2p_4^2.`$ (5.12) ($`\mathrm{\Delta }=|p_i|`$; $`ϵ_i=\text{sgn }p_i`$). This completes the calculation of $`\mathrm{\Gamma }_4(p_1,\mathrm{}p_4)`$. Let us notice that an alternative way to obtain these formulas is to use the Monge gauge in the expression for the minimal area: $`x_1=\sigma ,y=\tau ,\varphi _i=\varphi _i(\sigma ,\tau ),`$ $$A=\frac{d\tau }{\tau ^2}\sqrt{1+\stackrel{}{\varphi }_\tau ^{\mathrm{\hspace{0.17em}2}}+\stackrel{}{\varphi }_\sigma ^{\mathrm{\hspace{0.17em}2}}+\stackrel{}{\varphi }_\tau ^{\mathrm{\hspace{0.17em}2}}\stackrel{}{\varphi }_\sigma ^{\mathrm{\hspace{0.17em}2}}\left(\stackrel{}{\varphi }_\tau \stackrel{}{\varphi }_\sigma \right)^2}.$$ (5.13) The linear equation for $`\stackrel{}{\varphi }`$, $$_\tau \left(\frac{1}{\tau ^2}_\tau \stackrel{}{\varphi }\right)+\frac{1}{\tau ^2}_\sigma ^2\stackrel{}{\varphi }=0$$ has the solution $$\stackrel{}{\varphi }_{\text{cl}}(p,\tau )=(|p|\tau )^{3/2}K_{3/2}(|p|\tau )\stackrel{}{\varphi }(p)=(1+|p|\tau )e^{|p|\tau }\stackrel{}{\varphi }(p).$$ Expanding (5.13), we arrive once again at the expression (5.10). However, the Hamilton-Jacobi method has some general advantages. ## 6 The second derivative of the minimal area After the divergent part of (4.6) is absorbed into the mass renormalization of the test particle, we are left with the functional $$W[C]=e^{\sqrt{g_{YM}^2N}𝒜[C]}.$$ (6.1) Its second variational derivative has the form $$\frac{\delta ^2W}{\delta x_\mu (s)\delta x_\mu (s^{})}=\left(g_{YM}^2N\frac{\delta 𝒜}{\delta x_\mu (s)}\frac{\delta 𝒜}{\delta x_\mu (s^{})}\sqrt{g_{YM}^2N}\frac{\delta ^2𝒜}{\delta x_\mu (s)\delta x_\mu (s^{})}\right)W.$$ The first term in brackets has no singularity for $`ss^{}`$. This fact was considered in as a check that the loop equation is satisfied. Notice however that in this order an arbitrary functional $`𝒜`$ will pass this check. That is why it is necessary to consider the second term, which we proceed to calculate. The full second variational derivative of $`𝒜`$ consists of the longitudinal and transverse parts: $$\frac{\delta ^2𝒜}{\delta x_\mu (s)\delta x_\mu (s^{})}=\frac{\delta ^2𝒜}{\delta x_1(s)\delta x_1(s^{})}+\frac{\delta ^2𝒜}{\delta \stackrel{}{\varphi }(s)\delta \stackrel{}{\varphi }(s^{})}.$$ (6.2) The transverse part (in momentum representation) can be read off directly from (5). For the term quadratic in $`\varphi `$ we have $$\underset{p\mathrm{}}{lim}\frac{\delta ^2𝒜^{(2)}}{\delta \stackrel{}{\varphi }\left(\frac{q}{2}+p\right)\delta \stackrel{}{\varphi }\left(\frac{q}{2}p\right)}=(D1)\delta (q)\mathrm{\Gamma }_2(p)=(1D)\delta (q)|p|^3,$$ where $`\mathrm{\Gamma }_2(p)=|p|^3`$ is the finite part of $`\mathrm{\Gamma }_2(p,y)`$. Let us now consider the quartic term. The special structure in (5.11) leads to the formula $`(2\pi ){\displaystyle \frac{\delta ^2𝒜^{(4)}}{\delta \stackrel{}{\varphi }(k)\delta \stackrel{}{\varphi }(k^{})}}={\displaystyle \frac{1}{2}}{\displaystyle H(k,k^{},p_1,p_2)\left(\stackrel{}{\varphi }_{p_1}\stackrel{}{\varphi }_{p_2}\right)\delta (k+k^{}+p_1+p_2)𝑑p_1𝑑p_2},`$ $`H=(D+1)F(k,k^{},p_1,p_2)+(D3)\left[\mathrm{\Phi }(k,k^{})+\mathrm{\Phi }(p_1,p_2)\right]`$ $`(D1)[\mathrm{\Phi }(k,p_1)+\mathrm{\Phi }(k,p_2)+\mathrm{\Phi }(k^{},p_1)+\mathrm{\Phi }(k^{},p_2)].`$ (6.3) Let us first analyze the case $`k=k^{}`$, which corresponds to taking $`q=0`$ in (2.6). In this case the third term in the expression for $`H`$ in (6.3) drops out, because $`\mathrm{\Phi }(k,p)`$ is antisymmetric with respect to $`kk`$. The remaining terms have to be expanded for $`k\mathrm{}.`$ Because of homogeneity, it makes sense to express them as functions of $`x=k/p`$, where $`p_1=p_2=p`$, which amounts to taking $`p=1`$ in (5.12). We have: $`F(k,k,p,p)`$ $`=`$ $`|p|^5\left\{{\displaystyle \frac{x^4}{2(1+x)^3}}+{\displaystyle \frac{x^3}{2(1+x)}}+{\displaystyle \frac{x^2(1+4x+x^2)}{2(1+x)}}2x^2(1+x)\right\}`$ $`=`$ $`|p|^5\{{\displaystyle \frac{3}{2}}x^3x+O\left({\displaystyle \frac{1}{x}}\right)\},`$ $`\mathrm{\Phi }(k,k)+\mathrm{\Phi }(p,p)`$ $`=`$ $`|p|^5\left\{{\displaystyle \frac{x^2(1+x^2)}{2(1+x)}}{\displaystyle \frac{x^3}{2(1+x)}}{\displaystyle \frac{x^4}{2(1+x)^3}}\right\}`$ $`=`$ $`|p|^5\{{\displaystyle \frac{1}{2}}x^3x+2+O\left({\displaystyle \frac{1}{x}}\right)\}.`$ We substitute this into (6.3) and obtain the desired asymptotic expansion in $`k\mathrm{}`$: $`{\displaystyle \frac{\delta 𝒜^{(4)}}{\delta \stackrel{}{\varphi }(k)\delta \stackrel{}{\varphi }(k)}}`$ $`=`$ $`{\displaystyle \frac{1}{2\pi }}{\displaystyle }\{Dp^2|k|^3+(D1)p^4|k|`$ (6.4) $`+(3D)|p|^5\}\left(\stackrel{}{\varphi }_p\stackrel{}{\varphi }_p\right)dp+O\left({\displaystyle \frac{1}{k}}\right).`$ In general, when we do not assume that $`q=0`$ in (2.6), we have to take all the terms in (6.3) into account. The corresponding formula for the asymptotic expansion (of which (6.4) is a partial case) can be obtained by means of straightforward but lengthy calculations and has the form: $`{\displaystyle \frac{\delta ^2𝒜^{(4)}}{\delta \stackrel{}{\varphi }\left(\frac{q}{2}+k\right)\delta \stackrel{}{\varphi }\left(\frac{q}{2}k\right)}}={\displaystyle \frac{1}{2\pi }}{\displaystyle }\{Dp_1p_2|k|^3`$ $`+({\displaystyle \frac{D4}{2}}p_1p_2^3+{\displaystyle \frac{3D6}{2}}p_1^2p_2^2)|k|+((4D)p_1^2|p_2|^3+p_1p_2|p_2|^3)\}`$ $`\times \left(\stackrel{}{\varphi }_{p_1}\stackrel{}{\varphi }_{p_2}\right)\delta (p_1+p_2+q)dp_1dp_2+O\left({\displaystyle \frac{1}{k}}\right).`$ (6.5) To complete the calculation of (6.2), we have to compute the longitudinal part (the first term in (6.2)). This part is not immediately visible from (5) and has to be recovered by the use of the reparametrization invariance. The required identity can be easily derived from (5.4) and has the form: $$\frac{\delta ^2𝒜}{\delta x_1(s)\delta x_1(s^{})}=\dot{\varphi }_i(s)\dot{\varphi }_k(s^{})\frac{\delta ^2𝒜}{\delta \varphi _i(s)\delta \varphi _k(s^{})}\delta (ss^{})\dot{\stackrel{}{\varphi }}(s)\frac{d}{ds}\left(\frac{\delta A}{\delta \stackrel{}{\varphi }(s)}\right).$$ In the momentum representation: $`(2\pi ){\displaystyle \frac{\delta ^2𝒜}{\delta x_1\left(k\right)\delta x_1\left(k^{}\right)}}={\displaystyle 𝑑p𝑑p^{}pp^{}\varphi _i(p)\varphi _k(p^{})\frac{\delta ^2𝒜}{\delta \varphi _i(p+k)\delta \varphi _k(p^{}+k^{})}}`$ $`{\displaystyle 𝑑pp(p+k+k^{})\stackrel{}{\varphi }(p)\frac{\delta 𝒜}{\delta \stackrel{}{\varphi }(p+k+k^{})}}.`$ (6.6) In the approximation we are working with, only $`𝒜^{(2)}`$ contributes to the RHS of (6.6). The result is: $`{\displaystyle \frac{\delta ^2𝒜}{\delta x_1\left(\frac{q}{2}+k\right)\delta x_1\left(\frac{q}{2}k\right)}}`$ $`=`$ $`{\displaystyle \frac{1}{2\pi }}{\displaystyle }\{|{\displaystyle \frac{p_1p_2}{2}}+k|^3|p_2|^3\}`$ (6.7) $`\times p_1p_2\left(\stackrel{}{\varphi }_{p_1}\stackrel{}{\varphi }_{p_2}\right)\delta (p_1+p_2+q)dp_1dp_2`$ $`\underset{k\mathrm{}}{=}`$ $`{\displaystyle \frac{1}{2\pi }}{\displaystyle }\{|k|^3+{\displaystyle \frac{3}{4}}(p_1p_2)^2|k|p_1p_2|p_2|^3\}\times \mathrm{}`$ The final expression for the asymptotics of the second variational derivative is thus: $`{\displaystyle \frac{\delta ^2𝒜^{(4)}}{\delta x_\mu \left(\frac{q}{2}+k\right)\delta x_\mu \left(\frac{q}{2}k\right)}}=(1D)\delta (q)|k|^3+{\displaystyle \frac{1}{2\pi }}{\displaystyle }\{(1D)p_1p_2|k|^3`$ $`+({\displaystyle \frac{D1}{2}}p_1p_2^3+{\displaystyle \frac{3D9}{2}}p_1^2p_2^2)|k|+(4D)p_1^2|p_2|^3\}`$ $`\times \left(\stackrel{}{\varphi }_{p_1}\stackrel{}{\varphi }_{p_2}\right)\delta (p_1+p_2+q)dp_1dp_2+O\left({\displaystyle \frac{1}{k}}\right)+O(\varphi ^4).`$ (6.8) In the next section we will analyze this result. ## 7 Interpretation and discussion of the result Our main result is contained in the formula (6.8). It shows, first of all, that the second variational derivative has the expected form (2.4). Namely, dangerous terms $`\delta ^{\prime \prime }(ss^{})`$ (which would manifest themselves in (6.8) as terms $`k^2`$) cancel for all $`D`$. The presence of those terms would imply that our functional is *not* zigzag-invariant, that is not presentable in the form (2.7). The next result following from (6.8) concerns the loop operator $`\widehat{L}_q`$. Using (2.6) and picking up $`k^0`$ terms in (6.8), we get $$\widehat{L}_q𝒜=\frac{4D}{2\pi }p_1^2|p_2|^3\left(\stackrel{}{\varphi }_{p_1}\stackrel{}{\varphi }_{p_2}\right)\delta (p_1+p_2+q)𝑑p_1𝑑p_2.$$ (7.1) This shows that at $`D=4`$ the loop equation is satisfied (at least in our approximation)! We will discuss the significance of this fact in the next section. Now let us perform another test. Consider once again the OPE (2.3) and (2.4) and let us try to determine its contribution to (6.8). If we assume that in conformal theory in the WKB limit the field strength keeps its normal dimension 2 (this will be actually more of the conclusion than of the assumption), we get $`F_{\mu \lambda }(x_1)F_{\mu \sigma }(x_2)`$ $``$ $`C_1{\displaystyle \frac{\delta _{\lambda \sigma }}{|x_1x_2|^4}}+C_2{\displaystyle \frac{(x_1x_2)_\lambda (x_1x_2)_\sigma }{|x_1x_2|^6}}`$ (7.2) $`+C_3{\displaystyle \frac{(x_1x_2)_{(\lambda }(x_1x_2)_\mu F_{\mu \sigma )}(x)}{|x_1x_2|^4}}+C_4{\displaystyle \frac{(x_1x_2)_\mu _\mu F_{\lambda \sigma }(x)}{\left|x_1x_2\right|^2}}.`$ ($`x=\frac{x_1+x_2}{2}`$). The last two terms give no singular contribution to (2.2). The first two give after some calculations: $$\frac{\delta ^2W}{\delta x_\mu (s)\delta x_\mu (s^{})}\underset{ss^{}}{=}\frac{1}{|ss^{}|^4}\frac{C_1+C_2}{\dot{x}^2}+\frac{1}{|ss^{}|^2}\left(\frac{(C_1+C_2)(\dot{x}\stackrel{\mathrm{}}{x})}{12\dot{x}^4}+\frac{C_1\ddot{x}^2}{4\dot{x}^4}+\frac{C_2(\dot{x}\ddot{x})^2}{4\dot{x}^4}\right).$$ The derivatives in the RHS are taken at the point $`\overline{s}=\frac{s+s^{}}{2}.`$ In the wavy line approximation we get $$\frac{\delta ^2W}{\delta x_\mu (s)\delta x_\mu (s^{})}\underset{ss^{}}{=}\frac{1}{|ss^{}|^4}(C_1+C_2)(1\dot{\varphi }^2)+\frac{1}{|ss^{}|^2}\left(\frac{(C_1+C_2)}{12}\dot{\varphi }\stackrel{\mathrm{}}{\varphi }+\frac{C_1}{4}\ddot{\varphi }^2\right)+O(\varphi ^4).$$ (7.3) We have to compare this behavior with our formula (6.8). Picking up the terms $`|k|^3`$ and $`|k|`$ and using the Fourier transform identities $$|k|^3\frac{1}{|ss^{}|^4},|k|\frac{1}{6|ss^{}|^2}$$ to go back to the $`s`$-representation, we have $$\frac{\delta ^2𝒜}{\delta x_\mu (s)\delta x_\mu (s^{})}\frac{1}{|ss^{}|^4}(1D)(1\dot{\varphi }^2)+\frac{1}{|ss^{}|^2}\left(\frac{1D}{12}\dot{\varphi }\stackrel{\mathrm{}}{\varphi }+\frac{3D}{4}\ddot{\varphi }^2\right).$$ (7.4) We see that this formula can be put in complete agreement with (7.3) by taking $`C_1=D3`$, $`C_2=2`$. The second term in (7.4) may be modified if the theory contains a scalar operator of dimension 2. That can change the constants $`C_1`$ and $`C_2`$. At the same time, the structure of the first term cannot be modified by anything and provides a strong check for the consistency of our approach. Notice that it also *predicts* that the dimension of $`F_{\mu \nu }`$ is not renormalized. ## 8 Conclusions and outlook The main efforts of this work were directed towards the development of new techniques for checking the loop equations in string theory. We managed to apply the loop Laplacian to the minimal area in the Lobachevsky space and to show for the first time that the equations of motion of gauge theory are satisfied by string theory, at least in the WKB approximation. This point perhaps requires some clarifications. Namely, we looked at the Wilson loop (2.1) in conformal versions of the Yang-Mills theory. These versions unavoidably contain other fields. Hence we expect that $$_\mu F_{\mu \nu }=J_\nu 0$$ (where $`J_\nu `$ is the current generated by those fields). So, what is the meaning of finding that $`\widehat{L}A_{\mathrm{min}}=0`$ at $`D=4`$? There are several possible interpretations of this fact. Let us begin with the unpleasant one (in which we do not believe). It may be that our result is just a fluke, while if we proceed to higher orders in “waviness” of our contour, the loop equations will not be satisfied. Further progress will be difficult in this case. Let us take another, optimistic view. Various gauge theories presumably have a string representation with the background (1.1). If the theory is conformal, (1.1) must describe the AdS space. What distinguishes various theories is not the metric but other background fields and also the field content on the world sheet. However, in the WKB limit $`g_{YM}^2N\mathrm{}`$ the asymptotics of the Wilson loop is given by the formula (3.6) and is the same for all conformal theories. If this is the case, it means that the current in the above formula is negligible in the WKB limit, and our result *actually checks the universal law* (3.6). Notice also that we are considering the standard Wilson loop and not its modified version suggested in , . Once again, with the above philosophy this modification is irrelevant in the WKB limit. To verify the above assertions it is necessary to go beyond our wavy line approximation. We believe that this can be done by a more general treatment of the Hamilton-Jacobi equations. Conformal invariance of the loop Laplacian (which we discuss in the Appendix) should play an important role in this analysis. Alternatively, one can study the second variation of the functional (3.5) by developing the short distance expansion of the Green functions for equations (4.1). It is conceivable that by this method it will be possible to relate OPE on the world sheet and OPE in gauge theory. This brings us to a more difficult problem of going beyond WKB approximation. Again, the method of wavy lines may be useful here, but at the moment we do not know how to evaluate the action of $`\widehat{L}_q`$ on quantum corrections to our formula. In the case of non-conformal theories the metric has the form $$ds^2=f(\mathrm{log}y)\left(\frac{dy^2+d\stackrel{}{x}^2}{y^2}\right)$$ The problem for our analysis is not so much the function $`f`$ (it is easy to generalize our considerations to this case) as the absence of the WKB domain. It may be helpful to notice that instead of considering non-conformal case in 4d, one can, in the asymptotically free theories, shift to $`D=4+ϵ`$, when these theories become conformal. Perhaps, the classical part of $`\widehat{L}_qWϵ`$ will be canceled by the quantum fluctuations and will provide us with the “Holy Grail” of this subject — the space-time $`\beta `$-function. ## Acknowledgements We are grateful to Volodya Kazakov for his participation in the early stages of this project as well as for useful discussions. The work of A.P. was partially supported by NSF grant PHY-98-02484. ## Appendix: Conformal invariance of the loop equation The basic property of the functional $`𝒜`$ (and hence of $`W`$) is its conformal invariance: $$𝒜[C]=𝒜[f(C)]$$ (8.1) (where $`f`$ is a conformal transformation, say $`f_\mu (x)=x_\mu /x^2`$). This is true for any $`D`$. Although this invariance is to be expected, it is not entirely obvious, since isometries of the Lobachevsky space (corresponding to conformal transformations on the boundary) will change the cut-off $`ϵ`$ in (4.6). To check the invariance, we extend $`f`$ to the $`(D+1)`$-dimensional Lobachevsky isometry: $$(x_\mu ,y)\stackrel{F}{}(\frac{x_\mu }{x^2+y^2},\frac{y}{x^2+y^2}).$$ (8.2) Consider the minimal surface $`M`$ bounded by $`C`$. Since $`F`$ is an isometry, $`F(M)`$ is the corresponding surface for $`f(C)`$. Moreover, $$\text{Area}[M_ϵ]=\text{Area}[F(M_ϵ)]$$ (8.3) (where $`M_ϵ`$ is the surface $`M`$ cut off at the hight $`ϵ`$). For small $`ϵ`$, the LHS is equal to $`L[C]/ϵ+𝒜[C]`$. According to (8.2), the constant cut-off $`ϵ`$ is transformed by $`F`$ to the variable cut-off $`ϵ/x^2`$. To calculate the RHS of (8.3), it is convenient to introduce an auxiliary constant cut-off $`ϵ^{}`$ on $`F(M)`$, so that Area$`[F(M_ϵ)]`$ becomes equal to Area$`[F(M)_ϵ^{}]`$ plus the area of a narrow strip between the variable and constant cut-offs. Now (8.3) implies: $$\frac{L[C]}{ϵ}+𝒜[C]\underset{ϵ,ϵ^{}0}{=}\frac{L[f(C)]}{ϵ^{}}+𝒜[f(C)]+_C\left|\frac{df(x(s))}{ds}\right|𝑑s_{ϵ/x(s)^2}^ϵ^{}\frac{dy}{y^2}.$$ Calculating the integral, we see that the singular terms cancel, and we get (8.1). It is now natural to ask if the loop Laplacian transforms is conformally invariant, i.e. commutes with conformal transformations. This property can be written as the equality $$\widehat{L}(s)U_f[C]=\rho \left(\widehat{L}(s)U\right)[f(C)]$$ (8.4) valid for any (reparametrization invariant) functional $`U[C]`$ (where $`U_f[C]=U[f(C)]`$ is the functional $`U`$ transformed by a conformal transformation; $`\rho `$ is some factor). It turns out that (8.4) is true if and only if $`D=4`$. To prove this, consider the relation: $$\frac{\delta ^2U_f[x(s)]}{\delta x_\mu (s)\delta x_\mu (s^{})}=_\mu f_\lambda (x(s))_\mu f_\sigma (x(s^{}))\frac{\delta ^2U}{\delta f_\lambda (s)\delta f_\sigma (s^{})}+^2f_\lambda (x(s))\frac{\delta U}{\delta f_\lambda (s)}\delta (ss^{}).$$ For conformal transformations we have $$_\mu f_\lambda _\mu f_\sigma =\rho (f)\delta _{\lambda \sigma }.$$ Now, we must collect terms proportional to $`\delta (ss^{})`$. This gives $`\widehat{L}(s)U_f[C]\delta (ss^{})=\rho (f)\left(\widehat{L}(s)U\right)[f(C)]\delta (ss^{})`$ $`+^2f_\lambda {\displaystyle \frac{\delta U}{\delta f_\lambda }}\delta (ss^{})+_\mu f_\lambda (x(s))_\mu f_\sigma (x(s^{}))N_{[\lambda \sigma ]}\delta ^{}(ss^{}),`$ where $`N_{[\lambda \sigma ]}`$ is the coefficient in front of the $`\delta ^{}`$-function in the second variational derivative: $$\frac{\delta ^2U}{\delta f_\lambda (s)\delta f_\sigma (s^{})}=N_{[\lambda \sigma ]}\left(\frac{s+s^{}}{2}\right)\delta ^{}(ss^{})+\mathrm{}$$ From the condition of the reparametrization invariance $`\frac{dx_\lambda }{ds}\frac{\delta U}{\delta x_\lambda (s)}=0`$ we have the identity : $$\frac{\delta U}{\delta x_\lambda (s)}=N_{[\lambda \sigma ]}(s)\dot{x}_\sigma (s).$$ This gives the transformation law $`\widehat{L}(s)U_f[C]`$ $`=`$ $`\rho (f)\left(\widehat{L}(s)U\right)[f(C)]+\mathrm{\Omega }_{[\lambda \sigma ]\mu }(f)N_{[\lambda \sigma ]}\dot{x}_\mu (s),`$ $`\mathrm{\Omega }_{[\lambda \sigma ]\mu }`$ $`=`$ $`\left(^2f_{[\lambda }\right)\left(_\mu f_{\sigma ]}\right)\left(_\alpha _\mu f_{[\lambda }\right)\left(_\alpha f_{\sigma ]}\right).`$ (8.5) Substituting $`f_\mu =x_\mu /x^2`$, we find that $`\mathrm{\Omega }(f)(D4)`$, and thus (8.4) is true if and only if $`D=4`$. As a consequence of the above discussion, for $`D=4`$ the loop equation for the Wilson loop (6.1) is conformally invariant: $$\widehat{L}(s)W[f(C)]=\rho \widehat{L}(s)W[C].$$ (8.6) It follows that in order to check the equation at a point $`x(s)`$ of a contour $`C`$, we are allowed to first apply a conformal transformation, say with the purpose of simplifying the behavior of the contour at the point we are looking at. Although this observation does not play any significant role when working with wavy lines, it might become important for general contours. For $`D4`$ the presence of nonzero additional term in the RHS of the transformation law (8.5) implies that the loop equation for (6.1) *cannot* be satisfied in this case.
warning/0002/hep-ph0002121.html
ar5iv
text
# 1 Constraints on Angles of Unitarity Triangle ## 1 Constraints on Angles of Unitarity Triangle Constraints on unitarity triangle expressed in terms of Wolfenstin parameterization are given below: \] From charmless semileptonic B decays : $`\left|{\displaystyle \frac{V_{ub}}{V_{cb}}}\right|=0.08\pm 0.02`$ (1) which yields $`\left(\rho ^2+\eta ^2\right)^{1/2}=0.36\pm 0.09`$ (2) \] From $`B_d\overline{B}_d`$ mixing. Error is dominated by $`f_B`$ and the bag factor $`B_B`$: $`\left|V_{td}\right|=0.009\pm 0.003`$ (3) which yields $`\left|1\rho i\eta \right|=1.0\pm 0.3`$ (4) \] Value of $`ϵ`$ in $`K`$ system. Error is dominated by hadronic matrix elements: $`\eta \left(1\rho +0.35\right)=0.48\pm 0.20`$ (5) \] Constraints from $`B_s\overline{B}_s`$ $`\left(\mathrm{\Delta }m_s\right)>14.3ps^1\left(90\%CL\right)`$ (6) Using the relation from Box diagrams $`\left|{\displaystyle \frac{V_{td}}{V_{ts}}}\right|=\xi \left[{\displaystyle \frac{m_{B_s}\mathrm{\Delta }m_d}{m_{B_d}\mathrm{\Delta }m_s}}\right]^{1/2}`$ (7) where $`\xi ={\displaystyle \frac{f_{Bs}}{f_{Bd}}}\sqrt{{\displaystyle \frac{B_{Bs}}{B_{Bd}}}}`$ (8) Lattice calculations yield for $`\xi `$ the value $`\xi =1.15\pm 0.05`$ (9) This translates into the bound $`\left|V_{td}/V_{ts}\right|<0.214`$ (10) or $`\left|1\rho i\eta \right|<0.96`$ (11) The last constraint in particular implies $`\gamma <90^{}`$ and $`75^{}<\alpha <120^{}`$. This allowed range of $`\gamma `$ leads to unique estimate of errors in $`\alpha `$ as we shall see. ## 2 CP Violation Through Mixing Strategy to measure $`\beta `$ and $`\alpha `$ involve measuring time dependent asymmetry in $`B`$ decays to CP eigenstate. Defining the long and short lived eigenstates of $`B`$ as $`|B_{L,S}=p|B^{}\pm q|\overline{B}^{},`$ (12) the amplitudes for decays into CP eigenstates are defined as $`A=f_{CP}|H_w|B^{}`$ (13) $`\overline{A}=f_{CP}|H_w|\overline{B}^{}.`$ (14) The asymmetry is then defined by $`Asy(t)=[(1|\lambda |^2)\mathrm{cos}\left(\mathrm{\Delta }Mt\right)2Im(\lambda )\mathrm{sin}\left(\mathrm{\Delta }Mt\right)]/[|1+|\lambda |^2]`$ (15) where $`\lambda =(q/p)\left(\overline{A}/A\right)`$. In the standard model $`(q/p)=e^{2i\beta }`$. If $`A`$ is dependent on a single weak phase, $`\left(\overline{A}/A\right)=e^{2i\varphi _{\mathrm{weak}}}`$ (16) then we have the expression $`Asy(t)=Im(\lambda )\mathrm{sin}\left(\mathrm{\Delta }Mt\right)`$ (17) ### 2.1 Measurement of $`\beta `$ The mode that has the least theoretical uncertainty is $`B\psi Ks`$. The amplitude for this mode can be written in terms of Tree and Penguin contribution as $`A=V_{cb}V_{cs}^{}T+V_{tb}V_{ts}^{}P=V_{cb}V_{cs}^{}(TP)+V_{ub}V_{us}^{}P`$ (18) since $`\left|V_{ub}V_{us}^{}/V_{cb}V_{cs}^{}\right|`$ is $`1/50`$, and the Penguin contribution has predominantly $`\overline{c}c`$ in a color octet state, the contribution due to penguin diagram is less than 1%. If $`BD^+D^{}`$ mode is used instead, the penguin contribution is much larger, and there is no color suppression either. $`A=V_{cb}V_{cd}^{}T+V_{tb}V_{td}^{}P=V_{cb}V_{cd}^{}(TP)V_{ub}V_{ud}^{}P`$ (19) The value of $`\left|V_{ub}V_{ud}^{}/V_{cb}V_{cd}^{}\right|0.3`$, and although $`P`$ is suppressed compared to $`T`$ due to small Wilson coefficients, one can expect a contamination due to penguin of a few percent. ### 2.2 Measurement of $`\alpha `$ The mode $`B^{}\pi ^+\pi ^{}`$ lends itself to the earliest measurement of $`\alpha `$. For this mode the amplitude is $`A=V_{ub}V_{ud}^{}T+V_{tb}V_{td}^{}P=V_{ub}V_{ud}^{}(TP)+V_{cb}V_{cd}^{}P`$ (20) The value of $`\left|V_{cb}V_{cd}^{}/V_{ub}V_{ud}^{}\right|3`$ giving a crude estimate of around 15% for the penguin contamination. Gronau and London have presented a method of extracting $`\alpha `$ from measurements of $`B^{}\pi ^+\pi ^{}`$, $`\overline{B^{}}\pi ^+\pi ^{}`$, $`B^{}\pi ^{}\pi ^{}`$, $`\overline{B^{}}\pi ^{}\pi ^{}`$ and $`B^+\pi ^+\pi ^{}`$. However, the most recent theoretical estimates of $`B^{}\pi ^{}\pi ^{}`$ branching ratio are around $`5\times 10^7`$, making this method academic at present. However, we now discuss theoretical developments that may allow us to extract the correct $`\alpha `$ from measurements of asymmetry in $`B^{}\pi ^+\pi ^{}`$ alone. ## 3 Determination of $`\alpha `$ from $`B^{}\pi ^+\pi ^{}`$ This is based on recent work of Agashe and Deshpande . Recently, the CLEO collaboration has reported the first observation of the decay $`B\pi ^+\pi ^{}`$ .The effective Hamiltonian for $`B`$ decays is: $`_{eff}`$ $`=`$ $`{\displaystyle \frac{G_F}{\sqrt{2}}}[V_{ub}V_{ud}^{}(C_1O_1^u+C_2O_2^u).`$ (21) $`.+V_{cb}V_{cd}^{}(C_1O_1^c+C_2O_2^c)V_{tb}V_{td}^{}{\displaystyle \underset{i=3}{\overset{6}{}}}C_iO_i].`$ The $`C_i`$’s are the Wilson coefficients (WC’s). In a recent paper, Beneke et al. found that the matrix elements for the decays $`B\pi \pi `$, in the large $`m_b`$ limit, can be written as $`\pi \pi |O_i|B`$ $`=`$ $`\pi |j_1|B\pi |j_2|0`$ (22) $`\times [1+{\displaystyle }r_n\alpha _s^n(m_b)+O(\mathrm{\Lambda }_{QCD}/m_b)],`$ where $`j_1`$ and $`j_2`$ are bilinear quark currents. If the radiative corrections in $`\alpha _s`$ and $`O(\mathrm{\Lambda }_{QCD}/m_b)`$ corrections are neglected, then the matrix element on the left-hand side factorizes into a product of a form factor and a meson decay constant so that we recover the “conventional” factorization formula. These authors computed the $`O(\alpha _s)`$ corrections. In this approach, the strong interaction (final-state rescattering) phases are included in the radiative corrections in $`\alpha _s`$ and thus the $`O(\alpha _s)`$ strong interaction phases are determined . The matrix element for $`B\pi ^+\pi ^{}`$ is : $`i\overline{A}\left(\overline{B}_d\pi ^+\pi ^{}\right)`$ $`=`$ $`{\displaystyle \frac{G_F}{\sqrt{2}}}[V_{ub}V_{ud}^{}(a_1+a_4^u+a_6^ur_\chi ).`$ (23) $`+.V_{cb}V_{cd}^{}(a_4^c+a_6^cr_\chi )]\times X.`$ Here $$X=f_\pi \left(m_B^2m_\pi ^2\right)F_0^{B\pi ^{}}\left(m_\pi ^2\right),$$ (24) where $`f_\pi =131`$ MeV is the pion decay constant and $`F_0^{B\pi ^{}}`$ is a form factor. In the above equations, the $`a_i`$’s are (combinations of) WC’s with the $`O(\alpha _s)`$ corrections added. The values of the $`a_i`$’s are given in Table 1 . The imaginary parts of $`a_i`$’s are due to final-state rescattering. For the $`CP`$ conjugate processes, the CKM elements have to be complex-conjugated. We discuss two values of the form factors: $`F^{B\pi ^{}}=0.27`$ and $`0.33`$. Model calculations indicate that the $`SU(3)`$ breaking in the form factors is given by $`F^{BK^{}}1.13F^{B\pi ^{}}`$ . The large measured $`BR(BK\eta ^{})`$ requires $`F^{BK^{}}\stackrel{>}{}0.36`$ which, in turn, implies a larger value of $`F^{B\pi ^{}}`$ $`(0.33)`$. If $`F^{BK^{}}\stackrel{<}{}0.36`$, then we require a “new” mechanism to account for $`BR(BK\eta ^{})`$: high charm content of $`\eta ^{}`$ , QCD anomaly or new physics. Also, if $`F^{B\pi ^{}}<0.27`$, then the value of $`F^{BK}`$ is too small to explain the measured BR’s for $`BK\pi `$ . We use $`|V_{cb}|=0.0395`$, $`|V_{ud}|=0.974`$, $`|V_{cd}|=0.224`$, $`m_B=5.28`$ GeV and $`\tau _B=1.6`$ ps . In Fig. 1 we show the $`CP`$-averaged BR for $`B\pi ^+\pi ^{}`$ as a functions of $`\gamma `$ for $`F^{B\pi ^{}}=0.33`$ and $`0.27`$ and for $`|V_{ub}/V_{cb}|=0.1`$, $`0.08`$ and $`0.06`$. The CLEO measurement is $`B\pi ^+\pi ^{}=\left(4.7_{1.5}^{+1.8}\pm 0.6\right)\times 10^6`$ . If $`F^{B\pi ^{}}=0.33`$ and for $`\gamma \stackrel{<}{}90^{}`$, we see from the figures that smaller values of $`|V_{ub}/V_{cb}|0.06`$ are preferred: $`|V_{ub}/V_{cb}|=0.08`$ is still allowed at the $`2\sigma `$ level for $`\gamma 100^{}`$. The smaller value of $`|V_{ub}/V_{cb}|`$ leads to greater penguin contamination. However, if the smaller value of the form factor ($`0.27`$) is used, then the CLEO measurement is consistent with $`|V_{ub}/V_{cb}|0.08`$. We obtain similar results using “effective” WC’s ($`C^{eff}`$)’s and $`N=3`$ in the earlier factorization framework . Since the $`B_d\overline{B}_d`$ mixing phase is $`2\beta `$, if we neglect the (QCD) penguin operators, i.e., set $`a_{4,6}=0`$ in Eq. (23), we get $$\frac{\overline{A}}{A}=e^{i2\gamma }$$ (25) and $$\text{Im}\lambda =\mathrm{sin}\left(2(\beta +\gamma )\right)=\mathrm{sin}2\alpha .$$ (26) In the presence of the penguin contribution, however, $`\overline{A}/Ae^{i2\gamma }`$ so that $`\text{Im}\lambda \mathrm{sin}2\alpha `$. We define $$\text{Im}\lambda =\text{Im}\left(e^{i2\beta }\frac{\overline{A}}{A}\right)\mathrm{sin}2\alpha _{meas.}$$ (27) as the “measured” value of $`\mathrm{sin}2\alpha `$, i.e., $`\mathrm{sin}2\alpha _{meas.}=\mathrm{sin}2\alpha `$ if the penguin operators can be neglected. In Fig. 2 we plot the error in the measurement of $`\alpha `$, $`\mathrm{\Delta }\alpha \alpha _{meas.}\alpha `$, where $`\alpha _{meas.}`$ is obtained from Eq. (27) and $`\alpha `$ is obtained from $`\gamma `$ and $`|V_{ub}/V_{cb}|`$. Note that $`\mathrm{\Delta }\alpha `$ is independent of $`F^{B\pi ^{}}`$ since the form factor cancels in the ratio $`\overline{A}/A`$. We see that for the values of $`|V_{ub}/V_{cb}|0.06`$ preferred by the $`B\pi ^+\pi ^{}`$ measurement (if $`F^{B\pi ^{}}0.33`$), the error in the determination of $`\alpha `$ is large $`15^{}`$ (for $`\gamma 90^{}`$). If $`F^{B\pi ^{}}0.27`$, then $`|V_{ub}/V_{cb}|0.08`$ is consistent with the $`B\pi ^+\pi ^{}`$ measurement which gives $`\mathrm{\Delta }\alpha 10^{}`$ (for $`\gamma 90^{}`$). The computation of Beneke et al. includes final state rescattering phases, i.e., it is exact up to $`O(\mathrm{\Lambda }_{QCD}/m_b)`$ and $`O(\alpha _s^2)`$ corrections. Thus, the value of $`\mathrm{sin}2\alpha `$ “measured” in $`B\pi ^+\pi ^{}`$ decays (Eq. (27)) is a known function of $`\gamma `$ and $`|V_{ub}/V_{cb}|`$ only (in particular, there is no dependence on the phenomenological parameter $`\xi 1/N`$ and strong phases are included unlike in the earlier factorization framework ). Since, the “true” value of $`\alpha `$ can also be expressed in terms of $`\gamma `$ and $`|V_{ub}/V_{cb}|`$ , we can estimate the “true” value of $`\mathrm{sin}2\alpha `$ from the “measured” value of $`\mathrm{sin}2\alpha `$ for a given value of $`|V_{ub}/V_{cb}|`$ (of course, up to $`O(\mathrm{\Lambda }_{QCD}/m_b)`$ and $`O(\alpha _s^2)`$ corrections); this is shown in Fig. 3 where we have restricted $`\gamma `$ to be in the range $`(40^{},120^{})`$ as indicated by constraints on the unitarity triangle from present data. If $`0^{}\gamma 180^{}`$ is allowed, then there will be a discrete ambiguity in the determination of $`\mathrm{sin}2\alpha `$ from $`\mathrm{sin}2\alpha _{meas.}`$. ## 4 Conclusions We have shown how $`\alpha `$ can be obtained from the measured value of $`\mathrm{sin}2\alpha `$ inspite of large penguin effects. The theoretical work can be extended to $`K\pi `$ modes to obtain values of $`\gamma `$ from the measured branching ratios. Naive factorization suggest $`\gamma 100^{}`$ .
warning/0002/hep-th0002066.html
ar5iv
text
# 1 Plot of the holographic (continuous line) and proper (dashed line) beta functions versus the holographic coupling. NYU-TH/00/01/02 CERN-TH/2000-041 Bicocca-FT/00/04 hep-th/0002066 A NOTE ON THE HOLOGRAPHIC BETA AND $`C`$ FUNCTIONS D. Anselmi <sup>a</sup>, L. Girardello <sup>b,d</sup>, M. Porrati <sup>c</sup> and A. Zaffaroni <sup>d</sup> <sup>a</sup>CERN, Division Théorique, CH-1211, Geneva 23, Switzerland <sup>b</sup> Università di Milano-Bicocca, Dipartimento di Fisica <sup>c</sup> Department of Physics, NYU, 4 Washington Pl, New York NY 10012 <sup>d</sup>INFN - Sezione di Milano, Via Celoria 16, Milan, Italy Abstract The holographic RG flow in AdS/CFT correspondence naturally defines a holographic scheme in which the central charge $`c`$ and the beta function are related by the formula $`\dot{c}=2c\beta _a\beta _bG^{ab}`$, where $`G^{ab}`$ is the metric of the kinetic term of the supergravity scalars. In particular, the metric in the space of couplings is $`f^{ab}=2cG^{ab}`$. We perform some checks of that result and we compare it with the quantum field theory expectations. We discuss alternative definitions of the $`c`$-function. In particular, we compare, for a particular supersymmetric flow, the holographic $`c`$-function with the central charge computed directly from the two-point function of the stress-energy tensor. Conformal field theories in four dimensions have two main central charges, $`c`$ and $`a`$, which multiply the square of the Weyl tensor and the Euler density, respectively, in the trace anomaly. In a quantum field theory interpolating between UV and IR conformal fixed points the total flows of $`c`$ and $`a`$, i.e. the differences $`\mathrm{\Delta }c=c_{\mathrm{UV}}c_{\mathrm{IR}}`$ and $`\mathrm{\Delta }a=a_{\mathrm{UV}}a_{\mathrm{IR}}`$, give important physical information (see for instance ). The flows can be induced by dimensionful parameters, by the renormalization-group scale $`\mu `$, or by the combined effect of both. The RG flow, induced by $`\mu `$, is irreversible, which means that it satisfies the inequality $`\mathrm{\Delta }a0`$. The irreversibility of the RG flow is better studied when dimensionful parameters are absent. This means that the theory is conformal at the classical level. In a non-perturbative formula for the RG flow $`\mathrm{\Delta }a`$ was obtained and checked in perturbation theory. The flows induced by relevant deformations have a quantitatively different effect on $`\mathrm{\Delta }a,`$ although they still obey the inequality $`\mathrm{\Delta }a0`$. Nevertheless, there is a special class of theories in every even dimension, the theories interpolating between $`c=a`$ fixed points, where the formula for $`\mathrm{\Delta }a`$ (equal to $`\mathrm{\Delta }c`$) is universal . This universality also holds in two dimensions. The even-dimensional conformal field theories with $`c=a`$ share various properties with two dimensional conformal field theory . The “holographic” supergravity/gauge theory correspondence considers, in the 5-d gauged supergravity limit, precisely a class of $`c=a`$ conformal field theories , the simplest example being the N=4 supersymmetric Yang-Mills theory in the strongly coupled large-$`N`$ limit. Other examples have been constructed in the literature and need not be supersymmetric. In flows induced by massive deformations were considered in the context of this correspondence. We call them the “holographic” flows. On the basis of the considerations recalled above, we expect, and are indeed going to check in the present paper, that: 1) The “holographic” $`c`$-function defined in obeys a formula similar to the formula for the RG flow of the $`a`$-function in quantum field theory. This is our result (8). In particular, the holographic central charge $`c`$ is stationary at the fixed points. Observe that the stationarity of $`c`$ is not true in a general quantum field theory (with $`ca`$ at the fixed points) and is peculiar of the holographic flows; 2) Other definitions of $`c`$, not related to the equality $`c=a`$, but equally convenient in quantum field theory, for example the central function defined by the stress-tensor two-point function, should exhibit similar properties: monotonicity and stationarity at the fixed points. These facts are also peculiar of the holographic flows, because it is well known that $`c`$ does not even decrease in a general quantum field theory. We begin by discussing the properties of the $`c`$-function proposed in ref. and work out the general formula for its derivative $`\dot{c}`$ along the flow. Secondly, we directly compute the $`c`$-function using the correlator of two stress-energy tensors, always using the holographic correspondence, and compare the two definitions. We explain why the two definitions are compatible (in particular, both positive and interpolating monotonically between the critical values) even though they are not equal. A candidate $`c`$-function, decreasing along the holographic flow was proposed in . In the notation of the $`c`$-function is: $$c=\mathrm{const}.(T_{yy})^{3/2}=\left(\frac{\mathrm{d}\varphi }{\mathrm{d}y}\right)^3,$$ (1) Where $`\varphi `$ is the scale factor of the 5-d supergravity metric, and $`y`$ is its radial coordinate: $`ds^2=dy^2+\mathrm{exp}(2\varphi )dx_\mu dx^\mu `$. The equations for a holographic RG flow generated by one of the perturbations that can be studied within 5-d gauged supergravity are, in the notations of ref. <sup>1</sup><sup>1</sup>1 These perturbations have UV dimension 2 or 3; in gauge theory, they correspond to mass terms for scalars and/or fermions, and trilinear terms in the scalar potential. In supergravity, they correspond to VEVs of some of the 42 scalars in the 5-d, N=8 supergravity multiplet.: $$\frac{\mathrm{D}}{\mathrm{D}y}\left(e^{4\varphi }G_{ab}\frac{\mathrm{d}\lambda ^b}{\mathrm{d}y}\right)=e^{4\varphi }\frac{V}{\lambda ^a},6\left(\frac{\mathrm{d}\varphi }{\mathrm{d}y}\right)^2=\underset{ab}{}G_{ab}\frac{\mathrm{d}\lambda ^a}{\mathrm{d}y}\frac{\mathrm{d}\lambda ^b}{\mathrm{d}y}2V.$$ (2) Here $`\lambda ^a`$ denotes the 42 scalars of 5-d N=8 gauged supergravity, $`G_{ab}`$ denotes the metric of their kinetic term and D/D$`y`$ is the covariant derivative. From now on we will set for simplicity and with no loss of generality $`G_{ab}=\delta _{ab}`$. These equations imply, in particular, that the second derivative of $`\varphi `$ does not depend on the potential $`V`$: $$\frac{\mathrm{d}^2\varphi }{\mathrm{d}y^2}=\frac{2}{3}\underset{a}{}\left(\frac{\mathrm{d}\lambda _a}{\mathrm{d}y}\right)^2.$$ We also have $$\frac{\mathrm{d}c}{\mathrm{d}\varphi }=3\left(\frac{\mathrm{d}\varphi }{\mathrm{d}y}\right)^5\frac{\mathrm{d}^2\varphi }{\mathrm{d}y^2}=2c\underset{a}{}\left(\frac{\mathrm{d}\lambda _a}{\mathrm{d}\varphi }\right)^2.$$ To obtain quantitative agreement with QFT results (see ) and a consistent picture of the holographic RG flow we must set $$\varphi =\mathrm{ln}\mu ,\beta _a=\frac{\mathrm{d}\lambda _a}{\mathrm{d}\varphi }.$$ (3) Therefore: $$\dot{c}=\frac{\mathrm{d}c}{\mathrm{d}\varphi }=2c\underset{a}{}\beta _a^2.$$ (4) Let us recall a few other results from quantum field theory . Defining $$\mathrm{\Theta }=\beta _a𝒪_a$$ (5) and $$\mu \frac{\mathrm{d}}{\mathrm{d}\mu }\beta _a=\dot{\beta }_a=\mathrm{\Delta }_{ab}\beta _b,$$ a theorem proved in states that the critical value $`h_{}`$ of the $`\mathrm{\Theta }`$-anomalous dimension, $$\mathrm{\Theta }(x)\mathrm{\Theta }(0)=\frac{\mathrm{const}.}{|x|^{8+2h_{}}},$$ equals the minimal real part of the $`\mathrm{\Delta }`$-eigenvalues, in the IR limit, and the maximal real part of the $`\mathrm{\Delta }`$-eigenvalues in the UV limit. Note that $`h_{}`$ is also the anomalous dimension of the off-critical deformation of the theory, i.e. the operator $`\lambda _a𝒪_a`$ (the deformation being $`_{}=_{}+\lambda _a𝒪_a`$, where $`_{}`$ denotes the critical Lagrangian). Formula (4) implies, in particular, $$\frac{\ddot{c}}{2\dot{c}}=\frac{_{a,b}\beta _a\mathrm{\Delta }_{ab}\beta _b}{_a\beta _a^2}+\underset{a}{}\beta _a^2.$$ (6) At criticality the second term vanishes, while the first term selects the minimal- or maximal-real-part eigenvalue of the matrix $`\mathrm{\Delta }_{ab}`$, as we now show. Note that $`\mathrm{\Delta }_{ab}`$ is in general not symmetric. We can diagonalize it in a complex space. Let $`\mathrm{\Delta }=P^1DP`$, with $`D=\mathrm{diag}(\delta _a)`$, $`\delta _a`$ denoting the eigenvalues. Let us write, around the critical point, $$\beta _a(\lambda )=\mathrm{\Delta }_{ab}\lambda _b,\beta _a(\mu )=\mathrm{\Delta }_{ab}\mu ^{\mathrm{\Delta }_{bc}}k_c=(P^1D\mu ^DPk)_a,$$ $`k_c`$ denoting arbitrary constants. Now, in the UV limit ($`\mu \mathrm{}`$) the behavior of the first term of (6) is dominated by the eigenvalue of the matrix $`\mathrm{\Delta }_{ab}`$ with maximal real part. It is dominated by the eigenvalue with minimal real part in the IR limit ($`\mu 0`$). The imaginary parts of the eigenvalues are irrelevant phases. In conclusion, we have $$\frac{\ddot{c}}{2\dot{c}}=\mathrm{max}\text{Re }\delta _a\mathrm{in}\mathrm{the}\mathrm{UV},\frac{\ddot{c}}{2\dot{c}}=\mathrm{min}\text{Re }\delta _a\mathrm{in}\mathrm{the}\mathrm{IR}.$$ These are also the values of the anomalous dimensions of the operators $`𝒪_a`$ at criticality, as proved in . Therefore we have, in complete generality, $$h_{}=\underset{}{lim}\frac{\ddot{c}}{2\dot{c}},$$ (7) where the star denotes criticality. The “anomalous dimension” $`h_{}`$ denotes the deviation of the total dimension from the reference value 4, $`h_{}=\mathrm{\Delta }4`$ in the conventional notation. We can check, in complete generality, that this quantum field theoretical prediction is correctly reproduced by the holographic flows. Indeed, the second equation of (2) implies that, around a fixed point, $$\frac{\mathrm{d}\varphi }{\mathrm{d}y}=\frac{1}{R},$$ $`R`$ being the AdS radius, and the first of eqs. (2) gives $$\lambda \mathrm{const}.\mathrm{e}^{(4\mathrm{\Delta })y/R}=\mathrm{const}.\mathrm{e}^{(4\mathrm{\Delta })\varphi }$$ At this point, it is straightforward to see that (7) gives $`\mathrm{\Delta }4`$. The same can be see from the definition of $`\beta `$ in (3), confirming that the natural definition of holographic beta function works correctly. All the results described above generalize to non-canonical scalar metrics, in particular $$\dot{c}=2cG^{ab}\beta _a\beta _b.$$ (8) We must remark that our definition of $`c`$ is unique only at the critical points $`\dot{c}=0`$. Away from criticality, $`c`$ need not coincide with central functions defined in other ways; indeed, it need not coincide with other holographic definitions of $`c`$, as for instance that given in ref. . This non-uniqueness even within the holographic scheme follows from the ambiguity in the identification of $`\varphi `$ as a function of the scale $`\mu `$. Only at the critical points, $`\dot{c}=0`$, is the standard identification, $`\varphi =\mathrm{log}(\mu /\mu _0)`$, unique, because of the AdS/CFT correspondence. Away from criticality, uniqueness is lost. A canonical definition of $`c`$ as a function of the scale is obtained by computing the two point function of the stress-energy tensor using the equation $$T_{\mu \nu }(x)T_{\rho \sigma }(0)=\frac{1}{48\pi ^4}\underset{\mu \nu \rho \sigma }{\overset{(2)}{}}\left[\frac{c(x)}{x^4}\right]+\pi _{\mu \nu }\pi _{\rho \sigma }\left[\frac{f(x)}{x^4}\right],$$ (9) where $`\pi _{\mu \nu }=_\mu _\nu \eta _{\mu \nu }^2`$, and $`_{\mu \nu \rho \sigma }^{(2)}=2\pi _{\mu \nu }\pi _{\rho \sigma }3(\pi _{\mu \rho }\pi _{\nu \sigma }+\pi _{\mu \sigma }\pi _{\nu \rho })`$. We will call this $`c`$ the canonical $`c`$-function. For a generic flow, it is impossible to compute analytically this two-point function, even in the supergravity approximation. To the best of our knowledge, there are few exceptions, namely, the solutions describing the Coulomb branch of the N=4 supersymmetric gauge theory and the N=1 supersymmetric flows studied in ref. , which interpolate between the N=4 UV theory and an IR N=1 pure super Yang-Mills theory. Here, we mostly consider the flow to pure N=1 YM theory. Only a few modifications of the computation described below are required to study the N=4 Coulomb branch, which will be briefly discussed at the end of this paper. The flow we shall consider corresponds to an IR vacuum with zero gaugino condensate. By rescaling the AdS radius to $`R=1`$ and setting the IR singularity of the metric at $`y=0`$ (see ref. for details), the 5-d metric is completely specified by the scale factor $$e^{2\varphi (y)}=e^{2y}1.$$ (10) To compute the two-point function of the transverse-traceless part of the stress-energy tensor using the holographic correspondence, we need to solve the linearized equations of motion for the 5-d graviton on the background specified by eq. (10). These equations simplify dramatically for the transverse-traceless part, when they become identical with the equations of motion of a minimally-coupled massless scalar, denoted here by $`\chi (x)`$. By writing $`\chi (x)=\mathrm{exp}(ik_\mu x^\mu )\chi _k`$ we find $$\frac{\mathrm{d}^2}{dy^2}\chi _k4\frac{d\varphi }{dy}\frac{\mathrm{d}}{dy}\chi _k+k^2e^{2\varphi }\chi _k=0.$$ (11) With the change of variable $`x=\mathrm{exp}(2y)`$, a few elementary algebraic manipulations, and dropping the label $`k`$, eq. (11) reduces to a standard hypergeometric equation $$x(1x)\frac{\mathrm{d}^2}{dy^2}\chi (1+x)\frac{\mathrm{d}}{dy}\chi +a^2\chi =0,a^2\frac{k^2}{4},$$ (12) whose two solutions are (cfr. for notations) $`\chi _1`$ $`=`$ $`x^2F(a+2,a+2;3;x),`$ (13) $`\chi _2`$ $`=`$ $`x^2\mathrm{log}(x)F(a+2,a+2;3;x)+{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}[\psi (a+n+2)\psi (a+2)+\psi (a+n+2)+`$ (14) $`\psi (a+2)\psi (n+3)+\psi (3)\psi (n+1)+\psi (1)]{\displaystyle \frac{(a+2)_n(a+2)_n}{3_nn!}}x^{n+2}+`$ $`+{\displaystyle \frac{4}{a^2(a^21)}}x.`$ The linear combination of $`\chi _1,\chi _2`$ regular at $`x=1`$ and normalized to $`1`$ at $`x=0`$ is $$\chi =\frac{a^2(a^21)}{4}\{\chi _2[2\psi (a+2)\psi (a+2)+\psi (3)+\psi (1)]\chi _1\}.$$ (15) The two-point function of the stress-energy tensor is extracted from this expression in the usual manner . Namely, we compare eq. (15) with eq. (9), and we normalize the central charge in the UV using formula (32) of ref. . To simplify, we choose as in a Euclidean 4-momentum ($`k^20`$) oriented along the $`z`$ coordinate, and we find $$\stackrel{~}{T}_{xy}(k)T_{xy}(0)=\frac{N^2}{64\pi ^2}k^2(k^2+4)\text{Re }\psi (2+ik)+P(k^2).$$ (16) Here $`P(k^2)`$ denotes a polynomial in $`k^2`$ which only contributes to contact terms. In the UV this formula approaches, obviously, the pure-AdS form $$\stackrel{~}{T}_{xy}(k)T_{xy}(0)=\frac{N^2}{64\pi ^2}k^4\mathrm{log}(k^2)+\stackrel{~}{P}(k^2).$$ (17) The 2-point function of the transverse-traceless part of $`T_{\mu \nu }`$ is proportional to eq. (16), as we noticed above. From equation (9), we can read the $`c`$-function $$d^4xe^{ikx}\frac{c(x)}{|x|^4}=\frac{\pi ^2N^2}{4}\frac{k^2+4}{k^2}\text{Re }\psi (2+\frac{ik}{2})=\frac{\pi ^2N^2}{4}\underset{n=2}{\overset{\mathrm{}}{}}\frac{k^2+4}{n(4n^2+k^2)}$$ (18) where we used the series expansion for $`\psi `$. In the right hand side of this equation, we discarded any contact or trace term. The Fourier transform can be inverted (modulo contact terms) to give a closed expression for $`c(x)`$ $$c(x)=\frac{N^2}{2}\underset{n=2}{\overset{\mathrm{}}{}}(n^21)|x|^3K_1(2n|x|)=\frac{N^2}{2}|x|^3_0^{\mathrm{}}\frac{3e^{2x\mathrm{cosh}t}1}{(e^{2x\mathrm{cosh}t}1)^3}\mathrm{cosh}tdt$$ (19) Every candidate $`c`$-function has to satisfy some crucial requirements. First of all, it must be positive definite; this is manifest from equation (19). Second, it must coincide with the value of the central charge at the fixed points of the RG group. This is indeed the case. For small $`x`$, $`c(x)c_{UV}=N^2/8`$, while for large $`x`$, $`c(x)x^{5/2}e^{4x}c_{IR}=0`$, as appropriate for a confining theory. Finally, it must be monotonic. This can be checked by an explicit computation: $$\dot{c}=x\frac{dc}{dx}=\frac{3N^2x^3}{2}_0^{\mathrm{}}\frac{(3e^{2x\mathrm{cosh}t}1)(e^{2x\mathrm{cosh}t}1)4x\mathrm{cosh}te^{4x\mathrm{cosh}t}}{(e^{2x\mathrm{cosh}t}1)^4}\mathrm{cosh}tdt$$ (20) One can easily check that the integrand is negative definite. This is a non-trivial result that confirms the interpretation of supergravity solutions as description of quantum field theory RG flows. We can compute the first terms in the small $`x`$-expansion of $`c(x)`$ $$c(x)=N^2\left[\frac{1}{8}+\frac{x^2}{4}\mathrm{log}x+O(x^2)\right]$$ (21) Inserting this expansion into formula (7) we find correctly $`h_{\mathrm{UV}}=1`$, since the deformation is generated by a fermionic mass term ($`\mathrm{\Delta }=3`$). The holographic $`c`$-function for the flow to pure N=1 YM is easily computed from equation (10). With the naive identification $`\varphi =\mathrm{log}(\mu /\mu _0)`$, $`\mu _0=\text{constant}`$, one finds $$c_\mathrm{H}(\mu )=\frac{N^2}{8}\frac{\mu ^6}{(\mu ^2+\mu _0^2)^3}$$ (22) The $`c`$-function given by eq. (9), instead, depends non-analytically on $`\mu 1/x`$ already at small $`x`$, as shown by eqs. (21) and (22). This result does not mean that the two definitions are incompatible, rather, as pointed out before, it means that the identification of $`\mathrm{exp}(\varphi )`$ with the scale $`\mu `$ does not hold outside the critical points. The computation of the function $`f(x)`$ in eq. (9) would be interesting because, as noticed in ref. , it is related to the derivative of $`c(x)`$. Unfortunately, the computation of $`f(x)`$ can not be reduced to the one for a minimally coupled scalar field and requires the full stress-energy tensor two point function. We conclude with a few observations about the holographic scheme. i) An ansatz such as $`\dot{c}\beta ^k`$ with $`k2`$ would disagree with the AdS/CFT correspondence (it would not verify the check above). It would disagree also with quantum field theory . In this sense we have a consistent check of holography versus quantum field theory. ii) The metric $`f`$ in the space of couplings – i.e. the higher dimensional analogue of the Zamolodchikov metric – is the metric of the 5-d supergravity scalars times $`c`$ itself: $$f_{ab}=2cG_{ab}.$$ (23) Therefore, the monotonicity of $`c`$ is directly implied by the positivity of $`c`$ and $`G_{ab}`$, and vice versa. This is not completely surprising, because we can expect the metric $`f`$ to be related to the normalization of the two-point functions of scalar operators. From 5d supergravity, $$S=\sqrt{g}\left(\frac{R}{4}+G_{ab}\lambda ^a\lambda ^b\right)$$ (24) we can see by a simple scaling that, at least at the fixed points, where $`ds^2=R^2[dy^2+\mathrm{exp}(2y)_idx_i^2]`$ $`T(x)T(0)={\displaystyle \frac{c}{|x|^8}}`$ $``$ $`cR^3(\mathrm{\Lambda })^{3/2}`$ $`\lambda _a(x)\lambda _b(0)={\displaystyle \frac{f_{ab}}{|x|^{2\mathrm{\Delta }}}}`$ $``$ $`f_{ab}R^3G_{ab}cG_{ab}`$ (25) The first equation reproduces the known result for $`c`$ , the second one confirms equation (23). iii) Using the arguments of it is straightforward to show that (23) defines a consistent scheme choice, at least when $`c_{\mathrm{IR}}0`$. We call this scheme the “holographic scheme” and can be considered in the class of “proper” schemes of , in which the metric $`f`$ is set equal to a known, positive function: the identity in , $`2cG^{ab}`$ here. The choice $`f_{ab}=\delta _{ab}`$ defines the proper beta function $`\beta _\mathrm{P}`$ and relates the total $`c`$-flow to the area of the graph of the beta function. In the holographic scheme, instead, we have (for $`G^{ab}=\delta ^{ab}`$), $`f^{ab}=2c\delta ^{ab}`$, i.e. the total flow of $`\mathrm{ln}c`$ is (twice) the area of the graph of the holographic beta function. In Fig. 1 the two beta functions are compared for the model of , involving the flow to the (confining) pure N=1 super-Yang Mills theory. The holographic beta function tends to a costant in the IR, while the proper beta function better resembles an ordinary beta function. When $`c_{\mathrm{IR}}=0`$, as in our last example, it is natural to expect that the holographic scheme (23) is still consistent, because, although the holographic beta function tends to a constant, the metric $`f`$ is zero in the “null” IR theory. This is what our explicit computation of the $`T_{\mu \nu }`$ correlator shows. iv) In the presence of many couplings the formula $`\dot{c}=2c\beta _a\beta _bG^{ab}`$ does not give all the beta functions separately. Yet, the sum $`\beta _a\beta _bG^{ab}`$ is sufficient both to fix $`h_{}`$ and to identify the fixed points. In this sense we may call $$\beta _\mathrm{H}\sqrt{\beta _a\beta _bG^{ab}}=\sqrt{\frac{\dot{c}}{2c}}$$ (26) the holographic beta function, so that $`\dot{c}=2c\beta _\mathrm{H}^2`$. The proper beta function is instead $`\beta _\mathrm{P}=\beta _\mathrm{H}\sqrt{2c}`$, so that $`\dot{c}=\beta _\mathrm{P}^2`$. v) With obvious changes, various formulas above apply for the $`a`$-function of in the general case $`ca`$. Indeed, the relationship between the critical exponent $`h_{}`$ and the $`a`$-function does not require inputs from the AdS/CFT correspondence and holds purely in quantum field theory. This generalization is straightforward and left to the reader. vi) In refs. , explicit formulas for the two-point function of minimally-coupled, massless scalars in the Coulomb branch of N=4 supersymmetric gauge theory are given. From those formulas, one can extract a $`c`$-function using the same techniques described in this paper. As an example we now briefly discuss the case of a a 4-dimensional distribution of branes, giving rise to the two-point function described in eq. (25) of ref. . Following the same steps that led us to eq. (19) we find the central function: $$c(x)=\frac{N^2}{4}\underset{n=2}{\overset{\mathrm{}}{}}(2n1)\sqrt{n^2n}|x|^3K_1(2\sqrt{n^2n}|x|).$$ (27) This function is also positive and monotonic as shown in Figure 2. The holographic scheme is natural and simple. Other schemes and definitions for $`c`$-functions are less natural from the point of view of the AdS/CFT correspondence, but still have great interest in their own and give results for $`c=a`$ theories that share many properties with 2d conformal field theories. In particular, we considered the definition of a $`c`$-function from the two-point function of the stress-energy tensor. We computed such a $`c(x)`$ for a particular supersymmetric flow. The fact that it is monotonic is a highly non-trivial check of the AdS/CFT correspondence as well as of the fact that supergravity solutions may be interpreted as quantum field theory RG flows. Notice that the particular solution used in the computation is singular in the IR (as it happens for all the cases where analytical computations of two-point functions can be performed). Nevertheless, we obtained a sensible result, which indicates that the basic physical properties of such solutions are not completely spoiled by the IR singularity. We conclude by mentioning some possible extensions of this work that we find particularly interesting. 1) To compute the two-point function for a holographic flow between CFTs, as the one connecting the N=4 theory to an IR N=1 CFT, discussed in ref. . 2) To prove in full generality that the canonical $`c`$-function is always monotonic, as it happens for the holographic $`c`$-function. 3) To generalize formula (8) to the canonical $`c`$-function. AcknowledgmentsWe would like to thank D. Z. Freedman, M. Petrini and A. Starinets for useful discussions. L.G. and A. Z. are partially supported by INFN and MURST, and by the European Commission TMR program ERBFMRX-CT96-0045, wherein they are associated to the University of Torino. M.P. is supported in part by NSF grant no. PHY-9722083.
warning/0002/hep-ph0002166.html
ar5iv
text
# Strong coupling constant from 𝜏 decay within a renormalization scheme invariant treatment ## Abstract We extract a numerical value for the strong coupling constant $`\alpha _s`$ from the $`\tau `$-lepton decay rate into nonstrange particles. A new feature of our procedure is the explicit use of renormalization scheme invariance in analytical form in order to perform the actual analysis in a particular renormalization scheme. For the reference coupling constant in the $`\overline{\mathrm{MS}}`$-scheme we obtain $`\alpha _s(M_\tau )=0.3184\pm 0.0060_{exp}`$ which corresponds to $`\alpha _s(M_Z)=0.1184\pm 0.0007_{exp}\pm 0.0006_{hqmass}`$. This new numerical value is smaller than the standard value from $`\tau `$-data quoted in the literature and is closer to $`\alpha _s(M_Z)`$-values obtained from high energy experiments. preprint: MZ-TH/00-03 The physics of $`\tau `$-lepton hadronic decays is an important area of particle phenomenology where the theory of strong interaction (QCD) can be confronted with experiment to a very high precision. The central quantity of interest in this process is the spectral density of hadronic states related to the two-point correlator of hadronic currents with well established and simple analytic properties. The accuracy of experimental data for a variety of observables of the $`\tau `$-lepton system is rather good and is steadily improving . The spectral density itself (more precisely, the two-point correlator of hadronic currents in the Euclidean domain) has been calculated with a very high degree of accuracy within perturbation theory (see e.g. ). Nonperturbative corrections to the correlator are known to be small and under control within the operator product expansion and factorization approximation . The observables in the $`\tau `$ system are inclusive in nature which makes the comparison of experimental data with theoretical calculations very clean . Of some particular interest is the precise determination of the numerical value of the strong coupling constant at the low energy scale of the $`\tau `$-lepton mass. Within the renormalization group approach this number can then be evolved to high energies. This is a powerful consistency check of QCD since one is comparing hadron physics at a tremendous variety of scales, from one to hundreds of $`\mathrm{GeV}`$ (e.g. ). In the present note we provide a thorough analysis of the procedure of extracting numerical values of $`\alpha _s`$ from $`\tau `$-data in perturbation theory. On the theory side one expects a high degree of accuracy in the determination of $`\alpha _s`$ because of the existence of very accurate perturbation theory formulas and the simplicity of the renormalization group treatment of the massless quark case. However, the numerical value of the expansion parameter $`\alpha _s`$ is not small at the $`M_\tau `$ scale and the contribution of higher order terms in the perturbation theory series can be significant. Arguments have been brought forth that the accuracy of finite-order perturbation theory is already close to its asymptotic limit which makes the interpretation (usually called resummation) of the perturbation theory series in higher orders necessary . The resummation of contributions related to the running of the coupling constant is most advanced e.g. . The decisive new point of our analysis is the explicit use of renormalization group invariance in the analysis of the $`\tau `$-lepton decay rate within perturbation theory. Renormalization group invariance is a fundamental property of perturbation theory in quantum field theory which is related to the freedom in defining the subtraction procedure . It should be respected in any numerical analysis. Renormalization group invariance allows one to formally perform the numerical analysis in any renormalization scheme because all schemes are connected by a renormalization group transformation. However, in the finite-order perturbation theory approach this equivalence is only approximate due to the systematic omission of higher order terms in the perturbation theory expressions. This inroduces numerical differences into the results obtained in different renormalization schemes. Generally one can consider two ways of using perturbation theory calculations. One is to find relations between physical observables which are renormalization group invariant. Then perturbation theory calculations are just a purely intermediate step for finding relations between observables (see, e.g. ) and no numerical analysis for renormalization scheme noninvariant quantities is performed. Indeed, let the perturbation theory expressions for two observables $`𝒪_{1,2}`$ in a given scheme have the form $`𝒪_1`$ $`=`$ $`\alpha _s+r_1\alpha _s^2+O(\alpha _s^3),`$ (1) $`𝒪_2`$ $`=`$ $`\alpha _s+r_2\alpha _s^2+O(\alpha _s^3).`$ (2) Then the perturbation theory relation between observables $`𝒪_{1,2}`$ reads $$𝒪_2=𝒪_1+(r_2r_1)𝒪_1^2+O(𝒪_1^3)$$ (3) and is scheme-independent. The difference $`r_2r_1`$ takes the same value for calculations in any scheme. Another way of using perturbation theory calculations is to extract numerical values for renormalization scheme noninvariant quantities (as the coupling constant in a fixed scheme). These are then compared with the results of other experiments. In this case the truncation of the perturbation theory series leads to numerical violations of renormalization scheme invariance and plays an essential role. In our simple example this means that the relations in eq. (1) are treated as quadratic functions of $`\alpha _s`$ in some fixed scheme and the accuracy of extraction of the coupling constant value (and prediction of other observables) depends drastically on the scheme used, i.e. on the numerical values of the coefficients $`r_{1,2}`$. In the present paper we consider just this second application and extract a numerical value for the coupling constant which is not an immediate physical quantity. By convention the reference value of the coupling constant that is used to compare between different experiments is fixed to be the $`\overline{\mathrm{MS}}`$-scheme one. However, and this is our point in this paper, this does not necessarily mean that for its extraction from a given experiment the numerical analysis should be performed in the $`\overline{\mathrm{MS}}`$-scheme. It can be more convenient (and numerically accurate) to analyze the system in its internal scheme and after finding numerical values for the internal parameters translate them into the $`\overline{\mathrm{MS}}`$-scheme using renormalization scheme transformation. This program heavily uses explicit renormalization scheme covariance of the theory. However, expressions for the amplitudes are available only in perturbation theory as a truncated series in the coupling constant. For a truncated series the renormalization scheme invariance is only approximate with a precision of the order of the value of the first omitted term. Therefore numerical values obtained in the $`\overline{\mathrm{MS}}`$-scheme directly and through renormalization group transformations can differ. We discuss this problem and argue that the internal scheme results are most reliable physically and are more stable numerically than the results of the standard analysis in the $`\overline{\mathrm{MS}}`$-scheme. Then numerical values for the reference $`\overline{\mathrm{MS}}`$-scheme parameters can be obtained by a renormalization group “rotation” from the numerical values found in the internal schemes. Renormalization group “rotation” (the re-calculation of numerical values from one scheme to another) is a quite formal operation and can be easily controlled numerically. One example of such a “rotation” (the renormalization group scaling which is a one-parameter subgroup of the general renormalization group) is the evolution of the coupling constant to the reference scale $`M_Z`$. Below we give a detailed description of our approach. The normalized $`\tau `$-lepton decay rate into nonstrange hadrons $`h_{S=0}`$ is given by $`R_{\tau S=0}`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Gamma }(\tau h_{S=0}\nu )}{\mathrm{\Gamma }(\tau l\overline{\nu }\nu )}}`$ (4) $`=`$ $`N_c|V_{ud}|^2S_{EW}(1+\delta _P+\delta _{EW}+\delta _{NP})`$ (5) where $`N_c=3`$ is the number of colors. The first term in eq. (4) is the parton model result while the second term $`\delta _P`$ represents perturbative QCD effects. For the flavor mixing matrix element we use $`|V_{ud}|^2=0.9511\pm 0.0014`$ . The factor $`S_{EW}=1.0194`$ is an electroweak correction term and $`\delta _{EW}=0.001`$ is an additive electroweak correction . The nonperturbative corrections are rather small and consistent with zero; we use $`\delta _{NP}=0.003\pm 0.003`$ (see e.g. ). Note that recently the problem of duality violation for two-point correlators has been discussed . However, no established quantitative estimates of that violation are available yet. Considerations show that they can be rather large and can reach the level of few percents. This problem can affect the numerical value of the coupling extracted from the analysis because of the numerical change of the quantity $`\delta _P`$ extracted from eq. (4). In the present note we concentrate on the perturbative part of the decay rate and numerical uncertainties related to the renormalization scheme freedom of perturbation theory. In this respect new possible corrections do not qualitatively affect our analysis. The corrections due to duality violation are of a new nature and they can be added independently to eq. (4). They would only change the input numerical value for the $`\delta _P`$ within our approach. The value for the decay rate $`R_{\tau S=0}`$ has been measured by the ALEPH and OPAL collaborations with results very close to each other. For definiteness we use the ALEPH data and briefly comment on the OPAL data later on. With the experimental result $$R_{\tau S=0}^{exp}=3.492\pm 0.016$$ (6) one obtains from eq. (4) $$\delta _P^{exp}=0.203\pm 0.007.$$ (7) The basic object of the theoretical calculation is Adler’s $`D`$-function which is computable in perturbation theory in the Euclidean domain. In the $`\overline{\mathrm{MS}}`$-scheme the perturbative expansion for the $`D`$-function is given by $`D(Q^2)`$ $`=`$ $`1+{\displaystyle \frac{\alpha _s(Q)}{\pi }}+k_1\left({\displaystyle \frac{\alpha _s(Q)}{\pi }}\right)^2+k_2\left({\displaystyle \frac{\alpha _s(Q)}{\pi }}\right)^3`$ (9) $`+k_3\left({\displaystyle \frac{\alpha _s(Q)}{\pi }}\right)^4+O(\alpha _s(Q)^5)`$ with (see e.g. ) $`k_1`$ $`=`$ $`{\displaystyle \frac{299}{24}}9\zeta (3),`$ (10) $`k_2`$ $`=`$ $`{\displaystyle \frac{58057}{288}}{\displaystyle \frac{779}{4}}\zeta (3)+{\displaystyle \frac{75}{2}}\zeta (5).`$ (11) Here $`\zeta (x)`$ is Riemann’s $`\zeta `$-function. In the following we use the notation $$a_s(Q)=\frac{\alpha _s(Q)}{\pi }$$ (12) for the standard $`\overline{\mathrm{MS}}`$-coupling constant normalized at the scale $`\mu =Q`$. Numerically we find $`D(Q^2)`$ $`=`$ $`1+a_s(Q)+1.6398a_s(Q)^2+6.3710a_s(Q)^3`$ (14) $`+k_3a_s(Q)^4+O(a_s^5(Q)).`$ The coefficient $`k_3`$ is still unknown which prevents us from using the last term in eq. (14) for our analysis. We nevertheless list this term throughout the paper to obtain a feeling for the possible magnitude of the $`O(\alpha _s^4)`$ correction. The particular numerical value of $`k_325`$ is obtained on the basis of geometric series approximation for the series (14) and is often used in the literature . In our analysis we do not use any particular numerical value for $`k_3`$ and only give some illustrative results of the influence of this term on the numerical value of the coupling constant extracted from $`\tau `$-data. In the $`\overline{\mathrm{MS}}`$-scheme the perturbative correction $`\delta _P`$ is given by the perturbation theory expansion $`\delta _P^{th}`$ $`=`$ $`a_s+5.2023a_s^2+26.366a_s^3`$ (16) $`+(78.003+k_3)a_s^4+O(a_s^5)`$ where the $`\overline{\mathrm{MS}}`$-scheme coupling constant $`\alpha _s=\pi a_s`$ is taken at the scale of the $`\tau `$-lepton mass $`\mu =M_\tau =1.777\mathrm{GeV}`$. Usually one extracts a numerical value for $`\alpha _s(M_\tau )`$ by treating the first three terms of the expression in eq. (16) as an exact function – the cubic polynomial, i.e. one solves the equation $`a_s+5.2023a_s^2+26.366a_s^3=\delta _P^{exp}.`$ (17) The solution reads $$\pi a_s^{st}(M_\tau )\alpha _s^{st}(M_\tau )=0.3404\pm 0.0073_{exp}.$$ (18) We call this method the standard method. The quoted error is due to the error in the input value of $`\delta _P^{exp}`$. We retain some additional decimal points in the numerical expression for the coupling constant in order to use them for the evolution of the coupling constant to the scale $`M_Z`$. It is rather difficult to estimate the theoretical uncertainty of the procedure itself. The main problem is to estimate the quality of the approximation for the (asymptotic) series in eq. (16) given by the cubic polynomial in eq. (17). As a criterion of the quality of the approximation one can use the pattern of convergence of the series (16) which is $$\delta _P^{exp}=0.203=0.108+0.061+0.034+\mathrm{}$$ (19) One sees that the corrections provide a 100% change of the leading term. Another criterion is the order-by-order behavior of the extracted numerical value for the coupling constant. In consecutive orders of perturbation theory (LO - leading order, NLO - next-to-leading order, NNLO - next-next-to-leading order) one has $`\alpha _s^{st}(M_\tau )_{LO}=0.6377,\alpha _s^{st}(M_\tau )_{NLO}=0.3882,`$ (20) $`\alpha _s^{st}(M_\tau )_{NNLO}=0.3404.`$ (21) Formally we obtain a series for the numerical value of the coupling constant of the form $`\alpha _s^{st}(M_\tau )_{NNLO}`$ $`=`$ $`0.63770.24950.0478\mathrm{}`$ (22) Limiting ourselves to the next-to-next-to-leading order result (NNLO) we can take a half of the last term as an estimate of the theoretical uncertainty. It is only an indicative estimate. No rigorous justification can be given for such an assumption about the accuracy of the approximation without knowledge of the structure of the whole series. Nevertheless we stick to this definition for our purposes. The theoretical uncertainty obtained in such a way – $`\mathrm{\Delta }\alpha _s^{st}(M_\tau )_{th}=0.0239`$ – is much larger than the experimental uncertainty given in eq. (18). This is a challenge for the theory: the accuracy of theoretical formulas cannot compete with experimental precision at present. Assuming this theoretical uncertainty we have $$\alpha _s^{st}(M_\tau )_{NNLO}=0.3404\pm 0.0239_{th}\pm 0.0073_{exp}.$$ (23) Theory dominates the error even if the estimate for its precision $`\pm 0.0239_{th}`$ is not reliable (heuristic and only indicative). Thus the straightforward analysis in the $`\overline{\mathrm{MS}}`$-scheme is not stable numerically and the naive estimate of the theoretical uncertainty is large. The use of the $`\overline{\mathrm{MS}}`$-scheme is not obligatory for practical calculations. The $`\overline{\mathrm{MS}}`$-scheme has a history of success for massless calculations where its results look natural and the corrections are usually small. This is not the strict rule, however, and there are cases (like gluonic correlators ) where corrections dramatically depend on the quantum numbers of the operators. In fact, the $`\overline{\mathrm{MS}}`$-scheme is rather artificial. It is simply defined by convention (let us be remindful of the evolution from the MS-scheme to the $`\overline{\mathrm{MS}}`$-scheme which had its origin only in technical convenience ). From technical point of view, in practical calculations of massless diagrams of the propagator type, another scheme – the $`G`$-scheme – is the most natural one . It normalizes the basic quantity of the whole calculation within integration-by-parts technique – one loop masless scalar diagram – to unity . $`\beta `$-functions coincide in both schemes. It could have well happened that the $`G`$-scheme would be historically adopted as the reference scheme because corrections in this scheme are typically smaller than that in the $`\overline{\mathrm{MS}}`$-scheme. However, for the tau system the direct (standard) analysis in the $`G`$-scheme fails. Therefore different schemes used for the numerical analysis can produce rather different numerical results for the final reference quantity - the coupling constant in the $`\overline{\mathrm{MS}}`$-scheme. Note that strictly speaking any scheme is suitable for a given perturbative calculation. However, it can lead to unusual (or even unacceptable) results in a numerical analysis. The only criterion for the choice of scheme at present is the heuristic requirement of fast explicitl convergence: the terms of the series should decrease. Clearly this is a rather unreliable criterion. It does not provide strict quantitative constraints necessary for the level of precision usually claimed for the $`\tau `$-system analysis. In the following we suggest a new procedure for extracting $`\alpha _s`$ in the $`\overline{\mathrm{MS}}`$-scheme from the $`\tau `$ system without explicit use of eq. (16). This procedure is applicable to any observable in whatever scheme it was originally computed. The observation is that any perturbation theory observable generates a scale due to dimensional transmutation and this is its internal scale. It is natural for a numerical analysis (and is our suggestion) to determine this scale fisrt and then to transform the result into a $`\overline{\mathrm{MS}}`$-scheme parameter (or any other reference scheme) using the renormalization group invariance. We deliberately use the explicit renormalization scheme invariance of the theory to bring the result of the perturbation theory calculation into a special scheme first, then we perform a numerical analysis in this particular scheme. Only after that we transform the obtained numbers into the reference $`\overline{\mathrm{MS}}`$-scheme. The last step is done only for comparison with other experiments (or just for convenience; the system itself can be well described in its internal scheme without any reference to the $`\overline{\mathrm{MS}}`$-scheme). This is our suggestion for the resolution of the problem of numerical instability of extracting parameters from truncated perturbation theory expressions. A dimensional scale in QCD emerges as a boundary value parameterizing the evolution trajectoty of the coupling constant. The renormalization group equation $$\mu ^2\frac{d}{d\mu ^2}a(\mu ^2)=\beta (a(\mu ^2)),a=\frac{\alpha }{\pi }$$ (24) is solved by the integral $$\mathrm{ln}\left(\frac{\mu ^2}{\mathrm{\Lambda }^2}\right)=\mathrm{\Phi }(a(\mu ^2))+_0^{a(\mu ^2)}\left(\frac{1}{\beta (\xi )}\frac{1}{\beta _2(\xi )}\right)𝑑\xi $$ (25) where the indefinite integral $`\mathrm{\Phi }(a)`$ is normalized as follows $$\mathrm{\Phi }(a)=^a\frac{1}{\beta _2(\xi )}𝑑\xi =\frac{1}{a\beta _0}+\frac{\beta _1}{\beta _0^2}\mathrm{ln}\left(\frac{a\beta _0^2}{\beta _0+a\beta _1}\right).$$ (26) Here $`\beta _2(a)`$ and $`\beta (a)`$ denote the second order and full $`\beta `$ function, or as many terms as are available, given by $`\beta _2(a)`$ $`=`$ $`a^2(\beta _0+a\beta _1),`$ (27) $`\beta (a)`$ $`=`$ $`a^2(\beta _0+\beta _1a+\beta _2a^2+\beta _3a^3)+O(a^6),`$ (28) $`a`$ is a generic coupling constant. The four-loop $`\beta `$-function coefficient $`\beta _3`$ is now known in the $`\overline{\mathrm{MS}}`$-scheme $$\beta _3=\frac{140599}{4608}+\frac{445}{32}\zeta (3)=47.228\mathrm{}$$ (29) The integration constant in eq. (25) is adjusted such that the asymptotic expansion of the coupling constant at large momenta $`Q^2\mathrm{}`$ reads $`a(Q^2)`$ $`=`$ $`{\displaystyle \frac{1}{\beta _0L}}\left(1{\displaystyle \frac{\beta _1}{\beta _0^2}}{\displaystyle \frac{\mathrm{ln}(L)}{L^2}}\right)+O\left({\displaystyle \frac{1}{L^3}}\right),`$ (30) $`L`$ $`=`$ $`\mathrm{ln}\left({\displaystyle \frac{Q^2}{\mathrm{\Lambda }^2}}\right).`$ (31) This serves to define the parameter $`\mathrm{\Lambda }`$ (dimensional scale) for a generic coupling constant. $`\mathrm{\Lambda }_s`$ is the standard $`\overline{\mathrm{MS}}`$-scheme scale for the coupling constant $`a_s`$. The solution (25) of the renormalization group equation (24) describes the evolution trajectory of the coupling constant. This trajectory is parametrized by the scale parameter $`\mathrm{\Lambda }`$ and the coefficients of the $`\beta `$ function $`\beta _i`$ with $`i>2`$ (see e.g. ). The evolution is invariant under the renormalization group transformation $$aa(1+\kappa _1a+\kappa _2a^2+\kappa _3a^3+\mathrm{})$$ (32) with the simultaneous change $$\mathrm{\Lambda }^2\mathrm{\Lambda }^2e^{\kappa _1/\beta _0},$$ (33) $`\beta _{0,1}`$ left invariant and $`\beta _2`$ $``$ $`\beta _2\kappa _1^2\beta _0+\kappa _2\beta _0\kappa _1\beta _1`$ (34) $`\beta _3`$ $``$ $`\beta _3+4\kappa _1^3\beta _0+2\kappa _3\beta _0+\kappa _1^2\beta _12\kappa _1(3\kappa _2\beta _0+\beta _2).`$ (35) If this transformation was considered to be exact and the exact $`\beta `$-function corresponding to the new charge was used then it would be just a change of variable in a differential equation (24) or the exact reparametrization of the trajectory (25) and hence would lead to identical results. However, the renormalization group invariance of eq. (25) is violated in higher orders of the coupling constant because we consistently omit higher orders in the perturbation theory expressions for the $`\beta `$-functions. This is the point where the finite-order perturbation theory approximation for the respective $`\beta `$-functions is made. This is the source for different numerical outputs of analyses in different schemes. Our procedure for the extraction of $`\alpha _s`$ is heavily based on the formal renormalization group invariance of the theory. We claim that because of this invariance we can do our numerical analysis in any scheme. The reason for the choice of a particular scheme is only the quality of the convergence (which, of course, is subject to some personal taste). We have chosen the effective scheme because we consider it to be more consistent and more stable numerically. Technically we introduce an effective charge $`a_\tau `$ through the relation $$\delta _P^{th}=a_\tau \frac{\alpha _\tau }{\pi }$$ (36) and extract the parameter $`\mathrm{\Lambda }_\tau `$ which is associated with $`a_\tau `$ through eq. (25). This is just the internal scale associated with the physical observable $`R_\tau `$. The effective $`\beta `$-function is given by the expression $$\beta _\tau =a_\tau ^2(\beta _{\tau 0}+\beta _{\tau 1}a_\tau +\beta _{\tau 2}a_\tau ^2+\beta _{\tau 3}a_\tau ^3+\mathrm{})$$ (37) with $`\beta _{\tau 0}=\beta _0`$, $`\beta _{\tau 1}=\beta _1`$, and $$\beta _{\tau 2}=12.3204,\beta _{\tau 3}=182.719+\frac{9}{2}k_3.$$ (38) The extraction of the numerical value for the internal scale $`\mathrm{\Lambda }_\tau `$ is done from equation (25) with $`a_\tau (M_\tau )=\delta _P^{exp}`$. The coefficient $`\beta _{\tau 3}`$ does not enter the analysis. The parameter $`\mathrm{\Lambda }_s\mathrm{\Lambda }_{\overline{\mathrm{MS}}}`$ is found according to eq. (33). The $`\overline{\mathrm{MS}}`$ coupling at $`\mu =M_\tau `$ is obtained by solving eq. (25) for $`a_s(M_\tau )`$ with regard to $`\mathrm{ln}(M_\tau ^2/\mathrm{\Lambda }_s^2)`$ which is known if $`\mathrm{\Lambda }_s`$ is obtained; the $`\beta `$-function is taken in the $`\overline{\mathrm{MS}}`$-scheme. For consistency reasons we only use the $`\overline{\mathrm{MS}}`$-scheme $`\beta `$-function to three-loop order since the effective $`\beta `$-function $`\beta _\tau `$ is only known up to the second order, cf. eq. (38). A $`N^3LO`$ analysis is possible only if a definite value is chosen for $`k_3`$. We give some estimates later. Our procedure is based on renormalization group invariance and one can start from the expression for the decay rate obtained in any scheme. The only perturbative objects present are the $`\beta `$-functions. Both $`\beta _{\overline{\mathrm{MS}}}`$ and $`\beta _\tau `$, however, converge reasonably well which is the only perturbation theory restriction in our method. It also highlights the limit of precision within our procedure: the expansion for $`\beta _\tau `$ is believed to be asymptotic as any expansion in perturbation theory. The asymptotic expansion provides only limited accuracy for any given numerical value of the expansion parameter which cannot be further improved by including higher order terms. The expansion used is presumably rather close to its asymptotic limit as can be seen by taking a look at the expansion $`\beta _\tau (a_\tau )`$ $`=`$ $`a_\tau ^2({\displaystyle \frac{9}{4}}+4a_\tau 12.3204a_\tau ^2`$ (40) $`+a_\tau ^3(182.719+{\displaystyle \frac{9}{2}}k_3))+O(a_\tau ^6)`$ with $`a_\tau 0.2`$ at the scale $`M_\tau `$. The convergence of the series depends crucially on the numerical value of $`k_3`$. If $`k_3`$ had a value where the asymptotic growth starts at third order then further improvement of the accuracy within finite-order perturbation theory is impossible. At every order of the analysis we use the whole information of the perturbation theory calculation. Especially, the appropriate coefficient of the $`\beta _\tau `$-function is present. In the standard method the coefficient $`\beta _2`$ enters only at order $`O(\alpha _s^4)`$ of the $`\tau `$-lepton decay rate expansion. We call our procedure the renormalization scheme invariant extraction method (RSI) hoping that it is clear what is meant by this name from our explanations. Note also that $`\alpha _s`$ itself is not a physical object and is renormalization scheme noninvariant. In this respect we extract the noninvariant parameter $`\alpha _s`$ using invariance of the physics in order to perform the numerical analysis in the most suitable scheme. Then the output of the analysis is simply transformed into a numerical value for $`\alpha _s`$ according to the renormalization group transformation rules. For the coupling constant in the $`\overline{\mathrm{MS}}`$-scheme in NNLO we find $`\alpha _s^{RSI}(M_\tau )`$ $`=`$ $`0.3184\pm 0.0060_{exp}`$ (41) which is smaller than the corresponding value obtained within the standard procedure eq. (18). How to estimate the quality of this result? The parameter which is really extracted in consecutive orders of perturbation theory within our method is the scale $`\mathrm{\Lambda }_\tau `$. Because of the relation (see eqs. (16,32,33)) $$\mathrm{\Lambda }_s=\mathrm{\Lambda }_\tau e^{5.20232/2\beta _0}=0.3147\mathrm{\Lambda }_\tau $$ (42) we can look at $`\mathrm{\Lambda }_s`$ directly. We find $`\mathrm{\Lambda }_s|_{LO}=595\mathrm{MeV},\mathrm{\Lambda }_s|_{NLO}=288\mathrm{MeV},`$ (43) $`\mathrm{\Lambda }_s|_{NNLO}=349\mathrm{MeV}`$ (44) or, representing the NNLO result as a formal series, $$\mathrm{\Lambda }_s|_{NNLO}=595307+61\mathrm{}\mathrm{MeV}.$$ (45) Note that at leading order the scales (as well as charges) are equal in all schemes. Therefore the leading order result ($`\mathrm{\Lambda }_s|_{LO}=595\mathrm{MeV}`$) is not representative, only indicative. Assuming according to our convention that the uncertainty of $`\mathrm{\Lambda }_s`$ is given by the half of the last term of the series (45) we have $$\mathrm{\Lambda }_s=349\pm 31\mathrm{MeV}$$ (46) which leads to the numerical value for the $`\overline{\mathrm{MS}}`$-scheme coupling constant $$\alpha _s=0.3184_{+0.0160}^{0.0157}.$$ (47) This result is obtained from eq. (25) with three-loop $`\beta `$-function. Taking the average we find $$\alpha _s=0.3184\pm 0.0159.$$ (48) This is better than the theoretical error of the standard result eq. (23). Still the theoretical error should be considered as a guess rather than a well-justified estimate of the uncertainty. Let us briefly comment on the $`k_3`$ contribution. Clearly the estimate $`k_3=25`$ is rather speculative. We, therefore, use a different strategy in the analysis. We determine the range of $`k_3`$ which is safe for explicit convergence of perturbation theory. If the actual value of $`k_3`$ will be discovered in this range then perturbation theory is still valid and will give better accuracy in NNNLO. If not, the asymptotic growth of perturbation theory series is already reached and its accuracy cannot be improved. We require that the last term is equal to the half of the previous one. In the standard way (eq. 16) we have $$|(78+k_3)a_s|<\frac{1}{2}\mathrm{\hspace{0.17em}26.36}13$$ (49) which for $`a_s=0.1`$ gives $$208<k_3^{st}<52.$$ (50) In the RSI way (eq. 40) we have $$|(182+\frac{9}{2}k_3)a_\tau |<\frac{1}{2}\mathrm{\hspace{0.17em}12.32}6$$ (51) which for $`a_\tau =0.2`$ gives $$33.8<k_3^\tau <47.1.$$ (52) This range is much narrower than that in eq. (50). The effective scheme method is much more sensitive to the structure of the series as can be seen from eq. (40). The actual precision depends on the actual value chosen for $`k_3`$ and it is rather premature to speculate about numbers. Still we show the worst result (in the optimistic scenario that $`k_3`$ lies in the safe range) that can be expected within the RSI approach. In the RSI approach with $`k_3=47`$ we find the scale parameter in NNNLO $$\mathrm{\Lambda }_s|_{NNNLO}=334\mathrm{MeV}.$$ (53) With $`k_3=34`$ one has $$\mathrm{\Lambda }_s|_{NNNLO}=367\mathrm{MeV}.$$ (54) Taking the average we have $$\mathrm{\Lambda }_s=350\pm 17\mathrm{MeV}$$ (55) which is the best possible estimate if we require that the perturbation theory series for the $`\beta _\tau `$-function still converges (according to our quantitative criterion of convergence). That results in the numerical value for the $`\overline{\mathrm{MS}}`$-scheme coupling constant found with four-loop $`\beta `$-function from eq. (25) $$0.3133<\alpha _s<0.3314.$$ (56) Therefore our conservative estimate of the theoretical error in the optimistic scenario for the convergence of perturbation theory series in NNNLO reads $$\alpha _s=0.322\pm 0.009.$$ (57) While the estimation of the theoretical uncertainty is a tricky matter and can be considered as indicative the central numerical value of the coupling constant definitely becomes smaller as compared to the standard result. At present the reference value for the coupling constant is commonly given at the scale $`M_Z=91.187\mathrm{GeV}`$. The running to this reference scale is done with the four-loop $`\beta `$-function in the $`\overline{\mathrm{MS}}`$-scheme and three-loop matching conditions at the heavy quark (charm and bottom) thresholds . For the threshold parameters related to heavy quark masses we use $`\mu _c=\overline{m}_c(\mu _c)=(1.35\pm 0.15)\mathrm{GeV}`$ and $`\mu _b=\overline{m}_b(\mu _b)=(4.21\pm 0.11)\mathrm{GeV}`$ (e.g. ) where $`\overline{m}_q(\mu )`$ is the running mass of the heavy quark in the $`\overline{\mathrm{MS}}`$-scheme. Note that because of the truncation of matching conditions the result of the running slightly depends on at what scale the matching is actually performed. If the matching between the $`n_f=3`$ and $`n_f=4`$ effective theories is done directly at the scale $`M_\tau `$, which is possible, then the result is slighly smaller than in the case when the evolution within $`n_f=3`$ effective theory is done first to the scale $`\mu _c`$. In the following we stick to the procedure where the matching is performed precisely at the matching scales $`\mu _{c,b}`$. We first run the coupling constant within $`n_f=3`$ effective theory from the scale $`M_\tau `$ to $`\mu _c`$ then match the result to $`n_f=4`$ coupling constant, run it to $`\mu _b`$ and match to $`n_f=5`$ coupling constant. The last step is just evolution to $`M_Z`$. Note that the alternative would be to perform matching between $`n_f=3`$ and $`n_f=4`$ effective theories directly at the scale $`M_\tau `$ (because it is rather close to $`\mu _c`$) but in this case the final result is slightly smaller than in our present procedure. The running to the scale $`M_Z`$ gives the following result for the standard method estimate $$\alpha _s^{st}(M_Z)=0.1210\pm 0.0008_{exp}\pm 0.0006_c\pm 0.0001_b$$ (58) where the subscript $`exp`$ denotes the error originating from $`\delta _P^{exp}`$. The errors with subscripts $`c,b`$ arise from the uncertainty of the numerical values of the charm and bottom quark masses that enter the evolution analysis. These errors are rather small (we retain the additional decimal place in the result, which is not really justified from the precision of the experimental input, just to show these uncertainties). If the matching between the $`n_f=3`$ and $`n_f=4`$ effective theories is done directly at the scale $`M_\tau `$ one has to change the central value $`0.12100.1202`$ which shows the uncertainty related to the truncation of the matching conditions. The central value in eq. (58) is slightly higher than that calculated from high energy experiments . The theoretical perturbative expansions for observables in high energy experiments converge better numerically than expansions at low energies because the coupling, which is the parameter of the perturbative expansion, is smaller at higher energies due to the property of asymptotic freedom. This feature makes it less important to treat the higher order terms carefully in high energy applications as compared to the low energy $`\tau `$-lepton estimates. However, the experimental data in high energy experiments are usually less precise which leads to large errors in the $`\alpha _s`$ determination from high energy experiments. The fact that the value in eq. (58) is higher than that calculated from high energy experiments caused some discussion about the reliability of estimates from the $`\tau `$-lepton data. Our analysis resolves this problem. The running of $`\alpha _s^{RSI}(M_\tau )`$ given in eq. (48) to $`M_Z`$ with the four-loop $`\beta `$-function and with three-loop heavy quark matching accuracy gives $`\alpha _s^{RSI}(M_Z)`$ $`=`$ $`0.1184\pm 0.00074_{exp}`$ (60) $`\pm 0.00053_c\pm 0.00005_b`$ where we have kept five decimal places in order to exhibit the magnitude of different sources of uncertainty. Eq. (60) constitutes our main result for the coupling $`\alpha _s(M_Z)`$ derived from tau data. The OPAL collaboration has reported an experimental value of $`R_{\tau S=0}^{exp}=3.484\pm 0.024`$ . This leads to $`\delta _P^{exp}=0.200\pm 0.009_{exp}`$ and $$\alpha _s^{RSI}(M_\tau )=0.3158\pm 0.0078_{exp}$$ (61) which, when evolved to $`M_Z`$, gives $`\alpha _s^{RSI}(M_Z)`$ $`=`$ $`0.1181\pm 0.00097_{exp}`$ (63) $`\pm 0.00052_c\pm 0.00005_b.`$ This value is close to the one in eq. (60) based on the ALEPH data. The theoretical uncertainty comes mainly from the truncation of the perturbation theory series. Taking the result of the NNLO analysis eq. (48) we find $$\mathrm{\Delta }\alpha _s^{RSI}(M_Z)_{th}=0.0019$$ (64) In the most optimistic scenario with the NNNLO analysis eq. (57) one has $$\alpha _s^{RSI}(M_Z)_{N^3LO}=0.119\pm 0.001.$$ (65) As we have already noted the interpretation of the higher order terms in the perturbation theory expansion is numerically important for the analysis of the $`\tau `$-data. The regular method to resum higher order perturbation theory corrections is based on the direct integration of the renormalization group improved correlators over the contour in the complex $`Q^2`$ plane . This method allows one to resum corrections generated by the running of the coupling constant along the integration contour and is now widely used for the analysis of the $`\tau `$-data. We now briefly comment on the extraction of the strong coupling constant within resummed perturbation theory. As in ref. we fit the theoretical expression for the decay rate in the contour improved approach to the experimental result $`\delta _P^{exp}`$ eq. (7) and find $$\alpha _s^{CI}(M_\tau )=0.343\pm 0.009_{exp}$$ (66) within the renormalization scheme invariant extraction method described above i.e. with the introduction of the effective charge first. This value differs from the finite-order perturbation theory result eq. (41). Note that the two values extracted from finite-order perturbation theory analysis eq. (41) and the contour improved perturbation theory analysis eq. (66) do not overlap within their respective error bars given from the experimental uncertainty only. This situation was anticipated in where the resummed NNLO analysis had been first performed. The point is clear: resummation provides a specific estimate of higher order terms. In finite-order perturbation theory one adopts a model where all higher order terms have been neglected. In contour improved perturbation theory one adopts an explicit model with higher order terms generated by the running of the coupling constant along the integration contour. With present experimental accuracy one can already distinguish between these two possibilities. One should always keep in mind that the two determinations eq. (41) and eq. (66) result from different models and one should not mix their predictions. The numerical value of the coupling constant appropriate for high energy experiments is normally small (much smaller than for $`\tau `$-data) and perturbation theory converges faster (in similar kinematical situations). The resummation does not produce any big numerical changes. Therefore finite-order perturbation theory is normally used for the analysis of high energy experiments (resummation of the contour type can be done but produces a small numerical effect) and one usually quotes numerical values of the coupling constant extracted with finite-order perturbation theory. Or resummation of the sort different from that used for the $`\tau `$ system is used (like Coulomb type resummation for heavy quarks ). Therefore we suggest to use the finite-order perturbation theory prediction for the coupling constant extracted from $`\tau `$-data in order to compare it with the results of high energy experiments. To conclude, we have extracted the numerical value of the strong coupling constant from $`\tau `$-data within a procedure based on explicit use of renormalization scheme invariance. The numerical value for the coupling constant is systematically smaller than that derived by the standard treatment. When evolved to $`M_Z`$ our $`\overline{\mathrm{MS}}`$-scheme value for the coupling constant extracted in finite-order perturbation theory reads $$\alpha _s(M_Z)=0.1184\pm 0.0007_{exp}\pm 0.0006_{hqmass}.$$ (67) This central value is closer to the value of $`\alpha _s`$ derived from high energy experiments than previous determinations of $`\alpha _s`$ from $`\tau `$-data. The theoretical uncertainty of the result is still only indicative: it ranges from the conservative estimate in NNLO $`\mathrm{\Delta }\alpha _s(M_Z)_{th}=\pm 0.0019`$ to an optimistic one based on the assumption about NNNLO contribution $`\mathrm{\Delta }\alpha _s(M_Z)_{th}=\pm 0.001`$. The present work is supported in part by the Volkswagen Foundation under contract No. I/73611 and by the Russian Fund for Basic Research under contract 99-01-00091.
warning/0002/astro-ph0002201.html
ar5iv
text
# Chandra uncovers a hidden Low-Luminosity AGN in the radio galaxy Hydra A (3C 218) ## 1 Introduction Recent HST and ground-based optical observations provided strong evidence that many nearby galaxies harbor supermassive black holes (e.g., Kormendy & Richstone 1995), possibly accreting at sub-Eddington luminosities (Fabian & Rees 1995). About 40% of nearby early-type galaxies exhibit signs of mild nuclear activity, in the form of weak non-thermal radio cores (Sadler et al. 1989) and Low Ionization Emission Lines (LINERs; Heckman 1986; Ho, Fillippenko, & Sargent 1997a). These results collectively suggest that many nearby galaxies may harbor weak nuclear activity in the form of a Low-Luminosity Active Galactic Nucleus (LLAGN; e.g., Ho 1999a). Thanks to their high penetrating power, X-rays provide an optimal window to search for weak nuclear activity. Indeed, previous X-ray ROSAT images of a handful of galaxies show the presence of a central unresolved nucleus within the 5″ HRI resolution (e.g., Fabbiano 1996). Indirect clues are provided by ASCA spectral constraints: a heavily absorbed power law component is often measured at energies $`2`$ keV, with intrinsic luminosities L$`{}_{210keV}{}^{}10^{4042}`$ erg s<sup>-1</sup> , photon indices $`\mathrm{\Gamma }_{210keV}1.51.7`$, and a narrow Fe emission line at 6–7 keV in a few cases, suggestive of a LLAGN (Makishima et al. 1994; Ptak et al. 1999; Sambruna, Eracleous, & Mushotzky 1999; Terashima et al. 1999). Within the coarse angular resolution of these detectors, alternative scenarios can not be ruled out in many cases (e.g., a starburst or X-ray binaries). Unambiguous evidence for nuclear activity would be provided by the detection of a point source at the galaxy center. With its unprecedented angular resolution (0.5″), wide-band coverage (0.2–10 keV), and high sensitivity, Chandra is uniquely suited to this task. In this Letter, using recent Chandra calibration observations we discover a LLAGN in the LINER harbored by the nearby cD galaxy Hydra A ($`z`$=0.0537). Host of the powerful FR I radio source 3C 218, Hydra A is the dominant member of the poor cluster of galaxies A780, and famous for its twin-jet radio morphology (Taylor et al. 1990). The nuclear region contains a $``$ 6″ emission-line nebula (Baum et al. 1989; Heckman et al. 1989) and a disk of star formation (McNamara 1995; Melnick, Gopal-Krishna, & Terlevich 1997). At the position of the nucleus, a LINER-like optical and UV spectrum is observed (Hansen, Jørgensen, & Nørgaard-Nielsen 1995), which, together with the powerful lobe radio emission, led to the classification of Hydra A as a Weak Line Radio Galaxy (Tadhunter et al. 1998). The X-ray emission from Hydra A in previous Einstein, ROSAT, and ASCA images is dominated by the cluster and the inner ($``$ 1.5′) cooling flow, with $`L_{0.54.5keV}^{cluster}2\times 10^{44}`$ erg s<sup>-1</sup> (David et al. 1990; Ikebe et al. 1997). In the following, we assume a Friedman cosmology with $`H_0=75`$ km s<sup>-1</sup> Mpc<sup>-1</sup> and $`q_0=0.5`$. At the luminosity distance of Hydra A (217.46 Mpc), 1″=0.95 kpc. ## 2 Observations and Data analysis Hydra A was observed with Chandra ACIS-S for 20 ks on 1999 November 11 at the aimpoint of S3 in faint telemetry mode with 5 CCDs turned on. The data were analyzed using the `EventBrowser` tool at Penn State, and the `CIAO` software provided by the Chandra X-ray Center. Only events for ASCA grades 0, 2, 3, 4, and 6 were accepted. The S3 background was rather stable during the observation. Spectral analysis was performed within `XSPEC` v.10.0 using response files appropriate for the S3 aimpoint and for the epoch of the present observation (`ccd7_c0.7.15.32.rmf`, `s3_c1_middle.arf`). The spectra were rebinned over the energy range 0.2–6 keV to have a minimum of 20 counts in each bin, to validate the use of the $`\chi ^2`$ statistics. With a total count rate of $``$ 0.02 c/s, pileup is not a concern. At this time of writing, there appears to be an uncertainty of the order of 10% in the total effective area of the telescope mirror in the 1–2 keV energy range. Fortunately, for the present analysis the energy range of interest for the nuclear properties is restricted to energies above 2 keV, where the AGN component dominates. The errors in the on-axis effective area above 2 keV are of the order of 5% (Jerius et al. 2000). Hydra A was also observed with ACIS-I for 20 ks in 1999 October 30, after the CCDs suffered radiation damage resulting in a degradation of the gain and spectral resolution. To improve the signal-to-noise ratio we extracted the ACIS-I spectrum of the nucleus in a region of similar size as for the ACIS-S data, and analyzed the data using a position-dependent spectral response including an empirical CTI correction. The 2–6 keV ACIS-I spectrum of the nucleus was fitted simultaneously to the ACIS-S data leaving the intercalibration factors free to vary, giving a ratio of the two normalizations of $`N_{ACISS}/N_{ACISI}=1.09\pm 0.22`$. ## 3 Results Figure 1 shows the S3 image of the central regions of Hydra A in 0.2–10 keV, with the VLA radio contours overlayed (Taylor et al. 1990). The ACIS-S data were adaptively smoothed using a Gaussian function with a kernel of 3.5″ , and shifted by 3.6″ to align the X-ray and radio core and the other X-ray/radio morphology to better than 0.5″. This shift is well within the range of uncertainties on the astrometry (0.5″– 5″) measured in several other ACIS observations during the orbital activation and check-out phases. X-ray emission from the compact radio core is readily apparent, embedded in diffuse soft X-ray emission. In the 0.2–10 keV band, $``$ 900 counts are detected within 2.5″ from the position of the radio core and the radial profile of the X-ray source is extended. At energies $``$ 2 keV, the radial profile is point-like and a total of $``$ 150 counts are collected in 2–10 keV within 1.5″ (or 80% of the total encircled energy). The hard X-ray point source is also present in the 20 ks ACIS-I exposure. Thus, with Chandra we were able to detect a point-like X-ray source for the first time in Hydra A at hard energies, indicating that the core radio emission is due to a hidden AGN. Interestingly, no X-ray emission is detected from either the jets or lobes (Fig. 1). On the contrary, the extended radio structures appear to occupy regions relatively deficient in X-ray photons. This is opposite to what recently detected with Chandra in the FR II radio galaxy 3C 295 (Harris et al. 2000). Figure 2 shows the inner 10″ region around the nucleus in contour form, in 3.7–4.3 keV. The nuclear point source is apparent, together with extended faint emission elongated by $``$ 3″ in a N-W direction. Note the extension in the same direction in the radio contours (Fig. 1). At roughly this distance, a spot of enhanced blue emission was observed in previous $`B`$ and $`U`$ band images, identified as the bluest edge of a star forming disk (McNamara 1995; Melnick et al. 1997). A total of $``$ 260 counts are detected within 2″ over the 0.2–10 keV band for this region. The study of the starburst and of the large-scale X-ray emission in Hydra A are beyond the scope of this paper, and will not be discussed any further (see McNamara et al. 2000). The nuclear X-ray spectrum was extracted from a circular region with radius 1.5″. The background was extracted in an annulus centered on the nucleus with inner and outer radii 1′ and 1.3′, respectively. The cluster contribution in an area similar to the nuclear region is small, $``$ 6%, while the instrumental and cosmic background is negligible, $`2\times 10^5`$ ph s<sup>-1</sup> arcsec<sup>-2</sup>. The spectrum was fitted with a two-component model, including an absorbed power law at energies $``$ 2 keV, and a Raymond-Smith thermal plasma at softer energies. This model is a very good description of the ACIS-S data, with $`\chi ^2`$=17 for 23 degrees of freedom. The data and residuals are shown in Figure 3. All spectral components in the fit were absorbed by the Galactic column density, N$`{}_{H}{}^{Gal}=4.9\times 10^{20}`$ cm<sup>-2</sup> (Dickey & Lockman 1990). The fitted parameters of the absorbed power law are rest-frame column density N$`{}_{H}{}^{int}=2.8_{1.4}^{+3.0}\times 10^{22}`$ cm<sup>-2</sup> and photon index $`\mathrm{\Gamma }=1.75_{0.20}^{+1.14}`$ (uncertainties are 90% confidence for one parameter of interest, $`\mathrm{\Delta }\chi ^2`$=2.7). This is the spectrum of the hard X-ray point source coincident with the radio core in Figure 1. The slope is consistent within 1$`\sigma `$ with the average value measured with ASCA for other Weak Line Radio Galaxies (WLRGs), $`\mathrm{\Gamma }_{210keV}=1.49`$ and dispersion $`\sigma _\mathrm{\Gamma }=0.30`$ (Sambruna et al. 1999). The observed fluxes are F$`{}_{0.22keV}{}^{}9\times 10^{15}`$ and F$`{}_{210keV}{}^{}1.8\times 10^{13}`$ erg cm<sup>-2</sup> s<sup>-1</sup>. The intrinsic (absorption-corrected) luminosity is L$`{}_{210keV}{}^{}1.2\times 10^{42}`$ erg s<sup>-1</sup>, at the high-end of the distribution for WLRGs and LINERs. The data require a thermal component at soft energies, with fitted temperature $`kT=1.05_{0.14}^{+0.32}`$ keV, consistent with the value measured for other radio sources with ROSAT and ASCA (Worrall & Birkinshaw 1994; Sambruna et al. 1999). The abundance was fixed to the best-fit value, 0.1 solar. The observed fluxes are F$`{}_{0.22keV}{}^{}3\times 10^{14}`$ and F$`{}_{210keV}{}^{}5\times 10^{15}`$ erg cm<sup>-2</sup> s<sup>-1</sup>. A possible origin of the thermal component is the halo of the host galaxy. Indeed, the intrinsic luminosity is L$`{}_{0.22keV}{}^{}2\times 10^{41}`$ erg s<sup>-1</sup>, consistent with the values measured for elliptical galaxies. Alternatively, the thermal component could be due to the cooling flow. ## 4 Discussion and Conclusions Hydra A is the second powerful radio galaxy observed with Chandra, and the second instance where a nuclear X-ray point source is detected. A luminous AGN was also discovered in an ACIS-S 20 ks image of the FR II radio galaxy 3C 295 (Harris et al. 2000), with $`\mathrm{\Gamma }1.8`$ and $`L_{0.210keV}7\times 10^{43}`$ erg s<sup>-1</sup>. As in Hydra A, the AGN in 3C 295 resides in a cooling flow. These results confirm that optically-weak, narrow-emission line radio sources harbor true AGNs as their brighter, broad-emission line Seyfert-like counterparts. Thanks to Chandra, a detailed study of the central engines of these systems becomes possible for the first time. The large X-ray column density we measure in the nucleus of Hydra A, N$`{}_{H}{}^{int}3\times 10^{22}`$ cm<sup>-2</sup>, implies an optical extinction $`A_V=10`$, assuming galactic gas-to-dust ratios (Bohlin, Savage, & Drake 1978). This means that the optical and UV continuum from the AGN are strongly (factor $`300`$) suppressed, accounting for why the AGN was not previously detected at these wavelengths. The weak UV continuum detected in archival IUE spectra (Hansen et al. 1995 and our own inspection of an unpublished spectrum) cannot be due to the AGN; most likely candidates are the starburst and/or the cooling flow. Future HST observations should detect only an extended thermal component to the UV light. We now turn to the issue of whether the LLAGN is capable of photoionizing the emission-line nebula identified with the nucleus (Baum et al. 1989; Heckman et al. 1989). To make this assessment we first evaluate the total reddening to the nuclear emission-line source, $`E_{\mathrm{B}\mathrm{V}}`$. Since the nuclear spectrum resembles that of LINERs, we assume that the intrinsic Balmer decrement is H$`\alpha `$/H$`\beta =3.1`$ (e.g., Ho, Filippenko, & Sargent 1997b), leading to an estimate of $`E_{\mathrm{B}\mathrm{V}}0.33`$<sup>1</sup><sup>1</sup>1This is considerably higher than $`E_{\mathrm{B}\mathrm{V}}=0.15`$ given by Hansen et al. (1995) based on the Ly$`\alpha `$/H$`\beta `$ ratio. Hansen et al. have probably underestimated the reddening because the Ly$`\alpha `$ flux was measured in the large IUE aperture which very likely includes a contribution from the circumnuclear starburst region.. With this value of the reddening and the H$`\beta `$ flux measured by Hansen et al. (1995) we obtain the rate of emission of H$`\beta `$ photons from the nebula as $`Q_{\mathrm{H}\beta }=4.1\times 10^{51}`$ s<sup>-1</sup>. If the H$`\beta `$ emission is the result of case B recombination, then atomic physics implies $`Q_{\mathrm{H}\beta }=0.12Q_{\mathrm{ion}}`$ (Osterbrock 1989), where $`Q_{\mathrm{ion}}`$ is the ionizing photon rate from the active nucleus. Thus, the observed H$`\beta `$ flux requires that $`Q_{\mathrm{ion}}=(3.4\times 10^{52}/f_\mathrm{c})`$ s<sup>-1</sup>, where $`f_\mathrm{c}`$ is the covering fraction of the source by the nebula. In contrast, the observed ionizing photon rate, assuming that the X-ray power-law spectrum extends all the way through the UV band, is $`Q_{\mathrm{ion}}=1.1\times 10^{52}`$ s<sup>-1</sup>, which falls short of the required rate by a considerable factor, even if $`f_\mathrm{c}=1`$! In fact, if we model the spectral energy distribution (SED) of Hydra A after those observed in LINERs and other LLAGN, some of which are radio-loud (Ho 1999b), the problem persists. In order to balance the photon budget of the nebula, we must either postulate that the ionizing spectrum includes a UV bump or invoke and additional power source. If we assume there is a UV bump, we can parameterize it following Mathews & Ferland (1987) and normalize the SED according to the Chandra X-ray spectrum. We then find an ionizing photon output rate of $`Q_{\mathrm{ion}}=6.4\times 10^{53}`$ s<sup>-1</sup>, which is more than enough to power the emission lines. On the other hand, one or more other power sources may contribute to the ionization of the nebula. X-rays from shocks in the cooling flow are plausible (e.g., Heckman et al. 1989). Another possibility is mechanical interaction of the emission-line gas with the radio jets, which is particularly relevant since the images of Hansen et al. (1995) show superpositions of radio and emission-line knots. To get a handle on the nature of the accretion flow, we compared the X-ray luminosity to the limiting Eddington luminosity through a rough estimate of the central black hole mass in Hydra A, $`M_{BH}`$. The apparent $`V`$ magnitude of the host galaxy in Hydra A, $`m_V=13.7`$ (Sandage 1973), and Figure 8 of Magorrian et al. (1998) imply $`M_{BH}4\times 10^9M_{}`$, consistent with the values dynamically measured in other giant ellipticals (e.g., Ford et al. 1994). The Eddington luminosity is thus $`L_{Edd}5\times 10^{47}`$ erg s<sup>-1</sup>. Assuming $`L_X=0.1L_{total}`$, the luminosity relative to Eddington is $`L_{total}/L_{Edd}2\times 10^5`$. This places Hydra A in the regime of an Advection-Dominated Accretion Flow (ADAF; e.g., Narayan, Mahadevan, & Quataert 1998), similar to other WLRGs (Sambruna et al. 1999). A powerful diagnostic of the accretion flow structure in Hydra A will be afforded by future higher-sensitivity X-ray observations, such as delivered by XMM. The EPIC spectrum will allow us to study in more detail the nuclear X-ray properties, in particular whether an Fe line is present at 6–7 keV. The line energy and profile could allow us to discriminate between a standard Seyfert-like disk and an ADAF (e.g., Sambruna & Eracleous 1999). Finally, we note that the nuclear X-ray absorption in Hydra A is much larger (factor 10) than the optical/UV extinction to the emission-line nebula (see above), and consistent with the column measured in the radio toward the core (Taylor 1996). This suggests that the X-ray absorber lies further in than the emission-line regions, close to the central black hole. A possible candidate is the molecular torus on parsec scales postulated by unification models (Urry & Padovani 1995). Indeed, the radio and optical observations indicate an edge-on geometry for Hydra A (Taylor et al. 1990; Baum et al. 1989), such that the torus intercepts the line of sight to the nucleus. Our results thus support current unification models for radio-loud sources and in particular the presence of an obscuring torus in FR I radio galaxies. Clearly, large unbiased samples are needed to reach firmer conclusions, which we anticipate from Chandra in the next few years. ###### Acknowledgements. RMS acknowledges support from NASA contract NAS–38252. We are grateful to Gordon Garmire and the ACIS team for making these observations possible. We thank Joe Pesce for help with Figure 2, Niel Brandt for the `ASMOOTH` routine, and Pat Broos and Scott Koch for assistance with data retrieving and for the `TARA` software. Finally, we are grateful to the referee, Yuichi Terashima, for his prompt and thoughtful comments and suggestions. Figure Captions * Figure 1: Chandra ACIS-S image in 0.2–10 keV of the central regions of Hydra A, in a 20 ks calibration exposure. Radio VLA contours at 6 cm are overlayed (Taylor et al. 1990). The ACIS-S image was adaptively smoothed using a Gaussian function with a kernel of 3.5″, and shifted by 3.6″ in declination to align the X-ray and radio core. X-ray emission from the compact radio core is apparent, embedded in a diffuse halo, while the jets and lobes occupy regions deficient of X-rays. * Figure 2: Contours of the inner region of Hydra A at 4 keV. North is up and East is to the left. The intensity scale is logarithmic with contour interval 0.4 or a factor 2.5 in intensity. The nuclear point source is apparent. A faint extended structure is also present at $``$ 3″ N-W of the nucleus, roughly at the position of the star formation region previously detected in optical images (McNamara 1995; Melnick et al. 1997). * Figure 3: The Chandra spectrum of the nuclear region in Hydra A, extracted in a region of radius 1.5″. The top panel shows the data convolved with the best-fit model, a highly absorbed power law at energies $``$ 2 keV and a soft thermal component (dotted lines). The bottom panel are the residuals in the form of ratio of the data to the model. Crosses are ACIS-S data, asterisks are ACIS-I data.
warning/0002/physics0002020.html
ar5iv
text
# SIMULATION OF LASER-COMPTON COOLING OF ELECTRON BEAMS**footnote *This work was supported in part by the U.S. Department of Energy under Contract No. DE-AC03-76SF00098. ## 1 Introduction A novel method of electron beam cooling for future linear colliders was proposed by V.Telnov .$`^{\text{“sevenrm}\text{?}}`$ During head-on collisions with laser photons, the transverse distribution of electron beams remains almost unchanged and also the angular spread is almost constant. Because the Compton scattered photons follow the initial electron trajectory with a small additional spread due to much lower energy of photons (a few eV) than the energy of electrons (several GeV). The emittance $`ϵ_i=\sigma _i\sigma _i^{}`$ remains almost unchanged ($`i=x,y`$). At the same time, the electron energy decreases from $`E_0`$ to $`E_f`$. Thus the normalized emittances have decreased as follows $$ϵ_n=\gamma ϵ=ϵ_{n0}(E_f/E_0)=ϵ_{n0}/C,$$ (1) where $`ϵ_{n0}`$, $`ϵ_n`$ are the initial and final normalized emittances, the factor of the emittance reduction $`C=E_0/E_f`$. The method of electron beam cooling allows further reduction of the transverse emittances after damping rings or guns by 1-3 orders of magnitude .$`^{\text{“sevenrm}\text{?}}`$ In this paper, we have evaluated the effects of the laser-Compton interaction for transverse cooling using the Monte Carlo code CAIN .$`^{\text{“sevenrm}\text{?}}`$ The simulation calculates the effects of the nonlinear Compton scattering between the laser photons and the electrons during a multi-cooling stage. Next, we examine the optics for cooling with and without chromatic correction. The laser-Compton cooling for JLC/NLC $`^{\text{“sevenrm}\text{?}}`$ at $`E_0=2`$ GeV is considered in section 4. A summary of conclusion is given in section 5. ## 2 Laser-Electron Interaction ### 2.1 Laser-Electron Interaction In this section, we describe the main parameters for laser-Compton cooling of electron beams. A laser photon of energy $`\omega _L`$ (wavelength $`\lambda _L`$) is backward-Compton scattered by an electron beam of energy $`E_0`$ in the interaction point (IP). The kinematics of Compton scattering is characterized by the dimensionless parameter $`^{\text{“sevenrm}\text{?}}`$ $$x_0\frac{4E_0\omega _L}{m_e^2c^4}=0.019\frac{E_0[\mathrm{GeV}]}{\lambda _L[\mu \mathrm{m}]},$$ (2) where $`m_e`$ is electron mass. The parameters of the electron and laser beams for laser-Compton cooling are listed in Tables 1 and 1. The parameters of the electron beam with 2 GeV are given for JLC/NLC case in section 4. The parameters of that with 5 GeV are used for simulation in the next subsection. The wavelength of laser is assumed to be 0.5 $`\mu `$m. The parameters of $`x_0`$ with the electron energies 2 GeV and 5 GeV are 0.076 and 0.19, respectively. The required laser flush energy with $`Z_Rl_\gamma l_e`$ is $`^{\text{“sevenrm}\text{?}}`$ $$A=25\frac{l_e[\mathrm{mm}]\lambda _\mathrm{L}[\mu \mathrm{m}]}{E_0[\mathrm{GeV}]}(C1)[\mathrm{J}],$$ (3) where $`Z_R`$, $`l_\gamma (2\sigma _{L,z})`$, and $`l_e(2\sigma _z)`$ are the Rayleigh length of laser, and the bunch lengths of laser and electron beams. From this formula, the parameters of $`A`$ with the electron energies 2 GeV and 5 GeV are 56 J and 4 J, respectively. The nonlinear parameter of laser field is $`^{\text{“sevenrm}\text{?}}`$ $$\xi ^2=4.3\frac{\lambda _L^2[\mu \mathrm{m}^2]}{l_e[\mathrm{mm}]E_0[\mathrm{GeV}]}(C1).$$ (4) In this study, for the electron energies 2 GeV and 5 GeV, the parameters of $`\xi `$ are 2.2 and 1.5, respectively. The rms energy of the electron beam after Compton scattering is $`^{\text{“sevenrm}\text{?}}`$ $$\sigma _e=\frac{1}{C^2}\left[\sigma _{e0}^2[\mathrm{GeV}^2]+0.7x_0(1+0.45\xi )(C1)E_0^2[\mathrm{GeV}^2]\right]^{1/2}[\mathrm{GeV}],$$ (5) where the rms energy of the initial beam is $`\sigma _{e0}`$ and the ratio of energy spread is defined as $`\delta =\sigma _e/E_f`$. If the parameter $`\xi `$ or $`x_0`$ is larger, the energy spread after Compton scattering is increasing and it is the origin of the emittance growth in the defocusing optics, reacceleration linac, and focusing optics. The energy spreads $`\delta `$ for the electron energies 2 GeV and 5 GeV are 9.8% and 19%, respectively. The equilibrium emittances due to Compton scattering are $`^{\text{“sevenrm}\text{?}}`$ $$ϵ_{ni,\mathrm{min}}=\frac{7.2\times 10^{10}\beta _i[\mathrm{mm}]}{\lambda _L[\mu \mathrm{m}]}(i=x,y)[\mathrm{m}\mathrm{rad}],$$ (6) where $`\beta _i`$ are the beta functions at IP. From this formula we can see that small beta gives small emittance. However the large change of the beta functions between the magnet and the IP causes the emittance growth. Taking no account of the emittance growth, for the electron energies 2 GeV and 5 GeV, the equilibrium emittances are $`5.8\times 10^9`$ m$``$rad and $`1.4\times 10^{10}`$ m$``$rad, respectively. The equilibrium emittances depended on $`\xi `$ in the case $`\xi ^21`$ were calculated in Ref. ?. ### 2.2 Simulation of Laser-Electron Interaction For the simulation of laser-electron interaction, the electron beam is simply assumed to be a round beam in the case of $`E_0=5`$ GeV and $`C=5`$. Taking no account of the emittance growth of optics, the one stage for cooling consists two parts as follows: 1. The laser-Compton interaction between the electron and laser beams. 2. The reacceleration of electrons in linac. In the first part, we simulated the interactions by the CAIN code .$`^{\text{“sevenrm}\text{?}}`$ This simulation calculates the effects of the nonlinear Compton scattering between the laser photons and the electrons. We assume that the laser pulse interacts with the electron bunch in head-on collisions. The $`\beta _x`$ and $`\beta _y`$ at the IP are fixed to be 0.1 mm. The initial energy spread of the electron beams is 1%. The energy of laser pulse is 20 J. The difference of the pulse energy between the simulation and the formula depends on the transverse sizes of the electron beams at IP. The polarization of the electron and laser beams are $`P_e`$=1.0 and $`P_L`$=1.0 (circular polarization), respectively. When the $`x_0`$ and $`\xi `$ parameters are small, the spectrum of the scattered photons does not largely depend on the polarization combination. In order to accelerate the electron beams to 5 GeV for recovery of energy in the second part, we simply added the energy $`\mathrm{\Delta }E=5\text{GeV}E_{ave}`$ for reacceleration, where $`E_{ave}`$ is the average energy of the scattered electron beams after the laser-Compton interaction. Here we define the transverse, longitudinal, and 6D emittances in the simulation. The $`x,y`$-transverse emittance is $$ϵ_{n,i}=\sqrt{\sigma _i^2\sigma _i^{}^2(ii^{}ii^{})^2}(i=x,y),$$ (7) where the symbol $``$ means to take an average of all particles in a bunch. The longitudinal emittance is $$ϵ_{n,l}=\sqrt{\sigma _z^2\sigma _{p_z}^2(zp_zzp_z)^2}/(m_ec).$$ (8) The 6D emittance is defined as $$ϵ_{6N}=ϵ_{n,x}ϵ_{n,y}ϵ_{n,l}.$$ (9) Figure 2.2 shows the longitudinal distribution of the electrons after the first laser-Compton scattering. The average energy of the electron beams is 1.0 GeV and the energy spread $`\delta `$ is 0.19. The longitudinal distribution seems to be a boomerang. If we assume a short Rayleigh length of laser pulse, the energy loss of head and tail of beams is small. The number of the scattered photons per incoming particle and the average of the photon energy at the first stage are 40 and 96 MeV (rms energy 140 MeV), respectively. The transverse sizes of the electron beams in the multi-stage cooling are shown in Fig. 2.2. During collisions with the laser photons, the transverse distribution of the electrons remains almost unchanged. But they decrease when we focus them for the next laser-Compton interaction due to the lower normalized emittance and the fixed $`\beta `$-function at IP ($`\sigma _i=\sqrt{\beta _iϵ_{n,i}/\gamma }`$). The angles of the electron beams in the multi-stage cooling are shown in Fig. 2.2. As a result of reacceleration, the angles of the electrons decrease. They increase when we focus them for the next laser-Compton interaction. Finally the angles attain the average of Compton scattering angle and the effect of cooling saturates. Figure 2.2 shows the transverse emittances of the electron beams in the multi-stage cooling. From Eq.(6), $`ϵ_{ni,\mathrm{min}}=1.4\times 10^{10}`$ m$``$rad, and the simulation presents $`ϵ_{ni,\mathrm{min}}=1.2\times 10^9`$ m$``$rad. Figure 2.2 shows the longitudinal emittance of the electron beams in the multi-stage cooling. Due to the increase of the energy spread of the electron beams from 1% to 19%, the longitudinal emittance rapidly increases at the first stage. After the first stage, the normalized longitudinal emittance is stable. The 6D emittance of the electron beams in the multi-stage cooling is shown in Fig. 2.2. The second cooling stage has the largest reduction for cooling. The 8th or 9th cooling stages have small reduction for cooling. The initial and final 6D emittances $`ϵ_{6N}`$ are $`1.5\times 10^{13}`$ (m$``$rad)<sup>3</sup> and $`1.2\times 10^{19}`$ (m$``$rad)<sup>3</sup>, respectively. Figure 2.2 shows the polarization of the electron beams in the multi-stage cooling. The decrease of the polarization during the first stage is about 0.06. The final polarization $`P_e`$ after the multi-stage cooling is 0.54. ## 3 Optics Design for Laser-Compton Cooling ### 3.1 Optics without chromaticity correction There are three optical devices for the laser-Compton cooling of electron beams as follows: 1. The focus optics to the first IP. 2. The defocus optics from the first IP to the reacceleration linac. 3. The focus optics from the linac to the next IP. Figure 3.1 shows schematics of the laser-Compton cooling of electron beams. The optics 1 is focusing the electron beams from a few meters of $`\beta `$-function to several millimeters in order to effectively interact them with the laser beams. The optics 2 is defocusing them from several millimeters to a few meters for reacceleration of electron beams in linac. In a multi-stage cooling system, the optics 3 is needed for cooling in the next stage. The problem for the focus and defocus optical devices is the energy spread of electrons and the electron beams with a large energy spread are necessary to minimize or correct the chromatic aberrations avoiding emittance growth. In this subsection, we discuss the optics for laser-Compton cooling without chromatic corrections. For the focus and defocus of the beams, we use the final doublet system which is similar to that of the final focus system of the future linear colliders .$`^{\text{“sevenrm}\text{?}}`$ The pole-tip field of the final quadrupole $`B_T`$ is limited to 1.2 T and the pole-tip radius $`a`$ is greater than 3 mm. The strength of the final quadrupole is $`\kappa =B_T/(aB\rho )120/E[\mathrm{GeV}][\mathrm{m}^2],`$ (10) where $`B`$, $`\rho `$, and $`E`$ are the magnetic field, the radius of curvature and the energy of the electron beams. In our case, the electron energies in the optics 1, 2, and 3 are 5.0, 1.0, and 5.0 GeV, respectively and the limit of the strength of the quadrupole in laser cooling is much larger than that of the final quadrupole of the future linear colliders. Due to the low energy beams in laser cooling, the synchrotron radiation from quadrupoles and bends is negligible. The difference of three optical devices is the amount of the energy spread of the beams. In the optics 1,2, and 3, the beams have one, several tens, and a few % energy spread. In order to minimize the chromatic aberrations, we need to shorten the length between the final quadrupole and the IP. In this study, the length from the face of the final quadrupole to the IP, $`l`$ is assumed to be 2 cm. Here we estimated the emittance growth in the optics 2, because the chromatic effect in the optics 2 is the most serious. Figure 3.1 shows the defocus optics without chromaticity correction for laser-Compton cooling by the MAD code .$`^{\text{“sevenrm}\text{?}}`$ The input file is attached to Ref. ?. The parameters of the electron beam for laser-Compton cooling at $`E_0=5`$ GeV and $`C=5`$ are listed in Table 3.1. The initial $`\beta _x`$ and $`\beta _y`$ after laser-Compton interaction are 20 mm and 4 mm, respectively. The final $`\beta _x`$ and $`\beta _y`$ are assumed to be 2 m and 1 m, respectively. The initial and final $`\alpha _{x(y)}`$ with no energy spread $`\delta =0`$ are 0 in this optics. The strength $`\kappa `$ of the final quadrupole for the beam energy of 1 GeV from Eq. (10) is assumed to be 120 $`\mathrm{m}^2`$. In our case, the chromatic functions $`\xi _x`$ and $`\xi _y`$ are 18 and 148, respectively. The momentum dependence of the emittances in the defocus optics without chromaticity correction is shown in Fig. 3.1. In the paper ,$`^{\text{“sevenrm}\text{?}}`$ the analytical study by thin-lens approximation has been studied for the focusing system, and here the transverse emittances are calculated by a particle-tracking simulation. The 10000 particles are tracked for the transverse and longitudinal Gaussian distribution by the MAD code. The relative energy spread $`\delta `$ is changed from 0 to 0.4. Due to the larger chromaticity $`\xi _y`$, the emittance $`ϵ_y`$ is rapidly increasing with the energy spread $`\delta `$. If we set a limit of 200% for $`\mathrm{\Delta }ϵ_i/ϵ_i(i=x,y)`$, the permissible energy spread $`\delta _x`$ and $`\delta _y`$ are 0.11 and 0.012 which mean the momentum band widths $`\pm 22\%`$ and $`\pm 2.4\%`$, respectively. The results are not sufficient for cooling at $`E_0=5`$ GeV and $`C=5`$, because the beams through the defocusing optics have the energy spread of several tens %. On the one hand, the optics can be useful as the optics 1 and 3 with the energy spread of a few %. ### 3.2 Optics with chromaticity correction The optics without chromaticity correction for the optics 2 does not work as we saw in the previous subsection. In this subsection we apply the chromaticity correction for the optics 2. The lattice for cooling is designed referring to the final focus system of the future linear colliders by K. Oide .$`^{\text{“sevenrm}\text{?}}`$ The final doublet system is the same lattice as the optics before subsection. The method of chromaticity correction uses one family of sextupole to correct for vertical chromaticity and moreover we added two weak sextupoles in the lattice to correct for horizontal chromaticity. Figure 3.2 shows the defocus optics with chromaticity correction for the laser-Compton cooling. The input file is attached to Ref. ?. The total length of the lattice is about 63 m. The momentum dependence of the emittances in the defocus optics with chromaticity correction is shown in Fig. 3.2. The 10000 particles are tracked for the transverse and longitudinal Gaussian distribution by the MAD code. The relative energy spread $`\delta `$ is changed from 0 to 0.06 with the conservation $`\kappa _2\theta _B`$, where $`\kappa _2`$ and $`\theta _B`$ are the strength of the sextupole and the angle of the bending magnet. The initial $`\beta _x`$ and $`\beta _y`$ after laser-Compton interaction are 20 mm and 4 mm, respectively. The final $`\beta _x`$ and $`\beta _y`$ are assumed to be 2 m and 1 m, respectively. The initial and final $`\alpha _x(y)`$ with no energy spread $`\delta =0`$ are 0 in this optics. After the chromaticity correction, the chromaticity functions $`\xi _x`$ and $`\xi _y`$ are 9.3 and 1.6, respectively. If we set a limit of 200% for $`\mathrm{\Delta }ϵ_i/ϵ_i(i=x,y)`$, the permissible energy spread $`\delta _x`$ and $`\delta _y`$ are 0.040 and 0.023 which mean the momentum band widths $`\pm 8\%`$ and $`\pm 4.6\%`$, respectively. By the comparison with the results of the optics without chromaticity correction at a limit of $`200\%`$ for $`\mathrm{\Delta }ϵ_i/ϵ_i(i=x,y)`$, the $`ϵ_y`$ of the optics with chromaticity correction is about two times larger than that of the one before subsection, but the $`ϵ_x`$ of the optics with chromaticity correction is three times smaller than that of the one before. The results are still not sufficient for cooling with $`E_0=5`$ GeV and $`C=5`$. These results emphasize the need to pursue further ideas for plasma lens .$`^{\text{“sevenrm}\text{?}}`$ ## 4 Laser-Compton Cooling for JLC/NLC at $`𝑬_\mathrm{𝟎}\mathbf{=}\mathrm{𝟐}`$ GeV ### 4.1 Optics For the future linear colliders, the method of laser-Compton cooling is effective to reduce the transverse emittances after damping rings. Where can it be placed? There are two possibilities for JLC/NLC $`^{\text{“sevenrm}\text{?}}`$ as follows: 1. After the first bunch compressor (BC1) and before the pre-linac. $`E_0=2`$ GeV and $`\sigma _z=0.5`$ mm. 2. After the second bunch compressor (BC2) and before the main linac. $`E_0=10`$ GeV and $`\sigma _z=0.1`$ mm. Case 2 needs a large energy for recovery after Compton scattering and we consider case 1 in this study. The parameters of the electron and laser beams for laser-Compton cooling for JLC/NLC at $`E_0=2`$ GeV and $`C=10`$ are listed in Tables 1 and 1. The energy of laser pulse is 300 J. The simulation results of the laser-electron interaction by the CAIN code are summarized as follows. The energy spread of the electron beam is 11%. The decrease of the longitudinal polarization of the electron beam is 0.038 ($`P_e=1.0,P_L=1.0`$). The number of the scattered photons per incoming particle and the average of the photon energy are 200 and 8.9 MeV (rms energy 19 MeV), respectively. The electron energy after Compton scattering in case 2 is 0.2 GeV and the strength of the final quadrupole from Eq. (10) is 600 m<sup>-2</sup>. Table 4.1 lists the parameters of the defocusing optics for laser-Compton cooling for JLC/NLC at $`E_0=2`$ GeV and $`C=10`$. The final $`\beta _x`$ and $`\beta _y`$ are assumed to be 1 m and 0.25 m, respectively. The chromaticity functions $`\xi _x`$ and $`\xi _y`$ are 18 and 23, respectively. Using the MAD code, the emittance growth in the defocus optics are $$\mathrm{\Delta }ϵ_{n,x}^{\mathrm{defocus}}=ϵ_{n,x}ϵ_{n,x0}1.0ϵ_{n,x0}7.6\times 10^8[\mathrm{m}\mathrm{rad}],$$ (11) $$\mathrm{\Delta }ϵ_{n,y}^{\mathrm{defocus}}=ϵ_{n,y}ϵ_{n,y0}1.6ϵ_{n,y0}4.6\times 10^8[\mathrm{m}\mathrm{rad}],$$ (12) where the normalized emittances before and after the defocus optics are $`ϵ_{n,i0}`$ and $`ϵ_{n,i}`$ ($`i=x,y`$), respectively. The emittance growth in the other two focus optics are negligible. ### 4.2 Reacceleration Linac In the reacceleration linac, there are two major sources of the emittance increase $`^{\text{“sevenrm}\text{?}}`$ as follows: 1. The emittance growth due to the misalignment of the quadrupole magnet and the energy spread. 2. The emittance growth due to the cavity misalignment. The emittance growth due to these sources in the reacceleration linac (L-band linac) are formulated by K. Yokoya $`^{\text{“sevenrm}\text{?}}`$ $$\mathrm{\Delta }ϵ_{n,x}^{\mathrm{linac}}3.4\times 10^9[\mathrm{m}\mathrm{rad}]0.045ϵ_{n,x0},$$ (13) $$\mathrm{\Delta }ϵ_{n,y}^{\mathrm{linac}}3.4\times 10^9[\mathrm{m}\mathrm{rad}]0.12ϵ_{n,y0}.$$ (14) The final emittance growth and the final emittance with $`C=10`$ are $$\mathrm{\Delta }ϵ_{n,x}7.9\times 10^8[\mathrm{m}\mathrm{rad}]1.1ϵ_{n,x0}ϵ_{n,x}0.21ϵ_{n,x0},$$ (15) $$\mathrm{\Delta }ϵ_{n,y}4.9\times 10^8[\mathrm{m}\mathrm{rad}]1.7ϵ_{n,y0}ϵ_{n,y}0.27ϵ_{n,y0}.$$ (16) The total reduction factor of the 4D transverse emittance of the laser-Compton cooling for JLC/NLC at $`E_0=2`$ GeV is about 18. The decrease of the polarization of the electron beam is 0.038 due to the laser-Compton interaction. ## 5 Summary We have studied the method of laser-Compton cooling of electron beams. The effects of the laser-Compton interaction for cooling have been evaluated by the Monte Carlo simulation. From the simulation in the multi-stage cooling, we presented that the low emittance beams with $`ϵ_{6N}=1.2\times 10^{19}`$(m$``$rad)<sup>3</sup> can be achieved in our beam parameters. We also examined the optics with and without chromatic correction for cooling, but the optics are not sufficient for cooling due to the large energy spread of the electron beams. The laser-Compton cooling for JLC/NLC at $`E_0=2`$ GeV and $`C=10`$ was considered. The total reduction factor of the 4D transverse emittance of the laser-Compton cooling is about 18. The decrease of the polarization of the electron beam is 0.038 due to the laser-Compton interaction. Acknowledgments We would like to thank Y. Nosochkov, K. Oide, T. Takahashi, V. Telnov, M. Xie, and K. Yokoya for useful comments and discussions. References
warning/0002/hep-th0002231.html
ar5iv
text
# 1 Introduction ## 1 Introduction The Exact renormalization group (ERG) has been one of the analytical tools to investigate non-perturbative phenomena of field theories, (e.g. the chiral symmetry breaking etc.). The ERG flow equations are the functional differential equations for the Wilsonian effective actions $`S_\mathrm{\Lambda }[\varphi ]`$, where $`\mathrm{\Lambda }`$ is an ultra-violet (infra-red) momentum cutoff of the low energy modes $`\varphi (𝐪)`$ (the high energy modes which are already integrated out). The responce of the effective action under variation of the cutoff is exactly represented as $$\frac{}{\mathrm{\Lambda }}S_\mathrm{\Lambda }[\varphi ]=F[S_\mathrm{\Lambda }],$$ (1) where $`F[S_\mathrm{\Lambda }]`$ is a finite functional of the field $`\varphi `$. The explicit forms of $`F[S_\mathrm{\Lambda }]`$ are shown later. By solving the ERG flow equations toward to $`\mathrm{\Lambda }=0`$ with certain bare actions as the initial conditions, we can obtain the generating functionals of the connected Green’s functions. There have been also known another type of the ERG equations for the cutoff Legendre effective action (or the effective average action) $`\mathrm{\Gamma }_\mathrm{\Lambda }[\varphi ]`$. In this case the solutions of the ERG lead to the ordinary effective actions, or the generating functionals of the 1PI Green’s functions. Usually, we write down the ERG equations for the dimensionless parameters in the effective actions by scaling the parameters with the infra-red cutoff $`\mathrm{\Lambda }`$, since the energy unit used to represent the theories does not have any physical significance. The functional space of the dimensionless effective action is called the theory space. Through this manipulation the beta-functional $`F`$ becomes free from the scale $`\mathrm{\Lambda }`$. Among the RG flows of these dimensionless quantities, especially the fixed points, the critical surfacies and the renormalized trajectories, are of utility indispensable in investigating (statistical) continuum limit of field theories. It is important for the practical analyses that the ERG admits non-perturbative as well as systematic approximations; e.g. the derivative expansion , the momentum scale expansion . Though we used the word of ‘expansion’ here, the approximation schemes are not the series expansions with respect to some explicit small parameters. This is an essential distinction from the ordinary expansion schemes; $`ϵ`$-expansion, $`1/N`$-expansion and perturbation, which lead to the asymptotic series. The solutions of the ERG equations are expected to converge smoothly with the improvement of approximatioins. Furthermore, we may obtain fairly good non-perturbative results within the simple approximation schemes . The ERG flow equation depends on the cutoff schemes. Here the cutoff scheme means the profile of the cutoff function in the propagator. It is convenient to perform the cutoff of the infra-red region $`p^2<\mathrm{\Lambda }^2`$ by adding a momentum dependent mass, $$\mathrm{\Delta }S[\varphi ]d^dx\frac{\mathrm{\Lambda }^2}{2}\varphi C^1(^2/\mathrm{\Lambda }^2)\varphi ,$$ (2) where $`C`$ is a proper cutoff function satisfiying that $`C(x)0`$ as $`x0`$ and $`d`$ is the space-time dimensions. Then the propagator is multipied by the cutoff function $`\theta (x)=xC(x)/(1+xC(x))`$. In Fig. 1 the examples of the cutoff functions for various $`C(x)`$ are shown. The sharp cutoff scheme corresponds to the step function; $`\theta (p^2)=0`$ for $`p<\mathrm{\Lambda }`$ and $`\theta (p^2)=1`$ for $`p>\mathrm{\Lambda }`$. The effective actions treated by the ERG equations themselves are cutoff scheme dependent even after the cutoff is removed; $`\mathrm{\Lambda }0`$. While the physical quantities obtained from their continuum limit, or the renormalized trjectories, should not be affected by the regularization scheme. Therefore it will be important to see the cutoff scheme dependence of the RG flows, specially the renormalizaed trajectories, and to find out the scheme independent quantities at $`\mathrm{\Lambda }=0`$ which are the physical quantities obtained in the ERG approach. In this paper we discuss the basic strucure of the cutoff scheme dependence of the ERG equations and of thier solutions; the effective actions. Especially we look into the scheme dependence of the low energy effective actions, or the renormalized trajectories, in the asymptotic regions for massive theories. In such regions the Wilsonian effectives actions are found to suffer from strong scheme dependence, while the Legendre effective actions are free from such problems. Related with this we also examine the sharp cutoff limit of the so-called Polchinksi equation by taking care of the singular limit of the cutoff profiles. Moreover it will be shown that the Polchinksi equations turn out to be equivalent to the Wegner-Houghton equation in the sharp cutoff limit. This equivalence between these two formulations of the ERG has not been proved yet. This paper is organized as follows. In Sec. 2 we briefly overview the derivation of the ERG flow equations; the Wegner-Houghton equation, the Polchinski equation and the flow equation for the cutoff Legendre effective action. In Sec. 3 we will examine the general aspects of the scheme dependence of the effective actions. At first it will be shown that the variation of the cutoff function; $`\theta (x)\theta (x)+\delta \theta (x)`$, can be reinterpreted as the coordinate transformation on the theory space. From this observation it immediately follows that the critical exponents obtained by the ERG method are independent of the cutoff scheme. We also discuss the formulations using the cutoff mass functions depending on the wave function renormalization from this point of view. We will study also the cutoff scheme dependence of the renormalized trajectories in the infra-red asymptotic region of massive theories. The scheme dependence of the coefficient function $`V_k^i(\varphi )`$ of the Wilsonian effective action $`S_\mathrm{\Lambda }[\varphi ]`$ behaves as $`1/\mathrm{\Lambda }^{2k}`$ and, therefore, becomes so strong as to prevent from taking any physical information as the infra-red cutoff $`\mathrm{\Lambda }`$ is lowered. While the cutoff Legendre effective actions are free from such a strong scheme dependence. The sharp cutoff limit of the Polchinski equations and their equivalence to the Wegner-Houghton formulation will be clarified in Sec. 4 and in Sec. 5. Section 6 is devoted to the conclusions and some remarks. Throughout this paper, we restrict ourselves to the single scalar theories. This restriction does not loose the generality of the discussions. ## 2 ERG Equations To fix our conventions and to make this paper self-contained, we briefly overview the derivations of the ERG flow equations. Let us start from the generator of the connected Green’s functions $`W[J]`$ given by $$\mathrm{exp}\left(W[J]\right)=D\varphi \mathrm{exp}\left(\mathrm{\Delta }S_{\text{U.V.}}S_{\mathrm{\Lambda }_0}+J\varphi \right),$$ (3) where $`\mathrm{\Delta }S_{\text{U.V.}}`$ is the ultra-violet cutoff term regularizing the path-integral (3). We sometimes use the shorthand: $`J\varphi =d^dxJ(x)\varphi (x)`$ etc., where $`d`$ is the (euclidean) space-time dimensions. Now, we introduce the intermediate scale $`\mathrm{\Lambda }<\mathrm{\Lambda }_0`$ and formally integrate out the high energy modes $`\varphi _>(𝐪)`$ ($`\mathrm{\Lambda }<q\mathrm{\Lambda }_0`$). Then we get the effective action for the low energy modes $`\varphi (𝐪)`$ ($`q<\mathrm{\Lambda }`$), $$\mathrm{exp}\left(S_\mathrm{\Lambda }[\varphi ,J]\right)=D\varphi _>\mathrm{exp}\left(\mathrm{\Delta }S_\mathrm{\Lambda }^{\mathrm{\Lambda }_0}[\varphi _>]S_{\mathrm{\Lambda }_0}[\varphi +\varphi _>]+J(\varphi +\varphi _>)\right),$$ (4) where the cutoff action $`\mathrm{\Delta }S_\mathrm{\Lambda }^{\mathrm{\Lambda }_0}[\varphi _>]`$ is given by, $$\mathrm{\Delta }S_\mathrm{\Lambda }^{\mathrm{\Lambda }_0}[\varphi ]\frac{1}{2}\varphi \left[P_\mathrm{\Lambda }^{\mathrm{\Lambda }_0}\right]^1\varphi .$$ (5) The support of $`P_\mathrm{\Lambda }^{\mathrm{\Lambda }_0}(q)`$ is effectively restricted in the region $`\mathrm{\Lambda }<q<\mathrm{\Lambda }_0`$ by means of a certain smooth cutoff function. Furthermore, we set, $`\mathrm{\Delta }S_{\text{U.V.}}=\mathrm{\Delta }S_0^{\mathrm{\Lambda }_0}`$ and $`P_0^{\mathrm{\Lambda }_0}(q)=P_0^\mathrm{\Lambda }(q)+P_\mathrm{\Lambda }^{\mathrm{\Lambda }_0}(q)`$. This can be achieved by multiplying the partition of unity $`\theta _\mathrm{\Lambda }^{\mathrm{\Lambda }_0}(q)`$ to the propagator $`1/q^2`$, i.e. $`P_\mathrm{\Lambda }^{\mathrm{\Lambda }_0}(q)=\theta _\mathrm{\Lambda }^{\mathrm{\Lambda }_0}(q)/q^2`$. $`\theta _\mathrm{\Lambda }^{\mathrm{\Lambda }_0}(q)`$ approximately vanishes when $`q\mathrm{\Lambda }_0`$ or $`q\mathrm{\Lambda }`$, and $`\theta _\mathrm{\Lambda }^{\mathrm{\Lambda }_0}(q)1`$ for $`\mathrm{\Lambda }<q<\mathrm{\Lambda }_0`$. In the RG flow equation, we can safely forget the ultra-violet cutoff $`\mathrm{\Lambda }_0`$ by taking the limit $`\mathrm{\Lambda }_0\mathrm{}`$, since $`P_\mathrm{\Lambda }^{\mathrm{\Lambda }_0=\mathrm{}}(q)/\mathrm{\Lambda }`$ decays in both $`q0`$ and $`q\mathrm{}`$ sufficiently fast. Now, Eq. (3) can be rewritten in terms of $`S_\mathrm{\Lambda }`$ in Eq. (4), (See ref.) $$\mathrm{exp}\left(W[J]\right)=D\varphi _<\mathrm{exp}\left(\mathrm{\Delta }S_0^\mathrm{\Lambda }S_\mathrm{\Lambda }[\varphi ,J]\right).$$ (6) Putting $`J=0`$, Eq. (6) is nothing but the definition of the Wilsonian effective action $`\mathrm{\Delta }S_0^\mathrm{\Lambda }+S_\mathrm{\Lambda }[\varphi ,0]`$. Evidently, if we put $`\varphi _<=0`$ in Eq. (4) then $`S_\mathrm{\Lambda }`$ becomes the generator of the connected Green’s functions with the infra-red cutoff $`W_\mathrm{\Lambda }[J]=S_\mathrm{\Lambda }[0,J]`$ . By shifting $`\varphi _>\varphi _>\varphi _<`$ and setting $`J=0`$ in Eq. (4), we also find , $$W_\mathrm{\Lambda }[P_\mathrm{\Lambda }^1\varphi _<]=\frac{1}{2}\varphi _<P_\mathrm{\Lambda }^1\varphi _<S_\mathrm{\Lambda }[\varphi _<,0].$$ (7) Hereafter, we write $`S_\mathrm{\Lambda }[\varphi ]=S_\mathrm{\Lambda }[\varphi ,0]`$. The above cutoff $`\theta `$ is called the ‘multiplicative cutoff’, because we multiplied it to the kinetic term of $`\varphi `$. $`\mathrm{\Delta }S`$ is called ‘additive’, since the inverse cutoff propagator is given by $`C^1(q^2)`$ in Eq. (2) plus the ordinary kinetic term $`q^2`$. $`\theta (x)\theta _\mathrm{\Lambda }^{\mathrm{\Lambda }_0}(x)`$ above is written in terms of $`C(x)`$ as $`\theta (x)=xC(x)/(1+xC(x))`$. The relation between these two cutoff schemes: the multiplicative cutoff and the additive cutoff is given as follows. The bare actions of both schemes obviously satisfy $`S_{\mathrm{\Lambda }_0}^{\text{add}}=\frac{1}{2}d^dx(\varphi )^2+S_{\mathrm{\Lambda }_0}^{\text{multi}}`$, where $`S_{\mathrm{\Lambda }_0}^{\text{add}}`$ and $`S_{\mathrm{\Lambda }_0}^{\text{multi}}`$ are the bare actions with the additive cutoff and with the multiplicative cutoff respectively. By the definition, the generator of the connected Green’s functions $`W_\mathrm{\Lambda }[J]`$ is common for the both schemes. Thus, it immediately follows, $$S_\mathrm{\Lambda }^{\text{multi}}[P_\mathrm{\Lambda }J]\frac{1}{2}JP_\mathrm{\Lambda }J=S_\mathrm{\Lambda }^{\text{add}}[CJ]\frac{1}{2}JCJ.$$ (8) We will employ the multiplicative cutoff scheme, since it is convenient to investigate the sharp cutoff limit of the RG flow equations. ### 2.1 Flow equations Setting $`\varphi _<=0`$ and differentiating the boths sides of Eq. (4) with respect to $`\mathrm{\Lambda }`$, we get the RG flow equation for $`W_\mathrm{\Lambda }[J]`$, $$\frac{}{\mathrm{\Lambda }}W_\mathrm{\Lambda }=\frac{1}{2}W_\mathrm{\Lambda }^{}\frac{}{\mathrm{\Lambda }}P_\mathrm{\Lambda }^1W_\mathrm{\Lambda }^{}\frac{1}{2}\mathrm{𝐭𝐫}\frac{}{\mathrm{\Lambda }}P_\mathrm{\Lambda }^1W_\mathrm{\Lambda }^{\prime \prime },$$ (9) where the prime denotes the derivative with respect to the source $`J`$. By using Eq. (7), we also get, $$\frac{}{\mathrm{\Lambda }}S_\mathrm{\Lambda }=\frac{1}{2}S_\mathrm{\Lambda }^{}\frac{}{\mathrm{\Lambda }}P_\mathrm{\Lambda }S_\mathrm{\Lambda }^{}+\frac{1}{2}\mathrm{𝐭𝐫}\frac{}{\mathrm{\Lambda }}P_\mathrm{\Lambda }S_\mathrm{\Lambda }^{\prime \prime }.$$ (10) This is the famous ‘Polchinski equation’ . This equation may be represented diagramatically as in Fig. 2. Next, we scale all the dimensionful quantities in terms of the infra-red cutoff $`\mathrm{\Lambda }`$, i.e. $`\varphi =\mathrm{\Lambda }^{d_\varphi }\widehat{\varphi }`$, $`𝐩=\mathrm{\Lambda }\widehat{𝐩}`$ and $`_\mathrm{\Lambda }=\mathrm{\Lambda }^d\widehat{}_t`$, where $`d_\varphi =(d2)/2`$ is the canonical dimension of the field, $`t=\mathrm{ln}\mathrm{\Lambda }_0/\mathrm{\Lambda }`$ is the cutoff scale factor and $`\widehat{}_t`$ is the Lagrangian density. We also write the dimensionless Wilsonian effective action as $`\widehat{S}_t[\widehat{\varphi }]=d^d\widehat{x}\widehat{}_t(\widehat{\varphi })`$. Then, we get $$\mathrm{\Lambda }\frac{}{\mathrm{\Lambda }}S_\mathrm{\Lambda }=\mathrm{\Lambda }^d\left(\frac{}{t}+d_\varphi \mathrm{\Delta }_\varphi +\mathrm{\Delta }_{}d\right)\widehat{S}_t,$$ (11) where $`\mathrm{\Delta }_\varphi `$ and $`\mathrm{\Delta }_{}`$ count the degree of the fields and that of the derivatives $`_\mu `$ respectively. The initial boundary condition of Eq. (10) is given by the bare action, $`S_{\mathrm{\Lambda }_0}`$. We can derive the RG flow equation for the Legendre effective action with the infra-red cutoff $`\mathrm{\Gamma }_\mathrm{\Lambda }[\varphi ]`$ given by the Legendre transform of $`W_\mathrm{\Lambda }[J]`$. $$W_\mathrm{\Lambda }[J]=J\varphi \mathrm{\Gamma }_\mathrm{\Lambda }[\varphi ]+\frac{1}{2}\varphi \left(P_\mathrm{\Lambda }^1P_{\mathrm{\Lambda }=0}^1\right)\varphi ,$$ (12) where $`\varphi `$ is given by $`\varphi =\delta W_\mathrm{\Lambda }/\delta J`$. After the Legendre transformation, the RG flow equation for $`\mathrm{\Gamma }_\mathrm{\Lambda }[\varphi ]`$ can be read, $$\frac{}{\mathrm{\Lambda }}\mathrm{\Gamma }_\mathrm{\Lambda }[\varphi ]=\frac{1}{2}\mathrm{𝐭𝐫}\frac{}{\mathrm{\Lambda }}P_\mathrm{\Lambda }^1\left(P_\mathrm{\Lambda }^1P_{\mathrm{\Lambda }=0}^1+\mathrm{\Gamma }^{\prime \prime }\right)^1.$$ (13) The initial condition of Eq. (13) is given by $`\mathrm{\Gamma }_{\mathrm{\Lambda }_0}=S_{\mathrm{\Lambda }_0}`$, because all quantum corrections varnish at $`\mathrm{\Lambda }=\mathrm{\Lambda }_0`$. Since $`\mathrm{\Gamma }_\mathrm{\Lambda }[\varphi ]`$ is composed of the one particle irreducible diagrams, the diagrams corresponding to Eq. (13) contain no tree diagrams, as is shown in Fig. 3. By the definition, $`\mathrm{\Gamma }_\mathrm{\Lambda }[\varphi ]`$ coincides with the ordinary Legendre effective action at $`\mathrm{\Lambda }=0`$, i.e. $`\mathrm{\Gamma }_{\mathrm{\Lambda }=0}[\varphi ]=\mathrm{\Gamma }[\varphi ]`$. One can find the RG flow equation for the dimensionless effective action $`\widehat{\mathrm{\Gamma }}_t[\widehat{\varphi }]`$ by the same manner as we did for the Polchinski one. ### 2.2 Wegner-Houghton equation We start from the following partition function, $$Z=_{p\mathrm{\Lambda }}D\varphi \mathrm{exp}\left(\frac{1}{2}\varphi P^1\varphi S_\mathrm{\Lambda }[\varphi ]\right),$$ (14) where the support of $`\varphi `$ is restricted to $`p\mathrm{\Lambda }`$. We shift the quadratic part, the $`P^1`$ term, from the Wilsonian effective action, since it is subtracted also in $`S_\mathrm{\Lambda }`$ given by Eq. (4). Let us integrate out the modes with momenta $`\mathrm{\Lambda }\delta \mathrm{\Lambda }<p\mathrm{\Lambda }`$, we call these modes the ‘shell modes’ and write as $`\varphi _\text{s}`$. Expanding the action $`S_\mathrm{\Lambda }[\varphi +\varphi _\text{s}]`$ in the shell modes $`\varphi _\text{s}`$, we have $$S_\mathrm{\Lambda }[\varphi +\varphi _\text{s}]=S_\mathrm{\Lambda }[\varphi ]+\varphi _\text{s}S_\mathrm{\Lambda }^{(1)}[\varphi ]+\frac{1}{2}\varphi _\text{s}^2S_\mathrm{\Lambda }^{(2)}[\varphi ]+\mathrm{},$$ (15) where the superscript $`(n)`$ denotes the $`n`$-th functional derivative with respect to the shell mode. We can regard the Taylor coefficients $`S_\mathrm{\Lambda }^{(n)}[\varphi ]`$ to the field ($`\varphi `$) dependent vertices. The quantum fluctuations of the shell modes can be incorporated by perturbative expansion. For infinitesimal $`\delta \mathrm{\Lambda }`$, the leading corrections, i.e. of order $`\delta \mathrm{\Lambda }`$, come from less than or equal to one loop diagrams. The higher ($`n2`$) loops diagrams do not contribute to the leading order in $`\delta \mathrm{\Lambda }`$, since every loop integral introduces the factor $`\delta \mathrm{\Lambda }`$. Furthermore, each articulation line also brings the factor $`\delta \mathrm{\Lambda }`$. Hence, the leading corrections are found to be the Feynman diagrams with only one propagator which is either an articulation line or a loop one. Taking into account this constraint, the higher vertices $`S_\mathrm{\Lambda }^{(n3)}[\varphi ]`$ cannot appear and can be dropped in Eq. (15). After performing the Gaussian integration of the shell modes, we get a coarse-grained action $`S_{\mathrm{\Lambda }\delta \mathrm{\Lambda }}[\varphi ]`$ of the low energy modes $`\varphi (𝐩)`$ : $`p\mathrm{\Lambda }\delta \mathrm{\Lambda }`$, $$S_{\mathrm{\Lambda }\delta \mathrm{\Lambda }}[\varphi ]=S_\mathrm{\Lambda }[\varphi ]\frac{1}{2}\delta \mathrm{\Lambda }S_\mathrm{\Lambda }^{}\left(P^1+S_\mathrm{\Lambda }^{\prime \prime }\right)^1S_\mathrm{\Lambda }^{}+\frac{1}{2}\delta \mathrm{\Lambda }\mathrm{𝐭𝐫}\mathrm{ln}\left(P^1+S_\mathrm{\Lambda }^{\prime \prime }\right),$$ (16) where prime denotes the functional derivative with respect to $`\varphi _\text{s}`$. By letting $`\delta \mathrm{\Lambda }0`$, we find $$\frac{}{\mathrm{\Lambda }}S_\mathrm{\Lambda }=\frac{1}{2}S_\mathrm{\Lambda }^{}\left(P^1+S_\mathrm{\Lambda }^{\prime \prime }\right)^1S_\mathrm{\Lambda }^{}\frac{1}{2}\mathrm{𝐭𝐫}\mathrm{ln}\left(P^1+S_\mathrm{\Lambda }^{\prime \prime }\right).$$ (17) This is called the ‘Wegner-Houghton equation’ . The diagrams corresponding to Eq. (17) are shown in Fig. 4. ## 3 General Aspects of the Cutoff Scheme Dependence In this section, we discuss some general aspects on the cutoff scheme dependence of the ERG. In the first two subsections we show that the formal relation among the Wilsonian effective actions with different cutoff schemes can be egarded as the coordinate transformation on the theory space. Therefore it immediately follows that the critical exponents, which are the macroscopic physical quantities of the phase transition, do not suffer from the cutoff scheme dependence. In the remaining subsections the scheme dependence of the renormalized trajectories in the infra-red asymptotic region is examined. It is shown that the scheme dependence disappears form the cutoff Legendre effective action in this region, while not from the Wilsonian effectiveaction. ### 3.1 Coordinate Transformation on the Theory Space The Polchinski RG equation for a single scalar theory can be rewritten, $`\left({\displaystyle \frac{}{t}}+d_\varphi \varphi {\displaystyle \frac{\delta }{\delta \varphi }}+{\displaystyle \frac{d^dq}{\left(2\pi \right)^d}\varphi \left(𝐪\right)q_\mu \frac{}{q_\mu }\frac{\overline{\delta }}{\overline{\delta }\varphi \left(𝐪\right)}}\right)\mathrm{exp}\left(S_t\left[\varphi \right]\right)`$ $`={\displaystyle \frac{d^dq}{\left(2\pi \right)^d}\frac{\overline{\delta }}{\overline{\delta }\varphi \left(𝐪\right)}\left(\frac{}{q^2}\theta \left(q^2\right)\right)\frac{\overline{\delta }}{\overline{\delta }\varphi \left(𝐪\right)}\mathrm{exp}\left(S_t\left[\varphi \right]\right)},`$ (18) where, for the convenience, we write the Fourier transform of the functional derivative with respect to $`\varphi (x)`$ by $$\frac{\overline{\delta }}{\overline{\delta }\varphi \left(𝐪\right)}d^dxe^{i𝐪𝐱}\frac{\delta }{\delta \varphi \left(𝐱\right)}=(2\pi )^d\frac{\delta }{\delta \varphi \left(𝐪\right)}.$$ (19) In Eq. (18), $`\theta \left(q^2\right)`$ denotes the cutoff function, which is given in Fig. 1 for example. Note that, the momentum derivative operating to the effective action $`S_t\left[\varphi \right]`$ in the first line of Eq. (18) does not operate to the delta function $`\delta (\mathrm{\Sigma }𝐪_𝐢)`$ representing momentum conservation. Hence it operates to the effective action as $$\frac{d^dq}{\left(2\pi \right)^d}\varphi \left(𝐪\right)q_\mu \frac{}{q_\mu }\frac{\overline{\delta }}{\overline{\delta }\varphi \left(𝐪\right)}S_t\left[\varphi \right]=\left(\mathrm{\Delta }_{}d\right)S_t\left[\varphi \right],$$ where $`\mathrm{\Delta }_{}`$ counts the degree of derivatives. Let us consider the coordinate transformation of the theory space: $`S_t\left[\varphi \right]\stackrel{~}{S}_t\left[\varphi \right]`$, given by the following transformation: $$\mathrm{exp}\left(\stackrel{~}{S}_t\left[\varphi \right]\right)=\mathrm{exp}\left(\frac{1}{2}\delta D\right)\mathrm{exp}\left(S_t\left[\varphi \right]\right),$$ (20) where $`\delta D`$ is given by $$\delta D=\frac{d^dq}{\left(2\pi \right)^d}\frac{\overline{\delta }}{\overline{\delta }\varphi \left(𝐪\right)}\frac{1}{q^2}\delta \theta \left(q^2\right)\frac{\overline{\delta }}{\overline{\delta }\varphi \left(𝐪\right)}.$$ (21) Since $`\delta D`$ is independent of the cutoff scale $`t`$, this transformation is in fact a coordinate transformation on the theory space<sup>1</sup><sup>1</sup>1In this paper, we naively assume that the coordinate transformation given by Eq. (20) is well-defined. Since both infra-red and ultra-violet regions are regularized, the perturbative expansion of Eq. (20) is finite in all orders.. Therefore the critical exponents obtained by the RG technique are invariant under the transformation (20). (See Ref. .) Indeed, by this transformation the cutoff function $`\theta (q^2)`$ is changed to $`\theta (q^2)+\delta \theta (q^2)`$ in the Polchinski RG equation. Operating $`\mathrm{exp}(\delta D/2)`$ to both sides of Eq. (18) and using the commutation relation, $$[d_\varphi \varphi \frac{\delta }{\delta \varphi }+\frac{d^dq}{(2\pi )^d}\varphi \left(𝐪\right)q^\mu \frac{}{q^\mu }\frac{\overline{\delta }}{\overline{\delta }\varphi \left(𝐪\right)},\frac{\delta D}{2}]=\frac{d^dq}{(2\pi )^d}\frac{\overline{\delta }}{\overline{\delta }\varphi \left(𝐪\right)}\left(\frac{\delta \theta \left(q^2\right)}{q^2}\right)\frac{\overline{\delta }}{\overline{\delta }\varphi \left(𝐪\right)},$$ (22) one can realize that the effective action $`\stackrel{~}{S}_t[\varphi ]`$ just satisfies the Polchinski RGE with the cutoff scheme $`\theta (q^2)+\delta \theta (q^2)`$. Therefore, if $`S_t[\varphi ]`$ is a solution of Eq. (18) with cutoff scheme $`\theta (q^2)`$, then $`\stackrel{~}{S}_t[\varphi ]`$, defined by Eq. (20), is also a solution for another cutoff scheme $`\theta (q^2)+\delta \theta (q^2)`$. The transformation (20) maps the fixed point, the critical surface and the renormalized trajectories to those given in another scheme. (See Fig. 5.) However the critical exponents at the fixed points are scheme independent. ### 3.2 Wave-function Renormalization and Additive cutoff In order to extract the anomalous dimension it is more convenient to employ the additive cutoff instead of the multiplicative one for following two reasons. 1). In the multiplicative case the part of the kinetic term of the Wilsonian effective action is stolen by the inverse cutoff propagator $`q^2\left(\theta (q^2)\right)^1`$. 2). We can elminate the wave-function renormalization factor $`Z_\varphi `$ in RG equation by rescaling the cuotoff function $`C\left(q^2\right)`$ to $`Z_\varphi ^1C\left(q^2\right)`$ in the additive case, and the anomalous dimension $`(\eta )`$ of the field may be explicitly extracted. The additive cutoff is introduced by Eq. (2) and the Wilsonian effective actions with two cutoff schemes are related by the relation (8), i.e. $$S^{\text{multi}}[\varphi ]=S^{\text{add}}[(1^2C)\varphi ]+\frac{1}{2}d^dx\varphi ^2(1^2C)\varphi ,$$ (23) where $`S^{\text{multi}}[\varphi ]`$ is the effective action with the multiplicative cutoff. In the multiplicative cutoff case, we drop the part of the kinetic term of (the interaction part of) the effective action. It is rather convenient to include the kinetic term in the effective action completely in extracting the anomalous dimension. The additive cutoff $`C(x)`$ in Eq. (2) and the multiplicative cutoff $`\theta (x)`$ are related by $`\theta (x)=xC(x)/(1+xC(x))`$. Let us rescale the field $`\varphi `$ to $`\widehat{\varphi }=Z_\varphi ^{\frac{1}{2}}\varphi `$, where $`Z_\varphi `$ is the wave-function renormalization factor. If we also rescale the cutoff function as, $$C^1\left(q^2/\mathrm{\Lambda }^2\right)Z_\varphi C^1\left(q^2/\mathrm{\Lambda }^2\right),$$ (24) then the explicit $`Z_\varphi `$ dependence of the RG flow equation can be eliminated. The RG flow equation depends on $`Z_\varphi `$ only through the anomalous dimension $`\eta `$ defined by the consistency condition, i.e. the kinetic term should be unity at each scale. Consequently, $`\eta `$ becomes the function of the coupling constants. It means that the beta-function of $`Z_\varphi `$ is given by, $$\frac{}{t}Z_\varphi =\eta (g_i)Z_\varphi ,$$ (25) where $`\{g_i\}`$ is a coordinate system on the theory space. In this coordinate system, the RG beta-functions have the following structure: $$\mathrm{\Omega }_{ij}(g)=\frac{\beta _i}{g_j}(g)=0\mathrm{for}g_j=Z_\varphi ,$$ (26) because the beta-functions have no $`Z_\varphi `$ dependence except for $`\beta _Z`$. Such a parametrization is called the ‘perfect coordinate’ in Ref. . For the dimensionless Wilsonian effective action, the RG flow equation becomes, $`\left({\displaystyle \frac{}{t}}+d_\varphi \varphi {\displaystyle \frac{\delta }{\delta \varphi }}+{\displaystyle \frac{d^dq}{\left(2\pi \right)^d}\varphi \left(𝐪\right)q_\mu \frac{}{q_\mu }\frac{\overline{\delta }}{\overline{\delta }\varphi \left(𝐪\right)}}\right)\mathrm{exp}\left(S_t\left[\varphi \right]\right)`$ (27) $`={\displaystyle \frac{d^dq}{\left(2\pi \right)^d}\frac{\overline{\delta }}{\overline{\delta }\varphi \left(𝐪\right)}\left(q^2\frac{}{q^2}C+\frac{1}{2}(\eta 2)C\right)\frac{\overline{\delta }}{\overline{\delta }\varphi \left(𝐪\right)}\mathrm{exp}\left(S_t\left[\varphi \right]\right)},`$ where $`d_\varphi `$ and $`\eta =\dot{Z}_\varphi /Z_\varphi `$ are the physical scaling dimension of the field; $`d_\varphi =(d+\eta 2)/2`$ and the anomalous dimension of $`\varphi `$ respectively. One can easily extend the scheme dependence relations given by Eq. (22) etc. to this type of RGE. However one may wonder whether the coordinate transformation induced by Eq. (24) is well-defined or not. Suppose that $`C(x)`$ is a polynomial i.e. $`C(x)=x^k`$. Since the cutoff $`\mathrm{\Lambda }`$ appears only though the cutoff function, $$\mathrm{e}^{S_\mathrm{\Lambda }[\varphi ]}D\varphi _>\mathrm{exp}\left\{\frac{Z_\varphi }{2}\varphi _>\mathrm{\Lambda }^2C^1(^2/\mathrm{\Lambda }^2)\varphi _>S[\varphi _>+\varphi ]\right\},$$ (28) we can eliminate $`Z_\varphi `$ by shifting the cutoff; $`Z_\varphi \mathrm{\Lambda }^2C^1(\frac{p^2}{\mathrm{\Lambda }^2})=\mathrm{\Lambda }_{}^{}{}_{}{}^{2}C^1(\frac{p^2}{\mathrm{\Lambda }_{}^{}{}_{}{}^{2}})`$ where $`\mathrm{\Lambda }=Z_\varphi ^{1/(2k2)}\mathrm{\Lambda }^{}`$. Therefore the Wilsonian effective action $`\stackrel{~}{S}_\mathrm{\Lambda }[\varphi ]`$ with a cutoff scheme $`C^1`$ can be written in terms of $`S_\mathrm{\Lambda }[\varphi ]`$, $`\stackrel{~}{S}_\mathrm{\Lambda }^{}[\varphi ]=S_\mathrm{\Lambda }[\varphi ]=S_\mathrm{\Lambda }^{}[\varphi ]+{\displaystyle _\mathrm{\Lambda }^{}^\mathrm{\Lambda }}𝑑\overline{\mathrm{\Lambda }}{\displaystyle \frac{}{\overline{\mathrm{\Lambda }}}}S_{\overline{\mathrm{\Lambda }}}[\varphi ],`$ $`=S_\mathrm{\Lambda }^{}[\varphi ]+\delta f(Z_\varphi )\mathrm{\Lambda }^{}{\displaystyle \frac{}{\mathrm{\Lambda }^{}}}S_\mathrm{\Lambda }^{}[\varphi ]+{\displaystyle \frac{1}{2!}}(\delta f(Z_\varphi )\mathrm{\Lambda }^{})^2{\displaystyle \frac{^2}{\mathrm{\Lambda }_{}^{}{}_{}{}^{2}}}S_\mathrm{\Lambda }^{}[\varphi ]+\mathrm{},`$ (29) where $`\delta f(Z_\varphi )`$ is given by $`\mathrm{\Lambda }\mathrm{\Lambda }^{}=\delta f(Z_\varphi )\mathrm{\Lambda }^{}`$. In our case, $`\delta f(Z_\varphi )=Z_\varphi ^{1/(2k2)}1`$. The derivative with respect to $`\mathrm{\Lambda }^{}`$ in Eq. (29) will be eliminated by the RG flow equation. Thus one can find the coordinate transformation between $`\stackrel{~}{S}_\mathrm{\Lambda }[\varphi ]`$ and $`S_\mathrm{\Lambda }[\varphi ]`$. Since all the loop momentum integrals are regularized in the both regions of infra-red and ultra-violet, Eq. (29) gives the well-defined coordinate transformation at all orders in the Taylor expansion. In the more general case, we cannot eliminate $`Z_\varphi `$ by shifting the cutoff. However, since the change of $`\theta (x)=xC(x)/(1+xC(x))`$ induced by the change of the cutoff function $`C^1(x)`$ is concentrated in the finite region of the momentum $`x=p^2`$, it also gives the well-defined coordinate transformation equivalent to Eq. (20). Consequently, the critical exponents given by Eq. (18) and by Eq. (27) are completely the same, since these formulations can be understood as the difference of the coordinate systems on the theory space. ### 3.3 Asymptotic Region of the Renormalized Trajectory In this and the next subsections we discuss the cutoff scheme dependence of the renormalized trajectories in the ‘asymptotic region’ $`\mathrm{\Lambda }M_R`$, where $`M_R`$ is the renormalized mass of $`\varphi `$. We note that Eq. (20) may be rewritten as follows. Let $`S_t^{(n)}`$ be the vertices of the effective action $`S_t[\varphi ]`$, and $`\stackrel{~}{W}_t`$ be sum of the connected diagrams composed of the propagator $`(\delta P^1+S_t^{(2)})^1`$ and the vertices $`S_t^{(n)}`$ ( $`n>2`$ ), where $`\delta P(q^2)`$ is the cutoff propagator i.e. $`\delta P(q^2)\frac{1}{q^2}\delta \theta (q^2)`$. Then it is easily found that (See appendix), $$\mathrm{exp}\left(\frac{1}{2}\delta D[\delta /\delta \varphi ]\right)\mathrm{exp}\left(S_t\left[\varphi \right]\right)=\mathrm{exp}\left(\frac{1}{2}\varphi \delta P^1\varphi +\stackrel{~}{W}_t[\delta P^1\varphi ]\right).$$ (30) The cutoff scheme dependence of the renormalized trajectory is given by Eq. (30). (See Fig. 5.) If the RG flows of the dimensionful quantities freeze when $`t\mathrm{}`$, then all dimensionless coupling $`g_i(t)`$ should behave $`g_i(t)g_i^Re^{d_it}`$ as $`t\mathrm{}`$ with some finite dimensionful coupling $`g_i^R`$, where $`d_i`$ is the canonical dimension of $`g_i`$. Here, we take $`M_R`$ to be a unit of the mass scale: $`t=\mathrm{ln}M_R/\mathrm{\Lambda }`$. Hereafter we call such a region ‘the freezing region’, if it exists. As is aeen below the asymptotic region is not always the freezing region. In the asymptotic region, it can be realized that the loop diagrams in Eq. (30) are found to be suppressed compared with the tree diagrams. It is seen by the following arguments. Let us consider the Feynman diagram with $`N_I`$ internal lines, $`N_E`$ external legs and $`N_V`$ vertices. By comparing the canonical dimension of each operator in the both sides of Eq. (30), we find the following factor, $$\mathrm{exp}\{\mathrm{\Delta }t\}\mathrm{exp}\left\{\left(dN_Vd_\varphi (N_E+2N_I)N_D^{(1)}2N_I(dd_\varphi N_E)+N_D^{(2)}\right)t\right\},$$ (31) where $`N_D^{(1)}`$ is the total degree of the external momenta (or the derivatives) of the $`N_V`$ vertices in the Feynman diagram, and $`N_D^{(2)}`$ is that of the subset of $`N_D^{(1)}`$ derivatives which operate to the field $`\varphi (x)`$. The first three terms of $`\mathrm{\Delta }`$ come from the canonical dimensions of the vertices, the next one is from propagators and the last two terms are the dimension of $`N_E`$ point vertex with $`N_D^{(2)}`$ derivatives. $`N_D^{(1)}`$ and $`N_D^{(2)}`$ satisfy the relation $`N_D^{(1)}N_D^{(2)}`$, since the number of derivatives only decreases by the loop integration <sup>2</sup><sup>2</sup>2Some of the derivatives will be replaced by the loop momenta $`q1`$ which does not contribute to $`\mathrm{\Delta }`$.. We do not expand the propagators with respect to the external momenta, since we are interested only in the scaling behavior of the Feynman diagrams in which each propagator gives a negative power of the external momenta and brings the factor $`\mathrm{exp}(2t)`$ <sup>3</sup><sup>3</sup>3Namely we do not perform the derivative expansion here The asymptotic behavior in the derivative expansion will be discussed in § 6. . The loop integration does not change the factor $`\mathrm{\Delta }`$, because the support of $`\delta \theta (\widehat{p}^2)`$ is concentrated in the small region around $`\widehat{p}^21`$. The factor $`\mathrm{\Delta }`$ describes the response of the shrinkage of the loop momentum integral region. The diagrams with $`\mathrm{\Delta }<0`$ are dropped in Eq. (30). Nota Bene the massive field decouples in the asymptotic region not due to its large (dimensionless) mass but due to the shrinkage of the loop momentum integral region. One can rewrite $`\mathrm{\Delta }`$ by using the number of the loops $`L`$, $$\mathrm{\Delta }[L]=dL(N_D^{(1)}N_D^{(2)})dL,$$ (32) where we used the relation, $$N_IN_V+1=L.$$ (33) Hence all loop corrections i.e. ‘quantum’ corrections are relatively suppressed by the factor $`\mathrm{exp}(dLt)`$ compared with the tree diagrams. Since the number of derivatives is decreased by only loop integration, $`N_D^{(1)}=N_D^{(2)}`$ in the tree diagrams. Therefore all the tree diagrams survive. Now we can conclude that in the asymptotic region $`\stackrel{~}{W}_t`$ in Eq. (30) is consist of tree diagrams only. In Fig. 6 we show examples of the Feynman diagrams and their suppression factors. In these examples, the factor $`\mathrm{\Delta }`$ of first tree diagram is $`\mathrm{\Delta }=2(d4d_\varphi )2(d6d_\varphi )=0`$ and that of the second loop diagram is $`\mathrm{\Delta }=2d(4+6)d_\varphi 4(d6d_\varphi )=d<0`$. The later diagram is suppressed in comparison with the six points vertex itself. The above observation holds also for the RG flows, in which $`\delta P`$ is induced by lowering the cutoff $`\mathrm{\Lambda }`$. Hence the RG flows of the dimensionful couplings of the 1PI building blocks of the Wilsonian effective action freeze in the asymptotic region, however the Wilsonian effective action itself does not. The later conflicts to our assumption made before, i.e. $`g_i(t)g_i^Re^{d_it}`$ with fixed $`g_i^R`$. Thus there is no freezing region on the RG flow diagram of the Polchinski equation. The coordinate transformation $`\mathrm{\Delta }_{\delta P}[S_t]`$ in the asymptotic region is written by, $$\mathrm{\Delta }_{\delta P}[S_t[\varphi ]]=\frac{1}{2}\varphi \delta P^1\varphi \stackrel{~}{W}_{\mathrm{tree}}[\delta P^1\varphi ],$$ (34) where $`\stackrel{~}{W}_{\mathrm{tree}}`$ is the tree part of the connected diagrams. It can be easily realized that $`\stackrel{~}{W}_{\mathrm{tree}}`$ is given by the Legendre transform of the effective action $`S_t`$, since the 1PI part of $`\stackrel{~}{W}_{\mathrm{tree}}`$ is nothing but the ‘Legendre effective action’ $`\stackrel{~}{\mathrm{\Gamma }}_t`$. ($`\stackrel{~}{\mathrm{\Gamma }}_t`$ should not be confused with $`\mathrm{\Gamma }_\mathrm{\Lambda }`$ given in Eq. (12), which is equal to the effective action $`S_t`$ in the asymptotic region <sup>4</sup><sup>4</sup>4All the loop (quantum) corrections for the 1PI vertices are dropped in the asymptotic region. It means $`\stackrel{~}{\mathrm{\Gamma }}_t=S_t`$.. ) Therefore we find, $$\stackrel{~}{W}_{\mathrm{tree}}[J]=\overline{\varphi }JS_t[\overline{\varphi }]\frac{1}{2}\overline{\varphi }\delta P^1\overline{\varphi }.$$ (35) We add the last term of r.h.s. of Eq. (35) to the effective action, because $`\stackrel{~}{W}_t`$ in Eq. (30) consists of the connected diagrams with bare action $`S_t[\varphi ]+\frac{1}{2}\varphi \delta P^1\varphi `$. Here $`J`$ and $`\overline{\varphi }`$ satisfy the relations, $$J\delta P^1\overline{\varphi }=\frac{\delta }{\delta \overline{\varphi }}S_t,\overline{\varphi }=\frac{\delta }{\delta J}\stackrel{~}{W}_{\mathrm{tree}}.$$ (36) One can find the coordinate transformation $`\mathrm{\Delta }_{\delta P}`$ as $$\mathrm{\Delta }_{\delta P}[S_t[\varphi ]]=\frac{1}{2}S_t^{}[\overline{\varphi }]\delta PS_t^{}[\overline{\varphi }]+S_t[\overline{\varphi }],$$ (37) where the prime denotes the functional derivative with respect to $`\overline{\varphi }\left(\varphi \delta PS_t^{}[\overline{\varphi }]\right)`$. One can also easily check the following relations, $`\mathrm{\Delta }_{\delta P_1}[\mathrm{\Delta }_{\delta P_2}[S_t]]=\mathrm{\Delta }_{\delta P_1+\delta P_2}[S_t],`$ (38) $`\mathrm{\Delta }_{\delta P}[\mathrm{\Delta }_{\delta P}[S_t]]=S_t.`$ (39) The scheme dependence of the Wilsonian effective action may be understood as follows. As discussed in Ref. , the Wilsonian effective action is consist of two different elements. In the high energy region, the vertices of the effective action give the connected Green’s functions. In the low energy region, they coincide with those of the 1PI effective action, since all the articulation lines carry the infra-red cutoff. In the boundary of these regions, two quantities are connected to each other by the cutoff function. Therefore, the Wilsonian effective action is scheme dependent even after removing the cutoff. This scheme dependence turns out to be an obstacle in the approximated analyses. (See Sec. 6.) ### 3.4 Scheme independence of the cutoff Legendre effective action For the infinitesimal transformation $`\delta P1`$, the scheme dependence of the generator of the connected Green’s functions $`W_t[J]`$ becomes simpler. Now $`W_t[J]`$ is written in terms of the Wilsonian effective action $`S_t[\varphi ]`$ by $$W_t[J]=\frac{1}{2}JP_{\mathrm{\Lambda }(t)}JS_t[P_{\mathrm{\Lambda }(t)}J].$$ (40) By using Eq. (37) we find $$\delta W_t=\frac{1}{2}W_t^{}\frac{\delta P_{\mathrm{\Lambda }(t)}}{P_{\mathrm{\Lambda }(t)}^2}W_t^{}.$$ (41) Since all the loop corrections are suppressed, the 1PI parts of the cutoff connected Green’s functions in the asymptotic region are completely scheme independent. Indeed, by the Legendre transformation $$W_t[J]=J\varphi \mathrm{\Gamma }_t[\varphi ]\frac{1}{2}\varphi \left(P_{\mathrm{\Lambda }(t)}^1P_{\mathrm{\Lambda }(t)=0}^1\right)\varphi ,$$ (42) and Eq. (41), one can find $`\delta \mathrm{\Gamma }_t[\varphi ]=0`$. Namely, the renormalized trajectories of the 1PI vertices defined in the different cutoff schemes approaches to each other in the asymptotic region, as is schematically shown in Fig. 7. The coordinate transformation on the functinal space of the 1PI effective action $`\mathrm{\Gamma }_t[\varphi ]`$ induced by Eq. (20) maps the fixed points, the renormalized trajectories and the critical surfaces to those of the another scheme. Here we write this coordinate transformation as $`\mathrm{\Delta }_{(AB)}`$. Once the effective action of the scheme A ($`\mathrm{\Gamma }_t^A[\varphi ]`$) and that of the scheme B ($`\mathrm{\Gamma }_t^B[\varphi ]`$) satisfy the relation, $$\mathrm{\Gamma }_{t_1}^B[\varphi ]=\mathrm{\Delta }_{(AB)}[\mathrm{\Gamma }_{t_1}^A],$$ (43) at a certain scale $`t_1`$, then it holds also for the each scale $`t_2`$. (See Fig. 7.) Solving the RG flow equation, the cutoff Legendre effective actions finally arrive at the asymptotic region with maintaining the same relation (43). As discussed in the last subsection $`\mathrm{\Delta }_{(AB)}`$ reduces to the identical mapping in the asymptotic region. Note that the RG flow of the dimensionful cutoff Legendre effective action is frozen in the asymptotic region. The continuum limit of the field theories are found by tuning the initial boundary condition of the RG equation close to the critical surface or the fixed point and are discribed by the renormalized trajectories. As is seen above, the renormalized trajectories are cutoff scheme independent in the freezing region. This converging property of the RG flows of the cutoff Legendre effective action ensures that the solutions of the RG flow equations become cutoff scheme independent in the continuum limit. Needless to say, each theory must be specified by imposing the renormalization conditions for the renormalized couplings. Then other couplings are determined scheme independently. This structure should be compaired with the scheme dependence of the Wilsonian effective action (or the Polchinski RGE). It is an advantageous feature of the Legendre flow equations that the physically meaningful results can be obtained directly. ## 4 Sharp cutoff limit of Polchinski equation In this and next section, we confirm the equivalence between the sharp cutoff limit of the Polchinski equation and the Wegner-Houghton equation. It seems that the sharp cutoff RG equation, the Wegner-Houghton RG is quite different from the smooth cutoff one; the Polchinski RG. At first, we clarify the sharp cutoff limit of the Polchinski RGE in this section, and confirm the equivalence to the Wegner-Houghton equation in next section. Since we would like to consider the sharp cutoff limit, it is more convenient to write the cutoff propagator $`P_\mathrm{\Lambda }`$ in terms of the cutoff function $`\theta _\epsilon (q,\mathrm{\Lambda })`$. Here $`\theta _\epsilon `$ is a smooth function with respect to the momentum $`q`$, the cutoff $`\mathrm{\Lambda }`$ and a smoothness parameter $`\epsilon `$. In the limit of $`\epsilon 0`$, $`\theta _\epsilon (q,\mathrm{\Lambda })`$ becomes a step function $`\theta (q\mathrm{\Lambda })`$, therefore, $$P_\mathrm{\Lambda }(q)=\frac{1}{q^2}\theta _\epsilon (q,\mathrm{\Lambda })\stackrel{\epsilon 0}{}\frac{1}{q^2}\theta (q\mathrm{\Lambda }).$$ (44) If we also introduce $`\delta _\epsilon (q,\mathrm{\Lambda })`$ denoting the derivative of $`\theta _\epsilon `$ with respect to the cutoff $`\mathrm{\Lambda }`$, which satisfies $$\frac{}{\mathrm{\Lambda }}\theta _\epsilon (q,\mathrm{\Lambda })=\delta _\epsilon (q,\mathrm{\Lambda })\stackrel{\epsilon 0}{}\delta (q\mathrm{\Lambda }).$$ (45) It is pointed out in Ref. that one has to be careful for the behavior of $`\theta _\epsilon `$ and $`\delta _\epsilon `$ in the sharp cutoff limit. The non-trivial and universal behavior of $`\theta _\epsilon `$ and $`\delta _\epsilon `$ is $$\delta _\epsilon (q,\mathrm{\Lambda })f(\theta _\epsilon (q,\mathrm{\Lambda }),q,\mathrm{\Lambda })\stackrel{ϵ0}{}\delta (\mathrm{\Lambda }q)_0^1𝑑tf(t,q,\mathrm{\Lambda }).$$ (46) For the derivation of the above formula, see Ref. . Note that, $`\theta _\epsilon `$ with another momentum $`q^{}q`$ does not behave as Eq. (46). Let $`S_\mathrm{\Lambda }^{\epsilon =0}`$ be the effective action for the sharp cutoff case and $`S_\mathrm{\Lambda }^\epsilon `$ be one for the smooth cutoff case with $`\theta _\epsilon `$. These two effective actions should be related to each other by the formula (20), $$\mathrm{exp}\left(S_\mathrm{\Lambda }^\epsilon \left[\varphi \right]\right)=\mathrm{exp}\left(\frac{1}{2}\delta D\right)\mathrm{exp}\left(S_\mathrm{\Lambda }^{\epsilon =0}\left[\varphi \right]\right),$$ (47) where $$\delta D=\frac{d^dq}{\left(2\pi \right)^d}\frac{\overline{\delta }}{\overline{\delta }\varphi \left(𝐪\right)}\frac{1}{q^2}\left(\theta _\epsilon (q^2)\theta (q\mathrm{\Lambda })\right)\frac{\overline{\delta }}{\overline{\delta }\varphi \left(𝐪\right)}.$$ (48) Here, $`S_\mathrm{\Lambda }^\epsilon [\varphi ]`$ has the dependence of $`\theta _\epsilon `$ so that we have to take into account of Eq. (45). For the sake of simplicity, we write $`\theta _\epsilon (q_i^2)\theta (q_i\mathrm{\Lambda })\mathrm{\Delta }_i`$. The cutoff functions $`\theta _\epsilon `$ contributing to the non-trivial limit (46) should have common argument $`q`$ with that of $`\delta _\epsilon `$ in the RG flow equation. They lie only on the external legs. Diagrammatically, they can be found in Fig. 8. In this diagram, all the momentum of $`\theta _\epsilon (q)`$ are the same as that of $`\delta _\epsilon `$. Here, the filled circles in Fig. 8 correspond to the self energy $`\mathrm{\Sigma }_\mathrm{\Lambda }^\epsilon (q)`$. They are summed up and construct the ‘Full propagator’ times the inverse cutoff propagator i.e. $`(q^2/\mathrm{\Delta })/(q^2/\mathrm{\Delta }+\mathrm{\Sigma }_\mathrm{\Lambda }^\epsilon (q))`$. The pre-factor $`q^2/\mathrm{\Delta }`$, the inverse cutoff propagator, is required, since the argument of $`\stackrel{~}{W}_\mathrm{\Lambda }`$ in Eq. (30) is $`\delta P^1\varphi `$. In general, one can imagine other diagrams like Fig. 9 which the momentum $`p`$ flows in. In this figure, the filled circle denotes the multi-point 1PI vertex. Since the cutoff function $`\theta _\epsilon (𝐩+𝐪)`$ behaves differently at $`p=0`$ in the sharp cutoff limit, the discontinuous momentum dependence should appear in the beta-functional. If the field $`\varphi (𝐩)`$ is a smooth function, there are no finite contributions since the measure of the point $`p=0`$ is zero. However since $`\varphi (𝐩)`$ may have the distribution like a VEV $`\phi (2\pi )^d\delta ^d(𝐩)`$, we separate such distributions explicitly. Furthermore, since we would like to claim the equivalence between the Wegner-Houghton equation and the Polchinski equation in the sharp cutoff limit including these discontinuous momentum dependence, we also introduce the singularities like $`\delta ^d\left(𝐪𝐪_i\right)`$, $$\varphi \left(𝐪\right)(2\pi )^d\delta ^d\left(𝐪𝐪_i\right)\phi _i+\varphi \left(𝐪\right).$$ (49) Now, the effective action acquires $`\phi _i`$ dependence i.e. $`\widehat{S}_\mathrm{\Lambda }^\epsilon =\widehat{S}_\mathrm{\Lambda }^\epsilon [\varphi ,\phi _i]`$ and also satisfies the same formula (47). In this case, the Feynman diagrams like Fig. 9 contribute to the sharp cutoff flow equation via the combination $`\phi _i^{n_i}`$ with $`n_ip_i=0,n_i𝐍`$, that is a point $`p=0`$ of the diagram in Fig. 9. The corrections from $`\phi _i^{n_i}`$ terms can be absorbed by redefinition of the self energy $`\mathrm{\Sigma }_\mathrm{\Lambda }^\epsilon (q)`$. Consequently we regard the self energy $`\mathrm{\Sigma }_\mathrm{\Lambda }^\epsilon `$ as a $`\phi _i`$ dependent function. One may wonder if the self energy $`\mathrm{\Sigma }_\mathrm{\Lambda }^\epsilon (q)`$ has the discontinus momentum dependence, since there are $`\theta _\epsilon (q)`$ with momenta in common with $`\delta _\epsilon `$’s in the loop integrals giving $`\mathrm{\Sigma }^\epsilon `$. However, since the regions of the loop momentum integration have the vanishing measure, the above discontinuous momentum dependence does not contribute to Eq. (46) at all. Similarly, all the 1PI building blocks smoothly approach to those for the sharp cutoff. Hence, what we must check is only articulation lines. Furthermore, the internal lines with momenta $`𝐪+𝐩`$ vanish, since $`\mathrm{\Delta }(q+p)`$ goes to zero in the limit $`\epsilon 0`$. Therefore we need to care the external legs only. Finally, one can conclude that the relevant scheme dependence of $`n(>2)`$-point functions comes only from the external legs. Diagrammatically, they can be illustrated as in Fig. 10. The scheme dependence of the two-point function $`S_\mathrm{\Lambda }^{(2)}`$ is given by $`q^2/\mathrm{\Delta }`$ minus the ‘Full propagator’ times $`(q^2/\mathrm{\Delta })^2`$ and is different from those of other vertices. Therefore we separate the two point function in the effective action and write as $$S_\mathrm{\Lambda }[\varphi ]=\frac{1}{2}\varphi S_\mathrm{\Lambda }^{(2)}\varphi +\widehat{S}_\mathrm{\Lambda }[\varphi ],$$ (50) where $`\widehat{S}_\mathrm{\Lambda }[\varphi ]`$ is the part composed of $`n(>2)`$-point functions. Taking account of Eq. (30), we can extract the relevant scheme ($`\mathrm{\Delta }`$) dependence as follows. For the two-point function $`S_\mathrm{\Lambda }^{(2)}`$, $$\frac{\overline{\delta }^2S_\mathrm{\Lambda }^\epsilon }{\overline{\delta }\varphi (𝐩)\overline{\delta }\varphi (𝐪)}|_{\varphi =0}=\left(\frac{q^2}{\mathrm{\Delta }}\right)^2\left(\frac{\mathrm{\Delta }}{q^2}\frac{1}{q^2/\mathrm{\Delta }+\mathrm{\Sigma }_\mathrm{\Lambda }^\epsilon (q)}\right)(2\pi )^d\delta ^d(𝐩+𝐪),$$ (51) and for $`n(>2)`$-point functions $`\widehat{S}_\mathrm{\Lambda }`$, $$\widehat{S}_\mathrm{\Lambda }^\epsilon [\varphi ]=\underset{n2}{}\frac{1}{n!}\underset{i=1}{\overset{n}{}}\frac{d^dq_i}{\left(2\pi \right)^d}\left(\frac{q_i^2/\mathrm{\Delta }_i\varphi (𝐪_i)}{q_i^2/\mathrm{\Delta }_i+\mathrm{\Sigma }_\mathrm{\Lambda }^\epsilon }\right)S_\mathrm{\Lambda }^{\epsilon =0}(𝐪_1,\mathrm{},𝐪_n)+\mathrm{},$$ (52) where dots ‘$`\mathrm{}`$’ have no significant dependence on $`\mathrm{\Delta }`$ and vanish in the sharp cutoff limit. As mentioned above $`\mathrm{\Sigma }_\mathrm{\Lambda }^\epsilon `$ and $`S_\mathrm{\Lambda }^\epsilon `$ depend on $`\phi _i`$, $`e.g`$. $`\mathrm{\Sigma }_\mathrm{\Lambda }^\epsilon (q)=\mathrm{\Sigma }(q,\phi _i^{n_i})`$ and $`\widehat{S}_\mathrm{\Lambda }[\varphi ,\phi _i^{n_i}]`$. Let us first see the sharp cutoff limit of Eqs. (51) and (52). In this limit, $`\mathrm{\Delta }`$ vanishes and $`\mathrm{\Sigma }_\mathrm{\Lambda }^\epsilon (q)`$ can be replaced by $`\mathrm{\Sigma }_\mathrm{\Lambda }^{\epsilon =0}(q)`$ safely. For $`n>2`$ we easily find $$\widehat{S}_\mathrm{\Lambda }^{\epsilon =0}[\varphi ]=\underset{n2}{}\frac{1}{n!}\underset{i=1}{\overset{n}{}}\frac{d^dq_i}{\left(2\pi \right)^d}\varphi (𝐪_i)S_\mathrm{\Lambda }^{\epsilon =0}(𝐪_1,\mathrm{},𝐪_n).$$ (53) For the two-point function, we can rewrite Eq. (51) as, $$\mathrm{\Sigma }_\mathrm{\Lambda }^{\epsilon =0}(q)=q^2\mathrm{\Sigma }_\mathrm{\Lambda }^\epsilon /\left(q^2+\mathrm{\Delta }\mathrm{\Sigma }_\mathrm{\Lambda }^\epsilon \right),$$ (54) where $`\mathrm{\Sigma }_\mathrm{\Lambda }^{\epsilon =0}`$ corresponds to the two point function of the sharp cutoff effective action $`S_\mathrm{\Lambda }^{\epsilon =0}[\varphi ]`$, $$\frac{\overline{\delta }^2S_\mathrm{\Lambda }^{\epsilon =0}}{\overline{\delta }\varphi (𝐩)\overline{\delta }\varphi (𝐪)}|_{\varphi =0}=\mathrm{\Sigma }_\mathrm{\Lambda }^{\epsilon =0}(q,\phi _i^{n_i})(2\pi )^d\delta ^d(𝐩+𝐪).$$ (55) We must fix the $`\theta (0)`$ ambiguity before letting $`\epsilon 0`$ in the RG flow equation, since $`\mathrm{\Delta }`$ in Eqs. (51) and (52) is given by $`\mathrm{\Delta }(q)=\theta _\epsilon (q/\mathrm{\Lambda })\theta (q\mathrm{\Lambda })`$ and satisfy $$\delta _\epsilon (q/\mathrm{\Lambda })f(\mathrm{\Delta }(q))\stackrel{ϵ0}{}\delta (q\mathrm{\Lambda })_0^1𝑑tf(t\theta (0)).$$ (56) We simply set $`\theta (0)=0`$ here. It is different from the ordinary convention; $`\theta (0)=1/2`$. This is because, we implicitly used $`\theta (0)=0`$ to derive the Wegner-Houghton equation. The ‘shell modes’ $`\varphi _s`$ integrated out by the RG transformation have momenta $`\mathrm{\Lambda }\delta \mathrm{\Lambda }<q\mathrm{\Lambda }`$, lower than the scale $`\mathrm{\Lambda }`$ of the effective action $`S_\mathrm{\Lambda }`$. In the limit $`\delta \mathrm{\Lambda }0`$, the shell momentum $`q`$ reaches to $`\mathrm{\Lambda }`$ from below. To make this limit well-defined, we should employ the left semi-open interval $`\mathrm{\Lambda }\delta \mathrm{\Lambda }<q\mathrm{\Lambda }`$. Hence we can say that the fluctuations with $`q>\mathrm{\Lambda }`$ are incorporated in the Wilsonian effective action $`S_\mathrm{\Lambda }[\varphi ]`$, while that with $`q=\mathrm{\Lambda }`$ are not. It means that our infra-red cutoff $`\theta (q\mathrm{\Lambda })`$ satisfies $`\theta (0)=0`$! Since the ‘delta’-function $`\delta _\epsilon (q,\mathrm{\Lambda })`$ lies on the $`\mathrm{\Lambda }`$ derivative of the cutoff propagator, what we must check are the following two terms. One is a field $`\varphi (𝐪)`$ dependent term, $$\frac{\overline{\delta }S_\mathrm{\Lambda }}{\overline{\delta }\varphi (𝐪)}\left(\frac{}{\mathrm{\Lambda }}P_\mathrm{\Lambda }\right)\frac{\overline{\delta }S_\mathrm{\Lambda }}{\overline{\delta }\varphi (𝐪)}+\frac{}{\mathrm{\Lambda }}P_\mathrm{\Lambda }\frac{\overline{\delta }^2\widehat{S}_\mathrm{\Lambda }}{\overline{\delta }\varphi (𝐪)\overline{\delta }\varphi (𝐪)},$$ (57) where the first term corresponds to the ‘dumbbell’ diagram and the second term corresponds to the ‘ring’ diagram in Fig. 11. One can easily realize that the above equation is proportional to $$\frac{1}{q^2}\delta _\epsilon (q,\mathrm{\Lambda })\left(\frac{q^2/\mathrm{\Delta }}{q^2/\mathrm{\Delta }+\mathrm{\Sigma }_\mathrm{\Lambda }^\epsilon (q)}\right)^2\stackrel{\epsilon 0}{}\frac{\delta (q\mathrm{\Lambda })}{q^2+\mathrm{\Sigma }_\mathrm{\Lambda }^{\epsilon =0}(q)},$$ (58) where we were taking account of the following relations, $$\frac{\overline{\delta }S_\mathrm{\Lambda }^\epsilon }{\overline{\delta }\varphi (𝐪)}\frac{q^2}{q^2+\mathrm{\Delta }\mathrm{\Sigma }_\mathrm{\Lambda }^{\epsilon =0}}\frac{\overline{\delta }S_\mathrm{\Lambda }^{\epsilon =0}}{\overline{\delta }\varphi (𝐪)}+\text{no significant terms},$$ (59) and $$\frac{\overline{\delta }^2\widehat{S}_\mathrm{\Lambda }^\epsilon }{\overline{\delta }\varphi (𝐪)\overline{\delta }\varphi (𝐪)}\left(\frac{q^2}{q^2+\mathrm{\Delta }\mathrm{\Sigma }_\mathrm{\Lambda }^{\epsilon =0}(q)}\right)^2\frac{\overline{\delta }^2\widehat{S}_\mathrm{\Lambda }^{\epsilon =0}}{\overline{\delta }\varphi (𝐪)\overline{\delta }\varphi (𝐪)}+\text{no significant terms},$$ (60) for the functional derivative with respect to the field $`\varphi (𝐪)`$. Another is the $`\varphi (𝐪)`$ independent term which corresponds to the part evaluated in the Local Potential Approximation (LPA), $$\frac{}{\mathrm{\Lambda }}P_\mathrm{\Lambda }S_\mathrm{\Lambda }^{(2)}(q)=\frac{1}{q^2}\delta _\epsilon (q,\mathrm{\Lambda })\frac{q^2\mathrm{\Sigma }_\mathrm{\Lambda }^\epsilon }{q^2+\mathrm{\Delta }\mathrm{\Sigma }_\mathrm{\Lambda }^\epsilon }(2\pi )^d\delta ^d(𝐩+𝐪).$$ (61) By taking $`\epsilon 0`$, we find the sharp cutoff limit of this as $$\delta (q\mathrm{\Lambda })\mathrm{ln}\left(1+\mathrm{\Sigma }_\mathrm{\Lambda }^{\epsilon =0}/q^2\right)(2\pi )^d\delta ^d(𝐩+𝐪).$$ (62) They correspond to the diagrams given in Fig. 11. Using these results, the sharp cutoff limit of the Polchinski RG equation becomes, $`{\displaystyle \frac{}{\mathrm{\Lambda }}}S_\mathrm{\Lambda }^{\epsilon =0}={\displaystyle \frac{1}{2}}{\displaystyle \frac{d^dq}{(2\pi )^d}\frac{\delta (q\mathrm{\Lambda })}{q^2+\mathrm{\Sigma }_\mathrm{\Lambda }^{\epsilon =0}(q,\phi _i^{n_i})}\left\{\frac{\overline{\delta }S_\mathrm{\Lambda }^{\epsilon =0}}{\overline{\delta }\varphi (𝐪)}\frac{\overline{\delta }S_\mathrm{\Lambda }^{\epsilon =0}}{\overline{\delta }\varphi (𝐪)}\frac{\overline{\delta }^2\widehat{S}_\mathrm{\Lambda }^{\epsilon =0}}{\overline{\delta }\varphi (𝐪)\overline{\delta }\varphi (𝐪)}\right\}}`$ $`{\displaystyle \frac{1}{2}}\left(2\pi \right)^d\delta ^d(0){\displaystyle \frac{d^dq}{(2\pi )^d}\delta (q\mathrm{\Lambda })\mathrm{ln}\left(q^2+\mathrm{\Sigma }_\mathrm{\Lambda }^{\epsilon =0}(q,\phi _i^{n_i})\right)}.`$ (63) The canonical scaling of the momentum $`p_\mu /p_\mu `$ does not affect these results, since $`\mathrm{\Delta }/p_\mu 0`$ as $`\epsilon 0`$. ## 5 Comparison with the Wegner-Houghton equation To confirm the equivalence between Eq. (63) and the Wegner-Houghton RG equation, we substitute Eq. (49) to the Wegner-Houghton equation. Let us start from the following formula which gives the effective action $`S_{\mathrm{\Lambda }\delta \mathrm{\Lambda }}`$ up to $`O(\delta \mathrm{\Lambda }^2)`$. $$S_{\mathrm{\Lambda }\delta \mathrm{\Lambda }}=S_\mathrm{\Lambda }\frac{1}{2}\frac{\overline{\delta }S_\mathrm{\Lambda }}{\overline{\delta }\varphi _s}|(P_\mathrm{\Lambda }^1+\frac{\overline{\delta }^2S_\mathrm{\Lambda }}{\overline{\delta }\varphi _s\overline{\delta }\varphi _s}|)^1\frac{\overline{\delta }S_\mathrm{\Lambda }}{\overline{\delta }\varphi _s}|+\frac{1}{2}\mathrm{Tr}\mathrm{ln}(P_\mathrm{\Lambda }^1+\frac{\overline{\delta }^2S_\mathrm{\Lambda }}{\overline{\delta }\varphi _s\overline{\delta }\varphi _s}|),$$ (64) where $`\varphi _s`$ denotes the ‘shell mode’ whose support is given by the condition $`p^2=\mathrm{\Lambda }^2`$. Dot ($``$) denotes the integral $`_{\mathrm{\Lambda }\delta \mathrm{\Lambda }}^\mathrm{\Lambda }d^dq/(2\pi )^d`$. It can be realized that Eq. (64) involves the higher contribution of $`O(\delta \mathrm{\Lambda }^2)`$. First, we define $`\widehat{S}_\mathrm{\Lambda }`$ by, $$\frac{\overline{\delta }^2S_\mathrm{\Lambda }}{\overline{\delta }\varphi _s(𝐩)\overline{\delta }\varphi _s(𝐪)}|_{\varphi _s=0}\mathrm{\Sigma }_\mathrm{\Lambda }(q,\phi _i^{n_i})(2\pi )^d\delta ^d\left(𝐩+𝐪\right)+\frac{\overline{\delta }^2\widehat{S}_\mathrm{\Lambda }}{\overline{\delta }\varphi _s(𝐩)\overline{\delta }\varphi _s(𝐪)}|_{\varphi _s=0}.$$ (65) Here, $`\mathrm{\Sigma }_\mathrm{\Lambda }`$ is the same as $`\mathrm{\Sigma }_\mathrm{\Lambda }^{\epsilon =0}`$ given by Eq. (55) before. The second term of the r.h.s. of Eq. (65) is regular at $`𝐪=𝐩`$. Let us rewrite Eq. (64) in matrix notation. We define a matrix $`𝐌`$ by, $$𝐌_{p,q}\frac{\overline{\delta }^2\widehat{S}_\mathrm{\Lambda }}{\overline{\delta }\varphi _s(𝐩)\overline{\delta }\varphi _s(𝐪)}|_{\varphi _s=0}.$$ (66) The matrix $`𝐌_{p,q}`$ may have off-diagonal singularities, e.g. $`\delta (𝐩+𝐪+𝐤)`$ due to $`\phi _i`$. The first derivative of the effective action with respect to the field corresponds to a ‘vector’ $`𝐯`$; $$𝐯_q\frac{\overline{\delta }S_\mathrm{\Lambda }}{\overline{\delta }\varphi _s(𝐪)}|_{\varphi _s=0}.$$ (67) Using these notations, the r.h.s. of Eq. (64) can be expressed by the following equation, $$\frac{1}{2}𝐯^\text{T}\left((P^1+\mathrm{\Sigma })\mathrm{𝟏}+𝐌\right)^1𝐯\frac{1}{2}\mathrm{Tr}\mathrm{ln}\left((P^1+\mathrm{\Sigma })\mathrm{𝟏}+𝐌\right),$$ (68) where the unit matrix $`\mathrm{𝟏}`$ corresponds to $`(2\pi )^d\delta ^d\left(𝐩+𝐪\right)`$. We expand this with respect to the matrix $`𝐌`$. Since one momentum intergal $`_{\mathrm{\Lambda }\delta \mathrm{\Lambda }}^\mathrm{\Lambda }d^dq`$ brings a factor $`\delta \mathrm{\Lambda }`$, only the first few terms can contribute to the RG flow equation, therefore $$\frac{1}{2}𝐯^\text{T}\left((P^1+\mathrm{\Sigma })\mathrm{𝟏}\right)^1𝐯\frac{1}{2}\mathrm{Tr}\mathrm{ln}\left((P^1+\mathrm{\Sigma })\mathrm{𝟏}\right)\frac{1}{2}\mathrm{Tr}\left((P^1+\mathrm{\Sigma })\mathrm{𝟏}\right)^2𝐌+\mathrm{},$$ (69) where dots $`\mathrm{}`$ are the higher order in $`\delta \mathrm{\Lambda }`$. One of the higher order contribution is written as follows, $$\frac{\overline{\delta }S_\mathrm{\Lambda }}{\overline{\delta }\varphi _s}P_s\frac{\overline{\delta }^2\widehat{S}_\mathrm{\Lambda }}{\overline{\delta }\varphi _s\overline{\delta }\varphi _s}P_s\frac{\overline{\delta }S_\mathrm{\Lambda }}{\overline{\delta }\varphi _s},$$ (70) where $`P_s(q)`$ is the propagator of the shell mode $`\varphi _s`$ whose support is restricted to the region $`\mathrm{\Lambda }\delta \mathrm{\Lambda }<q\mathrm{\Lambda }`$. Eq. (70) corresponds to Fig. 12. Since the cross section of the integral region of $`p`$ and that of $`q`$ is $`O(\delta \mathrm{\Lambda }^2)`$, the contribution from the diagram given in Fig. 12 becomes $`O(\delta \mathrm{\Lambda }^2)`$. If $`k`$ vanishes, the volume of the integral region above becomes the first order of $`\delta \mathrm{\Lambda }`$, because two ‘spheres’ completely coincide with each other. Hence the value of the RG beta-function jumps at the point $`k=0`$. However since the field $`\varphi `$ is a smooth function of the momentum, not the distribution, we can drop it. It contributes through the distribution given in Eq. (49) by the combinations $`\phi _i^{n_i}`$ with $`n_ip_i=0`$. They are already taken in the beta-function by the $`\phi _i^{n_i}`$ dependence in the self energy $`\mathrm{\Sigma }_\mathrm{\Lambda }`$. Consequently, the Wegner-Houghton equation can be found as, $`S_\mathrm{\Lambda }S_{\mathrm{\Lambda }\delta \mathrm{\Lambda }}={\displaystyle \frac{\delta \mathrm{\Lambda }}{2}}{\displaystyle \frac{d^dq}{(2\pi )^d}\frac{\delta (q\mathrm{\Lambda })}{q^2+\mathrm{\Sigma }_\mathrm{\Lambda }(q,\phi _i^{n_i})}\left\{\frac{\overline{\delta }S_\mathrm{\Lambda }}{\overline{\delta }\varphi (𝐪)}\frac{\overline{\delta }S_\mathrm{\Lambda }}{\overline{\delta }\varphi (𝐪)}\frac{\overline{\delta }^2\widehat{S}_\mathrm{\Lambda }}{\overline{\delta }\varphi (𝐪)\overline{\delta }\varphi (𝐪)}\right\}}`$ $`{\displaystyle \frac{\delta \mathrm{\Lambda }}{2}}\left(2\pi \right)^d\delta ^d(0){\displaystyle \frac{d^dq}{(2\pi )^d}\delta (q\mathrm{\Lambda })\mathrm{ln}\left(q^2+\mathrm{\Sigma }_\mathrm{\Lambda }(q,\phi _i^{n_i})\right)}+O(\delta \mathrm{\Lambda }^2).`$ (71) This flow equation is nothing but the sharp cutoff limit of the Polchinski equation (63). ## 6 Conclusion and remarks In this article, we investigated the cutoff scheme dependence of the Wilsonian effective action. It can be reinterpleted as the coordinate transformation on the theory space. It is written formally by Eq. (20). We have studied it in two limiting cases. One is in the asymptotic region i.e. $`t\mathrm{}`$, and another is in the sharp cutoff limit i.e. $`\epsilon 0`$. In the both cases, we could write down the cutoff scheme dependence so simple as to explore the RG flows. As we have shown in Sec. 3, the scheme dependence of the renormalized trajectories in the asymptotic region $`t\mathrm{}`$ remains for the Wilsonian effective action. Besides, the RG flow of the Wilsonian effective action does not freeze in $`t\mathrm{}`$. The origin is as follows. The vertices of the Wilsonian effective action consist of two different quantities; the connected Green’s function at high energy region ($`p>\mathrm{\Lambda }`$) and the 1PI vertices at the low energy region ($`p<\mathrm{\Lambda }`$). The boundary between these regions are connected scheme dependently. (See also Ref. .) Therefore the Wilsonian effective action itself is not a physical quantity. Moreover, we have also shown the scheme independence of the Legendre effective action, or equivalently the 1PI building blocks of the Wilsonian effective action, on the renormalized trajectories. Recalling the statements in Sec. 3, we can say that if the RG flow of the dimensionful Legendre effective action $`\mathrm{\Gamma }_\mathrm{\Lambda }[\varphi ]`$ freezes on the renormalized trajectory in the asymptotic region, i.e. in the statistical continuum limit $`\mathrm{\Lambda }_0\mathrm{}`$, then our $`\mathrm{\Gamma }_\mathrm{\Lambda }[\varphi ]`$ should be scheme independent. In the perturbation theory, it can be easily realized. Indeed, all the cutoff scheme dependent contributions, i.e. the coefficients of the divergences, are completely absorbed into certain counterterms order by order, and remaining finite terms are scheme independent in the limit $`\mathrm{\Lambda }_0\mathrm{}`$. Of course, needless to say, we should insist on the common renormalization condition (or equivalently the subtruction rule). In the non-perturbative case, however, the problem will be more complicated, because the ordinary renormalization procedure will not work in general, e.g. for the theory around a non-Gaussian fixed point. Hence, the cancellation of divergences, and therefore the cutoff scheme dependent constants can be confirmed only case by case if possible. Turning to the Exact Renormalization Group, we can recapture it from another point of view. The scheme dependence of the counterterms corresponds to that of the fixed point and/or of the critical surface, and the scheme independence of the total solution can be appreciated by that of the renormalized trajectory in the asymptotic region. All these are described by a coordinate transformation (20). For our purpose, it is sufficient to investigate the asymptotic region of Eq. (20) without using the explicit solutions, since we have expected the asymptotic behavior $`g_i(t)g_i^Re^{d_it}`$ and do not need the explicit value $`g_i^R`$. Once we assume existence of the asymptotic region, then the scheme independence of $`\mathrm{\Gamma }_\mathrm{\Lambda }[\varphi ]`$ as $`\mathrm{\Lambda }0`$ is confirmed. (Recall the discussion in Sec. 3.) For massive theories, the scheme dependence of $`\mathrm{\Gamma }_\mathrm{\Lambda }[\varphi ]`$ decays like $`\mathrm{exp}(dt)(\mathrm{\Lambda }/M_R)^d`$ as $`t\mathrm{}`$. Instead, for the massless case, one may start from a massive case and then letting $`M_R0`$. We also confirm the equivalence between the Wegner-Houghton equation and the Polchinski equation in the sharp cutoff limit. It seems that these equations are much different from each other even though they are expected to be equivalent. We can prove equivalence of these two ERGs by help of Eq. (20) which describes the scheme dependence of the Wilsonian effective action. The superficial difference occurs by the strong scheme dependence of the Wilsonian effective action. As we showed, the crucial cutoff scheme dependence of the Wilsonian effective action lies in the external legs. Finally, we would like to comment on the cutoff scheme dependence of the approximate solutions. The ERG flow equations are approximated by projecting them onto smaller dimensional subspaces of the original theory space. In the derivative expansion, for example, we may employ these subspaces as the space of a finite number of the coefficient functions $`\{V_0,V_2,\mathrm{},V_k^i\}`$ defined by the following equation, $$S_\mathrm{\Lambda }[\varphi ]=d^dx\left\{V_0(\varphi )+\frac{1}{2}(_\mu \varphi )^2V_2(\varphi )+\frac{1}{2}(\mathrm{}\varphi )^2V_4^1(\varphi )+\mathrm{}\right\},$$ (72) The subscript $`k`$ of the coefficient function denotes the degrees of the derivatives and the superscript $`i`$ of it labels the independent $`k`$-th derivative vertices. Then, the ERG flow equations are reduced to the coupled partial differential equations for the coefficient functions $`V_k^i(\varphi )`$. One can easily improve the approximation systematically by enlarging the subspace $`\{V_0,V_2,\mathrm{},V_k^i\}`$ step by step. Especially, the approximation with $`k=0`$ is called the ‘local potential approximation’ (LPA). This procedure preserves the non-perturbative nature of the ERG flow equations. The scheme dependence given by Eq. (37) are infinitely enhanced in the derivative expansion. By dimensional analysis, the Taylor expansion of $`\delta P(q)`$, whose value changes rapidly near the infra-red cutoff $`q\mathrm{\Lambda }`$, is the expansion with respect to the combination $`q^2/\mathrm{\Lambda }^21`$. Therefore the scheme dependence of the coefficient functions $`V_i^k(\varphi )`$ in Eq. (72) become stronger and stronger as the infra-red cutoff $`\mathrm{\Lambda }`$ decreases. The scheme dependence of $`V_k^i(\varphi )`$ behaves as $`1/\mathrm{\Lambda }^{2k}`$. At last it diverges in the limit $`\mathrm{\Lambda }0`$. It means that the derivative expansion and the limit $`\mathrm{\Lambda }0`$ do not commute. Hence the cutoff scheme dependence of $`V_i^k(\varphi )`$ is strong enough to prevent the physical predictions. The physical information is completely washed off except for the potential part $`V_0(\varphi )`$ which is the 1PI effective potential. The RG beta-functionals of the coefficient functions of $`\mathrm{\Gamma }_\mathrm{\Lambda }[\varphi ]`$ like Eq. (72) are also cutoff scheme dependent in the region $`t\mathrm{}`$, since by the dimensional analysis, the expanding parameter there is $`/\epsilon /\mathrm{\Lambda }`$ where $`\epsilon `$ stands for the smoothness parameter given in Sec. 4. The higher derivative contributions finally overcome the suppression factor $`(\mathrm{\Lambda }/M_R)^d`$. Hence, the RG beta-functionals of the higher derivative operators blow up to infinity. In the limit $`t\mathrm{}`$, the cutoff scheme approaches towards to the sharp one, since $`\epsilon \mathrm{\Lambda }0`$. It is known that these diverging series can be summed up and lead to non-analytical momentum dependence of the vertices, e.g. $`\sqrt{p_\mu p_\mu }`$. This spurious scheme dependence can be avoided if we work on the sharp cutoff Legendre flow equation . ## Appendix Equation (30) in Sec. 3 is driven as follows. First, let us consider the positive definite deviation of the cutoff propagator $`\delta P(q)=\delta P_+(q)>0`$, since we call for the Gaussian integral with positive $`\delta P(q)`$. Then one can find $`\mathrm{exp}\left({\displaystyle \frac{1}{2}}\delta D_+[\delta /\delta \varphi ]\right)\mathrm{exp}\left(S_t\left[\varphi \right]\right)=\mathrm{exp}\left({\displaystyle \frac{1}{2}}\delta D_+[\delta /\delta \varphi ]\right)\mathrm{exp}\left(S_t\left[\varphi \right]+J\varphi \right)|_{J=0}`$ $`=\mathrm{exp}\left({\displaystyle \frac{1}{2}}\delta D_+[\delta /\delta \varphi ]\right)\mathrm{exp}\left(S_t\left[\delta /\delta J\right]\right)\mathrm{e}^{J\varphi }|_{J=0}`$ $`=\mathrm{exp}\left(S_t\left[\delta /\delta J\right]\right)\mathrm{exp}\left({\displaystyle \frac{1}{2}}\delta D_+[J]+J\varphi \right)|_{J=0}`$ $`\mathrm{exp}\left(S_t\left[\delta /\delta J\right]\right){\displaystyle D\varphi ^{}\mathrm{exp}\left(\frac{1}{2}\varphi ^{}\delta P_+^1\varphi ^{}+J(\varphi +\varphi ^{})\right)}|_{J=0}`$ $`={\displaystyle D\varphi ^{}\mathrm{exp}\left(\frac{1}{2}\varphi ^{}\delta P_+^1\varphi ^{}S_t\left[\varphi +\varphi ^{}\right]\right)}`$ $`=\mathrm{exp}\left({\displaystyle \frac{1}{2}}\varphi \delta P_+^1\varphi \right){\displaystyle D\varphi ^{}\mathrm{exp}\left(\frac{1}{2}\varphi ^{}\delta P_+^1\varphi ^{}S_t\left[\varphi ^{}\right]+\varphi ^{}\delta P_+^1\varphi \right)}`$ $`=\mathrm{exp}\left({\displaystyle \frac{1}{2}}\varphi \delta P_+^1\varphi +W_t[J=\delta P_+^1\varphi ]\right),`$ (73) where $`\delta D_+[J]`$ is given by, $$\delta D_+[J]=\frac{d^dq}{(2\pi )^d}J(q)\delta P_+(q)J(q).$$ (74) The negative part of the deviation $`\delta P_{}(q)`$ needs the special care, since the path-integral in the fourth line does not converge. However our final result can be hold also for the negative part $`\delta P_{}(q)`$, since Eq. (30) is the identity of $`\delta P`$. In other words, Eq. (30) means the graph by graph correspondence of the Feynman diagrams. It does not restrict our observations in Secs. 3-5, since we need the diagramatical representation of Eq. (73).
warning/0002/cond-mat0002253.html
ar5iv
text
# Effects of Electron-Electron Scattering on Electron-Beam Propagation in a Two-Dimensional Electron-Gas ## I Introduction The propagation of electron beams in the two-dimensional electron-gas (2DEG) of GaAs-(Al,Ga)As heterostructures was studied in a number of publications , and has proven to be a very sensitive tool for studying electron scattering-phenomena. In the first two Refs. , the emphasis was on the effects of electron-phonon scattering, where the beam was injected across tunnel barriers. These effects occur at relatively large excess energies of the electron beam, typically of the order of the optical phonon energy, some 30 meV. In later works , the effects of electron-electron scattering phenomena (occurring at much lower energies, typically below 10 meV) were analyzed, using opposite quantum point-contacts as injector and detector for the electron beam. In our paper, Ref. , we paid much attention to thermal beams in which the characteristic energy of beam electrons $`\epsilon `$, counted from Fermi level, is of the order of the sample temperature $`T_0`$. It was shown that electron-electron collisions played a main role in damping such beams. The overall behaviour of the signal attenuation could be reasonably understood using the formula of Giuliani and Quinn for the electron-electron scattering rate in a 2DEG, implying that the result of a single electron-electron collision is sufficient for an unequilibrium electron to escape from detection (relaxation time approximation). This conclusion was subsequently confirmed by other groups . In this work, we return to our studies of electron-electron scattering in a 2DEG system, equipped with a much more detailed framework of understanding of the dynamical scattering phenomena , which has first been used to explain the hydrodynamic electron flow-phenomena we observed a few years ago . These newly developed theories enable a much more refined analysis of the experimental data. Specifically targeting the theoretical predictions of Refs. we have performed a new series of electron-beam experiments for different samples at various temperatures and for a wide range of injection energies. In our experiments we can identify specific two-dimensional effects, as well as novel nonlinearities due to 2DEG-heating. Our results cast doubts on the interpretations in Refs. and . In the course of this paper, we will first present the experimental results and their qualitative explanation (Sec. II). Next we develop a theoretical approach for the electron-beam propagation in small systems, i.e. where the probability of secondary collision is negligible (Sec. III), and for the opposite case, the multi-collision limit (the propagation of a beam over long distances becomes possible due to specific two-dimensional effects) (Sec. IV). In Sec. V we consider non-linear phenomena which can play an essential role for the interpretation of an electron-beam signal and we analyse the experimental data in the framework of the here developed theory in Sec. VI. Throughout this paper we will use ’energy-units’ for temperature and potential differences, i.e. the Boltzmann constant $`k_B`$ and the electron charge $`e`$ are equal to one. ## II Experiment The experiments were performed on gate-defined nanostructures in conventional modulation doped GaAs-(Al,Ga)As heterojunctions. Typical values for the carrier density and mobility are $`n_e=2.45\times 10^{11}`$ cm<sup>-2</sup> and $`\mu 1\times 10^6`$ cm<sup>2</sup>(Vs)<sup>-1</sup>, corresponding to an impurity mean-free-path of $`l_{\mathrm{imp}}`$ 20 $`\mu `$m. A schematic topview of the sample gate-structure is given in Fig. 1. Schottky gates (grey areas) form two opposite quantum point-contacts, $`i`$ and $`d`$, at lithographical distances of $`L=`$ 0.6, 2, 3.4, and 4 $`\mu `$m for different samples. In the experiments, the electron beam was injected through the injector quantum point-contact $`i`$ by applying a dc voltage $`V_i=V_{12}`$ and detected as the non-local voltage $`V_d=V_{34}`$ across the detector point-contact $`d`$. The numbers 1, 2, 3, and 4 denote the Ohmic contacts to the 2DEG of the sample (Fig. 1, crossed squares). We stress that the use of all-dc techniques is very important for a proper interpretation of the observed signals. Differential resistance measurements with lock-in techniques will not elucidate the role of the thermovoltage background to the signal in full. Both injector and detector point-contacts were adjusted at the $`n=1`$ plateau i.e., both contain one transverse mode, and thus remain in the metallic regime, $`G_{\mathrm{QPC}}=n2e^2/h`$. (In other words, they do not act as tunnel barriers, as was the case in Refs. .) Thus, electrons of all possible energies, $`0<\epsilon <V_i`$, are present in the injected beam. In the presence of a magnetic field perpendicular to the 2DEG plane, the electron beam is deflected and the detector signal, $`V_d(B)`$ yields the beam profile (see Fig. 2). For the present a point-contact adjustment at $`n=1`$ and a injector-detector distance $`L=3.4`$ $`\mu `$m we obtain the characteristic opening angle of injector and detector which amounts to $`\varphi 18^{}`$ (cf. ). From Fig. 2, showing the $`V_d(B)`$-dependence for different injection energies ($`0.5V_i10`$ mV) at 1.6 K, one can see that the detector signal first increases with increasing injector voltage. Then for $`V_i>3`$ mV a strong increase of a isotropic background signal is observed while at the same time the beam profile broadens. For injection energies larger 10 meV a beam signal can hardly be resolved, while the background increases continuously. To investigate the effects of electron-electron scattering events on the beam propagation we are interested in the dependence of detector-signal on the injection energy at $`B=0`$. Fig. 3 a) presents the experimental results for the sample $`L=3.4`$ $`\mu `$m at three different sample temperatures, $`T_0`$ = 1.6, 8 and 11 K. Additional measurements (not shown here) were made at different sample (lattice) temperatures, $`T_0`$ = 2.2, 3.4, 5, 15, and 17 K and for different injector-detector distances. It can be seen from Fig. 3 a) that for low injection energies the detector signal increases linearly with $`V_i`$. For $`V_i>3`$ meV only for the lowest temperature (curve 1) a saturation and even a small decrease is observed. For high injection energies the $`V_d(V_i)`$ dependence increases for all temperatures. As we have seen from Fig. 2, for $`V_i>3`$ meV an increasing isotropic background signal is detected, which is not directly related to the ballistic electron-beam propagation. In order to extract the ballistic part of the detector signal we measured the isotropic background signal separately by repeating the experiment for high magnetic fields ($`B=50`$ mT) to ensure that the electron beam is totally deflected and ballistic beam electrons do not contribute to the detector signal \[Fig. 3 b)\]. Subtracting this background signal from the data measured for $`B=0`$ T we obtain the pure electron-beam contribution to the detector signal \[Fig. 3 c)\]. Now the result is similar for all temperatures: We observe first a linear increase of $`V_d`$ with increasing $`V_i`$ and then a saturation followed by a decrease for high injection energies, while with increasing sample temperature the maximum electron-beam signal decreases \[Fig. 3 c)\]. These experimental results can be understood from the following qualitative considerations. Let us assume, for simplicity, that the lattice temperature (the primary temperature of the system) is equal to zero. Then, for a nonequilibrium electron with excess energy $`\epsilon `$ above the Fermi energy $`\epsilon _F`$, the mean-free-path for collisions with equilibrium electrons decreases with increasing $`\epsilon `$, roughly speaking as $`ł_{ee}(\epsilon )\epsilon ^2\mathrm{ln}\epsilon ^1`$, $`\epsilon \epsilon _F`$ . Therefore, at sufficiently small $`V_i`$ all injected electrons will reach the detector, whose readout then is proportional to the number of injected electrons, $`V_dV_i`$, schematically shown in Fig. 4. (Electrons of all energies $`0\epsilon V_i`$ are present in the beam, with equal weight). This linear increase of $`V_d`$ with $`V_i`$ saturates for energies $`\epsilon \epsilon _0`$ when the electron- electron scattering mean free path length ($`l_{ee}`$) becomes comparable to $`L`$, the distance between injector and detector: $`V_i=\epsilon _0`$ for $`ł_{ee}\left(\epsilon _0\right)=L`$. Electrons with larger energies, $`\epsilon _0<\epsilon <V_i`$, will scatter and do not reach the detector. Thus, the signal $`V_d`$ is determined by an fraction of electrons which is completely saturated at $`V_i\epsilon _0`$ and should not change on a further increase of $`V_i`$. However, as is evident from Fig. 3 a), the signal, upon reaching a maximum, starts to decrease slightly. The only possible mechanism leading to such behaviour is heating of the 2DEG in between injector and detector point-contact by the electron beam. The heated 2DEG then leads to damping of the electron beam due to enhanced electron-electron scattering. At still higher $`V_i`$, $`V_d`$ shows again an increase \[Fig. 3 a)\]. This is due to the additional build-up of a thermovoltage across the detector point-contact \[Fig. 2 and 3 b)\], which is driven by the temperature difference between the heated 2DEG in between injector and detector and the still cold 2DEG behind the detector . As we will see below, the qualitative picture given above is fully confirmed by the theory described in this work. We will demonstrate that under our experimental conditions it is possible to separate the electrons of the beam into two groups, i.e. ”quasiballistic” ($`\epsilon <\epsilon _0`$) and ”heating electrons” ($`\epsilon >\epsilon _0`$), which greatly simplifies the interpretation of the experimental results. ## III One-Collision Approximation The detected signal $`V_d`$ is determined by the distribution function of nonequilibrium electrons $`f`$ in the vicinity of detector point-contact. For now, we neglect nonlinearities due to heating of the 2DEG in between injector and detector, which is a valid approximation for sufficiently low excess energies of the injected electrons. The linearized Boltzmann equation, describing the behaviour of the distribution function $`f`$, then has the form $$v_x\frac{f}{x}+v_y\frac{f}{y}=\widehat{J}f,f\left(x=0,y,𝐩\right)=f_0(y,𝐩)$$ (1) Here, $`f_0(y,𝐩)`$ is the beam profile at the exit from the injector, and the axis $`x`$ is directed from injector to detector (cf. Fig. 1). $`\widehat{J}f`$ is a linearized integral describing the electron-electron collisions. It is convenient to write it as $$\widehat{J}f=\nu f+𝑑𝐩^{}\nu _{\mathrm{𝐩𝐩}^{}}f_𝐩^{},\nu =𝑑𝐩^{}\nu _{𝐩^{}𝐩}.$$ (2) Here, the collision-integral kernel $`\nu _{𝐩^{}𝐩}`$ determines the probability of the appearance of a nonequilibrium electron ($`\nu _{𝐩^{}𝐩}>0`$) or hole ($`\nu _{𝐩^{}𝐩}<0`$) in state $`𝐩^{}`$ after the nonequilibrium electron has disappeared from state $`𝐩`$ (i.e., has been scattered into another state). The kernel has a complex structure and in the general case can not be presented in elementary functions. We have $$\nu _{𝐩^{}𝐩}=\frac{1}{n\left(\epsilon \right)}𝑑𝐩_1𝑑𝐩_2(2\mathrm{\Psi }_{𝐩^{}𝐩_1\mathrm{𝐩𝐩}_2}\mathrm{\Psi }_{𝐩^{}\mathrm{𝐩𝐩}_1𝐩_2}),$$ (3) where $`\mathrm{\Psi }_{𝐩^{}\mathrm{𝐩𝐩}_1𝐩_2}=W_{𝐩^{}\mathrm{𝐩𝐩}_1𝐩_2}(1n\left(\epsilon ^{}\right))n\left(\epsilon _1\right)n\left(\epsilon _2\right)\times `$ (4) $`\times \delta \left(𝐩^{}+𝐩𝐩_1𝐩_2\right)\delta \left(\epsilon ^{}+\epsilon \epsilon _1\epsilon _2\right).`$ (5) Here $`W_{𝐩^{}\mathrm{𝐩𝐩}_1𝐩_2}`$ is proportional to the square of the matrix element of the electron-electron interaction, and $`n\left(\epsilon \right)`$ is the equilibrium Fermi distribution function. We now introduce the angular scattering distribution function: $$g\left(\phi \right)=\nu ^1m𝑑\epsilon ^{}\nu _{𝐩^{}𝐩},$$ (6) where $`\phi `$ is the scattering angle. For simplicity, we assume parabolic bands ($`\epsilon =p^2/2m`$), which is a good approximation for the conduction band in GaAs-(Al,Ga)As heterostructures. At sufficiently small $`\epsilon `$ and $`T`$, the form of $`g\left(\phi \right)`$ is determined mainly by the phase-space restrains imposed by the two-dimensional character of the 2DEG . Roughly speaking, $`g\left(\phi \right)`$ consists of a narrow bunch of electrons flying forward in an angle range of the order of $`\pm \left(\epsilon +T\right)^{1/2}\epsilon _F^{1/2}`$ and a bunch of holes, of approximately the same width, flying backward (see Ref. ). Therefore, the electron-electron scattering is effectively a small-angular process . For the general case, Eq. (1) can not be solved. However, under conditions where the probability of collisions is small, i.e. $`ł_{ee}=v(\epsilon )\nu ^1(\epsilon )L`$, or $`\epsilon \epsilon _0`$, we can use perturbation theory for the collision integral. In the first order or one-collision approximation we then have $`f(x,y,𝐩)=\left(1{\displaystyle \frac{x\nu }{v_x}}\right)f_0(y{\displaystyle \frac{v_y}{v_x}}x,𝐩)+`$ (7) $`+{\displaystyle \frac{1}{v_x}}{\displaystyle \underset{0}{\overset{x}{}}}𝑑x^{}{\displaystyle 𝑑𝐩^{}\nu _{\mathrm{𝐩𝐩}^{}}f_0[y\frac{v_y}{v_x}x+\left(\frac{v_y}{v_x}\frac{v_y^{}}{v_x^{}}\right)x^{},𝐩^{}]}`$ (8) $`\left(1{\displaystyle \frac{x\nu }{v_x}}\right)f_0+\widehat{Q}f_0.`$ (9) The first term on the r.h.s of Eq. (6) describes the number of nonscattered particles reaching into the vicinity of a point $`(x,y)`$. The second (integral) term $`\widehat{Q}f_0`$ describes particles that reach the same spatial region, after having been scattered once. Note that for high-energy beams ($`\epsilon T`$), the probability of undergoing a second collision is approximately one order of magnitude lower than that of the first collision . This is connected with the fact that after collision with equilibrium (Fermi sea) electrons, the excess energy of a nonequilibrium electron ($`\epsilon `$) must be redistributed between three partners i.e., $`\overline{\epsilon }\epsilon /3`$, $`ł_{ee}(\overline{\epsilon })3^2ł_{ee}(\epsilon )`$, $`T\epsilon \epsilon _F`$, where $`\overline{\epsilon }`$ is the characteristic energy of the scattered electrons. Therefore, the one-collision approximation is valid for a relatively wide range of energies as long as $`ł_{ee}(\overline{\epsilon })L`$. On observing this and the fact that $`\widehat{Q}\nu f_0\nu \left(\epsilon \right)\widehat{Q}f_0\nu \widehat{Q}f_0\nu \left(\overline{\epsilon }\right)\widehat{Q}f_0`$, it is straightforward to build a new ”modified” one-collision approximation. After partial summation of the terms of the iteration series on the parameter $`x/l_{ee}(\epsilon )`$ of Eq. (1) one obtains the following expression in zero-eth order approximation for the parameter $`x/l_{ee}\left(\overline{\epsilon }\right)`$: $`fe^{\frac{\nu x}{v_x}}f_0(y{\displaystyle \frac{v_y}{v_x}}x,𝐩)+`$ (10) $`+{\displaystyle \frac{1}{v_x}}{\displaystyle \underset{0}{\overset{x}{}}}𝑑x^{}{\displaystyle 𝑑𝐩^{}\nu _{\mathrm{𝐩𝐩}^{}}e^{\frac{\nu (\epsilon ^{})x^{}}{v_x^{}}}}`$ (11) $`\times f_0[y{\displaystyle \frac{v_y}{v_x}}x+\left({\displaystyle \frac{v_y}{v_x}}{\displaystyle \frac{v_y^{}}{v_x^{}}}\right)x^{},𝐩^{}].`$ (12) This formula is valid when $`\epsilon <3\epsilon _0`$. The first term on the r.h.s. corresponds to the usual relaxation-time approximation, $`\widehat{J}f=\nu f`$. Note that the modified one-collision approximation Eq. (7) is based on an exact consideration of the first collision and not on perturbation theory. It does not take into account any further collisions. The experimentally measured voltage drop, $`V_d`$, is determined by the current passing the detector point-contact and can be calculated from $$V_d=e𝑑\epsilon 𝑑\phi \rho \left(\phi \right)v_xf\left(x=L,y=0,𝐩\right).$$ (13) Here $`\rho \left(\phi \right)`$ is the function characterizing the angular acceptance of the detector point-contact, which is positioned at $`(L,0)`$. For simplicity, we use in our numerical calculations Heaviside step-functions to represent the angular characteristics of injector and detector point-contacts: $`\rho \left(\phi \right)`$ $``$ $`\theta \left(\varphi /2\left|\phi \right|\right)`$ (14) $`f_0(y,𝐩)`$ $``$ $`\theta \left(\varphi /2\left|\phi \right|\right)\theta \left(V_i\epsilon \right).`$ (16) In this model, the behaviour of $`V_d\left(V_i\right)`$ is determined by two parameters, i.e. the angular injection (and acceptance) range of the point-contact $`\varphi `$ and the distance between injector and detector $`L`$. For more realistic models of the angular response of quantum point-contacts we refer to Ref. . The dependence $`V_d(V_i)`$ is calculated using Eqs. (7-9), including the expressions for the kernel $`\nu _{\mathrm{𝐩𝐩}^{}}`$ obtained in Ref. , setting $`\varphi =18^{}`$, and $`L`$ = 3.4 $`\mu `$m, i.e. close to the experimental conditions. The result is shown in Fig. 5 for the full expression of the modified one-collision approximation (MOCA) (curve 1) and the relaxation-time approximation (RTA) (curve 2). We clearly observe that the curves saturate with increasing $`V_i`$. Saturation occurs at a higher injection voltage, $`V_i`$, and a higher signal level, $`V_d`$, for the MOCA as compared with the RTA. The difference between these curves (about 15 %) characterizes the role of two-dimensional effects for the given parameters. This difference is due to the integral term in Eq. (7), which can be omitted when 2D effects are negligible. In the next section we show that the role of the two-dimensionality is much larger when $`\varphi >\left(\epsilon _0/\epsilon _F\right)^{1/2}`$. In this limit, a saturation of the curve at $`V_i<3\epsilon _0`$ does not take place at all. ## IV Multi-collision regimes In the limit where the electrons undergo a number of collisions on their way from injector to detector, it is impossible to obtain a completely analytical solution of the spatially-inhomogeneous problem of beam propagation. Instead, we will discuss below a simple qualitative theory that adequately describes this multi-collision regime. To obtain realistic numerical values of the angular relaxation rate we first consider the momentum relaxation in time for a spatially homogeneous distribution. For simplicity, we take the thermalized distribution $`f=(n/\epsilon )\chi (\phi ,t)`$, i.e. equilibrium is established in energy but not in momentum. In this case the kernel of the collision integral contains only differences in the angular variables, and the solution of the Boltzmann equation reduces to the calculation of a one-dimensional Fourier transform. Here, we use the numerical results of the angular distribution function $`g(\phi )`$ that were obtained in Ref. . (In the case of non-thermalized distributions the angular and energy variables are not separable and therefore the solution of the Boltzmann equation becomes a much more difficult problem.) Fig. 6 shows the results of a calculation for thermalized conditions at $`T_0=0.1\epsilon _F`$ and for different times $`t`$ after beam injection. From this figure follows that the beam remains narrow up to times of the order of $`10\tau _{ee}`$, whereas in the three-dimensional case a smooth drift-like distribution is already established after a delay of the order of one collision-time. From now on, we will use the convention $`0.1\epsilon _F\epsilon ^{}`$ to denote the characteristic energy of a beam, below which the specific features of two-dimensional relaxation essentially manifest themselves. Depending on the relative magnitude of $`\epsilon _0`$ and the temperature $`T`$ of the 2DEG, different multi-collision regimes are possible: 1. Let us start with the case of low temperatures, $`T\epsilon _0`$. We assume that $`L`$ is so large that $`\epsilon _0<\epsilon ^{}`$. In this case, the particles that undergo multiple collisions but still contribute to the electron beam signal, are those whose mean-free-path is considerably less than $`L`$ and whose scattering is small-angular: $$\epsilon _0<\epsilon <\epsilon ^{}.$$ (17) Note that after a few collisions the energy of such particles drops very rapidly to values close to $`\epsilon _0`$, upon which the particles will reach the detector without further collision. In contrast, the opening angle broadening $`\alpha `$ of the electron beam is determined by the first collision $`\alpha \sqrt{\epsilon /\epsilon _F}1`$. It is then straightforward to evaluate the contribution to the detector signal of the group of electrons with energies in the range $`(\epsilon _0,\epsilon )`$: $$V_d(\epsilon \epsilon _0)\frac{\lambda _F}{r_{}}\frac{\varphi }{\alpha }\epsilon _F\frac{\epsilon \epsilon _0}{\epsilon }\frac{\lambda _F}{L}\varphi .$$ (18) The transverse beam broadening is $`r_{}L\alpha `$. The detector width is chosen to be of the order of the electron Fermi wave-length $`\lambda _F`$ (corresponding to the occupation of one mode in the quantum point-contact). We have assumed an angle of acceptance $`\varphi \alpha `$ for the detector point-contact ; in the other case, if $`\varphi \alpha `$, the multiplier $`\varphi /\alpha `$ can be omitted for the above expression. The order of magnitude of the contribution of ballistic electrons to the detector signal can be estimated as $`V_d\epsilon _0\lambda _F(L\varphi )^1`$, where we assume identical characteristics for injector and detector quantum point-contacts. Therefore, the condition for a predominance of the group of non-ballistic electrons to the detector signal takes the form $$\varphi >\sqrt{\frac{\epsilon _0}{\epsilon _F}}.$$ (19) This inequality is satisfied more easily for samples with larger $`L`$ (i.e. smaller $`\epsilon _0`$) or larger acceptance angles $`\varphi `$. Under our experimental conditions the l.h.s. and r.h.s. in Eq. (12) coincide by an order of magnitude, and it turns out that the 2D effects lead to corrections of the order of unity. In case $`\varphi \sqrt{\epsilon _0/\epsilon _F}`$ it should be possible to observe the long-distance beam propagation as predicted in Ref. . In other words, one can detect an electron beam over a distance exceeding substantially the electron-electron mean-free-path $`l_{ee}`$ as a result of one-dimensional electron-hole diffusion. 2. For $`\epsilon _0T`$, ballistic electrons are practically absent. Roughly speaking, the number of quasi-particles reaching the detector without any collisions is exponentially small and proportional to $`\mathrm{exp}[L/l_{ee}(T)]=\mathrm{exp}(T^2\epsilon _0^2)`$. High-energy electrons with energies $`\epsilon _F>\epsilon >T`$ loose their excess energy very quickly, after a few collisions, and cool down to energies of the order of the lattice temperature $`T`$. Simultaneously, the beam acquires an angular broadening of the order of $`\sqrt{\epsilon /\epsilon _F}<1`$. After this initial relaxation, provided $`T<\epsilon ^{}`$, we still have a narrow distribution of electrons (and holes with opposite momenta) whose movement is a one-dimensional diffusion in coordinate space – an effect which is genuinely caused by the two-dimensionality of the electron system. The angular broadening in time of this specific group can be expressed as $`\alpha \sqrt{T/\epsilon _F}[t/\tau _{ee}(T)]^{1/4}`$ (cf. ). The contribution of this narrowly-directed group of electrons to the detected signal can be evaluated using Eq. (11). Taking into account, that, assuming one-dimensional diffusion, the time an electron needs to reach from injector to detector is of the order of $`v_F^1L^2l_{ee}^1(T)`$ we obtain $`\alpha \sqrt{\epsilon /\epsilon _F}+\sqrt{T/\epsilon _F}\left[L/l_{ee}(T)\right]^{1/2},`$ (20) (21) $`r_{}L\left\{\sqrt{\epsilon /\epsilon _F}+\sqrt{T/\epsilon _F}\left[L/l_{ee}(T)\right]^{3/2}\right\}.`$ (22) As one can see from these expressions, the result of Eq. (11) obtained above is retrieved for $`\epsilon >T^3\epsilon _0^2T^{}`$. At the same time, the contribution to the detector signal of electrons with energies $`T_0<\epsilon <T^{}`$ is given by $$V_d\epsilon \frac{\lambda _Fl_{ee}^2(T)}{L^3}\frac{\epsilon _F\varphi }{T}.$$ (23) According to Eqs. (11) and (14), the signal decreases with increasing $`L`$ according to a power-law, but not exponentially. This again is essentially a two-dimensional effect (cf. ) and should be well-pronounced in high-mobility samples with sufficiently large $`L`$. Thus, it is possible to create conditions in a two-dimensional electron-gas under which the electron-beam signal is determined rather by a higher-energy quasi-ballistic group of electrons which experience small-angle scattering than by purely ballistic electrons $`\epsilon \epsilon _0`$. ## V Non-linear effects and heating Due to the heating of the electron gas between injector and detector point-contacts for ’high’ excess energies the detector signal consists not only of quasiballistic beam electrons but also of an isotropic signal resulting in a thermovoltage across the detector. This causes the growth of $`V_d`$ for injection energies $`V_i>5`$ mV in our experiments \[Fig. 3 a) and b)\]. The contribution of the thermopower to the detector signal is given by $$\mathrm{\Delta }V_d=S(T)\mathrm{\Delta }T,\mathrm{\Delta }T=TT_0.$$ (24) Here, $`T`$ is the electron gas temperature between injector and detector and $`T_0`$ is the gas temperature beyond the detector (which is close to the lattice temperature), $`S(T)`$ is the Seebeck coefficient (thermopower) of the detector (heating of the 2DEG between injector and detector by the injected electron beam was already discussed by us in Ref. ). As discussed above (Sec. II), on increasing $`V_i`$, the increase of $`T`$ leads to an increase of the thermovoltage on the one hand, and to the decrease of the mean-free-path of quasiballistic electrons on the other hand, and therefore to the appearance of minimum in the dependence $`V_d(V_i)`$. For $`\epsilon _0>\epsilon ^{}`$ (the limit where specific 2D effects can be neglected), only ballistic electrons contribute to the beam signal. We then have as a rough estimate for the temperature dependence of the signal: $$V_d(T)\kappa e^{\frac{L}{l_{ee}(T)}}+S(T)\mathrm{\Delta }T,\kappa ^1L\lambda _F^1\varphi \left(V_i^1+\epsilon _0^1\right).$$ (25) An analysis of this expression shows that a minimum in $`V_d(T)`$ is always present when $`V_i>\epsilon _0`$. This statement holds as well in the multi-collision regime $`\epsilon _0<T<\epsilon ^{}`$ \[see Eq. (14)\]. It is evident that with increasing $`V_i`$ the temperature of the 2DEG between injector and detector point-contact increases. However, this does not necessarily imply that the $`V_d(V_i)`$ dependence replicates $`V_d(T)`$ of Eq. (16) qualitatively, because $`V_i`$ enters explicitly in Eq. (16) and not only through $`T(V_i)`$. One can only state definitely that the beam signal should decrease sooner or later on increasing $`V_i`$. Generally speaking, the theoretical determination of the dependence $`T(V_i)`$ requires solving a complex non-linear problem on the beam’s self-action. However, the essential dependence of the mean-free-path on excess energy allows considerable simplifications for sufficiently high $`V_i`$, i.e., the separation of the injected particles in (i) ”heating” (high-energy electrons which do not reach the detector) and (ii) quasiballistic electrons (which contribute mainly to the beam signal, but not to heating). Such a separation is undoubtedly possible at $`V_i>\epsilon _0>\epsilon ^{}`$. (If $`V_i\epsilon _0`$ we can neglect heating.) Thus, under certain conditions one can use the following quasi-linear approach: First, we find the electron gas heating $`\mathrm{\Delta }T(V_i)`$ due to the high-energy part of the beam, and then, using the electron temperature $`T`$ thus obtained, we determine the signal of the quasiballistic part. In particular, in the relaxation-time approximation we have: $$f=e^{\frac{L}{v_x\tau _{ee}(\epsilon ,T)}}f_0(yL\frac{v_y}{v_x},𝐩),$$ (26) where $`T=T_0+\mathrm{\Delta }T(V_i)`$. In fact, this means that the separation reduces the nonlinear problem to two linear equations. Finally, we want to consider the case when $`\epsilon _0<V_i<\epsilon ^{}`$ where, due to the specific two-dimensional effects, the injected particles slowly relax their directionality, but rapidly lose their excess energy . In this limit, it is not possible to separate heating particles from quasiballistic ones. The beam signal is proportional to $`l_{ee}^2(T)T^1`$ \[see Eq. (14)\] and, hence, it is sensitive to heating. This essentially nonlinear situation could be realized experimentally for high quality samples with a large distance $`L`$ between injector and detector. ## VI Discussion of the experiment In this section we want to compare the experimental results, Sec. II, with our theoretical results. A series of measurements for different sample temperatures, $`T_0=`$ 1.6, 2.2, 3.4, 5.0, 8.0, 11, 15, and 17 K were available for analysis (partly shown in Fig. 3). For temperatures $`T_0<8`$ K, the modified one-collision approximation \[Eqs. (7), (8) and (9)\] allows a proper description of the experiment for $`V_i<3`$ mV, see Fig. 4 (note that $`l_{ee}L=3.4`$ $`\mu `$m for $`V_i=\epsilon _02`$ mV). For higher values of $`V_i`$, where heating is essential, one can use the relaxation-time approximation \[Eq. (17)\], taking into account the $`T(V_i)`$ dependence\]. To compare theory with experiment we extract the heating $`\mathrm{\Delta }T(V_i)`$ in two different ways: First, by measuring the heating caused by the electron beam as a function of injection energy, using the thermovoltage across the detector quantum point-contact as a thermometer and second, by analysing the set of experimental data for the anisotropic part of the signal (Fig. 3), see below. The result of an electron-temperature measurement determined via thermopower for a lattice temperature $`T_0=1.6`$ K is displayed in the inset of Fig. 7. Here, the detector point-contact conductance was adjusted to yield a maximum thermopower ($`S20`$ $`\mu `$V/K) , where the conductance of the injector point-contact was fixed at one mode so that its thermopower is negligible compared with the detector. The measurements were done at small magnetic fields ($`B=50`$ mT) to prevent beam electrons from reaching the detector point-contact directly. Alternatively, the decrease of the detector signal due to the quasiballistic part of the electron beam allows for an estimate of the 2DEG beam heating. We assume that the part of these curves at values $`V_i`$ larger than the injector voltage at the maximum in $`V_d`$ (which we now denote as $`V_i^{\mathrm{max}}`$) describes the signal from the full narrowly-directed fraction of electrons as a function of 2DEG temperature, i.e. $`V_d(T_0+\mathrm{\Delta }T(V_i))`$. Thus, the curves in Fig. 3 are members of a one-parameter family which differ only by the value of the temperature $`T_0`$. For the validity of this statement, it is important that heating can be neglected at the local maximum of $`V_d(V_i)`$ for each curve, $`\mathrm{\Delta }T(V_i^{\mathrm{max}})0`$. From Fig. 3 b) it is evident that at $`V_i^{\mathrm{max}}`$ the thermovoltage is indeed negligible. Let us now consider any two curves $`T_{01}`$ and $`T_{02}`$ of this family and let $`T_{01}<T_{02}`$. Then curve $`T_{01}`$ has always larger values $`V_d`$ for a given $`V_i`$ than curve $`T_{02}`$, and curve $`T_{01}`$ decreases to a signal $`V_d`$, equal in size to the maximum of curve $`T_{02}`$ at a given, larger, value of $`V_i`$. Now, it is evident that $`T_{01}+\mathrm{\Delta }T(V_i)=T_{02}`$. In this manner, we are able to reconstruct the function $`\mathrm{\Delta }T(V_i)`$. Let us emphasize, that it is convenient to choose the local maximum $`V_i^{\mathrm{max}}`$ as a starting point for recovering $`\mathrm{\Delta }T(V_i)`$, since, in the vicinity of this point there is no need (i) to correct for the increase of the signal due to the quasiballistic group of electrons with increasing $`V_i`$, that takes place at low $`V_i`$ in the linear response regime, and (ii) to take heating effects into account. As an example, let us consider the curves 1 and 2 of Fig. 3, i.e., $`T_{01}=1.6`$ K and $`T_{02}=8`$ K. The maximum value of curve 2 is approximately 0.135 mV. The same value of $`V_d`$ for curve 1 is reached in decreasing part of the curve at $`V_i5.2`$ mV. Thus, we obtain a heating temperature of $`\mathrm{\Delta }TT_{02}T_{01}=6.4`$ K for $`V_i5.2`$ mV. The results obtained for the experiments at the lowest sample temperature $`T_0=1.6`$ K are shown in the inset of Fig. 7 (squares). It can be seen that the extracted heating temperatures agree well with 2DEG temperature measurements for applied magnetic fields. We therefore can use this $`T(V_i)`$-dependence for further considerations. Note that the heating temperature depends not only on $`V_i`$ but also on the initial sample temperature $`T_0`$: $`\mathrm{\Delta }T=\mathrm{\Delta }T(V_i,T_0)`$. For higher $`T_0`$, the 2DEG heating is less efficient, cf. Fig. 3. Additionally, the electron temperature can be estimated roughly from the heat-balance between the energies transfered from the electron beam into the 2DEG and removed by phonons . We then have $$v_F<\epsilon >\lambda _F\frac{V_i}{\epsilon _F}n_e\nu _{ep}s\mathrm{}k_F\frac{T}{\epsilon _F}n_e\mathrm{\Sigma }.$$ (27) Here, $`<\epsilon >V_i/2`$ is the average energy of the electron beam, $`n_eV_i/\epsilon _F`$ is the number of injected electrons, $`\nu _{ep}`$ is the frequency of electron-phonon collisions, $`s`$ is the sound velocity, $`k_F`$ is the Fermi wave vector, and $`\mathrm{\Sigma }`$ is the area of the heated 2DEG region. Thus, we obtain for the electron temperature $$T\frac{\lambda _Fl_{ep}}{\mathrm{\Sigma }}\frac{\epsilon _F}{s\mathrm{}k_F}<\epsilon >.$$ (28) In order to evaluate this expression we assume for the experimental situation the following values: $`s6\times 10^5`$ cm s<sup>-1</sup>, $`k_F=1.18\times 10^6`$ cm<sup>-1</sup>, $`\epsilon _F=9`$ mV, and $`\mathrm{\Sigma }`$ is taken to be of the order of the area between injector and detector, viz. 200 $`\mu `$m<sup>2</sup>. The mean-free-path for electron-phonon collisions $`l_{ep}`$ is estimated at 100 $`\mu `$m, yielding $`T(V_i=5.2`$ mV) $`7.3`$ K \[Eq. (19)\]. In spite of this very crude model, we thus find a remarkable agreement with the electron temperature obtain from the experimental data ($`\mathrm{\Delta }T(V_i5.2`$ mV$`)=6.4`$ K). As mentioned above, the experimental data can be approximated using the modified one-collision approximation \[Eq. (7)\] for injection energies $`V_i<V_i^{\mathrm{max}}`$ and the relaxation-time approximation \[Eq. (17)\] for $`V_i`$ which is sufficiently large in comparison with $`V_i^{\mathrm{max}}`$. At high $`T(V_i)`$ the scattering is not small-angular and leads to a more or less isotropic background, i.e. 2D effects are absent. According to this, we plotted in Fig. 7 the modified one-collision approximation (dashed line), experiment (solid line) for $`T_0=1.6`$ K and the relaxation-time approximation (RTA) (dotted line), which takes into account electron-heating effects. For the RTA we have use the asymptotic expression $$\tau _{ee}^1(\epsilon ,T)=\frac{\epsilon ^2+2\pi ^2T^2}{4\pi \mathrm{}\epsilon _F}\mathrm{ln}\frac{\epsilon _F}{T+\epsilon },$$ (29) which is valid for arbitrary ratio of small values of $`T/\epsilon _F`$ and $`\epsilon /\epsilon _F`$ . The coefficients for the theoretical calculations were chosen in such a way that coincidence is achieved for small injection energies, where the linear increase is observed and electron travel ballistically in any case. Thus, in fact, no additional fitting parameters are used. For this calculation, it is important that the detector size $`\lambda _FL\varphi `$. As one can see from Fig. 7, the divergence between the relaxation-time approximation and the experiment is only due to specific two-dimensional effects and reaches a maximum in the vicinity of $`V_i=3`$ mV, where on the one hand a number of the scattered particles is comparable with the number of pure ballistic, and on the other hand scattering is still small-angular. ## VII Conclusions We have studied the role of different groups of electrons on the propagation of electron beams in a high-mobility two-dimensional electron-gas for wide range of excess energies. We have observed a non-monotonic dependence of the detector signal on the excess energy of the injected electrons. This result can be explained in the framework of our model, separating the beam electrons into two groups, i.e. ”quasiballistic” electrons and ”heating” electrons (high-energy part of a beam). We have shown that due to the reduced dimensionality of the system the quasiballistic fraction consists not only of purely ballistic electrons but also of a significant number of electrons which have experienced small-angle electron-electron scattering events. The small-angle character of electron-electron scattering is essentially a two-dimensional effect, predicted earlier by us , which manifests itself in the experiments discussed here. In addition, we have formulated the conditions where 2D effects can be best observed and thus electron-beam propagation over very long distances should be possible. ## ACKNOWLEDGMENTS This work was supported by the Volkswagen-Stiftung (Grant No. I/72 531), and by the DFG MO 771/1-2.
warning/0002/nucl-th0002063.html
ar5iv
text
# Single-Particle and Collective Motion for Proton-Rich Nuclei in the Upper 𝑝⁢𝑓 Shell \[ ## Abstract Based on available experimental data, a new set of Nilsson parameters is proposed for proton-rich nuclei with proton or neutron numbers $`28N40`$. The resulting single-particle spectra are compared with those from relativistic and non-relativistic mean field theories. Collective excitations in some even–even proton-rich nuclei in the upper $`pf`$ shell are investigated using the Projected Shell Model with the new Nilsson basis. It is found that the regular bands are sharply disturbed by band crossings involving $`1g_{9/2}`$ neutrons and protons. Physical quantities for exploring the nature of the band disturbance and the role of the $`1g_{9/2}`$ single-particle are predicted, which may be tested by new experiments with radioactive beams. \] The nuclear shell model has been successful in the description of nuclear structure. Thanks to the increasing power of computation, exact diagonalizations in the full $`pf`$ shell has become possible in recent years . An immediate application has been seen in nuclear astrophysics , where knowledge about detailed nuclear structure is important in understanding the nuclear processes that govern those violent astrophysical phenomena such as nova and supernova explosions, or X-ray bursts. Nuclear structure information is thought to be important also in the study of the nuclear processes occurring on the astrophysical rapid proton capture or rp-process , which may be relevant to nova explosions or X-ray bursts. The rp-process path lies close to the proton drip line in the chart of nuclides. There, compound nuclei are formed at very low excitation energies and therefore at low level densities, which does not justify Hauser-Feshbach calculations. Thus, detailed nuclear structure information is required when studying the nuclear processes. One hopes that radioactive beams will provide us with the information eventually, but one has to rely on theoretical calculations at present. To obtain detailed nuclear structure, advanced shell model diagonalization methods, which can give explicitly spectroscopy and matrix elements for all kinds of nuclei (even-even, odd-A and odd-odd), are of particular importance. The Projected Shell Model (PSM) is a shell model diagonalization carried out in a projected space determined by a deformed Nilsson–BCS basis. This kind of shell model truncation is highly efficient if the single-particle (SP) basis is realistic because the basis already contains many correlations . It has been shown that the PSM can describe the spectra and electromagnetic transitions in normally deformed , superdeformed , and transitional nuclei . One can further calculate the nuclear matrix elements for astrophysical processes such as direct capture and decay rates. One advantage of the PSM is that it can easily handle heavy, well-deformed nuclear systems. This can be important for the structure study of significantly deformed nuclei ($`Z=3640`$) on the rp-process path, for which the current large-scale shell model diagonalizations are not feasible. As an initial step for a PSM study, a reliable Nilsson model calculation is required to build the projected space. The Nilsson model has been used widely in nuclear structure studies. Its “standard” set of parameters $`\kappa `$ and $`\mu `$ has been quite successful in describing the SP structure for stable nuclei across the whole chart of nuclides. However, previous work has shown that the standard Nilsson SP energies are no longer realistic for the neutron-rich region . Thus, an adjustment of these parameters is necessary for unstable nuclei. Experimental SP states above the $`N=28`$ shell closure can be read from the low-lying states of <sup>57</sup>Ni and <sup>57</sup>Cu , the nearest neighbors of the doubly magic nucleus <sup>56</sup>Ni. These low-lying $`2p_{3/2}`$, $`1f_{5/2}`$ and $`2p_{1/2}`$ states are well characterized as pure SP levels . According to Ref. , the neutron $`1g_{9/2}`$ state lies 3.7 MeV above the $`2p_{3/2}`$ orbital. For the proton $`1g_{9/2}`$ state, the best experimental information available is the observation of a low-lying $`\frac{9}{2}^+`$ level in <sup>59</sup>Cu and <sup>61</sup>Cu . We may reasonably assume this level to be the $`1g_{9/2}`$ state because of its unique parity. Taking the deformation of <sup>59,61</sup>Cu into account, the position of the proton SP $`1g_{9/2}`$ state can be estimated to lie 3.15 MeV above the $`2p_{3/2}`$ orbital. When compared with these data, the standard Nilsson parameterization produces energies for the $`1f_{5/2}`$, $`2p_{1/2}`$ and $`1g_{9/2}`$ orbits that are too high relative to the $`2p_{3/2}`$ orbit. To reproduce experimental data, one must reduce the strength of the spin–orbit interaction $`\kappa `$ for $`N=3`$ substantially from the standard value because of the observed smaller separation between $`2p_{1/2}`$ and $`2p_{3/2}`$ levels, and because of the position of $`1f_{5/2}`$ orbital. On the other hand, the pair of $`g`$-orbitals require a larger value of $`\kappa `$ for $`N=4`$ to position the $`1g_{9/2}`$ orbital properly. In Table I, we summarize the adjusted proton and neutron $`\kappa `$ and $`\mu `$ parameters for the $`N`$ = 3 and 4 shells that best reproduce the available data. The standard values for the $`N`$ = 2 shell are also displayed without modification. Relativistic Mean Field (RMF) theory with nonlinear self-interactions between mesons has been used in many studies of low-energy phenomena in nuclear structure. It has been extended recently to allow coupling between bound states and the continuum by the pairing force . In the RMF theory, the spin-orbit interaction arises naturally as a result of the Dirac structure of the nucleons. As discussed above, the spin–orbit interaction is one of the important factors to give a correct SP energies. Thus, it is relevant to consider the relation of the SP Nilsson spectrum to that of the RMF. Here, two typical interactions, NL1 and NL3, are used. The latter interaction is suitable also for nuclei away from the $`\beta `$-stability valley . For comparison, SP energies from the non-relativistic Hartree-Fock calculations with the Skyrme interactions (SHF) (see Ref. and references therein) are also presented. In Fig. 1, theoretical SP states (from the new and the standard Nilsson parameterizations, from SHF with SkM and SIII forces, and from RMF with NL1 and NL3 interactions) are compared with data. It is obvious that the standard Nilsson parameters produce a large spin-orbital splitting for $`2p_{1/2}`$ and $`2p_{3/2}`$, and high excitation energy for $`1f_{5/2}`$ while the new set of parameters reduces these values. The SHF and RMF calculate the SP states for <sup>56</sup>Ni. Without specific parameter adjustment for this mass region, SP levels of the RMF with NL3 for the $`p`$ and $`f`$ orbitals are found to be reasonably close to the data. However, the SHF with the SkM force produces rather wrong positions for all levels considered. In all the SHF and RMF calculations, the positions of both neutron and proton $`1g_{9/2}`$ orbital are much too high. The new Nilsson parameters should represent a better basis from which more sophisticated wave functions can be constructed. Therefore, we may test the new parameters by employing them in calculations that have a direct connection with measured collective spectra. For such a test, we shall employ the PSM to calculate the yrast bands of some even–even nuclei for which limited data are available for comparison: <sup>62,66</sup>Zn and <sup>64,66</sup>Ge. These nuclei lie on the rp-process path, and <sup>64</sup>Ge is a waiting point nucleus that was used as a test case for the recently proposed Quantum Monte Carlo Diagonalization Method (QMCD) . As we shall see, the new Nilsson SP states discussed above can modify substantially the level spacings, position and sharpness of band crossing, and electromagnetic transition properties along a level sequence. In PSM calculations of the collective states relevant here, the projected multi-quasiparticle (qp) space consists of 0-, 2-qp and 4-qp states for even–even nuclei, typically with a dimension of 50. This small shell model basis is sufficiently rich that the quality of the calculations is comparable to those from large-scale shell model diagonalizations . For the SP basis we use three full major shells: $`N`$ = 2, 3, and 4 for both neutrons and protons. This is the same size SP basis as employed in Ref. . The deformation parameters are taken from Ref. as follows: $`\epsilon _2=0.167`$ and $`\epsilon _4=0.020`$ for <sup>60</sup>Zn, $`\epsilon _2=0.192`$ and $`\epsilon _4=0.013`$ for <sup>62</sup>Zn, $`\epsilon _2=0.200`$ and $`\epsilon _4=0.047`$ for <sup>64</sup>Ge, and $`\epsilon _2=0.208`$ and $`\epsilon _4=0.067`$ for <sup>66</sup>Ge. The Hamiltonian is the usual quadrupole-quadrupole plus monopole pairing form, with quadrupole pairing included . The strength of the quadrupole-quadrupole force in the Hamiltonian is determined self-consistently, and the monopole pairing strength is the same as that in Ref. . For the four nuclei calculated in this paper, the ratio of quadrupole pairing to monopole pairing strength is fixed at 0.30. In Fig. 2, yrast band (lowest energy state for given spin) transitional energies $`E(I)E(I2)`$ are plotted as a function of spin $`I`$. It is obvious that the calculations employing the new set of Nilsson parameters reproduce the data very well, while those with the standard set of Nilsson parameters determined in the stability valley are in poorer agreement. In all the four nuclei, the standard set of Nilsson parameters gives too large level spacings for low-spin states, leading to the excitation energies that are too high. Note that the two calculations are performed with the same conditions except for different SP bases. It is the change in SP states that gives rise to the different results for the yrast spectrum. Following the yrast band in a nucleus in Fig. 2, one observes a sudden drop in $`E(I)E(I2)`$ at spin $`I=8`$ or 10, which corresponds to a backbending in the moment of inertia for the system . For these $`NZ`$ nuclei, neutron and proton Fermi levels are surrounded by orbits with the same Nilsson quantum numbers. Therefore, bands built on the neutron and proton $`1g_{9/2}`$ intruder orbits can have a similar probability to be the first that crosses the ground band and becomes a major part of the yrast band. From analysis of the wave functions, we find that the sudden drop is caused by such band crossings. The crossing bands have either 2-neutron $`1g_{9/2}`$ or 2-proton $`1g_{9/2}`$ configurations. Effects of the band crossing can be seen more clearly by looking at the reduced transition rates B(E2). In the B(E2) calculations, the effective charges used in this paper are 0.5e for neutrons and 1.5e for protons, which are the same as those used in previous work and in other shell models , and similar to those in the QMCD. For a comparison, our results for the first two transitions in <sup>64</sup>Ge (see Fig. 3a) are close to those obtained in the QMCD (B(E2; $`2^+0^+)=0.050(e^2b^2)`$ and B(E2; $`4^+2^+)=0.065(e^2b^2)`$). We emphasize that employment of different effective charges can modify the absolute values, but the essential spin dependence is determined by the wave functions. In Fig. 3a, the sudden drop in the B(E2) values at spin $`I=810`$ is consequence of band crossings discussed above. The drop indicates that the transition rate is sharply reduced by the structure change in the yrast band wave function. In , states in <sup>64</sup>Ge were calculated up to $`I=4`$. It would be an important comparison if the QMCD results could be extended beyond that spin. It is interesting to point out that our prediction for the crossings occurs exactly at the excited states where energy spectra measured to date have terminated (see Fig. 2). Thus the band crossing predictions may explain why the experimental measurements have not seen higher spin states. The gyromagnetic factor (g-factor) is the quantity most sensitive to the SP components in wave functions as well as to their interplay with collective degrees of freedom. Because of the intrinsically opposite signs of the neutron and proton $`g_s`$, a study of g-factors enables determination of the microscopic structure for underlying states. For example, variation of g-factors often is a clear indicator for a SP component that strongly influences the total wave function. In the calculations we use for $`g_l`$ the free values and for $`g_s`$ the free values damped by the usual 0.75 factor. The results are presented in Fig. 3b. Rather similar behavior is seen for all the four nuclei for spin states before band crossing: The g-factor values are close to the collective value of $`Z/A`$ and tend to increase slightly as a function of spin. However, the pattern diverges at band crossing. The g-factors of the two $`N=Z`$ nuclei (<sup>60</sup>Zn and <sup>64</sup>Ge) jump suddenly to a value near 1, while those of the two $`N=Z+2`$ nuclei (<sup>62</sup>Zn and <sup>66</sup>Ge) drop to zero. By analysis of the wave functions, we have found that this interesting behavior is determined by whether the proton or neutron $`1g_{9/2}`$ 2-qp states is lower in energy after band crossing. For $`N=Z`$ nuclei, proton $`1g_{9/2}`$ 2-qp states dominate the wave functions after band crossing, thus increasing the g-factors. For $`N=Z+2`$ nuclei, neutron $`1g_{9/2}`$ 2-qp states become lower due to different neutron shell fillings, leading to very low values of the g-factors. Similar band crossing pictures should be common for this mass region, and may also be seen in the neighboring odd-A and odd-odd nuclei. Successful measurement of the B(E2) values and g-factors before and after band crossing would test our PSM predictions as well as the Nilsson SP states proposed in this paper. We hope that recently developed techniques in combination with radioactive beams can permit such measurements. In summary, a new set of Nilsson parameters is proposed for proton-rich nuclei beyond <sup>56</sup>Ni. This new set of parameters can be employed for proton-rich nuclei with proton and number numbers $`N=2840`$. Our results suggest that improved SP structures may be necessary for this mass region in SHF and RMF theories. Available data for yrast bands in some even–even nuclei on the rp-process path are well reproduced by PSM calculations with the new set of Nilsson states as a basis. Distinct signatures for the role of neutron and proton $`1g_{9/2}`$ orbitals are seen in the band crossings and the electromagnetic transitions along the yrast bands. We conclude that precise positions of the $`1g_{9/2}`$ SP orbitals are very important for any realistic calculations. Predictions of the present paper concerning the $`1g_{9/2}`$ band crossings, B(E2) values, and the remarkably different behaviors of the g-factors at the crossings await future experimental tests. Research at the University of Tennessee is supported by the U. S. Department of Energy through Contract No. DE–FG05–96ER40983. This work was partially sponsored by the National Science Foundation of China under Project No. 19847002 and 19935030, and by SRF for ROCS, SEM, China.
warning/0002/hep-th0002155.html
ar5iv
text
# Contents ## Chapter 1 INTRODUCTION ### 1.1 Mathematics Imaginary numbers appeared in mathematics a long time ago. For example, Nicolas Chuquet (1445–1500) wrote “Triparty en la science des nombres” where he introduced an exponential notation, allowing positive, negative and zero powers. He showed that some equations lead to imaginary solutions but rejected them “tel nombre est ineperible”. Geronimo Cardano (1501–1576) wrote “Ars magna”, found solutions to polynomials which lead to square roots of negative quantities but also rejected them “as subtle as it is useless”. The first to consider imaginary numbers was Rafael Bombelli (1530–1590) who published “Algebra” and proposes the “wild idea” that one can use these square roots of negative numbers to get to the real solutions by using conjugation. Albert Girard (1595–1632) publishes “Invention nouvelle en l’algebra” retaining all imaginary roots because they show the general principles in the formation of an equation from its roots. Rene Descartes (1596–1650) coins the term “imaginary” for terms involving square roots of negative numbers but takes their existence as a sign that the problem is insoluble. Reviving some speculation, Gottfried Leibniz (1646–1716) says that imaginary numbers are halfway between existence and nonexistence. Sustaining algebra by geometry, John Wallis (1616–1703) was the first to represent complex numbers geometrically in his book “Algebra” published in 1673. Roger Cotes (1682–1716) deduces that $`\mathrm{exp}\left(\sqrt{1}a\right)=\mathrm{cos}\left(a\right)+\sqrt{1}\mathrm{sin}\left(a\right)`$ but his result was largely ignored. But then a new era begins, it was Leonhard Euler who brought complex numbers from the shadow to the daylight, he invents the symbol $`i`$ for $`\sqrt{1}`$ and works extensively with imaginary numbers, for example, he shows that a complex number to the power of a complex number is also a complex number. Jean d’Alembert’s (1717–1783) constructs functions of complex variables, obtaining what later is called the Cauchy–Riemann equation. Caspar Wessel (1745–1818) discovers that complex numbers can be represented graphically on a two dimensional plane, what we now call the “Argand” or “Guassian” representation of complex numbers. Modern complex analysis may be dated to the book of Augustin Cauchy (1789–1857) “Memoire sur les integrales definies, prises entre des limites imaginaires” which contains his integral theorems on residues. Then the work of Augustus de Morgan (1806–1871) and Carl Gauss (1777–1855) opens the way to what later becomes complex numbers analysis. So finally, what was rejected as useless quantities become the heart of mathematics. While the discovery and acceptance of complex numbers took a long time, the histroy of quaternions and octonions is much shorter. Quaternions were discovered by a single man , William Hamilton (1805–1865). Trying to generalize his “Theory of Algebraic Couples”, where he constructs a rigorous algebra of complex numbers as number pairs for the first time, he identifies $`x+iy`$ with its $`^2`$ coordinates $`(x,y)`$. After many years of trial and error, Hamilton discovers quaternions on Monday 16 October 1843 and defines a vector subspace $`ai+bj+ck`$ by elements which may be interpreted as an $`R^3`$ coordinate system $`(a,b,c)`$ but $`i,j,k`$ are not commutative. As early as 1845, Octonions were introduced by Arthur Cayley and John Graves independently . Quaternions and octonions may be presented as a linear algebra over the field of real numbers $``$ with a general element of the form $$Y=y_0e_0+y_ie_i,y_0,y_i$$ (1.1) where $`i=1,2,3`$ for quaternions $``$ and $`i=\mathrm{1..7}`$ for octonions $`𝕆`$. We always use Einstein’s summation convention. The $`e_i`$ are imaginary units, for quaternions $`e_ie_j`$ $`=`$ $`\delta _{ij}+ϵ_{ijk}e_k,`$ (1.2) $`e_ie_0`$ $`=`$ $`e_0e_i=e_i,`$ (1.3) $`e_0e_0`$ $`=`$ $`e_0,`$ (1.4) where $`\delta _{ij}`$ is the Kronecker delta and $`ϵ_{ijk}`$ is the three dimensional Levi–Cevita tensor, as $`e_0=1`$ when there is no confusion we omit it. Octonions have the same structure, only we must replace $`ϵ_{ijk}`$ by the octonionic structure constant $`f_{ijk}`$ which is completely antisymmetric and equal to one for any of the following three cycles $$123,\mathrm{\hspace{0.33em}145},\mathrm{\hspace{0.33em}176},\mathrm{\hspace{0.33em}246},\mathrm{\hspace{0.33em}257},\mathrm{\hspace{0.33em}347},\mathrm{\hspace{0.33em}\hspace{0.33em}365}.$$ (1.5) The important feature of real, complex, quaternions and octonions is the existence of an inverse for any non-zero element. For the generic quaternionic or octonionic element given in (1.1), we define the conjugate $`Y^{}`$ as an involution $`\left(Y^{}\right)^{}=Y`$, such that $$Y^{}=y_0e_0y_ie_i,$$ (1.6) introducing the norm as $`N\left(Y\right)Y=YY^{}=Y^{}Y`$ then the inverse is $$Y^1=\frac{Y^{}}{Y}.$$ (1.7) The Norm is nondegenerate and positively definite. We have the decomposition property $$XY=XY$$ (1.8) $`N\left(xy\right)`$ being nondegenerate and positive definite obeys the axioms of the scalar product and our algebra is called a normed algebra. The uniqueness and beauty of real, complex, quaternionic and octonionic numbers stem from Hurwitz’ theorem : * Each normed composition algebra with a unit element is isomorphic to one of the following algebras: to the algebra of real numbers, to the algebra of complex numbers, to the quaternion algebra or to the octonion algebra. Another important mathematical property of the ring algebra is the following: For the set $`𝒮`$defined by $$𝒮=\left\{X\right|X=1\}$$ (1.9) where $`X`$ is a ring division element then from the decomposition property (1.8), we have a closure structure $$X,Y𝒮Z=XY𝒮$$ (1.10) even for octonions which do not admit a group structure (group is defined for associative algebra). This beautiful closure can be extended to a generic ring division element by scaling. For any two generic ring division elements $`W,V`$ we construct $$\stackrel{~}{W}=\frac{W}{W},\stackrel{~}{V}=\frac{V}{V}$$ (1.11) hence $$\stackrel{~}{W}=\stackrel{~}{V}=\stackrel{~}{W}\stackrel{~}{V}=1.$$ (1.12) A geometric meaning of this closure is the parallelizability of ring division spheres. For $`X=1`$ $$\begin{array}{ccccccc}\hfill & & x_0^2+x_1^2\hfill & =1\hfill & defines\hfill & aunit\hfill & S^1\hfill \\ \hfill & & x_0^1+x_1^2+x_2^2+x_3^2\hfill & =1\hfill & defines\hfill & aunit\hfill & S^3\hfill \\ 𝕆\hfill & & x_0^1+x_1^2+x_2^2+x_3^2+x_4^2+x_5^2+x_6^2+x_7^2\hfill & =1\hfill & defines\hfill & aunit\hfill & S^7\hfill \end{array}$$ (1.13) the parallelizability means that there is such an $`X`$ that defines *globally* $`1,3,7`$ vector fields for $`S^1,S^3,S^7`$ respectively . Another importance of ring division algebras is their relevance to the classification of real Clifford algebras $`Cliff(m,n).`$ A task that had been achieved by Atiyah, Bott and Shapiro . Their appearance is clear through the Bott periodicity. We also wish to mention the work of Milnor<sup>1</sup><sup>1</sup>1According to legend, Milnor presented his first acheivements as an assignment. On one occasion, he was late for the class of Fox (the Father of the american Knot theory). During that lesson, Fox explained his way of doing research. Usually he writes the most difficult 10 questions and tries to solve any one of them. As an example, he wrote for his students the 10 questions that kept him busy at that time. Fortunatley, John Milnor came late that day, he saw the quetions, he thought they were homework. After the class he worked hard untill the next morning. Then before the class of the next day, he approached Fox expressing his desire to change his field of research because he only managed to solve one of the 10 questions. Fox was totally surprised. But then later Milnor continued to surprise the world especially by his Field Medal’s diffeomorphic non-homeomorphic structure., where for the first time in history a diffeomorphic non homeomorphic structure was found. There are 28 of such structures over $`S^7`$, the fake $`S^7`$ . As a matter of fact, these fake $`S^7`$ are the higher bundles of the four dimensional $`SO\left(4\right)`$ instanton solutions. Donaldson received in 1986 the Field Medal for his work about the infinite diffeomorphic non-homeomorphic $`^4`$. An idea that he got by carefully studying the space of solutions of the four dimensional $`n=1`$ “quaternionic” instanton. Finally, ring division algebras have a connection with homotopies, Hopf fibrarion and many other interesting topics. It is clear that the history of these other elements of the ring division algebra, quaternions and octonions, are much shorter than the complex one. Maybe, it has not been yet fully written. With the hindsight of complex numbers, these new hypercomplex numbers were immediately accepted, perhaps only the question of their utility in physics is still to be discovered and may yet involve much discussion and take a long time. ### 1.2 Physics Non associative algebra appeared for the first time in physics when Jordan, van Neuman and Wigner introduced commutative but non-associative operators – Jordan algebras – for the construction of a new quantum mechanics. More recently ,after the proposed eightfold way by Gell-Mann and Ne’eman, there were some octonionic rivals for $`SU\left(3\right)`$ such as $`G_2,SO\left(7\right),SO\left(8\right)`$ and others. Possible octonionic internal symmetries were considered by different people such as Souriau and Kastler , Pais , Tiomno , Gamba , and Penny . The inclusion $`SU\left(3\right)G_2SO\left(7\right)SO\left(8\right)`$ lead many authors to consider the relationship between the nuclear strong field and octonions. There are a lot of papers dealing with the $`SO\left(8\right)`$ symmetry . The potential significance of non-associative algebras for the generalization of classical dynamics was pointed out by Nambu . Generalizing the Liouville theorem, about the conservation of the phase space volume, Nambu introduced a new generalization of Poisson brackets which may be interpreted as an associator. Nambu in his paper made some remarks about non-associative algebras, octonions and Jordan algebras. As another application of octonions, allowing parastatistics and the paraquark model, Freund showed in that color gauging is only possible in the octonionic realization of the paraquark model. Octonions have even been applied to gravitation. Vollendorf constructed a bilocal field theory with the group $`SO(4,4)`$ instead of the Lorentz group and with a system of 24 coupled differential equations to explain the five conservation laws of charge, hypercharge, baryon number and the two lepton number then known. From the early seventies and up to the present time, octonions have been applied with some success to different important problems such as quark confinement, grand unified models (GUT). A serious step was taken by Günaydin and Gürsey when they present in their work a systematic study of the octonionic algebraic structure. If we follow the theory of observable states developed by Birkhoff and van Neumann , we can only have observable states in Hilbert spaces over associative normed algebras. The standard quantum mechanics explores the Hilbert space over complex numbers. The quaternion case with a quaternionic scalar product was developed by Finkelstein, Jauch and Emch . The general hope was to introduce isospin degrees of freedom by enlarging the quantum Hilbert space. But Jauch proved that quaternionic representations of the poincaré group did not generate any new states. The reader is referred to the book of Adler for a modern formulation of quaternionic field theory . Another interesting idea is the use of hypercomplex Hilbert space but with a complex scalar product, Goldstein, Horwitz and Biedenharn , considered octonionic Hilbert space with a complex scalar product as early as 1962. The abandon of associativity means the existence of nonobservable states. Günaydin and Gürsey used these nonobservable states to explain the quark confinement phenomena. Later on octonions entered the Grand Unified Theories era with different applications. In , Gürsey suggested that the exceptional group $`F_4`$ might be used to describe the internal charge space of particles. Even the other exceptional groups $`E_6,E_7,E_8`$ have been utilized to provide larger GUT models $$SU(3)_c\times SU(2)_L\times U(1)_YSU(5)SO(10)E_6E_8.$$ (1.14) As a matter of fact, nowadays, $`E_6\times U\left(1\right)`$ is the most promising scheme from the superstrings phenomenological point of view. Günaydin constructed an exceptional realization of the Lorentz group in . The exceptional $`SL(2,)`$ multiplets generate a non-associative algebra. Exceptional supergroups were also introduced and investigated in . Starting from the eighties, new applications of ring division algebras in physics were found. The instanton problem, supersymmetry, supergravity, superstrings and recently branes technology. We give references in the appropriate chapters of this thesis to the history of the first two topics. Application of quaternions and octonions to supergravity spontaneous compactification was a very important and active field of research during the mid eighties. Especially compactification of $`d=11`$ supergravity over $`S^7`$ to 4 dimensions. It is an impossible task to list all the relevant papers, so we direct the interested reader to the physics report written by Duff, Nilsson and Pope where a lot of references are given. We just mention that the first indication of the octonionic nature of this problem appeared in the Englert solution of $`d=11`$ supergravity compactification over $`S^7`$ and a systematic study along this line has been carried out in . The relations between superstrings (p-branes) and octonions had been considered from many different points of view, the reader may consult the references given in for details. ### 1.3 Outline of The Thesis Superstrings promise the possibility of unifying all the fundamental forces of nature. Abandoning the idea of point particle seems necessary to incorporate general relativity with renormalizable field theory. One of the puzzles of this string program is its double facet . On the one hand, we can work over the two dimensional string sheet where we can use the powerful methods of conformal field theory. On the other hand working with the ten dimensional space-time Green–Shwarz (GS) supersymmetric action, non–perturbative effects may be seen. Understanding the relations between these two different formalisms of string theory is important but many features of the GS formalism still have to be elucidated. For example, the the algebra is given on–shell and the action exists only in certain space–time dimensions, $`d=3,4,6,10`$. Only for ten dimensions, the quantum anomalies of the model cancel i.e. $`d=10`$ is a very special case of an already a very restrictive class. The relationship between gamma matrices needed for the existence of the GS picture is the same as that used to prove the existence of simple supersymmetric Yang-Mills models (containing only a gauge field and a spinor) in the same dimensions $`d=3,4,6,10`$ . A complete comprehension of this fact is important. At the quantum level, we know how to proceed in a perturbative fashion using Feynman diagrams. Starting from quantum chromodynamics QCD, and its quark confinement problem, theoretician searched for non–perturbative phenomena. Self–dual Yang–Mills solutions, *instantons* , can never be evaluated using perturbation theory. It is widely believed that in superstrings, non–perturbative solutions are of great significance and they may give interesting phenomenological applications that can be tested experimentally. In short, off–shell formalism of SSYM and self–duality are very critical topics which may improve our theoretical knowledge and can be used as toy models for testing new approaches. The object of this thesis is to investigate in a systematic way the relations between ring division algebras, off–shell SSYM and higher dimensional self–duality. The starting point is to understand how ring division algebras are specific representations of Clifford algebras. We present this analysis in chapter II. For complex numbers the discussion is simple. Even for quaternionic numbers, once the non-commutativity is taken into account, the formulation can be fully analyzed and understood easily. As octonions are non-associative numbers, we need to work harder to clarify many different subtleties. In chapter II we concentrate upon the Clifford structure that can be extracted from octonions. To make the picture clearer, we show what will happen if we go beyond the ring division algebras limit. The Clifford structure is no longer faithful. In chapter III, we continue our study of octonions, we show that they are endowed with additional useful characteristics. Fixing the direction of action, octonions exhibit soft Lie algebra properties which we call a soft seven sphere . Soft Lie algebras are elements that close under the action of the commutator with structure functions of coordinates system that parameterize a hidden space (the gauge manifold). We study this scheme in full details, we compute the structure functions explicitly with different degrees of complication. In Chapter IV, we start to investigate the physical applicability of the soft seven sphere. The self duality conditions play a fundamental role for any non–perturbative effects in point particle or string field theory . The higher dimensional self dual constraints have properties distinct from the four dimensional one. As a first exercise in the use of the soft seven sphere, we show how to reformulate a quartic eight dimensional self duality condition into a quadratic form. Thus, we put the Grossman–Kephart–Stasheff condition (GKS) into a form much similar to the four dimensional equation. In Chapter V, we will discuss supersymmetry. In particular, the off-shell simple supersymmetric Yang–Mills models SSYM. We shall show that important characteristics can be seen clearly only by using the ring division approach. The ten dimensional case will be very special. For example, we recover Berkovits formulation for the $`d=10`$ off–shell SSYM in a very transparent way. ## Chapter 2 Hypercomplex Structures The starting point of this chapter is to know how to translate some real $`n\times n`$ matrices $`(n)`$ to their corresponding complex, quaternionic and octonionic representations ( shaeffer bimodule representation for octonions ). It is well known from a mathematical point of view that any $`^{2n}`$ is trivially a $`^n`$ complex manifold and any $`^{4n}`$ is also a trivial quaternionic manifold $`^n`$. Furthermore, any $`^{8n}`$ is a trivial $`𝕆^n`$ octonionic manifold, in the sense that the seven sphere can always be embedded in $`^8`$. As any $`^n\times ^n`$ is isomorphic as a vector space to the space of $`n\times n`$ matrices $`(n)`$ , we would expect $`(n)\times (n)`$ $``$ $`(2n);`$ (2.1) $`(n)\times (n)`$ $``$ $`(4n);`$ (2.2) $`𝕆(n)\times 𝕆(n)`$ $``$ $`(8n).`$ (2.3) Even if in this thesis we only work with matrices, there is a hidden geometric and topological underlying structure behind this algebraic construction. Any hypercomplex manifold has a well defined local hypercomplex structure that can be put into the matrix form that we shall develop in this chapter. Lifting this local hypercomplex structure to a global one is not always possible. It amounts to dividing the manifold into local patches where the almost structure is well defined and gluing together these different patches insuring the existence of a (differentiable) structure function that transfers the local hypercomplex structure from one patch to another. If this can be achieved over all the manifold then our space admits a global hypercomplex structure. From the geometric point of view, one should prove the vanishing of the “Nehijinus tensors” . From the topological point of view, one should overcome global obstructions. The story is very similar to the existence of spinorial manifolds. Actually, our almost hypercomplex structures, when represented as matrices, close as Clifford algebra over certain Euclidean spaces. ### 2.1 Complex Structure For complex variables, one can represent any complex number z as an element of $`^2`$ $$z=z_0e_0+z_1e_1Z=\left(\begin{array}{c}z_0\\ z_1\end{array}\right).$$ The action of $`e_0=1`$ and $`e_1`$ induce the following matrix transformations on Z , $$e_0z=ze_0=z𝔼_0Z=\mathrm{𝟏}_2Z=Z,$$ where $`\mathrm{𝟏}_n`$ will always mean the $`n\times n`$ identity matrix, while $`e_1z`$ $`=`$ $`ze_1=z_0e_1z_1`$ (2.4) $``$ $`𝔼_1Z`$ (2.5) $`=`$ $`\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right)\left(\begin{array}{c}z_0\\ z_1\end{array}\right)=\left(\begin{array}{c}z_1\\ z_0\end{array}\right).`$ (2.12) Of course, we have $$\left(𝔼_0\right)^2=\mathrm{𝟏}_2,\left(𝔼_1\right)^2=\mathrm{𝟏}_2.$$ (2.13) Now, there is a problem, these two matrices $`𝔼_0`$ and $`𝔼_1`$ alone are not sufficient to form a basis for $`R(2)`$. The solution of our dilemma is straightforward. If we also take into account that $$z^{}=z_0z_1e_1Z^{}=\left(\begin{array}{c}z_0\\ z_1\end{array}\right)$$ we find $`Z^{}`$ $`=`$ $`\stackrel{~}{𝔼}_0Z`$ (2.14) $`=`$ $`\left(\begin{array}{cc}1& 0\\ 0& 1\end{array}\right)\left(\begin{array}{c}z_0\\ z_1\end{array}\right)=\left(\begin{array}{c}z_0\\ z_1\end{array}\right),`$ (2.21) and $`e_1z^{}`$ $`=`$ $`z^{}e_1=z_0e_1+z_1=e_1^{}z`$ (2.22) $``$ $`𝔼_1Z^{}=\stackrel{~}{𝔼}_1Z`$ (2.23) $`=`$ $`\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right)\left(\begin{array}{c}z_0\\ z_1\end{array}\right)=\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right)\left(\begin{array}{c}z_0\\ z_1\end{array}\right)=\left(\begin{array}{c}z_1\\ z_0\end{array}\right).`$ (2.34) Obviously with these four matrices $`\{𝔼_0,\stackrel{~}{𝔼}_0,𝔼_1,\stackrel{~}{𝔼}_1\}`$, $`(n)(2n)`$ is proved. ### 2.2 Sp(1) Structure The quaternionic algebra is given by $`e_i.e_j=\delta _{ij}+ϵ_{ijk}e_k[e_i,e_j]=2ϵ_{ijk}e_k,`$ where$`ϵ_{ijk}`$ is the three dimensional Levi-Civita tensor $`\left(ϵ_{123}=1\right)`$ and $`i,j,k=1,2,3`$. Being non commutative, one must distinguish between right and left multiplication. In 1989, writing a quaternionic Dirac equation , Rotelli introduced a “barred” momentum operator with right action of $`e_1`$ $$_\mu e_1$$ (2.35) such that $$[(_\mu e_1)\psi _\mu \psi e_1].$$ (2.36) In recent papers , partially barred quaternions $$q+pe_1[q,p],$$ (2.37) have been used to formulate a quaternionic quantum mechanics and quaternionic field theory. From the viewpoint of group structure, these barred operators are very similar to complexified quaternions $$q+p$$ (2.38) where the imaginary unit $``$ commutes with the quaternionic imaginary units ($`e_1,`$ $`e_2,`$ $`e_3`$), but in physical problems, like eigenvalue calculations, tensor products, relativistic equations solutions, they give different results. A complete generalization for quaternionic multiplication is represented by the following barred operators $$|=q_1+q_2e_1+q_3e_2+q_4e_3[q_{1,\mathrm{},4}],$$ (2.39) and was developed a long time ago. As early as 1912, they had been used by Conway and Silberstein to reformulate special relativity and electromagnetism in a pure quaternionic language. Look to Synge for a review. The set of $`|`$ numbers with its 16 linearly independent elements, form a basis of $`GL(4,)`$. They have been revived recently to write down a one-component Dirac equation . Let us now show how to represent these quaternionic operators as real $`4\times 4`$ matrices. Like in the complex case, we represent *any quaternionic number (as distinct from an “operator”)* as a column vector $$q=q_0+q_1e_1+q_2e_2+q_3e_3Q=\left(\begin{array}{c}q_0\\ q_1\\ q_2\\ q_3\end{array}\right),$$ then $$e_1.q=q_0e_1q_1+q_2e_3q_3e_2𝔼_1Q$$ (2.40) and so on forth. The canonical left quaternionic structures over $`^4`$ are $$𝔼_1=\left(\begin{array}{cccc}0& 1& 0& 0\\ 1& 0& 0& 0\\ 0& 0& 0& 1\\ 0& 0& 1& 0\end{array}\right);𝔼_2=\left(\begin{array}{cccc}0& 0& 1& 0\\ 0& 0& 0& 1\\ 1& 0& 0& 0\\ 0& 1& 0& 0\end{array}\right);$$ $$𝔼_3=\left(\begin{array}{cccc}0& 0& 0& 1\\ 0& 0& 1& 0\\ 0& 1& 0& 0\\ 1& 0& 0& 0\end{array}\right);$$ (2.41) such that $$𝔼_i𝔼_j=(\delta _{ij}+ϵ_{ijk}𝔼_k),(𝔼_i)^2=\mathrm{𝟏}_4,$$ (2.42) Using Rotelli’s notation, right action is given by $`(1|e_1)q`$ $`=`$ $`qe_1=q_0e_1q_1q_2e_3+q_3e_2`$ (2.43) $``$ $`1|𝔼_1Q`$ (2.44) and so on for $`1|e_2,1|e_3`$. Our canonical right quaternionic structures are $$1|𝔼_1=\left(\begin{array}{cccc}0& 1& 0& 0\\ 1& 0& 0& 0\\ 0& 0& 0& 1\\ 0& 0& 1& 0\end{array}\right);1|𝔼_2=\left(\begin{array}{cccc}0& 0& 1& 0\\ 0& 0& 0& 1\\ 1& 0& 0& 0\\ 0& 1& 0& 0\end{array}\right);$$ $$1|𝔼_3=\left(\begin{array}{cccc}0& 0& 0& 1\\ 0& 0& 1& 0\\ 0& 1& 0& 0\\ 1& 0& 0& 0\end{array}\right),$$ (2.45) and $$1|𝔼_i1|𝔼_j=(\delta _{ij}ϵ_{ijk}1|𝔼_k),(1|𝔼_i)^2=\mathrm{𝟏}_4.$$ (2.46) We can write these left/right quaternionic structures compactly as $`(𝔼_i)_{\mu \nu }`$ $`=`$ $`(\delta _{0\mu }\delta _{i\nu }\delta _{0\nu }\delta _{i\mu }ϵ_{i\mu \nu })`$ (2.47) $`(1|𝔼_i)_{\mu \nu }`$ $`=`$ $`(\delta _{0\mu }\delta _{i\nu }\delta _{0\nu }\delta _{i\mu }+ϵ_{i\mu \nu })`$ (2.48) where $`\mu ,\nu `$ run from 0 to 3 or explicitly $`𝔼_{i\mu \nu }`$ $`=`$ $`ϵ_{i\mu \nu }\text{ if }\mu ,\nu =1,2,3.`$ (2.49) $`𝔼_{i0\nu }`$ $`=`$ $`\delta _{i\nu },𝔼_{i\mu 0}=\delta _{i\mu },𝔼_{i00}=0.`$ (2.50) and $`1|𝔼_{i\mu \nu }`$ $`=`$ $`ϵ_{i\mu \nu }\text{ if }\mu ,\nu =1,2,3.`$ (2.51) $`1|𝔼_{i0\nu }`$ $`=`$ $`\delta _{i\nu },1|𝔼_{i\mu 0}=\delta _{i\mu },1|𝔼_{i00}=0.`$ (2.52) These mathematical quaternionic structures are the ’t Hooft eta symbols well known in physics and we can check that $$1|𝔼_{i\mu \nu }=\frac{1}{2}ϵ_{\mu \nu \alpha \beta }1|𝔼_{i\alpha \beta }and𝔼_{i\mu \nu }=\frac{1}{2}ϵ_{\mu \nu \alpha \beta }𝔼_{i\alpha \beta }.$$ (2.53) Having $`𝔼_i,1|𝔼_i`$ enables us to find any generic operator $`𝔼_i|𝔼_j`$ corresponding to $`e_i|e_j`$ $$\left(e_i|e_j\right)q=e_i\left(1|e_j\right)q=e_iqe_j\left(𝔼_i|𝔼_j\right)Q=𝔼_i\left(1|𝔼_j\right)Q,$$ (2.54) then we have the 16 base elements of the operator in $`|`$ $$\left\{\begin{array}{c}1,e_1,e_2,e_3,\\ 1|e_1,e_1|e_1,e_2|e_1,e_3|e_1,\\ 1|e_2,e_1|e_2,e_2|e_2,e_3|e_2,\\ 1|e_3,e_1|e_3,e_2|e_3,e_3|e_3\end{array}\right\}.$$ (2.55) And their corresponding matrix representations $$\left\{\begin{array}{c}\mathrm{𝟏}_4,𝔼_1,𝔼_2,𝔼_3,\\ 1|𝔼_1,𝔼_1|𝔼_1,𝔼_2|𝔼_1,𝔼_3|𝔼_1,\\ 1|𝔼_2,𝔼_1|𝔼_2,𝔼_2|𝔼_2,𝔼_3|𝔼_2,\\ 1|𝔼_3,𝔼_1|𝔼_3,𝔼_2|𝔼_3,𝔼_3|𝔼_3\end{array}\right\}.$$ (2.56) We can thus deduce the following group structure for our quaternionic operators * $`su(2)_{Left}`$ $`e_ie_j`$ $`=`$ $`\delta _{ij}+ϵ_{ijk}e_k,`$ (2.57) $`su(2)_{Left}`$ $``$ $`\{e_1,e_2,e_3\}\{𝔼_1,𝔼_2,𝔼_3\}.`$ (2.58) * $`su(2)_{Right}`$ $`1|e_j1|e_i`$ $`=`$ $`1|(e_ie_j)=\delta _{ij}ϵ_{ijk}1|e_k,`$ (2.59) $`su(2)_{Right}`$ $``$ $`\{a_1=1|e_2,a_2=1|e_1,a_3=1|e_3\}`$ (2.60) $``$ $`\left\{a_1=1\right|𝔼_2,a_2=1|𝔼_1,a_3=1|𝔼_3\}`$ such that <sup>1</sup><sup>1</sup>1We rename $`a_1=1|e_2`$ and $`a_2=1|e_1`$ so that equation (2.61) comes with the standard sign. $$a_ia_j=\delta _{ij}+ϵ_{ijk}a_k.$$ (2.61) * $`so(4)su(2)_{Left}\times su(2)_{Right}`$, as $$e_i.1|e_j=1|e_j.e_i=e_i|e_j,i.e.[e_i,1|e_j]=0,$$ (2.62) thus $$so(4)\{e_1,e_2,e_3,1|e_1,1|e_2,1|e_3\}.$$ (2.63) * $`spin(2,3)`$ \- and its subgroups - can be realized by a Clifford algebra construction, e.g. $`\gamma _1`$ $`=`$ $`e_3,\gamma _2=e_2,\gamma _3=e_1|e_1,\gamma _4=e_1|e_2,\gamma _5=e_1|e_3,`$ $$\{\gamma _\alpha ,\gamma _\beta \}=2diag(,,+,+,+).$$ (2.65) By explicit calculation, one finds (in the basis given above) $`spin(2,3)`$ $``$ $`thesetof[\gamma _\alpha ,\gamma _\beta ]\alpha ,\beta =\mathrm{1..5},`$ $``$ $`\{e_1,1|e_1,1|e_2,1|e_3,e_2|e_1,e_3|e_1,e_2|e_2,e_3|e_2,e_2|e_3,e_3|e_3\}.`$ The reason that eqn.(LABEL:gamma5) can lead to eqn.(2.65) is that $$e_ie_j\left(1|e_k\right)+e_je_i\left(1|e_k\right)=0.$$ This construction was first introduced by Synge to give a quaternionic formulation of special relativity ($`so(1,3)`$). * Also at the matrix level the full set $`|`$ closes as an algebra, indeed using the above equations we find $`1|e_ie_j|e_k`$ $`=`$ $`ϵ_{kil}e_j|e_l,`$ (2.67) $`e_ie_j|e_k`$ $`=`$ $`ϵ_{ijl}e_l|e_k,`$ (2.68) $`e_i|e_je_m|e_n`$ $`=`$ $`ϵ_{iml}ϵ_{njp}e_l|e_p.`$ (2.69) We have used Maple to prove that the 16 matrices of $`\left\{\right|\}`$ are linearly independent so that they can form a basis for any $`(4)`$ as we claimed in (2.2). Omitting the identity, the set of 15 elements $`\left\{\right|\}\backslash 1`$ closes an $`sl(4,)`$ algebra. ### 2.3 Octonionic Structure We now summarize our notation for the octonionic algebra. There is a number of equivalent ways to represent the octonions multiplication table. Fortunately, it is always possible to choose an orthonormal basis $`(e_0,\mathrm{},e_7)`$ such that $$\phi =\phi _0+\phi _me_m(\phi _{0,\mathrm{},7}),$$ (2.70) where $`e_m`$ are elements obeying the noncommutative and nonassociative algebra $$e_me_n=\delta _{mn}+f_{mnp}e_p(m,n,p=\mathrm{1..7}),$$ (2.71) with $`f_{mnp}`$ totally antisymmetric and equal to unity for the seven different three cycles $$123,\mathrm{\hspace{0.33em}145},\mathrm{\hspace{0.33em}176},\mathrm{\hspace{0.33em}246},\mathrm{\hspace{0.33em}257},\mathrm{\hspace{0.33em}347},\mathrm{\hspace{0.33em}\hspace{0.33em}365}$$ (each cycle represents a quaternionic subalgebra). We can define an associator as follows, for any three octonionic numbers $`a,b`$ and $`c`$, $$\{a,b,c\}(ab)ca(bc),$$ (2.72) where in each term on the right-hand we must, first of all, perform the multiplication in brackets. Note that for real, complex and quaternionic numbers the associator is trivially null. For octonionic imaginary units however we have $$\{e_m,e_n,e_p\}(e_me_n)e_pe_m(e_ne_p)=2C_{mnps}e_s,$$ (2.73) with $`C_{mnps}`$ totally antisymmetric and equal to unity for the seven combinations $$1247,\mathrm{\hspace{0.33em}1265},\mathrm{\hspace{0.33em}2345},\mathrm{\hspace{0.33em}2376},\mathrm{\hspace{0.33em}3146},\mathrm{\hspace{0.33em}3157},\mathrm{\hspace{0.33em}\hspace{0.33em}4567}.$$ Working with octonionic numbers the associator (2.73) is non-vanishing for any three elements which are not in the same three cycles, however, the “alternative condition” is always fulfilled $$\{a,b,c\}+\{c,b,a\}=0.$$ (2.74) Due to the non-associativity, representing any form of octonions by matrices seems impossible. Nevertheless, we overcome these problems by introducing left/right-octonionic operators and fixing the direction of action. We discuss in the next subsection their relation to $`GL(8,)`$ where we present our translation idea and give some explicit examples which allow us to establish the isomorphism between our left/right octonionic operators and $`GL(8,)`$. Let us summarize the main points of the translation idea leaving the details to the next subsection. Exactly as in the quaternionic case, it seems natural to define and investigate the existence of barred operators $$𝕆_0+𝕆_m|e_m[𝕆_{0,\mathrm{},7}\text{octonions}].$$ (2.75) We first observe that an octonionic barred operator, $`a|b`$, which acts on octonionic functions, $`\phi `$, $$a|b\phi a\phi b,$$ is not a well defined object. For $`ab`$ the triple product $`a\phi b`$ could be either $`(a\phi )b`$ or $`a(\phi b)`$. So, in order to avoid this ambiguity (due to the nonassociativity of the octonionic numbers) we need to introduce left/right-barred operators. We will define left-barred operators by $`a)b`$, with $`a`$ and $`b`$ which represent octonionic numbers . They act on octonionic functions $`\phi `$ as follows $$a)b\phi =(a\phi )b.$$ (2.76) In a similar way we can introduce right-barred operators $`a(b`$, defined by $$a(b\phi =a(\phi b).$$ (2.77) Obviously, there are barred-operators which are associative like $$1)a=1(a1|a.$$ Furthermore, because of the alternativity condition (2.74) , $$a)a=a(aa|a.$$ At first glance it seems that we must consider the following 106 barred-operators: | $`1,e_m,1|e_m`$ | (15 elements) , | | --- | --- | | $`e_m|e_m`$ | (7) , | | $`e_m)e_n`$ $`(mn)`$ | (42) , | | $`e_m(e_n`$ $`(mn)`$ | (42) , | | $`(m,n=1,\mathrm{},7).`$ | | Nevertheless, *it is possible to prove that each right-barred operator can be expressed by a suitable combination of left-barred operators.* For example, from eq. (2.74), by posing $`a=e_m`$ and $`c=e_n`$, we quickly obtain $$e_m(e_n+e_n(e_me_m)e_n+e_n)e_m.$$ (2.78) So we can represent the most general octonionic operator by only left-barred objects $$𝕆_0+\underset{m=1}{\overset{7}{}}𝕆_m)e_m[𝕆_{0,\mathrm{},7}\text{octonions}],$$ (2.79) reducing to 64 independent elements the previous 106. This number of 64 suggests a correspondence between our barred octonions (2.79) and $`GL(8,)`$. In subsection (2.3.2), we focus our attention on the group $`GL(4,)GL(8,)`$. In doing so, we will find that only particular combinations of octonionic barred operators give us suitable candidates for the $`GL(4,)`$ translation. #### 2.3.1 Octonionic Operators and $`8\times 8`$ Real Matrices In order to explain the idea of *translation*, let us look explicitly at the action of the operators $`1e_1`$ and $`e_2`$, on a generic octonionic function $`\phi `$ $$\phi =\phi _0+e_1\phi _1+e_2\phi _2+e_3\phi _3+e_4\phi _4+e_5\phi _5+e_6\phi _6+e_7\phi _7[\phi _{0,\mathrm{},7}].$$ (2.80) We have $`1|e_1\phi `$ $``$ $`\phi e_1=e_1\phi _0\phi _1e_3\phi _2+e_2\phi _3e_5\phi _4+e_4\phi _5+e_7\phi _6e_6\phi _7,`$ $`e_2\phi `$ $`=`$ $`e_2\phi _0e_3\phi _1\phi _2+e_1\phi _3+e_6\phi _4+e_7\phi _5e_4\phi _6e_5\phi _7.`$ If we represent our octonionic (“state”) function $`\phi `$ by the following real column vector $$\mathrm{\Phi }\left(\begin{array}{c}\phi _0\\ \phi _1\\ \phi _2\\ \phi _3\\ \phi _4\\ \phi _5\\ \phi _6\\ \phi _7\end{array}\right),$$ (2.83) we can rewrite eqs. (LABEL:opaLABEL:opa2) in matrix form, $`1|𝔼_1\mathrm{\Phi }`$ $`=`$ $`\left(\begin{array}{c}\phi _1\hfill \\ \phi _0\hfill \\ \phi _3\hfill \\ \phi _2\hfill \\ \phi _5\hfill \\ \phi _4\hfill \\ \phi _7\hfill \\ \phi _6\hfill \end{array}\right)=\left(\begin{array}{cccccccc}0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill \end{array}\right)\left(\begin{array}{c}\phi _0\hfill \\ \phi _1\hfill \\ \phi _2\hfill \\ \phi _3\hfill \\ \phi _4\hfill \\ \phi _5\hfill \\ \phi _6\hfill \\ \phi _7\hfill \end{array}\right)`$ (2.108) $`𝔼_2\mathrm{\Phi }`$ $`=`$ $`\left(\begin{array}{c}\phi _2\hfill \\ \phi _3\hfill \\ \phi _0\hfill \\ \phi _1\hfill \\ \phi _6\hfill \\ \phi _7\hfill \\ \phi _4\hfill \\ \phi _5\hfill \end{array}\right)=\left(\begin{array}{cccccccc}0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill \end{array}\right)\left(\begin{array}{c}\phi _0\hfill \\ \phi _1\hfill \\ \phi _2\hfill \\ \phi _3\hfill \\ \phi _4\hfill \\ \phi _5\hfill \\ \phi _6\hfill \\ \phi _7\hfill \end{array}\right)`$ (2.134) In this way we can immediately obtain a real matrix representation for the octonionic barred operators $`1e_1`$ and $`e_2`$. Following this procedure we can construct the complete set of translation rules for the imaginary unit operators $`e_m`$ and the barred operators $`1e_m`$ (appendix A). At first glance it seems that our translation doesn’t work. If we extract the matrices corresponding to $`e_1`$, $`e_2`$ and $`e_3`$, namely, $`𝔼_1`$ $`=`$ $`\left(\begin{array}{cccccccc}0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill \end{array}\right),`$ $`𝔼_2`$ $`=`$ $`\left(\begin{array}{cccccccc}0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill \end{array}\right),`$ $`𝔼_3`$ $`=`$ $`\left(\begin{array}{cccccccc}0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill \end{array}\right),`$ we find $$𝔼_1𝔼_2𝔼_3.$$ (2.139) In obvious contrast with the octonionic relation $$e_1e_2=e_3.$$ (2.140) This paradox is easily understood. In deducing our translation rules, we understand octonions as operators, and so they must be applied to a certain octonionic function, or state, $`\phi `$, and not upon another “operator”. So the octonionic relation $$e_3\phi =(e_1e_2)\phi $$ (2.141) is indeed translated by $$𝔼_3\phi ,$$ (2.142) whereas, $$e_1(e_2\phi )[e_3\phi ]$$ (2.143) becomes $$𝔼_1𝔼_2\phi [𝔼_3\phi ].$$ (2.144) For $`e_m`$ and $`1|e_n`$, we have simple multiplication rules. In fact, utilizing the associator properties we find $`e_m\left[e_n\phi \right]`$ $`=`$ $`(e_me_n)\phi +(e_m\phi )e_ne_m(\phi e_n),`$ (2.145) $`\left[\phi e_m\right]e_n`$ $`=`$ $`\phi (e_me_n)(e_m\phi )e_n+e_m(\phi e_n).`$ (2.146) Thus<sup>2</sup><sup>2</sup>2We have used square brackets on the L.H.S. in the previous two equations in order to avoid confusion between the L.H.S. and the last term on the R.H.S. in the following two equations which might occur if we had employed $`e_m\left(e_n\phi \right)`$ for the L.H.S. of (2.145) etc., $`e_m.e_n`$ $``$ $`\delta _{mn}+f_{mnp}e_p+e_m)e_ne_m(e_n,`$ (2.147) $`1|e_n\mathrm{\hspace{0.33em}\hspace{0.33em}.\hspace{0.33em}1}|e_m`$ $``$ $`\delta _{mn}+f_{mnp}e_pe_m)e_n+e_m(e_n.`$ (2.148) The previous relation can be immediately rewritten in matrix form as follows $`𝔼_m𝔼_n`$ $``$ $`\delta _{mn}+f_{mnp}𝔼_p+[1|𝔼_n,𝔼_m],`$ (2.149) $`1|𝔼_n1|𝔼_m`$ $``$ $`\delta _{mn}+f_{mnp}1|𝔼_p+[𝔼_m,\mathrm{\hspace{0.33em}1}|𝔼_n].`$ (2.150) Introducing a new matrix multiplication, “ $``$ ”, related to the standard matrix multiplication (row by column) by $$𝔼_m𝔼_n𝔼_m𝔼_n[1|𝔼_n,𝔼_m],$$ (2.151) we can quickly reformulate the nonassociative octonionic algebra (in terms of matrices this time) by $$𝔼_m𝔼_n=\delta _{mn}+f_{mnp}𝔼_p.$$ (2.152) Working with left/right barred operators we now show how the nonassociativity is realized with our matrix translation. Such operators enable us to reproduce the octonions nonassociativity by the matrix algebra. Consider for example $$e_3)e_1\phi (e_3\phi )e_1=e_2\phi _0e_3\phi _1+\phi _2e_1\phi _3e_6\phi _4e_7\phi _5+e_4\phi _6+e_5\phi _7.$$ (2.153) This equation will be translated into $$\left(\begin{array}{cccccccc}0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill \end{array}\right)\left(\begin{array}{c}\phi _0\hfill \\ \phi _1\hfill \\ \phi _2\hfill \\ \phi _3\hfill \\ \phi _4\hfill \\ \phi _5\hfill \\ \phi _6\hfill \\ \phi _7\hfill \end{array}\right)=\left(\begin{array}{c}\phi _2\hfill \\ \phi _3\hfill \\ \phi _0\hfill \\ \phi _1\hfill \\ \phi _6\hfill \\ \phi _7\hfill \\ \phi _4\hfill \\ \phi _5\hfill \end{array}\right)$$ (2.154) Whereas, $$e_3(e_1\phi e_3(\phi e_1)=e_2\phi _0e_3\phi _1+\phi _2e_1\phi _3+e_6\phi _4+e_7\phi _5e_4\phi _6e_5\phi _7,$$ (2.155) will become $$\left(\begin{array}{cccccccc}0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 1\hfill & 0\hfill & 0\hfill \end{array}\right)\left(\begin{array}{c}\phi _0\hfill \\ \phi _1\hfill \\ \phi _2\hfill \\ \phi _3\hfill \\ \phi _4\hfill \\ \phi _5\hfill \\ \phi _6\hfill \\ \phi _7\hfill \end{array}\right)=\left(\begin{array}{c}\phi _2\hfill \\ \phi _3\hfill \\ \phi _0\hfill \\ \phi _1\hfill \\ \phi _6\hfill \\ \phi _7\hfill \\ \phi _4\hfill \\ \phi _5\hfill \end{array}\right)$$ (2.156) The nonassociativity is then reproduced since left and right barred operators, like $$e_3)e_1\text{and}e_3(e_1$$ are represented by different matrices. The complete set of translation rules for left/right-barred operators is given in appendix A. The full matrix representation for left/right barred operators can be quickly obtained by suitable multiplications of the matrices $`𝔼_m`$ and $`1|𝔼_n`$. By direct calculations we can extract the matrices which correspond to the operators $$e_m)e_nande_m(e_n,$$ which we call, respectively, $$𝔼_m)𝔼_nand𝔼_m(𝔼_n.$$ Since our left/right barred operators can be represented by an ordered action of the operators $`e_m`$ and $`1e_n`$, we can relate the matrices $`𝔼_m)𝔼_n`$ and $`𝔼_m(𝔼_n`$ to the matrices $`𝔼_m`$ and $`1|𝔼_m`$: $`𝔼_m)𝔼_n`$ $``$ $`1|𝔼_n𝔼_m,`$ (2.157) $`𝔼_m(𝔼_n`$ $``$ $`𝔼_m\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}1}|𝔼_n.`$ (2.158) The previous discussions concerning the octonions nonassociativity and the isomorphism between $`GL(8,)`$ and barred octonions, can now be elegantly summarized as follows. 1 - Matrix representation for octonions nonassociativity acting on certain octonionic number. $$𝔼_m)𝔼_n𝔼_m(𝔼_n[1|𝔼_n𝔼_m𝔼_m1|𝔼_n\text{ for }mn].$$ 2 - Isomorphism between $`GL(8,)`$ and barred octonions. If we rewrite our 106 barred operators by real matrices: | $`1,𝔼_m,\mathrm{\hspace{0.33em}1}|𝔼_m`$ | (15 matrices) , | | --- | --- | | $`𝔼_m|𝔼_m𝔼_m\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}1}|𝔼_m=1|𝔼_m𝔼_m`$ | (7) , | | $`𝔼_m)𝔼_n1|𝔼_n𝔼_m`$ $`(mn)`$ | (42) , | | $`𝔼_m(𝔼_n𝔼_n\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}1}|𝔼_m`$ $`(mn)`$ | (42) , | | $`(m,n=1,\mathrm{},7);`$ | | we have two different basis for $`GL(8,)`$: | (1) | $`1,𝔼_m,1|𝔼_m,1|𝔼_n𝔼_m,\mathrm{\hspace{0.33em}\hspace{0.33em}1}|𝔼_m𝔼_m,`$ | | --- | --- | | (2) | $`1,𝔼_m,1|𝔼_m,𝔼_m\mathrm{\hspace{0.33em}\hspace{0.33em}1}|𝔼_n,1|𝔼_m𝔼_m.`$ | We now note some difficulties due to the nonassociativity of octonions. When we translate from barred octonions to $`8\times 8`$ real matrices there is no problem. For example, in the octonionic equation $$e_4\{[(e_6\phi )e_1]e_5\},$$ (2.159) we quickly recognize the following left/right octonionic operators, $$e_6)e_1followedbye_4(e_5.$$ Hence we can translate eq. (2.159) into $$\left[𝔼_4(𝔼_5][𝔼_6)𝔼_1\right]\phi .$$ (2.160) But in going from $`8\times 8`$ real matrices to octonions we must be careful of the ordering. For example, with A, B matrices $$AB\phi $$ (2.161) can be understood for translation purposes as $$(AB)\phi ,$$ (2.162) or $$A(B\phi ).$$ (2.163) In order to avoid confusion we translate eq. (2.161) by eq. (2.163). In general when brackets are absent we shall choose the convention that $$ABC\mathrm{}Z\phi A(B(C\mathrm{}(Z\phi )\mathrm{})).$$ (2.164) #### 2.3.2 Octonionic Operators and $`4\times 4`$ Complex Matrices Some complex groups play a critical role in physics. No one can deny the importance of $`U(1,)`$ and $`SU(2,)`$. In relativistic quantum mechanics, $`GL(4,)`$ is implicit in writing the Dirac equation. Starting from our $`GL(8,)`$, we should be able to extract its subgroup $`GL(4,)`$. Whence we should be able to translate the famous Dirac-gamma matrices and write down a four dimensional one-component octonionic Dirac equation . Let us show how we can extract our 32 basis of $`GL(4,)`$: Working with the symplectic decomposition of octonionic “states” $$\psi =\left(\begin{array}{c}\psi _1\\ \psi _2\\ \psi _3\\ \psi _4\end{array}\right)\psi _1+e_2\psi _2+e_4\psi _3+e_6\psi _4[\psi _{1,\mathrm{},4}(1,e_1)].$$ (2.165) we analyze the action of left-barred operators on our octonionic wave functions $`\psi `$. For example, we find $`1|e_1\psi `$ $`\psi e_1`$ $`=e_1\psi _1+e_2(e_1\psi _2)+e_4(e_1\psi _3)+e_6(e_1\psi _4),`$ (2.166) $`e_2\psi `$ $`=\psi _2+e_2\psi _1e_4\psi _4^{}+e_6\psi _3^{},`$ (2.167) $`e_3)e_1\psi `$ $`(e_3\psi )e_1`$ $`=\psi _2+e_2\psi _1+e_4\psi _4^{}e_6\psi _3^{}.`$ (2.168) Following the same methodology of the previous section, we can immediately note a correspondence between the complex matrix $`i\mathrm{𝟏}_{4\times 4}`$and the octonionic operator $`1e_1`$ $$\left(\begin{array}{cccc}i& 0& 0& 0\\ 0& i& 0& 0\\ 0& 0& i& 0\\ 0& 0& 0& i\end{array}\right)1|e_1.$$ (2.169) This translation does not work for all barred operators. Let us show this, explicitly. For example, we cannot find a $`4\times 4`$ complex matrix which, acting upon $$\left(\begin{array}{c}\psi _1\\ \psi _2\\ \psi _3\\ \psi _4\end{array}\right),$$ gives the column vector $$e_2\psi =\left(\begin{array}{c}\psi _2\\ \psi _1\\ \psi _4^{}\\ \psi _3^{}\end{array}\right)ore_3)e_1\psi =(\begin{array}{c}\psi _2\\ \psi _1\\ \psi _4^{}\\ \psi _3^{}\end{array})$$ and so we have not the possibility to relate $$e_2ore_3)e_1$$ with a complex matrix. Nevertheless, a combined action of these two operators gives us $$e_2\psi +(e_3\psi )e_1=2e_2\psi _1,$$ and allows us to represent the octonionic barred sum $$e_2+e_3)e_1,$$ (2.170) by the $`4\times 4`$ complex matrix $$\left(\begin{array}{cccc}0& 0& 0& 0\\ 2& 0& 0& 0\\ 0& 0& 0& 0\\ 0& 0& 0& 0\end{array}\right).$$ (2.171) Following this procedure we can represent our generic $`4\times 4`$ complex matrix by octonionic barred operators (but not necessarily the contrary). The explicit correspondence rules are given in appendix B. We conclude our discussion upon the relationship between barred operators and $`4\times 4`$ complex matrices, by noting that the 32 basis elements of $`GL(4,)`$ can be extracted in a different way from the 64 generators of $`GL(8,)`$. It is well known that any complex matrix can be rewritten as a real matrix by the following isomorphism ($`\sigma `$ are the standard Pauli matrices), $$1\mathrm{𝟏}_{2\times 2}\text{and}ii\sigma _2.$$ The situation at the lowest order is $`GL(2,)`$ generators : $`\mathrm{𝟏}_{2\times 2},\sigma _1,i\sigma _2,\sigma _3;`$ (2.172) $`GL(1,)`$ isomorphic : $`\mathrm{𝟏}_{2\times 2},i\sigma _2.`$ (2.173) In a similar way (choosing appropriate combinations of left-barred octonionic operators, in which only $`\pm \mathrm{𝟏}_{2\times 2}`$and $`\pm i\sigma _2`$appear) we can extract from $`GL(8,)`$ the 32 basis elements of $`GL(4,)`$. For further details see appendix B. ### 2.4 Beyond Octonions and Clifford Algebras Going to higher dimensions, we define “hexagonions” ($`𝕏`$) by introducing a new element $`e_8`$ such that $$\begin{array}{cccc}𝕏\hfill & =\hfill & 𝕆_1+𝕆_2e_8\hfill & \\ & =\hfill & x_0e_0+\mathrm{}+x_{16}e_{16}.\hfill & x_\mu \hfill \end{array}$$ (2.174) and $$e_ie_j=\delta _{ij}+C_{ijk}e_k.$$ (2.175) Now, we have to find a suitable form of $`C_{ijk}`$. Recalling how the structure constant is written for octonions $`𝕆`$ $`=`$ $`_1+_2e_4`$ (2.176) $`=`$ $`x_0e_0+\mathrm{}+x_7e_7,`$ where $``$ are quaternions, we have already chosen the convention $`e_1e_2=e_3`$ which is extendable to (2.176), we set $`e_1e_4=e_5`$, $`e_2e_4=e_6`$and $`e_3e_4=e_7`$, but we still lack the relationships between the remaining possible triplets, $`\{e_1,e_6,e_7\};`$ $`\{e_2,e_5,e_7\};`$ $`\{e_3,e_5,e_6\}`$ which can be fixed by using $$\begin{array}{c}e_1e_6=e_1(e_2e_4)=(e_1e_2)e_4=e_3e_4=e_7,\\ e_2e_5=e_2(e_1e_4)=(e_2e_1)e_4=+e_3e_4=+e_7,\\ e_3e_5=e_3(e_1e_4)=(e_3e_1)e_4=e_2e_4=e_6.\end{array}$$ These define all the structure constants for octonions. Returning to $`𝕏`$, we have the seven octonionic conditions, and the decomposition (2.174). We set $`e_1e_8=e_9,e_2e_8=e_A,e_3e_8=e_B,e_4e_8=e_C,e_5e_8=e_D,e_6e_8=e_E,e_7e_8=e_F`$ where $`A=10,B=11,C=12,D=13,E=14`$ and $`F=15`$. The other elements of the multiplication table may be chosen in analogy with (2.176). Explicitly, the 35 hexagonionic triplets are $$\begin{array}{ccccccc}(123),& (145),& (246),& (347),& (257),& (176),& (365),\\ (189),& (28A),& (38B),& (48C),& (58D),& (68E),& (78F),\\ (1BA),& (1DC),& (1EF),& (29B),& (2EC),& (2FD),& (3A9),\\ (49D),& (4AE),& (4BF),& (3FC),& (3DE),& (5C9),& (5AF),\\ (5EB),& (6FD),& (6CA),& (6BD),& (79E),& (7DA),& (7CB).\end{array}$$ This can be extended for any generic higher dimensional field $`𝔽^n`$. It can be shown by using some combinatorics that the number of such triplets $`N`$ for a general $`𝔽^n`$field is ($`n>1`$) $$N=\frac{\left(2^n1\right)!}{\left(2^n3\right)!3!},$$ (2.177) giving $$\begin{array}{cccc}𝔽^n& n& dim& N\\ & 2& 4& 1\\ 𝕆& 3& 8& 7\\ 𝕏& 4& 16& 35\\ & & andsoon.& \end{array}$$ One may notice that for any non-ring division algebra $`\left(𝔽,n>3\right)`$$`N>dim(𝔽^n)`$ except when dim = $`\mathrm{},`$ i.e. a functional Hilbert space with a Cliff(0,$`\mathrm{}`$) structure. Does this inequality have any connection with the ring division structure of the ($`S^1,S^3,S^7`$) spheres ? Yes, that is what we are going to show now. Following, the same translation idea projecting our algebra $`𝕏`$ over $`^{16}`$, any $`𝔼_i`$ is given by a relation similar to that for $``$ $$(𝔼_i)_{\alpha \beta }=\delta _{i\alpha }\delta _{\beta 0}\delta _{i\beta }\delta _{\alpha 0}+C_{i\alpha \beta }.$$ (2.178) But contrary to the quaternions and octonions, the Clifford algebra closes only for a subset of these $`E_i`$’s, namely $$\{𝔼_i,𝔼_j\}=2\delta _{ij}\text{for}i,j,k=1\mathrm{}8not\mathrm{\hspace{0.33em}1}\mathrm{}15.$$ (2.179) Because we have lost the ring division structure. By careful investigation, we find that another ninth $`𝔼_i`$ can be constructed, in agreement with the Clifford algebra classification . There is no standard<sup>3</sup><sup>3</sup>3Look to for a non standard representation. 16 dimensional representation for $`Cliff\left(15\right)`$. Following this procedure, we can give a simple way to write real Clifford algebras over any arbitrary dimensions. Sometimes, a specific multiplication table may be favored. For example in soliton theory, the existence of a symplectic structure related to the bihamiltonian formulation of integrable models is welcome. It is known from the Darboux theorem, that locally a symplectic structure is given up to a minus sign by $$𝒥_{dim\times dim}=\left(\begin{array}{cc}0& \mathrm{𝟏}_{\frac{dim}{2}}\\ \mathrm{𝟏}_{\frac{dim}{2}}& 0\end{array}\right),$$ (2.180) this fixes the following structure constants $`C_{\left(\frac{dim}{2}\right)1\left(\frac{dim}{2}+1\right)}=1,`$ (2.181) $`C_{\left(\frac{dim}{2}\right)2\left(\frac{dim}{2}+2\right)}=1,`$ (2.182) $`\mathrm{}`$ (2.183) $`C_{\left(\frac{dim}{2}\right)\left(\frac{dim}{2}1\right)\left(dim1\right)}=1,`$ (2.184) which is the decomposition that we have chosen in (2.176) for octonions $$C_{415}=C_{426}=C_{437}=1.$$ (2.185) Generally our symplectic structure is $$\left(1|𝔼_{\left(\frac{dim}{2}\right)}\right)_{\alpha \beta }=\delta _{0\alpha }\delta _{\beta \left(\frac{dim}{2}\right)}\delta _{0\beta }\delta _{\alpha \left(\frac{dim}{2}\right)}ϵ_{\alpha \beta \left(\frac{dim}{2}\right)}.$$ (2.186) Moreover some other choices may exhibit a relation with number theory and Galois fields . It is highly non-trivial how Clifford algebraic language can be used to unify many distinct mathematical notions such as Grassmanian , complex, quaternionic and symplectic structures. The main result of this section, the non-existence of 16 dimensional representation of $`Cliff(0,15)`$ is in agreement with the Atiyah–Bott–Shapiro classification of real Clifford algebras . In this context, the importance of ring division algebras can also be deduced from the Bott periodicity . Another interesting observation, if we interpret the complex, quaternions, octonions eigenfunctions as real spinors, we find that for $`complexZ^tE_1Z`$ $`=`$ $`0,`$ (2.187) $`quaternionsQ^t𝔼_iQ`$ $`=`$ $`0i=\mathrm{1..3},`$ (2.188) $`octonions\mathrm{\Phi }^t𝔼_i\mathrm{\Phi }`$ $`=`$ $`0i=\mathrm{1..7}.`$ (2.189) These states ($`Z`$, $`Q`$, $`\mathrm{\Phi }`$) are called pure spinors as first coined by Cartan. These pure spinors play an important role for minimal surfaces, integrable models, twistor calculus, and string theory . ## Chapter 3 The Soft Seven Sphere Ring division algebras play fundamental roles in mathematics from algebra to geometry and topology with many different applications. The applicability of real and complex numbers in physics is not in question. Quaternions which may be represented as Pauli matrices are also important. the use of octonions in physics is the problem. In the first section of this chapter, we introduce what we mean by the word “soft” algebra, generally we follow closely the presentation of Sohnius. Sohnius used soft algebras with structure functions that vary over space-time as well as over an internal gauge manifold. In this thesis we use only soft algebras with structure functions that vary over the internal gauge manifold (the fiber) not the base space-time manifold. In the second section, we introduce the seven sphere as a soft algebra , we calculate its structure functions explicitly and also discuss some relevant points such as the validity of the Jacobi identity. Furthermore, we emphasis some important features such as the pointwise reduction, closure, and some other consistency checks. In the last section, we reformulate some standard Lie group results putting them in a form suitable to subsequent applications. ### 3.1 Sohnuis’ Idea Trying to find a suitable framework for supersymmetric theories, Sohnius introduced the notion of soft gauge algebras i.e. algebras where the structure constants become structure functions of space-time and gauge-dependent fields. He proceeded as follows: For any Lie group, we know that a generic element can be written as $`exp(i\epsilon _iL_i)`$ where $`\epsilon `$ are finite numbers of parameters and $`L_i`$ are our Lie algebras elements. A field $`A`$, that lives in a certain representation of our algebras, transforms infinitesimally as $$AA+\delta _{\epsilon _1}A\text{with}\delta _{\epsilon _1}A=i[A,\epsilon _1^iL_i].$$ (3.1) Sohnuis considered the special case when the commutator of two successive transformations leads to $$[\delta _{\epsilon _1},\delta _{\epsilon _2}]\mathrm{\Phi }=i[\mathrm{\Phi },\epsilon _1^i\epsilon _2^jf_{ij}^k(\phi )L_k]=i[\mathrm{\Phi },\epsilon _3^kL_k],$$ (3.2) where $`\phi `$ is a coordinate system for the internal gauge manifold and $`f_{ij}^k(\phi )`$ are “the structure functions” of our soft algebra defined by $$[\delta _{\epsilon _1^i},\delta _{\epsilon _2^j}]=\delta _{f_k^{ij}(\phi )\epsilon _3^k}.$$ (3.3) In standard local gauge theory, we have $`\epsilon (x)`$ and we need a gauge field that transforms inhomogeneously: $$\delta _\epsilon A_\mu ^i=_\mu \epsilon (x)+A_\mu ^j\epsilon (x)^if_{ij}^k(\phi ).$$ (3.4) In this thesis, we will always assume $$_\mu f_{ijk}(\phi )=0.$$ (3.5) By the $`\phi `$ dependence of $`f_{ijk}(\phi )`$, we mean a dependence on the internal gauged space which is another manifold distinct from our space-time $`x`$. To develop a representation, we use $$\delta _\epsilon \mathrm{\Phi }=\epsilon (x)\delta (L_i)\mathrm{\Phi },$$ (3.6) and $$[\delta (L_i),\delta (L_j)]=af_{ij}^k(\phi )\delta (L_k)$$ (3.7) where $`a`$. In this thesis we set $`a=2`$. Let’s see how octonions may be treated as a soft algebra exactly in the sense of eq. (3.7). ### 3.2 Octonions and the Soft Seven Sphere We start by recalling the non-associative octonion algebra. A generic octonion number is $$\phi =\phi _0e_0+\phi _ie_i=\phi _\mu e_\mu .[i=\mathrm{1..7},\mu =\mathrm{0..7},\phi _\mu ]$$ (3.8) and its associator $$[e_i,e_j,e_k]=(e_ie_j)e_ke_i(e_je_k),$$ (3.9) is non-zero for any three elements that are not in the same three cycles and is completely antisymmetric The following formula or any of its generalization is thus ambiguous $$e_1e_5e_7=\{\begin{array}{c}(e_1e_5)e_7=e_3\\ \text{or}\\ e_1(e_5e_7)=e_3\end{array}$$ (3.10) so the best way is to define the action of the imaginary units in a certain direction. In the spirit of Englert, Sevrin, Troost, Van Proeyen and Spindel (also look at ) we define the left action of octonionic operators $`\delta _i`$ by $$\delta _i\phi =(e_i\phi )$$ (3.11) implying that the following equation is well defined $$\delta _i\delta _j\phi =\delta _i(\delta _j\phi )=\delta _i(e_j\phi )=e_i(e_j\phi ),$$ (3.12) then eq. (3.10) reads unambiguously as $$\delta _1\delta _5e_7=(e_1(e_5e_7))=e_3.$$ (3.13) Since octonions are also non-commutative, we must also differentiate between left and right action. Using the barred notation , we introduce right action as $$1|\delta _i\phi =(\phi e_i),$$ (3.14) for example $$(1|\delta _i)(1|\delta _j)\phi =(1|\delta _i)(1|\delta _j\phi )=1|\delta _i(\phi e_j)=((\phi e_j)e_i).$$ (3.15) As we shall see shortly, we can express the associator in terms of left and right operators. The imaginary octonionic units generate the seven sphere $`S^7`$which has many properties similar to Lie algebras and/or Lie groups. $`S^7`$and Lie groups are the only non-flat compact parallelizable manifolds . The important point for evaluating any Lie algebra is the commutator, so let’s examine $`[\delta _i,\delta _j]\phi `$ $`=`$ $`e_i(e_j\phi )e_j(e_i\phi )`$ (3.16) $`=`$ $`2f_{ijk}e_k\phi 2[e_i,e_j,\phi ]`$ (3.17) $`=`$ $`2f_{ijk}e_k\phi +2[e_i,\phi ,e_j].`$ (3.18) now the last term can be written as $`[e_i,\phi ,e_j]`$ $`=`$ $`(e_i\phi )e_je_i(\phi e_j)`$ (3.19) $`=`$ $`\delta _i(1|\delta _j)\phi +(1|\delta _j)\delta _i\phi `$ (3.20) $`=`$ $`[\delta _i,1|\delta _j]\phi ,`$ (3.21) thus our commutator can be rewritten as $`[\delta _i,\delta _j]\phi `$ $`=`$ $`2f_{ijk}e_k\phi +2[e_i,\phi ,e_j]`$ (3.22) $`=`$ $`2f_{ijk}\delta _k\phi 2[\delta _i,1|\delta _j]\phi .`$ (3.23) Note that right operators are necessary because the last term the associator can never be written in terms of left operators alone. After simple calculations, one concludes that the octonionic imaginary units are determined completely by (3.23) and the following equations $`[1|\delta _i,1|\delta _j]\phi `$ $`=`$ $`2f_{ijk}1|\delta _k\phi 2[\delta _i,1|\delta _j]\phi `$ (3.24) $`\{\delta _i,\delta _j\}\phi `$ $`=`$ $`2\delta _{ij}\phi `$ (3.25) $`\{1|\delta _i,1|\delta _j\}\phi `$ $`=`$ $`2\delta _{ij}\phi ,`$ (3.26) where the $`\delta _{ij}`$ in (3.25) and (3.26) are the standard Kronecker delta tensor. It has been proved in , using three different ways, that the $`\delta _i`$ algebra is associative. Thus a representation theory in terms of matrices should be, in principle, possible. Indeed, in the last chapter, we have derived an algebra completely isomorphic to (3.23,3.24,3.25,3.26) by exploiting the idea that octonions can be used as a basis for any $`8\times 8`$ real matrix. we have two sets of matrices, essentially, $$\begin{array}{ccccc}\delta _i& & (𝔼_i)_{\mu \nu }& =& \delta _{0\mu }\delta _{i\nu }\delta _{0\nu }\delta _{i\mu }f_{i\mu \nu },\\ 1|\delta _i& & (1|𝔼_i)_{\mu \nu }& =& \delta _{0\mu }\delta _{i\nu }\delta _{0\nu }\delta _{i\mu }+f_{i\mu \nu }.\end{array}$$ (3.27) The set of matrices $`𝔼_i`$ and $`1|𝔼_i`$have appeared in different octonionic works e.g. . Furthermore, they correspond, as we have already said, to the ’t Hooft eta symbols (2.47-2.51). We suggest that their most appropriate names should be *the canonical left and right octonionic structure at the north/south pole of the seven sphere*. By explicit calculation, one finds that $`[𝔼_i,𝔼_j]\phi `$ $`=`$ $`2f_{ijk}𝔼_k\phi 2[𝔼_i,1|𝔼_j]\phi `$ (3.28) $`[1|𝔼_i,1|𝔼_j]\phi `$ $`=`$ $`2f_{ijk}1|𝔼_k\phi 2[𝔼_i,1|𝔼_j]\phi `$ (3.29) $`\{𝔼_i,𝔼_j\}\phi `$ $`=`$ $`2\delta _{ij}\phi `$ (3.30) $`\{1|𝔼_i,1|𝔼_j\}\phi `$ $`=`$ $`2\delta _{ij}\phi `$ (3.31) where $`\phi `$ is represented by a column matrix $$\phi ^t=\left(\begin{array}{cccccccc}\phi _0\hfill & \phi _1\hfill & \phi _2\hfill & \phi _3\hfill & \phi _4\hfill & \phi _5\hfill & \phi _6\hfill & \phi _7\hfill \end{array}\right)$$ The word “isomorphic” above is justified since {$`\delta _i`$} is associative and the same holds obviously for our {$`𝔼_i`$ } as they are written in terms of matrices. Our Jacobian identities are $$[\delta _i,[\delta _j,\delta _k]]\phi +[\delta _j,[\delta _k,\delta _i]]\phi +[\delta _k,[\delta _i,\delta _j]]\phi =0,$$ (3.32) or $$[𝔼_i,[𝔼_j,𝔼_k]]\phi +[𝔼_j,[𝔼_k,𝔼_i]]\phi +[𝔼_k,[𝔼_i,𝔼_j]]\phi =0.$$ (3.33) We shall return to these identities again at the end of this section. Fixing the direction of the application for any imaginary octonionic units extracts a part of the algebra that respects associativity, but a certain price has to be paid. The presence of the $`\phi `$ is essential and either of {$`𝔼_i`$} or {$`1|𝔼_i`$} is an open algebra, they don’t close upon the action of the commutator. Here comes the second step, the soft Lie algebra idea. It is clear that the right hand side of (3.22 or 3.28) has a complicated $`\phi `$ dependence. Knowing that the seven sphere has a torsion that varies from one point to another and mimicking the Lie group case where the structure constants are proportional to the fixed group torsion, it is natural to propose that (3.22) may be redefined as $$[\delta _i,\delta _j]\phi =2f_{ijk}^{(+)}(\phi )\delta _k\phi ,$$ (3.34) where $`f_{ijk}^{(+)}(\phi )`$ are structure functions that vary over the whole $`S^7`$manifold. It is clear that our $`\delta _i`$ play the same role of the $`\delta (L_i)`$ defined in the previous subsection eq.(3.7). These structure functions $`f_{ijk}^{(+)}(\phi )`$ were computed previously using different properties of the associator and some other octonionic identities in . Here we use our matrix representation to give another alternative way to calculate $`f_{ijk}^{(+)}(\phi )`$ $$[𝔼_i,𝔼_j]\phi =2f_{ijk}^{(+)}(\phi )𝔼_k\phi .$$ (3.35) Let’s do it for the following example $$[𝔼_1,𝔼_2]\phi =2f_{12k}^{(+)}(\phi )𝔼_k\phi ,$$ (3.36) which is equivalent to the following eight equations , $`\phi _3`$ $`=`$ $`\phi _1f_{121}^{(+)}(\phi )+\phi _2f_{122}^{(+)}(\phi )+\phi _3f_{123}^{(+)}(\phi )`$ $`+\phi _4f_{124}^{(+)}(\phi )+\phi _5f_{125}^{(+)}(\phi )+\phi _6f_{126}^{(+)}(\phi )+\phi _7f_{127}^{(+)}(\phi ),`$ $`\phi _2`$ $`=`$ $`\phi _2f_{123}^{(+)}(\phi )\phi _0f_{121}^{(+)}(\phi )\phi _3f_{122}^{(+)}(\phi )`$ $`\phi _5f_{124}^{(+)}(\phi )+\phi _4f_{125}^{(+)}(\phi )+\phi _7f_{126}^{(+)}(\phi )\phi _6f_{127}^{(+)}(\phi ),`$ $`\phi _1`$ $`=`$ $`\phi _0f_{122}^{(+)}(\phi )\phi _3f_{121}^{(+)}(\phi )+\phi _1f_{123}^{(+)}(\phi )`$ $`+\phi _6f_{124}^{(+)}(\phi )+\phi _7f_{125}^{(+)}(\phi )\phi _4f_{126}^{(+)}(\phi )\phi _5f_{127}^{(+)}(\phi ),`$ $`\phi _0`$ $`=`$ $`\phi _2f_{121}^{(+)}(\phi )\phi _1f_{122}^{(+)}(\phi )+\phi _0f_{123}^{(+)}(\phi )`$ $`+\phi _7f_{124}^{(+)}(\phi )\phi _6f_{125}^{(+)}(\phi )+\phi _5f_{126}^{(+)}(\phi )\phi _4f_{127}^{(+)}(\phi ),`$ $`\phi _7`$ $`=`$ $`\phi _0f_{124}^{(+)}(\phi )\phi _5f_{121}^{(+)}(\phi )\phi _6f_{122}^{(+)}(\phi )`$ $`\phi _7f_{123}^{(+)}(\phi )+\phi _1f_{125}^{(+)}(\phi )+\phi _2f_{126}^{(+)}(\phi )+\phi _3f_{127}^{(+)}(\phi ),`$ $`\phi _6`$ $`=`$ $`\phi _7f_{122}^{(+)}(\phi )\phi _4f_{121}^{(+)}(\phi )\phi _6f_{123}^{(+)}(\phi )`$ $`+\phi _1f_{124}^{(+)}(\phi )\phi _0f_{125}^{(+)}(\phi )+\phi _3f_{126}^{(+)}(\phi )\phi _2f_{127}^{(+)}(\phi ),`$ $`\phi _5`$ $`=`$ $`\phi _7f_{121}^{(+)}(\phi )+\phi _4f_{122}^{(+)}(\phi )\phi _5f_{123}^{(+)}(\phi )`$ $`\phi _2f_{124}^{(+)}(\phi )+\phi _3f_{125}^{(+)}(\phi )+\phi _0f_{126}^{(+)}(\phi )\phi _1f_{127}^{(+)}(\phi ),`$ $`\phi _4`$ $`=`$ $`\phi _6f_{121}^{(+)}(\phi )\phi _5f_{122}^{(+)}(\phi )\phi _4f_{123}^{(+)}(\phi )`$ $`+\phi _3f_{124}^{(+)}(\phi )+\phi _2f_{125}^{(+)}(\phi )\phi _1f_{126}^{(+)}(\phi )\phi _0f_{127}^{(+)}(\phi ).`$ We now solve these equations for the seven unknown $`f_{12i}^{(+)}\left(\phi \right)`$. We find $$f_{121}^{(+)}\left(\phi \right)=f_{122}^{(+)}\left(\phi \right)=0,$$ (3.38) $$f_{123}^{(+)}\left(\phi \right)=\frac{\phi _{0}^{}{}_{}{}^{2}\phi _{6}^{}{}_{}{}^{2}\phi _{5}^{}{}_{}{}^{2}+\phi _{2}^{}{}_{}{}^{2}\phi _{4}^{}{}_{}{}^{2}+\phi _{1}^{}{}_{}{}^{2}+\phi _{3}^{}{}_{}{}^{2}\phi _{7}^{}{}_{}{}^{2}}{r^2}$$ and $`f_{124}^{(+)}\left(\phi \right)`$ $`=`$ $`+2{\displaystyle \frac{\phi _0\phi _7\phi _5\phi _2+\phi _6\phi _1+\phi _3\phi _4}{r^2}},`$ (3.39) $`f_{125}^{(+)}\left(\phi \right)`$ $`=`$ $`2{\displaystyle \frac{\phi _0\phi _6\phi _3\phi _5\phi _1\phi _7\phi _2\phi _4}{r^2}},`$ (3.40) $`f_{126}^{(+)}\left(\phi \right)`$ $`=`$ $`+2{\displaystyle \frac{\phi _0\phi _5\phi _1\phi _4+\phi _7\phi _2+\phi _3\phi _6}{r^2}},`$ (3.41) $`f_{127}^{(+)}\left(\phi \right)`$ $`=`$ $`2{\displaystyle \frac{\phi _0\phi _4+\phi _6\phi _2+\phi _1\phi _5\phi _3\phi _7}{r^2}},`$ (3.42) where $$r^2=(\phi _0^2+\phi _1^2+\phi _2^2+\phi _3^2+\phi _4^2+\phi _5^2+\phi _6^2+\phi _7^2).$$ (3.43) Along the same lines we can calculate all the structure functions, we give all of them in Appendix C. What we have just calculated is commonly called the (+) torsion, we can find the (–) torsion by replacing the left by right multiplication in (3.35) $$[1|𝔼_i,1|𝔼_j]\phi =2f_{ijk}^{()}(\phi )1|𝔼_k\phi .$$ (3.44) we find $$f_{121}^{()}\left(\phi \right)=f_{122}^{()}\left(\phi \right)=0,$$ (3.45) $$f_{123}^{()}\left(\phi \right)=\frac{\phi _{0}^{}{}_{}{}^{2}\phi _{6}^{}{}_{}{}^{2}\phi _{4}^{}{}_{}{}^{2}+\phi _{2}^{}{}_{}{}^{2}+\phi _{1}^{}{}_{}{}^{2}\phi _{7}^{}{}_{}{}^{2}\phi _{5}^{}{}_{}{}^{2}+\phi _{3}^{}{}_{}{}^{2}}{r^2},$$ and $`f_{124}^{()}\left(\phi \right)`$ $`=`$ $`+2{\displaystyle \frac{\phi _0\phi _7+\phi _5\phi _2\phi _6\phi _1\phi _3,\phi _4}{r^2}},`$ (3.46) $`f_{125}^{()}\left(\phi \right)`$ $`=`$ $`2{\displaystyle \frac{\phi _0\phi _6+\phi _3\phi _5+\phi _1\phi _7+\phi _2\phi _4}{r^2}},`$ (3.47) $`f_{126}^{()}\left(\phi \right)`$ $`=`$ $`+2{\displaystyle \frac{\phi _0\phi _5+\phi _1\phi _4\phi _7\phi _2\phi _3,\phi _6}{r^2}},`$ (3.48) $`f_{127}^{()}\left(\phi \right)`$ $`=`$ $`2{\displaystyle \frac{\phi _0\phi _4\phi _6\phi _2\phi _1\phi _5+\phi _3\phi _7}{r^2}},`$ (3.49) the remaining $`f_{ijk}^{()}\left(\phi \right)`$ are listed in appendix C. Let’s pause for a moment and note some of the evident features of these $`f_{ijk}^{(\pm )}(\phi )`$, * One notices immediately that at $`\phi ^t`$=(1,0,0,0,0,0,0,0) / (-1,0,0,0,0,0,0,0), the north / south pole (NP/SP), we recover the octonionic structure constants: $`f_{ijk}^{(+)}(NP/SP)=f_{ijk}^{()}(NP/SP)=f_{ijk}`$ and any non-standard cycles vanishes e.g. $`f_{567}^{(\pm )}(NP/SP)=0`$. * Our construction started from a given multiplication table and as there are different choices , we can have different families. * Restricting ourselves to $`S^3`$, we have the quaternionic structure constants i.e. for $`\phi _0^2+\phi _1^2+\phi _2^2+\phi _3^2=1`$, and $`\phi _4=\phi _5=\phi _6=\phi _7=0`$ we have $`f_{123}(\phi )=ϵ_{123}`$ given in eq. (2.42) and all other $`f`$’s vanish. * Our matrix representation $`\{𝔼_0,𝔼_i\}`$ is 8 dimensional and isomorphic to $`\{\delta _0,\delta _i\}`$. Of course constraining ourselves to a seven sphere of radius $`r,`$ ($`\phi ^\mu \phi _\mu =r^2)`$, $`𝔼_i`$ or $`1|𝔼_i`$ $`\left(i=\mathrm{1..7}\right)`$ are completely legitimate representation of the seven imaginary octonionic units as our $`𝔼_i`$ or $`1|𝔼_i`$ $`\left(i=1,2,3\right)`$ defined in the introduction are a representation of imaginary quaternion units. * Over S<sup>7</sup>, $`_{\text{S}^7}f_{ijk}^{(\pm )}(\phi )0.`$ This is a very important characteristic of the seven sphere. To manifest the $`\phi `$ dependence, let’s give some examples. To simplify the notations, we use here $`(i,j,k)^{(\pm )}`$ for $`f_{ijk}^{(\pm )}(\phi )`$ * As we said before at ($`\phi _0=1,\phi _i=0`$), the nonvanishing cocycles are $$(1,2,3)^{(+)}=(1,4,5)^{(+)}=(1,7,6)^{(+)}=(2,4,6)^{(+)}=(2,5,7)^{(+)}=1$$ $$(3,4,7)^{(+)}=(3,6,5)^{(+)}=1,$$ (3.50) $$(1,2,3)^{()}=(1,4,5)^{()}=(1,7,6)^{()}=(2,4,6)^{()}=(2,5,7)^{()}=1$$ $$(3,4,7)^{()}=(3,6,5)^{()}=1,$$ (3.51) and zero otherwise. Another non-trivial example is at $`\left(\phi _\mu =\frac{\mu +1}{\sqrt{204}}\right)`$, we find $$\begin{array}{c}\begin{array}{ccc}(1,2,3)^{(+)}=12/17,\hfill & (2,5,7)^{(+)}=4/51,\hfill & (1,5,6)^{(+)}=1/51,\hfill \\ (1,4,5)^{(+)}=6/17,\hfill & (1,7,6)^{(+)}=8/51,\hfill & (3,6,5)^{(+)}=0,\hfill \\ (4,3,7)^{(+)}=2/51,\hfill & (4,2,6)^{(+)}=3/17,\hfill & \end{array}\\ \begin{array}{ccc}(1,2,4)^{(+)}=4/17,\hfill & (1,5,2)^{(+)}=8/17,\hfill & (3,5,4)^{(+)}=44/51,\hfill \\ (5,6,7)^{(+)}=10/17,\hfill & (1,3,4)^{(+)}=5/17,\hfill & (4,1,6)^{(+)}=14/17,\hfill \\ (1,5,7)^{(+)}=40/51,\hfill & (3,5,7)^{(+)}=2/17,\hfill & (3,1,6)^{(+)}=14/51,\hfill \\ (2,3,5)^{(+)}=23/51,\hfill & (1,7,4)^{(+)}=4/17,\hfill & (4,5,6)^{(+)}=16/51,\hfill \\ (2,6,5)^{(+)}=38/51,\hfill & (1,6,2)^{(+)}=8/17,\hfill & (6,3,2)^{(+)}=22/51,\hfill \\ (1,3,5)^{(+)}=10/51,\hfill & (2,4,3)^{(+)}=2/17,\hfill & (4,3,6)^{(+)}=20/51,\hfill \\ (3,7,6)^{(+)}=13/17,\hfill & (1,2,7)^{(+)}=1/17,\hfill & (4,2,7)^{(+)}=16/17,\hfill \\ (2,7,3)^{(+)}=16/17,\hfill & (2,6,7)^{(+)}=4/51,\hfill & (7,1,3)^{(+)}=28/51,\hfill \\ (2,4,5)^{(+)}=2/17,\hfill & (4,7,5)^{(+)}=7/51,\hfill & (4,6,7)^{(+)}=10/51,\hfill \end{array}\end{array}$$ (3.52) and $$\begin{array}{c}\begin{array}{ccc}(1,2,3)^{()}=12/17,\hfill & (1,4,5)^{()}=6/17,\hfill & (1,7,6)^{()}=8/51,\hfill \\ (2,4,6)^{()}=3/17,\hfill & (3,5,7)^{()}=8/51,\hfill & (3,4,7)^{()}=2/51,\hfill \\ (4,6,7)^{()}=4/51,\hfill & (1,7,3)^{()}=2/3,\hfill & (1,4,3)^{()}=8/51,\hfill \end{array}\\ \begin{array}{ccc}(1,2,4)^{()}=4/51,\hfill & (1,5,6)^{()}=4/51,\hfill & (1,2,5)^{()}=31/51,\hfill \\ (1,2,7)^{()}=2/51,\hfill & (3,1,5)^{()}=2/51,\hfill & (1,4,6)^{()}=46/51,\hfill \\ (1,4,7)^{()}=3/17,\hfill & (2,5,7)^{()}=4/51,\hfill & (2,4,7)^{()}=50/51,\hfill \\ (2,5,3)^{()}=6/17,\hfill & (4,6,5)^{()}=8/51,\hfill & (1,7,5)^{()}=12/17,\hfill \\ (3,6,1)^{()}=3/17,\hfill & & (3,6,2)^{()}=10/17,\hfill \\ (2,4,5)^{()}=2/51,\hfill & (2,4,3)^{()}=0,\hfill & (3,4,5)^{()}=47/51\hfill \\ (3,4,6)^{()}=6/17,\hfill & (3,6,5)^{()}=0,\hfill & (2,5,6)^{()}=12/17,\hfill \\ (2,3,7)^{()}=3/17,\hfill & (2,6,7)^{()}=0,\hfill & (3,6,7)^{()}=12/17,\hfill \\ (1,2,6)^{()}=6/17,\hfill & (4,5,7)^{()}=0,\hfill & (6,5,7)^{()}=35/51.\hfill \end{array}\end{array}$$ (3.53) We have some kind of dynamical Lie algebra of seven generators with structure “constants” that change their values from one point to another. Let us emphasis the difference between considering $`𝔼_i`$ as an open algebra or as elements of a soft seven sphere, observe that $$\begin{array}{ccc}[𝔼_1,𝔼_2]\hfill & =2𝔼_32[𝔼_1,1|𝔼_2],\hfill & \end{array}$$ (3.54) but $$\begin{array}{ccc}[𝔼_1,𝔼_2]\mathrm{\Phi }\hfill & =2f_{123}^{(+)}\left(\phi \right)𝔼_3\mathrm{\Phi }\hfill & +2f_{124}^{(+)}\left(\phi \right)𝔼_4\mathrm{\Phi }\hfill \\ & +2f_{125}^{(+)}\left(\phi \right)𝔼_5\mathrm{\Phi }\hfill & +2f_{126}^{(+)}\left(\phi \right)𝔼_6\mathrm{\Phi }+2f_{127}^{(+)}\left(\phi \right)𝔼_7\mathrm{\Phi }.\hfill \end{array}$$ (3.55) At the NP $$\mathrm{\Phi }_{NP}^t=\left(\begin{array}{cccccccc}1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \end{array}\right)$$ (3.56) we still have $$[𝔼_1,𝔼_2]=2𝔼_32[𝔼_1,1|𝔼_2]$$ (3.57) whereas $`[𝔼_1,𝔼_2]\mathrm{\Phi }_{NP}`$ $`=`$ $`2f_{12k}^{(+)}(\phi _{NP})𝔼_k\mathrm{\Phi }_{NP}`$ $`=`$ $`2𝔼_3\mathrm{\Phi }_{NP}`$ $`\left(\begin{array}{cccccccc}0\hfill & 0\hfill & 0\hfill & 2\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 2\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 2\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 2\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 2\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 2\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 2\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 2\hfill & 0\hfill & 0\hfill & 0\hfill \end{array}\right)\left(\begin{array}{c}1\hfill \\ 0\hfill \\ 0\hfill \\ 0\hfill \\ 0\hfill \\ 0\hfill \\ 0\hfill \\ 0\hfill \end{array}\right)=\left(\begin{array}{c}0\hfill \\ 0\hfill \\ 0\hfill \\ 2\hfill \\ 0\hfill \\ 0\hfill \\ 0\hfill \\ 0\hfill \end{array}\right)`$ $`=`$ $`\left(\begin{array}{cccccccc}0\hfill & 0\hfill & 0\hfill & 2\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 2\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 2\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 2\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 2\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 2\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 2\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 2\hfill & 0\hfill & 0\hfill & 0\hfill \end{array}\right)\left(\begin{array}{c}1\hfill \\ 0\hfill \\ 0\hfill \\ 0\hfill \\ 0\hfill \\ 0\hfill \\ 0\hfill \\ 0\hfill \end{array}\right).`$ At $`\mathrm{\Phi }_W\left(\phi _\mu =\frac{\mu +1}{\sqrt{204}}\right),`$ we still have $$[𝔼_1,𝔼_2]=2𝔼_32[𝔼_1,1|𝔼_2]$$ (3.60) we find that $`[𝔼_1,𝔼_2]\mathrm{\Phi }_W`$ $`=`$ $`2f_{12k}^{(+)}(\phi _W)𝔼_k\mathrm{\Phi }_W`$ $`=`$ $`\left({\displaystyle \frac{24}{17}}𝔼_3+{\displaystyle \frac{8}{17}}𝔼_4+{\displaystyle \frac{16}{17}}𝔼_5+{\displaystyle \frac{16}{17}}𝔼_6{\displaystyle \frac{2}{17}}𝔼_7\right)\mathrm{\Phi }_W`$ $`\left(\begin{array}{cccccccc}0\hfill & 0\hfill & 0\hfill & 2\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 2\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 2\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 2\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 2\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 2\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 2\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 2\hfill & 0\hfill & 0\hfill & 0\hfill \end{array}\right)\left(\begin{array}{c}\frac{1}{\sqrt{204}}\hfill \\ \frac{2}{\sqrt{204}}\hfill \\ \frac{3}{\sqrt{204}}\hfill \\ \frac{4}{\sqrt{204}}\hfill \\ \frac{5}{\sqrt{204}}\hfill \\ \frac{6}{\sqrt{204}}\hfill \\ \frac{7}{\sqrt{204}}\hfill \\ \frac{8}{\sqrt{204}}\hfill \end{array}\right)=\left(\begin{array}{c}\frac{4}{51}\sqrt{51}\hfill \\ \frac{1}{17}\sqrt{51}\hfill \\ \frac{2}{51}\sqrt{51}\hfill \\ \frac{1}{51}\sqrt{51}\hfill \\ \frac{8}{51}\sqrt{51}\hfill \\ \frac{7}{51}\sqrt{51}\hfill \\ \frac{2}{17}\sqrt{51}\hfill \\ \frac{5}{51}\sqrt{51}\hfill \end{array}\right)`$ $`=`$ $`\left(\begin{array}{cccccccc}0\hfill & 0\hfill & 0\hfill & \frac{24}{17}\hfill & \frac{8}{17}\hfill & \frac{16}{17}\hfill & \frac{16}{17}\hfill & \frac{2}{17}\hfill \\ 0\hfill & 0\hfill & \frac{24}{17}\hfill & 0\hfill & \frac{16}{17}\hfill & \frac{8}{17}\hfill & \frac{2}{17}\hfill & \frac{16}{17}\hfill \\ 0\hfill & \frac{24}{17}\hfill & 0\hfill & 0\hfill & \frac{16}{17}\hfill & \frac{2}{17}\hfill & \frac{8}{17}\hfill & \frac{16}{17}\hfill \\ \frac{24}{17}\hfill & 0\hfill & 0\hfill & 0\hfill & \frac{2}{17}\hfill & \frac{16}{17}\hfill & \frac{16}{17}\hfill & \frac{8}{17}\hfill \\ \frac{8}{17}\hfill & \frac{16}{17}\hfill & \frac{16}{17}\hfill & \frac{2}{17}\hfill & 0\hfill & 0\hfill & 0\hfill & \frac{24}{17}\hfill \\ \frac{16}{17}\hfill & \frac{8}{17}\hfill & \frac{2}{17}\hfill & \frac{16}{17}\hfill & 0\hfill & 0\hfill & \frac{24}{17}\hfill & 0\hfill \\ \frac{16}{17}\hfill & \frac{2}{17}\hfill & \frac{8}{17}\hfill & \frac{16}{17}\hfill & 0\hfill & \frac{24}{17}\hfill & 0\hfill & 0\hfill \\ \frac{2}{17}\hfill & \frac{16}{17}\hfill & \frac{16}{17}\hfill & \frac{8}{17}\hfill & \frac{24}{17}\hfill & 0\hfill & 0\hfill & 0\hfill \end{array}\right)\left(\begin{array}{c}\frac{1}{\sqrt{204}}\hfill \\ \frac{2}{\sqrt{204}}\hfill \\ \frac{3}{\sqrt{204}}\hfill \\ \frac{4}{\sqrt{204}}\hfill \\ \frac{5}{\sqrt{204}}\hfill \\ \frac{6}{\sqrt{204}}\hfill \\ \frac{7}{\sqrt{204}}\hfill \\ \frac{8}{\sqrt{204}}\hfill \end{array}\right).`$ We are not making a projection but a reformulation of the algebra. This fact should always be kept in mind. The same happens in a non-trivial way for the Jacobi identity, i.e. $$\left(f_{ijm}(\phi )f_{mkt}(\phi )+f_{jkm}(\phi )f_{mit}(\phi )+f_{kim}(\phi )f_{mjt}(\phi )\right)𝔼_t\phi =0,$$ (3.63) but, in general, $$\left(f_{ijm}(\phi )f_{mkt}(\phi )+f_{jkm}(\phi )f_{mit}(\phi )+f_{kim}(\phi )f_{mjt}(\phi )\right)0.$$ (3.64) Another important feature is $$[𝔼_i,1|𝔼_j]=2f_{ijk}\left(\phi \right)𝔼_k+[𝔼_i,𝔼_j]=2f_{ijk}\left(\phi \right)1|𝔼_k+[1|𝔼_i,1|𝔼_j]$$ (3.65) which is equal to zero iff $`i=j`$ but for the soft seven sphere, $`[𝔼_i,1|𝔼_j]\phi =0`$ not only for $`i=j`$ but also at the NP/SP for any i,j. Lastly over any group manifold the left torsion equals minus the right torsion, but for $`S^7`$ this is not in general true. ### 3.3 More General Solutions In the previous section , we have used brute force to calculate $`f_{ijk}^{(+)}\left(\phi \right)`$. There is another way, smarter and easier. We have the following situation $$𝔼_i𝔼_j\phi =\left(\delta _{ij}+f_{ijk}^{(+)}\left(\phi \right)𝔼_k\right)\phi $$ (3.66) but our $`𝔼_i`$ defines what is called a pure spinor as we mentioned at the end of chapter II, $$\phi ^t𝔼_i\phi =0$$ (3.67) thus $$\phi ^t\left(𝔼_i𝔼_j\right)\phi =\phi ^t\left(\delta _{ij}\right)\phi ,$$ (3.68) using $$\left(𝔼_k\right)^1=𝔼_k$$ (3.69) we find $$\phi ^t\left(𝔼_k𝔼_i𝔼_j\right)\phi =\phi ^t\left(f_{ijk}^{(+)}\left(\phi \right)\right)\phi $$ (3.70) but $$\phi ^t\phi =r^2$$ (3.71) which gives us $$f_{ijk}^{(+)}\left(\phi \right)=\frac{\phi ^t\left(𝔼_k𝔼_i𝔼_j\right)\phi }{r^2}.$$ (3.72) and $$f_{ijk}^{()}\left(\phi \right)=\frac{\phi ^t\left(1|𝔼_k\mathrm{\hspace{0.33em}\hspace{0.33em}1}|𝔼_i\mathrm{\hspace{0.33em}\hspace{0.33em}1}|𝔼_j\right)\phi }{r^2}.$$ (3.73) There is another interesting property to note $$\phi ^t[𝔼_i,1|𝔼_j]\phi =0$$ (3.74) which may be the generalization of the standard Lie algebra relation, left and right action commute everywhere over the group manifold. The left and right torsions that we have constructed are not the only parallelizable torsions of S<sup>7</sup>. Our $`𝔼_i`$ and $`1|𝔼_i`$ are given in terms of the octonionic structure constants (3.27) i.e. the torsion at NP/SP. Considering two new points, we may define new sets of $`𝔼_i`$ and $`1|𝔼_i`$. As S<sup>7</sup> contains an infinity of points, practically, we have an infinity of parallelizable torsion. If our method is self contained and sufficient, we should be able to construct these infinity of pointwise structures. Indeed, $`𝔼_i\left(\phi \right)`$ and $`1|𝔼_i\left(\phi \right)`$ are in general $$\begin{array}{ccccc}\delta _i& & (𝔼_i(\phi ))_{\mu \nu }& =& \delta _{0\mu }\delta _{i\nu }\delta _{0\nu }\delta _{i\mu }f_{i\mu \nu }^{(+)}(\phi ),\\ 1|\delta _i& & (1|𝔼_i(\phi ))_{\mu \nu }& =& \delta _{0\mu }\delta _{i\nu }\delta _{0\nu }\delta _{i\mu }+f_{i\mu \nu }^{()}(\phi ),\end{array}$$ (3.75) in complete analogy with (3.23,3.24,3.25,3.26). Of course the soft Algebra idea should hold here as well as for the special $`(𝔼_i,1|𝔼_i)`$ constructed in terms of the north pole torsion. Repeating the calculation in terms of $`(𝔼_i(\phi ),1|𝔼_i(\phi ))`$. Let us introduce a new vector field $`\lambda `$, $$\lambda ^t=\left(\begin{array}{cccccccc}\lambda _0\hfill & \lambda _1\hfill & \lambda _2\hfill & \lambda _3\hfill & \lambda _4\hfill & \lambda _5\hfill & \lambda _6\hfill & \lambda _7\hfill \end{array}\right).$$ (3.76) We define two new generalized structure functions $`[𝔼_i(\phi ),𝔼_j(\phi )]\lambda `$ $`=`$ $`2f_{ijk}^{(++)}(\phi ,\lambda )𝔼_k(\phi )\lambda `$ (3.77) $`[1|𝔼_i(\phi ),1|𝔼_j(\phi )]\lambda `$ $`=`$ $`2f_{ijk}^{()}(\phi ,\lambda )1|𝔼_k(\phi )\lambda `$ (3.78) where $`f_{ijk}^{\left(\pm \pm \right)}(\phi ,\lambda )`$ have a very complicated structure, $`f_{ijk}^{\left(++\right)}(\phi ,\lambda )`$ $`=`$ $`{\displaystyle \frac{\lambda ^t\left(𝔼_k\left(\phi \right)𝔼_i\left(\phi \right)𝔼_j\left(\phi \right)\right)\lambda }{r^2}},`$ (3.79) $`f_{ijk}^{\left(\right)}(\phi ,\lambda )`$ $`=`$ $`{\displaystyle \frac{\lambda ^t\left(1|𝔼_k\left(\phi \right)\mathrm{\hspace{0.33em}\hspace{0.33em}1}|𝔼_i\left(\phi \right)\mathrm{\hspace{0.33em}\hspace{0.33em}1}|𝔼_j\left(\phi \right)\right)\lambda }{r^2}}`$ (3.80) as examples, we list four of them in Appendix D. We will use them later when we study some applications. ### 3.4 Some Group Theory An arbitrary octonion can be associated to $`^8=^7`$ where $``$ denotes the subspace spanned by the identity $`e_0=1`$. Octonions with unit length define the octonionic unit sphere $`S^7`$. The isometries of octonions is described by $`O(8)`$ which may be decomposed as $$O(8):HKE$$ (3.81) where $`H`$ is the 14 parameters $`G_2`$ algebra of the automorphism group of $`octonions,`$ K is the torsionful seven sphere $`SO(7)/G_2`$ and our E is the round seven sphere $`SO(8)/SO(7)`$. In fact the different three non-equivalent representation of O(8) - the vectorial so(8) and the two different spinorial $`spin^L(8)`$ and $`spin^R(8)`$, which are related by triality, can be realized by suitable left and right octonionic multiplication. The reduction of O(8) to O(7) induces $`so(8)so(7)1`$, $`spin^R(8)spin(7)`$ and $`spin^L(8)spin(7)`$. We would like to show how to generate these different Lie algebras entirely from our canonical left/right octonionic structures. We start from the $`8\times 8`$ gamma matrices $`\gamma _{\mu \nu }^i`$ in seven dimensions, using $`\delta _{ij}`$ as our flat Euclidean metric, $$\{\gamma ^i,\gamma ^j\}=2\delta ^{ij}\mathrm{𝟏}_8,$$ (3.82) where $`i,j,\mathrm{}=1,2,\mathrm{}7`$ and $`\mu ,\nu ,\mathrm{}=0,1,2,\mathrm{}7`$. We can use either of the following choices $$\gamma _+^j=i𝔼_jor\gamma _{}^j=i1|𝔼_j,$$ (3.83) of course the $`i`$ in the right hand sides is the imaginary complex unit. This relates our antisymmetric, Hermitian and hence purely imaginary gamma matrices to the canonical octonionic left/right structures. The antisymmetric product of two gamma matrices will be denoted by $$\gamma ^{ij}=\gamma ^{[i}\gamma ^{j]},$$ (3.84) and we have <sup>1</sup><sup>1</sup>1It is interesting to note that this equation may be used as an alternative definition for the octonionic multiplication table. $$\gamma ^i\gamma ^j\gamma ^k=\frac{1}{4!}ϵ^{ijklmnp}\gamma ^l\gamma ^m\gamma ^n\gamma ^p.$$ (3.85) The matrices $`\gamma ^{ij}`$ span the 21 generators $`J^{ij}`$of spin(7) in its eight-dimensional spinor representation. The spinorial representation of spin(7) can be enlarged to the left/right handed spinor representation of spin(8) by different ways. The easiest one is to include either of $`\pm 𝔼_i`$ or $`\pm 1|𝔼_i`$ defining $`J^i=J^{i0}`$, so(8) can be written as $`[J^i,J^i]`$ $`=`$ $`2J^{ij}`$ (3.86) $`[J^i,J^{mn}]`$ $`=`$ $`2\delta ^{im}J^n2\delta ^{in}J^m`$ (3.87) $`[J^{ij},J^{kl}]`$ $`=`$ $`2\delta ^{jk}J^{il}+2\delta ^{il}J^{jk}2\delta ^{ik}J^{jl}2\delta ^{jl}J^{ik}.`$ (3.88) The automorphism group of octonions is $`G_2SO(7)SO(8)`$. A suitable basis for $`G_2`$ is $$H_{ij}=f_{ijk}\left(𝔼_k1|𝔼_k\right)\frac{3}{2}[𝔼_i,1|𝔼_j],$$ (3.89) which implies the linear relations $$f_{ijk}H_{jk}=0,$$ (3.90) These constraints enforce $`H_{ij}`$ to generate the 14 dimensional vector space of $`G_2`$. There are different ways to represent the remaining seven generators, denoted here by K, $`{\displaystyle \frac{so(7)}{G_2}}`$ $`:`$ $`K_v^{\pm i}=\pm {\displaystyle \frac{1}{2}}\left(𝔼_i1|𝔼_i\right),`$ (3.91) $`{\displaystyle \frac{spin(7)}{G_2}}`$ $`:`$ $`K_s^{\pm i}=\pm \left({\displaystyle \frac{1}{2}}𝔼_i+1|𝔼_i\right),`$ (3.92) $`{\displaystyle \frac{\overline{spin(7)}}{G_2}}`$ $`:`$ $`\overline{K}_s^{\pm i}=\left(𝔼_i+{\displaystyle \frac{1}{2}}1|𝔼_i\right),`$ (3.93) Defining the conjugate representation<sup>2</sup><sup>2</sup>2n.b. this definition is not matrix conjugation. by $$\overline{𝔼}=1|𝔼\text{and }\overline{1|𝔼}=𝔼,$$ (3.94) (3.91) is self-conjugate while (3.92) is octonionic-conjugate to (3.93). The vector representation so(7) generated by $`H_{ij}K_v^{\pm i}`$ is seven dimensional because $`K_v^{\pm i}e_0=0`$ whereas the spin(7) representation generated by $`H_{ij}K_s^{\pm i}`$ is eight dimensional. To make apparent the role of the automorphism group $`G_2`$, the different commutators of $`𝔼`$ and $`1|𝔼`$ may be written as $`[𝔼_i,𝔼_j]`$ $`=`$ $`{\displaystyle \frac{1}{3}}\left(4H_{ij}+2f_{ijk}𝔼_k+4f_{ijk}1|𝔼_k\right),`$ (3.95) $`[1|𝔼_i,1|𝔼_j]`$ $`=`$ $`{\displaystyle \frac{1}{3}}\left(4H_{ij}4f_{ijk}𝔼_k2f_{ijk}1|𝔼_k\right),`$ (3.96) $`[𝔼_i,1|𝔼_j]`$ $`=`$ $`{\displaystyle \frac{1}{3}}\left(2H_{ij}+2f_{ijk}𝔼_k2f_{ijk}1|𝔼_k\right).`$ (3.97) or $`G_2`$ given by $$H_{ij}=\frac{1}{2}([𝔼_i,𝔼_j]+[1|𝔼_i,1|𝔼_j]+[𝔼_i,1|𝔼_j]).$$ (3.98) Thus as we promised, the $`𝔼`$ and $`1|𝔼`$ are the necessary and the sufficient building blocks for expressing the different Lie algebras and coset representations related to the seven sphere. Note that all the constructions given in this section start from the Clifford algebra relation (3.83), and the formulation holds equally for $`𝔼\left(\phi \right)`$ and $`1|𝔼\left(\phi \right)`$. ## Chapter 4 Soft Seven Sphere Self Duality The word instanton has been coined for solutions of elliptic non linear field equations in Euclidean space time, with boundary conditions at infinity in such a way that stable topological properties emerge . The study of Euclidean Yang-Mills fields involves many mathematical items falling under the headings : differential geometry (fiber bundles, connections), differential topology (characteristic classes, index theory), and algebraic geometry (twistors, holomorphic bundles) which makes it a rich and unique subject. We review the standard $`d=4`$ dimensional instanton solution in the first section. In higher dimensions $`d>4`$, there is no unique way to define self-duality. It is a matter of prejudice. One tries to conserve as much as possible of the four dimensions duality characteristics. Particularly in 8 dimensions, there have been a lot of proposals ,. There are some relationships between these apparently distinct constructions. Each author of these proposals concentrated on a certain aspect of the 4 dimensional self-duality which they considered a good starting point for the generalization to 8 dimensions. In the second section of this chapter, we list the main features of the GKS instanton, then in the last section we present the soft seven sphere instanton. We will show how the soft seven sphere can be used to reformulate the GKS in a way very similar to the four dimensional case. ### 4.1 The 4 Dimensional Instanton Consider an $`SU\left(2\right)`$ classical gauge field over four dimensional Euclidean space $`^4`$. Let the gauge potential be $`A_\mu =A_\mu ^at^a`$, while the field strength is defined as the commutator of the covariant derivative $`\left(D_\mu _\mu +A_\mu \right),`$ $$F_{\mu \nu }[D_\mu ,D_\nu ]=F_{\mu \nu }^at^a=_\mu A_\nu _\nu A_\mu +[A_\mu ,A_\nu ],$$ (4.1) where the generators of the Lie algebra are $`t^a=i\frac{\sigma ^a}{2},`$ and $`[t^a,t^b]=iϵ^{abc}t^c`$. We want to find the $`A_\mu ^a`$ which minimizes the Euclidean action $$S_E=d^4x\frac{1}{4}F_{\mu \nu }^aF^{a\mu \nu }=\frac{1}{2}d^4xtr\left(F_{\mu \nu }F^{\mu \nu }\right).$$ (4.2) We rewrite it as $$S_E=\frac{1}{8}d^4x\left\{\left[F_{\mu \nu }^a\pm {}_{}{}^{}F_{\mu \nu }^{a}\right]^22F_{\mu \nu }^a{}_{}{}^{}F_{}^{a\mu \nu }\right\}$$ (4.3) where the dual field strength $`{}_{}{}^{}F`$ is defined by (the four dimensional Levi–Cevita tensor $`ϵ_{0123}=1`$) $${}_{}{}^{}F_{\mu \nu }^{a}=\frac{1}{2}ϵ_{\mu \nu \alpha \beta }F^{a\alpha \beta }.$$ (4.4) Hence the lower bound for the action is $$S_E\frac{1}{4}d^4xF_{\mu \nu }^a{}_{}{}^{}F_{}^{a\mu \nu }$$ (4.5) The equality sign holds iff $$F_{\mu \nu }^a\pm {}_{}{}^{}F_{\mu \nu }^{a}=0$$ (4.6) which is the self/antiself duality condition (SD/ASD). The self/antiself duality condition when combined with the Bianchi identities $$D_\mu F_{\nu \rho }+D_\rho F_{\mu \nu }+D_\nu F_{\rho \mu }=0ϵ^{\omega \mu \nu \rho }D_\mu F_{\nu \rho }=0,$$ (4.7) yields the equations of motion $$ϵ^{\omega \mu \nu \rho }D_\mu F_{\nu \rho }=D_\mu ^{}F^{\mu \omega }=\pm D_\mu F^{\mu \omega }=0.$$ (4.8) Since the self/antiself duality equation is only first order in derivatives, it is much easier to solve than the second order field equations. In order for $`S_E`$ to remain finite, we require $$F_{\mu \nu }\stackrel{\left|x\right|\mathrm{}}{}0$$ (4.9) which is automatically valid if $$A_\mu \stackrel{\left|x\right|\mathrm{}}{}g^1\left(x\right)_\mu g\left(x\right)$$ (4.10) where $`g\left(x\right)SU\left(2\right)`$ is a gauge transformation. Thus we have the following situation: In general $`F_{\mu \nu }0`$ inside a volume $`^4`$, but vanishes at the infinite boundary $`E^4=S_{\mathrm{}}^3`$, a three dimensional sphere, where the gauge potential $`A_\mu `$ approaches a pure gauge. At the boundary $`xS_{\mathrm{}}^3`$, our gauge transformation $`g\left(x\right)SU\left(2\right),`$are mappings $$g:S_{\mathrm{}}^3SU\left(2\right)S^3.$$ (4.11) These mappings are classified according to the homotopy classes determined by the topological winding number $`w`$ $$\pi _3\left(SU\left(2\right)\right)\pi _3\left(S^3\right)\left\{w\right\}=.$$ (4.12) The topological number $`w`$ is related to the Chern number $$Ch_2=\frac{1}{8\pi ^2}d^4xϵ^{\mu \nu \alpha \beta }F_{\mu \nu }^aF_{\alpha \beta }^a.$$ (4.13) Defining the topological charge $$N=d^4xn(x)n(x)=\frac{e^2}{32\pi ^2}F_{\mu \nu }^a{}_{}{}^{}F_{}^{a\mu \nu }$$ (4.14) this is then an integer, since it is the topological number corresponding to the mapping of the three-dimensional sphere into the gauge group $`SU(2)`$. One can then get the following expression for the euclidean action if (4.6) is satisfied $$S_E=\frac{8\pi ^2}{e^2}N.$$ (4.15) The simplest solution to the ASD condition ($`F_{\mu \nu }=\frac{1}{2}ϵ_{\mu \nu \alpha \beta }F^{\alpha \beta }`$), corresponding to a value of the topological charge $`N=1`$, is given by: $$A_\mu ^a=\frac{x^2}{x^2+\lambda ^2}\left(g^1\left(x\right)_\mu g\left(x\right)\right)^a=\frac{2\left(𝔼^a\right)_{\mu \nu }x^\nu }{x^2+\lambda ^2}$$ (4.16) where $$g\left(x\right)=\frac{x_0\mathrm{𝟏}_4+ix.\sigma }{\sqrt{x^2}}$$ (4.17) and $`𝔼^a`$ are the canonical left quaternionic structures given in eq.(2.47). Leading to $$F_{\mu \nu }^a=\frac{4\lambda ^2\left(𝔼^a\right)_{\mu \nu }}{\left(x^2+\lambda ^2\right)^2}.$$ (4.18) Such a solution has a natural quaternionic formulation . Consider $$g\left(x\right)=\frac{x^\mu e_\mu }{\left|x\right|}$$ (4.19) where $`e_\mu `$ are our four quaternionic units, introducing the self dual $`SO\left(4\right)`$ basis $$\vartheta _{\mu \nu }=\frac{1}{2}\left(\overline{e}_\mu e_\nu \overline{e}_\nu e_\mu \right)$$ (4.20) such that $`\left(\overline{e}_0=e_0,\overline{e}_i=e_i\right)`$, we find $$\vartheta _{\mu \nu }=\frac{1}{2}ϵ_{\mu \nu \alpha \beta }\vartheta ^{\alpha \beta }$$ (4.21) and $$[\vartheta _{\mu \nu },\vartheta _{\alpha \beta }]=2\left(\delta _{\alpha \mu }\vartheta _{\beta \nu }\delta _{\alpha \nu }\vartheta _{\beta \mu }\delta _{\beta \mu }\vartheta _{\alpha \nu }+\delta _{\beta \nu }\vartheta _{\alpha \mu }\right).$$ (4.22) For $`A_\mu `$ given by $$A_\mu =\frac{x^2}{x^2+\lambda ^2}\left(g^1\left(x\right)_\mu g\left(x\right)\right)=\frac{2\vartheta _{\mu \nu }x^\nu }{x^2+\lambda ^2}$$ (4.23) we have $$F_{\mu \nu }=\frac{4\lambda ^2\vartheta _{\mu \nu }}{\left(x^2+\lambda ^2\right)^2}F_{\mu \nu }=\frac{1}{2}ϵ_{\mu \nu \alpha \beta }F^{\alpha \beta }.$$ (4.24) The self-dual solution $`F_{\mu \nu }=\frac{1}{2}ϵ_{\mu \nu \alpha \beta }F^{\alpha \beta }`$ can be found simply by replacing $`e_\mu 1|e_\mu `$ leading to $$g\left(x\right)=\frac{x^\mu 1|e_\mu }{\left|x\right|}.$$ (4.25) The corresponding $`SO\left(4\right)`$ antiself–dual basis are $$\overline{\vartheta }_{\mu \nu }=\frac{1}{2}\left(1|\overline{e}_\mu 1|e_\nu 1|\overline{e}_\nu 1|e_\mu \right)\overline{\vartheta }_{\mu \nu }=\frac{1}{2}ϵ_{\mu \nu \alpha \beta }\overline{\vartheta }^{\alpha \beta }.$$ (4.26) Then with the choice $$A_\mu =\frac{x^2}{x^2+\lambda ^2}\left(g^1\left(x\right)_\mu g\left(x\right)\right)=\frac{2\overline{\vartheta }_{\mu \nu }x^\nu }{x^2+\lambda ^2},$$ (4.27) we have $$F_{\mu \nu }=\frac{4\lambda ^2\overline{\vartheta }_{\mu \nu }}{\left(x^2+\lambda ^2\right)^2}F_{\mu \nu }=\frac{1}{2}ϵ_{\mu \nu \alpha \beta }F^{\alpha \beta }$$ (4.28) The higher $`n>1`$ instantons, Atiyah, Drinfield, Hitchin and Manin (ADHM) solutions, are given naturally in terms of quaternions . Lastly, we would like to show how to get static monopole solutions by field redefinition of the instanton problem. We consider only static monopoles, purely magnetic and which are solutions of an equation called the Bogomolny condition. The model is defined over $`^3`$. Such monopoles are called BPS states, they correspond to an $`su\left(2\right)`$ valued pair of a gauge field $`A_i`$ $`\left(i=1,2,3\right)`$ and a scalar field $`\varphi `$. The action of the monopole system is $$=\frac{1}{2}d^3x\left[B^iB_i+\left(D_i\varphi \right)\left(D^i\varphi \right)\lambda \left(\varphi ^2a^2\right)^2\right],$$ (4.29) $`B_i`$ and $`D_i\varphi `$ are defined by $$B_i=ϵ_{ijk}\left(^jA^k+A^jA^k\right),D_i\varphi =_i\varphi +[A_i,\varphi ].$$ (4.30) The field equations for this static system are $`ϵ_{ijk}D^jB^k`$ $`=`$ $`[\varphi ,D_i\varphi ],`$ (4.31) $`D^iD_i\varphi `$ $`=`$ $`2\lambda \varphi ^32\lambda a^2\varphi `$ (4.32) whereas the Bianchi identiy is $$D_iB_i=0,$$ (4.33) Being a second order system of partial differential equations (4.314.33), they are hard to solve. However, if we take the limit $`\varphi ^2=a^2`$, we impose a first order equation $$B_i=D_i\varphi $$ (4.34) which is the Bogomolny equation. We can relate the self duality condition (4.6) to the Bogomolny condition if we redefine $`\varphi `$ as a fourth component $`A_0`$ of the gauge field $`A_i`$ i.e. $`A_\mu \left(A_0=\varphi ,A_i\right)`$. We then have $$D_i\varphi =F_{i0},B_i=\frac{1}{2}ϵ_{ijk}F^{jk}$$ (4.35) by substitution in (4.34), we recover the self duality condition. ### 4.2 The Grossman–Kephart–Stasheff Instanton The four dimensional self–duality notion proved to be a very powerful tool both of physics and mathematics, so it is natural to investigate the occurrence of a similar condition in higher dimensions. As we have already said, there is no standard way to express the self duality equation in $`d>4`$ dimensions. In a generic $`d`$ dimensions, a $`p`$ form $`v`$ is $$v=\frac{1}{p!}v_{\alpha _1\mathrm{}\alpha _p}dx^{\alpha _1}\mathrm{}dx^{\alpha _p},\alpha _1,\mathrm{},\alpha _p=1\mathrm{}d$$ (4.36) and the dual form is $${}_{}{}^{}v=\frac{1}{d!}ϵ^{\alpha _1\mathrm{}\alpha _d}v_{\alpha _1\mathrm{}\alpha _p}dx^{\alpha _{p+1}}\mathrm{}dx^{\alpha _d}.$$ (4.37) which means, in $`d`$ dimensions the dual of a $`p`$ form is a $`dp`$ form. Knowing that the Yang-Mills field strength can be written as a two form $$F=\frac{1}{2}F_{\alpha _1\alpha _2}dx^{\alpha _1}dx^{\alpha _2},$$ (4.38) then the dual of a 2 form is another 2 form iff $`d=4`$. Constraining ourselves to Yang-Mills models, we have at disposal just the one form gauge field $`A=A_idx^i`$ or the 2 form field strength $`F.`$ In eight dimensions, it is not obvious how we should proceed. There have been different suggestions. For example: * The Fubini–Nicolai or the Corrigan-Devchand-Fairlie-Nuyts instanton: the authors insist on the existence of “squaring” i.e. the action can be written as the square of self-dual fields. * There exist also some promising generic higher dimensional self-duality conditions that are not just restricted to 8 dimensions but also go beyond that limit. For example, the Ivanon-Popov proposal where a Clifford-algebraic structure is used. Another example is the geometric Bais-Batenburg self-duality which is based upon hypercomplex structures over appropriate manifolds. Here, we would like to concentrate on another proposal. Grossman–Kephart-Stasheff suggested a condition, for eight dimensions, that has deep topological roots: The last Hopf map $`S^{15}\stackrel{S^7}{}S^8`$, conformal invariance and spin structure over $`S^8`$. Working over the 8 dimensional euclidean space $`^8`$. The (GKS) self duality condition is (the eight dimensional Levi–Cevita tensor $`ϵ_{01234567}=1`$) $$F_{\alpha _1\alpha _2}^aF_{\alpha _3\alpha _4}^a=\frac{1}{4!}ϵ_{\alpha _1\mathrm{}\alpha _8}F^{a\alpha _5\alpha _6}F^{a\alpha _7\alpha _8}.$$ (4.39) where there is summation over the Lie algebra indices $`a`$. Grossman, Kephart and Stasheff insisted upon the conformal invariance of the Yang-Mills action in 4 dimensions. In eight dimensions the Yang-Mills action is not conformally invariant hence they considered the functional $$𝒜=d^8x\left(F_{\mu _1\mu _2}^aF_{\mu _3\mu _4}^aF^{b\mu _1\mu _2}F^{b\mu _3\mu _4}\right)$$ (4.40) which upon the use of the self duality condition (4.39), takes a form similar to the fourth Chern class $$_{S^8}ϵ_{\alpha _1\mathrm{}\alpha _8}F^{a\alpha _1\alpha _2}F^{a\alpha _3\alpha _4}F^{b\alpha _5\alpha _6}F^{b\alpha _7\alpha _8}.$$ (4.41) In the search for solutions of the GKS duality condition and requiring that $`F_{\mu _1\mu _2}\stackrel{S_{\mathrm{}}^7\left\{\left|x\right|\mathrm{}\right\}}{}0`$, so that $`A_{\mu _1}`$ must be a pure gauge at infinity $$A_\mu \stackrel{S_{\mathrm{}}^7}{}g^1\left(x\right)_\mu g\left(x\right).$$ (4.42) where $`g\left(x\right)`$ is a gauge transformation. GKS assumed the following form for $`g\left(x\right)`$ $$g\left(x\right)=\frac{x^\mu 𝔼_\mu }{\sqrt{x^2}}$$ (4.43) where $`𝔼_\mu `$ are given by (3.27), whence, $$g^{}\left(x\right)g\left(x\right)=1andDet\left(g\left(x\right)\right)=1.$$ (4.44) Note that, for $`\widehat{x}^\mu =\frac{x^\mu }{\sqrt{x^2}},`$ we have $`\widehat{x}^\mu \widehat{x}_\mu =1`$ i.e. $`g\left(x\right)`$ parameterize a unit $`S^7.`$ For the boundary condition given above, we have the following situation: $$g\left(x\right):S_{\mathrm{}}^7S^7$$ (4.45) Such a class of maps are classified according to the seventh homotopy group of the seven sphere $$\pi _7\left(S^7\right)=n$$ (4.46) i.e. there is no map between solutions of different n. They lie in different classes. In particular, there can be no map between the trivial $`n=0`$ ($`A_\mu =0`$) field configuration and $`n>0`$ ($`A_\mu 0`$). Any $`n0`$ is stable and will never decay. GKS proposed the following ansatz for $`A_\mu `$ ($`n=1`$) solution $$A_\mu =\frac{x^2}{x^2+\lambda ^2}\left(g^1\left(x\right)_\mu g\left(x\right)\right).$$ (4.47) which solves (4.39). Now, let’s mention the difference between this GKS duality and the standard four dimensional duality considered in the previous section. * It is not derived from an action. $`𝒜`$ has no quadratic term in the derivatives i.e. there is no kinetic term. * The GKS duality is valid only for a specific representation $`8`$ of a specific group $`SO\left(8\right)`$ in contrast to the self–duality condition which is well defined for any representation of any simple Lie group. * The GKS solution is not a solution of the $`^8`$ Yang–Mills field equations. ### 4.3 Eight Dimensional Soft Self Duality Now we would like to reformulate in a quadratic form the GKS self duality condition. We work over $`^8`$. Contrary to the standard Yang-Mills gauge field, the soft gauge field strength carries a dependence upon the internal manifold. It is important to know where we stand over the $`S_{gauge}^7`$, as the structure functions vary from one point to another. For a soft gauge field $`A_\mu (x)`$ $`A_\mu ^i𝔼_i`$, the insertion of the $`\phi `$ is essential for the closure of the commutator $`[A_\mu (x),A_\nu (x)]\phi `$ $``$ $`A_\mu ^i(x)A_\nu ^j(x)[𝔼_i,𝔼_j]\phi `$ (4.48) $`=`$ $`2f_{ijk}^{(+)}\left(\phi \right)A_\mu ^i(x)A_\nu ^j(x)𝔼_k\phi .`$ (4.49) thus the field strength is given by $`F_{\mu \nu }(x,\phi )`$ $`=`$ $`F_{\mu \nu }(x)\phi `$ (4.50) $`=`$ $`F_{\mu \nu }^i\left(x\right)𝔼_i\phi `$ $`=`$ $`\left({\displaystyle \frac{A_\nu (x)}{x^\mu }}{\displaystyle \frac{A_\mu (x)}{x^\nu }}+[A_\mu (x),A_\nu (x)]\right)\phi .`$ The critical point for the self-duality condition is the existence of a fourth rank tensor. Adding a zero index to extend $`f_{ijk}^{(\pm )}(\phi )`$ from $`^7`$ to $`^8`$, we define a fourth rank tensor $`\eta _{\alpha \beta \mu \nu }(\phi )`$ which is equal to $$\eta _{0ijk}^{(\pm )}(\phi )=f_{ijk}^{(\pm )}(\phi ),$$ (4.51) and zero elsewhere. The proposed generalization of the four dimensional self duality is the following *soft self duality condition* $$F(x,\phi )=F(x,\phi ),$$ (4.52) or in terms of components $$F_{0i}(x,\phi )=\frac{1}{2}\eta _{0ijk}^{(\pm )}(\phi )F^{jk}(x,\phi ),$$ (4.53) note that $`\eta _{\alpha \beta \mu \nu }(\phi )`$ varies over the seven sphere. To proceed, we require the vanishing of $`F_{\mu \nu }`$ at the infinite $`S_{\mathrm{}}^7`$, thus $`A_\mu `$ (at $`S_{\mathrm{}}^7`$) must be a pure gauge $`A_\mu =g^1\left(x\right)_\mu g\left(x\right)`$, where our gauge transformation $`g\left(x\right)`$ is a map from the spatial $`S_{\mathrm{}}^7`$ to the gauge space $`S_{gauge}^7`$ <sup>1</sup><sup>1</sup>1Soft seven sphere gauge transformations reduce to the standard Yang-Mills theory at any single point over the seven sphere. The soft seven sphere gauge field $`A_\mu `$ transforms as $`A\mu \left(x\right)\phi g^1\left(A_\mu \left(x\right)+_\mu \right)g\phi `$ and the Field strength $`F_{\mu \nu }\left(x\right)`$ transforms as $`F_{\mu \nu }(x,\phi )g^1\left(F_{\mu \nu }\left(x\right)\right)g\phi .`$ The presence of the $`\phi `$ is essential.. Consider an $`S^7`$ element $$g(x)=\frac{𝔼_\mu x^\mu }{\sqrt{x^2}},$$ (4.54) the self-dual gauge solution of the self dual condition is exactly the GKS ansatz $$A_\mu ^{(+)}(x)=\frac{x^2}{x^2+\lambda ^2}g^1(x)_\mu g(x)=\frac{\mathrm{\Xi }_{\mu \nu }^{(+)}x^\nu }{x^2+\lambda ^2}$$ (4.55) where the $`\mathrm{\Xi }_{\mu \nu }^{(+)}`$ is given by $$\mathrm{\Xi }_{\mu \nu }^{(+)}=\frac{1}{2}\left(𝔼_\mu ^t𝔼_\nu 𝔼_\nu ^t𝔼_\mu \right).$$ (4.56) We call $`\mathrm{\Xi }_{\mu \nu }^{(+)}`$ the self dual tensor, because $$\mathrm{\Xi }_{oi}^{(+)}\phi =\frac{1}{2}\eta _{0ijk}^{(+)}(\phi )\mathrm{\Xi }^{(+)jk}\phi .$$ (4.57) After substituting $`A_\mu ^{(+)}(x)`$ into (4.50), we find $$F_{\mu \nu }^{(+)}(x,\phi )=2\frac{(\mathrm{\Xi }_{\mu \nu }^{(+)})\lambda ^2}{(\lambda ^2+x^2)^2}\phi $$ (4.58) which is obviously soft self dual (4.53). Another problem of the GKS instanton that can be overcame in the soft seven sphere framework is the compatibility of the equation of motion and the self duality condition. We find by explicit calculation that our solution (4.55) satisfies the Yang-Mills equation of motion for a soft gauge field $$D_\mu F_{\mu \nu }\left(x\right)\phi =_\mu F_{\mu \nu }\left(x\right)\phi +[A_\mu \left(x\right),F_{\mu \nu }\left(x\right)]\phi =0.$$ (4.59) Of course the four dimensional case is more powerful because the self duality is related directly to the Bianchi identity which does not hold in higher dimensions. However, in our case the soft self duality is compatible with the equation of motion whereas for Grossman-Kephart-Stasheff instanton, one must work over curved space–time with certain condition for the metric in order to satisfy both the self duality and the equation of motion. To construct a static seven dimensional monopole, we proceed by static dimensional reduction from $`^8`$ to $`^7`$. Identifying $`A_0`$ $`=`$ $`\varphi ,`$ (4.60) $`F_{ij}`$ $`=`$ $`f_{ijk}\left(\phi \right)B^ki,j,k=\mathrm{1..7}`$ (4.61) then the self duality will be reduced to $$\left(B_i\right)\phi =\left(D_i\varphi \right)\phi .$$ (4.62) Using the soft seven sphere, we can easily generate new solutions of GKS dualities. Working with $`𝔼\left(\phi \right)`$, replacing and $`g\left(x\right)`$ by $$g(x,\phi )=\frac{𝔼_\mu \left(\phi \right)x^\mu }{\sqrt{x^2}}$$ (4.63) the resulting gauge field given in terms $$\mathrm{\Xi }_{\mu \nu }^{\left(++\right)}\left(\phi \right)=\frac{1}{2}\left(𝔼_\mu ^t\left(\phi \right)𝔼_\nu \left(\phi \right)𝔼_\nu ^t\left(\phi \right)𝔼_\mu \left(\phi \right)\right)$$ (4.64) leads to $`A_\mu ^{(++)}(x)`$ $`=`$ $`{\displaystyle \frac{x^2}{x^2+\lambda ^2}}g^1(x,\phi )_\mu g(x,\phi )={\displaystyle \frac{\mathrm{\Xi }_{\mu \nu }^{\left(++\right)}\left(\phi \right)x^\nu }{x^2+\lambda ^2}}`$ (4.65) $`F_{\mu \nu }^{\left(++\right)}(x,\phi )`$ $`=`$ $`2{\displaystyle \frac{(\mathrm{\Xi }_{\mu \nu }^{\left(++\right)}\left(\phi \right))_{\mu \nu }\lambda ^2}{(\lambda ^2+x^2)^2}}`$ (4.66) which also satisfy (4.39) and (4.53). ## Chapter 5 Hypercomplex SSYM Models Day after day, supersymmetry consolidates its position in theoretical physics. Even if it was introduced more than 25 years ago, there are still problems with the geometric basis of extended $`\left(N>1\right)`$ supersymmetry. The situation of the extended superspace is far less satisfactory than the original N=1 superspace. At the level of the algebra the on-shell formalism closes up to modulo of the classical equations of motion. This fact seems odd at the quantum level since the equations of motion receive loop corrections<sup>1</sup><sup>1</sup>1Also, the supersymmetry transformations receive corrections and one should test the closure of the algebra order by order in perturbation theory.. The superspace introduces an elegant supermanifold with different enlarged superconnections, where some are truly integrable in the sense of having zero supercurvature. In principle, the extended superspace should be a very powerful tool for quantum calculations. Before starting, we feel obliged to mention something about the history of the following conjecture: Ring Division Algebras $`𝕂`${ real $`,`$complex $`,`$quaternions $`,`$octonions $`𝕆`$ } are relevant to simple supersymmetric Yang-Mills. The first hint, as mentioned by Schwarz comes from the number of propagating Bose and Fermi degrees of freedom which is one for $`d=3`$, two for $`d=4`$, four for $`d=6`$ and eight for $`d=10`$ suggesting a correspondence with real $``$, complex $``$, quaternions $``$ and octonions $`𝕆`$. Kugo and Townsend investigated in detail the relationship between $`𝕂`$ and the irreducible spinorial representation of the Lorentz group in $`d=3,4,6,10`$, building upon the following chain of isomorphisms $`so(2,1)`$ $``$ $`sl(2,)`$ $`so(3,1)`$ $``$ $`sl(2,)`$ $`so(5,1)`$ $``$ $`sl(2,).`$ They conjectured that $`so(9,1)sl(2,𝕆)`$, the correct relation turned out to be $$so(9,1)sl(2,𝕆)G_2$$ as has been shown by Chung and Sudbery , i.e. the dimension of $`Sl(2,𝕆)`$ is 31. Also in , a quaternionic treatment of the $`d=6`$ case is presented. Later, Evans made a systematic investigation of the relationship between SSYM and ring division algebra in a couple of papers. In the first , he simplified the construction of SSYM by proving a very important identity between gamma matrices by using the intrinsic triality of ring division algebra instead of the “tour de force” used originally by Brink, Scherck and Schwarz via Fierz identities generalized to $`d>4`$ dimensions. Then, in the second paper , Evans made the connection even clearer by showing how the auxiliary fields are really related to ring division algebras. For $`d=3,4,6,10`$ we need $`k=0,1,3,7`$ auxiliary fields respectively. An alternative approach for the octonionic case was introduced by Berkovits who invented a larger supersymmetric transformation called generalized supersymmetry in . There has also been a twistor attempt by Bengtsson and Cederwall . For more references about the octonionic case and ten dimensional physics one may consult references in and its extension to p-branes by Belecowe and Duff . The early work of Nilsson may be relevant too. As a first step towards an extended superspace, we address the point of the algebraic auxiliary fields for simple N=1 supersymmetric Yang-Mills (SSYM) definable only in $`d=3,4,6`$ and $`10`$ dimensions . The important point is: *While the physical fields couple to ring division left action the auxiliary ones couple to right action (or vice versa)*. To admit a closed off-shell supersymmetric algebra, left and right action must commute i.e. we should have a parallelizable associative algebra. For d = 6, quaternions work fine but for d = 10, the only associative seven dimensional algebra that is known is the soft seven sphere. We shall show below how this works. In this chapter, we use the same symbols $`\left(\text{left action}𝔼_j\text{, right action}1|𝔼_j\right)`$ for either complex, quaternionic or octonionic numbers and each case should be distinguished by the range of the indices $`j`$ which run from $`1`$ to $`(1,3,7)`$ for complex, quaternions and octonions respectively. ### 5.1 On–Shell SSYM in $`d=3,4,6,10`$ In this section we follow a notation which is a mixture between that of Evans and Supersolutions . The Minkowskian metric has signature $`\eta _{MN}(,+,\mathrm{},+)`$ and spinorial indices have range $`n`$. Our generalized gamma matrices $`\mathrm{\Gamma }`$ and $`\stackrel{~}{\mathrm{\Gamma }}`$ are *real symmetric* and we will never raise or lower the spinors indices. We define $`\left(\mathrm{\Gamma }_M\right)_{ab}`$ of lower spinorial indices, whereas the upper ones are defined by $`\left(\stackrel{~}{\mathrm{\Gamma }}_0\right)^{ab}=\left(\mathrm{\Gamma }_0\right)_{ab}`$ and $`\left(\stackrel{~}{\mathrm{\Gamma }}\right)^{ab}=\left(\mathrm{\Gamma }\right)_{ab},`$ whence $$\mathrm{\Gamma }^M\stackrel{~}{\mathrm{\Gamma }}^N+\mathrm{\Gamma }^N\stackrel{~}{\mathrm{\Gamma }}^M=2\eta ^{MN}$$ or in terms of components $$\left(\mathrm{\Gamma }^M\right)_{ab}\left(\stackrel{~}{\mathrm{\Gamma }}^N\right)^{bc}+\left(\mathrm{\Gamma }^N\right)_{ab}\left(\stackrel{~}{\mathrm{\Gamma }}^M\right)^{bc}=2\eta ^{MN}\delta _a^c.$$ (5.1) with $`a,b`$ of range $`n`$. Simple supersymmetric Yang-Mills models are composed of gauge fields $`A_M`$, spinor fields $`\mathrm{\Psi }^a`$ and the Lagrangian density is $$=\frac{1}{4}F_{MN}F^{MN}+\frac{1}{2}\mathrm{\Psi }^t\mathrm{\Gamma }^M_M\mathrm{\Psi },$$ (5.2) where $`_M_M+A_M;`$ $`F_{MN}[_M,_N]`$ and we have suppressed here the Lie algebra indices. We first ask in which dimensions $`d`$ and for what type of spinorial field $`\mathrm{\Psi }`$ the action (5.2) is supersymmetric? Assume the spinorial field has $`n`$ components. Upon quantization, the gauge field has $`d2`$ *physical degrees of freedom* while the spinor field has $`n/2`$ physical degrees of freedom. A supersymmetric model must have equal number of bosonic and fermionic degrees of freedom, hence $`d=2+\frac{n}{2}`$ and since for spinors $`n=2,4,8,16`$ this lead to $`d=3,4,6,10.`$ We work with *real bases for spinors and vectors* . Introducing an n components Grassmann variable $`\xi `$, we postulate the supersymmetry transformation $`\delta _\xi `$ to be $`\delta _\xi A_M`$ $`=`$ $`\xi ^a\left(\mathrm{\Gamma }_M\right)_{ab}\mathrm{\Psi }^b,`$ $`\delta _\xi \mathrm{\Psi }^a`$ $`=`$ $`{\displaystyle \frac{1}{2}}\xi ^b\left(\stackrel{~}{\mathrm{\Gamma }}^M\right)^{ac}\left(\mathrm{\Gamma }^N\right)_{cb}F_{MN}.`$ We have to check the invariance of the Lagrangian. The variation of $`F_{MN}`$ is $$\delta _\xi F_{MN}=\xi ^a\left(\mathrm{\Gamma }_N\right)_{ab}_M\mathrm{\Psi }^b\xi ^a\left(\mathrm{\Gamma }_M\right)_{ab}_N\mathrm{\Psi }^b,$$ (5.4) the variation of the first term of our $``$ gives $$\delta _\xi \left(\frac{1}{4}F_{MN}F^{MN}\right)=\xi ^a\left(\mathrm{\Gamma }_M\right)_{ab}F^{MN}_N\mathrm{\Psi }^b.$$ (5.5) Now, taking into account $`_M=_M+A_M`$, making the variation of the second term of the Lagrangian (5.2) and using as a basis of our Lie algebra $`\mathrm{\Psi }^zt^z`$ with $`[t^x,t^y]=c^{xyz}t^z,`$ $`\delta _\xi \left({\displaystyle \frac{1}{2}}\left(\mathrm{\Gamma }^M\right)_{ab}\mathrm{\Psi }^a_M\mathrm{\Psi }^b\right)`$ $`=`$ $`\xi ^a\left(\mathrm{\Gamma }^N\right)_{ab}F_{MN}^M\mathrm{\Psi }^b`$ (5.6) $`{\displaystyle \frac{1}{2}}\xi ^c\left(\mathrm{\Gamma }_M\right)_{ab}\left(\mathrm{\Gamma }^M\right)_{cd}\left(\mathrm{\Psi }^{az}c_{xyz}\mathrm{\Psi }^{dx}\mathrm{\Psi }^{by}\right)`$ $`+totalderivative.`$ We have used the Bianchi identity in the above derivation. The SSYM action (5.2) is invariant under (LABEL:trs) if $$\xi ^c\left(\mathrm{\Gamma }_M\right)_{ab}\left(\mathrm{\Gamma }^M\right)_{cd}\left(\mathrm{\Psi }^{az}c_{xyz}\mathrm{\Psi }^{dx}\mathrm{\Psi }^{by}\right)=0$$ (5.7) $$Q_{abcd}=\left(\mathrm{\Gamma }_M\right)_{ab}\left(\mathrm{\Gamma }^M\right)_{cd}+\left(\mathrm{\Gamma }_M\right)_{bd}\left(\mathrm{\Gamma }^M\right)_{ca}+\left(\mathrm{\Gamma }_M\right)_{da}\left(\mathrm{\Gamma }^M\right)_{cb}=0.$$ (5.8) We conclude from this calculation that $``$ is invariant under (LABEL:trs) in any dimension for abelian algebras with any spin representation because (5.7) is trivially zero for $`c_{xyz}=0`$. Now, we want to find the solution of (5.8). As our complex numbers, quaternions, octonions form a Clifford algebra of signature $`Cliff(0,1),Cliff(0,3),Cliff(0,7)`$ respectively, we have $`(𝔼_k)_{\mu \nu }(𝔼_j)_{\lambda \nu }+(𝔼_j)_{\mu \nu }(𝔼_k)_{\lambda \nu }`$ $`=`$ $`2\delta _{kj}\delta _{\mu \lambda },`$ $`(𝔼_k)_{\mu \nu }(𝔼_j)_{\mu \lambda }+(𝔼_j)_{\mu \nu }(𝔼_k)_{\mu \lambda }`$ $`=`$ $`2\delta _{kj}\delta _{\nu \lambda },`$ $`(𝔼_k)_{\mu \nu }(𝔼_k)_{\lambda \zeta }+(𝔼_k)_{\lambda \nu }(𝔼_k)_{\mu \zeta }`$ $`=`$ $`2\delta _{\mu \lambda }\delta _{\nu \zeta },`$ (5.9) and the same holds equally well for $`(1|𝔼_j)`$. As had been noticed by Evans, these are direct consequences of the ring division triality. We can construct immediately two sets of gamma matrices as follows $$\left(\mathrm{\Gamma }_0\right)=\left(\begin{array}{cc}\mathrm{𝟏}_{\frac{n}{2}}& 0\\ 0& \mathrm{𝟏}_{\frac{n}{2}}\end{array}\right);$$ $$\left(\mathrm{\Gamma }_j\right)=\left(\begin{array}{cc}0& 𝔼_j\\ 𝔼_j& 0\end{array}\right);\left(1|\mathrm{\Gamma }_j\right)=\left(\begin{array}{cc}0& 1|𝔼_j\\ 1|𝔼_j& 0\end{array}\right),$$ $`\left(\mathrm{\Gamma }_{d2}\right)`$ $`=`$ $`\left(\begin{array}{cc}0& \mathrm{𝟏}_{\frac{n}{2}}\\ \mathrm{𝟏}_{\frac{n}{2}}& 0\end{array}\right);\left(\mathrm{\Gamma }_{d1}\right)=\left(\begin{array}{cc}\mathrm{𝟏}_{\frac{n}{2}}& 0\\ 0& \mathrm{𝟏}_{\frac{n}{2}}\end{array}\right),`$ (5.14) where $`j=1..d3,`$ thus $$\mathrm{\Gamma }^M\stackrel{~}{\mathrm{\Gamma }}^N+\mathrm{\Gamma }^N\stackrel{~}{\mathrm{\Gamma }}^M=1|\mathrm{\Gamma }^M1|\stackrel{~}{\mathrm{\Gamma }}^N+1|\mathrm{\Gamma }^N1|\stackrel{~}{\mathrm{\Gamma }}^M=2\eta ^{MN}$$ or in terms of components $`\left(\mathrm{\Gamma }^M\right)_{ab}\left(\stackrel{~}{\mathrm{\Gamma }}^N\right)^{bc}+\left(\mathrm{\Gamma }^N\right)_{ab}\left(\stackrel{~}{\mathrm{\Gamma }}^M\right)^{bc}`$ $`=`$ $`\left(1|\mathrm{\Gamma }^M\right)_{ab}\left(1|\stackrel{~}{\mathrm{\Gamma }}^N\right)^{bc}+\left(1|\mathrm{\Gamma }^N\right)_{ab}\left(1|\stackrel{~}{\mathrm{\Gamma }}^M\right)^{bc}`$ (5.16) $`=`$ $`2\eta ^{MN}\delta _a^c.`$ automatically, our $`\mathrm{\Gamma }`$’s satisfy the identity (5.8) for both left and right actions. Indeed by direct calculation using (5.9), we see that $$\mathrm{\Gamma }_{Ma(b}\mathrm{\Gamma }_{cd)}^M=1|\mathrm{\Gamma }_{Ma(b}1|\mathrm{\Gamma }_{cd)}^M=0.$$ (5.17) Consequently, the spin representation decomposes into $`SPIN(1,2)`$ $``$ $`SL(2,R)`$ (5.18) $`SPIN(1,3)`$ $``$ $`SL(2,C)`$ (5.19) $`SPIN(1,5)`$ $``$ $`SL(2,H)`$ (5.20) and using the soft sphere, it seems that $`softSPIN(1,9)SL(2,softS^7)`$. We still have to show that the commutators of (LABEL:trs) close to a supersymmetric algebra. For any arbitrary field $`V`$ in our Lagrangian, we need to check that $`[\delta _\xi ,\delta _\chi ]V=2\xi ^a\chi ^b\left(\mathrm{\Gamma }^M\right)_{ab}_MV`$ . (5.21) As we are working on–shell, the algebra should closes modulo the fermionic equation of motion $$\left(\mathrm{\Gamma }^M\right)_{ab}_M\mathrm{\Psi }^a=0,b$$ (5.22) and gauge transformation. Using (LABEL:trs), we can easily check the closure for the gauge field $`A_M`$ $`[\delta _\xi ,\delta _\chi ]A_M`$ $`=`$ $`{\displaystyle \frac{1}{2}}\xi ^a\chi ^b\left(\left(\stackrel{~}{\mathrm{\Gamma }}^P\right)^{cd}\left(\mathrm{\Gamma }_M\right)_{bc}\left(\mathrm{\Gamma }^N\right)_{ad}+\left(\stackrel{~}{\mathrm{\Gamma }}^P\right)^{cd}\left(\mathrm{\Gamma }_M\right)_{ac}\left(\mathrm{\Gamma }^N\right)_{bd}\right)F_{PN}`$ $`=`$ $`2\xi ^a\chi ^b\left(\mathrm{\Gamma }^N\right)_{ab}F_{NM}`$ where we have used $`F_{PN}=F_{NP}`$. To check the close for the fermionic field is a little bit lengthy, but straightforward $`[\delta _\xi ,\delta _\chi ]\mathrm{\Psi }^c`$ $`=`$ $`\underset{¯}{\xi ^a\chi ^b\left(\stackrel{~}{\mathrm{\Gamma }}^M\right)^{cd}\left(\mathrm{\Gamma }^N\right)_{de}\left(\mathrm{\Gamma }_N\right)_{ab}_M\mathrm{\Psi }^e}\xi ^a\chi ^b\left(\stackrel{~}{\mathrm{\Gamma }}^M\right)^{cd}Q_{abde}_M\mathrm{\Psi }^e`$ $`=`$ $`2\xi ^a\chi ^b\left(\mathrm{\Gamma }^M\right)_{ab}_M\mathrm{\Psi }^c\xi ^a\chi ^b\left(\stackrel{~}{\mathrm{\Gamma }}^M\right)^{cd}Q_{abde}_M\mathrm{\Psi }^e,`$ where we have used the fermionic equation of motion to simplify the underlined term. Thus the supersymmetry closes iff $`Q_{abde}=0.`$ This is true both for the abelian and nonabelian cases. In summary to *close the algebra* *and* *to have an invariant Lagrangian* $`Q_{abde}`$ *must vanishes* for both the abelian and the non-abelian case. ### 5.2 Representation of the Supersymmetry Algebra For a theory to be supersymmetric, it is necessary that its particle content form a representation of the supersymmetry algebra. Using the gamma matrices representation given in the previous section, we show how to describe the representations of our supersymmetry algebra in $`d=3,4,6,10`$. From (5.21), we deduce our supersymmetry algebra as $$\overline{)\{Q_a,Q_b\}=2\left(\mathrm{\Gamma }\right)_{ab}_\mu 2\left(\mathrm{\Gamma }\right)_{ab}P_\mu }$$ (5.25) where $`Q_a`$ are the supersymmetry generators and transform as spin-half operators under the angular momentum algebra. Moreover, the supersymmetry generators commute with the momentum operator $`P_\mu `$ and hence, with $`P^2`$. Therefore, all states in a given representation of the algebra have the same mass. For our case we will be concerned with the massless representation only. For massless states, we can always go to a frame where $`P^\mu =M(1,..,1)`$. Then the supersymmetry algebra becomes $$\overline{)\{Q_a,Q_b\}=\left(\begin{array}{cc}0\hfill & \hfill 0\\ 0\hfill & \hfill 4M\end{array}\right)=2M\left(\mathrm{\Gamma }_+\right)_{ab}}.$$ where $$\left(\mathrm{\Gamma }_+\right)=\left(\mathrm{\Gamma }_0\right)+\left(\mathrm{\Gamma }_{d1}\right)=2\left(\begin{array}{cc}0\hfill & 0\hfill \\ 0\hfill & \mathrm{𝟏}_{\frac{n}{2}}\hfill \end{array}\right).$$ (5.26) It is convenient to rescale our generators as $$a_\mu =\frac{1}{\sqrt{2M}}Q_\mu ,$$ for $`\mu =0..(\frac{n}{2}1)`$, where $`(\frac{n}{2}1)=0,1,3,7`$ for $`d=3,4,6,10`$ respectively. Then, the supersymmetry algebra takes the form $$\{a_\mu ,a_\nu \}=\delta _{\mu \nu },$$ This is a Clifford algebra with $`\frac{n}{2}`$ generators. We can now proceed in two different ways: * To retrieve the standard complex representation of our supersymmetry algebra, we have to pair our generators $$\begin{array}{cccc}d=3\hfill & a_0\hfill & & \hfill \\ d=4\hfill & b=a_0+ia_1\hfill & b^{}=a_0ia_1\hfill & \hfill \\ d=6\hfill & \{\begin{array}{c}b_1=a_0+ia_1\hfill \\ b_1^{}=a_0ia_1\hfill \end{array}\hfill & \begin{array}{c}b_2=a_3+ia_4\hfill \\ b_2^{}=a_3ia_4\hfill \end{array}\hfill & \hfill \\ d=10\hfill & \{\begin{array}{c}\begin{array}{c}b_1=a_0+ia_1\hfill \\ b_1^{}=a_0ia_1\hfill \end{array}\\ \begin{array}{c}b_2=a_3+ia_4\hfill \\ b_2^{}=a_3ia_4\hfill \end{array}\end{array}\hfill & \begin{array}{c}\begin{array}{c}b_3=a_5+ia_6\hfill \\ b_3^{}=a_5ia_6\hfill \end{array}\hfill \\ \begin{array}{c}b_4=a_7+ia_8\hfill \\ b_4^{}=a_7ia_8\hfill \end{array}\hfill \end{array}\hfill & \hfill \end{array}$$ (5.27) leading to $`caseI`$ $$\begin{array}{cccccccc}d=3\hfill & a_0\hfill & & & & & & \hfill \\ d=4\hfill & \{b,b\}\hfill & =\hfill & \{b^{},b^{}\}\hfill & =0,\hfill & \{b,b^{}\}\hfill & =2\hfill & \hfill \\ d=6\hfill & \{b_i,b_j\}\hfill & =\hfill & \{b_i^{},b_j^{}\}\hfill & =0,\hfill & \{b_i,b_j^{}\}\hfill & =2\delta _{ij}\hfill & \begin{array}{c}\\ i,j=1,2\hfill \end{array}\hfill \\ d=10\hfill & \{b_i,b_j\}\hfill & =\hfill & \{b_i^{},b_j^{}\}\hfill & =0,\hfill & \{b_i,b_j^{}\}\hfill & =2\delta _{ij}\hfill & \begin{array}{c}\\ i,j=\mathrm{1..4}\hfill \end{array}\hfill \end{array}$$ (5.28) * We can work with hypercomplex numbers, then we have $`caseII`$ $$\begin{array}{cccc}d=3\hfill & a_0\hfill & \hfill & \\ d=4\hfill & \{\begin{array}{c}b=a_0+e_1a_1\hfill \\ b^{}=a_0e_1a_1\hfill \end{array}\hfill & \begin{array}{c}\\ e_1istheimaginarycomplexunit\hfill \end{array}\hfill & \\ d=6\hfill & \{\begin{array}{c}b_1=a_0+e_ia_i\hfill \\ b_1^{}=a_0e_ia_i\hfill \end{array}\hfill & \begin{array}{c}\\ e_iareimaginaryquaternionunits\hfill \\ i=1\mathrm{}3\hfill \end{array}\hfill & \\ d=10\hfill & \{\begin{array}{c}b_1=a_0+e_ia_i\hfill \\ b_1^{}=a_0e_ia_i\hfill \end{array}\hfill & \begin{array}{c}\\ e_iareimaginaryoctonionicunits\hfill \\ i=1\mathrm{}7\hfill \end{array}\hfill & \end{array}$$ (5.29) ### 5.3 The SSYM Auxiliary Fields Problem One may ask oneself: Why are auxiliary fields important? There are many convincing reasons. Let us mention jut five of them. 1- Only in the presence of auxiliary fields is the supersymmetry manifest. Indeed, when we use superspace, we can write our supersymmetric models in a form clearly invariant under Lorentz transformation as well as supersymmetry. The clearest example is the superspace formulation of $`N=1`$ supersymmetric theories. 2- As we saw in the first section of this chapter, the closure of the supersymmetric algebra is achieved only by using the field equations of motion. At the quantum level, the equations of motion get corrected and consequently the supersymmetric algebra will be realized in a highly non-trivial fashion, Look to . 3- The use of the Lagrangian formulation of field theory is usually advocated on the basis of symmetries arguments. Hence making the symmetry manifest is a priority. 4- Feynman diagrams with superfields explains naturally many of the “miracle” cancellations in supersymmetric models. 5- Supersymmetry is only constructed systematically when we use superspace. In principle, the superspace formulation should provide us with all the details, the supersymmetry transformations, the full interaction Lagrangian, even the constraints must be derived in agreement with the super Bianchi identities. Unfortunately, we don’t have a complete superspace treatment in $`d>4`$. The number of auxiliary fields can be counted easily. For SSYM, we have the following off–shell degrees of freedom (ofdf) $$\begin{array}{ccccccc}d\backslash ofdf\hfill & & \mathrm{\Psi }\hfill & & A_\mu \hfill & & \\ 3\hfill & & 2\hfill & \hfill & 2\hfill & =\hfill & 0\hfill \\ 4\hfill & & 4\hfill & \hfill & 3\hfill & =\hfill & 1\hfill \\ 6\hfill & & 8\hfill & \hfill & 5\hfill & =\hfill & 3\hfill \\ 10\hfill & & 16\hfill & \hfill & 9\hfill & =\hfill & 7\hfill \end{array}$$ (5.30) The $`d=3`$ case is trivial in the sense that it contains no auxiliary fields. For $`d=4`$, a superspace formalism based on $`SL(2,)`$ is needed to formulate our supersymmetric YM model in a manifestly invariant way. Such a superspace treatment provides us automatically with the needed single auxiliary field. In $`d=6,10`$, it is conjectured that an $`SL(2,)`$ and $`SL(2,𝕆)`$ are needed. Here, we try to support this conjecture by a different argument and we hope that the tools presented may help in the future to find the full superspace formulation. Using Evans ansatz , SSYM are composed of: Gauge fields $`A_M`$, spinors $`\mathrm{\Psi }^a`$, $`j(=1..d3)`$ algebraic auxiliary fields $`K^j`$. The gauge group indices will be suppressed in the following. The Lagrangian density is $$=\frac{1}{4}F_{MN}F^{MN}+\frac{i}{2}\mathrm{\Psi }^t\mathrm{\Gamma }^M_M\mathrm{\Psi }+\frac{1}{2}K^2,$$ (5.31) where $`_M_M+A_M;`$ $`F_{MN}[_M,_N]`$ and the $`\mathrm{\Gamma }`$ are given in (LABEL:gmm). The Lagrangian is invariant up to a total derivative iff (5.8) holds. Our supersymmetry transformations are<sup>2</sup><sup>2</sup>2Contrary to , we set $`\mathrm{\Lambda }_P=\stackrel{~}{\mathrm{\Lambda }}^P`$ from the strart. $`\delta _\eta A_M`$ $`=`$ $`i\eta \mathrm{\Gamma }_M\mathrm{\Psi },`$ $`\delta _\eta \mathrm{\Psi }^\alpha `$ $`=`$ $`{\displaystyle \frac{1}{2}}F_{MN}\left(\mathrm{\Gamma }_{MN}\eta \right)^\alpha +K^j\left(\mathrm{\Lambda }_j\right)_\beta ^\alpha \eta ^\beta ,`$ $`\delta _\eta K_j`$ $`=`$ $`i\left(\mathrm{\Gamma }^M_M\mathrm{\Psi }\right)_\alpha \left(\mathrm{\Lambda }_j\right)_\beta ^\alpha \eta ^\beta ,`$ where $`\mathrm{\Lambda }_j`$ are some real antisymmetric matrices $`\left(\mathrm{\Lambda }_j\right)^t=\left(\mathrm{\Lambda }_j\right)`$<sup>3</sup><sup>3</sup>3We will mention shortly how to relax this condition. and Lorentz transformations are generated by $`\mathrm{\Gamma }_{MN}\stackrel{~}{\mathrm{\Gamma }}_{[M}\mathrm{\Gamma }_{N]}`$. Imposing the closure of the supersymmetry infinitesimal transformations $$[\delta _ϵ,\delta _\eta ]=2iϵ^t\mathrm{\Gamma }^M\eta _M.$$ (5.33) The closure on $`A_M`$ yields $$\mathrm{\Gamma }_M\mathrm{\Lambda }_j\mathrm{\Lambda }_j\mathrm{\Gamma }_M=0.$$ (5.34) In addition to this condition the closure on $`K^j`$ also requires $$\mathrm{\Lambda }_j\mathrm{\Lambda }_h+\mathrm{\Lambda }_h\mathrm{\Lambda }_j=2\delta _{jh}.$$ (5.35) While closure on the fermionic field $`\mathrm{\Psi }^\alpha `$ holds iff $$\left(\mathrm{\Gamma }^M\right)_{\alpha \beta }\left(\stackrel{~}{\mathrm{\Gamma }}_M\right)^{\gamma \delta }=2\delta _{(\alpha }^\gamma \delta _{\beta )}^\delta +2\left(\mathrm{\Lambda }_j\right)_{(\alpha }^\gamma \left(\mathrm{\Lambda }_j\right)_{\beta )}^\delta .$$ Now, we continue in a different way to Evans. To construct $`\mathrm{\Lambda }_j`$, we first notice from (5.35) that the $`\mathrm{\Lambda }_j`$ form a real Clifford algebra, and from (5.34) that they commute with our space-time $`\mathrm{\Gamma }_M`$ Clifford algebra. The solution of the auxiliary field problem for $`d=3,4,6`$ dimensions is then simply $$\mathrm{\Lambda }_j=\left(\begin{array}{cc}1|𝔼_j& 0\\ 0& 1|𝔼_j\end{array}\right),$$ (5.36) because $$\left\{1\right|𝔼_j,1|𝔼_h\}=2\delta _{jh},$$ (5.37) and $$[𝔼_j,1|𝔼_h]=0.$$ (5.38) Of course this solution is not unique. For example, if someone had started with $`1|\mathrm{\Gamma }_M`$, he would have found $`\mathrm{\Lambda }_j=\left(\begin{array}{cc}𝔼_j\hfill & 0\hfill \\ 0\hfill & 𝔼_j\hfill \end{array}\right)`$. In general, we can relax the conditions of antisymmetricity of $`\mathrm{\Lambda }`$ and the symmetricity of $`\mathrm{\Gamma }.`$ One writes any $`\mathrm{\Gamma }`$ and expand it in terms left/right action $`(𝔼_{i,}1|𝔼_j,𝔼_m|𝔼_n)`$ then the $`\mathrm{\Lambda }`$ will be given in terms of $`(1|𝔼_{i,}𝔼_j,𝔼_n|𝔼_m)`$. For $`d=10`$, working with octonions the situation is different. From chapter 3, we know that octonionic left and right action commutes only when applied to $`\phi `$, $$\phi ^t[𝔼_j,1|𝔼_h]\phi =0,$$ (5.39) and $`\phi `$ is just an 8 dimensional column matrix. Up to now, we have not restricted $`\phi `$ by any other conditions. With two different $`\phi `$, $`(\phi ^{\left(1\right)},\phi ^{\left(2\right)})`$, we impose now the conditions that $`\phi ^{\left(i\right)}`$ be fermionic fields. We express our 16 dimensional Grassmanian variables $`ϵ,\eta `$ of eqn.(5.33) in terms of $`\phi `$, $`ϵ`$ $`=`$ $`\eta ^t`$ $``$ $`ϵ`$ $`=`$ $`\left(\begin{array}{cc}\phi ^{\left(1\right)}\hfill & \phi ^{\left(2\right)}\hfill \end{array}\right);\eta =\left(\begin{array}{c}\phi ^{\left(1\right)}\hfill \\ \phi ^{\left(2\right)}\hfill \end{array}\right)`$ (5.43) We now rederive (5.33) for the octonions. The closure conditions of our algebra, *without omitting the Grassmanian variables* are $`\eta ^t\left(\mathrm{\Gamma }_M\mathrm{\Lambda }_j\mathrm{\Lambda }_j\mathrm{\Gamma }_M\right)\eta `$ $`=`$ $`0,`$ $`\eta ^t\left(\mathrm{\Lambda }_j\mathrm{\Lambda }_h+\mathrm{\Lambda }_h\mathrm{\Lambda }_j\right)\eta `$ $`=`$ $`\eta ^t\left(2\delta _{jh}\right)\eta ,`$ $`\eta ^t\left(\left(\mathrm{\Gamma }^M\right)_{\alpha \beta }\left(\stackrel{~}{\mathrm{\Gamma }}_M\right)^{\gamma \delta }\right)\eta `$ $`=`$ $`\eta ^t\left(2\delta _{(\alpha }^\gamma \delta _{\beta )}^\delta +2\left(\mathrm{\Lambda }_j\right)_{(\alpha }^\gamma \left(\mathrm{\Lambda }_j\right)_{\beta )}^\delta \right)\eta ,`$ (5.44) which are satisfied for the octonionic representation $$\left(\mathrm{\Gamma }_j\right)_{ab}=\left(\begin{array}{cc}0& 𝔼_j\\ 𝔼_j& 0\end{array}\right),\mathrm{\Lambda }_j=\left(\begin{array}{cc}1|𝔼_j& 0\\ 0& 1|𝔼_j\end{array}\right).$$ (5.45) By interchanging left/right action, we have different solutions as in the quaternionic case. In summary, while the fermionic fields couple to left/right action through the gamma matrices, the auxiliary fields couple to right/left action through the $`\mathrm{\Lambda }`$. For the octonionic case *the presence of the Grassmanian variables is essential.* Contrary to the standard supersymmetry transformation, our Grassman variables are the same, which is identical to the result obtained by Berkovits in . According to Evans , the attractive feature of this scheme is that the Lagrangian (5.31) and the transformation (LABEL:dddd) are manifestly invariant under the generalized Lorentz group $`SO(1,9)`$. In our formulation, we can show some additional characteristic. In some cases, the (5.43) condition may be relaxed, for equal $`j`$ or $`h`$ (no summation) $$\begin{array}{c}\phi ^t𝔼_j[𝔼_j,1|𝔼_h]\phi \\ \phi ^t\mathrm{\hspace{0.33em}\hspace{0.33em}1}|𝔼_i[𝔼_j,1|𝔼_h]\phi \\ \phi ^tE_h[𝔼_j,1|𝔼_h]\phi \\ \phi ^t\mathrm{\hspace{0.33em}\hspace{0.33em}1}|E_h[𝔼_j,1|𝔼_h]\phi \end{array}\}=0.$$ (5.46) i.e. relating $`ϵ`$ and $`\eta `$ by an $`S^7`$ is also allowed. Now, Let us show what will happen to $`spin(1,9)`$ when we transform it to $`softspin(1,9)`$ $`softspin(1,9)`$ $``$ $`[\mathrm{\Gamma }_i,\mathrm{\Gamma }_j]\eta `$ (5.53) $`=`$ $`[\left(\begin{array}{cc}0& 𝔼_i\\ 𝔼_i& 0\end{array}\right),\left(\begin{array}{cc}0& 𝔼_j\\ 𝔼_j& 0\end{array}\right)]\left(\begin{array}{c}\phi ^{\left(1\right)}\hfill \\ \phi ^{\left(2\right)}\hfill \end{array}\right)`$ $`=`$ $`\left(\begin{array}{cc}0& [𝔼_i,𝔼_j]\\ [𝔼_i,𝔼_j]& 0\end{array}\right)\left(\begin{array}{c}\phi ^{\left(1\right)}\hfill \\ \phi ^{\left(2\right)}\hfill \end{array}\right)`$ (5.58) $`=`$ $`\left(\begin{array}{cc}0\hfill & f_{ijk}^{(+)}\left(\phi ^{\left(2\right)}\right)𝔼_k\hfill \\ f_{ijk}^{(+)}\left(\phi ^{\left(1\right)}\right)𝔼_k\hfill & 0\hfill \end{array}\right)\left(\begin{array}{c}\phi ^{\left(1\right)}\hfill \\ \phi ^{\left(2\right)}\hfill \end{array}\right).`$ (5.63) Lastly, let us make some comments about a possible superspace. It seems that the best way to find the $`d=6,10`$ superspace for SSYM is by defining some quaternionic and octonionic Grassmann variables that decompose the corresponding spinors into an $`SL(2,H)`$ and an $`SL(2,softS^7)`$ respectively $$\{\theta _\alpha ,\theta _\beta \}=\{\overline{\theta }_{\dot{\alpha }},\overline{\theta }_{\dot{\beta }}\}=\{\theta _\alpha ,\overline{\theta }_{\dot{\beta }}\}=0,$$ (5.65) where $`\alpha =1,2`$ over quaternions or octonions. We know that the supersymmetry generators $`Q_\alpha `$ are derived from right multiplication $`Q_\alpha `$ $`=`$ $`\left(_\alpha 1|\mathrm{\Gamma }_{\alpha \dot{\beta }}^\mu \overline{\theta }^{\dot{\beta }}P_\mu \right)`$ (5.66) $`Q^\alpha `$ $`=`$ $`\left(^\alpha +\overline{\theta }_{\dot{\beta }}1|\stackrel{~}{\mathrm{\Gamma }}^{\mu \dot{\beta }\alpha }P_\mu \right)`$ (5.67) also $$\overline{Q}^{\dot{\alpha }}=\left(^{\dot{\alpha }}1|\stackrel{~}{\mathrm{\Gamma }}^{\mu \dot{\alpha }\alpha }\theta _\alpha P_\mu \right)$$ (5.68) $$\overline{Q}_{\dot{\alpha }}=\left(_{\dot{\alpha }}+\theta ^\alpha 1|\mathrm{\Gamma }_{\alpha \dot{\alpha }}P_\mu \right)$$ (5.69) whereas the covariant derivative $`D_\alpha `$ are obtained by left action $`D_\alpha `$ $`=`$ $`\left(_\alpha +\mathrm{\Gamma }_{\alpha \dot{\beta }}^\mu \overline{\theta }^{\dot{\beta }}P_\mu \right)`$ (5.70) $`D^\alpha `$ $`=`$ $`\left(^\alpha \overline{\theta }_{\dot{\beta }}\stackrel{~}{\mathrm{\Gamma }}^{\mu \dot{\beta }\alpha }P_\mu \right)`$ (5.71) also $$\overline{D}^{\dot{\alpha }}=\left(^{\dot{\alpha }}+\stackrel{~}{\mathrm{\Gamma }}^{\mu \dot{\alpha }\alpha }\theta _\alpha P_\mu \right)$$ (5.72) $$\overline{D}_{\dot{\alpha }}=\left(_{\dot{\alpha }}\theta ^\alpha \mathrm{\Gamma }_{\alpha \dot{\alpha }}P_\mu \right)$$ (5.73) Leading to a result acceptable but different from the standard $`N=1`$, $`d=4`$ superspace, $`\{Q_\alpha ,\overline{Q}_{\dot{\alpha }}\}`$ $`=`$ $`2\left(1|\mathrm{\Gamma }_{\alpha \dot{\alpha }}^\mu \right)P_\mu ,`$ $`\{Q_\alpha ,Q_\beta \}`$ $`=`$ $`\{\overline{Q}_{\dot{\alpha }},\overline{Q}_{\dot{\beta }}\}=\mathrm{\hspace{0.33em}0},`$ $`\{D_\alpha ,\overline{D}_{\dot{\alpha }}\}`$ $`=`$ $`2\mathrm{\Gamma }_{\alpha \dot{\alpha }}^\mu P_\mu ,`$ $`\{D_\alpha ,D_\beta \}`$ $`=`$ $`\{\overline{D}_{\dot{\alpha }},\overline{D}_{\dot{\beta }}\}=\mathrm{\hspace{0.33em}0},`$ and iff left and right action commute, we restore $`\{Q_\alpha ,\overline{D}_{\dot{\alpha }}\}`$ $`=`$ $`\{D_\alpha ,\overline{Q}_{\dot{\alpha }}\}=0,`$ $`\{Q_\alpha ,D_\beta \}`$ $`=`$ $`\{\overline{D}_{\dot{\alpha }},\overline{Q}_{\dot{\beta }}\}=\mathrm{\hspace{0.33em}0}.`$ On the other hand for octonions we would have the weaker conditions, $`\left(\begin{array}{cc}\phi ^{\left(1\right)}\hfill & \phi ^{\left(2\right)}\hfill \end{array}\right)\{Q_\alpha ,\overline{D}_{\dot{\alpha }}\}\left(\begin{array}{c}\phi ^{\left(1\right)}\hfill \\ \phi ^{\left(2\right)}\hfill \end{array}\right)`$ $`=`$ $`\left(\begin{array}{cc}\phi ^{\left(1\right)}\hfill & \phi ^{\left(2\right)}\hfill \end{array}\right)\{D_\alpha ,\overline{Q}_{\dot{\alpha }}\}\left(\begin{array}{c}\phi ^{\left(1\right)}\hfill \\ \phi ^{\left(2\right)}\hfill \end{array}\right)=0,`$ (5.80) $`\left(\begin{array}{cc}\phi ^{\left(1\right)}\hfill & \phi ^{\left(2\right)}\hfill \end{array}\right)\{Q_\alpha ,D_\beta \}\left(\begin{array}{c}\phi ^{\left(1\right)}\hfill \\ \phi ^{\left(2\right)}\hfill \end{array}\right)`$ $`=`$ $`\left(\begin{array}{cc}\phi ^{\left(1\right)}\hfill & \phi ^{\left(2\right)}\hfill \end{array}\right)\{\overline{D}_{\dot{\alpha }},\overline{Q}_{\dot{\beta }}\}\left(\begin{array}{c}\phi ^{\left(1\right)}\hfill \\ \phi ^{\left(2\right)}\hfill \end{array}\right)=\mathrm{\hspace{0.33em}0}.`$ (5.87) The commutation of left and right actions is not just needed for associativity but for the invariance under supersymmetry transformation $$\delta _\xi \xi Q+\overline{\xi }\overline{Q}$$ (5.88) because only the associativity ensures $$\left(\begin{array}{cc}\phi ^{\left(1\right)}\hfill & \phi ^{\left(2\right)}\hfill \end{array}\right)[\delta _\xi ,D_\alpha ]\left(\begin{array}{c}\phi _1\hfill \\ \phi _2\hfill \end{array}\right)=\left(\begin{array}{cc}\phi ^{\left(1\right)}\hfill & \phi ^{\left(2\right)}\hfill \end{array}\right)[\delta _\xi ,\overline{D}_{\dot{\alpha }}]\left(\begin{array}{c}\phi _1\hfill \\ \phi _2\hfill \end{array}\right)=0,$$ (5.89) since $`\delta _\xi `$ is left action and $`D_\alpha `$ is right action which is a very important relation in the standard $`N=1`$ superspace for the invariance of the Lagrangian under supersymmetry transformation. We hope to return to this point in a future work. ## Chapter 6 Conclusions In this thesis, we have presented a systematic study of the hidden faithful Clifford algebraic structure in the different types of ring division algebras. This relationship had been elaborated by going beyond octonions to hexagonions. We have then dedicated a complete chapter to octonions. They are not as useless as often believed.. They may be safely employed once the non-associativity has been bypassed. The necessary ingredients are: * *Fixing the direction of action by introducing the* $`\delta `$ *operator.* * *Closing the* $`\delta `$ *algebra by using structure functions* $`f_{ijk}^{(+)}\left(\phi \right)`$*.* * *Matrix representation of the* $`\delta `$ *algebra. The* $`𝔼`$ *or* $`𝔼\left(\phi \right)`$ *can be found and their structure functions can be computed easily.* During this analysis, we have introduced and discussed the soft seven sphere. There maybe different applications of the soft seven sphere in physics . We have given two such cases where the ring division algebras occupies a special position. Self-duality and SSYM are two promising places where the soft seven sphere proves to be useful and indeed essential. In our formulation, *we find a new eight dimensional feature, that had never appeared before, the existence of an infinite family of dualities. By moving over the gauged seven sphere, we define new conditions and we have new solutions. We have parameterized all these conditions and solutions in terms of the coordinate system over the gauged seven sphere.* By defining the soft–duality condition, we have tried to retain as much as possible of the four dimensional case. We then showed how new solutions of the GKS condition can be easily found . *For SSYM, the new and old off–shell formulations can be rederived in a systematic and uniform fashion.* We believe that the interplay between left and right ring division operators is not a coincidence but an intrinsic property of supersymmetry that needs further study. By interchanging left and right action, we have many different solutions. Again the octonionic ten dimensional case is very special. It will be interesting to apply the ideas presented here into the GS context. We hope that this work constitute a first step in the correct direction for further application of the soft seven sphere algebra. ## Appendix A The First Appendix In this appendix we give the translation rules between octonionic left-right barred operators and $`8\times 8`$ real matrices. In order to simplify our presentation we introduce the following notation: $`\{a,b,c,d\}_{(1)}`$ $``$ $`\left(\begin{array}{cccc}a& 0& 0& 0\\ 0& b& 0& 0\\ 0& 0& c& 0\\ 0& 0& 0& d\end{array}\right),\{a,b,c,d\}_{(2)}\left(\begin{array}{cccc}0& a& 0& 0\\ b& 0& 0& 0\\ 0& 0& 0& c\\ 0& 0& d& 0\end{array}\right),`$ (A.9) $`\{a,b,c,d\}_{(3)}`$ $``$ $`\left(\begin{array}{cccc}0& 0& a& 0\\ 0& 0& 0& b\\ c& 0& 0& 0\\ 0& d& 0& 0\end{array}\right),\{a,b,c,d\}_{(4)}\left(\begin{array}{cccc}0& 0& 0& a\\ 0& 0& b& 0\\ 0& c& 0& 0\\ d& 0& 0& 0\end{array}\right),`$ (A.19) where $`a,b,c,d`$ and $`0`$ represent $`2\times 2`$ real matrices. As elsewhere by $`\sigma _1`$, $`\sigma _2`$, $`\sigma _3`$ we mean the standard Pauli matrices: $$\sigma _1=\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right),\sigma _2=\left(\begin{array}{cc}0& i\\ i& 0\end{array}\right),\sigma _3=\left(\begin{array}{cc}1& 0\\ 0& 1\end{array}\right).$$ (A.21) The only necessary translation rules that we need to know explicitly are the following | $`e_1`$ | $`\{`$ | $`i\sigma _2`$, | $`i\sigma _2`$, | $`i\sigma _2`$, | $`i\sigma _2\}_{(1)}`$ | | --- | --- | --- | --- | --- | --- | | $`e_2`$ | $`\{`$ | $`\sigma _3`$, | $`\sigma _3`$, | $`\mathrm{𝟏}`$, | $`\mathrm{𝟏}\}_{(2)}`$ | | $`e_3`$ | $`\{`$ | $`\sigma _1`$, | $`\sigma _1`$, | $`i\sigma _2`$, | $`i\sigma _2\}_{(2)}`$ | | $`e_4`$ | $`\{`$ | $`\sigma _3`$, | $`\mathrm{𝟏}`$, | $`\sigma _3`$, | $`\mathrm{𝟏}\}_{(3)}`$ | | $`e_5`$ | $`\{`$ | $`\sigma _1`$, | $`i\sigma _2`$, | $`\sigma _1`$, | $`i\sigma _2\}_{(3)}`$ | | $`e_6`$ | $`\{`$ | $`\mathrm{𝟏}`$, | $`\sigma _3`$, | $`\sigma _3`$, | $`\mathrm{𝟏}\}_{(4)}`$ | | $`e_7`$ | $`\{`$ | $`i\sigma _2`$, | $`\sigma _1`$, | $`\sigma _1`$, | $`i\sigma _2\}_{(4)}`$ | | $`1e_1`$ | $`\{`$ | $`i\sigma _2`$, | $`i\sigma _2`$, | $`i\sigma _2`$, | $`i\sigma _2\}_{(1)}`$ | | $`1e_2`$ | $`\{`$ | $`\mathrm{𝟏}`$, | $`\mathrm{𝟏}`$, | $`\mathrm{𝟏}`$, | $`\mathrm{𝟏}\}_{(2)}`$ | | $`1e_3`$ | $`\{`$ | $`i\sigma _2`$, | $`i\sigma _2`$, | $`i\sigma _2`$, | $`i\sigma _2\}_{(2)}`$ | | $`1e_4`$ | $`\{`$ | $`\mathrm{𝟏}`$, | $`\mathrm{𝟏}`$, | $`\mathrm{𝟏}`$, | $`\mathrm{𝟏}\}_{(3)}`$ | | $`1e_5`$ | $`\{`$ | $`i\sigma _2`$, | $`i\sigma _2`$, | $`i\sigma _2`$, | $`i\sigma _2\}_{(3)}`$ | | $`1e_6`$ | $`\{`$ | $`\sigma _3`$, | $`\sigma _3`$, | $`\sigma _3`$, | $`\sigma _3\}_{(4)}`$ | | $`1e_7`$ | $`\{`$ | $`\sigma _1`$, | $`\sigma _1`$, | $`\sigma _1`$, | $`\sigma _1\}_{(4)}`$ | The remaining rules can be easily constructed remembering that $`e_me_m`$ $`1|E_mE_m,`$ $`E_m\mathrm{\hspace{0.33em}\hspace{0.33em}1}|E_m,`$ $`e_m)e_n`$ $`1|E_nE_m,`$ $`e_m(e_n`$ $`E_m\mathrm{\hspace{0.33em}\hspace{0.33em}1}|E_n.`$ For example, $$\begin{array}{cccc}e_1e_1\hfill & \hfill & \left(\begin{array}{cccc}i\sigma _2& 0& 0& 0\\ 0& i\sigma _2& 0& 0\\ 0& 0& i\sigma _2& 0\\ 0& 0& 0& i\sigma _2\end{array}\right)\hfill & \left(\begin{array}{cccc}i\sigma _2& 0& 0& 0\\ 0& i\sigma _2& 0& 0\\ 0& 0& i\sigma _2& 0\\ 0& 0& 0& i\sigma _2\end{array}\right)\hfill \\ & & =\{\mathrm{𝟏},\mathbf{\hspace{0.33em}1},\mathbf{\hspace{0.33em}1},\mathbf{\hspace{0.33em}1}\}_{(1)},\hfill & \end{array}$$ $$\begin{array}{cccc}e_3)e_1\hfill & \hfill & \left(\begin{array}{cccc}i\sigma _2& 0& 0& 0\\ 0& i\sigma _2& 0& 0\\ 0& 0& i\sigma _2& 0\\ 0& 0& 0& i\sigma _2\end{array}\right)\hfill & \left(\begin{array}{cccc}0& \sigma _1& 0& 0\\ \sigma _1& 0& 0& 0\\ 0& 0& 0& i\sigma _2\\ 0& 0& i\sigma _2& 0\end{array}\right)\hfill \\ & & =\{\sigma _3,\sigma _3,\mathbf{\hspace{0.33em}1},\mathrm{𝟏}\}_{(2)},\hfill & \end{array}$$ and $$\begin{array}{cccc}e_3(e_1\hfill & \hfill & \left(\begin{array}{cccc}0& \sigma _1& 0& 0\\ \sigma _1& 0& 0& 0\\ 0& 0& 0& i\sigma _2\\ 0& 0& i\sigma _2& 0\end{array}\right)\hfill & \left(\begin{array}{cccc}i\sigma _2& 0^0& 0& \\ 0& i\sigma _2& 0& 0\\ 0& 0& i\sigma _2& 0\\ 0& 0& 0& i\sigma _2\end{array}\right)\hfill \\ & & =\{\sigma _3,\sigma _3,\mathrm{𝟏},\mathbf{\hspace{0.33em}1}\}_{(2)}.\hfill & \end{array}$$ Following this procedure any matrix representation of right/left barred operators can be obtained. Using Mathematica , we have proved the linear independence of the 64 elements which represent the most general octonionic operator $$𝕆_0+\underset{m=1}{\overset{7}{}}𝕆_m)e_m.$$ So our barred operators form a complete basis for any $`8\times 8`$ real matrix and this establishes the isomorphism between $`GL(8,)`$ and barred octonions. ## Appendix B The Second Appendix We have given the action of barred operators on the octonionic functions (states) $$\psi =\psi _1+e_2\psi _2+e_4\psi _3+e_6\psi _4[\psi _{1,\mathrm{},4}C(1,e_1)].$$ In the following we will use the notation $$e_2\{\psi _2,\psi _1,\psi _4^{},\psi _3^{}\},$$ to indicate $$e_2\psi =\psi _2+e_2\psi _1e_4\psi _4^{}+e_6\psi _3^{}.$$ As occurred in the previous appendix we need to know only the action of the barred operators $`e_m`$ and $`1e_m`$ | $`e_1`$ | $`\{`$ | $`e_1\psi _1`$, | $`e_1\psi _2`$, | $`e_1\psi _3`$, | $`e_1\psi _4\}`$ | | --- | --- | --- | --- | --- | --- | | $`e_2`$ | $`\{`$ | $`\psi _2`$, | $`\psi _1`$, | $`\psi _4^{}`$, | $`\psi _3^{}\}`$ | | $`e_3`$ | $`\{`$ | $`e_1\psi _2`$, | $`e_1\psi _1`$, | $`e_1\psi _4^{}`$, | $`e_1\psi _3^{}\}`$ | | $`e_4`$ | $`\{`$ | $`\psi _3`$, | $`\psi _4^{}`$, | $`\psi _1`$, | $`\psi _2^{}\}`$ | | $`e_5`$ | $`\{`$ | $`e_1\psi _3`$, | $`e_1\psi _4^{}`$, | $`e_1\psi _1`$, | $`e_1\psi _2^{}\}`$ | | $`e_6`$ | $`\{`$ | $`\psi _4`$, | $`\psi _3^{}`$, | $`\psi _2^{}`$, | $`\psi _1\}`$ | | $`e_7`$ | $`\{`$ | $`e_1\psi _4`$, | $`e_1\psi _3^{}`$, | $`e_1\psi _2^{}`$, | $`e_1\psi _1\}`$ | | $`1e_1`$ | $`\{`$ | $`e_1\psi _1`$, | $`e_1\psi _2`$, | $`e_1\psi _3`$, | $`e_1\psi _4\}`$ | | $`1e_2`$ | $`\{`$ | $`\psi _2^{}`$, | $`\psi _1^{}`$, | $`\psi _4^{}`$, | $`\psi _3^{}\}`$ | | $`1e_3`$ | $`\{`$ | $`e_1\psi _2^{}`$, | $`e_1\psi _1^{}`$, | $`e_1\psi _4^{}`$, | $`e_1\psi _3^{}\}`$ | | $`1e_4`$ | $`\{`$ | $`\psi _3^{}`$, | $`\psi _4^{}`$, | $`\psi _1^{}`$, | $`\psi _2^{}\}`$ | | $`1e_5`$ | $`\{`$ | $`e_1\psi _3^{}`$, | $`e_1\psi _4^{}`$, | $`e_1\psi _1^{}`$, | $`e_1\psi _2^{}\}`$ | | $`1e_6`$ | $`\{`$ | $`\psi _4^{}`$, | $`\psi _3^{}`$, | $`\psi _2^{}`$, | $`\psi _1^{}\}`$ | | $`1e_7`$ | $`\{`$ | $`e_1\psi _4^{}`$, | $`e_1\psi _3^{}`$, | $`e_1\psi _2^{}`$, | $`e_1\psi _1^{}\}`$ | From the previous correspondence rules we immediately obtain the others barred operators. We give, as an example, the construction of the operator $`e_4)e_7`$. We know that $$e_4\{\psi _3,\psi _4^{},\psi _1,\psi _2^{}\}$$ and $$1e_7\{e_1\psi _4^{},e_1\psi _3^{},e_1\psi _2^{},e_1\psi _1^{}\}.$$ (B.1) Combining these operators we find $$\{e_1(\psi _2^{})^{},e_1\psi _1^{},e_1(\psi _4^{})^{},e_1(\psi _3)^{}\},$$ and so $$e_4)e_7\{e_1\psi _2,e_1\psi _1^{},e_1\psi _4,e_1\psi _3^{}\}.$$ As remarked at the end of subsection IV-b, we can extract the 32 basis elements of $`GL(4,)`$ directly by suitable combinations of the 64 basis elements of $`GL(8,)`$. We must choose the combination which have only $`\mathrm{𝟏}_{2\times 2}`$ and $`i\sigma _2`$ as matrix elements. Nevertheless we must take care in manipulating our octonionic barred operators. If we wish to extract from $`GL(8,)`$ the 32 elements which characterize $`GL(4,)`$ we need to change the octonionic basis of $`GL(8,)`$. In fact, the natural choice for the symplectic octonionic representation $$\psi =(\phi _0+e_1\phi _1)+e_2(\phi _2+e_1\phi _3)+e_4(\phi _4+e_1\phi _5)+e_6(\phi _6+e_1\phi _7),$$ requires the following real counterpart $$\stackrel{~}{\phi }=\phi _0+e_1\phi _1+e_2\phi _2e_3\phi _3+e_4\phi _4e_5\phi _5+e_6\phi _6+e_7\phi _7.$$ whereas we used in subsection IV-a the following basis $$\phi =\phi _0+e_1\phi _1+e_2\phi _2+e_3\phi _3+e_4\phi _4+e_5\phi _5+e_6\phi _6+e_7\phi _7.$$ The changes in the signs of $`e_3\phi _3`$ and $`e_5\phi _5`$ implies a modification in the generators of $`GL(8,)`$. For example, $`e_2`$ and $`e_3)e_1`$ now read $$e_2\{\mathrm{𝟏},\mathbf{\hspace{0.33em}1},\sigma _3,\sigma _3\}_{(2)}\text{and}e_3)e_1\{\mathrm{𝟏},\mathbf{\hspace{0.33em}1},\sigma _3,\sigma _3\}_{(2)}.$$ i.e. the change of basis induces the following modifications $$\mathrm{𝟏}\sigma _3.$$ Their appropriate combination gives $$\frac{e_2+e_3)e_1}{2}\{0,\mathbf{\hspace{0.33em}1},\mathrm{\hspace{0.33em}0},\mathrm{\hspace{0.33em}0}\}_{(2)}\stackrel{complexifing}{}\left(\begin{array}{cccc}0& 0& 0& 0\\ 1& 0& 0& 0\\ 0& 0& 0& 0\\ 0& 0& 0& 0\end{array}\right),$$ as required by eq. (2.171). ## Appendix C The Third Appendix The seven standard cycles are given by $`f_{123}^{(+)}(\phi )`$ $`=`$ $`{\displaystyle \frac{\phi _{0}^{}{}_{}{}^{2}\phi _{6}^{}{}_{}{}^{2}\phi _{5}^{}{}_{}{}^{2}+\phi _{2}^{}{}_{}{}^{2}\phi _{4}^{}{}_{}{}^{2}+\phi _{1}^{}{}_{}{}^{2}+\phi _{3}^{}{}_{}{}^{2}\phi _{7}^{}{}_{}{}^{2}}{r^2}},`$ $`f_{145}^{(+)}(\phi )`$ $`=`$ $`{\displaystyle \frac{\phi _{0}^{}{}_{}{}^{2}\phi _{6}^{}{}_{}{}^{2}+\phi _{4}^{}{}_{}{}^{2}\phi _{2}^{}{}_{}{}^{2}+\phi _{1}^{}{}_{}{}^{2}\phi _{7}^{}{}_{}{}^{2}+\phi _{5}^{}{}_{}{}^{2}\phi _{3}^{}{}_{}{}^{2}}{r^2}},`$ $`f_{176}^{(+)}(\phi )`$ $`=`$ $`{\displaystyle \frac{\phi _{0}^{}{}_{}{}^{2}+\phi _{6}^{}{}_{}{}^{2}\phi _{4}^{}{}_{}{}^{2}\phi _{2}^{}{}_{}{}^{2}+\phi _{1}^{}{}_{}{}^{2}+\phi _{7}^{}{}_{}{}^{2}\phi _{5}^{}{}_{}{}^{2}\phi _{3}^{}{}_{}{}^{2}}{r^2}},`$ $`f_{246}^{(+)}(\phi )`$ $`=`$ $`{\displaystyle \frac{\phi _{0}^{}{}_{}{}^{2}+\phi _{6}^{}{}_{}{}^{2}+\phi _{4}^{}{}_{}{}^{2}+\phi _{2}^{}{}_{}{}^{2}\phi _{1}^{}{}_{}{}^{2}\phi _{7}^{}{}_{}{}^{2}\phi _{5}^{}{}_{}{}^{2}\phi _{3}^{}{}_{}{}^{2}}{r^2}},`$ $`f_{257}^{(+)}(\phi )`$ $`=`$ $`{\displaystyle \frac{\phi _{0}^{}{}_{}{}^{2}\phi _{6}^{}{}_{}{}^{2}\phi _{4}^{}{}_{}{}^{2}+\phi _{2}^{}{}_{}{}^{2}\phi _{1}^{}{}_{}{}^{2}+\phi _{7}^{}{}_{}{}^{2}+\phi _{5}^{}{}_{}{}^{2}\phi _{3}^{}{}_{}{}^{2}}{r^2}},`$ $`f_{347}^{(+)}(\phi )`$ $`=`$ $`{\displaystyle \frac{\phi _{0}^{}{}_{}{}^{2}\phi _{6}^{}{}_{}{}^{2}+\phi _{4}^{}{}_{}{}^{2}\phi _{2}^{}{}_{}{}^{2}\phi _{1}^{}{}_{}{}^{2}+\phi _{7}^{}{}_{}{}^{2}\phi _{5}^{}{}_{}{}^{2}+\phi _{3}^{}{}_{}{}^{2}}{r^2}},`$ $`f_{365}^{(+)}(\phi )`$ $`=`$ $`{\displaystyle \frac{\phi _{0}^{}{}_{}{}^{2}+\phi _{6}^{}{}_{}{}^{2}\phi _{4}^{}{}_{}{}^{2}\phi _{2}^{}{}_{}{}^{2}\phi _{1}^{}{}_{}{}^{2}\phi _{7}^{}{}_{}{}^{2}+\phi _{5}^{}{}_{}{}^{2}+\phi _{3}^{}{}_{}{}^{2}}{r^2}},`$ where $$r^2=(\phi _0^2+\phi _1^2+\phi _2^2+\phi _3^2+\phi _4^2+\phi _5^2+\phi _6^2+\phi _7^2)$$ (C.2) and the non-standard subset $`f_{124}^{(+)}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _0\phi _7\phi _5\phi _2+\phi _6\phi _1+\phi _3\phi _4}{r^2}},`$ $`f_{125}^{(+)}(\phi )`$ $`=`$ $`2{\displaystyle \frac{\phi _0\phi _6\phi _3\phi _5\phi _1\phi _7\phi _2\phi _4}{r^2}},`$ $`f_{126}^{(+)}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _0\phi _5\phi _1\phi _4+\phi _7\phi _2+\phi _3\phi _6}{r^2}},`$ $`f_{127}^{(+)}(\phi )`$ $`=`$ $`2{\displaystyle \frac{\phi _0\phi _4+\phi _6\phi _2+\phi _1\phi _5\phi _3\phi _7}{r^2}},`$ $`f_{143}^{(+)}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _0\phi _6+\phi _3\phi _5+\phi _2\phi _4\phi _1\phi _7}{r^2}},`$ $`f_{146}^{(+)}(\phi )`$ $`=`$ $`2{\displaystyle \frac{\phi _3\phi _0\phi _4\phi _7\phi _1\phi _2\phi _5\phi _6}{r^2}},`$ $`f_{175}^{(+)}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _3\phi _0\phi _1\phi _2+\phi _5\phi _6+\phi _4\phi _7}{r^2}},`$ $`f_{247}^{(+)}(\phi )`$ $`=`$ $`2{\displaystyle \frac{\phi _0\phi _1\phi _7\phi _6\phi _4\phi _5\phi _3\phi _2}{r^2}},`$ $`f_{147}^{(+)}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _2\phi _0\phi _4\phi _6+\phi _5\phi _7+\phi _1\phi _3}{r^2}},`$ $`f_{243}^{(+)}(\phi )`$ $`=`$ $`2{\displaystyle \frac{\phi _0\phi _5+\phi _1\phi _4+\phi _7\phi _2\phi _3\phi _6}{r^2}},`$ $`f_{253}^{(+)}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _0\phi _4\phi _1\phi _5+\phi _6\phi _2+\phi _3\phi _7}{r^2}},`$ $`f_{173}^{(+)}(\phi )`$ $`=`$ $`2{\displaystyle \frac{\phi _0\phi _5\phi _7\phi _2\phi _3\phi _6\phi _1\phi _4}{r^2}},`$ $`f_{245}^{(+)}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _3\phi _0+\phi _5\phi _6\phi _4\phi _7+\phi _1\phi _2}{r^2}},`$ $`f_{256}^{(+)}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _0\phi _1\phi _3\phi _2+\phi _7\phi _6+\phi _4\phi _5}{r^2}},`$ $`f_{361}^{(+)}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _0\phi _4+\phi _3\phi _7+\phi _1\phi _5\phi _6\phi _2}{r^2}},`$ $`f_{362}^{(+)}(\phi )`$ $`=`$ $`2{\displaystyle \frac{\phi _0\phi _7\phi _3\phi _4\phi _5\phi _2\phi _6\phi _1}{r^2}},`$ $`f_{345}^{(+)}(\phi )`$ $`=`$ $`2{\displaystyle \frac{\phi _2\phi _0\phi _5\phi _7\phi _1\phi _3\phi _4\phi _6}{r^2}},`$ $`f_{346}^{(+)}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _0\phi _1+\phi _3\phi _2+\phi _7\phi _6\phi _4\phi _5}{r^2}},`$ $`f_{367}^{(+)}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _2\phi _0+\phi _4\phi _6+\phi _5\phi _7\phi _1\phi _3}{r^2}},`$ $`f_{135}^{(+)}(\phi )`$ $`=`$ $`2{\displaystyle \frac{\phi _0\phi _7\phi _3\phi _4+\phi _6\phi _1+\phi _5\phi _2}{r^2}},`$ $`f_{156}^{(+)}(\phi )`$ $`=`$ $`2{\displaystyle \frac{\phi _2\phi _0+\phi _1\phi _3+\phi _4\phi _6\phi _5\phi _7}{r^2}},`$ $`f_{237}^{(+)}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _0\phi _6\phi _2\phi _4+\phi _3\phi _5+\phi _1\phi _7}{r^2}},`$ $`f_{267}^{(+)}(\phi )`$ $`=`$ $`2{\displaystyle \frac{\phi _3\phi _0\phi _5\phi _6+\phi _4\phi _7+\phi _1\phi _2}{r^2}},`$ $`f_{357}^{(+)}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _0\phi _1+\phi _4\phi _5\phi _7\phi _6+\phi _3\phi _2}{r^2}},`$ $`f_{456}^{(+)}(\phi )`$ $`=`$ $`2{\displaystyle \frac{\phi _0\phi _7+\phi _5\phi _2+\phi _3\phi _4\phi _6\phi _1}{r^2}},`$ $`f_{457}^{(+)}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _0\phi _6+\phi _2\phi _4+\phi _1\phi _7\phi _3\phi _5}{r^2}},`$ $`f_{467}^{(+)}(\phi )`$ $`=`$ $`2{\displaystyle \frac{\phi _0\phi _5+\phi _1\phi _4+\phi _3\phi _6\phi _7\phi _2}{r^2}},`$ $`f_{567}^{(+)}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _0\phi _4\phi _6\phi _2\phi _1\phi _5\phi _3\phi _7}{r^2}}.`$ For right actions, the standard cocycles are $`f_{123}^{()}(\phi )`$ $`=`$ $`{\displaystyle \frac{\phi _{0}^{}{}_{}{}^{2}\phi _{6}^{}{}_{}{}^{2}\phi _{4}^{}{}_{}{}^{2}+\phi _{2}^{}{}_{}{}^{2}+\phi _{1}^{}{}_{}{}^{2}\phi _{7}^{}{}_{}{}^{2}\phi _{5}^{}{}_{}{}^{2}+\phi _{3}^{}{}_{}{}^{2}}{r^2}},`$ $`f_{145}^{()}(\phi )`$ $`=`$ $`{\displaystyle \frac{\phi _{0}^{}{}_{}{}^{2}\phi _{6}^{}{}_{}{}^{2}+\phi _{4}^{}{}_{}{}^{2}\phi _{2}^{}{}_{}{}^{2}+\phi _{1}^{}{}_{}{}^{2}\phi _{7}^{}{}_{}{}^{2}+\phi _{5}^{}{}_{}{}^{2}\phi _{3}^{}{}_{}{}^{2}}{r^2}},`$ $`f_{176}^{()}(\phi )`$ $`=`$ $`{\displaystyle \frac{\phi _{0}^{}{}_{}{}^{2}+\phi _{6}^{}{}_{}{}^{2}\phi _{4}^{}{}_{}{}^{2}\phi _{2}^{}{}_{}{}^{2}+\phi _{1}^{}{}_{}{}^{2}+\phi _{7}^{}{}_{}{}^{2}\phi _{5}^{}{}_{}{}^{2}\phi _{3}^{}{}_{}{}^{2}}{r^2}},`$ $`f_{246}^{()}(\phi )`$ $`=`$ $`{\displaystyle \frac{\phi _{0}^{}{}_{}{}^{2}+\phi _{6}^{}{}_{}{}^{2}+\phi _{4}^{}{}_{}{}^{2}+\phi _{2}^{}{}_{}{}^{2}\phi _{1}^{}{}_{}{}^{2}\phi _{7}^{}{}_{}{}^{2}\phi _{5}^{}{}_{}{}^{2}\phi _{3}^{}{}_{}{}^{2}}{r^2}},`$ $`f_{257}^{()}(\phi )`$ $`=`$ $`{\displaystyle \frac{\phi _{0}^{}{}_{}{}^{2}\phi _{6}^{}{}_{}{}^{2}\phi _{4}^{}{}_{}{}^{2}+\phi _{2}^{}{}_{}{}^{2}\phi _{1}^{}{}_{}{}^{2}+\phi _{7}^{}{}_{}{}^{2}+\phi _{5}^{}{}_{}{}^{2}\phi _{3}^{}{}_{}{}^{2}}{r^2}},`$ $`f_{347}^{()}(\phi )`$ $`=`$ $`{\displaystyle \frac{\phi _{0}^{}{}_{}{}^{2}\phi _{6}^{}{}_{}{}^{2}+\phi _{4}^{}{}_{}{}^{2}\phi _{2}^{}{}_{}{}^{2}\phi _{1}^{}{}_{}{}^{2}+\phi _{7}^{}{}_{}{}^{2}\phi _{5}^{}{}_{}{}^{2}+\phi _{3}^{}{}_{}{}^{2}}{r^2}},`$ $`f_{365}^{()}(\phi )`$ $`=`$ $`{\displaystyle \frac{\phi _{0}^{}{}_{}{}^{2}+\phi _{6}^{}{}_{}{}^{2}\phi _{4}^{}{}_{}{}^{2}\phi _{2}^{}{}_{}{}^{2}\phi _{1}^{}{}_{}{}^{2}\phi _{7}^{}{}_{}{}^{2}+\phi _{5}^{}{}_{}{}^{2}+\phi _{3}^{}{}_{}{}^{2}}{r^2}},`$ while the non-standard cocycles are $`f_{124}^{()}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _0\phi _7+\phi _5\phi _2\phi _6\phi _1\phi _3\phi _4}{r^2}},`$ $`f_{125}^{()}(\phi )`$ $`=`$ $`2{\displaystyle \frac{\phi _0\phi _6+\phi _3\phi _5+\phi _1\phi _7+\phi _2\phi _4}{r^2}},`$ $`f_{126}^{()}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _0\phi _5+\phi _1\phi _4\phi _7\phi _2\phi _3\phi _6}{r^2}},`$ $`f_{127}^{()}(\phi )`$ $`=`$ $`2{\displaystyle \frac{\phi _0\phi _4\phi _6\phi _2\phi _1\phi _5+\phi _3\phi _7}{r^2}},`$ $`f_{143}^{()}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _0\phi _6\phi _3\phi _5\phi _2\phi _4+\phi _1\phi _7}{r^2}},`$ $`f_{146}^{()}(\phi )`$ $`=`$ $`2{\displaystyle \frac{\phi _3\phi _0+\phi _4\phi _7+\phi _1\phi _2+\phi _5\phi _6}{r^2}},`$ $`f_{175}^{()}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _3\phi _0+\phi _1\phi _2\phi _5\phi _6\phi _4\phi _7}{r^2}},`$ $`f_{247}^{()}(\phi )`$ $`=`$ $`2{\displaystyle \frac{\phi _0\phi _1+\phi _7\phi _6+\phi _4\phi _5+\phi _3\phi _2}{r^2}},`$ $`f_{147}^{()}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _2\phi _0+\phi _4\phi _6\phi _5\phi _7\phi _1\phi _3}{r^2}},`$ $`f_{243}^{()}(\phi )`$ $`=`$ $`2{\displaystyle \frac{\phi _0\phi _5\phi _1\phi _4\phi _7\phi _2+\phi _3\phi _6}{r^2}},`$ $`f_{253}^{()}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _0\phi _4+\phi _1\phi _5\phi _6\phi _2\phi _3\phi _7}{r^2}},`$ $`f_{173}^{()}(\phi )`$ $`=`$ $`2{\displaystyle \frac{\phi _0\phi _5+\phi _7\phi _2+\phi _3\phi _6+\phi _1\phi _4}{r^2}},`$ $`f_{245}^{()}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _3\phi _0\phi _5\phi _6+\phi _4\phi _7\phi _1\phi _2}{r^2}},`$ $`f_{256}^{()}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _0\phi _1+\phi _3\phi _2\phi _7\phi _6\phi _4\phi _5}{r^2}},`$ $`f_{361}^{()}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _0\phi _4\phi _3\phi _7\phi _1\phi _5+\phi _6\phi _2}{r^2}},`$ $`f_{362}^{()}(\phi )`$ $`=`$ $`2{\displaystyle \frac{\phi _0\phi _7+\phi _3\phi _4+\phi _5\phi _2+\phi _6\phi _1}{r^2}},`$ $`f_{345}^{()}(\phi )`$ $`=`$ $`2{\displaystyle \frac{\phi _2\phi _0+\phi _5\phi _7+\phi _1\phi _3+\phi _4\phi _6}{r^2}},`$ $`f_{346}^{()}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _0\phi _1\phi _3\phi _2\phi _7\phi _6+\phi _4\phi _5}{r^2}},`$ $`f_{367}^{()}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _2\phi _0\phi _4\phi _6\phi _5\phi _7+\phi _1\phi _3}{r^2}},`$ $`f_{135}^{()}(\phi )`$ $`=`$ $`2{\displaystyle \frac{\phi _0\phi _7+\phi _3\phi _4\phi _6\phi _1\phi _5\phi _2}{r^2}},`$ $`f_{156}^{()}(\phi )`$ $`=`$ $`2{\displaystyle \frac{\phi _2\phi _0\phi _1\phi _3\phi _4\phi _6+\phi _5\phi _7}{r^2}},`$ $`f_{237}^{()}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _0\phi _6+\phi _2\phi _4\phi _3\phi _5\phi _1\phi _7}{r^2}},`$ $`f_{267}^{()}(\phi )`$ $`=`$ $`2{\displaystyle \frac{\phi _3\phi _0+\phi _5\phi _6\phi _4\phi _7\phi _1\phi _2}{r^2}},`$ $`f_{357}^{()}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _0\phi _1\phi _4\phi _5+\phi _7\phi _6\phi _3\phi _2}{r^2}},`$ $`f_{456}^{()}(\phi )`$ $`=`$ $`2{\displaystyle \frac{\phi _0\phi _7\phi _5\phi _2\phi _3\phi _4+\phi _6\phi _1}{r^2}},`$ $`f_{457}^{()}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _0\phi _6\phi _2\phi _4\phi _1\phi _7+\phi _3\phi _5}{r^2}},`$ $`f_{467}^{()}(\phi )`$ $`=`$ $`2{\displaystyle \frac{\phi _0\phi _5\phi _1\phi _4\phi _3\phi _6+\phi _7\phi _2}{r^2}},`$ $`f_{567}^{()}(\phi )`$ $`=`$ $`+2{\displaystyle \frac{\phi _0\phi _4+\phi _6\phi _2+\phi _1\phi _5+\phi _3\phi _7}{r^2}}.`$ ## Appendix D The Fourth Appendix We give below some examples of the torsionful structure functions, $`f_{123}^{\left(++\right)}(\phi ,\lambda )=(y_0{}_{}{}^{2}x_{3}^{}{}_{}{}^{2}+y_0{}_{}{}^{2}x_{2}^{}{}_{}{}^{2}y_0{}_{}{}^{2}x_{5}^{}{}_{}{}^{2}y_0{}_{}{}^{2}x_{4}^{}{}_{}{}^{2}+y_0{}_{}{}^{2}x_{6}^{}{}_{}{}^{2}y_0{}_{}{}^{2}x_{7}^{}{}_{}{}^{2}`$ $`+y_{0}^{}{}_{}{}^{2}x_{1}^{}{}_{}{}^{2}+y_{7}^{}{}_{}{}^{2}x_{4}^{}{}_{}{}^{2}+y_{7}^{}{}_{}{}^{2}x_{6}^{}{}_{}{}^{2}4y_0y_6x_1x_7`$ $`y_{3}^{}{}_{}{}^{2}x_{4}^{}{}_{}{}^{2}y_0{}_{}{}^{2}x_{4}^{}{}_{}{}^{2}y_0{}_{}{}^{2}x_{6}^{}{}_{}{}^{2}y_0{}_{}{}^{2}x_{7}^{}^2`$ $`y_{3}^{}{}_{}{}^{2}x_{6}^{}{}_{}{}^{2}+4y_0y_4x_3x_74y_0y_7x_5x_24y_0y_7x_3x_4`$ $`+4y_0y_7x_6x_1+y_{1}^{}{}_{}{}^{2}x_{1}^{}{}_{}{}^{2}y_{1}^{}{}_{}{}^{2}x_{7}^{}{}_{}{}^{2}y_{1}^{}{}_{}{}^{2}x_{5}^{}{}_{}{}^{2}+y_{1}^{}{}_{}{}^{2}x_{2}^{}{}_{}{}^{2}`$ $`y_{1}^{}{}_{}{}^{2}x_{4}^{}{}_{}{}^{2}y_{1}^{}{}_{}{}^{2}x_{6}^{}{}_{}{}^{2}+y_{1}^{}{}_{}{}^{2}x_{3}^{}{}_{}{}^{2}4y_0y_5x_1x_4`$ $`4y_0y_5x_3x_6+4y_0y_4x_2x_6+4y_0y_6x_3x_5`$ $`4y_0y_6x_2x_4+4y_1y_4x_7x_2+4y_0y_4x_1x_5`$ $`+4y_1y_5x_1x_5+4y_1y_4x_1x_4y_{2}^{}{}_{}{}^{2}x_{4}^{}{}_{}{}^{2}4y_1y_5x_3x_7`$ $`+y_{2}^{}{}_{}{}^{2}x_{2}^{}{}_{}{}^{2}+4y_1y_6x_3x_4+4y_1y_6x_5x_2+y_{2}^{}{}_{}{}^{2}x_{1}^{}{}_{}{}^{2}`$ $`+4y_1y_6x_6x_1y_{2}^{}{}_{}{}^{2}x_{6}^{}{}_{}{}^{2}+4y_3y_5x_2x_4y_{7}^{}{}_{}{}^{2}x_{1}^{}{}_{}{}^{2}`$ $`y_{2}^{}{}_{}{}^{2}x_{5}^{}{}_{}{}^{2}+4y_3y_4x_3x_44y_3y_7x_6x_2+4y_3y_7x_3x_7`$ $`+4y_1y_7x_1x_74y_3y_4x_5x_2y_{6}^{}{}_{}{}^{2}x_{1}^{}{}_{}{}^{2}+y_{6}^{}{}_{}{}^{2}x_{5}^{}{}_{}{}^{2}`$ $`y_{4}^{}{}_{}{}^{2}x_{3}^{}{}_{}{}^{2}+4y_3y_4x_6x_1y_{6}^{}{}_{}{}^{2}x_{2}^{}{}_{}{}^{2}+4y_3y_5x_3x_5`$ $`+4y_3y_6x_7x_2+y_{6}^{}{}_{}{}^{2}x_{4}^{}{}_{}{}^{2}+4y_2y_7x_1x_4+y_{6}^{}{}_{}{}^{2}x_{6}^{}{}_{}{}^{2}`$ $`+y_{6}^{}{}_{}{}^{2}x_{7}^{}{}_{}{}^{2}4y_3y_6x_1x_4+4y_3y_6x_3x_6y_{6}^{}{}_{}{}^{2}x_{3}^{}{}_{}{}^{2}`$ $`+y_{5}^{}{}_{}{}^{2}x_{4}^{}{}_{}{}^{2}+y_{5}^{}{}_{}{}^{2}x_{6}^{}{}_{}{}^{2}+y_{5}^{}{}_{}{}^{2}x_{5}^{}{}_{}{}^{2}+4y_3y_5x_1x_7+y_{5}^{}{}_{}{}^{2}x_{7}^{}{}_{}{}^{2}`$ $`4y_1y_4x_3x_6+y_{2}^{}{}_{}{}^{2}x_{3}^{}{}_{}{}^{2}y_{3}^{}{}_{}{}^{2}x_{7}^{}{}_{}{}^{2}4y_1y_7x_2x_4`$ $`+4y_1y_7x_3x_5y_{2}^{}{}_{}{}^{2}x_{7}^{}{}_{}{}^{2}4y_1y_5x_6x_24y_3y_7x_1x_5`$ $`y_{5}^{}{}_{}{}^{2}x_{2}^{}{}_{}{}^{2}y_{5}^{}{}_{}{}^{2}x_{3}^{}{}_{}{}^{2}+4y_0y_5x_2x_74y_2y_6x_5x_1`$ $`+4y_2y_6x_6x_2+y_{3}^{}{}_{}{}^{2}x_{1}^{}{}_{}{}^{2}y_{3}^{}{}_{}{}^{2}x_{5}^{}{}_{}{}^{2}+4y_2y_5x_5x_2`$ $`+y_{7}^{}{}_{}{}^{2}x_{7}^{}{}_{}{}^{2}y_{7}^{}{}_{}{}^{2}x_{2}^{}{}_{}{}^{2}+y_{4}^{}{}_{}{}^{2}x_{5}^{}{}_{}{}^{2}y_{5}^{}{}_{}{}^{2}x_{1}^{}{}_{}{}^{2}+y_{7}^{}{}_{}{}^{2}x_{5}^{}{}_{}{}^{2}`$ $`+4y_2y_7x_3x_6y_{7}^{}{}_{}{}^{2}x_{3}^{}{}_{}{}^{2}+4y_2y_7x_7x_2+4y_2y_5x_6x_1`$ $`4y_2y_5x_3x_44y_2y_6x_3x_74y_2y_4x_7x_1+y_{4}^{}{}_{}{}^{2}x_{7}^{}{}_{}{}^{2}`$ $`4x_0y_1y_6x_74x_0y_3y_7x_4+4x_0y_2y_5x_7`$ $`4x_0y_3y_5x_6+x_{0}^{}{}_{}{}^{2}y_{0}^{}{}_{}{}^{2}+x_{0}^{}{}_{}{}^{2}y_{3}^{}{}_{}{}^{2}+x_{0}^{}{}_{}{}^{2}y_{1}^{}{}_{}{}^{2}x_{0}^{}{}_{}{}^{2}y_{6}^{}{}_{}{}^{2}`$ $`+x_{0}^{}{}_{}{}^{2}y_{2}^{}{}_{}{}^{2}x_{0}^{}{}_{}{}^{2}y_{4}^{}{}_{}{}^{2}x_{0}^{}{}_{}{}^{2}y_{7}^{}{}_{}{}^{2}x_{0}^{}{}_{}{}^{2}y_{5}^{}{}_{}{}^{2}+y_{4}^{}{}_{}{}^{2}x_{6}^{}{}_{}{}^{2}`$ $`+y_{4}^{}{}_{}{}^{2}x_{4}^{}{}_{}{}^{2}y_{4}^{}{}_{}{}^{2}x_{2}^{}{}_{}{}^{2}+4y_2y_4x_3x_5+4y_2y_4x_2x_4`$ $`y_{4}^{}{}_{}{}^{2}x_{1}^{}{}_{}{}^{2}+y_{3}^{}{}_{}{}^{2}x_{2}^{}{}_{}{}^{2}+y_{3}^{}{}_{}{}^{2}x_{3}^{}{}_{}{}^{2}4x_0y_2y_7x_5`$ $`+4x_0y_2y_4x_64x_0y_2y_6x_4+4x_0y_3y_6x_5`$ $`4x_0y_0y_7x_74x_0y_1y_5x_44x_0y_0y_5x_5`$ $`+4x_0y_3y_4x_74x_0y_0y_4x_4+4x_0y_1y_7x_6`$ $`4x_0y_0y_6x_6+4x_0y_1y_4x_5)/`$ $`((y_{0}^{}{}_{}{}^{2}+y_{1}^{}{}_{}{}^{2}+y_{2}^{}{}_{}{}^{2}+y_{3}^{}{}_{}{}^{2}+y_{4}^{}{}_{}{}^{2}+y_{5}^{}{}_{}{}^{2}+y_{6}^{}{}_{}{}^{2}+y_{7}^{}{}_{}{}^{2})`$ $`(x_{0}^{}{}_{}{}^{2}+x_{5}^{}{}_{}{}^{2}+x_{3}^{}{}_{}{}^{2}+x_{1}^{}{}_{}{}^{2}+x_{2}^{}{}_{}{}^{2}+x_{7}^{}{}_{}{}^{2}+x_{6}^{}{}_{}{}^{2}+x_{4}^{}{}_{}{}^{2}))`$ $`f_{127}^{\left(++\right)}(\phi ,\lambda )=2(y_0y_4x_5{}_{}{}^{2}2y_0y_6x_1x_32y_0y_5x_4x_5`$ $`+2y_0y_5x_6x_7+2y_0y_5x_2x_3+y_{1}^{}{}_{}{}^{2}x_6x_2+y_{1}^{}{}_{}{}^{2}x_5x_{1}^{}{}_{1}{}^{}x_32y_0y_5x_4x_5`$ $`2y_1y_6x_1x_2y_1y_5x_{3}^{}{}_{}{}^{2}y_0y_4x_{1}^{}{}_{}{}^{2}+y_0y_4x_{3}^{}{}_{}{}^{2}`$ $`2y_1y_6x_4x_7+2y_1y_6x_5x_6+2y_1y_4x_4x_5`$ $`+2y_1y_3x_1x_7+2y_1y_4x_6x_7+2y_1y_4x_2x_3`$ $`y_{1}^{}{}_{}{}^{2}x_3x_7+2y_2y_5x_5x_6y_{3}^{}{}_{}{}^{2}x_6x_2y_{3}^{}{}_{}{}^{2}x_1x_5`$ $`+2y_1y_3x_3x_52y_1y_3x_2x_4y_1y_5x_{1}^{}{}_{}{}^{2}+y_1y_5x_{5}^{}{}_{}{}^{2}`$ $`+y_1y_5x_{7}^{}{}_{}{}^{2}+y_2y_6x_{6}^{}{}_{}{}^{2}y_2y_6x_{4}^{}{}_{}{}^{2}y_1y_5x_{6}^{}{}_{}{}^{2}`$ $`y_1y_5x_{4}^{}{}_{}{}^{2}+y_1y_5x_{2}^{}{}_{}{}^{2}y_{0}^{}{}_{}{}^{2}x_3x_7+y_{0}^{}{}_{}{}^{2}x_5x_1`$ $`+y_{0}^{}{}_{}{}^{2}x_6x_22y_0y_6x_4x_62y_0y_6x_5x_7y_0y_4x_{2}^{}{}_{}{}^{2}`$ $`+y_{2}^{}{}_{}{}^{2}x_6x_2+y_2y_6x_{7}^{}{}_{}{}^{2}y_2y_6x_{2}^{}{}_{}{}^{2}y_2y_6x_{5}^{}{}_{}{}^{2}`$ $`y_2y_6x_{3}^{}{}_{}{}^{2}y_0y_4x_{7}^{}{}_{}{}^{2}+y_0y_4x_{6}^{}{}_{}{}^{2}y_0y_4x_{4}^{}{}_{}{}^{2}`$ $`+y_{2}^{}{}_{}{}^{2}x_5x_1y_{2}^{}{}_{}{}^{2}x_3x_7+2y_0y_3x_6x_1+2y_2y_3x_1x_4`$ $`2y_0y_3x_3x_42y_0y_3x_5x_22y_2y_4x_5x_7`$ $`+2y_2y_4x_4x_62y_2y_4x_1x_32y_4y_7x_6x_1`$ $`+2y_4y_7x_2x_52y_5y_7x_7x_12y_5y_7x_3x_5`$ $`2y_5y_7x_2x_42y_6y_7x_7x_2+2y_6y_7x_1x_4`$ $`+x_{0}^{}{}_{}{}^{2}y_0y_4+2y_2y_3x_7x_2+2y_2y_3x_3x_6+x_0y_{7}^{}{}_{}{}^{2}x_4`$ $`x_0y_{3}^{}{}_{}{}^{2}x_4x_0y_{4}^{}{}_{}{}^{2}x_4+x_0y_{2}^{}{}_{}{}^{2}x_4+x_0y_{0}^{}{}_{}{}^{2}x_4`$ $`x_0y_{5}^{}{}_{}{}^{2}x_42y_2y_5x_1x_2+2y_2y_5x_4x_7`$ $`2y_4y_7x_3x_4+y_2y_6x_{1}^{}{}_{}{}^{2}+y_{3}^{}{}_{}{}^{2}x_3x_7+y_3y_7x_{4}^{}{}_{}{}^{2}`$ $`+y_3y_7x_{6}^{}{}_{}{}^{2}y_3y_7x_{2}^{}{}_{}{}^{2}+y_3y_7x_{5}^{}{}_{}{}^{2}y_3y_7x_{3}^{}{}_{}{}^{2}`$ $`y_3y_7x_{1}^{}{}_{}{}^{2}+y_3y_7x_{7}^{}{}_{}{}^{2}y_{4}^{}{}_{}{}^{2}x_6x_2+y_{4}^{}{}_{}{}^{2}x_7x_3`$ $`y_{4}^{}{}_{}{}^{2}x_1x_5y_{5}^{}{}_{}{}^{2}x_6x_2y_{5}^{}{}_{}{}^{2}x_5x_1+y_{5}^{}{}_{}{}^{2}x_3x_7`$ $`+y_{7}^{}{}_{}{}^{2}x_6x_2+y_{7}^{}{}_{}{}^{2}x_5x_1y_{6}^{}{}_{}{}^{2}x_5x_1y_{6}^{}{}_{}{}^{2}x_2x_6`$ $`+y_{6}^{}{}_{}{}^{2}x_3x_7y_{7}^{}{}_{}{}^{2}x_3x_7+x_{0}^{}{}_{}{}^{2}y_2y_62y_6y_7x_3x_6`$ $`x_{0}^{}{}_{}{}^{2}y_3y_7+x_{0}^{}{}_{}{}^{2}y_1y_5x_0y_{6}^{}{}_{}{}^{2}x_4+x_0y_{1}^{}{}_{}{}^{2}x_4`$ $`2x_0y_0y_3x_72x_0y_4y_7x_72x_0y_1y_6x_3`$ $`2x_0y_2y_3x_5+2x_0y_5y_7x_6+2x_0y_0y_5x_1`$ $`+2x_0y_0y_6x_22x_0y_2y_4x_2+2x_0y_1y_3x_6`$ $`+2x_0y_2y_5x_32x_0y_6y_7x_52x_0y_1y_4x_1)/`$ $`((y_{0}^{}{}_{}{}^{2}+y_{1}^{}{}_{}{}^{2}+y_{2}^{}{}_{}{}^{2}+y_{3}^{}{}_{}{}^{2}+y_{4}^{}{}_{}{}^{2}+y_{5}^{}{}_{}{}^{2}+y_{6}^{}{}_{}{}^{2}+y_{7}^{}{}_{}{}^{2})`$ $`(x_{0}^{}{}_{}{}^{2}+x_{5}^{}{}_{}{}^{2}+x_{3}^{}{}_{}{}^{2}+x_{1}^{}{}_{}{}^{2}+x_{2}^{}{}_{}{}^{2}+x_{7}^{}{}_{}{}^{2}+x_{6}^{}{}_{}{}^{2}+x_{4}^{}{}_{}{}^{2}))`$
warning/0002/hep-ph0002081.html
ar5iv
text
# Lambda Polarization in Polarized Proton-Proton Collisions at RHIC ## I Introduction Measurements of the polarization dependent structure function, $`g_1`$, in deep inelastic scattering have inspired considerable experimental and theoretical effort to understand the spin structure of baryons. While most of these studies concern the spin structure of the nucleons, it has become clear that similar measurements involving other baryons would provide helpful, complementary information . The Lambda baryon plays a special role in this respect. It is an ideal testing ground for spin studies since it has a rather simple spin structure in the naive quark parton model. Furthermore, its self-analyzing decay makes polarization measurements experimentally feasible. Forthcoming experiments at RHIC could measure the polarization of Lambda hyperons produced in proton-proton collisions with one of the protons longitudinally polarized, $`p^{}p\mathrm{\Lambda }^{}X`$. The polarization dependent fragmentation function of quarks and gluons into Lambda hyperons can be extracted from such experiments. These fragmentation functions contain information on how the spin of the polarized quarks and gluons is transferred to the final state Lambda. The advantage of proton proton collisions, as opposed to $`e^+e^{}`$ annihilation, where $`\mathrm{\Lambda }`$ production and polarization is dominated by strange quark fragmentation, is that Lambdas at large positive rapidity are mainly fragmentation products of up and down valence quarks of the polarized projectile. Thus, the important question, intimately related to our understanding of the spin structure of baryons, of whether polarized up and down quarks can transfer polarization to the Lambda can be tested at RHIC . In a previous publication, we have shown that the hyperfine interaction, responsible for the $`\mathrm{\Delta }`$-$`N`$ and $`\mathrm{\Sigma }^0`$-$`\mathrm{\Lambda }`$ mass splittings leads to non-zero polarized non-strange quark fragmentation functions . These non-zero polarized up and down quark fragmentation functions give rise to sizeable positive $`\mathrm{\Lambda }`$ polarization in experiments where the strange quark fragmentation is suppressed. On the other hand, predictions based either on the naive quark model or on $`SU(3)`$ flavor symmetry predict zero or negative Lambda polarization . In section II, we briefly discuss fragmentation functions and show how the hyperfine interaction leads to polarized non-strange fragmentation functions. We fix the parameters of the model by fitting the data on $`\mathrm{\Lambda }`$ production in $`e^+e^{}`$ annihilation. In section III, we discuss $`\mathrm{\Lambda }`$ production in pp collisons at RHIC energies. We point out that the production of $`\mathrm{\Lambda }`$’s at high rapidities is dominated by the fragmentation of valence up and down quarks of the polarized projectile, and is ideally suited to test whether non-strange quarks transfer their polarization to the final state $`\mathrm{\Lambda }`$. We predict significant positive $`\mathrm{\Lambda }`$-polarization at large rapidities of the produced $`\mathrm{\Lambda }`$. ## II Fragmentation functions Fragmentation functions can be defined as light-cone Fourier transforms of matrix elements of quark operators $$\frac{1}{z}D_{q\mathrm{\Lambda }}^\mathrm{\Gamma }(z)=\frac{1}{4}\underset{n}{}\frac{d\xi ^{}}{2\pi }e^{iP^+\xi ^{}/z}\text{Tr}\{\mathrm{\Gamma }0|\psi (0)|\mathrm{\Lambda }(PS);n(p_n)\mathrm{\Lambda }(PS);n(p_n)|\overline{\psi }(\xi ^{})|0\},$$ (1) where, $`\mathrm{\Gamma }`$ is the appropriate Dirac matrix; $`P`$ and $`p_n`$ refer to the momentum of the produced $`\mathrm{\Lambda }`$ and of the intermediate system $`n`$; $`S`$ is the spin of the Lambda and the plus projections of the momenta are defined by $`P^+\frac{1}{\sqrt{2}}(P^0+P^3)`$. $`z`$ is the plus momentum fraction of the quark carried by the produced $`\mathrm{\Lambda }`$. Translating the matrix elements, using the integral representation of the delta function and projecting out the light-cone plus and helicity $`\pm `$ components we obtain $$\frac{1}{z}D_{q\mathrm{\Lambda }}^\pm (z)=\frac{1}{2\sqrt{2}}\underset{n}{}\delta [(1/z1)P^+p_n^+]|0|\psi _+^\pm (0)|\mathrm{\Lambda }(PS_{});n(p_n)|^2.$$ (2) Here, $`\psi _+^\pm =\frac{1}{2}\gamma _{}\gamma _+\frac{1}{2}(1\pm \gamma _5)\psi `$, and we have defined $`\gamma _\pm =\frac{1}{\sqrt{2}}(\gamma _0\pm \gamma _3)`$. The fragmentation function of an anti-quark into a $`\mathrm{\Lambda }`$ is given by Eq. (2), with $`\psi _+`$ replaced by $`\psi _+^{}`$: $$\frac{1}{z}D_{\overline{q}\mathrm{\Lambda }}^\pm (z)=\frac{1}{2\sqrt{2}}\underset{n}{}\delta [(1/z1)P^+p_n^+]|0|\psi _+^\pm (0)|\mathrm{\Lambda }(PS_{});n(p_n)|^2.$$ (3) $`D_{q\mathrm{\Lambda }}^\pm `$ can be interpreted as the probability that a quark with positive/negative helicity fragments into a $`\mathrm{\Lambda }`$ with positive helicity and similarly for antiquarks. The operator $`\psi _+`$ ($`\psi _+^{}`$) either destroys a quark (an antiquark) or it creates an antiquark (quark) when acting on the $`\mathrm{\Lambda }`$ on the right hand side in the matrix elements. Thus, whereas, in the case of quark fragmentation, the intermediate state can be either an anti-diquark state, $`\overline{q}\overline{q}`$, or a four-quark-antiquark state, $`q\overline{q}\overline{q}\overline{q}`$, in the case of antiquark fragmentation, only four-antiquark states, $`\overline{q}\overline{q}\overline{q}\overline{q}`$, are possible assuming that there are no antiquarks in the $`\mathrm{\Lambda }`$. (Production of $`\mathrm{\Lambda }`$’s through coupling to higher Fock states of the $`\mathrm{\Lambda }`$ is more complicated and involves higher number of quarks in the intermediate states. As a result it would lead to contributions at lower $`z`$ values.) Thus, we have * (1a) $`qqqq+\overline{q}\overline{q}=\mathrm{\Lambda }+\overline{q}\overline{q}`$ * (1b) $`qqqq+q\overline{q}\overline{q}\overline{q}=\mathrm{\Lambda }+q\overline{q}\overline{q}\overline{q}`$, for the quark fragmentation and * (2) $`\overline{q}qqq+\overline{q}\overline{q}\overline{q}\overline{q}=\mathrm{\Lambda }+\overline{q}\overline{q}\overline{q}\overline{q}`$, for the antiquark fragmentation. While, in case (1a), the initial fragmenting quark is contained in the produced Lambda, in case (1b) and (2), the Lambda is mainly produced by quarks created in the fragmentation process. Therefore, we not only expect that Lambdas produced through (1a) usually have larger momenta than those produced through (1b) or (2) but also that Lambdas produced through (1a) are much more sensitive to the flavor spin quantum numbers of the fragmenting quark than those produced through (1b) and (2). In the following we assume that (1b) and (2) lead to approximately the same fragmentation functions. In this case, the difference, $`D_{q\mathrm{\Lambda }}D_{\overline{q}\mathrm{\Lambda }}`$, responsible for leading particle production, is given by the fragmentation functions associated with process (1a). Similar observations also follow from energy-momentum conservation built in Eqs. (2) and (3). The delta function implies that the function, $`D_q(z)/z`$, peaks at $$z_{max}\frac{M}{M+M_n}.$$ (4) Here, $`M`$ and $`M_n`$ are the mass of the produced particle and the produced system, $`n`$, and we work in the rest frame of the produced particle. We see that the location of the maxima of the fragmentation function depends on the mass of the system $`n`$. While the high $`z`$ region is dominated by the fragmentation of a quark into the final particle and a small mass system, large mass systems contribute to the fragmentation at lower $`z`$ values. The maxima of the fragmentation functions from process (1a) are given by the mass of the intermediate diquark state and that of the the fragmentation functions from the processes (1b) and (2b) by the masses of intermediate four quark states. Thus, the contribution from process (1a) is harder than those from (1b) and (2). Energy-momentum conservation also requires that the fragmentation functions are not flavor symmetric. While the assertion of isospin symmetry, $`D_{u\mathrm{\Lambda }}=D_{d\mathrm{\Lambda }}`$, is well justified, SU(3) flavor symmetry is broken not only by the strange quark mass but also by the hyperfine interaction. Let us discuss the fragmentation of a $`u`$ (or $`d`$) quark and that of an $`s`$ quark into a Lambda through process (1a). While the intermediate diquark state is always a scalar in the strange quark fragmentation, it can be either a vector or a scalar diquark in the fragmentation of the non-strange quarks. The masses of the scalar and vector non-strange diquarks follow from the mass difference between the nucleon and the Delta , while those of the scalar and vector diquark containing a strange quark can be deduced from the mass difference between $`\mathrm{\Sigma }`$ and $`\mathrm{\Lambda }`$ . They are roughly $`m_s650`$ MeV and $`m_v850`$ MeV for the scalar and vector non-strange diquarks, and $`m_s^{}890`$ MeV and $`m_v^{}1010`$ MeV for scalar and vector diquarks containing strange quarks, respectively . According to Eq. (4), these numbers lead to soft up and down quark fragmentation functions and to hard strange quark fragmentation functions. Energy-momentum conservation, together with the splitting of vector and scalar diquark masses, has the further important consequence that polarized non-strange quarks can transfer polarization to the final state Lambda. To see this we note that the probabilities for the intermediate state to be a scalar or vector diquark state in the fragmentation of an up or down quark with parallel or anti-parallel spin to the spin of the Lambda can be obtained from the $`SU(6)`$ wave function of the $`\mathrm{\Lambda }`$ $`\mathrm{\Lambda }^{}`$ $`=`$ $`{\displaystyle \frac{1}{2\sqrt{3}}}[2s^{}(ud)_{0,0}+\sqrt{2}d^{}(us)_{1,1}d^{}(us)_{1,0}+d^{}(us)_{0,0}+`$ (6) $`\sqrt{2}u^{}(ds)_{1,1}+u^{}(ds)_{1,0}u^{}(ds)_{0,0}].`$ While the $`u`$ or $`d`$ quarks with spin anti-parallel to the spin of the $`\mathrm{\Lambda }`$ are always associated with a vector diquark, $`u`$ and $`d`$ quarks with parallel spin have equal probabilities to be accompanied by a vector or scalar diquark. The fragmentation functions of non-strange quarks with spin parallel to the $`\mathrm{\Lambda }`$ spin are harder than the corresponding fragmentation functions with anti-parallel spins. Thus, $`\mathrm{\Delta }D_{u\mathrm{\Lambda }}`$ is positive for large $`z`$ values and negative for small $`z`$. Their total contribution to polarized Lambda production might be zero or very small. Nevertheless, $`\mathrm{\Delta }D_{u\mathrm{\Lambda }}`$ and $`\mathrm{\Delta }D_{d\mathrm{\Lambda }}`$ can be sizable for large $`z`$ values, since both $`D_{u\mathrm{\Lambda }}`$ and $`\mathrm{\Delta }D_{u\mathrm{\Lambda }}`$ are dominated by the spin-zero component in the large $`z`$ limit. Furthermore, they will dominate polarized Lambda production whenever the production from strange quarks is suppressed. The matrix elements can be calculated using model wave functions at the scale relevant to the specific model and the resulting fragmentation functions can be evolved to a higher scale to compare them to experiments. In a previous paper , we calculated the fragmentation functions in the MIT bag model and showed that the resulting fragmentation functions give a very reasonable description of the data in $`e^+e^{}`$ annihilation. Since the mass of the intermediate states containing more than two quarks are not known we only calculate the contributions of the diquark intermediate states in the bag model. The other contributions have been determined by performing a global fit to the $`e^+e^{}`$ data. For this, we used the simple functional form $$D_{\overline{q}\mathrm{\Lambda }}(z)=N_{\overline{q}}z^\alpha (1z)^\beta $$ (7) to parameterize $`D_{\overline{q}\mathrm{\Lambda }}=D_{\overline{u}\mathrm{\Lambda }}=D_{\overline{d}\mathrm{\Lambda }}=\mathrm{}D_{\overline{b}\mathrm{\Lambda }}`$ and also set $`D_{g\mathrm{\Lambda }}=0`$ at the initial scale, $`\mu =0.25`$ GeV. The fragmentation functions have to be evolved to the scale of the experiment, $`\mu `$. The evolution of the non-singlet fragmentation functions in LO is given by $$\frac{d}{d\mathrm{ln}\mu ^2}[D_{q\mathrm{\Lambda }}D_{\overline{q}\mathrm{\Lambda }}](z,\mu ^2)=_z^1\frac{dz}{z^{}}P_{qq}(\frac{z}{z^{}})[D_{q\mathrm{\Lambda }}D_{\overline{q}\mathrm{\Lambda }}](z^{},\mu ^2).$$ (8) The singlet evolution equations are $`{\displaystyle \frac{d}{d\mathrm{ln}\mu ^2}}{\displaystyle \underset{q}{}}D_{q\mathrm{\Lambda }}(z,\mu ^2)`$ $`=`$ $`{\displaystyle _z^1}{\displaystyle \frac{dz^{}}{z^{}}}[P_{qq}({\displaystyle \frac{z}{z^{}}}){\displaystyle \underset{q}{}}D_{q\mathrm{\Lambda }}(z,\mu ^2)+2n_fP_{gq}({\displaystyle \frac{z}{z^{}}})D_{g\mathrm{\Lambda }}(z^{},\mu ^2)]`$ (9) $`{\displaystyle \frac{d}{d\mathrm{ln}\mu ^2}}D_{g\mathrm{\Lambda }}(z,\mu ^2)`$ $`=`$ $`{\displaystyle _z^1}{\displaystyle \frac{dz^{}}{z^{}}}[P_{qg}({\displaystyle \frac{z}{z^{}}}){\displaystyle \underset{q}{}}D_{q\mathrm{\Lambda }}(z,\mu ^2)+P_{gg}({\displaystyle \frac{z}{z^{}}})D_{g\mathrm{\Lambda }}(z^{},\mu ^2)],`$ (10) where the splitting functions are the same as those for the evolution of quark distributions. $`n_f`$ is the number of flavors. We used the evolution package of Ref. suitably modified for the evolution of fragmentation functions (interchanging the off-diagonal elements in the singlet case). The results of the LO fit to $`e^+e^{}`$ data are shown in Fig. 1. The parameters of our fits are given in Table I. The bag model calculations for $`D_{q\mathrm{\Lambda }}D_{\overline{q}\mathrm{\Lambda }}`$ and $`\mathrm{\Delta }D_{s\mathrm{\Lambda }}\mathrm{\Delta }D_{\overline{s}\mathrm{\Lambda }}`$ were parametrized using the functional form of Eq. (7). The fragmentation functions, $`\mathrm{\Delta }D_{u\mathrm{\Lambda }}\mathrm{\Delta }D_{\overline{u}\mathrm{\Lambda }}=\mathrm{\Delta }D_{d\mathrm{\Lambda }}\mathrm{\Delta }D_{\overline{d}\mathrm{\Lambda }}`$, change sign at some value of $`z`$, hence we parametrized them using the form $$\mathrm{\Delta }D_{q\mathrm{\Lambda }}(z)\mathrm{\Delta }D_{\overline{q}\mathrm{\Lambda }}(z)=N_{\overline{q}}z^\alpha (1z)^\beta (\gamma z).$$ (11) These parameters are also given in Table I. We also performed a fit using flavor symmetric fragmentation functions which we shall need for the discussion of Lambda production in pp collisions. The fit parameters are given in Table. II. In Fig. 2, we show the calculated fragmentation functions at $`Q^2=M_Z^2`$. We note that our fragmentation functions describe the asymmetry in leading and non-leading particle production, as well as the lambda polarization measured in $`e^+e^{}`$ annihilation at the $`Z`$ pole, very well — as has been shown in Ref. . ## III Polarized proton proton collision Spin-dependent fragmentation of quarks can be studied in proton-proton collisions with one of the protons polarized . Here, many subprocesses may lead to the final state Lambda so that one has to select certain kinematic regions to suppress the unwanted contributions. In particular, in order to test whether polarized up and down quarks do fragment into polarized Lambdas the rapidity of the produced Lambda has to be large, since at high rapidity, $`\mathrm{\Lambda }`$’s are mainly produced through valence up and down quarks. (We count positive rapidity in the direction of the polarized proton beam.) The difference of the cross sections to produce a Lambda with positive helicity through the scattering of a proton with positive/negative helicity on an unpolarized proton is given in leading order perturbative QCD (LO pQCD) by <sup>*</sup><sup>*</sup>*Since the relevant spin dependent cross sections on the parton level are only known in LO we perform a LO calculation here. $`E_C{\displaystyle \frac{\mathrm{\Delta }d\sigma }{d^3p_C}}(ABC+X)`$ $`=`$ $`E_C{\displaystyle \frac{d\sigma }{d^3p_C}}(A^{}BC^{}+X)E_C{\displaystyle \frac{d\sigma }{d^3p_C}}(A^{}BC^{}+X)`$ (12) $`=`$ $`{\displaystyle \underset{abcd}{}}{\displaystyle 𝑑x_a𝑑x_b𝑑z_c\mathrm{\Delta }f_{Aa}(x_a,\mu ^2)f_{Bb}(x_b,\mu ^2)\mathrm{\Delta }D_{cC}(z_c,\mu ^2)}`$ (14) $`{\displaystyle \frac{\widehat{s}}{\pi z_c^2}}{\displaystyle \frac{\mathrm{\Delta }d\sigma }{d\widehat{t}}}(abcd)\delta (\widehat{s}+\widehat{t}+\widehat{u}).`$ Here, $`\mathrm{\Delta }f_{Aa}(x_a,\mu ^2)`$ and $`f_{Bb}(x_b,\mu ^2)`$ are the polarized and unpolarized distribution functions of partons $`a`$ and $`b`$ in protons $`A`$ and $`B`$, respectively, at the scale $`\mu `$. $`x_a`$ and $`x_b`$ are the corresponding momentum fractions carried by partons $`a`$ and $`b`$. $`\mathrm{\Delta }D_{cC}(z_c,\mu ^2)`$ is the polarized fragmentation function of parton $`c`$ into baryon $`C`$, in our case $`C=\mathrm{\Lambda }`$. $`z_c`$ is the momentum fraction of parton $`c`$ carried by the produced Lambda. $`\mathrm{\Delta }d\sigma /d\widehat{t}`$ is the difference of the cross sections at the parton level between the two processes $`a^{}+bc^{}+d`$ and $`a^{}+bc^{}+d`$. The unpolarized cross section is given by Eq. (14) with the $`\mathrm{\Delta }`$’s dropped throughout. The Mandelstam variables at the parton level are given by $$\widehat{s}=x_ax_bs,\widehat{t}=x_ap_{}\sqrt{s}e^y/z_c,\widehat{u}=x_bp_{}\sqrt{s}e^y/z_c$$ (15) where, $`y`$ and $`p_{}`$ are the rapidity and transverse momentum of the produced Lambda and $`\sqrt{s}`$ is the total center of mass energy. The summation in Eq.(14) runs over all possible parton-parton combinations, $`qq^{}qq^{}`$, $`qgqg`$, $`q\overline{q}q\overline{q}`$… The elementary unpolarized and polarized cross sections can be found in Refs. . Performing the integration in Eq. (14) over $`z_c`$ one obtains $`E_C{\displaystyle \frac{\mathrm{\Delta }d\sigma }{d^3p_C}}(ABC+X)`$ $`=`$ $`{\displaystyle \underset{abcd}{}}{\displaystyle _{x_{amin}}^1}𝑑x_a{\displaystyle _{x_{bmin}}^1}𝑑x_b\mathrm{\Delta }f_{Aa}(x_a,\mu ^2)f_{Bb}(x_b,\mu ^2)\mathrm{\Delta }D_{cC}(z_c,\mu ^2)`$ (17) $`{\displaystyle \frac{1}{\pi z_c}}{\displaystyle \frac{\mathrm{\Delta }d\sigma }{d\widehat{t}}}(abcd)`$ with $$z_c=\frac{x_{}}{2x_b}e^y+\frac{x_{}}{2x_a}e^y,x_{bmin}=\frac{x_ax_{}e^y}{2x_ax_{}e^y},x_{amin}=\frac{x_{}e^y}{2x_{}e^y}$$ (18) where $`x_{}=2p_{}/\sqrt{s}`$. In order to elucidate the kinematics, in Fig. 3, we plotted $`z_c`$ as a function of $`x_a`$ and $`y`$ for two different transverse momenta, $`p_{}=10`$ GeV (left) and $`p_{}=30`$ GeV (right) and for two different values of $`x_b`$, $`x_b=x_{bmin}+0.01`$ (top) and and $`x_b=x_{bmin}+0.1`$ (bottom) in Fig. 3. Note, that $`z_c`$ is maximal both for $`x_b=x_{bmin}`$ and $`x_a=x_{amin}`$. With increasing rapidity, $`y`$, both the lower integration limit of $`x_a`$, $`x_{amin}`$, and the momentum fraction of the fragmenting quark transferred to the produced $`\mathrm{\Lambda }`$, $`z_c`$, increase. Hence, large rapidities probe the fragmentation of mostly valence quarks into fast Lambdas. The dependence on the transverse momenta is also shown in Fig. 3. With increasing $`p_{}`$, the kinematic boundary is shifted to smaller rapidities and the fragmentation of the valence quarks can be studied at lower rapidities. This is important since the available phase space is limited by the acceptance of the detectors at RHIC. However, the cross section also decreases with increasing transverse momenta, leading to lower statistics. In Fig. 4, we show the contributions of the various channels to the cross section for two different transverse momenta, both for inclusive Lambda (4a) and inclusive jet production (4b). $`gqgq`$ stands, for example, for the contribution to the cross section coming from the subprocess involving a gluon $`g`$ and a quark $`q`$ in the initial and final states. In $`qqqq`$, the quarks have different flavors and $`q\overline{q}q\overline{q}`$ is also included. Although the kinematics are not exactly the same for these two processes While there is only one integration variable, $`x_a`$, in inclusive jet production, once $`p_{}`$ and $`y`$ are fixed, both $`x_a`$ and $`x_b`$ have to be integrated over the allowed kinematic region in inclusive Lambda production, since the produced $`\mathrm{\Lambda }`$ carries only a fraction of the parton’s momentum. one can study the role played by the fragmentation functions by comparing inclusive Lambda and inclusive jet production. In particular, the contributions from channels containing two gluons in the final state are suppressed in inclusive Lambda production due to the smallness of $`D_{g\mathrm{\Lambda }}`$. We note that $`D_{g\mathrm{\Lambda }}`$ has been set to zero at the initial scale and is generated through evolution. Thus, while $`qgqg`$ is the dominant channel in inclusive Lambda production, both $`qgqg`$ and $`gggg`$ are equally important in inclusive jet production. There is some ambiguity due to our poor knowledge of the gluon fragmentation — larger probabilities for $`g\mathrm{\Lambda }`$ will enhance the contributions from gluons in inclusive Lambda production. However, the contribution from the process $`gggg`$ falls off faster than that from $`qgqg`$ with increasing rapidity, since $`g(x_a)`$ decreases faster than $`q(x_a)`$ with increasing $`x_a`$ and the Lambdas are produced mainly from valence up and down quarks at high rapidities. Our analysis of the kinematics and the various contributions to inclusive Lambda production already indicate that Lambda polarization measurements in pp collisions at high rapidities are ideally suited to test whether polarized up and down quarks may fragment into polarized $`\mathrm{\Lambda }`$’s. We calculated the Lambda polarization using our flavor asymmetric fragmentation functions for RHIC energies and for $`p_{}=10`$ GeV. Increasing the transverse momentum gives similar results, with the only difference that $`P_\mathrm{\Lambda }`$ starts to increase at lower rapidities. We used the standard set of GRSV LO quark distributions for the polarized parton distributions and the LO Cteq4 distributions for the unpolarized quark distributions . The scale, $`\mu `$, is set equal to $`p_{}`$. We also checked that there is only a very weak dependence on the scale by calculating the polarization using $`\mu =p_{}/2`$ and $`\mu =2p_{}`$. The predicted Lambda polarization is shown in Fig 5a. It is positive at large rapidities where the contributions of polarized up and down quarks dominates the production process. At smaller rapidities, where $`x_a`$ is small, strange quarks also contribute. However, since the ratios of the polarized to the unpolarized parton distributions are small at small $`x_a`$ the Lambda polarization is suppressed. The result also depends on the parameterization of the polarized quark distributions. In particular, the polarized gluon distribution is not well constrained. However, it is clear from the kinematics that the ambiguity associated with the polarized gluon distributions only effects the results at lower rapidities. This can be seen in Fig. 5b where we plot the contribution from gluons, up plus down quarks and strange quarks to the Lambda polarization. Next, we contrast our prediction with the predictions of various $`SU(3)`$ flavor symmetric models which use $$D_{u\mathrm{\Lambda }}=D_{d\mathrm{\Lambda }}=D_{s\mathrm{\Lambda }}.$$ (19) We fitted the cross sections in $`e^+e^{}`$ annihilation using Eq. (19) and the functional form given in Eq. (7). For the polarized fragmentation functions, we discuss two different scenarios: The model, $`SU(3)_A`$ (c.f. Fig. 5a), corresponds to the expectations of the naive quark model that only polarized strange quarks can fragment into polarized Lambdas $$\mathrm{\Delta }D_{u\mathrm{\Lambda }}=\mathrm{\Delta }D_{d\mathrm{\Lambda }}=0\mathrm{\Delta }D_{s\mathrm{\Lambda }}=D_{s\mathrm{\Lambda }}.$$ (20) It gives essentially zero polarization because the strange quarks contribute at low rapidities where the polarization is suppressed. Model, $`SU(3)_B`$ (c.f. Fig. 5a), which was proposed in Ref., is based on DIS data, and sets $$\mathrm{\Delta }D_{u\mathrm{\Lambda }}=\mathrm{\Delta }D_{d\mathrm{\Lambda }}=0.20D_{u\mathrm{\Lambda }}\mathrm{\Delta }D_{s\mathrm{\Lambda }}=0.60D_{s\mathrm{\Lambda }}.$$ (21) This model predicts negative Lambda polarization. Finally, we address the problem of Lambdas produced through the decay of other hyperons, such as $`\mathrm{\Sigma }^0`$ and $`\mathrm{\Sigma }^{}`$. In order to estimate the contribution of hyperon decays we assume, in the following, that (1) the $`\mathrm{\Lambda }`$’s produced through hyperon decay inherit the momentum of the parent hyperon (2) and that the total probability to produce $`\mathrm{\Lambda }`$, $`\mathrm{\Sigma }^0`$ or $`\mathrm{\Sigma }^{}`$ from a certain $`uds`$ state is given by the $`SU(6)`$ wave function and is independent of the mass of the produced hyperon. Further, in order to estimate the polarization transfer in the decay process we use the constituent quark model. The polarization can be obtained by noting that the boson emitted in both the $`\mathrm{\Sigma }^0\mathrm{\Lambda }\gamma `$ and the $`\mathrm{\Sigma }^{}\mathrm{\Lambda }\pi `$ decay changes the angular momentum of the nonstrange diquark from $`J=1`$ to $`J=0`$, while the polarization of the spectator strange quark is unchanged. Then, the polarization of the $`\mathrm{\Lambda }`$ is determined by the polarization of the strange quark in the parent hyperon, since the polarization of the $`\mathrm{\Lambda }`$ is exclusively carried by the strange quark in the naive quark model. First, let us discuss the case when the parent hyperon is produced by a strange quark. Since the strange quark is always accompanied by a vector $`ud`$ diquark, in both $`\mathrm{\Sigma }^0`$ and $`\mathrm{\Sigma }^{}`$ the fragmenation functions of strange quarks into these hyperons are much softer than the corresponding fragmentation function into a $`\mathrm{\Lambda }`$. Thus, in the high $`z`$ limit, the contributions from the processes, $`s\mathrm{\Sigma }^0\mathrm{\Lambda }`$ and $`s\mathrm{\Sigma }^{}\mathrm{\Lambda }`$, are negligible compared to the direct production, $`s\mathrm{\Lambda }`$. Furthermore, both channels, $`s\mathrm{\Sigma }^0\mathrm{\Lambda }`$ and $`s\mathrm{\Sigma }^{}\mathrm{\Lambda }`$, enhance the already positive polarization from the direct channel, $`s\mathrm{\Lambda }`$. This is different in the case when the parent hyperon is produced by an up or down quark. Both $`\mathrm{\Lambda }`$ and $`\mathrm{\Sigma }^0`$ can be produced by an up (down) quark and a scalar $`ds`$ ($`us`$) diquark — a process which dominates in the large $`z`$ limit. (The component with a vector diquark can be neglected in this limit). Furthermore, the up and down fragmentation function of the $`\mathrm{\Sigma }^{}`$ are as important as those of the $`\mathrm{\Lambda }`$ and $`\mathrm{\Sigma }^0`$ in the large $`z`$ limit. This is because the $`u`$ fragmentation function of $`\mathrm{\Sigma }^{}`$ peaks at about $`1385/(1010+1385)0.58`$ which is almost the same as the peak of the scalar components of the $`\mathrm{\Lambda }`$ and $`\mathrm{\Sigma }`$, which are $`1115/(890+1115)0.57`$ and $`1190/(890+1190)0.57`$, respectively. Thus, for the up and down quark fragmentation, it is important to include the $`\mathrm{\Lambda }`$’s from these decay processes. The relevant probabilities to produce a $`\mathrm{\Lambda }`$ with positive and negative polarization from a fragmenting up quark with positive polarization and an $`ds`$ diquark are shown in Table III. We assumed that all spin states of the $`ds`$ diquark are produced with equal probabilities. The final weights which are relevant in the large $`z`$ limit are set in bold. We find that if we include all channels, which survive in the large $`z`$ limit, the polarization of the $`\mathrm{\Lambda }`$ is reduced by a factor of $`10/27`$ compared to the case where only the directly produced $`\mathrm{\Lambda }`$’s are included. Since the $`\mathrm{\Sigma }^{}`$ decay is a strong decay it is sometimes included in the fragmentation function of the $`\mathrm{\Lambda }`$. Including only $`\mathrm{\Sigma }^{}`$, the suppression factor we obtain is $`49/81`$. (Note that our model predicts that $`u^{}(ds)_{0,0}\mathrm{\Lambda }`$, $`u^{}(ds)_{0,0}\mathrm{\Sigma }^0`$ and $`u^{}(ds)_{1,0}\mathrm{\Sigma }^{}`$ have approximately the same $`z`$ dependence and are approximately equal (up to the Clebsch-Gordon factors) since the ratios, $`M/(M+M_n)`$, have roughly the same numerical values. Thus, the effect of the $`\mathrm{\Sigma }^0`$ and $`\mathrm{\Sigma }^{}`$ decays can be taken into account by a multiplicative factor.) In order to illustrate the effect of these decays on the final $`\mathrm{\Lambda }`$ polarization, we multiplied our results with these factors. The results are shown in Fig. 6b as dotted lines. We note that our implementation of this correction relies on the assumptions that the produced $`\mathrm{\Lambda }`$ carries all the momentum of the parent hyperon and that all states are produced with equal probabilities. Since neither of these assumptions is strictly valid, we tend to overestimate the importance of hyperon decays. Note also that the inclusion of $`\mathrm{\Sigma }^0`$ decay in the $`SU(3)`$ symmetric models makes the resulting polarization more negative. As a result, even if effects of $`\mathrm{\Sigma }`$ decays are included, large discrepancies still persist between our predictions and those of $`SU(3)`$ symmetric models. ## IV Conclusions Measurements of the Lambda polarization at RHIC would provide a clear answer to the question of whether polarized up and down quarks can transfer polarization to the final state $`\mathrm{\Lambda }`$. We predict positive Lambda polarization at high rapidities, in contrast with models based on $`SU(3)`$ flavor symmetry and DIS which predict zero or negative Lambda polarization. Our prediction is based on the same physics which led to harder up than down quark distributions in the proton and to the $`\mathrm{\Delta }`$-N and $`\mathrm{\Sigma }`$-$`\mathrm{\Lambda }`$ mass splittings. We also estimated the importance of $`\mathrm{\Sigma }^0`$ and $`\mathrm{\Sigma }^{}`$ decays which tend to reduce the predicted $`\mathrm{\Lambda }`$ polarization. ###### Acknowledgements. This work was partly supported by the Australian Research Council. One of the authors \[JTL\] was supported in part by National Science Foundation research contract PHY-9722706. One author \[CB\] wishes to thank the Indiana University Nuclear Theory Center for its hospitality during the time part of this work was carried out.
warning/0002/astro-ph0002368.html
ar5iv
text
# Quasar variability: correlations with amplitude ## 1 Introduction One of the most important constraints on the structure of quasars is variability. Short term variations set limits on the size of the emitting region, and differences in the nature of the variation in the Xray, optical and radio domains give clues to the underlying structure. Variations on longer timescales of a few years are not so easily accounted for by the current black hole paradigm, and microlensing has been put forward as an alternative explanation for the observations. Although the most extensive monitoring of quasar variability has been done in optical wavebands, there is still little consensus even over the broad picture. One of the first problems is to parametrise the variation in such a way that it can be compared with models, or even more general expectations. The timescale of variation and the amplitude are the two parameters which have been mostly studied, although some authors have succeeded in confusing the two. This typically involves claiming that in a short run of data, objects varying on a short timescale will achieve a larger amplitude than those varying on a longer timescale, and so amplitude can be taken as an (inverse) measure of timescale (Hook et al. hoo94 (1994)). In this paper we shall confine our attention to the correlation of amplitude with other parameters. Amplitude is an easy parameter to measure, and involves none of the difficulties inherent in estimating timescale of variation. There are modes of variability in which some uncertainty will be introduced by the length of the run of observations, but this is a problem which can arise in any time series analysis. Most attention has so far been given to possible correlations of amplitude with redshift or luminosity (eg Trevese et al. tre89 (1989), Giallongo et al. gia91 (1991), Hook et al. hoo94 (1994), Cristiani et al. cri96 (1996)), but even in such straightforward situations there has been little agreement. The reasons for the lack of consensus are not easy to understand, but it is clear that any effect which is present is not large compared with the cosmic scatter in the data, which in the case of amplitude appears to be about one magnitude, much larger than any photometric errors. Perhaps the most likely cause of disagreement is in the selection of samples for analysis. In order to obtain a meaningful measure of correlation it is essential that unbiassed samples are used, and that they cover a large range of redshift and luminosity. In this paper particular attention is paid to the sample selection process, and by extending the analysis to high redshifts the baseline for the measurement of correlations is greatly extended over earlier samples. ## 2 Quasar samples ### 2.1 Observational material The parent sample of quasars for the analysis of variability amplitude comes from a large scale survey and monitoring programme being undertaken in the ESO/SERC field 287, at 21h 28m, -45. Quasars have been selected according to a number of criteria, including colour, variability, radio emission and objective prism spectra, or a combination of these techniques. Redshifts for over 600 quasars have so far been obtained, and several complete samples defined within specific limits of magnitude, redshift and position on the sky (Hawkins & Véron haw95 (1995)). A detailed description of the survey is given by Hawkins & Véron (haw95 (1995)) and Hawkins (haw96 (1996)). Briefly, a large set of UK 1.2 Schmidt plates spanning 20 years was scanned with the COSMOS and SuperCOSMOS measuring machines to provide a catologue of some 200,000 objects in the central 19 square degrees of the plate. These were calibrated with CCD frames to provide light curves in $`B_J`$ and $`R`$, and colours in $`UBVRI`$. Photometric errors have been discussed in detail in earlier papers (Hawkins 1996 and references therein) and are $`0.08`$ magnitudes for any individual machine measurement, and only weakly dependent on magnitude. There are approximately four plates in each year which reduces the error to $`0.04`$ magnitudes per epoch. This is small compared with the amplitudes of interest in this paper, and no attempt has been made to deconvolve it. ### 2.2 Sample selection For the analysis in this paper three samples will be defined. The first (UVX) is based only on position on the sky and ultraviolet excess. The area on the sky containing the sample is defined by a number of AAT AUTOFIB fields and a 2dF field (Folkes et al. fol99 (1999) and references therein), covering a total area of 7.0 square degrees. Within this area all objects with $`UB<0.2`$ and $`B_J<21.0`$ were observed. This was extended to $`B_J<21.5`$ in the 2dF field. The ultra-violet excess (UVX) cut although necessary to give a relatively clean sample of quasar candidates, has the well known limitation of only being effective for redshifts $`z<2.2`$. Beyond this redshift quasars become red in $`UB`$ as the Lyman forest enters the $`U`$ band. For samples at higher redshift, variability has been found to provide a very useful criterion for quasar selection (Hawkins haw96 (1996)). The second sample for consideration herei (VAR), was selected on this basis, with the requirement that the object should lie in the same area of sky as for the first sample with a magnitude limit $`B_J<21`$, or anywhere in the measured area of the plate with a magnitude limit $`B_J<19.5`$, and should have an amplitude $`\delta m>0.35`$. The defining epoch for the magnitudes was the year 1977. This sample was selected without any reference to colour and so can be used to measure trends over a large range of redshift ($`z<3.5`$). The third sample (AMP) was selected in a similar manner, but over the whole measured area of the plate (19 square degrees), and with an amplitude cut $`\delta m>1.1`$. There is some evidence that these large amplitude objects form a distinct group, which is discussed below. The development of fibre fed spectrographs has meant that every object included within a given set of search criteria can be observed to give a high completeness level. In the case of fields observed with AUTOFIB, the existence of forbidden regions in the 40 arcminute field, and the very variable throughput for different fibres aligned at different positions on the spectrograph slit meant that up to 20% of spectra did not have sufficient signal to provide an unambiguous redshift. These objects where re-observed later with the faint object spectrograph EFOSC on the ESO 3.6m telescope at La Silla. The 2dF observations where of very uniform quality, and the redshift measures or other classification were almost 100% successful from the fibre-feed spectra. Details of all three samples, containing a total of 384 quasars, are given in the Appendix. $`B`$ magnitudes are in the $`B_J`$ system defined by the IIIa-J emulsion and the GG395 filter, and refer to the year 1977. The amplitude $`\delta m`$ is the difference between maximum and minmum magnitude achieved over the 20 year run of data. The samples which each quasar belongs to are indicated by 1, 2 and 3 corresponding to VAR, UVX and AMP respectively, as defined above. Data for many of the quasars have already been published by Hawkins & Véron (haw95 (1995)), but are given again here for completeness. Any small differences in the parameters are the result of further refinement of the calibration, and more extended monitoring of the light curves. UVX selected samples have been used many times in the past for quasar surveys, and the constraints are quite well known. Variability selection has been less often used, and some additional comments are appropriate. In particular, there is the important question of completeness. Hawkins & Véron (haw95 (1995)) use a small sample of quasars in field 287 from Morris et al. (mor91 (1991)) selected by objective prism to test completeness. In 1993, 79% of the objective prism quasars would have also been detected according to the variability criterion $`\delta m>0.35`$; by 1997 this figure had risen to 93%, with two quasars remaining below the variability threshold. In fact both of these quasars are clearly variable, with amplitudes 0.32 and 0.29 magnitudes. Strangely, they lie at either extreme of the redshift and luminosity range of the sample, with redshifts 3.23, 0.29 and luminosities -27.55, -22.33 respectively. Fig. 1(a) shows the distribution of epoch at which quasars first satisfy the detection criterion given by Hawkins (haw96 (1996)). This is generally speaking equivalent to attaining an amplitude of 0.35 magnitudes. The distribution peaks between 2 and 4 years, and most quasars have satisfied the criterion after 7 years. Fig. 1(b) shows the cumulative distribution of detection epochs, illustrating the point that after 15 years nearly all quasars have varied sufficiently to put them in the variability selected sample. The distribution of amplitudes is best shown with the UVX sample, which was selected without any reference to variability. Fig. 2(a) is a histogram of amplitudes over a 20 year baseline, and shows a median amplitude of around 0.7 magnitudes. As would be expected from Fig. 1, nearly all of the sample lies above the threshold amplitude of the variability selected sample of 0.35 mags. With this in mind it is worth investigating the possibility of using the variability selected sample to look for correlations at higher redshift. Fig. 2(b) shows a histogram of amplitudes for the VAR sample, which from the way it was selected has a cut-off at an amplitude of 0.35 mags. More interestingly, the distribution as a whole peaks at an amplitude of about 0.7, similar to the UVX sample. This peak is well clear of the cut-off, and implies that only a small fraction of quasars are missed from the VAR sample. There may however be a problem detecting the most luminous quasars, ($`M_B<27`$) for which there is evidence for relatively small variations (Cristiani et al. cri96 (1996)). Nonetheless, the sample can be used with caution to extend quasar correlations to high redshift. ## 3 Amplitude correlations The top two panels in Fig. 3 show the distribution of amplitudes as a function of redshift and luminosity for the UVX sample. The most striking thing about these figures is the marked drop in amplitude towards low redshift and luminosity. This is illustrated more clearly in the top two panels of Fig. 4, where the data is binned in intervals of 0.5 in redshift and unit absolute magnitude, and the mean plotted with $`\sqrt{N}`$ error bars. The fall towards low redshift and luminosity is highly significanti (a 3-$`\sigma `$ effect), although it is not clear from these data alone whether it is primarily a luminosity or redshift effect. The correlation coeffecient for the top left panel of Fig. 3 with $`z<0.5`$ is 0.46, and for the top right panel with $`M_B>22`$ is -0.67, confirming the reality of the correlation. There is also some evidence for a decrease in amplitude for the most luminous quasars ($`M_B<25`$), and an even weaker decline for high redshift objects. Correlation coefficients for all the data in the top two panels are 0.12 and -0.10 for left and right panels respectively, emphasising the weakness of the effect. This is another manifestation of the well known degeneracy between redshift and luminosity. This point will be investigated further below by dividing the data into sub-samples. The bottom two panels in Figs 3 and 4 show similar plots for the VAR sample. The greater redshift range allowed by variability selection still shows little trend of amplitude with redshift, but the decrease in amplitude for luminous quasars is shown to continue to greater luminosities, although the effect is inevitably lessened by the absence of quasars with $`\delta m<0.35`$. It has been mentioned above that there is a degeneracy between redshift and luminosity. This results from the fact that in a magnitude limited sample high redshift quasars tend to be high luminosity objects, and vice versa. Thus a trend with one parameter will be mimicked by a trend with the other, and the true relation will be hard to disentangle. This degeneracy can in principle be broken byi binning the data in redshift and luminosity, and the result of doing this is shown in Figs. 5 and 6. The VAR sample was used as it covers a wider range of luminosity and redshift, making binning feasible. The left hand panel shows amplitude as a function of luminosity in two redshift bins, $`z<1.5`$ and $`z>1.5`$. Data for the two redshift ranges overlap nicely, and clearly show a decrease in the relation between amplitude and luminosity. The right hand panel shows amplitude as a function of redshift in two luminosity bins. In this case there is some evidence for an increase in the relation at low redshift, but it is essentially flat beyond $`z=0.5`$ for both luminosity ranges. It is however worth noting that the less luminous quasars have larger amplitudes, as expected from the data in the left hand panel. It would thus appear that while for $`M_B>25`$ amplitude does indeed decrease with luminosity, it does not change with redshift for $`z>0.5`$. Examination of Fig. 3, especially the bottom panels, suggests that there may be population of low luminosity and/or low redshift quasars distinct from the parent population. This is particularly evident in the bottom left hand panel of Fig. 3, which is uniformly populated between amplitudes of 1.1 and 1.8 up to a redshift of 2 at which point there is a sharp cut-off with no amplitudes greater than 1.1 at higher redshift. To investigate this population with better statistics all variables with $`\delta m>1.1`$ in the measured area of field 287 were observed on the 3.6m at La Silla to confirm their identification as quasars and measure redshifts. This became the AMP sample. Fig. 7 shows amplitude as a function of both redshift and luminosity, and it will be seen that there is indeed a cut-off in redshift at $`z2`$ and $`M_B25`$. It is clear that this cut-off must in fact be related to luminosity. If it were a redshift cut-off there is no reason why such objects should not be seen with greater luminosity. On the other hand if it were a luminosity cut-off, then this combined with a magnitude limit will indeed produce an effective cut-off in redshift. ## 4 Discussion Attempts to measure correlations of amplitude (or some related variability parameter) with redshift and luminosity have a long history, which is summarised by Hawkins (haw96 (1996)). Among recent work, a useful place to start is with the paper by Hook et al. (hoo94 (1994)). They analyse a sample of $`300`$ quasars in the SGP area from 12 UK 1.2m Schmidt plates taken in 5 separate yearly epochs spanning 16 years. They find a convincing anti-correlation between their variability parameter $`\sigma _v`$ (a measure of variation about the mean) and luminosity, but a much weaker anti-correlation with redshift. They attribute this to the degeneracy between redshift and luminosity in their sample. The same sample is re-analysed by Cid Fernandes et al. (cid96 (1996)) who use a variability index related to variance, and also one related to the structure function. They concur with Hook et al. (hoo94 (1994)) that there is an anti-correlation between their variability indices and luminosity, but claim a positive correlation with redshift. This effect is not apparent to the eye, but is interpreted as a variability-wavelength dependence rather than an intrinsic variability-redshift dependence. The net result in an un-binned sample cancels out with the luminosity variability relationship. Cid Fernandes et al.’s claim for a positive correlation between amplitude and redshift appears to be motivated at least in part by expectations arising from a paper by Di Clemente et al. (dic96 (1996)). This interesting paper examines the relation between their variability parameter $`S_1`$ (an amplitude based on the structure function) and wavelength. Their sample is composed of PG quasars (Schmidt and Green sch83 (1983)), which are mostly low redshift and relatively low luminosity objects. With the help of archival IUE observations they find that $`S_1`$ decreases with wavelength. This effect can clearly be seen in the light curve from the intensive monitoring programme of the Seyfert galaxy NGC 5548 (Clavel et al. cla91 (1991)), which has a larger amplitude at shorter wavelength. Fig. 8 shows the relation between amplitude and wavelength taken from the light curves, with a best-fit quadratic curve. The relation between rest wavelength and amplitude is essentially equivalent to the relation between redshift and amplitude, where one is seeing progressively shorter wavelengths at higher redshift. This is illustrated in the top panel of Fig. 9 which shows the UVX sample with two curves superimposed. The solid line is converted from Fig. 8 and does not appear to follow the trend of the data, but the large scatter makes it hard to construct a convincing test. Nonetheless it suggests that any relation which holds for Seyfert galaxies might have to be modified for quasars. The dotted line is from Di Clemente et al. (dic96 (1996)), and shows a very small effect, which does nonetheless follow the flat distribution of the data. The decrease of amplitude for the smallest redshift and luminosity seen in the top two panels of Fig. 4 has a number of possible explanations. It could be the effect of the underlying galaxy dominating any change in nuclear brightness for low luminosity objects, it could be a consequence of the small optical depth to microlensing at low redshift, or it could be a consequence of the wavelength dependence of variability (Cristiani et al. cri97). The present dataset is not adequate to settle the question, which is best done by looking at luminous quasars at very low redshift and Seyfert galaxies at high redshift. All the plots in Figs 3 and 4 show a trend of decreasing amplitude towards higher redshift or more luminous objects. Although the trend as a function of luminosity is more marked, the old problem of degeneracy makes it hard to say for certain that it is a luminosity effect which is being observed. However, if we look at Fig. 6 where the data are binned in luminosity and redshift we see that while there is no significant trend of amplitude with redshift in either luminosity bin, there is a marked inverse correlation between amplitude and luminosity. The relation between amplitude and luminosity is in agreement with that found in earlier work (Hook et al. hoo94 (1994), Hawkins haw96 (1996), Cristiani et al. cri96 (1996)) and may well turn out to be a useful way of distinguishing between various schemes for quasar variability. The evidence for a constant amplitude with redshift is more debatable. It would appear to be consistent with the early claim of Hook et al. (hoo94 (1994)) for a weak anti-correlation with redshift which they ascribed to degeneracy with luminosity. It is also in agreement with the results of Cristiani et al. in the observer’s frame. When they correct their structure function to the quasar rest frame they inevitably imprint a positive correlation between their variability parameter and redshift. To investigate this dependence of amplitude on redshift for the present sample, Fig. 10 shows the epoch at which quasars achieve their maximum amplitude as a function of redshift. Apart from very low redshifts ($`z<0.3`$) this relation is flat, implying that at least over the 21 years of the present dataset, time dilation effects will not bias the measurement of amplitude. Since the conclusions of Cid Fernandes et al. (cid96 (1996)) are largely based on a sample for which a time dilation correction has been applied, it is not feasible to make a direct comparison with the present work. However, it appears that the main difference between their results and those of Hook et al. is in the definition of a variability parameter and the method of analysis (both papers are based on the SGP sample). There are perhaps three currently discussed schemes for quasar variability. The least well constrained is the accretion disk model, where instabilities are propagated across the disk leading to variation in light. The details of this approach have proved hard to work out, especially in the context of the constraints imposed by existing observations, but it does not seem to lead to an inverse correlation between amplitude and luminosity. An interesting recent attempt to model variation on the basis of accretion disk instabilities by Kawaguchi et al. (kaw98 (1998)) may provide a means for producing the observed variations. It does however appear to predict variations which are either too asymmetric or of too small an amplitude to be consistent with the current observations. The timescales which they predict are also rather short, around 200 days for reasonable input parameters, and much shorter than the observed timescale of a few years. An alternative approach, developed by Terlevich and his collaborators, accounts for the variation by postulating that the quasar is powered by a series of supernova explosions. Qualitatively, this model can account for the observed relation between luminosity and amplitude, and works quite well for Seyfert galaxies (Aretxaga & Terlevich are94 (1994)). However, for quasars (Aretxaga et al. are97 (1997)) large numbers of supernovae are required to achieve the luminosity, which results in smaller variations. For example, even a relatively modest quasar with absolute magnitude $`M_B26`$ would need some 300 type II supernovae per year to power it, which given typical decay times would lead to very little variation at all. This clearly conflicts with observations in this paper which show that most quasars vary by around 0.5 to 1 magnitude on a timescale of a few years. The third way of explaining quasar variability is to invoke microlensing. This approach has been explored in several recent papers (Hawkins haw93 (1993), haw96 (1996), Hawkins & Taylor haw97 (1997)), and seems to account well for a number of statistical properties of quasar light curves. It can also explain the inverse correlationi between luminosity and amplitude in a natural way. It is well known that when a point source is microlensed by a population of compact bodies of significant optical depth, the lenses combine non-linearly to form a caustic pattern which produces sharp spikes in the resulting light curves (Schneider & Weiss sch87 (1987)). As the source becomes comparable in size with, or larger than, the Einstein radius of the lenses the amplitude decreases (Refsdal & Stabell ref91 (1991)). Thus if one assumes a uniform temperature for quasar disks, and that the luminosity is determined by the disk area, the larger more luminous disks will be amplified less. Using a relation between source size (in terms of Einstein radius) and amplitude given by Refsdal & Stabell (ref91 (1991)) (eqn 1), one can thus derive a relation between amplitude $`\delta m`$ and absolute magnitude $`M`$ of the form: $$\delta m=10^{0.2(M+c)}$$ where $`c`$ is a constant. The bottom panel of Fig. 9 shows a plot of amplitude versus absolute magnitude for the UVX sample with this relation superimposed. The constant $`c`$ was adjusted to allow the curve to track the upper envelope of the points, which has the effect of defining the quasar disk size. This ranges from 1.1 Einstein radii for $`M_B=23`$ to 11 for $`M_B=28`$. The points scatter downwards from the upper envelope because the quasars do not necessarily attain their maximum possible amplitude. Although one cannot say that the curve provides a fit to the data, the trend is certainly well represented. The possibility of a distinct population of large amplitude, low luminosity quasars suggested by Fig. 7 may also be used to test the models of quasar variability. Again, an accretion disk provides no obvious mechanism for such an effect. The Christmas Tree model certainly does imply large amplitude variations for low luminosity objects, where each event is of comparable brightness to the nucleus itself. Perhaps the most natural explanation comes from microlensing, where the nucleus of low luminosity quasars would plausibly become very small compared with the Einstein radii of the lenses, resulting in large amplifications from caustic crossing events. ## 5 Conclusions In this paper we have examined the dependance of amplitude on redshift and luminosity for large samples of quasars selected on the basis of ultra-violet excess and variability. The quasars span a redshift range $`0<z<3.5`$ and a luminosity range $`20>M_B>28`$. There is evidence for a correlation of amplitude with luminosity and/or redshift for the sample as a whole, but when it is binned in redshift the correlation with luminosity becomes significant. This result could be strengthened by the possible existence of a population of large amplitude low luminosity objects in the sample. No convincing evidence is found for a correlation between amplitude and redshift, either for the sample as a whole or when it is binned in luminosity. Various models of quasar variability are examined with respect to the observed correlations. It is concluded that any straightforward interpretation of an accretion disk model is incompatible with the data. The Christmas Tree model has some merits, especially in the regime of low luminosity quasars and Seyfert galaxies, but for luminous quasars the rate of supernovae required is too large to be compatible with the observed variability. The microlensing model can be used to explain all the data, although it does require that the Einstein radius of the microlensing bodies is comparable in size to the quasar nucleus. ###### Acknowledgements. I thank Andy Lawrence and Omar Almaini for making some excellent suggestions for improvements to the paper. ## References ## Appendix
warning/0002/nlin0002025.html
ar5iv
text
# Dynamic algorithm for parameter estimation and its applications ## I introduction An experimental observation often consists of reading a time series output from a dynamical system. Such a time series can contain information about the number as well as the form of the functions governing the evolution of the system variables including nonlinearities (if any) and the parameters . The estimation of parameter values from a given chaotic scalar time series of a nonlinear system is the topic of our interest here. We have recently given a method to dynamically estimate unknown parameters from the chaotic time series of a single phase space variable when the system equations are known . The method is based on a combination of synchronization and adaptive control similar to that used by John and Amritkar . The problem of parameter estimation in nonlinear dynamics has been considered earlier. Parlitz, Junge and Kocarev have given a static method based on minimization while Parlitz has developed a method based on auto-synchronization . Unlike our method, auto-synchronization method requires an ansatz for the parameter control loop and gives slower convergences in many cases. A method requiring a vector time series is given by Baker, Gollub and Blackburn and another method based on symbolic dynamics is discussed in Refs. . The effect of noise on parameter estimation was studied by us and recently by Goodwin, Brown and Junge . In contrast to many of these methods our method in Ref. works asymptotically so that an exact estimation of the parameters is in principle possible. The static methods based on minimization are computationally expensive because they take a longer time to run due to many iterations required for convergence and they also require annealing to eliminate the possibility of getting trapped into a local minimum. The dynamic method as described in Ref. requires only one time evolution of the system equations. The method also takes care of annealing in a dynamic way. In the first part of this paper we review our method for parameter estimation in brief. We then extend it to a case when the time series of a scalar function of phase space variables is given. We then go on to study the applicability of the method to a general quadratic flow in three dimensions. This system has a large number of parameters and we try to estimate some of them using our method. In the second part, we show that it is possible to extend our method to a more realistic situation, when the given time series is truncated after a finite time. We find that a repetitive use of the finite time series can be made to estimate the unknown parameters of the underlying system without altering the dynamic nature of the method. The accuracy of such an estimation increases with the increasing length of the given time series. We also see that the accuracy saturates with the number of times the finite time series is used. Lastly in the third part of this paper, we suggest an interesting application of parameter estimation method. Consider a situation where an unknown perturbation disturbs a known chaotic system. In many practical situations when the external perturbation is unknown, an ansatz function modeling the behaviour of the external perturbation is tried. We show that it is possible to use our parameter estimation method, to confirm the form of an ansatz function modeling the external perturbation. In section IIA we briefly introduce our method of parameter estimation and discuss its important features. In section IIB we extend it to a general situation when the given time series is obtained as a scalar function of the phase space variables. Section IIC deals with a general quadratic flow in three dimensions. In section III we extend the method to the case of finite time series and present two examples. Finally in section IV we give the application of the method in confirming the form of an unknown external perturbation to a known dynamical system. In section V we conclude with a summary of the results. ## II parameter estimation ### A The method Here, we briefly introduce our method for parameter estimation from a scalar time series. We would like to direct the reader to Ref. for a more detailed discussion. We start by considering an autonomous dynamical system of the form, $`\dot{𝐱}=𝐟(𝐱,\alpha ),`$ (1) where $`𝐱=(x_1,x_2,\mathrm{},x_n)`$ is an $`n`$-dimensional state vector whose evolution is described by the function $`𝐟=(f_1,\mathrm{},f_n)`$. We denote a set of $`m`$ unknown scalar parameters by $`\alpha =(\alpha _1,\alpha _2,\mathrm{},\alpha _m)`$. A possible appearance of any other parameters (assumed to be known) is not shown in Eq. (1). Without loss of generality we assume that a time series of the variable $`x_1`$ is given. The problem we consider is to estimate $`\alpha `$ from the given scalar time series of $`x_1`$ assuming the functional form of f to be known. In analogy with the control method used earlier by John and Amritkar , we combine synchronization with adaptive control to achieve our goal of estimating $`\alpha `$ in Eq. (1) as follows. We construct another system of variables $`𝐱^{}`$ having a structure identical to that of Eq. (1) with a linear feedback proportional to the difference $`x_1^{}x_1`$ added in the evolution of the variable $`x_1`$. Thus the system is given by, $`\dot{x}_1^{}`$ $`=`$ $`f_1(𝐱^{},\alpha ^{})ϵ(x_1^{}x_1)`$ (2) $`\dot{x}_j^{}`$ $`=`$ $`f_j(𝐱^{},\alpha ^{}),j=2,\mathrm{},n.`$ (3) where the function $`𝐟=(f_1,\mathrm{},f_n)`$ is the same as that in Eq. (1). The initial values of parameters $`\alpha ^{}`$ which correspond to the unknown parameters $`\alpha `$ in Eq. (1) are chosen randomly. The newly introduced parameter $`ϵ`$ is the feedback constant. It is known that if $`\alpha ^{}=\alpha `$ then the systems (1) and (3) synchronize after an initial transient, provided the conditional Lyapunov exponents (CLE’s) of the system (3) are all negative . The CLE’s are obtained from the eigenvalues of the Jacobian matrix $`J`$ whose elements are given by, $$J_{ij}=\frac{f_i}{x_j}ϵ\delta _{i1}\delta _{j1}$$ (4) Since the values $`\alpha =(\alpha _1,\mathrm{},\alpha _m)`$ are unknown, we need to set $`\alpha ^{}=(\alpha _1^{},\mathrm{},\alpha _m^{})`$ to random initial values and evolve them adaptively so that they converge to the values $`\alpha `$. Note that a good guess for the initial values of $`\alpha ^{}`$, can be useful in many cases. We first consider the case when $`\alpha `$ (and its counterpart $`\alpha ^{}`$) contains only a single element, i.e. the case when only a single parameter in Eq. (1) is unknown. For notational simplicity we now denote this single parameter by $`\alpha `$. We start with a random initial value for $`\alpha ^{}`$ and evolve it in a controlled fashion so that it converges to $`\alpha `$. This is achieved by raising $`\alpha ^{}`$ to the status of a variable which evolves as, $$\dot{\alpha }^{}=\delta (x_1^{}x_1)w\left(\frac{f_1}{\alpha ^{}}\right),$$ (5) where $`\delta `$ is called stiffness constant and $`w`$ is some suitably chosen function of $`f_1/\alpha ^{}`$. A simple choice for $`w`$ is $`w=f_1/\alpha ^{}`$ giving the adaptive evolution equation for $`\alpha ^{}`$ as, $$\dot{\alpha }^{}=\delta (x_1^{}x_1)\frac{f_1}{\alpha ^{}}.$$ (6) Eq. (5) or Eq. (6) when coupled with Eq. (3) constitutes our method of parameter estimation. A vector $`(𝐱^{},\alpha ^{})`$ initially set to random values asymptotically converges to a vector $`(𝐱,\alpha )`$ in Eq. (1) provided the conditional Lyapunov exponents (CLE’s) for the combined system (Eqs. (3) and (6)) are all negative. This facilitates the estimation of $`\alpha `$. Eq. (6) is equivalent to a dynamic algorithm for minimization of synchronization error between Eqs. (1) and (3) as discussed in Ref. . Note that if we assume in the above discussion that the unknown parameter $`\alpha `$ appears in the function $`f_1`$ corresponding to the variable $`x_1`$ for which the time series is given then the calculation of the factor $`f_1/\alpha ^{}`$ in Eq. (6) is straightforward. However this may not be necessarily the case. The parameter $`\alpha `$ may appear in any of the other system functions. If it appears in the functions for the variables for which the time series is not given, e.g. in any of the functions $`f_2,\mathrm{},f_n`$ in Eq. (1), then correspondingly the calculation of the factor $`f_1/\alpha ^{}`$ becomes nontrivial. To make this point clear we assume that the unknown parameter $`\alpha `$ appears in the function $`f_k(𝐱)`$ governing the evolution of variable $`x_k`$ with $`k1`$ while the time series of $`x_1`$ is given. In such a case Eq. (6) gets modified to, (See Ref. ) $$\dot{\alpha }^{}=\delta (x_1^{}x_1)\frac{f_1}{x_k^{}}\frac{f_k}{\alpha ^{}}.$$ (7) Further if the variable $`x_k`$ itself does not appear in the function $`f_1`$ then the complexity of the calculation still increases. This issue has been explained in detail with an example in Ref. . Next we consider the case when the set $`\alpha `$ of unknown parameters contains more than one element, say $`(\alpha _1,\alpha _2,\mathrm{})`$. Now we set up an adaptive evolution for each of the corresponding parameters $`(\alpha _1^{},\alpha ^{}2,\mathrm{})`$. For the case of two unknown parameters $`\alpha _1`$ and $`\alpha _2`$, appearing in functions $`f_k`$ and $`f_l`$ respectively, the adaptive evolution is given by, $`\dot{\alpha }_1^{}`$ $`=`$ $`\delta _1(x_1^{}x_1){\displaystyle \frac{f_1}{x_k^{}}}{\displaystyle \frac{f_k}{\alpha ^{}}}`$ (8) $`\dot{\alpha }_2^{}`$ $`=`$ $`\delta _2(x_1^{}x_1){\displaystyle \frac{f_1}{x_l^{}}}{\displaystyle \frac{f_l}{\alpha ^{}}},`$ (9) where $`\delta _1`$ and $`\delta _2`$ are two stiffness constants deciding the rates of convergence. For estimating the values of $`\alpha _1`$ and $`\alpha _2`$ Eqs. (9) can be coupled with Eqs. (3) which provide the necessary synchronization of system variables if the associated CLE’s are negative. In the next subsection we extend our method to a situation when a time series of a scalar function of phase space variables is given. We show that it is not only possible to build a synchronizing system but also to adaptively estimate an unknown parameter. ### B Parameter estimation using time series of a scalar function of variables In our discussion of parameter estimation in earlier subsection, we have assumed that time series of one of the phase space variables is given. This may not be the case in many practical applications and in general the observed quantity can be a function of the phase space variables, say $`s(𝐱)`$. It is possible to construct a synchronization scheme in such a situation . We consider the system given by Eq. (1) and assume that the time series $`s(𝐱)`$ which is a function of phase space variables is given. A synchronization scheme can be set up in this case by using a suitable modification of the feedback in Eq. (3) as follows . $`\dot{x}_1^{}`$ $`=`$ $`f_1(𝐱^{},\alpha )ϵ\mathrm{sgn}\left({\displaystyle \frac{\mathrm{s}^{}}{\mathrm{x}_1^{}}}\right)(\mathrm{s}^{}\mathrm{s}(𝐱))`$ (10) $`\dot{x}_j^{}`$ $`=`$ $`f_j(𝐱^{},\alpha )j=2,\mathrm{},n.`$ (11) where $`s^{}=s(𝐱^{})`$ and we give a feedback proportional to $`(s^{}s)`$ in the function $`f_1`$ with feedback constant $`ϵ`$. The function $`s(𝐱)`$ denotes the given time series. It can be shown that if the parameters $`\alpha `$ are assumed to be known, the above system of equations for $`x^{}`$ (Eqs.(11)) converges to $`x`$, provided the CLE’s are all negative . In Eqs. (11), we have assumed that $`s(𝐱)`$ has an explicit dependence on the variable $`x_1`$ so that $`s^{}/x_1^{}0`$. If this is not the case, we can choose any other variable for the feedback on which $`s(𝐱)`$ depends explicitly. The factor $`\mathrm{sgn}()`$ in Eq. (11) makes sure that the term provides a ‘negative feedback’ for all the time so that a convergence is feasible. To estimate parameter $`\alpha `$ in such a case, we set up a synchronization scheme combined with an adaptive control in analogy with Eqs. (3) and (5). This system can be written as, $`\dot{x}_1^{}`$ $`=`$ $`f_1(𝐱^{},\alpha ^{})ϵ\mathrm{sgn}\left({\displaystyle \frac{\mathrm{s}^{}}{\mathrm{x}_1^{}}}\right)(\mathrm{s}^{}\mathrm{s}(𝐱))`$ (12) $`\dot{x}_j^{}`$ $`=`$ $`f_j(𝐱^{},\alpha ^{})j=2,\mathrm{},n.`$ (13) $`\dot{\alpha }^{}`$ $`=`$ $`\delta \mathrm{sgn}\left({\displaystyle \frac{s^{}}{x_1^{}}}\right)(s^{}s(𝐱)){\displaystyle \frac{f_1}{\alpha ^{}}}.`$ (14) Eqs. (14) can be used for estimating $`\alpha `$ when a time series of $`s(𝐱)`$ is given. The condition for such an estimation of $`\alpha `$ to be possible is that the CLE’s associated with the system (14) are all negative. To demonstrate the above procedure, we consider the Lorenz system given by, $`\dot{x}`$ $`=`$ $`\sigma (yx)`$ (15) $`\dot{y}`$ $`=`$ $`rxyxz`$ (16) $`\dot{z}`$ $`=`$ $`xybz,`$ (17) where the variables $`(x,y,z)`$ define the state of the system while $`(\sigma ,r,b)`$ are the three parameters. We consider the case when the time series of $`s(x,y,z)=0.5x^2+1.1y`$ is given as an output of the above system and the parameter $`\sigma `$ is unknown. To estimate the value of $`\sigma `$, we form a system of variables $`(x^{},y^{},z^{},\sigma ^{})`$ similar to Eq. (14). The evolution equations are $`\dot{x}^{}`$ $`=`$ $`\sigma ^{}(y^{}x^{})ϵ\mathrm{sgn}(x^{})(s^{}s(x,y,z))`$ (18) $`\dot{y}^{}`$ $`=`$ $`rx^{}y^{}x^{}z^{}`$ (19) $`\dot{z}^{}`$ $`=`$ $`x^{}y^{}bz^{}`$ (20) $`\dot{\sigma }^{}`$ $`=`$ $`\delta \mathrm{sgn}(x^{})(s^{}s(x,y,z))(y^{}x^{}),`$ (21) where $`s^{}=0.5x^2+1.1y^{}`$. Figures 1(a)-(d) show the evolution of the differences $`x^{}x,y^{}y,z^{}z,\sigma ^{}\sigma `$ respectively (Eqs. (17) and (21)) as a function of time $`t`$. We see that these differences all go to zero as $`t\mathrm{}`$. This indicates that an unknown $`\sigma `$ can be estimated using Eq. (21). The CLE’s are obtained using the Jacobian matrix $`J`$ given by $$J=\left(\begin{array}{cccc}\sigma ϵ\mathrm{sgn}(x)x& \sigma 1.1ϵ\mathrm{sgn}(x)& 0& yx\\ rz& 1& x& 0\\ y& x& b& 0\\ \delta \mathrm{sgn}(x)x(yx)& 1.1\delta \mathrm{sgn}(x)(yx)& 0& 0\end{array}\right)$$ (22) We have verified that all the CLE’s are less than zero except one trivial CLE which is zero. We have performed simulations and successfully estimated unknown parameters in Lorenz system with other forms of the function $`s(x,y,z)`$. The function $`s(x,y,z)`$ should however be such that all the associated conditional Lyapunov exponents should be negative. ### C A general quadratic flow in 3-D Now we consider a quadratic flow in 3-D given by, $`\dot{x}=`$ $`a_0+a_1x+a_2y+a_3z+a_4x^2+a_5y^2`$ (23) $`+`$ $`a_6z^2+a_7xy+a_8yz+a_9xz`$ (24) $`\dot{y}=`$ $`b_0+b_1x+b_2y+b_3z+b_4x^2+b_5y^2`$ (25) $`+`$ $`b_6z^2+b_7xy+b_8yz+b_9xz`$ (26) $`\dot{z}=`$ $`c_0+c_1x+c_2y+c_3z+c_4x^2+c_5y^2`$ (27) $`+`$ $`c_6z^2+c_7xy+c_8yz+c_9xz,`$ (28) where $`(a_0,\mathrm{},a_9,b_0,\mathrm{},b_9,c_0,\mathrm{},c_9)`$ form a thirty dimensional parameter space and $`(x,y,z)`$ are the three variables. We have performed simulations in which we have assumed more than one of the thirty parameters of the system (28) to be unknown and tried to estimate them when a time series of one of the variables is given. To elaborate, we assume some of the thirty parameters to be unknown while the remaining to be known. Some of the known or unknown parameters may be zero thereby making the corresponding term absent from the system. To illustrate the procedure we consider a case when three parameters $`(a_1,a_2,a_7)`$ are unknown and a time series of $`x`$ is given, we set up a system of equations similar to Eq. (3) with the adaptive control loops similar to Eq. (9) for the three parameters $`(a_1^{},a_2^{},a_7^{})`$ as, $`\dot{a}_1^{}`$ $`=`$ $`\delta _1(x^{}x)x^{}`$ (29) $`\dot{a}_2^{}`$ $`=`$ $`\delta _2(x^{}x)y^{}`$ (30) $`\dot{a}_7^{}`$ $`=`$ $`\delta _3(x^{}x)x^{}y^{}.`$ (31) Eqs. (31) when coupled to the system of variables $`(x^{},y^{},z^{})`$ with an identical structure of evolution as Eq. (28) with a feedback term in the evolution of $`x^{}`$, can provide the necessary estimation of parameters when the CLE’s associated with the reconstructed system are all negative. In Fig.2(a)-(c) we plot the time evolution of the differences $`a_1^{}a_1,a_2^{}a_2,a_7^{}a_7`$ as a function of time. The correct value of $`a_7`$ was zero while the other two were non-zero. All the differences go to zero indicating the feasibility of simultaneous estimation of the three parameters $`(a_1,a_2,a_7)`$ even when the actual value of one of them is zero. This shows that the method does not falsely detect a term which is absent in the system. We have found cases when our method can be used successfully for the system (28) to simultaneously estimate as many as five parameters. (One such case is the set of parameters $`a_1,a_2,a_7,b_3,c_1`$, while the time series of $`x`$ is given.) Further we have also found that when any two of the thirty parameters in the system (28) are unknown, we can apply our method to simultaneously estimate them asymptotically to any desired accuracy when the time series of a suitably chosen variables is given. Our results suggest that the information about all the thirty parameters should in principle be contained in the time series of a single variable of the system, though at present we do not have any systematic approach to the simultaneous estimation of all of them. ## III parameter estimation using a finite time series ### A Algorithm for repetitive use In this section we discuss an algorithm for repetitive use of our method to impove the accuracy of parameter estimation when the given time series is of finite duration. Before going on to describe the algorithm it should be mentioned here that even if a finite time series is used repeatedly, we do not expect an exact estimation of the unknown parameter. A finite chaotic trajectory sets a limit on the accuracy to which the unknown parameter can be estimated. This can be seen as follows : We consider symbolic dynamics on the attractor which provides a generating partion of the attractor. It is well known that as the system evolves in time, a finer and finer coarse graining is required to specify a particular trajectory or alternatively, the trajectory gives us a finer coarse grained information about the attractor. The number of coarse grained partitions as a function of time goes as, $$n_pexp\{ht\},$$ (32) where $`h`$ is the Kolmogorov entropy . If $`\xi ^d`$ is the volume of a hypercube in a $`d`$ dimensional phase space and if the size of the attractor is normailized to unity, the number of hypercubes in a generating partion may be approximated as, $$n_p\frac{1}{\xi ^d}.$$ (33) Equations (32) and (33) indicate that the length scale of a hypercube in a generating partition goes as, $$\xi exp\{\frac{h}{d}t\}.$$ (34) It can be seen from Eq. (34) that as long as $`t`$ is finite, the volume of the hypercube in a coarse graining of the attractor will not reduce to zero. Thus a finite trajectory sets a limit on the accuracy to which any information can be extracted from it. This can be further related to Lyapunov exponents using the famous Kaplan-Yorke conjecture as, $$\xi exp\{\frac{\underset{\lambda >0}{}\lambda }{d}t\},$$ (35) where $`\lambda `$ is the characteristic Lyapunov exponent of the system. For a chaotic system with a single positive Lyapunov exponent denoted by $`\lambda ^+`$, eqn. (35) reduces to, $$\xi exp\{\frac{\lambda ^+}{d}t\}.$$ (36) Now we will discuss the algorithm for repetitive use of a finite time series to estimate an unknown parameter. Similar to the case considered in section IIA, we assume that the parameters $`\alpha =(\alpha _1,\alpha _2,\mathrm{},\alpha _m)`$ in Eq.(1) are unknown while the time series of $`x_1`$ is given. We further assume that the time series is truncated after a finite time $`T`$. For the time interval $`0tT`$, we can use the procedure identical to that described earlier (Eqs. (3) and (6)) to evolve variables $`(𝐱^{},\alpha ^{})`$ with random initial conditions. The given finite time series is fed in system (3) as in the earlier case. In this way we can get an approximate value of $`\alpha `$ which we denote as, $`\alpha ^1=\alpha ^{}(T)`$. Now at time $`t=T`$ we set the variables $`𝐱^{}`$ to exactly the same (randomly chosen earlier) initial values while $`\alpha ^{}=\alpha ^1`$ and feed the same finite time series $`\{x(t)|0tT\}`$ again into the system (3) through the feedback terms in Eqs. (3) and (6), i.e. we set $`x(t+T)=x(t)`$. We now evolve the variables $`(𝐱^{},\alpha ^{})`$ for the time interval $`Tt2T`$ to obtain a new estimated value of $`\alpha `$ which is $`\alpha ^2=\alpha ^{}(2T)`$. We repeat the procedure to get successive estimates for the value of $`\alpha `$ denoted by $`\alpha ^1,\alpha ^2,\alpha ^3,\mathrm{},\alpha ^N,\mathrm{}`$ at times $`t=T,2T,3T,\mathrm{},NT,\mathrm{}`$ respectively. Thus, starting from an initial guess for the value of $`\alpha `$ we obtain a sequence of estimates $`\alpha ^0,\alpha ^1,\mathrm{},\alpha ^N`$ after $`N`$ usages of the given finite time series. For large enough $`N`$ we get a better and better estimate of $`\alpha `$, although eventually the accuracy of such an estimate saturates as $`N`$ is increased further. The conditions for the method of parameter estimation using a finite time series to work successfully are : 1. The conditional Lyapunov exponents associated with the reconstructed system should be all negative. 2. The time $`T`$ after which the given time series is truncated should satisfy $`T>\tau `$ where $`\tau `$ denotes the transient time required for synchronization of the systems (1) and (3) with the parameter evolution given by Eq. (6). In the next subsection, we discuss two examples of parameter estimation from a finite time series, viz. Lorenz system and an electrical circuit of a phase converter. ### B Examples #### 1 Lorenz system As our first example we choose the Lorenz system given by Eq. (17) where we assume that the time series $`\{x(t)|0tT\}`$ is given and the value of $`\sigma `$ is to be estimated. We set up the following system of equations (see Eqs. (3) and (6)). $`\dot{x}^{}`$ $`=`$ $`\sigma (y^{}x^{})ϵ(x^{}x)`$ (37) $`\dot{y}^{}`$ $`=`$ $`rx^{}y^{}x^{}z^{}`$ (38) $`\dot{z}^{}`$ $`=`$ $`x^{}y^{}bz^{}`$ (39) $`\dot{\sigma }^{}`$ $`=`$ $`\delta (x^{}x)(y^{}x^{})`$ (40) where we feed the given time series in the evolution of $`x`$ for the interval $`0tT`$ to obtain the first estimate $`\sigma ^1`$. As described in the earlier subsection, we then go on repetitively feeding the same finite time series $`x(t)`$ in Eq.(40) to obtain successive estimates for the value of $`\sigma `$. Starting from a random initial value we denote this sequence of estimates by $`\sigma ^0,\sigma ^1,\mathrm{},\sigma ^N`$ where N denotes the number of times we use the given time series. In Fig.3 we plot the evolution of the difference $`\sigma ^{}\sigma `$ as a function of time $`t`$ during the time interval $`0t3T`$ where we use the time series $`x(t)`$ thrice. We see that the difference decreases as we increase the number of times the finite time series is used. We also observe that shortly after each resetting of the initial vector $`(x^{},y^{},z^{})`$ which is done at times $`T,2T`$, the synchronization weakens and fluctuations are present. This is due to the random resetting of the $`y`$ and $`z`$ components which gives a transient before the synchronization is recovered. An appropriate feedback constant $`ϵ`$ may be chosen to lessen this transient in every usage of the time series. In Fig.4 we plot the successive differences $`\sigma ^N\sigma `$ as a function of $`N`$, the number of times we use the given finite time series. We see that the difference $`\sigma ^N\sigma `$ goes on decreasing with increasing $`N`$. However, as $`N`$ is increased further, it saturates to a constant finite value depending on the length of the time series used for the calculations. This is consistent with our expectations that finite time series can contain only finite information about the system as discussed in the previous subsection, e.g. using $`\lambda ^+0.9`$, the finest length scale that can be obtained using a finite time series with $`T=30`$ is estimated to be $`0.05`$ (Eq. (36)) which means an accuracy of about $`10^3`$. This is also the order of magnitude of the accuracy of parameter estimation. The three curves in Fig. 4 correspond to three different values of $`T=T_1<T_2<T_3`$. We see that an increasing $`T`$ gives better estimate of the parameter. This is natural since a very long time series corresponding to $`T\mathrm{}`$ is expected to give an exact estimation of the unknown parameter. We have similarly implemented our method to estimate other parameters of Lorenz system using finite time series of either $`x`$ or $`y`$. The method fails to estimate any of the parameters when the time series of $`z`$ is given. The reason for this is that one of the associated conditional Lyapunov exponents is critically zero and the convergences are slow. #### 2 A phase converter circuit As our next example, we consider the set of equations describing an electrical circuit for a phase converter system in a dimensionless form given by, $`\dot{x}_1`$ $`=`$ $`x_2`$ (41) $`\dot{x}_2`$ $`=`$ $`kx_2{\displaystyle \frac{x_1}{4}}\left(x_1^2+3x_3^2\right)`$ (42) $`\dot{x}_3`$ $`=`$ $`x_4`$ (43) $`\dot{x}_4`$ $`=`$ $`kx_4{\displaystyle \frac{x_3}{4}}\left(x_1^2+3x_3^2\right)+B\mathrm{cos}t`$ (44) where $`k`$ and $`B`$ are the two parameters. Here we consider the time series $`\{x_2(t)|0tT\}`$ to be given. Notice that the system (44) has a simple time dependent term making it a non-autonomous system. Such a system is equivalent to an autonomous system in higher dimensions. We have successfully estimated any one of the parameters $`k`$ or $`B`$ (or both) using finite time series of $`x_2(t)`$. Figure 5(a) shows a schematic diagram of the circuit for the phase converter. The system in known to exhibit a chaotic behaviour due to period doubling bifurcations, codimension two bifurcations etc. Figure 5 (b) shows a chaotic attractor in the $`x_1x_2`$ plane of the phase space. Figure 6 shows the plot of the successive differences $`k^Nk`$ as a function of $`N`$, the number of times we use the given time series for two different values of the truncation time T. As expected the accuracy of the estimation increases with increasing $`T`$ while showing a saturation with increasing number of repeated usages. Thus, we have shown how the method of parameter estimation can be used when a finite time series is given. The method works when the CLE’s associated are all negative and the time series given is of longer duration than the transient time required for synchronization. ## IV Form of a model perturbation Here we describe an interesting application of our method to test a function modeling an unknown external source of perturbation to a known chaotic system. In many practical situations when an external source of disturbance in not known, a trial function is used to model the perturbation. We imagine a situation when it is required to verify a proposed trial model form for the perturbation. We denote the actual perturbation by a function $`F(𝐱,\mu )`$ and the trial function by $`G(𝐱^{},\mu ^{})`$ where $`\mu `$ and $`\mu ^{}`$ are parameters. In the following, we demonstrate the use of our method of parameter estimation to confirm the form of the trial function. Note that here we do not deal with the issue of obtaining the form of the model function. Now if the proposed trial function $`G`$ models the external perturbation $`F`$ correctly, then a scheme based on synchronization combined with adaptive control should produce synchronization of variables and make the parameters $`\mu ^{}`$ converge (to $`\mu `$). Thus a successful synchronization should then indicate a correctly chosen model function. In this manner we can use the method to distinguish between a correct model and a wrong model for an external perturbation. We elaborate on this application further using the example of Lorenz system. Consider the Lorenz system perturbed by a sinusoidal term $`F=A\mathrm{sin}(\omega x)`$, $`\dot{x}`$ $`=`$ $`\sigma (yx)+A\mathrm{sin}(\omega x)`$ (45) $`\dot{y}`$ $`=`$ $`rxyxz`$ (46) $`\dot{z}`$ $`=`$ $`xybz,`$ (47) where we assume the unperturbed Lorenz system to be known. The function $`F=A\mathrm{sin}(\omega x)`$ is the external perturbation. We assume that the time series of $`x`$ is given as an output of the system (47). To set up the required scheme we construct a system of variables $`(x^{},y^{},z^{})`$ and their evolution as, $`\dot{x}^{}`$ $`=`$ $`\sigma (y^{}x^{})+G(x^{},y^{},z^{},\mu ^{})ϵ(x^{}x)`$ (48) $`\dot{y}^{}`$ $`=`$ $`rx^{}y^{}x^{}z^{}`$ (49) $`\dot{z}^{}`$ $`=`$ $`x^{}y^{}bz^{},`$ (50) $`\dot{\mu }^{}`$ $`=`$ $`\delta (x^{}x){\displaystyle \frac{G}{\mu ^{}}}.`$ (51) where $`G(x^{},y^{}z^{})`$ is the trial perturbation function. We feed the time series $`x(t)`$ obtained from system (47) into the model system (51). Now if $`G`$ models the behaviour of $`F`$ correctly then the two systems should exhibit synchronization while the parameters should show convergence to the correct values. In our simulations we have tried several different forms for the trial function $`G`$. Figures 7(a)-(c) show the time evolution of $`x^{}x,\mu _1`$ and $`\mu _2`$ respectively while the feedback is given into $`x`$ and the trial function is $`G=\mu _1x^2+\mu _2`$. It can be clearly seen that there is no synchronization of variables. The trial function $`G=\mu _1x^2+\mu _2`$ thus fails to produce synchronization and hence can be discarded as a plausible model for $`F`$. We also note that the parameters $`\mu _1^{}`$ and $`\mu _2^{}`$ do not show convergence. In Figs. 8 and 9, we plot similar graphs for two more choices of the trial function. In Fig. 8(a)-(c) we use $`G=\mu _1x\mu _2x^3`$ and plot $`x^{}x,\mu _1`$ and $`\mu _2`$ respectively. We choose this form of $`G`$ since it represents the two leading terms in the series expansion of the the function $`F=A\mathrm{sin}(\omega x)`$. We can see from Fig. 8 that such an approximation fails to produce synchronization and also the convergence of parameters. As a third choice we use $`G=\mu _1\mathrm{sin}(\mu _2x)`$ in Eq. (51) and plot the time evolution of $`x^{}x,\mu _1`$ and $`\mu _2`$ in Fig.9(a)-(c) respectively. The difference $`x^{}x`$ goes to zero as time increases showing synchronization. The parameters $`\mu _1`$ and $`\mu _2`$ converge to the correct values $`A`$ and $`\omega `$ respectively. The variables $`y^{}`$ and $`z^{}`$ also synchronize with $`y`$ and $`z`$ respectively. This confirms that this trial function correctly models the function $`F`$. Now as a last consideration, we use the form $`G=\mu _1\mathrm{sin}(\mu _2x)`$ again but unlike in Eq. (51) we perturb a wrong variable in the model system, i.e. we choose to add the trial perturbation in the evolution of say $`y^{}`$. The feedback is given in $`x`$. The evolution equations are $`\dot{x}^{}`$ $`=`$ $`\sigma (y^{}x^{})ϵ(x^{}x)`$ (52) $`\dot{y}^{}`$ $`=`$ $`rx^{}y^{}x^{}z^{}+G(x^{},y^{},z^{},\mu ^{})`$ (53) $`\dot{z}^{}`$ $`=`$ $`x^{}y^{}bz^{},`$ (54) $`\dot{\mu }^{}`$ $`=`$ $`\delta (x^{}x){\displaystyle \frac{G}{\mu ^{}}}.`$ (55) In Fig.10 (a)-(c) we plot the time evolution of $`x^{}x,\mu _1`$ and $`\mu _2`$ respectively. We see that even if $`G`$ correctly models $`F`$, synchronization does not take place. This shows that along with the form of $`F`$ we can also confirm a guess about the perturbed variable. Thus, the results presented in this section suggest that the method which we use for estimating parameters can be used to distinguish between a correct trial function and the wrong trial functions for an unknown external perturbation to a known system . ## V summary and conclusions We have described dynamic method of parameter estimation from a given chaotic time series of a phase space variable of a dynamical system . Further, We have generalized the method for the case when the quantity for which the time series is given is a scalar function of the phase space variables. We have shown that it is not only possible to synchronize two systems using the time series of the scalar function but also to asymptotically estimate unknown parameters adaptively to any desired accuracy. This is done by providing a linear feedback in the evolution of one of the variables on which the scalar function explicitly depends. The method works successfully provided the function for which the time series is given is such that the associated conditional Lyapunov exponents are all negative. We have also applied our method to a system with a large number of parameters, i.e. a general quadratic flow in 3-D. We have observed that a simultaneous estimation of a few parameters is possible provided the condition of convergence as stated in Ref. is satisfied i.e. all the CLE’s are negative. As a next consideration, we have extended our method to a realistic situation when the given series is truncated after a finite time. We have shown that repetitive use of a finite time series can be made to estimate an unknown parameter of the system. The accuracy of the parameter estimation saturates as the given finite time series is used more and more number of times. The accuracy increases with the increasing length of the given time series. In the end we have demonstrated an important application of our method in confirming the correctness of a trial model function for an unknown external perturbation to a known system. We see that a perfect synchronization between a perturbed system and its dynamical copy using a model for the perturbation is possible only when the form of the trial function is correctly guessed. These results indicate that our method can be used as a test for the trial model for an unknown external perturbation to a known system. Another possible application (not discussed in the paper) is as follows. Our method may be employed to experimentally measure the unknown value of a component added to a known circuit. In such a situation the equations governing the circuit are known, and can be used to estimate the unknown component value accurately. This is feasible due to the asymptotic convergences in our method. ###### Acknowledgements. One of the authors (A.M.) would like to thank U.G.C.(India) for financial support.
warning/0002/cond-mat0002097.html
ar5iv
text
# Charge localization and phonon spectra in hole doped La2NiO4 \[ ## Abstract The in-plane oxygen vibrations in La<sub>2</sub>NiO<sub>4</sub> are investigated for several hole-doping concentrations both theoretically and experimentally via inelastic neutron scattering. Using an inhomogeneous Hartree-Fock plus RPA numerical method in a two-dimensional Peierls-Hubbard model, it is found that the doping induces stripe ordering of localized charges, and that the strong electron-lattice coupling causes the in-plane oxygen modes to split into two subbands. This result agrees with the phonon band splitting observed by inelastic neutron scattering in La<sub>2-x</sub>Sr<sub>x</sub>NiO<sub>4</sub>. Predictions of strong electron-lattice coupling in La<sub>2</sub>NiO<sub>4</sub>, the proximity of both oxygen-centered and nickel-centered charge ordering, and the relation between charged stripe ordering and the splitting of the in-plane phonon band upon doping are emphasized. \] There is currently great interest in the importance of charge localization and ordering tendencies in a variety of doped transition metal oxides: including nickelates, bismuthates, cuprates, and manganites. Recent experiments have suggested nanoscale coexistence of charge and spin ordering, as well as related multiscale dynamics . The cuprates have been widely investigated, both theoretically and experimentally, as this inhomogeneity may be related to high-temperature superconductivity. The nickelates are considered strong electron-lattice (e-l) coupling systems, which helps stabilize charge-ordering in the form of ”stripe” phases . For the commensurate 1/3 doping case of La<sub>1.67</sub>Sr<sub>0.33</sub>NiO<sub>4</sub>, it has been shown in optical absorption and Raman scattering experiments that new phonon modes appear when the temperature is lowered below the stripe-ordering temperature ($`T_{so}`$=240 K); this is a signature of the stripe formation . Until now, only the temperature dependence and the apical oxygen (Ni-O(2)) vibrations have been investigated. However, the doping dependence and the in-plane oxygen vibrations (Ni-O(1) stretching modes) are also very important for the properties of the quasi-two dimensional nickelate materials. In this study, we use an inhomogeneous Hartree-Fock (HF) plus random-phase approximation (RPA) numerical method for a two-dimensional (2D), four-band Peierls-Hubbard model, to interpret the inelastic neutron scattering spectra. This reveals specific signatures of the stripe patterns in the in-plane oxygen phonons. Our main results are: (i) There is agreement between the results from our multiband model including electron-electron and e-l interactions and the inelastic neutron scattering spectra for the in-plane oxygen vibrations with various commensurate hole-doping concentrations; (ii) The theoretical results predict new vibrational modes (“edge modes”) which are associated with oxygen motions near localized holes or in the vicinity of stripes; (iii) The e-l coupling strength, at which the best agreement between our model and the inelastic neutron scattering data is achieved, is close to the transition from an oxygen-centered stripe phase to a nickel-centered one. This suggests that the nickelates may be in a mixed state of both stripe phases, and sensitive to temperature, pressure and magnetic field. The inelastic neutron scattering spectra were measured on polycrystalline La<sub>2-x</sub>Sr<sub>x</sub>NiO<sub>4</sub> for various doping concentrations, $`x=`$ 0, 1/8, 1/4, 1/3, 1/2. Time-of-flight neutron scattering measurements were performed on the Low Resolution Medium Energy Chopper Spectrometer at Argonne National Laboratory’s Intense Pulsed Neutron Source. For all measurements, an incident neutron energy of 120 meV was chosen and data were summed over all scattering angles from $`2^{}120^{}`$. Detailed information on the experiment, as well as the preparation of the samples used, can be found in Ref. . The focus is on the (neutron-scattering-weighted) generalized density-of-states (GDOS) of phonons and particular attention is given to the in-plane oxygen vibrations, i.e., the Ni-O(1) stretching modes. Figure 1 shows the experimental GDOS for several hole concentrations at $`T`$=10 K for phonon modes in the range from 50-100 meV. From the analysis of lattice dynamical shell model calculations, it is known that the in-plane Ni-O(1) oxygen stretching modes are separated in frequency from other vibrations, and the phonon intensity above 65 meV is associated entirely with these vibrations. For the low doping samples ($`x=1/8`$ and $`1/4`$, not shown), there is little change in the $``$81 meV phonon band , but for the $`x=1/3`$ and $`x=1/2`$ samples, a new peak appears around 75 meV along with a slight hardening of the main band. This feature is interpreted as a splitting of the 81 meV Ni-O(1) stretching band into two phonon bands centered at approximately 83 meV and 75 meV. In order understand this dependence of the Ni-O(1) stretching modes on hole concentration and its possible reflection of stripe ordering, we have performed a calculation of the phonon spectrum in a minimal Peierls-Hubbard model in 2D. Due to the strong e-l coupling expected in the nickelates, we resort to modeling with an inhomogeneous HF plus RPA numerical approach . This has proven to be a very robust method for studying charge localization and stripe formation, especially when electron-lattice coupling is strong, obviating subtle many-body effects and quantum fluctuations . We use a 2D four-band extended Peierls-Hubbard model of a doped NiO<sub>2</sub> plane, which includes both electron-electron and e-l interactions . Here, for nickelate oxides, besides the $`d_{x^2y^2}`$ orbital used in the cuprate oxide models , the $`d_{3z^21}`$ Ni $`d`$ orbitals must be included to account for the higher spin state ($`S=1`$) at half-filling (i.e. undoped). Our model Hamiltonian is : $`H`$ $`=`$ $`{\displaystyle \underset{ij,m,n,\sigma }{}}t_{im,jn}(u_{ij})(c_{im\sigma }^{}c_{jn\sigma }+\mathrm{H}.\mathrm{c}.)`$ (1) $`+`$ $`{\displaystyle \underset{i,m,\sigma }{}}ϵ_mc_{im\sigma }^{}c_{im\sigma }+{\displaystyle \underset{ij}{}}{\displaystyle \frac{1}{2}}K_{ij}u_{ij}^2+H_c,`$ (2) where $`c_{im\sigma }^{}`$ creates a hole with spin $`\sigma `$ at site $`i`$ in orbital $`m`$ (Ni $`d_{x^2y^2}`$, $`d_{3z^21}`$, or O $`p`$). The Ni-O hopping $`t_{im,jn}`$ has two values: $`t_{pd}`$ between $`d_{x^2y^2}`$ and $`p`$ and $`\pm t_{pd}/\sqrt{3}`$ between $`d_{3z^21}`$ and $`p`$. The O-site electronic energy is $`ϵ_p`$, and Ni-site energies are $`ϵ_d`$ and $`ϵ_d+E_z`$ for $`ϵ_m`$, with $`E_z`$ the crystal-field splitting on the Ni site. $`H_c`$ describes the electron correlations in the Ni orbitals, $`H_c`$ $`=`$ $`{\displaystyle \underset{im}{}}(U+2J)n_{im}n_{im}{\displaystyle \underset{i,mn}{}}2J𝐒_{im}𝐒_{in}`$ (3) $`+`$ $`{\displaystyle \underset{i,mn,\sigma ,\sigma ^{}}{}}(UJ/2)n_{im\sigma }n_{in\sigma ^{}}`$ (4) $`+`$ $`{\displaystyle \underset{i,m,n}{}}Jc_{im}^{}c_{im}^{}c_{in}c_{in}`$ (5) The electron-electron interactions include the on-site Ni Coulomb repulsions ($`U`$) as well as the Hund interaction ($`J`$) at the same Ni site to account for the high spin situation. (The interplay of the two orbitals can also lead to pseudo Jahn-Teller distortions, but these are not our focus here). We emphasize that, due to the large spin at the nickel site, Hund’s rule leads to ferromagnetic exchange coupling $`2J`$, and $`𝐒_{\mathrm{𝐢𝐦}}=\frac{1}{2}_{\tau \tau ^{}}c_{im\tau }^{}𝝈_{𝝉𝝉^{\mathbf{}}}c_{im,\tau ^{}}`$,with $`𝝈`$ the Pauli matrix. For the e-l interaction, we consider that the Ni-O hopping is modified linearly by the O-ion displacement $`u_{ij}=u_\mathrm{O}`$ as $`t_{im,jn}(u_{ij})=t_{im,jn}(1\pm \alpha u_\mathrm{O})`$, where the $`+`$ ($``$) applies if the bond shrinks (stretches) with positive $`u_\mathrm{O}`$. For the lattice terms, we study only the motion of O ions along the Ni-O bonds—other oxygen (or Ni) distortion modes can readily be included if necessary. It is known that for the nickelate oxides the e-l coupling is stronger than in cuprate oxides , and is therefore likely to play an even more decisive role in the formation, localization, and nature of stripe phases. We adopt the following representative paremeters for the nickelate materials : $`t_{pd}=1`$, $`\mathrm{\Delta }=ϵ_pϵ_d=9`$, $`U=4`$, $`J=1`$, $`E_z=1`$ and $`K=32t_{pd}/\AA ^2`$ (all in units of $`t_{pd}`$). In real oxides , $`t_{pd}`$ is estimated to be in the range 1.3 eV $``$ 1.5 eV. The electron-lattice coupling strength is varied to achieve a best fit to the neutron scattering data; we find $`\alpha 3.0`$. The commensurate doping cases are examined in a 4x4 unit supercell for $`x=0`$, 1/2 and a 3x3 unit supercell for $`x=1/3`$. Periodic boundary conditions are used. The densities-of-states (DOS) of in-plane phonons were calculated from our model at $`x=0`$, 1/3, 1/2 and are shown in Fig. 2. For the undoped case, as the ground state is spatially homogeneous, only one oxygen phonon band appears, centered around 80.5 meV. When holes are added into the NiO<sub>2</sub> plane at $`x=1/3`$, the ground state is found to be a stripe pattern with more holes accumulating along the (1,1,0) direction, forming an antiphase domain wall within the original antiferromagnetic background. This is consistent with many neutron, optical and Raman scattering experiments . Interestingly, we find that a new phonon band appears centered at 75 meV. In addition, there is a hardening of the main phonon band (to 83 meV) corresponding to an overall splitting of 8 meV. From examination of the eigenvectors, the main character of this band is local oxygen vibrations in the vicinity of the stripe, i.e. having the nature of localized ”edge modes” (see for more details). For the higher doping $`x=1/2`$, the charge-ordering takes on a commensurate checker-board pattern. A similar splitting of the 85 meV phonon band into two phonon bands around 83 and 75 meV is again found. At this half doping, the checker-board ground state is in essence a commensurate charge-density-wave (CDW) system with the nature of an ordered binary alloy. The above results are in agreement with the GDOS data obtained from inelastic neutron scattering, which are shown in Fig. 1. In addition to the splitting energy, even the slight hardening of the main band observed experimentally is accounted for in the model. Besides the new phonon modes centered at 75 meV appearing for the 1/3 and 1/2 doping, another low intensity phonon mode around 65 meV is also predicted in our model at $`x=1/2`$ (Fig. 2). The signature for these modes are weak, however, so that they may be difficult to detect in the current experiment. In so far as we have included only a small subset of the possible oxygen displacement patterns and wavevectors in the model (whereas the neutron scattering experiment samples all wavevectors and polarizations), the relative intensities and widths of the bands obtained by experiment and theory cannot be usefully compared. We emphasize that the excellent agreement between our model and the GDOS experimental data is achieved by varying the e-l coupling strength to match the positions of the phonon bands. As noted, the choice of $`\alpha 3`$ best fits the data; our model calculations predict a variation of the phonon splitting with $`\alpha `$ which is quite large ($`\mathrm{\Delta }\omega _{split}/\mathrm{\Delta }\alpha `$ 7 meV). Most strikingly, as illustrated in Fig. 3, an O-centered stripe is found as the ground state at small $`\alpha 2.2`$, while for a larger $`\alpha 3.0`$, a Ni-centered stripe is found as the ground state. The transition region includes $`\alpha 3.0`$, where the best agreement between Fig. 2 based on our model and Fig. 1 on the inelastic neutron scattering spectra is achieved. It has been suggested from various experimental data that stripe formation for $`x=1/3`$ can not be simply assigned as Ni-centered or O-centered , but is also dependent on temperature. Our comparison of theory and experiment provides a possible explanation on the sensitivity of stripe formation; they suggest that La<sub>1.67</sub>Sr<sub>0.33</sub>NiO<sub>4</sub> may be in a mixed stripe phase state, and also in a region of sensitivity to temperature, pressure, magnetic field, etc. In conclusion, we have made a study of oxygen breathing lattice vibrations in La<sub>2-x</sub>Sr<sub>x</sub>NiO<sub>4</sub> via inelastic neutron scattering compared with predictions of a 2D four-band model, including both electron-lattice and electron-electron interactions. The in-plane oxygen vibrations above 65 meV were thoroughly investigated. The splitting of the in-plane 81 meV band upon doping into two subbands centered around 75 meV and 83 meV is observed experimentally and predicted theoretically, and interpreted in terms of new localized phonon modes (“edge modes” at charge localized stripes). The excellent agreement between the experiment and the model strongly supports the view that strong electron-lattice coupling in this kind of material plays a decisive role on the charge localization and mesoscopic stripe formation. For the doping at $`x=1/3`$, at which stripes are found both in experiments and our model, our results suggest that there may be a mixed state of O- and Ni-centered stripe phases, and sensitivity to temperature, pressure, and magnetic field. Our model calculations also predict distinctive dispersion of the phonon bands, as well as inhomogeneous magnetoelastic coupling along the boundaries between charge-rich and magnetic nanophase domains . These predictions require additional experiments for their confirmation and consequences. We have benefitted from valuable discussions with Dr. J. T. Gammel. This work is supported (in part) by the U. S. Department of Energy under contract W-7405-Eng-36 with the University of California. This work has benefited from the use of the Intense Pulsed Neutron Source at Argonne National Laboratory. This facility is funded by the U. S. Department of Energy, BES-Materials Science, under contract W-31-109-Eng-38.
warning/0002/astro-ph0002522.html
ar5iv
text
# 1 Introduction ## 1 Introduction The last years of the past century are marked by huge efforts of the community of astronomers, physicists and astrophysicists devoted to determine the most fundamental parameters of our Universe, the cosmological parameters. The most important among them are the mass densities of baryons $`\mathrm{\Omega }_b`$ (in units of the critical density) and of cold dark matter $`\mathrm{\Omega }_{cdm}`$, the neutrino rest masses $`m_\nu `$ and their total density $`\mathrm{\Omega }_\nu `$, the value of cosmological term $`\mathrm{\Lambda }`$ (or $`\mathrm{\Omega }_\mathrm{\Lambda }`$), the Hubble constant $`H_0`$, the spatial curvature parameter $`\mathrm{\Omega }_k`$ and the slopes $`n`$ and amplitudes $`A`$ of the primordial power spectra of scalar and tensor fluctuations. The primordial ratio of the number of deuterium to hydrogen nuclei (D/H) created in Big Bang nucleosynthesis is the most sensitive measure of the cosmological density of baryons $`\mathrm{\Omega }_b`$. Quasar absorption systems give definite measurements of the primordial deuterium and the most accurate value of baryon density obtained recently in this way is $`\mathrm{\Omega }_bh^2=0.019\pm 0.0024`$ . The measurements of the neutrino rest mass is not so certain, unfortunately. Up-to-day we have only some indications for the range where it may be found. The oscillations of solar and atmospheric neutrinos registered by the SuperKamiokande experiment show that the difference of rest masses between $`\tau `$ and $`\mu `$-neutrinos is $`0.02<\mathrm{\Delta }m_{\tau \mu }<0.08eV`$ . This also provides a lower limit for the neutrino mass, $`m_\nu |\mathrm{\Delta }m|`$ and does not exclude models with cosmologically significant values $`120eV`$. Therefore, at least two species of neutrinos can have approximately equal masses in this range. Some versions of elementary particle theories predict $`m_{\nu _e}m_{\nu _\tau }2.5eV`$ and $`m_{\nu _\mu }m_{\nu _s}10^5eV`$, where $`\nu _e`$, $`\nu _\tau `$, $`\nu _\mu `$ and $`\nu _s`$ denote the electron, $`\tau `$, $`\mu `$ and sterile neutrinos accordingly (e.g. ). The strongest upper limit for the neutrino mass comes from the observed large scale structure of our Universe: $`_im_{\nu _i}/94\mathrm{e}\mathrm{V}0.3h^2`$. Since observations give for the Hubble parameter an upper limit of $`h=0.8`$ one gets $`_im_{\nu _i}18eV`$. It is interesting to note that this upper limit coincides roughly with the upper limit for the electron neutrino mass obtained from the supernova explosion SN1987A and tritium $`\beta `$-decay experiments. Important conclusions about measurements of matter density $`\mathrm{\Omega }_m`$ ($`\mathrm{\Omega }_b+\mathrm{\Omega }_{cdm}+\mathrm{\Omega }_\nu `$) come from the Supernova Cosmology Project and the High-z Supernova Search. In particular, the relation of observed brightness vs. redshift for SNeIa shows that distant supernovae are fainter than expected for a decelerating Universe, and, thus, more distant. This can be interpreted as an accelerated expansion rate, or $`\mathrm{\Omega }_\mathrm{\Lambda }>0`$. The best-fit value is $`\mathrm{\Omega }_\mathrm{\Lambda }=\frac{4}{3}\mathrm{\Omega }_m+\frac{1}{3}\pm 0.1`$ (1$`\sigma `$ error) and $`\mathrm{\Omega }_m=1`$ models are ruled out at the 8$`\sigma `$ level . For a flat Universe $`\mathrm{\Omega }_m+\mathrm{\Omega }_\mathrm{\Lambda }=1`$ ($`\mathrm{\Omega }_k=0`$) the best-fit values are $`\mathrm{\Omega }_m=0.25\pm 0.1`$ and $`\mathrm{\Omega }_\mathrm{\Lambda }=0.75\pm 0.1`$ . An upper limit of $`\mathrm{\Omega }_\mathrm{\Lambda }<0.7`$ (95% C.L.) follows from gravitational lensing statistics, just consistent with distant supernovas results. Strong evidence against an open Universe can be derived from recent measurements of the position of the first acoustic peak in the cosmic microwave background (CMB) power spectrum by the Boomerang experiment. The $`1\sigma `$ range for the curvature parameter derived from this experiment is $`0.25\mathrm{\Omega }_k0.15`$ and the mean value is close to the flat Universe, $`\mathrm{\Omega }_k0`$. Currently there are a few completely independent and broad routes to the determination of the Hubble constant $`H_0`$. The direct experiments can be divided into three groups: the gravitational lens time delay methods, the Sunyaev-Zel’dovich method for clusters and extra-galactic distance measurements. Almost all observations yield values of $`H_0`$ in the range 50-80 km/sec/Mpc. Other independent methods for the determination of cosmological parameters are based on large scale structure (LSS) observations. Their advantage is that all parameters mentioned above can be determined together because the form and amplitude of the power spectrum of density fluctuations are rather sensitive to all of them. Their disadvantage is that they are model dependent. This approach has been carried out in several papers (e.g. and references therein) and it is also the goal of this paper. The papers on this subject differ by the number of parameters and the set of observational data included into the analysis. In this paper a total of 23 measurements from sub-galaxy scales (Ly-$`\alpha `$ clouds) over cluster scales up to horizon scale (CMB quadrupole) is used to determine eight cosmological parameters, namely the tilt of the primordial spectrum $`n`$, the densities of cold dark matter $`\mathrm{\Omega }_{cdm}`$, hot dark matter $`\mathrm{\Omega }_\nu `$, baryons $`\mathrm{\Omega }_b`$ and cosmological constant $`\mathrm{\Omega }_\mathrm{\Lambda }`$, the number of massive neutrino species $`N_\nu `$, the Hubble parameter $`H_0`$ and, in addition, the bias parameter $`b_{cl}`$ for rich clusters of galaxies. We restrict ourselves to the analysis of spatially flat cosmological models with $`\mathrm{\Omega }_\mathrm{\Lambda }+\mathrm{\Omega }_m=1`$ ($`\mathrm{\Omega }_k=0`$) and to an inflationary scenario without tensor mode. We also neglect the effect of a possible early reionization which could reduce the amplitude of the first acoustic peak in the CMB anisotropy spectrum. In comparison to the companion paper the influence of the uncertainties in the normalization of the scalar mode amplitude caused by experimental errors on determination of cosmological parameters is also taken into account here. ## 2 The experimental data set and our methods We use the power spectrum of Abell-ACO clusters , measured in the range $`0.03k0.2h/`$Mpc, as observational input. Its amplitude and slope at lower and larger scales are quite sensitive to baryon content $`\mathrm{\Omega }_b`$, Hubble constant $`h`$, neutrino mass $`m_\nu `$ and number of species of massive neutrinos $`N_\nu `$ . The total number of Abell-ACO data points with their errors used for minimization is 13, but not all of these points can be considered as independent measurements. Since we can accurately fit the power spectrum by an analytic expression depending on three parameters only (the amplitude at large scales, the slope at small scales and the scale of the bend); we assign to the power spectrum 3 effective degrees of freedom. The second observational data set which we use are the position and amplitude of the first acoustic peak derived from the data on the angular power spectrum of CMB temperature fluctuations. To determine the position and amplitude of the first acoustic peak we use a 6-th order polynomial fit to the data set on CMB temperature anisotropy, accumulated in Table 2 of our accompanied paper , 51 data points in total. The amplitude $`A_p`$ and position $`\mathrm{}_p`$ of first acoustic peak determined from this fit are $`79.6\pm 16.5\mu \mathrm{K}`$ and $`253\pm 70`$ correspondingly. The statistical errors are estimated by edges of the $`\chi ^2`$-hyper-surface in the space of polynomial coefficients which corresponds to 68.3% ($`1\sigma `$) probability level under the assumption of Gaussian statistics. Also the mean weighted bandwidth of each experiment around $`\mathrm{}_p`$ is added to obtain total $`\mathrm{\Delta }\mathrm{}_p`$. A constraint on the amplitude of the matter density fluctuation power spectrum at cluster scale can be derived from the cluster mass and X-ray temperature functions. It is usually formulated in terms of the density fluctuation in a top-hat sphere of 8$`h^1`$ Mpc radius, $`\sigma _8`$, which can be easily calculated for the given initial power spectrum. According to the recent optical determination of the mass function of nearby galaxy clusters and taking into account the results from other authors (for references see ) we use the value $`\stackrel{~}{\sigma }_8\stackrel{~}{\mathrm{\Omega }}_m^{0.460.09\mathrm{\Omega }_m}=0.60\pm 0.08`$. ¿From the existence of three most massive clusters of galaxies observed at $`z>0.5`$ a further constraint has been established by Bahcall & Fan : $`\stackrel{~}{\sigma }_8\stackrel{~}{\mathrm{\Omega }}_m^\alpha =0.8\pm 0.1,`$ where $`\alpha =0.24`$ if $`\mathrm{\Omega }_\mathrm{\Lambda }=0`$ and $`\alpha =0.29`$ if $`\mathrm{\Omega }_\mathrm{\Lambda }>0`$ with $`\mathrm{\Omega }_\mathrm{\Lambda }+\mathrm{\Omega }_m=1.`$ A constraint on the amplitude of the linear power spectrum of density fluctuations in our vicinity comes from the study of galaxy bulk flow, the mean peculiar velocity of galaxies in sphere of radius $`50h^1`$Mpc around our position. We use the data given by Kollat & Dekel , $`\stackrel{~}{V}_{50}=(375\pm 85)`$ km/s. A further essential constraint on the linear power spectrum of matter clustering at galactic and sub-galactic scales $`k(240)h/`$Mpc can be obtained from the Ly-$`\alpha `$ forest of absorption lines seen in quasar spectra . Assuming that the Ly-$`\alpha `$ forest is formed by discrete clouds of a physical extent near Jeans scale in the reionized inter-galactic medium at $`z24`$, Gnedin has obtained a constraint on the value of the r.m.s. linear density fluctuations $`1.6<\stackrel{~}{\sigma }_F(z=3)<2.6`$ (95% C.L.) at Jeans scale for $`z=3`$ equal to $`k_F38\mathrm{\Omega }_m^{1/2}h/`$Mpc . The procedure to recover the linear power spectrum from the Ly-$`\alpha `$ forest has been elaborated by Croft et al.. Analyzing the absorption lines in a sample of 19 QSO spectra, they have obtained the following 95% C.L. constraint on the amplitude and slope of the linear power spectrum at $`z=2.5`$ and $`k_p=1.5\mathrm{\Omega }_m^{1/2}h/`$Mpc $$\stackrel{~}{\mathrm{\Delta }}_\rho ^2(k_p)k_p^3P(k_p)/2\pi ^2=0.57\pm 0.26,$$ (1) $$\stackrel{~}{n}_p\frac{\mathrm{\Delta }\mathrm{log}P(k)}{\mathrm{\Delta }\mathrm{log}k}_{k_p}=2.25\pm 0.1.$$ (2) In addition to the power spectrum measurements we use the constraints on the value of Hubble constant $`\stackrel{~}{h}=0.65\pm 0.15`$ which is a compromise between measurements made by two groups: and . We also employ the nucleosynthesis constraints on the baryon density of $`\stackrel{~}{\mathrm{\Omega }_bh^2}=0.019\pm 0.0024`$ (95% C.L.) . In order to find the best fit model we must evaluate the above mentioned quantities for a given cosmological model. To this end we use the accurate analytic approximations of the MDM transfer function $`T(k;z)`$ depending on the parameters $`\mathrm{\Omega }_m`$, $`\mathrm{\Omega }_b`$, $`\mathrm{\Omega }_\nu `$, $`N_\nu `$, $`h`$ by Eisenstein & Hu . The linear power spectrum of matter density fluctuations is $$P(k;z)=Ak^nT^2(k;z)D_1^2(z)/D_1^2(0),$$ (3) where $`A`$ is the normalization constant and $`D_1(z)`$ is the growth factor, useful analytical approximation for which has been given by Carrol et al.. We normalize the spectra using the 4-year COBE data which can be expressed by the value of the density perturbation at the horizon crossing scale, $`\delta _h`$ . The normalization constant is related to $`\delta _h`$ by $$A=2\pi ^2\delta _h^2(3000/h)^{3+n}\mathrm{Mpc}^{3+n}.$$ (4) The Abell-ACO power spectrum is given by the matter power spectrum at $`z=0`$ multiplied by a linear and scale independent cluster biasing parameter $`b_{cl}`$, which we include as a free parameter $$P_{A+ACO}(k)=b_{cl}^2P(k;0).$$ (5) For a given set of parameters $`n`$, $`\mathrm{\Omega }_m`$, $`\mathrm{\Omega }_b`$, $`h`$, $`\mathrm{\Omega }_\nu `$, $`N_\nu `$ and $`b_{cl}`$ the theoretical value of $`P_{A+ACO}(k_j)`$ can now be calculated for each observed scale $`k_j`$. Let’s denote these values by $`y_j`$ ($`j=1,\mathrm{},13`$). The dependence of position and amplitude of the first acoustic peak in the CMB power spectrum on cosmological parameters has been investigated using the public code CMBfast by Seljak & Zaldarriaga. As expected, these characteristics are independent on the hot dark matter content. We determine the values $`\mathrm{}_p`$ and $`A_p`$ for given parameters ($`n`$, $`h`$, $`\mathrm{\Omega }_b`$ and $`\mathrm{\Omega }_\mathrm{\Lambda }`$) on a 4-dimensional grid for parameter values in between the grid points we determine $`\mathrm{}_p`$ and $`A_p`$ by linear interpolation. We denote $`\mathrm{}_p`$ and $`A_p`$ by $`y_{14}`$ and $`y_{15}`$ respectively. The theoretical values of the other experimental constraints are obtained as follows: The density fluctuation $`\sigma _8`$ is calculated according to $$\sigma _8^2=\frac{1}{2\pi ^2}_0^{\mathrm{}}k^2P(k;0)W^2(8\mathrm{M}\mathrm{p}\mathrm{c}k/h)𝑑k,$$ (6) with $`P(k;z)`$ from Eq. (3). We set $`y_{16}=\sigma _8\mathrm{\Omega }_m^{0.460.09\mathrm{\Omega }_m}`$ and $`y_{17}=\sigma _8\mathrm{\Omega }^\alpha `$, where $`\alpha =0.24`$ for $`\mathrm{\Omega }_\mathrm{\Lambda }=0`$ and $`\alpha =0.29`$ for $`\mathrm{\Omega }_\mathrm{\Lambda }>0`$, respectively. The r.m.s. peculiar velocity of galaxies in a sphere of radius $`R=50h^1`$Mpc is $$V_{50}^2=\frac{1}{2\pi ^2}_0^{\mathrm{}}k^2P^{(v)}(k)e^{k^2R_f^2}W^2(50\mathrm{M}\mathrm{p}\mathrm{c}k/h)𝑑k,$$ (7) where $`P^{(v)}(k)`$ is the density-weighted power spectrum for the velocity field , $`W(50\mathrm{M}\mathrm{p}\mathrm{c}k/h)`$ is a top-hat window function, and $`R_f=12h^1`$Mpc is the radius of a Gaussian filter used for smoothing of the raw data. For the scales considered $`P^{(v)}(k)(\mathrm{\Omega }^{0.6}H_0)^2P(k;0)/k^2`$. We denote the r.m.s. peculiar velocity by $`y_{18}`$. The value of the r.m.s. linear density perturbation from the formation of Ly-$`\alpha `$ clouds at redshift $`z`$ and scale $`k_F`$ is given by $$\sigma _F^2(z)=\frac{1}{2\pi ^2}_0^{\mathrm{}}k^2P(k;z)e^{(k/k_F)^2}𝑑k.$$ (8) We set $`\sigma _F^2(z=3)y_{19}`$. The value of $`\mathrm{\Delta }_\rho ^2(k_p,z)`$ and the slope $`n(z)`$ are obtained from the linear power spectrum $`P(k;z)`$ by Eq. (1) and Eq. (2) at $`z=2.5`$ and $`k_p=0.008H(z)/(1+z)(\mathrm{km}/\mathrm{s})^1`$, and are denoted by $`y_{20}`$ and $`y_{21}`$ accordingly. For all tests except Gnedin’s Ly-$`\alpha `$ test we use the density weighted transfer function $`T_{cb\nu }(k,z)`$ from . For $`\sigma _F`$ the function $`T_{cb}(k,z)`$ is used according to the prescription given by Gnedin . Note, however, that even in the model with maximal $`\mathrm{\Omega }_\nu `$ ($`0.2`$) the difference between $`T_{cb}(k,z)`$ and $`T_{cb\nu }(k,z)`$ is less than $`12\%`$ for $`kk_p`$. Finally, the values of $`\mathrm{\Omega }_bh^2`$ and $`h`$ are denoted by $`y_{22}`$ and $`y_{23}`$ respectively. Under the assumption that the errors on the data points are Gaussian, the deviations of the theoretical values from their observational counterparts can be characterized by $`\chi ^2`$: $$\chi ^2=\underset{j=1}{\overset{23}{}}\left(\frac{\stackrel{~}{y}_jy_j}{\mathrm{\Delta }\stackrel{~}{y}_j}\right)^2,$$ (9) where $`\stackrel{~}{y}_j`$ and $`\mathrm{\Delta }\stackrel{~}{y}_j`$ are the experimental data and their dispersions, respectively. The set of parameters $`n`$, $`\mathrm{\Omega }_m`$, $`\mathrm{\Omega }_b`$, $`h`$, $`\mathrm{\Omega }_\nu `$, $`N_\nu `$ and $`b_{cl}`$ are then determined by minimizing $`\chi ^2`$ using the Levenberg-Marquardt method . The derivatives of the predicted values w.r.t the search parameters required by this method are calculated numerically using a relative step size of $`10^5`$. This method has been tested and has proven to be reliable, independent on the initial values of parameters and it has good convergence. ## 3 Results The determination of the parameters $`n`$, $`\mathrm{\Omega }_m`$, $`\mathrm{\Omega }_b`$, $`h`$, $`\mathrm{\Omega }_\nu `$, $`N_\nu `$ and $`b_{cl}`$ by the Levenberg-Marquardt $`\chi ^2`$ minimization method is realized in the following way: we vary the set of parameters $`n`$, $`\mathrm{\Omega }_m`$, $`\mathrm{\Omega }_b`$, $`h`$, $`\mathrm{\Omega }_\nu `$ and $`b_{cl}`$ and find the minimum of $`\chi ^2`$, using all observational data described in previous section. Since the $`N_\nu `$ is discrete, we repeat the procedure three times for $`N_\nu `$=1, 2, and 3. The lowest of the three minima is the minimum of $`\chi ^2`$ for the complete set of free parameters. The results are presented in the Table 1. The errors in the determined parameters are the square roots of diagonal elements of the covariance matrix of standard errors. More information about the accuracy of the determination of parameters and their sensitivity to the data used can be obtained from the contours of confidence levels presented in Fig. 1 for the tilted $`\mathrm{\Lambda }`$MDM model with parameters from Table 1 (case $`N_\nu =1`$). These contours show the confidence regions which contain 68.3% (solid line), 95.4% (dashed line) and 99.73% (dotted line) of the total probability distribution in the two dimensional sections of the six-dimensional parameter space, if the probability distribution is Gaussian. Since the number of degrees of freedom is 7 they correspond to $`\mathrm{\Delta }\chi ^2=`$8.2, 14.3 and 21.8 respectively. The parameters not shown in a given diagram are set to their best-fit value. As one can see in Fig.1a the iso-$`\chi ^2`$ surface is rather prolate from the low-$`\mathrm{\Omega }_m`$ \- high-$`n`$ corner to high-$`\mathrm{\Omega }_m`$ \- low-$`n`$. This indicates a degeneracy in $`n\mathrm{\Omega }_m`$ parameter plane. Within the $`1\sigma `$ the ’maximum likelihood ridge’ in this plane can be approximated by the equation $`n\sqrt{\mathrm{\Omega }_m}=0.73`$. A similar degeneracy is observed in the $`\mathrm{\Omega }_\nu \mathrm{\Omega }_m`$ plane in the range $`0\mathrm{\Omega }_\nu 0.17`$, $`0.25\mathrm{\Omega }_m0.6`$ (Fig.1c). The equation for the ’maximum likelihood ridge’ or ’degeneracy equation’ has here the form: $`\mathrm{\Omega }_\nu =0.0230.44\mathrm{\Omega }_m+1.3\mathrm{\Omega }_m^2`$. The important question is: which is the confidence limit of each parameter marginalized over the other ones. The straightforward answer is the integral of the likelihood function over the allowed range of all the other parameters. But for a 6-dimensional parameter space this is computationally time consuming. Therefore, we have estimated the 1$`\sigma `$ confidence limits for all parameters in the following way. By variation of all parameter we determine the 6-dimensional $`\chi ^2`$ surface which contains 68.3% of the total probability distribution. We then project the surface onto each axis of parameter space. Its shadow on the parameter axes gives us the 1$`\sigma `$ confidence limits on cosmological parameters. For the best $`\mathrm{\Lambda }`$MDM model with one sort of massive neutrinos the 1$`\sigma `$ confidence limits on parameters obtained in this way are presented in Table 2. It must be noted that, using the observational data described above, the upper $`1\sigma `$ edge for $`h`$ is equal 1.08 when we marginalized over all other parameters. But this contradicts the age of the oldest globular clusters $`t_0=13.2\pm 3.0`$. Thus we have included this value into the marginalization procedure for the upper limit of $`h`$. We then have 8 degrees of freedom (24 data points) and the 6-dimensional $`\chi ^2`$ surface which contains 68.3% of the probability is confined by the value 13.95. We did not use the age of oldest globular cluster for searching of best fit parameters in general case because it is only a lower limit for age of the Universe. The errors given in Table 2 represent 68% likelihood, of course, only when the probability distribution is Gaussian. As one can see from Fig.1 (all panels without degeneracy) the ellipticity of the likelihood contours in most of planes is close to what is expected from a Gaussian distribution. This indicates that around their maxima the likelihood functions are close to Gaussian. However, the asymmetry of the error bars obtained, shows that away from the maxima this is no longer the case. Therefore, our estimates of the confidence limits have to be taken with a grain of salt. The errors define the range of each parameter within which the best-fit values obtained for the remaining parameters lead to $`\chi _{min}^212.84`$. Of course, the best-fit values of the remaining parameters lay within the range given in Table 2. However clearly not every set of parameters from these ranges satisfies the condition, $`\chi _{min}^212.84`$. For example, standard CDM model ($`\mathrm{\Omega }_m=1`$, $`h=0.5`$, $`\mathrm{\Omega }_b=0.05`$, $`n=1`$ and best-fit value of cluster biasing parameter $`b_{cl}=2.17`$ ($`\sigma _8=1.2`$)) has $`\chi _{min}^2=142`$ (!), which excludes it at very high confidence level, $`>99.999\%`$. When we use the baryon density inferred from nucleosynthesis ($`h^2\mathrm{\Omega }_b=0.019`$ ($`b_{cl}=2.25`$, $`\sigma _8=1.14`$)) the situation does not improve much, $`\chi _{min}^2=112`$. Furthermore, even if we leave $`h`$ as free parameter we still find $`\chi _{min}^2=16`$ ($`>1\sigma `$) with the best-fit values $`h=0.37`$ and $`b_{cl}=3.28`$ ($`\sigma _8=0.74`$); this variant of CDM is ruled out by direct measurements of the Hubble constant. The standard MDM model ($`\mathrm{\Omega }_m=1`$, $`h=0.5`$, $`\mathrm{\Omega }_b=0.5`$, $`n=1`$, $`\mathrm{\Omega }_\nu =0.2`$, $`N_\nu =1`$ with a best value of the cluster biasing parameter $`b_{cl}=2.74`$ ($`\sigma _8=0.83`$)) does significantly better: it has $`\chi _{min}^2=23.1`$ ($`99\%`$ C.L.) which is out of the $`2\sigma `$ confidence contour but inside $`3\sigma `$. With the nucleosynthesis constraint the situation does not change: $`\chi _{min}^2=22`$; also if we leave $`h`$ as free parameter: $`\chi _{min}^2=21`$, $`h=0.48`$. But if, in addition, we let vary $`\mathrm{\Omega }_\nu `$, we obtain $`\chi _{min}^2=13`$ with best-fit values of $`\mathrm{\Omega }_\nu =0.09`$, $`h=0.43`$, $`b_{cl}=3.2`$ ($`\sigma _8=0.73`$). This means that the model is ruled out by the data set considered in this work at $`70\%`$ confidence level only. But also here the best-fit value for $`h`$ is very low. If we fix it at lower observational limit $`h=0.5`$ then $`\chi _{min}^2=18.9`$ (the best fit values are: $`\mathrm{\Omega }_\nu =0.15`$, $`b_{cl}=2.8`$ ($`\sigma _8=0.83`$)), which corresponds to a confidence level of 95% . Therefore, we conclude that the observational data set used here rules out CDM models with $`h0.5`$, a scale invariant primordial power spectrum ($`n=1`$) and $`\mathrm{\Omega }_k=\mathrm{\Omega }_\mathrm{\Lambda }=0`$ at very high confidence level, $`>99.99\%`$. MDM models with $`h0.5`$, $`n=1`$ and $`\mathrm{\Omega }_k=\mathrm{\Omega }_\mathrm{\Lambda }=0`$ are ruled out at $`95\%`$ C.L. One can see the model with one sort of massive neutrinos provides the best fit to the data, $`\chi _{min}^24.6`$. Note, however, that there are only marginal differences in $`\chi _{min}^2`$ for $`N_\nu =1,2,3`$. Therefore, with the given accuracy of the data we cannot conclude whether – if massive neutrinos are present at all – their number of species is one, two, or three. The number of degrees of freedom is $`N_F=N_{\mathrm{exp}}N_{\mathrm{par}}=7`$. The $`\chi _{min}^2`$ for all cases is within the expected range, $`N_F\sqrt{2N_F}\chi _{\mathrm{min}}^2N_F+\sqrt{2N_F}`$ for the given number of degrees of freedom. This means that the cosmological paradigm which has been assumed is consistent with the data. One important question is how each point of the data influences our result. To estimate this we have excluded some data points from the searching procedure. Excluding any part of observable data results only in a change of the best-fit values of $`n`$, $`\mathrm{\Omega }_m`$ and $`h`$ within the range of their corresponding standard errors. This indicates that the data are mutually in agreement, implying consistent cosmological parameters (within the still considerable error bars). The small scale constraints, the Ly-$`\alpha `$ tests reduce the hot dark matter content from $`\mathrm{\Omega }_\nu 0.22`$ to $`0.075`$. The $`\sigma _8`$-tests further reduce $`\mathrm{\Omega }_\nu `$ to $`0.06`$. Including of the Abell-ACO power spectrum in the search procedure, tends to enhance $`\mathrm{\Omega }_\nu `$ slightly. The most crucial test for the baryon content is of course the nucleosynthesis constraint. Its $`6\%1\sigma `$-accuracy safely keeps $`h^2\mathrm{\Omega }_b`$ near its median value 0.019. The parameter $`\mathrm{\Omega }_b`$ in turn is only known to $`36\%`$ accuracy due to the large errors of other experimental data used here, especially of the Hubble constant. The accuracy of $`h`$ ($`17\%`$) is better than the one assumed from direct measurements, $`23\%`$. Summarizing, we conclude that all data points used here are important for searching the best-fit cosmological parameters and do not contradict each other. Up to this point we ignored the uncertainties in the COBE normalization. Indeed, the statistical uncertainty of the fit to the four-year COBE data, $`\delta _h`$, is 7% (1$`\sigma `$) and we want to study how this uncertainty influences the accuracy of cosmological parameters which we determine? Varying $`\delta _h`$ in the 1$`\sigma `$ range we found that the best-fit values of all parameters except $`\mathrm{\Omega }_\nu `$ do not vary by more than 2% from the values presented in Table 1. Only $`\mathrm{\Omega }_\nu `$ varies in a range of 12% . These uncertainties are significantly smaller than the standard errors given in Table 1 and neglecting them is thus justified. The normalization constant $`A`$, which for best model with one species of massive neutrinos, $`A=4.6810^7(h^1\text{ Mpc })^{3+n}`$, varies in a range of 8%, and not 14% as one might expect from (4) at the first sight. The reason is that variation of $`\delta _h`$ is somewhat compensated by correlated variation of $`n`$ and $`h`$. Moreover, if we disregard the COBE normalization and treat the normalization constant $`A`$ as a free parameter to be determined like the others, its best-fit value becomes $`4.8210^7(h^1\text{ Mpc })^{3+n}`$ (for $`N_\nu =1`$), consistent with COBE normalization. The best-fit values of the other parameters correspondingly do not vary substantially: $`n=1.09`$, $`\mathrm{\Omega }_m=0.40`$, $`\mathrm{\Omega }_\nu =0.052`$, $`\mathrm{\Omega }_b=0.041`$, $`h=0.68`$ and $`b_{cl}=2.19`$ (this is less than 5% except for $`\mathrm{\Omega }_\nu `$, which is reduced by $`12\%`$). This implies that determinations of the amplitude of scalar fluctuations by the COBE measurement of the large scale CMB anisotropies and by large scale structure data at much smaller scales are in good agreement. It also indicates that a possible tensor component in the COBE data cannot be very substantial. ## 4 Conclusions We summarize, that the observational data of the LSS of the Universe considered here can be explained by a tilted $`\mathrm{\Lambda }`$MDM inflationary model without tensor mode. The best fit parameters are: $`n=1.12\pm 0.09`$, $`\mathrm{\Omega }_m=0.41\pm 0.11`$, $`\mathrm{\Omega }_\nu =0.06\pm 0.028`$, $`\mathrm{\Omega }_b=0.039\pm 0.014`$ and $`h=0.70\pm 0.12`$. All predictions of measurements are close to the experimental values given above and within the error bars of the data. The CDM density parameter is $`\mathrm{\Omega }_{cdm}=0.31\pm 0.12`$ and $`\mathrm{\Omega }_\mathrm{\Lambda }`$ is moderate, $`\mathrm{\Omega }_\mathrm{\Lambda }=0.59\pm 0.11`$. The neutrino matter density corresponds to a neutrino mass $`m_\nu =94\mathrm{\Omega }_\nu h^22.7\pm 1.2`$ eV. The value of the Hubble constant is close to the measurements by Madore et al.. The age of the Universe for this model equals 12.3 Gyrs which is in good agreement with the age of the oldest objects in our galaxy . The spectral index coincides with the COBE prediction. The relation between the matter density $`\mathrm{\Omega }_m`$ and the cosmological constant $`\mathrm{\Omega }_\mathrm{\Lambda }`$ agrees well with the independent measurements of cosmic deceleration and global curvature based on the SNIa observation. The 1$`\sigma `$ (68.3%) confidence limits on each cosmological parameter, obtained by marginalizing over the other parameters, are $`0.82n1.39`$, $`0.19\mathrm{\Omega }_m1`$, $`0\mathrm{\Omega }_\mathrm{\Lambda }0.81`$, $`0\mathrm{\Omega }_\nu 0.17`$, $`0.021\mathrm{\Omega }_b0.13`$ and $`0.38h0.85`$. The observational data set used here rules out the CDM models with $`h0.5`$, scale invariant primordial power spectrum $`n=1`$ and $`\mathrm{\Omega }_\mathrm{\Lambda }=\mathrm{\Omega }_k=0`$ at very high confidence level, $`>99.99\%`$. Also pure MDM models are ruled out at $`95\%`$ C.L. It is remarkable also that this data set determines the value of normalization constant for scalar fluctuations which approximately equals the value deduced from COBE four-year data. It indicates that a possible tensor component in the COBE data cannot be very substantial. The coincidence of the values of cosmological parameters obtained by different methods indicates that a wide set of cosmological measurements are correct and that their theoretical interpretation is consistent. However, we must also note that the accuracy of present observational data on the large scale structure of the Universe is still insufficient to determine a set of cosmological parameters with high accuracy.
warning/0002/nucl-th0002020.html
ar5iv
text
# 1 Introduction ## 1 Introduction Results of the CERES experiment show a significantly higher dilepton yield in the invariant mass region between about 200 and 600 MeV compared to that from a hadronic cocktail based on known reaction channels. After adding in the – expected – secondary $`\pi \pi \rho `$ yield the observed yield is still underestimated by a hadronic cocktail by a ’sensitivity factor’ of about 3. While there is also an attempt to explain these observations in terms of standard ’classical’ hadronic sources by exploiting their experimental uncertainties, most theoretical explanations invoke a change of the properties of the vector meson masses and widths. Some of these are based on ’classical’ collision broadening whereas others involve $`\rho a_1`$ mixing as a precursor to chiral symmetry restoration . However, all these attempts have in common that in a first step they determine the in-medium properties in hot and dense equilibrium and then use these locally either in transport calculations or in more simplified expansion scenarios. The latter procedure is obviously not a-priori correct for an ultrarelativistic heavy-ion collision which – at least in its early stages – proceeds far away from equilibrium. It is worthwhile to realize that even in ultrarelativistic heavy-ion collisions with their high peak densities about 60 % of all dileptons are produced at densities below $`2\rho _0`$ (see Fig. 7 in ). This is so because the observed yield inherently contains an integration over the whole reaction history and thus the late stages of the reaction with high pion densities contribute via $`\pi ^++\pi ^{}e^++e^{}`$ a significant part to the total observed dilepton yield. It is, therefore, intriguing to ask if other reactions involving more elementary probes on nuclei that proceed closer to equilibrium can be used to investigate the question of in-medium changes of hadronic properties. While the density in such reactions obviously is always below $`\rho _0`$, this disadvantage may be overcome by the ’cleaner’ and more stable environment in which dilepton production proceeds in such reactions. In this talk I show that indeed pion- and photon-induced dilepton production on nuclei is as sensitive to possible in-medium changes as are heavy-ion reactions. I will also – in the last part of this talk – show how in-medium changes of vector mesons are expected to show up not only in dilepton spectra, but also in photo-absorption processes. Such a connection, if it can be firmly established, clearly adds to the consistency of our picture of in-medium changes of hadrons. Details to all the points I am going to discuss in this talk can be found in a number of recent publications and in particular in the PhD thesis of Martin Effenberger . ## 2 Model The calculations shown in this talk are all based on a new development in transport theory that allows one to also tranport broad resonances. While former calculations (for a review see ) had to employ simplifiying assumptions the calculations of Effenberger et al. , for the first time, implemented a transport-theoretical treatment of broad mesons and treated the widths of the vector mesons consistently. The theoretical breakthrough here was to transport not the phase-space density $`f(\stackrel{}{x},\stackrel{}{p},t)`$, but a ’spectral phase space density’ $`F(\stackrel{}{x},\stackrel{}{p},\mu ,t)`$ that contains also information about the spectral function of the resonance. In (see also ) we have discussed in detail that a consistent formulation of the scattering term, that takes the spectral function of the particles into account, puts automatically intrinsically broad, collision-broadened resonances (such as, e.g. the $`\rho `$ meson) – because of their frequent decay and reformation as they traverse the nucleus – back on their free spectral function when they leave the nucleus. A problem arises, however, when particles with intrinsically sharp spectral functions, such as the nucleon or the $`\omega `$ meson, get collision broadened. In such a case the repetitive decay and reformation is not frequent enough so that these particles would emerge from the collision still collision-broadened. In order to surpress this unphysical behavior we introduced an ad-hoc potential that drives all particles back on mass-shell when they leave the nucleus . It is amusing that this potential, that we intuitively guessed in ref. , can – for sharp resonances, where $`\mathrm{\Gamma }_{tot}\mathrm{\Gamma }_{coll}`$ – actually be derived from the Kadanoff-Baym equations . The results presented here are based on this method (for more details see ). The equation of motion for the spectral phase space distribution reads $$\left(\frac{}{t}+\stackrel{}{}_pH_i\stackrel{}{_x}\stackrel{}{}_xH_i\stackrel{}{}_p\right)F_i=G_iA_iL_iF_i$$ (1) with the spectral phase-space-density $$F_i((\stackrel{}{x},\stackrel{}{p},\mu ,t)=A_i(\stackrel{}{x},\stackrel{}{p},\mu ,t)f(\stackrel{}{x},\stackrel{}{p},\mu ,t),$$ (2) the spectral function $$A_i(\mu )=\frac{2}{\pi }\frac{\mu ^2\mathrm{\Gamma }_{\mathrm{tot}}(\mu )}{\left(\mu ^2M_i^2\right)^2+\mu ^2\mathrm{\Gamma }_{\mathrm{tot}}^2(\mu )},$$ (3) and the usual phase-space density $`f(\stackrel{}{x},\stackrel{}{p},\mu ,t)`$. As stressed in particular by Knoll it is essential to maintain the consistency condition that the total width appearing in the spectral function is related to the loss-term $`L`$ in the transport equation $$\mathrm{\Gamma }_{\mathrm{tot}}=\gamma L.$$ (4) Since this condition presents a major consistency problem it has been fulfilled in the calculations reported here only for the vector mesons. In it was shown that the two-body collisional width does not broaden the nucleon resonances significantly. For all further details I refer the reader to . ## 3 Pion-Induced Dileptons With the availability of a $`\pi ^{}`$ beam at GSI and the detectorsystem HADES it will be possible to explore dilepton production in pion-induced reactions on nuclei. We have, therefore, performed extensive studies of such reactions . Input to these calculations are cross sections predicted by the Manley coupled channel analysis of $`\pi N`$ data on which our analysis is based. The transport calculations allow one to follow the collision history of all vector mesons produced. With realistic $`VN`$ cross sections we obtain a collision width $`\mathrm{\Gamma }_{\mathrm{coll}}=\rho \sigma _{VN}v_{VN}`$ of up to 500 MeV for the $`\rho `$ meson and 70 MeV for the $`\omega `$, both at normal nuclear matter density . The overall sensitivity to these effects is about a factor 2, which is close to the expected systematic uncertainty of these transport calculations. However, this sensitivity can be enhanced by performing appropriate cuts on the dilepton cm momenta. Fig. 1 shows that with a cut that selects only slow vector mesons the sensitivity to various in-medium scenarios, in particular for the $`\omega `$ meson, is clearly strong enough to be seen experimentally. ## 4 Photon-Induced Dileptons Pions have one possible disadvantage for such in-medium studies: they experience a strong initial-state interaction and thus a pion-beam illuminates only a part of the nucleus. This is visible in Fig. 2 which shows in the two lower figures the location of the first collision of the pion with the nucleons of the target (left) and all meson-baryon collisions (right). The upper two figures give the same information for a photon beam. The left part shows clearly how the photon illuminates the whole nuclear volume and not just the surface. Subsequent collisions remember this initial state as shown in the right part of this figure. Thus, we expect that photon-induced experiments will be even more sensitive to in-medium changes of hadronic properties than experiments with pions. That the theory, with the proper, consistent treatment of in-medium broadening describes the ’normal’ photonuclear processes very well is shown in a comparison with data for photo-pion, photo-2pion and photon-eta production data at energies up to 800 MeV . We are thus quite confident that we have these photonuclear processes well under control and can thus use these methods to predict and investigate the sensitivity of photonuclear dilepton production data to in-medium changes of vector meson properties. Photon-induced dilepton data have the problem that they necessarily contain a big contribution from the so-called Bethe-Heitler process in which the incoming photon radiates already a dilepton pair before any coupling to the hadrons takes place. As is well known, however, this contribution can be surpressed by proper kinematical cuts , both for the incoherent and the coherent part . Fig. 3 then shows the sensitivity of the dilepton spectra to various in-medium changes of the vector mesons. It is immediately apparent, that this sensitivity is just as large as for the ultrarelativistic heavy-ion collisions. In particular, again the $`\omega `$ properties are reflected quite clearly in the (still to be measured) dilepton spectra. As in the case of pions one can enrich the in-medium sensitivity by making a cut on the dilepton cm momentum. ## 5 Photoabsorption All the in-medium effects on vector meson spectral functions are connected with a significant broadening and corresponding shift of strength to lower masses. This has an interesting consequence for the decay of the N(1520) resonance that I have already pointed out in in 1997. According to the particle data group listing, which is based on Manley’s analysis in this respect, this $`D_{13}`$ resonance decays on resonance with a probability of 15 - 25 % into $`N\rho `$ and this, even though this decay is energetically closed if one works with the pole mass of the (broad) $`\rho `$. Thus the decay can proceed only through the low-mass tails of the $`\rho `$-meson spectral function; the rather large partial decay width in turn indicates a large coupling constant for the vertex $`N(1520)N\rho `$. If the $`\rho `$ meson’s spectral function in nuclear matter indeed gains strength at low masses then the partial $`\rho `$ decay width of the $`N(1520)`$ resonance would increase dramatically. This effect may already have been experimentally seen. The data for photoabsorption on nuclei exhibit a universal behavior that – beginning with rather small mass numbers $`A`$ – scales with $`A`$. This universal cross section exhibits the $`\mathrm{\Delta }`$ resonance – though somewhat broadened –, but the second and third resonance regions around 1500 and 1650 MeV, respectively, that are well seen in photoabsorption on the nucleon, have disappeared in nuclei . Fermi-motion alone is the more effective in smearing out resonance strength the higher the energy of the resonance is. This alone explains the absence of the third resonance region . The second resonance region, around 1500 MeV, however, survives the effects of Fermi-smearing. We have shown in that this finds a natural explanation in the opening of the $`\rho `$-decay channel of the $`N(1520)`$ resonance. With $`\rho `$ strength moving in medium down to lower masses the formerly nearly closed decay opens up dramatically thus leading to a very large (several hundred MeV) total width of this resonance. Crucial for this argument is the high partial decay width of the $`D_{13}`$ resonance into the $`N\rho `$ channel. This has so far, not been directly verified by experiment; the cited large partial decay width is based only on a theoretical coupled channel analysis . Although it is gratisfying to see that a similar result also emerges now from the CC calculations of Friman, Lutz and Wolf , a direct experimental verification of this decay branch would necessitate a detailed partial-wave analysis of ($`\gamma ,2\pi `$) reactions on the nucleon which so far does not exist. It is, however, encouraging, that the DAPHNE collaboration and the TAPS collaboration , both at MAMI, have found experimental evidence for pion invariant mass spectra in such reactions that show clear deviations from phase-space. ## 6 Summary In this talk I have shown that pion- and photon-induced reactions show a sensitivity to in-medium changes of hadrons that is as large as that observed in ultrarelativistic heavy ion collisions. This result was originally unexpected because of the much lower baryon densities probed in such elementary reactions on nuclei ($`<1`$), compared with those obtained in ultrarelativistic heavy-ion collisions ($`>5`$). It finds its explanation in the fact that all experimentally observed dilepton spectra contain an integration over the whole reaction history, which, in the case of heavy-ions, runs from $`1\rho _0`$ to more than $`5\rho _0`$ and then back down again to nearly $`0\rho _0`$ while the pion- or photon-induced reaction on nuclei always proceeds close to $`\rho _0`$, i.e. much closer to equilibrium. In my opinion, an unequivocal identification of the origins of changes of hadronic spectral functions observed in ultrarelativistic heavy-ion collisions will not be possible as long as the more elementary reactions on nuclei have not been studied as well. It is obvious that hadronic in-medium effects, if they are observed in such reactions at or below $`\rho _0`$, have nothing to do with any transition to the QGP and even a possible connection with chiral symmetry restoration, a very popular keyword in this field, will be hard to establish (of course, duality arguments can always be used to keep this connection alive!). What we can learn, however, are interaction rates of hadrons with baryons, also for unstable hadrons, and this alone will be an exciting prospect. Experiments with HADES using the pion beam at GSI and experiments with photon beams, starting now with TAPS at MAMI, will help us to clarify the issue of in-medium changes of hadrons. For example, a close comparison of dilepton spectra obtained in pion-induced reactions with those obtained with photons should allow one to sample quite different density regions and thus in-medium effects, in particular for the $`\omega `$ meson. This is illustrated without further words in Fig. 4. This talk is mainly based on the PhD thesis work of Martin Effenberger to whom I am grateful for many years of exciting collaboration. The work has been supported by BMBF, DFG and GSI Darmstadt.
warning/0002/hep-ph0002051.html
ar5iv
text
# 1 Example of a diagram for quantum electrodynamics, which makes a contribution ∼𝑁! to the 𝑁th order of perturbation theory []. ## 1. There have been numerous semi-intuitive assertions which reduce to the notion that instantons do not exhaust all of physics. This thesis is correct as long as it is understood correctly, but in the present case it is not relevant. Historically, instantons first appeared when the saddle-point approximation was employed in the original integral (1). It was substantiated only in a narrow region of parameters, and thus instantons did not, in fact, exhaust all of physics. In the Lipatov technique the situation changed, because the saddle-point approximation is used not in the integral (1), but in the expression (3) for the expansion coefficients. Since only large values of $`N`$ are considered, only a limited role is assigned to instantons from the onset; however, the saddle-point method is now always applicable, <sup>5</sup><sup>5</sup>5 Of course, instantons exist only in a part of the region of parameters, but this is not a restriction in the Lipatov technique: the values of $`a`$, $`b`$, and $`c`$ in the asymptotics (5) are calculated exactly, and they allow analytic continuation as functions of the physical parameters. and there is a basis to assume that everything is determined by instantons. Let us illustrate the foregoing statements in the example of the Schrödinger equation with a random potential $`V(x)`$: $$[\widehat{p}^2/2m+V(x)]\mathrm{\Psi }(x)=E\mathrm{\Psi }(x).$$ $`(15)`$ At large negative values of $`E`$ its eigenfunctions (Fig. 5) are localized on the infrequent fluctuations of the random potential (a and b), at large positive values of $`E`$ they are similar to plane waves (d and e), and in the vicinity of the bare spectrum edge at $`E=0`$ they have a highly broken, fractal character (c). The problem of investigating Eq. (15) can be reformulated in the language of an effective field theory, viz., $`\phi ^4`$ theory with the “incorrect” sign for $`g`$ (Refs. and ). In this case the typical wave functions of localized states are described by instantons. The changes in the situation observed as $`E`$ increases can be described in the following manner in terms of instantons: at first instantons have a small radius and a sparse distribution, i.e., they form an ideal gas (a); then the radius of the instantons increases, and their density rises, i.e., interactions between them appear (b); then condensation of the instantons takes place (c), and an instanton crystal forms (d and e). Only the case in Fig. 5,a corresponds to applicability of the saddle-point method in an integral of the type (1), and thus the standard instanton approximation is very poor. Let us examine this situation from the standpoint of perturbation theory with respect to the random potential $`V(x)`$. An ideal instanton crystal (e) corresponds to a plane wave, i.e., the zeroth order of perturbation theory. In a nonideal crystal (d) the higher orders have an increased role; in the vicinity of the bare spectrum edge (c) all the diagrams are of the same order of magnitude, so that the high and low orders of perturbation theory are equally significant. In the region of localized states (a and b) the dominant role shifts to the high orders: these states are not manifested in any finite order of perturbation theory, and discarding the low-order contributions does not influence their properties in any way. We see that Lipatov’s conception (high orders are determined by instantons) fits excellently into the existing physical picture. Thus, the status of instantons in the integral (1) and the integral (3) differs significantly. In our opinion, this accounts for the position taken by ’t Hooft, since he is a classic on instantons specifically in the original integral (1). ## 2. Relationship to the logarithmic situation . Renormalons exist only in renormalizable theories, but not in super-renormalizable theories. If a theory is super-renormalizable, an upper bound of the type $`a^Ng^N`$ can be obtained for the contribution of an individual diagram, and the appearance of the factor $`N!`$ in the asymptotics (5) can be associated only with a factorially large number of diagrams. Renormalons and, thus, a new mechanism for the appearance of factorial contributions appear in renormalizable theories. It can be expected that this mechanism is associated with the formation of the tails in the integral (3) and is not taken into account in the Lipatov technique. In this argument everything except the last conclusion is correct. We can illustrate this in the case of $`\phi ^4`$ theory, which is renormalizable for $`d=4`$ and super-renormalizable for $`d<4`$. Among the large set of integrations concealed in the symbol $`D\phi `$ in the integral (3), we can single out one for which the limit $`d4`$ is associated with qualitative changes: it is the integration over the instanton radius $`R`$ (Fig. 6). For $`d`$ significantly smaller than 4 (for example, $`d=3`$), the integrand $`\mathrm{exp}(S\{\phi \})`$ has a sharp maximum as a function of $`R`$ and allows saddle-point integration; when $`d=4ϵ`$ the maximum becomes gently sloping, and when $`d=4`$ the instanton action $`S\{\phi _c\}`$ does not depend on $`R`$ at all. In the latter case the integral diverges, leading to the logarithmic situation. We see (see the curve for $`d=4ϵ`$ in Fig. 6) that the “activation” of renormalon contributions is, in fact, related to the appearance of slow tails in the integral (3), but these tails are taken into account in the Lipatov technique . One can be puzzled by a following question. If the Lipatov technique is based on a saddle-point method, then how can it cover the definitely nonsaddle-point situation for $`d=4ϵ`$? The fact is that the saddle point in a functional integral practically never reduces to a simple maximum achieved at a single point: the maximum is degenerate in a certain space of finite dimensionality. Accordingly, a finite number of integrations should be performed exactly, rather than in the saddle-point approximation. However, if the integration is performed exactly over a certain variable (for example, $`R`$), it is of no significance whether the degeneracy is exact ($`d=4`$) or approximate ($`d=4ϵ`$). Nevertheless, in the latter case technical difficulties arise, and the corresponding methods (constrained instantons ) have been poorly developed hitherto . It is thus clear that even in cases where slow tails actually appear in the integral (3), the Lipatov procedure is sufficiently flexible and contains broad possibilities for dealing with them. ## 3. The limit $`n\mathrm{}`$. There is an opinion that the significance of renormalon contributions can easily be proved by treating the $`n`$-component $`\phi ^4`$ theory in the limit $`n\mathrm{}`$ (and the analogous models in QCD and electrodynamics) : the factor $`n`$ corresponds to a closed loop, and renormalon graphs containing the maximum possible number of loops are singled out by the large parameter $`n`$. Although diagrams of the same order, but with a smaller number of loops, can make comparable contributions at large $`N`$ due to the combinatorial factors, they have a slower dependence on $`n`$; therefore, the renormalons cannot be cancelled identically. This argument is valid in any finite order of $`1/n`$. However, a detailed investigation of the structure of the $`1/n`$-expansion reveals the presence of numerous cancellations, and although the situation cannot be totally elucidated, the question is not resolved on the level of simple arguments of the type indicated. It is not difficult to identify the crux of the problem here. As an example, let us consider the self-energy $`\mathrm{\Sigma }(p,m)`$ of $`\phi ^4`$ theory; it is clear from a diagrammatic analysis for $`m=0`$ and values of the momentum $`p`$ close to the ultraviolet cutoff $`\mathrm{\Lambda }`$ that the $`(N+1)`$th expansion coefficient for $`\mathrm{\Sigma }(p,0)\mathrm{\Sigma }(0,0)`$ has the form of a polynomial in $`n`$ $$p^2\{A_N(N)n^N+A_{N1}(N)n^{N1}+\mathrm{}+A_1(N)n+A_0(N)\},$$ $`(16)`$ in which the coefficient $`A_N(N)`$ is specified by renormalon graphs: $$A_N(N)=\mathrm{const}\left(\frac{1}{16\pi ^2}\right)^NN!.$$ $`(17)`$ If it is assumed that the renormalon graphs are contained in the instanton contribution, the expression (16) should transform into the Lipatov asymptotics at large $`N`$ \[see Eq. (130) in Ref. for $`M=1`$ and $`p\mathrm{\Lambda }`$\]: $$p^2\alpha \beta ^nN^{(n+6)/2}\left(\frac{1}{8\pi ^2}\right)^NN!\left[\mathrm{\Gamma }\left(\frac{n+2}{2}\right)\right]^1\underset{0}{\overset{\mathrm{}}{}}𝑑yy^{(n+5)/3}K_1(y)^2,$$ $`(18)`$ where $`\alpha ,\beta 1`$, and $`K_1(x)`$ is the McDonald function. It is easily seen that an equality between (16) and (18) is impossible when $`n\mathrm{}`$. This is a manifestation of the “noncancelability” of renormalons. However, the usual condition for applicability of the Lipatov technique, $`N1`$ at large $`n`$ is, generally speaking, replaced by a more rigid condition, for example, $`Nn`$, and $`n`$ then has a bound of the type $$n\mathrm{¡}\mathrm{}n_0(N),$$ $`(19)`$ which precludes going to the limit $`n\mathrm{}`$. If it is taken into account that the Lipatov asymptotics has limited accuracy ($`1/N`$ in relative units), the correct formulation of the question is as follows. Can we construct an interpolation polynomial of type (16) with the high-order coefficient (17) which would approximate the function (18) within an assigned accuracy in the interval $`0nn_0(N)`$, where $`n_0(N)\mathrm{}`$ as $`N\mathrm{}`$? The answer to this question is positive (see the Appendix); therefore, the assumption that the renormalon graphs are contained in the instanton contribution does not lead to contradictions. ## 4. Relationship to a Landau pole . It is easy to see (Fig. 2,c) that the summation of a sequence of renormalon diagrams corresponds to “dressing” the interaction. The relationship between the renormalized charge $`g`$ and the bare charge $`g_0`$ is then given by the familiar expression $$g_0=\frac{g}{1\beta _0g\mathrm{ln}(\mathrm{\Lambda }^2/m^2)},$$ $`(20)`$ which contains a pole at the point $$\mathrm{\Lambda }_c^2=m^2e^{1/\beta _0g}.$$ $`(21)`$ Under the literal understanding of this pole in the spirit of the early papers by Landau and Pomeranchuk , a simple physical interpretation can be given to renormalon singularities (Refs. and ). <sup>6</sup><sup>6</sup>6 It is assumed below that $`\beta _0>0`$. For asymptotically free theories, in which $`\beta _0<0`$, similar arguments are valid in regard to so-called infrared renormalons. The latter are obtained from integrals of the type (13) with $`m=1,0,1,2,\mathrm{}`$ in the region of small momenta. The dependence of the perturbation series on the cutoff parameter $`\mathrm{\Lambda }`$ has the structure $$c_1\mathrm{\Lambda }^2+c_0\mathrm{ln}\mathrm{\Lambda }^2+c_1\mathrm{\Lambda }^2+c_2\mathrm{\Lambda }^4+\mathrm{}+c_n\mathrm{\Lambda }^{2n}+\mathrm{}.$$ $`(22)`$ The first two terms are eliminated by a renormalization procedure, and the rest terms, in principle, remain, but vanish in the limit $`\mathrm{\Lambda }\mathrm{}`$. Because of the pole in (20), values of $`\mathrm{\Lambda }`$ greater than $`\mathrm{\Lambda }_c`$ are inaccessible in principle, and unremovable uncertainties of the type $$\mathrm{\Lambda }_c^{2n}=m^2e^{n/\beta _0g}$$ $`(23)`$ appear in the theory. Similar uncertainties are generated by renormalon singularities, whose existence on the positive semiaxis leads to ambiguity in the choice of the integration contour in the Borel integral (6). The contour can be drawn to the right or the left of the $`n`$th singularity (Fig. 7), producing an uncertainty in reconstructing the function from its Borel transform: $$\delta F(g)\underset{zz_n}{}𝑑ze^{z/g}B(z)e^{z_n/g},$$ $`(24)`$ which, with consideration of the equality $`z_n=n/\beta _0`$ coincides with (23). Of course, the literal interpretation of the Landau pole seems archaic, but after some modification of the argument presented, a real meaning can be assigned to it. It is well known , that the dependence of the charge $`g`$ on the distance scale $`\mathrm{\Lambda }^1`$ is given by the equation $$\frac{dg}{d\mathrm{ln}\mathrm{\Lambda }^2}=\beta (g)=\beta _0g^2+\beta _1g^3+\mathrm{},$$ $`(25)`$ whose solution depends drastically on the behavior of the Gell-Mann–Low function $`\beta (g)`$. The pole in (20) is eliminated, if $`\beta (g)`$ changes sign or behaves as $`g^\alpha `$ with $`\alpha <1`$ at large $`g`$. If, on the other hand, $`\beta (g)`$ is positive and increases as $`g^\alpha `$ with $`\alpha >1`$ when $`g\mathrm{}`$, the pole is preserved, and the theory is internally inconsistent because it is impossible to determine $`g(\mathrm{\Lambda })`$ for all $`\mathrm{\Lambda }`$ (Ref. ). In the latter case the position of the pole is given by the equation $$\mathrm{ln}\frac{\mathrm{\Lambda }_c^2}{m^2}=\underset{g}{\overset{\mathrm{}}{}}\frac{dx}{\beta (x)},$$ $`(26)`$ which, for small values of $`g`$, leads to the result $$\mathrm{\Lambda }_c^2=\mathrm{const}m^2e^{1/\beta _0g},$$ $`(27)`$ which differs from (21) only by an insignificant constant factor. Thus, the existence of renormalon singularities seems fairly convincing for internally inconsistent theories. Conversely, there is no reason for them in “good” theories. <sup>7</sup><sup>7</sup>7 In particular, the result (27) is valid when the expansion (25) is truncated at a finite number of terms, provided the polynomial obtained is positive. On these grounds it is easy to draw the erroneous conclusion that the high-order terms of the expansion of the $`\beta `$ function are insignificant. Parisi’s arguments regarding the momentum dependence of Borel transforms exactly follow this line of reasoning. In fact, the character of the solution of Parisi’s equations depends significantly on the behavior of $`\beta (g)`$ at infinity. In particular, they are easily solved for the model function $`\beta (g)=\beta _0g^2/(1+\lambda g)`$ with $`\lambda 1`$ and lead to a result which differs qualitatively from its one-loop analog. Since the behavior of the function $`\beta (g)`$ at $`g\mathrm{¿}\mathrm{}1`$ is unknown, the presence or absence of renormalon singularities is a matter of belief. We stress, however, the following point. Factorial contributions of individual diagrams exist in all field theories in which the expansion of the $`\beta `$ function (25) begins from the quadratic term: then the interaction on the $`k^1`$ scale is described by a formula of the type (20) with the replacement of $`\mathrm{\Lambda }`$ by $`k`$, whose expansion gives $`(\beta _0\mathrm{ln}k^2)^N`$ in the $`N`$th order \[see (13)\]. <sup>8</sup><sup>8</sup>8 The concrete sequences of the renormalon diagrams can differ somewhat in different theories. For example, in $`\phi ^4`$ theory the significant diagrams do not reduce to chains of “bubbles” (Fig. 2,c), but form a so-called parquet . To resolve the question of the internal inconsistency of a theory, one should know all the coefficients in the expansion (25). Therefore, it would be incorrect to consider the formal existence of renormalon contributions as an indication of the internal inconsistency of a theory. 3. ANALYTIC PROPERTIES OF THE BOREL TRANSFORMS OF $`\phi ^4`$ THEORY 3.1. Expansion of the class of Borel transformations For the ensuing treatment it is convenient to expand the class of Borel transformations, setting $$F(g)=\underset{0}{\overset{\mathrm{}}{}}𝑑xe^xx^{b_01}B(gx),$$ $$B(g)=\underset{N=0}{\overset{\mathrm{}}{}}\frac{F_N}{\mathrm{\Gamma }(N+b_0)}g^N$$ $`(28)`$ with the arbitrary parameter $`b_0>0`$, instead of (6) and (7). If $`B(g)`$ and $`\stackrel{~}{B}(g)`$ are the Borel transforms corresponding to the parameters $`b_0`$ and $`b_1`$ (for definiteness, we set $`b_1>b_0`$), it is not difficult to derive the conversion formula: $$\stackrel{~}{B}(g)=\frac{1}{\mathrm{\Gamma }(b_1b_0)}\underset{0}{\overset{\mathrm{}}{}}𝑑x\frac{x^{b_1b_01}}{(1+x)^{b_1}}B\left(\frac{g}{1+x}\right).$$ $`(29)`$ We define the analyticity region of $`B(g)`$ by constructing the so-called Mittag–Leffler star , i.e., by drawing cuts from all singular points to infinity along rays drawn through these points from the origin of coordinates. If $`g`$ lies in the analyticity region of $`B(g)`$, the integration path in (29) does not pass through its singularities and $`\stackrel{~}{B}(g)`$ is also analytic. If $`g_c`$ is a singular point of $`B(g)`$, the integration path in (29) for $`g=g_c`$ unavoidably passes through $`g_c`$, generating a singularity in $`\stackrel{~}{B}(g)`$. For the interesting case of power-law singularities we have the correspondence rules $$B(g)=A\mathrm{\Gamma }(\beta )\left(\frac{g_cg}{g_c}\right)\stackrel{~}{B}(g)=A\mathrm{\Gamma }(\beta b_1+b_0)\left(\frac{g_cg}{g_c}\right)^{\beta +b_1b_0}$$ $`(30)`$ for noninteger $`\beta +b_1b_0`$ and $$B(g)=A\mathrm{\Gamma }(\beta )\left(\frac{g_cg}{g_c}\right)^\beta \stackrel{~}{B}(g)=A\frac{(1)^{n+1}}{n!}\left(\frac{g_cg}{g_c}\right)^n\mathrm{ln}\left(\frac{g_cg}{g_c}\right),$$ $`(31)`$ if $`\beta +b_1b_0=n`$ is an integer. We see that the analyticity region for all the Borel transforms is identical and that it is sufficient to establish it for any fixed $`b_0`$. The choice $`b_0=1/2`$ is convenient for investigating functional integrals, since a simple result is obtained in that case for the Borel transform of an exponential function: $$F(g)=g^gB(g)=\frac{\mathrm{cos}(2\sqrt{g})}{\sqrt{\pi }}=\frac{1}{2\sqrt{\pi }}\{\mathrm{exp}(2i\sqrt{g}+\mathrm{c}.\mathrm{c}.\},$$ $`(32)`$ which preserves its exponential form. This permits writing an explicit expression for the Borel transform of the functional integral (1): $$B_I(g)=\frac{1}{2\sqrt{\pi }}D\phi \mathrm{exp}(S_0\{\phi \})[\mathrm{exp}\left(2i\sqrt{gS_{\mathrm{int}}\{\phi \}}\right)+\mathrm{c}.\mathrm{c}.].$$ $`(33)`$ The integrand is a regular function, and the analyticity region of $`B_I(g)`$ is determined by the condition for convergence of the integral. 3.2. Analyticity outside the negative semiaxis For simplicity, let us consider scalar $`\phi ^4`$ theory. Generalization to the $`n`$-component case is trivial and reduces to only some complication of the notation. We assume that $`m^2>0`$, bearing in mind the subsequent analytic continuation to arbitrary complex $`m^2`$. The integral (33) for $`\phi ^4`$ theory is defined well for positive values of $`g`$, since its convergence is determined by an exponential function of $`S_0\{\phi \}`$ and is obvious after the Fourier transformation of $`\phi (x)`$: $$S_0\{\phi \}=\frac{1}{2}d^dx\{(\phi )^2+m^2\phi ^2\}=\frac{1}{2}\underset{k}{}(k^2+m^2)|\phi _k|^2.$$ $`(34)`$ For the analytic continuation to complex $`g`$ we turn the integration path in (33), setting $$g=\stackrel{~}{g}e^{i\mathrm{\Psi }},\phi =\stackrel{~}{\phi }e^{i\mathrm{\Psi }/4},$$ $`(35)`$ where $`\stackrel{~}{g}`$ and $`\stackrel{~}{\phi }`$ are real, and $`\stackrel{~}{g}>0`$. Then the integral in (33) takes the form $$D\stackrel{~}{\phi }\mathrm{exp}(S_0\{\stackrel{~}{\phi }\}e^{i\mathrm{\Psi }/2})[\mathrm{exp}\left(2i\sqrt{\stackrel{~}{g}S_{\mathrm{int}}\{\stackrel{~}{\phi }\}}\right)+\mathrm{c}.\mathrm{c}.]$$ $`(36)`$ and converges for $`\pi <\mathrm{\Psi }<\pi `$. Thus, the Borel transform is analytic outside the negative semiaxis. 3.3. Analyticity within a circle We utilize the formal technique used in Refs. and and introduce the function $$R\{\phi \})=\frac{S_0\{\phi \}^2}{4S_{\mathrm{int}}\{\phi \}}.$$ $`(37)`$ Then (33) is rewritten in the form $$B_I(g)=\frac{1}{2\sqrt{\pi }}D\phi \mathrm{exp}\left(\left[1i\left(\frac{g}{R\{\phi \}}\right)^{1/2}\right]S_0\{\phi \}\right)+\mathrm{c}.\mathrm{c}.,$$ $`(38)`$ and after the replacement of $`R\{\phi \}`$ by the constant $`R_0`$, it is analytic within the circle $`|g|<R_0`$. Let us now suppose that $$R\{\phi \}R_0$$ $`(39)`$ for all $`\phi `$, i.e., $`R_0`$ is the exact lower bound of $`R\{\phi \}`$. Setting $`g=|g|e^{i\gamma }`$ ($`\pi \gamma \pi `$), we have the inequality $$|B_I(g)|\frac{1}{2\sqrt{\pi }}D\phi \left\{\mathrm{exp}\left(\left[1\left|\frac{g}{R\{\phi \}}\right|^{1/2}\mathrm{cos}\frac{\gamma }{2}\right]S_0\{\phi \}\right)+\mathrm{exp}(S_0\{\phi \})\right\},$$ $`(40)`$ which ensures convergence of the integral in (33) and, consequently, its analyticity within the circle $`|g|<R_0`$. To find $`R_0`$, we consider the variational problem of minimizing $`R\{\phi \}`$. It yields the equation $$\mathrm{\Delta }\phi (x)+m^2\phi (x)C\phi ^3(x)=0.$$ $`(41)`$ where $$C=\frac{S_0\{\phi \}}{2S_{\mathrm{int}}\{\phi \}},$$ which, after the replacement $`\phi (x)\phi (x)/\sqrt{C}`$, transforms into the standard equation of an instanton of $`\phi ^4`$ theory. Using it, we can easily show that $`R_0=S\{\phi _c\}`$, which establishes the required analyticity region (Fig. 3,b). Questions concerning the absence of instantons in massive four-dimensional theory are discussed in Sec. 3.5. Apart from the integral (1), some other functional integrals containing products of the type $`\phi (x_1)\phi (x_2)\mathrm{}\phi (x_M)`$ in the preexponential factor are of interest. The presence of such products does not influence the convergence, and all the proofs performed remain unchanged. 3.4. Invariance relative to algebraic operations As ’t Hooft pointed out , the singularities of Borel transforms are not shifted when algebraic operations are performed on the original functions. This can easily be proved for a modified definition of the Borel transform (10), which differs from (6) and (28), since $$F(g)=F_0+F_1g+F_2g^2+F_3g^3+\mathrm{},B(z)=F_0\delta (z)+\frac{F_1}{0!}+\frac{F_2}{1!}z+\frac{F_3}{2!}z^2+\mathrm{}$$ $`(42)`$ and $`B(z)`$ contains a $`\delta `$-function singularity at zero. The transformation of (10) by means of the replacement $`g1/z`$ reduces to a Laplace transformation and allows inversion. It can be used to express the Borel transform of the product $`F_3(g)=F_1(g)F_2(g)`$ in terms of the known Borel transforms of the factors: $$B_3(z)=\underset{0}{\overset{z}{}}𝑑z^{}B_1(z^{})B_2(zz^{}).$$ $`(43)`$ It can easily be seen that the $`\delta `$-function singularity in $`B_3(z)`$ corresponds to the definition (42) and that the singular points for finite $`z`$ coincide with the singular points of $`B_1(z)`$ and $`B_2(z)`$ (see the analogous reasoning in Sec. 3.1). In particular, the Borel transform $`g^n`$ is the function $`z^{n1}/\mathrm{\Gamma }(n)`$, which is analytic for integer values of $`n`$, and multiplication of the function by $`g^n`$ does not alter its analytic properties in the Borel plane. If $`F_2(z)=1/F_1(z)`$, then $$\delta (z)=\underset{0}{\overset{z}{}}𝑑z^{}B_1(z^{})B_2(zz^{})$$ $`(44)`$ and the $`\delta `$-function singularity on the left-hand side cancels out with the $`\delta `$-function singularities in $`B_1(z)`$ and $`B_2(z)`$. At finite values of $`z`$ the right-hand side contains singularities corresponding to singular points of $`B_1(z)`$ and $`B_2(z)`$, which are absent on the left-hand side and, therefore, compensate one another. This is possible only if $`B_2(z)`$ has singularities at the same points as $`B_1(z)`$. The proof of the analogous statements for linear operations, viz., summation, differentiation, integration, etc., is trivial. The standard definition of the Borel transform (6) is obtained from (10) and (42) when $`F_0=0`$ after the replacement $`F(g)gF(g)`$. In this case the $`\delta `$-function singularities disappear, and the remaining singularities are preserved at the same points due to the insignificance of the multiplier $`g`$. The definition (6) corresponds to the definition (28) with $`b_0=1`$, and, by virtue of Sec. 3.1, the analysis performed can be extended to arbitrary $`b_0`$. Since all the quantities entering into the theory, viz. the Green’s functions, vertex parts, etc., can be expressed in terms of functional integrals with identical analytic properties (see the end of Sec. 3.3) using a finite number of algebraic operations, their singular points in the Borel plane are the same as for the integral (1). 3.5. Renormalization procedure The absence of ultraviolet divergences was implicitly assumed above. In $`\phi ^4`$ theory this is correct for $`d<2`$. For $`2d4`$ a continual theory without divergences can be constructed by introducing counterterms into the Lagrangian . In the simple case where only renormalization of the mass is required ($`2d<4`$) the corresponding term in (4) is rewritten in the form $$m_0^2\phi ^2=(m^2+\mathrm{\Delta }m^2)\phi ^2=(m^2+Ag+Bg^2+Cg^3+\mathrm{})\phi ^2,$$ $`(45)`$ where the coefficients $`A,B,C,\mathrm{}`$ are chosen so as to cancel the divergences. When counterterms are present, the analytic properties of integrals of the type (1) become more complicated, since the coupling constant appears not only in the combination $`g\phi ^4`$, but also in the form of $`g\phi ^2`$, $`g^2\phi ^2`$, etc. One of the directions of renormalon-related activity involved specifically the introduction of additional terms into the Lagrangian and tracing the renormalon singularities appearing . A question arises in regard to the cancellation of singularities in the case when the coefficients in front of the additional terms are selected so as to remove divergences, for which an unequivocal answer could not be obtained. A simpler route is to explicitly introduce regularization and to use renormalization-group equations. Here we have in mind the so-called cutoff scheme : the vertex parts are calculated perturbatively as functions of the bare charge $`g_0`$ and the cutoff parameter $`\mathrm{\Lambda }`$, then scaling functions which depend only on $`g_0`$ are obtained, and, finally, renormalized vertices, which depend on the renormalized charge $`g`$, are constructed . In this case the explicit introduction of counterterms is not required, but all the details associated with their presence are taken into account, since the fundamental possibility of eliminating the divergences is essentially used to write the renormalization-group equations. The simplest way of regularization consists of substituting $`\phi ϵ(\widehat{p})\phi `$, where $`\widehat{p}`$ is the momentum operator, for the term $`(\phi )^2`$ in (4), which is brought into the form $`\phi \mathrm{\Delta }\phi =\phi \widehat{p}^2\phi `$. If $$ϵ(p)=ϵ(p),ϵ(p)0,$$ $`(46)`$ then both the entire structure of the instanton calculations and the proofs presented above are preserved. The only change occurs in the equation of the instanton (41), which is brought into the form $$ϵ(\widehat{p})\phi (x)+m^2\phi (x)\phi ^3(x)=0.$$ $`(47)`$ When the regularization $$ϵ(p)=p^2+p^6/\mathrm{\Lambda }^4$$ $`(48)`$ is employed, the dependence of the action $`S\{\phi \}`$ on the instanton radius $`R`$ in four-dimensional $`\phi ^4`$ theory has the form shown in Fig. 8. <sup>9</sup><sup>9</sup>9 This dependence can easily be obtained by describing an instanton by two parameters, viz., its radius and amplitude, and performing a variational estimation of the action. In the theory of disordered systems this corresponds to the optimal-fluctuation method . If $`\mathrm{\Lambda }=\mathrm{}`$, there is degeneracy with respect to the instanton radius in the massless theory , while there are no instantons in the massive theory because of the monotonic dependence of $`S\{\phi \}`$ on $`R`$. At finite values of $`\mathrm{\Lambda }`$ a minimum appears on the plot of $`S\{\phi \}`$ versus $`R`$ at $`m^2>0`$ (the dashed curves in Fig. 8), and instantons appear in the massive theory. Their action $`S\{\phi _c\}`$ determines the positions of the singularities in the Borel plane. For $`\mathrm{\Lambda }\mathrm{}`$ and arbitrary $`m^2>0`$ the value of $`S\{\phi _c\}`$ tends to the instanton action of the massless theory, and the positions of the singularities do not depend on $`m`$. <sup>10</sup><sup>10</sup>10 For dimensionalities $`2d<4`$ the influence of $`\mathrm{\Lambda }`$ on the properties of instantons is insignificant, and the role of renormalizations reduces to the fact that the instanton equation contains the renormalized mass . The dependence of $`S\{\phi _c\}`$ on $`m`$ is preserved in this case. The renormalization-group equations (in the Callan–Symanzik form) are valid for the vertices $`\mathrm{\Gamma }^{L,N}`$ with $`N`$ free tails and $`L`$ two-line insertions : $$\left[\frac{}{\mathrm{ln}\mathrm{\Lambda }^2}+\beta (g_0)\frac{}{g_0}+\left(L\frac{N}{2}\right)\eta (g_0)L\eta _2(g_0)\right]\mathrm{\Gamma }^{L,N}(g_0,\mathrm{\Lambda })=0.$$ $`(49)`$ Writing out three such equations with different $`L`$ and $`N`$, we can express the scaling functions $`\beta (g_0)`$, $`\eta (g_0)`$, and $`\eta _2(g_0)`$ in terms of the vertices $`\mathrm{\Gamma }^{L,N}(g_0,\mathrm{\Lambda })`$ using algebraic operations, which do not shift the positions of the singularities in the Borel plane. In the limit $`\mathrm{\Lambda }\mathrm{}`$, where Eq. (49) is valid, the dependence of the scaling functions on $`\mathrm{\Lambda }`$ disappears , and their singularities in the Borel plane correspond to the massless theory. The Gell-Mann–Low function $`\beta (g_0)`$ defines the relationship between the renormalized charge $`g`$ and the bare charge $`g_0`$. Let the functions $`F_0`$ and $`F_1`$ be such that $`F_0(g_0)F_1(g)`$. The relationship between the corresponding Borel transforms $`B_0`$ and $`B_1`$ \[in the sense of the definition (10)\] can easily be found for an infinitesimal charge transformation, $`g_0=g+2\beta (g)\delta \mathrm{\Lambda }/\mathrm{\Lambda }`$ \[see (25)\]: $$B_1(z)=B_0(z)+\frac{2\delta \mathrm{\Lambda }}{\mathrm{\Lambda }}\underset{0}{\overset{z}{}}𝑑y[B_0(y)+yB_0^{}(y)]B_\beta (zy),$$ $`(50)`$ where $`B_\beta (z)`$ is the Borel transform of the function $`\beta (g)/g`$. Equation (50) is analogous to Eq. (43); therefore, the analytic properties do not change in the course of the charge transformation. The vertices $`\mathrm{\Gamma }^{L,N}`$ diverge as $`\mathrm{\Lambda }\mathrm{}`$, but they become finite after separation of the divergent $`Z`$ factors from them and the transition from the bare to the renormalized charge. Since the $`Z`$ factors are, in turn, expressed in terms of the vertices $`\mathrm{\Gamma }^{L,N}`$ (Ref. ), the renormalized vertices have the required analytic properties. The dependence of the scaling functions on the renormalization scheme is given by the following conversion formulas : $$\stackrel{~}{\beta }(q(g))=\beta (g)\frac{dq(q)}{dg},$$ $$\stackrel{~}{\eta }(q(g))=\eta (g)\beta (g)\frac{d\mathrm{ln}p(q)}{dg},$$ $$\stackrel{~}{\eta }_2(q(g))=\eta _2(g)\beta (g)\frac{d\mathrm{ln}p_2(q)}{dg}.$$ $`(51)`$ The conversion functions $`q(g)`$, $`p(g)`$, and $`p_2(g)`$ for standard renormalization schemes (subtraction, cutoff etc.) are expressed in terms of the vertices $`\mathrm{\Gamma }^{L,N}`$. Therefore, the analytic properties of the scaling functions are identical in all the schemes. In the general case the analytic properties of the conversion functions require additional investigation. 4. CONCLUSION The results in Sec. 3 rule out the existence of renormalon singularities in $`\phi ^4`$ theory. If the arguments in Sec. 2 regarding the relationship to a Landau pole are considered convincing, $`\phi ^4`$ theory cannot be internally inconsistent. The same conclusion can be drawn on the basis of solid-state applications: a reasonable model of a disordered system reduces exactly to $`\phi ^4`$ theory , and the internal inconsistency of $`\phi ^4`$ theory would signify the impossibility, in principle, of obtaining a mathematical description of this model. Therefore, a revision of the results in Refs. and , in which indications of the internal inconsistency of $`\phi ^4`$ theory were obtained on the basis of an approximate reconstruction of the Gell-Mann–Low function, is urgently needed. The results of Sec. 3 refer only to $`\phi ^4`$ theory and cannot be extended directly to other field theories; however, along with the qualitative arguments in Sec. 2, they demonstrate the nonconstructiveness of the conception of renormalons as a whole. Therefore, it would be of interest to generalize the method of proof used in Sec. 3 to other cases. In quantum chromodynamics (QCD) the renormalon doctrine presently prevails . However, the specific details of QCD in this context have never been stressed. For example, ’t Hooft , speaking about QCD, gives explanations within $`\phi ^4`$ theory and quantum electrodynamics. In recent publications the term “naive nonabelianization” appeared, which essentially means neglecting the specific details of QCD. On the other hand, in QCD there is a special reason for the belief in renormalons, which has a purely phenomenological character. The analysis of experimental data leads to the conclusion that the contribution of the high orders has a momentum dependence $`1/q^4`$ (Ref. ). This dependence can easily be obtained from renormalon graphs, but, as is generally assumed, it cannot be obtained within the instanton method. The latter is based on the results in Refs. and , according to which the instanton contribution is proportional to $`1/q^{18}`$. However, it can easily be seen that a term $`1/q^4`$ appeared in Refs. and , but contained divergences which the authors found difficult to eliminate; <sup>11</sup><sup>11</sup>11 Such divergences also appear in $`\phi ^4`$ theory, and a procedure for eliminating them is well known . this term was “transported” to the renormalon sector with the reasoning that it “contributes to the renormalon singularity …rather than to the instanton one” (Ref. , p. 287). If there are no renormalon singularities, this contribution was simply discarded; therefore, no real calculation of the Lipatov asymptotics were made for QCD. This work was stimulated by lengthy discussions with P. G. Sil’vestrov, whom we thank for opposing the renormalon doctrine, his critical remarks, and general assistance in acquaintance with the situation. We also thank B. L. Ioffe, L. N. Lipatov, and the participants in the seminars at the Institute of Physical Problems, the P. N. Lebedev Physics Institute, the Institute of Theoretical and Experimental Physics, and the St. Petersburg Nuclear Physics Institute for their interest in this work and useful discussions. This work was carried out with financial support from the INTAS (Grant 96-0580) and the Russian Foundation for Basic Research (Project 96-02-19527). Appendix Construction of interpolation polynomial The polynomial of degree $`N`$ which coincides with the function $`f(x)`$ at the points $`x_0,x_1,x_2,\mathrm{},x_N`$, is defined by the Lagrange formula : $$P_N(x)=\underset{k=0}{\overset{N}{}}\frac{f(x_k)}{\psi ^{}(x_k)}\frac{\psi (x)}{(xx_k)},$$ $`(A1)`$ $$\psi (x)=(xx_0)(xx_1)(xx_2)\mathrm{}(xx_N),$$ $`(A2)`$ and the interpolation error is given by the expression $$R_N(x)=f(x)P_N(x)=\frac{f^{(N+1)}(\xi )}{(N+1)!}\psi (x),$$ $`(A3)`$ where $`\xi `$ belongs to the interval $`(x_0,x_N)`$. The function (18), which is of interest to us, behaves as $`q^n`$ at $`n\mathrm{¡}\mathrm{}N`$ with slowly varying $`q`$, so that $`\mathrm{ln}q\mathrm{ln}N`$. Neglecting these slow variations and omitting the common multiplier in (16) and (18), we have $$f(x)=q^x,A_N\frac{1}{6^NN^3}.$$ $`(A4)`$ Taking into account that $`|\psi (x)|\mathrm{\Delta }^{N+1}`$ in the interval $`0x\mathrm{\Delta }`$, we obtain $$|R_N(x)|\frac{(\mathrm{ln}q)^{N+1}q^\mathrm{\Delta }}{(N+1)!}\mathrm{\Delta }^{N+1},$$ $`(A5)`$ and the interpolation error is small for $$\mathrm{\Delta }\mathrm{¡}\mathrm{}N/\mathrm{ln}N.$$ $`(A6)`$ To investigate the dependence of the coefficient $`A_N`$ on the positions of the points $`x_k`$, we set $`\psi (x)=Re\psi (x+i0)`$, and calculating $`\mathrm{ln}\psi (x+i0)`$ using the Euler–MacLaurin formula, for $`\psi (x)`$ we obtain the expression $$\psi (x)=F(x)\mathrm{sin}G(x).$$ $`(A7)`$ In particular, for the power-law arrangement of points $$x_k=(k/N)^\alpha \mathrm{\Delta },k=0,1,\mathrm{},N,$$ $`(A8)`$ at $`\alpha 1`$ we have $$F(x)=(1)^N\sqrt{x(\mathrm{\Delta }x)}\mathrm{exp}\left\{N\left[\alpha (x/\mathrm{\Delta })^{1/\alpha }+\mathrm{ln}\mathrm{\Delta }\alpha \right]\right\},$$ $$G(x)=\pi N(x/\mathrm{\Delta })^{1/\alpha }.$$ $`(A9)`$ For a high-order coefficient of the polynomial (A.1) we obtain $$A_N=\underset{k=0}{\overset{N}{}}\frac{f(x_k)}{\psi ^{}(x_k)}\mathrm{exp}\{(\alpha \mathrm{ln}\mathrm{\Delta })N\}$$ $`(A10)`$ (the sum is determined by the term with $`k=1`$), and for $`\alpha \mathrm{ln}N`$ the coefficient $`A_N`$ can be factorially small or factorially large, depending on the relationship between $`\alpha `$ and $`\mathrm{ln}\mathrm{\Delta }`$, so that the required value of (A4) falls in the range of variation. Thus, the required polynomial (16) exists in the interval $`0nn_0`$, where $`n_0N/\mathrm{ln}N`$. Translated by P. Shelnitz
warning/0002/cond-mat0002306.html
ar5iv
text
# Conductivity of a clean one-dimensional wire ## Abstract We study the low-temperature low-frequency conductivity $`\sigma `$ of an interacting one dimensional electron system in the presence of a periodic potential. The conductivity is strongly influenced by conservation laws, which, we argue, need be violated by at least two non-commuting Umklapp processes to render $`\sigma `$ finite. The resulting dynamics of the slow modes is studied within a memory matrix approach, and we find exponential increase as the temperature is lowered, $`\sigma (\mathrm{\Delta }n)^2e^{T_0/(NT)}`$ close to commensurate filling $`M/N`$, $`\mathrm{\Delta }n=nM/N1`$, and $`\sigma e^{(T_0^{}/T)^{2/3}}`$ elsewhere. The finite-temperature conductivity of a clean one-dimensional wire is a fundamental and much studied question. Clearly the “bulk” conductivity of a wire in the absence of a periodic potential is infinite even at finite temperatures $`T`$. In this case the conductance is independent of the length of the wire and is determined by the contacts only. Surprisingly, much less is known about the conductivity in the presence of Umklapp scattering induced by a periodic potential. There is not even an agreement whether it is finite or infinite at finite temperatures for generic systems . We shall show that the correct answer emerges when all relevant (weakly violated) conservation laws are taken into account. Those conservation laws are exact at the Fermi surface and are violated by Umklapp terms away from it. We shall study the associated slow modes by means of a memory matrix formalism able to keep track of their dynamics. It will allow us to calculate reliably the low temperature, low frequency conductivity. The topology of the Fermi surface of a $`1d`$ metal determines its low-energy excitations. Two well defined Fermi-points exist at momenta $`k=\pm k_F`$, allowing us to define left and right moving excitations, to be described by $`\mathrm{\Psi }_{L/R,\sigma =}`$. We shall include in the fields momentum modes extending to the edge of the Brillouin zone, usually omitted in treatments that concentrate on physics very close to the Fermi-surface. The Hamiltonian, including high energy processes, is $`H=H_{LL}+H_{\text{irr}}+{\displaystyle \underset{n,m}{\overset{\mathrm{}}{}}}H_{n,m}^U.`$ (1) $`H_{LL}`$ is the well-known Luttinger liquid Hamiltonian capturing the low energy behavior, $`H_{LL}`$ $`=`$ $`v_F{\displaystyle \left(\mathrm{\Psi }_{L\sigma }^{}i_x\mathrm{\Psi }_{L\sigma }\mathrm{\Psi }_{R\sigma }^{}i_x\mathrm{\Psi }_{R\sigma }\right)}+g{\displaystyle \rho (x)^2}`$ (2) $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \frac{dx}{2\pi }\underset{\nu =\sigma ,\rho }{}v_\nu \left(K_\nu (_x\theta _\nu )^2+\frac{1}{K_\nu }(_x\varphi _\nu )^2\right)}`$ (3) $`v_F`$ is the Fermi velocity, $`g>0`$ measures the strength of interactions, $`\rho =\rho _L+\rho _R`$ is the sum of the left and right moving electron densities. In the second line we wrote the bosonized version of the Hamiltonian. Here $`v_\sigma `$, $`v_\rho `$ are the spin and charge velocities, and the interactions determine the Luttinger parameters $`K_\nu `$ with $`v_\nu K_\nu =v_F`$, $`v_\rho /K_\rho =v_F+g/\pi `$, $`v_\sigma /K_\sigma =v_Fg/\pi `$. The high energy processes are captured in the subsequent terms which are formally irrelevant at low energies (we consider only systems away from a Mott transition, i.e. away from half filling). Some of them, however, determine the low-frequency behavior of the conductivity at any finite $`T`$, since they induce the decay of the conserved modes of $`H_{LL}`$ (they are “dangerously irrelevant”). We classify these irrelevant terms with the help of two operators which will play the central role in our discussion. The first one is the translation operator $`P_T`$ of the right- and left-moving fields, the second one, $`J_0=N_RN_L`$, is the difference of the number of right- and left-moving electrons, and is up to $`v_F`$, the charge current of $`H_{LL}`$: $`P_T`$ $`=`$ $`{\displaystyle \underset{\sigma }{}}{\displaystyle 𝑑x\left(\mathrm{\Psi }_{R\sigma }^{}(i_x)\mathrm{\Psi }_{R\sigma }+\mathrm{\Psi }_{L\sigma }^{}(i_x)\mathrm{\Psi }_{L\sigma }\right)}`$ (4) $`J_0`$ $`=`$ $`N_RN_L={\displaystyle \underset{\sigma }{}}{\displaystyle 𝑑x\left(\mathrm{\Psi }_{R\sigma }^{}\mathrm{\Psi }_{R\sigma }\mathrm{\Psi }_{L\sigma }^{}\mathrm{\Psi }_{L\sigma }\right)}`$ (5) Both $`P_T`$ and $`J_0`$ are conserved by $`H_{LL}`$; their importance for transport properties is due to the fact that both stay approximately conserved in any one dimensional metal (away from half filling): processes which change $`J_0`$ are forbidden close to the Fermi surface by momentum conservation. The linear combination $`P_0=P_T+k_FJ_0`$ can be identified with the total momentum of the full Hamiltonian $`H`$ and is therefore also approximately conserved. We proceed to the classification of the formally irrelevant terms in the Hamiltonian. This classification allows us to select all those terms (actually few in number) that determine the current dynamics. $`H_{\text{irr}}`$ includes all terms in $`HH_{LL}`$ which commute with both $`P_T`$ and $`J_0`$, such as corrections due to the finite band curvature, due to finite-range interactions and similar terms. We will not need their explicit form. The Umklapp terms $`H_{n,m}^U`$ ($`n,m=0,1,\mathrm{}`$) convert $`n`$ right-movers to left-movers (and vice versa) picking up lattice momentum $`m2\pi /a=mG`$, and do not commute with either $`P_T`$ or $`J_0`$. Leading terms are of the form, $`H_{0,m}^U`$ $``$ $`g_{0,m}^U{\displaystyle e^{i\mathrm{\Delta }k_{0,m}x}(\rho _L+\rho _R)^2}+h.c.`$ (6) $`H_{1,m}^U`$ $``$ $`g_{1,m}^U{\displaystyle \underset{\sigma }{}}{\displaystyle e^{i\mathrm{\Delta }k_{1,m}x}\mathrm{\Psi }_{R\sigma }^{}\mathrm{\Psi }_{L\sigma }\rho _\sigma }+h.c.`$ (7) $`H_{2,m}^U`$ $``$ $`g_{2,m}^U{\displaystyle e^{i\mathrm{\Delta }k_{2,m}x}\mathrm{\Psi }_R^{}\mathrm{\Psi }_R^{}\mathrm{\Psi }_L\mathrm{\Psi }_L}+h.c.`$ (8) with momentum transfer $`\mathrm{\Delta }k_{n,m}=n2k_FmG`$. A process transfering $`n>1`$ electrons with total spin $`n_s/2`$ pointing in the $`z`$-direction can be neatly expressed as $`H_{n,m}^U={\displaystyle \frac{g_{n,m,n_s}^U}{(2\pi \alpha )^n}}{\displaystyle e^{i\mathrm{\Delta }k_{n,m}x}e^{i\sqrt{2}(n\varphi _\rho +n_s\varphi _\sigma )}}+h.c.,`$ (9) $`\alpha `$ being a cut-off, of the order of the lattice spacing. In fermionic variables the integrand takes the form $`_{j=0}^{n/21}\mathrm{\Psi }_R^{}(x+j\alpha )\mathrm{\Psi }_R^{}(x+j\alpha )\mathrm{\Psi }_L(x+j\alpha )\mathrm{\Psi }_L(x+j\alpha )`$ (for $`n_s=0`$ and even $`n`$). Note, though, that any single term $`H_{nm}^U`$ conserves a linear combination of $`J_0`$ and $`P_T`$, $`[H_{nm}^U,\mathrm{\Delta }k_{nm}J_0+2nP_T]=0.`$ (10) Indeed, a term of the form (9) would appear in a continuum model without Umklapp scattering, but with a Fermi momentum $`\stackrel{~}{k}_F=\mathrm{\Delta }k_{nm}/(2n)`$. In such a model, $`\mathrm{\Delta }k_{nm}J_0/(2n)+P_T`$ is the total momentum of the system and therefore conserved. The importance of this simple but essential conservation law has to our knowledge not been sufficiently realized in previous calculations of the conductivity. Due to this conservation law a single Umklapp term can never induce a finite conductivity! At least two independent Umklapp terms are required to lead to a complete decay of the current. Further, two incommensurate Umklapp terms suffice to generate the rest. To calculate the conductivity it is necessary to keep track of the nearly conserved quantities and their relation to the current. We will develop a description of the slowest variables using the Mori-Zwanzig memory functional . Approximations within this scheme amount to short-time expansions. In general, the short time decay of a quantity carries little information on its long-time behavior; this, however, is not the case for the slowest variables in the system, where the short time and hydrodynamic behavior coincide. To set up the formalism we define a scalar product $`(A|B)`$ in the space of operators, $`\left(A(t)|B\right)`$ $``$ $`{\displaystyle \frac{1}{\beta }}{\displaystyle _0^\beta }𝑑\lambda A(t)^{}B(i\lambda ),`$ (11) where we use the usual Heisenberg picture with $`A(t)=e^{iHt}Ae^{iHt}`$. We choose a set “slow” operators $`j_1,j_2,\mathrm{}j_N`$ which includes $`j_1=J`$, the full current operator. Standard arguments lead to the electric conductivity, $`\sigma (\omega ,T)`$ $`=`$ $`\left[\left(\widehat{M}(\omega ,T)i\omega \right)^1\widehat{\chi }(T)\right]_{11}.`$ (12) Here $`\widehat{\chi }_{pq}=\beta (j_p|j_q)`$ is the matrix of the static $`j_pj_q`$ susceptibilities (as usually defined), and $`\widehat{M}`$ is the matrix of memory functions given by the projected correlation functions of time-derivatives of the “slow” operators, $`\widehat{M}_{pq}(\omega )=\beta {\displaystyle \underset{r}{}}\left(_tj_q\left|Q{\displaystyle \frac{i}{\omega QLQ}}Q\right|_tj_r\right)(\widehat{\chi }^1)_{rp}.`$ (13) The Liouville “super”-operator, $`L`$, is defined by $`LA=[H,A]`$ and $`Q`$ is the projection operator on the space perpendicular to the slowly varying variables $`j_p`$, $`Q=1{\displaystyle \underset{pq}{}}|j_q)\beta (\widehat{\chi }^1)_{qp}(j_p|.`$ (14) We assumed for simplicity that all $`j_p`$ have the same signature under time reversal. The perturbative expansion of the memory matrix $`\widehat{M}`$ is accompanied by factors $`1/\omega `$ guaranteeing it is always valid at short times. It is also valid for small frequencies provided the slowly evolving degrees of freedom are projected out (by the operator $`Q`$). Unlike the conductivity it is expected to be a smooth function of the coupling constants which can be perturbatively evaluated. We first consider a situation where some linear combinations of the $`j_p`$ are conserved by $`H`$, in which case an infinite conductivity is expected. We introduce $`𝒫_c`$, the projection operator on the space of conserved currents, and carry out the required matrix inversion to find, $`\sigma (\omega 0,T>0)`$ $`=`$ $`\sigma _{\text{reg}}(\omega ,T)+i{\displaystyle \frac{(\widehat{\chi }\widehat{\chi }_c^1\widehat{\chi })_{11}}{\omega +i0}},`$ (15) where $`\widehat{\chi }_c^1=𝒫_c(𝒫_c\widehat{\chi }𝒫_c)^1𝒫_c`$. Within any simple (short-time) approximation, $`\sigma _{\text{reg}}(\omega ,T)`$ as defined above, is regular (this approximation fails e.g. if some conserved current $`\widehat{j}`$ is not included in $`j_1\mathrm{}j_N`$). Hence the Drude weight $`D(T)`$ is finite at finite temperatures, $`\text{Re}\sigma (\omega 0)=2\pi D(T)\delta (\omega )=\pi (\widehat{\chi }\widehat{\chi }_c^1\widehat{\chi })_{11}\delta (\omega )`$. It is determined by the “overlap” of the physical current operator $`J`$ with the conserved quantities $`\chi _{1s}`$, $`s`$ labeling the conserved currents. Remarkably, our perturbative approximation is in accord with an exact inequality for the Drude weight, $`D(T)\frac{1}{2}(\widehat{\chi }\widehat{\chi }_c^1\widehat{\chi })_{11}`$. Note that $`\widehat{\chi }`$ can be calculated to an arbitrary degree of precision around a Luttinger liquid and that the lower bound can be improved by including more conserved quantities . Now consider the more realistic situation where the previously conserved currents decay slowly (via Umklapp processes), in which case a finite conductivity is expected. We restrict ourselves to the two-dimensional space spanned by $`v_FJ_0`$ and $`P_T`$, which we argue have the longest decay rate and dominate the transport. Here we approximate $`Jv_FJ_0`$ to keep the presentation simple. This affects only the high frequency behavior of the conductivity . There is a large number of other nearly conserved quantities. For example $`H_{LL}+H_{21}^U`$, the relevant low-energy model close to half filling, is integrable and therefore is characterized by an infinite number of conservation laws. We can, however, neglect them at low $`T`$ if our initial model is not integrable, expecting that practically all conservation laws are destroyed by (formally irrelevant) terms close to the Fermi surface leading to decay rates proportional to some power of $`T`$. This is to be compared to $`J_0`$ and $`P_T`$ which commute with all scattering processes at the Fermi surface, leading to exponentially large lifetimes. We now proceed to calculate the Memory matrix. To leading order in the perturbations we can replace $`L`$ in (13) by $`L_{LL}=[H_{LL},.]`$ , since $`_tv_FJ_0`$ and $`_tP_T`$ are already linear in $`g_{n,m}^U`$. As $`L_{LL}P_T=L_{LL}J_0=0`$, there is no contribution from the projection operator $`Q`$. The memory matrix takes the form, $`\widehat{M}`$ $``$ $`{\displaystyle \underset{nm}{}}M_{nm}(\omega ,T)\left(\begin{array}{cc}v_F^2(2n)^2& 2nv_F\mathrm{\Delta }k_{nm}\\ 2nv_F\mathrm{\Delta }k_{nm}& (\mathrm{\Delta }k_{nm})^2\end{array}\right)\widehat{\chi }^1`$ (18) where, $`\widehat{\chi }`$ $``$ $`\left(\begin{array}{cc}2v_F/\pi & 0\\ 0& \frac{\pi T^2}{3}\left(\frac{1}{v_\rho ^3}+\frac{1}{v_\sigma ^3}\right)\end{array}\right)`$ (21) $`M_{nm}`$ $``$ $`(g_{nm}^U)^2M_n(\mathrm{\Delta }k_{n,m},\omega ){\displaystyle \frac{F;F_\omega ^0F;F_{\omega =0}^0}{i\omega }}.`$ (22) Here $`F=[J_0,H_{nm}^U]/(2n)`$ (for simplicity we drop the indices $`n,m`$ on F), and $`F;F_\omega ^0`$ is the retarded correlation function of $`F`$ calculated with respect to $`H_{LL}`$. The memory function $`M_2`$ of the $`4k_FQ`$ process $`H_{21}^U`$ was calculated by Giamarchi , (not considering the matrix structure of $`\widehat{M}`$ required by the conservation laws.) Higher Umklapps are considered in . For $`n_s=0`$ and even $`n`$ the memory function due to the term (9) can be analytically calculated, $`M_n(\mathrm{\Delta }k,\omega )=`$ $`{\displaystyle \frac{2\mathrm{sin}2\pi K_\rho ^n}{\pi ^4\alpha ^{2n2}v_\rho }}\left[{\displaystyle \frac{2\pi \alpha T}{v_\rho }}\right]^{4K_\rho ^n2}{\displaystyle \frac{1}{i\omega }}\times `$ (23) $`\times [B(K_\rho ^n`$ $`iS_+,12K_\rho ^n)B(K_\rho ^niS_+,12K_\rho ^n)`$ (24) $`B(K_\rho ^n`$ $`iS_+^0,12K_\rho ^n)B(K_\rho ^niS_+^0,12K_\rho ^n)]`$ (26) $`{\displaystyle \frac{\alpha ^{22n}}{\pi ^2\mathrm{\Gamma }^2(2K_\rho ^n)v_\rho T}}\left({\displaystyle \frac{\alpha \mathrm{\Delta }k}{2}}\right)^{4K_\rho ^n2}e^{v_\rho \mathrm{\Delta }k/(2T)}`$ where $`K_\rho ^n=(n/2)^2K_\rho `$, $`B(x,y)=\mathrm{\Gamma }(x)\mathrm{\Gamma }(y)/\mathrm{\Gamma }(x+y)`$ and $`S_\pm =(\omega \pm v_\rho \mathrm{\Delta }k)/(4\pi T)`$, $`S_\pm ^0=S_\pm (\omega =0)`$. The last line is valid for $`\omega =0`$ and $`Tv_\rho \mathrm{\Delta }k`$. The origin of the exponential factor is as follows: processes involving momentum transfer $`\mathrm{\Delta }k`$ are associated with initial and final states of energies $`v|\mathrm{\Delta }k|/2`$, which are exponentially suppressed. If only charge degrees of freedom are involved $`v=v_\rho `$, otherwise $`v=\text{min}(v_\sigma ,v_\rho )=v_\sigma `$ . For $`Tv_\sigma \mathrm{\Delta }k_{nm}`$, $`n_s>0`$ and $`\omega =0`$, we have, $`M_n(\mathrm{\Delta }k){\displaystyle \frac{\left(\alpha T/v_\rho \right)^{n^2K_\rho 1}\left(\alpha \mathrm{\Delta }k\right)^{n_s^2K_\sigma 2}}{\mathrm{\Gamma }^2(n_s^2K_\sigma /2)v_\sigma ^2\alpha ^{2n3}}}e^{v_\sigma \mathrm{\Delta }k/(2T)},`$ (27) while for $`Tv_\rho \mathrm{\Delta }k_{nm}`$: $`M_nT^{n^2K_\rho +n_s^2K_\sigma 3}`$. Using the above expressions with only one Umklapp term leads to a finite Drude weight (cf eq(15)), $`D(T){\displaystyle \frac{v_\rho K_\rho }{\pi }}{\displaystyle \frac{1}{1+T^2\frac{2\pi ^2n^2K_\rho }{3(v_\rho \mathrm{\Delta }k_{nm})^2}\left(1+\frac{v_\rho ^3}{v_\sigma ^3}\right)}}.`$ (28) in accord with the observation that one process $`H_{nm}^U`$ is not sufficient to degrade the current. Only in the presence of a second incommensurate process $`H_{n^{}m^{}}^U`$ is the dc conductivity finite, $`\sigma (T,\omega =0)={\displaystyle \frac{(\mathrm{\Delta }k_{nm})^2/M_{n^{}m^{}}+(\mathrm{\Delta }k_{n^{}m^{}})^2/M_{nm}}{\pi ^2(n\mathrm{\Delta }k_{n^{}m^{}}n^{}\mathrm{\Delta }k_{nm})^2}}`$ (29) Note that the slowest process determines the low-$`T`$ conductivity. The frequency and temperature dependence of the conductivity in the case of two competing Umklapp terms is shown in Fig. 1. The commensurate situation $`\mathrm{\Delta }k_{nm}=0`$ requires extra considerations. Whether the dominant scattering process $`H_{nm}`$ will completely relax the current $`J`$ depends according to (15) on the overlap $`\chi _{JP_T}`$ ($`[P_T,H_{nm}]=0`$). Using the continuity equation for the charge, $`\chi _{JP_T}`$ can be related to the deviation $`\mathrm{\Delta }\rho =2\mathrm{\Delta }n/a`$ of the electron density from commensurate filling with the remarkable identity $`\chi _{JP_T}=2\mathrm{\Delta }n/a+O(e^{\beta ϵ_F})`$. In a 3d lattice of 1d wires, $`\mathrm{\Delta }n`$ is fixed by charge neutrality and is $`T`$ independent, in a single wire with contacts $`\mathrm{\Delta }n`$ varies at low $`T`$ with $`\mathrm{\Delta }n(T)T^2/(mv^3)`$, where the mass $`m`$ is a measure of the breaking of particle-hole symmetry, e.g. due to a band-curvature $`k^2/2m`$. In this case it is important to replace $`\mathrm{\Delta }k_{nm}=0`$ in Eqn. (28) or (29) by $`G\mathrm{\Delta }n(T)`$. Which of the various scattering processes will eventually dominate at lowest $`T`$? At intermediate temperatures, certainly low-order (small $`n`$) scattering events win, being less suppressed by Pauli blocking. At lower temperature the exponential factors in (27) prevail and the processes with the smallest $`\mathrm{\Delta }k_{nm}`$ are favored. We first analyze the situation close to a commensurate point $`k_FGM_0/(2N_0)`$. The two dominant processes are $`H_{N_0M_0}^U`$ with $`\mathrm{\Delta }k_{N_0M_0}0`$ and $`H_{N_1M_1}^U`$ with $`\mathrm{\Delta }k_{N_1M_1}=\pm G/N_0`$ (or $`N_1M_0=\pm 1\text{ mod }N_0`$). The integer $`N_1`$ of order $`N_0`$, $`N_1=\gamma _1N_0`$, depends strongly on the precise values of $`N_0`$ and $`M_0`$. We thus find that the d.c. conductivity at low $`T`$ is largest close to commensurate points with, $`\sigma (k_FGM_0/(2N_0))(\mathrm{\Delta }n(T))^2\mathrm{exp}[\beta vG/(2N_0)]`$ (30) but $`\sigma T^{N_0^2K_\rho (N_0\text{ mod }2)^2K_\sigma +3}`$ if the density is exactly commensurate with $`|\mathrm{\Delta }n(T)|<e^{\beta Gv/(4N_0)}`$. To estimate the conductivity at a typical “incommensurate” point or at commensurate points at temperatures not too low, we have to balance algebraic and exponential suppression in (27) by minimizing $`\beta vG/(2N)+(\gamma _1N)^2K\mathrm{log}[T]`$ in a saddle-point approximation to the sum over all Umklapp processes in $`\widehat{M}`$. Up to logarithmic corrections we obtain $`N_{\text{max}}^3\beta vG/(\gamma _1)^2`$ and therefore for a “typical” incommensurate filling, $`\sigma _{\text{typical}}\mathrm{exp}[c(\beta vG)^{2/3}]`$ (31) where $`c`$ is a number depending logarithmically on $`T`$. At present we cannot rule out that various logarithmic corrections sum up to modify the power law in the exponent. We argue, however, that due to the exponential increase (30) of $`\sigma `$ at commensurate fillings with exponents proportional to $`1/N_0`$, the conductivity at small $`T`$ at any incommensurate point is smaller than any exponential (but is larger than any power since any single process is exponentially suppressed). In Fig. 2 we show schematically the conductivity as a function of filling becoming more and more “fractal-like ” for lower $`T`$. Can the effects we predict be seen experimentally? The complicated structures as a function of filling shown in Fig. 2 are not observable in practice as they occur only at exponentially large conductivities. The $`T`$-dependence of the conductivity at intermediate temperatures, however, should be accessible, e.g. by comparing the conductivities of clean wires of different length. Perhaps more importantly, it is straightforward to apply our method to a large number of other relevant situation, e.g. close to a Mott transition or in the presence of 3d phonons, as we will discuss in a forthcoming paper. We thank R. Chitra, A.J. Millis, E. Orignac, A.E. Ruckenstein, S. Sachdev, and P. Wölfle for helpful discussions. Part of this work was supported by the A. v. Humboldt Foundation and NSF grant DMR9632294 (A.R.).
warning/0002/quant-ph0002081.html
ar5iv
text
# A group of invariance transformations for nonrelativistic quantum mechanics ## 1 Introduction. The rise or full development of almost all major physical theories –such as classical mechanics, electromagnetism, special relativity theory and general relativity theory– saw the introduction of a new group of transformations, with respect to which an invariant formulation of physical laws was required: Galilean transformations for classical mechanics, Lorentz-Poincaré transformations for electromagnetism and special relativity, and diffeomorphisms for general relativity. Quantum mechanics is a remarkable exception to this rule, since no special group of transformations is associated with it. This observation is an incentive to search for such a group, in the hope that its discovery will allow to formulate the theory more elegantly and without the many paradoxes and problems that currently affect it. With this aim in mind, the group of asymptotically Euclidean transformations, i.e., transformations which are equivalent to Euclidean transformations at space-time infinity, is defined on the Galilean space-time. A formulation of nonrelativistic quantum mechanics (Schrödinger’s equation) which is invariant with respect to such transformations is then proposed. The first part of this paper develops the corresponding mathematical theory; the second part discusses its physical interpretation. In the mathematical section, a demonstration will be provided only for the most significant lemmas and theorems. This paper is a development a previous paper . ## 2 Mathematical theory. ### 2.1 Asymptotically Euclidean transformations. For the sake of brevity, let $`G`$ designate the Galilean space-time $`R\times R^3`$. The transformations usually considered on $`G`$ are the Galilean transformations, which have the form $`f(t,𝐱)=(t+t_0,R𝐱𝐯_0t+𝐱_0)`$, where $`R`$ is an orthogonal matrix. The group of Galilean transformations is composed of three subgroups: translations $`f(t,𝐱)=(t+t_0,𝐱+𝐱_0)`$, rotations $`f(t,𝐱)=(t,R𝐱)`$, and boosts $`f(t,𝐱)=(t,𝐱𝐯_0t)`$. Galilean transformations separately preserve space distances and time distances. Given a set $`AG`$, let $`A(t)R^3`$ be the space cross-section of the set $`A`$ at the time $`t`$, i.e., $`A(t):=\{𝐱R^3|(t,𝐱)A\}`$. A semitrajectory $`\gamma `$ is the graph in $`G`$ of a continuous curve from $`[t_0,+\mathrm{})`$ to $`R^3`$. In accordance with the above defined notation, $`\gamma (t)R^3`$ is the space point occupied by the semitrajectory at the time t. The point $`(t_0,\gamma (t_0))`$ is called origin of the semitrajectory. Clearly, this definition of semitrajectory implies that it is temporally headed into the future. This directionality, although undeclared, will always be implied in all the definitions, lemmas and theorems of this paper. #### 2.1.1 Asymptotic velocity. Let us now introduce the important notion of asymptotic velocity of a semitrajectory. This notion, applied to classical trajectories, is already in use within classical scattering theory . ###### Definition 1 (a) The asymptotic velocity of a semitrajectory is the limit of $`\gamma (t)/t`$ for $`t+\mathrm{}`$, when it exists. (b) A semitrajectory allowing asymptotic velocity is said to be asymptotically regular. Let $`V`$ be the space of asymptotic velocities (which is isomorphic with $`R^3`$); let $`\mathrm{\Gamma }`$ be the set of asymptotically regular semitrajectories; and finally, let $`\omega :\mathrm{\Gamma }V`$ be the application that associates each asymptotically regular semitrajectory with its asymptotic velocity. ###### Lemma 1 (a) If there is a $`𝐯V`$ such that $`\gamma (t)𝐯t`$ is bounded for $`tt_0`$, then $`\omega (\gamma )=𝐯`$; in particular, if $`\gamma (t)`$ is bounded, then $`\omega (\gamma )=0`$. (b) If $`\gamma (t)`$ admits first derivative and $`\dot{\gamma }(t)𝐯`$ for $`t+\mathrm{}`$ , then $`\omega (\gamma )=𝐯`$. (c) If $`f(t,𝐱)=(t+t_0,R𝐱𝐯_0t+𝐱_0)`$ is a Galilean transformation, then $`\omega [f(\gamma )]=R\omega (\gamma )𝐯_0`$. ###### Examples 1 (a) If $`\gamma (t)=𝐯t+𝐱_0\mathrm{sin}\omega t`$, then $`\omega (\gamma )=𝐯`$. (b) If $`\gamma (t)=𝐚|t|^{(1ϵ)}+𝐱_0`$, with $`0<ϵ<1`$, then $`\omega (\gamma )=0`$. (c) The curve $`\gamma (t)=𝐯t\mathrm{sin}\omega t`$ is not asymptotically regular. The converse of Lemma 1a does not hold, i.e., it is possible to have $`\omega (\gamma )=0`$ while $`\gamma (t)`$ is not bounded: consider example 1c. Unbounded trajectories having vanishing asymptotic velocity are also termed almost bounded trajectories. Hereafter, unless explicitly stated, the term trajectory is used to designate an asymptotically regular trajectory. #### 2.1.2 Causal transformations. Let us now study homeomorphic transformations on $`G`$, for which we will use the more generic term transformations. This paper studies a group of transformations which is broader than Galilean transformations. One property that these transformation will still be required to have, however, is causality, i.e., they must maintain the time ordering of events. Let us specify this property by using first of all the following notation: if $`(t,𝐱)G`$, we define the projectors $`P_T(t,𝐱):=t`$ and $`P_X(t,𝐱):=𝐱`$. ###### Definition 2 A transformation $`f`$ on $`G`$ is causal if, for every $`u,wG`$ such that $`P_T(u)<P_T(w)`$, one has $`P_T[f(u)]<P_T[f(w)]`$. ###### Lemma 2 (a) If $`f`$ is causal, $`P_T(u)=P_T(w)`$ implies $`P_T[f(u)]=P_T[f(w)]`$. (b) Causal transformations form a group. (c) A causal transformation is of the form $`f(t,𝐱)=(f_T(t),f_X(t,𝐱))`$, where $`f_T:RR`$ is a monotonically increasing homeomorphism, and $`f_X(t,):R^3R^3`$ is a homeomorphism for every $`t`$. (d) If $`f`$ is causal and $`AG`$, one has $`f(A)[f_T(t)]=f_X[t,A(t)]`$. (e) A transformation $`f`$ is causal if and only if it preserves semitrajectories. Hereafter, unless otherwise specified, the term transformation is used to designate a causal transformation. #### 2.1.3 Asymptotic homeomorphisms. The following definitions meaningfully characterize the asymptotic behavior of a transformation. ###### Definition 3 (a) A transformation $`f`$ is asymptotically regular at the point $`𝐯V`$ if there is a $`𝐯^{}V`$ such that for every semitrajectory $`\gamma `$ for which the relation $`\omega (\gamma )=𝐯`$ holds, one has $`\omega [f(\gamma )]=𝐯^{}`$. (b) A transformation $`f`$ is asymptotically regular if it is asymptotically regular in every point $`𝐯V`$. (c) If $`f`$ is asymptotically regular, the application $`f^+:VV`$ defined by setting $`f^+[\omega (\gamma )]:=\omega [f(\gamma )]`$ is termed asymptotic transform of f. (d)$`f`$ is an asymptotic homeomorphism if $`f`$ and $`f^1`$ are asymptotically regular. Clearly, the definition of asymptotic regularity can be considered the asymptotic equivalent of the usual definition of regularity of a function, which says that a function is regular at a point if it admits a limit at that point. The definition of asymptotic homeomorphism is also similar to the usual nonasymptotic definition. ###### Examples 2 (a) A Galilean transformation $`f(t,𝐱)=(t+t_0,R𝐱𝐯_0t+𝐱_0)`$ is an asymptotic homeomorphism, and one has $`f^+(𝐯)=R𝐯𝐯_0`$. This is derived trivially from Lemma 1c. The asymptotic transform of a Galilean transformation is therefore a Euclidean transformation on the space of asymptotic velocities: rotations remain rotations, boosts become translations, and translations disappear. (b) If $`f(t,𝐱)=(at,a𝐱)`$, where $`a>0`$ is a multiplicative constant, $`f`$ is asymptotically regular and the relation $`f^+(𝐯)=𝐯`$ holds. (c) The transformation $`f(t,𝐱)=(t,𝐱+𝐯_0t\mathrm{sin}\omega t)`$ is not asymptotically regular at any point. ###### Theorem 1 (a) If $`f`$ is asymptotically regular, $`f^+`$ is continuous. (b) If $`f`$ is an asymptotic homeomorphism, then $`f^+`$ is a homeomorphism on V and one has $`(f^1)^+=(f^+)^1`$. (c) If $`f`$ and $`g`$ are two asymptotic homeomorphisms, then $`gf`$ also is an asymptotic homeomorphism, and the relation $`(gf)^+=g^+f^+`$ holds. In order to prove the above theorem it is convenient to modify the function $`f`$ so that asymptotic regularity is changed into normal regularity. In order to easily carry out this modification, instead of considering transformations on $`G`$, we will consider transformations on $`G^+:=[a,+\mathrm{})\times R^3G`$, where $`a`$ is a positive constant. Such a restriction is irrelevant to the remainder, since the theorem relate to the future asymptotic properties of transformations. Let us define $`F^+:=(0,1/a]\times V`$, $`F_0:=\{0\}\times V`$ and $`F_0^+:=F^+F_0`$. If $`(s,𝐯)F_0^+`$, let us define the projectors $`P_S(s,𝐯):=s`$ and $`P_V(s,𝐯):=𝐯`$. Furthermore, let us define the application $`h:G^+F^+`$ by setting $`h(t,𝐱):=(1/t,𝐱/t)`$. The application $`h`$ is a homeomorphism between $`G^+`$ and $`F^+`$, and carries the time infinity of $`G^+`$ onto $`F_0`$. If $`f`$ is a transformation on $`G^+`$, we define $`\widehat{f}:F^+F^+`$ by setting $`\widehat{f}:=hfh^1`$. The function $`\widehat{f}`$ is a causal homeomorphism on $`F^+`$, i.e., $`P_S(u)<P_S(w)`$ implies $`P_S[\widehat{f}(u)]<P_S[\widehat{f}(w)]`$. If $`g`$ is a homeomorphism on $`F^+`$ and is regular in $`F_0`$, we define $`g_0:VV`$ by setting $`g_0(𝐯):=P_V[lim_{(s,𝐯^{})(0,𝐯)}g(s,𝐯^{})]`$; moreover, $`P_S[lim_{(s,𝐯^{})(0,𝐯)}g(s,𝐯^{})]=0`$, because the limit must be an accumulation point for $`F^+`$, but not belonging to it. The connection between the asymptotic behavior of a transformation $`f`$ on $`G^+`$ and the behavior of the transformation $`\widehat{f}`$ on $`F^+`$ in the proximity of $`F_0`$ is described by the following lemma: ###### Lemma 3 $`f`$ is asymptotically regular if and only if $`\widehat{f}`$ is regular in $`F_0`$, and the relation $`f^+=\widehat{f}_0`$ holds. Lemma 1 allows us to obtain results regarding the asymptotic behavior of a homeomorphism on $`G^+`$ by studying the behavior of a homeomorphism on $`F^+`$ in the proximity of $`F_0`$. If $`g`$ is a homeomorphism on $`F^+`$ which is regular in $`F_0`$, we define $`\overline{g}:F_0^+F_0^+`$, setting $$\overline{g}(s,𝐯):=\{\begin{array}{cc}g(s,𝐯)\text{ if }s>0\hfill & \\ (0,g_0(𝐯))\text{ if }s=0\hfill & \end{array}.$$ We have the following lemma: ###### Lemma 4 (a) If $`g`$ is a homeomorfism on $`F^+`$ and it is regular in $`F_0`$, then the functions $`g_0`$ and $`\overline{g}`$ are continuous. (b) if $`g`$ and $`g^1`$ are regular in $`F_0`$, then $`g_0`$ is invertible and the relation $`(g_0)^1=(g^1)_0`$ holds; furthermore, $`\overline{g}`$ is a homeomorphism. Proof. We only give the proof that $`g_0`$ is continuous. Consider a sequence $`𝐯_n𝐯_0V`$. Due to the regularity of $`g`$ at every point $`(0,𝐯_n)`$, for every $`n`$ it is possible to find a point $`(s_n,𝐰_n)F^+`$ such that $`(s_n,𝐰_n)(0,𝐯_n)1/n`$ and $`g(s_n,𝐰_n)(0,g_0(𝐯_n))1/n`$. Therefore $`(s_n,𝐰_n)(0,𝐯_0)`$ and $`g(s_n,𝐰_n)(0,g_0(𝐯_0))`$. Since $`g_0(𝐯_n)g_0(𝐯_0)=(0,g_0(𝐯_n))(0,g_0(𝐯_0))=(0,g_0(𝐯_n))(0,g_0(𝐯_0))+g(s_n,𝐰_n)g(s_n,𝐰_n)g(s_n,𝐰_n)(0,g_0(𝐯_n))+g(s_n,𝐰_n)(0,g_0(𝐯_0))`$, on also has that $`g_0(𝐯_n)g_0(𝐯_0)`$. QED. By combining Lemma 3 with Lemma 4, and by taking into account that $`(f^1)\widehat{}=(\widehat{f})^1`$ and that $`(gf)\widehat{}=\widehat{g}\widehat{f}`$, one can prove Theorem 1. The details of the proof are omitted. As in the nonasymptotic case, it is possible for a transformation to be asymptotically regular while its inverse is not. Consider for instance the transformation $`f(t,𝐱):=(t,𝐱/t)`$, for which $`f^+(𝐯)=0`$ holds, and whose inverse $`f^1(t,𝐱)=(t,t𝐱)`$ is not asymptotically regular at any point. It does not appear to be trivial to find a rigorous proof to the following reasonable statement, which is therefore proposed here as a conjecture: ###### Conjecture 1 If $`f`$ is asymptotically regular and $`f^+`$ is a homeomorphism on $`V`$, then $`f`$ is an asymptotic homeomorphism. Hereafter, unless otherwise stated, the term transformation will be used to designate a causal transformation which is also an asymptotic homeomorphism. #### 2.1.4 Aymptotically Euclidean transformations. The following classes of transformations are particularly important: ###### Definition 4 (a) An asymptotically identical transformation is a transformation $`f`$ such that $`f^+(𝐯)=𝐯`$. (b) An asymptotically Euclidean transformation is a transformation $`f`$ such that $`f^+(𝐯)=R𝐯𝐯_0`$. ###### Lemma 5 (a) Asymptotically identical transformations and asymptotically Euclidean transformations are groups. (b) An asymptotically Euclidean transformation $`f`$ can univocally be factorized as $`f_1f_2f_3`$, where $`f_3`$ is a rotation, $`f_2`$ is asymptotically identical and $`f_1`$ is a boost. (c) If $`\gamma _1`$ and $`\gamma _2`$ are two semitrajectories, the two following statements are equivalent: (i) $`\omega (\gamma _1)=\omega (\gamma _2)`$; (ii) there exists an asymptotically identical transformation $`f`$ such that $`\gamma _2=f(\gamma _1)`$. #### 2.1.5 N-bigbang. Let us now consider the case of $`N`$ semitrajectories, whose particularities with respect to the case of a single semitrajectory it is convenient to note. Furthermore, for reasons that will become clear later, let us require that the N semitrajectories share a common origin. So let us give the following definition: ###### Definition 5 N-bigbang is the union of $`N`$ semitrajectories that are disjoined everywhere but in the origin, which is common. The term N-bigbang derives from the fact that, at the origin, all the particles are concentrated at a single point. The reason for requiring the semitrajectories to be disjoined everywhere else is that this allows to demonstrate the following Theorem 2. All the transformations on $`G`$ preserve the N-bigbangs, since they can neither separate the semitrajectories at the origin nor transform disjoined trajectories into intersecting trajectories. The notion of N-bigbang is relevant to this paper because it will represent the ideal model of universe that will be proposed. Let $`B`$ be the set of N-bigbangs. The asymptotic velocity of an N-bigbang is a vector belonging to $`V^N`$. We again use the symbol $`\omega `$ to indicate the application $`\omega :BV^N`$ which associates with every N-bigbang its asymptotic velocity. If $`g`$ is a transformation on $`V`$ and $`v=(𝐯_\mathrm{𝟏},\mathrm{},𝐯_N)V^N`$, then $`g(v)`$ is the vector $`(g(𝐯_1),\mathrm{},g(𝐯_N))`$; if $`𝐯_0V`$, then $`Rv𝐯_0`$ is the vector $`(R𝐯_1𝐯_0,\mathrm{},R𝐯_N𝐯_0)`$. The following theorem that holds for N-bigbangs is analogous to Lemma 5c for semitrajectories. ###### Theorem 2 Let $`\beta _1`$ and $`\beta _2`$ be two N-bigbangs. The following two statements are equivalent: (i) $`\omega (\beta _1)=\omega (\beta _2)`$; (ii) There exists an asymptotically identical transformation $`f`$ such that $`\beta _2=f(\beta _1)`$. Proof. The implication (ii)$``$ (i) is obvious. An intuitive proof of the implication (i)$``$ (ii) is given. By using the induction principle, let us suppose that the theorem holds for an N-bigbang and prove that it also holds for an (N+1)-bigbang. Let $`\beta `$ and $`\beta ^{}`$ be two (N+1)-bigbangs for which the equation $`\omega (\beta )=\omega (\beta ^{})`$ holds. Due to our hypothesis of induction, there exists a transformation $`f`$ such that $`\gamma _i^{}=f(\gamma _i)`$, for $`i=1,\mathrm{},N`$, where $`\gamma _i`$ and $`\gamma _i^{}`$ are the semitrajectories that form the (N+1)-bigbangs. It is therefore sufficient to prove that given an N-bigbang $`\beta `$ and two semitrajectories $`\gamma `$ and $`\gamma ^{}`$ such that $`\omega (\gamma )=\omega (\gamma ^{})`$, there exists an asymptotically identical transformation $`f`$ such that $`f(\beta )=\beta `$ and $`f(\gamma )=\gamma ^{}`$. To intuitively accept the existence of such a transformation, consider two points $`\gamma (t)`$ and $`\gamma ^{}(t)`$ which evolve in three-dimensional space along with the segment that joins them. If the segment intersects one of the other points of the N-bigbang, it is allowed to fold slightly in order to avoid it (note that in two-dimensional space this is not possible). Therefore, there exists at every given instant a broken line which joins the two points and whose length is less than $`2\gamma (t)\gamma ^{}(t)`$. Consider now the “tube” of radius $`ϵ`$ that surrounds the segment, i.e. the set of points whose distance from the segment is less than $`ϵ`$ . At every instant it is possible to determine $`ϵ`$ so that the tube does not intersect any of the remaining $`N`$ points. Finally, consider for every $`t`$ a space transformation $`f_X(t,)`$ which keeps the space outside the tube unmodified and deforms it inside the tube, bringing the point $`\gamma (t)`$ onto the point $`\gamma ^{}(t)`$. The space-time transformation required will therefore be $`f(t,𝐱):=(t,f_X(t,𝐱))`$. Since for every $`t`$ the relation $`f_X(t,𝐱)𝐱2\left(\gamma (t)\gamma ^{}(t)+ϵ\right)`$ holds, one can easily prove that $`f`$ is asymptotically identical. QED. #### 2.1.6 Asymptotic intervals. This section proves a theorem and a corollary which will be subsequently used to demonstrate the invariance of the asymptotic quantum measure under AET. First of all, let us introduce some notations: if $`\mathrm{\Delta }V`$, let us define $`\mathrm{\Delta }^c:=\{(t,𝐯t)G|t0,𝐯\mathrm{\Delta }\}`$; in particular, if $`𝐯V`$, then $`𝐯^c:=\{𝐯\}^c`$ is the semitrajectory $`\{𝐯t\}_{t0}`$. One can easily see that if $`f`$ is asymptotically identical, then $`f(𝐯^c)(t)𝐯^c(t)/t0`$ for $`t+\mathrm{}`$ (one should bear in mind that an asymptotically identical transformation preserves the asymptotic limit of semitrajectories). The following theorem extends this property: ###### Theorem 3 Let $`\mathrm{\Delta }V`$ be a bounded set and let $`f`$ be an asymptotically identical transformation. Then $`sup_{𝐯\mathrm{\Delta }}f(𝐯^c)(t)𝐯^c(t)/t0`$ for $`t+\mathrm{}`$. Proof. In this case as well, for the sake of simplicity, let us think of $`f`$ as a causal transformation on $`G^+`$. Since $`sup_{𝐯\mathrm{\Delta }}f(𝐯^c)(t)𝐯^c(t)sup_{𝐯\overline{\mathrm{\Delta }}}f(𝐯^c)(t)𝐯^c(t)`$, we can directly assume that $`\mathrm{\Delta }`$ is compact. Let us set $`d(t):=f(𝐯^c)(t)𝐯^c(t)/t`$. Obviously, $`d(t)0`$ if and only if $`d(f_T(t))0`$. Due to Lemma 2d, one finds that $`d(f_T(t))=sup_{𝐯\mathrm{\Delta }}f_X(t,𝐯t)/f_T(t)𝐯`$. Since $`\widehat{f}(1/t,𝐯)=(1/f_T(t),f_X(t,𝐯t)/f_T(t))`$ (where $`\widehat{f}`$ is defined in section 2.1.3), therefore $`d(f_T(t))=sup_{𝐯\mathrm{\Delta }}g_{(_1/t)}(𝐯)𝐯`$, where $`g_s(𝐯):=P_V[\widehat{f}(s,𝐯)]`$. Due to the well known property of uniformly convergent functions, if one proves that for $`s0`$ the function $`g_s(𝐯)`$ uniformly tends to $`𝐯`$ in $`\mathrm{\Delta }`$ , the theorem is proven. In order to prove the uniform convergence of $`g_s(𝐯)`$, we use the fact that a continuous function on a compact set is also uniformly continuous. From Lemma 4a, the function $$\overline{g}(s,𝐯):=\{\begin{array}{cc}\widehat{f}(s,𝐯)\text{ for }s>0\hfill & \\ (0,𝐯)\text{ for }s=0\hfill & \end{array}$$ is continuous on $`F_0^+`$, and therefore uniformly continuous on $`[0,c]\times \mathrm{\Delta }F_0^+`$, where $`c`$ is any constant greater than $`0`$. Hence, for any fixed $`ϵ`$, there exists a $`\delta `$ such that if $`(s_1,𝐯_1),(s_2,𝐯_2)[0,c]\times \mathrm{\Delta }`$ and $`(s_1,𝐯_1)(s_2,𝐯_2)<\delta `$ then $`\overline{g}(s_1,𝐯_1)\overline{g}(s_2,𝐯_2)<ϵ`$. In particular, if one chooses $`s_2=0`$ and $`𝐯_1=𝐯_2=𝐯`$, if $`(s,𝐯)(0,𝐯)=s<\delta `$, then $`\overline{g}(s,𝐯)(0,𝐯)<ϵ`$, $`𝐯\mathrm{\Delta }`$. Since $`g_s(𝐯)𝐯\overline{g}(s,𝐯)(0,𝐯)`$, the theorem is proven. QED. Given $`𝐚,𝐛V`$, with $`a_i<b_i`$, $`i=1,2,3`$, let $`I:=(𝐚,𝐛]=(a_1,b_1]\times (a_2,b_2]\times (a_3,b_3]`$ be a half-open interval of $`V`$; furthermore, if $`0<ϵ<\mathrm{min}\{(b_1a_1),(b_2a_2),(b_3a_3)\}/2`$, let $`I_ϵ`$ and $`I_{+ϵ}`$ be the intervals $`(𝐚+ϵ,𝐛ϵ]`$ and $`(𝐚ϵ,𝐛+ϵ]`$, respectively. ###### Corollary 1 Let $`f`$ be asymptotically identical and let $`I=(𝐚,𝐛]`$ be an interval of $`V`$. For every $`ϵ`$ such that $`0<ϵ<\mathrm{min}\{(b_1a_1),(b_2a_2),(b_3a_3)\}/2`$, there exists a $`t_0`$ such that if $`tt_0`$, then $`I_ϵ^c(t)f(I^c)(t)I_{+ϵ}^c(t)`$. Proof. From Theorem 3, for every $`ϵ>0`$ there exists a $`t_0`$ such that for $`tt_0`$ the relation $`sup_{𝐯I}f(𝐯^c)(t)𝐯^c(t)ϵ`$ holds. This means that $`I_ϵ^c(t)f(I^c)(t)I_{+ϵ}^c(t)`$. QED. The above corollary can easily be generalized to the intervals of $`V^N`$. ### 2.2 Classical Trajectories. Suppose that $`N`$ masses $`m_1,\mathrm{},m_N`$ and $`N(N1)/2`$ potentials $`V_{i,j}(r)=V_{j,i}(r)`$, $`i,j1,\mathrm{},N`$ are given which we assume to be bounded, differentiable and vanishing for $`r\mathrm{}`$ with standard conditions (which include long-range potentials such as the Coulomb potential; see for details). ###### Definition 6 An N-bigbang (not necessarily an asymptotically regular one) is classical if it satisfies the equations of motion deriving from the Lagrangian $$L(t)=\underset{i=1}{\overset{N}{}}\frac{1}{2}m_i\dot{\gamma }_i^2(t)+\underset{i<j}{}V_{i,j}\left(\gamma _i(t)\gamma _j(t)\right),$$ (1) where $`\gamma _i`$ are the semitrajectories that make up the N-bigbang. Let $`B_C`$ be the set of the classical N-bigbangs. Obviously, the set $`B_C`$ is invariant under Galilean transformations. The following theorem is derived from an important theorem of classical scattering theory, which states that classical trajectories are asymptotically regular: ###### Theorem 4 Classical N-bigbangs are asymptotically regular. Proof. See . In physical terms, particles asymptotically tend to separate into a certain number of independent clusters. The velocities of the centers of mass of the clusters tend to a constant value. The positions of the particles within each individual cluster with respect to the center of mass of the cluster are bounded or almost bounded, and in any case their asymptotic velocity is the limit velocity of the cluster’s center of mass. Let us set $`\mathrm{\Delta }_C:=\omega (B_C)V^N`$. The set $`\mathrm{\Delta }_C`$ is invariant under Euclidean transformations on $`V^N`$. In general, the equation $`\mathrm{\Delta }_C=V^N`$ does not hold (see next section). It is useful to define on $`B_C`$ the following equivalence relations: ###### Definition 7 (a) Two classical N-bigbangs $`\beta _1`$ and $`\beta _2`$ are said to be G-equivalent if there exists on $`G`$ a Galilean transformation $`f`$ such that $`\beta _2=f(\beta _1)`$. (b) Two classical N-bigbangs $`\beta _1`$ and $`\beta _2`$ are said to be $`\omega E`$-equivalent if there exists on $`V^N`$ a Euclidean transformation $`g`$ such that $`\omega (\beta _2)=g[\omega (\beta _1)]`$. In general, there can be $`\omega E`$-equivalent N-bigbangs which are not $`G`$-equivalent. In particular, there can be N-bigbangs with the same asymptotic velocity which are not connected by a translation. This can be easily understood if one thinks of asymptotic velocity as the limit of the proper boundary condition of Hamilton’s action principle (see next section). The operation of taking the limit for $`t+\mathrm{}`$ induces a further degeneration, due to the fact that all the bound states of a cluster merge into the same asymptotic velocity. #### 2.2.1 Asymptotic velocities and asymptotic boundary conditions. Consider the classical N-bigbangs with origin at the point $`(0,0)`$ and traveling through the point $`(t,x),xR^{3N}`$. These are the proper boundary conditions of Hamilton’s action principle, and they are known to generally define more than one trajectory. Suppose we set $`x=vt`$ and take $`t`$ to infinity. We will say that the velocity $`v`$ was set as asymptotic boundary condition for the N-bigbang. The following questions naturally arise: In which domain $`\mathrm{\Delta }_BV^N`$ are the asymptotic boundary conditions defined, i.e., for which values of $`v`$ can one take the limit? Is the asymptotic velocity of the N-bigbangs obtained in this way equal to $`v`$? What happens if $`v\overline{)}\mathrm{\Delta }_C`$? The following simple example helps to show what might happen. Let us consider, for the sake of simplicity, a one-dimensional particle of mass m subject to the potential $$V(x)=\{\begin{array}{cc}V_0>0\text{ for }|x|a\hfill & \\ 0\text{ for }r>a\hfill & \end{array}.$$ (2) The semitrajectory $`x(t)`$ that starts at $`(0,0)`$ with the initial velocity $`v_I>0`$ is $$x(t)=\{\begin{array}{cc}v_It\text{ for }ta/v_I\hfill & \\ a+(ta/v_I)\sqrt{v_I^2+2V_0/m}\text{ for }t>a/v_I\hfill & \end{array}.$$ (3) An analogous equation defines the trajectory with $`v_I<0`$, while if $`v_I=0`$ the trajectory is simply $`x(t)=0`$. One can easily see that $`\mathrm{\Delta }_C=(\mathrm{},\sqrt{2V_0/m})\{0\}(\sqrt{2V_0/m},+\mathrm{})`$. Let us now set the boundary condition $`x(t)=vt`$, then we will take $`t`$ to infinity and we will compute $`v_I`$ as a function of $`v`$. If $`v>0`$, by setting the boundary condition $`x(t)=vt`$, for large enough values of $`t`$ one obtains from (3) the equation $$a+(ta/v_I)\sqrt{v_I^2+2V_0/m}=vt,$$ (4) which is a fourth-degree equation in $`v_I`$. One can easily obtain a solution for $`t+\mathrm{}`$ if one assumes that $`v_I(t)v_I\mathrm{}`$ , i.e., if one allows that $`v_I(t)`$ admits a limit, for $`t+\mathrm{}`$. One has to distinguish between two cases: $`v_I\mathrm{}0`$ and $`v_I\mathrm{}=0`$. In the former, for very large values of $`t`$ one can disregard $`a`$ and $`a/v_I`$ as compared to $`t`$ in expression (4), and obtain $$\sqrt{v_I^2+2V_0/m}=v,$$ hence $$v_I\mathrm{}=\sqrt{v^22V_0/m}.$$ (5) This solution is acceptable only if $`v>\sqrt{2V_0/m}`$. By replacing $`v_I\mathrm{}`$ in expression (3), one finds that the asymptotic velocity of the trajectory is $`v`$ and is therefore equal to the asymptotic boundary condition. In the case $`v_I\mathrm{}=0`$, again for very large values of $`t`$, one can disregard the term $`v_I^2`$ with respect to the term $`2V_0/m`$ in expression (4). One obtains $$a+(ta/v_I)\sqrt{2V_0/m}=vt,$$ hence $$v_I=\frac{a\sqrt{2V_0/m}}{t\left(\sqrt{2V_0/m}v\right)+a}.$$ (6) This solution, which holds only for $`0v<\sqrt{2V_0/m}`$ because $`v_I`$ must be $`0`$, confirms the correctness of the Ansatz $`v_I\mathrm{}=0`$. The case $`v=\sqrt{2V_0/m}`$ is a limit case, and by straightforward reasoning one can deduce that it forces $`v_I\mathrm{}`$ to be null. In a similar way one can calculate the dependence of $`v_I\mathrm{}`$ upon $`v`$ in the case of $`v<0`$, while $`v=0`$ trivially implies $`v_I\mathrm{}=0`$. In conclusion, one can say that in this example any value of the velocity $`v`$ is allowed as asymptotic boundary condition; therefore the set of asymptotic boundary conditions $`\mathrm{\Delta }_B`$ is different and wider than the set of asymptotic velocities $`\mathrm{\Delta }_C`$. Any value $`|v|>\sqrt{2V_0/m}`$ determines a trajectory with asymptotic velocity $`v`$, while all the values $`|v|\sqrt{2V_0/m}`$ determine the same trajectory with null initial velocity and null asymptotic velocity. One could therefore say that for $`v_I=0`$ there is a “degeneration” of the asymptotic boundary conditions. Notice that the point $`v_I=0`$ is a discontinuity point of the application $`\omega _V:V_IV`$ that associates the initial velocity with the asymptotic velocity: $$\omega _V(v_I)=\{\begin{array}{cc}\text{sign }(v_I)(|v_I|+\sqrt{2V_0/m})\text{ for }v_I0\hfill & \\ 0\text{ if }v_I=0\hfill & \end{array}.$$ (7) This simple example gives a glimpse of the possibility of developing an interesting theory on asymptotic boundary conditions. In particular, this theory should: (i) Provide a definition of asymptotic boundary conditions which is more rigorous and general than the one provided above, which is likely to become inadequate in the presence of bound states; (ii) Define the existence domain $`\mathrm{\Delta }_B`$ of the asymptotic boundary conditions <sup>2</sup><sup>2</sup>2One thing that can be easily said about $`\mathrm{\Delta }_B`$ is that it is invariant for Euclidean transformations on $`V^N`$. ; (iii) Describe the relation between asymptotic velocity and asymptotic boundary condition, proving for example that if the asymptotic boundary condition belongs to $`\mathrm{\Delta }_C`$, then it is equal to the asymptotic velocity; (iv) Describe the relation between degeneration points of asymptotic boundary conditions and discontinuity points of the function $`\omega _V`$. I am not aware of the existence of such a theory, and it will not be developed in this paper. ### 2.3 Asymptotic quantum measure. This section is devoted to quantum mechanics. The first part defines and studies quantum asymptotic velocity. This operator will be used for the definition of the asymptotic quantum measure in the second part of this section. The third part will prove the invariance of asymptotic quantum measure under AET, which is probably the main mathematical result of this paper. Consider an N-particle quantum system. Its state is described by the vector $`\psi L^2(R^{3N})`$; $`Q=(𝐐_1,\mathrm{},𝐐_N)=(Q_{1x},Q_{1y},Q_{1z},\mathrm{},Q_{Nx},Q_{Ny},Q_{Nz})`$ is the position vector operator and $`P`$ (of analogous structure) is the momentum operator. The Hamiltonian is $$H=\underset{i=1}{\overset{N}{}}\frac{𝐏_i^2}{2m_i}+\underset{i<j}{}V_{i,j}(𝐐_i𝐐_j),$$ (8) where the same considerations of section 2.2 apply to the potentials $`V_{i,j}`$. #### 2.3.1 Quantum asymptotic velocity ###### Definition 8 We call quantum asymptotic velocity the vector operator $$V^+:=s\underset{t+\mathrm{}}{lim}e^{iHt}\frac{Q}{t}e^{iHt}.$$ (9) Quantum asymptotic velocity also is an important concept used in quantum scattering theory, and its existence is assured by a theorem which is analogous to the classical case: ###### Theorem 5 The limit (9) exists for a dense subset of $`L^2(R^{3N})`$. $`V^+`$ is a vector of Hermitian operators which commute with one another and with the Hamiltonian. Proof. See . Let us give the explicit expression for $`V^+`$ in two simple cases. For a free particle one has: $$𝐕^+=\frac{𝐏}{m}.$$ (10) One can easily obtain equation (10) by using the relation $`e^ABe^A=B+[A,B]+\frac{1}{2!}[A,[A,B]]+\mathrm{}`$. The physical interpretation of this result is simple: for very large times, measuring the position of a particle at the time $`t`$ and dividing it by $`t`$ is equivalent to measuring its velocity. For a particle in a central potential which admits the Møller operators $`\mathrm{\Omega }^\pm `$, one has $$𝐕^+=\mathrm{\Omega }^{}\frac{𝐏}{m}\mathrm{\Omega }^{}E_S,$$ (11) where $`E_S`$ is the projector over the scattering states of the Hamiltonian. See for the proof. Note that the bounded states of the Hamiltonian belong to the eigenvalue $`0`$ of $`𝐕^+`$. This result holds in general. We derive now the transformation rules for asymptotic velocity. First of all, let us summarize the Galilean transformations and their generators for an N-particle quantum system: (i) Time translations: $`e^{itH}`$. (ii) Space translations: $`e^{i𝐏𝐱}`$, where $`𝐏=_i𝐏_i`$ is the total momentum; one has: $`e^{i𝐏𝐱}Qe^{i𝐏𝐱}=Q+𝐱`$. The momentum and the Hamiltonian are invariant under space translations. (iii) Rotations: $`e^{i𝐋\alpha }`$, where $`𝐋=_i𝐋_i`$ is the total angular momentum and $`\alpha `$ is the vector associated with a rotation; one has $`e^{i𝐋\alpha }Qe^{i𝐋\alpha }=R_\alpha Q`$, where $`R_\alpha `$ is the orthogonal matrix associated with the vector $`\alpha `$ . The Hamiltonian is invariant under rotations. (iv) Boosts: $`e^{im𝐐𝐯}`$, where $`m𝐐=_im_i𝐐_i`$; we have $`e^{im𝐐𝐯}𝐏_ie^{im𝐐𝐯}=𝐏_im_i𝐯`$ and $`e^{im𝐐𝐯}He^{im𝐐𝐯}=H+𝐯𝐏+\frac{1}{2}𝐯^2_im_i`$. ###### Lemma 6 The operators $`V^+`$ have the following transformation rules: $`e^{iHt}V^+e^{iHt}`$ $`=`$ $`V^+,`$ $`e^{i𝐏𝐱}V^+e^{i𝐏𝐱}`$ $`=`$ $`V^+,`$ (12) $`e^{i𝐋\alpha }V^+e^{i𝐋\alpha }`$ $`=`$ $`R_\alpha V^+,`$ $`e^{im𝐐𝐯}V^+e^{im𝐐𝐯}`$ $`=`$ $`V^+𝐯.`$ Proof. The proof can be obtained by simple calculations based on the transformation rules of $`Q`$, $`P`$, and $`H`$, and taking into account that $`V^+`$ commutes with the Hamiltonian. #### 2.3.2 Asymptotic quantum measure. Let $`E^+(\mathrm{\Delta })`$ be the spectral measure on $`V^N`$ associated with the operators $`V^+`$, where $`\mathrm{\Delta }`$ belongs to the Borel $`\sigma `$-algebra $``$ of $`V^N`$. Choosing a state $`\psi `$, one can induce on $`V^N`$ a measure $`\mu `$ by defining $`\mu (\mathrm{\Delta }):=\psi |E^+(\mathrm{\Delta })|\psi `$. The state we choose to define the quantum measure $`\mu _Q`$ is the following: if $`𝐱R^3`$, $`𝐱_N`$ is the vector $`(𝐱_1,\mathrm{},𝐱_N)R^{3N}`$ such that $`𝐱_i=𝐱,i=1,\mathrm{},N`$. The vector $`𝐱_N`$ describes all the particles concentrated at the point $`𝐱`$. The state $`|𝐱_N`$ is the improper eigenvector with eigenvalue $`𝐱_N`$ of the position operator $`Q`$. ###### Definition 9 The asymptotic quantum measure (or more simply quantum measure) is the measure $`\mu _Q`$ on $`V^N`$ defined as $$\mu _Q(\mathrm{\Delta }):=𝐱_N|E^+(\mathrm{\Delta })|𝐱_N.$$ (13) Example. One can easily calculate $`\mu _Q`$ when there is no potential and the equation $`V^+=(𝐏_1/m_1,\mathrm{},𝐏_N/m_N)`$ holds. If $`E^P()`$ is the spectral measure of $`P`$, then $`E^+(\mathrm{\Delta })=E^P(m\mathrm{\Delta })`$, where $`m\mathrm{\Delta }:=\{(m_1𝐯_1,\mathrm{},m_N𝐯_N)R^{3N}|(𝐯_1,\mathrm{},𝐯_N)\mathrm{\Delta }\}`$. Therefore $`\mu _Q(\mathrm{\Delta })`$ $`=`$ $`𝐱_N|E^+(\mathrm{\Delta })|𝐱_N={\displaystyle 𝐱_N|p_1𝑑p_1p_1|E^P(m\mathrm{\Delta })|p_2𝑑p_2p_2|𝐱_N}=`$ $`=`$ $`{\displaystyle \frac{1}{(2\pi )^{3N}}}{\displaystyle \mathrm{exp}[i𝐱_N(p_2p_2)]\chi _{m\mathrm{\Delta }}(p_1)\delta (p_2p_1)𝑑p_1𝑑p_2}=`$ $`=`$ $`{\displaystyle \frac{1}{(2\pi )^{3N}}}{\displaystyle \chi _{m\mathrm{\Delta }}(p)𝑑p}={\displaystyle \frac{\mu _L(m\mathrm{\Delta })}{(2\pi )^{3N}}},`$ where $`\chi _{m\mathrm{\Delta }}`$ is the characteristic function of the set $`m\mathrm{\Delta }`$ and $`\mu _L`$ is the Lebesgue measure. The measure $`\mu _Q`$ in the absence of a potential is therefore proportional to the Lebesgue measure on the momentum space. The following lemma describe the invariance properties of the measure $`\mu _Q`$: ###### Lemma 7 (a) $`\mu _Q`$ does not depend on the point $`𝐱`$. (b) $`\mu _Q`$ is invariant under Euclidean transformations on $`V^N`$, i.e., $`\mu _Q(\{R\mathrm{\Delta }𝐯_0\})=\mu _Q(\mathrm{\Delta })`$, where $`\{R\mathrm{\Delta }𝐯_0\}:=\{(Rv𝐯_0)V^N|v\mathrm{\Delta }\}`$. Proof. Here also the proof can be obtained by simple calculations based on the transformation rules of $`V^+`$. The invariance of the measure $`\mu _Q`$ under Euclidean transformations allows to define a quotient measure $`\stackrel{~}{\mu }_Q`$ on the quotient $`\sigma `$-algebra $`\stackrel{~}{}`$, which is formed by the measurable sets of $`V^N`$ which are also invariant under Euclidean transformations. See the Appendix. #### 2.3.3 Invariance of the quantum measure for asymptotically Euclidean transformations. ###### Lemma 8 The equation $$\mu _Q(\mathrm{\Delta })=\underset{t+\mathrm{}}{lim}\underset{\mathrm{\Delta }^c(t)}{}|K(𝐱_N,y,t)|^2dy$$ (15) holds, where $`K(𝐱_N,y,t)`$ is the Feynman propagator between points $`(0,𝐱_N)`$ and $`(t,x)`$, and where $`\mathrm{\Delta }^c`$ has been defined in section 2.1.6. Proof. If $`E_x^Q`$ is used to indicate the spectral family of $`Q`$, the equation $$e^{iHt}\frac{Q}{t}e^{iHt}=e^{iHt}\frac{1}{t}\left[x𝑑E_x^Q\right]e^{iHt}=vd(e^{iHt}E_{vt}^Qe^{iHt})$$ allows one to deduce that, for the spectral family $`E_v^+`$ of $`V^+`$, the equation $$E_v^+=s\underset{t+\mathrm{}}{lim}e^{iHt}E_{vt}^Qe^{iHt}$$ (16) holds, and therefore for the spectral measure $`E^+(\mathrm{\Delta })`$ one has that $$E^+(\mathrm{\Delta })=s\underset{t+\mathrm{}}{lim}e^{iHt}E^Q[\mathrm{\Delta }^c(t)]e^{iHt}$$ (17) holds. Equation (15) can be easily derived from equations (13) and (17). QED. By expressing the Feynman propagator in terms of a sum over paths, one has $$\mu _Q(\mathrm{\Delta })=\underset{t+\mathrm{}}{lim}\underset{\mathrm{\Delta }^c(t)}{}dy\underset{(0,𝐱_N)}{\overset{(t,y)}{}}𝒟\beta _1𝒟\beta _2S(\beta _1)S^{}(\beta _2),$$ (18) where $`\beta _1`$ and $`\beta _2`$ are two paths with extremes $`(0,𝐱_N)`$ and $`(t,y)`$, and $$S(\beta _i):=\mathrm{exp}\left[i_0^tL(\beta _i(\tau ),\dot{\beta }_i(\tau ))d\tau \right],$$ where $`L`$ is the Lagrangian of the system. The following theorem establishes the invariance of the quantum measure under AET: ###### Theorem 6 If $`f`$ is an asymptotically Euclidean transformation, then the equation $$\mu _Q(\mathrm{\Delta })=\underset{t+\mathrm{}}{lim}\underset{\mathrm{\Delta }^c(t)}{}dy\underset{(0,𝐱_N)}{\overset{(t,y)}{}}𝒟\beta _1𝒟\beta _2S[f(\beta _1)]S^{}[f(\beta _2)]$$ (19) holds. Proof. Since we have already proved the invariance of $`\mu _Q`$ under translations of the point $`𝐱`$ and under Galilean transformations, for the sake of simplicity we demonstrate equation (19) only in the case of a transformation $`f`$ which is asymptotically identical and leaves the point $`(0,𝐱)`$ unmodified. One has $`{\displaystyle \underset{\mathrm{\Delta }^c(t)}{}}dy{\displaystyle \underset{(0,𝐱_N)}{\overset{(t,y)}{}}}𝒟\beta _1𝒟\beta _2S[f(\beta _1)]S^{}[f(\beta _2)]=`$ $`{\displaystyle \underset{f_X[t,\mathrm{\Delta }^c(t)]}{}}dy{\displaystyle \underset{(0,𝐱_N)}{\overset{(f_T(t),y)}{}}}𝒟\beta _1𝒟\beta _2S(\beta _1)S^{}(\beta _2)=`$ (20) $`{\displaystyle \underset{f(\mathrm{\Delta }^c)[f_T(t)]}{}}|K(𝐱_N,y,f_T(t))|^2dy,`$ where $`f_T`$ and $`f_X`$ were defined in section 2.1.2, and Lemma 2d has been used. If we disregard border effects, which as we will see are nullified by the condition of asymptotic identity, equations (2.3.3) point out the invariance of the Feynman path integrals under a generic causal homeomorphism. Since $`f_T(t)`$ is monotonically increasing, in order to prove equation (19) is enough to prove that $$\underset{t+\mathrm{}}{lim}\underset{f(\mathrm{\Delta }^c)(t)}{}|K(𝐱_N,y,t)|^2dy=\underset{t+\mathrm{}}{lim}\underset{\mathrm{\Delta }^c(t)}{}|K(𝐱_N,y,t)|^2dy.$$ Let $`I`$ be a half-open interval of $`V^N`$, as defined in section 2.1.6. For the sake of brevity, let $$\mu _{Qt}(I):=\underset{I^c(t)}{}|K(𝐱_N,y,t)|^2dy.$$ Owing to Corollary 1, one has, for large enough values of $`t`$, $$\mu _{Qt}(I_ϵ)\mu _{Qt}[f(I)]\mu _{Qt}(I_{+ϵ});$$ if one takes the limit for $`t+\mathrm{}`$, $$\mu _Q(I_ϵ)\underset{t+\mathrm{}}{lim\; inf}\mu _{Qt}[f(I)]\underset{t+\mathrm{}}{lim\; sup}\mu _{Qt}[f(I)]\mu _Q(I_{+ϵ})$$ (21) Furthermore, one has: $$\underset{ϵ0}{lim}\left[\mu _Q(I_{+ϵ})\mu _Q(I_ϵ)\right]=\underset{ϵ0}{lim}\mu _Q\left[I_{+ϵ}I_ϵ\right]=\mu _Q\left[\underset{ϵ>0}{}(I_{+ϵ}I_ϵ)\right]=\mu _Q(I).$$ (22) where $`I`$ is the boundary of the interval $`I`$. The Lebesgue measure of $`I`$ is null, but this does not ensure that the quantum measure is null as well, because it could be singular with respect to the Lebesgue measure. However, one can prove that $`\mu _Q(I)=0`$ if the measure $`\mu _Q`$ is $`\sigma `$-finite; we will assume this without proof. Consider a single side of the interval $`I`$, for instance $`D:=\{a_{1x}\}\times (a_{1y},b_{1y}]\times \mathrm{}\times (a_{Nz},b_{Nz}]`$, and suppose that $`\mu _Q(D)>0`$. The set $`D_v:=\{D(v,0,0)\}`$ is obtained from set $`D`$ by means of a translation along the x-axis; therefore, due to the invariance of $`\mu _Q`$, one has that $`\mu _Q(D)=\mu _Q(D_v)`$. If $`(v_n,0,0)`$ is a bounded sequence of elements of $`V`$ with distinct values, one has that $`_nD_{v_n}`$ is a bounded set; however, since all the $`D_{v_n}`$ are disjoined, one has that $`\mu _Q(_nD_{v_n})=_n\mu _Q(D_{v_n})=+\mathrm{}`$ , which is in contrast with the hypothesis that $`\mu _Q`$ is $`\sigma `$-finite. If $`\mu _Q(I)=0`$, then from relation (22), and since $`I_ϵII_{+ϵ}`$ , one obtains $$\underset{ϵ0}{lim}\mu _Q(I_{+ϵ})=\underset{ϵ0}{lim}\mu _Q(I_ϵ)=\mu _Q(I),$$ while from relation (21) one obtains $$\underset{t+\mathrm{}}{lim}\mu _{Qt}[f(I)]=\mu _Q(I).$$ (23) We have demonstrated that equation (19) holds when $`\mathrm{\Delta }`$ is an half-open interval of $`V^N`$. The demonstration is completed by using the fact that according to a well-known theorem of measure theory, if two measures on $`R^N`$ agree on half-open intervals, then they are equal. QED. ### 2.4 Asymptotically Euclidean manifolds. The fact that the quantum measure is invariant under AET’s means that its definition does not require a space-time provided with metrics, i.e., it demonstrates the non-metric nature of such a measure. Formulating the AET theory in a coordinate-free context, generated through a mechanism similar to the one used for differential manifolds, points out this property more clearly. This section therefore defines and studies asymptotically Euclidean manifolds. ###### Definition 10 (a) Let $`M`$ be a set. We call Galilean reference frame for $`M`$ a bijective application $`\phi :GM`$. (b) Two Galilean reference frames $`\phi _1`$ and $`\phi _2`$ are said to be $`\omega `$E-equivalent if $`\phi _2^1\phi _1:GG`$ is an asymptotically Euclidean transformation. (c) An asymptotically Euclidean manifold is the pair $`(M,𝒜)`$, where $`𝒜`$ is a class of the $`\omega E`$-equivalence relation. The definitions of asymptotically regular semitrajectories and N-bigbangs, of asymptotic homeomorphism, asymptotically identical transformation and, finally, asymptotically Euclidean transformation can be transferred trivially from $`G`$ to the manifold $`M`$ through any one of its reference frames. For instance: a subset $`\widehat{\gamma }M`$ is an asymptotically regular semitrajectory on $`M`$ if $`\phi ^1(\widehat{\gamma })`$ is an asymptotically regular semitrajectory on $`G`$ for any reference frame $`\phi 𝒜`$. However, the notion of classical N-bigbang cannot be transferred on $`M`$, because it is not invariant under AET’s. In order to indicate the objects on $`M`$ we use the same symbols used for the corresponding objects on $`G`$, but writing them with hat. For instance, $`\widehat{B}`$ is the set of the asymptotically regular N-bigbangs on $`M`$. We will now construct the space of asymptotic velocities for the manifold $`(M,𝒜)`$. ###### Definition 11 (a) Two semitrajectories $`\widehat{\gamma }_1`$ and $`\widehat{\gamma }_2`$ on $`M`$ are said to be $`\omega `$-equivalent if, for any reference frame $`\phi `$, one has $`\omega [\phi ^1(\widehat{\gamma }_1)]=\omega [\phi ^1(\widehat{\gamma }_2)]`$. This definition does not depend on the chosen reference frame. (b) The space of asymptotic velocities on M, indicated by $`\widehat{V}`$, is the quotient space of the $`\omega `$-equivalence relation. The space $`\widehat{V}`$ can be considered the asymptotic correspondent of the tangent vector space of differential manifolds. Let us construct the class of reference frames for $`\widehat{V}`$. Let $`\widehat{\omega }:\widehat{\mathrm{\Gamma }}\widehat{V}`$ be the application that associates each semitrajectory wich its $`\omega `$-equivalence class. A reference frame $`\phi `$ of $`M`$ induces an application $`\varphi _\phi :V\widehat{V}`$ which is defined by setting $`\varphi _\phi (𝐯):=\widehat{\omega }[\phi (\gamma )]`$, where $`\gamma `$ is any semitrajectory such that $`\omega (\gamma )=𝐯`$. Let us set $`𝒜_V:=\{\varphi _\phi |\phi 𝒜\}`$. ###### Lemma 9 (a) The applications of $`𝒜_V`$ are bijective. (b) If $`\varphi _1,\varphi _2𝒜_V`$, then $`\varphi _2^1\varphi _1`$ is a Euclidean transformation on $`V`$. (c) If $`\varphi 𝒜_V`$ and if $`g`$ is a Euclidean transformation on $`V`$, then $`\varphi g𝒜_V`$ as well. The set $`𝒜_V`$ is therefore the class of reference frames for $`\widehat{V}`$. Class $`𝒜_V`$ induces on $`\widehat{V}`$ both a structure of affine space, and the definition of Euclidean transformation. The space $`\widehat{V}^N`$ can equally be defined as the Cartesian product of $`\widehat{V}`$ or as the $`\omega `$-equivalence class of the N-bigbangs of $`\widehat{B}`$. A reference frame for $`\widehat{V}^N`$ is derived trivially from a reference frame for $`\widehat{V}`$. Let $`\varphi _1`$ and $`\varphi _2`$ be two reference frames for $`\widehat{V}^N`$, and let $`\mathrm{\Delta }V^N`$ be invariant under Euclidean transformations; since $`\varphi _2^1\varphi _1`$ is a Euclidean transformation on $`V^N`$, one has that $`\varphi _1(\mathrm{\Delta })=\varphi _2(\mathrm{\Delta })`$. This means that an invariant subset of $`V^N`$ can be transferred without ambiguities onto $`\widehat{V}^N`$. In particular, let us define $`\widehat{\mathrm{\Delta }}_C:=\varphi (\mathrm{\Delta }_C)`$ and $`\widehat{\mathrm{\Delta }}_B:=\varphi (\mathrm{\Delta }_B)`$, where $`\mathrm{\Delta }_C`$ and $`\mathrm{\Delta }_B`$, defined in sections 2.2 and 2.2.1, are, respectively, the set of asymptotic velocities and the set of asymptotic boundary conditions for the classical N-bigbangs. On $`\widehat{V}^N`$ one can define the quantum measure $`\widehat{\mu }_Q`$, setting $`\widehat{\mu }_Q(\widehat{\mathrm{\Delta }}):=\mu _Q[\varphi ^1(\widehat{\mathrm{\Delta }})]`$; in this case also, the definition does not depend on the reference frame chosen. The definition of a classical reference frame is now given: ###### Definition 12 (a) Given an N-bigbang $`\widehat{\beta }`$ on M, a reference frame $`\phi `$ is said to be a classical reference frame for $`\widehat{\beta }`$ if $`\phi ^1(\widehat{\beta })`$ is a classical N-bigbang on $`G`$. (b) The classical image of an N-bigbang $`\widehat{\beta }`$ is the set $`\{\phi ^1(\widehat{\beta })B|\phi `$ is a classical reference frame for $`\widehat{\beta }\}`$. ###### Lemma 10 (a) An N-bigbang $`\widehat{\beta }`$ admits a classical reference frame if and only if $`\widehat{\omega }(\widehat{\beta })\widehat{\mathrm{\Delta }}_C`$. (b) If there exists on $`M`$ an asymptotically Euclidean transformation $`\widehat{f}`$ such that for two N-bigbangs $`\widehat{\beta }_1`$ and $`\widehat{\beta }_2`$ one has $`\widehat{\beta }_2=f(\widehat{\beta }_1)`$, then $`\widehat{\beta }_1`$ and $`\widehat{\beta }_2`$ have the same classical image. (c) The classical image of an N-bigbang is a class of the $`\omega E`$-equivalence relation on $`B_C`$ (defined in section 2.2). Proof. Only the proof of point (a) is given. (i) $`\widehat{\beta }`$ admits a classical reference frame $`\widehat{\omega }(\widehat{\beta })\widehat{\mathrm{\Delta }}_C`$: if $`\phi `$ is a classical reference frame for $`\widehat{\beta }`$, the relation $`\omega [\phi ^1(\widehat{\beta })]\mathrm{\Delta }_C`$ must hold; since $`\widehat{\omega }(\widehat{\beta })=\varphi _\phi [\omega (\phi ^1(\widehat{\beta }))]`$, the relation $`\widehat{\omega }(\widehat{\beta })\widehat{\mathrm{\Delta }}_C`$ also must hold. (ii) $`\widehat{\omega }(\widehat{\beta })\widehat{\mathrm{\Delta }}_C\widehat{\beta }`$ admits a classical reference frame: let $`\phi `$ be a generic reference frame; then $`\omega [\phi ^1(\widehat{\beta })]\mathrm{\Delta }_C`$. Let $`\beta _CB_C`$ be a classical N-bigbang such that $`\omega (\beta _C)=\omega [\phi ^1(\widehat{\beta })]`$; by virtue of Theorem 2, there exists on $`G`$ an asymptotically identical transformation $`f`$ such that $`f(\beta _C)=\phi ^1(\widehat{\beta })`$. The reference frame $`\phi _C:=\phi f`$ is therefore a classical reference frame for $`\widehat{\beta }`$. QED. One can derive from Lemma 10c that two classical N-bigbangs not connected by a Galilean transformation can belong to the classical image of the same N-bigbang $`\widehat{\beta }`$ on $`M`$. The following definition will be useful later: ###### Definition 13 Let $`\widehat{\beta }`$ be an N-bigbang on $`M`$. The generalized classical image of $`\widehat{\beta }`$ is the set of the classical N-bigbangs on $`G`$ defined by the set of asymptotic boundary conditions $`\{\omega [\phi ^1(\widehat{\beta })]V^N|\phi 𝒜\}`$. Similarly to what happens for the classical image, one can prove that the generalized classical image of an N-bigbang $`\widehat{\beta }`$ is not empty if and only if $`\widehat{\omega }(\widehat{\beta })\widehat{\mathrm{\Delta }}_B`$. The theory mentioned in section 2.2.1 should prove that if $`\widehat{\omega }(\widehat{\beta })\widehat{\mathrm{\Delta }}_B\widehat{\mathrm{\Delta }}_C`$, its classical image and its generalized classical image coincide. If $`\widehat{\omega }(\widehat{\beta })\widehat{\mathrm{\Delta }}_B\widehat{\mathrm{\Delta }}_C`$, then $`\widehat{\beta }`$ admits a generalized classical image but admits no classical reference frame. ## 3 Physical interpretation. This second part of the paper discusses the physical and conceptual aspects of the mathematical theory developed in the first part. ### 3.1 An ideal model of universe. This section proposes an ideal model of universe based on asymptotically Euclidean manifolds. Since this is a nonrelativistic structure, and since the quantum measure is based on bare Schrödinger’s equation (which is inadeguate to explain many atomic phenomena, see section 3.3.4), this can only be a schematic and ideal model, and one can draw only interpretative and conceptual conclusions from it. For the sake of brevity, hereafter it will be referenced as asymptotic model. The asymptotic model can be described by the following three statements: 1. The evolution of the particles of the universe is represented by an N-bigbang $`\widehat{\beta }`$ on an asymptotically Euclidean manifold. Two N-bigbangs connected by an asymptotically Euclidean transformation represent the same evolution. 2. (a) The N-bigbang $`\widehat{\beta }`$ appears to us as if it were seen from a classical reference frame. (b) All the classical reference frames for $`\widehat{\beta }`$ are equivalent from an observational standpoint. 3. The quantum measure $`\widehat{\mu }_Q`$ on $`\widehat{V}^N`$ rules the choice of the asymptotic velocity that defines the N-bigbang, in the same way in which, in a probability space, the probability measure rules the choice of an elementary outcome. Let us examine these three statements. Statements 1 and 2a imply renouncing what we might consider the paradigm of classical mechanics, i.e., the existence of a metrics defined a priori on which a law of motion is based which determines the motion of particles starting from the initial conditions <sup>3</sup><sup>3</sup>3More in general, one can certainly say that this paradigm is, to a large extent, at the core of our very notion of physical reality.. Statements 1 and 2a instead state that there is neither a metrics defined a priori nor a law of motion, and that it is the reference frame, i.e., our perception of space distances and time intervals, that structures itself so that particle motion appears to be ruled by a law. It should be noted that Hamilton’s action principle already favors an interpretation according to which particle motion is not determined starting from the initial conditions alone, but rather from the initial and final conditions together, which in the asymptotic model are respectively constituted by the initial constrain of the bigbang and by the asymptotic velocity. The asymptotic model takes a further step forward by stating that Hamilton’s action principle does not determine the trajectory of the particles but rather the metric. To paraphrase Wheeler et al., who in wrote: “Time is defined so that motion looks simple”, one could state that: “Space-time is defined so that motion looks simple”. From the observational standpoint, the classical paradigm and this new interpretation are equivalent, since in both cases the particles travel (or appear to travel) along classical trajectories. This new interpretation is proposed here both because it is suggested by the mathematical formalism and because it will make statement 3 more acceptable together with the extremely counterintuitive consequences that it implies, as we will see. Statement 2a requires the N-bigbang $`\widehat{\beta }`$ to admit a classical reference frame, i.e., it requires its asymptotic velocity to belong to $`\widehat{\mathrm{\Delta }}_C`$. Taking into account statement 3, this condition would be certainly verified if the equation $$\mu _Q(V^N\mathrm{\Delta }_C)=0$$ (24) were to hold. The validity of (24) obviously depends upon the Lagrangian. Even though the asymptotic model does not provide a Lagrangian, by taking into account the example of section 2.2.1, condition (24) seems exceedingly restrictive. In the mentioned example, the set of asymptotic boundary conditions $`\mathrm{\Delta }_B`$ is wider than the set $`\mathrm{\Delta }_C`$, so that the condition $$\mu _Q(V^N\mathrm{\Delta }_B)=0$$ (25) would seem more acceptable \[notice that equation (25) includes the condition $`\mathrm{\Delta }_B=V^N`$ as a special case\]. We will therefore conjecture that for a “reasonable” Lagrangian, condition (25) should be satisfied. If so, the N-bigbang $`\widehat{\beta }`$ admits a generalized classical image, and it is possible to maintain the logical consistency of the asymptotic model by extending statement 2a in the following way: the N-bigbang $`\widehat{\beta }`$ appears to us as a classical N-bigbang which belongs to its generalized classical image. Let us move on to statement 2b. Owing to Lemma 10c, this statement can be expressed in an equivalent formulation by saying that two classical N-bigbangs that belong to the same $`\omega E`$-equivalence class are indistinguishable to our perception. This statement can be broken into the following two statements: (i) two N-bigbangs connected by a Galilean transformation are indistinguishable, and (ii) two N-bigbangs not connected by a Galilean transformation but having the same asymptotic velocity are indistinguishable. Let us examine these two statements separately. Statement (i) is obvious, since what we perceive are the relative positions and velocities of particles (which are invariant under Galilean transformations), while we are certainly not able to perceive neither the position and velocity of the center of mass of the universe, nor its space orientation (which are not invariant). Statement (ii) instead is novel and requires deeper analysis. Consider two different classical N-bigbangs having a common origin, for instance the point $`(0,0)`$, and having equal asymptotic velocities $`v`$. The fact that they are indistinguishable could be accounted for physically by stating that they differ only microscopically and therefore that they are macroscopically indistinguishable. To accept this statement, one should consider that the condition of equal asymptotic velocity can be exceedingly fragile, since it is not met if even a single particle of the universe has a different asymptotic velocity within the two N-bigbangs. For instance, if a photon is emitted in two different space directions (or if it is emitted in one case and not in the other), then the two N-bigbangs will have different asymptotic velocities. It is therefore reasonable to admit that a macroscopic difference between the two N-bigbangs, by irreversibly propagating into the surrounding environment, will lead them to have different asymptotic velocities. Vice versa, two N-bigbangs with the same asymptotic velocity will never be able to differ macroscopically from one another; therefore they will not be distinguishable to our perception. Two N-bigbangs of this kind could correspond, for example, to the two trajectories followed, in the two-slit experiment, by a particle passing through one slit or the other (see section 3.3.2). In other words, the macroscopic events of the evolution are determined by the asymptotic boundary conditions of the universe. For instance, two different asymptotic boundary conditions correspond to the fact that I blinked or not while I was writing this sentence. This is not particularly surprising. One should bear in mind that an even stricter rule holds for the initial conditions (position and velocity): even microscopically different events correspond to different initial conditions of the universe. Let us return to the general discussion of the asymptotic model. We have seen that according to statement 2a, particles travel (or appear to travel) along classical trajectories. Therefore, in the asymptotic model the wave function is not used to describe particle motion. So how does it explain the quantum phenomena observed in nature? This explanation resides in statement 3 and in the phenomenon of nonchaotic distribution of initial conditions that derives from it. The sections that follow are devoted to the study of these topics. Let us make a final remark on the big-bang structure of the asymptotic model. Although this type of model for the universe is almost unanimously accepted, its use outside a general-relativistic context might appear simple-minded or strained. However I would like to point out that this structure has played a fundamental role in defining the quantum measure $`\mu _Q`$ and in proving its invariance under AET’s. It is nonetheless reasonable to expect that a much more realistic and elegant big-bang model may be obtained by formulating the theory of asymptotic transformations starting from Einstenian space-time rather from Galilean space-time. ### 3.2 The universe as a probability space. Statement 3 of the asymptotic model likens the universe to a probability space. A probability space is a tern $`(S,,\nu )`$, where $`S`$ is the set of elementary outcomes (also called sample space), $``$ is the $`\sigma `$-algebra of the events and $`\nu `$ is the probability measure. Probability spaces are used in physics to represent a sequence of repetitions of the same experiment, using the following correspondences: the space $`S`$ corresponds to the possible elementary outcomes of the experiment, the $`\sigma `$-algebra $``$ defines the possible measurable events, and finally the measure of an event, i.e., its probability, corresponds to its relative frequency of occurrence in the sequence. In this paper, a probability space will be used to represent the universe, employing the following correspondences: * The space $`S`$ is the set $`B_C`$ of the classical N-bigbangs. * The $`\sigma `$-algebra describes our ability to distinguish the N-bigbangs; therefore its structure must take into account statement 2b of the asymptotic model. This aspect will be considered in detail in section 3.2.2. * The probability measure derives from the quantum measure $`\widehat{\mu }_Q`$. However, the correspondence between probability and relative frequency has to be redefined, since universe evolves only once. This redefinition can be achieved in a simple way, by introducing the concept of sequence of equivalent events and by making use of the law of large numbers. This new correspondence will be illustrated by a simple example in the next section. The example will furthermore show that not only is it possible to consider the universe as a probability space, but it is necessary to do so in order to account for the statistical regularities of evolution. The need for a measure in addition to the law of motion to account for the statistical regularities of evolution was noted Popper and Landé , although they used this conclusion as a criticism of determinism. #### 3.2.1 A simple example. Consider a universe represented by a dynamical system whose phase-space is the interval $`[0,1)`$ of the real axis, and whose law of motion is Bernoulli’s application $`f(x):=2x`$ (mod. 1). In this universe, a trajectory is a sequence $`\{a_n\}`$ of numbers in the interval $`[0,1)`$ that can be obtained by repeatedly applying the law of motion to an initial condition $`a_0`$. There is, therefore, a one-to-one correspondence between the trajectories and the initial conditions, i.e., the elements of the interval $`[0,1)`$. The time of this universe is represented by discrete ticks, which correspond to the natural numbers. It is very easy to prove that the law of motion alone cannot account for all the observable phenomena of the evolution of this universe. Suppose we measure, over long tick sequences, the relative frequency with which the state of the system is less than $`\frac{1}{2}`$. Naturally we expect the result to be very close to $`\frac{1}{2}`$. However, this prediction cannot be deduced from the law of motion, due to the simple fact that there are trajectories which satisfy the law of motion and that have a relative frequency different from $`\frac{1}{2}`$: consider, for instance, the trajectory generated by the initial condition $`\frac{1}{7}`$, which determines a relative frequency of $`\frac{2}{3}`$. In order to obtain a given relative frequency, and particularly the frequency $`\frac{1}{2}`$, we need to define a measure on the trajectories which says that there are “many more” trajectories giving a relative frequency of $`\frac{1}{2}`$ than trajectories yielding other relative frequencies, and that therefore it is “almost certain” that the trajectory our universe is traveling along is one of those that yields a relative frequency of $`\frac{1}{2}`$. Let us construct, therefore, a probability space for this universe. The space $`S`$ is the space of the trajectories of the system, the $`\sigma `$-algebra and the measure $`\nu `$ on $`S`$ derive from Borel’s $`\sigma `$-algebra and from the Lebesgue measure of the interval $`[0,1)`$ through the one-to-one correspondence that exists between the two sets (see Appendix). Events are subsets of trajectories and can be described by statements such as: “at the i-th tick, the state of the system is less than $`\frac{1}{2}`$,” or: “between the i-th and j-th ticks, the state of the system is greater than $`\frac{1}{2}`$”. In order to establish a correspondence between probability and frequency, let us introduce the notion of sequence of equivalent events: we say that a sequence of events $`E_1,\mathrm{},E_n`$ is a sequence of equivalent events if all the events have the same probability and are independent, i.e., $`\nu (E_1\mathrm{}E_n)=\nu (E_1)\mathrm{}\nu (E_n)`$. For instance, if $`E_i`$ is the event “at the i-th tick, the state of the system is less than $`\frac{1}{2}`$”, any sequence composed of events $`E_i`$ is a sequence of equivalent events with probability $`\frac{1}{2}`$. Given a sequence of equivalent events, we associate with every trajectory $`\{a_i\}`$ the number $`\eta (\{a_i\})`$ that indicates how many events of the sequence are verified by the trajectory (i.e., how many events it belongs to). By the law of large numbers, for any $`ϵ>0`$, $$\nu \{\{a_i\}S|ϵ|\eta (\{a_i\})/nP|\}0\mathrm{for}n\mathrm{},$$ (26) where $`n`$ is the number of events in the sequence and $`P`$ is their probability. Expression (26) says that if $`n`$ is large, then there are very “few” trajectories which determine, for the relative frequency of events, a value which is significantly different from their probability. Thanks to espression (26), one can therefore set the following correspondence between probability and frequency: ###### Proposition 1 The relative frequency with which the events of a long sequence of equivalent events occur corresponds to their probability <sup>4</sup><sup>4</sup>4 Actually, the passage from espression (26) to Proposition 1 is based on the implicit assumption that an event having very small probability does not occur. This assumption is accepted without discussing its validity or its conceptual implications.. From this statement and from the definition of the events $`E_i`$ one easily obtains the value $`\frac{1}{2}`$ for the relative frequency described at the beginning of the section. #### 3.2.2 Probability space and measurement theory for the asymptotic model. Let us now build, by analogy with the Bernoulli system, a probability space for the asymptotic model. The sample space is the set $`B_C`$ of classical N-bigbangs. The $`\sigma `$-algebra and the measure are transferred to $`B_C`$ from $`V^N`$ through the application $`\omega _C:=\omega |_{B_C}:B_CV^N`$. The following are defined on $`V^N`$: (i) Borel’s $`\sigma `$-algebra $``$ and the quantum measure $`\mu _Q`$, and (ii) the quotient $`\sigma `$-algebra $`\stackrel{~}{}`$ and the quotient measure $`\stackrel{~}{\mu }_Q`$ (see section 2.3.2 and the Appendix). The $`\sigma `$-algebra on $`B_C`$ must take into account statement 2b of the asymptotic model; therefore two disjoined events, i.e., two distinguishable events, which contain two N-bigbangs belonging to the same class of $`\omega E`$-equivalence will not be allowed. This requirement is certainly satisfied if the $`\sigma `$-algebra on $`B_C`$ is derived from the $`\sigma `$-algebra $`\stackrel{~}{}`$, because in this way the atoms of the derived $`\sigma `$-algebra are the classes of $`\omega E`$-equivalence themselves. The “probability” space that represent the universe will therefore be the tern $$(B_C,\stackrel{~}{},\stackrel{~}{\nu }_Q),$$ (27) where $`\stackrel{~}{}:=\omega _C^1(\stackrel{~}{})`$, and $`\stackrel{~}{\nu }_Q:=\omega _C^1(\stackrel{~}{\mu }_Q)`$ <sup>5</sup><sup>5</sup>5 Actually, this method for transferring the measure $`\mu _Q`$ from $`V^N`$ onto $`B_C`$ or onto $`V_I^N`$ does not give correct results at the discontinuity points of the function $`\omega _V`$. See Appendix A for more details. I preferred to use this method anyway in order to simplify the exposition, but the reader should bear in mind that this method should be replaced everywhere with the method described in Appendix A.. Note that if the $`\sigma `$-algebra on $`B_C`$ were derived from $``$ rather than from $`\stackrel{~}{}`$, one would eliminate the fact that two classical N-bigbangs connected by a Galilean transformation are indistinguishable but not the fact that two N-bigbangs having the same asymptotic velocity are indistinguishable; the latter feature is structural to the asymptotic model and cannot be eliminated. Actually, the tern (27) does not represent a true probability space, since the measure $`\stackrel{~}{\nu }_Q`$ is not bounded, and therefore it cannot be normalized (one should bear in mind that there is no limit to the energy of the N-bigbangs of $`B_C`$). This fact will not prevent the construction of a consistent measurement theory: as we shall see, the probability of the result of a measurement will be defined as a ratio of measures. In order to build a measurement theory for the asymptotic model (to which we will refer as asymptotic theory of measurement), consider two kinds of events: the experiment event and the result event. The experiment-event includes all those trajectories of $`B_C`$ which are macroscopically compatible with the preparation of a precise experiment occurring in a given space-time region. The result event is a subset of the experiment event, containing all the trajectories compatible with a certain result of the experiment. It is natural to define the probability $`P(R|E)`$ of a result $`R`$ of the experiment $`E`$ as $$P(R|E):=\frac{\stackrel{~}{\nu }_Q(R)}{\stackrel{~}{\nu }_Q(E)}.$$ (28) As in the previous example, it is not difficult to establish a correlation between probability and frequency: given an experiment $`E`$, we say that a sequence of results $`R_1,\mathrm{},R_n`$ of the experiment is equivalent if $`\stackrel{~}{\nu }_Q(R_1)=\mathrm{}=\stackrel{~}{\nu }_Q(R_n)`$, and furthermore $`\stackrel{~}{\nu }_Q(R_1\mathrm{}R_n)=\stackrel{~}{\nu }_Q(R_1)\mathrm{}\stackrel{~}{\nu }_Q(R_n)`$. For instance, if the experiment consists of the emission of a large number of particles toward a screen, equivalent results are those in which different particles hit the same region of the screen. In this case also, the law of large numbers allows one to deduce that if $`n`$ is large, the probability that the relative frequency of the results $`R_i`$ differs significantly from their probability is almost null. One should bear in mind that in the asymptotic model equation (28) determines the probability of the results of all experiments of a statistical nature; for instance, it determines both the probability that tossing a coin will produce a given result and the probability that in the two-slit experiment a particle will hit the screen at a given point. We will see, in the latter case, how the phenomenon of the nonchaotic distribution of initial conditions allows to account for the quantum effect of interference fringes even within a classical dynamical context. #### 3.2.3 Comparison between asymptotic and quantum measurement theory. The aim of this section is to compare asymptotic theory with the quantum theory of measurement, showing that if one makes a suitable assumption regarding the quantum measurement process they are compatible. This section is mathematically less rigorous out of necessity and for the sake of simplicity. As a first simplification, we will consider as a probability space for the asymptotic model the tern $$(C,,\nu _Q)$$ (29) instead of the tern (27), where $`:=\omega _C^1()`$ and $`\nu _Q:=\omega _C^1(\mu _Q)`$ . Hence the requirement that two N-bigbang connected by a Galilean transformation be indistinguishable has been disregarded. Therefore equation (28) becomes: $$P(R|E)=\frac{𝐱_N|E^+[\omega _C(R)]|𝐱_N}{𝐱_N|E^+[\omega _C(E)]|𝐱_N}.$$ (30) As regards quantum theory, we will use the Everett-DeWitt Many-Worlds theory for the comparison, as it is the most convenient for this purpose. In this theory, the state of the entire universe is represented by a wave function $`\psi (x,t)`$, and by obvious analogy with relation (13) we assume that $`\psi (x,0)=|𝐱_N`$ . During evolution, every time a measurement (or an equivalent process) takes place, the state of the universe subdivides into branches which differ macroscopically from one another. Consider a measurement made at the time $`t`$ which involves $`n`$ possible distinct results; let $`\psi `$ be the state of the universe at the time of the measurement, and let $`\psi _1,\mathrm{},\psi _n`$ be the states into which it decomposes. Obviously the equation $`_i\psi _i=\psi `$ holds. The quantum probability $`P_i`$ of the i-th result is given by the classical formula $$P_i=\frac{|\psi |\psi _i|^2}{\psi |\psi \psi _i|\psi _i}.$$ (31) One can achieve a tight link between the two measurement theories by making a conjecture for which we prepare a few definitions. Let the space support of a state $`\psi `$ be the support of the measure $`\psi |E^Q()|\psi `$ , and let the asymptotic support be the support of the measure $`\psi |E^+()|\psi `$ . We say that two states $`\psi _1`$ and $`\psi _2`$ are spatially (asymptotically) disjoined if their space (asymptotic) supports are disjoined. Finally, we say that the two states are definitively disjoined if $`e^{iHt}\psi _1`$ and $`e^{iHt}\psi _2`$ are spatially disjoined for every $`t0`$. Let us now make the following assumption: ###### Proposition 2 The states into which the wave function of the universe subdivides due to a measurement are definitively disjoined. The fact that states that correspond to different results of a measurement are disjoined is widely accepted in the literature . This conclusion trivially derives from the fact that the two states must represent measurement instruments whose pointers occupy different positions (note that here, too, in order for two states of the universe to be spatially disjoined it is sufficient for them to describe a single particle in two different positions). The fact they are also definitively disjoined can be justified by a line of argument which is similar to the one used in section 3.1 to justify the fact that two macroscopically distinct trajectories have distinct asymptotic velocities. An assumption perfectly analogous to Proposition 2 was made by Bohm . Unfortunately, after presenting the physical reasons why Proposition 2 must hold, we must acknowledge that from a mathematical standpoint, at least in the context developed in this paper, the proposition is impossible. In fact, it is well-known that a wave function that evolves according to Schr’́odinger’s equation and has a compact space support at a given instant, is spread to the entire space at any subsequent instant. Proposition 2 can therefore be at most “almost true”. We will replace it with the following weaker but more precise assumption: ###### Proposition 3 The states into which the wave function of the universe subdivides due to a measurement are asymptotically disjoined. From Proposition 3 and from the definition of asymptotic support one can derive that $`\psi _i=E^+(\mathrm{\Delta }_i)\psi `$ , where $`\mathrm{\Delta }_i`$ is the asymptotic support of $`\psi _i`$. In turn, the state $`\psi `$ derives from an alternate sequence of free evolutions and subdivisions $$\psi =e^{iH(tt_n)}E^+(\mathrm{\Delta }^{(n)})e^{iH(t_nt_{n1})}\mathrm{}E^+(\mathrm{\Delta }^{(1)})e^{iHt_1}|𝐱_N,$$ where $`\mathrm{\Delta }^{(k)}`$ is the asymptotic support that corresponds to the branching which took place at the k-th instant. Since the projectors $`E^+(\mathrm{\Delta }^{(k)})`$ commute with the Hamiltonian, one obtains $$\psi =E^+(\mathrm{\Delta }^{(n)}\mathrm{}\mathrm{\Delta }^{(1)})e^{iHt}|𝐱_N=E^+(\mathrm{\Delta }_\psi )e^{iHt}|𝐱_N,$$ (32) where $`\mathrm{\Delta }_\psi :=\mathrm{\Delta }^{(n)}`$ is the asymptotic support of $`\psi `$ . By replacing equation (32) in equation (31) one obtains $$P_i=\frac{𝐱_N|E^+(\mathrm{\Delta }_i)|𝐱_N}{𝐱_N|E^+(\mathrm{\Delta }_\psi )|𝐱_N}.$$ (33) This exactly corresponds to equation (30), provided that one identifies the asymptotic supports of the branches into which the wave function subdivides with the sets of asymptotic velocities of the experiment events and result events of the asymptotic theory of measurement. ### 3.3 Nonchaotic distribution of initial conditions. An important difference between the measure on the trajectories of the Bernoulli system and the measure on the N-bigbangs of the asymptotic model is that the former derives from a measure defined on the initial conditions, while the latter derives from a measure defined on the asymptotic velocities, i.e., on the final conditions. This feature admits, in the asymptotic model, the phenomenon of nonchaotic distribution of initial conditions (NCDIC), i.e., a distribution of the initial conditions that depends on the future evolution of the trajectories. The physical understanding of this extremely counterintuitive phenomenon is fundamental, in order to understand how typically quantum-like phenomena such as the two-slit experiment can be explained in a context where particle motion is described by classical trajectories. We will also see how this phenomenon is tightly linked to the EPR paradox. The phenomenon of NCDIC is remarkably similar to the phenomenon of preinteractive correlations, on which Price has written extensively. One can illustrate this phenomenon in general in the following way. The law of motion allows to determine univocally the asymptotic velocity of a classical N-bigbang starting from its initial velocity, i.e., from the velocities of the particles at the origin of the N-bigbang. Let $`V_I^N=R^{3N}`$ be the space of initial velocities and let $`\omega _V:V_I^NV^N`$ be the application that describes this correspondence. We assume without demonstration that $`\omega _V`$ is measurable. On $`V_I^N`$ we define the measure $`\pi _C`$ (where $`C`$ stands for “classical”) by setting $$\pi _C(\mathrm{\Delta }):=\frac{\mu _L(m\mathrm{\Delta })}{(2\pi )^{3N}},$$ (34) where $`\mu _L`$ is the Lebesgue measure. The measure $`\pi _C`$ describes a uniform (or chaotic) distribution of the initial momenta \[the factor $`(2\pi )^{3N}`$ is introduced for consistency with the quantum measure in the case of lack of potential, see equation (2.3.2)\]. On $`V_I^N`$ one can also define the quantum measure $`\pi _Q`$ by deriving it through the application $`\omega _V`$ from the quantum measure $`\mu _Q`$ on the asymptotic velocities: $$\pi _Q:=\omega _V^1(\mu _Q).$$ (35) The measure $`\pi _Q`$ is defined on the $`\sigma `$-algebra $`_\omega :=\omega _V^1()`$, which in general is different from $``$; since $`\omega _V`$ is Borel-measurable, we have $`_\omega `$. One should also consider note 4. The measure $`\pi _Q`$ describes the distribution that the initial velocities of the particles must have in order for the distribution of their asymptotic velocities to be $`\mu _Q`$. In general, one will have $$\pi _Q(\mathrm{\Delta })\pi _C(\mathrm{\Delta }),\mathrm{\Delta }_\omega .$$ (36) Formula (36) expresses the fact that the asymptotic model has a nonchaotic distribution of the initial momenta of the particles. The structure of $`\pi _Q`$ will be extremely complex and intricate, because a minute difference in the initial velocities entails a large difference in the final asymptotic velocities. The measure $`\mu _C`$, defined by $$\mu _C:=\omega _V(\pi _C),$$ (37) instead describes the distribution of asymptotic velocities that one would obtain if the distribution of initial momenta were uniform. The Appendix shows an equation which compares, albeit in an approximate way, the measures $`\mu _C`$ and $`\mu _Q`$. We will now show how the possibility of an NCDIC radically modifies the physical interpretation of some major quantum experiments. #### 3.3.1 The scattering process. Consider a flux of particles of the same type and with the same velocity oriented toward the positive direction of the $`z`$-axis and headed toward a target represented by a central potential $`V(r)`$. Let $`\mathrm{\Theta }`$ be the scattering angle and let $`\varphi `$ be the angle that the plane of the trajectory forms with the $`x`$-axis. The scattering cross-section $`\sigma (\mathrm{\Theta },\varphi )`$ is defined as $$\sigma (\mathrm{\Theta },\varphi )\mathrm{d}\mathrm{\Omega }=\frac{\mathrm{number}\mathrm{of}\mathrm{particles}\mathrm{scattered}\mathrm{for}\mathrm{unit}\mathrm{time}\mathrm{in}\mathrm{the}\mathrm{solid}\mathrm{angle}\mathrm{d}\mathrm{\Omega }}{\mathrm{incoming}\mathrm{flux}\mathrm{intensity}},$$ (38) where $`\mathrm{d}\mathrm{\Omega }`$ is the solid angle $`\mathrm{sin}\mathrm{\Theta }\mathrm{d}\mathrm{\Theta }\mathrm{d}\varphi `$, and the incoming flux intensity is the number of incoming particles per unit time divided by the total impact area, which we assume to be finite yet as large as one chooses. The calculation of the classical cross-section is based on the statistical hypothesis that the flux of incoming particles is uniform in the incidence plane. Let us now develop a derivation of the cross-section which highlights this hypothesis, keeping it separate from the dynamical part of the calculation. The trajectory of a particle is univocally determined by the impact parameter $`s`$ and by the angle $`\varphi `$. For the time being, we are not making any hypotheses on the distribution of the incoming flux and we describe it generically by means of a density $`\rho _I(s,\varphi )`$, so that if $`A`$ is a region in the incident plane, the quantity $`_A\rho _I(s,\varphi )sdsd\varphi `$ is the number of particles per unit time that crosses the region $`A`$. Likewise, let $`\rho _S(\mathrm{\Theta },\varphi )`$ be the density of the scattered particles, so that $`_\mathrm{\Omega }\rho _S(\mathrm{\Theta },\varphi )\mathrm{sin}\mathrm{\Theta }\mathrm{d}\mathrm{\Theta }\mathrm{d}\varphi `$ is the number of particles scattered in the unit time within the solid angle $`\mathrm{\Omega }`$. We will therefore have $`\sigma (\mathrm{\Theta },\varphi )=\rho _S(\mathrm{\Theta },\varphi )/I`$, where $`I`$ is the incoming intensity. In order to make the calculation more realistic, let us suppose that the flux of incoming particles originates from a point-like source, located on the $`z`$-axis at the point $`z_0`$, which is at a great distance from the target. All the particles are emitted with the same energy and so that the direction of the velocities is distributed according to a given distribution $`\rho _E(\theta ,\varphi )`$, where $`\theta `$ is the angle formed with the $`z`$-axis. Obviously, the equation $`s=\theta |z_0|`$ holds, so that $`\rho _I(s,\varphi )=\rho _E(s/|z_0|,\varphi )`$. Therefore the distribution $`\rho _I(s,\varphi )`$ derives from the minute fluctuations of the emission angle $`\theta `$, which are completely beyond the control of the experimenter. The dynamics allows to express $`\mathrm{\Theta }`$ as a function of $`s`$; for the sake of simplicity, let us assume that this correspondence is invertible, as certainly happens in the case of a repulsive potential. Clearly, $$\rho _S(\mathrm{\Theta },\varphi )\mathrm{sin}\mathrm{\Theta }\mathrm{d}\mathrm{\Theta }\mathrm{d}\varphi =\rho _I(s,\varphi )s\mathrm{d}s\mathrm{d}\varphi =\rho _E(s/|z_0|,\varphi )s\mathrm{d}s\mathrm{d}\varphi $$ holds. Since $`\mathrm{d}s=|s/\mathrm{\Theta }|\mathrm{d}\mathrm{\Theta }`$, we have: $$\rho _S(\mathrm{\Theta },\varphi )=\rho _I(s,\varphi )\frac{s}{\mathrm{sin}\mathrm{\Theta }}\left|\frac{s}{\mathrm{\Theta }}\right|.$$ (39) Equation (39) is purely dynamical, i.e., it does not have any embedded statistical hypothesis. The classical cross-section is obtained by using the statistical hypothesis that the flux is uniform, i.e., $$\rho _I(s,\varphi )=\rho _E(s/|z_0|,\varphi )=I.$$ (40) Hypothesis (40) is normally considered so obvious that it is very difficult to find in the literature a line of reasoning which supports it or justifies it. It exactly corresponds to the hypothesis that the distribution of the velocities of the particles emitted by the source is chaotic and therefore does not depend on the subsequent interactions of the particles. It is well-known that there are potentials for which the classical cross-section is not correct. It is therefore evident that either equation (39) or hypothesis (40) are not correct. The current interpretation states that equation (39), i.e., the dynamics, is incorrect, and that it must be replaced with a quantum dynamics, while hypothesis (40) remains substantially true. The interpretation that derives from the asymptotic model instead states the opposite, i.e., that equation (39) is true and hypothesis (40) is false. In the asymptotic model, it is not the distribution $`\rho _S`$ that derives from the distribution $`\rho _E`$ through equation (39), but vice versa it is the distribution $`\rho _E`$ that, through the same equation, derives from $`\rho _S`$, which in turn derives from the measure on the asymptotic velocities of the particles of the universe \[note the analogy of this procedure with the one used in the previous section to derive the measure $`\pi _Q`$ from the measure $`\mu _Q`$; in this example, the application $`\mathrm{\Theta }(s)`$ plays the role of the application $`\omega _V`$\]. This reversal of perspective renders the distribution of the velocities with which particles are released by the source nonchaotic and allows it to depend on the form of the potential. #### 3.3.2 The two-slit experiment. Particle diffraction in the two-slit experiment is probably, along with the tunnel effect, the phenomenon that best convinced physicists that they needed to abandon classical trajectories to describe particle motion. See for instance Heisenberg . However, we see now that NCDIC allows an interpretation of this experiment which does not entail renouncing classical trajectories. Consider the experimental apparatus of Figure 1, which is the apparatus by which the two-slit experiment with electrons has been actually performed. Fig. 1 Here $`S`$ is an electron source, $`F`$ is a tiny conducting wire which crosses the plane of the figure at right angles and is set to a positive electric potential with respect to the two electrodes $`E_1`$ and $`E_2`$; $`H`$ is a screen constituted by a photographic plate. Due to the electrostatic field generated by the wire, the electrons emitted by the source are deflected and produce interference fringes on the screen. If the electrostatic field is turned off, the interference fringes disappear. This experiment is completely equivalent to the scattering experiment: there is a distribution $`\rho _E`$ of the directions of the velocities with which electrons are emitted by the source, and there is a distribution $`\rho _S`$ of the locations where they hit the plate. If one maintains the assumption that between the source and the screen the particles travel along classical trajectories, the two distributions $`\rho _E`$ and $`\rho _S`$ are mutually linked by a dynamical equation which is analogous to equation (39). The paradox arises from the implicit assumption that the distribution that “controls” the experiment is $`\rho _E`$, that it is chaotic, and that therefore it cannot depend on the fact that an electrostatic field is acting or not. All paradoxes disappear, however, if one admits that the “controlling” distribution is $`\rho _S`$, which is derived from the quantum measure of the asymptotic boundary conditions of the universe, and that the distribution $`\rho _E`$ derives from it. A typical question regarding the two-slit experiment is the following: through which of the two slits (in this case, on which side of the wire) does the particle pass? Let us see how the asymptotic model answers this question. The figure shows the two classical trajectories which make the particles hit the plate at a same point $`P`$. Those two trajectories, intended as overall trajectories of the universe, do not differ macroscopically from one another, since the electrons darken the same silver grain on the plate. One can therefore reasonably assume that they have the same asymptotic velocity. Therefore the two trajectories correspond to two different classical reference frames of the same N-bigbang on $`M`$, which is the actual real physical object. In this sense one can state that the particle travels through both slits. #### 3.3.3 The EPR paradox. The demonstration of Bell’s inequality also implicitly uses the hypothesis that the distribution of initial conditions is chaotic. Without this hypothesis, Bell’s inequality, and therefore the EPR paradox, can no longer be demonstrated. Let us now show where, in Bell’s classical demonstration , this hypothesis is used. A source emits pairs of particles in opposite directions towards two measurement instruments, each able to measure two distinct observables. In total, therefore, one can make four different measurements on each pair of particles. On the basis of the well-known criteria of local realism, one can deduces that every pair can be assigned, when it is emitted by the source, a hidden variable $`\lambda `$ which univocally determines the result of each one of the four measurements. The demonstration considers the execution of four distinct experiments: in each one, one of the four measurements is made on N pairs. Let $`\rho _i(\lambda ),i=1,\mathrm{},4`$ be the distributions of the variable $`\lambda `$ in the four experiments. Bell derives his inequality by assuming that the distributions $`\rho _i(\lambda )`$ are the same in the four experiments. This is exactly the chaotic hypothesis. If the chaotic hypothesis is not used, one can admit that the distributions $`\rho _i(\lambda )`$ depend on the measurement made, in the same way in which in scattering the distribution of the incoming particles can depend on the form of the potential. It is straightforward to prove that without the condition that the distributions $`\rho _i(\lambda )`$ are equal to one another the inequality can no longer be demonstrated. Despite this conclusion, the asymptotic model cannot describe the EPR experiment in terms of classical trajectories, because it cannot deal with particles with spin. This result could possibly be attained by extending the model. #### 3.3.4 Atomic levels and the tunnel effect. It is easy to realize that from a formal standpoint the bare Schrödinger equation cannot explain atomic levels. In fact it does not require an orbital electron to be in an eigenstate of the Hamiltonian, nor does it allow the electron to decay from one energy level to another. In order to achieve a more accurate representation of these phenomena one needs to add to the Hamiltonian a term which represents interaction with the electromagnetic field. The asymptotic model, by construction, inevitably shares this limitation with Schrödinger’s equation. An asymptotic model more suitable to represent those phenomena could possibly be obtained by extending this model to quantum field theory. Let us now consider the tunnel effect. Consider for instance an $`\alpha `$-decay, where an atom of radium emits an $`\alpha `$ particle and becomes an atom of radon. The ideal representation of this phenomenon describes the $`\alpha `$ particles as being confined in a potential well, as in Figure, Fig. 2 which is determined by the short-range nuclear attractive forces and by the weaker long-range Coulomb repulsive force. The kinetic energy of the emitted $`\alpha `$ particles is lower than the peak potential $`V_p`$; this is classically considered impossible. Indeed, even according to the asymptotic model, in a universe consisting of a single $`\alpha `$ particle subjected to the potential of Figure 2 the tunnel effect would not be observable. In fact, by using a line of argument similar to the one used in section 2.2.1, one can demonstrate that by setting as asymptotic boundary condition a velocity less than $`\sqrt{2V_p/m_\alpha }`$ one obtains for the $`\alpha `$ particle the limit trajectory which at the time $`t=+\mathrm{}`$ reaches the peak of the potential. However, in a real experiment, the situation is much more complex: the $`\alpha `$ particle interacts with the other particles in the atomic nucleus; furthermore, around the nucleus there are orbital electrons and finally the atom interacts with other atoms within the fissionable material. It is therefore possible for a suitable concurrence of this complex set of interactions to produce a classical trajectory which is compatible with the observed behavior. For instance, an orbital electron suitably hit by other orbital electrons or by a nearby atom could come close enough to the nucleus to lower the peak potential and allow the escape of the $`\alpha `$ particle having the required energy. This “conspiracy” of interactions appears absurd and unacceptable only if one does not take into account that here, too, there is the same reversal of perspective mentioned earlier during the analysis of the scattering experiment and of the two-slit experiment: it is not the distribution of the initial conditions that rules the experiment, but rather the distribution of the asymptotic conditions; this distribution entails high probabilities for asymptotic conditions that correspond to the tunnel effect and the trajectories conform to those conditions, no matter how particular and “conspiratorial” the interactions taking place during the evolution might appear. One has to take into account, however, that this representation of the tunnel effect by the asymptotic model is unavoidably affected by its inadequacy in representing atomic phenomena, as explained at the beginning of the section. The compatibility of the tunnel effect with classical trajectories was also stated in . ## 4 Conclusion. In order to understand whether the approach proposed in this paper is valid and fruitful, one must succeed in extending it to curved space-time and to quantum field theory. In my opinion the connection it highlights between a class of space-time transformations and a very general quantization mechanism such as Feynman path integrals makes it worthwhile to proceed with research in this direction. ## Appendix Transfer of measures. Let $`A`$ and $`B`$ be two sets and let $`f:AB`$ be an application. We describe two ways of transferring $`\sigma `$-algebras and measures between $`A`$ and $`B`$: * Let a $`\sigma `$-algebra $`_B`$ and a measure $`\mu _B`$ be defined on $`B`$. They can be transferred on $`A`$ by defining $`_A:=f^1(_B):=\{f^1(\mathrm{\Delta }_B)|\mathrm{\Delta }_B_B\}`$ and $`\mu _A(\mathrm{\Delta }_A):=[f^1(\mu _B)](\mathrm{\Delta }_A):=\mu _B[f(\mathrm{\Delta }_A)],\mathrm{\Delta }_A_A`$. * Let two $`\sigma `$-algebras $`_A`$ and $`_B`$ be defined on $`A`$ and $`B`$, respectively, let the measure $`\mu _A`$ be defined on $`A`$ and let the application $`f`$ be measurable, i.e., $`\mathrm{\Delta }_B_B`$ implies $`f^1(\mathrm{\Delta }_B)_A`$. The measure $`\mu _A`$ can be transferred on $`B`$ by defining $`\mu _B(\mathrm{\Delta }_B):=[f(\mu _A)](\mathrm{\Delta }_B):=\mu _A(f^1[\mathrm{\Delta }_B)]`$. Notice that if $`A`$ and $`B`$ have the $`\sigma `$-algebras $`_A`$ and $`_B`$, and $`f`$ is measurable, then $`f^1(_B)_A`$. Transfer of the measure $`\mu _Q`$. In section 3.3, the measure $`\mu _Q`$ was transferred from $`V^N`$ onto $`V_I^N`$ by means of the application $`\omega _V`$, by defining $$\pi _Q:=\omega _V^1(\mu _Q).$$ (41) This definition is not correct for sets containing discontinuity points of the function $`\omega _V`$. The correct definition is the following: consider the application $`X_t:V_I^NR^{3N}`$ which associates with a velocity $`vV_I^N`$ the position at time t of the particles of a classical N-bigbang with origin at $`(0,\mathrm{𝟎})`$ and initial velocity $`v`$. We furthermore indicate by $`\eta _{Qt}`$ the measure $$\eta _{Qt}(\mathrm{\Delta }_X)=\mathrm{𝟎}_N|e^{iHt}E^Q[\mathrm{\Delta }_X]e^{iHt}|\mathrm{𝟎}_N,\mathrm{\Delta }_XR^{3N},$$ (42) hence $$\mu _Q(\mathrm{\Delta })=\underset{t+\mathrm{}}{lim}\eta _{Qt}[\mathrm{\Delta }^c(t)],\mathrm{\Delta }V^N.$$ (43) The measure $`\pi _Q`$ should be defined as $$\pi _Q(\mathrm{\Delta }_I):=\underset{t+\mathrm{}}{lim}\eta _{Qt}[X_t(\mathrm{\Delta }_I)],\mathrm{\Delta }_IV_I^N.$$ (44) One can easily appreciate the difference between the definitions (41) and (44) by applying them to the example in section 2.2.1. Choosing $`b>0`$, we set $`\mathrm{\Delta }_b:=[b,b]V_I`$; the interval $`\mathrm{\Delta }_b`$ embeds the point $`v=0`$ which is a discontinuity point for the function $`\omega _V`$ defined by (7). Since $$\omega _V(\mathrm{\Delta }_b)=[\sqrt{2V_0/m+b^2},\sqrt{2V_0/m}]\{0\}[\sqrt{2V_0/m},\sqrt{2V_0/m+b^2}].$$ by using definition (41) one has $$\pi _Q(\mathrm{\Delta }_b)=\mu _Q\left([\sqrt{2V_0/m+b^2},\sqrt{2V_0/m}]\{0\}[\sqrt{2V_0/m},\sqrt{2V_0/m+b^2}]\right).$$ On the other hand, since for $`t0`$ one has $$X_t(\mathrm{\Delta }_b)[t\sqrt{2V_0/m+b^2},t\sqrt{2V_0/m+b^2}],$$ by using the second definition one obtains $$\pi _Q(\mathrm{\Delta }_b)=\mu _Q\left([\sqrt{2V_0/m+b^2},\sqrt{2V_0/m+b^2}]\right).$$ The application $`\alpha :B_CV_I^N`$ that associates with every classical N-bigbang its own initial velocity allows to define on $`B_C`$ the measure $`\nu _Q`$: $$\nu _Q:=\alpha ^1(\pi _Q).$$ (45) Equations (45) describe the correct definition of $`\nu _Q`$, to be used instead of the one of section 3.2.2. The quotient measure $`\stackrel{~}{\nu }_Q`$ is defined in the same way. Quotient Measure. Let $`A`$ be a set equipped with a $`\sigma `$-algebra $``$ and with a measure $`\mu `$, on which there acts a group of transformations $`G`$, and let the measure $`\mu `$ be invariant under such transformations, i.e., $`\mu (\mathrm{\Delta })=\mu [g(\mathrm{\Delta })],gG`$. The relation of $`G`$-equivalence is defined on $`A`$, so that $`a_1,a_2A`$ are $`G`$-equivalent if there exists a $`gG`$ such that $`a_2=g(a_1)`$. Let $`\stackrel{~}{A}`$ be the quotient space of this relation. The $`\sigma `$-algebra $`\stackrel{~}{}`$ is defined for $`\stackrel{~}{A}`$ by setting $`\stackrel{~}{}:=𝒫(\stackrel{~}{A})`$, where $`𝒫(\stackrel{~}{A})`$ is the set of parts of $`\stackrel{~}{A}`$. The $`\sigma `$-algebra $`\stackrel{~}{}`$ is formed by the $``$-measurable sets that are invariant under the transformations of the group G, and can also be thought as a $`\sigma `$-subalgebra of $``$. The aim here is to define on $`\stackrel{~}{A}`$ a measure $`\stackrel{~}{\mu }`$ which corresponds to the measure $`\mu `$ on $`A`$. Let $`BA`$ be an $``$-measurable set which represents the set $`\stackrel{~}{A}`$, i.e., which contains one and only one element for every $`G`$-equivalence class, and let $`h:\stackrel{~}{A}B`$ be the corresponding projection. The application $`k:B\times GA`$ is defined by setting $`k(a,g):=g(a)`$. By Haar’s theorem , on $`G`$ there exists a single measure $`\mu _H`$ (defined modulo a proportionality factor) which is left-invariant, i.e., such that $`\mu _H(g\mathrm{\Delta }_G)=\mu _H(\mathrm{\Delta }_G)`$ for every $`\mathrm{\Delta }_GG`$. Let us define measure $`\stackrel{~}{\nu }`$ on $`B`$ by setting $$\stackrel{~}{\nu }(\mathrm{\Delta }_B)=\frac{\mu [k(\mathrm{\Delta }_B\times \mathrm{\Delta }_G)]}{\mu _H(\mathrm{\Delta }_G)},$$ (46) where $`\mathrm{\Delta }_GG`$ is a set such that $`0<\mu _H(\mathrm{\Delta }_G)<+\mathrm{}`$. One can prove that $`\stackrel{~}{\nu }`$ depends neither on the choice of the set $`B`$ nor on the choice of the set $`\mathrm{\Delta }_G`$. Finally, the quotient measure $`\stackrel{~}{\mu }`$ is defined by setting $$\stackrel{~}{\mu }:=h^1(\stackrel{~}{\nu }).$$ (47) Example. Let $`A=R^2`$, let $`\mu `$ be Lebesgue’s measure and let $`G`$ be the group of translations along $`x:g(x,y)=(x+x_g,y)`$. The elements of the quotient set $`\stackrel{~}{A}`$ are the straight lines that are parallel to the $`x`$-axis, while the elements of $`\stackrel{~}{}`$ are the measurable “strips” that are parallel to the $`x`$-axis. The set $`B`$ can be, for instance, the straight line $`x=0`$. The group $`G`$ is isomorphic to $`R`$, and its Haar measure is the Lebesgue measure on $`R`$. If $`\mathrm{\Delta }_B`$ is a subset of $`B`$, for instance an interval, and $`\mathrm{\Delta }_G`$ is another interval, then $`k(\mathrm{\Delta }_B\times \mathrm{\Delta }_G)`$ is a rectangle with base $`\mathrm{\Delta }_G`$ and height $`\mathrm{\Delta }_B`$, $`\mu [k(\mathrm{\Delta }_B\times \mathrm{\Delta }_G)]`$ its area, and $`\stackrel{~}{\nu }(\mathrm{\Delta }_B)`$ is the length of the interval $`\mathrm{\Delta }_B`$. Comparison between classical and quantum measure. The classical measure $`\mu _C`$ and the quantum measure $`\mu _Q`$ on $`V^N`$ \[see equations (13) and (37)\] can be derived by taking to the limit $`t+\mathrm{}`$ the following two measures: * Classical measure: $`\eta _{Ct}(\mathrm{\Delta }_X):=\pi _C[X_t^1(\mathrm{\Delta }_X)],\mathrm{\Delta }_XR^{3N}`$, where $`X_t:V_I^NR^{3N}`$ is defined above in the appendix, and $`\pi _C`$ is the uniform measure on momenta defined by (34). * Quantum measure: $`\eta _{Qt}(\mathrm{\Delta }_X):=\mathrm{𝟎}_N|e^{iHt}E^Q[\mathrm{\Delta }_X]e^{iHt}|\mathrm{𝟎}_N,\mathrm{\Delta }_XR^{3N}`$. One can easily see that for both measures $$\mu _{C(Q)}(\mathrm{\Delta })=\underset{t+\mathrm{}}{lim}\eta _{C(Q)t}[\mathrm{\Delta }^c(t)],\mathrm{\Delta }V^N$$ (48) holds. Let us indicate with $`\rho _C(t,x)`$ and $`\rho _Q(t,x)`$ the densities that correspond to the measures $`\eta _{Ct}`$ and $`\eta _{Qt}`$. In it is demonstrated that $$\rho _C(t,x)=\frac{1}{(2\pi )^{3N}}\underset{i}{}\left|det\left(\frac{W_i}{x_1x_2}\right)\right|(0,0,t,x),$$ (49) where $`W_i(t_1,x_1,t_2,x_2)`$ is the action of the $`i`$-th classical trajectory that joins the points $`(t_1,x_1)`$ and $`(t_2,x_2)`$. The same paper demonstrates that in a semi-classical approximation, $`\rho _Q`$ is given by $$\rho _Q(t,x)=\frac{1}{(2\pi )^{3N}}\left|\underset{i}{}\left|det\left(\frac{W_i}{x_1x_2}\right)\right|^{1/2}\mathrm{exp}\left\{i\left(W_i\frac{M_i\pi }{2}\right)\right\}\right|^2(0,0,t,x),$$ (50) where $`M_i`$ is a phase factor. From (49) and from (50) one obtains $$\rho _C(t,x)=\rho _Q(t,x)+I(t,x),$$ (51) where $`I(t,x)`$ is the following interference term: $`I(t,x)`$ $`=`$ $`{\displaystyle \underset{ij}{}}\left|{\displaystyle \frac{W_i}{x_1x_2}}|^{1/2}\right|{\displaystyle \frac{W_j}{x_1x_2}}|^{1/2}\times `$ (52) $`\times \mathrm{exp}\left[i\left((W_iW_j){\displaystyle \frac{(M_iM_j)\pi }{2}}\right)\right](0,0,t,x).`$
warning/0002/math-ph0002048.html
ar5iv
text
# 1 Introduction ## 1 Introduction At present there exists an interest to the so-called $`M`$-theory (see, for example, -). This theory is “supermembrane” analogue of superstring models in $`D=11`$. The low-energy limit of $`M`$-theory after a dimensional reduction leads to models governed by a Lagrangian containing a metric, fields of forms and scalar fields. These models contain a large variety of the so-called $`p`$-brane solutions (see - and references therein). In it was shown that after the dimensional reduction on the manifold $`M_0\times M_1\times \mathrm{}\times M_n`$ when the composite $`p`$-brane ansatz for fields of forms is considered the problem is reduced to the gravitating self-interacting $`\sigma `$-model with certain constraints imposed. (For electric $`p`$-branes see also .) This representation may be considered as a tool for obtaining different solutions with intersecting $`p`$-branes. In the Majumdar-Papapetrou type solutions (see ) were obtained (for non-composite case see ). These solutions corresponding to Ricci-flat factor-spaces $`(M_i,g^i)`$, ($`g^i`$ is metric on $`M_i`$) $`i=1,\mathrm{},n`$, were also generalized to the case of Einstein internal spaces . Earlier some special classes of these solutions were considered in . The obtained solutions take place, when certain (block-)orthogonality relations (on couplings parameters, dimensions of ”branes”, total dimension) are imposed. In this situation a class of cosmological and spherically-symmetric solutions was obtained . Special cases were also considered in . The solutions with the horizon were studied in details in . In models under consideration there exists a large variety of Toda-chain solutions, when certain intersection rules are satisfied . Cosmological and spherically symmetric solutions with $`p`$-branes and $`n`$ internal spaces related to $`A_m`$ Toda chains were previously considered in and . Recently in a family of $`p`$-brane solutions depending on one harmonic function with nearly arbitrary (up to some restrictions) intersection rules were obtained. These solutions are defined up to solutions of Laplace and Toda-type equations and correspond to null-geodesics of the sigma-model target-space metric. Here we consider a family of spherically-symmetric and cosmological type solutions from (see Sect. 2) and single out a new subclass of black-hole configurations related to Toda-type equations with certain asymptotical conditions imposed (Sect. 3). These black hole solutions are governed by functions $`H_s(z)>0`$ defined on the interval $`(0,(2\mu )^1)`$ ($`\mu >0`$) and obeying a set of second order non-linear differential equations $$\frac{d}{dz}\left(\frac{(12\mu z)}{H_s}\frac{d}{dz}H_s\right)=\overline{B}_s\underset{s^{}S}{}H_s^{}^{A_{ss^{}}},$$ (1.1) with the following boundary conditions imposed: $`(𝐢)H_s((2\mu )^10)=H_{s0}(0,+\mathrm{});`$ (1.2) $`(\mathrm{𝐢𝐢})H_s(+0)=1,`$ (1.3) $`sS`$. In (1.1) $`\overline{B}_s>0`$, $`sS`$, and $`(A_{ss^{}})`$ is a ”quasi-Cartan” matrix ($`A_{ss}=2`$, $`sS`$) coinciding with the Cartan one when intersections are related to Lie algebras. Equations (1.1) are equivalent to Toda-type equations (see Sect. 2). For positively defined scalar field metric $`(h_{\alpha \beta })`$ all $`p`$-branes in this solution should contain a time manifold (see Proposition 1 from Sect. 3). This agrees with Theorem 3 from Ref. (for orthogonal case, see also ). In Sect. 4 we suggest a hypothesis: the functions $`H_s`$ are polynomials when intersection rules correspond to semisimple Lie algebras. This hypothesis (Conjecture ) is proved for Lie algebras: $`A_m`$, $`C_{m+1}`$, $`m=1,2,\mathrm{}`$. It is also confirmed by special black-hole “block orthogonal” solutions considered earlier in . An analogue of this conjecture for extremal black holes was considered earlier in . In Sect. 5 explicit formulas for the solution corresponding to the algebra $`A_2`$ are obtained. These formulas are illustrated by two examples of $`A_2`$-dyon solutions: a dyon in $`D=11`$ supergravity (with $`M2`$ and $`M5`$ branes intersecting at a point) and Kaluza-Klein dyon. ## 2 The model and Toda-type solutions We consider a model governed by the action $`S={\displaystyle d^Dx\sqrt{|g|}\left\{R[g]h_{\alpha \beta }g^{MN}_M\phi ^\alpha _N\phi ^\beta \underset{a\mathrm{}}{}\frac{\theta _a}{n_a!}\mathrm{exp}[2\lambda _a(\phi )](F^a)^2\right\}}`$ (2.1) where $`g=g_{MN}(x)dx^Mdx^N`$ is a metric, $`\phi =(\phi ^\alpha )\text{R}^l`$ is a vector of scalar fields, $`(h_{\alpha \beta })`$ is a constant symmetric non-degenerate $`l\times l`$ matrix $`(l\text{N})`$, $`\theta _a=\pm 1`$, $$F^a=dA^a=\frac{1}{n_a!}F_{M_1\mathrm{}M_{n_a}}^adz^{M_1}\mathrm{}dz^{M_{n_a}}$$ (2.2) is a $`n_a`$-form ($`n_a1`$), $`\lambda _a`$ is a 1-form on $`\text{R}^l`$: $`\lambda _a(\phi )=\lambda _{\alpha a}\phi ^\alpha `$, $`a\mathrm{}`$, $`\alpha =1,\mathrm{},l`$. In (2.1) we denote $`|g|=|det(g_{MN})|`$, $$(F^a)_g^2=F_{M_1\mathrm{}M_{n_a}}^aF_{N_1\mathrm{}N_{n_a}}^ag^{M_1N_1}\mathrm{}g^{M_{n_a}N_{n_a}},$$ (2.3) $`a\mathrm{}`$. Here $`\mathrm{}`$ is some finite set. Let us consider a family of one-variable sector solutions to field equations corresponding to the action (2.1) and depending upon one variable $`u`$ . These solutions are defined on the manifold $$M=(u_{},u_+)\times M_1\times M_2\times \mathrm{}\times M_n,$$ (2.4) where $`(u_{},u_+)`$ is an interval belonging to R. The solutions read $`g=\left({\displaystyle \underset{sS}{}}[f_s(u)]^{2d(I_s)h_s/(D2)}\right)\{[f_1(u)]^{2d_1/(1d_1)}\mathrm{exp}(2c^1u+2\overline{c}^1)`$ (2.5) $`\times [wdudu+f_1^2(u)g^1]+{\displaystyle \underset{i=2}{\overset{n}{}}}\left({\displaystyle \underset{sS}{}}[f_s(u)]^{2h_s\delta _{iI_s}}\right)\mathrm{exp}(2c^iu+2\overline{c}^i)g^i\},`$ $`\mathrm{exp}(\phi ^\alpha )=\left({\displaystyle \underset{sS}{}}f_s^{h_s\chi _s\lambda _{a_s}^\alpha }\right)\mathrm{exp}(c^\alpha u+\overline{c}^\alpha ),`$ (2.6) $`F^a={\displaystyle \underset{sS}{}}\delta _{a_s}^a^s,`$ (2.7) $`\alpha =1,\mathrm{},l`$. In (2.5) $`w=\pm 1`$, $`g^i=g_{m_in_i}^i(y_i)dy_i^{m_i}dy_i^{n_i}`$ is a Ricci-flat metric on $`M_i`$, $`i=2,\mathrm{},n`$, the space $`(g^1,M_1)`$ is an Einstein space of non-zero curvature: $$R_{mn}[g^1]=\xi ^1g^1,$$ (2.8) $`\xi ^10`$, and $$\delta _{iI}=\underset{jI}{}\delta _{ij}$$ (2.9) is the indicator of $`i`$ belonging to $`I`$: $`\delta _{iI}=1`$ for $`iI`$ and $`\delta _{iI}=0`$ otherwise. The $`p`$-brane set $`S`$ is by definition $`S=S_eS_m,S_v=_a\mathrm{}\{a\}\times \{v\}\times \mathrm{\Omega }_{a,v},`$ (2.10) $`v=e,m`$ and $`\mathrm{\Omega }_{a,e},\mathrm{\Omega }_{a,m}\mathrm{\Omega }`$, where $`\mathrm{\Omega }=\mathrm{\Omega }(n)`$ is the set of all non-empty subsets of $`\{2,\mathrm{},n\}`$. Hence all $`p`$-branes do not “live” in $`M_1`$. Any $`p`$-brane index $`sS`$ has the form $`s=(a_s,v_s,I_s),`$ (2.11) where $`a_s\mathrm{}`$, $`v_s=e,m`$ and $`I_s\mathrm{\Omega }_{a_s,v_s}`$. The sets $`S_e`$ and $`S_m`$ define electric and magnetic $`p`$-branes correspondingly. In (2.6) $`\chi _s=+1,1`$ (2.12) for $`sS_e,S_m`$ respectively. In (2.7) forms $$^s=Q_s\left(\underset{s^{}S}{}f_s^{}^{A_{ss^{}}}\right)du\tau (I_s),$$ (2.13) $`sS_e`$, correspond to electric $`p`$-branes and forms $$^s=Q_s\tau (\overline{I}_s),$$ (2.14) correspond to magnetic $`p`$-branes; $`Q_s0`$, $`sS`$. In (2.14) and in what follows $$\overline{I}\{1,\mathrm{},n\}I.$$ (2.15) All the manifolds $`M_i`$, $`i>1`$, are assumed to be oriented and connected and the volume $`d_i`$-forms $$\tau _i\sqrt{|g^i(y_i)|}dy_i^1\mathrm{}dy_i^{d_i},$$ (2.16) are well–defined for all $`i=1,\mathrm{},n`$. Here $`d_i=\mathrm{dim}M_i`$, $`i=1,\mathrm{},n`$ (in spherically symmetric case $`M_1=S^{d_1}`$), $`d_1>1`$, $`D=1+_{i=1}^nd_i`$, and for any $`I=\{i_1,\mathrm{},i_k\}\mathrm{\Omega }`$, $`i_1<\mathrm{}<i_k`$, we denote $`\tau (I)\tau _{i_1}\mathrm{}\tau _{i_k},`$ (2.17) $`M_IM_{i_1}\times \mathrm{}\times M_{i_k},`$ (2.18) $`d(I)\mathrm{dim}M_I={\displaystyle \underset{iI}{}}d_i.`$ (2.19) The parameters $`h_s`$ appearing in the solution satisfy the relations $$h_s=K_s^1,K_s=B_{ss},$$ (2.20) where $`B_{ss^{}}d(I_sI_s^{})+{\displaystyle \frac{d(I_s)d(I_s^{})}{2D}}+\chi _s\chi _s^{}\lambda _{\alpha a_s}\lambda _{\beta a_s^{}}h^{\alpha \beta },`$ (2.21) $`s,s^{}S`$, with $`(h^{\alpha \beta })=(h_{\alpha \beta })^1`$. Here we assume that $$(𝐢)B_{ss}0,$$ (2.22) for all $`sS`$, and $$(\mathrm{𝐢𝐢})\mathrm{det}(B_{ss^{}})0,$$ (2.23) i.e. the matrix $`(B_{ss^{}})`$ is a non-degenerate one. In (2.13) another non-degenerate matrix (“a quasi-Cartan” matrix) appears $$(A_{ss^{}})=\left(2B_{ss^{}}/B_{s^{}s^{}}\right).$$ (2.24) Here some ordering in $`S`$ is assumed. This matrix also appears in the relations for $$f_s=\mathrm{exp}(q^s),$$ (2.25) where $`(q^s)=(q^s(u))`$ is a solution to Toda-type equations $$\ddot{q^s}=B_s\mathrm{exp}(\underset{s^{}S}{}A_{ss^{}}q^s^{}),$$ (2.26) with $$B_s=2K_sA_s,A_s=\frac{1}{2}\epsilon _sQ_s^2,$$ (2.27) $`sS`$. Here $$\epsilon _s=(\epsilon [g])^{(1\chi _s)/2}\epsilon (I_s)\theta _{a_s},$$ (2.28) $`sS`$, $`\epsilon [g]signdet(g_{MN})`$. More explicitly (2.28) reads: $`\epsilon _s=\epsilon (I_s)\theta _{a_s}`$ for $`v_s=e`$ and $`\epsilon _s=\epsilon [g]\epsilon (I_s)\theta _{a_s}`$, for $`v_s=m`$. In (2.5) $`f_1(u)=Rsh(\sqrt{C_1}u),C_1>0,\xi _1w>0;`$ (2.29) $`R\mathrm{sin}(\sqrt{|C_1|}u),C_1<0,\xi _1w>0;`$ (2.30) $`Rch(\sqrt{C_1}u),C_1>0,\xi _1w<0;`$ (2.31) $`\left|\xi _1(d_11)\right|^{1/2}u,C_1=0,\xi _1w>0,`$ (2.32) where $`C_1`$ is constant and $`R=|\xi _1(d_11)/C_1|^{1/2}`$. Vectors $`c=(c^A)=(c^i,c^\alpha )`$ and $`\overline{c}=(\overline{c}^A)`$ satisfy the linear constraints $`U^s(c)={\displaystyle \underset{iI_s}{}}d_ic^i\chi _s\lambda _{a_s\alpha }c^\alpha =0,`$ (2.33) $`U^s(\overline{c})={\displaystyle \underset{iI_s}{}}d_i\overline{c}^i\chi _s\lambda _{a_s\alpha }\overline{c}^\alpha =0,`$ (2.34) $`sS`$, $`U^1(c)=c^1+{\displaystyle \underset{j=1}{\overset{n}{}}}d_jc^j=0,`$ (2.35) $`U^1(\overline{c})=\overline{c}^1+{\displaystyle \underset{j=1}{\overset{n}{}}}d_j\overline{c}^j=0,`$ (2.36) and $$C_1\frac{d_1}{d_11}=2E_{TL}+h_{\alpha \beta }c^\alpha c^\beta +\underset{i=2}{\overset{n}{}}d_i(c^i)^2+\frac{1}{d_11}\left(\underset{i=2}{\overset{n}{}}d_ic^i\right)^2,$$ (2.37) where $$E_{TL}=\frac{1}{4}\underset{s,s^{}S}{}h_sA_{ss^{}}\dot{q^s}\dot{q^s^{}}+\underset{sS}{}A_s\mathrm{exp}(\underset{s^{}S}{}A_{ss^{}}q^s^{}),$$ (2.38) is an integration constant (energy) for the solutions from (2.26). We note that the eqs. (2.26) correspond to the Toda-type Lagrangian $$L_{TL}=\frac{1}{4}\underset{s,s^{}S}{}h_sA_{ss^{}}\dot{q^s}\dot{q^s^{}}\underset{sS}{}A_s\mathrm{exp}(\underset{s^{}S}{}A_{ss^{}}q^s^{}).$$ (2.39) Remark 1. Here we identify notations for $`g^i`$ and $`\widehat{g}^i`$, where $`\widehat{g}^i=p_i^{}g^i`$ is the pullback of the metric $`g^i`$ to the manifold $`M`$ by the canonical projection: $`p_i:MM_i`$, $`i=1,\mathrm{},n`$. An analogous agreement will be also kept for volume forms etc. Due to (2.13) and (2.14), the dimension of $`p`$-brane worldsheet $`d(I_s)`$ is defined by $`d(I_s)=n_{a_s}1,d(I_s)=Dn_{a_s}1,`$ (2.40) for $`sS_e,S_m`$ respectively. For a $`p`$-brane: $`p=p_s=d(I_s)1`$. The solutions are valid if the following restrictions on the sets $`\mathrm{\Omega }_{a,v}`$ are imposed. These restrictions guarantee the block-diagonal structure of the stress-energy tensor, like for the metric, and the existence of $`\sigma `$-model representation (see also ). We denote $`w_1\{i|i\{2,\mathrm{},n\},d_i=1\}`$, and $`n_1=|w_1|`$ (i.e. $`n_1`$ is the number of 1-dimensional spaces among $`M_i`$, $`i=1,\mathrm{},n`$). Restriction 1. Let 1a) $`n_11`$ or 1b) $`n_12`$ and for any $`a\mathrm{}`$, $`v\{e,m\}`$, $`i,jw_1`$, $`i<j`$, there are no $`I,J\mathrm{\Omega }_{a,v}`$ such that $`iI`$, $`jJ`$ and $`I\{i\}=J\{j\}`$. Restriction 2. Let 2a) $`n_1=0`$ or 2b) $`n_11`$ and for any $`a\mathrm{}`$, $`iw_1`$ there are no $`I\mathrm{\Omega }_{a,m}`$, $`J\mathrm{\Omega }_{a,e}`$ such that $`\overline{I}=\{i\}J`$. These restrictions are satisfied in the non-composite case : $`|\mathrm{\Omega }_{a,e}|+|\mathrm{\Omega }_{a,m}|=1`$, (i.e when there are no two $`p`$-branes with the same color index $`a`$, $`a\mathrm{}`$.) Restriction 1 and 2 forbid certain intersections of two $`p`$-branes with the same color index for $`n_12`$ and $`n_11`$ respectively. Restriction 2 is satisfied identically if all $`p`$-branes contain a common manifold $`M_j`$ (say, time manifold). This solution describes a set of charged (by forms) overlapping $`p`$-branes ($`p_s=d(I_s)1`$, $`sS`$) “living” on submanifolds of $`M_2\times \mathrm{}\times M_n`$. ### 2.1 $`U^s`$-vectors and scalar products Here we consider a minisuperspace covariant form of constraints and corresponding scalar products that will be used in the next section. The linear constraints (2.33)-(2.36) may be written in the following form $$U^r(c)=U_A^rc^A=0,U^r(\overline{c})=U_A^r\overline{c}^A=0,$$ (2.41) $`r=s,1`$, where $`(U_A^s)=(d_i\delta _{iI_s},\chi _s\lambda _{\alpha a_s}),`$ (2.42) $`s=(a_s,v_s,I_s)S`$, and $`(U_A^1)=(\delta _i^1+d_i,0),`$ (2.43) $`A=(i,\alpha )`$. The quadratic constraint (2.37) reads $$E=E_1+E_{TL}+\frac{1}{2}\widehat{G}_{AB}c^Ac^B=0,$$ (2.44) where $`C_1=2E_1(U^1,U^1)`$, $$(U^1,U^1)=1/d_11,$$ (2.45) ($`d_1>1`$) and $`(\widehat{G}_{AB})=\left(\begin{array}{cc}G_{ij}& 0\\ 0& h_{\alpha \beta }\end{array}\right),`$ (2.48) is the target space metric with $`G_{ij}=d_i\delta _{ij}d_id_j,`$ (2.49) $`i,j=1,\mathrm{},n`$. In (2.45) a scalar product appears $`(U,U^{})=\widehat{G}^{AB}U_AU_B^{},`$ (2.50) where $`U=U_Az^A`$, $`U^{}=U_A^{}z^A`$ are linear functions on $`\text{R}^{n+l}`$, and $`(\widehat{G}^{AB})=(\widehat{G}_{AB})^1`$. The scalar products (2.50) for co-vectors $`U^s`$ from $`(\text{2.42})`$ were calculated in $$(U^s,U^s^{})=B_{ss^{}},$$ (2.51) $`s,s^{}S`$ (see (2.21)). It follows from (2.23) and (2.51) that the vectors $`U^s`$, $`sS`$, are linearly independent. Hence, the number of the vectors $`U^s`$ should not exceed the dimension of the dual space $`(𝐑^{n+l})^{}`$, i.e. $$|S|n+l.$$ (2.52) We also get $$(U^s,U^1)=0,$$ (2.53) for all $`sS`$. This relation takes place, since all $`p`$-branes do not live in $`M_1`$: $`I_s\{2,\mathrm{},n\}`$. Intersection rules. From (2.20), (2.21) and (2.24) we get the intersection rules corresponding to the quasi-Cartan matrix $`(A_{ss^{}})`$ $$d(I_sI_s^{})=\frac{d(I_s)d(I_s^{})}{D2}\chi _s\chi _s^{}\lambda _{a_s}\lambda _{a_s^{}}+\frac{1}{2}K_s^{}A_{ss^{}},$$ (2.54) where $`\lambda _{a_s}\lambda _{a_s^{}}=\lambda _{\alpha a_s}\lambda _{\beta a_s^{}}h^{\alpha \beta }`$, $`s,s^{}S`$. The contravariant components $`U^{rA}=\widehat{G}^{AB}U_B^r`$ reads $$U^{si}=G^{ij}U_j^s=\delta _{iI_s}\frac{d(I_s)}{D2},U^{s\alpha }=\chi _s\lambda _{a_s}^\alpha ,$$ (2.55) $$U^{1i}=\frac{\delta _1^i}{d_1},U^{1\alpha }=0,$$ (2.56) $`sS`$. Here (as in ) $$G^{ij}=\frac{\delta ^{ij}}{d_i}+\frac{1}{2D},$$ (2.57) $`i,j=1,\mathrm{},n`$, are the components of the matrix inverse to $`(G_{ij})`$ from (2.49). The contravariant components (2.55) and (2.56) occur as powers in relations for the metric and scalar fields in (2.5) and (2.6). We note that the solution under consideration for the special case of the $`A_m`$ Toda chain was obtained earlier in . Special $`A_1\mathrm{}A_1`$ Toda case, when vectors $`U^s`$ are mutually orthogonal, was considered earlier in (for non-composite case see also ). For a (general) block-orthogonal set of vectors $`U^s`$ special solutions were considered in . ## 3 Black holes solutions ### 3.1 The choice of parameters Here we consider the spherically symmetric case: $$w=1,M_1=S^{d_1},g^1=d\mathrm{\Omega }_{d_1}^2,$$ (3.1) where $`d\mathrm{\Omega }_{d_1}^2`$ is the canonical metric on a unit sphere $`S^{d_1}`$, $`d_12`$. In this case $`\xi ^1=d_11`$. We also assume that $$M_2=\text{R},g^2=dtdt,$$ (3.2) i.e. $`M_2`$ is a time manifold. We put $`C_10`$. In this case relations (2.29)-(2.32) read $`f_1(u)=\overline{d}{\displaystyle \frac{sh(\sqrt{C_1}u)}{\sqrt{C_1}}},C_1>0,`$ (3.3) $`\overline{d}u,C_1=0.`$ (3.4) Here and in what follows $$\overline{d}=d_11.$$ (3.5) Let us consider the null-geodesic equations for the light “moving” in the radial direction (following from $`ds^2=0`$): $`{\displaystyle \frac{dt}{du}}=\pm \mathrm{\Phi },`$ (3.6) $`\mathrm{\Phi }=f_1^{d_1/(1d_1)}e^{(c^1c^2)u+\overline{c}^1\overline{c}^2}{\displaystyle \underset{sS}{}}f_s^{h_s\delta _{2I_s}},`$ (3.7) equivalent to $$tt_0=\pm _{u_0}^u𝑑\overline{u}\mathrm{\Phi }(\overline{u}),$$ (3.8) where $`t_0,u_0`$ are constants. Let us consider solutions (defined on some interval $`[u_0,+\mathrm{})`$) with a horizon at $`u=+\mathrm{}`$ satisfying $$_{u_0}^+\mathrm{}𝑑u\mathrm{\Phi }(u)=+\mathrm{}.$$ (3.9) Here we restrict ourselves to solutions with $`C_1>0`$ and linear asymptotics at infinity $$q^s=\beta ^su+\overline{\beta }^s+o(1),$$ (3.10) $`u+\mathrm{}`$, where $`\beta ^s,\overline{\beta }^s`$ are constants, $`sS`$. This relation gives us an asymptotical solution to Toda type eqs. (2.26) if $$\underset{s^{}S}{}A_{ss^{}}\beta ^s^{}>0,$$ (3.11) for all $`sS`$. In this case the energy (2.38) reads $$E_{TL}=\frac{1}{4}\underset{s,s^{}S}{}h_sA_{ss^{}}\beta ^s\beta ^s^{}.$$ (3.12) Remark 2. For positive-definite matrices $`(h_sA_{ss^{}})`$ and $`(h_{\alpha \beta })`$ we get from (2.37) and (3.12): $`E_{TL}0`$, $`C_10`$. (For the extremal case $`E_{TL}=C_1=0`$ see Sect. 7.) According to Lemma 2 from black hole solutions can only exist for $`C_10`$ and the horizon is then at $`u=\mathrm{}`$. For the function (3.7) we get $$\mathrm{\Phi }(u)\mathrm{\Phi }_0e^{\beta u},u+\mathrm{},$$ (3.13) where $`\mathrm{\Phi }_00`$ is constant, $$\beta =c^1c^2+\sqrt{C_1}h_1+\underset{sS}{}\beta _sh_s\delta _{2I_s},$$ (3.14) and $$h_1=(U^1,U^1)^1=\frac{d_1}{1d_1}.$$ (3.15) Horizon at $`u=+\mathrm{}`$ takes place if and only if $$\beta 0.$$ (3.16) Let us introduce dimensionless parameters $$b^s=\beta ^s/\sqrt{C_1},b^A=c^A/\sqrt{C_1},$$ (3.17) where $`sS`$, $`A=(i,\alpha )`$, $`C_1>0`$. Thus, a horizon at $`u=+\mathrm{}`$ corresponds to a point $`b=(b^s,b^A)\text{R}^{|S|+n+l}`$ satisfying the relations following from (2.41), (2.44), (3.11), (3.12) and (3.14)-(3.17): $`U_A^rb^A=0,r=s,1;sS,`$ (3.18) $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{s,s^{}S}{}}h_sA_{ss^{}}b^sb^s^{}+\widehat{G}_{AB}b^Ab^B=|h_1|,`$ (3.19) $`{\displaystyle \underset{s^{}S}{}}A_{ss^{}}b^s^{}>0,`$ (3.20) $`f(b)b^1b^2+{\displaystyle \underset{sS}{}}b_sh_s\delta _{2I_s}|h_1|.`$ (3.21) Proposition 1. Let matrix $`(h_{\alpha \beta })`$ be positively defined. Then the point $`b=(b^s,b^A)`$ satisfying relations (3.18)-(3.21) exists only if $$2I_s,sS,$$ (3.22) (i. e. all p-branes have a common time direction $`t`$) and is unique: $`b=b_0`$, where $`b_0^A=\delta _2^A+h_1U^{1A}+{\displaystyle \underset{sS}{}}h_sb_0^sU^{sA},`$ (3.23) $`b_0^s=2{\displaystyle \underset{s^{}S}{}}A^{ss^{}},`$ (3.24) where $`sS`$, $`A=(i,\alpha )`$, and the matrix $`(A^{ss^{}})`$ is inverse to the matrix $`(A_{ss^{}})=(2(U^s,U^s^{})/(U^s^{},U^s^{}))`$. Proof. Let $``$ be a manifold described by relations (3.18)-(3.19). This manifold is an ellipsoid. Indeed, due to positively definiteness of $`(h_{\alpha \beta })`$ the matrix $`\widehat{G}_{AB}`$ has a signature $`(,+,\mathrm{},+)`$, since the matrix $`(G_{ij})`$ from (2.49) has a signature $`(,+,\mathrm{},+)`$ . Due to relations $`(U^1,U^1)<0`$, $`(U^1,U^s)=0`$, $`(U^s,U^s)0`$ for all $`sS`$, and (2.23) the matrices $`(B_{ss^{}})`$ and $`(A_{ss^{}})`$ are positively defined and all $`h_s>0`$, $`sS`$. Then, the quadratic form in (3.19) has a pseudo-Euclidean signature. Due to $`(U^1,U^1)<0`$ the intersection of the hyperboloid (3.19) with the (multidimensional) plane $`U_A^1z^A=0`$ gives us an ellipsoid. Its intersection with the planes $`U_A^sz^A=0`$, $`sS`$, give us to an ellipsoid, coinciding with $``$. Let us consider a function $`f_|:\text{R}`$ that is a restriction of the linear function (3.21) on $``$. Let $`b_{}`$ be a point of maximum of $`f_|`$. Using the conditional extremum method and the fact that $``$ is ellipsoid we prove that that $`b_{}^A=\delta _2^A+h_1U^{1A}+{\displaystyle \underset{sS}{}}h_sb_{}^sU^{sA},`$ (3.25) $`b_{}^s=2{\displaystyle \underset{s^{}S}{}}A^{ss^{}}\delta _{2I_s^{}},`$ (3.26) $`sS`$, $`A=(i,\alpha )`$. Let us consider the function $`\overline{f}(b,\lambda )f(b)\lambda _1U_A^1b^A{\displaystyle \underset{sS}{}}\lambda _sU_A^sb^A\lambda _0\left({\displaystyle \underset{s,s^{}S}{}}{\displaystyle \frac{h_s}{2}}A_{ss^{}}b^sb^s^{}+\widehat{G}_{AB}b^Ab^B+h_1\right),`$ (3.27) where $`\lambda =(\lambda _0,\lambda _1,\lambda _s)`$ is a vector of Lagrange multipliers. The points of extremum for the function $`\overline{f}`$ from (3.27) have the form $`(\lambda _0b_{},\lambda )`$ with $`b_{}`$ from (3.25) and $`\lambda _0=\pm 1,\lambda _1=1/(d_11),\lambda _s=2{\displaystyle \underset{s^{}S}{}}h_sA^{ss^{}}\delta _{2I_s^{}},`$ (3.28) $`sS`$. Then, the points $`b_{}`$ and $`b_{}`$ are the points of maximum and minimum, respectively, for the function $`f_|`$ defined on the ellipsoid $``$. Since $`f(b_{})=|h_1|`$, the only point satisfying the restriction $`f(b)|h_1|`$ is $`b=b_{}`$. From (3.20) we get $`{\displaystyle \underset{s^{}S}{}}A_{ss^{}}b^s^{}=2\delta _{2I_s}>02I_s,`$ (3.29) for all $`sS`$. The proposition is proved. We introduce a new radial variable $`R=R(u)`$ by relations $`\mathrm{exp}(2\overline{\mu }u)=1{\displaystyle \frac{2\mu }{R^{\overline{d}}}}=F,\overline{\mu }=\sqrt{C_1},\mu =\overline{\mu }/\overline{d}>0,`$ (3.30) $`u>0`$, $`R^{\overline{d}}>2\mu `$ ($`\overline{d}=d_11`$). We put $`\overline{c}^A=0,`$ (3.31) $`q^s(0)=0.`$ (3.32) $`A=(i,\alpha )`$, $`sS`$. These relations guarantee the asymptotical flatness (for $`R+\mathrm{}`$) of the $`(2+d_1)`$-dimensional section of the metric. Let us denote $$H_s=f_se^{\overline{\mu }b_0^su},$$ (3.33) $`sS`$. Then, solutions (2.5)-(2.7) may be written as follows $`g=\left({\displaystyle \underset{sS}{}}H_s^{2h_sd(I_s)/(D2)}\right)\{F^1dRdR+R^2d\mathrm{\Omega }_{d_1}^2`$ (3.34) $`\left({\displaystyle \underset{sS}{}}H_s^{2h_s}\right)Fdtdt+{\displaystyle \underset{i=3}{\overset{n}{}}}\left({\displaystyle \underset{sS}{}}H_s^{2h_s\delta _{iI_s}}\right)g^i\},`$ $`\mathrm{exp}(\phi ^\alpha )={\displaystyle \underset{sS}{}}H_s^{h_s\chi _s\lambda _{a_s}^\alpha },`$ (3.35) $`F^a={\displaystyle \underset{sS}{}}\delta _{a_s}^a^s,`$ (3.36) where $$^s=\frac{Q_s}{R^{d_1}}\left(\underset{s^{}S}{}H_s^{}^{A_{ss^{}}}\right)dR\tau (I_s),$$ (3.37) $`sS_e`$, $$^s=Q_s\tau (\overline{I}_s),$$ (3.38) $`sS_m`$. Here $`Q_s0`$, $`h_s=K_s^1`$; parameters $`K_s0`$ and the non-degenerate matrix $`(A_{ss^{}})`$ are defined by relations (2.54) and $`(A_{ss})=2`$, $`sS`$. Functions $`H_s>0`$ obey the equations $$R^{d_1}\frac{d}{dR}\left(R^{d_1}\frac{F}{H_s}\frac{dH_s}{dR}\right)=B_s\underset{s^{}S}{}H_s^{}^{A_{ss^{}}},$$ (3.39) $`sS`$, where $`B_s0`$ are defined in (2.27) and (2.28). These equations follow from Toda-type equations (2.26) and the definition (3.30) and (3.33). It follows from (3.10), (3.17) and (3.33) that there exist finite limits $$H_sH_{s0}0,$$ (3.40) for $`R^{\overline{d}}2\mu `$, $`sS`$. We note, that in this case the metric (3.34) does really have a horizon at $`R^{\overline{d}}=2\mu `$. From (3.32) we get. $$H_s(R=+\mathrm{})=1,$$ (3.41) $`sS`$. The metric (3.34) has a regular horizon at $`R^{\overline{d}}=2\mu `$. The Hawking temperature corresponding to the solution is (see also for orthogonal case) found to be $$T_H=\frac{\overline{d}}{4\pi (2\mu )^{1/\overline{d}}}\underset{sS}{}H_{s0}^{h_s},$$ (3.42) where $`H_{s0}`$ are defined in (3.40). The boundary conditions (3.40) and (3.41) play a crucial role here, since they single out, generally speaking, only few solutions to eqs. (3.39). Moreover for some values of parameters $`\mu =\overline{\mu }/\overline{d}`$, $`\epsilon _s`$ and $`Q_s^2`$ the solutions to eqs. (3.39)-(3.41) do not exist. Indeed, from (2.27), (2.38), (3.12), (3.17), (3.24), (3.30) and (3.32 we get $$E_{TL}=\overline{\mu }^2\underset{s,s^{}S}{}h_sA^{ss^{}}=\frac{1}{4}\underset{s,s^{}S}{}h_sA_{ss^{}}\dot{q^s}(0)\dot{q^s^{}}(0)+\underset{sS}{}\frac{1}{2}\epsilon _sQ_s^2.$$ (3.43) Let the matrix $`(h_sA_{ss^{}})`$ be positive-definite (in this case matrix $`(B_{ss^{}})`$ is positive-definite too and all $`h_s>0`$). Then $`E_{TL}>0`$ and $$\overline{\mu }^2\underset{s,s^{}S}{}h_sA^{ss^{}}\underset{sS}{}\frac{1}{2}\epsilon _sQ_s^2.$$ (3.44) If the parameters obey the relation $$0<\overline{\mu }^2\underset{s,s^{}S}{}h_sA^{ss^{}}<\underset{sS}{}\frac{1}{2}\epsilon _sQ_s^2,$$ (3.45) e.g. for $`\epsilon _s=+1`$ and big enough $`Q_s^2`$, the solution under consideration does not exist. We note that the solution to eqs. (3.39)-(3.41) may not be unique. The simplest example occurs in the case of one $`p`$-brane, when $`h_s>0`$, $`\epsilon _s=+1`$ and $`\overline{\mu }^2h_s>Q_s^2`$. In this case we have two solutions to (3.39)-(3.41) corresponding to two possible values of $`\dot{q^s}(0)`$. Hypothesis. For positive-definite matrix $`(h_sA_{ss^{}})`$ and $`\epsilon _s=1`$, $`sS`$, the solution to (3.39)-(3.41) is uniquely defined. This hypothesis will be a subject of a future investigation. It implies a ”no-hair theorem” for black hole solutions under consideration. Thus, we obtained a family of black hole solutions up to solutions of radial equations (3.39) with the boundary conditions (3.40) and (3.41). In the next sections we consider several exact solutions to eqs. (3.39)-(3.41). Remark 3. Let $`M_i=\text{R}`$ and $`g^i=d\overline{t}d\overline{t}`$ for some $`i3`$. Then the metric (3.34) has no a horizon with respect to the “second time” $`\overline{t}`$ for $`R^{\overline{d}}2\mu `$. Thus, we a led to a “single-time” theorem from . Relation (3.22) from Proposition 1 coincides with the “no-hair” theorem from . ## 4 Polynomial structure of $`H_s`$ for Lie algebras ### 4.1 Conjecture on polynomial structure Now we deal with solutions to second order non-linear differential equations (3.39) that may be rewritten as follows $$\frac{d}{dz}\left(\frac{F}{H_s}\frac{d}{dz}H_s\right)=\overline{B}_s\underset{s^{}S}{}H_s^{}^{A_{ss^{}}},$$ (4.1) where $`H_s(z)>0`$, $`F=12\mu z`$, $`\mu >0`$, $`z=R^{\overline{d}}`$, $`\overline{B}_s=B_s/\overline{d}^20`$. Eqs. (3.41) and (3.40) read $`H_s((2\mu )^10)=H_{s0}(0,+\mathrm{}),`$ (4.2) $`H_s(+0)=1,`$ (4.3) $`sS`$. (Here we repeat equations (1.1)-(1.3)) for convenience.) It seems rather difficult to find the solutions to a set of eqs. (4.1)-(4.3) for arbitrary values of parameters $`\mu `$, $`\overline{B}_s`$, $`sS`$ and quasi-Cartan matrices $`A=(A_{ss^{}})`$. But we may expect a drastically simplification of the problem under consideration for certain class of parameters and/or $`A`$-matrices. In general we may try to seek solutions of (4.1) in a class of functions analytical in a disc $`|z|<L`$ and continuous in semi-interval $`0<z(2\mu )^1`$. For $`|z|<L`$ we get $$H_s(z)=1+\underset{k=1}{\overset{\mathrm{}}{}}P_s^{(k)}z^k,$$ (4.4) where $`P_s^{(k)}`$ are constants, $`sS`$. Substitution of (4.4) into (4.1) gives us an infinite chain of relations on parameters $`P_s^{(k)}`$ and $`\overline{B}_s`$. In general case it seems to be impossible to solve this chain of equations. Meanwhile there exist solutions to eqs. (4.1)-(4.3) of polynomial type. The simplest example occurs in orthogonal case , when $$(U^s,U^s^{})=B_{ss^{}}=0,$$ (4.5) for $`ss^{}`$, $`s,s^{}S`$. In this case $`(A_{ss^{}})=\mathrm{diag}(2,\mathrm{},2)`$ is a Cartan matrix for semisimple Lie algebra $`A_1\mathrm{}A_1`$ and $$H_s(z)=1+P_sz,$$ (4.6) with $`P_s0`$, satisfying $$P_s(P_s+2\mu )=\overline{B}_s,$$ (4.7) $`sS`$. In this solution was generalized to a block orthogonal case: $`S=S_1\mathrm{}S_k,S_iS_j=\mathrm{},ij,`$ (4.8) $`S_i\mathrm{}`$, i.e. the set $`S`$ is a union of $`k`$ non-intersecting (non-empty) subsets $`S_1,\mathrm{},S_k`$, and $`(U^s,U^s^{})=0`$ (4.9) for all $`sS_i`$, $`s^{}S_j`$, $`ij`$; $`i,j=1,\mathrm{},k`$. In this case (4.6) is modified as follows $$H_s(z)=(1+P_sz)^{b_0^s},$$ (4.10) where $`b_0^s`$ are defined in (3.24) and parameters $`P_s`$ are coinciding inside blocks, i.e. $`P_s=P_s^{}`$ for $`s,s^{}S_i`$, $`i=1,\mathrm{},k`$. Parameters $`P_s0`$ satisfy the relations $$P_s(P_s+2\mu )=\overline{B}_s/b_0^s,$$ $`b_0^s0`$, and parameters $`\overline{B}_s/b_0^s`$ are also coinciding inside blocks, i.e. $`\overline{B}_s/b_0^s=\overline{B}_s^{}/b_0^s^{}`$ for $`s,s^{}S_i`$, $`i=1,\mathrm{},k`$. In this case $`H_s`$ are analytical in $`|z|<L`$, where $`L=\mathrm{min}(|P_s|^1,sS`$). Let $`(A_{ss^{}})`$ be a Cartan matrix for a finite-dimensional semisimple Lie algebra $`𝒢`$. In this case all powers in (3.24) are natural numbers $$b_0^s=2\underset{s^{}S}{}A^{ss^{}}=n_s\text{N},$$ (4.11) and hence, all functions $`H_s`$ are polynomials, $`sS`$. Integers $`n_s`$ coincide with the components of twice the dual Weyl vector in the basis of simple coroots (see Sect. 13.7 in ). Conjecture. Let $`(A_{ss^{}})`$ be a Cartan matrix for a semisimple finite-dimensional Lie algebra $`𝒢`$. Then the solution to eqs. (4.1)-(4.3) (if exists) is a polynomial $$H_s(z)=1+\underset{k=1}{\overset{n_s}{}}P_s^{(k)}z^k,$$ (4.12) where $`P_s^{(k)}`$ are constants, $`k=1,\mathrm{},n_s`$, integers $`n_s=b_0^s`$ are defined in (4.11) and $`P_s^{(n_s)}0`$, $`sS`$. In extremal case ($`\mu =+0`$) an a analogue of this conjecture was suggested previously in . ### 4.2 Proof of Conjecture for $`A_m`$ and $`C_{m+1}`$ First, we prove the Conjecture for simple Lie algebras $`A_m=sl(m+1)`$, $`m1`$. Let us consider exact solutions to equations of motion of a Toda-chain corresponding to the Lie algebra $`A_m`$ , $$\ddot{q}^s=B_s\mathrm{exp}\left(\underset{s^{}=1}{\overset{m}{}}A_{ss^{}}q^s^{}\right),$$ (4.13) where $$\left(A_{ss^{}}\right)=\left(\begin{array}{cccccc}2& 1& 0& \mathrm{}& 0& 0\\ 1& 2& 1& \mathrm{}& 0& 0\\ 0& 1& 2& \mathrm{}& 0& 0\\ \multicolumn{6}{c}{}\\ 0& 0& 0& \mathrm{}& 2& 1\\ 0& 0& 0& \mathrm{}& 1& 2\end{array}\right)$$ (4.14) is the Cartan matrix of the Lie algebra $`A_m`$ and $`B_s>0`$, $`s,s^{}=1,\mathrm{},m`$. Here we put $`S=\{1,\mathrm{},m\}`$. The equations of motion (4.13) correspond to the Lagrangian $$L_T=\frac{1}{2}\underset{s,s^{}=1}{\overset{m}{}}A_{ss^{}}\dot{q}^s\dot{q}^s^{}\underset{s=1}{\overset{m}{}}B_s\mathrm{exp}\left(\underset{s^{}=1}{\overset{m}{}}A_{ss^{}}q^s^{}\right).$$ (4.15) This Lagrangian may be obtained from the standard one by separating a coordinate describing the motion of the center of mass. Using the result of A. Anderson we present the solution to eqs. (4.13) in the following form $$C_s\mathrm{exp}(q^s(u))=\underset{r_1<\mathrm{}<r_s}{\overset{m+1}{}}v_{r_1}\mathrm{}v_{r_s}\mathrm{\Delta }^2(w_{r_1},\mathrm{},w_{r_s})\mathrm{exp}[(w_{r_1}+\mathrm{}+w_{r_s})u],$$ (4.16) $`s=1,\mathrm{},m`$, where $$\mathrm{\Delta }(w_{r_1},\mathrm{},w_{r_s})=\underset{i<j}{\overset{s}{}}\left(w_{r_i}w_{r_j}\right);\mathrm{\Delta }(w_{r_1})1,$$ (4.17) denotes the Vandermonde determinant. The real constants $`v_r`$ and $`w_r`$, $`r=1,\mathrm{},m+1`$, obey the relations $$\underset{r=1}{\overset{m+1}{}}v_r=\mathrm{\Delta }^2(w_1,\mathrm{},w_{m+1}),\underset{r=1}{\overset{m+1}{}}w_r=0.$$ (4.18) In (4.16) $$C_s=\underset{s^{}=1}{\overset{m}{}}B_s^{}^{A^{ss^{}}},$$ (4.19) where $`A^{ss^{}}={\displaystyle \frac{1}{m+1}}\mathrm{min}(s,s^{})[m+1\mathrm{max}(s,s^{})],`$ (4.20) $`s,s^{}=1,\mathrm{},m`$, are components of a matrix inverse to the Cartan one, i.e. $`(A^{ss^{}})=(A_{ss^{}})^1`$ (see Sect. 7.5 in ). Here $$v_r0,w_rw_r^{},rr^{},$$ (4.21) $`r,r^{}=1,\mathrm{},m+1`$. We note that the solution with $`B_s>0`$ may be obtained from the solution with $`B_s=1`$ (see ) by a certain shift $`q^sq^s+\delta ^s`$. The energy reads $$E_T=\frac{1}{2}\underset{s,s^{}=1}{\overset{m}{}}A_{ss^{}}\dot{q}^s\dot{q}^s^{}+\underset{s=1}{\overset{m}{}}B_s\mathrm{exp}\left(\underset{s^{}=1}{\overset{m}{}}A_{ss^{}}q^s^{}\right)=\frac{1}{2}\underset{r=1}{\overset{m+1}{}}w_r^2.$$ (4.22) If $`B_s>0`$, $`sS`$, then all $`w_r,v_r`$ are real and, moreover, all $`v_r>0`$, $`r=1,\mathrm{},m+1`$. In a general case $`B_s0`$, $`sS`$, relations (4.16)-(4.19) also describe real solutions to eqs. (4.13) for suitably chosen complex parameters $`v_r`$ and $`w_r`$. These parameters are either real or belong to pairs of complex conjugate (non-equal) numbers, i.e., for example, $`w_1=\overline{w}_2`$, $`v_1=\overline{v}_2`$. When some of $`B_s`$ are negative, there are also some special (degenerate) solutions to eqs. (4.13) that are not described by relations (4.16)-(4.19), but may be obtained from the latter by certain limits of parameters $`w_r`$. For the energy (2.38) we get $$E_{TL}=\frac{1}{2K}E_T=\frac{h}{4}\underset{r=1}{\overset{m+1}{}}w_r^2.$$ (4.23) Here $$K_s=K,h_s=h=K^1,$$ (4.24) $`sS`$. Thus, in the $`A_m`$ Toda chain case eqs. (4.16)-(4.24) should be substituted into relations (2.25) and (2.37). Now we consider $`A_m`$-solutions with asymptotics (3.10). In this case all $`w_1,\mathrm{},w_{m+1}`$ are real and without loss of generality $`w_1<\mathrm{}<w_{m+1}`$. For $`b_0^s=n_s`$ from (3.24) we get $$n_s=b_0^s=s(ms+1),$$ (4.25) $`s=1,\mathrm{},m`$, or explicitly $$b_0^1=m,b_0^2=2(m1),\mathrm{},b_0^m=m.$$ (4.26) From (3.10), (3.17), $`\overline{\mu }=\sqrt{C_1}`$ and (4.16) we get ($`w_1<\mathrm{}<w_{m+1}`$) $`\overline{\mu }b_0^1=\overline{\mu }m=w_{m+1},`$ (4.27) $`\overline{\mu }b_0^2=2\overline{\mu }(m1)=w_m+w_{m+1},`$ (4.28) $`\mathrm{}`$ $`\overline{\mu }b_0^m=\overline{\mu }m=w_2+\mathrm{}+w_{m+1}.`$ (4.29) These relations imply $$w_{m+1}=\overline{\mu }m,w_m=\overline{\mu }(m2),\mathrm{},w_1=\overline{\mu }m,$$ (4.30) or, $$w_j=(2jm2)\overline{\mu },$$ (4.31) $`j=1,\mathrm{},m+1`$. From (4.16) and (4.30) we get $$f_s=e^{q^s}=\alpha _s^{(0)}e^{n_s\overline{\mu }u}+\alpha _s^{(1)}e^{(n_s2)\overline{\mu }u}+\mathrm{}+\alpha _s^{(n_s)}e^{n_s\overline{\mu }u},$$ (4.32) where $`\alpha _s^{(k)}`$ are constants, $`k=1,\mathrm{},n_s`$, $`\alpha _s^{(n_s)}0`$. Hence, due to (3.30), (3.33) we obtain the relations $$H_s=e^{q^sn_s\overline{\mu }u}=\alpha _s^{(0)}+\alpha _s^{(1)}F+\mathrm{}+\alpha _s^{(n_s)}F^{n_s},$$ (4.33) equivalent to (4.12) ($`\alpha _s^{(0)}+\alpha _s^{(1)}+\mathrm{}+\alpha _s^{(n_s)}=1)`$ with $`\alpha _s^{(n_s)}=P_s^{(n_s)}0`$, $`s=1,\mathrm{},m`$. Thus, the Conjecture is proved for the Lie algebras $`𝒢=A_m`$, $`m1`$. Now we prove the Conjecture for simple Lie algebras $`C_{m+1}=sp(m+1)`$, $`m1`$. (Remind that for $`m=1`$: $`C_2=B_2=so(5)`$). The Cartan matrix for the Lie algebra $`C_{m+1}`$ ($`m1`$) reads $$\left(A_{ss^{}}\right)=\left(\begin{array}{cccccc}2& 2& 0& \mathrm{}& 0& 0\\ 1& 2& 1& \mathrm{}& 0& 0\\ 0& 1& 2& \mathrm{}& 0& 0\\ \multicolumn{6}{c}{}\\ 0& 0& 0& \mathrm{}& 2& 1\\ 0& 0& 0& \mathrm{}& 1& 2\end{array}\right)$$ (4.34) $`s,s^{}=0,\mathrm{},m`$. The set of equations (4.1) with the Cartan matrix (4.34) and $`s=0,\mathrm{},m`$, may be embedded into a set of equations (4.13) corresponding to the Cartan matrix of the Lie algebra $`A_{2m+1}`$ (see (4.14)) with $`s=m,\mathrm{},0,\mathrm{},m`$, if the following identifications : $`\overline{B}_k=\overline{B}_k`$ and $`H_k=H_k`$, $`k=1,\mathrm{},m`$, are adopted. This proves the Conjecture for $`C_{m+1}`$, since it was proved for $`A_{2m+1}`$. ## 5 Some examples ### 5.1 Solution for $`A_2`$ Here we consider some examples of solutions related to the Lie algebra $`A_2=sl(3)`$. According to the results of previous section we seek the solutions to eqs. (4.1)-(4.3) in the following form (see (4.12); here $`n_1=n_2=2`$): $$H_s=1+P_sz+P_s^{(2)}z^2,$$ (5.1) where $`P_s=P_s^{(1)}`$ and $`P_s^{(2)}0`$ are constants, $`s=1,2`$. The substitution of (5.1) into equations (4.1) and decomposition in powers of $`z`$ lead us to the relations $`P_s(P_s+2\mu )+2P_s^{(2)}=\overline{B}_s,`$ (5.2) $`2P_s^{(2)}(P_s+4\mu )=P_{s+1}\overline{B}_s,`$ (5.3) $`2P_s^{(2)}(\mu P_s+P_s^{(2)})=P_{s+1}^{(2)}\overline{B}_s,`$ (5.4) corresponding to powers $`z^0,z^1,z^2`$ respectively, $`s=1,2`$. Here we denote $`s+1=2,1`$ for $`s=1,2`$ respectively. For $`P_1+P_2+4\mu 0`$ the solutions of (5.2)-(5.4) read $`P_s^{(2)}={\displaystyle \frac{P_sP_{s+1}(P_s+2\mu )}{2(P_1+P_2+4\mu )}},`$ (5.5) $`\overline{B}_s={\displaystyle \frac{P_s(P_s+2\mu )(P_s+4\mu )}{P_1+P_2+4\mu }},`$ (5.6) $`s=1,2`$. For $`P_1+P_2+4\mu =0`$ there exist also a special solution with $`P_1=P_2=2\mu ,2P_s^{(2)}=\overline{B}_s>0,\overline{B}_1+\overline{B}_2=4\mu ^2.`$ (5.7) Thus, in the $`A_2`$-case the solution is described by relations (3.34)-(3.38) with $`S=\{s_1,s_2\}`$, intersection rules (2.54), or, equivalently, $`d(I_{s_1}I_{s_2})={\displaystyle \frac{d(I_{s_1})d(I_{s_2})}{D2}}\chi _{s_1}\chi _{s_2}\lambda _{a_{s_1}}\lambda _{a_{s_2}}{\displaystyle \frac{1}{2}}K,`$ (5.8) $`d(I_{s_i}){\displaystyle \frac{(d(I_{s_i}))^2}{D2}}+\lambda _{a_{s_i}}\lambda _{a_{s_i}}=K,`$ (5.9) where $`K=K_{s_i}0`$, and functions $`H_{s_i}=H_i`$ are defined by relations (5.1) and (5.5)-(5.7) with $`z=R^{\overline{d}}`$, $`i=1,2`$. ### 5.2 $`A_2`$-dyon in $`D=11`$ supergravity Consider the “truncated” bosonic sector of $`D=11`$ supergravity (“truncated” means without Chern-Simons term). The action (2.1) in this case reads $`S_{tr}={\displaystyle _M}d^{11}z\sqrt{|g|}\left\{R[g]{\displaystyle \frac{1}{4!}}F^2\right\}.`$ (5.10) where $`\mathrm{rank}F=4`$. In this particular case, we consider a dyonic black-hole solutions with electric $`2`$-brane and magnetic $`5`$-brane defined on the manifold $$M=(2\mu ,+\mathrm{})\times (M_1=S^2)\times (M_2=\text{R})\times M_3\times M_4,$$ (5.11) where $`dimM_3=2`$ and $`dimM_4=5`$. The solution reads, $`g=H_1^{1/3}H_2^{2/3}\{{\displaystyle \frac{dRdR}{12\mu /R}}+R^2d\mathrm{\Omega }_2^2`$ (5.12) $`H_1^1H_2^1(1{\displaystyle \frac{2\mu }{R}})dtdt+H_1^1g^3+H_2^1g^4\},`$ $`F={\displaystyle \frac{Q_1}{R^2}}H_1^2H_2dRdt\tau _3+Q_2\tau _1\tau _3,`$ (5.13) where metrics $`g^2`$ and $`g^3`$ are Ricci-flat metrics of Euclidean signature, and $`H_s`$ are defined as follows $$H_s=1+\frac{P_s}{R}+\frac{P_s^{(2)}}{R^2},$$ (5.14) where parameters $`P_s`$, $`\mu >0`$ and $`P_s^{(2)}`$, $`\overline{B}_s=B_s=2Q_s^2`$, $`s=1,2`$, satisfy relations (5.5) and (5.6). The solution describes $`A_2`$-dyon consisting of electric $`2`$-brane with world sheet isomorphic to $`(M_2=\text{R})\times M_3`$ and magnetic $`5`$-brane with worldsheet isomorphic to $`(M_2=\text{R})\times M_4`$. The “branes” are intersecting on the time manifold $`M_2=\text{R}`$. Here $`K_s=(U^s,U^s)=2`$, $`\epsilon _s=1`$ for all $`sS`$. The $`A_2`$ intersection rule reads (see (2.54)) $$25=1$$ (5.15) Here and in what follows $`(p_1p_2=d)(d(I)=p_1+1,d(J)=p_2+1,d(IJ)=d)`$. The solution (5.12), (5.13) satisfies not only equations of motion for the truncated model, but also the equations of motion for $`D=11`$ supergravity with the bosonic sector action $`S=S_{tr}+c{\displaystyle _M}AFF`$ (5.16) ($`c=\mathrm{const}`$, $`F=dA`$), since the only modification related to “Maxwells” equations $`dF=\mathrm{const}FF,`$ (5.17) is trivial due to $`FF=0`$ (since $`\tau _i\tau _i=0`$). This solution in a special case $`H_1=H_2=H^2`$ ($`P_1=P_2`$, $`Q_1^2=Q_2^2`$) was considered in . The 4-dimensional section of the metric (5.12) in this special case coincides with the Reissner-Nordström metric. For the extremal case, $`\mu +0`$, and multi-black-hole generalization see also . ### 5.3 $`A_2`$-dyon in Kaluza-Klein model Let us consider $`4`$-dimensional model $$S=_Md^4z\sqrt{|g|}\left\{R[g]g^{\mu \nu }_\mu \phi _\nu \phi \frac{1}{2!}\mathrm{exp}[2\lambda \phi ]F^2\right\}$$ (5.18) with scalar field $`\phi `$, two-form $`F=dA`$ and $$\lambda =\sqrt{3/2}.$$ (5.19) This model originates after Kaluza-Klein reduction of $`5`$-dimensional gravity. The 5-dimensional metric in this case reads $$g^{(5)}=\varphi g_{\mu \nu }dx^\mu dx^\nu +\varphi ^2(dy+𝒜)(dy+𝒜),$$ (5.20) where $$𝒜=\sqrt{2}A=\sqrt{2}A_\mu dx^\mu ,\varphi =\mathrm{exp}(2\phi /\sqrt{6}).$$ (5.21) We consider the dyonic black-hole solution carrying electric charge $`Q_1`$ and magnetic charge $`Q_2`$, defined on the manifold $$M=(2\mu ,+\mathrm{})\times (M_1=S^2)\times (M_2=\text{R}).$$ (5.22) This solution reads $`g=\left(H_1H_2\right)^{1/2}\left\{{\displaystyle \frac{dRdR}{12\mu /R}}+R^2d\mathrm{\Omega }_2^2H_1^1H_2^1\left(1{\displaystyle \frac{2\mu }{R}}\right)dtdt\right\},`$ (5.23) $`\mathrm{exp}(\phi )=H_1^{\lambda /2}H_2^{\lambda /2},`$ (5.24) $`F=dA={\displaystyle \frac{Q_1}{R^2}}H_1^2H_2dRdt+Q_2\tau _1,`$ (5.25) where functions $`H_s`$ are defined by relations (5.1), (5.5) and (5.6) with $`\overline{B}_s=2Q_s^2`$, $`z=R^1`$, $`s=1,2`$; where $`\tau _1`$ is volume form on $`S^2`$. For 5-metric we obtain from (5.20)-(5.24) $`g^{(5)}=H_2\left\{{\displaystyle \frac{dRdR}{12\mu /R}}+R^2d\mathrm{\Omega }_2^2H_1^1H_2^1\left(1{\displaystyle \frac{2\mu }{R}}\right)dtdt\right\}`$ (5.26) $`+H_1H_2^1(dy+𝒜)(dy+𝒜),`$ $`d𝒜=\sqrt{2}F`$. For $`Q_20`$ we get the black hole version of Dobiash-Maison solution from and for $`Q_10`$ we are led to the black hole version of Gross-Perry-Sorkin monopole solution from , see . The solution coincides with Gibbons-Wiltshire dyon solution . Our notations are related to those from ref. , as following : $`H_1R^2=B`$, $`H_2R^2=A`$, $`R^22\mu R=\mathrm{\Delta }`$, $`Q_1=\sqrt{2}q`$, $`Q_2=\sqrt{2}p`$, $`R\mu =rm`$, $`\mu ^2=m^2+d^2p^2q^2`$, $`(P_2P_1)/2(P_2+1)=d/(d\sqrt{3}m)`$). (For general spherically symmetric configurations see also ref. .) We note that, quite recently, in the KK dyon solution was used for constructing the dyon solution in $`D=11`$ supergravity (5.12)-(5.13) for flat $`g^3`$ and $`g^4`$ and its rotating version. ## 6 Conclusions Thus here we obtained a family of black hole (BH) solutions with intersecting $`p`$-branes with nearly arbitrary intersection rules, see relations (2.22), (2.23) and Restriction 1. (Restriction 2 is satisfied, since all $`p`$-branes have a common time manifold.) These BH solutions are given by relations (3.34)-(3.41). The metric of solutions contains $`n1`$ Ricci-flat “internal” space metrics. The solutions are defined up to a set of (”moduli”) functions $`H_s`$ obeying a set of equations (3.39) with boundary conditions (3.40)-(3.41) (or, equivalently, eqs. (1.1)-(1.3)). These solutions are new and generalize a lot of special classes of BH solutions considered earlier in the literature. It is not necessary in future investigations to consider special models and setups, find spherically symmetric solutions and single out BH ones. All this program is fulfilled in this paper (with the use of results of ref. ). What we only need is to find explicit relations for moduli functions $`H_s`$, when matrix $`A`$ is fixed, i.e. to solve equations (1.1) with boundary conditions (1.2)-(1.3) imposed. The problem (1.1)-(1.3) seems to be rather difficult and may be of interest from the pure mathematical point of view, regardless to possible physical applications. Here we suggested a conjecture on polynomial structure of $`H_s`$ for intersections related to semisimple Lie algebras and proved it for $`A_m`$ and $`C_{m+1}`$ algebras, $`m1`$. This result may be interesting, since any appearance of polynomials in mathematical physics, especially related to Lie algebras, is always a rather attractive for mathematicians (and physicists, as well). Here we also obtained explicit relations for the solutions in the $`A_2`$-case and considered two examples of $`A_2`$-dyon solutions: one in $`D=11`$ supergravity (with $`M2`$\- and $`M5`$-branes intersecting at a point ) and another in $`5`$-dimensional Kaluza-Klein theory (Gibbons-Wiltshire solution). Explicit relations for $`H_s`$ corresponding to other examples of Lie algebras (e.g. $`A_3`$, $`B_2`$ etc) will be considered in future publications. Acknowledgments This work was supported in part by the Russian Ministry for Science and Technology, Russian Foundation for Basic Research, and project SEE.
warning/0002/astro-ph0002307.html
ar5iv
text
# 1 Introduction to Cosmic Magnetic Fields ## 1 Introduction to Cosmic Magnetic Fields A large number of spiral galaxies, including the Milky Way, carry magnetic fields . With few exceptions, the galactic field strengths are measured to be a few times $`10^6`$ G. This particular value has also been found at a redshift of $`z=0.395`$ and between the galaxies in clusters. Studies of the polarisation of synchrotron radiation emitted by galaxies with a face-on view, such as M51, reveal that their magnetic fields are aligned with spiral arms and density waves in the disk. A plausible explanation is that galactic magnetic fields were created by a mean-field dynamo mechanism , in which a much smaller seed field was exponentially amplified by the turbulent motion of ionised gas in conjunction with the differential rotation of the galaxy. For the dynamo to work, the initial seed field must be correlated on a scale of $`100`$ pc, corresponding to the largest turbulent eddy . The required strength of the seed field is subject to large uncertainties; past authors have quoted $`10^{21\pm 2}`$ G as the lower bound at the time of completed galaxy formation. This would present a problem for most particle-physics and field-theory inspired mechanisms of magnetic field generation. However, in recent work with A.-C. Davis and M.J. Lilley , I have shown that the lower bound on the dynamo seed field can be significantly relaxed if the universe is flat with a cosmological constant, as is suggested by recent supernovae observations . In particular, for the same dynamo parameters that give a lower bound of $`10^{20}`$ G for $`\mathrm{\Omega }_0=1`$, $`\mathrm{\Omega }_\mathrm{\Lambda }=0`$, we obtain $`10^{30}`$ G for $`\mathrm{\Omega }_0=0.2=1\mathrm{\Omega }_\mathrm{\Lambda }`$, implying that particle-physics mechanisms could still be viable. The observation at redshift $`z=0.395`$ can also be accounted for with these parameters, but requires a seed field of at least $`10^{23}`$ G . The dynamo amplifies the magnetic field until its energy reaches equipartition with the kinetic energy of the ionised gas, $`B^2/2=\rho v^2/2`$, when further growth is suppressed by dynamical back reaction. Thus a final field of $`B_010^6`$ G results for any seed field of sufficient strength. In order to explain the galactic field strength without a dynamo mechanism, one would require a strong primordial field of $`10^3(\mathrm{\Omega }_0h^2)^{1/3}`$ G at the epoch of radiation decoupling $`t_{\mathrm{dec}}`$, corresponding to a field strength $`10^9(\mathrm{\Omega }_0h^2)^{1/3}`$ G on comoving scales of $`1`$ Mpc. Future precision measurements of the CMB will put severe constraints on such a primordial field . Moreover, magnetic fields on Mpc scales have been probed by observations of the Faraday rotation of polarised light from distant luminous sources, which give an upper bound of about $`10^9`$ G . The observation of micro-Gauss fields between galaxies in clusters presents an interesting dilemma. Because such regions are considerably less dense than galaxies, it is doubtful whether a dynamo could have been operative. Thus the intra-cluster magnetic fields, unless somehow ejected from galaxies, have formed directly from a primordial field stronger than $`10^3(\mathrm{\Omega }_0h^2)^{1/3}`$ G at $`t_{\mathrm{dec}}`$. Such a field would certainly leave a signature in future CMB data . Particle-physics inspired models, which typically produce weak seed fields, lead to precise predictions and there have an advantage over astrophysical mechanisms, where the magnetic field strength is determined by complicated nonlinear dynamics, or solutions of general relativity with a magnetic field , where the field strength must be fixed by observations. With the possible exception of the last-mentioned model, there is no compelling scenario that produces a primordial field strong enough to eliminate the need for a dynamo. Seed fields for the dynamo can be astrophysical or primordial. In the former category there is the important possibility that a seed field may arise spontaneously due to non-parallel gradients of pressure and charge density during the collapse of a protogalaxy . For the rest of this talk, however, I shall assume that the seed field is primordial. ## 2 Primordial Seed Fields It is useful to distinguish between primordial seed fields that are produced with correlation length smaller than vs. larger than the horizon size. Subhorizon-scale seed fields typically arise in first-order phase transitions and from causal processes involving defects. For example, magnetic fields may be created on the surface of bubble walls due to local charge separation induced by baryon-number gradients. The magnetic fields are then amplified by plasma turbulence near the bubble wall. This possibility has been explored for the QCD as well as for the electroweak phase transition. The production of magnetic fields in collisions of expanding true-vacuum bubbles will be discussed in Sec. 4. Fields can also be generated in the wakes of, or due to the wiggles of, GUT-scale cosmic strings during structure formation, resulting in a large correlation length . Joyce and Shaposhnikov have shown that an asymmetry of right-handed electrons, possibly generated at the GUT scale, would become unstable to the generation of a hypercharge magnetic field shortly before the electroweak phase transition , leading to a correlation length of order $`10^6/T`$. Horizon-scale seed fields emerge naturally in second-order phase transitions of gauge theories from the failure of covariant derivatives of the Higgs field to correlate on superhorizon scales . Superhorizon-scale seed fields can arise as a solution of the Einstein equations for axisymmetric universes and in inflationary or pre-Big Bang (superstring) scenarios. In the latter case, vacuum fluctuations of the field tensor are amplified by the dynamical dilaton field . Inflationary models produce extremely weak magnetic fields unless conformal invariance is explicitly broken , but even then great difficulties remain. An exciting new possibility is that magnetic fields may be produced via parametric resonance with an oscillating field e.g. during preheating after inflation. Because the inflaton is initially coherent on superhorizon scales, large correlations can arise without violating causality. A similar proposal involves charged scalar particles, minimally coupled to gravity, that are created from the vacuum due to the changing space-time geometry at the end of inflation. The particles give rise to fluctuating electric currents which are claimed to produce superhorizon-scale (indeed, galactic-scale) fields of sufficient strength to satisfy the galactic dynamo bound . This mechanism deserves further investigation. ## 3 Evolution of Primordial Magnetic Fields A serious problem with many particle-physics and field-theory scenarios for producing primordial magnetic fields is that the resulting correlation length $`\xi `$ is very small. If the fields are produced at the QCD phase transition or earlier with sub-horizon correlation length, then the expansion of the universe cannot stretch $`\xi `$ to more than 1 pc today (see Fig. 1). This is far short of the galactic dynamo lower bound of 5-10 kpc (comoving), corresponding to 100 pc in a virialised galaxy . Nevertheless, many authors have argued that the correlation length will grow more rapidly due to magnetohydrodynamic (MHD) turbulence and inverse cascade, which transfers power from small-scale to large-scale Fourier modes. Several non-relativistic models for this evolution are analysed in Fig. 1. The most conservative estimate is obtained by assuming that the magnetic field strength on the scale of one correlation length at any time equals the volume average of fields that were produced on smaller scales but have since decayed . This leads to a growth $`\xi t^{7/10}`$ (obtained from the Minkowski-space growth $`\xi t^{2/5}`$ via the substitution $`t\tau =t^{1/2}`$ and multiplication by the scale factor). The most optimistic estimate corresponds to the case when the magnetic field has maximal helicity in relation to the energy density . As magnetic helicity is approximately conserved in the high-conductivity early-universe environment, one obtains the growth law $`\xi t^{5/6}`$. Turbulence ends, freezing the growth (in comoving coordinates) when the kinetic Reynolds number drops below unity at the $`e^+e^{}`$ annihilation or later, depending on the model and the parameters of the initial field. An intermediate and rather plausible estimate has been given by Dimopoulos and Davis , who use the fact that the magnetic flux enclosed by a (sufficiently large) comoving closed curve is conserved. The correlation length here increases at a rate given by the Alfvén velocity, so that $`\xi t^{3/4}`$. As Fig. 1 shows, only the most optimistic of these growth laws leads to a correlation length today that satisfies the galactic dynamo bound. This occurs for fields correlated over the horizon scale at the QCD phase transition. Beware, however, that the growth laws were derived using non-relativistic MHD equations assuming that the magnetic field energy density remains in equipartition with the kinetic energy density $`\rho \overline{v}^2/2`$, where $`\overline{v}`$ is the presumed non-relativistic “bulk velocity” of the ultra-relativistic plasma. It seems plausible that a relativistic treatment could alter the predicted evolution dramatically. In this light, I find it too early to reject the idea that also subhorizon fields might evolve into fields sufficiently correlated to seed the galactic dynamo. At the same time, Fig. 1 demonstrates the intrinsic advantage of superhorizon field generation mechanisms. For these, the principal problem is not the correlation length, but to achieve sufficient strength of the magnetic field. ## 4 Magnetic Fields From Bubble Collisions First-order phase transitions in the early universe proceed through the nucleation of bubbles , which subsequently expand and collide. In order to study the generation of magnetic fields, the initial field strength is assumed to vanish. One may then choose a gauge in which the vector potential $`V_\mu `$ is initially zero. In this gauge, the nucleation probability is peaked around bubbles with constant orientation (phase) of the Higgs field. We consider first a U(1) toy model. Let the Higgs field in two colliding bubbles be given by $`\varphi _1=\rho _1(x)e^{i\theta _1}`$ and $`\varphi _2=\rho _2(x)e^{i\theta _2}`$, respectively, where $`\theta _1\theta _2`$. When the bubbles meet, the phase gradient establishes a gauge-invariant current $`j_k=iq[\varphi ^{}D_k\varphi (D_k\varphi )^{}\varphi ]`$ across the surface of intersection of the two bubbles, where $`D_k=_kiqV_k`$. This current, in turn, gives rise to a ring-like flux of the field strength $`F_{ij}=_iV_j_jV_i`$, which takes the shape of a girdle encircling the bubble intersection region. In recent work we have obtained accurate, but rather complicated, analytical solutions for the field evolution in U(1) bubble collisions using an analytical expression for the bubble-wall “bounce” profile . A simpler analytical solution was found by Kibble and Vilenkin \[KV\] , who made three rather crude approximations: (1) The bubble walls move through the plasma without friction, (2) the modulus of the Higgs field in the interior of the bubbles equals a constant, and (3) the phase $`\theta `$ of the Higgs field is a step function at the moment of collision. The first of these assumptions leads to a simple equation of motion for the bubble wall, which endows the system with a dynamical O(1,2) symmetry . All quantities are then functions only of two coordinates: $`z`$, the position along the axis through the bubble centres, and $`\tau =\sqrt{t^2x^2y^2}`$, combining time with the perpendicular directions. The solutions can be written down explicitly in terms of Bessel functions and, despite the crudeness of the approximations, capture correctly the qualitative behaviour of the fields in the bubble overlap region . The simplicity of the KV approach makes it ideal for attacking the more complicated problem of non-Abelian bubble collisions. These could occur in a first-order electroweak phase transition (e.g. in the MSSM for $`M_h\stackrel{<}{}116`$ GeV) or in a GUT phase transition. Field strengths created in an early phase transition naturally project onto the electromagnetic U(1) subgroup at the electroweak transition. I will here concentrate on SU(2)$`\times `$U(1) $``$U(1)<sub>EM</sub> and the minimal Standard Model as a solvable example. The initial Higgs field in the two bubbles can be written in the form $$\mathrm{\Phi }_1=\mathrm{exp}(i\theta _0𝒏𝝈)\left(\begin{array}{c}0\\ \rho _1(x)\end{array}\right),\mathrm{\Phi }_2=\mathrm{exp}(i\theta _0𝒏𝝈)\left(\begin{array}{c}0\\ \rho _2(x)\end{array}\right).$$ (1) As the bubbles collide, non-Abelian currents $`j_k^A=i[\mathrm{\Phi }^{}T^AD_k\mathrm{\Phi }(D_k\mathrm{\Phi })^{}T^A\mathrm{\Phi }]`$ develop across the surface of their intersection, where $`T^A=(g^{}/2,g\sigma ^a/2)`$, $`D_k=_kiW_k^AT^A`$ and $`W_k^A=(Y_k,W_k^a)`$. In analogy with the U(1) case, one obtains a ring-like flux of non-Abelian fields. The recipe for projecting out the electromagnetic field amongst the non-Abelian fields in an arbitrary gauge has been given elsewhere . In the special cases $`𝒏=(0,0,\pm 1)`$ and $`𝒏=(n_1,n_2,0)`$ it is known that the non-Abelian flux consists of pure Z and W vector fields, respectively. The absence of an electromagnetic field has its explanation in the fact that the normalised Higgs field $`\widehat{\mathrm{\Phi }}\mathrm{\Phi }/(\mathrm{\Phi }^{}\mathrm{\Phi })^{1/2}`$ maps to a geodesic on the Higgs vacuum manifold and the gauge fields map to a line in the Lie algebra spanned by the generator of that same geodesic: $$\widehat{\mathrm{\Phi }}(x)=\mathrm{exp}(i\theta (x)𝒏𝝈)(01)^{},𝐀\backslash _\mu (x)W_\mu ^A(x)T^A=f_\mu (x)𝒏𝝈.$$ (2) When $`n_30,\pm 1`$, both $`Z`$ and $`W`$ fields are excited. Because they have unequal masses $`M_WM_Z`$, they evolve differently and the fields $`\widehat{\mathrm{\Phi }}`$ and $`𝐀\backslash _\mu `$ stray from the geodesic and its tangent, producing an electromagnetic current. I have used a KV approach to derive a perturbative analytical solution for the evolution of gauge fields in an electroweak bubble collision, valid as long as the fields are small and higher-order nonlinearities can be neglected. The space allotted here allows me only to indicate the structure of the solution for the electromagnetic field, which is $`F^{\alpha z}`$ $`=`$ $`x^\alpha {\displaystyle \frac{\mathrm{sin}\theta _\mathrm{w}}{g}}n_3(1n_3^2){\displaystyle _0^z}dz^{}(\text{}h_3(\tau ,z^{})`$ (3) $`+`$ $`{\displaystyle _{t_\mathrm{c}}^\tau }d\tau ^{\prime \prime }{\displaystyle _{z^{}\tau ^{\prime \prime }+t_\mathrm{c}}^{z^{}+\tau ^{\prime \prime }t_\mathrm{c}}}dz^{\prime \prime }h_1(\sqrt{(\tau \tau ^{\prime \prime })^2(z^{}z^{\prime \prime })^2})h_2(\tau ^{\prime \prime },z^{\prime \prime })\text{}),`$ where $`t_\mathrm{c}`$ is the time of collision and the functions $`h_i`$ contain products of $`i`$ Bessel functions. As expected, the resulting field strength is of the order of $`M_W^2/g`$ with a correlation length $`\xi M_W^1`$. However, when plasma friction and conductivity are taken into account, the magnetic field spreads over the interior of a bubble leading to an appreciable increase in correlation length. ### Acknowledgments The author is supported by the European Commission’s TMR programme under Contract No. ERBFMBI-CT97-2697.
warning/0002/hep-th0002013.html
ar5iv
text
# References Effective action in general chiral superfield model A.Yu.Petrov Department of Theoretical Physics, Tomsk State Pedagogical University Tomsk 634041, Russia According to the superstring theory the low-energy elementary particle models contain as ingredient the multiplets of chiral and antichiral superfields action of which is given in terms of kählerian effective potential $`K(\overline{\mathrm{\Phi }},\mathrm{\Phi })`$ and chiral $`W(\mathrm{\Phi })`$ and antichiral $`\overline{W}(\overline{\mathrm{\Phi }})`$ potentials. These potentials are found in explicit and closed form within string perturbation theory (see f.e. ). Phenomenological aspects of such models have been studied in recent papers . In quantum theory one can expect an appearance of quantum corrections to the potentials $`K(\overline{\mathrm{\Phi }},\mathrm{\Phi })`$ and $`W(\mathrm{\Phi })`$. As a result we face a problem of calculating effective action in models with arbitrary functions $`K(\overline{\mathrm{\Phi }},\mathrm{\Phi })`$ and $`W(\mathrm{\Phi })`$. The remarkable features of the massless theories with $`N=1`$ chiral superfields are the possibilities of obtaining the chiral quantum corrections. A few years ago West pointed out that finite two-loop chiral contribution to effective action really arises in massless Wess-Zumino model (see also ). In this talk we consider the general problem of calculating leading quantum correction to chiral potential and kählerian potential in theory with arbitrary potentials $`K(\overline{\mathrm{\Phi }},\mathrm{\Phi })`$ and $`W(\mathrm{\Phi })`$, $`\overline{W}(\overline{\mathrm{\Phi }})`$. The remarkable result we obtain here is that despite the theory under consideration is non-renormalizable at arbitrary $`K(\overline{\mathrm{\Phi }},\mathrm{\Phi })`$, $`W(\mathrm{\Phi })`$, the lower (two-loop) chiral correction to effective action is always finite. We consider $`N=1`$ supersymmetric field theory with action $$S[\overline{\mathrm{\Phi }},\mathrm{\Phi }]=d^8zK(\overline{\mathrm{\Phi }},\mathrm{\Phi })+(d^6zW(\mathrm{\Phi })+h.c.)$$ (1) where $`\mathrm{\Phi }(z)`$ and $`\overline{\mathrm{\Phi }}(z)`$ are chiral and antichiral superfields respectively. As well known, the real function $`K(\overline{\mathrm{\Phi }},\mathrm{\Phi })`$ is called kählerian potential and holomorphic function $`W(\mathrm{\Phi })`$ is called chiral potential . The partial cases of the theory (1) are Wess-Zumino model with $`K(\overline{\mathrm{\Phi }},\mathrm{\Phi })=\mathrm{\Phi }\overline{\mathrm{\Phi }}`$, $`W(\mathrm{\Phi })\mathrm{\Phi }^3`$ and $`N=1`$ supersymmetric four-dimensional sigma-model with $`W(\mathrm{\Phi })=0`$. The action (1) is a most general one constructed from chiral and antichiral superfields which does not contain the higher derivatives at component level. Therefore it is natural to call the theory (1) a general chiral superfield model. Let $`\mathrm{\Gamma }[\overline{\mathrm{\Phi }},\mathrm{\Phi }]`$ be effective action in the model (1). Within a momentum expansion the effective action can be presented as a series in supercovariant derivatives $`D_A=(_a,D_\alpha ,\overline{D}_{\dot{\alpha }})`$ in the form $`\mathrm{\Gamma }[\overline{\mathrm{\Phi }},\mathrm{\Phi }]`$ $`=`$ $`{\displaystyle d^8z_{eff}(\mathrm{\Phi },D_A\mathrm{\Phi },D_AD_B\mathrm{\Phi };\overline{\mathrm{\Phi }},D_A\overline{\mathrm{\Phi }},D_AD_B\overline{\mathrm{\Phi }})}+`$ (2) $`+`$ $`({\displaystyle }d^6z_{eff}^{(c)}(\mathrm{\Phi })+h.c.)+\mathrm{}`$ Here $`_{eff}`$ is called general effective lagrangian, $`_{eff}^{(c)}`$ is called chiral effective lagrangian. Both these lagrangians are the series in supercovariant derivatives of superfields and can be written as follows $`_{eff}`$ $`=`$ $`K_{eff}(\overline{\mathrm{\Phi }},\mathrm{\Phi })+\mathrm{}=K(\overline{\mathrm{\Phi }},\mathrm{\Phi })+{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}K_{eff}^{(n)}(\overline{\mathrm{\Phi }},\mathrm{\Phi })`$ $`_{eff}^{(c)}`$ $`=`$ $`W_{eff}(\mathrm{\Phi })+\mathrm{}=W(\mathrm{\Phi })+{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}W_{eff}^{(n)}(\mathrm{\Phi })+\mathrm{}`$ (3) Here dots mean terms depending on covariant derivatives of superfields $`\mathrm{\Phi },\overline{\mathrm{\Phi }}`$. Here $`K_{eff}(\overline{\mathrm{\Phi }},\mathrm{\Phi })`$ is called kahlerian effective potential, $`W_{eff}(\mathrm{\Phi })`$ is called chiral (or holomorphic) effective potential, $`K_{eff}^{(n)}`$ is a $`n`$-th correction to kahlerian potential and $`W_{eff}^{(n)}`$ is a $`n`$-th correcton to chiral (holomorphic) potential $`W`$. To consider the effective lagrangians $`_{eff}`$ and $`_{eff}^{(c)}`$ we use path integral representation of the effective action $`\mathrm{exp}({\displaystyle \frac{i}{\mathrm{}}}\mathrm{\Gamma }[\overline{\mathrm{\Phi }},\mathrm{\Phi }])`$ $`=`$ $`{\displaystyle }𝒟\varphi 𝒟\overline{\varphi }\mathrm{exp}({\displaystyle \frac{i}{\mathrm{}}}S[\overline{\mathrm{\Phi }}+\sqrt{\mathrm{}}\overline{\varphi },\mathrm{\Phi }+\sqrt{\mathrm{}}\varphi ]`$ (4) $``$ $`({\displaystyle }d^6z{\displaystyle \frac{\delta \mathrm{\Gamma }[\overline{\mathrm{\Phi }},\mathrm{\Phi }]}{\delta \mathrm{\Phi }(z)}}\varphi (z)+h.c.))`$ Here $`\mathrm{\Phi },\overline{\mathrm{\Phi }}`$ are the background superfields and $`\varphi ,\overline{\varphi }`$ are the quantum ones. The effective action can be written as $`\mathrm{\Gamma }[\overline{\mathrm{\Phi }},\mathrm{\Phi }]=S[\overline{\mathrm{\Phi }},\mathrm{\Phi }]+\stackrel{~}{\mathrm{\Gamma }}[\overline{\mathrm{\Phi }},\mathrm{\Phi }]`$, where $`\stackrel{~}{\mathrm{\Gamma }}[\overline{\mathrm{\Phi }},\mathrm{\Phi }]`$ is a quantum correction. Eq. (4) allows to obtain $`\stackrel{~}{\mathrm{\Gamma }}[\overline{\mathrm{\Phi }},\mathrm{\Phi }]`$ in form of loop expansion $`\stackrel{~}{\mathrm{\Gamma }}[\overline{\mathrm{\Phi }},\mathrm{\Phi }]=_{n=1}^{\mathrm{}}\mathrm{}^n\mathrm{\Gamma }^{(n)}[\overline{\mathrm{\Phi }},\mathrm{\Phi }]`$ and hence, to get loop expansion for the effective lagrangians $`_{eff}`$ and $`_{eff}^{(c)}`$. To find loop corrections $`\mathrm{\Gamma }^{(n)}[\overline{\mathrm{\Phi }},\mathrm{\Phi }]`$ in explicit form we expand right-hand side of eq. (4) in power series in quantum superfields $`\varphi `$, $`\overline{\varphi }`$. As usual, the quadratic part of expansion of $`\frac{1}{\mathrm{}}S[\overline{\mathrm{\Phi }}+\sqrt{\mathrm{}}\overline{\varphi },\mathrm{\Phi }+\sqrt{\mathrm{}}\varphi ]`$ $`S_2={\displaystyle \frac{1}{2}}{\displaystyle }d^8z\left(\begin{array}{cc}\varphi & \overline{\varphi }\end{array}\right)\left(\begin{array}{cc}K_{\mathrm{\Phi }\mathrm{\Phi }}& K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}\\ K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}& K_{\overline{\mathrm{\Phi }}\overline{\mathrm{\Phi }}}\end{array}\right)\left(\begin{array}{c}\varphi \\ \overline{\varphi }\end{array}\right)+[{\displaystyle }d^6z{\displaystyle \frac{1}{2}}W^{^{\prime \prime }}\varphi ^2+h.c.]`$ (10) defines the propagators and the higher terms of expansion define the vertices. Here $`K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}=\frac{^2K(\overline{\mathrm{\Phi }},\mathrm{\Phi })}{\mathrm{\Phi }\overline{\mathrm{\Phi }}}`$, $`K_{\mathrm{\Phi }\mathrm{\Phi }}=\frac{^2K(\overline{\mathrm{\Phi }},\mathrm{\Phi })}{\mathrm{\Phi }^2}`$ etc, $`W^{^{\prime \prime }}=\frac{d^2W}{d\mathrm{\Phi }^2}`$. Let us consider holomorphic effective potential $`W_{eff}(\mathrm{\Phi })`$. It was noted by West that non-renormalization theorem (see f.e. ) does not forbid an existence of the finite chiral corrections to the effective action. The matter is the theories with chiral superfields admit the loop corrections of the form $`d^8zf(\mathrm{\Phi })\left(\frac{D^2}{4\mathrm{}}\right)g(\mathrm{\Phi })=d^6zf(\mathrm{\Phi })g(\mathrm{\Phi })`$ where $`f(\mathrm{\Phi }),g(\mathrm{\Phi })`$ are some functions of chiral superfield $`\mathrm{\Phi }`$. We note that this equation shows the superfield $`\mathrm{\Phi }`$ is not a constant. It is easy to prove that the chiral contributions to effective action can be generated by supergraphs containing massless propagators only. To find chiral corrections to effective action we put $`\overline{\mathrm{\Phi }}=0`$ in eqs. (4,10). Therefore here and further all derivatives of $`K`$, $`W`$ and $`\overline{W}`$ will be taken at $`\overline{\mathrm{\Phi }}=0`$. Under this condition the action of quantum superfields $`\varphi ,\overline{\varphi }`$ in external superfield $`\mathrm{\Phi }`$ looks like $`S[\overline{\varphi },\varphi ,\mathrm{\Phi }]={\displaystyle \frac{1}{2}}{\displaystyle d^8z\left(\begin{array}{cc}\varphi & \overline{\varphi }\end{array}\right)\left(\begin{array}{cc}K_{\mathrm{\Phi }\mathrm{\Phi }}& K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}\\ K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}& K_{\overline{\mathrm{\Phi }}\overline{\mathrm{\Phi }}}\end{array}\right)\left(\begin{array}{c}\varphi \\ \overline{\varphi }\end{array}\right)}+{\displaystyle d^6z\frac{1}{2}W^{^{\prime \prime }}\varphi ^2}+\mathrm{}`$ (16) The dots here denote the terms of third, fourth and higher orders in quantum superfields. We call the theory massless if $`W^{^{\prime \prime }}|_{\mathrm{\Phi }=0}=0`$. Further we consider only massless theory. To calculate the corrections to $`W(\mathrm{\Phi })`$ we use supergraph technique (see f.e. ). For this purpose one splits the action (16) into sum of free part and vertices of interaction. As a free part we take the action $`S_0=d^8z\varphi \overline{\varphi }`$. The corresponding superpropagator is $`G(z_1,z_2)=\frac{D_1^2\overline{D}_2^2}{16\mathrm{}}\delta ^8(z_1z_2)`$. And the term $`S[\overline{\varphi },\varphi ,\mathrm{\Phi }]S_0`$ will be treated as vertices where $`S[\overline{\varphi },\varphi ,\mathrm{\Phi }]`$ is given by eq. (16). Our purpose is to find first leading contribution to $`W_{eff}(\mathrm{\Phi })`$. As we will show, chiral loop contributions are began with two loops. Therefore we keep in eq. (16) only the terms of second, third and fourth orders in quantum fields. Non-trivial corrections to chiral potential can arise only if $`2L+1n_{W^{^{\prime \prime }}}n_{V_c}=0`$ where $`L`$ is a number of loops, $`n_{W^{^{\prime \prime }}}`$ is a number of vertices proportional to $`W^{^{\prime \prime }}`$, $`n_{V_c}`$ is that one of vertices of third and higher orders in quantum superfields, otherwise corresponding contribution will either vanish or lead to singularity in infrared linit. In one-loop approximation this equation leads to $`n_{W^{^{\prime \prime }}}+n_{V_c}=3`$. However, all supergraphs satisfying this condition have zero contribution. Therefore first correction to chiral effective potential is two-loop one. In two-loop approximation this equation has the form $`n_{W^{^{\prime \prime }}}+n_{V_c}=5`$. Since the number of purely chiral (antichiral) vertices independent of $`W^{^{\prime \prime }}`$ in two-loop supergraphs can be equal to 0, 1 or 2, number of external vertices $`W^{^{\prime \prime }}`$ takes values from 3 to 5. We note that non-trivial contribution to holomorphic effective potential from any diagram can arise only if number of $`D^2`$-factors is more by one than the number of $`\overline{D}^2`$-factors (see details in ). The only Green function in the theory is a propagator $`<\varphi \overline{\varphi }>`$. Therefore total number of quantum chiral superfields $`\varphi `$ corresponding to all vertices must be equal to that one of antichiral ones $`\overline{\varphi }`$. As a result we find that the only two-loop supergraph contributing to chiral effective potential looks like Here internal line is a propagator $`<\varphi \overline{\varphi }>`$ depending on background chiral superfields which has the form $$<\varphi \overline{\varphi }>=\overline{D}_1^2D_2^2\frac{\delta ^8(z_1z_2)}{16K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}(z_1)\mathrm{}}.$$ (17) We note that the superfield $`K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}`$ is not constant here. Double external lines are $`W^{\prime \prime }`$. After $`D`$-algebra transformations and loop integrations we find that two-loop contribution to holomorphic effective potential in this model looks like $$W^{(2)}=\frac{6}{(16\pi ^2)^2}\zeta (3)\overline{W}^{{}_{}{}^{\prime \prime \prime }2}\left\{\frac{W^{\prime \prime }(z)}{K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}^2(z)}\right\}^3$$ (18) One reminds that $`\overline{W}^{^{\prime \prime \prime }}=\overline{W}^{^{\prime \prime \prime }}(\overline{\mathrm{\Phi }})|_{\overline{\mathrm{\Phi }}=0}`$ and $`K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}(z)=\frac{^2K(\overline{\mathrm{\Phi }},\mathrm{\Phi })}{\mathrm{\Phi }\overline{\mathrm{\Phi }}}|_{\overline{\mathrm{\Phi }}=0}`$ here. We see that the correction (18) is finite and does not require renormalization. Now we turn to studying of quantum contributions to kählerian effective potential depending only on superfields $`\mathrm{\Phi }`$, $`\overline{\mathrm{\Phi }}`$ but not on their derivatives. The one-loop diagrams contributing to kahlerian effective potential are Double external lines correspond to alternating $`W^{\prime \prime }`$ and $`\overline{W}^{\prime \prime }`$. Internal lines are $`<\varphi \overline{\varphi }>`$-propagators of the form $$G_0<\varphi \overline{\varphi }>=\frac{\overline{D}^2D^2}{16K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}\mathrm{}}$$ Supergraph of such structure with $`2n`$ legs represents itself as a ring containing of $`n`$ links of the following form The total contribution of all these diagrams after $`D`$-algebra transformations, summarizing, integration over momenta and subtraction of divergences is equal to $$K^{(1)}=d^4\theta \frac{1}{32\pi ^2}\frac{W^{\prime \prime }\overline{W}^{\prime \prime }}{K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}^2}\mathrm{ln}\left(\frac{W^{\prime \prime }\overline{W}^{\prime \prime }}{\mu ^2K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}^2}\right)$$ (19) We can find also two-loop correction to kählerian potential. Since kählerian effective potential depends only on superfields $`\mathrm{\Phi },\overline{\mathrm{\Phi }}`$ but not on their derivatives supergraphs contributing to it must include equal number of $`D^2`$ and $`\overline{D}^2`$ factors with all vertices rewritten in the form of an integral over whole superspace. Components of matrix superpropagator for the case of constant superfields looks like $`G_{++}`$ $`=`$ $`{\displaystyle \frac{\overline{W}^{^{\prime \prime }}}{K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}\mathrm{}+|W^{^{\prime \prime }}|^2}}{\displaystyle \frac{\overline{D}_1^2}{4}}\delta _{12};G_+={\displaystyle \frac{1}{K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}\mathrm{}+|W^{^{\prime \prime }}|^2}}{\displaystyle \frac{\overline{D}_1^2D_2^2}{16}}\delta _{12}`$ $`G_+`$ $`=`$ $`{\displaystyle \frac{1}{K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}\mathrm{}+|W^{^{\prime \prime }}|^2}}{\displaystyle \frac{D_1^2\overline{D}_2^2}{16}}\delta _{12};G_{}={\displaystyle \frac{W^{^{\prime \prime }}}{K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}\mathrm{}+|W^{^{\prime \prime }}|^2}}{\displaystyle \frac{D_1^2}{4}}\delta _{12}`$ (20) The all diagrams with equal number of $`D^2`$ and $`\overline{D}^2`$-factors are given in this picture. Here Two-loop correction to kählerian potential is a sum of all these contributions. After $`D`$-algebra transformations, loop integrations and subtraction of one-loop and two-loop divergences it is equal to $`K^{(2)}`$ $`=`$ $`{\displaystyle \frac{1}{(16\pi ^2)^2}}({\displaystyle \frac{|W^{\prime \prime }|^2}{K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}^{}{}_{}{}^{4}}}({\displaystyle \frac{1}{4}}\mathrm{ln}^2{\displaystyle \frac{|W^{\prime \prime }|^2}{\mu ^2K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}}}+{\displaystyle \frac{3\gamma }{2}}\mathrm{ln}{\displaystyle \frac{|W^{\prime \prime }|^2}{\mu ^2K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}}}+{\displaystyle \frac{3}{2}}(\gamma 1)+`$ $`+`$ $`{\displaystyle \frac{1}{4}}(\gamma ^2+\zeta (2)))\left)\right\{{\displaystyle \frac{1}{6}}(K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}^{}{}_{}{}^{3}|W^{\prime \prime \prime }|^2+(\overline{W}^{\prime \prime }W^{\prime \prime \prime }K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}^{}{}_{}{}^{2}K_{\mathrm{\Phi }\mathrm{\Phi }\overline{\mathrm{\Phi }}}+h.c.)+`$ $`+`$ $`W^{\prime \prime }\overline{W^{\prime \prime }}K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}K_{\mathrm{\Phi }\mathrm{\Phi }\overline{\mathrm{\Phi }}}K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}\overline{\mathrm{\Phi }}}+`$ $`+`$ $`K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}^{}{}_{}{}^{2}|W^{\prime \prime }|^2|K_{\mathrm{\Phi }\mathrm{\Phi }\overline{\mathrm{\Phi }}}|^2)+{\displaystyle \frac{1}{2}}K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}^{}{}_{}{}^{2}|W^{\prime \prime }|^2K_{\mathrm{\Phi }\mathrm{\Phi }\overline{\mathrm{\Phi }}}K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}\overline{\mathrm{\Phi }}}\}+{\displaystyle \frac{1}{(16\pi ^2)^2}}\times `$ $`\times `$ $`{\displaystyle \frac{|W^{\prime \prime }|^4}{K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}^{}{}_{}{}^{4}}}(\gamma 1+\mathrm{ln}{\displaystyle \frac{|W^{\prime \prime }|^2}{\mu ^2K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}}})^2\left(K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}^{}{}_{}{}^{2}K_{\mathrm{\Phi }\mathrm{\Phi }\overline{\mathrm{\Phi }}}K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}\overline{\mathrm{\Phi }}}+K_{\mathrm{\Phi }\overline{\mathrm{\Phi }}}^{}{}_{}{}^{2}K_{\mathrm{\Phi }\mathrm{\Phi }\overline{\mathrm{\Phi }}\overline{\mathrm{\Phi }}}\right)`$ The value of normalization point $`\mu `$ can be fixed by imposing of suitable normalization condition. To conclude, we have solved the problem of calculating leading holomorphic correction to superfield effective action in general chiral superfield model (1) with arbitrary potentials $`K(\overline{\mathrm{\Phi }},\mathrm{\Phi })`$ and $`W(\mathrm{\Phi })`$. The result has the universal form (18) and it is finite independently if the functions $`K(\overline{\mathrm{\Phi }},\mathrm{\Phi })`$, $`W(\mathrm{\Phi })`$ correspond to renormalizable theory or no. We have also calculated one-loop and two-loop contributions to kählerian effective potential. We note that results (18,19,S0.Ex9) reproduces the known results for Wess-Zumino model at $`W=\frac{\lambda }{3!}\mathrm{\Phi }^3`$, $`K=\overline{\mathrm{\Phi }}\mathrm{\Phi }`$ . Acknowledgements. The work was carried out under partial support of INTAS, project INTAS–96–0308; RFBR – DFG, project No. 99-02-04022; RFBR, project No. 99-02-16617; grant center of St. Peterburg University, project No. 97–6.2–34.
warning/0002/nucl-th0002006.html
ar5iv
text
# Faddeev approach to confined three-quark problems ## I introduction The Faddeev approach has for a long time been a very efficient tool for solving three-body problems in nuclear physics. Especially in the case of short-range forces the Faddeev equations can nowadays be solved to a high precision both for bound and scattering problems. The Faddeev theory offers a number of advantages over other approaches, such as solving the Schrödinger differential equation. In particular one can immediately include all relevant channels with the proper boundary conditions; in the cases of either two or three identical particles, a straightforward scheme for antisymmetrization is implemented; and finally a wide class of (local and nonlocal) two- and three-body interactions can be dealt with. However, with the Faddeev integral equations one encounters difficulties whenever the interactions are long-ranged. There is the notorious problem of solving the three-body Coulomb problem. Similarly, the solution of the constituent quark models poses practical difficulties. Irrespective of the type of confinement, one is confronted with complications due to the fact that the potentials are rising infinitely, as a function of inter-quark distance. In the past the Faddeev theory has already been successfully applied to the calculation of nonrelativistic constituent quark models. In this case, when the confinement in general remains relatively weak, a direct solution of the usual Faddeev equations can be achieved with a sufficient accuracy, and a variety of results for baryon spectra has been obtained . The situation becomes much more difficult for semirelativistic constituent quark models (i.e. with the kinetic-energy operator taken in relativistic form); in this case the confinement is considerably enhanced, to a strength practically consistent with the string tension of quantum chromodynamics. The convergence of the partial wave series gets slow, a relatively large number of channels needs to be included, and one may also easily pick up spurious contributions in the Faddeev components, which get wiped out only in the total three-body wave functions. The corresponding problems are, of course, already there in a nonrelativistic framework and they were thoroughly studied (e.g. in Ref. ) for the case of the harmonic-oscillator potential, which sometimes is also used as a simplified model for confinement. In the attempt to make the Faddeev approach directly amenable to any three-body problem with confining interactions, be it in a nonrelativistic or semirelativistic framework, one of us has previously suggested a different set-up of the Faddeev formalism . The proposed method exploits the asymptotic filtering property of the Faddeev scheme by adopting a different splitting of the full three-body Hamiltonian. Already in the nonrelativistic case considered in Ref. this has led to a more efficient solution of the (modified) Faddeev equations. Here we further elaborate on this method and demonstrate its efficiency even in the case of semirelativistic constituent quark models. Usually the full Hamiltonian $`H`$ of the system under consideration is split into a free Hamiltonian (kinetic-energy operator) $`H^0`$ and an interaction part $`H^I`$; the latter contains all two and three-body forces. However, this is not absolutely necessary. One may equally well adopt a different splitting of $`H`$, where part of the interactions are put into a modified Hamiltonian $`\stackrel{~}{H}^0`$ and the residual ones are retained in $`\stackrel{~}{H}^I`$. This turns out to be especially useful whenever there are interactions that cause difficulties in the original Faddeev scheme. It is usually the case for interactions that do not fall off fast enough at large distances, e.g., the Coulomb forces and notably also infinite confinement potentials. In the following section we outline the method that can treat confined three-body problems in an efficient and accurate manner. Beyond solving nonrelativistic problems we demonstrate how the method can also be applied to semirelativistic constituent quark models, which rely on a kinetic-energy operator in relativistic form. In Section 3 we show how to solve the modified Faddeev equations, and we prove the good performance of the method in Section 4, taking the specific examples of the Goldstone-boson-exchange (GBE) constituent quark model in the nonrelativistic and semirelativistic versions. ## II Treatment of confinement in the Faddeev approach We start from the total Hamiltonian of a nonrelativistic or a semirelativistic three-quark system, which can be written as $$H=H^0+v_\alpha +v_\beta +v_\gamma ,$$ (1) where $`H^0`$ is the three-body kinetic-energy operator and $`v_\delta =v_\delta ^c+v_\delta ^{hf}`$ represents the mutual quark-quark interactions containing both the confinement ($`v_\delta ^c`$) and hyperfine ($`v_\delta ^{hf}`$) potentials in the subsystems $`\delta =\alpha ,\beta ,\gamma `$. For the moment we leave out genuine three-quark forces, which, however, can easily be included into the formalism and the method we follow. In the nonrelativistic case we may express the kinetic-energy operator by four equivalent forms $`H^0`$ $`=`$ $`{\displaystyle \frac{p_\alpha ^2}{2\mu _\alpha }}+{\displaystyle \frac{q_\alpha ^2}{2M_\alpha }}={\displaystyle \frac{p_\beta ^2}{2\mu _\beta }}+{\displaystyle \frac{q_\beta ^2}{2M_\beta }}={\displaystyle \frac{p_\gamma ^2}{2\mu _\gamma }}+{\displaystyle \frac{q_\gamma ^2}{2M_\gamma }}`$ (2) $`=`$ $`{\displaystyle \underset{i=1}{\overset{3}{}}}{\displaystyle \frac{k_i^2}{2m_i}},`$ (3) i.e. either through individual particle momenta $`\stackrel{}{k}_i`$ in the center-of-mass system or in terms of relative momenta $`\stackrel{}{p}_\delta `$ and $`\stackrel{}{q}_\delta `$ conjugate to the usual Jacobi coordinates $`\stackrel{}{x}_\delta `$ and $`\stackrel{}{y}_\delta `$, respectively ($`\delta =\alpha ,\beta ,\gamma `$). In Eq. (3), $`m_i`$ denotes the individual particle mass, $`\mu _\delta `$ the reduced mass in the two-body subsystem $`\delta `$, and $`M_\delta `$ the reduced mass of this subsystem with the third particle $`\delta `$. In the semirelativistic case the kinetic-energy operator takes the form $$H^0=\underset{i=1}{\overset{3}{}}\sqrt{k_i^2+m_i^2},$$ (4) where again $`\stackrel{}{k}_i`$ are the individual particle three-momenta in the frame with total three-momentum $`\stackrel{}{P}=_{i=1}^3\stackrel{}{k}_i=0`$. We note that a Hamiltonian as in Eq. (1) together with the relativistic kinetic-energy operator (4) represents an allowed mass operator in the point-form formalism of Poincaré-invariant quantum mechanics, irrespective of the dynamical origin of the interactions . In the conventional Faddeev treatment the total Hamiltonian (1) is first split into the free Hamiltonian $`H^0`$ and the interaction Hamiltonian $`H^I=v_\alpha +v_\beta +v_\gamma `$. Then a decomposition of the total wave function $$|\mathrm{\Psi }=|\psi _\alpha +|\psi _\beta +|\psi _\gamma $$ (5) is carried out such that each of the three components satisfies the Faddeev integral equations of the type $$|\psi _\alpha =G_\alpha (E)v_\alpha (|\psi _\beta +|\psi _\gamma )$$ (6) with $`\alpha ,\beta ,\gamma `$ a cyclic permutation. Herein the so-called channel resolvent $$G_\alpha (E)=(EH^0v_\alpha )^1$$ (7) occurs, which contains the interactions in subsystem $`\alpha `$ but is otherwise characterized by a free motion of the third particle relative to the two-body subsystem. This scheme works well for short-range interactions and makes it possible to easily incorporate the correct asymptotic behavior. Strictly speaking the standard Faddeev scheme applies only for potentials falling of fast enough at large distances. When the interactions are long-ranged, however, there is never a free motion of the third particle relative to the two-body subsystem. This makes it necessary to modify the Faddeev formalism. Otherwise one risks unpleasant properties in the Faddeev components. In particular, for infinitely rising potentials spurious contributions are picked up and also the partial-wave series becomes slowly convergent. One can circumvent these difficulties by modifying the decomposition (5) such that all the long-range potentials are included in a modified channel Green’s operator. Specifically, in our case at least the long-range parts of the confinement interactions in all subsystems $`\alpha ,\beta ,`$ and $`\gamma `$ should be included in the modified channel resolvent . One can attain this goal by adopting a different splitting of the total Hamiltonian into $$H=H^c+\stackrel{~}{v}_\alpha +\stackrel{~}{v}_\beta +\stackrel{~}{v}_\gamma ,$$ (8) where $$H^c=H^0+\stackrel{~}{v}_\alpha ^c+\stackrel{~}{v}_\beta ^c+\stackrel{~}{v}_\gamma ^c$$ (9) contains, besides the kinetic energy, the long-range parts $`\stackrel{~}{v}_\delta ^c`$ of the confining interactions $`v_\delta ^c`$ in all subsystems. The potentials $`\stackrel{~}{v}_\delta `$ are the residual interactions containing the hyperfine potentials and the short-range parts of the confinement. Based on Eqs. (8) and (9) we now decompose the total wave function into $$|\mathrm{\Psi }=|\stackrel{~}{\psi }_\alpha +|\stackrel{~}{\psi }_\beta +|\stackrel{~}{\psi }_\gamma ,$$ (10) where the modified Faddeev components are defined as $$|\stackrel{~}{\psi }_\alpha =G^c(E)\stackrel{~}{v}_\alpha |\mathrm{\Psi }$$ (11) with $$G^c(E)=(EH^c)^1.$$ (12) They fulfill the integral equations $$|\stackrel{~}{\psi }_\alpha =G_\alpha ^c(E)\stackrel{~}{v}_\alpha (|\stackrel{~}{\psi }_\beta +|\stackrel{~}{\psi }_\gamma ),$$ (13) with $`\alpha ,\beta ,\gamma `$ again a cyclic permutation. As compared to Eq. (7) the new channel resolvent gets modified to $$G_\alpha ^c(E)=(EH^c\stackrel{~}{v}_\alpha )^1.$$ (14) It exhibits just the desired property of including the long-range confining interactions in all subsystems $`\alpha ,\beta ,\gamma `$. Only the short-range potential $`\stackrel{~}{v}_\alpha `$ remains in the modified Faddeev equations (13). Specifically, since now $`G_\alpha ^c`$ contains also the long-range parts $`\stackrel{~}{v}_\beta ^c+\stackrel{~}{v}_\gamma ^c`$ of the confinement interactions in channels $`\beta `$ and $`\gamma `$, the dependence of the component $`|\stackrel{~}{\psi }_\alpha `$ on the Jacobi coordinate $`\stackrel{}{y}_\delta `$ can never become a free motion. Rather the proper confinement-type asymptotic conditions are imposed on $`|\stackrel{~}{\psi }_\alpha `$. As a result, spurious contributions are avoided in the individual Faddeev components, and at the same time the partial-wave expansion converges much faster. The splitting of the interactions in Eqs. (8) and (9) has to be done with care. In general, the interaction parts put into $`H^c`$ must not produce any bound state. Otherwise the proper behavior of the Faddeev components $`|\stackrel{~}{\psi }_\alpha `$ would again be spoiled. Suppose the potentials contained in $`H^c`$ would produce bound states. Then at the corresponding energies, the resolvent $`G^c(E)`$ would become singular. Consequently, according to Eq. (11), any large Faddeev component $`|\stackrel{~}{\psi }_\alpha `$ could be generated even if the full solution $`|\mathrm{\Psi }`$ remains infinitesimally small. Therefore, besides the true physical solutions of the Hamiltonian $`H`$, Eqs. (13) would also produce spurious solutions associated with the discrete eigenstates of the Hamiltonian $`H^c`$ . These spurious solutions would occur for any $`\stackrel{~}{v}_\alpha `$, thus having no bearing for the physical spectrum of $`H`$. Of course, when adding up the three individual Faddeev components these spurious solutions would cancel out. However, they would cause numerical instabilities in the practical calculations. Therefore they should be avoided by not allowing $`H^c`$ to produce any bound states. In the case of confinement interactions the above requirement cannot strictly be met, since even the longest-range parts of the infinitely rising potential generate bound states. However, there is a practical way out: one needs to eliminate the bound states generated by $`H^c`$ only in the region of physical interest. Outside that domain, i.e. reasonably far above the physical spectrum, they do not matter. In practice, upon splitting the interactions in the Hamiltonian (8) an auxiliary short-range potential is introduced with no effect on the physically interesting states. It only serves the purpose of cutting off the confinement interaction at short and intermediate distances thus avoiding low-lying bound states of $`H^c`$. In the solution of the Faddeev equations below we take a Gaussian form for the auxiliary potential as used already in Ref. . ## III Solution of the modified Faddeev Equations We solve Eqs. (13) along the Coulomb-Sturmian (CS) separable expansion approach . For a two-body system in an angular momentum state $`l`$ one makes use of the CS functions defined by $$r|n=\left[\frac{n!}{(n+2l+1)!}\right]^{1/2}(2br)^{l+1}\mathrm{exp}(br)L_n^{2l+1}(2br)$$ (15) or $$p|n=\frac{(n+l+1)l!\sqrt{2n!}}{\sqrt{\pi (n+2l+1)!}}\frac{b(4bp)^{l+1}}{(p^2+b^2)^{2l+2}}G_n^{l+1}\left(\frac{p^2b^2}{p^2+b^2}\right)$$ (16) in configuration and momentum spaces, respectively. Here, $`L`$ are the Laguerre and $`G`$ the Gegenbauer polynomials, and $`b`$ is a parameter. The CS functions form a complete set $$\mathrm{𝟏}=\underset{N\mathrm{}}{lim}\underset{n=0}{\overset{N}{}}|\stackrel{~}{n}n|=\underset{N\mathrm{}}{lim}\mathrm{𝟏}_N,$$ (17) where, in configuration space, $`|\stackrel{~}{n}`$ is defined by $`r|\stackrel{~}{n}=r|n/r`$. In three-body Hilbert space one extends the basis by defining the direct product $$|n\nu _\alpha =\{|n_\alpha |\nu _\alpha \},(n,\nu =0,1,2,\mathrm{}),$$ (18) where the states $`|n_\alpha `$ and $`|\nu _\alpha `$ are associated with the Jacobi coordinates $`x_\alpha `$ and $`y_\alpha `$, respectively. The curly brackets stand for the angular-momentum coupling including orbital angular momentum, spin and isospin. The completeness relation now takes the form $$\mathrm{𝟏}=\underset{N\mathrm{}}{lim}\underset{n,\nu =0}{\overset{N}{}}|\stackrel{~}{n\nu }_\alpha _\alpha n\nu |=\underset{N\mathrm{}}{lim}\mathrm{𝟏}_N^\alpha .$$ (19) It should be noted that we can introduce three analogous bases, which belong to different fragmentations $`\alpha `$, $`\beta `$, and $`\gamma `$. In the practical solution of the modified Faddeev equations (13) we terminate the sums in Eq. (18) at some finite $`N`$. Of course, the basis can always be made large enough to produce any desired accuracy. Thereby we provide for a separable expansion of the short-range potential $`\stackrel{~}{v}_\alpha `$ in the three-body Hilbert space $$\stackrel{~}{v}_\alpha \mathrm{𝟏}_N^\alpha \stackrel{~}{v}_\alpha \mathrm{𝟏}_N^\beta =\underset{n,\nu ,n^{},\nu ^{}=0}{\overset{N}{}}|\stackrel{~}{n\nu }_\alpha \underset{¯}{\overset{~}{v}}_{\alpha \beta }_\beta \stackrel{~}{n^{}\nu ^{}}|,$$ (20) where $`\underset{¯}{\overset{~}{v}}_{\alpha \beta }=_\alpha n\nu |\stackrel{~}{v}_\alpha |n^{}\nu ^{}_\beta `$. Inserting expression (20) into Eq. (13) yields $$|\stackrel{~}{\psi }_\alpha =G_\alpha ^c(E)[\mathrm{𝟏}_N^\alpha \stackrel{~}{v}_\alpha \mathrm{𝟏}_N^\beta |\stackrel{~}{\psi }_\beta +\mathrm{𝟏}_N^\alpha \stackrel{~}{v}_\alpha \mathrm{𝟏}_N^\gamma |\stackrel{~}{\psi }_\gamma ].$$ (21) By applying the CS states $`_\alpha \stackrel{~}{n\nu }|`$ from the left one obtains a linear system of homogeneous equations for the coefficients of the Faddeev components $`\underset{¯}{\overset{~}{\psi }}_{\alpha n\nu }=_\alpha \stackrel{~}{n\nu }|\stackrel{~}{\psi }_\alpha `$: $$\{[\underset{¯}{G}^c(E)]^1\underset{¯}{\overset{~}{v}}\}\underset{¯}{\overset{~}{\psi }}=0.$$ (22) A unique solution exists if and only if $$det\{[\underset{¯}{G}^c(E)]^1\underset{¯}{\overset{~}{v}}\}=0.$$ (23) This condition determines the energy eigenvalues. The matrices $`\underset{¯}{G}^c(E)`$ and $`\underset{¯}{\overset{~}{v}}`$ have block structure and their matrix elements are given by $$\underset{¯}{G}_{\alpha n\nu n^{}\nu ^{}}^c=\delta _{\alpha \beta }_\alpha \stackrel{~}{n\nu }|G_\alpha ^c(E)|\stackrel{~}{n^{}\nu ^{}}_\alpha $$ (24) and $$\underset{¯}{\overset{~}{v}}_{\alpha \beta n\nu n^{}\nu ^{}}=(1\delta _{\alpha \beta })_\alpha n\nu |\stackrel{~}{v}_\alpha |n^{}\nu ^{}_\beta .$$ (25) Notice that the matrix elements of the resolvent are needed only between states of one and the same partition $`\alpha `$ whereas the matrix elements of the potentials are always to be taken between states of different partitions $`\alpha `$ and $`\beta `$. While the latter can easily be evaluated (numerically) either in configuration or in momentum space, the matrix elements $`\underset{¯}{G}_\alpha ^c`$ are more involved. In Eq. (22) the inverse of the matrix $`\underset{¯}{G}_\alpha ^c(E)`$ is needed and thus we make use of the approximation $$(_\alpha \stackrel{~}{n\nu }|G_\alpha ^c(E)|\stackrel{~}{n^{}\nu ^{}}_\alpha )^1_\alpha n\nu |(EH^c\stackrel{~}{v}_\alpha )|n^{}\nu ^{}_\alpha ,$$ (26) i.e. the inverse of the matrix containing the matrix elements of an operator is replaced by the matrix containing the matrix elements of the inverse operator. Since $`H^c`$ contains all the confining interactions, this approximation does not modify the character of the spectrum of $`\underset{¯}{G}_\alpha ^c`$ and, like the approximation in Eq. (20), it becomes exact in the limit $`N\mathrm{}`$. The matrix elements on the r.h.s. of Eq. (26) can be calculated in a straightforward manner, as the matrix element of $`E`$ and the ones of the nonrelativistic kinetic-energy operator are known in closed analytical form. The potential matrix elements can easily be calculated numerically. In the case of the relativistic kinetic energy (4) the matrix elements of the corresponding operator are most conveniently calculated numerically in momentum space. ## IV Performance of the method In order to demonstrate the efficiency of our method we present nonrelativistic and semirelativistic three-quark calculations for light baryons. For the quark-quark dynamics we choose the Goldstone-boson-exchange chiral quark model . The total quark-quark interaction is thus given by $$v=v^c+v^\chi $$ (27) where the confinement is of the linear form $$v^c(r)=V_0+Cr.$$ (28) The chiral interaction is derived from Goldstone-boson exchange and is represented by the spin-spin component of pseudoscalar meson exchange (for the corresponding formulae see Refs. ). The parameters of the nonrelativistic GBE chiral quark model can be found in Ref. , whereas the semirelativistic version is specified in Refs. . For the nonrelativistic version of this model the solutions have previously been obtained in good agreement by both the standard Faddeev approach and the stochastic variational method (SVM) , whereas for the semirelativistic model only SVM solutions have existed up to now. Here we are completing the solutions with the Faddeev integral-equation results also for the semirelativistic version of the GBE quark model. In the splitting of the short- and long-range potential parts in the confinement interaction we introduce an auxiliary potential in Gaussian form such that $$\stackrel{~}{v}^c(r)=V_0+Cr+a_0\mathrm{exp}[(r/r_0)^2]$$ (29) and $$\stackrel{~}{v}^\chi (\stackrel{}{r})=v^\chi (\stackrel{}{r})a_0\mathrm{exp}[(r/r_0)^2].$$ (30) As explained in the previous section, this guarantees that the bound states of $`H^c`$ are shifted far beyond the energy range of physical interest (i.e. the light-baryon ground states and resonances). The parameters of the auxiliary potential have been taken as $`a_0=3\text{fm}^1`$ and $`r_0=1\text{fm}`$. By this choice of the parameter values any bound states of $`H^c`$ are avoided below $`2`$ GeV. The values of $`a_0`$ and $`r_0`$ also influence the rate of convergence but not the final (converged) results. With the quark-quark interactions prepared in this way we can go ahead and calculate the $`N`$ and $`\mathrm{\Delta }`$ spectra by solving Eq. (22). The results for the ground states and the excitation spectra are shown in Tables I-VI for both the nonrelativistic and semirelativistic versions of the GBE quark model. We have also displayed the rate of convergence with respect to the number of channels included. The convergence rate turns out to be essentially the same for the nonrelativistic and semirelativistic cases. It is practically not influenced by the strength of confinement, which is more than three times stronger in the latter case. One observes that only a few angular-momentum states are needed to attain satisfactory convergence. All this is essentially due to the modified splitting of the Hamiltonian according to Eq. (8). The convergence is faster for the $`\mathrm{\Delta }`$ ground state (Table V). Also for its first positive- and negative-parity excitations only a rather small number of channels need to be included. In case of the nucleon and its resonances at most $`10`$ channels are necessary. For the separable expansion of the potentials, according to Eq. (20), about $`20`$ terms are perfectly sufficient. In all cases satisfactory agreement with the results obtained with the SVM is reached. In the present work we have also succeeded in treating semirelativistic constituent quark models, whose kinetic-energy operator is of the form (4). Technically the calculation of the semirelativistic case is more involved, as the matrix elements of $`H^0`$ must also be calculated numerically. Otherwise, however, the procedure is the same as in the nonrelativistic case. Finally we remark that we can also calculate wave functions by putting together the solutions for the Faddeev components. The total wave functions are then expressed in terms of CS functions. This fact may facilitate their use in further applications. ## V Summary We have proposed an efficient method for solving the three-body Faddeev equations for the case of infinitely rising confinement interactions. The method basically consists in splitting of the total Hamiltonian in a way different from the usual approach: the long-range parts of the interactions are incorporated into a modified resolvent, which takes the place of the usual free resolvent. Consequently the Faddeev decomposition of the total three-body wave function is carried out in a different manner, and one derives a set of modified Faddeev equations. In solving them one can avoid the difficulties associated with long-range confining interactions. We have demonstrated the performance of our method in the case of the Goldstone-boson-exchange chiral quark model both for its nonrelativistic and semirelativistic versions. We have obtained a rapid convergence with the number of angular-momentum states included, irrespective of the strength of confinement. Our results agree perfectly with the ones obtained before . In particular, for the semirelativistic case we have now confirmed within the Faddeev approach the results that have heretofore only been available from a variational method. It is comfortable to see that the two solution methods, which are quite distinct in nature, produce exactly the same answers for the light baryon spectra. ###### Acknowledgements. This work was supported by the Austrian-Hungarian Scientific-Technical Cooperation within project A-14/1998 and by Hungarian OTKA grants No. T026233 and T029003.
warning/0002/hep-th0002079.html
ar5iv
text
# The 𝜈⁢𝜈⁢𝛾 Amplitude in an External Homogeneous Electromagnetic Field ## I Introduction Neutrino-photon interactions are of interest in astrophysics and cosmology. It is well known that in the standard model, neutrino-photon interactions appear at the one-loop level. It is also well known that in a vacuum the reaction $`\nu \nu \gamma `$ is forbidden. The processes with two photons, for example, neutrino-photon scattering $`\gamma \nu \gamma \nu `$, turn out to be highly suppressed. In Gell-Mann showed that in the four-Fermi limit of the standard model the amplitude is exactly zero to order $`G_F`$ because, according to Yang’s theorem , two photons cannot couple to the $`J=1`$ state. Therefore the amplitude is suppressed by the additional factors of $`\omega /m_W`$, where $`\omega `$ is the photon energy and $`m_W`$ is the $`W`$ mass . For example, in the case of massless neutrinos, the amplitude for $`\gamma \nu \gamma \nu `$ in the standard model is suppressed by the factor $`1/m_W^2`$ . As a result, the cross section is exceedingly small. So we see that neutrino-photon processes with one or two photons do not occur or are suppressed in the vacuum. The presence of a medium or external electromagnetic field drastically changes the situation. It induces an effective coupling between photons and neutrinos, which contributes to the $`\nu \nu \gamma `$ process and cross-related reactions. Furthermore, it was shown in that in the presence of an external magnetic field, cross sections for neutrino-photon processes such as $`\gamma \gamma \nu \overline{\nu }`$ and $`\nu \gamma \nu \gamma `$ are enhanced by the factor $`\left(m_W/m_e\right)^4\left(B/B_c\right)^2`$ for $`\omega m_e`$ and $`BB_c`$. Later this result was extended to very strong magnetic fields and arbitrary $`\omega `$ . In this paper we will deal with $`\nu \nu \gamma `$ and $`\gamma \nu \overline{\nu }`$ reactions in the presence of an external electromagnetic field. The photon decay process $`\gamma \nu \overline{\nu }`$ in presence of a magnetic field was studied by several authors . The Cherenkov process $`\nu \nu \gamma `$ is the cross process to photon decay and was studied in a crossed field and in a magnetic field in . The aim of this paper is to consider the case of an arbitrary homogeneous electromagnetic field. This case was also considered recently in but, from our point of view, the expression obtained is unfinished insofar as the very short tensorial structure of the V-A two-point amplitude, which we derive in (24), is nowhere visible in . Therefore it was not possible to compare our expressions to Schubert’s formula (5.15). Since this topic abounds in errors and controversies ( see for a thorough discussion of the literature and its critique), we will devote this article mainly to a careful analysis of this expression and will postpone physical applications to another forthcoming paper. ## II The $`\nu \nu \gamma `$ Process in Presence of External Fields Let us begin with the effective Lagrangian. At energies very much smaller than the W- and Z-boson masses, both processes are described by an effective four-fermion interaction, $$_{eff}=\frac{G_F}{\sqrt{2}}\overline{\nu }\gamma ^\mu \left(1+\gamma _5\right)\nu \overline{E}\gamma _\mu \left(g_V+g_A\gamma _5\right)E.$$ (1) Here, $`E`$ stands for the electron field, $`\gamma _5=i\gamma ^0\gamma ^1\gamma ^2\gamma ^3`$, $`g_V=1\frac{1}{2}\left(14\mathrm{sin}^2\theta _W\right)`$ and $`g_A=1\frac{1}{2}`$ for $`\nu _e`$ , where the first terms in $`g_V`$ and $`g_A`$ are the contributions from the W exchange diagram and the second one from the Z exchange diagram. Also, $`g_V=2\mathrm{sin}^2\theta _W\frac{1}{2}`$ and $`g_A=\frac{1}{2}`$ for $`\nu _{\mu ,\tau }`$. Then the amplitude for the diagram in Fig.1 is given by $$=\frac{G_F}{\sqrt{2}e}\epsilon _\mu \overline{\nu }\gamma _\nu (1+\gamma _5)\nu (g_V\mathrm{\Pi }^{\mu \nu }+g_A\mathrm{\Pi }_5^{\mu \nu }),$$ (2) where $`\mathrm{\Pi }^{\mu \nu }`$ is the well-known polarisation operator of QED. It was calculated earlier in . Our aim is to calculate $`\mathrm{\Pi }_5^{\mu \nu }`$. For this purpose we will use the technique developed in . In this approach we begin with $$\mathrm{\Pi }_5^{\mu \nu }=ie^2M_5^{\mu \nu },$$ (3) where $$M_5^{\mu \nu }Sp0\left|\gamma ^\mu \frac{1}{𝒫/m+i\epsilon }\gamma ^\nu \gamma ^5\frac{1}{𝒫/k/m+i\epsilon }\right|0,$$ (4) with $`𝒫_\mu =i_\mu eA_\mu `$, and we have to calculate the mean value over the states x=0 : $`0\left|\mathrm{}\right|0`$. For the analysis it is convenient to use a special representation for the field tensor: $$F_{\mu \nu }=aC_{\mu \nu }+bB_{\mu \nu },$$ (5) with the tensors $`C_{\mu \nu }`$ and $`B_{\mu \nu }`$ defined by $$C_{\mu \nu }=\frac{1}{a^2+b^2}\left(aF_{\mu \nu }+bF_{\mu \nu }^{}\right);B_{\mu \nu }=\frac{1}{a^2+b^2}\left(bF_{\mu \nu }aF_{\mu \nu }^{}\right),$$ (6) where $$a,b=\sqrt{(^2+𝒢^2)^{\frac{1}{2}}\pm };=\frac{1}{4}F_{\mu \nu }F^{\mu \nu },𝒢=\frac{1}{4}F_{\mu \nu }^{}F^{\mu \nu }.$$ (7) Then<sup>*</sup><sup>*</sup>*We always use the metric $`g=\mathrm{diag}(+)`$. $$\left(C^2\right)_{\mu \nu }=\frac{1}{a^2+b^2}\left(F_{\mu \nu }^2+b^2g_{\mu \nu }\right);\left(B^2\right)_{\mu \nu }=\frac{1}{a^2+b^2}\left(F_{\mu \nu }^2a^2g_{\mu \nu }\right).$$ (8) These tensors satisfy very useful identities: $$\left(CB\right)_{\mu \nu }=0;\left(C^3\right)_{\mu \nu }=C_{\mu \nu };\left(B^3\right)_{\mu \nu }=B_{\mu \nu }.$$ (9) Now the result of our calculation is given by $`M_5^{\mu \nu }`$ $`=2i{\displaystyle \frac{\pi ^2}{(2\pi )^4}}\{2ebC^{\mu \nu }2eaB^{\mu \nu }+{\displaystyle _1^1}dv{\displaystyle _0^{\mathrm{}}}sds{\displaystyle \frac{ie^2ab}{\mathrm{sin}(ebs)\mathrm{sinh}(eas)}}\mathrm{exp}[i\mathrm{\Psi }]e^{ism^2}`$ (10) $`[(Ck)^\mu (C^2k)^\nu M_1+(Bk)^\mu (B^2k)^\nu M_2+`$ (11) $`[(C^2k)^\mu (Bk)^\nu +(Bk)^\mu (C^2k)^\nu +(kC^2k)B^{\mu \nu }]M_3`$ (12) $`[(B^2k)^\mu (Ck)^\nu +(Ck)^\mu (B^2k)^\nu +(kB^2k)C^{\mu \nu }]M_4]\},`$ (13) where $$\mathrm{\Psi }=\frac{(kC^2k)}{2ea}\frac{\mathrm{cosh}(eas)\mathrm{cosh}(easv)}{\mathrm{sinh}(eas)}+\frac{(kB^2k)}{2eb}\frac{\mathrm{cos}(ebs)\mathrm{cos}(ebsv)}{\mathrm{sin}(ebs)},$$ (15) with $`M_1`$ $`=`$ $`\mathrm{sin}(ebs)(\mathrm{cosh}(eas)\mathrm{cosh}(easv)){\displaystyle \frac{1}{\mathrm{sinh}^2(eas)}}`$ (16) $`M_2`$ $`=`$ $`\mathrm{sinh}(eas)(\mathrm{cos}(ebs)+\mathrm{cos}(ebsv)){\displaystyle \frac{1}{\mathrm{sin}^2(ebs)}}`$ (17) $`M_3`$ $`=`$ $`{\displaystyle \frac{1}{2}}(\mathrm{cos}(ebsv)\mathrm{cosh}(easv)\mathrm{coth}(eas)+\mathrm{cos}(ebsv){\displaystyle \frac{1}{\mathrm{sinh}(eas)}}+`$ (19) $`\mathrm{cot}(ebs)\mathrm{sin}(ebsv)\mathrm{sinh}(easv))`$ $`M_4`$ $`=`$ $`{\displaystyle \frac{1}{2}}(\mathrm{cos}(ebsv)\mathrm{cosh}(easv)\mathrm{cot}(ebs)\mathrm{cosh}(easv){\displaystyle \frac{1}{\mathrm{sin}(ebs)}}+`$ (21) $`\mathrm{coth}(eas)\mathrm{sin}(ebsv)\mathrm{sinh}(easv)).`$ As in the previous calculations in presence of a magnetic field , this result is not gauge invariant. We also see that our amplitude is ultraviolet convergent. Nevertheless, in order to preserve gauge invariance, we must regularize it using a gauge invariant regularization. This phenomenon should be familiar from quantum electrodynamics. For example, the photon-photon scattering amplitude, though formally converging, must be regularized to get a gauge invariant result . A convenient way of restoring gauge invariance in diagrams with pseudovector coupling is to use the Pauli-Villars regularization. Then it is easy to show that to obtain the final gauge invariant expression for the amplitude we must just omit the first two terms in (10), because they are cancelled by regulator terms. Finally, we obtain $`M_5^{\mu \nu }`$ $`=`$ $`{\displaystyle \frac{1}{8\pi ^2}}{\displaystyle _1^1}dv{\displaystyle _0^{\mathrm{}}}sds{\displaystyle \frac{e^2ab}{\mathrm{sin}(ebs)\mathrm{sinh}(eas)}}\mathrm{exp}[i\mathrm{\Psi }]e^{ism^2}\{(Ck)^\mu (C^2k)^\nu M_1+`$ (24) $`(Bk)^\mu (B^2k)^\nu M_2+[(C^2k)^\mu (Bk)^\nu +(Bk)^\mu (C^2k)^\nu +(kC^2k)B^{\mu \nu }]M_3+`$ $`[(B^2k)^\mu (Ck)^\nu +(Ck)^\mu (B^2k)^\nu +(kB^2k)C^{\mu \nu }]M_4\}.`$ This method of getting a gauge invariant expression allows us, unlike the authors of , to do without an anomaly cancellation mechanism at this stage of calculation. Let us note that the whole contribution of the triangle diagram is still present in (24). This fact can be easily confirmed by an independent calculation of the triangle graph contribution. To deal with unwanted anomaly terms we must, as usual, take into account all fermions in a generation. In the following this will always be assumed. Equation (24) is the main result of this article. Let us now consider various checks of this expression. First, in the limit of zero electric field it reproduces the results of . A second nontrivial check of formula (24) is to consider the following transformation: $$aib;bia;C_{\mu \nu }iB_{\mu \nu };B_{\mu \nu }iC_{\mu \nu };F_{\mu \nu }F_{\mu \nu }.$$ (25) Because of $`F_{\mu \nu }F_{\mu \nu }`$, $`F_{\mu \nu }^{}F_{\mu \nu }^{}`$, the tensor $`M_5^{\mu \nu }`$ must be invariant under this transformation. From (24) we see that it is indeed invariant and that under this transformation, the first term is interchanged with the second, and the third term is interchanged with the fourth, thus demonstrating the nontriviality of this check. The next interesting limit is the case of a crossed field, where $`\stackrel{}{E}`$ and $`\stackrel{}{B}`$ are not only orthogonal but are also of equal modulus. This causes the two invariants $``$ and $`𝒢`$ to vanish. The field dependence is therefore completely described by the invariant $`(F^{\mu \nu }k_\nu )^2`$ . The correct procedure of finding the amplitude is to set $`a=b`$ first, and then take the limit $`a=b0`$. Using this procedure we find following expression: $`M_5^{\mu \nu }`$ $`=`$ $`{\displaystyle \frac{e}{8\pi ^2}}\{{\displaystyle \frac{1}{2m^2}}[k^2F^{\mu \nu }+(F^{}k)^\mu k^\nu +(F^{}k)^\nu k^\mu ]`$ (28) $`{\displaystyle _0^1}𝑑v\left(1v^2\right)u{\displaystyle _0^{\mathrm{}}}𝑑z\mathrm{exp}(i(zu+{\displaystyle \frac{z^3}{3}}))+`$ $`{\displaystyle \frac{4}{(kF^2k)}}(F^{}k)^\mu (F^2k)^\nu \{{\displaystyle _0^1}dvu{\displaystyle _0^{\mathrm{}}}dz\mathrm{exp}(i(zu+{\displaystyle \frac{z^3}{3}}))+i\}\},`$ where $$u=\left(\frac{e^2(kF^2k)}{16m^6}\right)^{\frac{1}{3}}(1v^2)^{\frac{2}{3}}.$$ (29) Earlier the amplitudes in crossed fields were considered in . We can compare only parts of (28) with those existing in the literature, because all previous results were presented in contracted form and with photons on the mass shell. Taking this into account we can say that our result contradicts that of and confirms that of . ## III Conclusions We have presented the $`\nu \nu \gamma `$ amplitude in the presence of an external homogeneous constant electromagnetic field. The result is gauge invariant and reproduces the known results for an external magnetic field. We further identified the crossed-field limit. Let us also emphasize that the general case of electric and magnetic fields is in need when considering ( low-frequency ) multiphoton one-loop processes with and without external fields. The situation is similar to processes where the Heisenberg-Euler lagrangian turns out to be useful as e.g., in the study of photon-splitting or in photon-neutrino processes as discussed in the recent article by Dicus and Repko . Hence we conclude that our results allow for a variety of applications that will also be discussed in a future publication. ## Acknowledgments I would like to thank Professor W. Dittrich for helpful discussions and for carefully reading the manuscript. I am also grateful to Dr H. Gies for valuable comments and suggestions. This work was supported by Deutsche Forschungsgemeinschaft under DFG Di 200/5-1.
warning/0002/cond-mat0002121.html
ar5iv
text
# Comment on: ”Auger decay, Spin-exchange, and their connection to Bose-Einstein condensation of excitons in 𝐶⁢𝑢₂⁢𝑂 ” by G.M. Kavoulakis, A. Mysyrowicz ( cond-mat/0001438 ) In the very recent work a new mechanism of the interconversion of the triplet excitons into singlet excitons in $`Cu_2O`$ has been suggested. In accordance with it, two triplet excitons with the opposite (internal) angular momenta may collide and interconvert into a pair of the singlet excitons. Estimates presented in show that such an interconversion is the most effective channel for the decrease of the triplet exciton density. This questions the commonly accepted view that the Auger decay is the primary channel for the decay (see in ). Furthermore, it has been pointed out in that the actual rate of the Auger decay must be several orders of magnitude less than it was previously calculated. In this comment, it is suggested that the mechanism leads to a verifiable prediction about the rate of the decrease of the triplet exciton density as function of the polarization of the incident laser light inducing the two-photon creation of the triplet excitons. As a matter of fact, this rate can be greatly reduced if the polarization of the laser light is properly chosen. First, it is worth noting that a necessary condition for the triplet-singlet interconversion is that before the collision a pair of the triplet excitons has total angular momentum $`J=0`$, that is, they form a singlet state (invariant under the point group rotations). Thus, the rate of the interconversion should be extremely sensitive to the initial state of the excitonic ensemble. If this state is a thermal mixture with random orientation of the excitonic spins, then each exciton can easily find another one with the opposite $`J_z`$, so that the collision between them would result in a pair of the singlet excitons . On the contrary, if initially excitonic spins were aligned, the collision induced interconversion will be completely suppressed because of the conservation of the angular momentum of the colliding pairs. Thus, aligning angular momenta of the triplet excitons in one way or another should prevent the triplet excitons from transforming into the singlet excitons in accordance with the mechanism . This property can be used as a test for the mechanism . One way for preparing triplet excitons with preferential orientation of their spins is a direct creation of the coherent triplet excitons employed in . In this work, the triplet excitons have been created by the two-photon direct transitions: The incoming laser field was tuned to the half of the excitonic frequency, so that the two-photon transition was in exact resonance with the triplet excitons. Such a method allows to create a dense and coherent cloud of the triplet excitons. This is practically a direct mean of creating a condensate of the triplet excitons. However, fast collision induced decay of the triplet excitons may destroy this coherence on the scale of few nanoseconds. On one hand, this is the case if the Auger decay is responsible for the triplet exciton depletion because this channel is not sensitive to the orientation of the angular momenta of the colliding pairs. On the other hand, if the primary channel for the decay is the mechanism , it should be possible to use such a polarization of the cloud that the created condensate of the triplet excitons is stable on much longer time scale. It is possible to employ general symmetry considerations, and make a suggestion for the choice of the orientation of the incoming laser fields in the geometry of the experiment . Indeed, the two-photon process of creation of the triplet exciton corresponds to the interaction term in the energy density $`H_{le}={\displaystyle \underset{a=\pm 1,0;i,j}{}}\psi _{(a)}^{}Q_{ij}^{(a)}E_iE_j+H.c.`$ (1) where $`\psi _{(a)}`$ stands for the triplet exciton Bose field which has three projections $`a=\pm 1`$ and $`a=0`$ of the (internal) angular momenta; $`E_j`$ denotes three space components of the incoming laser field $`E_j\mathrm{exp}(i\omega t)`$ which is taken in the rotating wave approximation; $`Q_{ij}^{(a)}`$ are corresponding matrices representing the point symmetry (including spins) of $`Cu_2O`$ in such a way that (1) is invariant under this symmetry. The interaction term responsible for the decay can be represented in the contact form (S-wave channel) as $`H_{op}=g_{op}\psi ^{}\psi ^{}(\psi _{(+1)}\psi _{(1)}+\psi _{(1)}\psi _{(+1)}+`$ (2) $`+\psi _{(0)}\psi _{(0)})+H.c.`$ (3) where $`\psi `$ is the field of the singlet excitons, and $`g_{op}`$ is the interaction constant such that the rate estimated in is $`g_{op}^2`$. It is worth noting that the term in the brackets in (2) is invariant under the symmetry group (including spins) of $`Cu_2O`$, where the excitonic states are formed on the total angular momenta states of $`Cu`$ (see in ). If the triplet excitons are created in such a manner that this invariant is zero, the interconversion process will be suppressed. The induced fields $`\psi _{(a)}`$ are given from (1) as $`\psi _{(a)}_{i,j}Q_{ij}^{(a)}E_iE_j`$. If substituted into (2), this will result in the term describing four-photon production of the singlet excitons which in general should be significant as long as the mechanism is dominant, provided the density $`|\psi _{(a)}|^2|_{i,j}Q_{ij}^{(a)}E_iE_j|^2`$ of the induced triplet excitons is large enough. In fact, the symmetry of $`Q_{ij}^{(a)}`$ is the same as that of the tensors of the direct quadrupole transitions for the triplet excitons. Using this, it is possible to find the energy density (2) as $`H_{op}g_{op}\psi ^{}\psi ^{}(E_x^2E_y^2+E_x^2E_z^2+E_y^2E_z^2)+H.c.`$ (4) where $`E_x,E_y,E_z`$ refer to the components of the laser field with respect to the principal cubic axes of $`Cu_2O`$. Accordingly, the requirement $`E_x^2+E_y^2+E_z^2=0`$ (5) insures that the interconversion process described by (2), (4) is zero in the dominant s-wave channel as long as no thermalization of the created triplet excitons occurs. A solution of (5) for the six components of $`E_j=E_j^{}+iE_j^{\prime \prime }`$, where $`E_j^{}`$ and $`E_j^{\prime \prime }`$ stand for the real and imaginary parts of $`E_j`$, respectively, can be represented as follows $`E_j={\displaystyle \frac{1}{ϵ_j^{}+iϵ_j^{\prime \prime }}}`$ (6) in terms of the two auxiliary real vectors $`ϵ_j^{},ϵ_j^{\prime \prime }`$ which are arbitrary except for the conditions $`{\displaystyle \underset{j}{}}ϵ_j^2={\displaystyle \underset{j}{}}ϵ_j^{\prime \prime 2},{\displaystyle \underset{j}{}}ϵ_j^{}ϵ_j^{\prime \prime }=0.`$ (7) The interpretation of these conditions is straightforward: the complex vector $`E_j`$ represented by (6) should be chosen in such a way that the two auxiliary vectors $`ϵ_j^{}`$ and $`ϵ_j^{\prime \prime }`$ are equal in magnitude to each other and are mutually orthogonal. For the case of the incidence of the light along the direction (1,1,1), the solution of (7), (6) gives that $`_jE_j^{}E_j^{\prime \prime }=0`$ and $`_jE_j^2=_jE^{\prime \prime 2}`$. Note that, given this, the interconversion rate $`k(T=0)=0`$, as opposed to the rate of the Auger decay which is not sensitive to temperature and the mutual orientation of the excitonic angular momenta of the colliding pairs. At finite temperatures $`T0`$, the normal component - thermal triplet excitons - is present in addition to the condensate. This component should be characterized by zero net spin polarization due to the interaction with phonons. Thus, the interconversion process will take place. Its rate $`k(T)`$ is proportional to the normal density $`n^{}`$. Thus, it must be strongly temperature dependent. It is straightforward to find an estimate for $`k(T)`$ in the temperature range $`\mu (0)<TT_0`$, where $`\mu (0)=4\pi an_0`$ and $`a,n_0`$ stand for the excitonic scattering length and the exciton condensate (spin-polarized) density, respectively; and $`T_0`$ denotes the temperature of the excitonic Bose-Einstein condensation. Indeed, in this range the normal component behaves almost as an ideal gas. Thus, the estimate follows from Eqs.(8-10) of Ref. where the total density is replaced by the density of the normal component $`n^{}=n_0(T/T_0)^{3/2}`$ . Accordingly, the ratio of the rate of the interconversion $`k(T)`$ at $`T0`$ for the spin-polarized excitonic condensate to the rate $`1/\tau _{o,p}`$ estimated for the case of the non-polarized cloud is $`k(T)\tau _{o,p}{\displaystyle \frac{n^{}}{n_0}}=\left({\displaystyle \frac{T}{T_0}}\right)^{3/2}1`$ (8) for the temperatures under consideration. At temperatures $`T<\mu (0)`$, further significant reduction of the rate should occur due to the interaction between the triplet exciton condensate and the normal component.
warning/0002/hep-ph0002182.html
ar5iv
text
# References The flavor–asymmetry of the light–quark sea in the nucleon has attracted a lot of attention and many attempts were undertaken to explain the origin and calculate its magnitude (e.g., Ref. and references therein). In the present article we study this issue inspired by the suggestion that this asymmetry is related to the Pauli exclusion principle (‘Pauli blocking’). Our proposed implementation of this idea is summarized by the phenomenological ansatz for the unpolarized and polarized antiquark distributions $$\overline{d}(x,Q_0^2)/\overline{u}(x,Q_0^2)=u(x,Q_0^2)/d(x,Q_0^2)$$ (1) and $$\mathrm{\Delta }\overline{d}(x,Q_0^2)/\mathrm{\Delta }\overline{u}(x,Q_0^2)=\mathrm{\Delta }u(x,Q_0^2)/\mathrm{\Delta }d(x,Q_0^2),$$ (2) respectively, with $`Q_0^2`$ being some low resolution scale, e.g., the one in . These are our basic relations for the flavor–asymmetries of the unpolarized and polarized light sea densities which imply that $`u>d`$ determines $`\overline{u}<\overline{d}`$, etc. This is in accordance with the suggestion of Feynman and Field that, since there are more $`u`$\- than $`d`$–quarks in the proton, $`u\overline{u}`$ pairs in the sea are suppressed more than $`d\overline{d}`$ pairs by the exclusion principle. It should be emphasized that our suggested regularity in (1) is entirely of empirical origin, and our anticipated relation (2) has of course to be tested by future polarized experiments. Both relations require obviously the idealized situation of maximal Pauli–blocking and hold approximately in some effective field theoretic (mesonic, bag and chiral) models as we shall see below. In Table I we present $`\overline{d}(x,Q_0^2)/\overline{u}(x,Q_0^2)`$ calculated according to Eq. (1) from the (fitted) $`d,u`$ input distributions of GRV98 , as compared to the actual fitted values of this ratio. The good agreement lends support to the phenomenological ansatz in Eq. (1) and thus also to the experimentally so far unknown polarized antiquark flavor–asymmetry implied by Eq. (2). The predictions for $`\mathrm{\Delta }\overline{d}/\mathrm{\Delta }\overline{u}`$ according to Eq. (2) are shown in Table II utilizing the most recent LO AAC distributions $`\mathrm{\Delta }u(x,Q_0^2)`$ and $`\mathrm{\Delta }d(x,Q_0^2)`$ which compare favorably with the predictions of the relativistic field theoretical chiral quark–soliton model for $`\mathrm{\Delta }\overline{d}/\mathrm{\Delta }\overline{u}`$, as well as with a recent analysis based on the statistical parton model . The latter flavor–asymmetry for $`\mathrm{\Delta }\overline{u}`$ and $`\mathrm{\Delta }\overline{d}`$ can also be studied by replacing the common constraint $`\mathrm{\Delta }\overline{u}=\mathrm{\Delta }\overline{d}\mathrm{\Delta }\overline{q}`$ by our present Eq. (2). Using the recent analysis of , for example, one obtains the LO results for $`\mathrm{\Delta }\overline{u}(x,Q_0^2)`$, $`\mathrm{\Delta }\overline{d}(x,Q_0^2)`$ and their difference presented in Figs. 1 and 2, respectively . These predictions refer to an input scale of $`Q_0^2=1`$ GeV<sup>2</sup> . At the somewhat lower dynamical input scales $`Q_0^2=0.30.4`$ GeV<sup>2</sup> , the maxima/minima of the curves shown in Figs. 1 and 2 move slightly to the right, i.e. to slightly larger values of $`x`$. Strictly speaking a more consistent study of the antiquark asymmetry should be done within the framework of the ‘valence’ scenario where $`\mathrm{\Delta }s(x,Q_0^2)=\mathrm{\Delta }\overline{s}(x,Q_0^2)=0`$. This, however, is expected to modify the present results only marginally. The NLO analysis of the polarized antiquark asymmetry affords a direct implementation of Eq. (2) in the fit procedure due to the enhanced sensitivity of the NLO calculation of $`g_1^{p,n}(x,Q^2)`$ to the polarized gluon distribution which is affected by modifications of the polarized quark and antiquark distributions. Again, no qualitative changes of our present results are expected. It is interesting to note that our results for the flavor–asymmetry of the polarized sea distributions at $`Q_0^2=1`$ GeV<sup>2</sup> in Figs. 1 and 2 are comparable to those obtained in chiral quark–soliton model calculations and in the previously mentioned statistical parton model which correctly reproduced the flavor–asymmetry of the unpolarized sea ! A further direct test of our phenomenological ansatz (2) must await the polarized version of the Drell–Yan $`\mu ^+\mu ^{}`$ pair production experiments which provided the information on the flavor–asymmetry of the unpolarized sea distributions in Eq. (1). Future polarized semi–inclusive DIS experiments at CERN (COMPASS) and DESY (HERMES) could also become relevant for measuring possible flavor–asymmetries of polarized light–quark sea distributions . The statistics of present SMC and HERMES measurements is not sufficient for testing our expectations in Figs. 1 and 2, for example, despite the fact that rather stringent model assumptions have been made for the analyses of these experiments in order to improve the statistical significance. Finally it should be noted that the data select the solution of Eq. (2) which satisfies $$\mathrm{\Delta }q(x,Q_0^2)\mathrm{\Delta }\overline{q}(x,Q_0^2)>0$$ (3) where $`q=u,d`$. This can be understood as a consequence of the expected predominant pseudoscalar configuration of the quark–antiquark pairs in the nucleon sea. In fact, Eqs. (1) and (2) can be rewritten as $`u_+\overline{u}_++u_{}\overline{u}_{}`$ $`=`$ $`d_+\overline{d}_++d_{}\overline{d}_{}f_p`$ (4) $`u_+\overline{u}_{}+u_{}\overline{u}_+`$ $`=`$ $`d_+\overline{d}_{}+d_{}\overline{d}_+f_a`$ (5) where the common helicity densities are given by $`\underset{\pm }{\overset{()}{q}}=(\stackrel{()}{q}\pm \mathrm{\Delta }\stackrel{()}{q})/2`$ and, for brevity, we dropped the $`x`$–dependence everywhere. A predominant (pseudo)scalar configuration of the $`(q\overline{q})`$ pairs in the nucleon sea implies, via Pauli–blocking, that the aligned quark–quark configurations $`q_+(q_+\overline{q}_{})`$ and $`q_{}(q_{}\overline{q}_+)`$ are suppressed relatively to the antialigned $`q_+(q_{}\overline{q}_+)`$ and $`q_{}(q_+\overline{q}_{})`$ ‘cloud’ configurations, i.e. $`f_p>f_a`$ which yields the result in Eq. (3). We would like to thank M. Stratmann and W. Vogelsang for helpful discussions and comments. This work has been supported in part by the ‘Bundesministerium für Bildung, Wissenschaft, Forschung und Technologie’, Bonn. Table I. The predicted flavor–asymmetry $`\overline{d}(x,Q_0^2)/\overline{u}(x,Q_0^2)`$ according to Eq. (1) using the GRV98 distributions $`u(x,Q_0^2)`$ and $`d(x,Q_0^2)`$ at the LO– and NLO–QCD input scales $`Q_0^2\mu _{\mathrm{LO}}^2(\mu _{\mathrm{NLO}}^2)=0.26`$ GeV$`{}_{}{}^{2}(0.4`$ GeV$`{}_{}{}^{2})`$. The actually fitted ratio $`(\overline{d}/\overline{u})_{\mathrm{fit}}`$ is shown as well for comparison. The NLO results are shown in parentheses. | x | 0.01 | 0.05 | 0.1 | 0.2 | 0.3 | | --- | --- | --- | --- | --- | --- | | $`\overline{d}/\overline{u}`$ | 1.12 (1.13) | 1.25 (1.24) | 1.34 (1.36) | 1.64 (1.70) | 2.04 (2.11) | | $`(\overline{d}/\overline{u})_{\mathrm{fit}}`$ | 1.03 (1.04) | 1.16 (1.19) | 1.50 (1.53) | 2.00 (1.84) | 1.98 (1.65) | Table II. The predicted polarized flavor–asymmetry $`\mathrm{\Delta }\overline{d}(x,Q_0^2)/\mathrm{\Delta }\overline{u}(x,Q_0^2)`$ according to Eq. (2) using the LO AAC input distributions $`\mathrm{\Delta }u(x,Q_0^2)`$ and $`\mathrm{\Delta }d(x,Q_0^2)`$ at $`Q_0^2=1`$ GeV<sup>2</sup> . The NLO results are similar. The chiral quark–soliton predictions for $`\mathrm{\Delta }\overline{d}/\mathrm{\Delta }\overline{u}`$ of Wakamatsu and Kubota are shown for comparison as well which refer to a scale $`Q_0^20.36`$ GeV<sup>2</sup>. | | x | 0.01 | 0.05 | 0.1 | 0.2 | 0.3 | | --- | --- | --- | --- | --- | --- | --- | | $`\frac{\mathrm{\Delta }\overline{d}}{\mathrm{\Delta }\overline{u}}`$ | AAC | 1.55 | 1.76 | 1.95 | 2.26 | 2.59 | | | soliton | 1.73 | 2.05 | 2.15 | 1.94 | 1.66 | Figure Captions * The predictions for the polarized sea distributions $`\mathrm{\Delta }\overline{u}(x,Q_0^2)`$ and $`\mathrm{\Delta }\overline{d}(x,Q_0^2)`$ according to Eq. (2) with the LO results for $`\mathrm{\Delta }u(x,Q_0^2)`$ and $`\mathrm{\Delta }d(x,Q_0^2)`$ taken from Ref. at $`Q_0^2=1`$ GeV<sup>2</sup>. * The same as in Fig. 1 but for $`x\mathrm{\Delta }\overline{u}(x,Q_0^2)x\mathrm{\Delta }\overline{d}(x,Q_0^2)`$ at $`Q_0^2=1`$ GeV<sup>2</sup>.
warning/0002/math0002061.html
ar5iv
text
# Is Bootstrap Really Helpful in Point Process Statistics? ## 1 Introduction Recently, bootstrap is a popular tool in many branches of statistics, also for stochastic processes. Thus it is natural to ask whether bootstrap techniques could be helpful also in point process statistics. Indeed, some authors have developed statistical procedures using bootstrap techniques, see e.g. , , and . All these papers deal with the estimation of the accuracy of estimators of point process characteristics. In the first three papers estimation of variance of pair correlation function estimators is treated. The last one presents a procedure to determine confidence regions for the intensity function of an inhomogeneous Poisson process. The fundamental idea of bootstrap to resample given data to obtain ‘new’ pseudo data appears also in statistics of stochastic processes, in particular in the analysis of time series, see e.g. . In some variants of the method, called the blockwise bootstrap, the time series is partitioned into several parts, which are then resampled. A similar idea is also applied in in the statistical analysis of a planar random set. Clearly, the partioning procedure can also be adapted to point process statistics. However, partition can destroy point structures or add new artificial structures to the point pattern. In the one-dimensional case the error resulting from this loss of information may be still acceptable, but in higher dimensions it will be serious. Thus in spatial point process statistics another method is used which is quite similar to the application of bootstrap in case of classical statistics: the points of the process (including their places, which are assumed to be pairwise different) are resampled. The pseudo pattern then consists of $`n`$ points $`x_1^{},\mathrm{},x_n^{}`$ which are obtained by sampling randomly with replacement $`n`$ times from the original data $`\{x_1,\mathrm{},x_n\}`$. Naturally, the pseudo patterns generated by this method have always multiple points. Thus they have a character different to that of the original, which does not have multiple points. Consequently, it would be surprising if quantities of such point processes would produce good estimators for quantities of the original point process. This paper analyses the pointwise resampling technique for some examples of point process statistics. Section 2 discusses the main ideas of the paper which presents a procedure for estimating the standard error of an estimator of the pair correlation function. In Section 3 a method (drawn from ) to determine confidence intervals for the intensity function of an inhomogeneous Poisson process is considered. Finally, an easier method is presented which yields confidence regions for the intensity function of an inhomogeneous Poisson process without bootstrap. ## 2 Variance of estimators of the product density This part discusses the main ideas of , where bootstrap techniques are used to approximate the standard error of a pair correlation function estimator. The calculations are presented in an abridged form; the complete calculations are given in the Appendix. ### 2.1 Fundamentals Let $`\mathrm{\Phi }`$ be a stationary and isotropic point process, see, for example, for definitions. A standard second order characteristic of $`\mathrm{\Phi }`$ is the product density function $`\varrho ^{(2)}(r)`$. This function can be interpreted heuristically as follows. If $`B_1`$ and $`B_2`$ are two infinitesimally small disjoint Borel sets of volumes d$`V_1`$ and d$`V_2`$ and if $`x_1B_1`$ and $`x_2B_2`$ are points of distance $`x_1x_2=r`$ then $`\varrho ^{(2)}(r)\text{d}V_1\text{d}V_2`$ is the probability that $`\mathrm{\Phi }`$ has a point in each of $`B_1`$ and $`B_2`$. A simple estimator of $`\varrho ^{(2)}`$ without any border correction is given by $$\widehat{\varrho }(r)=\frac{1}{2\pi r\nu (W)}\underset{x,y\mathrm{\Phi }W}{\overset{}{\stackrel{}{}}}K(rxy).$$ The summation goes over all point pairs with different members, $`W`$ denotes the window of observation and $`K`$ is a kernel function. This situation can be generalized to the case of any ‘two-point estimator’ $$\widehat{\theta }=\underset{x,y\mathrm{\Phi }}{\overset{}{\stackrel{}{}}}f(x,y)$$ (1) with $`f`$ being symmetrical in its arguments and of the form $$f(x,y)=\mathrm{𝟏}_W(x)\mathrm{𝟏}_W(y)h(x,y)$$ with some function $`h`$. As the special form of $`f`$ leading to $`\widehat{\varrho }(r)`$ is unimportant, the following calculations are carried out for a general $`\widehat{\theta }`$. The quantity of interest is the variance of $`\widehat{\theta }`$ which is given by $$𝐕\widehat{\theta }=𝐄\widehat{\theta }^2(𝐄\widehat{\theta })^2=s_4+4s_3+2s_2(𝐄\widehat{\theta })^2$$ (2) with $$s_i=\varrho ^{(i)}(x_1,\mathrm{},x_i)f(x_1,x_2)f(x_{i1},x_i)\text{d}x_1\mathrm{}\text{d}x_i,$$ where $`\varrho ^{(i)}`$ is the $`i`$th order product density function of $`\mathrm{\Phi }`$, see the Appendix. ### 2.2 Bootstrap version of $`\widehat{\theta }`$ Assume that a sample of $`\mathrm{\Phi }`$ is given which consists of $`n`$ points $`x_1,\mathrm{},x_n`$ in the observation window $`W`$. It is resampled $`N`$ times to obtain $`N`$ ‘new’ point patterns. Each pseudo pattern consists of $`n`$ points $`x_1^{},\mathrm{},x_n^{}`$ which are obtained by sampling randomly with replacement $`n`$ times from $`\{x_1,\mathrm{},x_n\}`$. Thus it happens that in the pseudo samples some points of the original point pattern do not occur while others occur twice or even more. Let the number of occurrences of $`x_i`$ in the $`k`$th sample be $`w_k(i)`$. Then the $`k`$th sample can be represented by the vector $`w_k=(w_k(1),\mathrm{},w_k(n))`$ which has a multinomial distribution. This distribution depends only on $`n`$. In the limiting case $`n\mathrm{}`$ the components $`w_k(i)`$ of $`w_k`$ are independent and Poisson distributed with mean $`\mu =1`$. The bootstrap estimate for the $`k`$th pseudo sample is $$\widehat{\theta _k^{}}=\underset{i,j=1}{\overset{n}{\stackrel{}{}}}f(x_i,x_j)w_k(i)w_k(j),k=1,\mathrm{},N$$ where the summation goes over all pairs $`(i,j)`$ with $`ij`$. The variance of $`\widehat{\theta }`$ is estimated by the usual variance estimator corresponding to the $`\widehat{\theta _k^{}}`$’s, $$\widehat{v_N^{}}=\frac{1}{N1}\underset{k=1}{\overset{N}{}}\left(\widehat{\theta _k^{}}\frac{1}{N}\underset{i=1}{\overset{N}{}}\widehat{\theta _i^{}}\right)^2.$$ Since the $`\widehat{\theta _k^{}}`$ are (conditionally on $`x_1,\mathrm{},x_n`$) independent and identically distributed, it is $`\underset{N\mathrm{}}{lim}\widehat{v_N^{}}`$ $`=`$ $`𝐕\widehat{\theta _1^{}}`$ $`=`$ $`𝐄\widehat{\theta _1^{}}^2(𝐄\widehat{\theta _1^{}})^2`$ $`=`$ $`\alpha _4{\displaystyle \underset{i,j,k,l=1}{\overset{n}{\stackrel{}{}}}}f(x_i,x_j)f(x_k,x_l)`$ $`+4\alpha _3{\displaystyle \underset{i,j,k=1}{\overset{n}{\stackrel{}{}}}}f(x_i,x_j)f(x_i,x_k)`$ $`+2\alpha _2{\displaystyle \underset{i,j=1}{\overset{n}{\stackrel{}{}}}}\left(f(x_i,x_j)\right)^2`$ with $`\alpha _4`$ $`=`$ $`\left[𝐄w_1(1)w_1(2)w_1(3)w_1(4)\left(𝐄w_1(1)w_1(2)\right)^2\right]`$ $`\alpha _3`$ $`=`$ $`\left[𝐄(w_1(1))^2w_1(2)w_1(3)\left(𝐄w_1(1)w_1(2)\right)^2\right]`$ $`\alpha _2`$ $`=`$ $`\left[𝐄(w_1(1)w_1(2))^2\left(𝐄w_1(1)w_1(2)\right)^2\right],`$ where the expectations are conditionally on fixed $`x_1,\mathrm{},x_n`$. All the $`\alpha _i`$ can be calculated numerically and depend only on $`n`$ (see the Appendix). Thus the result of the whole bootstrap procedure for $`N\mathrm{}`$ can be simply obtained by direct computation. ### 2.3 Expectation of $`\widehat{v_N^{}}`$ The futility of $`\widehat{v_N^{}}`$ is demonstrated by the fact that it does neither estimate what is hoped (the variance of $`\widehat{\theta }`$) nor a multiple with a fixed factor. To show this, the unconditional expectation of $`\widehat{v_N^{}}`$ is determined, see the Appendix. Since the result is not very transparent, here an approximation is given which makes it possible to characterize the quality of $`\widehat{v_N^{}}`$. Assume that the $`w_k(i)`$ are independent and Poisson distributed with parameter $`\mu =1`$; this simplifying assumption is exact in the limiting case $`n\mathrm{}`$, see above. This leads to a result which is close to the exact value for large $`n`$ and is easy to interpret. By the way, the simplification is equivalent to replacement of $`n`$ by $`n^{}`$ in each pseudo sample where $`n^{}`$ is a Poisson distributed number with mean $`\mu =n`$. In this scheme each pseudo sample consists of a random number of points. The result is $$\underset{N\mathrm{}}{lim}𝐄\widehat{v_N^{}}=𝐄\underset{N\mathrm{}}{lim}\widehat{v_N^{}}=4s_3+6s_2,$$ (4) see the Appendix, while the desired result, given by (2), is $$𝐕\widehat{\theta }=s_4+4s_3+2s_2\left(𝐄\widehat{\theta }\right)^2.$$ Remark: The formulae suggest that the bootstrap result (4) can considerably differ from the true variance of $`\widehat{\theta }`$. Nevertheless, the bootstrap procedure may make sense. In some cases $`s_4`$ converges to $`\left(𝐄\widehat{\theta }\right)^2`$ with growing $`W`$ and $`s_3`$ is small compared with $`s_2`$. Then the bootstrap result (4) may approximate three times the true variance, see . ## 3 Confidence regions for the intensity function of an inhomogeneous Poisson process The paper presents a procedure which uses bootstrap techniques to determine confidence regions for the intensity function of an inhomogeneous Poisson process. The confidence regions are estimated using a kernel estimator. The following discusses the main idea of that paper and shows that, as above, it is not necessary to carry out the bootstrap procedure. ### 3.1 Fundamentals For simplicity, the following calculations are carried out for an one-dimensional point process, but they could be easily generalized to higher-dimensional processes. Consider an inhomogeneous Poisson point process $`\mathrm{\Phi }`$ with unknown intensity function $`\lambda (x)`$ in the interval $`(0,1)`$, with points $`0<x_1x_2\mathrm{}x_n<1`$. A kernel estimator for $`\lambda (x)`$ is used as $$\widehat{\lambda }(x)=\frac{1}{h}\underset{i=1}{\overset{n}{}}K\left(\frac{xx_i}{h}\right),x(0,1),$$ where $`K`$ is a kernel function and $`h`$ bandwidth (see, for example, ). Define $$T(x)=\frac{\widehat{\lambda }(x)𝐄\widehat{\lambda }(x)}{\sqrt{\widehat{\lambda }(x)}},0<x<1,$$ and, for a given $`\alpha `$ with $`0\alpha 1`$, $$t_\alpha (x)=\underset{t^+}{\mathrm{min}}\left\{t:𝐏\left\{\left|T(x)\right|t\right\}1\alpha \right\},0<x<1.$$ Then an estimate of a confidence region for $`\lambda (x)`$ of level $`1\alpha `$ is the interval $$C(x)=[\widehat{\lambda }(x)t_\alpha (x)\sqrt{\widehat{\lambda }(x)},\widehat{\lambda }(x)+t_\alpha (x)\sqrt{\widehat{\lambda }(x)}],0<x<1,$$ where the left border is set on 0 if it is negative. ### 3.2 Bootstrap versions Since the distribution of $`T`$ is not available (because the intensity function is unknown) it is approximated by simulation of pseudo data, see . A set of pseudo data is obtained by drawing $`x_1^{},\mathrm{},x_n^{}^{}`$ by sampling randomly with replacement $`n^{}`$ times from $`\{x_1,\mathrm{},x_n\}`$, where $`n^{}`$ has a Poisson distribution with mean $`n`$ (this is method 2 in and similar to the simplified case in Section 2.3). The number of occurrences of $`x_i`$ in the $`k`$th sample is a random variable, denoted as above by $`w_k(i)`$. All the $`w_k(i)`$ are independent and Poisson distributed with mean $`\lambda =1`$ for $`i=1,\mathrm{},n`$. For given $`\alpha `$ with $`0\alpha 1`$ the bootstrap versions of the quantities defined above are $`\widehat{\lambda _k^{}}(x)`$ $`=`$ $`{\displaystyle \frac{1}{h}}{\displaystyle \underset{i=1}{\overset{n}{}}}K\left({\displaystyle \frac{xx_i}{h}}\right)w_k(i),`$ $`T_k^{}(x)`$ $`=`$ $`{\displaystyle \frac{\widehat{\lambda _k^{}}(x)\widehat{\lambda }(x)}{\sqrt{\widehat{\lambda _k^{}}(x)}}},x(0,1),`$ $`t_\alpha ^{}(x)`$ $`=`$ $`\underset{t^+}{\mathrm{min}}\left\{t:𝐏^{}\left\{\left|T^{}(x)\right|t\right\}1\alpha \right\},`$ $`C^{}(x)`$ $`=`$ $`[\widehat{\lambda }(x)t_\alpha ^{}(x)\sqrt{\widehat{\lambda }(x)},\widehat{\lambda }(x)+t_\alpha ^{}(x)\sqrt{\widehat{\lambda }(x)}],`$ where $`𝐏^{}()=𝐏(|\{x_1,\mathrm{},x_n\})`$ is the distribution conditionally on $`\{x_1,\mathrm{},x_n\}`$. The determination of $`t_\alpha ^{}(x)`$ can be carried out by simulation. However, a faster and simpler possibility uses the well-known fact that the sum of independent Poisson distributed random variables is also Poisson distributed. It is demonstrated here for the simple rectangular kernel function $$K(x)=\frac{1}{2}\mathrm{𝟏}_{[1,1]}(x).$$ For other kernels, similar calculations are possible. Let $`p(x)`$ be the number of observed points in the interval $`[xh,x+h]`$. Then its bootstrap version $`p^{}(x)`$ is a random variable which is Poisson distributed with mean $`p(x)`$. Its cumulative disribution function is denoted by $`𝐅^{}`$. Thus, for given $`\alpha `$, $`t_\alpha ^{}(x)`$ $`=`$ $`\underset{t^+}{\mathrm{min}}\left\{t:𝐏^{}\left\{\left|T_1^{}(x)\right|t\right\}1\alpha \right\}`$ $`=`$ $`\underset{t^+}{\mathrm{min}}\left\{t:𝐏^{}\left\{\left|{\displaystyle \underset{i=1}{\overset{p(x)}{}}}w_1(i)a_x(t)\right|b_x(t)\right\}1\alpha \right\}`$ $`=`$ $`\underset{t^+}{\mathrm{min}}\left\{t:𝐅^{}\left(a_x(t)+b_x(t)\right)𝐅^{}\left(a_x(t)b_x(t)\right)1\alpha \right\},`$ with $`a_x(t)`$ $`=`$ $`p(x)+ht^2,`$ $`b_x(t)`$ $`=`$ $`t\sqrt{2hp(x)+h^2t^2}`$ The calculations are of an elementary nature using that $`\widehat{\lambda _k^{}}`$ is equal to $`\frac{1}{2h}_{1=1}^{p(x)}w_k(i)`$. ### 3.3 Confidence regions without bootstrap Of course, in case of an inhomogeneous Poisson process it is easy to build confidence regions without the bootstrap methodology. Assume that the intensity function $`\lambda (x)`$ is approximately linear in the interval $`[xh,x+h]`$. Then $`2h\widehat{\lambda }(x)`$ using the rectangular kernel is Poisson distributed with mean $`2h\lambda (x)`$. (The rectangular kernel could be replaced by another kernel; then the corresponding calculations become a bit more difficult.) Therefore, known confidence regions for the Poisson parameter can be used, see for example . Thus it is easy to build the desired confidence region for $`\lambda (x)`$. This result corresponds to a general observation. If a parametric statistic problem is given, then parametric estimators lead usually to better results than bootstrap techniques. Acknowledgements: I am most grateful to Dietrich Stoyan for his great encouragement and for helpful discussions. ## Appendix Here the derivation of some equations of Section 2 is given. ### Equation (2) For $$\widehat{\theta }=\underset{x,y\mathrm{\Phi }}{\overset{}{\stackrel{}{}}}f(x,y)=\underset{x,y\mathrm{\Phi }}{\overset{}{\stackrel{}{}}}\mathrm{𝟏}_W(x)\mathrm{𝟏}_W(y)h(x,y)$$ it is $`𝐄\widehat{\theta }^2`$ $`=`$ $`𝐄\left({\displaystyle \underset{x,y\mathrm{\Phi }}{\overset{}{\stackrel{}{}}}}f(x,y)\right)^2`$ $`=`$ $`𝐄{\displaystyle \underset{w,x,y,z\mathrm{\Phi }}{\overset{}{\stackrel{}{}}}}f(w,x)f(y,z)`$ $`+4𝐄{\displaystyle \underset{x,y,z\mathrm{\Phi }}{\overset{}{\stackrel{}{}}}}f(x,y)f(x,z)`$ $`+2𝐄{\displaystyle \underset{x,y\mathrm{\Phi }}{\overset{}{\stackrel{}{}}}}\left(f(x,y)\right)^2`$ $`=`$ $`{\displaystyle \varrho ^{(4)}(x_1,x_2,x_3,x_4)f(x_1,x_2)f(x_3,x_4)\text{d}x_1\text{d}x_2\text{d}x_3\text{d}x_4}`$ $`+4{\displaystyle \varrho ^{(3)}(x_1,x_2,x_3)f(x_1,x_2)f(x_1,x_3)\text{d}x_1\text{d}x_2\text{d}x_3}`$ $`+2{\displaystyle \varrho ^{(2)}(x_1,x_2)\left(f(x_1,x_2)\right)^2\text{d}x_1\text{d}x_2}`$ $`=`$ $`s_4+4s_3+2s_2`$ with $$s_i=\varrho ^{(i)}(x_1,\mathrm{},x_i)f(x_1,x_2)f(x_{i1},x_i)\text{d}x_1\mathrm{}\text{d}x_i.$$ ### Equation (2.2) For $$\widehat{\theta _1^{}}=\underset{i,j=1}{\overset{n}{\stackrel{}{}}}f(x_i,x_j)w_1(i)w_1(j)$$ it is $`𝐄\widehat{\theta _1^{}}^2`$ $`=`$ $`𝐄\widehat{\theta _1^{}}^2𝐄\widehat{\theta _1^{}}\widehat{\theta _2^{}}`$ $`=`$ $`𝐄{\displaystyle \underset{i,j,k,l=1}{\overset{n}{\stackrel{}{}}}}f(x_i,x_j)f(x_k,x_l)w_1(i)w_1(j)w_1(k)w_1(l)`$ $`+4𝐄{\displaystyle \underset{i,j,k=1}{\overset{n}{\stackrel{}{}}}}f(x_i,x_j)f(x_i,x_k)(w_1(i))^2w_1(j)w_1(k)`$ $`+2𝐄{\displaystyle \underset{i,j=1}{\overset{n}{\stackrel{}{}}}}\left(f(x_i,x_j)\right)^2(w_1(i)w_1(j))^2`$ $`=`$ $`{\displaystyle \underset{i,j,k,l=1}{\overset{n}{\stackrel{}{}}}}f(x_i,x_j)f(x_k,x_l)𝐄w_1(i)w_1(j)w_1(k)w_1(l)`$ $`+4{\displaystyle \underset{i,j,k=1}{\overset{n}{\stackrel{}{}}}}f(x_i,x_j)f(x_i,x_k)𝐄(w_1(i))^2w_1(j)w_1(k)`$ $`+2{\displaystyle \underset{i,j=1}{\overset{n}{\stackrel{}{}}}}\left(f(x_i,x_j)\right)^2𝐄(w_1(i)w_1(j))^2`$ $`=`$ $`𝐄w_1(1)w_1(2)w_1(3)w_1(4){\displaystyle \underset{i,j,k,l=1}{\overset{n}{\stackrel{}{}}}}f(x_i,x_j)f(x_k,x_l)`$ $`+4𝐄(w_1(1))^2w_1(2)w_1(3){\displaystyle \underset{i,j,k=1}{\overset{n}{\stackrel{}{}}}}f(x_i,x_j)f(x_i,x_k)`$ $`+2𝐄(w_1(1)w_1(2))^2{\displaystyle \underset{i,j=1}{\overset{n}{\stackrel{}{}}}}\left(f(x_i,x_j)\right)^2`$ and $`𝐄\widehat{\theta _1^{}}\widehat{\theta _2^{}}`$ $`=`$ $`𝐄{\displaystyle \underset{i,j,k,l=1l}{\overset{n}{\stackrel{}{}}}}f(x_i,x_j)f(x_k,x_l)w_1(i)w_1(j)w_2(k)w_2(l)`$ $`+4𝐄{\displaystyle \underset{i,j,k=1}{\overset{n}{\stackrel{}{}}}}f(x_i,x_j)f(x_i,x_k)w_1(i)w_1(j)w_2(i)w_2(k)`$ $`+2𝐄{\displaystyle \underset{i,j=1}{\overset{n}{\stackrel{}{}}}}\left(f(x_i,x_j)\right)^2w_1(i)w_1(j)w_2(i)w_2(j)`$ $`=`$ $`\left(𝐄w_1(1)w_1(2)\right)^2{\displaystyle \underset{i,j,k,l=1}{\overset{n}{\stackrel{}{}}}}f(x_i,x_j)f(x_k,x_l)`$ $`+4\left(𝐄w_1(1)w_1(2)\right)^2{\displaystyle \underset{i,j,k=1}{\overset{n}{\stackrel{}{}}}}f(x_i,x_j)f(x_i,x_k)`$ $`+2\left(𝐄w_1(1)w_1(2)\right)^2{\displaystyle \underset{i,j=1}{\overset{n}{\stackrel{}{}}}}\left(f(x_i,x_j)\right)^2.`$ This yields $`\underset{N\mathrm{}}{lim}\widehat{v_N^{}}`$ $`=`$ $`[𝐄w_1(1)w_1(2)w_1(3)w_1(4)\left(𝐄w_1(1)w_1(2)\right)^2]`$ $`{\displaystyle \underset{w,x,y,z\mathrm{\Phi }}{\overset{}{\stackrel{}{}}}}f(w,x)f(y,z)`$ $`+4[𝐄(w_1(1))^2w_1(2)w_1(3)\left(𝐄w_1(1)w_1(2)\right)^2]`$ $`{\displaystyle \underset{x,y,z\mathrm{\Phi }}{\overset{}{\stackrel{}{}}}}f(x,y)f(x,z)`$ $`+2[𝐄(w_1(1)w_1(2))^2\left(𝐄w_1(1)w_1(2)\right)^2]`$ $`{\displaystyle \underset{x,y\mathrm{\Phi }}{\overset{}{\stackrel{}{}}}}\left(f(x,y)\right)^2`$ $`=`$ $`\alpha _4{\displaystyle \underset{i,j,k,l=1}{\overset{n}{\stackrel{}{}}}}f(x_i,x_j)f(x_k,x_l)`$ $`+4\alpha _3{\displaystyle \underset{i,j,k=1}{\overset{n}{\stackrel{}{}}}}f(x_i,x_j)f(x_i,x_k)`$ $`+2\alpha _2{\displaystyle \underset{i,j=1}{\overset{n}{\stackrel{}{}}}}\left(f(x_i,x_j)\right)^2`$ with $`\alpha _4`$ $`=`$ $`(4n^2+10n6)/n^3`$ $`\alpha _3`$ $`=`$ $`(n^37n^2+12n6)/n^3`$ (6) $`\alpha _2`$ $`=`$ $`(3n^311n^2+14n6)/n^3`$ ### Equation (4) The expectation value of $`\widehat{v^{}}`$ is $`𝐄\widehat{v}^{}`$ $`=`$ $`\alpha _4{\displaystyle \varrho ^{(4)}(x_1,x_2,x_3,x_4)f(x_1,x_2)f(x_3,x_4)\text{d}x_1\text{d}x_2\text{d}x_3\text{d}x_4}`$ $`+4\alpha _3{\displaystyle \varrho ^{(3)}(x_1,x_2,x_3)f(x_1,x_2)f(x_1,x_3)\text{d}x_1\text{d}x_2\text{d}x_3}`$ $`+2\alpha _2{\displaystyle \varrho ^{(2)}(x_1,x_2)\left(f(x_1,x_2)\right)^2\text{d}x_1\text{d}x_2}`$ $`=`$ $`s_4\alpha _4+4s_3\alpha _3+2s_2\alpha _2`$ In the limiting case ($`n\mathrm{}`$) it is $`𝐄\widehat{v^{}}`$ $`=`$ $`4s_3+6s_2`$ (see Equation (Equation ())).
warning/0002/cond-mat0002210.html
ar5iv
text
# Suppression and enhancement of the critical current in multiterminal S/N/S mesoscopic structures. ## Abstract We analyse the measured critical current $`I_{m\text{ }}`$ in a mesoscopic 4-terminal S/N/S structure. The current through the S/N interface is shown to consist not only of the Josephson component $`I_c\mathrm{sin}\phi ,`$ but also a phase-coherent part $`I_{sg}\mathrm{cos}\phi `$ of the subgap current. The current $`I_m`$ is determined by the both components $`I_c`$ and $`I_{sg},`$ and depends in a nonmonotonic way on the voltage $`V`$ between superconductors and normal reservoirs reaching a maximum at $`V\mathrm{\Delta }/e`$. The obtained theoretical resultas are in qualitative agreement with recent experimental data. Recent achievements in nanotechnology have revived interest in the study of nonequilibrium and phase-coherent phenomena in superconductor-normal metal (S/N) structures. One of the most remarkable, discovered recently , was the observation of the sign reversal of the Josephson critical current $`I_c`$ (the so-called $`\pi `$-junction) in a multi-terminal mesoscopic Nb/Au/Nb structure under nonequilibrium conditions. By passing an additional current through the N layer or, in another words, by applying a voltage $`V`$ to the normal reservoirs (see Fig.1) with respect to the superconductors, one can create a nonequlibrium electron-hole distribution, or at least one can shift this distribution with respect to the electron-hole distribution in the superconductors. Under this condition, the critical current $`I_c`$ decreases with $`V`$ and changes sign at a certain value of the applied voltage $`V`$. This effect was predicted first in Ref. where a ballistic 3-terminal structure was considered (for more details, see also Refs.). In diffusive 4-terminal S/N/S structures, the sign-reversal effect has been considered in Refs. (see also ). The sign-reversal effect and switching of the $`\pi `$-junction into a state where $`\phi =\pi `$ has much in common with an instability of an uniform superconductor with a nonequilibrium distribution function . In multi-terminal S/N/S structures one can observe not only the sign reversal effect, but also a number of other interesting phenomena. For example, the conductance of a normal wire between N reservoirs oscillates with varying phase difference $`\phi `$ (see review articles ). In addition, as shown in Refs. , the measured critical current $`I_m`$ depends on the geometry of a particular structure and instead of decreasing may also increase with increasing voltage $`V`$. In particular one can observe Josephson-like effects (plateau on the $`I_3(V_S)`$ curve, oscillations of the measured critical current $`I_m`$ in a magnetic field etc) even if the Josephson coupling between superconductors under equilibrium conditions is negligable. The reason for these effects is that the current $`I_m`$ in a multi-terminal S/N/S structure is determined not only by the Josephson component $`I_c\mathrm{sin}\phi `$, but also by the phase-dependent subgap current $`I_{sg}\mathrm{cos}\phi `$ through the S/N interface. Therefore even in the case of a small $`I_c`$, the current $`I_m`$ can be altered by varying the phase $`\varphi `$. An increase of the critical current was observed in the recent paper where a mesoscopic three-terminal S/N/S structure was studied. The authors used a third superconductor as a reservoir the electric potential of which was shifted with respect to the other two superconductors by the voltage $`V`$. The measured critical current reached its maximal value when the magnitude of $`V`$ was comparable with $`\mathrm{\Delta }.`$ At some, not too low temperatures $`T`$ the measured critical current $`I_m`$ exceeds its magnitude in the equilibrium state: $`I_m(V)>I_m(0)`$. In the present paper we show that the enhancement of the supercurrent observed in Ref. is most likely caused by the mechanism mentioned above. In Refs. the model case of gapless superconductors was considered where there is no singularity in the density-of states in superconductors at $`ϵ=\mathrm{\Delta }.`$ Here we will consider the case of ordinary superconductors with an energy gap $`\mathrm{\Delta }`$ and show that the enhancement of the critical current reaches a maximum for $`V`$ of order $`\mathrm{\Delta }.`$ The voltage dependence $`I_m`$ $`(V)`$ calculated for different temperatures is in qualitative agreement with the experimental data. We consider the structure shown in Fig.1 which differs from the structure studied experimentally. However in our opinion, this difference is not essential and allows us to give at least qualitative explanation for the phenomena observed in Ref.. First, we assume for simplicity that the structure under consideration is symmetrical, i.e. it has four terminals and not three as in the experiment. Secondly, we consider normal reservoirs in order to avoid complications which would arise in case of superconducting reservoirs (ac Josephson effects when the finite voltage is applied to the S reservoir). We also assume for simplicity that the contacts between the N wire and N reservoirs are good (the resistance of the N wire/N reservoir interface is much smaller than the resistance of the N wire), whereas the S/N interface resistance is finite (larger or less than the resistance of the N wire). We will study the diffusive case which corresponds to the experiment . In order to find the dependence of the effective critical current $`I_m(V)`$ ( the definition of $`I_m(V)`$ will be given later), we need to determine two distribution function $`f_+`$ and $`f_{}`$. Both these functions are isotropic in space. The function $`f_+`$ is related to a symmetrical part of the distribution function in the electron-hole space:$`f_+(ϵ)=1(n_{}(ϵ)+p_{}(ϵ))=1(n_{}(ϵ)+p_{}(ϵ))`$, here $`p_{}(ϵ)=1n_{}(ϵ)`$ is the hole distribution function. It determines the critical current $`I_c.`$ The function $`f`$ <sub>-</sub> describes the electron-hole imbalance and determines the electric potential and current: $`f(ϵ)=(n_{}(ϵ)p_{}(ϵ))=(n_{}(ϵ)p_{}(ϵ))`$. Equations for $`f_+`$ and $`f_{}`$ are obtained from an equation for the matrix Keldysh function $`\stackrel{}{G}`$ (see, for example For the structure shown in Fig.1 they can be written in the form $$L_x(M_{}_xf_{}(x)+J_sf_+J_{an}_xf_+(x))=r[A_{}\delta (xL_1)+\stackrel{\mathrm{\_}}{A_{}}\delta (x+L_1)].$$ (1) $$L_x(M_+_xf_+(x)+J_sf_{}+J_{an}_xf_{}(x))=r[A_+\delta (xL_1)+\stackrel{\mathrm{\_}}{A_+}\delta (x+L_1)]$$ (2) Here all the coefficients are expressed through the retarded (advanced) Green’s functions $`\stackrel{}{G^R}=G^R\widehat{\sigma }_z+\stackrel{}{F^R}`$ and are equal to $`M_\pm =(1G^RG^A(\stackrel{}{F^R}\stackrel{}{F^A})_1)/2;`$ $`J_{an}=\stackrel{}{(F^R}\stackrel{}{F^A})_z/2,J_s=(1/2)(\stackrel{}{F^R}\stackrel{}{_xF^R}\stackrel{}{F^A}\stackrel{}{_xF^A})_z,`$ $`A_{}=(\nu \nu _S+g_{1+})f_{}(g_zf_{eq}+g_{z+}f_+);`$ $`A_+=(\nu \nu +g_1)(f_+f_{eq})g_zf_{};g_{1\pm }=(1/4)[(\stackrel{}{F^R}\stackrel{}{\pm F^A})(\stackrel{}{F_S^R}\pm \stackrel{}{F_S^A})]_1;`$ $`g_{z\pm }=(1/4)[(\stackrel{}{F^R}\stackrel{}{F^A})(\stackrel{}{F_S^R}\pm \stackrel{}{F_S^A})]_z;`$ The coefficient $`r=R/R_b`$, $`R=\rho L/d`$ is the resistance of the N film per unit length in the z-direction,$`\rho `$ is the specific resistivity of the N film, $`d`$ is the thickness of the N film, $`R_b`$ is the S/N interface resistance; the functions $`\stackrel{\mathrm{\_}}{A_{}}`$ and $`\underset{+}{\overset{\mathrm{\_}}{A}}`$ coincide with $`A_{}`$ and $`A_+`$ if we make a substitution $`\phi \phi `$. We introduced above the following notations $`(\stackrel{}{F^R}\stackrel{}{F^A})_1=Tr(\stackrel{}{F^R}\stackrel{}{F^A})/2,`$ $`(\stackrel{}{F^R}\stackrel{}{F^A})_z=Tr(\widehat{\sigma }_z\stackrel{}{F^R}\stackrel{}{F^A})/2`$ etc.; $`\nu `$ and $`\nu _S`$ are the density-of states in the N film at $`x=L_1`$ and in the superconductors. The boundary conditions for $`f_{+\text{ }}`$and $`f_{}`$ are: $`f_{+\text{ }}(L)=F_{V+}`$ and $`f_{}(L)=F_V`$; the functions $`F_{V\pm }`$ are the corresponding distribution functions in the normal reservoirs: $`F_{V\pm }=[\mathrm{tanh}((ϵ+eV)\beta )\pm \mathrm{tanh}((ϵeV)\beta )]/2`$. We set the electrical potential at the superconductors equal to zero and assumed that the width of the S/N interfaces $`w`$ is small compared to $`L_{1,2}`$. Eq.(1) describes the conservation of the electric current (at a given energy). The term in the brackets on the left is the total partial current in the N wire, consisting of the quasiparticle current (the first term), the supercurrent in the interval $`(L_1,L_1)`$ (the second term) and a ”nonequilibrium supercurrent” (the third term). The coefficient $`M`$ is a quantity which is proportional to the diffusion coefficient renormalised due to proximity effect. The right hand side is the partial current through the S/N interface; the term $`(\nu \nu _S+g_{1+})f_{}`$ is the quasiparticle current above ($`\nu \nu _Sf_{}`$) and below ($`g_{1+}f_{}`$) the gap. The term ($`g_zf_{eq}+g_{z+}f_+`$) is the Josephson current in nonequilibrium conditions. Eq.(2) describes the conservation of the energy flux (at a given energy). The coefficient $`A_+`$ is zero below the gap ( complete Andreev reflection) as the difference ($`F_S^RF_S^A`$) equals zero at $`ϵ<\mathrm{\Delta }`$. The solutions of Eqs.(1)-(2) can be found exactly and expressed in terms of the retarded (advanced) Green’s functions which obey the Usadel equation. First we note that the expressions in brackets in the left hand side of Eqs.(1)-(2) in the regions ($`0,L_1`$) and ($`L_1,L`$) are equal to the constants of integration $`C_{1,2\pm }`$. The constants $`C_{1,2}`$ relate to partial currents $`J_{1,2}`$ ($`C_{1,2}=eJ_{1,2}\rho /d)`$. The partial currents $`J_{1,2}`$ are the currents per unit energy and connected with the electrical currents $`I_{1,2}`$ via the relation $$I_{1,2}=_0^{\mathrm{}}𝑑ϵJ_{1,2}(ϵ)$$ (3) Our aim is to find the current $`I_3`$ and express it in terms of the control current $`I_2`$ (or voltage $`V`$) and the phase difference $`\phi `$. We note that the distribution functions $`f_\pm (x)`$ are constants in the region $`x(0,L_1)`$ and vary in the region $`x(L_1,L)`$ reaching $`F_{V\pm }`$ at $`x=L`$. Dropping details of calculations, we present final results for limiting cases. a) Large interface resistance: $`r<<1`$. One can show that in this case $`f_+(0)(F_{V+}+f_{eq}(r_2\nu \nu _s))/(1+r_2\nu \nu _s)`$ and $`f_{}(0)F_V/(1+r_2\nu \nu _s)`$, where $`r_2=r(L_2/L).`$ The current $`I_3`$ through the S/N interface consists of three terms $$I_3(V)=I_o(V)I_c(V)\mathrm{sin}\phi +I_{sg}(V)\mathrm{cos}\phi $$ (4) Two of them ($`I_o,I_{sg}\mathrm{cos}\phi `$) are the quasiparticle currents and one ($`I_c\mathrm{sin}\phi `$) is the Josephson current. This expression shows that at a given control voltage $`V`$ and zero voltage difference between the superconductors ($`\phi `$ is constant in time) the current $`I_3`$ may vary with changing $`\phi `$ in the limits:$`I_3(V)I_o(V)I_m(V)`$. This means a plateau on the $`V_S(I_3)`$ characteristics (see ); here $`V_S=(\mathrm{}/2e)_t\phi `$ is the voltage difference between superconductors. We can write the phase-dependent part of $`I_3`$ in the form $`I_{3\phi }=I_m\mathrm{sin}(\phi +\alpha )`$, where $`I_m=\sqrt{I_c^2+I_{sg}^2}`$ is the measured critical current, $`\mathrm{cos}\alpha =I_c/I_m`$. In the considered limit of high interface resistance, we have for $`I_c`$ and $`I_{sg}`$ $`I_c(eR_b)=_0^{\mathrm{}}𝑑ϵ\{Im(F_S(F_yF_x))f_o(0)+ReF_SIm(F_yF_x)f(0)\}`$ $`I_{sg}(eR_b)=_0^{\mathrm{}}𝑑ϵg_{sg}f(0)=_0^{\mathrm{}}𝑑ϵImF_SIm(F_yF_x)f(0);`$ Here $`\theta =k_ϵL,\theta _2=k_ϵL_2,k_ϵ=\sqrt{(2iϵ+\gamma )/D},`$ $`\gamma `$ and $`D`$ is the damping rate and diffusion coefficient in the N film, $`g_{sg}=g_{1+}`$ is the normalised subgap conductance (see the expression for $`A`$). The functions $`F_y,F_x`$ are the components of the retarded Green’s function in the N film: $`\stackrel{}{F^R}=F_xi\widehat{\sigma }_x+F_yi\widehat{\sigma }_y,`$ and $`F_S=\mathrm{\Delta }/\sqrt{(ϵ+i\mathrm{\Gamma })^2\mathrm{\Delta }^2}`$ is the amplitude of the retarded Green’s function in the superconductors. If we linearise the Usadel equation, we obtain $`F_yF_x=2F_S\mathrm{sinh}^2\theta _2/(\theta \mathrm{sinh}2\theta )`$. We note that the numerical solution of the Usadel equation shows that the linearised solution is a good approximation even if $`r1`$ (at $`r=1`$ the difference between the exact and the linearized solutions at the characteristic energy $`ϵ=ϵ_L=D/L^2`$ is less than 5 percent). In Fig.2 we plot the $`V`$ dependence of $`I_c,I_{sg}`$ and $`I_m`$ where we see that the real critical current $`I_c`$ decreases and changes sign with increasing $`V`$, whereas the measured critical current $`I_m`$ first decreases and then increases. Its maximum may exceed $`I_c(0).`$ The reason for such a behaviour of $`I_m`$ is the third term on the right side in Eq.(5) which describes a contribution of the phase-dependent part of the subgap quasiparticle current $`I_{sg}`$ through the S/N interface to the current $`I_3`$. The current $`I_{sg}`$ is zero at $`V=0`$ and increases with $`V;`$ this current leads to a low and high temperature peak in the conductance. Its phase dependence was measured in Ref. and discussed in many papers (see review articles ). One can see from Fig.2 that due to the current $`I_{sg}`$ the measured critical current $`I_m`$ remains finite when $`I_c(V)`$ turns to zero. Fig.3 shows the dependence of the measured critical current $`I_m`$ on the control voltage $`V`$ for different temperatures. Our results qualitatively agree with the experimental data of Ref. ; that is, the current $`I_m`$ reaches a maximum at $`V\mathrm{\Delta }/e`$ and this maximum exceeds the equilibrium value of $`I_c=I_c(0)`$ when the temperature is not too low. One can see, in agreement with the experimental results of Ref. , the maximal value of $`I_m`$ depends on the temperature much weaker than $`I_c`$. Although it is difficult to carry out a quantitative comparision between theory and experiment because in the experiment the width $`w`$ and the interface resistance $`R_{b\text{ }}`$were comparable with $`L_{1,2}`$ and $`R`$ respectively, and a superconducting reservoir was used instead of a normal one (therefore, strictly speaking, one must take into account ac Josephson effects). An important point to note is that our results do not mean that the sign reversal of the real critical current $`I_c`$ can not be identified directly. Consider for example a fork-shape circuit; this means that two vertical superconducting leads in Fig.1 are attached to a T-shape (inverted) superconducting lead. Analysing the stability of the state with negative $`I_c`$, one can easily show that the state with $`\phi =0`$ is unstable with respect to fluctuations of $`\phi `$ and the system switches to a state with a circulating current. Indeed, taking into account the fluctuating voltage at the superconductor $`V_S=\mathrm{}_t\phi /2e,`$ we replace $`V`$ in Eq.(4) by $`VV_S`$. We then write down the equation for the current $`\overline{I_3}`$ in the lead attached to the left superconductor; this equation coincides with Eq.(4) if $`\phi `$ is replaced by $`\phi `$. Subtracting these equations for $`I_3`$ and $`\overline{I_3}`$, we arrive at the equation for a circulating current $`I_{cir}=(I_3`$ $`\overline{I_3})/2`$: $$I_{cir}=I_c(V)\mathrm{sin}\phi +V_S(R_0+R_{sg}\mathrm{cos}\phi ).$$ (5) where $`R_0=I_o/V`$ and $`R_{sg}=I_{sg}/V`$. Fluctuations of $`I_{cir}`$ lead to a magnetic flux $`\mathrm{\Phi }=I_{cir}L/c`$ in the loop which is related to $`\phi `$: $`\mathrm{\Phi }=\mathrm{\Phi }_o\phi `$, here $`\mathrm{\Phi }_o`$ is the magnetic flux quantum and we assumed the absence of flux in the ground state. We find readily from Eq.(5) that the state with $`\phi =0`$ is unstable if $`I_c(V)<0`$ and $`I_c(V)>c\mathrm{\Phi }_o/L,`$ where $`L`$ is the loop inductance . b) Small interface resistance. One can show that in this case the function $`f_{}(0)`$ is zero in the main approximation with respect to the parameter $`(r\theta )^1`$ (this means that the condition $`r^2>>\mathrm{\Delta }/ϵ_L`$ should be satisfied; here $`ϵ_L=D/L^2`$ is the Thouless energy). The function $`f_+`$ , which determines the Josephson current, in the main approximation is equal to $`F_{V+}`$ at $`ϵ<\mathrm{\Delta }`$ and to $`f_{eq}`$ at $`ϵ>\mathrm{\Delta }`$. Therefore the dependence $`I_c(V)`$ is similar to that found numerically in Ref. for another geometry (for small interface resistance); that is, the critical current $`I_c(V)`$ changes sign with increasing $`V`$ at $`V`$ of the order of the Thouless energy. As to the current $`I_2`$, it does not depend on the phase difference in the main approximation. Indeed, in order to find $`I_2`$ we need to solve the Usadel equation in the region $`x(L_1,L)`$ with boundary condition which is reduced to $`\stackrel{}{G^R}=\stackrel{}{G_S^R}`$. Making the gauge transformation $`\stackrel{}{G_S^R}`$ $`\stackrel{}{S}`$ $`\stackrel{}{G_S^R}`$ $`\stackrel{}{S^+}`$, we can exclude the phase (here $`\stackrel{}{S}=\mathrm{cos}(\phi /2)+i\widehat{\sigma }_z\mathrm{sin}(\phi /2)`$). Therefore in the main approximation the third term in Eq.(4) is zero. In conclusion, we have studied the dependence of the measured critical current $`I_m`$ on the voltage $`V`$ between normal reservoirs and superconductors in a 4-terminal S/N mesoscopic structure. The current $`I_m`$ is shown to decrease with increasing $`V`$, then to increase reaching a maximum at $`V\mathrm{\Delta }/e.`$ Our results qualitatively agree with experimental data obtained in the recent paper . We are grateful to the EPSRC for their financial support.
warning/0002/astro-ph0002342.html
ar5iv
text
# The discovery of a low mass, pre-main-sequence stellar association around 𝛾-Velorum ## 1 Introduction As little as 15 years ago it was believed that high mass stars formed mainly in OB associations and that low mass stars formed mainly in T associations. In the last few years it has become increasingly recognized that the majority of low mass stars in the solar neighbourhood are in fact likely to have formed in OB associations. The mass function in OB associations may match popular log-normal parameterisations of the general field population such as those proposed by Miller & Scalo (1979), with many low mass stars formed for every O/B star. The main reason for this shift in paradigm is the discovery of numerous low mass pre-main sequence (PMS) stars in young OB associations by virtue of their high levels of X-ray activity (see for example Walter et al. 1994, 1999; Naylor & Fabian 1999, Preibisch & Zinnecker 1999 and references therein). Recently, some evidence has been found that these X-ray selected groups can be quite small, concentrated around just one or two high mass stars, which are themselves part of larger associations. Examples include $`\beta `$ Cru (Feigelson & Lawson 1997), $`\sigma `$ Ori (Walter, Wolk & Sherry 1998) and $`\eta `$ Cha (Mamajek, Lawson & Feigelson 1999). The investigation of these low mass PMS stars is of prime importance in establishing the form of the initial mass function. Although OB associations are by definition unbound, they are young enough that dynamical effects such as mass segregation and preferential evaporation will not have occurred; and they are usually free of the heavy, variable extinction that plagues observations of active star forming regions. If the majority of low mass stars are born in OB associations it is crucial to establish the influence that high mass neighbours have on the formation and evolution of their lower mass siblings. The winds and ionizing radiation of hot stars could influence the mass function and circumstellar disc lifetimes of the lower mass stars, with implications for angular momentum evolution and planet formation. These ideas have gained currency with the discovery of evaporating discs around PMS stars in the Orion nebula (McCaughrean & O’Dell 1996) and theoretical studies showing that discs could be ionized and evaporated by the UV radiation fields of O stars (Johnstone, Hollenbach & Bally 1998). In more extreme circumstances, nearby supernova explosions have been invoked both as a means of terminating low mass star formation (Walter et al. 1994) or triggering it (Preibisch & Zinnecker 1999). In this paper we report the discovery, by X-ray selection, of a low mass stellar population that seems likely to be associated with the nearest example of a Wolf-Rayet (WR) star, $`\gamma ^2`$ Velorum (HD 68273, HIP 39953, WR11). Like about half of the $`200`$ galactic WR stars known, it is a binary system (WC8+O8) with an orbital period of 78.5 days and a massive, interacting stellar wind. There is currently some controversy concerning the distance to $`\gamma ^2`$ Vel, which impacts upon the deduced luminosities, masses and mass loss rates from the system. It is important to get these parameters right because, as the nearest WR, $`\gamma ^2`$ Vel is an extreme test of stellar evolution models and calibrates the absolute magnitudes of WR stars. The Hipparcos parallax yields a distance of $`258_{31}^{+41}`$ pc to $`\gamma ^2`$ Vel (Schaerer, Schmutz & Grenon 1997, van der Hucht et al. 1997), in marked contrast to previous distance estimates which place it at 350-450 pc. The larger distance is in better agreement with the mean distance to the Vela OB2 association ($`410\pm 12`$ pc), of which $`\gamma ^2`$ Vel is the most massive proper-motion member (de Zeeuw et al. 1999). The presence of a low mass stellar association in the extreme environment around $`\gamma ^2`$ Velorum offers an excellent empirical test of the possible influence of winds and ionizing radiation on low mass stars and their discs. Additionally it gives us a chance to measure the distance to $`\gamma ^2`$ Vel by matching PMS isochrones to the low mass stars. ## 2 X-Ray observations We suspected the presence of a low mass association around $`\gamma ^2`$ Vel from the large surrounding population of X-ray point sources seen in ROSAT images, taken for a programme to investigate its interacting stellar winds (see Willis, Schild & Stevens 1995 and Fig.1). The X-ray observations of the WR were retrieved from the ROSAT public archive and consisted of 10 Position Sensitive Proportional Counter (PSPC) datasets and 2 datasets taken with the High Resolution Imager. In this initial paper we discuss only the more sensitive PSPC results, which contain the vast majority of the X-ray point sources. We used the Starlink-distributed asterix data reduction package in our X-ray analysis (Allan & Vallance 1995). The 10 PSPC datasets were sorted into 1$`{}_{}{}^{}\times 1^{}`$ images, selecting pulse height channels 11 to 240 (approximately 0.1-2.4 keV photons) and excluding times with anomalously high background rates. The images were centered on $`\gamma ^2`$ Vel (the brightest source in each dataset) and then summed. The resulting image has dimensions of $`720\times 720`$ 5 arcsecond pixels and an effective on-axis exposure time of 25.3 ks. We used the Point Source Searching (PSS - Allan 1992) algorithm to search for sources by Cash-statistic maximisation. By assuming the background to be zero we obtained a preliminary list of 104 X-ray sources which were masked out of the image. The masked image was then patched and smoothed with a 75 arcsec FWHM gaussian to create a background map. We then executed the PSS algorithm again and found 109 sources above a pseudo-gaussian significance level of 4.5$`\sigma `$, which corresponds roughly to 1 spurious detection in the X-ray image. The positions of the X-ray sources were corrected for errors in the satellite aspect solution by comparing the optical and X-ray positions of $`\gamma ^2`$ Vel and four other bright stars from the CDS SIMBAD database. We applied shifts of 4.5 arcsec in RA and 13.8 arcsec in Dec to the X-ray positions. ## 3 Optical photometry CCD photometry of the X-ray field was obtained on 8 February 1999 with the 0.9-m telescope at the Cerro Tololo Inter-American Observatory. A Tek 2048x2048 CCD was used to give a 13.5x13.5 arcmin<sup>2</sup> field of view. Eight overlapping fields around $`\gamma ^2`$-Vel were surveyed in BVI with short (20,10,10s) and long (200,100,100s) exposures, together with five fields from Landolt (1992) to determine zeropoints, colour terms and extinction coefficients. Several standards with $`VI_c>2.5`$ were observed. Figure 1 illustrates the location of the fields around $`\gamma ^2`$ Vel. The external accuracy of our photometry was determined to be around 0.02 mag on the $`BVI_\mathrm{c}`$ system. Photometry was performed for each of the eight fields using an optimal aperture algorithm (Naylor 1998; Totten et al. in preparation), and the results combined to give an optical catalogue of sources. Astrometry was performed by comparison with star positions in version A2.0 of the USNO catalogue (Monet 1998) and we estimate positional accuracies of around 0.3 arcsec. The final catalogue contains 20617 individual objects with their magnitudes, colours and positions. Figure 2a shows the $`V`$ versus $`VI_\mathrm{c}`$ colour-magnitude diagram (CMD). The location of a possible PMS low-mass star association is clearly visible above the bulk of the background contamination. To this optical catalogue, we added the positions of bright ($`V<11`$) stars found in this region from the SIMBAD database. To establish an appropriate cross-correlation radius to use between the X-ray and optical source lists we modelled the cumulative number of X-ray sources that were correlated with an optical source with $`V<19`$ (see Jeffries, Thurston & Pye 1997 for details). Assuming a uniform spread of optical sources, we determined that there were 77 correlations (from 83 X-ray sources inside the CCD survey) within 10 arcsec of X-ray positions, that 66 of these would be true counterparts to X-ray sources, 11 would be spurious correlations and that the $`1\sigma `$ X-ray error circle was 3.7 arcsecs. Figure 2b shows 75 sources that have an optical counterpart within 10 arcsecs (another two are bright stars without $`VI`$ colours). Clearly the X-ray emitting population coincides with the proposed PMS population in the CMD. Indeed, if we were to consider just a subset of the optical catalogue consisting of a broad strip containing all these PMS sources, we would only expect 1 of these correlations to be spurious. We have calculated the X-ray properties of this population and these along with the HRI observations will be reported in a subsequent publication. Briefly, the X-ray to bolometric flux ratio of these objects lies in the range $`10^5`$ to $`10^2`$, broadly what we would expect from a population of young PMS stars. The cut-off in the PMS X-ray correlations at $`V18`$ is almost certainly due to the X-ray sensitivity. To be detected in X-rays, fainter objects would have to have higher than feasible X-ray to bolometric flux ratios. However, it is clear that the PMS sequence we have found extends down to the limits of our optical survey at $`V20.5`$. ## 4 Discussion The appearance of Fig.2 should leave the reader in no doubt that we have found a young and exceptionally rich population of low mass active stars in the direction of $`\gamma ^2`$ Vel. The central question to be answered is whether these sources are physically close to $`\gamma ^2`$ Vel and/or whether they are background members of the Vela OB2 association. We can tackle this problem in a number of ways. ### 4.1 Are the PMS stars and $`\gamma ^2`$ Vel aligned? Figure 3a shows the spatial distribution of PMS stars selected from the CMD in a strip enclosing the bulk of the X-ray sources (marked with a dashed box in Fig.2). This box was chosen to avoid background contamination. We determined the radial distribution of these stars, centred on $`\gamma ^2`$ Vel. The distribution is normalized using the radial distribution of background stars with a similar $`V`$ magnitude range, but a colour range of 0.8$`<VI_\mathrm{c}<`$1.7, under the assumption that the background stars are uniformly distributed. The resulting radial distribution is shown in Fig. 3b, which exhibits a small but significant (at the $`3\sigma `$ level) peak within 5 arcmin of the centre of the CCD survey and $`\gamma ^2`$ Vel. The point closest to $`\gamma ^2`$ Vel is missing because the 30 arcsec region immediately surrounding $`\gamma ^2`$ Vel is swamped by its light and no accurate photometry was obtained there. In a similar fashion we can show that the X-ray sources are also marginally concentrated toward the centre of the field even after correction for the PSPC vignetting function. There is thus some evidence that the PMS stars and $`\gamma ^2`$ Vel are spatially correlated, although this does not rule out a chance alignment of $`\gamma ^2`$ Vel with a background cluster of low mass stars in Vela OB2. In particular, it is possible that any concentration we see could be associated with $`\gamma ^1`$ Vel, a B2III, single lined spectroscopic binary which is a common proper motion companion to $`\gamma ^2`$ Vel. $`\gamma ^1`$ Vel is also a likely member of the Vela OB2 association, has a spectroscopic parallax of $`450`$ pc and is located only 41 arcsecs at a PA of 220 from $`\gamma ^2`$ Vel. ### 4.2 Are the PMS stars and $`\gamma ^2`$ Vel at the same distance? The isochrones plotted in Fig.2 come from the D’Antona & Mazzitelli (1997) low mass evolutionary models. We have converted from bolometric luminosity and effective temperature to magnitude and colour using empirical bolometric corrections as a function of colour and a colour-effective temperature relationship derived by forcing the well studied low mass stars in the Pleiades cluster to fit a 125 Myr isochrone (see Jeffries & Tolley 1998). The assumed distance and reddening are those appropriate for the Vela OB2 association of 410 pc and $`E(VI)0.06`$. It is clear that if the PMS stars are at the mean distance of the Vela OB2 association they appear to be about $`4\pm 2`$ Myr old (taking into account the likely presence of unresolved binary systems). If however, the PMS association were at the Hipparcos distance to $`\gamma ^2`$ Vel, the isochrones would be shifted upwards by 1 magnitude. In that case we would deduce that the PMS population had an age $`>10`$ Myr. The age of $`\gamma ^2`$ Vel, based upon the mass of the O star ($`30`$M) at the Hipparcos distance, is less than 5 Myr (Schaerer et al. 1997; de Marco & Schmutz 1999). Thus if the Hipparcos distance is adopted, the PMS stars and $`\gamma ^2`$ Vel cannot be at the same distance and coeval. However, if $`\gamma ^2`$ Vel were at the mean distance of the Vela OB2 association, it would be more massive (see below), slightly younger and could easily be coeval with the PMS population. An age range of 2-6 Myr would be compatible with distances between 490 and 360 pc. ### 4.3 How far away is $`\gamma ^2`$ Vel? The evidence that $`\gamma ^2`$ Vel is as close as 258 pc from the Hipparcos parallax should be treated with some caution. de Zeeuw et al. (1999) examine the high mass membership of the Vela OB2 association on the basis of both proper motions and parallaxes. The association is well defined by proper motions and has an angular radius of about 6 degrees. The mean parallax from 93 members corresponds to $`410\pm 12`$ pc. The parallax dispersion can be quite well modelled as a gaussian with a $`\sigma =0.68`$ mas, such that $`95\%`$ of the members appear to lie between 260 pc and 900 pc. The dispersion is reasonably consistent with the errors on the individual points and not inconsistent with the idea that all the stars are at nearly the same distance. Indeed, if we were to assume that the front to back size of Vela OB2 were similar to its diameter on the sky, then all the stars should be contained within $`\pm 40`$ pc. We would therefore interpret the parallax to $`\gamma ^2`$ Vel simply as a $`2\sigma `$ deviation and that its distance was $`410\pm 40`$ pc. The companionship of $`\gamma ^1`$ Vel is also evidence for a distance closer to the mean Vela OB2 distance. Although these objects have a common proper motion, they were too close together on the sky for Hipparcos to obtain independent parallaxes. The distance to $`\gamma ^1`$ Vel from its spectral type and photometry is almost certainly 400-500 pc (e.g. Abt et al. 1976; Hernández & Sahade 1980). The idea that $`\gamma ^2`$ Vel is a foreground object randomly placed within 1 arcminute of another bright member of the Vela OB2 association has a probability of only $`10^3`$. There is also reasonable evidence that the radial velocities of the two systems are similar. Hernández & Sahade (1980) quote $`9.7\pm 1.0`$ km s<sup>-1</sup> for $`\gamma ^1`$ Vel and Niemelä & Sahade (1980) give $`12\pm 1`$ km s<sup>-1</sup> for $`\gamma ^2`$ Vel, although it is likely that the accuracy of the latter result is exaggerated (Schmutz et al. 1997). ### 4.4 Can the PMS stars be at a range of distances? Preibisch & Zinnecker (1999) observed a very similar PMS CMD in the Upper Sco OB association, with a similar vertical scatter of about $`\pm 0.6`$ mag about the isochrones. They showed that if one takes into account unresolved binaries, photometric errors and allowed a $`10\%`$ range in distance, that the scatter around the PMS isochrones was exactly as expected for a coeval population at $`5`$ Myr. We therefore similarly conclude that the CMD in Fig. 2 shows no evidence for a large spread in either the age or distance of the PMS population we have found. In particular, we can rule out any distance modulus spread in a coeval population that is larger than a few tenths of a magnitude, or any age spread in a co-spatial population of more than a Myr or so (for a mean age of $`4`$ Myr). Thus unless there is a conspiracy to place older stars closer to us, the PMS association seems likely to have a relatively narrow spread around an age and distance that are incompatible with the deduced age for $`\gamma ^2`$ Vel and its Hipparcos distance. ### 4.5 Is a larger distance to $`\gamma ^2`$ Vel consistent with its physical properties? Our findings challenge the conclusions of several recent papers which use the Hipparcos parallax of $`\gamma ^2`$ Vel and its error to derive: the absolute magnitude of the system and its components; system masses from the interferometric binary separation of Hanbury-Brown (1970); the O star luminosity, mass and age from stellar evolution models and hence the orbital inclination and further mass estimates from radial velocity curves (see van der Hucht et al. 1997; Schaerer et al. 1997, Schmutz et al. 1997, de Marco & Schmutz 1999). A distance as large as 410 pc for $`\gamma ^2`$ Vel significantly changes the system parameters deduced in these papers. The system luminosity increases by a factor 2.5. The effective temperature and luminosity deduced for the O star would then give it a mass $`>40`$$`\text{M}_{}`$ and an age $`<3`$ Myr, compared with the values of 30 $`\text{M}_{}`$ and 3.6 Myr quoted by de Marco & Schmutz (1999). At this larger distance, an age of 2-3 Myr could be compatible with the low mass PMS stars we have found. The absolute magnitude of the O star would decrease to $`6.0\pm 0.3`$ (van der Hucht et al. 1997), which argues for a supergiant rather than a giant classification. Van der Hucht et al. comment that this would be in better agreement with published spectra of Conti & Smith (1972) and Niemelä & Sahade (1980). As the mass ratio from the radial velocity curves is fixed, the WR mass increases by a similar fraction. The total system mass, based on a binary separation of $`4.3\pm 0.5`$ mas, increases from $`30\pm 10`$$`\text{M}_{}`$ to $`120\pm 40`$$`\text{M}_{}`$ (now comfortably exceeding the minimum mass from radial velocity curves – Niemelä & Sahade 1980) and the binary inclination is reduced to around $`50^{}`$ to explain the radial velocity curves. ## 5 Conclusions From our discussion there seem to be two possible scenarios. (1) That the PMS stars are approximately at the same distance and age as $`\gamma ^2`$ Vel, and that this distance places $`\gamma ^2`$ Vel within the Vela OB2 association at 360-490 pc. (2) That the PMS stars are part of the Vela OB2 association, possibly surrounding $`\gamma ^1`$ Vel, but that $`\gamma ^2`$ Vel is an isolated foreground object with no surrounding low mass stars at a similar age. We believe that (1) is far more plausible than (2) because of the dispersion in the Vela OB2 Hipparcos parallaxes and the likely association of $`\gamma ^1`$ and $`\gamma ^2`$ Vel. Recently, the idea that $`\gamma ^2`$ Vel could form in isolation without accompanying low mass stars has also been challenged by the near IR detection of a K-type PMS companion only 4.7 arcsec distant (Tokovinin et al. 1999). If the low mass PMS stars we have found are truly in the vicinity of $`\gamma ^2`$ Vel, they represent an exciting opportunity to explore the influence of adjacent high mass loss stars and ionizing UV radiation fields on the mass function and circumstellar disc lifetimes of low mass stars. It will be interesting to compare the frequencies of T-Tauri discs around these stars with the frequencies found in T associations and OB associations with similar ages. The PMS stars in Fig.2 have masses, found from the D’Antona & Mazzitelli (1997) models, down to (an age dependent) mass of about 0.15 $`\text{M}_{}`$. The mass function will be addressed when we have a better census of the association membership. ## ACKNOWLEDGMENTS This research has made use of ROSAT data obtained from the Leicester Database Archive Service at the Department of Physics and Astronomy, Leicester University, UK. The Cerro Tololo Interamerican Observatory is operated by the Association of Universities for Research in Astronomy, Inc., under contract to the US National Science Foundation. TN was supported by a UK Particle and Physics and Astronomy Research Council (PPARC) Advanced Fellowship. SH was supported by a Nuffield Foundation Undergraduate Research Bursary (NUF-URB98). EJT is a PPARC postdoctoral research associate. MK was supported by an undergraduate research bursary from Keele University. Computational work for this paper was performed on the Keele node of the Starlink network funded by PPARC.
warning/0002/cond-mat0002262.html
ar5iv
text
# 1 First and second-order terms in the diagram expansion of Σ^𝑅. Quantum, Multi-Body Effects and Nuclear Reaction Rates in Plasmas V. I. Savchenko <sup>+</sup> Princeton University, PPPL, Forrestal Campus, Princeton, N. J. 08543 ## Abstract Detailed calculations of the contribution from off-shell effects to the quasiclassical tunneling of fusing particles are provided. It is shown that these effects change the Gamow rates of certain nuclear reactions in dense plasma by several orders of magnitude. Methods of field theory do not straightforwardly apply to systems at finite temperature. Naive perturbation theory fails in the infrared region of energy at high temperature but can be fixed by the partial HTL resummation . This approach was generalized in , and it was shown that fields without dispersion relation are necessary for consistent formulation of the theory. The particle distribution over momenta, $`f(𝐩)`$, acquires non-Maxwellian tail in thermodynamic equilibrium due to quantum effects, while the particle distribution over energies, $`\stackrel{~}{f}(\omega )`$, remains Maxwellian. An interesting suggestion was made in Ref. , that the rates of various processes, including nuclear reactions, may be significantly increased by the quantum tail of the momentum distribution. Non-linear integral equation for the rate of inelastic processes as well as a final formula for the ionization rate was obtained in . The formula for the nuclear reaction rates was found in under assumption, that it is correct to substitute the particle momentum in the form $`ϵ_𝐩=p^2/2m`$ into the known quasiclassical cross-section $`\sigma (\omega )`$ and then average it over the particle distribution over momenta $`f(𝐩)`$ rather than the distribution over energies $`\stackrel{~}{f}(\omega )`$. This procedure becomes even more unclear , if we take into account the fact, that the momentum, $`𝐩`$, and energy, $`\omega `$, of a colliding particle are independent variables not connected by the usual dispersion relation $`\delta (\omega 𝐩^2/2m)`$ . Note, that the averaging of $`\sigma (\omega )`$ over the distribution over energies $`\stackrel{~}{f}(\omega )`$, which is Maxwellian, would give the usual reaction rates rather than the rates found in , which are accelerated by many orders of magnitude in certain regimes. One can argue that averaging over momentum may not be completely wrong, because the dependence of the scattering amplitude on energy can be neglected in the gaseous approximation . However, the connection between the scattering amplitude and the rate of nuclear reactions is not clear. Neither is it obvious that this argument can straightforwardly lead to correct answer in the case of essentially non-linear processes of nuclear reactions. It was shown that non-linearities can be very subtle, subject to various non-trivial cancellations . It is therefore important to find the nuclear reaction rates from the first principles. In this paper we rigorously solve the quasiclassical problem of tunneling through the potential barrier. Our result demonstrates that the cancellations of the type found in do not play a role in the case under consideration. We discuss this question in more details in the future publication . The tunneling particles undergo simultaneous collisions with other particles of the plasma maintained in thermodynamic equilibrium. The plasma is assumed to be fully ionized. We use the Green–function technique, introduced by Keldysh and Korenman , and do not rely on any ad hoc assumptions about any averaging procedure. Our final result for the nuclear reaction rate agrees with that postulated in , apart from the coefficient depending on the masses of colliding and reacting particles. A number of different off-shell processes in plasmas were investigated within a framework of thermal field theory. Emission rate of soft photons from hot plasmas was calculated in . Using full spectral functions allowed to eliminate unphysical divergences and extend the range of validity of the results obtained by HTL resummation by taking into account off-shell particle propagation. Landau damping was shown to lead to anomalous relaxation of the particle distribution function . Off-shell effects due to particle collisions were treated within Born approximation in . The relaxation rate, calculated with the help of full spectral functions, is slower approximately by a factor of 2 as compared to the rates calculated by solving the Boltzman equation . This relaxation rate is, however, an integral quantity, insensitive to the details of changes of the kinetic Green function due to inclusion of off-shell effects. In contrast to the relaxation rate, inherent non-linearity of the tunneling problem makes the average nuclear reaction rate very sensitive to the behaviour of this function at large momenta. This explaines why we should expect off-shell effects to lead to significant corrections to the nuclear reaction rates. Indeed, we find that certain reactions are accelerated by many orders of magnitude as compared to the calculations, performed using spectral functions with zero width . We solve the tunneling problem, neglecting Salpeter corrections due to screening, and assume that, the nuclear transformations are unaffected by the medium. Kinetic properties of the particles are described by the Green function $`G^+(X_1,X_2)`$, where $`X=(t,𝐱)`$ . Equations for $`G^+`$ include the interaction between the particles as well as their interaction with the external field. Therefore, the rate of change of this function, evaluated at any point after the barrier, will give the tunneling rate $`K(𝐱,t)`$ for the particles colliding simultaneously with other particles : $`K(𝐱,t)=\left[(_{t_1}_{t_2})G^+(t_1𝐱_\mathrm{𝟏};t_2𝐱_\mathrm{𝟐})\right]_{X_1X_2=X}`$ (1) Now we will proceed with solution of the kinetic equations for $`G^+(X_1,X_2)`$. They can be written as Dyson equations for $`G_{12}^{\alpha \beta }`$ : $$G_{12}^{\alpha \beta }=G_{12}^{(0)\alpha \beta }+G_{14}^{(0)\alpha \gamma }\mathrm{\Sigma }_{43}^{\gamma \delta }G_{32}^{\delta \beta }d^4X_4d^4X_3,$$ (2) where we use subscripts $`\mathrm{1..4}`$ to denote dependence on $`X_1..X_4`$. Dependence on $`G_{12}^{(0)\alpha \beta }`$ can be eliminated by acting on both sides of the Eq. (2) with the operator $`\widehat{G}_1^1=i{\displaystyle \frac{}{t_1}}+{\displaystyle \frac{\mathrm{\Delta }_1}{2m}}U(z_1)i{\displaystyle \frac{}{t_1}}+{\displaystyle \frac{\mathrm{\Delta }_1^{}}{2m}}+\widehat{L}(z_1),`$ where $`\mathrm{\Delta }_1^{}_{x_1}^2+_{y_1}^2`$. Since we are interested in the steady state solution, one can see that, the properties of this operator can be fully accounted for if we introduce the function $`g(z,k)`$, such that $$\left[\frac{1}{2m}\frac{^2}{z^2}U(z)\right]g(z,k)=ϵ_kg(z,k),$$ (3) where $`ϵ_k=k^2/2m`$. Since the operator $`\widehat{L}`$ is Hermitian, we will use the property of its eigen-functions $`g(z,k)`$: $$_{\mathrm{}}^{\mathrm{}}g(z,k_1)g^{}(z,k_2)𝑑z=2\pi \delta (k_1k_2)$$ (4) Now let us solve for $`G_{12}^{(0)+}`$, which satisfies $$\widehat{G}_1^1G_{12}^{(0)+}=0.$$ (5) It is easier to find $`G_{12}^{(0)+}`$ by using its definition $$G_{12}^{(0)+}=i\widehat{\psi }_2^{}\widehat{\psi }_1,$$ (6) where $`\widehat{\psi }^{},\widehat{\psi }`$ are the Heisenberg operators of creation and annihilation of the particles at a point $`X`$, and $`<..>`$ means quantum and statistical averaging. Since $`\widehat{\psi }`$ evolves according to $`\widehat{G}^1\widehat{\psi }=0`$, one can see, that it is equal to $$\widehat{\psi }(X)=\underset{𝐪_{}k}{}\widehat{a}_{𝐪_{}k}g(z,k)e^{iϵ_{𝐪_{}k}t+i𝐪_{}𝐱^{}}$$ (7) Here $`\widehat{a}_{𝐪_{}k}`$ is the annihilation operator of the particle with momentum $`(𝐪_{},k)`$, and $`ϵ_{𝐪_{}k}𝐪_{}^2/2m+k^2/2m`$. We substitute $`\psi (X)`$ from (7) into Eq. (6) and obtain $`G_{\omega 𝐪_{}}^{(0)+}(z_1,z_2)=2\pi i{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{dk}{2\pi }}`$ $$\delta (\omega ϵ_{𝐪_{}k}+\mu )n(𝐪_{}k)g(z_1,k)g^{}(z_2,k),$$ (8) where we denote $`a_{𝐪_{}k}^{}a_{𝐪_{}k}n(𝐪_{}k)`$; $`\mu `$ is the chemical potential. It is easy to see that, in thermodynamic equilibrium we are considering, the function $`n(𝐪_{}k)`$ is equal to Fermi distribution $`n_F(ϵ_{𝐪_{}k})`$. To convince ourselves we take $`G_{\omega 𝐪_{}}^{(0)+}(z_1,z_2)`$ at such values of $`z_1,z_2`$, when $`U(z)`$ is negligible and $`g(z,k)exp(ikz)`$. One more Fourier transform $`exp[iq(z_1z_2)]d(z_1z_2)`$ will give $`G_{\omega 𝐪}^{(0)+}`$, which has to coincide with the known expression $`G_{\omega 𝐪}^{(0)+}=2\pi i\delta (\omega ϵ_𝐪+\mu )n_F(ϵ_𝐪)`$. Before proceeding further with solving for $`G^+(X_1,X_2)`$ we first find $`K^{(0)}`$, which is the tunneling rate obtained by using $`G_{12}^{(0)+}`$ in Eq. (1). Since we are interested in $`K^{(0)}`$ behind the barrier, where $`U(z)=0`$, we can exchange $`_{t_1}`$ for $`i\mathrm{\Delta }_1/2m`$ as is clear from the form of the operator $`\widehat{G}_1^1`$ and Eq. (5). Then the formula for the rate $`K^{(0)}`$ takes the form: $$K^{(0)}(z)=\frac{i}{2m}\frac{d\omega }{2\pi }\frac{d^2𝐪_{}}{(2\pi )^2}\left[\frac{^2}{\zeta ^2}G_{\omega 𝐪_{}}^{(0)+}(z,\zeta )\right]_{\zeta =0},$$ (9) where $`z=(z_1+z_2)/2`$, $`\zeta =z_1z_2`$. In deriving (9) we expressed $`G_{12}^{(0)+}`$ through its Fourier transform $`G_{\omega 𝐪_{}}^{(0)+}(z,\zeta )`$ with respect to “fast” space $`𝐫_{}=𝐫_1^{}𝐫_2^{}`$ and time $`\tau =t_1t_2`$ variables and took the limit $`𝐫,\tau 0`$ as prescribed by Eq. (1). Finally we need to know the function $`g(z,k)`$ in the region behind the barrier. We will use the standard quasiclassical expression $$g(z,k)=\frac{C_k}{\stackrel{~}{q}(z,k)^{1/2}}exp\left[i_z_{}^z\stackrel{~}{q}(z^{},k)𝑑z^{}\right]$$ (10) $$C_k=A_kexp\left[_{z_a}^{z_b(k)}\stackrel{~}{q}(z^{},k)𝑑z^{}\right]$$ (11) $$\stackrel{~}{q}(z,k)=\sqrt{2m(ϵ_kU(z))}$$ (12) The square of the exponential factor in Eq. (11) is the tunneling coefficient, $`\stackrel{~}{W}(k)`$ . We multiply it by a factor $`A_k=(S(ϵ_k)/ϵ_k)^{1/2}`$, which takes into account nuclear physics effects, not considered here. $`S(ϵ_k)`$ is the astrophysical factor . To find $`K^{(0)}`$ from Eq. (9) we substitute Eqs. (10)-(12) in Eq. (8), use $`z_1=z+\zeta /2`$, $`z_2=z\zeta /2`$ and take $`_\zeta ^2`$ derivative in the quasiclassical sense keeping only zero-th order terms. The result of the differentiation is $$\left[\frac{^2}{\zeta ^2}\left[g(z_1,k)g^{}(z_2,k)\right]\right]_{\zeta =0}=\stackrel{~}{q}(z,k)C_k^2$$ (13) Note, that $`\stackrel{~}{q}(z,k)=k`$ is true behind the barrier. The problem of tunneling collision between particles $`m_1`$ and $`m_2`$ can be reduced to the one-dimensional motion of a particle with reduced mass $`m_r=m_1m_2/(m_1+m_2)`$ in the external potentianl $`U(r)`$ . We make this reduction in the original Hamiltonian and then second quantize it. We need to perform calculations of $`G^{(0)+}`$ and $`K^0`$ similar to those outlined above. However, now we have to do 3-D calculations, using spherical coordinates. So we replace $`\widehat{L}(z)`$-operator with $`\widehat{L}_𝐱=\mathrm{\Delta }_𝐱/2m_r`$, denote its radial part in spherical coordinates by $`\widehat{L}_r`$ and introduce the function $`g(𝐱)`$ according to $`g(𝐱)={\displaystyle \frac{k^2dk}{(2\pi )^3}g_k(𝐱)}`$ (14) $`g_k(𝐱)={\displaystyle \underset{lm}{}}g_{kl}(r)Y_{lm}(\theta \varphi )`$ (15) $`\left(\widehat{L}_rU_l(r)\right)g_{kl}(r)=ϵ_𝐤g_{kl}(r),`$ (16) where $`Y_{lm}(\theta \varphi )`$ is a spherical harmonic and $`U_l(r)=U(r)+\frac{l(l+1)}{2m_rr^2}`$. Quite analogously to Eq. (7) we express $`\psi `$ through $`g(𝐱)`$ as $$\psi (𝐱_1,t_1)=\underset{k_1l_1m_1}{}\widehat{a}_{k_1l_1}g_{k_1l_1}(r_1)Y_{l_1m_1}(\theta _1\varphi _1)e^{iϵ_{𝐤_1}t_1}.$$ (17) Now using Eqs. (6), (16), (17) and taking into account the contribution only from spherically symmetric modes, we obtain $`G^{(0)+}`$ $`G_\omega ^{(0)+}(𝐱_1,𝐱_2)=2\pi i{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{k^2dk}{(2\pi )^3}}`$ $$\delta (\omega ϵ_𝐤+\mu )n_F(ϵ_𝐤)g_k(𝐱_1)g_k^{}(𝐱_2).$$ (18) Rewrite the rate $`K^0`$, Eq. (9), as $$K^{(0)}(𝐱)=\frac{i}{2m_r}\frac{d\omega }{2\pi }\left[\mathrm{\Delta }_{𝐱_1}G_\omega ^{(0)+}(𝐱_1,𝐱_2)\right]_{𝐱_1𝐱_2},$$ (19) In the WKB approximation we get for the function $`g_k(r)`$ the following formula $$g_k(r)=\frac{1}{r}\chi (r)$$ (20) $$\chi (r)=\frac{\overline{C}_k}{\stackrel{~}{q}(r,k)^{1/2}}exp\left[i_{r_0}^r\stackrel{~}{q}(r^{},k)𝑑r^{}\right]$$ (21) $$W(k)=exp\left[_{r_0}^{\alpha /\overline{ϵ}_𝐤}\stackrel{~}{q}(r^{},k)𝑑r^{}\right]$$ (22) $$\stackrel{~}{q}(r,k)=\sqrt{2m_r(\overline{ϵ}_kU(r))},$$ (23) where the bar denotes dependence on the reduced mass, $`\overline{ϵ}_k=k^2/2m_r`$, $`\overline{C}_k^2=\overline{A}_k^2W(k)`$, $`\alpha Z_1Z_2e_1e_2`$ and $`\overline{A}_k^2=r_0^2S(\overline{ϵ}_k)/\overline{ϵ}_𝐤`$. We evaluate the derivative of Eq. (19) in the quasiclassical sense $$\left[g_k^{}(𝐱_2)\mathrm{\Delta }_{𝐱_1}g_k(𝐱_1)\right]_{𝐱_1𝐱_2=r_0}=k\frac{\overline{C}_k^2}{r_0^2}$$ (24) and obtain the final result for $`K^{(0)}`$ $$K^{(0)}=n_0_0^{\mathrm{}}\frac{4\pi k^2dk}{(2\pi )^3}\overline{n}_M(k)\frac{k}{m_r}\frac{S(\overline{ϵ}_k)}{\overline{ϵ}_k}W(k),$$ (25) where $`\overline{n}_M(k)`$ is the Maxwell distribution which depends on $`m_r`$. It agrees with the answer obtained in by averaging the quasiclassical tunneling factor $`W(k)`$ over the Maxwell distribution, $`n_M(v)\sigma (m_rv^2/2)v`$. Hence, we see, that the dispersion relation $`\delta (\omega ϵ_𝐪)`$ was implicitly used in , which corresponds to an approximation of instantaneous, two-body collisions. Now we will not make this assumption and proceed with finding the tunneling rate, $`K`$, of the particles, which always collide and hence are never “in between collisions”. By using , we find $`\left({\displaystyle \frac{q}{m}}{\displaystyle \frac{}{z}}+U^{}(z){\displaystyle \frac{}{q}}\right)G_{\omega 𝐪}^+(z)=\gamma _{\omega 𝐪}(z)G_{\omega 𝐪}^+(z)`$ $$\gamma _{\omega 𝐪}(z)n_F(\omega )\left(G_{\omega 𝐪}^R(z)G_{\omega 𝐪}^A(z)\right),$$ (26) where we denote $`\mathrm{\Sigma }^R\mathrm{\Sigma }^Ai\gamma _{\omega 𝐪}(z)i\gamma _{\omega 𝐪_{}}(q,z)`$. Note, that $`\gamma _{\omega 𝐪}(z)`$ is a non-linear function of $`G^{\alpha \beta }`$ . We now analyze Eq. (26) qualitatively, which will help us to find that part of the solution, which makes the largest contribution to $`K`$. In the region away from the barrier we can neglect the LHS of Eq. (26), and obtain: $$\stackrel{~}{G}_{\omega 𝐪}^+(z)=n_F(\omega )\left(G_{\omega 𝐪}^R(z)G_{\omega 𝐪}^A(z)\right)$$ (27) Therefore, Eq. (26) will allow us to propagate this solution into the region behind the barrier. We can find the first integral and formally integrate Eq. (26) along trajectories, which will lead to a sum of solutions of homogeneous and ”inhomogeneous” equations (due to the last term in (26)). As can be seen from Eq. (26), ”inhomogeneous” solution will involve the integral of the product of $`\gamma _{\omega 𝐪}(z)`$ and $`G_{\omega 𝐪}^R(z)`$ or $`G^R=G^A`$. Qualitatively, it means, that in the region of $`z`$ behind the barrier it will be proportional to a product of at least two exponential factors $`W(k)`$. This is so, because both $`\gamma _{\omega 𝐪}(z)`$ and $`G_{\omega 𝐪}^R(z)`$ will depend on the integral of the product $`g(z_1,k)g^{}(z_2,k)`$ and such product is $`W(k)`$ as can be seen from Eqs. (10)-(12). Explicitely, the expression for the retarded Green function can be obtained through the usual resummation $$G_{\omega 𝐪}^R(z)=\frac{1}{\omega ϵ_𝐪\mathrm{\Sigma }_{\omega 𝐪}^R(z)}$$ (28) Since the advanced function, $`G^A`$, is related to $`G^R`$ by $`G^A=G^R`$, we find that their difference is $`G^RG^A`$ $`=`$ $`2Im\left({\displaystyle \frac{1}{\left(G_0^R\right)^1\mathrm{\Sigma }^R}}\right)`$ (29) $`=`$ $`{\displaystyle \frac{\gamma }{\left(G_0^R\right)^2+\left(Im\mathrm{\Sigma }^R\right)^2}}`$ (30) Now we will show that $`\gamma `$ is $`W(k)`$, which will allow us to neglect the “inhomogeneous” term in Eq. (26), as being second order in $`W(k)`$ since $`G^RG^A`$ is also proportional to $`W(k)`$. First, note that Eq. (15) is written in the Furry representation. The Furry representation is needed, since we are interested in the behaviour of all the functions in the external field $`U(z)`$. This means that the diagrammatic expansion of the operator $`\gamma =2Im\mathrm{\Sigma }^R`$ in Eq. (26) is built upon zero-th order functions, which incorporate dynamics in the external potential and hence are exponentially suppressed in the region behind the barrier. This is in contrast to the usual expansion which utilizes zero-th order functions of free particles. (Such representation is usually called Furry representation in the literature.) In other words, these functions resolve the operator $$\left(\frac{q}{m}\frac{}{z}+U^{}(z)\frac{}{q}\right)$$ (31) which is the left hand side of Eq. (26). Therefore, they are exponentially suppressed in the region of $`z`$ behind the barrier. This can be also seen from the explicit expressions for the Green functions. The function $`G^{(0)+}`$, is written in Eq. (8), and is proportional to the product $`g(z_1,k)g^{}(z_2,k)`$. As discussed above, this product is exponentially suppressed behind the barrier. Analogously, the function $`G^{(0)+}`$ can be obtained from $`G^{(0)+}`$ by substitution $`n(𝐪_{}k)(1n(𝐪_{}k))`$, and is also suppressed exponentially, since its dependence on $`z`$ is the same. This is an important property, which we will make use of, while analyzing the behaviour of $`\gamma `$ in the region behind the barrier. Now we can write explicitely expression of the first Born diagram for $`\gamma _{\omega 𝐪_{}}(z_1,z_2)`$ in the Furry representation from the Fig. 1 b) of Appendix A. There is a slight difference from the case of free zero-th order functions. Namely, there are additional integrals over intermediate $`z_3,z_4`$-points in the loop (see Appendix A for notations): $`i\mathrm{\Sigma }_{\omega 𝐩_{}}^+(z_1,z_2)={\displaystyle G_a^+}`$ $`(\omega ,𝐩_{}𝐪_{};z_1,z_2)G_b^+(\omega ,𝐩_1^{};z_4,z_3)G_b^+(\omega ,𝐩_1^{}𝐪^{};z_3,z_4)`$ (33) $`V(𝐪_{};z_1z_4)V(𝐪_{};z_2z_3){\displaystyle \frac{d\omega _{𝐩_1}d𝐩_1^{}}{(2\pi )^3}}{\displaystyle \frac{d\omega _𝐪d𝐪_{}}{(2\pi )^3}}dz_3dz_4,`$ Since all the Green functions in Eq. (33) are exponentially suppressed behind the barrier, we will obtain the largest contribution, if we restrict integration over $`z_3`$ and $`z_4`$ to the region away from the barrier, $`z>z_{}`$. The functions $`G_b^+,G_b^+`$ are not exponentially suppressed there. However, note that $`\mathrm{\Sigma }_{\omega 𝐩_{}}^+(z_1,z_2)`$ depends on $`z_1,z_2`$ through the function $`G_a^+`$ in Eq. (33). One can see that behaviour of $`\mathrm{\Sigma }_{\omega 𝐪_{}}^+(q,z)`$ in the region of $`z`$ behind the barrier will be determined by the behaviour of $`G_a^+(\omega ,𝐩_{}𝐪_{};z_1,z_2)`$, Wigner-transformed as in Eq. (34). The same argument applies to $`\mathrm{\Sigma }^+`$, which will be determined by $`G_a^+`$. Therefore, $`\gamma =\mathrm{\Sigma }^+\mathrm{\Sigma }^+`$, which is $$\gamma _{\omega 𝐪_{}}(q,z)=exp[iq(z_1z_2)]\left(\mathrm{\Sigma }_{\omega 𝐪_{}}^+(z_1,z_2)\mathrm{\Sigma }_{\omega 𝐪_{}}^+(z_1,z_2)\right)d(z_1z_2)$$ (34) will depend on the behaviour of these Green functions behind the barrier. But we know that, since both $`G_a^{(0)+}`$ and $`G_a^{(0)+}`$ resolve the operator (31), they are exponentially suppressed behind the barrier. (This is also proven by their explicit expressions, as explained above). Therefore, $`\gamma _{\omega 𝐪_{}}(q,z)`$ is also exponentially suppressed there. More formally, we can obtain explicit expression for the $`\gamma `$ as far as its dependence on $`z`$ is concerned. First, we have to simplify Eq. (33); then apply similar arguments to $`\mathrm{\Sigma }^+`$, which will allow to find $`\gamma `$. We are interested in the region of $`z(z_1+z_2)/2`$ behind the barrier. We can make this region rather small, compared to the region away from the barrier, by choosing potential $`U(z)`$ to be infinite for $`z<z_0`$. Here $`z_0`$ is an arbitrary, finite point behind the barrier. This procedure does not restrict generality of this argument, since the actual region behind the Coulomb barrier of a fusing particle is much smaller than the region away from the barrier. Then, we can approximate $`V(𝐪_{},z_1z_4)`$ by $`V(𝐪_{},zz_4)`$, as well as $`V(𝐪_{},z_2z_3)`$ by $`V(𝐪_{},zz_3)`$ in Eq. (33). As we explained above, the main contribution to the integral over $`z_3,z_4`$ comes from the region of $`z>z_{}`$ (the region away from the barrier). In this region, the functions $`g_b(z_3,k),g_b(z_4,k)`$, entering $`G_b^{(0)+},G_b^{(0)+}`$ in Eq. (33), can be well approximated by $`exp(ikz_3),exp(ikz_4)`$. After this approximation, the region of integration can be extended from $`\mathrm{}`$ to $`\mathrm{}`$, which will introduce only an error in the numerical coefficient on the order of unity. By using the explicit forms of $`G^{(0)+},G^{(0)+}`$ from Eq. (8), we can carry explicitely integrations over frequencies. Then we find $`i\mathrm{\Sigma }_{\omega 𝐩_{}}^+(z_1,z_2)=`$ $`4\pi i{\displaystyle \frac{d𝐩_1}{(2\pi )^3}\frac{d𝐪}{(2\pi )^3}\underset{\mathrm{}}{\overset{\mathrm{}}{}}\frac{dk}{(2\pi )}n_{𝐩_{}𝐪_{}k}^an_{𝐩_1}^bg_a(z_1,k)g_a^{}(z_2,k)}`$ (36) $`V_𝐪^2\delta (\omega _pϵ_{𝐩_1}^b+ϵ_{𝐩_1𝐪}^bϵ_{𝐩_1^{}𝐪^{},k}^a)`$ In a similar manner, we find the answer for the $`\mathrm{\Sigma }^+`$: $`i\mathrm{\Sigma }_{\omega 𝐩_{}}^+(z_1,z_2)=`$ $`4\pi i{\displaystyle \frac{d𝐩_1}{(2\pi )^3}\frac{d𝐪}{(2\pi )^3}\underset{\mathrm{}}{\overset{\mathrm{}}{}}\frac{dk}{(2\pi )}(1n_{𝐩_{}𝐪_{}k}^an_{𝐩_1}^b)n_{𝐩_1𝐪}^bg_a(z_1,k)g_a^{}(z_2,k)}`$ (38) $`V_𝐪^2\delta (\omega _pϵ_{𝐩_1}^b+ϵ_{𝐩_1𝐪}^bϵ_{𝐩_1^{}𝐪^{},k}^a)`$ The expression for $`\gamma `$ follows from Eqs. (34, 36, 38). We see that its dependence on $`z`$ is determined by functions $`g(z,k)`$, which are exponentially suppressed behind the barrier. This completes the proof and allows us to neglect the “inhomogeneous” term in the Eq. (26). The solution of the ”homogeneous” equation, $`G_f^+`$, involves only one factor of $`W(k)`$, since $`G_f^+gg^{}`$, see Eq. (26). Therefore, to obtain $`K`$ we will find only the solution to the ”homogeneous” equation. We supplement this homogeneous equation with the boundary condition, $`\stackrel{~}{G}^+`$ from Eq. (27) imposed at $`z=z_{}`$ away from the barrier. Since the largest contribution to $`K`$ is made by $`G_f^+`$ one can see, that we can use Eq. (9) to find $`K`$ if we substitute $`G_f^+`$ instead of $`G^{(0)+}`$ in (9). Therefore, we find $`G_f^+`$ with dependence on $`z_1,z_2`$ from the very begining by acting on Eq. (2) with the operator $$g(z_1,k_1)g^{}(z_2,k_2)\frac{dk_1}{2\pi }\frac{dk_2}{2\pi }$$ (39) We use the boundary condition, Eq. (27) and obtain the following answer for $`G_f^+`$: $$in_F(\omega )g_k(z_1)\delta _\gamma (\omega ϵ_{𝐪_{}k})g_k^{}(z_2)𝑑k$$ (40) where $`g_k(z)g(z,k)`$ and we used for $`\delta _\gamma `$ $`2ImG_{\omega 𝐪_{}}^R(k)={\displaystyle \frac{\gamma _{\omega 𝐪_{}}(k)}{(\omega ϵ_{𝐪_{}k})^2+(\gamma _{\omega 𝐪_{}}(k))^2}}`$ (41) Comparing Eq. (8) with (40), (41) we see, that the effect of collisions, which makes the largest contribution to $`K`$ can be described as $`\delta (\omega ϵ_{𝐪_{}k})\delta _\gamma (\omega ϵ_{𝐪_{}k})`$, as can be expected on the intuitive grounds. Now we substitute Eqs. (40), (41) into (9) and integrate $`𝑑\omega /2\pi `$ by using the result of for large momenta, $`qq_T`$: $`{\displaystyle \frac{d\omega }{2\pi }n_M(\omega )\delta _\gamma (\omega ϵ_{𝐪_{}k})}`$ $`=`$ $`n_M(ϵ_{𝐪_{}k})+\delta n_\gamma (𝐪_{}k),`$ (42) where $`\delta n_\gamma (𝐪_{}k)`$ is the “quantum tail”. To describe the tunneling of particles $`m_1`$ and $`m_2`$ we perform the same steps as those leading to Eq. (25). Note that this procedure implies quantization of the hypothetical particle with the mass equal to the reduced mass of fusing particles. Strictly speaking, this approximation is correct in the limit when one particle is light and the other is heavy. Keeping this in mind, we arrive at the answer for $`K`$: $`K=n_0{\displaystyle \frac{d\omega _𝐩}{2\pi }_0^{\mathrm{}}\frac{4\pi p^2dp}{(2\pi )^3}\frac{S(\overline{ϵ}_p)}{\overline{ϵ}_p}}`$ $$\frac{p}{m_r}W(p)\overline{n}_M(\omega )\delta _\gamma (\omega ϵ_𝐩)$$ (43) We see that it involves integrals over both the energy and momentum. The width of the resonance, $`\gamma (\omega _𝐩,𝐩)`$, is, in general, a complicated function of energy and momentum. It obeys under certain assumptions the integral equation which follows from the Dyson diagram expansion. More simplified calculations, based on the use of the zero-th order Green functions, lead to the analytical answer found in (see Appendix A). We use this answer for $`\gamma `$ and obtain $`K=n_0{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{4\pi p^2dp}{(2\pi )^3}}{\displaystyle \frac{S(\overline{ϵ}_p)}{\overline{ϵ}_p}}`$ $$\left[\overline{n}_M(p)+\delta \overline{n}_\gamma (p)\right]K_M+K_\gamma .$$ (44) Finally, we evaluate $`\mathrm{\Sigma }^R`$ through the collision frequency, $`\nu _T`$, carry out the integral in (44) and form the ratio $`r_{12}=K_\gamma /K_M`$ (see Appendices A and B): $$r_{12}=\frac{3^{19/2}}{8\pi ^{3/2}}\underset{j}{}\frac{h\nu _T(m_{coll})}{T}\left(\frac{m_{coll}}{m_r}\right)^2\frac{e^{\tau _{12}}}{\tau _{12}^8}$$ (45) Here $`m_{coll}=m_rm_j/(m_r+m_j)`$, $`m_r=m_1m_2/(m_1+m_2)`$, $`\tau _{12}=3(\pi /2)^{2/3}\left(\frac{100Z_1^2Z_2^2A_{12}}{T_k}\right)^{1/3}`$, $`A_{12}=A_1A_2/(A_1+A_2)`$, $`T_k`$ is the temperature in keV; $`A_1`$, $`A_2`$ are the atomic numbers of tunneling particles, $`m_j`$ is the mass of the background particles colliding with tunneling particles. Summation is over all species of colliding background particles, including tunneling species. This result, (44), (45), supports the idea of averaging of $`\sigma v`$ over momentum distribution postulated in . Note an important feature of the result (45): this ratio goes to zero as $`\mathrm{}\nu _𝐩/T0`$. As we show in Appendix A, the width, $`\gamma `$, can be approximated as $`\mathrm{}\nu _𝐩`$ . Therefore, this means that in the limit of negligible width, the correction to the rate becomes very small, and the reaction rate coincides with the usual Gamow rate. This is expected, since the smaller the width, the closer the particle to the mass shell. And the Gamow rate is derived for on-shell particles. On the other hand, the ratio becomes large, if $`\mathrm{}\nu _𝐩/T`$ is not very small, yet still $`\mathrm{}\nu _𝐩/T<1`$. It is this condition, which allows us to use the result of for $`\gamma `$ in Eqs. (44), (45). In the opposite case, $`\mathrm{}\nu _𝐩/T1`$, corresponding to more strongly coupled plasma, the theory presented here does not apply. Consider a $`DT`$ plasma of density $`\rho =10\mathrm{g}/\mathrm{cm}^3`$ and temperature $`T=0.1`$ keV. The ratio $`\mathrm{}\nu _𝐩/T=0.0075`$ for these conditions. Then Eq. (45) will lead to the rate accelerated by five orders of magnitude as compared to the conventional answer. We find the ratio $`r_{DT}=K_{DT}^\gamma /K_{DT}^M=1.810^5`$, with $`K_{DT}^M=2.610^{30}\mathrm{cm}^3/\mathrm{sec}`$, $`K_{DT}^\gamma =4.810^{25}\mathrm{cm}^3/\mathrm{sec}`$. It is also interesting to consider how the reaction rates are changed under conditions relevant to astrophysics, e.g. in the Sun interior. Modification of several reaction rates and their influence on neutrino fluxes from the Sun was considered in . Here we give examples of $`p+Be^7`$ and $`p+p`$ reactions, occuring at the center of the Sun, $`r=0`$, where the temperature is $`T=1.336`$ keV and the density $`\rho =156\mathrm{g}/\mathrm{cm}^2`$, according to the solar model of . The estimate for $`\gamma _{17}/T`$ is $`\mathrm{}\nu _T/T0.0056`$, while the rates and the ratio, $`r_{17}`$, are $`K_{17}^M=9.610^{36}\mathrm{cm}^3/\mathrm{sec}`$, $`K_{17}^\gamma =4.010^{30}\mathrm{cm}^3/\mathrm{sec}`$, $`r_{17}=4.1710^5`$. In case of heavier elements and not very small $`\mathrm{}\nu _T/T`$, the ratio $`r_{ij}`$ becomes much higher. On the other hand, the corresponding numbers for the $`p+p`$ reaction under the same conditions are: $`\gamma /T\mathrm{}\nu _T/T0.00054`$, $`K_{11}^M=1.410^{43}\mathrm{cm}^3/\mathrm{sec}`$, $`r_{11}=310^3`$. The reaction rate is essentially the same. Note that $`\mathrm{}\nu _T/T0.00054`$ is rather small. This confirms the conclusion of the discussion above, that small $`\gamma `$ lead to insignificant changes in the reaction rates. Note, however, that we need to have some gauge to determine which $`\gamma `$ is small. Naturally, such a gauge should be a ratio of the height of the Coulomb barrier to the average particle energy. As we see from Eq. (45) this ratio enters the answer for $`r_{12}`$ through $`\tau _{12}`$. In conclusion, we derived the momentum distribution in equilibrium through real-time technique of Keldysh and Korenman which coincides with the distribution, obtained in imaginary time . This distribution takes into account off-shell effects which lead to the quantum tail with power-like dependence on momentum in contrast to the distribution for on-shell particles which is Fermi-Dirac or Maxwellian. We calculated the rates of fusion reactions for particles in a weakly coupled plasma and showed that off-shell effects may significantly enhance the reaction rates. I would like to acknowledge helpful discussions with N. J. Fisch, D. Spergel, A. N. Starostin, A. Demura, A. A. Panteleev. This work was supported by NSF–DOE grant DE-FG02-97ER54436. E-mail: vsavchenpppl.gov APPENDIX A CALCULATION OF THE MASS OPERATOR, $`\mathrm{\Sigma }^R`$, AND THE DISTRIBUTION FUNCTION OVER MOMENTUM Galitski and Yakimets found how off-shell processes change the distribution function over momentum in equilibrium, using imaginary time technique . In this Appendix we show how to find the same result through real time diagram method. We also generalize it slightly by allowing for collisions between different types of particles. Dyson equations for the function $`G^+`$ can be written as $`\left(i{\displaystyle \frac{}{t}}+i{\displaystyle \frac{𝐩}{m}}{\displaystyle \frac{}{𝐑}}\right)G^+(𝐑,t;\omega ,𝐩)=`$ (46) $`\mathrm{\Sigma }^+G^++\mathrm{\Sigma }^+G^+`$ (47) To simplify notation we do not write explicitely the variables the functions $`\mathrm{\Sigma }`$ and $`G`$ depend upon. We set to zero the LHS of this equation in equilibrium and obtain $$G^+=\frac{\mathrm{\Sigma }^+}{\mathrm{\Sigma }^+}\left(G^RG^A+G^+\right),$$ (48) which follows from the relation between Green functions . The result for $`G^+`$ follows by substituting Eq. (48) into (28): $$G^+=2\pi in_F(\omega )\frac{Im\mathrm{\Sigma }^R}{(\omega ϵ_𝐩Re\mathrm{\Sigma }^R)^2+(Im\mathrm{\Sigma }^R)^2}$$ (49) We use this equation to find the distribution function over momentum which is $$f(𝐩)=\underset{\mathrm{}}{\overset{\mathrm{}}{}}\frac{d\omega _𝐩}{2\pi }G^+(\omega _𝐩,𝐩).$$ (50) First we calculate the mass-operator $`\mathrm{\Sigma }^R`$. It can be expressed using the relation $$\mathrm{\Sigma }^R=\frac{1}{2}(\mathrm{\Sigma }^{}\mathrm{\Sigma }^{++})\frac{1}{2}(\mathrm{\Sigma }^+\mathrm{\Sigma }^+)$$ (51) The first term in this equation corresponds to the real part $`Re\mathrm{\Sigma }^R`$, while the second term is imaginary part $`Im\mathrm{\Sigma }^R`$. Since it is the imaginary part of $`\mathrm{\Sigma }^R`$ which is responsible for the quantum tail we calculate only $`Im\mathrm{\Sigma }^R`$ and ignore $`Re\mathrm{\Sigma }^R`$ altogether. The first two terms of the series for the operators $`\mathrm{\Sigma }^{\alpha \beta }`$ are shown in the Fig. 1. One has to assign all possible combination of signs $`,+`$ to the vertices, keeping in mind that the sign does not change along the dashed line . In a more rigorous approach, based on the resummation of the series of graphs, one can show that, solid lines in the diagrams of Fig. 1 correspond to exact Green-functions, while the dashed lines represent scattering amplitudes of the particles in the media. However, such resummations lead to intractable non-linear integro-differential equations. As a first approximation we assume that, the solid lines correspond to zero-order Green functions, while the dashed lines correspond to the interaction potential between the particles. These diagrams represent quite different physical phenomena. Diagram “a)” describes the screening effects and formation of the quasiparticles in the plasma. Therefore, an imaginary part of this graph represents decay of the quasi-particles. This term is conventionally put into the LHS of the kinetic equation, signifying that it has nothing to do with the collisional integral . Diagram “b)”, on the other hand, if evaluated with zero-order Green-functions, gives the usual Boltzman collision terms . Therefore, the imaginary parts of the diagrams “b)” and “c)” describe damping of the oscillators used for the second quantization of the interacting particles. We evaluate the “b)” and “c)”-type of diagrams for $`\mathrm{\Sigma }^+`$, assuming that collisions occur between particles of type $`a`$ and $`b`$. Therefore, we assign the index $`a`$ to the Green function, represented by the horizontal line in the diagram “b)”, and the index $`b`$ to the Green functions of the loop. Note that since the diagram “c)” corresponds to the exchange processes only single index can be assigned to its Green functions. This obviously reflects the fact that only identical particles are subject to exchange. Taking into account these remarks, we write the analytical expression for the diagramm b) in the form $`i\mathrm{\Sigma }^+={\displaystyle G_a^+}`$ $`(PQ)G_b^+(P_1)G_b^+(P_1Q)`$ (53) $`\left|V(𝐪)\right|^2{\displaystyle \frac{d\omega _{𝐩_1}d𝐩_1}{(2\pi )^4}}{\displaystyle \frac{d\omega _𝐪d𝐪}{(2\pi )^4}},`$ where capital letters denote four-momentum, e.g. $`P=(\omega _𝐩,𝐩)`$ and $`V(𝐪)`$ is the interaction potential between the particles. We substitute here equilibrium zero-th order Green functions $$G^+(\omega _p𝐩)=2\pi in_𝐩\delta (\omega _pϵ_𝐩+\mu ),$$ (54) $$G^+(\omega _p𝐩)=2\pi i(1n_𝐩)\delta (\omega ϵ_𝐩+\mu ).$$ (55) Analogously, we obtain an expression for the diagram c). For a moment, we omit the indices $`a`$ and $`b`$ to simplify the formulas, but write them explicitely in the final formulas. Then the result for $`\mathrm{\Sigma }^+`$ is $`\mathrm{\Sigma }^+(\omega _𝐩,𝐩)=(2\pi i){\displaystyle \frac{d𝐩_1}{(2\pi )^3}\frac{d𝐪}{(2\pi )^3}}`$ (56) $`\left(2V_𝐪^2V_𝐪V_{𝐪𝐩𝐩_1}\right)n_{𝐩𝐪}n_{𝐩_1𝐪}`$ (57) $`\delta (\omega _𝐩ϵ_{𝐩𝐪}+ϵ_{𝐩_1}ϵ_{𝐩_1𝐪})`$ (58) Proceeding in a similar fashion with $`\mathrm{\Sigma }^+`$ and using Eq. (51) we obtain the final result for $`Im\mathrm{\Sigma }^R\gamma (\omega _p,𝐩)`$: $`\gamma `$ $`(\omega _𝐩,𝐩)=\pi i{\displaystyle \frac{d𝐪}{(2\pi )^3}\frac{d𝐩_1}{(2\pi )^3}V_𝐪(2V_𝐪V_{𝐪𝐩𝐩_1})}`$ (61) $`\left[n_{𝐩𝐪}(n_{𝐩_1𝐪}n_{𝐩_1})+n_{𝐩_1}(1n_{𝐩_1𝐪})\right]`$ $`\delta (\omega _𝐩ϵ_{𝐩𝐪}+ϵ_{𝐩_1}ϵ_{𝐩_1𝐪})`$ This answer coincides with the result of obtained in the imaginary-time technique. Note that we could improve this relation by substituting $`\delta _\gamma `$ of Eq. (41) instead of the delta-function. This, however, would lead to an integral equation for $`\gamma `$, and its first iteration will be just Eq. (61). Now we can find the distribution function over momentum from Eqs. (50), (61), (49) $$f(𝐩)=\underset{\mathrm{}}{\overset{\mathrm{}}{}}\frac{d\omega _𝐩}{(2\pi )^2}\frac{n_F(\omega )\gamma (\omega _𝐩,𝐩)}{(\omega _𝐩ϵ_𝐩)^2+\gamma ^2(\omega _𝐩,𝐩)}$$ (62) by substituting Eq. (61) into Eq. (49<sup>*</sup><sup>*</sup>*From now on we follow closely the work of . The integral $`𝑑\omega `$ can be rewritten as a sum of two terms $`f(𝐩)=`$ $`{\displaystyle \underset{\mathrm{}}{\overset{ϵ_0}{}}}{\displaystyle \frac{d\omega _𝐩}{2\pi }}n_F(\omega )\delta _\gamma (\omega _𝐩ϵ_𝐩)+`$ (64) $`{\displaystyle \underset{ϵ_0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{d\omega _𝐩}{2\pi }}n_F(\omega )\delta _\gamma (\omega _𝐩ϵ_𝐩),`$ where $`ϵ_0`$ is the characteristic energy. In the degenerate case, $`T\mu `$, we have $`ϵ_0=\mu `$, in the nondegenerate case, $`T\mu `$ and $`ϵ_0=T`$. We define $`p_0`$ as $`ϵ_0=p_0^2/2m`$. Consider the case of large momenta, $`pp_0`$. The function $`\delta _\gamma (\omega _𝐩ϵ_𝐩)`$ has a maximum at $`\omega _{}=ϵ_𝐩ϵ_0`$, so that $`\omega _{}[ϵ_0,\mathrm{}]`$, and the contribution of the second integral can be approximated as $`n_𝐩`$. In the region of integration of the first integral in Eq. (64) we can set $`n_F(\omega )1`$. Then Eq. (64) becomes for $`pp_0`$ $$f(𝐩)=n_𝐩+\underset{\mathrm{}}{\overset{ϵ_0}{}}\frac{d\omega _𝐩}{2\pi }\frac{\gamma (\omega _𝐩,𝐩)}{(\omega _𝐩ϵ_𝐩)^2}n_𝐩+\delta n_\gamma (𝐩),$$ (65) We carry out the integral in the second term of (65), use Eq. (61) and obtain $`\delta n_\gamma (𝐩)={\displaystyle \frac{d𝐪}{(2\pi )^3}\frac{d𝐩_1}{(2\pi )^3}V_𝐪(2V_𝐪V_{𝐪𝐩𝐩_1})}`$ (66) $`{\displaystyle \frac{\left[n_{𝐩𝐪}(n_{𝐩_1𝐪}n_{𝐩_1})+n_{𝐩_1}(1n_{𝐩_1𝐪})\right]}{(ϵ_𝐩ϵ_{𝐩𝐪}+ϵ_{𝐩_1}ϵ_{𝐩_1𝐪})^2}},`$ (67) with the condition on momenta $$\mathrm{}<ϵ_{𝐩𝐪}ϵ_{𝐩_1}+ϵ_{𝐩_1𝐪}<ϵ_0$$ (68) due to the integration of the delta-function from $`\gamma (\omega _𝐩,𝐩)`$. The integral in Eq. (67) should be carried out numerically. We can, however, obtain a reasonable estimate in the case of large momentum, $`pp_0`$. Observe that the first two terms in the brackets $`[..]`$ should have the same dependence on $`p`$. To obtain an estimate consider the product of the first term in $`(..)`$, $`V_𝐪^2`$, and the first term in $`[..]`$ which involves the product $`n_{𝐩𝐪}n_{𝐩_1𝐪}`$. We denote it $`I_1(𝐩)`$. Note that there is a region of non-exponential contribution to the integral, which is limited to the region of momenta $$\left|𝐩𝐪\right|\left|𝐩_1𝐪\right|p_0,pp_1qp_0$$ (69) The inequality (68) is satisfied which, in turn, means that the dominant contribution to this product can be approximated as $$I_1(𝐩)=\frac{d𝐪}{(2\pi )^3}\frac{d𝐩_1}{(2\pi )^3}2V_𝐪^2\frac{n_{𝐩𝐪}^{(a)}n_{𝐩_1𝐪}^{(b)}}{\stackrel{~}{ϵ}_𝐩^2},$$ (70) where $`\stackrel{~}{ϵ}_𝐩𝐩^2/2m_{ab}`$ and we reinstalled indices $`a`$ and $`b`$. Consider non-degenerate case. We change variables, $`𝐩𝐪=𝐪^{}`$, $`𝐩_1𝐪=𝐩_1^{}`$ and obtain an estimate $$I_1(𝐩)=e_a^2e_b^2\frac{(2m_{ab})^2}{p^8}\left(\frac{m_a}{2\beta \pi }\right)^{3/2}\left(\frac{m_b}{2\beta \pi }\right)^{3/2}e^{\beta (\mu _a+\mu _b)}$$ (71) We substitute the relation $$e^{\beta (\mu _a+\mu _b)}=n^{(a)}\lambda _a^3n^{(b)}\lambda _b^3$$ (72) $$\lambda _a^3=\left(\frac{m_a}{2\beta \pi }\right)^{3/2},\lambda _b^3=\left(\frac{m_b}{2\beta \pi }\right)^{3/2}$$ (73) into Eq. (71) and find $$I_1(𝐩)=\frac{(2m_{ab})^2}{p^8}(e_ae_b)^2n^{(a)}n^{(b)}$$ (74) Since the physical reason for non-zero $`\gamma (\omega _p,𝐩)`$ is collisions, we rewrite (74) in terms of the collision frequency $$\nu _p=\left(\frac{\pi }{2}\right)^{3/2}n^{(a)}(e_ae_b)^2m_{ab}^{1/2}T^{3/2}\frac{(2m_{ab}T)^{3/2}}{p^3}$$ (75) as $$I_1(𝐩)=\frac{2}{\pi ^{3/2}}\frac{\mathrm{}\nu _pT}{\stackrel{~}{ϵ}_𝐩^2}\frac{\stackrel{~}{p}_T}{p}\frac{n^{(b)}}{\stackrel{~}{p}_T^3},$$ (76) where $`\stackrel{~}{p}_T(2m_{ab}T)^{1/2}`$. This is essentially the result of for the degenerate case, sligtly generalized to the case of colliding particles with different masses. In the non-degenerate case we can keep only the term linear in $`n_𝐩`$ in Eq. (67), because the quadratic terms are smaller. We denote this term by $`I_3`$. Since there is no exponential suppression in the $`q`$-integral for this term, we have to evaluate the limits of the $`q`$-integration carefully. We can obtain these limits from Eq. (68). After some algebra it leads to $$\frac{q^2}{2m_{ab}}qx\left|\frac{𝐩}{m_a}+\frac{𝐩_1}{m_b}\right|>ϵ_𝐩T,$$ (77) where $`x`$ is the cosine of the angle between $`𝐪`$ and $`\frac{𝐩}{m_a}+\frac{𝐩_1}{m_b}`$. We resolve this inequality in the limit of large $`pp_T`$, neglecting all terms on the order of $`p_T^2`$ in comparison to $`p^2`$. Note that since the term in quesiton, $`I_3`$, is proportional to $`n_{𝐩_1}`$, the contribution from the region $`p_1p_T`$ is exponentially suppressed. Since we are interested in the non-exponential contribution only, we can assume that $`p_1p_T`$ in Eq. (77). Then we finally arrive at $$q>q_{}=\frac{m_{ab}}{m_a}p(1+\sqrt{1+\frac{m_a}{m_{ab}}})$$ (78) To obtain an estimate for $`I_3`$, we carry out the integral of the linear term in $`n_{𝐩_1}`$ from Eq. (67), multiplied by the first term in $`(..)`$, $`V_𝐪^2`$: $`I_3(𝐩)=2\pi {\displaystyle 𝑑𝐩_1\underset{1}{\overset{1}{}}𝑑x\underset{q_{}}{\overset{\mathrm{}}{}}q^2𝑑q}`$ (79) $`{\displaystyle \frac{(e_ae_b)^2}{q^4}}{\displaystyle \frac{e^{p_1^2/2m_bT}e^{\beta (\mu _a+\mu _b)}}{\left(\frac{q^2}{2m_{ab}}qx\left|\frac{𝐩}{m_a}+\frac{𝐩_1}{m_b}\right|\right)^2}}`$ (80) After performing the integral over $`x`$, we are left with the integral over $`y`$, which is the cosine of the angle between $`𝐩`$ and $`𝐩_1`$. The $`y`$-dependent term becomes $$𝑑y\frac{1}{\left(\frac{q^2}{2m_{ab}}\right)^2q^2\left(\frac{p^2}{m_a^2}+\frac{p_1^2}{m_b^2}+\frac{2pp_1y}{m_am_b}\right)}$$ (81) We carry out this integral, neglect all the terms of the order of $`p_T^2`$ in comparison to $`p^2`$ and obtain $`2m_a^2/(q^2p^2)`$. The remaining integrals over $`\underset{0}{\overset{\mathrm{}}{}}𝑑p_1`$ and $`\underset{q_{}}{\overset{\mathrm{}}{}}𝑑q`$ are trivial. Using Eq. (78) we get the result for $`I_3`$ $`I_3(p)={\displaystyle \frac{2}{3}}{\displaystyle \frac{1}{(2\pi )^3}}{\displaystyle \frac{m_a^2}{p^5}}\left({\displaystyle \frac{m_a}{m_{ab}}}\right)^3{\displaystyle \frac{(e_ae_b)^2n^{(b)}e^{\beta \mu _a}}{\left(1+\sqrt{1+\frac{m_a}{m_{ab}}}\right)^3}}`$ Now we can see that different terms in Eq. (67) have a different dependence on $`p`$. Note however that the answers for these terms are really estimates. To find the correct answer one should solve non-linear equation for $`\gamma `$ , which is possible only through numerical methods. It is also possible to obtain an estimate through dimensional analysis which leads to $$\gamma =2Im\mathrm{\Sigma }^R2\mathrm{}\nu _𝐩$$ (82) This expression has to be substituted into Eq. (65). Note that we have to carry out the integral over energy from $`\mathrm{}`$ to $`T`$. Since we do not know how $`\gamma `$ behaves at negative energies, we can only make an assumption that, this region does not contribute significantly to the integral, and the main contribution comes from $`\omega _𝐩T`$. Then the integral is equal roughly to $$I(p)=\frac{\mathrm{}\nu _pT}{2\pi ϵ_𝐩^2}\frac{n^{(a)}}{p_T^3}.$$ (83) We can simplify it to $$I(p)=\frac{\mathrm{}\nu _pT}{2\pi ϵ_𝐩^2}e^{\beta \mu _a}$$ (84) which is the result for the non-degenerate case found in . We will use this estimate in our calculations of the reaction rate because of its simplicity. APPENDIX B REACTION RATE: CALCULATION OF THE INTEGRAL OVER $`\delta n_\gamma (𝐩)`$ In this Appendix we find the rate $`K_\gamma `$ of Eq. (44). Observe, that in the main text we used approximation of a hypothetical particle with the mass equal to the reduced mass of fusing particles, $`m_r`$. It is this particle which undergoes collisions with the background, leading to its power-like distribution over momentum. We will need the result (83) for these particles. We write as $$I(𝐩)=\frac{\mathrm{}\nu _pT}{2\pi \stackrel{~}{ϵ}_𝐩^2}\left(\frac{m_r}{m_{coll}}\right)^{3/2}e^{\beta \mu _a}$$ (85) taking into account that $$\frac{n^{(a)}}{\stackrel{~}{p}_T^3}=\left(\frac{m_r}{m_{coll}}\right)^{3/2}e^{\beta \mu _a}.$$ (86) Therefore, we substitute Eq. (85) into Eq. (44) and obtain $`K_\gamma `$ $`K_\gamma =\left({\displaystyle \frac{m_r}{m_{coll}}}\right)^{3/2}(2m_r)^2{\displaystyle \frac{\mathrm{}\nu _T(m_{coll})T}{2\pi }}`$ (87) $`S(ϵ^{})(2m_{coll}T)^{3/2}{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}e^{\pi p_G/p}{\displaystyle \frac{1}{p^6}}𝑑p,`$ (88) where $`S(ϵ^{})`$ is the astrophysical factor at some fixed value of energy. The integral is trivial and is equal to $$\underset{0}{\overset{\mathrm{}}{}}e^{\pi p_G/p}\frac{1}{p^6}𝑑p=\frac{24}{\pi ^5p_G^5}.$$ (89) For the sake of completeness, we also find here the usual Gamow rate: $$K_M=4\pi S(ϵ^{})_0^{\mathrm{}}e^{\pi p_G/pp^2/2m_rT}p𝑑p.$$ (90) The result, obtained by using the steepest descend method, is $$K_M=S(ϵ^{})\frac{2}{3}\pi ^{1/2}(m_rT)\tau ^{1/2}e^\tau .$$ (91) Strictly speaking, the astrophysical factors in Eqs. (88) and (91) need not be the same. In fact, one should consider a generalized astrophysical factor, $`S(\omega _p,ϵ_𝐩)`$, depending both on energy and momenta. Such factors would take into account the influence of the off-shell effects on nuclear transformations. However, as far as we know, they have not been studied. Therefore, we assume that the factors in Eqs. (88), (91) are the same. With this in mind, we can form the ratio $$r_{12}=\frac{K_\gamma }{K_M}=\frac{3^{19/2}}{2^2\pi ^{3/2}}\frac{\mathrm{}\nu _T(m_{coll})}{T}\left(\frac{m_{coll}}{m_r}\right)^2\frac{e^\tau }{\tau ^8}$$ (92)
warning/0002/astro-ph0002045.html
ar5iv
text
# Conservation of circulation in magnetohydrodynamics ## I Introduction Kelvin’s theorem on the conservation of circulation of a simple perfect fluid has played an important role in the development of hydrodynamics. For instance, it shows that potential flows are possible, that isolated vortices can exist, that they obey the Helmholtz laws, etc. Kelvin’s theorem is valid only for flows in which the body force per unit mass is a gradient; mostly this includes incompressible or isentropic flows of one–component fluids. Most flows in geophysics and astrophysics are more complicated. In particular, many fluids in the real world carry magnetic fields: they are magnetofluids. Yet the Lorentz force per unit mass on a magnetofluid is almost never a perfect gradient. Thus the circulation theorem in its original form is almost never true in magnetohydrodynamics (MHD). Must we then surrender the many insights that Kelvin’s theorem conferred on pure hydrodynamics ? Not necessarily. One might speculate that a suitable combination of fluid velocity $`𝐯`$ and magnetic induction $`𝐁`$ may inherit the property of having a “circulation” on a closed curve which is preserved as that curve is dragged with the magnetofluid. Such conserved circulation might play as useful a role in MHD as has Kelvin’s circulation in pure fluid dynamics. For example, it might help characterize a set of magnetoflows as being potential in some sense, with consequent simplification of this intricate subject. Or it might help to characterize a new type of vortex, a hybrid vorticity–magnetic rope. In view of the importance of the vortex phenomenon in contemporary physics, this last possibility is by itself ample reason to delve into the subject. Two decades ago, E. Oron discovered, with the formalism of relativistic perfect MHD, a circulation theorem of the above kind. Although some of its consequences for new helicity conservation laws have been explored, this new conserved circulation has remained obscure. Contributing to this, no doubt, is the fact that it has only been derived relativistically, and that this derivation is an intricate one, even for relativistic MHD. In addition, Oron’s derivation assumes both stationary symmetry and axisymmetry, while it is well known that Kelvin’s theorem requires neither of these. In the present paper we use the least action principle to give a rather straightforward existence proof for a generically conserved hybrid velocity–magnetic field circulation within the framework of perfect MHD which does not depend on spacetime symmetries. We do this both at the Newtonian (Sec.II) and general relativistic (Sec.III) levels; the importance of MHD effects in pulsars, active galactic nuclei and cosmology underscores that this last arena is not just of academic importance. As mentioned, we approach the whole problem not from equations of motion, but from the least action principle. Lagrangians for nonrelativistic pure perfect flow have been proposed by Herivel, Eckart, Lin, Seliger and Witham, Mittag, Stephen and Yourgrau and others. Many of the proposed Lagrangians necessarily imply irrotational flow, i.e. not to generic flow, a deficiency which is often missed by the authors. Lin introduced a device that allows vortical flows to be encompassed. This device is used by Seliger and Witham. Lagrangians for nonrelativistic perfect MHD flow in Eulerian coordinates have been proposed by Eckart, Henyey, Newcomb, Lundgren and others. In special relativity Penfield proposed a perfect fluid Lagrangian which admits vortical isentropic flow. The early general relativistic Lagrangian of Taub as well as the more recent one by Kodama et. al describe only irrotational perfect fluid flows. The Lin device is incorporated by Schutz, whose perfect fluid Lagrangian admits vortical as well as irrotational flows in general relativity. Carter introduced Lagrangians for particle-like motions from which can be inferred the properties of fluid flows, including vortical ones. Achterberg proposed a general relativistic MHD action, which, however, describes only “irrotational” flows. Thompson used this Lagrangian in the extreme relativistic limit. Heyl and Hernquist modified it to include QED effects. In this paper we follow mostly Seliger and Witham and Schutz. In Sec. II.A we propose a nonrelativistic MHD Lagrangian, and show in Sec. II.B and II.C that it gives rise to the correct equations of motion for the density, entropy, velocity and magnetic fields in Newtonian MHD. In Sec. II.D we derive from it the conserved circulation, defined in terms of a new vector field $`𝐑`$, and discuss its invariance under redefinition of $`𝐑`$. Sec. II.E furnishes two examples of the conserved circulation in action. In Sec. III.A we collect all the equations of motion of general relativistic MHD, and propose a general relativistic MHD Lagrangian in Sec. III.B. Secs. III.C and III.D recover all the relativistic MHD equations of motion from it. Finally in Sec. III.E we generalize the conserved MHD circulation to the general relativistic case. ## II Variational Principle in Eulerian Coordinates ### A The Lagrangian Density Perfect MHD describes situations where the flow is nondissipative, and, in particular, when the magnetoflow has “infinite conductivity”, and where Maxwell’s displacement current may be neglected in Ampere’s equation. We shall adopt this approximation. We work in eulerian coordinates: all physical quantities are functions of coordinates $`x_i`$ or $`𝐫`$ which describe a fixed point in space. We first summarize the MHD equations. We work in units for which $`c=1`$. First of all, the fluid obeys the equation of continuity ($`_t/t)`$ $$_t\rho +\left(\rho 𝐯\right)=0,$$ (1) where $`\rho (𝐫,t)`$ is the mass density per unit volume of the fluid and $`𝐯(𝐫,t)`$ is the fluid’s velocity field. Second, since there is no dissipation, $`s`$, the entropy per unit mass, must be conserved along the flow: $$Ds_ts+𝐯s=0.$$ (2) Here we have defined the convective derivative $`D`$, which in Cartesian coordinates has the same form for scalars or vectors. With the help of Eq. (1) this equation can be written as $$_t(\rho s)+\left(\rho s𝐯\right)=0.$$ (3) Third, “infinite conductivity” implies that $`𝐄+(𝐯/c)\times 𝐁=0`$, where $`𝐄`$ and $`𝐁`$ are the electric and magnetic fields, respectively. Combining this with Faraday’s equation yields the so called field-freezing equation $$_t𝐁=\times \left(𝐯\times 𝐁\right),$$ (4) which implies Alfven’s law of conservation of the magnetic flux through a closed loop moving with the flow. Finally, the evolution of the velocity field is governed by the MHD Euler equation, $$\rho D𝐯=p\rho U+\frac{(\times 𝐁)\times 𝐁}{4\pi },$$ (5) where $`p`$ is the fluid’s pressure (here assumed isotropic), and $`U(𝐫,t)`$ is the gravitational potential. The least action principle is in general $$\delta S[f_a]\delta 𝑑td^3r(f_a,_tf_a,f_a)=0.$$ (6) Here the action $`S`$ is a functional of various fields $`f_a(𝐫,t)`$, $`a=1,2,\mathrm{}`$. One varies each $`f_a`$, transfers time and space derivatives of each variation $`\delta f_a`$ to the adjacent factor by integration by parts, and sets to zero the overall coefficient of the bare $`\delta f_a`$. This gives us the Lagrange–Euler equation $$_t\left(\frac{}{(_tf_a)}\right)+\left(\frac{}{f_a}\right)\frac{}{f_a}=0.$$ (7) It is usually more convenient to get the equation for each $`f_a`$ ab initio by the above procedure, rather than by using Eq. (7). We now propose the following Lagrangian density for MHD flow of perfect infinitely conducting fluid which incorporates Eqs.(1-4), as three Lagrange constraints $``$ $`=`$ $`\rho 𝐯^2/2\rho ϵ(\rho ,s)\rho U𝐁^2/(8\pi )+`$ (8) $`+`$ $`\varphi [_t\rho +\left[\rho 𝐯\right)]+\eta [_t(\rho s)+\left(\rho s𝐯\right)]`$ (9) $`+`$ $`\lambda \left[_t(\rho \gamma )+\left(\rho \gamma 𝐯\right)\right]+𝐊\left[_t𝐁\times \left(𝐯\times 𝐁\right)\right].`$ (10) In the above $`ϵ(\rho ,s)`$ is the thermodynamic internal energy per unit mass; in the total Lagrangian the corresponding total internal energy enters as a potential energy. The magnetic energy, the volume integral of $`𝐁^2/(8\pi )`$, also enters the total Lagrangian as a potential energy. In Eq. (10) $`\varphi `$, $`\eta `$ are Lagrange multiplier fields which locally enforce the conservation laws (1-2), as may be verified by varying with respect to these multipliers. $`𝐊`$ is a triplet of Lagrange multiplier fields which enforce the field–freezing constraint Eq. (4): varying with respect to $`𝐊`$ reproduces Eq. (4) at every point and time. Finally, $`\lambda `$ is a Lagrange multiplier field which enforces the Lin constraint on a new field, $`\gamma `$: $$_t(\rho \gamma )+(\rho 𝐯\gamma )=0\mathrm{or}D\gamma =0.$$ (11) Here we have used Eq. (1) to reduce to the second form. Lin’s field $`\gamma `$, like $`s`$, is conserved along the flow, but unlike $`s`$ it does not occur elsewhere in the Lagrangian. Lin interprets $`\gamma (𝐫,t)`$ as one of the three initial Lagrangian coordinates which label each fluid element. But whatever the interpretation, the condition (11) is essential so that the flow can be vortical also in the limit $`𝐁0`$. This matter is further discussed in the following section. ### B The Equations of Motion Can our proposed Lagrangian density reproduce all the equations of motion of perfect MHD flow ? We have already seen that it does reproduce Eqs. (1-2) and (4). Let us now vary $`\gamma `$ to get $$D\lambda =0,$$ (12) so that $`\lambda `$, like $`\gamma `$, is conserved with the flow. Both this and Eq. (11) will be essential in demonstrating the existence of the new conserved circulation. Next we vary $`s`$; remembering that $`(ϵ/s)_\rho `$ is just the fluid’s temperature $`T`$, we have $$D\eta =T,$$ (13) which establishes that $`\eta `$ decreases along the flow. The next variation is one with respect to $`\rho `$. Recalling that $`(ϵ/\rho )_s=p/\rho ^2`$, introducing the enthalpy per unit mass $`w=ϵ+p/\rho `$, and using Eqs. (12-13) we get $$D\varphi =v^2/2w+TU.$$ (14) When we vary $`𝐯`$ in the action we may take advantage of the identity $`(𝐀\times 𝐁)=𝐁\times 𝐀𝐀\times 𝐁`$ and Gauss’ theorem to flip the curl operation from $`\delta 𝐯\times 𝐁`$ onto $`𝐊`$. Then the identity $`𝐀𝐁\times 𝐂=𝐁𝐀\times 𝐂`$ helps to shift the $`\delta 𝐯`$ into the position of a factor in a scalar product. We may then factor out the common $`\delta 𝐯`$ and isolate the vector equation $`𝐯`$ $`=`$ $`\varphi +\gamma \lambda +s\eta +𝐐`$ (15) $`𝐐`$ $``$ $`𝐁\times 𝐑/\rho .`$ (16) where $`𝐑\times 𝐊`$. This is neither a solution for $`𝐯`$ ($`\lambda `$ and $`\eta `$ not known), nor an equation of motion ($`𝐯`$ appears undifferentiated). In the next subsection we show that this prescription for $`𝐯`$ leads to the MHD Euler equation (5). Expression (15) shows the importance of including Lin’s field $`\gamma `$. For suppose we consider an unmagnetized fluid in isentropic ($`s=`$ const.) flow. Without $`\gamma `$ the expression for $`𝐯`$ is a perfect gradient, which means the proposed Lagrangian density describes only irrotational flows, a small subset of all possible ones. It is well known that this problem does not appear when one couches the problem in Lagrangian coordinates because one gets then an equation, not for $`𝐯`$, but for the fluid’s acceleration. Lin’s way out of this difficulty is to remember that the initial coordinates of the fluid element are maintained throughout its flow. These coordinates “label” the element, and this can be interpreted as a triplet of constraints (one for each coordinate) of the form $`\lambda _i\left(b_i/t+b_i\right),`$ where $`𝐛`$ is the initial vector coordinate for the element in question. Lundgren used this triplet form for the MHD case. It was later shown (see for example) that the triplet can be reduced to a single constraint with the help of Pfaff’s theorem. One thus returns to form (10) of the Lagrangian density and Eq. (15) for the fluid velocity. The vorticity is now (still excluding $`𝐁`$) $$\omega =\times 𝐯=\gamma \times \lambda +s\times \eta ,$$ (17) so we see that isentropic vortical flow is possible. In the MHD case, the magnetic term in Eq. (15) contributes to the vorticity. Henyey, who suggested a Lagrangian density similar to ours, occasionally dropped the Lin term in the MHD case. However, we shall retain the Lin term throughout. It might seem peculiar at first that adding a constraint like Lin’s permits the appearance of solutions (vortical) which were forbidden before it was imposed. But we must remember that we add to the Lagrangian not only a constraint, but also a new degree of freedom, $`\gamma (𝐫,t)`$, and it is natural that with more degrees of freedom the class of allowed flows will expand. Finally, we vary $`𝐁`$ in the action; by similar manipulation to those which gave Eq. (15) we get $$_t𝐊=𝐯\times 𝐑𝐁/(4\pi ).$$ (18) Taking the curl of this equation we get the more convenient one $$_t𝐑=\times [𝐯\times 𝐑𝐁/(4\pi )]=\times (𝐯\times 𝐑)𝐉.$$ (19) Here $`𝐉=\times 𝐁/4\pi `$ is the electric current density coming from Ampere’s equation. Notice the similarity between Eq. (19) and (4). Eq. (19) says that the rate of change of the flux of $`𝐑`$ through the surface spanning a closed curve carried with the flow equals minus the flux of the electric current density through that curve. ### C The MHD Euler equation We now show that the Lagrangian density (10) yields the correct MHD Euler equation. We first operate with the convective derivative $`D`$ on Eq. (15) remembering that $`Ds=0`$ and $`D\gamma =0`$: $$D𝐯=D\varphi +\gamma D\lambda +sD\eta +D𝐐.$$ (20) We now use the identity $$D=D(𝐯),$$ (21) where in Cartesian coordinates $$[(𝐯)]_i\underset{j}{}\frac{v_j}{x_i}\frac{}{x_j},$$ (22) in conjunction with Eqs. (12-14) to transform Eq. (20) into $`D𝐯`$ $`=`$ $`(v^2/2w+TsU)sTs(𝐯)\eta `$ (23) $``$ $`(𝐯)\varphi \gamma (𝐯)\lambda +D𝐐.`$ (24) From the thermodynamic identity $`dw=Tds+dp/\rho `$ we infer $$w+Ts=p/\rho ,$$ (25) and we also have $`v^2/2=(𝐯)𝐯`$, where the meaning of the right hand side is clear by analogy with Eq. (22). Thus Eq. (24) turns into $$D𝐯=p/\rho U+(𝐯)(𝐯\varphi s\eta \gamma \lambda )+D𝐐.$$ (26) Finally comparing with Eq. (15) we see that the last brackets stand for $`𝐐`$ so that $$D𝐯=p/\rho U+(𝐯)𝐐+D𝐐.$$ (27) Thus, magnetic term aside, we have recovered the Euler equation (5). We now go on to calculate the $`𝐐`$ dependent terms. We may rewrite the equation of continuity (1) as $$D\rho =\rho 𝐯,$$ (28) With this, the Gauss law $`𝐁=0`$ and the identity $`\times (𝐀\times 𝐁)=𝐁𝐀𝐁𝐀𝐀𝐁+𝐀𝐁`$, Eq. (4) may be recast in the well known form $$D(𝐁/\rho )=\left((𝐁/\rho )\right)𝐯.$$ (29) Analogously, because $`𝐑=0`$, Eq. (19) may be put in the form $$D𝐑=(𝐑)𝐯𝐑𝐯𝐉.$$ (30) Therefore, $`(𝐯)𝐐+D𝐐`$ $`=`$ $`(𝐯)(𝐑\times 𝐁/\rho )D𝐑\times 𝐁/\rho 𝐑\times D(𝐁/\rho )`$ (31) $`=`$ $`(𝐯)(𝐑\times 𝐁/\rho )\left((𝐑)𝐯\right)\times (𝐁/\rho )`$ (32) $`+`$ $`(𝐯)𝐑\times (𝐁/\rho )𝐑\times \left((𝐁/\rho )\right)𝐯+𝐉\times 𝐁/\rho .`$ (33) The four terms in the second version of Eq. (33) involving derivatives of $`𝐯`$ can be shown to cancel out by expanding them out in Cartesian coordinates. Hence, Eq. (27) is the magnetic Euler equation with the usual Lorentz force per unit mass, $`𝐉\times 𝐁/\rho `$, in addition to the pure fluid terms. The fact that we obtain the correct MHD equations (1-2), (4) and (5) is testament to the correctness of our proposed Lagrangian density Eq. (10). Note that Lin’s field $`\gamma `$ has disappeared from the final equation of motion. ### D Circulation Conservation Law With the help of the above formalism, we can now prove the existence of a generalization of Kelvin’s circulation theorem applicable to perfect MHD. Let us calculate the line integral of the vector $$𝐙=𝐯+𝐑\times 𝐁/\rho $$ (34) along a closed curve $`𝒞`$ drifting with the fluid: $$\mathrm{\Gamma }=_𝒞𝐙𝑑𝐫.$$ (35) According to Eq. (15) this integral is $$\mathrm{\Gamma }=_𝒞\varphi d𝐫+_𝒞\gamma \lambda d𝐫+_𝒞s\eta d𝐫.$$ (36) The term involving $`\varphi `$ obviously vanishes (we assume all the Lagrange multipliers are single valued). For like reason so does the term involving $`\eta `$ in the isentropic ($`s=`$ const ) case as $`s`$ can be taken out of the integral. The middle integral can be written $`_𝒞\gamma 𝑑\lambda `$, where $`d\lambda \lambda d𝐫`$. But Eqs. (11-12) tell us that both $`\gamma `$ and $`\lambda `$ are conserved along the flow. Hence $`\mathrm{\Gamma }`$ remains constant as $`𝒞`$ drifts along with the flow. Since in the limit $`𝐁0`$, $`\mathrm{\Gamma }`$ becomes Kelvin’s circulation, we have found an extension of Kelvin’s theorem to perfect MHD. Obviously the conservation of $`\mathrm{\Gamma }`$ implies the conservation of the flux of $`\times 𝐙`$ through $`𝒞`$. The vector field $`𝐑`$ is not unique for a given physical situation. For example, the change $`𝐑𝐑+k𝐁`$ ($`k`$ a real constant) leaves invariant all equations of motion, Eqs. (11-16), (19), and (27), as well as the conserved circulation expressions (34-35). In addition, suppose that at time $`t=0`$ we define an arbitrary solenoidal (divergence–free) field $`𝐛`$ all over the flow, and then evolve it in time as a passive vector, i.e., in accordance with the frozen–in field equation (4). Comparing with Eq. (19) we see that $`𝐑+k𝐛`$ and $`𝐑`$ obey the same equation, and both are permanently solenoidal \[this property is obviously preserved by Eqs. (4) and (19) in the MHD approximation\]. If in $`𝐙`$ we use $`𝐑+k𝐛`$ in lieu of $`𝐑`$ to construct the conserved circulation, $`\mathrm{\Gamma }`$ gets the additional contribution $$\mathrm{\Delta }\mathrm{\Gamma }=k_𝒞\left(𝐛\times 𝐁/\rho \right)𝑑𝐫=k_𝒞𝐁\left(d𝐫\times 𝐛/\rho \right)$$ (37) Here we have used a well known vector identity. Now by analogy with $`𝐁`$, $`𝐛`$ obeys Eq. (29) which tells us that any two elements of the fluid permanently lie on one and the same line of $`𝐛/\rho `$, and their distance, if small, is proportional to $`|𝐛|/\rho `$. We can always make $`𝐛`$ small. Then $`d𝐫\times 𝐛/\rho `$ is a vectorial element of area of a narrow closed strip carried along by the fluid, one of whose edges coincides with $`𝒞`$. The integral in Eq. (37) is just the flux of magnetic induction through this strip (not through the space bounded by the strip), and we know this is conserved by virtue of Alfven’s law. Thus with the change $`𝐑𝐑+k𝐛`$ we added some conserved magnetic flux to $`\mathrm{\Gamma }`$, and did not get a new conserved circulation. The MHD flow $`\{𝐁,𝐯,\rho ,p\}`$ is evidently unchanged because the MHD Euler equation (5) does not contain $`𝐑`$, so we must conclude that in the expression for $`𝐯`$, Eq. (15-16), the change of the $`𝐐`$ term must be compensated by suitable changes in the Lagrange multipliers $`\varphi +s\eta `$ and $`\lambda `$ (recall that we are working with $`s=\mathrm{const}.`$). Indeed, the initial choice of $`𝐛`$ involves a choice of two functions because of the $`𝐛=0`$ constraint, so that the two functions $`\varphi +s\eta `$ and $`\lambda `$ are just enough to absorb the change $`𝐑𝐑+k𝐛`$ thus generated and leave $`𝐯`$ unchanged. It is not possible to eliminate $`𝐑`$ altogether by the change $`𝐑𝐑+k𝐛`$ because $`𝐑`$ and $`𝐛`$ obey different equations. This means the circulation conservation law we have found cannot be reduced to an Alfven type law; it is a new law. In Sec.III E we shall discuss the freedom inherent in $`𝐑`$ by a covariant procedure. Fixing the freedom is a necessary step in any attempt to exhibit explicitly the conserved circulation. ### E Examples First consider a situation where the fluid is isentropic but not flowing: $`𝐯=0`$. It follows from Eq. (1) that $`\rho =\rho _0(𝐫)`$, and from Eq. (4) that $`𝐁=𝐁_0(𝐫)`$. From these facts and Eq. (19) we see that $$𝐑=t\times 𝐁_0(𝐫)/(4\pi )+𝐑_0(𝐫).$$ (38) Although the physical quantities are stationary, $`𝐑`$ is not. This is so because like the electromagnetic potential, $`𝐑`$ is not a measurable quantity, being subject to “gauge changes” $`𝐑𝐑+𝐛`$ as already discussed. According to Eq. (34) the conserved circulation (around a contour fixed in space because $`𝐯=0`$) should be $$\mathrm{\Gamma }=t_𝒞\frac{(\times 𝐁_0)\times 𝐁_0}{4\pi \rho _0}𝑑𝐫+_𝒞\frac{𝐑_0\times 𝐁_0}{4\pi \rho _0}𝑑𝐫.$$ (39) On the face of it, the time dependence of the first term in this simple situation puts the claimed circulation conservation law in jeopardy. However, according to the magnetic Euler equation (5), the first integrand here is equal to $`U+p/\rho _0`$ which, by virtue of Eq. (21) and the isentropic nature of the fluid, is a perfect gradient (for isentropic flow $`p/\rho _0=w`$). Hence the first integral vanishes, and the circulation is indeed time independent as required by our theorem. As a second example consider an axisymmetric differentially rotating fluid exhibiting a purely poloidal magnetic field. Let the flow also be isentropic and stationary. We choose to work in cylindrical coordinates $`\{\varrho ,\varphi ,z\}`$; the hat symbol will denote a unit vector in the stated direction. It then follows that $`\rho =\rho _0(𝐫)`$, $`𝐁=𝐁_0(𝐫)`$ and $`𝐯=\mathrm{\Omega }\varrho \widehat{\varphi }`$, where $`\mathrm{\Omega }(\varrho ,z)`$ is the angular velocity of the fluid. It is well known that for axisymmetric fields the curl of a poloidal field is a toroidal one, and the toroidal field has only a $`\widehat{\varphi }`$ component. Therefore, the electric current density $`𝐉=\times 𝐁/(4\pi )`$, is everywhere collinear with $`𝐯`$ and time independent. Since the problem is stationary, $`\mathrm{\Omega }`$ satisfies Ferraro’s law of isorotation $`𝐁\mathrm{\Omega }=0`$. In addition the field must be torque-free i.e. no Lorentz force in the $`\widehat{\varphi }`$ direction. This condition is identically satisfied for a purely poloidal field. Combining all of the above we get the following solution of Eq. (19): $$𝐑=t𝐉$$ (40) According to Eq. (34) the conserved circulation should be $$\mathrm{\Gamma }=_𝒞\mathrm{\Omega }\varrho ^2𝑑\varphi t_𝒞\frac{𝐉\times 𝐁_0}{4\pi \rho _0}𝑑𝐫$$ (41) where we have exploited the axisymmetry to rewrite the first term. We now verify that this circulation is indeed conserved. Because of the differential rotation, the contour $`𝒞`$ is gradually deformed in the azimuthal direction. The difference $`d\varphi `$ in the azimuthal coordinates between two infinitesimally close fluid elements lying on $`𝒞`$ can be written as $`d\varphi =d\varphi _0+td\mathrm{\Omega }`$ where $`d\varphi _0`$ is the initial difference in azimuthal coordinates while $`d\mathrm{\Omega }`$ is the difference between the elements’ angular velocities. Hence we have $$_𝒞\mathrm{\Omega }\varrho ^2𝑑\varphi =_𝒞\mathrm{\Omega }\varrho ^2𝑑\varphi _0+t_𝒞\mathrm{\Omega }\varrho ^2𝑑\mathrm{\Omega }.$$ (42) Note that the first term is time independent while the second one is linear in time. The magnetic Euler equation (5) in cylindrical coordinate reads $$\mathrm{\Omega }^2\varrho \widehat{\varrho }=\frac{p}{\rho _0}U+\frac{𝐉\times 𝐁_0}{4\pi \rho _0}.$$ (43) Again, by the isentropic condition we can write $`p/\rho _0=w`$. Taking the integral round $`𝒞`$ of both sides of Eq. (43) we have $$_𝒞\mathrm{\Omega }^2\varrho 𝑑\varrho =_𝒞\frac{𝐉\times 𝐁_0}{4\pi \rho _0}𝑑𝐫.$$ (44) Substituting from Eq. (44) and Eq. (42) into Eq. (41) we get $`\mathrm{\Gamma }`$ $`=`$ $`{\displaystyle _𝒞}\mathrm{\Omega }\varrho ^2𝑑\varphi _0+t{\displaystyle _𝒞}\mathrm{\Omega }\varrho ^2𝑑\mathrm{\Omega }+t{\displaystyle _𝒞}\mathrm{\Omega }^2\varrho 𝑑\varrho `$ (45) $`=`$ $`{\displaystyle _𝒞}\mathrm{\Omega }\varrho ^2𝑑\varphi _0+{\displaystyle \frac{t}{2}}{\displaystyle _𝒞}d(\mathrm{\Omega }^2\varrho ^2)={\displaystyle _𝒞}\mathrm{\Omega }\varrho ^2𝑑\varphi _0,`$ (46) and $`\mathrm{\Gamma }`$ is indeed time independent. Note that it is possible to add to $`𝐑`$ in Eq. (40) an arbitrary time independent solenoidal vector field $`𝐑_0(𝐫)`$ which satisfies $`𝐑_0\times 𝐯=\chi `$. However, as already stressed in the previous subsection, this will only add to $`\mathrm{\Gamma }`$ a time independent quantity. It is important to note that although the example specifically relates to an axisymmetric problem, Eq. (40) applies to all stationary MHD flows which have $`𝐉`$ collinear with $`𝐯`$. Accordingly, $`\mathrm{\Gamma }`$ will be conserved in all such flows. ## III Relativistic Variational principle In this section we formulate a Lagrangian density for MHD flow in the framework of general relativity (GR). Greek indices run from 0 to 3. The coordinates are denoted $`x^\alpha =(x^0,x^1,x^2,x^3)`$; $`x^0`$ stands for time. A comma denotes the usual partial derivative; a semicolon covariant differentiation. Our signature is $`\{,+,+,+\}`$. We continue to take $`c=1`$. ### A Relativistic MHD Equations The general relativistic (GR) equations for MHD were formulated by Lichnerowicz, Novikov and Thorne, Carter, Bekenstein and Oron and others. The role of the mass conservation equation (1) is taken over by the law of particle number conservation, $$N^\alpha {}_{;\alpha }{}^{}=\left(nu^\alpha \right)_{;\alpha }=0,$$ (47) where $`N^\alpha `$ is the particle number 4–current density, $`n`$ the particle proper number density and $`u^\alpha `$ the fluid 4–velocity field normalized by $`u^\alpha u_\alpha =1`$. If $`s`$ represents the entropy per particle (not per unit mass as in Sec. II), then the assumption of ideal adiabatic flow, Eq. (2), can be put in the form $$\left(sN^\alpha \right)_{;\alpha }=0\mathrm{or}u^\alpha s_{,\alpha }=0.$$ (48) Because the flow is assumed adiabatic, the energy momentum tensor for the magnetized fluid is that of an ideal fluid augmented by the electromagnetic energy–momentum tensor: $$T^{\alpha \beta }=pg^{\alpha \beta }+(p+\rho )u^\alpha u^\beta +(F^{\alpha \gamma }F^\beta {}_{\gamma }{}^{}\frac{1}{4}F^{\gamma \delta }F_{\gamma \delta }g^{\alpha \beta })/(4\pi ).$$ (49) Here $`\rho `$ represents the fluid’s energy proper density (including rest masses) and $`p`$ the scalar pressure (again assumed isotropic), while $`F^{\alpha \beta }`$ denotes the electromagnetic field tensor. As usual the covariant divergence $`T^{\alpha \beta }_{;\beta }`$ must vanish (energy–momentum conservation). In consequence $`T^{\alpha \beta }{}_{;\beta }{}^{}+u^\alpha u_\gamma T^{\gamma \beta }{}_{;\beta }{}^{}=0`$ which can be recast as $$(\rho +p)u^\beta u^\alpha {}_{;\beta }{}^{}=(g^{\alpha \beta }+u^\alpha u^\beta )p_{,\beta }+F^{\alpha \beta }F_\beta {}_{}{}^{\gamma }{}_{;\gamma }{}^{}/(4\pi ).$$ (50) The term $`a^\alpha u^\beta u^\alpha _{;\beta }`$ stands for the fluid’s acceleration 4–vector. The effects of gravitation are automatically included by the appeal to curved metric and covariant derivatives. This equation parallels Eq. (5);as usual in GR the pressure contributes alongside the energy density to the inertia. The electromagnetic field tensor is subject to Maxwell’s equations $`F^{\alpha \beta }_{;\beta }`$ $`=`$ $`4\pi j^\alpha `$ (51) $`F_{\alpha \beta ,\gamma }+F_{\beta \gamma ,\alpha }+F_{\gamma \alpha ,\beta }`$ $`=`$ $`0.`$ (52) where $`j^\alpha `$ denotes the electric 4–current density. Putting all this together we have the GR MHD Euler equation $$(\rho +p)a^\alpha =h^{\alpha \beta }p_{,\beta }+F^{\alpha \beta }j_\beta ,$$ (53) where we have introduced the projection tensor $$h^{\alpha \beta }g^{\alpha \beta }+u^\alpha u^\beta .$$ (54) The above equations do not completely specify MHD flow (as opposed to flow of a generic magnetofluid). For any flow carrying an electromagnetic field, the (antisymmetric) Faraday tensor $`F_{\alpha \beta }`$ may be split into electric and magnetic vectors with respect to the flow: $`E_\alpha `$ $`=`$ $`F_{\alpha \beta }u^\beta `$ (55) $`B_\alpha `$ $`=`$ $`{}_{}{}^{}F_{\beta \alpha }^{}u^\beta {\displaystyle \frac{1}{2}}ϵ_{\beta \alpha \gamma \delta }F^{\gamma \delta }u^\beta .`$ (56) Here $`ϵ_{\alpha \beta \gamma \delta }`$ is the Levi-Civita totally antisymmetric tensor ($`ϵ_{0123}=(g)^{1/2}`$ with $`g`$ denoting the determinant of the metric $`g_{\alpha \beta }`$) and $`{}_{}{}^{}F_{\alpha \beta }^{}`$ is the dual of $`F_{\alpha \beta }`$. In the frame moving with the fluid, these 4–vectors have only spatial parts which correspond to the usual $`𝐄`$ and $`𝐁`$, respectively. The inversion of Eqs. (55-56) is $$F_{\alpha \beta }=u_\alpha E_\beta u_\beta E_\alpha +ϵ_{\alpha \beta \gamma \delta }u^\gamma B^\delta $$ (57) For an infinitely conducting (perfect MHD) fluid, the electric field in the fluid’s frame must vanish, i.e., $$E_\alpha =F_{\alpha \beta }u^\beta =0.$$ (58) This corresponds to the usual MHD condition $`𝐄+𝐯\times 𝐁=0`$. ### B Relativistic Lagrangian Density Inspired by Schutz’s Lagrangian density for pure fluids in GR, we now propose a Lagrangian density for GR MHD flow which reproduces Eqs. (47-48), (51-53) and (58). Like Schutz we include Lin’s term, which proves essential to our subsequent proof of the existence of a circulation theorem. The proposed Lagrangian density is $$=\rho (n,s)F_{\alpha \beta }F^{\alpha \beta }/(16\pi )+\varphi N^\alpha {}_{;\alpha }{}^{}+\eta \left(sN^\alpha \right)_{;\alpha }+\lambda \left(\gamma N^\alpha \right)_{;\alpha }+\tau ^\alpha F_{\alpha \beta }N^\beta .$$ (59) Now in GR the scalar density $`(g)^{1/2}`$ replaces $``$ in the action (6), and is what enters in the Euler–Lagrange equations (7). The covariant derivatives cause no problem; for example $`(g)^{1/2}\varphi N^\alpha {}_{;\alpha }{}^{}=\varphi [(g)^{1/2}N^\alpha ]_{,\alpha }`$, whose variation with respect to $`N^\alpha `$ is easily integrated by parts. As in the nonrelativistic case, $`\varphi `$ is the Lagrange multiplier associated with the conservation of particle number constraint, Eq. (47), $`\eta `$ is that multiplier associated with the adiabatic flow constraint, Eq. (48), and $`\lambda `$ is that associated with the conservation along the flow of Lin’s quantity $`\gamma `$. We view $`\gamma `$, $`N^\alpha `$ and $`s`$ as the independent fluid variables, while $`n`$ and $`u^\alpha `$ are determined by the obvious relations $$N^\alpha N_\alpha =n^2;u^\alpha =n^1N^\alpha .$$ (60) Strictly speaking, one should include in $``$ a new Lagrange multiplier times the constrained expression $`N^\alpha N_\alpha +n^2`$. Rather than clutter up $``$ further, we enforce this constraint below by hand. As usual, we view the vector potential $`A_\alpha `$, rather than the electromagnetic field tensor $`F_{\alpha \beta }=A_{\beta ;\alpha }A_{\alpha ;\beta }=A_{\beta ,\alpha }A_{\alpha ,\beta }`$, as the independent electromagnetic variable. In consequence, the Maxwell Eqs. (52) are satisfied as identities. The last term in $``$ enforces the “vanishing of electric field” constraint, Eq. (58); $`\tau ^\alpha `$ is a Lagrange multiplier 4–vector field. Because here we enforce the “vanishing of electric field” rather than the equivalent flux freezing condition (4), the $`\tau ^\alpha `$ is more like $`𝐑`$ of Sec. II.B. than like $`𝐊`$. Not all of $`\tau ^\alpha `$ is physically meaningful. For suppose we add an arbitrary function $`f(x^\beta )`$ multiplied by $`N^\alpha `$ to $`\tau ^\alpha `$. This increments the Lagrangian density by $`fN^\alpha F_{\alpha \beta }N^\beta `$ which vanishes identically by the antisymmetry of $`F_{\alpha \beta }`$. So $`\tau ^\alpha `$ and $`\tau ^\alpha +fN^\alpha `$ are physically equivalent. We shall exploit this to substract from $`\tau ^\alpha `$ its component along $`u^\alpha `$. So henceforth we may take it that $`\tau ^\alpha u_\alpha =0`$. Much freedom is still left in $`\tau ^\alpha `$. Suppose we add to it a term proportional to $`n^1B^\alpha `$. By Eqs. (55-57), this adds to the Lagrangian density the term $`E_\alpha B^\alpha `$. Of course we cannot take this to vanish at the Lagrangian’s level because we have not yet obtained the freezing-in condition (58) from it. However, it is known that $`B^\alpha E_\alpha =\frac{1}{4}ϵ^{\alpha \beta \gamma \delta }F_{\alpha \beta }F_{\gamma \delta }`$. By introducing the potential $`A_\alpha `$ we can write this as $`\frac{1}{2}\left[ϵ^{\alpha \beta \gamma \delta }F_{\alpha \beta }A_\gamma \right]_{;\delta }\frac{1}{2}ϵ^{\alpha \beta \gamma \delta }F_{\alpha \beta ;\delta }A_\gamma `$, where we have used the fact that $`ϵ^{\alpha \beta \gamma \delta }`$ has vanishing covariant derivatives. Obviously the last term vanishes by virtue of Maxwell’s equations (52) which are identities in the present approach. When multiplied by $`(g)^{1/2}`$, the first term becomes a perfect derivative. Such term, when added to the integral forming the Lagrangian, is known not to affect its physical content. Thus $`\tau ^\alpha `$ and $`\tau ^\alpha +\mathrm{const}.\times n^1B^\alpha `$ are physically equivalent, and this transformation respects the condition $`\tau _\alpha u^\alpha =0`$ because $`B_\alpha u^\alpha =0`$ \[see (56)\]. However, there is not enough freedom in the constant to allow us to eliminate the component of $`\tau ^\alpha `$ along $`B^\alpha `$. But in Sec.III E we shall exploit what we have just found. ### C Equations of Motion We can now derive the equations of motion. Variation of $`\varphi `$ recovers the conservation of particles $`N^\alpha _{;\alpha }`$. Variation of $`\lambda `$ with subsequent use of the previous result yields $$\gamma _{,\alpha }u^\alpha =0.$$ (61) If we vary $`\gamma `$ we get $$\lambda _{,\alpha }u^\alpha =0.$$ (62) These two results are precise analogs of Eqs. (11) and (12); they inform us that $`\gamma `$ and $`\lambda `$ are both locally conserved with the flow. In view of the thermodynamic relation $`n^1(\rho /s)_n=T`$, with $`T`$ the locally measured fluid temperature, variation of $`s`$ gives $$u^\alpha \eta _{,\alpha }=T;$$ (63) this is the analogue of Eq. (13). We now vary $`N^\alpha `$ using the obvious consequence of Eq. (60), $$\delta n=u_\alpha \delta N^\alpha ,$$ (64) together with the thermodynamic relation involving the specific enthalpy $`\mu `$, $$\mu (\rho /n)_s=n^1(\rho +p);$$ (65) we get the GR analog of Eq. (15), $$\mu u_\alpha =\varphi _{,\alpha }+s\eta _{,\alpha }+\gamma \lambda _{,\alpha }+\tau ^\beta F_{\alpha \beta }.$$ (66) The importance of Lin’s $`\gamma `$ is again clear here; in the pure isentropic fluid case ($`F^{\alpha \beta }=0`$ and $`s=`$ const.), the Khalatnikov vorticity tensor given by $$\omega _{\alpha \beta }=\left(\mu u_\beta \right)_{,\alpha }\left(\mu u_\alpha \right)_{,\beta }=\left(\gamma \lambda _{,\beta }\right)_{,\alpha }\left(\gamma \lambda _{,\alpha }\right)_{,\beta }$$ (67) would vanish in the absence of $`\gamma `$, thus constraining us to discuss only irrotational flow. By contracting Eq. (66) with $`u^\alpha `$ and using $`u_\alpha u^\alpha =1`$ as well as Eqs. (58) and (62-63), we get the following GR version of Eq. (14): $$\varphi _{,\alpha }u^\alpha =\mu +Ts.$$ (68) Thus the proper time rate of change of $`\varphi `$ along the flow is just minus the specific Gibbs energy or chemical potential. The apparent difference between the result here and Eq. (14) stems from the fact that proper time rate (here) and coordinate time rate (there) differ by gravitational redshift and time dilation effects. These effects are not noticeable when one compares Eqs. (63) with (13) because the first refers to locally measured temperature and the second to global temperature; these two temperatures differ by the same factors as do proper and coordinate time. Turn now to the variation of $`A_\alpha `$. Because of the antisymmetry of $`F_{\alpha \beta }`$, the last term of the Lagrangian, Eq. (59), can be written as $`(\tau ^\beta N^\alpha \tau ^\alpha N^\beta )A_{\alpha ,\beta }`$. The variation of $`A_\alpha `$ in the corresponding term in the action produces, after integration by parts, the term $`\left[(g)^{1/2}(\tau ^\alpha N^\beta \tau ^\beta N^\alpha )\right]_{,\beta }\delta A_\alpha `$. Because for any antisymmetric tensor $`t^{\alpha \beta }`$, $`(g)^{1/2}t^{\alpha \beta }{}_{;\beta }{}^{}=[(g)^{1/2}t^{\alpha \beta }]_{,\beta }`$, this finally leads to the equation $$F^{\alpha \beta }{}_{;\beta }{}^{}=4\pi (\tau ^\alpha N^\beta \tau ^\beta N^\alpha )_{;\beta }.$$ (69) Comparison with Eq. (51) shows that this result gives us a representation of the electric current density $`j^\alpha `$ as the divergence of the bivector $`\tau ^\alpha N^\beta \tau ^\beta N^\alpha `$. Such representation makes the conservation of charge ($`j^\alpha {}_{;\alpha }{}^{}=0)`$ an identity because the divergence of the divergence of any antisymmetric tensor vanishes. This equation is the GR analogue of Eq (19). Formally Eq. (69) determines the Lagrange multiplier 4–vector $`\tau ^\alpha `$, modulo the freedom inherent in it. ### D MHD Euler Equation in General Relativity Our central task now is to show that the equations in Sec. III.C lead uniquely to the GR MHD Euler equation (53). We begin by writing the Khalatnikov vorticity $`\omega _{\beta \alpha }`$ in two forms, $$\omega _{\beta \alpha }=\mu _{,\beta }u_\alpha \mu _{,\alpha }u_\beta +\mu u_{\alpha ;\beta }\mu u_{\beta ;\alpha },$$ (70) as well as by means of Eq. (66) $`\omega _{\beta \alpha }`$ $`=`$ $`s_{,\beta }\eta _{,\alpha }s_{,\alpha }\eta _{,\beta }+\gamma _{,\beta }\lambda _{,\alpha }\gamma _{,\alpha }\lambda _{,\beta }`$ (71) $`+`$ $`\tau ^\delta {}_{;\beta }{}^{}F_{\alpha \delta }^{}\tau ^\delta {}_{;\alpha }{}^{}F_{\beta \delta }^{}+\tau ^\delta F_{\alpha \delta ;\beta }\tau ^\delta F_{\beta \delta ;\alpha }.`$ (72) Contracting the left hand side of the first with $`N^\alpha `$, recalling Eq. (60) and that by normalization $`u^\alpha u_{\alpha ;\beta }=0`$ whereas $`u^\beta u_{\alpha ;\beta }=a_\alpha `$, the fluid’s 4–acceleration, we get $$\omega _{\beta \alpha }N^\alpha =n\mu _{,\beta }n\mu _{,\alpha }u^\alpha u_\beta n\mu a_\beta =nh_\beta {}_{}{}^{\alpha }\mu _{,\alpha }^{}n\mu a_\beta .$$ (73) Now contracting Eq. (71) with $`N^\alpha `$ and using Eqs. (61-63) and (58) to drop a number of terms we get $$\omega _{\beta \alpha }N^\alpha =nTs_{,\beta }\tau ^\delta {}_{;\alpha }{}^{}F_{\beta \delta }^{}N^\alpha +\tau ^\delta F_{\alpha \delta ;\beta }N^\alpha \tau ^\delta F_{\beta \delta ;\alpha }N^\alpha .$$ (74) By virtue of Eq. (48), $`nTs_{,\beta }`$ is the same as $`nTh_\beta {}_{}{}^{\alpha }s_{,\alpha }^{}`$. It is convenient to use the thermodynamic identity $`d\mu =n^1dp+Tds`$, which follows from Eq. (65) and the first law $`d(\rho /n)=Tdspd(1/n)`$, to replace in Eq. (74) $`nTs_{,\beta }`$ by $`h_\beta {}_{}{}^{\alpha }(n\mu _{,\alpha }+p_{,\alpha })`$. Equating our two expressions for $`\omega _{\beta \alpha }N^\alpha `$ gives, after a cancellation, $$\left(n\mu a_\beta +h_\beta {}_{}{}^{\alpha }p_{,\alpha }^{}\right)=\tau ^\delta {}_{;\alpha }{}^{}F_{\beta \delta }^{}N^\alpha +\tau ^\delta F_{\alpha \delta ;\beta }N^\alpha \tau ^\delta F_{\beta \delta ;\alpha }N^\alpha .$$ (75) The last two terms in this equation can be combined into a single one by virtue of Eq. (52), which, as well known, can be written with covariant as well as ordinary derivatives. Further, by Eq. (65) we may replace $`n\mu `$ by $`\rho +p`$. In this manner we get $$\left(\rho +p\right)a_\beta =h_\beta {}_{}{}^{\alpha }p_{,\alpha }^{}+F_{\beta \alpha ;\delta }\tau ^\delta N^\alpha +F_{\beta \delta }\tau ^\delta {}_{;\alpha }{}^{}N_{}^{\alpha }.$$ (76) The term $`\tau ^\delta {}_{;\alpha }{}^{}N_{}^{\alpha }`$ here can be replaced by two others with help of Eq. (69) if we take into account that $`N^\beta {}_{;\beta }{}^{}=0`$: $$(\rho +p)a_\beta =h_\beta {}_{}{}^{\alpha }p_{,\alpha }^{}+F_{\beta \delta }F^{\delta \alpha }{}_{;\alpha }{}^{}/(4\pi )+F_{\beta \alpha ;\delta }\tau ^\delta N^\alpha +F_{\beta \delta }\left(\tau ^\alpha N^\delta \right)_{;\alpha }.$$ (77) We note that the last two terms on the right hand side combine into $`\left(F_{\beta \alpha }N^\alpha \tau ^\delta \right)_{;\delta }`$ which vanishes by Eq. (58). Now substituting from the Maxwell equations (51) we arrive at the final equation $$\left(\rho +p\right)a_\beta =h_\beta {}_{}{}^{\alpha }p_{,\alpha }^{}+F_{\beta \delta }j^\delta ,$$ (78) which is the correct GR MHD Euler equation (53). We have not used any information about $`\tau ^\alpha `$ beyond Eq. (69); hence Euler’s equation is valid for all choices of $`\tau ^\alpha `$. Since we are able to obtain all equations of motion for GR MHD from our Lagrangian density, we may regard it as correct, and go on to look at some consequences. ### E New Circulation Conservation Law Eqs. (66) and (61-62) allow us to generalize the conserved circulation of Sec. II.D to relativistic perfect MHD. Let $`\mathrm{\Gamma }`$ be the line integral $$\mathrm{\Gamma }=_𝒞z_\alpha 𝑑x^\alpha ,$$ (79) where $`𝒞`$ is, again, a closed curve drifting with the fluid, and $$z_\alpha \mu u_\alpha \tau ^\beta F_{\alpha \beta }.$$ (80) According to Eq. (66), $`z_\alpha =\varphi _{,\alpha }+s\eta _{,\alpha }+\gamma \lambda _{,\alpha }`$. Since $`\varphi _{,\alpha }`$ is a gradient, its contribution to $`\mathrm{\Gamma }`$ vanishes. Likewise, for isentropic flow ($`s=`$ const.) the term involving $`s\eta _{,\alpha }`$ makes no contribution to $`\mathrm{\Gamma }`$. Thus $$\mathrm{\Gamma }=_𝒞\gamma \lambda _{,\alpha }𝑑x^\alpha =_𝒞\gamma 𝑑\lambda .$$ (81) By Eqs. (61-62) both $`\gamma `$ and $`\lambda `$ are conserved with the flow. Thus $`\mathrm{\Gamma }`$ is conserved along the flow. Note that by virtue of $`\gamma `$’s presence, $`\mathrm{\Gamma }`$ need not vanish. In the absence of electromagnetic fields and in the nonrelativistic limit $`(\mu m`$ where $`m`$ is a fluid particle’s rest mass), $`\mathrm{\Gamma }`$ for a curve $`𝒞`$ taken at constant time reduces to Kelvin’s circulation. On this ground our result can be considered a generalization of Kelvin’s circulation theorem to general relativistic MHD. We have gone here beyond Oron’s original result in that no symmetry is necessary for the circulation to be conserved. To manifestly exhibit the conserved circulation, one has to know $`\tau ^\alpha `$ explicitly. The first step is to understand the freedom left in $`\tau ^\alpha `$ beyond that discussed in Sec. III B. The second is to determine $`\tau ^\alpha `$ in a specific flow exploiting for this the symmetries and other information. Below we address the first step; the second is left mainly to future publications. Given a specific MHD flow as background, let us at define a generic test field $`f_{\alpha \beta }=f_{\beta \alpha }`$ which satisfies Maxwell’s homogeneous equations (52) as well as the freezing-in condition (58), e.g. $`e_\alpha f_{\alpha \beta }u^\beta =0`$. We think of $`f_{\alpha \beta }`$ as very weak, so that it does not disturb the MHD flow or the spacetime geometry; it is a passive tensor. Because $`f_{\alpha \beta }u^\beta =0`$, $`f_{\alpha \beta }`$ has only three independent components. Therefore, its full content is reflected in the “magnetic 4-vector” $`b_\alpha \frac{1}{2}ϵ_{\beta \alpha \gamma \delta }f^{\gamma \delta }u^\beta `$, which is obviously orthogonal to $`u_\alpha `$. The transformation $`\tau ^\alpha \tau ^\alpha +kn^1b^\alpha `$ ($`k`$ a real constant) is not a symmetry of the Lagrangian. However, it does not disturb the inhomogeneous Maxwell equations (51,69). This is because the change in $`\tau ^\alpha `$ merely adds to the electric current the term $`\left(b^\alpha u^\beta b^\beta u^\alpha \right)_{;\beta }=(g)^{1/2}\left[(g)^{1/2}(b^\alpha u^\beta b^\beta u^\alpha )\right]_{,\beta }`$. Because of the condition $`e^\alpha =0`$, we may easily invert the analog of (57) to get $`b^\alpha u^\beta b^\beta u^\alpha =\frac{1}{2}ϵ^{\alpha \beta \gamma \delta }f_{\gamma \delta }`$. But since $`(g)^{1/2}ϵ^{\alpha \beta \gamma \delta }`$ is just the constant antisymmetric symbol, our assumed equations $`f_{\alpha \beta ,\gamma }+f_{\gamma \alpha ,\beta }+f_{\beta \gamma ,\alpha }=0`$ imply that $`\left(b^\alpha u^\beta b^\beta u^\alpha \right)_{;\beta }=0`$ so that the Maxwell equations (69) are invariant under $`\tau ^\alpha \tau ^\alpha +kn^1b^\alpha `$. So is the Euler equation since its derivation used only the information about $`\tau ^\alpha `$ inherent in (69). The expression for $`u_\alpha `$, (66), does seem to change under $`\tau ^\alpha \tau ^\alpha +kn^1b^\alpha `$, and we also note that $`\mathrm{\Gamma }\mathrm{\Gamma }+kn^1b^\beta F_{\alpha \beta }𝑑x^\alpha `$. Now since the “magnetic 4-vector” $`b^\alpha `$ is frozen in, like any such infinitesimal field, it evolves in such a way that $`n^1b^\alpha `$ gives for all time that part of the spacetime separation of two neighboring fluid elements which is orthogonal to $`u^\alpha `$, c.f. discussion after Eq. (37). Thus $`n^1b^\alpha `$ can be employed to define a thin closed strip dragged with the fluid such that one of its edges coincides with the curve $`𝒞`$. Therefore, the increment $`n^1b^\beta F_{\alpha \beta }𝑑x^\alpha `$ is just the conserved magnetic flux through this strip. Evidently the transformation $`\tau ^\alpha \tau ^\alpha +kn^1b^\alpha `$ has not changed the nature of the conservation law for $`\mathrm{\Gamma }`$, but only added a trivially conserved quantity to it.. Now the MHD flow $`\{B^\alpha ,u^\alpha ,n,\rho ,\mu \}`$ is evidently unchanged because neither the MHD Euler equation (50) nor Maxwell’s equations were changed, so we must conclude that in the expression for $`u^\alpha `$, Eq. (66), the change of the $`\tau ^\beta F_{\alpha \beta }`$ term must be compensated by suitable changes in the pair of Lagrange multipliers $`\varphi +s\eta `$ and $`\lambda `$ (since we are assuming $`s=\mathrm{const}.`$). They are capable of this because $`b^\alpha `$ has only two independent components. For the condition $`b^\alpha u_\alpha =0`$ eliminates one of the four. In addition $`b^\alpha `$ comes from $`f_{\alpha \beta }`$ which satisfies Eqs. (52); in particular, $`f_{12,3}+f_{31,2}+f_{23,1}=0`$ in the chosen coordinates. But since no time derivatives appear in it, this last equation serves as an initial constraint on $`b^\alpha `$ just as the Gauss equation $`𝐁=0`$ does for the true magnetic field. Accordingly, one further relation exists between components of $`b^\alpha `$ so that the generic $`b^\alpha `$ contains only two free functions. Thus the change in $`\tau ^\beta F_{\alpha \beta }`$ can be taken up by changes in the two functions $`\varphi +s\eta `$ and $`\lambda `$ so that $`\mu u_\alpha `$ is unchanged. Note that it is not possible to “get rid” of $`\tau ^\alpha `$ by means of the transformation $`\tau ^\alpha \tau ^\alpha +kn^1b^\alpha `$ because, as we shall make clear presently, $`\tau ^\alpha `$ and $`b^\alpha `$ obey different equations of motion. Thus there must be a residual part of $`\tau ^\alpha `$ which is not changed by the transformations. It is this part which is responsible for the conserved circulation, so that the conservation of $`\mathrm{\Gamma }`$ cannot be reduced to magnetic flux conservation. The following algorithm can be used to find $`\tau ^\alpha `$. Maxwell’s inhomogeneous equations (69) which say that the divergence of a certain tensor vanishes can always be “solved” by the prescription $$F^{\alpha \beta }4\pi (\tau ^\alpha N^\beta \tau ^\beta N^\alpha )=\frac{1}{2}ϵ^{\alpha \beta \gamma \delta }_{\gamma \delta }$$ (82) where the new field $`_{\gamma \delta }`$ just has to satisfy Maxwell’s homogeneous equations (52), i.e. $`_{\gamma \delta }𝒜_{\delta ,\gamma }𝒜_{\gamma ,\delta }`$. Taking the dual of Eq. (82) with help of the identity $`ϵ_{\gamma \delta \alpha \beta }ϵ^{\alpha \beta \mu \nu }=2(\delta _\gamma {}_{}{}^{\mu }\delta _{\delta }^{}{}_{}{}^{\nu }\delta _\gamma {}_{}{}^{\nu }\delta _{\delta }^{}{}_{}{}^{\mu })`$ gives $${}_{}{}^{}F_{\gamma \delta }^{}4\pi ϵ_{\gamma \delta \alpha \beta }\tau ^\alpha N^\beta =_{\gamma \delta }$$ (83) Contracting this equation with $`u^\gamma `$ gives the further requirement on $`_{\alpha \beta }`$: $$_{\delta \gamma }u^\gamma =B_\delta ,$$ (84) where we have used Eq. (56). The $`_{\delta \gamma }`$ can always be solved for: because of gauge freedom there are three independent components in $`𝒜_\alpha `$, and this is enough to find a solution for an arbitrary field $`B_\delta `$ obeying $`B_\alpha u^\alpha =0`$ (thus three components at most). If fact, $`B_\delta `$ does not determine $`_{\delta \gamma }`$ uniquely: if one adds to this last one of the frozen-in $`f_{\gamma \delta }`$ we discussed earlier in this section (which also satisfy the homogeneous Maxwell equations), Eq. (84) is still satisfied because $`f_{\delta \gamma }u^\gamma =0`$. We get $`\tau ^\alpha `$ by contracting Eq. (82) by $`u_\beta `$ and remembering that $`F^{\alpha \beta }u_\beta =0`$ and $`\tau ^\beta u_\beta =0`$. Thus $$\tau ^\alpha =(8\pi n)^1ϵ^{\alpha \beta \gamma \delta }_{\gamma \delta }u_\beta $$ (85) It is interesting that $`B_\delta `$ plays the role of electric part of $`_{\delta \gamma }`$ while $`\tau ^\alpha `$ enters like the magnetic part of this tensor, c.f. Eq. (56) (but because $`_{\delta \gamma }u^\gamma 0`$, $`\tau ^\alpha `$ evolves differently from a magnetic type field like $`B^\alpha `$ or the $`b^\alpha `$). It should also be clear now that the freedom in redefining $`_{\gamma \delta }_{\gamma \delta }+f_{\gamma \delta }`$ is equivalent to the changes $`\tau ^\alpha \tau ^\alpha +kn^1b^\alpha `$ we considered earlier in this section. This freedom can be exploited together with the symmetries to simplify the problem of solving explicitly for $`\tau ^\alpha `$ in any specific MHD flow.
warning/0002/hep-th0002199.html
ar5iv
text
# Supergravity Predictions on Conformal Field TheoriesTalk delivered by G. Dall’Agata. ## 1 Introduction Over two years have passed since the proposal of Maldacena of a correspondence between supergravity and string theories on Anti de Sitter (AdS) spaces and conformal field theories (CFT) on their boundary, in large $`N`$ limits. During this period, this conjectured relation has been expressed more precisely , it has been investigated under many aspects, partially verified in various cases and also extended in different directions . One of the tests which has been carried out in great depth, giving also some unexpected new results, is the matching of the spectra of conformal operators on the CFT side with the Kaluza–Klein (KK) excitations in the compactified supergravity. The AdS/CFT correspondence indeed predicts a fixed relation between scaling dimensions and KK mass modes, which can be tested in many examples. This matching was first proposed and used in , where it was called the “comparison to experiment” of the AdS/CFT conjecture. In a first stage, it had been performed only for the maximal supersymmetric cases (i.e. compactifications on spheres) and for the lower supersymmetric models deriving from orbifold compactifications . Our group has extensively focused on the generalization of the spectra matching test to lower supersymmetric models obtained by supergravity compactifications on the product of AdS space with various Einstein manifolds . Due to the presence of extra global symmetries inherited from the isometries of the internal manifold, beside the $`R`$-symmetries, these models have a far richer structure and thus yield much more probing proofs of the AdS/CFT conjecture. In spite of the greater technical complexity of lower (super)symmetric cases, we have chosen to engage in their thorough study because we had at our disposal quite powerful tools for supergravity analysis, such as harmonic expansion on coset manifolds, that were developed in the old days in the context of KK reduction of supergravity models . We would like to collect here our main results and provide a brief resumé of the lessons we have learned by exploring this subject. ## 2 A test of the correspondence In the investigation of supergravity theories with lower supersymmetry given by compactifications on coset manifolds, one encounters a very interesting and elaborate multiplet structure which makes possible some non–trivial checks in the correspondence with the spectra of conformal operators of the boundary field theory. In fact, differently from the spheres, where all KK modes belong to short representations of supersymmetry and thus have mass values that are protected against quantum corrections, for less symmetric cosets one also finds long and semilong representations, that in principle do not have any protection mechanism to prevent them from running with the couplings. It is thus quite remarkable that one can nevertheless establish a full map between each kind of KK multiplet and appropriate families of conformal operators and their descendants. We have essentially explored two directions: the correspondence $`AdS_5/CFT_4`$ and $`AdS_4/CFT_3`$. The $`AdS_5/CFT_4`$ case is more directly relevant from a physical point of view, since it involves four dimensional gauge theories, but also the $`AdS_4/CFT_3`$ one offers some intriguing challenges which could give us more insight in the formulation of the conjecture. For spontaneous compactifications of type IIB supergravity on a five dimensional coset manifold, there is only one space preserving some supersymmetry: $$T^{11}\frac{SU(2)\times SU(2)}{U(1)}$$ where the $`U(1)`$ factor is embedded diagonally in the two $`SU(2)`$. We have determined the full KK spectrum on this manifold (extending some previous partial results ) and then tested the AdS/CFT map , by matching it against the spectrum of primary conformal operators of the dual CFT constructed in . In this example we have not only shown that the duality works, but we have also some new hints on the CFT behaviour. The extension of such study to $`M`$–theory compactifications on seven–manifolds is much more complicated. It is indeed known that three dimensional CFT’s are difficult to analyze because they emerge in non–perturbative limits of conventional gauge field theories. Moreover, if for type IIB on $`T^{11}`$ we had a well defined CFT to be used towards the comparison, for the $`M`$–theory compactifications a well established CFT was not available. Thus we have used the correspondence, at first to guess these CFT’s, and then to verify by matching the spectra whether they were well–defined. $`M`$–theory allows a variety of supersymmetric compactifications down to four dimensions. The $`𝒩=2`$ examples can be divided into two categories : toric ones $$M^{111}\frac{SU(3)\times SU(2)\times U(1)}{SU(2)\times U(1)\times U(1)}\text{and }$$ $$Q^{111}\frac{SU(2)\times SU(2)\times SU(2)}{U(1)\times U(1)}$$ and non–toric ones $$V_{(5,2)}\frac{SO(5)}{SO(3)}.$$ While for the first, toric geometry helps in the definition of the CFT , in the non–toric case one has to deal with even harder difficulties . Summing up, in this analysis we have met three main features which are worth describing in some detail: i) the agreement between the CFT expectations and the supergravity results, ii) the existence of long multiplets with rational energy quantum numbers predicted by supergravity, and iii) the identification of the baryonic symmetries as those deriving from the well known presence of Betti multiplets in the compactified supergravity. ## 3 Matching the spectra The $`AdS/CFT`$ correspondence can be used in two ways: either to control the validity of the CFT by predicting properties of the supergravity, such as the mass spectrum, or to obtain information from tree level calculations in supergravity on the strong coupling CFT behaviour. Not only the fixed relation required by the $`AdS/CFT`$ map between the anomalous dimension of the various boundary conformal fields and the masses of the bulk KK modes holds for lower supersymmetry as for the highest symmetric cases , but there exists a full correspondence between all the KK modes and the conformal operators of preserved scaling dimension. In order to give a taste of how this works, we turn to the simplest non–trivial example, that is type IIB compactification on $`AdS_5\times T^{11}`$. The dual four–dimensional $`CFT`$ was given in as an $`𝒩=1`$ Yang–Mills theory with a flavor symmetry $`G=SU(2)\times SU(2)`$. It should describe the physics of a large number ($`N`$) of $`D3`$–branes placed at the singular point of the cone over the $`T^{11}`$ manifold in the decoupling limit. The “singleton” degrees of freedom of the CFT, called $`A`$ and $`B`$, are each a doublet of the $`G`$ factor groups and have a conformal anomalous dimension $`\mathrm{\Delta }_{A,B}=3/4`$. The gauge group $`𝒢`$ is $`SU(N)\times SU(N)`$ and the $`A`$ and $`B`$ chiral multiplets transform in the $`(N,\overline{N})`$ and $`(\overline{N},N)`$ of $`𝒢`$ respectively. The gauge potentials lie in the adjoint of one of the two $`SU(N)`$ groups, and their field–strength in superfield notation is given by $`W_\alpha `$. They are singlet of the matter groups, with $`R`$–symmetry charge $`r=1`$ and $`\mathrm{\Delta }=3/2`$. There is also a superpotential given by $$𝒲=\lambda ϵ^{ij}ϵ^{kl}Tr(A_iB_kA_jB_l),$$ (1) which has $`\mathrm{\Delta }=3`$, $`r=2`$. It plays an important role in the discussion in that it determines to some extent both the chiral spectrum and the marginal deformations of the SCFT. Knowing the fundamental degrees of freedom of the conformal field theory, one could try to write the conformal operators by simply combining the above fields into all possible products while respecting the symmetries of the theory. The first operators one can build in this way are the chiral operators $$Tr(AB)^k$$ (2) which are those with the lowest possible dimension for a given $`R`$–charge (they have indeed $`\mathrm{\Delta }=\frac{3}{2}r=\frac{3}{2}k`$). We notice that in the (2) operators we can freely permute all the $`A`$’s and $`B`$’s by using the equations for a critical point of the superpotential $$B_1A_kB_2=B_2A_kB_1,A_1B_lA_2=A_2B_lA_1.$$ (3) Next to these, one could also have an operator given by $`Tr[W_\alpha (AB)^k]`$ or $`Tr[W^2(AB)^k]`$, and so on. But which are the operators with protected dimension? This is a crucial question, since only the protected operators find a matching state among the KK fields, while those that suffer from quantum corrections are to be found within the full string theory. It is a well–established result that operators with protected conformal dimension correspond to the short representations of the supergroup which they belong to. For $`T^{11}`$, this supergroup is $`SU(2,2|1)`$ , while for the previously mentioned $`M`$–theory cases it is $`OSp(4|2)`$. More generally, for $`N`$ supersymmetries the four dimensional case involves $`SU(2,2|𝒩)`$ whose shortening conditions in terms of superfields have been explained in , while the generic three–dimensional case involves $`OSp(4|𝒩)`$ whose shortenings have been recently discussed in . In the $`T^{11}`$ example ($`\alpha `$, $`\dot{\alpha }`$ are spinor indices. $`x`$,$`\theta `$ and $`\overline{\theta }`$ are the bosonic and fermionic coordinates) we have only three types of such operators, namely the chiral $$\overline{D}^{\dot{\alpha }}S_{\alpha _1\mathrm{}\alpha _{2J}}=0,$$ (4) conserved $`\overline{D}^{\dot{\alpha }_1}J_{\alpha _1\mathrm{}\alpha _{2J_1},\dot{\alpha }_1\mathrm{}\dot{\alpha }_{2J_2}}`$ $`=`$ $`0`$ (5) $`\text{and }D^{\alpha _1}J_{\alpha _1\mathrm{}\alpha _{2J_1},\dot{\alpha }_1\mathrm{}\dot{\alpha }_{2J_2}}`$ $`=`$ $`0`$ (6) and semi–conserved superfields $`\overline{D}^{\dot{\alpha }_1}L_{\alpha _1\mathrm{}\alpha _{2J_1},\dot{\alpha }_1\mathrm{}\dot{\alpha }_{2J_2}}(x,\theta ,\overline{\theta })`$ $`=`$ $`0,`$ (7) $`(\overline{D}^2L_{\alpha _1\mathrm{}\alpha _{2J_1}}=0\text{for}J_2=0).`$ These differential constraints imply that these fields satisfy certain specific restrictions on their quantum numbers. As a consequence, their anomalous dimension is fixed in terms of their spin and $`R`$–symmetry charge. These constraints are respectively: $$r=\frac{2}{3}\mathrm{\Delta },$$ (8) for chiral ones, $$r=\frac{2}{3}(\mathrm{\Delta }22J_2)$$ (9) for semiconserved ones and $$\begin{array}{ccc}\hfill r& =& \frac{2}{3}(J_1J_2),\hfill \\ \hfill \mathrm{\Delta }& =& 2+J_1+J_2,\hfill \end{array}$$ (10) for conserved ones. It is easy to relate operators of different type by superfield multiplication. The product of a chiral $`(J_1,0)`$ and an anti–chiral $`(0,J_2)`$ primary gives a generic superfield with $`(J_1,J_2)`$, $`\mathrm{\Delta }=\mathrm{\Delta }^c+\mathrm{\Delta }^a`$ and $`r=\frac{2}{3}(\mathrm{\Delta }^c\mathrm{\Delta }^a)`$. By multiplying a conserved current superfield $`J_{\alpha _1\mathrm{}\alpha _{2J_1},\dot{\alpha }_1\mathrm{}\dot{\alpha }_{2J_2}}`$ by a chiral scalar superfield one gets a semi–conserved superfield with $`\mathrm{\Delta }=\mathrm{\Delta }^c+2+J_1+J_2`$ and $`r=\frac{2}{3}(\mathrm{\Delta }22J_2)`$. These are the basic rules to construct operators with protected dimensions beside the chiral ones, and they also apply in superconformal field theories of lower or higher dimensions. Since the anomalous dimensions of the protected operator is fixed in terms of their spin and $`R`$–symmetry, it must be given by a rational number. This condition severely restricts the search for the corresponding supergravity states, as it imposes strong constraints on the allowed masses and matter group quantum numbers. We find in our analysis that the requirement for the anomalous dimensions to be rational implies that one must look for dual states also having rational masses . The virtue of KK harmonic analysis on a coset space hinges on the possibility of reducing the computation of the mass eigenvalues of the various kinetic differential operators to a completely algebraic problem, while it allows to eliminate completely any explicit dependence on the coordinates of the internal manifold. Harmonics are uniquely identified by $`G`$ quantum numbers, and they are acted upon by derivatives that are reduced to algebraic operators. Such elegant technique can be quite cumbersome for complicated cosets , but it is quite straightforward for the simple $`T^{11}`$ manifold, where it leads beyond the computation of the scalar laplacian eigenvalues , or of specific sectors of the mass spectrum . By diagonalizing different operators for fields of various spin, we have found that all the masses have a fixed dependence on the scalar laplacean eigenvalue $$H_0(j,l,r)=6[j(j+1)+l(l+1)1/8r^2]$$ (11) where $`(j,l,r)`$ refer to the $`SU(2)\times SU(2)`$ and to the $`R`$–symmetry quantum numbers. This gives us a new element in the analysis as we will soon see, since besides the $`SU(2,2|1)`$ quantum numbers, we have also to match those of the matter group. The full analysis reveals that the supergravity theory has one long graviton multiplet with conformal dimensions $$\mathrm{\Delta }=1+\sqrt{H_0(j,l,r)+4},$$ (12) four long gravitino multiplets with $$\begin{array}{ccc}\hfill \mathrm{\Delta }& =& 1/2+\sqrt{H_0(j,l,r\pm 1)+4},\hfill \\ \hfill \mathrm{\Delta }& =& 5/2+\sqrt{H_0(j,l,r\pm 1)+4},\hfill \end{array}$$ (13) and four long vector multiplets, with $`\mathrm{\Delta }=2+\sqrt{H_0(j,l,r)+4},`$ $`\mathrm{\Delta }=4+\sqrt{H_0(j,l,r)+4},`$ (14) $`\mathrm{\Delta }=1+\sqrt{H_0(j,l,r\pm 2)+4}.`$ Beside these long ones, there are the shortened supermultiplets. The above formulae clearly show that the conformal dimensions become rational when the square roots assume rational values $$H_0+4\text{}^2.$$ (15) This equation is found to admit some special solutions for $`j=l=|r/2|,`$ (16) $`j=l1=|r/2|\mathrm{or}l=j1=|r/2|.`$ (17) Given these strong constraints on the possible $`SU(2,2|1)`$ quantum numbers as well as on the $`SU(2)\times SU(2)`$ ones, it becomes an easy task to build the appropriate conformal operators satisfying such constraints and find the relevant bulk supermultiplets. While referring to for all details, we list some interesting examples of conformal operators. The chiral operators of the conformal field theory are given by $`S^k`$ $`=`$ $`Tr(AB)^k`$ (18) $`\mathrm{\Phi }^k`$ $`=`$ $`Tr\left[W^2(AB)^k\right]`$ (19) $`T^k`$ $`=`$ $`Tr\left[W_\alpha (AB)^k\right]`$ (20) and are shown to correspond to hyper–multiplets containing massive recursions of the dilaton or the internal metric (18 and 19) or to tensor multiplets (20). Even more interesting are the towers of operators associated to the semi–conserved currents. Some of them are $`J_{\alpha \dot{\alpha }}^k`$ $`=`$ $`Tr(W_\alpha e^V\overline{W}_{\dot{\alpha }}e^V(AB)^k),`$ (21) $`J^k`$ $`=`$ $`Tr(Ae^V\overline{A}e^V(AB)^k),`$ (22) which lead to short multiplets whose highest state is a spin 2 and spin 1 field respectively, with masses given by $`M_{J_{\alpha \dot{\alpha }}^k}=\sqrt{{\displaystyle \frac{3}{2}}k\left({\displaystyle \frac{3}{2}}k+4\right)},`$ (23) $`\text{and }M_{J^k}=\sqrt{{\displaystyle \frac{3}{2}}k\left({\displaystyle \frac{3}{2}}k+2\right)}.`$ (24) These bulk states correspond to massive recursion of the graviton and of the gauge bosons of the matter groups. It has been explained that under certain conditions the semi–conserved superfields can become conserved, and this is indeed the case. If we set $`k=0`$ we retrieve the conserved currents related to the stress–energy tensor and the matter isometries . In fact $`M_{J_{\alpha \dot{\alpha }}^0}=M_{J^0}=0`$ are the massless graviton and gauge bosons of the supergravity theory. The above analysis can be carried out for $`M`$–theory compactifications, where again a full correspondence can be established for the short operators on the CFT and the short multiplets of the supergravity theory. We must say however that, while in the $`T^{11}`$ case the superpotential gives us a rule to discard all the sets of operators which are not related to any KK state, for the $`M`$–theory KK spectra to agree with the CFT operators one has to uncover some unknown quantum mechanism or the existence of some highly non trivial superpotential that would eliminate the mismatching states. Up to now we have checked the AdS/CFT correspondence as far as what the conformal field theory imposes on the bulk states, but what can we learn on the CFT from the analysis of the supergravity states? ## 4 Supergravity predictions There are essentially two aspects of the supergravity theory which can give us new insight on the dual CFT. The first is the presence of long multiplets that nevertheless have rational scaling dimensions, which could provide us with new non–renormalization theorems (at least in the large $`N`$, $`g_sN`$ limit). The other is is the existence of the so–called Betti multiplets, which give rise to additional symmetries in the boundary theory. Let us now turn to the first aspect. We have shown that the conformal operators with protected dimension are given by chiral ones or by their products with the conserved currents. The surprising output of the supergravity analysis is that there exist some conformal operators that in spite of not being protected by supersymmetry, still have rational conformal dimension. Confining ourselves to the $`T^{11}`$ case, if we take the chiral operator $$Tr(W^2(AB)^k),$$ we can make it non–chiral by simply inserting into the trace an antichiral combination of the gauge field–strength $$Tr(W^2e^V\overline{W}^2e^V(AB)^k).$$ This operator then corresponds to a long multiplet in the bulk theory and one should expect its scaling dimension to be generically renormalized to an irrational number. If we search for the corresponding vector multiplet in the supergravity theory, we see that its anomalous dimension is instead rational and matches exactly the naive sum of the dimensions of the operators inside the trace. We find this to be the case for all the lowest non–chiral operators of general towers with irrational scaling dimension. For instance, the towers of operators $`Tr\left[W_\alpha (Ae^V\overline{A}e^V)^n(AB)^k\right]`$ (25) $`Tr\left[e^V\overline{W}_{\dot{\alpha }}e^V(Ae^V\overline{A}e^V)^n(AB)^k\right]`$ (26) have an irrational value of $`\mathrm{\Delta }`$ for generic $`n`$, but when $`n=1`$ we have found that they do have rational anomalous dimension $`\mathrm{\Delta }=5/2+3/2k`$. When $`n=0`$ we retrieve the chiral, or semi–conserved operators with protected $`\mathrm{\Delta }`$. This is a highly non–trivial prediction of the correspondence on the CFT which comes only from the computation of the spectrum on the KK side and we hope it could receive in the future an explanation from the CFT point of view. If we restrict our attention to the protected operators, we could say that the above peculiar feature arises also in the $`AdS_4/CFT_3`$ case. However, we have a true one–to–one map and full agreement on the two sides only for a specific seven–dimensional compactification, that is the Stiefel manifold $`SO(5)/SO(3)`$ (see the summary table therein).The latter seems to be rather different from the other $`𝒩=2`$ compactifications of . Indeed, although the spectra look very similar, it seems that in the examples dealt with in , for some of the supergravity states it is not easy to identify the related CFT operator. ## 5 Betti multiplets The second AdS prediction on the CFT is the existence of baryon symmetries. As pointed out by Witten , the existence of such baryon symmetries is related to non–trivial Betti numbers of the internal manifold. Moreover, from the supergravity point of view, the non trivial value of such numbers implies the appearance of extra massless multiplets, the Betti multiplets . It is then quite natural to propose a relation between the existence of Betti multiplets and of baryon symmetries. Let’s see how this works. The non–trivial $`b_2`$ and $`b_3`$ numbers of the $`T^{11}`$ manifold imply the existence of closed non–exact 2–form $`Y_{ab}`$ and 3-form $`Y_{abc}`$. These forms must be singlets under the full isometry group, and thus they signal the presence of new additional massless states in the theory than those connected to the $`SU(2)\times SU(2)\times U_R(1)`$ isometry. From the KK expansion of the complex rank 2 $`A_{MN}`$ and real rank 4 $`A_{MNPQ}`$ tensors of type IIB supergravity we learn that we should find in the spectrum a massless vector (from $`A_{\mu abc}`$), a massless tensor (from $`A_{\mu \nu ab}`$) and two massless scalars (from the complex $`A_{ab}`$). This implies the existence of the so called Betti vector, tensor and hyper–multiplets, the last two being a peculiar feature of the $`AdS_5`$ compactification . The additional massless vector can be seen to be the massless gauge boson of an additional $`U_B(1)`$ symmetry of the theory. From the boundary point of view we need now to find an operator counterpart for such vector multiplet and look for an interpretation of the additional symmetry. The task of finding the conformal operator is very easy, once we take into account that it must be a singlet of the full isometry group and must have $`\mathrm{\Delta }=3`$. The only operator we can write is $$\begin{array}{c}𝒰=Tr\left(Ae^V\overline{A}e^V\right)Tr\left(Be^V\overline{B}e^V\right)\hfill \\ (D^2𝒰=\overline{D}^2𝒰=0),\hfill \end{array}$$ (27) which represents the conserved current of a baryon symmetry of the boundary theory under which the $`A`$ and $`B`$ field transform with opposite phase. We have shown that the occurrence of such Betti multiplets is indeed due to the existence of non–trivial two and three–cycles on the $`T^{11}`$ manifold. This implies that, from the stringy point of view, we can wrap the $`D3`$–branes of type IIB superstring theory around such 3–cycles and the wrapping number coincides with the baryon number of the low–energy CFT . We would like to point out that this feature of some manifolds can be used to check the right dimension of the singleton fields as done in . One can indeed compute the conformal dimension of the CFT operator coupling to the baryon field obtained by a $`Dp`$–brane wrapping a non–trivial $`p`$–cycle and match it with its mass, which should be proportional to the volume of the same cycle. ## 6 A puzzle An interesting case where the baryonic symmetry does not appear to be simply related to the Betti multiplets is that of type IIA compactification on $`AdS_4\times \text{}^3`$. This gives a supergravity theory which should be dual to an $`𝒩=6`$ CFT in three dimensions. It has been conjectured that the supergravity spectrum should be the same for $`M`$–theory on $`AdS_4\times S^7/\text{}_k`$ (for $`k>3`$) and for the Hopf reduction of $`AdS_4\times S^7`$ on $`AdS_4\times \text{}^3`$ . It can indeed be shown that the surviving states of the $`M`$–theory expansion on $`AdS_4\times S^7/\text{}_k`$ are the same as those of the $`𝒩=8`$ theory which are neutral under the $`U(1)`$ along which we Hopf reduce $`S^7`$ to $`\text{}^3`$. These are exactly the same as those appearing in the harmonic expansion of type IIA theory on $`AdS_4\times \text{}^3`$. From these facts, one should deduce that the massless vector of the additional $`U(1)`$ baryon symmetry is simply the KK vector deriving from the reduction of the eleven dimensional metric on the ten dimensional space $`AdS_4\times \text{}^3`$. But here comes the puzzle. Type IIA theory has a three–form $`C`$ which should give rise to Betti vector multiplets when expanded on the internal manifold $`\text{}^3`$. The complex projective space $`\text{}^3`$ has indeed a non–trivial Betti two–form: the complex structure $`J_{ab}`$. This implies that the expansion of $`C_{\mu ab}(x,y)`$ in terms of the harmonics of the internal manifold contains a vector $`c_\mu ^0`$ coupled to this form: $$C_{\mu ab}(x,y)=\underset{I}{}c_\mu ^I(x)Y_{ab}^I(y)+c_\mu ^0(x)J_{ab}.$$ (28) This again could be interpreted as the massless vector of the baryon symmetry, but we know we have only one such vector. The solution lies in the fact that this $`c_\mu ^0`$ is non–physical. It is actually a pure gauge mode as we will shortly see. Usually, type IIA supergravity is described by a one–form $`A`$, a two–form $`B`$ a three–form $`C`$ and a dilaton $`\mathrm{\Phi }`$ with field–strengths: $`F`$ $`=`$ $`dA,`$ (29) $`H`$ $`=`$ $`dB,`$ (30) $`G`$ $`=`$ $`dC+AdB.`$ (31) If we define $$C^{}C+AB,$$ (32) then $`dC^{}=dC+AdBdAB`$ and the four–form definition becomes $$G=dC^{}+FB.$$ (33) At this point $`G`$ is trivially invariant under $$\delta C^{}=dK,\text{ and }\{\begin{array}{ccc}\hfill \delta A& =& d\mathrm{\Lambda }\hfill \\ \hfill \delta C^{}& =& 0\hfill \end{array},$$ (34) while $`\delta B=d\mathrm{\Sigma }`$ requires $`\delta C=F\mathrm{\Sigma }`$. This implies that the physical invariance of $`B_{\mu \nu }(x)`$, $`\delta B_{\mu \nu }(x)=2_{[\mu }\mathrm{\Sigma }_{\nu ]}(x)`$ requires $`C_{\mu ab}^{}`$ to transform according to $$\delta C_{\mu ab}^{}(x,y)=F_{ab}\mathrm{\Sigma }_\mu .$$ (35) Keeping only linear terms in (35), we get $$\delta C_{\mu ab}^{}(x)=J_{ab}\mathrm{\Sigma }_\mu (x),$$ (36) which tells us, by comparison with (28) now applied to $`C^{}`$, that the generic mode $`c_\mu ^I`$ is invariant $`\delta _\mathrm{\Sigma }c_\mu ^I=0`$, while $`\delta _\mathrm{\Sigma }c_\mu ^0=\mathrm{\Sigma }_\mu (x)`$ is a pure gauge field. ### Acknowledgements. We are glad to thank the local organizers for a stimulating conference and their very warm hospitality. S.F. is supported by the DOE under grant DE-FG03-91ER40662, Task C. This work is also supported by the European Commission TMR programme ERBFMRX-CT96-0045 (University and Politecnico of Torino and Frascati nodes).
warning/0002/hep-ph0002184.html
ar5iv
text
# Finite width effects and gauge cancellations in W- and Z-boson production in framework of modified perturbation theory ## 1 Introduction In many applications of the Standard-Model, connected with the present and future collider experiments, one should take into account the effects of the instability of W- and Z-bosons (as well as of Higgs boson, top quark etc.) . In quantum field theory the conventional way to take into account an instability consists in the Dyson resummation of the self-energies of unstable particles . This procedure avoids nonintegrable phase-space singularities caused by the processes of production and decay of intermediate unstable particles. However, Dyson resummation leads to a deviation from the scheme of fixed-order calculations in the framework of perturbation theory (PT). In gauge theories this results in the violation of the Ward identities (WI) and loss of gauge invariance . This fact leads to loss of the control of the high-energy behavior of the theory and to the emergence of large errors in the description of particular processes. In the case of single Z-boson production (LEP1) and within the precision defined by one-loop corrections to the vertex functions, the problem of the gauge invariance was solved ad hoc (in fact, the mentioned precision implies the approximation of next-to-leading order, NLO). The most consistent scheme of calculations was described in , where only the gauge-invariant contributions to the self-energy were Dyson resummed, while the gauge-dependent ones were considered by the conventional PT. As a result, the amplitude could be presented as a product of gauge-independent factors, two vertex and one resonant. Nevertheless, this result is not universal. Anyway, now it is not clear whether this holds within the next order of precision determined by two-loop corrections to the vertex functions. In the case of a pair production of unstable particles (LEP2) the amplitude cannot be presented in the framework of Dyson resummation in a completely gauge invariant form<sup>1</sup><sup>1</sup>1It should be noted that there is an alternative approach called the pole scheme, described in and then elaborated in numerous papers. The gauge invariance in this scheme is initially maintained, but, unfortunately, an algorithm for the evaluation of corrections is not developed completely (see and discussion in Sect.8).. The hopes, nevertheless, for any further progress in such calculations are usually connected with the rather general idea of the determination of the minimal set of Feynman diagrams that are necessary for compensating the gauge violation by the Dyson resummed self-energies. For this purpose within the NLO approximation the fermion-loop scheme was proposed . Nevertheless, the bosonic corrections were not taken into consideration in this approach. In order to solve this problem a generalization of the fermion-loop scheme was proposed , defined in terms of the formalism of the background-field method. It solved the problem of the bosonic corrections and, moreover, it remained in force also beyond the one-loop approximation. However, this approach could not solve another problem, which was in the fermion-loop scheme, too. This problem concerns the incompleteness of the description within the declared precision of description because of the loss of one-loop order of PT in the resonance region . Let us consider in more detail the latter phenomenon. The point is that the denominator of unstable-particle propagator in the resonance region, $`p^2M^2=O(g^2)`$, is of order $`O(g^2)`$, but not $`O(1)`$. So, in this region the Dyson resummed one-loop self-energy actually makes a contribution in the leading-order, but not in NLO. Therefore, in order to complete the NLO approximation the two-loop correction to the self-energy should be Dyson-resummed. However, the Dyson resummation of the two-loop correction, made without taking into consideration the two-loop corrections to the vertex functions, hardly leaves a chance to maintain WI. Consideration of all two-loop corrections maybe solves the problem. Nevertheless, this solution is certainly impractical . Besides, the question remains unclear how the two-loop corrections to the vertex functions, which in fact contribute in NNLO, can compensate the gauge violation that occurs in NLO. So, anyway the problem of the loss of one-loop PT order has to be solved in a practical fashion. The present paper proposes the solution to this problem in the framework of the modified PT . Its basic idea is the expansion of direct probabilities in powers of the coupling constant instead of amplitudes. (The amplitudes prior to calculation of the probabilities are considered to be full.) Such an order of operations, taken together with ideas of the theory of generalized functions , allows one to trace the fundamental connection between the origin of the phase-space non-integrable singularity and the loss of one-loop PT order in the resonance region. (In fact this loss manifests itself as the emergence of an extra singularity in the coupling constant instead of the phase-space singularity.) Moreover, while considering the probability the contribution of the Dyson resummed two-loop correction (in the amplitude) may be reproduced within the given precision in the form of an additive anomalous term. Owing to the additivity of this term, it becomes possible to give an independent proof of the gauge cancellations in the probability. It is worth mentioning that this result does not mean that the inclusion of two-loop corrections only to denominators of “unstable” propagators (without the inclusion to vertices) does not lead to violation of WI in the amplitude. This means that contributions that violate WI in the amplitude turn out to be beyond the given precision in the probability. By its nature this phenomenon is connected with the effect of changing of the order of individual contributions in the probability due to the emergence of a singularity in the coupling constant instead of the phase-space singularities. Below we elaborate the above-stated ideas in any order of the modified PT and give a general proof of the gauge cancellations in the probability within the given precision. Notice, that the latter outcome was practically anticipated in pioneering work of , presenting the modified PT. However the reasoning of this work in the part that concerns the gauge cancellations was not complete. Indeed, it was based on a comparison with results of the conventional approach, but it overlooked the problem of losing one-loop PT order in the resonance region. Besides, it omitted the problem of the difference between remainders of the expansions of the amplitude and the corresponding probability. Nevertheless, proceeding from only the amplitude it is impossible to estimate in a mathematically correct way the remainder of the expansion of the probability, since in the resonance region the expansion of amplitude faces the ambiguity of the expression 0/0. In order to investigate in detail the above-mentioned problem we first perform the systematic study of the modified PT approach. Then we proceed with applying the results in the theory of electroweak interactions. In general case we find a practical way to keep a fixed precision without violating gauge cancellations in the framework of the background-field method. Within NLO we also find the solution in the usual formalism, basing ourselves on the results of the fermion-loop scheme. This paper is organized as follows. In Sect. 2 a statement of the problem is discussed by a simplified example of a single unstable particle production. The basic formulas of the modified PT are derived in Sect.3 (as a whole, the content of this section follows ). The properties of the expansion of the unstable propagator squared are studied in Sect.4. Section 5 discusses briefly the soft-photon problem. In Sect.6 the general proof is given of the gauge cancellations in processes mediated by unstable particle production. Section 7 is devoted to construction of the generalization of the fermion-loop scheme in NLO approximation. In Sect.8 the results are discussed. ## 2 Unstable propagator in AO: <br>statement of the problem In this section we consider the structure of the denominator of an unstable propagator. We reject, for a moment, all factors in the numerator. Then let us consider the propagator in the following form: $$\mathrm{\Delta }(\alpha ;\tau )=\frac{1}{M^2p^2\mathrm{\Sigma }}=\frac{1}{\tau \alpha h(\tau )\text{i}\alpha f(\tau )}.$$ (1) Here $`\alpha =g^2/(4\pi )`$ is the coupling constant squared, $`\tau =M^2p^2`$ is the kinematic variable, $`M`$ and $`p`$ are the mass and momentum of the unstable particle (the mass squared is considered without the conventional $`\text{i}0`$), $`\mathrm{\Sigma }`$ is the renormalized self-energy,<sup>2</sup><sup>2</sup>2In the case of vector bosons the self-energy includes two structures, $`g_{\mu \nu }`$ and $`p_\mu p_\nu `$. Formula (1) represents the contribution of the first structure only. The effect of presence of the $`p_\mu p_\nu `$-structure will be discussed latter on. $`\alpha h`$ and $`\alpha f`$ are its real and imaginary parts with the extracted factor $`\alpha `$. Here we do not fix the scheme of UV renormalization because the results will not depend on it (see below). By definition, the property of instability means that $`f0`$ in some neighborhood of the point $`\tau =0`$. Owing to causality we assume that $`f>0`$ in this neighborhood. In what follows we assume that the size of this neighborhood is of the order of $`O(\alpha )`$ and that it contains a solution to the equation $`\tau \alpha h(\tau )=0`$. Generally speaking, the function $`h(\tau )`$ may be nonzero at $`\tau =0`$. The probability of the process of the production and decay of an unstable particle is defined by an integral over some region of the kinematic variable of the propagator squared $`\text{W}(\alpha ;\tau )`$, multiplied by some weight function, $$P(\alpha )=\text{d}\tau \phi (\tau )\text{W}(\alpha ;\tau ),$$ (2) $$\text{W}(\alpha ;\tau )\left|\mathrm{\Delta }(\tau )\right|^{\mathrm{\hspace{0.17em}2}}=\frac{1}{\left[\tau \alpha h(\tau )\right]^2+\alpha ^2f^2(\tau )}.$$ (3) The weight function $`\phi (\tau )`$ stands for the complementary part of the unitarity diagram with respect to the given propagator squared. It includes the photons radiation from the initial and final charged states. In the general case $`\phi (\tau )`$ includes also hardware factors of experimental devices. By virtue of the property of $`f0`$ the function $`\text{W}(\alpha ;\tau )`$ is finite and, therefore, is integrable in a neighborhood of $`\tau =0`$. However, in the limit $`\alpha 0`$ the non-integrable singularity $`1/\tau ^2`$ appears in $`\text{W}(\alpha ;\tau )`$. This fact means the impossibility of direct application of the conventional PT in the case of processes mediated by unstable particles. Nevertheless, the expansion in the coupling constant does exist in the probability $`P(\alpha )`$. Actually, the origin of non-integrable singularity means that the *result of integration* of $`\text{W}(\alpha ;\tau )`$ with a weight function $`\phi (\tau )`$ includes a singularity in $`\alpha `$ at $`\alpha 0`$. If one extracts this singularity, then the expansion of the remaining integral becomes possible. In the case of a power-like singularity, this expansion will be a Laurent expansion, but not a Taylor one. (Let us note that the weight function $`\phi (\tau )`$ actually depends on the parameter $`\alpha `$, including it, in particular, as a factor. Therefore, the expansion of the integral ultimately may take a Taylor form. However, a priori the kind of the singularity is not known. So in order to study the problem, we first consider $`\phi (\tau )`$ as a function independent from $`\alpha `$. We also assume that $`\phi (\tau )`$ is a rather smooth function and generally is non zero at $`\tau =0`$.) Let us study the kind of singularity of the integral. Since the singularity originates from integration over small $`\tau `$, we may keep in the functions $`h(\tau )`$ and $`f(\tau )`$ only their leading terms of the asymptotic expansion at $`\tau 0`$, i.e. approximate them by $`h_0=h(0)`$ and $`f_0=f(0)`$. As a result, up to inessential corrections (which will be calculated later on), we get the approximation as follows: $$\text{W}(\alpha ;\tau )\frac{1}{[\tau \alpha h_0]^2+\alpha ^2f_0^2}.$$ (4) By virtue of the homogeneity one can deduce from (4) that the integral of $`\text{W}(\alpha ;\tau )`$ with the weight function $`\phi (\tau )`$ has a singularity of $`1/\alpha `$. Indeed, let us divide the range of integration into $`|\tau |>\text{const}\times \alpha `$ and $`|\tau |<\text{const}\times \alpha `$ with large enough “const”. The integration over the first range gives a finite contribution, while the integration over the second range gives the above-mentioned singularity (this may be verified by a change of the integration variable). Moreover, the coefficient at the singularity is proportional to $`f_0^1`$ and does not depend on $`h_0`$. In fact, $`f_0`$ may be included into the normalization of $`\alpha `$, whereas $`h_0`$ does not make contribution in the leading order. Indeed, setting $`h_0=0`$ does not lead to a singularity in $`\tau `$. For similar reasons the weight function $`\phi (\tau )`$ gives a contribution to the leading-order term as a trivial factor $`\phi (0)`$. So, despite the fact that the expansion of the integrand is an incorrect operation, the expansion of the *result of integration* is worthwhile. Moreover, some properties of such expansion may be determined before the actual calculation of the integral. For systematic determination of the properties there is a special method called the asymptotic operation, AO . Its key point is the transition to an extended interpretation of the integrand as a product of a kernel of *generalized* function on a *test* function . In fact this means that the integral is interpreted as a continuous linear functional on a test function. When the integral is well defined (*before* expanding the integrand) the above-mentioned generalization does not lead to any modification. However, after the formal expansion of the integrand the new interpretation allows one to give meaning to the non-integrable terms of the expansion. Thus, the problem of the expansion of the integrand may be solved basically within the method of generalized functions. Next, it is necessary to consider the asymptotic properties of the expansion. For this purpose an ambiguity of the extension of the interpretation of nonintegrable functions in the integrand is used. Generally, it is well-known (e.g. from experience of the UV renormalizations) that elimination of divergences may be accompanied by ambiguities. When an integral is determined by the method of generalized functions, the ambiguities are described by means of so-called counterterms, which are proportional to the delta-function or its derivatives located strictly in the point of the non-integrable singularity.<sup>3</sup><sup>3</sup>3 Let us emphasize that the introduction of counterterms is a general place in the theory of generalized functions (see e.g. ). Actually this idea was used by N.N.Bogoliubov for establishing the R-operation. In the AO context the term “counterterms” was introduced in order to emphasize the analogy with the theory of UV renormalizations. In the AO framework the coefficients at counterterms are unambiguously fixed by the requirement of the reconstruction of the result which should be obtained from an expansion of the initial integral. (Let us recall that before expansion of the integrand the integral was well defined and there were no ambiguities in it.) Moreover, AO gives a practical recipe of the calculation of these coefficients in each order of the expansion prior to a calculation of the integral. The resulting counterterms contain complete information about the terms which are singular in the parameter of the expansion. Simultaneously the counterterms may also contain some non-singular contributions which correct the asymptotic property of the expansion. In the above example the counterterm that describes the leading term of the asymptotic expansion of $`\text{W}(\alpha ;\tau )`$ is $`c/(\alpha f_0)\times \delta (\tau )`$ with $`c`$ is some numerical factor. In the given case the value of $`c`$, as well as the appearance of this counterterm, follows from the formula which is well-known in the theory of generalized functions, $$\underset{\alpha 0}{lim}\frac{\alpha }{\tau ^2+\alpha ^2}=\pi \delta (\tau ).$$ (5) Thanks to this formula the expansion of $`\text{W}(\alpha ;\tau )`$ up to an $`O(1)`$ correction is defined unambiguously and turns out to be the delta-function only. Up to the $`O(\alpha )`$ correction the expansion is non-trivial. Its most general form is $$\text{W}(\alpha ;\tau )=\pi \mathrm{/}\left(\alpha f_0\right)\delta (\tau )+\left[\mathrm{\hspace{0.17em}1}\mathrm{/}\tau ^2\right]+c_0\delta (\tau )+c_1()\delta ^{}(\tau )+O(\alpha ).$$ (6) Here $`\left[\mathrm{\hspace{0.17em}1}\mathrm{/}\tau ^2\right]`$ is the generalized function defined with some prescription. (The most common but not mandatory prescription is the principal value. Later on we omit the square brackets.) The last two terms in (6) are counterterms that correct the contributions of the first two terms. In the general case the complete determination of the generalized function $`1/\tau ^2`$ may be done by the following way . First, one makes two subtractions in the test function $`\phi (\tau )`$ in some neighborhood of $`\tau =0`$ by replacing $`\phi (\tau )`$ to $`\phi (\tau )\phi (0)\tau \phi ^{}(0)`$. As a result, the non-integrable singularity in $`1/\tau ^2`$ becomes compensated. (In fact this is one of possible prescriptions.) Then in order to describe the ambiguity emerging with this subtraction, one has to add two counterterms to $`1/\tau ^2`$. One counterterm is proportional to the delta-function and the other one is proportional to its first derivative (both counterterms correspond to the above subtractions). The coefficients at the counterterms must be determined in such a way as to guarantee the asymptotic properties of the expansion of the integral at order $`O(1)`$. Their values depend on $`h`$ and $`f`$, and on the choice of the prescription in $`1/\tau ^2`$, but the sum of all terms will not depend on the prescription. The above process may be continued. The next term of the formal expansion of $`\text{W}(\alpha ;\tau )`$ is $`2\alpha h(\tau )\times 1\mathrm{/}\tau ^3`$. For its complete determination one needs three counterterms which include the delta-function, its first and its second derivatives. The coefficients at counterterms are fixed by the requirement of the asymptotic properties of the expansion. The practical recipe of the calculation of the coefficients is presented in the next section. ## 3 Calculation of counterterms Let us show the technique of the calculation of counterterms on the example of the AO expansion of $`\text{W}(\alpha ;\tau )`$ up to $`O(\alpha ^2)`$. Since the leading term of this expansion is of order $`O(\alpha ^1)`$, the mentioned precision is sufficient for the determination of the next-to-next-to-leading order (NNLO) approximation. Such a precision seems to be sufficient for most of the practical applications. It is sufficient also for understanding the properties of the AO expansion of $`\text{W}(\alpha ;\tau )`$. The general structure of the AO expansion of $`\text{W}(\alpha ;\tau )`$ within the considered precision is as follows: $$\text{W}(\alpha ;\tau )=\frac{1}{\tau ^2}+\frac{2h(\tau )}{\tau ^3}\alpha +E(\tau )+O(\alpha ^2).$$ (7) Here the first two terms represent the result of the formal expansion of $`\text{W}(\alpha ;\tau )`$. For definiteness, we treat the poles with respect to $`\tau `$ in the sense of principal value. Below we recall two equivalent definitions of the principal value: $$VP\frac{1}{\tau ^n}=\frac{1}{2}\left[\frac{1}{(\tau +\text{i}0)^n}+\frac{1}{(\tau \text{i}0)^n}\right]=\frac{()^{n1}}{(n1)!}\frac{\text{d}^n}{\text{d}\tau ^n}\mathrm{ln}\left|\tau \right|.$$ (8) The derivatives in the last expression are understood in the sense of generalized functions, i.e. they must be switched to the test function via formal integration by parts. The quantity $`E(\tau )`$ in (7) represents the sum of counterterms. In this case they are proportional to the delta-function, its first and its second derivatives: $$E(\tau )=\underset{n=\mathrm{\hspace{0.17em}0}}{\overset{2}{}}\frac{()^nc_n}{n!}\delta ^{(n)}(\tau ).$$ (9) In what follows we assume that $`h(\tau )`$ and $`f(\tau )`$ together with their second derivatives are regular functions in some neighborhood of $`\tau =0`$.<sup>4</sup><sup>4</sup>4 If the unstable particle is able to interact with massless particles (photons), this property does not occur. Nevertheless, supplying the photons with mass we restore the analyticity inside a neighborhood defined by the generated mass gap. This is enough for our purposes (see discussion in Sect. 5). At first, for simplicity we suppose that the functions $`h`$ and $`f`$ include one-loop contributions only. A generalization to the multi-loop case is considered below in this section. We wish to define a procedure for the determination of the coefficients $`c_n`$, $`n=0,1,2`$, for an arbitrary $`\phi (\tau )`$ function, which decreases rapidly enough at infinity. This procedure should provide the following: $$\underset{\mathrm{}}{\overset{+\mathrm{}}{}}\text{d}\tau \phi (\tau )\left[\text{W}(\alpha ;\tau )\frac{1}{\tau ^2}\frac{2h(\tau )}{\tau ^3}\alpha E(\tau )\right]=O(\alpha ^2).$$ (10) Note, that for the solution of this problem we do not have to know all information about the function $`h(\tau )`$ in the third term in square brackets. We need only have knowledge about three terms of its asymptotic expansion, $$h(\tau )=h_0+\tau h_0^{}+(\tau ^2/2)h_0^{\prime \prime }+o(\tau ^2).$$ (11) This property follows from the fact that the remainder $`o(\tau ^2)`$ cancels the nonintegrable singularity $`1/\tau ^3`$ in (10). Now let us substitute (9) into (10) and, following , present the test function $`\phi (\tau )`$ as a linear combination of three basis functions $`\phi _n(\tau )`$, $`n=0,1,2`$, such that $`\phi _n^{(k)}(0)=\delta _n^k`$, where $`\delta _n^k`$ is the Kronecker symbol. As a result we get $$c_n=\underset{\mathrm{}}{\overset{+\mathrm{}}{}}\text{d}\tau \phi _n(\tau )\left[\text{W}(\alpha ;\tau )\frac{1}{\tau ^2}\frac{2h(\tau )}{\tau ^3}\alpha \right]+O(\alpha ^2).$$ (12) It is obvious from (12) that within the given precision the coefficients $`c_n`$ do not depend on the choice of test functions. Indeed, any other test functions $`\stackrel{~}{\phi }_n(\tau )`$ possessing the same property, $`\stackrel{~}{\phi }_n^{(k)}(0)=\delta _n^k`$, lead to the coefficients $`\stackrel{~}{c}_n`$ instead of $`c_n`$. However, the difference $`\stackrel{~}{c}_nc_n`$ amounts to a quantity of order $`O(\alpha ^2)`$, because the difference between the corresponding integrals is controlled by the test function $`\stackrel{~}{\phi }_n(\tau )\phi _n(\tau )`$, which equals zero at $`\tau =0`$ together with its first and second derivatives. It is easy to see that the integral in formula (12) with such a weight function is of order $`O(\alpha ^2)`$. Due to this property, without loss of generality one may choose the test functions to be the step-like ones. Namely, we may set $`\phi _n(\tau )=\tau ^n\times \theta (|\tau |<\mathrm{\Lambda })`$. Then, with $`n=0,1,2`$, we have $$c_n=\underset{\mathrm{\Lambda }}{\overset{+\mathrm{\Lambda }}{}}\text{d}\tau \tau ^n\left[\text{W}(\alpha ;\tau )\frac{1}{\tau ^2}\frac{2h(\tau )}{\tau ^3}\alpha \right]+O(\alpha ^2).$$ (13) Formula (13) basically solves the problem stated above. However, it is still too complicated for practical usage since, generally speaking, through $`\text{W}(\alpha ;\tau )`$ it contains the dependence on the unknowns functions $`h(\tau )`$ and $`f(\tau )`$. Moreover, (13) contains a lot of superfluous information because the integral in r.h.s. includes contributions beyond the given precision. In particular, the dependence on the cutoff parameter $`\mathrm{\Lambda }`$ is of this kind. The problem may be solved by homogenization procedure , i.e. by the specific transformations in the integrand. Namely, let us do the substitutions $`\tau \xi \tau `$, $`\alpha \xi \alpha `$. Then the result we expand in powers of $`\xi `$, and in the end we set $`\xi =1`$. Each term of this (secondary) expansion is proved to be a homogeneous function of $`\tau `$ and $`\alpha `$. So it leads to a strictly definite contribution in the power in $`\alpha `$ to the integral, which now may be considered without the cutoff. The first term $`c_n^{(0)}`$ of this expansion gives the leading-order contribution to the coefficient $`c_n`$, $$c_n^{(0)}=\underset{\mathrm{}}{\overset{+\mathrm{}}{}}\text{d}\tau \tau ^n\left[\frac{1}{(\tau \alpha h_0)^2+\alpha ^2f_0^2}\frac{1}{\tau ^2}\frac{2h_0}{\tau ^3}\alpha \right].$$ (14) The next term of the expansion of homogenization gives the correction term $`c_n^{(1)}`$, etc. On counting of powers we have $`c_n^{(0)}\alpha ^{n1}`$, $`c_n^{(1)}\alpha ^n`$, etc. Adding the necessary number of $`c_n^{(i)}`$ we get $`c_n`$ with the required precision. Let us emphasize, that the integral in (14) is convergent at infinity. The convergence takes place also for other $`c_n^{(i)}`$. Moreover, this property is a general result of the application of the homogenization . It is worth recalling that the singular terms $`1/\tau ^2`$ and $`1/\tau ^3`$ in (14) are defined in the sense of principal value. Nevertheless, the prescription may be changed (in all above-derived formulas). As a result the coefficients $`c_n^{(i)}`$ will change, but the asymptotic properties of the expansion (7) will be conserved. After the corresponding calculations we come to the following result (for the first time obtained in ): $`c_0`$ $`=`$ $`{\displaystyle \frac{\pi }{\alpha f_0}}+{\displaystyle \frac{\pi \left(h_0^{}f_0h_0f_0^{}\right)}{f_0^2}}`$ $`+{\displaystyle \frac{\pi \left(h_{}^{}{}_{0}{}^{2}f_0^2+h_0^2f_{}^{}{}_{0}{}^{2}+h_0h_0^{\prime \prime }f_0^22h_0h_0^{}f_0f_0^{}\frac{1}{2}h_0^2f_0f_0^{\prime \prime }\frac{1}{2}f_0^3f_0^{\prime \prime }\right)}{f_0^3}}\alpha ,`$ $`c_1`$ $`=`$ $`{\displaystyle \frac{\pi h_0}{f_0}}+{\displaystyle \frac{\pi \left(2h_0h_{}^{}{}_{0}{}^{}f_0h_0^2f_{}^{}{}_{0}{}^{}f_0^2f_{}^{}{}_{0}{}^{}\right)}{f_0^2}}\alpha ,c_2={\displaystyle \frac{\pi \left(h_0^2f_0^2\right)}{f_0}}\alpha .`$ (15) Here the subscript 0 means that the corresponding quantities are defined at $`\tau =0`$, while the superscript means we are taking derivatives. For example, $`h_0^{}=\text{d}h(\tau )/\text{d}\tau |_{\tau =0}`$, etc. The derived result may be written in a more compact form if the quantities $`c_n`$ in (9) are understood as functions of $`\tau `$. In this case, instead of (15), one can obtain $$c_0=\frac{\pi }{\alpha f},c_1=\frac{\pi h}{f},c_2=\alpha \frac{\pi \left(h^2f^2\right)}{f}.$$ (16) Here $`c_{\mathrm{\hspace{0.17em}0},1,2}`$, $`h`$ and $`f`$ are functions of the $`\tau `$. The equivalence of these two forms, (15) and (16), follows from the relations $`f(\tau )\delta ^{}(\tau )=f_0\delta ^{}(\tau )f_0^{}\delta (\tau )`$ and $`f(\tau )\delta ^{\prime \prime }(\tau )=f_0\delta ^{\prime \prime }(\tau )2f_0^{}\delta ^{}(\tau )+f_0^{\prime \prime }\delta (\tau )`$. The above-derived results may easily be extended to the case of multi-loop contributions in $`h(\tau )`$ and $`f(\tau )`$. In this case one should perform the conventional expansion in $`\alpha `$ in formulas (15) or (16), in which $`h`$ and $`f`$ are understood with the presence of multi-loop contributions . The simplest way to prove this statement is as follows: assuming that $`h`$ and $`f`$ are the full functions we repeat all the same reasoning as was done above, except that during homogenization we modify the scaling by setting $`\alpha ^n\xi \alpha ^n`$ in the $`n`$-loop contributions (instead of $`\alpha ^n\xi ^n\alpha ^n`$). As a result the higher-loop contributions become identified with the one-loop ones, and formulas (15) and (16) are restored automatically. ## 4 The properties of AO expansion of $`\text{W}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}`$ Let us discuss now the properties of the AO expansion of $`\text{W}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}`$. At first we are interested in the characteristic which will be useful in proving gauge cancellations in the probability. It should be emphasized that a solution to this problem is not obvious a priori due to the specificity of the Feynman rules in the modified PT approach (see formulas (7), (9) and (15)). Then we will examine the self-consistency properties of the expansion, such as a non-sensitivity of the results to the ambiguities in the definition of the gauge-bosons masses , and consistency with the requirements of UV renormalization. We start from a rather methodological problem consisting in the explicit demonstration of the independence of the results of an AO expansion from the sequence of the expansion in the higher-loops. In other words, we show that the result of the AO expansion of $`\text{W}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}`$ will be the same if instead of (1) one will start from the following formula of the incomplete Dyson resummation: $$𝚫\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{=}\frac{\mathrm{𝟏}}{𝝉\mathbf{}𝜶𝚺_\mathrm{𝟏}\mathbf{}𝜶^\mathrm{𝟐}𝚺_\mathrm{𝟐}\mathbf{}𝜶^\mathrm{𝟑}𝚺_\mathrm{𝟑}\mathbf{+}\mathbf{}}\mathbf{=}\frac{\mathrm{𝟏}}{𝝉\mathbf{}𝜶𝚺_\mathrm{𝟏}}\mathbf{+}\frac{𝜶^\mathrm{𝟐}𝚺_\mathrm{𝟐}\mathbf{+}𝜶^\mathrm{𝟑}𝚺_\mathrm{𝟑}}{\mathbf{\left(}𝝉\mathbf{}𝜶𝚺_\mathrm{𝟏}\mathbf{\right)}^\mathrm{𝟐}}\mathbf{+}\mathbf{}\mathbf{.}$$ (17) Here $`𝜶^𝒏𝚺_𝒏\mathbf{(}𝝉\mathbf{)}`$ is the $`𝒏`$-loop contribution to self-energy. By squaring (17) we get the incomplete expansion for $`\text{W}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}`$, each term of which is integrable in the conventional sense: $`\begin{array}{c}\text{W}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{=}\text{W}_\mathrm{𝟏}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{+}\mathbf{\left[}\mathbf{\left(}𝜶^\mathrm{𝟐}𝚺_\mathrm{𝟐}\mathbf{+}𝜶^\mathrm{𝟑}𝚺_\mathrm{𝟑}\mathbf{\right)}\text{W}_{\mathrm{𝟏𝟏}}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{+}𝜶^\mathrm{𝟒}\mathbf{\left(}𝚺_\mathrm{𝟐}\mathbf{\right)}^\mathrm{𝟐}\text{W}_{\mathrm{𝟏𝟐}}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{+}\text{h.c.}\mathbf{\right]}\\ \\ \mathbf{+}𝜶^\mathrm{𝟒}\mathbf{\left|}𝚺_\mathrm{𝟐}\mathbf{\right|}^\mathrm{𝟐}\text{W}_\mathrm{𝟏}^\mathrm{𝟐}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{+}𝑶\mathbf{(}𝜶^\mathrm{𝟐}\mathbf{)}\mathbf{.}\end{array}`$ Here we have introduced three new functions: $`\text{W}_{\mathrm{𝟏𝟏}}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{=}\text{W}_\mathrm{𝟏}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}𝚫_\mathrm{𝟏}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}`$, $`\text{W}_{\mathrm{𝟏𝟐}}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{=}\text{W}_\mathrm{𝟏}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}𝚫_\mathrm{𝟏}^\mathrm{𝟐}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}`$, and $`\text{W}_\mathrm{𝟏}^\mathrm{𝟐}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{=}\text{W}_\mathrm{𝟏}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\text{W}_\mathrm{𝟏}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}`$, where $`\text{W}_\mathrm{𝟏}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}`$ and $`𝚫_\mathrm{𝟏}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}`$ are the same functions as $`\text{W}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}`$ and $`𝚫\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}`$, but with only one-loop self-energies Dyson resummed. By considering these new functions as generalized functions, each term in (18) may be in turn completely AO expanded. Repeating the reasonings of Sect.2, one can show that the leading term of AO expansion of $`\text{W}_{\mathrm{𝟏𝟏}}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}`$ has the behavior of $`\mathrm{𝟏}\mathbf{/}𝜶^\mathrm{𝟐}`$. So, after integration with the weight function $`𝝋\mathbf{(}𝝉\mathbf{)}`$ the first term in the square brackets in (18) makes a contribution of order $`𝑶\mathbf{(}\mathrm{𝟏}\mathbf{)}`$. The leading terms in $`\text{W}_{\mathrm{𝟏𝟐}}`$ and $`\text{W}_\mathrm{𝟏}^\mathrm{𝟐}`$ both are $`𝑶\mathbf{(}\mathrm{𝟏}\mathbf{/}𝜶^\mathrm{𝟑}\mathbf{)}`$. So, the second term in the square brackets and the last term in (18) are of order $`𝑶\mathbf{(}𝜶\mathbf{)}`$. By a similar reasoning one can show that the neglected terms in (18) are of the order $`𝑶\mathbf{(}𝜶^\mathrm{𝟐}\mathbf{)}`$. So in view of $`\text{W}_\mathrm{𝟏}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{=}𝑶\mathbf{(}𝜶^\mathbf{}\mathrm{𝟏}\mathbf{)}`$, formula (18) approximates $`\text{W}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}`$ within the NNLO precision. It should be noted, that the above reasoning is valid only as long as the functions $`𝚺_𝒍\mathbf{(}𝝉\mathbf{)}`$, $`𝒍\mathbf{=}\mathrm{𝟐}\mathbf{,}\mathrm{𝟑}\mathbf{,}\mathbf{}`$, and several of their derivatives are regular functions in some neighborhood of $`𝝉\mathbf{=}\mathrm{𝟎}`$. If this is not the case, then the products of $`𝚺_𝒍\mathbf{(}𝝉\mathbf{)}`$ on $`\text{W}_{\mathrm{𝟏}𝒏}^𝒎`$ must be considered as new generalized functions, of which the properties should additionally be investigated. Such situation occurs when the unstable particle interacts with massless particles (photons). This difficulty may be eliminated by the inclusion of the regularizing mass for massless particles, because then functions $`𝚺_𝒍\mathbf{(}𝝉\mathbf{)}`$ become regular in a vicinity of $`𝝉\mathbf{=}\mathrm{𝟎}`$. Assuming this trick let us now consider the complete AO expansion of $`\text{W}_{\mathrm{𝟏}𝒏}^𝒎`$. For brevity we omit the corresponding derivation since it is similar to the one considered in the previous section. The desirable accuracy of the expansion is controlled by formula (18), so that $`\text{W}_{\mathrm{𝟏𝟏}}`$ should be expanded up to $`𝑶\mathbf{(}\mathrm{𝟏}\mathbf{)}`$ corrections, and $`\text{W}_{\mathrm{𝟏𝟐}}`$ and $`\text{W}_\mathrm{𝟏}^\mathrm{𝟐}`$ up to $`𝑶\mathbf{(}𝜶^\mathbf{}\mathrm{𝟐}\mathbf{)}`$ corrections. However, in further calculations major precision is required. So, let us consider $$\text{W}_{\mathrm{𝟏𝟏}}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{=}𝑬\mathbf{(}𝝉\mathbf{)}\mathbf{+}\frac{\mathrm{𝟏}}{𝝉^\mathrm{𝟑}}\mathbf{+}𝑶\mathbf{(}𝜶\mathbf{)}\mathbf{,}\text{W}_{\mathrm{𝟏𝟐}}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{=}𝑬\mathbf{(}𝝉\mathbf{)}\mathbf{+}𝑶\mathbf{(}\mathrm{𝟏}\mathbf{)}\mathbf{,}\text{W}_\mathrm{𝟏}^\mathrm{𝟐}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{=}𝑬\mathbf{(}𝝉\mathbf{)}\mathbf{+}𝑶\mathbf{(}\mathrm{𝟏}\mathbf{)}\mathbf{.}$$ (18) Here in all three cases $`𝑬\mathbf{(}𝝉\mathbf{)}`$ is defined by (9), but with different coefficients $`𝒄_𝒏`$. In the compact form, when $`𝒄_𝒏`$ are defined as functions on $`𝝉`$, we get: | $`\text{W}_{\mathrm{𝟏𝟏}}\mathbf{:}`$ | $`𝒄_\mathrm{𝟎}\mathbf{=}\frac{\text{i}𝝅}{\mathrm{𝟐}𝜶^\mathrm{𝟐}𝒇^\mathrm{𝟐}}\mathbf{,}`$ | $`𝒄_\mathrm{𝟏}\mathbf{=}\frac{𝝅\mathbf{\left(}\text{i}𝒉\mathbf{+}𝒇\mathbf{\right)}}{\mathrm{𝟐}𝜶𝒇^\mathrm{𝟐}}\mathbf{,}`$ | $`𝒄_\mathrm{𝟐}\mathbf{=}\frac{𝝅\mathbf{\left(}\text{i}𝒉^\mathrm{𝟐}\mathbf{+}\text{i}𝒇^\mathrm{𝟐}\mathbf{+}\mathrm{𝟐}𝒉𝒇\mathbf{\right)}}{\mathrm{𝟐}𝒇^\mathrm{𝟐}}\mathbf{;}`$ | (20) | | --- | --- | --- | --- | --- | | $`\text{W}_{\mathrm{𝟏𝟐}}\mathbf{:}`$ | $`𝒄_\mathrm{𝟎}\mathbf{=}\mathbf{}\frac{𝝅}{\mathrm{𝟒}𝜶^\mathrm{𝟑}𝒇^\mathrm{𝟑}}\mathbf{,}`$ | $`𝒄_\mathrm{𝟏}\mathbf{=}\frac{𝝅\mathbf{\left(}\text{i}𝒇\mathbf{}𝒉\mathbf{\right)}}{\mathrm{𝟒}𝜶^\mathrm{𝟐}𝒇^\mathrm{𝟑}}\mathbf{,}`$ | $`𝒄_\mathrm{𝟐}\mathbf{=}\frac{𝝅\mathbf{\left(}\mathrm{𝟐}\text{i}𝒉𝒇\mathbf{+}𝒇^\mathrm{𝟐}\mathbf{}𝒉^\mathrm{𝟐}\mathbf{\right)}}{\mathrm{𝟒}𝜶𝒇^\mathrm{𝟑}}\mathbf{;}`$ | (21) | | $`\text{W}_\mathrm{𝟏}^\mathrm{𝟐}\mathbf{:}`$ | $`𝒄_\mathrm{𝟎}\mathbf{=}\frac{𝝅}{\mathrm{𝟐}𝜶^\mathrm{𝟑}𝒇^\mathrm{𝟑}}\mathbf{,}`$ | $`𝒄_\mathrm{𝟏}\mathbf{=}\frac{𝝅𝒉}{\mathrm{𝟐}𝜶^\mathrm{𝟐}𝒇^\mathrm{𝟑}}\mathbf{,}`$ | $`𝒄_\mathrm{𝟐}\mathbf{=}\frac{𝝅\mathbf{\left(}𝒉^\mathrm{𝟐}\mathbf{+}𝒇^\mathrm{𝟐}\mathbf{\right)}}{\mathrm{𝟐}𝜶𝒇^\mathrm{𝟑}}\mathbf{.}`$ | (22) | Notice here that the coefficients singular in $`𝜶`$ do not depend on the prescription for the poles in $`𝝉`$, since in the examples considered above the non-integrable terms (for which the prescription is needed) are non-singular in $`𝜶`$. Based on the derived results we formulate the following properties of the expansion. (All of them can be verified by direct complete expanding and comparing the results.) Property 1. The incomplete expansion (18) is equivalent within the given precision to the complete AO expansion of $`\text{W}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}`$. The next properties are non-trivial and some additional argument should be given. Property 2. Within the given precision an incomplete expansion of $`\text{W}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}`$ is equivalent to its complete AO expansion if this incomplete expansion includes as Dyson resummed all contribution non-zero at $`𝝉\mathbf{=}\mathrm{𝟎}`$ to $`\text{Im}𝚺_\mathrm{𝟏}\mathbf{(}𝝉\mathbf{)}`$. All other contributions to $`𝚺_\mathrm{𝟏}\mathbf{(}𝝉\mathbf{)}`$ may be transferred from denominators to numerators in the sense of a conventional expansion. Moreover, only a finite number of terms of the latter expansion are relevant within the given precision (see remark at the end of the proof). We perform the proof in two steps. At first we will show that all contributions to $`𝚺_\mathrm{𝟏}\mathbf{(}𝝉\mathbf{)}`$ zero at $`𝝉\mathbf{=}\mathrm{𝟎}`$ without loss of precision may be expanded in the conventional sense. Then we will show that the same operation may be done also for the whole of the real part of $`𝚺_\mathrm{𝟏}\mathbf{(}𝝉\mathbf{)}`$. So, let $`𝚺_\mathrm{𝟏}\mathbf{(}𝝉\mathbf{)}\mathbf{=}𝚺_{\mathrm{𝟎𝟏}}\mathbf{(}𝝉\mathbf{)}\mathbf{+}\stackrel{\mathbf{~}}{𝚺}_\mathrm{𝟏}\mathbf{(}𝝉\mathbf{)}`$, where by definition $`\stackrel{\mathbf{~}}{𝚺}_\mathrm{𝟏}\mathbf{(}\mathrm{𝟎}\mathbf{)}\mathbf{=}\mathrm{𝟎}`$, but $`𝚺_{\mathrm{𝟎𝟏}}\mathbf{(}\mathrm{𝟎}\mathbf{)}\mathbf{}\mathrm{𝟎}`$. Since in some neighborhood of $`𝝉\mathbf{=}\mathrm{𝟎}`$ $`\stackrel{\mathbf{~}}{𝚺}_\mathrm{𝟏}\mathbf{(}𝝉\mathbf{)}`$ is a correction to $`𝚺_{\mathrm{𝟎𝟏}}\mathbf{(}𝝉\mathbf{)}`$, its contribution may be expanded like $`𝚺_𝒏\mathbf{(}𝝉\mathbf{)}`$ with $`𝒏\mathbf{>}\mathrm{𝟏}`$ in (18): $$\begin{array}{c}\\ \text{W}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{=}\text{W}_\mathrm{𝟏}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{+}\mathbf{\left[}\begin{array}{c}\mathbf{\left(}𝜶\stackrel{\mathbf{~}}{𝚺}_\mathrm{𝟏}\mathbf{+}𝜶^\mathrm{𝟐}𝚺_\mathrm{𝟐}\mathbf{+}𝜶^\mathrm{𝟑}𝚺_\mathrm{𝟑}\mathbf{\right)}\text{W}_{\mathrm{𝟏𝟏}}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\\ \mathbf{+}\mathbf{\left(}𝜶^\mathrm{𝟐}\stackrel{\mathbf{~}}{𝚺}_\mathrm{𝟏}^\mathrm{𝟐}\mathbf{+}\mathrm{𝟐}𝜶^\mathrm{𝟑}\stackrel{\mathbf{~}}{𝚺}_\mathrm{𝟏}𝚺_\mathrm{𝟐}\mathbf{+}𝜶^\mathrm{𝟒}𝚺_\mathrm{𝟐}^\mathrm{𝟐}\mathbf{\right)}\text{W}_{\mathrm{𝟏𝟐}}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{+}\text{h.c.}\end{array}\mathbf{\right]}\\ \mathbf{+}\mathbf{\left[}𝜶^\mathrm{𝟐}\mathbf{\right|}\stackrel{\mathbf{~}}{𝚺}_\mathrm{𝟏}\mathbf{|}^\mathrm{𝟐}\mathbf{+}𝜶^\mathrm{𝟑}\mathbf{(}\stackrel{\mathbf{}}{𝚺}_\mathrm{𝟏}\underset{\mathrm{𝟐}}{\overset{\mathbf{}}{𝚺}}\mathbf{+}\text{h.c.}\mathbf{)}\mathbf{+}𝜶^\mathrm{𝟒}\mathbf{\left|}𝚺_\mathrm{𝟐}\mathbf{|}^\mathrm{𝟐}\mathbf{\right]}\text{W}_\mathrm{𝟏}^\mathrm{𝟐}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{+}𝑶\mathbf{(}𝜶^\mathrm{𝟐}\mathbf{)}\mathbf{.}\end{array}$$ (23) Here symbol “$`\mathbf{}`$” means complex conjugation. In the denominators of $`\text{W}_{\mathrm{𝟏}𝒏}^𝒎`$ only the $`𝚺_{\mathrm{𝟎𝟏}}\mathbf{(}𝝉\mathbf{)}`$ is Dyson resummed. The remainder in (23) is estimated in the AO sense. The AO expansions of $`\text{W}_\mathrm{𝟏}`$, $`\text{W}_{\mathrm{𝟏𝟏}}`$, $`\text{W}_{\mathrm{𝟏𝟐}}`$ and $`\text{W}_\mathrm{𝟏}^\mathrm{𝟐}`$ within the required precision were previously described in this section. The proof of (23) may be given noticing that within the given precision the quantity $`𝜶\stackrel{\mathbf{~}}{𝚺}_\mathrm{𝟏}`$ gives a non-zero contribution only if it is raised to not more than the second power. In fact, against the background of the regular terms this property is obvious. If $`\stackrel{\mathbf{~}}{𝚺}_\mathrm{𝟏}^\mathrm{𝟐}`$ (or $`\mathbf{|}\stackrel{\mathbf{~}}{𝚺}_\mathrm{𝟏}\mathbf{|}^\mathrm{𝟐}`$) is multiplied by a counterterm originating from $`\text{W}_{\mathrm{𝟏}𝒏}^𝒎`$, the result is non-zero only in the presence of second- or higher-order derivatives of the delta-function in the counterterm (otherwise there acts the property $`\stackrel{\mathbf{~}}{𝚺}_\mathrm{𝟏}\mathbf{(}\mathrm{𝟎}\mathbf{)}\mathbf{=}\mathrm{𝟎}`$). Consider the functions $`\text{W}_{\mathrm{𝟏}𝒏}^𝒎\mathbf{=}\mathbf{\left[}\text{W}_\mathrm{𝟏}\mathbf{\right]}^𝒎𝚫_\mathrm{𝟏}^𝒏`$, $`𝒏\mathbf{}\mathrm{𝟎}`$, $`𝒎\mathbf{}\mathrm{𝟎}`$, $`𝒏\mathbf{+}𝒎\mathbf{}\mathrm{𝟏}`$ is the number of self-energy insertions on one side of the cut of the diagram of unitarity, $`𝒎\mathbf{}\mathrm{𝟏}`$ is the number on the other side. In these functions such counterterms appear in order $`𝜶^{\mathbf{}\mathbf{(}𝒏\mathbf{+}\mathrm{𝟐}𝒎\mathbf{}\mathrm{𝟏}\mathbf{)}}\mathbf{\times }𝜶^\mathrm{𝟐}`$, and only those functions $`\text{W}_{\mathrm{𝟏}𝒏}^𝒎`$ are to be multiplied by $`\stackrel{\mathbf{~}}{𝚺}_\mathrm{𝟏}^\mathrm{𝟐}`$ (or $`\mathbf{|}\stackrel{\mathbf{~}}{𝚺}_\mathrm{𝟏}\mathbf{|}^\mathrm{𝟐}`$) which satisfy the condition $`𝒏\mathbf{+}\mathrm{𝟐}𝒎\mathbf{}\mathrm{𝟐}\mathbf{}\mathrm{𝟐}`$. Note that in both insertions of $`\stackrel{\mathbf{~}}{𝚺}_\mathrm{𝟏}`$, each gives a factor $`𝜶`$. Of the other remaining $`𝒏\mathbf{+}\mathrm{𝟐}𝒎\mathbf{}\mathrm{𝟒}`$ insertions of $`\stackrel{\mathbf{~}}{𝚺}_𝒌`$, $`𝒌\mathbf{}\mathrm{𝟐}`$, each gives a factor not less than $`𝜶^\mathrm{𝟐}`$. Thus all considered terms result in $`𝑶\mathbf{(}𝜶\mathbf{)}`$ contribution. While extending the above reasoning to the third and the higher powers of $`𝜶\stackrel{\mathbf{~}}{𝚺}_\mathrm{𝟏}`$, we easily see that they give nonzero contributions of $`𝑶\mathbf{(}𝜶^\mathrm{𝟐}\mathbf{)}`$ only. The above result may be generalized to the real part of $`\stackrel{\mathbf{~}}{𝚺}_\mathrm{𝟏}\mathbf{(}𝝉\mathbf{)}`$. In this case $`\stackrel{\mathbf{~}}{𝚺}_\mathrm{𝟏}\mathbf{(}𝝉\mathbf{)}`$ in (23) is to be defined by $`\text{Im}\stackrel{\mathbf{~}}{𝚺}_\mathrm{𝟏}\mathbf{(}\mathrm{𝟎}\mathbf{)}\mathbf{=}\mathrm{𝟎}`$ with $`\text{Im}𝚺_{\mathrm{𝟎𝟏}}\mathbf{(}\mathrm{𝟎}\mathbf{)}\mathbf{}\mathrm{𝟎}`$. The basis for this generalization is the observation that the real part of the self-energy does not contribute to the leading-order term of the AO expansion of $`\text{W}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}`$. Let us remark, that at first sight formula (23) with this modification should look much more complicated, because from a formal point of view it should contain an infinite series of terms of $`\mathbf{[}𝜶\text{Re}𝚺_\mathrm{𝟏}\mathbf{(}𝝉\mathbf{)}\mathbf{]}^𝒏\mathbf{\times }\text{W}_{\mathrm{𝟏}𝒏}`$, each of which is of order $`𝑶\mathbf{(}𝜶^\mathbf{}\mathrm{𝟏}\mathbf{)}`$, etc. However, forming the groups with other functions $`\text{W}_{\mathrm{𝟏}𝒏}^𝒎`$ all superfluous terms must mutually cancel. The mentioned groups will be formed by virtue of relations of the type $`\mathrm{𝟐}\text{Re}\text{W}_{\mathrm{𝟏𝟐}}\mathbf{+}\text{W}_\mathrm{𝟏}^\mathrm{𝟐}\mathbf{=}𝑶\mathbf{(}𝜶^\mathbf{}\mathrm{𝟏}\mathbf{)}`$ \[not $`𝑶\mathbf{(}𝜶^\mathbf{}\mathrm{𝟑}\mathbf{)}`$\], etc. It is worth noticing that the above result permits one to expand the bosonic corrections to self-energy of W- and Z-bosons within the *finite* number of terms. (This is because the mentioned bosonic corrections possess the property $`\text{Im}𝚺_\mathrm{𝟏}^{\text{bos}}\mathbf{(}\mathrm{𝟎}\mathbf{)}\mathbf{=}\mathrm{𝟎}`$ .) This is a nontrivial result since according to the formal power counting all these terms are of the same order in $`𝜶`$ and, consequently, should be explicitly taken into consideration. Property 3. There is the following approximation of $`\text{W}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}`$ up to $`𝑶\mathbf{(}𝜶^𝒏\mathbf{)}`$ corrections: $$\text{W}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{=}\text{W}_{\mathbf{[}𝒏\mathbf{]}}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{}𝜶^{𝒏\mathbf{}\mathrm{𝟏}}\frac{\text{Im}𝚺_{𝒏\mathbf{+}\mathrm{𝟏}}\mathbf{(}\mathrm{𝟎}\mathbf{)}}{\mathbf{\left[}\text{Im}𝚺_\mathrm{𝟏}\mathbf{(}\mathrm{𝟎}\mathbf{)}\mathbf{\right]}^\mathrm{𝟐}}𝝅𝜹\mathbf{(}𝝉\mathbf{)}\mathbf{+}𝑶\mathbf{(}𝜶^𝒏\mathbf{)}\mathbf{.}$$ (24) Here $`\text{W}_{\mathbf{[}𝒏\mathbf{]}}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}`$ stands for the propagator squared with Dyson resummed contributions to the self-energy up to $`𝒏`$-loops. The second term includes the $`\mathbf{(}𝒏\mathbf{+}\mathrm{𝟏}\mathbf{)}`$-loop correction which is necessary (with the appropriate factors) within the indicated precision. Formula (24) follows immediately from the obvious generalization of (18) to the case of any $`𝒏`$ taking into consideration the first result in (20). It is worth noticing that the second term in (24) makes a contribution in the highest order within the given precision. Consequently, any $`𝑶\mathbf{(}𝜶\mathbf{)}`$-variation of the mass is worthless in this term, because the effect may be referred to the neglected term of order $`𝑶\mathbf{(}𝜶^𝒏\mathbf{)}`$. Actually this is a very important observation, since in the case of electroweak theory it permits one to leave out of account the ambiguities in the definition of the gauge-bosons masses . In particular, one may consider the factor $`\text{Im}𝚺_{𝒏\mathbf{+}\mathrm{𝟏}}\mathbf{(}\mathrm{𝟎}\mathbf{)}`$ to be defined at the complex-value pole position in the spirit of the pole-scheme (as a result $`\text{Im}𝚺_{𝒏\mathbf{+}\mathrm{𝟏}}\mathbf{(}\mathrm{𝟎}\mathbf{)}`$ becomes completely gauge-invariant), or simply as the $`𝒏`$-loop correction to the width of the unstable particle divided by its mass. Formula (24) represents the quantitative characteristic of the property of the loss of one-loop PT order in the resonance region. Indeed, as long as $`\text{W}_{\mathbf{[}𝒏\mathbf{]}}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{}𝜶^\mathbf{}\mathrm{𝟏}`$ at $`𝜶\mathbf{}\mathrm{𝟎}`$, the second term in (24), represents the $`𝒏`$-th order correction, but not the $`\mathbf{(}𝒏\mathbf{+}\mathrm{𝟏}\mathbf{)}`$-th one, as might naively be expected. Since the second term in (24) cannot be obtained from the analysis of an amplitude, it is pertinent to call it the *anomalous additive term*. Property 4. The above expansions possess the following transformation property when the argument $`𝝉`$ is shifted by a quantity of order $`𝑶\mathbf{(}𝜶\mathbf{)}`$: $$\stackrel{\mathbf{~}}{\text{W}}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{=}\stackrel{\mathbf{~}}{\text{W}}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{}𝜶𝒎^\mathrm{𝟐}\mathbf{)}_{\mathbf{|}\genfrac{}{}{0pt}{}{𝒉\mathbf{\left(}𝝉\mathbf{}𝜶𝒎^\mathrm{𝟐}\mathbf{\right)}\mathbf{}𝒉\mathbf{\left(}𝝉\mathbf{\right)}\mathbf{}𝒎^\mathrm{𝟐}}{𝒇\mathbf{\left(}𝝉\mathbf{}𝜶𝒎^\mathrm{𝟐}\mathbf{\right)}\mathbf{}𝒇\mathbf{\left(}𝝉\mathbf{\right)}}}\mathbf{.}$$ (25) Here $`\stackrel{\mathbf{~}}{\text{W}}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}`$ stands for any incomplete or complete AO expansion of the considered above functions; the quantity $`𝒎^\mathrm{𝟐}`$ is of order $`𝑶\mathbf{(}\mathrm{𝟏}\mathbf{)}`$. Property (25) means the non-sensitivity of the entire formalism with respect to variation of the mass shell within $`𝑶\mathbf{(}𝜶\mathbf{)}`$. Moreover, it means independence of the formalism from the UV renormalization scheme. Indeed, at the one-loop level the transition, for example, from the MS scheme to the on-mass-shell (OMS) scheme is described by <sup>5</sup><sup>5</sup>5 Remember, we do not consider contributions to the numerators of propagators. Therefore, we disregard the multiplicative wave function renormalization. $$\begin{array}{c}\\ 𝚺_{𝑶𝑴𝑺}\mathbf{(}𝒑^\mathrm{𝟐}\mathbf{)}\mathbf{=}𝚺\mathbf{(}𝒑^\mathrm{𝟐}\mathbf{)}\mathbf{}\text{Re}𝚺\mathbf{(}𝑴_{𝑶𝑴𝑺}^\mathrm{𝟐}\mathbf{)}\mathbf{}\mathbf{(}𝒑^\mathrm{𝟐}\mathbf{}𝑴_{𝑶𝑴𝑺}^\mathrm{𝟐}\mathbf{)}\mathbf{\times }\text{Re}𝚺^{\mathbf{}}\mathbf{(}𝑴_{𝑶𝑴𝑺}^\mathrm{𝟐}\mathbf{)}\mathbf{,}\\ 𝑴_{𝑶𝑴𝑺}^\mathrm{𝟐}\mathbf{=}𝑴^\mathrm{𝟐}\mathbf{}\text{Re}𝚺\mathbf{(}𝑴_{𝑶𝑴𝑺}^\mathrm{𝟐}\mathbf{)}\mathbf{.}\end{array}$$ (26) It is obvious that (26) belongs to the class of transformations covered by (25). The transformation at the multi-loop level is controlled by formula (18). So this can be done in accordance with the standard recipes of the UV renormalization, which do not depend on presence of the “infrared” counterterms located on the mass shell (see also and references therein). Let us note that the transformation to another scheme of UV renormalization may be done (speculatively) before the squaring of the amplitude and the AO expansion. The consequent AO expansion by no means “feels” in what scheme the Green functions have been renormalized. The property (25) is trivial in cases of non-expanded $`\stackrel{\mathbf{~}}{\text{W}}`$ or its formal expansions. The nontrivial aspect is that it remains valid for the counterterms $`𝑬\mathbf{(}𝝉\mathbf{)}`$, too. However, this also can be understood if one notes that the transformation $`𝝉\mathbf{}𝝉\mathbf{}𝜶𝒎^\mathrm{𝟐}`$ does not affect the structure of the homogenization at the scaling $`𝝉\mathbf{}𝝃𝝉`$, $`𝜶\mathbf{}𝝃𝜶`$ (see Sect.3). As a result, the property (25) is valid in the most general case. Finally, we present one more property which represents doubtless independent interest. Property 5. From (25) follows the next recurrent formula for the coefficients $`𝒄_𝒏`$, $$𝒄_{𝒏\mathbf{}\mathrm{𝟏}}\mathbf{=}\frac{\mathrm{𝟏}}{𝒏}\mathbf{\left[}\frac{\mathrm{𝟏}}{𝜶}\frac{\mathbf{}}{\mathbf{}𝒉_\mathrm{𝟎}}\mathbf{}\underset{𝒓\mathbf{=}\mathrm{𝟎}}{\overset{𝑵\mathbf{}𝒏}{\mathbf{}}}\mathbf{\left(}𝒉_\mathrm{𝟎}^{\mathbf{(}𝒓\mathbf{+}\mathrm{𝟏}\mathbf{)}}\frac{\mathbf{}}{\mathbf{}𝒉_\mathrm{𝟎}^{\mathbf{(}𝒓\mathbf{)}}}\mathbf{+}𝒇_\mathrm{𝟎}^{\mathbf{(}𝒓\mathbf{+}\mathrm{𝟏}\mathbf{)}}\frac{\mathbf{}}{\mathbf{}𝒇_\mathrm{𝟎}^{\mathbf{(}𝒓\mathbf{)}}}\mathbf{\right)}\mathbf{+}\frac{\mathbf{}}{\mathbf{}𝑴^\mathrm{𝟐}}\mathbf{\right]}𝒄_𝒏\mathbf{.}$$ (27) Here $`𝒄_𝒏\mathbf{=}𝒄_𝒏\mathbf{(}𝑴^\mathrm{𝟐}\mathbf{;}𝜶\mathbf{;}𝒉_\mathrm{𝟎}\mathbf{,}\mathbf{}𝒉_\mathrm{𝟎}^{\mathbf{(}𝑵\mathbf{}𝒏\mathbf{)}}\mathbf{;}𝒇_\mathrm{𝟎}\mathbf{,}\mathbf{}𝒇_\mathrm{𝟎}^{\mathbf{(}𝑵\mathbf{}𝒏\mathbf{)}}\mathbf{)}`$ are the constants independent from $`𝝉`$, $`𝒏`$ runs values $`\mathrm{𝟎}\mathbf{}𝒏\mathbf{}𝑵`$, where $`𝑵`$ is the maximum degree of the derivative of the delta-functions in $`𝑬\mathbf{(}𝝉\mathbf{)}`$. Formula (27) is written down with possible dependence of the coefficients $`𝒄_𝒏`$ on the parameter $`𝑴^\mathrm{𝟐}`$. An essential point for the derivation of (27) is the expansion in powers of $`𝜶`$ of the delta-function $`𝜹\mathbf{(}𝝉\mathbf{}𝜶𝒎^\mathrm{𝟐}\mathbf{)}`$ and its derivatives in r.h.s. of (25). The practical value of (27) consists in the opportunity to check the results of calculation of counterterms, or to determine the “lower” coefficient $`𝒄_{𝒏\mathbf{}\mathrm{𝟏}}`$ to within $`𝑶\mathbf{(}𝜶^𝑳\mathbf{)}`$ if the “higher”coefficient $`𝒄_𝒏`$ is known to within $`𝑶\mathbf{(}𝜶^{𝑳\mathbf{+}\mathrm{𝟏}}\mathbf{)}`$. ## 5 Massless particles exchange The problem of taking into consideration massless particles (photons) requires a special analysis because massless-particle contributions to the self-energy of a massive (unstable) particle involve the singularity of $`𝝉\mathbf{\times }\mathrm{𝐥𝐧}\mathbf{(}𝝉\mathbf{}\text{i}\mathrm{𝟎}\mathbf{)}`$. The first derivative of this expression is not defined at zero. So already the first corrections to coefficients (15) become uncertain. One way to solve this problem is to introduce a regularizing mass for massless particles (a soft-photon mass). Then the non-analyticity at $`𝝉\mathbf{=}\mathrm{𝟎}`$ disappears, and this fact opens a way to use without problems the above-derived formulas of the AO expansions. In a properly defined probability, taking into account radiation of the real soft photons, all contributions singular in the photon mass should cancel among themselves. That leads to continuity of probability with respect to the photon mass. So, while solving qualitative problems, one could not worry about the presence of the photon mass, because in the final results the dependence on it may be eliminated by the ordinary passage to the zero-mass limit. It should be stressed that in the above discussion the photon mass should not necessarily be infinitesimal. Moreover, it may be chosen to be so large as far as is needed for the solution of a particular qualitative problem. In what follows we use the photon-mass insertion for the consistent determination of the orders of expansion in the coupling constant of the photon contributions to the probability. After completion of all cancellations and after the passage to the zero-mass limit these orders will not change. It is worth noticing that the photon mass may be introduced in such a way as to guarantee validity of WI. This can be seen by consideration of the problem in the Stueckelberg formalism.<sup>6</sup><sup>6</sup>6 The author is grateful to A.A.Slavnov for indication of this fact Moreover, a photon mass (and a gluon mass, as well) may be introduced in a totally gauge-invariant fashion . So, the photon mass will not brake down gauge cancellations. Another way to solve the problem of massless particles is based on usage of the regularization property of the parameter $`𝜶`$. This method is deeply involved in the context of the AO expansion. It should be noted, however, that it is able to cure only those IR divergences of which the origin is connected with the emergence of an additional singularity in $`𝜶`$ at $`𝜶\mathbf{}\mathrm{𝟎}`$. In reality such divergences appear only in the diagrams where the soft momenta of massless particles come into the propagator of unstable particle considered on shell. Let us note that cancellation of these IR divergences means cancellation of the corresponding singularities in the coupling constant, and vice versa. The idea of the method of consists of a stepwise expansion of the full Green functions squared: first in the contributions of the massless particles only, and then in other vertices using the AO technique. (The first step will not lead to an infinite series of equal orders in $`𝜶`$ in the resonance region owing to Property 2 and the property $`\text{Im}𝚺_\mathrm{𝟏}^{\text{bos}}\mathbf{(}\mathrm{𝟎}\mathbf{)}\mathbf{=}\mathrm{𝟎}`$ for the massless-particle contributions.) Acting in the framework of the method of , after the first step one gets the modified Green functions with unstable propagators not containing contributions of the massless particles. However, this simplification is not indisputable, since there will appear a lot of configurations with the explicit contributions of photons for which some special counterterms will be needed. (In fact they will regularize the products $`𝚺_𝒍\mathbf{\times }\text{W}_{\mathrm{𝟏}𝒏}^𝒎`$; see the discussion in the previous section.) Basing ourselves on considerations of unitarity one can justify cancellation of the considering class of IR divergences . However, the property of the gauge cancellations still remain to be proved. So this method of handling the IR divergences seems not too good, especially since it does not provide a complete solution of the problem and requires a lot of additional efforts. ## 6 The gauge cancellations Now we are ready to give a proof of the gauge cancellations in any order of the modified PT. We present our argumentation in a rather general form, applied basically to any unstable particle in electroweak theory. The main idea is to separate in the probability the contributions certainly possessing the property of gauge cancellations from the problematic contributions. That allows us to concentrate on study the problematic contributions only. Let us begin with some preliminary notes. First of all we determine the photon-mass regularization for IR divergences. The advantages of this method have been discussed in the previous section. Here we stress that this method permits one to consider soft photons like ordinary massive particles. That means that their contributions either may be referred completely to vertex blocks, or, in other cases, they suppress the resonant behaviour of the corresponding contributions. Really, an exchange by a massive particle suppresses the *on-shell* propagating of unstable particles both before and after the emission of the massive particle. As a result, the problem of instability is reduced solely to configurations which include only the pairs of unstable propagators of equal mass and momenta, placed on both sides of the cut of unitarity. It is worth remembering that the mass insertion for photons will not violate gauge cancellations (see previous section). Moreover, since the probability is a continuous function of the photon mass the dependence on it may be eliminated in the final results by the ordinary passage to the zero-mass limit. The next note concerns the unphysical pole contribution to the vector boson propagators. Assuming the parameterization for the self-energy $$𝚺_{𝝁𝝂}\mathbf{(}𝒑\mathbf{)}\mathbf{=}𝚺𝒈_{𝝁𝝂}\mathbf{+}𝚺_𝑳𝒑_𝝁𝒑_𝝂\mathbf{,}$$ (28) we come to the following explicit expression for the full propagator in $`𝑹_𝝃`$-gauge: $$𝑫_{𝝁𝝂}\mathbf{(}𝒑\mathbf{)}\mathbf{=}\frac{𝒈_{𝝁𝝂}\mathbf{}𝒑_𝝁𝒑_𝝂\mathbf{/}𝒑^\mathrm{𝟐}}{𝒑^\mathrm{𝟐}\mathbf{}𝑴^\mathrm{𝟐}\mathbf{+}𝚺}\mathbf{+}𝝃\frac{𝒑_𝝁𝒑_𝝂\mathbf{/}𝒑^\mathrm{𝟐}}{𝒑^\mathrm{𝟐}\mathbf{}𝝃𝑴^\mathrm{𝟐}\mathbf{+}𝝃\mathbf{\left(}𝚺\mathbf{+}𝒑^\mathrm{𝟐}𝚺_𝑳\mathbf{\right)}}\mathbf{.}$$ (29) Here the first term represents the product of the spin factor on the scalar propagator $`𝚫\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}`$ introduced in (1). The second term describes the unphysical contribution. In the conventional PT it is cancelled due to WI by contributions of the other unphysical states. Consequently, the second term in (29) does not lead to non-integrable singularity in the phase space. In the modified PT this fact allows us to take into account the second term in the conventionally expanded form. Then, the property of gauge cancellations will be controlled by the properties of the first term, taken in an absolute value and having been squared. Note that the cross terms resulting from the squaring of the propagator (29) will not lead to non-integrable singularities in the phase space. So they also may be taken into account as conventionally expanded. Now let us show that the first term in (29), being taken into consideration in framework of the modified PT, does not break the gauge cancellations in the probability. Remember that in accordance with (24) its contribution may be presented in the form of a sum of two terms, where one is conventional and the other is anomalous. Let us start from analysis of the anomalous term contribution. Due to its additivity the simplest way to perform the analysis is to assume for a moment that the squared unstable propagator consists in this anomalous term only. Firstly we consider the case with a single unstable particle production. Let us note that the complementary part of diagram of unitarity with respect to the given propagator squared, by virtue of the delta-function *without a derivative* in the anomalous term, is taken strictly on mass shell. Moreover, the details of the definition of the mass shell are inessential here and may be omitted. (This is because these details have meaning of corrections, but the anomalous term describes the highest order within the given precision.) From these two facts follows the conclusion that the sum of the complementary parts is gauge invariant, since actually it coincides (up to factors) with the product of two $`𝑺`$-matrix elements squared, with one describing the on-shell production of the unstable particle and another one describing its decay. Furthermore, in the anomalous term itself one may neglect the gauge-dependent contributions, if in this case they are in $`\text{Im}𝚺_{𝒏\mathbf{+}\mathrm{𝟏}}\mathbf{(}\mathrm{𝟎}\mathbf{)}`$, since they are beyond the given precision (see Property 3 and the discussion therein). So the total sum of unitarity diagrams that include the anomalous term within the given precision is gauge invariant. In the case of multiple production of unstable particles let us single out any unstable particle and, again, consider its contribution of an anomalous term. Since the anomalous term is of the highest order, the contributions of all other unstable particles should be considered in the leading order of the AO expansion. By virtue of (7), (9), and (15) the leading-order contributions are determined by the delta-function *without a derivative* with a coefficient including the one-loop $`𝒇_\mathrm{𝟎}`$ only as a non-trivial factor. In view of the gauge invariance of the one-loop $`𝒇_\mathrm{𝟎}`$ (see, for instance, ) follows, once again, the gauge invariance of the total sum of complementary parts of the diagrams of unitarity. So, the gauge invariance in this case occurs, too. Due to additivity, a similar reasoning may be repeated for any other unstable particle of multiple production, and in each case we will reach the property of gauge invariance of the corresponding contribution to the probability. Now let us proceed to the analysis of all other contributions, which do not at all include an anomalous term. Such contributions are controlled by the first term in (24). Also, all non-resonant contributions belong to this class of contributions. As a whole, these contributions are described by the amplitude squared determined in the conventional approach with Dyson resummation. The property of gauge cancellations in the amplitude so determined has already been shown in the framework of the background-field formalism. Let us emphasize that the only condition which has to be observed when considering these contributions is the Dyson resummation of all corrections to the self-energies up to $`𝒏`$-loops together with taking into consideration all $`𝒏`$-loop corrections to the vertices. So, treating $`\text{W}_{\mathbf{[}𝒏\mathbf{]}}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}`$ in (24) in such a manner, we automatically get the property of gauge cancellations in the probability.<sup>7</sup><sup>7</sup>7 It should be noted that within the background-field formalism after all gauge cancellations there may remain some residual dependence on the quantum gauge parameter. This phenomenon has been considered in the original work of , where it has been stressed also that this residual dependence will not affect the high-energy behaviour of the amplitude. An idea for how to solve the problem of the residual gauge dependence is discussed in Sect.8. The above discussion completes a construction which provides both the gauge cancellations and the necessary precision of the description in the sense of an expansion in the coupling constant. Basing ourselves on this result one may do the next step: proceeding to the complete AO expansion of the probability. Then the property of gauge cancellations within the given precision must take place as well, because AO only expands in powers of the coupling constant the expression which possesses the property of gauge cancellations. With this note we complete the general proof of gauge cancellations in the approach of the modified PT. ## 7 Generalization of the fermion-loop scheme If the analysis may be restricted by NLO precision, for example for the case of W-pair production studied at LEP2, the gauge cancellations may be proved within the usual formalism, without applying the background-field method. The key point is the well-known result on gauge cancellations in the so-called fermion-loop scheme . Recall that it consists of including all fermionic one-loop corrections in tree-level amplitudes and Dyson resumming self-energies. The difficulty of this scheme is twofold. First, it is not known how to incorporate the one-loop bosonic corrections into this scheme without spoiling the gauge invariance. Second, there is a problem with gauge invariance while taking into account the two-loop corrections to self-energies in the denominators of unstable propagators, which is necessary in the resonance region. Both these problems may be solved within the modified PT approach with usage of the AO technique. Let us begin with the problem of the two-loop corrections. As we have seen above, they can be taken into consideration without breaking gauge invariance by adding the anomalous terms to the probability. In view of (24) that may be done by means of the formula $$\text{W}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{=}\text{W}_{\mathbf{[}\mathrm{𝟏}\mathbf{]}}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{}\frac{\text{Im}𝚺_\mathrm{𝟐}\mathbf{(}\mathrm{𝟎}\mathbf{)}}{\mathbf{\left[}\text{Im}𝚺_\mathrm{𝟏}\mathbf{(}\mathrm{𝟎}\mathbf{)}\mathbf{\right]}^\mathrm{𝟐}}𝝅𝜹\mathbf{(}𝝉\mathbf{)}\mathbf{+}𝑶\mathbf{(}𝜶\mathbf{)}\mathbf{.}$$ (30) Here $`\text{W}_{\mathbf{[}\mathrm{𝟏}\mathbf{]}}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}`$ represents the unstable propagator squared with Dyson resummed all one-loop corrections. Remember that $`\text{W}_{\mathbf{[}\mathrm{𝟏}\mathbf{]}}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{=}𝑶\mathbf{(}𝜶^\mathbf{}\mathrm{𝟏}\mathbf{)}`$ at $`𝜶\mathbf{}\mathrm{𝟎}`$. Therefore, the second term in (30) describes the highest-order (NLO) correction. So, any ambiguities in its definition, in particular those which concern the problem of the gauge invariance, may be referred to the discarded terms of order $`𝑶\mathbf{(}𝜶\mathbf{)}`$. (See Property 3 and the discussion therein.) The problem of one-loop bosonic corrections may be solved, too. Let us group these corrections into two classes. To the first class we refer the corrections to self-energy of unstable particles. To the second class we refer the corrections to the vertex factors, the corrections due to exchange processes, and due to the real (soft) photons. The corrections of the first class can easily be taken into consideration by means of the fact that the imaginary parts of the on-shell bosonic corrections to the self-energies of W- and Z-bosons are zero . Owing to this fact and Property 2 of Sect. 4 these corrections may be transferred from the denominators of unstable propagators to their numerators without loss of precision. Moreover, in an OMS-like scheme of UV renormalization, where the renormalized self-energies satisfy the condition $`\text{Re}𝚺_\mathrm{𝟏}\mathbf{(}\mathrm{𝟎}\mathbf{)}\mathbf{=}\text{Re}𝚺_\mathrm{𝟏}^{\mathbf{}}\mathbf{(}\mathrm{𝟎}\mathbf{)}\mathbf{=}\mathrm{𝟎}`$, one has the following approximation: $$\text{W}_{\mathbf{[}\mathrm{𝟏}\mathbf{]}}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{=}\text{W}_{\mathrm{𝟏}𝑭}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{+}𝑶\mathbf{(}𝜶\mathbf{)}\mathbf{.}$$ (31) Here $`\text{W}_{\mathrm{𝟏}𝑭}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}`$ represents the propagator squared where only the fermionic one-loop corrections are Dyson resummed. Substituting (31) into (30) we come to the formula which leads to gauge cancellations. In order to see this, we should only repeat the reasoning of the previous section keeping in mind the gauge invariance in the fermion-loop scheme. However, the bosonic corrections of the second class have not yet been taken into account. In order to do that let us make use the fact that within the NLO approximation and in the presence of the *corrections*, the quantity $`\text{W}_{\mathrm{𝟏}𝑭}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}`$ must be taken into consideration in the leading-order approximation only. Remember that in this approximation $`\text{W}_{\mathrm{𝟏}𝑭}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{=}𝝅\mathbf{/}\mathbf{(}𝜶𝒇_{\mathrm{𝟎}𝑭}\mathbf{)}\mathbf{\times }𝜹\mathbf{(}𝝉\mathbf{)}`$, and this expression is explicitly gauge invariant. Moreover, since the delta-function enters without derivatives, the sum of all factors appearing in its presence in the unitarity diagrams are also gauge invariant (see previous section). Thus, we come to the following recipe of the generalization. We formulate it having in mind the total cross-section for the typical LEP2 processes CC10, CC11 and CC20, which have been studied in the framework of the fermion-loop scheme . In fact, the generalization consists of adding to the probability two types of corrections. The corrections of the first type describe the anomalous contributions. In the NLO approximation they look like the cross-section of the pair on-shell production of unstable particles taken in the leading-order approximation, times the leading-order decay blocks of unstable particles, and times the “anomalous” factor. Indeed, the presence of the anomalous factor means that all other factors should be taken in the leading order. However, in the leading-order approximation only the double-resonant subprocesses contribute to $`𝒆^\mathbf{+}𝒆^{\mathbf{}}\mathbf{}\mathrm{𝟒}𝒇`$. (In fact these subprocesses are of CC03 class.) That follows from the fact that only these subprocesses include the factor $`\mathrm{𝟏}\mathbf{/}𝜶^\mathrm{𝟐}`$ originating from the product of two unstable propagators squared. Now, let us remember that the additive anomalous term in (30) includes the delta-function as well as $`\text{W}_{\mathrm{𝟏}𝑭}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}`$ does in the leading-order approximation. So, the anomalous terms always contribute “on-shell”. The corrections of the second type may be represented, again, as the cross-section of the pair on-shell production of unstable particles, multiplied by their decay blocks. However, instead of “anomalous” factors, they include the bosonic corrections to the vertex blocks and the corrections due to the real soft-photons. The examples of the unitarity diagrams which generate contributions of this type are shown in Fig. 1. With the non-zero mass of the photons all these diagrams include exactly *two* pairs of unstable propagators of identical mass and momenta in both sides of the cut. (The only exception, configuration (1.e), is discussed below.) Therefore, they include *two* delta-functions of the corresponding kinematic variables, which make these configurations on shell. The sum of all such configurations represents the product of the cross-section of the pair on-shell production of unstable particles and their decay blocks. Since these quantities are continuous functions of the photon-mass, the dependence on it may be eliminated in the very end of the calculation by the ordinary limiting procedure. Notice that the mentioned property of continuity follows directly from the well-known theorem for cross-sections with the real soft-photon contributions. Now let us discuss the mentioned above exceptional configuration (1.e). Strictly speaking it should not be considered to belong to the unitarity diagrams, since it describes the self-energy correction to the unstable propagator, which has already been taken into consideration by formula (31). Nevertheless, while considering the cross-section of the pair on-shell production, one has to take into account the virtual soft-photon insertion to the external legs, which is due to the wave function renormalization. The configuration (1.e) was added to the list of diagrams of Fig. 1 only in order to indicate this fact. It should be noted that among the diagrams of Fig. 1 there are both factorizable and non-factorizable configurations in the sense of the classification of . Nevertheless, at the intermediate step, when the soft-photons are supplied with the mass, the non-factorizable configurations of Fig. 1 become factorizable. All other non-factorizable corrections, which do not provide this property, produce such configurations that include *less* than two pairs of unstable propagators of identical mass and momenta in both sides of the cut. Therefore, they do not include the leading-order factor $`\mathrm{𝟏}\mathbf{/}𝜶^\mathrm{𝟐}`$. In view of the presence of the additional factor $`𝜶`$ due to the bosonic corrections they contribute beyond the NLO approximation. The examples of configurations of this type are shown in Fig. 2. It is worth noticing that they may not be considered as corrections to the cross-section discussed above of the pair on-shell production of unstable particles, or to their decay blocks. The above discussion leads us to the following formulas for the total cross-section: $$𝝈\mathbf{(}𝒔\mathbf{)}\mathbf{=}\underset{\mathrm{𝟎}}{\overset{\mathrm{𝟏}}{\mathbf{}}}\text{d}𝒛\mathit{\varphi }\mathbf{(}𝒛\mathbf{;}𝒔\mathbf{)}\widehat{𝝈}\mathbf{(}𝒛𝒔\mathbf{)}\mathbf{,}$$ (32) $$\widehat{𝝈}\mathbf{(}𝒔\mathbf{)}\mathbf{=}\underset{\mathrm{𝟎}}{\overset{𝒔}{\mathbf{}}}\text{d}𝒔_\mathbf{+}\underset{\mathrm{𝟎}}{\overset{\mathbf{(}\sqrt{𝒔}\mathbf{}\sqrt{𝒔_\mathbf{+}}\mathbf{)}^\mathrm{𝟐}}{\mathbf{}}}\text{d}𝒔_{\mathbf{}}\widehat{𝝈}_\mathrm{𝟎}\mathbf{(}𝒔\mathbf{;}𝒔_\mathbf{+}\mathbf{,}𝒔_{\mathbf{}}\mathbf{)}\mathbf{.}$$ (33) Here $`𝝈\mathbf{(}𝒔\mathbf{)}`$ is the experimentally measured cross-section at the center-of-mass energy squared $`𝒔`$. $`\mathit{\varphi }\mathbf{(}𝒛\mathbf{;}𝒔\mathbf{)}`$ is the “flux” function<sup>8</sup><sup>8</sup>8In general case, $`\varphi (z;s)`$ includes experimental cuts and hardware factors. describing the contributions of the initial- and final-state photon radiations with large IR and collinear logarithms, $`\widehat{𝝈}\mathbf{(}𝒔\mathbf{)}`$ stands for the hard scattering improved Born cross-section at the reduced center-of-mass energy squared (see and references therein). It should be emphasized that $`\widehat{𝝈}\mathbf{(}𝒔\mathbf{)}`$ contributes like a distribution to $`𝝈\mathbf{(}𝒔\mathbf{)}`$, since $`\widehat{𝝈}\mathbf{(}𝒔\mathbf{)}`$ is smeared by the flux function over a wide range of the kinematic variable ($`\mathit{\varphi }\mathbf{(}𝒛\mathbf{;}𝒔\mathbf{)}`$ is peaked at $`𝒛\mathbf{=}\mathrm{𝟏}`$ and has a tail till a cut at lower values of $`𝒛`$). Expression (33) represents $`\widehat{𝝈}\mathbf{(}𝒔\mathbf{)}`$ in a form with explicit integration over the virtualities of unstable particles (over the invariant masses of the corresponding final states). The quantity $`\widehat{𝝈}_\mathrm{𝟎}\mathbf{(}𝒔\mathbf{)}`$ reads $`\widehat{𝝈}_\mathrm{𝟎}\mathbf{(}𝒔\mathbf{;}𝒔_\mathbf{+}\mathbf{,}𝒔_{\mathbf{}}\mathbf{)}\mathbf{=}\widehat{𝝈}_\mathrm{𝟎}^{\text{fermion-loop-scheme}}\mathbf{(}𝒔\mathbf{;}𝒔_\mathbf{+}\mathbf{,}𝒔_{\mathbf{}}\mathbf{)}`$ $`\mathbf{}\widehat{𝝈}_\mathrm{𝟎}^{\text{on-shell,tree}}\mathbf{(}𝒔\mathbf{;}𝑴_\mathbf{+}\mathbf{,}𝑴_{\mathbf{}}\mathbf{)}\mathbf{\times }\mathrm{𝟐}𝜶\text{Im}𝚺_\mathrm{𝟐}\mathbf{(}\mathrm{𝟎}\mathbf{)}\mathbf{/}\text{Im}𝚺_\mathrm{𝟏}\mathbf{(}\mathrm{𝟎}\mathbf{)}\mathbf{\times }{\displaystyle \underset{𝜿\mathbf{=}\mathbf{\pm }}{\mathbf{}}}𝜹\mathbf{(}𝒔_𝜿\mathbf{}𝑴_𝜿^\mathrm{𝟐}\mathbf{)}\mathbf{\times }𝑩𝑹_𝜿^{\text{tree}}`$ $`\mathbf{+}\widehat{𝝈}_\mathrm{𝟎}^{\text{on-shell,tree}}\mathbf{(}𝒔\mathbf{;}𝑴_\mathbf{+}\mathbf{,}𝑴_{\mathbf{}}\mathbf{)}\mathbf{\times }{\displaystyle \underset{𝜿\mathbf{=}\mathbf{\pm }}{\mathbf{}}}𝜹\mathbf{(}𝒔_𝜿\mathbf{}𝑴_𝜿^\mathrm{𝟐}\mathbf{)}\mathbf{\times }𝑩𝑹_𝜿^{\text{tree/boson-one-loop+real-photon}}`$ $`\mathbf{+}\widehat{𝝈}_\mathrm{𝟎}^{\text{on-shell, boson-one-loop + real-photon}}\mathbf{(}𝒔\mathbf{;}𝑴_\mathbf{+}\mathbf{,}𝑴_{\mathbf{}}\mathbf{)}\mathbf{\times }{\displaystyle \underset{𝜿\mathbf{=}\mathbf{\pm }}{\mathbf{}}}𝜹\mathbf{(}𝒔_𝜿\mathbf{}𝑴_𝜿^\mathrm{𝟐}\mathbf{)}\mathbf{\times }𝑩𝑹_𝜿^{\text{tree}}\mathbf{.}`$ (34) Here the first term represents the result of the conventional fermion-loop scheme. All other terms describe its generalization. Factor 2 in the second term reflects the presence of two intermediate unstable particles of equal masses; $`𝑩𝑹_\mathbf{\pm }^{\text{tree}}`$ denotes their on-shell branching in the tree approximation. In the third term one $`𝑩𝑹`$ is determined with the bosonic corrections to the partial width. (In fact the third term includes the sum of two subterms, with the modified $`𝑩𝑹`$ for one of two unstable particle.) In (34) we have used the relation $`𝜶\text{Im}𝚺_\mathrm{𝟏}\mathbf{(}\mathrm{𝟎}\mathbf{)}\mathbf{=}𝑴𝚪_\mathrm{𝟎}\mathbf{(}𝑴\mathbf{)}`$, with $`𝚪_\mathrm{𝟎}\mathbf{(}𝑴\mathbf{)}`$ being the total on-shell width at tree level. This relation follows from unitarity and may be verified by direct calculation . Formula (34), in principle, may be further simplified by carrying out the complete AO expansion in its first term. The key formula for this simplification, written in an OMS-like scheme of UV renormalization, is as follows: $$\text{W}_\mathrm{𝟏}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}\mathbf{=}\mathbf{\left[}𝜶\text{Im}𝚺_\mathrm{𝟏}\mathbf{(}\mathrm{𝟎}\mathbf{)}\mathbf{\right]}^\mathbf{}\mathrm{𝟏}𝝅𝜹\mathbf{(}𝝉\mathbf{)}\mathbf{+}𝑽𝑷\frac{\mathrm{𝟏}}{𝝉^\mathrm{𝟐}}\mathbf{+}𝑶\mathbf{(}𝜶\mathbf{)}\mathbf{.}$$ (35) Substituting (35) into (34) we finally get $`\widehat{𝝈}_\mathrm{𝟎}\mathbf{(}𝒔\mathbf{;}𝒔_\mathbf{+}\mathbf{,}𝒔_{\mathbf{}}\mathbf{)}\mathbf{=}𝑽𝑷\widehat{𝝈}_\mathrm{𝟎}^{\text{on/off-shell, tree}}\mathbf{(}𝒔\mathbf{;}𝑴_\mathbf{+}\mathbf{,}𝒔_{\mathbf{}}\mathbf{)}\mathbf{\times }𝜹\mathbf{(}𝒔_\mathbf{+}\mathbf{}𝑴_\mathbf{+}^\mathrm{𝟐}\mathbf{)}\mathbf{\times }𝑩𝑹_\mathbf{+}^{\text{tree}}`$ $`\mathbf{+}𝑽𝑷\widehat{𝝈}_\mathrm{𝟎}^{\text{off/on-shell, tree}}\mathbf{(}𝒔\mathbf{;}𝒔_\mathbf{+}\mathbf{,}𝑴_{\mathbf{}}\mathbf{)}\mathbf{\times }𝜹\mathbf{(}𝒔_{\mathbf{}}\mathbf{}𝑴_{\mathbf{}}^\mathrm{𝟐}\mathbf{)}\mathbf{\times }𝑩𝑹_{\mathbf{}}^{\text{tree}}`$ $`\mathbf{}\widehat{𝝈}_\mathrm{𝟎}^{\text{on-shell,tree}}\mathbf{(}𝒔\mathbf{;}𝑴_\mathbf{+}\mathbf{,}𝑴_{\mathbf{}}\mathbf{)}\mathbf{\times }\mathrm{𝟐}𝜶\text{Im}𝚺_\mathrm{𝟐}\mathbf{(}\mathrm{𝟎}\mathbf{)}\mathbf{/}\text{Im}𝚺_\mathrm{𝟏}\mathbf{(}\mathrm{𝟎}\mathbf{)}\mathbf{\times }{\displaystyle \underset{𝜿\mathbf{=}\mathbf{\pm }}{\mathbf{}}}𝜹\mathbf{(}𝒔_𝜿\mathbf{}𝑴_𝜿^\mathrm{𝟐}\mathbf{)}\mathbf{\times }𝑩𝑹_𝜿^{\text{tree}}`$ $`\mathbf{+}\widehat{𝝈}_\mathrm{𝟎}^{\text{on-shell,tree}}\mathbf{(}𝒔\mathbf{;}𝑴_\mathbf{+}\mathbf{,}𝑴_{\mathbf{}}\mathbf{)}\mathbf{\times }{\displaystyle \underset{𝜿\mathbf{=}\mathbf{\pm }}{\mathbf{}}}𝜹\mathbf{(}𝒔_𝜿\mathbf{}𝑴_𝜿^\mathrm{𝟐}\mathbf{)}\mathbf{\times }𝑩𝑹_𝜿^{\text{tree/boson-one-loop+real-photon}}`$ $`\mathbf{+}\widehat{𝝈}_\mathrm{𝟎}^{\text{on-shell, tree + one-loop + real-photon}}\mathbf{(}𝒔\mathbf{;}𝑴_\mathbf{+}\mathbf{,}𝑴_{\mathbf{}}\mathbf{)}\mathbf{\times }{\displaystyle \underset{𝜿\mathbf{=}\mathbf{\pm }}{\mathbf{}}}𝜹\mathbf{(}𝒔_𝜿\mathbf{}𝑴_𝜿^\mathrm{𝟐}\mathbf{)}\mathbf{\times }𝑩𝑹_𝜿^{\text{tree}}\mathbf{.}`$ (36) Here the first two terms and the leading-order tree contribution in the last term originate from the fermion-loop-scheme term in formula (34). Namely, the first two terms accumulate the contributions originating from CC03 subprocesses and simultaneously from subprocesses of single vector boson production. In the case of CC03 subprocesses the symbol $`𝑽𝑷`$ means that the corresponding off-shell unstable propagator squared is approximated, in accordance with (35), by $`𝑽𝑷\mathbf{\hspace{0.17em}1}\mathbf{/}𝝉^\mathrm{𝟐}`$. Another unstable particle in this case is considered as real, produced on shell. In the case of single production the vector (unstable) boson is considered as real, too. The symbol $`𝑽𝑷`$ in this case is superfluous and has to be omitted. It should be noted that in the above discussion the usage of prescription of the principal value in the case of the CC03 subprocesses is necessary, because the usage of any other prescription for $`\mathrm{𝟏}\mathbf{/}𝝉^\mathrm{𝟐}`$ in (31) and (35) may change the results in (36). This remark, however, does not concern the anomalous term, since it arises from the contribution singular in $`𝜶`$ to $`\text{W}_{\mathrm{𝟏𝟏}}\mathbf{(}𝜶\mathbf{;}𝝉\mathbf{)}`$ (see (20) and the note after (22)). The above results may be easily generalized to other processes with unstable particles production, including the processes of the “NC”-type and of the “mixed” type. In the latter case there appear the additional terms in (34) and (36) of almost the same structure, that describe the ZZ-contributions (of course, configurations of Fig. 1 are not relevant in the ZZ-case). The generalization for the case of the differential cross-sections should not lead to difficulties either. However, the discussion of this question is beyond the scope of the present paper. ## 8 Discussion In this paper we have found a practical method that ensures both the gauge cancellations and a fixed precision of description of processes mediated by the production and decay of unstable particles in electroweak theory. The solution has been found in the framework of the modified PT which consists of an expansion in the coupling constant directly of probabilities instead of amplitudes . In the general case and within any fixed precision the proposed method is based on the results obtained earlier in the framework of the background-field formalism. Within NLO precision and in the case of W- and/or Z-pair production we have found a way also in the usual formalism, basing ourselves on results of the fermion-loop scheme . But in contrast to the fermion-loop scheme we perform the description explicit taking into account the bosonic corrections and the two-loop corrections to the self-energy of unstable particles in the resonance region. From the practical point of view the latter result apparently is the most important one obtained in this paper. It should be noted that the result on gauge cancellations in the completely expanded probability generally was expected since the appearance of . This work demonstrated that calculation of the probability of a process mediated by unstable particle production may be reduced to the regular fixed-order calculation. The latter fact gave a reason to think that the problem of the gauge invariance was solved, too . However, on closer examination of the problem a large distance to obtaining the result was found. Indeed, the Feynman rules in the modified PT and in the conventional PT coincide only out of the mass shell. So, the gauge invariance a priori occurs only in this area of kinematic variable, but not in a neighborhood of the mass shell, where one has the non-standard contributions of the modified PT. However, due to the delta-functions these non-standard contributions do not vanish irrespectively of that as far as the neighborhood is small. Moreover, the properties of these non-standard contributions are unknown in advance. In particular, it is not known whether they are gauge invariant or not. This becomes especially clear in the higher orders of the AO expansion, beginning with NLO, where the non-standard contributions include the derivatives of the delta-functions. In the present paper we avoid this difficulty by making use of incomplete AO expansions of the unstable propagators squared. This method allows us to carry out a comparison of the results of the modified PT with those of the conventional PT with Dyson resummation. The special role in this scheme is assigned to the anomalous additive term that corrects the result of incomplete expansions of the propagator squared in the vicinity of mass shell. We draw attention to a particular problem that arises when comparing the results obtained in the modified PT with the results obtained in the conventional approach with use of the background-field formalism. (It should be emphasized that there is no such problem in the conventional formalism on using the fermion-loop scheme.) This is the problem of the residual dependence on the quantum gauge parameter which still remains in the background-field formalism after all gauge cancellations . In the general case this dependence may pass on to the results of the modified PT. Nevertheless, *in the probability* owing to the phenomenon of changing of the order of particular contributions, which in turn is a corollary of the singularity in the coupling constant in the propagator squared, one may expect that this dependence will drop out within the given precision. Note that we indicate only an opportunity of giving a solution to the problem, but not the actual solution. However, earlier it was not clear even how to initiate a solution . Now let us discuss the generalization of the fermion-loop scheme. Remember that it provides both the NLO precision and the gauge invariance. As a matter of fact the generalization consists in the instruction on how to use the already known results of the former calculations. Indeed, the first term in formula (34) and the last two terms are already known . The remaining second term is actually known too, since $`\text{Im}𝚺_\mathrm{𝟐}\mathbf{(}\mathrm{𝟎}\mathbf{)}`$ in it may be replaced by a one-loop correction to the width of W-boson divided by its mass (see Property 3 and discussion thereof). The last two terms in (34) present exhaustive description of the bosonic corrections, which formerly were a “stumbling block” for gauge invariance. In the proposed generalization they are located in the $`𝑺`$-matrices squared, with the one $`𝑺`$-matrix is for the on-shell production of unstable particles and the other ones are for their decays. The key observation that leads us to this simple result is the absence in the NLO approximation of non-factorizable corrections of Fig. 2 (see Sect. 7). This property demonstrates one of the differences of the approach offered with the well known double-pole approximation (DPA) . In fact there are more serious differences, too. The most obvious one is as follows. By definition, DPA does not take into account the single-pole contributions, while the fermion-loop scheme, and consequently our generalization, does. The mentioned missed contributions are of order $`𝚪_𝑾\mathbf{/}𝑴_𝑾`$, which amounts to 2–3%. This is too much compared with the current accuracy of LEP2 results. In view of this problem, DPA usually is applied for the calculation of the radiative correction only, but not of the Born term. The Born term is considered in the framework of the conventional PT with “by-hand” substitution of Breit–Wigner’s propagators instead of the free propagators for unstable particles . However, the latter operation violates the gauge invariance and leads to some error, too, and a certain estimate of this error does not exist. All that it is possible to say is that this error most likely is somewhat less than the shift of the amplitude caused by this substitution, but it is not clear how far less. Nevertheless, the latter quantity again is $`𝑶\mathbf{(}𝚪_𝑾\mathbf{/}𝑴_𝑾\mathbf{)}`$. So, the discussed approximation of the Born term cannot be considered satisfactory. Another serious difficulty of DPA is its inapplicability in the vicinity of phase-space boundaries. This effect arises as a result of the “mapping” procedure in the phase space or the “analytic continuation” of the constant residues at the double pole in the amplitude . First of all, it manifests itself near the threshold region, where the DPA uncertainties are blowing up. Another evident restriction of applicability of DPA is the region lying far off from resonance where the pole-scheme expansion cannot be viewed as an effective expansion in powers of $`𝚪_𝑾\mathbf{/}𝑴_𝑾`$. Remember that the proposed generalization of the fermion-loop scheme does not need the mapping procedure and does not use the pole expansion. Consequently, it should be free from these difficulties. In conclusion, let us discuss two problems that arise in the modified PT approach. In fact, the topic is about a comparison of the results of the modified PT with Breit–Wigner’s parameterization of unstable particles. The first problem concerns the definition of the “physical” mass and the width of unstable particles. In this connection it should be noted, first of all, that both these quantities are pseudo-observables that are to be determined on the basis of realistic observables . Next, let us note that in the framework of the modified PT both these quantities are secondary ones, which should be determined on the basis of primary objects, such as the renormalized Lagrangian mass, coupling constant, etc. Apparently, the most radical way consists of the identification of both these quantities with the position of the pole in the complex plane of the full unstable propagator (in the spirit of the pole scheme). This procedure is equivalent to finding a solution to the equation $`𝑴^\mathrm{𝟐}\mathbf{}𝒔_𝒑\mathbf{}𝚺\mathbf{(}𝒔_𝒑\mathbf{)}\mathbf{=}\mathrm{𝟎}`$ with $`𝒔_𝒑\mathbf{=}𝑴_𝒑^\mathrm{𝟐}\mathbf{}\text{i}𝑴_𝒑𝚪_𝒑`$. This operation may be done perturbatively . The second problem concerns the presence of the delta-functions and VP’s in the results of the complete AO expansion. (The actuality of this problem is not so high while making use of incomplete AO expansions as employed in this paper.) The problem may be designated as an illusory discrepancy between the presence of these singular functions in the amplitude squared and the notion of a continuity of the physical results as functions of the kinematic variables. This problem has been stated and discussed in . The essence of the solution is reduced to the observation that actually there is not a direct identity between the squared amplitude (that is usually calculated) and the genuine probability (that is observed). In fact, a formal expression for the probability, calculated on the basis of the unitarity diagrams, should necessarily be subjected to the operation of *integration*, or “*smearing*” with some weight function before it becomes an observable quantity.<sup>9</sup><sup>9</sup>9 It is worth noticing, that with absence of “smearing” there is no problem of non-integrable singularities in the phase space. In the latter case all calculations actually are carried out in the framework of the conventional PT. In processes with light and charged initial and/or final states such a smearing function is, first of all, the flux function that extracts and exponentiates the infrared and collinear singularities from the cross-section (see formula (32)). In the case of W- and/or Z-production the smearing is large enough: on some distance from threshold its final effect is about ten or more percent and it covers an area which is certainly greater than the unstable-particle width (see formulas (32) and (33)). Consequently, after the convolution the contributions of the delta-functions and VP’s become strongly distributed over a wide area. Mathematically, this fact means that the smearing may be considered as an effect of order $`𝑶\mathbf{(}\mathrm{𝟏}\mathbf{)}`$, which is required in the AO. Therefore, even in the case of ideal energy resolution in a given experiment, the presence of the singular functions in the results of the AO expansion becomes invisible, at least within the given precision of the description. The above conclusion should hold both for the total and differential cross-sections, since the differential cross-sections are determined through the convolution procedure, too . *Acknowledgements.* The author is grateful to D.Yu. Bardin, L.V. Kalinovskaya, and F.V. Tkachov for discussions of various problems connected with applications of the electroweak theory. Special thanks to F.V. Tkachov for the invitation to study the problems of instability and consultation on the results of his work . The author would also like to thank A.A. Slavnov for the valuable note about WI in presence of a photon mass, and E.E. Boos for the indication of the importance of the single-resonant subprocesses in W-pair production. It is a pleasure to thank B.A. Arbuzov, V.E. Rochev and S.R. Slabospitsky for support and useful discussions. This work was supported in part by the Russian Foundation for Basic Research, grant No. 99-02-18365.
warning/0002/hep-th0002061.html
ar5iv
text
# NON-ABELIAN FIELD THEORY OF STABLE NON-BPS BRANES ## 1 Introduction Non-BPS branes in type II string theory have recently attracted considerable attention. On one hand they are of interest because the are stable, non-perturbative states in string theory without preserving any supersymmetry. These states are therefore important to gain a deeper understanding of the non-perturbative spectrum of string theory. On the other hand, BPS D-branes have played a key role in the recent success of describing non-perturbative properties of supersymmetric Yang-Mills theory in terms of string theory . The hope is now that non-BPS states could play a similarly important role in the analysis of non-perturbative properties of non-supersymmetric Yang-Mills theory. Previous work in this direction has focused on D-branes in non-supersymmetric type 0 theory . A successful generalisation of the AdS/CFT correspondence to these non-supersymmetric backgrounds of a stack of D-branes has however been hampered by the tachyonic instability in the closed string sector of that theory<sup>2</sup><sup>2</sup>2For recent progress though see . In Sen’s non-BPS branes of type II theory such instabilities in the closed string sector are absent. Instead one has to deal with a tachyon in the open string sector. However, it turns out that these instabilities can be cured by considering type II theory on an appropriate orbifold. In doing so stable non-BPS branes have been obtained . While the possible application of these objects for a string-theoretic description of non-supersymmetric, non-abelian field theory has been suggested we are not aware of any concrete progress in this direction. In this paper we initiate a systematic analysis of the low energy dynamics of a stack of non-BPS branes in type II theory. In particular we construct the low energy effective field theory on a stack of non-BPS 3-branes. Before proceeding it is perhaps necessary to clarify some subtleties that arise in the present situation. It has been argued that generically there is a force between two non-BPS branes generated at loop one loop in open string theory . If so, a stack of non-BPS branes is unstable. In such a situation it is not clear that the effective field theory is appropriate. Indeed one motivation for our work is to address this question in a concrete example. In this paper we shall determine the low energy interactions of the light fields from tree-level string theory<sup>3</sup><sup>3</sup>3See also for a similar calculation of the tachyon potential for the $`D\overline{D}`$ system in type II theory.. The effective action is then simply defined to be a local quantum field theory of these light modes which reproduces the correct amplitudes. It turns out that, as in the supersymmetric case, many of the couplings such as the gauge couplings can be inferred solely from geometrical considerations and T-duality. In addition, the Yukawa couplings are restricted by general properties of the string S-matrix such as winding number conservation. In this way one can fix all couplings apart from the potential for the scalars $`\chi _i`$ originating in the tachyonic sector. These in turn will be fixed below by an explicit calculation of the four-tachyon amplitude in the presence of non-BPS branes. The potential obtained in this way shows some interesting features. The four scalar interaction ($`[\chi _i,\chi _j]^2`$) always present in supersymmetric theories arises also in the potential for $`\chi _i`$, but with the opposite sign! The corresponding instability is however cured by another $`\chi ^4`$-term that comes with exactly the right coefficient to transform this instability into a flat direction of the potential. Note however, that contrary to supersymmetric theories, these new flat directions do not lie in the commuting Cartan subalgebra. As a consequence, unlike in the supersymmetric case, at a generic point in this branch of the moduli space all scalars corresponding to the separations of the branes acquire masses. This in turn leads to a condensation of the non-BPS branes. Nevertheless, the gauge symmetry is spontaneously broken. We then discuss how quantum corrections affect these flat directions within the field theory approximation. In vacua where $`\chi _i=0`$ the one-loop correction to the potential for the scalars $`\varphi _I`$ (corresponding to separating the non-BPS branes) leads to a repulsive force, unless the orbifold is tuned in such a way that the masses of the scalars from the tachyonic sector vanish. This is in agreement with a one-loop open string calculation and is a result of a Bose-Fermi degeneracy. However we will see that this degeneracy does not persist at the level of the interacting field theory and therefore we suspect that the effective potential will not vanish at two loops at the critical radii. We also consider the one loop correction to the effective potential for the scalars $`\chi _i`$, which removes all flat directions in the corresponding branch when $`m_i=0`$. The new minimum prefers a non-vanishing, non-abelian expectation value of $`\chi _i`$. This, in turn induces a mass term for other scalars $`\varphi _I`$ resulting in an attractive force between the branes. This raises the possibility that a stack of non-BPS branes can condense even when the orbifold is not tuned precisely to criticality. This is encouraging in view of a possible existence of a gravitational background dual to the field theory on the branes. Indeed, a necessary condition for a successful description of non-Abelian YM-theory in terms of these non-BPS branes is, of course, that a stack of such branes is stable, i.e. does not fly apart. Our result for the field theory action of these branes provides sufficient information to further verify this condition at higher order. We leave this lengthy but, in principle straightforward computation for future work. The rest of this paper is organised as follows. In section two we review some basic features of non-BPS 3-branes in type IIA string theory. In section three we determine the low energy field theory on parallel branes. Lastly in section four we calculate the one-loop corrections to the effective potentials for the scalar fields. ## 2 Non-BPS 3-branes In this section we review the basic features of stable non-BPS branes in type II string theory. We will consider a non-BPS 3-brane in type IIA string theory for the sake of clarity. However, the generalisation to other branes is clear. We use a “mostly plus” metric and the conventions that indices $`m,n=0,1,2,\mathrm{},9`$ run over all of ten-dimensional space-time, $`\mu ,\nu =0,1,2,3`$ run over the 3-brane world volume, $`i,j=6,7,8,9`$ and $`I,J=4,5`$ label the transverse directions. Group indices are labeled by $`a,b`$ and we choose a hermitian basis with $`\text{Tr}(t^at^b)=\delta ^{ab}`$. We will suppress all spinor indices. As explained in the excitations of a single non-BPS D-brane are carried by two types of open strings with CP factor $`I`$ and $`\sigma _1`$ respectively. The CP $`I`$-sector is precisely the same as for a BPS D-brane and gives rise to massless $`N=4`$ vector multiplet $`A_\mu ,\varphi ^I,\chi ^i,\lambda `$ on the brane. Here $`\lambda `$ is a ten-dimensional spinor which, as a consequence of the GSO projection, is chiral (in a ten-dimensional sense). If there are $`N`$ branes then the gauge group is $`U(N)`$. The CP $`\sigma _1`$ sector has GSO-projection $`(1)^F=1`$ and hence contains a tachyonic state in its spectrum. The lightest modes are therefore a (real) tachyon $`\tau `$ with $`m^2=\frac{1}{2\alpha ^{}}`$ from the NS ground state and a massless a ten-dimensional fermion $`\psi `$ from the R ground state. The two fermions $`\lambda `$ and $`\psi `$ have opposite ten-dimensional chirality. All fields are in the adjoint of $`U(N)`$. The tachyonic instability of the non-BPS brane can be removed by compactifying the directions $`x^ix^i+R_i`$ and introducing an orbifold $`T^4/_4(1)^{F_L}`$, where $`_4:x^ix^i`$ and $`F_L`$ is the left-moving spacetime fermion number. The effect of this orbifold in the $`I`$-sector is simply to remove the scalars $`\varphi ^i`$ and project onto six-dimensional chiral fermions $`\lambda _+`$, $`\mathrm{\Gamma }^{6789}\lambda _+=\lambda _+`$. This leaves an $`N=2`$ vector multiplet in four dimensions. In the $`\sigma _1`$-sector the $`_4`$ also projects on to six-dimensional chiral fermions $`\psi _{}`$, $`\mathrm{\Gamma }^{6789}\psi _{}=\psi _{}`$. In addition $`_4`$ reverses the sign of the tachyon winding modes and the orbifold keeps only those components which are odd under $`_4`$. Thus after the orbifold the lightest field in the NS sector with CP-factor $`\sigma _1`$ originates in the ground state of strings with winding number one around the coordinate $`x_i`$. As there are four compact directions we will have eight different scalar fields $`\tau _i^\pm `$ associated with these modes, where $`\tau _i^\pm `$ is the component of the tachyon with winding number $`\pm 1`$ around $`x^i`$. In particular only the combinations $$\chi _i=\frac{\tau _i^+\tau _i^{}}{\sqrt{2}},$$ (1) survive. Their mass is given by $$m_i^2=\left(\frac{R_i}{\alpha ^{}}\right)^2\frac{1}{2\alpha ^{}},$$ (2) where $`R_i`$ is the radius of the compact direction $`x^i`$. Thus, so as long as $`R_iR_c=\sqrt{\frac{\alpha ^{}}{2}}`$, these lowest mass states are not tachyons and hence a non-BPS brane is stable. On the other hand we can tune the radii so that the $`m_i^2`$ are small compared to the string scale. Therefore these states are relevant for the low energy dynamics and should be included in the field action of the brane. To summarise, after the orbifold, the light fields on the non-BPS branes consist of an $`N=2`$ vector multiplet $`(A_\mu ,\varphi ^I,\lambda _+)`$ with gauge group $`U(N)`$ coupled to a massless six-dimensional fermion $`\psi _{}`$ and four massive scalars $`\chi _i`$ all in the adjoint representation. Note that at the critical radius we obtain the field content of an $`N=4`$ super-Yang-Mills gauge theory. However we will shortly see that the interactions of these fields are never supersymmetric. ## 3 Non-Abelian action for non-BPS 3-branes In Sen proposed a world volume action for a single non-BPS p-brane of the form $$S=T_pd^4\sigma \sqrt{|G_{ab}+B_{ab}+2\pi \alpha F_{ab}|}[\underset{i}{}(\chi )^2V(\chi )]+\mathrm{},$$ (3) where $`V(\chi )=1+m^2\chi ^2+O(\chi ^3)`$ and the ellipses denote fermionic terms. We now want to find the non-Abelian action describing a stack of non-BPS 3-branes of type IIA string theory on the orbifold $`T^4/_4(1)^{F_L}`$. For this we first recall that the fields originating in the CP-factor $`I`$ are the same as for a BPS-brane. Therefore the action for these fields is the same as for the BPS case, i.e. an $`N=2`$, $`U(N)`$ gauge theory. For the fields originating in the $`\sigma _1`$-sector some parts of the action (namely those up to $`O(\chi ^2)`$) can, in fact, be obtained from geometric considerations and T-duality alone as we shall now demonstrate. Consider for example a pair of non-BPS D0-branes, separated by a distance $`\mathrm{\Delta }X_1`$, and a $`\sigma _1`$-string starting at one $`D0`$ wrapping around the orbifold and ending on the other $`D0`$. The mass of such a string is given by $$M_i^2=\left(\frac{\mathrm{\Delta }X}{\alpha ^{}}\right)^2+\frac{N\frac{1}{2}}{\alpha ^{}},$$ (4) where $`N`$ is the oscillator number of the string state. Now, from (Fig. 1) $`(\mathrm{\Delta }X)^2=(R_i)^2+(\mathrm{\Delta }X_i)^2`$. This mass must then be reproduced within the world volume theory. A similar argument applies to the mass of $`\psi _{}`$ which is non-zero if the $`\sigma _1`$ string is stretched (with no $`\frac{1}{2}`$ in (4)). Thus, after rescaling the fields to the canonical dimensions in field theory we conclude that the effective action has the form $`S`$ $`=`$ $`\mathrm{Tr}{\displaystyle }\text{d}^4x[{\displaystyle \frac{1}{4}}F_{\mu \nu }F^{\mu \nu }+{\displaystyle \frac{1}{2}}(D_\mu \varphi _I)(D^\mu \varphi ^I){\displaystyle \frac{g_{\text{YM}}^2}{4}}([\varphi ^I,\varphi ^J])^2`$ $`ig_{\text{YM}}\overline{\lambda }_+\mathrm{\Gamma }^\mu D_\mu \lambda _++g_{\text{YM}}\overline{\lambda }_+\mathrm{\Gamma }^I[\varphi ^I,\lambda _+]`$ $`+{\displaystyle \frac{1}{2}}(D_\mu \chi _i)(D^\mu \chi ^i)+{\displaystyle \frac{1}{2}}m_i^2\chi _i^2{\displaystyle \frac{g_{\text{YM}}^2}{2}}[\varphi ^I,\chi _i][\varphi ^I,\chi ^i]`$ $`ig_{\text{YM}}\overline{\psi }_{}\mathrm{\Gamma }^\mu D_\mu \psi _{}+g_{\text{YM}}\overline{\psi }_{}\mathrm{\Gamma }^I[\varphi ^I,\psi _{}]+_{\sigma _1}],`$ where $`_{\sigma _1}`$ denotes additional terms involving fields from the $`\sigma _1`$-sector and $`D_\mu =_\mu +ig_{\text{YM}}[A_\mu ,]`$ is the gauge covariant derivative. For simplicity we continue to label the fermions in terms of ten dimensional spinors and therefore the $`\mathrm{\Gamma }`$-matrices are also ten-dimensional. In this notation $`\lambda _+`$ and $`\psi _{}`$ are constrained to have opposite ten-dimensional chiralities and the subscript $`\pm `$ refers to their chirality under $`\mathrm{\Gamma }^{012345}`$. This action reproduces the correct masses for the scalars $`\varphi _I`$ in the CP factor $`I`$ and the tachyons $`\chi _i`$ in the CP factor $`\sigma _1`$. The covariant derivative arises because the $`\sigma _1`$-sector fields are in the adjoint. This also follows via T-duality of the couplings $`[\varphi _I,\chi _i][\varphi ^I,\chi ^i]`$ and $`\overline{\psi }_{}\mathrm{\Gamma }^I[\varphi ^I,\psi _{}]`$ in (3). Alternatively, one can deduce the $`([\varphi _I,\chi _i])^2`$ and $`\overline{\psi }_{}[\varphi _I,\psi _{}]`$ couplings from T-duality and fact that $`\chi `$ and $`\psi `$ are in the adjoint. Next we determine the remaining terms $`_{\sigma _1}`$ that appear in (3). Let us first discuss the fermion-tachyon coupling. In addition to the terms $$\overline{\psi }_{}\mathrm{\Gamma }^I[\varphi _I,\psi _{}]$$ (6) in (3) one might expect, in view of the similarity of (3) with $`N=4`$ Yang-Mills, further Yukawa couplings involving the tachyons $`\chi _i`$. If both fermions originate in the same sector then this coupling is excluded as it involves a trace over an odd number of $`\sigma _1`$’s. On the other hand this does not exclude coupling of the form $$\overline{\lambda }_+\mathrm{\Gamma }^k[\chi _k,\psi _{}].$$ (7) Indeed precisely such a coupling occurs in an $`N=4`$ super-Yang-Mills action. However, this coupling is excluded in the field theory limit because, due to winding number conservation, one of the fermions would have to have a non-zero winding number and therefore a mass of $`O(\alpha ^{})`$. The absence of these terms is a first indication that (3) is never supersymmetric in contrast to the effective action for BPS D3-branes. The only remaining undetermined term in the effective action is the potential for the scalars originating in the tachyonic NS sector $$V(\chi _i)=\frac{1}{2}m_i^2\chi _i^2_{\sigma _1}.$$ (8) Since have already determined the mass term in $`V(\chi _i)`$ let us concentrate on the terms $`O(\chi ^3)`$ and $`O(\chi ^4)`$. To determine the exact form of these terms one can analyse disk three-and four point functions respectively. In fact the three-point function must vanish, as can be seen in a number ways. One way is to note that the trace over an odd number of $`\sigma _1`$ vanishes. In addition winding number conservation implies that processes involving an odd number of $`\chi _i`$ fields must vanish. This leads to a $`𝐙_2^4`$ symmetry of the effective theory generated by $`\chi _i\chi _i`$. Thus we are left with the four-point function $`S_D(k_1,\mathrm{},k_4)`$. The relevant graphs for this are given in Fig. 2. The rest of this section is devoted to calculating the corresponding amplitude and deducing the form of the four-tachyon term in the effective action. If we denote by $`k`$ the “field theory” momentum (i.e. momentum tangent to the brane) then the full ten-dimensional momentum $`K`$ of the i-th tachyon winding mode is $$K=k+\frac{1}{\alpha ^{}}\stackrel{}{w}_i.$$ (9) Here $`\stackrel{}{w}_i`$ is the winding vector of the tachyon around the compact directions $`x^i`$. The lowest mass states $`\tau _i^\pm `$ left after the orbifold have $`\stackrel{}{w_6}=(\pm R_6,0,0,0)`$, $`\stackrel{}{w_7}=(0,\pm R_7,0,0)`$, $`\stackrel{}{w_8}=(0,0,\pm R_8,0)`$ and $`\stackrel{}{w_9}=(0,0,0,\pm R_9)`$. To fix the gauge we need to fix three of the bosonic coordinates and two of the fermionic coordinates in this correlator. The fixed (-1 picture) and the integrated (0-picture) vertex operators are $$V_\tau ^{(0)}=i2\alpha ^{}gc(x)\left(K\psi (x)\right)e^{iKX(x)}\text{and}V_\tau ^{(1)}=i\sqrt{2\alpha ^{}}gc(x)e^{\varphi (x)}e^{iKX(x)},$$ (10) respectively. The corresponding amplitude is found to be $`S_D(k_1,\mathrm{},k_4)`$ $`=`$ $`4\alpha ^{}g^2(2\pi )^4\delta _{{\scriptscriptstyle \stackrel{}{w}_i}}\delta ^4({\displaystyle \underset{i=1}{\overset{4}{}}}k_i)\text{Tr}_f[t^{a_1},\mathrm{},t^{a_4}]`$ $`{\displaystyle \frac{x_{12}x_{23}x_{13}}{x_{12}x_{34}}}K_3K_4{\displaystyle \underset{\mathrm{}}{\overset{\mathrm{}}{}}}\text{d}x_4{\displaystyle \underset{i<j}{}}|x_{ij}|^{2\alpha ^{}K_iK_j}+(K_2K_3),`$ where $`\delta _{{\scriptscriptstyle \stackrel{}{w}_i}}`$ ensures the winding number conservation. In terms of the ten-dimensional Mandelstam variables $$S=(K_1+K_2)^2\text{,}T=(K_1+K_3)^2\text{,}U=(K_1+K_4)^2,$$ (12) with $$S+T+U=2/\alpha ^{},$$ (13) the amplitude takes the form $`S_D(k_1,\mathrm{},k_4)`$ $`=`$ $`2g^2(2\pi )^4{\displaystyle \underset{p=1}{\overset{4}{}}}\delta _{{\scriptscriptstyle \stackrel{}{w_i}}}\delta ^4({\displaystyle \underset{i=1}{\overset{4}{}}}k_i)`$ $`[(1+\alpha ^{}U)\text{Tr}(t^{a_3}t^{a_4}t^{a_2}t^{a_1}+t^{a_2}t^{a_4}t^{a_3}t^{a_1})B(\alpha ^{}S,\alpha ^{}T)`$ $`+(1+\alpha ^{}T)\text{Tr}(t^{a_2}t^{a_3}t^{a_4}t^{a_1}+t^{a_3}t^{a_2}t^{a_1}t^{a_4})B(\alpha ^{}S,\alpha ^{}U)`$ $`+(1+\alpha ^{}S)\text{Tr}(t^{a_3}t^{a_2}t^{a_4}t^{a_1}+t^{a_2}t^{a_3}t^{a_1}t^{a_4})B(\alpha ^{}T,\alpha ^{}U)],`$ where $`B(a,b)`$ is the Euler beta function. The amplitude (3) has no pole at $`\alpha ^{}S=\frac{1}{2}`$. In the bosonic string this pole is related to the tachyonic intermediary state. Its absence is the consistent with the absence of a three tachyon coupling, as explained above. After orbifolding the full four-tachyon amplitude is the given by the sum over all diagrams in figure 2. Winding number conservation implies that the external states must be of the form $`\tau _i^+,\tau _i^{},\tau _j^+,\tau _j^{}`$. There are two cases to consider: either $`ij`$ or $`i=j`$. Let us first consider diagrams of Fig. 2 with $`ij`$. Here we find that there are four separate graphs which contribute to $`\chi _i,\chi _i,\chi _j,\chi _j`$ scattering. It is helpful here to introduce the Mandelstam variables for the field theory momenta $$s=(k_1+k_2)^2,t=(k_1+k_3)^2,u=(k_1+k_4)^2.$$ (15) For the four graphs of interest we find that $$S=1/\alpha ^{}+s(m_i^2+m_j^2),T=1/\alpha ^{}+t(m_i^2+m_j^2),U=u.$$ (16) We may now expand the amplitude (3) to lowest order in $`\alpha ^{}`$, noting that $`s,t`$ and $`u`$ are all $`O(1)`$. In this way we find $`S_D^{ij}(k_1,\mathrm{},k_4)`$ $`=`$ $`g^2(2\pi )^4\delta ^4({\displaystyle \underset{i=1}{\overset{4}{}}}k_i)(\text{Tr}({\displaystyle \frac{1}{2}}[t^{a_2},t^{a_3}][t^{a_1},t^{a_4}]+\{t^{a_2},t^{a_4}\}\{t^{a_1},t^{a_3}\}`$ (17) $`+\left({\displaystyle \frac{ts}{2u}}\right)\text{Tr}([t^{a_2},t^{a_3}][t^{a_1},t^{a_4}])).`$ The pole corresponds to an exchange of a gauge boson in the u-channel. Note that, for $`ij`$, the incoming states can only annihilate to a zero winding number state such as a gauge boson in the u-channel. Let us now consider diagrams in Fig. 2 where $`i=j`$. Winding number conservation now implies that there are six graphs that must be summed over. These graphs come in three pairs with $$S=s,T=t,U=2/\alpha ^{}+u4m_i^2,$$ (18) and similarly for the other two pairs. If we now sum of all these contributions we find $`S_D^{ii}(k_1,\mathrm{},k_4)`$ $`=`$ $`g^2(2\pi )^4\delta ^4({\displaystyle \underset{i=1}{\overset{4}{}}}k_i)(\left({\displaystyle \frac{ts}{u}}\right)\text{Tr}([t^{a_2},t^{a_3}][t^{a_1},t^{a_4}])`$ (19) $`+\left({\displaystyle \frac{su}{t}}\right)\text{Tr}([t^{a_1},t^{a_3}][t^{a_2},t^{a_4}])`$ $`+\left({\displaystyle \frac{tu}{s}}\right)\text{Tr}([t^{a_1},t^{a_2}][t^{a_3},t^{a_4}])).`$ Thus here there are only pole contributions corresponding to the exchange of a gauge boson. Note that for $`i=j`$ it is possible for the incoming states to annihilate into a zero winding state in any of the three channels. Comparing (17) and (19) with the tree-level field theory amplitude we find after a lengthy but straightforward computation that the string tree-level amplitude is reproduced by the following potential for the scalars $`\chi `$ originating in the tachyonic sector $$V(\chi )=\text{Tr}\left(\frac{1}{2}\underset{i}{}m_i^2\chi _i^2+\frac{g_{\text{YM}}^2}{4}\underset{ij}{}([\chi _i,\chi _j])^2+g_{\text{YM}}^2\underset{ij}{}(\chi _i)^2(\chi _j)^2\right).$$ (20) This is the main result of this paper. Before discussing its implications a comment about the uniqueness of the effective potential (20) is in order. As is well known the effective action for tachyons is ambiguous due to the possibility of replacing $`m^2T`$ by $`^2T`$. In the present situation this ambiguity can in principle arise for the $`\chi ^4`$ term and should be kept in mind for the cases when $`m_i0`$. Let us briefly describe the classical vacuum moduli space of the theory. Of course without supersymmetry we do not expect any of these vacuum moduli to persist in the quantum theory and hence we must interpret the term moduli space loosely. We note that the complete potential for all the fields can be written as $$\text{Tr}\left[\frac{1}{2}g_{YM}^2([\varphi _4,\varphi _5])^2\frac{1}{2}g_{YM}^2\underset{I,i}{}([\varphi _I,\chi _i])^2+\frac{1}{2}\underset{i}{}m_i^2\chi _i^2+\frac{g_{YM}^2}{4}\underset{ij}{}(\{\chi _i,\chi _j\})^2\right].$$ (21) Clearly if $`m_i^2<0`$ for some $`i`$, say $`i=1`$, then the potential is not bounded from below because we may take only $`\chi _1`$ to be non-vanishing and thereby make the potential as negative as we wish. On the other hand if $`m_i^20`$ for all $`i`$ then since we have chosen a hermitian basis for the Lie algebra it is not hard to see that all the terms appearing in (21) are positive. Thus the potential is bounded below by zero. For generic values of the orbifold radii the potential is minimized by $`\chi _i=0`$ and $`[\varphi _4,\varphi _5]=0`$. In this case the moduli space of vacua are given by the Cartan subalgebra of $`U(N)`$. This corresponds to a Coulomb branch with $`N`$ massless $`U(1)`$ gauge fields. If $`m_i^2=0`$ then there are additional branches of the vacuum moduli space that can arise where $`\varphi _I=0`$. We are then left with the requirement $`\{\chi _i,\chi _j\}=0`$ for $`ij`$, $`m_i^2=m_j^2=0`$ (and $`\chi _i=0`$ if $`m_i^2>0`$), i.e. this vacuum is parameterised by anti-commuting scalar vevs. Generically in these branches the gauge group is completely higgsed leaving only the massless fields from the $`U(1)`$ corresponding to the overall translation invariance. Note that here the scalars $`\varphi _I`$ become massive so that the non BPS-branes are bound together. In addition the $`𝐙_2^4`$ symmetry of the theory is spontaneously broken. ## 4 One-Loop Effective Potential In this section we compute within the the field theory approximation the one-loop corrections to the tree level potentials obtained in the previous section. For simplicity we take all the radii to be equal and assume that there are only two non-BPS branes. It has been shown using open string theory that there is a one-loop repulsive force between the two non-BPS branes which vanishes at the critical radius $`R=R_c`$. We will reproduce this force in the field theory approximation where it corresponds to the lifting of the Coulomb branch by a one-loop effective potential for the scalars $`\varphi _I`$, when the $`\sigma _1`$-scalars $`\chi _i`$ are set to zero. We will also discuss the lifting of the Higgs branch discussed above when $`m=0`$ and $`\varphi _I=0`$. As we shall see, the one-loop quantum corrections remove the $`\chi _i=0`$ vacuum, indicating that the branes can condense. Let us first describe the lifting of the Coulomb branch resulting in a force between the branes when $`R_i>R_c`$ and the scalars from the $`\sigma _1`$-sector are massive. A quick way to compute the potential is to first consider the massless limit. In this case the field content becomes the same as that of $`N=4`$ super-Yang-Mills. In addition the interaction terms which break supersymmetry do not contribute to the $`\varphi _I`$ effective potential at one loop. Therefore the effective potential is the same as for an $`N=4`$ super Yang-Mills theory and hence vanishes on account of the Bose-Fermi degeneracy. Then, because the mass term only appears in the fluctuation determinant of the tachyon<sup>4</sup><sup>4</sup>4We choose an $`N=2`$ supersymmetric version of the $`R_\xi `$-gauge , the effective potential at non-zero $`m`$ is simply given by $$V_{eff}(X_1)=2\left[\mathrm{log}det(M_m)\mathrm{log}det(M_0)\right],$$ (22) where $`_m=\frac{1}{2}g_{ab}\delta _{ij}(^2+m_i^2)`$ is the fluctuation operator for the scalars in the $`\sigma _1`$-sector. We use the $`\zeta `$-function definition of the regularised functional determinants. The one-loop effective potential is then found to be $`(\varphi _I^a\delta ^{a3}\delta _{I5}F+\varphi _I^a)`$ $`2\pi ^2V_{eff}(F)`$ $`=`$ $`(m^2+4g_{\text{YM}}^2F^2)^2\left[{\displaystyle \frac{3}{2}}\mathrm{ln}\left({\displaystyle \frac{m^2+4g_{\text{YM}}^2F^2}{\mathrm{\Lambda }^2}}\right)\right]`$ $`+(4g_{\text{YM}}^2F^2)^2\left[{\displaystyle \frac{3}{2}}\mathrm{ln}\left({\displaystyle \frac{4g_{\text{YM}}^2F^2}{\mathrm{\Lambda }^2}}\right)\right]m^4\left[{\displaystyle \frac{3}{2}}\mathrm{ln}\left({\displaystyle \frac{m^2}{\mathrm{\Lambda }^2}}\right)\right].`$ Of physical interest to us is the force between to branes $`{\displaystyle \frac{V_{eff}(F)}{F}}`$ $`=`$ $`{\displaystyle \frac{32g_{\text{YM}}^4}{\pi ^2}}F^3\{\left({\displaystyle \frac{m^2}{4g_{\text{YM}}^2F^2}}\right)[1\mathrm{log}\left({\displaystyle \frac{m^2+4g_{\text{YM}}^2F^2}{\mathrm{\Lambda }^2}}\right)]`$ (24) $`\mathrm{log}\left({\displaystyle \frac{m^2+4g_{\text{YM}}^2F^2}{4g_{\text{YM}}^2F^2}}\right)\},`$ where $`\mathrm{\Lambda }`$ is the UV cut-off which should be set to the characteristic string scale ($`\alpha ^{}`$) in the present situation . Here we take the one-loop mass of $`m_\varphi `$ to vanish at<sup>5</sup><sup>5</sup>5for $`\mu \mathrm{\Lambda }`$ the effective mass will then be given by $`m_\varphi ^2=\frac{g_{\text{YM}}^2m^2}{\pi ^2}\mathrm{log}(\mu /\mathrm{\Lambda })`$. $`\mu =\mathrm{\Lambda }`$. From (4) $$4g_{\text{YM}}^2F^2=\left(\frac{\mathrm{\Delta }X^1}{2\pi \alpha ^{}}\right)^2\left(\frac{X}{\alpha ^{}}\right)^2,$$ (25) so that $`{\displaystyle \frac{V_{eff}(X)}{X}}`$ $`=`$ $`A\left[r\left(c\mathrm{log}(1+r^2)\right)r^3\mathrm{log}\left({\displaystyle \frac{1+r^2}{r^2}}\right)\right],`$ where $$r=\frac{X}{m\alpha ^{}}\text{,}A=\frac{2m^3}{\alpha ^{}\pi ^2}\text{and}c=1\mathrm{log}(m^2\alpha ^{}).$$ (26) Note that in the field theory approximation we have $`X^2<<\alpha ^{}`$ and $`m^2\alpha ^{}<<1`$ always. Nevertheless we can distinguish between $`r<<1`$ and $`r>>1`$. In the first case the force between two non-BPS branes is approximated by $$\frac{V_{eff}(X)}{X}Acr,$$ (27) whereas in the second case the leading behaviour is $$\frac{V_{eff}(X)}{X}Ar\mathrm{log}\left(\frac{X^2}{\alpha ^{}}\right).$$ (28) In particular the force is repulsive in both limits. This is in agreement with the string theory result . Of course the effective potential is bounded from below however the minimum is outside the validity of our one-loop and field theory approximations. Nevertheless one does expect a stable minimum to arise in the field theory. Finally let us turn to the effective potential for the $`\sigma _1`$-scalars $`\chi _i`$. Due to the absence of the corresponding Yukawa couplings in (20) only scalar and vector loops contribute to the effective potential. For simplicity, rather than a general background we consider the ansatz $$\chi _6=\frac{v_1}{2g_{\text{YM}}}\sigma _1,\chi _7=\frac{v_2}{2g_{\text{YM}}}\sigma _2\text{and}\chi _8=\chi _9=0,$$ (29) where $`\sigma _1,\sigma _2`$ and $`\sigma _3`$ are Pauli matrices. The corresponding one-loop effective potential is given by $`16\pi ^2V(v_1,v_2)`$ $`=`$ $`6(v_1^2+v_2^2)^2\left[{\displaystyle \frac{3}{2}}\mathrm{log}\left({\displaystyle \frac{v_1^2+v_2^2}{\mathrm{\Lambda }^2}}\right)\right]`$ $`8v_1^4\left[{\displaystyle \frac{3}{2}}\mathrm{log}\left({\displaystyle \frac{v_1^2}{\mathrm{\Lambda }^2}}\right)\right]8v_2^4\left[{\displaystyle \frac{3}{2}}\mathrm{log}\left({\displaystyle \frac{v_2^2}{\mathrm{\Lambda }^2}}\right)\right].`$ The minimum of this potential is at $`v_1^2=v_2^2=O(\mathrm{\Lambda }^2)`$. Therefore it favours a non-abelian expectation value for $`\chi `$. This, in turn, induces a mass term for $`\varphi _I`$ and hence an attraction between the non-BPS branes at tree-level in $`\varphi _I`$. The loop correction in the $`\varphi `$-channel in a non-vanishing $`\chi `$-background will be repulsive but this is should be of sub-leading order in $`g_{\text{YM}}`$. Of course, this result is to be taken with a grain of salt as the predicted expectation value of $`\chi _i`$ is beyond the range of validity of both the one-loop approximation in field theory and furthermore the field theory approximation altogether ($`\mathrm{\Lambda }=(\alpha ^{})^{\frac{1}{2}}`$). A more reliable result should come from a two-loop computation in the field theory or, better still, a string loop computation in a $`\chi `$-background. We leave this challenge for future work and highlight here the possibility of brane condensation due to the existence of a non-trivial, non-abelian minima in the potential for $`\chi _i`$. Acknowledgments We would like to thank Peter West for discussions. I.S. would like to thank Kings College London, where this project was started, for hospitality. We are also indebted to the referee for pointing out a mistake in our review of non-BPS branes. N.D.L. was supported by the EU grant ERBFMRX-CT96-0012 and I.S. was supported in parts by Swiss Government TMR Grant, BBW Nr. 970557.
warning/0002/hep-ph0002216.html
ar5iv
text
# Long-distance effects in 𝐵→𝑉⁢𝛾 radiative weak decays ## I Introduction Rare radiative decays of $`B`$ mesons received considerable theoretical attention due to their special sensitivity to physics beyond the Standard Model. Alternatively, if the validity of the SM is taken for granted, these decays can offer useful information on the magnitude of the CKM parameters. So far, most of the theoretical effort has been concentrated on aspects of inclusive decays $`bs\gamma `$, which can be treated with the help of an operator product expansion (OPE) combined with the heavy quark expansion . The corresponding strong interaction effects can be quantified in terms of a few nonperturbative matrix elements. Experimental results on the branching ratio for inclusive $`BX_s\gamma `$ decays appear to be in agreement with the predictions of the Standard Model. On the other hand, the treatment of exclusive radiative decays such as $`BK^{}\gamma `$ or $`B\rho \gamma `$ is considerably more difficult. This is due to the fact that bound state effects must be taken into account, which are essentially nonperturbative in nature. In addition to the matrix element of the penguin operator (short-distance component), a nonlocal contribution must be included too, arising from a product of the electromagnetic coupling with the weak nonleptonic Hamiltonian (long-distance component). These effects have been estimated using various methods, such as perturbative QCD combined with the quark model , dispersion methods , vector meson dominance and QCD sum rules . Recently the CLEO collaboration measured the branching ratios for several exclusive radiative decays $`BV\gamma `$, with $`V=K^0,K^+`$ , and gave upper limits on the $`V=\rho ,\omega `$ modes. A more detailed study of these decays is therefore well motivated. In this paper we propose a different approach to the calculation of the dominant long distance contributions to $`BV\gamma `$ decays (with $`V`$ a member of the SU(3) octet of vector mesons), arising from weak annihilation and $`W`$-exchange topologies. We start by parametrizing all these decays in terms of a few common amplitudes (for each photon helicity $`\lambda =L,R`$), using SU(3) symmetry. It can be shown that the structure of these universal amplitudes simplifies very much to the leading order of a twist expansion, in powers of $`\mathrm{\Lambda }/E_\gamma `$ and $`\mathrm{\Lambda }/p_+=\mathrm{\Lambda }/m_B`$, with $`p_+`$ the light-cone component of the vector meson momentum. Since the photon energy in $`B\rho (K^{})\gamma `$ decays is $`E_\gamma =2.6`$ GeV, the higher twist corrections can be expected to be well under control. We find that, to leading twist, the photons emitted in $`\overline{B}`$ decay are predominantly left-handed. Second, the nonfactorizable corrections turn out to be computable in terms of the light-cone wavefunctions of the $`B`$ and $`V`$ mesons. Furthermore, in the chiral limit, the nonfactorizable corrections vanish exactly to one-loop order. The remaining factorizable contribution can be determined in a model-independent way using experimental data on radiative leptonic $`B`$ decays $`B\gamma e\nu `$. These results can be used to assess the feasibility of the determination of the ratio $`|V_{td}/V_{ts}|`$ from a comparison of the $`BK^{}\gamma `$ and $`B\rho \gamma `$ decays. The Appendix describes certain constraints on radiative decay form-factors following from a Ward identity expressing the conservation of the electromagnetic current. ## II Flavor SU(3) predictions for $`\overline{B}V\gamma `$ decay amplitudes In the Standard Model, the radiative $`\overline{B}V\gamma `$ decays (with $`V`$ a member of the SU(3) octet of vector mesons) are dominated by the operators ($`q=s,d`$) (see, e.g., ) $`_{rad}`$ $`=`$ $`{\displaystyle \frac{4G_F}{\sqrt{2}}}\lambda _t^{(s)}{\displaystyle \underset{i=7,8}{}}C_i(\mu )𝒪_i^{(s)}(\mu )+(sd),`$ (1) $`𝒪_7^{(q)}`$ $`=`$ $`{\displaystyle \frac{e}{16\pi ^2}}F^{\mu \nu }\overline{q}\sigma _{\mu \nu }(m_bP_R+m_qP_L)b,`$ (2) $`𝒪_8^{(q)}`$ $`=`$ $`{\displaystyle \frac{g_s}{16\pi ^2}}\overline{q}\sigma _{\mu \nu }(m_bP_R+m_qP_L)G^{\mu \nu }b,P_{R,L}={\displaystyle \frac{1}{2}}(1\pm \gamma _5)`$ (3) We denote here and in the following the combination of CKM matrix elements $`\lambda _q^{(q^{})}=V_{qb}V_{qq^{}}^{}`$. In addition to this, there are also contributions mediated by the usual weak nonleptonic Hamiltonian $`_W`$ $`=`$ $`{\displaystyle \frac{4G_F}{\sqrt{2}}}\{\lambda _u^{(s)}[C_1(\overline{u}b)_{VA}(\overline{s}u)_{VA}+C_2(\overline{s}b)_{VA}(\overline{u}u)_{VA}]`$ (4) $`+`$ $`\lambda _c^{(s)}[C_1(\overline{c}b)_{VA}(\overline{s}c)_{VA}+C_2(\overline{s}b)_{VA}(\overline{c}c)_{VA}]\lambda _t^{(s)}{\displaystyle \underset{i=3}{\overset{10}{}}}c_iQ_i^{(s)}+(sd)\}`$ (5) with $`(\overline{q}_1q_2)_{VA}(\overline{q}_3q_4)_{VA}=[\overline{q}_1\gamma _\mu P_Lq_2][\overline{q}_3\gamma ^\mu P_Lq_4]`$. In these diagrams, the photon attaches to the internal quark lines with the usual electromagnetic coupling. We will neglect in the following the contributions of the penguin operators $`Q_{310}`$, which are suppressed relative to those of the current-current operators by their smaller Wilson coefficients. In a quark diagram language, all $`BV\gamma _\lambda `$ decay amplitudes are parametrized by nine independent $`SU(3)`$ amplitudes, for each individual photon helicity $`\lambda =L,R`$. They appear together in fewer combinations, such that a certain predictive power is still retained. They include penguin-type amplitudes $`P_{u,c,t}`$ – corresponding to topologies with a $`u`$-, $`c`$\- and $`t`$-quark running in the loop, the weak annihilation $`(WA)`$-type amplitude $`A`$, the $`W`$-exchange amplitude $`E`$, penguin-annihilation amplitudes $`PA`$, and amplitudes with one insertion of the gluonic penguin operator $`𝒪_8`$ which will be denoted $`M_i`$ (see Fig. 1). In the penguin $`P_{u,c}^{(i)}`$ and $`M^{(i)}`$ amplitudes we distinguish between diagrams with the photon coupling to the loop quark or the other emerging light quark $`(i=1)`$, and diagrams with the photon coupling to the spectator quark $`(i=2)`$. We adopt the usual phase conventions for the vector states $`(\rho ^+,\rho ^0,\rho ^{})=(u\overline{d},\frac{1}{\sqrt{2}}(d\overline{d}u\overline{u}),d\overline{u})`$, $`(\overline{K}^0,K^{})=(s\overline{d},s\overline{u})`$, $`(K^+,K^0)=(u\overline{s},d\overline{s})`$, $`\varphi ^{(8)}=\frac{1}{\sqrt{6}}(2s\overline{s}u\overline{u}d\overline{d})`$, and for the heavy mesons $`(B^+,B^0,B_s)=(\overline{b}u,\overline{b}d,\overline{b}s)`$, $`(B^{},\overline{B}^0,\overline{B}_s)=(b\overline{u},b\overline{d},b\overline{s})`$. The weak annihilation amplitudes $`A^{(i)}`$ contribute only to the $`B^\pm `$ radiative decays, where they appear always in the same combination $`A\frac{2}{3}A^{(1)}+\frac{2}{3}A^{(2)}\frac{1}{3}A^{(3)}`$ (see Fig. 1(a)). Writing explicitly the contributions of the penguin amplitudes one has $`A(B^{}\rho ^{}\gamma _\lambda )=\lambda _u^{(d)}(P_{u\lambda }^{(1)}+Q_uP_{u\lambda }^{(2)}+A_\lambda )+\lambda _c^{(d)}(P_{c\lambda }^{(1)}+Q_uP_{c\lambda }^{(2)})`$ (6) $`+\lambda _t^{(d)}(P_{t\lambda }+M_\lambda ^{(1)}+Q_uM_\lambda ^{(2)})`$ (7) $`A(B^{}K^{}\gamma _\lambda )=\lambda _u^{(s)}(P_{u\lambda }^{(1)}+Q_uP_{u\lambda }^{(2)}+A_\lambda )+\lambda _c^{(s)}(P_{c\lambda }^{(1)}+Q_uP_{c\lambda }^{(2)})`$ (8) $`+\lambda _t^{(s)}(P_{t\lambda }+M_\lambda ^{(1)}+Q_uM_\lambda ^{(2)}).`$ (9) The penguin amplitude with an internal $`t`$-quark $`P_{t\lambda }`$ (arising from the penguin operator $`𝒪_7`$) is usually called the short-distance contribution, in contrast to the long-distance amplitude, arising from nonlocal insertions of the weak Hamiltonian (4) with one photon attachment. In a hadronic language, the penguin amplitudes $`P_u^{(i)}`$ receive contributions from rescattering effects of the form $`B^\pm \{\rho ^\pm \rho ^0\}\rho ^\pm \gamma `$, and the charmed penguins $`P_c^{(i)}`$ arise from rescattering processes $`B^\pm \{D^\pm \overline{D}^0\},\{\psi \rho ^0\}\rho ^\pm \gamma `$. The $`\overline{B}^0`$ decay amplitudes contain the $`W`$-exchange topology, parametrized by the graphical amplitudes $`E^{(i)}`$ (Fig. 1(b)). All the decays discussed below contain these amplitudes in the same combination $`E\frac{2}{3}E^{(2)}+\frac{2}{3}E^{(3)}\frac{1}{3}E^{(1)}`$ $`A(\overline{B}^0\rho ^0\gamma _\lambda )={\displaystyle \frac{1}{\sqrt{2}}}\lambda _u^{(d)}(P_{u\lambda }^{(1)}+Q_dP_{u\lambda }^{(2)}E_\lambda (Q_uQ_d)PA_{u\lambda })`$ (10) $`+{\displaystyle \frac{1}{\sqrt{2}}}\lambda _c^{(d)}(P_{c\lambda }^{(1)}+Q_dP_{c\lambda }^{(2)}(Q_uQ_d)PA_{c\lambda })+{\displaystyle \frac{1}{\sqrt{2}}}\lambda _t^{(d)}(P_{t\lambda }+M_\lambda ^{(1)}+Q_dM_\lambda ^{(2)})`$ (11) $`A(\overline{B}^0\overline{K}^0\gamma _\lambda )=\lambda _u^{(s)}(P_{u\lambda }^{(1)}+Q_dP_{u\lambda }^{(2)})+\lambda _c^{(s)}(P_{c\lambda }^{(1)}+Q_dP_{c\lambda }^{(2)})+\lambda _t^{(s)}(P_{t\lambda }+M_\lambda ^{(1)}+Q_dM_\lambda ^{(2)})`$ (12) $`A(\overline{B}^0\varphi ^{(8)}\gamma _\lambda )={\displaystyle \frac{1}{\sqrt{6}}}\lambda _u^{(d)}(P_{u\lambda }^{(1)}+Q_dP_{u\lambda }^{(2)}+E_\lambda +(Q_u+Q_d2Q_s)PA_{u\lambda })`$ (13) $`{\displaystyle \frac{1}{\sqrt{6}}}\lambda _c^{(d)}(P_{c\lambda }^{(1)}+Q_dP_{c\lambda }^{(2)}+(Q_u+Q_d2Q_s)PA_{c\lambda }){\displaystyle \frac{1}{\sqrt{6}}}\lambda _t^{(d)}(P_{t\lambda }+M_\lambda ^{(1)}+Q_dM_\lambda ^{(2)}).`$ (14) Finally, the corresponding $`\overline{B}_s`$ decays are written in terms of the amplitudes previously introduced as $`A(\overline{B}_sK^0\gamma _\lambda )=\lambda _u^{(d)}(P_{u\lambda }^{(1)}+Q_sP_{u\lambda }^{(2)})\lambda _c^{(d)}(P_{c\lambda }^{(1)}+Q_sP_{c\lambda }^{(2)})`$ (15) $`\lambda _t^{(d)}(P_{t\lambda }+M_\lambda ^{(1)}+Q_sM_\lambda ^{(2)})`$ (16) $`A(\overline{B}_s\rho ^0\gamma _\lambda )={\displaystyle \frac{1}{\sqrt{2}}}\lambda _u^{(s)}(E_\lambda +(Q_uQ_d)PA_{u\lambda })+{\displaystyle \frac{1}{\sqrt{2}}}\lambda _c^{(s)}(Q_uQ_d)PA_{c\lambda }`$ (17) $`A(\overline{B}_s\varphi ^{(8)}\gamma _\lambda )={\displaystyle \frac{1}{\sqrt{6}}}\lambda _u^{(s)}\left(2P_{u\lambda }^{(1)}+2Q_sP_{u\lambda }^{(2)}E_\lambda (Q_u+Q_d2Q_s)PA_{u\lambda }\right)`$ (18) $`{\displaystyle \frac{1}{\sqrt{6}}}\lambda _c^{(s)}(2P_{c\lambda }^{(1)}+2Q_sP_{c\lambda }^{(2)}(Q_u+Q_d2Q_s)PA_{c\lambda })\sqrt{{\displaystyle \frac{2}{3}}}\lambda _t^{(s)}(P_{t\lambda }+M_\lambda ^{(1)}+Q_sM_\lambda ^{(2)}).`$ (19) Note that the long-distance penguin-type amplitudes with an internal $`u`$ and $`c`$-quark are different in $`B^\pm `$ and $`B^0,B_s`$ decays, due to the different electric charge of the spectator quark in the two cases. This is different from the conclusion reached in ; present experimental data from CLEO (see Eqs. (75), (76) below) appear to confirm the presence of these long-distance contributions at the $`1\sigma `$ level. Decays to the SU(3) singlet vector meson $`\varphi ^{(1)}=\frac{1}{\sqrt{3}}(u\overline{u}+d\overline{d}+s\overline{s})`$ introduce new amplitudes $`S_\lambda `$, which arise from diagrams similar to $`PA_\lambda `$ (see Fig. 1(d)) but with the photon attaching to the quark in the loop or to the spectator in the $`\overline{B}`$ meson $`A(\overline{B}^0\varphi ^{(1)}\gamma _\lambda )`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{3}}}\lambda _u^{(d)}(P_{u\lambda }^{(1)}+Q_dP_{u\lambda }^{(2)}+E_\lambda +S_{u\lambda })`$ (21) $`{\displaystyle \frac{1}{\sqrt{3}}}\lambda _c^{(d)}(P_{c\lambda }^{(1)}+Q_dP_{c\lambda }^{(2)}+S_{c\lambda }){\displaystyle \frac{1}{\sqrt{3}}}\lambda _t^{(d)}(P_{t\lambda }^{(1)}+M_\lambda ^{(1)}+Q_dM_\lambda ^{(2)})`$ $`A(\overline{B}_s\varphi ^{(1)}\gamma _\lambda )`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{3}}}\lambda _u^{(s)}(P_{u\lambda }^{(1)}+Q_dP_{u\lambda }^{(2)}+E_\lambda +S_{u\lambda })`$ (23) $`+{\displaystyle \frac{1}{\sqrt{3}}}\lambda _c^{(s)}(P_{c\lambda }^{(1)}+Q_dP_{c\lambda }^{(2)}+S_{c\lambda })+{\displaystyle \frac{1}{\sqrt{3}}}\lambda _t^{(s)}(P_{t\lambda }^{(1)}+M_\lambda ^{(1)}+Q_dM_\lambda ^{(2)}).`$ The decay amplitudes into the physical states $`\varphi `$, $`\omega `$ are obtained by combining the octet and singlet amplitudes with a mixing angle $`\theta _V`$ as $`\varphi `$ $`=`$ $`\varphi ^{(8)}\mathrm{cos}\theta _V\varphi ^{(1)}\mathrm{sin}\theta _V`$ (24) $`\omega `$ $`=`$ $`\varphi ^{(8)}\mathrm{sin}\theta _V+\varphi ^{(1)}\mathrm{cos}\theta _V.`$ (25) The physical value of the mixing angle is close to $`\mathrm{tan}\theta _V=\frac{1}{\sqrt{2}}`$, corresponding to ideal mixing $`\varphi =\overline{s}s`$, $`\omega =\frac{1}{\sqrt{2}}(u\overline{u}+d\overline{d})`$. Combining the amplitudes for decays into SU(3) eigenstates one finds the following amplitudes for decay into physical states $`\omega ,\varphi `$ $`A(\overline{B}^0\omega \gamma _\lambda )`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}\lambda _u^{(d)}(P_{u\lambda }^{(1)}+Q_dP_{u\lambda }^{(2)}+E_\lambda +{\displaystyle \frac{1}{3}}PA_{u\lambda }+{\displaystyle \frac{2}{3}}S_{u\lambda })`$ (27) $`{\displaystyle \frac{1}{\sqrt{2}}}\lambda _c^{(d)}(P_{c\lambda }^{(1)}+Q_dP_{c\lambda }^{(2)}+{\displaystyle \frac{1}{3}}PA_{c\lambda }+{\displaystyle \frac{2}{3}}S_{c\lambda }){\displaystyle \frac{1}{\sqrt{2}}}\lambda _t^{(d)}(P_{t\lambda }^{(1)}+M_\lambda ^{(1)}+Q_dM_\lambda ^{(2)})`$ $`A(\overline{B}^0\varphi \gamma _\lambda )`$ $`=`$ $`{\displaystyle \frac{1}{3}}\lambda _u^{(d)}(PA_{u\lambda }+S_{u\lambda })+{\displaystyle \frac{1}{3}}\lambda _c^{(d)}(PA_{c\lambda }+S_{c\lambda }),`$ (28) and $`A(\overline{B}_s\omega \gamma _\lambda )`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}\lambda _u^{(s)}(E_\lambda +{\displaystyle \frac{1}{3}}PA_{u\lambda }+{\displaystyle \frac{2}{3}}S_{u\lambda })+{\displaystyle \frac{1}{\sqrt{2}}}\lambda _c^{(s)}({\displaystyle \frac{1}{3}}PA_{c\lambda }+{\displaystyle \frac{2}{3}}S_{c\lambda })`$ (29) $`A(\overline{B}_s\varphi \gamma _\lambda )`$ $`=`$ $`\lambda _u^{(s)}(P_{u\lambda }^{(1)}+Q_sP_{u\lambda }^{(2)}{\displaystyle \frac{1}{3}}PA_{u\lambda }+{\displaystyle \frac{1}{3}}S_{u\lambda })`$ (31) $`\lambda _c^{(s)}(P_{c\lambda }^{(1)}+Q_sP_{c\lambda }^{(2)}{\displaystyle \frac{1}{3}}PA_{c\lambda }+{\displaystyle \frac{1}{3}}S_{c\lambda })\lambda _t^{(s)}(P_{t\lambda }+M_\lambda ^{(1)}+Q_sM_\lambda ^{(2)}).`$ The radiative decay widths are given, in terms of these amplitudes, by $$\mathrm{\Gamma }(\overline{B}V\gamma )=\frac{E_\gamma }{8\pi m_B^2}\underset{\lambda =L,R}{}|A(\overline{B}V\gamma _\lambda )|^2.$$ (32) ### A The short-distance amplitude Due to the smallness of the light quark masses $`m_d`$, $`m_s`$ appearing in the short-distance Hamiltonian (1), relative to the $`b`$-quark mass, it couples predominantly to left-handed photons. The amplitude for emitting right-hand polarized photons is suppressed relative to the one for left-handed photons by the quark mass ratio $`|P_{tR}/P_{tL}|=𝒪(m_{d,s}/m_b)`$. Therefore, the right-handed amplitude $`P_{tR}`$ can be neglected to a good approximation. The left-handed short-distance amplitude can be expressed in terms of the $`BV`$ form factors of the tensor current, defined as $`V(p^{},ϵ)|\overline{q}\sigma _{\mu \nu }b|\overline{B}(p)`$ $`=`$ $`g_+(q^2)\epsilon _{\mu \nu \lambda \sigma }ϵ_\lambda ^{}(p+p^{})_\sigma +g_{}(q^2)\epsilon _{\mu \nu \lambda \sigma }ϵ_\lambda ^{}(pp^{})_\sigma `$ (34) $`+h(q^2)\epsilon _{\mu \nu \lambda \sigma }(p+p^{})_\lambda (pp^{})_\sigma (ϵ^{}p).`$ Using the relation $`\sigma _{\mu \nu }\gamma _5=\frac{i}{2}\epsilon _{\mu \nu \alpha \beta }\sigma ^{\alpha \beta }`$ (corresponding to $`\epsilon ^{0123}=1`$), one finds $`V(p^{},ϵ)|\overline{q}\sigma _{\mu \nu }\gamma _5b|\overline{B}(p)=`$ (35) $`ig_+(q^2)[ϵ_\mu ^{}(p+p^{})_\nu ϵ_\nu ^{}(p+p^{})_\mu ]ig_{}(q^2)[ϵ_\mu ^{}(pp^{})_\nu ϵ_\nu ^{}(pp^{})_\mu ]`$ (36) $`ih(q^2)(ϵ^{}p)[(p+p^{})_\mu (pp^{})_\nu (p+p^{})_\nu (pp^{})_\mu ].`$ (37) The relevant quantity is the formfactor $`g_+`$ at the kinematical point $`q^2=0`$. This has been computed using light-cone QCD sum rules (LCSR) and lattice QCD , with results which are in reasonable agreement with each other. The most recent LCSR calculation gave $`g_+^{(\rho )}(0)=0.29\pm 0.04`$ and $`g_+^{(K^{})}(0)=0.38\pm 0.06`$ , which compares well with the lattice result $`g_+^{(K^{})}(0)=0.32_{0.02}^{+0.04}`$ . A somewhat larger value $`g_+^{(K^{})}(0)=0.4`$ was recently extracted from the $`DK^{}e\nu `$ semileptonic form factors using heavy quark symmetry relations . Using the latter value for $`g_+(0)`$, one finds for the short-distance penguin amplitudes with an internal $`t`$ quark $`P_{tL}`$ $`=`$ $`8\sqrt{2}G_F{\displaystyle \frac{e}{16\pi ^2}}C_7(m_b)m_bm_BE_\gamma g_+(q^2=0)`$ (38) $`=`$ $`(1.8\times 10^6)\left({\displaystyle \frac{g_+(0)}{0.4}}\right)\left({\displaystyle \frac{m_b\text{ (GeV)}}{4.2}}\right)\text{ GeV}`$ (39) and $`P_{tR}0`$. We used in this numerical estimate the leading-log value for the Wilson coefficient $`C_7(m_b)=0.31`$. In the following section we will prove a similar suppression of the right-handed amplitude relative to the left-handed one for the weak annihilation ($`A`$) and the $`W`$-exchange ($`E`$) amplitudes in $`\overline{B}`$ decays, to leading order in an expansion in powers of $`1/E_\gamma `$. ### B Weak annihilation and $`W`$-exchange amplitudes We start by assuming factorization; the corrections to this approximation will be discussed below. The factorized weak annihilation amplitude contributing to the $`B^{}\rho ^{}\gamma `$ decay is written as $`A_\lambda ={\displaystyle \frac{G_F}{\sqrt{2}}}a_1\left\{f_Bp_{B\mu }\rho ^{}\gamma _\lambda |(\overline{d}u)_{VA}^\mu |0+m_\rho ^{}f_\rho ^{}(ϵ_\rho ^{})_\mu \gamma _\lambda |(\overline{u}b)_{VA}^\mu |B^{}\right\},`$ (40) corresponding to the photon coupling to the $`B^{}`$ or to the $`\rho ^{}`$ constituent quarks respectively. To tree level, the Wilson coefficient $`a_1`$ is given by $`a_1=C_1(m_b)+C_2(m_b)/N_c=1.02`$ . The first term can be computed exactly in the chiral limit with the result $`p_{B\mu }\rho ^{}\gamma _\lambda |(\overline{d}u)_{VA}^\mu |0=\rho ^{}\gamma _\lambda |i_\mu (\overline{d}u)_{VA}^\mu |0=em_\rho ^{}f_\rho ^{}(ϵ_\rho ^{}ϵ_\lambda ^{})+𝒪(m_u,m_d).`$ (41) We used here the following relations for the divergence of the isovector and isoaxial currents in the presence of an electromagnetic field $`𝒜`$ (with $`e>0`$) $`i_\mu (\overline{q}\gamma _\mu \lambda ^iq)`$ $`=`$ $`e𝒜_\lambda \overline{q}[\widehat{𝒬},\lambda ^i]\gamma _\lambda q\overline{q}[\widehat{m},\lambda ^i]q`$ (42) $`i_\mu (\overline{q}\gamma _\mu \gamma _5\lambda ^iq)`$ $`=`$ $`e𝒜_\lambda \overline{q}[\widehat{𝒬},\lambda ^i]\gamma _\lambda \gamma _5q\overline{q}\{\widehat{m},\lambda ^i\}\gamma _5q.`$ (43) $`\widehat{𝒬}=\text{diag }(\frac{2}{3},\frac{1}{3},\frac{1}{3})`$ and $`\widehat{m}=\text{diag }(m_u,m_d,m_s)`$ are the quarks’ electric charge and mass matrix, respectively. The result (41) is analogous to a similar one obtained in for the long-distance contribution to the decay $`B\pi e^+e^{}`$. The remaining contribution can be expressed in terms of the two form factors $`f_{V,A}(E_\gamma )`$ parametrizing the radiative leptonic decay $`B^{}\gamma e\overline{\nu }`$ $`{\displaystyle \frac{1}{\sqrt{4\pi \alpha }}}\gamma (q,ϵ_\lambda )|\overline{q}\gamma _\mu (1\gamma _5)b|\overline{B}(v)=`$ (44) $`i\epsilon (\mu ,ϵ_\lambda ^{},v,q)f_V(E_\gamma )+[ϵ_{\lambda \mu }^{}(vq)q_\mu (ϵ_\lambda ^{}v)]f_A(E_\gamma ){\displaystyle \frac{1}{E_\gamma }}(Q_qQ_b)f_Bm_B(vϵ_\lambda ^{})v_\mu ,`$ (45) where $`v`$ denotes the $`B`$ meson velocity ($`p_B=m_Bv`$). The last term is present only for charged $`B`$ mesons, and is required for the gauge invariance of the complete decay amplitude, including also the lepton part. Although it is not relevant for real photon processes, it does contribute to decays involving virtual photons, which can probe the $`\mu =0`$ component of the electromagnetic current. A detailed derivation of this term is presented in the Appendix. Combining the two terms in (40) one finds the following result for the weak annihilation helicity amplitudes $`A_\lambda `$ $`A_{L,R}={\displaystyle \frac{G_F}{\sqrt{2}}}a_1em_\rho ^{}f_\rho ^{}\left[f_B+E_\gamma (f_A^{(B^{})}(E_\gamma )f_V^{(B^{})}(E_\gamma ))\right].`$ (46) Eventually the form factors $`f_{V,A}(E_\gamma )`$ will be extracted from the doubly differential spectrum $`d^2\mathrm{\Gamma }/dE_edE_\gamma `$ in the radiative leptonic decay $`B^\pm \gamma e\nu `$, which will allow a model-independent calculation of the weak annihilation amplitudes $`A_{L,R}`$. A considerable simplification can be achieved if the $`WA`$ amplitudes are expanded in powers of $`\mathrm{\Lambda }/E_\gamma `$ using the methods of perturbative QCD for exclusive processes . It has been shown in that the leading terms in the expansion (of order $`O(1/E_\gamma )`$) of $`f_V(E_\gamma )`$ and $`f_A(E_\gamma )`$ are related, and can be expressed in terms of the valence light-cone wave function of the $`B`$ meson as $`f_V^{(B^\pm )}(E_\gamma )=\pm f_A^{(B^\pm )}(E_\gamma )={\displaystyle \frac{f_Bm_B}{2E_\gamma }}\left(Q_uR{\displaystyle \frac{Q_b}{m_b}}\right)+𝒪\left({\displaystyle \frac{\mathrm{\Lambda }^2}{E_\gamma ^2}}\right).`$ (47) $`R`$ is a hadronic parameter given by an integral over the $`B`$ meson light-cone wave function $`R=_0^{\mathrm{}}𝑑k_+\psi _B(k_+)/k_+`$ (with the normalization $`_0^{\mathrm{}}𝑑k_+\psi _B(k_+)=1`$). A similar result is obtained in the quark model, with the identification $`R1/m_u`$, with $`m_u350`$ MeV the constituent light quark mass . The relation (47) among $`f_{V,A}(E_\gamma )`$ implies that a measurement of the photon spectrum in $`B^\pm \gamma e\nu `$ should be sufficient for their extraction (without any knowledge of the $`B`$ meson light-cone wavefunction). For the moment, in the absence of such data, the parameter $`R`$ can be estimated using model wavefunctions, which results in typical values around $`R=2.5\pm 0.5`$ GeV<sup>-1</sup> . We will use this central value in our estimates below. It is possible to give a model-independent lower bound for the $`R`$ parameter, in terms of the $`B`$ meson binding energy $`\overline{\mathrm{\Lambda }}=m_Bm_b`$. At tree-level this bound reads $`R3/(4\overline{\mathrm{\Lambda }})`$ . This can be used together with (46) and (47) to derive a lower bound on the magnitude of the $`A_L`$ amplitude (corresponding to $`B^{}\rho ^{}\gamma `$ decays) $`A_L\sqrt{2}G_Fa_1em_\rho ^{}f_\rho ^{}{\displaystyle \frac{3Q_uf_Bm_B}{8\overline{\mathrm{\Lambda }}}}=(0.54\times 10^6)\left({\displaystyle \frac{a_1}{1.0}}\right)\left({\displaystyle \frac{f_B\text{ (MeV)}}{175}}\right)\left({\displaystyle \frac{350}{\overline{\mathrm{\Lambda }}\text{ (MeV)}}}\right)\text{ GeV}.`$ (48) We used in this estimate $`f_\rho ^{}=216`$ MeV, as determined from the leptonic decay width of the $`\rho ^0`$ meson. To leading order in $`1/E_\gamma `$, the right-handed amplitude $`A_R`$ receives contributions only from the first term in (46) $`A_R={\displaystyle \frac{G_F}{\sqrt{2}}}a_1em_\rho ^{}f_\rho ^{}f_B=(0.07\times 10^6)\text{ GeV},`$ (49) where we used the same values for the hadronic parameters as above. Comparing with (48) one can see that the left/right-handed WA amplitude ratio is enhanced by a factor of more than 8. The $`W`$-exchange amplitude $`E`$ is given by an expression analogous to (40), with a different phenomenological factorization coefficient $`a_2=C_2(m_b)+C_1(m_b)/N_c=0.17`$ . The result corresponding to the decay $`\overline{B}^0\rho ^0\gamma `$ is written as $`E_\lambda =G_Fa_2\left\{f_Bp_{B\mu }\rho ^0\gamma _\lambda |(\overline{u}u)_{VA}^\mu |0+m_{\rho ^0}f_{\rho ^0}(ϵ_\rho ^{})_\mu \gamma _\lambda |(\overline{d}b)_{VA}^\mu |\overline{B}^0\right\}.`$ (50) In the chiral limit only the second term contributes, which gives (with $`f_{\rho ^0}=\frac{1}{\sqrt{2}}f_{\rho ^\pm }`$) $`E_{L,R}={\displaystyle \frac{G_F}{\sqrt{2}}}a_2em_{\rho ^0}f_{\rho ^+}E_\gamma (f_A^{(\overline{B}^0)}(E_\gamma )f_V^{(\overline{B}^0)}(E_\gamma )).`$ (51) The form-factors appearing on the RHS are given by an expression analogous to (47) (the sign change is due to our phase convention for the $`\overline{B}^0`$ state) $`f_V^{(\overline{B}^0)}(E_\gamma )=f_A^{(\overline{B}^0)}(E_\gamma )={\displaystyle \frac{f_Bm_B}{2E_\gamma }}\left(Q_dR{\displaystyle \frac{Q_b}{m_b}}\right)+𝒪\left({\displaystyle \frac{\mathrm{\Lambda }^2}{E_\gamma ^2}}\right).`$ (52) To the leading order in $`\mathrm{\Lambda }/E_\gamma `$, one finds the lower bound $`E_L\sqrt{2}G_Fa_2em_{\rho ^0}f_{\rho ^+}{\displaystyle \frac{3|Q_d|f_Bm_B}{8\overline{\mathrm{\Lambda }}}}=(0.05\times 10^6)\left({\displaystyle \frac{a_2}{0.175}}\right)\left({\displaystyle \frac{f_B\text{ (MeV)}}{175}}\right)\left({\displaystyle \frac{350}{\overline{\mathrm{\Lambda }}\text{ (MeV)}}}\right)\text{ GeV}.`$ (53) The $`W`$-exchange amplitude $`E`$ is suppressed relative to the WA amplitude $`A`$ by color-suppression in the ratio $`a_2/a_1`$ and by an additional factor $`|Q_d/Q_u|=1/2`$. The right-handed $`E_R`$ amplitude receives contributions only from the higher twist terms in (52) and is therefore suppressed relative to the left-handed amplitude $`E_R`$ by a factor of $`\mathrm{\Lambda }/E_\gamma 0.15`$, corresponding to a photon energy in $`B\rho (K^{})\gamma `$ decays $`E_\gamma =2.6`$ GeV. This gives the estimate $`E_R0.15E_L>(0.007\times 10^6)`$ GeV. The remaining amplitudes $`P_{u,c}`$ are considerably more difficult to evaluate. A great deal of effort has been put into attempts to calculate them, with methods such as simple vector meson dominance , dispersion techniques combined with Regge theory and light-cone QCD sum rules . In the following section we will revisit some of these estimates, putting them into perspective relative to the effects computed in this section (see Table I). | Photon helicity | $`|P_{t\lambda }|`$ | $`|P_{c\lambda }|`$ | $`|P_{u\lambda }|`$ | $`|A_\lambda |`$ | $`|E_\lambda |`$ | | --- | --- | --- | --- | --- | --- | | $`\lambda =L`$ | $`1.8`$ | $`0.16`$ | $`0.03`$ | $`0.6`$ | $`0.05`$ | | $`\lambda =R`$ | $`0`$ | $`0.04`$ | $`0.007`$ | $`0.07`$ | $`0.007`$ | > Table I. Estimates of the short-distance and long-distance amplitudes in $`B\rho \gamma `$ decays (in units of $`10^6`$ GeV). The estimates of the $`WA`$ and $`W`$-exchange amplitudes $`A_\lambda `$ and $`E_\lambda `$ used $`R=2.5`$ GeV<sup>-1</sup>. ### C Long-distance amplitudes with internal $`c`$ and $`u`$-quark loops The long-distance amplitude $`P_{c\lambda }`$ induced by the $`bc\overline{c}q`$ part of the weak Hamiltonian (4) is usually assumed to be dominated by the diagram wherein the photon couples to the charm quark loop. In a hadronic language, this contribution arises from the weak decay of the $`B`$ meson into $`V\psi ^{(n)}`$ (with $`\psi ^{(n)}`$ a $`\overline{c}c`$ state with the quantum numbers of the photon), followed by the annihilation of the $`\psi ^{(n)}`$ state into a photon . For an arbitrary photon virtuality $`q^2`$, this long-distance amplitude can be written as a sum over intermediate states $$A(\overline{B}V\psi ^{(n)}V\gamma _\lambda )=Q_ce\underset{n,\epsilon _n}{}\frac{0|\overline{c}\gamma \epsilon _\lambda ^{}c|\psi ^{(n)}(q,\epsilon _n)A(\overline{B}V\psi ^{(n)})}{q^2M_n^2+iM_n\mathrm{\Gamma }_n}.$$ (54) The lowest-lying state contributing to the sum is the $`J/\psi `$, with a mass of 3.097 GeV. Therefore, for a real photon, the $`J/\psi `$ propagator denominator is large, such that its contribution can be expected to be strongly suppressed. These heuristic arguments are supported by an explicit QCD sum rule calculation , where the charmed penguin amplitude $`P_{c\lambda }`$ is found to be less than 5% of the dominant short-distance amplitude $`P_{tL}`$. This result is confirmed by existing experimental data, as discussed below. In the following we will substantiate these qualitative arguments with an estimate of the sum (54), truncating it to a few lowest states. For definiteness we consider the decay $`B^{}\rho ^{}\gamma `$. The weak decay amplitude in (54) can be estimated using factorization, with the result $`A(B^{}\rho ^{}\psi ^{(n)})`$ $`=`$ $`{\displaystyle \frac{G_F}{\sqrt{2}}}\lambda _c^{(d)}a_2f_{\psi ^{(n)}}m_{\psi ^{(n)}}\{{\displaystyle \frac{2V(m_{\psi ^{(n)}}^2)}{m_B+m_\rho }}i\epsilon (ϵ_\psi ^{},p_B,p_\rho ,ϵ_\rho ^{})`$ (56) $`(m_B+m_\rho )A_1(m_{\psi ^{(n)}}^2)(ϵ_\psi ^{}ϵ_\rho ^{})+{\displaystyle \frac{2A_2(m_{\psi ^{(n)}}^2)}{m_B+m_\rho }}(ϵ_\rho ^{}q)(ϵ_\psi ^{}p_B)\}`$ The form-factors appearing here are defined in the usual way (with $`q=p^{}p`$) $`\rho ^{}(p^{},ϵ)|\overline{u}\gamma _\mu b|B^{}(p)={\displaystyle \frac{2V(q^2)}{m_B+m_\rho }}i\epsilon (\mu ,p,p^{},ϵ^{})`$ (57) $`\rho ^{}(p^{},ϵ)|\overline{u}\gamma _\mu \gamma _5b|B^{}(p)=2m_\rho A_0(q^2){\displaystyle \frac{ϵ^{}q}{q^2}}q_\mu +(m_B+m_\rho )A_1(q^2)\left(ϵ_\mu ^{}{\displaystyle \frac{ϵ^{}q}{q^2}}q_\mu \right)`$ (58) $`A_2(q^2){\displaystyle \frac{ϵ^{}q}{m_B+m_\rho }}\left(p_\mu +p_\mu ^{}{\displaystyle \frac{m_B^2m_\rho ^2}{q^2}}q_\mu \right),`$ (59) and the decay constant of the $`\psi `$ state is given by $`0|\overline{c}\gamma _\mu c|\psi (q,ϵ)=m_\psi f_\psi ϵ_\mu `$. Inserting (56) into the sum (54) and performing the sum over the $`\psi ^{(n)}`$ polarization one finds the following result for the long-distance charmed penguin contribution to the $`B^{}\rho ^{}\gamma `$ amplitude $`P_{c\lambda }(B^{}\rho ^{}\gamma _\lambda )`$ $`=`$ $`Q_ce{\displaystyle \frac{G_F}{\sqrt{2}}}a_2{\displaystyle \frac{2m_B}{m_B+m_\rho }}{\displaystyle \underset{n}{}}f_{\psi ^{(n)}}^2`$ (61) $`\times \left\{V(0)i\epsilon (q,ϵ_\lambda ^{},ϵ_\rho ^{},v)A_2(0)[(vq)(ϵ_\rho ^{}ϵ_\lambda ^{})(vϵ_\lambda ^{})(qϵ_\rho ^{})]\right\}.`$ To put the result into this (explicitly gauge-invariant) form, one must assume the following relation between the form factors $`A_1`$ and $`A_2`$ at $`q^2=0`$ $`A_1(0)=A_2(0){\displaystyle \frac{m_Bm_\rho }{m_B+m_\rho }}.`$ (62) While the approximate numerical values of these form factors $`V^{B\rho }(0)=0.34,A_1^{B\rho }(0)=0.26,A_2^{B\rho }(0)=0.22`$ do not satisfy exactly the relation (62), it is worth noting that it does hold true to leading order in a large energy expansion in powers of $`1/E_\rho `$. In , the form-factors for the decay of a $`B`$ meson into a light pseudoscalar (vector) meson have been analyzed using an effective theory for the final state quarks . To leading twist, all these form factors can be expressed in terms of two universal functions $`\zeta _{}(E_\rho )`$ and $`\zeta _{}(E_\rho )`$ $`V(q^2)`$ $`=`$ $`\left(1+{\displaystyle \frac{m_\rho }{m_B}}\right)\zeta _{}(E_\rho )`$ (63) $`A_1(q^2)`$ $`=`$ $`{\displaystyle \frac{2E_\rho }{m_B+m_\rho }}\zeta _{}(E_\rho )`$ (64) $`A_2(q^2)`$ $`=`$ $`\left(1+{\displaystyle \frac{m_\rho }{m_B}}\right)\left[\zeta _{}(E_\rho ){\displaystyle \frac{m_\rho }{E_\rho }}\zeta _{}(E_\rho )\right].`$ (65) The leading twist form factors satisfy certain relations, which are very similar (but not completely identical) to relations previously derived in the quark model . This is not completely unexpected, considering that to leading twist only the meson’s valence degrees of freedom are relevant, which esentially coincides with the quark model picture. Using the relations (64), (65) one can see that the condition (62) is indeed satisfied, up to corrections of higher order in $`m_\rho /E_\rho `$, proportional to $`\zeta _{}`$. The relations (63), (65) imply also that only left-handed photons are emitted to leading twist. The relevant helicity amplitudes are given by $`P_{cL,R}`$ $`=`$ $`Q_ce{\displaystyle \frac{G_F}{\sqrt{2}}}a_2{\displaystyle \frac{2m_B}{m_B+m_\rho }}E_\gamma (V(0)\pm A_2(0)){\displaystyle \underset{n}{}}f_{\psi ^{(n)}}^2.`$ (66) Using the realistic values for form-factors quoted above, one finds that the left/right-handed helicity amplitudes ratio is enhanced by about a factor of 5. Absolute values for the $`|P_{c\lambda }|`$ helicity amplitudes obtained by keeping only the lowest two states in (54) are quoted in Table I. The corresponding decay constants can be extracted from their leptonic widths and are given by $`f_{\psi ^{(1)}}=395`$ MeV, $`f_{\psi ^{(2)}}=293`$ MeV . A similar analysis can be performed for the long-distance amplitude $`P_{u\lambda }`$ describing quark diagrams with $`u`$-quark loops, which is given by a sum analogous to (54) $$A(\overline{B}VV^{}V\gamma _\lambda )=e\underset{V^{}=\rho ^0,\omega ,\mathrm{}}{}\frac{0|j_{\mathrm{e}.\mathrm{m}.}\epsilon _\lambda ^{}|V^{}(q,\epsilon )A_C(\overline{B}VV^{})}{q^2M_V^{}^2+iM_V^{}\mathrm{\Gamma }_V^{}}.$$ (67) The weak amplitudes $`A(\overline{B}VV^{})`$ receive, in general, both color-allowed and color-suppressed contributions (in contrast to the decays $`BV\psi `$ which are purely color-suppressed). For example, in $`B^{}\rho ^{}\gamma `$ decay, the color-allowed component gives a WA-type contribution to the radiative decay. To avoid double counting of amplitudes, we will keep therefore only the color-suppressed component of the weak amplitude in (67) (symbolized by the subscript $`C`$). Our analysis is different in this respect from other VMD estimates of this amplitude . Keeping only the contributions of the lowest two states $`V^{}=\rho ^0,\omega `$, one finds $`P_{uL,R}={\displaystyle \frac{G_F}{\sqrt{2}}}ea_2f_{\rho _0}m_{\rho _0}{\displaystyle \frac{2m_B}{m_B+m_\rho }}E_\gamma \left({\displaystyle \frac{f_{\rho _0}}{m_\rho }}+{\displaystyle \frac{f_\omega }{m_\omega }}\right)(V(0)\pm A_2(0)).`$ (68) In writing this result we used the nonet symmetry relation $`0|\overline{u}\gamma _\mu u|\rho ^0=0|\overline{u}\gamma _\mu u|\omega `$, valid in the large $`N_c`$ limit or the quark model. The corresponding numerical results are shown in Table I. The enhancement of the left-handed photon amplitudes relative to the right-handed ones in the long-distance contributions we found is somewhat surprising. Calculations of these amplitudes using different methods such as QCD sum rules and perturbative QCD find a similar result. Such an enhancement does not necessarily hold in the presence of new physics. In certain extensions of the Standard Model, such as the left-right symmetric model, the right-handed helicity amplitude can become large (proportional to the virtual heavy fermion mass) . Such effects lead to observable consequences such as significant mixing-induced CP asymmetries in weak radiative $`BV\gamma `$ decays . To summarize the results of this section, the rare radiative decays are dominated by the amplitudes corresponding to left-handed photons. The dominant long-distance contribution to the $`B^{}\rho ^{}\gamma `$ decay comes from the $`WA`$ mechanism; the corresponding amplitude $`A_\lambda `$ is calculable in a model-independent way from experimental data. The left-handed amplitudes satisfy an approximate hierarchy of sizes $`EP_u0.4P_c,P_c0.2A,A0.3P_t.`$ (69) The relative magnitude of the components with different CKM coefficients can be obtained by combining these estimates for the amplitudes with the corresponding CKM factors. ### D Nonfactorizable corrections The leading corrections to the factorization result (40) for the WA amplitudes come, in perturbation theory, from the graphs in Fig. 2, with one gluon exchanged between the $`B`$ and $`\rho `$ quarks. Let us take for definiteness the final vector meson to be moving along the positive $`z`$ axis. Momentum conservation $`m_Bv=q+p`$ is expressed in terms of light-cone components as $`p_+=m_B`$ and $`m_B=2E_\gamma +p_{}`$ (we define the light-cone coordinates as $`p_\pm =p^0\pm p^3`$). Then to leading order $`O(1)`$ in an expansion in powers of $`\mathrm{\Lambda }/p_+=\mathrm{\Lambda }/m_B`$, the $`\rho `$ meson can be represented by its valence component $`|q(p_1)\overline{q}(p_2)`$ with parallel momenta $`p_1=(yp_+,0_{},0_{})`$ and $`p_2=((1y)p_+,0_{},0_{})`$. In a tensor language this corresponds to the $`\rho `$ wave function $$(\mathrm{\Psi }_\rho )_{\alpha \beta }=\frac{\sqrt{N_c}}{8\sqrt{2}}p_+(\gamma _{}\mathit{\epsilon ̸})_{\alpha \beta }\varphi _{}(y)$$ (70) where $`\varphi _{}(y)`$ is the twist-2 chiral-odd structure function as appropriate for a transversely polarized $`\rho `$ meson. It satisfies the normalization condition $$_0^1𝑑y\varphi _{}(y)=\sqrt{\frac{2}{N_c}}f_\rho ^T,$$ (71) where the decay constant $`f_\rho ^T=(160\pm 10)`$ MeV is defined as $$0|\overline{u}\sigma _{\mu \nu }d|\rho ^{}(p,\epsilon )=if_\rho ^T(\epsilon _\mu p_\nu \epsilon _\nu p_\mu ).$$ (72) The $`B`$ meson is represented by the tensor wavefunction $`(\mathrm{\Psi }_B)_{\alpha \beta }={\displaystyle \frac{\sqrt{N_c}}{\sqrt{2}k_+}}\psi _B(k_+,\stackrel{}{k}_{})\left\{(k_++\stackrel{}{\alpha }_{}\stackrel{}{k}_{})\mathrm{\Lambda }_+{\displaystyle \frac{1+\mathit{}}{2}}\gamma _5\right\}_{\alpha \beta },\mathrm{\Lambda }_+={\displaystyle \frac{\gamma _{}\gamma _+}{4}}.`$ (73) In contrast to the light mesons $`\pi `$ or $`\rho `$, the leading twist description of a heavy meson involves in general (beyond tree level) also the transverse motion of the heavy quark. Each of the four diagrams in Fig. 2a-d contains IR divergences; however, an explicit calculation shows that they cancel in the sum. Therefore, (at least to one-loop order) the leading-twist nonfactorizable contribution is computable in perturbation theory as a convolution of the $`B`$ and $`\rho `$ light-cone wave functions with a IR-finite hard scattering amplitude $`T_H`$ $`\delta _{n.f.}A_L={\displaystyle _0^{\mathrm{}}}𝑑k_+{\displaystyle _0^1}𝑑y\varphi _{}(y,\mu )\psi _B(k_+,\mu )T_H(k_+,y,\mu )+\text{higher twist}.`$ (74) (This is analogous to a result recently found in for nonleptonic decays $`B\pi \pi `$.) However, it is easy to see that the hard-scattering amplitude vanishes in the limit of massless final quarks $`T_H=O(m_{u,d})`$. Technically, in this limit, the $`\rho `$ side of the matrix element contains only traces of an odd number of $`\gamma `$ matrices, which vanish. None of these conclusions depends on the fact that the photon is on-shell $`q^2=0`$, such that a similar result can be derived for $`B\rho _{}e^+e^{}`$ decays, mediated by a virtual photon. However, a nontrivial nonfactorizable correction is obtained for a longitudinally polarized $`\rho `$ or a pion in the final state. In conclusion, the one-loop nonfactorizable correction vanishes identically in the chiral limit and to leading-twist order. Therefore the factorized result (40) can be expected to give a very good approximation to the weak annihilation amplitude $`A`$. ## III Applications ### A Implications from experimental data The CLEO Collaboration recently reported the branching ratios for the $`BK^{}\gamma `$ exclusive modes $`(B^\pm K^\pm \gamma )`$ $`=`$ $`(3.76_{0.83}^{+0.89}\pm 0.28)\times 10^5`$ (75) $`(B^0K^0\gamma )`$ $`=`$ $`(4.55_{0.68}^{+0.72}\pm 0.34)\times 10^5.`$ (76) These results give the following ratio for the charge-averaged radiative decay widths $`{\displaystyle \frac{\mathrm{\Gamma }(B^0K^0\gamma )}{\mathrm{\Gamma }(B^\pm K^\pm \gamma )}}={\displaystyle \frac{\tau _{B^\pm }}{\tau _{B^0}}}{\displaystyle \frac{(B^0K^0\gamma )}{(B^\pm K^\pm \gamma )}}=1.29\pm 0.55.`$ (77) We used here the lifetime ratio of charged and neutral $`B`$ mesons $`\tau (B^\pm )/\tau (B^0)=1.066\pm 0.024`$ . Isospin symmetry constrains the short-distance amplitude $`P_{tL}`$ to be the same in both decays (75), (76). Therefore the deviation of the ratio (77) from 1 can only come from the long-distance contribution $`M^{(2)}P_c^{(2)}`$ arising from the photon coupling to the spectator quark. Using the unitarity of the CKM matrix one can rewrite $`\lambda _c^{(s)}=\lambda _u^{(s)}\lambda _t^{(s)}`$ in (8) and (12). Furthermore, neglecting the small CKM factor $`\lambda _u^{(s)}`$, these amplitudes can be written to a good approximation as $`A(B^{}K^{}\gamma )`$ $`=`$ $`\lambda _t^{(s)}\left(P_t+(M^{(1)}P_c^{(1)})+{\displaystyle \frac{2}{3}}(M^{(2)}P_c^{(2)})\right)`$ (78) $`A(\overline{B}^0\overline{K}^0\gamma )`$ $`=`$ $`\lambda _t^{(s)}\left(P_t+(M^{(1)}P_c^{(1)}){\displaystyle \frac{1}{3}}(M^{(2)}P_c^{(2)})\right).`$ (79) The short-distance amplitude $`P_{tL}`$ alone would give the following branching ratios $`_{\mathrm{s}.\mathrm{d}.}(B^\pm K^\pm \gamma )`$ $`=`$ $`4.6\times 10^5,_{\mathrm{s}.\mathrm{d}.}(B^0K^0\gamma )=4.5\times 10^5,`$ (80) where we used the estimate for $`P_{tL}`$ shown in Table I together with $`\lambda _t^{(s)}=0.039`$. The $`B`$ meson lifetimes were taken as $`\tau _{B^\pm }=1.64\pm 0.02`$ ps, $`\tau _{B^0}=1.55\pm 0.03`$ ps . Assuming that the central values of the experimental numbers (75), (76) will be confirmed by future, more precise determinations, one is led to conclude therefore that the long-distance effects interfere among themselves destructively in $`B^0`$ decays, whereas in $`B^\pm `$ they could be significant, and contribute up to 10% of the total amplitude. This agrees with the theoretical estimates of and disfavor a non-SM explanation of the isospin breaking in the ratio (77) based on an enhanced gluonic penguin as proposed in . In the SU(3) limit the long-distance contributions to $`B^0`$ and $`B_s`$ decays are the same (see (10)-(18)). Therefore one can use the observed $`B^0K^0\gamma `$ branching ratio to make predictions for other strangeness-changing radiative decays of these mesons. Neglecting the OZI-suppressed amplitudes $`S_{c\lambda }`$ and $`PA_{c\lambda }`$, one obtains the following prediction (together with $`\tau _{B_s}=1.46\pm 0.06`$ ps ) $`(B_s\varphi \gamma )`$ $`=`$ $`{\displaystyle \frac{\tau _{B^0}}{\tau _{B_s}}}(B^0K^0\gamma )4.8\times 10^5.`$ (81) Using the estimate for $`E_L`$ in Table I, and neglecting again the contributions of the OZI-suppressed annihilation penguin $`PA`$ and SU(3) singlet $`S`$ amplitudes, one can predict very small rates for the $`B_s`$ modes (we used here $`|V_{ub}|=1.6\times 10^2`$) $`(B_s\rho ^0\gamma )(B_s\omega \gamma )0.3\times 10^8.`$ (82) ### B Determining $`|V_{td}|`$ The main interest in observing the exclusive charmless radiative decay $`B^\pm \rho ^\pm \gamma `$ is connected with the possibility of determining the CKM matrix element $`|V_{td}|`$. The amplitude for this decay is related by $`U`$-spin symmetry to that for the $`bs\gamma `$ mode $`B^\pm K^\pm \gamma `$ $`A(B^{}\rho ^{}\gamma _L)`$ $`=`$ $`\lambda _u^{(d)}a_L+\lambda _t^{(d)}p_L,A(B^{}K^{}\gamma _L)=\lambda _u^{(s)}a_L^{}+\lambda _t^{(s)}p_L^{}.`$ (83) In the SU(3) symmetry limit $`a_L=a_L^{}=A_LP_{cL},p_L=p_L^{}=P_{tL}+M_LP_{cL}`$. In the absence of the $`a,a^{}`$ terms, the ratio of the two amplitudes is equal to $`|\lambda _t^{(d)}/\lambda _t^{(s)}|=|V_{td}/V_{ts}|`$ (up to SU(3) breaking corrections), which offers a way for determining $`|V_{td}|`$. In the general case, however, the $`(\rho ^{}\gamma _L)`$ amplitude can be written as $`A(B^{}\rho ^{}\gamma _L)`$ $`=`$ $`\lambda _t^{(d)}p(1{\displaystyle \frac{|\lambda _u^{(d)}|}{|\lambda _t^{(d)}|}}e^{i\alpha }\epsilon _Ae^{i\varphi _A}),`$ (84) with $`\epsilon _Ae^{i\varphi _A}a/pA_L/P_{tL}`$ (where, following the estimates of Sec. II we neglected the charmed penguin $`P_{cL}`$ and gluonic penguin $`M_L`$ amplitudes relative to the short-distance amplitude $`P_{tL}`$ and the $`WA`$ amplitude $`A_L`$). The estimates of the preceding section give $`\epsilon _A0.3`$, and $`\varphi _A=0`$ to leading twist. Furthermore, global analyses of the unitarity triangle suggest that the ratio of CKM factors appearing in this amplitude is subunitary $$\frac{|\lambda _u^{(d)}|}{|\lambda _t^{(d)}|}\left|\frac{V_{ub}}{V_{cb}}\right|\left|\frac{V_{ts}}{V_{td}}\right|0.4.$$ (85) The estimates of Sec. II.B show that the total rate for $`B^{}`$ ($`B^+`$) radiative decay is dominated by left-hand (right-hand) polarized photons. Neglecting the small right-handed (left-handed) component, which gives only a second order contribution to the ratio, one finds for the ratio of the charge-averaged rates $`{\displaystyle \frac{(B^{}\rho ^{}\gamma )+(B^+\rho ^+\gamma )}{(B^{}K^{}\gamma )+(B^+K^+\gamma )}}`$ (86) $`=\left({\displaystyle \frac{|p|}{|p^{}|}}{\displaystyle \frac{|V_{td}|}{|V_{ts}|}}\right)^2\left\{1{\displaystyle \frac{|\lambda _u^{(d)}|}{|\lambda _t^{(d)}|}}\epsilon _A\mathrm{cos}\alpha \mathrm{cos}\varphi _A+\left({\displaystyle \frac{|\lambda _u^{(d)}|}{|\lambda _t^{(d)}|}}\epsilon _A\right)^2\right\}.`$ (87) The factor in the braces is bounded from above and below by using the inequality $$1+x+x^21x\mathrm{cos}\alpha \mathrm{cos}\varphi _A+x^21x+x^2,x\frac{|\lambda _u^{(d)}|}{|\lambda _t^{(d)}|}\epsilon _A0.12.$$ (88) Assuming, conservatively, $`x=0.2`$ (which is on the upper side of the estimates for this parameter) gives an uncertainty of about 20% in the determination of $`|V_{td}|`$ arising from the factor in braces in (86). The SU(3) breaking factor in the ratio of the penguin amplitudes $`|p|/|p^{}|g_+^{(\rho )}(0)/g_+^{(K^{})}(0)=0.76\pm 0.22`$ is known with a uncertainty of about 30% (we used here the results of the LCSR calculation ). Future lattice QCD calculations will hopefully improve the precision with which this ratio is known. Adding these errors in quadrature one finds a theoretical error in the determination of $`|V_{td}|`$ from the ratio (86) of about $`35\%`$. The leading twist result $`\varphi _A=0`$ implies that the CP asymmetry in $`B^\pm \rho ^\pm \gamma `$ can be expected to be very small $`(A_{CP}\mathrm{sin}\varphi _A)`$. The above estimate for the long-distance contribution $`\epsilon _A`$ is in general agreement with the quark model of and QCD sum rules calculations. The advantage of our approach will become apparent once experimental data on radiative leptonic $`B`$ decays become available. Such data could be used, as discussed in Sec. II, to actually determine the WA amplitude $`A_L`$, and thereby the long-distance contamination $`\epsilon _A`$. ## IV Conclusions Present experimental data from CLEO on exclusive radiative weak decays are sufficiently precise such that the long-distance effects are beginning to be observable. A good control over the magnitude of these effects is important for an assessment of the uncertainty they induce into the extraction of $`|V_{td}|`$ by combining $`B^{}\rho ^{}\gamma `$ and $`B^{}K^{}\gamma `$ decays . Unfortunately, these effects have proved notoriously difficult to treat in any systematic way. In the present paper we focus on a certain class of long-distance contributions, arising from weak annihilation and $`W`$-exchange quark diagram topologies. We argue that the former gives the dominant long-distance correction to the amplitude for $`B^{}\rho ^{}\gamma `$ decays. These corrections can be computed reliably using factorization, in terms of form-factors observable in radiative leptonic decays $`B^+\gamma e^+\nu `$. The nonfactorizable corrections are shown to be very small, as they appear only at nonleading twist. A similar method can be used to compute the $`W`$-exchange-type long-distance contribution, relevant for the weak radiative $`B_d`$ and $`B_s`$ decays. Furthermore, to the leading order of an expansion in powers of $`\mathrm{\Lambda }/E_\gamma `$, one can show that the coupling of left-handed photons in the long-distance amplitude is greatly enhanced relative to that of right-handed photons, just as in the short-distance part of the amplitude. Such an enhancement had been previously observed in QCD sum rule calculations of the long-distance amplitude ; our approach clarifies the theoretical limit in which this enhancement holds and quantifies the magnitude of the corrections to it. Our results should allow one to reduce the model-dependence of the leading long-distance effects in $`B^{}\rho ^{}\gamma `$ decays, and achieve a better control over the theoretical error in the corresponding determination of $`|V_{td}|`$. ###### Acknowledgements. D. P. would like to thank Jon Rosner for discussions of the quark diagram approach and Vivek Sharma for comments about the experimental feasibility of some of the methods discussed here. We are grateful to Ahmed Ali and Hai-Yang Cheng for comments on the manuscript and to Detlef Nolte for many discussions. This work has been supported by the National Science Foundation. ## A Ward identities for long-distance matrix elements We present in this Appendix a set of constraints on certain long-distance contributions to weak radiative decays of $`B`$ mesons. These constraints follow from a Ward identity expressing the conservation of the electromagnetic current. Let us consider the following matrix element of a local operator $`𝒪`$, to first order in electromagnetism and to all orders in the strong coupling $`\gamma (q,ϵ)f|𝒪(0)|B(v)=ieϵ_\mu ^{}{\displaystyle \text{d}^4xe^{iqx}f|\text{T}j_\mu ^{\mathrm{e}.\mathrm{m}.}(x)𝒪(0)|B(v)},`$ (A1) for any hadronic final state $`f`$. The operator $`𝒪`$ can be any quark bilinear including gluon fields, or even a four-quark operator. The electromagnetic current includes contributions from both the light and heavy quarks $`j_\mu ^{\mathrm{e}.\mathrm{m}.}=\overline{q}\widehat{𝒬}\gamma _\mu q+\frac{2}{3}\overline{c}\gamma _\mu c\frac{1}{3}\overline{b}\gamma _\mu b`$. A relation similar to (A1) can be written for the matrix element with the photon replaced with a dilepton pair in the final state, coupling through a virtual photon to the hadronic system. The conservation of the electromagnetic current implies, in the standard way, a Ward identity for the matrix element of the time-ordered product in (A1) $`iq_\mu {\displaystyle \text{d}^4xe^{iqx}f|\text{T}j_\mu ^{\mathrm{e}.\mathrm{m}.}(x)𝒪(0)|B(v)}={\displaystyle \text{d}^3xe^{i\stackrel{}{q}\stackrel{}{x}}f|[j_0^{\mathrm{e}.\mathrm{m}.}(\stackrel{}{x}),𝒪(\stackrel{}{0})]|B(v)}.`$ (A2) The commutator on the RHS is nonvanishing only if the operator $`𝒪`$ carries an electric charge, as in the case of decays induced by the weak charged current such as $`B^+\gamma e^+\nu `$, $`B\pi (\rho )\gamma e^+\nu `$ or $`\overline{B}D^{()}\gamma e^+\nu `$. In the following we analyze the simplest such case, the radiative leptonic decays of a $`B`$ meson, for which $`𝒪=\overline{b}\gamma _\nu \gamma _5q`$ and $`|f=|0`$. The equal-time commutator on the RHS of (A2) can be computed explicitly, with the result $`iq_\mu {\displaystyle \text{d}^4xe^{iqx}0|\text{T}j_\mu ^{\mathrm{e}.\mathrm{m}.}(x)(\overline{b}\gamma _\nu \gamma _5q)(0)|B(v)}`$ $`=`$ $`(Q_bQ_q)0|\overline{b}\gamma _\nu \gamma _5q|B(v)`$ (A3) $`=`$ $`(Q_bQ_q)f_Bm_Bv_\mu .`$ (A4) The most general parametrization of the matrix element on the LHS can be written in terms of five form-factors $`f_i(q^2,vq)`$ $`i{\displaystyle \text{d}^4xe^{iqx}0|\text{T}j_\mu ^{\mathrm{e}.\mathrm{m}.}(x)(\overline{b}\gamma _\nu \gamma _5q)(0)|B(v)}=f_1g_{\mu \nu }+f_2v_\mu v_\nu +f_3q_\mu q_\nu +f_4q_\mu v_\nu +f_5v_\mu q_\nu .`$ (A5) The Ward identity (A3) implies two constraints on these form-factors $`(vq)f_2+q^2f_4=(Q_bQ_q)f_Bm_B,f_1+q^2f_3+(vq)f_5=0.`$ (A6) For the case of a real photon $`q^2=0`$, these constraints fix uniquely the form-factor $`f_2(0,vq)`$, and relate $`f_1(0,vq)`$ and $`f_5(0,vq)`$. From Eq. (A1) one finds thus the following result for the matrix element of the axial weak current $`\gamma (q,ϵ)f|\overline{b}\gamma _\mu \gamma _5q|B(v)=f_5[(vq)ϵ_\mu ^{}(vϵ^{})q_\mu ]+(vϵ^{})v_\mu {\displaystyle \frac{1}{vq}}(Q_bQ_q)f_Bm_B,`$ (A7) which is the same as the result presented in text Eq. (44), with the identification $`f_5=f_A`$.
warning/0002/hep-ph0002004.html
ar5iv
text
# Gravitino production in hybrid inflationary models ## I Introduction Low energy effective $`N=1`$ supergravity is a predictive theory , which could predict an inflationary potential flat enough to provide adequate density perturbations . So far, such viable inflationary models were constrained from observations by fixing the height of the potential, which essentially determines the amplitude of the COBE normalization. The first and the second derivative of the potential determines the tilt in the power spectrum, and the Yukawa couplings of the inflaton to other particles determines the reheat temperature of the Universe. The higher the coupling constant is, the higher is the temperature of the thermal bath and so the creation of the gravitinos from the collisions or the decay of other particles. The gravitinos decay very late and depending on their mass their life time could be long enough to disrupt the synthesis of light elements, via hadronic shower or by altering the entropy density of the baryons. Due to these reasons there is a strong constraint on the reheat temperature, for a review see Ref. . However, there is also a non-thermal phase of the Universe, just after the end of slow-roll inflation when the scalar field begins to oscillate coherently at the bottom of the potential. During this era an explosive production of particles, both bosons and fermions , may take place due to the non-perturbative decay of the inflaton to other fields. It has also been shown that it is possible to create super-massive bosons and fermions. However, fermionic production is always saturated by the Pauli blocking. As a matter of fact creation of heavy non-thermal bosons can be a good candidate for weakly interacting massive particles, known as WIMPS . The decay of super-massive bosons can explain the ultra-high energy cosmic rays . Super-massive fermions can be used in leptogenesis, mainly from the decay of right-handed neutrinos into Higgs and leptons, which explicitly violates the CP conserving phase . It is worth mentioning that the non-perturbative technique of decaying inflaton to other particles has given a new paradigm shift in understanding the hot big bang universe from the ultra-cold inflationary regime. Recently preheating in the context of global supersymmetric theories has been considered and, as a natural extension it has been necessary to consider a local version of the supersymmetric theory and discuss the non-perturbative aspects of particle production and their consequences to nucleosynthesis. The local version of supersymmetry, known as supergravity, naturally accommodates the graviton and its superpartner the gravitino, a spin $`3/2`$ particle. Quantization of spin $`3/2`$ particles in the presence of an external background is plagued by consistency problems. It has been known for a long time that quantization of spin $`3/2`$ particles in scalar, electromagnetic, or gravitational backgrounds can give rise to acausal behaviour . Supergravity is the only set-up where such problems do not occur, provided the background fields also satisfy the corresponding equations of motion . Nevertheless, the complicated form of the Rarita-Schwinger equation makes it extremely difficult to extract any explicit results even in a simple background . The problem was first addressed in Ref. , where the authors quantized spin $`3/2`$ particles in a non-vanishing cosmological background, almost two decades ago. The slightest generalization of quantizing spin $`3/2`$ has been done very recently in literature, in Ref. . The authors have extended the calculation of quantizing spin $`3/2`$ in presence of a time-varying homogeneously oscillating scalar field in a cosmological background. This is the first result where the non-perturbative decay of the inflaton to gravitinos during preheating has been taken into account. The authors have explicitly shown the production of a particular helicity, the $`\pm 3/2`$ components of gravitino, in a particular new-inflationary type model . It has been noticed that the non-perturbative result can give rise to a larger abundance compared to the perturbative decay. The gravitino to photon number density has been found to be $`n_{3/2}/n_\gamma 10^{12}`$ . This abundance is $`3`$ orders of magnitude larger than the thermal abundance for a reheat temperature $`10^5`$ GeV . Such over production of gravitinos has been the first proof of the non-thermal production of gravitinos with helicity $`\pm 3/2`$, which demands constraining the reheat temperature in any supergravity motivated inflationary model. However, a massive gravitino has $`4`$ degrees of freedom, the other two degrees of freedom due to the helicity $`\pm 1/2`$ components of the gravitino. The production of helicity $`\pm 1/2`$ gravitinos is directly related to the problem of super-Higgs mechanism in supergravity models, which was studied in Ref. in the context of a non-vanishing cosmological constant, and in Ref. with a vanishing cosmological constant. In the presence of a time-varying scalar field there is an additional source of supersymmetry breaking via the non-vanishing time derivative of the homogeneous scalar field. This plays an important role in the context of cosmology when the scalar field is recognized as the inflaton, oscillating coherently at the bottom of the scalar potential. Due to the presence of such a field, supersymmetry is always broken at the minima of the potential and the initially massless gravitino which possesses only the helicity $`\pm 3/2`$ components “eats” the goldstino to gain the other $`\pm 1/2`$ components. At this point, one may wonder how to generalize the super-Higgs mechanism in such a scenario. In fact the problem turns out to be quite complicated and it has been addressed in two seminal papers . These papers also study for the first time the production mechanism of the helicity $`\pm 1/2`$ components of gravitino ( see other papers in the similar context ). In all these papers the authors have assumed the existence of a unitary gauge, where the physical Lagrangian is free from the goldstino field. We will explicitly show that it is possible to choose such a gauge, where the gravitino equation of motion is free from the goldstino field. Our calculation is valid for more than one chiral field as well. Though we shall give a proof for the $`F`$-type supersymmetry breaking, this can be extended to $`D`$-type supersymmetry breaking also. It has been noticed in Refs. that the production of the two helicity states are completely different. The helicity $`\pm 1/2`$ components are produced copiously compared to the helicity $`\pm 3/2`$. In the case of helicity $`\pm 3/2`$, conformal invariance is broken due to the presence of a time-varying gravitino mass, which is usually Planck mass suppressed, whereas for helicity $`\pm 1/2`$ the breaking of conformal invariance is related to the presence of a massive Goldstone fermion , whose time-varying mass is not suppressed by the Planck mass. Moreover in the high momentum limit the helicity $`\pm 1/2`$ gravitino behaves like a fermion, the goldstino, as it is stated by the equivalence principle . This has been studied in a single-chiral field scenario, where the source of conformal breaking can be directly related to the mass of Goldstino. Unfortunately in the multi-chiral field scenario the quantization scheme becomes more involved, and the relation between conformal breaking and Goldstino mass is not so straightforward. The situation has been briefly discussed in Ref. , and an attempt of a perturbative scheme has been suggested. However, it would be nice to discuss a non-perturbative scheme. Our paper fills that gap and, as we shall see, we can discuss non-perturbative production of the helicity $`\pm 1/2`$ gravitinos in the multi-chiral case. The best example to study the multi-chiral field scenario is in the context of a general class of supersymmetric hybrid inflation model . There are essentially two scalar fields, one is responsible for inflation, and the other field is responsible for the phase transition which results in terminating the inflationary era. Unlike the non-supersymmetric version of hybrid inflation model , the supersymmetric version considered in this paper has only one coupling constant in the potential. This leads to a single natural frequency of oscillations. This gives us an ample opportunity to use techniques to explore gravitino production in a similar spirit as in the case of a single-chiral field. Fermionic production in this case is very much different compared to that of a quadratic inflationary potential. In the hybrid models the effective mass term for the fermions is always positive and as a result the production can never be completed in just a few oscillations, rather the occupation number gradually increases and depending on the model parameters the rate of production could be slowed down significantly. For example, in order to reach the Fermi saturation it may be necessary more than $`100`$ oscillations. Due to such a slow rate of production, the issue of back-reaction becomes important. It is very likely that simultaneous non-perturbative production of bosons can change the picture quite significantly. In some sense, hybrid model can be considered to be the safest of all the supergravity motivated inflationary models, because the gravitino production can stop due to back-reaction effects coming from other newly created bosons or fermions. In other models, such as in the chaotic models, where most of the particle creation takes place in the very first few oscillations, the issue of back-reaction hardly plays any significant role. The layout of the paper is as follows; in section $`2`$, we establish the supergravity Lagrangian and discuss the gauge fixing mechanism for the multi-chiral field scenario; this is a mere generalization of the $`R_\zeta `$ gauge usually used in pure gauge theories as well as an alternative way of explaining the existence of a unitary gauge in supergravity theories . In section $`3`$, we discuss the quantization procedure for the gravitinos. In section $`4`$, we briefly describe a general class of supersymmetric hybrid inflationary models we use. Analytical results and discussions based on our numerical results are presented in section $`5`$. The implications of these results on the reheat temperature are discussed in section $`6`$. We give a detailed discussion on fermionic creation in hybrid model in the appendix. ## II General supergravity Lagrangian and gauge-fixing In this section we describe the supergravity Lagrangian. For the sake of brevity, and for our purpose, we concentrate upon the chiral-supermultiplet which contains the bosonic part, fermionic part and the interaction terms between the fermions. We consider minimal Kähler potential with $`K=\varphi ^i\varphi _i`$, where the scalar fields $`\varphi _i`$ are taken to be real. The superpotential is denoted by $`W`$, and the Kähler function is given as usual by $`G=K+\mathrm{ln}|W|^2`$. The choice of minimal Kähler potential also ensures $`G_j^i=\delta _j^i`$. The total Lagrangian is as follows : $`e^1_{\mathrm{SUGRA}}={\displaystyle \frac{R}{2}}+_\mu \varphi _i^\mu \varphi ^i+e^G\left(3G_iG^i\right)`$ (1) $``$ $`{\displaystyle \frac{e^1}{2}}ϵ^{\mu \nu \rho \sigma }\overline{\psi }_\mu \gamma _5\gamma _\nu D_\rho \psi _\sigma +{\displaystyle \frac{i}{2}}\overline{\chi }_i\gamma ^\mu D_\mu \chi _i+{\displaystyle \frac{i}{2}}e^{G/2}\overline{\psi }_\mu \sigma ^{\mu \nu }\psi _\nu `$ (2) $`+`$ $`({\displaystyle \frac{i}{2}}G^j\overline{\chi }_i\overline{)}\varphi _j\chi ^i+{\displaystyle \frac{1}{\sqrt{2}}}\overline{\psi }_\mu \overline{)}\varphi ^i\gamma ^\mu \chi _i+{\displaystyle \frac{i}{\sqrt{2}}}e^{G/2}G^i\overline{\psi }_\mu \gamma ^\mu \chi _i`$ (3) $``$ $`{\displaystyle \frac{1}{2}}e^{G/2}(G^{ij}+G^iG^j)\overline{\chi }_i\chi _j+h.c.),`$ (4) where indices $`i,j,\mathrm{},`$ represent species of the chiral multiplets and $`\overline{)}\gamma ^\lambda _\lambda `$. The derivatives $`D_\mu =_\mu \frac{i}{4}\omega _{\mu mn}\sigma ^{mn}`$ are covariantized with respect to gravity, where $`\omega _{\mu mn}`$ is the standard spin connection . The gravitino field is represented by $`\psi _\mu `$, $`\chi _i`$ represents the fermions and $`e`$ is the determinant of the vierbein $`e_\nu ^\mu `$. We have also used $`\sigma ^{\mu \nu }\frac{i}{2}[\gamma ^\mu ,\gamma ^\nu ]`$. The complete Lagrangian would also contain the fermion Yukawa couplings, four fermion terms, and numerous non-renormalizable terms, details can be found in the literature . Here we have neglected the torsion term in the covariantized derivative, assuming that the gravitino production is small, so their back reaction can be neglected. However, we mention that consistency of supergravity theory would demand the presence of the torsion term in the covariantized derivative. These terms are suppressed by the Planck mass square and we can safely neglect them in the linearized gravitino equation of motion. The Planck mass $`M_P`$ has been taken to be unity and the convention shall be retained, unless otherwise stated. The above Lagrangian is invariant under local supersymmetric transformation laws . For spontaneous supersymmetry breaking to occur at least one of the field vacuum expectation values should be non-zero. In particular, for $`F`$-term type breaking of local supersymmetry one requires $`0|\delta _\xi \chi _i|0=i\overline{)}\varphi _i\xi e^{G/2}G_i\xi 0,`$ (5) where $`\xi `$ is the infinitesimal Grassmann-odd parameter. The right-hand side of Eq. (5) has two explicit terms which can break local supersymmetry. The second term is the usual $`F`$-term of the scalar field whose non-vanishing vacuum expectation value induces supersymmetry breaking. The first term will give a non-zero contribution in case of a time-varying scalar background field. This will be the situation if we identify the background fields $`\varphi _i`$ with the oscillating inflaton fields. In this case local supersymmetry is always broken during the oscillations. The goldstino can be identified as usual from Eq. (5) $`\eta =\theta _i\chi ^i,`$ (6) where, $`\theta _i=i\overline{)}\varphi _ie^{G/2}G_i,`$ (7) and we follow the notations introduced in Ref. . We are only interested in homogeneous scalar fields $`\varphi _i`$ which are solely function of time. The goldstino is then “eaten” by the gravitino in local supersymmetric theories. This also ensures that the gravitino gains the helicity $`\pm 1/2`$ components other than the $`\pm 3/2`$ components, and becomes massive. In the high energy limit it is possible to relate the helicity $`\pm 1/2`$ components of the gravitino to the goldstino via the equivalence principle . In the limit when $`M_\mathrm{P}\mathrm{}`$, the helicity $`\pm 1/2`$ components retain the memory of the goldstino contribution and this is the reason why the two helicities behave differently and have different production rates . As it can be realized by inspecting Eq. (1), the gravitino is coupled to the fermions through the following mixing terms $$\frac{1}{\sqrt{2}}\overline{\psi }_\mu \left(ie^{G/2}G_i+\overline{)}\varphi _i\right)\gamma ^\mu \chi ^i.$$ (8) However, it is possible to get rid of these mixing terms by adding a gauge-fixing term to the Lagrangian as it has already been discussed in various places . In this section we mainly concentrate on the mere existence of such a gauge-fixing term in supersymmetric theories. Regarding this, we extend the previous calculation made on gauge-fixing for a single-chiral field<sup>*</sup><sup>*</sup>*In Ref. , the authors have only considered the equation of motion for the goldstino, because they were more interested in establishing the high-energy equivalence between helicity $`\pm 1/2`$ gravitinos and goldstinos. . Before we move onto specifying the gauge, let us introduce the projection operators, which we need to use later on . $`𝒫_{ij}^{}`$ $`=`$ $`\delta _{ij}{\displaystyle \frac{\theta _i^{}}{\theta ^{}\theta }}\theta _j,`$ (9) $`𝒫_{ij}^{}`$ $`=`$ $`{\displaystyle \frac{\theta _i^{}}{\theta ^{}\theta }}\theta _j.`$ (10) The modulus of $`\theta _i`$ is given by $`\theta ^{}\theta =e^GG^iG_i+\dot{\varphi }^i(t)\dot{\varphi }_i(t)=\rho +3e^G,`$ (11) where derivative with respect to time is denoted by dot, $`\rho =\dot{\varphi }^i\dot{\varphi }_i+V`$ is the energy density, and the scalar potential is $`V=e^G(G^iG_i3)`$. Accordingly, the fermion $`\chi _i`$ can be split into two components by using the projection operators Eqs. (9) and (10) $`\chi _i`$ $`=`$ $`\chi _i^{}+{\displaystyle \frac{\theta _i^{}}{\theta ^{}\theta }}\eta ,`$ (12) $`\chi _i^{}`$ $`=`$ $`𝒫_{ij}^{}\chi ^j.`$ (13) Now, with the help of Eqs. (7) and (12) the mixing terms between the gravitino and the chiral fermions can be recast as $$\frac{i}{\sqrt{2}}\overline{\psi }_\mu \theta _i\gamma ^\mu \frac{\theta ^i}{\theta ^{}\theta }\eta ,$$ (14) where we have used $`\theta _i\gamma ^\mu \chi ^i=0.`$ (15) In a similar way, one can also reduce the complex conjugate part of the mixing terms. Eq. (14) tells us about the direct coupling of the gravitino to the goldstino. The equation of motion for the gravitino $`\psi _\mu `$ that follows from the Lagrangian Eq. (1) is then $`e^1ϵ^{\mu \nu \rho \sigma }\gamma _5\gamma _\nu D_\rho \psi _\sigma +{\displaystyle \frac{1}{2}}e^{G/2}[\gamma ^\mu ,\gamma ^\nu ]\psi _\nu `$ (16) $`{\displaystyle \frac{i}{\sqrt{2}}}{\displaystyle \frac{\theta _i\gamma ^\mu \theta ^i}{\theta ^{}\theta }}\eta =0,`$ (17) with an explicit term depending on the goldstino field due to the mixing. However, we notice that the contribution of the mixing term in Eq. (14) can be canceled by adding to the Lagrangian the following gauge-fixing term $`i\zeta \overline{F}\overline{)}DF,`$ (18) where $`\overline{)}D=\gamma ^\lambda D_\lambda `$, $`\overline{F}=F^{}\gamma _0`$, and the gauge-fixing function is given by $`F(\psi ,\eta )={\displaystyle \frac{\theta ^i\gamma ^\mu \theta _i^{}}{\theta ^{}\theta }}\psi _\mu +{\displaystyle \frac{1}{\sqrt{2}\zeta }}{\displaystyle \frac{1}{\overline{)}D}}\eta ,`$ (19) This is a mere generalization of the $`R_\zeta `$ gauge used in pure gauge theories. Similar gauge-fixing has been initially introduced in Ref. in the static case, where the scalar field has been taken to its value at the minimum of the potential. Therefore, once we fix the gauge, the equation of motion for the gravitino is completely free from the goldstino. The limit $`\zeta 0`$ corresponds to the unitary gauge, and it implies $`\eta 0`$. This is equivalent to demand that no goldstino be present in the physical spectrum. From here onwards we will work in the unitary gauge. Using the gauge-fixing condition, $`F(\psi ,\eta )=0`$, and demanding $`\zeta 0`$, the final equation of motion for the gravitino, namely Eq. (16) in the unitary gauge can be written as $`e^1ϵ^{\mu \nu \rho \sigma }\gamma _5\gamma _\nu D_\rho \psi _\sigma +{\displaystyle \frac{1}{2}}e^{G/2}[\gamma ^\mu ,\gamma ^\nu ]\psi _\nu =0.`$ (20) Here the mass term for the gravitino $`e^{G/2}`$ depends on time, due to presence of oscillating background scalar field $`\varphi _i(t)`$, whose dynamics we shall discuss later on. From now onwards we express the mass term as $`m(t)`$ $`m(t)=e^{G/2}e^{K/2}|W|,`$ (21) where $`W`$ is the model dependent superpotential. Detailed discussion will be given in the coming sections. The above demonstration of removing the goldstino dependence from the gravitino equation of motion in the unitary gauge suggets that it is possible to generalize $`R_\zeta `$ gauge for a multi-chiral time-varying scalar background. Imposing the unitary gauge simplifies the gravitino field equation of motion in general and now we will be interested in quantizing gravitino field in a cosmological background. ## III Difference between helicity $`1/2`$ and $`3/2`$ We have seen in the preceeding section that we can recognize the goldstino component which couples to the gravitino field in a scenario when the dynamics of the background scalar field is also taken into consideration. By using the gauge-fixing term in the Lagrangian, we have shown that it is possible to cancel the gravitino-goldstino coupling term appropriately. The final equation of motion for the gravitino field, thus free from the goldstino, can be used to study the gravitino production in a dynamical background dominated by an oscillating scalar field. In this section we do not attempt to re-derive the equations of motion, which have already been discussed in several papers , rather we discuss few issues and the main equations. By studying the equation of motion for the Rarita-Schwinger field, one notices that there is a free index left, which in principle can be contracted by at best two possible ways, say $`\gamma _\mu `$ or $`D_\mu `$, thus giving rise to two constraint equations for the whole system. In presence of a cosmological constant, the equations of motion for both the helicities look alike, with two simple constraint equations, namely $`\gamma ^\mu \psi _\mu =0,D^\mu \psi _\mu =0`$. However, this is not correct in any arbitrary gravitational background. These constraints do not hold true for the helicity $`\pm 1/2`$ case in an oscillating scalar background , even though these constraints continue to hold for the helicity $`\pm 3/2`$ case in the same oscillating background . The helicity $`\pm 1/2`$ components gain an effective mass during the oscillations of the inflaton, but the helicity $`\pm 3/2`$ do not seem to see the effect of curvature at all. This gain in mass is purely due to presence of a non-trivial background curvature, which suggests that the different helicities couple to gravity differently. We follow the notations of Ref. in order to write down the equations of motions for helicity $`1/2`$ and $`3/2`$ for a momentum mode $`(0,0,0,|k|)`$ projected along the $`z`$ direction. $`\left(i\gamma ^0_0+i{\displaystyle \frac{5}{2}}{\displaystyle \frac{a^{}}{a}}\gamma ^0ma+k\gamma ^3\right)\psi _{3/2}(\tau ,x)=0,`$ (22) $`\left(i\gamma ^0_0+i{\displaystyle \frac{5}{2}}{\displaystyle \frac{a^{}}{a}}\gamma ^0ma+kG\gamma ^3\right)\psi _{1/2}(\tau ,x)=0.`$ (23) The equations have been written in conformal time $`d\tau =dt/a`$, where $`a`$ is the scale factor. Prime denotes derivative with respect to $`\tau `$, and $`m`$ is the gravitino mass Eq. (21). Here $`\psi _{1/2}`$ and $`\psi _{3/2}`$ are Majorana spinors which would correspond in the flat limit case to the $`1/2`$ and $`3/2`$ helicity states respectively (for details see Ref. ). The matrix $`G`$ in Eq. (23) can be expressed in terms of $`A`$ and $`B`$ functions $`G=A+i\gamma ^0B={\displaystyle \frac{p3m^2}{\rho +3m^2}}+i\gamma ^0{\displaystyle \frac{2m^{}a^1}{\rho +3m^2}},`$ (24) where $`\rho `$ and $`p`$ are denoted by $`\rho `$ $`=`$ $`{\displaystyle \underset{i}{}}|\dot{\varphi }_i|^2+V(\varphi _i),`$ (25) $`p`$ $`=`$ $`{\displaystyle \underset{i}{}}|\dot{\varphi }_i|^2V(\varphi _i),`$ (26) with $`V(\varphi _i)`$ $`=`$ $`e^K\left(|_iW+\varphi _iW|^23|W|^2\right).`$ (28) Here again dot means time derivative with respect to physical time. The potential energy Eq. (28) is given in terms of the Kähler potential $`K`$ and the superpotential $`W`$. In the limit when $`\varphi _i1`$ (in units of the Planck mass), $`A`$ and $`B`$ can be expressed in a simpler form $`G=A+i\gamma ^0B={\displaystyle \frac{p}{\rho }}+i\gamma ^0{\displaystyle \frac{2\dot{W}}{\rho }}.`$ (29) It is important to point out that in general $`|G|^2=A^2+B^2`$ is time dependent, and $`|G|1`$. Only in the case of a single-chiral field $`|G|=1`$. For multi-chiral field scenario, in particular in the case of a supersymmetric hybrid model, $`|G|`$ departs from $`1`$. This makes the quantization scheme slightly more involved than simple scenarios where $`|G|=1`$. To proceed with the quantization we redefine $`G`$ in terms of the conformal time $`G=A(\tau )+i\gamma ^0B(\tau )=e^{{\scriptscriptstyle \alpha 𝑑\tau }}e^{2i\gamma ^0{\scriptscriptstyle \mu 𝑑\tau }},`$ (30) where the coefficient in front of the overall phase represents $`|G|`$. We concentrate upon the helicity $`1/2`$ case, helicity $`3/2`$ being a simpler generalization of that. We expand $`\psi _{1/2}`$ in terms of the mode functions $`\psi _{1/2}`$ $`=`$ $`a^{5/2}{\displaystyle \frac{d^3k}{(2\pi )^{3/2}}e^{i\stackrel{}{k}.\stackrel{}{x}}e^{i\gamma ^0{\scriptscriptstyle \mu 𝑑\tau }}}`$ (32) $`\times {\displaystyle \underset{r=1,2}{}}(u^r(\tau ,\stackrel{}{k})a_k^r+v^r(\tau ,\stackrel{}{k})b_k^r),`$ where, $`v^r(\tau ,\stackrel{}{k})=u^{r^C}(\tau ,\stackrel{}{k})`$. The spinor $`u^r(\tau ,\stackrel{}{k})`$ satisfies the following equations of motion $`u_\pm ^{}`$ $`=`$ $`im_{\mathrm{eff}}u_\pm +ik|G|u_{},`$ (33) $`m_{\mathrm{eff}}`$ $`=`$ $`ma+\mu ,`$ (34) where $`u^T=(u_+,u_{})`$, and for the gamma matrices we have used the representation in which $$\gamma ^0=\left(\begin{array}{cc}\mathrm{𝟏}& 0\\ 0& \mathrm{𝟏}\end{array}\right),\gamma ^3=\left(\begin{array}{cc}0& \mathrm{𝟏}\\ \mathrm{𝟏}& 0\end{array}\right).$$ (35) It is possible to write down a second order differential equation from the set of equations in Eq. (33). $`u_\pm ^{\prime \prime }{\displaystyle \frac{|G|^{}}{|G|}}u_\pm ^{}+(k^2|G|^2+m_{\mathrm{eff}}^2\pm i{\displaystyle \frac{|G|^{}}{|G|}}m_{\mathrm{eff}}`$ (36) $`im_{\mathrm{eff}}^{})u_\pm =0.`$ (37) Eq. (36) can be further reduced by redefining $`u_\pm e^{{\scriptscriptstyle \alpha /2𝑑\tau }}u_\pm `$ $`u_\pm ^{\prime \prime }+\left(k^2|G|^2+\mathrm{\Omega }^2i\mathrm{\Omega }^{}\right)u_\pm =0,`$ (38) $`\mathrm{\Omega }=m_{\mathrm{eff}}+i{\displaystyle \frac{\alpha }{2}}.`$ (39) or, by redefining a new time in Eq. (36) $`\frac{d}{d\tau }=|G|\frac{d}{dz}`$, $`{\displaystyle \frac{d^2u_\pm }{dz^2}}+\left(k^2+\left({\displaystyle \frac{m_{\mathrm{eff}}}{|G|}}\right)^2i{\displaystyle \frac{d}{dz}}\left({\displaystyle \frac{m_{\mathrm{eff}}}{|G|}}\right)\right)u_\pm =0,`$ (40) where $`m_{\mathrm{eff}}`$ is defined in Eq. (33), and the equation is formally analogous to the evolution equation for a spin-1/2 fermion in a time-varying background. It is important to notice that all the three equations Eq. (36), Eq. (38) and Eq. (40) are equivalent, expressed in different forms. For our numerical results we have used Eq. (36), and for our analytical treatment we can consider any of these three equations. In the next section it will become clear that $`|G|1`$ for any general model, so that the factor $`|G|^2`$ in front of the momentum squared cannot give rise to any acausal behavior. Nevertheless, we note that particle creation could take place due to time variation in $`|G|`$. However, from Eq. (40) it is clear that this requires a non-vanishing effective mass term $`m_{\mathrm{eff}}`$, even if it is merely a constant. Particle production takes place due to the breaking of the conformal invariance. In our case, the violation of conformal invariance is not only due to presence of the gravitino mass but also due to $`\mu `$, which can be related to the goldstino mass . In order to evaluate the occupation number we first evaluate the hamiltonian $`H(\tau )={\displaystyle d^3k\underset{r}{}E_k(\tau )(a_r^{}a_rb_rb_r^{})}+F_k(\tau )b_ra_r`$ (41) $`+F_k^{}(\tau )a_r^{}b_r^{}),`$ (42) in which the momentum $`k`$ is along the third axis, and $`E_k`$ $`=`$ $`2k|G|Re(u_+^{}u_{})+m_{\mathrm{eff}}(|u_{}|^2|u_+|^2),`$ (43) $`F_k`$ $`=`$ $`2m_{\mathrm{eff}}u_+u_{}+k|G|(u_+^2u_{}^2),`$ (44) $`E_k^2`$ $`+`$ $`|F_k|^2=m_{\mathrm{eff}}^2+|G|^2k^2.`$ (45) The Hamiltonian can be diagonalized with the help of a Bogolyubov transformation. The new set of creation and annihilation operators which diagonalize the Hamiltonian can be defined as $`\widehat{a}(k,\tau )=\alpha _k(\tau )a(k)+\beta _k(\tau )b^{}(k),`$ (46) $`\widehat{b}^{}(\tau )=\beta _k^{}(\tau )a(k)+\alpha _k^{}(\tau )b^{}(k),`$ (47) where $`\alpha _k`$, and $`\beta _k`$ are the normalized Bogolyubov coefficients, $`{\displaystyle \frac{\alpha _k}{\beta }}`$ $`=`$ $`{\displaystyle \frac{E_k+\omega }{F_k^{}}},`$ (48) $`|\beta _k|^2`$ $`=`$ $`{\displaystyle \frac{|F_k|^2}{2\omega (\omega +E_k)}}={\displaystyle \frac{\omega E_k}{2\omega }},`$ (49) $`\omega ^2`$ $`=`$ $`m_{\mathrm{eff}}^2+|G|^2k^2.`$ (50) Now, the time dependent occupation number can be written in terms of the vacuum expectation value of the number operator $`n(\tau )=0|N|0={\displaystyle \frac{1}{\pi ^2a^3(\tau )}}{\displaystyle 𝑑kk^2|\beta _k|^2}.`$ (51) In order to solve Eq. (38), one needs to specify the boundary conditions. Usually they are defined such that at the beginning, when $`\tau 0`$, we have $`|\beta _k|^2=0`$, and the occupation number $`n(0)=0`$, suggesting that there is no particle density at the initial time: $`u_\pm (0)`$ $`=`$ $`\sqrt{{\displaystyle \frac{\omega m_{\mathrm{eff}}}{\omega }}},`$ (52) $`u_\pm ^{}(0)`$ $`=`$ $`im_{\mathrm{eff}}u_\pm (0)+ik|G(0)|u_{}(0).`$ (53) Now, we have all the tools necessary to study the production of gravitinos in the hybrid inflationary model. Next we describe the hybrid model and how to estimate the occupation number. ## IV Hybrid Inflation The hybrid inflation potential can be derived from the following superpotential $`W=\lambda \varphi (N^2N_0^2),`$ (54) where $`\varphi `$ plays the role of inflaton. During inflation the other field $`N`$ is trapped in its false vacuum, $`N=0`$, while the $`\varphi `$ field rolls down towards the critical value determined by the value of $`N`$ at the global minimum, $`N_0`$, and the coupling constant $`\lambda `$ between $`\varphi `$ and $`N`$. At this point the $`N`$ field rolls down from its zero value towards the global minima and begins to oscillate around $`N_0`$, while $`\varphi `$ oscillates around zero. This will enable the preheating phase of the Universe. During this period an effective potential for the fields $`\varphi `$ and $`N`$ can be derived $`V`$ $`=`$ $`|W_\varphi |^2+|W_N|^2,`$ (55) $`=`$ $`\lambda ^2(N^2N_0^2)^2+4\lambda ^2\varphi ^2N^2,`$ (56) where the subscripts denote the derivative of $`W`$ with respect to the fields. The superpotential in Eq. (54) also ensures a non-vanishing constant vacuum energy during inflation $`V(0)=\lambda ^2N_0^4`$. We should also mention that $`\lambda `$ and $`N_0`$ act as free parameters of the model, but they are constrained to some extent from the COBE normalization and the tilt in the power spectrum $`\lambda N_01.27\times 10^{15}|\eta _{}|GeV,`$ (57) where we have taken$`\eta _{}`$ is one of the two slow-roll parameters, usually defined in inflationary cosmology as $`\eta _{}=(m_P^2/(8\pi ))(V^{\prime \prime }/V(0))`$; the “star” index means that it is evaluated at least 60 e-foldings before the end of inflation. This should not be confused with the goldstino; we have already defined goldstino with the same notation earlier. $`|\eta _{}|0.01`$ in our analysis, which is a reasonable assumption in order not to generate a sharp tilt in the power spectrum of the density perturbation during inflation. The present constraint on the spectral index is $`|n1|<0.2`$ . Hybrid inflationary model derived from such a superpotential is known as F-term hybrid inflation. Slightly different version of hybrid inflation popularly known as D-term inflation can be derived from the Fayet-Illiopoulus term appearing from an anomalous $`U(1)`$ symmetry, which could provide the necessary potential energy during inflation. Whatsoever be the cause of such a potential, our argument for the gravitino production is quite generic and will not depend upon a particular origin of the vacuum energy. Due to presence of a single mass scale $`\lambda N_0`$ in the model, which is related to the supersymmetric breaking scale during the inflationary era, we will have a single frequency of oscillations during the preheating phase, and effectively a single scalar field oscillating. In our case this can be taken as $`N`$, related to the other field $`\varphi `$ by $`\varphi ={\displaystyle \frac{N_0N(t)}{\sqrt{2}}}.`$ (58) By solving the equation of motion for $`N(t)`$ we get the approximate solution when the field is around the bottom of the potential $`{\displaystyle \frac{N(t)}{N_0}}1+\mathrm{\Sigma }(t)\mathrm{cos}(\nu _\varphi t),`$ (59) where $`\nu _\varphi =2\lambda N_0`$ provides the natural frequency of the oscillations, and $`\mathrm{\Sigma }(t)`$ is the amplitude of the oscillations decreasing in time as $`1/t`$ The actual decrease in amplitude over an oscillation period depends on the ratio $`H/\nu _\varphi `$, $`H`$ being the Hubble parameter.. The above expression is valid when $`|N(t)/N_01|1/3`$, with $`\mathrm{\Sigma }(0)1/3`$. What makes the hybrid scenario interesting is that $`N(t)`$ never vanishes and that causes the fermionic creation to be completely different than the usual chaotic type inflationary potentials. Now, with the present knowledge we can evaluate $`|G|`$ with the help of Eq. (29) $`|G|^2`$ $`=`$ $`{\displaystyle \frac{(\rho 2V)^2}{\rho ^2}}+{\displaystyle \frac{4|\dot{W}|^2}{\rho ^2}},`$ (60) $`=`$ $`14{\displaystyle \frac{|W_\varphi \dot{N}W_N\dot{\varphi }|^2}{\rho ^2}},`$ (61) $`=`$ $`1{\displaystyle \frac{4\lambda ^2N_0^2}{\rho ^2}}|\dot{N}|^2(NN_0)^2,`$ (62) where the subscript in $`W_\varphi `$, $`W_N`$ means derivative with respect to the corresponding field, and dot denotes as usual physical time derivative. The last equation has been written using Eq. (58), knowing that $`|\dot{\varphi }|^2+|\dot{N}|^2=3/2|\dot{N}|^2`$. Here it is important to notice that the departure from $`1`$ is quite obvious, even though there is effectively a single scalar field oscillating. In fact, we can easily generalize Eq. (61) to any arbitrary number of scalar fields $`\varphi _i`$ contributing to the energy density. Using Eqs. (24-28), without assuming $`\varphi _i1`$, we get $$|G|^2=1\frac{4e^K}{(\rho +3m^2)^2}\underset{i<j}{}|𝒟_iW\dot{\varphi }_j𝒟_jW\dot{\varphi }_i|^2,$$ (63) where $`𝒟_iW=W_i+K_iW`$, and the subscript $`i`$ denotes derivative with respect to the homogeneous scalar field $`\varphi _i`$. This shows explicitly that $`|G|1`$. ## V Analytical estimation of occupation number ### A Helicity $`1/2`$ It is possible to analytically estimate the occupation number, and the number density $`n(t)`$ for both the helicities. Assuming that $`|\beta _k|^21`$, for a given momentum $`k`$, our task reduces to estimate the cut-off momentum $`k`$. For this we need to solve Eq. (38), which will then lead us to study the occupation number for the helicity $`1/2`$ gravitinos in hybrid model. In the case of helicity $`1/2`$ states, the contribution of $`|G|^2`$ can be important, and we need to evaluate it before we could estimate the abundance of helicity $`1/2`$ states. Later on we will discuss the abundance of helicity $`3/2`$ states, where $`|G|^2=1`$. To carry out our calculation we need to know the dominant contribution to $`\mathrm{\Omega }`$ appearing in Eq. (38), $`\mathrm{\Omega }`$ $`=`$ $`ma+\mu +i{\displaystyle \frac{\alpha }{2}},`$ (64) where $`\mu `$ $`=`$ $`{\displaystyle \frac{A^{}}{2B}}+{\displaystyle \frac{A}{2B}}{\displaystyle \frac{|G|^{}}{|G|}},`$ (65) $`\alpha `$ $`=`$ $`{\displaystyle \frac{|G|^{}}{|G|}}.`$ (66) With the help of Eq. (59) and Eq. (60) it is possible to evaluate all the terms in Eq. (64). Here we simply quote the final results in terms of the physical time: $`m(t)`$ $`=`$ $`{\displaystyle \frac{\nu _\varphi ^3\mathrm{\Sigma }^2(t)}{4\sqrt{2}\lambda ^2M_\mathrm{p}^2}}\left(1+\mathrm{cos}(2\nu _\varphi t)\right)+𝒪(1/t^3),`$ (67) $`|G|^2`$ $`=`$ $`1{\displaystyle \frac{1}{9}}\mathrm{sin}^2(2\nu _\varphi t)+𝒪(1/t),`$ (68) $`{\displaystyle \frac{|\dot{G}|}{|G|}}`$ $`=`$ $`{\displaystyle \frac{\nu _\varphi }{9}}\mathrm{sin}(4\nu _\varphi t)\left(1+{\displaystyle \frac{\mathrm{sin}^2(2\nu _\varphi t)}{9}}\right)+𝒪(1/t),`$ (69) $`{\displaystyle \frac{\dot{A}}{2B}}`$ $`=`$ $`{\displaystyle \frac{3\nu _\varphi }{2\sqrt{2}}}\left(1+{\displaystyle \frac{5\mathrm{\Sigma }(t)}{4}}\mathrm{cos}(\nu _\varphi t)\right)+𝒪(1/t^2),`$ (70) $`{\displaystyle \frac{A}{2B}}{\displaystyle \frac{|\dot{G}|}{|G|}}`$ $`=`$ $`{\displaystyle \frac{\nu _\varphi }{6\sqrt{2}}}\mathrm{cos}^2(2\nu _\varphi t)\left(1+{\displaystyle \frac{\mathrm{sin}^2(2\nu _\varphi t)}{9}}\right)+𝒪(1/t).`$ (71) It is apparent that the only term which dominates $`\mathrm{\Omega }(t)`$ in Eq. (64) is due to Eq. (70). The mass term $`m(t)`$ is subdominant due to the Planck suppression ( here we have explicitly written the Planck mass). The amplitude of the oscillations from Eq. (71) is one order of magnitude smaller compared to that in Eq. (70). As a first approximation we can consider $`\dot{A}/(2B)`$ to be the effective mass term for the helicity $`1/2`$ states. A general discussion of production of massive fermions in an oscillatory background is given in the Appendix. We will use the results quoted there in order to estimate the occupation number for the helicity $`1/2`$ gravitinos. By taking $`\mathrm{\Omega }(t)\dot{A}/(2B)`$, and $`|G|^21`$, we notice that Eq. (38) mimics Eq. (77) in the appendix for a time-varying mass which is denoted by $`\mathrm{\Omega }(t)`$ in our present situation. First of all we notice that the effective mass term $`\dot{A}/(2B)`$ never vanishes at any point of the scalar field oscillations. With $`\mathrm{\Sigma }(t)1/3`$, the effective mass is always positive. This situation is typical of any hybrid model. Comparing Eq. (70) with Eq. (78) in the Appendix, we have $`\nu _\varphi =m_\varphi `$, $`m_X=3\nu _\varphi /(2\sqrt{2})`$ and $`g\varphi (0)=5\nu _\varphi /(8\sqrt{2})`$. It is evident that $`m_X>g\varphi (0)`$, and as a result $`m(t)`$ is always positive, quite similar to the case of bosonic production, where the mass term appears always as squared. However, when studying fermionic creation for a chaotic type potential there is a possibility of having $`m(t)`$ vanishing at some point during the oscillations . In this case fermion production takes place in the first few oscillations, since the adiabatic condition is violated maximally when the inflaton field passes through the point where the effective mass vanishes, and it is possible to create very heavy fermions in the process. In hybrid models and for the helicity 1/2 gravitinos, because the effective mass remains always positive, the adiabatic condition is broken when the effective mass reaches its $`minimal`$ (non zero) value, and the mass of the fermions created never exceeds the mass of inflaton. In this case it is also important to note that the amplitude of the oscillations decays slowly and it takes roughly $`2030`$ oscillations to make any significant change in the initial amplitude. As a result the production of fermions takes place at each and every oscillation, and since the degree of violation of the adiabaticity is much weaker compared to the former, the production process takes a longer time to saturate the Fermi band. This is quite evident from our numerical results, see Fig. (1) and Fig. (2). The choice of model parameters in Fig. (1) leads to a higher inflationary scale compared to that of Fig. (2). However, the rate of production is exactly the same in both the cases. The reason is that the occupation number essentially depends on $`\nu _\varphi `$, which is exactly the same in both the models we have considered. This can be vividly seen by comparing the number of peaks in the occupation number. In Fig. 2 it is easy to see the particle creation taking place in bursts and in regular intervals. The peaks remind us the violation of the adiabatic condition. This is a perfect example of particle creation in a broad resonance regime. In the narrow resonance, particles are always created throughout the evolution of the scalar field , they are not produced in bursts like in Fig. 2. It is visible that the particle number remains constant for some time and then there is a sudden jump in the occupation number. Physically one can understand the situation from the comparison with an interacting quantum field theory. Usually in field theory, at extreme past/future we consider the plane wave solution of the free equation of motion. However, in between extreme past and future the interaction is switched on. Here also as we can see that the occupation number remains constant for a while, which represents the ingoing wave and then the interaction switches on, which is depicted by a jump in the occupation number. The outgoing wave in this picture is the newly filled occupation number. In fact the Bugolyubov coefficients can be calculated by estimating the reflection and the transmission coefficients of the plane wave passing through a barrier. Most of the important information concerning the production of helicity $`1/2`$ gravitinos can be extracted from Fig. 2. The production of helicity $`1/2`$ gravitinos builds up in each and every oscillation. The production does not saturate the Fermi level in the first few oscillations, but it takes several oscillations to reach the Fermi level. This behaviour is in stark contrast to the fermionic creation in a quadratic potential , where the Fermi level is reached in a few oscillations. It is also noticeable that there is no stochastic behaviour in the occupation number, and it is always increasing until it reaches the Fermi level. The main reason is that the effect of expansion is felt very slowly in hybrid models, when compared to the typical period of oscillation, that is, $`H_0/\nu _\varphi N_0/M_P1`$, where $`H_0`$ is the value of the Hubble constant at the end of inflation. Therefore, the amplitude decreases very slowly in most of the cases, almost adiabatically. There is also a slight change in frequency, as it can be noticed from the plot (for details see Ref. ). However this small change in frequency is not going to affect our analytical estimation. We have plotted the spectrum for the helicity $`1/2`$ gravitinos in Figs. (3) and (4), for two different set of model parameters. In both the cases Fermi-level is saturated for $`k_{\mathrm{max}}\nu _\varphi `$. This confirms our analytical study in the preheating section, see Eq. (91), which reduces to numerically observed $`k_{\mathrm{max}}`$ for $`m_Xm_\varphi `$. The number density of helicity $`1/2`$ can be obtained from Eq. (51) $`n_{1/2}{\displaystyle \frac{1}{4\pi ^3}}{\displaystyle d^3kn(k)}{\displaystyle \frac{k_{\mathrm{max}}^3}{3\pi ^2}}{\displaystyle \frac{\nu _\varphi ^3}{3\pi ^2}}.`$ (72) Here $`k_{\mathrm{max}}`$ has been taken to be a comoving momentum. It is evident from Fig. (1) and Fig. (2) that the occupation number grows gradually and saturates after many oscillations depending on the choice of $`\lambda `$ and $`N_0`$. ### B Helicity $`3/2`$ We can follow similar arguments to evaluate the occupation number for helicity $`3/2`$ gravitinos. Notice that Eq. (22) for helicity $`3/2`$ reduces exactly to Eq. (38) with $`|G|=1`$ and $`\mathrm{\Omega }=m_{\mathrm{eff}}=ma`$, . Hence, the single time-varying mass scale appearing in the problem is now given by Eq. (67). Comparing Eq. (67) with Eq. (78), it becomes clear that the bare mass $`m_X`$ itself is time-varying for the helicity $`3/2`$ case, but as we have noted before, the amplitude of the oscillations is almost constant, especially in the hybrid model we are interested in. The parameters $`m_X`$ and $`g\varphi (0)`$ are equal to each other, and both are Planck mass suppressed. But this does not mean that we cannot excite helicity 3/2 gravitinos. The main point is that the effective mass always vanishes at some point during each and every oscillation. As a matter of fact we can excite them precisely due to this. What matters is the violation of adiabaticity, and that takes place precisely at those points where $`m(t)`$ vanishes. Production of helicity $`3/2`$ mimics the first scenario we have discussed in the appendix, but now $`m_Xm_\varphi =2\nu _\varphi `$. However, since the effective mass $`m(t)`$ vanishes in each and every oscillation, so the gravitino production takes place in each and every oscillation. The spectrum of helicity $`3/2`$ gravitinos is plotted in Fig. (5), for the model parameters given in Fig. (3). The spectrum preserves the essential features obtained before, but most importantly Fermi level never gets saturated and the production is extremely subdominant compared to helicity $`1/2`$ case. Nevertheless, the helicity $`3/2`$ gravitino abundance may also pose a strong bound on the model parameters and the reheating temperature, and thus it is necessary to study them as well . Our task is to estimate $`k_{\mathrm{max}}`$, and with the help of Eq. (85) we get $`k_{\mathrm{max},3/2}^3{\displaystyle \frac{\nu _\varphi ^5\mathrm{\Sigma }^2(t)}{2\sqrt{2}\lambda ^2M_\mathrm{p}^2}}.`$ (73) By taking $`M_\mathrm{p}10^{18}`$ GeV, and $`\nu _\varphi 10^{14}`$ GeV, we get $`k_{\mathrm{max}}0.1\nu _\varphi `$, which matches very well with our numerical result. In Fig. (5), the spectrum peaks around $`0.3\nu _\varphi `$. It is straightforward to estimate the occupation number for the helicity $`3/2`$ gravitinos $`n_{3/2}{\displaystyle \frac{|\beta |^2\nu _\varphi ^5\mathrm{\Sigma }^2(t)}{6\sqrt{2}\pi ^2\lambda ^2M_\mathrm{p}^2}}.`$ (74) Here we have assumed $`\beta _k`$ in Eq.(51) to be a constant, $`0|\beta |^21`$, for the cut-off momentum $`k_{\mathrm{max}}`$. The actual value of $`\beta `$ is difficult to estimate analytically, and we do not attempt to analyse it here. We can also estimate the abundance ratio for the two helicities $`{\displaystyle \frac{n_{1/2}}{n_{3/2}}}={\displaystyle \frac{2\sqrt{2}\lambda ^2M_\mathrm{p}^2}{|\beta |^2\nu _\varphi ^2\mathrm{\Sigma }^2(t)}}.`$ (75) We should mention that in Eqs. (73-75), $`\mathrm{\Sigma }(t)`$ can be taken to be $`1/3`$, since the amplitude of the oscillations remains unchanged for many oscillations. For the numerical values we have considered, we would get the production of helicity $`1/2`$ gravitinos to be roughly $`4`$ orders of magnitude larger than that of the helicity $`3/2`$ states, if we naively assumed that helicity $`3/2`$ saturated the Fermi-level ($`|\beta |^2=1`$, which is an over estimated production). It is evident from the numerical example, that the production is suppressed by at least $`8`$ orders of magnitude compared to that of helicity $`1/2`$, see Fig. (3) and Fig. (5). ## VI Implications on $`T_{\mathrm{reh}}`$ Through parametric resonance, other particles can also be created, and especially if their couplings are not suppressed by the Planck mass, they are perhaps produced more abundantly than gravitinos. Assuming that the end of reheating gives rise to a thermal bath with a final temperature $`T_{\mathrm{reh}}`$, it is possible to estimate the ratio $`n/s`$, where $`s`$ is the entropy density, and $`n`$ represents the number density of gravitinos after the end of reheating. Both $`n`$ and $`s`$ scales like $`a^3`$, and we can also assume that the background scalar field density behaves on average like matter, $`\rho _\varphi a^3`$, during the oscillatory period. It is possible then to estimate the final ratio by noticing that $`\rho _i\lambda ^2N_0^4`$ and $`n(t_\mathrm{i})\lambda ^3N_0^3`$, where the subscript denotes the initial values $`{\displaystyle \frac{n}{s}}={\displaystyle \frac{n(t_\mathrm{i})}{\rho _i}}T_{\mathrm{reh}}{\displaystyle \frac{\lambda }{N_0}}T_{\mathrm{reh}}.`$ (76) Here the left hand side represents the final abundance of gravitinos during nucleosynthesis. Gravitinos are weakly coupled to gauge bosons and its gaugino partners and their lifetime, $`\tau _{\mathrm{decay}}M_\mathrm{p}^2/m_{3/2}^2`$, is very long. For a TeV mass gravitino it could be around $`10^410^5`$ seconds, and it would pose a genuine threat to nucleosynthesis. However, this statement is strictly correct only for the helicity $`3/2`$ component, since they can decay to gauge bosons and its gaugino partner through a dimension $`5`$ operator. At high energies the interaction channels are governed by $`3/2`$ component rather than $`1/2`$ component gravitinos. In particular, helicity $`3/2`$ may be produced with a mass close to a TeV range, so they decay very late and they are the ones which survive till late to cause problems for nucleosynthesis. However, the same can not be said with confident about the helicity $`1/2`$ components. As we have seen, the violation of conformal invariance is not the same for both the helicities, and the helicity $`1/2`$ gains an effective mass which is of the order of $`\nu _\varphi m_{3/2}`$. Essentially, helicity $`1/2`$ gravitinos are in an oscillatory scalar background with a frequency similar to their effective mass. There is no good reason to believe that the decay rate of helicity $`1/2`$ to gauge bosons and to gauge fermions would mimic the decay rate similar to that in a flat background. So far, a detailed study is lacking in this area, but there is sufficient hint that the decay rate of helicity $`1/2`$ is much smaller than the Hubble parameter. We do not repeat the argument, rather we refer the reader to Ref. . The detailed calculation of the decay rate seems to be quite involved and we leave that for our future investigation. The important point to realize is that once the Universe reheats and thermalizes, the effective mass of the helicity $`1/2`$ gravitinos becomes similar to that of the helicity $`3/2`$, and as a result the decay rate would essentially be given by the usual decay rate in a flat background limit. Whatsoever be the detailed analysis, we must mention that while deriving Eq. (76), we have implicitly assumed that the initial abundance of gravitinos produced remain frozen until thermalization. Since the decay rate is smaller than the Hubble rate, the gravitinos produced during preheating will be able to survive until the end of reheating. The final abundance will mainly depend on the model parameters such as $`\lambda `$ and $`N_0`$, see Eq. (76). Therefore, the constraint on the reheat temperature derived from Eq. (76) is clearly model dependent. We also mention that we have not included the effect of back-reaction coming from newly created bosons and fermions. In hybrid models, the production of the quanta associated with $`N`$ and $`\varphi `$ is very efficient and can take place just in the first few oscillations , much before the gravitinos have saturated the Fermi level. Therefore, back-reaction effects due to these quanta will quickly change the frequency and also the amplitude of the oscillating fields. We strongly suspect that especially in the hybrid scenario gravitino production will be affected due to such considerations and hence our current estimation of gravitino abundance will not hold anymore. However, even though the non-thermal production of gravitinos may stop after a while, there is a possibility to create them through scattering processes. This is beyond the scope of the present discussion and we shall hope to come back to these issues elsewhere . ## VII Discussion and conclusions We have carried out the calculation for the gravitino production in multi-chiral field scenario, in particular in the context of hybrid model. As we have shown, it is possible to add a gauge-fixing term to the supergravity Lagrangian to get rid of the mixing between the goldstino and the gravitino field. Our method is a generalization of $`R_\zeta `$ gauge studied in various contexts. We choose to work in a unitary gauge, in which the goldstino is completely removed from the physical spectrum. Our study emphasizes major points in the non-perturbative production mechanism of gravitinos, analytically and numerically for the multi-chiral field models. We also give detailed analysis of the fermionic creation in general. We have observed that the fermionic creation in hybrid model is quite different from other chaotic inflationary models. The effective fermionic mass in hybrid models never vanishes, and as a result the particle production does not take place in first few oscillations, rather it builds up gradually. This makes it more interesting as far as the gravitino production is concerned. If we really want nucleosynthesis to be preserved in the context of supergravity inflationary models, we believe models based on hybrid inflation with low scales are probably going to be the only saviour. The reason is very simple; in other models there is no way we can argue the back-reaction due to the creation of other particles would stop creating gravitinos, but in hybrid models there is a scope where the back-reaction due to non-perturbative creation of bosons could affect the coherent oscillations of the inflaton and halt the particle production completely. This gives us a new hope to understand the abundance of gravitinos during nucleosynthesis and we leave these important issues to be investigated in near future. ## Acknowledgements A.M. is supported by Inlaks foundation and the authors are thankful to Lev Kofman, David Lyth, Antonio Lopez Maroto, and Subir Sarkar for many useful suggestions. ## Estimation of $`k_{\mathrm{max}}`$ In this section we briefly discuss non-perturbative production of massive fermions. This discussion is general and self sufficient. We begin with the second order Dirac equation represented in terms of the two mode functions. It reads in conformal time $`u_\pm ^{\prime \prime }+\left(k^2+(ma)^2\pm i(ma)^{}\right)u_\pm =0,`$ (77) where $`u_\pm `$ are the mode functions, $`k`$ is the momentum, prime denotes the derivative with respect to the conformal time, $`a`$ denotes the scale factor, and $`m`$ is an effective time-dependent mass. This mass can be written in terms of the physical time as $`m(t)`$ $`=`$ $`m_X+g\varphi (t),`$ (78) $`\varphi (t)`$ $`=`$ $`\mathrm{\Phi }(t)cos(m_\varphi t),`$ (79) where $`m_X`$ is the mass of the fermion we are interested in, $`\varphi (t)`$ is the scalar field oscillating with some initial amplitude $`\mathrm{\Phi }(0)`$, and $`m_\varphi `$ is its mass. The Yukawa coupling between the fermion and the scalar is determined by $`g`$. Inspecting the mode equation Eq. (77), it is obvious that it mimics harmonic oscillator with a time-varying imaginary frequency $`\omega ^2=k^2+(ma)^2\pm i(ma)^{}.`$ (80) Particle creation occurs due to non-adiabatic evolution of the total frequency $`\omega (t)`$, that is, whenever $`{\displaystyle \frac{d\omega (t)}{dt}}\omega ^2,`$ (81) where we have expressed the total frequency in physical time. We shall explore here two possible scenarios for fermionic production, depending on whether or not the effective mass for the fermion vanishes at some point of the scalar field oscillations. To start our discussion, we consider first the situation when the amplitude of the scalar field initially satisfies $`|\mathrm{\Phi }(0)|>m_X/g`$, so that the effective mass goes through zero at some point during the oscillations. Since the particle production takes place within a time $`\mathrm{\Delta }tH^1`$, we can safely neglect the effect of expansion during the production of particles, especially in a broad resonance regime<sup>§</sup><sup>§</sup>§In the hybrid model scenario presented in this paper, it is evident from Fig.(2) that the gravitino production is taking place in bursts within a short interval of time. However, there is another possibility, a narrow resonance regime, where the particle production takes place through out the evolution of the scalar field .. However, we have to keep in mind that the amplitude is gradually decreasing due to the effect of expansion. Once the amplitude drops below a critical value, the production of the fermions will soon stop and their number density will be frozen in time . The critical amplitude is given approximately by equating Eq.(78) to zero $`\mathrm{\Phi }_{}{\displaystyle \frac{m_X}{g}}.`$ (82) Below this value, the effective mass remains always positive. It is also important to mention that the production enhances near the regime when the effective mass vanishes, or, in other words, when there is a maximum violation of adiabaticity condition. With the above information it can be possible to estimate the typical momentum $`k`$ required to violate the adiabatic condition when the amplitude of the field is close to $`\mathrm{\Phi }_{}`$. This will occur when $`\mathrm{cos}(m_\varphi t_{})m_X/(g\mathrm{\Phi }_{})`$, or $`t_{}\pi /m_\varphi `$. Our condition Eq. (81) implies $`2g^2\dot{\varphi }(t)^2\left(m_X+g\varphi (t)\right)^2+{\displaystyle \frac{1}{2}}g^2\ddot{\varphi }(t)^2`$ (83) $`\left((k^2+(m_X+g\varphi (t))^2)^2+g^2\dot{\varphi }(t)^2\right)^{3/2}.`$ (84) Since, $`\dot{\varphi }(t_{})=0`$, and $`\ddot{\varphi }(t_{})m_\varphi ^2m_X/g`$, we can estimate the left-hand side of Eq. (83) around $`\varphi (t_{})`$, and the final condition translates to a simpler form, $`{\displaystyle \frac{1}{2}}m_Xm_\varphi ^2k_{\mathrm{max}}^3.`$ (85) We have assumed $`\omega k_{\mathrm{max}}`$ in the final derivation. This result confirms similar result already obtained in Ref. . We can express Eq. (85) in terms of an effective $`q`$ parameter $`k_{\mathrm{max}}\left({\displaystyle \frac{m_\varphi ^4}{m_X}}q_{}\right)^{1/3},`$ (86) where $`q(t)=g^2\mathrm{\Phi }^2(t)/m_\varphi ^2`$. In the above equation the value of $`q`$ is evaluated at $`\mathrm{\Phi }\mathrm{\Phi }_{}`$. The $`q`$ dependence in $`k_{\mathrm{max}}`$ is quite different from the bosonic production. This is mainly due to the presence of the imaginary part of the frequency, which has a significant contribution to the violation of the adiabatic condition. At this point one may be able to estimate the maximum mass $`m_X`$ allowed by the violation of the adiabatic condition. With the help of the $`q`$ parameter, it is possible to re-express $`m(t)`$ as $`m(t)=m_X+{\displaystyle \frac{\sqrt{q(0)}}{t}}\mathrm{cos}(m_\varphi t),`$ (87) where we have written explicitly the time-dependence of the amplitude. Hence, the maximum mass is achieved when $`t_{}=\pi /m_\varphi `$, and this gives $`m_X{\displaystyle \frac{m_\varphi }{\pi }}\sqrt{q(0)}.`$ (88) For reasonable values of the coupling constant $`g`$, it is possible to achieve very high values of the fermion mass $`m_Xm_\varphi `$. This suggests that such production of supermassive fermions is indeed non-thermal and non-perturbative in nature. It is also important to notice that for values of $`q`$ which are of the order of tens, the maximum fermionic mass obtained is of the order of the mass of the oscillating field. The above analysis suggests that creating super-massive fermions is possible because the effective mass $`m(t)`$ vanishes around $`\varphi _{}`$, and this is where the adiabatic condition is violated maximally. This happens quite naturally in quadratic inflationary potentials because the initial amplitude of the oscillations for the inflaton is large enough to pass through the point where the effective mass $`m(t)`$ vanishes . A similar situation arises in the case of the helicity $`3/2`$ states for the gravitino, and we find that its mass term Eq. (67) vanishes in each and every oscillation of the $`N`$ field, defined earlier. This means that the adiabatic condition is violated maximally at those points. However, in the hybrid case the amplitude of the oscillations and the mass of the helicity $`3/2`$ states are exactly the same $`m_X=g\mathrm{\Phi }(t)`$. This suggests that we can not create very massive $`3/2`$ states, although we can create them with mass $`m_Xm_\varphi `$. Therefore, the hybrid situation is slightly different from the chaotic inflationary scenario with quadratic potential. Next, we study a scenario where the effective mass of the fermion never vanishes at any point, that is $`g\mathrm{\Phi }(0)<m_X`$. This is the situation which arises in the hybrid inflationary scenario for the helicity $`1/2`$ gravitino states. In this case the adiabatic condition is violated maximally near the point where $`m(t)`$ is minimal but non zero ( then $`|g\varphi (t)|`$ is maximal), and this again happens when $`\dot{\varphi }(t)0`$. Using Eq. (81), we can then easily estimate the upper limit on $`|\varphi (t)|`$ $`\left({\displaystyle \frac{m_\varphi ^2g\varphi (t)}{\sqrt{2}}}\right)^{2/3}(m_X|g\varphi (t)|)^2k^2,`$ (89) where we have replaced $`\ddot{\varphi }(t)=m_\varphi ^2\varphi (t)`$. Notice that we have explicitly taken the negative sign for $`\varphi (t)`$, in order to extremise the left-hand side of the above equation. We would like to see the range of the violation of the adiabatic condition for small $`k`$, $`{\displaystyle \frac{m_X}{3}}\stackrel{<}{_{}}|g\varphi (t)|\stackrel{<}{_{}}m_X.`$ (90) Hence, in general the production could continue as long as the field amplitude follows the above condition Eq. (90). The maximal range of momenta for which the fermions are produced is obtained when $`|g\varphi (t)|m_X`$, and is given by $`\left({\displaystyle \frac{m_\varphi ^2m_X}{\sqrt{2}}}\right)^{1/3}k_{\mathrm{max}},`$ (91) and for $`m_Xm_\varphi `$, the above expression reduces to $`k_{\mathrm{max}}m_X`$. This result could have been easily derived from Eq. (85) by taking the masses to be almost equal. In the hybrid model, which we have considered, the amplitude of the oscillations is small, never exceeding $`m_X/g`$. Hence we do not expect to produce super-massive fermions. The above result can be directly applied to the production mechanism of helicity $`1/2`$ gravitinos. We have mentioned in the main section that the effective time-varying mass for the helicity $`1/2`$ states does not vanish. The effective mass of the $`1/2`$ states is almost equal to that of the oscillating frequency, which corresponds to having $`m_Xm_\varphi `$ in our present discussion. In addition the adiabatic condition is broken for a narrow range of field values, see Eq. (90). This means that the occupation number for the helicity $`1/2`$ builds up in each and every oscillation and the gravitinos are produced in bursts.
warning/0002/hep-ph0002148.html
ar5iv
text
# I Introduction ## I Introduction The main goal of present and future experiments with ultrarelativistic heavy ions is to produce and study in the laboratory a new state of strongly interacting matter, the Quark-Gluon Plasma (QGP). In achieving this goal one is facing two major problems. First, the phase structure of QCD is not fully understood yet. Second, the matter evolution in the course of an ultrarelativistic heavy–ion collision may be out of thermodynamical equilibrium. Most calculations of the QCD phase diagram are made under the assumption of thermal and chemical equilibrium. The QCD lattice calculations at zero baryon chemical potential reveal a second order phase transition or a rapid crossover at temperatures around 140–160 MeV. Recent calculations based on different effective models show the possibility of a first order phase transition at finite baryon densities and moderate temperatures. The predicted phase diagram in the $`T\mu `$ plane contains a first order transition line terminating at a critical point $`(T_c,\mu _c)`$ with $`T_c`$ 120 MeV and a finite $`\mu _c`$ . Possible signatures of this point in heavy–ion collisions were discussed recently in Ref. . The problem is, however, that the matter produced in central heavy–ion collisions at very high energies has rather low net baryon density. If such a baryon–free matter would expand following (locally) an equilibrium path, it would miss the first order transition line. In this case no clear signatures of the phase transition would be observed. Moreover, it is most likely that the matter evolution in ultrarelativistic heavy–ion collisions does not follow thermodynamical equilibrium. The reason is that the matter produced in such collisions expands very fast. As has been already observed in heavy–ion experiments at the SPS energies (see e.g. ), the expansion velocities along the beam direction are close to the speed of light and the transverse velocities are close to $`0.5c`$ . A strong collective expansion of matter is expected also at RHIC and LHC energies. Under such conditions one may expect significant deviations from thermodynamical equilibrium , especially from the chemical equilibration. This may happen on both the partonic and the hadronic stages of the reaction. One can mention at least two mechanisms which may lead to the deviation from chemical equilibrium on the partonic stage. First, it is believed that multiple color strings are produced initially in hard nucleon–nucleon collisions. Later on they decay via the Schwinger mechanism into quark–antiquark pairs whose abundances are determined by quark masses and a string tension constant. At this stage, multiplicities of secondary quarks and antiquarks may be different from their values in thermodynamical equilibrium. Second, simple fits of QCD lattice data indicate that gluons may acquire a large effective mass around the deconfinement transition point. Therefore, even if the ideal QGP were created at some intermediate stage of the reaction, gluons would subsequently decay into lighter quark–antiquark pairs. Hence, the abundances of different quark species may deviate strongly from their equilibrium values. In particular, an overpopulation of the light quark–antiquark pairs may be expected. The phase diagram of such a chemically nonequilibrated matter may be very different from the predictions based on the equilibrium concepts. Another motivation to study chemically nonequilibrated quark–antiquark systems comes from the hadronic spectroscopy. It is well known that some mesonic resonances do not fall into the classification scheme based on the constituent quark model. They cannot be interpreted as conventional $`q\overline{q}`$ bound states. Rather they could be either bag–like $`(qq\overline{q}\overline{q})`$ states or meson–meson bound states $`(q\overline{q}q\overline{q})`$. Well known examples include the $`f_0(980)`$ ($`K\overline{K}`$ bound state close to the threshold), the $`f_1(1420)`$ ($`K\overline{K}^{}`$), the $`f_0(1500)`$ and $`f_2(1565)`$ ($`\omega \omega `$ and $`\rho \rho `$) etc. . A legitimate question is: what will happen if more and more $`q\overline{q}`$ pairs will be put together? Such multi-$`q\overline{q}`$ systems might be even more bound due to the reduced surface energy as compared with the bulk one. The mesonic substructure will most likely melt away and such states will look like multi-$`q\overline{q}`$ bags. We call this hypothetical state of matter as “meso-matter” and its finite droplets at “mesoballs”. The analogous state of hadronic matter, bound multipion droplets, has been considered in Ref. . Since the direct application of QCD at moderate temperatures and nonzero chemical potentials is not possible at present, more simple effective models respecting some basic symmetry properties of QCD are commonly used. One of the most popular models of this kind, which is dealing with constituent quarks and respects chiral symmetry, is the Nambu–Jona-Lasinio (NJL) model . In recent years this model has been widely used for describing hadron properties (see reviews ), phase transitions in dense matter and multiparticle bound states . In the previous paper we have used the NJL model to study properties of the quark–antiquark plasma out of chemical equilibrium. In fact, we considered a system with independent densities of quarks and antiquarks. To our surprise, we have found not only first order transitions but also deep bound states even in the baryon–free matter with equal densities of quarks and antiquarks. The consideration in Ref. was limited to systems composed of either light ($`u,d`$) or strange ($`s`$) quarks and antiquarks. In the present paper we extend the model to arbitrary mixtures of light and strange quarks. The emphasis is put on investigating the possibility of bound states in such systems at various flavour compositions of quarks and antiquarks. As a special case we consider the baryon–rich quark matter in chemical equilibrium, in particular, the possibility of bound states in strange quark matter, i.e. strangelets. Thermal properties of meso-matter and strange quark matter are also studied. The paper is organized as follows: in Sect. II a generalized NJL model including flavour–mixing terms is formulated in the mean–field approximation. Then in Sect. III the model is used to study the bound states in $`q\overline{q}`$ systems with different strangeness contents. The model predictions for strangelets are discussed in Sect. IV. The characteristics of bound states at zero temperature are summarized in Sect. V. Effects of finite temperatures are considered in Sect. VI. Possible decay modes of new bound states are discussed in Sect. VII. Main results of the present paper are summarized in Sect. VIII. ## II Formulation of the model We proceed from the SU(3)–flavour version of the NJL model suggested in Ref. . The corresponding Lagrangian is written as ($`\mathrm{}=c=1`$) $`=\overline{\psi }(i/\widehat{m}_0)\psi `$ $`+`$ $`G_S{\displaystyle \underset{j=0}{\overset{8}{}}}\left[\left(\overline{\psi }{\displaystyle \frac{\lambda _j}{2}}\psi \right)^2+\left(\overline{\psi }{\displaystyle \frac{i\gamma _5\lambda _j}{2}}\psi \right)^2\right]`$ (1) $``$ $`G_V{\displaystyle \underset{j=0}{\overset{8}{}}}\left[\left(\overline{\psi }\gamma _\mu \frac{\lambda _j}{2}\psi \right)^2+\left(\overline{\psi }\gamma _\mu \frac{\gamma _5\lambda _j}{2}\psi \right)^2\right]`$ (2) $``$ $`K\left[\mathrm{det}_f\left(\overline{\psi }(1\gamma _5)\psi \right)+\mathrm{det}_f\left(\overline{\psi }(1+\gamma _5)\psi \right)\right].`$ (3) Here $`\psi `$ is the column vector consisting of three single–flavour spinors $`\psi _f`$, $`f=u,d,s`$, $`\lambda _1,\mathrm{},\lambda _8`$ are the SU(3) Gell-Mann matrices in flavour space, $`\lambda _0\sqrt{2/3}𝑰`$, and $`\widehat{m}_0=\mathrm{diag}(m_{0u},m_{0d},m_{0s})`$ is the matrix of bare (current) quark masses. At $`\widehat{m}_0=0`$ this Lagrangian is invariant with respect to $`\mathrm{SU}_\mathrm{L}(3)\mathrm{SU}_\mathrm{R}(3)`$ chiral transformations. The second and third terms in Eq. (3) correspond, respectively, to the scalar–pseudoscalar and vector–axial-vector 4–fermion interactions. The last 6–fermion interaction term breaks the $`U_A(1)`$ symmetry and gives rise to the flavour mixing effects. In particular, this term is responsible for the large $`\eta ^{}`$ mass . In the mean–field approximation the Lagrangian (3) is reduced to $`_{\mathrm{mfa}}`$ $`=`$ $`{\displaystyle \underset{f}{}}\overline{\psi }_f(iD/m_f)\psi _f`$ (4) $``$ $`{\displaystyle \frac{G_S}{2}}{\displaystyle \underset{f}{}}\rho _{Sf}^2+{\displaystyle \frac{G_V}{2}}{\displaystyle \underset{f}{}}\rho _{Vf}^{}{}_{}{}^{2}+4K{\displaystyle \underset{f}{}}\rho _{Sf},`$ (5) where $`D/=/+i\gamma _0G_V\rho _{Vf}`$ and $`\rho _{Sf}`$ $`=`$ $`<\overline{\psi }_f\psi _f>,`$ (6) $`\rho _{Vf}`$ $`=`$ $`<\overline{\psi }_f\gamma _0\psi _f>`$ (7) are scalar and vector densities of quarks with flavour $`f`$ . Angular brackets correspond to the quantum–statistical averaging. The constituent quark masses, $`m_f`$, are determined by the coupled set of gap equations $$m_f=m_{0f}G_S\rho _{Sf}+2K\underset{f^{}f}{}\rho _{Sf^{}}.$$ (8) The NJL model is non-renormalizable, because its coupling constants have non-trivial dimensions: $`G_S,G_V[\mathrm{mass}]^2`$ and $`K[\mathrm{mass}]^5`$. As a consequence, the contribution of negative energy states of the Dirac sea are divergent, and one must introduce an ultraviolet cut–off. In this respect the NJL model is an effective model, aimed at describing the non-perturbative regime of QCD at low energies. Following common practice, we introduce the 3–momentum cut–off $`\mathrm{\Lambda }`$ to regularize divergent integrals. The structure of the fermionic vacuum within the NJL model is shown schematically in Fig. 1(a). Only “active” levels of the Dirac sea, i.e. with $`p<\mathrm{\Lambda }`$ are included in calculations. The model parameters $`m_{0f},G_S,K,\mathrm{\Lambda }`$ can be fixed by reproducing the observed masses of $`\pi ,K`$ , and $`\eta ^{}`$ mesons as well as the pion decay constant $`f_\pi `$. As shown in Ref. , a reasonable fit is achieved with the following values: $`m_{0u}=m_{0d}`$ $`=`$ $`5.5\mathrm{MeV},m_{0s}=140.7\mathrm{MeV},`$ (9) $`G_S=20.23\mathrm{GeV}^2,`$ $`K=155.9\mathrm{GeV}^5,\mathrm{\Lambda }=0.6023\mathrm{GeV}.`$ (10) Motivated by the discussions in in Refs. , we choose the following “standard” value of the vector coupling constant<sup>1</sup><sup>1</sup>1This value is somewhat different from the one used in our previous paper , although the ratio $`G_V/G_S`$ is the same. $$G_V=0.5G_S=10.12\mathrm{GeV}^2.$$ (11) It should be stressed that the vector and axial–vector terms cannot simply be ignored in the effective Lagrangian (3), as it is often done. These terms are necessary for correctly describing the vector meson properties , for adjusting the nucleon axial charge $`g_A`$ , etc. In context of the present study, the vector interaction is important, because it generates a net repulsive contribution in asymmetric matter, i.e. when the numbers of quarks and antiquarks are not equal. Let us consider homogeneous, thermally (but not, in general, chemically) equilibrated quark–antiquark matter at temperature $`T`$ . Let $`a_{𝒑,\lambda }`$ ( $`b_{𝒑,\lambda }`$) and $`a_{𝒑,\lambda }^+`$ ( $`b_{𝒑,\lambda }^+`$) be the destruction and creation operators of a quark (an antiquark) in the state $`𝒑,\lambda `$ , where $`𝒑`$ is the 3-momentum and $`\lambda `$ is the discrete quantum number denoting spin and flavour (color indices are suppressed). By using the plane wave decomposition of quark spinors in Eq. (5), it can be shown that quark and antiquark phase–space occupation numbers coincide with the Fermi–Dirac distribution functions: $`<a_{𝒑,\lambda }^+a_{𝒑,\lambda }>n_{𝒑f}`$ $`=`$ $`\left[\mathrm{exp}\left({\displaystyle \frac{E_{𝒑f}\mu _{Rf}}{T}}\right)+1\right]^1,`$ (12) $`<b_{𝒑,\lambda }^+b_{𝒑,\lambda ,f}>\overline{n}_{𝒑f}`$ $`=`$ $`\left[\mathrm{exp}\left({\displaystyle \frac{E_{𝒑f}\overline{\mu }_{Rf}}{T}}\right)+1\right]^1,`$ (13) where $`E_{𝒑f}=\sqrt{m_f^2+𝒑^2}`$ and $`\mu _{Rf},\overline{\mu }_{Rf}`$ denote the reduced chemical potentials of quarks and antiquarks: $`\mu _{Rf}`$ $`=`$ $`\mu _fG_V\rho _{Vf},`$ (14) $`\overline{\mu }_{Rf}`$ $`=`$ $`\overline{\mu }_f+G_V\rho _{Vf}.`$ (15) In our calculations we consider the chemical potentials $`\mu _f`$ and $`\overline{\mu }_f`$ as independent variables. The assumption of chemical equilibrium with respect to creation and annihilation of $`q\overline{q}`$ pairs would lead to the conditions $$\overline{\mu }_f=\mu _f,f=u,d,s.$$ (16) The explicit expression for the vector density can be written as $$\rho _{Vf}=\rho _f\overline{\rho }_f,$$ (17) where $$\rho _f=\nu \frac{\mathrm{d}^3p}{(2\pi )^3}n_{𝒑f},\overline{\rho }_f=\nu \frac{\mathrm{d}^3p}{(2\pi )^3}\overline{n}_{𝒑f}$$ (18) are, respectively, the number densities of quarks and antiquarks of flavour $`f`$ and $`\nu =2N_c=6`$ is the spin–color degeneracy factor. The net baryon density is obviously defined as $$\rho _B=\frac{1}{3}\underset{f}{}\rho _{Vf}.$$ (19) The physical vacuum ($`\rho _f=\overline{\rho }_f=0`$) corresponds to the limit $`n_{𝒑f}=\overline{n}_{𝒑f}=0`$ . Within the NJL model the energy density and pressure of matter as well as the quark condensates $`\rho _{Sf}`$ contain divergent terms originating from the negative energy levels of the Dirac sea. As noted above, these terms are regularized by introducing the 3–momentum cutoff $`\theta (\mathrm{\Lambda }|𝒑|)`$, where $`\theta (x)\frac{1}{2}(1+\mathrm{sgn}x)`$. Then the scalar density is expressed as $$\rho _{Sf}=\nu \frac{\mathrm{d}^3p}{(2\pi )^3}\frac{m_f}{E_{𝒑f}}\left[n_{𝒑f}+\overline{n}_{𝒑f}\theta (\mathrm{\Lambda }p)\right].$$ (20) The energy density and pressure are obtained in a standard way from the energy-momentum tensor corresponding to the Lagrangian (5). They can be decomposed into several parts as $`e`$ $`=`$ $`e_K+e_D+e_S+e_V+e_{FM}+e_0,`$ (21) $`P`$ $`=`$ $`P_K+P_D+P_S+P_V+P_{FM}+P_0,`$ (22) These expressions include: the “kinetic” terms $`e_K`$ $`=`$ $`\nu {\displaystyle \underset{f}{}}{\displaystyle \frac{\mathrm{d}^3p}{(2\pi )^3}E_{𝒑f}\left(n_{𝒑f}+\overline{n}_{𝒑f}\right)},`$ (23) $`P_K`$ $`=`$ $`{\displaystyle \frac{\nu }{3}}{\displaystyle \underset{f}{}}{\displaystyle \frac{\mathrm{d}^3p}{(2\pi )^3}\frac{𝒑^2}{E_{𝒑f}}\left(n_{𝒑f}+\overline{n}_{𝒑f}\right)},`$ (24) the “Dirac sea” terms $$e_D=P_D=\nu \underset{f}{}\frac{\mathrm{d}^3p}{(2\pi )^3}E_{𝒑f}\theta (\mathrm{\Lambda }p),$$ (25) the scalar interaction terms $$e_S=P_S=\frac{G_S}{2}\underset{f}{}\rho _{Sf}^{\mathrm{\hspace{0.17em}2}},$$ (26) the vector interaction terms $$e_V=P_V=\frac{G_V}{2}\underset{f}{}\rho _{Vf}^{\mathrm{\hspace{0.17em}2}}$$ (27) and the flavour mixing terms $$e_{FM}=P_{FM}=4K\underset{f}{}\rho _{Sf}.$$ (28) A constant $`e_0=P_0`$ is introduced in Eqs. (21) and (22) in order to set the energy density and pressure of the physical vacuum equal to zero. This constant can be expressed through the vacuum values of constituent masses, $`m_f^{\mathrm{vac}}`$, and quark condensates, $`\rho _{Sf}^{\mathrm{vac}}`$. These values are obtained by selfconsistently solving the gap equations (8) in vacuum, i.e. at $`n_{𝒑f}=\overline{n}_{𝒑f}=0`$ . For a system with independent chemical potentials for quarks ($`\mu _f`$) and antiquarks ($`\overline{\mu }_f`$) one can use the thermodynamic identity $$e=\underset{f}{}(\mu _f\rho _f+\overline{\mu }_f\overline{\rho }_f)P+sT.$$ (29) Then one can obtain the standard expression for the entropy density, $$s=_TP_K=\nu \underset{f}{}\frac{\mathrm{d}^3p}{(2\pi )^3}\left[n_{𝒑f}\mathrm{ln}n_{𝒑f}+(1n_{𝒑f})\mathrm{ln}(1n_{𝒑f})+n_{𝒑f}\overline{n}_{𝒑f}\right].$$ (30) By using Eqs. (8), (12)–(30) one can also show that the differential relation $$dP=\underset{f}{}\left(\rho _fd\mu _f+\overline{\rho }_fd\overline{\mu }_f\right)+sdT$$ (31) holds for any thermally (but not necessarily chemically) equilibrated process. ## III Symmetric quark–antiquark matter In this section we study the multi–quark–antiquark systems at zero temperature. Let us consider a symmetric system with equal numbers of quarks and antiquarks for each flavour. This requirement enforces the chemical potentials of quarks and antiquarks to be equal, $$\overline{\mu }_f=\mu _f.$$ (32) In this case the net vector density is automatically zero for each flavour, $`\rho _{Vf}=0`$. The net baryon number and electric charge are also zero. Such systems are especially interesting because the contribution of repulsive vector interaction, Eq. (27), vanishes in this case. One can view such systems as a compressed meson gas where mesons are melted to their elementary constituents, quarks and antiquarks. The single particle states of quarks and antiquarks in such a system are schematically shown in Fig. 1(b). To understand this picture one should simply realize that antiquarks are holes in the Dirac sea. Then one can imagine that a certain number of quarks from the negative energy states are collectively excited into the positive energy states, producing an equal number of holes. Due to the presence of valence quarks and antiquarks the mass gap will be reduced. In this situation one can expect the appearance of bound states. To characterize the flavour composition we introduce the strangeness fraction parameter $$r_s=\frac{N_s+N_{\overline{s}}}{N_u+N_{\overline{u}}+N_d+N_{\overline{d}}+N_s+N_{\overline{s}}},$$ (33) where $`N_{f(\overline{f})}`$ is the number of quarks (antiquarks) of flavour $`f`$ . For simplicity we consider only the isospin–symmetric mixtures where $`N_u=N_d`$ and $`N_{\overline{u}}=N_{\overline{d}}`$. Fig. 2 shows the energy per particle, $`ϵ=e/\rho _{\mathrm{tot}}`$, as a function of total density of valence quarks and antiquarks, $`\rho _{\mathrm{tot}}=\underset{f}{}(\rho _f+\overline{\rho }_f)`$, for different $`r_s`$. At low densities $`ϵ`$ tends to the sum of the constituent quark and antiquark masses in vacuum, weighted according to $`r_s`$, $$ϵ(\rho _{\mathrm{tot}}0,r_s)=m_q^{\mathrm{vac}}(r_s)=(1r_s)m_u^{\mathrm{vac}}+r_sm_s^{\mathrm{vac}}.$$ (34) With growing density, $`ϵ`$ first decreases due to the attractive scalar interaction. At higher densities $`ϵ`$ starts to increase, approaching slowly the limit of ideal ultrarelativistic Fermi gas. For each $`r_s`$ one can see the appearance of a nontrivial minimum corresponding to a bound multiparticle state with $`ϵ_{\mathrm{min}}(r_s)<m_q^{\mathrm{vac}}(r_s)`$. The density of $`q\overline{q}`$ pairs in these bound states varies between 0.7 and 1.4 fm<sup>-3</sup>, depending on $`r_s`$. It is easy to see that the minimum in $`ϵ`$ at any $`r_s`$ corresponds to zero pressure. Thus, finite droplets of such matter could be in mechanical equilibrium with vacuum. A more detailed behaviour of $`ϵ`$ in the plane $`\rho _{u+d}\rho _s`$ is shown in Fig. 3(a) (here $`\rho _{u+d}=\rho _u+\rho _d=2\rho _u`$). The condition of fixed $`r_s`$ corresponds to a straight line with slope $`r_s/(1r_s)`$ starting from the origin. By inspecting the figure one can notice the valley of local minima<sup>2</sup><sup>2</sup>2This valley is shown by the dotted line in Fig. 3(a). starting from $`ϵ=0.482`$ GeV per particle in the pure $`s\overline{s}`$ system ($`r_s=1`$) and descending to $`ϵ=0.304`$ GeV per particle in pure $`u\overline{u}+d\overline{d}`$ matter ($`r_s=0`$). There is no potential barrier on the way from $`r_s=1`$ to $`r_s=0`$. Thus, a droplet with any $`r_s0`$ will eventually “roll down” in the state with $`r_s=0`$. Possible decay modes of the considered bound states are discussed in Sect. VII. The constituent masses of $`u`$ and $`s`$ quarks as functions of $`\rho _{tot}`$ are shown in Fig. 4 for different $`r_s`$. As expected, the constituent masses decrease with density, signaling the gradual restoration of chiral symmetry. It is interesting to note that the constituent mass $`m_f`$ is more sensitive to the density of the same flavour $`f`$. For instance, in pure $`u\overline{u}+d\overline{d}`$ matter, i.e. at $`r_s=0`$, the $`s`$-quark mass does not change much with density. In this case $`m_s`$ varies entirely due to the flavour–mixing interaction. At densities corresponding to the bound states (indicated by dots on the respective curves) the constituent masses drop significantly as compared to their vacuum values: $`m_{u,d}0.1m_{u,d}^{\mathrm{vac}}`$ , $`m_s0.3m_s^{\mathrm{vac}}`$. For zero bare masses, $`m_{0f}=0`$, the constituent quark masses in bound states would be practically zero. This means that the bound states correspond to the Wigner phase where chiral symmetry is restored. Fig. 5 shows the chiral condensate $`\overline{u}u`$ which, in our notation (see Eq. (6)), coincides with the scalar density $`\rho _{Su}`$ (notice the minus sign on the vertical axis). Its behaviour is strongly correlated with the $`u`$–quark constituent mass (compare with Fig. 4). As before, the variation of the $`\overline{u}u`$ condensate is mainly sensitive to the density of valence $`u`$–quarks and antiquarks. At $`r_s=1`$, when this density is zero, the change in the $`\overline{u}u`$ condensate is caused by the flavour–mixing interaction. The behaviour of the $`\overline{s}s`$ condensate is qualitatively similar, but due to the larger bare mass, $`m_{0s}140`$ MeV, its density dependence is weaker. The $`u`$ and $`s`$ chemical potentials are shown in Fig. 6. It is interesting that initially they drop and then rise with density. One can easily understand this behaviour by analyzing the explicit expression for $`\mu _f`$, $$\mu _f=\sqrt{m_f^{\mathrm{\hspace{0.17em}2}}+p_{Ff}^{\mathrm{\hspace{0.17em}2}}},$$ (35) where $`p_{Ff}=\left(6\pi ^2\rho _f/\nu \right)^{1/3}`$ is the Fermi momentum of quarks with flavour $`f`$ (antiquarks have the same Fermi momentum in symmetric matter). At low densities, when $`p_{Ff}m_f`$ and $`\mu _fm_f`$, all chemical potentials decrease with density together with the respective constituent mass $`m_f`$ (see Fig. 4). At high densities, when $`p_{Ff}m_f`$, the chemical potential grows with density as in the free relativistic Fermi–gas, $`\mu _fp_{Ff}\rho _f^{1/3}`$. In fact, this nontrivial behaviour of $`\mu _f`$ is responsible for the first order phase transition discussed below. ## IV Strange quark matter Let us turn now to quark matter with nonzero net baryon density at $`T=0`$. We assume that only valence quarks are present, i.e. the density of valence antiquarks is zero for each flavour ($`\overline{\rho }_f=0`$). In other words, this means that all levels in the Dirac sea are filled up, and additionally some levels in the Fermi sea are occupied by the valence quarks. This situation is schematically shown in Fig. 1(c). In contrast to the symmetric $`q\overline{q}`$ matter discussed above, now the symmetry between positive and negative energy states is lost due to the presence of repulsive vector interaction. Fig. 7 shows the energy per baryon $`ϵ`$ as a function of baryon density, $`\rho _B=\frac{1}{3}\underset{f}{}\rho _f`$ . Different curves correspond to different $`r_s`$, which in this case is simply the ratio of the strange quark density to the total density of all quarks. At $`\rho _B0`$ the energy per quark tends to the corresponding vacuum mass given by Eq. (34). With growing density both the attractive scalar and repulsive vector interactions contribute to $`ϵ`$. It is interesting that at $`r_s0.7`$ the attractive interaction is strong enough to produce a nontrivial local minimum at a finite $`\rho _B`$. In the pure $`u,d`$ matter ($`r_s=0`$) this minimum is unbound by about 20 MeV as compared to the vacuum masses of $`u`$ and $`d`$ quarks. On the other hand, it is located at a baryon density of about $`1.8\rho _0`$, which is surprisingly close to the saturation density of normal nuclear matter. Of course, the location of this minimum depends on the model parameters. Nevertheless, one can speculate that nucleon–like 3–quark correlations, not considered in the mean–field approach, will turn this state into the correct nuclear ground state. When $`r_s`$ grows from 0 to about 0.4, the local minimum is getting more pronounced and the corresponding baryon density increases to about $`3.2\rho _0`$. At larger $`r_s`$, the minimum again becomes more shallow and disappears completely at $`r_s0.7`$. At $`0.2<r_s<0.6`$ the minima correspond to the true bound states, i.e. the energy per quark is lower than the respective vacuum mass. But in all cases these bound states are rather shallow: even the most strongly bound state at $`r_s0.4`$ is bound only by about 5 MeV per quark or 15 MeV per baryon. Nevertheless, the appearance of local minima signifies the possibility for finite droplets to be in mechanical equilibrium with the vacuum at $`P=0`$. It is natural to identify such droplets with strangelets, which are hypothetical objects made of light and strange quarks . Similar multiparticle bound states made of nucleons and hyperons were also discussed . It should be emphasized here that $`\beta `$–equilibrium is not required in the present approach (see the discussion below). That is why our most bound strangelets are predicted to be more rich in strange quarks ($`r_s>1/3`$) than in the approaches assuming $`\beta `$–equilibrium , which give $`r_s<1/3`$. As a result, these strangelets will be negatively charged<sup>3</sup><sup>3</sup>3Negatively–charged strangelets have been also considered in Refs. .. Indeed, the ratio of the charge $`Q`$ to the baryon number $`B`$ is expressed through $`r_s`$ as $$\frac{Q}{B}=\frac{2}{3}\frac{\rho _u}{\rho _B}\frac{1}{3}\frac{\rho _d}{\rho _B}\frac{1}{3}\frac{\rho _s}{\rho _B}=\frac{1}{2}(13r_s).$$ (36) For $`r_s0.4`$ this gives $`Q/B0.1`$. In light of recent discussions (see e.g. Ref. ) concerning possible dangerous scenarios of the negatively–charged strangelet production at RHIC, we should emphasize that the strangelets predicted here are not absolutely bound<sup>4</sup><sup>4</sup>4The analogous conclusion has been made in Ref. ., i.e. their energy per baryon is higher than that for the normal nuclear matter. Hence, the spontaneous conversion of normal nuclear matter to strange quark matter is energetically not possible. The behaviour of the energy per quark for arbitrary mixtures of light and strange quarks is shown in Fig. 3(b). One can clearly see that the valley corresponding to the local minima has always a positive slope in $`r_s`$. There is no potential barrier separating the states with $`r_s0`$ and $`r_s=0`$. Thus, a strangelet with any $`r_s0`$ will freely roll down to the normal nuclear matter state with $`r_s=0`$. It is interesting to note that a barrier in $`r_s`$ may appear if attractive scalar interaction is arbitrarily enhanced in the strange sector . Fig. 8 shows the constituent masses of $`u`$ and $`s`$ quarks as functions of baryon density. The dropping masses manifest again a clear tendency to the restoration of chiral symmetry at high densities. As before, the dots indicate the masses at the local minima in the respective energies per baryon shown in Fig. 7. These mass values are somewhat larger than in the symmetric $`q\overline{q}`$ matter discussed above (see Fig. 4). Note that the stronger is reduction of constituent masses the deeper are the corresponding bound states. For the metastable state at $`r_s=0`$, which is a candidate for the nuclear ground state, the masses of $`u`$ and $`s`$ quarks are equal to 0.3 and 0.9 of their vacuum values respectively. For the most bound state at $`r_s0.4`$ the respective mass ratios are reduced to 0.15 and 0.6. The behaviour of the $`u`$ and $`s`$ chemical potentials is shown in Fig. 9. One can notice the differences as compared with symmetric $`q\overline{q}`$ matter (Fig. 6). In particular, the curves for different $`r_s`$ do not intersect. Due to the additional contribution of the vector interaction, which is linear in $`\rho _B`$ (see Eq. (14) where $`\mu _{Rf}=\sqrt{m_f^{\mathrm{\hspace{0.17em}2}}+p_{Ff}^{\mathrm{\hspace{0.17em}2}}}`$ at $`T=0`$), the minima in $`\mu _u`$ and $`\mu _s`$ are less pronounced. At $`r_s>0.7`$ the curves have no minima at all. Therefore the chiral phase transition will be not as strong in this case as in symmetric matter. ## V Systematics of bound states The properties of the multiparticle bound states discussed in preceding sections are summarized in Figs. 10–11. Fig. 10 shows the binding energy per quark, $`m_q^{\mathrm{vac}}(r_s)ϵ_{\mathrm{min}}(r_s)`$, where $`m_q^{\mathrm{vac}}`$ is the energy at $`\rho _{tot}=0`$ and $`ϵ_{\mathrm{min}}(r_s)`$ is the energy at a local minimum, both taken for a given $`r_s`$ value. To avoid misunderstanding, Figs. 11(a) and 11(b) show $`ϵ_{\mathrm{min}}(r_s)`$ separately. Although the absolute minimum of $`ϵ_{\mathrm{min}}`$ corresponds to $`r_s=0`$, the maximum bindings are realized at nonzero $`r_s`$. For symmetric $`\overline{q}q`$ systems the maximum value of about 90 MeV is reached for $`r_s0.7`$. This is indeed very strange and very bound matter! For asymmetric systems, where $`\overline{\rho }_f=0`$, the maximum binding energy is much smaller, about 5 MeV per quark at $`r_s0.4`$. One should bear in mind, however, that in this case $`ϵ_{\mathrm{min}}`$ results from a strong cancellation between the attractive scalar and repulsive vector interactions. Therefore, it is very sensitive to their relative strengths. The results presented above are obtained for $`G_V=0.5G_S`$. For comparison in Figs. 10 and 11(b) we also present the model predictions for $`G_V=0`$. In this case the maximum binding energy increases to about 30 MeV per quark and the corresponding $`r_s`$ value shifts to about 0.6. It is interesting to note that for $`G_V=0`$ the bound state appears even in the pure $`u,d`$ matter. The corresponding binding energy is about 7 MeV per quark, i.e. 21 MeV per baryon. Figures 10 and 11 reveal some differences compared to the previous calculations in Ref. . This is mainly due to a different set of model parameters used there. In particular, the present set of parameters gives higher quark constituent masses in the vacuum, $`m_{u,d}^{\mathrm{vac}}368`$ MeV and $`m_s^{\mathrm{vac}}550`$ MeV, as compared to $`m_{u,d}^{\mathrm{vac}}=300`$ MeV and $`m_s^{\mathrm{vac}}=520`$ MeV in Ref. . This leads to an overall upward shift in the energy per particle by 30–60 MeV. On the other hand, inclusion of the flavour–mixing interaction lowers the energy. As a result, all binding energies increase slightly. A local minimum in $`ϵ`$ appears now even in the pure $`u,d`$ matter. The dots in Fig. 11 indicate the positions of some conventional mesons (a) and baryons (b). Their empirical masses are rescaled (by factor $`1/2`$ for mesons and $`1/3`$ for baryons) and shown at $`r_s`$ values corresponding to their flavour compositions. By inspecting the figure, one can make a few interesting observations. According to Fig. 11(a), mesoballs lie lower in energy than conventional vector mesons, but higher than the pseudoscalar mesons (see discussion in Sect. VIII). As one can see in Fig. 11(b), conventional baryons are more bound than strangelets even at $`G_V=0`$ . This is an indication that baryon–like 3–quark correlations might be indeed very important in the baryon–rich quark matter. ## VI Quark matter at finite temperatures In this section we study properties of deconfined matter at finite temperatures. The calculations for this case can be done by using general formulas of Sect. II with the quark and antiquark occupation numbers given by Eqs. (12)–(13). First we discuss thermal properties of meso-matter where the chemical potentials obey the conditions (32). A typical behaviour of the equation of state for this case is illustrated in Fig. 12(a). It shows the pressure isotherms for symmetric $`q\overline{q}`$ matter with strangeness content $`r_s=1/3`$. One can clearly see a strong first order phase transition which is signaled by the appearance of isotherms with negative slopes, $`_\rho P<0`$ (spinodal instability). The corresponding critical temperature is about 100 MeV. Another important feature is that the zero-pressure states persist up to temperatures as high as 70 MeV. This means that a finite droplet of this matter which has cooled down to this temperature and still remaining at high density of $`q\overline{q}`$ pairs (around 4 $`\rho _0`$ in this case), will be trapped in a bound state. At later times it will further cool down by emitting hadrons from the surface (see below). The critical temperatures for a phase transition and for the appearance of a bound state are shown in Fig. 13(a) as functions of $`r_s`$. One can conclude that this dependence is rather weak: the first critical temperature changes between 90 and 110 MeV while the second one varies around 70 MeV. Now let us consider chemically–equilibrated quark matter at finite temperatures. In this case the chemical potentials of quarks obey the conditions (16). Fig. 12(b) represents the pressure isotherms for the case $`\mu _s=0`$ . This condition implies equal numbers of strange quarks and antiquarks, i.e. zero net strangeness. It is appropriate for fast processes where strangeness is conserved, e.g. in relativistic nuclear collisions. As before one can see a region of spinodal instability, $`_\rho P<0`$, which is characteristic for a first order phase transition. However, this phase transition is much weaker than in the case of symmetric $`q\overline{q}`$ matter (note the different scales in Figs. 12(a) and 12(b)). The corresponding critical temperature is about 35 MeV in this case. The zero–pressure states exist only at temperatures below 15 MeV. Generally, the equation of state of the chemically equilibrated quark matter is characterized by two quantities: net baryon charge, $`B`$, and net strangeness, $`S`$. Therefore, it is interesting to study thermal properties of this matter at $`S0`$ . Such states can be reached in neutron stars. They can also be realized via the distillation mechanism accompanying a QCD phase transition in heavy–ion collisions . Leaving a detailed study of this question for a future publication, here we only present results which complement our discussion concerning Fig. 13(a). Fig. 13(b) shows the critical temperatures for the phase transition and for the bound states in the equilibrated matter as functions of $`r_s^{}=S/3B`$<sup>5</sup><sup>5</sup>5We introduce this new notation in order to distinguish this quantity from $`r_s`$ defined in Eq. (33). In chemically equilibrated matter at $`T=0`$ (no antiquarks) these two quantities coincide.. One can see that both temperatures first grow with $`r_s^{}`$ and then drop to zero at $`r_s^{}0.8`$. The maximal values of respectively 50 MeV and 30 MeV are realized at some intermediate $`r_s^{}`$ around 0.4. As demonstrated earlier in Fig. 10, this value of $`r_s^{}`$ corresponds to the most bound state of strange quark matter at $`T=0`$. So we see an obvious correlation: the deeper is a bound state at $`T=0`$ the stronger is a phase transition at finite temperatures. It should be emphasized here again that the thermal properties of asymmetric baryon–rich quark matter are very sensitive to the relative strength of scalar and vector interactions. To illustrate this point we have calculated the phase diagrams for several values of $`G_V`$. The results for $`\mu _s=0`$ are presented in Fig. 14. It shows the boundaries of two–phase coexistence regions calculated by using standard Gibbs conditions . If we take $`G_V=0`$, as in most calculations in the literature, the coexistence region becomes wider and the corresponding critical temperature increases to about 70 MeV. On the other hand, if one takes $`G_V=0.65G_S`$, the phase transition becomes weaker: the critical temperature moves down to 20 MeV and the zero–pressure states disappear completely. The calculation shows that there is no phase transition at $`G_V>0.71G_S`$ . It is interesting to note that in all cases this first order phase transition occupies the region of densities around the normal nuclear density $`\rho _0`$ . For instance, at $`G_V=0.5G_S`$ the coexistence region at $`T=0`$ extends from 0.2 to 1.7 $`\rho _0`$. These results demonstrate that this chiral phase transition is rather similar to a liquid–gas phase transition in normal nuclear matter. The critical temperature and baryon density in the present case ($`T_c30`$ MeV, $`\rho _{Bc}\rho _0`$) are not so far from the values predicted by the conventional nuclear models ($`T_c20`$ MeV, $`\rho _{Bc}0.5\rho _0`$) . One may expect that in a more realistic approach, taking into account nucleonic correlations, the “chiral” transition may turn into the ordinary “liquid–gas” phase transition. If this would be the case, one should be doubtful about the possibility of any other QCD phase transition of the liquid–gas type at a higher baryon density. At least only one phase transition of this type is predicted within the NJL model. ## VII Discussion of decay modes Let us discuss briefly possible decay channels of the novel states of quark matter described above. Naively one could think that the symmetric $`q\overline{q}`$ matter would be extremely unstable with respect to the annihilation of quarks and antiquarks. But, in fact, many annihilation channels are closed or do not exist in dense $`q\overline{q}`$ matter. From the first sight, one may think that two–pion annihilation, $`\overline{q}+q\pi +\pi `$, should be the strongest decay channel in the bulk (one–pion annihilation is not allowed by the energy–momentum conservation). But simple arguments show that this might be not true. Indeed, the bound states of $`q\overline{q}`$ matter appear at such high densities when chiral symmetry is practically restored. In this case the pion looses its special nature as a Goldstone boson and inside this matter it should be as heavy as other mesons. Moreover, due to the Mott transition , it is unlikely that conventional mesonic states survive in this dense medium. Therefore, one can expect that the hadronic channels of the $`q\overline{q}`$ annihilation simply do not exist in the bulk. Another possible decay channel is the strong flavour conversion, $`\overline{s}s\overline{u}u`$ or $`\overline{d}d`$. In principle, this process can proceed through the one–gluon (color octet) intermediate state. However, it may be shown that due to color neutrality the corresponding matrix elements vanish after summation over the color indices. Within the present approach, the 4–fermion interaction terms are diagonal in the flavour space (see Eq. (5)) and, therefore, cannot change the flavour. Strong flavour conversion is possible only through the 6–fermion flavour–mixing interaction, $`\overline{s}s\overline{u}u+\overline{d}d`$ . This channel is only open when $`\mu _s>\mu _u+\mu _d`$. By inspecting Fig. 6 one can see that this condition is never fulfilled. Therefore we conclude that strong flavour conversion is not possible or at least strongly suppressed in the bulk. In a finite droplet, hadronic annihilation, $`\overline{q_f}+q_f^{}h_1+h_2+\mathrm{}`$, is certainly possible at the surface, if $`\overline{\mu }_f+\mu _f^{}>m_{h_1}+m_{h_2}+\mathrm{}`$ . It is clear from Fig. 11(a) that the emission of pions in annihilation of light $`q\overline{q}`$ pairs is energetically more favorable than the emission of heavier mesons. Moreover, many annihilation channels of strange quarks and antiquarks are simply closed e.g. $`\overline{s}+sK+\overline{K}`$ . As a result of the pion emission, the strangeness content of the daughter droplet will increase<sup>6</sup><sup>6</sup>6Similar processes leading to strangeness distillation have been considered in Refs. .. Finally, it will contain predominantly the $`s\overline{s}`$ pairs, similarly to a system composed of $`\varphi `$–mesons (“$`\varphi `$–ball”). Further hadronization may be slowed down by several reasons . According to Fig. 10, at $`r_s1`$ the energy available in the annihilation of a $`s\overline{s}`$ pair is $`\mu _s+\overline{\mu }_s0.95`$ GeV. This is lower than the threshold energies of the hadronic states $`\varphi (1020)`$ and $`K\overline{K}(990)`$. The only open channels in this case are $`\rho \pi (910)`$ and $`3\pi (420)`$. Their partial width in the $`\varphi `$-meson decay is about 0.7 MeV. If this would be the dominant decay mode, one could expect a life time for these $`\varphi `$–balls of about 280 fm/c, i.e much longer than the typical duration of a heavy–ion collision. In strange quark matter (without antiquarks), flavour conversion is only possible through weak decays. As follows from Fig. 9, at densities corresponding to zero pressure, the condition $`\mu _s>\mu _u`$ holds. This means that weak processes of the types $`su+e^{}+\overline{\nu }_e`$ , $`s+uu+d`$ are allowed. Since there is no local barrier at any $`r_s`$ (see Fig. 11(b)), a system produced initially at some $`r_s0`$ will subsequently reduce this $`r_s`$ value by a cascade of weak decays. As a result, the system will roll down along the line of local minima shown in Fig. 11(b) (see also Fig. 3(b)). Finally, all $`s`$ quarks will be converted into light $`u,d`$ quarks. Two steps of this conversion process are shown schematically in Figs. 15(a) and 15(b). This picture is very different as compared to the one based on the MIT bag model. Within that model the pressure and energy density are given by simple expressions $$P_{\mathrm{MIT}}=P_{\mathrm{id}}B,e_{\mathrm{MIT}}=e_{\mathrm{id}}+B,$$ (37) where $`B`$ is a bag constant, $`P_{\mathrm{id}}`$ and $`e_{\mathrm{id}}`$ are the pressure and energy density for a mixture of ideal gases of $`u`$, $`d`$ and $`s`$ quarks with constant (bare) masses, $`m_{u,d}=m_{0u,d}0`$, $`m_s=m_{0s}150`$ MeV. By properly choosing $`B`$ one can always get a zero pressure point and, accordingly, a minimum in the energy per baryon. Because the quark masses are kept constant, the condition $`\mu _s>m_s`$ will be first satisfied at a relatively low baryon density $`m_{0s}^3`$. At higher densities a certain fraction of $`s`$ quarks will always be present in a $`\beta `$–equilibrated matter. In contrast, in the NJL model the $`s`$ quark mass is a function of both baryon density and strangeness content. As one can see from Fig. 9, at any given $`r_s`$ the condition $`\mu _{u,d}=\mu _s`$ can be satisfied only at sufficiently high baryon densities which correspond to positive pressure. On the other hand, at the points of zero pressure we always have $`\mu _{u,d}<\mu _s`$. Therefore, weak decays will proceed until a strangelet reaches $`r_s=0`$ (see Fig. 15(c)). ## VIII Conclusions In the present paper we have investigated properties of deconfined matter at different densities of quarks and antiquarks, not necessarily constrained by the conditions of chemical equilibrium. All calculations are carried out within the SU(3)–flavour NJL model including scalar, vector and flavour–mixing interactions. We have demonstrated the possibility of strongly bound states in symmetric $`q\overline{q}`$ systems consisting of equal numbers of quarks and antiquarks (“mesoballs”). The maximal binding energies of mesoballs, of about 90 MeV per particle or 180 MeV per $`q\overline{q}`$ pair, are realized for flavour compositions with about 70% of $`s\overline{s}`$ pairs. These systems remain bound up to the temperatures $`T\stackrel{<}{}\mathrm{\hspace{0.17em}70}`$ MeV. The lifetimes of mesoballs may be long enough due to the suppression of annihilation into hadrons in the bulk. The model predicts a strong first order phase transition in chemically nonequilibrated meso-matter, i.e. at zero net baryon density. The critical temperature of this phase transition is in the range of 90–110 MeV depending of the relative abundance of $`s\overline{s}`$ pairs. As discussed in Ref. , formation of mesoballs in high–energy heavy–ion collisions may be observed through the event–by–event analysis of $`\pi ,K,\varphi `$ spectra. In particular, narrow bumps in the hadron rapidity distributions may be generated by hadronizing mesoballs. By using the same model, we have also investigated the equation of state of chemically equilibrated deconfined matter at various temperatures, baryon densities and strangeness contents. The model predicts the existence of loosely bound, negatively–charged strangelets with maximal binding energies of about 20 MeV per baryon at $`r_s0.4`$ . Similarly to Ref. , no absolutely stable strange quark matter has been found. It is shown that properties of baryon–rich quark matter are very sensitive to the relative magnitude of the vector and scalar interactions. At the standard value of vector and scalar couplings, $`G_V/G_S=0.5`$, the metastable bound states of chemically equilibrated matter exist at $`T<15`$ MeV, while at $`G_V=0`$ this temperature increases to 40 MeV. The first order chiral phase transition in equilibrated baryon–rich quark matter is much weaker as compared to the symmetric $`q\overline{q}`$ matter. In the case of zero net strangeness the critical temperature is in the range of 30 MeV and the critical baryon density is around $`\rho _0`$. We believe that this phase transition is reminiscent of the ordinary liquid–gas phase transition in nuclear matter. One can use the present model to study the chiral phase transition at nonzero net strangeness, which is appropriate e.g. for neutron star matter. We have found that the maximal critical temperature $`T_c50`$ MeV is reached for the ratio of net strangeness to baryon charge $`S/B1.2`$ . It should be emphasized that all calculations in this paper have been made with constant values of coupling constants $`G_S,G_V,K`$ . We think that, this is a reasonable approximation at intermediate densities and temperatures. On the other hand, the asymptotic freedom of QCD requires all effective interactions to vanish at high densities and temperatures. Therefore, in a more realistic approach the coupling constants should decrease with density and temperature. Hopefully, this behaviour can be simulated by introducing finite range form factors. ## Acknowledgments The authors thank M. Belkacem and J. Schaffner-Bielich for useful discussions. Two of us (I.N.M. and L.M.S.) thank the Institut für Theoretische Physik, J.W. Goethe Universität, Frankfurt am Main, for the kind hospitality. This work has been supported by the Alexander von Humboldt Stiftung, the Graduiertenkolleg “Experimentelle und Theoretische Schwerionenphysik”, GSI, BMBF and DFG. ## Figure captions FIG. 1. Schematic picture of energy levels of quarks and antiquarks in vacuum (a), symmetric $`q\overline{q}`$ droplet (b) and pure quark droplet (c). Occupied quark states are shown by full dots while open dots represent holes (antiquarks). $`R`$ denotes the radius of a droplet, $`m`$ and $`V`$ are, respectively, constituent mass and vector potential of quarks. The boundaries of vacuum mass gap are shown by dashed lines. FIG. 2. Energy per particle $`ϵ`$ in symmetric $`q\overline{q}`$ matter vs. total particle density at different values of strangeness fraction $`r_s`$ (shown near the corresponding curves). Points indicate local minima of $`ϵ`$. $`\rho _0=0.17\mathrm{fm}^3`$ is normal nuclear density. FIG. 3. Contours of energy per particle $`ϵ`$ (in GeV) for symmetric (a) and asymmetric (b) matter in the $`\rho _{u+d}\rho _s`$ plane. Dotted lines represent local extrema of $`ϵ`$ . Shading shows regions with negative pressure. FIG. 4. Constituent masses of $`u`$ and $`s`$ quarks in symmetric $`q\overline{q}`$ matter as functions of total particle density. Figures in the box show values of strangeness fraction $`r_s`$. Dots correspond to minima of energy per particle at given $`r_s`$. FIG. 5. The same as in Fig. 4, but for condensate densities of $`u`$ and $`s`$ quarks. FIG. 6. The same as in Fig. 4, but for chemical potentials of $`u`$ and $`s`$ quarks. FIG. 7. The same as in Fig. 2, but for asymmetric $`u,d,s`$ matter without antiquarks. FIG. 8. The same as in Fig. 4, but for asymmetric quark matter. FIG. 9. The same as in Fig. 6, but for asymmetric quark matter. FIG. 10. Binding energies per particle in symmetric $`q\overline{q}`$ matter (solid line) and asymmetric quark matter as functions of strangeness fraction $`r_s`$. Dotted line shows the results of calculations when the vector interaction is switched off ($`G_V=0`$). FIG. 11. Minimal energies per particle in symmetric (a) and asymmetric (b) matter as functions of strangeness fraction $`r_s`$. Dotted lines show energy per particle in the limit of zero particle densities, Eq. (34). Different parts of the solid line in the lower panel correspond to metastable (AB and CD) or bound (BC) states. Triangles and squares show masses of lightest mesons and baryons devided by the total number of constituents (two in mesons and three in baryons). Symbol $`<K>`$ represents the spin averaged mass of $`K`$ and $`K^{}(892)`$ mesons, i.e. $`(3m_K^{}+m_K)/8`$. Dashed line in the lower plot shows the results in the limit $`G_V0`$. FIG. 12. Pressure isotherms for symmetric $`q\overline{q}`$ matter (a) and equilibrated quark matter with $`\mu _s=0`$ (b). Temperatures are given in MeV near the corresponding curves. Boundaries of spinodal regions are shown by the dashed lines. FIG. 13. Critical temperatures for existence of bound states ($`P<0`$) and phase transition ($`_\rho P<0`$) in symmetric $`q\overline{q}`$ matter (a) and equilibrium quark matter (b) as functions of strangeness fraction. FIG. 14. Phase diagrams of equilibrium quark matter with zero net strangeness at various ratios of vector and scalar coupling constants. FIG. 15. Schematic pictures of energy levels (shown by shading) occupied by light and strange quarks in cold quark matter at different strangeness fractions $`r_s`$. Hatched boxes represent the constituent quark masses. Arrows show weak decay processes $`su+e+\overline{\nu }_e`$ and $`s+uu+d`$.
warning/0002/astro-ph0002300.html
ar5iv
text
# Rotational modes of non-isentropic stars and the gravitational radiation driven instability ## 1 Introduction Since the recent discovery of the gravitational radiation driven instability of the $`r`$-modes by Andersson (1998) and Friedman & Morsink (1998), a large number of papers have been published on the properties of the rotational modes of rotating stars because of their possible importance in astrophysics (e.g., Kojima 1998, Andersson, Kokkotas & Schutz 1998, Kokkotas & Stergioulas 1998, Lindblom & Isper 1998, Beyer & Kokkotas 1999, Lindblom, Mendell & Owen 1999, Kojima & Hosonuma 1999, Lockitch & Friedman 1999, Yoshida & Lee 2000, Lockitch 1999, Yoshida et al 1999). Here, we have used the term of “rotational mode” to refer to both $`r`$-mode and inertial mode of rotating stars. As an important consequence of the instability, for example, Lindblom, Owen & Morsink (1998) argued that due to the instability, the maximum angular rotation velocity of hot neutron stars is strongly restricted. Owen et al. (1998) also suggested that the gravitational radiation emitted from hot young neutron stars due to the $`r`$-mode instability would be one of the potential sources for the gravitational wave detectors, e.g. LIGO. The dominant restoring force for the rotational modes is the Coriolis force and the characteristic frequency is comparable to the angular rotation frequency $`\mathrm{\Omega }`$ of the star (e.g., Greenspan 1964, Pedlosky 1987). The $`r`$-modes are characterized by the properties that the toroidal motion dominates the spheroidal motion in the velocity field, and that the oscillation frequency $`\omega `$ observed in the corotating frame of the star is given by $`\omega =2m\mathrm{\Omega }/(l^{}(l^{}+1))`$ in the limit of $`\mathrm{\Omega }0`$, where $`m`$ and $`l^{}`$ are the indices of the spherical harmonic function $`Y_l^{}^m`$ representing the toroidal component of the velocity field of the mode (e.g., Papaloizou & Pringle 1978 , Provost, Berthomieu & Rocca 1981, Saio 1982). On the other hand, the inertial modes have the comparable toroidal and spheroidal components of the velocity field and therefore they are not necessarily represented by a single spherical harmonic function (e.g, Lee, Strohmayer & Van Horn 1992, Yoshida & Lee 2000). Note that the inertial mode of this kind is also called “rotation mode” or “generalized $`r`$-mode” (Lindblom & Ipser 1998, Lockitch & Friedman 1999). Most of the recent studies on the $`r`$-mode instability driven by the gravitational radiation reaction are restricted to the case of neutron stars with isentropic structure, since the barotropic (one-parameter) equation of state, $`p=p(\rho )`$ is usually assumed for both the interior structure and the small amplitude oscillation. In isentropic stars the specific entropy of the fluid in the interior is constant and the buoyant force has no effects on the oscillation. However, stars in general have non-isentropic interior structure, which may be divided into two kinds of layers with fluid stratification stable or unstable against convection. It is well know that the property of the oscillation modes of non-isentropic stars is substantially different from that of isentropic stars, since the buoyant force comes into play as a restoring force for low frequency oscillations like $`g`$-mode propagating in the layers with the stable stratification (see, e.g., Unno et al 1989). This is also the case for rotational modes, which have low oscillation frequencies comparable to the spin rate $`\mathrm{\Omega }`$ of the star. For example, we note that the $`r`$-mode associated with $`l^{}=|m|`$ having the nodeless eigenfunction in the radial direction is the only $`r`$-mode permitted in an isentropic star for a given $`m`$ (e.g, Lockitch & Friedman 1999, Yoshida & Lee 2000), but in non-isentropic stars with the stable fluid stratification all the $`r`$-modes with $`l^{}|m|`$ are possible for a given $`m`$, and there are in principle an infinite number of $`r`$-modes for given $`l^{}`$ and $`m`$ (e.g., Provost et al 1981, Saio 1982, Lee & Saio 1997). We therefore consider it important to investigate the modal properties of the rotational modes in non-isentropic stars, particularly in order to answer the question whether the property of the $`r`$-mode instability for neutron star models with non-isentropic structure remains the same as that for isentropic neutron star models. In this paper, we study the $`r`$-mode and the inertial mode oscillations of slowly rotating, non-isentropic, Newtonian stars, where uniform rotation and adiabatic oscillation are assumed to simplify the mathematical treatments. The effects of the rotational deformation of the equilibrium structure on the eigenfunction and eigenfrequency are included in the analysis, using the formulation by Yoshida & Lee (2000) (see also Lee & Saio 1986, Lee 1993). We employ simple polytropic models with the polytropic index $`n=1`$ as the background neutron star models. For the non-isentropic models we consider only two cases, that is, the models that have stable fluid stratification in the whole interior and those that are in convective equilibrium. In this paper, following the conventions used in the field of stellar evolution and pulsation, we simply call these two kinds of models “radiative” and “convective” models, respectively, although no energy transport in the interior is considered. The plan of this paper is as follows. In §2, we briefly describe the basic equations employed in the linear modal analysis of slowly rotating stars. In §3, we show numerical results for the rotational modes of non-isentropic stars, and discuss the properties of the eigenfrequency and eigenfunction of the modes. In §4, we examine the stability of simple neutron star models against the $`r`$-modes, computing their growth or damping timescales due to the gravitational radiation reaction and some viscous damping processes. §5 is for conclusions. ## 2 Method of Solutions The method of solutions used in this paper is exactly the same as that applied in Yoshida & Lee (2000). Here, therefore, we briefly describe the basic equations and the method of calculation. The details are given in Yoshida & Lee (2000). ### 2.1 Equilibrium Configurations We consider oscillations of uniformly rotating Newtonian stars. Thus the angular rotation frequency of the star, $`\mathrm{\Omega }`$ is constant. We assume the polytropic relation for the equilibrium structure of the model, and the relation between the pressure $`p`$ and the mass density $`\rho `$ is given by $$p=K\rho ^{1+\frac{1}{n}},$$ (1) where $`n`$ and $`K`$ are the polytropic index and the structure constant determined by giving the mass and the radius of the model, respectively. In this investigation, assuming slow rotation, we employ the Chandrasekhar-Milne expansion (see, e.g, Tassoul 1978) to obtain the equilibrium structure of rotating stars. In this technique the effects of the centrifugal force and the equilibrium deformation are treated as small perturbations to the non-rotating spherical symmetric structure of the star. The small expansion parameter due to rotation is chosen as $`\overline{\mathrm{\Omega }}=\mathrm{\Omega }(R^3/GM)^{1/2}`$, where $`R`$ and $`M`$ are the radius and the mass of the non-rotating star, respectively. The equilibrium configurations are constructed with accuracy up to the order of $`\overline{\mathrm{\Omega }}^2`$. For simplicity, we consider a sequence of slowly rotating stars whose central density is the same as that of the non-rotating star. ### 2.2 Perturbation Equations In the perturbation analysis, we introduce a parameter $`a`$ that is constant on the distorted effective potential surface of the star. In practice, the parameter $`a`$ is defined such that $$\mathrm{\Psi }(r,\theta )=\mathrm{\Psi }_0(a),$$ (2) where $`\mathrm{\Psi }_0`$ and $`\mathrm{\Psi }`$ are the effective potentials for the non-rotating and slowly rotating models, respectively. Note that $`\mathrm{\Psi }_0`$ is equal to the gravitational potential and $`\mathrm{\Psi }`$ is the sum of the gravitational potential and the centrifugal potential. With this parameter $`a`$, the distorted equi-potential surface may be given by $$r=a\{1+ϵ(a,\theta ,\phi )\},$$ (3) where $`ϵ`$ is the function of the order of $`\overline{\mathrm{\Omega }}^2`$, and is obtained by giving the functional forms of $`\mathrm{\Psi }`$ and $`\mathrm{\Psi }_0`$ using the Chandrasekhar-Milne expansion (Yoshida & Lee 2000). Here we have used spherical polar coordinates $`(r,\theta ,\phi )`$. However, we employ hereafter the parameter $`a`$ as the radial coordinate, in stead of the polar radial coordinate $`r`$. The governing equations of nonradial oscillations of a rotating star are obtained by linearizing the basic equations used in fluid mechanics (e.g., Unno et al 1989). In the following, we use $`\delta `$ and $`\mathrm{\Delta }`$ to denote the Eulerian and the Lagrangian changes of physical quantities, respectively. Since the equilibrium state is assumed to be stationary and axisymmetric, the perturbations may be represented by a Fourier component proportional to $`e^{i(\sigma t+m\phi )}`$, where $`\sigma `$ is the frequency observed in an inertial frame and $`m`$ is the azimuthal quantum number. The continuity equation may be linearized to be $$\delta \rho =_i(\rho \xi ^i),$$ (4) where $`\xi ^i`$ is the Lagrangian displacement vector, and we have made use of $`\delta v^i=i(\sigma +m\mathrm{\Omega })\xi ^ii\omega \xi ^i`$ with $`\omega `$ being the oscillation frequency observed in the corotating frame of the star. Here $`_i`$ denotes the covariant derivative derived from the metric $`g_{ij}`$. In this study, we consider only adiabatic oscillations. Thus, the perturbed equation of state is given by $$\delta p=\mathrm{\Gamma }p\left(\frac{\delta \rho }{\rho }+\xi ^iA_i\right),$$ (5) where $`\mathrm{\Gamma }`$ is the adiabatic exponent defined as $$\mathrm{\Gamma }=\left(\frac{\mathrm{ln}p}{\mathrm{ln}\rho }\right)_{ad},$$ (6) and $`A_i`$ is the generalized Schwarzschild discriminant<sup>1</sup><sup>1</sup>1Since the Schwarzschild discriminant $`A_i`$ is proportional to $`_is`$ with $`s`$ being the specific entropy, a model with $`A_i=0`$ throughout the interior is called isentropic model in this paper. Using the same barotropic (one-parameter) equation of state, $`p=p(\rho )`$ for both the equilibrium structure and the perturbation leads to isentropic models. given by $$A_i\frac{1}{\rho }_i\rho \frac{1}{\mathrm{\Gamma }p}_ip,$$ (7) and the positive and negative values of $`A_i`$ indicate the fluid stratification unstable and stable against convection, respectively. The linearized Euler’s equation is $$\omega ^2\xi _i+_i\left(\frac{\delta p}{\rho }+\delta \mathrm{\Phi }\right)+\left(\frac{\delta p}{\rho }g_{ij}+\xi _j\frac{1}{\rho }_ip\right)A^j+2i\omega \mathrm{\Omega }\xi ^j_j\phi _i=0,$$ (8) where $`\mathrm{\Phi }`$ denotes the gravitational potential, and $`\phi ^i`$ is the rotational Killing vector, with which the 3-velocity of the equilibrium fluid of a rotating star is given as $`v^i=\mathrm{\Omega }\phi ^i`$. The perturbed Poisson equation is $$_i^i\delta \mathrm{\Phi }=4\pi G\delta \rho ,$$ (9) where $`G`$ is the gravitational constant. Physically acceptable solutions of the linear differential equations are obtained by imposing boundary conditions at the inner and outer boundaries of the star. The inner boundary conditions are the regularity condition of the perturbed quantities at the stellar center. The outer boundary conditions at the surface of the star are $`\mathrm{\Delta }p/\rho =0`$ and the continuity of the perturbed gravitational potential at the surface to the solution of $`_i^i\delta \mathrm{\Phi }=0`$ that vanishes at the infinity. In order to solve the system of partial differential equations given above, we employ the series expansion in terms of spherical harmonics to represent the angular dependence of the perturbed quantities. The Lagrangian displacement vector, $`\xi ^i`$ is expanded in terms of the vector spherical harmonics as $$\xi ^a=\underset{l|m|}{\overset{\mathrm{}}{}}aS_l(a)Y_l^m(\theta ,\phi )e^{i\sigma t},$$ (10) $$\xi ^\theta =\underset{l,l^{}|m|}{\overset{\mathrm{}}{}}\left\{H_l(a)Y_{l}^{m}{}_{,\theta }{}^{}+T_l^{}(a)\frac{1}{\mathrm{sin}\theta }Y_{l^{}}^{m}{}_{,\phi }{}^{}\right\}e^{i\sigma t},$$ (11) $$\xi ^\phi =\frac{1}{\mathrm{sin}^2\theta }\underset{l,l^{}|m|}{\overset{\mathrm{}}{}}\left\{H_l(a)Y_{l}^{m}{}_{,\phi }{}^{}T_l^{}(a)\mathrm{sin}\theta Y_{l^{}}^{m}{}_{,\theta }{}^{}\right\}e^{i\sigma t}$$ (12) (Regge & Wheeler 1957, see also Thorne 1980). The perturbed scalar quantity such as $`\delta \mathrm{\Phi }`$ is expressed as $$\delta \mathrm{\Phi }=\underset{l|m|}{\overset{\mathrm{}}{}}\delta \mathrm{\Phi }_l(a)Y_l^m(\theta ,\phi )e^{i\sigma t}.$$ (13) Here, the expansion coefficients $`T_l^{}`$ are called “axial” (or “toroidal”) and the other coefficients are called “polar” (or “spheroidal”). It is known that these two set of components are decoupled when $`\mathrm{\Omega }=0`$. Substituting the perturbed quantities expanded in terms of the spherical harmonics into the linearized basic equations (4), (5), (8) and (9), we obtain an infinite system of coupled linear ordinary differential equations for the expansion coefficients. Notice that the system of differential equations employed in this paper is arranged to give eigenfunctions and eigenfrequencies consistent up to the order of $`\overline{\mathrm{\Omega }}^3`$ for slow rotation. The explicit expressions for the governing equations are given in Appendix of Yoshida & Lee (2000). Since the equilibrium state of a rotating star is invariant under the parity transformation defined by $`\theta \pi \theta `$, the linear perturbations have definite parity for that transformation. In this paper, a set of modes whose scalar perturbations such as $`\delta \mathrm{\Phi }`$ are symmetric with respect to the equator is called “even” modes, while the other set of modes whose scalar perturbations are antisymmetric with respect to the equator is called “odd” modes (see, e.g., Berthomieu et al. 1978). For positive integers $`j=1,2,3,\mathrm{}`$, we have $`l=|m|+2j2`$ and $`l^{}=l+1`$ for even modes, and $`l=|m|+2j1`$ and $`l^{}=l1`$ for odd modes, where the symbols $`l`$ and $`l^{}`$ have been used to denote the spheroidal and toroidal components of the displacement vector $`\xi ^i`$, respectively (see equations (10)–(13)). Note that we do not use the term of the even (the odd) parity mode to refer to the polar (the axial) mode, although this terminology is used traditionally in relativistic perturbation theory (e.g., Regge & Wheeler 1957). Oscillations of rotating stars are also divided into prograde and retrograde modes. When observed in the corotating frame of the star, wave patterns of the prograde modes are propagating in the same direction as rotation of the star, whereas those of the retrograde modes are propagating in the opposite direction to the rotation. For positive values of $`m`$, the prograde and retrograde modes have negative and positive $`\omega `$, respectively. For numerical calculations, we truncate the infinite set of linear ordinary differential equations to obtain a finite set by discarding all the expansion coefficients associated with $`l`$ higher than $`l_{max}`$. This truncation determines the number of the expansion coefficients kept in the spherical harmonic expansion of each perturbed quantity. We denote this number as $`k_{max}`$. Our basic equation becomes a system of $`4\times k_{max}`$-th order linear ordinary differential equations, which, together with the boundary conditions, are numerically solved as an eigenvalue problem of $`\omega \sigma +m\mathrm{\Omega }`$ by using a Henyey type relaxation method (e.g., Unno et al . 1989, see also Press et al. 1992). We cannot determine a priori $`k_{max}`$, the number of the expansion coefficients kept in the eigenfunctions. In this paper, we start with an appropriate trial value of $`k_{\text{max}}`$ to compute an oscillation mode at a given value of $`\overline{\mathrm{\Omega }}`$. Then, we increase the value of $`k_{\text{max}}`$, until the eigenfrequency converges within an appropriate numerical error, where the mode identification is done by examining the form and the number of nodes of the dominant expansion coefficients. We consider that the mode thus obtained with the sufficient number of the expansion coefficients is the correct oscillation mode we want for the rotating model at $`\overline{\mathrm{\Omega }}`$. In order to check our numerical code, we calculate eigenfrequencies of the $`r`$-modes of the $`n=3`$ and $`\mathrm{\Gamma }=5/3`$ polytropic model at $`\overline{\mathrm{\Omega }}=0.1`$. In Table 1, our results are tabulated together with those calculated by Saio (1982) by using a semi-analytical perturbative approach. As shown by Table 1, our results are in good agreement with Saio’s results. ## 3 Rotational Modes of Slowly Rotating Non-isentropic Stars Because it is believed that isentropic structure is a good approximation for young and hot neutron stars, in this study we concentrate our attention to the case that the deviation of the non-isentropic models from isentropic structure is sufficiently small. In this case, we may assume that the adiabatic exponent $`\mathrm{\Gamma }`$ is given by $`{\displaystyle \frac{1}{\mathrm{\Gamma }}}={\displaystyle \frac{n}{n+1}}+\gamma ,`$ (14) where $`\gamma `$ is a constant giving the deviation from the isentropic structure. By substituting equation (14) into equation (7), we obtain an explicit form of $`A_i`$ as: $`A_i`$ $`=`$ $`\gamma _i\mathrm{ln}p.`$ (15) In this study, we only consider oscillations of polytropic models with the index $`n=1`$. Positive and negative values of $`\gamma `$ are respectively for convective and radiative models. Isentropic models are given by $`\gamma =0`$. Since our formulation for stellar oscillations of rotating stars is accurate only up to the order of $`\overline{\mathrm{\Omega }}^3`$, we shall apply our method to the oscillation modes in the range of $`\overline{\mathrm{\Omega }}0.1`$. For convenience of later discussions, it may be useful to review the properties of $`r`$-modes and inertial modes of isentropic models with $`\gamma =0`$ (see, e.g., Yoshida & Lee 2000). Because of the asymptotic behavior of the eigenfrequency of the rotational modes in the limit of $`\mathrm{\Omega }0`$, it is convenient to introduce a non-dimensional frequency $`\kappa `$ as follows: $$\frac{\omega }{\mathrm{\Omega }}=\kappa .$$ (16) Since the rotational deformation is of the order of $`\overline{\mathrm{\Omega }}^2`$, the effects on the frequency of the rotational modes appear as the terms of the order of $`\overline{\mathrm{\Omega }}^3`$. Therefore, we may expand the dimensionless eigenfrequency $`\kappa `$ of the rotational modes in powers of $`\overline{\mathrm{\Omega }}`$ as follows: $$\kappa =\kappa _0+\kappa _2\overline{\mathrm{\Omega }}^2+O(\overline{\mathrm{\Omega }}^4),$$ (17) where $$\kappa _0=\frac{2m}{l^{}(l^{}+1)}$$ (18) for the $`r`$-modes associated with the indices $`l^{}`$ and $`m`$ of vector harmonics with the axial parity (see, e.g. Papaloizou & Pringle 1978). For the inertial modes, there is no analytic expression for $`\kappa _0`$ in general. As suggested by Provost et al. (1981) and Saio (1982), the $`r`$-mode associated with $`l^{}=|m|`$ having the nodeless eigenfunction $`T_l^{}(r)`$ is the only $`r`$-mode permitted for isentropic stars for a given $`m`$. In the limit of $`\mathrm{\Omega }0`$, we have $`\kappa \kappa _0=2m/l^{}(l^{}+1)`$. Among the expansion coefficients of the eigenfunctions the toroidal component $`T_{l^{}=|m|}`$ is dominating. For the inertial modes, the value of $`\kappa `$ in the limit of $`\mathrm{\Omega }0`$ is also finite but different from $`\kappa _0=2m/l^{}(l^{}+1)`$. The inertial modes are characterized by the two angular quantum numbers $`l_0`$ and $`m`$, where the value of $`l_0`$ is the maximum index $`l`$ of spherical harmonics associated with the dominant expansion coefficients of the eigenfunctions (see Lockitch & Friedman 1999, Yoshida & Lee 2000). Note that our definition of $`l_0`$ is not the same as that of Lockitch & Friedman (1999), but the same as that of Lindblom & Ipser (1998) and Yoshida & Lee (2000). Note also that odd (even) parity modes have odd (even) values of $`l_0|m|`$, and that the odd parity mode having $`l_0|m|=1`$ is the $`r`$-mode associated with $`l^{}=|m|`$. The value of $`\kappa _0`$ for the inertial modes depends not only on the indices $`m`$ and $`l_0`$, but also on the equilibrium structure of stars (e.g., Yoshida & Lee 2000). For given values of $`l_0`$ and $`m`$, the number of different inertial modes is equal to $`l_0|m|(2)`$. However, there seems to be no good quantum number to classify the different inertial modes associated with the same $`l_0`$ and $`m`$. We here introduce an ordering integer $`n`$ to use the labellings such as $`i_n`$ and $`\omega _n`$ for these inertial modes. Since the frequencies of the inertial modes with the same $`l_0`$ and $`m`$ are found in a sequence given by, in the case of $`m>0`$, $$\omega _{(l_0|m|1)/2}<\mathrm{}<\omega _1<0\omega _0<\omega _1<\mathrm{}<\omega _{(l_0|m|1)/2}$$ (19) for odd parity modes, and $$\omega _{(l_0|m|)/2}<\mathrm{}<\omega _1<0<\omega _1<\mathrm{}<\omega _{(l_0|m|)/2}$$ (20) for even parity modes, we may define the ordering number $`n`$ as an integer that satisfies the inequality $`|l_0|m|1|/2|n|0`$ for odd parity inertial modes, and $`|l_0|m||/2|n|>0`$ for even parity inertial modes (see Lockitch & Friedman 1998, Yoshida & Lee 2000). We apply this classification to the inertial modes of both isentropic and non-isentropic stars in the following discussions. ### 3.1 $`r`$-modes For the radiative models, all the $`r`$-modes associated with $`l^{}|m|`$ are possible for a given $`m`$, and there are in principle a countable infinite number of $`r`$-modes for given $`l^{}`$ and $`m`$. The $`r`$-modes with the same $`l^{}`$ and $`m`$ may be classified by counting the number of nodes of the axial components $`T_l^{}`$ in the radial direction. For them the values of $`\kappa _2`$ are different, but the values of $`\kappa _0`$ are the same and equal to $`2m/l^{}(l^{}+1)`$. It is important to note that this classification of the $`r`$-modes is applicable only in the limit of $`\overline{\mathrm{\Omega }}0`$ and $`\kappa \kappa _0=2m/l^{}(l^{}+1)`$, since the components associated with the other values of $`l^{}`$ will come into play for large $`\overline{\mathrm{\Omega }}`$ (see, e.g. Provost et al. 1981, Saio 1982). Let us introduce an integer $`k`$ to denote the number of nodes of the toroidal eigenfunction $`T_l^{}`$, where the node at the center of the star is excluded from the count. Then, the $`r`$-modes are completely specified by the three quantum numbers $`m`$, $`l^{}`$ and $`k`$. In this paper, for given $`l^{}`$ and $`m`$, we employ the notation of $`r_k`$ to denote the $`r`$-mode whose toroidal eigenfunction $`T_l^{}(r)`$ has $`k`$ nodes in the radial direction. In Figures 1 to 3, the dimensionless eigenvalues $`\kappa `$ are given versus $`\overline{\mathrm{\Omega }}`$ for several low-radial-order $`r`$-modes with $`m=1`$, 2, and 3 for the $`n=1`$ polytropic model with $`\gamma =10^4`$, where the $`r`$-modes with $`l^{}=|m|`$, $`l^{}=|m|+1`$ and $`l^{}=|m|+2`$ are given in each figure. Modes with $`l^{}=|m|+1`$ are even parity modes, and those with $`l^{}=|m|`$ and $`l^{}=|m|+2`$ are odd parity modes. In the figures, the eigenvalues $`\kappa `$ of the $`i_0`$-modes with $`l_0=|m|+2k+1`$, calculated at $`\overline{\mathrm{\Omega }}=0.1`$ for the $`n=1`$ isentropic model, are designated by the filled circles, where the integer $`k`$ corresponds to the number in the subscript of the notation $`r_k`$. Figures 1 to 3 clearly show that the qualitative behavior of the eigenfrequencies as functions of $`\overline{\mathrm{\Omega }}`$ is not strongly dependent on the azimuthal quantum number $`m`$. From these figures we also understand that the $`r`$-modes of the radiative model with $`\gamma <0`$ are divided into three distinctive classes according to the asymptotic behavior of the eigenvalues at large $`\overline{\mathrm{\Omega }}`$. The first class of the $`r`$-modes consists of the nodeless $`r_0`$-modes associated with $`l^{}=|m|`$. The eigenvalue $`\kappa `$ of the $`r`$-modes of this class is well approximated by equations (17) and (18) at any values of $`\overline{\mathrm{\Omega }}`$ examined in this paper. Note that the $`r_0`$-mode with $`m=l^{}=1`$ is a peculiar mode in the sense that $`\kappa `$ is almost constant since $`\kappa _2=0`$ and $`\kappa =\kappa _0+O(\overline{\mathrm{\Omega }}^4)`$ (see Saio 1982, Greenspan 1964). The second class consists of the $`r_k`$-modes with $`l^{}=|m|`$ and $`k>0`$. Equations (17) and (18) can give good values for the eigenvalue $`\kappa `$ of the $`r`$-modes of this class only when $`\overline{\mathrm{\Omega }}`$ is sufficiently small. As $`\overline{\mathrm{\Omega }}`$ increases, $`\kappa `$ tends to the $`i_0`$ inertial modes with $`l_0=|m|+2k+1`$ of the isentropic model. Note that the $`i_0`$-modes are odd parity inertial modes (see equations (19) and (20)). We therefore consider that at large $`\overline{\mathrm{\Omega }}`$ the $`r_k`$-modes of this class become identical with the $`i_0`$-modes with $`l_0=|m|+2k+1`$, for which $`\kappa `$ is almost constant. The third class of the $`r`$-modes consists of the $`r_k`$-modes with $`l^{}>|m|`$ and $`k0`$. The eigenvalue $`\kappa `$ of the $`r`$-modes of this class is well represented by equations (17) and (18) only when $`\overline{\mathrm{\Omega }}`$ is sufficiently small, and it becomes small monotonically as $`\overline{\mathrm{\Omega }}`$ increases. The $`r`$-modes of this class may have no corresponding inertial modes of the isentropic model. As proved by Friedman & Schutz (1978), oscillation modes whose frequency $`\sigma `$ observed in an inertial frame satisfies the condition $`\sigma (\sigma +m\mathrm{\Omega })=\sigma \omega <0`$ are unstable to the gravitational radiation reaction. The condition for the instability may be rewritten as $`0<\kappa <m`$ for retrograde modes with $`\omega >0`$ and $`m>0`$. Figures 1 to 3 show that the $`r`$-mode with $`l^{}2`$ satisfies the inequality $`0<\kappa <m`$ for the range of $`\overline{\mathrm{\Omega }}`$ examined in this paper. It is also found that the $`r_k`$-modes with $`l^{}=1`$ and $`k>0`$ also satisfy the inequality $`0<\kappa <m`$ for $`\overline{\mathrm{\Omega }}>0`$. The reason for the instability of the $`r_k`$-modes with $`l^{}=1`$ and $`k>0`$ may be understood by noting that the polar components with the index $`l=2`$ can be responsible for the emission of gravitational radiation. We may conclude that the $`r_0`$-mode with $`l^{}=|m|=1`$ is the only $`r`$-mode that is stable against the gravitational radiation reaction. In Figure 4, the eigenfrequencies $`\kappa `$ of three $`r`$-modes with $`m=2`$ are plotted as functions of $`|\gamma |`$, where we have assumed $`\overline{\mathrm{\Omega }}=0.01`$. The three modes are respectively the $`r_0`$-mode with $`l^{}=m=2`$, the $`r_1`$-mode with $`l^{}=m=2`$, and the $`r_0`$-mode with $`l^{}=3`$ and $`m=2`$. These three $`r`$-modes belong to the first, second, and third classes of the $`r`$-modes discussed in the previous paragraph. The solid lines and the dotted lines are used to denote the cases of negative and positive $`\gamma `$, respectively. Note that although we can obtain the three $`r`$-modes without any difficulty for the radiative models with negative $`\gamma `$, we cannot always calculate the $`r`$-modes of the second and the third classes for the convective models with positive $`\gamma `$. Let us first discuss the case of the radiative models. As shown by Figure 4, the eigenvalue $`\kappa `$ of the $`r_0`$-mode with $`l^{}=|m|=2`$ is practically independent of $`|\gamma |`$ and has values approximately equal to $`\kappa _0=2m/l^{}(l^{}+1)=2/3`$. However, $`\kappa `$ of the $`r_1`$-mode with $`l^{}=|m|=2`$ is dependent on $`|\gamma |`$. At small $`|\gamma |`$, the $`r_1`$-mode has the eigenvalue close to $`\kappa _0=0.517`$, which corresponds to the inertial mode with $`l_0|m|=3`$ of the isentropic model, but it tends to $`2m/l^{}(l^{}+1)=2/3`$ as $`|\gamma |`$ increases. The $`r_0`$-mode with $`l^{}=3`$ and $`m=2`$ also has the eigenfrequency $`\kappa `$ which depends on $`|\gamma |`$. Although the value of $`\kappa `$ is close to $`\kappa _0=2m/l^{}(l^{}+1)=1/3`$ at large $`|\gamma |`$, it tends to zero as $`|\gamma |0`$. This behavior of $`\kappa 0`$ in the limit of $`|\gamma |0`$ is consistent with the fact that the $`r`$-modes associated with $`l^{}>|m|`$ do not exist in isentropic stars, for which $`\gamma =0`$. Let us next discuss the case of the convective models with $`\gamma >0`$. The eigenvalue $`\kappa `$ of the $`r_0`$-mode with $`l^{}=|m|=2`$ is only weakly dependent on $`\gamma `$, as in the case of the radiative models. Note that the curves for the $`r_0`$-mode with $`l^{}=|m|=2`$ for negative and positive $`\gamma `$ are almost indistinguishable in Figure 4. The $`r_1`$-mode with $`l^{}=|m|=2`$ for the convective models behaves differently than that for the radiative models as $`|\gamma |`$ increases. In fact, the value of $`\kappa `$ decreases rapidly with increasing $`\gamma `$ and the $`r_1`$-mode cannot be found at large values of $`\gamma >0`$. No $`r_0`$-mode with $`l^{}=3`$ and $`m=2`$ are found for the convective models. Let us discuss the properties of the eigenfunctions of the $`r`$-modes for the radiative models. Since the amplitude of the displacement vector is much larger than that of the scalar perturbations, in the following discussions we may show only the displacement vector of the modes. In Figures 5 to 7, we show the expansion coefficients $`S_l`$, $`H_l`$ and $`iT_l^{}`$ as functions of the fractional radius $`a/R`$ for the $`r`$-modes with $`m=2`$, where we have assumed $`\gamma =10^4`$, and the amplitude is normalized at the surface such that $`iT_l^{}=1`$. Here, $`l=|m|`$ and $`l^{}=l+1`$ for even modes, and $`l=|m|+1`$ and $`l^{}=l1`$ for odd modes. The modes given in Figures 5 to 7 are respectively the $`r_0`$-mode with $`l^{}=2`$, the $`r_1`$-mode with $`l^{}=2`$, and the $`r_0`$-mode with $`l^{}=3`$. In each figure, the eigenfunctions for the cases of $`\overline{\mathrm{\Omega }}=10^3`$ and $`\overline{\mathrm{\Omega }}=10^1`$ are shown in panels $`(a)`$ and $`(b)`$, respectively. The panels $`(a)`$ of Figures 5 to 7 show that the toroidal component $`iT_l^{}`$ of the $`r`$-modes is dominating the other expansion coefficients when $`\overline{\mathrm{\Omega }}`$ is small. However, as shown by the panels (b), although $`iT_l^{}`$ of the $`r_0`$-mode with $`l^{}=2`$ remains dominant even at $`\overline{\mathrm{\Omega }}=0.1`$, the toroidal components $`iT_l^{}`$ of the $`r_1`$-mode with $`l^{}=2`$ and the $`r_0`$-mode with $`l^{}=3`$ are not necessarily dominating any more at $`\overline{\mathrm{\Omega }}=0.1`$. The properties of the eigenfunctions of the $`r_1`$-mode with $`l^{}=m=2`$ at large $`\overline{\mathrm{\Omega }}`$ may be regarded as those of the inertial modes. It is interesting to note that the displacement vector of the $`r_0`$-mode with $`l^{}=3`$ has large amplitude in the deep interior and almost vanishing amplitude near the stellar surface at large $`\overline{\mathrm{\Omega }}`$. This suggests that because of the low frequencies $`\omega `$ in the corotating frame the buoyant force is as important as the Coriolis force to determine the mode character at large $`\overline{\mathrm{\Omega }}`$. ### 3.2 Inertial modes For the radiative models with the polytropic index $`n=1`$ and $`\gamma =10^4`$, the eigenfrequencies $`\overline{\omega }\omega /(GM/R^3)^{1/2}`$ of low frequency oscillations are shown as functions of $`\overline{\mathrm{\Omega }}`$ for $`m=1`$, 2, and 3 in Figures 8 to 10, where even and odd parity modes are displayed in panels (a) and (b), respectively. The straight dashed lines drawn in the figures are given by $`\overline{\omega }=0`$ and $`\overline{\omega }=m\overline{\mathrm{\Omega }}`$ (see below). In the figures, the inertial modes of the $`n=1`$ isentropic model computed at $`\overline{\mathrm{\Omega }}=0.1`$ are designated by the filled circles. We note that the oscillation modes in a mode sequence are regarded as inertial modes when $`\overline{\mathrm{\Omega }}`$ is large, and as internal gravity ($`g`$-) modes when $`\overline{\mathrm{\Omega }}`$ is small. If we follow one of the continuous mode sequences from large $`\overline{\mathrm{\Omega }}`$ to small $`\overline{\mathrm{\Omega }}`$, we may physically understand the behavior of the low frequency oscillation in the radiative models as a function of $`\overline{\mathrm{\Omega }}`$. At large $`\overline{\mathrm{\Omega }}`$, the Coriolis force is dominating as the restoring force and the oscillation mode is identified as an inertial mode. As $`\overline{\mathrm{\Omega }}`$ decreases, however, the buoyant force becomes dominating the Coriolis force and the mode becomes identical with a $`g`$-mode in the limit of $`\overline{\mathrm{\Omega }}0`$. Because of the effects of the buoyant force in the radiative model, the frequency $`\overline{\omega }`$ of the inertial modes deviates from that of the inertial modes of the isentropic model as $`\overline{\mathrm{\Omega }}`$ decreases. The modes within the two straight dashed lines depicted in Figures 8, 9 and 10 are unstable against the gravitational radiation driven instability in the sense that they satisfy the frequency condition $`\sigma (\sigma +m\mathrm{\Omega })<0`$ (Friedman & Schutz 1978). We can see from Figure 8 that the $`i_k`$-modes with $`k0`$ and $`m=1`$ are stable against the instability for all $`\overline{\mathrm{\Omega }}`$. Figures 9 and 10 show that as $`\overline{\mathrm{\Omega }}`$ increases all the retrograde $`g`$-modes with $`|m|2`$, which are stable against the instability when $`\overline{\mathrm{\Omega }}0`$, become retrograde inertial modes, which are unstable against the instability at large $`\overline{\mathrm{\Omega }}`$. There are several cases where we find it practically impossible to obtain completely continuous mode sequences between the $`g`$-modes at small $`\overline{\mathrm{\Omega }}`$ and the inertial modes at large $`\overline{\mathrm{\Omega }}`$. In such sequences, there exist a domain of $`\overline{\mathrm{\Omega }}`$, in which there appear a lot of modes having a large number of nodes of the dominant expansion coefficients, and it is difficult to identify the same oscillation modes as those computed in the previous steps in $`\overline{\mathrm{\Omega }}`$. In Figures 8 to 10, the domains of $`\overline{\mathrm{\Omega }}`$ with this difficulty are indicated by the dotted lines in the mode sequences. Unfortunately, we cannot remove the difficulty by increasing $`k_{max}`$. One of the reasons for the difficulty may be attributed to the method of calculation we employ in this paper. As discussed in Unno et al (1989), since we have truncated the infinite system of differential equations to obtain a finite system of equations, the numerical procedure used here inevitably contains a process of inverting matrices which are singular at some value of the ratio $`2\mathrm{\Omega }/\omega 1`$, depending on the oscillation mode we are interested in. In this paper, we assume that there always exist a continuous mode sequence between the $`g`$-mode and inertial mode as a function of $`\overline{\mathrm{\Omega }}`$, even if the mode sequence contains a domain of $`\overline{\mathrm{\Omega }}`$ in which the modes we are interested in cannot be obtained numerically. ### 3.3 Connection rules between the $`g`$-modes, $`r`$-modes and inertial modes In order to find the connection rule between the $`g`$-modes at small $`\overline{\mathrm{\Omega }}`$ and the inertial modes at large $`\overline{\mathrm{\Omega }}`$, we tabulate first the frequencies of the $`g`$-modes at $`\overline{\mathrm{\Omega }}=0`$ for the radiative model with the index $`n=1`$ and $`\gamma =10^4`$ in Table 2. To obtain a clear understanding of the global correspondence between the $`g`$-, $`r`$-, and inertial modes, with the help of Table 2 and Figures 8 to 10, we give schematic diagrams for the connection rules between the several low-radial-order modes as follows: For odd parity modes we have $`\left[\begin{array}{cccc}& & & \mathrm{}\hfill \\ & & i_2(l_0=|m|+5)\hfill & \mathrm{}\hfill \\ & i_1(l_0=|m|+3)\hfill & i_1(l_0=|m|+5)\hfill & \mathrm{}\hfill \\ i_0(l_0=|m|+1)\hfill & i_0(l_0=|m|+3)\hfill & i_0(l_0=|m|+5)\hfill & \mathrm{}\hfill \\ & i_1(l_0=|m|+3)\hfill & i_1(l_0=|m|+5)\hfill & \mathrm{}\hfill \\ & & i_2(l_0=|m|+5)\hfill & \mathrm{}\hfill \\ & & & \mathrm{}\hfill \end{array}\right]`$ $``$ $`\left[\begin{array}{cccc}& & & \mathrm{}\hfill \\ & & g_1(l_{}=|m|+3)\hfill & \mathrm{}\hfill \\ & g_1(l_{}=|m|+1)\hfill & g_2(l_{}=|m|+1)\hfill & \mathrm{}\hfill \\ r_0(l_{}^{}=|m|)\hfill & r_1(l_{}^{}=|m|)\hfill & r_2(l_{}^{}=|m|)\hfill & \mathrm{}\hfill \\ & g_1(l_{}=|m|+1)\hfill & g_2(l_{}=|m|+1)\hfill & \mathrm{}\hfill \\ & & g_1(l_{}=|m|+3)\hfill & \mathrm{}\hfill \\ & & & \mathrm{}\hfill \end{array}\right]`$ $`,`$ (36) and for even parity modes we have $`\left[\begin{array}{cccc}& & & \mathrm{}\hfill \\ & & i_3(l_0=|m|+6)\hfill & \mathrm{}\hfill \\ & i_2(l_0=|m|+4)\hfill & i_2(l_0=|m|+6)\hfill & \mathrm{}\hfill \\ i_1(l_0=|m|+2)\hfill & i_1(l_0=|m|+4)\hfill & i_1(l_0=|m|+6)\hfill & \mathrm{}\hfill \\ i_1(l_0=|m|+2)\hfill & i_1(l_0=|m|+4)\hfill & i_1(l_0=|m|+6)\hfill & \mathrm{}\hfill \\ & i_2(l_0=|m|+4)\hfill & i_2(l_0=|m|+6)\hfill & \mathrm{}\hfill \\ & & i_3(l_0=|m|+6)\hfill & \mathrm{}\hfill \\ & & & \mathrm{}\hfill \end{array}\right]`$ $``$ $`\left[\begin{array}{cccc}& & & \mathrm{}\hfill \\ & & g_1(l_{}=|m|+4)\hfill & \mathrm{}\hfill \\ & g_1(l_{}=|m|+2)\hfill & g_2(l_{}=|m|+2)\hfill & \mathrm{}\hfill \\ g_1(l_{}=|m|)\hfill & g_2(l_{}=|m|)\hfill & g_3(l_{}=|m|)\hfill & \mathrm{}\hfill \\ g_1(l_{}=|m|)\hfill & g_2(l_{}=|m|)\hfill & g_3(l_{}=|m|)\hfill & \mathrm{}\hfill \\ & g_1(l_{}=|m|+2)\hfill & g_2(l_{}=|m|+2)\hfill & \mathrm{}\hfill \\ & & g_1(l_{}=|m|+4)\hfill & \mathrm{}\hfill \\ & & & \mathrm{}\hfill \end{array}\right]`$ $`,`$ (54) where the two corresponding modes occupy the same position in the two matrices connected with an arrow. From the diagrams, we may find the connection rule between the $`g_{\pm k}`$-modes with $`l`$ and $`m`$ and the inertial modes $`i_{\pm j}`$ with $`l_0`$ and $`m`$: $$g_{\pm k}(l=l^j)i_{\pm j}(l_0=l^j+2k),$$ (55) where $`l^j=|m|+2j2`$ for even modes and $`l^j=|m|+2j1`$ for odd modes, and $`j`$ and $`k`$ are positive integers. The symbol $`g_{\pm k}`$ denotes the $`g`$-mode with $`k`$ radial nodes of the expansion coefficient $`S_l`$, having positive and negative frequencies $`\omega `$ observed in the corotating frame of the star when $`\overline{\mathrm{\Omega }}0`$. The connection rule between the $`r`$-modes associated with $`l^{}=|m|`$ and the inertial modes $`i_0`$ with $`l_0`$ and $`m`$ may be given by $$r_{j1}(l^{}=|m|)i_0(l_0=|m|+2j1),$$ (56) where $`j`$ is a positive integer. Note that the $`i_0`$-mode with $`l_0=|m|+1`$ is exactly the same as the $`r_0`$-modes with $`l^{}=|m|`$. To establish the connection rules given above, we have neglected the effects of avoided crossings between the modes associated with different values of $`l`$ and/or $`k`$, since the mode property is carried over at the crossings. ## 4 Dissipation Timescales for $`r`$-modes of Non-isentropic Stars For the rotational modes of non-isentropic models, we have found that all the $`r`$-modes except the $`r_0`$-mode with $`l^{}=|m|=1`$ are unstable to the gravitational radiation reaction in the sense that the frequency $`\sigma `$ satisfies the condition $`\sigma (\sigma +m\mathrm{\Omega })<0`$. Let us determine the stability of the $`r`$-modes of the non-isentropic models against the gravitational radiation driven instability, taking account of the dissipative processes due to the shear and bulk viscosity, where we will follow the method of calculation almost the same as that employed in the previous stability analyses of the $`r_0`$-mode with $`l^{}=|m|`$ and the inertial modes for isentropic models (Lindblom et al 1998, Owen et al 1998, Andersson et al 1998, Kokkotas & Stergioulas 1998, Lindblom et al 1999, Lockitch & Friedman 1999, Yoshida & Lee 2000). The damping timescale of the $`r_0`$-mode associated with $`l^{}=|m|`$ for sufficiently small $`\overline{\mathrm{\Omega }}`$ may be estimated by (Lindblom et al 1998) $`{\displaystyle \frac{1}{\tau }}`$ $`=`$ $`{\displaystyle \frac{1}{\stackrel{~}{\tau }_S}}\left({\displaystyle \frac{10^9K}{T}}\right)^2+{\displaystyle \frac{1}{\stackrel{~}{\tau }_B}}\left({\displaystyle \frac{T}{10^9K}}\right)^6\left({\displaystyle \frac{\mathrm{\Omega }^2}{\pi G\overline{\rho }}}\right)`$ (57) $`+`$ $`{\displaystyle \underset{l=2}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{\stackrel{~}{\tau }_{J,l}}}\left({\displaystyle \frac{\mathrm{\Omega }^2}{\pi G\overline{\rho }}}\right)^{l+1}+{\displaystyle \underset{l=2}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{\stackrel{~}{\tau }_{D,l}}}\left({\displaystyle \frac{\mathrm{\Omega }^2}{\pi G\overline{\rho }}}\right)^{l+2},`$ where $`\overline{\rho }`$ is the average density of the star. The dissipative timescales $`\stackrel{~}{\tau }`$ are calculated for each oscillation mode in the limit of $`\overline{\mathrm{\Omega }}0`$ so that they are independent of the rotation frequency $`\mathrm{\Omega }`$. Here, the first, second, third and fourth terms in the right-hand side of equation (57) are contributions from the shear viscosity, the bulk viscosity, the current multipole radiation and the mass multipole radiation, respectively. The expression (57) of the damping timescales has been derived on the assumptions that the frequency of the modes is well represented by a linear function of $`\mathrm{\Omega }`$ and the toroidal component of the displacement vector is dominating. Although the assumptions are reasonable for the $`r_0`$-modes with $`l^{}=|m|`$ (see Yoshida and Lee 2000), they are not necessarily correct for the $`r_k`$-modes with $`l^{}=|m|`$ and $`k>0`$, for which we use $`{\displaystyle \frac{1}{\tau }}`$ $`=`$ $`{\displaystyle \frac{1}{\stackrel{~}{\tau }_S}}\left({\displaystyle \frac{10^9K}{T}}\right)^2+{\displaystyle \frac{1}{\stackrel{~}{\tau }_B}}\left({\displaystyle \frac{T}{10^9K}}\right)^6\left({\displaystyle \frac{\pi G\overline{\rho }}{\mathrm{\Omega }^2}}\right)`$ (58) $`+`$ $`{\displaystyle \underset{l=2}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{\stackrel{~}{\tau }_{J,l}}}\left({\displaystyle \frac{\mathrm{\Omega }^2}{\pi G\overline{\rho }}}\right)^{l+1}+{\displaystyle \underset{l=2}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{\stackrel{~}{\tau }_{D,l}}}\left({\displaystyle \frac{\mathrm{\Omega }^2}{\pi G\overline{\rho }}}\right)^{l+2},`$ because these $`r`$-modes become inertial modes at large $`\overline{\mathrm{\Omega }}`$. The difference between these two expressions is found in the second terms in the right-hand side of the two equations. When $`\overline{\mathrm{\Omega }}`$ is small, equation (58) gives the correct expression of $`\tau ^1`$ for inertial modes of isentropic stars (See, Lockitch & Friedman 1999, Yoshida & Lee 2000). To determine the damping timescales $`\stackrel{~}{\tau }`$ in (58) in the limit of $`\overline{\mathrm{\Omega }}0`$, we assume that the dissipative timescale $`\stackrel{~}{\tau }^1`$ depends on $`\overline{\mathrm{\Omega }}`$ as $`a\overline{\mathrm{\Omega }}^2+b`$ for each dissipative process, where the two unknown constants $`a`$ and $`b`$ are determined by calculating $`\stackrel{~}{\tau }^1`$ and $`d\stackrel{~}{\tau }^1/d\overline{\mathrm{\Omega }}`$ at $`\overline{\mathrm{\Omega }}=0.1`$. The $`\stackrel{~}{\tau }^1`$ in the limit of $`\overline{\mathrm{\Omega }}0`$ is given by $`\stackrel{~}{\tau }^1=b`$. We note that equation (58) is not necessarily a correct approximation for non-isentropic stars. We use the expression (58) simply because we want to give rough estimates of the dissipative timescales, in order to compare the timescales with those of the isentropic case. In Table 3, we tabulate the dissipation timescales $`\stackrel{~}{\tau }`$ in the unit of second for the various dissipative processes for the $`r_k`$-modes with $`l^{}=|m|`$, where the radius and the mass of the $`n=1`$ polytropic neutron star model at $`\mathrm{\Omega }=0`$ are chosen to be $`R=12.57\text{km}`$ and $`M=1.4M_{\mathrm{}}`$, respectively. Note that in Table 3 we have not included the $`r_k`$-modes with $`l^{}>|m|`$, since the frequency $`\omega `$ becomes vanishingly small at large $`\overline{\mathrm{\Omega }}`$ (see Figures 1 to 3) and hence the magnitude of the gravitational radiation driven instability, which is proportional to $`\omega `$, becomes quite small. For the $`r_0`$-modes with $`l^{}=|m|`$, in order to see the effects of the buoyancy force on the stability, the results for the four different values of $`\gamma `$ are tabulated. As shown by Table 3, the gravitational radiation driven instability for the $`r_0`$-modes with $`l^{}=|m|`$ is not affected by introducing the deviation from isentropic structure and remains strong even for the non-isentropic models. As for the $`r_k`$-modes with $`l^{}=|m|`$ and $`k>0`$, it is found that the dissipative timescales are comparable to those of the corresponding inertial modes (compare Table 4 of Yoshida & Lee 2000). Thus, we may conclude that the $`r`$-mode instability driven by the gravitational radiation reaction is dominated by the $`r_0`$-mode with $`l^{}=|m|=2`$ both in isentropic and in non-isentropic stars. ## 5 Conclusions In this paper, we have numerically investigated the properties of rotational modes for slowly rotating, non-isentropic, Newtonian stars, where the effects of the centrifugal force are taken into consideration to calculate the modes with accuracy up to the order of $`\overline{\mathrm{\Omega }}^3`$. Estimating the dissipation timescales due to the gravitational radiation and the shear and bulk viscosity for simple polytropic neutron star models with the index $`n=1`$, we have shown that the gravitational radiation driven instability of the $`r_0`$-mode with $`l^{}=|m|`$ is hardly affected by introducing the deviation from the isentropic structure, and that the instability of the $`r_0`$-modes with $`l^{}=|m|=2`$ remains strongest even for the non-isentropic models. Note that the $`r_0`$-modes with $`l^{}=|m|`$ are the only $`r`$-modes possible in isentropic stars and their modal property does not very much depend on $`\mathrm{\Omega }`$ and the deviation from isentropic structure. We thus conclude that the result of the stability analysis of the $`r_0`$-modes with $`l^{}=|m|`$ for isentropic stars remains valid even for non-isentropic stars. We have numerically found that for the radiative models the inertial modes at large $`\overline{\mathrm{\Omega }}`$ become identical with the $`r_k`$-modes with $`l^{}=|m|`$ and $`k>0`$ and $`g`$-modes in the limit of $`\overline{\mathrm{\Omega }}0`$ because of the effect of the buoyant force, where the integer $`k`$ denotes the number of node of the eigenfunction with dominant amplitude. This behavior of the $`g`$-modes or the inertial modes as functions of $`\mathrm{\Omega }`$ has been suggested by, for example, Saio (1999) and Lockitch & Friedman (1999). We find the connection rules between the $`r_k`$-modes with $`l^{}=|m|`$ and $`k>0`$ and $`g`$-modes at small $`\overline{\mathrm{\Omega }}`$ and the inertial modes at large $`\overline{\mathrm{\Omega }}`$. We also suggest that the $`r_k`$-modes with $`l^{}>|m|`$ and $`k0`$, whose frequency in the corotating frame becomes vanishingly small as $`\overline{\mathrm{\Omega }}`$ increases, have at large $`\overline{\mathrm{\Omega }}`$ no corresponding inertial modes of the isentropic models. Good progress has been made in the understanding of the properties of the $`r`$-mode instability due to the gravitational radiation since its discovery. Recently, for example, Yoshida et al (1999) have shown that the results of the $`r`$-mode instability obtained by using the slow rotation approximation, as employed in this paper, are consistent with the results obtained by calculating directly the $`r`$-mode oscillations of rotating stars without the approximation. Kojima (1998) and Kojima & Hosonuma (1999) have studied the time evolutional behavior of $`r`$-modes by using the Laplace transformation technique for slowly rotating relativistic models. Lockitch (1999) also obtained $`r`$-modes and rotation modes for slowly rotating, incompressible, relativistic stars. However, we believe it necessary to show that the gravitational radiation driven instability of $`r`$-modes remains sufficiently robust in more realistic and unrestricted situations, taking account of the effects of differential rotation, magnetic field and general relativity, for example. To investigate the $`r`$-mode instability in more general situations is one of our future studies. We would like to thank Prof. H. Saio for useful comments and discussions. S.Y. was supported by Research Fellowship of the Japan Society for the Promotion of Science for Young Scientists.
warning/0002/gr-qc0002057.html
ar5iv
text
# Cauchy-characteristic matching for a family of cylindrical solutions possessing both gravitational degrees of freedom ## 1 Introduction This paper is part of a long term program to Cauchy-characteristic matching (CCM) codes as investigative tools in numerical tools in numerical relativity. For a review of CCM see the article of Winicour. The approach has two distinct features: (i) it dispenses with an outer boundary condition and replaces this with matching conditions at an interface between the Cauchy and characteristic regions, and (ii) by employing a compactified coordinate, it proves possible to generate global solutions. This paper is continuation of work investigating CCM in systems with vacuum cylindrical symmetry (as a prototype system containing only one spatial degree of freedom). In we developed the necessary machinery and in we applied the approach to the exact solution of Weber and Wheeler which has one degree of polarization and so only possesses one gravitational degree of freedom. An agreement with the exact solution was found with a maximum error better then 1 part in $`10^3`$. In the same paper we also investigated the propagation of Gaussian wave packets possessing two gravitational degrees of freedom, but this work did not involve a check against any exact solution. Such a test is likely to be a more rigorous one because, unlike the Weber-Wheeler case, it would involve passing derivative information across the interface. One of the major problems in numerical relativity is that there are very few exact solutions known which can be used to test numerical codes. In this paper CCM is applied to an exact two-parameter family of cylindrically symmetric vacuum solutions possessing both gravitational degrees of freedom due to Piran, Safier and Katz . This family is somewhat unphysical because the rotational degree of freedom diverges at future null infinity. Nonetheless, it can be used to test the CCM cylindrical code more rigorously because it involves passing derivative information across the interface. The previous CCM cylindrical code was constructed specifically for solutions which (after using Geroch decomposition to factor out the $`z`$-direction) are asymptotically flat and therefore can not be used for processing the Piran et al family. However, by working directly in terms of the Geroch potential it proves possible to develop a modified version of the code which is able to process the family. In section 2 we present the family of solutions and briefly discuss its properties. In section 3 we review the derivation of the field equations in the characteristic region for solutions which, after factoring out the $`z`$-direction, are asymptotically flat. In section 4 we discuss the asymptotic limit of the Piran et al family and derive the modified field equations and the interface equations. An appendix discusses the regular singular nature of two of the resulting equations. In section 5 we briefly discuss the results of the modified code. ## 2 The two-parameter family of solutions Piran et al derive their metric by an indirect method starting from the Kerr line element in Boyer-Lindquist coordinates $`(\stackrel{~}{t},\stackrel{~}{r},\stackrel{~}{\theta },\stackrel{~}{\varphi })`$, namely $$ds^2=\frac{\mathrm{\Delta }}{\rho ^2}(d\stackrel{~}{t}\stackrel{~}{a}\mathrm{sin}^2\stackrel{~}{\theta }d\stackrel{~}{\varphi })^2+\frac{\mathrm{sin}^2\stackrel{~}{\theta }}{\rho ^2}[(\stackrel{~}{r}^2+\stackrel{~}{a}^2)d\stackrel{~}{\varphi }\stackrel{~}{a}d\stackrel{~}{t}]^2+\frac{\rho ^2}{\mathrm{\Delta }}d\stackrel{~}{r}^2+\rho ^2d\stackrel{~}{\theta }^2,$$ (1) where $`\rho ^2`$ $`=`$ $`\stackrel{~}{r}^2+\stackrel{~}{a}^2\mathrm{cos}^2\stackrel{~}{\theta },`$ (2) $`\mathrm{\Delta }`$ $`=`$ $`\stackrel{~}{r}^22m\stackrel{~}{r}+\stackrel{~}{a}^2.`$ (3) They first transform to a new isotropic radial coordinate $`R`$ where $$\stackrel{~}{r}=m+R+\frac{m^2\stackrel{~}{a}^2}{4R},$$ (4) and then to cylindrical coordinates $`(\stackrel{~}{t},\stackrel{~}{\rho },\stackrel{~}{Z},\stackrel{~}{\varphi })`$ where $`\stackrel{~}{\rho }`$ $`=`$ $`R\mathrm{sin}\stackrel{~}{\theta },`$ (5) $`\stackrel{~}{Z}`$ $`=`$ $`R\mathrm{cos}\stackrel{~}{\theta }.`$ (6) Employing the complex trick of sending $$\stackrel{~}{t}iZ,\stackrel{~}{\rho }\stackrel{~}{\rho },\stackrel{~}{Z}i\stackrel{~}{t},\stackrel{~}{\varphi }\stackrel{~}{\varphi },\stackrel{~}{a}ia,$$ (7) they then introduce new cylindrical coordinates $`(t,r,z,\varphi )`$ given by $`t`$ $`=`$ $`{\displaystyle \frac{\stackrel{~}{t}}{\alpha }}\left[1+{\displaystyle \frac{m^2+a^2}{4(\stackrel{~}{\rho }^2\stackrel{~}{t}^2)}}\right],`$ (8) $`r`$ $`=`$ $`{\displaystyle \frac{\stackrel{~}{\rho }}{\alpha }}\left[1{\displaystyle \frac{m^2+a^2}{4(\stackrel{~}{\rho }^2\stackrel{~}{t}^2)}}\right],`$ (9) $`z`$ $`=`$ $`Z2a^1m(m+\sqrt{m^2+a^2})\stackrel{~}{\varphi },`$ (10) $`\varphi `$ $`=`$ $`\alpha \stackrel{~}{\varphi },`$ (11) where $$\alpha =\frac{\sqrt{m^2+a^2}}{a},$$ (12) and the resulting line element is in the Jordan-Ehlers-Kompaneets form for a cylindrically symmetric spacetime , $$ds^2=e^{2\gamma 2\psi }(dt^2dr^2)+e^{2\psi }(dz+\omega d\varphi )^2+e^{2\psi }r^2d\varphi ^2.$$ (13) If we finally introduce coordinates $`u`$ $`=`$ $`tr,`$ (14) $`v`$ $`=`$ $`t+r,`$ (15) then we can express the two-parameter family of solutions in the form $$ds^2=e^{2\gamma 2\psi }dudv+e^{2\psi }(dz+\omega d\varphi )^2+\frac{e^{2\psi }(vu)^2}{4}d\varphi ^2,$$ (16) where explicitly $`e^{2\gamma }`$ $`=`$ $`{\displaystyle \frac{(\lambda _u+\lambda _v)^2+\alpha ^2(1\lambda _u\lambda _v)^2}{(1+\lambda _u^2)(1+\lambda _v^2)}},`$ (17) $`e^{2\psi }`$ $`=`$ $`{\displaystyle \frac{\alpha ^2(1\lambda _u\lambda _v)^2+(\lambda _u+\lambda _v)^2}{\alpha ^2\mathrm{\Xi }^2+(\lambda _u\lambda _v)^2}},`$ (18) $`\omega `$ $`=`$ $`{\displaystyle \frac{2a\sqrt{\alpha ^21}}{\alpha }}(\alpha +\sqrt{\alpha ^21})`$ $`{\displaystyle \frac{a\sqrt{\alpha ^21}\mathrm{\Xi }(\lambda _u+\lambda _v)^2}{\sqrt{\lambda _u\lambda _v}[\alpha ^2(1\lambda _u\lambda _v)^2+(\lambda _u+\lambda _v)^2]}},`$ (19) with $`\mathrm{\Xi }`$ $`=`$ $`1+\lambda _u\lambda _v+2\alpha ^1\sqrt{\alpha ^21}\sqrt{\lambda _u\lambda _v},`$ (20) $`\lambda _u`$ $`=`$ $`\left(\sqrt{a^2+u^2}u\right)/a,`$ (21) $`\lambda _v`$ $`=`$ $`\left(\sqrt{a^2+v^2}+v\right)/a.`$ (22) The coordinates $`u`$ and $`v`$ are null, $`\varphi `$ is the canonical azimuthal coordinate and $`z`$ lies along the axis of symmetry. The two parameters are $`a`$ $`(0a<\mathrm{})`$ which is a length scale and $`\alpha `$ $`(1\alpha <\mathrm{})`$ which is a measure of the total energy of the system, with $`\alpha =1`$ corresponding to flat space . The solution is regular everywhere in the coordinate range $`\mathrm{}<`$ $`u`$ $`<\mathrm{},`$ (23) $`\mathrm{}<`$ $`v`$ $`<\mathrm{},`$ (24) $`0`$ $`\varphi `$ $`<2\pi ,`$ (25) $`\mathrm{}<`$ $`z`$ $`<\mathrm{}.`$ (26) There is no conical singularity on the axis of symmetry and the solution reduces to Minkowski spacetime at past and future infinity. However, the solution is conical at spatial infinity and singular at both past and future null infinity. More precisely, at future null infinity (cf Eq. (10) in ) $$e^{2\psi }1,e^{2\gamma }\frac{1+\alpha ^2\lambda _u^2}{1+\lambda _u^2},\omega \mathrm{},$$ (27) as $`v\mathrm{}`$, with analogous behaviour at past null infinity on interchanging $`u`$ and $`v`$. (Note that Eq. (10) in includes a typographical error and that $`u`$ and $`v`$ should be transposed). Since Piran et al only derived this family of solutions by the indirect method described above, we tried to confirm that the family is vacuum by a direct computation of the Ricci tensor using both of the computer algebra systems SHEEP and GRTENSOR . Unfortunately, neither system was able to complete the calculation because the metric involves nested radicals and algebra systems are notorious for the difficulties such quantities present. However, we were able to confirm the indirect derivation described above. ## 3 The standard field equations The standard treatment of the Jordan-Ehlers-Kompaneets cylindrically symmetric line element (13) in the Cauchy region $`t_0`$ $``$ $`tt_f,`$ (28) $`0`$ $``$ $`r1,`$ (29) is described in . In particular, it is shown that the independent set of dynamical equations for the variables $`\psi `$, $`\omega `$, $`L_z^\varphi `$ and $`\stackrel{~}{L}`$ are $`\psi _{,t}`$ $`=`$ $`{\displaystyle \frac{1}{r}}\stackrel{~}{L},`$ (30) $`\omega _{,t}`$ $`=`$ $`2e^{4\psi }L_z^\varphi ,`$ (31) $`L_{z,t}^\varphi `$ $`=`$ $`{\displaystyle \frac{1}{r}}e^{4\psi }({\displaystyle \frac{1}{2}}\omega _{,r}{\displaystyle \frac{1}{2}}r\omega _{,rr}2r\psi _{,r}\omega _{,r}),`$ (32) $`\stackrel{~}{L}_{,t}`$ $`=`$ $`{\displaystyle \frac{1}{r}}[r^2\psi _{,rr}+r\psi _{,r}{\displaystyle \frac{1}{2}}e^{4\psi }\omega _{,r}^2+2e^{4\psi }(L_z^\varphi )^2],`$ (33) where $`L_z^\varphi `$ and $`\stackrel{~}{L}`$ are defined in terms of the mixed components of the extrinsic curvature $`K_\nu ^\mu `$ by $`L_z^\varphi `$ $`=`$ $`r^2e^{\gamma \psi }K_z^\varphi ,`$ (34) $`\stackrel{~}{L}`$ $`=`$ $`r^2e^{\gamma \psi }(K_\varphi ^\varphi \omega K_z^\varphi ).`$ (35) These equations are augmented by the constraint equation $$\gamma _{,r}=\frac{1}{4r}e^{4\psi }\omega _{,r}^2r\psi _{,r}^2+\frac{1}{r}[\stackrel{~}{L}^2+e^{4\psi }(L_z^\varphi )^2],$$ (36) which serves to determine $`\gamma `$ once the main variables are known. In the rest of this section we review the treatment in the characteristic region since this will need modification to cope with the Piran et al family of solutions. We introduce the compactified coordinate $$y=\frac{1}{\sqrt{r}}=\frac{\sqrt{2}}{\sqrt{vu}},$$ (37) in which case the line element becomes $$ds^2=e^{2(\gamma \psi )}du^2+\frac{4}{y^3}e^{2(\gamma \psi )}dudy+e^{2\psi }(dz+\omega d\varphi )^2+\frac{e^{2\psi }}{y^4}d\varphi ^2,$$ (38) where the metric functions $`\psi `$, $`\omega `$ and $`\gamma `$ are now to be regarded as functions of $`u`$ and $`y`$. The characteristic region then consists of $`u_0`$ $``$ $`uu_f,`$ (39) $`0`$ $``$ $`y1,`$ (40) where the interface is taken to be at $`r=y=1`$ and future null infinity is given by $`y=0`$. Instead of working directly with the metric functions $`\psi `$ and $`\omega `$, we use the related quantities $`m`$ and $`w`$ which are defined by $`m`$ $`=`$ $`{\displaystyle \frac{e^{2\psi }1}{y}},`$ (41) $`w`$ $`=`$ $`{\displaystyle \frac{o}{y}}={\displaystyle _F^P}\lambda ^2(\frac{1}{2}y^4\omega _{,y}+y\omega _{,u})𝑑u+{\displaystyle \frac{1}{y}}{\displaystyle _F^P}\lambda ^2y^2\omega _{,y}𝑑y,`$ (42) where $$\lambda =e^{2\psi }=1+my,$$ (43) $`o`$ is the Geroch potential , and the integration is along any path connecting a fixed point $`F`$ on the interface (we choose $`(u,y)=(u_0,1)`$) to a general point $`P`$ in the characteristic region (see Fig. 1). With these definitions the vacuum equations can be written in the succinct form $`M`$ $`=`$ $`m_{,u}/\lambda ,`$ (44) $`W`$ $`=`$ $`w_{,u}/\lambda ,`$ (45) $`M_{,y}`$ $`=`$ $`f(y,m,m_{,y},m_{,yy},w,w_{,y},W),`$ (46) $`W_{,y}`$ $`=`$ $`g(y,m,m_{,y},w,w_{,y},w_{,yy},M),`$ (47) where $`f`$ and $`g`$ are explicit functions of their arguments . Eqs. (44) and (45) serve to define $`M`$ and $`W`$ and (46) and (47) provide coupled propagation equations for $`M`$ and $`W`$ along the null rays ruling the hypersurfaces $`u=\text{constant}`$. The initial data consists of specifying $`m`$ and $`w`$ on the initial hypersurface $`\{u=u_0,0y1\}`$ and the interface $`\{u_0uu_f,y=1\}`$, together with $`\gamma `$ at their intersection $`(u,y)=(u_0,1)`$. The essence of the iterative integration scheme is to use (46) and (47) to determine $`M`$ and $`W`$ on the interior null hypersurface $`u=u_i`$ and then (44) and (45) to determine $`m_{,u}`$ and $`w_{,u}`$ on $`u=u_i`$, which in turn determines $`m`$ and $`w`$ on the next neighbouring null slice. Finally $`\gamma `$ is determined from the constraint equation (Eq. (42) in ) which has the form $$\gamma _{,y}=h(y,m,m_{,y},w,w_{,y}),$$ (48) where $`h`$ is an explicit function of its arguments. The CCM method requires the exchange of values of the metric functions and their derivatives at the interface at each iteration, the details of which are given in . ## 4 The modified field equations The approach described in the last section breaks down for the Piran et al solution in the characteristic region because $`\omega \mathrm{}`$ as $`y0`$. Let us consider the behaviour of the various functions on the hypersurface $`u=u_i`$ in the asymptotic limit $`y0`$. We have, $`\lambda _u`$ $`=`$ $`(\sqrt{a^2+u_i^2}u_i)/a,`$ (49) $`\lambda _v`$ $`=`$ $`\overline{\lambda }_v/y^2,`$ (50) $`\mathrm{\Xi }`$ $`=`$ $`\overline{\mathrm{\Xi }}/y^2,`$ (51) where we set $`\overline{\lambda }_v`$ $``$ $`\left(2+u_iy^2+\sqrt{4+4u_iy^2+(a^2+u_i^2)y^4}\right)/a=4/a+O(y^2),`$ (52) $`\overline{\mathrm{\Xi }}`$ $``$ $`\lambda _u\overline{\lambda }_v+2y\sqrt{(1\alpha ^2)\lambda _u\overline{\lambda }_v}+y^2=4\lambda _u/a+O(y).`$ (53) Thus, asymptotically, $`e^{2\psi }`$ $`=`$ $`1+O(y),`$ (54) $`e^{2\gamma }`$ $`=`$ $`(1+\alpha ^2\lambda _u^2)/(1+\lambda _u^2)+O(y^2),`$ (55) $`\omega `$ $`=`$ $`\left({\displaystyle \frac{2\sqrt{(\alpha ^21)a\lambda _u}}{1+\alpha ^2\lambda _u^2}}\right){\displaystyle \frac{1}{y}}+{\displaystyle \frac{2a\sqrt{\alpha ^21}}{\alpha }}\left(\alpha +\sqrt{\alpha ^21}\right)+O(y),`$ (56) and $`m`$ $`=`$ $`{\displaystyle \frac{2\alpha \sqrt{(\alpha ^21)a\lambda _u^3}}{1+\alpha ^2\lambda _u^2}}+O(y),`$ (57) $`w`$ $`=`$ $`{\displaystyle \frac{b}{y}}+{\displaystyle \frac{2\sqrt{(\alpha ^21)a\lambda _u}}{1+\alpha ^2\lambda _u^2}}+O(y),`$ (58) $`o`$ $`=`$ $`b+{\displaystyle \frac{2\sqrt{(\alpha ^21)a\lambda _u}}{1+\alpha ^2\lambda _u^2}}y+O(y^2),`$ (59) where $$b=o(u_i,0).$$ (60) It is clear that $`\lambda _v`$, $`\mathrm{\Xi }`$, $`\omega `$ and $`w`$ are all divergent as $`y0`$. However, the ancillary quantities $`\overline{\lambda }_v`$ and $`\overline{\mathrm{\Xi }}`$ as well as $`m`$ and $`o`$, the Geroch potential, are all regular. This suggests rewriting the system (44)–(48) in terms of these last two variables. We find explicitly that $`M`$ $`=`$ $`{\displaystyle \frac{m_{,u}}{1+my}},`$ (61) $`O`$ $`=`$ $`{\displaystyle \frac{o_{,u}}{1+my}},`$ (62) $`M_{,y}`$ $`=`$ $`{\displaystyle \frac{o_{,y}}{y(1+my)}}O+{\displaystyle \frac{1}{4(1+my)}}[y(m+y^2m_{,yy}+3ym_{,y})`$ $`+{\displaystyle \frac{y^2}{1+my}}(m^2+2ymm_{,y}+y^2m_{,y}^2o_{,y}^2)],`$ (63) $`O_{,y}`$ $`=`$ $`{\displaystyle \frac{1}{y}}O+{\displaystyle \frac{yo_{,y}}{(1+my)}}M{\displaystyle \frac{y^2}{4(1+my)}}(yo_{,yy}+o_{,y})`$ $`+{\displaystyle \frac{y^3}{2(1+my)^2}}(mo_{,y}+ym_{,y}o_{,y}),`$ (64) $`\gamma _{,y}`$ $`=`$ $`{\displaystyle \frac{y}{8(1+my)^2}}\left((m+ym_{,y})^2+o_{,y}^2\right).`$ (65) Although the equations (63) and (64) are coupled singular equations for $`M`$ and $`O`$, they are regular singular equations and the solutions remain regular as $`y0`$ (see the Appendix). In fact, it is clear from the defining equations (61) and (62), together with (57) and (59), that both $`M`$ and $`O`$ are of order unity in the limit $`y0`$. These are the modified equations on which the new code is based for investigating the Piran et al solutions. We need to augment the equations with the interface equations which relate the two sets of variables in the two coordinate systems on the interface. The relationships required for injection, i.e. going from the characteristic to the Cauchy region are, $`\psi `$ $`=`$ $`{\displaystyle \frac{1}{2}}\mathrm{ln}(1+my),`$ (66) $`\psi _{,r}`$ $`=`$ $`{\displaystyle \frac{1}{2}}yM{\displaystyle \frac{1}{4}}{\displaystyle \frac{y^3}{\lambda }}(ym)_{,y},`$ (67) $`\psi _{,rr}`$ $`=`$ $`{\displaystyle \frac{1}{2}}yM_{,u}+{\displaystyle \frac{1}{2}}y^3(yM)_{,y}{\displaystyle \frac{1}{8}}{\displaystyle \frac{y^6}{\lambda ^2}}[(ym)_{,y}]^2`$ $`+{\displaystyle \frac{1}{8}}{\displaystyle \frac{y^5}{\lambda }}\left\{2(ym)_{,y}+[y(ym)_{,y}]_{,y}\right\},`$ (68) $`\stackrel{~}{L}`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \frac{M}{y}},`$ (69) $`\omega _{,r}`$ $`=`$ $`{\displaystyle \frac{O}{y^2\lambda }},`$ (70) $`\omega _{,rr}`$ $`=`$ $`{\displaystyle \frac{O_{,u}}{y^2\lambda }}+{\displaystyle \frac{MO}{y\lambda }}+{\displaystyle \frac{O}{\lambda }}+{\displaystyle \frac{1}{2}}{\displaystyle \frac{yO(ym)_{,y}}{\lambda ^2}}{\displaystyle \frac{1}{2}}{\displaystyle \frac{yO_{,y}}{\lambda }},`$ (71) $`L_z^\varphi `$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \frac{\lambda O}{y^2}}+{\displaystyle \frac{1}{4}}yo_{,y}.`$ (72) Similarly, the relationships required for extraction, i.e. going from the Cauchy to the characteristic region are, $`\lambda `$ $`=`$ $`e^{2\psi },`$ (73) $`m`$ $`=`$ $`r^{1/2}(e^{2\psi }1),`$ (74) $`m_{,y}`$ $`=`$ $`r^2e^{2\psi }(4\psi _{,t}+4\psi _{,r}+r^1)+r,`$ (75) $`m_{,yy}`$ $`=`$ $`2r^{7/2}e^{2\psi }(4\psi _{,tt}+8\psi _{,tr}+8\psi _{,t}^2+10r^1\psi _{,t}+16\psi _{,t}\psi _{,r}+4\psi _{,rr}`$ $`+8\psi _{,r}^2+10r^1\psi _{,r}+r^2)2r^{3/2},`$ (76) $`M`$ $`=`$ $`2r^{1/2}\stackrel{~}{L},`$ (77) $`o`$ $`=`$ $`{\displaystyle _F^I}r^1e^{4\psi }\omega _{,r}𝑑t,`$ (78) $`o_{,y}`$ $`=`$ $`4r^{1/2}L_z^\varphi 2r^{1/2}e^{4\psi }\omega _{,r},`$ (79) $`o_{,yy}`$ $`=`$ $`2r^2e^{4\psi }(8\psi _{,t}\omega _{,r}+2\omega _{,tr}+r^1\omega _{,r}+8\psi _{,r}\omega _{,r}+2\omega _{,rr}+r^1\omega _{,t})`$ $`8r^2(L_{z,t}^\varphi +L_{z,r}^\varphi ),`$ (80) $`O`$ $`=`$ $`r^1e^{2\psi }\omega _{,r}.`$ (81) where the integration in (78) is along the interface from the point $`F:(t,r)=(t_0,1)`$ to the point $`I:(t,r)=(t_i,1)`$ as shown in Fig. 1. ## 5 Results In order to test the CCM code against the Piran et al solution we will first compare the metric quantity $`\psi `$ in both the Cauchy and characteristic regions. Eq. (18) provides the exact solution in both cases when $`\lambda _u`$ and $`\lambda _v`$, Eqs. (21) and (22), are written in terms of $`(t,r)`$ coordinates and $`(u,y)`$ coordinates for the Cauchy and characteristic regions respectively. It is not possible to use $`\omega `$ for a similar comparison since from (56) we have seen that it diverges as $`y0`$. We use, instead, the Geroch potential, $`o`$, which is computed directly by the code in the characteristic region, but must be constructed in the Cauchy region using its definition (cf Eq. (24) of ), $$o=r^1e^{4\psi }\omega _{,r}𝑑t+r^1e^{4\psi }\omega _{,t}𝑑r.$$ (82) It is straightforward to evaluate the integrals in the Cauchy region using finite-difference representations of $`\omega _{,r}`$ and $`\omega _{,t}`$. For the exact solution $`\omega _{,r}`$ and $`\omega _{,t}`$ can be evaluated by differentiation, however the integral must be evaluated numerically to find a semi-analytic $`o`$. A similar approach is required for the ‘exact’ value of $`o`$ in the characteristic region using Eq. (42). As mentioned previously, the two parameters $`a`$ and $`\alpha `$ of the Piran et al solution represent the length scale and strength of the gravitational wave respectively. For $`t<0`$ the wave moves in from infinity, reaches its highest concentration at $`t=0`$, where it rebounds from the $`r=0`$ axis, and is outgoing for $`t>0`$. CCM allows us to evolve initial data containing an ingoing gravitational wave, something which is difficult to do for a Cauchy only evolution. Since we have set the interface at $`r=1`$ the value of $`a`$ can be chosen such that a substantial fraction of the the wave will move onto the Cauchy region and will be well resolved with the numbers of grid points we choose. Changing the interface position while keeping the ‘interface distance to $`a`$’ ratio the same results in an identical evolution on the Cauchy portion of the grid. The primary variable we used for comparison are $`\psi `$ and $`o`$ (since $`\gamma `$ is a derived quantity). We considered grid resolution numbers of $`N=301,601`$ and $`1201`$, where the error is defined as $$ϵ(\psi )=\psi _E\psi _C_2/\psi _E_2,$$ (83) where $`\psi _E`$ is the exact value, $`\psi _C`$ is the code computed value and $`||||_2`$ denotes the $`L_2`$ norm. We ran the modified code for a whole range of values in the $`a,\alpha `$ parameter space and found an error in $`\psi `$ of no more than 0.01% and an error in $`o`$ of less than 0.2%. However, in this version of the code, although the convergence rate of the solution starts as second-order it deteriorates to first order after long time evolution. In some recent work, colleagues in the Southampton Numerical Relativity group have modified the code so that it is now fully second order and very long time stable . The improvement was achieved by dispensing with variables which are related by exponential or logarithmic functions at the interface, using the Geroch potential both in the Cauchy and characteristic regimes and also using an implicit method at the interface. This version of the code is currently being applied successfully to modelling dynamic cosmic strings. Surface plots are shown for evolutions of the metric quantity $`e^{2\psi }`$ in Fig. 2, the Geroch potential $`o`$ in Fig. 3, and the radial derivative $`\gamma _{,r}`$ (which indicates the distribution of energy within the wave)in Fig. 4, where the parameters are $`a=0.5`$ and $`\alpha =10`$. The surface plots use a radial coordinate $`z`$ defined by, $$z=\{\begin{array}{cc}r\hfill & \text{for }0r1\hfill \\ 2y\hfill & \text{for }0y1\hfill \end{array}$$ (84) so that there is a change in the coordinate system at $`r=y=z=1`$. Note that the value of $`e^{2\psi }`$ remains equal to 1 at null infinity, but is very close to zero at the axis $`r=0`$. The incoming wave hits the axis at $`t=0`$ and rebounds. This can be seen very clearly in the surface plot of $`\gamma _{,r}`$. The large peak of Fig. 2c grows without bound as $`\alpha `$ is increased. For comparison purposes figures 5, 6 and 7 shows the same quantities, but for parameter values $`a=0.5`$ and $`\alpha =1.01`$, which represents an almost flat spacetime. Again the wave reaches its maximum concentration on the $`r=0`$ axis at $`t=0`$, however the peak is much less dominant in this case. We would like to thank James Vickers for helpful discussions. This work has been supported by PPARC grant number GR/K44510. ## Appendix A Regular-singular behaviour of modified propagation equations Let $$F=\frac{o_{,y}}{(1+my)},$$ (85) which by (57) and (59) is regular as $`y0`$. The homogeneous part of the modified propagation equations (63) and (64) is $`M_{,y}+{\displaystyle \frac{F}{y}}O`$ $`=`$ $`0,`$ (86) $`O_{,y}yFM{\displaystyle \frac{1}{y}}O`$ $`=`$ $`0.`$ (87) Differentiating (87) with respect to $`y`$ and using (87) to eliminate $`M`$ and (86) to eliminate $`M_{,y}`$ we get $$O_{,yy}\left(\frac{2}{y}+\frac{F_{,y}}{F}\right)O_{,y}+\left(\frac{2}{y^2}+\frac{F_{,y}}{yF}+F^2\right)O=0,$$ (88) which has dominant singular part $$O_{,yy}\frac{2}{y}O_{,y}+\frac{2}{y^2}O=0.$$ (89) Substituting in the trial solution $`y^k`$ we obtain the auxiliary equation $$k^23k+2=0$$ (90) which has roots $`k=1,2`$ and so gives rise to the regular independent solutions $`y`$ and $`y^2`$. ## References
warning/0002/math0002080.html
ar5iv
text
# On the K-property of quantized Arnold cat maps ## I Introduction The concept of K-system is very important in ergodic theory. Narnhofer and Thirring introduced a non-commutative analogue of this notion. In V.Ya. Golodets and the author proved the following sufficient condition for the K-property: a W-system $`(M,\varphi ,\alpha )`$ is an entropic K-system if there exists a W-subalgebra $`M_0`$ of $`M`$ such that $`M_0\alpha (M_0)`$, $`_n\alpha ^n(M_0)=1`$, $`_n(\alpha ^n(M_0)^{}\alpha ^n(M_0))`$ is weakly dense in $`M`$. This condition and the observation that a subsystem of a K-system invariant under the modular group is a K-system too allow to construct a large class of quantum K-systems (see, in particular, ). We know only one class of quantum systems for which the K-property is obtained by different arguments. This is quantized Arnold cat maps. This result was formulated in Narnhofer’s paper . The decompositions contructed in the course of the proof there are not strictly local, that leads to a factor that again could only be controlled by using asymptotic abelian arguments. So the essential interest lies in the construction of a completely positive map that is strictly local and can be well controlled and generalized in a larger context. ## II The K-property of quantized cat maps Let $`G`$ be a discrete abelian group, $`\omega :G\times G𝕋`$ a bicharacter. Consider the twisted group C-algebra $`C^{}(G,\omega )`$ generated by unitaries $`u_g`$, $`gG`$, such that $$u_gu_h=\omega (g,h)u_{g+h}.$$ The canonical trace $`\tau `$ on $`C^{}(G,\omega )`$ is given by $`\tau (u_g)=0`$ for $`g0`$. It is known that the uniqueness of the trace is equivalent to the simplicity of $`C^{}(G,\omega )`$, and is also equivalent to the non-degeneracy of the pairing $`(g,h)\omega (g,h)\overline{\omega }(h,g)`$. In particular, if $`G`$ is countable and the pairing is non-degenerate, then $`\pi _\tau (C^{}(G,\omega ))^{\prime \prime }`$ is the hyperfinite II<sub>1</sub>-factor. Each $`\omega `$-preserving automorphism $`T`$ of $`G`$ defines an automorphism $`\alpha _T`$ of $`C^{}(G,\omega )`$, $`\alpha _T(u_g)=u_{Tg}`$. The non-commutative torus $`A_\theta `$ ($`\theta [0,1)`$) is the algebra $`C^{}(^2,\omega _\theta )`$, where $$\omega _\theta (g,h)=e^{i\pi \theta \sigma (g,h)},\sigma (g,h)=g_1h_2g_2h_1.$$ The following theorem was formulated in . Theorem 1. Let $`T\mathrm{SL}_2()`$, $`\mathrm{Spec}T=\{\lambda ,\lambda ^1\}`$. Suppose $`|\lambda |>1`$ (so that $`\lambda `$ is real) and $`\theta [0,1)(2\lambda ^2+2)`$. Then $`(\pi _\tau (A_\theta )^{\prime \prime },\tau ,\alpha _T)`$ is an entropic K-system. We will prove the following more general result. Theorem 2. Let $`T`$ be an aperiodic $`\omega `$-preserving automorphism of $`G`$. Suppose that $$\underset{n}{}|1\omega (g,T^nh)|<\mathrm{}g,hG.$$ Then $`(\pi _\tau (C^{}(G,\omega ))^{\prime \prime },\tau ,\alpha _T)`$ is an entropic K-system. It was proved in \[1, Theorem 3.8\] that under the assumptions of Theorem 1, for any $`g,h^2`$, we have $$|1\omega (g,T^nh)|C|\lambda |^{|n|},$$ so Theorem 1 is really follows from Theorem 2. The key observation for that estimate was the equality $$T^nh=\frac{1}{\lambda ^21}\underset{i=0}{\overset{2}{}}(\lambda ^{n+i}+\lambda ^{ni})\overline{h}_i+\lambda ^n\overline{h}(n),$$ where $`\overline{h}_i^2`$ and $`\overline{h}^2`$ depend only on $`h`$ and $`T`$, which is obtained by computations in a basis diagonalizing $`T`$. Since $`\sigma (g,\overline{h}_i)`$, $`\sigma (g,h)`$, $`\lambda ^{n+i}+\lambda ^{ni}=\text{Tr}T^{n+i}`$ are all integers and $`\theta 2s(\lambda ^21)\text{mod}\mathrm{\hspace{0.17em}2}`$ for some $`s`$, we have $$\theta \sigma (g,T^nh)\lambda ^n2s(\lambda ^21)\sigma (g,\overline{h})\text{mod}\mathrm{\hspace{0.17em}2},$$ whence $`|1\omega (g,T^nh)||\lambda |^n2\pi |s|(\lambda ^21)|\sigma (g,\overline{h})|`$. Starting the proof of Theorem 2, consider a unital completely positive mapping $`\gamma :A\pi _\tau (C^{}(G,\omega )^{\prime \prime }`$ of a finite-dimensional C-algebra $`A`$. By definition , we have to prove that $$\underset{n\mathrm{}}{lim}\underset{k\mathrm{}}{lim}\frac{1}{k}H_\tau (\gamma ,\alpha _T^n\gamma ,\mathrm{},\alpha _T^{n(k1)}\gamma )=H_\tau (\gamma ).$$ For a finite set $`X`$, we denote by $`\mathrm{Mat}(X)`$ the C-algebra of linear operators on $`l^2(X)`$. Let $`\{e_{xy}\}_{x,yX}`$ be the canonical system of matrix units in $`\mathrm{Mat}(X)`$. For $`XG`$, we define a unital completely positive mapping $`i_X:\mathrm{Mat}(X)C^{}(G,\omega )`$ by $$i_X(e_{xy})=\frac{1}{|X|}u_xu_y^{}=\frac{\overline{\omega }(xy,y)}{|X|}u_{xy}.$$ As follows from (see Lemmas 5.1 and 6.1 there), there exist a net $`\{X_i\}_i`$ of finite subsets in $`G`$ and, for each $`i`$, a unital completely positive mapping $`j_{X_i}:C^{}(G,\omega )\mathrm{Mat}(X_i)`$ such that $`(i_{X_i}j_{X_i})(a)a\underset{i}{}0`$ $`aC^{}(G,\omega )`$. ¿From this we may conclude that any partition of unit in $`\pi _\tau (C^{}(G,\omega ))^{\prime \prime }`$ can be approximated in strong operator topology by a partition of the form $`\{i_X(a_k)\}_k`$, where $`\{a_k\}_k`$ is a partition of unit in $`\mathrm{Mat}(X)`$. Hence, for any $`\epsilon >0`$, there exist a finite subset $`XG`$ and a finite partition of unit $`1=_{iI}a_i`$ in $`\mathrm{Mat}(X)`$ such that, for $`b_i=i_X(a_i)`$, we have $$H_\tau (\gamma )<\epsilon +\underset{i}{}\eta \tau (b_i)+\underset{i}{}S(\tau (\gamma ()),\tau (\gamma ()b_i)),$$ where $`\eta x=x\mathrm{log}x`$. Set $`X_{nk}=_{l=1}^kT^{n(l1)}(X)`$. The following lemma was proved in for $`G=^2`$. Lemma. Let $`G`$ be a discrete abelian group, $`T`$ an aperiodic endomorphism of $`G`$, $`\mathrm{Ker}T=0`$, $`Y`$ a finite subset of $`G`$, $`0Y`$. Then there exists $`n_0`$ such that if $$\underset{l=1}{\overset{k}{}}T^{n(l1)}y_l=0$$ (1) for some $`y_1,\mathrm{},y_kY`$, $`nn_0`$, $`k`$, then $`y_1=\mathrm{}=y_k=0`$. Proof. First consider the case where $`G`$ is finitely generated. Then the periodic part of $`G`$ is finite. Since $`T`$ acts on it aperiodically, it is trivial, so $`G^n`$ for some $`n`$. Then $`T`$ is defined by a non-degenerate matrix with integral entries, which we denote by the same letter $`T`$. It is known that the aperiodicity is equivalent to $`𝕋\mathrm{Spec}T=\mathrm{}`$. Let $`\mathrm{Spec}T=\{\lambda _1,\mathrm{},\lambda _m\}`$, $`V_i^n`$ be the root space corresponding to $`\lambda _i`$, and $`P_i`$ the projection onto $`V_i`$ along $`_{ji}V_j`$. Then (1) is equivalent to the system of equalities $$\underset{l=1}{\overset{k}{}}T^{n(l1)}P_iy_l=0,$$ (2) $`i=1,\mathrm{},m`$. Fix $`i`$. Suppose, for definiteness, that $`|\lambda _i|<1`$, and choose $`\delta `$, $`0<\delta <1|\lambda _i|`$. Since $`T|_{V_i}`$ is a sum of Jordan cells, there exists a constant $`C`$ such that $$||T^n|_{V_i}||C(|\lambda _i|+\delta )^nn.$$ There exists also a constant $`M>0`$ such that, for $`yY`$, we have either $`P_iy=0`$ or $`M^1P_iyM`$. Finally, choose $`n_i`$ such that $$\underset{n=n_i}{\overset{\mathrm{}}{}}MC(|\lambda _i|+\delta )^n<M^1.$$ Then if the equality (2) holds with $`nn_i`$, then $`P_iy_1=0`$. Since $`\mathrm{Ker}T=0`$, we can rewrite (2) as $`_{l=1}^{k1}T^{n(l1)}P_iy_{l+1}=0`$. Thus we sequentially obtain $`P_iy_1=\mathrm{}=P_iy_k=0`$. So we may take $`n_0=\mathrm{max}_in_i`$. We prove the general case by induction on $`|Y|`$ using the same method as in to reduce the proof to the case considered above. Let $`H_0`$ be the group generated by $`Y,TY,T^2Y,\mathrm{}`$. Set $`H_n=T^nH_0`$, $`H_{\mathrm{}}=_nH_n`$, $`Y^{}=YH_{\mathrm{}}`$. Suppose $`Y^{}Y`$. There exists $`n_1`$ such that $`Y^{}=YH_{n_1}`$. If the equality (1) holds with $`nn_1`$, then $`y_1H_{n_1}Y=Y^{}H_{\mathrm{}}`$. Then $`_{l=2}^kT^{n(l1)}y_lH_{\mathrm{}}`$. Since $`\mathrm{Ker}T=0`$ and $`TH_{\mathrm{}}=H_{\mathrm{}}`$, we conclude that $`_{l=1}^{k1}T^{n(l1)}y_{l+1}H_{\mathrm{}}`$. Thus we sequentially obtain that $`y_1,\mathrm{},y_kY^{}`$. Since $`|Y^{}|<|Y|`$, we may apply the inductive assumption. If $`Y^{}=Y`$, then $`YH_1`$, hence there exists $`n`$ such that if $`\overline{H}`$ is the group generated by $`Y,TY,\mathrm{},T^nY`$, then $`YT\overline{H}`$. Then $`\overline{H}`$ is a finitely generated group, $`T^1`$ an aperiodic endomorphism of $`\overline{H}`$. For this case Lemma is already proved. Applying Lemma to the set $`Y=XX`$ we see that the mapping $$X^kX_{nk},(x_1,\mathrm{},x_k)\underset{l=1}{\overset{k}{}}T^{n(l1)}x_l,$$ is a bijection for all $`k`$ and for all $`n`$ sufficiently large. This bijection induces an isomorphism of $`\mathrm{Mat}(X^k)`$ onto $`\mathrm{Mat}(X_{nk})`$. Composing it with $`i_{X_{nk}}:\mathrm{Mat}(X_{nk})C^{}(G,\omega )`$ and identifying $`\mathrm{Mat}(X^k)`$ with $`\mathrm{Mat}(X)^k`$ we obtain a unital completely positive mapping $$\sigma _{nk}:\mathrm{Mat}(X)^kC^{}(G,\omega ).$$ Set $`b(n,k)_{i_1\mathrm{}i_k}=\sigma _{nk}(a_{i_1}\mathrm{}a_{i_k})`$. By definition , we obtain $`{\displaystyle \frac{1}{k}}H_\tau (\gamma ,\alpha _T^n\gamma ,\mathrm{},\alpha _T^{n(k1)}\gamma )`$ $$\frac{1}{k}\underset{i_1,\mathrm{},i_k}{}\eta \tau (b(n,k)_{i_1\mathrm{}i_k})+\frac{1}{k}\underset{l=1}{\overset{k}{}}\underset{i_l}{}S(\tau \left(\gamma ()\right),\tau \left(\gamma ()\alpha _T^{n(l1)}(b(n,k)_{i_l}^{(l)})\right)),$$ where $`b(n,k)_{i_l}^{(l)}={\displaystyle \underset{i_1,\mathrm{},\widehat{i}_l,\mathrm{},i_k}{}}b(n,k)_{i_1\mathrm{}i_k}`$. If we denote by $`\tau _Y`$ the unique tracial state on $`\mathrm{Mat}(Y)`$, then $`\tau _Y=\tau i_Y`$, so that $`\tau \sigma _{nk}=\tau _X^k`$, whence $$\tau (b(n,k)_{i_1\mathrm{}i_k})=\underset{l=1}{\overset{k}{}}\tau _X(a_{i_l})=\underset{l=1}{\overset{k}{}}\tau (b_{i_l}).$$ So the first term in the inequality above is equal to $`_i\eta \tau (b_i)`$, and in order to prove Theorem it remains to show that $$\alpha _T^{n(l1)}(b(n,k)_{i_l}^{(l)})b_{i_l}\underset{n\mathrm{}}{}0$$ uniformly on $`k,l`$ ($`lk`$) and $`i_lI`$. Let $`\theta _l`$ be the embedding of $`\mathrm{Mat}(X)`$ into $`\mathrm{Mat}(X)^k`$ defined by $$\theta _l(a)=\underset{l1}{\underset{}{1\mathrm{}1}}a\underset{kl}{\underset{}{1\mathrm{}1}}.$$ Then $`b(n,k)_{i_l}^{(l)}=(\sigma _{nk}\theta _l)(a_{i_l})`$. Thus we just have to estimate $$\alpha _T^{n(l1)}\sigma _{nk}\theta _li_X.$$ Using the facts that $`\omega `$ is bilinear and $`T`$-invariant we obtain $`\sigma _{nk}(e_{x_1x_1}\mathrm{}e_{x_{l1}x_{l1}}e_{xy}e_{x_{l+1}x_{l+1}}\mathrm{}e_{x_kx_k})=`$ $`=`$ $`{\displaystyle \frac{1}{|X|^k}}\overline{\omega }(T^{n(l1)}(xy),T^{n(l1)}y+{\displaystyle \underset{i=1,il}{\overset{k}{}}}T^{n(i1)}x_i)u_{T^{n(l1)}(xy)}`$ $`=`$ $`\left({\displaystyle \underset{i=1,il}{\overset{k}{}}}{\displaystyle \frac{\overline{\omega }(xy,T^{n(il)}x_i)}{|X|}}\right){\displaystyle \frac{\overline{\omega }(xy,y)}{|X|}}u_{T^{n(l1)}(xy)},`$ so that $`(\alpha _T^{n(l1)}\sigma _{nk}\theta _li_X)(e_{xy})=`$ $`=`$ $`{\displaystyle \frac{1}{|X|}}\left|{\displaystyle \underset{x_1,\mathrm{},\widehat{x}_l,\mathrm{},x_k}{}}\left({\displaystyle \underset{i=1,il}{\overset{k}{}}}{\displaystyle \frac{\overline{\omega }(xy,T^{n(il)}x_i)}{|X|}}\right)1\right|`$ $`=`$ $`{\displaystyle \frac{1}{|X|}}\left|{\displaystyle \underset{i=1,il}{\overset{k}{}}}\left({\displaystyle \frac{1}{|X|}}{\displaystyle \underset{zX}{}}\overline{\omega }(xy,T^{n(il)}z)\right)1\right|.`$ We must show that the latter expression tends to zero as $`n\mathrm{}`$ uniformly on $`k,l`$ ($`lk`$). This follows from $$\underset{n}{}\left|1\frac{1}{|X|}\underset{zX}{}\omega (xy,T^nz)\right|<\mathrm{}.$$ So the proof of Theorem 2 is complete. ## III Classical case If $`\omega 1`$, then $`C^{}(G,\omega )=C(\widehat{G})`$, the algebra of continuous functions on the dual group $`\widehat{G}`$. It is known that an automorphism $`T`$ of $`G`$ is aperiodic iff the dual automorphism of $`\widehat{G}`$ is ergodic. Thus we obtain a classical Rohlin’s result stating that ergodic automorphisms of compact abelian groups have completely positive entropy. Note that in this case we have $$b(n,k)_{i_1\mathrm{}i_k}=b_{i_1}\alpha _T^n(b_{i_2})\mathrm{}\alpha _T^{n(k1)}(b_{i_k}),$$ so what is really necessary for the proof is Lemma above and the possibility of approximating in mean measurable partitions of unit by partitions consisting from trigonometric polynomials, which can be proved by elementary methods without appealing to Voiculescu’s completely positive mappings. Institute for Low Temperature Physics & Engineering Lenin Ave 47 Kharkov 310164, Ukraine neshveyev@ilt.kharkov.ua
warning/0002/hep-ph0002205.html
ar5iv
text
# 1 Introduction ## 1 Introduction The Standard Model (SM) effective potential is unstable beyond a scale, $`\mathrm{\Lambda }_{SM}`$, which solely depends upon the value of the (yet undiscovered) Higgs mass $`M_H`$ <sup>1</sup><sup>1</sup>1The original studies considered $`\mathrm{\Lambda }_{SM}`$ as a function of $`M_H`$ and $`M_t`$ (the top-quark mass) . In this paper we will fix $`M_t`$ to its experimental mean value and disregard the effect due to the experimental error $`\mathrm{\Delta }M_t`$.. This instability is a drawback of the Standard Model, which is unable to describe physics at scales beyond $`\mathrm{\Lambda }_{SM}`$ and, then, requires the presence of new physics to stabilize the SM vacuum. For large values of the Higgs mass the scale $`\mathrm{\Lambda }_{SM}`$ is larger than the Planck scale and thus has no impact on the physics at present and future accelerators. However, for the case of a light Higgs (i.e. $`M_H\stackrel{<}{_{}}140`$ GeV) $`\mathrm{\Lambda }_{SM}`$ can be closer to values that could directly, or indirectly, be detected at future accelerators, thus having an impact on the physics at the corresponding scales. In particular, and to guarantee the absolute stability of the electroweak vacuum, new physics must be introduced. The direct, or indirect, detection of the new physics depends on the mass, $`M`$, of the new involved fields, as deduced from the stability requirement. This is the issue that will be considered in this paper. Since the presence of extra bosonic (fermionic) degrees of freedom tend to stabilize (destabilize) the SM Higgs potential , it is clear that the simplest and most economic SM extension that could circumvent the instability problem is the addition of just scalar fields coupled to the SM Higgs field . This will be the model considered in the present paper. A similar analysis of this kind of model was performed by Hung and Sher in Ref. , where the effective potential was considered in the one-loop approximation. However improving the effective potential by the renormalization group equations (RGE) yields very important corrections and must be taken into account in any realistic calculation. Obviously, other SM extensions, such as the MSSM, can be much more realistic, but, for the previous reasons, the simple model under consideration basically represents the most efficent way to cure the SM instability, thus giving reliable upper bounds on the scale of new physics, i.e. on the mass of the extra particles. Improving the effective potential in a SM extension generally implies considering a multi-scale problem. In our case, we have to consider two different mass scales: the SM scale, say $`\mu _t`$, that can be identified with the top-quark mass <sup>2</sup><sup>2</sup>2For the sake of simplicity we will consider $`\mu _t`$ as the only SM scale, and will not make a distinction with the other (nearby) SM scales, as the masses of the Higgs and the gauge bosons., and the scale of new physics, say $`\mu _\mathrm{\Phi }`$, that should be identified with the mass of the new scalar fields ($`\mathrm{\Phi }`$). A general formalism to evaluate the effective potential in this kind of multi-scale scenario was presented by the authors in Ref. , where the general procedure for decoupling fields was incorporated. We will use this formalism throughout the paper. The outline of the paper is as follows: In section 2 we will briefly summarize our general proposal of multi-scale effective potential. In section 3 we will apply the ideas and results of section 2 to the model we are considering to remove the SM instability. In section 4 we will present our numerical results and in section 5 our conclusions. ## 2 Multi-scale effective theory We will consider the effective potential of a theory with $`N`$ different scales in a mass-independent renormalization scheme, as e.g. the $`\overline{\mathrm{MS}}`$ scheme, where the decoupling is not automatically incorporated in the theory. The usual procedure is to decouple every field at a given scale, $`\mu _i`$, that is normally associated to its mass, by means of e.g. a step function. However, the scale invariance of the complete effective potential indicates that the results should not depend on the details of the chosen decoupling scale $`\mu _i`$, thus implying independence of the (complete) effective action with respect to the scale $`\mu _i`$ (similar to the usual scale invariance). This will give rise to a set of RGE with respect to the scales $`\mu _i`$, as well as with respect to $`\mu `$, which is the ordinary $`\overline{\mathrm{MS}}`$ renormalization scale. On top of that we will use a simple step function for decoupling <sup>3</sup><sup>3</sup>3In addition, at a given threshold the corresponding symmetry of the system may change, in which case the corresponding matching conditions have to be taken into account at the corresponding thresholds. A typical example is the threshold of supersymmetric particles in the MSSM. Below it, the theory is non-supersymmetric and beyond it supersymmetric. This kind of complications will not appear, however, in the model at consideration in the present paper., although it could be easily smoothed. These ideas were presented in Ref. and similar ones can be found in Refs. and . In this decoupling approach, when computing the one-loop $`\beta `$-functions corresponding to $`\lambda _a`$ \[where $`\lambda _a`$ denotes all dimensionless and dimensional couplings of the theory, including the bosonic and fermionic fields\], the contribution from decoupled fields should not be counted. This translates into the decomposition: $$\mu \frac{d\lambda _a}{d\mu }_\mu \beta _{\lambda _a}=\underset{i=1}{\overset{N}{}}{}_{\mu _i}{}^{}\stackrel{~}{\beta }_{\lambda _a}^{}\theta (\mu \mu _i)$$ (2.1) which provides the definition of the factors $`{}_{\mu _i}{}^{}\stackrel{~}{\beta }_{\lambda _a}^{}`$. Invariance of the complete effective potential with respect to the $`\overline{\mathrm{MS}}`$ scale $`\mu `$ simply reads $$\mu \frac{dV_{\mathrm{eff}}}{d\mu }=0$$ (2.2) On the other hand, the one-loop effective potential can be written as: $$V_{\mathrm{eff}}=V_0+V_1$$ (2.3) where $`V_0`$ is the tree-level potential, and $$V_1=\frac{1}{64\pi ^2}\underset{i=1}{\overset{N}{}}()^{2s_i}M_i^4\left[\mathrm{log}\frac{M_i^2}{\mu _i^2}+\theta (\mu \mu _i)\mathrm{log}\frac{\mu _i^2}{\mu ^2}C_i\right]$$ (2.4) where $`s_i`$ is the spin of the $`ith`$ field, $`M_i`$ are the (tree-level) mass eigenvalues and $`C_i`$ depends on the renormalization scheme. In the $`\overline{\mathrm{MS}}`$-scheme it is equal to 3/2 (5/6) for scalar bosons and fermions (gauge bosons). Notice that for $`\mu >\mu _i`$ (for all $`i`$) the potential (2.4) coincides with the usual $`\overline{\mathrm{MS}}`$ effective potential when there are no decoupled particles. For $`\mu <\mu _i`$ the term $`M_i^4\mathrm{log}(M_i^2/\mu _i^2)`$ can be taken (at the one-loop level) as frozen at the scale $`\mu =\mu _i`$, so it does not run with respect to $`\mu `$. On the other hand, the invariance of the effective potential with respect to $`\mu _i`$, $$\mu _i\frac{dV_{\mathrm{eff}}}{d\mu _i}=0$$ (2.5) leads to the running of the parameters with respect to the scale $`\mu _i`$: $$\mu _i\frac{d\lambda _a}{d\mu _i}_{\mu _i}\beta _{\lambda _a}=_{\mu _i}\stackrel{~}{\beta }_{\lambda _a}\theta (\mu _i\mu )$$ (2.6) From Eqs. (2.2) and (2.6) it follows that $$_\mu \beta _{\lambda _a}+\underset{i=1}{\overset{N}{}}{}_{\mu _i}{}^{}\beta _{\lambda _a}^{}=\underset{i=1}{\overset{N}{}}{}_{\mu _i}{}^{}\stackrel{~}{\beta }_{\lambda _a}^{}=\beta _{\lambda _a}^{\overline{\mathrm{MS}}}$$ (2.7) where $`\beta _{\lambda _a}^{\overline{\mathrm{MS}}}`$ is the usual (complete) $`\beta `$-function in the $`\overline{\mathrm{MS}}`$-scheme. The invariance of the effective potential with respect to $`\mu `$ and $`\mu _i`$ allows to choose any values for these scales. These can be constant, as it is usually done , or field-dependent <sup>4</sup><sup>4</sup>4This is similar to the ordinary $`\overline{\mathrm{MS}}`$-scheme where the scale $`\mu `$ can be fixed to a field dependent value. In the case of the SM this value is usually $`M_t`$, in order to improve the validity of perturbation theory.. A choice that is particularly convenient to greatly improve the validity of perturbation theory is $`\mu _iM_i`$ and $`\mu \stackrel{<}{_{}}\mathrm{min}\{\mu _i\}`$ . This gets rid of all the dangerous logarithms in Eq. (2.4). Notice here that since $`M_i`$ are in general functions of the fields, so the preferred value of the $`\mu _i`$ scales are. In addition, the evaluation of the effective potential requires to evaluate all the $`\lambda _a`$ parameters at the same values of $`\mu _i`$ (note that $`\lambda _a`$ run with the $`\mu _i`$ scales). This implies a knowledge of the $`{}_{\mu _i}{}^{}\stackrel{~}{\beta }_{\lambda _a}^{}`$ functions defined in Eq. (2.1). It is interesting to note that the previous decoupling approach can be obtained starting with the bare lagrangian written in an appropriate renormalization scheme. The latter is a generalization of the so-called multi-scale renormalization scheme proposed in Refs. . Given a set of bosonic $`\varphi _i`$ and fermionic $`\psi _j`$ fields, the bare lagrangian, in terms of renormalized fields $`(\varphi _i,\psi _j)`$ and renormalized couplings and masses $`\lambda _b`$, can be written as: $$_{\mathrm{Bare}}=\underset{i}{}_{\mathrm{kin},\varphi _i}+\underset{j}{}_{\mathrm{kin},\psi _j}+_{\mathrm{Bare},\mathrm{int}}$$ (2.8) where we have included in $`_{\mathrm{Bare},\mathrm{int}}`$ all interaction and potential terms. More precisely, $`_{\mathrm{kin},\varphi _i}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\stackrel{~}{\mu }_i^{\epsilon /2}f_{\varphi _i}(\stackrel{~}{\mu }_{\mathrm{}})Z_{\varphi _i}\left(_\mu \varphi _i\right)^2`$ $`_{\mathrm{kin},\psi _j}`$ $`=`$ $`i\stackrel{~}{\mu }_j^{\epsilon /4}f_{\psi _j}(\stackrel{~}{\mu }_{\mathrm{}})Z_{\psi _j}\overline{\psi }_j_\mu \gamma ^\mu \psi _j`$ $`_{\mathrm{Bare},\mathrm{int}}`$ $`=`$ $`{\displaystyle \underset{b}{}}Z_{\lambda _b}f_{\lambda _b}(\stackrel{~}{\mu }_{\mathrm{}})\lambda _bO_b(Z_{\varphi _i}^{1/2}\varphi _i,Z_{\psi _j}^{1/2}\psi _j)`$ (2.9) where $`Z_{\varphi _i}`$ ($`Z_{\psi _j}`$) and $`Z_{\lambda _b}`$ are the bosonic (fermionic) wave function renormalizations and the coupling renormalizations respectively, and $`O_b(\varphi _i,\psi _j)`$ represent interaction operators between the fields. The $`\stackrel{~}{\mu }_i`$ scales are appropriate combinations of the independent scales $`\mu `$ and $`\mu _i`$, namely $$\mathrm{log}\stackrel{~}{\mu }_i=\mathrm{log}\mu \theta (\mu \mu _i)+\mathrm{log}\mu _i\theta (\mu _i\mu )$$ (2.10) Finally, the functions $`f_{\varphi _i}`$, $`f_{\psi _j}`$ and $`f_{\lambda _b}`$ are dimensionless functions of the ratios $`\stackrel{~}{\mu }_i/\stackrel{~}{\mu }_j`$ which are constant and finite and can be expanded as: $`f=1+𝒪(\mathrm{})`$. They correspond to finite wave-function and coupling renormalizations. They should be chosen so that the $`\beta `$ and $`\gamma `$-functions obtained from the bare lagrangian satisfy the decomposition given by Eqs. (2.1) and (2.6). Let us notice that for $`\mu <\mu _i`$ (all $`i`$), we have $`\stackrel{~}{\mu }_i=\mu _i`$, while for $`\mu >\mu _i`$ (all $`i`$), we have $`\stackrel{~}{\mu }_i=\mu `$. In the latter region, we recover the ordinary $`\overline{\mathrm{MS}}`$-scheme. At intermediate values of $`\mu `$, the situation is also intermediate: some of the $`\stackrel{~}{\mu }_i`$ become equal to the $`\overline{\mathrm{MS}}`$-scale $`\mu `$. It can be checked that the one-loop effective potential obtained from the lagrangian (2.8) has precisely the form of the proposed one-loop effective potential of Eq. (2.4). In terms of the $`\stackrel{~}{\mu }_i`$ scales defined in Eq. (2.10), it simply reads $$V_1=\frac{1}{64\pi ^2}\underset{i=1}{\overset{N}{}}()^{2s_i}M_i^4\left[\mathrm{log}\frac{M_i^2}{\stackrel{~}{\mu }_i^2}C_i\right]$$ (2.11) In the next section we will apply this approach to our simple extension of the Standard Model. ## 3 Effective theory of the Standard Model extension The presence of extra bosonic (fermionic) degrees of freedom tends to stabilize (destabilize) the SM Higgs potential. Consequently, when the SM Higgs potential presents an instability at a certain scale, $`\mathrm{\Lambda }_{SM}`$, the most economic cure is the presence of just extra bosonic fields. Consequently, in this section we will apply the results of section 2 to a very simple extension of the Standard Model defined by the lagrangian $$=_{\mathrm{SM}}+\frac{1}{2}\left(_\mu \stackrel{}{\mathrm{\Phi }}\right)^2\frac{1}{2}\delta |H|^2\stackrel{}{\mathrm{\Phi }}^2\frac{1}{2}M^2\stackrel{}{\mathrm{\Phi }}^2\frac{1}{4!}\lambda _s\stackrel{}{\mathrm{\Phi }}^4$$ (3.1) where $`_{\mathrm{SM}}`$ is the SM lagrangian, $`H`$ the SM Higgs doublet and $`\stackrel{}{\mathrm{\Phi }}`$ $`N`$ real scalar fields transforming under the vector representation of $`O(N)`$. $`M`$ is the invariant mass of $`\stackrel{}{\mathrm{\Phi }}`$, $`\lambda _s`$ its quartic coupling and $`\delta `$ provides the mixing between the Higgs and $`\stackrel{}{\mathrm{\Phi }}`$ fields. Since we are assuming that $`\stackrel{}{\mathrm{\Phi }}`$ is a SM singlet, $`\delta `$ is the only coupling that involves the SM with the new physics and will play a relevant role in our analysis. Of course, this model may be unrealistic, but, for the previous reasons, it basically represents the most efficent way to cure a SM instability, thus giving reliable upper bounds on the scale of the required new physics, i.e. the mass of the extra particles. In the present case we have two relevant scales: the SM scale, which can be conventionally chosen to be that corresponding to the top-quark, $`\mu _t`$, and the scale of the new physics, $`\mu _\mathrm{\Phi }`$. In the background of the Higgs field, $`H^0=h_c/\sqrt{2}`$, the one-loop effective potential in the decoupling approach explained in the previous section can be decomposed as in Eq. (2.3) with $$V_0=\frac{1}{2}m^2h_c^2+\frac{1}{4}\lambda h_c^4+\mathrm{\Lambda }_c$$ (3.2) where $`m^2`$, $`\lambda `$ and $`\mathrm{\Lambda }_c`$ are the SM mass term, quartic coupling and vacuum energy, respectively, and $`V_1`$ as given by (2.4) \[or, equivalently, by (2.11)\] $`V_1`$ $`=`$ $`{\displaystyle \frac{1}{64\pi ^2}}[12M_t^4(h_c)(\mathrm{log}{\displaystyle \frac{M_t^2(h_c)}{\mu _t^2}}+\theta (\mu \mu _t)\mathrm{log}{\displaystyle \frac{\mu _t^2}{\mu ^2}}{\displaystyle \frac{3}{2}})`$ (3.3) $`+`$ $`NM_\mathrm{\Phi }^4(h_c)(\mathrm{log}{\displaystyle \frac{M_\mathrm{\Phi }^2(h_c)}{\mu _\mathrm{\Phi }^2}}+\theta (\mu \mu _\mathrm{\Phi })\mathrm{log}{\displaystyle \frac{\mu _\mathrm{\Phi }^2}{\mu ^2}}{\displaystyle \frac{3}{2}})].`$ The masses of the top-quark and the $`\stackrel{}{\mathrm{\Phi }}`$ field are defined by $`M_t^2(h_c)`$ $`=`$ $`{\displaystyle \frac{\lambda _th_c^2}{\sqrt{2}}}`$ $`M_\mathrm{\Phi }^2(h_c)`$ $`=`$ $`M^2+\delta h_c^2,`$ (3.4) $`\lambda _t`$ being the SM top-quark Yukawa coupling. We are neglecting in Eq. (3.3) the contribution to the one-loop effective potential of all SM fields, except the top-quark field, as it is usually done in SM studies. To evaluate the potential given by Eqs. (3.2, 3.3) we also need to know the $`\lambda _a`$ parameters (i.e. all the masses, couplings and fields) at the corresponding values of the $`\mu ,\mu _t,\mu _\mathrm{\Phi }`$ scales. Hence we need the $`\beta `$ and $`\gamma `$-function decomposition defined in (2.1) and (2.6) for the relevant parameters of our model. This is given by $`{}_{\mu _t}{}^{}\stackrel{~}{\beta }_{\lambda }^{}`$ $`=`$ $`\beta _\lambda ^{\mathrm{SM}}{}_{\mu _\mathrm{\Phi }}{}^{}\stackrel{~}{\beta }_{\lambda }^{}={\displaystyle \frac{2N}{\kappa }}\delta ^2`$ (3.5) $`{}_{\mu _t}{}^{}\stackrel{~}{\beta }_{m^2}^{}`$ $`=`$ $`\beta _{m^2}^{\mathrm{SM}}{}_{\mu _\mathrm{\Phi }}{}^{}\stackrel{~}{\beta }_{m^2}^{}={\displaystyle \frac{2N}{\kappa }}\delta M^2`$ (3.6) $`{}_{\mu _t}{}^{}\stackrel{~}{\beta }_{\mathrm{\Lambda }}^{}`$ $`=`$ $`\beta _\mathrm{\Lambda }^{\mathrm{SM}}{}_{\mu _\mathrm{\Phi }}{}^{}\stackrel{~}{\beta }_{\mathrm{\Lambda }}^{}={\displaystyle \frac{N}{2\kappa }}M^4`$ (3.7) $`{}_{\mu _t}{}^{}\stackrel{~}{\gamma }_{h}^{}`$ $`=`$ $`\gamma _h^{\mathrm{SM}}{}_{\mu _\mathrm{\Phi }}{}^{}\stackrel{~}{\gamma }_{h}^{}=0`$ (3.8) for the SM couplings, where $`\kappa =16\pi ^2`$, and $`\beta ^{\mathrm{SM}}`$, $`\gamma ^{\mathrm{SM}}`$ are the SM $`\beta `$ and $`\gamma `$-functions. For couplings corresponding to new physics the $`\stackrel{~}{\beta }`$-functions are: $`{}_{\mu _t}{}^{}\stackrel{~}{\beta }_{\delta }^{}`$ $`=`$ $`2\gamma _h\delta +{\displaystyle \frac{24}{\kappa }}\lambda \delta `$ $`{}_{\mu _\mathrm{\Phi }}{}^{}\stackrel{~}{\beta }_{\delta }^{}`$ $`=`$ $`{\displaystyle \frac{1}{\kappa }}\left(8\delta ^2+{\displaystyle \frac{1}{3}}(N+2)\lambda _s\delta \right)`$ (3.9) $`{}_{\mu _t}{}^{}\stackrel{~}{\beta }_{\lambda _s}^{}`$ $`=`$ $`{\displaystyle \frac{48}{\kappa }}\delta ^2`$ $`{}_{\mu _\mathrm{\Phi }}{}^{}\stackrel{~}{\beta }_{\lambda _s}^{}`$ $`=`$ $`{\displaystyle \frac{1}{3\kappa }}(N+8)\lambda _s^2`$ (3.10) $`{}_{\mu _t}{}^{}\stackrel{~}{\beta }_{M^2}^{}`$ $`=`$ $`{\displaystyle \frac{8}{\kappa }}\delta m^2`$ $`{}_{\mu _\mathrm{\Phi }}{}^{}\stackrel{~}{\beta }_{M^2}^{}`$ $`=`$ $`{\displaystyle \frac{1}{3\kappa }}\lambda _s(N+2)M^2`$ (3.11) As noted in the previous section, and is apparent from (3.3), for scales $`\mu >\mu _t,\mu _\mathrm{\Phi }`$, the effective potential (3.3) reduces to the ordinary $`\overline{\mathrm{MS}}`$ effective potential. However, for an improved evaluation of the potential it is much more convenient to choose $`\mu \mu _tM_t(h_c)`$, $`\mu _\mathrm{\Phi }M_\mathrm{\Phi }(h_c)`$, thus getting rid of dangerous logarithms. (In the next section we will be more precise about the exact values of these choices.) These values belong to the range $`\mu _t\stackrel{<}{_{}}\mu <\mu _\mathrm{\Phi }`$, where the one-loop effective potential can be written as: $`V_1`$ $`=`$ $`V_1^{\mathrm{SM}}+V_1^\mathrm{\Phi }`$ $`V_1^\mathrm{\Phi }`$ $`=`$ $`{\displaystyle \frac{N}{64\pi ^2}}M_\mathrm{\Phi }^4(h_c)\left(\mathrm{log}{\displaystyle \frac{M_\mathrm{\Phi }^2(h_c)}{\mu _\mathrm{\Phi }^2}}{\displaystyle \frac{3}{2}}\right)`$ (3.12) Here $`V_1^{\mathrm{SM}}`$ is the usual SM one loop-effective potential in the $`\overline{\mathrm{MS}}`$-scheme, which depends explicitly on the RGE scale $`\mu `$, while $`V_1^\mathrm{\Phi }`$ corresponds to the contribution of the decoupled field $`\stackrel{}{\mathrm{\Phi }}`$, which runs with $`\mu _\mathrm{\Phi }`$. Expressions (3.2, 3) will be used in the next section to evaluate explicitly the effective potential. To finish this section, it is interesting to write the explicit form of the bare lagrangian in a multiscale renormalization scheme \[see Eqs.(2.8, 2.9)\] which leads to the effective potential (3) and the set (3.5)–(3.11) of $`\stackrel{~}{\beta }`$-functions. In the interesting range of scales, $`\mu _t\stackrel{<}{_{}}\mu <\mu _\mathrm{\Phi }`$ \[which, by Eq. (2.10) implies $`\stackrel{~}{\mu }_t=\mu `$, $`\stackrel{~}{\mu }_\mathrm{\Phi }=\mu _\mathrm{\Phi }`$\], this is explicitly given by $`_{\mathrm{Bare}}`$ $`=`$ $`_{\mathrm{Bare}}^{\mathrm{SM}}`$ $`+`$ $`{\displaystyle \frac{1}{2}}\mu _\mathrm{\Phi }^{\epsilon /2}Z_\mathrm{\Phi }\left(_\mu \stackrel{}{\mathrm{\Phi }}\right)^2{\displaystyle \frac{1}{2}}Z_\mathrm{\Phi }Z_{M^2}M^2\stackrel{}{\mathrm{\Phi }}^2{\displaystyle \frac{1}{4!}}Z_\mathrm{\Phi }^2Z_{\lambda _s}\lambda _s\stackrel{}{\mathrm{\Phi }}^4{\displaystyle \frac{1}{2}}Z_\mathrm{\Phi }Z_HZ_\delta f_\delta \delta |H|^2\stackrel{}{\mathrm{\Phi }}^2`$ where $`_{\mathrm{Bare}}^{\mathrm{SM}}`$ is the SM bare lagrangian, all couplings and fields are renormalized, and all constant factors $`Z_a`$ have the form $`Z_a=1+z_a/\epsilon +\mathrm{}`$, where the $`z_a`$ factors are those of the $`\overline{\mathrm{MS}}`$ scheme. The introduction of the finite renormalization of the coupling $`\delta `$, $$f_\delta =\left(\frac{\mu }{\mu _\mathrm{\Phi }}\right)^{4\delta /\kappa }$$ (3.14) is necessary in particular to satisfy the RGE given by Eqs. (3.9). The term in $`f_\delta `$, when expanded to one-loop order, $`f=1+(4\delta /\kappa )\mathrm{log}\frac{\mu }{\mu _\mathrm{\Phi }}+\mathrm{}`$ generates a (finite counterterm) contribution to the effective potential in the presence of background fields $`h_c`$ and $`\mathrm{\Phi }`$ as, $$\mathrm{\Delta }_cV(h_c,\mathrm{\Phi })=\frac{1}{16\pi ^2}\delta ^2h_c^2\stackrel{}{\mathrm{\Phi }}^2\mathrm{log}\frac{\mu }{\mu _\mathrm{\Phi }}$$ (3.15) which grows logarithmically with the scales ratio $`\mu /\mu _\mathrm{\Phi }`$. In principle, this is worrying, as it represents the kind of drawback of the pure $`\overline{\mathrm{MS}}`$-scheme in the presence of several scales that we wanted to avoid with our approach. Besides, this seems to contradict the form of the effective potential (3), which is free of such dangerous logarithms. However, when turning the background field $`\mathrm{\Phi }`$ on and computing the one-loop diagram with external legs $`|H|^2\stackrel{}{\mathrm{\Phi }}^2`$ and internal propagators corresponding to a Higgs and a $`\stackrel{}{\mathrm{\Phi }}`$ field <sup>5</sup><sup>5</sup>5In fact this is the only non-trivial one-loop diagram, in the sense that it contains internal lines corresponding to the SM and to new physics. This diagram is proportional to $`\delta ^2`$ and plays a major role in our construction., the contribution of the latter (after renormalization in the $`\overline{\mathrm{MS}}`$-scheme) precisely cancels that of the counterterm in Eq. (3.15) and, altogether, the presence of the dangerous $`\mathrm{log}(\mu /\mu _\mathrm{\Phi })`$ term, leading to the one-loop effective potential given in Eq. (3). ## 4 Numerical results In this section we will study the potential given by Eqs. (3.2, 3) and analyze the conditions for stability of the electroweak minimum at the vacuum expectation value (VEV) $`h_c=v=246`$ GeV, and the non-appearance of a (destabilizing) deeper minimum at larger values of the field. For given (fixed) values of the Higgs VEV $`v`$ and the Higgs mass squared $`m_H^2`$, the minimum conditions of the potential read<sup>6</sup><sup>6</sup>6With the definition of Eq.(4) the physical (pole) squared Higgs mass $`M_H^2`$ is equal to $`m_H^2`$ plus some (small) radiative corrections, which have been taken into acount in the numerical computations. $`{\displaystyle \frac{dV_{\mathrm{eff}}}{dh_c}}|_{h_c=v}`$ $`=`$ $`0`$ $`{\displaystyle \frac{d^2V_{\mathrm{eff}}}{dh_c^2}}|_{h_c=v}`$ $`=`$ $`m_H^2`$ (4.1) Using these conditions, the effective potential parameters, $`m^2`$ and $`\lambda `$, can be traded by $`v`$ and $`m_H^2`$ as: $`m^2`$ $`=`$ $`m_{\mathrm{SM}}^2{\displaystyle \frac{N\delta ^2}{\kappa }}v^2+{\displaystyle \frac{N\delta }{\kappa }}M^2\left(\mathrm{log}{\displaystyle \frac{M_\mathrm{\Phi }^2}{\mu _\mathrm{\Phi }^2}}1\right)`$ $`\lambda `$ $`=`$ $`\lambda _{\mathrm{SM}}{\displaystyle \frac{N\delta ^2}{\kappa }}\mathrm{log}{\displaystyle \frac{M_\mathrm{\Phi }^2}{\mu _\mathrm{\Phi }^2}}`$ (4.2) where $`m_{\mathrm{SM}}^2`$, $`\lambda _{\mathrm{SM}}`$ represent the corresponding values as obtained in the pure SM $`m_{\mathrm{SM}}^2`$ $`=`$ $`{\displaystyle \frac{1}{2}}m_H^2+{\displaystyle \frac{3\lambda _t^4}{\kappa }}v^2`$ $`\lambda _{\mathrm{SM}}`$ $`=`$ $`{\displaystyle \frac{m_H^2}{2v^2}}+{\displaystyle \frac{3\lambda _t^4}{\kappa }}\mathrm{log}{\displaystyle \frac{M_t^2}{\mu ^2}}.`$ (4.3) From the second equality in Eq. (4) we see that in order to preserve the validity of perturbation theory, a choice of the scale, $`\mu _\mathrm{\Phi }M_\mathrm{\Phi }`$, must be done, as expected. Furthermore, from the first equality in Eq. (4), we see that for $`M^2m_H^2`$ (which is the usual case) the third term of the right hand side will amount to a huge contribution, which must be fine-tuned with the value of $`m^2`$, in order to keep the right scale for $`m_H`$. This technical problem, which reflects a hierarchy problem, is avoided by choosing $`\mu _\mathrm{\Phi }`$ in such a way that the annoying term in (4) is cancelled, i.e. $$\mathrm{log}\frac{M_\mathrm{\Phi }^2}{\mu _\mathrm{\Phi }^2}=1.$$ (4.4) Then the parameter fixing (4) becomes, $`m^2`$ $`=`$ $`m_{\mathrm{SM}}^2{\displaystyle \frac{N\delta ^2}{\kappa }}v^2`$ $`\lambda `$ $`=`$ $`\lambda _{\mathrm{SM}}{\displaystyle \frac{N\delta ^2}{\kappa }}.`$ (4.5) Still perturbation theory can be spoiled, along with the SM vacuum, for values of $`N`$ and $`\delta `$ such that $`N\delta ^2\stackrel{>}{_{}}\kappa `$ so we will restrict ourselves to the range of values such that $`N\delta ^2\kappa `$. An immediate consequence of the choice (4.4) for $`\mu _\mathrm{\Phi }`$ is that all the couplings of the theory, $`\lambda _a`$, acquire an implicit $`h_c`$-dependence through their dependence on $`\mu _\mathrm{\Phi }`$. In fact this dependence can be obtained from the second equalities of Eqs. (3.53.11) using $$\frac{d\lambda _a(h_c)}{d\mathrm{log}h_c}=\left[1\frac{M^2}{M_\mathrm{\Phi }^2(h_c)}\right]_{\mu _\mathrm{\Phi }}\stackrel{~}{\beta }_{\lambda _a}\theta (\mu _\mathrm{\Phi }\mu ).$$ (4.6) We will consider now the effective potential given by Eqs. (3.2) and (3) and will improve it by the RGE (3.5) to (3.11) and (4.6) with the choice (4.4) of the $`\mu _\mathrm{\Phi }`$ scale. The initial conditions for all parameters will be taken at the boundary scales: $`\mu _0^2`$ $`=`$ $`M_t^2(v)`$ $`\mu _{\mathrm{\Phi },0}^2`$ $`=`$ $`M_\varphi ^2(v)/e`$ (4.7) and those for $`m^2`$ and $`\lambda `$ will be fixed by (4). The system of partial differential equations (3.5) to (3.11) is solved by a step-wise procedure, which allows to evaluate the effective potential for any value of $`h_c`$. In Fig. 1 we show a plot of the effective potential $`V_{\mathrm{eff}}`$ for values of the Higgs and top-quark masses, $`m_H=100`$ GeV and $`M_t(v)=175`$ GeV, for the SM (solid line), which shows an instability for values of the field $`h_c125`$ TeV $`\mathrm{\Lambda }_{SM}`$. The dotted line corresponds to $`V_{\mathrm{eff}}`$ for the SM extension with $`\delta =1`$, $`N=10`$ and $`M=400`$ TeV, which shows how the instability is cured by the new physics. Smaller values of $`M`$ also work. Conversely, the SM results are recovered in the limit when $`M\mathrm{}`$ or $`\delta 0`$. This is illustrated by the dashed line, which corresponds to $`M=420`$ TeV. The previous example explicitely shows that the scale of new physics, $`M`$, responsible for the cure of a SM instability, can be larger than the scale $`\mathrm{\Lambda }_{SM}`$ at which the SM instability develops. It cannot be, however, arbitrarily larger. Fig. 2 shows (solid line) the ratio $`M_{max}/\mathrm{\Lambda }_{SM}`$ for $`\delta =1`$, as a function of the number of extra degrees of freedom, $`N`$ (recall that $`N\delta ^2\kappa `$ to preserve perturbativity). Clearly, $`M`$ could be as large as $`10\mathrm{\Lambda }_{SM}`$, which puts an upper bound on the scale of new physics, $`M`$. For a typical value of the multiplicity, $`N=𝒪(10)`$, we get the conservative bound $`M\stackrel{<}{_{}}4\mathrm{\Lambda }_{SM}`$. This is e.g. the case of the MSSM, where the relevant multiplicity is $`N=12`$, corresponding to the stops. The dashed line corresponds to the result of Ref. , obtained in a cruder approximation (one-loop instead of RGE improved). Our results confirm the trend observed in that paper, but show that the ratio $`M_{max}/\mathrm{\Lambda }_{SM}`$ can be substantially larger than the one estimated there. Finally, Fig. 3 shows the smooth increasing of both, $`\mathrm{\Lambda }_{SM}`$ and $`M_{max}`$, with the Higgs mass, $`M_H`$. Also, there appears a lower bound on $`M`$, coming from the requirement of perturbativity up to the $`\mathrm{\Lambda }_{SM}`$ scale. This lower bound may seem paradoxical. Actually, it is a feature of the particular SM extension we have chosen: the lower $`M`$, the sooner the new physics enters, which, due to the RGE (3.5), may drive more quickly the quartic Higgs coupling $`\lambda `$ into a non-perturbative regime. Other SM extensions, in particular supersymmetric extensions, do not present such lower limits on $`M`$. On the other hand, the upper bound on $`M`$ is quite robust for any conceivable SM extension, as has been explained at the beginning of section 3. The numerical results presented in Figs. 1–3 show the relation between the scale $`\mathrm{\Lambda }_{SM}`$, at which the SM develops a instability, and the maximum value allowed for the scale of new physics required to cure it. With the present experimental lower bounds on $`M_H`$, $`M_H\stackrel{>}{_{}}105`$ GeV, it is clear that the possible new physics could easily escape detection in the present and future accelerators. ## 5 Conclusions The possible detection of a relatively light Higgs ($`M_H\stackrel{<}{_{}}140`$ GeV) would imply an instability of the Higgs effective potential, thus signaling the existence of new physics able to cure it. It is, therefore, a relevant question what is the relation between the Higgs mass (or, equivalently, the scale at which the instability develops, $`\mathrm{\Lambda }_{SM}`$) and the maximum allowed value of the scale of the new physics, $`M_{max}`$. In this paper we have examined this question in a rigorous way. This requires, in the first place, a reliable approach to evaluate the effective potential when several different mass scales are present. We have followed the decoupling approach exposed in a previous paper , which can also be re-formulated as a multi-scale renormalization approach, similar to those of Refs. . Then, we have considered a simple extension of the Standard Model, consisting of $`N`$ extra scalar fields with an invariant mass $`M`$, coupled to the Higgs field with a coupling $`\delta `$. This model, although unrealistic, arguably represents the most efficent way to cure a SM instability, thus giving reliable upper bounds on the scale of the required new physics, i.e. the mass of the extra particles. The numerical results, presented in section 4, in particular in Figs. 1–3, show the relation between $`\mathrm{\Lambda }_{SM}`$ and $`M_{max}`$. More precisely, for $`\delta =𝒪(1)`$ and $`N=𝒪(10)`$ (similar to the stop sector in the MSSM case), we obtain that $`M_{max}4\mathrm{\Lambda }_{SM}`$, which puts an upper bound on the scale of new physics. Unfortunately, the present lower bounds on the Higgs mass, $`M_H\stackrel{>}{_{}}105`$ GeV, imply that $`\mathrm{\Lambda }_{SM}`$ is at least $`10^2`$ TeV. This fact, toghether with the previous upper bound on $`M`$, implies that the new physics could easily (but not necessarily) escape detection in the present and future accelerators. ## Acknowledgments We thank F. Feruglio, H. Haber and F. Zwirner for very useful discussions.
warning/0002/math0002088.html
ar5iv
text
# On the size of Diophantine 𝑚-tuples ## 1 Introduction Let $`n`$ be a nonzero integer. A set of $`m`$ positive integers $`\{a_1,a_2,\mathrm{},a_m\}`$ is said to have the property $`D(n)`$ if $`a_ia_j+n`$ is a perfect square for all $`1i<jm`$. Such a set is called a Diophantine $`m`$-tuple (with the property $`D(n)`$), or $`P_n`$-set of size $`m`$. Diophantus found the quadruple $`\{1,\mathrm{\hspace{0.17em}33},\mathrm{\hspace{0.17em}68},\mathrm{\hspace{0.17em}105}\}`$ with the property $`D(256)`$. The first Diophantine quadruple with the property $`D(1)`$, the set $`\{1,\mathrm{\hspace{0.17em}3},\mathrm{\hspace{0.17em}8},\mathrm{\hspace{0.17em}120}\}`$, was found by Fermat (see ). Baker and Davenport proved that this Fermat’s set cannot be extended to the Diophantine quintuple, and a famous conjecture is that there does not exist a Diophantine quintuple with the property $`D(1)`$. The theorem of Baker and Davenport has been recently generalized to several parametric families of quadruples , but the conjecture is still unproved. On the other hand, there are examples of Diophantine quintuples and sextuples like $`\{1,\mathrm{\hspace{0.17em}33},\mathrm{\hspace{0.17em}105},\mathrm{\hspace{0.17em}320},\mathrm{\hspace{0.17em}18240}\}`$ with the property $`D(256)`$ and $`\{99,\mathrm{\hspace{0.17em}315},\mathrm{\hspace{0.17em}9920},\mathrm{\hspace{0.17em}32768},\mathrm{\hspace{0.17em}44460},\mathrm{\hspace{0.17em}19534284}\}`$ with the property $`D(2985984)`$ . The purpose of this paper is to find some upper bounds for the numbers $`M_n`$ defined by $$M_n=sup\{|S|:S\text{ has the property }D(n)\},$$ where $`|S|`$ denotes the number of elements in the set $`S`$. Considering congruences modulo $`4`$, it is easy to prove that $`M_{4k+2}=3`$ for all integers $`k`$ (see ). In we proved that if $`n2(mod4)`$ and $`n\{4,3,1,\mathrm{\hspace{0.17em}3},\mathrm{\hspace{0.17em}5},\mathrm{\hspace{0.17em}8},\mathrm{\hspace{0.17em}12},\mathrm{\hspace{0.17em}20}\}`$, then $`M_n4`$. Recently, we were able to prove that $`M_18`$ (see ). (As we said before, the conjecture is that $`M_1=4`$.) Since a set with the property $`D(4)`$ may contain at most two odd elements, this result implies $`M_410`$. Since the number of integer points on the elliptic curve $$y^2=(a_1x+n)(a_2x+n)(a_3x+n)$$ (1) is finite, we conclude that there does not exist an infinite set with the property $`D(n)`$. However, bounds for the size and for the number of solutions of (1) depend not only on $`n`$ but also on $`a_1,a_2,a_3`$. On the other hand, we may consider the hyperelliptic curve $$y^2=(a_1x+n)(a_2x+n)(a_3x+n)(a_4x+n)(a_5x+n)$$ (2) of genus $`g=2`$. Caporaso, Harris and Mazur proved that the Lang conjecture on varieties of general type implies that for $`g2`$ the number $`B(g,𝐊)=\mathrm{max}_C|C(𝐊)|`$ is finite. Here $`C`$ runs over all curves of genus $`g`$ over a number field $`𝐊`$, and $`C(𝐊)`$ denotes the set of all $`𝐊`$-rational points on $`C`$. However, even the question whether $`B(2,𝐐)<\mathrm{}`$ is still open. An example of Keller and Kulesz shows that $`B(2,𝐐)588`$ (see also ). Since $`M_n5+B(2,𝐐)`$ (by we have also $`M_n4+B(4,𝐐(\sqrt{n})`$), we see that the Lang conjecture implies that $$M=sup\{M_n:n𝐙\{0\}\}$$ is finite. At present we are able to prove only the weaker result that $`M_n`$ is finite for all $`n𝐙\{0\}`$. In the proof of this result we will try to estimate the number of ”large” (greater than $`|n|^3`$), ”small” (between $`n^2`$ and $`|n|^3`$) and ”very small” (less that $`n^2`$) elements of a set with the property $`D(n)`$. Let us introduce the following notation: $`A_n`$ $`=`$ $`sup\{|S[|n|^3,+\mathrm{}|:S\text{ has the property }D(n)\},`$ $`B_n`$ $`=`$ $`sup\{|Sn^2,|n|^3|:S\text{ has the property }D(n)\},`$ $`C_n`$ $`=`$ $`sup\{|S[1,n^2]|:S\text{ has the property }D(n)\}.`$ In estimating the number of ”large” elements, we used a theorem of Bennett on simultaneous approximations of algebraic numbers and a very useful gap principle. We proved ###### Theorem 1 $`A_n21`$ for all nonzero integers $`n`$. For the estimate of the number of ”small” elements we used a ”weak” variant of the gap principle and we proved ###### Theorem 2 $`B_n<0.65\mathrm{log}|n|+2.24`$ for all nonzero integers $`n`$. Finally, in the estimate of the number of ”very small” elements we used a large sieve method due to Gallagher and we proved ###### Theorem 3 $`C_n<265.55\mathrm{log}|n|(\mathrm{log}\mathrm{log}|n|)^2+9.01\mathrm{log}\mathrm{log}|n|`$ for $`|n|400`$. Since we checked that $`C_n5`$ for $`|n|400`$, we may combine Theorems 1, 2 and 3 to obtain ###### Theorem 4 $`M_n`$ $``$ $`32\text{for }|n|400,`$ $`M_n`$ $`<`$ $`267.81\mathrm{log}|n|(\mathrm{log}\mathrm{log}|n|)^2\text{for }|n|400.`$ ## 2 Large elements Assume that the set $`\{a,b,c,d\}`$ has the property $`D(n)`$. Let $`ab+n=r^2`$, $`ac+n=s^2`$, $`bc+n=t^2`$, where $`r,s,t`$ are nonegative integers. Eliminating $`d`$ from the system $$ad+1=x^2,bd+1=y^2,cd+1=z^2$$ we obtain the following system of Pellian equations $`az^2cx^2`$ $`=`$ $`n(ac),`$ (3) $`bz^2cy^2`$ $`=`$ $`n(bc).`$ (4) We will apply the following theorem of Bennett on simultaneous approximations of square roots of two rationals which are very close to $`1`$. ###### Theorem 5 () If $`c_i`$, $`p_i`$, $`q`$ and $`L`$ are integers for $`0i2`$, with $`c_0<c_1<c_2`$, $`c_j=0`$ for some $`0j2`$, $`q`$ nonzero and $`L>M^9`$, where $$M=\mathrm{max}\{|c_0|,|c_1|,|c_2|\},$$ then we have $$\underset{0i2}{\mathrm{max}}\left\{\left|\sqrt{1+\frac{c_i}{L}}\frac{p_i}{q}\right|\right\}>(130L\gamma )^1q^\lambda $$ where $$\lambda =1+\frac{\mathrm{log}(33L\gamma )}{\mathrm{log}\left(1.7L^2_{0i<j2}(c_ic_j)^2\right)}$$ and $$\gamma =\{\begin{array}{cc}\frac{(c_2c_0)^2(c_2c_1)^2}{2c_2c_0c_1}\hfill & \text{if }c_2c_1c_1c_0,\hfill \\ \frac{(c_2c_0)^2(c_1c_0)^2}{c_1+c_22c_0}\hfill & \text{if }c_2c_1<c_1c_0.\hfill \end{array}$$ We will apply Theorem 5 to the numbers $`\theta _1`$ $`=`$ $`{\displaystyle \frac{s}{a}}\sqrt{{\displaystyle \frac{a}{c}}}=\sqrt{{\displaystyle \frac{ac+n}{ac}}}=\sqrt{1+{\displaystyle \frac{n}{ac}}}=\sqrt{1+{\displaystyle \frac{nb}{abc}}},`$ $`\theta _2`$ $`=`$ $`{\displaystyle \frac{t}{b}}\sqrt{{\displaystyle \frac{b}{c}}}=\sqrt{{\displaystyle \frac{bc+n}{bc}}}=\sqrt{1+{\displaystyle \frac{n}{bc}}}=\sqrt{1+{\displaystyle \frac{na}{abc}}}.`$ ###### Lemma 1 Assume that $`a<b<c`$ and $`ac>n`$. Then all positive integer solutions $`x,y,z`$ of the system (3) and (4) satisfy $$\mathrm{max}(|\theta _1\frac{sbx}{abz}|,|\theta _2\frac{zay}{abz}|)<\frac{c|n|}{a}z^2.$$ Proof. We have $$\left|\frac{s}{a}\sqrt{\frac{a}{c}}\frac{sbx}{abz}\right|=\frac{s}{az\sqrt{c}}|z\sqrt{a}x\sqrt{c}|=\frac{s}{az\sqrt{c}}\frac{|n(ca)|}{z\sqrt{a}+x\sqrt{c}}.$$ If $`n<0`$, then $`s=\sqrt{ac|n|}<\sqrt{ac}`$ and we obtain $$|\theta _1\frac{sbx}{abz}|<\frac{\sqrt{ac}|n|c}{a\sqrt{ac}z^2}=\frac{c|n|}{a}z^2.$$ If $`n>0`$, then $`x\sqrt{c}>z\sqrt{a}`$ and we obtain $$|\theta _1\frac{sbx}{abz}|<\frac{\sqrt{ac+n}nc}{2a\sqrt{ac}z^2}=\sqrt{1+\frac{n}{ac}}\frac{cn}{2a}z^2<\frac{cn}{a}z^2.$$ In the same manner, we obtain $`|\theta _2\frac{tay}{abz}|<\frac{c|n|}{b}z^2<\frac{c|n|}{a}z^2`$. ###### Lemma 2 Let $`\{a,b,c,d\}`$, $`a<b<c<d`$, be a Diophantine quadruple with the property $`D(n)`$. If $`c>b^{11}|n|^{11}`$, then $`dc^{131}`$. Proof. Let $`r,s,t,x,y,z`$ be defined as in the beginning of this section. We will apply Theorem 5 with $`\{c_0,c_1,c_2\}=\{0,na,nb\}`$, $`L=abc`$, $`M=|nb|`$, $`q=abz`$, $`p_1=sbx`$, $`p_2=tay`$. Since $`abc>|n|^9b^9`$, the condition $`L>M^9`$ is satisfied. For the quantity $`\gamma `$ from Theorem 5 we have $`\gamma =\frac{b^2(ba)^2}{2ba}|n|^3`$ if $`b2a`$ and $`\gamma =\frac{a^2b^2}{a+b}|n|^3`$ if $`a<b2a`$. In both cases we have $$\frac{b^3}{6}|n|^3\gamma <\frac{b^3}{2}|n|^3.$$ For the quantity $`\lambda `$ from Theorem 5 we have $$\lambda =1+\frac{\mathrm{log}(33abc\gamma )}{\mathrm{log}(1.7c^2(ba)^2n^6)}=2\lambda _1,$$ where $$\lambda _1=\frac{\mathrm{log}\frac{1.7c}{33ab(ba)^2n^6\gamma }}{\mathrm{log}(1.7c^2(ba)^2n^6)}.$$ Theorem 5 and Lemma 1 imply $$\frac{c|n|}{az^2}>(130abc\gamma )^1(abz)^{\lambda _12}>(130abc\gamma )^1a^2b^2z^{\lambda _12}.$$ This implies $$z^{\lambda _1}<130a^2b^3c^2|n|\gamma $$ and $$\mathrm{log}z<\frac{\mathrm{log}(130a^2b^3c^2|n|\gamma )\mathrm{log}(1.7c^2(ba)^2n^6)}{\mathrm{log}(\frac{1.7c}{33ab(ba)^2n^6\gamma })}.$$ (5) Let us estimate the right hand side of (5). We have $$130a^2b^3c^2|n|\gamma <65a^2b^6c^2n^4<c^3\frac{65a^2}{b^5|n|^7}<c^3,$$ unless $`n=1`$, $`a=1`$, $`b=2`$. However, in it was proved that the Diophantine pair $`\{1,2\}`$ with the property $`D(1)`$ cannot be extended to a Diophantine quadruple. The same result implies also that if $`|n|=1`$, then $`ba>1`$. Therefore $$1.7c^2(ba)^2n^6<c^2.$$ Finally, $$\frac{1.7c}{33ab(ba)^2n^6\gamma }>0.103a^1b^6cn^9>c^{\frac{1}{11}}\frac{b^4|n|}{9.71a}>c^{\frac{1}{11}}.$$ The last estimate shows that $`\lambda _1>0`$, what we implicitly used in (5). Putting these three estimates in (5), we obtain $$\mathrm{log}z<\frac{3\mathrm{log}c2\mathrm{log}c}{\frac{1}{11}\mathrm{log}c}=66\mathrm{log}c.$$ Hence, $`z<c^{66}`$ and $$d=\frac{z^2n}{c}\frac{z^2+|n|}{c}<\frac{c^{132}+c^{\frac{1}{11}}}{c}<c^{131}+1.$$ Now we will develop a very useful gap principle for the elements of a Diophantine $`m`$-tuple. The principle is based on the following construction which generalizes the constructions of Arkin, Hoggatt and Strauss and Jones for the case $`n=1`$. ###### Lemma 3 If $`\{a,b,c\}`$ is a Diophantine triple with the property $`D(n)`$ and $`ab+n=r^2`$, $`ac+n=s^2`$, $`bc+n=t^2`$, then there exist integers $`e,x,y,z`$ such that $$ae+n^2=x^2,be+n^2=y^2,ce+n^2=z^2$$ and $$c=a+b+\frac{e}{n}+\frac{2}{n^2}(abe+rxy).$$ Proof. Define $$e=n(a+b+c)+2abc2rst.$$ Then $`(ae+n^2)(atrs)^2`$ $`=`$ $`an(a+b+c)+2a^2bc2arst+n^2`$ $`a^2(bc+n)+2arst(ab+n)(ac+n)=0.`$ Hence we may take $`x=atrs`$, and analogously $`y=bsrt`$, $`z=crst`$. We have $`abe+rxy`$ $`=`$ $`abn(a+b+c)+2a^2b^2c2abrst`$ $`+`$ $`abrsta(ab+n)(bc+n)b(ab+n)(ac+n)+rst(ab+n)`$ $`=`$ $`abcnn^2(a+b)+rstn,`$ and finally $$a+b+\frac{e}{n}+\frac{2}{n^2}(abe+rxy)=2a+2b+c+\frac{2abc}{n}\frac{2rst}{n}\frac{2abc}{n}2a2b+\frac{2rst}{n}=c.$$ ###### Lemma 4 If $`\{a,b,c,d\}`$ is a Diophantine quadruple with the property $`D(n)`$ and $`|n|^3a<b<c<d`$, then $$d>\frac{3.847bc}{n^2}.$$ Proof. We apply Lemma 3 to the triple $`\{a,c,d\}`$. Since $`ce+n^2`$ is a perfect square, we have that $`ce+n^20`$. On the other hand, the assumption is that $`c>|n|^3`$. Hence, if $`e1`$, then $`ce+n^2<|n|^3+n^2<0`$, a contradiction. Since $`e`$ is an integer, we have $`e0`$. If $`e=0`$, then $`d=a+c+2s`$. If $`e1`$, then $$d>a+c+\frac{2ac}{n^2}+\frac{2s\sqrt{ac}}{n^2}>\frac{2ac}{n^2}.$$ (6) (Note that if $`n>0`$ then $`x<0`$, $`y<0`$, and if $`n<0`$ and $`b>|n|`$ then $`x>0`$, $`y>0`$.) Analogously, applying Lemma 3 to the triple $`\{b,c,d\}`$ we obtain that $`d=b+c+2t`$ or $`d>b+c+\frac{2bc}{n^2}+\frac{2t\sqrt{bc}}{n^2}`$. However, $`d=b+c+2t`$ is impossible since $`b+c+2t>a+c+2s`$ and $$b+c+2tb+c+2\sqrt{c(c1)+n}<4c\frac{2ac}{n^2},$$ unless $`a<2n^2`$. But if $`|n|^3a<2n^2`$, then $`|n|=1`$, $`a=1`$, and in that case we have $$a+c+\frac{2ac}{n^2}+\frac{2s\sqrt{ac}}{n^2}>3c+2\sqrt{c(c1)}>4c.$$ Hence we proved that $$d>b+c+\frac{2bc}{n^2}+\frac{2t\sqrt{bc}}{n^2}.$$ (7) From we know that the triples $`\{1,2,3\}`$ and $`\{1,2,4\}`$ cannot be extended to Diophantine quadruples. Thus $`bc10`$ and it implies $$t^2=bc+nbc|n|>bc\sqrt[6]{bc}>0.853bc.$$ If we put this in (7), we obtain $`d>\frac{3.847bc}{n^2}`$. Proof of Theorem 1. Assume that $`\{a_1,a_2,\mathrm{},a_{22}\}`$ has the property $`D(n)`$ and $`|n|^3a_1<a_2<\mathrm{}<a_{22}`$. By Lemma 4 we find that $`a_4>{\displaystyle \frac{a_2^2}{n^2}},a_5>{\displaystyle \frac{a_2^3}{n^4}},a_6>{\displaystyle \frac{a_2^5}{n^8}},a_7>{\displaystyle \frac{a_2^8}{n^{14}}},`$ $`a_8>{\displaystyle \frac{a_2^{13}}{n^{24}}},a_9>{\displaystyle \frac{a_2^{21}}{n^{40}}},a_{10}>{\displaystyle \frac{a_2^{34}}{n^{66}}},a_{11}>{\displaystyle \frac{a_2^{55}}{n^{108}}}.`$ Since $`a_2>|n|^3`$, we have $`\frac{a_2^{55}}{n^{108}}>a_2^{11}|n|^{11}`$, and we may apply Lemma 2 with $`a=a_1`$, $`b=a_2`$, $`c=a_{11}`$. We conclude that $`a_{22}a_{11}^{131}`$. However, Lemma 4 implies $`a_{12}>|n|a_{11},a_{13}>{\displaystyle \frac{a_{11}^2}{|n|}},a_{14}>{\displaystyle \frac{a_{11}^3}{n^2}},a_{15}>{\displaystyle \frac{a_{11}^5}{|n|^5}},`$ $`a_{16}>{\displaystyle \frac{a_{11}^8}{|n|^9}},a_{17}>{\displaystyle \frac{a_{11}^{13}}{n^{16}}},a_{18}>{\displaystyle \frac{a_{11}^{21}}{|n|^{27}}},a_{19}>{\displaystyle \frac{a_{11}^{34}}{|n|^{45}}},`$ $`a_{20}>{\displaystyle \frac{a_{11}^{55}}{n^{74}}},a_{21}>{\displaystyle \frac{a_{11}^{89}}{|n|^{121}}},a_{22}>{\displaystyle \frac{a_{11}^{144}}{|n|^{197}}}.`$ Since $`a_{11}>a_2^{11}|n|^{11}>n^{44}`$, we obtain $$a_{22}>\frac{a_{11}^{144}}{|n|^{197}}a_{11}^{144\frac{197}{44}}>a_{11}^{139}>a_{11}^{131},$$ a contradiction. ## 3 Small elements ###### Lemma 5 If $`\{a,b,c,d\}`$ is a Diophantine quadruple with the property $`D(n)`$, $`|n|1`$, and $`n^2a<b<c<d`$, then $`c>3.88a`$ and $`d>4.89c`$. Proof. We will apply Lemma 3. Since $`b>n^2`$, we have $`e0`$. Thus Lemma 3 implies that $$ca+b+2r.$$ Since $`|n|1`$ we have $`ab20`$ and $`r^2ab\sqrt[4]{ab}>0.89ab>0.89a^2`$. Hence, $`c>3.88a`$. Since $`db+c+2t>a+c+2s`$, from (6) we conclude that $$d>a+c+\frac{2ac}{n^2}+\frac{2s\sqrt{ac}}{n^2}.$$ We have $`ac24`$ and $`s^2ac\sqrt[4]{ac}>0.9ac`$. Therefore $$d>a+c+\frac{3.89ac}{n^2}>4.89c.$$ Proof of Theorem 2. We may assume that $`|n|2`$ since $`B_1=B_1=0`$. Let $`\{a_1,a_2,\mathrm{},a_m\}`$ be a Diophantine $`m`$-tuple with the property $`D(n)`$ and $`n^2<a_1<a_2<\mathrm{}<|n|^3`$. By Lemma 5 we have $`a_3>3.88a_1,a_4>3.884.89a_1,\mathrm{},a_m>3.884.89^{m3}a_1.`$ Therefore $$3.884.89^{m3}n^2<|n|^3$$ and from $`m3<\frac{\mathrm{log}\frac{|n|}{3.88}}{\mathrm{log}4.89}`$ we obtain $`m<0.65\mathrm{log}|n|+2.24`$. ## 4 Very small elements We are left with the task to estimate the number of ”very small” elements in a Diophantine $`m`$-tuple. Let $`\{a_1,a_2,\mathrm{},a_m\}`$ be a Diophantine $`m`$-tuple with the property $`D(n)`$ and assume that $`a_1<a_2<\mathrm{}<a_mN`$, where $`N`$ is a positive integer. Let $`1k<m`$. Then $`x=a_{k+1},\mathrm{},x=a_m`$ satisfy the system $$a_1x+n=\mathrm{},a_2x+n=\mathrm{},\mathrm{},a_kx+n=\mathrm{},$$ (8) where $`\mathrm{}`$ denotes a square of an integer. Denote by $`Z_k(N)`$ the number of solutions of system (8) satisfying $`1xN`$. Motivated by the observations from the introduction of , we will apply a sieve method based on the following theorem of Gallagher (see also \[25, p.29\]): ###### Theorem 6 () If all but $`g(q)`$ residue classes $`(\mathrm{mod}q)`$ are removed for each prime power $`q`$ in a finite set $`𝒮`$, then the number of integers which remain in any interval of length $`N`$ is at most $$\left(\underset{q𝒮}{}\mathrm{\Lambda }(q)\mathrm{log}N\right)/\left(\underset{q𝒮}{}\frac{\mathrm{\Lambda }(q)}{g(q)}\mathrm{log}N\right)$$ provided the denominator is positive. Here $`\mathrm{\Lambda }(q)=\mathrm{log}p`$ for $`q=p^\alpha `$. We will use Theorem 6 to estimate the number $`Z_k(N)`$. For this purpose, we will take $$𝒮=\{p:p\text{ is prime, }83pQ\text{}\mathrm{gcd}(a_1a_2\mathrm{}a_k,p)=1\},$$ where $`Q`$ is sufficiently large. For a prime $`p𝒮`$ we may remove all residue classes $`(\mathrm{mod}p)`$ such that $`\left(\frac{a_ix+n}{p}\right)=1`$ for some $`i\{1,\mathrm{},k\}`$. Here $`(\frac{}{p})`$ denotes the Legendre symbol. Let $`1lk`$. Then $`g(p)`$ $``$ $`|\{x𝐅_p:\left({\displaystyle \frac{a_ix+n}{p}}\right)=0\text{or }1\text{, for }i=1,\mathrm{},l\}|`$ $``$ $`l+|\{x𝐅_p:\left({\displaystyle \frac{x+n\overline{a_i}}{p}}\right)=\left({\displaystyle \frac{\overline{a_i}}{p}}\right),\text{for }i=1,\mathrm{},l\}|.`$ Here $`a_i\overline{a_i}1(modp)`$. Using estimates for character sums (see \[29, p.325\]), we obtain $$g(p)l+\frac{p}{2^l}+\left(\frac{l2}{l}+\frac{1}{2^l}\right)\sqrt{p}+\frac{l}{2}.$$ Assume that $`k=\mathrm{log}_2Q`$. We may take $`l=\mathrm{log}_2p`$. Then we have $$\frac{p}{2^l}+\frac{\sqrt{p}}{2^l}+\frac{3l}{2}<2+\frac{2}{\sqrt{p}}+\frac{3\mathrm{log}_2p}{2}<\sqrt{p}$$ for $`p179`$. Hence $$l+\frac{p}{2^l}+\left(\frac{l2}{2}+\frac{1}{2^l}\right)\sqrt{p}+\frac{l}{2}<\frac{l}{2}\sqrt{p}<\frac{\mathrm{log}_2p}{2}\sqrt{p}<0.722\sqrt{p}\mathrm{log}p$$ for $`p179`$, and we may check directly that $`l+\frac{p}{2^l}+\left(\frac{l2}{2}+\frac{1}{2^l}\right)+\frac{l}{2}<0.722\sqrt{p}\mathrm{log}p`$ for $`83p173`$. Therefore we proved that $$g(p)<0.722\sqrt{p}\mathrm{log}p.$$ By Theorem 6, we have $`Z_k(N)\frac{E}{F}`$, where $$E=\underset{p𝒮}{}\mathrm{log}p\mathrm{log}N,F=\underset{p𝒮}{}\frac{1}{0.722\sqrt{p}}\mathrm{log}N.$$ By \[33, Theorem 9\], we have $`E<_{83pQ}\mathrm{log}p<\theta (Q)<1.01624Q`$. Assume that at least $`\frac{4}{5}\pi (Q)`$ primes less than $`Q`$ satisfy the condition $`\mathrm{gcd}(a_1a_2\mathrm{}a_k,p)=1`$. Then we have $`F`$ $``$ $`{\displaystyle \frac{1}{0.722\sqrt{Q}}}|𝒮|\mathrm{log}N{\displaystyle \frac{1}{0.722\sqrt{Q}}}\left({\displaystyle \frac{4}{5}}\pi (Q)23\right)\mathrm{log}N`$ (9) $`>`$ $`1.108{\displaystyle \frac{\sqrt{Q}}{\mathrm{log}Q}}{\displaystyle \frac{31.86}{\sqrt{Q}}}\mathrm{log}N.`$ Since $`F`$ must be positive in the applications of Theorem 6, we will choose $`Q`$ of the following form $$Q=c_1\mathrm{log}^2N(\mathrm{log}\mathrm{log}N)^2,$$ (10) where $`c_1`$ is a constant. We have to check whether our assumption is correct. Suppose that $`a=a_1a_2\mathrm{}a_k`$ is divisible by at least one fifth of the primes $`Q`$. Then $`ap_1p_2\mathrm{}p_{\frac{1}{5}\pi (Q)}`$, where $`p_i`$ denotes the $`i^{\mathrm{th}}`$ prime. By \[33, p.69\], we have $$p_{\frac{1}{5}\pi (Q)}>\frac{1}{5}\pi (Q)\mathrm{log}(\frac{1}{5}\pi (Q))>\frac{1}{5}\frac{Q}{\mathrm{log}Q}\mathrm{log}\left(\frac{1}{5}\frac{Q}{\mathrm{log}Q}\right):=R.$$ Therefore, by \[33, p.70\], $$\mathrm{log}a>\underset{pR}{}\mathrm{log}p>R\left(1\frac{1}{\mathrm{log}R}\right).$$ Assume that $`Q210^4`$. Then $`\frac{1}{5}\frac{Q}{\mathrm{log}Q}>Q^{0.605}`$ and $`R>0.128Q`$. Furthermore, $`\mathrm{log}R>7.793`$ and therefore $$\mathrm{log}a>0.105Q.$$ On the other hand, $`a<N^k`$ and $`\mathrm{log}a<k\mathrm{log}N\mathrm{log}_2Q\mathrm{log}N`$. Assume that $`N1.610^5`$ and $`c_180`$. Then $`Q\mathrm{log}^{4.498}N`$. In order to obtain a contradiction, it suffices to check that $$0.105c_1\mathrm{log}^2N(\mathrm{log}\mathrm{log}N)^2>\frac{4.498}{\mathrm{log}2}\mathrm{log}N\mathrm{log}\mathrm{log}N$$ or $$c_1\mathrm{log}N\mathrm{log}\mathrm{log}N>61.81,$$ and this is certainly true for $`N1.610^5`$ if we choose $`c_12.08`$. Thus we may continue with estimating the quantity $`F`$. We are working under assumptions that (10) holds with $`2.08c_180`$, $`Q210^4`$ and $`N1.610^5`$. We would like to have the estimate of the form $$F>\frac{\sqrt{Q}}{c_2\mathrm{log}Q}.$$ (11) This estimate will lead to $$Z_k(N)<1.01624c_2\sqrt{Q}\mathrm{log}Q<4.572c_2\sqrt{c_1}\mathrm{log}N(\mathrm{log}\mathrm{log}N)^2.$$ (12) In order to fulfill (11), it suffice to check $$\frac{31.86}{\sqrt{Q}}+\mathrm{log}N<\frac{\sqrt{Q}}{\mathrm{log}Q}\left(1.108\frac{1}{c_2}\right).$$ Since $`Q>210^4`$ we have $`\frac{31.86}{\sqrt{Q}}<0.016\frac{\sqrt{Q}}{\mathrm{log}Q}`$. Furthermore, $$\frac{\mathrm{log}N\mathrm{log}Q}{\sqrt{Q}}<\frac{4.498\mathrm{log}N\mathrm{log}\mathrm{log}N}{\sqrt{c_1}\mathrm{log}N\mathrm{log}\mathrm{log}N}=\frac{4.498}{\sqrt{c_1}}.$$ Hence $`c_2>1/(1.092\frac{4.498}{\sqrt{c_1}})`$. Thus if we choose $`c_1=68`$, then we may take $`c_2=1.83`$ and from (12) we obtain $$Z_k(N)<69\mathrm{log}N(\mathrm{log}\mathrm{log}N)^2.$$ (13) Note that with this choice of $`c_1`$, $`N1.610^5`$ implies $`Q>60222>210^4`$. Proof of Theorem 3. Let $`\{a_1,a_2,\mathrm{},a_m\}`$ be a Diophantine $`m`$-tuple with the property $`D(n)`$ and $`a_1<a_2<\mathrm{}<a_mn^2`$. Then for any $`k\{1,2,\mathrm{},m\}`$ we have $$mk+Z_k(n^2).$$ Let $`k=\mathrm{log}_2Q`$, where $`Q=68\mathrm{log}^2n^2(\mathrm{log}\mathrm{log}n^2)^2`$. Since $`|n|400`$, we have $`n^21.610^5`$ and we may apply formula (13) to obtain $$Z_k(n^2)<69\mathrm{log}n^2(\mathrm{log}\mathrm{log}n^2)^2<265.55\mathrm{log}|n|(\mathrm{log}\mathrm{log}|n|)^2,$$ (14) Furthermore, $$k<\frac{1}{\mathrm{log}_2}\mathrm{log}(\mathrm{log}^{4.489}n^2)<9.01\mathrm{log}\mathrm{log}|n|,$$ (15) and combining (14) and (15) we finally obtain $$m<265.55\mathrm{log}|n|(\mathrm{log}\mathrm{log}|n|)^2+9.01\mathrm{log}\mathrm{log}|n|.$$ ###### Remark 1 In Katalin Gyarmati recently considered the more general problem. She estimated $`\mathrm{min}\{|𝒜|,||\}`$, where $`𝒜,\{1,2,\mathrm{},N\}`$ satisfy the condition that $`ab+n`$ is a perfect square for all $`a𝒜`$, $`b`$. Using her approach, it can be deduced that if $`\{a_1,a_2,\mathrm{},a_m\}`$ has the property $`D(n)`$, where $`n>0`$ and $`a_1<a_2<\mathrm{}<a_mN`$, then $`m2n\mathrm{log}N`$. This yields $`C_n4n\mathrm{log}n`$ for $`n2`$. ###### Remark 2 Let us mention that Rivat, Sárközy and Stewart recently used Gallagher’s ”larger sieve” method is estimating the size of a set $`Z`$ of integers such that $`z+z^{}`$ is a perfect square whenever $`z`$ and $`z^{}`$ are distinct elements of $`Z`$. They proved that if $`Z\{1,2,\mathrm{},N\}`$, where $`N`$ is greater that an effectively computable constant, then $`|Z|<37\mathrm{log}N`$. Largest known set with the above property is a set with six elements found by J. Lagrange . Maybe this may be compared with our situation where the largest known Diophantine $`m`$-tuples are Diophantine sextuples found by Gibbs . Proof of Theorem 4. Since $`M_nA_n+B_n+C_n`$, the second part of the theorem follows directly from Theorems 1, 2 and 3. For $`|n|400`$, Theorem 2 gives $`B_n6`$. It is easy to verify with a computer that for $`|n|400`$ it holds $`C_n5`$. More precisely, $`C_n=5`$ if and only if $`n\{299,255,256,400\}`$. These two estimates together with Theorem 1 imply $`M_n32`$. ## 5 Concluding remarks It is not surprising that in Theorem 4 the main contribution comes from $`C_n`$. Namely, if we define $`C=sup\{C_n:n𝐙\{0\}\}`$, then we have $`M=C`$. Indeed, if $`\{a_1,a_2,\mathrm{},a_m\}`$ is a Diophantine $`m`$-tuple with the property $`D(n)`$, then $`\{a_1c,a_2c,\mathrm{},a_mc\}`$ has the property $`D(nc^2)`$ and for sufficiently large $`c`$ we have $`a_ic(nc^2)^2`$, $`i=1,2,\mathrm{},m`$. It means that in order to prove $`M<\mathrm{}`$, it suffices to prove $`C<\mathrm{}`$. The above argumentation shows that it suffice to prove that for some $`\epsilon >0`$ it holds $$\underset{n0}{sup}sup\{|S[1,n^{0.5+\epsilon }]|:S\text{ has the property }D(n)\}<\mathrm{}.$$ We may define also $`A=sup\{A_n:n𝐙\{0\}\}`$ and $`B=sup\{B_n:n𝐙\{0\}\}`$. Gibbs’ example mentioned in introduction shows that $`C6`$ and $`M6`$. If $`n=a^2`$, $`a5`$, then $`B_n3`$ since $`\{a^2+1,a^2+2a+1,\mathrm{\hspace{0.17em}4}a^2+4a+4\}`$ has the property $`D(a^2)`$. Hence $`B3`$. Finally, since $`\{k,k+2,\mathrm{\hspace{0.17em}4}k+4,\mathrm{\hspace{0.17em}16}k^3+48k^2+44k+12\}`$ has the property $`D(1)`$ we have $`AA_14`$. Department of Mathematics, University of Zagreb, Bijenička cesta 30, 10000 Zagreb, Croatia E-mail address: duje@math.hr
warning/0002/hep-ph0002175.html
ar5iv
text
# Anomalous triple and quartic gauge boson couplings ## 1 Introduction Triple and quartic gauge boson couplings arise in the Standard Model due to the non-abelian nature of the theory and are, therefore, a fundamental prediction. The study of these couplings is mainly motivated by the hope that some new physics may result in a modification of the couplings. If the new physics occurs at an energy scale well above that being probed experimentally, it is possible to integrate it out. The result is an effective theory with unknown coefficients of the various operators in the Lagrangian. Any theory beyond the Standard Model should be able to predict these coefficients. However, lacking a serious candidate for such a theory, they are simply treated as anomalous couplings. In most analyses of the past (see e.g. reference for a review) the Lagrangian was required to conserve $`C`$ and $`P`$ separately. This was mainly motivated by the fact that this requirement leads to a reduction of unknown parameters. Furthermore, the analyses concentrated on triple gauge boson couplings, since the obtainable constraints for these couplings are much stronger. It is not the aim of this article to review these analyses, but rather to present two new possibilities to look for anomalous couplings. The first is concerned with the search for $`CP`$-violating triple gauge boson couplings and the second with anomalous quartic couplings. Both are done for LEP and they will be discussed in sections 2 and 3 respectively. An extension of the study of anomalous quartic couplings to the Tevatron is presented in section 4. In the case of a hadron collider, the search for anomalous couplings is complicated by the fact that form factors have to be introduced. Indeed, since the inclusion of anomalous couplings spoils the gauge cancellation in the high energy limit, the effective theory will lead to violation of unitarity for increasing partonic center of mass energy $`\widehat{s}`$. In order to prevent this, a suppressing factor is needed. The form of this factor and the scale of new physics associated with it is to a large extent arbitrary. This introduces an unwanted dependence of limits on anomalous couplings on the precise form of the form factor. In section 5 we investigate to which extent this arbitrariness in measurements of anomalous couplings at hadron colliders can be avoided. ## 2 $`CP`$-violating gauge boson couplings at LEP The study of the helicity of the intermediate state $`W`$ bosons gives direct access to a model independent test of the Standard Model. The values of both $`CP`$ conserving and $`CP`$ violating anomalous Trilinear Gauge boson couplings can be directly measured by comparing the W helicity properties with those predicited in the Standard Model. According to the most general Lorentz invariant Lagrangian , there are 14 independent couplings describing the WWV vertex $`(\mathrm{V}=\mathrm{Z},\gamma )`$. Within the $`SU(2)_L\times U(1)_Y`$ theory, $`CP`$ violation is only present in the $`\mathrm{e}^+\mathrm{e}^{}\mathrm{W}^+\mathrm{W}^{}`$ process via the Kobayashi-Maskawa phase , which affects it at the two loop order only. $`CP`$ violating terms for the trilinear $`\gamma `$WW and $`\mathrm{Z}^0`$WW interactions are, however, easily included in the $`SU(2)_L\times U(1)_Y`$ Lagrangian . There are then 4 couplings which violate $`P`$ and $`CP`$ invariance, $`\stackrel{~}{\kappa }_V`$ and $`\stackrel{~}{\lambda }_V`$, and 2 which violate $`C`$ and $`CP`$ invariance $`g_4^V`$. Within the Standard Model all these couplings are zero but a linear realization of the basic $`SU(2)_L\times U(1)_Y`$ symmetry gives the following relations between the $`CP`$ violating couplings: $$\stackrel{~}{\kappa }_Z=\mathrm{tan}^2\theta _w\stackrel{~}{\kappa }_\gamma ;\stackrel{~}{\lambda }_Z=\stackrel{~}{\lambda }_\gamma ;g_4^Z=g_4^\gamma $$ (1) One method of measuring the $`CP`$-violating couplings is the Spin Density Matrix (SDM) analysis. The two-particle joint SDM completely describes the helicity of the $`W`$ bosons produced in the triple gauge boson interaction. The matrix elements are observables directly related to the polarisation of the $`W`$ bosons and so their measurement will give direct access to the underlying physics of the WW production process and allow a model independent test of the TGCs. This method of analysis is extremely desirable for investigating the $`CP`$-violating couplings, because a number of the SDM elements’ coefficients are particularly sensitive to the $`CP`$-violating couplings while being unaffected by changes in the $`CP`$-conserving couplings. The matrix elements are normalised products of the helicity amplitudes of the $`W^+`$ and $`W^{}`$. The matrix is hermitian and therefore has 80 independent elements if the off-diagonal elements are complex. This results in 80 independent coefficients to be experimentally measured. The diagonal elements are purely real and are equivalent to the probability of producing a final WW state with helicity $`\tau _{}\tau _+`$ (where $`\tau _{}`$ and $`\tau _+`$ are the helicity states of the $`W^+`$ and $`W^{}`$). The off-diagonal elements are the cross terms from the interference of all the possible final states. The number of independent elements can be further reduced by only considering the decay of one $`W`$ and summing over all possible helicity states of the other. This single $`W`$ SDM has only nine elements. The single $`W`$ SDM matrix is hermitian, the off-diagonal components of which are once again complex, leading to the nine independent SDM coefficients. The diagonal elements are real and can be interpreted as the probability of producing a $`W`$ boson of the respective helicity, $`\tau `$. Therefore, they are normalised to unity. The imaginary SDM coefficients are extremely sensitive to $`CP`$ violation at the three gauge boson vertex but completely insensitive to $`CP`$ conserving anomalous couplings. However, in a theory with no $`CP`$ violation at the vertex, any deviation from zero in the imaginary SDM coefficients could only be due to loop effects. The unnormalised single $`W`$ SDM elements can be extracted from the data of the decay product angles by integrating with suitable spin projection operators that reflect the standard V-A coupling of fermions to the $`W`$ boson in the $`W`$ decay. The theoretical predictions for the single $`W`$ SDM elements as a function of anomalous couplings can be derived from the analytical expressions of the helicity amplitudes . The single $`W`$ matrix elements can be extracted using the three-fold differential cross section . This extraction method uses the data event by event, so each event is analysed individually and then the sum of all events in the bin is taken. Certain projection operators are symmetric under the transformation $`\mathrm{cos}\theta ^{}\mathrm{cos}\theta ^{}`$, $`\varphi ^{}\varphi ^{}+\pi `$ and so a number of the SDM elements (or combinations thereof), can be extracted from the folded angular distribution of the hadronically decaying $`W`$ in the semileptonic event, where differentiation between the particle and anti-particle decay product is extremely difficult. Figure 1 shows the SDM elements for the Standard Model (solid), $`\stackrel{~}{\lambda }_Z=+1`$ (dotted) and $`\stackrel{~}{\lambda }_Z=1`$ (dashed). This also means using suitable combinations of projection operators, certain combinations of the joint particle SDM elements, can be extracted from the 5 fold differential cross section. The SDM elements are directly related to the polarisation of the W bosons, so they can be used to extract the polarised differential cross sections from the data. Figure 2 shows the above polarised differential cross section for the Standard Model (solid), $`\stackrel{~}{\kappa }_Z=+1`$ (dotted) and $`\stackrel{~}{\kappa }_Z=1`$ (dashed). A study of the $`CP`$-violating couplings has been performed at OPAL using W pair events which decay semileptonically, from the data recorded in 1998, at a centre-of-mass energy of 189 GeV with an integrated luminosity of 183 pb<sup>-1</sup>. ## 3 Anomalous quartic couplings at LEP The lowest dimension operators which lead to genuine quartic couplings where at least one photon is involved are of dimension 6 . The two most commonly studied are $`_0`$ $`=`$ $`{\displaystyle \frac{e^2}{16\mathrm{\Lambda }^2}}a_0F^{\mu \nu }F_{\mu \nu }\stackrel{}{W^\alpha }\stackrel{}{W_\alpha }`$ (2) $`=`$ $`{\displaystyle \frac{e^2}{16\mathrm{\Lambda }^2}}a_0\left[2(p_1p_2)(AA)+2(p_1A)(p_2A)\right]`$ $`\times \left[2(W^+W^{})+(ZZ)/\mathrm{cos}^2\theta _w\right],`$ $`_c`$ $`=`$ $`{\displaystyle \frac{e^2}{16\mathrm{\Lambda }^2}}a_cF^{\mu \alpha }F_{\mu \beta }\stackrel{}{W^\beta }\stackrel{}{W_\alpha }`$ (3) $`=`$ $`{\displaystyle \frac{e^2}{16\mathrm{\Lambda }^2}}a_c[(p_1p_2)A^\alpha A_\beta +(p_1A)A^\alpha p_{2\beta }.`$ $`.+(p_2A)p_1^\alpha A_\beta (AA)p_1^\alpha p_{2\beta }]`$ $`\times \left[W_\alpha ^{}W^{+\beta }+W_\alpha ^+W^\beta +Z_\alpha Z^\beta /\mathrm{cos}^2\theta _w\right].`$ both giving anomalous contributions to the $`VV\gamma \gamma `$ vertex, with $`VV`$ either being $`W^+W^{}`$ or $`Z^0Z^0`$, where $`p_1`$ and $`p_2`$ are the photon momenta and $`\stackrel{}{W_\mu }=\left(\begin{array}{c}\frac{1}{\sqrt{2}}(W_\mu ^++W_\mu ^{})\\ \frac{i}{\sqrt{2}}(W_\mu ^+W_\mu ^{})\\ \frac{Z_\mu }{\mathrm{cos}\theta _w}\end{array}\right).`$ (7) Anomalous $`VZ\gamma \gamma `$ vertices can in principle also be considered. Since the sensitivity to those is much smaller we restrict ourselves in this analysis to the two anomalous parameters $`a_0`$ and $`a_c`$. For a complete set of operators of this type see Ref. . The anomalous scale parameter $`\mathrm{\Lambda }`$ that appears in the above anomalous contributions has to be fixed. In practice, $`\mathrm{\Lambda }`$ can only be meaningfully specified in the context of a specific model for the new physics giving rise to the quartic couplings. However, in order to make our analysis independent of any such model, we choose to fix $`\mathrm{\Lambda }`$ at a reference value of $`M_W`$, following the conventions adopted in the literature. Any other choice of $`\mathrm{\Lambda }`$ (e.g. $`\mathrm{\Lambda }=1`$ TeV) results in a trivial rescaling of the anomalous parameters $`a_0`$ and $`a_c`$. It follows from the Lagrangian that any anomalous contribution is linear in the photon energy $`E_\gamma `$. This means that it is the hard tail of the photon energy distribution that is most affected by the anomalous contributions, but unfortunately the cross sections here are very small. In the following numerical studies we will impose a lower energy photon cut of $`E_\gamma ^{\mathrm{min}}=20`$ GeV. Similarly, there is also no anomalous contribution to the initial-state photon radiation, and so the effects are largest for centrally-produced photons. We therefore impose an additional cut of $`|\eta _\gamma |<2`$<sup>1</sup><sup>1</sup>1Obviously in practice these cuts will be tuned to the detector capabilities.. We do not include any branching ratios or acceptance cuts on the decay products of the produced $`W^\pm `$ and $`Z^0`$ bosons, since we assume that at $`e^+e^{}`$ colliders the efficiency for detecting these is high. Figure 3 shows the contour in the $`(a_0,a_c)`$ plane that corresponds to a $`+3\sigma `$ deviation of the $`WW\gamma `$ and $`Z^0\gamma \gamma `$ SM cross sections at $`\sqrt{s}=200`$ GeV with $`=150`$ pb<sup>-1</sup>. The key features in determining the sensitivity for a given process, apart from the fundamental process dynamics, are the available photon energy $`E_\gamma `$, the ratio of anomalous diagrams to SM ‘background’ diagrams, and the polarisation state of the weak bosons . A high-energy linear collider ($`\sqrt{s}5001000`$ GeV), would allow more phase space for photon emission, and would give significantly tighter bounds on the coupling, see Ref. . At LEP2 energies $`Z^0\gamma \gamma `$ benefits kinematically from producing only one massive boson, which leaves more energy for the photons as well as having fewer ‘background’ diagrams. On the other hand $`W^+W^{}\gamma `$ production at this collision energy suffers from the lack of phase space available for energetic photon emission, although this is partially compensated by the production of longitudinal bosons, which gives rise to higher sensitivity to the anomalous couplings. Finally, it is important to emphasise that in our study we have only considered ‘genuine’ quartic couplings from new six-dimensional operators. We have assumed that all other anomalous couplings are zero, including the trilinear ones. Since the number of possible couplings and correlations is so large, it is in practice very difficult to do a combined analysis of all couplings simultaneously. In fact, it is not too difficult to think of new physics scenarios in which effects are only manifest in the quartic interactions. One example would be a very heavy excited $`W`$ resonance produced and decaying as in $`W^+\gamma W^{}W^+\gamma `$. ## 4 Anomalous quartic couplings at the Tevatron Motivated by a request from experimentalists at the Workshop, we investigated the sensitivity of the processes $`p\overline{p}W^+W^{}\gamma `$ and $`Z\gamma \gamma `$ to the above anomalous quartic couplings, $`a_0`$ and $`a_c`$. We consider a Tevatron scenario of $`\sqrt{s}=2`$ TeV with an integrated luminosity $`=2`$ fb<sup>-1</sup> and impose a transverse momentum cut $`p_\gamma >10`$ GeV and a rapidity cut of $`|\eta _\gamma |<2.5`$ on the final-state photon(s). It can be seen from Figure 4 that the mean partonic centre-of-mass energy is $`250`$ GeV and hence it is possible to perform the analysis without the need to introduce a form factor. For ease of comparison with the LEP results, we again choose the anomalous scaling parameter $`\mathrm{\Lambda }=M_W`$. For purposes of illustration we only consider here the sensitivity of the cross sections to one of the anomalous parameters, $`a_0`$, since this one has the highest sensitivity. Thus Figure 4 shows the partonic centre-of-mass spectrum corresponding to $`a_0=0,100,500`$, with $`a_c=0`$. Again for the purpose of illustration we have chosen here to display the results for the process $`W^+W^{}\gamma `$ only. Similar results are found for $`Z\gamma \gamma `$ production. In Figure 4 we also show the impact of the anomalous parameter $`a_0`$ on the transverse momentum of the photon. As anticipated above, it is the hard tail of the photon spectrum that is particularly sensitive to the anomalous contributions and this observable therefore offers a means to search directly for such anomalous contributions. Finally, we have studied the impact on the total cross sections of the processes $`p\overline{p}W^+W^{}\gamma `$ and $`Z\gamma \gamma `$. Figure 5 shows the contour in the $`(a_0,a_c)`$ plane corresponding to the $`+3\sigma `$ deviation of the SM cross section. Just as in the LEP2 study, $`Z\gamma \gamma `$ production promises a better discovery potential, again due to the higher photon energy available. Comparing Figures 3 and 5, we conclude that the Tevatron should be able to set much tighter limits on the anomalous couplings $`a_0`$ and $`a_c`$ than LEP. The basic difference is the significantly higher subprocess scattering energies available at the hadron collider. ## 5 Measurements of anomalous couplings at hadron colliders At $`e^+e^{}`$ colliders anomalous couplings are directly investigated at a fixed centre of mass energy $`\sqrt{s}`$. This results in bounds of the anomalous couplings $`\alpha _{\mathrm{ac}}`$ as a function of $`s`$. At hadron colliders, the center of mass energy of the colliding partons $`\sqrt{\widehat{s}}`$ is not fixed and there will be events where $`\sqrt{\widehat{s}}`$ is very large. In order to avoid problems with violation of unitarity, form factors $`f`$ are introduced, i.e. the anomalous couplings $`\alpha _{\mathrm{ac}}`$ are replaced by $`\alpha _{\mathrm{ac}}f`$. The precise form of $`f`$ as well as the associated scale for new physics $`\mathrm{\Lambda }`$ are to a big extent arbitrary. A common choice is $$f=\left(1+\frac{\widehat{s}}{\mathrm{\Lambda }^2}\right)^n$$ (8) where $`n`$ is chosen big enough to ensure unitarity for $`\widehat{s}\mathrm{}`$. This procedure has the unpleasant consequence that all bounds on the anomalous couplings depend on $`\mathrm{\Lambda }`$ and the precise form of $`f`$. In order to improve the situation it is desirable to get bounds directly on $`\alpha _{\mathrm{ac}}(\sqrt{\widehat{s}})`$ also at hadron colliders. In some cases, this is straightforward to do. As an example we mention $`Z\gamma `$ production which probes the anomalous couplings $`h_3^\gamma ,h_4^\gamma ,h_3^Z`$ and $`h_4^Z`$. In this process $`\widehat{s}`$ can be fully reconstructed. This allows to investigate the anomalous couplings in different regions of $`\widehat{s}`$ and get separate bounds in each region. At the LHC the statistics should be good enough to allow such an analysis. The situation is more difficult in processes, where $`\widehat{s}`$ can not be fully reconstructed, such as $`ppW\gamma \mathrm{}\nu \gamma X`$. In these cases an observable quantity has to be found which has a very strong correlation to $`\widehat{s}`$. Then the analysis could be done using this quantity instead of $`\widehat{s}`$ without introducing a large error. There are several possibilities, such as the transverse mass $`M_T`$ or the cluster mass $`M_C`$. They are defined as follows: $`M_T^2`$ $``$ $`(\sqrt{p_T^2(\mathrm{}\gamma )+m^2(\mathrm{}\gamma )}+|p_T(\nu )|)^2p_T^2(\mathrm{}\gamma \nu )`$ (9) $`M_C^2`$ $``$ $`(\sqrt{p_T^2(\mathrm{}\gamma )+m^2(\mathrm{}\gamma )}+|p_T(\nu )|)^2`$ (10) Note that at leading order $`M_T=M_C`$. Another possibility is to take $`\widehat{s}_{\mathrm{min}}`$ which is defined as follows: assuming the $`W`$ to be on-shell and identifying the missing transverse momentum with $`p_T(\nu )`$ it is possible to reconstruct the full kinematics with a twofold ambiguity. This result in two possible values of $`\widehat{s}`$ and $`\widehat{s}_{\mathrm{min}}`$ is by definition the smaller of the two. In Figure 6 the distribution of the true $`\widehat{s}`$ is shown for two particular bins of the observed quantity. The curves have been obtained for $`pp`$ collision at $`\sqrt{s}`$=14 TeV with some appropriate cuts on the rapidity and transverse momentum of the leptons. For plots (a) and (b) we have $`150\mathrm{GeV}<𝒬<200\mathrm{GeV}`$ whereas for plots (c) and (d) we have $`600\mathrm{GeV}<𝒬<650\mathrm{GeV}`$ where the observed quantity $`𝒬\{M_T,M_C,\widehat{s}_{\mathrm{min}}\}`$. To make the comparison of the various correlations easier, the histogram has been normalized to one in the first bin. The by far strongest correlations are obtained for $`\widehat{s}_{\mathrm{min}}`$. Even if we include unrealistically large anomalous couplings the correlation is preserved, if not enhanced. This can be seen in Figures 6 (b) and (d), where we show the correlations with $`\mathrm{\Delta }\kappa ^\gamma =0.8,\lambda =0.2`$ and the ususal dipole form factor with a scale $`\mathrm{\Lambda }=1`$ TeV. These results have been obtained with a Monte Carlo program including next-to-leading order QCD corrections . The large correlation between $`\widehat{s}`$ and $`\widehat{s}_{\mathrm{min}}`$ should allow for a similar analysis as in the case of $`Z\gamma `$ production by simply replacing $`\widehat{s}`$ by $`\widehat{s}_{\mathrm{min}}`$. ## References
warning/0002/hep-ph0002021.html
ar5iv
text
# 1 Lagrangian with Majorana neutrinos. ## 1 Lagrangian with Majorana neutrinos. During the last few years a lot of attention was paid to the possibility of creating a baryon asymmetry through leptogenesis. The proposed schemes introduce heavy Majorana neutrinos with CP–violating couplings, where both the so–called “direct” and “indirect” contributions to leptogenesis were considered. In calculating the lepton asymmetry, the authors consider the Standard Model (SM) with the usual particle content plus three Majorana neutrinos, which are singlets under the weak SU(2)–group \- . The part of the Lagrangian with Majorana neutrinos consists of the Majorana mass term and the Yukawa interactions of these neutrinos with leptons and Higgs bosons: $$L=\frac{1}{2}\underset{i}{}M_i\overline{N_i}N_i+\underset{\alpha ,i}{}h_{\alpha i}\overline{l_{L\alpha }}\varphi P_RN_i+\underset{\alpha ,i}{}h_{\alpha i}^{}\overline{N_i}P_Ll_{L\alpha }\varphi ^++h.c.$$ (1) In this Lagrangian $`l_{L\alpha }`$ are left-handed lepton doublets of the SM, $`\varphi `$ is the Higgs doublet of the SM, $`\alpha ,i=1,2,3`$ denote the index of fermion generation and $`N_i`$ is the self-conjugate Majorana field: $$N_i=N_{Ri}+(N_{Ri})^c=\left(\begin{array}{c}0\\ N_{Ri}\end{array}\right)+\left(\begin{array}{c}(N_{Ri})^c\\ 0\end{array}\right)=\left(\begin{array}{c}(N_{Ri})^c\\ N_{Ri}\end{array}\right).$$ (2) One should remember here that $`(N_{Ri})^c`$ is a left-handed antiparticle $`(N_{Ri})^c=(N_i^c)_L`$. Another (more explicit) way of writing this Lagrangian is used in : $$\begin{array}{cc}L=\underset{i}{}M_i\left[\overline{(N_{Ri})^c}N_{Ri}+\overline{N_{Ri}}(N_{Ri})^c\right]+\underset{\alpha ,i}{}h_{\alpha i}\overline{l_{L\alpha }}N_{Ri}\varphi +\underset{\alpha ,i}{}h_{\alpha i}^{}\overline{N_{Ri}}l_{L\alpha }\varphi ^+\hfill & \\ +\underset{\alpha ,i}{}h_{\alpha i}\overline{(N_{Ri})^c}(l_{L\alpha })^c\varphi +\underset{\alpha ,i}{}h_{\alpha i}^{}\overline{(l_{L\alpha })^c}(N_{Ri})^c\varphi ^+.\hfill & \end{array}$$ (3) One can easily show that the two Lagrangians (1) and (3) are identical. In fact, by the definition of charge–conjugate fields one can show that $$\overline{(l_{L\alpha })^c}\varphi ^+(N_{Ri})^c=(l_{L\alpha })^TC^1\varphi ^+C\overline{N_{Ri}}{}_{}{}^{T}=(l_{L\alpha })^T\varphi ^+\overline{N_{Ri}}{}_{}{}^{T}=\overline{N_{Ri}}\varphi ^+l_{L\alpha }.$$ (4) Notice that the mass term in (1) violates lepton number by two units and the Yukawa interaction terms violate the CP-symmetry. In general, it is desirable to have a gauge theory with a symmetry responsible for lepton number conservation. In such a theory, the Majorana mass term is generated as the result of spontaneous breaking of lepton number. This theory can be the goal of future investigations; meanwhile, in the model under consideration, the asymmetry is believed to be generated at temperatures bigger than the electroweak symmetry scale, but lower than the scale where the Majorana mass is created. Thus the Lagrangian (1) is an “intermediate–energy” effective Lagrangian. At the high energies considered here, the vacuum expectation value for the Higgs condensate is very small, so that the masses (here we mean vacuum masses and neglect temperature contributions) of the charged leptons $`m_\alpha `$ and Higgs particles $`m_\varphi `$ are negligibly small or zero. ## 2 Indirect contribution to leptogenesis. The Lagrangian (1) was considered several times \- , where the so-called “direct” contribution to leptogenesis was computed. Here we are interested in the “indirect” contribution to leptogenesis, calculated in references and reviewed recently in . The indirect contribution is described by the self-energy diagram in Fig.1. This diagram was investigated in the context of two different approaches. In the first one the self–energy diagram is considered to be the intermediate state of some physical process (for example, lepton–Higgs scattering ). In this case the four-momentum of the Majorana neutrino is off the mass shell and is determined by the four-momenta of the incoming (outgoing) lepton and Higgs boson. We wish to emphasize is that there are no restrictions on the value of neutrino four-momentum. In this approach the lepton asymmetry appears in the scattering process $`l^c\varphi l\varphi ^+`$, through interference of the tree-level diagram with the one-loop diagram . The second approach, presented earlier in , uses the notions of transition $`N_jN_i`$, originating from the one–loop diagram mentioned already. If $`i=j`$ then the diagram is analogous to any self-energy diagram and describes the correction to the fermion (Majorana neutrino in our case) mass. The precise role and calculation of this sort of diagrams is described in textbooks (see, for example, corrections to the electron mass in . A new situation appears for $`ij`$. Now we have a free stable particle with mass $`M_i`$ as incoming particle and it is impossible to transform it into a stable particle with mass $`M_j`$, because the process does not conserve energy and momentum. This transition can happen for unstable particles provided their mass difference is comparable to their widths. In addition the Majorana states are in a thermal bath interacting with the other fields so that energy and momentum is continuously exchanged with the background fields. If the transition probabilities are smooth functions of energies and widths, then it is a good approximation to calculate the transition amplitudes and probabilities per unit time and volume for unstable Majorana neutrinos and then introduce the results into the Boltzmann equations. In the latter step we also introduce the particle densities for initial and final states, as dictated by the thermodynamics of the early universe. A prototype of the thermodynamic calculation is described in ref. , and recent articles . The interaction with the background fields is an integral part in the generation of the lepton asymmetry. In this article we calculate how the transition probabilities and the asymmetry are generated for unstable particles. Before we address this topic, we review how the mixing occurs in a few known cases. The mixing of the $`K^0`$ with $`\overline{K}^0`$ is described by the box–diagram. In this case the $`K^0`$–mesons are treated as composite particles made up of quarks. The mixing diagrams are computed at the quark level, which are the fields of the basic Lagrangian (Standard Model). In the diagrams appear vertices of the form $`\overline{u}sW`$, because the quarks mix in the Lagrangian through the Cabibbo–Kobayashi–Maskawa matrix. Thus the mixings in the above case originate from the mixings at vertices between quarks of different charges and manifest themselves as the mixing of mesonic states through the interaction of their constituent quarks. In the above case energy and momentum is conserved. For the mesonic states the mass difference is comparable to the widths. For K–mesons $$|\mathrm{\Delta }M_K|=\frac{1}{2}|\mathrm{\Delta }\mathrm{\Gamma }_K|\frac{1}{2}\mathrm{\Gamma }_K$$ and for the B–mesons $$\left(\frac{\mathrm{\Delta }M}{\mathrm{\Gamma }}\right)_{B_d}=0.734\pm 0.035.$$ The neutrinos have charged current couplings to the leptons where the lepton number is conserved to a large degree of accuracy. The Majorana couplings, on the other hand, may have lepton number violating charged couplings as seen in the Lagrangian of eqs. (1) or (3). The mass matrix of Majorana neutrinos in eq. (1) can be made real and at the classical (tree) level the generations do not mix. However, it was noted in ref. that the one–loop effects can mix the Majorana states, i.e. the states defined at tree level are not the physical ones. To state the result in another way, we find that Majorana neutrinos of different generation mix, while the electron and the muon do not mix. In this paper we show that the instability of the Majorana neutrinos is of principal importance for the mixing phenomenon. The neutrinos have finite widths and can be treated, through the uncertainty principle, as “fuzzy states” on their mass shell. This allows us to define the mixing of physical states in the case when the difference of the tree–level neutrino masses is less or about equal to their widths: $$|M_iM_j|<max\{\mathrm{\Gamma }_i,\mathrm{\Gamma }_j\}.$$ (5) Condition (5) can evidently be applied to any other particles. For electron and muon, however, $`m_\mu m_e\mathrm{\Gamma }_\mu `$ and consequently the mixing is very small. ## 3 S-matrix for the mixing of unstable particles. The general S-matrix theory necessarily uses the notion of asymptotic states. This means that states of incoming, $`|in(t\mathrm{})`$, and outgoing, $`|out(t+\mathrm{})`$, particles are defined as states of free (noninteracting) particles at infinite times. The second principal element of the $`S`$matrix theory presupposes that interactions do not exist at all “infinite” times, but are “turned on adiabatically” from $`t\mathrm{}`$ to $`t=0`$ and also “turn off” adiabatically from $`t=0`$ to $`t+\mathrm{}`$. One, of course, understands that “infinite” here means macroscopically large in comparison to the time of interaction. This general argument can naturally be applied to the case of unstable particles. In this case the time of “turning on” and “off” the interaction is easily estimated to be of the order of the inverse particle width. We try to express this in a phenomenological definition (the idea is similar to that of Breit and Wigner when considering the cross section at resonance) of field operators for unstable particles. For neutrinos we introduce $$N_i(x)=\underset{\stackrel{}{k},\lambda }{}\left[u^i(\stackrel{}{k},\lambda )a_{\stackrel{}{k},\lambda }^ie^{i\sqrt{\stackrel{}{k}{}_{}{}^{2}+M_i^2}t}e^{i\stackrel{}{k}\stackrel{}{x}}+v^i(\stackrel{}{k},\lambda )a_{\stackrel{}{k},\lambda }^{i+}e^{i\sqrt{\stackrel{}{k}{}_{}{}^{2}+M_i^2}t}e^{i\stackrel{}{k}\stackrel{}{x}}\right]e^{\mathrm{\Gamma }_i|t|}$$ (6) The indices $`i=1,2`$ denote the generation of the neutrino and for the sake of simplicity we consider only two generations. $`a_{\stackrel{}{k},\lambda }^{i+}`$ and $`a_{\stackrel{}{k},\lambda }^i`$ are the fermionic creation and annihilation operators. Formula (6) reflects the fact that neutrinos “disappear” (decay) for infinite time ($`N_i0`$ at $`t\pm \mathrm{}`$). We assume that the widths are generated by the Lagrangian in eq. (1) and calculated by the decay diagrams or from the absorptive part of the diagonal terms of the self–energy. Eq. (6) is the only step which is not proved from field theory, since field theory does not consider asymptotic states. We think, however, that the physical meaning of eq. (6) is rather clear and will help us understand the mixing of Majorana neutrinos. In this paper we will follow the idea of an earlier paper and will introduce an “effective Hamiltonian” with non-diagonal terms in the mass matrix. Initially they are calculated as elements of the S-matrix to second order in the Yukawa–couplings by using perturbation theory. Let $`i`$ and $`j`$ be neutrinos of definite flavors, then $$S_{ij}=\frac{1}{2}d^4xd^4x^{}i|L_{int}(x)L_{int}(x^{})|j,$$ (7) with the neutrino fields defined by eq. (6). ## 4 Loop calculation with unstable Majorana neutrinos. As mentioned in the previous section, we shall use perturbation theory. To second order in the Yukawa couplings the $`S`$matrix element, corresponding to the one-loop diagram, is given by (7). ¿From the product of the interaction Lagrangians $$\begin{array}{c}i|\{h_{\alpha m}\overline{l_{L\alpha }}(x)\varphi (x)P_RN_m(x)+h_{\alpha m}^{}\overline{N_m}(x)P_L\varphi ^+(x)l_{L\alpha }(x)\}\times \hfill \\ \times \left\{h_{\alpha n}\overline{l_{L\alpha }}(x^{})\varphi (x^{})P_RN_n(x^{})+h_{\alpha n}^{}\overline{N_n}(x^{})P_L\varphi ^+(x^{})l_{L\alpha }(x^{})\right\}|j\hfill \end{array}$$ two terms contribute to (7): $$\begin{array}{c}2h_{\alpha i}^{}h_{\alpha j}\overline{N_i}(x)P_L\varphi ^+(x)l_{L\alpha }(x)\overline{l_{L\alpha }}(x^{})\varphi (x^{})P_RN_j(x^{})\hfill \\ +2h_{\alpha i}h_{\alpha j}^{}P_RN_i(x)\varphi (x)\overline{l_{L\alpha }}(x)\overline{N_j}(x^{})P_L\varphi ^+(x^{})l_{L\alpha }(x^{}).\hfill \end{array}$$ (8) The new aspect of our calculation is the form of the field operators for unstable neutrinos given in eq. (6). We calculate in detail the first term, using Wick’s expansion for the product of operators $$\begin{array}{c}S_{ij}=\frac{1}{2}d^4xd^4x^{}2h_{\alpha i}^{}h_{\alpha j}\overline{u}_{p\sigma }^iP_Le^{i\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_i^2}t\mathrm{\Gamma }_i|t|i\stackrel{}{p}\stackrel{}{x}}\frac{d^4q}{(2\pi )^4}\frac{e^{iq(xx^{})}}{\mathit{}m_\alpha +i\epsilon }\hfill \\ \frac{d^4k}{(2\pi )^4}e^{ik(xx^{})}\left(\frac{1}{k^2m_\varphi ^2+i\epsilon }\frac{1}{k^2\mathrm{\Lambda }^2+i\epsilon }\right)e^{i\sqrt{\stackrel{}{p}{}_{}{}^{}{}_{}{}^{2}+M_j^2}t^{}\mathrm{\Gamma }_j|t^{}|+i\stackrel{}{p}{}_{}{}^{}\stackrel{}{x}_{}^{}}P_Ru_{p^{}\sigma ^{}}^j,\hfill \end{array}$$ (9) where $`\mathrm{\Lambda }`$ is a parameter for the Pauli–Villars regularization.<sup>1</sup><sup>1</sup>1Another regularization is equally possible. After integration over $`d^3x`$, $`d^3k`$, $`d^3x^{}`$ one obtains: $$\begin{array}{c}S_{ij}=\frac{1}{2}𝑑t𝑑t^{}2h_{\alpha i}^{}h_{\alpha j}\overline{u}_{p\sigma }^iP_L\frac{d^4q}{(2\pi )^4}\frac{\mathit{}+m_\alpha }{q^2m_\alpha ^2+i\epsilon }\frac{dk^0}{2\pi }\frac{1}{k^0{}_{}{}^{2}(\stackrel{}{p}\stackrel{}{q})^2m_\varphi ^2+i\epsilon }\hfill \\ e^{i(\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_i^2}k^0q^0)t\mathrm{\Gamma }_i|t|}e^{i(\sqrt{\stackrel{}{p}{}_{}{}^{}{}_{}{}^{2}+M_j^2}k^0q^0)t^{}\mathrm{\Gamma }_j|t^{}|}\delta (\stackrel{}{p}\stackrel{}{p}^{})(2\pi )^3P_Ru_{p^{}\sigma ^{}}^j.\hfill \end{array}$$ (10) with a similar expression for the terms with the Pauli-Villars regularization parameter. An important feature of this formula is the product of the two integrals $$J_1=𝑑te^{i(\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_i^2}k^0q^0)t\mathrm{\Gamma }_i|t|}J_2=𝑑t^{}e^{i(\sqrt{\stackrel{}{p}^{}{}_{}{}^{2}+M_j^2}k^0q^0)t^{}\mathrm{\Gamma }_j|t^{}|}.$$ (11) In the case of vanishing widths $`\mathrm{\Gamma }_i=\mathrm{\Gamma }_j=0`$ each of the integrals is equal to a $`\delta `$function and their product $`\delta (\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_i^2}k^0q^0)\delta (\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_j^2}k^0q^0)`$ is zero except when $`M_i=M_j`$ (with $`\stackrel{}{p}=\stackrel{}{p}^{}`$). This result is a mathematical demonstration of physical arguments, discussed in section 2. On the other hand, for non–zero widths the integrals $`J_1`$ and $`J_2`$ lead to “bell-shaped” Lorentzian functions: $$J_1J_2=(2\pi )^2\frac{\mathrm{\Gamma }_i/\pi }{\mathrm{\Gamma }_i^2+(\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_i^2}k^0q^0)^2}\frac{\mathrm{\Gamma }_j/\pi }{\mathrm{\Gamma }_j^2+(\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_j^2}k^0q^0)^2}.$$ (12) This product is non–zero and large when condition (5) is satisfied; its non–zero value can be graphically understood as the “overlap” of two “bells”. So the limit $`|M_iM_j|\mathrm{\Gamma }_{i(j)}`$, which is often used in the calculation of asymmetry, leads to very small overlap functions. Another remark concerns the three-dimensional $`\delta `$function with zero argument $`(2\pi )^3\delta (\stackrel{}{p}\stackrel{}{p}^{})=(2\pi )^3\delta (\stackrel{}{0})`$, which appears in eq. (10). We will keep it throughout the section and it corresponds to the volume element where the interaction takes place. When we substitute (12) into (10), and in addition integrate over $`k^0`$, we obtain the first order expression for $`\mathrm{\Gamma }_{i(j)}`$: $$\begin{array}{c}S_{ij}=h_{\alpha i}^{}h_{\alpha j}2\pi \delta (\stackrel{}{p}\stackrel{}{p})(2\pi )^3\overline{u}_{p\sigma }^iP_L\frac{d^4q}{(2\pi )^4}\frac{\mathit{}+m_\alpha }{q^2m_\alpha ^2+i\epsilon }P_Ru_{p\sigma ^{}}^j\hfill \\ \{\frac{\mathrm{\Gamma }_j}{\pi }\frac{1}{(\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_i^2}\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_j^2})^2+\mathrm{\Gamma }_j^2}\times \hfill \\ \times \left(\frac{1}{(\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_i^2}q^0)^2(\stackrel{}{p}\stackrel{}{q})^2m_\varphi ^2}\frac{1}{(\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_i^2}q^0)^2(\stackrel{}{p}\stackrel{}{q})^2\mathrm{\Lambda }^2}\right)\hfill \\ +\frac{\mathrm{\Gamma }_i}{\pi }\frac{1}{(\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_j^2}\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_i^2})^2+\mathrm{\Gamma }_i^2}\times \hfill \\ \times (\frac{1}{(\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_j^2}q^0)^2(\stackrel{}{p}\stackrel{}{q})^2m_\varphi ^2}\frac{1}{(\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_j^2}q^0)^2(\stackrel{}{p}\stackrel{}{q})^2\mathrm{\Lambda }^2})\}\hfill \end{array}$$ (13) This expression, as is easily seen, reproduces the usual expression for the loop integral after one introduces the zero-component of four momentum $`p^0`$: $$\begin{array}{c}S_{ij}=h_{\alpha i}^{}h_{\alpha j}2\pi \delta (\stackrel{}{p}\stackrel{}{p})(2\pi )^3\overline{u}_{p\sigma }^iP_L\frac{d^4q}{(2\pi )^4}\frac{\mathit{}+m_\alpha }{q^2m_\alpha ^2+i\epsilon }P_Ru_{p\sigma ^{}}^j\times \hfill \\ \times \left\{\frac{\mathrm{\Gamma }_j}{\pi }\frac{1}{(p^0\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_j^2})^2+\mathrm{\Gamma }_j^2}\frac{1}{(p^0q^0)^2(\stackrel{}{p}\stackrel{}{q})^2m_\varphi ^2}\right|^{p^0=\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_i^2}}\hfill \\ +\frac{\mathrm{\Gamma }_i}{\pi }\frac{1}{(p^0\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_i^2})^2+\mathrm{\Gamma }_i^2}\frac{1}{(p^0q^0)^2(\stackrel{}{p}\stackrel{}{q})^2m_\varphi ^2}|{}_{}{}^{p^0=\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_j^2}}\}\hfill \end{array}$$ (14) Integrals of this type are standard and the result for the limit $`m_\alpha 0`$, $`m_\varphi 0`$ is well known: $$\begin{array}{c}Int=\underset{m_\alpha ,m_\varphi 0}{lim}\frac{d^4q}{(2\pi )^4}\frac{\mathit{}+m_\alpha }{q^2m_\alpha ^2+i\epsilon }\left(\frac{1}{(p^\mu q^\mu )^2m_\varphi ^2+i\epsilon }\frac{1}{(p^\mu q^\mu )^2\mathrm{\Lambda }^2+i\epsilon }\right)=\\ =i\gamma ^\mu p_\mu \left(g_{dis}\frac{i}{2}g_{abs}\right),g_{dis}=\frac{1}{16\pi ^2}\left(\frac{1}{2}\mathrm{ln}\frac{\mathrm{\Lambda }^2}{p^2}+\frac{3}{4}\right),g_{abs}=\frac{1}{16\pi }\end{array}$$ It is recognized that the dispersive part of this integral ($`g_{dis}`$) can be reabsorbed in the definition of coupling constants, while only the absorptive part ($`g_{abs}`$) survives and has physical consequences; for more details see and references therein. Now the matrix element $`S_{ij}`$ is given by $$\begin{array}{c}S_{ij}=h_{\alpha i}^{}h_{\alpha j}2\pi \delta (\stackrel{}{p}\stackrel{}{p})(2\pi )^3(i)\times \hfill \\ \times \{\overline{u}_{p\sigma }^iP_L[\gamma ^0(\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_i^2})\stackrel{}{\gamma }\stackrel{}{p}]\frac{\mathrm{\Gamma }_j}{\pi }\frac{i/2g_{abs}}{(\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_i^2}\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_j^2})^2+\mathrm{\Gamma }_j^2}P_Ru_{p\sigma ^{}}^j\hfill \\ +\overline{u}_{p\sigma }^iP_L[\gamma ^0(\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_j^2})\stackrel{}{\gamma }\stackrel{}{p}]\frac{\mathrm{\Gamma }_i}{\pi }\frac{i/2g_{abs}}{(\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_j^2}\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_i^2})^2+\mathrm{\Gamma }_i^2}P_Ru_{p\sigma ^{}}^j\}\hfill \end{array}$$ (15) We transform this expression with the help of the Dirac equation $$\begin{array}{c}\overline{u}_{p\sigma }^iP_L\left[\gamma ^0\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_i^2}\stackrel{}{\gamma }\stackrel{}{p}\right]=\overline{u}_{p\sigma }^iP_RM_i,\hfill \\ \left[\gamma ^0\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_j^2}\stackrel{}{\gamma }\stackrel{}{p}\right]P_Ru_{p\sigma ^{}}^j=M_jP_Lu_{p\sigma ^{}}^j,\hfill \\ \overline{v}_{p\sigma ^{}}^jP_L\left[\gamma ^0\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_j^2}\stackrel{}{\gamma }\stackrel{}{p}\right]=\overline{v}_{p\sigma ^{}}^jP_RM_j,\hfill \\ \left[\gamma ^0\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_i^2}\stackrel{}{\gamma }\stackrel{}{p}\right]P_Rv_{p\sigma }^i=M_iP_Lv_{p\sigma }^i.\hfill \end{array}$$ (16) Finally we arrive at $$\begin{array}{c}S_{ij}^{(I)}=h_{\alpha i}^{}h_{\alpha j}2\pi \delta (\stackrel{}{p}\stackrel{}{p})(2\pi )^3(i)\frac{1}{2}\overline{u}_{p\sigma }^i\{M_i\frac{\mathrm{\Gamma }_j}{\pi }\frac{i/2g_{abs}}{(\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_i^2}\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_j^2})^2+\mathrm{\Gamma }_j^2}\hfill \\ +M_j\frac{\mathrm{\Gamma }_i}{\pi }\frac{i/2g_{abs}}{(\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_j^2}\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_i^2})^2+\mathrm{\Gamma }_i^2}\}u_{p\sigma ^{}}^j\hfill \end{array}$$ (17) This is the final expression for the S–matrix element originating from the first term in eq. (8). The second term from eq. (8) leads us to the similar expression with $`v`$spinors $$\begin{array}{c}S_{ij}^{(II)}=h_{\alpha i}h_{\alpha j}^{}2\pi \delta (\stackrel{}{p}\stackrel{}{p})(2\pi )^3(i)\frac{1}{2}\overline{v}_{p\sigma ^{}}^j\{M_i\frac{\mathrm{\Gamma }_j}{\pi }\frac{i/2g_{abs}}{(\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_i^2}\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_j^2})^2+\mathrm{\Gamma }_j^2}\hfill \\ +M_j\frac{\mathrm{\Gamma }_i}{\pi }\frac{i/2g_{abs}}{(\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_j^2}\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_i^2})^2+\mathrm{\Gamma }_i^2}\}v_{p\sigma }^i\hfill \end{array}$$ (18) As mentioned already, $`\delta (\stackrel{}{p}\stackrel{}{p})(2\pi )^3`$ represents the volume element $`V`$. Similarly the time of interaction also appears as multiplicative factor. $$J_{ij}=\underset{\mathrm{}}{\overset{\mathrm{}}{}}𝑑te^{i\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_i^2}t\mathrm{\Gamma }_i|t|}e^{i\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_j^2}t\mathrm{\Gamma }_j|t|}=2\pi \frac{(\mathrm{\Gamma }_i+\mathrm{\Gamma }_j)/\pi }{(\mathrm{\Gamma }_i+\mathrm{\Gamma }_j)^2+(\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_i^2}\sqrt{\stackrel{}{p}{}_{}{}^{2}+M_j^2})^2}.$$ (19) The new feature of this calculation is the presence of the masses and the decay widths in the S–matrix. The new terms define the time intervals when the interaction takes place. To obtain the transition amplitudes per unit volume and unit time we must divide by the factor $`VJ_{ij}`$ where $`V`$ is the volume element and $`J_{ij}`$ the time interval of the interaction. In other words, we shall work with the “$`T`$–Matrix” defined through the equation $$S_{fi}=1+iT_{fi}(2\pi )^4\delta ^4(p_fp_i)$$ (20) where the energy conserving $`\delta `$–function will be substituted by the expression $$(2\pi )\delta (E_fE_i)\frac{2(\mathrm{\Gamma }_i+\mathrm{\Gamma }_j)}{(E_fE_i)^2+(\mathrm{\Gamma }_f+\mathrm{\Gamma }_i)^2}.$$ (21) One easily arrives at $`T_{ij}={\displaystyle \frac{S_{ij}^{(I)}+S_{ij}^{(II)}}{VJ_{ij}}}`$ $`={\displaystyle \frac{(1)}{2}}{\displaystyle \frac{1}{16\pi }}`$ $`(h_{\alpha i}^{}h_{\alpha j}\overline{u}_{p\sigma }^i[M_i{\displaystyle \frac{\mathrm{\Gamma }_j}{\mathrm{\Gamma }_i+\mathrm{\Gamma }_j}}P_R+M_j{\displaystyle \frac{\mathrm{\Gamma }_i}{\mathrm{\Gamma }_i+\mathrm{\Gamma }_j}}P_L]u_{p\sigma ^{}}^j`$ (22) $`+h_{\alpha i}h_{\alpha _j}^{}\overline{v}_{p\sigma ^{}}^j[M_j{\displaystyle \frac{\mathrm{\Gamma }_i}{\mathrm{\Gamma }_i+\mathrm{\Gamma }_j}}P_R+M_i{\displaystyle \frac{\mathrm{\Gamma }_j}{\mathrm{\Gamma }_i+\mathrm{\Gamma }_j}}P_L]v_{p\sigma }^i).`$ Notice the S–matrix element (17), (18) vanish in the case $`|M_iM_j|\mathrm{\Gamma }_{ij}`$ since the Lorentzian representation of the Delta–function (21) tends to zero. It is easy to see that the same occurs for the transition probability $$\frac{w(N_iN_j)}{VJ_{ij}}=|T_{ij}|^2(2\pi )^4\delta (\stackrel{}{p}\stackrel{}{p}^{})\frac{(\mathrm{\Gamma }_i+\mathrm{\Gamma }_j)/\pi }{(E_fE_i)^2+(\mathrm{\Gamma }_f+\mathrm{\Gamma }_i)^2}.$$ Let us summarize our results. We have shown that neutrinos of different generations can mix through one–loop diagrams even if they have different masses. The physical reason of this possibility is related to the final widths of neutrinos : their mass shells are “fuzzy” and energy is conserved with the accuracy comparable to their width. The closer the masses of neutrinos, the larger is the transition $`N_iN_j`$ probability; and for $`|M_iM_j|\mathrm{\Gamma }_{i(j)}`$ the mixing has no physical meaning. As we already know, neutrino mixing provides an indirect contribution to leptogenesis. As it was previously done in , we are working in terms of an effective Hamiltonian, or, what is the same, in terms of an effective mass matrix. From the transition matrix elements and the mass term in eq. (1) we can now proceed to calculate an effective mass matrix and the physical states. We consider the expression for $`T_{ij}`$ as the first order contribution for mass matrix elements. One should notice that the contributions are different for $`R`$ and $`L`$ parts of our 4–spinor $`N_i`$ in eq. (2), i.e. they are different for $`N_{Ri}`$ and $`(N_{Ri})^c`$. Evidently, an overall factor of $`(i)`$ and the spinors originate from the definition of the S–matrix and do not appear on the mass matrix. A factor of 1/2 is also omitted in order to be consistent with the Lagrangian in eq. (1), which has 1/2 in front of the mass term. So the corrections to the $`M_i\overline{(N_{Ri})^c}N_{Ri}`$ mass term (see Lagrangian in the form (3)) are $$H_{ij}=\frac{(i)}{2}\frac{1}{8\pi }\left(h_{\alpha i}^{}h_{\alpha j}\frac{M_i\mathrm{\Gamma }_j}{\mathrm{\Gamma }_i+\mathrm{\Gamma }_j}+h_{\alpha i}h_{\alpha j}^{}\frac{M_j\mathrm{\Gamma }_i}{\mathrm{\Gamma }_i+\mathrm{\Gamma }_j}\right)$$ and the corrections to the $`M_i\overline{N_{Ri}}(N_{Ri})^c`$ mass term are $$\stackrel{~}{H}_{ij}=\frac{(i)}{2}\frac{1}{8\pi }\left(h_{\alpha i}^{}h_{\alpha j}\frac{M_j\mathrm{\Gamma }_i}{\mathrm{\Gamma }_i+\mathrm{\Gamma }_j}+h_{\alpha i}h_{\alpha j}^{}\frac{M_i\mathrm{\Gamma }_j}{\mathrm{\Gamma }_i+\mathrm{\Gamma }_j}\right)$$ When $`\mathrm{\Gamma }_i=\mathrm{\Gamma }_j`$, this result coincides with those obtained in ref. . Our account of finite widths of Majorana neutrinos slightly corrects the result which is not very essential. What is important, is that the definition in eq. (6) enables one to give physical meaning to the mixing of particles with different masses. ## 5 Effective Contribution to the Mass Term and Physical Neutrino States. We shall work in a vector space with four basis vectors ($`N_{R1}^c,N_{R1},N_{R2}^c,N_{R2}`$) which are the states occurring in the interaction term of eq. (1). The transitions among these states introduce an effective matrix $$m=\left(\begin{array}{cccc}0& M_1+H_{11}& 0& H_{12}\\ M_1+H_{11}& 0& \stackrel{~}{H}_{12}& 0\\ 0& H_{12}& 0& M_2+H_{22}\\ \stackrel{~}{H}_{12}& 0& M_2+H_{22}& 0\end{array}\right)$$ (23) with $`H_{ij}`$ $`=`$ $`2\left[h_{\alpha i}^{}h_{\alpha j}{\displaystyle \frac{M_i\mathrm{\Gamma }_j}{\mathrm{\Gamma }_i+\mathrm{\Gamma }_j}}+h_{\alpha i}h_{\alpha j}^{}{\displaystyle \frac{M_j\mathrm{\Gamma }_i}{\mathrm{\Gamma }_i+\mathrm{\Gamma }_j}}\right]\left({\displaystyle \frac{i}{2}}g_{ab}\right)`$ (24) $`\stackrel{~}{H}_{ij}`$ $`=`$ $`H_{ij}^{}\mathrm{and}g_{ab}={\displaystyle \frac{1}{16\pi }}`$ (25) The $`2\times 2`$ matrices in the upper left and lower right corners of eq. (23) are lepton–number violating but flavor conserving, while the $`2\times 2`$ matrices along the off–diagonal are lepton– and flavor–number violating. All terms should be present and we found no way to reduce it to a $`2\times 2`$ matrix. It remains to diagonalize the matrix and find the wavefunctions. It is instructive to present a perturbative solution of the problem and then discuss the exact solution. We split the mass matrix into the dominant term and a perturbation $$m=H_0+\lambda V$$ with $$H_0=\left(\begin{array}{cccc}0& M_1+H_{11}& 0& 0\\ M_1+H_{11}& 0& 0& 0\\ 0& 0& 0& M_2+H_{22}\\ 0& 0& M_2+H_{22}& 0\end{array}\right)$$ (26) and $$\lambda V=\left(\begin{array}{cccc}0& 0& 0& \lambda H_{12}\\ 0& 0& \lambda H_{12}^{}& 0\\ 0& \lambda H_{12}& 0& 0\\ \lambda H_{12}^{}& 0& 0& 0\end{array}\right)$$ (27) We introduce a small parameter $`\lambda `$ in order to keep track of the perturbative corrections and at the very end set $`\lambda =1`$. Applying perturbation theory we find the eigenfunctions to $`0(\lambda ^2)`$ $$U_1=\left(\begin{array}{c}1\\ 1\\ X\\ Y\end{array}\right)\mathrm{and}U_2=\left(\begin{array}{c}X^{}\\ Y^{}\\ 1\\ 1\end{array}\right)$$ (28) with the eigenvalues $`\lambda _1=M_1i\frac{\mathrm{\Gamma }_{11}}{2}`$, $`\lambda _2=M_2i\frac{\mathrm{\Gamma }_{22}}{2}`$, and $`H_{ii}=\frac{i}{2}\mathrm{\Gamma }_{ii}`$. They have the time dependence $`U_ie^{i(M_ii\frac{\mathrm{\Gamma }_{ii}}{2})t}`$, with $`i=1`$ and 2 which is consistent with the time dependence introduced in eq. (6). The physical states are $`\psi _1`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{N}}}\left[|N_{R1}+|N_{R1}^c+X|N_{R2}^c+Y|N_{R2}\right]`$ (29) $`\psi _2`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}\left[|N_{R2}+|N_{R2}^c+X^{}|N_{R1}^c+Y^{}|N_{R1}\right]`$ (30) with $`X=\frac{H_{12}M_1+H_{12}^{}M_2}{M_1^2M_2^2}`$, $`Y=\frac{H_{12}M_2+H_{12}^{}M_1}{M_1^2M_2^2}`$ and similar formulas for $`X^{}`$ and $`Y^{}`$. The results are the same as in ref. with an additional dependence on the widths. The interesting result is that the definition in eq. (6) makes it possible to give a physical meaning to the mixing of particles with different masses. We can finally calculate the asymmetry for the decay of each state. The width for the decay of $`\psi _1`$ into leptons is $$\mathrm{\Gamma }(\psi _1\mathrm{}+\mathrm{})\underset{\alpha }{}|h_{\alpha 1}+h_{\alpha _2}Y|^2$$ (31) and into antileptons $$\mathrm{\Gamma }(\psi _1\overline{\mathrm{}}+\mathrm{})\underset{\alpha }{}|h_{\alpha _1}^{}+h_{\alpha _2}^{}X|^2.$$ (32) The lepton asymmetry, defined as $$\delta _1=\frac{\mathrm{\Gamma }_{\mathrm{\Psi }_1l\varphi ^+}\mathrm{\Gamma }_{\mathrm{\Psi }_1l^c\varphi }}{\mathrm{\Gamma }_{\mathrm{\Psi }_1l\varphi ^+}+\mathrm{\Gamma }_{\mathrm{\Psi }_1l^c\varphi }},$$ (33) and it is straightforward to calculate it $$\delta _1=\frac{1}{8\pi }\frac{M_1M_2}{M_2^2M_1^2}\frac{Im(h_{\alpha 1}^{}h_{\alpha 2})^2}{|h_{\alpha 1}|^2+|h_{\alpha 2}X|^2+Reh_{\alpha 1}^{}h_{\alpha 2}(Y+X^{})}.$$ (34) In the above calculation we considered the case $`|M_2M_1|H_{12}`$. In the case that the two Majorana neutrinos are nearly degenerate we must diagonalize exactly the matrix in eq. (23) and then we recover the resonance phenomenon introduced in eq. (17) of ref. . Our study so far considered the neutrinos as free particles. In reality, however, they are in a backgound of fields interacting many times with the other particles. These interactions at a finite temperature can modify their masses and widths . In the present analysis we have found that within the range of validity of eqs. (18) and (19) the $`T`$–matrix in eq. (22) is a smooth function of masses and the widths, and for this reason we can introduce the matrix elements in the Boltzmann equations in order to study the development of the asymmetry. This approach was followed in ref. . ## Summary We have shown that finite widths of Majorana neutrinos play a principal role in producing neutrino mixing via self–energy diagrams. In our approach the widths of neutrinos are treated as “fuzzy” states on their mass shell and the mixing is understood as an “overlap” of two fuzzy mass shell states. For the mathematical realization of these ideas we changed phenomenologically the definition of asymptotic neutrino states, including their widths as in eq. (6). Concerning the results, we showed that these changes do not influence significantly the expressions for the effective mass matrix and for the asymmetry generated in the case of small mass difference. For large mass differences the asymmetry is suppressed. ## Acknowledgement One of us (O.L.) expresses her thanks to the members of the chair Theoretical Physics III of the University of Dortmund for their hospitality, where the largest part of this work was completed, and to Prof. G. Vereshkov for useful discussions. This work was supported in part by the Bundesministerium für Bildung, Wissenschaft, Forschung und Technologie, Bonn (05 HT9PEB 8), and in part by NATO Collaborative Research Grant No. 97 1470.
warning/0002/cond-mat0002333.html
ar5iv
text
# Gap inhomogeneities and the density of states in disordered d-wave superconductors \[ ## Abstract We report on a numerical study of disorder effects in 2D $`d`$-wave BCS superconductors. We compare exact numerical solutions of the Bogoliubov-deGennes (BdG) equations for the density of states $`\rho (E)`$ with the standard T-matrix approximation. Local suppression of the order parameter near impurity sites, which occurs in self-consistent solutions of the BdG equations, leads to apparent power law behavior $`\rho (E)|E|^\alpha `$ with non-universal $`\alpha `$ over an energy scale comparable to the single impurity resonance energy $`\mathrm{\Omega }_0`$. We show that the novel effects arise from static spatial correlations between the order parameter and the impurity distribution. \] In spite of strong electronic correlations in the normal state, the superconducting state of high $`T_c`$ materials seems to be accurately described by a conventional BCS-like phenomenology. The debate over the $`k`$-space structure of the BCS order parameter $`\mathrm{\Delta }_𝐤`$ has now been resolved in favour of pairing states with $`d`$-wave symmetry. Since this symmmetry implies that $`\mathrm{\Delta }_𝐤`$ changes sign under rotation by $`\pi /2`$, there are necessarily points on the 2D Fermi surface at which $`\mathrm{\Delta }_𝐤`$ vanishes. The unique low-energy properties of high $`T_c`$ superconductors are determined by the quasiparticle excitations in the vicinity of these nodal points. For conventional $`s`$-wave superconductors, the density of states (DOS) $`\rho (E)`$ has a well defined gap and is largely unaffected by non-magnetic disorder. In contrast, $`\rho (E)|E|`$ in clean d-wave superconductors, and can be substantially altered by disorder. Much of the current understanding of disorder effects comes from perturbative theories, such as the widely-used self-consistent T-matrix approximation (SCTMA). In particular, the SCTMA is exact in the limit of a single impurity, and has been used in studies of the local DOS near an isolated scatterer. For sufficiently strong scatterers, an isolated impurity introduces a pair of resonances at energies $`\pm \mathrm{\Omega }_0`$, where $`\mathrm{\Omega }_0<\mathrm{\Delta }`$ is a function of of the impurity potential $`u_0`$ and of the band asymmetry. Analytic expressions for $`\mathrm{\Omega }_0(u_0)`$ have been given for a symmetric band, and in this special instance the unitary limit $`\mathrm{\Omega }_00`$ coincides with $`u_0\mathrm{}`$. For a realistic (asymmetric) band, the relationship is more complex. For a finite concentration of impurities $`n_i`$, the SCTMA predicts that the impurity resonances broaden, with tails which overlap at the Fermi energy, leading to a finite residual DOS $`\rho (0)`$. The region over which $`\rho (E)\rho (0)`$, the “impurity band”, and has a width comparable to the scattering rate $`\gamma `$ at $`E=0`$. In the Born limit, the $`\pm \mathrm{\Omega }_0`$ resonances are widely separated in energy, and the overlap of their tails is exponentially small. In the strong-scattering limit, however, the overlap is substantial and $`\gamma \sqrt{\mathrm{\Gamma }\mathrm{\Delta }_d}`$, where $`\mathrm{\Delta }_d`$ is the magnitude of the d-wave gap, $`\mathrm{\Gamma }=n_i/\pi N_0`$ is the scattering rate in the normal state, and $`N_0`$ is the 2D normal-state DOS at the Fermi level. Several recent experiments have studied quasiparticles in the impurity band in Zn-doped high $`T_c`$ materials. Of particular note are attempts to verify a provocative prediction that transport coefficients on energy scales $`\omega ,T<\gamma `$ have a universal value, independent of $`\mathrm{\Gamma }`$. The SCTMA which forms the basis of this world-view has several limitations. It is an effective medium theory, in which one solves for the eigenstates of an isolated impurity in the presence of a homogeneous mean-field representing all other impurities. This approach ignores multiple impurity scattering processes which are responsible for localization physics in metals and may lead to novel effects in 2D d-wave superconductors. Another limitation of most SCTMA calculations is the use of a $`\delta `$-function potential as an impurity model. While this simplifies the calculation substantially, numerical results hint that the detailed structure of the impurity potential may be important. A related issue, which will be discussed at length in this Letter, is inhomogeneous order-parameter suppression. It is well-known that d-wave superconductivity is destroyed locally near a strong scatterer, and in the single-impurity limit, the additional scattering was found to renormalise $`\mathrm{\Omega }_0`$ but, surprisingly, to leave most other details of the scattered eigenstates unchanged. Here we show, using exact numerical solutions of the Bogoliubov-deGennes (BdG) equations, that additional novel physics does arise in the many-impurity case. The main results of this work are summarised in Fig. 1. For a homogeneous order parameter, $`\rho (E)`$ saturates at a constant value as $`E0`$ (in agreement with SCTMA), down to a mesoscopic energy scale $`1/\rho L^2`$ where level repulsion across the Fermi surface induces a gap. This small gap may be a precursor to a regime associated with strong localization in the thermodynamic limit as discussed by Senthil et al.. These authors predicted asymptotic power-law behavior in $`\rho (E)`$ over an exponentially small energy scale $`E_21/\rho (0)\xi _L^2`$, where $`\rho (0)`$ is the residual density of states in the impurity band plateau and $`\xi _L(v_F/\gamma )\mathrm{exp}(v_F/v_\mathrm{\Delta }+v_\mathrm{\Delta }/v_F)`$ is the quasiparticle localization length. ($`v_F`$ is the Fermi velocity and $`v_\mathrm{\Delta }`$ is the gradient of $`\mathrm{\Delta }_k`$ along the Fermi surface at the gap node). The actual value of the DOS at $`E=0`$ is consistent with zero but not determined in our work; theoretically this point is still controversial. Figure 1 shows that when the order parameter is determined self-consistently from the BdG equations $`\rho (E)`$ is quite different. At low energies the DOS can be fit to a power law $`\rho (E)|E|^\alpha `$ with nonuniversal $`\alpha `$ (Fig. 2). The power-law is the result of spatial correlations between the order-parameter and the impurity potential, and is therefore fundamentally different from those of Nersesyan et al. and Senthil et al, where asymptotic power-laws were found, with $`\alpha =1/7`$ and $`\alpha =1`$ respectively. Unlike the DOS, the dimensionless conductance and inverse participation ratio (Fig. 3) are not changed significantly by self-consistency. Finally, we remark that the energy scale for the low-energy regime is > Ω0 > subscriptΩ0\vbox{\hbox{\lower 10.79993pt\hbox{$>$}} \break\hbox{\lower 2.39996pt\hbox{$\sim$}} }\Omega_{0}, which is orders of magnitude larger than $`E_2`$ for realistic parameters. We employ a one-band lattice model with nearest neighbour hopping amplitude $`t`$ and a nearest neighbour attractive interaction $`V`$. Substitutional impurities are represented by a change in the on-site atomic energy. The Hamiltonian is $``$ $`=`$ $`t{\displaystyle \underset{i,j}{}}{\displaystyle \underset{\sigma }{}}c_{i\sigma }^{}c_{j\sigma }{\displaystyle \underset{i,\sigma }{}}[\mu U_i]c_{i\sigma }^{}c_{i\sigma }`$ (1) $``$ $`{\displaystyle \underset{i,j}{}}\{\mathrm{\Delta }_{ij}c_i^{}c_j^{}+h.c.\},`$ (2) where the angle-brackets indicate that site indices $`i`$ and $`j`$ are nearest neighbours, $`U_i`$ is the impurity potential which takes the value $`u_0`$ at a fraction $`n_i`$ of the sites and is zero elsewhere, and $`\mathrm{\Delta }_{ij}=Vc_jc_i`$ is the mean-field order parameter, determined self-consistently by diagonalizing Eq. (2). Throughout this work, energies are measured in units of $`t`$, where $`t`$ is of order 100 meV for high $`T_c`$ materials. In non-self-consistent calculations the OP has the familiar $`k`$-space form $`\mathrm{\Delta }_k=\mathrm{\Delta }_d[\mathrm{cos}(k_x)\mathrm{cos}(k_y)]`$, where $`\mathrm{\Delta }_d=\frac{1}{2}_\pm (\mathrm{\Delta }_{ii\pm x}\mathrm{\Delta }_{ii\pm y})`$ is independent of $`i`$. Unless otherwise stated, $`V=2.3`$ and $`\mu =1.2`$ which yields $`\mathrm{\Delta }_d=0.4`$ (corresponding to $`v_F/v_\mathrm{\Delta }5`$) in the absence of disorder. Self-consistent solutions show that $`\mathrm{\Delta }_{ij}`$ is suppressed within a few lattice constants of each strongly-scattering impurity. Throughout this Letter, curves marked BdG and BdG+OP refer to the neglect or inclusion of self-consistency in $`\mathrm{\Delta }_{ij}`$. The DOS is $`\rho (E)=L^2_\alpha \delta (EE_\alpha )`$, where $`L`$ is the linear system size and $`E_\alpha `$ are the discrete eigenenergies of $``$. Our numerical calculations were performed on periodically continued systems with $`L45`$. Typical DOS curves were obtained for $`L=25`$ by averaging $`\rho (E)`$ over $`50500`$ impurity configurations and $`50100`$ $`k`$-vectors in the supercell Brillouin-zone. For system sizes $`L35`$, computational constraints restricted us to real periodic and anti-periodic boundary conditions. An important motivation for the present study is the need for a test of the reliability of SCTMA predictions. Thus, we use the same disordered lattice model for the SCTMA as for the BdG calculations. We would also like to model order-parameter suppression within a self-consistent T-matrix approximation (SCTMA+OP), and follow the ansatz of that the off-diagonal potential is $`\delta \mathrm{\Delta }_{ij}=\mathrm{\Delta }_{ij}[\delta _{i,0}+\delta _{j,0}]`$. This term appears in the off-diagonal block of the effective potential. In both cases, the T-matrix is a $`2\times 2`$ matrix in particle-hole space which satisfies $$T_{ij}(E)=U_{ij}+\underset{R,R^{}}{}U_{iR}G(RR^{},E)T_{R^{}j}(E),$$ (4) where $`G`$ has the Fourier transform, $$G_k(E)=\left[E\tau _0ϵ_k\tau _3\mathrm{\Delta }_k\tau _1\mathrm{\Sigma }_k(E)\right]^1,$$ (5) $`ϵ_k`$ is the tight-binding dispersion, $`\tau _i`$ are the Pauli matrices, and $`\mathrm{\Sigma }_k(E)=n_iT_{kk}(E)`$. Equation (4) is solved in real-space to take advantage of the short range of the effective potential $`U`$. Finally, $`\rho (E)=\pi ^1L^2\text{Im }_k\text{Tr }G_k(E)`$, where the trace is over particle-hole indices. Except for the mesoscopic gap discussed previously, the BdG curve in Fig. 1 is quantitatively similar to the SCTMA (inset). In contrast, the BdG+OP curve vanishes smoothly as $`E0`$, indicating that qualitatively new physics has been introduced by the inclusion of order parameter suppression. To emphasize that this result is unexpected, we point to the popular “Swiss cheese” model, in which it is assumed that pair-breaking causes a pocket of normal metal of radius $`\xi _0`$ (the coherence length) to form around each impurity. Within this model, $`\rho (0)`$ is enhanced relative to the SCTMA by $`n_iN_0\xi _0^2`$. We emphasize that the correct spatial correlations between an impurity configuration and its self-consistent off-diagonal potentials must be preserved for the BdG+OP results to arise. This point is illustrated with a simple numerical calculation (inset of Fig. 1), in which $`\mathrm{\Delta }_{ij}`$ is found self-consistently for random impurity distributions which are different from those appearing in the diagonal block of $``$. The system therefore has two distinct types of impurity, one of which is purely off-diagonal, with uncorrelated distributions. It is striking that there is no hint of the correct low-energy behaviour in this calculation. It is instructive to compare the BdG+OP result with the SCTMA+OP (inset, Fig. 1), since they include off-diagonal scattering from the order parameter at different levels of approximation. Furthermore, the SCTMA+OP does preserve the correlation between impurity location and order-parameter suppression. It has been used succesfully to describe shifts in $`\mathrm{\Omega }_0`$ due to off-diagonal scattering but, here, fails to reproduce the correct low energy DOS in the bulk disordered case. Instead, $`\rho (E)`$ is quantitatively similar to the SCTMA, which is a direct result of the relative smallness of the off-diagonal potential $`\mathrm{\Delta }_d/u_00.04`$. In Fig. 2 we illustrate the dependence of the low energy DOS on both $`n_i`$ and $`u_0`$. In Fig. 2(a) a series of curves shows how the low-energy regime scales towards zero-width as $`n_i0`$ for the near-unitary scattering potential $`u_0=5`$. That the size of the regime should scale faster than $`\gamma `$ with $`n_i`$ is consistent with our earlier assertion that the novel behavior stems from a multiple-impurity effect. The details of the dilute impurity limit depend on the particular value of $`u_0`$ however: at $`2\%`$ impurity concentration, the low-energy regime is significantly larger for $`u_0=10`$, which lies farther from unitarity, than for $`u_0=5`$. For fixed $`n_i`$, we find that the low-energy regime is > Ω0 > subscriptΩ0\vbox{\hbox{\lower 10.79993pt\hbox{$>$}} \break\hbox{\lower 2.39996pt\hbox{$\sim$}} }\Omega_{0}. In Fig. 2(b), a logarithmic plot of $`\rho (E)`$ reveals that the low energy DOS has an apparent power-law dependence on $`E`$, with non-universal exponent $`0<\alpha <1`$ \[Fig. 2(c),(d)\]. At low impurity concentrations, $`\alpha `$ is a strong function of both $`n_i`$ and $`u_0`$, with $`\alpha `$ a minimum for unitary scatterers. For larger $`n_i`$, $`\alpha `$ appears to saturate at a value which is independent of $`u_0`$. We assert that this power law is fundamentally different from those reported elsewhere , since it is only observed when off-diagonal scattering is present and, as we will see next, is unrelated to strong quasiparticle localization. We have studied the scaling of both the inverse participation ratio $`a(E)`$ and the Thouless number $`g(E)`$ with system size. The inverse participation ratio is defined in the usual way, $`a(E)=x_4(E)/x_2(E)^2`$, where $`x_m(E)=[\rho (E)L^2]^1_{n,r}[|u_n(r)|^m+|v_n(r)|^m]\delta (EE_n)`$ and $`u_n(r)`$ and $`v_n(r)`$ are the particle and hole eigenfunctions, and is plotted in Fig. 3 for several system sizes. $`a(E)`$ is a direct measure of the spatial extent of the wavefunction; it scales as $`1/L^2`$ for extended states and is constant for states with localization length $`\xi _L<L`$. The scaling and energy dependence of $`a(E)`$ is in general agreement with ref. where calculations were made with similar parameters—there is a crossover in scaling between $`E<\gamma `$ and $`E>\gamma `$ consistent with a shorter localization length in the impurity band. Unlike ref. , we do not find saturation in $`a_<`$ \[the average of $`a(E<\gamma )`$\] for $`L45`$. For larger disorder concentrations (not shown) where $`\xi _L<45`$ sites, we find that $`a(E)`$ saturates at $`E=0`$ first indicating that $`\xi _L(E)`$ is an increasing function of $`E`$ in the impurity band. The most important result for this work is that $`a(E)`$, and therefore $`\xi _L`$, is not significantly different in the BdG and BdG+OP calculations despite a substantial difference in $`\rho (E)`$. In the calculation which is shown, $`a(E)`$ is actually decreased slightly by self-consistency, corresponding to a slight increase in the localisation length. We have also studied the Thouless number, defined as $`g(E)=_n[E_n(\pi )E_n(0)]\delta (EE_n)`$, where the argument of $`E_n`$ refers to the application of periodic or anti-periodic boundary conditions in the $`x`$-direction. As $`L`$ is increased, $`g(E)`$ is expected to cross over from a constant in the diffusive regime to exponential scaling indicative of strong localization. For $`L45`$ we find no significant scaling of $`g(E)`$ with $`L`$, consistent with what is found for $`a(E)`$. Most significantly, we find that $`g(E)`$ is nearly identical in the BdG and BdG+OP calculations, even for low-energy states where substantial changes in $`\rho (E)`$ occur. This behaviour is reminiscent of what is seen in Hartree-Fock studies of interacting electrons in disordered conductors. There, the Coulomb interaction enforces spatial correlations between the disorder and charge distributions and leads to the formation of a gap in $`\rho (E)`$, yet leaves the dimensionless conductance (a two-particle property related to the Thouless number) unchanged. The power-law DOS we observe here may have a similar origin; it is certainly clear that $`\rho (E)`$ depends crucially on the spatial correlations between the impurity potential and d-wave order parameter. In the current work, however, it is the pairing interaction which is relevant and the BdG equations provide a mean-field description of the pairing interaction which is analogous to the Hartree-Fock description of the Coulomb interaction. We speculate that the short-ranged pairing interaction produces spatial correlations between distant impurities via the overlap of the long range tails of the single impurity resonances. In this work we have shown that spatial correlations between order parameter and impurity distributions in d-wave superconductors lead to apparent power-laws in $`\rho (E)`$ at low energies. These results are potentially relevant to quasi-2D superconductors like BSCCO-2212. Unfortunately, most disorder studies have been performed on the anisotropic 3D YBCO system, where correlation effects are expected to be less pronounced. Indeed there is considerable evidence, particularly in Zn-substituted YBCO, that disorder does indeed induce a finite DOS at the Fermi level and somewhat weaker evidence that it scales with disorder in accordance with the SCTMA. Our work should therefore provide a strong motivation to study Zn doping, and other types of planar disorder, in the quasi-2D BSCCO-2212 system at low temperatures. This work is supported by NSF grants NSF-DMR-9974396 and NSF-DMR9714055. The authors would like to thank M.P.A. Fisher, K. Muttalib, S. Vishveshwara and K. Ziegler for helpful discussions.
warning/0002/hep-ph0002279.html
ar5iv
text
# Acknowledgments ## Acknowledgments The author would like to thank Y. Okada, J. Hisano and A. Akeroyd for reading manuscript and useful comments. He also thanks K. Okumura and S. Kiyoura for discussions.
warning/0002/math-ph0002024.html
ar5iv
text
# Algebraic Investigation of the Soft Seven Sphere ## 1 Octonions and the Soft Seven Sphere We start by recalling the non-associative octonion algebra. A generic octonion number is $$\phi =\phi _0e_0+\phi _ie_i=\phi _\mu e_\mu .[i=\mathrm{1..7},\mu =\mathrm{0..7},\phi _\mu ]$$ (1) such that $`e_0=1`$ and the other seven imaginary units satisfy $`e_ie_j=\delta _{ij}+f_{ijk}e_k[e_i,e_j]=2f_{ijk}e_k`$ where $`f_{ijk}`$ is completely antisymmetric and equals one for any of the following three-cycles (123), (145), (246), (347), (176), (257), (365) and the associator $$[e_i,e_j,e_k]=(e_ie_j)e_ke_i(e_je_k),$$ (2) is non-zero for any three elements that are not in the same three cycles and is completely antisymmetric The following formula or any of its generalization is thus ambiguous $$e_1e_5e_7=\{\begin{array}{c}(e_1e_5)e_7=e_3\\ \text{or}\\ e_1(e_5e_7)=e_3\end{array}$$ (3) so the best way is to define the action of the imaginary units in a certain direction. In the spirit of Englert, Sevrin, Troost, Van Proeyen and Spindel (also look at and ) we define the left action of octonionic operators $`\delta _i`$ by $$\delta _i\phi =(e_i\phi )$$ (4) implying that the following equation is well defined $$\delta _i\delta _j\phi =\delta _i(\delta _j\phi )=\delta _i(e_j\phi )=e_i(e_j\phi ),$$ (5) then eq. (3) reads unambiguously as $$\delta _1\delta _5e_7=(e_1(e_5e_7))=e_3.$$ (6) Since octonions are also non-commutative, we must also differentiate between left and right action. Using the barred notation , we introduce right action as $$1|\delta _i\phi =(\phi e_i),$$ (7) for example $$(1|\delta _i)(1|\delta _j)\phi =(1|\delta _i)(1|\delta _j\phi )=1|\delta _i(\phi e_j)=((\phi e_j)e_i).$$ (8) As we shall see shortly, we can express the associator in terms of left and right operators. The imaginary octonionic units generate the seven sphere $`S^7`$which has many properties similar to Lie algebras and/or Lie groups. $`S^7`$and Lie groups are the only non-flat compact parallelizable manifolds . The important point for evaluating any Lie algebra is the commutator, so let’s examine $`[\delta _i,\delta _j]\phi `$ $`=`$ $`e_i(e_j\phi )e_j(e_i\phi )=2f_{ijk}e_k\phi 2[e_i,e_j,\phi ]`$ (9) $`=`$ $`2f_{ijk}e_k\phi +2[e_i,\phi ,e_j].`$ now the last term can be written as $`[e_i,\phi ,e_j]`$ $`=`$ $`(e_i\phi )e_je_i(\phi e_j)=\delta _i(1|\delta _j)\phi +(1|\delta _j)\delta _i\phi `$ (10) $`=`$ $`[\delta _i,1|\delta _j]\phi ,`$ thus our commutator can be rewritten as $$[\delta _i,\delta _j]\phi =2f_{ijk}e_k\phi +2[e_i,\phi ,e_j]=2f_{ijk}\delta _k\phi 2[\delta _i,1|\delta _j]\phi .$$ (11) Note that right operators are necessary because the last term the associator can never be written in terms of left operators alone. After simple calculations, one concludes that the octonionic imaginary units are determined completely by (11) and the following equations $`[1|\delta _i,1|\delta _j]\phi `$ $`=`$ $`2f_{ijk}1|\delta _k\phi 2[\delta _i,1|\delta _j]\phi `$ (12) $`\{\delta _i,\delta _j\}\phi `$ $`=`$ $`2\delta _{ij}\phi `$ (13) $`\{1|\delta _i,1|\delta _j\}\phi `$ $`=`$ $`2\delta _{ij}\phi ,`$ (14) where the $`\delta _{ij}`$ in (13) and (14) are the standard Kronecker delta tensor. It has been proved in , using three different ways, that the $`\delta _i`$ algebra is associative. Thus a representation theory in terms of matrices should be possible. Indeed, in , we have derived an algebra completely isomorphic to (1114) by exploiting the idea that octonions can be used as a basis for any $`8\times 8`$ real matrix. we have two sets of matrices, essentially, $$\begin{array}{ccccc}\delta _i& & (𝔼_i)_{\mu \nu }& =& \delta _{0\mu }\delta _{i\nu }\delta _{0\nu }\delta _{i\mu }f_{i\mu \nu },\\ 1|\delta _i& & (1|𝔼_i)_{\mu \nu }& =& \delta _{0\mu }\delta _{i\nu }\delta _{0\nu }\delta _{i\mu }+f_{i\mu \nu }.\end{array}$$ (15) The set of matrices $`𝔼_i`$ and $`1|𝔼_i`$have appeared in different octonionic works e.g. . We suggest that their most appropriate names should be *the canonical left and right octonionic structure at the north/south pole of the seven sphere*. By explicit calculation, one finds that $`[𝔼_i,𝔼_j]\phi `$ $`=`$ $`2f_{ijk}𝔼_k\phi 2[𝔼_i,1|𝔼_j]\phi `$ (16) $`[1|𝔼_i,1|𝔼_j]\phi `$ $`=`$ $`2f_{ijk}1|𝔼_k\phi 2[𝔼_i,1|𝔼_j]\phi `$ (17) $`\{𝔼_i,𝔼_j\}\phi `$ $`=`$ $`2\delta _{ij}\phi `$ (18) $`\{1|𝔼_i,1|𝔼_j\}\phi `$ $`=`$ $`2\delta _{ij}\phi `$ (19) where $`\phi `$ is represented by a column matrix $$\phi ^t=\left(\begin{array}{cccccccc}\phi _0\hfill & \phi _1\hfill & \phi _2\hfill & \phi _3\hfill & \phi _4\hfill & \phi _5\hfill & \phi _6\hfill & \phi _7\hfill \end{array}\right)$$ The word “isomorphic” above is justified since {$`\delta _i`$} is associative and the same holds obviously for our {$`𝔼_i`$ } as they are written in terms of matrices. Our Jacobian identities are $$[\delta _i,[\delta _j,\delta _k]]\phi +[\delta _j,[\delta _k,\delta _i]]\phi +[\delta _k,[\delta _i,\delta _j]]\phi =0,$$ (20) or $$[𝔼_i,[𝔼_j,𝔼_k]]\phi +[𝔼_j,[𝔼_k,𝔼_i]]\phi +[𝔼_k,[𝔼_i,𝔼_j]]\phi =0.$$ (21) We shall return to these identities again at the end of this section. Fixing the direction of the application for any imaginary octonionic units extracts a part of the algebra that respects associativity, but a certain price has to be paid. The presence of the $`\phi `$ is essential and either of {$`𝔼_i`$} or {$`1|𝔼_i`$} is an open algebra, they don’t close upon the action of the commutator. Here comes the second step, the soft Lie algebra idea. It is clear that the right hand side of (11) or (16) has a complicated $`\phi `$ dependence. Knowing that the seven sphere has a torsion that varies from one point to another and mimicking the Lie group case where the structure constants are proportional to the fixed group torsion, it is natural to propose that (11) may be redefined as $$[\delta _i,\delta _j]\phi =2f_{ijk}^{(+)}(\phi )\delta _k\phi ,$$ (22) where $`f_{ijk}^{(+)}(\phi )`$ are structure functions that vary over the whole $`S^7`$manifold. These structure functions $`f_{ijk}^{(+)}(\phi )`$ were computed previously using different properties of the associator and some other octonionic identities in . Here we use our matrix representation to give another alternative way to calculate $`f_{ijk}^{(+)}(\phi )`$ $$[𝔼_i,𝔼_j]\phi =2f_{ijk}^{(+)}(\phi )𝔼_k\phi .$$ (23) Let’s do it for the following example $$[𝔼_1,𝔼_2]\phi =2f_{12k}^{(+)}(\phi )𝔼_k\phi ,$$ (24) which is equivalent to the following eight equations , $`\begin{array}{c}_{\phi _3=\phi _1f_{121}^{(+)}(\phi )+\phi _2f_{122}^{(+)}(\phi )+\phi _3f_{123}^{(+)}(\phi )+\phi _4f_{124}^{(+)}(\phi )+\phi _5f_{125}^{(+)}(\phi )+\phi _6f_{126}^{(+)}(\phi )+\phi _7f_{127}^{(+)}(\phi ),}\hfill \\ _{\phi _2=\phi _2f_{123}^{(+)}(\phi )\phi _0f_{121}^{(+)}(\phi )\phi _3f_{122}^{(+)}(\phi )\phi _5f_{124}^{(+)}(\phi )+\phi _4f_{125}^{(+)}(\phi )+\phi _7f_{126}^{(+)}(\phi )\phi _6f_{127}^{(+)}(\phi ),}\hfill \\ _{\phi _1=\phi _0f_{122}^{(+)}(\phi )\phi _3f_{121}^{(+)}(\phi )+\phi _1f_{123}^{(+)}(\phi )+\phi _6f_{124}^{(+)}(\phi )+\phi _7f_{125}^{(+)}(\phi )\phi _4f_{126}^{(+)}(\phi )\phi _5f_{127}^{(+)}(\phi ),}\hfill \\ _{\phi _0=\phi _2f_{121}^{(+)}(\phi )\phi _1f_{122}^{(+)}(\phi )+\phi _0f_{123}^{(+)}(\phi )+\phi _7f_{124}^{(+)}(\phi )\phi _6f_{125}^{(+)}(\phi )+\phi _5f_{126}^{(+)}(\phi )\phi _4f_{127}^{(+)}(\phi ),}\hfill \\ _{\phi _7=\phi _0f_{124}^{(+)}(\phi )\phi _5f_{121}^{(+)}(\phi )\phi _6f_{122}^{(+)}(\phi )\phi _7f_{123}^{(+)}(\phi )+\phi _1f_{125}^{(+)}(\phi )+\phi _2f_{126}^{(+)}(\phi )+\phi _3f_{127}^{(+)}(\phi ),}\hfill \\ _{\phi _6=\phi _7f_{122}^{(+)}(\phi )\phi _4f_{121}^{(+)}(\phi )\phi _6f_{123}^{(+)}(\phi )+\phi _1f_{124}^{(+)}(\phi )\phi _0f_{125}^{(+)}(\phi )+\phi _3f_{126}^{(+)}(\phi )\phi _2f_{127}^{(+)}(\phi ),}\hfill \\ _{\phi _5=\phi _7f_{121}^{(+)}(\phi )+\phi _4f_{122}^{(+)}(\phi )\phi _5f_{123}^{(+)}(\phi )\phi _2f_{124}^{(+)}(\phi )+\phi _3f_{125}^{(+)}(\phi )+\phi _0f_{126}^{(+)}(\phi )\phi _1f_{127}^{(+)}(\phi ),}\hfill \\ _{\phi _4=\phi _6f_{121}^{(+)}(\phi )\phi _5f_{122}^{(+)}(\phi )\phi _4f_{123}^{(+)}(\phi )+\phi _3f_{124}^{(+)}(\phi )+\phi _2f_{125}^{(+)}(\phi )\phi _1f_{126}^{(+)}(\phi )\phi _0f_{127}^{(+)}(\phi ).}\hfill \end{array}`$ (33) (34) We now solve these equations for the seven unknown $`f_{12i}^{(+)}\left(\phi \right)`$. We find $$\begin{array}{c}f_{121}^{(+)}\left(\phi \right)=f_{122}^{(+)}\left(\phi \right)=0,\hfill \\ f_{124}^{(+)}\left(\phi \right)=+2\frac{\phi _0\phi _7\phi _5\phi _2+\phi _6\phi _1+\phi _3\phi _4}{r^2},\hfill \\ f_{125}^{(+)}\left(\phi \right)=2\frac{\phi _0\phi _6\phi _3\phi _5\phi _1\phi _7\phi _2\phi _4}{r^2},\hfill \\ f_{126}^{(+)}\left(\phi \right)=+2\frac{\phi _0\phi _5\phi _1\phi _4+\phi _7\phi _2+\phi _3\phi _6}{r^2},\hfill \\ f_{127}^{(+)}\left(\phi \right)=2\frac{\phi _0\phi _4+\phi _6\phi _2+\phi _1\phi _5\phi _3\phi _7}{r^2},\hfill \\ f_{123}^{(+)}\left(\phi \right)=\frac{\phi _{0}^{}{}_{}{}^{2}\phi _{6}^{}{}_{}{}^{2}\phi _{5}^{}{}_{}{}^{2}+\phi _{2}^{}{}_{}{}^{2}\phi _{4}^{}{}_{}{}^{2}+\phi _{1}^{}{}_{}{}^{2}+\phi _{3}^{}{}_{}{}^{2}\phi _{7}^{}{}_{}{}^{2}}{r^2},\hfill \end{array}$$ where $$r^2=(\phi _0^2+\phi _1^2+\phi _2^2+\phi _3^2+\phi _4^2+\phi _5^2+\phi _6^2+\phi _7^2).$$ Along the same lines we can calculate all the structure functions, we give all of them in Appendix A. What we have just calculated is commonly called the (+) torsion, we can find the (–) torsion by replacing the left by right multiplication in (23) $$[1|𝔼_i,1|𝔼_j]\phi =2f_{ijk}^{()}(\phi )1|𝔼_k\phi .$$ (35) we find $$\begin{array}{c}f_{121}^{()}\left(\phi \right)=f_{122}^{()}\left(\phi \right)=0,\hfill \\ f_{124}^{()}\left(\phi \right)=+2\frac{\phi _0\phi _7+\phi _5\phi _2\phi _6\phi _1\phi _3,\phi _4}{r^2},\hfill \\ f_{125}^{()}\left(\phi \right)=2\frac{\phi _0\phi _6+\phi _3\phi _5+\phi _1\phi _7+\phi _2\phi _4}{r^2},\hfill \\ f_{126}^{()}\left(\phi \right)=+2\frac{\phi _0\phi _5+\phi _1\phi _4\phi _7\phi _2\phi _3,\phi _6}{r^2},\hfill \\ f_{127}^{()}\left(\phi \right)=2\frac{\phi _0\phi _4\phi _6\phi _2\phi _1\phi _5+\phi _3\phi _7}{r^2},\hfill \\ f_{123}^{()}\left(\phi \right)=\frac{\phi _{0}^{}{}_{}{}^{2}\phi _{6}^{}{}_{}{}^{2}\phi _{4}^{}{}_{}{}^{2}+\phi _{2}^{}{}_{}{}^{2}+\phi _{1}^{}{}_{}{}^{2}\phi _{7}^{}{}_{}{}^{2}\phi _{5}^{}{}_{}{}^{2}+\phi _{3}^{}{}_{}{}^{2}}{r^2},\hfill \end{array}$$ the remaining $`f_{ijk}^{()}\left(\phi \right)`$ are listed in appendix A. Let’s pause for a moment and note some of the evident features of these $`f_{ijk}^{(\pm )}(\phi )`$, * One notices immediately that at $`\phi ^t`$=(1,0,0,0,0,0,0,0) / (-1,0,0,0,0,0,0,0), the north / south pole (NP/SP), we recover the octonionic structure constants: $`f_{ijk}^{(+)}(NP/SP)=f_{ijk}^{()}(NP/SP)=f_{ijk}`$ and any non-standard cycles vanishes e.g. $`f_{567}^{(\pm )}(NP/SP)=0`$. * Our construction started from a given multiplication table and as there are different choices , we can have different families. * Over S<sup>7</sup>, $`_{\text{S}^7}f_{ijk}^{(\pm )}(\phi )0.`$ This is a very important characteristic of the seven sphere. To manifest the $`\phi `$ dependence, let’s give a non–tivial example. To simplify the notations, we use here $`(i,j,k)^{(\pm )}`$ for $`f_{ijk}^{(\pm )}(\phi ).`$ At $`\left(\phi _\mu =\frac{\mu +1}{\sqrt{204}}\right)`$, we find $`\begin{array}{c}\begin{array}{ccc}(1,2,3)^{(+)}=12/17,\hfill & (2,5,7)^{(+)}=4/51,\hfill & (1,5,6)^{(+)}=1/51,\hfill \\ (1,4,5)^{(+)}=6/17,\hfill & (1,7,6)^{(+)}=8/51,\hfill & (3,6,5)^{(+)}=0,\hfill \\ (4,3,7)^{(+)}=2/51,\hfill & (4,2,6)^{(+)}=3/17,\hfill & \end{array}\\ \begin{array}{ccc}(1,2,4)^{(+)}=4/17,\hfill & (1,5,2)^{(+)}=8/17,\hfill & (3,5,4)^{(+)}=44/51,\hfill \\ (5,6,7)^{(+)}=10/17,\hfill & (1,3,4)^{(+)}=5/17,\hfill & (4,1,6)^{(+)}=14/17,\hfill \\ (1,5,7)^{(+)}=40/51,\hfill & (3,5,7)^{(+)}=2/17,\hfill & (3,1,6)^{(+)}=14/51,\hfill \\ (2,3,5)^{(+)}=23/51,\hfill & (1,7,4)^{(+)}=4/17,\hfill & (4,5,6)^{(+)}=16/51,\hfill \\ (2,6,5)^{(+)}=38/51,\hfill & (1,6,2)^{(+)}=8/17,\hfill & (6,3,2)^{(+)}=22/51,\hfill \\ (1,3,5)^{(+)}=10/51,\hfill & (2,4,3)^{(+)}=2/17,\hfill & (4,3,6)^{(+)}=20/51,\hfill \\ (3,7,6)^{(+)}=13/17,\hfill & (1,2,7)^{(+)}=1/17,\hfill & (4,2,7)^{(+)}=16/17,\hfill \\ (2,7,3)^{(+)}=16/17,\hfill & (2,6,7)^{(+)}=4/51,\hfill & (7,1,3)^{(+)}=28/51,\hfill \\ (2,4,5)^{(+)}=2/17,\hfill & (4,7,5)^{(+)}=7/51,\hfill & (4,6,7)^{(+)}=10/51.\hfill \end{array}\end{array}`$ (50) (51) We have some kind of dynamical Lie algebra of seven generators with structure “constants” that change their values from one point to another. Let us emphasis the difference between considering $`𝔼_i`$ as an open algebra or as elements of a soft seven sphere, observe that $$\begin{array}{cc}[𝔼_1,𝔼_2]\hfill & =2𝔼_32[𝔼_1,1|𝔼_2],\hfill \end{array}$$ (52) but $$\begin{array}{ccc}[𝔼_1,𝔼_2]\mathrm{\Phi }\hfill & =2f_{123}^{(+)}\left(\phi \right)𝔼_3\mathrm{\Phi }\hfill & +2f_{124}^{(+)}\left(\phi \right)𝔼_4\mathrm{\Phi }\hfill \\ & +2f_{125}^{(+)}\left(\phi \right)𝔼_5\mathrm{\Phi }\hfill & +2f_{126}^{(+)}\left(\phi \right)𝔼_6\mathrm{\Phi }+2f_{127}^{(+)}\left(\phi \right)𝔼_7\mathrm{\Phi }.\hfill \end{array}$$ (53) At the NP $$\mathrm{\Phi }_{NP}^t=\left(\begin{array}{cccccccc}1\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \end{array}\right)$$ (54) we still have $$[𝔼_1,𝔼_2]=2𝔼_32[𝔼_1,1|𝔼_2]$$ (55) whereas $`[𝔼_1,𝔼_2]\mathrm{\Phi }_{NP}`$ $`=`$ $`2f_{12k}^{(+)}(\phi _{NP})𝔼_k\mathrm{\Phi }_{NP}`$ $`=`$ $`2𝔼_3\mathrm{\Phi }_{NP}`$ $`\left(\begin{array}{cccccccc}0\hfill & 0\hfill & 0\hfill & 2\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 2\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 2\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 2\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 2\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 2\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 2\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 2\hfill & 0\hfill & 0\hfill & 0\hfill \end{array}\right)\left(\begin{array}{c}1\hfill \\ 0\hfill \\ 0\hfill \\ 0\hfill \\ 0\hfill \\ 0\hfill \\ 0\hfill \\ 0\hfill \end{array}\right)=\left(\begin{array}{c}0\hfill \\ 0\hfill \\ 0\hfill \\ 2\hfill \\ 0\hfill \\ 0\hfill \\ 0\hfill \\ 0\hfill \end{array}\right)`$ $`=`$ $`\left(\begin{array}{cccccccc}0\hfill & 0\hfill & 0\hfill & 2\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 2\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 2\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 2\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 2\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 2\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 2\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 2\hfill & 0\hfill & 0\hfill & 0\hfill \end{array}\right)\left(\begin{array}{c}1\hfill \\ 0\hfill \\ 0\hfill \\ 0\hfill \\ 0\hfill \\ 0\hfill \\ 0\hfill \\ 0\hfill \end{array}\right).`$ At $`\mathrm{\Phi }_W\left(\phi _\mu =\frac{\mu +1}{\sqrt{204}}\right),`$ we still have $$[𝔼_1,𝔼_2]=2𝔼_32[𝔼_1,1|𝔼_2]$$ we find that $`[𝔼_1,𝔼_2]\mathrm{\Phi }_W`$ $`=`$ $`2f_{12k}^{(+)}(\phi _W)𝔼_k\mathrm{\Phi }_W`$ $`=`$ $`\left({\displaystyle \frac{24}{17}}𝔼_3+{\displaystyle \frac{8}{17}}𝔼_4+{\displaystyle \frac{16}{17}}𝔼_5+{\displaystyle \frac{16}{17}}𝔼_6{\displaystyle \frac{2}{17}}𝔼_7\right)\mathrm{\Phi }_W`$ $`\left(\begin{array}{cccccccc}0\hfill & 0\hfill & 0\hfill & 2\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 2\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 2\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 2\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 2\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 2\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 0\hfill & 2\hfill & 0\hfill & 0\hfill \\ 0\hfill & 0\hfill & 0\hfill & 0\hfill & 2\hfill & 0\hfill & 0\hfill & 0\hfill \end{array}\right)\left(\begin{array}{c}\frac{1}{\sqrt{204}}\hfill \\ \frac{2}{\sqrt{204}}\hfill \\ \frac{3}{\sqrt{204}}\hfill \\ \frac{4}{\sqrt{204}}\hfill \\ \frac{5}{\sqrt{204}}\hfill \\ \frac{6}{\sqrt{204}}\hfill \\ \frac{7}{\sqrt{204}}\hfill \\ \frac{8}{\sqrt{204}}\hfill \end{array}\right)=\left(\begin{array}{c}\frac{4}{51}\sqrt{51}\hfill \\ \frac{1}{17}\sqrt{51}\hfill \\ \frac{2}{51}\sqrt{51}\hfill \\ \frac{1}{51}\sqrt{51}\hfill \\ \frac{8}{51}\sqrt{51}\hfill \\ \frac{7}{51}\sqrt{51}\hfill \\ \frac{2}{17}\sqrt{51}\hfill \\ \frac{5}{51}\sqrt{51}\hfill \end{array}\right)`$ $`=`$ $`\left(\begin{array}{cccccccc}0\hfill & 0\hfill & 0\hfill & \frac{24}{17}\hfill & \frac{8}{17}\hfill & \frac{16}{17}\hfill & \frac{16}{17}\hfill & \frac{2}{17}\hfill \\ 0\hfill & 0\hfill & \frac{24}{17}\hfill & 0\hfill & \frac{16}{17}\hfill & \frac{8}{17}\hfill & \frac{2}{17}\hfill & \frac{16}{17}\hfill \\ 0\hfill & \frac{24}{17}\hfill & 0\hfill & 0\hfill & \frac{16}{17}\hfill & \frac{2}{17}\hfill & \frac{8}{17}\hfill & \frac{16}{17}\hfill \\ \frac{24}{17}\hfill & 0\hfill & 0\hfill & 0\hfill & \frac{2}{17}\hfill & \frac{16}{17}\hfill & \frac{16}{17}\hfill & \frac{8}{17}\hfill \\ \frac{8}{17}\hfill & \frac{16}{17}\hfill & \frac{16}{17}\hfill & \frac{2}{17}\hfill & 0\hfill & 0\hfill & 0\hfill & \frac{24}{17}\hfill \\ \frac{16}{17}\hfill & \frac{8}{17}\hfill & \frac{2}{17}\hfill & \frac{16}{17}\hfill & 0\hfill & 0\hfill & \frac{24}{17}\hfill & 0\hfill \\ \frac{16}{17}\hfill & \frac{2}{17}\hfill & \frac{8}{17}\hfill & \frac{16}{17}\hfill & 0\hfill & \frac{24}{17}\hfill & 0\hfill & 0\hfill \\ \frac{2}{17}\hfill & \frac{16}{17}\hfill & \frac{16}{17}\hfill & \frac{8}{17}\hfill & \frac{24}{17}\hfill & 0\hfill & 0\hfill & 0\hfill \end{array}\right)\left(\begin{array}{c}\frac{1}{\sqrt{204}}\hfill \\ \frac{2}{\sqrt{204}}\hfill \\ \frac{3}{\sqrt{204}}\hfill \\ \frac{4}{\sqrt{204}}\hfill \\ \frac{5}{\sqrt{204}}\hfill \\ \frac{6}{\sqrt{204}}\hfill \\ \frac{7}{\sqrt{204}}\hfill \\ \frac{8}{\sqrt{204}}\hfill \end{array}\right).`$ We are not making a projection but a reformulation of the algebra. This fact should always be kept in mind. The same happens in a non-trivial way for the Jacobi identity, i.e. $$\left(f_{ijm}(\phi )f_{mkt}(\phi )+f_{jkm}(\phi )f_{mit}(\phi )+f_{kim}(\phi )f_{mjt}(\phi )\right)𝔼_t\phi =0,$$ (60) but, in general, $$\left(f_{ijm}(\phi )f_{mkt}(\phi )+f_{jkm}(\phi )f_{mit}(\phi )+f_{kim}(\phi )f_{mjt}(\phi )\right)0.$$ (61) Another important feature is $$[𝔼_i,1|𝔼_j]=2f_{ijk}\left(\phi \right)𝔼_k+[𝔼_i,𝔼_j]=2f_{ijk}\left(\phi \right)1|𝔼_k+[1|𝔼_i,1|𝔼_j]$$ (62) which is equal to zero iff $`i=j`$ but for the soft seven sphere, $`[𝔼_i,1|𝔼_j]\phi =0`$ not only for $`i=j`$ but also at the NP/SP for any i,j. Lastly over any group manifold the left torsion equals minus the right torsion, but for $`S^7`$ this is not in general true. ## 2 More General Solutions In the previous section , we have used brute force to calculate $`f_{ijk}^{(+)}\left(\phi \right)`$. There is another way, smarter and easier. We have the following situation $$𝔼_i𝔼_j\phi =\left(\delta _{ij}+f_{ijk}^{(+)}\left(\phi \right)𝔼_k\right)\phi $$ (63) but one can check that our $`𝔼_i`$ defines what Cartan calls pure spinors , $$\phi ^t𝔼_i\phi =0$$ (64) thus $$\phi ^t\left(𝔼_i𝔼_j\right)\phi =\phi ^t\left(\delta _{ij}\right)\phi ,$$ (65) using $$\left(𝔼_k\right)^1=𝔼_k$$ (66) we find $$\phi ^t\left(𝔼_k𝔼_i𝔼_j\right)\phi =\phi ^t\left(f_{ijk}^{(+)}\left(\phi \right)\right)\phi $$ (67) but $$\phi ^t\phi =r^2$$ (68) which gives us $$f_{ijk}^{(+)}\left(\phi \right)=\frac{\phi ^t\left(𝔼_k𝔼_i𝔼_j\right)\phi }{r^2}.$$ (69) and $$f_{ijk}^{()}\left(\phi \right)=\frac{\phi ^t\left(1|𝔼_k\mathrm{\hspace{0.33em}\hspace{0.33em}1}|𝔼_i\mathrm{\hspace{0.33em}\hspace{0.33em}1}|𝔼_j\right)\phi }{r^2}.$$ (70) There is another interesting property to note $$\phi ^t[𝔼_i,1|𝔼_j]\phi =0$$ (71) which may be the generalization of the standard Lie algebra relation, left and right action commute everywhere over the group manifold. The left and right torsions that we have constructed are not the only parallelizable torsions of S<sup>7</sup>. Our $`𝔼_i`$ and $`1|𝔼_i`$ are given in terms of the octonionic structure constants (15) i.e. the torsion at NP/SP. Considering two new points, we may define new sets of $`𝔼_i`$ and $`1|𝔼_i`$. As S<sup>7</sup> contains an infinity of points, practically, we have an infinity of parallelizable torsion. If our method is self contained and sufficient, we should be able to construct these infinity of pointwise structures. Indeed, $`𝔼_i\left(\phi \right)`$ and $`1|𝔼_i\left(\phi \right)`$ are in general $$\begin{array}{ccccc}\delta _i& & (𝔼_i(\phi ))_{\mu \nu }& =& \delta _{0\mu }\delta _{i\nu }\delta _{0\nu }\delta _{i\mu }f_{i\mu \nu }^{(+)}(\phi ),\\ 1|\delta _i& & (1|𝔼_i(\phi ))_{\mu \nu }& =& \delta _{0\mu }\delta _{i\nu }\delta _{0\nu }\delta _{i\mu }+f_{i\mu \nu }^{()}(\phi ),\end{array}$$ (72) in complete analogy with (11,12,13,14). Of course the soft Algebra idea should hold here as well as for the special $`(𝔼_i,1|𝔼_i)`$ constructed in terms of the north pole torsion. Repeating the calculation in terms of $`(𝔼_i(\phi ),1|𝔼_i(\phi ))`$. Let us introduce a new vector field $`\lambda `$, $$\lambda ^t=\left(\begin{array}{cccccccc}\lambda _0\hfill & \lambda _1\hfill & \lambda _2\hfill & \lambda _3\hfill & \lambda _4\hfill & \lambda _5\hfill & \lambda _6\hfill & \lambda _7\hfill \end{array}\right).$$ (73) We define two new generalized structure functions $`[𝔼_i(\phi ),𝔼_j(\phi )]\lambda `$ $`=`$ $`2f_{ijk}^{(++)}(\phi ,\lambda )𝔼_k(\phi )\lambda `$ (74) $`[1|𝔼_i(\phi ),1|𝔼_j(\phi )]\lambda `$ $`=`$ $`2f_{ijk}^{()}(\phi ,\lambda )1|𝔼_k(\phi )\lambda `$ (75) where $`f_{ijk}^{\left(\pm \pm \right)}(\phi ,\lambda )`$ have a very complicated structure, $`f_{ijk}^{\left(++\right)}(\phi ,\lambda )`$ $`=`$ $`{\displaystyle \frac{\lambda ^t\left(𝔼_k\left(\phi \right)𝔼_i\left(\phi \right)𝔼_j\left(\phi \right)\right)\lambda }{r^2}},`$ (76) $`f_{ijk}^{\left(\right)}(\phi ,\lambda )`$ $`=`$ $`{\displaystyle \frac{\lambda ^t\left(1|𝔼_k\left(\phi \right)\mathrm{\hspace{0.33em}\hspace{0.33em}1}|𝔼_i\left(\phi \right)\mathrm{\hspace{0.33em}\hspace{0.33em}1}|𝔼_j\left(\phi \right)\right)\lambda }{r^2}}`$ (77) ## 3 Some Group Theory An arbitrary octonion can be associated to $`^8=^7`$ where $``$ denotes the subspace spanned by the identity $`e_0=1`$. Octonions with unit length define the octonionic unit sphere $`S^7`$. The isometries of octonions is described by $`O(8)`$ which may be decomposed as $$O(8):HKE$$ (78) where $`H`$ is the 14 parameters $`G_2`$ algebra of the automorphism group of $`octonions,`$ K is the torsionful seven sphere $`SO(7)/G_2`$ and our E is the round seven sphere $`SO(8)/SO(7)`$. In fact the different three non-equivalent representation of O(8) - the vectorial so(8) and the two different spinorial $`spin^L(8)`$ and $`spin^R(8)`$, which are related by triality, can be realized by suitable left and right octonionic multiplication. The reduction of O(8) to O(7) induces $`so(8)so(7)1`$, $`spin^R(8)spin(7)`$ and $`spin^L(8)spin(7)`$. We would like to show how to generate these different Lie algebras entirely from our canonical left/right octonionic structures. We start from the $`8\times 8`$ gamma matrices $`\gamma _{\mu \nu }^i`$ in seven dimensions, using $`\delta _{ij}`$ as our flat Euclidean metric, $$\{\gamma ^i,\gamma ^j\}=2\delta ^{ij}\mathrm{𝟏}_8,$$ (79) where $`i,j,\mathrm{}=1,2,\mathrm{}7`$ and $`\mu ,\nu ,\mathrm{}=0,1,2,\mathrm{}7`$. We can use either of the following choices $$\gamma _+^j=i𝔼_jor\gamma _{}^j=i1|𝔼_j,$$ (80) of course the $`i`$ in the right hand sides is the imaginary complex unit. This relates our antisymmetric, Hermitian and hence purely imaginary gamma matrices to the canonical octonionic left/right structures. The antisymmetric product of two gamma matrices will be denoted by $$\gamma ^{ij}=\gamma ^{[i}\gamma ^{j]},$$ (81) and we have <sup>3</sup><sup>3</sup>3It is interesting to note that this equation may be used as an alternative definition for the octonionic multiplication table. $$\gamma ^i\gamma ^j\gamma ^k=\frac{1}{4!}ϵ^{ijklmnp}\gamma ^l\gamma ^m\gamma ^n\gamma ^p.$$ (82) The matrices $`\gamma ^{ij}`$ span the 21 generators $`J^{ij}`$of spin(7) in its eight-dimensional spinor representation. The spinorial representation of spin(7) can be enlarged to the left/right handed spinor representation of spin(8) by different ways. The easiest one is to include either of $`\pm 𝔼_i`$ or $`\pm 1|𝔼_i`$ defining $`J^i=J^{i0}`$, so(8) can be written as $`[J^i,J^i]`$ $`=`$ $`2J^{ij}`$ (83) $`[J^i,J^{mn}]`$ $`=`$ $`2\delta ^{im}J^n2\delta ^{in}J^m`$ (84) $`[J^{ij},J^{kl}]`$ $`=`$ $`2\delta ^{jk}J^{il}+2\delta ^{il}J^{jk}2\delta ^{ik}J^{jl}2\delta ^{jl}J^{ik}.`$ (85) The automorphism group of octonions is $`G_2SO(7)SO(8)`$. A suitable basis for $`G_2`$ is $$H_{ij}=f_{ijk}\left(𝔼_k1|𝔼_k\right)\frac{3}{2}[𝔼_i,1|𝔼_j],$$ (86) which implies the linear relations $$f_{ijk}H_{jk}=0,$$ (87) These constraints enforce $`H_{ij}`$ to generate the 14 dimensional vector space of $`G_2`$. There are different ways to represent the remaining seven generators, denoted here by K, $`{\displaystyle \frac{so(7)}{G_2}}`$ $`:`$ $`K_v^{\pm i}=\pm {\displaystyle \frac{1}{2}}\left(𝔼_i1|𝔼_i\right),`$ (88) $`{\displaystyle \frac{spin(7)}{G_2}}`$ $`:`$ $`K_s^{\pm i}=\pm \left({\displaystyle \frac{1}{2}}𝔼_i+1|𝔼_i\right),`$ (89) $`{\displaystyle \frac{\overline{spin(7)}}{G_2}}`$ $`:`$ $`\overline{K}_s^{\pm i}=\left(𝔼_i+{\displaystyle \frac{1}{2}}1|𝔼_i\right),`$ (90) Defining the conjugate representation<sup>4</sup><sup>4</sup>4n.b. this definition is not matrix conjugation. by $$\overline{𝔼}=1|𝔼\text{and }\overline{1|𝔼}=𝔼,$$ (91) (88) is self-conjugate while (89) is octonionic-conjugate to (90). The vector representation so(7) generated by $`H_{ij}K_v^{\pm i}`$ is seven dimensional because $`K_v^{\pm i}e_0=0`$ whereas the spin(7) representation generated by $`H_{ij}K_s^{\pm i}`$ is eight dimensional. To make apparent the role of the automorphism group $`G_2`$, the different commutators of $`𝔼`$ and $`1|𝔼`$ may be written as $`[𝔼_i,𝔼_j]`$ $`=`$ $`{\displaystyle \frac{1}{3}}\left(4H_{ij}+2f_{ijk}𝔼_k+4f_{ijk}1|𝔼_k\right),`$ (92) $`[1|𝔼_i,1|𝔼_j]`$ $`=`$ $`{\displaystyle \frac{1}{3}}\left(4H_{ij}4f_{ijk}𝔼_k2f_{ijk}1|𝔼_k\right),`$ (93) $`[𝔼_i,1|𝔼_j]`$ $`=`$ $`{\displaystyle \frac{1}{3}}\left(2H_{ij}+2f_{ijk}𝔼_k2f_{ijk}1|𝔼_k\right).`$ (94) or $`G_2`$ given by $$H_{ij}=\frac{1}{2}([𝔼_i,𝔼_j]+[1|𝔼_i,1|𝔼_j]+[𝔼_i,1|𝔼_j]).$$ (95) Thus as we promised, the $`𝔼`$ and $`1|𝔼`$ are the necessary and the sufficient building blocks for expressing the different Lie algebras and coset representations related to the seven sphere. Note that all the constructions given in this section start from the Clifford algebra relation (80), and the formulation holds equally for $`𝔼\left(\phi \right)`$ and $`1|𝔼\left(\phi \right)`$. I am grateful to C. Imbimbo, P. Rotelli, G. Thompson, A. Van Proeyen for useful comments. Also, I would like to thank S. Marchiafava and F. Englert for encouragements. ## Appendix A The structure function of the soft seven sphere The seven standard cycles are given by $$\begin{array}{c}f_{123}^{(+)}(\phi )=f_{123}^{()}(\phi )=\frac{\phi _{0}^{}{}_{}{}^{2}\phi _{6}^{}{}_{}{}^{2}\phi _{5}^{}{}_{}{}^{2}+\phi _{2}^{}{}_{}{}^{2}\phi _{4}^{}{}_{}{}^{2}+\phi _{1}^{}{}_{}{}^{2}+\phi _{3}^{}{}_{}{}^{2}\phi _{7}^{}{}_{}{}^{2}}{r^2},\hfill \\ f_{145}^{(+)}(\phi )=f_{145}^{()}(\phi )=\frac{\phi _{0}^{}{}_{}{}^{2}\phi _{6}^{}{}_{}{}^{2}+\phi _{4}^{}{}_{}{}^{2}\phi _{2}^{}{}_{}{}^{2}+\phi _{1}^{}{}_{}{}^{2}\phi _{7}^{}{}_{}{}^{2}+\phi _{5}^{}{}_{}{}^{2}\phi _{3}^{}{}_{}{}^{2}}{r^2},\hfill \\ f_{176}^{(+)}(\phi )=f_{176}^{()}(\phi )=\frac{\phi _{0}^{}{}_{}{}^{2}+\phi _{6}^{}{}_{}{}^{2}\phi _{4}^{}{}_{}{}^{2}\phi _{2}^{}{}_{}{}^{2}+\phi _{1}^{}{}_{}{}^{2}+\phi _{7}^{}{}_{}{}^{2}\phi _{5}^{}{}_{}{}^{2}\phi _{3}^{}{}_{}{}^{2}}{r^2},\hfill \\ f_{246}^{(+)}(\phi )=f_{246}^{()}(\phi )=\frac{\phi _{0}^{}{}_{}{}^{2}+\phi _{6}^{}{}_{}{}^{2}+\phi _{4}^{}{}_{}{}^{2}+\phi _{2}^{}{}_{}{}^{2}\phi _{1}^{}{}_{}{}^{2}\phi _{7}^{}{}_{}{}^{2}\phi _{5}^{}{}_{}{}^{2}\phi _{3}^{}{}_{}{}^{2}}{r^2},\hfill \\ f_{257}^{(+)}(\phi )=f_{257}^{()}(\phi )=\frac{\phi _{0}^{}{}_{}{}^{2}\phi _{6}^{}{}_{}{}^{2}\phi _{4}^{}{}_{}{}^{2}+\phi _{2}^{}{}_{}{}^{2}\phi _{1}^{}{}_{}{}^{2}+\phi _{7}^{}{}_{}{}^{2}+\phi _{5}^{}{}_{}{}^{2}\phi _{3}^{}{}_{}{}^{2}}{r^2},\hfill \\ f_{347}^{(+)}(\phi )=f_{347}^{()}(\phi )=\frac{\phi _{0}^{}{}_{}{}^{2}\phi _{6}^{}{}_{}{}^{2}+\phi _{4}^{}{}_{}{}^{2}\phi _{2}^{}{}_{}{}^{2}\phi _{1}^{}{}_{}{}^{2}+\phi _{7}^{}{}_{}{}^{2}\phi _{5}^{}{}_{}{}^{2}+\phi _{3}^{}{}_{}{}^{2}}{r^2},\hfill \\ f_{365}^{(+)}(\phi )=f_{365}^{()}(\phi )=\frac{\phi _{0}^{}{}_{}{}^{2}+\phi _{6}^{}{}_{}{}^{2}\phi _{4}^{}{}_{}{}^{2}\phi _{2}^{}{}_{}{}^{2}\phi _{1}^{}{}_{}{}^{2}\phi _{7}^{}{}_{}{}^{2}+\phi _{5}^{}{}_{}{}^{2}+\phi _{3}^{}{}_{}{}^{2}}{r^2},\hfill \end{array}$$ and the non-standard subset $$\begin{array}{cc}f_{124}^{(+)}(\phi )=+2\frac{\phi _0\phi _7\phi _5\phi _2+\phi _6\phi _1+\phi _3\phi _4}{r^2},\hfill & f_{125}^{(+)}(\phi )=2\frac{\phi _0\phi _6\phi _3\phi _5\phi _1\phi _7\phi _2\phi _4}{r^2},\hfill \\ f_{126}^{(+)}(\phi )=+2\frac{\phi _0\phi _5\phi _1\phi _4+\phi _7\phi _2+\phi _3\phi _6}{r^2},\hfill & f_{127}^{(+)}(\phi )=2\frac{\phi _0\phi _4+\phi _6\phi _2+\phi _1\phi _5\phi _3\phi _7}{r^2},\hfill \\ f_{143}^{(+)}(\phi )=+2\frac{\phi _0\phi _6+\phi _3\phi _5+\phi _2\phi _4\phi _1\phi _7}{r^2},\hfill & f_{146}^{(+)}(\phi )=2\frac{\phi _3\phi _0\phi _4\phi _7\phi _1\phi _2\phi _5\phi _6}{r^2},\hfill \\ f_{175}^{(+)}(\phi )=+2\frac{\phi _3\phi _0\phi _1\phi _2+\phi _5\phi _6+\phi _4\phi _7}{r^2},\hfill & f_{247}^{(+)}(\phi )=2\frac{\phi _0\phi _1\phi _7\phi _6\phi _4\phi _5\phi _3\phi _2}{r^2},\hfill \\ f_{147}^{(+)}(\phi )=+2\frac{\phi _2\phi _0\phi _4\phi _6+\phi _5\phi _7+\phi _1\phi _3}{r^2},\hfill & f_{243}^{(+)}(\phi )=2\frac{\phi _0\phi _5+\phi _1\phi _4+\phi _7\phi _2\phi _3\phi _6}{r^2},\hfill \\ f_{253}^{(+)}(\phi )=+2\frac{\phi _0\phi _4\phi _1\phi _5+\phi _6\phi _2+\phi _3\phi _7}{r^2},\hfill & f_{173}^{(+)}(\phi )=2\frac{\phi _0\phi _5\phi _7\phi _2\phi _3\phi _6\phi _1\phi _4}{r^2},\hfill \\ f_{245}^{(+)}(\phi )=+2\frac{\phi _3\phi _0+\phi _5\phi _6\phi _4\phi _7+\phi _1\phi _2}{r^2},\hfill & f_{256}^{(+)}(\phi )=+2\frac{\phi _0\phi _1\phi _3\phi _2+\phi _7\phi _6+\phi _4\phi _5}{r^2},\hfill \\ f_{361}^{(+)}(\phi )=+2\frac{\phi _0\phi _4+\phi _3\phi _7+\phi _1\phi _5\phi _6\phi _2}{r^2},\hfill & f_{362}^{(+)}(\phi )=2\frac{\phi _0\phi _7\phi _3\phi _4\phi _5\phi _2\phi _6\phi _1}{r^2},\hfill \\ f_{345}^{(+)}(\phi )=2\frac{\phi _2\phi _0\phi _5\phi _7\phi _1\phi _3\phi _4\phi _6}{r^2},\hfill & f_{346}^{(+)}(\phi )=+2\frac{\phi _0\phi _1+\phi _3\phi _2+\phi _7\phi _6\phi _4\phi _5}{r^2},\hfill \\ f_{367}^{(+)}(\phi )=+2\frac{\phi _2\phi _0+\phi _4\phi _6+\phi _5\phi _7\phi _1\phi _3}{r^2},\hfill & f_{135}^{(+)}(\phi )=2\frac{\phi _0\phi _7\phi _3\phi _4+\phi _6\phi _1+\phi _5\phi _2}{r^2},\hfill \\ f_{156}^{(+)}(\phi )=2\frac{\phi _2\phi _0+\phi _1\phi _3+\phi _4\phi _6\phi _5\phi _7}{r^2},\hfill & f_{237}^{(+)}(\phi )=+2\frac{\phi _0\phi _6\phi _2\phi _4+\phi _3\phi _5+\phi _1\phi _7}{r^2},\hfill \\ f_{267}^{(+)}(\phi )=2\frac{\phi _3\phi _0\phi _5\phi _6+\phi _4\phi _7+\phi _1\phi _2}{r^2},\hfill & f_{357}^{(+)}(\phi )=+2\frac{\phi _0\phi _1+\phi _4\phi _5\phi _7\phi _6+\phi _3\phi _2}{r^2},\hfill \\ f_{456}^{(+)}(\phi )=2\frac{\phi _0\phi _7+\phi _5\phi _2+\phi _3\phi _4\phi _6\phi _1}{r^2},\hfill & f_{457}^{(+)}(\phi )=+2\frac{\phi _0\phi _6+\phi _2\phi _4+\phi _1\phi _7\phi _3\phi _5}{r^2},\hfill \\ f_{467}^{(+)}(\phi )=2\frac{\phi _0\phi _5+\phi _1\phi _4+\phi _3\phi _6\phi _7\phi _2}{r^2},\hfill & f_{567}^{(+)}(\phi )=+2\frac{\phi _0\phi _4\phi _6\phi _2\phi _1\phi _5\phi _3\phi _7}{r^2}.\hfill \end{array}$$ For right actions, the non-standard cocycles are $$\begin{array}{cc}f_{124}^{()}(\phi )=+2\frac{\phi _0\phi _7+\phi _5\phi _2\phi _6\phi _1\phi _3\phi _4}{r^2},\hfill & f_{125}^{()}(\phi )=2\frac{\phi _0\phi _6+\phi _3\phi _5+\phi _1\phi _7+\phi _2\phi _4}{r^2},\hfill \\ f_{126}^{()}(\phi )=+2\frac{\phi _0\phi _5+\phi _1\phi _4\phi _7\phi _2\phi _3\phi _6}{r^2},\hfill & f_{127}^{()}(\phi )=2\frac{\phi _0\phi _4\phi _6\phi _2\phi _1\phi _5+\phi _3\phi _7}{r^2},\hfill \\ f_{143}^{()}(\phi )=+2\frac{\phi _0\phi _6\phi _3\phi _5\phi _2\phi _4+\phi _1\phi _7}{r^2},\hfill & f_{146}^{()}(\phi )=2\frac{\phi _3\phi _0+\phi _4\phi _7+\phi _1\phi _2+\phi _5\phi _6}{r^2},\hfill \\ f_{175}^{()}(\phi )=+2\frac{\phi _3\phi _0+\phi _1\phi _2\phi _5\phi _6\phi _4\phi _7}{r^2},\hfill & f_{247}^{()}(\phi )=2\frac{\phi _0\phi _1+\phi _7\phi _6+\phi _4\phi _5+\phi _3\phi _2}{r^2},\hfill \\ f_{147}^{()}(\phi )=+2\frac{\phi _2\phi _0+\phi _4\phi _6\phi _5\phi _7\phi _1\phi _3}{r^2},\hfill & f_{243}^{()}(\phi )=2\frac{\phi _0\phi _5\phi _1\phi _4\phi _7\phi _2+\phi _3\phi _6}{r^2},\hfill \\ f_{253}^{()}(\phi )=+2\frac{\phi _0\phi _4+\phi _1\phi _5\phi _6\phi _2\phi _3\phi _7}{r^2},\hfill & f_{173}^{()}(\phi )=2\frac{\phi _0\phi _5+\phi _7\phi _2+\phi _3\phi _6+\phi _1\phi _4}{r^2},\hfill \\ f_{245}^{()}(\phi )=+2\frac{\phi _3\phi _0\phi _5\phi _6+\phi _4\phi _7\phi _1\phi _2}{r^2},\hfill & f_{256}^{()}(\phi )=+2\frac{\phi _0\phi _1+\phi _3\phi _2\phi _7\phi _6\phi _4\phi _5}{r^2},\hfill \\ f_{361}^{()}(\phi )=+2\frac{\phi _0\phi _4\phi _3\phi _7\phi _1\phi _5+\phi _6\phi _2}{r^2},\hfill & f_{362}^{()}(\phi )=2\frac{\phi _0\phi _7+\phi _3\phi _4+\phi _5\phi _2+\phi _6\phi _1}{r^2},\hfill \\ f_{345}^{()}(\phi )=2\frac{\phi _2\phi _0+\phi _5\phi _7+\phi _1\phi _3+\phi _4\phi _6}{r^2},\hfill & f_{346}^{()}(\phi )=+2\frac{\phi _0\phi _1\phi _3\phi _2\phi _7\phi _6+\phi _4\phi _5}{r^2},\hfill \\ f_{367}^{()}(\phi )=+2\frac{\phi _2\phi _0\phi _4\phi _6\phi _5\phi _7+\phi _1\phi _3}{r^2},\hfill & f_{135}^{()}(\phi )=2\frac{\phi _0\phi _7+\phi _3\phi _4\phi _6\phi _1\phi _5\phi _2}{r^2},\hfill \\ f_{156}^{()}(\phi )=2\frac{\phi _2\phi _0\phi _1\phi _3\phi _4\phi _6+\phi _5\phi _7}{r^2},\hfill & f_{237}^{()}(\phi )=+2\frac{\phi _0\phi _6+\phi _2\phi _4\phi _3\phi _5\phi _1\phi _7}{r^2},\hfill \\ f_{267}^{()}(\phi )=2\frac{\phi _3\phi _0+\phi _5\phi _6\phi _4\phi _7\phi _1\phi _2}{r^2},\hfill & f_{357}^{()}(\phi )=+2\frac{\phi _0\phi _1\phi _4\phi _5+\phi _7\phi _6\phi _3\phi _2}{r^2},\hfill \\ f_{456}^{()}(\phi )=2\frac{\phi _0\phi _7\phi _5\phi _2\phi _3\phi _4+\phi _6\phi _1}{r^2},\hfill & f_{457}^{()}(\phi )=+2\frac{\phi _0\phi _6\phi _2\phi _4\phi _1\phi _7+\phi _3\phi _5}{r^2},\hfill \\ f_{467}^{()}(\phi )=2\frac{\phi _0\phi _5\phi _1\phi _4\phi _3\phi _6+\phi _7\phi _2}{r^2},\hfill & f_{567}^{()}(\phi )=+2\frac{\phi _0\phi _4+\phi _6\phi _2+\phi _1\phi _5+\phi _3\phi _7}{r^2}.\hfill \end{array}$$
warning/0002/astro-ph0002271.html
ar5iv
text
# Contributions to the Power Spectrum of Cosmic Microwave Background from Fluctuations Caused by Clusters of Galaxies ## 1 Introduction The power spectrum of the cosmic microwave radiation (CMBR) carries much cosmological information about primordial density fluctuations in the early Universe. As photons leave the last scattering surface and travel across the Universe, however, these brightness fluctuations are modified by intervening structures, causing secondary fluctuations, which we would expect to become more important on small angular scales. The power spectrum of the CMBR alone can be used to determine cosmological parameters. Recently it has been shown, however, that a geometrical degeneracy effect prevents some combinations of cosmological parameters from being disentangled by the power spectrum alone (Bond, Efstathiou & Tegmark (1997); Efstathiou & Bond (1998); Zaldarriaga, Spergel & Seljak (1997)). The primordial density fluctuations and matter content determine the positions and magnitudes of the Doppler peaks at the last scattering surface. These fluctuations are transferred to apparent angular scales determined by their angular diameter distance. As a result, cold dark matter (CDM) models with the same primordial density fluctuations, matter content, and angular diameter distance can not be distinguished. The models are “effectively degenerate” in the sense that their power spectrum is degenerate for parameter determination on intermediate and small scales. Although the observed power spectrum also depends on the time variation of the metric, via the integrated Sachs-Wolfe effect, this breaks the degeneracy only at large angular scales. Unfortunately observations of the power spectrum do not provide strong constraints on models at large scales, since this is where the statistics of the data are dominated by the cosmic variance due to the fact that we have only one realization of our cosmological model, the Universe itself, and we encounter a sampling problem). Therefore we can determine, for example, only combinations such as $`\mathrm{\Omega }_0h^2`$ and $`\mathrm{\Omega }_bh^2`$ (where $`\mathrm{\Omega }_0`$ and $`\mathrm{\Omega }_b`$ are the $`z=0`$ matter and baryon density parameters and $`h`$ is the dimensionless Hubble parameter). It has been noticed that gravitational lensing can break the geometric degeneracy at small angular scales, $`\mathrm{}2000`$, in such a way that the cosmological parameters can be determined separately (Metcalf & Silk (1998); Stompor & Efstathiou (1998)). The effect of gravitational lensing on the CMBR was studied by several authors (see for example Blanchard & Schneider (1987); Metcalf & Silk (1997); Seljak (1997)). Static gravitational lenses do not change a smooth CMBR, but the fluctuations get distorted by lensing. As a result, power from the acoustic peaks is transferred to small angular scales, conserving the variance of the spectrum. The amount of power transferred depends on the cosmological model, thus, in principle, we can determine separately $`\mathrm{\Omega }_0`$, $`\mathrm{\Omega }_b`$, and $`h`$. This method makes use of the small angular scale part of the power spectrum, where the amplitude of primordial fluctuations is declining and secondary fluctuations are becoming more important. The question naturally arises: How do contributions to the power spectrum from secondary fluctuations influence parameter determination based on the small scale CMBR power spectrum? The most important secondary fluctuations are the thermal static and kinematic Sunyaev-Zel’dovich (SSZ and KSZ) effects (Sunyaev & Zel’dovich (1980)), the Rees-Sciama (RS) effect (Rees & Sciama (1968)), the moving cluster of galaxies (MCG) effect (Birkinshaw & Gull (1983); Gurvits & Mitrofanov (1986) ; Pyne & Birkinshaw (1993)), point sources (Toffolatti et al. (1999)), and, if the Universe was re-ionized at some early stage, the Ostriker-Vishniac effect (Ostriker & Vishniac (1986); Vishniac (1987)). In this paper we concentrate on secondary effects caused by clusters of galaxies. Since detailed reviews are available on the SZ and the MCG effects (Rephaeli (1995); Birkinshaw (1998)), here we just summarize their major features. The thermal SZ effect is a change in the CMBR via inverse Compton scattering by electrons in the hot atmosphere of an intervening cluster of galaxies. We use the terms kinematic or static SZ effect depending on whether or not the intracluster gas possesses bulk motion. To date only the static thermal SZ effect has been detected (Birkinshaw (1998)). The MCG effect is a special type of RS effect, due to the time-varying gravitational field of a cluster of galaxies as it moves relative to the rest frame of the CMBR. Unlike the original RS effect, the MCG effect is not caused by intrinsic variation of the gravitational field, so that in the rest frame of the cluster, the photons fall into and climb out of the same gravitational field. However, in the rest frame of the cluster the CMBR is not isotropic, but has a dipole pattern, being brighter in the direction of the cluster peculiar velocity vector. Photons passing the cluster are deflected towards its center. Thus in the direction of the cluster peculiar velocity vector (ahead of the cluster) one can see a cooler part of the dipole pattern. Towards the tail of the cluster, one can see a brighter part of the dipole (ahead of the cluster). When transferring back to the rest frame of the CMBR, we transfer the dipole out, but the fluctuations remain, showing a bipolar pattern of positive and negative peaks. At cluster center there is no deflection, thus there is no effect. The amplitude of the MCG effect is proportional to the product of the gravitational deflection angle and the peculiar velocity of the cluster. The most important characteristics of the SSZ, KSZ, and MCG effects in the context of cosmology is that their amplitudes do not depend on the redshift of the clusters causing the effect. Using thermodynamic temperature units, their maximum amplitude are about 500 $`\mu `$K, 20 $`\mu `$K, and 10 $`\mu `$K, respectively. The SSZ and KSZ effects have the same spatial dependence as the line of sight optical depth, the MCG effect has a unique bipolar pattern. Assuming a King approximation for the total mass and an isothermal beta model for the intracluster gas, the full width at half maximum (FWHM) of the SSZ and KSZ effects $`(24)r_c`$, where $`r_c`$ is the core radius, depending on $`\beta `$, which is typically between 2/3 and 1. The MCG effect has a much larger spatial extent, with FWHM for each part of the bipolar distribution $`10r_c`$. The spectra of the effects are also important: the SSZ effect has a unique spectrum which changes sign from negative to positive at about 218 GHz. The KSZ and MCG effects have the same frequency dependence as the primordial fluctuations. The most important difference between the SZ effect and the MCG effect is that the SZ effect is caused by intracluster gas, the MCG effect is caused by gravitational lensing by the total mass regardless the physical nature of that mass. Therefore the SZ effect only arises from clusters with intracluster gas. Clusters can produce significant MCG effects even if devoid of intracluster gas. The effects of clusters of galaxies on the CMBR in a given cosmology have been a subject of intensive research since the late 1980s. There are several different ways of extracting information from these effects. Source counts of the SSZ effect were estimated by using the Press-Schechter mass function (PSMF) and scaling relations (Cole & Kaiser (1988); Markevitch et al. (1992); 1994; Makino & Suto (1993); Bartlett & Silk (1994), De Luca, Desert & Puget 1995, Colafrancesco et al. (1994); 1997, Suto et al. (1999)). The importance of the SSZ effect was demonstrated and it was shown that thousands of detections are expected with the next generation of satellites. Contributions to the CMBR from the RS and the KSZ effects were derived by Tuluie, Laguna & Anninos (1996) and Seljak (1996) for CDM models with zero cosmological constant. Tuluie et al. used N-body simulations and a ray-tracing technique, Seljak used N-body simulations and second order perturbation theory. Contributions from the SSZ and KSZ effects originating from large scale mass concentrations (superclusters) were studied by Persi et al. (1995). Bersanelli et al. (1996), in their extensive study of the CMBR for the Planck mission, estimated the contribution to the power spectrum from the SSZ and KSZ effects. Aghanim et al. (1998) estimated the effects of the KSZ and MCG effects on the CMBR including their contributions to the CMBR power spectrum. Aghanim et al. simulated $`12.5^{}\times 12.5^{}`$ maps with pixel size of $`1.5^{}\times 1.5^{}`$. They used the PSMF normalized to X-ray data (assuming an X-ray luminosity-mass relation). The total mass was assumed to have a Navarro-Frenk-White profile (Navarro, Frenk & White (1997)), and the intracluster gas was assumed to follow an isothermal beta model distribution. The time evolution of the electron temperature and the core radius were assumed to follow models of Bartlett & Silk (1994), which are based on self-similar models of Kaiser (1986). According to Aghanim et al. (1998)’s results, the KSZ effect is many orders of magnitude stronger than the primordial CMBR on small angular scales, and therefore the effect would prevent the use of the power spectrum to break the geometric degeneracy. Atrio-Barandela & Mucket (1998) estimated the power spectra of the SSZ effect in a standard dark matter dominated model with different lower mass cut-offs. Contributions to the power from the Ostriker-Vishniac effect in CDM models were estimated by Jaffe & Kamionkowski (1998). In this paper we estimate the contributions to the CMBR power spectrum from the SSZ, KSZ, and MCG effects on small angular scales adopting cold dark matter dominated models. In our models we assumed scale invariant primordial fluctuations with a processed spectrum having a power law form on cluster scales with a power law index of $`n_P=1.4`$ (Bahcall & Fan 1998). This maybe used as a first approximation as long as contributions from very low and/or very high mass clusters are small (cf. our discussion about mass cut offs at section 3). We use three representative models in our study: Model 1, open CDM (OCDM) model: a low density open model with $`\mathrm{\Omega }_0=0.2`$, $`\mathrm{\Lambda }=0`$, $`\sigma _8=1.2`$; Model 2, flat lambda CDM ($`\mathrm{\Lambda }`$CDM) model: a low density flat model with $`\mathrm{\Omega }_0=0.2`$, $`\mathrm{\Lambda }=0.8`$, $`\sigma _8=1.35`$; Model 3, standard CDM (SCDM) model: a flat model with $`\mathrm{\Omega }_0=1`$, $`\mathrm{\Lambda }=0`$, and $`\sigma _8=0.65`$. In Section 2 we outline our method of estimating the power spectra of secondary fluctuations caused by clusters of galaxies and discuss our normalization method for the PSMF. Sections 3 and 4 describe how we used the PSMF and the scaling relations to obtain masses and other physical parameters of clusters. In section 5 we present the spherical harmonic expansion of the SSZ, KSZ and MCG effects, and our method of estimating their power spectra. Section 6 describes our simulations to evaluate the integrals over clusters. Sections 7 and 8 present our results and discuss the differences from previous work. ## 2 Outline of the method We used the Press-Schechter Mass Function (PSMF, Press & Schechter (1974)) as a distribution function for cluster masses. We used $`n=1.4`$ as indicated by observations (Bahcall and Fan 1998, and references therein). We used observationally determined cluster abundances as a constraint on the PSMF. Where necessary, we altered the model parameters resulting from the usual top hat spherical collapse model since that model is only an approximation. In our SCDM model we changed only the overall normalization of the PSMF by multiplying it by 0.23 (a similar normalization was used by De Luca et al. (1995)). This procedure is inconsistent with the interpretation of the PSMF as a probability distribution (it does not integrate to unity), but we use results for the SCDM model only as a comparison to the other two models. In our Lambda-CDM model we multiplied the critical density threshold, $`\delta _c^0`$, obtained from the spherical collapse model (equation ), by 1.23 (which is equivalent to changing the $`\sigma _8`$ normalization) and made no other changes. Our OCDM model needed no adjustments. With these changes all three models agree well with the present day ($`z=0`$) observed mass spectrum (Figure 1; Bahcall & Cen 1993). For high masses, the first two models (OCDM, and $`\mathrm{\Lambda }`$CDM) also agree with the observed $`z`$ dependence of the high mass cumulative mass function (Figure 2; Bahcall, Fan & Cen 1997). The $`\mathrm{\Lambda }`$CDM and the SCDM models agree with CMBR constraints, while the OCDM model is rejected by these constraints (Lineweaver & Barbosa (1998)). As a cautionary note, it is useful to keep it in mind that taking all data into account none of these models are acceptable. We assumed that the total mass distribution follows a truncated King profile (King (1962)). For the intra-cluster gas we assumed an isothermal beta model (Cavaliere & Fusco-Femiano 1976). Isothermal beta model fits to X-ray images of clusters give $`\beta 2/3`$ (Jones & Forman (1984)). Determinations of the $`\beta `$ parameter based on spectroscopy suggest $`\beta 1`$ (Girardi et al. (1996); Lubin & Bahcall (1993)). Numerical simulations imply a range for $`\beta `$ from 1 to about 1.3 ($`\beta =1.05`$: Evrard, Metzler & Navarro (1996)); $`\beta 1.3`$: Bryan & Norman (1998); Frenk et al. (1999) suggest $`\beta =1.17`$). We follow the precepts of Eke et al. (1998) and adopt $`\beta =1`$. This choice provides a mass temperature function which is in a good agreement with the observed function (Horner et al. (1999)). The fitted X-ray spatial distribution is highly dependent on the X-ray structure of the core, and may be expected to be less reliable in the outer regions of clusters. The SZ effect is more sensitive to the outer regions (the SZ effect is proportional to the electron density as opposed to thermal bremsstrahlung, which is proportional to density squared). Choosing the spectroscopically derived $`\beta =1`$ gives a smaller SZ effect, and so our choice of $`\beta `$ should provide a conservative estimate of the contribution of the SZ effect to the power spectrum. A slightly larger $`\beta `$ would not change our results significantly. Although the beta model describes the inner intra-cluster gas well, for more accurate SZ work an improved cluster model, which fits the outer regions better, will be needed. The other physical parameters were determined using the virial theorem, a spherical collapse model, and models of the intra-cluster gas by Colafrancesco & Vittorio (1994). We assumed a Maxwellian distribution for the peculiar velocities, $`v_{pec}`$, and used results of N-body simulations by Gramann et al. (1995) to normalize the distribution. We took velocity bias into account, and assumed that the peculiar velocities are isotropically distributed. We used analytical approximations to calculate the contributions of the SSZ, KSZ and MCG effects to the CMBR power spectrum. These contributions are important only at small angular scales, where we can neglect the overlap between cluster images and ignore the weak cluster-cluster correlation, and therefore we can approximate the resulting power spectrum by summing the contributions from individual clusters (similar methods were used by Cole & Kaiser (1988); Bartlett & Silk (1994); and Atrio-Barandela & Mucket (1998)). We expanded the SSZ, KSZ, and MCG effects as Laplace series (i.e., series of spherical harmonics), then determined the individual cluster contributions and summed over the clusters (for a detailed description see Molnar (1998)). Our approximation breaks down at large angular scales, but at these scales primordial fluctuations dominate the CMBR, and the cluster contribution is only a minor perturbation, so that only a rough indication of the cluster effect is needed. ## 3 Total Mass Distribution According to the Press-Schechter method, the co-moving number density of clusters of total mass $`M`$ at redshift $`z`$ (the PSMF) is given by $$\frac{dn_c(M,z)}{d\mathrm{ln}M}=\sqrt{\frac{2}{\pi }}\frac{\mathrm{\Omega }_0\rho _c}{M}\nu (M,z)\left(\frac{d\mathrm{ln}\sigma }{d\mathrm{ln}M}\right)e^{\nu (M,z)^2/2},$$ (1) where $`\mathrm{\Omega }_0`$ is the matter density today in units of the critical density, $`\rho _c=1.88\times 10^{29}\mathrm{g}\mathrm{cm}^3`$ is the current critical density of the universe (we adopt a dimensionless Hubble parameter $`h_{100}=0.5`$ in our work); $`\nu (M,z)=\delta _c^0(\mathrm{\Omega }_0,z)/\sigma (M)`$, where the present mass variance for a power law power spectrum, $`P(k)k^n`$, is $$\sigma (M)=\sigma _8\left(\frac{M}{M_8}\right)^\alpha ,$$ (2) $`\alpha =(n+3)/6`$, $`M_8=6\times 10^{14}\mathrm{\Omega }_0h^1M_{}`$ is the mass within an $`R_8=8h^1`$ Mpc sphere, and $`\sigma _8`$ is the normalization (Lacey & Cole 1993, Press & Schechter 1974). The over-density threshold linearly extrapolated to the present may be expressed as (Lacey & Cole (1993); Navarro, Frenk & White (1997)) $$\delta _c^0(\mathrm{\Omega }_0,z)=\{\begin{array}{cc}\frac{3}{20}(12\pi )^{2/3}(1+z)\hfill & \mathrm{\Omega }_0=1\text{}\mathrm{\Lambda }=0\hfill \\ \frac{3}{2}D(\mathrm{\Omega }_0,0)\left[\left(\frac{2\pi }{\mathrm{sinh}\eta \eta }\right)^{2/3}+1\right]\hfill & \mathrm{\Omega }_0<1\text{}\mathrm{\Lambda }=0\hfill \\ 0.15(12\pi )^{2/3}\mathrm{\Omega }_m^{0.0055}D_\mathrm{\Lambda }(\mathrm{\Omega }_0,0)/D_\mathrm{\Lambda }(\mathrm{\Omega }_0,z)\hfill & \mathrm{\Omega }_0<1\text{}\mathrm{\Lambda }=1\mathrm{\Omega }_0\hfill \end{array},$$ (3) where the conformal time for open models is $$\eta =\mathrm{cosh}^1\left[\frac{2}{\mathrm{\Omega }_m(z)}1\right].$$ (4) For open models with cosmological constant $`\mathrm{\Lambda }=0`$ the linear growth factor is (Peebles 1980) $$D(\mathrm{\Omega }_0,z)=1+\frac{3}{w}+\frac{3(1+w)^{1/2}}{w^{3/2}}\mathrm{ln}\left((1+w)^{1/2}w^{1/2}\right),$$ (5) where $`w(\mathrm{\Omega }_0,z)=\mathrm{\Omega }_m^1(z)1`$, and the density parameter $`\mathrm{\Omega }_m(z)`$ is $$\mathrm{\Omega }_m(z)=\frac{\mathrm{\Omega }_0(1+z)}{\mathrm{\Omega }_0(1+z)+(1\mathrm{\Omega }_0)}.$$ (6) For models with a non-zero cosmological constant the integral can not be done analytically, thus we use an approximation (Lahav et al 1991; Caroll, Press & Turner 1992) to obtain the equivalent expression to (5), $$D_\mathrm{\Lambda }(\mathrm{\Omega }_0,z)=(1+z)^1\frac{5\mathrm{\Omega }_f(z)}{2}\left\{\mathrm{\Omega }_f(z)^{4/7}\mathrm{\Omega }_\mathrm{\Lambda }(z)+\left[1+\frac{\mathrm{\Omega }_f(z)}{2}\right]\left[1+\frac{\mathrm{\Omega }_\mathrm{\Lambda }(z)}{70}\right]\right\}^1.$$ (7) The density parameters, $`\mathrm{\Omega }_f(z)`$ and $`\mathrm{\Omega }_\mathrm{\Lambda }(z)`$, for spatially flat Universes with $`\mathrm{\Lambda }=1\mathrm{\Omega }_0`$, are $`\mathrm{\Omega }_f(z)`$ $`={\displaystyle \frac{\mathrm{\Omega }_0(1+z)^3}{\mathrm{\Omega }_0(1+z)^3+1\mathrm{\Omega }_0}}`$ (8) $`\mathrm{\Omega }_\mathrm{\Lambda }(z)`$ $`={\displaystyle \frac{1\mathrm{\Omega }_0}{\mathrm{\Omega }_0(1+z)^3+1\mathrm{\Omega }_0}}.`$ (9) The normalization of these growth functions is chosen so that at high redshifts they approximately match the time variation of density contrast in an Einstein-de Sitter ($`\mathrm{\Omega }_0=1`$) Universe, which is a good approximation to the early Universe whatever its density parameter today. The total number of clusters at redshift $`z`$ (in a redshift interval of $`dz`$) is $$N(z)dz=_{M_{low}}^{M_{up}}\frac{dn_c(M,z)}{dM}\frac{dV}{dz}𝑑M𝑑z,$$ (10) where $`M_{low}`$ and $`M_{up}`$ are the lower and upper mass cut offs for clusters. We used $`M_{low}=10^{13}M_{}`$ for the SSZ and KSZ effects, $`M_{low}=10^{12}M_{}`$ for the MCG effect, and $`M_{up}=1\times 10^{16}M_{}`$ for all effects. The lower cut off, $`M_{low}`$, for SZ effects signifies the lowest cluster mass for which we expect a well-developed intracluster atmosphere. In the case of the MCG effect, $`M_{low}`$ is the mass limit from which we consider a mass concentration as a cluster (“formation”). We found that low mass clusters do not contribute substantially to the power spectrum, thus the lower cut off, $`M_{low}`$, has little effect on our results. The upper cut off, $`M_{up}`$, has no effect on our results (as long as it is large enough): the probability of getting such a large cluster is negligible, so that the contribution from more massive clusters is negligible. ## 4 Other physical parameters of clusters of galaxies We assumed a truncated King profile for the total mass distribution $$\rho (r)=\{\begin{array}{cc}\rho _0\left(1+\frac{r^2}{r_c^2}\right)^{\frac{3}{2}}\hfill & r<pr_c\hfill \\ 0\hfill & rpr_c\hfill \end{array}$$ (11) where $`\mathrm{r}_\mathrm{c}`$ is the core radius, $`pr_c`$ is the cut off, and an isothermal $`\beta `$ model for the intra-cluster gas $$n(r)=\{\begin{array}{cc}n_0\left(1+\frac{r^2}{r_c^2}\right)^{\frac{3\beta }{2}}\hfill & r<pr_c\hfill \\ 0\hfill & rpr_c\hfill \end{array},$$ (12) where $`n(r)`$ and $`n_0`$ are the electron number density at radius $`r`$ and at the center of the cluster (Cavaliere & Fusco-Femiano 1976). Analytical studies and numerical simulations show that the gas density profile scales with the total density, and that the gas central electron density may be expressed as $$n_0=f_g\frac{2\rho _0}{m_p(1+X)},$$ (13) where $`X=0.69`$ is the average Hydrogen mass fraction, $`f_g`$ is the intra-cluster gas mass fraction. $`\rho _0`$, the central mass density, is determined from the total mass by integrating equation (11). Little is known about the total mass and redshift dependence of the intra-cluster gas from observations. Here, we adopt Colafrancesco & Vittorio (1994)’s model which assumes that changes in the intra-cluster gas are driven by entropy variation and/or shock compression and heating. According to their model, the gas mass fraction may be expressed as $$f_g=f_{g0}\left(\frac{M}{10^{15}h^1M_{}}\right)^\eta (1+z)^s,$$ (14) where the normalization, $`f_{g0}=0.1`$, is based on local rich clusters, and we used $`\eta =0.5`$ and $`s=1`$, which are consistent with available data. Using the virial radius to express the core radius, $`r_c=R_v/p`$, and assuming spherical collapse, we obtain $$r_c(\mathrm{\Omega }_0,M,z)=\frac{1.69h^1\mathrm{Mpc}}{p}\left[\left(\frac{M}{10^{15}h^1M_{}}\right)\frac{178}{\mathrm{\Omega }_0\mathrm{\Delta }_c(\mathrm{\Omega }_0,z)}\right]^{1/3}\frac{1}{1+z},$$ (15) where $`\mathrm{\Delta }_c(z)\rho _v(z)/\rho _b(z)`$ is the overdensity of the cluster relative to the background (Colafrancesco & Blasi (1998)). For $`\mathrm{\Lambda }=0`$ models (our OCDM and SCDM models) the over density may be expressed as $$\mathrm{\Delta }_c(\mathrm{\Omega },z)=4\pi ^2Q^2\left[(Q^2+2Q)^{1/2}\mathrm{ln}\left(1+Q+(Q^2+2Q)^{1/2}\right)\right],$$ (16) where $`Q=2(1\mathrm{\Omega }_0)/(\mathrm{\Omega }_0(1+z))`$ (Oukbir & Blanchard 1997). For spatially flat models with finite cosmological constant (our $`\mathrm{\Lambda }`$CDM model) we have $$\mathrm{\Delta }_c(\mathrm{\Omega },z)=18\pi ^2\left[1+0.4093\left(\mathrm{\Omega }_f(z)^11\right)^{0.9052}\right],$$ (17) where we used the approximation of Kitayama & Suto (1996). Numerical models of cluster formation show that cluster temperature scales with total mass. Using the virial theorem and assuming spherical collapse with a recent-formation approximation in a standard CDM model, the electron temperature, $`T_e`$, becomes $$kT_e=7.76\beta ^1\left(\frac{M}{10^{15}h^1M_{}}\right)^{2/3}(1+z)\mathrm{keV},$$ (18) where $`\mathrm{\Delta }_c`$ is the density contrast of a spherical top-hat perturbation relative to the background density just after virialization (cf. for example Eke, Cole & Frenk 1996). The recent-formation approximation, however is valid only for $`\mathrm{\Omega }_0=1`$. For our low matter density open model (Model 1, OCDM), we use a model which takes into account accretion during the evolution of clusters, and leads to the following scaling: $$kT_e=2.76\beta ^1\frac{1\mathrm{\Omega }_0}{\mathrm{\Omega }_0^{2/3}}\left(\frac{M}{10^{15}h^1M_{}}\right)^{2/3}\left[\left(\frac{2\pi }{\mathrm{sinh}\eta \eta }\right)^{2/3}+\frac{n_P+3}{5}\right]\mathrm{keV}.$$ (19) This was derived for open models with zero cosmological constant (Voit & Donahue (1998)), but since structure formation evolves similarly in a low density model with the same matter density and a zero cosmological constant, we use it for our Model 2 ($`\mathrm{\Lambda }`$CDM) as an approximation. We assumed a Maxwellian for the cluster peculiar velocity distribution, $`v_{pec}`$, as expected from a Gaussian initial density field: $$P(v_{pec},z)dv_{pec}v_{pec}^2\mathrm{exp}\left\{v_{pec}^2/2\sigma _p(z)^2\right\}dv_{pec},$$ (20) where $`\sigma _p(z)`$ is the Maxwellian width of the peculiar velocity distribution. The rms peculiar velocity from linear theory, smoothed with a top-hat window function of radius $`R`$, $`W_R`$, is given by $$v^2_R(z)=H^2(z)a^2(z)f^2(\mathrm{\Omega }_0,\mathrm{\Lambda })\sigma _1(R),$$ (21) where $`a(z)`$ is the scale factor, and the moments, $`\sigma _j(R)`$, are defined as $$\sigma _j(R)=\frac{1}{2\pi ^2}_0^{\mathrm{}}k^{2j+2}P(k)W(kR)𝑑k,$$ (22) where $`P(k)`$ is the Fourier transform of the power spectrum and equation (21) uses the moment $`j=1`$ (Peebles (1980)). The velocity factor, $`f(z)d\mathrm{ln}\delta /d\mathrm{ln}a`$, can be approximated as (Peebles (1980); 1984) $$f(z)\{\begin{array}{cc}\mathrm{\Omega }_m^{0.6}(z)\hfill & \mathrm{\Lambda }=0\hfill \\ \mathrm{\Omega }_f(z)\left[\frac{5}{2(1+z)D_\mathrm{\Lambda }(\mathrm{\Omega }_0,z)}\frac{3}{2}\right]\hfill & \mathrm{\Lambda }=1\mathrm{\Omega }_0\hfill \end{array}.$$ (23) The cluster peculiar velocity rms differs from this since we assume that clusters form at the peaks of the density distribution, and with this bias may be expressed as $$v_p^2_R(z)=v^2_R(z)\left[1\frac{\sigma _0^4(R)}{\sigma _1^2(R)\sigma _1^2(R)}\right]$$ (24) (Bardeen et al 1986). Colberg et al. (1998) calculated the velocity bias in a series of CDM models using a top-hat filter and processed CDM power spectra. The correction factor has a weak dependence on $`\mathrm{\Omega }_0`$: it is about 0.8 for low density and flat CDM models. We obtain the Maxwellian width in equation (20) from the rms peculiar velocity from averaging a Maxwellian: $$v_p^2=\frac{_0^{\mathrm{}}v^4\mathrm{exp}\{v_{pec}^2/2\sigma _p(z)^2\}𝑑v}{_0^{\mathrm{}}v^2\mathrm{exp}\{v_{pec}^2/2\sigma _p(z)^2\}𝑑v}=3\sigma _p^2.$$ (25) We expressed $`\sigma _p`$ as $`\sigma _p=norm\times [H(z)a(z)f(z))]/[H(0)a(0)f(0)]`$. The normalization at $`z=0`$ was determined by using results on the peculiar velocity distribution from numerical simulations (Gramann et al. (1995)). Thus we obtain the following expression for the Maxwellian width of the peculiar velocities, $`\sigma _p(\mathrm{\Omega }_0,\mathrm{\Lambda },z)`$, with velocity bias for models with no cosmological constant (OCDM, SCDM, $`\mathrm{\Lambda }=0`$): $$\sigma _p(\mathrm{\Omega }_0,0,z)=(410\mathrm{km}\mathrm{s}^1)\mathrm{\Omega }_m^{0.6}(z)\left(\mathrm{\Omega }_0(1+z)+1\mathrm{\Omega }_0\right)^{1/2}.$$ (26) For our $`\mathrm{\Lambda }`$CDM model ($`\mathrm{\Lambda }=1\mathrm{\Omega }_0`$) we obtain $`\sigma _p(\mathrm{\Omega }_0,1\mathrm{\Omega }_0,z)`$ $`=`$ $`410\mathrm{km}\mathrm{s}^1{\displaystyle \frac{1+z}{(\mathrm{\Omega }_0(1+z)^2+1\mathrm{\Omega }_0)^{1/2}}}{\displaystyle \frac{D_\mathrm{\Lambda }(\mathrm{\Omega }_0,0)}{D_\mathrm{\Lambda }(\mathrm{\Omega }_0,z)}}\times `$ (27) $`\times `$ $`\left[{\displaystyle \frac{53(1+z)D_\mathrm{\Lambda }(\mathrm{\Omega }_0,z)}{53D_\mathrm{\Lambda }(\mathrm{\Omega }_0,0)}}\right],`$ This normalization is significantly larger than some recent measurements suggest (Bahcall & Oh 1996), but it is a good match to others (Gramann 1998). This uncertainty should be remembered when interpreting our final results. ## 5 Power Spectra of SSZ, KSZ and MCG Effects Ignoring the correlation between clusters, the power spectrum becomes $$C_{\mathrm{}}^X=𝑑z𝑑M\frac{dn_c(M,z)}{dM}G_{\mathrm{}}^X\frac{dV}{dz},$$ (28) where $`G_{\mathrm{}}^X`$ is the contribution from clusters with total mass $`M`$ at redshift $`z`$, and $`X`$ denotes the SSZ, KSZ or MCG effects. $`dV/dz`$ is the differential volume element (assumed isotropy) $$\frac{dV}{dz}=r(z)^2\frac{4\pi c}{H_0}\left[\mathrm{\Omega }_0(1+z)^3+(1\mathrm{\Omega }_0\mathrm{\Lambda })(1+z)^2+\mathrm{\Lambda }\right]^{1/2},$$ (29) where the effective distance $`r(z)`$ is $$r(z)=\{\begin{array}{cc}\frac{2c}{H_0}\left[\frac{\mathrm{\Omega }_0z+(\mathrm{\Omega }_02)(\sqrt{1+\mathrm{\Omega }_0z}1)}{\mathrm{\Omega }_0^2(1+z)}\right]\hfill & \mathrm{\Lambda }=0\hfill \\ \frac{c}{H_0}_0^z𝑑x\left[\mathrm{\Omega }_0(1+x)^3+1\mathrm{\Omega }_0\right]^{1/2}\hfill & \mathrm{\Lambda }=1\mathrm{\Omega }_0\hfill \end{array}$$ (30) (Peebles 1993). In general, the coefficients $`G_{\mathrm{}}`$ may be determined by calculating the spherical harmonic expansion of the cluster image by averaging out the azimuthal parameter, $`m`$, $$G_{\mathrm{}}=\frac{1}{2\mathrm{}+1}\underset{m=\mathrm{}}{\overset{\mathrm{}}{}}|a_\mathrm{}m|^2.$$ (31) Our task is to determine the $`a_\mathrm{}m`$ coefficients. The SSZ and KSZ effects are cylindrically symmetric for spherical clusters, therefore we may describe them using only one coordinate, the angular distance from the cluster center. We separate the effects into amplitudes and geometrical form factors which carry their spatial dependence. The SSZ and KSZ effects in thermodynamic temperature units may be expressed as $`\left({\displaystyle \frac{\mathrm{\Delta }T}{T}}\right)_S(x,\theta )\mathrm{\Delta }_S`$ $`=`$ $`\mathrm{\Delta }_S^0(x)\zeta (\theta )`$ (32) $`\left({\displaystyle \frac{\mathrm{\Delta }T}{T}}\right)_K(x,\theta )\mathrm{\Delta }_K`$ $`=`$ $`\mathrm{\Delta }_K^0(x)\zeta (\theta ),`$ (33) where the central effects for the SSZ and KSZ effects are $$\mathrm{\Delta }_S^0(x,\mathrm{\Theta })=\left[Y_0(x)+\mathrm{\Theta }Y_1(x)+\mathrm{\Theta }^2Y_2(x)+\mathrm{\Theta }^3Y_3(x)+\mathrm{\Theta }^4Y_4(x)\right]\mathrm{\Theta }\tau _0,$$ (34) and $`\mathrm{\Delta }_K^0(x,\mathrm{\Theta })`$ $`=`$ $`\{[{\displaystyle \frac{1}{3}}Y_0+\mathrm{\Theta }({\displaystyle \frac{5}{6}}Y_0+{\displaystyle \frac{2}{3}}Y_1)]\beta ^2[1+\mathrm{\Theta }C_1(x)+\mathrm{\Theta }^2C_2(x)]\beta P_1(\alpha )`$ $`+`$ $`[D_0(x)+\mathrm{\Theta }D_1(x)]\beta ^2P_2(\alpha )\}\tau _0.`$ In these expressions $`P_k`$ is the Legendre polynomial of order of $`k`$, $`x=h\nu /k_BT_{CB}`$ is the dimensionless frequency, $`\mathrm{\Theta }=k_BT_e/(m_ec^2)`$ is the dimensionless temperature, $`\alpha `$ is the angle between the cluster’s peculiar velocity vector and its position vector, $`h`$, $`\nu `$, $`k_B`$ and $`T_{CB}`$ are the Planck constant, frequency, Boltzmann constant, and temperature of the CMBR, $`T_{CB}=2.728\pm 0.002`$ K (Fixsen et al. (1996)), and the lengthy expressions for the spectral functions $`Y_i(x)`$, $`C_i(x)`$ and $`D_i(x)`$ may be found in Nozawa, Itoh & Kohyama (1998). These functions arise from an expansion of the Boltzmann equation and although they are inaccurate for high temperature clusters, their precision is sufficient for our purposes here (for a discussion see Molnar & Birkinshaw (1999) and references therein). The optical depth through the cluster center for gas model (12) is $$\tau _0=2\sigma _Tn_0r_cI_p(3\beta /2,p),$$ (36) and the geometrical form factor is $$\zeta (\theta )=(n_e/n_0)𝑑z=\left(1+d_c^2\right)^{\frac{1}{2}\frac{3\beta }{2}}j(\beta ,p,d_c),$$ (37) where the function $`j(\beta ,p,d_c)`$ is defined as $$j(\beta ,p,d_c)=\frac{I_p(3\beta /2,\sqrt{p^2d_c^2})}{I_p(3\beta /2,p)},$$ (38) $`d_c=\theta /\theta _c`$ holds in the small angle approximation, and the integral, $`I_p(\alpha ,q)`$, is $$I_p(\alpha ,q)=\frac{\sqrt{\pi }}{2}\frac{\mathrm{\Gamma }(\alpha \frac{1}{2})}{\mathrm{\Gamma }(\alpha )}(q+1)^\alpha (q1)^{1\alpha }\frac{\mathrm{\Gamma }(2\alpha 1)}{\mathrm{\Gamma }(2\alpha )}F(\alpha ;1;2\alpha ;2q),$$ (39) where $`\mathrm{\Gamma }`$ is the gamma function, $`F(x;y;w;z)`$ is Gauss’ hyper-geometric function, and $`\alpha `$ must be greater than 1/2 (Gradshteyn & Ryzhik (1980)). The geometrical form factor is normalized to one at the cluster center ($`\zeta (0)=1`$). We may expand the SSZ and KSZ effects in Legendre series as $$\mathrm{\Delta }_{SZ}=\mathrm{\Delta }_{SZ}^0\underset{\mathrm{}=0}{\overset{\mathrm{}}{}}\frac{2\mathrm{}+1}{4\pi }\zeta _{\mathrm{}}P_{\mathrm{}},$$ (40) where $`\zeta _{\mathrm{}}`$ are Legendre coefficients of $`\zeta `$, and $`SZ`$ refers to the SSZ or the KSZ effect. We determine the Legendre coefficients using a small angle approximation, as $$\zeta _{\mathrm{}}=2\pi _0^{p\theta _c}\zeta (\theta )J_0[(\mathrm{}+\frac{1}{2})\theta ]\theta 𝑑\theta ,$$ (41) where we used the approximation $$P_{\mathrm{}}J_0[(\mathrm{}+\frac{1}{2})\theta ],$$ (42) where $`J_0`$ is a Bessel function of the first kind and zero order (Peebles (1980)). We can convert Legendre coefficients to Laplace coefficients by expressing the Laplace series of such a function as $$\underset{\mathrm{}}{}a_\mathrm{}0Y_{\mathrm{}}^0(\theta ,\phi )=\underset{\mathrm{}}{}a_\mathrm{}0\sqrt{\frac{2\mathrm{}+1}{4\pi }}P_{\mathrm{}}(\mu )=\underset{\mathrm{}}{}\frac{2\mathrm{}+1}{4\pi }b_{\mathrm{}}P_{\mathrm{}}(\mu ),$$ (43) where $`\mu =\mathrm{cos}\theta `$, and $`Y_{\mathrm{}}^m`$ and $`P_{\mathrm{}}`$ are the spherical harmonics and Legendre polynomials. Therefore the conversion can be done as $$a_\mathrm{}0=\left(\frac{2\mathrm{}+1}{4\pi }\right)^{1/2}b_{\mathrm{}}.$$ (44) Thus the Laplace series of the SSZ and KSZ effects become $$\mathrm{\Delta }_{SZ}(\theta )=\mathrm{\Delta }_{SZ}^0\underset{\mathrm{}}{}\left(\frac{2\mathrm{}+1}{4\pi }\right)^{1/2}\zeta _{\mathrm{}}Y_{\mathrm{}}^0(\theta ).$$ (45) Using equation (45), the contribution of one cluster to the power spectrum of the SSZ and KSZ effects becomes $$G_{\mathrm{}}^{SZ}=\frac{1}{4\pi }(\mathrm{\Delta }_{SZ}^0)^2\zeta _{\mathrm{}}^2,$$ (46) where $`\mathrm{\Delta }_{SZ}^0`$ and $`\zeta _{\mathrm{}}`$ are given by equations (34), (5), and (41). Similarly, the MCG effect may be expressed as $$\left(\frac{\mathrm{\Delta }T}{T}\right)_M(x,\theta ,\phi )\mathrm{\Delta }_M=\mathrm{\Delta }_M^{max}\xi (\theta ,\phi ),$$ (47) where $`\xi (\theta ,\phi )`$ is the geometrical form factor. $`\theta `$ is the angle of the line of sight relative to the center of the cluster. The azimuthal angle, $`\phi `$, is measured in the plane of the sky from the direction of the tangential component of the peculiar velocity. The maximum of the MCG effect is $$\mathrm{\Delta }_M^{max}=(v_p/c)\mathrm{sin}\alpha \delta _{max},$$ (48) where $`\delta _{max}`$ is the maximum deflection angle, $`\alpha `$ is the angle between the cluster’s peculiar velocity vector, $`v_p`$, and its position vector, and $`c`$ is the speed of light in vacuum. For a spherically symmetric thin lens the deflection angle is given by $$\delta (d)=\frac{4G}{c^2d}(d),$$ (49) where $`d`$ is the impact parameter at the source, and $``$ is the mass enclosed by a cylindrical volume with axis parallel to the line of sight and radius equal to the impact parameter $`d`$ (cf. for example Schneider, Ehlers & Falco 1992). Using the King approximation for the density distribution (equation ), the total mass in cylindrical coordinates, $`(R,\psi ,z)`$, becomes $$=\rho _0r_c^3_0^{2\pi }𝑑\psi _{d_c}^{d_c}𝑑z_0^p\frac{RdR}{\left(1+R^2+z^2\right)^{\frac{3}{2}}},$$ (50) where $`d_cd/r_c\theta /\theta _c`$ in the small angle approximation. A straightforward integration and equation (49) lead to $$\delta (d_c,p)=\frac{4G}{c^2r_cg(p,p)}\frac{g(d_c,p)}{d_c},$$ (51) where the function $`g(x,p)`$ is $$g(x,p)=\mathrm{ln}(1+x^2)+\mathrm{ln}\left[\frac{p+\sqrt{1+p^2}}{p+\sqrt{1+p^2+x^2}}\right].$$ (52) Thus the geometrical form factor in our case becomes $$\xi (\theta ,\phi )=\frac{g(d_c,p)}{b_md_c}\mathrm{cos}\phi ,$$ (53) where $`b_m`$ is the maximum value of the function $`g(x,p)/x`$. The MCG effect depends only on $`\mathrm{cos}\phi `$, therefore we need to determine only the $`m=\pm 1`$ terms in the spherical harmonic expansion. Expressing the spherical harmonics by associated Legendre polynomials, equation (53) expands as $$\xi (\theta ,\phi )=1\frac{2\mathrm{}+1}{4\pi }\frac{(\mathrm{}1)!}{(\mathrm{}+1)}\underset{\mathrm{}}{}P_{\mathrm{}}^1\left[\xi _{\mathrm{}}^1e^{i\phi }\xi _{\mathrm{}}^1e^{i\phi }\right],$$ (54) where we used the identity $$P_{\mathrm{}}^m=(1)^m\frac{(\mathrm{}m)!}{(\mathrm{}+m)}P_{\mathrm{}}^m.$$ (55) In order to obtain a real function, the imaginary terms must vanish, therefore we must have $$\xi _{\mathrm{}}^1=\xi _{\mathrm{}}^1.$$ (56) Using orthogonality, expressing the spherical harmonics in terms of associated Legendre polynomials and using equations (56) and (54), we obtain $$\xi _{\mathrm{}}^1=\frac{k_{\mathrm{}}}{b_{max}}_{\mathrm{\Omega }_{\widehat{x}^{}}}𝑑\mathrm{\Omega }_{\widehat{x}^{}}\frac{g(d_c,p)}{d_c}P_{\mathrm{}}^1(\mathrm{cos}\theta )\mathrm{cos}^2\phi ,$$ (57) where $$k_{\mathrm{}}=\sqrt{\frac{2\mathrm{}+1}{4\pi }}\sqrt{\frac{(\mathrm{}1)!}{(\mathrm{}+1)!}}.$$ (58) The $`\phi `$ integral can be performed since $`g(x,p)`$ and $`P_{\mathrm{}}^1`$ do not depend on $`\phi `$ giving $$\xi _{\mathrm{}}^1=\frac{\pi k_{\mathrm{}}}{b_{max}}_1^1𝑑\mu \frac{g(d_c,p)}{d_c}P_{\mathrm{}}^1(\mu ),$$ (59) from which we find $$\xi _{\mathrm{}}^1=\frac{\pi k_{\mathrm{}}}{b_{max}}(\mathrm{}+1/2)\theta _c_0^{p\theta _c}g(\theta ,p)J_1[(\mathrm{}+\frac{1}{2})\theta ]𝑑\theta .$$ (60) Here we used the small angle approximation for the associated Legendre polynomials: $$P_{\mathrm{}}^1(\mu )(\mathrm{}+1/2)J_1[(\mathrm{}+\frac{1}{2})\theta ],$$ (61) where $`J_1`$ is the Bessel function of the first kind and order 1 (for a derivation see Appendix). Thus the Laplace series of the MCG effect is $$\mathrm{\Delta }_M=\mathrm{\Delta }_M^{max}\underset{\mathrm{}}{}\xi _{\mathrm{}}^1\left(Y_{\mathrm{}}^1Y_{\mathrm{}}^1\right),$$ (62) where $`\xi _{\mathrm{}}^1`$ is given by equation (60). For the power spectrum of the MCG effect, using equation (62), we obtain $$G_{\mathrm{}}^M=\frac{2}{2\mathrm{}+1}\left(\mathrm{\Delta }_M^{max}\right)^2|\xi _{\mathrm{}}^1|^2.$$ (63) The observed effects are calculated by convolving the theoretical fluctuation pattern with the telescope’s point spread function (PSF). One advantage of using the spherical harmonic coefficients is that this convolution is just a multiplication in spherical harmonic space. Assuming an axially symmetric PSF, $`R`$, its Legendre polynomial expansion may be expressed as $$R(\widehat{x}\widehat{x}^{})=\underset{\mathrm{}}{}\frac{2\mathrm{}+1}{4\pi }R_{\mathrm{}}P_{\mathrm{}}(\widehat{x}\widehat{x}^{}),$$ (64) where the unit vectors, $`\widehat{x}`$ and $`\widehat{x}^{}`$, point to an arbitrary direction (where we want to evaluate the expansion) and to the center of the PSF. Assuming a non-axially symmetric effect, its spherical harmonic expansion can be written as $$f(\widehat{x})=\underset{\mathrm{},m}{}f_{\mathrm{}}^mY_{\mathrm{}}^m(\widehat{x}),$$ (65) where $`\mathrm{}`$ runs from zero to infinity and $`m`$ runs from -$`\mathrm{}`$ to $`\mathrm{}`$. Using the addition theorem for spherical harmonics and their orthogonality, the convolution of these two functions, $`M=Rf`$, becomes $$M(\widehat{x})=\underset{\mathrm{},m}{}R_{\mathrm{}}f_{\mathrm{}}^mY_{\mathrm{}}^m(\widehat{x}).$$ (66) ## 6 Simulation of clusters of galaxies We used Monte Carlo simulations to generate an ensemble of clusters of galaxies with masses sampled from the PSMF (equation 1) with parameters those of our OCDM, $`\mathrm{\Lambda }`$CDM and SCDM models. We obtained the central electron number density and temperature, and the cluster core radius from scaling relations (equations 13, 18, 19, and 15). We choose to sample the PSMF using a rejection method. The magnitude of the peculiar velocity may be sampled using an inversion method on the Maxwellian (20), and yields $$v_p=\sqrt{2}\sigma _p\gamma ^1(3/2,RN)^1,$$ (67) where $`\gamma ^1(order,x)`$ is the inverse of the incomplete gamma function and $`\sigma _p`$ can be determined by using equations (25), (26) and (27). $`RN`$ is a uniformly distributed random number in (0,1). We assumed an isotropic distribution in space for the directions of the peculiar velocity vectors, and ignored correlations between cluster peculiar velocities. The tangential and radial peculiar velocities are distributed as projections of equation (67) accordingly. As an illustration, in Figure 3 we show results from one simulation using our SCDM model projected on a grid. The observational mass function (Figure 1) is specified by $`M_{1.5}`$, the mass contained within co-moving radius of 1.5 Mpc. To convert $`M_{1.5}`$ to the virial mass, $`M_v`$, which we use in the PSMF, we assume that the mass profile near 1.5 Mpc can be approximated with $`M_v(R)R^q`$. We obtain $$M_v=\left(\frac{\mathrm{\Delta }_c(0)}{\mathrm{\Delta }_c\mathrm{\Omega }_m(z)}\frac{M_{1.5}}{\mathrm{\Delta }_c(0)(4\pi /3)\rho _c(0)(1.5h^1\mathrm{Mpc})^3}\right)^{\frac{q}{3q}}M_{1.5},$$ (68) or, substituting the numerical values, $$M_v=\left(\frac{178}{\mathrm{\Delta }_c\mathrm{\Omega }_m(z)}\frac{M_{1.5}}{6.98\times 10^{14}h^1M_{}}\right)^{\frac{q}{3q}}M_{1.5}.$$ (69) We used q = 0.64 (Carlberg, Yee & Ellingson (1997)) to obtain curves shown in Figure 2. The power spectrum for an ensemble of clusters may be determined by summing the individual contributions of the simulated clusters (equation ). We binned clusters by their apparent core radii, $`\theta _c`$, then we summed the amplitudes in each bin. The numerical evaluation of integral (28), in this case, can be performed as $$C_{\mathrm{}}^{SZ}=\frac{1}{4\pi }\underset{i}{}(\zeta _{\mathrm{}})_i^2\underset{cl}{}(\mathrm{\Delta }_{SZ}^0)_{cl}^2,$$ (70) and $$C_{\mathrm{}}^M=\frac{2}{2\mathrm{}+1}\underset{i}{}(\xi _{\mathrm{}}^1)_i^2\underset{cl}{}(\mathrm{\Delta }_M^{max})_{cl}^2,$$ (71) where the index $`cl`$ runs over clusters whose core radii fall within the $`i^{th}`$ bin. ## 7 Results Our results for the power spectra (more exactly the dimensionless $`𝒞(\mathrm{})\mathrm{}(\mathrm{}+1)C_{\mathrm{}}`$) from our Model 1 (OCDM) are shown on Figure 4. Figures 5 and 6 show our results for Model 2, ($`\mathrm{\Lambda }`$CDM) and Model 3, (SCDM) respectively. As a comparison, in each figure, we plot the corresponding primordial CMBR power spectrum (solid line) with $`COBE`$ normalization including the effects of gravitational lensing calculated by using a new version of CMBFAST (Zaldarriaga & Seljak (1998); Zaldarriaga, Spergel & Seljak (1997)). On large angular scales (up to about $`\mathrm{}10`$) the cosmic variance dominates (not shown). On small angular scales the shape of the power spectra depends on the apparent angular sizes of the clusters and the amplitudes of the effects. The apparent angular size depends on how the core radius and the angular diameter distance change with redshift, while the amplitude is sensitive to the gas content, gas temperature and total mass as a function of redshift. Figures 4-6 demonstrate that for small angular scales ($`\mathrm{}3000`$) the contribution to the power spectrum from the SSZ effect exceeds that of the primordial CMBR in all our models. The contributions of the KSZ and MCG effects become important only on small scales, but, at those scales, they may dominate over the lensed primordial fluctuations. Due to the early structure formation, there are more clusters at high redshift in our OCDM and $`\mathrm{\Lambda }`$CDM than in our SCDM simulations. Therefore the contribution to the power spectrum from clusters in a low matter density model is substantially larger than in a SCDM model. Also, most clusters are closer to us in a SCDM model, thus the contribution from clusters to the power spectrum peaks at higher angular scales (lower $`\mathrm{}`$) than in low matter density models. The KSZ and MCG effects have their coherence length (peak contributions) at $`\mathrm{}10000`$. The coherence length of the SSZ effect is about $`\mathrm{}1000`$ for our SCDM model and at about $`\mathrm{}2000`$ for our OCDM and $`\mathrm{\Lambda }`$CDM models. In general, the contributions to the power spectrum from the SSZ effect are about 2 and 3 orders of magnitude greater than those from the KSZ and MCG effects. At very small angular scales, $`\mathrm{}>10000`$, the contribution to the power spectrum from the MCG effect might exceed that of the SSZ or KSZ effects, and even the primordial fluctuations in the CMBR, but this depends on the details of the evolution of cluster atmospheres. Our simulations give somewhat different results for the KSZ and MCG effects than those of Aghanim et al. (1998). In our simulations the amplitudes of the KSZ and MCG effects for low matter density models are about an order of magnitude greater than those for our SCDM model, and have a coherence length of about $`\mathrm{}=10000`$, while rising monotonically at smaller $`\mathrm{}`$. According to Aghanim et al.’s simulations, with similar cut off to ours, $`M_{low}=10^{13}M_{}`$, contributions from the KSZ effect in all models constantly grow and show no signs of leveling off, and their amplitude has a very weak dependence on cosmological model. Contributions from the MCG effect on the other hand show a plateau in all Aghanim et al.’s models for $`\mathrm{}>1000`$, and for the SCDM model, the MCG effect is larger than for the other two models. Quantitatively, our models show cluster-related effects that are weaker by a factor of 10 for the MCG effect in our SCDM model and a factor of 100 for the KSZ effect for all models. We attribute these differences mostly to the different evolution models for the intracluster gas. The ratio between the overall amplitudes of the KSZ and MCG effects in our calculations is about the same as in Aghanim et al. (1998)’s results. Our results show that the power spectra of the KSZ and MCG effects do not exceed the gravitationally lensed primordial power spectrum up to $`\mathrm{}10000`$, while the power spectrum of the SSZ effect becomes dominant at $`\mathrm{}7000`$ in all our models. Contributions to the power spectrum from the SSZ and KSZ effects based on Aghanim et al.’s model would exceed those from the CMBR at $`\mathrm{}5000`$ even if one takes gravitational lensing of the primordial CMBR into account. Our results are comparable to those obtained by Tuluie et al. (1996) and Seljak (1996). Persi et al. (1995)’s results for contributions to the power spectrum from the SSZ (KSZ) effect are about the same as (an order of magnitude higher than) our results suggest. ## 8 Discussion An observed power spectrum is made up from the sum of all astrophysical effects and noise. We rely on the different frequency and/or power spectra of the secondary effects to separate these foregrounds from the primordial CMBR signal (Tegmark (1998)). Of the effects discussed here, it should be easy to separate the SSZ effect by using multi-frequency measurements of its unique spectrum. The separation of primordial fluctuations in the CMBR and fluctuations caused by the KSZ and MCG effects is more difficult since their frequency spectra are the same. Optimal filters have been designed to separate the KSZ effect (Haehnelt & Tegmark (1996); Aghanim et al. (1997)): here it helps to know the SSZ effect for the same cluster, since that would give us a position and even an estimate for the expected amplitude of the effect. Aghanim et al. (1998) discussed methods to separate the MCG effect: this is facilitated by its unique dipole pattern with sharp peaks (Figure 3). Primordial fluctuations are usually assumed to be gaussian, where the probability of getting such a strongly peaked bipolar pattern is small, and we would expect the strong small angular scale gradient near a known cluster of galaxies to be a definite indication of the presence of the MCG effect. Also, knowing the position of clusters helps to find the effect. However, contributions from other effects, such as early ionization and discrete radio sources causes further confusion, and may be expected to make it difficult to determine the power from the SZ or MCG effects. We analyzed the contributions to the power spectrum from the SSZ, KSZ and MCG effects to check their impact on the determination of cosmological parameters, especially at large $`\mathrm{}`$ where gravitational lensing may break the geometric degeneracy. In Figure 8 we show the small scale lensed primordial fluctuation power spectra of our three models (OCDM, SCDM, $`\mathrm{\Lambda }`$CDM; solid lines) with power spectra resulting from the sum of fluctuations due to the lensed primordial CMBR and the SSZ effect (long dashed lines), and from the sum of the lensed primordial CMBR, the KSZ and the MCG effects (short dashed lines). According to our models, if the fluctuations due to the SSZ effect are fully separated, the KSZ and MCG effects do not prevent the use of this part of the power spectrum to break the geometric degeneracy and distinguish between different CDM models. Note that normalization at the first Doppler peak, rather than the usual COBE normalization, would lower the contributions to the power spectrum from primordial fluctuations in a $`\mathrm{\Lambda }`$CDM model relative to those from a SCDM model, and thus secondary effects would become more important relative to the primordial CMBR fluctuations. Our simulations also show that the power spectrum of the SSZ effect may itself be used to break the geometric degeneracy. Since the separation of the SSZ effect from other secondary effects should be straightforward, we should be able to determine the power spectrum of the SSZ effect alone. As can be seen from Figure 7, this power spectrum depends on $`\mathrm{\Omega }_0h^2`$ and $`\mathrm{\Omega }_bh^2`$, providing an additional constraint on these parameters. We note, however, that the amplitude of the SSZ effect is model dependent. Since the contributions to the power spectrum from the SZ and MCG effects are model dependent, to evaluate fully their power spectra we need a better observationally-supported model for the intracluster gas. Sensitive, high-resolution all-sky, X-ray observations could map the emission from intracluster gas up to high redshift providing strong constraints on gas formation and evolution and thus a good basis for modeling the SSZ and KSZ effects (Jahoda (1997)). Number counts of clusters based on their SSZ effect can also be used to constrain cosmological models (Suto et al. (1999)). There are many possibilities of using observations to break the geometric degeneracy. For example measurements of the CMBR polarization, the Hubble constant, or light curves of Type Ia supernovae have been discussed (Zaldarriaga et al. 1997; Eisenstein, Hu & Tegmark (1998); Tegmark et al. (1998)). Also, combination of measurements of the SSZ effect and thermal bremsstrahlung (X-ray) emission from clusters can be used to determine the Hubble constant for a large number of clusters, providing a statistical sample which might enable us to determine the Hubble constant, and perhaps the acceleration parameter, to good accuracy (Birkinshaw (1998)). Secondary fluctuations introduce non-gaussianity into the primordial spectrum at small scales. This non-gaussianity should be taken into account when estimating CMBR non-gaussianity at these scales. Winitzky (1998) estimated the effect of lensing and concluded that Planck may observe non-gaussianity due to lensing near the angular scale of maximum effect, $`10^{}`$. Other processes, including the SSZ, KSZ, and especially the MCG effect, introduce a highly non-gaussian signal as is easily seen for the MCG effect on Figure 3. A similar non-gaussian pattern arises from moving cosmic strings (the Kaiser-Stebbins effect, compare our Figure 3 to Figure 6a of Magueijo & Lewin (1997)). Our results indicate that at $`\mathrm{}10^4`$ the MCG effect might be comparable in strength to the primordial fluctuations. Evidence for non-gaussianity has been reported by Ferreira, Magueijo & Gorski (1998) and Gaztanaga, Fosalba, & Elizalde (1997) at angular scales $`\mathrm{}16`$ and $`\mathrm{}150`$. They do not exclude the possibility that this non-gaussianity has been introduced by foregrounds, but our results show that clusters can not introduce detectable non-gaussianity on such scales (Figure 7). We convolved our theoretical results (Figures 4 \- 6) with the expected point spread functions (PSFs) of instruments on the MAP and Planck missions to estimate the level of the secondary fluctuations caused by clusters of galaxies on the observable power spectrum. The observed $`C_{\mathrm{}}`$ values become $$C_{\mathrm{}}^{obs}=C_{\mathrm{}}W_{\mathrm{}},$$ (72) where the $`W_{\mathrm{}}`$ values are the Legendre coefficients of the PSF. For an assumed gaussian response function and in the small angle approximation, the $`W_{\mathrm{}}`$ coefficients are $$W_{\mathrm{}}=e^{\sigma ^2(\mathrm{}+1/2)^2},$$ (73) where $`\sigma =h/2\sqrt{\mathrm{ln}4}`$, and $`h`$ is the FWHM of the beam (for a detailed description of window functions and $`W_{\mathrm{}}`$ coefficients, see White & Srednicki 1995). The observed rms fluctuations then become $$\mathrm{\Delta }T/T_0_{rms}^2=\underset{\mathrm{}}{}\frac{2\mathrm{}+1}{4\pi }C_{\mathrm{}}W_{\mathrm{}}.$$ (74) In general, contributions from unresolved cluster static effects add to provide a cumulative contribution to the CMBR power spectrum. Contributions from the KSZ and MCG effects from unresolved sources tend to cancel. In the case of the MCG effect this is because each unresolved source contribution would be zero owing to the dipole spatial pattern of the effect. For small-scale KSZ effects there are several sources in the field of view of the telescope, and different sources have positive or negative contributions depending on the sign of their line of sight peculiar velocity, and therefore they tend to cancel each other. The larger the beam size, the more effective is the cancellation of the MCG and KSZ effects. Note that the spatial extension of the MCG effect is much larger than that of the KSZ effect, so many clusters may be unresolved in their KSZ and resolved in their MCG effect. The MCG effect might be relatively more important at high redshifts, since it does not require a well-developed cluster atmosphere. In Figures 9 and 10 we show the contributions to the power spectrum from primordial fluctuations, and the SSZ, KSZ and MCG effects, convolved with the PSF of the planned $`\nu =94`$ GHz receiver on MAP, and the planned $`\nu =353`$ GHz bolometer on Planck. The amplitude of the fluctuations from the the SSZ effect is negative at $`\nu =94`$ GHz and positive at $`\nu =353`$ GHz, but only the the absolute value of the effect contributes to the power spectrum. The different maximum $`\mathrm{}`$ values for the MAP and Planck systems ($`\mathrm{}_{max}`$ 1000 and 2000, respectively) can clearly be seen on Figures 9 and 10. Because of these cutoffs, the observable power spectrum is dominated by primordial fluctuations at all $`\mathrm{}`$ for these missions. According to our results, the SSZ effect may cause a 1% enhancement in the amplitude of the Doppler peaks, which is at the limit of the sensitivity of the MAP and Planck missions. From the analysis of the power spectrum, this would lead to an overestimation of the parameter $`\mathrm{\Omega }_0h^2`$ by about 1%,. The shift in the position of peaks as a function of $`\mathrm{}`$ caused by the SSZ effect is less important since the spectrum of the SSZ effect has only a weak dependence on $`\mathrm{}`$. In Table 1 we show the $`(\mathrm{\Delta }T/T)_{rms}`$ values of the contributions to the CMBR from the SSZ, KSZ, and MCG effects convolved with the the 94 GHz MAP and 353 GHz Planck receivers for our three models (OCDM; $`\mathrm{\Lambda }`$CDM; SCDM). As a comparison, we display the corresponding rms values of the primordial fluctuations. The rms values of all these secondary effects are an order of magnitude smaller than rms values from primordial fluctuations. The most important contribution is that of the SSZ effect at these frequencies. The KSZ and MCG effects give similar contributions with the KSZ effect being about a factor of two stronger. Aghanim et al. (1998)’s results for the rms values of the MCG effect is a factor of 10 (SCDM) or a factor of 3 (OCDM and $`\mathrm{\Lambda }`$CDM) larger than our results. Note, however, that rms values give only a crude estimate of the magnitude of the effects. At large angular scales the primordial fluctuations are about 100 (for the SSZ effect) or $`10^510^7`$ (KSZ, MCG effects) times stronger than the secondary fluctuations. An ideal observation to measure the contribution to the power spectrum from the SSZ effect would use high angular resolution ($`\mathrm{}7000`$) and high frequency ($`\nu 250`$ GHz). SuZIE probes the power spectrum at angular scale $`\mathrm{}7500`$ at 140 GHz. The 2$`\sigma `$ upper limit on the power at this scale from SuZIE is $`\mathrm{}(\mathrm{}+1)C_{\mathrm{}}1.4\times 10^9`$ (Ganga et al. (1997)). Unfortunately our models suggest that at this frequency the primordial contribution to the power spectrum is about 10 times stronger than that from the SSZ effect. A promising experiment is SCUBA, which probes the anisotropies at angular scale $`\mathrm{}10000`$ and frequency 348.4 GHz. Their preliminary 2$`\sigma `$ upper limit on the power spectrum at this scale is $`\mathrm{}(\mathrm{}+1)C_{\mathrm{}}4.7\times 10^8`$. Much further work is planned, and should lower this limit by a factor of 3-10 (Borys, Chapman & Scott (1998)). SMM is grateful to Bristol University for a full scholarship, where most of this work was done. This work was finished while SMM held a National Research Council Research Associateship at NASA Goddard Space Flight Center. We thank N. Aghanim for comments on an earlier version of the manuscript, and our referee, Dr Bartlett, for his detailed comments and for helping us to clarify some aspects of our approximations. We thank U. Seljak and M. Zaldarriaga for making the CMBFAST code available. Appendix: Small Angle Approximation for $`P_{\mathrm{}}^1`$ We derive a small angle approximation for the associated Legendre polynomial $`P_{\mathrm{}}^1`$. We express the associated Legendre polynomial $`P_{\mathrm{}}^1`$ in terms of the Legendre polynomial $`P_{\mathrm{}}`$ (see for example Arfken & Weber (1995)) as $$P_{\mathrm{}}^1(\mu )=\sqrt{1\mu ^2}\frac{d}{d\mu }P_{\mathrm{}}(\mu ).$$ (1) Introducing a new variable, $`x=\mathrm{}(\mathrm{}+1)\theta `$, using the chain rule and a small angle approximation, we get $$P_{\mathrm{}}^1(\mu )=\sqrt{1\mu ^2}\frac{dP_{\mathrm{}}}{dx}\frac{dx}{d\theta }\frac{d\theta }{d\mu }\mathrm{}(\mathrm{}+1)\frac{dP_{\mathrm{}}}{dx}.$$ (2) Using a small angle approximation for $`P_{\mathrm{}}`$ (Peebles 1980) we obtain $$\mathrm{}(\mathrm{}+1)\frac{dP_{\mathrm{}}}{dx}\mathrm{}(\mathrm{}+1)\frac{d}{dx}J_0(x).$$ (3) Using the relation $$J_1(x)=\frac{d}{dx}J_0(x)$$ (4) for Bessel functions, we finally obtain $$P_{\mathrm{}}^1(\mu )\mathrm{}(\mathrm{}+1)\frac{d}{dx}J_0(x)=\mathrm{}(\mathrm{}+1)J_1(x).$$ (5)
warning/0002/nucl-th0002007.html
ar5iv
text
# Interrelation between the isoscalar octupole phonon and the proton-neutron mixed-symmetry quadrupole phonon in near spherical nuclei ## I Introduction The phonon concept is a useful and simple concept in nuclear structure physics . The low-lying collective $`J^\pi =2^+`$ and $`3^{}`$ excitations in near-spherical nuclei can be considered as quadrupole and octupole vibrations, which represent the most important vibrational degrees of freedom at low energies. The bosonic phonon concept suggests the occurrence of multiphonon states, which can decay collectively to states with one phonon less by the annihilation of one phonon. In a harmonic picture $`n`$-phonon states have excitation energies of $`n`$ times the one-phonon energy. Two of the appealing aspects of the study of multiphonon excitations high above the yrast-line is the investigation of their anharmonicities reflecting the influence of the Pauli principle and higher order residual interactions as well as the interaction of multiphonon modes with single particle degrees of freedom. Certainly these interactions can cause possible fragmentation of the idealized multiphonon modes. This fragmentation or the eventual observation of almost unfragmented multiphonon states can tell how well the considered phonon is an eigenmode of the nuclear dynamical system. To judge the deviation from the idealized multiphonon picture in reality obviously needs the formulation of this idealized situation and the investigation in how far the collective picture can account for the data. Multiphonon and two-phonon states are at present a very actively investigated topic in nuclear structure physics. A two-quadrupole-phonon ($`2^+2^+`$) triplet with states having spin and parity quantum numbers $`J^\pi =0^+`$, $`2^+`$, $`4^+`$ is usually known in near spherical nuclei. The absence of $`J^\pi =1^+,3^+`$ members of this two-phonon multiplet points at the bosonic character of the identical quadrupole phonons involved. Multi-quadrupole-phonon states (with $`n>2`$) have also been identified, see e.g. . The investigation and identification of double giant dipole resonances (2GDR), ($`3^{}3^{}`$) double-octupole states and ($`2_\gamma ^+2_\gamma ^+`$) double-gamma-vibrations in deformed nuclei are actively debated in the literature, see e.g. Refs. . The mixed-multipolarity ($`2^+3^{}`$) quadrupole-octupole coupled quintuplet with spin and parity quantum numbers $`J^\pi =1^{}`$$`5^{}`$ is a very interesting example of the coupling of two different phonons. Members of the ($`2^+3^{}`$) two-phonon quintuplet should decay by definition by collective $`E2`$ and $`E3`$ transitions to the $`3_1^{}`$ octupole phonon state and to the $`2_1^+`$ quadrupole phonon state, respectively. The ($`2^+3^{}`$) two-phonon states are expected to occur close to the sum energy $`E_{2^+3^{}}=E(2_1^+)+E(3_1^{})`$. There are only a few examples, e.g. <sup>144</sup>Sm , <sup>144</sup>Nd , <sup>142</sup>Ce and <sup>112</sup>Cd , where besides the two-phonon $`1^{}`$ state other multiplet members have been identified experimentally. The reason might be that all the other states from the $`2^+3^{}`$ quintuplet, except the $`1^{}`$, are strongly mixed with non-collective excitations, losing their pure quadrupole–octupole nature. This is supported by shell-model calculations performed recently for $`N=82`$ nuclei . Thus, the most complete data, which can be used as a testing ground for the concept of quadrupole-octupole coupled two-phonon excitations, are the transition strengths from the mostly unfragmented $`1^{}`$ member of the quintuplet. The $`J^\pi =1^{}`$ two-phonon state has been very well investigated in magic and near spherical nuclei. In many nuclei, it has been identified close to the sum energy $`E_{2^+3^{}}`$, indicating a quite harmonic phonon coupling . According to its definition, the two-phonon $`1^{}`$ state should decay by a collective one-phonon $`E2`$ transition to the $`3^{}`$ octupole phonon state. This has been confirmed experimentally . The collective $`1_{2^+3^{}}^{}2_1^+`$ $`E3`$ transition has not yet been measured due to the possibly dominating $`E1`$ and $`M2`$ contaminations in this transition. A particularly interesting feature of the two-phonon $`1^{}`$ state is the existence of a relatively strong $`E1`$ transition to the ground state, which can be sensitively measured in photon scattering experiments . The strength of this two-phonon annihilating $`E1`$ transition approximately equals the strength of the $`3_1^{}2_1^+`$ two-phonon changing $`E1`$ transition which strongly supports quadrupole-octupole-collective origin of these $`E1`$ transitions. All the low-energy one-phonon and two-phonon states mentioned above involve isoscalar phonons. Recently the isovector quadrupole excitation in the valence shell, the $`2_{\mathrm{ms}}^+`$ state, has been identified in some nuclei from the measurement of absolute transition strengths, see, e.g., . Isovector excitations in the valence shell form a whole class of low-lying collective states called, more precisely, mixed-symmetry states . The fundamental $`2_{\mathrm{ms}}^+`$ state decays by a weakly collective $`E2`$ transition to the ground state and by a strong $`M1`$ transition to the isoscalar $`2_1^+`$ state, with an $`M1`$ matrix element of the order of one nuclear magneton . In an harmonic phonon coupling scheme one can expect also the existence of mixed-symmetry two-phonon multiplets, that involve at least one excitation of the mixed-symmetry quadrupole phonon. Two symmetric–mixed-symmetric ($`2_1^+2_{\mathrm{ms}}^+`$) two-quadrupole phonon excitations with positive parity, namely a $`J^\pi =1^+`$ state (the scissors mode), and a $`J^\pi =3^+`$ state have been identified already in the near spherical nucleus <sup>94</sup>Mo . Similarly one can expect the existence of a ($`2_{\mathrm{ms}}^+3^{}`$) mixed-symmetry quadrupole-octupole coupled multiplet with negative parity. In a naive phonon coupling picture members of this quintuplet should show a very complex but, nevertheless, simple to understand decay pattern, which can be predicted from the decay properties of the one-phonon excitations (weakly-collective $`E2`$ decay and strong $`M1`$ decay of the $`2_{\mathrm{ms}}^+`$ state, and collective $`E3`$ and relatively strong $`E1`$ decay of the $`3_1^{}`$ octupole vibration). For instance, the $`J^\pi =1_{2_{\mathrm{ms}}^+3^{}}^{}`$ state should decay by relatively strong $`E1`$, strong $`M1`$, weakly-collective $`E2`$, and collective $`E3`$ transitions to the ground and to the $`1^+`$ scissors state, to the ordinary, isoscalar ($`2^+3^{}`$) multiplet, and to the one-phonon $`3^{}`$ and $`2_{\mathrm{ms}}^+`$ states, respectively (see Fig. 1). Recently, $`\gamma `$ transitions between the $`3^{}`$ octupole phonon state and the $`2_{\mathrm{ms}}^+`$ state have been observed experimentally in near spherical nuclei. For an understanding of these new observations it is necessary to investigate the interrelation of the octupole phonon state and the mixed-symmetry $`2_{\mathrm{ms}}^+`$ state theoretically. That is one aim of the present article. Another purpose of this article is the proposal of possible two-phonon states generated by the coupling of the octupole degree of freedom with the proton-neutron degree of freedom and the quantitative investigation of the properties of such constructions. A convenient and sometimes successful approach to new and complicated physical phenomena is the investigation of symmetries and the algebraic solution of symmetric Hamiltonians. In low-energy nuclear structure physics this task can be achieved by application of the algebraic interacting boson model . Since we deal simultaneously with the isoscalar octupole vibration and the isovector quadrupole vibration in the valence shell of near spherical nuclei, we need a model which incorporates them both. We apply for the first time the $`sdf`$-IBM-2, which considers $`l^\pi =0^+`$ monopole ($`s`$) bosons, $`l^\pi =2^+`$ quadrupole ($`d`$) bosons, $`l^\pi =3^{}`$ octupole ($`f`$) bosons, and the proton-neutron ($`\pi `$ \- $`\nu `$) degree of freedom . In this paper we develop the formalism of the $`sdf`$-IBM-2 in two particular dynamical symmetry limits relevant for the description of vibrational and $`\gamma `$-unstable nuclei. In section II we discuss the group reduction chains of the $`sdf`$-IBM-2 and the quantum numbers, which are necessary to classify the eigenstates of the dynamically symmetric Hamiltonians. We derive analytical formulae for $`\gamma `$ transition rates. The $`E1`$ transitions are calculated with a two-body operator. Quadrupole-octupole coupled two-body operators have turned out to be essential for the description of $`E1`$ transitions generated by quadrupole-octupole collectivity . ## II $`sdf`$-IBM-2 The building blocks of the $`sdf`$-IBM-2 are $`2\times 13`$ creation and $`2\times 13`$ annihilation operators $$b_{\rho lm}^+,b_{\rho lm}$$ (1) where $`\rho =\pi ,\nu `$, $`l=0,2,3`$, $`m=0,\pm 1,\mathrm{},\pm l`$, which by definition satisfy standard boson commutation relations: $$[b_{\rho lm},b_{\rho ^{}l^{}m^{}}^+]=\delta _{\rho \rho ^{}}\delta _{ll^{}}\delta _{mm^{}}.$$ (2) The total number of bosons is conserved for a given nucleus, i.e. $$N=N_\pi +N_\nu =n_{s\pi }+n_{d\pi }+n_{f\pi }+n_{s\nu }+n_{d\nu }+n_{f\nu }.$$ (3) The integers $`n_{l\rho }`$ in Eq. (3) are the eigenvalues of the boson number operators $`\widehat{n}_{l\rho }=\left(b_{\rho l}^+\stackrel{~}{b}_{\rho l}\right)`$, where the dot denotes the scalar product and $`\stackrel{~}{b}_{\rho lm}=(1)^{lm}b_{\rho l,m}`$. The $`2\times 13^2=338`$ bilinear combinations of proton and neutron boson operators of the type $$b_{\rho lm}^+b_{\rho l^{}m^{}}$$ (4) generate the Lie algebra U<sub>π</sub>(13) $``$ U<sub>ν</sub>(13). ### A Group reductions, quantum numbers and wave functions The U<sub>π</sub>(13) $``$ U<sub>ν</sub>(13) symmetry algebra has a rich substructure. In the present study we are interested in the $`F`$-spin symmetric dynamical symmetry limits of the model, which can be characterized by the reduction chains $$\begin{array}{ccccccccccccccc}\text{U}_\pi (13)& & \text{U}_\nu (13)& & \text{U}_{\pi \nu }(13)& & \text{U}_{\pi \nu }(6)& & \text{U}_{\pi \nu }(7)& \mathrm{}& \text{SO}_{\pi \nu }^d(3)& & \text{SO}_{\pi \nu }^f(3)& & \text{SO}_{\pi \nu }(3)\\ & & & & & & & & & & & & & & \\ [N_\pi ]& & [N_\nu ]& & [N_1,N_2]& & [n_1,n_2]& & [m_1,m_2]& \mathrm{}& L_d& & L_f& & L\end{array}$$ (5) where the dots denote alternative group reductions discussed later. Below each group we indicate the quantum numbers, associated with it. In total, the complete classification of the basis states (5) requires 26 quantum numbers corresponding to the $`2\times 13`$ occupation numbers $`n_{\rho lm}`$ in the $`m`$-scheme. The $`F`$-spin quantum number relates in terms of Young tableaux to the one- or two-rowed U<sub>πν</sub>(13) representation as $$F=\frac{N_1N_2}{2}.$$ The maximum value of the $`F`$-spin quantum number $$F_{\mathrm{max}}=\frac{N}{2}$$ corresponds to the totally symmetric representations of the U<sub>πν</sub>(13) group. States with smaller $`F`$-spin quantum numbers $`F<F_{\mathrm{max}}`$ are called mixed-symmetry states and correspond to non-symmetric representations of the U<sub>πν</sub>(13) group. Since the $`sdf`$-IBM-2 contains an octupole degree of freedom, it enlarges the diversity of mixed-symmetry states compared to the standard $`sd`$-IBM-2. This fact can be recognized from the branching rules of U<sub>πν</sub>(13) $``$ U<sub>πν</sub>(6) $``$ U<sub>πν</sub>(7) (see Appendix A). The simplest two-rowed representation of U<sub>πν</sub>(13) reduces to the U<sub>πν</sub>(6) $``$ U<sub>πν</sub>(7) representations in the following way: $`\mathrm{U}_{\pi \nu }(13)`$ $``$ $`\mathrm{U}_{\pi \nu }(6)\mathrm{U}_{\pi \nu }(7)`$ (7) $`[N1,1]`$ $`=`$ $`[n1,1][m]`$ () $`[n][m1,1]`$ $`[n][m]`$ where, $`n=n_1+n_2`$ is the total number of $`s`$ and $`d`$ bosons, $`m=m_1+m_2`$ is the number of $`f`$ bosons, giving thus $`N=n+m`$. As follows from Eq. (6), three types of the mixed-symmetry states arise. 1. The $`[n1,1][m]`$ decomposition corresponds to the usual mixed-symmetry states in the $`sd`$-space, coupled to $`m`$ octupole-bosons with a wave function, which is symmetric in the $`f`$-sector. Examples are the fundamental $`2_{\mathrm{ms}}^+`$ state and the $`1_{sc}^+`$ “scissors” state with no $`f`$-boson at all, which belong to the $`\mathrm{U}_{\pi \nu }(6)\mathrm{U}_{\pi \nu }(7)`$ representation $`[n1,1][0]`$, or the new negative-parity mixed-symmetry two-phonon states which we denote as $`(2_{\mathrm{ms}}^+3_1^{})`$ because they are generated by the coupling of the lowest $`2_{\mathrm{ms}}^+`$ state in the $`sd`$-sector and one symmetric $`f`$-boson, which belong to the $`\mathrm{U}_{\pi \nu }(6)\mathrm{U}_{\pi \nu }(7)`$ representation $`[n1,1][1]`$. Note that in the presence of one $`f`$-boson due to the boson number conservation only $`n=N1`$ bosons remain in the $`sd`$-sector, which can belong to the symmetric $`\mathrm{U}_{\pi \nu }(6)`$ representation $`[n]=[N1]`$ or to the mixed-symmetry representations, i.e., to the lowest $`[n1,1]=[N2,1]`$. Therefore, the representation $`[n1,1][1]`$ can be considered as $`(2_{\mathrm{ms}}^+3_1^{})`$ two-phonon states. It turns out from the application of the branching rules given in Appendix A that this $`(2_{\mathrm{ms}}^+3_1^{})`$ two-phonon coupling will generate mixed-symmetry states with $`F`$-spin quantum numbers $`F=F_{\mathrm{max}}1`$ and $`F=F_{\mathrm{max}}2`$, because the representation $`[N2,1][1]`$ is also present in the U<sub>πν</sub>(13) representation $`[N2,2]`$, which has a total $`F`$-spin quantum number $`F=F_{\mathrm{max}}2`$. In this article, we restrict ourselves to mixed-symmetry states with $`F=F_{\mathrm{max}}1`$. 2. The $`[n][m1,1]`$ reduction describes the mixed-symmetry states which appear due to mixed-symmetry in the $`f`$-sector. We will not consider such kind of states in the present paper. 3. Finally, $`[n][m]`$ states are symmetric separately in the $`sd`$\- and in the $`f`$-sectors, however, they are coupled in a non-symmetric way within the full $`sdf`$-space. The simplest example for a mixed-symmetry state if this type is the $`3_{\mathrm{ms}}^{}`$ state, which consists of $`m=1`$ $`f`$-boson and $`n=N1`$ $`s`$-bosons. This $`3_{\mathrm{ms}}^{}`$ state is the mixed-symmetry analogue state of the symmetric octupole phonon, the $`3_1^{}`$ state. This can be clearly seen from the explicit wave functions given in Table I. Higher excited mixed-symmetry states of this type can be obtained by replacing in a symmetric way $`s`$-bosons with $`d`$-bosons in the $`3_{\mathrm{ms}}^{}`$ wave function along with proper normalization. We denote the lowest energy example of such mixed-symmetry states schematically as $`(2^+3_{\mathrm{ms}}^{})`$. The operator which controls the excitation energy of mixed-symmetry states with $`F`$-spin quantum numbers $`F<F_{\mathrm{max}}`$ is the Majorana operator. By definition, this is the most general operator which annihilates any totally symmetric state. Within the U<sub>πν</sub>(13) $``$ U<sub>πν</sub>(6) $``$ U<sub>πν</sub>(7), there are three important types of Majorana operators. The well known Majorana operator $`\widehat{M}_6`$ is associated with the U(6) group generated by the $`s`$\- and $`d`$-bosons only. It has the form $$\widehat{M}_6=[s_\nu ^+\times d_\pi ^+s_\pi ^+\times d_\nu ^+]^{(2)}[s_\nu \times \stackrel{~}{d}_\pi s_\pi \times \stackrel{~}{d}_\nu ]^{(2)}2\underset{k=1,3}{}[d_\nu ^+\times d_\pi ^+]^{(k)}[\stackrel{~}{d}_\nu \times \stackrel{~}{d}_\pi ]^{(k)},$$ (8) which influences the states whose wave functions are non-symmetric in the $`sd`$-sector of the model space. In Eq. (8) and below, $`\times `$ denotes the standard tensor product of two irreducible tensorial operators. It relates to the $`\mathrm{U}_{\pi \nu }(6)`$ quadratic Casimir operator as $$C_2[\text{U}_{\pi \nu }(6)]=n(n+5)2\widehat{M}_6.$$ (9) The Majorana operator $`\widehat{M}_7`$ is associated with the U(7) group generated by the $`f`$-bosons only. It has the form $$\widehat{M}_7=2\underset{k=1,3,5}{}[f_\nu ^+\times f_\pi ^+]^{(k)}[\stackrel{~}{f}_\nu \times \stackrel{~}{f}_\pi ]^{(k)}$$ (10) which is responsible for pushing up the mixed-symmetry states of the second type, i.e non-symmetric in the $`f`$-sector. It relates to the $`\mathrm{U}_{\pi \nu }(7)`$ quadratic Casimir invariant as $$C_2[\text{U}_{\pi \nu }(7)]=m(m+6)2\widehat{M}_7.$$ (11) Finally, the Majorana operator which is associated with the full group U(13), reads $`\widehat{M}_{13}`$ $`=`$ $`\widehat{M}_6+\widehat{M}_7+[s_\nu ^+\times f_\pi ^+s_\pi ^+\times f_\nu ^+]^{(3)}[s_\nu \times \stackrel{~}{f}_\pi s_\pi \times \stackrel{~}{f}_\nu ]^{(3)}`$ (13) $`+{\displaystyle \underset{k=1,2,3,4,5}{}}(1)^{k+1}[d_\nu ^+\times f_\pi ^+d_\pi ^+\times f_\nu ^+]^{(k)}[\stackrel{~}{d}_\nu \times \stackrel{~}{f}_\pi \stackrel{~}{d}_\pi \times \stackrel{~}{f}_\nu ]^{(k)},`$ $`=`$ $`\left[F_{\mathrm{max}}\left(F_{\mathrm{max}}+1\right)\widehat{F}^2\right].`$ (14) This operator acts on all three types of mixed-symmetry states. $`\widehat{M}_{13}`$ relates to the $`\mathrm{U}_{\pi \nu }(13)`$ quadratic Casimir operator as $$C_2[\text{U}_{\pi \nu }(13)]=N(N+12)2\widehat{M}_{13}.$$ (15) The most general Hamiltonian constructed from linear and quadratic Casimir operators of the group U<sub>πν</sub>(6) $``$ U<sub>πν</sub>(7) reads $$\begin{array}{cc}H\hfill & =H_0+\lambda \widehat{M}_{13}+\lambda ^{}\widehat{M}_6+\lambda ^{\prime \prime }\widehat{M}_7+ϵ_dC_1[\text{U}_{\pi \nu }(5)]+\alpha C_2[\text{U}_{\pi \nu }(5)]+\kappa ^{}C_2[\text{SU}_{\pi \nu }(3)]\hfill \\ & +\eta C_2[\text{SO}_{\pi \nu }(6)]+\beta C_2[\text{SO}_{\pi \nu }(5)]+\gamma _dC_2[\text{SO}_{\pi \nu }^d(3)]+ϵ_fC_1[\text{U}_{\pi \nu }(7)]\hfill \\ & +\kappa C_2[\text{SO}_{\pi \nu }(7)]+\xi C_2[\text{G}_{2\pi \nu }]+\gamma _fC_2[\text{SO}_{\pi \nu }^f(3)]+\gamma C_2[\text{SO}_{\pi \nu }(3)].\hfill \end{array}$$ (16) The definitions of the Casimir invariants of Lie groups used here are given in Appendix B. The structure of the Majorana operators has interesting consequences. The $`sdf`$-IBM-2 reduces to the simpler version $`sdf`$-IBM-1, if the strength constant $`\lambda `$ of the Majorana operator $`\widehat{M}_{13}`$ is set to infinity. Using a finite value of $`\lambda `$, but $`\lambda ^{\prime \prime }=\mathrm{}`$ will remove only those states which have mixed-symmetry in the $`f`$-sector alone. We will use this choice throughout the paper. For the ordinary mixed-symmetry states in the $`sd`$-sector, which are known already from the $`sd`$-IBM-2, the sum $`\lambda +\lambda ^{}`$ plays the role of the ordinary Majorana parameter in the $`sd`$-IBM-2. The excitation energies of mixed-symmetry states, which contain one $`f`$-boson are in addition also functions of the quantity $`\lambda \lambda ^{}`$. This parameter controls the excitation energy of mixed-symmetry states with negative parity. Since we are going to study vibrational or $`\gamma `$-unstable nuclei, we are interested in two well-known dynamical symmetry limits of the positive parity $`sd`$-subchain , namely, $$\text{U}_{\pi \nu }(6)\left\{\begin{array}{c}\text{U}_{\pi \nu }(5)\\ \text{SO}_{\pi \nu }(6)\end{array}\right\}\text{SO}_{\pi \nu }(5)\text{SO}_{\pi \nu }(3)$$ (17) which we consider in the following subsections. The negative parity subchain has a unique reduction : $$\begin{array}{ccccccc}\text{U}_{\pi \nu }(7)& & \text{SO}_{\pi \nu }(7)& & \text{G}_{2\pi \nu }& & \text{SO}_{\pi \nu }(3)\\ & & & & & & \\ [m_1,m_2]& & (\omega _1,\omega _2)& & (u_1,u_2)& \{\beta _j\}& L_f\end{array}$$ (18) where the quantum numbers $`\{\beta _j\}`$ ($`j=1,2,3,4`$) are the so-called missing labels, which are necessary to classify completely the G$`{}_{2}{}^{}`$ SO(3) reduction. Unlike the case of $`sdf`$-IBM-1 where only the totally symmetric representations are important , we also deal here with two-rowed representations of the algebras in (5) and (18). This requires some modifications. For example, we have to take into account the exceptional group G<sub>2</sub>. For the reduction of not fully symmetric representations of SO(7) the group G<sub>2</sub> helps to resolve the labelling problem. The missing label, which is in general necessary to uniquely define the SO(7) $``$ G<sub>2</sub> reduction, is redundant in our case of only two-rowed representations. #### 1 The U<sub>πν</sub>(1)$``$U<sub>πν</sub>(5)$``$U<sub>πν</sub>(7) dynamical symmetry limit In this dynamical symmetry limit the reduction chain of the positive parity U<sub>πν</sub>(6) subalgebra is $$\begin{array}{ccccccc}\text{U}_{\pi \nu }(6)& & \text{U}_{\pi \nu }(5)& & \text{SO}_{\pi \nu }(5)& & \text{SO}_{\pi \nu }^d(3)\\ & & & & & & \\ [n_1,n_2]& & \{n_{d_1},n_{d_2}\}& & [\tau _1,\tau _2]& \{\alpha _i\}& L_d\end{array}$$ (19) where $`\alpha _i`$ ($`i=1,2`$) are missing labels, necessary to completely classify the SO(5) $``$SO(3) reduction. We denote here $`n=n_1+n_2`$ the total number of $`s`$ and $`d`$ bosons, while the number of $`d`$ bosons only is denoted as $`n_d=n_{d_1}+n_{d_2}`$. Thus, the total wave function can be written as $$|[N_\pi ][N_\nu ];[N_1,N_2][n_1,n_2]\{n_{d_1},n_{d_2}\}(\tau _1,\tau _2)\{\alpha _i\}L_d;[m_1,m_2](\omega _1,\omega _2)(u_1,u_2)\{\beta _j\}L_f;LM$$ (20) The quantum numbers of the lowest states are given in Table I. In the U<sub>πν</sub>(5)$``$U<sub>πν</sub>(7) dynamical symmetry limit it is rather simple to construct explicitly the wave functions. We shall do this in order to make clear the variety of mixed-symmetry states appearing in the model. In Table I we present the wave functions for some states of interest. We discussed above already the one-$`f`$-boson octupole states. Of particular interest are here also the lowest $`1^{}`$ states. From the coupling of one $`d`$-boson quadrupole phonon and one $`f`$-boson octupole phonon four $`1^{}`$ states emerge. This is evident in the $`m`$-scheme in $`F`$-space because the four basis wave functions $`|d_\pi f_\pi `$, $`|d_\nu f_\pi `$, $`|d_\pi f_\nu `$, and $`|d_\nu f_\nu `$ can be formed. These basis states can be combined to states with U<sub>πν</sub>(13) symmetry and good $`F`$-spin. One obtains one symmetric $`1^{}`$ state corresponding to the totally symmetric $`[N]`$ irrep of U<sub>πν</sub>(13) with $`F`$-spin quantum number $`F=F_{\mathrm{max}}`$, and three mixed-symmetry states: two of them with $`F`$-spin quantum numbers $`F=F_{\mathrm{max}}1`$ and one with $`F=F_{\mathrm{max}}2`$. One $`F=F_{\mathrm{max}}1`$ state corresponds to the $`[N1,1][N2,1][1]`$ reduction and the other to the $`[N1,1][N1][1]`$ reduction in the U<sub>πν</sub>(13)$``$ U<sub>πν</sub>(6)$``$ U<sub>πν</sub>(7), chain. The last one arises from the $`[N2,2][N2,1][1]`$ reduction. According to the discussion above we denote them as $`1^{}=(2^+3^{})_1^{}`$, $`1_{\mathrm{ms}}^{}=(2_{\mathrm{ms}}^+3^{})_1^{}`$, $`1_{\mathrm{ms}}^{}=(2^+3_{\mathrm{ms}}^{})_1^{}`$ and $`1_{\mathrm{ms}}^{\prime \prime }`$, respectively. The U<sub>πν</sub>(5)$``$U<sub>πν</sub>(7) dynamical symmetry Hamiltonian expressed in terms of Casimir invariants of first and second order of the relevant groups reads $$\begin{array}{cc}H\hfill & =H_0+\lambda \widehat{M}_{13}+\lambda ^{}\widehat{M}_6+\lambda ^{\prime \prime }\widehat{M}_7+ϵ_dC_1[\text{U}_{\pi \nu }(5)]\hfill \\ & +\alpha C_2[\text{U}_{\pi \nu }(5)]+\beta C_2[\text{SO}_{\pi \nu }(5)]+\gamma _dC_2[\text{SO}_{\pi \nu }^d(3)]+ϵ_fC_1[\text{U}_{\pi \nu }(7)]\hfill \\ & +\kappa C_2[\text{SO}_{\pi \nu }(7)]+\xi C_2[\text{G}_{2\pi \nu }]+\gamma _fC_2[\text{SO}_{\pi \nu }^f(3)]+\gamma C_2[\text{SO}_{\pi \nu }(3)].\hfill \end{array}$$ (21) The eigenvalue problem for the dynamical symmetry Hamiltonian (21) can be solved analytically. The solution of the full U(5) Hamiltonian reads $$\begin{array}{cc}E\hfill & =E_0+\lambda \left[F_{\mathrm{max}}\left(F_{\mathrm{max}}+1\right)F(F+1)\right]+\lambda ^{}n_2(n_1+1)+\lambda ^{\prime \prime }m_2(m_1+1)\hfill \\ & +ϵ_d(n_{d_1}+n_{d_2})+\alpha [n_{d_1}(n_{d_1}+4)+n_{d_2}(n_{d_2}+2)]+\beta [\tau _1(\tau _1+3)+\tau _2(\tau _2+1)]\hfill \\ & +\gamma _dL_d(L_d+1)+ϵ_f(m_1+m_2)+\kappa [\omega _1(\omega _1+5)+\omega _2(\omega _2+3)]\hfill \\ & +\xi [u_1(u_1+5)+u_2(u_2+4)+u_1u_2]+\gamma _fL_f(L_f+1)+\gamma L(L+1).\hfill \end{array}$$ (22) To construct the spectrum one needs to know the branching rules for the groups involved (see Appendix). An energy spectrum corresponding to (21) is shown in Fig. 2. The parameters of the Hamiltonian are specified in the figure caption. #### 2 The SO<sub>πν</sub>(6)$``$U<sub>πν</sub>(7) dynamical symmetry limit In this dynamical symmetry limit the reduction chain of the positive parity U<sub>πν</sub>(6) subalgebra is $$\begin{array}{ccccccc}\text{U}_{\pi \nu }(6)& & \text{SO}_{\pi \nu }(6)& & \text{SO}_{\pi \nu }(5)& & \text{SO}_{\pi \nu }^d(3)\\ & & & & & & \\ [n_1,n_2]& & \sigma _1,\sigma _2& & [\tau _1,\tau _2]& \{\alpha _i\}& L_d\end{array}$$ (23) and the total wave function is as follows: $$|[N_\pi ][N_\nu ];[N_1,N_2][n_1,n_2]\sigma _1,\sigma _2(\tau _1,\tau _2)\{\alpha _i\}L_d;[m_1,m_2](\omega _1,\omega _2)(u_1,u_2)\{\beta _j\}L_f;LM$$ (24) The quantum numbers for the lowest states in this dynamical symmetry are presented in Table II. The SO<sub>πν</sub>(6)$``$U<sub>πν</sub>(7) dynamical symmetry Hamiltonian reads $$\begin{array}{cc}H\hfill & =H_0+\lambda \widehat{M}_{13}+\lambda ^{}\widehat{M}_6+\lambda ^{\prime \prime }\widehat{M}_7+\zeta C_2[\text{SO}_{\pi \nu }(6)]\hfill \\ & +\beta C_2[\text{SO}_{\pi \nu }(5)]+\gamma _dC_2[\text{SO}_{\pi \nu }^d(3)]+ϵ_fC_1[\text{U}_{\pi \nu }(7)]\hfill \\ & +\kappa C_2[\text{SO}_{\pi \nu }(7)]+\xi C_2[\text{G}_{2\pi \nu }]+\gamma _fC_2[\text{SO}_{\pi \nu }^f(3)]+\gamma C_2[\text{SO}_{\pi \nu }(3)].\hfill \end{array}$$ (25) and has the eigenvalues $$\begin{array}{cc}E\hfill & =E_0+\lambda \left[F_{\mathrm{max}}\left(F_{\mathrm{max}}+1\right)F(F+1)\right]+\lambda ^{}n_2(n_1+1)+\lambda ^{\prime \prime }m_2(m_1+1)\hfill \\ & +\zeta [\sigma _1(\sigma _1+4)+\sigma _2(\sigma _2+2)]+\beta [\tau _1(\tau _1+3)+\tau _2(\tau _2+1)]+\gamma _dL_d(L_d+1)\hfill \\ & +ϵ_f(m_1+m_2)+\kappa [\omega _1(\omega _1+5)+\omega _2(\omega _2+3)]\hfill \\ & +\xi [u_1(u_1+5)+u_2(u_2+4)+u_1u_2]+\gamma _fL_f(L_f+1)+\gamma L(L+1).\hfill \end{array}$$ (26) The branching rules for the corresponding groups can be found in Ref. and in the Appendix below. A typical energy spectrum corresponding to Eq. (26) is shown in Fig. 3. The parameters are specified in the caption. ### B Electromagnetic transitions In this subsection we discuss electromagnetic transitions between low-lying states in the $`sdf`$-IBM-2 with particular emphasis of those transitions which are new in the present model with respect to the standard $`sd`$-IBM-2 and $`sdf`$-IBM-1 species. For the study of the interrelation between the $`F=F_{\mathrm{max}}`$ octupole phonon and the $`F=F_{\mathrm{max}}1`$ mixed-symmetry quadrupole phonon and possible two-phonon states generated by the coupling of them, we are interested in $`E1,M1,E2`$, and $`E3`$ transition strengths. The reduced transition probabilities $$B(\mathrm{\Pi }\lambda ;J_iJ_f)=\frac{1}{2J_i+1}|J_f||T(\mathrm{\Pi }\lambda )||J_i|^2$$ (27) are calculated from the matrix elements of the transition operators $`T(\mathrm{\Pi }\lambda )`$, where $`\mathrm{\Pi }`$ stands for the radiation character, either $`E`$ or $`M`$, and $`\lambda `$ denotes the multipolarity. The electric quadrupole operator reads $$T(E2)=e_\pi Q_\pi ^{(2)}+e_\nu Q_\nu ^{(2)},$$ (28) where $$Q_\rho ^{(2)}=[d_\rho ^+\times s_\rho +s_\rho ^+\times \stackrel{~}{d}_\rho ]^{(2)}+\chi _\rho [d_\rho ^+\times \stackrel{~}{d}_\rho ]^{(2)}+\chi _\rho ^{}[f_\rho ^+\times \stackrel{~}{f}_\rho ]^{(2)}.$$ (29) The magnetic dipole operator has the form $$T(M1)=\sqrt{\frac{3}{4\pi }}\left(g_\pi L_\pi ^d+g_\nu L_\nu ^d+g_\pi ^{}L_\pi ^f+g_\nu ^{}L_\nu ^f\right),$$ (30) where $$L_\rho ^d=\sqrt{10}[d_\rho ^+\times \stackrel{~}{d}_\rho ]^{(1)},L_\rho ^f=2\sqrt{7}[f_\rho ^+\times \stackrel{~}{f}_\rho ]^{(1)}.$$ (31) According to the Eqs. (2831) the addition of the $`f`$ boson enlarges the structure of electromagnetic operators of the standard $`sd`$-IBM-2. The standard $`sd`$-IBM-2, however, works well for the description of $`E2`$ and $`M1`$ transitions. In order to reduce the number of model parameters in the $`sdf`$-IBM-2 we, therefore, neglect below those parts of the $`E2`$ and $`M1`$ transition operators which involve the $`f`$-boson. Thus, we assume $`\chi _\rho ^{}=0`$ and $`g_\rho ^{}=0`$. In contrast, the $`f`$-boson is essential for $`E3`$ and $`E1`$ transitions, which are new with respect to the $`sd`$-IBM-2. Since there is at present no experimental evidence for the existence of states with mixed-symmetry in the $`f`$-boson sector, we shall consider here only $`F`$-scalar octupole transitions, i.e., we use the following symmetric octupole operator: $$T(E3)=e_3\left(O_\pi ^{(3)}+O_\nu ^{(3)}\right),$$ (32) where $`O_\rho ^{(3)}`$ stands for $$O_\rho ^{(3)}=[s_\rho ^+\times \stackrel{~}{f}_\rho +f_\rho ^+\times s_\rho ]^{(3)}+\chi [d_\rho ^+\times \stackrel{~}{f}_\rho +f_\rho ^+\times \stackrel{~}{d}_\rho ]^{(3)}.$$ (33) The general one-body electric dipole operator has the form $$T(E1)=\alpha _\pi D_\pi ^{(1)}+\alpha _\nu D_\nu ^{(1)},$$ (34) where $$D_\rho ^{(1)}=\left[d_\rho ^+\times \stackrel{~}{f}_\rho +f_\rho ^+\times \stackrel{~}{d}_\rho \right]^{(1)}.$$ (35) While the one-body $`M1`$, $`E2`$, and $`E3`$ transition operators work satisfactorily, the one-body $`E1`$-transition operator fails to provide an adequate description of low-lying dipole transition strengths in vibrational nuclei . This applies for near-spherical nuclei also in the case when a $`p`$-boson is considered as it will be discussed below. The operator (34) has very strict selection rules, conserving the total number of $`d`$\- and $`f`$-bosons $`\mathrm{\Delta }(n_d+n_f)=0`$. This contradicts the experimental findings where the quadrupole-octupole collective $`E1`$ transitions, including the ground state decay of the quadrupole-octupole two-phonon $`1^{}`$ state, are even found to be enhanced, see e.g. Refs. . Two approaches exist to avoid this problem. Following the first, the $`sdf`$-IBM is enlarged with a $`p`$ boson ($`l^\pi =1^{}`$), thus leading to the so-called $`spdf`$-IBM . The inclusion of the $`p`$ boson in the IBM corresponds to the consideration of a collective dipole vibration. There exists no unambiguous interpretation of the $`p`$-boson origin. Some authors point out the origin of the $`p`$-boson from the collective low-lying valence shell excitation after removing the center-of-mass motion . Others have associated the $`p`$-boson with the GDR . The $`p`$ boson induced one-body electric dipole operator works succesfully for the description of dipole transition strengths in deformed nuclei . For near-spherical nuclei, where the present article is focussing on, the inclusion of the $`p`$ boson does not lift all discrepancies . The lowest $`1^{}`$ state is the only strong low-lying $`E1`$ excitation in many near-spherical nuclei . This fact strongly suggests that this state is identified with the collective negative-parity dipole mode in the case of considering a low-energy $`p`$ boson. This view, however, renders the well-established agreement of the $`1^{}`$ excitation energy with the quadrupole-octupole two-phonon sum energy a sheer coincidence. Also the enhanced $`E1`$ strengths from other members of the quadrupole-octupole coupled quintuplet with spin quantum numbers $`J1`$ are not at all simple to understand. On the other hand some other experimental facts indicate different origin of strong $`E1`$ transitions in near-spherical nuclei than sheer mixing with the GDR. For example, studies of $`E1`$-strength distribution in <sup>140</sup>Ce report that the lowest $`1^{}`$ state exhibits an enhanced dipole transition to the ground state, while transitions from a number of higher lying $`1^{}`$ states are considerably weaker although they lie closer to the GDR and although they are calculated to have predominantly a simple 1p1h-structure, which would even favor a mixing with the GDR. Thus, the observed concentration of $`E1`$-strength in the lowest $`1^{}`$ states should not be simply attributed to admixtures of the GDR into the states wave function if not at the same time the smaller $`E1`$ strengths of higher lying $`1^{}`$ states are explained, too. It is even more striking that a model which can treat all these $`E1`$ excitations on the same footing, the microscopic Quasiparticle Phonon Model, supports the interpretation of enhanced low-lying $`E1`$ transitions due to quadrupole-octupole collectivity . Therefore, another approach is preferable. The multiphonon structure of excitations in vibrational nuclei as well as an octupole or quadrupole-octupole coupled origin of negative-parity states, naturally suggests an $`E1`$-transition operator which contains a quadrupole-octupole coupled two-body part. Such a picture is also supported by the empirical correlation between $`E1`$-transition strengths of the $`1_1^{}0_1^+`$ and $`3_1^{}2_1^+`$ transitions found recently for heavy near-spherical nuclei and by the fore-mentioned microscopic calculations . Furthermore, it avoids the problems in connection with the introduction of a $`p`$ boson discussed above. We choose, therefore, to work with a quadrupole-octupole coupled two-body part in the $`E1`$ operator. The two-body dipole operator has already been introduced within the simple quadrupole and octupole phonon model and earlier within the $`sdf`$-IBM-1 . This ansatz is in good agreement with the available data. The $`E1`$ transition operator considered here has the form $$\begin{array}{c}T(E1)=\alpha _\pi D_\pi ^{(1)}+\alpha _\nu D_\nu ^{(1)}+\\ \frac{\beta }{2N}\left[\left(Q_\pi ^{(2)}+\eta Q_\nu ^{(2)}\right)\times \left(O_\pi ^{(3)}+O_\nu ^{(3)}\right)+\left(O_\pi ^{(3)}+O_\nu ^{(3)}\right)\times \left(Q_\pi ^{(2)}+\eta Q_\nu ^{(2)}\right)\right]^{(1)},\end{array}$$ (36) where $`D_\rho ^{(1)}`$, $`Q_\rho ^{(2)}`$, and $`O_\rho ^{(3)}`$ are given by the Eqs. (35), (29) with $`\chi _\rho ^{}=0`$, and (33), respectively, and $`\eta =e_\nu /e_\pi `$. We stress that we introduce only one free parameter more ($`\beta `$) in the $`E1`$ operator with respect to the one-body operator from Eq. (34). Below we analyze $`E1`$-transitions within two particular dynamical symmetry limits. #### 1 The U<sub>πν</sub>(1)$``$U<sub>πν</sub>(5)$``$U<sub>πν</sub>(7) dynamical symmetry limit Using the wave functions (20) we can obtain the matrix elements of the transition operators (28), (30), (32) and (36). The technique for the calculation of such matrix elements is described in detail in Ref. . In this particular dynamical symmetry limit, it is more convenient to use the explicit expressions of the wave functions constructed from boson operators following the method described in (see, e.g., Ref. ). The $`E2`$ and $`M1`$ transition strengths between those states, whose wave functions do not involve any $`f`$ bosons, are exactly the same as in the case of the $`sd`$-IBM-2 and can be found in . The advantage of the $`sdf`$-IBM-2 is that it provides the description of $`E2`$ and $`M1`$ transitions between negative parity states and of $`E1`$ and $`E3`$ transitions between the states of different parity and different $`F`$-spin. Here we are particularly interested in the electric dipole transitions which connect the octupole $`3_1^{}`$ state with the symmetric $`2_1^+`$ state and with the mixed-symmetry $`2_{\mathrm{ms}}^+`$ state, as well as in those which connect the $`(2^+3^{})`$ two-phonon negative parity multiplet with the ground state band. We also consider here the new mixed-symmetry quadrupole-octupole coupled dipole states $`1_{\mathrm{ms}}^{}=(2_{\mathrm{ms}}^+3^{})_1^{}`$ and $`1_{\mathrm{ms}}^{}=(2^+3_{\mathrm{ms}}^{})_1^{}`$ and their electromagnetic properties (including $`M1`$ strengths) as examples for mixed-symmetry states with negative parity. Analytical expressions of reduced transition probabilities are summarized in Table III. It is worthwhile to look closely to the analytical expressions in Table III in order recognize the fingerprints of the simple phonon picture the consideration of which has guided us to the present model. The $`E1`$ transition strengths from the $`3_1^{}`$ state to the symmetric and mixed-symmetry one-phonon quadrupole excitations, the $`2_1^+`$ state and the $`2_{\mathrm{ms}}^+`$ state, are both of the order of $`(1/N)^0`$, i.e., they can be strong on an absolute scale and comparable. The $`3_1^{}2_{\mathrm{ms}}^+`$ $`F`$-vector $`E1`$ strength can be even larger than that of the strong $`F`$-scalar $`3_1^{}2_1^+`$ $`E1`$ transition if the effective electric dipole charges $`\alpha _\pi `$ and $`\alpha _\nu `$ have opposite signs. Similarly, the $`E1`$ ground state excitation strengths to the $`1_1^{}`$ state and the $`1_{\mathrm{ms}}^{}`$ state are both of the order of $`(1/N)^0`$ due to the two-body part of the $`E1`$ operator. However, here the $`F`$-scalar $`0_1^+1_1^{}`$ $`E1`$ strength must be expected to be larger than the $`F`$-vector $`E1`$ strength of the $`0_1^+1_{\mathrm{ms}}^{}`$ transition because the parameter $`\eta `$ usually takes values between 0 and 1 leading to a reduction of the $`F`$-vector $`E1`$ strength which is proportional to the factor $`(1\eta )^2`$. It is also interesting to compare the properties of the new mixed-symmetry $`1^{}`$ states, $`1_{\mathrm{ms}}^{}`$ and $`1_{\mathrm{ms}}^{}`$. The direct $`E1`$ ground state excitation strengths to the $`1_{\mathrm{ms}}^{}`$ state is only of the order of $`(1/N)^2`$ and therefore much smaller than the $`E1`$ strength to the $`1_{\mathrm{ms}}^{}`$ state, which is of the order of $`(1/N)^0`$. The $`1_{\mathrm{ms}}^{}3_1^{}`$ $`E2`$ strength is of the order of N, which indicates quadrupole collectivity. Since it is proportional to the square of the difference of the effective quadrupole boson charges $`(e_\pi e_\nu )^2`$ this transition should be weakly collective like the known $`2_{\mathrm{ms}}^+0_1^+`$ one-quadrupole phonon decay. In contrast, the $`1_{\mathrm{ms}}^{}3_1^{}`$ $`E2`$ strength is of the order of $`1/N`$ and cannot be large. Very interesting are also the $`M1`$ transitions between symmetric and mixed-symmetry negative parity states. The $`M1`$ strengths from the $`1_{\mathrm{ms}}^{}`$ state to the quadrupole-octupole coupled $`1^{}`$ and $`2^{}`$ two-phonon states are of the order of $`(1/N)^0`$ and should be comparable to the $`2_{\mathrm{ms}}^+2_1^+`$ strong $`M1`$ transition. The $`1_{\mathrm{ms}}^{}1^{},2^{}`$ $`M1`$ transitions are of the order of $`(1/N)^2`$ if the $`d`$-boson part alone is used in the $`M1`$ transition operator (30). If one allows for finite $`g`$-factors $`g_\rho ^{}`$ for the $`f`$-boson parts in the $`M1`$ operator (30), they will be responsible for the leading term of the order of $`(1/N)^0`$ which was therefore included in Table III. This resembles the fact that the $`1_{\mathrm{ms}}^{}`$ state originates from the non symmetric coupling of an $`f`$-boson to the $`sd`$-sector. In summary the $`1_{\mathrm{ms}}^{}`$ state shows the behavior, which is mentioned in the introduction to be expected for a two-phonon state formed by the coupling of the fundamental phonons which generate the $`3_1^{}`$ state and the $`2_{\mathrm{ms}}^+`$ state. #### 2 The SO<sub>πν</sub>(6)$``$U<sub>πν</sub>(7) dynamical symmetry limit Using the wave functions (24) we have calculated the matrix elements of the transition operators (28), (30), (32) and (36). We use here the quadrupole operator (29) with $`\chi _\rho =0`$ as is conventional for the SO(6) limit. The analytical expressions for the $`E2`$ and $`M1`$ transitions between the states which are formed only from $`s`$ and $`d`$ bosons are identical to those of the $`sd`$-IBM-2 and can be found in Ref. . Analytical expressions for $`E2`$ and $`M1`$ transition strengths between negative parity states and for $`E1`$ and $`E3`$ transition strengths are summarized in Table IV. ## III Summary The relation and coupling of the collective octupole degree of freedom to proton-neutron mixed-symmetry states is discussed in the framework of the interacting boson model. For this purpose $`s`$-bosons, $`d`$-bosons, and $`f`$-bosons as well as the proton-neutron degree of freedom had to be taken into account leading to the hitherto not considered $`sdf`$-IBM-2 version of the model. We developed the formalism of the $`sdf`$-IBM-2 considering two $`F`$-spin symmetric dynamical symmetry limits which are denoted by U<sub>πν</sub>(5)$``$U<sub>πν</sub>(7) and SO<sub>πν</sub>(6)$``$U<sub>πν</sub>(7). These analytically solvable limits of the model can be relevant for the description of near spherical, vibrational and $`\gamma `$-unstable nuclei, respectively. The full dynamically symmetric Hamiltonians are formulated, their eigenstates are specified and the analytical expressions for their excitation energies are given. Low-lying, collective, negative parity states are discussed including for the first time negative parity mixed-symmetry states. Analytical expressions for $`E1`$, $`M1`$, $`E2`$, and $`E3`$ transition strengths were evaluated. The $`E1`$ operator considered here contains a quadrupole-octupole coupled two-body part besides the one-body term. Evidence for the possible dominance of $`F`$-vector $`E1`$ transitions over $`F`$-scalar $`E1`$ transitions is found. The decay properties of $`1_{\mathrm{ms}}^{}`$ states in vibrational nuclei are predicted. ###### Acknowledgements. We thank C. Fransen for informing us about transition strengths in <sup>94</sup>Mo prior to publication. We gratefully acknowledge discussions with P. von Brentano, R. F. Casten, C. Fransen, A. Gelberg, F. Iachello, J. Jolie, R. V. Jolos, T. Otsuka, A. Wolf, N.V. Zamfir, and A. Zilges. N.A.S., N.P. and P.V.I. thank the Institute for Nuclear Theory, University of Washington for hospitality. This work was partially supported by the Competitive Center at the St. Petersburg State University, by Grant-in-Aid for Scientific Research (B)(2)(10044059) from the Ministry of Education, Science and Culture, by the U.S. DOE under Contract No. DE-FG02-91ER-40609, and by the Deutsche Forschungsgemeinschaft under Contracts No. Br 799/9-1 and Pi 393/1-1. ## A U(13)$``$ U(6)$``$U(7) branching rules In order to provide a complete classification scheme we need to reduce the representations of U(13) group to the Kronecker product U(6)$``$U(7). Here we present the branching rules only for the totally symmetric $`[N]`$ and the lowest mixed-symmetry $`[N1,1]`$ and $`[N2,2]`$ irreducible representations of U(13). The U(6)$``$U(7) representations are denoted as $`[\alpha ][\beta ]`$. Those U(6)$``$U(7) decompositions, which contain $`1^{}`$ states discussed in the text, are underlined. $$\begin{array}{cc}[N]=\hfill & [N][0],\underset{¯}{[N1][1]},\mathrm{},[1][N1],[0][N]\hfill \\ [N1,1]=\hfill & [N1,1][0],\underset{¯}{[N2,1][1]},\mathrm{},[2,1][N3],[1,1][N2]\hfill \\ & \underset{¯}{[N1][1]},[N2][2],\mathrm{},[2][N2],[1][N1]\hfill \\ & [N2][1,1],[N3][2,1],\mathrm{},[1][N2,1],[0][N1,1]\hfill \\ [N2,2]=\hfill & [N2,2][0],[N3,2][1],\mathrm{},[3,2][N5],[2,2][N4]\hfill \\ & \underset{¯}{[N2,1][1]},[N3,1][2],\mathrm{},[3,1][N4],[2,1][N3]\hfill \\ & [N2][2],[N3][3],\mathrm{},[3][N3],[2][N2]\hfill \\ & [N3,1][1,1],[N4,1][2,1],\mathrm{},[2,1][N4,1],[1,1][N3,1]\hfill \\ & [N3][2,1],[N4][3,1],\mathrm{},[2][N3,1],[1][N2,1]\hfill \\ & [N4][2,2],[N5][3,2],\mathrm{},[1][N3,2],[0][N2,2]\hfill \end{array}$$ (A1) The branching rules for the chains U(6)$``$ U(5)$``$ SO(5) $``$SO(3) and U(6)$``$ SO(6)$``$ SO(5) $``$SO(3) reduction chains can be found elsewhere . The partial classification scheme for the chain U(7)$``$ SO(7)$``$G<sub>2</sub> $``$SO(3) is given in Table VII. ## B Casimir invariants The Casimir invariants of Lie algebras exploited here can be expressed in terms of boson creation and annihilation operators as follows : $$\begin{array}{cc}C_1\left[\text{U(7)}\right]=\hfill & \left(f^+\stackrel{~}{f}\right),\hfill \\ C_2\left[\text{U(7)}\right]=\hfill & \underset{K=0}{\overset{6}{}}\left[f^+\times \stackrel{~}{f}\right]^{(K)}\left[f^+\times \stackrel{~}{f}\right]^{(K)},\hfill \\ C_1\left[\text{U(6)}\right]=\hfill & \left(s^+s\right)+\left(d^+\stackrel{~}{d}\right),\hfill \\ C_2\left[\text{U(6)}\right]=\hfill & (s^+s)(s^+s)+\underset{K=0}{\overset{4}{}}\left[d^+\times \stackrel{~}{d}\right]^{(K)}\left[d^+\times \stackrel{~}{d}\right]^{(K)},\hfill \\ C_2\left[\text{SU(3)}\right]=\hfill & 2\left(\left[d^+\times s+s^+\times \stackrel{~}{d}\right]^{(2)}\frac{\sqrt{7}}{2}\left[d^+\times \stackrel{~}{d}\right]^{(2)}\right)\left(\left[d^+\times s+s^+\times \stackrel{~}{d}\right]^{(2)}\frac{\sqrt{7}}{2}\left[d^+\times \stackrel{~}{d}\right]^{(2)}\right)\hfill \\ & +\frac{15}{2}\left[d^+\times \stackrel{~}{d}\right]^{(1)}\left[d^+\times \stackrel{~}{d}\right]^{(1)},\hfill \\ C_2\left[\text{SO(7)}\right]=\hfill & \underset{K=1,3,5,7}{}\left[f^+\times \stackrel{~}{f}\right]^{(K)}\left[f^+\times \stackrel{~}{f}\right]^{(K)},\hfill \\ C_2\left[\text{SO(6)}\right]=\hfill & \left[d^+\times s+s^+\times \stackrel{~}{d}\right]^{(2)}\left[d^+\times s+s^+\times \stackrel{~}{d}\right]^{(2)}+2\underset{K=1,3}{}\left[d^+\times \stackrel{~}{d}\right]^{(K)}\left[d^+\times \stackrel{~}{d}\right]^{(K)},\hfill \\ C_2\left[\text{SO(5)}\right]=\hfill & \underset{K=1,3}{}\left[d^+\times \stackrel{~}{d}\right]^{(K)}\left[d^+\times \stackrel{~}{d}\right]^{(K)},\hfill \\ C_2\left[\text{SO}\text{d}\text{(3)}\right]=\hfill & 10\left[d^+\times \stackrel{~}{d}\right]^{(1)}\left[d^+\times \stackrel{~}{d}\right]^{(1)},\hfill \\ C_2\left[\text{SO}\text{f}\text{(3)}\right]=\hfill & 28\left[f^+\times \stackrel{~}{f}\right]^{(1)}\left[f^+\times \stackrel{~}{f}\right]^{(1)},\hfill \\ C_2\left[\text{G}_2\right]=\hfill & \underset{K=1,5}{}\left[f^+\times \stackrel{~}{f}\right]^{(K)}\left[f^+\times \stackrel{~}{f}\right]^{(K)}.\hfill \end{array}$$ (B1) The exceptional algebra G<sub>2</sub> has been introduced in physics by Racah . The eigenvalues of the Casimir invariants from (B1) in an irreducible representation given by a $`[f_1f_2\mathrm{}f_n]`$ Young tableau are $$\begin{array}{c}C_1\left[\text{U}(n)\right]=\underset{i=1}{\overset{n}{}}f_i,\hfill \\ C_2\left[\text{U}(n)\right]=\underset{i=1}{\overset{n}{}}f_i(f_i+n+12i),\hfill \\ C_2\left[\text{SU(3)}\right]=\lambda ^2+\mu ^2+\lambda \mu +3\lambda +3\mu ,(\lambda =f_1f_2,\mu =f_2)\hfill \\ C_2\left[\text{SO}(2n+1)\right]=\underset{i=1}{\overset{n}{}}2f_i(f_i+2n+12i),\hfill \\ C_2\left[\text{SO}(2n)\right]=\underset{i=1}{\overset{n}{}}2f_i(f_i+2n2i),\hfill \\ C_2\left[\text{G}_2\right]=f_1(f_1+5)+f_2(f_2+4)+f_1f_2.\hfill \end{array}$$ (B2)
warning/0002/astro-ph0002044.html
ar5iv
text
# 1 Origin of the Cosmic Background Radiation ## 1 Origin of the Cosmic Background Radiation Our present understanding of the beginning of the universe is based upon the remarkably successful theory of the Hot Big Bang. We believe that our universe began about 15 billion years ago as a hot, dense, nearly uniform sea of radiation a minute fraction of its present size (formally an infinitesimal singularity). If inflation occurred in the first fraction of a second, the universe became matter dominated while expanding exponentially and then returned to radiation domination by the reheating caused by the decay of the inflaton. Baryonic matter formed within the first second, and the nucleosynthesis of the lightest elements took only a few minutes as the universe expanded and cooled. The baryons were in the form of plasma until about 300,000 years after the Big Bang, when the universe had cooled to a temperature near 3000 K, sufficiently cool for protons to capture free electrons and form atomic hydrogen; this process is referred to as recombination. The recombination epoch occurred at a redshift of 1100, meaning that the universe has grown over a thousand times larger since then. The ionization energy of a hydrogen atom is 13.6 eV, but recombination did not occur until the universe had cooled to a characteristic temperature (kT) of 0.3 eV (Padmanabhan, 1993). This delay had several causes. The high entropy of the universe made the rate of electron capture only marginally faster than the rate of photodissociation. Moreover, each electron captured directly into the ground state emits a photon capable of ionizing another newly formed atom, so it was through recombination into excited states and the cooling of the universe to temperatures below the ionization energy of hydrogen that neutral matter finally condensed out of the plasma. Until recombination, the universe was opaque to electromagnetic radiation due to scattering of the photons by free electrons. As recombination occurred, the density of free electrons diminished greatly, leading to the decoupling of matter and radiation as the universe became transparent to light. The Cosmic Background Radiation (CBR) released during this era of decoupling has a mean free path long enough to travel almost unperturbed until the present day, where we observe it peaked in the microwave region of the spectrum as the Cosmic Microwave Background (CMB). We see this radiation today coming from the surface of last scattering (which is really a spherical shell of finite thickness) at a distance of nearly 15 billion light years. This Cosmic Background Radiation was predicted by the Hot Big Bang theory and discovered at an antenna temperature of 3K in 1964 by Penzias & Wilson (1965). The number density of photons in the universe at a redshift $`z`$ is given by (Peebles, 1993) $$n_\gamma =420(1+z)^3cm^3$$ (1) where $`(1+z)`$ is the factor by which the linear scale of the universe has expanded since then. The radiation temperature of the universe is given by $`T=T_0(1+z)`$ so it is easy to see how the conditions in the early universe at high redshifts were hot and dense. The CBR is our best probe into the conditions of the early universe. Theories of the formation of large-scale structure predict the existence of slight inhomogeneities in the distribution of matter in the early universe which underwent gravitational collapse to form galaxies, galaxy clusters, and superclusters. These density inhomogeneities lead to temperature anisotropies in the CBR due to a combination of intrinsic temperature fluctuations and gravitational blue/redshifting of the photons leaving under/overdense regions. The DMR (Differential Microwave Radiometer) instrument of the Cosmic Background Explorer (COBE) satellite discovered primordial temperature fluctuations on angular scales larger than $`7^{}`$ of order $`\mathrm{\Delta }T/T=10^5`$ (Smoot et al., 1992). Subsequent observations of the CMB have revealed temperature anisotropies on smaller angular scales which correspond to the physical scale of observed structures such as galaxies and clusters of galaxies. ### 1.1 Thermalization There were three main processes by which this radiation interacted with matter in the first few hundred thousand years: Compton scattering, double Compton scattering, and thermal bremsstrahlung. The simplest interaction of matter and radiation is Compton scattering of a single photon off a free electron, $`\gamma +e^{}\gamma +e^{}`$. The photon will transfer momentum and energy to the electron if it has significant energy in the electron’s rest frame. However, the scattering will be well approximated by Thomson scattering if the photon’s energy in the rest frame of the electron is significantly less than the rest mass, $`h\nu m_ec^2`$. When the electron is relativistic, the photon is blueshifted by roughly a factor $`\gamma `$ in energy when viewed from the electron rest frame, is then emitted at almost the same energy in the electron rest frame, and is blueshifted by another factor of $`\gamma `$ when retransformed to the observer’s frame. Thus, energetic electrons can efficiently transfer energy to the photon background of the universe. This process is referred to as Inverse Compton scattering. The combination of cases where the photon gives energy to the electron and vice versa allows Compton scattering to generate thermal equilibrium (which is impossible in the Thomson limit of elastic scattering). Compton scattering conserves the number of photons. There exists a similar process, double Compton scattering, which produces (or absorbs) photons, $`e^{}+\gamma e^{}+\gamma +\gamma `$. Another electromagnetic interaction which occurs in the plasma of the early universe is Coulomb scattering. Coulomb scattering establishes and maintains thermal equilibrium among the baryons of the photon-baryon fluid without affecting the photons. However, when electrons encounter ions they experience an acceleration and therefore emit electromagnetic radiation. This is called thermal bremsstrahlung or free-free emission. For an ion $`X`$, we have $`e^{}+Xe^{}+X+\gamma `$. The interaction can occur in reverse because of the ability of the charged particles to absorb incoming photons; this is called free-free absorption. Each charged particle emits radiation, but the acceleration is proportional to the mass, so we can usually view the electron as being accelerated in the fixed Coulomb field of the much heavier ion. Bremsstrahlung is dominated by electric-dipole radiation (Shu, 1991) and can also produce and absorb photons. The net effect is that Compton scattering is dominant for temperatures above 90 eV whereas bremsstrahlung is the primary process between 90 eV and 1 eV. At temperatures above 1 keV, double Compton is more efficient than bremsstrahlung. All three processes occur faster than the expansion of the universe and therefore have an impact until decoupling. A static solution for Compton scattering is the Bose-Einstein distribution, $$f_{BE}=\frac{1}{e^{x+\mu }1}$$ (2) where $`\mu `$ is a dimensionless chemical potential (Hu, 1995). At high optical depths, Compton scattering can exchange enough energy to bring the photons to this Bose-Einstein equilibrium distribution. A Planckian spectrum corresponds to zero chemical potential, which will occur only when the number of photons and total energy are in the same proportion as they would be for a blackbody. Thus, unless the photon number starts out exactly right in comparison to the total energy in radiation in the universe, Compton scattering will only produce a Bose-Einstein distribution and not a blackbody spectrum. It is important to note, however, that Compton scattering will preserve a Planck distribution, $$f_P=\frac{1}{e^x1}.$$ (3) All three interactions will preserve a thermal spectrum if one is achieved at any point. It has long been known that the expansion of the universe serves to decrease the temperature of a blackbody spectrum, $$B_\nu =\frac{2h\nu ^3/c^2}{e^{h\nu /kT}1},$$ (4) but keeps it thermal (Tolman, 1934). This occurs because both the frequency and temperature decrease as $`(1+z)`$ leaving $`h\nu /kT`$ unchanged during expansion. Although Compton scattering alone cannot produce a Planck distribution, such a distribution will remain unaffected by electromagnetic interactions or the universal expansion once it is achieved. A non-zero chemical potential will be reduced to zero by double Compton scattering and, later, bremsstrahlung which will create and absorb photons until the number density matches the energy and a thermal distribution of zero chemical potential is achieved. This results in the thermalization of the CBR at redshifts much greater than that of recombination. Thermalization, of course, should only be able to create an equilibrium temperature over regions that are in causal contact. The causal horizon at the time of last scattering was relatively small, corresponding to a scale today of about 200 Mpc, or a region of angular extent of one degree on the sky. However, observations of the CMB show that it has an isotropic temperature on the sky to the level of one part in one hundred thousand! This is the origin of the Horizon Problem, which is that there is no physical mechanism expected in the early universe which can produce thermodynamic equilibrium on superhorizon scales. The inflationary universe paradigm (Guth, 1981; Linde, 1982; Albrecht & Steinhardt, 1982) solves the Horizon Problem by postulating that the universe underwent a brief phase of exponential expansion during the first second after the Big Bang, during which our entire visible Universe expanded out of a region small enough to have already achieved thermal equilibrium. ## 2 CMB Spectrum The CBR is the most perfect blackbody ever seen, according to the FIRAS (Far InfraRed Absolute Spectrometer) instrument of COBE, which measured a temperature of $`T_0=2.726\pm 0.010`$ K (Mather et al., 1994). The theoretical prediction that the CBR will have a blackbody spectrum appears to be confirmed by the FIRAS observation (see Figure 1). But this is not the end of the story. FIRAS only observed the peak of the blackbody. Other experiments have mapped out the Rayleigh-Jeans part of the spectrum at low frequency. Most are consistent with a 2.73 K blackbody, but some are not. It is in the low-frequency limit that the greatest spectral distortions might occur because a Bose-Einstein distribution differs from a Planck distribution there. However, double Compton and bremsstrahlung are most effective at low frequencies so strong deviations from a blackbody spectrum are not generally expected. Spectral distortions in the Wien tail of the spectrum are quite difficult to detect due to the foreground signal from interstellar dust at those high frequencies. For example, broad emission lines from electron capture at recombination are predicted in the Wien tail but cannot be distinguished due to foreground contamination (White et al., 1994). However, because the energy generated by star formation and active galactic nuclei is absorbed by interstellar dust in all galaxies and then re-radiated in the far-infrared, we expect to see an isotropic Far-Infrared Background (FIRB) which dominates the CMB at frequencies above a few hundred GHz. This FIRB has now been detected in FIRAS data (Puget et al., 1996; Burigana & Popa, 1998; Fixsen et al., 1998) and in data from the COBE DIRBE instrument (Schlegel et al., 1998; Dwek et al., 1998). Although Compton, double Compton, and bremsstrahlung interactions occur frequently until decoupling, the complex interplay between them required to thermalize the CBR spectrum is ineffective at redshifts below $`10^7`$. This means that any process after that time which adds a significant portion of energy to the universe will lead to a spectral distortion today. Neutrino decays during this epoch should lead to a Bose-Einstein rather than a Planck distribution, and this allows the FIRAS observations to set constraints on the decay of neutrinos and other particles in the early universe (Kolb & Turner, 1990). The apparent impossibility of thermalizing radiation at low redshift makes the blackbody nature of the CBR strong evidence that it did originate in the early universe and as a result serves to support the Big Bang theory. The process of Compton scattering can cause spectral distortions if it is too late for double Compton and bremsstrahlung to be effective. In general, low-frequency photons will be shifted to higher frequencies, thereby decreasing the number of photons in the Rayleigh-Jeans region and enhancing the Wien tail. This is referred to as a Compton-y distortion and it is described by the parameter $$y=\frac{T_e(t)}{m_e}\sigma n_e(t)𝑑t.$$ (5) The apparent temperature drop in the long-wavelength limit is $$\frac{\delta T}{T}=2y.$$ (6) The most important example of this is Compton scattering of photons off hot electrons in galaxy clusters, called the Sunyaev-Zel’dovich (SZ) effect. The electrons transfer energy to the photons, and the spectral distortion results from the sum of all of the scatterings off electrons in thermal motion, each of which has a Doppler shift. The SZ effect from clusters can yield a distortion of $`y10^510^3`$ and these distortions have been observed in several rich clusters of galaxies. The FIRAS observations place a constraint on any full-sky Comptonization by limiting the average $`y`$-distortion to $`y<2.5\times 10^5`$ (Hu, 1995). The integrated $`y`$-distortion predicted from the SZ effect of galaxy clusters and large-scale structure is over a factor of ten lower than this observational constraint (Refregier et al., 1998) but that from “cocoons” of radio galaxies (Yamada et al., 1999) is predicted to be of the same order. A kinematic SZ effect is caused by the bulk velocity of the cluster; this is a small effect which is very difficult to detect for individual clusters but will likely be measured statistically by the Planck satellite. ## 3 CMB Anisotropy The temperature anisotropy at a point on the sky $`(\theta ,\varphi )`$ can be expressed in the basis of spherical harmonics as $$\frac{\mathrm{\Delta }T}{T}(\theta ,\varphi )=\underset{\mathrm{}m}{}a_\mathrm{}mY_\mathrm{}m(\theta ,\varphi ).$$ (7) A cosmological model predicts the variance of the $`a_\mathrm{}m`$ coefficients over an ensemble of universes (or an ensemble of observational points within one universe, if the universe is ergodic). The assumptions of rotational symmetry and Gaussianity allow us to express this ensemble average in terms of the multipoles $`C_{\mathrm{}}`$ as $$a_\mathrm{}m^{}a_\mathrm{}^{}m^{}C_{\mathrm{}}\delta _{\mathrm{}^{}\mathrm{}}\delta _{m^{}m}.$$ (8) The predictions of a cosmological model can be expressed in terms of $`C_{\mathrm{}}`$ alone if that model predicts a Gaussian distribution of density perturbations, in which case the $`a_\mathrm{}m`$ will have mean zero and variance $`C_{\mathrm{}}`$. The temperature anisotropies of the CMB detected by COBE are believed to result from inhomogeneities in the distribution of matter at the epoch of recombination. Because Compton scattering is an isotropic process in the electron rest frame, any primordial anisotropies (as opposed to inhomogeneities) should have been smoothed out before decoupling. This lends credence to the interpretation of the observed anisotropies as the result of density perturbations which seeded the formation of galaxies and clusters. The discovery of temperature anisotropies by COBE provides evidence that such density inhomogeneities existed in the early universe, perhaps caused by quantum fluctuations in the scalar field of inflation or by topological defects resulting from a phase transition (see Kamionkowski & Kosowsky, 1999 for a detailed review of inflationary and defect model predictions for CMB anisotropies). Gravitational collapse of these primordial density inhomogeneities appears to have formed the large-scale structures of galaxies, clusters, and superclusters that we observe today. On large (super-horizon) scales, the anisotropies seen in the CMB are produced by the Sachs-Wolfe effect (Sachs & Wolfe, 1967). $$\left(\frac{\mathrm{\Delta }T}{T}\right)_{SW}=𝐯𝐞|_o^e\mathrm{\Phi }|_o^e+\frac{1}{2}_o^eh_{\rho \sigma ,0}n^\rho n^\sigma 𝑑\xi ,$$ (9) where the first term is the net Doppler shift of the photon due to the relative motion of emitter and observer, which is referred to as the kinematic dipole. This dipole, first observed by Smoot et al. (1977), is much larger than other CMB anisotropies and is believed to reflect the motion of the Earth relative to the average reference frame of the CMB. Most of this motion is due to the peculiar velocity of the Local Group of galaxies. The second term represents the gravitational redshift due to a difference in gravitational potential between the site of photon emission and the observer. The third term is called the Integrated Sachs-Wolfe (ISW) effect and is caused by a non-zero time derivative of the metric along the photon’s path of travel due to potential decay, gravitational waves, or non-linear structure evolution (the Rees-Sciama effect). In a matter-dominated universe with scalar density perturbations the integral vanishes on linear scales. This equation gives the redshift from emission to observation, but there is also an intrinsic $`\mathrm{\Delta }T/T`$ on the last-scattering surface due to the local density of photons. For adiabatic perturbations, we have (White & Hu, 1997) an intrinsic $$\frac{\mathrm{\Delta }T}{T}=\frac{1}{3}\frac{\delta \rho }{\rho }=\frac{2}{3}\mathrm{\Phi }.$$ (10) Putting the observer at $`\mathrm{\Phi }=0`$ (the observer’s gravitational potential merely adds a constant energy to all CMB photons) this leads to a net Sachs-Wolfe effect of $`\mathrm{\Delta }T/T=\mathrm{\Phi }/3`$ which means that overdensities lead to cold spots in the CMB. ### 3.1 Small-angle anisotropy Anisotropy measurements on small angular scales ($`0.^{}1`$ to $`1^{}`$) are expected to reveal the so-called first acoustic peak of the CMB power spectrum. This peak in the anisotropy power spectrum corresponds to the scale where acoustic oscillations of the photon-baryon fluid caused by primordial density inhomogeneities are just reaching their maximum amplitude at the surface of last scattering i.e. the sound horizon at recombination. Further acoustic peaks occur at scales that are reaching their second, third, fourth, etc. antinodes of oscillation. Figure 2 (from Hu et al., 1997) shows the dependence of the CMB anisotropy power spectrum on a number of cosmological parameters. The acoustic oscillations in density (light solid line) are sharp here because they are really being plotted against spatial scales, which are then smoothed as they are projected through the last-scattering surface onto angular scales. The troughs in the density oscillations are filled in by the 90-degree-out-of-phase velocity oscillations (this is a Doppler effect but does not correspond to the net peaks, which are best referred to as acoustic peaks rather than Doppler peaks). The origin of this plot is at a different place for different values of the matter density and the cosmological constant; the negative spatial curvature of an open universe makes a given spatial scale correspond to a smaller angular scale. The Integrated Sachs-Wolfe (ISW) effect occurs whenever gravitational potentials decay due to a lack of matter dominance. Hence the early ISW effect occurs just after recombination when the density of radiation is still considerable and serves to broaden the first acoustic peak at scales just larger than the horizon size at recombination. And for a present-day matter density less than critical, there is a late ISW effect that matters on very large angular scales - it is greater in amplitude for open universes than for lambda-dominated because matter domination ends earlier in an open universe for the same value of the matter density today. The late ISW effect should correlate with large-scale structures that are otherwise detectable at $`z1`$, and this allows the CMB to be cross-correlated with observations of the X-ray background to determine $`\mathrm{\Omega }`$ (Crittenden & Turok, 1996; Kamionkowski, 1996; Boughn et al., 1998; Kamionkowski & Kinkhabwala, 1999) or with observations of large-scale structure to determine the bias of galaxies (Suginohara et al., 1998). For a given model, the location of the first acoustic peak can yield information about $`\mathrm{\Omega }`$, the ratio of the density of the universe to the critical density needed to stop its expansion. For adiabatic density perturbations, the first acoustic peak will occur at $`\mathrm{}=220\mathrm{\Omega }^{1/2}`$ (Kamionkowski et al., 1994). The ratio of $`\mathrm{}`$ values of the peaks is a robust test of the nature of the density perturbations; for adiabatic perturbations these will have ratio 1:2:3:4 whereas for isocurvature perturbations the ratio should be 1:3:5:7 (Hu & White, 1996). A mixture of adiabatic and isocurvature perturbations is possible, and this test should reveal it. As illustrated in Figure 2, the amplitude of the acoustic peaks depends on the baryon fraction $`\mathrm{\Omega }_b`$, the matter density $`\mathrm{\Omega }_0`$, and Hubble’s constant $`H_0=100h`$ km/s/Mpc. A precise measurement of all three acoustic peaks can reveal the fraction of hot dark matter and even potentially the number of neutrino species (Dodelson et al., 1996). Figure 2 shows the envelope of the CMB anisotropy damping tail on arcminute scales, where the fluctuations are decreased due to photon diffusion (Silk, 1967) as well as the finite thickness of the last-scattering surface. This damping tail is a sensitive probe of cosmological parameters and has the potential to break degeneracies between models which explain the larger-scale anisotropies (Hu & White, 1997b; Metcalf & Silk, 1998). The characteristic angular scale for this damping is given by $`1.8^{}\mathrm{\Omega }_B^{1/2}\mathrm{\Omega }_0^{3/4}h^{1/2}`$ (White et al., 1994). There is now a plethora of theoretical models which predict the development of primordial density perturbations into microwave background anisotropies. These models differ in their explanation of the origin of density inhomogeneities (inflation or topological defects), the nature of the dark matter (hot, cold, baryonic, or a mixture of the three), the curvature of the universe ($`\mathrm{\Omega }`$), the value of the cosmological constant ($`\mathrm{\Lambda }`$), the value of Hubble’s constant, and the possibility of reionization which wholly or partially erased temperature anisotropies in the CMB on scales smaller than the horizon size. Available data does not allow us to constrain all (or even most) of these parameters, so analyzing current CMB anisotropy data requires a model-independent approach. It seems reasonable to view the mapping of the acoustic peaks as a means of determining the nature of parameter space before going on to fitting cosmological parameters directly. ### 3.2 Reionization The possibility that post-decoupling interactions between ionized matter and the CBR have affected the anisotropies on scales smaller than those measured by COBE is of great significance for current experiments. Reionization is inevitable but its effect on anisotropies depends significantly on when it occurs (see Haimann & Knox, 1999 for a review). Early reionization leads to a larger optical depth and therefore a greater damping of the anisotropy power spectrum due to the secondary scattering of CMB photons off of the newly free electrons. For a universe with critical matter density and constant ionization fraction $`x_e`$, the optical depth as a function of redshift is given by (White et al., 1994) $$\tau 0.035\mathrm{\Omega }_Bhx_ez^{3/2},$$ (11) which allows us to determine the redshift of reionization $`z_{}`$ at which $`\tau =1`$, $$z_{}69\left(\frac{h}{0.5}\right)^{\frac{2}{3}}\left(\frac{\mathrm{\Omega }_B}{0.1}\right)^{\frac{2}{3}}x_e^{\frac{2}{3}}\mathrm{\Omega }^{\frac{1}{3}},$$ (12) where the scaling with $`\mathrm{\Omega }`$ applies to an open universe only. At scales smaller than the horizon size at reionization, $`\mathrm{\Delta }T/T`$ is reduced by the factor $`e^\tau `$. Attempts to measure the temperature anisotropy on angular scales of less than a degree which correspond to the size of galaxies could have led to a surprise; if the universe was reionized after recombination to the extent that the CBR was significantly scattered at redshifts less than 1100, the small-scale primordial anisotropies would have been washed out. To have an appreciable optical depth for photon-matter interaction, reionization cannot have occurred much later than a redshift of 20 (Padmanabhan, 1993). Large-scale anisotropies such as those seen by COBE are not expected to be affected by reionization because they encompass regions of the universe which were not yet in causal contact even at the proposed time of reionization. However, the apparently high amplitiude of degree-scale anisotropies is a strong argument against the possibility of early ($`z50`$) reionization. On arc-minute scales, the interaction of photons with reionized matter is expected to have eliminated the primordial anisotropies and replaced them with smaller secondary anisotropies from this new surface of last scattering (the Ostriker-Vishniac effect and patchy reionization, see next section). ### 3.3 Secondary Anisotropies Secondary CMB anisotropies occur when the photons of the Cosmic Microwave Background radiation are scattered after the original last-scattering surface (see Refregier, 1999 for a review). The shape of the blackbody spectrum can be altered through inverse Compton scattering by the thermal Sunyaev-Zel’dovich (SZ) effect (Sunyaev & Zeldovich, 1972). The effective temperature of the blackbody can be shifted locally by a doppler shift from the peculiar velocity of the scattering medium (the kinetic SZ and Ostriker-Vishniac effects, Ostriker & Vishniac, 1986) as well as by passage through the changing gravitational potential caused by the collapse of nonlinear structure (the Rees-Sciama effect, Rees & Sciama, 1968) or the onset of curvature or cosmological constant domination (the Integrated Sachs-Wolfe effect). Several simulations of the impact of patchy reionization have been performed (Aghanim et al., 1996; Knox et al., 1998; Gruzinov & Hu, 1998; Peebles & Juszkiewicz, 1998). The SZ effect itself is independent of redshift, so it can yield information on clusters at much higher redshift than does X-ray emission. However, nearly all clusters are unresolved for $`10^{}`$ resolution so higher-redshift clusters occupy less of the beam and therefore their SZ effect is in fact dimmer. In the 4.5 channels of Planck this will no longer be true, and the SZ effect can probe cluster abundance at high redshift. An additional secondary anisotropy is that caused by gravitational lensing (see e.g. Cayon et al., 1993, 1994; Metcalf & Silk, 1997; Martinez-Gonzalez et al., 1997). Gravitational lensing imprints slight non-Gaussianity in the CMB from which it might be possible to determine the matter power spectrum (Seljak & Zaldarriaga, 1998; Zaldarriaga & Seljak, 1998). ### 3.4 Polarization Anisotropies Polarization of the Cosmic Microwave Background radiation (Kosowsky, 1994; Kamionkowski et al., 1997; Zaldarriaga & Seljak, 1997) arises due to local quadrupole anisotropies at each point on the surface of last scattering (see Hu & White, 1997a for a review). Scalar (density) perturbations generate curl-free (electric mode) polarization only, but tensor (gravitational wave) perturbations can generate divergence-free (magnetic mode) polarization. Hence the polarization of the CMB is a potentially useful probe of the level of gravitational waves in the early universe (Seljak & Zaldarriaga, 1997; Kamionkowski & Kosowsky, 1998), especially since current indications are that the large-scale primary anisotropies seen by COBE do not contain a measurable fraction of tensor contributions (Gawiser & Silk, 1998). A thorough review of gravity waves and CMB polarization is given by Kamionkowski & Kosowsky (1999). ### 3.5 Gaussianity of the CMB anisotropies The processes turning density inhomogeneities into CMB anisotropies are linear, so cosmological models that predict gaussian primordial density inhomogeneities also predict a gaussian distribution of CMB temperature fluctuations. Several techniques have been developed to test COBE and future datasets for deviations from gaussianity (e.g. Kogut et al., 1996b; Ferreira & Magueijo, 1997; Ferreira et al., 1997). Most tests have proven negative, but a few claims of non-gaussianity have been made. Gaztañaga et al. (1998) found a very marginal indication of non-gaussianity in the spread of results for degree-scale CMB anisotropy observations being greater than the expected sample variances. Ferreira et al. (1998) have claimed a detection of non-gaussianity at multipole $`\mathrm{}=16`$ using a bispectrum statistic, and Pando et al. (1998) find a non-gaussian wavelet coefficient correlation on roughly $`15^{}`$ scales in the North Galactic hemisphere. Both of these methods produce results consistent with gaussianity, however, if a particular area of several pixels is eliminated from the dataset (Bromley & Tegmark, 1999). A true sky signal should be larger than several pixels so instrument noise is the most likely source of the non-gaussianity. A different area appears to cause each detection, giving evidence that the COBE dataset had non-gaussian instrument noise in at least two areas of the sky. ### 3.6 Foreground contamination Of particular concern in measuring CMB anisotropies is the issue of foreground contamination. Foregrounds which can affect CMB observations include galactic radio emission (synchrotron and free-free), galactic infrared emission (dust), extragalactic radio sources (primarily elliptical galaxies, active galactic nuclei, and quasars), extragalactic infrared sources (mostly dusty spirals and high-redshift starburst galaxies), and the Sunyaev-Zel’dovich effect from hot gas in galaxy clusters. The COBE team has gone to great lengths to analyze their data for possible foreground contamination and routinely eliminates everything within about $`30^{}`$ of the galactic plane. An instrument with large resolution such as COBE is most sensitive to the diffuse foreground emission of our Galaxy, but small-scale anisotropy experiments need to worry about extragalactic sources as well. Because foreground and CMB anisotropies are assumed to be uncorrelated, they should add in quadrature, leading to an increase in the measurement of CMB anisotropy power. Most CMB instruments, however, can identify foregrounds by their spectral signature across multiple frequencies or their display of the beam response characteristic of a point source. This leads to an attempt at foreground subtraction, which can cause an underestimate of CMB anisotropy if some true signal is subtracted along with the foreground. Because they are now becoming critical, extragalactic foregrounds have been studied in detail (Toffolatti et al., 1998; Refregier et al., 1998; Gawiser & Smoot, 1997; Sokasian et al., 1998; Gawiser et al., 1998). The Wavelength-Oriented Microwave Background Analysis Team (WOMBAT, see Gawiser et al., 1998; Jaffe et al., 1999) has made Galactic and extragalactic foreground predictions and full-sky simulations of realistic CMB skymaps containing foreground contamination available to the public (see http://astro.berkeley.edu/wombat). One of these CMB simulations is shown in Figure 3. Tegmark et al. (1999) used a Fisher matrix analysis to show that simultaneously estimating foreground model parameters and cosmological parameters can lead to a factor of a few degradation in the precision with which the cosmological parameters can be determined by CMB anisotropy observations, so foreground prediction and subtraction is likely to be an important aspect of future CMB data analysis. Foreground contamination may turn out to be a serious problem for measurements of CMB polarization anisotropy. While free-free emission is unpolarized, synchrotron radiation displays a linear polarization determined by the coherence of the magnetic field along the line of sight; this is typically on the order of 10% for Galactic synchrotron and between 5 and 10% for flat-spectrum radio sources. The CMB is expected to show a large-angular scale linear polarization of about 10%, so the prospects for detecting polarization anisotropy are no worse than for temperature anisotropy although higher sensitivity is required. However, the small-angular scale electric mode of linear polarization which is a probe of several cosmological parameters and the magnetic mode that serves as a probe of tensor perturbations are expected to have much lower amplitude and may be swamped by foreground polarization. Thermal and spinning dust grain emission can also be polarized. It may turn out that dust emission is the only significant source of circularly polarized microwave photons since the CMB cannot have circular polarization. ## 4 Cosmic Microwave Background Anisotropy Observations Since the COBE DMR detection of CMB anisotropy (Smoot et al., 1992), there have been over thirty additional measurements of anisotropy on angular scales ranging from $`7^{}`$ to $`0.^{}3`$, and upper limits have been set on smaller scales. The COBE DMR observations were pixelized into a skymap, from which it is possible to analyze any particular multipole within the resolution of the DMR. Current small angular scale CMB anisotropy observations are insensitive to both high $`\mathrm{}`$ and low $`\mathrm{}`$ multipoles because they cannot measure features smaller than their resolution and are insensitive to features larger than the size of the patch of sky observed. The next satellite mission, NASA’s Microwave Anisotropy Probe (MAP), is scheduled for launch in Fall 2000 and will map angular scales down to $`0.^{}2`$ with high precision over most of the sky. An even more precise satellite, ESA’s Planck, is scheduled for launch in 2007. Because COBE observed such large angles, the DMR data can only constrain the amplitude $`A`$ and index $`n`$ of the primordial power spectrum in wave number $`k`$, $`P_p(k)=Ak^n`$, and these constraints are not tight enough to rule out very many classes of cosmological models. Until the next satellite is flown, the promise of microwave background anisotropy measurements to measure cosmological parameters rests with a series of ground-based and balloon-borne anisotropy instruments which have already published results (shown in Figure 4) or will report results in the next few years (MAXIMA, BOOMERANG, TOPHAT, ACE, MAT, VSA, CBI, DASI, see Lee et al., 1999 and Halpern & Scott, 1999). Because they are not satellites, these instruments face the problems of shorter observing times and less sky coverage, although significant progress has been made in those areas. They fall into three categories: high-altitude balloons, interferometers, and other ground-based instruments. Past, present, and future balloon-borne instruments are FIRS, MAX, MSAM, ARGO, BAM, MAXIMA, QMAP, HACME, BOOMERANG, TOPHAT, and ACE. Ground-based interferometers include CAT, JBIAC, SUZIE, BIMA, ATCA, VLA, VSA, CBI, and DASI, and other ground-based instruments are TENERIFE, SP, PYTHON, SK, OVRO/RING, VIPER, MAT/TOCO, IACB, and WD. Taken as a whole, they have the potential to yield very useful measurements of the radiation power spectrum of the CMB on degree and subdegree scales. Ground-based non-interferometers have to discard a large fraction of data and undergo careful further data reduction to eliminate atmospheric contamination. Balloon-based instruments need to keep a careful record of their pointing to reconstruct it during data analysis. Interferometers may be the most promising technique at present but they are the least developed, and most instruments are at radio frequencies and have very narrow frequency coverage, making foreground contamination a major concern. In order to use small-scale CMB anisotropy measurements to constrain cosmological models we need to be confident of their validity and to trust the error bars. This will allow us to discard badly contaminated data and to give greater weight to the more precise measurements in fitting models. Correlated noise is a great concern for instruments which lack a rapid chopping because the $`1/f`$ noise causes correlations on scales larger than the beam in a way that can easily mimic CMB anisotropies. Additional issues are sample variance caused by the combination of cosmic variance and limited sky coverage and foreground contamination. Figure 4 shows our compilation of CMB anisotropy observations without adding any theoretical curves to bias the eye<sup>2</sup><sup>2</sup>2This figure and our compilation of CMB anisotropy observations are available at http://mamacass.ucsd.edu/people/gawiser/cmb.html; CMB observations have also been compiled by Smoot & Scott (1998) and at http://www.hep.upenn.edu/~max/cmb/experiments.html and http://www.cita.utoronto.ca/~knox/radical.html. It is clear that a straight line is a poor but not implausible fit to the data. There is a clear rise around $`\mathrm{}=100`$ and then a drop by $`\mathrm{}=1000`$. This is not yet good enough to give a clear determination of the curvature of the universe, let alone fit several cosmological parameters. However, the current data prefer adiabatic structure formation models over isocurvature models (Gawiser & Silk, 1998). If analysis is restricted to adiabatic CDM models, a value of the total density near critical is preferred (Dodelson & Knox, 1999). ### 4.1 Window Functions The sensitivity of these instruments to various multipoles is called their window function. These window functions are important in analyzing anisotropy measurements because the small-scale experiments do not measure enough of the sky to produce skymaps like COBE. Rather they yield a few “band-power” measurements of rms temperature anisotropy which reflect a convolution over the range of multipoles contained in the window function of each band. Some instruments can produce limited skymaps (White & Bunn, 1995). The window function $`W_{\mathrm{}}`$ shows how the total power observed is sensitive to the anisotropy on the sky as a function of angular scale: $$Power=\frac{1}{4\pi }\underset{\mathrm{}}{}(2\mathrm{}+1)C_{\mathrm{}}W_{\mathrm{}}=\frac{1}{2}(\mathrm{\Delta }T/T_{CMB})^2\underset{\mathrm{}}{}\frac{2\mathrm{}+1}{\mathrm{}(\mathrm{}+1)}W_{\mathrm{}}$$ (13) where the COBE normalization is $`\mathrm{\Delta }T=27.9\mu `$K and $`T_{CMB}=2.73`$K (Bennett et al., 1996). This allows the observations of broad-band power to be reported as observations of $`\mathrm{\Delta }T`$, and knowing the window function of an instrument one can turn the predicted $`C_{\mathrm{}}`$ spectrum of a model into the corresponding prediction for $`\mathrm{\Delta }T`$. This “band-power” measurement is based on the standard definition that for a “flat” power spectrum, $`\mathrm{\Delta }T=(\mathrm{}(\mathrm{}+1)C_{\mathrm{}})^{1/2}T_{CMB}/(2\pi )`$ (flat actually means that $`\mathrm{}(\mathrm{}+1)C_{\mathrm{}}`$ is constant). The autocorrelation function for measured temperature anisotropies is a convolution of the true expectation values for the anisotropies and the window function. Thus we have (White & Srednicki, 1995) $$\frac{\mathrm{\Delta }T}{T}(\widehat{n}_1)\frac{\mathrm{\Delta }T}{T}(\widehat{n}_2)=\frac{1}{4\pi }\underset{\mathrm{}=1}{\overset{\mathrm{}}{}}(2\mathrm{}+1)C_{\mathrm{}}W_{\mathrm{}}(\widehat{n}_1,\widehat{n}_2),$$ (14) where the symmetric beam shape that is typically assumed makes $`W_{\mathrm{}}`$ a function of separation angle only. In general, the window function results from a combination of the directional response of the antenna, the beam position as a function of time, and the weighting of each part of the beam trajectory in producing a temperature measurement (White & Srednicki, 1995). Strictly speaking, $`W_{\mathrm{}}`$ is the diagonal part of a filter function $`W_{\mathrm{}\mathrm{}^{}}`$ that reflects the coupling of various multipoles due to the non-orthogonality of the spherical harmonics on a cut sky and the observing strategy of the instrument (Knox, 1999). It is standard to assume a Gaussian beam response of width $`\sigma `$, leading to a window function $$W_{\mathrm{}}=\mathrm{exp}[\mathrm{}(\mathrm{}+1)\sigma ^2].$$ (15) The low-$`\mathrm{}`$ cutoff introduced by a 2-beam differencing setup comes from the window function (White et al., 1994) $$W_{\mathrm{}}=2[1P_{\mathrm{}}(\mathrm{cos}\theta )]\mathrm{exp}[\mathrm{}(\mathrm{}+1)\sigma ^2].$$ (16) ### 4.2 Sample and Cosmic Variance The multipoles $`C_{\mathrm{}}`$ can be related to the expected value of the spherical harmonic coefficients by $$\underset{m}{}a_\mathrm{}m^2=(2\mathrm{}+1)C_{\mathrm{}}$$ (17) since there are $`(2\mathrm{}+1)`$ $`a_\mathrm{}m`$ for each $`\mathrm{}`$ and each has an expected autocorrelation of $`C_{\mathrm{}}`$. In a theory such as inflation, the temperature fluctuations follow a Gaussian distribution about these expected ensemble averages. This makes the $`a_\mathrm{}m`$ Gaussian random variables, resulting in a $`\chi _{2\mathrm{}+1}^2`$ distribution for $`_ma_\mathrm{}m^2`$. The width of this distribution leads to a cosmic variance in the estimated $`C_{\mathrm{}}`$ of $`\sigma _{cv}^2=(\mathrm{}+\frac{1}{2})^{\frac{1}{2}}C_{\mathrm{}}`$, which is much greater for small $`\mathrm{}`$ than for large $`\mathrm{}`$ (unless $`C_{\mathrm{}}`$ is rising in a manner highly inconsistent with theoretical expectations). So, although cosmic variance is an unavoidable source of error for anisotropy measurements, it is much less of a problem for small scales than for COBE. Despite our conclusion that cosmic variance is a greater concern on large angular scales, Figure 4 shows a tremendous variation in the level of anisotropy measured by small-scale experiments. Is this evidence for a non-Gaussian cosmological model such as topological defects? Does it mean we cannot trust the data? Neither conclusion is justified (although both could be correct) because we do in fact expect a wide variation among these measurements due to their coverage of a very small portion of the sky. Just as it is difficult to measure the $`C_{\mathrm{}}`$ with only a few $`a_\mathrm{}m`$, it is challenging to use a small piece of the sky to measure multipoles whose spherical harmonics cover the sphere. It turns out that limited sky coverage leads to a sample variance for a particular multipole related to the cosmic variance for any value of $`\mathrm{}`$ by the simple formula $$\sigma _{sv}^2\left(\frac{4\pi }{\mathrm{\Omega }}\right)\sigma _{cv}^2,$$ (18) where $`\mathrm{\Omega }`$ is the solid angle observed (Scott et al., 1994). One caveat: in testing cosmological models, this cosmic and sample variance should be derived from the $`C_{\mathrm{}}`$ of the model, not the observed value of the data. The difference is typically small but will bias the analysis of forthcoming high-precision observations if cosmic and sample variance are not handled properly. ### 4.3 Binning CMB data Because there are so many measurements and the most important ones have the smallest error bars, it is preferable to plot the data in some way that avoids having the least precise measurements dominate the plot. Quantitative analyses should weight each datapoint by the inverse of its variance. Binning the data can be useful for display purposes but is dangerous for analysis, because a statistical analysis performed on the binned datapoints will give different results from one performed on the raw data. The distribution of the binned errors is non-Gaussian even if the original points had Gaussian errors. Binning might improve a quantitative analysis if the points at a particular angular scale showed a scatter larger than is consistent with their error bars, leading one to suspect that the errors have been underestimated. In this case, one could use the scatter to create a reasonable uncertainty on the binned average. For the current CMB data there is no clear indication of scatter inconsistent with the errors so this is unnecessary. If one wishes to perform a model-dependent analysis of the data, the simplest reasonable approach is to compare the observations with the broad-band power estimates that should have been produced given a particular theory (the theory’s $`C_{\mathrm{}}`$ are not constant so the window functions must be used for this). Combining full raw datasets is superior but computationally intensive (see Bond et al. 1998a). A first-order correction for the non-gaussianity of the likelihood function of the band-powers has been calculated by Bond et al. (1998b) and is available at http://www.cita.utoronto.ca/~knox/radical.html. ## 5 Combining CMB and Large-Scale Structure Observations As CMB anisotropy is detected on smaller angular scales and large-scale structure surveys extend to larger regions, there is an increasing overlap in the spatial scale of inhomogeneities probed by these complementary techniques. This allows us to test the gravitational instability paradigm in general and then move on to finding cosmological models which can simultaneously explain the CMB and large-scale structure observations. Figure 5 shows this comparison for our compilation of CMB anisotropy observations (colored boxes) and of large-scale structure surveys (APM - Gaztañaga & Baugh 1998, LCRS - Lin et al. 1996, Cfa2+SSRS2 - Da Costa et al. 1994, PSCZ - Tadros et al. 1999, APM clusters - Tadros et al. 1998) including measurements of the dark matter fluctuations from peculiar velocities (Kolatt & Dekel, 1997) and the abundance of galaxy clusters (Viana & Liddle, 1996; Bahcall et al., 1997). Plotting CMB anisotropy data as measurements of the matter power spectrum is a model-dependent procedure, and the galaxy surveys must be corrected for redshift distortions, non-linear evolution, and galaxy bias (see Gawiser & Silk 1998 for detailed methodology.) Figure 5 is good evidence that the matter and radiation inhomogeneities had a common origin - the standard $`\mathrm{\Lambda }`$CDM model with a Harrison-Zel’dovich primordial power spectrum predicts both rather well. On the detail level, however, the model is a poor fit ($`\chi ^2`$/d.o.f.=2.1), and no cosmological model which is consistent with the recent Type Ia supernovae results fits the data much better. Future observations will tell us if this is evidence of systematic problems in large-scale structure data or a fatal flaw of the $`\mathrm{\Lambda }`$CDM model. ## 6 Conclusions The CMB is a mature subject. The spectral distortions are well understood, and the Sunyaev-Zeldovich effect provides a unique tool for studying galaxy clusters at high redshift. Global distortions will eventually be found, most likely first at very large $`l`$ due to the cumulative contributions from hot gas heated by radio galaxies, AGN, and galaxy groups and clusters. For gas at $`10^610^7`$ K, appropriate to gas in galaxy potential wells, the thermal and kinematic contributions are likely to be comparable. CMB anisotropies are a rapidly developing field, since the 1992 discovery with the COBE DMR of large angular scale temperature fluctuations. At the time of writing, the first acoustic peak is being mapped with unprecedented precision that will enable definitive estimates to be made of the curvature parameter. More information will come with all-sky surveys to higher resolution (MAP in 2000, PLANCK in 2007) that will enable most of the cosmological parameters to be derived to better than a few percent precision if the adiabatic CDM paradigm proves correct. Degeneracies remain in CMB parameter extraction, specifically between $`\mathrm{\Omega }_0`$, $`\mathrm{\Omega }_b`$ and $`\mathrm{\Omega }_\mathrm{\Lambda }`$, but these can be removed via large-scale structure observations, which effectively constrain $`\mathrm{\Omega }_\mathrm{\Lambda }`$ via weak lensing. The goal of studying reionization will be met by the interferometric surveys at very high resolution ($`l10^310^4`$). Polarization presents the ultimate challenge, because the foregrounds are poorly known. Experiments are underway to measure polarization at the 10 percent level, expected on degree scales in the most optimistic models. However one has to measure polarisation at the 1 percent level to definitively study the ionization history and early tensor mode generation in the universe, and this may only be possible with long duration balloon or space experiments. CMB anisotropies are a powerful probe of the early universe. Not only can one hope to extract the cosmological parameters, but one should be able to measure the primordial power spectrum of density fluctuations laid down at the epoch of inflation, to within the uncertainties imposed by cosmic variance. In combination with new generations of deep wide field galaxy surveys, it should be possible to unambiguously measure the shape of the predicted peak in the power spectrum, and thereby establish unique constraints on the origin of the large-scale structure of the universe.
warning/0002/math-ph0002031.html
ar5iv
text
# References Linear Odd Poisson Bracket on Grassmann Algebra Vyacheslav A. Soroka<sup>*</sup><sup>*</sup>*E-mail: vsoroka@kipt.kharkov.ua Kharkov Institute of Physics and Technology 310108 Kharkov, Ukraine ## Abstract A linear odd Poisson bracket realized solely in terms of Grassmann variables is suggested. It is revealed that with the bracket, corresponding to a semi-simple Lie group, both a Grassmann-odd Casimir function and invariant (with respect to this group) nilpotent differential operators of the first, second and third orders are naturally related and enter into a finite-dimensional Lie superalgebra. 1. Recently a linear degenerate odd Poisson bracket built only of Grassmann variables has been introduced . It was constructed for this bracket, in contrast with the non-degenerate odd bracket having the only Grassmann-odd nilpotent differential $`\mathrm{\Delta }`$-operator of the second order, at once three Grassmann-odd nilpotent $`\mathrm{\Delta }`$-like differential operators of the first, the second and the third orders with respect to Grassmann derivatives. It was also shown that these $`\mathrm{\Delta }`$-like operators together with a Grassmann-odd nilpotent Casimir function of this degenerate odd bracket form a finite-dimensional Lie superalgebra. Following to , in the present report we extend the above-mentioned results to the case of an arbitrary linear odd Poisson bracket, which is also realized solely in terms of the Grassmann variables and corresponds to a semi-simple Lie group. 2. There is a well-known linear even Poisson bracket given in terms of the commuting (Grassmann-even) variables $`X_\alpha `$ $`(g(X_\alpha )=0)`$ $$\{X_\alpha ,X_\beta \}_0=c_{\alpha \beta }^{}{}_{}{}^{\gamma }X_\gamma ,(\alpha ,\beta ,\gamma =1,\mathrm{},N),$$ $`\left(1\right)`$ where $`c_{\alpha \beta }^{}{}_{}{}^{\gamma }`$ are Grassmann-even $`(g(c_{\alpha \beta }^{}{}_{}{}^{\gamma })=0)`$ constants which, because of the main properties of the even Poisson bracket: $$\{A,B+C\}_0=\{A,B\}_0+\{A,C\}_0,$$ $`\left(2\right)`$ $$g\left(\{A,B\}_0\right)=g\left(A\right)+g\left(B\right)\left(mod2\right),$$ $`\left(3\right)`$ $$\{A,B\}_0=\left(1\right)^{g\left(A\right)g\left(B\right)}\{B,A\}_0,$$ $`\left(4\right)`$ $$\underset{\left(ABC\right)}{}\left(1\right)^{g\left(A\right)g\left(C\right)}\{A,\{B,C\}_0\}_0=0,$$ $`\left(5\right)`$ $$\{A,BC\}_0=\{A,B\}_0C+\left(1\right)^{g\left(A\right)g\left(B\right)}B\{A,C\}_0,$$ $`\left(6\right)`$ are antisymmetric in the two lower indices $$c_{\alpha \beta }^{}{}_{}{}^{\gamma }=c_{\beta \alpha }^{}{}_{}{}^{\gamma }$$ $`\left(7\right)`$ and obey the conditions $$c_{\alpha \lambda }^{}{}_{}{}^{\delta }c_{\beta \gamma }^{}{}_{}{}^{\lambda }+c_{\beta \lambda }^{}{}_{}{}^{\delta }c_{\gamma \alpha }^{}{}_{}{}^{\lambda }+c_{\gamma \lambda }^{}{}_{}{}^{\delta }c_{\alpha \beta }^{}{}_{}{}^{\lambda }=0.$$ $`\left(8\right)`$ A sum with the symbol $`(ABC)`$ in (5) means a summation over cyclic permutations of the quantities $`A,B,C`$. In relations (2)-(6) $`A,B,C`$ are functions of the variables $`X_\alpha `$ and $`g(A)`$ is a Grassmann parity of the quantity $`A`$. The linear even bracket (1) plays a very important role in the theory of Lie groups, Lie algebras, their representations and applications (see, for example, ). The bracket (1) can be realized in a canonical even Poisson bracket $$\{A,B\}_0=A\underset{\alpha =1}{\overset{N}{}}(\underset{q^\alpha }{\overset{}{}}\underset{p_\alpha }{\overset{}{}}\underset{p_\alpha }{\overset{}{}}\underset{q^\alpha }{\overset{}{}})B,$$ where $`\stackrel{}{}`$ and $`\stackrel{}{}`$ are the right and left derivatives and $`_{x^A}\frac{}{x^A}`$, on the following bilinear functions of coordinates $`q^\alpha `$ and momenta $`p_\alpha `$ $$X_\alpha =c_{\alpha \beta }^{}{}_{}{}^{\gamma }q^\beta p_\gamma $$ if $`c_{\alpha \beta }^{}{}_{}{}^{\gamma }`$ satisfy the conditions (7), (8) for the structure constants of a Lie group. As in the Lie algebra case, we can define a symmetric Cartan-Killing tensor $$g_{\alpha \beta }=g_{\beta \alpha }=c_{\alpha \gamma }^{}{}_{}{}^{\lambda }c_{\beta \lambda }^{}{}_{}{}^{\gamma }$$ $`\left(9\right)`$ and verify with the use of relations (8) an anti-symmetry property of a tensor $$c_{\alpha \beta \gamma }=c_{\alpha \beta }^{}{}_{}{}^{\delta }g_{\delta \gamma }=c_{\alpha \gamma \beta }.$$ $`\left(10\right)`$ By assuming that the Cartan-Killing metric tensor is non-degenerate $`det(g_{\alpha \beta })0`$ (this case corresponds to the semi-simple Lie group), we can define an inverse tensor $`g^{\alpha \beta }`$ $$g^{\alpha \beta }g_{\beta \gamma }=\delta _\gamma ^\alpha ,$$ $`\left(11\right)`$ with the help of which we are able to build a quantity $$C=X_\alpha X_\beta g^{\alpha \beta },$$ that, in consequence of relation (10), is for the bracket (1) a Casimir function which annihilates the bracket (1) and is an invariant of the Lie group with the structure constants $`c_{\alpha \beta }^{}{}_{}{}^{\gamma }`$ and the generators $`T_\alpha `$ $$\{X_\alpha ,C\}_0=c_{\alpha \beta }^{}{}_{}{}^{\gamma }X_\gamma _{X_\beta }C=T_\alpha C=0.$$ 3. Now let us replace in expression (1) the commuting variables $`X_\alpha `$ by Grassmann variables $`\mathrm{\Theta }_\alpha `$ $`(g(\mathrm{\Theta }_\alpha )=1)`$. Then we obtain a binary composition $$\{\mathrm{\Theta }_\alpha ,\mathrm{\Theta }_\beta \}_1=c_{\alpha \beta }^{}{}_{}{}^{\gamma }\mathrm{\Theta }_\gamma ,$$ $`\left(12\right)`$ which, due to relations (7) and (8), meets all the properties of the odd Poisson brackets: $$\{A,B+C\}_1=\{A,B\}_1+\{A,C\}_1,$$ $`\left(13\right)`$ $$g\left(\{A,B\}_1\right)=g\left(A\right)+g\left(B\right)+1\left(mod2\right),$$ $`\left(14\right)`$ $$\{A,B\}_1=\left(1\right)^{\left(g\left(A\right)+1\right)\left(g\left(B\right)+1\right)}\{B,A\}_1,$$ $`\left(15\right)`$ $$\underset{\left(ABC\right)}{}\left(1\right)^{\left(g\left(A\right)+1\right)\left(g\left(C\right)+1\right)}\{A,\{B,C\}_1\}_1=0,$$ $`\left(16\right)`$ $$\{A,BC\}_1=\{A,B\}_1C+\left(1\right)^{\left(g\left(A\right)+1\right)g\left(B\right)}B\{A,C\}_1.$$ $`\left(17\right)`$ It is surprising enough that the odd bracket can be defined solely in terms of the Grassmann variables as well as an even Martin bracket . On the following bilinear functions of canonical variables commuting $`q^\alpha `$ and Grassmann $`\theta _\alpha `$ $$\mathrm{\Theta }_\alpha =c_{\alpha \beta }^{}{}_{}{}^{\gamma }q^\beta \theta _\gamma $$ a canonical odd Poisson bracket $$\{A,B\}_1=A\underset{\alpha =1}{\overset{N}{}}(\underset{q^\alpha }{\overset{}{}}\underset{\theta _\alpha }{\overset{}{}}\underset{\theta _\alpha }{\overset{}{}}\underset{q^\alpha }{\overset{}{}})B$$ is reduced to the bracket (12) providing that $`c_{\alpha \beta }^{}{}_{}{}^{\gamma }`$ obey the conditions (7), (8). On functions $`A,B`$ of Grassmann variables $`\mathrm{\Theta }_\alpha `$ the bracket (12) has the form $$\{A,B\}_1=A\underset{\mathrm{\Theta }_\alpha }{\overset{}{}}c_{\alpha \beta }^{}{}_{}{}^{\gamma }\mathrm{\Theta }_\gamma \underset{\mathrm{\Theta }_\beta }{\overset{}{}}B,$$ The bracket (12) can be either degenerate or non-degenerate in the dependence on whether the matrix $`c_{\alpha \beta }^{}{}_{}{}^{\gamma }\mathrm{\Theta }_\gamma `$ in the indices $`\alpha ,\beta `$ is degenerate or not. Raising and lowering of the indices $`\alpha ,\beta `$, the non-degenerate metric tensors (9), (11) relate with each other the adjoint and co-adjoint representations which are equivalent for a semi-simple Lie group $$\mathrm{\Theta }^\alpha =g^{\alpha \beta }\mathrm{\Theta }_\beta ,_{\mathrm{\Theta }^\alpha }=g_{\alpha \beta }_{\mathrm{\Theta }_\beta }.$$ Hereafter only the non-degenerate metric tensors (11) will be considered. 4. By contracting the indices in a product of the Grassmann variables with the upper indices and of the successive Grassmann derivatives, respectively, with the lower indices in (8), we obtain the relations $$\mathrm{\Theta }^\alpha \mathrm{\Theta }^\beta \left(c_{\alpha \beta }^{}{}_{}{}^{\lambda }c_{\lambda \gamma }^{}{}_{}{}^{\delta }+2c_{\gamma \alpha }^{}{}_{}{}^{\lambda }c_{\lambda \beta }^{}{}_{}{}^{\delta }\right)=0,\mathrm{\Theta }^\alpha \mathrm{\Theta }^\beta \mathrm{\Theta }^\gamma c_{\alpha \beta }^{}{}_{}{}^{\lambda }c_{\lambda \gamma }^{}{}_{}{}^{\delta }=0,$$ $`(18a,b)`$ $$\left(c_{\alpha \beta }^{}{}_{}{}^{\lambda }c_{\lambda \gamma }^{}{}_{}{}^{\delta }+2c_{\gamma \alpha }^{}{}_{}{}^{\lambda }c_{\lambda \beta }^{}{}_{}{}^{\delta }\right)_{\mathrm{\Theta }_\alpha }_{\mathrm{\Theta }_\beta }=0,c_{\alpha \beta }^{}{}_{}{}^{\lambda }c_{\lambda \gamma }^{}{}_{}{}^{\delta }_{\mathrm{\Theta }_\alpha }_{\mathrm{\Theta }_\beta }_{\mathrm{\Theta }_\gamma }=0,$$ $`(19a,b)`$ which will be used later on many times. In particular, taking into account relation (18b), we can verify that the linear odd bracket (12) has the following Grassmann-odd nilpotent Casimir function $$\mathrm{\Delta }_{+3}=\frac{1}{\sqrt{3!}}\mathrm{\Theta }^\alpha \mathrm{\Theta }^\beta \mathrm{\Theta }^\gamma c_{\alpha \beta \gamma },\left(\mathrm{\Delta }_{+3}\right)^2=0,$$ $`\left(20\right)`$ which is an invariant of the Lie group $$\{\mathrm{\Theta }_\alpha ,\mathrm{\Delta }_{+3}\}_1=\mathrm{\Theta }_\gamma c_{\alpha \beta }^{}{}_{}{}^{\gamma }_{\mathrm{\Theta }_\beta }\mathrm{\Delta }_{+3}=S_\alpha \mathrm{\Delta }_{+3}=0$$ $`\left(21\right)`$ with the generators $`S_\alpha `$ obeying the Lie algebra permutation relations<sup>1</sup><sup>1</sup>1Note that below $`[A,B]=ABBA`$ and $`\{A,B\}=AB+BA`$. $$[S_\alpha ,S_\beta ]=c_{\alpha \beta }^{}{}_{}{}^{\gamma }S_\gamma .$$ $`\left(22\right)`$ It is a well-known fact that, in contrast with the even Poisson bracket, the non-degenerate odd Poisson bracket has one Grassmann-odd nilpotent differential $`\mathrm{\Delta }`$-operator of the second order, in terms of which the main equation has been formulated in the Batalin-Vilkovisky scheme for the quantization of gauge theories in the Lagrangian approach. In a formulation of Hamiltonian dynamics by means of the odd Poisson bracket with the help of a Grassmann-odd Hamiltonian $`\overline{H}`$ $`(g(\overline{H})=1)`$ this $`\mathrm{\Delta }`$-operator plays also a very important role being used to distinguish the Hamiltonian dynamical systems, for which the Liouville theorem is valid $`\mathrm{\Delta }\overline{H}=0`$, from those ones, for which this theorem takes no place $`\mathrm{\Delta }\overline{H}0`$ <sup>2</sup><sup>2</sup>2Note also applications of the odd bracket to the integrability problem . Now let us try to build the $`\mathrm{\Delta }`$-operator for the linear odd bracket (12). It is remarkable that, in contrast with the canonical odd Poisson bracket having the only $`\mathrm{\Delta }`$-operator of the second order, we are able to construct at once three $`\mathrm{\Delta }`$-like Grassmann-odd nilpotent operators which are differential operators of the first, the second and the third orders respectively $$\mathrm{\Delta }_{+1}=\frac{1}{\sqrt{2}}\mathrm{\Theta }^\alpha \mathrm{\Theta }^\beta c_{\alpha \beta }^{}{}_{}{}^{\gamma }_{\mathrm{\Theta }^\gamma },\left(\mathrm{\Delta }_{+1}\right)^2=0;$$ $`\left(23\right)`$ $$\mathrm{\Delta }_1=\frac{1}{\sqrt{2}}\mathrm{\Theta }_\gamma c_{\alpha \beta }^{}{}_{}{}^{\gamma }_{\mathrm{\Theta }_\alpha }_{\mathrm{\Theta }_\beta },\left(\mathrm{\Delta }_1\right)^2=0;$$ $`\left(24\right)`$ $$\mathrm{\Delta }_3=\frac{1}{\sqrt{3!}}c_{\alpha \beta \gamma }_{\mathrm{\Theta }_\alpha }_{\mathrm{\Theta }_\beta }_{\mathrm{\Theta }_\gamma },\left(\mathrm{\Delta }_3\right)^2=0.$$ $`\left(25\right)`$ The nilpotency of the operators $`\mathrm{\Delta }_{+1}`$ and $`\mathrm{\Delta }_1`$ is a consequence of relations (18b) and (19b). The operator $`\mathrm{\Delta }_{+1}`$ is proportional to the second term in a BRST charge $$Q=\mathrm{\Theta }^\alpha G_\alpha \frac{1}{2}\mathrm{\Theta }^\alpha \mathrm{\Theta }^\beta c_{\alpha \beta }^{}{}_{}{}^{\gamma }_{\mathrm{\Theta }^\gamma },$$ where $`\mathrm{\Theta }^\alpha `$ and $`_{\mathrm{\Theta }^\alpha }`$ represent the operators for ghosts and antighosts respectively. $`Q`$ itself will be proportional to the operator $`\mathrm{\Delta }_{+1}`$ if we take the representation $`S_\alpha `$ (21) for group generators $`G_\alpha `$. The operator $`\mathrm{\Delta }_1`$, related with the divergence of a vector field $`\{\mathrm{\Theta }_\alpha ,A\}_1`$ $$_{\mathrm{\Theta }_\alpha }\{\mathrm{\Theta }_\alpha ,A\}_1=_{\mathrm{\Theta }_\alpha }S_\alpha A=\sqrt{2}\mathrm{\Delta }_1A,$$ is proportional to the true $`\mathrm{\Delta }`$-operator for the bracket (12). It is also interesting to reveal that these $`\mathrm{\Delta }`$-like operators together with the Casimir function $`\mathrm{\Delta }_{+3}`$ (20) are closed into the finite-dimensional Lie superalgebra, in which the anticommuting relations between the quantities $`\mathrm{\Delta }_\lambda `$ $`(\lambda =3,1,+1,+3)`$ (20), (23)-(25) with the nonzero right-hand side are $$\{\mathrm{\Delta }_1,\mathrm{\Delta }_{+1}\}=Z,$$ $`\left(26\right)`$ $$\{\mathrm{\Delta }_3,\mathrm{\Delta }_{+3}\}=N3Z,$$ $`\left(27\right)`$ where $$N=c^{\alpha \beta \gamma }c_{\alpha \beta \gamma }$$ is a number of values for the indices $`\alpha ,\beta ,\gamma `$ $`(\alpha ,\beta ,\gamma =1,\mathrm{},N)`$ and $$Z=DK$$ $`\left(28\right)`$ is a central element of this superalgebra $$[Z,\mathrm{\Delta }_\lambda ]=0,\left(\lambda =3,1,+1,+3\right).$$ $`\left(29\right)`$ In (28) $$D=\mathrm{\Theta }^\alpha _{\mathrm{\Theta }^\alpha }$$ $`\left(30\right)`$ is a ”dilatation” operator for the Grassmann variables $`\mathrm{\Theta }_\alpha `$, which distinguishes the $`\mathrm{\Delta }_\lambda `$-operators with respect to their uniformity degrees in $`\mathrm{\Theta }`$ $$[D,\mathrm{\Delta }_\lambda ]=\lambda \mathrm{\Delta }_\lambda ,\left(\lambda =3,1,+1,+3\right)$$ $`\left(31\right)`$ and is in fact a representation for a ghost number operator, and the quantity $`K`$ has the form $$K=\frac{1}{2}\mathrm{\Theta }^\alpha \mathrm{\Theta }^\beta c_{\alpha \beta }^{}{}_{}{}^{\lambda }c_{\lambda \gamma \delta }_{\mathrm{\Theta }_\gamma }_{\mathrm{\Theta }_\delta }.$$ $`\left(32\right)`$ The operator $`Z`$ is also a central element of the Lie superalgebra which contains both the operators $`\mathrm{\Delta }_\lambda `$ (20), (23)-(25), $`Z`$ (28) and the operator $`D`$ (30) $$[Z,D]=0.$$ $`\left(33\right)`$ We can add to this superalgebra the generators $`S_\alpha `$ (21) with the following commutation relations: $$[S_\alpha ,\mathrm{\Delta }_\lambda ]=0,[S_\alpha ,Z]=0,[S_\alpha ,D]=0,$$ $`\left(34\right)`$ which indicate that both the Casimir function $`\mathrm{\Delta }_{+3}`$ and the operators $`\mathrm{\Delta }_\lambda `$ $`(\lambda =3,1,+1)`$, $`Z`$ and $`D`$ are invariants of the Lie group with the generators $`S_\alpha `$. In order to prove the permutation relations for the Lie superalgebra (20)-(34), we have to use relations (18) and (19). Note that the central element $`Z`$ (28) coincides with the expression for a quadratic Casimir operator of the Lie algebra (22) for the generators $`S_\alpha `$ given in the representation (21) $$S_\alpha S_\beta g^{\alpha \beta }=Z.$$ $`\left(35\right)`$ 5. Thus, we see that both the even and odd linear Poisson brackets are internally inherent in the Lie group with the structure constants subjected to conditions (7) and (8). However, only for the linear odd Poisson bracket realized in terms of the Grassmann variables and only in the case when this bracket corresponds to the semi-simple Lie group, there exists the Lie superalgebra (20)-(34) for the $`\mathrm{\Delta }`$-like operators of this bracket. Note that in the case of the degenerate Cartan-Killing metric tensor (9), relation (10) remains valid and we can construct only two $`\mathrm{\Delta }`$-like Grassmann-odd nilpotent operators: $`\mathrm{\Delta }_1`$ (24) and $`\mathrm{\Delta }_3`$ (25), which satisfy the trivial anticommuting relation $$\{\mathrm{\Delta }_1,\mathrm{\Delta }_3\}=0.$$ Note also that anticommuting relations for the operators $$\underset{1}{\overset{i}{\mathrm{\Delta }}}=\frac{1}{\sqrt{2}}\mathrm{\Theta }_\gamma \stackrel{i}{c_{\alpha \beta }^{}{}_{}{}^{\gamma }}_{\mathrm{\Theta }_\alpha }_{\mathrm{\Theta }_\beta },\left(\underset{1}{\overset{i}{\mathrm{\Delta }}}\right)^2=0,$$ corresponding to the Lie algebras with structure constants $`\stackrel{i}{c_{\alpha \beta }^{}{}_{}{}^{\gamma }}`$ $`(i=1,2)`$, vanish provided that $`\stackrel{i}{c_{\alpha \beta }^{}{}_{}{}^{\gamma }}`$ satisfy compatibility conditions $$\underset{\left(\alpha \beta \gamma \right)}{}\stackrel{\{i}{c_{\alpha \beta }^{}{}_{}{}^{\lambda }}\stackrel{k\}}{c_{\lambda \gamma }^{}{}_{}{}^{\delta }}=0,$$ where $`\{ik\}`$ denotes the symmetrization of the indices $`i`$ and $`k`$. The Lie superalgebra (20)-(34), naturally connected with the linear odd Poisson bracket (12), may be useful for the subsequent development of the Batalin-Vilkovisky formalism for the quantization of gauge theories. Indeed, very similar to (12) odd Poisson brackets on the Grassmann algebra are used in a generalization of the triplectic formalism which is a covariant version of the $`Sp(2)`$-symmetric quantization of general gauge theories. We should therefore expect that the Lie superalgebra (20)-(34), closely related with the linear odd bracket (12), will also find the application for the further development of the above-mentioned generalization of the triplectic formalism. Let us note that the superalgebra (20)-(34) can also be used in the theory of representations of the semi-simple Lie groups. The author is sincerely thankful to V.D. Gershun, D.A. Leites and S.L. Lyakhovich for useful discussions and is indebted to J.D. Stasheff for illuminating remarks. The author wishes to thank J. Wess for kind hospitality at the University of Munich where this work was completed. This work was supported in part by the Ukrainian State Foundation of Fundamental Researches, Grant No 2.5.1/54 and by Grant INTAS No 93-127 (Extension).
warning/0002/astro-ph0002164.html
ar5iv
text
# Is There a Detectable Vishniac Effect? ## 1 Introduction While recombination at $`z1100`$ marked the end of ionized hydrogen from the viewpoint of a linearly evolving universe, the nonlinear evolution of small-scale perturbations resulted in the reionization of the intergalactic medium at much lower redshifts. The fact that quasar spectra show an absence of an absorption trough from Ly$`\alpha `$ resonant scattering by neutral H atoms distributed diffusely along the line of sight, the Gunn-Peterson effect (Gunn & Peterson 1965), means that this reionization must have occurred with a high degree of efficiency before a redshift of 5. One of the necessary consequences of this reionization is the presence of secondary anisotropies in the cosmic microwave background (CMB) due to the scattering of photons off ionized electrons. These secondary fluctuations can be divided into two classes: anisotropies due to nonlinear structures and linear anisotropies. Nonlinear secondary anisotropies are of several types. Some of the more studied of these include the scattering of photons off the hot intracluster medium of galaxy clusters (Sunyaev & Zel’dovich 1970, 1972; or for more recent treatments see, e.g., Evrard & Henry 1991; Colfrancesco et al. 1994; Aghanim et al. 1997), gravitational lensing (see, e.g., Linder 1997; Metcalf & Silk 1997), the impact of inhomogeneous reionization (Aghanim et al. 1995; Peebles & Juszkiewicz 1998; Knox, Scoccimarro, & Dodelson 1998), and the Rees-Sciama effect due to the bulk motions of collapsing nonlinear structures (see, e.g., Rees & Sciama 1968; Kaiser 1982; Seljak 1996). Small-scale linear anisotropies come in fewer flavors. Detailed analyses of linear perturbations have uncovered a single dominant effect known as the Vishniac or Ostriker-Vishniac effect (Hu, Scott, & Silk 1994; Dodelson & Jubas 1995; Hu & White 1995; Hu & Sugiyama 1996). The level of these perturbations has been calculated by several authors (Ostriker & Vishniac 1985; Vishniac 1987; Jaffe & Kamionkowski 1998, hereafter JK). These investigations raise the question of whether a detectable Vishniac effect even exists since nonlinear structures must exist on some length scale at the time of secondary scattering of CMB photons, as it is only by the formation of nonlinear objects that the universe is able to reionize itself. If these scales are comparable to those making the dominant contribution to the Vishniac effect, then a linear analysis is inappropriate and a calculation of secondary anisotropies must incorporate nonlinear effects. In this work we determine the minimum length scale, $`R_V`$, which must remain linear in order for a linear approach to scattering by ionized regions with varying bulk motions to be accurate for the range of angular scales over which one can hope to measure secondary fluctuations. In hierarchical scenarios of structure formation, such as the Cold Dark Matter (CDM) model, smaller structures assemble at early times, later merging to form larger objects. This allows us to place limits on the time between the formation of structures large enough to reionize the universe and the time at which $`R_V`$ becomes nonlinear. At that point, while peculiar velocities of ionized gas continue to be imprinted on the microwave background, the nature of this signature is qualitatively different and is best interpreted from another perspective. The structure of this work is as follows. In Sec. 2 we describe the Vishniac effect and determine the physical length scale on which it depends. In Sec. 3 we compare this to the scale of reionizing objects in different reionization scenarios and discuss the applicability of linear theory. In Sec. 4 we examine how the Vishniac effect is distinguished from other effects. Conclusions are given in Sec. 5, and the various cosmological expressions used throughout are summarized in the appendix. ## 2 Analysis ### 2.1 Approximations The Vishniac effect is caused by the scattering of CMB photons by ionized regions with varying bulk motions. The temperature fluctuations induced along a line of sight are given by $$\frac{\mathrm{\Delta }T}{T}(\stackrel{}{\theta })=_0^{t_0}𝑑t\sigma _Te^{\tau (\stackrel{}{\theta },t)}n_e(\stackrel{}{\theta },t)\widehat{\theta }𝐯(\stackrel{}{\theta },t),$$ (1) where $`\tau (\stackrel{}{\theta },t)`$, $`n_e(\stackrel{}{\theta },t)`$, and $`𝐯(\stackrel{}{\theta },t)`$ are the optical depth along the line of sight, electron density, and bulk velocity, $`\sigma _T`$ is the cross section for Thomson scattering, $`t`$ is the age of the universe, and $`t_0`$ is the present age. Following JK, we choose a coordinate system in which $`\widehat{\theta }`$ represents a three-dimensional unit vector along the line of sight, $`\stackrel{}{\theta }`$ represents a two-dimensional unit vector in the plane perpendicular to it, and bold letters represent fully three-dimensional vectors. Thus $`𝐯=(v_x,v_y,v_z)`$, $`\stackrel{}{\theta }=(\theta _1,\theta _2,0)`$, and $`\widehat{\theta }=(\theta _1,\theta _2,\sqrt{1\theta _1^2\theta _2^2})(\theta _1,\theta _2,1)`$, the validly of the approximation deriving from the small-scale nature of the effect. Note that $`n_e`$, $`𝐯`$, and $`\tau `$ are all functions of position, the optical depth being given by $`\tau (\stackrel{}{\theta },t)=_t^{t_0}\sigma _Tn_e(\stackrel{}{\theta },t)c𝑑t^{}`$, where $`c`$ is the speed of light. If we decompose the density field into average and fluctuating components, we obtain, to leading order, $$\frac{\mathrm{\Delta }T}{T}(\stackrel{}{\theta })=\frac{\sigma _Tn_0}{c}_0^1𝑑wa_0^3\frac{x_e(\widehat{\theta }w_{\mathrm{ang}},w)}{a(w)^2}(1+\delta (\widehat{\theta }w_{\mathrm{ang}},w)\mathrm{\Delta }\tau (\widehat{\theta }w_{\mathrm{ang}},w))\widehat{\theta }𝐯(\widehat{\theta }w_{\mathrm{ang}},w)e^{\tau _0(w)},$$ (2) where $`x_e(𝐱,w)`$ is the ionization fraction, $`\tau _0`$ and $`\mathrm{\Delta }\tau (𝐱,w)`$ are the optical depths due to the average and fluctuating density components respectively, $`a(w)`$ is the scale factor with $`a_0a(0)`$, $`\delta (𝐱,w)`$ is the overdensity field defined such that $`\delta (𝐱,w)\rho (𝐱,w)/\overline{\rho }(w)1`$ where $`\rho (𝐱,w)`$ is the density field and $`\overline{\rho }(w)`$ the average density as a function of comoving distance, $`n_0`$ is the present average electron density, and $`w_{\mathrm{ang}}`$ is the comoving angular distance, given by Eq. (27). Note that we have replaced time by $`w`$, the comoving distance defined by $`dwcdt/a(t)`$, and rewritten $`𝐯`$, and $`\tau `$ in comoving coordinates, $`𝐱`$. Taking the mass fraction of He to be $`25\%`$, and approximating helium reionization as simultaneous to that of hydrogen, $`n_0=\mathrm{\Omega }_b\rho _c/m_p\times 7/8=9.9\times 10^6\mathrm{\Omega }_bh^2\mathrm{cm}^3`$, where $`\rho _c`$ is the critical density, $`m_p`$ is the mass of the proton, $`h`$ is Hubble’s constant normalized to $`100\mathrm{km}\mathrm{s}^1\mathrm{Mpc}^1`$, and $`\mathrm{\Omega }_b`$ is the present baryonic matter density in units of the critical density. The Vishniac effect arises by considering the homogeneously-reionizing, low optical depth case. In this case $`x_e(𝐱,w)`$ is independent of position and $`\mathrm{\Delta }\tau `$ can be ignored. In practice, realistic reionization scenarios lead to low values of optical depth and hence dropping $`\mathrm{\Delta }\tau `$ is a safe assumption (Hu, Scott, & Silk 1994). Indeed, the measurement of CMB fluctuations at large scales precludes the high degree of damping that would be caused by a high optical depth (Scott & White 1994; Hancock et al. 1998). The homogeneity of reionization, however, is a question of scales and thus represents a basic assumption on which the Vishniac analysis is based. In this limit Eq. (2) becomes $$\frac{\mathrm{\Delta }T}{T}(\stackrel{}{\theta })=_0^1𝑑wg(w)\widehat{\theta }\left(𝐯(\widehat{\theta }w_{\mathrm{ang}},w)+𝐪(\widehat{\theta }w_{\mathrm{ang}},w)\right),$$ (3) where $`𝐪(𝐱,t)𝐯(𝐱,w)\delta (𝐱,w)`$ and $`g(w)`$ is the visibility function $$g(w)\frac{a_0^3n_0x_e(w)}{ca(w)^2}\sigma _Te^\tau =\frac{.121\mathrm{\Omega }_bh}{c}(1+z(w))^2x_e(w)e^\tau ,$$ (4) with our conventions for the scale factor as in Appendix A. This gives the probability of scattering off reionized electrons and is a slowly-varying function of $`w`$. Finally, we must approximate both $`𝐯`$ and $`𝐪`$ using linear theory. In this case, the density contrast at a comoving coordinate $`𝐱`$ and comoving distance $`w`$ from the observer is a random field with Fourier transform given by $`\stackrel{~}{\delta }(𝐤,w)d^3𝐱\mathrm{exp}(i𝐤𝐱)\delta (𝐱,w).`$ The spatial and time dependence of $`\stackrel{~}{\delta }(𝐤,w)`$ can be factorized, so $`\stackrel{~}{\delta }(𝐤,w)=\stackrel{~}{\delta }_0(𝐤)D(w)/D_0`$ where $`\stackrel{~}{\delta }_0(𝐤)\stackrel{~}{\delta }(𝐤,0)`$, $`D(w)`$ is the linear growth factor, given by Eq. (28), and $`D_0D(0)`$. The power spectrum is then defined by the relation $$\stackrel{~}{\delta _0}(𝐤)\stackrel{~}{\delta _0}(𝐤^{})=\stackrel{~}{\delta _0}(𝐤)\stackrel{~}{\delta _0}^{}(𝐤^{})=(2\pi )^3\delta ^3(𝐤+𝐤^{})P(k),$$ (5) where $`\delta ^3(𝐤+𝐤^{})`$ denotes the three-dimensional Dirac delta function. This completely specifies the probability density functional from which $`\delta (𝐤)`$ is drawn in gaussian theories. In the CDM cosmogony, $`P(k)`$ is given by Eq. (31) and Eq. (33), and is dependent on the ‘shape parameter’ $`\mathrm{\Gamma }`$, which is given as a function of cosmological parameters by Eq. (32) and constrained by observations of the galaxy correlation function to be $`0.23_{0.034}^{+0.042}`$ (Viana and Liddle 1996). The overall normalization of $`P(k)`$ can be fixed by the amplitude of mass fluctuations on the 8 $`h^1`$ Mpc scale as defined in Eq. (34). The linear velocity field is simply related to the density field by the continuity equation $`𝐯(𝐱)=a(w)\dot{\delta }(𝐱,w)`$ which in Fourier space gives $$\stackrel{~}{𝐯}(𝐤,t)=\frac{ia(w)}{k^2}\frac{\dot{D}(w)}{D_0}𝐤\stackrel{~}{\delta }_0(𝐤),$$ (6) where $`\dot{D}(w)`$ denotes the derivative of $`D`$ with respect to time rather than $`w`$, and is given by Eq. (29). As the velocity always points along the direction of $`𝐤`$, only $`𝐤`$ modes with large values along the line of sight can have large peculiar velocities in the $`\widehat{\theta }`$ direction. But these modes are varying with wavelengths much smaller than variations in the window function and cancel when projected along the line of sight. It is only the second order ($`𝐪`$) term then, that contributes to the integral in Eq. (3). A simple real-space argument gives a different way of understanding this cancellation. As gravitational perturbations to a pressureless fluid can be written in terms of the gradient of a scalar potential field ($`𝐯\varphi `$ where $`\varphi `$ is the gravitational potential) the curl of the velocity field is zero to all orders. Thus a line integral of $`𝐯`$ along a line of sight is approximately the integral of a gradient and is zero except for a small contribution at the end points. From all the terms in Eq. 2, only the $`\delta \times \widehat{\theta }𝐯`$ survives to contribute to $`\frac{\mathrm{\Delta }T}{T}`$. ### 2.2 Power Spectrum Dropping the velocity term from Eq. 3, we can analytically construct the power spectrum of the angular fluctuations in the linear limit. Let us define $`\stackrel{~}{\frac{\mathrm{\Delta }T}{T}}(\stackrel{}{\kappa })`$ as the Fourier transform of the temperature fluctuations such that $`\stackrel{~}{\frac{\mathrm{\Delta }T}{T}}(\stackrel{}{\kappa })=d^2\stackrel{}{\theta }\mathrm{exp}(i\stackrel{}{\kappa }\stackrel{}{\theta })\frac{\mathrm{\Delta }T}{T}(\stackrel{}{\theta }),`$ with the angular power spectrum defined as $`P_{\mathrm{ang}}(\kappa _1)(2\pi )^2\delta ^2(\stackrel{}{\kappa }_1+\stackrel{}{\kappa }_2)\left(\stackrel{~}{\frac{\mathrm{\Delta }T}{T}}(\stackrel{}{\kappa }_1)\stackrel{~}{\frac{\mathrm{\Delta }T}{T}}(\stackrel{}{\kappa }_2)\right)`$. At the small angular scales appropriate to the Vishniac effect, $`P_{\mathrm{ang}}(\kappa )`$ is simply related to the usual $`C_{\mathrm{}}`$s used to express CMB fluctuations by $`C_{\mathrm{}}=P_{\mathrm{ang}}(\kappa =\mathrm{})`$ (JK). Several authors (Vishniac 1987; Kaiser 1992; JK) have derived expressions for the Vishniac $`C_{\mathrm{}}`$s. Here we provide a new approach that, unlike other techniques, is easily extended to calculate higher-order moments, as is shown in §4. Our method is a simple extension of the usual formalism used to calculate single-point moments of the $`\frac{\mathrm{\Delta }T}{T}(\stackrel{}{\theta })`$ distribution. The simplest quantity of this sort is the second moment $`\left[\frac{\mathrm{\Delta }T_B}{T}(0)\right]^2`$, where we use $`B`$ to denote convolution with a beam profile. Given such a profile in Fourier space $`B(\stackrel{}{\kappa })`$, the second moment can be calculated as $$\left[\frac{\mathrm{\Delta }T_B}{T}(0)\right]^2=\frac{d^2\stackrel{}{\kappa }}{(2\pi )^2}B(\stackrel{}{\kappa })^2P_{\mathrm{ang}}(\kappa ).$$ (7) Suppose, however, that instead of asking about the second moment calculated from a single map observed with a beam $`B(\stackrel{}{\kappa })`$, we instead compute the single-point function as calculated from the convolution of two maps, observed by beams $`B_1(\stackrel{}{\kappa })`$ and $`B_2(\stackrel{}{\kappa })`$. As these profiles are arbitrary, we are free to take $$B_1(\stackrel{}{\kappa })(2\pi )^2\delta ^2(\stackrel{}{\kappa }\stackrel{}{\kappa }_1)B_2(\stackrel{}{\kappa })(2\pi )^2\delta ^2(\stackrel{}{\kappa }\stackrel{}{\kappa }_2),$$ (8) where $`\delta ^2(\stackrel{}{\kappa })`$ is the two-dimensional delta function. In this case we find $$\frac{\mathrm{\Delta }T_{B_1}}{T}(0)\frac{\mathrm{\Delta }T_{B_2}}{T}(0)=(2\pi )^2P_{\mathrm{ang}}(\kappa _1)\delta ^2(\stackrel{}{\kappa }_1+\stackrel{}{\kappa }_2),$$ (9) recovering the angular power spectrum. Thus the power spectrum of the Vishniac effect can be computed from the single-point correlation if we are careful to express the beam profiles in sufficient generality. Let us consider then $`{\displaystyle \frac{\mathrm{\Delta }T_{B_1}}{T}}(0){\displaystyle \frac{\mathrm{\Delta }T_{B_2}}{T}}(0)=`$ $`{\displaystyle _0^1}𝑑w_1g(w_1){\displaystyle _0^1}𝑑w_2g(w_2){\displaystyle \frac{d^3𝐤_𝐚}{(2\pi )^3}\frac{d^3𝐤_𝐛}{(2\pi )^3}}`$ (10) $`B_1(\stackrel{}{k}_aw_{1,\mathrm{ang}})B_2(\stackrel{}{k}_bw_{2,\mathrm{ang}})e^{ik_{a,z}w_1+ik_{b,z}w_2}\stackrel{~}{q}_z(𝐤_𝐚,w_1)\stackrel{~}{q}_z(𝐤_𝐛,w_2),`$ where $`\stackrel{~}{𝐪}(𝐤,w)`$ is the Fourier transform of $`𝐪(𝐱,w)`$. Substituting in the expression for $`\stackrel{~}{𝐪}`$ in terms of $`\stackrel{~}{\delta }`$: $$\stackrel{~}{𝐪}(𝐤,w)=\frac{ia(w)\dot{D}(w)D(w)}{D_0^2}\frac{d^3𝐤^{}}{(2\pi )^3}\stackrel{~}{\delta }_0(𝐤^{})\stackrel{~}{\delta }_0(𝐤𝐤^{})\frac{𝐤^{}}{k^2},$$ (11) Eq. (10) becomes $`{\displaystyle \frac{\mathrm{\Delta }T_{B_1}}{T}}(0){\displaystyle \frac{\mathrm{\Delta }T_{B_2}}{T}}(0)=`$ $``$ $`{\displaystyle _0^1}𝑑w_1G(w_1){\displaystyle _0^1}𝑑w_2G(w_2){\displaystyle \underset{j=1}{\overset{4}{}}}\left[{\displaystyle \frac{d^3𝐤_𝐣}{(2\pi )^3}}\right]B_1((\stackrel{}{k}_1+\stackrel{}{k}_2)w_{1,\mathrm{ang}})`$ (12) $`B_2((\stackrel{}{k}_3+\stackrel{}{k}_4)w_{2,\mathrm{ang}})e^{i(k_1+k_2)w_1+i(k_3+k_4)w_2}{\displaystyle \frac{k_{1,z}}{k_1^2}}{\displaystyle \frac{k_{3,z}}{k_3^2}}{\displaystyle \underset{l=1}{\overset{4}{}}}\stackrel{~}{\delta }_0(𝐤_𝐥),`$ where $$G(w)\frac{g(w)a(w)D(w)\dot{D}(w)}{D_0^2}.$$ (13) As $`G(w)`$ is slowly varying, we can follow Kaiser (1992) in dividing the integrals over comoving distance into $`N`$ statistically independent intervals of width $`\mathrm{\Delta }w`$, over each of which $`G(w)`$ is well approximated by a constant. In this case $`{\displaystyle \frac{\mathrm{\Delta }T_{B_1}}{T}}(0){\displaystyle \frac{\mathrm{\Delta }T_{B_2}}{T}}(0)={\displaystyle \underset{n=1}{\overset{N}{}}}G(w_n)^2\mathrm{\Delta }w^2{\displaystyle \underset{j=1}{\overset{4}{}}}\left[{\displaystyle \frac{d^3𝐤_𝐣}{(2\pi )^3}}\right]`$ $`B_1((\stackrel{}{k}_1+\stackrel{}{k}_2)w_{n,\mathrm{ang}})B_2((\stackrel{}{k}_3+\stackrel{}{k}_4)w_{n,\mathrm{ang}})`$ $`j_0\left({\displaystyle \frac{(k_{1,z}+k_{2,z})\mathrm{\Delta }w}{2}}\right)j_0\left({\displaystyle \frac{(k_{3,z}+k_{4,z})\mathrm{\Delta }w}{2}}\right){\displaystyle \frac{k_{1,z}}{k_1^2}}{\displaystyle \frac{k_{3,z}}{k_3^2}}{\displaystyle \underset{l=1}{\overset{4}{}}}\stackrel{~}{\delta }_0(𝐤_𝐥).`$ (14) As the density fluctuations are taken to be gaussian, we can expand the expectation value of the product of overdensities by Wick’s theorem, keeping only the terms in which $`k_1`$ is paired with $`k_3`$ or $`k_4`$. If we then define $`k_2^{}k_1+k_2`$, we find $`{\displaystyle \frac{\mathrm{\Delta }T_{B_1}}{T}}(0){\displaystyle \frac{\mathrm{\Delta }T_{B_2}}{T}}(0)={\displaystyle \underset{n=1}{\overset{N}{}}}G(w_n)^2\mathrm{\Delta }w^2{\displaystyle \frac{d^3𝐤_\mathrm{𝟏}}{(2\pi )^3}\frac{d^3𝐤_\mathrm{𝟐}^{}}{(2\pi )^3}B_1(\stackrel{}{k}_2^{}w_{n,\mathrm{ang}})B_2(\stackrel{}{k}_2^{}w_{n,\mathrm{ang}})}`$ $`\left[{\displaystyle \frac{k_{1,z}^2}{k_1^4}}+{\displaystyle \frac{k_{1,z}(k_{2,z}^{}k_{1,z})}{k_1^2𝐤_\mathrm{𝟐}^{}𝐤_\mathrm{𝟏}^2}}\right]P(k_1)P(𝐤_\mathrm{𝟐}^{}𝐤_\mathrm{𝟏})j_0^2\left({\displaystyle \frac{k_{2,z}^{}\mathrm{\Delta }w}{2}}\right).`$ (15) The Bessel function has a width $`\delta k_z1/\mathrm{\Delta }w`$ so $`k_{2,z}<<\kappa /w_{\mathrm{ang}}`$ wherever $`j_0^2`$ is appreciable. We can neglect terms that are smaller by a factor of $`w_{\mathrm{ang}}^2\kappa ^2/\mathrm{\Delta }w^2`$ to obtain $`{\displaystyle \frac{\mathrm{\Delta }T_{B_1}}{T}}(0){\displaystyle \frac{\mathrm{\Delta }T_{B_2}}{T}}(0)`$ $`=`$ $`{\displaystyle \underset{n=1}{\overset{N}{}}}G(w_n)^2\mathrm{\Delta }w{\displaystyle \frac{d^3𝐤_\mathrm{𝟏}}{(2\pi )^3}\frac{d^2\stackrel{}{k}_2^{}}{(2\pi )^2}B_1(\stackrel{}{k}_2^{}w_{\mathrm{ang}})B_2(\stackrel{}{k}_2^{}w_{\mathrm{ang}})}`$ (16) $`{\displaystyle \frac{k_{1,z}^2}{k_1^4}}{\displaystyle \frac{k_{1,z}^2}{k_1^2(\stackrel{}{k}_2^{},0)𝐤_\mathrm{𝟏}^2}}P(k_1)P((\stackrel{}{k}_2^{},0)𝐤_\mathrm{𝟏}).`$ Taking $$B_1(\stackrel{}{\kappa })=B_2(\stackrel{}{\kappa })=2\pi \sigma ^2e^{\frac{\sigma ^2\kappa ^2}{2}}$$ (17) gives the second moment as observed by a beam of gaussian width $`\sigma `$, while choosing beam profiles as given in Eq. (8) yields the angular power spectrum. In this case $$C_{\mathrm{}}=P_{\mathrm{ang}}(\kappa =\mathrm{})=_0^1𝑑w\frac{G(w)^2}{w_{\mathrm{ang}}^2}P_\mathrm{V}(\mathrm{}/w_{\mathrm{ang}}),$$ (18) where $$P_\mathrm{V}(k)=\frac{d^3𝐤_\mathrm{𝟏}}{(2\pi )^3}P(k^{})P(||((k,0,0)𝐤_\mathrm{𝟏}||)[\frac{k_{1,z}^{}_{}{}^{}2}{k_1^{}_{}{}^{}4}\frac{k_{1,z}^{}_{}{}^{}2}{k_1^{}_{}{}^{}2(k,0,0)𝐤_\mathrm{𝟏}^2}],$$ (19) which, choosing a coordinate system in which the $`z^{}`$ axis points along the direction of $`(k,0,0)`$ becomes $$P_\mathrm{V}(k)=\frac{k}{8\pi ^2}_0^{\mathrm{}}𝑑x_1^1𝑑\mu P(kx)P(k\sqrt{12x\mu +x^2})(1\mu ^2)\left[1\frac{x^2}{12x\mu +x^2}\right].$$ (20) This is equivalent to the usual expression for the Vishniac power spectrum $$P_\mathrm{V}(k)=\frac{k}{8\pi ^2}_0^{\mathrm{}}𝑑x_1^1𝑑\mu P(kx)P(k\sqrt{12x\mu +x^2})\frac{(1\mu ^2)(12x\mu )^2}{(12x\mu +x^2)^2},$$ (21) as can be seen by rewriting both integrals in rectangular coordinates and applying an origin shift. Thus the $`C_{\mathrm{}}`$s are dependent on an integral along the line of sight of a term, $`P_\mathrm{V}(k)`$, that is independent of redshift and arises from the convolution of the $`𝐪`$ fields. We note in passing that our results are in agreement with Dodelson and Jubas (1995) and are twice the values found in JK. ### 2.3 Physical Scales Having outlined the approximations which are used to calculate this effect and constructed the resulting power spectrum of fluctuations, we now examine which physical scales contribute most to $`P_\mathrm{V}(k)`$. Typically, Eq. (18) is used to calculate the Vishniac effect by integrating over a particular matter power spectrum. To study the dependence of the effect on physical scale, we replace the power spectrum that appears in Eq. (21) with $`P(k)W^2(kR)`$ where $`P(k)`$ is the CDM power spectrum and $`W(x)`$ is the spherical top-hat window function, given by Eq. (35). Here we consider a $`\mathrm{\Lambda }`$CDM model in which the current nonrelativistic matter, vacuum, and baryonic densities in units of the critical density are $`\mathrm{\Omega }_0`$ = 0.35, $`\mathrm{\Omega }_\mathrm{\Lambda }`$ = 0.65, and $`\mathrm{\Omega }_b`$ = 0.06 respectively, and the “tilt” in the power spectrum as parameterized in Eq. (31) is taken to be flat, $`n=1.0`$. If the other parameters are taken to be $`h`$ =0.65, $`\sigma _8`$ = 1.05, $`\mathrm{\Gamma }=0.2`$ and reionization occurs instantaneously and completely at $`z_{\mathrm{re}}=18`$, this results in $`P_\mathrm{V}(k)`$ and $`C_{\mathrm{}}`$ as plotted in Fig. 1. Note that these values correspond to $`\frac{\mathrm{\Delta }T}{T}`$s a full order of magnitude smaller than large-scale primary anisotropies, pointing out the experimental challenges that must be overcome before secondary anisotropies can be measured (see, e.g., Subrahmanyan et al. 1993; Church et al. 1997). For reference we also plot the COBE-normalized primary fluctuations as computed by the CMBFAST code V2.4.1 (Seljak & Zaldarriaga 1996; Hu et al. 1998; Zaldarriaga & Seljak 1998), for the same cosmological model. Note that the position of this line is sensitive only to the Silk damping scale and thus is largely independent of the shape of the primordial power spectrum. In this figure we see that even when filtered at the $`1`$ Mpc/$`h`$ scale, the power spectrum is appreciably changed, especially at higher wavenumbers. While low $`\mathrm{}`$ fluctuations are not affected, a comparison with the solid line shows that these fluctuations are lost in the CMB primary signal and are very difficult to measure. Note that the damping shown in this graph represents not only a loss of power in linear theory, but also an increase of nonlinear power. Thus, at the point that the $`1`$ Mpc/$`h`$ scale has become nonlinear, the Vishniac effect is competing with nonlinear effects over the range of angular scales in which it is able to be detected. By the time the $`2`$ Mpc/$`h`$ scale becomes nonlinear, however, the peak wavelength of the Vishniac effect has been shifted by a factor of $`1/2`$ and $`50\%`$ of the power of $`P_\mathrm{V}(k)`$ has been lost. At this point, linear calculations are unlikely to be reliable at measurable $`\mathrm{}`$ values $`4000`$, and a more careful theoretical approach becomes necessary. These length scales are inversely proportional to the ‘shape parameter’ $`\mathrm{\Gamma }`$ which is $`0.2`$ for this model. Thus we can conservatively fix $`R_V=`$ 2 Mpc/$`h\mathrm{\Gamma }_{0.2}`$ where $`\mathrm{\Gamma }_{0.2}\mathrm{\Gamma }/0.2`$, as the maximum nonlinear length scale that still allows a linear analysis to be appropriate. ## 3 Redshift of applicability Having determined $`R_V`$, we now consider what scenarios of reionization are compatible with a Vishniac effect. These scenarios can be roughly divided into two classes: those in which the dominant source of ionizing photons is due to stars formed in dwarf galaxies with halo masses $`10^6M_{}`$ (Couchman & Rees 1986; Fukugita & Kawakasi 1994; Shapiro, Giroux, & Babul 1994; Haiman & Loeb 1997) and models in which reionization occurs due to active nuclei in galaxies with halo masses $`10^9M_{}`$ (Efstathiou & Rees 1988; Haehnelt & Rees 1993; Aghanim et al. 1995; Haiman & Loeb 1998; Valageas & Silk 1999). See also, however, the issues raised in Madau, Haardt, & Rees (1998) and Miralda-Escudé, Haehnelt, & Rees (1998), and the more exotic scenarios of reionization described in Scott, Rees, & Sciama (1991) and Adams, Sarkar, & Sciama (1998). Models in which smaller objects are the most important predict redshifts of reionization of $`z_{\mathrm{re}}20`$ while models in which reionization is due to objects on scales $`10^9M_{}`$ predict more modest values of $`z_{\mathrm{re}}`$. All models of reionization, however, are constrained by the lack of a Gunn-Peterson absorption trough in the spectra of high-redshift quasars, implying that the intergalactic medium was highly deficient in neutral hydrogen at redshifts $`5.`$ Thus, high-mass reionization scenarios can only be successful in cosmologies in which the parameters are such that relatively large objects became nonlinear at high redshifts, precisely those models where the linear approximation is on the most shaky ground. Using our value for $`R_V`$ from Sec. 2, we are able to determine a redshift of applicability, $`z_V`$, below which the Vishniac approximation is invalid over the measurable range of $`\mathrm{}`$ scales. The number density of objects above a critical over-density, $`\delta _c`$, is given by Press-Schechter theory (Press & Schechter 1974) as $$\frac{dn(M,z)}{dM}=\sqrt{\frac{2}{\pi }}\frac{\rho (z)}{M}\frac{\delta _cD_0}{\sigma (M)^2D(z)}\mathrm{exp}\left(\frac{\delta _c^2D_0^2}{2\sigma ^2(M)D(z)^2}\right)\frac{d\sigma }{dM}(M),$$ (22) where $`n(M,z)`$ is the number density of collapsed objects per unit mass at a redshift of $`z`$, $`\rho (z)`$ is the comoving density of the universe at a redshift of $`z`$, $`D(z)`$ is the linear growth factor of fluctuations, and $`\sigma (M)`$ is the level of fluctuations on the mass scale corresponding to a sphere containing a mass $`M`$, which can be computed from Eq. (34). Typically, this formula is used to determine the number density of virialized halos. In this case $`\delta _c(z)`$ is a weak function of $`z`$ for open models and $`\mathrm{\Lambda }`$ models and a fixed value of $`3(12\pi )^{2/3}/201.69`$ in the $`\mathrm{\Omega }_0=1`$ case (Kitayama & Suto 1996). As a rough rule of thumb we can assume that reionization takes place when the 2$`\sigma `$ fluctuations at the relevant scale have collapsed. In this case, $`D(z_{\mathrm{re}})/D_0=\delta _c(z_{\mathrm{re}})/(2\sqrt{2}\sigma (M))`$ which is $`0.6/\sigma (M)`$ in the flat case. We take the linear approximation to be valid up the point at which the 1$`\sigma `$ scale fluctuations at the $`R_V`$ scale have reached an overdensity of 1. This gives $`D(z_V)/D_0=1/(\sqrt{2}\sigma (R_V))0.7/\sigma (R_V).`$ In Fig. 2 we plot both $`z_{\mathrm{re}}`$ and $`z_V`$ as functions of mass, as it is the mass scale rather than the length scale that is most easily identified with different reionization scenarios. We consider three cosmologies, representative of parameters that favor both low mass-scale and high mass-scale scenarios of reionization. In order to compare with a scenario that is representative of stellar reionization, we consider a flat model normalized at the 8 Mpc/$`h`$ scale (Viana & Liddle 1996). Here $`\mathrm{\Omega }_0`$ = 1.0, $`\mathrm{\Omega }_\mathrm{\Lambda }`$ = 0.0, $`\mathrm{\Omega }_b`$ = 0.07, $`h`$ = 0.5, $`\sigma _8`$ = 0.60, and $`n`$ = 1.0. In this scenario $`\mathrm{\Gamma }=0.44`$, shifting the CDM line and decreasing $`R_V`$ by a factor of $`2`$. Note that this value of $`\mathrm{\Gamma }`$ is incompatible with the observed galaxy correlation function. Typical of high-mass reionization scenarios, we consider the “concordance model” of Ostriker & Steinhart (1995), which was used by Haiman & Loeb (1998a) in their modeling of reionization by quasars. In this case the parameters are taken to be $`\mathrm{\Omega }_0`$ = 0.35, $`\mathrm{\Omega }_\mathrm{\Lambda }`$ = 0.65, $`\mathrm{\Omega }_b`$ = 0.04, $`h`$ = 0.65, $`\sigma _8`$ = 0.87, $`\mathrm{\Gamma }=0.20`$, and $`n`$ = 0.96. Finally, we examine an open model with $`\mathrm{\Omega }_0=0.35`$, again normalized at 8 Mpc/$`h`$. In this case $`\mathrm{\Omega }_\mathrm{\Lambda }`$ = 0.0, $`\mathrm{\Omega }_b`$ = 0.04, $`h`$ = 0.65, $`\sigma _8`$ = 1.02, $`\mathrm{\Gamma }=0.20`$, and $`n`$ = 1.0. Let us first consider the $`\mathrm{\Omega }_0=1`$ model, represented by the lowest pair of lines. In this model $`\sigma _8`$ is the lowest of all the cosmologies considered and the perturbations evolve the most quickly. The combination of these two effects moves the redshift of applicability down to a value of $`z_V2`$, and thus one might expect to find an appreciable Vishniac effect. The problem, however, is that the low normalization and rapid evolution of perturbations also lowers the collapse redshift of the objects responsible for reionization. In this scenario, the $`2\sigma `$, $`10^9M_{}`$ peaks collapse at $`z_{\mathrm{re}}7`$. As reionization must have occurred with a high degree of efficiency by $`z=5`$, this scenario is only marginally consistent with quasar absorption-line observations. Thus a flat cosmology is most compatible with low-mass reionization scenarios. If the collapse of $`10^6M_{}`$ halos is responsible for reionization, then $`z_{\mathrm{re}}13`$ for this model, yielding a larger range of redshifts over which the Vishniac effect could be imprinted on the microwave background. This redshift is comparable to the revised values calculated in the stellar reionization scenario of Haiman and Loeb (1997; 1998b), although they consider a somewhat higher range of $`\sigma _8`$ values. To quantify this further, in Figs. 3 and 4 we replot Fig. 2, replacing the vertical axis with $`_0^{w(z)}𝑑wG(w)^2/w_{\mathrm{ang}}^2P_V(\mathrm{}/w_{\mathrm{ang}})`$, the contribution to $`C_{\mathrm{}}`$ due to bulk motions within a redshift of $`z`$. We take $`\mathrm{}=4000`$ in Fig. 3 and $`\mathrm{}=12000`$ in Fig. 4, and normalize the $`y`$ axis such that the integral is equal to 1 at a redshift of $`z_{\mathrm{re}}(10^6M_{})`$. The magnitude of the Vishniac $`C_{\mathrm{}}`$ is then directly proportional to the vertical width of the gap between $`_0^{w(z_{\mathrm{res}})}𝑑wG(w)^2/w_{\mathrm{ang}}^2P_V(\mathrm{}/w_{\mathrm{ang}})`$, and $`_0^{w(z_V)}𝑑wG(w)^2/w_{\mathrm{ang}}^2P_V(\mathrm{}/w_{\mathrm{ang}})`$, allowing us to judge the linear and nonlinear contributions to Eq. (18) in arbitrary scenarios of reionization at a glance. From this point of view, reionization by $`10^6M_{}`$ objects results in only a marginal improvement the accuracy of a linear treatment. While 35% of the $`\mathrm{}=4000`$ integral is nonlinear if $`M_{\mathrm{re}}=10^9M_{}`$, the $`M_{\mathrm{re}}=10^6M_{}`$ case is still 25% inaccurate. These numbers are somewhat lower in the high-$`\mathrm{}`$ case, in which 18% of the integral is nonlinear in the low mass case, and 25% in the high mass. Note however, that our definition of $`R_V=2\mathrm{Mpc}(h\mathrm{\Gamma }/0.2)`$ was based on the damping of perturbations at $`\mathrm{}=4000.`$ From Fig. 1 we see that perturbations at $`\mathrm{}=12000`$ are largely damped when the matter power spectrum is filtered at the $`R=1\mathrm{Mpc}(h\mathrm{\Gamma }/0.2)`$ scale, and a more fair comparison between Figs. 3 and 4 would be to shift the vertical lines in Fig. 4 to masses lower by a factor of 8, yielding much the same numbers as in the $`\mathrm{}=4000`$ case. More typical of high-mass reionization scenarios is the $`\mathrm{\Lambda }`$CDM model represented by the dotted lines in Figs. 2-4. In this model, $`\sigma _8`$ is slightly higher than in the flat case and the evolution of $`D(z)`$ is slowed. These effects raise the collapse redshift of $`10^9M_{}`$ peaks to $`z_{\mathrm{re}}11`$, easily compatible with Gunn-Peterson tests. Note that our crude estimate of the redshift of reionization is almost the same as the redshift of $`12`$ calculated by Haiman & Loeb (1998a) for the same set of cosmological parameters, using a more sophisticated Press-Schechter based argument for the ionizing flux from quasars. In this scenario $`z_V`$ is also pushed back, although this is lessened by the shift of $`\mathrm{\Gamma }`$ as compared to the flat case. Thus $`z_V2.5`$. One might imagine that in this case, a much wider margin of redshifts would lead to an accurate linear calculation. Fig. 3 indicates otherwise. In this case almost 20% of the low-mass and over 25% of the high-mass integral comes from redshifts at which $`R_V`$ is nonlinear. The high-$`\mathrm{}`$ values are slightly lower, with 12% of the $`10^6M_{}`$ and 17% of the $`10^9M_{}`$ integral taking place when $`R_V`$ is nonlinear, but again these numbers become roughly the same as the $`\mathrm{}=4000`$ case for a more fair comparison. The most extreme case we consider is the cluster-normalized open model, in which $`\sigma _8`$ is the highest and $`D(z)`$ the most slowly evolving. In this case $`z_V4`$, the $`10^9M_{}`$ and the $`10^6M_{}`$ schemes reionize at $`z_{\mathrm{re}}22`$ and $`z_{\mathrm{re}}33`$ respectively. This cosmology yields the largest regime of redshift space over which a linear analysis is valid and the most accurate results. Here nonlinear $`R_V`$ scale fluctuations contaminate 15% of the low-mass and 20% of the high-mass $`C_{4000}`$ integrals. Again these numbers are lower at higher $`\mathrm{}`$ but roughly the same after accounting for the smaller filtering scale of $`R=1\mathrm{Mpc}(h\mathrm{\Gamma }/0.2)`$. As a final check of the validity of our analysis, we construct $`C_{\mathrm{},\mathrm{filtered}}`$ defined as the angular power spectrum as given by Eq. (18) but replacing $`P(k)`$ with $`P(k)W^2(kR_{\mathrm{nl}}(z))`$ where $`R_{\mathrm{nl}}(z)`$ is now the nonlinear length scale at each redshift as in Fig. 2 ($`D(z)/D_0=0.7/\sigma (R_{\mathrm{nl}}(z))`$). In Fig. 5 we plot the ratio of $`C_{\mathrm{},\mathrm{filtered}}`$ to $`C_{\mathrm{}}`$ calculated from the unfiltered CDM power spectrum. While only a few reionization scenarios are represented in this graph, this nevertheless gives us some feel of the accuracy of the linear treatment over different scales, and unlike Figs. 2 - 4, is completely independent of our definition of $`R_V`$. Here we see that in the range of $`\mathrm{}`$ values at which the effect is most likely to be measured ($`4000\mathrm{}120000`$) this estimate is in good agreement with the accuracies given in the previous figures. Note also that a linear treatment becomes increasingly inaccurate with $`\mathrm{}`$, and thus measurements of fluctuations at angular scales just below this Silk damping scale will be most easy to interpret. From these results we can safely conclude that even given our present ignorance as to the cosmological parameters, no more than $`85\%`$ of the contribution to Eq. (18) for measurable $`\mathrm{}`$ values can be calculated by linear theory in a CDM cosmogony. This depends only on the shape of the CDM power spectrum and the lack of diffuse L$`\alpha `$ absorption in quasar spectra out to $`z5`$. ## 4 Is There a Detectable Vishniac Effect, Really? At this point, one may raise the objection that our argument is a bit semantic, as there will still be scattering due to bulk motions even when linear theory breaks down. Indeed, the Kinetic Sunyaev-Zel’dovich effect, which is due to the peculiar velocities of clusters, can be viewed as a nonlinear counterpart to the Vishniac effect. Is there not, then, a sort of detectable Vishniac effect in cosmological scenarios in which $`R_V`$ is nonlinear during the epoch of reionization, albeit under a different name? The problem with this point of view is that it overlooks some important physical distinctions between these effects. The Vishniac effect is due to the presence of a redshift regime in which $`G(w)`$ is slowly varying and a delicate cancellation takes place due to the lack of curl in the peculiar velocity field. As it can be calculated precisely, it provides a unique probe of the reionization history of the universe that is not available from measurements of nonlinear fluctuations. Furthermore, the Vishniac effect displays a distinct signature of higher-order moments that allows it to be distinguished from other contributions. In order to understand why this occurs, let us consider the bispectrum $`B(\stackrel{}{\kappa }_1,\stackrel{}{\kappa }_2)(2\pi )^2\delta ^2(\stackrel{}{\kappa }_1+\stackrel{}{\kappa }_2+\stackrel{}{\kappa }_3)\stackrel{~}{\frac{\mathrm{\Delta }T}{T}}(\stackrel{}{\kappa }_1)\stackrel{~}{\frac{\mathrm{\Delta }T}{T}}(\stackrel{}{\kappa }_2)\stackrel{~}{\frac{\mathrm{\Delta }T}{T}}(\stackrel{}{\kappa }_3)`$. We can apply our $`\delta `$-function beam approach in the linear regime to calculate this as $`\stackrel{~}{{\displaystyle \frac{\mathrm{\Delta }T}{T}}}(\stackrel{}{\kappa }_1)\stackrel{~}{{\displaystyle \frac{\mathrm{\Delta }T}{T}}}(\stackrel{}{\kappa }_2)\stackrel{~}{{\displaystyle \frac{\mathrm{\Delta }T}{T}}}(\stackrel{}{\kappa }_3)=i{\displaystyle \underset{i=1}{\overset{3}{}}}\left[{\displaystyle _0^{w_0}}𝑑w_iG(w_i)\right]{\displaystyle \underset{j=1}{\overset{6}{}}}\left[{\displaystyle \frac{d^3𝐤_𝐣}{(2\pi )^3}}\right]`$ $`(2\pi )^6\delta ^2((\stackrel{}{k}_1+\stackrel{}{k}_2)w_{1,\mathrm{ang}}\stackrel{}{\kappa }_1)\delta ^2((\stackrel{}{k}_3+\stackrel{}{k}_4)w_{2,\mathrm{ang}}\stackrel{}{\kappa }_2)\delta ^2((\stackrel{}{k}_5+\stackrel{}{k}_6)w_{3,\mathrm{ang}}\stackrel{}{\kappa }_3)`$ $`e^{i(k_{1,z}+k_{2,z})w_1+i(k_{3,z}+k_{4,z})w_2+i(k_{5,z}+k_{6,z})w_3}{\displaystyle \frac{k_{1,z}}{k_1^2}}{\displaystyle \frac{k_{3,z}}{k_3^2}}{\displaystyle \frac{k_{5,z}}{k_5^2}}{\displaystyle \underset{l=1}{\overset{6}{}}}\stackrel{~}{\delta _0}(𝐤_𝐥,w_i).`$ (23) If $`G(w)`$ is slowly varying, we can again follow Kaiser (1992) in dividing the integrals over comoving distance into $`N`$ statistically independent intervals of width $`\mathrm{\Delta }w`$, $`\stackrel{~}{{\displaystyle \frac{\mathrm{\Delta }T}{T}}}(\stackrel{}{\kappa }_1)\stackrel{~}{{\displaystyle \frac{\mathrm{\Delta }T}{T}}}(\stackrel{}{\kappa }_2)\stackrel{~}{{\displaystyle \frac{\mathrm{\Delta }T}{T}}}(\stackrel{}{\kappa }_3)=i{\displaystyle \underset{n=1}{\overset{N}{}}}G(w_n)^3\mathrm{\Delta }w^3{\displaystyle \underset{j=1}{\overset{6}{}}}\left[{\displaystyle \frac{d^3𝐤_𝐣}{(2\pi )^3}}\right](2\pi )^6\delta ^2((\stackrel{}{k}_1+\stackrel{}{k}_2)w_{n,\mathrm{ang}}\stackrel{}{\kappa }_1)`$ $`\delta ^2((\stackrel{}{k}_3+\stackrel{}{k}_4)w_{n,\mathrm{ang}}\stackrel{}{\kappa }_2)\delta ^2((\stackrel{}{k}_5+\stackrel{}{k}_6)w_{n,\mathrm{ang}}\stackrel{}{\kappa }_3)j_0\left({\displaystyle \frac{(k_{1,z}+k_{2,z})\mathrm{\Delta }w}{2}}\right)`$ $`j_0\left({\displaystyle \frac{(k_{3,z}+k_{4,z})\mathrm{\Delta }w}{2}}\right)j_0\left({\displaystyle \frac{(k_{5,z}+k_{6,z})\mathrm{\Delta }w}{2}}\right){\displaystyle \frac{k_{1,z}}{k_1^2}}{\displaystyle \frac{k_{3,z}}{k_3^2}}{\displaystyle \frac{k_{5,z}}{k_5^2}}{\displaystyle \underset{l=1}{\overset{6}{}}}\stackrel{~}{\delta _0}(𝐤_𝐥,w_n).`$ (24) As there are an odd number of $`k_z`$ terms, there is no pairing of density fields that does not result in an odd $`k_z`$ term. As all the $`k`$ integrals are even, $`\stackrel{~}{\frac{\mathrm{\Delta }T}{T}}(\stackrel{}{\kappa }_1)\stackrel{~}{\frac{\mathrm{\Delta }T}{T}}(\stackrel{}{\kappa }_2)\stackrel{~}{\frac{\mathrm{\Delta }T}{T}}(\stackrel{}{\kappa }_3)`$ = 0. This cancellation can be understood from a more general perspective. $`B(\stackrel{}{\kappa }_1,\stackrel{}{\kappa }_2)`$ is generated by the expectation value of the triple product of the field $`𝐪`$. As $`𝐪`$ is an isotropic vector field, $`\stackrel{~}{q}_i(𝐤_\mathrm{𝟏})\stackrel{~}{q}_j(𝐤_\mathrm{𝟐})\stackrel{~}{q}_k(𝐤_\mathrm{𝟑})`$ can depend on no vectors other than the $`𝐤`$ vectors themselves and must therefore be proportional to at least one of the them (Monin & Yaglom 1971). This means that the $`i=j=l=z`$ component of this product must be proportional to $`k_{1,z}`$, $`k_{2,z}`$, or $`k_{3,z}`$. Thus the odd $`k_z`$ term that results in $`B(\stackrel{}{\kappa }_1,\stackrel{}{\kappa }_2)=0`$ is due to the isotropy of the density fluctuations. By a similar argument, all odd moments of the temperature fluctuations must be zero as these also depend on the expectation value of the product of an odd number of $`𝐪`$’s. Note that this is true independent of the gaussianity of the probability distribution functional of $`\stackrel{~}{\delta }(𝐤)`$. This cancellation does not apply to the even moments, however, as an even number of $`𝐪`$’s can be arranged in a way that is not proportional to one of the $`𝐤`$ vectors. Thus the Vishniac effect is unique in that it is nongaussian independent of the gaussianity of the probability distribution functional of $`\stackrel{~}{\delta }(𝐤)`$, but this nongaussianity is expressed only in the even higher-order moments. This alternation of zero-and-nonzero higher-order moments provides a unique signal that distinguishes the Vishniac effect from other secondary anisotropies, and provides us with the opportunity to use nongaussian statistics as a discriminator between these contributions. Note however, that the difficulty of measuring secondary anisotropies may make such an analysis difficult to apply in practice. ## 5 Conclusions Due to the tremendous predictive power of linear theory, comparisons between linear predictions and large-scale cosmic microwave background measurements promise to constrain cosmological parameters to the order of a few percent (Jungman et al. 1996; Bond, Efstathiou, & Tegmark 1997; Zaldarriaga, Spergel, & Seljak 1997). The natural extension of this approach is to try to measure small-scale secondary anisotropies and match them to linear predictions to study the reionization history of universe. The situation in this case is more muddled, however, as a number of nonlinear secondary effects also contribute at these scales. The dominant secondary linear anisotropy is a second-order contribution known as the Vishniac or Ostriker-Vishniac effect. As this effect can be predicted accurately as a function of cosmological parameters, several authors have proposed that its measurement will prove to be a sensitive probe of the reionization history of the universe. Reionization occurs by the formation of nonlinear structures, however, raising the question of whether a regime of redshift space exists in which these objects have collapsed but a linear analysis is still appropriate. In this work, we have determined the relevant physical scales that give rise to the Vishniac effect in a Cold Dark Matter cosmogony, showing that approximations are already compromised when $`1`$ Mpc/($`h\mathrm{\Gamma }_{0.20}`$) scales have become nonlinear, and break down when $`2`$ Mpc/($`h\mathrm{\Gamma }_{0.20}`$) dark matter halos reach overdensities of 1. The width of the redshift regime over which the effect can be imprinted on the CMB is dependent on the cosmological parameters and the reionizing mass scale. Schemes in which reionization is due to radiation from active galactic nuclei associated with dark matter halos of masses $`10^9M_{}`$ are limited by the absence of a Gunn-Peterson absorption trough. As reionization must have occurred with a high degree of efficiency before a redshift of 5, such models are successful only if one assumes a large value of $`\sigma _8`$, or considers open models with slowly-changing linear growth factors. Both these assumptions push back the redshift at which $`R_V`$ becomes nonlinear, limiting the range over which a linear analysis is appropriate. Scenarios in which reionization is due to much smaller objects, such as stars formed in dwarf galaxies associated with dark matter halos of masses $`10^6M_{}`$, are able to reionize the universe at much larger redshifts even in cosmologies in which $`\sigma _8`$ is small and $`D(z)`$ quickly evolving. This represents only a marginal gain however, as the high redshift contribution to the Vishniac integral is roughly proportional to comoving distance, and comoving distances are small at high redshifts. Thus low-mass scenarios of reionization are more compatible with a linear analysis not so much because they reionize earlier as because they allow $`R_V`$ to become nonlinear more recently without violating Gunn-Peterson limits. The Vishniac effect arises from physical processes that are distinct from nonlinear secondary anisotropies. Its detection indicates the presence of a redshift regime in which a delicate cancellation takes place due to the lack of curl in the peculiar velocity field and slow variations in $`G(w)`$. This leaves a unique signature in the higher-order moments of the temperature fluctuations that is absent from its nonlinear counterparts. Furthermore, due to the predictive power of linear theory, it represents a sensitive probe of the reionization history not available from measurements of nonlinear contributions. As with measurements of large angular scale anisotropies, small-scale microwave background anisotropy measurements have the potential to uncover much about the history of our universe. Also as with large-scale measurements, whether this potential will be realized remains to be seen. While the Vishniac effect represents a possible probe of the reionization epoch, the analysis will, as always, be more involved than first suggested. Ultimately it will only be through the measurement and analysis of small-scale microwave background anisotropies that we will be able to know if there is a detectable Vishniac effect. I wish to thank Nabila Aghanim, François Bouchet, Rychard Bouwens, Andrew Jaffe, Douglas Scott, and Naoshi Sugiyama for helpful discussions and am particularly indebted to Joseph Silk, whose comments and suggestions have been invaluable during the preparation of this work. I thank Uroš Seljak and Matias Zaldarriaga for the use of CMBFAST and acknowledge partial support by the NSF. ## Appendix In this appendix, we provide explicit expressions for the cosmological factors used throughout this paper. We allow both $`\mathrm{\Omega }_0`$ and $`\mathrm{\Omega }_\mathrm{\Lambda }`$ to be free. In this case the Friedman equations for the evolution of the scale factor of the Universe, $`a(z)`$ are $$\frac{\dot{a}}{a}=H_0E(z)H_0\sqrt{\mathrm{\Omega }_0(1+z)^3+\mathrm{\Omega }_\mathrm{\Lambda }+(1\mathrm{\Omega }_0\mathrm{\Omega }_\mathrm{\Lambda })(1+z)^2},$$ (25) and $$\frac{\ddot{a}}{a}=H_0^2[\mathrm{\Omega }_\mathrm{\Lambda }\mathrm{\Omega }_0(1+z)^3/2],$$ (26) where $`H_0=100h`$ km sec<sup>-1</sup> Mpc<sup>-1</sup> is the Hubble constant, and the overdot denotes a derivative with respect to time. We choose the scale factor such that $`a_0H_0=2c`$. If we are located at the origin, $`w=0`$, then an object at redshift $`z`$ is at a comoving distance, $`w(z)=\frac{1}{2}_0^z𝑑z^{}E^1(z^{})`$ and at a time $`t(z)=\frac{1}{H_0}_z^{\mathrm{}}𝑑z^{}(1+z^{})^1E^1(z^{}).`$ Note that this is different than conformal time, defined by $`d\eta =dt/a`$, such that $`c\eta (z)=w(\mathrm{})w(z)`$. For any flat universe the angular size distance $`w_{\mathrm{ang}}=w`$, and in an open universe $$w_{\mathrm{ang}}=\frac{\mathrm{sinh}(2w\sqrt{1\mathrm{\Omega }_0\mathrm{\Omega }_\mathrm{\Lambda }})}{2\sqrt{1\mathrm{\Omega }_0\mathrm{\Omega }_\mathrm{\Lambda }}}.$$ (27) The growth factor as a function of redshift is $$D(z)=\frac{5\mathrm{\Omega }_0E(z)}{2}_z^{\mathrm{}}\frac{1+z^{}}{[E(z^{})]^3}𝑑z^{},$$ (28) while $$\frac{\dot{D}}{D}=\frac{\ddot{a}}{\dot{a}}\frac{\dot{a}}{a}+\frac{5\mathrm{\Omega }_0}{2}\frac{\dot{a}}{a}\frac{(1+z)^2}{[E(z)]^2D(z)}.$$ (29) The evolution of $`\mathrm{\Omega }`$ is given by $$\mathrm{\Omega }(z)=\mathrm{\Omega }_0(1+z)^3E^2(z)$$ (30) where $`\mathrm{\Omega }_0=\mathrm{\Omega }(0).`$ For the power spectrum, we use $`P(k)`$ $`=`$ $`{\displaystyle \frac{2\pi ^2}{8}}\delta _H^2(k/2)^nT^2(k_p\mathrm{Mpc}/h\mathrm{\Gamma }),`$ (31) where $`T(q)`$ is the CDM transfer function, $`k_p=k/a_0=kH_0/2c`$ is the physical wavenumber, and the shape parameter $`\mathrm{\Gamma }`$ is defined as (Sugiyama 1995) $$\mathrm{\Gamma }\mathrm{\Omega }_0(h/0.5)\mathrm{exp}(\mathrm{\Omega }_b\mathrm{\Omega }_b/\mathrm{\Omega }_0).$$ (32) The factor of 8 in the denominator in Eq. (31) arises because we are using $`a_0H_0=2c`$. For the transfer function, we use the analytic fit given by Bardeen et al. (1986) for the CDM cosmogony, $$T(q)=\frac{\mathrm{ln}(1+2.34q)/(2.34q)}{[1+3.89q+(16.1q)^2+(5.46q)^3+(6.71q)^4]^{1/4}}.$$ (33) The normalization factor $`\delta _H`$ can be fixed by specifying $`\sigma (8\mathrm{M}\mathrm{p}\mathrm{c}/h)`$, where the variance of the mass enclosed in a sphere of radius $`R`$ is given by $$\sigma ^2(R)=\frac{1}{2\pi ^2}_0^{\mathrm{}}k^2𝑑kP(k)W^2(k_pR).$$ (34) Here $`W(x)`$ is the spherical top-hat window function, defined in Fourier space as $$W(x)3\left[\frac{\mathrm{sin}(x)}{x^3}\frac{\mathrm{cos}(x)}{x^2}\right].$$ (35)
warning/0002/hep-ph0002046.html
ar5iv
text
# Higher Twist Contributions To 𝑅-Hadron Phenomenology In The Light Gluino Scenario ## Introduction The gluino is the supersymmetric partner of the gluon. It is an electromagnetically neutral, adjoint fermion with the same color structure as its boson counterpart. As yet, no clear experimental evidence of supersymmetric particles has been found. The most likely reason for this is the large expected mass of the supersymmetric particles ($`\mathrm{\Lambda }_{\mathrm{SUSY}}1`$ TeV). However, an intriguing scenario exists whereby the gluino is not only the lightest supersymmetric particle but also very light compared to the SUSY scale, $`m_{\stackrel{~}{g}}100`$ GeV. This possibility arises naturally in a number of quite attractive models characterized by special boundary conditions at the grand unification scale and in certain models of gauge-mediated supersymmetry breaking . Light gluinos are predicted to form relatively light bound states of quarks or gluons and gluinos called $`R`$-hadrons . The lightest predicted $`R`$-hadrons include mesinos ($`q\overline{q}\stackrel{~}{g}`$), two barinos, $`R^+`$($`uud\stackrel{~}{g}`$) and $`S^0`$($`uds\stackrel{~}{g}`$), gluinoballs ($`\stackrel{~}{g}\stackrel{~}{g}`$), and the glueballino or $`R^0`$ ($`\stackrel{~}{g}g`$). The properties of $`R`$-hadrons including their mass, decay modes, and lifetimes depend strongly on the mass of the gluino. There have been many theoretical and experimental attempts to find evidence for and/or exclude the light gluino scenario. Searches for $`R`$-hadrons produced in fixed target experiments have been performed for a number of the predicted $`R`$-hadron decay channels . Effects of a light gluino on QCD observables have been analyzed . Stable particle searches, $`\mathrm{{\rm Y}}`$ decays, beam dump experiments etc. all have potential sensitivity to the presence of a light gluino or the $`R`$-hadrons. A brief summary of the various possible resulting constraints on a light gluino is given in Ref. . In addition, Ref. claims that $`m_{\stackrel{~}{g}}>2.53\mathrm{GeV}`$ is excluded on the basis of their analysis of OPAL data. Although these various analyses are, in combination, potentially sensitive to most regions of light gluino mass, all rely on model-dependent inputs. As a result, we believe that at present it is impossible to definitively exclude any gluino mass below $`45\mathrm{GeV}`$. Thus, it is of great interest to find additional approaches for discovering and/or constraining light gluinos and the $`R`$-hadrons. In this paper, we will explore the possibility of detecting $`R`$-hadrons at large $`x_F`$ in $`pp`$ and $`pA`$ fixed-target interactions. Our calculations will be restricted to the $`m_{\stackrel{~}{g}}1.25`$ GeV region where we can be confident that the semi-perturbative techniques that we employ are reliable. This region is of particular phenomenological interest because of the analogy that can be drawn between heavy quark and light gluino production. Indeed, if the gluino and heavy quark masses are comparable, one might anticipate observation of hard gluino production analogous to that already observed in high-$`x_F`$ charm hadroproduction . The leading-twist pQCD predictions for charm production in $`pp`$ and $`pA`$ collisions fail to account for many features of the high-$`x_F`$ data. These include unexpectedly large production rates and anomalies such as flavor correlations between the produced hadrons and the valence spectators, manifested as leading charm and a strong $`D^+/D^{}`$ asymmetry in $`\pi ^{}A`$ interactions , double $`J/\mathrm{\Psi }`$ production at large $`x_F`$ , and Feynman scaling of $`J/\mathrm{\Psi }`$ production in $`pA`$ interactions , all of which suggest a breakdown of factorization at large $`x_F`$. The anomalies and cross section enhancement may be partly explained by higher twist terms in the operator product expansion (OPE) on the light cone associated with the dynamics of the QCD bound state. Analgous terms should be present for light gluinos. The intrinsic charm model (IC) approximates non-perturbative higher twist Fock-state contributions of heavy quarks in hadronic wave functions. The phenomenological predictions of IC directly address the above puzzles in charm hadroproduction . For example, IC provides a coalescence mechanism whereby final state hadrons can share valence quarks with the projectile, naturally producing leading particles. In analogy with leading charm, we study $`R`$-hadron distributions using “intrinsic gluinos” (I$`\stackrel{~}{\mathrm{G}}`$) in regions of phase space where the gluino mass and momentum fractions conspire so that higher twist effects cannot be ignored. In this paper, we calculate enhancements over the leading twist $`R`$-hadrons $`x_F`$ distributions with gluino masses $`m_{\stackrel{~}{g}}=1.2`$, $`1.5`$, $`3.5`$, and $`5.0`$ GeV. Both $`pp`$ and $`pA`$ interactions at $`p_{\mathrm{lab}}=800`$ GeV are considered. ## pQCD Light Gluino Hadroproduction In pQCD, gluinos are produced in pairs by $`gg`$ fusion and $`q\overline{q}`$ annihilation, $`gg,q\overline{q}\stackrel{~}{g}\stackrel{~}{g}`$, as well as quark-gluon scattering to squark and gluino, $`qg\stackrel{~}{q}\stackrel{~}{g}`$. Precision Z-pole data has constrained the squark mass to be greater than $`100`$ GeV, quite large compared to the light gluino masses considered here. Therefore, we expect that the $`qg`$ contribution with the virtual squark in the $`t`$-channel will be small compared to the other contributions, particularly at fixed-target energies. The leading twist inclusive $`R`$-hadron $`x_F`$ distribution at leading order is obtained from the gluino $`x_F`$ distribution ($`x_F=(2m_T/\sqrt{s})\mathrm{sinh}y`$) which has the factorized form in pQCD $$\frac{d\sigma }{dx_F}=\underset{i,j}{}\frac{\sqrt{s}}{2}𝑑z_3𝑑y_2d^2p_T\frac{1}{E_1}\frac{D_{H/\stackrel{~}{g}}(z_3)}{z_3}f_i^A(x_a)f_j^B(x_b)\frac{1}{\pi }\frac{d\widehat{\sigma }_{ij}}{d\widehat{t}}.$$ (1) Here $`a`$ and $`b`$ are the initial partons from projectile and target hadrons $`A`$ and $`B`$, $`1`$ and $`2`$ are the produced gluinos, and $`3`$ is the final-state $`R`$-hadron. The sum over $`i`$ and $`j`$ extends over all partonic gluino production subprocesses. A $`K`$ factor of 2.5 is included to account for NLO corrections. Since the $`K`$ factor is approximately constant with $`x_F`$ for charm production except as $`x_F1`$, we assume that the $`K`$ factor for gluino production is also independent of $`x_F`$. The fragmentation functions, $`D_{H/\stackrel{~}{g}}(z)`$ with $`z=x_H/x_{\stackrel{~}{g}}`$, describe the collinear fragmentation of final state $`R`$-hadrons from the produced gluinos. For simplicity, a delta function was used for hadronization, $`D_{H/\stackrel{~}{g}}(z)=\delta (z1)`$. This assumption results in the hardest $`x_F`$ distribution at leading twist since the $`R`$-hadron carries all of the gluino’s momentum. Other fragmentation functions would soften these distributions. Note that for any fragmentation function to factorize, it must be independent of the initial state (i.e. it only depends on $`z_3`$ and not $`x_a`$). Thus, regardless of the fragmentation function used, all $`R`$-hadrons will be decoupled from the initial state to leading twist. The partonic cross sections for gluino production in Eq. (1) are $`{\displaystyle \frac{d\widehat{\sigma }_{gg\stackrel{~}{g}\stackrel{~}{g}}}{d\widehat{t}}}`$ $`=`$ $`{\displaystyle \frac{9\pi \alpha _s^2}{4\widehat{s}^2}}[{\displaystyle \frac{2(m_{\stackrel{~}{g}}^2\widehat{t})(\widehat{u}m_{\stackrel{~}{g}}^2)}{\widehat{s}^2}}+{\displaystyle \frac{m_{\stackrel{~}{g}}^2(\widehat{s}4m_{\stackrel{~}{g}}^2)}{(m_{\stackrel{~}{g}}^2\widehat{t})(\widehat{u}m_{\stackrel{~}{g}}^2)}}`$ (4) $`+{\displaystyle \frac{(m_{\stackrel{~}{g}}^2\widehat{t})(\widehat{u}m_{\stackrel{~}{g}}^2)2m_{\stackrel{~}{g}}^2(m_{\stackrel{~}{g}}^2+\widehat{t})}{(m_{\stackrel{~}{g}}^2\widehat{t})^2}}+{\displaystyle \frac{(m_{\stackrel{~}{g}}^2\widehat{t})(\widehat{u}m_{\stackrel{~}{g}}^2)+m_{\stackrel{~}{g}}^2(\widehat{u}\widehat{t})}{\widehat{s}(m_{\stackrel{~}{g}}^2\widehat{t})}}`$ $`+{\displaystyle \frac{(m_{\stackrel{~}{g}}^2\widehat{u})(\widehat{t}m_{\stackrel{~}{g}}^2)2m_{\stackrel{~}{g}}^2(m_{\stackrel{~}{g}}^2+\widehat{u})}{(m_{\stackrel{~}{g}}^2\widehat{u})^2}}+{\displaystyle \frac{(m_{\stackrel{~}{g}}^2\widehat{u})(\widehat{t}m_{\stackrel{~}{g}}^2)+m_{\stackrel{~}{g}}^2(\widehat{t}\widehat{u})}{\widehat{s}(m_{\stackrel{~}{g}}^2\widehat{u})}}]`$ $`{\displaystyle \frac{d\widehat{\sigma }_{q\overline{q}\stackrel{~}{g}\stackrel{~}{g}}}{d\widehat{t}}}`$ $`=`$ $`{\displaystyle \frac{\pi \alpha _s^2}{\widehat{s}^2}}[{\displaystyle \frac{8}{3}}{\displaystyle \frac{(m_{\stackrel{~}{g}}^2\widehat{t})^2+(\widehat{u}m_{\stackrel{~}{g}}^2)^2+2m_{\stackrel{~}{g}}^2\widehat{s}}{\widehat{s}^2}}+{\displaystyle \frac{32}{27}}{\displaystyle \frac{(m_{\stackrel{~}{g}}^2\widehat{t})^2}{(m_{\stackrel{~}{q}}^2\widehat{t})^2}}+{\displaystyle \frac{32}{27}}{\displaystyle \frac{(\widehat{u}m_{\stackrel{~}{g}}^2)^2}{(\widehat{u}m_{\stackrel{~}{q}}^2)^2}}`$ (6) $`+{\displaystyle \frac{8}{3}}{\displaystyle \frac{(m_{\stackrel{~}{g}}^2\widehat{t})^2+m_{\stackrel{~}{g}}^2\widehat{s}}{\widehat{s}(\widehat{t}m_{\stackrel{~}{q}}^2)}}+{\displaystyle \frac{8}{27}}{\displaystyle \frac{m_{\stackrel{~}{g}}^2\widehat{s}}{(m_{\stackrel{~}{q}}^2\widehat{t})(\widehat{u}m_{\stackrel{~}{q}}^2)}}+{\displaystyle \frac{8}{3}}{\displaystyle \frac{(\widehat{u}m_{\stackrel{~}{g}}^2)^2+m_{\stackrel{~}{g}}^2\widehat{s}}{\widehat{s}(\widehat{u}m_{\stackrel{~}{q}}^2)}}]`$ $`{\displaystyle \frac{d\widehat{\sigma }_{gq\stackrel{~}{q}\stackrel{~}{g}}}{d\widehat{t}}}`$ $`=`$ $`{\displaystyle \frac{\pi \alpha _s^2}{\widehat{s}^2}}[{\displaystyle \frac{4}{9}}{\displaystyle \frac{m_{\stackrel{~}{g}}^2\widehat{t}}{\widehat{s}}}+{\displaystyle \frac{(m_{\stackrel{~}{g}}^2\widehat{t})\widehat{s}+2m_{\stackrel{~}{g}}^2(m_{\stackrel{~}{q}}^2\widehat{t})}{(m_{\stackrel{~}{g}}^2\widehat{t})^2}}+{\displaystyle \frac{4}{9}}{\displaystyle \frac{(\widehat{u}m_{\stackrel{~}{g}}^2)(\widehat{u}+m_{\stackrel{~}{q}}^2)}{(\widehat{u}m_{\stackrel{~}{q}}^2)^2}}`$ (10) $`{\displaystyle \frac{(\widehat{s}m_{\stackrel{~}{q}}^2+m_{\stackrel{~}{g}}^2)(m_{\stackrel{~}{q}}^2\widehat{t})m_{\stackrel{~}{g}}^2\widehat{s}}{\widehat{s}(m_{\stackrel{~}{g}}^2\widehat{t})}}+{\displaystyle \frac{1}{18}}{\displaystyle \frac{\widehat{s}(\widehat{u}+m_{\stackrel{~}{g}}^2)+2(m_{\stackrel{~}{q}}^2m_{\stackrel{~}{g}}^2)(m_{\stackrel{~}{g}}^2\widehat{u})}{\widehat{s}(\widehat{u}m_{\stackrel{~}{q}}^2)}}`$ $`+{\displaystyle \frac{1}{2}}{\displaystyle \frac{(m_{\stackrel{~}{q}}^2\widehat{t})(2\widehat{u}+m_{\stackrel{~}{g}}^2+\widehat{t})}{2(\widehat{t}m_{\stackrel{~}{g}}^2)(\widehat{u}m_{\stackrel{~}{q}}^2)}}+{\displaystyle \frac{1}{2}}{\displaystyle \frac{(m_{\stackrel{~}{g}}^2\widehat{t})(\widehat{s}+2\widehat{t}2m_{\stackrel{~}{q}}^2)}{2(\widehat{t}m_{\stackrel{~}{g}}^2)(\widehat{u}m_{\stackrel{~}{q}}^2)}}`$ $`+{\displaystyle \frac{1}{2}}{\displaystyle \frac{(\widehat{u}m_{\stackrel{~}{g}}^2)(\widehat{t}+m_{\stackrel{~}{g}}^2+2m_{\stackrel{~}{q}}^2)}{2(\widehat{t}m_{\stackrel{~}{g}}^2)(\widehat{u}m_{\stackrel{~}{q}}^2)}}].`$ We calculate leading twist pQCD gluino distributions for $`800`$ GeV $`pp`$ interactions. Figure 1 shows the gluino distributions using the MRS D-’ parton distributions in the proton with $`m_{\stackrel{~}{g}}=1.2,1.5,3.5`$, and $`5.0`$ GeV and $`m_{\stackrel{~}{q}}=100`$ GeV. The characteristic falloff at large $`x_F`$ is similar to heavy quark production. Choosing a larger squark mass would only marginally decrease the total cross section because the $`gq`$ channel is suppressed by the large squark mass. The gluino production cross section is a strong function of mass. The cross section is largest for $`m_{\stackrel{~}{g}}=1.2`$ GeV and decreases by a factor of 3 for $`m_{\stackrel{~}{g}}=1.5`$ GeV. There is then a drop of 250 to the $`m_{\stackrel{~}{g}}=3.5`$ GeV gluino cross section and another factor of 20 between the $`3.5`$ and $`5`$ GeV cross sections. Additionally, the falloff of the cross section with $`x_F`$ becomes steeper as $`m_{\stackrel{~}{g}}`$ is increased. Charm hadroproduction phenomenology has taught us that higher twist contributions can become comparable to leading twist in certain parts of phase space, introducing correlations between the initial and final states. These effects will be addressed in the next section. ## Intrinsic Contribution to Higher Twist In deep inelastic scattering, higher twist terms in the OPE are suppressed by a factor of $`1/Q^{2n}`$. These terms are essentially irrelevant when $`Q^2`$ is large. Analogously, in hadroproduction, a similar suppression of $`1/M^2`$ typically renders higher-twist effects unimportant except in regions where pQCD is seemingly inapplicable (i.e. where $`M^2`$ is small). However, it has been shown that in the simultaneous $`M^2\mathrm{}`$ and $`x1`$ limit with $`M^2(1x)`$ fixed, a new hard scale emerges where higher twist contributions to the cross section become comparable to leading twist . In the case of heavy quark production, this new scale can be associated with either the resolution of the transverse size of the intrinsic heavy quark pair or with the transverse resolution of any “pre-coalesced” hadrons inside the parent hadron. The heavy quark fluctuations can carry a large fraction of the projectile’s forward momentum since the constituents of the bound state move with the same velocity. The Fock state may be broken up by an interaction with soft gluons in the target, producing a leading hadron containing a heavy parton. The bound state wave function for a state containing higher-twist contributions can be obtained from the Bethe-Saltpeter formalism evaluated at equal “time” on the light cone : $`(M^2\mathrm{\Sigma }_{i=1}^n{\displaystyle \frac{\widehat{m}_i^2}{x_i}})\mathrm{\Psi }(x_i,k_{T_i})={\displaystyle _0^1}[dy]{\displaystyle \frac{[d^2l_T]}{16\pi ^2}\stackrel{~}{K}(x_i,k_{T_i};y_i,l_{T_i};M^2)\mathrm{\Psi }(y_i,l_{T_i})}`$ (11) where $`M`$ is the mass of the projectile hadron. The transverse mass of an individual parton is defined by $`\widehat{m_i}^2=k_{T_i}^2+m_i^2`$, where $`k_{T_i}`$ is the transverse momentum of the $`i^{\mathrm{th}}`$ parton in the $`n`$-particle Fock state, $`|q_1,\mathrm{},q_i,\mathrm{},q_n`$. The momentum fraction of the $`i^{\mathrm{th}}`$ parton in the Fock state is $`x_i`$, $`[dy]=\mathrm{\Pi }_{i=1}^ndy_i\delta (1\mathrm{\Sigma }_{i=1}^ny_i)`$ is a longitudinal momentum conserving metric and $`[d^2l_T]=\mathrm{\Pi }_{i=1}^nd^2l_{T_i}\delta ^2(\mathrm{\Sigma }_{i=1}^n\stackrel{}{l}_{T_i})`$. The interaction kernel is $`\stackrel{~}{K}`$. The simplest way to create final state hadron distributions from a specific Fock state wave function is now described. The vertex function on the right hand side of Eq. (11) is assumed to be slowly varying with momentum. The operator on the left hand side of the equation is then evaluated at the average transverse momentum of each parton, $`k_{T_i}^2`$, with the constraint $`\mathrm{\Sigma }_i^n\stackrel{}{k}_{T_i}=0`$. With these assumptions, the transverse mass of each parton is fixed and the vertex function becomes constant. The probability distribution is then proportional to the square of the wave function which is now inversely proportional to the off-shell parameter $`\mathrm{\Delta }=M^2\mathrm{\Sigma }_{i=1}^n\widehat{m}_i^2/x_i`$ where $`\widehat{m}_i^2`$ is the average transverse mass squared of the $`i^{\mathrm{th}}`$ parton. After longitudinal momentum conservation is specified by $`\delta (1\mathrm{\Sigma }_{i=1}^nx_i)`$, the probability distribution becomes $`{\displaystyle \frac{d^nP_n(x_1,\mathrm{},x_n)}{\mathrm{\Pi }_{i=1}^ndx_n}}=N_n\delta (1\mathrm{\Sigma }_{i=1}^nx_i)\mathrm{\Delta }^2`$ (12) where $`N_n`$ is the normalization constant for an $`n`$-particle distribution. The probability distributions as a function of $`x`$ for any final state hadron can be generated by integrating Eq. (12) including final state coalescence constraints. The characteristic shape of the longitudinal momentum distribution of the final state hadron can now be obtained up to an overall normalization constant. The important feature of this model is that final state particles are not “produced” in a collision, as such, but are rather “intrinsic” to the projectile’s Fock state and are liberated after a soft interaction with the target. This intrinsic source of final state particles acts as a perturbation to the dominant parton fusion mechanism. However, unlike parton fusion, it incorporates flavor correlations between the initial and final states. This mechanism will dominate the total cross section in the limit $`x_F1`$ since $`x_Fx`$ when the final-state hadron evolves directly from the projectile wave function. In this paper, we assume that the model developed for heavy quark hadroproduction at higher twist can be applied to gluino production in the proton wave function. Final-state $`R`$-hadron production from I$`\stackrel{~}{\mathrm{G}}`$ states is described in the remainder of this section along with its relationship to IC production. The characteristic shapes of the intrinsic distributions in the proton were generated for the gluino alone and for the $`R^+(uud\stackrel{~}{g})`$ and $`S^0(uds\stackrel{~}{g})`$, and the $`R^0(g\stackrel{~}{g})`$. In all cases, the “minimal Fock state” was used to generate the final state coalescence. This emphasizes the most leading final states. The gluino can fragment into a $`R`$-hadron, just as in pQCD production. In this uncorrelated case , the hadron $`x_F`$ distribution is $`{\displaystyle \frac{dP_{i\stackrel{~}{g}}^{kF}}{dx_H}}=N_k{\displaystyle \mathrm{\Pi }_{j=1}^k𝑑x_j𝑑z\delta (1\underset{i=1}{\overset{k}{}}x_i)\frac{D_{H/\stackrel{~}{g}}(z)}{z}\delta (x_Hzx_{\stackrel{~}{g}})\mathrm{\Delta }^2},`$ (13) where $`k`$ indicates the order of the Fock state containing the intrinsic gluinos (i.e. the $`x_{\stackrel{~}{g}}`$’s are included among the $`x_i`$). Gluinos are produced in pairs because other supersymmetric vertices involving squarks and photinos are highly suppressed due to their much greater masses. The minimal proton Fock state with a gluino pair then has five particles, $`|uud\stackrel{~}{g}\stackrel{~}{g}`$. Fragmentation of other, higher, Fock states will have a smaller production probability and produce gluinos with lower average momentum. For consistency with the coalescence production described below, we include fragmentation of six and seven particle Fock states with $`R^0`$ and $`S^0`$ production respectively. $`R`$-hadron production by coalescence is specific to each hadron. The intrinsic gluino Fock states are fragile and can easily collapse into a new hadronic state through a soft interaction with the target, as is the case for IC states. The coalescence function is assumed to be a delta function. The momentum fraction of the of the final state hadron is the sum of the momentum fractions of the of the $`R`$-hadron valence constituents from the proton wave function. The three $`R`$-hadrons we consider are all calculated from only the minimal Fock state required for their production by coalescence. Thus, only the most leading configuration is used. As in the fragmentation case in Eq. (13), including higher Fock components does not significantly increase the total rate because the other Fock state probabilities are smaller and also does not enhance the yield at large $`x_F`$ because the average $`x_F`$ of coalescence is reduced relative to that from the minimal Fock state. The five-particle Fock state $`|uud\stackrel{~}{g}\stackrel{~}{g}`$ produces the most leading $`R`$-hadron, the $`R^+`$, because the $`R^+`$ is generated from four of the five constituents of the Fock state. $`{\displaystyle \frac{dP_{i\stackrel{~}{g}}^{5C}}{dx_{R^+}}}=N_5P_C^5{\displaystyle \mathrm{\Pi }_{j=1}^5𝑑x_j\delta (1\underset{i=1}{\overset{5}{}}x_i)\delta (x_{R^+}x_ux_ux_dx_{\stackrel{~}{g}})\mathrm{\Delta }^2}.`$ (14) Here, $`P_C^5`$ is a factor incorporating the coalescence probability given the five-constituent Fock state. Note that in this case, the $`R^+`$ $`x_F`$ distribution is proportional to the gluino distribution in Eq. (13), obtained by setting $`D_{H/\stackrel{~}{g}}(z)=\delta (1z)`$, with $`k=5`$ evaluated at $`1x_F`$. The $`R^0`$ is generated from a six-particle Fock state, $`|uudg\stackrel{~}{g}\stackrel{~}{g}`$. Unlike the gluinos, single gluons can be included in the higher-twist Fock state since one gluon can couple to two quarks in the Fock state. The six-particle state is the most leading state for $`R^0`$ production. The coalescence of $`R^0`$ hadrons is described by $`{\displaystyle \frac{dP_{i\stackrel{~}{g}}^{6C}}{dx_{R^0}}}=N_6P_C^6{\displaystyle \mathrm{\Pi }_{j=1}^6𝑑x_j\delta (1\underset{i=1}{\overset{6}{}}x_i)\delta (x_{R_0}x_gx_{\stackrel{~}{g}})\mathrm{\Delta }^2}.`$ (15) The last $`R`$-hadron we consider is the $`S^0`$ which, since it contains an $`s`$ quark, must be produced from a seven-particle Fock state, $`|uuds\overline{s}\stackrel{~}{g}\stackrel{~}{g}`$. The $`S^0`$ will have a harder $`x_F`$ distribution than the $`R^0`$ even though the average momentum fraction of each constituent in the seven-particle state is smaller than those of the six-particle state. This harder $`x_F`$ distribution is due to the greater number of $`S^0`$ constituents, four, rather than the two $`R^0`$ constituents. In this case, $`{\displaystyle \frac{dP_{i\stackrel{~}{g}}^{7C}}{dx_{S^0}}}=N_7P_C^7{\displaystyle \mathrm{\Pi }_{j=1}^7𝑑x_j\delta (1\underset{i=1}{\overset{7}{}}x_i)\delta (x_{S_0}x_ux_dx_sx_{\stackrel{~}{g}})\mathrm{\Delta }^2}.`$ (16) In what follows, the coalescence probabilities $`P_C^{5,6,7}`$ appearing in Eqs. (14), (15),(16) are taken to be unity. That is, it is assumed that the gluinos will always coalesce. Figure 2 shows (using arbitrary normalization) the characteristic $`x`$ dependence of the probability distributions in Eqs. (13)-(16) with $`m_{\stackrel{~}{g}}=1.2`$ GeV. The single gluino distribution is calculated using $`k=5`$ and $`D_{H/\stackrel{~}{g}}(z)=\delta (1z)`$ in Eq. (13). $`R`$-hadrons produced by uncorrelated fragmentation have the softest $`x_F`$ distributions, $`x_{\stackrel{~}{g}}=0.24`$ when $`k=5`$. Contributions from progressively higher single gluino Fock states have smaller relative probabilities, as we discuss below, and a decreased $`x_{\stackrel{~}{g}}`$, which would eventually build up a gluino sea in the proton. The distributions from coalescence are all forward of the single gluino distribution. As expected, since the $`R^+`$ takes all three of the proton valence quarks, it is the most leading $`R`$-hadron with $`x_{R^+}=0.76`$. The distributions for the other final state particles, the $`S_0`$ and the $`R_0`$, are softer with $`x_{S^0}=0.56`$ and $`x_{R^0}=0.35`$ respectively. We have shown the results with the lowest gluino mass we consider. Increasing the mass increases the average $`x_F`$ of the gluino distribution of uncorrelated fragmentation, Eq. (13), but leaves the average $`x_F`$ of the $`R`$ hadrons unchanged in the mass range we consider. The intrinsic gluino production cross section for $`R`$-hadrons, from an $`n`$-particle Fock state is written by analogy with the IC cross section $`\sigma _{i\stackrel{~}{g}}^n(pp)=G_CP_{i\stackrel{~}{g}}^n\alpha _s^4(m_{\stackrel{~}{g}\stackrel{~}{g}})\sigma _{pp}^{\mathrm{in}}{\displaystyle \frac{\mu ^2}{4\widehat{m}_{\stackrel{~}{g}}^2}},`$ (17) where $`G_C`$ is a color factor. The inelastic $`pp`$ cross section is $`35`$ mb at 800 GeV. The ratio $`\mu ^2/4\widehat{m}_{\stackrel{~}{g}}^2`$ sets the scale at which the higher and leading twist contributions are comparable. We use $`\mu ^20.2`$ GeV<sup>2</sup>, consistent with attributing the diffractive fraction of the total $`J/\psi `$ production cross section to IC . There is a factor of $`\alpha _s^4`$ because the intrinsic state couples to two of the projectile valence quarks. The higher-twist contribution then contains two more powers of $`\alpha _s`$ than the leading-twist contribution. This factor is included in the cross section rather than in the probability distributions as done previously to more explicitly show the effect of this dependence on the cross section when the mass of the intrinsic state is changed. Since the intrinsic charm cross section is $`\sigma _{ic}^n(pp)=P_{ic}^n\alpha _s^4(m_{c\overline{c}})\sigma _{pp}^{\mathrm{in}}{\displaystyle \frac{\mu ^2}{4\widehat{m}_c^2}},`$ (18) the two cross sections are related by $`{\displaystyle \frac{\sigma _{i\stackrel{~}{g}}^n(pp)}{\sigma _{ic}^n(pp)}}={\displaystyle \frac{G_CP_{i\stackrel{~}{g}}^n}{P_{ic}^n}}{\displaystyle \frac{\widehat{m}_c^2}{\widehat{m}_{\stackrel{~}{g}}^2}}{\displaystyle \frac{\alpha _s^4(m_{\stackrel{~}{g}\stackrel{~}{g}})}{\alpha _s^4(m_{c\overline{c}})}}`$ (19) The relative color factor between intrinsic gluinos and intrinsic charm, represented by $`G_C`$, may enhance the I$`\stackrel{~}{\mathrm{G}}`$ contribution over that of IC because of the color octet nature of the gluino. However, in this work, to isolate mass effects, we assume the color factors for I$`\stackrel{~}{\mathrm{G}}`$ are the same as IC, setting $`G_C=1`$. Changing $`G_C`$ would effectively scale the cross section ratio in Eq. (19) by a constant factor. The overall effect of changing $`G_C`$ is small relative to the leading-twist cross section unless $`G_C`$ is very large. The intrinsic charm mass is used as the scale from which to approximately evolve the intrinsic gluino cross section as previously done for intrinsic beauty . Note that when $`G_C=1`$, if $`\widehat{m}_{\stackrel{~}{g}}=\widehat{m}_c`$, the I$`\stackrel{~}{\mathrm{G}}`$ and IC cross sections are the same. The I$`\stackrel{~}{\mathrm{G}}`$ cross sections are normalized by scaling $`P_{i\stackrel{~}{g}}`$ in proportion to $`P_{ic}`$, as described below. A limit of $`P_{ic}^5=0.31`$% was placed on the intrinsic charm probability in the five-particle state $`|uudc\overline{c}`$ by charm structure function data . The higher Fock state probabilities were obtained from an estimate of double $`J/\mathrm{\Psi }`$ production , resulting in $`P_{icc}^74.4\%P_{ic}^5`$ . Mass scaling was used to obtain the mixed intrinsic charm probabilities, $`P_{iqc}^7(\widehat{m}_c/\widehat{m}_q)^2P_{icc}^7`$ . To obtain the $`n`$-particle gluino Fock state probabilities, $`P_{i\stackrel{~}{g}}^n`$, we assume that the same relationships hold for the gluino states. The five-particle gluino state then scales as $`P_{i\stackrel{~}{g}}^5={\displaystyle \frac{\widehat{m}_c^2}{\widehat{m}_{\stackrel{~}{g}}^2}}P_{ic}^5.`$ (20) Assuming $`P_{i\stackrel{~}{g}\stackrel{~}{g}}^7=4.4\%P_{i\stackrel{~}{g}}^7`$, the seven-particle Fock state probabilities are $`P_{iq\stackrel{~}{g}}^7={\displaystyle \frac{\widehat{m}_c^2}{\widehat{m}_q^2}}P_{i\stackrel{~}{g}\stackrel{~}{g}}^7.`$ (21) Thus, if $`\widehat{m}_{\stackrel{~}{g}}=\widehat{m}_c`$, $`P_{ic}^5=P_{i\stackrel{~}{g}}^5`$ and $`P_{icc}^7=P_{i\stackrel{~}{g}\stackrel{~}{g}}^7`$. For simplicity, the probability for the mixed gluon-gluino proton six-particle Fock state was set equal to the seven-particle mixed probability with $`\widehat{m}_g=\widehat{m}_q`$. The effective transverse masses used were $`\widehat{m}_q=\widehat{m}_g=0.45`$ GeV, $`\widehat{m}_s=0.71`$ GeV, and $`\widehat{m}_c=1.8`$ GeV. The transverse mass of the gluino, $`\widehat{m}_{\stackrel{~}{g}}`$, is fixed to the values of $`m_{\stackrel{~}{g}}`$ used in the leading twist calculation. ## Composite Model Predictions In this section, we calculate the total $`x_F`$ distribution of final-state $`R`$-hadrons including both leading- and higher-twist contributions. The model predictions for $`R^+`$, $`R^0`$ and $`S^0`$ production on proton and nuclear targets are then given at 800 GeV. The final state $`d\sigma /dx_F`$ distribution is the sum of the leading twist pQCD distribution and the higher twist intrinsic contributions. Since many experiments use a nuclear target, the characteristic $`A`$ dependence of each contribution is included, $`{\displaystyle \frac{d\sigma }{dx_F}}=A{\displaystyle \frac{d\sigma _{lt}}{dx_F}}+A^\beta {\displaystyle \frac{d\sigma _{i\stackrel{~}{g}}}{dx_F}}.`$ (22) The first term is the leading twist term whereas the second term is the higher twist I$`\stackrel{~}{\mathrm{G}}`$ contribution. Leading twist necessarily involves single parton interactions between the target and the projectile and thus cannot account for collective nuclear effects. Thus, the leading twist cross section scales linearly with the number of nucleons in the target modulo nuclear shadowing effects. The nuclear dependence of $`J/\psi `$ production in $`pA`$ interactions shows that if the nuclear dependence is parameterized by $`A^\alpha `$, $`\alpha 2/3`$ as $`x_F1`$ . The emergence of this surface effect at large $`x_F`$ is consistent with spectators in the projectile coupling to soft gluons from the front face of the target rather than the volume. The NA3 collaboration extracted the $`A`$ dependence of $`J/\psi `$ production at large $`x_F`$ and obtained $`\beta =0.71`$ in Eq. (22) . We use the same value of $`\beta `$ for charm production since the available data on the charm $`A`$ dependence leads us to expect a similar $`A`$ dependence for charm and $`J/\psi `$ production at large $`x_F`$. The intrinsic gluino contribution to $`R`$-hadron production includes contributions from both hadronization of single gluinos by uncorrelated fragmentation, Eq. (13), and coalescence into final-state $`R`$-hadrons, described in Eqs. (14)-(16). That is, $`{\displaystyle \frac{dP_{i\stackrel{~}{g}}^n}{dx_F}}=\xi _1{\displaystyle \frac{dP_{i\stackrel{~}{g}}^{nF}}{dx_F}}+\xi _2{\displaystyle \frac{dP_{i\stackrel{~}{g}}^{nC}}{dx_F}}`$ (23) where $`P_{i\stackrel{~}{g}}^{nF}`$ and $`P_{i\stackrel{~}{g}}^{nC}`$ are the I$`\stackrel{~}{\mathrm{G}}`$ contributions from fragmentation and coalescence respectively. The parameters $`\xi _1`$ and $`\xi _2`$ allow adjustment of the relative gluino fragmentation and coalescence contributions. We used single gluino fragmentation from the same Fock state as the coalesced hadron. That is, for $`R^+`$, $`k=5`$ in Eq. (13), while $`k=6`$ for $`R^0`$ and $`k=7`$ for $`S^0`$. We fix $`\xi _1=\xi _2=0.5`$ for simplicity. For a more realistic accounting of all possible contributions to Eq. (23) for charm production, see Ref. for relative charm hadron production probabilities in the proton. The respective fragmentation and coalescence probability distributions in Eq. (23) are converted to cross sections using Eq. (17) and added to the leading twist cross section as in Eq. (22). We calculate $`R`$-hadron production at 800 GeV in $`pp`$, $`p`$Be, and $`p`$Cu interactions with $`m_{\stackrel{~}{g}}=\widehat{m}_{\stackrel{~}{g}}=1.2`$, $`1.5`$, $`3.5`$, and $`5.0`$ GeV. Delta function fragmentation was used for single intrinsic gluino production by uncorrelated fragmentation and for leading twist hadronization. That is, we take $`D_{H/\stackrel{~}{g}}(z)=\delta (1z)`$ in Eqs. (1) and (13). Figure 3 shows the normalized $`R`$-hadron $`x_F`$ distributions calculated according to Eq. (17) in $`pp`$ interactions with $`m_{\stackrel{~}{g}}=1.2`$ GeV. The difference in the yields as $`x_F0`$ is due to the difference in probability for the five, six, and seven particle Fock states. The $`R^0`$ and $`S^0`$ cross sections are similar at low $`x_F`$ because we have assumed $`P_{ig\stackrel{~}{g}}^6=P_{iq\stackrel{~}{g}}^7`$, as described in the previous section. However, the shapes are different at low $`x_F`$ because the probability distribution for uncorrelated fragmentation has a smaller average $`x_F`$ when $`k=7`$ in Eq. (13). The $`R^+`$ has the largest cross section of the three hadrons. Its distribution is symmetric around $`x_F=0.5`$ because the fragmentation yield and the $`R^+`$ yield from coalescence are symmetric in the five particle Fock state. The $`S^0`$ yield increases near $`x_F0.25`$ due to the forward peak of the $`S^0`$ coalescence distribution seen in Fig. 2. The yield at low $`x_F`$ is relatively reduced because the fragmentation calculation with $`k=7`$ is narrower so that the two peaks are effectively separated in Fig. 2. Since the fragmentation peak for $`k=6`$ and the $`R^0`$ coalescence distribution lie close together, they blend into a broad peak for the $`R^0`$ $`x_F`$ distribution. Figures 45, and 6 show the predicted $`R^+`$, $`S^0`$, and $`R^0`$ $`x_F`$ distributions per nucleon in $`pp`$, $`p`$Be, and $`p`$Cu interactions at 800 GeV calculated according to Eq. (22). Each figure includes all four gluino masses. As $`x_F0`$ the $`x_F`$ distributions of all targets are equal for a given $`m_{\stackrel{~}{g}}`$. This indicates the dominance of leading twist production at low $`x_F`$, independent of the final state. As $`x_F1`$ the higher twist terms begin to contribute. These higher twist effects are suppressed in nuclear targets because of their slower relative growth as a function of $`A`$ compared to the leading twist $`A`$ dependence. Although larger mass gluinos are more difficult to create, the relative contribution to the total cross section from higher-twist production in Eq. (22) increases with gluino mass because of the slower decrease of the intrinsic gluino contribution relative to the mass suppression of the leading twist cross section. The greater mass suppression of the leading twist cross section also influences the value of $`x_F`$ where the higher twist contribution begins to appear. Increasing the gluino mass leads to intrinsic gluino effects appearing at lower $`x_F`$. This effect is seen in Figs. 4-6. When $`m_{\stackrel{~}{g}}=1.2`$, I$`\stackrel{~}{\mathrm{G}}`$ effects become obvious near $`x_F0.5`$ while I$`\stackrel{~}{\mathrm{G}}`$ contributions begin to appear for $`x_F0.2`$ in $`R^0`$ production when $`m_{\stackrel{~}{g}}=5.0`$ GeV. Dramatic leading effects are predicted for the $`R^+`$ which, as pointed out above, shares three valence quarks with the proton in a minimal five-particle Fock state configuration. This characteristic “hardening” of the $`x_F`$ distribution for $`x_F>0.6`$ should be clear in a successful $`R^+`$ search. However, the leading effects are also present for the other particles. The $`S^0`$ is the next hardest distribution, sharing two valence quarks with the proton while the $`R^0`$ tends to be the softest, since no projectile valence quarks are shared. For a clearer comparison of the leading effects predicted for each final state $`R`$-hadron, Figs. 7-10 show the $`R^+`$, $`S^0`$, and $`R^0`$ distributions together in $`pp`$ interactions with $`m_{\stackrel{~}{g}}=1.2,1.5,3.5`$, and $`5.0`$ GeV respectively. The leading twist gluino distribution is also shown for comparison. In each case, the intrinsic contribution begins to emerge from the leading twist calculation between $`x_F0.2`$ and $`x_F0.4`$. In Fig. 8, with $`m_{\stackrel{~}{g}}=1.5`$ GeV, the predicted $`R^+`$ enhancement at $`x_F0.8`$ is about $`700`$ times larger than the leading twist prediction. At the same value of $`m_{\stackrel{~}{g}}`$ and $`x_F`$, the $`S^0`$ contribution is about $`40`$ times greater while the $`R^0`$ is just under $`6`$ times greater. When the gluino mass is increased to $`m_{\stackrel{~}{g}}=5.0`$ GeV, shown in Fig. 10, the $`R^0`$ dominates $`R`$-hadron yields for $`x_F<0.6`$. This is a consequence of the increased $`x_F`$ for single gluino fragmentation at the larger mass. Although the cross sections are small at $`m_{\stackrel{~}{g}}=5.0`$ GeV since the gluino mass is comparable to the bottom mass, the predicted enhancements over the leading-twist baseline are quite large: $`2.5\times 10^3`$ for the $`R^+`$, $`1.6\times 10^3`$ for the $`S^0`$, and $`281`$ for the $`R^0`$. The enhancements are in fact larger than those with smaller gluino masses due to the greater mass suppression of the leading twist cross section. ## Conclusions The light gluino window opens the possibility of non-trivial higher twist gluino contributions to the proton wave function. In analogy to charm hadroproduction, intrinsic gluino Fock components contribute to final state $`R`$-hadron formation, enhancing gluino production over leading twist parton fusion in the forward $`x_F`$ region. In this work, we have studied a “maximally leading” scenario for final state $`R`$-hadrons in $`pp`$ and $`pA`$ interactions at $`800`$ GeV. Our model predicts that the contributions of higher-twist intrinsic states lead to strong flavor correlations between initial and final states for $`x_F>0.6`$. The large intrinsic gluino enhancements at high $`x_F`$ over the leading-twist predictions imply that this region of phase space could be especially appropriate for $`R`$-hadron searches in the light gluino scenario. For $`m_{\stackrel{~}{g}}`$ in the $`15\mathrm{GeV}`$ range, a mass region where substantial evidence for the analogous intrinsic heavy quark states exists and for which our computational techniques should be most reliable, the enhancements are very significant (factors of several hundred to several thousand being common). The magnitudes we predict for these enhancements may even be conservative since the increased color factor associated with intrinsic gluinos compared to intrinsic charm has been neglected. Acknowledgements We would like to thank S.J. Brodsky for discussions.
warning/0002/nucl-th0002023.html
ar5iv
text
# Matter induced 𝜌-𝛿 mixing : a source of dileptons ## Abstract We study the possibility of $`\rho \delta `$ mixing via N-N excitations in dense nuclear matter. This mixing induces a peak in the dilepton spectra at an invariant mass equal to that of the $`\delta `$. We calculate the cross section for dilepton production through the mixing process and we compare its size with that of $`\pi \pi `$ annihilation. In-medium masses and mixing angles are also calculated. PACS numbers: 25.75.-q, 25.75Dw, 24.10Cn Heavy ion physics has recently seen a considerable effort being devoted to the study of the properties of hadrons in a hot and/or dense nuclear medium. Those activities were stimulated in part by the suggestion that in the nuclear medium, the vector meson masses would drop from their values in free space and that this could be interpreted as a precursor phenomenon of chiral symmetry restoration. Several attempts have been made to highlight and understand the in-medium behaviour of vector mesons, both in theory and experiment . In this respect electromagnetic signals constitute valuable probes, especially lepton pairs. This owes to the fact that the leptons couple to hadrons via vector mesons and therefore hadronic processes involving $`e^+e^{}`$ in the final channel are expected to reveal their properties in the dilepton spectra. Furthermore, the $`e^+e^{}`$ pairs suffer minimum final state interactions and are thus likely to bring information to the detectors essentially unscathed. Several experiments have measured, or are planning to measure, the lepton pairs produced in nucleus-nucleus collisions. They have been carried out by the DLS at LBL , and by HELIOS and CERES at CERN. Two new initiatives that will focus on electromagnetic probes will be PHENIX at RHIC and HADES at GSI . The density-dependent characteristics of vector mesons can also be highlighted through experiments performed at TJNAF . The last two projects will involve measurements performed in environments where the possible modifications from vacuum properties will mostly be density-driven. It is with those in mind that we have performed the theoretical estimates about which we report in this paper. While several theoretical studies have sought to investigate the in-medium properties of the vector mesons (mainly the $`\rho `$), their possible mixing with other mesons has only started to receive attention in the context of dense baryonic matter. An exception is the case of $`\rho `$-$`\omega `$ . This specific mixing can be omitted when dealing with symmetric nuclear matter, as we will here. The popularity of the $`\rho `$ meson resides in the fact that in nuclear collisions a substantial contribution to the dilepton spectra comes from $`\pi `$-$`\pi `$ annihilation which proceeds through $`\rho `$ as an intermediate state. This fact can also be stated as the dilepton spectrum sampling the in-medium vector meson spectral function . We explore here the possibility of $`\rho `$-$`\delta `$ (or $`a_0`$ as listed in ) mixing via nucleon(n)-nucleon(n) excitations in nuclear matter. Such a mixing, in effect, is similar to the known $`\omega `$-$`\sigma `$ mixing . This is a pure density-dependent effect and is forbidden in free space on account of Lorentz symmetry. We will show that such a mixing opens up a new channel for the dilepton productions and induces an additional peak in the $`\varphi `$ mass region. The interaction Lagrangian we will use can be written as $`_{int}=g_\sigma \overline{\psi }\varphi _\sigma \psi +g_\delta \overline{\psi }\varphi _{\delta ,a}\tau ^a\psi +g_{\omega NN}\overline{\psi }\gamma _\mu \psi \omega ^\mu +g_\rho [\overline{\psi }\gamma _\mu \tau ^\alpha \psi +{\displaystyle \frac{\kappa _\rho }{2m_n}}\overline{\psi }\sigma _{\mu \nu }\tau ^\alpha ^\nu ]\rho _\alpha ^\mu ,`$ (1) where $`\psi `$, $`\varphi _\sigma `$, $`\varphi _\delta `$, $`\rho `$ and $`\omega `$ correspond to nucleon, $`\sigma `$, $`\delta `$ , $`\rho `$ and $`\omega `$ fields, and $`\tau _a`$ is a Pauli matrix. The values used for the coupling parameters are obtained from Ref. . The polarization vector through which the $`\delta `$ couples to $`\rho `$ via the n-n loop is given by $`\mathrm{\Pi }_\mu (q_0,|\stackrel{}{q}|)`$ $`=`$ $`2ig_\delta g_\rho {\displaystyle \frac{d^4k}{(2\pi )^4}\text{Tr}[G(k)\mathrm{\Gamma }_\mu G(k+q)]}.`$ (2) where 2 is an isospin factor and the vertex for $`\rho `$-nn coupling is: $`\mathrm{\Gamma }_\mu =\gamma _\mu {\displaystyle \frac{\kappa _\rho }{2m_n}}\sigma _{\mu \nu }q^\nu .`$ (3) $`G(k)`$ is the in-medium nucleon propagator given by $`G(k_0,|\stackrel{}{k}|)=G_F(k)+G_D(k_0,\stackrel{}{k})`$ (4) with $`G_F(k)={\displaystyle \frac{(k/+m_n^{})}{k^2m_n^2+iϵ}},`$ (5) and $`G_D(k_0,|k|)=(k/+m_n^{}){\displaystyle \frac{i\pi }{E_k^{}}}\delta (k_0E_k^{})\theta (k_F|𝐤|),`$ (6) where $`E_k^{}=\sqrt{k^2+m_n^2}`$. The second term ($`G_D`$) deletes on-mass shell propagation of nucleons having momenta below the Fermi momentum $`k_F`$. In Eq. (4), the subscripts $`F`$ and $`D`$ refer to the free and density-dependent part of the propagator. In the subsequent equations $`m_n^{}`$ denotes the effective nucleon mass evaluated at the mean field level . With the evaluation of the trace and after a little algebra Eq. (2) could be cast into a suggestive form: $`\mathrm{\Pi }_\mu (q_0,|q|)={\displaystyle \frac{g_\rho g_\delta }{\pi ^3}}2q^2(2m_n^{}{\displaystyle \frac{\kappa q^2}{2m_n}}){\displaystyle _0^{k_F}}{\displaystyle \frac{d^3k}{E^{}(k)}}{\displaystyle \frac{k_\mu \frac{q_\mu }{q^2}(kq)}{q^44(kq)^2}}.`$ (7) This immediately leads to two conclusions. First, it respects the current conservation condition, viz. $`q^\mu \mathrm{\Pi }_\mu =0=\mathrm{\Pi }_\nu q^\nu `$. Secondly, there are only two components which would survive after the integration over azimuthal angle. In fact this guarantees that it is only the longitudinal component of the $`\rho `$ meson which couples to the scalar meson while the transverse mode remains unaltered. Furthermore, current conservation implies that out of the two non-zero components of $`\mathrm{\Pi }_\mu `$, only one is independent. It should be noted here that the tensor interaction, as evident from Eq. (7), inhibits the mixing. In presence of mixing the combined meson propagator might be written in a matrix form where the dressed propagator would no longer be diagonal: $`𝒟=𝒟^0+𝒟^0\mathrm{\Pi }𝒟.`$ (8) It is to be noted that the free propagator is diagonal and has the form $$𝒟^0=\left(\begin{array}{cc}D_{\mu \nu }^0& 0\\ 0& \mathrm{\Delta }_0\end{array}\right).$$ (9) In Eq. (9) the noninteracting propagator for the $`\delta `$ and $`\rho `$ are given respectively by $`\mathrm{\Delta }_0(q)`$ $`=`$ $`{\displaystyle \frac{1}{q^2m_\delta ^2+iϵ}},`$ (10) $`D_{\mu \nu }^0(q)`$ $`=`$ $`{\displaystyle \frac{g_{\mu \nu }+\frac{q_\mu q_\nu }{q^2}}{q^2m_\rho ^2+iϵ}},`$ (11) The mixing is characterised by the polarization matrix which contains non-diaginal elements $$\mathrm{\Pi }=\left(\begin{array}{cc}\mathrm{\Pi }_{\mu \nu }^\rho (q)& \mathrm{\Pi }_\nu (q)\\ \mathrm{\Pi }_\mu (q)& \mathrm{\Pi }^\delta (q)\end{array}\right).$$ (12) In the above expression, $`\mathrm{\Pi }^\delta `$ and $`\mathrm{\Pi }_{\mu \nu }^\rho `$ refer to the diagonal self-energies of the $`\delta `$ and $`\rho `$ meson induced by the n-n polarization: $`\mathrm{\Pi }^\delta (q_0,|\stackrel{}{q}|)`$ $`=`$ $`2ig_\delta ^2{\displaystyle \frac{d^4k}{(2\pi )^4}\text{Tr}[G(k)G(k+q)]}`$ (13) $`\mathrm{\Pi }_{\mu \nu }^\rho (q_0,|\stackrel{}{q}|)`$ $`=`$ $`2ig_\rho ^2{\displaystyle \frac{d^4k}{(2\pi )^4}\text{Tr}[G(k)\mathrm{\Gamma }_\mu G(k+q)\stackrel{~}{\mathrm{\Gamma }}_\nu ]}.`$ (14) It should be mentioned that, unlike mixing, Eqs. (13) and (14) also involve a free part stemming from the $`G_F(k)G_F(k)`$ combination which is divergent. This therefore needs to be regularized. The regularization condition we employ is $`^n\mathrm{\Pi }^F(q^2)/(q^2)^n|_{m_n^{}m,q^2=m_v^2}=0(n=0,1,2\mathrm{},\mathrm{})`$ . For $`\rho `$ the results may be found in Ref. which we do not present here and for $`\delta `$ the free part of the self-energy is given by : $`\mathrm{\Pi }^\delta (q^2)={\displaystyle \frac{3g_\delta ^2}{2\pi ^2}}[3(m_n^2m_n^2)4(m_n^{}m_n)m_n(m_n^2m_n^2){\displaystyle _0^1}dx\mathrm{ln}\left[{\displaystyle \frac{m_n^2x(1x)q^2}{m_n^2}}\right]`$ (15) $`{\displaystyle _0^1}𝑑x(m_n^2x(1x)q^2)\mathrm{ln}\left[{\displaystyle \frac{m_n^2x(1x)q^2}{m_n^2x(1x)q^2}}\right].`$ (16) It might be worthwhile to say here that the $`\rho `$ meson, being a vector, the collective oscillations set by its propagation through matter would have longitudinal (L) and transverse (T) components depending upon whether its spin is aligned along or perpendicular to the direction of propagation. Accordingly, with a special choice of z-axis along the direction of the momenta $`(\stackrel{}{q})`$, one can define the longitudinal and transverse polarization as $`\mathrm{\Pi }_L=\mathrm{\Pi }_{00}+\mathrm{\Pi }_{33}`$ and $`\mathrm{\Pi }_T=\mathrm{\Pi }_{11}=\mathrm{\Pi }_{22}`$ respectively . To determine the collective modes, one defines the dielectric function as : $`ϵ(q_0,|\stackrel{}{q}|)`$ $`=`$ $`det(1𝒟^0\mathrm{\Pi })=ϵ_T^2\times ϵ_{mix},`$ (17) where $`ϵ_T`$ corresponds to two identical transverse (T) modes and $`ϵ_{mix}`$ correspond to the longitudinal mode with the mixing. The latter, of course, also characterizes the mode relevant for the $`\delta `$ meson propagation. $`ϵ_T`$ $`=`$ $`1d_0\mathrm{\Pi }_T,d_0={\displaystyle \frac{1}{q^2m_\delta ^2+iϵ}}`$ (18) $`ϵ_{mix}`$ $`=`$ $`(1d_0\mathrm{\Pi }_L)(1\mathrm{\Delta }_0\mathrm{\Pi }_s){\displaystyle \frac{q^2}{|\stackrel{}{q}|^2}}\mathrm{\Delta }_0d_0(\mathrm{\Pi }_0)^2`$ (19) The zeros of the dielectric functions characterize the dispersion relation for the meson propagation. Fig. 1 shows the relevant dispersion curves with and without mixing at density $`\rho `$=2.5$`\rho _0`$. As only the L mode mixes with the scalar mode, we do not consider the T mode. The latter in fact is the same as presented in Ref. for $`\rho `$ meson. The effect of mixing on the pole masses, as evident from Fig. 1, are found to be very small. However, the mixing could be large when the mesons involved go off-shell. It should be noted that the modes with mixing move away from each other compared to what one obtains without mixing. This can be understood in terms of “level-level” repulsion driven by the off-diagonal terms of the dressed propagator . Fig. 2 shows the dependence of the invariant masses ($`M_i=\sqrt{q_{0(i)}^2|\stackrel{}{q}_i|^2}`$) ( i = $`\rho ,\delta `$ ) on nuclear densities where $`q_0`$’s are determined from the zeros of the dielectric function, Eq. (19). It is evident that the difference of the invariant masses first decreases with density reaching a minima and then again starts increasing. This behavior arises from the non-monotonic density dependence of the polarization functions. To calculate the mixing angle, one diagonalizes the mass matrix with the mixing and obtains $`\theta _{mix}={\displaystyle \frac{1}{2}}\mathrm{arctan}({\displaystyle \frac{2\mathrm{\Pi }_{mix}^{\rho \delta }}{m_\delta ^2m_\rho ^2\mathrm{\Pi }_L^\rho +\mathrm{\Pi }^\delta }})`$ (20) In Eq. (20) $`\mathrm{\Pi }_{mix}^{\rho \delta }=M_i/|\stackrel{}{q}|\mathrm{\Pi }_0`$ which increases with density. $`\mathrm{\Pi }_0`$ is the zero component of Eq. (7). Eq. (20) clearly shows that the mixing angle depends not only on the mixing amplitude ($`\mathrm{\Pi }_0`$) but also on the “energy denominator”. The latter, as seen in Fig. 2, first decreases as a function of density then again shows an increase characterizing the density dependence of the mixing angle as presented in Fig. 3. The mixing angle in Fig. 3 corresponds to $`|\stackrel{}{q}|`$=0.3 GeV/c. The momentum dependence for a density 2.5 times higher than the normal nuclear matter density is shown in the right panel of the same figure. This shows that for momenta beyond $`|\stackrel{}{q}|0.2`$ GeV/c the mixing is quite appreciable which affects the dilepton yield substantially as shown later. It should also be noted that the mixing angle vanishes at $`|\stackrel{}{q}|`$ = 0 or at $`\rho `$ = 0, as it should. The $`\rho `$-$`\delta `$ mixing opens a new channel viz $`\pi +\eta e^++e^{}`$ in dense nuclear matter through n-n excitations. The Feynman diagram of the process is depicted in Fig. 4. The cross-section for this process might be expressed in terms of the mixing amplitude ($`\mathrm{\Pi }_0`$) $`\sigma _{\pi \eta e^+e^{}}={\displaystyle \frac{4\pi \alpha ^2}{3q_z^2M}}{\displaystyle \frac{g_{\delta \pi \eta }^2}{g_\rho ^2}}{\displaystyle \frac{m_\rho ^4}{(M^2m_\rho ^2)^2+m_\rho ^2\mathrm{\Gamma }_\rho ^2(M)}}{\displaystyle \frac{1}{(M^2m_\delta ^2)^2+m_\delta ^2\mathrm{\Gamma }_\delta ^2(M)}}{\displaystyle \frac{1}{\sqrt{M^24m_\pi ^2}}}\mathrm{\Pi }_0^2.`$ (21) For the decay widths we consider the invariant mass dependence as presented below: $`\mathrm{\Gamma }_\rho (M)={\displaystyle \frac{g_{\rho \pi \pi }^2}{6\pi }}{\displaystyle \frac{(\frac{M^2}{4}m_\pi ^2)^{\frac{3}{2}},}{M^2}}`$ (22) and $`\mathrm{\Gamma }_\delta (M)={\displaystyle \frac{g_{\delta \pi \eta }^2}{16\pi }}{\displaystyle \frac{\sqrt{(M^2(m_\pi +m_\eta )^2)(M^2(m_\pi m_\eta )^2)}}{M^3}}.`$ (23) To describe the $`\pi \delta \eta `$ vertex we use $`_{\delta \pi \eta }=f_{\delta \pi \eta }{\displaystyle \frac{m_\delta ^2m_\eta ^2}{m_\pi }}\varphi _\eta \stackrel{}{\varphi }_\pi \stackrel{}{\varphi }_\delta ,`$ (24) where for later convenience we define $`g_{\pi \delta \eta }=f_{\delta \pi \eta }(m_\delta ^2m_\eta ^2)/m_\pi `$. Of course, there is an uncertainty involved with the coupling parameter $`f_{\delta \pi \eta }`$ as discussed in Refs.. This arises from the fact that $`\delta `$ (or $`a_0`$) lies close to the opening of the $`K\overline{K}`$ channel leading to a cusp-like behavior in the resonant amplitude, therefore a naive Breit-Wigner form for the decay width is inadequate. Furthermore, as mentioned before, there is also uncertainty involved with the $`\delta `$NN coupling which renders the precise extraction of $`\delta `$\- $`\pi `$-$`\eta `$ coupling even more difficult . We take a value for $`f_{\delta \eta \pi }`$=0.44 from Ref. which gives $`\mathrm{\Gamma }_{\delta \pi \eta }`$($`m_\delta `$)=$`59`$ MeV, while the experimental vacuum width of $`\delta `$ is between $`50100`$ MeV . One can notice in Fig. 5 that the process, $`\pi +\eta e^++e^{}`$, at densities higher than $`\rho _\mathrm{𝟎}`$, not only enhances the overall production of lepton pairs but also induces an additional peak near the $`\varphi `$ mass region. The contribution at the $`\delta `$ mass is comparable to that of $`\pi +\pi e^++e^{}`$ near the $`\rho `$ peak, for densities higher than $`\rho _\mathrm{𝟎}`$. Fig. 5 also shows that as the density goes even higher the dilepton yield arising out of the mixing also increases further. The cross-section increases with increasing momenta of the mesons in keeping with the mixing angle as shown in Fig. 3. We have highlighted the possibility of $`\rho `$-$`\delta `$ mixing in dense nuclear matter. We observe the appearance of an additional peak at a dilepton invariant mass that corresponds to that of the $`\delta `$. With sufficient experimental resolution, this effect could be observable. Probably not as an individual peak, because of the $`\delta `$’s vacuum width which is already not small, but more realistically as a shoulder in the $`\varphi `$ spectrum. This feature is then exclusively density-dependent. Our aim here was to establish the existence of the signal. Our calculation can, and will be improved upon: further studies are in progress to assess finite temperature effects and to self-consistently incorporate the necessary many-body machinery. For example, the characteristics of the $`\rho `$ can be modified in the nuclear medium and the in-medium behaviour of the $`\delta `$ needs to be addressed. We have verified that the inclusion of hadronic form factors does not change the conclusions we reach in this work. Detailed results will be presented elsewhere. ###### Acknowledgements. This work was supported in part by the Natural Sciences and Engineering Research Council of Canada and in part by the Fonds FCAR of the Québec Government.
warning/0002/math0002210.html
ar5iv
text
# 1 Introduction and basic definitions ## 1 Introduction and basic definitions There are many applications of the Lie (super)algebra cohomology in mathematics: characteristic classes of foliations; invariant differential operators; MacDonald-type combinatorial identities, etc. (see<sup>?</sup> for details). The use of cohomology in theoretical and mathematical physics can be illustrated by the following applications: * construction of the central extensions and deformations for Lie superalgebras; * construction of the supergravity equations for $`N`$-extended Minkowski superspaces and search for possible models for these superspaces; * study of stability for nonholonomic systems like ballbearings, gyroscopes, electro-mechanical devices, waves in plasma, etc.; * description of an analogue of the curvature tensor for nonlinear nonholonomic constraints;<sup>?</sup> * new methods for the study of integrability of dynamical systems; * construction of so-called higher order Lie algebras<sup>?</sup> which allow in turn to construct the Nambu mechanics<sup>?</sup> generalizing the ordinary Hamiltonian mechanics; * construction of possible invariant effective actions of Wess-Zumino-Witten type and the study of anomalies.<sup>?</sup> General definitions and properties of cohomology of Lie algebras and superalgebras are described in book.<sup>?</sup> Let us recall briefly some basic definitions. A Lie superalgebra is a $`\mathrm{𝖹𝖹}_2`$-graded algebra over a commutative ring $`K`$ with a unit: $$L=L_{\overline{0}}L_{\overline{1}},uL_\alpha ,vL_\beta ,\alpha ,\beta \mathrm{𝖹𝖹}_2=\{\overline{0},\overline{1}\}[u,v]L_{\alpha +\beta }$$ The elements of $`L_{\overline{0}}`$ and $`L_{\overline{1}}`$ are called even and odd, respectively. By definition, the Lie product (shortly, bracket) $`[,]`$ satisfies the following axioms $`[u,v]=(1)^{p(u)p(v)}[v,u],`$ $`skewsymmetry,`$ $`[u,[v,w]]=[[u,v],w]+(1)^{p(u)p(v)}[v,[u,w]],`$ $`\text{Jacobi identity},`$ where $`p(a)`$ is the parity of element $`aL_{p(a)}`$. We shall assume that $`K`$ is a field. If $`K`$ is a field of characteristic $`2`$ or $`3`$, extra axioms are needed: $`[u,u]=0`$ for even $`u`$ in characteristic $`2`$ and $`[v,[v,v]]=0`$ for odd $`v`$ in characteristic $`3`$. To provide connection with enveloping algebra, characteristic $`2`$ requires also the existence of a quadratic operator $`q`$ mapping odd elements of the algebra into even ones such that $`q(\alpha u)=\alpha ^2q(u),`$ $`[u,v]=q(u+v)q(u)q(v),`$ $`[u,[u,v]]=[q(u),v],`$ where $`\alpha K`$ and $`u,v`$ are odd. A module over a Lie superalgebra $`A`$ is a vector space $`M`$ (over the same field $`K`$) with a mapping $`A\times MM`$, such that $`[a_1,a_2]m=a_1(a_2m)(1)^{p(a_1)p(a_2)}a_2(a_1m)`$, where $`a_1,a_2A`$, $`mM`$. The most important (and easy for program implementation) are trivial ($`M`$ is arbitrary vector space, e.g., $`M=K`$; $`am=0`$), adjoint ($`M=A;am=[a,m]`$) and coadjoint ($`M=A^{};am=\{a,m\}`$ is coadjoint action) modules. A cochain complex is a sequence of linear spaces $`C^k`$ with linear mappings $`d^k`$ $$0C^0\stackrel{d^0}{}\mathrm{}\stackrel{d^{k2}}{}C^{k1}\stackrel{d^{k1}}{}C^k\stackrel{d^k}{}C^{k+1}\stackrel{d^{k+1}}{}\mathrm{},$$ (1) where the linear space $`C^k=C^k(A;M)`$ is a super skew-symmetric $`k`$-linear mapping $`A\times \mathrm{}\times AM`$, $`C^0=M`$ by definition. The super skew-symmetry means symmetry w.r.t. transpositions of odd adjacent elements of $`A`$ and antisymmetry for all other transpositions of adjacent elements. Elements of $`C^k`$ are called cochains. The linear mapping $`d^k`$ (or, briefly, $`d`$) is called a differential and satisfies the following property: $`d^kd^{k1}=0`$ (or $`d^2=0`$). The cochains mapped into zero by the differential are called cocycles, i.e., the space of cocycles is $$Z^k=\mathrm{Ker}d^k=\{C^k|dC^k=0\}.$$ The cochains which can be represented as differentials of other cochains are called coboundaries, i.e., the space of coboundaries is $$B^k=\mathrm{Im}d^{k1}=\{C^k|C^k=dC^{k1}\}.$$ Any coboundary is, obviously, a cocycle. The non-trivial cocycles, i.e., those which are not coboundaries, form the cohomology. In other words, the cohomology is the quotient space $$H^k(A;M)=Z^k/B^k.$$ As is seen from the initial part of cochain complex (1) the basis elements of module $`M`$ can be considered as non-trivial 0-cocycles. The explicit form of the differential for a Lie superalgebra is $`dC(e_0,\mathrm{},e_q;O_{q+1},\mathrm{},O_k)=`$ $`{\displaystyle \underset{i<j}{\overset{q}{}}}(1)^jC(e_0,\mathrm{},e_{i1},[e_i,e_j],\mathrm{},\widehat{e_j},\mathrm{},e_q;O_{q+1},\mathrm{},O_k)+`$ $`(1)^{q+1}{\displaystyle \underset{i=0}{\overset{q}{}}}{\displaystyle \underset{j=q+1}{\overset{k}{}}}C(e_0,\mathrm{},e_{i1},[e_i,O_j],\mathrm{},e_q;O_{q+1},\mathrm{},\widehat{O_j},\mathrm{},O_k)+`$ $`(1)^{i+1}{\displaystyle \underset{i=q+1}{\overset{k1}{}}}{\displaystyle \underset{j=q+2}{\overset{k}{}}}C(e_0,\mathrm{},e_q;O_{q+1},\mathrm{},O_{i1},[O_i,O_j],\mathrm{},\widehat{O_j},\mathrm{},O_k)`$ $`+{\displaystyle \underset{i=0}{\overset{q}{}}}(1)^{i+1}e_iC(e_0,\mathrm{},\widehat{e_i},\mathrm{},e_q;O_{q+1},\mathrm{},O_k)`$ $`+(1)^q{\displaystyle \underset{i=q+1}{\overset{k}{}}}O_iC(e_0,\mathrm{},e_q;O_{q+1},\mathrm{},\widehat{O_i},\mathrm{},O_k).`$ Here $`e_i`$ and $`O_i`$ are even and odd elements of the algebra, respectively, and the hat “$`\widehat{}`$” marks the omitted elements. Here are some properties and statements we use in the sequel. An algebra and a module are called graded if they can be presented as sums of homogeneous components in a way compatible with the algebra bracket and the action of the algebra on the module: $$A=_{gG}A_g,M=_{gG}M_g,[A_{g_1},A_{g_2}]A_{g_1+g_2},A_{g_1}M_{g_2}M_{g_1+g_2},$$ where $`G`$ is some abelian (semi)group. We assume $`G=\mathrm{𝖹𝖹}`$ in this paper. To avoid confusion, we use in the sequel the terms grade and degree for element of $`G`$ and number of cochain arguments, respectively. The grading in the algebra and module induces a grading on cochains and, hence, in the cohomology: $$C^{}(A;M)=_{gG}C_g^{}(A;M),H^{}(A;M)=_{gG}H_g^{}(A;M).$$ This property allows one to compute the cohomology separately for different homogeneous components; this is especially useful when the homogeneous components are finite-dimensional. If there is an element $`a_0A`$, such that eigenvectors (with the same eigenvalues for a given grade) of the operator $`a[a_0,a]`$ form a (topological) basis of algebra $`A`$, then $`H^{}(A)H_0^{}(A).`$ In other words, all the non-trivial cocycles of the cohomology in the trivial module lie in the zero grade component. The element $`a_0`$ is called an internal grading element. If eigenvectors of the operator $`ma_0m`$ form also a topological basis of module $`M`$, then the same statement holds for the cohomology in the module $`M`$: $`H^{}(A;M)H_0^{}(A;M).`$ In the case of trivial module, the exterior multiplication of cochains provides the cohomology with a structure of graded ring, i.e., if $`C^k`$ and $`C^m`$ are cocycles then $`C^{k+m}=C^kC^m`$ is also a cocycle. If algebra $`A`$ contains a central element $`Z`$, i.e., $`[Z,a]=0`$ for any $`aA`$, then cochain $`C^{k+1}(a_1,\mathrm{},a_k,Z)=C^k(a_1,\mathrm{},a_k)C^1(Z)`$ is a cocycle provided that $`C^k`$ is a cocycle: $`dC^k=0dC^{k+1}=0`$ because $`dC(a,Z)C([Z,a])=0`$ and the differential $`d`$ acts on a product of cochains as a (super)differentiation. Due to this fact any cocycle $`C^k(a_1,\mathrm{},a_k)`$ for algebras with odd center leads to an infinite set of cocycles of the form $`C^{k+m}(a_1,\mathrm{},a_k,\underset{m}{\underset{}{Z,\mathrm{},Z}}).`$ We shall encounter such situation later in our computations. There are also another multiplicative structures in the cohomology theory, but we shall not use them in this work. ## 2 Lie superalgebras of vector fields Below a list of the main Lie superalgebras of formal vector fields is given.<sup>?</sup> We consider some sets of even ($`x_i,q_i,p_i,t`$) and odd (called also Grassmann) variables ($`X_i,T`$). In many cases the vector fields can be expressed in terms of generating functions. The coordinates of vector fields and generating functions are assumed to be formal power series in the even and odd variables. Note that all the algebras depending only on the odd variables are finite-dimensional. All these algebras are graded due to a prescribed grading of the variables. There are some standard gradings for the variables: all variables $`x_i,q_i,p_i,X_i`$ have grade 1 and the separate variables $`t,T`$ have grade 2. Non-standard gradings with zero or negative grades for some odd variables are possible (and useful, as we show below) too. The divergence-free algebras are called special. The symbol $`𝐙`$ denotes the 1-dimensional center of an algebra consisting of constants in terms of generating function, i. e., $`𝐙=\mathrm{Span}(\{1\})`$. 1. General vectorial superalgebra $`𝐖(𝐧𝐦)`$ or $`\mathrm{𝐯𝐞𝐜𝐭}(𝐧𝐦)`$ Variables: $`x_1,\mathrm{},x_n;X_1,\mathrm{},X_m`$ The bracket denotes the supercommutator of vector fields of the form $`\underset{i=1}{\overset{n}{}}f_i\frac{}{x_i}+\underset{k=1}{\overset{m}{}}g_k\frac{}{X_k}`$ 2. Special vectorial superalgebra $`𝐒(𝐧𝐦)`$ or $`\mathrm{𝐬𝐯𝐞𝐜𝐭}(𝐧𝐦)`$ consists of the elements from $`𝐖(𝐧𝐦)`$ satisfying the divergence-free condition $`\underset{i=1}{\overset{n}{}}\frac{f_i}{x_i}+\underset{k=1}{\overset{m}{}}(1)^{p(g_k)}\frac{g_k}{X_k}=0`$ 3. Poisson superalgebra $`\mathrm{𝐏𝐨}(\mathrm{𝟐}𝐧𝐦)`$ Variables: $`p_1,\mathrm{},p_n,q_1,\mathrm{},q_n;X_1,\mathrm{},X_m`$ Bracket: $`\{f,g\}_{Pb}=\underset{i=1}{\overset{n}{}}\left(\frac{f}{p_i}\frac{g}{q_i}\frac{f}{q_i}\frac{g}{p_i}\right)(1)^{p(f)}\underset{k=1}{\overset{m}{}}\frac{f}{X_k}\frac{g}{X_k}`$ Sometimes it is more convenient to redenote the odd variables $`X_k`$ and set $`P_k=\frac{1}{2}(X_kX_{r+k}),Q_k=\frac{1}{2}(X_k+X_{r+k})`$ for $`kr=[m/2],`$ ($`U=X_{2r+1}`$ for $`m`$ odd) and the last sum in the bracket takes the form $`\underset{k=1}{\overset{r}{}}(\frac{f}{P_k}\frac{g}{Q_k}+\frac{f}{Q_k}\frac{g}{P_k})`$ (one should add the term $`\frac{f}{U}\frac{g}{U}`$ for $`m`$ odd). Hamiltonian superalgebra is $`𝐇(\mathrm{𝟐}𝐧𝐦)=\mathrm{𝐏𝐨}(\mathrm{𝟐}𝐧𝐦)/𝐙`$ Special Hamiltonian superalgebra $`\mathrm{𝐒𝐇}(\mathrm{𝟎}𝐦)`$ is a simple ideal of codimension one in $`𝐇(\mathrm{𝟎}𝐦)`$ 4. Contact superalgebra $`𝐊(\mathrm{𝟐}𝐧+\mathrm{𝟏}𝐦)`$ Variables: $`t,p_1,\mathrm{},p_n,q_1,\mathrm{},q_n;X_1,\mathrm{},X_m`$ Bracket: $`\{f,g\}_{Kb}=\delta (f)\frac{g}{t}\frac{f}{t}\delta (g)\{f,g\}_{Pb}`$ $`\delta (f)=2fE(f),E=\underset{i=1}{\overset{n}{}}\left(p_i\frac{}{p_i}+q_i\frac{}{q_i}\right)+\underset{k=1}{\overset{m}{}}X_k\frac{}{X_k}`$ 5. Buttin superalgebra $`𝐁(𝐧)`$ Variables: $`x_1,\mathrm{},x_n;X_1,\mathrm{},X_n`$ Bracket: $$\{f,g\}_{Bb}=\underset{i=1}{\overset{n}{}}\left(\frac{f}{x_i}\frac{g}{X_i}+(1)^{p(f)}\frac{f}{X_i}\frac{g}{x_i}\right)$$ (2) Leites superalgebra is $`\mathrm{𝐋𝐞}(𝐧)=𝐁(𝐧)/𝐙`$ 6. Special Buttin superalgebra $`\mathrm{𝐒𝐁}(𝐧)`$ is subalgebra of $`𝐁(𝐧)`$ subject to the constraint $`\mathrm{\Delta }f=0`$ for generating function, where $$\mathrm{\Delta }=\underset{i=1}{\overset{n}{}}\frac{^2}{x_iX_i}.$$ (3) Special Leites superalgebra is $`\mathrm{𝐒𝐋𝐞}(𝐧)=\mathrm{𝐒𝐁}(𝐧)/𝐙`$ 7. Odd contact superalgebra $`𝐌(𝐧)`$ Variables: $`x_1,\mathrm{},x_n;T,X_1,\mathrm{},X_n`$ Bracket: $`\{f,g\}_{Mb}=\delta (f)\frac{g}{T}+(1)^{p(f)}\frac{f}{T}\delta (g)\{f,g\}_{Bb}`$ $`\delta (f)=2fE(f),E=\underset{i=1}{\overset{n}{}}\left(x_i\frac{}{x_i}+X_i\frac{}{X_i}\right)`$ 8. Special odd contact superalgebra $`\mathrm{𝐒𝐌}(𝐧)`$ is subalgebra of $`𝐌(𝐧)`$ subject to the constraint $`(1E)\frac{f}{T}\mathrm{\Delta }f=0`$ for generating function. ## 3 Outline of algorithm and its implementation To compute the cohomology one needs to solve the equation $$dC^k=0,$$ (4) and throw away those solutions of (4) which can be expressed in the form $$C^k=dC^{k1}.$$ In the case of finite-dimensional Lie superalgebras determining equation (4) is a system of finite-dimensional homogeneous linear algebraic equations. In the case of infinite-dimensional graded Lie superalgebras, such as Lie superalgebras of vector fields with even variables, equation (4) is a system of linear homogeneous functional equations with integer arguments. Unfortunately there is no general method for solving such systems in closed form, though in a few exceptional cases such solutions are known. If the grading leads to finite-dimensional space of cochains in a given grade, one can proceed just as in the case of finite-dimensional algebra. Unfortunately the very important case of computation of cohomology in adjoint module for infinite-dimensional algebras can not be reduced to the set of finite-dimensional tasks at any choice of grading: in the case of adjoint module the cochains contain both elements of algebra and dual elements, these elements inevitably should have opposite grades. Nevertheless, there are important problems (such as the Spencer cohomology playing an essential role in the formal theory of differential equations<sup>?</sup>) requiring computation of cohomology in adjoint module with respect to finite-dimensional subalgebras of infinite-dimensional algebras. There are several packages for computing cohomology of Lie algebras and superalgebras written in Reduce <sup>?,?</sup> and Mathematica.<sup>?</sup> Some new results were obtained completely or partially with the help of these packages.$`^\text{a}`$footnotetext: $`^\text{a}`$In particular, D. Leites informed us that A. Shapovalov discovered one of the cocycles from the cohomology of special Hamiltonian superalgebra $`\mathrm{SH}(0|4)`$ with the help of the Mathematica program<sup>?</sup> written by P. Grozman. This cocycle was missed by D. Leites and D. Fuchs when they investigated this cohomology<sup>?</sup> by hand. However, abilities of these packages are restricted by rather small problems. We wrote a more advanced program<sup>?</sup> which allows us to consider more difficult and real problems. The C code of the program, of total length near 14200 lines, contains about 400 functions realizing top level algorithms, simplification of indexed objects, working with Grassmannian objects, exterior calculus, linear algebra, substitutions, list processing, input and output, etc. All operations with scalar coefficients, including input and output, are localized in 17 functions which formats do not depend on the nature of field of scalars $`K`$. Reading from the input which field should be used, the program assigns the suitable function addresses to the corresponding function pointer variables. This feature of the C language allows to carry out computations over arbitrary fields without recompiling and any loss of efficiency. Up to now we have implemented rational numbers of arbitrary precision, i.e., the field $`\mathrm{Q}`$, its complex extension $`\text{ }\mathrm{Q}[i]`$,$`^\text{b}`$footnotetext: $`^\text{b}`$Note that the fields $`\mathrm{I}\mathrm{R}`$ and $`\mathrm{C}`$ being non-constructive objects do not admit a computer implementation at all. rational functions of arbitrary parameters (for classification problems) and the fields $`\mathrm{𝖹𝖹}_p`$.$`^\text{c}`$footnotetext: $`^\text{c}`$For efficiency reasons the prime $`p`$ should not exceed $`46337`$ on 32bit and $`3037000493`$ on 64bit computers. Of course, other fields can be easily added if necessary. We represent Grassmann monomials by integer numbers using one-to-one correspondence between (binary codes of) non-negative integers and Grassmann monomials. This representation allows one efficiently to implement the operations with Grassmann monomials by means of the basic computer commands. The program performs sequentially the following steps: 1. Reading input information. 2. Constructing a basis for the algebra. The basis can be read from the input file; otherwise the program constructs it from the definition of the algebra. Non-trivial computations at this step arise only in the case of divergence-free algebras. The basis elements of such algebras should satisfy some conditions. In fact, we should construct the basis elements of the subspace given by a system of linear equations. The task is thereby reduced to some problem of linear algebra combined with shifts of indices. For example, among the divergence-free conditions for the special Buttin algebra $`\mathrm{SB}(3)`$ there are the following two equations $$ia_{ijk;XY}(k+1)a_{i1,j,k+1;YZ}=0,$$ $$ia_{ijk;XZ}+(j+1)a_{i1,j+1,k;YZ}=0.$$ Here $`a_{ijk;XY},\mathrm{}`$ are coefficients at the monomials $`x^iy^jz^kXY,\mathrm{}`$ in the generating function; $`x,y,z`$ and $`X,Y,Z`$ are even and odd variables, respectively. First of all, we have to shift indices $`j`$ and $`k`$ in the second equation to reduce the last terms of both equations to the same multiindices. Then, using some simple tricks of linear algebra, we can easily construct the corresponding basis element $$O_{ijk}^1=kx^iy^jz^{k1}XYjx^iy^{j1}z^kXZ+ix^{i1}y^jz^kYZ.$$ As a result for $`\mathrm{SB}(3)`$ we have the basis: $`(1)E^1`$ $`=`$ $`XYZ`$ $`(2)E_{ijk}^2`$ $`=`$ $`kx^iy^jz^{k1}Xix^{i1}y^jz^kZ`$ $`(3)E_{ijk}^3`$ $`=`$ $`kx^iy^jz^{k1}Yjx^iy^{j1}z^kZ`$ $`(4)O_{ijk}^1`$ $`=`$ $`kx^iy^jz^{k1}XYjx^iy^{j1}z^kXZ+ix^{i1}y^jz^kYZ`$ $`(5)O_{ijk}^2`$ $`=`$ $`x^iy^jz^k`$ 3. Constructing the commutator table for the algebra (if this table has not been read from the input file). We illustrate this step by non-zero commutators of $`\mathrm{SB}(3)`$ generated by the program: $$\begin{array}{cccc}(1)\hfill & [E_{ijk}^2,E_{lmn}^2]\hfill & =\hfill & (nilk)E_{i+l1,j+m,k+n1}^2\hfill \\ (2)\hfill & [E_{ijk}^2,E_{lmn}^3]\hfill & =\hfill & \frac{nkjmk^2+mk}{n+k1}E_{i+l,j+m1,k+n1}^2\hfill \\ & & & +\frac{n^2inlkni}{n+k1}E_{i+l1,j+m,k+n1}^3\hfill \\ (3)\hfill & [E_{ijk}^3,E_{lmn}^3]\hfill & =\hfill & (njmk)E_{i+l,j+m1,k+n1}^3\hfill \\ (4)\hfill & [E_{ijk}^2,O_{lmn}^1]\hfill & =\hfill & (nilk)O_{i+l1,j+m,k+n1}^1\hfill \\ (5)\hfill & [E_{ijk}^3,O_{lmn}^1]\hfill & =\hfill & (njmk)O_{i+l,j+m1,k+n1}^1\hfill \\ (6)\hfill & [E^1,O_{ijk}^2]\hfill & =\hfill & O_{ijk}^1\hfill \\ (7)\hfill & [E_{ijk}^2,O_{lmn}^2]\hfill & =\hfill & (nilk)O_{i+l1,j+m,k+n1}^2\hfill \\ (8)\hfill & [E_{ijk}^3,O_{lmn}^2]\hfill & =\hfill & (njmk)O_{i+l,j+m1,k+n1}^2\hfill \\ (9)\hfill & [O_{ijk}^1,O_{lmn}^2]\hfill & =\hfill & \frac{njmk}{n+k}E_{i+l,j+m1,k+n}^2+\frac{ni+lk}{n+k}E_{i+l1,j+m,k+n}^3\hfill \end{array}$$ 4. Creating the general form of expressions for coboundaries and determining equations for cocycles. 5. Transition to a particular grade in general expressions. At this step expressions for coboundaries take the form $`𝐱=\mathrm{𝐛𝐭}`$, equations for cocycles take the form $`\mathrm{𝐙𝐱}=\mathrm{𝟎}`$, where vector $`𝐱`$ corresponds to $`C^k`$, parameter vector $`𝐭`$ corresponds to $`C^{k1}`$, matrices $`𝐙,𝐛`$ correspond to the differential $`d`$. All these vector spaces are finite-dimensional for any particular grade. 6. Computing the quotient space $`H^k(A;M)=Z^k/B^k`$. Here the cocycle subspace $`Z^k`$ is given by relations $`\mathrm{𝐙𝐱}=\mathrm{𝟎}`$, and the coboundary subspace $`B^k`$ is given parametrically by $`𝐱=\mathrm{𝐛𝐭}`$. Substeps: 1. Eliminate $`𝐭`$ from $`𝐱=\mathrm{𝐛𝐭}`$ to get equations $`\mathrm{𝐁𝐱}=\mathrm{𝟎}`$ 2. Reduce both relations $`\mathrm{𝐁𝐱}=\mathrm{𝟎}`$ and $`\mathrm{𝐙𝐱}=\mathrm{𝟎}`$ to the canonical (row echelon) form by Gauss elimination. If $`\mathrm{rank}𝐁=\mathrm{rank}𝐙`$, then there is no non-trivial cocycle; otherwise go to Substep (6c). 3. Set $`\mathrm{𝐁𝐱}=𝐲`$ and substitute these relations into $`\mathrm{𝐙𝐱}=\mathrm{𝟎}`$ to get relations $`\mathrm{𝐀𝐲}=\mathrm{𝟎}`$. The parametric (non-leading) $`y^{}`$s of the last relations are non-trivial cocycles; that is, they form a basis of the cohomology. In fact, the above procedure is based on the relation for quotient spaces $$Z/B=\frac{Y/B}{Y/Z},$$ where $`Y`$ is an artificially introduced space, combining the above $`x^{}`$s and $`y^{}`$s. 7. Output the non-trivial cocycles. The program can output results in 2D ASCII, and standard for usual computer algebra systems 1D forms. The last form of output is useful for investigating the structure of cohomology ring with the help of the systems like Maple, Mathematica or Reduce. The operations in such investigations (multiplications and comparisons of cochains) are not difficult from computational point of view, but interactive abilities of the above systems are very convenient for the analysis of the cohomology ring. To split the whole task to smaller ones Steps (47) are executed separately for even and odd parts of the cochain complex. ### 3.1 Example of output file: Computation of $`H_0^5(\mathrm{SLe}(2))`$ The below output demonstrates computation of 5-cocycle in grade 0 from cohomology in trivial module for special Leites superalgebra of vector fields in superdimension $`(2|2)`$. Here are some explanations to this output. The brackets $`<\mathrm{}>`$ include a comment in input file. Some elements of input are optional. If they are omitted, the program either constructs some standard ones or asks to input them from keyboard. $`g(a)`$ is a $`\mathrm{𝖹𝖹}`$-grade of element $`a`$. $`E_{ij}`$ and $`O^1,O_{ij}^2`$ are even and odd basis elements of superalgebra. $`{}_{}{}^{}E_{ij}^{},^{}O^1,^{}O_{ij}^2`$ are dual elements to corresponding basis elements. The vertical and horizontal dots mean that we have omitted for brevity some long (sub)expressions in this illustrative example. Output of determining equations for cocycles may be useful. Sometimes (for a low degree cohomology) one can see the general solution for these equations. Almost all elements of output (excepting the resulting cocycle) can be suppressed by corresponding settings in the initiating file. $`t_i`$ are arbitrary parameters describing the space of coboundaries. This computation gives one basis element of the cohomology, but the program produces also its four equivalent forms: one can choose any of them (or their linear combination) in order to get more compact or symmetrical expression. Note that in the case of several basis elements of cohomology, these alternative forms may be linear combinations of the original ones. ``` Input file: D:\Kornyak\LieCohomology\In\test.in Input data: <* Special Leites superalgebra SLe(n) = SB(n)/Z *> Even variables: x; y. < Optional > Grading for even variables: 1; 1. < Optional > Odd variables: X; Y. < Optional > Grading for odd variables: -1; -1. < Optional > Module type: Trivial. < Coadjoint Adjoint> Special Leites superalgebra: 2. Cohomology number: 5. < Optional > Grade: 0. < Optional > ``` Even variables of vector field:$`xy`$;$`g(x)=1g(y)=1.`$ Odd variables of vector field: $`XY`$;$`g(X)=1g(Y)=1.`$ Basis elements of Lie superalgebra: $$\begin{array}{cccccc}(1)\hfill & E_{ij}\hfill & =\hfill & jx^iy^{j1}Xix^{i1}y^jY;\hfill & g(E_{ij})=i+j2;\hfill & i0,j0.\hfill \\ (2)\hfill & O^1\hfill & =\hfill & XY;\hfill & g(O^1)=2\hfill & \\ (3)\hfill & O_{ij}^2\hfill & =\hfill & x^iy^j;\hfill & g(O_{ij}^2)=i+j;\hfill & i0,j0.\hfill \end{array}$$ Non-zero commutators of Lie superalgebra: $$\begin{array}{cccc}(1)\hfill & [E_{ij},E_{kl}]\hfill & =\hfill & (likj)E_{i+k1,j+l1}\hfill \\ (2)\hfill & [E_{ij},O_{kl}^2]\hfill & =\hfill & (likj)O_{i+k1,j+l1}^2\hfill \\ (3)\hfill & [O^1,O_{ij}^2]\hfill & =\hfill & E_{ij}\hfill \end{array}$$ Expression for even coboundaries: $`dC_{\overline{0}}^4`$ $`=`$ $`\{(roqp)C(E_{ij},E_{kl},E_{mn},E_{o+q1,p+r1})+\mathrm{}`$ $`\mathrm{}+(ri+qj)C(O_{i+q1,j+r1}^2,O_{kl}^2,O_{mn}^2,O_{op}^2)\}`$ $`{}_{}{}^{}E_{ij}^{}^{}O_{kl}^2^{}O_{mn}^2^{}O_{op}^2^{}O_{qr}^2`$ Expression for odd coboundaries: $`dC_{\overline{1}}^4`$ $`=`$ $`\{(pm+on)C(E_{ij},E_{kl},E_{m+o1,n+p1},O^1)+\mathrm{}`$ $`\mathrm{}+C(E_{op},O_{ij}^2,O_{kl}^2,O_{mn}^2)\}^{}O^1^{}O^2_{ij}^{}O^2_{kl}^{}O^2_{mn}^{}O^2_{op}`$ Determining equation for even cocycles: $`dC_{\overline{0}}^5`$ $`=`$ $`\{(tq+sr)C(E_{ij},E_{kl},E_{mn},E_{op},E_{q+s1,r+t1})+\mathrm{}`$ $`\mathrm{}+C(E_{qr},O_{ij}^2,O_{kl}^2,O_{mn}^2,O_{op}^2)\}^{}O^1^{}O^2_{ij}^{}O^2_{kl}^{}O^2_{mn}^{}O^2_{op}^{}O^2_{qr}`$ Determining equation for odd cocycles: $`dC_{\overline{1}}^5`$ $`=`$ $`\{(roqp)C(E_{ij},E_{kl},E_{mn},E_{o+q1,p+r1},O^1)+\mathrm{}`$ $`\mathrm{}+(ti+sj)C(O_{i+s1,j+t1}^2,O_{kl}^2,O_{mn}^2,O_{op}^2,O_{qr}^2)\}`$ $`{}_{}{}^{}E_{ij}^{}^{}O_{kl}^2^{}O_{mn}^2^{}O_{op}^2^{}O_{qr}^2^{}O_{st}^2`$ Coboundary component expressions in grade 0: $$\begin{array}{ccc}C(E_{01},E_{02},E_{03},E_{10},E_{12})& =& 2t_13t_3\\ \mathrm{}& \mathrm{}& \mathrm{}\\ C(E_{10},O^1,O_{10}^2,O_{10}^2,O_{10}^2)& =& 0\end{array}$$ where $$\begin{array}{ccc}t_1& =& C(E_{01},E_{02},E_{03},E_{11})\\ \mathrm{}& \mathrm{}& \mathrm{}\\ t_{246}& =& C(O^1,O^1,O_{20}^2,O_{20}^2)\end{array}$$ Even cocycles in grade 0 are trivial. Coboundary component expressions in grade 0: $$\begin{array}{ccc}C(E_{01},E_{02},E_{03},E_{04},O^1)& =& 0\\ \mathrm{}& \mathrm{}& \mathrm{}\\ C(O^1,O^1,O_{10}^2,O_{10}^2,O_{20}^2)& =& 4t_{239}+2t_{245}\end{array}$$ where $$\begin{array}{ccc}t_1& =& C(E_{01},E_{02},E_{05},O^1)\\ \mathrm{}& \mathrm{}& \mathrm{}\\ t_{245}& =& C(E_{20},O^1,O_{10}^2,O_{10}^2)\end{array}$$ Odd cocycles in grade 0: $`(1)a_0^5`$ $`=`$ $`C(E_{02},E_{10},E_{11},E_{20},O_{01}^2)2C(E_{10},E_{11},O^1,O_{02}^2,O_{10}^2)`$ $`C(E_{11},E_{20},O^1,O_{01}^2,O_{01}^2)`$ $`=`$ $`C(2yX,Y,xXyY,2xY,y)2C(Y,xXyY,XY,y^2,x)`$ $`C(xXyY,2xY,XY,y,y)`$ and also: $`(1)`$ $`C(E_{01},E_{02},E_{10},E_{11},O_{20}^2)+4C(E_{10},E_{11},O^1,O_{01}^2,O_{11}^2)`$ $`2C(E_{10},E_{11},O^1,O_{02}^2,O_{10}^2)`$ $`=C(X,2yX,Y,xXyY,x^2)+4C(Y,xXyY,XY,y,xy)`$ $`2C(Y,xXyY,XY,y^2,x)=a_0^5`$ $`\mathrm{}`$ $`(4)`$ $`C(E_{01},E_{10},E_{11},E_{20},O_{02}^2)4C(E_{10},E_{11},O^1,O_{01}^2,O_{11}^2)`$ $`C(E_{11},E_{20},O^1,O_{01}^2,O_{01}^2)`$ $`=C(X,Y,xXyY,2xY,y^2)4C(Y,xXyY,XY,y,xy)`$ $`C(xXyY,2xY,XY,y,y)=a_0^5`$ ## 4 Buttin vector fields and related algebras The Poisson and Hamiltonian (super)algebras are very important algebras of vector fields. In the papers on the deformation quantization<sup>?,?</sup> it was proven that the Poisson bracket is an unique structure providing deformation of a commutative algebra of differentiable functions on a manifold to a new noncommutative but associative algebra. In paper<sup>?</sup> we present some results about the structure of cohomology rings for Poisson, Hamiltonian and related algebras obtained with the help of our program. Whereas the Poisson bracket, defined on a $`2n`$-dimensional symplectic manifold, has an old history, its counterpart called Buttin bracket (or odd Poisson bracket, or antibracket) and defined on a $`(n|n)`$-dimensional odd symplectic supermanifold$`^\text{d}`$footnotetext: $`^\text{d}`$Such manifolds possess interesting and unusual geometrical properties.<sup>?,?</sup> is a comparatively new construction. The first example of such bracket has appeared in Schouten’s paper<sup>?</sup> as an extension of the Lie bracket on vector fields to an bracket on skew-symmetric contravariant (i.e., tangent) tensor fields (multivectors). A more abstract formulation for this bracket was given by Buttin.<sup>?</sup> Since 1981 antibrackets are very popular in theoretical physics, because they play the crucial role in the Batalin-Vilkovisky (BV) covariant method for quantizing general gauge theories.<sup>?</sup> This method called the BV (or antibracket, or field-antifield) formalism$`^\text{e}`$footnotetext: $`^\text{e}`$The antibracket in BV formalism is an odd symplectic form on the (infinite-dimensional) space of fields and antifields playing the role of even and odd variables respectively, and the partial derivatives in formulas (2) and (3) should be replaced by variational derivatives and the summation by integration. being currently a most powerful procedure for quantizing gauge theories is applied also in string and topological field theories.<sup>?</sup> Therefore investigation of properties of Buttin and related algebras is a problem of interest for physics. We should stress also importance of the special subalgebras $`\mathrm{SB}()`$ and $`\mathrm{SLe}()`$, because $`\mathrm{\Delta }`$-operator (3) plays an essential role in the BV formalism: so-called master equation in the BV formalism is defined via this operator. ### 4.1 Computations Here we present some results of computations of cohomologies in the trivial module for the algebras Buttin $`\mathrm{B}(1)`$, special Buttin $`\mathrm{SB}(1)`$ and their centerless quotients $`\mathrm{Le}(1)`$ and $`\mathrm{SLe}(1)`$. We present also the results for odd contact algebra $`\mathrm{M}(1)`$ and its special subalgebra $`\mathrm{SM}(1)`$. Our computations are restricted with values for cohomology degree ($`10`$) and grade ($`10`$). We computed also some cocycles for the case $`n1`$. Here we mention only the most regular of them: $`a_n^1=C(X_1\mathrm{}X_n)`$ for the algebras $`\mathrm{SB}(n)`$ and $`\mathrm{SLe}(n)`$, this cocycle generates an infinite number of its wedged powers if $`n`$ is even, and $`a_0^2=C(X_1,x_1)=\mathrm{}=C(X_n,x_n)`$ for the algebras $`\mathrm{Le}(n)`$ and $`\mathrm{SLe}(n).`$ Any cocycle for the algebra $`\mathrm{SB}(n)`$ generates also new cocycles due to presence of odd center in this algebra. We use the following grading for the variables: $`g(x_i)=1,g(X_i)=1,g(T)=0.`$ With this grading the algebras $`\mathrm{B}(n),\mathrm{Le}(n)`$ and $`\mathrm{M}(n)`$ contain an internal grading element at any $`n`$, i. e., there is no need to compute cohomology in grades different from zero for this algebras. Unfortunately, there is no good grading providing internal grading element for the special subalgebras.$`^\text{f}`$footnotetext: $`^\text{f}`$There are grading elements for the algebras $`\mathrm{SB}(n),\mathrm{SLe}(n)`$ and $`\mathrm{SM}(n)`$ for $`n`$ even at the grading: $`g(x_i)=1,g(X_i)=1,1i\frac{n}{2};g(x_i)=1,g(X_i)=1,\frac{n}{2}<in;`$ but in this case the space of cochains in zero grade becomes infinite-dimensional. One can see also that the even generating functions correspond to the odd elements of algebra and vice versa. In particular, the element $`1`$ is an odd central element in the algebras $`\mathrm{B}(n)`$ and $`\mathrm{SB}(n).`$ In the below formulas all indices $`i,j0`$, but for the algebras $`\mathrm{Le}(1)`$ and $`\mathrm{SLe}(1)`$ the central element $`O_0=1`$ should be excluded. #### 4.1.1 $`H^k(\mathrm{B}(1))`$ and $`H^k(\mathrm{Le}(1))`$ Basis elements: $$\begin{array}{ccccc}(1)\hfill & E_i\hfill & =\hfill & x^iX;\hfill & g(E_i)=i+1;\hfill \\ (2)\hfill & O_i\hfill & =\hfill & x^i;\hfill & g(O_i)=i.\hfill \end{array}$$ Non-zero commutators: $$\begin{array}{cccc}(1)\hfill & [E_i,E_j]\hfill & =\hfill & (ij)E_{i+j1}\hfill \\ (2)\hfill & [E_i,O_j]\hfill & =\hfill & jO_{i+j1}\hfill \end{array}$$ Generating cocycles for $`H^k(\mathrm{B}(1)):`$ $`a^3`$ $`=`$ $`C(X,xX,x^2X)`$ $`b^3`$ $`=`$ $`C(X,xX,x){\displaystyle \frac{1}{2}}C(X,x^2X,1)`$ The ring $`H^{}(\mathrm{B}(1))`$ contains also all cocycles of the form $`a^3C(1)\mathrm{}C(1)`$ and $`b^3C(1)\mathrm{}C(1).`$ $`H^k(\mathrm{Le}(1))`$ contains only 3 non-trivial cocycles for $`k20:`$ the above cocycle $`a^3`$ and the cocycles $`b^3=C(X,xX,x)`$ and $`a^2=C(X,x).`$ The last cocycle describes the central extension of $`\mathrm{Le}(1)`$ to $`\mathrm{B}(1).`$ #### 4.1.2 $`H^k(\mathrm{M}(1))`$ Basis elements: $$\begin{array}{ccccc}(1)\hfill & E_i^1\hfill & =\hfill & x^iX;\hfill & g(E_i^1)=i1;\hfill \\ (2)\hfill & E_i^2\hfill & =\hfill & x^iT;\hfill & g(E_i^2)=i;\hfill \\ (3)\hfill & O_i^1\hfill & =\hfill & x^iTX;\hfill & g(O_i^1)=i1;\hfill \\ (4)\hfill & O_i^2\hfill & =\hfill & x^i;\hfill & g(O_i^2)=i.\hfill \end{array}$$ Non-zero commutators: $$\begin{array}{cccc}(1)\hfill & [E_i^1,E_j^1]\hfill & =\hfill & (ji)E_{i+j1}^1\hfill \\ (2)\hfill & [E_i^1,E_j^2]\hfill & =\hfill & (i+1)E_{i+j}^1+jE_{i+j1}^2\hfill \\ (3)\hfill & [E_i^2,E_j^2]\hfill & =\hfill & (ji)E_{i+j}^2\hfill \\ (4)\hfill & [E_i^1,O_j^1]\hfill & =\hfill & (ji)O_{i+j1}^1\hfill \\ (5)\hfill & [E_i^2,O_j^1]\hfill & =\hfill & (ji+1)O_{i+j}^1\hfill \\ (6)\hfill & [E_i^1,O_j^2]\hfill & =\hfill & jO_{i+j1}^2\hfill \\ (7)\hfill & [E_i^2,O_j^2]\hfill & =\hfill & (j2)O_{i+j}^2\hfill \\ (8)\hfill & [O_i^1,O_j^2]\hfill & =\hfill & (j+2)E_{i+j}^1+jE_{i+j1}^2\hfill \end{array}$$ We have found only one non-trivial cocycle $$\begin{array}{ccc}a^3\hfill & =\hfill & C(T,TX,x)\frac{1}{2}C(T,xTX,1)\hfill \\ & =\hfill & C(X,xX,xT)\frac{1}{2}C(T,xTX,1)\hfill \\ & =\hfill & C(X,T,xT)\frac{1}{2}C(T,xTX,1)\hfill \\ & =\hfill & C(X,xTX,x)+\frac{1}{2}C(x^2X,TX,1)+\frac{1}{2}C(xT,TX,1)\hfill \\ & =\hfill & C(xX,TX,x)\frac{1}{2}C(T,xTX,1).\hfill \end{array}$$ #### 4.1.3 $`H_g^k(\mathrm{SB}(1))`$ and $`H_g^k(\mathrm{SLe}(1))`$ Basis elements: $$\begin{array}{ccccc}(1)\hfill & E\hfill & =\hfill & X;\hfill & g(E)=1;\hfill \\ (2)\hfill & O_i\hfill & =\hfill & x^i;\hfill & g(O_i)=i.\hfill \end{array}$$ Non-zero commutators: $$\begin{array}{cccc}(1)\hfill & [E,O_i]\hfill & =\hfill & iO_{i1}\hfill \end{array}$$ The algebras $`\mathrm{SB}(1)`$ and $`\mathrm{SLe}(1)`$ contain the odd centers $`Z=\mathrm{Span}(\{1\})`$ and $`Z=\mathrm{Span}(\{x\})`$ respectively. The sets of non-trivial cocycles consist of some generating cocycles and their consequences obtained by multiplication of these cocycles by arbitrary wedged powers of $`C(1)`$ and $`C(x)`$ for $`\mathrm{SB}(1)`$ and $`\mathrm{SLe}(1)`$ respectively. Table 1 presents all generating non-trivial cocycles in the limits for cohomology degree $`k10`$ and grade $`g10`$ for algebras $`\mathrm{SB}(1)`$ and $`\mathrm{SLe}(1).`$ The presence of generating cocycle for $`\mathrm{SLe}(1)`$ is marked with * in the table. We give here the explicit expressions up to 3-cocycles for $`\mathrm{SB}(1)`$ (The cocycles for $`\mathrm{SLe}(1)`$ can be obtained by deleting terms with argument 1 from these expressions). $`a_{\text{-}1}^1`$ $`=`$ $`C(X)`$ $`a_1^3`$ $`=`$ $`C(X,1,x^2)C(X,x,x)`$ $`a_3^3`$ $`=`$ $`C(X,1,x^4)4C(X,x,x^3)+3C(X,x^2,x^2)`$ $`a_5^3`$ $`=`$ $`C(X,1,x^6)6C(X,x,x^5)+15C(X,x^2,x^4)10C(X,x^3,x^3)`$ $`a_7^3`$ $`=`$ $`C(X,1,x^8)8C(X,x,x^7)+28C(X,x^2,x^6)56C(X,x^3,x^5)`$ $`+35C(X,x^4,x^4)`$ $`a_9^3`$ $`=`$ $`C(X,1,x^{10})10C(X,x,x^9)+45C(X,x^2,x^8)120C(X,x^3,x^7)`$ $`+210C(X,x^4,x^6)126C(X,x^5,x^5)`$ #### 4.1.4 $`H_g^k(\mathrm{SM}(1))`$ Basis elements: $$\begin{array}{ccccc}(1)\hfill & E_i\hfill & =\hfill & ix^{i1}T+(i+2)x^iX;\hfill & g(E_i)=i1;\hfill \\ (2)\hfill & O^1\hfill & =\hfill & TX;\hfill & g(O^1)=1;\hfill \\ (3)\hfill & O_i^2\hfill & =\hfill & x^i;\hfill & g(O_i^2)=i.\hfill \end{array}$$ Non-zero commutators: $$\begin{array}{cccc}(1)\hfill & [E_i,E_j]\hfill & =\hfill & (2j2i)E_{i+j1}\hfill \\ (2)\hfill & [E_i,O_j^2]\hfill & =\hfill & (2j2i)O_{i+j1}^2\hfill \\ (3)\hfill & [O^1,O_i^2]\hfill & =\hfill & E_i\hfill \end{array}$$ Our computations in the limits $`k,g10`$ show that the only non-trivial cocycles take the form $`a_{k\text{-}2}^k`$ excepting the case $`k=2`$. We give below the explicit expressions for the first three these cocycles. The cocycle $`a_{\text{-}1}^1`$ generates an infinite set of cocycles $`a_{\text{-}k}^k=a_{\text{-}1}^1\mathrm{}a_{\text{-}1}^1=C(TX,\mathrm{},TX).`$ $`a_{\text{-}1}^1`$ $`=`$ $`C(TX)`$ $`a_1^3`$ $`=`$ $`C(T+xX,2xT,1)2C(TX,x,x)`$ $`=`$ $`C(2X,T+xX,x^2)2C(TX,x,x)`$ $`=`$ $`C(2X,2xT,x)+2C(TX,1,x^2)+2C(TX,x,x)`$ $`a_2^4`$ $`=`$ $`C(T+xX,2xT,1,x){\displaystyle \frac{1}{6}}C(T+xX,3x^2Tx^3X,1,1){\displaystyle \frac{4}{3}}C(TX,x,x,x)`$ $`=`$ $`C(2X,T+xX,x,x^2){\displaystyle \frac{1}{6}}C(T+xX,3x^2Tx^3X,1,1){\displaystyle \frac{4}{3}}C(TX,x,x,x)`$ $`=`$ $`{\displaystyle \frac{1}{2}}C(2X,2xT,x,x){\displaystyle \frac{1}{6}}C(T+xX,3x^2Tx^3X,1,1)+2C(TX,1,x,x^2)`$ $`+{\displaystyle \frac{2}{3}}C(TX,x,x,x)`$ ## 5 Conclusion Mathematicians regard the problem of computation of cohomology as solved if they construct the full cohomology ring. The structure of such rings may be rather complicated including many non-trivial relations between the cocycles in contrast to the examples in this paper where the rings are commutative. To get a clear idea about the structure of cohomology ring one should compute usually the cocycles up to degrees high enough. Unfortunately the computation of cohomology is a typical problem with the combinatorial explosion. Nevertheless, some results can be obtained with the help of computer having an efficient enough program. On the other hand, physicists are interested mainly in the second cohomologies describing the central extensions and deformations. Such cohomologies can be computed rather easily even for large algebras. Some essential possibilities remain for increasing the efficiency of the program and we hope to implement the corresponding improvements in future. Acknowledgment I would like to thank D. Leites for initiating this work and helpful communications. I am also grateful to V. Gerdt and O. Khudaverdian for fruitful discussions and useful advises. This work was supported in part by INTAS project No. 96-0842 and RFBR project No. 98-01-00101. References
warning/0002/quant-ph0002054.html
ar5iv
text
# Towards the elusive Efimov state of the 4He3 molecule through a new atom-optical state-selection technique ## Abstract Excited states and excitation energies of weakly bound systems, e. g. atomic few-body systems and clusters, are difficult to study experimentally. For this purpose we propose a new and very general atom-optical method which is based on inelastic diffraction from transmission gratings. The technique is applicable to the recently found helium trimer molecule <sup>4</sup>He<sub>3</sub>, allowing for the first time an investigation of the possible existence of an excited trimer state and determination of its excitation energy. This would be of fundamental importance for the famous Efimov effect. Already in 1935 Thomas discovered a surprising property of three-body systems. He considered short-range two-particle potentials which supported just one single bound state with an arbitrarily weak binding energy. He then found that for the three-body system there could exist a much more tightly bound ground state and that the binding energy could approach $`\mathrm{}`$ when the range of the two-particle potential approached zero. Thirty-five years later Efimov obtained a striking generalization of this result. He predicted that when one weakened the two-body interaction the number of three-body excited states could increase to infinity. An excited state which appears under weakening of the two-particle potential is called an Efimov state. Conversely, under strengthening of the two-particle potential an Efimov state disappears into the continuum. Intuitively one can understand the Efimov effect in a three-body system by imagining a weakly bound, and therefore spatially very extended, two-particle subsystem. This subsystem can then act on the third particle with a force whose range is given by the spatial extent of the subsystem . This range therefore increases more and more when the two-particle potential is decreased. As opposed to a short-range potential , however, a long-range potential can have infinitely many excited states. Whether the Efimov effect does occur in nature is still an open question. In nuclear physics no generally accepted examples of Efimov states have been found . For systems of neutral atoms, however, they might exist, and an excellent candidate is the helium trimer, <sup>4</sup>He<sub>3</sub>, since the dimer, <sup>4</sup>He<sub>2</sub>, is believed to have an extremely weak binding energy ($`1.3`$ mK) and no excited states. A few years ago <sup>4</sup>He<sub>2</sub> has been observed by Luo et al. and independently by Schöllkopf and Toennies , who also observed <sup>4</sup>He<sub>3</sub>. There has been a lot of theoretical work on the existence of an Efimov state in the helium trimer, with sometimes conflicting results . The overall theoretical picture presently indicates the existence of a ground state and a single excited state, denoted here by <sup>4</sup>He<sub>3</sub> and <sup>4</sup>He$`{}_{}{}^{}{}_{3}{}^{}`$, respectively. The latter is believed to be an Efimov state. Both are $`s`$ states with respective energy around $`E_0=0.1`$ K and $`E_1=2`$ mK . Experimentally, <sup>4</sup>He$`{}_{}{}^{}{}_{3}{}^{}`$ has not yet been seen . The Efimov property as well as the very existence of the excited state <sup>4</sup>He$`{}_{}{}^{}{}_{3}{}^{}`$ depend sensitively on the detailed form of the two-atom interaction, and even small retardation corrections to the potential can significantly affect the <sup>4</sup>He<sub>2</sub> binding energy and bond length . A precise knowledge of this potential is also necessary for understanding liquid helium droplets and superfluidity. Therefore conclusive experimental evidence of an excited state <sup>4</sup>He$`{}_{}{}^{}{}_{3}{}^{}`$ and determination of its binding energy would not only be a crucial step towards establishing the existence of an Efimov state but would also give important information on the He-He potential. Here we propose a new and very general method to both detect and select an excited state of <sup>4</sup>He<sub>3</sub>, or of other systems, as well as to determine its excitation energy. Excitation can be achieved by scattering a beam from a solid surface or, in our case, from many small surfaces. Taking for the latter the bars of a microfabricated transmission grating one can achieve excitation and separation at the same time, as will now be shown. The state-selective property stems from two conservation laws. If the incident molecular center of mass momentum is denoted by $`P^{}`$ and the final one by $`P`$, energy conservation implies $$P^2/2M=P^2/2M+\mathrm{\Delta }E_{\mathrm{int}}$$ (1) where $`M`$ is the molecular mass and $`\mathrm{\Delta }E_{\mathrm{int}}`$ accounts for a possible change of the internal molecular state. The other conservation law comes from the discrete translational invariance of the grating in the 2 direction (period $`d`$, cf. Fig. 1). This implies the conservation of the initial momentum component $`P_2^{}`$ up to a reciprocal lattice vector , i. e. $$P_2=P_2^{}+n2\pi \mathrm{}/d$$ (2) where $`n=0,\pm 1,\pm 2,\mathrm{}`$. With $`P_2=P\mathrm{sin}\phi `$ (see Fig. 1) this yields a relation between the angle of incidence $`\phi ^{}`$ and the allowed angle, $`\phi _n`$, of the $`n`$-th order diffraction peak. With $`\lambda ^{}=2\pi \mathrm{}/P^{}`$ and $`\lambda =2\pi \mathrm{}/P`$ the initial and final de Broglie wave length, the relation can be written, after a short calculation, as $$\mathrm{sin}\phi _n=\frac{\lambda }{\lambda ^{}}\mathrm{sin}\phi ^{}+n\frac{\lambda }{d}.$$ (3) Eq. (3) holds for any molecule-grating interaction potential. A wave theoretical interpretation of Eq. (3) can be given as follows (see Fig. 1). First, a wave with angle of incidence $`\phi ^{}`$ is refracted in the plane $`A`$, with a change of wave length from $`\lambda ^{}`$ to $`\lambda `$. The refraction angle $`\phi _0`$ and the incidence angle $`\phi ^{}`$ are related as in Snell’s law through $`\mathrm{sin}\phi _0/\mathrm{sin}\phi ^{}=\lambda /\lambda ^{}`$. Secondly, in the plane $`B`$, the new wave of wave length $`\lambda `$ is diffracted by the slits as in classical optics . Combining this with Snell’s law gives Eq. (3) . For an elastic process $`\lambda `$ coincides with the initial $`\lambda ^{}`$ and then Eq. (3) reduces to the condition for grating diffraction of a de Broglie wave as obtained from classical optics . But if the molecule is excited by the interaction with the grating, $`\lambda `$ differs from $`\lambda ^{}`$ by a factor of $`1/\sqrt{1(E_1E_0)/E_{\mathrm{kin}}^{}}1+(E_1E_0)/2E_{\mathrm{kin}}^{}`$ where $`E_{\mathrm{kin}}^{}P^2/2M`$ is the initial kinetic energy. For <sup>4</sup>He<sub>3</sub> three different kinds of processes can occur, namely elastic scattering (<sup>4</sup>He$`{}_{3}{}^{}_{}^{4}`$He<sub>3</sub>), excitation (<sup>4</sup>He$`{}_{3}{}^{}_{}^{4}`$He$`{}_{}{}^{}{}_{3}{}^{}`$) and breakups (<sup>4</sup>He$`{}_{3}{}^{}_{}^{4}`$He$`{}_{2}{}^{}+_{}^{4}`$He or <sup>4</sup>He$`{}_{3}{}^{}_{}^{4}`$He$`+^4`$He$`+^4`$He) where the latter have diffuse scattering angles. For the elastic case ($`\lambda =\lambda ^{}`$) the diffraction term $`n\lambda /d`$ in Eq. (3) has been used previously to separate molecules of different mass . In order to separate the equal mass particles <sup>4</sup>He<sub>3</sub> and <sup>4</sup>He$`{}_{}{}^{}{}_{3}{}^{}`$ we propose here to use the first term in Eq. (3) (Snell’s law) and its dependence on the incidence angle as follows. For normal incidence of <sup>4</sup>He<sub>3</sub> the low order diffraction-peak angles of <sup>4</sup>He<sub>3</sub> and <sup>4</sup>He$`{}_{}{}^{}{}_{3}{}^{}`$ differ by micro radians only and are practically indistinguishable ($`\phi ^{}=0`$ in Eq. (3)), but by rotating the grating ($`\phi ^{}0`$) the peaks will separate. This allows the identification of <sup>4</sup>He$`{}_{}{}^{}{}_{3}{}^{}`$ and determination of $`E_1E_0`$. For example, the zeroth order <sup>4</sup>He$`{}_{}{}^{}{}_{3}{}^{}`$ diffraction angle gives $$E_1E_0=\frac{P^2}{2M}\left(1\frac{\mathrm{sin}^2\phi ^{}}{\mathrm{sin}^2\phi _0}\right).$$ (4) To quantitatively check the feasibility of our proposal we have calculated diffraction patterns of a <sup>4</sup>He<sub>3</sub> beam (<sup>4</sup>He$`{}_{3}{}^{}_{}^{4}`$He<sub>3</sub>, <sup>4</sup>He$`{}_{3}{}^{}_{}^{4}`$He$`{}_{}{}^{}{}_{3}{}^{}`$) incident under various angles on a 100 nm period silicon nitride (SiN<sub>x</sub>) transmission grating and for a typical nozzle temperature of 6 K . For this we have applied the approach of Refs. and to an incident three-body system with the trimer wave functions of Fig. 2. The wave functions have been obtained by means of the momentum space Faddeev approach and the unitary pole approximation (see e. g. Ref. ) using the Tang, Toennies, Yiu (TTY) potential and they are sufficiently accurate to yield $`E_0=0.1`$ K and $`E_1=2.1`$ mK, comparable to the results of the adiabatic hyperspherical approach in Ref. . In the calculation of the diffraction pattern in the Fraunhofer regime the trapezoidal cross section of the grating bars with a wedge angle of $`\beta =9^{}`$ (see Fig. 1) and the helium-silicon nitride van der Waals interaction potential of Ref. have been included. As can be seen in Fig. 3 a <sup>4</sup>He$`{}_{}{}^{}{}_{3}{}^{}`$ signal appears in the form of side peaks on the elastic diffraction peaks for angles of incidence $`\phi ^{}>10^{}`$. The energy difference $`E_1E_0`$ is very small compared to the initial kinetic energy and results in small angle differences between <sup>4</sup>He<sub>3</sub> and <sup>4</sup>He$`{}_{}{}^{}{}_{3}{}^{}`$ diffraction peaks. But such small angle differences are easily resolvable in present-day experiments. Fig. 4 reveals the role played by the attractive van der Waals interaction between the molecules and the grating material. While the elastic processes qualitatively follow the predictions of classical optics, the total transmitted <sup>4</sup>He$`{}_{}{}^{}{}_{3}{}^{}`$ intensity (<sup>4</sup>He$`{}_{3}{}^{}_{}^{4}`$He$`{}_{}{}^{}{}_{3}{}^{}`$) is only slowly varying over a wide range of incidence angles $`\phi ^{}`$. Therefore, the excitation of the molecules depends only weakly on the slit projection orthogonal to the incidence direction and will take place mainly at the slit boundaries. If the angle of incidence approaches the wedge angle of the grating bars (see Fig. 1) the <sup>4</sup>He$`{}_{}{}^{}{}_{3}{}^{}`$ transmission probability reaches its maximum value. In this case of grazing incidence on the surface of one side of the bars the van der Waals interaction strongly affects the excitation process and leads to a gain in the total transmitted <sup>4</sup>He$`{}_{}{}^{}{}_{3}{}^{}`$ intensity. Theory shows the excitation process to be the more efficient the larger the overlap region of the two wave functions. Fig. 3 indicates that a realization of the proposed experiment requires the measurement of intensities over five orders of magnitude. With conventional electron impact ionization mass spectrometer detectors this accuracy has been achieved so far solely for helium-atom beams because the estimated detection efficiency is only $`10^6`$. A helium-atom detector with an efficiency of about $`10^4`$ and an improved <sup>4</sup>He<sub>3</sub> source are presently under study . One of them, or both, should make it possible to measure, for example, the zeroth order excitation peaks in Fig. 3 and thus demonstrate the existence of an excited helium trimer state. The principle of the proposed method of selecting internal states and determining the energy difference of the respective internal transitions by means of transmission gratings is not restricted to the particular <sup>4</sup>He<sub>3</sub> molecule. An extension to other systems with low-energy internal transitions like other helium clusters, Rydberg atoms or even more complicated systems can be envisaged. Also, the method could provide pure beam sources of particular internal states and might be a useful spectroscopic tool for the investigation of internal atomic and molecular transitions which are not easily accessible to laser light. We would like to thank R.E. Grisenti, W. Schöllkopf and J.P. Toennies for stimulating discussions.
warning/0002/gr-qc0002087.html
ar5iv
text
# References Macroscopic Einstein equations for a system of interacting particles and their cosmological applications. A. V. Zakharov Dept.Gen.Rel.& Grav. Kremljevskaja Str. 18 Kazan Stata University 420008 Kazan, Russia tel.:(+7)(8432)315586 e-mail: Alexei.Zakharov@ksu.ru SUMMARY One of the possible applications of macroscopic Einstein equations has been considered. So, the nonsingular isotropic and uniform cosmological model is built. The cosmological consequences of this model are agree with conclusions of standard hot model of the Universe. Macroscopic Einstein equations for a system of interacting particles and their cosmological applications. A. V. Zakharov 1. Macroscopic system of Einstein equations As it is known the macroscopic Maxwell equation for continuous media can be obtained also from the microscopic Maxwell equations by ensemble averaging the latter (refer to.-). The Einstein equations, which right-hand sides contain the energy-momentum tensor of matter, are phenomenological equations. It is natural to suppose that the Einstein equations (or their generalizations) for continuous media can be also obtained from the microscopic Einstein equations, i.e., Einstein equations which right-hand side contain the sum of the energy - momentum tensors of individual particles. However, due to the nonlinearity of the Einstein equations left-hand side, the averaging of the microscopic Einstein equations is more complicated than one of the microscopic Maxwell equations (refer to. - ). In Ref. - were obtained the macroscopic Einstein equations for a system of interaction particles up to second order in the interaction constant. They have the forms: $$G_{ij}+Z_{ij}=\chi T_{ij},$$ (1) Here $`G_{ij}`$ is the Einstein’s tensor of the Riamannian space with macroscopic metric $`g_{ij}`$, $`T_{ij}`$ is the macroscopic energy - momentum tensor, $`\chi =8\pi k/c^4`$ is Einstein constant, $`k`$ is the gravitational constant, $`c`$ is the velosity of light. The Einstein equations of the gravitational field for continium media differ from the classical Einstein equations by the presence of additional tensor $`Z_{ij}`$ in the left-hand side. It caused by particle interaction. The form of this tensor is given in Ref. in the case of gravitationally interaction of particles and in Ref. in the case of electromagnetically interaction of particles. The tensor $`Z_{ij}`$ is the traceless tensors with zero divergence: $$g^{ij}Z_{ij}=0,Z_{;j}^{ij}=0.$$ (2) where the semicolon denote a covariant derivative in a spase with the macroscopic metric $`g_{ij}`$. In the case of gravitationally interacting particles the tensor $`Z_{ij}`$ has the form (Ref. ): $$Z_{ij}=\phi _{ij;k}^k+\mu _{ij},$$ (3) where $$\phi _{ij}^k=\underset{bc}{}\frac{\chi ^3m_b^3m_c^3c^9}{8(2\pi )^3}\frac{d^3p^{}}{p^0\sqrt{(g)}}\frac{d^3p^{\prime \prime }}{p^{\prime \prime 0}\sqrt{(g)}}[\frac{1}{2}g^{fk}u_i^{\prime \prime }u_j^{\prime \prime }+$$ $$+u^k(u^{}u^{\prime \prime })(\delta _j^fu_i^{\prime \prime }+\delta _i^fu_j^{\prime \prime })\left]\right((u^{}u^{\prime \prime })^2\frac{1}{2})$$ $$K_{f\alpha }(u^{},u^{\prime \prime })\left(F_c^{\prime \prime }\frac{F_b^{}}{p_\alpha ^{}}F_b^{}\frac{F_c^{\prime \prime }}{p_\alpha ^{\prime \prime }}\right),$$ (4) $$\mu _{ij}=\underset{bc}{}\frac{\chi ^3m_b^3m_c^3c^9}{8(2\pi )^3}\frac{d^3p^{}}{p^0\sqrt{(g)}}\frac{d^3p^{\prime \prime }}{p^{\prime \prime 0}\sqrt{(g)}}\{[(z^2+\frac{1}{2})(u_i^{\prime \prime }u_j^{\prime \prime }+u_i^{}u_j^{})+$$ $$+(z^2\frac{1}{2})g_{ij}2z(u_i^{}u_j^{\prime \prime }+u_i^{\prime \prime }u_j^{})]g^{qr}2(z^2\frac{1}{2})\delta _i^q\delta _j^r\}$$ $$F_c^{\prime \prime }\frac{}{p_f^{}}\left\{F_b^{}\left[(z^2\frac{1}{2})\delta _f^m+(z^2+\frac{1}{2})u_f^{}u^m2zu_f^{\prime \prime }u^m\right]\right\}J_{rqm}(u^{},u^{\prime \prime }).$$ (5) Here $`p^i`$ is the momentum of particles of species $`b`$, $`p^{\prime \prime i}`$ is the momentum of particles of spesies $`c`$ $`u^i=p^i/m_bc`$, $`u^{\prime \prime i}=p^{\prime \prime i}/m_cc`$, $`F_b^{}`$ is the one - particle distribution function of particles of spesies $`b`$, $`F_c^{\prime \prime }`$ is the one - particle distribution function of spesies $`c`$, $`z=u^iu_i^{\prime \prime }`$, $`m_b`$ and $`m_c`$ are the mass of particles of spesies $`b`$ and $`c`$ respectively, $$\frac{d^3p^{}}{p^0\sqrt{(g)}}and\frac{d^3p^{\prime \prime }}{p^{\prime \prime 0}\sqrt{(g)}}$$ are the invariant volume elements in tree - dimensional momentum space of particles spesies ”b” and ”c” respectively. The greek index $`\alpha `$ in (4) takes the values 1,2 and 3 only (the spartial index). The derivative with respect to $`p_f^{}`$ in (5) should be calculated as all four components of momentum are independent. The dependence of $`p_0^{}`$ on $`p_\alpha ^{}`$ is taken into account after differentiation with respect $`p_f^{}`$ is completed. The tensors $`K_{ij}(u^{},u^{\prime \prime })`$ and $`J_{ijk}(u^{},u^{\prime \prime })`$ have the form: $$K_{ij}(u^{},u^{\prime \prime })=\frac{4\pi ^2}{k_{min}^2[(u^{}u^{\prime \prime })^21]^{3/2}}\{[(u^{}u^{\prime \prime })^21]g_{ij}$$ $$u_i^{}u_j^{}u_i^{\prime \prime }u_j^{\prime \prime }+(u^{}u^{\prime \prime })(u_i^{}u_j^{\prime \prime }+u_i^{\prime \prime }u_j^{})\}$$ (6) $$J_{ijk}(u^{},u^{\prime \prime })=A[(g_{ij}u_k^{}+g_{ik}u_j^{}+g_{jk}u_i^{})z(g_{ij}u_k^{\prime \prime }+g_{ik}u_j^{\prime \prime }+g_{jk}u_i^{\prime \prime })$$ $$(u_i^{}u_j^{\prime \prime }u_k^{\prime \prime }+u_i^{\prime \prime }u_j^{}u_k^{\prime \prime }+u_i^{\prime \prime }u_j^{\prime \prime }u_k^{})+3zu_i^{\prime \prime }u_j^{\prime \prime }u_k^{\prime \prime }]+$$ $$+C[u_i^{}u_j^{}u_k^{}z(u_i^{}u_j^{}u_k^{\prime \prime }+u_i^{}u_j^{\prime \prime }u_k^{}+u_i^{\prime \prime }u_j^{}u_k^{})+$$ $$+z^2(u_i^{}u_j^{\prime \prime }u_k^{\prime \prime }+u_i^{\prime \prime }u_j^{}u_k^{\prime \prime }+u_i^{\prime \prime }u_j^{\prime \prime }u_k^{})z^3u_i^{\prime \prime }u_j^{\prime \prime }u_k^{\prime \prime }],$$ (7) where $$A=\frac{2\pi \sqrt{2}}{k_{min}}\left[\frac{(z2)}{(z1)^2(z+1)^{1/2}}+\frac{(2z1)}{(z+1)(z1)^{5/2}}\mathrm{ln}\left(z+\sqrt{z^21}\right)\right],$$ (8) $$C=\frac{2\pi \sqrt{2}}{k_{min}}\left[\frac{(z6)}{(z1)^3(z+1)^{3/2}}+\frac{(6z1)}{(z+1)^2(z1)^{7/2}}\mathrm{ln}\left(z+\sqrt{z^21}\right)\right].$$ (9) In these expressions $$k_{min}=\frac{1}{r_{max}},$$ where $`r_{max}`$ is the size of the correlation region in the case of gravitationally interacting particles. In Refs. , there are estimates for $`r_{max}`$ in the case where the average metric $`g_{ij}`$ is the metric of isotropic cosmological model. 2. Applications of the theory. Nonsingular isotropic and uniform cosmological models in macroscopic theory of gravity. Let’s consider received equations for the ambience, base at local thermodynamic balance. In this cause the distribution function of each sorts of particles has the type ((refer to. ): $$F_a(p_\alpha )=A_aexp[c(v_ip^i)/(k_BT)].$$ (10) Here $`v_i`$ is the macroscopic four-velocities of ambience, $`k_B`$ is the Boltzman’s constant, $`T`$ is the temperature, $`A_a`$ is the normalizing constant. In this case tensor $`\phi _{ij}^k`$ becomes zero, but tensor $`\mu _{ij}`$ has the form: $$\mu _{ij}=\chi ϵ_1\left(\frac{4}{3}v_iv_j\frac{1}{3}g_{ij}\right)$$ (11) This is correct both for gravitational and electromagnetic interactions. If move $`\mu _{ij}`$ from left-hand side to the right-hand side , the macroscopic equation effectively change into usual Einstein equations with the additional tensor of energy-momentum in the right-hand side. Moreover the ”additional tensor of energy-momentum” has the form of energy-momentum tensor of ideal liquid with the state equation $`P_1=ϵ_1/3`$, but with negative ”density of energy” $`(ϵ_1)`$ and negative ”pressure” $`(P_1)`$ . Via $`ϵ_1`$ and $`P_1`$ their absolute values are marked. In the case of gravitational interactions in nonrelativistic limit when $`mc^2>>k_BT`$ , the absolute value of this ”negative density of energy” is $$ϵ_1=\underset{ab}{}\frac{4k^2r_{max}}{k_BT}m_a^2m_b^2N_aN_b,$$ (12) where $`N_a`$ and $`N_b`$ is the number of particles of species $`a`$ and $`b`$ recpectively densities, $`m_a`$ and $`m_b`$ are their masses. It is proportional to the square of gravitational constant and square of number of particles density. Consequently the additional terms can play significant role in ambiences of sufficiently high density. Such density is possible on the early stages of the Universe evolution. So, naturally, one can use the first exhibits of the received equations in the theories of the early stages of the Universe evolution. Thereby there is a real possibility of manifestation an interaction on early stages of the Universe evolution. Let’s turn to the building of uniform and isotropic cosmological models within the framework of macroscopic theory of gravity. Let’s write the metricses of these models in a form: $$(ds)^2=a^2(\eta )\left((d\eta )^2(dr)^2\varphi ^2(r)((d\theta )^2+\mathrm{sin}^2(\theta )(d\phi )^2)\right).$$ (13) Here $`\varphi (r)=r`$ , $`\varphi (r)=\mathrm{sin}r`$, $`\varphi (r)=\mathrm{sinh}r`$ for the flat, close and open models respectively. The system of Einstein equations for these metricses is reduced to two equations (refer to ,for instance, ): $$a^2+\xi a^2=\frac{8\pi k}{3c^4}a^4\stackrel{~}{ϵ},$$ (14) $$\frac{d\stackrel{~}{ϵ}}{\stackrel{~}{ϵ}+\stackrel{~}{P}}=3\frac{da}{a}.$$ (15) Here stroke under $`a`$ marks derivative on time variable $`\eta `$; $`\stackrel{~}{ϵ}`$ and $`\stackrel{~}{P}`$ is a density of energy and pressure of matter , $`\xi =0,+1,1`$ for the flat, closed and open models respectively. The macroscopic Einstein equations are distinguish from (14) - (15) by following. In the first, $`\stackrel{~}{ϵ}`$ in the right-hand side of (14) is replased by the difference of usual energy density $`ϵ`$ and absolute value ”density of energy” $`ϵ_1`$, stipulated by the particles interaction of ambience. At the present moment of the Universe evolution the expression (12) for $`ϵ_1`$ is correct. In the second, the equation (15) is changed by two similar: $$\frac{dϵ}{ϵ+P}=3\frac{da}{a},$$ (16) $$\frac{dϵ_1}{ϵ_1+P_1}=3\frac{da}{a}.$$ (17) Here $`P`$ is a usual pressure of matter, $`P_1=ϵ_1/3`$ is the ”pressure”, stipulated by the interaction of particles. Equations (16) and (17) follows from the fact, that divergency of the usual energy - momentum tensor and divergency of the ”additional energy - momentum tensor”, stipulated by the interaction, equals to zero. From (12) it follow that in present the value $`ϵ_1`$ and the relict radiation energy density are of the same order of value (see to below). So, if we take into account a contribution $`ϵ_1`$ to $`\stackrel{~}{ϵ}`$, we must take into account a contribution of the relict radiation energy density to $`\stackrel{~}{ϵ}`$ density. So, let’s put in the right part of (14) $$\stackrel{~}{ϵ}=ϵ_m\overline{ϵ}.$$ (18) Here $`ϵ_m`$ is the density of energy of material disregarding density of energy of relict radiation, but under $`\overline{ϵ}`$ we shall understand a difference between the absolute value ”density of energy”, stipulated by the interaction, and density of energy of relict radiation. This is suitable, since equations of condition for relict radiation and for ”energy - momentum tensor”, stipulated by the interaction alike: pressure is one third from density of energy. So that the equation (16) and (17) coused from the following $`\overline{ϵ}`$ dependency on the scale factor: $$\overline{ϵ}=\frac{3c^4a_1^2}{8\pi ka^4}.$$ (19) Here $`a_1=const`$. For density of energy of rest material from (16) we have: $$ϵ_m=\frac{3c^4a_0}{4\pi ka^3},$$ (20) where $`a_0=const`$ . We substitute (18), (19) and (20) into (14). As a result we obtain the equation: $$a^2+\xi a^2=2a_0aa_1^2$$ (21) The solution of this equation for the scale factor $`a(\eta )`$ one can write in the form: $$a=a_m+\frac{1}{2}a_0\eta ^2$$ (22) for flat ($`\xi =0`$) models ( here $`a_m=a_1^2/2a_0`$), $$a=a_m+(a_0a_m)(1\mathrm{cos}\eta )$$ (23) for closed ($`\xi =+1`$) models (here $`a_m=a_0\sqrt{(a_0^2a_1^2)}`$), and $$a=a_m+(a_0+a_m)(\mathrm{cosh}\eta 1)$$ (24) for open ($`\xi =1`$) models (here $`a_m=\sqrt{(a_0^2+a_1^2)}a_0`$ ). As it is seen, all three models are nonsingular: under $`\eta =0`$ the scale factor $`a(\eta )`$ does not apply to the zero as in standard cosmologycal models, but takes a minimum value $`a_m`$. Let us calculate a value of density $`\rho _m`$ of nonrelativistic matter at the moment, when scale factor takes a minimum value $`a_m`$: $$\rho _m=\rho _0\frac{a^3(\eta _0)}{a_m^3}.$$ (25) Here $`\rho _0`$ is the value of usual matter density at present moment. This moment corresponds a time coordinate value $`\eta _0`$. Substituting (22) — (24) into (25) we obtain the expression for $`\rho _m`$ via $`\rho _0`$, $`\eta _0`$ and ratio $`a_0/a_m`$. Let us write down the expressions for $`\mathrm{\Omega }=(ϵ_m\overline{ϵ})_{\eta =\eta _0}/(c^2\rho )`$, and $`\lambda =(\overline{ϵ}/ϵ_m)_{\eta =\eta _0}`$ via $`\eta _0`$ and ratio $`a_0/a_m`$. Here $`\rho =3H_0^2/8\pi k`$ is the critical density at the present moment, $`\mathrm{\Omega }`$ is the densities parameter, $`\lambda `$ is a ratio of ”energy density”, stipulated by the interaction absolute value, and density of relict radiation energy difference to the density of usual matter energy at the present moment. From (14), (18) - (20) we obtain $$\mathrm{\Omega }=\left(1+\xi \frac{a^2}{a^2}\right)_{\eta =\eta _0}\lambda =\frac{a_1^2}{2a_0a(\eta _0)}$$ (26) Substituting (22) — (24) into (26) we obtain the expression for $`\mathrm{\Omega }`$ and $`\lambda `$ via, $`\eta _0`$ and ratio $`a_0/a_m`$. Expressing from received correlations $`\eta _0`$ and $`a_0/a_m`$ via $`\mathrm{\Omega }`$ and $`\lambda `$ and then substituting then received expressions for $`\eta _0`$ and $`a_0/a_m`$ in (25) , we obtain the following result: $$\rho _m=\frac{\rho _0}{\lambda ^3}\left(\frac{1}{2}+\sqrt{\frac{1}{4}\lambda (1\lambda )\frac{(\mathrm{\Omega }1)}{\mathrm{\Omega }}}\right)^3$$ (27) Result (27) is correct for flat $`(\mathrm{\Omega }=1)`$ , open $`(\mathrm{\Omega }<1)`$, and closed $`(\mathrm{\Omega }>1)`$ models. The received models can be realized, if absolute value of ”energy density” (12), stipulated by the interaction, exceeds at present moment the energy density of relict radiation. The ”energy density” (12) , stipulated by interaction, is created basically galaxies accumulations with the mass $`m10^{15}10^{16}M_{}`$. This ”energy density” is estimated as $$\frac{12k^2t_0\rho _0^2m}{<v>}.$$ When getting this correlation we put $`3k_BT=m<v>^2`$, where $`<v>`$ is an average chaotic velocity of galaxies accumulations, $`N=\rho _0/m`$ , but parameter $`r_{max}`$ is estimated as $`<v>t_0`$, where $`t_0`$ is the modern age of the Universe (refer to. , ). Substituting into the obtained ratio the numerical values of fundamental constants, cosmological time, average density of matter in the Universe at present moment (refer to ,for instance, ), we obtain under $`<v>/c=10^6`$, $`m10^{15}M_{}`$: $$ϵ_1710^{13}erg/cm^3$$ It is approximately the value of relict radiation energy density we have at present moment (refer to.). Consequently, realization of such situation is possible, when absolute value of ”energy density”, stipulated by interaction, exeeds the density of relict radiation energy. And it is the case considered in the paper when $`\lambda >0`$. On initial stage of the Universe evolution the density $`\rho _m\rho _0/\lambda ^3`$ is sufficiently great , so usual scenario of the Universe hot models preserves. For a moment when scale factor takes a minimum value and the density of matter is maximum let us introduse the folloing. If at this moment a temperature does not exceed the temperatures of nucleon - antinucleons couple annihilation, the density of nonrelativistic mater can be evaluated as $$\frac{m_p10^8\sigma T^4}{k_BT},$$ where $`\sigma `$ is the constant of Stefan - Boltzman, $`m_p`$ is the mass of proton. Comparing this expression with $`\rho _0/\lambda ^3`$ one can make a conclusion, that minimum value of scale factor will be reach on leptones stage of the Universe evolution when parameter $`\lambda `$ is within $`10^{10}`$ \- $`10^{12}`$, and the temperature will be about $`10^{10}K^o`$ \- $`10^{12}K^o`$. 3. Domain of the theory application. When concluding the macroscopic Einstein equations we suggest the distant collision to be the prevalent. It is correct if $$\gamma =\frac{E_p}{E_k}1,$$ where $`E_p`$ is the potential energy of two particles interaction, $`E_k`$ is the kinetic energy of particles. For the system of electromagnetical interacting particles in a condition of local thermodynamic balance we can put $`E_kkT`$, $`E_pe^2n^{1/3}`$. So the condition of validity takes the form $$\gamma =\frac{e^2n^{1/3}}{kT}1.$$ (28) Here $`e,n,T,k`$ is the charge, density, temperature and Boltzman constant correspondingly. Let us verify (28) in constructing model under $`t=t_m`$, when the Universe reaches the maximum density. The majority of electromagnetical interacting particles on this stage compose the electron-positron pairs. It’s density is compared on the order of value with the one of relict photons and can be estimated as $`\sigma T^4/kT=\sigma T^3/k`$, where $`\sigma `$ is the Stephan-Boltzman constant. Then for parameter $`\gamma `$ we have estimation: $$\gamma =\frac{e^2\sigma ^{1/3}}{k^{4/3}}10^21.$$ Consequently the application condition of the macroscopic Einstein equations in this model is not violated.
warning/0002/cond-mat0002289.html
ar5iv
text
# 1 Introduction ## 1 Introduction An olfactory system must solve the problems of odor detection, recognition, and segmentation. Segmentation is necessary because the odor environment often contains two or more odor objects. The system must be able to identify these objects separately and signal their presence to higher brain areas. An odor object is defined as an odor entity (which, e.g., the smell of a cat, often contains fixed proportions of multiple types of odor molecules) that enters the environment independently of other odors. Therefore, two odor objects usually do not enter the environment at the same time although they often stay together in the environment afterwards. In cases when different odors do enter the environment together as a mixture, human subjects have great difficulty identifying the components. In this paper we present a model which performs odor segmentation temporally. First one odor object is detected and recognized, then the system adapts to this specific odor so a subsequent one can be detected and recognized. The odor specificity of this adaptation is the key feature of the operation of the system. This specificity can not be achieved with simple single-unit fatigue mechanisms because of the highly distributed nature of odor pattern representations in the olfactory system: fatiguing neurons that respond to one odor would strongly reduce their response to another one, thereby distorting the pattern evoked by the second odor. In our model a delayed inhibitory feedback signal is directed to the input units in such a way as to cancel out the current input, leaving the system free to respond to new odors as if the first one were not there. Our model is not intended as a faithful representation of any particular animal olfactory system. Present anatomical and physiological knowledge do not permit such detailed modelling. Rather, our focus is on the computations performed by different groups of neurons, based on general biological findings, which we review briefly here. In animals, different odor molecules produce different, distributed activity patterns across the neurons of the olfactory nerve, which provide the input to the olfactory bulb . We do not model this part of the processing. We will simply represent different odors as different but overlapping input patterns to the bulb. They are temporally modulated by the animal’s sniff cycle (typically 2-4 sniffs per second), i.e., active only during and immediately after inhalation. The main cell types of the mammalian bulb are the excitatory mitral cells and the inhibitory granule cells. The mitral cells receive the odor input and excite the granule cells, which in turn inhibit them. The outputs of the bulb are carried to the olfactory cortex by the mitral cell axons. In vertebrate animals, odors evoke oscillatory bulbar activity in the 35-90 Hz range, which may be detected by surface EEG electrodes . Different parts of the bulb have the same dominant frequency but different amplitudes and phases , and this oscillation pattern is odor-specific . These oscillations are an intrinsic property of the bulb, persisting after central connections to the bulb are cut . (In invertebrates, oscillations exist without odor input but are modulated by odors .) Upon repeated presentation of a conditioned odor stimulus, the bulbar oscillations weaken markedly . Since olfactory receptor neurons exhibit only limited adaptation , this adaptation must originate either in the bulb or in cortical structures. The pyriform or primary olfactory cortex receives bulbar outputs via the lateral olfactory tract, which distibutes outputs from each mitral cell over many cortical locations . The signals are conveyed to the (excitatory) pyramidal cells of the cortex, both directly and via feedforward inhibitory cells in the cortex. The pyramidal cells send axon collaterals to each other and to feedback interneurons which, in turn, inhibit them. There is thus excitatory-inhibitory circuitry as in the bulb, and oscillatory responses to odors are observed in the cortex, too. However, the cortex differs from the bulb in the much greater spatial range of the excitatory connections and in the presence (or at least the greater extent) of excitatory-to-excitatory connections. This anatomical structure has led a number of workers to model the olfactory cortex as an associative memory for odors . Furthermore, the oscillations in the cortex require input from the bulb; they do not occur spontaneously. Cortical output, including the feedback to the bulb, is from pyramidal cells . Some of the feedback is direct, while some of it is via other cortical centers, notably the entorhinal cortex. Most central feedback to the bulb is to the granule cells . Cooling the cortex, presumably reducing or removing the central feedback, enhances the bulbar responses . The basic features outlined here constrain our model: we employ coupled excitatory and inhibitory populations in both bulb and cortex, we wire the network so that odors evoke oscillations in the bulb, which drive similar cortical oscillations through excitatory and inhibitory connections, and we send the central feedback to reduce the bulbar responses. We will neglect many known features of animal olfactory systems, such as (to name a few) the patterns of connectivity from receptors to mitral cells, the dendrodentritic character of the mitral-granule synapses, and the differing spatial range of connectivity in bulb and cortex. Indeed, the model has no geometry: “location” and “distance” have no meaning here. We retain only the basic elements necessary to illustrate the basic operation of the system, in order not to obscure the functions we focus on (detection, recognition, and segmentation). We will also hypothesize features of the system, in particular the nature of the feedback signal from the cortex to the bulb, for which there is not yet experimental evidence (though they are not incompatible with present knowledge). These assumptions will be necessary in order to make an explicit model that can be tested computationally. Some details of its implementation are neither crucial to the computational function of the model nor intended as explicit neurophysiological predictions. However, the basic framework of the model and the dynamical properties we find for it are subject to experimental test. In the next section we present the model: its equations of motion and how it detects, recognizes, and segments odor inputs. The following section demonstrates how it works in simulations. In the final section we discuss the implications of our work, including potential experimental tests for this and related models and how they can help us understand the functioning of the olfactory system. ## 2 The model Our model consists of two modules, a bulb and a cortex, with feedforward and feedback connections between them. It is depicted schematically in Fig. 1. The bulb encodes odor inputs as patterns of oscillation. These form the input to the cortex, which acts as an associative memory for odor objects, recognizing them by resonant oscillation in an odor-specific pattern when the input from the bulb matches one of its stored odor memories. The odor-specific resonant cortical activity pattern is transformed to a feedback signal to the bulb, which approximately cancels the effect of the odor input that elicited it. The system is then able to respond to a newly arrived odor superposed on the previous one. In this way it segments temporally the different odor objects in the environment. The model is a rate-model network , in which we associate each unit with a local population of cells that share common synaptic input (mitral cells for the excitatory units, granule cells for the inhibitory ones). The output (activation) of a unit, representing the average firing rate within the corresponding population, is modeled as a sigmoidal function of the net synaptic input. In both the bulb and cortex modules, the units occur in pairs, one unit excitatory and the other inhibitory. In the absence of coupling between different such pairs, they form independent damped local oscillators. The coupling between pairs leads to oscillation patterns across the modules, with specific amplitudes for the individual local oscillators and specific phase relations between them. The odor input makes these oscillatory patterns different from odor to odor; thus, these patterns form the internal encoding of the odors. The sizes of the local populations corresponding to our formal units are different for excitatory and inhibitory units; this difference is accounted for in the model by appropriate scaling of the synaptic strengths. We turn now to the explicit mathematical description of the two modules and the coupling between them. ### 2.1 the bulb The bulb model we employ was introduced by Li and Hopfield (1989) . For completeness, we review it here. The odor input to (mitral) unit $`i`$ is denoted $`I_i`$. (We will also use a vector notation, in which the entire input pattern is denoted $`𝐈`$.) Adding to this the synaptic input from granule cells within the bulb, we obtain an equation of motion $$\dot{x}_i=\alpha x_i\underset{j}{}H_{ij}^0g_y(y_j)+I_i$$ (1) for the (local population average) membrane potential $`x_i`$. Here $`\alpha ^1`$ is the membrane time constant, $`g_y()`$ is the (sigmoidal) activation function of the granule units, $`y_j`$ is the membrane potential for granule unit $`j`$, and $`H_{ij}^0`$ is the inhibitory synaptic strength from granule unit $`j`$ to mitral unit $`i`$. All the $`H_{ij}^0`$ are non-negative; the inhibitory nature of the granule cells is represented by the negative sign in the second term on the right-hand side. The signal the bulb sends on to the cortex is carried by the mitral unit outputs $`g_x(x_i)`$ (with $`g_x(.)`$ their activation function). We have not included mitral-mitral connections here, because the experimental evidence for them is weak, but including them would not change the properties of the model qualitatively. For the inhibitory units, representing local populations of granule cells, we have, similarly to (1), $$\dot{y}_i=\alpha y_j+\underset{j}{}W_{ij}^0g_x(x_j)+I_i^\mathrm{c},$$ (2) with the mitral-to-granule synaptic matrix $`W_{ij}^0`$. Here the external input $`I_i^\mathrm{c}`$ represents the centrifugal input (from the cortex), which contains the feedback signal that implements the odor-specific adaptation. In describing the response to an initial odor, it can be neglected or taken as a constant background input. To see how this network produces oscillatory excitation patterns in response to an odor, start by taking the input $`𝐈`$ to be static. It determines a fixed point $`\overline{x}_i`$ and $`\overline{y}_i`$ of the equations, i.e., $`\dot{x}_i=\dot{y}_i=0`$ at $`\overline{x}_i`$ and $`\overline{y}_i`$, which increase with odor input $`𝐈`$. Taking the deviation from this fixed point as $`x_i\overline{x}_ix_i`$ and $`y_i\overline{y}_iy_i`$, linearizing and eliminating the $`y_i`$ leads to $$\ddot{x}_i+2\alpha \dot{x}_i+\alpha ^2x_i+\underset{j}{}A_{ij}x_j,=0,$$ (3) where the matrix $`𝖠=\mathrm{𝖧𝖶}`$, with $`H_{ij}=H_{ij}^0g_y^{}(\overline{y}_j)`$ and $`W_{ij}=W_{ij}^0g_x^{}(\overline{x}_j)`$. This equation describes a coupled oscillator system, with a coupling matrix $`𝖠`$. Denoting the eigenvectors and eigenvalues of this matrix by $`𝐗_k`$ and $`\lambda _k`$, respectively, (3) has solutions $`𝐱=_kc_k𝐗_k\mathrm{exp}[\alpha t\pm \mathrm{i}(\sqrt{\lambda _k}t+\varphi _k)]`$, with $`c_k`$ and $`\varphi _k`$ the amplitude and phase of the $`k^{th}`$ mode. If $`𝖠`$ is not symmetric (the general case), $`\lambda _k`$ is complex, and the mode has oscillation frequencies $`\omega _k\mathrm{Re}(\sqrt{\lambda _k})`$. The amplitude for mode $`k`$ will grow exponentially (in this linearized theory) if $`\pm \mathrm{Im}(\sqrt{\lambda }_k)>\alpha `$. Its growth will be limited by nonlinearities, and it will reach a steady-state saturation value. In this spontaneously oscillating state, the fastest-growing mode, call it the $`1^{st}`$ mode, will dominate the output. The whole bulb will oscillate with a single frequency $`\omega _1`$ (plus its higher harmonics), and the oscillation amplitudes and phases may be approximated by the complex vector $`𝐗_1`$. Thus, the olfactory bulb encodes the olfactory input via the following steps: (1) the odor input $`𝐈`$ determines the fixed point $`(\overline{𝐱},\overline{𝐲})`$, which in turn (2) determines the matrix $`𝖠`$, which then (3) determines whether the bulb will give spontanous oscillatory outputs and, if it does, the oscillation amplitude and phase pattern $`𝐗_1`$ and its frequency $`\omega _1`$. Strictly speaking, this description only applies to very small oscillations. For larger amplitudes, nonlinearities make the problem in general intractable. However, we will suppose that the present analysis gives a decent qualitative guide to the dynamics, checking this assumption later with simulations of the network. In this model, oscillations arise strictly as a consequence of the asymmetry of the matrix $`𝖠`$. The model could be generalized to add intrinsic single-unit oscillatory properties, and these might enhance the network oscillations. However, a model with symmetric $`𝖠`$ and intrinsic oscillatory properties only at the single-unit level can not support oscillation patterns in which the phase varies across the units in the network. We will return to this point in the Discussion section. A word about timescales: The odor input varies on the timescale of a sniff: 300-500 ms. The oscillations are in the 40 Hz range, so the input $`𝐈`$ hardly changes at all over a few oscillation periods ($`25`$ ms). We may therefore treat periods of several oscillations as if the input were static within them, and do the above analysis separately for each such period (adiabatic approximation). With inhalation, the increasing input $`𝐈`$ pushes the fixed point membrane potentials $`\overline{x}_i`$ from their initial values (where the activation function $`g(x)`$ has low gain) through a range of increasing gains, thereby increasing the size of some of the elements of the matrix $`𝖠`$ (recall the definition of $`𝖠`$ above). This increases the magnitude of both the real and imaginary parts of the eigenvalues $`\lambda _k`$, until the threshold where $`|\mathrm{Im}(\sqrt{\lambda }_k)|=\alpha `$, where oscillations appear. These oscillations increase in amplitude as the input increases further, until the animal stops inhaling and the input $`𝐈`$ decreases. Then the oscillations shrink and disappear as the system returns toward its resting state. This rise and fall of oscillations within each sniff cycle give the bulb outputs both a slowly-varying component (2-4 Hz) and a high frequency (25-60 Hz) one, as observed experimentally . It is not known how the synaptic connections represented in the model by the matrices $`𝖧^0`$ and $`𝖶^0`$ develop in the real olfactory bulb, and we do not attempt to model this process here. It is possible that the real bulb acts, to some degree, as an associative memory as a result of this learning. However, our conclusions will not depend on this. Similarly, our analysis does not depend on details of the synaptic matrices, such as their range and degree of connectivity. We require only that the connections lead to distinct oscillation patterns for different odors, with dissimilar patterns evoked by dissimilar odors. ### 2.2 the cortex Our cortical module is structurally similar to that of the bulb. However, there are the following significant differences: (1) The cortex receives an oscillatory input from the bulb, while the bulb receives non-oscillatory (at the time scale of the cortical oscillation) input; (2) The cortex has excitatory-to-excitatory connections, while our bulb module does not. We focus on the local excitatory (pyramidal) and feedback inhibitory interneuron populations. The units that represent them obey equations of motion similar to those for the mitral and granule units of the bulb: $`\dot{u}_i`$ $`=`$ $`\alpha u_i\beta ^0g_v(v_i)+{\displaystyle \underset{j}{}}J_{ij}^0g_u(u_j){\displaystyle \underset{j}{}}\stackrel{~}{H}_{ij}^0g_v(v_j)+I_i^\mathrm{b},`$ (4) $`\dot{v}_i`$ $`=`$ $`\alpha v_i+\gamma ^0g_u(u_i)+{\displaystyle \underset{j}{}}\stackrel{~}{W}_{ij}^0g_u(u_j).`$ (5) Here $`u_i`$ represent the the average membrane potentials of the local excitatory populations and $`v_i`$ those of the inhibitory populations. The synaptic matrix $`𝖩^0`$ is excitatory-to-excitatory connections, $`\stackrel{~}{𝖧}^0`$ is inhibitory-to-excitatory connections, and $`\stackrel{~}{𝖶}^0`$ is excitatory-to-inhibitory connections. For later convenience, we have written the local terms (the effect of $`v_i`$ on $`u_i`$ and vice versa) explicitly, so $`\stackrel{~}{𝖧}^0`$ and $`\stackrel{~}{𝖶}^0`$ have no diagonal elements. We also assume $`J_{ii}^0=0`$. $`I_i^\mathrm{b}`$ are the net inputs from the bulb, both directly and indirectly via the feedforward inhibitory units (see later for the description of this pathway). Like the bulb activity itself, these contain in general both a slow part $`I_i^{\mathrm{b0}}`$, varying with the sniff cycle, and an oscillating ($`\gamma `$-band) part $`\delta I_i^\mathrm{b}`$, i.e., $`I_i^bI_i^{\mathrm{b0}}+\delta I_i^\mathrm{b}`$. We can carry out the same analysis as in the bulb, taking the fixed point as $`(\overline{𝐮},\overline{𝐯})`$, which are determined by $`𝐈^{\mathrm{b0}}`$, i.e., $`\dot{𝐮}=\dot{𝐯}=0`$ at $`(\overline{𝐮},\overline{𝐯})`$ when $`I_i^b=I_i^{b0}`$ with $`\delta I_i^b=0`$. Taking $`𝐮𝐮\overline{𝐮}`$, $`𝐯𝐯\overline{𝐯}`$, linearizing and eliminating the $`v_i`$, we obtain $`\ddot{u}_i`$ $`+`$ $`{\displaystyle \underset{j}{}}[2\alpha \delta _{ij}J_{ij}]\dot{u}_j`$ (6) $`+`$ $`{\displaystyle \underset{j}{}}[(\alpha ^2+\beta _i\gamma _i)\delta _{ij}\alpha J_{ij}+\gamma _i\stackrel{~}{H}_{ij}+\beta _i\stackrel{~}{W}_{ij}+{\displaystyle \underset{k}{}}\stackrel{~}{H}_{ik}\stackrel{~}{W}_{kj}]u_j=(_t+\alpha )\delta I_i^\mathrm{b}.`$ Here $`\beta _i=\beta ^0g_v^{}(\overline{v}_i)`$, $`\gamma _i=\gamma ^0g_u^{}(\overline{u}_i)`$, $`J_{ij}=J_{ij}^0g_u^{}(\overline{u}_j)`$, $`\stackrel{~}{H}_{ij}=\stackrel{~}{H}_{ij}^0g_v^{}(\overline{v}_j)`$, and $`\stackrel{~}{W}_{ij}=\stackrel{~}{W}_{ij}^0g_u^{}(\overline{u}_j)`$. Thus this is a system of driven oscillators coupled by connections $`𝖩`$, $`\stackrel{~}{𝖧}`$, and $`\stackrel{~}{𝖶}`$ and driven by an external oscillatory signal $`\delta \dot{𝐈}^\mathrm{b}+\alpha \delta 𝐈^\mathrm{b}`$, which is proportional to $`\delta 𝐈^\mathrm{b}`$ for a purely sinusoidal oscillation. A single dissipative oscillator driven by an oscillatory force will resonate to it if the frequency of the driving force matches the intrinsic frequency of the oscillator. A system of coupled oscillators has its intrinsic oscillation patterns — the normal modes determined by the coupling. Analogously, it will also resonate to the input when the driving force, a complex vector proportional to $`\delta 𝐈^\mathrm{b}`$, matches one of the intrinsic modes, also a complex vector, in frequency and in its pattern of oscillation amplitudes and phases. It is apparent from Eq. (6) that the matrices $`\stackrel{~}{𝖧}`$ and $`\stackrel{~}{𝖶}`$ play the same roles. Therefore, for simplicity, we will drop the inhibitory-to-excitatory couplings $`\stackrel{~}{𝖧}`$ from now on, thinking of the fact that the real anatomical long-range connections appear to come predominantly from excitatory cells. #### Odor selectivity and sensitivity In our model, the olfactory cortex functions as an associative memory, as described and modeled by a number of authors . It is similar to a Hopfield model, but instead of stationary patterns it stores oscillating patterns which vary in phase as well as magnitude across the units of the network. The memory pattern for the $`\mu ^{th}`$ odor is described by a complex vector $`\xi ^\mu `$, whose component $`\xi _i^\mu `$ describes both the relative amplitude and phase of the oscillation in the $`i^{th}`$ unit. The cortex stores the memories about the odours in the synaptic weights $`𝖩^\mathrm{𝟢}`$ and $`\stackrel{~}{𝖶}^\mathrm{𝟢}`$, or, effectively, the coupling between oscillators. It then recognizes the input odors, as coded by the oscillating input patterns $`\delta 𝐈^\mathrm{b}`$ (which are linearly related to the bulbar oscillatory outputs), by resonating to them, giving high-amplitude oscillatory responses itself. If, however, the input $`\delta 𝐈^\mathrm{b}`$ does not match one of the stored odor patterns $`\xi ^\mu `$ closely enough, the cortex will fail to respond appreciably. In the present model the memory pattern $`\xi _i^\mu `$ for odors $`\mu =1,2,\mathrm{}`$ are designed into the synaptic connections $`𝖩`$ and $`\stackrel{~}{𝖶}`$. Let $`\omega `$ be the oscillation frequency, $`\delta I_i^\mathrm{b}e^{\mathrm{i}\omega t}`$. Once the oscillation reaches a steady amplitude $`u_ie^{\mathrm{i}\omega t}`$, we have $`\dot{u}_i=\mathrm{i}\omega u_i`$, $`\ddot{u}_i=\mathrm{i}\omega \dot{u}_i`$, so we get $$\dot{u}_i=[2\alpha \frac{\mathrm{i}}{\omega }(\beta _i\gamma _i+\alpha ^2)]u_i+\underset{j}{}[J_{ij}\frac{\mathrm{i}}{\omega }(\beta _i\stackrel{~}{W}_{ij}\alpha J_{ij})]u_j+\frac{\mathrm{i}}{\omega }(\mathrm{i}\omega +\alpha )\delta I_i^\mathrm{b}.$$ (7) The second term \[…\] on the right hand side gives an effective coupling between the oscillators. From now on in this analysis we will make the approximation that the different local oscillators have the same natural frequencies, i.e. $`\beta _i\gamma _i`$ is independent of $`i`$. Assuming further that the oscillation frequencies for different odors are nearly the same, the odor patterns can then be stored in the matrices in a generalized Hebb-Hopfield fashion as $$M_{ij}[J_{ij}\frac{\mathrm{i}}{\omega }(\beta \stackrel{~}{W}_{ij}\alpha J_{ij})]=J\underset{\mu }{}\xi _i^\mu \xi _j^\mu ,$$ (8) or, with $`\xi _i^\mu `$ expressed in terms of amplitudes and phases as $`|\xi _i^\mu |\mathrm{exp}(\mathrm{i}\varphi _i^\mu )`$, $`J_{ij}`$ $`=`$ $`J{\displaystyle \underset{\mu }{}}|\xi _i^\mu ||\xi _j^\mu |\mathrm{cos}(\varphi _i^\mu \varphi _j^\mu )`$ (9) $`\beta \stackrel{~}{W}_{ij}`$ $`=`$ $`J{\displaystyle \underset{\mu }{}}|\xi _i^\mu ||\xi _j^\mu |[\omega \mathrm{sin}(\varphi _i^\mu \varphi _j^\mu )+\alpha \mathrm{cos}(\varphi _i^\mu \varphi _j^\mu )].`$ (10) Note that here both kinds of connections, $`𝖩`$ (excitatory-to-excitatory) and $`\stackrel{~}{𝖶}`$ (excitatory-to-inhibitory), are used to store the amplitude and phase patterns of the oscillation. $`𝖩`$ is symmetric, while $`\stackrel{~}{𝖶}`$ is not. These connections can be obtained by an online algorithm, a simplified version of the full Hebbian learning treated by Liljenström and Wu . Suppose the cortex has effective oscillatory input $`\delta 𝐈^b=\xi ^\mu e^{\mathrm{i}\omega t}+\xi ^\mu e^{\mathrm{i}\omega t}`$ during learning of the $`\mu ^{th}`$ pattern. Here we make explicit the real nature of the signals. Suppose also that the $`𝖩`$ and $`\stackrel{~}{𝖶}`$ connections inactive, consistent with the picture proposed by Wilson, Bower and Hasselmo , who suggested that learning occurs when the long-range intracortical connections are weakened by neuromodulatory effects. Then the linearized (4) and (5) are simply $`\dot{u}_i+\alpha u_i`$ $`=`$ $`\beta v_i+\xi _i^\mu e^{\mathrm{i}\omega t}+\xi _i^\mu e^{\mathrm{i}\omega t}`$ $`\dot{v}_i+\alpha v_i`$ $`=`$ $`\gamma u_i,`$ (11) with solution $`u_i(t)`$ $`=`$ $`{\displaystyle \frac{\mathrm{i}\omega +\alpha }{\omega ^2+\alpha ^2+\beta \gamma 2\mathrm{i}\alpha \omega }}\xi _i^\mu e^{\mathrm{i}\omega t}+\mathrm{c}.\mathrm{c}.`$ $`v_i(t)`$ $`=`$ $`{\displaystyle \frac{\gamma }{\omega ^2+\alpha ^2+\beta \gamma 2\mathrm{i}\alpha \omega }}\xi _i^\mu e^{\mathrm{i}\omega t}+\mathrm{c}.\mathrm{c}.`$ (12) where c.c. denotes complex conjugate. In other words, the cortical activities are clamped by the inputs. For Hebbian learning, $`\dot{J}_{ij}u_i(t)u_j(t)`$, and, after time averaging, $`\delta J_{ij}_0^{2\pi /\omega }u_i(t)u_j(t)dt`$, leading to $`\delta J_{ij}`$ $``$ $`{\displaystyle \frac{\omega ^2+\alpha ^2}{|\omega ^2+\alpha ^2+\beta \gamma 2\mathrm{i}\alpha \omega |^2}}(\xi _i^\mu \xi _j^\mu +\xi _i^\mu \xi _j^\mu )`$ (13) $`=`$ $`2{\displaystyle \frac{\omega ^2+\alpha ^2}{|\omega ^2+\alpha ^2+\beta \gamma 2\mathrm{i}\alpha \omega |^2}}|\xi _i^\mu ||\xi _j^\mu |\mathrm{cos}(\varphi _i^\mu \varphi _j^\mu ).`$ Similarly, $`\delta W_{ij}_0^{2\pi /\omega }v_i(t)u_j(t)dt`$ leading to $`\delta W_{ij}`$ $`{\displaystyle \frac{\gamma }{|\omega ^2+\alpha ^2+\beta \gamma 2\mathrm{i}\alpha \omega |^2}}[(\mathrm{i}\omega +\alpha )\xi _i^\mu \xi _j^\mu +(\mathrm{i}\omega +\alpha )\xi _i^\mu \xi _j^\mu ]`$ (14) $`=`$ $`{\displaystyle \frac{2\gamma }{|\omega ^2+\alpha ^2+\beta \gamma 2\mathrm{i}\alpha \omega |^2}}[\omega |\xi _i^\mu ||\xi _j^\mu |\mathrm{sin}(\varphi _i^\mu \varphi _j^\mu )+\alpha |\xi _i^\mu ||\xi _j^\mu |\mathrm{cos}(\varphi _i^\mu \varphi _j^\mu )].`$ Then, if the relative learning rates for $`𝖩`$ and $`\stackrel{~}{𝖶}`$ are tuned appropriately, we simply recover the formulae (9) and (10). In actual online learning, we can use high-pass versions of $`𝐮`$ and $`𝐯`$ to learn $`𝖩`$ and $`\stackrel{~}{𝖶}`$ to remove the baseline value, i.e., the operation point $`\overline{𝐮}`$ and $`\overline{𝐯}`$, which does not contain odor information. To see the selective resonance explicitly, suppose that different patterns $`\xi ^\mu `$ are orthogonal to each other. Let us denote the overlap $`(1/N)_i\delta I_i^\mathrm{b}\xi _i^\lambda `$ of the input $`\delta I_i^\mathrm{b}`$ with the stored pattern $`\xi _i^\lambda `$ by $`\delta I^\lambda `$. Then, multiplying (7) by $`\xi _i^\lambda `$ and summing on $`i`$, we find that at steady oscillatory state, the response $`u^\lambda (1/N)_iu_i\xi _i^\lambda `$ to pattern $`\xi ^\lambda `$ obeys $$\dot{u}^\lambda =(2\alpha J)u^\lambda \frac{\mathrm{i}}{\omega }(\beta \gamma +\alpha ^2)u^\lambda +\frac{\mathrm{i}}{\omega }(\mathrm{i}\omega +\alpha )\delta I^\lambda $$ (15) This is like an oscillator with oscillation frequency $`(\beta \gamma +\alpha ^2)/\omega `$ and an effective oscillation decay rate $`2\alpha J`$. It resonates to external oscillatory input of frequency $`\omega \sqrt{\beta \gamma +\alpha ^2}`$ with a steady state amplitude $$u^\lambda =\frac{(\mathrm{i}\omega +\alpha )\delta I^\lambda }{\beta \gamma +\alpha ^2\omega ^2\mathrm{i}\omega (2\alpha J)}\frac{(1+i\alpha /\omega )\delta I^\lambda }{2\alpha J},$$ (16) However, for an input $`\delta I_i^\mathrm{b}`$ orthogonal to all the stored patterns, $`\delta I^\lambda =0`$ for all $`\lambda `$, and the resonance will be washed out when $`J<2\alpha `$. For $`J>2\alpha `$, the network will support spontaneous oscillations analogous to those in the bulb, but not as observed in the cortex. The effect of the long-range couplings, through the parameter $`J`$, is to reduce the damping in the circuit from $`2\alpha `$ to $`2\alpha J`$ when the input matches a stored pattern, thereby sharpening the resonance as $`J2\alpha `$ while we keep $`J<2\alpha `$. On the other hand, the resonant driving frequency depends only on the single-oscillator parameters $`\alpha `$, $`\beta `$ and $`\gamma `$. This oscillatory associative memory enjoys the usual properties that characterize Hopfield networks , including rapid convergence (a few oscillation cycles if the presented pattern has reasonable overlap with a stored one), robustness with respect to noise and corrupted input, and a storage capacity of the order of $`N`$ random patterns, where $`N`$ is the network size. ### 2.3 Coupling between bulb and cortex The model has both feedforward (bulb-cortex) and feedback (cortex-bulb) connections. The former transmit the bulbar encoding of the input odors to the cortex for recognition, while the latter permit segmentation by producing adaptation to recognized odor objects. #### bulb to cortex As mentioned in the Introduction, in the real cortex, the excitatory cells receive input from the bulb both directly from the fibers of the lateral olfactory tract and in a slower pathway via feedforward inhibitory interneurons in the cortex. We model this in the following way. The synapses from local bulb populations $`j`$ to local cortical populations $`i`$ are specified by a matrix $`C_{ij}^{\mathrm{b}\mathrm{c}}`$. The values of these connections are not important in the model, and very little is know about them, so we will take them to be random. The resulting signals are then fed to the excitatory cells, both directly and, with the opposite sign, through a parallel low-pass filter, representing the effect of the feedforward inhibitory cells; see Fig. 1. Details are given in the appendix. The combination of the direct excitatory and low-pass filtered inhibitory signals makes the feedforward pathway act as a high-pass filter, partially cancelling the slow part $`𝐈^{\mathrm{b0}}`$ of the bulb output from the cortical input. Consequently, the net input to the cortical excitatory units is dominated by the oscillatory component of the bulb activity, which encodes information about the odor input. (We do not know how well such a cancellation is actually achieved in real olfactory systems, but this could be tested experimentally.) #### cortex to bulb The odor-specific adaptation that forms the basis for odor segmentation in our model is implemented using a feedback signal from the cortex to the granule units of the bulb. We do not know how such a signal is generated in animals, or even whether it is, although anatomically such a pathway exists. If the signal does exist, it likely also involves areas such as entorhinal cortex, which contributes to the centrifugal input to the bulb. These areas lie outside the scope of the present model, so we will simply construct a suitable signal and explore the consequences. In exploratory computations, we have found that this form of feedback control only works if the signal is slowly varying in time (on the order of the sniff cycle time or slower). Merely feeding back the oscillating cortical activities does not appear to permit any kind of robust stimulus-specific adaptation. Thus, we generate the feedback signal in the following ad hoc fashion: First each excitatory cortical output $`g_u(u_i)`$ is run through a threshold-linear element to remove its non-oscillatory part, which carries no odor information. Then the output of this element is run through a low-pass filter. The time constants of this filter are on the order of the sniff cycle or longer. The net result is a signal pattern which takes a sniff-cycle time or so to grow to full strength. The signal component from excitatory unit $`i`$ will be proportional to the amplitude of the oscillation of that unit, so this signal will contain information about the odor that evoked the cortical oscillation pattern. The explicit form of the equations used to generate the feedback signal in the simulations is given in the Appendix. Since we rectify and low-pass only the excitatory cortical outputs $`g_u(u_i)`$, the feedback signal includes only the odor information coded in the amplitude but not in the phase pattern of the cortical oscillation. Phase information could be included by (for example) feeding the difference signals $`g_u(u_i)g_u(u_j)`$ through the rectification and low-pass processes. However, we have not explored such mechanisms in this work. The granule units in the bulb respond to the feedback signals by changing their activities proportional to it. These changes are then transmitted to the mitral cells by the synaptic matrix $`𝖧`$. As shown by Li , a feedback signal $$𝐅𝖧^1𝐈,$$ (17) will, when transmitted onward to the mitral units, cancel the odor inputs to the bulb (in linear approximation). In our model we want to make this cancellation work for all the odor patterns stored in the cortex. Denoting by $`G_j^\mu `$ the rectified and low-passed cortical output when the system is stimulated by odor pattern $`I_k^\mu `$, this can be achieved by a Hebbian feedback connection matrix $`𝖢^{\mathrm{c}\mathrm{b}}`$ that maps $`𝐆^\mu `$ to feedback signal $`𝐅^\mu `$ for each odor $`\mu `$ in a single layer network: $$C_{ij}^{\mathrm{c}\mathrm{b}}\underset{\mu }{}F_i^\mu G_j^\mu =\underset{k}{}H_{ik}^1\underset{\mu }{}I_k^\mu G_j^\mu .$$ (18) ## 3 Simulations We have simulated a network with bulb and cortical modules each consisting of 50 excitatory and 50 inhibitory units. They were coupled as described in Sect. 2.3 and the Appendix. The coupled differential equations were integrated using a fourth-order Runge-Kutta routine from Numerical Recipes . We used three random odor input patterns $`I_i^\mu `$. Their elements were drawn independently for each $`i`$ and $`\mu `$ from a uniform distribution on (0,1\]. The elements of the granule-to-mitral synaptic matrix $`𝖧`$ were taken to have the form $`H_{ij}=\mathrm{const}.\delta _{ij}`$. We designed the mitral-to-granule matrix $`𝖶`$ so as to make the bulb oscillate in response to the three input patterns, taking $`W_{ij}\mathrm{Im}_{\mu =1}^3\zeta _j^\mu \zeta _j^\mu `$. Here the $`\zeta _i^\mu `$ are complex, with amplitudes resembling the input odor patterns $`I_i^\mu `$ and with random phases. Since $`𝖶`$ should have non-negative elements, we simply zeroed out the negative $`W_{ij}`$ in the construction.<sup>1</sup><sup>1</sup>1In the bulb model of Li and Hopfield, the idea was that extensive asymmetric random synapses would, for a large network, automatically generate a distributed encoding of odors in the amplitudes and phases of oscillation patterns in the network. Here we will be more concerned with how these patterns are processed by the cortex, so, for convenience, we have engineered particular amplitude patterns in through the bulbar W matrix in this fashion. However, the particular forms used for $`𝖧`$ and $`𝖶`$ are not important for the problem that we are studying here, as long as $`𝖠\mathrm{𝖧𝖶}`$ is sufficiently asymmetric. This dilution did not affect the bulb oscillations qualitatively. Other parameters were set as in , so the evoked oscillations were in the 40-Hz range. The cortical design followed Sect. 2.2. The local couplings $`\beta ^0`$ and $`\gamma ^0`$ were chosen so that the cortical oscillation frequency roughly matched the bulbar one, i.e., $`\beta ^0\gamma ^0+\alpha ^2\overline{\omega }^2`$ (see equation (16), where $`\overline{\omega }`$ is the average oscillation frequency in the bulbar outputs. The inhibitory units had the sigmoidal activation function used in the model of the bulb . In some of our simulations, the activation function of the excitatory units also had this form. In obtaining the results presented here, however, we used a piecewise linear activation function with gains of 1 and 2, respectively, in the low- and high-input regions above threshold. This choice was made only for convenience in analyzing the nonlinear dynamics and is not essential for the function of the network. The cortical synaptic matrices $`𝖩`$ and $`\stackrel{~}{𝖶}`$ were designed to store oscillation patterns for two of the three odor input patterns, in the following way. For each of the two odors, we stimulated the bulb with its input pattern $`I_i^\mu `$ and fed the resulting oscillatory bulb output through the bulb-to-cortex matrix $`C_{ij}^{\mathrm{b}\mathrm{c}}`$ and the subsequent high-pass filtering operation to the cortex, with the intracortical connections $`𝖩^0`$ and $`\stackrel{~}{𝖶}^0`$ set to zero. The resulting oscillation patterns in the cortical units for the two odors were then used as $`\xi ^\mu `$ in constructing $`𝖩`$ and $`\stackrel{~}{𝖶}`$. We modified the Hebb rule (eq. (8) or eqs. (9) and (10)) slightly, using, instead, a pseudoinverse formula $$M_{ij}=J\underset{\mu }{}\xi _i^\mu \eta _j^\mu ,$$ (19) where $`_i\eta _i^\mu \xi _i^\nu =N\delta _{\mu \nu }`$. This was done only to reduce finite-size effects due to mutual overlaps (of order $`\sqrt{N}`$) between patterns, and would be inessential in sufficiently large networks. As explained in section 2.3 and the Appendix, the slowly-varying feedback signal used for the odor-specific adaptation was generated by a threshold-linear rectification, followed by a pair of simple linear filters. The time constants of these (3 and 0.3 sec respectively) would made it take 10-12 256-ms sniff cycles to generate a full strength feedback signal if the cortical signal were held constant. Similarly, the adaptation takes just as long to wear off after the stimulus is removed. Like the intracortical $`𝖬`$ matrix, the cortex-to-bulb matrix $`𝖢^{\mathrm{c}\mathrm{b}}`$ was modified using the projection-rule algorithm to eliminate finite-size overlap effects between the cortical oscillation patterns of different odors. Thus, in the formula (18), we replaced the rectified and low-pass-filtered cortical patterns $`G_j^\mu `$ by $`\stackrel{~}{G}_j^\mu `$, where $`\stackrel{~}{𝐆}^\mu `$ are vectors such that $`\stackrel{~}{𝐆}^\mu 𝐆^\nu =N\delta ^{\mu \nu }`$. Fig. 2 shows the bulbar and cortical oscillatory response patterns evoked on 5 of the 50 mitral or cortical excitatory units by three odors: A, B, and C. Only odors A and B are stored in the cortical memory in the $`𝖩`$ and $`\stackrel{~}{𝖶}`$ matrices. Different amplitude response patterns to different odors are apparent. The cortex resonates appreciably to only odors A or B, but not to C, demonstrating the selectivity of the cortical response. Fig. 3 demonstrates odor adaptation to odor A. The response amplitudes decay quicky in successive sniffs, although the oscillation patterns do not change appreciably before the amplitudes decay to zero. The way this comes about is that the feedback signal generated by A, when relayed by the granule cells to the mitral ones, creates an effective extra input signal $`\overline{\mathrm{A}}`$ (anti-A), and by the third sniff $`\mathrm{A}+\overline{\mathrm{A}}0`$. To quantify the similarity between oscillation patterns, we extract an $`N`$=50 dimensional complex vector $`𝐎`$ from the temporal Fourier transform of the activity of the cortical excitatory units during the sniff cycle, with the component $`O_i`$ specifying the amplitude and phase of the oscillations in excitatory unit $`i`$. We can measure the similarity between $`𝐎`$ and $`𝐎^{}`$ by the normalized overlap $`S_{\mathrm{OO}^{}}=|𝐎|𝐎^{}/(|𝐎||𝐎^{}|)`$, which is 1 for $`𝐎𝐎^{}`$ and near zero ($`\mathrm{O}(1/\sqrt{N})`$) for two unrelated patterns. Calling the pattern vectors $`𝐀^0`$, $`𝐀^1`$, $`𝐀^2`$, and $`𝐀^3`$ for cortical response to odor A without adaptation and during the first, second, and third sniff cycles of the adaptation respectively, we find $`S_{\mathrm{A}^0\mathrm{A}^1}=0.9997`$, $`S_{\mathrm{A}^0\mathrm{A}^2}=0.992`$, and $`S_{\mathrm{A}^0\mathrm{A}^3}=0.74`$, with response amplitudes $`|𝐀^1|/|𝐀^0|=0.97`$, $`|𝐀^2|/|𝐀^0|=0.3`$, $`|𝐀^3|/|𝐀^0|=0.08`$. Thus, the strength of the response is already significantly weakened after one sniff, but its cortical pattern of variation remains undistorted through several sniffs. The way this adaptation varies in successive sniffs depends on both the time constants in the feedback circuitry (as discussed above) and the strength of the filtered signal fed back to the bulb. In the simulations shown here, the latter was strong enough that even after one sniff, a large fraction of the input signal is cancelled by the feedback, and after two sniffs the cancellation was nearly complete. A smaller feedback strength and a correspondingly longer time constant of the feedback circuitry would make it take longer for the adaptation to set in. Similarly, the time it takes for the adaptation, once established, to wear off is set by the same time constants (for the values used here, around 3 s or 12 sniff cycles). Fig. 4a demonstrates the segmentation capability of the system. The response $`𝐁^{\mathrm{seg}}`$ to the odor mixture A+B at the third sniff after the first 2 sniffs of odor A is quite similar to that, $`𝐁^0`$, to odor B alone: $`S_{\mathrm{B}^{\mathrm{seg}}\mathrm{B}^0}=0.993`$, and $`|𝐁^{\mathrm{seg}}|/|𝐁^0|=0.91`$. Thus, although A is still present, so is the anti-A, so the net signal to the mitral units is $`\mathrm{A}+\overline{\mathrm{A}}+\mathrm{B}\mathrm{B}`$. This demonstrates odor-specific adaptation in the model. The system responds with the activity pattern characterizing the new odor, essentially undistorted by the existing odors in the environment, thus effectively achieving odor segmentation. Odor B can be segmented as long as it enters after the adaptation to A is established, in this model at the 3rd or any subsequent sniffs. However, if odor A is suddenly withdrawn at the start of the 3rd sniff, when odor B is introduced, the system response to odor B is weakened and distorted (this is particularly noticable in the bulbar responses). The reason for this is that the effective total input is now $`\mathrm{B}+\overline{\mathrm{A}}\mathrm{B}\mathrm{A}`$, which is not at all like $`\mathrm{B}`$ (Fig. 4, b and c). This corresponds to the psychophysically observed cross-adaptation — after sniffing one odor, another odor at next sniff smells less strong than it normally would and may even smell different . In the normal olfactory environment, however, such sudden and near complete withdrawal of an odor seldom happens. Let $`𝐁^{\mathrm{cross}}`$ and $`\stackrel{~}{𝐁}^{\mathrm{cross}}`$ be the cortical response vectors to cross adapted odor B and odor 1.5B. Comparing with the response to odor B alone, we find $`S_{\mathrm{B}^0\mathrm{B}^{\mathrm{cross}}}=0.94`$, $`|\mathrm{B}^{\mathrm{cross}}|/|\mathrm{B}^0|=0.23`$; $`S_{\mathrm{B}^0\stackrel{~}{\mathrm{B}}^{\mathrm{cross}}}=0.97`$, $`|\stackrel{~}{\mathrm{B}}^{\mathrm{cross}}|/|\mathrm{B}^0|=0.74`$. We can understand these results in the following way. The feedback input $`\overline{\mathrm{A}}\mathrm{A}`$ acts to move the bulb operating point $`\overline{x}_i`$ to lower gain values (for units where $`I_i^\mathrm{A}`$ is strong), thereby weakening the overall response. For normal-strength B, most of the mitral units in the bulb do not respond much, so the cortical response is correspondingly weak and distorted relative to that to B in the absence of adaptation. The stronger input 1.5B evokes a stronger bulb response, however, and the cortical response is stronger and better (but still imperfectly) correlated with the unadapted pattern. Since the olfactory bulb is nonlinear, the odor mixture A+B does not induce a bulbar response equal to the sum of the responses to A and B individually. Consequently, the unadapted cortical response to it (Fig. 5, left panel) is weaker than that to A or B (the bulb response to the mixture is not embedded in the cortical connections) and not strongly correlated with the responses to the pure odors. The situation is similar to that for any other unstored odor, such as C (Fig. 2, panel C), to which there is almost no adaptation in the bulb because there is almost no cortical signal to feed back. The unadapted cortical response to A+B is stronger than that to C because the nonlinearity in the bulb here is not strong enough to completely destroy correlations between its reponses to the individual odors A and B and that to their mixture. Nevertheless, the weakness of the cortical response reduces the feedback to the bulb significantly, and the system does not adapt to the mixture as effectively as to individual odors, as shown in the middle and right panels of Fig. 5. We also note that because the feedback is weak, the attenuation of the signals in both bulb and cortex, is also weaker than for pure stored odors (cf Fig. 3). Thus, the cortical response to the mixed odor, while initially weaker than that to pure stored ones, lasts longer. ## 4 Discussion We have presented a computational model for an olfactory system that can detect, recognize and segment odors. Detection is performed in the bulb, which encodes odors in oscillatory activity patterns. Recognition is carried out by the cortex using a resonant associative memory mechanism. Finally, segmentation is implemented by a slowly-varying feedback signal which acts to cancel the specific input that evoked the resonant cortical response. The model is constrained by a few basic anatomical and physiological facts: Odors evoke oscillatory activity in populations of excitatory and inhibitory neurons in both bulb and cortex, these two structures are coupled by both feedforward and feedback connections, reducing the cortical feedback enhances the bulbar responses, and the system exhibits odor-specific adaptation. Within these constraints, we have tried to build a minimal model. We have taken the bulb module from earlier work by one of us and augmented it with a model of the pyriform cortex and with feedforward and feedback connections between it and the bulb. We have ignored many further known details of real olfactory systems that do not bear directly on the fundamental property of stimulus-specific adaptation, and when we have had to go beyond current knowledge (as in constructing the feedback signal) we have done so in a purely phenomenological way, avoiding hypothesizing specific details unrelated to the function of the system. From the analysis of the model and the simulations we can see how the basic computations necessary for olfactory segmentation might be carried out by the neural networks of the bulb and cortex. But do real olfactory systems actually function in this way? This can be tested at the level of both the assumptions we put into the model and the properties we find for it. First of all, we have assumed that the feedback from cortex to bulb is slowly varying (i.e. that firing rates for the feedback fibers vary on the timescale of the sniff cycle, but not of the oscillations found in both the bulb and cortex). Furthermore, we have assumed that this feedback is odor-specific. While the existence of some feedback is well-established, neither of these specific hypotheses has been tested experimentally. However they both could be. Properties we find in the model, beyond the fact that it successfully implements segmentation, can also be tested. These include the following: First, the fact that the feedback signal requires strong cortical activity to drive it means that unfamiliar (unlearnt) odors will not be adapted to as strongly as familiar ones, so they will not be so easily segmented from subsequently presented ones. As we saw in Fig. 5, this expectation also applies to unfamiliar mixtures of familiar odors. Furthermore, as we also noted, we expect the weakening of the (initially weaker) responses with adaptation to be slower for such mixtures than for familiar odors. Second, cross-adaptation, as illustrated in Fig. 4, is a necessary consequence of the slow feedback: The effective bulb input $`\overline{\mathrm{A}}\mathrm{A}`$, from the previous presence of the adapting stimulus, will be present for some time (depending on the time constants of the feedback circuitry) whether odor A remains in the environment or not. Thus the total input to the bulb with A still present will be very different from that with A suddenly removed. If there is odor-specific adaptation of the kind necessary to perform segmentation when A remains in the environment (A cancelled by $`\overline{\mathrm{A}}`$), then a different response must occur when A is withdrawn. Present evidence on cross-adaptation is rather limited, but psychophysical and electrophysiological investigation of this phenomenon would be helpful in pinning down quantitatively the time constants of the circuitry involved in segmentation. If odor-specific adaptation is not implemented using our cortical feedback mechanism, how else might it be done? One possibility to consider is single-unit-level adaptation (or fatigue), which can be implemented in a network like ours by making the threshold for each unit dependent on its own recent activity. In a model with the structure of ours (with bulb and cortical modules) but without feedback, such fatigue would have to be implemented in the bulb; otherwise the activity there would not exhibit adaptation. This presents a problem if the activity patterns of different odors overlap significantly – it is not evident that one can avoid changing the response to a new odor when some of the units active in the normal response to it are to be fatigued. Indeed, in investigations of simple oscillatory associative memories with such adaptation , temporal segmentation has been found only for patterns with rather weak mutual overlap. This overlap will be weak for sparse patterns, but it is not clear how sparse real evoked bulb and cortical activity patterns are, when looked at at the level of resolution of the units in our model. This problem is not present for the mechanism we propose, in which bulb units themselves are not fatigued. Rather, the mechanism cancels the input to bulb units in exactly the degree that they receive input from the adapting odor. It is as if every receptor activated by an odor became adapted by an amount exactly equal to its initial response. In our model, the feedback connections to the inhibitory bulb units have to have just the right values to produce the necessary cancellation. In real olfactory systems, the strengths of the centrifugal synapses on granule cells are presumably determined by some learning mechanism, and for our model to apply it is necessary that this mechanism find the right values for them. As we know nothing about this mechanism, here in our model we just assumed the necessary form. This form has a degree of plausibility because it is Hebbian, but very little is known yet about learning in these synapses. Investigations could shed important light on the validity of this key element of the model. Another plausible mechanism, which could implement odor-specific adaptation in the bulb in more or less the right manner, is adaptation of receptor-bulb synapses in such a way that the inputs to bulb capture mainly the transient but not static odor inputs. This would reduce the input signal for the adapting odor directly, at just the right places, and so does not suffer from the problems that single-unit fatigue in the bulb does. However, there is a simple difference between the predictions of such a model and ours, since in ours the cortex, functioning as an associative memory, only sends its feedback to the bulb (or only sends it at full strength) for learnt odors. The receptor-mitral synaptic adaptation model would exhibit the same degree of odor-specific adaptation for all odors, learnt or not. Of course, both mechanisms could be present, and the difference could be large or small according to the relative size of the two contributions. The fact that we have employed both excitatory-to-excitatory ($`𝖩`$) and excitatory-to-inhibitory ($`\stackrel{~}{𝖶}`$) cortical connections enhances the associative memory function by permitting oscillation patterns which differ in phase as well as amplitude. This is of no help for selective adaptation in the model as described here, since phase information is lost in the generation of the feedback signal, but this information could be retained using more elaborate mechanisms, as mentioned in Sect. 2.3. It is not clear whether real olfactory systems code odors in the phases of their oscillation patterns. However, in any case, a restricted version of our cortex, without $`\stackrel{~}{𝖶}`$, could function with only amplitude-modulated patterns, similarly to the model of Wang et al . The addition of intrinsic oscillatory properties for individual units or, implicitly, the individual neurons in the populations they represent, would not change the properties of such a network qualitatively. The three tasks carried out by the system – detection, recognition, and segmentation – are computationally linked. For example, even if an ambiguous or weak odor is “recognized” by the pyriform cortex in the sense that a characteristic oscillatory response is evoked there, that response may be too weak to suppress further bulbar response. Then the system will continue to respond to the odor in the same way as if it had not recognized it; that is, the odor-specific adaptation necessary for segmentation can be seen as part of the recognition process. While our units correspond to functional groups of neurons in real olfactory systems, our model is of higher resolution than that of Ambros-Ingerson et al . While we emphasize the coding of odor information in distributed oscillation patterns, their model contains no explicit treatment of dynamics on the 40-hz timscale or of the temporal segmentation problem. They address instead a higher-level problem (hierarchical odor classification) with a higher-level model. In such more complex situations, cortex-to-bulb feedback could be a more general, active phenomenon than in the limited-scope problem we consider, but we do not address such issues here. Our network performs what might be called “the simplest cognitive computation”. It is natural to expect that evolution has employed elaborations on this structure in other sensory systems and in central processing. For example, hippocampal processing also employs oscillations, long-range intra-area associative connections, and feedback . In another context, work by one of us on visual processing suggests a function for slow feedback to inhibitory neurons from higher areas in modulating the computations carried out in area V1. Our hope is that studying and modeling the olfactory system in the way we have done here will lead to insights into aspects of top-down/bottom-up interactions in other cognitive computations. ## Appendix: Bulb-cortex coupling: implementation details ### Feedforward In the feedforward pathway from bulb to cortex, the mitral unit outputs $`g_x(x_i)`$ are fed both directly to the excitatory cortical units and in parallel, indirectly via feedforward inhibitory units. The process, as indicated schematically in Fig. 1, is described by the equations $`L_i`$ $`=`$ $`{\displaystyle \underset{j}{}}C_{ij}^{\mathrm{b}\mathrm{c}}g_x(x_j)`$ (20) $`\dot{z}_i`$ $`=`$ $`\alpha _{\mathrm{ff}}z_i+L_i`$ (21) $`I_i^\mathrm{b}`$ $`=`$ $`L_i\sigma g_z(z_i).`$ (22) Here $`L_i`$ is the input signal to the cortical location $`i`$, $`𝖢^{\mathrm{b}\mathrm{c}}`$ is the connection matrix that transforms the mitral outputs to the cortical inputs, $`z_i`$ are the membrane potentials of the feedforward inhibitory units, $`g_z(.)`$ is their activation function and $`\alpha _{\mathrm{ff}}^1`$ is their time constant. $`I_i^\mathrm{b}`$ is then the total input signal to the $`i`$-th cortical excitatory unit in Eqn. (4). In general, this input contains both slowly-varying and rapidly-oscillating components. The pathway via the inhibitory feedforward units acts like a low-pass filter. Thus, the net effect is that the rapidly-varying or high frequency components, which contain the odor information, are transmitted to the cortex. In the simulations reported in Sect. 3, we took $`g_z(.)`$ to have two regions of different gain values, with a smaller gain at smaller input. We designed $`\sigma `$ and the parameters of $`g_z(.)`$ so that the net slow component of $`I_i^\mathrm{b}`$ pushed the cortical operation points $`\overline{u}_i`$ and $`\overline{v}_i`$ to stable values close to, but below, their high gain region. Thus the cortex had a stable operating point, enabling it to carry out its associative memory function more cleanly that without this engineering refinement. We make no claims about biological realism for the details of the feedfoward mechanism. However, some kind of effective high-pass filter is essential to the robust functioning of the model. Further experimental investigation of the dynamical properties of the feedfoward pathway would be important for understanding how it actually works. ### Feedback To generate the half-wave rectified, low-passed feedback signal to the bulb from the cortical excitatory unit outputs, we use three successive groups of units followed by a synaptic matrix, as shown in Fig. 6: $`\dot{p}_i`$ $`=`$ $`\alpha _{\mathrm{fast}}p_i+g_u(u_i),\dot{q}_i=\alpha _{\mathrm{slow}}q_i+g_p(p_i),\dot{r}_i=\alpha _{\mathrm{slow}}^{}r_i+q_i,`$ (23) $`I_i^\mathrm{c}`$ $`=`$ $`m(t){\displaystyle \underset{j}{}}C_{ij}^{\mathrm{c}\mathrm{b}}g_r(r_j),`$ (24) where $`m(t)`$ is a modulating signal that synchronizes with breathing, increasing during inhalation and decreasing during exhalation. With a short time constant $`1/\alpha _{\mathrm{fast}}`$ and a strong nonlinear $`g_p`$, the $`p_i`$ unit has a output $`g_p(p_i)`$ which is effectively $`g_u(u_i)`$ thresholded above the average signal level. This “rectified” output is then transformed by the two subsequent units $`q_i`$ and $`r_i`$, both with long time constants $`1/\alpha _{\mathrm{slow}}`$ and $`1/\alpha _{\mathrm{slow}}^{}`$, into a slowly-varying signal, which is modulated by a function $`m(t)`$ (representing the breathing rhythm of the animal) and fed back via the connections $`𝖢^{\mathrm{c}\mathrm{b}}`$ to produce the centrifugal input $`𝐈^\mathrm{c}`$ to the granule units in the bulb. It is not necessary to use two low-pass filter operations; the model works qualitatively the same with just one. However, adding the second one delays the feedback signal somewhat, giving the oscillations time to establish themselves before the feedback begins to act. In a more complete model, the large time constants $`1/\alpha _{\mathrm{slow}}`$ and $`1/\alpha _{\mathrm{slow}}^{}`$ could emerge as a dynamic network property of secondary olfactory areas. Similarly, the modulating signal $`m(t)`$ could arise from additional signals from other parts of the brain.
warning/0002/physics0002015.html
ar5iv
text
# Density Functional Study of adsorption of molecular hydrogen on graphene layers ## I INTRODUCTION The adsorption of hydrogen by different forms of carbon has been studied by different groups. Dillon et al were the first to study the storage of molecular hydrogen by assemblies of single wall carbon nanotubes (SWCNT) and porous activated carbon. They pointed out that the attractive potential of the walls of the pores makes it possible a high density storage. From temperature-programmed desorption experiments Dillon et al concluded that those forms of carbon are promising candidates for hydrogen storage, although the density of hydrogen is still low in order to meet the requirements of the DOE Agency for novel hydrogen storage systems. More recently Levesque et al , Ye et al , and Liu et al also studied the adsorption of molecular hydrogen on SWCNT at different temperatures and pressures. Chambers et al have reported obtaining an extraordinary storage capacity by some graphite nanofibers but Wang and Johnson have tried unsuccessfully to confirm the high storage capacity by graphite nanofibers (slit pores) and SWCNT. Hynek et al investigated ten carbon sorbents but only one of them could augment the capacity of compressed hydrogen gas storage vessels. The improvement was marginal at 190 K and 300 K but nonexistent at 80 K. The storage capacity of carbon nanotubes and graphitic fibers has been enhanced by doping with lithium and other alkali elements . The alkali atoms seem to have a catalytic effect in dissociating the $`H_2`$ molecule and promoting atomic adsorption. An advantage is that the doped systems can operate at moderate temperatures and ambient pressure. Some of the authors cited above have also performed computer simulations of the adsorption of molecular hydrogen inside, outside and in the interstices of an array of SWCNT and in idealized carbon slit pores using model pair potentials to describe the interactions. Wang and Johnson adopted the semiempirical pair potential of Silvera and Goldman for the $`H_2H_2`$ interaction and the $`H_2C`$ interaction was modelled by a potential derived by Crowell and Brown by averaging an empirical Lennard-Jones $`H_2C`$ potential over a graphite plane. In the simulations Wang and Johnson used a hybrid path integral-Monte Carlo method. Johnson also studied the influence of electrical charging of the tubes. Stan and Cole performed calculations based on a sum of isotropic Lennard-Jones interactions between the molecule and the C atoms of the tube. They calculated the adsorption potential of a hydrogen molecule, considered as a spherically symmetric entity, as a function of distance from the axis of a SWCNT, along radial lines upon the center of an hexagon of carbon atoms and upon a carbon atom respectively. Those simulations give useful insight to interpret the results of the experiments. However the description of the interaction between $`H_2`$ and the graphitic surfaces of the SWCNT or the slit pores in those works is too simple. Simplicity is a necessary requirement for massive simulations involving several hundred (or several thousand) $`H_2`$ molecules and an assembly of SWCNT of realistic size, but one can expect more realistic results if the interaction potential is derived from an ab initio calculation. The adsorption of ”atomic” hydrogen on a planar graphene sheet, that is a planar layer exfoliated from graphite, has been studied previously . Bercu and Grecu used a semiempirical molecular orbital LCAO treatment at the INDO (intermediate neglect of differential overlap) unrestricted Hartree-Fock level and Jeloaica and Sidis used the Density Functional formalism (DFT) . In both works the description of the graphene layers was simplified by modelling this layer by a finite cluster C<sub>24</sub>-H<sub>12</sub>, where the hydrogen atoms saturate the dangling bonds on the periphery of the planar cluster. But, as mentioned above, hydrogen is adsorbed in molecular form by graphitic surfaces (SWCNT and slit pores), so in this work we study the interaction of an $`H_2`$ molecule with a planar graphene layer. Since the graphene layers interact weakly in bulk graphite, the interaction of $`H_2`$ with a graphitic surface is a localized phenomenon restricted to the outermost plane. For this reason our calculations have relevance for understanding the adsorption of $`H_2`$ on the walls of slit pores in graphite, and also for the case of adsorption by SWCNT, since these differ from a graphene layer only in the curvature of the layer. ## II THEORETICAL METHOD AND TESTs To calculate the interaction between $`H_2`$ and a planar graphene layer we use the ab initio fhi96md code, developed by Scheffler et al . This code uses the DFT to compute the electronic density and the total energy of the system, and we have chosen the local density approximation (LDA) for exchange and correlation. Only the two electrons of the $`H_2`$ molecule and the four external electrons ($`2s^2\mathrm{\hspace{0.17em}2}p^2`$) of each carbon atom are explicitly included in the calculations, while the $`1s^2`$ core of carbon is replaced by a pseudopotential. For this purpose we use the nonlocal norm-conserving pseudopotentials of Hamann et al . Nonlocality in the pseudopotential is restricted to $`\mathrm{}=2`$, and we take as local part of the pseudopotential the $`s`$ component. The code employs a supercell geometry and a basis of plane waves to expand the electronic wave functions. First we have tested the method for pure graphite. By minimization of the total energy with respect to the interatomic distances we obtained an in-plane $`CC`$ bond length equal to 2.66 a.u. and a distance between planar graphitic layers of 6.275 a.u. The corresponding experimental values are 2.68 a.u. and 6.34 a.u. respectively. The small ($`1\%`$) underestimation of bond lengths is characteristic of the LDA. Next we have studied an isolated graphene layer. Since the computer code uses a periodic supercell method, the cell axis has to be large in the z-direction to avoid the interaction between graphene sheets in different cells. Table I gives the calculated energy of the graphene layer as a function of the length c of the unit cell in the z direction, or in other words, as a function of the distance between parallel graphene layers. Results given for c = 20, 25, 30 and 35 a.u. show that the energy is well converged for those layer separations and that for c = 20 a.u. the error in the energy per atom is only about 1 in $`10^5`$. A cutoff of 40 Ry was used in all the calculations. We also tested the method by calculating the energy of the $`H_2`$ molecule, that was placed at the center of a simple cubic supercell. The total energy obtained for a plane wave cut-off energy of 40 Ry and supercell lattice constants of 18 a.u. and 20 a.u. is the same, -2.247 Ry as well as the bondlength, 1.48 a.u. Notice that this bond length is small compared to the $`CC`$ bond length. Anticipating the geometry to be used in the study of the interaction between $`H_2`$ and graphene, another set of calculations were performed for the energy of $`H_2`$ by placing the molecule in the superlattice described above in the study of the graphene layer, but this time without graphene. Calculations for distances between the imaginary graphene planes ranging from 20 a.u. to 35 a.u. (the plane wave cut-off was again 40 Ry) gave energies for the $`H_2`$ molecule, identical to the energies obtained for the cubic superlattice geometry. ## III INTERACTION BETWEEN $`H_2`$ AND THE GRAPHENE LAYER For the periodicity of the system we have selected a unit cell with eight carbon atoms and one hydrogen molecule (see Fig. 1). If we place a hydrogen molecule at any point of the cell, the distance from this molecule to other in the nearest cells is 9.224 a.u. This separation is large compared to the bond length of $`H_2`$ (1.480 a.u.), and we have verified that there is no interaction between two hydrogen molecules separated by that distance. The interaction of the $`H_2`$ molecule with the graphene sheet has been studied by performing static calculations for two orientations of the axis of the molecule: axis perpendicular to the graphene plane and axis parallel to that plane. Three possibles configurations, called A, B and C below, have been selected for the perpendicular approach of the molecule to the plane: (A) upon one carbon atom, (B) upon the center of a carbon-carbon bond, and (C) upon the center of an hexagon of carbon atoms. On the other hand, for the parallel approach the molecule is placed upon the center of an hexagon of carbon atoms with the molecular axis perpendicular to two parallel sides of the hexagon, and this is called configuration D. These four configurations are given in the bottom panel of Fig. 1. To obtain the interaction energy curve for each of those four cases, the distance between the hydrogen molecule and the graphene layer was varied while maintaining the relative configuration. In these calculations the bond length of the $`H_2`$ was held fixed at 1.48 a.u., the bondlength of the free molecule. This is expected to be valid in the relevant region of the interaction. This constraint will, however, be relaxed in simulations described at the end of this section. Calculations were first performed in the parallel configuration (D) for a superlattice such that the distance between graphene layers is 30 a.u. The plane wave cut-off was 40 Ry. The interaction energy curve is plotted in Figure 2 and the curve has a minimum at 5.07 a.u. For separations larger than this value the energy rises fast and reaches its asymptotic value for 10 - 11 a.u. The energy at the maximum possible separation between the center of mass of the $`H_2`$ molecule and the graphene plane for this superlattice, 15 a.u., was taken as the zero of energy. The figure also gives the results of a similar calculation for a smaller superlattice, such that the distance between graphene layers is 20 a.u. The corresponding energy curve, referred to the same zero of energy as above, is practically indistinguishable from the former curve. The calculations also show that for all practical purposes the energy curve has reached its asymptotic value for a distance of 10 a.u., that is the longest separation allowed for the superlattice of 20 a.u. This indicates that calculations using the smaller superlattice are enough for our purposes of studying the $`H_2`$ \- graphene interaction. Then, the results of calculations corresponding to configurations A, B, C and D for a superlattice of 20 a.u. are given in Figure 3. The potential energy curves for the perpendicular approach (A, B, C) rapidly merge with each other for large $`H_2`$ \- graphene separation, becoming indistinguishable from one another beyond 6.5 a.u. Actually, curves A and B are very close in the whole range of separations although B is marginally more attractive. The common value of the energy of curves A, B and C at separation 10 a.u. is taken as zero of energy in Figure 3. The predicted equilibrium positions and the binding energies (depth of the minimum) of the different curves are given in Table II. The small magnitude of the binding energies, less than 0.1 eV, shows that the system is in the physisorption regime. Comparison of the four curves reveals that the most favorable position for the $`H_2`$ molecule is physisorbed in a position above the center of a carbon hexagon, and that the parallel configuration is slightly more favorable than the perpendicular one. We have verified that different orientations of the molecular axis with respect to the underlying carbon hexagon in the parallel configuration lead, in all cases, to the same curve D plotted in Fig. 3. The differences in binding energy shown in Table II are very small. For instance, configuration D and A only differ by 16 meV, and configuration D and C by 3 meV. Figures 4a and 4b give the electron density of the pure graphene layer in two parallel planes, 5 and 3 a.u. above the plane of the nuclei, respectively. The former one is very close to the preferred distance of approach for the $`H_2`$ molecule in configuration D. First of all one can note that the values of the electron density in that plane are very small, of the order of $`10^5`$ e/$`(\mathrm{a}.\mathrm{u}.)^3`$, so the plane is in the tail region of the electron density distribution. Nevertheless the densities clearly reveal the topography of the graphene layer. Electron density contours on top of carbon atoms surround other contours representing the large hexagonal holes. Densities are larger in the other plane, closer to the plane of nuclei. In each plane the density is larger in the positions above carbon atoms and lower above the hexagons. A plot that complements this view is given in Figure 5, that gives the electron density in a plane perpendicular to the graphene layer through a line containing two adjacent carbon atoms, labelled $`C_1`$ and $`C_2`$ in the figure. Then, points labelled M and X represent the midpoint of a carbon-carbon bond and the center of an hexagon respectively. The most noticeable feature is the existence of depressions of electron density in the regions above the centers of carbon hexagons. These hollow regions are separated by regions of larger density that delineate the skeleton of carbon-carbon bonds. In this figure the density of the most external contour is $`\rho =1.11\times 10^2`$ e/$`(\mathrm{a}.\mathrm{u}.)^3`$ and the interval between contours $`\mathrm{\Delta }\rho =1.11\times 10^2`$ e/$`(\mathrm{a}.\mathrm{u}.)^3`$ These observations correlate with the features in Fig. 3, and lead to the following interpretation of the potential energy curves. Each curve can be seen as arising from two main contributions, one attractive and one repulsive. The attractive contribution is rather similar for all the configurations (notice the similarity of the potential energy curves beyond 6 a.u.) and is mainly due to exchange and correlation effects. Neglecting correlation for the purposes of simplicity, the exchange contribution to the total energy is given, in the LDA, by the functional $$E_x^{LDA}[\rho ]=C_x\rho (𝐫)^{\frac{4}{3}}\text{ }d^3r,$$ (1) where $`C_x`$ is a well known negative constant. In the regime of weakly overlapping densities, and assuming no density rearrangements due to the close-shell character of $`H_2`$, the contribution of exchange to the interaction energy becomes $$\mathrm{\Delta }E_x=C_x[[\rho _{H_2}(𝐫)+\rho _g(𝐫)]^{\frac{4}{3}}d^3r\rho _{H_2}(𝐫)^{\frac{4}{3}}d^3r\rho _g(𝐫)^{\frac{4}{3}}d^3r],$$ (2) where $`\rho _g`$ and $`\rho _{H_2}`$ represent the tail densities of the graphene layer and $`H_2`$ molecule respectively. A net ”bonding” contribution arises from the nonlinearity of the exchange energy functional. On the other hand the sharp repulsive wall is due to the short-range repulsion between the close electronic shell of the $`H_2`$ molecule and the electron gas of the substrate. This contribution is very sensitive to the local electron density sampled by the $`H_2`$ molecule in its approach to the graphene layer and explains the correlation between the position and depth of the different minima in Fig. 3 and the features of the substrate electron density in Figs. 4 and 5. Similar arguments explain the physisorption of noble gas atoms on metallic surfaces and the weak bonding interaction between noble gases. At very large separation the interaction energy curves should approach the Van der Waals interaction, that is not well described, however, by the LDA. An interesting point concerns the comparison of the minima of the curves C and D of Fig. 3. That of curve D is deeper and occurs at a shorter $`H_2`$ \- graphene separation. The reason is that the surfaces of constant electron density of the $`H_2`$ molecule have the shape of slightly prolate ellipsoids instead of simple spheres. Consequently, for a given distance d between the center of mass of $`H_2`$ and the graphene plane, the molecule with the perpendicular orientation (C) penetrates more deeply into the electronic cloud of the substrate than in the parallel orientation (D). In other words, the repulsive wall is reached earlier, that is for larger d, in the perpendicular configuration (C). If we consider an electronic density contour in $`H_2`$ with a value $`\rho =0.018`$ e/$`(\mathrm{a}.\mathrm{u}.)^3`$, then the two semiaxes have lengths of 2.07 and 1.71 a.u. respectively and the difference between these two lengths is 0.36 a.u. This value is in qualitative agrement with the difference between the $`H_2`$ \- graphene separations for the two minima of curves C and D, which is 0.20 a.u. This shape effect is usually neglected in the phenomenological approaches, that treat $`H_2`$ simply as a spherical molecule. Figure 6 gives a plot of the charge density difference $$\rho _{diff}(𝐫)=\rho _{tot}(𝐫)(\rho _g(𝐫)+\rho _{H_2}(𝐫)),$$ (3) where $`\rho _{tot}(𝐫)`$ is the calculated density of the total system, that is the $`H_2`$ molecule physisorbed in orientation D at a distance of 5 a.u. above the graphene layer, whereas $`\rho _g`$ \+ $`\rho _{H_2}`$ is the simple superposition of the densities of the pure graphene layer and $`H_2`$ molecule placed also in orientation D, 5 a.u. above the graphene layer. That density difference $`\rho _{diff}(𝐫)`$ is given in the same plane, perpendicular to the graphene layer, used in Fig. 5. $`\rho _{diff}(𝐫)`$ has positive and negative regions. The positive region is the area bound by the contour labelled P. This region has the shape of two lobes joined by a narrow neck. Contour P has a value $`\rho _{diff}=2.36\times 10^5`$ e/$`(\mathrm{a}.\mathrm{u}.)^3`$ and $`\rho _{diff}`$ increases in this positive region as we move towards inner contours in the lobes. The innermost contour shown has a value $`\rho _{diff}=2.87\times 10^4`$ e/$`(\mathrm{a}.\mathrm{u}.)^3.`$ The $`H_2`$ molecule sits above the neck, so the figure reveals that the repulsive interaction produced by the close electronic shell of $`H_2`$ pushes some charge from the region immediately below the molecule (the neck region) to form the lobes of positive $`\rho _{diff}(𝐫).`$ This displacement of electronic charge is nevertheless quantitatively very small. Notice that $`\rho _g`$ takes values between $`1.6\times 10^3`$ and $`4.1\times 10^3`$ e/$`(\mathrm{a}.\mathrm{u}.)^3`$ in a plane 3 a.u. above the graphene layer, while $`\rho _{diff}`$ has values of the order $`10^510^4`$ e/$`(\mathrm{a}.\mathrm{u}.)^3`$ in the same plane. The smallness of $`\rho _{diff}`$ justifies the argument given in eq. (2) for the attractive exchange-correlation contribution to the interaction potential. The static calculations discussed above have been complemented with dynamical simulations in which the $`H_2`$ molecule was initially placed in different orientations at distances of 4 - 6 a.u. from the graphene layer and was left to evolve under the influence of the forces on the H atoms. The $`H_2`$ bondlength was allowed to adjust in the process. The simulations confirm the results of the static calculations, in the sense that the $`H_2`$ molecules end up in positions above the center of an hexagon at the end of the simulations. The binding energies and $`H_2`$ \- graphene layer distances practically coincide with those in Table II. Marginally small differences in separation or binding energy are due to very small changes of the bondlength of $`H_2`$, always smaller than $`0.3\%`$. The result of one of the simulations is worth to be mentioned. A configuration intermediate between those labelled C and D above was obtained: the center of mass of the molecule was 5.10 a.u. above the center of a carbon hexagon, with the molecular axis forming an angle of about $`30^{}`$ with the graphene plane. The binding energy in this new configuration was only 1 meV larger than in the parallel configuration D. In summary, the picture arising from the calculations is rather clear. The $`H_2`$ molecules prefer the hollow sites above the centers of carbon hexagons where the background electron density is lower than in channels on top of the skeleton of carbon-carbon bonds. The exchange-correlation contribution provides the weak attraction responsible for physisorption, but the preferred distance of approach is determined by the repulsive part of the interaction potential. That repulsive contribution is due to the close-shell electronic structure of $`H_2`$. We have performed static calculations of the barrier for the diffusion of a molecule, initially in the parallel configuration D at the preferred distance of 5.07 a.u. above the graphene plane, to an equivalent configuration D above an adjacent hexagon. The initial configuration of the molecule, with its axis perpendicular to two parallel carbon-carbon bonds, can be seen in the bottom panel of Fig. 1. The molecule was then forced to follow a path across one of those bonds, allowing for the reorientation of the molecular axis at each step in order to minimize the energy of the system. Although the molecule begins with the axis parallel to the graphene plane, the orientation of the axis changes as the molecule approaches the carbon-carbon bond. In fact, when the center of mass of the molecule is precisely above that bond, the molecular axis becomes perpendicular to the graphene plane, that is the molecule adopts configuration B, as indicated also in Fig. 1. The energy difference between this saddle configuration and the starting one gives a calculated diffusional barrier of 14 meV. A temperature of 163 K is enough to surpass this barrier. The conclusions from the calculations are, in our view, general enough that one can make some extrapolations to the case of adsorption of $`H_2`$ by carbon nanotubes. When adsorption occurs on the outside wall of an isolated nanotube, the predictions of Fig. 3 will be valid, with a minor influence of the nanotube curvature. If the tubes form a parallel bundle and we consider the interstitial channels between tubes, the effects seen in Figure 3 will be smoothed out because of the addition of contributions of different graphitic surfaces non in registry. Addition of these contributions will give rise to an interstitial channel with a potential energy nearly independent of z, if we call z the direction parallel to the tube axis. Finally, the same smoothing effect will occur in the inner channel of a tube if the tube diameter is not large. In summary we predict very easy diffusion of the $`H_2`$ molecule in arrangements of parallel tubes along the direction parallel to the tube axis, both inside the tube cavity and in the interstitial channels. The present adsorption results can be partialy compared with those of Stan and Cole. They considered the $`H_2`$ molecule as a spherically symmetric entity and calculated the adsorption potential inside zigzag (13,0) nanotubes (radius $`=`$ 9.62 a.u.) based on a sum of isotropic Lennard-Jones interactions between the molecule and the carbon atoms of the tube. Our calculation and that of Stan and Cole agree in that the smallest binding energy is obtained for the $`H_2`$ upon one carbon atom and the largest one for the $`H_2`$ upon the center of the hexagon of carbon atoms. However Stan and Cole do not distinguish between parallel and perpendicular orientations because they considered an spherical molecule. Their Fig. 1 shows a binding energy about 0.079 eV for adsorption in front of the center of an hexagon of carbon atoms and that the equilibrium distance between the molecule and the nanotube wall is 5.7 a.u. This distance is consistent but a little larger than those reported in our Table II. On the other hand, the value 0.079 eV for the binding energy is also consistent with the binding energies in Table II. Notice, however, that the binding energy for a tube of larger radius, or for a planar graphene sheet, will be a little smaller because the curvature of the tube increases the number of nearest neighbor carbon atoms. In fact, Wang and Johnson calculated an adsorption binding energy near 0.050 eV for molecular hydrogen in an idealized carbon slit pore with a pore width of 17.4 a.u. ## IV CONCLUSIONS By performing DFT calculations we confirm that physisorption of $`H_2`$ on graphitic layers is possible. The differences between the binding energies corresponding to different positions (on top of carbon atoms, on top of carbon-carbon bonds, on top of hexagonal holes) are small, and the diffusional barriers are also small, so easy diffusion is expected at low temperature. The nonsphericity of the $`H_2`$ molecule has some influence on the preferred orientation of the molecular axis with respect to the graphene plane. These small effects associated to different positions and orientations of the physisorbed molecule are expected to average out inside carbon nanotubes or in the interstitial channels in parallel arrays of carbon nanotubes. ## acknowledgements Work supported by DGES(Grant PB95-0720-C02-01), Junta de Castilla y León (Grant VA28/99) and European Community (TMR Contract ERBFMRX-CT96-0062-DG12-MIHT). L.M.M. is greatful to DGES for a Predoctoral Grant. J.S.A. wishes to thank the hospitality of Universidad de Valladolid during his sabbatical leave and grants given by Universidad Autónoma Metropolitana Azcapotzalco and by Instituto Politécnico Nacional (México). Finally we thank the referee for constructive suggestions.
warning/0002/hep-th0002039.html
ar5iv
text
# 1 Introduction ## 1 Introduction The AdS/CFT correspondence relates string theories on anti-de Sitter space ($`\text{AdS}_{d+1}`$) backgrounds to $`d`$-dimensional conformal field theories (CFTs) . One simple example of this correspondence, which is the only example treated in this paper, is type IIB string theory on $`\text{AdS}_5\times \text{S}^5`$ and $`𝒩=4,d=4`$ SU(N) Super Yang Mills (SYM) theory. In the large N limit, with $`g_{YM}^2N`$ held fixed and very large, the supergravity (SUGRA) approximation of type IIB string theory is valid, thus providing a perturbative way of understanding SYM theory at strong coupling. Correlation functions provide an important way of studying the correspondence. Calculations of 2- and 3-point functions have already provided evidence that the correspondence is correct , but 4-point functions, as their form is not completely fixed by conformal invariance, can provide more detailed information about the CFT at strong coupling. Previous studies of 4-point correlators which were peripherally useful in preparing this paper include -. Recently in , for type IIB SUGRA on $`\text{AdS}_5\times \text{S}^5`$, the first realistic 4-point functions $$\varphi (x_1)\varphi (x_2)\varphi (x_3)\varphi (x_4),C(x_1)C(x_2)C(x_3)C(x_4),$$ and $$\varphi (x_1)C(x_2)\varphi (x_3)C(x_4)$$ were calculated.<sup>1</sup><sup>1</sup>1 These results were elaborated in where the original calculations were simplified and in where the logarithmic singularities were explained. The operators $`\varphi `$ and $`C`$ correspond to the dilaton and axion supergravity fields. As was pointed out in , the dilaton and axion fields correspond to the operators $`\varphi Tr(F^2+\mathrm{})`$ and $`CTr(F\stackrel{~}{F}+\mathrm{})`$ in $`𝒩=4`$ SYM theory. We attempt to expand upon the results in by considering the corresponding CFT 4-point functions. To make contact with CFT, it is convenient to expand these AdS 4-point functions as a power series in $`x_{13}^2`$, $`x_{24}^2`$, and $`x_{13}x_{24}`$ in the “direct” or t-channel limit $`|x_{13}|,|x_{24}||x_{12}|`$.<sup>2</sup><sup>2</sup>2 Let $`x_{ij}x_ix_j`$. The power law singular terms in this series are identical for all three 4-point functions and come solely from graviton exchange. Because of the proposed AdS/CFT correspondence, we expect to be able to reconstruct the AdS 4-point amplitudes in terms of a double operator product expansion (OPE) in CFT. We express $`\varphi (x)\varphi (y)`$ and $`C(x)C(y)`$ in terms of their OPEs. Multiplying two such OPEs together and taking the 2-point functions of operators in the expansion should recover the scattering amplitude. In the limit where $`|x_{13}|`$ and $`|x_{24}|`$ are very small compared to $`|x_{12}|`$, we expect to be able to represent the 4-point amplitude schematically as $$𝒪_1(x_1)𝒪_2(x_2)𝒪_3(x_3)𝒪_4(x_4)=\underset{n,m}{}\frac{\alpha _n𝒪_n(X_{13})𝒪_m(X_{24})\beta _m}{x_{13}^{\mathrm{\Delta }_1+\mathrm{\Delta }_3\mathrm{\Delta }_m}x_{24}^{\mathrm{\Delta }_2+\mathrm{\Delta }_4\mathrm{\Delta }_n}}$$ (1) where $`𝒪_p`$ is some operator of dimension $`\mathrm{\Delta }_p`$. We have defined $`X_{ij}(x_i+x_j)/2`$. The expression is schematic because in general the operator $`𝒪_p`$ may be a tensor. Because the leading order terms in the three 4-point functions come from graviton exchange and because of the proposed AdS/CFT correspondence, we expect and it was shown in that the leading order term in (1) comes from the 2-point function of the energy momentum tensor with itself, $`T_{ab}(X_{13})T_{cd}(X_{24})`$. In this paper, we go further and show precisely how, from a CFT point of view, all singular terms in the t-channel limit of the three 4-point functions arise from exchange of $`T_{ab}`$ and its descendants. We also investigate how the 4-point functions change as we move from strong to weak coupling. The work proceeds in four parts. First we look at the leading order singular terms in the AdS 4-point functions in the t-channel limit, the same terms that in the next section we will be able to compute using conformal invariance. In the third and fourth sections, we compute the equivalent 4-point function in the weak coupling limit of the CFT, which is essentially the case of electricity and magnetism, in order to try to understand how the 4-point function changes as we move from strong to weak coupling. This investigation will reveal the existence of a new nonchiral primary operator in the weak limit which, if the AdS/CFT correspondence is to hold, acquires a large anomalous dimension as we move to strong coupling and hence does not contribute to the AdS singular terms. ## 2 AdS 4-point Functions It turns out that in the t-channel limit, all three of the AdS 4-point functions calculated in have the same singular power law terms. Moreover, the singular terms come only from t-channel graviton exchange. The singular terms in the amplitude are $$I_{\text{grav}}|_{\text{sing}}=\frac{2^{10}}{35\pi ^6}\frac{1}{x_{13}^8x_{24}^8}\left[s(7t^2+6t^4)+s^2(7+3t^2)8s^3\right].$$ (2) The variables $`s`$ and $`t`$ are conformally invariant functions of the $`x`$’s: $`s`$ $``$ $`{\displaystyle \frac{1}{2}}{\displaystyle \frac{x_{13}^2x_{24}^2}{x_{12}^2x_{34}^2+x_{14}^2x_{23}^2}},`$ (3) $`t`$ $``$ $`{\displaystyle \frac{x_{12}^2x_{34}^2x_{14}^2x_{23}^2}{x_{12}^2x_{34}^2+x_{14}^2x_{23}^2}}.`$ (4) To make contact with (1), we can expand $`I_{\text{grav}}`$ in one of two ways. We can carry out an asymmetric expansion in powers of $`x_{12}`$ or we can perform a symmetric expansion in powers of $`w(X_{13}X_{24})`$. We consider only the symmetric expansion because terms odd in powers of $`w`$ will not appear: $`s`$ $`=`$ $`{\displaystyle \frac{x_{13}^2x_{24}^2}{4w^4}}{\displaystyle \frac{1}{g(w,x_{13},x_{24})}},`$ $`t`$ $`=`$ $`{\displaystyle \frac{x_{13}J(w)x_{24}}{w^2}}\left[1+{\displaystyle \frac{1}{4}}{\displaystyle \frac{x_{13}x_{24}}{w^2}}{\displaystyle \frac{x_{13}^2+x_{24}^2}{x_{13}J(w)x_{24}}}\right]{\displaystyle \frac{1}{g(w,x_{13},x_{24})}}.`$ In the above, $`J_{ij}=\delta _{ij}2x_ix_j/x^2`$ is the Jacobian tensor of the conformal inversion $`x_i^{}=x_i/x^2`$, and we have defined $`g(w,x_{13},x_{24})`$ $``$ $`[1+{\displaystyle \frac{1}{2w^2}}(x_{13}^2+x_{24}^2){\displaystyle \frac{1}{w^4}}((wx_{13})^2+(wx_{24})^2)+`$ $`+{\displaystyle \frac{1}{16w^4}}(x_{13}^4+x_{24}^4+2x_{13}^2x_{24}^2+4(x_{13}x_{24})^2)].`$ Armed with these expressions for $`s`$ and $`t`$, we can expand $`I_{\text{grav}}`$ to the third nontrivial order in $`w`$, i.e. we consider terms of order $`w^n`$ where $`n=8,\mathrm{\hspace{0.17em}10},\text{and}\mathrm{\hspace{0.17em}12}`$. The amplitude can be written order by order as $$I_{\text{grav}}|_{8^{\text{th}}}=\frac{2^6}{5\pi ^6}\frac{1}{x_{13}^6x_{24}^6w^8}\left[4(x_{13}J(w)x_{24})^2x_{13}^2x_{24}^2\right],$$ (5) $`I_{\text{grav}}|_{10^{\text{th}}}={\displaystyle \frac{2^6}{5\pi ^6}}{\displaystyle \frac{1}{x_{13}^6x_{24}^6w^{10}}}[`$ (6) $`6(x_{13}J(w)x_{24})^2\left[(x_{13}J(w)x_{13})+(x_{24}J(w)x_{24})\right]+`$ $`+2x_{13}x_{24}(x_{13}^2+x_{24}^2)(x_{13}J(w)x_{24})+`$ $`+x_{13}^2x_{24}^2[(x_{13}J(w)x_{13})+(x_{24}J(w)x_{24})]],`$ $`I_{\text{grav}}|_{12^{\text{th}}\text{, asym}}={\displaystyle \frac{2^6}{5\pi ^6}}{\displaystyle \frac{1}{x_{13}^6x_{24}^6w^{12}}}[`$ (7) $`6(x_{13}J(w)x_{13})^2(x_{13}J(w)x_{24})^2+`$ $`3x_{13}^2(x_{13}x_{24})(x_{13}J(w)x_{13})(x_{13}J(w)x_{24})+`$ $`{\displaystyle \frac{3}{4}}x_{13}^4(x_{13}J(w)x_{24})^2{\displaystyle \frac{3}{4}}x_{13}^2x_{24}^2(x_{13}J(w)x_{13})^2+`$ $`+{\displaystyle \frac{1}{4}}(x_{13}x_{24})^2x_{13}^4+{\displaystyle \frac{1}{8}}x_{13}^6x_{24}^2+(x_{13}x_{24})],`$ $`I_{\text{grav}}|_{12^{\text{th}}\text{, sym}}={\displaystyle \frac{2^6}{5\pi ^6}}{\displaystyle \frac{1}{x_{13}^6x_{24}^6w^{12}}}[`$ (8) $`+{\displaystyle \frac{24}{7}}(x_{13}J(w)x_{24})^4+12(x_{13}J(w)x_{24})^2(x_{13}J(w)x_{13})(x_{24}J(w)x_{24})+`$ $`{\displaystyle \frac{3}{2}}x_{13}^2x_{24}^2(x_{13}J(w)x_{13})(x_{24}J(w)x_{24})+`$ $`3x_{13}x_{24}(x_{13}J(w)x_{24})\left[x_{13}^2(x_{24}J(w)x_{24})+x_{24}^2(x_{13}J(w)x_{13})\right]+`$ $`3(x_{13}J(w)x_{24})^2({\displaystyle \frac{5}{14}}x_{13}^2x_{24}^2+(x_{13}x_{24})^2)+`$ $`+x_{13}^2x_{24}^2(x_{13}x_{24})^2{\displaystyle \frac{1}{28}}x_{13}^4x_{24}^4].`$ The abbreviations sym and asym in the previous two equations symbolize that we have split the twelfth order term into pieces with equal and unequal numbers of $`x_{13}`$ and $`x_{24}`$ respectively. As these two pieces come from different 2-point functions in the double OPE (1), this separation will be important. Note that the $`st^4`$, $`s^2t^2`$, and $`s^3`$ terms contribute only to the symmetric twelfth order term, (8). ## 3 A General CFT As is well known, conformal invariance alone does not completely specify the form of a 4-point function. For two pairs of scalars of dimension four, the most we can say is that $$𝒪^{}(x_1)𝒪(x_2)𝒪^{}(x_3)𝒪(x_4)=\frac{1}{x_{13}^8x_{24}^8}F(s,t)$$ where $`F`$ is some unknown function of the conformally invariant variables $`s`$ and $`t`$. If we know the primary operators that occur in the OPEs of $`𝒪O`$ and $`𝒪^{}O^{}`$ along with their coefficients, then we can specify $`F`$ completely. As was shown in , the assumption that $`T_{ab}`$ appears in the OPE reproduces the leading order term in the 4-point function in the t-channel limit, (5). We will show that this assumption actually reproduces all the terms in (2). In the CFT literature, equations have been derived that describe the conformal block contribution, up to an overall normalization factor, of an arbitrary tensor primary operator exchanged in a 4-point interaction of four arbitrary scalars. We have tried unsuccessfully to use Eq. 3.15 of and suspect there may be some normalization problem. A similar equation can be found in but appears to be more difficult to apply and was discovered only after the following work had been completed. We shall use a brute force approach that has the advantage of showing us precisely how $`T_{ab}`$ and its descendants arise in the OPE of $`𝒪O`$. Our approach is similar to methods for calculating 4-point functions that can be found in the CFT literature . We may write the symmetric OPE schematically as $`𝒪\left({\displaystyle \frac{x}{2}}\right)𝒪\left({\displaystyle \frac{x}{2}}\right)`$ (9) $`A{\displaystyle \frac{x_ax_b}{x^6}}\left[T_{ab}(0){\displaystyle \frac{1}{2}}x_ix_jT_{abij}^{(2)}(0)+{\displaystyle \frac{1}{24}}x_ix_jx_kx_lT_{abijkl}^{(4)}(0)+\mathrm{}\right]`$ where $`A`$ is an overall constant and $`T_{abij}^{(2)}`$ and $`T_{abijkl}^{(4)}`$ are second and fourth order descendants of $`T_{ab}`$. From conformal invariance, we know that the descendants can be expressed as derivatives of $`T_{ab}`$. More specifically, we can write the second order descendant as $$T_{abij}^{(2)}(x)=\mu _i_jT_{ab}(x)+\nu \delta _{ij}\mathrm{}T_{ab}(x),$$ (10) and the fourth order descendant can be written correspondingly in terms of fourth order derivatives of $`T_{ab}`$. Terms of odd order in $`x`$ would be inconsistent with the symmetry of the expansion. As a warm up, we consider the contribution of $`T_{ab}`$ alone to the 4-point function. The leading nontrivial term in the OPE of two scalars is by assumption $`T_{ab}`$ and as $`T_{ab}`$ has dimension 4, the two point function of $`T_{ab}`$ with itself that appears inside the 4-point function must have a $`w^8`$ in the denominator. In other words, the leading term in the power series $`F`$ must be of the form $`\alpha s^2+\beta st^2+\gamma t^4`$. In fact we can do better. From conformal invariance (see for example ), we know that the 2-point function of $`T_{ab}`$ with itself must be $$T_{ab}(x_1)T_{cd}(x_2)=\frac{C_T}{x_{12}^8}[J_{ac}(x_{12})J_{bd}(x_{12})+J_{ad}(x_{12})J_{bc}(x_{12})\frac{2}{d}\delta _{ab}\delta _{cd}].$$ (11) We can at this stage show agreement between the CFT and the AdS 4-point functions at leading order, as was done in . Using (9) and (11), we see that (1) agrees with the leading order term (5) provided $`A^2C_T=2^7/(5\pi ^6)`$. Another way of understanding this calculation is to say that (11) is consistent with $`F`$ only if $`\alpha =\beta `$ and if $`\gamma =0`$. Note that in $`F`$, terms odd in powers of $`t`$ are not allowed as they are not consistent with the structure of the symmetric OPE of two scalars. Therefore, next order terms will be of the form $`s^3`$, $`s^2t^2`$, $`st^4`$, and possibly $`t^6`$. To proceed further in our calculation of the double OPE, note that one class of 2-point functions that need to be calculated involve $`T_{ab}`$ in one OPE with $`T_{ab}`$ and any of its descendants in the other OPE. We can obtain this entire class easily from or , where the authors use conformal invariance to show that the 3-point function of two scalars with $`T_{ab}`$ must take the form $$T_{ab}(x_1)𝒪(x_2)𝒪(x_3)=\frac{a}{x_{12}^dx_{13}^dx_{23}^{2\eta d}}t_{ab}(X_{23})$$ where $$t_{ab}(X)=\left(\frac{X_aX_b}{X^2}\frac{1}{d}\delta _{ab}\right)$$ and where $`X_{23}`$ $`=`$ $`{\displaystyle \frac{x_{21}}{x_{21}^2}}{\displaystyle \frac{x_{31}}{x_{31}^2}},`$ $`X_{23}^2`$ $`=`$ $`{\displaystyle \frac{x_{23}^2}{x_{21}^2x_{31}^2}}.`$ In the above expression, $`a`$ is an as yet undetermined constant, $`d`$ is the dimension of space, and $`\eta `$ is the dimension of $`𝒪`$. To read off the 2-point functions of interest, we consider the limit $`x_{23}0`$, and we expand the three point function in the variables $`x=x_{23}`$ and $`y=(x_{12}+x_{13})/2`$. In the case where $`d=4=\eta `$, the resulting somewhat cumbersome expression for the 3-point function is $`T_{ab}(x_1)𝒪(x_2)𝒪(x_3)={\displaystyle \frac{a}{x^6y^{12}}}\left(1+{\displaystyle \frac{x^2}{2y^2}}{\displaystyle \frac{(xy)^2}{y^4}}+{\displaystyle \frac{x^4}{16y^4}}\right)^3`$ (12) $`[4(xy)^2y_ay_bxy(2y^2+{\displaystyle \frac{1}{2}}x^2)(x_ay_b+x_by_a)+`$ $`+(y^4+{\displaystyle \frac{1}{2}}x^2y^2+{\displaystyle \frac{1}{16}}x^4)x_ax_b+`$ $`{\displaystyle \frac{1}{4}}\delta _{ab}x^2(y^4+{\displaystyle \frac{1}{2}}x^2y^2(xy)^2+{\displaystyle \frac{1}{16}}x^4)]`$ Now the lowest order term in the above expression corresponds to the 2-point function of $`T_{ab}`$ with itself, which was discussed previously. As a check on the calculations so far, one may verify that (11) is completely consistent with the highest order term in (12) if $`AC_T=a/2`$. The second order term in (12) corresponds to the 2-point function of $`T_{ab}`$ with the descendant operator $`T_{abij}^{(2)}`$. One finds that up to permutations of the indices $`(abij)`$, the two point function takes the form $`T_{abij}^{(2)}(x)T_{cd}(0)`$ $`=`$ $`{\displaystyle \frac{2C_T}{x^{10}}}[3J_{ac}J_{bd}J_{ij}+`$ (13) $`+{\displaystyle \frac{1}{2}}\delta _{ij}(\delta _{ac}J_{bd}+\delta _{bd}J_{ac})+{\displaystyle \frac{1}{2}}\delta _{ab}\delta _{cd}J_{ij}]`$ where all the $`J`$ take $`x`$ as an argument. Given the same condition on $`A`$ and $`C_T`$ as above, we have agreement between (12) and the higher order term (6) in the 4-point function. The third order term in (12) corresponds to the 2-point function of $`T_{ab}`$ with the descendant operator $`T_{abijkl}^{(4)}`$. We have indeed checked that this third order term agrees with the asymmetric term (7) in the 4-point function given the same conditions on $`A`$ and $`C_T`$. Note that (13) is consistent with (10) only if $`\mu =1/28`$ and $`\nu =1/28`$. Calculating the 2-point function of $`T_{abij}^{(2)}`$ with itself then becomes a simple matter of taking derivatives of (11). We have checked that $`T_{abij}^{(2)}T_{cdkl}^{(2)}`$ agrees with the symmetric term (8) in the 4-point function. This calculation fixes the coefficients of $`st^4`$, $`s^2t^2`$, and $`s^3`$ and also shows that $`t^6`$ does not appear in the singular terms. There are other types of expansions one could use to compare the AdS results with CFT. For example, one could have taken a limit in which only two scalars approach one another. Then, instead of 2-point functions, one considers the set of 3-point functions involving the two other scalars and the operators in the OPE of the two neighboring scalars. We have followed the lead of and used the double OPE method. ## 4 $`𝒩=4`$ SYM at Weak Coupling Conformal invariance implies that the coordinate dependence of the 2- and 3-point functions in AdS/CFT correspondence does not change as we move from weak to strong coupling. In addition, nonrenormalization theorems are thought to keep the coefficients of 2- and 3-point correlation functions involving chiral primaries independent of coupling . However, the coordinate dependence of 4-point functions can and does change as we vary the coupling. Thus a calculation and comparison of 4-point functions in the two coupling regimes is likely to be a much more enlightening way of seeing how the theory changes as we move from weak to strong coupling. So far, we have only looked at the strong coupling regime. In this section, we begin a consideration of $`𝒩=4`$ SYM theory in the weak coupling limit which will culminate in the next section with a computation of the connected dilaton 4-point function at weak coupling to leading order in $`\lambda =g_{YM}^2N`$. Essentially, this correlation function is equivalent to the 4-point function of $`F^2`$ in electricity and magnetism as the difference between the two only appears at subleading order in $`\lambda `$. To be more specific, the dilaton and axion operators take the form $`\varphi Tr(F^2+\mathrm{})`$ and $`CTr(F\stackrel{~}{F}+\mathrm{})`$.<sup>3</sup><sup>3</sup>3 To be completely precise, $$\stackrel{~}{F}_{ab}=\frac{1}{2}ϵ_{abcd}F_{cd}.$$ As shown in in the case of the dilaton, the higher order terms will involve three or more of the operators $`F_{ab}^{kl},X_a^{kl}`$, and $`\mathrm{\Theta }_\alpha ^{kl}`$.<sup>4</sup><sup>4</sup>4$`k,l,\mathrm{}`$ are SU(N) indices, $`a,b,\mathrm{}`$ are spatial indices, and $`\alpha ,\beta ,\mathrm{}`$ are spinor indices. The 2-point function for $`F_{ab}`$ is: $$F_{ab}^{kl}(x_1)F_{cd}^{mn}(x_2)=\frac{c}{x_{12}^4}\delta ^{kn}\delta ^{lm}[J_{ac}(x_{12})J_{bd}(x_{12})J_{ad}(x_{12})J_{bc}(x_{12})]$$ (14) where $`J_{ab}`$ is as defined above and $`cg_{YM}^2`$ is a constant. In general, the two point function of an operator with itself will contain these same Kronecker delta functions of the SU(N) indices. From this fact and Wick’s Theorem, it is not difficult to see that the higher order terms in $`\varphi `$ and $`C`$ involving three or more operators produce corrections to the correlation functions which are higher order in $`g_{YM}^2N`$. From hereon, we suppress the SU(N) indices and consider only the leading order terms in $`\varphi `$ and $`C`$. As described in the introduction, scattering in the t-channel limit can be represented in terms of a double OPE. Thus, first we express $`\varphi (x)\varphi (y)`$ and $`C(x)C(y)`$ in terms of their OPEs. It turns out that up to terms with no contractions, the dilaton and axion have the same OPE. We present first an intermediate result: $`F^2(x)F^2(0)F\stackrel{~}{F}(x)F\stackrel{~}{F}(0)`$ (15) $`{\displaystyle \frac{48c^2}{x^8}}{\displaystyle \frac{32cx_ax_b}{x^6}}\left[F_{ac}(x)F_{bc}(0){\displaystyle \frac{1}{4}}\delta _{ab}F_{cd}(x)F_{cd}(0)\right].`$ This expression is reminiscent of the energy momentum tensor which takes the form $$T_{ab}(x)=K[F_{ac}F_{bc}\frac{1}{4}\delta _{ab}F^2](x)$$ where $`K`$ another constant.<sup>5</sup><sup>5</sup>5 The constants in this section are related to those in the previous section by $`C_T=2c^2K^2`$ and $`A=32c/K`$. Thus, the symmetric OPE of the dilaton and axion can be written as $`F^2\left({\displaystyle \frac{x}{2}}\right)F^2\left({\displaystyle \frac{x}{2}}\right){\displaystyle \frac{48c^2}{x^8}}+`$ (16) $`{\displaystyle \frac{32cx_ax_b}{Kx^6}}\left[T_{ab}(0){\displaystyle \frac{1}{2}}x_ix_j(T_{abij}^{(2)}(0)+P_{abij}(0))\right]+\mathrm{}`$ where the second order descendant $`T_{abij}^{(2)}`$ takes the same form as in (10) and we have found potentially a new primary with the complicated form $$P_{abij}(x)\frac{3}{14}_i_jT_{ab}(x)\frac{1}{28}\delta _{ij}\mathrm{}T_{ab}(x)+T_{abij}^{}(x)$$ (17) where $$T_{abij}^{}(x)K\left[(_iF_{ac})(_jF_{bc})(x)\frac{1}{4}\delta _{ab}(_iF_{cd})(_jF_{cd})(x)\right].$$ Using Wick’s Theorem, we have checked that the 2-point function of $`P_{abij}`$ with $`T_{ab}`$ and with $`F^2`$ is zero. In the next section, we will see from evaluating the dilaton 4-point function, that the 2-point function of $`P_{abij}`$ with itself is nonzero, and therefore that $`P_{abij}`$ does not vanish by the equations of motion. To show definitively that $`P_{abij}`$ is a primary operator, it would be nice to demonstrate that it transforms appropriately under the conformal group and more specifically under inversion. Preliminary results suggest that $`P_{abij}`$ does transform as a primary under inversion, but the full calculation is lengthy and has not been completed. From the index structure, it is clear that $`P_{abij}`$ has spin four, and at least in the weak coupling regime, the dimension of this new primary operator is six. If the operator $`P_{abij}`$ were a chiral primary, its dimension would be algebraically protected, and the operator should contribute to the 4-point functions equally at weak and strong coupling. However, as we have seen, $`P_{abij}`$ does not contribute to the leading singular terms at strong coupling and, as we will see in the next section, $`P_{abij}`$ does contribute at weak coupling. Moreover, our new primary does not correspond to any of the known chiral primaries of dimension six. The logical conclusion is that $`P_{abij}`$ is nonchiral and acquires a large anomalous dimension in the strong coupling regime: That there seem to be no nonchiral fields on AdS space with a mass below the string scale suggests that our nonchiral primary has a dimension which grows at least as fast as $`(g_{YM}^2N)^{1/4}`$ in the strong coupling limit, $`g_{YM}^2N\mathrm{}`$ . ## 5 The Dilaton 4-point Function in the Weak Coupling Limit of SYM We calculate the four dilaton amplitude in the weak coupling limit of $`𝒩=4,d=4`$ SYM theory, which as noted in the previous section, is essentially equivalent, at leading order in $`g_{YM}^2N`$, to the 4-point function of $`F^2`$ of electricity and magnetism, i.e. $$M_4F^2(x_1)F^2(x_2)F^2(x_3)F^2(x_4).$$ Let $$W_{1324}F_{ab}(x_1)F_{cd}(x_3)F_{cd}(x_3)F_{ef}(x_2)F_{ef}(x_2)F_{gh}(x_4)F_{gh}(x_4)F_{ab}(x_1)$$ and let $$W_{12}F_{ab}(x_1)F_{cd}(x_2)F_{cd}(x_2)F_{ab}(x_1)=\frac{24c^2}{x_{12}^8}.$$ Then $$M_4=16(W_{1234}+W_{1324}+W_{1342})+4(W_{12}W_{34}+W_{13}W_{24}+W_{14}W_{23}).$$ The terms containing $`W_{ab}`$ do not concern us as they describe the disconnected pieces of the 4-point function. Note that by definition $`W_{abcd}=W_{adcb}`$. We proceed with a calculation of the amplitude $`W_{1324}`$. If we define $$A_{ab}J_{ac}(x_{13})J_{ce}(x_{23})J_{eg}(x_{24})J_{gb}(x_{14}),$$ then the amplitude takes the form $$W_{1324}=\frac{8c^4}{x_{13}^4x_{24}^4x_{23}^4x_{14}^4}[(\text{tr}A)^2\text{tr}A^2].$$ Let $`\lambda _i`$ be the four eigenvalues of $`A`$. Clearly $$Ch(A)(\text{tr}A)^2\text{tr}A^2=2\underset{i<j}{}\lambda _i\lambda _j.$$ A brief consideration of $`Ch(A)`$ reveals that it is invariant under the conformal group. In particular, we consider the case in which we use a translation to set $`x_3=0`$ and then an inversion to send it off to infinity. In this case, $$J_{ac}(x_{13})J_{ce}(x_{23})=\delta _{ae}+O(|x_3|^1).$$ If we then choose a basis in which $`x_{24}=(a,0,0,0)`$ and $`x_{14}=(b\mathrm{sin}\varphi ,b\mathrm{cos}\varphi ,0,0)`$, the matrix $`A`$ becomes effectively two dimensional: Two of the $`\lambda _i`$ equal one, and the remaining two can be obtained by diagonalizing the matrix $$\left(\begin{array}{cc}\mathrm{cos}2\varphi & \mathrm{sin}2\varphi \\ \mathrm{sin}2\varphi & \mathrm{cos}2\varphi \end{array}\right)$$ giving $`\lambda _\pm =\mathrm{exp}(\pm 2\varphi i)`$. Hence $$Ch(A)=4(1+4\mathrm{sin}\varphi ).$$ As $`Ch(A)`$ is invariant under the conformal group, it must be expressible as a function of $`s`$ and $`t`$. In the limit $`x_3\mathrm{}`$ $`s`$ $``$ $`{\displaystyle \frac{1}{2}}{\displaystyle \frac{x_{24}^2}{x_{12}^2+x_{14}^2}},`$ $`t`$ $``$ $`{\displaystyle \frac{x_{12}^2x_{14}^2}{x_{12}^2+x_{14}^2}}.`$ It is now a straightforward matter to express $`Ch(A)`$ in terms of $`s`$ and $`t`$: $$Ch(A)=\frac{4}{s(1t)}(s+t^23st+4s^2).$$ Indeed, Mathematica was used to verify that this expression is correct. It follows immediately that $$W_{1324}=\frac{2^9c^4}{x_{13}^8x_{24}^8}\frac{s(s+t^23st+4s^2)}{(1t)^3}.$$ Repeating the calculation for the other two $`W_{abcd}`$, we obtain the connected 4-point function $`M_4|_{\text{connected}}`$ $`=`$ $`{\displaystyle \frac{2^{14}c^4}{x_{13}^8x_{24}^8}}{\displaystyle \frac{1}{(1t^2)^3}}[s(s+t^2+4s^212st^2+3t^4+12s^2t^23st^4)+`$ $`+2^3s^4(316s+t^2+16s^2)].`$ As one can see, the leading order terms $`s^2`$ and $`st^2`$ are in agreement with (2). However, the coefficients of the higher order terms are quite different from those in (2), thus demonstrating that a new primary, or primaries, appear at this level, as we indeed saw in the previous section. ## 6 Discussion, Conclusions, and Ideas for Future Work We have successfully reproduced all of the singular terms in the four dilaton, four axion, and two dilaton-two axion 4-point functions calculated in using only the assumptions of conformal invariance and the presence of $`T_{ab}`$ in the OPE of two scalars. Moreover, we have developed a better understanding of the structure of the descendants of $`T_{ab}`$ in this OPE. Comparison of this strongly coupled result to the weakly coupled limit revealed the presence of a nonchiral primary $`P_{abcd}`$ (see equation 17). At weak coupling, this nonchiral primary has spin four and dimension six. However, this primary is believed to have an anomalous dimension which grows at least as fast as $`(g_{YM}^2N)^{1/4}`$, thus effectively disappearing as an exchanged operator in the 4-point functions calculated by . Some lines of inquiry remain open. It would be nice to know the exact form of $`T_{abcdef}^{(4)}`$, and knowing this form would allow the work done here to be extended to the order $`n=16`$, the first order in the strongly coupled 4-point functions calculated by where new chiral primaries are expected to appear. The difference between the conformal block contribution of $`T_{ab}`$ and the correlation functions as calculated by would then presumably give us some insight as to the precise nature of these additional operators, allowing, perhaps, a better understanding of the logarithms that appear in these 4-point functions. The present method of calculating descendants order by order becomes extremely cumbersome to apply at higher order, so a more efficient approach may be to use variants of equations given in and . Although, as was mentioned above, we have had trouble using Eq. 3.15 of directly, we do have a guess as to how to modify the equation, and the modified version matches the AdS results in a way we would expect. ## Note Added for Publication A year has elapsed between the writing of this paper and its submission for publication. During the interim several papers have appeared which use results derived here and which answer some of the questions raised in this paper. For example, in , more explicit formulae for the conformal block contribution of higher spin operators to scalar four point functions were derived. These results hopefully clarify the confusion in this paper concerning the work of . Another important work is where, using some results derived here, the program suggested in the conclusion of this paper was successfully carried out. ## Appendix Here are some useful identities involving the tensor $$J_{ab}(x)\delta _{ab}\frac{2x_ax_b}{x^2}.$$ In what follows, the argument of $`J_{ab}`$ will be $`x`$. First, here are some elementary properties of the tensor: $$J_{ab}J_{bc}=\delta _{ac};J_{aa}=2;x_aJ_{ab}=x_b.$$ Next, here are some trivial index rearrangements: $$x_iJ_{ab}=x_bJ_{ia}+x_i\delta _{ab}x_b\delta _{ia},$$ $$J_{cd}J_{ij}=J_{ic}J_{jd}+\delta _{ij}J_{cd}+\delta _{cd}J_{ij}\delta _{ic}J_{jd}\delta _{jd}J_{ic}+\delta _{ic}\delta _{jd}\delta _{cd}\delta _{ij}.$$ Finally, here are some derivatives of $`J_{ij}`$: $$_mJ_{ij}=\frac{2}{x^2}(x_mJ_{ij}x_m\delta _{ij}+x_i\delta _{jm}+x_j\delta _{im}),$$ $`_n_mJ_{ij}`$ $`=`$ $`{\displaystyle \frac{2}{x^2}}(2J_{mn}J_{ij}2\delta _{ij}J_{mn}\delta _{mn}J_{ij}+`$ $`+\delta _{jm}J_{in}+\delta _{im}J_{jn}+\delta _{jn}J_{im}+\delta _{in}J_{jm}+`$ $`\delta _{im}\delta _{jn}\delta _{jm}\delta _{in}+\delta _{mn}\delta _{ij}).`$ ## Acknowledgments The author would like to acknowledge many useful discussions with Igor Klebanov. He is also grateful for several useful conversations with Leonardo Rastelli. Thanks also go to A. Petkou, who, after the first electronic submission of this paper, brought to our attention relevant literature concerning O(N) non-linear $`\sigma `$-models.
warning/0002/math0002127.html
ar5iv
text
# Wavelets having the Translation Invariance Property of Order n ## 1. Introduction The study of the interplay between wavelets and operator theory has proven useful recently. This paper considers part of that study, in particular the interaction of certain wavelets and the translation operator. Every wavelet can be associated to a GMRA, which is a nested sequence of subspaces of $`L^2()`$ with other properties. The central subspace, called the *core subspace* has the property that it is, along with the larger spaces in this “ladder”, invariant under integer translations. In this paper, we shall show the existence of wavelets such that some, but not all, of the smaller spaces in this ladder are also invariant under integer translations. In this context, these wavelets can be considered generalizations of MSF wavelets. We begin with several definitions. A wavelet for dilation $`d`$ is a square integrable function such that the collection $$\{d^{\frac{n}{2}}\psi (d^nxl):n,l\}$$ forms an orthonormal basis of $`L^2()`$. Define unitary operators on $`L^2()`$ by $`Tf(x)=f(x1)`$ and $`Df(x)=\sqrt{d}f(dx)`$ for $`d,d2`$, called the translation and dilation operator, respectively. As in , define a nonsingular *Generalized Multiresolution Analysis* (GMRA) as a sequence of subspaces $`\{V_j\}_j`$ with the following five properties: 1. $`V_jV_{j+1}`$, 2. $`_jV_j`$ is dense in $`L^2()`$, 3. $`_jV_j=\{0\}`$, 4. $`D(V_j)=V_{j+1}`$ 5. $`V_0`$ is invariant under $`T^l`$ for $`l`$. Given a wavelet $`\psi `$, define $`V_j`$ as the closed span of $`\{D^nT^l\psi :n<j,l\}`$. It is routine to verify that this sequence of subspaces of $`L^2()`$ form a GMRA. Lastly, define a dual sequence of closed subspaces, denoted $`W_j`$, by $`V_{j+1}=V_jW_j`$. Note that $`W_0=\overline{span}\{T^l\psi :l\}`$, i.e. $`W_0`$ is where $`\psi `$ lives. The commutation relation between $`D`$ and $`T`$, $`TD=DT^d`$, yields that for any $`j`$, if $`V_j`$ is invariant under $`T`$, then so is $`V_{j+1}`$. Hence, for a GMRA, the subspaces $`V_j`$ are invariant under $`T`$ for $`j0`$. We will show by construction that, given a negative integer $`n`$, there exists a wavelet such that its associated GMRA has the property that the subspaces $`V_j`$ are invariant under $`T^l`$ if and only if $`jn`$. Such a wavelet is said to have the translation invariance property of order $`n`$. We shall denote by $`_n`$ the collection of all wavelets such that $`V_n`$ is invariant under integral translations. Note that these collections are nested: $$_0_1_2\mathrm{}_n_{n+1}\mathrm{}_{\mathrm{}}$$ Define $`_n=_n_{n+1}`$, with $`_{\mathrm{}}=_{\mathrm{}}`$, the collection of all wavelets such that $`V_j`$ is invariant under translations for all $`j`$. It is proven in that the Minimally Supported Frequency (MSF) wavelets coincide with $`_{\mathrm{}}`$. A MSF wavelet is a wavelet $`\psi `$ such that $`|\widehat{\psi }|`$ is the indicator function of some set, usually called a wavelet set. MSF wavelets shall be a starting point for constructing the wavelets we seek. The goal of this paper is to prove the following theorem: ###### Theorem 1. For all positive integers $`n`$, there exists a wavelet with translation invariance property of order $`n`$, i.e. the collections $`_n`$ are non-empty. This is true for all $`n`$ and for all positive integer dilation factors. The proof is the content of section 3. For the purposes of this paper, we shall say that a set $`E`$ is *partially self-similar* with respect to $`\alpha `$ if there exists a set $`F`$ of non-zero measure such that both $`F`$ and $`F+\alpha `$ are subsets of $`E`$. Additionally, if $`G,H`$ are two subsets of $``$, we shall say that $`G`$ is $`2\pi `$ translation congruent to $`H`$ if there exists a measurable partition $`G_n`$ of $`G`$ such that the collection $`\{G_n+2n\pi :n\}`$ forms a partition of $`H`$, modulo sets of measure zero. Define a mapping $`\tau :[0,2\pi )`$ such that $`\tau (x)x=2\pi k`$ for some integer $`k`$. ## 2. A Characterization In this section we shall develop a characterization of wavelets that have the translation invariance property of order $`n`$. This characterization comes from considering the support of the Fourier transform of the wavelet in question. Fix a dilation factor $`d`$. Denote by $`T_\alpha `$ the unitary operator $`T_\alpha f(x)=f(x\alpha )`$. $`T`$ is to be understood as $`T_1`$. Note that $`\widehat{T_\alpha }=M_{e^{i\alpha \xi }}`$. We shall consider the groups of translations $`𝒢_n=\{T_{\frac{m}{d^n}}:m\}`$. The following proposition will allow us to consider the action of translation operators on $`W_0`$ instead of on the $`V_j`$’s. ###### Proposition 2. The following are equivalent: 1. the space $`V_n`$ for $`\psi `$ is invariant under the action of translations, 2. $`V_0`$ is invariant under the action of $`T_{\frac{m}{d^n}}`$, 3. $`W_0`$ is invariant under the action of $`T_{\frac{m}{d^n}}`$. ###### Proof. By definition, $`fV_n`$ if and only if $`D^nfV_0`$; this fact combined with the commutation relation $`DT=T_{\frac{1}{d}}D`$ establishes the equivalence of $`1`$ and $`2`$. Furthermore, these two facts show that if $`V_0`$ is invariant under the action of $`T_{\frac{m}{d^n}}`$, then so is $`V_1`$, whence $`W_0`$ is as well. Finally, suppose that $`W_0`$ is invariant under $`T_{\frac{m}{d^n}}`$. Then for $`k>0`$, $`W_k`$ is invariant under $`T_{\frac{m}{d^n}}`$, whence $`V_0^{}=_{k=0}^{\mathrm{}}W_k`$ is invariant under $`𝒢_n`$. It follows that $`V_0`$ is invariant under $`𝒢_n`$. ∎ If $`fW_0`$, then we can write $`f=_kc_kT^k\psi `$, so taking the Fourier transform of both sides yields $`\widehat{f}=g\widehat{\psi }`$ for some $`gL^2([0,2\pi ))`$. Hence, we can describe $`\widehat{W}_0`$ by $`\{g(\xi )\widehat{\psi }(\xi ):gL^2([0,2\pi ))\}`$. Suppose $`\psi `$ is a wavelet that is in $`_n`$, by proposition 2, this is equivalent to $`T_{\frac{m}{d^n}}fW_0`$. By taking the Fourier Transform, we have $`e^{i\frac{m}{d^n}}\widehat{\psi }\widehat{W}_0`$. Hence, $`\psi _n`$ is equivalent to the condition that $`e^{i\frac{m}{d^n}\xi }\widehat{\psi }(\xi )=g(\xi )\widehat{\psi }(\xi )`$. For ease of notation, define: $$_\psi =\{k:supp(\widehat{\psi })\text{ is partially self similar w.r.t. }2\pi k\}.$$ ###### Theorem 3. Let $`\psi `$ be a wavelet. Then $`\psi _n`$ if and only if every element of $`_\psi `$ is divisible by $`d^n`$. ###### Only If. We prove by contradiction. Let $`\psi _n`$. Hence, $$e^{i\frac{1}{d^n}\xi }\widehat{\psi }(\xi )=g(\xi )\widehat{\psi }(\xi )$$ for some $`gL^2([0,2\pi ))`$. Let $`E=supp(\widehat{\psi })`$, for $`\xi E`$, $`e^{i\frac{1}{d^n}\xi }=g(\xi )`$. Suppose that $`k_\psi `$ is not divisible by $`d^n`$, and $`F`$ is a set of non-zero measure such that both $`F`$ and $`F+2k\pi `$ are subsets of $`E`$. Let $`\xi F`$; we have, $`e^{i\frac{\xi +2k\pi }{d^n}}\widehat{\psi }(\xi +2k\pi )`$ $`=g(\xi +2k\pi )\widehat{\psi }(\xi +2k\pi )`$ $`e^{i\frac{\xi }{d^n}}e^{i\frac{2k\pi }{d^n}}\widehat{\psi }(\xi +2k\pi )`$ $`=g(\xi )\widehat{\psi }(\xi +2k\pi )`$ $`e^{i\frac{2k\pi }{d^n}}`$ $`=1`$ a contradiction. *If.* It suffices to show that $$e^{i\frac{1}{d^n}\xi }\widehat{\psi }(\xi )=g(\xi )\widehat{\psi }(\xi )$$ for some $`gL^2([0,2\pi ))`$. Again, let $`E=supp(\widehat{\psi })`$. Let $`FE`$ be such that $`\tau :F[0,2\pi )`$ is a bijection. (It can be easily shown that $`\tau :E[0,2\pi )`$ is a surjection.) The injective property of $`\tau `$ can be assured in the following manner: for each $`\xi [0,2\pi )`$, define the set $`Z_\xi =\{m_\xi :\xi +2m_\xi \pi E\}`$, then for $`\xi `$ choose $`k_\xi `$ to be 0 if $`\xi E`$, if not, choose $`k_\xi =min\{m>0:mZ_\xi \}`$, else choose $`k_\xi =max\{m<0:mZ_\xi \}`$. Let $`F=\{\xi +2k_\xi \pi :\xi [0,2\pi )\}`$. Note that by construction, $`F`$ is $`2\pi `$ translation congruent to $`[0,2\pi )`$. Hence, $$e^{i\frac{1}{d^n}\xi }\chi _F(\xi )=g(\xi )$$ where $`g(\xi )L^2(F)`$ and is $`2\pi `$ periodic. Thus, for $`\xi F`$, $$e^{i\frac{1}{d^n}\xi }\widehat{\psi }(\xi )=g(\xi )\widehat{\psi }(\xi ).$$ For almost any $`\xi EF`$, there exists a $`\xi ^{}F`$ and an integer $`l_\xi `$ such that $`\xi \xi ^{}=2l_\xi \pi `$. Moreover, by hypothesis, $`l_\xi `$ is a multiple of $`d^n`$, since $`l_\xi _\psi `$. Since $`e^{i\frac{1}{d^n}\xi }`$ is $`d^n\pi `$ periodic, we have that for $`\xi EF`$, $`e^{i\frac{1}{d^n}\xi }\widehat{\psi }(\xi )`$ $`=e^{i\frac{1}{d^n}(\xi ^{}+2l_\xi \pi )}\widehat{\psi }(\xi ^{}+2l_\xi \pi )`$ $`=e^{i\frac{1}{d^n}\xi ^{}}\widehat{\psi }(\xi ^{}+2l_\xi \pi )`$ $`=g(\xi ^{})\widehat{\psi }(\xi ^{}+2l_\xi \pi )`$ $`=g(\xi )\widehat{\psi }(\xi ).`$ This completes the proof. ∎ As an immediate corollary, we have the following characterization of wavelets in $`_n`$. ###### Corollary 4. Let $`\psi `$ be a wavelet. Then $`\psi _n`$ if and only if every element of $`_\psi `$ is divisible by $`d^n`$, but there exists an element of $`_\psi `$ that is not divisible by $`d^{n+1}`$ ## 3. Wavelets of Higher Order Translation Invariance In this section, we arrive at the crux of the paper, constructing wavelets with the translation invariance property of order $`n`$ . When finished, we will have proven theorem 1. The basic procedure is to consider a special class of *wavelet multiplicity functions* (see ), from which we can construct two generalized scaling sets. Those in turn give rise to two wavelet sets which admit operator interpolation. The resulting interpolated wavelet has the desired support in the frequency domain (proposition 2). The multiplicity functions depend on whether the dilation factor is 2 or another integer; we shall first consider dilation by 2. Recall that $`\psi `$ is a MSF wavelet if $`|\widehat{\psi }|=\chi _W`$, where $`W`$ is referred to as a *wavelet set*. Such a wavelet shall be denoted by $`\psi _W`$. We refer the reader to for background information on wavelet sets. A set $`E`$ is called a *generalized scaling set* if the set $`W=dEE`$ is a wavelet set (). We refer the reader to \[5, ch. 4\] for a detailed description of operator interpolation. We shall give here the necessary results concerning operator interpolation. Let $`\psi _{W_1}`$ and $`\psi _{W_2}`$ be MSF wavelets. By lemma 4.3 in \[5, pg. 41\], $`W_1`$ is $`2\pi `$ translation congruent to $`W_2`$. If $`\sigma :W_1W_2`$ is determined by this translation congruence, then $`\sigma `$ can be extended to a measurable bijection of $``$ by defining $`\sigma (x)=d^n\sigma (d^nx)`$ where $`n`$ is such that $`d^nxW_1`$. If $`\sigma `$ is *involutive*, i.e. $`\sigma ^2`$ is the identity, and if $`h_1`$ and $`h_2`$ are measurable, essentially bounded, d-dilation periodic functions (i.e. $`h_1(dx)=h_1(x)`$), then $`\psi `$ defined by $$\widehat{\psi }=h_1\widehat{\psi }_{W_1}+h_2\widehat{\psi }_{W_2}$$ is again a wavelet provided the matrix $$\left(\begin{array}{cc}h_1& h_2\\ h_2\sigma ^1& h_1\sigma ^1\end{array}\right)$$ (1) is unitary almost everywhere. (Since $`\sigma ^1`$ is d-homogeneous, and the $`h_i`$’s are d-dilation periodic, in general it suffices to check this condition on $`W_1`$.) Note that the interpolated wavelet $`\psi `$ has the property that $`supp(\widehat{\psi })W_1W_2`$. Furthermore, since $`\sigma `$ on $`W_1`$ is given by translations by integral multiples of $`2\pi `$, $`\sigma `$ completely describes the partial self similarity of $`W_1W_2`$ with respect to multiples of $`2\pi `$. In order to find the wavelet sets and do the interpolation between them, we need the following theorem, which provides a technique for constructing generalized scaling sets. The proof can be found in . ###### Theorem 5 (Merrill). Given a multiplicity function $`m:[\pi ,\pi )\{0,1,\mathrm{},\mathrm{}\}`$ which satisfies the consistency equation: $$m(\omega )+1=\underset{i=0}{\overset{d1}{}}m\left(\frac{\omega }{d}+\frac{2\pi i}{d}\right)$$ (2) $`m`$ is associated with a (wavelet set $`W`$) wavelet if and only if $``$ a set E such that: 1. $`m(\omega )=_k\chi _E(\omega +2\pi k)`$ 2. $`EdE`$ 3. $`_jd^jE\text{ contains a neighborhood of 0.}`$ Indeed, $`W=dEE`$ is such a wavelet set. Baggett constructed an entire family of multiplicity functions for $`d=2`$ that satisfy the consistency equation and produce wavelet sets, and hence wavelets. Merrill discovered that these functions could be constructed using the idea of periodic points, and the breaks in these multiplicity functions occurred at these periodic points. These periodic points are found by solving the following equation: $$d^ka=a+2\pi j$$ where $`k1`$ is the highest value of $`m(\omega )`$. Solving for $`a`$ yields $$a=\frac{2\pi j}{d^k1}.$$ For a dilation factor of 2, Baggett constructed the following family of multiplicity functions, where $`k2`$ and the highest value of $`m(\omega )`$ is $`k1`$. $$m_k^2(\omega )=\{\begin{array}{cc}1,\hfill & \pi \omega <\frac{(2^k2)\pi }{2^k1};\hfill \\ 0,\hfill & \frac{(2^k2)\pi }{2^k1}\omega <\frac{2^{k1}\pi }{2^k1};\hfill \\ kj,\hfill & \frac{2^j\pi }{2^k1}\omega <\frac{2^{j1}\pi }{2^k1};\hfill \\ k1,\hfill & \frac{2\pi }{2^k1}\omega <\frac{2\pi }{2^k1};\hfill \\ kj,\hfill & \frac{2^{j1}\pi }{2^k1}\omega <\frac{2^j\pi }{2^k1};\hfill \\ 0,\hfill & \frac{2^{k1}\pi }{2^k1}\omega <\frac{(2^k2)\pi }{2^k1};\hfill \\ 1,\hfill & \frac{(2^k2)\pi }{2^k1}\omega <\pi ,\hfill \end{array}$$ for $`j=2,3,\mathrm{},k1`$. A proof that these multiplicity functions satisfy the consistency equation can be found in . Now, by Merrill’s Theorem, we need to find a set $`E`$ that satisfies the conditions of the theorem. To construct $`E`$, we take the support of $`m_k^2`$, call it $`E_{m_k^2}`$ and put that in $`E`$. Now, we take the interval where $`m_k^2(\omega )kj`$ for $`j=1,2,\mathrm{}k2`$ and shift the negative half, $`[\frac{2^j\pi }{2^k1},0)`$ to the right $`2^j\pi `$, and call this $`E_j^+`$. We then take the positive half, $`[0,\frac{2^j\pi }{2^k1})`$ and shift it to the left $`2^j\pi `$ and call this part $`E_j^{}`$. Then let $`E_j=E_j^{}E_j^+`$. Define, $$E=E_{m_k^2}\underset{j=1}{\overset{k2}{}}E_j;$$ we have: $$\begin{array}{c}E=_{j=1}^{k2}[2^j\pi ,2^j\pi +\frac{2^j\pi }{2^k1})[\pi ,\frac{(2^k2)\pi }{2^k1})[\frac{2^{k1}\pi }{2^k1},\frac{2^{k1}}{2^k1}\pi )\hfill \\ \hfill [\frac{(2^k2)\pi }{2^k1},\pi )_{j=1}^{k2}[2^j\pi \frac{2^j\pi }{2^k1},2^j\pi ).\end{array}$$ By definition, $`E`$ satisfies the first condition of Merrill’s theorem. The other two conditions are easily verified. Hence, there is a wavelet set $`W=2EE`$: $$\begin{array}{c}W=[2^{k1}\pi ,\frac{(2^{2k1}2^k)\pi }{2^k1})[\frac{2^k\pi }{2^k1},\pi )[\frac{(2^k2)\pi }{2^k1},\frac{2^{k1}\pi }{2^k1})\hfill \\ \hfill [\frac{2^{k1}\pi }{2^k1},\frac{(2^k2)\pi }{2^k1})[\pi ,\frac{2^k\pi }{2^k1})[\frac{(2^{2k1}2^k)\pi }{2^k1},2^{k1}\pi ).\end{array}$$ ###### Proof of theorem 1 (for dilation factor 2). Now we shall construct a second wavelet set to form an interpolation pair. Take the same multiplicity function, $`m_k^2(\omega )`$ and construct an $`E^{}`$ similar to $`E`$, except we will take the interval $`[\frac{2^{k1}}{2^k1},\frac{\pi }{2})`$ and shift it right $`2^{k1}\pi `$. The result is $$\begin{array}{c}E^{}=_{j=1}^{k2}[2^j\pi ,2^j\pi +\frac{2^j\pi }{2^k1})[\pi ,\frac{(2^k2)\pi }{2^k1})\hfill \\ \hfill [\frac{\pi }{2},\frac{2^{k1}}{2^k1})[\frac{(2^k2)\pi }{2^k1},\pi )_{j=1}^{k2}[2^j\pi \frac{2^j\pi }{2^k1},2^j\pi )\\ \hfill [2^{k1}\pi \frac{2^{k1}}{2^k1},2^{k1}\pi \frac{\pi }{2}).\end{array}$$ This new $`E^{}`$ also satisfies the conditions of Merrill’s theorem, so there is a wavelet set $`W^{}`$: $$\begin{array}{c}W^{}=[2^{k1}\pi ,\frac{(2^{2k1}2^k)\pi }{2^k1})[\frac{(2^k2)\pi }{2^k1},\frac{\pi }{2})\hfill \\ \hfill [\frac{2^{k1}\pi }{2^k1},\frac{(2^k2)\pi }{2^k1})[\pi ,\frac{2^k\pi }{2^k1})[2^{k1}\pi \frac{\pi }{2},2^{k1}\pi )\\ \hfill [2^k\pi \frac{2^k\pi }{2^k1},2^k\pi \pi )\end{array}$$ $$\begin{array}{c}WW^{}=[2^{k1}\pi ,\frac{(2^{2k}2^k)\pi }{2^k1})[\frac{(2^k2)\pi }{2^k1},\frac{2^{k1}\pi }{2^k1})\hfill \\ \hfill [\frac{2^{k1}\pi }{2^k1},\frac{(2^k2)\pi }{2^k1})[\pi ,\frac{2^k\pi }{2^k1})[2^{k1}\pi \frac{\pi }{2},2^{k1}\pi )\end{array}$$ $$WW^{}=[\frac{2^k\pi }{2^k1},\pi )[\frac{(2^{2k1}2^k)\pi }{2^k1},2^{k1}\pi \frac{\pi }{2})=A_1A_2$$ $$W^{}W=[\frac{2^{k1}\pi }{2^k1},\frac{\pi }{2})[2^k\pi \frac{2^k\pi }{2^k1},2^k\pi \pi )=B_1B_2$$ We claim that the wavelet sets $`W`$ and $`W^{}`$ form an interpolation pair. Then $`\sigma :WW^{}`$ is defined as follows: $$\sigma (\xi )=\{\begin{array}{cc}\xi ,\hfill & \xi WW^{}\hfill \\ \xi +2^k\pi ,\hfill & \xi [\frac{2^k\pi }{2^k1},\pi ),(A_1B_2)\hfill \\ \xi 2^{k1}\pi ,\hfill & \xi [\frac{(2^{2k1}2^k)\pi }{2^k1},2^{k1}\frac{\pi }{2}),(A_2B_1)\hfill \end{array}$$ We need to check that $`\sigma `$ is involutive (see ). Notice that $`B_2=2A_2`$ and that $`2B_1=A_1`$. For $`\xi A_1,\sigma (\xi )B_2,\text{ so }\frac{1}{2}\sigma (\xi )A_2`$. This yields the following: $`\sigma ^2(\xi )`$ $`=2\sigma ({\displaystyle \frac{1}{2}}\sigma (\xi ))`$ $`=2\sigma ({\displaystyle \frac{1}{2}}(\xi +2^k\pi ))`$ $`=2\sigma ({\displaystyle \frac{1}{2}}\xi +2^{k1}\pi )`$ $`=2({\displaystyle \frac{1}{2}}\xi +2^{k1}\pi 2^{k1}\pi )`$ $`=\xi `$ Similarly, for $`\xi A_2`$, we get that $`\sigma ^2(\xi )=\xi `$, whence $`\sigma `$ is involutive. Finally, to interpolate between these two wavelets, define the functions $`h_1`$ on $`W`$ and $`h_2`$ on $`W^{}`$ as follows: $`h_1(\xi )`$ $`=\{\begin{array}{cc}1,\hfill & \xi WW^{},\hfill \\ \frac{1}{\sqrt{2\pi }},\hfill & \xi [\frac{(2^{2k1}2^k)\pi }{2^k1},2^{k1}\frac{\pi }{2}),\hfill \\ \frac{1}{\sqrt{2\pi }},\hfill & \xi [\frac{2^k\pi }{2^k1},\pi ),\hfill \end{array}`$ $`h_2(\xi )`$ $`=\{\begin{array}{cc}1,\hfill & \xi WW^{},\hfill \\ \frac{1}{\sqrt{2\pi }},\hfill & \xi [\frac{(2^k2^{k1})\pi }{2^k1},\frac{\pi }{2}),\hfill \\ \frac{1}{\sqrt{2\pi }},\hfill & \xi [2^k\pi \frac{2^k\pi }{2^k1},2^k\pi \pi ),\hfill \end{array}`$ and extend via 2-dilation periodicity. It is also routine to verify that these functions satisfy the matrix condition 1. ###### Claim 1. Let $`W`$, $`W^{}`$, $`h_1`$ and $`h_2`$ be as above. Then the wavelet defined by $$\widehat{\psi }=h_1\widehat{\psi }_W+h_2\widehat{\psi }_W^{}$$ is an element of $`_{k2}`$ ###### Proof. As defined above, $`\widehat{\psi }`$ is a wavelet. Note that $`\sigma `$ above yields that $`2^{k2},2^{k1}_\psi `$, whence the claim is established by corollary 4. ∎ Consequently, there exist wavelets in all $`_n`$’s, for dilation by 2. ∎ ###### Example 1 (k=3). The following is a concrete example of the above procedure. The resulting wavelet is in $`_1`$, so that the corresponding subspace $`V_1`$ is invariant under translations. Additionally, this example will be of use in section 4. The multiplicity function is: $$m(\omega )=\{\begin{array}{cc}2,\hfill & \omega [\frac{2\pi }{7},\frac{2\pi }{7}),\hfill \\ 0,\hfill & \omega [\frac{6\pi }{7},\frac{4\pi }{7})[\frac{4\pi }{7},\frac{6\pi }{7}),\hfill \\ 1,\hfill & \text{for all other }\omega [\pi ,\pi ).\hfill \end{array}$$ Note that this is the multiplicity function for the Journé wavelet. $`W`$ $`=[4\pi ,{\displaystyle \frac{24\pi }{7}})[{\displaystyle \frac{8\pi }{7}},\pi )[{\displaystyle \frac{6\pi }{7}},{\displaystyle \frac{4\pi }{7}})`$ $`[{\displaystyle \frac{4\pi }{7}},{\displaystyle \frac{6\pi }{7}})[\pi ,{\displaystyle \frac{8\pi }{7}})[{\displaystyle \frac{24\pi }{7}},4\pi )`$ $`W^{}`$ $`=[4\pi ,{\displaystyle \frac{24\pi }{7}})[{\displaystyle \frac{6\pi }{7}},{\displaystyle \frac{\pi }{2}})[{\displaystyle \frac{4\pi }{7}},{\displaystyle \frac{6\pi }{7}})`$ $`[\pi ,{\displaystyle \frac{8\pi }{7}})[{\displaystyle \frac{7\pi }{2}},4\pi )[{\displaystyle \frac{48\pi }{7}},7\pi )`$ We have: $$\sigma (\xi )=\{\begin{array}{cc}\xi ,\hfill & \xi WW^{},\hfill \\ \xi 4\pi ,\hfill & \xi [\frac{24\pi }{7},\frac{7\pi }{2}),\hfill \\ \xi +8\pi ,\hfill & \xi [\frac{8\pi }{7},\pi ),\hfill \end{array}$$ therefore interpolation as above yields a wavelet in $`_1`$. We have found wavelet sets for dilation by 2 and interpolated between them. Now we would like to generalize this argument and be able to do this for any dilation. It turns out that 2 is the only dilation factor that has symmetric multiplicity functions, the multiplicity functions associated to the others factors are asymmetric. We will now look at integer dilation factors greater than or equal to 3, such factors will be denoted by $`d`$. ###### Theorem 6. There exists a family of multiplicity functions on $`[\pi ,\pi )`$, $`\{m_k^d\}`$ for $`k2`$, $`d3`$ that satisfy the consisten cy equation 2, given as follows: $$m_k^d(\omega )=\{\begin{array}{cc}0,\hfill & \pi \omega <\frac{d^{k2}(d1)(2\pi )}{d^k1},(I_1)\hfill \\ kj,\hfill & \frac{d^{j1}(d1)(2\pi )}{d^k1}\omega <\frac{d^{j2}(d1)(2\pi )}{d^k1},(I_2)\hfill \\ k1,\hfill & \frac{(d1)(2\pi )}{d^k1}\omega <\frac{2\pi }{d^k1},(I_3)\hfill \\ kj,\hfill & \frac{d^{j2}(2\pi )}{d^k1}\omega <\frac{d^{j1}(2\pi )}{d^k1},(I_4)\hfill \\ 0,\hfill & \frac{d^{k2}(2\pi )}{d^k1}\omega <\frac{2\pi }{d}\frac{(d1)(2\pi )}{d(d^k1)},(I_5)\hfill \\ 1,\hfill & \frac{2\pi }{d}\frac{(d1)(2\pi )}{d(d^k1)}\omega <\frac{2\pi }{d}+\frac{2\pi }{d(d^k1)},(I_6)\hfill \\ 0,\hfill & \frac{2\pi }{d}+\frac{2\pi }{d(d^k1)}\omega <\pi ,(I_7)\hfill \end{array}$$ for, $`j=2,3,\mathrm{},k1`$. Furthermore, for each of the functions $`m_k^d`$, there is an associated generalized scaling set $`E_k^d`$ and a wavelet set $`W_k^d`$. ###### Proof. Fix $`d`$ and $`k`$. We shall first give a brief outline to show that $`m_k^d(\omega )`$ satisfies the consistency equation 2. For a complete proof of this fact see . Let $`\omega I_1`$, then $`\frac{\omega }{d}I_2`$ where $`m(\omega )=k(k1)`$ and $`\frac{\omega }{d}+\frac{2\pi }{d}I_5`$, and $`\frac{\omega }{d}+\frac{4\pi }{d},\mathrm{},\frac{\omega }{d}+\frac{(d1)2\pi }{d}`$ are in either $`I_1`$ or $`I_7`$. Consistency equation: $`0+1=k(k1)+0\mathrm{}+0`$. A similar argument is used for $`\omega I_m,\mathrm{\hspace{0.25em}2}m7`$ to prove that $`m_k^d`$ satisfies the consistency equation. Here is how to construct the set $`E_k^d`$. Let $`E_{m_k^d}`$ = support of $`m_k^d`$ so $`E_{m_k^d}`$ $`=`$ $`[{\displaystyle \frac{d^{k2}(d1)(2\pi )}{d^k1}},{\displaystyle \frac{d^{k2}(2\pi )}{d^k1}})`$ $`[{\displaystyle \frac{2\pi }{d}}{\displaystyle \frac{(d1)(2\pi )}{d(d^k1)}},{\displaystyle \frac{2\pi }{d}}+{\displaystyle \frac{2\pi }{d(d^k1)}})`$ Now take the set where $`m_k^dkj,\text{ for }j=1,\mathrm{}k2`$ and shift it to the right by $`2\pi (d^{j1})`$, call this set $`E_{j1}`$. Thus, $$E_{j1}=[2\pi (d^{j1})\frac{d^{j1}(d1)(2\pi )}{d^k1},2\pi (d^{j1})+\frac{d^{j1}(2\pi )}{d^k1})$$ and let $$E_k^d=E_{m_k^d}_{j=1}^{k2}E_{j1}.$$ It is easy to show that $$E_k^dd(E_k^d)$$ so by Merrill’s theorem, $`W_k^d`$ $`=`$ $`d(E_k^d)E_k^d`$ $`=`$ $`[{\displaystyle \frac{d^{k1}(d1)2\pi }{d^k1}},{\displaystyle \frac{d^{k2}(d1)2\pi }{d^k1}})`$ $`[{\displaystyle \frac{d^{k2}(2\pi )}{d^k1}},{\displaystyle \frac{2\pi }{d}}{\displaystyle \frac{(d1)2\pi }{d(d^k1)}})`$ $`[2\pi (d^{k2}){\displaystyle \frac{d^{k2}(d1)2\pi }{d^k1}},2\pi (d^{k2})+{\displaystyle \frac{d^{k2}(2\pi )}{d^k1}})`$ is the desired wavelet set. ∎ ###### Proof of theorem 1 (for dilation factor greater than 2). The above is one way of constructing $`E_k^d`$ and $`W_k^d`$. Now, we will construct a second generalized scaling set $`E`$ and a second wavelet set $`W`$ that have the same multiplicity function by moving the interval $`[\frac{d^{k2}(d1)2\pi }{d^k1},\frac{2\pi }{d^2})`$ to the right $`2\pi (d^{k2})`$. $$\begin{array}{c}\stackrel{~}{E_k^d}=[\frac{2\pi }{d^2},\frac{d^{k2}(2\pi )}{d^k1})[\frac{2\pi }{d}\frac{(d1)2\pi }{d(d^k1)},\frac{2\pi }{d}+\frac{2\pi }{d(d^k1)})\hfill \\ \hfill _{j=1}^{k2}E_{j1}[2\pi (d^{k2})\frac{d^{k2}(d1)2\pi }{d^k1},2\pi (d^{k2})\frac{2\pi }{d^2})\end{array}$$ $$\begin{array}{c}\stackrel{~}{W_k^d}=[\frac{2\pi }{d},\frac{2\pi }{d^2})[\frac{d^{k2}(2\pi )}{d^k1},\frac{2\pi }{d}\frac{(d1)2\pi }{d(d^k1)})\hfill \\ \hfill [2\pi (d^{k2})\frac{2\pi }{d^2},2\pi (d^{k2})+\frac{d^{k2}(2\pi )}{d^k1})\\ \hfill [2\pi (d^{k1})\frac{d^{k1}(d1)2\pi }{d^k1},2\pi (d^{k1})\frac{2\pi }{d})\end{array}$$ Consider the wavelet sets $`W_k^d`$ and $`\stackrel{~}{W_k^d}`$, then we have the following: $$\begin{array}{c}W_k^d\stackrel{~}{W_k^d}=[\frac{2\pi }{d},\frac{d^{k2}(d1)2\pi }{d^k1})\hfill \\ \hfill [\frac{d^{k2}(2\pi )}{d^k1},\frac{2\pi }{d}\frac{(d1)2\pi }{d(d^k1)})[2\pi d^{k2}\frac{2\pi }{d^2},2\pi d^{k2}+\frac{d^{k2}2\pi }{d^k1})\end{array}$$ $$\begin{array}{c}W_k^d\stackrel{~}{W_k^d}=[\frac{d^{k1}(d1)2\pi }{d^k1},\frac{2\pi }{d})\hfill \\ \hfill [2\pi d^{k2}\frac{d^{k2}(d1)2\pi }{d^k1},2\pi d^{k2}\frac{2\pi }{d^2})\end{array}$$ $$\begin{array}{c}\stackrel{~}{W_k^d}W_k^d=[\frac{d^{k2}(d1)2\pi }{d^k1},\frac{2\pi }{d^2})\hfill \\ \hfill [2\pi d^{k1}\frac{d^{k1}(d1)2\pi }{d^k1},2\pi d^{k1}\frac{2\pi }{d})\end{array}$$ This gives us the following map $`\sigma (\xi ):W_k^d\stackrel{~}{W_k^d}`$: $$\sigma (\xi )=\{\begin{array}{cc}\xi ,\hfill & \xi W_k^d\stackrel{~}{W_k^d},\hfill \\ \xi +d^{k1}2\pi ,\hfill & \xi [\frac{d^{k1}(d1)2\pi }{d^k1},\frac{2\pi }{d}),\hfill \\ \xi d^{k2}2\pi ,\hfill & \xi [2\pi d^{k2}\frac{d^{k2}(d1)2\pi }{d^k1},2\pi d^{k2}\frac{2\pi }{d^2}).\hfill \end{array}$$ As with the case of dilation by two, it is routine to check that $`\sigma `$ is involutive. Additionally, we need to define functions $`h_1`$ and $`h_2`$ for the interpolation; define the functions $`h_1`$ on $`W_k^d`$ and $`h_2`$ on $`\stackrel{~}{W_k^d}`$ as follows: $`h_1(\xi )`$ $`=\{\begin{array}{cc}1,\hfill & \xi W_k^d\stackrel{~}{W_k^d},\hfill \\ \frac{1}{\sqrt{2\pi }},\hfill & \xi [\frac{d^{k1}(d1)2\pi }{d^k1},\frac{2\pi }{d}),\hfill \\ \frac{1}{\sqrt{2\pi }},\hfill & \xi [2\pi d^{k2}\frac{d^{k2}(d1)2\pi }{d^k1},2\pi d^{k2}\frac{2\pi }{d^2}),\hfill \end{array}`$ $`h_2(\xi )`$ $`=\{\begin{array}{cc}1,\hfill & \xi W_k^d\stackrel{~}{W_k^d},\hfill \\ \frac{1}{\sqrt{2\pi }},\hfill & \xi [\frac{d^{k2}(d1)2\pi }{d^k1},\frac{2\pi }{d^2}),\hfill \\ \frac{1}{\sqrt{2\pi }},\hfill & \xi [2\pi d^{k1}\frac{d^{k1}(d1)2\pi }{d^k1},2\pi d^{k1}\frac{2\pi }{d}),\hfill \end{array}`$ and extend via d-dilation periodicity. ###### Claim 2. Let $`W_k^d`$ and $`\stackrel{~}{W_k^d}`$, $`h_1`$ and $`h_2`$ be as above. Then the wavelet defined by $$\widehat{\psi }=h_1\widehat{\psi }_W+h_2\widehat{\psi }_W^{}$$ is an element of $`_{k2}`$ ###### Proof. As defined above, $`\widehat{\psi }`$ is a wavelet. Note that $`\sigma `$ above yields that $`d^{k2},d^{k1}_\psi `$, whence the claim is established by corollary 4. ∎ Hence, there exist wavelets in all $`_n`$’s, for dilation by $`d>2`$. We have now completed the proof of theorem 1. ∎ ###### Example 2. : $`d=3,k=4`$ $$m_4^3(\omega )=\{\begin{array}{cc}0,\hfill & \pi \omega <\frac{18\pi }{40},\hfill \\ 1,\hfill & \frac{18\pi }{40}\omega <\frac{6\pi }{40},\hfill \\ 2,\hfill & \frac{6\pi }{40}\omega <\frac{2\pi }{40},\hfill \\ 3,\hfill & \frac{2\pi }{40}\omega <\frac{\pi }{40},\hfill \\ 2,\hfill & \frac{\pi }{40}\omega <\frac{3\pi }{40},\hfill \\ 1,\hfill & \frac{3\pi }{40}\omega <\frac{9\pi }{40},\hfill \\ 0,\hfill & \frac{9\pi }{40}\omega <\frac{26\pi }{40},\hfill \\ 1,\hfill & \frac{26\pi }{40}\omega <\frac{27\pi }{40},\hfill \\ 0,\hfill & \frac{27\pi }{40}\omega <\pi .\hfill \end{array}$$ From the construction in the proof above, we get: $$E_4^3=[\frac{18\pi }{40},\frac{9\pi }{40})[\frac{26\pi }{40},\frac{27\pi }{40})[\frac{78\pi }{40},\frac{81\pi }{40})[\frac{234\pi }{40},\frac{243\pi }{40})$$ Since $$W_4^3=3E_4^3E_4^3,$$ the resulting wavelet set is: $$W_4^3=[\frac{54\pi }{40},\frac{18\pi }{40})[\frac{9\pi }{40},\frac{26\pi }{40})[\frac{702\pi }{40},\frac{729\pi }{40}).$$ If we take the same multiplicity function and do the second construction, we get: $$\begin{array}{c}\stackrel{~}{E_4^3}=[\frac{2\pi }{9},\frac{9\pi }{40})[\frac{26\pi }{40},\frac{27\pi }{40})[\frac{78\pi }{40},\frac{81\pi }{40})\hfill \\ \hfill [\frac{234\pi }{40},\frac{243\pi }{40})[\frac{702\pi }{40},\frac{160\pi }{9})\end{array}$$ The second wavelet set is: $$\stackrel{~}{W_4^3}=[\frac{2\pi }{3},\frac{2\pi }{9})[\frac{9\pi }{40},\frac{26\pi }{40})[\frac{160\pi }{9},\frac{729\pi }{40})[\frac{2106\pi }{40},\frac{160\pi }{3}).$$ Now we want to find a map $`\sigma (\xi ):W_4^3\stackrel{~}{W_4^3}`$. First, we will identify the intersection of the wavelet sets. $$W_4^3\stackrel{~}{W_4^3}=[\frac{2\pi }{3},\frac{18\pi }{40})[\frac{9\pi }{40},\frac{26\pi }{40})[\frac{160\pi }{9},\frac{729\pi }{40})$$ $$W_4^3\stackrel{~}{W_4^3}=[\frac{54\pi }{40},\frac{2\pi }{3})[\frac{702\pi }{40},\frac{160\pi }{9})$$ $$\stackrel{~}{W_4^3}W_4^3=[\frac{18\pi }{40},\frac{2\pi }{9})[\frac{2106\pi }{40},\frac{160\pi }{3})$$ We have: $$\sigma (\xi )=\{\begin{array}{cc}\xi ,\hfill & \xi W_4^3\stackrel{~}{W_4^3},\hfill \\ \xi +54\pi ,\hfill & \xi [\frac{54\pi }{40},\frac{2\pi }{3}),\hfill \\ \xi 18\pi ,\hfill & \xi [\frac{702\pi }{40},\frac{160\pi }{9}),\hfill \end{array}$$ Hence, by interpolating as above, we get a wavelet that is an element of $`_2`$ for dilation by 3. ## 4. Topology and Concluding Remarks We have shown the existence of wavelets with higher order translation invariance properties. Not much is known about them, other than what is known about the MSF wavelets ($`_{\mathrm{}}`$). For instance, the MSF wavelets are relatively closed in the collection of all wavelets in the $`L^2()`$ norm. We have the following analogue of the this fact: the collections $`_n`$ are relatively closed, but the collections $`_n`$ are not. We shall conclude the paper with some possible research directions. ###### Theorem 7. Suppose that $`\psi _n`$ is a sequence of wavelets that converges to a wavelet $`\psi `$. Suppose further that $`\{\psi _n\}_n_j`$. Then $`\psi _j`$. ###### Proof. We shall prove by contrapositive. Suppose that $`\psi _j`$, then there exists an $`l_\psi `$ that is not divisible by $`d^j`$. We shall show that there exists a positive integer $`N`$ such that for $`nN`$, $`\psi _n_j`$, by showing that $`l_{\psi _n}`$. Note that $`\psi _n`$ converges to $`\psi `$ in measure. Let $`Fsupp(\widehat{\psi })`$ be such that $`F+2\pi lsupp(\widehat{\psi })`$. Choose $`GF`$, a set of non-zero measure, such that for $`\xi GG+2\pi l`$, $`|\widehat{\psi }(\xi )|>ϵ_1`$ for some $`ϵ_1>0`$. Let $`ϵ_2`$ be the measure of $`G`$. Choose $`ϵ<min(\frac{ϵ_1}{2},\frac{ϵ_2}{2})`$. Let $`N`$ be such that for $`nN`$, $`m(\{\xi :|\widehat{\psi }(\xi )\widehat{\psi }_n(\xi )|ϵ\})<ϵ`$. Let $`H=\{\xi G:|\widehat{\psi }(\xi )\widehat{\psi }_n(\xi )|<ϵ\}`$; note that $`m(H)>m(G)ϵ>\frac{m(G)}{2}`$. Likewise, define $`H^{}=\{\xi G+2\pi l:|\widehat{\psi }(\xi )\widehat{\psi }_n(\xi )|<ϵ\}`$; $`m(H^{})>\frac{m(G)}{2}`$. We have that $`H+2\pi lG+2\pi l`$ and $`H^{}G+2\pi l`$, and both sets have measure greater than half of $`G+2\pi l`$, whence they must intersect. Hence, if $`H`$ and $`H^{}`$ are subsets of $`supp(\widehat{\psi }_n)`$, then $`\widehat{\psi }_n_j`$. For $`\xi HH^{}`$, $$|\widehat{\psi }(\xi )||\widehat{\psi }_n(\xi )||\widehat{\psi }(\xi )\widehat{\psi }_n(\xi )|<ϵ<ϵ_1<|\widehat{\psi }(\xi )|,$$ therefore, $`|\widehat{\psi }_n(\xi )|>0`$, as required. ∎ We shall now demonstrate that the $`_n`$’s are not closed. In the following example, we show a sequence of wavelets in $`_1`$ that converges to a MSF wavelet in $`L^2()`$. This technique can be extended to any of the interpolation pairs we used in this paper. Recall from Example 1 that a wavelet was constructed out of an interpolation pair and the resulting wavelet was in $`_1`$. By perturbing one of the wavelet sets in an appropriate manner, a sequence of interpolation pairs can be constructed. The end result is a sequence of wavelets in $`_1`$ which converges to an MSF wavelet. We shall use the first wavelet set unchanged: $$\begin{array}{c}W=[4\pi ,\frac{24\pi }{7})[\frac{8\pi }{7},\pi )[\frac{6\pi }{7},\frac{4\pi }{7})\hfill \\ \hfill [\frac{4\pi }{7},\frac{6\pi }{7})[\pi ,\frac{8\pi }{7})[\frac{24\pi }{7},4\pi ).\end{array}$$ Let $`a_n`$ be a sequence such that $`\frac{4\pi }{7}<a_n<\frac{\pi }{2}`$ and $`lim_n\mathrm{}a_n=\frac{4\pi }{7}`$. Consider the sequence of generalized scaling sets (they satisfy the conditions of Merrill’s theorem): $$\begin{array}{c}E_n=[2\pi ,\frac{12\pi }{7})[\pi ,\frac{6\pi }{7})[a_n,\frac{4\pi }{7})\hfill \\ \hfill [\frac{6\pi }{7},\pi )[\frac{12\pi }{7},2\pi )[\frac{24\pi }{7},4\pi +a_n).\end{array}$$ The corresponding wavelet sets are: $$\begin{array}{c}W_n=[4\pi ,\frac{24\pi }{7})[2a_n,\pi )[\frac{6\pi }{7},a_n)\hfill \\ \hfill [\frac{4\pi }{7},\frac{6\pi }{7})[\pi ,\frac{8\pi }{7})[4\pi +a_n,4\pi )[\frac{48\pi }{7},8\pi +2a_n).\end{array}$$ It is easy to verify that for all $`n`$, $`W`$ and $`W_n`$ form an interpolation pair, where $`\sigma `$ is given by: $$\sigma (\xi )=\{\begin{array}{cc}\xi ,\hfill & \xi WW_n\hfill \\ \xi 4\pi ,\hfill & \xi [\frac{24\pi }{7},4\pi +a_n)\hfill \\ \xi +8\pi ,\hfill & \xi [\frac{8\pi }{7},2a_n)\hfill \end{array}$$ as in example 1. Note that $`W_nW`$ as $`n\mathrm{}`$, in the symmetric difference metric, whence the limit of the sequence of interpolated wavelets is precisely $`\psi _W`$. There are many open questions regarding wavelets with higher order translation invariance properties. The first is whether there exist “nice” wavelets with these properties. The wavelets constructed in this paper are non-MRA wavelets and are discontinuous in the frequency domain. As such, these particular wavelets are not suitable for applications. A second consideration is the extension of this to several dimensions. The characterization theorem (4) has a natural extension, but operator interpolation in several dimensions is difficult as finding interpolation pairs is non-trivial.
warning/0002/cond-mat0002006.html
ar5iv
text
# Onset of sliding friction in incommensurate systems ## Abstract We study the dynamics of an incommensurate chain sliding on a periodic lattice, modeled by the Frenkel Kontorova hamiltonian with initial kinetic energy, without damping and driving terms. We show that the onset of friction is due to a novel kind of dissipative parametric resonances, involving several resonant phonons which are driven by the (dissipationless) coupling of the center of mass motion to the phonons with wavevector related to the modulating potential. We establish quantitative estimates for their existence in finite systems and point out the analogy with the induction phenomenon in Fermi-Ulam-Pasta lattices. The possibility of measuring friction at the atomic level provided by the Lateral Force Microscopes and Quartz Crystal Microbalance has stimulated intense research on this topic. Phonon excitations are the dominant cause of friction in many cases. Most studies are carried out for one-dimensional non-linear lattices and in particular for the Frenkel-Kontorova (FK) model, where the surface layer is modeled by a harmonic chain and the substrate is replaced by a rigid periodic modulation potential. The majority examines the steady state of the dynamical FK model in presence of dissipation representing the coupling of phonons to other, undescribed degrees of freedom. We study the dynamics of an undriven incommensurate FK chain. Our aim is to ascertain whether the experimentally observed superlubricity can be due to the blocking of the phonon channels caused by an incommensurate contact of the two sliding surfaces. Therefore we do not include any explicit damping of the phonon modes, since we wish to find out if they can be excited at all by the motion of the center of mass (CM). In an earlier study, Shinjo and Hirano found a superlubric regime for this model, where the chain would slide indefinitely without dynamic friction but with a recurrent exchange of kinetic energy between CM and a single internal mode. We will show that their finding is oversimplified by either too short simulation times or too small system sizes. The inherent non-linear coupling of the CM to the phonons leads to an irreversible decay of the CM velocity, albeit with very long time scales in some windows. The dissipative mechanism is driven by the coupling of the CM to the modes with modulation wavevector $`q`$ or its harmonics, $`\omega _{nq}`$, and consists in a novel kind of parametric resonances with much wider windows of instabilities than those deriving from the standard Mathieu equation. The importance of harmonic resonances at $`\omega _{nq}`$ has been pointed out before, with the suggestion that they could be absent in finite systems due to the discreteness of the phonon spectrum. However, it has not been realized that they act as a driving term for the onset of dissipation via subsequent complex parametric excitations which we shall describe, establishing quantitative estimates for their existence in finite systems. A related mechanism has recently been identified in the resonant energy transfer in the induction phenomenon in Fermi-Ulam-Pasta lattices. We start with the FK hamiltonian $$=\underset{n=1}{\overset{N}{}}\left[\frac{p_n^2}{2}+\frac{1}{2}\left(u_{n+1}u_nl\right)^2+\frac{\lambda }{2\pi }\mathrm{sin}(\frac{2\pi u_n}{m})\right]$$ (1) where $`u_n`$ are the lattice positions and $`l`$ is the equilibrium spacing of the chain for $`\lambda =0`$, $`\lambda `$ being the strength of the coupling scaled to the elastic spring constant. We take an incommensurable ratio of $`l`$ to the period $`m`$ of the periodic potential, namely $`m=1`$, $`l=(1+\sqrt{5})/2`$. We consider chains of $`N`$ atoms with periodic boundary conditions $`u_{N+1}=Nl+u_1`$. Hence, in the numerical implementation, we have to choose commensurate approximations for $`l`$ so that $`l\times N=M\times 1`$ with $`N`$ and $`M`$ integer, i.e. we express $`l`$ as ratio of consecutive Fibonacci numbers. The ground state of this model displays the so-called Aubry transition from a modulated to a pinned configuration above a critical value $`\lambda _c=0.14`$. Here we just note that in the limit of weak coupling ($`\lambda <<\lambda _c`$), deviations from equidistant spacing $`l`$ in the ground state are modulated with the substrate modulation wavevector $`q=2\pi l/m`$ as due to the frozen-in phonon $`\omega _q`$. Higher harmonics $`nq`$ have amplitudes which scale with $`\lambda ^n`$. We define the CM position and velocity as $`Q=\frac{1}{N}_nu_n`$, $`P=\frac{1}{N}_np_n`$. By writing $`u_n=nl+x_n+Q`$, the equations of motion for the deviations from a rigid displacement $`x_n`$ read $$\ddot{x}_n=x_{n+1}+x_{n1}2x_n+\lambda \mathrm{cos}(qn+2\pi x_n+2\pi Q)$$ (2) We integrate by a Runge Kutta algorithm the $`N`$ Eqs. (2) with initial momenta $`p_n=P_0`$ and $`x_n(t=0)`$ corresponding to the ground state. For a given velocity $`P`$, particles pass over maxima of the potential with frequency $`\mathrm{\Omega }=2\pi P`$, the so-called washboard frequency. In Fig. 1 we show the time evolution of the CM momentum for $`\lambda \lambda _c/3`$ and several values of $`P_0`$. According to the phase diagram of Ref. a superlubric behavior should be observed for this value of $`\lambda `$ and $`P_00.1`$. We find instead a non trivial time evolution with oscillations of varying period and amplitude and, remarkably, a very fast decay of the CM velocity for $`P_0\omega _q/(2\pi )`$ despite the absence of a damping term in Eq. (2). A similar, but much slower, decay is found for $`nP_0\omega _{nq}/(2\pi )`$. In the study of the driven underdamped FK it is shown that, at these superharmonic resonances, the differential mobility is extremely low. Here, we work out an analytical description in terms of the phonon spectrum which explains this complex time evolution and identifies the dissipative mechanism which is triggered by these resonances. In the limit of weak coupling $`\lambda `$ it is convenient to go from real to reciprocal space by defining Fourier transformed coordinates $`x_k=\frac{1}{N}_ne^{ikn}x_n`$ and $`x_n=_ke^{ikn}x_k`$, where $`k=2\pi n/N`$ and the normalization is chosen to remove the explicit $`N`$-dependence in the equations of motion, which become: $`\ddot{x}_k`$ $`=`$ $`\omega _k^2x_k+{\displaystyle \frac{\lambda }{2N}}{\displaystyle \underset{n}{}}e^{ikn}[e^{iqn}e^{i2\pi Q}e^{i2\pi x_n}+c.c.]`$ (4) $`\ddot{Q}`$ $`=`$ $`{\displaystyle \frac{\lambda }{2N}}{\displaystyle \underset{n}{}}[e^{i2\pi Q}e^{iqn}e^{i2\pi x_n}+c.c.]`$ (5) with $`\omega _k=2|\mathrm{sin}(k/2)|`$. We expand Eq. (4) in $`x_n`$ as: $$\ddot{x}_k=\omega _k^2x_k+\frac{\lambda }{2}\underset{m=0}{\overset{\mathrm{}}{}}\frac{(i2\pi )^m}{m!}\underset{k_1\mathrm{}k_m}{}\left[e^{i2\pi Q}x_{k_1}\mathrm{}x_{k_m}\delta _{k_1+\mathrm{}+k_m,q+k}+(1)^me^{i2\pi Q}x_{k_1}\mathrm{}x_{k_m}\delta _{k_1+\mathrm{}+k_m,q+k}\right]$$ (6) Since in the ground state the only modes present in order $`\lambda `$ are $`x_q=x_q=\lambda /2\omega _q^2`$ the CM is coupled only to these modes up to second order in $`\lambda `$: $`\ddot{Q}`$ $`=`$ $`i\lambda \pi \left(e^{i2\pi Q}x_qe^{i2\pi Q}x_q\right)`$ (8) $`\ddot{x}_q`$ $`=`$ $`\omega _q^2x_q+{\displaystyle \frac{\lambda }{2}}e^{i2\pi Q}`$ (9) $`\ddot{x}_q`$ $`=`$ $`\omega _q^2x_q+{\displaystyle \frac{\lambda }{2}}e^{i2\pi Q}`$ (10) In Fig. 2 we compare the behavior of $`P(t)=\dot{Q}(t)`$, obtained by solving the minimal set of Eqs. (1) with the appropriate initial conditions $`Q(t=0)=0`$, $`P(t=0)=P_0`$, $`x_q(t=0)=\lambda /(2\omega _q^2)`$, $`\dot{x}_q(t=0)=0`$, with the one obtained from the full system of Eqs. (2). Eqs. (1) reproduce very well the initial behavior of the CM velocity which displays oscillations of frequency $`\mathrm{\Delta }`$ around the value $`\mathrm{\Omega }/2\pi `$ but do not predict the decay occurring at later times because, as we show next, this is due to coupling to other modes. To this aim, we analyze the relation between the initial CM velocity $`P_0`$ and $`\mathrm{\Omega }/2\pi `$, respectively $`\mathrm{\Delta }_\pm `$. Take as an ansatz for the CM motion: $$Q(t)=\frac{\mathrm{\Omega }}{2\pi }t+\alpha _+\mathrm{sin}(\mathrm{\Delta }_+t)+\alpha _{}\mathrm{sin}(\mathrm{\Delta }_{}t)$$ (11) Inserting the ansatz (11) in the coupled set of Eqs. (1) keeping only terms linear in $`\alpha _\pm `$, we find that both $`\mathrm{\Delta }_\pm `$ are roots of: $$\mathrm{\Delta }^2=\lambda ^2\pi ^2\left(2Z(0)Z(\mathrm{\Delta })Z(\mathrm{\Delta })\right)$$ (12) $`Z(\mathrm{\Delta })`$ being the impedance $$Z(\mathrm{\Delta })=\frac{1}{\omega _q^2(\mathrm{\Omega }+\mathrm{\Delta })^2}\text{.}$$ (13) In general Eq. (12) has (besides the trivial solution $`\mathrm{\Delta }=0`$) indeed two solutions, related to the sum and difference of the two basic frequencies in the system, $`\omega _q`$ and $`\mathrm{\Omega }`$ : $$\mathrm{\Delta }_\pm |\omega _q\pm \mathrm{\Omega }+\frac{\lambda ^2\pi ^2}{2\omega _q(\mathrm{\Omega }\pm \omega _q)^2}+\mathrm{}|$$ (14) Close to resonance, $`\mathrm{\Omega }\omega _q`$, the amplitude $`\alpha _{}`$ dominates (see below) and the CM oscillates with a single frequency $`\mathrm{\Delta }=\mathrm{\Delta }_{}`$ (see Fig. 2). Very close to resonance (more precisely $`\omega _q<\mathrm{\Omega }<\omega _q+(2\lambda ^2\pi ^2/\omega _q)^{\frac{1}{3}}`$), the root $`\mathrm{\Delta }_{}`$ becomes imaginary, signaling an instability. In fact the system turns out to be bistable as it can be seen in Fig. 3 by the jump in $`\mathrm{\Omega }(P_0)`$ as $`P_0`$ passes through $`\omega _q/(2\pi )`$. Analytically, the relation between $`\mathrm{\Omega }`$ and $`P_0`$ and the amplitudes $`\alpha _\pm `$, is determined by matching the ansatz (11) with the initial condition: $$\alpha _\pm =\frac{\lambda ^2\pi \mathrm{\Omega }}{2\omega _q^2}\frac{1}{(\omega _q\pm \mathrm{\Omega })^3}$$ (15) $$P_0=\frac{\mathrm{\Omega }}{2\pi }+\frac{\lambda ^2\pi \mathrm{\Omega }}{2\omega _q^2}\left[\frac{1}{(\mathrm{\Omega }+\omega _q)^2}+\frac{1}{(\mathrm{\Omega }\omega _q)^2}\right]+\mathrm{}$$ (16) The fact that Eq. (16) has multiple solutions for $`\mathrm{\Omega }`$ when $`P_0\omega _q/2\pi `$ is in accordance with the jump seen in Fig. 3. However, Eq. (16), which is derived by keeping only linear terms in $`\alpha _{}`$, is not accurate enough to describe in detail the instability in the above range around $`\omega _q`$ where $`\alpha _{}`$ diverges. An initial behavior similar to that for $`P_0\omega _q/2\pi `$ is observed in Fig. 1 for $`nP_0\omega _{nq}/n2\pi `$. We examine the case $`n=2`$. Eq. (6) shows that $`x_{2q}`$ is driven in next order in $`\lambda `$ by $`x_q`$: $$\ddot{x}_{2q}=\omega _{2q}^2x_{2q}+i2\lambda \pi e^{i2\pi Q}x_q$$ (17) When $`2\pi Q\mathrm{\Omega }t`$, $`x_q`$ will be $`\lambda e^{i\mathrm{\Omega }t}`$, so that $`x_{2q}`$ is forced with amplitude $`\lambda ^2`$ and frequency $`2\mathrm{\Omega }`$, yielding resonance for $`2\mathrm{\Omega }=\omega _{2q}`$. Since $`x_{2q}`$ couples back to $`x_q`$, we have a set of equations similar to Eqs. (1), but at order $`\lambda ^2`$. We now come to the key issue, namely the onset of friction causing the decay of the CM velocity seen in Fig. 1 at later times, which cannot be explained by the coupling of the CM to the main harmonics $`nq`$. Since $`x_q`$ is by far the largest mode in the early stage, we consider second order terms involving $`x_q`$ in Eq. (6): $$\ddot{x}_k=\left[\omega _k^2+2\lambda \pi ^2\left(e^{i2\pi Q}x_q+e^{i2\pi Q}x_q\right)\right]x_k$$ (18) Insertion of the solution obtained above for $`x_q`$ (Eq. (9)) and $`Q`$ (Eq. (11)) yields $$\ddot{x}_k=\left[\omega _k^2+A+B\mathrm{cos}(\mathrm{\Delta }t)\right]x_k$$ (19) with $`A=2(\lambda \pi )^2/Z(0)`$ and $`B\alpha _{}`$. Clearly, Eq. (19) is a Mathieu parametric resonance for mode $`x_k`$. The relevance of parametric resonances has been recently stressed. However here, due to the coupling of the CM to the modulation mode $`q`$, resonances are not with the washboard frequency $`\mathrm{\Omega }`$ but with $`\mathrm{\Delta }\mathrm{\Omega }\omega _q`$. Hence, we find instability windows around $`\omega _k^2+A=(n\mathrm{\Delta }/2)^2`$. Since $`\mathrm{\Delta }`$ is small close to resonance, one expects to find instabilities for acoustic modes with $`k`$ small. Indeed, as shown in Fig. 4a, we find by solving Eq. (2) that the decay of the CM is accompanied by the exponential increase of the modes $`k=2,3,4`$ and, with a longer rise time, $`k=1`$. However, the instability windows resulting from Eq. (19), shown in Fig. 4b, cannot explain the numerical results of Fig. 4a, i.e. the Mathieu formalism cannot explain the observed instability. In Eq. (6), the only linear terms left out in Eq. (18) are couplings with $`x_{k\pm q}`$, which are much higher order in $`\lambda `$. Nevertheless, these terms are crucial since they may cause new instabilities due to the fact that, for $`k`$ small, they are also close to resonance. We have solved the coupled set of equations for mode $`x_{\pm k}`$ and $`x_{k\pm q}`$ : $`\ddot{x}_k`$ $`=`$ $`\left[\omega _k^2+2\lambda \pi ^2\left(e^{i2\pi Q}x_q+e^{i2\pi Q}x_q\right)\right]x_k+i\lambda \pi \left(e^{i2\pi Q}x_{kq}e^{i2\pi Q}x_{k+q}\right)`$ (21) $`\ddot{x}_{k\pm q}`$ $`=`$ $`\left[\omega _{k\pm q}^2+2\lambda \pi ^2\left(e^{i2\pi Q}x_q+e^{i2\pi Q}x_q\right)\right]x_{k\pm q}\pm i\lambda \pi e^{\pm i2\pi Q}x_k`$ (22) together with Eqs. (1) for continuous $`k`$. Indeed, we find a wider range of instabilities, giving a detailed account of the numerical result as shown in Fig. 4b. This mechanism where a parametric resonance is enhanced by coupling to near resonant modes is quite general in systems with a quasi continuous spectrum of excitations and is related to the one proposed in explaining instabilities in the FPU chain in a different physical context. The number of particles in the chain is an important parameter. When this number is very small, the chain is in fact commensurate and the phase of the CM is locked (the gap scales as $`\lambda ^N`$ due to Umklapp terms). Next, one enters a stage of apparent superlubric behavior due to the fact that the spectrum is still discrete on the scale of the size of the instability windows discussed above. For $`N=144`$ and $`\lambda =\frac{1}{3}\lambda _c`$ (Fig. 1) we only begin to see the decay for values of $`P_0`$ close to resonances. The experimentally observed superlubricity in could then be due either to the finiteness of the system or to the low sliding velocities. The above described multiple parametric excitation gives rise to an effective damping for the system via a cascade of couplings to more and more modes via the non-linear terms in Eq. (6). It remains an open question if this mechanism will eventually lead to a full or partial equilibrium distribution of energy over the normal modes although our preliminary results support the former hypothesis even at weak couplings. In summary, we have described in detail the mechanism which gives rise to friction during the sliding of a harmonic system onto an incommensurate substrate. The onset of friction occurs in two steps: the resonant coupling of the CM to modes with wavevector related to the substrate modulation leads to long wavelength oscillations which in turn drive a complex parametric resonance involving several resonant modes. This mechanism is robust in that it leads to wide instability windows and represents a quite general mechanism for the onset of energy transfer in systems with a quasi continuous spectrum of excitations. We are grateful to Ted Janssen for many constructive discussions and for his support.
warning/0002/hep-th0002132.html
ar5iv
text
# Geodesics and Newton’s Law in Brane Backgrounds ## 1 Introduction It has recently been suggested by Randall and Sundrum that four-dimensional gravity can arise at long distances on a brane embedded in a five-dimensional anti-de Sitter space. In their model the fifth dimension is non-compact. An effective dimensional reduction occurs because the metric perturbations admit a bound state solution which looks like a four-dimensional graviton bound to the brane. Earlier work appeared in . This interesting alternative to compactification has been discussed in a number of recent papers . The metric of the Randall-Sundrum (RS) background has the form $$ds^2=e^{2k|y|}\eta _{\mu \nu }dx^\mu dx^\nu +dy^2,$$ (1) where $`\eta _{\mu \nu }=\text{diag}(1,1,1,1)`$ ($`\mu ,\nu =0,1,2,3`$). It was argued in that the Kaluza-Klein excitations, although they are light, are suppressed near the brane and almost decouple from the matter fields. Moreover, it is assumed that matter fields are trapped to the brane by a certain mechanism. If the brane describes our Minkowski space-time, then there should exist trajectories for free massive particles located on the brane only. However, this is not true for the metric (1). The trajectory of the free particle in the metric (1) has the form (see Sec. 2 for details, here we take $`y_0=\dot{y}_0=0`$) $$x^\mu =x_0^\mu +v^\mu t,y=\frac{1}{2k}\mathrm{ln}(1v^2k^2t^2)$$ (2) Therefore, free massive ($`v^2<0`$) particles cannot move in Minkowski space-time without being ineviatably expelled into the $`y`$-dimension. This seems rather undesirable, even if for a very small (Planck scale) $`k`$ the time needed for a significant deviation in the fifth direction will be rather large. Thus, in the RS background, some other, non-gravitational mechanism is needed in order to trap matter on the brane. The simplest way to obtain an attractive brane would be to change the sign of the brane tension, although one might argue that this would imply other undesirable features. This alternative was considered in . In the present paper, we shall study the RS background and the alternative possibility, whose background metrics are given by $$ds^2=\mathrm{e}^{2k|y|}\eta _{\mu \nu }dx^\mu dx^\nu +dy^2.$$ (3) The upper or lower signs correspond to the RS and alternative backgrounds, respectively. This convention shall be used throughout this paper, and $`k`$ is positive, $`k>0`$. The metric (3) is a solution to Einstein’s equation for the action $$S=d^4x𝑑y\sqrt{g}(R2\mathrm{\Lambda })+\sigma _{y=0}d^4x\sqrt{g_B},$$ (4) where the cosmological constant and brane tension are $$\mathrm{\Lambda }=6k^2,\sigma =12k.$$ (5) We shall now give a brief outline of the rest of the paper. First, in Sec. 2 we study the geodesics in the two backgrounds and find that only in the alternative background gravity provides a mechanism for the trapping of matter on the brane. Secondly, in Sec. 3 we consider the linearized gravity equations, which shall be used in Sec. 4 to derive the Newtonian limit for gravity on the brane. We find that in both backgrounds the gravitational potential for a static point source will be $`1/r`$, and we find an exact formula for the corrections. In Sec. 5, we give expressions for the graviton modes in the both backgrounds. Space-like modes are absent in the RS background, but our results are inconclusive for the alternative background. ## 2 Geodesics In this section, we explicitely solve the geodesic equation in the RS and alternative backgrounds. We shall find that, in the RS background, ordinary matter will be expelled from the brane, but in the alternative background, the brane is attractive. Using the zeroth order terms of the connections given in the appendix, the geodesic equation takes the form $`{\displaystyle \frac{d^2x^\mu }{d\theta ^2}}2k\mathrm{sgn}y{\displaystyle \frac{dx^\mu }{d\theta }}{\displaystyle \frac{dy}{d\theta }}`$ $`=0,`$ (6) $`{\displaystyle \frac{d^2y}{d\theta ^2}}\pm k\mathrm{sgn}y\mathrm{e}^{2k|y|}\eta _{\mu \nu }{\displaystyle \frac{dx^\mu }{d\theta }}{\displaystyle \frac{dx^\nu }{d\theta }}`$ $`=0.`$ (7) We start by integrating eqn. (6), which yields $$\frac{dx^\mu }{d\theta }=v^\mu \mathrm{e}^{\pm 2k|y|},$$ (8) where $`v^\mu `$ is a constant four-vector. Eqn. (7) is explicitely solved by the first integral of the geodesic equation, $$\left(\frac{dy}{d\theta }\right)^2+\mathrm{e}^{2k|y|}\eta _{\mu \nu }\frac{dx^\mu }{d\theta }\frac{dx^\nu }{d\theta }=C,$$ where $`C`$ is a constant. Hence, inserting eqn. (8), we find $$\left(\frac{dy}{d\theta }\right)^2+\mathrm{e}^{\pm 2k|y|}v^2=C,$$ (9) where $`v^2=v^\mu v^\nu \eta _{\mu \nu }`$. It is convenient to change the parameterization to a non-affine parameter $`t`$ such that $$\frac{dt}{d\theta }=\mathrm{e}^{\pm 2k|y|}.$$ Then, eqn. (8) becomes (notation $`\dot{x}=dx/dt`$) $$\dot{x}^\mu =v^\mu x^\mu =x_0^\mu +v^\mu t,$$ (10) which shows that we can choose a reference frame such that $`t=x^0`$, i.e. $`t`$ is the time on the brane. Moreover, eqn. (9) becomes $$\dot{y}^2\mathrm{e}^{\pm 4k|y|}+\mathrm{e}^{\pm 2k|y|}v^2=C,$$ (11) and we can determine $`C`$ from the initial data: $$C=\dot{y}_0^2\mathrm{e}^{\pm 4k|y_0|}+v^2\mathrm{e}^{\pm 2k|y_0|}.$$ (12) Before integrating eqn. (11), we note that it depends only on $`|y|`$, as long as we do not pass through $`y=0`$. Therefore, it is sufficient to consider $`y>0`$; replacing $`y`$ with $`|y|`$ at the end will take care of the case $`y<0`$. For $`y>0`$, we can integrate eqn. (11) and find $$\sqrt{Cv^2\mathrm{e}^{\pm 2ky}}=\pm v^2kt+\sqrt{Cv^2\mathrm{e}^{\pm 2ky_0}},$$ where we have again expressed the integration constant by the initial data. Notice, that the $`\pm `$ sign in front of the term containing $`t`$ on the right hand side is not (yet) related to the sign in the exponentials, but stems from the ambiguity of taking a square root. After some simple steps involving the substitution of $`C`$ from eqn. (12) we obtain $$\mathrm{e}^{\pm 2ky_0}\mathrm{e}^{\pm 2ky}=v^2k^2t^2\pm |\dot{y}_0|2k\mathrm{e}^{\pm 2ky_0}t.$$ From the initial data we can deduce that we have to replace $`\pm |\dot{y}_0|`$ by $`\dot{y}_0`$. Thus, the final result is $$\mathrm{e}^{\pm 2k|y|}=\mathrm{e}^{\pm 2k|y_0|}\pm 2k\dot{y}_0\mathrm{e}^{\pm 2k|y_0|}tv^2k^2t^2.$$ (13) The solution (13) is valid, as long as $`|y|0`$. If we hit the brane at $`y=0`$, we have to match a solution for $`y>0`$ with a solution for $`y<0`$. However, from eqn. (7) we see that the velocity $`\dot{y}`$ must be continuous at $`y=0`$, since the second term in that equation is finite. Thus, the brane will not deflect particles gravitationally, but we might expect that non-gravitational interactions with matter on the brane might do so. The interpretation of the solution (13) is rather simple: In the RS background, which corresponds to the upper sign, ordinary matter ($`v^2<0`$) is expelled from the brane. On the other hand, tachyonic particles ($`v^2>0`$) are attracted to the brane, whereas massless particles are not affected by its presence. Using the lower sign, the brane attracts ordinary matter. This is sketched in Fig. 1. Another interesting fact is that, for the alternative background, $`\mathrm{e}^{2k|y|}=0`$ corresponds to $`|y|=\mathrm{}`$. Thus, tachyonic particles will be expelled to $`|y|=\mathrm{}`$ in finite brane time, as will massless particles with the right initial conditions. Moreover, there exist initial conditions for ordinary particles, which will yield $`|y|=\mathrm{}`$ in finite brane time. ## 3 Linearized Gravity In this section, we shall study the linearized gravity equations with two applications in mind: The derivation of Newton’s law on the brane and the study of graviton modes, which will be carried out in Secs. 4 and 5, respectively. For our purpose, we introduce a matter perturbation on the brane, $$\delta T_{00}=\delta (y)t_{00}(x),$$ (14) and solve the linearized gravity equations for this source. The form (3) of the background metric suggests to use the time slicing formalism for calculating metric perturbations, although we do not slice with respect to time, but with respect to the transverse coordinate $`y`$. Let us first give some useful formulae. In the time slicing formalism, we split up the metric tensor as $$g_{ab}=\left(\begin{array}{cc}g_{\mu \nu }& n_\nu \\ n_\mu & n_\mu n^\mu +n^2\end{array}\right),g^{ab}=\frac{1}{n^2}\left(\begin{array}{cc}n^2g^{\mu \nu }+n^\mu n^\nu & n^\nu \\ n^\mu & 1\end{array}\right),$$ (15) where $`a,b=0,1,2,3,5`$, $`x^5=y`$, and $`g_{\mu \nu }(x,y)`$ are the induced metrics in the hypersurfaces with internal coordinates $`x^\mu `$. The quantities $`n`$ and $`n^\mu `$ are called lapse function and shift vector, respectively, and are fixed to their respective background values in the radiation gauge: $$n^\mu =0,n=1.$$ (16) Thus, we consider a metric of the form $$ds^2=g_{\mu \nu }dx^\mu dx^\nu +dy^2.$$ (17) Then, the second fundamental form measuring the extrinsic curvature on the hypersurfaces is given by $$H_{\mu \nu }=\frac{1}{2n}\left(\frac{}{y}g_{\mu \nu }_\mu n_\nu _\nu n_\mu \right)=\frac{1}{2}\frac{}{y}g_{\mu \nu },$$ (18) where $`_\mu `$ is the covariant derivative on the hypersurfaces, and the second equality holds in the radiation gauge. Einstein’s equation is $$R_{ab}\frac{1}{2}g_{ab}R=g_{ab}\mathrm{\Lambda }+8\pi T_{ab},$$ (19) where $`T_{ab}=\overline{T}_{ab}+\delta T_{ab}`$, and $`\overline{T}_{ab}`$ is the background from the brane, whose non-zero components are found from eqn. (4) as $$\overline{T}_{\mu \nu }=\frac{3k}{4\pi }\delta (y)g_{\mu \nu }.$$ (20) One can observe from eqns. (14) and (20) that $`T_{a5}=0`$. Therefore, by virtue of the Gauss-Codazzi equations , the normal and mixed components of eqn. (19) become $`\widehat{R}+H_\nu ^\mu H_\mu ^\nu H^2`$ $`=2\mathrm{\Lambda },`$ (21) $`_\mu H_\nu H_\mu ^\nu `$ $`=0,`$ (22) respectively, where $`H=H_\mu ^\mu `$, and $`\widehat{R}`$ is the intrinsic scalar curvature of the hypersurfaces. The tangential components of eqn. (19) simply read $`R_{\mu \nu }{\displaystyle \frac{1}{2}}g_{\mu \nu }R`$ $`=g_{\mu \nu }\mathrm{\Lambda }+8\pi T_{\mu \nu }.`$ (23) Eqn. (23) is the equation of motion for $`g_{\mu \nu }`$, whereas eqns. (21) and (22) are constraints. We linearize eqns. (21)–(23) around the background (3), for which purpose we use an induced metric of the form $$g_{\mu \nu }=\mathrm{e}^{2k|y|}(\eta _{\mu \nu }+\gamma _{\mu \nu }),$$ (24) where $`\gamma `$ is a small perturbation compared to $`\eta `$. The indices of $`\gamma `$ shall be raised and lowered using the Lorentz metric $`\eta `$. Some useful expressions for the connections and curvatures are given in the appendix. Eqns. (21)–(23) take the forms $`\mathrm{e}^{\pm 2k|y|}(\gamma _{}^{\mu \nu }{}_{,\mu \nu }{}^{}\mathrm{}\gamma )\pm 3k\mathrm{sgn}y\gamma _{,y}`$ $`=0,`$ (25) $`{\displaystyle \frac{1}{2}}_y(\gamma _{,\mu }\gamma _{}^{\nu }{}_{\mu ,\nu }{}^{})`$ $`=0,`$ (26) $`{\displaystyle \frac{1}{2}}\left(\gamma _{}^{\mu }{}_{\rho ,\mu \nu }{}^{}+\gamma _{}^{\mu }{}_{\nu ,\mu \rho }{}^{}\mathrm{}\gamma _{\nu \rho }\gamma _{,\nu \rho }\right){\displaystyle \frac{1}{2}}\eta _{\nu \rho }\left(\gamma _{}^{\mu \lambda }{}_{,\mu \lambda }{}^{}\mathrm{}\gamma \right)`$ $`+\mathrm{e}^{2k|y|}\left[{\displaystyle \frac{1}{2}}\gamma _{\nu \rho ,yy}+{\displaystyle \frac{1}{2}}\eta _{\nu \rho }\gamma _{,yy}\pm 2k\mathrm{sgn}y(\gamma _{\nu \rho ,y}\eta _{\nu \rho }\gamma _{,y})\right]`$ $`=8\pi \delta T_{\nu \rho },`$ (27) where the background has been cancelled using eqns. (5) and (20). Let us start by solving the constraints. First, from eqn. (26) we find $$\gamma _{}^{\nu }{}_{\mu ,\nu }{}^{}=\gamma _{,\mu }+\xi _\mu (x),$$ (28) where $`\xi _\mu `$ are functions of the brane coordinates $`x^\mu `$ only. Secondly, after substituting eqn. (28), eqn. (25) leads to $$\mathrm{e}^{\pm 2k|y|}\xi _{}^{\mu }{}_{,\mu }{}^{}\pm 3k\mathrm{sgn}y\gamma _{,y}=0.$$ (29) Thus, integrating eqn. (29) yields the trace $`\gamma `$ as $$\gamma =\frac{1}{6k^2}\xi _{}^{\mu }{}_{,\mu }{}^{}\left(\mathrm{e}^{\pm 2k|y|}1\right),$$ (30) where we have used the residual gauge freedom to impose $`\gamma =0`$ on the brane. We see from eqn. (30) that $`\gamma `$ is unbounded for the RS background, if $`\xi _{}^{\mu }{}_{,\mu }{}^{}0`$. Moreover, this observation is independent of whether we use the residual gauge freedom as indicated or not. We shall see later that we do not have the choice of setting $`\xi ^\mu =0`$, if a matter perturbation is present on the brane. This indicates that, in the RS background, the linear approximation is not consistent. However, this might be an artifact of the particular choice of gauge. For further discussion of this problem, see . Finally, substituting eqns. (14), (28) and (30) into the equation of motion (27), we obtain the equation $`\mathrm{}\gamma _{\nu \rho }+_y\left(\mathrm{e}^{2k|y|}\gamma _{\nu \rho ,y}\right)2k\mathrm{sgn}y\mathrm{e}^{2k|y|}\gamma _{\nu \rho ,y}\xi _{\nu ,\rho }\xi _{\rho ,\nu }`$ (31) $`+{\displaystyle \frac{1}{3}}\eta _{\nu \rho }\xi _{}^{\mu }{}_{,\mu }{}^{}+{\displaystyle \frac{1}{6k^2}}\left(\mathrm{e}^{\pm 2k|y|}1\right)\xi _{}^{\mu }{}_{,\mu \nu \rho }{}^{}\pm {\displaystyle \frac{2}{3k}}\delta (y)\eta _{\nu \rho }\xi _{}^{\mu }{}_{,\mu }{}^{}`$ $`=16\pi \delta (y)t_{\nu \rho }.`$ We can take the trace of eqn. (31) and find $$\xi _{}^{\mu }{}_{,\mu }{}^{}=8\pi kt,$$ (32) where $`t=t_\mu ^\mu `$. Thus, as indicated earlier, the four-divergence $`\xi _{}^{\mu }{}_{,\mu }{}^{}`$ is fixed by the content of matter perturbation on the brane. As the next step, we consider the discontinuity of eqn. (31) at $`y=0`$. One easily finds $$16\pi t_{\nu \rho }=\gamma _{\nu \rho ,y}|_{y=+0}\gamma _{\nu \rho ,y}|_{y=0}\pm \frac{2}{3k}\eta _{\nu \rho }\xi _{}^{\mu }{}_{,\mu }{}^{}.$$ (33) However, as the perturbation is symmetric around the brane, we need only look for solutions which are even in $`y`$. Therefore, we can consider eqn. (31) in the region $`y>0`$, and eqns. (33) and (32) provide the Neumann boundary condition $$\gamma _{\nu \rho ,y}|_{y=+0}=8\pi \left(t_{\nu \rho }\frac{1}{3}\eta _{\nu \rho }t\right).$$ (34) Consider now eqn. (31) for $`y>0`$. First, let us choose the vector $`\xi ^\mu `$ as $$\xi _\mu =8\pi k\frac{1}{\mathrm{}}_\mu t,$$ (35) which is consistent with the condition (32). Then, we shall write $$\gamma _{\nu \rho }=\pm \frac{4\pi }{3k}\left[\frac{1}{\mathrm{}}t_{,\nu \rho }\left(\mathrm{e}^{\pm 2ky}1\right)+2k^2\left(\eta _{\nu \rho }\frac{1}{\mathrm{}}t\frac{4}{\mathrm{}^2}t_{,\nu \rho }\right)\right]+\stackrel{~}{\gamma }_{\nu \rho }$$ (36) in order to obtain from eqn. (31) the following homogeneous equation for $`\stackrel{~}{\gamma }_{\nu \rho }`$: $$\mathrm{}\stackrel{~}{\gamma }_{\nu \rho }+_y\left(\mathrm{e}^{2ky}\stackrel{~}{\gamma }_{\nu \rho ,y}\right)2k\mathrm{e}^{2ky}\stackrel{~}{\gamma }_{\nu \rho ,y}=0.$$ (37) Moreover, from eqn. (30) we find that the trace $`\stackrel{~}{\gamma }0`$, and the Neumann boundary condition (34) yields $$\stackrel{~}{\gamma }_{\nu \rho ,y}|_{y=+0}=8\pi \left(t_{\nu \rho }\frac{1}{3}\eta _{\nu \rho }t+\frac{1}{3\mathrm{}}t_{,\nu \rho }\right).$$ (38) Notice that a trivial (zero) solution to the homogeneous equation (37) is not consistent with this boundary condition. In order to solve eqn. (37), let us Fourier transform with respect to the brane coordinates and change variables to $`z=\mathrm{e}^{\pm 2ky}`$. Then, eqn. (37) becomes $$\left(z^2_z^2z_z\frac{p^2}{4k^2}z\right)\stackrel{~}{\gamma }_{\nu \rho }=0,$$ (39) whose solution can be expressed in terms of Bessel functions . ## 4 Newton’s Law In order to derive Newton’s law on the brane, we have to look for a unique solution to the linearized Einstein equations in the presence of a static point source on the brane. In the last section, we presented the general formalism of linearized gravity. Let us now continue the solution for static point source. In order to obtain the Newtonian limit, we have to calculate $`\gamma _{00}(x,0)`$, since the gravitational potential is given by $$V=\frac{m}{2}\gamma _{00}.$$ (40) We need a second boundary condition for eqn. (37) in order to obtain a unique solution. We shall simply use $$\stackrel{~}{\gamma }_{\nu \rho }|_{y=+\mathrm{}}=0.$$ (41) For static potentials, we have $`p_0=0`$, and therefore $`p^20`$ in eqn. (39). In fact, we need only consider $`p^2>0`$, as the solution for $`p^2=0`$ can be reconstructed as the limit $`p^20`$. The solution of eqn. (39) for $`p^2>0`$ is $$\stackrel{~}{\gamma }_{\nu \rho }(p,y)=c_{\nu \rho }(p)\mathrm{e}^{\pm 2ky}\{\begin{array}{cc}\mathrm{K}_2\left(\mathrm{e}^{\pm ky}|p|/k\right),\hfill & \\ \mathrm{I}_2\left(\mathrm{e}^{\pm ky}|p|/k\right),\hfill & \end{array}$$ (42) where the choice between the two possible solutions is dictated by eqn. (41). We easily see that this amounts to choosing the solution with the $`\mathrm{K}`$ function for the RS background (upper sign), and the solution with the $`\mathrm{I}`$ function for the alternative background. Moreover, from eqn. (42) we find the first derivative, $$\stackrel{~}{\gamma }_{\nu \rho ,y}(p,y)=\pm |p|c_{\nu \rho }(p)\mathrm{e}^{\pm 3ky}\{\begin{array}{cc}\mathrm{K}_1\left(\mathrm{e}^{\pm ky}|p|/k\right),\hfill & \\ \mathrm{I}_1\left(\mathrm{e}^{\pm ky}|p|/k\right),\hfill & \end{array}$$ (43) which, combined with the boundary condition (38), yields the coefficients $$c_{\nu \rho }=\frac{8\pi }{|p|}\left[t_{\nu \rho }\frac{1}{3}t\left(\eta _{\nu \rho }\frac{p_\nu p_\rho }{p^2}\right)\right]\{\begin{array}{cc}[\mathrm{K}_1(|p|/k)]^1\hfill & \text{for }\sigma <0\text{,}\hfill \\ [\mathrm{I}_1(|p|/k)]^1\hfill & \text{for }\sigma >0\text{.}\hfill \end{array}$$ (44) Thus, inserting eqns. (42) and (44) into eqn. (36), we obtain the solution for the metric perturbation $`\gamma _{\nu \rho }(p,y)`$ $`=\pm {\displaystyle \frac{4\pi }{3k}}\left[{\displaystyle \frac{p_\nu p_\rho }{p^2}}\left(\mathrm{e}^{\pm 2ky}1\right)2k^2\left({\displaystyle \frac{\eta _{\nu \rho }}{p^2}}{\displaystyle \frac{4p_\nu p_\rho }{p^4}}\right)\right]t`$ (45) $`+{\displaystyle \frac{8\pi }{|p|}}\left[t_{\nu \rho }{\displaystyle \frac{1}{3}}t\left(\eta _{\nu \rho }{\displaystyle \frac{p_\nu p_\rho }{p^2}}\right)\right]\{\begin{array}{cc}\mathrm{e}^{2ky}\frac{\mathrm{K}_2\left(\mathrm{e}^{ky}|p|/k\right)}{\mathrm{K}_1(|p|/k)}\hfill & \text{for }\sigma <0\text{,}\hfill \\ \mathrm{e}^{2ky}\frac{\mathrm{I}_2\left(\mathrm{e}^{ky}|p|/k\right)}{\mathrm{I}_1(|p|/k)}\hfill & \text{for }\sigma >0\text{.}\hfill \end{array}`$ Let us now use a static point source, $`t_{00}(p)=2\pi \delta (p_0)a`$, where $`a=M/M_{Pl(5)}^3`$, and solve for $`\gamma _{00}(x,0)`$. Consider first the RS background. We can use the recursion formula for modified Bessel functions, $$\mathrm{K}_2(z)=\frac{2}{z}\mathrm{K}_1(z)+\mathrm{K}_0(z),$$ in order to separate the divergent term for $`|p|0`$ in the Fourier integral. Then, from eqn. (45) we find $`\gamma _{00}(x,0)`$ $`={\displaystyle \frac{2ka}{3r}}+{\displaystyle \frac{8ka}{3r}}+\underset{y0}{lim}{\displaystyle \frac{d^3p}{(2\pi )^3}\mathrm{e}^{ipx}\frac{16\pi a}{3|p|}\frac{\mathrm{K}_0(\mathrm{e}^{ky}|p|/k)}{\mathrm{K}_1(|p|/k)}}`$ $`={\displaystyle \frac{2ka}{r}}+{\displaystyle \frac{8a}{3\pi r^2}}\underset{y0}{lim}{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}𝑑s\mathrm{sin}s{\displaystyle \frac{\mathrm{K}_0(\mathrm{e}^{ky}s/kr)}{\mathrm{K}_1(s/kr)}}.`$ (46) It is interesting to note that the inhomogeneous terms of the solution would yield Newton’s law with a wrong sign, but the homogeneous part, whose presence is necessary because of the Neuman boundary condition, takes care of this. In fact, the importance of the boundary conditions for obtaining the gravity on the brane has already been pointed out in . Moreover, we observe that the integral in eqn. (46) is well-defined only for $`y>0`$. It is for this reason that we take the limit $`y0`$ after the integration. The non-zero $`y`$ acts as a regulator of the integral for large $`s`$. The second term in eqn. (46) is a correction to Newton’s law, which we shall demonstrate in a moment. Let us consider now the alternative background. From eqn. (45) we obtain $`\gamma _{00}(x,0)`$ $`={\displaystyle \frac{2ka}{3r}}+\underset{y0}{lim}{\displaystyle \frac{d^3p}{(2\pi )^3}\mathrm{e}^{ipx}\frac{16\pi a}{3|p|}\frac{\mathrm{I}_2(\mathrm{e}^{ky}|p|/k)}{\mathrm{I}_1(|p|/k)}}`$ $`={\displaystyle \frac{2ka}{3r}}+{\displaystyle \frac{8a}{3\pi r^2}}\underset{y0}{lim}{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}𝑑s\mathrm{sin}s{\displaystyle \frac{\mathrm{I}_2(\mathrm{e}^{ky}s/kr)}{\mathrm{I}_1(s/kr)}}.`$ (47) The first term in eqn. (47) represents Newton’s law, and the second term corrections, as we shall demonstrate now. Consider an integral of the form $`_0^{\mathrm{}}𝑑s\mathrm{sin}sf(s/z,y)`$, where $`f`$ is a differentiable and integrable function. From we know that the integrands in eqns. (46) and (47) satisfy this property for any $`y>0`$. Given the integrability of $`f`$, we can rewrite the integral as $$\underset{0}{\overset{\mathrm{}}{}}𝑑s\mathrm{sin}sf(s/z,y)=\underset{k=0}{\overset{\mathrm{}}{}}\underset{\pi }{\overset{\pi }{}}𝑑s(\mathrm{sin}s)f(\frac{\pi (2k+1)+s}{z},y).$$ (48) At this point, we can take the limit $`y0`$, and we shall write $`f(x,0)=f(x)`$. For large $`z`$, the argument of $`f`$ will change little in one period of the $`\mathrm{sin}`$ function, and we can write $`{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}𝑑s\mathrm{sin}sf(s/z)`$ $`{\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}{\displaystyle \underset{\pi }{\overset{\pi }{}}}𝑑s\mathrm{sin}s\left[f\left({\displaystyle \frac{\pi (2k+1)}{z}}\right)+{\displaystyle \frac{s}{z}}f^{}\left({\displaystyle \frac{\pi (2k+1)}{z}}\right)\right]`$ $`{\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{2\pi }{z}}f^{}\left({\displaystyle \frac{\pi (2k+1)}{z}}\right).`$ Here, we observe that the $`z\mathrm{}`$ limit exists, namely it is just the integral $$\underset{0}{\overset{\mathrm{}}{}}𝑑xf^{}(x)=f(0)f(\mathrm{}).$$ For both integrands under consideration we have $`f(0)=0`$ and $`f(\mathrm{})=1`$ (for $`y>0`$ we would have the stronger $`f(\mathrm{})=0`$). Thus, we find that the second terms in eqns. (46) and (47) go to zero at least as fast as $`1/r^2`$ for large $`r`$ and are in fact corrections to Newton’s law. Finally, from eqn. (40) we obtain the gravitational potential $$V(r)=\frac{kmM}{M_{Pl(5)}^3}\{\begin{array}{cc}1/r+𝒪(1/r^2)\hfill & \text{for }\sigma <0\text{,}\hfill \\ 1/(3r)+𝒪(1/r^2)\hfill & \text{for }\sigma >0\text{.}\hfill \end{array}$$ (49) Thus, one can read off the four-dimensional Planck masses as $$M_{Pl(4)}^2=\{\begin{array}{cc}M_{Pl(5)}^3/k\hfill & \text{for }\sigma <0\text{,}\hfill \\ 3M_{Pl(5)}^3/k\hfill & \text{for }\sigma >0\text{.}\hfill \end{array}$$ (50) Our result for the RS background is in agreement with the result given in . ## 5 Graviton Modes It is our objective in this section to study graviton modes. The equation of motion for gravitons in the gauge $`\delta g_{a5}=0`$ is the homogeneous equation (37) with the boundary condition $$\gamma _{\nu \rho ,y}|_{y=0}=0.$$ (51) This boundary condition stems from eqn. (34) and implies that we consider only modes which are even in $`y`$. This would be natural for the orbifold $`S^1/_2`$, but in general one would have to match a solution for $`y>0`$ with a solution for $`y<0`$. Let us now give the graviton solutions. As we consider only modes which are even in $`y`$, we restrict ourselves to $`y0`$. Using eqn. (39), the solutions to eqn. (37) satisfying the boundary condition (51) are $$\gamma _{\mu \nu }(p,y)=c_{\mu \nu }(p)\{\begin{array}{cc}\mathrm{e}^{\pm 2ky}\left[\mathrm{I}_2(\alpha \mathrm{e}^{\pm ky})\mathrm{K}_1(\alpha )+\mathrm{K}_2(\alpha \mathrm{e}^{\pm ky})\mathrm{I}_1(\alpha )\right]\hfill & \text{for }p^2>0\text{,}\hfill \\ 1\hfill & \text{for }p^2=0\text{,}\hfill \\ \mathrm{e}^{\pm 2ky}\left[\mathrm{J}_2(\alpha \mathrm{e}^{\pm ky})\mathrm{N}_1(\alpha )\mathrm{N}_2(\alpha \mathrm{e}^{\pm ky})\mathrm{J}_1(\alpha )\right]\hfill & \text{for }p^2<0\text{.}\hfill \end{array}$$ (52) Here, we have set $`\alpha =\sqrt{|p^2|}/k`$. Let us discuss the normalizability of these modes. For the RS background, one finds that space-like modes ($`p^2>0`$) are not normalizable, since $`\mathrm{I}_2`$ diverges as $`\mathrm{e}^{\mathrm{e}^y}`$ for large $`y`$. On the other hand, zero-like and time-like modes are normalizable. On the other hand, for the alternative background the $`y`$ integral diverges in all three cases just as the volume integral. Thus, just as in flat space, we might assume that we can form wave packets which describe normalizable wave functions. Alternatively, one might regularize integrals like the norm of wavefunctions or the effective action by dividing by the total volume of space. We would like to leave these points open to further research. ## 6 Conclusions In this paper, we have studied the geodesics and derived Newton’s law for the Randall-Sundrum and an alternative brane background. We found that matter will be expelled from the brane in the RS background. A similar behaviour was observed by Rubakov *et al.* in a different context. Therefore, the RS background is classically unstable, and it must be supplemented with a mechanism for confinement of matter. On the other hand, gravity provides a natural mechanism for the trapping of matter in the alternative background. It would be interesting to study a realization of the alternative background in supergravity. The RS background in supergravity has been discussed in . Our derivation of gravity on the brane revealed the validity of Newton’s law to leading order in both backgrounds, but the corresponding Planck masses on the brane are different. Moreover, we found exact formulas, eqns. (46) and (47), for the corrections. In the RS background there are no normalizable space-like modes, which certainly is in favour of this background. For the alternative background, all modes are non-integrable, but a thorough discussion of the confinement or non-confinement of gravity remains an open problem. It seems that the confinement of gravity cannot be discussed separately from the confinement of matter. ## Acknowledgements This research was partly supported by NSERC. I. V. is grateful to the Physics Department of Simon Fraser University for its kind hospitality. Moreover, we are grateful to I. Ya. Aref’eva and M. G. Ivanov for their help in the early stage of this work and to A. Linde and V. A. Rubakov for their critical remarks. ## Appendix We state here some expressions for the connections and curvature up to first order in the perturbations $`\gamma _{\mu \nu }`$. The only non-zero connections for the metric $`g_{ab}`$ are $`\mathrm{\Gamma }_{}^{\mu }{}_{\nu \lambda }{}^{}`$ $`={\displaystyle \frac{1}{2}}\left(\gamma _{}^{\mu }{}_{\nu ,\lambda }{}^{}+\gamma _{}^{\mu }{}_{\lambda ,\nu }{}^{}\gamma _{\nu \lambda }^{}{}_{}{}^{,\mu }\right),`$ (53) $`\mathrm{\Gamma }_{}^{y}{}_{\nu \lambda }{}^{}`$ $`=\pm k\mathrm{sgn}yg_{\nu \lambda }{\displaystyle \frac{1}{2}}\mathrm{e}^{2k|y|}\gamma _{\nu \lambda ,y},`$ (54) $`\mathrm{\Gamma }_{}^{\nu }{}_{\lambda y}{}^{}=\mathrm{\Gamma }_{}^{\nu }{}_{y\lambda }{}^{}`$ $`=k\mathrm{sgn}y\delta _\lambda ^\nu +{\displaystyle \frac{1}{2}}\gamma _{}^{\nu }{}_{\lambda ,y}{}^{}.`$ (55) Moreover, we find from eqn. (18) $$H_\nu ^\mu =k\mathrm{sgn}y\delta _\nu ^\mu +\frac{1}{2}\gamma _{}^{\mu }{}_{\nu ,y}{}^{},$$ (56) and some expressions for the curvatures are ($`\mathrm{}=\eta ^{mu\nu }_\mu _\nu `$) $`\widehat{R}_{\nu \rho }`$ $`={\displaystyle \frac{1}{2}}\left(\gamma _{}^{\mu }{}_{\nu ,\rho \mu }{}^{}+\gamma _{}^{\mu }{}_{\rho ,\nu \mu }{}^{}\gamma _{,\nu \rho }\mathrm{}\gamma _{\nu \rho }\right),`$ (57) $`\widehat{R}`$ $`=\mathrm{e}^{\pm 2k|y|}\left(\gamma _{}^{\mu \nu }{}_{,\mu \nu }{}^{}\mathrm{}\gamma \right),`$ (58) $`R_{}^{y}{}_{\nu y\rho }{}^{}`$ $`=k^2g_{\nu \rho }\pm 2k\delta (y)g_{\nu \rho }+\mathrm{e}^{2k|y|}\left(\pm k\mathrm{sgn}y\gamma _{\nu \rho ,y}{\displaystyle \frac{1}{2}}\gamma _{\nu \rho ,yy}\right),`$ (59) $`R_{\mu \rho }`$ $`=\widehat{R}_{\mu \rho }+H_\mu ^\nu H_{\nu \rho }HH_{\mu \rho }+R_{}^{y}{}_{\mu y\rho }{}^{},`$ (60) $`R`$ $`=\widehat{R}20k^2\pm 16k\delta (y)\pm 5k\mathrm{sgn}y\gamma _{,y}\gamma _{,yy}.`$ (61)
warning/0002/quant-ph0002038.html
ar5iv
text
# 1 Introduction ## 1 Introduction Since more than ten years, interference phenomena in open quantum systems are studied theoretically in the framework of different models. Common to all these studies is the appearance of different time scales as soon as the resonance states start to overlap (see and references in these papers to older ones). Some of the states align with the decay channels and become short-lived while the remaining ones decouple to a great deal from the continuum of decay channels and become long-lived. The wavefunctions show permanent changes: they are mixed strongly in the basic wavefunctions of the corresponding closed system. In many-body systems the interaction is caused, above all, by two-body forces between the constituents of the system. The additional interaction connected with avoided level crossings is believed, usually, to lead only to an exchange of the wavefunctions but not to permanent changes of their structure. This conclusion results from many spectroscopic studies on closed systems with discrete states. Recent investigations in the framework of a schematical model showed however that, in the case of collective resonance states, permanent changes in the wavefunctions occur due to the interplay between the real and imaginary parts of the different coupling matrix elements. The mixing of the resonance states of a microwave resonator is not caused by two-body forces. A mixing of the states may occur only as a result of avoided level crossings. It is therefore an interesting question whether or not some permanent mixing in the wavefunctions of a microwave cavity can arise. In , the changes in the structure of the wavefunctions at avoided crossings in a strongly driven (closed) square potential well system are studied. The avoided crossings are shown to lead, in some cases, to temporary and, in other cases, to permanent changes as a function of driving field strength. The avoided level crossings are related to exceptional points in the complex plane . The coupling induced by avoided level crossings is therefore surely connected with the coupling matrix elements of the discrete states of the closed system to the continuum. In the continuum shell model, these coupling matrix elements are complex . Interferences appear when more than one channel are open. It is possible therefore that interferences of different type may provide eventually a permanent mixing of the wavefunctions of the resonance states. A similar study for microwave cavities does not exist. It is the aim of the present paper to study the resonance picture of an open microwave resonator in detail. We show that its characteristic features are the same as those which are known from open many-body quantum systems. This means that not only two-body forces play a role for the interaction among the resonance states but also the interaction $`W`$ via the continuum is important. Neither $`\mathrm{}(W)`$ nor $`\mathrm{}(W)`$ can be neglected, generally. They are important near avoided crossings in the complex plane and their interplay can not be neglected when $`\mathrm{}(W)`$ and $`\mathrm{}(W)`$ are of the same order of magnitude. As a result, permanent changes in the structure of the wavefunctions appear, as a rule. The basic assumptions of the statistical theory (random matrix theory) are fulfilled only when the interferences caused by the hermitian and anti-hermitian parts of the Hamiltonian can be neglected to a good approximation. This is the case e.g. for a Gaussian orthogonal ensemble coupled weakly to the continuum. In Section 2 of the present paper, the Hamiltonian of an open quantum system and the relation of its eigenvalues to the poles of the $`S`$-matrix is considered. The formalism can be applied to a many-body system as well as to a microwave resonator. The Hamiltonian is non-hermitian and the eigenvalues provide the energies as well as the widths of the states. In Section 3, the avoided crossing of two resonance states is traced. The differences between the mixing of the states due to the hermitian and the anti-hermitian part of $``$ are the central point of discussion. The hermitian part causes an equilibration of the states in relation to the time scale which is accompanied by level repulsion along the real axis (energy). In contrast to this, the anti-hermitian part leads to an attraction of the levels in energy and to a bifurcation of the widths (formation of different time scales). These processes are characteristic for the interplay among resonances which takes place locally in more complicated systems . In Section 4, the resonance structure of a rectangular microwave resonator coupled to one lead is studied. Inside the resonator is a circular scatter. Level repulsion in the complex plane appears. It can be seen sometimes as level repulsion along the real energy axis. In other cases, a bifurcation of the widths occurs. The changes in the structure of the wavefunctions are permanent, as a rule. Collective states are formed at strong coupling to the lead. The structure of their wavefunctions has almost nothing in common with the structure of the wavefunctions of these states at small coupling. Together with the collective states, long-lived trapped states appear. The conductance of the microwave resonator is studied after coupling it to a second lead. The conductance peaks are determined by the poles of the $`S`$-matrix, which move as a function of the coupling strength between cavity and leads. The results are discussed in Section 5 and some conclusions are drawn in the last section. ## 2 The Hamilton operator of an open quantum system The function space of an open quantum system consists of two parts: the subspace of discrete states ($`Q`$-subspace) and the subspace of scattering states ($`P`$-subspace). The discrete states are the states of the closed system which are embedded into the continuum of scattering states. Due to the coupling of the discrete states to the continuum, they can decay and get a finite lifetime. Let us define two sets of wavefunctions by solving first the Schrödinger equation $`(H^{\mathrm{cl}}E_R^{\mathrm{cl}})\mathrm{\Phi }_R^{\mathrm{cl}}=0`$ for the discrete states of the closed system and secondly the Schrödinger equation $`(H^{cc}E^{(+)})\xi _E^c=0`$ for the scattering states of the environment. Note that the closed system can be a many-particle quantum system or a system like a microwave resonator. The only condition is that it can be described quantum mechanically by the hermitian Hamilton operator $`H^{\mathrm{cl}}`$. In the case of the flat microwave resonator, this is possible by using the analogy to the Helmholtz equation. Then, the $`Q`$ and $`P`$ operators can be defined by $`Q={\displaystyle \underset{R=1}{\overset{N}{}}}|\mathrm{\Phi }_R^{\mathrm{cl}}\mathrm{\Phi }_R^{\mathrm{cl}}|P={\displaystyle \underset{c=1}{\overset{\mathrm{\Lambda }}{}}}{\displaystyle _{ϵ_c}^{\mathrm{}}}𝑑E|\xi _E^c\xi _E^c|`$ (1) and $`Q\xi _E^c=0;P\mathrm{\Phi }_R^{\mathrm{cl}}=0`$. In order to perform spectroscopic studies, we do not use any statistical assumptions (for details see ). Assuming $`Q+P=1`$, we can determine a third wavefunction by solving the scattering problem $`(H^{cc}E^{(+)})\omega _R=_c\gamma _{Rc}\xi _E^c`$ with source term. The source term is determined by the coupling matrix elements $`\gamma _{Rc}`$ between the two subspaces. Further, we identify $`H^{\mathrm{cl}}`$ with $`_0QHQ`$ where $`(HE)\mathrm{\Psi }=0`$ is the Schrödinger equation in the total function space $`P+Q`$. Then, the solution $`\mathrm{\Psi }=P\mathrm{\Psi }+Q\mathrm{\Psi }`$ in the total function space is $`\mathrm{\Psi }`$ $`=`$ $`\xi _E^c+{\displaystyle \frac{1}{2\pi }}{\displaystyle \underset{R=1}{\overset{N}{}}}{\displaystyle \underset{R^{}=1}{\overset{N}{}}}(\mathrm{\Phi }_R^{\mathrm{cl}}+\omega _R)\mathrm{\Phi }_R^{\mathrm{cl}}|{\displaystyle \frac{1}{E}}|\mathrm{\Phi }_R^{}^{\mathrm{cl}}\gamma _{R^{}c}`$ (2) $`=`$ $`\xi _E^c+{\displaystyle \underset{R=1}{\overset{N}{}}}\stackrel{~}{\mathrm{\Phi }}_R{\displaystyle \frac{\stackrel{~}{\gamma }_{Rc}}{E\stackrel{~}{E}_R+\frac{i}{2}\stackrel{~}{\mathrm{\Gamma }}_R}}.`$ Here, $`=_0+W`$ (3) is the effective Hamilton operator appearing in the $`Q`$-subspace due to the coupling to the continuum, $`\stackrel{~}{\mathrm{\Phi }}_R`$ are the eigenfunctions of $``$ and $`\stackrel{~}{}_R\stackrel{~}{E}_R\frac{i}{2}\stackrel{~}{\mathrm{\Gamma }}_R`$ its eigenvalues. They provide the wavefunctions, energies and widths, respectively, of the resonance states. The $`\gamma _{Rc}`$ are the coupling matrix elements between the discrete states $`\mathrm{\Phi }_R^{\mathrm{cl}}`$ and the continuum of scattering states $`\xi _E^c`$, while the $`\stackrel{~}{\gamma }_{Rc}`$ are those between the resonance states $`\stackrel{~}{\mathrm{\Phi }}_R`$ and the continuum. The matrix elements of $`W`$ consist of the principal value integral and the residuum , $`W_{R^{}R}^{\mathrm{ex}}`$ $`=`$ $`{\displaystyle \frac{1}{2\pi }}{\displaystyle \underset{c=1}{\overset{\mathrm{\Lambda }}{}}}𝒫{\displaystyle \underset{ϵ_c}{\overset{\mathrm{}}{}}}𝑑E^{}{\displaystyle \frac{\gamma _{Rc}\gamma _{R^{}c}}{EE^{}}}{\displaystyle \frac{i}{2}}{\displaystyle \underset{c=1}{\overset{\mathrm{\Lambda }}{}}}\gamma _{Rc}\gamma _{R^{}c}.`$ (4) Here, $`c=1,\mathrm{},\mathrm{\Lambda }`$ are the channels which open at the energies $`ϵ_c`$. They describe the external mixing of two states via the continuum of decay channels. As a rule, both parts $`\mathrm{}(W)`$ and $`\mathrm{}(W)`$ are non-vanishing. Note, the expressions (2), (3) and (4) follow by formal rewriting the Schrödinger equation $`(HE)\mathrm{\Psi }=0`$ with the only condition that $`Q`$ and $`P=1Q`$ are defined in such a manner that the channel wavefunctions of the $`P`$ subspace are uncoupled . Otherwise, the eigenvalues and eigenfunctions of $``$ have no physical meaning. The $`\stackrel{~}{E}_R,\stackrel{~}{\mathrm{\Gamma }}_R,\stackrel{~}{\gamma }_{Rc}`$ and $`\stackrel{~}{\mathrm{\Phi }}_R`$ are energy dependent functions, generally. The resonance part of the $`S`$-matrix is $`S_{cc^{}}^{(\mathrm{res})}=i{\displaystyle \underset{R=1}{\overset{N}{}}}{\displaystyle \frac{\stackrel{~}{\gamma }_{Rc^{}}\stackrel{~}{\gamma }_{Rc}}{E\stackrel{~}{E}_R+\frac{i}{2}\stackrel{~}{\mathrm{\Gamma }}_R}}.`$ (5) We underline that the $`\stackrel{~}{\gamma }_{Rc},\stackrel{~}{E}_R`$, $`\stackrel{~}{\mathrm{\Gamma }}_R`$ and $`\stackrel{~}{\mathrm{\Phi }}_R`$ are functions which are calculated inside the formalism. They contain the contributions of $`\mathrm{}(W)`$ and of $`\mathrm{}(W)`$. The $`\stackrel{~}{\gamma }_{Rc}`$ and $`\stackrel{~}{\mathrm{\Phi }}_R`$ are complex. ## 3 Avoided crossing of two resonance states ### 3.1 Schematical study In order to illustrate the mutial influence of two neighboured resonance states, we consider the following Hamilton operator $`^{(v)}`$ $`=`$ $`\left(\begin{array}{cc}ϵ_1& v_{\mathrm{in}}\\ v_{\mathrm{in}}& ϵ_2\end{array}\right)\left(\begin{array}{cc}E_1& v_{\mathrm{in}}\\ v_{\mathrm{in}}& E_2\end{array}\right){\displaystyle \frac{i}{2}}\left(\begin{array}{cc}\mathrm{\Gamma }_1& 0\\ 0& \mathrm{\Gamma }_2\end{array}\right)`$ (12) which describes two resonance states lying at the energies $`E_1`$ and $`E_2`$. These two states have the widths $`\mathrm{\Gamma }_1`$ and $`\mathrm{\Gamma }_2`$, respectively, and are coupled by $`v_{\mathrm{in}}`$ (where $`v_{\mathrm{in}}`$ is real). The eigenvalues are: $$_\pm ^{(v)}=\frac{ϵ_1+ϵ_2}{2}\pm \frac{1}{2}\sqrt{(ϵ_1ϵ_2)^2+4v_{\mathrm{in}}^2}.$$ (13) When $`\mathrm{\Gamma }_1\mathrm{\Gamma }_2`$, the coupling $`v_{\mathrm{in}}`$ of the two states leads to level repulsion along the real axis. Numerical results for $`\mathrm{\Gamma }_1\mathrm{\Gamma }_2`$ are shown in Figure 1.a. Let us now consider the Hamiltonian with the coupling $`iw_{\mathrm{ex}}`$ ($`w_{\mathrm{ex}}`$ is real) of the two states via the continuum, $`^{(w)}`$ $`=`$ $`\left(\begin{array}{cc}ϵ_1& iw_{\mathrm{ex}}\\ iw_{\mathrm{ex}}& ϵ_2\end{array}\right)\left(\begin{array}{cc}E_1& 0\\ 0& E_2\end{array}\right){\displaystyle \frac{i}{2}}\left(\begin{array}{cc}\mathrm{\Gamma }_1& 2w_{\mathrm{ex}}\\ 2w_{\mathrm{ex}}& \mathrm{\Gamma }_2\end{array}\right).`$ (20) In this case, the eigenvalues are: $`_\pm ^{(w)}={\displaystyle \frac{ϵ_1+ϵ_2}{2}}\pm {\displaystyle \frac{1}{2}}\sqrt{(ϵ_1ϵ_2)^24w_{\mathrm{ex}}^2}.`$ (21) For $`E_1E_2`$, the coupling via the continuum due to $`iw_{\mathrm{ex}}`$ leads to repulsion along the imaginary axis (bifurcation of the widths), i.e. to resonance trapping. Numerical results for $`E_1E_2`$ are shown in Figure 1.d. As it is well known and can be seen from eq. (13), two interacting discrete states ($`v_{\mathrm{in}}0`$) can not cross. In the complex plane, however, the conditions for crossing of two resonance states may be fulfilled. From $`_+=_{}`$, it follows $`R`$ $``$ $`(E_1E_2)^2{\displaystyle \frac{1}{4}}(\mathrm{\Gamma }_1\mathrm{\Gamma }_2)^2+4(v_{\mathrm{in}}^2w_{\mathrm{ex}}^2)=0`$ $`I`$ $``$ $`(E_1E_2)(\mathrm{\Gamma }_1\mathrm{\Gamma }_2)+8v_{\mathrm{in}}w_{\mathrm{ex}}=0`$ (22) for the general case of a complex interaction $`v_{\mathrm{in}}+iw_{\mathrm{ex}}`$. These conditions define the critical values of the coupling strength ($`v_{\mathrm{in}}^{\mathrm{cr}}`$ and $`w_{\mathrm{ex}}^{\mathrm{cr}}`$, respectively) at which the $`S`$-matrix has a branch point . It is also possible that two resonance states cross along the real or imaginary axis while the crossing is avoided along the other axis. The conditions for such a case with $`I=0`$ are $`R<0`$ for crossing along the real axis and $`R>0`$ for crossing along the imaginary axis. The crossing in the complex plane is avoided, in any case. The trajectories for the motion of the eigenvalues in the complex plane as a function of the interaction $`v_{\mathrm{in}}`$ and $`w_{\mathrm{ex}}`$, respectively, (Figures 1.a and d) show the avoided crossing of the two resonance states in the complex plane. It occurs at a certain critical value of the coupling strength. Here and in its neighbourhood a redistribution between the two states takes place. It is accompanied by the biorthogonality of the eigenfunctions $`\stackrel{~}{\mathrm{\Phi }}_i`$ of $``$, $`B(1/2)_{i=1,2}\stackrel{~}{\mathrm{\Phi }}_i|\stackrel{~}{\mathrm{\Phi }}_i>1`$. The wavefunctions of the two resonance states become mixed: $`\stackrel{~}{\mathrm{\Phi }}_i=b_{ij}\mathrm{\Phi }_j^0`$ where the $`\mathrm{\Phi }_j^0`$ are the eigenfunctions of $`^0`$, which is the Hamilton operator with vanishing non-diagonal matrix elements ($`v_{\mathrm{in}}=0`$ and $`w_{\mathrm{ex}}=0`$, respectively). In Figures 1.b, c and e, f, the coefficients $`b_{ij}`$ are shown as a function of the coupling strength $`v_{\mathrm{in}}`$ and $`w_{\mathrm{ex}}`$, respectively. The results can be summarized in the following manner. * Weak interaction $`v_{\mathrm{in}}<v_{\mathrm{in}}^{\mathrm{cr}}`$ or $`w_{\mathrm{ex}}<w_{\mathrm{ex}}^{\mathrm{cr}}`$: the resonance states are isolated, their positions and widths are almost independent of the value of $`v_{\mathrm{in}}`$ and $`w_{\mathrm{ex}}^{\mathrm{cr}}`$, respectively. $`B1`$. * Starting at a certain critical value of the coupling strength, the behaviour of the system is different for the two cases with $`v_{\mathrm{in}}`$ and $`w_{\mathrm{ex}}`$. $`v_{\mathrm{in}}v_{\mathrm{in}}^{\mathrm{cr}}`$: the states start to repel in energy and to attract each other along the imaginary axis (if their widths at small $`v_{\mathrm{in}}`$ are different). $`w_{\mathrm{ex}}w_{\mathrm{ex}}^{\mathrm{cr}}`$: the widths bifurcate and the levels attract each other along the real axis (if their positions in energy at small $`w_{\mathrm{ex}}`$ are different). In both cases $`B1`$. The wavefunctions of the two states become mixed. * $`v_{\mathrm{in}}>v_{\mathrm{in}}^{\mathrm{cr}}`$: the widths are almost independent of $`v_{\mathrm{in}}`$: (equilibration in relation to the time scale) and the states repel each other along the real axis (level repulsion). $`w_{\mathrm{ex}}>w_{\mathrm{ex}}^{\mathrm{cr}}`$: the energies are almost independent of $`w_{\mathrm{ex}}`$ (the two states remain close to one another in their position) and resonance trapping occurs (the width of one of the states increases with increasing $`w_{\mathrm{ex}}`$, while the width of the other state decreases). In both cases $`B1`$. From a mathematical point of view, the properties of the system at an avoided crossing in the complex plane (i.e. in the region of the critical coupling strength $`v_{\mathrm{in}}^{\mathrm{cr}}`$ and $`w_{\mathrm{ex}}^{\mathrm{cr}}`$, respectively) are almost the same: repulsion of the eigenvalues along one axis and attraction along the other axis. The physical meaning is, however, very different: $`v_{\mathrm{in}}`$ causes equilibrium (in relation to the lifetime) and level repulsion along the real axis while $`iw_{\mathrm{ex}}`$ creates different time scales (bifurcation of the widths) and level attraction along the real axis. When the coupling contains both $`v_{\mathrm{in}}`$ and $`iw_{\mathrm{ex}}`$, then it depends on the ratio between the two parts whether level repulsion or attraction along the real axis dominates. In any case, the crossing of states is avoided in the complex plane and results in a complicated interference picture. The wavefunctions of the resonance states are mixed permanently in the set of the eigenfunctions of the Hamiltonian $`^0`$ of the corresponding closed system. ### 3.2 Realistic cases In , the behaviour of poles of the $`S`$-matrix in an open two-dimensional regular microwave billiard connected to a single waveguide is studied. As a function of the coupling strength between the resonator and the waveguide, the position of the correponding resonance poles, the wavefunctions of the resonance states and the Wigner-Smith time delay function are calculated. The poles are calculated on the basis of the exterior complex scaling method. The energy of the incoming wave is chosen so that only the channel corresponding to the first transversal mode in the lead is open. In , the bifurcation of the widths (resonance trapping) can be seen very clearly, indeed. In particular, the contraintuitive result that the lifetimes of certain resonance states increase with increasing coupling to the continuum can be traced not only in the motion of the poles of the $`S`$-matrix in the complex plane. It can be seen also in the wavefunctions of the resonance states and, above all, in the measurable time-delay function. In the case of three interfering resonance states, the wavefunction of (at most) one of the long-lived trapped states may be pure in relation to the bound states of the closed resonator . More exactly: $`b_{ii}=1`$ at small and large coupling strength. Some mixing of all three wavefunctions appears in the critical region where the wavefunctions are biorthogonal ($`B>1`$). Another example is the motion of the poles of the $`S`$-matrix by varying the coupling strength between the states of an atom by means of a laser. In , the positions and widths of two resonances in the vicinity of an autoionizing state coupled to another autoionizing one (or a discrete state) by a strong laser field are considered. For different atomic parameters, the trajectories in the complex energy plane are traced by fixing the field frequency $`\omega `$ but varying the intensity $`I`$ of the laser field. The states are coupled directly as well as via a common continuum and the ratio of these couplings is defined by the Fano parameter $`Q`$ . Most interesting is the region of avoided resonance crossing where the motion of each eigenvalue trajectory is influenced strongly by the motion of the other one. This occurs at a certain critical intensity $`I_{\mathrm{cr}}`$. When furthermore the frequency is equally to the critical value $`\omega _{\mathrm{cr}}`$, then laser induced degenerate states arising at the double pole of the $`S`$-matrix are formed. The strong correlation between the two states for intensities around $`I_{\mathrm{cr}}`$ reflects itself in the strong changes of the shape parameters of the resonances in the cross section. It can therefore be traced. In the limit of vanishing direct coupling ($`Q0`$), the widths bifurcate at $`I=I_{\mathrm{cr}}`$ as in other open quantum systems. That means, the width of one of the resonance states increases with increasing $`I>I_{\mathrm{cr}}`$ while the width of the other decreases relatively to the first one. In the limit $`I\mathrm{}`$, the ratio between the width of the long-lived and short-lived state approaches zero (resonance trapping). This corresponds to the situation shown in Fig. 1.d. In the other limiting case, the coupling via the continuum vanishes ($`Q`$ value large). Here, the levels repel in their energetic positions when $`II_{\mathrm{cr}}`$. This corresponds to Fig. 1.a. In any case, i.e. for all $`Q`$ values, the two resonance states start to repel each other in the complex energy plane at $`I=I_{\mathrm{cr}}`$. The repulsion of the eigenvalues in the complex plane is an expression of the strong mutual influence of one state on the other one in the critical region around $`I_{\mathrm{cr}}`$. In the transition region ($`Q`$ values of the order of magnitude 1), the trajectories show a complicated behaviour. Here, population trapping may appear, i.e. the width of one of the resonance states may vanish at a certain finite intensity $`I_{\mathrm{pt}}>I_{\mathrm{cr}}`$. It appears, generally, if the process is neither pure level repulsion on the real axis nor pure resonance trapping but the amplitudes of both processes (i.e. the direct coupling of the two states and their coupling via the continuum) are of comparable importance and interfere. Thus, the results obtained in for two interacting atomic levels confirm qualitatively the results of the schematical study represented in Section 3.1, although not only the non-diagonal matrix elements of $`H^{\mathrm{eff}}`$ but also the width $`\mathrm{\Gamma }_1`$ itself depends on $`I`$. These results and those for the microwave cavity discussed above show very clearly that individual resonance states can mix not only due to the two-body forces between the substituents of the system but also as a consequence of avoided resonance crossings. Other realistic cases are considered in . These are the resonance doublet $`J^\pi =2^+,T=0,\mathrm{\hspace{0.33em}1}`$ in the nucleus <sup>8</sup>Be and the $`\rho \omega ,T=1,\mathrm{\hspace{0.33em}0}`$ doublet of mesons. ## 4 Spectroscopic properties of an open microwave resonator ### 4.1 Calculations for the open microwave resonator The calculations are performed for a rectangular flat resonator coupled to a waveguide. Inside the cavity, a circular scatter is placed. We use the Dirichlet boundary condition, $`\mathrm{\Phi }=0`$, on the border of the billiard and of the waveguide. The waveguide has a width equal to $`1`$ and is attached to the resonator through a slide with an adjustable opening (which is described also by the Dirichlet boundary condition). For $`w=0.5`$ the resonator and the waveguide are disconnected, while $`w=0`$ represents the maximal coupling (opening). The cavity has a minimum area $`3\times 3`$ which is determined by $`x_r=1.5`$ and $`y_d=3`$ (compare Figure 2). The area is varied by varying $`x_r`$ or $`y_d`$ while both the position of the lead and the scatter inside the cavity remain unchanged. We solve the equation $`\mathrm{\Delta }\mathrm{\Phi }=E\mathrm{\Phi }.`$ Inside the waveguide, the wavefunction has the asymptotic form $`\mathrm{\Phi }=\left(e^{iky}R(E)e^{iky}\right)u(x).`$ Here $`u(x)`$ is the transversal mode in the waveguide, $`k`$ is the wave number and $`R(E)`$ is the reflection coefficient. The energies and widths of the resonance states are given by the poles of the coefficient $`R(E)`$ analytically continued into the lower complex plane. They are identical to the poles of the $`S`$ matrix when the fixed point equations for the $`\stackrel{~}{E}_R`$ and $`\stackrel{~}{\mathrm{\Gamma }}_R`$ are solved (see Section 2). To find the poles we use the method of exterior complex scaling . For details see . ### 4.2 Resonances as a function of the area of the resonator We studied the motion of the poles of the $`S`$ matrix as a function of the area of the resonator by changing both its length $`y`$ and width $`x`$. The changes of the corresponding wavefunctions $`\mathrm{\Phi }_R`$ are traced. We studied the energy region between the two thresholds at $`\pi ^210`$ and $`(2\pi )^240`$ where only one channel is open. In Figure 3, the eigenvalue picture is shown for $`w=0.15`$ (aperture partly closed by the slide) and $`w=0`$ (aperture fully open) for different values of the length and width of the resonator. In all cases, oscillations of the widths as a function of $`y_d`$ or $`x_r`$ in the energy region considered can be seen. The amplitudes of the oscillations are larger for larger widths. At $`w=0.15`$, all states corresponding to different $`y_d`$ ($`3>y_d>6`$) and lying around 24 have small widths. This is caused, obviously, by certain symmetry properties of the wavefunctions in relation to the channel (since this energy is in the middle between the two thresholds). The minimum in the widths vanishes when the wavefunctions are strongly mixed via the continuum ($`w=0`$). This shows that the coupling to the channel washes out some spectroscopic properties of the closed system. For $`w=0.15`$, the energies and widths of the states lying in the energy region around 24 are shown in Figure 4 as a function of $`y_d`$. The $`E_R(y_d)`$ show typical avoided crossings while the picture of the $`\mathrm{\Gamma }_R(y_d)`$ is more complicated. For the three states denoted by diamonds, stars and circles, respectively, the wavefunctions are shown in Figure 5 for 10 different neighboured values of $`y_d`$. Two states ($`B`$ and $`C`$) cross freely in the energy at $`E_R23`$. The wavefunctions of the two states $`B`$ and $`C`$ are very different from one another and the interaction due to $`\mathrm{}(W)`$ between them is obviously small. The wavefunctions of both states do almost not change in the crossing region. Only in the widths, some repulsion can be seen caused obviously by $`\mathrm{}(W)`$. This can be seen from Figure 6 where the results are shown from a calculation with smaller steps in $`y_d`$ around the free crossing. Around $`E_R=27`$, the state $`B`$ avoids crossing in energy with the other state ($`A`$) at some value $`y_d^{\mathrm{cr}}`$ (around -3.63). In this region, the wave functions of both states become strongly mixed, their widths become comparable and cross. The repulsion in their energies can be seen. The avoided crossing is caused mainly by $`\mathrm{}(W)`$. Beyond the critical region, the wavefunctions of the two states remain mixed although some hint to their exchange can be seen. These results show that avoided level crossing in the complex plane can be seen in the projection onto the energy axis or in the projection onto the width axis. In the first case, $`\mathrm{}(W)`$ dominates while in the second case, the mixing of the states occurs mainly due to $`\mathrm{}(W)`$. According to the oscillations of the widths and the varying number of states as a function of the length or width of the resonator, the sum of the widths of all states, lying between the two thresholds, fluctuates as a function of these values. In Figure 7 (bottom), we show the number $`N`$ of states as a function of $`y_d`$ (for $`w=0.15`$). This number increases since the number of states moving from above into the energy region considered is larger than the number of states leaving it to get bound. On the average, $`_R\mathrm{\Gamma }_R`$ is constant for a fixed value of $`w`$ with fluctuations smaller than 10 %. This can be seen from the example with $`w=0.15`$ shown in Figure 7 (top). The coupling of the cavity to the lead is therefore characterized by $`w`$ but not by the area of the cavity. ### 4.3 Resonances as a function of the coupling strength to the lead In Figure 8, we show the eigenvalue picture obtained by varying $`w`$ from 0.4 (almost closed aperture) to 0 (fully open aperture). The width of the resonator is determined by $`x_r=1.5`$ and the length by the two neighboured values $`y_d=3.34`$ (Figure 8 top) and $`y_d=3.28`$ (Figure 8 bottom). In both cases, collective states are formed. They are formed in regions where the level density is comparably high. Even at full opening of the aperture, the collective states belonging to the different groups do not overlap. Thus, they scarcely mix via the continuum. In Figure 9, we show the wavefunctions of the collective states from the lower part of Figure 8. Although the wavefunctions of the collective states are very different from one another at small opening of the aperture $`(w=0.4)`$, they are similar at full opening $`(w=0)`$ (Figures 9.d,f) where they have large amplitudes near to the aperture. The state shown in the middle (Figure 9.e) is trapped by the state to the left (Figure 9.d) at a comparably large opening (compare Figure 8 bottom). The wavefunctions of the collective states at $`w=0`$ in the long and in the broad resonators ($`x_r4.0,y_d6.0`$) are also similar to those shown in Figures 9.d and 9.f. ### 4.4 Resonances and conductance of the resonator The conductance of the resonator is described by the matrix elements $`S_{cc^{}}`$, equation (5), where $`c`$ is the channel of the incoming wave and $`c^{}`$ that of the outgoing wave. In our calculations, the second lead is on the lower right corner of the cavity, symmetrically to the first lead on the upper left corner. It is $`x_r=1.5`$ and $`y_d=3`$. In Figure 10, the conductance at three different coupling strengths to the leads is shown together with the eigenvalue picture. In the eigenvalue picture, one can see the formation of two short-lived states at large opening (small $`w`$) in each group. This corresponds to the coupling of the resonator to two leads. It can be seen from the wavefunctions of the states that both short-lived states of each group are coupled strongly to both leads. The conductance is therefore large at large opening. At low opening ($`w=0.4`$), the conductance peaks coincide with the resonance peaks. At larger opening ($`w=0.2`$ and 0), the conductance is an interference picture created by the overlapping resonances. The influence of the short-lived resonances onto the conductance can clearly be seen. In Figure 11, the conductance is integrated over the energy of each group ($`15E25`$ and $`25E40`$) and plotted as a function of $`w`$. The conductance at the higher energy increases quite rapidly in a small region of $`w`$ which corresponds to the critical region around $`w_{\mathrm{cr}}`$ (compare ). ## 5 Discussion of the results As demonstrated in Sections 3 and 4, the wavefunctions of a quantum system mix under the influence of the hermitian as well as the anti-hermitian part of the Hamiltonian. If the hermitian part of the Hamiltonian is dominant then avoided crossing can be seen along the energy axis (level repulsion). If the anti-hermitian part of the coupling via the continuum becomes important, then resonance trapping (bifurcation of the widths) appears. In general, both types of mixing appear and may interfere. Note that this interaction between different states of a quantum system via the continuum does not require two-body forces between the constituents of the system. The states whose wavefunctions are shown in Figure 5 lie in the energy region around 24 where the coupling to the continuum is small. The mixing of the states is varied by means of varying the area of the resonator. In the upper part of the related Figure 4, we see typical avoided level crossings in the energies $`E_R(y_d)`$. Here, the widths of the two states become comparable with one another. This implies that $`\mathrm{}(W)`$ is decisive for the process. In this case, the results are similar to those known very well from studies on closed systems with discrete states (see Section 3.1). We see, however, also the opposite case: the crossing of the states $`B`$ and $`C`$ in Figure 4 is free along the real axis while the widths repel each other. In this case, $`\mathrm{}(W)`$ is obviously small (the wavefunctions of the two states are very different from one another, Figure 5). Therefore, $`\mathrm{}(W)`$ is decisive, and the levels can, according to Section 3.1, cross along the real axis. The variation of the widths of the resonance states as a function of the coupling strength to the continuum is traced in Figure 8. In each group of overlapping states, one collective state is formed whose structure is determined by the channel wavefunction. This can be seen very clearly by comparing the wavefunctions of the different collective states which are similar to one another but have almost nothing in common with the original wavefunctions of these states at small opening of the aperture (Figure 9). Here, the variation of the external mixing occurs mainly in the $`\mathrm{}(W)`$: the approaching of the states of a group in their positions as well as the trapping of all but one state inside each group due to enlarging $`\mathrm{}(W)`$ can be seen very clearly in Figure 8. It is interesting to compare Figure 8 with the results for a slightly changed geometry of the cavity. In , the disk is smaller and all states between the two thresholds belong to one group. According to this, only one broad state is formed at full opening of the slide. The avoided crossing of the two broad states in the lower part of Figure 8 occurs according to the schematical picture with $`iw_{\mathrm{ex}}`$ (Figure 1.d, level attraction and width bifurcation) with the only difference that not only the non-diagonal matrix elements of $``$ depend on the coupling strength determined by $`w_{\mathrm{ex}}`$, but also the diagonal ones. This case is studied in detail analytically and numerically in the framework of a schematical model in . In , avoided level crossings in a closed resonator under the influence of a driving field are studied. The results show avoided level crossings with and without permanent mixing of the wavefunctions in a similar manner as in the open resonator studied by us. The relation between the peaks in the conductance, the Wigner delay times and the positions of the states in the closed resonator is studied in . The results of the present paper show that the conductance peaks are related to the positions of the resonance states in the open resonator. The peaks are, generally, the result of interferences between the resonance states. Altogether, the interplay between $`\mathrm{}(W)`$ and $`\mathrm{}(W)`$ leads, as a rule, to permanent mixings of the wavefunctions. Level repulsion along the real axis is caused by $`\mathrm{}(W)`$ while bifurcation of the widths (resonance trapping) is caused by $`\mathrm{}(W)`$. Both processes may interfere. As a result, different time scales may appear and the energy dependency of the conductance changes with the degree of opening of the system in a non-trivial manner. ## 6 Conclusions The interaction $`W`$ of resonance states via the continuum of decay channels consists of the hermitian part $`\mathrm{}(W)`$ and the anti-hermitian part $`\mathrm{}(W)`$. Both terms have to be considered not only in a many-body system but also in the micro-wave billiard as shown in the present paper. Some results show the dominance of $`\mathrm{}(W)`$, others the dominance of $`\mathrm{}(W)`$. The avoided crossing of the resonance states in the complex plane may appear, under certain conditions, as a free crossing along the real axis or along the imaginary axis. The interplay between the hermitian and anti-hermitian parts of the coupling operator between two resonance states via a common continuum may lead, in some cases, to a bifurcation of the widths (resonance trapping). In other cases, it may lead to a repulsion of the states along the real energy axis. The interaction $`W`$ introduces, as a rule, permanent changes in the wavefunctions of the resonance states. Under certain conditions, the system may be stabilized dynamically . The resonance picture of a microwave resonator shows all the characteristic features which are known from open quantum systems with two-body forces between the constituents. This result means that the interaction between the resonance states at the avoided level crossings in the complex plane plays an important role for the mixing of the wavefunctions. As an example, the wavefunctions of the collective short-lived states are strongly mixed in the set of wavefunctions of the closed system. They are quite different from those of the original states at small coupling to the continuum. The statistical theory (random matrix theory) describes resonance states of an almost closed system. The poles of the $`S`$-matrix are near to the real axis and $`\mathrm{}(W)`$ is small. The effective Hamilton operator is $`=_0+\mathrm{}(W)+\mathrm{}(W)=\mathrm{}()iVV^{}`$ where the $`V`$ are the coupling vectors between discrete and scattering states . The level repulsion along the real energy axis is embodied in $`\mathrm{}()`$ by choosing, e.g., the Gaussian orthogonal ensemble. Under these conditions, the effects caused by the interplay between $`\mathrm{}()`$ and $`\mathrm{}()`$ can be neglected to a good approximation. The results obtained in the present paper show, however, that the situation is different when the system is really open, i.e. when $`\mathrm{}()`$ and $`\mathrm{}()`$ are of the same order of magnitude. In this case, the interplay between the two parts of $``$ causes non-negligible effects which are not considered in the statistical theory. The avoided crossing of resonance states in the complex plane embodies the interplay between resonance trapping and level repulsion along the real axis. Acknowledgment: Valuable discussions with J. Burgdörfer, M. Müller, J. Nöckel, K. Richter and M. Sieber are gratefully acknowledged.
warning/0002/hep-ph0002074.html
ar5iv
text
# 1 Introduction ## 1 Introduction In recent times the interest in understanding hard processes involving nuclei, like electron nucleus collisions ($`e+A`$) and hadron nucleus collisions ($`h+A`$), has rapidly increased. The relativistic heavy ion collider (RHIC) will provide a new tool for experimental studies of $`h+A`$ and $`A+A`$ physics. This experimental advance provides a challenge to theorists for developing a description of such reactions strictly in terms of the underlying fundamental field theory, i.e. quantum chromodynamics (QCD). The usual way to proceed in QCD is to use factorization theorems that permit a rigorous treatment of hard processes. In this approach, incalculable soft contributions are factorized into universal, process independent distribution functions. Measurement of the distribution functions in one process then allows the prediction of cross sections for other processes. An example is the Drell-Yan (DY) process where two hadrons scatter to produce a lepton pair of large invariant mass $`Q^2`$. If $`Q^2`$ is large enough the factorization theorems state, that the inclusive Drell Yan cross section can be expressed as the product of a calculable hard part and universal twist-2 parton distribution functions that can be measured, for instance in deeply inelastic lepton nucleon scattering. Such universal distribution functions have been successfully extracted from existing data. The natural question arises how this picture is changed for hadron-nucleus collisions. Differences between the twist-2 distribution of a single nucleon and those of a nucleus of mass number $`A>1`$ were first discovered by the European Muon Collaboration. This phenomenon is known as the EMC effect at large values of the Bjorken scaling variable $`x`$ and as shadowing at lower values of $`x`$ . However, strong interaction dynamics cannot solely be described in terms of twist-2 structure functions. These functions are simple one-particle correlators and have the interpretation of a probability to find a parton in the hadron with certain momentum fraction $`x`$. These distribution functions are measured when the hard probe scatters off a single parton in the nucleon. Of course, the probe may scatter off more than one parton in the target. Such processes are related to multi-parton correlators and describe corrections to the single-parton scattering process. In general they are suppressed by powers of the large scale, and are called contributions of higher twist. If we consider processes involving large nuclei it is quite obvious that multiple scattering will be enhanced and even dominate in the limit $`A\mathrm{}`$. The classical example for enhancement of multiple scattering in $`h+A`$ processes is the Cronin effect, i.e. the anomalous $`A`$-dependence of the Drell-Yan cross section at large values of the transverse momentum of the lepton pair . Instead of a scaling of the cross section with the volume $`A`$ as expected in the twist-2 case, one observes an $`A`$-dependence parametrized by $`\sigma ^{h+A}=A^\alpha \sigma ^{h+p}`$ with $`\alpha 4/3`$ . This is a clear signal for multiple scattering and shows that higher-twist multiple scattering is grossly enhanced at large momentum transfer. The literature contains a wide variety of approaches dealing with multiple scattering, see e.g. , but a rigorous formulation in the language of perturbative QCD was attempted only recently. In a series of papers, Luo, Qiu and Sterman (LQS) and recently also Guo, showed how to extend the factorization theorem beyond leading twist to get a handle on double scattering in terms of perturbative QCD . Their twist-4 matrix elements are correlators of two partons in the nucleus and can describe scattering off partons from different nucleons. This approach predicts a stronger dependence on the size of the nucleus than the linear scaling with $`A`$ of single scattering. LQS argue that — due to the geometrical structure of the matrix elements — some of these twist-4 nuclear matrix elements pick up an additional factor $`A^{1/3}`$, compensating for some of the inherent suppression of higher twist corrections. This nuclear enhancement is able to describe the observed scaling with $`A^{4/3}`$ in certain kinematical regions. In this paper we discuss nuclear enhanced double scattering contributions to DY pair production off nuclear targets $$h(P_2)+A(P_1A)l^+l^{}+X$$ (1) in the framework of the LQS approach. It was already shown some time ago by Guo that higher-twist effects in the process (1) are enhanced at large transverse momentum. Extending the calculations of Guo , we have calculated the full kinematical dependence of the cross section $`\mathrm{d}\sigma /\mathrm{d}Q^2\mathrm{d}q_{}^2\mathrm{d}y\mathrm{d}\mathrm{\Omega }`$ including the angular distribution of the lepton pair and evaluated it, in particular, for $`p+A`$ scattering at RHIC. Our most important results were already published as a letter . In this contribution we present the complete analytical results together with a brief review of the LQS approach and some technical details. We also give some numerical results not contained in and an extended discussion. We will argue that nuclear enhanced double scattering corrections cannot be neglected at RHIC. In Fig. 1 we show typical Feynman diagrams contributing to the DY process. Single scattering means that, e.g., a quark from the single hadron and an antiquark from the nucleus annihilate into a virtual photon in a hard process. To obtain large transverse momentum an additional unobserved parton, here a gluon, has always to be radiated. Double scattering in our example implies that the quark can pick up an additional gluon from the nucleus. The generalized twist-4 factorization theorem of LQS states that two distinct scattering reactions occur for double scattering. One is referred to as double-hard (DH) scattering. Here the parton from the single hadron undergoes two subsequent independent hard scattering reactions in the target. The other one is called soft-hard (SH) scattering. Here the parton from the single hadron interacts with the collective soft gluon field in the nucleus (it can e.g. be thought of as feeling the color-magnetic Lorentz force) before producing the virtual photon via the final hard scattering. At large transverse momentum interference of these two mechanisms can be neglected. The double-hard reaction resembles the classical double scattering picture, while soft-hard scattering may have an analogy in plasma physics, where a hard scattering event can be accompanied by initial- or final-state interaction with the mean field. The first process might be accounted for by event generators in simulations of heavy-ion reactions whereas the implementation of soft-hard scattering has yet to be formulated in a fully consistent manner (see for a conceptual discussion). It would be helpful to find observables with vanishing or at least very small contributions from single scattering. The measurement of such observables would permit the identification of twist-4 effects at larger values of $`Q^2`$ where perturbation theory is more reliable and thus provide a clearer evidence for the occurrence of double scattering. We find these observables in the angular distribution of DY pairs, namely the helicity amplitude $`W_\mathrm{\Delta }`$ and the Lam-Tung relation $`W_{\mathrm{\Delta }\mathrm{\Delta }}=W_\mathrm{L}/2`$ , both introduced in section 2. The Lam-Tung relation is analogous to the Callan-Gross relation in deep-inelastic electron nucleon scattering (DIS). Furthermore it would be helpful to find observables where double-hard and soft-hard contributions can be disentangled. These can also be found in the angular distribution of DY pairs. While SH scattering leads to non-trivial angular patterns, DH scattering does not. Any anomalous $`A`$-dependent angular distribution therefore would be a clear sign of non trivial twist-4 correlations. Although we will focus on $`h+A`$ collisions, our work has been primarily motivated by the fact that higher-twist effects or multiple scattering should be most prominent in nucleus-nucleus ($`A+A`$) collisions at high energy, as they will be routinely studied at RHIC. Two microscopic, QCD-inspired descriptions of nucleus-nucleus collisions at RHIC energies and beyond have been proposed in recent years. In one, the parton cascade picture , the collision is modeled as a sequence of incoherent binary scattering events among partons (“minijets”), which can be cast in a probabilistic formulation as a multiple scattering process. In the other one, the random mean-field model the collision is described as a nonlinear interaction among coherent color fields carried by the colliding nuclei. We hope that the study of the interplay between soft-hard and multiple-hard processes in a simpler framework, such as $`h+A`$ interactions, will help to elucidate the competition between these processes in $`A+A`$ collisions. The determination of multiparton correlation functions in nuclei is also an important step towards the application of QCD to nuclear physics in general. The generic enhancement of higher-twist effects in reactions involving nuclei facilitates their extraction and give the chance for a deeper theoretical understanding. They encode the essential aspects of the difference in the quark-gluon structure between isolated nucleons and nuclei. From the practical point of view a better understanding of such medium effects is necessary to be able to interpret the so called “hard probes” of high-energy heavy-ion collisions in a convincing manner. The unintegrated Drell-Yan cross section is actually the classic example for a typical two-scale process. Both $`Q^2`$ and $`q_{}`$, the mass and transverse momentum of the photon, are detected. The perturbative treatment then gives rise to logs of the type $`\mathrm{log}^2(Q^2/q_{}^2)`$ which can become large if the scales differ substantially. In such a case they have to be resummed. Resumming large logs in the presence of additional soft gluons contained in the twist-4 correlators is a delicate task that could not be solved yet. We therefore follow a less ambitious procedure and require that both scales are of the same order, thus reducing the original two-scale problem to an effective one-scale problem. In addition the LQS formalism in its present form requires $`q_{}^2`$ to be not much smaller than $`Q^2`$ in order to neglect interference terms that are not calculated up to now and which would introduce additional twist-4 matrix elements. Recently Guo, Qiu and Zhang proposed a method to describe the region of low transverse momentum and how to match it to the perturbative calculation at high $`q_{}`$ . Finally we would like to comment on the experimental situation. There exist data for angular coefficients on $`\pi +A`$ scattering, e.g. from the NA10 and E615 experiments which are not yet completely understood in terms of conventional QCD-treatment. These experiments measure pair masses $`Q4\mathrm{}8.5\mathrm{GeV}`$ and transverse momenta $`q_{}<3\mathrm{GeV}`$ which is certainly at the limit of where we could still trust our calculation. We will focus on future RHIC perspectives where we could hopefully reach higher transverse momenta. The paper is organized as follows: In section 2 we give the general definition of the Drell-Yan kinematics, we introduce our observables — helicity amplitudes and angular coefficients — and explain possible choices of the photon rest frame. In section 3 we remind the reader of the known twist-2 results for the helicity amplitudes. Section 4 is devoted to the calculation of the double-scattering contribution. We explain some essential features of the factorization procedure of LQS. In particular we discuss the distinction between double-hard and soft-hard scattering and give an introduction of how to calculate them practically. Finally section 5 contains our numerical results along with a discussion and we close with a summary of our final conclusions. In the appendices we have collected some technical details referred to in the text. ## 2 General Definitions ### 2.1 Kinematics Let a nucleus with momentum $`P_1^\mu `$ per nucleon collide with a hadron with momentum $`P_2^\mu `$. We study lepton pair decay of virtual photons with momentum $`q^\mu `$ and mass $`Q^2=q^2`$. In the hadron center-of-mass (c.m.) frame we take $$P_1^\mu =\frac{1}{2}(\sqrt{S},0,0,\sqrt{S}),P_2^\mu =\frac{1}{2}(\sqrt{S},0,0,\sqrt{S})$$ (2) where $`\sqrt{S}`$ is the total center-of-mass energy. The differential cross section for Drell-Yan pair production is then given by $$d\sigma =\frac{\alpha _{em}^2}{2SQ^4}L_{\mu \nu }W^{\mu \nu }\frac{\mathrm{d}^4q}{(2\pi )^4}\mathrm{d}\mathrm{\Omega }.$$ (3) The solid angle $`\mathrm{\Omega }`$ is parametrized by the polar and azimuthal angles $`\theta `$ and $`\varphi `$ of one decay lepton in the rest frame of the virtual photon. We have the freedom to choose the axes in the photon rest frame conveniently. We will discuss this point after eq. (16). The hadronic tensor follows the standard definition and is given by $$W_{\mu \nu }=\mathrm{d}^4xe^{iqx}P_1P_2\left|j_\mu (x)j_\nu (0)\right|P_1P_2.$$ (4) The leptonic tensor is $$L^{\mu \nu }=2\left(l_1^\mu l_2^\nu +l_2^\mu l_1^\nu \right)g^{\mu \nu }Q^2,$$ (5) where $`l_1`$ and $`l_2`$ are the momenta of the leptons. We rewrite the $`q`$-dependence of the cross section in terms of rapidity $`y`$, transverse momentum $`q_{}`$ and mass $`Q`$ of the virtual photon in the hadron c.m. frame. Introducing the hadronic Mandelstam invariants $`S`$ $`=`$ $`(P_1+P_2)^2`$ $`T`$ $`=`$ $`(P_1q)^2`$ (6) $`U`$ $`=`$ $`(P_2q)^2`$ the rapidity $`y`$ of the photon is $$y=\frac{1}{2}\mathrm{ln}\left(\frac{q^0+q^3}{q^0q^3}\right)=\frac{1}{2}\mathrm{ln}\left(\frac{Q^2U}{Q^2T}\right).$$ (7) Likewise we can express the transverse momentum of the lepton pair in the hadron c.m. system by $$q_{}^2=\frac{(Q^2U)(Q^2T)}{S}Q^2.$$ (8) and we can rewrite the differential cross section as: $$\frac{\mathrm{d}\sigma }{\mathrm{d}Q^2\mathrm{d}q_{}^2\mathrm{d}y\mathrm{d}\mathrm{\Omega }}=\frac{\alpha _{em}^2\pi }{4SQ^4}L_{\mu \nu }W^{\mu \nu }\frac{1}{(2\pi )^4}$$ (9) ### 2.2 Helicity amplitudes and their frame dependence To discuss the angular distribution of the lepton pair it is convenient to introduce helicity amplitudes $`W_{\sigma ,\sigma ^{}}`$ defined as $$W_{\sigma ,\sigma ^{}}=ϵ_\mu ^{(\sigma )}ϵ_\nu ^{(\sigma ^{})}W^{\mu \nu }$$ (10) where $`ϵ_\mu ^{(\sigma )}`$, $`\sigma \{\pm 1,0\}`$ is a set of polarization vectors of the virtual photon with respect to the axes defined by the spatial unit vectors X, Y, Z in its rest frame: $$\epsilon _\mu ^{(\pm )}=\frac{1}{\sqrt{2}}(XiY)_\mu ,\epsilon _\mu ^{(0)}=Z_\mu $$ (11) Following it is advantageous to introduce a complete set of structure functions $$\begin{array}{cc}\hfill W_\mathrm{T}& =W_{1,1}\hfill \\ \hfill W_\mathrm{L}& =W_{0,0}\hfill \\ \hfill W_\mathrm{\Delta }& =\frac{1}{\sqrt{2}}\left(W_{1,0}+W_{0,1}\right)\hfill \\ \hfill W_{\mathrm{\Delta }\mathrm{\Delta }}& =W_{1,1}\hfill \end{array}$$ (12) which are referred to as the transverse, longitudinal, spin flip, and double spin flip helicity amplitudes. Due to parity conservation other combinations are related to these by $`W_{\sigma ,\sigma ^{}}=()^{\sigma +\sigma ^{}}W_{\sigma ,\sigma ^{}}`$. In terms of these helicity structure functions the cross section can be written as $$\begin{array}{c}\frac{\mathrm{d}\sigma }{\mathrm{d}Q^2\mathrm{d}q_{}^2\mathrm{d}y\mathrm{d}\mathrm{\Omega }}=\frac{\alpha _{em}^2}{64\pi ^3SQ^2}\hfill \\ \hfill \left(W_{\mathrm{TL}}\left(1+\mathrm{cos}^2\theta \right)+W_\mathrm{L}\left(\frac{1}{2}\frac{3}{2}\mathrm{cos}^2\theta \right)+W_\mathrm{\Delta }\left(\mathrm{sin}2\theta \mathrm{cos}\varphi \right)+W_{\mathrm{\Delta }\mathrm{\Delta }}\left(\mathrm{sin}^2\theta \mathrm{cos}2\varphi \right)\right).\end{array}$$ (13) The angular dependence can be integrated out which reduces the cross section to $$\frac{\mathrm{d}\sigma }{\mathrm{d}Q^2\mathrm{d}q_{}^2\mathrm{d}y}=\frac{\alpha _{em}^2}{64\pi ^3SQ^2}\left(\frac{16}{3}\pi \right)W_{\mathrm{TL}}$$ (14) with $`W_{\mathrm{TL}}=W_\mathrm{T}+\frac{1}{2}W_\mathrm{L}=\frac{1}{2}W_\mu ^\mu `$. The ratio of the angular dependent and angular integrated cross sections is usually parametrized in two standard forms : $$\begin{array}{cc}\hfill \frac{16\pi }{3}\left(\frac{\frac{\mathrm{d}\sigma }{\mathrm{d}Q^2\mathrm{d}q_{}^2\mathrm{d}y\mathrm{d}\mathrm{\Omega }}}{\frac{\mathrm{d}\sigma }{\mathrm{d}Q^2\mathrm{d}q_{}^2\mathrm{d}y}}\right)& =(1+\mathrm{cos}^2\theta )+A_0\left(\frac{1}{2}\frac{3}{2}\mathrm{cos}^2\theta \right)+A_1\mathrm{sin}2\theta \mathrm{cos}\varphi +\frac{A_2}{2}\mathrm{sin}^2\theta \mathrm{cos}2\varphi =\hfill \\ & =\frac{4}{\lambda +3}\left(1+\lambda \mathrm{cos}^2\theta +\mu \mathrm{sin}2\theta \mathrm{cos}\varphi +\frac{\nu }{2}\mathrm{sin}^2\theta \mathrm{cos}2\varphi \right)\hfill \end{array}$$ (15) in terms of two sets of angular coefficients $`A_0`$ $`={\displaystyle \frac{W_\mathrm{L}}{W_{\mathrm{TL}}}}`$ $`A_1`$ $`={\displaystyle \frac{W_\mathrm{\Delta }}{W_{\mathrm{TL}}}}`$ $`A_2`$ $`={\displaystyle \frac{2W_{\mathrm{\Delta }\mathrm{\Delta }}}{W_{\mathrm{TL}}}},`$ (16) $`\lambda `$ $`={\displaystyle \frac{23A_0}{2+A_0}}`$ $`\mu `$ $`={\displaystyle \frac{2A_1}{2+A_0}}`$ $`\nu `$ $`={\displaystyle \frac{2A_2}{2+A_0}}.`$ (17) At this point it is necessary to comment on the proper choice of coordinate axes in the photon rest system with respect to which the polar and azimuthal angles $`\theta `$ and $`\varphi `$ are defined. Clearly the helicity amplitudes and angular coefficients depend on the chosen frame. Two prominent examples are the Collins-Soper frame (CS) and the Gottfried-Jackson frame (GJ) which are discussed in detail in . When we boost from the hadron c.m. frame to a lepton c.m. frame, the collinearity of the hadron momenta $`𝐏_1`$ and $`𝐏_2`$ is lost and they span a plane which we identify with the $`𝐗`$-$`𝐙`$-plane. We are still free to fix $`𝐙`$ within this plane. In the Collins-Soper frame we choose $`𝐙`$ to bisect the angle between $`𝐏_1`$ and $`𝐏_2`$, see Fig. 2. The angle between $`𝐙`$ and $`𝐏_1`$ and between $`𝐙`$ and $`𝐏_2`$ is called $`\gamma _{\mathrm{CS}}`$ In the Gottfried-Jackson frame $`𝐙`$ lies in the direction of $`𝐏_1`$. We have introduced the the angle $`\gamma _{\mathrm{GJ}}`$ between $`𝐏_1`$ and $`𝐏_2`$ for short notations in that case. Obviously we have $`2\gamma _{\mathrm{CS}}+\gamma _{\mathrm{GJ}}=\pi `$. See Appendix 7.1 for more details. Also note that for $`q_{}=0`$ both frames are equal and $`\varphi `$ is no longer defined. In order to obtain the results in both frames in a straightforward way we use the projectors given by Mirkes $`P_{\mathrm{TL}}^{\mu \nu }`$ $`=g^{\mu \nu }`$ $`P_{\mathrm{L1}}^{\mu \nu }`$ $`={\displaystyle \frac{P_1^\mu P_1^\nu }{E_1^2}}`$ $`P_{\mathrm{L12}}^{\mu \nu }`$ $`={\displaystyle \frac{P_1^\mu P_2^\nu +P_2^\mu P_1^\nu }{E_1E_2}}`$ $`P_{\mathrm{L2}}^{\mu \nu }`$ $`={\displaystyle \frac{P_2^\mu P_2^\nu }{E_2^2}}`$ (18) where $`E_1`$ and $`E_2`$ are the energies of the hadrons in the lepton c.m. frame. In terms of hadronic Mandelstam variables and the photon momentum these are $$E_1=\frac{Q^2T}{2Q}=\frac{e^y}{2Q}\sqrt{S(Q^2+q_{}^2)},E_2=\frac{Q^2U}{2Q}=\frac{e^y}{2Q}\sqrt{S(Q^2+q_{}^2)}.$$ (19) Contracting the hadronic tensor with these projectors gives amplitudes $`T_\alpha =P_\alpha ^{\mu \nu }W_{\mu \nu }`$, $`\alpha \{\mathrm{TL},\mathrm{L1},\mathrm{L2},\mathrm{L12}\}`$ which are just linear combinations of the helicity amplitudes in the different frames. These can be found by linear transformations $$\left(\begin{array}{c}W_{\mathrm{TL}}^{(\beta )}\\ W_\mathrm{L}^{(\beta )}\\ W_\mathrm{\Delta }^{(\beta )}\\ W_{\mathrm{\Delta }\mathrm{\Delta }}^{(\beta )}\end{array}\right)=M_{(\beta )}\left(\begin{array}{c}T_{\mathrm{TL}}\\ T_\mathrm{L}1\\ T_\mathrm{L}2\\ T_\mathrm{L}12\end{array}\right),$$ (20) with $`\beta \{\mathrm{CS},\mathrm{GJ}\}`$. The transformation matrices $`M_{(\beta )}`$ follow from simple geometry and are given in Appendix 7.1 for the CS and GJ frames. ## 3 Leading twist results To make this paper self contained we give the leading order results for the angular coefficients. Due to twist-2 factorization the hadronic tensor is given by a convolution of a perturbatively calculable partonic tensor and two twist-2 distribution functions: $$W_{\mu \nu }=g^2e_q^2\underset{a,b}{}\frac{\mathrm{d}\xi _1}{\xi _1}\frac{\mathrm{d}\xi _2}{\xi _2}H_{\mu \nu }^{a+b}(\xi _1,\xi _2)f_{a/A}(\xi _1)f_{b/h}(\xi _2)2\pi \delta (s+t+uQ^2).$$ (21) The sum runs over partons $`a`$ and $`b`$ in the nucleus $`A`$ and the single hadron $`h`$ respectively and $`\xi _1`$, $`\xi _2`$ are their momentum fractions. We denote the strong coupling by $`g`$ and $`e_q`$ is the fractional electric charge of the quark flavour coupling to the virtual photon. The $`f(\xi )`$ correspond to twist-2 parton distribution functions defined in the usual way as $$\begin{array}{cc}\hfill f_{q/h}(\xi )& =\frac{1}{P^+}\frac{\mathrm{d}y^{}}{2\pi }P^+e^{i\xi P^+y^{}}\frac{1}{2}P\left|\overline{q}(0)\gamma ^+𝒫q(y^{})\right|P\hfill \\ \hfill f_{g/h}(\xi )& =\frac{1}{\xi P^+}\frac{\mathrm{d}y^{}}{2\pi }e^{i\xi P^+y^{}}P\left|F_a^{+\nu }(0)𝒫F_{a\nu }^+(y^{})\right|P\hfill \end{array}$$ (22) for quarks and gluons, respectively in a hadron moving with momentum $`P^+`$ along the light cone . $`q`$ and $`F`$ denote quark fields and gluon field strength tensors and $`𝒫`$ is the path-ordered exponential $$𝒫=P\mathrm{exp}\left\{ig_y^{}^0dx^{}A^+(x^{})\right\}$$ (23) which makes the definition gauge invariant. In the following we will not write this path-ordered exponential explicitly, but it is assumed to be inserted between all factors in a product of operators referring to different space-time points. In the $`A^+=0`$ gauge this term will disappear but we will always use covariant gauge. Note that we also suppressed the dependence of the distribution function on the factorization scale $`\mu ^2`$. The partonic tensor $`H_{\mu \nu }^{a+b}`$ represents the hard part of the process. We have introduced partonic Mandelstam variables by $`s`$ $`=`$ $`(\xi _1P_1+\xi _2P_2)^2=\xi _1\xi _2S`$ $`t`$ $`=`$ $`(\xi _1P_1q)^2=\xi _1(TQ^2)+Q^2`$ (24) $`u`$ $`=`$ $`(\xi _2P_2q)^2=\xi _2(UQ^2)+Q^2.`$ The $`\delta `$-function in eq. (21) stems from the on-shell condition for the unobserved emitted parton with momentum $`l`$. It can be used for trivial integration of $`\xi _2`$ via $$\delta (l^2)=\delta (s+t+uQ^2)=\frac{1}{\xi _1S+UQ^2}\delta \left(\xi _2+\frac{Q^2+\xi _1(TQ^2)}{\xi _1S+UQ^2}\right).$$ (25) We can decompose the partonic tensor in a way completely analogous to the decomposition of the hadronic tensor $`W_{\mu \nu }`$ in terms of helicity amplitudes. Therefore we define the four projections $`H_{\mathrm{TL}}^{a+b},H_\mathrm{L}^{a+b},H_\mathrm{\Delta }^{a+b},H_{\mathrm{\Delta }\mathrm{\Delta }}^{a+b}`$ which are obtained in the same way as in eqs. (10,12). This allows us to write hadronic tensor and helicity amplitudes as $$W_\alpha =8\pi ^2\alpha _se_q^2\underset{a,b}{}\underset{B}{\overset{1}{}}\frac{\mathrm{d}\xi _1}{\xi _1}\frac{1}{\xi _1(Q^2T)Q^2}f_{a/A}(\xi _1)f_{b/h}(\overline{\xi }_2)H_\alpha ^{a+b}(\xi _1,\overline{\xi }_2).$$ (26) with $`B=\frac{U}{S+TQ^2}`$, $`\overline{\xi }_2=\frac{Q^2+\xi _1(TQ^2)}{\xi _1S+UQ^2}`$, $`\alpha _s=g^2/4\pi `$ and $`\alpha \{\mu \nu ,TL,L,\mathrm{\Delta },\mathrm{\Delta }\mathrm{\Delta }\}`$. There are three kinds of partonic processes contributing to the leading-twist process at large transverse momentum: quark-antiquark annihilation ($`\overline{q}+q`$) and Compton processes with a gluon from the single hadron ($`q+g`$) or the nucleus ($`g+q`$). They are shown in Fig. 3. The cross section for the leading twist DY pair production at large transverse momentum is well-known . For completeness we give the results in the Collins-Soper frame in Appendix 7.2. The relation $`H_{\mathrm{\Delta }\mathrm{\Delta }}=\frac{1}{2}H_\mathrm{L}`$ is precisely the so-called Lam-Tung relation . It can be expressed in terms of angular coefficients as $`A_0=A_2`$ or $`2\nu (1\lambda )=0`$. This relation holds for all partonic subprocesses and is not specific for the Collins-Soper frame but holds in any photon rest frame as long as the $`𝐙`$-axis lies in the reaction plane. Corrections arise at next-to-leading order (NLO) . Note that $`H_\mathrm{\Delta }^{a+b}`$ picks up an additional sign compared with $`H_\mathrm{\Delta }^{b+a}`$ in contrast to the other helicity amplitudes. This is due to the inversion of the orientation of the angle $`\gamma _{\mathrm{CS}}`$ if we interchange the roles of $`P_1`$ and $`P_2`$. For symmetric collisions, e. g. $`p+p`$, this leads to the leading-twist result $`W_\mathrm{\Delta }=0`$. ## 4 Twist-4 Factorisation ### 4.1 Soft-hard and double-hard scattering For the calculation of the double scattering contributions we use the approach of Luo, Qiu and Sterman which is presented e. g. in . We consider it worthwile to repeat some aspects here. In Fig. 4 we illustrate the forward diagrams that need to be calculated for double scattering. The box denotes all possible partonic subprocesses. We have to introduce a cut to obtain the corresponding hadronic tensors. This is not only possible in the middle but we also have to include interference graphs which appear if the forward diagram is cut in an asymmetrical way. In Fig. 5 we give concrete examples for graphs which involve a quark-gluon matrix element in the nucleus and an antiquark coming from the single hadron. The first diagram is cut in the middle (M), the others are cut left (L) and right (R). The starting point for the calculation of the hadronic tensor is the forward scattering amplitude in fourth order of the strong coupling constant $`g`$: $$\begin{array}{cc}\hfill T_{\mu \nu }& =\frac{g^4e_q^2}{4!}\mathrm{d}^4x\mathrm{d}^4x_1\mathrm{d}^4x_2\mathrm{d}^4x_3\mathrm{d}x_4e^{iqx}P_1P_2|T[J_\rho (x_3)J_\alpha (x_1)j_\mu (0)j_\nu (x)\hfill \\ & J_\beta (x_2)J_\sigma (x_4)A^\rho (x_3)A^\alpha (x_1)A^\beta (x_2)A^\sigma (x_4)+\mathrm{}]|P_1P_2\hfill \end{array}$$ (27) where $`j`$ ($`J`$) are electromagnetic (color) quark currents and $`A`$ denotes gluon fields. $`T`$ is the time-ordering operator. Dots indicate further contributions involving more gluon fields with gluon self interaction and less quark currents such that the power of $`g`$ remains four. By application of Wick’s theorem we obtain all diagrams which potentially contribute to the forward process. We classify diagrams by their partonic subprocess $`ab+c`$ where $`a`$ and $`b`$ are partons from the nucleus and $`c`$ comes from the single hadron. Our notation for the parton momenta is shown in Fig. 4. We concentrate first on $`qg+\overline{q}`$ which can be realized e.g. by the diagrams of Fig. 5. As a first step towards factorization we perform Fierz transformations in color and Dirac space for the fermion lines entering and leaving the blobs: $$q_\alpha ^i(0)\overline{q}_\beta ^j(z)=\frac{1}{4}\left(\gamma _\mu \right)_{\alpha \beta }\overline{q}^j^{}(z)\gamma ^\mu q^i^{}(0)\left(\frac{1}{N_c}\delta _{j^{}i^{}}\delta _{ij}+2t_{j^{}i^{}}^At_{ij}^A\right)+\mathrm{}.$$ (28) Here the $`t^A`$ are the $`SU(N_c)`$ Gell-Mann matrices, and $`N_c`$ is the number of colors. Latin characters are color indices, greek one are Dirac indices. We keep only Dirac structures which give leading contributions to the unpolarized Drell-Yan process. After Fourier transformations of the fields we end up with $$\begin{array}{c}T_{\mu \nu }^{qg+\overline{q}}=\frac{\mathrm{d}^4l}{(2\pi )^4}\frac{\mathrm{d}^4r_1}{(2\pi )^4}\frac{\mathrm{d}^4r_2}{(2\pi )^4}\frac{\mathrm{d}^4K}{(2\pi )^4}\frac{\mathrm{d}^4K^{}}{(2\pi )^4}\mathrm{d}^4y\mathrm{d}^4z_1\mathrm{d}^4z_3\mathrm{d}^4z_4\hfill \\ \hfill (2\pi )^4\delta ^{(4)}(r_1+r_2+Klq)e^{ir_1z_1}e^{ir_2y}e^{iKz_3}e^{iK^{}z_4}\frac{1}{2}P_1\left|\overline{q}(0)\gamma ^\eta q(z_1)A^\rho (z_3)A^\sigma (z_4)\right|P_1\\ \hfill \left(\frac{1}{2}\right)P_2\left|\overline{q}(0)\gamma ^\tau q(y)\right|P_2e_q^2g^4\frac{1}{N_c^2}\frac{\delta ^{RS}}{N_c^21}\underset{d}{}S(d)_{\nu \sigma \tau \rho \mu \eta }^{RS}.\end{array}$$ (29) We used the fact that forward matrix elements only depend on the relative coordinates. So, after shifting positions in each matrix element separately, we introduced a new set of space-time coordinates ($`y`$,$`z_1`$,$`z_3`$,$`z_4`$). $`l`$ is the momentum of the unobserved radiated parton. The index $`d`$ runs over all diagrams of the class $`qg+\overline{q}`$ and $`S(d)_{\rho \mu ϵ\nu \sigma \tau }^{RS}`$ contains all perturbative propagators and vertices of the hard part of diagram $`d`$, with $`\gamma _\eta `$ ($`\gamma _\tau `$) inserted for the upper (lower) blob, and traces over color and Dirac indices. For the diagram of Fig. 5(a) we have e. g. $$\begin{array}{c}\hfill S_{\nu \sigma \tau \rho \mu \eta }^{RS}=\frac{1}{4}\mathrm{Tr}[(i\gamma _\nu )\frac{i(\mathit{}_1+\mathit{}\mathit{}^{}\mathit{})}{(r_1+KK^{}q)^2+iϵ}(i\gamma _\beta t^B)\frac{i(\mathit{}_2\mathit{}^{})}{(r_2+K^{})^2+iϵ}(i\gamma _\sigma t^S)\\ \hfill \gamma _\tau (i\gamma _\rho t^R)\frac{i(\mathit{}_2\mathit{})}{(r_2+K)^2+iϵ}(i\gamma _\alpha t^A)\frac{i(\mathit{}_1\mathit{})}{(r_1q)^2+iϵ}(i\gamma _\mu )\gamma _\eta ]\frac{i(g^{\alpha \beta })\delta ^{AB}}{l^2+iϵ}.\end{array}$$ (30) $`R`$ and $`S`$ are color indices of gluon fields. Note that we have coupled quark and gluon field operators in the nuclear matrix elements separately to color singlets. We have omitted a term proportional to $`(d^{RES}+if^{RES})P_1\left|\overline{q}\gamma ^\eta t^EqA^{R\rho }A^{S\sigma }\right|P_1`$ which would induce color forces between quark and gluon in the nucleus. This would not allow to factorize quarks and gluons in color space to test the size of the nucleus. Such contributions are not expected to show nuclear enhancement, i.e. scaling by $`A^{4/3}`$. Next we perform the light cone expansion, i.e. we use that the partons move almost collinear to the hadron in the infinite momentum frame. We introduce the parton momentum fractions by setting $`r_1=\xi _1P_1`$, $`r_2=\xi _2P_2`$, $`K=xP_1+K_{}`$ and $`K^{}=x^{}P_1+K_{}^{}`$. $`r_1^{}=r_1+KK^{}`$ is fixed by momentum conservation in the upper forward matrix element. Note that we have to allow one parton (called $`b`$ here) from the nucleus to be soft with essentially zero longitudinal momentum. Therefore we must take into account transverse momenta $`K_{}`$ and $`K_{}^{}`$ to get a contribution in higher orders of the collinear expansion as we will see later. Now one can integrate the trivial degrees of freedom in eq. (29). To complete the factorization we perform a Sudakov decomposition of the $`\gamma `$-matrices in the forward matrix elements on the light cone and retain only the leading terms $`\gamma ^\eta =\frac{P_1^\eta }{P_1^+}\gamma ^+`$, $`\gamma ^\tau =\frac{P_2^\tau }{P_2^{}}\gamma ^{}`$. Then we can introduce twist-2 distribution functions $`f_{c/h}(\xi _2)`$, given in eq. (22), for the lower blob. Finally we cut the forward diagram. This sets the radiated unobserved parton with momentum $`l`$ on the mass shell and in the hard part all expressions right to the cut receive a complex conjugation. Hence after these first steps we have factorized the hadronic tensor in a straightforward way. We get $$\begin{array}{c}\hfill W_{\mu \nu }^{ab+c}=(4\pi \alpha _s)^2e_q^2\frac{\mathrm{d}\xi _1}{\xi _1}\frac{\mathrm{d}\xi _2}{\xi _2}dxP_1^+\mathrm{d}^2K_{}dx^{}P_1^+\mathrm{d}^2K_{}^{}f_{c/h}(\xi _2)(2\pi )\delta (l^2)\\ \hfill \overline{T}_{ab}(\xi _1,x,x^{},K_{},K_{}^{})\overline{H}_{\mu \nu }^{ab+c}(\xi _1,\xi _2,x,x^{},K_{},K_{}^{}).\end{array}$$ (31) We have derived eq. 31 for our examples of Fig. 5, but this equation holds for all processes shown in Fig.4. We have separated the soft part involving the twist-2 distribution function $`f_{c/h}`$ and the nuclear matrix element $`\overline{T}_{ab}`$ from the hard part $`\overline{H}_{\mu \nu }^{ab+c}`$. For our example $`qg+\overline{q}`$ we are left with a quark-gluon matrix element in the nucleus $$\begin{array}{c}\hfill \overline{T}_{qg}(\xi _1,x,x^{},K_{},K_{}^{})=\frac{\mathrm{d}z_1^{}}{2\pi }\frac{\mathrm{d}z_3^{}\mathrm{d}^2z_3^{}}{(2\pi )^3}\frac{\mathrm{d}z_4^{}\mathrm{d}^2z_4^{}}{(2\pi )^3}e^{ir_1^+z_1^{}}e^{iK^+z_3^{}}e^{iK_{}z_3^{}}e^{iK_{}^{}{}_{}{}^{+}z_4^{}}e^{iK_{}^{}z_4^{}}\\ \hfill \frac{1}{2}P_1\left|\overline{q}(0)\gamma ^+q(z_1^{})A^\rho (z_3^{},z_3^{})A^\sigma (z_4^{},z_4^{})\right|P_1.\end{array}$$ (32) For other classes of processes we also need the quark-quark and gluon-gluon correlators $`\overline{T}_{gg}`$ and $`\overline{T}_{qq}`$ which are given by eq. (32) with the second line replaced by $$\begin{array}{c}\hfill \frac{1}{\xi _1P_1^+}P_1\left|F^{\omega +}(0)F_\omega ^+(z_1^{})A^\rho (z_1^{},z_1^{})A^\sigma (z_4^{},z_4^{})\right|P_1\text{for }\overline{T}_{gg}\\ \hfill \frac{1}{2}P_1\left|\overline{q}(0)\gamma ^+q(z_1^{})\overline{q}(z_3^{},z_3^{})\gamma ^\kappa q(z_4^{},z_4^{})\right|P_1\text{for }\overline{T}_{qq}.\end{array}$$ (33) Again here the field operators of each single parton form color singlets separately. We have arranged the first two field operators referring to the nuclear parton $`a`$ in the way they appear in the corresponding twist-2 structure function while the remaining two field operators of parton $`b`$ are still in their original form. The hard part for the $`qg+\overline{q}`$-process is $$\overline{H}_{\mu \nu }^{qg+\overline{q}}(\xi _1,\xi _2,x,x^{},K_{},K_{}^{})=\frac{1}{N_c^2}\frac{1}{N_c^21}(g^{\alpha \beta })r_1^\eta r_2^\tau \frac{1}{4}\underset{d}{}S(d)_{\nu \sigma \tau \rho \mu \eta }^{RR}|_{\mathrm{cut}}.$$ (34) Hard parts for other processes are obtained analogously. The delta function in eq. (31) can be rewritten for the cases in which the diagrams are cut in the middle (M), right (R) or left (L) as $`(\mathrm{M}):\delta (l^2)`$ $`=`$ $`(\xi _2S+TQ^2)^1\delta (\xi _1+x\stackrel{~}{\xi }_b)`$ $`(\mathrm{R}):\delta (l^2)`$ $`=`$ $`(\xi _2S+TQ^2)^1\delta (\xi _1+xx^{}\xi _c)`$ (35) $`(\mathrm{L}):\delta (l^2)`$ $`=`$ $`(\xi _2S+TQ^2)^1\delta (\xi _1\xi _c)`$ where $$\stackrel{~}{\xi }_b=\frac{\xi _2(UQ^2)+Q^22qK_{}+K_{}^2}{\xi _2S+TQ^2}\text{ and }\stackrel{~}{\xi }_c=\stackrel{~}{\xi }_b|_{K_{}=0}.$$ (36) As pointed out in various papers there are in principle two contributions to double scattering: There can be two hard scattering reactions of the parton $`c`$ from the single hadron or the outgoing unobserved parton on one side and partons $`a`$ and $`b`$ from the nucleus on the other side. This is called double-hard scattering. Furthermore there can occur processes where the parton $`c`$ from the single hadron or the outgoing parton picks up one soft nuclear parton $`b`$ in addition to one hard scattering off the nuclear parton $`a`$. This is called soft-hard scattering. These are not only qualitative notions but they can be rigorously derived from the equations given above. One can understand them in terms of poles in the hard part of the scattering tensor. In eq. (31) we have four momentum integrations $`\mathrm{d}\xi _1\mathrm{d}\xi _2\mathrm{d}x\mathrm{d}x^{}`$ and four propagators. We will perform the $`\xi _1`$-integration by making use of the delta function $`\delta (l^2)`$. The propagators provide poles with respect to $`x`$ and $`x^{}`$. In general we expect two poles in $`x`$ and two poles in $`x^{}`$. Application of the residue theorem will fix the momenta and it is easy to check that two poles give large momenta $`x=x_{\mathrm{hard}}=x^{}`$, respectively and the other ones give small momenta $`x=x_{\mathrm{soft}}=x^{}`$ with $`x_{\mathrm{hard}}x_{\mathrm{soft}}1`$ as long as $`q_{}Q`$. To be more precise the pole structure we expect for our example in Fig. 5(a) is $$W_{\mu \nu }dx\frac{L(x)}{(xx_{\mathrm{soft}}+iϵ)(xx_{\mathrm{hard}}+iϵ)}dx^{}\frac{R(x^{})}{(x^{}x_{\mathrm{soft}}iϵ)(x^{}x_{\mathrm{hard}}iϵ)},$$ (37) where $`L`$ and $`R`$ stand for the numerators of the amplitudes left and right of the cut depending on $`x`$ and $`x^{}`$. The residue theorem implies $$W_{\mu \nu }\left[\frac{L(x=x_{\mathrm{soft}})}{x_{\mathrm{soft}}x_{\mathrm{hard}}}\frac{L(x=x_{\mathrm{hard}})}{x_{\mathrm{soft}}x_{\mathrm{hard}}}\right]\left[\frac{R(x^{}=x_{\mathrm{soft}})}{x_{\mathrm{soft}}x_{\mathrm{hard}}}\frac{R(x^{}=x_{\mathrm{hard}})}{x_{\mathrm{soft}}x_{\mathrm{hard}}}\right]$$ (38) Soft-hard and double-hard scattering correspond to the quadratic terms proportional to $`L(x=x_{\mathrm{soft}})R(x^{}=x_{\mathrm{soft}})`$ and $`L(x=x_{\mathrm{hard}})R(x^{}=x_{\mathrm{hard}})`$ respectively. We follow the arguments of reference and neglect the remaining mixed terms of eq. (38) which are interference terms between soft and hard rescattering. This is permitted if $`q_{}`$ is not too small. In the case $`q_{}^2Q^2`$ however we obtain $`x_{\mathrm{soft}}x_{\mathrm{hard}}`$. The interference will then be important and eventually spoil nuclear enhancement due to the negative signs in eq. (38) . Since the integrated Drell-Yan cross section does not exhibit nuclear enhancement, we even expect a suppression effect at low values of transverse momenta . Most diagrams we naively get from Wick’s theorem for double scattering show a different behaviour. For them the propagators do not provide four but only three or two poles — but at least one in each variable $`x`$ and $`x^{}`$. This implies that only soft-hard, only double-hard or neither of these contributions can occur. For example it is easy to understand from this analysis that only symmetrically (M) cut diagrams can contribute to double-hard scattering. Consider the diagram of Fig. 5(b). The propagator on the right-hand side has denominator $`[(r_1^{}q)^2iϵ]`$, from which $`x`$ and $`x^{}`$ dependences cancel out. On the left-hand side, the propagator at the bottom of the diagram contributes a denominator $`(r_2+K)^2+iϵ=\xi _2S(x+\frac{K_{}^2}{\xi _2S}+iϵ)`$. This is a soft pole which fixes $`x`$ to $`x_{\mathrm{soft}}=\frac{K_{}^2}{\xi _2S}1`$. The remaining two propagators on the left-hand side fix $`x^{}`$ to either a soft or a hard value. Therefore we have two contributions: (i) $`x=x^{}`$ is small. This means that parton $`b`$ is soft — in Fig. 5(b) this is the gluon from the nucleus. Together with the production of the virtual photon which is always a hard process, this constitutes a soft-hard double scattering process. (ii) $`x`$ is soft and $`x^{}`$ is hard. This is what we called a soft-hard interference because parton $`b`$ has different momenta on the left and the right side of the diagram, i.e. momentum is transferred between partons $`a`$ and $`b`$. We omit this contribution as explained above. We conclude that there is no double-hard scattering contribution from the diagram in Fig. 5(b) since $`x`$ cannot be hard. On the other hand the diagram in Fig. 5(a) is an example in which all four poles occur, indicated by small circles. We now discuss another technical point which has some relevance for nuclear enhancement. Consider again eq. (37) and the example of Fig. 6(a) and note that the numerators $`L`$ and $`R`$ contain exponential functions $`e^{ixP_1^+z_3^{}}e^{i(\stackrel{~}{\xi }_bx)P_1^+z_1^{}}`$ and $`e^{ix^{}P_1^+z_4^{}}`$. To apply the residue theorem we have to close the integration contours. For $`z_3^{}z_1^{}>0`$ we have to close the $`x`$-integration in the upper complex $`x`$-plane with no poles encircled and therefore vanishing integral. On the other hand for $`z_3^{}z_1^{}<0`$ we must close in the lower complex $`x`$-plane with soft and hard pole encircled. Therefore these poles always introduce additional theta functions in coordinate space, in our example $`\mathrm{\Theta }(z_1^{}z_3^{})`$ from the $`x`$ integration and $`\mathrm{\Theta }(z_4^{})`$ from the $`x^{}`$ integral. The presence of these $`\mathrm{\Theta }`$ functions in the matrix elements is an essential ingredient of the proof of the following rule about the cancellation of final state interactions for soft-hard scattering: If at least one soft parton from the twist-4 matrix element is directly coupled to the outgoing unobserved parton line these graphs do not contribute to soft-hard scattering. These graphs exhibit more than one possibility to set the cut and the nuclear enhancement cancels between the contributions with different cuts due to the particular arrangement of $`\mathrm{\Theta }`$-functions in that case . Hence, after a careful discussion of the pole structure of each diagram the number of diagrams we are finally forced to evaluate can be reduced. Figs. 6 and 7 show the diagrams that contribute to double-hard and soft-hard scattering, respectively. We marked the poles used to fix momenta. Also we omitted soft-hard contributions with soft quarks since they should be suppressed compared to their counterparts with soft gluons. ### 4.2 Calculation of the double-hard processes First we investigate the case of double-hard scattering. Since all parton momenta are hard we can work in the collinear limit and set $`K_{}=0=K_{}^{}`$ immediately. We perform the integrations over $`x`$ and $`x^{}`$ only taking into account contributions from hard poles. These poles are marked with circles in Fig. 6. They fix momenta to the values $`x=x_h=x^{}`$ and $`\xi _1=x_a`$ with $`x_a`$ $`={\displaystyle \frac{Q^2}{Q^2T}}`$ (39) $`x_h`$ $`=\stackrel{~}{\xi }_ax_a={\displaystyle \frac{\xi _2(UQ^2)+Q^2}{\xi _2S+TQ^2}}{\displaystyle \frac{Q^2}{Q^2T}}`$ (40) Note that $`x_a`$ is determined by kinematical variables only whereas $`x_h`$ depends on the integration variable $`\xi _2`$. Two additional $`\mathrm{\Theta }`$-functions in the light cone space-time coordinates show up as explained above. The field operators of the second parton in the nuclear matrix element are selected in such a way that we get the usual twist-2 light cone operator $`\overline{q}\gamma ^+q`$ for quarks and field strength tensors $`F^{\omega +}F_\omega ^+`$ for gluons. In the latter case we have to turn two factors of $`P_1^+`$ into light cone derivatives $`(ix_h)^1_{z_3^{}}`$, $`(ix_h)^1_{z_4^{}}`$ and perform partial integrations to let them act on the gluon fields in the matrix element. Hard gluons carry only physical polarizations, this allows us to set $`^+A^\rho (z_3^{})^+A^\sigma (z_4^{})=F^{\omega +}(z_3^{})F_\omega ^+(z_4^{})\frac{1}{2}(g^{\rho \sigma }+\overline{n}^\rho n^\sigma +n^\rho \overline{n}^\sigma )`$ with the usual light cone basis vectors $`\overline{n}^\rho =P_1^\rho /P_1^+`$, $`\overline{n}^\rho =P_2^\rho /P_2^{}`$. We are now able to give the final result for double-hard scattering. For a partonic subprocess of the type $`ab+c`$ we have $$W_\alpha ^{ab+c}=\frac{(2\pi )^2(4\pi \alpha _s)^2e_q^2}{Q^2T}_B^1\frac{\mathrm{d}\xi _2}{\xi _2}\frac{1}{\xi _2S+TQ^2}f_{c/H}(\xi _2)T_{ab}^{\mathrm{DH}}(x_a,x_h)\frac{1}{x_h}H_\alpha ^{\mathrm{DH}/ab+c}(x_a,x_h,\xi _2)$$ (41) with $`\alpha \{\mu \nu ,TL,L,\mathrm{\Delta },\mathrm{\Delta }\mathrm{\Delta }\}`$. Note that parton $`a`$ is always a quark or antiquark for double-hard scattering and $`e_q`$ denotes its electric charge. As in the helicity amplitudes which arise from contractions of the hadronic tensor $`W_{\mu \nu }`$ defined in eq. 10, here we use the same contractions for the hard partonic part $`H_{\mu \nu }`$, defining amplitudes $`H_{\mathrm{TL}}`$, $`H_\mathrm{L}`$, $`H_\mathrm{\Delta }`$ and $`H_{\mathrm{\Delta }\mathrm{\Delta }}`$ depending on the frame. The hard parts for the relevant diagrams can be written as $`H_{\mu \nu }^{\mathrm{DH},qg+\overline{q}}`$ $`=𝒞{\displaystyle \frac{1}{Q^2}}{\displaystyle \frac{1}{4}}\mathrm{Tr}\left[(\mathit{}_1\mathit{})R_{\sigma \beta }\mathit{}_2L_{\rho \alpha }(\mathit{}_1\mathit{})\gamma _\mu \mathit{}_1\gamma _\nu \right]_{\begin{array}{c}\xi _1=x_a\\ x=x^{}=x_h\end{array}}{\displaystyle \frac{1}{2}}g^{\alpha \beta }g^{\rho \sigma }`$ $`H_{\mu \nu }^{\mathrm{DH},qg+g}`$ $`=𝒞{\displaystyle \frac{1}{Q^2}}{\displaystyle \frac{1}{4}}\mathrm{Tr}\left[(\mathit{}_1\mathit{})R_{\sigma \beta }\mathit{}L_{\rho \alpha }(\mathit{}_1\mathit{})\gamma _\mu \mathit{}_1\gamma _\nu \right]_{\begin{array}{c}\xi _1=x_a\\ x=x^{}=x_h\end{array}}{\displaystyle \frac{1}{2}}g^{\alpha \beta }g^{\rho \sigma }`$ (42) $`H_{\mu \nu }^{\mathrm{DH},q\overline{q}+g}`$ $`=𝒞{\displaystyle \frac{1}{Q^2}}{\displaystyle \frac{1}{4}}\mathrm{Tr}\left[(\mathit{}_1\mathit{})R_{\sigma \beta }\mathit{}L_{\rho \alpha }(\mathit{}_1\mathit{})\gamma _\mu \mathit{}_1\gamma _\nu \right]_{\begin{array}{c}\xi _1=x_a\\ x=x^{}=x_h\end{array}}{\displaystyle \frac{1}{2}}g^{\alpha \beta }g^{\rho \sigma }.`$ $`L_{\rho \alpha }`$ and $`R_{\sigma \beta }`$ indicate the sum over all possible tree level diagrams for the blobs at the left and the right side of the corresponding diagram in Fig. 6. Each diagram provides an individual color factor which is indicated symbolically by the factor $`𝒞`$. For partonic processes with three participating fermions we have to distinguish between different possible combinations of quark and antiquarks and different flavours. In principle there are contributions written in our short notation as $`q\overline{q}+q`$, $`q\overline{q}+\overline{q}`$ and $`qq+\overline{q}`$. Each of these contributes four diagrams contained in Fig. 6(d). In addition, there are processes of the type $`q\overline{q}+q^{}`$, $`qq^{}+\overline{q}`$ and $`qq^{}+\overline{q}^{}`$, which are each represented by a single diagram. In our notation $`q`$ is a quark or antiquark, $`\overline{q}`$ the corresponding antiparticle and $`q^{}`$ denotes a quark or antiquark of a different flavour than $`q`$. As an example we give the hard part for the $`q\overline{q}+q`$ subprocess: $$\begin{array}{c}H_{\mu \nu }^{\mathrm{DH},q\overline{q}+q}=𝒞\frac{1}{Q^2}\frac{1}{8}\{\mathrm{Tr}\left[(\mathit{}_1\mathit{})\gamma _\sigma \mathit{}\gamma _\beta (\mathit{}_1\mathit{})\gamma _\mu \mathit{}_1\gamma _\nu \right]\mathrm{Tr}\left[\mathit{}_2\gamma _\rho \mathit{}\gamma _\alpha \right](r_1+Kq)^4\hfill \\ \hfill \mathrm{Tr}\left[(\mathit{}_1\mathit{})\gamma _\sigma \mathit{}\gamma _\beta \mathit{}_2\gamma _\rho \mathit{}\gamma _\alpha (\mathit{}_1\mathit{})\gamma _\mu \mathit{}_1\gamma _\nu \right](r_1+Kq)^2(r_2+K)^2+\\ \hfill +(Kl)\}_{\begin{array}{c}\xi _1=x_a\\ x=x^{}=x_h\end{array}}g^{\alpha \beta }g^{\rho \sigma }.\end{array}$$ (43) Note the relative sign that is due to the different number of fermion loops in the corresponding diagrams. The $`T_{ab}^{\mathrm{DH}}(x_a,x_h)`$ are the universal nuclear matrix elements for double-hard scattering introduced in . They depend on the momentum fractions $`x_a`$, $`x_h`$ of both hard partons from the nucleus and are given by $$\begin{array}{cc}\hfill T_{qg}^{\mathrm{DH}}(x_a,x_h)& =\frac{1}{x_h}\frac{1}{2}dz_4^{}\frac{\mathrm{d}z_3^{}}{2\pi }\frac{\mathrm{d}z_1^{}}{2\pi }\mathrm{\Theta }(z_1^{}z_3^{})\mathrm{\Theta }(z_4^{})e^{ix_aP_1^+z_1^{}}e^{ix_hP_1^+(z_3^{}z_4^{})}\hfill \\ & P_1\left|F^{\omega +}(z_4^{})F_\omega ^+(z_3^{})\overline{q}(0)\gamma ^+q(z_1^{})\right|P_1\hfill \\ \hfill T_{q\overline{q}}^{\mathrm{DH}}(x_a,x_h)& =\frac{1}{4}dz_4^{}P_1^+\frac{\mathrm{d}z_3^{}}{2\pi }\frac{\mathrm{d}z_1^{}}{2\pi }\mathrm{\Theta }(z_1^{}z_3^{})\mathrm{\Theta }(z_4^{})e^{ix_aP_1^+z_1^{}}e^{ix_hP_1^+(z_3^{}z_4^{})}\hfill \\ & P_1\left|\overline{q}(0)\gamma ^+q(z_1^{})\overline{q}(z_3^{})\gamma ^+q(z_4^{})\right|P_1\hfill \\ \hfill T_{qq}^{\mathrm{DH}}(x_a,x_h)& =\frac{1}{4}dz_4^{}P_1^+\frac{\mathrm{d}z_3^{}}{2\pi }\frac{\mathrm{d}z_1^{}}{2\pi }\mathrm{\Theta }(z_1^{}z_3^{})\mathrm{\Theta }(z_4^{})e^{ix_aP_1^+z_1^{}}e^{ix_hP_1^+(z_3^{}z_4^{})}\hfill \\ & P_1\left|\overline{q}(0)\gamma ^+q(z_1^{})\overline{q}(z_4^{})\gamma ^+q(z_3^{})\right|P_1.\hfill \end{array}$$ (44) We are now ready to explain the origin of nuclear enhancement in these matrix elements. As mentioned above the fields of both partons are coupled separately to colour singlets. This allows to probe distances up to the diameter of a large nucleus. Each spatial integration of the matrix element arises from a Fourier transformation and we expect it to be accompanied by a rapidly oscillating exponential function (since $`P_1^+\mathrm{}`$). At first glance, this seems to rule out coherence over the whole nucleus. However only $`z_1^{}`$ and $`z_3^{}z_4^{}`$ in eq. (44) are restricted by phase factors while $`z_3^{}+z_4^{}`$ is not. Hence, we can set $`\mathrm{d}(z_3^{}+z_4^{})`$ to be proportional to the nuclear radius $`R_Ar_0A^{1/3}`$ and obtain an additional factor of $`A^{1/3}`$. Here $`r_01.1\mathrm{fm}`$ is the radius of a nucleon. To find a reasonable model for the double-hard matrix elements imagine the insertion of a complete set of states $`_P|PP|/(2P^+V)=1`$ where $`V`$ is the volume of the nucleus. Neglecting for a moment the $`\mathrm{\Theta }`$-function and the integration over $`z_3^{}+z_4^{}`$ we can compare these definitions with the common definitions of twist-2 structure functions in eq. (22). Assuming that the sum over states is saturated by the lowest lying nucleon state double hard matrix elements reduce to a product of two twist-2 structure functions times a dimensional factor which scales like $`A^{4/3}`$ $$T_{ab}^{\mathrm{DH}}(x_a,x_h)=CA^{4/3}f_{a/A}(x_a)f_{b/A}(x_h).$$ (45) $`f_{a/A}`$ is the nuclear parton distribution per nucleon in a nucleus with mass number $`A`$ and proton number $`Z`$, $`N=AZ`$, given by $$f_{a/A}(x)=\frac{Z}{A}f_{a/p}(x)+\frac{N}{A}f_{a/n}(x).$$ (46) In a primitive model one can simply set the remaining free integration with respect to $`z_4^{}+z_3^{}`$ equal to $`R_A`$ and obtains $`C=3/(8\pi r_0^2)0.005\mathrm{GeV}^2`$. Although the factorization of the double-hard matrix element into a product of two twist-2 distribution functions and the scaling by $`A^{4/3}`$ seems to provide a reasonable model, the choice of the normalization factor $`C`$ is a hot matter of discussion, see section 5. The $`\mathrm{\Theta }`$-functions provide an ordering of both single scattering events on the light cone. They simply express the fact that scattering off parton $`b`$ takes place before the scattering off parton $`a`$. This corresponds to $`z_4^{}<0`$ on the left side and to $`z_3^{}<z_1^{}`$ on the right side of each diagram. Now we give the final results for the hard parts. For the evaluation it is important also to include ghosts, whenever one can close gluon lines through the non-perturbative blobs. For $`H_{\mathrm{TL}}^{\mathrm{DH}}=\frac{1}{2}H_\mu ^{\mathrm{DH}\mu }`$, which enters the angular integrated cross section, we reproduce the results of . We have listed them in Appendix 7.3 for all subprocesses. The helicity amplitude $`W_{\mathrm{TL}}`$ is independent of the frame. Due to the special form of all diagrams contributing to double-hard scattering we easily see that the contractions $`P_{\mathrm{L1}}^{\mu \nu }H_{\mu \nu }^{\mathrm{DH}}`$ and $`P_{\mathrm{L12}}^{\mu \nu }H_{\mu \nu }^{\mathrm{DH}}`$ in eq. (4.2) always vanish, and for the last contraction we get $$P_{\mathrm{L2}}^{\mu \nu }H_{\mu \nu }^{\mathrm{DH}}=\frac{Q^2S\left[(Q^2T)(Q^2U)Q^2S\right]}{(Q^2T)^2(Q^2U)^2}H_{\mathrm{TL}}^{\mathrm{DH}}=\frac{Q^2q_{}^2}{(Q^2+q_{}^2)^2}H_{\mathrm{TL}}^{\mathrm{DH}}$$ (47) for all processes. For the Collins-Soper frame this leads to the simple relations $`W_\mathrm{L}^{\mathrm{DH}}=`$ $`\mathrm{sin}^2\gamma _{\mathrm{CS}}W_{\mathrm{TL}}^{\mathrm{DH}}`$ $`={\displaystyle \frac{q_{}^2}{Q^2+q_{}^2}}W_{\mathrm{TL}}^{\mathrm{DH}}`$ (48) $`W_\mathrm{\Delta }^{\mathrm{DH}}=`$ $`\mathrm{sin}\gamma _{\mathrm{CS}}\mathrm{cos}\gamma _{\mathrm{CS}}W_{\mathrm{TL}}^{\mathrm{DH}}`$ $`={\displaystyle \frac{Qq_{}}{Q^2+q_{}^2}}W_{\mathrm{TL}}^{\mathrm{DH}}`$ (49) $`W_{\mathrm{\Delta }\mathrm{\Delta }}^{\mathrm{DH}}=`$ $`{\displaystyle \frac{1}{2}}W_L^{\mathrm{DH}}.`$ (50) Immediately one observes that the Lam-Tung relation holds in the CS frame and all other frames considered here. For the Gottfried-Jackson frame the results get particularly simple. The double-hard contributions for deviations from the simple $`1+\mathrm{cos}^2\theta `$ behaviour vanish completely: $$W_\mathrm{L}^{\mathrm{DH}}=0,W_\mathrm{\Delta }^{\mathrm{DH}}=0,W_{\mathrm{\Delta }\mathrm{\Delta }}^{\mathrm{DH}}=0.$$ (51) At first glance this seems to be rather surprising but it is a direct consequence of the special form of all double-hard contributions and therefore has a simple explanation. The double hard process resembles the classical double scattering picture. To understand this we consider again the process depicted in Fig. 1. The antiquark from the hadron undergoes a first scattering with a hard gluon from the nucleus. By subsequently radiating a gluon with large transverse momentum the antiquark returns to the mass shell, but now with transverse momentum $`q_{}`$. The second scattering then corresponds to the lowest-order annihilation process of two massless quarks. Therefore the double-hard process can be factored as the product of two single scattering cross sections, supported by our model that the double-hard matrix element can in good approximation be written as a product of two twist 2 distribution functions. This process naturally fulfills the Lam-Tung relation and depends only on the structure function $`W_{\mathrm{TL}}W_\mu ^\mu `$. Other helicity amplitudes differ only by kinematical factors, i.e. can be obtained from $`W_{\mathrm{TL}}`$ in the Gottfried-Jackson frame by simple rotations. In this probabilistic picture scaling by $`A^{4/3}`$ appears to be very natural. The cross section of the first scattering obviously scales by the volume $`A`$ of the nucleus. Afterwards the projectile parton travels further through the nucleus and has to penetrate the nuclear matter. The probability whether a second interaction takes place on the way to the boundary of the nucleus scales then by the nuclear radius $`A^{1/3}`$. ### 4.3 Calculation of the soft-hard processes For soft-hard scattering we have to keep $`K_{}`$ in the expansion of the gluon momenta. The limit $`K_{}=0`$ would lead to $`K=0`$ which gives no contribution to physical double scattering but just a contribution of the eikonal phase. Note that we are always using covariant gauge where these factors are present. Starting from equation (31) we do a collinear expansion $$\stackrel{~}{H}=\stackrel{~}{H}|_{\begin{array}{c}K_{}=0\\ K_{}^{}=0\end{array}}+K_{}^\lambda \frac{\mathrm{d}}{\mathrm{d}K_{}^\lambda }\stackrel{~}{H}|_{K_{}=0}+\frac{1}{2!}K_{}^\lambda K_{}^\kappa \frac{\mathrm{d}^2}{\mathrm{d}K_{}^\lambda \mathrm{d}K_{}^\kappa }\stackrel{~}{H}|_{K_{}=0}+\mathrm{}.$$ (52) of the terms depending on intrinsic transverse momentum $$\stackrel{~}{H}=\frac{\mathrm{d}x\mathrm{d}x^{}}{\xi _1}e^{ir_1^+z_1^{}}e^{iK^+z_3^{}}e^{iK_{}^{}{}_{}{}^{+}z_4^{}}\overline{H}(\xi _1,\xi _2,x,x^{},K_{},K_{}^{}).$$ (53) We have used $`K_{}=K_{}^{}`$ which allows us to write $$K_{}^\lambda \frac{\mathrm{d}}{\mathrm{d}K_{}^\lambda }=K_{}^\lambda \frac{}{K_{}^\lambda }+K_{}^{}{}_{}{}^{\lambda }\frac{}{K_{}^{}{}_{}{}^{\lambda }}.$$ (54) The mixed terms of second order proportional to $`K_{}^\lambda K_{}^{}{}_{}{}^{\kappa }`$ give the leading contribution to soft hard scattering. The factor $`K_{}^\lambda K_{}^{}{}_{}{}^{\kappa }`$ can be converted into partial derivatives with respect to $`z_3^{}`$, $`z_4^{}`$. A partial integration and an expansion $`A^\rho =A^+P_1^\rho /P_1^++\mathrm{}`$ for soft gluon fields in the nuclear matrix elements turns them into field strength tensors. The $`K_{}`$ and $`K_{}^{}`$ integrals can then be carried out. Note that having soft quarks with transverse momentum and converting the corresponding $`K_{}`$ into covariant derivatives acting on the quark fields would lead to operators contributing only at the level of twist-6 . We perform the pole integrations for the symmetrically cut graphs (M) in Figs.7 and the corresponding asymmetrically cut ones (R) and (L). The residues fix the momentum fractions of the soft gluons to $`x=x_s=x^{}`$ while the hard partons carry $`\xi _1=x_b`$ for the symmetric diagrams and $`\xi _1=x_c`$ for the asymmetric diagrams and $`x_s`$ $`={\displaystyle \frac{k_{}^2}{\xi _2S}},`$ (55) $`x_b`$ $`=\stackrel{~}{\xi }_bx_s={\displaystyle \frac{\xi _2(UQ^2)+Q^22qK_{}k_{}^2}{\xi _2S+TQ^2}}{\displaystyle \frac{k_{}^2}{\xi _2S}},`$ (56) $`x_c`$ $`=\stackrel{~}{\xi }_c={\displaystyle \frac{\xi _2(UQ^2)+Q^2}{\xi _2S+TQ^2}}.`$ (57) We have introduced the notation $`k_{}^2=K_{}^20`$. When we take a forward diagram for soft-hard scattering then the sum of differently cut diagrams (M), (R) and (L) can be divided into two terms. This procedure is described in detail in . In the first term, phase factors and $`\mathrm{\Theta }`$-functions arrange in such a way that this term is bounded in space-time by geometrical arguments and cannot pick up an additional factor $`A^{1/3}`$. This contribution shows no nuclear enhancement. The second term is not bounded and only depends on the hard part of the symmetric diagram (M). We therefore neglect the first term and arrive at $$\begin{array}{c}W_\alpha ^{\mathrm{SH}/ab+c}=(2\pi )^2(4\pi \alpha _s)^2e_q^2_B^1\frac{\mathrm{d}\xi _2}{\xi _2}\frac{1}{\xi _2S+TQ^2}f_{c/H}(\xi _2)\hfill \\ \hfill \left(\frac{g^{\lambda \kappa }}{4}\right)\frac{\mathrm{d}^2}{\mathrm{d}K_{}^\lambda \mathrm{d}K_{}^\kappa }|_{K_{}=0}T_{ab}^{\mathrm{SH}}(x_b)\frac{1}{x_b}H_\alpha ^{\mathrm{SH},ab+c}(x_b,x_s,\xi _2)\end{array}$$ (58) where $`\alpha \{\mu \nu ,\mathrm{TL},\mathrm{L},\mathrm{\Delta },\mathrm{\Delta }\mathrm{\Delta }\}`$. The hard part for the $`qg+\overline{q}`$ process shown in Fig. 7(a) is $$H_{\mu \nu }^{\mathrm{SH}/qg+\overline{q}}=𝒞\frac{1}{(\xi _2S)^2}\frac{1}{4}\mathrm{Tr}\left[R_{\nu \beta }(\mathit{}_2\mathit{})\mathit{}_1\mathit{}_2\mathit{}_1(\mathit{}_2\mathit{})L_{\alpha \mu }\gamma _\mu \mathit{}_1\gamma _\nu \right]_{\begin{array}{c}\xi _1=x_b\\ x=x^{}=x_s\end{array}}(g^{\alpha \beta })$$ (59) $`L_{\alpha \mu }`$ and $`R_{\nu \beta }`$ stand for the tree-level diagrams on the left- and the right-hand side of the diagram in Fig. 7(a) indicated by the blobs. $`𝒞`$ symbolizes the color factor of each single diagram. Hard parts for $`qg+g`$ and $`gg+q`$ subprocesses are obtained in an analogous fashion. The universal matrix elements appearing for soft-hard scattering depend only on the momentum fraction $`x_b`$ of the hard parton and are given by $$\begin{array}{cc}\hfill T_{qg}^{\mathrm{SH}}(x_b)& =\frac{1}{2}dz_4^{}\frac{\mathrm{d}z_1^{}}{2\pi }\frac{\mathrm{d}z_3^{}}{2\pi }\mathrm{\Theta }(z_1^{}z_3^{})\mathrm{\Theta }(z_4^{})e^{ix_bP_1^+z_1^{}}\hfill \\ & P_1\left|F^{\omega +}(z_4^{})\overline{q}(0)\gamma ^+q(z_1^{})F_\omega ^+(z_3^{})\right|P_1\hfill \\ \hfill T_{gg}^{\mathrm{SH}}(x_b)& =\frac{1}{x_bP_1^+}dz_4^{}\frac{\mathrm{d}z_1^{}}{2\pi }\frac{\mathrm{d}z_3^{}}{2\pi }\mathrm{\Theta }(z_1^{}z_3^{})\mathrm{\Theta }(z_4^{})e^{ix_bP_1^+z_1^{}}\hfill \\ & P_1\left|F^{\omega +}(z_4^{})F_\omega ^+(z_3^{})F^{\lambda +}(0)F_\lambda ^+(z_1^{})\right|P_1\hfill \end{array}$$ (60) for scattering off a quark-gluon and a gluon-gluon pair in the nucleus, respectively. We can repeat the arguments given below eq. (44) for the nuclear enhancement. Since the longitudinal momentum fraction of the gluons is essentially zero, we even encounter only one phase factor $`e^{ix_bP_1^+z_1^{}}`$ contrary to the case of double hard scattering discussed before. One therefore might expect that soft-hard matrix elements lead to an enhancement factor proportional to the square of the nuclear radius $`A^{2/3}`$ since only one out of three integrations is limited by phase factors. However, the fields at positions $`z_3^{}`$ and $`z_4^{}`$ are correlated by color forces. The anticipated absence of long-range color correlations in a nucleus implies that integration over $`z_4^{}z_3^{}`$ is still restricted to a region of the size of a nucleon. To build a model for the soft hard matrix elements one again can compare eq. (60) with the definition of the usual twist-2 matrix elements eq. (22). Neglecting for a moment the $`\mathrm{\Theta }`$-functions we see that the twist-4 pieces differ only by an additional $`x_b`$ independent renormalization. Assuming that this can be factored into a term proportional to the nuclear radius and a (possibly scale dependent) constant $`\lambda ^2`$ $`A^{1/3}\lambda ^2`$ $`{\displaystyle dz_4^{}\frac{\mathrm{d}z_3^{}}{2\pi }F^{\omega +}(z_4^{})F_\omega ^+(z_3^{})}`$ (61) the $`x_b`$ dependence of the twist-4 matrix element can be modeled by the usual twist-2 structure function. $$T_{ab}^{\mathrm{SH}}(x)=\lambda ^2A^{\frac{4}{3}}f_{a/A}(x)$$ (62) Eq. (61) can be interpreted as the collective color-magnetic Lorentz force experienced by the projectile parton while traveling through nuclear matter . The only $`K_{}`$-dependence of the hard part will be implicit in $`x_b`$ and $`x_s`$. Therefore we can rewrite the derivatives in the form $$\begin{array}{cc}\hfill W_\alpha ^{\mathrm{SH},ab+c}=& (2\pi )^2(4\pi \alpha _s)^2e_q^2_B^1\frac{\mathrm{d}\xi _2}{\xi _2}\frac{1}{\xi _2S+TQ^2}f_{c/H}(\xi _2)\hfill \\ & \frac{1}{2}[\frac{^2}{x_c^2}\left(T_{ab}^{\mathrm{SH}}(x_c)\frac{1}{x_c}H_\alpha ^{\mathrm{SH},ab+c}(x_c,\xi _2)\right)\frac{2q_{}^2}{(\xi _2S+TQ^2)^2}+\hfill \\ & +\frac{}{x_c}\left(T_{ab}^{\mathrm{SH}}(x_c)\frac{1}{x_c}H_\alpha ^{\mathrm{SH},ab+c}(x_c,\xi _2)\right)\frac{2(Q^2T)}{\xi _2S(\xi _2S+TQ^2)}+\hfill \\ & +T_{ab}^{\mathrm{SH}}(x_c)\frac{1}{x_c}(H_\alpha ^{\mathrm{SH},ab+c})^{}(x_c,\xi _2)\frac{2}{\xi _2S}]\hfill \end{array}$$ (63) where $`H(x_c,\xi _2)=H(x_c,x_s=0,\xi _2)`$ and $`H^{}(x_c,\xi _2)=\frac{}{x_s}H(x_c,x_s=0,\xi _2)`$. Note that in the last term containing a derivative with respect to $`x_s`$ is absent since it turns out that $`H_{\mathrm{TL}}^{\mathrm{SH}}`$ does not depend explicitly on $`x_s`$. We give the results for the hard parts in the Collins-Soper frame in Appendix 7.4. There are some subtleties in removing the unphysical degrees of freedom for processes of the class $`qg+g`$. There we have soft quarks from the nucleus with dominating component $`A^+`$ and for the hard gluons from the single hadron we have to take only the transverse polarizations. For $`W_{\mathrm{TL}}^{\mathrm{SH}}`$ we reproduce the results of . Unlike for double-hard scattering the amplitudes $`W_L^{\mathrm{SH}}`$, $`W_\mathrm{\Delta }^{\mathrm{SH}}`$ and $`W_{\mathrm{\Delta }\mathrm{\Delta }}^{\mathrm{SH}}`$ cannot be obtained in a trivial way from $`W_{\mathrm{TL}}`$. Related to that is an obvious violation of the Lam-Tung relation. ## 5 Numerical Results In order to cover different energies $`S`$ we calculated helicity amplitudes and angular coefficients for two different settings: for $`252\mathrm{GeV}`$ $`\pi ^{}`$ on tungsten as in the Fermilab E615 experiment and for future $`p+A`$ collider experiments at RHIC energies with $`250\mathrm{GeV}`$ protons and $`A\times 100\mathrm{GeV}`$ large nuclei. The main task before doing the numerics is to insert a suitable model for the twist-4 matrix elements in the nucleus. The models introduced in the last sections for the double-hard matrix elements $$T_{ab}^{\mathrm{DH}}(x_a,x_h)=CA^{\frac{4}{3}}f_{a/A}(x_a)f_{b/A}(x_h).$$ (64) and for the soft-hard matrix elements $$T_{ab}^{\mathrm{SH}}(x)=\lambda ^2A^{\frac{4}{3}}f_{a/A}(x)$$ (65) are already given by LQS . We tried to argue above that their dependence on the parton momentum fractions and their scaling behaviour with respect to $`A`$ seem to be quite natural. Nevertheless the correct normalization of the models is not yet clear. The hope is that $`C`$ and $`\lambda `$ are universal constants. In the early days of the model a value of $`C=0.072\mathrm{GeV}^2`$ was taken from theory while $`\lambda ^2`$ was taken from comparison with experiment. This was done by analyzing dijet photoproduction on nuclei leading to a value $`\lambda ^2=0.050.1\mathrm{GeV}^2`$. Later an analysis of $`q_{}`$-broadening in hadron-nucleon Drell-Yan processes implied a value of $`\lambda ^2=0.01\mathrm{GeV}^2`$. Obviously there is a discrepancy, which may either point to a strong scale dependence or questions the universality of the soft-hard nuclear matrix elements. Note that the soft interaction besides the hard process in soft-hard double scattering is a final state interaction for dijet photoproduction but an initial state interaction for the Drell-Yan process. In their latest work Guo, Qiu and Zhang used the fact that there is a formal connection between soft-hard and double-hard matrix elements via the limit $$\underset{x_g0}{lim}x_gT_{qg}^{\mathrm{DH}}(x_q,x_g)=T_{qg}^{\mathrm{SH}}(x_q)$$ (66) which leads with the models of eqs. (64),(65) to $$C\frac{\lambda ^2}{x_gg(x_g)|_{x_g0}}.$$ (67) They take $`x_gg(x_g)|_{x_g0}3`$ as the gluon momentum density at $`x_g0`$. Taking into account the later value of $`\lambda ^2`$ this leads to a normalization factor $`C`$ which is an order of magnitude below the former theoretical value. But it is more consistent with the value of $`C0.005\mathrm{GeV}^2`$ which we obtained in section 4 from a naive normalization argument. In another, more geometrical, model one can estimate the two-parton area density that can be seen by the projectile parton in the nucleus by the product of two single-parton area densities. Integration of the two-parton area density over the effective transverse area of the nucleus yields exactly the right-hand side of eq. (64) with a constant $`C=9/(8\pi r_0^2)0.01\mathrm{GeV}^2`$. Obviously $`C`$ is presently basically unknown and a true improvement can only be expected if comparisons of a larger number of observables with more precise experimental data are available. For soft-hard scattering we use the smaller value of $`\lambda ^2=0.01\mathrm{GeV}^2`$ obtained from Drell-Yan data to maintain consistency. For double-hard scattering we have chosen the maximal value $`C=0.072\mathrm{GeV}^2`$ to give a feeling how big the effects of double scattering might be. We take the twist-2 distribution functions from the CTEQ3M parameterization for the nucleons and from the recent GRS fits for the pion . Where nuclear effects on twist-2 distribution functions are taken into account the EKS98 parametrization is used. Let us first address the question how this approach works with the pion data from E615. In Fig. 8 we show our results for the angular coefficients $`\lambda `$, $`\mu `$ and $`\nu `$ in the Gottfried-Jackson frame for the process $`\pi ^{}+W`$ at E615 energy. We also give experimental data from the E615 collaboration. We can fairly well describe the data for $`\lambda `$ at low transverse momentum but fail for $`\mu `$ and $`\nu `$. It seems to be an essential feature that corrections due to nuclear enhanced twist-4 matrix elements can be quite large for the helicity amplitudes but the effects largely cancel for the angular coefficients which are ratios thereof. Therefore changes of the angular coefficients in the range $`q_{}Q`$ are small compared with the pure twist-2 calculation. Experimental values have large error bars above $`q_{}=2\mathrm{GeV}`$. We already pointed out that data at $`q_{}/Q1`$ would provide a much better test for the twist-4 calculation because of the absence of soft-hard interference terms and the absence of double logarithmic corrections of the type $`\mathrm{log}^2(Q^2/q_{}^2)`$ which have to be resummed. We have to conclude that nuclear enhanced double scattering contributions are not able to yield substantial improvement for the old pion data. Deviations of the $`\pi W`$ data from the naive perturbative QCD-predictions might be explained by other approaches like the Berger-Brodsky mechanism which takes into account higher twist contributions of the pion wave function , in terms of non-trivial spin correlations in the QCD-vacuum or via contributions of the chiral-odd $`T`$-odd distribution function $`h_1^{}`$ . We turn to $`p+Au`$ collisions at RHIC with $`S=10^5\mathrm{GeV}^2`$ per nucleon pair. We give helicity amplitudes instead of angular coefficients since these provide a clearer measure of nuclear enhanced higher twist effects as explained above. To make numerical results of helicity amplitudes $`W_\alpha `$, $`\alpha \{\mathrm{TL},\mathrm{L},\mathrm{\Delta },\mathrm{\Delta }\mathrm{\Delta }\}`$ more meaningful, we introduce the prefactor given in eq. (13), $$w_\alpha =\frac{\alpha ^2}{64\pi ^3SQ^2}W_\alpha ,$$ (68) to convert the quantities $`W_\alpha `$ into differential cross sections. In Fig. 9 we give single scattering, double-hard and soft-hard results for the helicity amplitudes $`w_{\mathrm{TL}}`$, $`w_{\mathrm{TL}}`$ and $`w_\mathrm{\Delta }`$ as functions of rapidity $`y`$ in the Collins-Soper frame. One can read off the main results already given in . First the nuclear enhanced twist-4 contribution is absolutely important at RHIC energies. Even at quite large values of $`Q=5\mathrm{GeV}`$ and $`q_{}=4\mathrm{GeV}`$ used for the calculation its effect can exceed the leading-twist result for the parameters used. Second the double-hard contribution is by far larger than the soft-hard part in the kinematical range under consideration. This can be seen for all helicity amplitudes in Fig. 9. Even if we use smaller values of $`C`$ proposed above double-hard scattering would reach the same order of magnitude as single scattering. We also give results with EKS98 modifications to nuclear parton distributions. These parametrizations take into account shadowing, antishadowing and other nuclear effects on parton distributions, first described by the European Muon Collaboration. The numerical effect of these modifications on the single scattering results are small, but they turn out to be quite large for double-hard scattering. Note that it is not clear whether these nuclear effects for twist-2 parton distributions should be carried over to the models for twist-4 matrix elements. We give results for modified and unmodified distributions to indicate the further uncertainty of the twist-4 matrix elements. The sensitivity of double-hard scattering to these EKS98 modifications is simply related to the fact that they involve a product of two nuclear parton distributions. The general question about the correct small-$`x`$ behaviour of the twist-4 nuclear matrix elements is very interesting since it has a big influence at RHIC energies. We return to the question of the Lam-Tung relation. We already saw analytically that double-hard scattering respects this relation while soft-hard does not. In Fig. 10 we show the difference $`A_0A_2`$. It only picks up contributions from soft-hard scattering and therefore this quantity is very small. That means that violation of the Lam-Tung sum rule is a tiny effect for the process under consideration. The characteristic feature of the rapidity distribution is the peak of the double-hard distribution at negative rapidities (i.e. for the photon emitted in direction of the initial single hadron). The reason for this is that negative rapidity prefers small Bjorken-$`x`$ for partons from the nucleus. Parton distributions at small parton momentum fractions are large particularly for gluons and they enter quadratically for double-hard scattering. In Fig. 11 we separated the different partonic subprocesses contributing to double-hard and soft-hard scattering to enable a deeper analysis. One can see that double-hard processes with a quark and a gluon from the nucleus dominate over those with two quarks at negative rapidities (i.e. small $`x_h`$). On the other hand $`qg+\overline{q}`$ is more important than $`qg+g`$ in this region since momentum fraction $`\xi _2`$ of the parton from the single hadron is large and quarks dominate over gluons in that case. All the rapidity spectra can be understood in this way. Fig. 12 shows dependence of $`w_{\mathrm{TL}}`$ on photon mass $`Q`$ and transverse momentum $`q_{}`$. It can be seen that soft-hard and double-hard contributions as expected for higher-twist processes fall off more rapidly with $`Q`$ and $`q_{}`$ than (twist-2) single-scattering. ## 6 Summary We have calculated nuclear enhanced double-scattering contributions to Drell-Yan pair production in hadron-nucleus collisions using the framework of Luo, Qiu and Sterman. We found that the nuclear enhancement at RHIC energies and large transverse momentum makes double scattering a contribution that is equally important as single scattering. Measurements of angular coefficients and helicity amplitudes can distinguish double-hard and soft-hard scattering. Double-hard scattering has the interpretation of two sequential hard scatterings of onshell partons and is therefore related to probabilistic pictures of multiple scattering that are widely used in the phenomenology of heavy-ion reactions. In this spirit further comparision with alternative approaches to multiple scattering would be very interesting. Double-hard scattering respects the Lam-Tung relation and, in contrast to soft-hard scattering, does not show non-trivial deviations from the $`1+\mathrm{cos}^2\theta `$ angular distribution. Numerically double-hard scattering dominates over the soft-hard contribution. Soft-hard scattering violates the Lam-Tung relation but this violation is an effect of order $`10^3`$ for the model used. Double-hard scattering can be modeled as the product of two twist-2 structure functions which give the probability for finding two partons in the nucleus. The soft-hard matrix elements on the other hand encodes essential non-perturbative properties of the nucleus, describing the influence of the collective color field on the transverse momentum of the produced lepton pair. Inserting reasonable models for the corresponding matrix elements we predict, that collective color effects are negligible compared to the effects of subsequent independent hard scatterings in the kinematical range considered here. However the absolute normalization of the twist-4 matrix elements in these models is not yet clear. The double-hard scattering shows an interesting asymmetry in the rapidity distribution of the Drell-Yan pair. The pair is predominantly produced in the projectile, i.e. single hadron rapidity region. Although the absolute counting rates for the DY pairs at large transverse momentum are not high the predicted effect fits perfectly well in the PHENIX detector acceptance . This asymmetry therefore should be observable. An analysis of $`W_L`$ and $`W_\mathrm{\Delta }`$, particularly in the Gottfried-Jackson frame, as well as the measurement of $`W_{\mathrm{TL}}`$ (i.e. the angular integrated cross section) as a function of rapidity can check whether double-hard scattering is the dominant contribution. A possible violation of the Lam-Tung sum rule provides a further possibility to check the picture of soft-hard and double-hard scattering in general. If our predictions with respect to the ratio of soft-hard and double-hard scattering are confirmed by experiment it would provide support for the description of heavy ion reactions in terms of incoherent multiple scattering of free partons. Note that the prediction of a trivial frame dependence of double-hard scattering and the conservation of the Lam-Tung relation are independent of the models for the twist-4 distribution functions and provide good tests for the entire LQS approach. Acknowledgements. We are grateful to X. Guo for helpful correspondence. In course of the work we benefitted from discussions with V. Braun, L. Syzmanowsky, and O. Teryaev. E. S. thanks G. Sterman and X.N. Wang for useful discussions. ## 7 Appendix ### 7.1 Transformation of amplitudes to different frames Here we list the matrices $`M_{(\beta )}`$ defined in eq. (20) that map the invariant projections of the hadronic tensor to the set of helicity amplitudes ($`W_{\mathrm{TL}}`$, $`W_L`$, $`W_\mathrm{\Delta }`$, $`W_{\mathrm{\Delta }\mathrm{\Delta }}`$). They are $$\begin{array}{cc}\hfill M_{\mathrm{CS}}& =\left(\begin{array}{cccc}\frac{1}{2}& 0& 0& 0\\ 0& \frac{1}{4\mathrm{cos}^2\gamma _{\mathrm{CS}}}& \frac{1}{4\mathrm{cos}^2\gamma _{\mathrm{CS}}}& \frac{1}{4\mathrm{cos}^2\gamma _{\mathrm{CS}}}\\ 0& \frac{1}{4\mathrm{sin}\gamma _{\mathrm{CS}}\mathrm{cos}\gamma _{\mathrm{CS}}}& \frac{1}{4\mathrm{sin}\gamma _{\mathrm{CS}}\mathrm{cos}\gamma _{\mathrm{CS}}}& 0\\ \frac{1}{2}& \frac{1+\mathrm{cos}^2\gamma _{\mathrm{CS}}}{8\mathrm{cos}^2\gamma _{\mathrm{CS}}\mathrm{sin}^2\gamma _{\mathrm{CS}}}& \frac{1+\mathrm{cos}^2\gamma _{\mathrm{CS}}}{8\mathrm{cos}^2\gamma _{\mathrm{CS}}\mathrm{sin}^2\gamma _{\mathrm{CS}}}& \frac{13\mathrm{cos}^2\gamma _{\mathrm{CS}}}{8\mathrm{cos}^2\gamma _{\mathrm{CS}}\mathrm{sin}^2\gamma _{\mathrm{CS}}}\end{array}\right)\hfill \\ \hfill M_{\mathrm{GJ}}& =\left(\begin{array}{cccc}\frac{1}{2}& 0& 0& 0\\ 0& 1& 0& 0\\ 0& \mathrm{cot}\gamma _{\mathrm{GJ}}& 0& \frac{1}{2\mathrm{sin}\gamma _{\mathrm{GJ}}}\\ \frac{1}{2}& \frac{1+\mathrm{cos}^2\gamma _{\mathrm{GJ}}}{2\mathrm{sin}^2\gamma _{\mathrm{GJ}}}& \frac{1}{\mathrm{sin}^2\gamma _{\mathrm{GJ}}}& \frac{\mathrm{cos}\gamma _{\mathrm{GJ}}}{\mathrm{sin}^2\gamma _{\mathrm{GJ}}}\end{array}\right)\hfill \end{array}$$ (69) for the Collins-Soper and the Gottfried-Jackson frame, respectively. $`\gamma _{\mathrm{CS}}`$ denotes the angle between $`𝐏_1`$ and the $`𝐙`$ axis in the photon rest frame and $`\gamma _{\mathrm{GJ}}`$ is the angle between $`𝐏_1`$ and $`𝐏_2`$. Obviously $`2\gamma _{\mathrm{CS}}+\gamma _{\mathrm{GJ}}=\pi `$ and in terms of hadronic Mandelstam variables we have $`\mathrm{cos}\gamma _{\mathrm{CS}}=\sqrt{{\displaystyle \frac{Q^2S}{(Q^2T)(Q^2U)}}},`$ $`\mathrm{sin}\gamma _{\mathrm{CS}}=\sqrt{1{\displaystyle \frac{Q^2S}{(Q^2T)(Q^2U)}}}`$ (70) $`\mathrm{cos}\gamma _{\mathrm{GJ}}=1{\displaystyle \frac{2Q^2S}{(Q^2T)(Q^2U)}},`$ $`\mathrm{sin}\gamma _{\mathrm{GJ}}={\displaystyle \frac{2Q^2S}{(Q^2T)(Q^2U)}}\sqrt{{\displaystyle \frac{(Q^2T)(Q^2U)}{Q^2S}}1}.`$ ### 7.2 Single-scattering results Here we list the results for the hard parts $`H_\alpha `$ of single-scattering helicity amplitudes in the Collins-Soper frame. They are defined in eq. (26). We distinguish annihilation and Compton processes. $$\begin{array}{cc}\hfill H_{\mathrm{TL}}^{q+\overline{q}}& =𝒞_1\frac{1}{tu}\left((Q^2t)^2+(Q^2u)^2\right)\hfill \\ \hfill H_\mathrm{L}^{q+\overline{q}}& =𝒞_1\left(\frac{Q^2t}{Q^2u}+\frac{Q^2u}{Q^2t}\right)\hfill \\ \hfill H_\mathrm{\Delta }^{q+\overline{q}}& =𝒞_1\sqrt{\frac{Q^2s}{tu}}\left(\frac{Q^2u}{Q^2t}\frac{Q^2t}{Q^2u}\right)\hfill \\ \hfill H_{\mathrm{\Delta }\mathrm{\Delta }}^{q+\overline{q}}& =\frac{1}{2}H_L^{q+\overline{q}}\hfill \\ \hfill H_{\mathrm{TL}}^{q+g}& =𝒞_2\frac{(Q^2s)^2+(Q^2t)^2}{st}\hfill \\ \hfill H_\mathrm{L}^{q+g}& =𝒞_2\frac{u}{s}\frac{(Q^2+s)^2+(Q^2t)^2}{(Q^2t)(Q^2u)}\hfill \\ \hfill H_\mathrm{\Delta }^{q+g}& =𝒞_2\sqrt{\frac{Q^2u}{st}}\frac{2(Q^2t)^2(Q^2u)^2}{(Q^2t)(Q^2u)}\hfill \\ \hfill H_{\mathrm{\Delta }\mathrm{\Delta }}^{q+g}& =\frac{1}{2}H_L^{q+g}\hfill \\ \hfill H_{\mathrm{TL}}^{g+q}& =𝒞_2\frac{(Q^2s)^2+(Q^2u)^2}{su}\hfill \\ \hfill H_\mathrm{L}^{g+q}& =𝒞_2\frac{t}{s}\frac{(Q^2+s)^2+(Q^2u)^2}{(Q^2t)(Q^2u)}\hfill \\ \hfill H_\mathrm{\Delta }^{g+q}& =𝒞_2\sqrt{\frac{Q^2t}{su}}\frac{2(Q^2u)^2(Q^2t)^2}{(Q^2t)(Q^2u)}\hfill \\ \hfill H_{\mathrm{\Delta }\mathrm{\Delta }}^{g+q}& =\frac{1}{2}H_L^{g+q}.\hfill \end{array}$$ (71) $`𝒞_1=C_F/N_c=4/9`$ and $`𝒞_2=1/(2N_c)=1/6`$ are essentially the color factors. ### 7.3 Double-hard results In this section we give the hard parts $`H_{\mathrm{TL}}`$ of the projections $`W_{\mathrm{TL}}^{\mathrm{DH}}`$ for double-hard processes. They do not depend on the frame. $$\begin{array}{cc}\hfill H_{\mathrm{TL}}^{DH,qg+\overline{q}}& =\frac{2}{27}\left(\frac{(Q^2T)}{\xi _2S}+\frac{\xi _2S}{Q^2T}\right)+\frac{1}{6}\frac{(Q^2T)^2+(\xi _2S)^2}{\xi _2S+TQ^2}\hfill \\ \hfill H_{\mathrm{TL}}^{DH,qg+g}& =\frac{1}{36}\left(\frac{\xi _2S+TQ^2}{Q^2T}+\frac{Q^2T}{\xi _2S+TQ^2}\right)\frac{1}{16}\frac{(Q^2T)^2+(\xi _2S+TQ^2)^2}{(\xi _2S)^2}\hfill \\ \hfill H_{\mathrm{TL}}^{DH,q\overline{q}+g}& =\frac{2}{27}\left(\frac{\xi _2S+TQ^2}{\xi _2S}+\frac{\xi _2S}{\xi _2S+TQ^2}\right)+\frac{1}{6}\frac{(\xi _2S)^2+(\xi _2S+TQ^2)^2}{(Q^2T)^2}\hfill \\ \hfill H_{\mathrm{TL}}^{DH,q\overline{q}+q}& =\frac{2}{27}\left(\frac{(\xi _2S)^2+(\xi _2S+TQ^2)^2}{(Q^2T)^2}+\frac{(Q^2T)^2+(\xi _2S+TQ^2)^2}{(\xi _2S)^2}\right)+\hfill \\ & +\frac{4}{81}\frac{(\xi _2S+TQ^2)^2}{\xi _2S(Q^2T)}\hfill \\ \hfill H_{\mathrm{TL}}^{DH,q\overline{q}+\overline{q}}& =\frac{2}{27}\left(\frac{(\xi _2S)^2+(\xi _2S+TQ^2)^2}{(Q^2T)^2}+\frac{(Q^2T)^2+(\xi _2S)^2}{(\xi _2S+TQ^2)^2}\right)\hfill \\ & \frac{4}{81}\frac{(\xi _2S)^2}{(Q^2T)(\xi _2S+TQ^2)}\hfill \\ \hfill H_{\mathrm{TL}}^{DH,qq+\overline{q}}& =\frac{2}{27}\left(\frac{(Q^2T)^2+(\xi _2S+TQ^2)^2}{(Q^2T)^2}+\frac{(Q^2T)^2+(\xi _2S)^2}{(\xi _2S+TQ^2)^2}\right)+\hfill \\ & +\frac{4}{81}\frac{(Q^2T)^2}{\xi _2S(\xi _2S+TQ^2)}\hfill \\ \hfill H_{\mathrm{TL}}^{DH,q\overline{q}+q^{}}& =\frac{2}{27}\frac{(\xi _2S)^2+(\xi _2S+TQ^2)^2}{(Q^2T)^2}\hfill \\ \hfill H_{\mathrm{TL}}^{DH,qq^{}+\overline{q}^{}}& =\frac{2}{27}\frac{(Q^2T)^2+(\xi _2S+TQ^2)^2}{(\xi _2S)^2}\hfill \\ \hfill H_{\mathrm{TL}}^{DH,qq^{}+\overline{q}}& =\frac{2}{27}\frac{(Q^2T)^2+(\xi _2S)^2}{(\xi _2S+TQ^2)^2}\hfill \end{array}$$ (72) ### 7.4 Soft-hard results Here we list the hard parts $`H(T_\alpha )`$ of the invariant projections $`T_{\mathrm{TL}}^{\mathrm{SH}}`$, $`T_{\mathrm{L1}}^{\mathrm{SH}}`$, $`T_{\mathrm{L2}}^{\mathrm{SH}}`$ and $`T_{\mathrm{L12}}^{\mathrm{SH}}`$ of the hadronic tensor, defined in section 2.2, for soft-hard processes. The results can be projected into a specific frame via eq. (20) and the matrices given in Appendix 7.1. This is mathematically trivial but it complicates the expressions substantially. For this reason we give only the invariant projections. Note that especially $`H_{\mathrm{TL}}=H(T_{\mathrm{TL}})/2`$ for every frame described in section 2.2. $$\begin{array}{cc}\hfill H(T_{\mathrm{TL}}^{qg+\overline{q}})& =\frac{4}{27}\left(\frac{(Q^2T)(x_cx_a)}{x_c(\xi _2S+TQ^2)}+\frac{2x_a\xi _2S}{(x_cx_a)(\xi _2S+TQ^2)}+\frac{x_c(\xi _2S+TQ^2)}{(Q^2T)(x_cx_a)}\right)\hfill \\ \hfill H(T_{\mathrm{L1}}^{qg+\overline{q}})& =\frac{8}{27}\frac{x_a\xi _2S}{x_c(Q^2T)}\hfill \\ \hfill H(T_{\mathrm{L2}}^{qg+\overline{q}})& =\frac{8}{27}(\frac{\xi _2S+TQ^2}{(x_c+x_s)S+UQ^2}\frac{x_sQ^2S}{(Q^2U)^2}\frac{x_a}{x_ax_c}\frac{x_cS((x_c+x_s)S+UQ^2)}{(Q^2U)^2}\hfill \\ & \frac{x_a}{x_ax_c}\frac{S^2}{(Q^2U)^2}\frac{(x_c+x_s)((x_cx_s)(Q^2T)+\xi _2(Q^2U)x_c\xi _2S)(x_cx_s)Q^2}{\xi _2S+TQ^2})\hfill \\ \hfill H(T_{\mathrm{L12}}^{qg+\overline{q}})& =\frac{8}{27}(\frac{x_a}{x_ax_c}\frac{\xi _2S^2}{(Q^2T)(Q^2T)}\frac{(x_cx_s)(Q^2T)+\xi _2(Q^2U)Q^2x_c\xi _2S}{\xi _2S+TQ^2}+\hfill \\ & +\frac{x_as}{x_c(Q^2U)}\frac{(x_c+x_s)(\xi _2S+TQ^2)+\xi _2((x_c+x_s)S+UQ^2)Q^2+x_c\xi _2S}{\xi _2S+TQ^2})\hfill \\ \hfill H(T_{\mathrm{TL}}^{qg+g})& =\frac{1}{8}\left(\frac{x_c\xi _2S}{(x_cx_a)(Q^2T)}\frac{2x_a(\xi _2S+TQ^2)}{\xi _2S(x_cx_a)}+\frac{(x_cx_a)(Q^2T)}{x_c\xi _2S}\right)\hfill \\ \hfill H(T_{\mathrm{L1}}^{qg+g})& =\frac{1}{4}\frac{x_a(\xi _2S+TQ^2)}{x_c(Q^2T)}\hfill \\ \hfill H(T_{\mathrm{L2}}^{qg+g})& =\frac{1}{4}(\frac{x_sQ^2}{x_c(Q^2U)}\frac{(x_c+x_s)S+UQ^2}{\xi _2(Q^2U)}\hfill \\ & \frac{x_a}{x_ax_c}\frac{(x_c+x_s)S+UQ^2}{\xi _2(Q^2U)}\frac{(x_cx_s)(Q^2T)+\xi _2(Q^2U)Q^2x_c\xi _2S}{Q^2U}\hfill \\ & \frac{x_ax_c}{x_ax_c}\frac{((x_c+x_s)S+UQ^2)(\xi _2S+TQ^2)+(Q^2T)(Q^2U)Q^2S}{\xi _2(Q^2U)^2})\hfill \\ \hfill H(T_{\mathrm{L12}}^{qg+g})& =\frac{1}{4}(\frac{x_a}{x_ax_c}\frac{\xi _2S+TQ^2}{Q^2T)}\frac{(x_c+x_s)(Q^2T)\xi _2(Q^2U)Q^2+x_c\xi _2S}{\xi _2(Q^2U)}+\hfill \\ & +\frac{x_a}{x_c}\frac{(x_c+x_s)(\xi _2S+TQ^2)+\xi _2((x_c+x_s)S+UQ^2)+Q^2x_c\xi _2S}{\xi _2(Q^2U)})\hfill \\ \hfill H(T_{\mathrm{TL}}^{qg+g})& =\frac{1}{18}\left(\frac{\xi _2S}{\xi _2S+TQ^2}\frac{2Q^2(Q^2T)(x_cx_a)}{x_c^2\xi _2S(\xi _2S+TQ^2)}\frac{\xi _2S+TQ^2}{\xi _2S}\right)\hfill \\ \hfill H(T_\mathrm{L}^{qg+g})& =\frac{2}{9}\frac{x_a(x_ax_c)}{x_c^2}\hfill \\ \hfill H(T_\mathrm{\Delta }^{qg+g})& =\frac{1}{9}\frac{Q^2}{Q^2U}(\frac{(x_c+x_s)S+UQ^2}{\xi _2(Q^2U)}+\frac{x_sS}{(Q^2U)}\frac{S}{\xi _2S+TQ^2}\hfill \\ & \frac{(x_cS+UQ^2)((x_c+x_s)(\xi _2S+TQ^2)+\xi _2((x_c+x_s)S+UQ^2)+Q^2x_c\xi _2S)}{x_c\xi _2(Q^2U)(\xi _2S+TQ^2)}\hfill \\ & \frac{2x_sQ^2((x_c+x_s)S+UQ^2)}{x_c^2\xi _2(Q^2U)(\xi _2S+TQ^2)})\hfill \\ \hfill H(T_{\mathrm{\Delta }\mathrm{\Delta }}^{qg+g})& =\frac{1}{9}\frac{x_a}{\xi _2x_c}(\frac{2x_c\xi _2(Q^2U)+x_c(x_c+x_s)(Q^2T)+x_c^2\xi _2S+2x_2Q^2}{Q^2U}+\hfill \\ & +\frac{\xi _2((x_c+x_s)S+UQ^2)(2Q^2+x_c(Q^2T))+3x_cQ^2(Q^2T)+2Q^4+x_c^2\xi _2^2S^2}{(Q^2U)(\xi _2S+TQ^2)})\hfill \end{array}$$
warning/0002/gr-qc0002065.html
ar5iv
text
# GRAVITY QUANTA, ENTROPY AND BLACK HOLES ## I INTRODUCTION. The Heisenberg uncertainty principle is the basis for quantum mechanics, through the use of Planck’s constant. On the other hand, quantum gravity is not yet an established theory. One would like to have a parallel treatment for gravity with a gravitational Planck’s constant for this case. We derive here the value for such a constant. It turns out to be of the same order of magnitude as the one proposed by Zecca() but following a very different argument. A step forward in quantum gravity would have important cosmological implications, and here we link black hole properties with the size of the Universe, an important cosmological parameter. The hope is to get some light in the direction of a quantum treatment of gravitation. Based upon the gravitational uncertainty principle we derive the value for the quantum of gravity. Using this we define the entropy of any system, from a gravitational point of view, and get an equivalent to the 2<sup>nd</sup> law of thermodynamics. Entropy is defined in a cosmological context. In particular, our definition for entropy keeps its extensive quality. One can then derive the primordial black hole entropy. In fact, for the entropy of any black hole we get a result in contrast with the Bekenstein () and Hawking () formulation. In our case the extensive property for the entropy is kept valid. Finally we prove that with these new concepts the radiation from black holes is many orders of magnitude less than the Hawking radiation formula. ## II A GRAVITATIONAL UNCERTAINTY PRINCIPLE. Elementary particles of typical mass $`m_p`$ have a characteristic size $`r_p`$ of the order of the Compton wavelength $$r_p\frac{\mathrm{}}{m_pc}210^{13}cm$$ (1) where $`\mathrm{}`$ is Planck’s constant and $`c`$ the speed of light. For the numerical value we have used the proton mass. If these particles emit gravity quanta of characteristic wavelength $`\lambda `$ one must have $`\lambda r_p`$ . i.e., one must have the same order of magnitude between the antenna size and the wavelength of the emitted radiation. Then, the gravity quantum equivalent mass $`m_1`$ must be related to a gravitational Planck’s constant $`H`$ in the following way, $$\lambda r_p\frac{\mathrm{}}{m_pc}\frac{H}{m_1c}$$ (2) where $`\mathrm{}H`$ and $`m_pm_1.`$ For gravity quanta one expects for $`H`$ to play the role of $`\mathrm{}`$. Let us see the rate of emission of gravity quanta. The characteristic speed must be the speed of light, and sizes being of the order of $`r_p`$ as in (1) the rate of emission $`R=1/\tau `$ must be $$R=\frac{1}{\tau }=\frac{c}{r_p}\frac{m_pc^2}{\mathrm{}}\frac{m_1c^2}{H}$$ (3) Let us introduce gravitation explicitly in the picture. We know that the gravitational field has negative energy. In fact the gravitational potential between two masses $`M`$ and $`m`$ at a distance $`r`$ is $$\frac{GMm}{r}$$ (4) When the distance $`r`$ tends to infinity this energy tends to zero as it should. As the distance $`r`$ becomes less and less, the gravitational potential decreases, becoming more negative. Applying an order of magnitude estimate for the energy equation, considering a hydrogen atom of total energy $`E`$ and rest energy $`E_0`$ falling radial in the gravitational field of a much larger mass $`M`$ one has $$E\frac{GM\left(\frac{E}{c^2}\right)}{r}=E_0$$ (5) As the atom falls in the gravitational field of $`M`$, i.e. approaches $`M`$ more and more if geometrically possible ($`M`$ concentrated enough), $`E`$ becomes relativistic due to the increase in kinetic energy and one can then neglect the term $`E_0`$ in (5). Hence, there is a limit $`r_{bh}`$ in the distance of approach given by $$r_{bh}\frac{GM}{c^2}$$ (6) This is the order of magnitude of the gravitational radius of the mass $`M`$. It is the limit of approach that any body can have going towards the mass $`M`$. It is well known that if $`M`$ has a size of this order or less, it is then a black hole. From an outside observer’s point of view the body never reaches $`M`$. It would take an infinite time for that. From the point of view of the body time runs differently and it crosses the gravitational radius in a never returning way. We see then the meaning of the gravitational radius as a characteristic size of the corresponding black hole when the mass is all inside this radius. Emission of gravity quanta implies emission of negative energy, and therefore an increase in the positive mass of the source, the particle. Having found the rate of emission (3) of gravity quanta of mass $`m_1`$ one can equate the rate of mass emission to the ratio $`m_p/t`$, where $`t`$ is the age of the Universe, i.e., $$\frac{m_1}{\tau }\frac{m_p}{t}$$ (7) Using (3) and (7) one gets the equivalent set $$\frac{m_1^2c^2}{H}\frac{m_p}{t}$$ $$m_1\frac{H}{\mathrm{}}m_p$$ (8) and solving for $`m_1`$ and $`H`$ we finally have $$m_1\frac{\mathrm{}}{c^2t}=\frac{\mathrm{}}{m_pc}\frac{1}{ct}m_p10^{41}m_p$$ $$H\frac{\mathrm{}}{m_pc}\frac{1}{ct}\mathrm{}10^{41}\mathrm{}$$ (9) where we have used the present age of the Universe $`t1.510^{10}`$ years. In 1975 Zecca() proposed a gravitational Planck’s constant, derived from a very different argument, arriving at a value of this order of magnitude for $`H`$ too. The quantum world is governed by Planck’s constant. But quantum mechanics does not include gravitation. A quantum gravity approach should have the gravitational Planck’s constant derived here as the fundamental constant for gravitation, particularly for quantum radiation. The parallel gravitational uncertainty principle must be formulated with the corresponding gravitational Planck’s constant $`H`$. For example, the radiation of a quantum of gravity of energy $`E_0=m_1c^2`$ must be governed by the gravitational uncertainty relation $$\tau E_0H$$ i.e. $$\frac{\mathrm{}}{m_pc^2}m_1c^2H$$ (10) which is equivalent to the relation found in (9). ## III ENTROPY We may see now that entropy and mass are proportional if we take the emission of $`m_1`$, the quantum of gravity, as the production of a unit of entropy. There is a deep physical meaning here. We are claiming that the basic quantum of gravitation is $`m_1`$, and that it represents an entity by itself: we just count the number of these emitted entities to get the entropy. Hence we are counting items of deep physical significance, the gravity quanta, to get the entropy. Since $`m_p`$ is of the order of $`10^{41}m_1`$ we can take the number of emitted gravity quanta as the entropy of the particle, about $`10^{41}`$. This is the number of parts that the particle has contributed to the Universe in the form of gravity quanta. This means that the entropy of the seeable Universe (size $`ct`$) is this number times the number of particles in it (we approximately know the density and size of the Universe, so that we know the equivalent number of protons): $`10^{41}10^{79}=10^{120}`$ . Entropy is an extensive quantity, and under this view it is just the total mass of the system in $`m_1`$ units: it is not a question of expressing entropy in a particular unit. It is a question of taking the emission of gravity quanta as the accounting for the creation of entropy. Of course we are expressing the entropy $`S`$ in non-dimensional form, $`S/k`$, where $`k`$ is the Boltzmann constant. We can express the entropy of the Universe as the product of he number of parts each particle has contributed to the Universe times he number of particles $`N_p`$: $$\frac{S_u}{k}t\frac{m_pc^2}{\mathrm{}}N_p$$ (11) In this sense the entropy of the Universe is the number of “tics” of each particle times the number of particles in the Universe. Entropy increases linearly with time and is very high today just because the Universe is old. In standard cosmology there is a problem with the entropy. Usually entropy is taken as the number of photons in the background blackbody radiation (at $`2.73^{}K`$) about $`10^{88}`$; no one knows where it comes from or why is so high. Now we know where the entropy comes from. Here we see that the entropy of the Universe is a linear function of time. That is why is so high today. The argument for the entropy of a system to be the maximum number of parts it consists of is here reproduced: the maximum number of parts in the Universe is the number of gravity quanta. Hence, any mass $`M`$ has an entropy $`S`$ as $$\frac{S}{k}\frac{M}{m_1}=M\frac{c^2t}{\mathrm{}}=\frac{M}{m_p}\frac{ct}{r_p}$$ (12) i.e. the number of particles of mass $`m_p`$ times the factor $`ct/r_p`$ $`10^{41}`$ today. ## IV PRIMORDIAL BLACK HOLE ENTROPY. We may define a primordial black hole of mass $`M_p`$ as the black hole that has the size of an elementary particle, $`r_p`$: $$\frac{GM_p}{c^2}=r_p=\frac{\mathrm{}}{m_pc}$$ (13) Hence the mass of a primordial black hole so defined is $$M_p=\frac{\mathrm{}c}{Gm_p}=\frac{\mathrm{}c}{Gm_p^2}m_p10^{38}m_p10^{79}m_1$$ (14) Should these objects be responsible for the dark matter in the Universe there would be about $`10^{41}`$ of them. The entropy of one of them is clearly about $`10^{79}`$, the number of gravity quanta emitted. The entropy of an elementary particle $`S_p`$, a primordial black hole $`S_{pbh}`$ and the Universe $`S_u`$ are then given by $$\frac{S_p}{k}10^{41}$$ $$\frac{S_{pbh}}{k}10^{79}$$ (15) $$\frac{S_u}{k}10^{120}$$ ## V THE BLACK HOLE ENTROPY. Black hole entropy as related to mass was theoretically established many years ago (Bekenstein, Hawking). Its value was given in terms of the area of the event horizon, the square of the gravitational radius of the black hole of mass $`M_{bh}`$, in units of Planck’s length. We know that from the physical parameters speed of light $`c`$, Planck’s constant $`\mathrm{}`$ , and the gravitational constant $`G`$, one can define Planck’s units as $$l_{}=\left(\frac{G\mathrm{}}{c^3}\right)^{\frac{1}{2}}=1.610^{33}cm$$ $$m_{}=\left(\frac{hc}{G}\right)^{\frac{1}{2}}=2.210^5gr$$ (16) $$t_{}=\left(\frac{G\mathrm{}}{c^5}\right)^{\frac{1}{2}}=5.410^{44}s$$ Then, the Bekenstein-Hawking formula for the entropy of a black hole can be expressed as $$\frac{S_{bh}}{k}\left(\frac{GM_{bh}}{c^2}\right)\frac{1}{l_{}^2}\frac{G}{\mathrm{}c}M_{bh}^2$$ (17) However, there are objections to this expression as given by Dunning-Davies and Lavenda(). In particular it is argued that this expression is not extensive. We see here that the Bekenstein-Hawking expression (17) is different from the one we propose in (13). If we apply (17) to the case of the primordial black hole we get an entropy of about $`10^{38}`$. In our case this number is $`10^{79}`$. The two concepts are very different indeed. Our expression (12) is extensive, as it should be for entropy. One of the main ideas of the Bekenstein-Hawking findings on the entropy associated to a black hole is the horizon area. This means that entropy can be linked to the area of the black hole (the horizon area) as measured in Planks units. A Planck cell is of fundamental meaning in the vacuum concept. Again, we are taking the entropy in dimensionless units, $`S/k`$. We can count the number of gravity quanta or the number of Plancks cells in the horizon area. In this way we can find an equivalent area $`A_e`$ in Plancks units (Bekenstein-Hawking idea in expression (17)) to our relation in (12). Then one has now the new area $`A_e`$ that in Plancks units expresses the value of the entropy: $$\frac{A_e}{l_{}^2}=\frac{M_{bh}}{m_p}\frac{ct}{r_p}M_{bh}\frac{c^2t}{\mathrm{}}$$ (18) and using (16) we get for the equivalent area $$A_e\frac{GM_{bh}}{c^2}ct$$ (19) The equivalent area has a size equal to the geometric mean of the gravitational radius of the black hole, $`GM_{bh}/c^2`$, and the size of the Universe ct. Hence, our entropy expression can be defined in this way as $$\frac{S_{bh}}{k}\frac{A_e}{l_{}^2}\frac{\frac{GM_{bh}}{c^2}ct}{l_{}^2}$$ (20) If we use Planck’s length as the unit of length, then in units of the Boltzmann constant the entropy of a black hole of mass $`M_{bh}`$ is the equivalent area formed by the product of its gravitational radius times the size of the Universe. In this way we keep the entropy as an extensive property. Also, the cosmological implication is here included through the factor $`ct`$, an effect similar to the Mach’s principle. This principle relates the cosmological property of the mass of the Universe with the local inertia properties of mass, i.e., cosmology with local objects. Relation (20) has a cosmological parameter $`ct`$, the size of the Universe, related to the local black hole. In this sense it is a Machean formula. ## VI BLACK HOLE RADIATION. The characteristic wavelength of radiation from a black hole is about its size, the gravitational radius $`GM_{bh}/c^2`$ . In the Bekenstein-Hawking expression (17) the horizon area, the square of the gravitational radius, is the emitting source. In our case we propose the equivalent area $`A_e`$ in (20) as the emitting source. In other words, we propose as the characteristic wavelength of radiation $`\lambda `$ from a black hole the size of the equivalent area: $$\lambda \sqrt{A_e}\sqrt{\frac{GM_{bh}}{c^2}ct}$$ (21) which is much higher than the Hawking proposal. In particular, for the primordial black holes one has $$\lambda \sqrt{\frac{GM_{pbh}}{c^2}ct}10^8cm$$ (22) to be compared with $`10^{12}`$ $`cm`$ deduced from the Hawking expression, which is in the gamma ray range. Then, the wavelength being so large in our case, the radiation of energy is much lower than in the Hawking case, about $`10^{40}`$ times lower. This means that black holes radiate much less than thought and therefore they last longer in the Universe. If primordial black holes were an important part of the dark matter they are still around us. The fact that no gamma ray radiation is observed to account for them, a high radiation predicted by the Hawking expression, is then no proof of their absence. In our case the prediction is that if any, there should be a radiation peak of a wavelength close to $`10^8cm`$. Interferometric methods would be adequate to detect it, if there are enough primordial black holes in the Universe. Then, for radiation from a black hole the antenna size is given by (21), a number much higher than its gravitational radius, and linked to the Universe in a style similar to the Mach’s principle. ## VII CONCLUSION. A parallel gravitational uncertainty principle can be formulated using a gravitational Planck’s constant. Then, gravitational phenomena can be treated as in the quantum mechanical case probably with very similar formulation. The scale change is of the order of $`10^{40}`$ times smaller for the gravity case today. The quantum of gravitational action has a mass about $`10^{65}`$ grams. Entropy can be defined as an extensive property in terms of the equivalent mass of the system using the gravity quanta. Hence, in any system a total entropy concept for gravitational properties can be formulated by adding all equivalent mass of its different energy components, in units of the quantum of gravity. A corresponding black hole entropy can then be defined. Applied to the primordial black holes one concludes that they radiate energy at a much lower rate than in the Hawking model, $`10^{40}`$ times smaller. And therefore they are still a candidate for the dark matter in the Universe. The second law of thermodynamics for gravity, according to our view, is summarized as follows: entropy increases linearly with time, the total amount of entropy being the emitted equivalent number of gravity quanta through the age of the Universe. Speculating about the future of this Universe, taking physics beyond the theoretical, the time will stop when no more gravity quanta are emitted, if the entropy arrives at a maximum and remains constant.
warning/0002/math-ph0002054.html
ar5iv
text
# The Reeh-Schlieder Property for Quantum Fields on Stationary Spacetimes ## 1 Introduction For the analysis of quantum field theory in curved spacetime it has turned out that the framework of algebraic quantum field theory (see ) is most suitable for analyzing the problems connected with the non-uniqueness of the quantization of linear fields (see e.g. and the references therein). The problem reduces to finding appropriate representations of the field algebra which can straightforwardly be constructed on manifolds (see ). This is the same as specifying vacuum-like states over the algebra. On stationary spacetimes it is possible to distinguish states as ground- or KMS-states with respect to the canonical time translations. For free fields satisfying certain wave-equations it was recently shown in that such passive quasifree states satisfy the microlocal spectrum condition (see ) which is believed to be a substitute for the usual spectrum condition in Minkowski spacetime. For the case of the scalar field on a 4-dimensional globally hyperbolic spacetime it was shown in that the microlocal spectrum condition is equivalent to the requirement that the 2-point function is of Hadamard form (see e.g. ), which allows for a renormalization of the stress energy tensor (). It therefore seems reasonable to consider ground- or KMS-states for free quantum fields on stationary spacetimes as good substitutes for the vacuum in flat spacetime. In the Minkowski space theory the vacuum vector turns out to be cyclic for field algebras associated to nonvoid open regions (). This Reeh-Schlieder property of the vacuum vector holds also for thermal quantum field theories (). By the lack of symmetry in general spacetimes it is not clear whether physically reasonable vacuum states have this property as well. It was shown in that the quasifree ground-state of the massive scalar field on an ultrastatic spacetime is of this kind. The proof uses an anti-locality property of the square root of the Laplace operator. Using similar methods it was possible to obtain a Reeh-Schlieder-type property for solutions to the Dirac equation on an ultrastatic spacetime with compact Cauchy surface in , and the Reeh-Schlieder property for ground- and KMS-states of the free Dirac field on a static globally hyperbolic 4-dimensional spacetime in . We will show in this paper that the Reeh-Schlieder property of ground- and KMS- states holds for a large class of free fields on stationary spacetimes. We introduce the notions of (classical) linear fermionic and bosonic field theories on a spacetime and show that these notions lead via canonical quantization to quantum field theories on this spacetime. We note that our approach does not use an initial data formulation. In the last section we show how the most common free fields fit into this framework. Our main theorem states that as soon as the classical field fulfills a certain hyperbolic partial differential equation, a state over the field algebra of the quantized theory, which is a (quasifree in the bosonic case) ground- or KMS-state with respect to the group of time translations, has the Reeh-Schlieder property. In order to show this we combine the Gelfand-Maurin theorem on generalized eigenvectors with classical results on the strong unique continuation property of solutions of certain second order elliptic differential equations. As a result we obtain a curved spacetime analog to the timelike tube theorem of Borchers (). Using standard arguments this yields the Reeh-Schlieder property. The class of fields which fulfill our assumptions contains the Dirac field, the Proca field and the scalar field on arbitrary connected 4-dimensional globally hyperbolic stationary Lorentzian manifolds. We note that as a consequence the Hartle-Hawking state for the Klein-Gordon field on the external Schwarzschild spacetime has the Reeh-Schlieder property (see ). This is a new result which so far has not been obtained by the methods previously employed. The Reeh-Schlieder property serves as the starting point for the use of the Tomita-Takesaki-theory within quantum field theory. The application of this theory to the analysis of quantum field theory in Minkowski spacetime was very fruitful (see e.g. ) and there might be an impact as well on curved space quantum physics (see ,). ## 2 Classical linear field theories In the following and throughout the text a smooth manifold will always be Hausdorff and separable as a topological space. If we are given a smooth vector bundle $`E`$ over a smooth manifold $`M`$ one can endow the space of smooth sections $`\mathrm{\Gamma }(E)`$ and the space of compactly supported smooth sections $`\mathrm{\Gamma }_0(E)`$ with locally convex topologies (see ) in a similar way as for $`C^{\mathrm{}}(^n)`$ and $`C_0^{\mathrm{}}(^n)`$. These locally convex vector spaces turn out to be nuclear (see , or for properties of nuclear spaces). We denote the topological dual of the space $`\mathrm{\Gamma }_0(E)`$ by $`𝒟^{}(M,E^{})`$, calling it the space of distributions with values in the dual bundle $`E^{}`$. For some open $`𝒪M`$ the subspaces $`\mathrm{\Gamma }_0(E,𝒪)`$ and $`\mathrm{\Gamma }(E,𝒪)`$ of sections supported in $`𝒪`$ are closed. We call a smooth vector bundle $`E`$ a $`G`$-vector bundle for some Lie-group $`G`$ if there is a smooth action of $`G`$ on the base space $`M`$ and on $`E`$, such that the bundle projection $`EM`$ is equivariant and the left action $`q:E_xE_{qx}`$ is linear for all $`qG`$ and $`xM`$. In this case we have a canonical action of $`G`$ on $`\mathrm{\Gamma }_0(E)`$ and $`\mathrm{\Gamma }(E)`$, for which we use the notation $`G\times \mathrm{\Gamma }(E)\mathrm{\Gamma }(E),(q,f)qf`$. These representations of $`G`$ are continuous in the corresponding locally convex topologies. A smooth n-dimensional Lorentzian manifold $`(M,g)`$ is a smooth n-dimensional manifold with smooth metric $`g`$ of constant signature $`(1,1,\mathrm{},1)`$ and $`n2`$ (see e.g. ). For any open set $`𝒪`$ we denote the open causal complement of $`𝒪`$, i.e. the interior of the set of points in $`M`$ which cannot be joined to a point in $`𝒪`$ by a causal curve, by $`𝒪^{}`$. If $`E`$ is a smooth vector bundle over a Lorentzian manifold $`(M,g)`$ we say a second order differential operator $`P:\mathrm{\Gamma }(E)\mathrm{\Gamma }(E)`$ has metric principal part, if in local coordinates the principal part is of the form $`\mathrm{𝟏}g^{ik}_i_k`$. ### 2.1 The fermionic case Let $`(M,g)`$ be a smooth n-dimensional Lorentzian manifold. We denote the identity component of the group of isometries of $`M`$ by $`G`$ and its universal covering group by $`\stackrel{~}{G}`$. Note that $`\stackrel{~}{G}`$ is a Lie-group which acts smoothly on $`M`$. ###### Definition 2.1. A linear fermionic field theory on $`M`$ is a 5-tuple $`(,𝒦,E,\eta ,\rho )`$, where $``$ is a complex Hilbert space with conjugation $`𝒦`$, $`E`$ a smooth complex $`\stackrel{~}{G}`$-vector bundle, $`\eta `$ a linear map $`\eta :\mathrm{\Gamma }_0(E)`$ with dense range, and $`\rho `$ a unitary representation of $`\stackrel{~}{G}`$ on $``$ commuting with $`𝒦`$, such that the following conditions are satisfied: 1. Covariance: $`\eta (qf)=\rho (q)(\eta (f)),f\mathrm{\Gamma }_0(E),q\stackrel{~}{G}`$. 2. Causality: $`𝒪_1𝒪_2^{}`$ implies $`(𝒪_1)(𝒪_2)`$, where $`(𝒪)`$ denotes the closure of the image of $`\mathrm{\Gamma }_0(E,𝒪)`$ under $`\eta `$. 3. Continuity: $`\eta `$ is weakly continuous, i.e. $`v,\eta ()`$ is in $`𝒟^{}(M,E^{})`$ for all $`v`$. 4. $`𝒦`$ is local, i.e. $`𝒦((𝒪))=(𝒪)`$. ###### Proposition 2.2. For any linear fermionic field theory $`(,𝒦,E,\eta ,\rho )`$ on $`M`$ the map $`\eta `$ is norm continuous. Moreover, the representation $`\rho `$ is strongly continuous. ###### Proof. The sesquilinear form $`B(f,g):=\eta (f),\eta (g)`$ on $`\mathrm{\Gamma }_0(E)\times \mathrm{\Gamma }_0(E)`$ is separately continuous, and as a consequence of the nuclearity of $`\mathrm{\Gamma }_0(E)`$ it is jointly continuous. This gives the norm continuity of $`\eta `$. Since the action of $`\stackrel{~}{G}`$ on $`E`$ is smooth, the action of $`\stackrel{~}{G}`$ on $`\mathrm{\Gamma }_0(E)`$ is continuous. As a consequence we obtain the strong continuity of $`\rho `$. ∎ ### 2.2 The bosonic case Let $`(M,g)`$ be again a smooth n-dimensional Lorentzian manifold and $`G`$ be the identity component of the group of isometries of $`M`$. ###### Definition 2.3. A linear bosonic field theory on M is a 5-tuple $`(𝒲,\sigma ,E,\eta ,\rho )`$, where $`𝒲`$ is a symplectic vector space with symplectic form $`\sigma `$, $`E`$ a smooth real $`G`$-vector bundle, $`\eta `$ a surjective linear map $`\eta :\mathrm{\Gamma }_0(E)𝒲`$, and $`\rho `$ a representation of $`G`$ on $`𝒲`$ by symplectomorphisms, such that the following conditions are satisfied: 1. Covariance: $`\eta (qf)=\rho (q)(\eta (f)),f\mathrm{\Gamma }_0(E),qG`$. 2. Causality: $`𝒪_1𝒪_2^{}`$ implies $`𝒲(𝒪_1)𝒲(𝒪_2)`$, where $`𝒲(𝒪)`$ denotes the image of $`\mathrm{\Gamma }_0(E,𝒪)`$ under $`\eta `$. 3. Continuity: For all $`v𝒲`$ the map $`\mathrm{\Gamma }_0(E),f\sigma (v,\eta (f))`$ is continuous, i.e. defines a distribution in $`𝒟^{}(M,E^{})`$. ## 3 Canonical Quantization In this section $`(M,g)`$ will be a smooth n-dimensional Lorentzian manifold. We will work in the framework of algebraic quantum field theory (). A quantum field theory on $`M`$ will be defined by a net of local field algebras, i.e. a map $`𝒪(𝒪)`$ from the relatively compact open subsets of $`M`$ to the set of closed $``$-subalgebras of a $`C^{}`$\- or $`W^{}`$-algebra $``$ which is isotone, i.e. $`(𝒪_1)(𝒪_2)`$ whenever $`𝒪_1𝒪_2`$. ### 3.1 Quantization of a linear fermionic field theory We show how a linear fermionic field theory gives rise to a quantum field theory on $`M`$ given by a net of local field algebras. Given a linear fermionic field theory $`(,𝒦,E,\eta ,\rho )`$, the field algebra $``$ is the (self-dual) CAR-algebra $`\text{CAR}(,𝒦)`$ (see ). This is the $`C^{}`$-algebra with unit generated by symbols $`B(v)`$ with $`v`$ and the relations $$vB(v)\text{is complex linear,}$$ (1) $$B(v)^{}=B(𝒦v),$$ (2) $$\{B(v_1),B(v_2)\}=B(v_1)B(v_2)+B(v_2)B(v_1)=𝒦v_1,v_2.$$ (3) $``$ has a natural $`_2`$-grading. The even/odd parts are spanned by those products $`B(v_1)\mathrm{}B(v_k)`$ with an even/odd number of generators. For each relatively compact subset $`𝒪M`$ we define the local algebra $`(𝒪)`$ to be the closed unital $``$-subalgebra generated by the symbols $`B(\eta (f))`$ with $`f\mathrm{\Gamma }_0(E,𝒪)`$. The representation $`\rho `$ of the group $`\stackrel{~}{G}`$ on $``$ gives rise to a representation $`\tau `$ of $`\stackrel{~}{G}`$ by strongly continuous Bogoliubov automorphisms of the algebra (see ). It is not difficult to check the following properties of the net $`𝒪(𝒪)`$: 1. Isotony: $`𝒪_1𝒪_2`$ implies $`(𝒪_1)(𝒪_2)`$. 2. Causality: if $`𝒪_1𝒪_2^{}`$, then $`\{(𝒪_1),(𝒪_2)\}=\{0\}`$, where $`\{,\}`$ denotes the graded commutator. 3. Covariance: $`\tau (q)(𝒪)=(q𝒪)q\stackrel{~}{G}`$. Moreover, $``$ is the quasilocal algebra of the net $`𝒪(𝒪)`$ (see , proposition 5.2.6). Hence, this defines a reasonable quantum field theory on the manifold $`M`$. Note that $``$ is not the algebra of observables. The algebra of observables $`𝒜`$ should be a \*-subalgebra of $`_{\text{even}}`$ consisting of elements $`a`$ for which $`\tau _{g_1}a=\tau _{g_2}a`$ whenever $`p(g_1)=p(g_2)`$, where $`p:\stackrel{~}{G}G`$ is the covering map. Usually $`𝒜`$ is the $``$-subalgebra of elements which are invariant under the action of a gauge group. ### 3.2 Quantization of a linear bosonic field theory Each linear bosonic field theory $`(𝒲,\sigma ,E,\eta ,\rho )`$ gives rise to a quantum field theory on $`M`$. The field algebra $``$ is defined to be the CCR-algebra $`\text{CCR}(𝒲,\sigma )`$ (see ). This is the $`C^{}`$-algebra generated by symbols $`W(v)`$ with $`v𝒲`$ and the relations $$W(v)=W(v)^{},$$ (4) $$W(v_1)W(v_2)=e^{i\sigma (v_1,v_2)/2}W(v_1+v_2).$$ (5) We define for each relatively compact open subset $`𝒪M`$ the local field algebra $`(𝒪)`$ to be the closed $``$-subalgebra generated by the symbols $`W(v)`$ with $`v𝒲(𝒪)`$. The representation $`\rho `$ of $`G`$ gives rise to a representation $`\tau `$ of $`G`$ by Bogoliubov automorphisms of the algebra $``$ (see e.g.), and the net $`𝒪(𝒪)`$ has the following properties: 1. Isotony: $`𝒪_1𝒪_2`$ implies $`(𝒪_1)(𝒪_2)`$. 2. Causality: if $`𝒪_1𝒪_2^{}`$, then $`[(𝒪_1),(𝒪_2)]=\{0\}`$. 3. Covariance: $`\tau (q)(𝒪)=(q𝒪)qG`$. Moreover, $``$ is the quasilocal algebra of the net $`𝒪(𝒪)`$ (see , proposition 5.2.10). Hence, this defines a reasonable quantum field theory on the manifold $`M`$. Unlike the fermionic case the representation $`\tau `$ fails to be strongly continuous whenever it is nontrivial. We therefore need to pass to certain representations of the field algebra to obtain a net of von Neumann algebras on which $`\tau `$ extends to a $`\sigma `$-weakly-continuous representation. In order to avoid complications we specialize to the so-called quasifree states. Assume that we are given a scalar product $`\mu `$ on $`𝒲`$ which dominates $`\sigma `$, i.e. satisfies the estimate $$|\sigma (v_1,v_2)|^24\mu (v_1,v_1)\mu (v_2,v_2)v_1,v_2𝒲.$$ (6) In this case the linear functional $`\omega _\mu :`$, defined by $$\omega _\mu (W(v)):=e^{\mu (v,v)/2}v𝒲,$$ (7) is a state (see ). The states over $``$ which can be realized in this way are called quasifree states. A quasifree state $`\omega _\mu `$ gives rise to a one particle structure (Proposition 3.1 in ), that is a map $`K_\mu :𝒲H_\mu `$ to some complex Hilbert space $`H_\mu `$, such that 1. the complexified range of $`K_\mu `$, (i.e. $`K_\mu 𝒲+iK_\mu 𝒲`$), is dense in $`H_\mu `$, 2. $`K_\mu v_1,K_\mu v_2=\mu (v_1,v_2)+\frac{i}{2}\sigma (v_1,v_2)`$. This structure is unique up to equivalence. A one particle structure $`(K_\mu ,H_\mu )`$ for a quasifree state allows one to construct the GNS-triple $`(\pi _{\omega _\mu },_{\omega _\mu },\mathrm{\Omega }_{\omega _\mu })`$ explicitly (see ). Namely, one takes $`_{\omega _\mu }`$ to be the bosonic Fock space over $`H_\mu `$ with Fock vacuum $`\mathrm{\Omega }_{\omega _\mu }`$, and defines $`\pi _{\omega _\mu }(W(v))=\text{exp}((\overline{\widehat{a}^{}(Kv)\widehat{a}(Kv)}))`$, where $`\widehat{a}^{}()`$ and $`\widehat{a}()`$ are the usual creation and annihilation operators. One clearly has the following ###### Proposition 3.1. Let $`\omega _\mu `$ be a quasifree state over the CCR-algebra $`=\text{CCR}(𝒲,\sigma )`$ and let $`(\pi _{\omega _\mu },_{\omega _\mu },\mathrm{\Omega }_{\omega _\mu })`$ be its GNS-triple. If $`V𝒲`$ is a subspace which is dense in $`𝒲`$ in the topology defined by $`\mu `$, then the $``$-algebra generated by the set $$\{\pi _{\omega _\mu }(W(v)),vV\}\pi _{\omega _\mu }()$$ is strongly dense in the von Neumann algebra $`\pi _{\omega _\mu }()^{\prime \prime }`$. ###### Definition 3.2. Let $``$ be a field algebra constructed from a linear bosonic field theory $`(𝒲,\sigma ,E,\eta ,\rho )`$ on $`M`$. We call a quasifree state $`\omega _\mu `$ over $``$ continuous if the map $$\mathrm{\Gamma }_0(E)\times \mathrm{\Gamma }_0(E),(f_1,f_2)\mu (\eta (f_1),\eta (f_2))$$ is continuous and hence defines a distribution in $`𝒟^{}(M\times M,E^{}E^{})`$, where $`E^{}E^{}`$ is the direct product of the bundle $`E^{}`$ with itself over the base space $`M\times M`$. This is clearly a necessary and sufficient condition for Wightman 2-point function $`w_2(,):=K\eta (),K\eta ()`$ to be a distribution. For the class of continuous quasifree states we can circumvent the problems connected with the non-continuity of the representation $`\tau `$ of $`G`$ on $``$. Since the action of $`G`$ is continuous on $`\mathrm{\Gamma }_0(E)`$ and $`\rho `$ leaves $`\sigma `$ and $`\mu `$ invariant, there exists a unique strongly continuous representation $`\stackrel{~}{U}`$ of $`G`$ on the one particle Hilbert space $`H_\mu `$, such that $`K_\mu \rho (q)=\stackrel{~}{U}(q)K_\mu `$ for all $`qG`$. Second quantization gives a strongly continuous unitary representation $`U`$ of $`G`$ on $`_{\omega _\mu }`$, such that $`\pi _{\omega _\mu }(\rho (q)a)=U(q)\pi _{\omega _\mu }(a)U^1(q)`$ for all $`qG`$ and $`a`$. Hence, one gets the following proposition: ###### Proposition 3.3. Let $`\omega _\mu `$ be a continuous $`G`$-invariant quasifree state over the field algebra $``$ constructed from a linear bosonic field theory $`(𝒲,\sigma ,E,\eta ,\rho )`$ on $`M`$. Let $`(\pi _{\omega _\mu },_{\omega _\mu },\mathrm{\Omega }_{\omega _\mu })`$ be its GNS-triple and $`U`$ be the unitary representation of $`G`$ on $`_{\omega _\mu }`$ induced by $`\tau `$. Then $`U`$ is strongly continuous and hence $`\tau `$ can be continued to a $`\sigma `$-weakly-continuous representation $`\widehat{\tau }`$ of $`G`$ by automorphisms of the von Neumann algebra $`\widehat{}:=\pi _{\omega _\mu }()^{\prime \prime }`$. Therefore, given a continuous quasifree state $`\omega _\mu `$ we can construct the net of von Neumann algebras $`𝒪\widehat{}(𝒪):=\pi _{\omega _\mu }((𝒪))^{\prime \prime }`$. This assignment is isotone, causal and covariant, and the representation $`\widehat{\tau }`$ of $`G`$ is $`\sigma `$-weakly-continuous. It gives rise to a quantum field theory on $`M`$ with reasonable physical properties. ## 4 The Reeh-Schlieder property for quantized linear fields ### 4.1 Stationary spacetimes Let $`(M,g)`$ be an n-dimensional time-oriented Lorentzian manifold which admits a one parameter group $`h_t`$ of isometries, smooth in $`t`$, with timelike orbits giving rise to a timelike Killing vector field $`\xi `$. Such a manifold is called stationary. For later considerations we need a special class of charts. ###### Lemma 4.1. For each point $`pM`$ there exists an open neighbourhood $`𝒪`$ and a chart $`\varphi :𝒪^n`$ with coordinates $`(x_0,\mathrm{},x_{n1})`$, such that in local coordinates 1. $`\xi =\frac{}{x_0}`$, 2. the $`(n1)\times (n1)`$ matrix $`g^{\alpha \beta }(x),\alpha ,\beta =1,\mathrm{},n1`$ is positive for all $`x\varphi (𝒪)`$. ###### Proof. One can always choose a neighbourhood $`𝒪_1`$ of $`p`$ and a chart $`\varphi :𝒪_1^n`$ with coordinates $`(x_0,\mathrm{},x_{n1})`$, such that $`\xi =\frac{}{x_0}`$ and the dual metric tensor is diagonal in the point $`p`$, i.e. $$g^{ik}(\varphi (p))=\text{diag}_n(g(\xi ,\xi ),1,\mathrm{},1).$$ (8) $`g^{\alpha \beta }(\varphi (p)),\alpha ,\beta =1,\mathrm{},n1`$ is then a positive matrix. Since the matrix-valued function $`g^{\alpha \beta }`$ is continuous, there exists a neighbourhood $`\varphi (𝒪)`$ of $`\varphi (p)`$, on which it is positive. ∎ ### 4.2 Free quantum fields on stationary spacetimes Since $`h_t`$ is a group of isometries it defines a group homomorphism $`G`$ which lifts uniquely to a group homomorphism $`\stackrel{~}{G}`$. Hence, in case we have a net of field algebra $`𝒪(𝒪)`$ constructed from a linear fermionic or bosonic field theory we canonically get a one parameter group of automorphisms $`\tau _t`$ which acts covariantly, i.e. $`\tau _t(𝒪)=(h_t𝒪)`$. We call this group the group of canonical time translations induced by $`h_t`$. In the interesting cases one can realize $``$ on a Hilbert space, such that $`\tau _t`$ is $`\sigma `$-weakly-continuous and hence extends to an automorphism group $`\widehat{\tau }_t`$ of the von Neumann algebra $`^{\prime \prime }`$ (see e.g. proposition 3.3). It is then possible to distinguish vacuum states over the field algebra $``$ as ground- or KMS-states with respect to the group of time translations $`\tau _t`$. ###### Definition 4.2. Let $`𝒜`$ be a $`W^{}`$-algebra and $`\alpha _t`$ be a $`\sigma `$-weakly-continuous one-parameter group of $``$-automorphisms of $`𝒜`$. An $`\alpha _t`$ invariant normal state $`\omega `$ is called ground-state with respect to $`\alpha _t`$ if the generator of the corresponding strongly continuous unitary group $`U(t)`$ on the GNS-Hilbert space is positive. ###### Definition 4.3. Let $`𝒜`$ and $`\alpha _t`$ be as above. A normal state $`\omega `$ is called KMS-state with inverse temperature $`\beta >0`$ with respect to $`\alpha _t`$ if for any pair $`A,B𝒜`$ there exists a complex function $`F_{A,B}`$ which is analytic in the strip $$𝒟_\beta :=\{z;0<Im(z)<\beta \}$$ and bounded and continuous on $`\overline{𝒟_\beta }`$, such that $$F_{A,B}(t)=\omega (A\alpha _t(B)),$$ $$F_{A,B}(t+i\beta )=\omega (\alpha _t(B)A).$$ ###### Definition 4.4. Let $`(M,g,h_t)`$ be a connected stationary Lorentzian manifold. Let $`\{(𝒪)\}`$ be the net of local field algebras constructed from a linear fermionic field theory $`(,𝒦,E,\eta ,\rho )`$ or from a linear bosonic field theory $`(𝒲,\sigma ,E,\eta ,\rho )`$. Denote the group of canonical time translations by $`\tau _t`$. Let $`\omega `$ be a $`\tau _t`$-invariant state over the quasilocal algebra $``$, which we assume to be quasifree and continuous in the bosonic case, and denote by $`(\pi _\omega ,_\omega ,\mathrm{\Omega }_\omega )`$ the corresponding GNS-triple. We say that $`\omega `$ is a ground or KMS state, if the unique normal extension of $`\omega `$ over $`\pi _\omega ()^{\prime \prime }`$ is a ground or KMS state with respect to the unique $`\sigma `$-weakly-continuous extension of the group of time translations $`\tau _t`$. In the fermionic case the existence of ground states is always guaranteed (). Moreover, there exists a unique quasifree KMS-state with inverse temperature $`\beta >0`$ (see ). In the bosonic case the construction of continuous quasifree ground- and KMS-states seems problematic in the general case. See e.g. for conditions that allow the construction of a continuous quasifree ground state for the Klein-Gordon quantum field on a 4-dimensional stationary spacetime. ### 4.3 A density theorem and a tube theorem ###### Theorem 4.5. Let $`(M,g,h_t)`$ be a connected stationary Lorentzian manifold and $`E`$ a smooth complex $`h_t`$-vector bundle. Let $`H`$ be a complex Hilbert space and $`\rho _t`$ a strongly continuous unitary one-parameter group on $`H`$. Assume that we have a linear strongly continuous map $`\widehat{\eta }:\mathrm{\Gamma }_0(E)H`$ with dense range which is covariant, i.e. $`\widehat{\eta }(h_tf)=\rho _t\widehat{\eta }(f)`$ for all $`t,f\mathrm{\Gamma }_0(E)`$. Assume furthermore that $`\widehat{\eta }P=0`$ for some second order differential operator $`P`$ with metric principal part. Then $`\widehat{\eta }(\mathrm{\Gamma }_0(E,h_{}𝒪))`$ is dense in $`H`$ for each nonvoid open set $`𝒪M`$. ###### Proof. We introduce the following notations: $$V:=\widehat{\eta }(\mathrm{\Gamma }_0(E,h_{}𝒪)^{},$$ (9) $$p_V\mathrm{}\text{orthogonal projection onto}V,$$ (10) $$\mathrm{\Phi }:=\text{Ran}(p_V\widehat{\eta }).$$ (11) $`V`$ is clearly a $`\rho _t`$-invariant subspace and $`\mathrm{\Phi }`$ is dense in $`V`$. Identifying $`\mathrm{\Phi }`$ with $`\mathrm{\Gamma }_0(E)/\text{ker}(p_V\widehat{\eta })`$ we can endow $`\mathrm{\Phi }`$ with the locally convex quotient topology. Since $`\mathrm{\Gamma }_0(E)`$ is nuclear and $`p_V\widehat{\eta }`$ is continuous, $`\mathrm{\Phi }`$ is a nuclear space (see ), and clearly the inclusion map $`\mathrm{\Phi }V`$ is continuous. We denote the dual of $`\mathrm{\Phi }`$ by $`\mathrm{\Phi }^{}`$. It follows that $$\mathrm{\Phi }V\mathrm{\Phi }^{}$$ is a Gelfand triple. We denote the selfadjoint generator of the group $`\rho _t|_V`$ by $`A`$. Clearly, $`\mathrm{\Phi }𝒟(A)`$ and $`A`$ restricts to a continuous operator $`\mathrm{\Phi }\mathrm{\Phi }`$. Moreover $`\mathrm{\Phi }`$ is invariant under the action of $`\rho _t|_V`$. As a consequence $`A`$ is essentially selfadjoint on $`\mathrm{\Phi }`$. Hence, there exists a complete set of generalized eigenvectors (see ) for $`A`$, i.e. a family $`v_\lambda \mathrm{\Phi }^{}`$ indexed by a subset $`I`$, such that $$v_\lambda (Af)=\lambda v_\lambda (f),f\mathrm{\Phi },$$ (12) $$v_\lambda (f)=0\text{for all }\lambda If=0.$$ (13) By continuity $$\psi _\lambda ():=v_\lambda (p_V\widehat{\eta }())$$ (14) defines for each $`\lambda I`$ a distribution in $`𝒟^{}(M,E^{})`$, such that $$\psi _\lambda (\mathrm{\Gamma }_0(E,h_{}𝒪))=\{0\},$$ (15) $$\psi _\lambda (_\xi f)=i\lambda \psi _\lambda (f),$$ (16) $$\psi _\lambda (P)=0,$$ (17) where $`\xi `$ is the timelike Killing vector field induced by $`h_t`$ and $`_\xi `$ the Lie derivative on $`\mathrm{\Gamma }_0(E)`$ defined by $`_\xi f=lim_{t0}\frac{h_tff}{t}`$. For each point $`pM`$ there exists an open contractible neighbourhood $`𝒰`$ and a chart mapping $`𝒰`$ to $`^n`$ which we can choose to be of the form constructed in lemma 4.1. The restriction of $`E`$ to $`𝒰`$ is trivial and we can identify $`\mathrm{\Gamma }_0(E,𝒰)`$ with $`C^{\mathrm{}}(𝒰)^N`$ in such a way that $`_\xi f=\frac{}{x_0}f`$ for both functions and sections. We consider the distribution $`\psi _\lambda `$ in such a chart. We have $`P^{}\psi _\lambda =0`$, where $`P^{}`$ is the adjoint operator. Moreover, equation (16) reads $`\frac{}{x_0}\psi _\lambda =i\lambda \psi _\lambda `$. Note that the principal part of $`P^{}`$ has the form $$g^{00}\frac{^2}{x_0^2}+g^{0\alpha }\frac{^2}{x_0x_\alpha }+g^{\alpha \beta }\frac{^2}{x_\alpha x_\beta }$$ (18) $$\alpha ,\beta =1,2,\mathrm{},n1.$$ Replacing the $`x_0`$-derivatives by $`i\lambda `$ and adding the term $`\frac{^2}{x_0^2}\lambda ^2`$ we obtain an elliptic second order differential operator $`P_e`$ with $`P_e\psi _\lambda =0`$. Hence, $`\psi _\lambda `$ is smooth (see e.g.) and since $`P_e`$ has scalar principal part the classical result of Arozajn (see especially Remark 3)<sup>1</sup><sup>1</sup>1 For a detailed treatment see section 17.1 of and the references therein implies that $`\psi _\lambda =0`$ in each such chart in which $`\psi _\lambda `$ vanishes in an open nonvoid set, in particular in each such chart intersecting with $`h_{}𝒪`$. Since $`M`$ is connected this implies $`\psi _\lambda =0`$ on $`M`$. The set of generalized eigenvectors $`v_\lambda `$ was complete and therefore $`V`$ equals $`\{0\}`$ and the theorem is proved. ∎ As a consequence one gets a result similar to the timelike tube theorem () in Minkowski spacetime. ###### Theorem 4.6. Let $`(M,g,h_t)`$ be a connected stationary Lorentzian manifold. Let $`\{(𝒪)\}`$ be a net of local field algebras constructed from a linear fermionic field theory $`(,𝒦,E,\eta ,\rho )`$ or from a linear bosonic field theory $`(𝒲,\sigma ,E,\eta ,\rho )`$. Assume that $`\eta P=0`$ for some second order differential operator $`P`$ with metric principal part. One has 1. The fermionic case. The $``$-subalgebra of $``$ generated by the subset $`_t(h_t𝒪)`$ is norm dense in $``$ for each nonvoid relatively compact open set $`𝒪M`$. 2. The bosonic case Let $`\omega _\mu `$ be a quasifree and continuous state over the quasilocal algebra $``$ . Assume that $`\omega _\mu `$ is invariant under the automorphism group $`\tau _t`$ induced by the Killing flow $`h_t`$. Denote its GNS-triple by $`(\pi _{\omega _\mu },_{\omega _\mu },\mathrm{\Omega }_{\omega _\mu })`$. The $``$-subalgebra of $`\pi _{\omega _\mu }()`$ generated by the subset $`_t\pi _{\omega _\mu }((h_t𝒪))`$ is strongly dense in $`\pi _{\omega _\mu }()`$ for each nonvoid relatively compact open set $`𝒪M`$. ###### Proof. We start with the fermionic case. Let $`_1`$ be the subspace of $``$ generated by the set $$\{\eta (h_tf);f\mathrm{\Gamma }_0(E,𝒪),t\},$$ and let $`_1`$ be the unital $``$-subalgebra of $``$ generated by the set $`\{B(v),v_1\}`$. Clearly, $`_1`$ is equal to the $``$-subalgebra of $``$ generated by $`_t(h_t𝒪)`$. The group of time translations $`\rho _t`$ is a strongly continuous one parameter group on $``$, and with $`\widehat{\eta }=\eta `$ we can apply theorem 4.5 above. It follows that $`_1`$ is dense in $``$ and hence $`_1`$ is norm dense in $``$. In the bosonic case let $`𝒲_{}`$ be the complexification of $`𝒲`$ and take the complexification of $`\mu `$ as a scalar product on $`𝒲_{}`$. We complete this space and obtain a Hilbert space $`H`$. We complexify the real vector bundle $`E`$ and obtain the complex vector bundle $`E_{}EE`$ with a canonical action of $`h_t`$. We can extend the map $`\eta `$ to a map $`\eta _{}`$ which maps from the section of $`E_{}`$ to $`H`$. By construction $`\eta _{}`$ has dense range. Since $`\mu `$ is invariant under the action of the time translations $`\rho _t`$ we get a strongly continuous unitary action of the group of time translations on $`H`$ such that $`\eta _{}`$ is equivariant. Let $`H_1`$ be the complex subspace of $`H`$ generated by the set $$\{\eta _{}(h_tf)H;f\mathrm{\Gamma }_0(E_{},𝒪),t\}.$$ We see that all the assumptions for theorem 4.5 are fulfilled and hence $`H_1`$ is dense in $`H`$. Using proposition 3.1 one therefore concludes that the $``$-subalgebra of $`\pi _{\omega _\mu }()`$ generated by the set $`\{W(\eta (h_tf)),f\mathrm{\Gamma }_0(E,𝒪),t\}`$ is strongly dense in $`\pi _{\omega _\mu }()`$. ∎ ### 4.4 The Reeh-Schlieder property for ground- and KMS-states ###### Definition 4.7. Let $`\{(𝒪)\}`$ be a net of local field algebras indexed by the relatively compact open subsets of a manifold $`M`$. Let $`\omega `$ be a state over the quasilocal field algebra $``$ and $`(\pi _\omega ,_\omega ,\mathrm{\Omega }_\omega )`$ its GNS-triple. We say that $`\omega `$ has the Reeh-Schlieder property if $`\mathrm{\Omega }_\omega `$ is cyclic for the von Neumann algebra $`\pi _\omega ((𝒪))^{\prime \prime }`$ for each nonvoid relatively compact open set $`𝒪M`$. Our main theorem is: ###### Theorem 4.8. Let $`(M,g,h_t)`$ be a connected stationary Lorentzian manifold. Let $`\{(𝒪)\}`$ be the net of local field algebras constructed from a linear fermionic field theory $`(,𝒦,E,\eta ,\rho )`$ or from a linear bosonic field theory $`(𝒲,\sigma ,E,\eta ,\rho )`$. Assume that $`\eta P=0`$ for some second order differential operator $`P`$ with metric principal part. Let $`\omega `$ be a state over the quasilocal algebra $``$ which we assume to be quasifree and continuous in the bosonic case. If $`\omega `$ is a ground- or KMS-state with respect to the automorphism group $`\tau _t`$ induced by the Killing flow $`h_t`$, then $`\omega `$ has the Reeh-Schlieder property. We postpone the proof for a moment. Let $`𝒪M`$ be a nonvoid relatively compact open set and $`(𝒪)(𝒪)`$ the $``$-subalgebra consisting of those elements $`a(𝒪)`$ for which there exists a neighbourhood $`I`$ of $`0`$, such that $`\tau _I(a)(𝒪)`$. Let $`𝒪_1M`$ be another nonvoid open set such that $`\overline{𝒪_1}𝒪`$. We clearly have the inclusions $$(𝒪_1)(𝒪)(𝒪).$$ (19) We denote the strongly continuous unitary group, implementing $`\tau _t`$ on the GNS-Hilbert space, by $`\stackrel{~}{U}(t)`$. One has ###### Lemma 4.9. Let $`:=\overline{\pi _\omega ((𝒪))\mathrm{\Omega }_\omega }`$. Then $``$ is invariant under the action of $`\stackrel{~}{U}(t)`$, i.e. $`\stackrel{~}{U}()`$. ###### Proof. Let $`\psi ^{}`$. For each $`a\pi _\omega ((𝒪))`$ we then have at least for some open neighbourhood $`I`$ of 0 $$f(t):=\psi ,\stackrel{~}{U}(t)a\mathrm{\Omega }_\omega =0tI.$$ Since $`\omega `$ is a KMS-state ($`\beta >0`$) or a ground-state ($`\beta =\mathrm{}`$), $`f(t)`$ is the boundary value of a function $`F(z)`$ which is analytic on the strip $$𝒟_{\beta /2}=\{z;0<Im(z)<\beta /2\}$$ and bounded and continuous on $`\overline{𝒟_{\beta /2}}`$. By the Schwartz reflection principle $`f(t)`$ vanishes on the whole real axis. Therefore $`\stackrel{~}{U}(t)\psi ,a\mathrm{\Omega }_\omega =0`$ for all $`t`$. Hence, $`^{}`$ is invariant under the action of $`\stackrel{~}{U}(t)`$. ∎ We are now able to give the proof of the main theorem: Proof of theorem 4.8. We denote by $``$ the von Neumann algebra $`_t\stackrel{~}{U}(t)\pi _\omega ((𝒪_1))\stackrel{~}{U}(t)^{}`$. Lemma 4.9 implies that $``$ is invariant under the action of $``$, i.e. $``$. By theorem 4.6 $`=\pi _\omega ()^{\prime \prime }`$. Hence $`\mathrm{\Omega }_\omega `$ is cyclic for $``$ and therefore $`=_\omega `$. By the inclusions (19) we have $`\overline{\pi _\omega ((𝒪))\mathrm{\Omega }_\omega }`$ and hence $`\mathrm{\Omega }_\omega `$ is cyclic for $`\pi _\omega ((𝒪))`$. ## 5 Examples of linear field theories In this section $`(M,g)`$ will be an oriented time-oriented 4-dimensional Lorentzian manifold which is globally hyperbolic in the sense that it admits a smooth Cauchy surface. For some subset $`𝒪M`$ the set of points, which can be reached by future/past directed causal curves emanating from $`𝒪`$, will be denoted by $`J^\pm (𝒪)`$. The free Dirac field (see ). It is known that $`M`$ possesses a trivial spin structure given by a $`\text{Spin}^+(3,1)`$-principal bundle $`SM`$ and a two-fold covering map $`SMFM`$ onto the bundle $`FM`$ of oriented time-oriented orthonormal frames. One can now construct the Dirac bundle $`DM`$ which is associated to $`SM`$ by the spinor representation and is a natural module for the Clifford algebra bundle Cliff($`TM`$) (see and ). Furthermore, the Levi-Civita connection on $`TM`$ induces a connection on $`DM`$ with covariant derivative $$:\mathrm{\Gamma }(SM)\mathrm{\Gamma }(SMT^{}M).$$ Given a vector field $`n`$ we write as usual $`/n`$ for the section in the Clifford algebra bundle $`\gamma (n)`$, or in local coordinates $`\gamma ^in_i`$. There exists an antilinear bijection $`\mathrm{\Gamma }(DM)\mathrm{\Gamma }(DM^{}):uu^+`$, the Dirac conjugation, which in the standard representation in a local orthonormal spin frame has the form $`u^+=\overline{u}\gamma ^0`$, where the bar denotes complex conjugation in the dual frame. We use the symbol <sup>+</sup> also for the inverse map. Canonically associated with the Dirac bundle there is the Dirac operator which in a frame takes the form $$/=\gamma ^i_{e_i}.$$ The Dirac equation for mass $`m0`$ is $$(i/+m)u=0,u\mathrm{\Gamma }(DM).$$ (20) The Dirac equation has unique advanced and retarded fundamental solutions $`S^\pm :\mathrm{\Gamma }_0(DM)\mathrm{\Gamma }(DM)`$ satisfying $$(i/+m)S^\pm =S^\pm (i/+m)=\mathrm{id}\text{on}\mathrm{\Gamma }_0(DM),$$ $$\text{supp}(S^\pm f)J^\pm (\text{supp}(f)).$$ We can define the operator $`S:=S^+S^{}`$ and form the pre-Hilbert space $$H:=\mathrm{\Gamma }_0(DM)/\text{ker}(S)$$ (21) with inner product $$[u_1],[u_2]_H:=i_M(u_1^+,Su_2)(x)𝑑\mu (x),$$ (22) where $`(,)`$ denotes the dual pairing between the fibres of $`DM`$ and $`DM^{}`$ and $`\mu (x)`$ the canonical (pseudo)-Riemannian measure on $`M`$. $`[f]`$ denotes the equivalence class containing $`f`$. Analogously we form the pre-Hilbert space $$H^+:=\mathrm{\Gamma }_0(DM^{})/(\text{ker}(S))^+$$ (23) with inner product $$[v_1],[v_2]_{H^+}:=[v_2^+],[v_1^+]_H.$$ (24) We note that $`S`$ maps $`H`$ onto the space of smooth solutions to the Dirac equation whose supports have compact intersections with any Cauchy surface. The construction of the Dirac field starts with the Hilbert space $`:=\overline{HH^+}`$ and the antiunitary involution $$𝒦:,uvv^+u^+.$$ (25) By construction we have a map $$\eta :\mathrm{\Gamma }_0(DMDM^{});fg[f][g]$$ with dense range. Let $`G`$ be the identity component of the group of isometries of $`M`$ and $`\stackrel{~}{G}`$ its universal covering group. $`\stackrel{~}{G}`$ acts canonically on $`FM`$ and since $`\stackrel{~}{G}`$ is simply connected this actions lifts uniquely to a smooth action on $`SM`$, yielding a smooth equivariant action on the bundles $`DM`$ and $`DM^{}`$, making them $`\stackrel{~}{G}`$-vector bundles. The Dirac operator $`/`$ and its conjugate $`/^+:={}_{}{}^{+}/^+`$ are both invariant under these actions and hence the representation of $`\stackrel{~}{G}`$ on $`\mathrm{\Gamma }_0(DMDM^{})`$ gives rise to a unitary representation $`\rho `$ of $`\stackrel{~}{G}`$ on $``$. Taking $`E:=DMDM^{}`$ one shows that $`(,𝒦,E,\eta ,\rho )`$ is a linear fermionic field theory on $`M`$. Moreover, the differential operator $`P=//^+:\mathrm{\Gamma }_0(E)\mathrm{\Gamma }_0(E)`$ maps to the kernel of $`\eta `$, i.e. $`\eta P=0`$. The square of $`P`$ has metric principal part. The real scalar field (see ). The construction of the Klein-Gordon field starts with the Klein-Gordon operator for mass $`m0`$: $$P:=\mathrm{}_g+m^2,$$ (26) where $`\mathrm{}_g=g^{ik}_i_k`$ and $``$ is the Levi-Civita covariant derivative. This operator acts on the real-valued smooth functions with compact support $`C_{0r}^{\mathrm{}}(M)`$. It has unique advanced and retarded fundamental solutions $`F_s^\pm :C_{0r}^{\mathrm{}}(M)C_r^{\mathrm{}}(M)`$ satisfying $$PF_s^\pm =F_s^\pm P=\mathrm{id}\text{on}C_{0r}^{\mathrm{}}(M),$$ $$\text{supp}(F_s^\pm f)J^\pm (\text{supp}(f)).$$ With $`F_s:=F_s^+F_s^{}`$, $`\widehat{\sigma }(f_1,f_2):=_Mf_1F_s(f_2)w`$ defines an antisymmetric bilinear form on $`C_{0r}^{\mathrm{}}(M)\times C_{0r}^{\mathrm{}}(M)`$, where $`w`$ is the pseudo-Riemannian volume form on $`M`$. Defining $`𝒲:=C_{0r}^{\mathrm{}}(M)/\text{ker}(F_s)`$ with quotient map $`\eta `$, the bilinear form $`\sigma (\eta (f_1),\eta (f_2)):=\widehat{\sigma }(f_1,f_2)`$ on $`𝒲`$ is symplectic. We have a canonical linear action of the group $`G`$ on $`C_{0r}^{\mathrm{}}(M)`$ which leaves $`\text{ker}(F_s)`$ and $`\widehat{\sigma }`$ invariant and hence gives a representation $`\rho `$ of $`G`$ on $`𝒲`$ by symplectomorphisms. Taking the trivial bundle $`M\times `$ for $`E`$, one shows that $`(𝒲,\sigma ,E,\eta ,\rho )`$ is a linear bosonic field theory on $`M`$ and $`\eta P=0`$. Note that $`P`$ has metric principal part. ###### Remark 5.1. Since the complex scalar field consists of two independent real scalar fields it fits into this framework as well. The Proca field (see ). Let $`d`$ be the exterior derivative of differential forms, $``$ the Hodge star operator and $`\delta =d`$. Take the cotangent bundle for $`E`$. For mass $`m>0`$ the Proca equation for sections $`f\mathrm{\Gamma }_0(E)`$ is $$(\delta d+m^2)f=0,$$ (27) which is equivalent to the hyperbolic system $$(\mathrm{}_g+m^2)f=(\delta d+d\delta +m^2)f=0,$$ (28) $$\delta f=0.$$ (29) We define the differential operator $`\stackrel{~}{P}:\mathrm{\Gamma }_0(E)\mathrm{\Gamma }_0(E)`$ by $$\stackrel{~}{P}:=\mathrm{}_g+m^2.$$ (30) It has unique advanced and retarded fundamental solutions $`\stackrel{~}{F}_p^\pm :\mathrm{\Gamma }_0(E)\mathrm{\Gamma }(E)`$ satisfying $$\stackrel{~}{P}\stackrel{~}{F}_p^\pm =\stackrel{~}{F}_p^\pm \stackrel{~}{P}=\mathrm{id}\text{on}\mathrm{\Gamma }_0(E),$$ $$\text{supp}(\stackrel{~}{F}_p^\pm f)J^\pm (\text{supp}(f)).$$ We define the operators $`F_p^\pm :=(m^2d\delta +1)\stackrel{~}{F}_p^\pm `$. It is not difficult to see that the $`F_p^\pm `$ are the unique fundamental solutions for the operator $`P=\delta d+m^2`$ with the above properties. With $`F_p:=F_p^+F_p^{}`$, $`\widehat{\sigma }(f_1,f_2):=_Mf_1F_p(f_2)`$ defines an antisymmetric bilinear form on $`\mathrm{\Gamma }_0(E)\times \mathrm{\Gamma }_0(E)`$. Taking $`𝒲:=\mathrm{\Gamma }_0(E)/\text{ker}(F_p)`$ with quotient map $`\eta `$, the bilinear form $`\sigma (\eta (f_1),\eta (f_2)):=\widehat{\sigma }(f_1,f_2)`$ on $`𝒲`$ is symplectic. The pullback of forms induces a linear action of $`G`$ on $`\mathrm{\Gamma }_0(E)`$ which leaves $`\text{ker}(F_p)`$ and $`\widehat{\sigma }`$ invariant and hence gives rise to a representation $`\rho `$ of $`G`$ on $`𝒲`$ by symplectomorphisms. Again one can show that $`(𝒲,\sigma ,E,\eta ,\rho )`$ is a linear bosonic field theory and moreover, $`\eta \stackrel{~}{P}=\eta P=0`$. One gets the following corollary. ###### Corollary 5.2. Let $`(M,g,h_t)`$ be a connected stationary globally hyperbolic oriented time-oriented 4-dimensional Lorentzian manifold. Let $`𝒪(𝒪)`$ be a net of field algebras for one of the following free fields: * The real or complex scalar field for mass $`m0`$, * The Proca field for mass $`m>0`$, * The Dirac field for mass $`m0`$, Assume that $`\omega `$ is a state over the field algebra $``$ which we require to be quasifree and continuous in the bosonic case and which is a ground- or KMS-state with respect to the canonical time translations. Then $`\omega `$ has the Reeh-Schlieder property. ## 6 Acknowledgements The author would like to thank Prof. M. Wollenberg and Dr. R. Verch for useful discussions and comments. This work was supported by the Deutsche Forschungsgemeinschaft within the scope of the postgraduate scholarship programme “Graduiertenkolleg Quantenfeldtheorie” at the University of Leipzig.
warning/0002/hep-th0002128.html
ar5iv
text
# 1 Introduction ## 1 Introduction Scattering processes in the curved backgrounds of p-brane configurations of M theory and string theory have been extensively studied over the past few years. Motivation for these studies is the fact that the low energy absorption cross-sections for different fields yield information about the two-point correlation functions in strongly coupled gauge theories via AdS/CFT correspondence . The extremal D3-brane background is of special interest, since the correspondence there is to D=4 super Yang-Mills theory. By now, there have been extensive studies of the scattering processes at low-energies for the minimally-coupled scalar field (dilaton-axion) from the gravity perspective, as well as the analyses of the correlation functions on the field-theory side (see, for example, and references therein), thus leading to important insights into the AdS/CFT correspondence. In particular, the supersymmetry constraints imply that the two-point correlation functions do not get renormalized by higher-order corrections on the field theory side, and a precise agreement between low-energy absorption cross-sections for all the partial waves of the dilaton field and the (weak coupling) field-theory calculations of the corresponding correlation functions was obtained . On the other hand, the scattering of other fields has been explored to a lesser extent, and that only for low-energies (see, for example, and references therein); nevertheless it is expected that non-renormalization theorems on the field theory side ensure a precise agreement between the low energy absorption cross-sections and the corresponding weak coupling calculation of the n-point correlation functions on the field-theory side (see, for example, and references therein). This paper addresses several issues. We provide an analysis of the absorption cross-section in the extremal D3-brane background for a broad class of massless modes and for the whole energy range. In particular, we uncover a pattern in the energy dependence of the absorption cross-sections for both integer and half-integer spins; certain half-integer and integer spin pairs have identical absorption cross-sections, thus providing an evidence on the gravity side that such pairs couple on the dual field theory side to the pairs of operators forming supermultiplets of strongly coupled gauge theory. The paper is organized as follows. In Section 2, we cast the wave equations of various fields into Schrödinger form, and obtain the effective Schrödinger potentials. We show that effective Schrödinger potentials for certain field are identical and hence the absorption for these fields is the same. In other cases where Schrödinger potentials are not the same, we argue that the different potentials are dual and yield the same absorption probabilities, with numerical results supporting these claims. In Section 3, we obtain numerical results for the absorption of a large class of fields for the whole energy range in the extremal D3-brane background. The method for the numerical evaluation of the absorption probabilities is given in Appendix A, while the calculation of the high energy absorption cross-section in the geometrical optics limit is given in Appendix B. ## 2 Effective potentials The D3-brane of the type IIB supergravity is given by $`ds_{10}^2`$ $`=`$ $`H^{1/2}(fdt^2+dx_1^2+dx_2^2+dx_3^2)+H^{1/2}(f^1dr^2+r^2d\mathrm{\Omega }_5^2),`$ $`G_{\left(5\right)}`$ $`=`$ $`d^4xdH^1+(d^4xdH^1).`$ (2.1) where $$H=1+\frac{R^4}{r^4},f=1\frac{2m}{r^4}.$$ (2.2) Here $`R`$ specifies the D3-brane charge and $`m`$ is the non-extremality parameter (defined for convenience as $`m\mu R^4/\sqrt{1+\mu }`$). We shall primarily concentrate on the extreme limit $`m=\mu =0`$. (See, however, Appendix B for the discussion of the high energy limit of the absorption cross-section in the non-extreme D3-background.) The low energy absorption probabilities for various bosonic linearly-excited massless fields under this background were obtained in . The low energy absorption probabilities for the gravitino and the two-form field are given in and , respectively (and for massive minimally coupled modes in ). In the following subsections we study the absorption probabilities for the whole energy range and uncover completely parallel structures. In particular, we shall cast the wave equation for different modes into Schrödinger form and discuss the pattern of the the Schrödinger potentials. We also provide a conjectured form of the dual potentials which, in turn, yield the same absorption probabilities. In the subsequent section we confirm the pattern with numerical results. ### 2.1 Dilaton-axion The axion and dilaton of the type IIB theory are decoupled from the D3-brane. Thus, in the D3-brane background, they satisfy the minimally-coupled scalar wave equation $$\frac{1}{\sqrt{g}}_\mu \sqrt{g}g^{\mu \nu }_\nu \varphi =0.$$ (2.3) It follows from (2.1) that the radial wave equation of a dilaton-axion in the spacetime of an extremal D3-brane is given by $$\left(\frac{1}{\rho ^5}\frac{}{\rho }\rho ^5\frac{}{\rho }+H\frac{\mathrm{}(\mathrm{}+4)}{\rho ^2}\right)\varphi (\rho )=0,$$ (2.4) where $$H=1+\frac{e^4}{\rho ^4}$$ (2.5) and $`\mathrm{}=0,1,\mathrm{}`$ corresponds to the $`\mathrm{}^{th}`$ partial wave. The quantity $`e`$ and $`\rho `$ are dimensionless energy and radial distance parameters: $`e=\omega R`$ and $`\rho =\omega r`$. The leading order and sub-leading order cross-sections of the minimally-coupled scalar by the D3-brane background were obtained in by matching inner and outer solutions of the wave equations. It was observed in that if one performs the following change of variables $$\rho =eExp(z),\varphi (r)=Exp(2z)\psi (r),$$ (2.6) the wave equation (2.4) becomes $$\left[\frac{^2}{z^2}+2e\mathrm{cosh}(2z)(\mathrm{}+2)^2\right]\psi (z)=0,$$ (2.7) which is precisely the modified Mathieu equation. One can then obtain analytically the absorption probability order by order in terms of dimensionless energy $`e`$ . In this paper, we shall express the wave equation in Schrödinger form, and study the characteristics of the Schrödinger effective potential. By the substitution $$\varphi =\rho ^{5/2}\psi ,$$ (2.8) we render (2.4) into Schrödinger form $$\left(\frac{^2}{\rho ^2}V_{\mathrm{eff}}\right)\psi =0,$$ (2.9) where $$V_{\mathrm{eff}}(\mathrm{})=H+\frac{(\mathrm{}+3/2)(\mathrm{}+5/2)}{\rho ^2}V_{\mathrm{dilaton}}(\mathrm{}).$$ (2.10) Factors shared by the incident and outgoing parts of the wave function cancel out when calculating the absorption probability. Thus, the absorption probability of $`\varphi `$ and $`\psi `$ are the same. Technically, $`V_{\mathrm{eff}}`$ cannot be interpreted as an effective potential, since it depends on the particle’s incoming energy. It is straightforward to use a coordinate transformation to put the equation in the standard Schrödinger form, where $`V_{\mathrm{eff}}`$ is independent on the energy. However, for our purposes of analyzing and comparing the form of the wave equations for various fields, this is of no consequence. Note that the first term in (2.10) represents the spacetime geometry of the extremal D3-brane, whereas the second term represents the angular dynamics (partial modes) of the particles. We shall see presently that some particles have effective potentials which contain terms mixed with both “geometrical” and “angular” dynamical contributions. ### 2.2 Antisymmetric tensor from 4-form For two free indices of the 4-form along $`S^5`$ and two free indices in the remaining 5 directions, the radial wave equation for the antisymmetric tensor derived from the 4-form is $$\left(\frac{1}{\rho }\frac{}{\rho }\rho \frac{}{\rho }+H\frac{(\mathrm{}+2)^2}{\rho ^2}\right)\varphi (\rho )=0,$$ (2.11) where $`\mathrm{}=1,2,\mathrm{}`$. By the substitution $$\varphi =\rho ^{1/2}\psi ,$$ (2.12) we render (2.11) into Schrödinger form with $$V_{4\mathrm{form}}(\mathrm{})=V_{\mathrm{dilaton}}(\mathrm{}).$$ (2.13) ### 2.3 Dilatino The radial wave equation for the dilatino on an extremal D3-brane was found by inserting the following spherical wave decomposition form for the dilatino field $`\lambda `$ into the covariant Dirac equation: $$H^{\frac{1}{8}}\lambda =e^{i\omega t}r^{\frac{5}{2}}\left(F(r)\mathrm{\Psi }_{\mathrm{}}^\pm +iG(r)\left(\frac{\mathrm{\Gamma }^0\mathrm{\Gamma }^ix_i}{r}\right)\mathrm{\Psi }_{\mathrm{}}^\pm \right),$$ (2.14) where $`\mathrm{\Gamma }^i`$ are field-independent gamma-matrices and $`i=4,\mathrm{},9`$ runs normal to the brane. $`\mathrm{\Psi }_{\mathrm{}}^\pm `$ is the eigenspinor of the total angular momentum with $`\mathrm{\Sigma }_{ij}L_{ij}=\mathrm{}`$, $`\mathrm{\Gamma }^{0123}=\pm i`$. The spatial momenta tangential to the branes can be made to vanish via the Lorentz transformations and the radial wave equations are obtained : $`\omega H^{\frac{1}{2}}F+\left({\displaystyle \frac{d}{dr}}+{\displaystyle \frac{\mathrm{}+5/2}{r}}\pm {\displaystyle \frac{1}{4}}(\mathrm{ln}H)^{}\right)G`$ $`=`$ $`0`$ (2.15) $`\omega H^{\frac{1}{2}}G+\left({\displaystyle \frac{d}{dr}}{\displaystyle \frac{\mathrm{}+5/2}{r}}{\displaystyle \frac{1}{4}}(\mathrm{ln}H)^{}\right)F`$ $`=`$ $`0`$ (2.16) Decoupling (2.15) and (2.16) yields second-order differential equations for $`F`$ and $`G`$. Let us first consider the case of positive eigenvalue, i.e., $`\mathrm{\Gamma }^{0123}=+i`$. In this case, the second-order wave equation for $`F`$ can be cast into Schrödinger form by the substitution $`F=H^{1/4}\psi `$, giving rise to the effective Schrödinger potential $$V_{+\mathrm{dilatino}}^F(\mathrm{})=V_{\mathrm{dilaton}}(\mathrm{}).$$ (2.17) Thus, the absorption probability for $`F`$ is identically the same as that for the dilaton-axion. The wave equation for $`G`$ can also be cast into Schrödinger form by the substitution $`G=H^{1/4}\psi `$. The corresponding Schrödinger potential, on the other hand, takes a different form, given by $$V_{+\mathrm{dilatino}}^G(\mathrm{})=V_{\mathrm{dilaton}}(\mathrm{})+\frac{3e^82e^8\mathrm{}10e^4\rho ^4+5\rho ^8+2\rho ^8\mathrm{}}{H^2\rho ^{10}}.$$ (2.18) Thus, we see that the effective potentials for the two components $`F`$ and $`G`$ are quite different. However, we expect that these two potentials, although different, yield the same absorption probability: they form a dual pair of potentials. While it is not clear to us how to present a rigorous analytical proof, our numerical calculation (in section 4) confirm that, indeed, they yield the same absorption probability. Similar results are obtained for the the negative eigenvalue solutions, i.e., $`\mathrm{\Gamma }^{0123}=i`$: $`V_{\mathrm{dilatino}}^G(\mathrm{})`$ $`=`$ $`V_{\mathrm{dilaton}}(\mathrm{}+1),`$ $`V_{\mathrm{dilatino}}^F(\mathrm{})`$ $`=`$ $`V_{\mathrm{dilaton}}(\mathrm{}+1)+{\displaystyle \frac{5\rho ^82\rho ^8\mathrm{}10e^4\rho ^4+7e^8+2e^8\mathrm{}}{H^2\rho ^{10}}}.`$ (2.19) Again, the numerical results in section 4 indicate that the above two potentials yield the same absorption probabilities and hence form a dual pair. In the above discussion of the dilatino scattering equation, we encountered three different potentials, namely $$V_{+\mathrm{dilatino}}^G(\mathrm{})\stackrel{\mathrm{dual}}{}V_{+\mathrm{dilatino}}^F(\mathrm{})=V_{\mathrm{dilatino}}^G(\mathrm{}1)\stackrel{\mathrm{dual}}{}V_{\mathrm{dilatino}}^F(\mathrm{}1)$$ (2.20) which all yield the same absorption probability. The above structure of the dual potentials can be cast in a more general form. Namely, the above dual potential pairs for the dilatino can be cast into the following form: $$V(\mathrm{})=H+\frac{(2\mathrm{}+\alpha )(2\mathrm{}+\alpha \pm 2)}{4\rho ^2},$$ (2.21) and $$V^{\mathrm{dual}}(\mathrm{})=V(\mathrm{})+\frac{\pm [(2\mathrm{}+\alpha \pm 2)e^8(2\mathrm{}+\alpha )\rho ^8]10e^4\rho ^4}{\rho ^{10}H^2},$$ (2.22) where $`\alpha =5`$ and we use in the $`\pm `$ sign in (2.22) for the $``$ eigenvalue, respectively. We have found numerically that (2.21) and (2.22) are dual potentials for integer values of $`\alpha `$. In particular, for odd values of $`\alpha `$, (2.21) can be identified with $`V_{\mathrm{dilaton}}(\mathrm{}+(\alpha 4\pm 1)/2)`$, in which case the absorption can be found analytically, since the wave equation is that of Mathieu equation. Thus, in all the subsequent examples when the potential is of the form (2.22), we are now able to identify a dual potential (2.21) of a simpler form with a corresponding wave equation which can be solved analytically. ### 2.4 Scalar from the two-form For the free indices of the two-form taken to lie along the $`S^5`$, the radial wave equation is $$\left(\frac{H}{\rho }\frac{}{\rho }\frac{\rho }{H}\frac{}{\rho }+H\frac{(\mathrm{}+2)^2}{\rho ^2}\frac{4e^4}{H\rho ^6}(\mathrm{}+2)\right)\varphi (\rho )=0,$$ (2.23) where again $`\mathrm{}=1,2,\mathrm{}`$ correspond to the $`\mathrm{}^{th}`$ partial wave. The sign $`\pm `$ corresponds to the sign in the spherical harmonics involved in the partial wave expansion. By the substitution $$\varphi =\rho ^{1/2}H^{1/2}\psi ,$$ (2.24) we render (2.23) into Schrödinger form. For the positive eigenvalue, the effective potential is of the form given by (2.22), with $`\alpha =5`$ and a positive sign. This is, in fact, the same as that for the dilatino with negative eigenvalue. Thus, the dual potential is $$V_{+\mathrm{scalar}}(\mathrm{})\stackrel{\mathrm{dual}}{}V_{\mathrm{dilaton}}(\mathrm{}+1).$$ (2.25) For the negative eigenvalue, the effective potential is of the form given by (2.22), with $`\alpha =3`$ and a negative sign. Thus, the dual potential is $$V_{\mathrm{scalar}}(\mathrm{})\stackrel{\mathrm{dual}}{}V_{\mathrm{dilaton}}(\mathrm{}1).$$ (2.26) ### 2.5 Two-form from the antisymmetric tensor The equations for two-form perturbations polarized along the D3-brane are coupled. For s-wave perturbations, they can be decoupled , and the radial wave equation is $$\left(\frac{1}{\rho ^5H}\frac{}{\rho }\rho ^5H\frac{}{\rho }+H\frac{16e^8}{\rho ^{10}H^2}\right)\varphi (\rho )=0,$$ (2.27) where $`\mathrm{}=0,1,\mathrm{}.`$ By the substitution $$\varphi =\rho ^{5/2}H^{1/2}\psi ,$$ (2.28) we render (2.27) into Schrödinger form given by (2.22), with $`\alpha =5`$, $`\mathrm{}=0`$ and the positive sign. Thus, the dual potential is $$V_{2\mathrm{form}}(\mathrm{}=0)\stackrel{\mathrm{dual}}{}V_{\mathrm{dilaton}}(\mathrm{}=1).$$ (2.29) ### 2.6 Vector from the two-form We now consider one free index of the two-form along $`S^5`$ and one free index in the remaining 5 directions. For the tangential components of this vector field, the radial wave equation is $$\left(\frac{1}{\rho ^3}\frac{}{\rho }\rho ^3\frac{}{\rho }+H\frac{(\mathrm{}+1)(\mathrm{}+3)}{\rho ^2}\right)\varphi (\rho )=0,$$ (2.30) where $`\mathrm{}=1,2,\mathrm{}.`$ By the substitution $$\varphi =\rho ^{3/2}\psi ,$$ (2.31) we render (2.30) into Schrödinger form with $$V_{\mathrm{tangential}\mathrm{vector}}(\mathrm{})=V_{\mathrm{dilaton}}(\mathrm{}).$$ (2.32) The radial $`a_r`$ and time-like $`a_0`$ components of the vector field are determined by the following coupled first order differential equations : $$i\frac{}{\rho }a_o=\left[1\frac{(\mathrm{}+1)(\mathrm{}+3)}{\rho ^2H}\right]a_r$$ (2.33) and $$i\frac{1}{\rho }\frac{}{\rho }\left(\frac{\rho }{H}a_r\right)=a_o,$$ (2.34) where $`\mathrm{}=1,2,\mathrm{}.`$ Eqs. (2.33) and (2.34) can be decoupled and the wave equation for $`a_r`$ is $$\left(H\frac{}{\rho }\frac{1}{\rho }\frac{}{\rho }\frac{\rho }{H}+H\frac{(\mathrm{}+1)(\mathrm{}+3)}{\rho ^2}\right)a_r=0.$$ (2.35) By the substitution $$a_r=\rho ^{1/2}H\psi ,$$ (2.36) we render (2.35) into Schrödinger form with $$V_{\mathrm{radial}\mathrm{vector}}(\mathrm{})=V_{\mathrm{dilaton}}(\mathrm{}).$$ (2.37) The wave equation for $`a_0`$ is $$\left(\frac{1}{\rho }\frac{}{\rho }\frac{\rho ^3}{\rho ^2H(\mathrm{}+1)(\mathrm{}+3)}\frac{}{\rho }+1\right)a_0=0.$$ (2.38) By the substitution $$a_0=\rho ^{3/2}\sqrt{\rho ^2H(\mathrm{}+1)(\mathrm{}+3)}\psi ,$$ (2.39) we render (2.38) into Schrödinger form with $$V_{0\mathrm{vector}}(\mathrm{})=V_{\mathrm{dilaton}}(\mathrm{})+\frac{3e^810e^4\rho ^4+4\rho ^6(\mathrm{}+1)(\mathrm{}+3)\rho ^8}{\rho ^{10}\left(H\frac{(\mathrm{}+1)(\mathrm{}+3)}{\rho ^2}\right)^2}.$$ (2.40) Note that this effective potential is nonsingular only for $`e^2>(\mathrm{}+1)(\mathrm{}+3)/2`$. We have numerically confirmed that $`V_{0\mathrm{vector}}(\mathrm{})`$ is dual to $`V_{\mathrm{dilaton}}(\mathrm{})`$. Working directly from Eqs. (2.33)-(2.34), we have numerically confirmed that $`a_0`$ shares the same absorption probability with the dilaton-axion, for all values of $`e`$. ## 3 Qualitative features of absorption by the extremal D3-brane We have obtained numerical absorption probabilities by a method described in Appendix A. Fig. 1 shows the s-wave absorption probabilities for all the particles that we have studied on an extremal D3-brane vs. the energy of incoming particles measured in dimensionless units of $`e\omega R`$. (a) is the s-wave absorption probability of the dilaton-axion, dilatino with positive total angular momentum eigenvalue and scalar from the two-form with a negative sign in the spherical harmonic. (b) is the s-wave absorption probability of the dilatino with negative total angular momentum eigenvalue, two-form from the antisymmetric tensor, antisymmetric tensor from the four-form and the longitudinal and tangential components of the vector from the two-form. (c) is the s-wave absorption probability of the scalar from the two-form with a positive sign in the spherical harmonic. These absorption probabilities are simply related by $`\mathrm{}\mathrm{}\pm 1`$. This structure is suggestive of a particle supermultiplet structure– namely, different multiplets. This demonstrates the similar and surprisingly simple structure of absorption probabilities between the various particles, which is not apparent from previous analytical low-energy absorption probabilities. These numerical results also support the idea of dual potentials. Fig. 2 shows the partial absorption probabilities for a dilaton-axion on an extremal D3-brane vs. the energy of incoming particles. There is no absorption at zero energy and total absorption is approached at high energy. Thus, the high energy absorption cross-section is found by setting $`P=1`$ in (A.4): $$\sigma ^{(\mathrm{})}=\frac{8\pi ^2(\mathrm{}+3)(\mathrm{}+2)^2(\mathrm{}+1)}{3(\omega R)^2}$$ (3.1) In fact, for all types of waves absorbed by all branes and black holes studied thus far , total partial wave absorption occurs at high energy. It is conjectured that this is a general property of absorption for all objects that have an event horizon. Results for objects other than D3-branes will be published shortly by the present authors. The partial absorption probabilities of a massive minimally-coupled scalar on an extremal D3-brane vs. the energy of the incoming particles is shown in Fig. 3. As the mass of the scalar is increased, the partial absorption probabilities occur at lower energies. Physically, there is greater absorption at low energy due to the additional gravitational attraction that is present from the nonzero mass. The low energy absorption probability for this case were obtained in . Fig. 4 shows the numerical s-wave absorption cross-section for a dilaton-axion on an extremal D3-brane (continuous line), super-imposed with previously obtained low energy semi-analytical results (short dashes) , as well as high energy total absorption (long dashes) vs. the energy of incoming particles. Throughout the rest of this paper, absorption cross-sections are plotted in units of $`R^5`$. The resonance roughly corresponds to the region in which there is a transition from zero absorption to total absorption. As evident from Fig. 5, the energies at which there is a resonance in the partial absorption cross-section are proportional to the partial-wave number. Also, the magnitude of the peak of each partial absorption cross-section decreases with the partial-wave number. These characteristics seem reasonable if one considers particle dynamics; in terms of radial motion, rotational kinetic energy counteracts gravitational attraction. In Fig. 6, we plot the partial-wave effective potentials vs. radial distance (in dimensionless units) for a dilaton-axion on an extremal D3-brane. These are plotted at the energies of the maxima of the corresponding partial-wave absorption cross-sections. For the s-wave absorption cross-section, the maximum is at $`e=1.4`$. The maxima of higher partial-wave absorption cross-sections are separated by energy gaps of approximately $`\mathrm{\Delta }e=.75`$, with increasing $`\mathrm{}`$. As can be seen, the incoming particle must penetrate an effective barrier in order to be absorbed by the D3-brane. Once absorbed, the waves inhabit quasi-bound states until they quantum tunnel to asymptotically flat spacetime. As is shown, the heights of the partial-wave effective potential barriers at the energies of the maxima of the partial-wave absorption cross-sections are roughly equal, which partly explains the similar structure of the partial absorption probabilities of different partial-wave numbers. Thus, the decreasing maximum values of the partial-wave absorption cross-sections with increasing $`\mathrm{}`$ arises purely as a result of the phase-factors. The superposition of maxima of partial-wave absorption cross-section leads to the oscillatory character of the total absorption cross-section with respect to the energy of the incoming particles. This is shown in Fig. 7 for the case of the dilaton-axion, dilatino and scalar from the two-form for comparison. (a) is the total absorption cross-section of the dilaton-axion on an extremal D3-brane, (b) is that of the dilatino with positive total angular momentum eigenvalue, (c) is that of the scalar from the two-form with positive sign in the spherical harmonic, and (d) is that of the dilatino with negative total angular momentum eigenvalue. The amplitude of oscillation decreases exponentially with energy. For the dilaton-axion, the total absorption cross-section converges to the geometrical optics limit at high energy, which we have calculated in a previous section. The oscillatory behavior is shared by all total absorption cross-sections that have been studied as of this time , for various particles in various spacetime backgrounds. The oscillatory structure of the absorption cross-section of a scalar on a Schwarzschild black hole has been noted by Sanchez . It is interesting to note that extinction cross-sections which arise in the field of optics have similar oscillatory properties . Fig. 8 shows the total absorption cross-section of the scalar from the two-form with negative total angular momentum eigenvalue. As already noted and shown in Fig. 3, as the mass of the minimally-coupled scalar is increased, the partial-wave absorption probabilities occur at lower energies. This causes the drastic qualitative difference between the massless and massive scalar cases at low energy, i.e., the divergent total absorption cross-section at zero energy for the massive scalar, as is shown in Fig. 9. ## Acknowledgments We would like to thank C. Pope for useful discussions and J.V.P. would like to thank A. Maarouf for his assistance in formating the plots. ## Appendix A Outline of Numerical Method We will now outline the numerical method for finding the absorption cross-section. We will use the well-studied case of the dilaton-axion on an extremal D3-brane. In this case, the radial wave equation is given by (2.9) and (2.10). We take the wave close to the horizon, at $`\rho =.01`$, to be purely incoming: $$\psi (\rho )=\rho \mathrm{exp}\left(\frac{i(\omega R)^2}{\rho }\right)$$ (A.1) The solution in the far region is: $$\psi (\rho )=A_{in}\mathrm{exp}(i\rho )+A_{out}\mathrm{exp}(i\rho ).$$ (A.2) We use Mathematica to numerically integrate (2.9) with the boundary condition given by (A.1). At $`\rho =45`$, we match the result with (A.2) to find $`A_{in}`$ and $`A_{out}`$. The absorption probability is given by $$P=1|\frac{A_{out}}{A_{in}}|^2.$$ (A.3) The absorption cross-section for a scalar is $$\sigma ^{(\mathrm{})}=\frac{8\pi ^2(\mathrm{}+3)(\mathrm{}+2)^2(\mathrm{}+1)}{3(\omega R)^2}P^{(\mathrm{})}$$ (A.4) Numerical integration for other types of particles incident on other objects with event-horizons is straight-forward. ## Appendix B High-energy absorption cross-section for a dilaton-axion on a nonextremal D3-brane In this appendix, we obtain the analytical high energy absorption cross-section for a dilaton-axion in a nonextremal D3-brane background, by employing the geometrical optics limit for the classical motion of a particle . For a nonextremal D3-brane, the metric takes the form (2.1). The classical Lagrangian for a particle is of the form: $$L=\frac{1}{2}g_{\alpha \beta }\dot{x}^\alpha \dot{x}^\beta .$$ (B.1) $`\dot{x}^\alpha dx^\alpha /d\lambda `$, where $`\lambda `$ is an affine parameter. The Euler-Lagrange equations are $$\frac{d}{d\lambda }\left(\frac{L}{\dot{x}^\alpha }\right)\frac{L}{x^\alpha }=0.$$ (B.2) The equation of motion for $`\theta `$ is $$\frac{d}{d\lambda }(H^{1/2}r^2\dot{\theta })=H^{1/2}r^2\mathrm{sin}\theta \mathrm{cos}\theta (\dot{\varphi _3}^2\dot{\varphi }^2\mathrm{cos}^2\varphi \dot{\varphi _1}^2\mathrm{sin}^2\varphi \dot{\varphi _2}^2).$$ (B.3) The solution of (B.3) is $`\theta =\pi /2`$ and $`\dot{\theta }=0`$. The equations of motion for $`\varphi _3`$ and $`t`$ are of the form $$\frac{d}{d\lambda }(H^{1/2}r^2\dot{\varphi _3})=0$$ (B.4) and $$\frac{d}{d\lambda }(H^{1/2}f\dot{t})=0.$$ (B.5) (B.4) and (B.5) each imply a constant of motion: $$H^{1/2}r^2\dot{\varphi _3}=\mathrm{constant}\mathrm{}$$ (B.6) and $$H^{1/2}f\dot{t}=\mathrm{constant}E,$$ (B.7) where $`\mathrm{}`$ and $`E`$ are interpreted as the angular momentum and energy of the particle, respectively. Also, since the particle scatters only in a direction transverse to the D3-brane (i.e., it does not travel along the D3-brane), $`\dot{x_i}=0`$, for $`i=1,2,3`$. Substituting our solutions for $`\theta `$ and $`\dot{x_i}`$ into the Lagrangian yields $$2L=H^{1/2}f\dot{t}^2+H^{1/2}f^1\dot{r}^2+H^{1/2}r^2\dot{\varphi _3}^2.$$ (B.8) Instead of finding a rather complicated equation of motion for $`r`$, we use the fact that, for a massless particle, $$2L=g_{\alpha \beta }p^\alpha p^\beta =0,$$ (B.9) where $`p^\alpha =\dot{x}^\alpha `$. Substituting our results for $`\dot{\varphi _3}`$ and $`\dot{t}`$ into the previous equation yields $$\dot{r}^2=E^2\frac{\mathrm{}^2}{r^2}\frac{f}{H}.$$ (B.10) We introduce a new parameter, $`\lambda ^{}\mathrm{}\lambda `$, so that $$\left(\frac{dr}{d\lambda ^{}}\right)^2=\frac{1}{b^2}V_{\mathrm{effective}},$$ (B.11) where $`b\mathrm{}/E`$ is the impact parameter and $$V_{\mathrm{effective}}=\frac{1}{r^2}\frac{f}{H}.$$ (B.12) The absorption cross-section for particles at high energy can be obtained by determining the classical trajectory of the scattered particle and using the optical limit result: $$\sigma _{\mathrm{abs}}=\frac{8}{15}\pi ^2b_{\mathrm{crit}}^5,$$ where the critical impact parameter separating absorption from scattering orbits is given by $`1/b_{\mathrm{crit}}^2=V_{\mathrm{maximum}}`$. Thus, $`\sigma _{\mathrm{abs}}`$ $`=`$ $`{\displaystyle \frac{8}{15}}\pi ^2({\displaystyle \frac{1}{2}}R^4+3m+{\displaystyle \frac{1}{2}}\sqrt{(R^4+6m)^2+8mR^4})^{5/4}\times `$ (B.13) $`=`$ $`\left({\displaystyle \frac{\frac{3}{2}R^4+3m+\frac{1}{2}\sqrt{(R^4+6m)^2+8mR^4}}{\frac{1}{2}R^4+m+\frac{1}{2}\sqrt{(R^4+6m)^2+8mR^4}}}\right)^{5/2}.`$ In the extremal limit, i.e. $`m=0`$, this result reduces to: $$\sigma _{\mathrm{abs}}=\frac{32\sqrt{2}}{15}\pi ^2R^5.$$ (B.14)
warning/0002/nucl-th0002029.html
ar5iv
text
# 1 Introduction ## 1 Introduction Experimental efforts at Brookhaven’s Relativistic Heavy Ion Collider will soon bring to fruition laboratory studies of nuclear systems heated and compressed to energy densities far exceeding that of the proton. Quantum chromodynamics (QCD), which is believed be the appropriate description in the subatomic domain, predicts that matter produced in this regime could have its subhadronic degrees of freedom liberated from hadronic boundaries and allowed to move about as a plasma of quarks and gluons. Identification of the quark-gluon plasma (QGP) concerns much of the current activities as models and interpretation seem to point toward a common conception regarding the nature of the transition or crossover from ordinary hadronic matter to QGP. However, many issues remain unsettled providing a wealth of opportunities in the vibrant and rapidly advancing research field of ultra-relativistic heavy ion physics. Possibly the best available tool for studying nuclear systems in such states of excitation are electromagnetic probes of photons and lepton pairs. Each has its kinematical advantages, but both share the property that once produced in the system, undeflected flight to detectors ensues. Since virtual photons couple directly to the neutral component of the vector hadronic current, dileptons have been raised to a level of premier importance as they allow study of in-medium properties of the vector mesons. In the low mass sector, effects of the $`\rho `$ meson clearly appear in proton-nucleus reactions whereas in heavy-ion experiments a significant broadening of the “$`\rho `$” distribution is observed. The importance of broadening effects through relatively high scattering rates were highlighted several years ago in Ref. . Since then sophisticated hadronic models have been developed which seek to include in a consistent formalism these and other effects into vector meson spectral functions. The emerging lore favors a $`\rho `$ distribution in matter which is broadened essentially beyond recognition. In the higher mass sector, $`J/\psi `$ appears as a promising tool for spectroscopy. Within conventional hadronic scenarios it is expected to contribute a measurable muon-pair signal, whereas in QGP scenarios color screening effectively inhibits $`c\overline{c}`$ binding and limits $`J/\psi `$ survival probabilities. $`J/\psi `$ yields are therefore expected to be severely suppressed when plasma is produced. The notion has recently gained a great deal of attention as $`J/\psi `$ yields from Pb+Pb reactions at 158 $`A`$GeV show “anomalous” suppression compared simple extrapolations from lighter projectiles’ results. Models of absorption on hadronic comovers are able to provide interpretation for the lighter system, but not consistently for the lead results. Very intriguing analyses involving QGP scenarios have been put forward which are able to explain the yields. Meanwhile, purely hadronic approaches have been proposed too, but necessary input of $`J/\psi `$ scattering cross sections with light hadrons are up to now quite uncertain. We report here on results of effective field theoretical methods of estimating cross sections for $`J/\psi `$ with light hadrons and we use them to construct a spectral function for $`J/\psi `$ in hot hadronic matter. ## 2 Effective Lagrangian Hadronic interactions are here modeled with meson exchange, consequently a symmetry embodying strangeness and charm is needed. We therefore begin with SU(4) and introduce pseudoscalar ($`\varphi `$ = $`\phi _a\lambda _a`$) and vector (V<sup>μ</sup> = $`v_a^\mu \lambda _a`$) meson matrices, where $`\phi _a`$ and $`v_a^\mu `$ are pseudoscalar and vector multiplets and the $`\lambda `$s are SU(4) generators. The large charm quark mass and its symmetry breaking effects are duly noted, but since we use physical mass eigenstates and we incorporate empirical constraints on the model where possible, we expect reasonably reliable results as evidenced by the coupling constants’ near universality. The free meson Lagrangian is written as $`\text{L}_0=\mathrm{Tr}(^\mu \varphi ^{}_\mu \varphi )\mathrm{Tr}\left(\left(_\mu V_\nu ^{}\right)\left(^\mu V^\nu ^\nu V^\mu \right)\right)+\mathrm{mass}\mathrm{terms}.`$ (1) The pseudoscalar meson mass matrix that leads to properly normalized mass terms is $`\varphi =\left(\begin{array}{cccc}\frac{\pi ^0}{\sqrt{2}}+\frac{\eta }{\sqrt{6}}+\frac{\eta _c}{\sqrt{12}}& \pi ^+& K^+& \overline{D}^0\\ \pi ^{}& \frac{\pi ^0}{\sqrt{2}}+\frac{\eta }{\sqrt{6}}+\frac{\eta _c}{\sqrt{12}}& K^0& D^{}\\ K^{}& \overline{K}^0& \eta \sqrt{\frac{2}{3}}+\frac{\eta _c}{\sqrt{12}}& D_s^{}\\ D^0& D^+& D_s^+& 3\frac{\eta _c}{\sqrt{12}}\end{array}\right),`$ (6) while that for the vector multiplet is (suppressing the Lorentz index) $`V=\left(\begin{array}{cccc}\frac{\rho ^0}{\sqrt{2}}+\frac{\omega }{\sqrt{6}}+\frac{J/\psi }{\sqrt{12}}& \rho ^+& K_{}^{}{}_{}{}^{+}& \overline{D^{}}^0\\ \rho ^{}& \frac{\rho ^0}{\sqrt{2}}+\frac{\omega }{\sqrt{6}}+\frac{J/\psi }{\sqrt{12}}& K_{}^{}{}_{}{}^{0}& D_{}^{}{}_{}{}^{}\\ K_{}^{}{}_{}{}^{}& \overline{K^{}}^0& \omega \sqrt{\frac{2}{3}}+\frac{J/\psi }{\sqrt{12}}& D_{}^{}{}_{s}{}^{}\\ D_{}^{}{}_{}{}^{0}& D_{}^{}{}_{}{}^{+}& D_{}^{}{}_{s}{}^{+}& 3\frac{J/\psi }{\sqrt{12}}\end{array}\right).`$ (11) We then introduce interactions through a gauge covariant minimal substitution $`_\mu \text{f}\text{D}_\mu \text{f}\text{=}_\mu \text{f}\text{+ [}\text{A}_\mu \text{,}\text{}\text{]}`$, where $`\text{A}_\mu `$ = -$`ig/2V_\mu `$. As an aside remark, we note that the appearance of the $`g/2`$ instead of a mere $`g`$ is nothing but a choice of convention. Since the model is calibrated to data, the choice is irrelevant. We must collect terms up to order $`g^{\mathrm{\hspace{0.17em}2}}`$ for a consistently gauge invariant description for the hadronic currents. They are (since $`\varphi ^{}`$ = $`\varphi `$ and V = V) $`\text{L}_{\mathrm{int}}`$ $`=`$ $`ig\text{Tr}(\varphi V^\mu _\mu \varphi ^\mu \varphi V_\mu \varphi )+{\displaystyle \frac{1}{2}}g^{\mathrm{\hspace{0.17em}2}}\text{Tr}(\varphi V^\mu V_\mu \varphi \varphi V^\mu \varphi V_\mu )`$ (12) $`+`$ $`ig\text{Tr}\left(^\mu V^\nu [V_\mu ,V_\nu ]+[V^\mu ,V^\nu ]_\mu V_\nu \right)+g^{\mathrm{\hspace{0.17em}2}}\text{Tr}\left(V^\mu V^\nu [V_\mu ,V_\nu ]\right).`$ Carrying out the matrix algebra, we arrive at the totality of interaction terms allowed in the symmetry group. The model is calibrated to observed hadronic decays, tuned to respect vector dominance, or as a last resort, to respect SU(4) symmetry. Details can be found in Ref. . ### 2.1 The $`𝑱\mathbf{/}𝝍\mathbf{+}𝒉`$ cross section Of particular importance are cross sections for a light hadron to knock apart the charmonium leaving D, D or antiparticles when appropriate for conservation of relevant quantities. We present in Fig. 1 the results for the $`\pi `$-, K-, and $`\rho `$-induced dissociation. Purely elastic $`J/\psi +hJ/\psi +h`$ are found to be of order femtobarns for pions, microbarns for rho and nanobarns for kaons, clearly too small for any significance. ## 3 The $`𝑱\mathbf{/}`$$`𝝍`$ spectral function We next place $`J/\psi `$ in a strongly interacting medium and explore possible modifications in spectral properties as compared to the vacuum. From previous studies on vector mesons, one has seen that two-loop effects are much stronger than one-loop. We have checked to see this trend continue for charmonium. We here adopt the approach to neglect one-loop effects and use free masses throughout. We calculate the expected broadening of the $`J/\psi `$ distribution from collisions with light pseudoscalar and vector mesons in the hot hadronic matter. The extra width induced in the distribution by a reaction of type $`J/\psi \mathrm{\hspace{0.17em}2}34`$, where 2, 3, and 4 are arbitrary species is . $`\mathrm{\Gamma }(\omega ,\stackrel{}{p})`$ $`=`$ $`{\displaystyle \frac{1}{2\omega }}{\displaystyle 𝑑\mathrm{\Omega }n_2(E_2)(1+n_3(E_3))(1+n_4(E_4))|\overline{(J/\psi 234)}|^2},`$ (13) where $`\omega =\sqrt{\stackrel{}{p}^2+m_{J/\psi }^2}`$, $`\stackrel{}{p}`$ being the three-vector of the $`J/\psi `$. Note that 3 or 4 can be a $`J/\psi `$. In Eq. (13), $`d\mathrm{\Omega }`$ $`=`$ $`d\overline{p}_2d\overline{p}_3d\overline{p}_4(2\pi )^4\delta (p+p_2p_3p_4),`$ (14) where we have used shorthand notation $`d\overline{p}_i`$ $`=`$ $`{\displaystyle \frac{d^3p_i}{(2\pi )^3\mathrm{\hspace{0.17em}2}E_i}}.`$ (15) A direct connection can be drawn between the rate in Eq. (13) and the eventual structure of the spectral function, utilizing at the intermediate stages such field theoretical concepts as the $`J/\psi `$ propagator and imaginary part of the self-energy. In an on-shell approximation, one has for the full spectral function $`A_{J/\psi }(\omega ,\stackrel{}{p})={\displaystyle \frac{2m_{J/\psi }\mathrm{\Gamma }_{J/\psi }}{(p^2m_{J/\psi }^2)^2+m_{J/\psi }^2\mathrm{\Gamma }_{J/\psi }^2}},`$ (16) where $`\mathrm{\Gamma }_{J/\psi }`$ contains the vacuum width as well as contributions from elastic and inelastic collisions. Discussion of Eq. (16) in the context of more general formalism will be published elsewhere. Now we consider $`J/\psi `$ in a finite temperature gas consisting of $`\pi `$s, $`K`$s, $`\rho `$s and $`K^{}`$s. The spectral function at 150 MeV is shown below in Fig. 2. Severe broadening of the spectral distribution with an accompanying suppression of the peak is immediately clear already for temperatures on the order of the pion mass. Studies of temperature dependence as well as specific channels’ contributions will also be included in Ref. . ## 4 Effects of form factors Up to now pointlike hadrons have been considered. Effective lagrangian approaches are not considered complete until finite size effects are incorporated, typically accomplished with vertex form factors. Here we outline the procedure for implementation. Each $`t`$-channel Feynman graph is given a product of monopoles, one for each vertex, having the following structure $`h(t)`$ $`=`$ $`\left({\displaystyle \frac{\mathrm{\Lambda }_1+t_{max}m_\alpha ^2}{\mathrm{\Lambda }_1+t_{max}t}}\right)\left({\displaystyle \frac{\mathrm{\Lambda }_2+t_{max}m_\alpha ^2}{\mathrm{\Lambda }_2+t_{max}t}}\right),`$ (17) where $`\mathrm{\Lambda }_1`$ and $`\mathrm{\Lambda }_2`$ are cutoff parameters depending on the species of on-shell particle, and $`m_\alpha `$ is the mass of the exchanged meson. Two cases will be examined: first we choose a fixed cutoff for all vertices of 2, 3 and 4 GeV, and second we allow each vertex to be given its own $`\mathrm{\Lambda }_i`$ = $`m_i`$, where $`m_i`$ is the mass of the on-shell charmed meson entering or leaving the vertex. Exchange graphs, or $`u`$-channels, are given similar form-factor structure. Let us call the function $`g(u)`$. Finally, contact graphs are given general Lorentz structure with several expansion coefficients. The idea is then to determine the coefficients (in terms of Lorentz invariants and $`h`$ and $`g`$) which result in a manifestly gauge invariant amplitude for the most general case of functions $`h`$ and $`g`$. Due to space limitation, we do not include here the full expression. More complete details will be published elsewhere. As an example of form-factor effects, we show in Fig. 3 the dissociation cross section for one of the pion channels ($`K`$ and $`\rho `$ cross sections are similarly affected). We are inclined to point to the results for $`\mathrm{\Lambda }_\alpha =m_\alpha `$ as likely being the most consistent, although means of constraining the form factors are under investigation. ## 5 Conclusions We have studied the spectral function for $`J/\psi `$ modified from its vacuum structure due to interactions with a gas of light mesons. Results suggest that the spectral function gets considerably modified due mostly to dissociation-type reactions. An environment much like the one we studied here is known to occur in ultrarelativistic heavy ion collisions. We stress that the approach we have taken is fully consistent in terms of conserving currents and respecting gauge invariance. Some issues will require further attention before we can take the next step of employing a spacetime model to ultimately compare with heavy ion data. Since the $`SU`$(4) symmetry is known to be only approximately valid, some coupling strengths are not completely pinned down. Also, the final results will depend fairly sensitively on what we do with the form factors. Work on these and related issues is in progress. Acknowledgment K.H. is supported by the National Science Foundation under Grant No. 9814247, and C.G. is supported by the Natural Science and Engineering Research Council of Canada and by the Fonds FCAR of the Quebec Government.
warning/0002/nlin0002016.html
ar5iv
text
# Growth and decay of discrete nonlinear Schrödinger breathers interacting with internal modes or standing-wave phonons ## I Introduction The concept of nonlinear self-localization is of importance for many physical phenomena, and has appeared in a number of different contexts since the pioneering work by Landau on the polaron problem in the 1930s. In recent years, much attention has been devoted to studies of spatially localized and time-periodic vibrational modes in anharmonic lattices (see e.g. , for recent reviews). The general existence of such modes, which have been termed discrete breathers, or intrinsic localized modes, as robust solutions to nonlinear (and in general non-integrable) lattice-equations was suggested in 1988 by Takeno et al . Later, their existence was rigorously proven under rather general conditions by MacKay and Aubry by considering the limit of uncoupled oscillators (the so called anticontinuous or anti-integrable limit). By means of the implicit function theorem, they showed that the trivial solution of a single-site localized vibration at the uncoupled limit could be continued into a localized breather solution for non-zero coupling between the oscillators, provided that the individual oscillators are anharmonic, and that no multiples of the breather frequency resonate with the bands of linear excitations (phonons). As was demonstrated first in , the ideas of the rigorous proof can be turned into an efficient numerical scheme to calculate breather solutions to any desired accuracy. Since the discrete breathers appear under very general conditions in anharmonic lattices and provide efficient means of energy localization, they have been proposed as candidates to explain experimentally observed localization of energy in many different physical areas, e.g. DNA dynamics . Although, from a fundamental and mathematical viewpoint, the existence theorems for discrete breathers provide an important cornerstone for understanding the dynamics of anharmonic lattices, it is probably of even larger physical importance to understand the behaviour of a system close to an exact breather solution. By linearizing the lattice-equations around the exact solution, one can obtain an approximate description of the dynamics of weakly perturbed breathers, and in particular the linear stability properties determining whether small perturbations will grow exponentially or not. It was shown in , that the simplest, single-site, breathers are generally linearly stable close to the uncoupled limit, and numerical investigations using standard Floquet analysis (see e.g. ) have shown that linearly stable breathers typically exist also for rather large values of the inter-site coupling. However, when considering time-scales large compared to the breather period, the mere linear stability of a breather does no longer guarantee the eternal existence of the breather in the presence of small perturbations, and there are still many questions remaining concerning the different mechanisms by which breathers may grow or decay, or possibly finally be destroyed. If the breathers have a finite life-time, the determination of this life-time is of large importance for understanding the role of breathers in real systems. It is the purpose of this paper to investigate in more detail some mechanisms for breather growth and decay in a simple model system, the discrete nonlinear Schrödinger (DNLS) equation. The DNLS equation is generic in the sense that it describes slowly (in time) varying modulational waves in discrete systems in a ’rotating-wave’ approximation (see e.g. ); however, due to its extra symmetry properties (see Sec. II) it exhibits some nongeneric features among discrete systems, e.g. exact quasiperiodic breathers . The single-site breathers of the DNLS equation are stationary states which are linearly stable for all inter-site coupling, and which reduce to the NLS soliton in the continuum limit (see e.g. ). An important application appears in nonlinear optics, where the single-site DNLS-breather describes a discrete spatial soliton in an array of weakly coupled waveguides ; recent experimental observations confirm the successful use of the DNLS-model in this context. Some recent numerical investigations have shown that DNLS-breathers can be spontaneously created from noisy backgrounds, in a similar manner as was previously observed for Klein-Gordon and FPU lattices. Typically, this spontaneous energy localization was observed to occur in two steps. In the first step, a large number of small breathers are created as a result of the modulational instability of travelling plane waves occurring for certain wave number regimes. The second step proceeds by inelastic collisions between the breathers, in which systematically the big breathers grow at expense of the smaller ones. Thus, the outcome will be a small number of large breathers, together with some remaining background of small-amplitude (phonon) oscillations. However, it can generally not be concluded from the numerical simulations that this is the true final state of the system, and actually long-time simulations for FPU-chains revealed also a third step, in which the interaction with the phonon oscillations leads to the final destruction of the breather and equipartition of energy. Thus, to elucidate the nature of the final states for typical initial conditions in anharmonic chains, it is necessary to obtain a better understanding of the mechanisms for interactions between breathers and small-amplitude perturbations. In this paper, we take the following approach. As initial state, we consider an exact single-site breather solution, and add a small perturbation corresponding to a time-periodic eigensolution to the equations of motion linearized around the breather. These solutions, which can be either localized or extended in space, constitute a complete set in which an arbitrary initial perturbation can be expanded. The localized solutions correspond to internal modes of the breather , while the excitation of an extended solution corresponds to a standing-wave (i.e., non-propagating) phonon interacting with the breather. In Sec. II we describe the model and outline the perturbational approach which forms the analytical backbone for the interpretation of the numerical results presented in Secs. III and IV. Sec. III discusses the long-time consequences of the interaction between the breather and its internal modes, while Sec. IV concerns the interaction between the breather and small-amplitude standing-wave phonons of different wave vectors. We will find that in both cases, a simple argument based on the conservation laws can be used to obtain a sufficient condition for breather growth. In Sec. V we discuss another type of mechanism for breather growth and destruction which becomes appreciable when the amplitude of the standing wave is non-negligible (and consequently the perturbational approach can be expected to fail), and which has its origin in the recently discovered oscillatory instabilities of the nonlinear standing-wave phonon themselves . Finally, we make some concluding remarks in Sec. VI. Concerning the numerical simulations of the dynamics presented in this paper, unless otherwise stated they always apply for a system of infinite size (finite size systems are considered only in Sec. V). The simulations have been performed either by using very large system sizes or by appending damping regions of various sizes to the boundaries; in all cases we have carefully checked that the boundary conditions have no essential influence on our results. ## II Model and framework for the perturbational approach ### A Model We consider the following form of the DNLS Hamiltonian with canonical conjugated variables $`\{i\psi _n\},\{\psi _n^{}\}`$: $$(\left\{i\psi _n\right\},\left\{\psi _n^{}\right\})=\underset{n}{}\left(C|\psi _{n+1}\psi _n|^2\frac{1}{2}|\psi _n|^4\right)\underset{n}{}_n.$$ (1) This yields the DNLS equation $$i\dot{\psi _n}=\frac{}{\psi _n^{}}=C(\psi _{n+1}+\psi _{n1}2\psi _n)|\psi _n|^2\psi _n,$$ (2) which, in addition to the Hamiltonian (1), also conserves the total excitation norm (or power in nonlinear optics applications), $$𝒩=\underset{n}{}|\psi _n|^2\underset{n}{}𝒩_n.$$ (3) The conservation laws for the norm and Hamiltonian are, through Noether’s theorem, related to the invariance of the DNLS equation (or, more precisely, of its corresponding action integral) under infinitesimal transformations in phase ($`\psi _n\psi _ne^{iϵ}`$) and time ($`tt+ϵ`$), respectively. Defining the ’norm density’ $`𝒩_n`$ and ’Hamiltonian density’ $`_n`$ as in Eqs. (3) resp. (1), the conservation laws can be expressed in terms of continuity equations as $$\frac{\mathrm{d}𝒩_n}{\mathrm{d}t}+\left(J_𝒩\right)_n\left(J_𝒩\right)_{n1}=0,$$ (4) $$\frac{\mathrm{d}_n}{\mathrm{d}t}+\left(J_{}\right)_n\left(J_{}\right)_{n1}=0,$$ (5) with the (norm) current density $$J_𝒩=2C\mathrm{Im}\left[\psi _n^{}\psi _{n+1}\right]$$ (6) and the Hamiltonian flux density $$J_{}=2C\mathrm{Re}\left[\dot{\psi }_{n+1}(\psi _{n+1}^{}\psi _n^{})\right],$$ (7) respectively. These conservation laws are discrete analogs to those existing for the continuous NLS equations with general nonlinearities (see e.g. ), however there is no discrete counterpart to the momentum conservation law since the discrete equation has no continuous translational symmetry in space. Furthermore, we note that the transformation $`CC`$ in (2) is equivalent to $`\psi _n(1)^ne^{i4Ct}\psi _n`$, and thus we will for the rest of this paper only consider $`C>0`$ without loss of generality. The single-site DNLS-breather is a stationary-state solution to Eq. (2) of the form $$\psi _n(t)=\varphi _n(\mathrm{\Lambda })e^{i\mathrm{\Lambda }t},$$ (8) where the time-independent shape $`\{\varphi _n\}`$ depends on the frequency $`\mathrm{\Lambda }`$ and is spatially localized with a single maximum at a lattice site. The breather exists for all $`\mathrm{\Lambda }/C>0`$; the limit $`C0`$ (or $`\mathrm{\Lambda }\mathrm{}`$ ) corresponding to the anticontinuous limit, where $`\{\varphi _n\}`$ is localized at a single lattice-site, while the limit $`\mathrm{\Lambda }/C0`$ corresponds to the continuous limit, where $`\{\varphi _n\}`$ approaches the NLS soliton. The single-site breather is a ground state solution to Eq. (2) in the sense that it minimizes the Hamiltonian (1) for a fixed value of the norm (3), i.e., $`\delta +\mathrm{\Lambda }\delta 𝒩=0`$, where the frequency $`\mathrm{\Lambda }`$ appears as the Lagrange multiplier (see e.g. ). The norm for the single-site breather, $`𝒩_\varphi `$, is known to be a monotonously increasing function of $`\mathrm{\Lambda }`$, while the Hamiltonian, $`_\varphi `$, is negative and monotonously decreasing (see e.g. ). From the minimization condition, these functions will be related as $$\frac{\mathrm{d}_\varphi }{\mathrm{d}\mathrm{\Lambda }}=\mathrm{\Lambda }\frac{\mathrm{d}𝒩_\varphi }{\mathrm{d}\mathrm{\Lambda }}.$$ (9) To describe the dynamics close to the breather (8), we introduce the following perturbation expansion: $$\psi _n(t)=\left\{\varphi _n+\lambda ϵ_n(t)+\lambda ^2\eta _n(t)+\lambda ^3\xi _n(t)+\lambda ^4\theta _n(t)+\mathrm{}\right\}e^{i{\scriptscriptstyle \mathrm{\Lambda }𝑑t}},$$ (10) where $`ϵ_n(0)`$ is the initial perturbation and $`\eta _n(0)=\xi _n(0)=\theta _n(0)=\mathrm{}=0`$. Thus, as for the usual stability analysis of stationary states (see e.g. ), the perturbation is applied in a frame rotating with the breather frequency $`\mathrm{\Lambda }`$. Substituting into Eq. (2) and identifying coefficients for consecutive powers of the small parameter $`\lambda `$ yields an infinite set of equations, which from 0th to 4th order read: $`\mathrm{\Lambda }\varphi _n+C(\varphi _{n+1}+\varphi _{n1}2\varphi _n)+|\varphi _n|^2\varphi _n=0`$ (11) $`(\mathrm{\Lambda })\{ϵ_n\}\{i\dot{ϵ}_n+C(ϵ_{n+1}+ϵ_{n1}2ϵ_n)+2|\varphi _n|^2ϵ_n+\varphi _n^2ϵ_n^{}\mathrm{\Lambda }ϵ_n\}=0`$ (12) $`(\mathrm{\Lambda })\{\eta _n\}=\varphi _n^{}ϵ_n^22\varphi _n|ϵ_n|^2`$ (13) $`(\mathrm{\Lambda })\{\xi _n\}=2\varphi _n^{}ϵ_n\eta _n2\varphi _n\left(ϵ_n^{}\eta _n+ϵ_n\eta _n^{}\right)|ϵ_n|^2ϵ_n`$ (14) $`(\mathrm{\Lambda })\{\theta _n\}=2\varphi _n\left(ϵ_n\xi _n^{}+ϵ_n^{}\xi _n+|\eta _n|^2\right)\varphi _n^{}\left(2ϵ_n\xi _n+\eta _n^2\right)ϵ_n^2\eta _n^{}2|ϵ_n|^2\eta _n,`$ (15) where the operator $`(\mathrm{\Lambda })`$ (which is linear over the field of real numbers) is defined from the first equality in Eq. (12). The 0th order equation (11) gives the breather shape $`\{\varphi _n\}`$ (which for the single-site breather can be assumed real and positive without loss of generality), while the 1st order equation (12) is the linearization of the DNLS equation around the breather. ### B Solutions to the linearized equations To obtain the solutions to the linearized equations (12), we proceed in a similar way as is usually done for continuous generalized NLS models (see e.g.) and introduce a substitution of the form $$ϵ_n(t)ϵ_n^{(r)}(t)+iϵ_n^{(i)}(t)=\frac{1}{2}a\left(U_n+W_n\right)e^{i\omega _pt}+\frac{1}{2}a^{}\left(U_n^{}W_n^{}\right)e^{i\omega _pt},$$ (16) so that $$ϵ_n^{(r)}(t)=\mathrm{Re}\left(ϵ_n(t)\right)=\mathrm{Re}\left(aU_ne^{i\omega _pt}\right),ϵ_n^{(i)}(t)=\mathrm{Im}\left(ϵ_n(t)\right)=\mathrm{Im}\left(aW_ne^{i\omega _pt}\right).$$ (17) Substituting (16) into Eq. (12) and assuming $`\varphi _n`$ real yields $`_0W_n`$ $``$ $`C(W_{n+1}+W_{n1}2W_n)\varphi _n^2W_n+\mathrm{\Lambda }W_n=\omega _pU_n`$ (18) $`_1U_n`$ $``$ $`C(U_{n+1}+U_{n1}2U_n)3\varphi _n^2U_n+\mathrm{\Lambda }U_n=\omega _pW_n,`$ (19) where the operators $`_0`$ and $`_1`$ are Hermitian. Thus, we can obtain the eigenfrequencies $`\omega _p`$ and the corresponding eigenvectors $`(\{U_n\},\{W_n\})`$ from matrix diagonalization, $$𝐌^{(0)}\left(\begin{array}{c}\{U_n\}\\ \{W_n\}\end{array}\right)\left(\begin{array}{c}0_0\\ _1\mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}\end{array}\right)\left(\begin{array}{c}\{U_n\}\\ \{W_n\}\end{array}\right)=\omega _p\left(\begin{array}{c}\{U_n\}\\ \{W_n\}\end{array}\right).$$ (20) To make connection to earlier work , we remark that the vector $`(\{ϵ_n^{(r)}\},\{ϵ_n^{(i)}\})=(\{U_n\},\{iW_n\})`$ is an eigenvector of the Floquet matrix with eigenvalue $`e^{i\omega _pT}`$, where the time-period $`T`$ here is arbitrary since the operators $`_0`$ and $`_1`$ are time-independent. (The symplectic Floquet matrix is $`e^{𝐌_FT}`$, where $`𝐌_F`$ is obtained from $`𝐌^{(0)}`$ by changing $`_1`$ into $`_1`$.) Thus, (16) is the linear combination of two complex conjugated Floquet eigensolutions which makes $`ϵ_n^{(r)}`$ and $`ϵ_n^{(i)}`$ real. For the single-site breather, all eigenvalues $`\omega _p`$ of $`𝐌^{(0)}`$ are always real, implying the linear stability of the breather for all parameter values $`\mathrm{\Lambda }/C>0`$ . Accordingly, we can also choose the eigenvectors $`(\{U_n\},\{W_n\})`$ of $`𝐌^{(0)}`$ to be real and normalized, in which case the phase of the amplitude $`a`$ describes the symmetry properties of the solution (16) under time reversal: choosing $`a`$ real yields a time-symmetric solution, $`ϵ_n(t)=ϵ_n^{}(t)`$, while choosing $`a`$ purely imaginary yields a time-antisymmetric solution, $`ϵ_n(t)=ϵ_n^{}(t)`$. For an infinite system, the spectrum of the (non-Hermitian) matrix $`𝐌^{(\mathrm{𝟎})}`$ can generally be divided into a continuous (phonon) part, corresponding to extended eigenvectors, and a point spectrum corresponding to localized eigenvectors. The phonon spectrum for any localized solution $`\{\varphi _n\}`$ is easily obtained from the limit $`|n|\mathrm{}`$, since the condition $`\varphi _n0`$ implies that the operators $`_0`$ and $`_1`$ become identical and Eqs. (18) - (19) reduce into two uncoupled equations for the linear combinations $`a_nU_n+W_n`$ resp. $`b_nU_nW_n`$. Assuming $`a_ne^{\pm iq_an}`$ and $`b_ne^{\pm iq_bn}`$, respectively, yields the dispersion relations $`\omega _p`$ $`=`$ $`\mathrm{\Lambda }2C(\mathrm{cos}q_a1),`$ (21) $`\omega _p`$ $`=`$ $`\mathrm{\Lambda }+2C(\mathrm{cos}q_b1),`$ (22) from Eqs. (18) - (19). Thus, the continuous spectrum of the matrix $`𝐌^{(\mathrm{𝟎})}`$ consists of two branches, symmetrically located around $`\omega _p=0`$, and since $`\mathrm{\Lambda }>0`$ for the single-site breather these two branches never overlap. Note also that two eigenvectors with eigenvalues $`\pm \omega _p`$ correspond to the same solution to Eq. (12) (changing the sign of $`\omega _p`$ in Eq. (16) is equivalent to changing $`U_nU_n^{}`$, $`W_nW_n^{}`$, $`aa^{}`$), and therefore it is enough to consider e.g. $`\omega _p>0`$, in which case $`b_n=U_nW_n`$ always vanishes exponentially as $`n\pm \mathrm{}`$. When $`\mathrm{\Lambda }/C`$ is not too large, the linear spectrum around the single-site breather contains also two pairs of nonzero isolated eigenvalues $`\omega _p`$, which correspond to the two internal modes of the breather . One of these modes is a spatially symmetric, ’breathing’, mode, while the other is a spatially antisymmetric, ’translational’ or ’pinning’ mode. Numerically, it has been found that the breathing mode exists for $`0<\mathrm{\Lambda }/C1.7`$, while the pinning mode exists for $`0<\mathrm{\Lambda }/C1.1`$. The numerically calculated internal mode frequencies $`\omega _p`$ as a function of breather frequency $`\mathrm{\Lambda }`$ are shown in Fig. 1. Note that as $`\mathrm{\Lambda }/C0`$, the breathing mode frequency approaches the lower edge of the phonon band (but always stays outside the band ), while the pinning mode frequency approaches zero (but always stays nonzero). This is consistent with the fact that the soliton solution of the continuous NLS equation has no breathing mode (due to its exact integrability), and has a translational mode with zero frequency due to the translational symmetry of the NLS equation. To obtain a complete set of solutions to Eq. (12) in which an arbitrary initial perturbation $`ϵ_n(0)`$ can be expanded, we must include also the zero-frequency solutions, which generally can be written as a superposition of two fundamental modes. One of these modes (’phase mode’ ) is the solution $`W_n=\varphi _n`$ to the homogeneous equation (18), $`_0W_n=0`$. The corresponding perturbation $`ϵ_n=i\varphi _n`$ describes a rotation of the overall phase of the breather. The second mode (’growth mode’ ) is obtained by solving the inhomogeneous equation $`_1U_n=\varphi _n`$ , which has the solution $`U_n=\varphi _n/\mathrm{\Lambda }`$. The corresponding solution to Eq. (12) is $`ϵ_n=\varphi _n/\mathrm{\Lambda }+i\varphi _nt`$, and corresponds to a time-linear growth of the perturbation representing a small change in the breather frequency. Although the set of eigensolutions (16) together with the two zero-frequency modes forms a basis for the space of solutions to Eq. (12) (there are no bifurcations, which could result in additional ’marginal modes’ with time-linear growth at degenerate eigenvalues), this basis is in general not orthogonal using the ordinary scalar product, and typically there is a considerable overlap between the solution corresponding to the internal breathing mode and the zero-frequency modes. However, in analogy with e.g. , we can define a ’pseudo-scalar’ product between any two vectors $`(\{U_n^{(1)}\},\{W_n^{(1)}\})`$ and $`(\{U_n^{(2)}\},\{W_n^{(2)}\})`$ by $$\underset{n}{}\left(U_n^{(1)}W_n^{(2)^{}}+W_n^{(1)}U_n^{(2)^{}}\right).$$ (23) This product is formally not a true scalar product, since for the general case the product of a vector with itself as defined by (23) is not necessarily positive. However, when $`(\{U_n\},\{W_n\})`$ is a real eigenvector of $`𝐌^{(\mathrm{𝟎})}`$ we have $`_nU_nW_n=\frac{1}{\omega _p}_nW_n_0W_n`$ from (18), and the operator $`_0`$ is positive definite for all $`W_n\varphi _n`$ . With this product, it follows from (18) - (19) that all eigensolutions with different (real) eigenfrequencies $`\omega _p`$ are ’orthogonal’ in the sense that $$\left(\omega _p^{(1)}\omega _p^{(2)}\right)\underset{n}{}\left(U_n^{(1)}W_n^{(2)^{}}+W_n^{(1)}U_n^{(2)^{}}\right)=0,$$ (24) and the only nonzero product involving the zero-frequency modes is the cross-product between the phase mode and the growth mode : $$\underset{n}{}\varphi _n\frac{\varphi _n}{\mathrm{\Lambda }}=\frac{1}{2}\frac{\mathrm{d}𝒩_\varphi }{\mathrm{d}\mathrm{\Lambda }}>0,$$ (25) where $`𝒩_\varphi `$ is the norm (3) of the breather with frequency $`\mathrm{\Lambda }`$. We also remark that, since the product (23) multiplied by a factor $`i`$ is just the symplectic product between the two vectors $`(\{ϵ_n^{(r)(1)}\},\{ϵ_n^{(i)(1)}\})`$ and $`(\{ϵ_n^{(r)(2)}\},\{ϵ_n^{(i)(2)}\})`$, the sign of the product of an eigenvector with itself can be interpreted as the negative of the Krein signature of the corresponding pair of Floquet eigenvalues . ### C Strategy for the perturbational approach We now consider as initial state a single-site breather perturbed in the direction of a single eigenmode (16) (localized or extended) of the linearized equations (12), and wish to describe qualitatively the long-time effects of this perturbation using the expansion (10). In general, taking into account terms up to order $`p`$ in this expansion yields a solution to the DNLS equation which is correct to $`O(\lambda ^{p+1})`$, i.e., for long but finite time-scales for small initial perturbations (note that the expansion parameter $`\lambda `$ plays the same role as the mode amplitude $`a`$.). As is wellknown however, this kind of expansion in general diverges due to resonances between solutions to the homogeneous equation (12) and the inhomogeneous terms appearing in the right-hand sides of Eqs. (13)-(15) and the corresponding higher-order equations. A resonance with a solution belonging to the continuous spectrum results in a bounded but nonlocalized solution corresponding to outgoing radiation, while a resonance with an eigenfunction belonging to the discrete spectrum gives a spatially localized response which diverges linearly in time. However, up to any finite order these divergences can be systematically removed by allowing a slow time dependence of the independent variables, which in our case are taken to be the mode amplitude $`a`$ and the breather frequency $`\mathrm{\Lambda }`$. This procedure adds additional terms to the equations, which can be tuned so that the divergent parts of the response disappear. In other words, these two quantities are used as collective variables which, together with the outgoing radiation fields, are expected to describe the main features of the asymptotic dynamics if the initial perturbation is sufficiently small. The second order correction is given by the inhomogeneous Eq. (13), which with the substitution (16) becomes (choosing $`\varphi _n`$, $`U_n`$, and $`W_n`$ real without loss of generality): $`(\mathrm{\Lambda })\{\eta _n\}={\displaystyle \frac{1}{2}}\varphi _n[|a|^2(3U_n^2+W_n^2)+(3U_n^2W_n^2)\mathrm{Re}\left(a^2e^{2i\omega _pt}\right)`$ (26) $`+2iU_nW_n\mathrm{Im}\left(a^2e^{2i\omega _pt}\right)].`$ (27) Thus, the right-hand side contains one static part and one part involving the frequencies $`\pm 2\omega _p`$. It acts as a periodic force with frequencies 0 and $`2\omega _p`$, and since all terms contain the factor $`\varphi _n`$ this force is localized at the breather region. The response to this force will remain bounded and localized unless the corresponding homogeneous equation (12) has a solution with frequency 0 resp. $`2\omega _p`$ which is non-orthogonal to the corresponding part of the right-hand side in (26). As will be shown in Sec. III B, a non-zero overlap between the static part of (26) and the zero-frequency solutions of (12) yields a (time-independent) shift of the breather frequency. Moreover, if $`\mathrm{\Lambda }<2|\omega _p|<\mathrm{\Lambda }+4C`$, so that $`2\omega _p`$ is inside the phonon band of the homogeneous equation, a resonance will generally occur, resulting in radiation with frequency $`2\omega _p`$ emitted from the breather region. The strength of the radiation field is determined by the (generally nonzero) overlap between the $`2\omega _p`$-part of (26) and the corresponding homogeneous solution (see Sec. III B). In a similar way, we obtain that the right-hand-side of the third order equation (14) contains the frequencies $`\omega _p`$ and $`3\omega _p`$, the fourth order equation (15) contains the frequencies 0, $`2\omega _p`$, and $`4\omega _p`$, and in general the pth order equation contains as its highest harmonic the frequency $`p\omega _p`$. Accordingly, we conclude that if $$\mathrm{\Lambda }<p|\omega _p|<\mathrm{\Lambda }+4C,$$ (28) so that $`p\omega _p`$ belongs to the phonon band, the perturbed breather will radiate to pth order. The consequences of this radiation for the breather itself will be discussed in Secs. III and IV for the cases of localized and extended perturbations $`\{ϵ_n\}`$, respectively. ## III Breather interacting with internal modes With the initial perturbation $`ϵ_n(0)`$ of the single-site breather corresponding to a spatially localized eigenmode of the linearized equations (12), it is clear from the discussion in Sec. II C that higher order radiation always will be created, since for any internal mode frequency $`\omega _p`$ there is always an integer $`p`$ such that $`p\omega _p`$ belongs to the phonon band and (28) is fulfilled. Moreover, from the numerical results presented in Fig. 1 we find that the spatially symmetric breathing mode always radiates to second order, since (28) always is fulfilled for $`p=2`$, while the antisymmetric pinning mode radiates to second order only when $`\mathrm{\Lambda }/C0.480`$. Thus, due to this radiation from the breather the total norm contained in any finite region around the breather(i.e., the total norm of breather + internal mode) will always decrease with time. However, the main concern here is the long-time effect of the internal-mode excitation on the breather itself, and thus we must investigate whether there will also be some transfer of energy between the breather and its internal mode. We will first (Sec. III A) show results from direct numerical integration of Eq. (2); then we will give two alternative approaches to the analytical interpretation of these results based on the higher-order equations (13)-(15) (Sec. III B) resp. the conservation laws (4)-(5) (Sec. III C). ### A Numerical simulations In Fig. 2, we show a typical example on the long-time evolution of a breather when the initial perturbation is taken in the direction of its internal breathing mode. As is seen from Fig. 2 (a), the amplitude of the breathing mode decays slowly with time as a consequence of the losses due to generation of second order radiation, and a careful study of its envelope $`|a(t)|`$ indicates that it decays as $$|a(t)|\frac{|a(0)|}{\sqrt{1+\gamma |a(0)|^2t}},$$ (29) where $`\gamma >0`$ is a constant. This is consistent with a similar result obtained for the continuum NLS equation with generalized (non-cubic) nonlinearity ; the analytical motivation for this result (which is analogous to that of the continuum model given in Ref. ) is given in the following subsections. However, the main result of this section is illustrated in Fig. 2 (b) and (c). Fig. 2 (b) is obtained by calculating the time-average of the central-site intensity as $$|\psi _{n_0}|^2_{t=t_K}=\frac{1}{K}\underset{k=1}{\overset{K}{}}|\psi _{n_0}(t_k)|^2,$$ (30) where $`t_k`$ is a set of closely spaced time-instants. It is clear that the interaction between the breather and its internal mode asymptotically leads to an increase of the average peak intensity, i.e., to breather growth. The same phenomenon is illustrated also in Fig. 2 (c), where we have plotted the difference between the instantaneous breather frequency calculated at time $`t`$, $`\mathrm{\Lambda }(t)`$, and the frequency of the unperturbed breather $`\mathrm{\Lambda }_0`$. From this figure, we can also conclude that there are two different mechanisms causing the shift of breather frequency. Firstly, there is an initial (almost instantaneous) rather large frequency shift, which can be interpreted as an adaption of the initially perturbed breather to the breather which is ’closest’ to the initial condition. As is shown below in Sec. III B, this time-independent frequency shift, which is observed to be always positive for the breathing mode, is a consequence of the overlap between the static part of the right-hand side of the second-order equation (26) and the zero-frequency modes. Secondly, there is the slow, continuous increase of the breather frequency which corresponds to the slow increase of $`|\psi _{n_0}|^2_t`$ in Fig. 2 (b), indicating a continuous transfer of norm from the internal mode to the breather. It is described by the static part of the right-hand side of the fourth-order equation (15) (see Sec. III B). When the initial perturbation of the breather is taken in the direction of its internal pinning mode we observe, just as for the breathing mode, that the breather-internal mode interaction asymptotically always leads to breather growth. An example is shown in Fig. 3, where the parameter values have been chosen so that the lowest harmonic that enters the phonon band is $`3\omega _p`$ ($`\mathrm{\Lambda }/C=0.45<0.480`$), We observe two qualitative differences compared to the case with breathing mode excitation. Firstly, since in this case the first phonon resonance occurs only in the third order equation (14), the decay of the internal mode amplitude will be slower, and a good fit is obtained by $`|a(t)||a(0)|\left(1+\gamma |a(0)|^4t\right)^{1/4}`$. This agrees with the general result when $`p\omega _p`$ is the lowest harmonic that enters the phonon band obtained for the continuum generalized NLS equation in Ref. ; the derivation of the corresponding result for the discrete case (see Eq. (58)) is given in Sec. III C. As a consequence of the slower decay of the internal mode amplitude, the breather growth will also be slower when $`p>2`$, as can be seen from Fig. 3 (c) by comparing the time-scales with those of Fig. 2. Secondly, the initial shift of the breather frequency for a pinning mode excitation is always much smaller than for the breathing mode excitation (also when $`2\omega _p`$ is in the phonon band), and is when $`\mathrm{\Lambda }/C0.55`$ also observed to be negative. The explanation for this is given in Sec. III B. However, it is important to stress that also in the cases where the initial frequency shift is negative, we find that the continuous breather growth always will give an asymptotic frequency shift which is positive. ### B Analysis of higher-order equations Here, we will analyze the higher-order equations (13)-(15) by making use of the strategy of systematically removing the appearing divergent parts as outlined in Sec. II C (in analogy with the treatment of the continuous NLS-type equations in e.g. Refs. ). First, we show how the dominating contribution to the time-independent frequency shift observed in the numerical simulations above can be calculated from the static part of the right-hand side of the second-order equation (13). This frequency shift can be explicitly taken into account by replacing $`\mathrm{\Lambda }`$ in Eq. (10) with $`\mathrm{\Lambda }_0+\lambda ^2\mathrm{\Lambda }_2`$, where $`\mathrm{\Lambda }_0`$ is the unperturbed breather frequency and $`\mathrm{\Lambda }_2`$ the second-order shift to be determined. This implies that the additional term $`\mathrm{\Lambda }_2\varphi _n`$ will be added to the right-hand side of Eq. (26). Writing the response to the static part of (26) as $`\eta _n^{(s)}=|a|^2(u_n^{(s)}+iw_n^{(s)})`$ with real $`u_n^{(s)}`$ and $`w_n^{(s)}`$ then yields $$𝐌^{(0)}\left(\begin{array}{c}\left\{u_n^{(s)}\right\}\\ \left\{w_n^{(s)}\right\}\end{array}\right)=\left(\begin{array}{c}\left\{0\right\}\\ \left\{\varphi _n\left(\frac{\mathrm{\Lambda }_2}{|a|^2}\frac{1}{2}\left(3U_n^2+W_n^2\right)\right)\right\}\end{array}\right),$$ (31) with $`𝐌^{(0)}`$ as defined by Eq. (20). If the expansion of the right-hand side of (31) in the complete set of vectors consisting of the eigenvectors of $`𝐌^{(0)}`$ (including the phase mode) and the growth mode contains some component on either of the two zero-frequency modes, the response $`\eta _n^{(s)}`$ will not remain bounded but diverge linearly with time. Thus, in order to remove this divergency, the frequency shift $`\mathrm{\Lambda }_2`$ must be chosen so that both these components are identically zero. The component on the growth mode is trivially zero, while the component on the phase mode is obtained by applying the pseudo-scalar product (23) with the vector corresponding to the growth mode and using (25). Demanding this component to be zero yields $$\mathrm{\Lambda }_2=\frac{|a|^2}{\frac{\mathrm{d}𝒩_\varphi }{\mathrm{d}\mathrm{\Lambda }}}\underset{n}{}\varphi _n\frac{\varphi _n}{\mathrm{\Lambda }}\left(3U_n^2+W_n^2\right).$$ (32) This is typically positive for the breathing mode, since the dominating contribution to the sum in (32) comes from the central site $`n_0`$, and $`\frac{\varphi _{n_0}}{\mathrm{\Lambda }}`$ is always positive. For the spatially antisymmetric pinning mode, there is no contribution to this sum from the central site, since $`U_{n_0}`$ and $`W_{n_0}`$ are zero. The change from a positive to a negative frequency shift when increasing $`\mathrm{\Lambda }`$ in this case is related to a qualitative change of the nature of the growth mode $`\frac{\varphi _n}{\mathrm{\Lambda }}`$: for $`\mathrm{\Lambda }0.55`$ we find that $`\frac{\varphi _n}{\mathrm{\Lambda }}<0`$ for all $`nn_0`$, so that all terms in the sum in Eq. (32) are negative, while $`\frac{\varphi _n}{\mathrm{\Lambda }}`$ becomes positive also for sites in the neighborhood of $`n_0`$ for smaller $`\mathrm{\Lambda }`$. For the rest of the analysis in this subsection, we assume for calculational simplicity that the internal mode frequency is such that $`2\omega _p`$ is inside the phonon band (and thus it is not applicable for the pinning mode excitation when $`\mathrm{\Lambda }/C0.480`$). Then, the non-static part of the right-hand side of the second order equation (26) will generally give rise to a non-localized response, which can be written in the form $`\eta _n^{(rad)}=\frac{1}{2}a^2\left(u_n^{(2)}+w_n^{(2)}\right)e^{2i\omega _pt}+\frac{1}{2}a^2\left(u_n^{(2)^{}}w_n^{(2)^{}}\right)e^{2i\omega _pt}`$. This response corresponds to the radiation field going out from the breather region, and since the right-hand side of (26) is spatially localized and symmetric, this field should asymptotically correspond to two identical linear waves propagating to the left (right) for $`n\mathrm{}`$ ($`+\mathrm{}`$). Thus, the boundary conditions can be written as $$u_n^{(2)},w_n^{(2)}r_2e^{\pm iq_2n},n\pm \mathrm{},$$ (33) with $`q_2=\mathrm{arccos}\left(1\frac{2\omega _p\mathrm{\Lambda }}{2C}\right)`$ according to (21). Defining for general $`\omega `$ the matrix $`𝐌^{(\omega )}`$ (cf. (20)) as $$𝐌^{(\omega )}\left(\begin{array}{c}\omega _0\hfill \\ _1\omega \hfill \end{array}\right),$$ (34) the functions $`u_n^{(2)}`$ and $`w_n^{(2)}`$ are seen from (26) to be determined by $$𝐌^{(2\omega _p)}\left(\begin{array}{c}\{u_n^{(2)}\}\\ \{w_n^{(2)}\}\end{array}\right)=\frac{\varphi _n}{2}\left(\begin{array}{c}\{2U_nW_n\}\\ \{3U_n^2W_n^2\}\end{array}\right).$$ (35) Since for general $`\omega `$ every eigenvector of $`𝐌^{(\omega )}`$ with eigenvalue $`\mu `$ is also an eigenvector of $`𝐌^{(0)}`$ with eigenvalue $`\mu +\omega `$, the right-hand side can be expanded on the basis of eigenvectors of $`𝐌^{(0)}`$ (including the zero-frequency modes). The strength of the radiation field is then given by the expansion coefficient for the (continuous spectrum) eigenvector of $`𝐌^{(0)}`$ with eigenvalue $`2\omega _p`$, since this corresponds to the eigenvalue $`\mu =0`$ of $`𝐌^{(2\omega _p)}`$, and thus a spatially non-bounded response in (35). Using the orthogonality relation (24), this coefficient is simply the ’overlap’ between the right-hand side of (35) and the eigenvector of $`𝐌^{(0)}`$ with eigenvalue $`2\omega _p`$ calculated with the pseudo-scalar product (23), which is generally nonzero. Next, we show how the dominating contribution to the decay of the internal mode amplitude as given by Eq. (29) is obtained from the condition that $`\xi _n`$ in the third-order equation (14) should remain bounded. To this end, we assume a slow time-dependence of the internal mode amplitude of the form $`a=a(\lambda ^2t)`$, and consider the response to the terms with frequency $`\omega _p`$ in the right-hand side of Eq. (14) ’corrected’ by the additional terms $`(i\dot{a}+\mathrm{\Lambda }_2a)(U_n+W_n)e^{i\omega _pt}+(i\dot{a}^{}+\mathrm{\Lambda }_2a^{})(U_nW_n)e^{i\omega _pt}`$ appearing as a consequence of including the time-dependence of $`a`$ and the second-order frequency shift $`\mathrm{\Lambda }_2`$ from Eq. (32) in the perturbation expansion (10). Writing the response to this part as $`\xi _n^{(\omega _p)}=\frac{1}{2}|a|^2a\left(u_n^{(3)}+w_n^{(3)}\right)e^{i\omega _pt}+\frac{1}{2}|a|^2a^{}\left(u_n^{(3)^{}}w_n^{(3)^{}}\right)e^{i\omega _pt}`$ yields $`𝐌^{(\omega _p)}\left(\begin{array}{c}\{u_n^{(3)}\}\\ \{w_n^{(3)}\}\end{array}\right)`$ (38) $`=`$ $`\left(\begin{array}{c}\{i\frac{\dot{a}}{|a|^2a}U_n+\varphi _n\left(U_nw_n^{(2)}W_nu_n^{(2)}+2W_nu_n^{(s)}\right)+\frac{1}{4}\left(U_n^2W_n+3W_n^3\right)\frac{\mathrm{\Lambda }_2}{|a|^2}W_n\}\\ \{i\frac{\dot{a}}{|a|^2a}W_n+\varphi _n\left(3U_nu_n^{(2)}+W_nw_n^{(2)}+6U_nu_n^{(s)}\right)+\frac{1}{4}\left(3U_n^3+U_nW_n^2\right)\frac{\mathrm{\Lambda }_2}{|a|^2}U_n\}\end{array}\right).`$ (41) A bounded response for $`\xi _n^{(\omega _p)}`$ exists only if the vector on the right-hand side of (38) has no component in the direction of the internal mode eigenvector $`(\{U_n\},\{W_n\})`$, since this is the eigenvector corresponding to the eigenvalue zero of $`𝐌^{(\omega _p)}`$. Using the orthogonality relation (24), this component is obtained by application of the product (23) with the vector $`(\{U_n\},\{W_n\})`$, and the condition that this component must be zero determines the time-evolution of $`a`$. Considering only the absolute value $`|a|^2`$, the resulting equation has the form $`\frac{\mathrm{d}}{\mathrm{d}t}\left(|a|^2\right)+\gamma |a|^4=0`$, which has the desired solution (29). The constant $`\gamma `$ is given by $$\gamma =\frac{_n\varphi _n\left[2U_nW_n\mathrm{Im}\left(w_n^{(2)}\right)+\left(3U_n^2W_n^2\right)\mathrm{Im}\left(u_n^{(2)}\right)\right]}{_nU_nW_n}=\frac{8C\omega _p|r_2|^2\mathrm{sin}q_2}{_nW_n_0W_n}>0,$$ (42) where the second equality is obtained using Eqs. (18), (33) and (35), and the positivity of $`\gamma `$ follows from the fact that, as mentioned in Sec. II B, the operator $`_0`$ is positive definite for all $`W_n\varphi _n`$ . Finally, we show how the continuous increase of the breather frequency appears from the divergent response to the static part of the right-hand side of the fourth-order equation (15). In its unmodified form, this part is given by $`R_n^{(4s)}{\displaystyle \frac{1}{2}}\varphi _n|a|^4\{6U_n\mathrm{Re}\left(u_n^{(3)}\right)+2W_n\mathrm{Re}\left(w_n^{(3)}\right)+2iU_n\mathrm{Im}\left(w_n^{(3)}\right)2iW_n\mathrm{Im}\left(u_n^{(3)}\right)`$ (43) $`+6(u_n^{(s)})^2+3|u_n^{(2)}|^2+|w_n^{(2)}|^2+2i\mathrm{Im}\left(u_n^{(2)}w_n^{(2)}\right)\}`$ (44) $`{\displaystyle \frac{1}{4}}|a|^4\{2(3U_n^2+W_n^2)u_n^{(s)}+(3U_n^2W_n^2)\mathrm{Re}\left(u_n^{(2)}\right)+2U_nW_n\mathrm{Re}\left(w_n^{(2)}\right)`$ (45) $`2iU_nW_n\mathrm{Im}\left(u_n^{(2)}\right)+i(U_n^23W_n^2)\mathrm{Im}\left(w_n^{(2)}\right)\}.`$ (46) Now, it is clear from (32) and (29) that the time-dependence of $`a`$ will induce a time-dependence of the second-order frequency shift $`\mathrm{\Lambda }_2`$ of the form $`\mathrm{\Lambda }_2(\lambda ^2t)`$, so that we can express the total breather frequency up to order $`\lambda ^4`$ as $`\mathrm{\Lambda }(t)=\mathrm{\Lambda }_0+\lambda ^2\mathrm{\Lambda }_2(\lambda ^2t)+\lambda ^4\mathrm{\Lambda }_4`$, where a 4th order correction $`\mathrm{\Lambda }_4`$ has also been included. Then, we must also take into account the time-dependence of the breather shape $`\varphi _n`$ by writing $`\varphi _n(\mathrm{\Lambda }(t))`$. As a consequence, the term $`i\frac{\varphi _n}{\mathrm{\Lambda }}\dot{\mathrm{\Lambda }}_2+\mathrm{\Lambda }_4\varphi _n`$ will be added to the expression (46) for $`R_n^{(4s)}`$ in the right-hand side of (15), and writing the response to this total force as $`\theta _n^{(s)}=|a|^4\left(u_n^{(4s)}+iw_n^{(4s)}\right)`$ with real $`u_n^{(4s)}`$ and $`w_n^{(4s)}`$ yields $$𝐌^{(0)}\left(\begin{array}{c}\{u_n^{(4s)}\}\\ \{w_n^{(4s)}\}\end{array}\right)=\frac{1}{|a|^4}\left(\begin{array}{c}\left\{\frac{\varphi _n}{\mathrm{\Lambda }}\dot{\mathrm{\Lambda }}_2\mathrm{Im}\left(R_n^{(4s)}\right)\right\}\\ \left\{\mathrm{\Lambda }_4\varphi _n\mathrm{Re}\left(R_n^{(4s)}\right)\right\}\end{array}\right).$$ (47) The response $`\theta _n^{(s)}`$ will be bounded in time only if the right-hand side of (47) has no component either on the growth mode or on the phase mode, which gives two conditions for the determination of $`\dot{\mathrm{\Lambda }}_2`$ and $`\mathrm{\Lambda }_4`$. Using (23)-(25), we obtain by demanding the expansion coefficient for the growth mode to be zero: $`\dot{\mathrm{\Lambda }}_2={\displaystyle \frac{|a|^4}{2\frac{\mathrm{d}𝒩_\varphi }{\mathrm{d}\mathrm{\Lambda }}}}{\displaystyle \underset{n}{}}\{4\varphi _n^2[W_n\mathrm{Im}\left(u_n^{(3)}\right)U_n\mathrm{Im}\left(w_n^{(3)}\right)+\mathrm{Im}\left(u_n^{(2)}w_n^{(2)}\right)]`$ (48) $`+\varphi _n[2U_nW_n\mathrm{Im}\left(u_n^{(2)}\right)+(3W_n^2U_n^2)\mathrm{Im}\left(w_n^{(2)}\right)]\}.`$ (49) (Similarly, $`\mathrm{\Lambda }_4`$ is obtained by demanding the component on the phase mode to be zero.) Thus, the dominating contribution to the frequency growth should be of order $`\dot{\mathrm{\Lambda }}_2|a|^4`$, so that with the approximate time-dependence (29) of the internal-mode amplitude we obtain qualitatively $$\mathrm{\Lambda }(t)\mathrm{\Lambda }_0|a(0)|^2\left(C_1C_2\frac{1}{1+\gamma |a(0)|^2t}\right),$$ (50) which is in good agreement with the numerically observed time-dependence of the frequency shift as shown in Fig. 2 (c). However, the positivity of $`\dot{\mathrm{\Lambda }}_2`$ is not easily seen from the expression (48), and therefore we will in the following subsection derive an alternative expression from which the positivity follows immediately, using the conservation laws for the norm resp. Hamiltonian. ### C Approach using conservation laws We consider first the conservation law (4) for the total norm (3) contained in any large but finite region around the breather. Averaging over a time-interval $`[t,t+2\pi /\omega _p]`$ and using Eqs. (10), (16), we can write the time-averaged norm to second order in $`|a|`$ (putting $`\lambda =1`$) as $$𝒩_t(t)\underset{n}{}\left[\varphi _n^2(\mathrm{\Lambda }(t))+\frac{|a(t)|^2}{2}\left(U_n^2+W_n^2\right)\right].$$ (51) (Note that there will be no contribution at order 2 from the static second-order correction $`u_n^{(s)}`$, since the renormalization of the breather frequency $`\mathrm{\Lambda }`$ according to (32) yields $`_n\varphi _nu_n^{(s)}=0`$.) In the case when $`2\omega _p`$ belongs to the phonon band, we obtain the following balance equation, which is correct up to order $`|a|^4`$: $$\frac{\mathrm{d}𝒩_t}{\mathrm{d}t}=\frac{\mathrm{d}𝒩_\varphi }{\mathrm{d}\mathrm{\Lambda }}\dot{\mathrm{\Lambda }}+\frac{1}{2}\underset{n}{}\left(U_n^2+W_n^2\right)\frac{\mathrm{d}|a|^2}{\mathrm{d}t}=J_𝒩(\mathrm{})J_𝒩(+\mathrm{})=4C|a|^4|r_2|^2\mathrm{sin}q_2,$$ (52) where we have used Eqs. (4), (6) and (33), and $`q_2`$ is as defined below Eq. (33). Similarly, we can use the conservation law (5) for the Hamiltonian (1) together with the general expression for the Hamiltonian flux density (7) for a small-amplitude plane wave $`\psi _n=Ae^{i(qn\omega (q)t)}`$, $$J_{}=2|A|^2C\omega (q)\mathrm{sin}q,$$ (53) to write the balance equation for the total time-averaged Hamiltonian in the same region for the case of second order radiation: $$\frac{\mathrm{d}_t}{\mathrm{d}t}=\frac{\mathrm{d}_\varphi }{\mathrm{d}\mathrm{\Lambda }}\dot{\mathrm{\Lambda }}+\frac{_t}{|a|^2}\frac{\mathrm{d}|a|^2}{\mathrm{d}t}=J_{}(\mathrm{})J_{}(+\mathrm{})=4C|a|^4|r_2|^2(2\omega _p\mathrm{\Lambda })\mathrm{sin}q_2,$$ (54) which is also correct to order $`|a|^4`$. The lowest order contribution to the derivative $`\frac{_t}{|a|^2}`$ can be obtained using the first equality in the equation of motion (2) and its complex conjugate as follows: $`{\displaystyle \frac{_t}{|a|^2}}={\displaystyle \frac{1}{a^{}}}{\displaystyle \frac{_t}{a}}={\displaystyle \frac{1}{a^{}}}{\displaystyle \underset{n}{}}{\displaystyle \frac{}{\psi _n}}{\displaystyle \frac{\psi _n}{a}}+{\displaystyle \frac{}{\psi _n^{}}}{\displaystyle \frac{\psi _n^{}}{a}}_t`$ (55) $`={\displaystyle \frac{1}{2}}\mathrm{\Lambda }{\displaystyle \underset{n}{}}\left(U_n^2+W_n^2\right)+\omega _p{\displaystyle \underset{n}{}}U_nW_n+𝒪(|a|^2),`$ (56) which is always negative for an internal mode excitation since $`|\omega _p|<\mathrm{\Lambda }`$. Thus, we can combine the two balance equations (52) and (54), and using Eqs. (9) and (56) we obtain the expression (29) for $`|a(t)|`$ with $`\gamma `$ as in (42), together with the following expression for the frequency growth rate from which its positivity is immediately seen: $$\dot{\mathrm{\Lambda }}=\frac{4C|a|^4|r_2|^2\mathrm{sin}q_2}{\frac{\mathrm{d}𝒩_\varphi }{\mathrm{d}\mathrm{\Lambda }}}\left(\frac{_n\left(U_n^2+W_n^2\right)}{_nU_nW_n}1\right)>0.$$ (57) This approach also has the advantage that it is easily generalized to the case where the lowest harmonic that enters the phonon band is $`p\omega _p`$ with $`p>2`$, i.e., for the pinning mode excitation when $`\mathrm{\Lambda }/C0.480`$. Then, we can write the boundary conditions at the infinities to lowest order in $`a`$ as $`\psi _na^pr_pe^{i[\pm q_pn(p\omega _p\mathrm{\Lambda })t]}`$, $`n\pm \mathrm{}`$, where $`r_p1`$ and $`q_p=\mathrm{arccos}\left(\frac{\mathrm{\Lambda }p\omega _p}{2C}+1\right)`$ according to (21). Consequently, we can proceed exactly as above, writing down the balance equations for the norm and Hamiltonian to order $`|a|^{2p}`$ just by modifying the right-hand sides of Eqs. (52) resp. (54) by replacing $`|a|^4`$ with $`|a|^{2p}`$, $`r_2`$ with $`r_p`$, $`q_2`$ with $`q_p`$, and $`2\omega _p`$ with $`p\omega _p`$. Combining the balance equations yields the following general expression for the time-dependence of the internal mode amplitude, $$|a(t)|=\frac{|a(0)|}{\left[1+(p1)\gamma _p|a(0)|^{2p2}t\right]^{1/(2p2)}},\gamma _p=\frac{4pC|r_p|^2\mathrm{sin}q_p}{_nU_nW_n}>0,$$ (58) which is the analog to the expression obtained with similar arguments in for the continuum NLS models. And most importantly, we obtain a general expression for the breather frequency growth rate which is positive for all $`p`$: $$\dot{\mathrm{\Lambda }}=\frac{4C|a|^{2p}|r_p|^2\mathrm{sin}q_p}{\frac{\mathrm{d}𝒩_\varphi }{\mathrm{d}\mathrm{\Lambda }}}\left(\frac{p_n\left(U_n^2+W_n^2\right)}{2_nU_nW_n}1\right)>0.$$ (59) Thus, integrating Eq. (59) using the time-dependence (58) of the internal-mode amplitude, we obtain that the dominating contribution to the frequency growth generally can be written qualitatively as $$\mathrm{\Lambda }(t)\mathrm{\Lambda }_0|a(0)|^2\left(C_1C_2\frac{1}{1+(p1)\gamma _p|a(0)|^{2p2}t}\right)^{1/(p1)}.$$ (60) Let us finally point out that the approach used in this section provides a simple physical interpretation of why the interaction between the breather and its internal mode in addition to generate radiation also should lead to breather growth. Since the expression (53) for the Hamiltonian flux density of a small-amplitude plane wave is positive when $`\omega (q)`$ and $`q`$ have the same sign, the Hamiltonian energy for a plane wave always propagates in the same direction as the wave itself. Thus, the second (or higher) order radiation emitted from the breather region will always carry away a positive amount of the Hamiltonian energy, or, equivalently, negative Hamiltonian energy will flow into the breather region. Moreover, from Eq. (56) it is clear that the contribution to the Hamiltonian from the internal mode always is negative and a monotonously decreasing function of its amplitude, and thus the decay of the internal mode would cause an increase of the Hamiltonian in the breather region. Consequently, since the Hamiltonian of the pure breather is a monotonously decreasing function of the breather frequency, the breather should grow so that the total Hamiltonian of (breather+internal mode) decreases. A similar mechanism was recently found to cause soliton growth in the parametrically driven continuum NLS equation in the regime of oscillatory instability , and this type of argument has also been used to explain the ’quasi-collapse’ of a broad excitation to a narrow localized state in the two-dimensional DNLS equation . ## IV Breather interacting with standing-wave phonons We choose, as in the previous section, the initial perturbation $`ϵ_n(0)`$ to be an eigensolution of the linearized equations (12), but now we consider the case of a spatially extended perturbation. Choosing without loss of generality a solution of the form (16) with positive frequency $`\omega _p`$ yields the asymptotic behaviour $$U_n,W_n\mathrm{cos}(qn\pm \delta );n\pm \mathrm{},$$ (61) where the wave vector $`q`$ $`(0q\pi )`$ is determined by the dispersion relation (21), and $`\delta `$ is the phase shift across the breather. Thus, the excitation of an extended eigenmode corresponds to an interaction between the breather and a non-propagating, standing-wave phonon with small amplitude $`a`$. As was mentioned in the introduction, these standing-wave phonons are generally unstable, but since the instabilities become exponentially weak in the small-amplitude limit they are expected to have very little effect on the breather for the perturbation sizes and time-scales considered in this section. We will return to the effects on the breather of these instabilities in Sec. V, where larger perturbations are considered. In contrast to the case of excitation of a localized internal mode discussed in Sec. III, where higher order radiation always was emitted from the breather region, the condition (28) yields that for the standing-wave perturbation the breather will radiate to higher order only if $`\mathrm{\Lambda }\omega _p<\mathrm{\Lambda }/2+2C`$, so that both $`\omega _p`$ and $`2\omega _p`$ are inside the phonon band. In terms of the phonon wave vector $`q`$, this means that there is a critical value $`q_c`$, $$q_c=\mathrm{arccos}\left(\frac{\mathrm{\Lambda }}{4C}\right),$$ (62) such that for $`0q<q_c`$ second order radiation will be emitted from the breather region, while for $`q_c<q\pi `$ ($`\mathrm{\Lambda }/2+2C<\omega _p\mathrm{\Lambda }+4C`$) all multiples of $`\omega _p`$ are outside the phonon band, and no higher order radiation is emitted. As will be shown below, these two regions yield qualitatively different scenarios for the long-time evolution of the perturbed breather. We also note that for $`\mathrm{\Lambda }>4C`$ we have $`q_c=0`$, so that for the highly localized, high-frequency breathers no phonons can generate higher-order radiation (note also that there are no internal modes in this regime). Furthermore, we always have $`q_c<\pi /2`$, so that the regime of higher-order radiation generation is a subset of the regime $`0<q<\pi /2`$ where modulational instability for travelling plane waves occurs . Let us first discuss the case $`q<q_c`$. A typical example of the long-time evolution of a breather interacting with a small-amplitude standing-wave phonon with $`q<q_c`$ is illustrated by Fig. 4. As is seen from Fig. 4 (a), the amplitude of the oscillations remains essentially constant in time, but a closer inspection reveals that the average value of $`|\psi _{n_0}|^2`$ asymptotically increases with an apparently constant rate (see inset in Fig. 4 (b)). Similarly, Fig. 4 (b) shows that also the total norm contained in any finite region around the breather asymptotically increases linearly with time. We find that these results are generic for all cases when the phonon wave vector $`q<q_c`$ (the spatial symmetry of the phonon is not important for the asymptotic behaviour), and thus we conclude that in this regime, the breather can ’pump’ energy from the phonon (which is infinite for an infinite system), and thereby grow. In the same spirit as for the internal mode excitation in Sec. III C, we can give a simple argument based on the conservation laws to motivate why the generation of second order radiation should lead to breather growth. To this end, we assume that the initial standing-wave phonon is infinitely extended, and that far away from the breather a stationary regime will be reached corresponding to the following boundary conditions $$\psi _n\left[(ae^{iqn}+re^{\pm iqn})e^{i\omega _pt}+r_2e^{i(\pm q_2n2\omega _pt)}\right]e^{i\mathrm{\Lambda }t};n\pm \mathrm{}.$$ (63) Thus, we have taken into account the second-order radiation with frequency $`2\omega _p`$ generated at the breather region but neglected possible higher-order radiation; moreover the resonance at the original phonon frequency $`\omega _p`$ in the third-order equation (14) has been taken into account by allowing the incoming and outgoing complex amplitudes $`a`$ and $`r`$ to be different. We can then, in analogy with Eqs. (52) and (54), write the balance equations for the total norm and Hamiltonian contained in a region around the breather averaged over a time-interval $`[t,t+2\pi /\omega _p]`$ as $$\frac{\mathrm{d}𝒩_t}{\mathrm{d}t}=\frac{\mathrm{d}𝒩_\varphi }{\mathrm{d}\mathrm{\Lambda }}\dot{\mathrm{\Lambda }}=J_𝒩(\mathrm{})_tJ_𝒩(+\mathrm{})_t=4C\left[(|a|^2|r|^2)\mathrm{sin}q|r_2|^2\mathrm{sin}q_2\right],$$ (64) and $`{\displaystyle \frac{\mathrm{d}_t}{\mathrm{d}t}}={\displaystyle \frac{\mathrm{d}_\varphi }{\mathrm{d}\mathrm{\Lambda }}}\dot{\mathrm{\Lambda }}`$ $`=`$ $`J_{}(\mathrm{})_tJ_{}(+\mathrm{})_t`$ (65) $`=`$ $`4C[(|a|^2|r|^2|)(\omega _p\mathrm{\Lambda })\mathrm{sin}q|r_2|^2(2\omega _p\mathrm{\Lambda })\mathrm{sin}q_2],`$ (66) respectively. Here, we have used the facts that the time-average of the norm current density (6) and the Hamiltonian flux density (7) are additive quantities for small-amplitude plane waves, and that in the stationary regime, the mode amplitudes $`a`$, $`r`$ and $`r_2`$ are time-independent. Combining Eqs. (64) and (66) and using (9), we obtain that the breather frequency grows with a constant rate given by $$\dot{\mathrm{\Lambda }}=\frac{4C|r_2|^2\mathrm{sin}q_2}{\frac{\mathrm{d}𝒩_\varphi }{\mathrm{d}\mathrm{\Lambda }}}>0.$$ (67) The physical interpretation of this result is, similarly as for the case of internal mode excitation, that the generation of higher-order radiation results in a net flow of negative Hamiltonian energy into the breather region, which is absorbed by the breather by increasing its frequency and maximum amplitude. This process is similar to the one observed for the two-channel phonon scattering on breathers in a Klein-Gordon model with a Morse potential in Ref. ; however in the latter case the second outgoing wave resulted from a resonance in the linearized equations and were therefore of the same order of magnitude as the incoming wave, and moreover the outcome in this case was breather decay since the energy of a Klein-Gordon breather is an increasing function of its amplitude. Considering now the regime $`q>q_c`$ where all multiples of $`\omega _p`$ are outside the phonon band, the most important conclusion from our extensive numerical investigations is that we never observe breather growth. Instead, we sometimes (but not always) observe a very slow decrease of $`|\psi _{n_0}|^2_t`$, and an increase of the fluctuations around this mean-value. This behaviour is illustrated by Fig. 5 (a) and (b). A possible interpretation of these results is that, since the higher harmonics which are created by the breather-phonon interaction cannot propagate, they stay trapped around the breather. Thus, this could lead to a transfer of energy from the ’pure’ breather, which acquires more and more internal frequencies and becomes a ’chaotic breather’ . Another possible interpretation is that the increase of the oscillation amplitude is connected with the oscillatory instabilities of the standing waves; as we will show in the next section these instabilities provide a mechanism for breather decay. However in some cases, illustrated by Fig. 5 (c) and (d), the oscillation amplitude as well as its average value apparently approaches a constant limit value. We have at present no explanation for this behaviour (as will be discussed in the next section, there exist exact ’phonobreather’ solutions which could be candidates for such a final state, but they are unstable); it is possible that the time ranges that we were able to study with sufficient numerical accuracy in these cases simply were too short to observe the scenario described by Fig. 5 (a) and (b). To conclude this section, we repeat our main result that breather growth is observed if and only if $`q<q_c`$, where $`q_c`$ is given by (62). The fact that $`q_c=0`$ for $`\mathrm{\Lambda }>4C`$ thus implies the existence of an upper limit beyond which the breather cannot grow with the type of perturbations considered here. We would also like to relate our results to recent numerical simulations of breathers interacting with propagating phonons in Klein-Gordon and FPU lattices. For the Klein-Gordon lattice with a (soft) Morse on-site potential, phonons with small wave vector $`q`$ were observed to yield breather growth, while phonons with large $`q`$ caused breather decay. For a FPU lattice with hard anharmonicity, the opposite situation was observed, i.e., small-$`q`$ phonons caused breather decay and large-$`q`$ phonons breather growth. The fact that the situation for the hard FPU lattice was opposite to that of the soft Klein-Gordon lattice could be expected, since in the former case the modulational instability occurs for large $`q`$, whereas soft Klein-Gordon and DNLS lattices with $`C>0`$ are modulationally unstable for small $`q`$. However, we stress that the relation between plane-wave modulational instability and breather growth is nontrivial, and at least for the case considered in this paper the critical value $`q_c`$ for breather growth from interaction with standing-wave phonons differs from the critical value $`q=\pi /2`$ for modulational instability of travelling waves. ## V Breather growth and destruction from standing-wave instabilities In this section, with the aim at describing the interaction between a breather and a standing-wave phonon with non-negligible amplitude, we will take a slightly different point of view than in the preceding sections. Instead of choosing as initial condition an exact breather solution and adding a perturbation corresponding to an eigenmode of the linearized equations, we will here consider initial conditions which are exact phonobreather (or nanopteron) solutions. By definition, a phonobreather consists of a spatially localized breather on top of a spatially extended tail which is a nonlinear, standing-wave phonon. (There are also solutions where the tail is a propagating wave , but we will not discuss them further here.) Phonobreathers exist generically for nonlinear lattice-equations (see e.g. ), but their existence normally requires an integer relationship between the breather and phonon frequencies. However, for the DNLS equation phonobreathers exist for any (rational or irrational) relation between the two frequencies , as a consequence of the additional invariance of the equation under global phase transformations. Since the phonobreathers are exact solutions consisting of a breather part and a standing-wave part, one could expect them to be attractors for the initial conditions considered in Sec. IV. However, as was shown recently , generically for soft Klein-Gordon and DNLS models with $`C>0`$ all phonobreathers with phonon wave vector $`q\pi `$ will be linearly unstable if, for fixed phonon amplitude $`a`$, the linear coupling $`C`$ is larger than some threshold value $`C_{cr}(a,q)`$ (i.e., away from the anticontinuous limit). (For lattices with hard potentials and DNLS with $`C<0`$, the stable phonobreather has $`q=0`$.) These instabilities are caused by an oscillatory instability of the standing-wave phonon itself, which can be understood by considering the construction of a nonlinear standing wave with wave vector $`q`$ close to $`\pi `$ at the anticontinuous limit $`C=0`$ by introducing a periodic array of discommensurations or ’defects’ in the nonlinear phonon with wave vector $`\pi `$ and amplitude $`a`$, $`\psi _n=a(1)^ne^{i(4C|a|^2)t}`$, which is linearly stable for all $`a`$ and $`C>0`$ . In the anticontinuous limit $`C=0`$, each defect consists when $`\pi /2<q<\pi `$ of one extra site with $`\psi _n=0`$ added to the $`\pi `$-phonon, which consequently suffers an additional phase shift of $`\pi `$ across each defect. For $`0<q<\pi /2`$, each defect consists of several consecutive zero-amplitude sites with associated phase shifts (a general method for generating the anticontinuous coding sequence for standing-wave phonons from a circle map is described in ); in this case it is also useful to consider the periodic repetition of sites with $`\psi _n=\pm a`$ as defects of the zero-amplitude state. The limit case of one isolated zero-amplitude defect, which is a discrete counterpart of the dark-soliton solution of the continuum NLS equation, was investigated in . The linear stability analysis of this mode showed that, although it is stable close to the anticontinuous limit, it suffers a bifurcation for $`C/|a|^2=C_c0.0765`$ where two pairs of eigenvalues of the eigenvalue problem (20) go out in the complex plane. The resulting oscillatory instability occurs due to a resonance between a mode localized around the defect (the defect pinning mode) and linear radiation modes. It was shown that for finite systems, the mode recovers its stability above some upper critical value of $`C/|a|^2`$ (since the wavelength of the resonating linear modes becomes larger than the system size); however this critical value increases with system size so that in the limit of an infinite system, the instability persists for all $`C/|a|^2>C_c`$ but with a growth-rate that decreases in an exponential-like fashion when approaching the continuum limit $`C/|a|^2\mathrm{}`$. This instability was shown to result in the defect becoming mobile (in NLS terms, the stationary ’black’ soliton with zero minimum intensity transforms into a moving ’grey’ soliton with non-zero minimum intensity) and radiation being emitted. In terms of the phase dynamics, this describes a moving, slowly spreading phase kink. The instability scenario for the standing-wave phonons is basically the same as for the isolated defect, with the essential difference that the localized pinning modes associated with the individual, periodically repeated defects now will form a continuous ’defect band’. In general, the Krein signature of this defect band is opposite to that of the bands associated with the non-zero amplitude sites (for $`0<q<\pi /2`$ there are generally several defect bands, but they will have the same Krein signature), and as a consequence resonances between the bands will occur if the linear coupling $`C`$ is large enough, giving rise to similar oscillatory instabilities as described above (details are given in ). We remark that earlier analysis of standing waves in nonlinear lattices, based on a quasi-continuum approximation, did not reveal these instabilities since their origin is the discrete nature of the lattice. Let us now return to the main objective of this section, namely to study the effect of the oscillatory standing-wave instabilities on the phonobreathers. We find that the families of phonobreathers which are stable close to the anticontinuous limit and whose tails approach harmonic standing waves in the small-amplitude limit can be constructed from anticontinuous standing-wave solutions at $`C=0`$, placing the breather at a zero-amplitude site of the phonon and adjusting it so that the resulting solution is either symmetric or antisymmetric around the breather site. Denoting the anticontinuous breather amplitude by $`b`$ (the phase of the breather site is unimportant when $`|b||a|`$ ), this yields the following possibilities: (i) For $`q>\pi /2`$ the antisymmetric anticontinuous solution (here $`q=2\pi /3`$) $$\{\psi _n(0)\}=\{\mathrm{}a,a,0,a,a,0,a,a,b,a,a,0,a,a,0\mathrm{}\},$$ (68) with the asymptotic behaviour $`\psi _n(0)a\mathrm{sin}(qn),n\pm \mathrm{}`$ in the continuum limit $`C/|a|^2\mathrm{}`$. (Note that this solution is antisymmetric only at $`C=0`$.) (ii) From (68) we can construct a symmetric solution for $`q>\pi /2`$ by introducing an additional phase shift of $`\pi `$ at one side of the breather site, giving for $`q=2\pi /3`$ $$\{\psi _n(0)\}=\{\mathrm{}a,a,0,a,a,0,a,a,b,a,a,0,a,a,0\mathrm{}\},$$ (69) with the asymptotic behaviour $`\psi _n(0)a\mathrm{cos}(qn\pm \pi /2),n\pm \mathrm{}`$ in the continuum limit. (iii) For $`q=M\pi /N<\pi /2`$, $`N`$ even, the number of consecutive zero-amplitude sites is odd and the antisymmetric anticontinuous solution is for e.g. $`q=\pi /4`$ $$\{\psi _n(0)\}=\{\mathrm{}0,0,0,a,0,0,0,a,0,0,0,a,0,b,0,a,0,0,0,a,0,0,0,a\mathrm{}\}$$ (70) behaving as $`\psi _n(0)a\mathrm{cos}(qn),n\pm \mathrm{}`$ in the continuum limit. (iv) From (70) the symmetric solution for $`q=M\pi /N<\pi /2`$, $`N`$ even, is constructed by a phase shift as above, giving for $`q=\pi /4`$ $$\{\psi _n(0)\}=\{\mathrm{}0,0,0,a,0,0,0,a,0,0,0,a,0,b,0,a,0,0,0,a,0,0,0,a\mathrm{}\}$$ (71) with the asymptotic behaviour $`\psi _n(0)a\mathrm{sin}(qn\pm \pi /2),n\pm \mathrm{}`$ in the continuum limit. (v) For $`q=M\pi /N<\pi /2`$, $`N`$ odd, the number of consecutive zero-amplitude sites is even and we must add an extra site to obtain the antisymmetric anticontinuous solution, which for e.g. $`q=\pi /3`$ becomes $$\{\psi _n(0)\}=\{\mathrm{}0,0,a,0,0,a,0,0,a,0,b,0,a,0,0,a,0,0,a,0,0,a\mathrm{}\}$$ (72) behaving as $`\psi _n(0)a\mathrm{cos}(q(n+1/2)\pm q/2),n\pm \mathrm{}`$ in the continuum limit. (A solution with similar properties is obtained by instead removing one zero-amplitude site; however its symmetric counterpart is always unstable.) (vi) The symmetric counterpart of (72) for $`q=M\pi /N<\pi /2`$, $`N`$ odd, is constructed by a phase shift as above, giving for $`q=\pi /3`$ $$\{\psi _n(0)\}=\{\mathrm{}0,0,a,0,0,a,0,0,a,0,b,0,a,0,0,a,0,0,a,0,0,a\mathrm{}\}$$ (73) with the asymptotic behaviour $`\psi _n(0)a\mathrm{sin}(q(n+1/2)\pm (q+\pi )/2),n\pm \mathrm{}`$ in the continuum limit. A typical example on the time-evolution for an initially very weakly perturbed phonobreather with phonon wave vector $`q>\pi /2`$ and phonon amplitude small but non-negligible compared to the breather amplitude is illustrated in Fig. 6. (The example in the figure belongs to type (i), but similar dynamics is observed also for the spatially symmetric states of type (ii).) We can clearly distinguish two different steps leading to the final breather destruction. The first step is the linear oscillatory instability described above, which leads to the generation of new internal frequencies of the breather, and to the movement of the defect sites in a similar way as for the case of an isolated defect. In the second step, the moving defects start interacting, and a close inspection of Fig. 6 (a) shows that neighboring defects tend to merge and create regions of accumulated phase fluctuations travelling around in the lattice. These will interact with the breather, and apparently cause its decay. When the breather has decayed sufficiently to have an excitable pinning mode, it will start to move in the lattice but with rapidly decreasing amplitude, and it will finally be destroyed. We have at present no complete understanding for the mechanism by which the interaction of the breather with the moving ’phase kinks’ cause its decay, but we remark that a similar scenario was observed when adding to the DNLS equation an external, parametric white noise term . In that case, the white-noise approximation allowed a qualitative understanding of the breather decay as a consequence of phase fluctuations by using a collective coordinate approach. However, to observe this scenario for breather destruction it is necessary (at least for a finite system) that the phonon amplitude is not too small compared to the breather amplitude. If we increase the breather amplitude in Fig. 6 (or decrease the phonon amplitude) sufficiently, we find that although the oscillatory instability develops, the fluctuations created in the second step will be too weak to cause the breather to decay, and it will live seemingly forever as a ’chaotic phonobreather’. The absence of decay for small perturbations can be viewed as a consequence of the fact that the single-site DNLS breather is nonlinearly (Lyapunov) stable for norm-conserving perturbations, in the sense that $`|\psi _n(t)|`$ remains arbitrarily close to the breather for all times if the initial perturbation is small enough . Thus, it is clear that for finite systems, the breather cannot be destroyed unless the phonon amplitude exceeds some critical value, while nothing can be said about the infinite system since any infinitely extended phonon obviously has an infinite norm. With $`0<q<\pi /2`$, the first step resulting from the oscillatory instability occurs in a similar way as for $`q>\pi /2`$: the breather acquires new frequencies and the small-amplitude sites of the phonon start moving. By instead making the interpretation that the sites with non-zero amplitude at the anticontinuous limit are defects in the zero-amplitude state, their movement can be seen as a consequence of the repulsive interaction between spatially separated, small-amplitude breathers with opposite phases observed e.g. in . We also observe, similarly as for $`q>\pi /2`$, the merging of neighboring defects, but in this case their interaction with the breather will not lead to breather decay, but rather to breather growth if the original amplitude of the breather is not too large. A typical example is illustrated in Fig. 7. It can be seen by a careful inspection of Fig. 7 (a) how the merging of small-amplitude sites results in localized humps of larger amplitude reminiscent of small-amplitude moving breathers travelling around in the lattice. The interaction of these humps with the original breather leads to growth of the latter in a similar way as observed in . However, this growth stops when the breather amplitude has reached a critical value which is close to (but apparently smaller than) that corresponding to the limit value $`\mathrm{\Lambda }=4C`$ for small-amplitude perturbations found in Sec. IV (the latter corresponds to $`|\psi _{n_0}|^25.65`$). The final state appears also here to be a ’chaotic phonobreather’; we have followed the time evolution of this kind of state for times up to $`10^6`$ without seeing any signs of decay. Also, if the initial breather frequency is chosen above the critical value $`\mathrm{\Lambda }=4C`$, we typically do not observe breather growth; instead the mean value of the chaotic amplitude oscillations resulting from the oscillatory instability remains close to the initial amplitude. ## VI Concluding remarks Investigating the interaction between discrete nonlinear Schrödinger breathers and small perturbations, we have found firstly that exciting an internal mode of the breather always leads to a slow energy transfer to the breather, i.e., to breather growth. Furthermore, we found that a DNLS-breather can pump energy from a small-amplitude standing-wave phonon, provided that the phonon wave vector is smaller than the critical value $`q_c`$ given by Eq. (62). In both cases, the mechanism for breather growth involves the higher-order generation of radiating modes. Since this mechanism disappears at the threshold value $`\mathrm{\Lambda }=4C`$ of the breather frequency, it is impossible for a breather to grow beyond this value with the type of small-amplitude perturbations considered here. To analyze the interaction between breathers and standing-wave phonons of small but non-negligible amplitude, we considered the long-time evolution of weakly perturbed exact phonobreather solutions. The instabilities of these, originating in oscillatory instabilities of the nonlinear phonons, where shown to lead to propagating inhomogeneities whose interaction with the breather provided a mechanism for breather decay and destruction (when the phonon wave vector $`q>\pi /2`$) or growth ($`q<\pi /2`$ and $`\mathrm{\Lambda }<4C`$). As was mentioned already in the introduction, the existence of the two conserved quantities (1) and (3) makes the DNLS equation non-generic among nonlinear lattice equations, and it is therefore necessary to investigate to what extent the results obtained in this paper apply also for Klein-Gordon and FPU lattices. We plan to address these questions in a forthcoming publication, but let us stress already here that the perturbational approach used here for the DNLS equation needs to be modified to account for the fact that generically, the dynamics of the breather involves also higher harmonics of its fundamental frequency. Moreover, the approach used in Secs. III C and IV based on the conservation laws cannot be directly applied in the absence of a second conserved quantity. However, in view of the wide applicability of the DNLS equation (and in particular its appearance as a limit case of general lattice equations as mentioned in the introduction), we believe that the mechanisms for breather growth and destruction described in this paper are essential ingredients also for the corresponding processes in general lattice models. ###### Acknowledgements. We thank Yu. S. Kivshar for giving us an early preprint of Ref. , I. V. Barashenkov for directing our attention to Ref. , and A. M. Morgante for discussions on phonobreathers and standing-wave instabilities. M. J. acknowledges a Marie Curie Research Training Grant from the European Community. A preliminary version of these results was presented at the conference Nonlinearity ‘99 (Heraklion, May 10-14, 1999).
warning/0002/astro-ph0002521.html
ar5iv
text
# Distances to Cepheid Open Clusters Via Optical and K-Band Imaging ## 1 Introduction Determining the value of Hubble’s constant, H, has been a challenge to astronomers since the discovery of the universal expansion in 1927. It is sometimes argued that we are now at the fine tuning stage and many measurements give values for H which lie between the hotly argued values of 50kms<sup>-1</sup>Mpc<sup>-1</sup> (Sandage) and 100kms<sup>-1</sup>Mpc<sup>-1</sup> (de Vaucouleurs), e.g. ? calculated H=67$`\pm 7`$kms<sup>-1</sup>Mpc<sup>-1</sup>. However, many of these measurements are based on secondary indicator methods which in turn are dependent on the accuracy of primary indicators of distance such as the Cepheid Period-Luminosity (P-L) relation. The well-studied LMC P-L relation is usually calibrated via the distance modulus to the LMC and the previously accepted value was around 18.50. However, this has been recently challenged in a paper by ? who determined the distance modulus to the LMC as 18.70$`\pm 0.1`$. This small difference in the distance modulus causes a 10$`\%`$ decrease in estimates of the Hubble’s Constant. This discrepancy has further motivated us to check the Galactic zeropoint of the P-L relation. We do this by checking the values of the distance modulus and reddening of the 11 Galactic clusters that contain Cepheids via zero age main sequence fitting (ZAMS). Previous work on measuring the reddening and distance to young open clusters which contain Cepheids via ZAMS fitting has been done using photoelectric and photographic measurements in optical wavebands. It is time consuming to observe a large number of stars using photoelectric observations as each star has to be observed individually. Photographic data can give relatively inaccurate magnitudes and colours. However, CCD’s now make it possible to observe a large number of stars in many different wavebands quickly and accurately. Although CCD’s have already been used for open cluster studies e.g. ?, ?, ?, these have mainly been carried out in BVRI. Recently CCD’s with improved U-band sensitivity have become available and U-band CCD data is included in this study. Infra-red imaging detectors are also now available and although some of the detectors used here do not cover as wide an area as optical CCD’s, observing the full extent of an open cluster with a mosaic of pointings is a practical proposition. Until fairly recently, good quality infra-red measurements of the Cepheids themselves were not available and the Cepheid P-L relation has been primarily calibrated in the V-band. Laney and Stobie (1993,1994) present infra-red along with V-band magnitudes for a large number of Southern Hemisphere Galactic Cepheids. Using data in the literature to obtain values for the distance modulus and reddening to the clusters they calibrated the Cepheid P-L relation in the V and K-band. Any errors in the determination of the distance modulus and the reddening in the previous work would cause an error in the PL relation as determined by Laney and Stobie. The layout of this paper is as follows. In section 2 we present the observational data and we test the accuracy of the photometry and calibration of the data. In section 3.1 we describe how the reddenings and distances to the open clusters are obtained and in section 3.2 we discuss each cluster individually. In section 4 we use these values with the magnitudes of the Cepheids to calibrate the Cepheid Period-Luminosity relation, In section 5 we discuss the implications of the results, particularly for the clusters whose U-B:B-V diagrams do not appear to follow the canonical locus. In section 6 we draw conclusions. ## 2 Data ### 2.1 Observations The observations of the Galactic Open clusters were taken during five observing runs on the JKT, UKIRT, at CTIO, at Calar Alto and on the WHT over a two year period. The spread in declination of the clusters and the multi-wavelength nature of the study meant that many different telescopes were required. #### 2.1.1 JKT Optical imaging of eight open clusters was obtained during an observing run from the 16/9/1997 to the 22/9/1997. The observations were carried out using the 1024$`\times `$1024 Tektronix CCD with pixel scale of 0.33 arcsec pixel<sup>-1</sup>. Typical seeing was around 1.3<sup>′′</sup>. Short exposures of 5s in V, 10s in B and 20s in U were observed for calibration purposes but the main imaging observations were typically 6x120s in the V-band, 6x180s in the B-band and 6x300s in the U-band. Due to the Southerly declination of some of the objects, they had to be observed at high air mass. However these observations were normally used to obtain relative photometry and calibration frames were observed at as low an airmass as possible or during a later observing run at CTIO. The pointings are given in Table 2. For calibration purposes, standard stars from ? were used. On the one fully photometric night (21/9/97) six Landolt fields were observed at regular intervals throughout the night, most of these fields containing several standard stars. #### 2.1.2 CTIO The observations were made using the CTIO 0.9-m during an observing run from the 24/9/98 to 29/9/98. These observations were carried out using the 2048x2048 Tek#3 CCD with pixel scale 0.384 arcsec pixel<sup>-1</sup>. The Tek#3 CCD has low readout noise (4 electrons) and good quantum efficiency in the U-band. The average seeing during the observations was around 1.4<sup>′′</sup>. Approximate pointings are again given in Table 2. Standard E-region fields from ? and standard stars from ? were observed throughout the night. The only photometric night was the 28/9/1998 and on this night six E-region standard fields and one Landolt standard field containing four standard stars were observed throughout the night. Three new clusters were observed at CTIO and further observations of clusters observed at the JKT were made in cases where the clusters had been observed in non-photometric conditions only. #### 2.1.3 UKIRT The infra-red data was mostly observed at UKIRT during the four nights 16-19/6/97 using the IRCAM3 near-IR imaging camera with a 256x256 detector. The pixel scale used was 0.286 arcsec pixel<sup>-1</sup> giving a field of view of 73<sup>′′</sup>. To cover a sufficient area of each open cluster we therefore had to mosaic images. Generally a mosaic of 9x7 images was observed. Each image was overlapped by half in both the x and y direction so the final image had an approximate area of 6$`{}_{}{}^{}\times `$ 5. Observations were generally taken using the ND-STARE mode, where the array is reset and read immediately and then read again after the exposure which reduces the readout noise to 35e<sup>-</sup>. The exposures were 60$`\times `$2 seconds and the centre of the mosaic is approximately in the same position as the corresponding optical frame. Standards from the UKIRT faint standards list were observed throughout the nights. Around 10 standards were observed on the three photometric nights, some of which were observed early on in the night, half way through and at the end of the night. The seeing throughout the run was typically 0.6<sup>′′</sup>. Observations were also taken in the J and H-band. #### 2.1.4 Calar Alto NGC129 lies further north than the declination limit of UKIRT so infra red observations were instead taken at Calar Alto during another observing run. Observations were done using the Rockwell 1k$`\times `$1k Hawaii detector with pixel scale 0.396 arcsec pixel<sup>-1</sup>. This gives a 6.6 field of view so there was no need for the mosaicing technique used at UKIRT. The exposure time was 10$`\times `$1.5s, the seeing was better than 1<sup>′′</sup> and the exposure was centred on star 113 in ?. Observations were made in the K<sub>short</sub>-band. UKIRT faint standards and standards of Hunt et al (1998) were observed for calibration. #### 2.1.5 WHT NGC7790 also lies further north than the UKIRT declination limit and so infra red observations were made on the 4.2m WHT during another observing run. The observations were made on the 1/9/1996. The WHIRCAM 256$`\times `$256 detector which was situated at the Nasmyth focus and behind the MARTINI instrument was used for the observations, without MARTINI tip-tilt in operation. The WHIRCAM detector was the IRCAM detector previously used at UKIRT. The pixel size was 0.25 arcsec pixel<sup>-1</sup> and the field-of-view was therefore 64<sup>′′</sup>, centred on star E in ?. The K<sub>short</sub> filter was used and UKIRT faint standards were observed for calibration. #### All the observations are summarized in Table 1. The date of each observation, the wavebands observed for each cluster and the airmass are given in columns 2, 5 and 6 respectively. Column 7 gives the exposure time of the frames used for imaging. For some of the clusters a calibration frame was observed at CTIO and the exposure time of these clusters are also given in column 7. Column 8 indicates where a cluster was observed on a photometric night and hence where an independent zero point was obtained. ### 2.2 Data Reduction #### 2.2.1 JKT Removal of the bias introduced into the data and trimming of the frames to remove the overscan region was done on all the frames using the IRAF task CCDPROC. At least eight U-band sky flats and six B and V-band sky flats were observed on each of the nights so a separate flat field was created for every night using a combination of dust and dawn sky flats. This was created within FLATCOMBINE using a median combining algorithm and a 3$`\sigma `$ clipping to remove any cosmic rays. The residual gradient in the flat fields is around 1%. The task CCDPROC then applies the flat fields to all the images. The same flat fields were used in the reduction of the standard star frames. Many images of the same cluster were observed. These were all combined together by aligning the images with linear shifts using the task IMSHIFT. Generally these shifts were small (a few pixels either way) as the observation were done one after each other and in some cases no shifts were required. The images were combined using IMCOMBINE and were averaged together using a 5$`\sigma `$ clipping. #### 2.2.2 CTIO The data was obtained at CTIO using the four amplifier readout mode. A package called QUADPROC within IRAF corrects for the different bias levels in the four quadrants of the CCD. The frames were also trimmed using QUADPROC. In a previous observing run, ? had found that there was a residual gradient of 5% in the dome flat fields so sky flats were used. At least three sky flats were observed on each night in the B and V-band and at least 5 sky flats were observed on each night in the U-band. The resulting flat fields were flat to better then 1%. The equivalent version of FLATCOMBINE in the QUAD package was used to median combine the flat fields using a 3$`\sigma `$ clipping. The E-region standards and observations of Landolt standards were again reduced in the same manner. Multiple images of the same cluster were again combined using IMCOMBINE with the same settings as for the JKT data and where any offset shifts appeared between the data frames they were again corrected for using IMSHIFT. #### 2.2.3 UKIRT The UKIRT data was reduced using a program called STRED within the package IRCAMDR. This is a fairly automated routine which reads in the data frames, subtracts of the dark count and creates a flat field frame by median filtering the image frames. Then the program flat fields the dark subtracted object images, corrects for any bad pixels and finally creates a mosaic. All the data frames were median combined to create a flat field for each night. There was no evidence of a large scale gradient greater than about 1% in the flat fields. To create the final image, all the individual frames have to be mosaiced together. STRED reads in the offset from the data header, however, these offsets were not accurate enough. By creating a separate offsets file the mosaicing could be done more accurately. To create the offsets file, one of the corner frames was fixed and the offset required for the neighbour frame were found by eye. This was built up over the whole frame, however once one offset had been determined, all the other offsets in the x and y direction from frame to frame were the same. The offsets for the standard stars were more accurate and could be used to create the mosaic. #### 2.2.4 Calar Alto Basic IRAF routines such as IMCOMBINE and IMARITH were used to reduce the Calar Alto data. A flat field was created by median combining all the data frames. A more detailed description of the data reduction can be found in ?. We found that the best results were obtained by first subtracting a sky frame from each image, as follows: A sky frame for each individual image was created using IMCOMBINE with a $`5\sigma `$ clipping to median filter four data frames that were local in time to the image frame to create a sky frame for each data frame. The sky frame was then subtracted off the image before IMCOMBINE was used again to combine the data frames, forming one final image frame. #### 2.2.5 WHT The data from the WHT was reduced for us as part of another project. Dome flats were used to flat field the data and this was divided into the science frame using IMARITH. Sky subtraction was also required. A sky frame was created by combining dedicated sky frames observed locally in time to the science frame and this was then subtracted from the science frame also using IMARITH. ### 2.3 Image Alignment In order to be able to produce colour-magnitude diagrams, the magnitude of each star in all the different wavebands is required. To do this, all the frames need to be aligned. Aligning the optical frames was easy as there were only linear shifts between each waveband. Rather than altering the data, the alignment was just done by applying small corrections to the x and y positions of the stars. Aligning the optical data with the K-band data was more complicated though as there were shifts, shears and rotations between the frames. We used the IRAF routine GEOMAP to calculate the best spatial transformation function between any two images thus allowing the optical and K-band data to be aligned. The mapping was only used to transform coordinates and we did not perform photometry on the resampled images. ### 2.4 Photometric Calibration #### 2.4.1 JKT Observations of the standard stars of ? were taken at regular intervals on each of the six nights at the JKT. The CCD frames of the standard stars were reduced in the same manner as the data frames with the same flat fields etc as discussed in section 2.2.1. The aperture size used to measure the magnitude of the standard stars was 15<sup>′′</sup>, large enough to determine accurately the total magnitude of the star but not so large that sky subtraction errors dominate the magnitude measurement. Only the sixth night (21/9/97) was fully photometric and the zero point, airmass coefficient and colour equation are shown in Figure 1 and given below with the rms scatter. $$U_{\mathrm{jkt}}=U_{\mathrm{ldt}}+3.76+0.49sec(z)0.063(U_{\mathrm{ldt}}B_{\mathrm{ldt}})\pm 0.038$$ $$B_{\mathrm{jkt}}=B_{\mathrm{ldt}}+2.00+0.27sec(z)0.013(B_{\mathrm{ldt}}V_{\mathrm{ldt}})\pm 0.033$$ $$V_{\mathrm{jkt}}=V_{\mathrm{ldt}}+1.99+0.19sec(z)\pm 0.026$$ where the subscript ldt stands for the Landolt standard star magnitude and the subscript jkt stands for the instrumental magnitude, z represents the zenith distance. The errors on the airmass are $`\pm `$0.0025, 0.0012 and 0.0008 in the U, B and V and the errors on the colour equation are $`\pm `$0.0007 and 0.0005 in the U and B-bands respectively. There are no Landolt magnitudes available for the data frames so the colour term has to be translated in instrumental magnitudes. The colour term is negligible in the V-band calibration so the instrumental V-band magnitudes come directly from the above equations. The B<sub>ldt</sub>-V<sub>ldt</sub> and U<sub>ldt</sub>-B<sub>ldt</sub> colours are given by $$(B_{\mathrm{ldt}}V_{\mathrm{ldt}})=1.013((B_{\mathrm{jkt}}V_{\mathrm{jkt}})0.010.08sec(z))$$ $$(U_{\mathrm{ldt}}B_{\mathrm{ldt}})=1.067((U_{\mathrm{jkt}}B_{\mathrm{jkt}})1.760.22sec(z))$$ we assume that the contribution from the colour term in the calibration of the B-band is negligible when determining the U<sub>ldt</sub>-B<sub>ldt</sub> colour. These exact colour terms are used to correct the instrumental magnitudes in order to make the colour-colour and colour-magnitude diagrams in Figures 11, 12 and 13. The clusters NGC7790, NGC6664 and Trumpler 35 were observed on the one photometric night. Short exposure observations of NGC6649 and M25 were made at CTIO in order to obtain an independent zero point for these frames. Around 20 bright (brighter than V$``$15), fairly uncrowded, unsaturated stars were taken as standard stars to identify the relation between the zero point from the JKT data and from CTIO data to an accuracy of 0.01 mags. The agreement between the zero point found via this method and previous, photoelectric calibrations is good (see Table 3) with only small offsets in each case. The remaining three clusters were observed in non-photometric conditions only so previous work had to be relied upon for the calibration. For NGC6823 the photoelectric observations from Table 1 of ? were used. Some of these stars were saturated on the CCD frame and the area of overlap between the two images was not identical but thirteen stars were suitable for calibration purposes. ? compares his photoelectric data with that of previous work and finds good agreement. Two sources of photoelectric data are available for the cluster NGC129, (?) and (?). There are 9 stars in common with the Arp photometry and 13 stars in common with the Turner photometry. For these samples of stars, we find that the U-band zero point obtained from Arp is 0.04$`\pm `$0.03 mags brighter than that of Turner. In the B-band the difference is 0.02$`\pm 0.01`$ mags in the sense that Arp is brighter than Turner and there is 0.01$`\pm 0.01`$ mag difference in the same sense in the V-band. These differences are mainly caused by the stars in the sample with V fainter than 14 mag. We therefore use the average of the brightest two stars from Arp and Turner to calibrate NGC129. The zero point from this method agrees very well with the zero point obtained using Turner’s photometry, which is shown in Figure 16. Finally the photoelectric work of ? was used to calibrate the cluster containing the Cepheid WZ Sgr. The magnitudes of the stars on the data frames on all nights were measured using a 5<sup>′′</sup> aperture. The standard star magnitudes were measured using a 15<sup>′′</sup> aperture so for the photometric night (21/9/97) an aperture correction had to be applied to the data. The correction in the U, B and V-bands are -0.165$`\pm `$0.02, -0.11$`\pm `$0.025 and -0.11$`\pm `$0.024 magnitudes respectively as determined by comparing the magnitudes of the standard stars with a 5<sup>′′</sup> and 15<sup>′′</sup> aperture. No aperture correction was required for those cases where the calibration was tied to a photometric sequence in the cluster itself. For the clusters NGC6649 and M25 the CTIO aperture correction was required. #### 2.4.2 CTIO E-region standards from ? and Landolt standards were observed for the photometric calibration of the optical CTIO data. ? compared the zero points and colour differences found from using the two different standard star studies and found the offsets between the two to be small, 0.004$`\pm `$0.0095 offset in the sense E regions - Landolt. and similar sized offsets in the U-B and B-V colours. Any offsets are within the quoted error. As before, these standard frames were reduced in the same manner as the data frames and a 15<sup>′′</sup> aperture was used to determine the magnitude. Again, only one night was photometric and this was the last night (28/9/98). The zero points, airmass coefficients and colour equations for each waveband, U, B and V, are shown in Figure 2 and given below again with the rms scatter. $$U_{\mathrm{ctio}}=U_{\mathrm{std}}+4.61+0.47sec(z)0.036(U_{\mathrm{std}}B_{\mathrm{std}})\pm 0.035$$ $$B_{\mathrm{ctio}}=B_{\mathrm{std}}+3.15+0.21sec(z)+0.099(B_{\mathrm{std}}V_{\mathrm{std}})\pm 0.015$$ $$V_{\mathrm{ctio}}=V_{\mathrm{std}}+2.93+0.12sec(z)0.018(B_{\mathrm{std}}V_{\mathrm{std}})\pm 0.007$$ where the subscript std stands for the E-field or Landolt standard magnitude and ctio stands for the instrumental magnitude. The errors on the airmass are $`\pm `$0.003, 0.0011 and 0.0009 in the U, B and V and the errors on the colour equation are $`\pm `$0.0006, 0.0007 and 0.001 in the U, B and V-bands respectively. There is generally good agreement between the values for the airmass coefficients and colour equations found in this work and in ?. Again, a 5<sup>′′</sup> aperture was used for the data frames so an aperture correction for the photometric night was required. The aperture corrections, in the U, B and V-bands, are -0.17$`\pm `$0.03, -0.17$`\pm `$0.02 and -0.13$`\pm `$0.02 magnitudes respectively. Again, for the data frames the colour terms have to be found in terms of CCD magnitudes rather than standard magnitudes. The colour term in the B-band is in this case non-negligible so V-band magnitudes have to be used in the U-band calibration. The standard colours are given by $$(B_{\mathrm{std}}V_{\mathrm{std}})=1.088((B_{\mathrm{ctio}}V_{\mathrm{ctio}})0.220.09sec(z))$$ $$\begin{array}{c}\hfill (U_{\mathrm{std}}B_{\mathrm{std}})=1.037((U_{\mathrm{ctio}}B_{\mathrm{ctio}})1.460.26sec(z)\\ \hfill +0.099(1.088((B_{\mathrm{ctio}}V_{\mathrm{ctio}})0.220.09sec(z))))\end{array}$$ again these colours are used in all to calculate the instrumental magnitudes which are used in the colour-colour and colour-magnitude diagrams in Figures 11, 12 and 13. This night provided independent zero points for the clusters NGC6067, Lynga 6 and vdBergh1, which were not observed at JKT, and also provided a zero point for NGC6649 and M25, which were only observed in non-photometric conditions at the JKT. #### 2.4.3 UKIRT The standards stars observed at UKIRT were taken from the faint standards list available from the UKIRT Web page. The standard stars were observed as a mosaic of five frames, a central frame with an overlapping frame in each direction. A 15<sup>′′</sup> aperture was used to determine the magnitude of the standard. Three of the nights were photometric, night1,2 and 4 (16,17,19/6/97) so all the clusters were observed in photometric conditions. The zero points for each of the nights are shown in Figure 3 and the airmass coefficient and colour equation are shown in Figure 4 and given below for each night. 16/06/1997; $$K_\mathrm{U}=K_{\mathrm{std}}+6.940.082sec(z)+0.005(J_{\mathrm{std}}K_{\mathrm{std}})\pm 0.024$$ 17/06/1997; $$K_\mathrm{U}=K_{\mathrm{std}}+6.950.082sec(z)+0.005(J_{\mathrm{std}}K_{\mathrm{std}})\pm 0.033$$ 18/06/1997; $$K_\mathrm{U}=K_{\mathrm{std}}+6.960.082sec(z)+0.005(J_{\mathrm{std}}K_{\mathrm{std}})\pm 0.037$$ where again the subscript std refers to the standard stars. The difference between each of the calibrations is just a small change in the zero point. The error on the airmass coefficient is $`\pm `$0.002 and the error on the colour term is $`\pm `$0.0003. The airmass coefficient agrees well with the values given in ? and those found on the UKIRT web page. An aperture correction of -0.19$`\pm `$0.018 magnitudes is required for the magnitudes of the data frames measured using a 5<sup>′′</sup> aperture on each night. As the colour term is very small, it was assumed negligible in the V-K:V CMD’s. #### 2.4.4 Calar Alto and WHT UKIRT faint standards and ? standards were observed in order to calibrate the Calar Alto and WHT data. The calibration was provided for us by Nigel Metcalfe as part of another project (see McCracken, Metcalfe and Shanks, in preparation, for more details). ### 2.5 Photometry Automated aperture photometry was done using PHOT within IRAF’s DAOPHOT package. A small, 5<sup>′′</sup>, aperture was used to minimise any crowding problems. PHOT was also used to obtain the magnitudes of the standard stars, the magnitude of each standard star was measured individually using a 15<sup>′′</sup> aperture. First we establish the depth of the photometry. We define the limiting depth of the observations to be where the Poisson error in the electron counts is 5%, i.e. when $`\sqrt{N_{obj}+N_{sky}}/N_{obj}`$ = 0.05. The depth of the JKT data is 19.9 mags for a 1200s U-band exposure, 19.7 mags for a 360s B-band exposure and 20.4 mags for a 240s V-band exposure. The depths of the CTIO data are similar. The depth of a typical K-band exposure of 120s observed was around 19.2 magnitudes. The accuracy of the zero points were tested by comparing our photometry to photoelectric observations in the literature. Table 3 shows the residual when our magnitudes for typically 10 stars from the calibrated data frames (see section 2.4) are compared to the magnitudes found from previous photoelectric studies. In all cases except for NGC6067, the residual is small (less than 0.03 mags). The errors in the comparison of our magnitudes to other photoelectric observations are slightly smaller than the errors in the zero point found from the standard stars in this work. This is probably because the stars considered when comparing the previous work were slightly brighter. Unfortunately for the cluster NGC6067 there is little photoelectric data available and only 6 stars can be compared in the B and V-bands due to saturation and different parts of the cluster being observed. In the U-band there are only 2 stars in common. Given that the offsets between the work here and previous work are small for all the other clusters, we assume that the zero point obtained for NGC6067 is accurate. Finally, we test how the crowding of the field effects the photometry. Twenty “simulated stars” of each magnitude shown in Figure 5 were placed in a 300s U-band image of the cluster M25 then PHOT was used to determine how well the magnitudes could be recovered. Shown in Figure 5 are the results. The x axis shows the true magnitude of the stars and the y axis shows the deviation of the mean magnitude from the true magnitude. This plot shows the total error (Poisson errors, read noise and also errors due to crowding) at each magnitude (full errorbar) and the Poisson error, found as described above, as the smaller, wider, errorbar. The total error for this exposure of 300s is less than 0.03 magnitudes down to U=17 mags. For a U-band exposure of 1800s, assuming that the crowding errors remain the same and then correcting the error shown in Figure 5 for the reduced Poisson error, the total error is estimated to be $`\pm `$0.17 at U=20, reducing to $`\pm `$0.08 at 18th mag and $`\pm `$0.03 at 17th mag. Figure 5 also shows that although the errors increase for fainter magnitudes there is no systematic trend. ## 3 Reddening and Distance ### 3.1 Method In order to work out the Cepheid Period-Luminosity (P-L) relation, the distance to the cluster needs to be known. However, there is generally significant dust absorption along the line of sight to the cluster which must be corrected for. The method for determining both of these parameters is done via ZAMS fitting to colour-magnitude and colour-colour diagrams. The ZAMS used in this study is a combination of ? for the optical data, the intrinsic colours of near IR-band stars from the UKIRT Web page (?) for the K-band data and nearby local stars taken from the Strasbourg Catalogue to give an indication of the acceptable spread in the ZAMS. The ZAMS of ? and ? have also been tested and would give the same results as the ZAMS of ?. The reddening is obtained using the U-B:B-V diagram which is independent of the distance to the cluster. The U-B:B-V diagram has always been the favoured method for estimating the reddening; however, previously, the accuracy of the reddening determination was limited by the depth of the U-band photoelectric photometry. The improved U sensitivity of the current generation of CCD’s should allow a potential improvement in the accuracy of the reddening estimated from U-B:B-V diagrams and this is the route we have adopted here. The reddening law assumed in this work is from ?. $$\frac{\mathrm{E}(UB)}{\mathrm{E}(BV)}=0.72+0.05\mathrm{E}(BV)$$ (1) We choose to fit the ZAMS to the ridge-line of the O and B stars rather than the least reddened envelope, because it helps take account of differential reddening in some of the clusters. There is evidence for differential reddening in NGC6823 and TR35 (see Figure 11) as the main sequence in the colour-colour diagram is substantially broadened. By fitting to the centre of the data we measure the average reddening for the cluster which we can then apply to colour-magnitude diagrams which are uncorrected for differential absorption to obtain distances. To determine the error we measure the standard deviation of the O and B type stars from the ZAMS via least squares fitting. The errors quoted on the values for the reddening (see Table 4) are typically 0.1 mags and include any error in the calibration. In quite a few cases, the ZAMS does not fit the colour-colour data well over the whole range of B-V colours. This is particularly problematic in some cases and these cases are discussed below. There is also the problem that the Cepheid could have a different reddening to the cluster, caused by differential reddening across the cluster or by the location of the Cepheid away from the cluster. This is discussed further in section 4. We use both the B-V:V and V-K:V colour-magnitude diagrams to determine the cluster distance. V-K:V diagrams have the advantage that the slope of the ZAMS is flatter than at B-V:V, possibly allowing more accurate distance estimates but V-K:V diagrams are available for only 7 of the clusters and the scatter in these diagrams is greater than in the B-V:V CMD’s, particularly in the case of NGC6664 as there were difficulties aligning the V and K frames due to a shift in the telescope position half way through creating the K-band mosaic for this cluster. There are also few points on the V-K:V diagram for NGC7790 due to the small size of the IRCAM detector. This diagram was also made for us before the V-band observations for this cluster were available as a test for the feasibility of this project. The V-band data therefore comes from ?, however there is a good match between our V-band data and that of Romeo et al. (see Figure 17). The distance modulus which best fits the B-V:V CMD around the position of the A0V stars, using the method of least squares, is taken to be the distance modulus, $`\mu _0`$, of the cluster. The errors on the distance modulus were found by measuring the standard deviation away from the ZAMS of A0V type stars over the range -0.1$`<`$B-V $`<`$0.1 in the dereddened ZAMS which covers a range of approximately 4 mags in V. The distance modulus found from the B-V:V CMD is then checked against the V-K:V CMD for the 7 clusters with such a CMD for consistency. In all cases the distance modulus found from the B-V:V diagram was consistent with the V-K:V CMD within the errors. No attempt has been made to remove foreground and background stars. Only stars which lie clearly off the main sequence (off in B-V by more than 1 mag for example) were removed. There is no clear recipe for how to remove the contaminating stars from the colour-magnitude and colour-colour diagrams so the data shown in Figures 11, 12 and 13 contain non-cluster members. The observations in this study only cover a field-of-view of $`6^{}`$ and are pointed at the cluster centre so contamination may not be as much of an issue as for the wider field photographic plates. As an example, we discuss this further for the cluster NGC7790 in section 3.2. All of the U-B:B-V diagrams, the B-V:V and the V-K:V diagrams are given in Figures 11, 12 and 13 respectively and our results for the reddening and distances obtained are given in Table 4 together with previous results as summarised by ?. All of the clusters are individually discussed below. ### 3.2 Discussion of Individual Clusters NGC6649 has been studied previously by, for example, ? and ?. The agreement between the photometry of this study and the photoelectric data of ? is good (see Table 3), with only small offsets in the U and B-band. ? find E(B-V) = 1.37 (no quoted error) for the reddening towards the cluster. The distance modulus $`\mu _{}`$ from ? is 11.15$`\pm `$0.7. ? uses the photometry of ? and that of ? to study NGC6649 and find the cluster suffers from differential reddening. However for stars close to the cluster centre a value of E(B-V)=1.38 is appropriate. The distance modulus is found to be 11.06$`\pm `$0.03 when individual stars are dereddened. ? used U, B and V-band CCD data to study NGC6649. Agreement between the photoelectric data of ? and ? was found to be better than 0.03 mags in the V-band. ? do not measure the reddening of the cluster due to the claims of differential reddening by Turner. ? deredden each star individually to find a distance modulus of 11.00$`\pm `$0.15. Rather than correct for differential reddening, we fit the ZAMS line to the centre of the U-B:B-V and B-V:V diagrams to try and measure the average values of the reddening. We find E(B-V)=1.37$`\pm `$0.07 and $`\mu _{}`$=11.22$`\pm `$0.32, both values are consistent with previous work. The distance modulus used in ? is slightly higher, 11.278 and the reddening,E(B-V), slightly lower, 1.35 but again these values are well within the errors. ? found the radial velocity of the Cepheid V367 Sct to be -20$`\pm `$6 kms<sup>-1</sup> and the radial velocity of the cluster NGC6649 to be -14$`\pm `$5 kms<sup>-1</sup>. This is taken to be evidence for the cluster membership of the Cepheid. M25 The photometry used in this study and that of ? is compared in Figure 14 and 15. The agreement in all the wavebands is good, less that 0.03 mags different from that of Sandage. M25 has a U-B:B-V diagram where the ZAMS fit is good over a wide range of B-V colours. The value for E(B-V) of 0.49$`\pm `$0.08 is in excellent agreement with the work of ? who obtained 0.49$`\pm `$0.05. ? stated that E(B-V) lay in the range 0.4 to 0.56 and ? obtained E(B-V)=0.51$`\pm `$0.01. The distance modulus obtained for this cluster is 9.05$`\pm 0.43`$ which is slightly higher than that of Sandage, who found $`\mu _{}`$=8.78$`\pm `$0.15. The difference in the distance modulus is probably due to where to fit was made, we fit to the A type stars to obtain 9.05$`\pm 0.43`$, Sandage’s work contains brighter stars which may lie slightly off the main sequence. Wampler et al (1960) obtained 9.08$`\pm 0.2`$ and ? found $`\mu _{}`$= 9.0$`\pm `$0.3 which agree well. There is also good agreement between our values and those used by ?, as given in Table 4. ? studied M25 by measuring radial velocities and spectra for stars within M25. He obtained radial velocities of around 4 kms<sup>-1</sup> for the Cepheid U Sgr and for 35 stars in the cluster M25 indicating that U Sgr is a member of M25. NGC6664 has been studied previously by ?. The agreement between his photometry and ours is good, with only small offsets between the two data sets as given in Table 3. Figure 11 shows that the cluster NGC6664 has an anomalous U-B:B-V diagram. This is a clear case where we fit the O and B type stars rather than trying to fit the F and G type stars as the F and G type stars maybe more affected by metallicity. The value for E(B-V) is then 0.66$`\pm `$0.08 which is slightly larger than 0.6 obtained by ?. No error is quoted by Arp. In the previous work by Arp, the observations were not deep enough to see if the anomalous shape of the U-B:B-V diagram would have been detected or not. Unfortunately the only previous source of photometry for more than a handful of stars is that of Arp so no other zero point comparisons can be made. The distance modulus obtained is 11.01$`\pm `$0.37 which is larger than 10.8 from ?, mostly due to the increased value for the reddening estimated here. There is an larger discrepancy between the value of the distance modulus used by ? who quote a reddening of 0.64 but a distance modulus of 10.405. It is not clear why their distance modulus is so low as ? use 10.88. The radial velocity work of ? shows EV Sct is a member of NGC6664. WZ Sgr The first problem with WZ Sgr is that its membership of an open cluster is questionable. ? discusses the membership of WZ Sgr to the cluster C1814-190 in some detail and concludes that the strongest evidence for membership of WZ Sgr to an open cluster comes from the fact that often the Cepheid in a cluster is around 4 magnitudes more luminous than the B-type stars on the main sequence. This essentially means that the age of the Cepheid is consistent with the age of the cluster. ? assume WZ Sgr to be a cluster member. The cluster C1814-190 is also only comparatively sparsely populated with only around 35 members brighter than B$``$16 ?. There is also patchiness in the dust obscuration which would cause differential reddening (?). These two factors could possibly explain the odd shape of the U-B:B-V colour-colour diagram particularly in the range 0.5 $`<`$ B-V $`<`$ 1. Contamination from foreground and background stars could also be the source of the unusual U-B:B-V diagram. By measuring the reddening of the O and B type stars though we get a value for E(B-V)=0.56$`\pm `$0.20 which is in agreement with ? and consistent with E(B-V)=0.57 used by ?. The B-V:V diagram is surprisingly tight, giving a distance modulus of 11.15$`\pm `$0.49 which is again in agreement with ? who obtained 11.16$`\pm `$0.1. The distance modulus used by ? is slightly larger, 11.219 but again well within the errors of our result. Lynga 6 The photometry of Lynga 6 has been checked against that of ? (shown in Figure 14 and 15) and also against that of ?. Offsets between the different sets of photometry are less than 0.03 mags in each waveband. Like NGC6649, Lynga 6 is also very heavily reddened which makes it very difficult to observe stars over a large range of B-V and U-B colours. E(B-V) is estimated to be 1.36$`\pm `$0.17 which is consistent with previous measurements by ? who obtained 1.34$`\pm `$0.01, by ? who obtained 1.37$`\pm `$0.03 and the value of 1.34 used by ?. The distance modulus obtained here is around 11.10$`\pm `$0.45 which is consistent with 11.15$`\pm `$0.3 obtained by ? but is discrepant with the value used by ?. They have a distance modulus of 11.429 which agrees reasonably well with the value of ? who obtained ($`VM_V`$)=16.2$`\pm `$0.5 for the apparent distance modulus which, if a value of 3.2 is assumed for the extinction coefficient, roughly implies 11.8$`\pm `$0.5 for the distance modulus (see section 4 for more details of the extinction coefficient used here). The difference between the value of Madore and the value for the distance found here is that Madore tended to fit the edge of the ZAMS. The Cepheid TW Nor lies close to the centre of the cluster Lynga 6 and has a very similar value of the reddening. This is taken as evidence of the membership of the Cepheid to the cluster (?). NGC6067 There is very little photoelectric data for this cluster. The B and V-bands have been compared to the data of ? but there are only 7 stars in common and in the U-band there are only two stars in common. The comparison shows that the zero point used here is at least consistent with previous the previous work. As the photometry obtained at CTIO for other clusters such as Lynga 6 and vdBergh1 agrees well with previous results, we have to assume that the photometry for NGC6067 is also good. ? finds good agreement between their B and V-band CCD data and the photoelectric data of ?. NGC6067 has the lowest value for the reddening, of the clusters in this study, with E(B-V) estimated as 0.42$`\pm `$0.02. The U-B:B-V diagram presented here has a main sequence which agrees fairly well with the ZAMS, although there is a spread around B-V=0.8. The value of 0.42 is slightly higher than previous values, ? obtained 0.35$`\pm `$0.1 and ? obtained 0.33 with no quoted error. As the reddening has been estimated to be slightly higher than previously, the distance modulus is also increased to 11.17$`\pm `$0.35 as opposed to 11.05$`\pm `$0.1 from Walker. However ? found the distance modulus of the cluster to be around 11.3. This estimate is higher than our estimate, despite a smaller measured reddening, as the ZAMS fit was made to the edge of the CMD. ? use 0.35 for the reddening and 11.13 for the distance modulus to the cluster. The membership of the Cepheids to this cluster is discussed in detail by ?. Eggen noted that the Cepheid V340 Nor is centrally located in the cluster and has the same reddening as the cluster so is assumed to be a member. The Cepheid QZ Nor lies at a distance of two cluster radii out from the cluster centre (?) but is still assumed to be a cluster member by ? vdBergh1 Our photometry is tested against the photoelectric data of of ? in Table 3. There is no significant offset between the two data sets and no evidence of any scale dependent error. ? has compared their photometry to that of ? and find good agreement. The cluster vdBergh1 was only given a short exposure of 300s in U and 90s in B and V. The U-B:B-V diagram is therefore not very well populated at faint U magnitudes. There is a fair amount of scatter in the U-B:B-V diagram for this cluster, the average value is E(B-V)=0.9$`\pm `$0.18. This is larger than the value of previous estimates. ? obtained a minimum value of E(B-V)=0.66, this clearly fits the the edge of the B-type stars. ? obtained 0.76 (no error) but the spread in the U-B:B-V diagram is such that the larger value would also have been acceptable. 0.77 was used for the cluster reddening by ? The distance modulus, $`\mu _{}`$ obtained here is 11.4$`\pm `$0.65. This is larger than previous results, 10.94 obtained by ? and 11.08 obtained by ?. This is mostly due to the measurement of increased reddening. The value used by ? was 11.356. The membership of CV Mon to the cluster has been determined by ? using radial velocity measurements, evolutionary arguments and by it’s location in the cluster. Trumpler 35 There is very little photoelectric data available for this cluster. We compare our photometry against the photoelectric observations of ? and find reasonable agreement in the B and V-bands. The agreement in the U-band is less good but as there is only one source of comparison and the JKT photometry appears to agree well for other clusters we suggest that the U-band data photometry is accurate. The data in the U-B:B-V diagram for the cluster Trumpler 35 (TR35) shows quite a large spread. This is probably caused by differential reddening. Rather than trying to correct for this, a mean value of E(B-V)=1.19$`\pm `$0.10 is taken for the reddening. Shown on the TR35 panel in Figure 11 are two lines. The dashed shows the value of E(B-V)=1.03 taken from ? and the solid line shows the value of 1.19 adopted here. Using this, the distance modulus to the cluster is then estimated as 11.3$`\pm `$0.53 by fitting up the centre of the data. This is different to the value of ? who obtained 11.6$`\pm 0.16`$ as Turner fitted the edge of the B-V:V diagram rather than the centre. The values for the reddening used by ? (E(B-V)=0.92, $`\mu _{}`$=11.56) are in better agreement with those of ? rather than those obtained here. The membership of RU Sct to Trumpler 35 using the reddening and evolutionary status of the Cepheid is discussed and supported by ?. However RU Sct does lie 15 away from TR35 (?). NGC6823 was only observed in non-photometric conditions. We use the photoelectric data of ? for calibration purposes. ? has compared his photometry with that of Hiltner (1956) and Hoag (1961) and finds that there are only small offsets of less than 0.04mags in U-B between the different data sets. ? finds good agreement with the photometry of Hiltner (1956) which agrees well with the photometry of ? used for calibration here. NGC6823 suffers from differential reddening (?), perhaps to an even greater extent than Trumpler 35. Again shown in the panel for the cluster NGC6823 in Figure 11 are two lines. One is for E(B-V)=0.53 taken from ? and the other is E(B-V)=0.85 which fits the centre of the B-type stars. The value of E(B-V)=0.85$`\pm `$0.09 is assumed here. The distance modulus with this reddening is then 11.20$`\pm `$0.55, lower than the value of 11.81 found by ? who again fitted the edge rather than the centre of the B-V:V diagram. ? assume a reddening of E(B-V)=0.44 and $`\mu _{}`$=11.787. These values again agree fairly well with those of ? rather than the values found here. The Cepheid SV Vul is only thought to be associated with the cluster NGC6823 (?) as it lies a few arcmins away from the cluster centre. NGC129 was only observed in non-photometric conditions so previous work had to be relied upon for the calibration, as discussed in detail in section 2.4.1. The reddening for this cluster is E(B-V)=0.57$`\pm `$0.06. This value is slightly larger than previous values, ? obtained 0.47 but fits to the least reddened edge of the O and B stars rather than to the centre of these stars. ? found 0.53 (with no error) for the reddening which is in agreement with the value found here. The distance modulus obtained assuming a value of 0.57 for the reddening is $`\mu _{}`$=10.90$`\pm `$0.37. This value is lower than the value of 11.11 obtained by ? who again fits the edge of the main sequence rather than the centre. ? find 11.0$`\pm `$0.15 for the distance modulus of the cluster, in agreement with the value found here. DL Cas is not included in the study by ? due to the Northerly latitude of the cluster NGC129. The values for the reddening and the distance modulus used by ? agree better with the values of ? rather than those obtained here. ? found a measurement of -14$`\pm `$3 kms<sup>-1</sup> for the radial velocity of the cluster NGC129. He found that the Cepheid itself had a radial velocity of -11km$`s^1`$. Given that the error on any individual measurement is estimated to be around 1.5kms<sup>-1</sup>, DL Cas is assumed to be a member of NGC129. NGC7790 The U-B:B-V diagram for this cluster appears quite clean and well defined. However, Fig. 11 shows that NGC7790 has a U-B:B-V diagram where the data poorly fits the ZAMS line. We show one ZAMS shifted to fit the OB stars which implies E(B-V)= 0.59$`\pm `$0.04 and another shifted to fit the F stars which would imply E(B-V)=0.43$`\pm `$0.04. Most previous estimates are closer to that for the OB stars. This poor fit of the ZAMS to the U-B:B-V data in the case of this cluster is particularly significant since it contains three Cepheids (see Table 1). The U-B:B-V diagram for NGC7790 has a history of controversy. The original UBV photoelectric photometry of Sandage (1958) of 33 11$`<`$V$`<`$15 stars was criticised by Pedreros et al (1984). A check of 16 stars with the KPNO CCD seemed to confirm that Sandage’s U-B and B-V colours were too blue by +0.075 and +0.025 mag respectively, although few details were given of errors etc. However we find excellent agreement between the work here and the photometry of Sandage, see Table 3, Fig. 14 and Fig 15. A direct comparison of the photoelectric observations of Sandage to the photographic observations in Pedreros (see Table 1 in Pedreros et al. 1984) implies that the differences in the colours of Sandage are too blue by $``$ 0.04 mags for both the U-B and B-V colours and when we compare our CCD photometry to 22 of the brightest photographically observed stars from Pedreros, we find similar colour differences of around 0.04 mags in both the U-B and B-V colours. The U-B:B-V diagram in ? seemed to give the same sort of ill fitting ZAMS, throughout the range 0.3$`<`$B-V$`<`$1.2, as found in this work. The suggestion was that the problem might lie in Sandage’s photometry which Pedreros et al had used for calibration. However, we have tested various zero points for the U-B and B-V colours. With our own zero points for the colours we find the ill fitting ZAMS and even if we apply the corrections suggested by Pedreros to the offsets found between our photometry and the photographic data of Pedreros we still find an ill fitting ZAMS. Romeo et al (1984) used CCD data to obtain BVRI photometry for this cluster to V=20. In the absence of U, their only route to E(B-V) was via fitting the shape of the B-V:V and V-I:V CMD and they obtained E(B-V)=0.54$`\pm `$0.04. They used the Sandage (1958) photometry for calibration in B and V, subject to the small re calibration in these bands by Pedreros et al. and checked against the B,V photoelectric photometry of 10 stars by Christian et al (1985). They tested the faint photographic photometry of Pedreros et al and found scale errors at B$`>`$17 and V$`>`$15 in the sense that Pedreros et al were too bright. In B-V, however they claimed better agreement with B-V<sub>Pedreros</sub> being $``$0.1mag too red. A comparison of our photometry and that of Romeo suggests that there is a scale error for B$`>`$17 and V$`>`$16 but the extent of this is less than in the comparison of the Pedreros et al. data with the Romeo data, at V=17 our data is brighter than Romeo’s by 0.1 mag whereas the Pedreros et al. data is brighter by 0.2 mags and similar differences are found in the B data. We conclude that the photometry in this study agrees very well with the photometry of ? and is better agreement with the CCD data of ? than that of Pedreros, (see Fig. 17 for more details). We have also attempted to check to see if contamination is causing the ill fitting ZAMS. Figure 6 shows the B-V:V and the U-B:B-V diagrams for the cluster NGC7790. The B-V:V diagram has been trimmed so that only the stars that lie very close to the ZAMS remain. As these stars have the correct combination of distance and reddening to lie almost on the main sequence then it is likely that they are main sequence stars. The same stars are then used to produce the U-B:B-V colour-colour diagram. The same UV deficit around the F-type stars that appears in Figure 11, when all the stars in the field are used, appears in Figure 6. Figure 7 shows the colour-colour diagram found from this work with the points from Sandage and Pedreros included. The Sandage points do not go deep enough to test the shape of the data but the Pedreros et al. points do and the poor match of the data to the ZAMS is seen. Therefore, we believe the ill fitting ZAMS to the U-B:B-V diagram is not caused by contamination from foreground or background stars or by errors in the photometry but is a real feature in the data. Thus our estimate of the NGC7790 reddening based on the U-B:B-V colours of OB stars, E(B-V)=0.59$`\pm `$0.05, is between the E(B-V)=0.52 $`\pm `$0.04 of Sandage (1958) and the E(B-V)=0.63$`\pm `$0.05 of Pedreros et al(1984) who used similar techniques. Assuming the estimate of E(B-V)=0.59$`\pm `$0.05 from the OB stars, the distance modulus which fits V:B-V is then $`\mu _{}`$=12.72$`\pm `$0.11 which is close to the original value of 12.8$`\pm `$0.15 found by Sandage(1958), although this is in the opposite direction that would be expected from the difference in reddening and slightly more than 12.65 obtained by ? which is roughly in line with their obtaining E(B-V)=0.54 for the reddening. ? obtained only 12.3 for the distance modulus, no fits are shown in the paper and the reason they obtain this low value is unclear. The values quoted by ? are similar to those of ?. ? states that the membership of CF Cas, CEa Cas and CEb Cas to NGC7790 is almost certain due to the position of the Cepheids on the CMD. ## 4 P-L Relation Using the values for the distance modulus and the reddening towards the cluster, we proceed to determine the P-L relation. As well as the reddening and the distance modulus of the cluster, the apparent magnitude of each of the Cepheids is required. Where possible these come from LS (1993, 1994) who have high quality V and K-band measurements for most of the Cepheids in this study. The clusters NGC7790 and NGC129 lie at northerly latitudes so are unobservable from SAAO and so there are no magnitudes from Laney and Stobie for these Cepheids. The K-band data for DL Cas and CF Cas comes therefore from ? and the V-band data and the periods are taken from ?. The reddening obtained from the ZAMS fitting is that of the cluster OB stars. ? found that when the effect of the colour difference between the OB stars and the Cepheid is taken into account, $`\mathrm{E}(BV)_{\mathrm{ceph}}`$ $`=`$ $`\mathrm{E}(BV)_{\mathrm{clus}}[0.98`$ (2) $`0.09(<B_o><V_o>)_{\mathrm{ceph}}]`$ gives a good approximation to the reddening of the Cepheid. The values for $`<B_o><V_o>)_{\mathrm{ceph}}`$ come from ?. Figure 8 shows the M<sub>V</sub>-M<sub>K</sub> \- Log(P) relation with the reddenings found for the Galactic Cepheids (filled circles) found from the cluster reddenings obtained in this work via equation 2. The triangles show the same relation for the LMC Cepheids and the squares are for the SMC Cepheids. These are taken from Tables 2 and 3 in ?. Three of the Cepheids, (from left to right in Figure 8) QZ Nor, RU Sct and SV Vul seem to have the wrong M<sub>v</sub>-M<sub>k</sub> colours for their periods. The clusters containing RU Sct and SV Vul suffer both from the presence of differential reddening and from the fact that the Cepheids lie at some distance away from the cluster. Figure 8 indicates that the reddening local to RU Sct and SV Vul may be somewhat different from the average value of the cluster reddening. We correct for this using the space reddenings given in ? which are more local to the Cepheids (see for example Turner 1980). Note that because there is difficulty obtaining the Cepheids true reddening from the cluster reddening for these two Cepheids we do not include them in the best sample (later in this section). ? notes that the Cepheid QZ Nor also lies away from the centre of the cluster NGC6067, at a distance of two cluster radii so again the reddening of the cluster may not be appropriate for the reddening of the Cepheid. ? take the value for the Cepheid reddening from ? of E(B-V) = 0.265, derived from BVI<sub>c</sub> reddenings. This is the reddening used to calculate the position of the open circle in Figure 8 and which we assume for QZ Nor henceforth. However, the effect of changing the reddening of the cluster to E(B-V) = 0.265 changes the distance modulus to the LMC by less than 0.02, less than the error quoted here. To determine the absolute magnitude of the Cepheid, the apparent magnitude has to be corrected for reddening and distance. First of all the extinction coefficient is required. We follow ? and use $$\mathrm{}(\mathrm{ceph})=3.07+0.28(BV)_{}+0.04\mathrm{E}(BV)_{\mathrm{ceph}}$$ (3) to take into account the effect of Cepheid colour on the ratio of total to selective extinction. The Cepheid reddening comes from Eq. 2 except for the three corrected values. Then the reddening free magnitudes of the Cepheids are $`V_{}`$ $`=`$ $`V\mathrm{}(\mathrm{ceph})\mathrm{E}(BV)_{\mathrm{ceph}}`$ $`K_{}`$ $`=`$ $`V_{}V+K+{\displaystyle \frac{\mathrm{}(\mathrm{ceph})\mathrm{E}(BV)_{\mathrm{ceph}}}{1.1}}`$ (4) The expression for $`K_{}`$ has the form given above as the extinction coefficient in the K-band is one tenth of that in the V-band. To obtain finally the absolute magnitude, the distance modulus, $`\mu _{}`$, given in Table 4 has to be subtracted off. The P-L relation can now be determined. We consider two samples, one where we consider all the Cepheids available to us and another where the Cepheids RU Sct and SV Vul are removed due to the problem of differential reddening and the question of cluster membership. The zero points are obtained by fixing the slope and obtaining the least squares solution using the Galactic Cepheids in this study. These are summarised in Table 6. The slopes that are considered are the slopes from ? which are the best fitting slopes to all the Cepheid data (Galactic open cluster Cepheids, LMC and SMC Cepheids) in their study. The slopes are -2.874 in the V-band and -3.443 in the K-band. Also considered is -2.81 in the V-band as this is the slope of the LMC Cepheids and the slope used by ?. Once the slope and zero point of the PL relation is fixed, the distance modulus to the LMC can be calculated. ? give the period and the dereddened V and K-band magnitude, V and K of 45 LMC Cepheids. The distance to the LMC is then given by $$<\mu _{}>=\frac{1}{\mathrm{N}}\underset{i}{\overset{\mathrm{N}}{}}(m_{}\delta \mathrm{log}(\mathrm{P}))\rho $$ (5) where $`m_{}`$ represents the dereddened apparent magnitude in each waveband and $`\delta `$ and $`\rho `$ are the values for the slope and zero point as given in Table 6. Our PL(V) and PL(K) relations are shown in Figure 9 along with the equivalent relation for the LMC Cepheids from ? using the distance moduli in Table 6 to determine the absolute magnitude. Taking the value for the PL(V) zeropoint for the best sample with the -2.874 slope used by Laney & Stobie (1994) gives $`\rho `$=-1.279$`\pm `$0.33 which is in good agreement with the value of $`\rho `$=-1.197$`\pm `$0.09 found by these authors and implies a distance modulus of 18.57 for the LMC as compared to 18.50 found by Laney & Stobie. Figs. 8(a,b) show that the best fitting zeropoint gives a somewhat poor fit to the majority of the Galactic Cepheids which lie at log(P)$`<`$1.0. This is because the 12 Galactic Cepheids in the best sample give a slope of $`\delta `$=-1.85$`\pm `$0.33 which is much flatter than the LMC data which gives $`\delta `$=-2.79$`\pm `$0.1, close to -2.81. Indeed, if only the 9 Galactic Cepheids with log(P)$`<`$1.0 are used, the zeropoint rises to $`\rho `$=-1.418$`\pm `$0.19 and the LMC distance modulus would rise to 18.70, indicating why the errors on the PL(V) zeropoint are as large as they appear in Table 5. Our PL(V) zeropoint is also consistent with the zeropoint and LMC distance obtained from an analysis of Hipparcos trigonometrical parallaxes of nearby Galactic Cepheids by ?. They obtained $`\rho `$=-1.43$`\pm `$0.1 for the Galactic PL(V) zeropoint for an assumed slope of $`\delta `$=-2.81. This can be compared to the $`\rho `$=-1.332$`\pm `$0.32 obtained for our best sample with the same slope. They used the same 45 Laney & Stobie (1994) Cepheids as used here to obtain a metallicity corrected LMC distance modulus $`\mu _{}`$=18.70$`\pm `$0.10. We note that their semi-theoretical metallicity correction to the LMC Cepheid V magnitudes increases the distance to the LMC, which is in the opposite sense to most empirically determined estimates of the effects of metallicity on Cepheids (eg Kennicutt et al 1998). Subtracting their metallicity correction leads to an LMC distance modulus $`\mu _{}`$=18.66$`\pm `$0.10 which can be directly compared with our best value of $`\mu _{}`$=18.55$`\pm `$0.036 from Table 6. We conclude that our PL(V) estimates the LMC distance modulus are between those of Laney & Stobie and Feast & Catchpole but have too little statistical power to discriminate between these previous estimates. The K-band P-L relation is tighter for the LMC Cepheids and for the Galactic Cepheids, see Figure 9(c), and the slopes are closer with the LMC Cepheids giving $`\delta `$=-3.27$`\pm `$0.04 and the best sample of Galactic Cepheids giving $`\delta `$=-2.81$`\pm `$0.21. We note in passing that the slope of the Galactic Cepheid PL(K) relation is now much flatter than the $`\delta `$=-3.79$`\pm `$0.1 slope found in the Galactic Cepheid sample of Laney & Stobie. Assuming the -3.443 slope used by Laney & Stobie our best sample in Table 5 gives a PL(K) zeropoint of $`\rho `$=-2.200$`\pm `$0.29 which implies an LMC distance of $`\mu _{}`$=18.47$`\pm `$0.29 which remains in good agreement with the value $`\mu _{}`$=18.56$`\pm `$0.07 found by Laney & Stobie. (1994). The smaller error of Laney & Stobie is due to their larger numbers of calibrators although it must be said that many of their extra calibrators (8/12) are in associations rather than clusters and frequently given half-weight in P-L fits. Indeed, Hipparcos proper motion data has shown that one of their further cluster Cepheids, S Nor, is also unlikely to be a member of its cluster (Haguenau conference, September 1998). Moreover, they have not included the 4 cluster Cepheids in NGC129 and NGC7790. Therefore we believe that our result supercedes the Laney & Stobie result with our bigger error estimate perhaps being a more realistic indication of the actual errors. Certainly in our best sample the biggest changes in M<sub>K</sub> between Laney & Stobie and ourselves, which contribute most to our 50% increased scatter, are for EV Sct and TW Nor (see Table 4) where the reasons for the distances used by Laney & Stobie are unclear. Finally, as our overall estimate of the LMC distance, we take the average of the PL(V) and PL(K) estimates in the best sample of Table 5 which gives $`\mu _{}`$=18.51$`\pm `$0.3. We conclude that although in the case of individual clusters we have markedly improved the distance and reddening estimates, our new estimates of the zeropoint of the PL relation and thus the distance to the LMC are close to previous values. ## 5 Discussion We now discuss the most intriguing new result in this study, which is that the Solar metallicity ZAMS may not always fit the U-B:B-V data in individual clusters. This is not the first time an effect like this has been seen. ? saw poorly fitting U-B:B-V ZAMS for the open cluster Roslund 3. Turner interpreted this as evidence for the young, B-type stars having a cocoon of circumstellar dust around them. This cocoon of dust then increases the reddening of the B-type stars as compared to the F and G type stars, causing the ill-fitting U-B:B-V ZAMS. The O and B type stars in the U-B:B-V diagram of Roslund 3 show a large spread around the ZAMS as the amount of excess dust would probably vary from star to star. Excess reddening could perhaps also be so strong that it was causing some O and B stars to be so reddened that they appeared as F type stars. However, the O and B type stars in the U-B:B-V diagram for NGC7790 are very tight so the shape of the U-B:B-V diagram is unlikely to be caused by excess dust. The next possibility we consider is that the effect might be due to stellar evolution, However, the CMD for the clusters look unevolved even at AOV as might be expected for clusters which have Cepheid variables and are expected to be less than 10<sup>8</sup> years old. We also considered whether the discrepancy between the main sequence fitted distance and the Hipparcos parallax to the Pleiades could explain our result. If it were assumed that all open clusters had roughly the same composition then the different forms for U-B:B-V that we find might be taken as evidence that the colours of main sequence stars may not be unique. This possibility has also been discussed as an explanation of the problem with the MS fitted distance to the Pleiades (?) and if it proves relevant in that case it will certainly also be worthy of further consideration here. If the reddening vector in U-B:B-V varied as a function of Galactic position then this would also affect our results. However, at least in the case of NGC7790 it seems that for whatever relative shift in U-B and B-V, the ZAMS still has the wrong shape to fit the observed colour-colour relation. The final possibility is that metallicity is affecting the F stars’ U-B colours in some of these clusters. Qualitatively there is some evidence supporting this suggestion. First, ‘line blanketing’ is well known to redden the U-B colours of metal rich stars at F and G and low metallicity sub-dwarfs are known to show UV- excess as the reverse of this case (e.g. ?). Second, there is some suggestion that the cluster, NGC7790, that shows a UV excess lies outside the solar radius while NGC6664 which is redder in U-B at F tends to lie inside (see Figure 10). M25 which lies closest to the Sun also fits the solar metallicity U-B:B-V diagram as well as any of the clusters. Given that metallicity in the Galaxy is known to decrease with Galactocentric radius, this is suggestive of a metallicity explanation. Many of the other clusters’ U-B:B-V diagrams are either too noisy due to differential reddening (Tr35, NGC6823) or too obscured to reach the F stars (NGC6649, Lynga 6, vdBergh1) to further test this hypothesis. However, NGC6067 forms a counter-example to any simple gradient explanation, since it seems to have a normal UBV plot and lies inside the solar position This would have to be accommodated by allowing a substantial variation on top of any average metallicity gradient. However, quantitatively the case for metallicity is less clear. The size of the UV excess seen is much larger in the case of NGC7790 than expected on the basis of previous metallicity estimates of these clusters, or of any measurement of the amplitude of the Galactic metallicity gradient. Using the Fe/H vs $`\mathrm{\Delta }`$ U-B relations of ? or ? it would be concluded that NGC7790 showed $`\mathrm{\Delta }`$ U-B $``$ 0.2mag which corresponds to Fe/H$``$-1.5. Thus clusters which on the basis of their Main Sequences and the presence of Cepheids, must be less than 10<sup>8</sup> yr old, would be implied to have near halo metallicity. Previously, ? find Fe/H$``$-0.3 for these 2 clusters. Also ? find Fe/H=-0.2$`\pm `$0.02 for NGC7790 while finding Fe/H=-0.37$`\pm `$0.03 for NGC6664 based on spectroscopy of the Cepheids in these clusters themselves. Also according to the Galactocentric metallicity gradient which is usually taken to lie in the range -0.02-0.1dex kpc<sup>-1</sup> (?), there should only be on the average $`\mathrm{\Delta }`$ Fe/H $``$ 0.3 in the range of metallicity covering these clusters. On the other hand, it should be noted that Panagia & Tosi’s Fe/H estimates are based on more poorly measured estimates of UV excess than those presented here and also that there is little agreement between the metallicity estimates of Fry and Carney and those of Panagia & Tosi. Measuring the metallicity of the Cepheids themselves as attempted by Fry and Carney is difficult since the effective temperature is a function of the light curve phase and a small difference in estimated temperature can make a large difference in metallicity. Also the Galactocentric metallicity gradient at least as measured for open clusters depends on relatively poor U-B photometry at the limit of previous data from ?, ? and ?. In any case, it is well accepted that the dispersion in metallicity around the mean gradient is indeed high, with the range -0.6$`<`$Fe/H$`<`$+0.3 at the Solar position. Further ? using Washington photometry to estimate metallicity also found an example of a cluster, NGC 2112, only $``$ 0.8kpc outside the solar radius with Fe/H=-1.2, although this cluster is older than those discussed here. However, it would also seem that the tightness of the P-L relations in Fig. 9 could form a final argument against the idea that NGC7790 has Fe/H$``$-1.5. If the metallicity of the cluster NGC7790 was really Fe/H$``$-1.5 then at given B-V, MS stars would be sub-dwarfs with $``$1mag fainter absolute V magnitudes than normal solar metallicity main sequence stars (see Cameron, 1984, Figs. 5,6). Thus since we have used a normal Main Sequence to derive the distance to NGC7790, is it not surprising that the Cepheids in these clusters lie so tight on the P-L relation when they should be a magnitude too bright if the low metallicity hypothesis is correct. The only way that the low metallicity hypothesis for NGC7790 could survive this argument is if it were postulated that the effect of metallicity on Main sequence star magnitude and Cepheid magnitude were the same - then the effect of our derived distance modulus being $``$1 magnitude too high would be cancelled out by the fact that the Cepheid is actually sub-luminous by 1 magnitude because of metallicity which would leave the Cepheid tight on the P-L relation as observed. This might not be too contrived if a low metallicity Cepheid prefers to oscillate about its subdwarf, rather than solar metallicity, zero-age luminosity (at fixed effective temperature) on the Main Sequence. This would lead to a strong implied metallicity effect on the Cepheid PL(V) and PL(K) zeropoints; the implication would be that $`\frac{\delta M}{\delta Fe/H}`$0.66 in the sense that lower metallicity Cepheids are fainter. This coefficient is within the range that has been discussed for the empirical effects of metallicity on Cepheids by ? and ? although the most recent work by ? appears to give a lower value of $`\frac{\delta M}{\delta Fe/H}`$0.24$`\pm `$0.16 again in the same sense. The immediate effect on the distance to the LMC with Fe/H=-0.3 is that our estimate of its distance modulus would decrease from 18.5 to 18.3. However, since all that is determined at the LMC is the slope of the P-L relation, then the zeropoints we have derived in Table 5 from the Galactic Cepheids would still refer the P-L relation to the Galactic zeropoint. The ultimate effect on H would then be decided by the metallicity of the Cepheids, for example, in the galaxies observed by the HST for the Distance Scale Key project ?, by ? in the case of the Leo I Group and by ? in the case of SNIa. ? have measured metallicities in these galaxies already but if the dispersion in Cepheid metallicity is as large as it is implied to be in the Galaxy then there may be some signature in a wider dispersion in the Cepheid P-L relations in at least the high metallicity cases. The possibility of detecting this signature is currently being investigated (Shanks et al 2000 in prep.) Obviously the most direct route to checking the metallicity explanation for the anomalous behaviour seen in the U-B:B-V diagrams is to obtain medium-high dispersion spectroscopy for a sample of F stars in NGC7790 and NGC6664 to determine the metallicity directly for these main sequence stars. Currently proposals are in to use WHT ISIS spectrograph for this purpose. ## 6 Conclusions We have presented colour-colour diagrams and colour-magnitude diagrams of a sample of galactic clusters which contain or are associated with Cepheids. All the clusters have been observed using similar methods and the data reduction and extraction has also been done with similar techniques. The use of the improved U-band data has allowed powerful new checks of previous E(B-V) estimates over a wide range of magnitudes. In order to estimate the reddenings and distance moduli, we have fitted all the clusters in the same manner and have not attempted to correct for differential reddening but instead taken a simpler approach and fitted the average reddening value of the cluster. In most cases the differences that we have found between values for the reddening and distance modulus are small and where there are significant differences these can mostly be explained by comparing whether the ZAMS fit was made to the centre or to the edge of the colour-colour and colour-magnitude diagrams. The Cepheid P-L relations found from fitting the best sample are M<sub>V</sub>=-2.81$`\times `$log(P)-1.332 and M<sub>K</sub>=-3.44$`\times `$log(P)-2.20 and a distance modulus to the LMC of 18.54$`\pm `$0.32 in the V-band and 18.46$`\pm `$0.29 in the K-band giving an overall distance modulus to the LMC of 18.50$`\pm `$0.3, ignoring any possible effect of metallicity. These results for both the PL relations and the LMC distance are consistent with the previous results of Laney & Stobie (1994) although the improved distances and reddenings have increased the errors over what was previously claimed. These increased errors mean that our result for the PL(V) relation are also consistent with the result of Feast & Catchpole (1997) from Hipparcos measurements of Cepheid parallaxes, although this gives rise to an LMC distance modulus of $`\mu _{}`$= 18.66$`\pm `$0.1 as opposed to our $`\mu _{}`$=18.51$`\pm `$0.3. With the improved U-band data, we find that for at least two of the clusters, the data in the U-B:B-V two-colour diagram is not well fitted by the solar metallicity ZAMS. One possibility is that significant metallicity variations from cluster to cluster may be affecting the U-B colours of F- and G-type stars. The problem is that the metallicity variations this would require are much larger than expected for young, open clusters. More work is therefore required to determine the metallicity of the individual main sequence stars in each of the clusters NGC7790 and NGC6664. If metallicity is proven to be the cause of the anomalous U-B:B-V relations, then it would imply that the Cepheid P-L relation in both the visible and the near-infrared is strongly affected by metallicity. ## Acknowledgments FH acknowledges the receipt of a PPARC studentship. We thank Floor van Leeuwen for useful discussions and we thank Henry McCracken and Nigel Metcalfe for help with the Calar Alto data reduction and Patrick Morris for the WHT reduction. We thank the PATT telescope committee for the time on the UKIRT, JKT and WHT telescopes. We thank CTIO for supporting this project and the Calar Alto time allocation committee.
warning/0002/hep-ph0002094.html
ar5iv
text
# LA-UR-00-600 ## I Introduction Recently, Isgur has re-emphasized the experimental fact that spin-orbit splittings in meson and baryon systems, which might be expected to originate from one-gluon-exchange (OGE) effects between quarks, are absent from the observed spectrum. He conjectures that this is due to a fairly precise, but accidental, cancellation between OGE and Thomas precession effects, each of which has “splittings of hundreds of MeV” . Taking the point of view that precise cancellations reflect symmetries rather than accidents, we have examined what dynamical requirements would lead to such a result. One of us recently observed that a relativistic symmetry is the origin of pseudospin degeneracies first observed in nuclei more than thirty years ago. We find that a close relative of that dynamics can account for the spin degeneracies observed in hadrons composed of one light quark (antiquark) and one heavy antiquark (quark). Below, we first elucidate the experimental evidence for small spin-orbit splittings. Then we identify the symmetry involved in terms of potentials in the Dirac Hamiltonian for heavy-light quark systems, and note the relation to the symmetry for pseudospin. We show that the former symmetry predicts that the Dirac momentum space wavefunctions will be identical for the two states in the doublet, leading to a proposed experimental test. Finally, we argue that the required relation between the potentials may be plausible from known features of QCD. ## II Experimental and Lattice QCD Spectrum In the limit where the heavy (anti)quark is infinitely heavy, the angular momentum of the light degrees of freedom, $`j`$, is separately conserved. The states can be labelled by $`l_j`$, where $`l`$ is the orbital angular momentum of the light degrees of freedom. In non–relativistic models of conventional mesons the splitting between $`l_{l+\frac{1}{2}}`$ and $`l_{l\frac{1}{2}}`$ levels, e.g. the $`p_{\frac{3}{2}}`$ and $`p_{\frac{1}{2}}`$ or $`d_{\frac{5}{2}}`$ and $`d_{\frac{3}{2}}`$ levels, can only arise from spin-orbit interactions. The $`p_{\frac{1}{2}}`$ level corresponds to two degenerate broad states with different total angular momenta $`J=j\pm s_Q`$ (here $`j=\frac{1}{2})`$, where $`s_Q`$ is the spin of the heavy (anti)quark. For example, in the case of $`D`$-mesons, $`s_Q=\frac{1}{2}`$ and the two states are called $`D_0^{}`$ and $`D_1^{}`$. There are also two degenerate narrow $`p_{\frac{3}{2}}`$ states $`D_1`$ and $`D_2^{}`$. The degenerate states separate as one moves slightly away from the heavy quark limit, and their spin-averaged mass remains approximately equal to the mass before separation. For the $`D`$–mesons, the CLEO collaboration claims a broad $`J^P=1^+`$ state at $`2461_{34}^{+41}\pm 10\pm 32`$ MeV, belonging to the $`p_{\frac{1}{2}}`$ level, in close vicinity to the $`D_2^{}`$ at $`2459\pm 2`$ MeV, belonging to the $`p_{\frac{3}{2}}`$ level, indicating a remarkable $`p_{\frac{3}{2}}`$-$`p_{\frac{1}{2}}`$ spin-orbit degeneracy of $`2\pm 50`$ MeV. It is appropriate to extract the spin-orbit splitting this way since: Firstly, the charm quark behaves like a heavy quark. Secondly, the difference between the $`D_1^{}`$ and $`D_2^{}`$ levels is the best indicator of the $`p_{\frac{3}{2}}`$-$`p_{\frac{1}{2}}`$ splitting in the absence of experimental data<sup>*</sup><sup>*</sup>*The FOCUS collaboration preliminarily found $`D_0^{}`$ at a mass of $`2420`$ MeV . The error on the mass was not reported. on the $`D_0^{}`$, as opposed to the difference between the $`D_1^{}`$ and spin-averaged $`p_{\frac{3}{2}}`$ level at $`2446\pm 2`$ MeV. Spin-averaged masses are determined from experiment. For the $`K`$-mesons, the $`p_{\frac{1}{2}}`$ level is at $`1409\pm 5`$ MeV, with $`p_{\frac{3}{2}}`$ nearby at $`1371\pm 3`$ MeV, corresponding to a $`p_{\frac{3}{2}}`$-$`p_{\frac{1}{2}}`$ splitting of $`38\pm 6`$ MeV. The splitting between the higher-lying $`d_{\frac{5}{2}}`$ and $`d_{\frac{3}{2}}`$ levels is $`4\pm 14`$ MeV or $`41\pm 13`$ MeV, depending on how the states are paired into doublets. These results indicate a near spin-orbit degeneracy if the strange quark can be treated as heavy, although it has certainly not been established that such a treatment is valid. For $`B`$-mesons, both L3 and OPAL have performed analyses, using input from theoretical models and heavy quark effective theory, to determine that the $`p_{\frac{3}{2}}`$-$`p_{\frac{1}{2}}`$ splitting is $`97\pm 11`$ MeV (L3) or $`109\pm 14`$ MeV (OPAL). Note that these are not model-independent experimental results. In the same analyses the mass difference between the $`B_2^{}`$ and $`B_0^{}`$, an approximate indicator of the $`p_{\frac{3}{2}}`$-$`p_{\frac{1}{2}}`$ splitting, is $`110\pm 11`$ MeV (L3) or $`89\pm 14`$ MeV (OPAL). The L3 result agrees with lattice QCD estimates of $`155_{13}^{+9}\pm 32`$ MeV and $`183\pm 34`$ MeV. However, according to other estimates, the splitting is less than 100 MeV, and consistent with zero. Recently, $`31\pm 18`$ MeV was calculated . One lattice QCD study found evidence for a change of sign in the splitting somewhere between the charm and bottom quark masses, albeit with large error bars . A splitting of 40 MeV serves as a typical example of model predictions , although there is variation in the range -155 to 72 MeV , summarized in ref. . In order to more quantitatively measure the spin-orbit splitting, define $$r=\frac{(p_{\frac{3}{2}}p_{\frac{1}{2}})}{((4p_{\frac{3}{2}}+2p_{\frac{1}{2}})/6s_{\frac{1}{2}})},$$ (1) where all entries refer to masses. The experimental data on $`D`$, $`K`$ and $`B`$ mesons give respectively $`r=0.00\pm 0.10,0.06\pm 0.00`$ and $`0.23\pm 0.04`$ (L3) or $`0.23\pm 0.03`$ (OPAL). For the Dirac equation with arbitrary vector and scalar Coulomb potentials, the only cases for which the relevant analytic solutions are known, $`0.7\stackrel{<}{}r\stackrel{<}{}0.6`$. It is hence evident that the spin-orbit splittings extracted from experimental results are indeed small. There is also evidence in light quark mesons and baryonic systems that the spin-orbit interaction is small. In non-relativistic models, meson and “two-body” baryon spin-orbit interactions are related and, for a specific class of baryons, the spin-orbit interaction is small for exactly the same reasons that it is small in mesons. ## III A Dynamical Symmetry for the Dirac Hamiltonian If we consider a system of a (sufficiently) heavy antiquark (quark) and light quark (antiquark), the dynamics may well be represented by the motion of the light quark (antiquark) in a fixed potential provided by the heavy antiquark (quark). Let us assume that both vector and scalar potentials are present. Then the Dirac Hamiltonian describing the motion of the light quark is $$H=\stackrel{}{\alpha }\stackrel{}{p}+\beta (m+V_S)+V_V+M,$$ (2) where we have set $`\mathrm{}=c=1`$, $`\stackrel{}{\alpha }`$, $`\beta `$ are the usual Dirac matrices, $`\stackrel{}{p}`$ is the three-momentum, $`m`$ is the mass of the light quark and $`M`$ is the mass of the heavy quark. This one quark Dirac Hamiltonian follows from the two-body Bethe-Salpeter equation in the equal time approximation, the spectator (Gross) equation with a simple kernel, and a two quark Dirac equation, in the limit that $`M`$ is large. If the vector potential, $`V_V(\stackrel{}{r})`$, is equal to the scalar potential plus a constant potential, $`U`$, which is independent of the spatial location of the light quark relative to the heavy one, i.e., $`V_V(\stackrel{}{r})=V_S(\stackrel{}{r})+U`$, then the Dirac Hamiltonian is invariant under a spin symmetry, $`[H,\widehat{S}_i]=0`$, where the generators of that symmetry are given by, $$\widehat{S}_i=\left(\genfrac{}{}{0pt}{}{\widehat{s}_i}{0}\genfrac{}{}{0pt}{}{0}{\widehat{\stackrel{~}{s}}_i}\right).$$ (3) where $`\widehat{s}_i=\sigma _i/2`$ are the usual spin generators, $`\sigma _i`$ the Pauli matrices, and $`\widehat{\stackrel{~}{s}}_i=U_p\widehat{s}_iU_p`$ with $`U_p=\frac{\stackrel{}{\sigma }\stackrel{}{p}}{p}`$. Thus Dirac eigenstates can be labeled by the orientation of the spin, even though the system may be highly relativistic, and the eigenstates with different spin orientation will be degenerate. For spherically symmetric potentials, $`V_V(\stackrel{}{r})=V_V(r),V_S(\stackrel{}{r})=V_s(r)`$, the Dirac Hamiltonian has an additional invariant algebra; namely, the orbital angular momentum, $$\widehat{L}_i=\left(\genfrac{}{}{0pt}{}{\widehat{\mathrm{}}_i}{0}\genfrac{}{}{0pt}{}{0}{\widehat{\stackrel{~}{\mathrm{}}}_i}\right),$$ (4) where $`\widehat{\stackrel{~}{\mathrm{}}}_i=U_p\widehat{\mathrm{}}_i`$ $`U_p`$ and $`\widehat{\mathrm{}}_i=(\stackrel{}{r}\times \stackrel{}{p})_i`$. This means that the Dirac eigenstates can be labeled with orbital angular momentum as well as spin, and the states with the same orbital angular momentum projection will be degenerate. Thus, for example, the $`n_rp_{1/2}`$ and $`n_rp_{3/2}`$ states will be degenerate, where $`n_r`$ is the radial quantum number. Thus, we have identified a symmetry in the heavy-light quark system which produces spin-orbit degeneracies independent of the details of the potential. If this potential is strong, the heavy-light quark system will be very relativistic; that is, the lower component for the light quark will be comparable in magnitude to the upper component of the light quark. It is remarkable that non-relativistic behaviour of energy levels can arise for such fully relativistic systems. This symmetry is similar to the relativistic symmetry identified as being responsible for pseudospin degeneracies observed in nuclei. In contrast to spin symmetry, pseudospin symmetry has the pairs of states $`((n_r1)s_{1/2},n_rd_{3/2})`$, $`((n_r1)p_{3/2},n_rf_{5/2})`$, etc. degenerate, making the origin of this symmetry less transparent. The pseudospin generators are $$\widehat{\stackrel{~}{S}}_i=\left(\genfrac{}{}{0pt}{}{\widehat{\stackrel{~}{s}}_i}{0}\genfrac{}{}{0pt}{}{0}{\widehat{s}_i}\right).$$ (5) For pseudospin symmetry, the nuclear mean scalar and vector potential must be equal in magnitude and opposite in sign, up to a constant, $`V_V=V_S+U`$. Relativistic mean field representations of the nuclear potential do have this property; that is, $`V_SV_V`$. We will return later to the question of whether the relation $`V_V=V_S+U`$ arises in QCD. It has previously been observed that pseudospin symmetry improves with increasing energy of the states, for various potentials. A similar behaviour may be expected for spin symmetry, consistent with the experimental observations that spin–orbit splittings decrease for higher mass states. The Dirac Hamiltonian (2) encompasses the effects of the OGE and Thomas precession spin-dependent terms customarily included in non-relativistic models. ## IV Experimental Test In the spin symmetry limit, the radial wavefunctions of the upper components of the Dirac wavefunction of the two states in the spin doublet will be identical, behaving “non-relativistically”, whereas the lower components will have different radial wavefunctions. This follows from the form of the spin generators given in Equation (3). The $`(1,1)`$ entry of the operator matrix is simply the non-relativistic spin operator which relates the upper component of the Dirac wavefunction of one state in the doublet to the upper component of the other state in the doublet. Since this operator does not affect the radial wavefunction, the two radial wavefunctions must be the same. By contrast, the lower component wavefunction is operated on by $`U_p`$ which does operate on the radial wavefunction because of the momentum operator. As an example, we show in Figure 1 the upper and lower components for Dirac wavefunctions of the $`p_{1/2}p_{3/2}`$ doublet. The scalar and vector potentials were determined by matching the available spectral data of the $`D`$-mesons, assuming a $`p_{\frac{3}{2}}p_{\frac{1}{2}}`$ splitting at the lower end of the range defined by the experimental value of $`2\pm 50`$ MeV. This maximizes the wavefunction differences. In this realistic case, $`V_VV_S+U`$, so the radial wavefunctions for the upper components are not exactly identical but are very close, whereas the radial wavefunctions for the lower components are very different. Likewise the momentum space wavefunctions for the upper components will be very similar, as seen in Figure 2, again because the spin operator does not affect the wavefunction. However, since $`U_p`$ depends only on the angular part of the momentum, $`\widehat{p}=\frac{\stackrel{}{p}}{p}`$, it does not affect the radial momentum space wavefunction. In Figure 2 we see that the radial momentum space wavefunctions are very similar for the lower components as well. This prediction of the symmetry can be tested in the following experiment. The annihilation $`e^+e^{}D_0^{}D_0^{}`$, $`D_0^{}D_2^{}`$ and $`D_2^{}D_2^{}`$ allows for the extraction of the $`D_0^{}`$ and $`D_2^{}`$ electromagnetic static form factors and the $`D_0^{}`$ to $`D_2^{}`$ electromagnetic transition form factor. The photon interaction ensures that all radial wavefunctions of the light quark are accessed. When spin symmetry is realised, there are only two independent radial momentum space wavefunctions, which should enable the prediction of one of the three form factors in terms of the other two. This should enable the verification of the predictions of spin symmetry. On the other hand, non-relativistic models, with no lower components for the wavefunctions, have only one independent radial wavefunction, which will lead to the prediction of two of the form factors in terms of the remaining one. This might be too restrictive. The proposed experiment can be carried out at the Beijing Electron Positron Collider at an energy of approximately 1 GeV above the $`\psi (4040)`$ peak in the final state $`DD\pi \pi `$. An equivalent experiment for K-mesons would involve detection of the $`KK\pi \pi `$ final state, which has already been measured. The wavefunctions of $`K`$-mesons fitting the experimental spectrum show similar behaviour to the $`D`$-mesons, with the $`p_{\frac{3}{2}}`$ and $`p_{\frac{1}{2}}`$ wavefunctions even more similar than in Figures 1a and 2. If B-mesons do also exhibit spin symmetry, one can do equivalent experiments around 1 GeV above the $`\mathrm{{\rm Y}}`$(3S) peak at the SLAC, KEK or CESR B-factories. ## V QCD Origins If such a dynamical symmetry can explain the suppression of spin-orbit splitting in the hadron spectrum, the question remains as to why it might be expected to appear in QCD. To address this, we first recall the ongoing argument as to whether confinement corresponds to a vector or scalar potential. The first natural expectation was that confinement reflected the infrared growth of the QCD coupling constant, enhancing the color-Coulomb interaction at large distances, see e.g. Ref.(). An involved two- (or multi-) gluon effect has been proposed to account for the origin of a scalar confining potential. The existence of one or the other of these vector and scalar potentials is not necessarily exclusionary – they may both be realised. The arguments in Ref.() suggest further that they are related, with the scalar exceeding the vector by an amount which may be approximately constant as one saturates into the linear confining region at large separations. We very briefly reiterate the basic argument of Ref.() here. The starting point is to accept the standard approach that renormalization-group-improved single-gluon-exchange produces a linearly increasing vector potential between a quark and an antiquark. One then considers what to expect for multiple gluon exchange, starting with two gluons. Since two gluons are attracted to each other in a color singlet channel, and also have a zero mass threshold (as for massless quark-antiquark pairs), it is reasonable to conclude that a (Lorentz and color) scalar gluonic condensate develops, along with a mass gap for a glueball state. These developments are indeed observed in lattice QCD calculations. Ref.() goes on to argue that renormalization-group-improved single-glueball-exchange involves the square of the QCD coupling and so, despite the massiveness of the object exchanged, also leads to a (now scalar) confining potential between quarks and antiquarks. This further implies that the ratio of the slopes of the two potentials in their common linear (confining) region is given by the square of the ratio of the QCD scale for growth of the coupling constant to the value of the mass gap of the condensate formation. This ratio may be expected to be of order one as both quantities are determined by the underlying QCD scale. If the two potentials do indeed have similar slopes in the region outside that dominated by the color Coulomb interaction, they would necessarily differ only by an approximately constant value, in that region. Thus, the origin of the dynamical symmetry may not be unreasonable, and may indeed be a natural outcome of non-perturbative QCD. On the other hand, identically equal vector and scalar potentials, except for a constant difference, would appear to be coincidental. An ameliorating effect is that to produce an approximation to the spin symmetry of Eq. (2) this condition need only hold in regions where the wavefunctions are substantial. The determination of QCD potentials, from models like the minimal area law, stochastic vacuum model, or dual QCD, and from lattice QCD, is hampered by the problem of rigorously defining the concept of a potential from QCD when one quark is light. It suffices to say that there is no agreement on the mixed Lorentz character of the potential even between two heavy quarks, where the potential can be rigorously defined, although lattice QCD results are consistent with simply a vector Coulomb and scalar linear potential. ## VI Summary The observation of “accidental” spin-orbit degeneracies observed in heavy-light quark mesons can be explained by a relativistic symmetry of the Dirac Hamiltonian which occurs when the vector and scalar potentials exerted on the light quark by the heavy antiquark differ approximately by a constant, $`V_VV_S+U`$. Conversely, if future experiments determine that spin-orbit splittings are small not only for the lowest excited states in mesons but are small throughout the meson spectrum, this experimental fact dictates that the effective QCD vector and scalar potentials between a quark and antiquark are approximately equal up to a constant, which would be a significant observation about the nature of non-perturbative QCD. Furthermore, the approximate symmetry predicts that the spatial Dirac wavefunction for the spin doublets will be approximately equal in momentum space, a feature which can be tested in electron-positron annihilation. We have argued that $`V_VV_S+U`$ may occur in QCD, particularly for regions of space dominated by the light quark wavefunction. Work is in progress to extend this symmetry to purely light quark systems. This research is supported by the Department of Energy under contract W-7405-ENG-36. FIGURES Figure 1: (a) The square of the Dirac radial wavefunction of the upper component times $`r^2`$. (b) The square of the Dirac radial wavefunction of the lower component times $`r^2`$. $`p_{\frac{3}{2}}`$ is the solid line and $`p_{\frac{1}{2}}`$ is the dashed line. Note that the lower component is comparable to the upper component. The wavefunctions are solutions of the Dirac equation (see Eq. (2)) with Coulomb potentials $`V_S(r)=\frac{\alpha _S}{r}+U_S`$ and $`V_V(r)=\frac{\alpha _V}{r}+U_V`$, where $`\alpha _S=1.279,U_S=506`$ MeV, $`\alpha _V=0.779,U_V=515`$ MeV, $`m=330`$ MeV and $`M=1480`$ MeV. This corresponds to a $`p_{\frac{3}{2}}p_{\frac{1}{2}}`$ splitting of -52 MeV. Figure 2: (a) The square of the Dirac momentum space wavefunction of the upper component times $`q^2`$. (b) The square of the Dirac momentum space wavefunction of the lower component times $`q^2`$. Other conventions are the same as in Figure 1.
warning/0002/gr-qc0002037.html
ar5iv
text
# Untitled Document On Schwarzschild-Like Solutions in Curvature-Quadratic Gravity Paul Federbush Department of Mathematics University of Michigan Ann Arbor, MI 48109-1109 (pfed@math.lsa.umich.edu) Abstract Partial results are obtained for Schwarzschild-like solutions in a gravity theory with action density $`c(g)^{1/2}[R_{ik}^2+bR^2]`$. A seven parameter family of implicit solutions is found. A number of explicit solutions are also exhibited. In a previous paper, , we have proposed the curvature-quadratic action $$cd^4x(g)^{1/2}\left[R^{ik}R_{ik}+bR^2\right]$$ (1) as the basis for quantum gravity. If $`b=\frac{1}{3}`$ then this action is of the conformal Weyl theory. Classical solutions of the Euler-Lagrange equations for this conformal theory have been studied by Mannheim and Kazanas, in particular for Schwarzsc hild-like solutions, . In this paper we study the problem of such solutions for the action (1) and arbitrary value of $`b`$. (In there are some results on the linearized equations.) We thus seek static spherically symmetric metrics. By a coordinate change these may be put in form $$ds^2=d^2(r)\left[c(r)dt^2+\frac{1}{c(r)}(dr)^2+r^2d\mathrm{\Omega }\right]$$ (2) (as shown in ). For the Einstein action $$cd^4x(g)^{1/2}R$$ there is the solution of the form (2) as follows: $$d(r)=1,c(r)=1+\frac{\alpha }{r}.$$ (3) In the conformal Weyl theory, the solution metrics are of the form: $$c(r)=1\beta (23\beta \gamma )\frac{1}{r}3\beta \gamma +\gamma rkr^2$$ (4) (See .) In this case, of course, $`d(r)`$ is arbitrary. For the action (1), general $`b`$, our results are partial. There are several explicit solutions: explicit solution 1: $$c(r)=1+\frac{\gamma }{r},d(r)=1+(\frac{\alpha }{\gamma })\mathrm{}n(1+\frac{\gamma }{r})+\beta $$ (5) explicit solution 2: $$c(r)=1,d(r)=1+\frac{\alpha }{r}+\beta $$ $`(5^{})`$ This is a limit of the solution in equation (5). explicit solution 3: $$c(r)=1+\frac{\alpha }{r}+\beta r^2,d(r)=1+\gamma .$$ (6) This is the well-known de Sitter solution. explicit solution 4: $$c(r)=1+\alpha ^2r^2,d(r)=\frac{\beta }{(\alpha r)}\mathrm{arctan}(\alpha r)$$ (7) explicit solution 5: $`c(r)`$ $`=`$ $`12\alpha r+4\alpha ^2r^2`$ $`d(r)`$ $`=`$ $`\beta \left({\displaystyle \frac{1}{2\alpha r}}\right){\displaystyle \frac{2}{3}}\sqrt{3}\mathrm{arctan}\left({\displaystyle \frac{\sqrt{3}}{3}}(4\alpha r1)\right)`$ (8) Beyond these, there is the formal solution. $$(c(r),d(r))=(1,1)+\underset{n=1}{\overset{\mathrm{}}{}}t^n\stackrel{}{u}_n$$ (9) $$\stackrel{}{u}_1=\underset{j=1}{\overset{7}{}}a_j\stackrel{}{\varphi }_j$$ $`(9^{})`$ $$\stackrel{}{\varphi }_1=(0,1)$$ $`(10a)`$ $$\stackrel{}{\varphi }_2=(r^2,0)$$ $`(10b)`$ $$\stackrel{}{\varphi }_3=(0,r^2)$$ $`(10c)`$ $$\stackrel{}{\varphi }_4=(\frac{1}{r},0)$$ $`(10d)`$ $$\stackrel{}{\varphi }_5=(0,\frac{1}{r})$$ $`(10e)`$ $$\stackrel{}{\varphi }_6=(2r,r)$$ $`(10f)`$ $$\stackrel{}{\varphi }_7=(4rlnr,2rlnr\frac{1}{2}\frac{3+8b}{3b+1}r)$$ $`(10g)`$ The $`\stackrel{}{u}_i`$ may be found inductively, each is a finite sum of terms, polynomial in the $`a_i,r,1/r,lnr`$. The sum in (9) is essentially a perturbation expansion. The convergence may be described as follows: for each finite interval, $`\alpha <r<\beta `$, and choice of the $`a_i`$, there is an $`\epsilon `$ such that the series in (9) converges if $`0|t|<\epsilon `$, and represents a solution of the Euler-Lagrange equations. The Euler-Lagrange equations were derived as follows. A metric of the form $$ds^2=\left(d^2(r)c(r)+B(r)\right)dt^2+\left(d^2(r)+A(r)\right)r^2d\mathrm{\Omega }+\left(\frac{d^2(r)}{c(r)}+A(r)+r^2C(r)\right)dr^2$$ (11) was substituted into the action (1); and three Euler-Lagrange equations $`X`$ $`=`$ $`0`$ $`Y`$ $`=`$ $`0`$ (12) $`Z`$ $`=`$ $`0`$ were obtained by requiring the action to be invariant under variations, with respect to $`A(r),B(r)`$ and $`C(r)`$ respectively (evaluated at $`A=B=C=0`$). This was done using Maple software. Each of the equations in (12) has approximately 500 terms, the al gebra could not be done by hand. There are thus three equations to be satisfied by the two functions $`c(r)`$ and $`d(r)`$. It was then checked (in Maple) that (5), (6), (7), and (8) satisfied the equations (12). (The exact form in which the variations $`A,B`$, and $`C`$ were put in (11) was a largely arbitrary choice.) Under a coordinate change $$rr+\epsilon \varphi (r)$$ (13) the action (1) remains invariant, but the $`A,B`$, and $`C`$ change as follows: $`\delta A`$ $`=`$ $`\epsilon \left[\varphi {\displaystyle \frac{(d^2(r))}{r}}+2d^2(r){\displaystyle \frac{1}{r}}\varphi \right]`$ (14) $`\delta B`$ $`=`$ $`\epsilon \varphi {\displaystyle \frac{}{r}}\left(d^2(r)c(r)\right)`$ (15) $`\delta C`$ $`=`$ $`\epsilon [\varphi {\displaystyle \frac{1}{r^2}}{\displaystyle \frac{}{r}}\left({\displaystyle \frac{d^2(r)}{c(r)}}\right)+2{\displaystyle \frac{1}{r^2}}{\displaystyle \frac{\varphi }{r}}{\displaystyle \frac{d^2(r)}{c(r)}}`$ (16) $``$ $`{\displaystyle \frac{1}{r^2}}(\varphi {\displaystyle \frac{}{r}}(d^2(r))+2d^2(r){\displaystyle \frac{1}{r}}\varphi )].`$ This leads to the relation between $`X,Y,Z`$ of (12) as follows: $`({\displaystyle \frac{d}{dr}}(d^2(r))`$ $`+`$ $`2d^2(r){\displaystyle \frac{1}{r}})X+{\displaystyle \frac{d}{dr}}(d^2(r)c(r))Y`$ $`+({\displaystyle \frac{1}{r^2}}{\displaystyle \frac{d}{dr}}\left({\displaystyle \frac{d^2(r)}{c(r)}}\right)`$ $``$ $`{\displaystyle \frac{1}{r^2}}{\displaystyle \frac{d}{dr}}\left(d^2(r)\right){\displaystyle \frac{2}{r^3}}d^2(r))Z`$ $``$ $`{\displaystyle \frac{d}{dr}}\left(2{\displaystyle \frac{d^2(r)}{c(r)}}{\displaystyle \frac{1}{r^2}}Z\right)=0.`$ We can see from (17) that if the last two equations of (12) hold, the first automatically is satisfied. The terms in (10) arise as a complete set of seven linearly independent solutions of the linear terms in the equations $`Y=0,Z=0`$, the first of these is a fourth order differential equation, the second a third order. The linear portions of $`Y=0`$ and $`Z=0`$ are easily verified to be a non-singular linear system (that of $`X=0`$ and $`Z=0`$ would not be). This ensures that during the iterative con struction of the $`\stackrel{}{u}_i`$ no compatibility requirements are met. i.e. once $`\stackrel{}{u}_1,\mathrm{},\stackrel{}{u}_N`$ have been found, the construction of $`\stackrel{}{u}_{N+1}`$ does not impose restrictions on the $`\stackrel{}{u}_1,\mathrm{},\stackrel{}{u}_N`$ already developed. It is of course of interest to see if other explicit solutions may be found. Acknowledgment: I would like to thank Professor Hans-Juergen Schmidt for helpful comments. References * \] P. Federbush, “A Speculative Approach to Quantum Gravity”, Symposium in Honor of Eyvind H. Wichmann, University of California, Berkeley, June 1999. * \] P.D. Mannheim and D. Kazanas, “Exact Vacuum Solution to Conformal Weyl Gravity and Galactic Rotation Curves”, Astrophys. J. 342 635 (89). * \] M. Roberts, “Galactic Rotation Curves and Quadratic Lagrangians”, Monthly Notices of the Royal Astronomical Society, Vol. 249, p. 339 (91).
warning/0002/astro-ph0002412.html
ar5iv
text
# Expected Number and Flux Distribution of Gamma-Ray-Burst Afterglows with High Redshifts ## 1 Introduction The past decade has been marked by major observational breakthroughs concerning the properties of the Gamma Ray Burst (GRB) sources. The Burst and Transient Source Experiment (BATSE) on board the Compton Gamma Ray Observatory (Meegan et al. 1992) showed that the GRB population is distributed isotropically across the sky, and that there is a deficiency of faint GRBs relative to a Euclidean distribution. These were the first observational clues indicating a cosmological distance scale for GRBs. The localization of GRBs by X-ray observations with the BeppoSAX satellite (Costa et al. 1997) allowed the detection of afterglow emission at optical (e.g., van Paradijs et al. 1997) and radio (e.g., Frail et al. 1997; Frail, Waxman & Kulkarni 1999) wavelengths up to more than a year following the events (Fruchter et al. 1999; Frail, Waxman & Kulkarni 1999). The afterglow emission is characterized by a broken power-law spectrum with a peak frequency that declines with time. The radiation can be well modeled as synchrotron emission from a decelerating blast wave, created by the GRB explosion in an ambient medium, plausibly the interstellar medium of galaxies (Waxman 1997; Wijers & Galama 1999; Mészáros 1999; but see also Chevalier & Li 1999). The detection of spectral features, such as metal absorption lines (Metzger et al. 1997), in some optical afterglows allowed an unambiguous identification of the cosmological distance scale. By now, the redshift of almost a dozen GRBs has been identified either through the detection of absorption features in the afterglow spectra or of emission lines from host galaxies (see Fig. 3 in Kulkarni et al. 2000). The central engine of GRBs is still unknown. Since the inferred energy release is comparable to or higher than that in supernovae, most popular models relate GRBs to stellar remnants, such as neutron stars or black holes (see, e.g., Eichler et al. 1989; Paczyński 1991; Usov 1992; Mochkovitch et al. 1993; Paczyński 1998; MacFadyen & Woosley 1999). Recently it has been claimed that the late evolution of some rapidly declining optical afterglows shows a component which is possibly associated with supernova emission (e.g., Bloom et al. 1999; Reichart 1999). If the supernova association will be confirmed by detailed spectra of future afterglows, the GRB phenomenon will be linked to the terminal evolution of massive stars. Any association of GRBs with the formation of single compact stars implies that the GRB rate should trace the star formation history of the universe (Totani 1997; Sahu et al. 1997; Wijers et al. 1998; but see Krumholz, Thorsett & Harrison 1998). Owing to their high brightness, GRB afterglows might be detected out to exceedingly high redshifts. Similarly to quasars, the broad band emission of GRB afterglows can be used to probe the absorption properties of the intergalactic medium (IGM) out to the epoch when it was reionized at a redshift $`z10`$ (Miralda-Escudé & Rees 1998; Loeb 1999). Lamb & Reichart (1999) have extrapolated the observed gamma-ray and afterglow spectra of known GRBs to high redshifts and emphasized the important role that their detection might play in probing the IGM. Simple scaling of the long-wavelength spectra and temporal evolution of afterglows with redshift implies that at a fixed time lag after the GRB in the observer’s frame, there is only mild change in the observed flux at infrared or radio wavelengths with increasing redshift. This results in part from the fact that afterglows are brighter at earlier times, and that a given observed time refers to an earlier intrinsic time in the source frame as the source redshift increases. The mild dependence of the long-wavelength flux on redshift is in contrast with other high-redshift sources such as galaxies or quasars, which fade rapidly with increasing redshift (Haiman & Loeb 1998; 1999). The “apparent brightening” of GRB afterglows with redshift could be further enhanced by the expected increase in the mean density of the interstellar medium of galaxies at increasing redshifts (Wood & Loeb 1999). It therefore appears natural to use GRBs as an important tool in probing the high-redshift universe and its star formation history (Blain & Natarajan 1999). Since GRBs are rare, all-sky searches for their early $`\gamma `$-ray emission are needed before follow-up observations at much longer wavelength are conducted. In this paper we model the emission from GRB afterglows and follow their number counts at different wavelengths as a function of redshift. For simplicity, we assume that the ambient gas surrounding GRB sources is the interstellar medium of their host galaxies. In §2 we review our model for both the relativistic and non-relativistic stages of the afterglow emission, as well as a simple prescription for the redshift evolution of the interstellar medium of galaxies. The numerical results and their dependence on model assumptions are discussed in §3. Finally, §4 summarizes our main conclusions and describes the implications of our results. ## 2 Afterglow Emission ### 2.1 Relativistic Regime GRB afterglows can be modeled as synchrotron emission from a decelerating relativistic blast wave, created by the GRB explosion in an external medium. For a point explosion in a uniform medium, the shock structure in the highly-relativistic regime is described by the Blandford & McKee (1976) self-similar solution. The synchrotron emission originates from a power-law population of shock-accelerated electrons in a strong magnetic field (Waxman 1997; Mészáros, Rees, & Wijers 1998; Sari, Piran & Narayan 1998), with both the electrons and the magnetic field having close to equipartition energy densities (see Medvedev & Loeb 1999 for the possible origin of the magnetic field). For simplicity, we focus our discussion on a spherically-symmetric explosion in a uniform ambient medium, assumed to be the interstellar medium of the host galaxy, although at least some GRB blast waves might be expanding into the stellar wind of the progenitor star (Dai & Lu 1998; Mészáros, Rees & Wijers 1998; Chevalier & Li 1999). The dependence of our results on the ambient medium density will be discussed in § 3.3. We assume that the shock-accelerated electrons have a power-law distribution of Lorentz factor, $`\gamma _e`$, with a minimum Lorentz factor $`\gamma _m`$ (see, e.g., Sari, Piran & Narayan 1998). We also define $`\gamma _c`$ as the threshold Lorentz factor below which electrons do not lose a significant fraction of their energy to radiation. In the fast cooling regime, when $`\gamma _m>\gamma _c`$, all of the electrons cool rapidly down to a Lorentz factor $`\gamma _c`$ and the flux observed at the frequency $`\nu `$ is given by, $$F_\nu =F_{\nu _m}\{\begin{array}{cc}(\nu /\nu _c)^{1/3}\hfill & \nu _c>\nu ,\hfill \\ (\nu /\nu _c)^{1/2}\hfill & \nu _m>\nu \nu _c,\hfill \\ (\nu _m/\nu _c)^{1/2}(\nu /\nu _m)^{p/2}\hfill & \nu \nu _m,\hfill \end{array}$$ (1) where $`F_{\nu _m}`$ is the observed peak flux, $`\nu _c=\nu (\gamma _c)`$ and $`\nu _m=\nu (\gamma _m)`$ are the characteristic synchrotron frequency calculated at $`\gamma _c`$ and $`\gamma _m`$ respectively and $`p`$ is the power-law index of the electron energy distribution. A typical value of $`p2.5`$ often fits both the GRB and the afterglow observations (Sari, Piran & Narayan 1998; Kumar & Piran 1999). When $`\gamma _c>\gamma _m`$, only electrons with $`\gamma _e>\gamma _c`$ cool efficiently. In this slow cooling regime the observed flux is, $$F_\nu =F_{\nu _m}\{\begin{array}{cc}(\nu /\nu _m)^{1/3}\hfill & \nu _m>\nu ,\hfill \\ (\nu /\nu _m)^{(p1)/2}\hfill & \nu _c>\nu \nu _m,\hfill \\ (\nu _c/\nu _m)^{(p1)/2}(\nu /\nu _c)^{p/2}\hfill & \nu \nu _c.\hfill \end{array}$$ (2) Typically, synchrotron self-absorption results in an additional break, at a frequency $`\stackrel{<}{}5`$GHz. We will not consider the low-frequency regime below this break in our discussion. Assuming a fully adiabatic shock (Sari, Piran & Narayan 1998), $$\nu _c=2.7\times 10^{12}ϵ_B^{3/2}E_{52}^{1/2}n_1^1t_d^{1/2}(1+z)^{1/2}\mathrm{Hz},$$ (3) $$\nu _m=5.7\times 10^{14}ϵ_B^{1/2}ϵ_e^2E_{52}^{1/2}t_d^{3/2}(1+z)^{1/2}\mathrm{Hz},$$ (4) $$F_{\nu _m}=1.1\times 10^5ϵ_B^{1/2}E_{52}n_1^{1/2}d_{28}^2(1+z)\mu \mathrm{Jy}.$$ (5) The redshift dependence results from the fact that the radiation emitted by a source at a redshift $`z`$ at the frequency $`\nu _s`$ over a time $`\mathrm{\Delta }t_s`$, will be observed at $`z=0`$ at a frequency $`\nu _o=\nu _s/(1+z)`$ over a time $`\mathrm{\Delta }t_o=(1+z)\mathrm{\Delta }t_s`$. Here $`ϵ_B`$ and $`ϵ_e`$ are the fraction of the shock energy that is converted to magnetic fields and accelerated electrons, respectively, for which we adopt the values $`ϵ_B=0.1`$ and $`ϵ_e=0.2`$ (see, e.g., Waxman 1997); $`E=E_{52}10^{52}`$ erg is the energy of the spherical shock; $`n=n_1`$ 1 cm<sup>-3</sup> is the mean number density of the ambient gas; $`t=t_d`$ 1 day is the time lag since the GRB trigger, as measured in the observer frame, and $`d_L=d_{28}10^{28}`$ cm is the cosmology-dependent luminosity distance. For a flat universe ($`\mathrm{\Omega }_0=\mathrm{\Omega }_m+\mathrm{\Omega }_\mathrm{\Lambda }=1`$), the luminosity distance can be written as: $$d_L=(1+z)_0^z(1+z^{})\left|\frac{dt}{dz^{}}\right|𝑑z^{},$$ (6) where $$\left(\frac{dt}{dz}\right)^1=H_0(1+z)\sqrt{(1+\mathrm{\Omega }_mz)(1+z)^2\mathrm{\Omega }_\mathrm{\Lambda }(2z+z^2)}.$$ (7) $`H_0=100h`$ Km s<sup>-1</sup> Mpc<sup>-1</sup> is the current Hubble constant. Throughout the paper we adopt a flat cosmology with $`h=0.65`$, density parameters $`\mathrm{\Omega }_m=0.35`$ and $`\mathrm{\Omega }_\mathrm{\Lambda }=0.65`$, and a Cold Dark Matter power spectrum of density fluctuations with power-law index $`n=0.96`$, an amplitude $`\sigma _8=0.87`$ and baryon density parameter $`\mathrm{\Omega }_b=0.04`$ (Bahcall et al. 1999). The explosion energy, $`E`$, is highly uncertain. Based on X-ray afterglow data, Freedman & Waxman (1999) inferred explosion energies in the range $`10^{51.5}10^{53.5}\mathrm{ergs}(\mathrm{\Delta }\mathrm{\Omega }/4\pi )`$, where $`\mathrm{\Delta }\mathrm{\Omega }`$ is the solid angle into which the energy is channeled. If one assumes that GRB explosions are isotropic, then in some events the radiated energy is estimated to be in excess of $`10^{53}`$ ergs, reaching a value of $`3.4\times 10^{54}`$ erg for GRB 990123 (see, e.g., Kulkarni et al. 1999). It has been suggested that these values are reduced by orders of magnitude due to beaming (Kulkarni et al. 1999; Mészáros & Rees 1999), although the relatively low efficiency for converting internal shock energy into radiation might imply an energy budget as high as $`10^{54}`$ ergs even after the beaming correction (Kumar 1999). Unless otherwise noted, we adopt the value $`E=10^{52}`$ ergs, which is intermediate for the range inferred by Freedman & Waxman (1999). In § 3.3 we will discuss the effect of beaming and different choices of the explosion energy on our results. ### 2.2 Non-Relativistic Regime As the blast wave decelerates, the fireball eventually makes a transition from relativistic to sub-relativistic ($`sr`$) expansion (Huang, Dai & Lu 1998; Woods & Loeb 1999; Frail, Waxman & Kulkarni 1999; Wei & Lu 1999). In this regime, the evolution of the GRB remnants is well described by the Sedov-Taylor solution (Taylor 1950; Sedov 1959). The time when this transition takes place in the observer frame, $`t_{sr}`$, can be approximately derived by setting the shock velocity in the Sedov-Taylor solution to be equal to the speed of light, $`c`$. This gives: $$t_{sr}=\left(1+z\right)\left[\frac{E}{m_pn}\left(\frac{0.47}{c}\right)^5\right]^{1/3}=1.8\times 10^2(1+z)E_{52}^{1/3}n_1^{1/3}\mathrm{days}.$$ (8) In the sub-relativistic regime, the shock radius and velocity evolve as, $$r(t)=r(t_{sr})(t/t_{sr})^{2/5},$$ (9) $$v(t)=c(t/t_{sr})^{3/5},$$ (10) where all times are in the observer frame and $`r(t_{sr})`$ is the shock radius at time $`t_{sr}`$. For a strong non-relativistic shock, the post-shock particle density, energy density, and magnetic field are, $$n^{}=4n,$$ (11) $$u^{}=\frac{9}{8}nm_pv^2,$$ (12) $$B^{}=(8\pi ϵ_Bu^{})^{1/2}.$$ (13) The observed flux requires a substitution of the values of $`\nu _c`$, $`\nu _m`$ and $`F_{\nu _m}`$ in the sub-relativistic regime into equations (1) and (2). The synchrotron formulae can be used since the emitting electrons are still ultra-relativistic for the relevant remnant ages in this regime. The characteristic synchrotron frequency is given by, $$\nu (\gamma )=\gamma ^2\frac{eB^{}}{2\pi m_ec}.$$ (14) Recalling that $`\gamma _mB^2u^{}/n^{}`$, we find $`\gamma _mv^2t^{6/5}`$, and $`\gamma _c(B^2t)^1t^{1/5}`$. Equation (14) then provides an expression for $`\nu _c`$ and $`\nu _m`$ in the sub-relativistic regime. In particular $`\nu _ct^{1/5}`$ and $`\nu _mt^3`$. During the relativistic expansion the observed peak flux is constant in time; however, in the sub-relativistic regime it varies as $`F_{\nu _m}r^3B^{}t^{3/5}`$. In the UV regime one needs to take account of absorption by the intergalactic medium (IGM). At redshifts greater than the reionization redshift, $`z_{reion}`$, the neutral IGM is optically thick to photon energies above the Ly$`\alpha `$ resonance. Recent models for the reionization of the IGM predict $`z_{reion}`$ in the range of 7-12 (Haiman & Loeb 1997; Miralda-Escudé, Haehnelt & Rees 1998; Valageas & Silk 1999; Chiu & Ostriker 1999; Gnedin 1999; Ciardi et al. 2000). Here we adopt $`z_{reion}=8`$; the precise choice of which has only a minor effect on our results since the Ly$`\alpha `$ forest yields a high opacity at the redshifts of interest even if most of the IGM is ionized. For $`z\stackrel{<}{}5`$, we use the updated Ly$`\alpha `$ forest opacity derived by Haardt & Madau (1996) and Madau, Haardt & Rees (1999), which is based on the observed absorber distribution in the spectra of high-redshift quasars. The continuum depression blue-ward of the Ly$`\alpha `$ resonance is already close to unity at $`z5`$ (see Figure 13 in Stern & Spinrad 1999), and so the treatment of the Ly$`\alpha `$ absorption at $`5\stackrel{<}{}z\stackrel{<}{}8`$ has little effect on our results. Next, we derive an expression for the ambient gas density as a function of redshift. For simplicity, we assume that the ambient gas has the mean density of the interstellar medium of the host galaxy, as inferred for GRB970508 and GRB971214 (Waxman 1997; Wijers & Galama 1999). ### 2.3 Evolution of the Gas Density in Galactic Disks The phenomenological modeling of many afterglow lightcurves implies values of the ambient medium density $`1\mathrm{cm}^3`$, which are of order the mean density of the interstellar medium of disk galaxies (Waxman 1997; Wijers & Galama 1999; Mészáros 1999). In popular hierarchical models of galaxy formation (Kauffmann, White & Guiderdoni 1993; Mo, Mao & White 1998), the mass and size of galactic disks evolves with redshift (Barkana & Loeb 2000). Hence, in modeling the statistical properties of GRBs within disk galaxies at different redshifts, we need to follow the evolution of the average density of the interstellar gas within these galaxies as a function of cosmic time. We model a disk galaxy as a gaseous disk embedded within a dark matter halo. The disk is assumed to be self-gravitating and radially exponential with a scale radius, $`r_d`$. Its scale height at any radius is dictated by the balance between its self-gravity and the gas pressure. For simplicity, we assume the disk to be isothermal. Its mass density profile is then (Spitzer 1942), $$\rho (\zeta ,r)=\rho _0\mathrm{e}^{r/r_d}\mathrm{sech}^2(\xi /\sqrt{2}),$$ (15) where $`\zeta `$ and $`r`$ are the vertical and radial coordinates of the disk, respectively; $`\rho _0=`$const, $`\xi =\zeta (4\pi G\rho _0\mathrm{e}^{r/r_d}/c_s^2)^{1/2}`$, and $`c_s`$ is the effective sound speed (including the possible contribution from turbulence). We adopt a sound speed of $`c_s10`$ km s<sup>-1</sup>, corresponding to a temperature of $`10^4\mathrm{K}`$ below which atomic cooling is highly inefficient. At redshifts $`z\stackrel{<}{}15`$ the formation of galaxies with virial temperatures below $`10^4\mathrm{K}`$ is expected to be suppressed due to various feedback effects (Barkana & Loeb 1999; Ciardi et al. 2000). The scale-radius of the disk is dictated by the angular momentum per unit mass of the gas and can be expressed as (Mo, Mao & White 1998), $$r_d=\frac{1}{\sqrt{2}}\frac{j_d}{m_d}\lambda r_{vir}.$$ (16) Here $`j_d=J_d/J`$ and $`m_d=M_d/M`$, where $`J`$ and $`M`$ are the halo angular momentum and mass and the subscript $`d`$ refers to the disk; $`\lambda `$ and $`r_{vir}`$ are the halo spin parameter and virial radius respectively. We assume that the disk mass fraction is the cosmic baryonic fraction, $`M_d/M=(\mathrm{\Omega }_b/\mathrm{\Omega }_0)`$, since gas cooling is efficient at the high redshifts of interest. The characteristic size distribution of local galactic disks is obtained by adopting $`j_d/m_d=1`$ (Fall & Efstathiou 1980). Based on numerical simulations, we parametrize the distribution of the spin parameter $`\lambda `$ by a log-normal form (Mo, Mao & White 1998, and references therein), $$p(\lambda )d\lambda =\frac{1}{\sigma _\lambda \sqrt{2\pi }}\mathrm{exp}\left[\frac{\mathrm{ln}^2(\lambda /\overline{\lambda })}{2\sigma _\lambda ^2}\right]\frac{d\lambda }{\lambda },$$ (17) with $`\overline{\lambda }=0.05`$ and $`\sigma _\lambda =0.5`$. The halo virial radius is given by (Barkana & Loeb 1999), $$r_{vir}=7.56\left[h^2\frac{M}{10^8M_{}}\frac{\mathrm{\Omega }(z)}{\mathrm{\Omega }_m}\frac{200}{\delta _c}\right]^{1/3}(1+z)^1\mathrm{kpc},$$ (18) with $`\delta _c=18\pi ^2`$ and, $$\mathrm{\Omega }(z)=\frac{\mathrm{\Omega }_m(1+z)^3}{\mathrm{\Omega }_m(1+z)^3+\mathrm{\Omega }_\mathrm{\Lambda }+(1\mathrm{\Omega }_m\mathrm{\Omega }_\mathrm{\Lambda })(1+z)^2}.$$ (19) The total mass of the disk, $`M_d`$, can be derived by integrating equation (15). This yields the relation, $$\rho _0=\frac{GM_d^2}{128\pi c_s^2r_d^4}.$$ (20) The star formation rate depends on the local surface density of the gaseous disk. Schmidt’s law implies that the star formation rate is proportional to the local surface density of the disk to the power of 1-2 (Kennicutt 1998). Hence we first model the probability for a GRB occurrence within a volume element $`2\pi rdrd\zeta `$ as being proportional to the density squared, $`P_{GRB}=𝒜n^2(\zeta ,r)`$; while in §3.3 we consider the sensitivity of our results to other power-laws scalings. The proportionality constant $`𝒜`$ is chosen to normalize the integral of $`P_{GRB}`$ to unity. Consequently, we can write the average of the ambient density probed by GRBs as $$\rho =_0^{\mathrm{}}_0^{\mathrm{}}\rho (\zeta ,r)P_{GRB}2\pi r𝑑r𝑑\zeta =0.29\rho _0.$$ (21) The number density of atoms in the medium is thus $`n=\rho /\mu m_p`$, where $`\mu `$ is the mean molecular weight. Similarly, we can derive $`\sqrt{n}`$ and $`1/n`$ and substitute them into equations (3)-(5). Even if the central engine of all GRBs were standard, the dependence of the ambient medium density on the properties (mass and size) of the host galaxy would have introduced significant scatter and evolution to the flux distribution of GRB afterglows. ## 3 Results ### 3.1 Flux Evolution Figure 1 shows the observed flux for a GRB hosted by the average halo mass as a function of redshift. This average is calculated based on the Press & Schechter (1974) formalism, which provides the mass function of dark matter halos as a function of redshift. The two sets of curves correspond to two observed wavelengths: in Figure 1a the thick lines correspond to $`\lambda _{\mathrm{obs}}=2\mu `$m and the thin lines to $`5000`$Å, while in Figure 1b the thick lines correspond to $`\lambda _{\mathrm{obs}}=10`$ cm and the thin lines to 1 mm. Within each set, the lines correspond to an observed time of 1 hr (solid), 1 day (dotted) and 10 days (dashed). While the optical and infrared emission peak at observed times shorter than an hour, the radio emission has not reached its maximum even after 50 days. At observed frequencies below the Ly$`\alpha `$ resonance frequency, $`\nu _\alpha (z)=2.47\times 10^{15}/(1+z)`$ Hz, the emitted flux is not absorbed by the intervening intergalactic medium at a redshift $`z`$. Figure 1 shows that at these frequencies the afterglow flux is only weakly dependent on redshift. This can be crudely understood from the scaling laws implied by equations (1) and (2), by considering the simple Einstein–de Sitter cosmology ($`\mathrm{\Omega }_m=1`$), for which $`d_L(1+z)[1(1+z)^{1/2}]`$. Recalling that $`n(1+z)^4`$ for a fixed host galaxy mass \[see eqs. (18)- (21)\] and substituting these scaling laws into equations (1) and (2), we find for the case of fast cooling, $$F_\nu \{\begin{array}{cc}(1+z)^{5/2}[1(1+z)^{1/2}]^2\hfill & \nu _c>\nu ,\hfill \\ (1+z)^{5/4}[1(1+z)^{1/2}]^2\hfill & \nu _m>\nu \nu _c,\hfill \\ (1+z)^{(p6)/4}[1(1+z)^{1/2}]^2\hfill & \nu \nu _m;\hfill \end{array}$$ (22) while for slow cooling, $$F_\nu \{\begin{array}{cc}(1+z)^{5/6}[1(1+z)^{1/2}]^2\hfill & \nu _m>\nu ,\hfill \\ (1+z)^{(p3)/4}[1(1+z)^{1/2}]^2\hfill & \nu _c>\nu \nu _m,\hfill \\ (1+z)^{(p6)/4}[1(1+z)^{1/2}]^2\hfill & \nu \nu _c.\hfill \end{array}$$ (23) Although these scaling laws ignore the evolution of the characteristic galaxy mass with redshift, they demonstrate that the observed flux does not decline rapidly with increasing redshift, but might rather increase with redshift at low frequencies (e.g. in the millimeter or radio regimes). This behavior results from three causes: (i) for a fixed observed time-lag after the GRB, the cosmic time-dilation implies that the higher the source redshift is, the earlier the emission time is, and the brighter the intrinsic afterglow luminosity is; (ii) the spectral slope of GRB afterglows results in a favorable K-correction for the redshifts and wavelengths of interest; and (iii) the gas density in galaxies increases with redshift (due to the corresponding increase in the mean density of the universe). ### 3.2 Number Count In a snapshot mode of observations, the total number of GRBs with observed flux greater than $`F`$ at an observed frequency $`\nu `$ is $$N(>F,\nu )=_0^{\mathrm{}}_{M_{min}}^{\mathrm{}}r_{GRB}(z,M)\frac{t_F(z,M,\nu )}{(1+z)}\frac{dn}{dM}(z)\frac{dV}{dz}dMdz.$$ (24) where $`r_{GRB}`$ denotes the GRB rate in a galaxy halo of mass $`M`$ at a redshift $`z`$, $`t_F`$ is the observed time over which a GRB event from this halo yields a flux brighter than $`F`$ at a frequency $`\nu `$, $`dn/dM`$ is the comoving number density of halos with masses between $`M`$ and $`M+dM`$ (based on the Press-Schechter formalism), and $`M_{min}`$ is the minimum galaxy mass in which stars form at a redshift $`z`$. The minimum halo mass inside which star formation occurs, is related to a minimum virial temperature of $`10^4\mathrm{K}`$ below which atomic cooling of the gas is suppressed and fragmentation into stars is inhibited (Haiman, Rees & Loeb 1997; Ciardi et al. 2000). This minimum virial temperature implies $`M_{\mathrm{min}}=4.4\times 10^9M_{}(1+z)^{1.5}h^1`$ (Padmanabhan 1993), and is also consistent with our choice for the sound speed of the disk (see § 2.3). The comoving volume element per unit redshift, $`dV/dz`$, is given by, $$\frac{dV}{dz}=\frac{4\pi cd_L^2}{(1+z)}\left|\frac{dt}{dz}\right|,$$ (25) where $`d_L`$ and $`dt/dz`$ are given in equations (6) and (7), respectively. Next we consider the value of the GRB rate, $`r_{GRB}`$. If GBRs are related to the final stages in the evolution of massive stars, then one may assume that the GRB rate, $`r_{GRB}`$, is proportional to the star formation rate (SFR). This assumption is justified as long as the time delay between the formation of a massive star and a GRB event is short compared to the Hubble time at the redshift of interest. We adopt this as our working hypothesis. Wijers et al. (1998) have estimated the constant of proportionality between $`r_{GRB}`$ and the cosmic SFR for an Eistein-de Sitter cosmology, but their results do not change significantly for other values of the cosmological parameters (Bagla 1999, private communication). Using the observed redshift history of the SFR per comoving volume in the universe (Lilly et al. 1996; Madau et al. 1996) and the observed distribution of $`\gamma `$-ray flux of GRBs, Wijers et al. (1998) have calibrated the GRB rate per comoving volume. Assuming that GRBs are standard candles, they have found a GRB rate at $`z=0`$ of $`R_{GRB}(0)=0.14\pm 0.02`$ Gpc<sup>-3</sup> yr<sup>-1</sup>. This estimate ignores recent corrections to the cosmic SFR. The same calculation, performed with a SFR per comoving volume that flattens at high redshift, as suggested by more recent corrections for dust extinction, still leads to similar results (Bagla 1999, private communication). Assuming that the proportionality constant does not vary with redshift, we write, $$r_{GRB}(z,M)=\frac{R_{GRB}(0)}{\dot{\rho }_{}(0)}\dot{M}_{}(z,M).$$ (26) Here $`\dot{\rho }_{}(0)`$ is the star formation rates per comoving volume in units of M yr<sup>-1</sup> Mpc<sup>-3</sup>, and $`\dot{M}_{}(z,M)=\alpha (z)\mathrm{\Omega }_bM`$ is the SFR within a particular halo in units of M yr<sup>-1</sup>. We calibrate the star formation efficiency as a function of redshift, $`\alpha (z)`$, so as to match the observed cosmic SFR, $`\dot{\rho }_{}`$, at $`z\stackrel{<}{}4`$, $$\dot{\rho }_{}(z)=_{M_{min}}^{\mathrm{}}\dot{M}_{}(z)\frac{dn}{dM}(z)𝑑M.$$ (27) An analytical fit for the observed $`\dot{\rho }_{}(z)`$ (assuming a Salpeter IMF and an extinction correction of A$`{}_{1500}{}^{}=1.2`$ mag) was derived by Madau & Pozzetti (1999), $$\dot{\rho }_{}(z)=\frac{0.23\mathrm{e}^{3.4z}}{\mathrm{e}^{3.8z}+44.7}\mathrm{M}_{}\mathrm{yr}^1\mathrm{Mpc}^3.$$ (28) At $`z>4`$ the SFR within a particular halo is calculated using the common prescription $`\alpha (z)t_{dyn}^1(z)`$, where $`t_{dyn}(z)`$ is the mean dynamical time of the halo. The constant of proportionality is chosen so as to match the observed cosmic SFR at $`z=4`$ (see Ciardi et al. 2000, for more details). Figure 2 presents the total number counts based on equation (24). The set of curves corresponds to different observed wavelengths. From the right to the left, $`\lambda _{obs}`$ is equal to 10 cm, 1 mm, 2 $`\mu `$m and 5000 Å. At low fluxes the number counts approach an asymptotic value. Out of the entire population of faint GRB afterglows, we predict that there are $`15`$ GRBs from redshifts $`z\stackrel{>}{}5`$ across the sky which are brighter than $`100`$ nJy at an observed wavelength of $`2\mu `$m. Figure 3 illustrates the contribution to the total number counts from different redshift bins, centered at $`z`$=2 \[Fig. (3a)\], 6 \[Fig. (3b)\] and 10 \[Fig. (3c)\]. Infrared or radio afterglows are detectable out to redshifts as high as 10, while optical afterglows are strongly absorbed by the intervening IGM at $`z\stackrel{>}{}5`$. Figure 4 shows the differential number count distribution per logarithmic flux interval and redshift interval of all sources at a given redshift. At high frequencies, for which the observed flux decreases with increasing redshift, low-redshift events dominate the counts at most fluxes. However, at low frequencies (i.e. millimetric wavelength) the high-redshift events dominate the counts at sufficiently low fluxes. ### 3.3 Dependence of Results on Model Assumptions Next we consider the dependence of our results on various model parameters. Figure 5a shows the observed flux at $`\lambda _{obs}=2\mu `$m and $`t=1`$ day for different choices of the mean ambient density. The standard case with $`P_{GRB}n^2`$ (solid line) is compared to two other cases; one where $`P_{GRB}n`$ (dot-dashed line) and the other where the ambient density is kept constant and equal to 1 cm<sup>-3</sup> (dotted line). The figure indicates that for observations in the infrared, all three cases provide identical results at $`z\stackrel{>}{}5`$, when one enters a regime where the observed flux becomes independent of the ambient density \[either the fast cooling regime with $`\nu >\nu _m`$ or the slow cooling regime with $`\nu >\nu _c`$, given by Eqs. (1) and (2)\]. Outside this regime the observed flux is roughly proportional to $`n^{5/6}`$ and $`\sqrt{n}`$ in the fast and slow cooling cases, respectively. In the slow cooling regime, for example, the condition $`\nu \nu _c`$ leads to a minimum redshift above which the flux is independent of ambient density $`(1+z)0.32E_{52}^1n_1^2t_d^1\left(\lambda _{\mathrm{obs}}/2\mu \mathrm{m}\right)^2`$. Hence, at long wavelengths or very low densities, the density independent regime is not reached. Figure 5a shows that the difference between the $`P_{GRB}n^2`$ and $`P_{GRB}n`$ cases, is small since they follow a similar evolution of the gas density with redshift. Figure 5b shows the number count of GRBs per logarithmic redshift interval at $`z=6`$, both for the standard case with $`P_{GRB}n^2`$ (solid line) and the case of a constant ambient density equal to 1 cm<sup>-3</sup> (dotted line). The difference between these cases is small. Figure 6a shows the dependence on the energy output of the afterglow number count from a redshift bin centered at $`z=6`$ for an observed wavelength of 2$`\mu `$m. We consider two cases that bracket the standard case ($`E=10^{52}`$ ergs); namely $`E=10^{53}`$ erg (dotted) and $`E=5\times 10^{50}`$ erg (dashed). The increase in the observed flux as the energy output increases \[see eqs. (1)–(5)\], is reflected in the afterglow number count. So far we have assumed that the GRB explosions are spherically-symmetric. If, however, the energy release is beamed, then the afterglow lightcurve is expected to evolve differently than we assumed. At early times, as long as the expansion Lorentz factor, $`\gamma `$, is still larger than the inverse of the jet opening angle, $`\theta ^1`$, the expansion behaves as if it were spherically symmetric with the same energy output per solid angle. Once the jet decelerates to a Lorentz factor, $`\gamma \stackrel{<}{}\theta ^1`$, the lightcurve declines more rapidly, since the jet starts to expand sideways and to reduce the mean energy output per solid angle (Rhoads 1999a,b; Panaitescu & Meszaros 1999). Finally, the isotropization of the energy ends when the expansion becomes sub-relativistic, at which time the lightcurve recovers the spherically-symmetric Taylor-Sedov evolution for the actual total energy output. (Synchrotron radiation losses are usually negligible in the late afterglow phase.) For an observing point which is aligned with the GRB jet axis, the afterglow lightcurve may start with an isotropic-equivalent energy of $`E`$ and end in the non-relativistic regime with the actual energy output of $`E_{sr}=\eta E`$ where $`\eta =(\mathrm{\Delta }\mathrm{\Omega }/4\pi )=2(\pi \theta ^2/4\pi )1`$ is the fraction of sky around the GRB source which is illuminated by two opposing jets of angular radius, $`\theta `$. Hence, the afterglow makes a transition between the lightcurve corresponding to a highly-relativistic explosion with a total energy $`E`$ to the lightcurve of a non-relativistic explosion with an energy $`\eta E`$. To examine the effect of beaming on our results, we model this transition using a linear interpolation between the two lightcurves. We start with an isotropic equivalent of $`E=10^{53}`$ergs, and adopt $`\theta =0.1`$ so that the actual energy output is $`E_{sr}=5\times 10^{50}`$ erg (since $`\gamma \stackrel{>}{}10^2\theta ^1`$, initially). As expected, the afterglow fluxes and source number counts make a transition between their values for the two boundary energies. Figures 6a-c show the transition between the two regimes with a dot-dashed line. We find that, for $`\lambda _{\mathrm{obs}}=2\mu `$m, the number counts are dominated by sources in the relativistic phase. However, at longer wavelengths, $`\lambda _{obs}=1`$ mm or 10 cm \[Fig. 6b or Fig. 6c, respectively\], the bulk of the emission takes place at late times when the expansion is in its mildly or sub-relativistic phase. Obviously, the effect of beaming can be recovered from the shape of the afterglow lightcurve in individual GRB events. Figure 6 shows that the existence of beaming might also be inferred from the number count statistics. ## 4 Discussion We have calculated the expected fluxes and number counts of high-redshift GRB afterglows, assuming that their rate of occurrence is proportional to the star formation rate. We have computed the observed properties of GRB lightcurves at different wavelengths ranging from the optical to the radio, treating both the relativistic and sub-relativistic phases of the expansion of their remnants. Our main result is that at a fixed observed time after the GRB event, the characteristic afterglow flux is not decreasing rapidly with increasing redshift (Figs. 1a, 1b), in contrast with other high-redshift sources, such as galaxies or quasars. Hence, the broad-band spectrum of GRB afterglows is ideally suited for probing the ionization state and metal content of the intergalactic medium at high redshifts, particularly during the epoch of reionization (Loeb 1999; Lamb & Reichart 1999). The only difficulty in using GRBs as probes of the high-redshift universe is that they are rare, and hence their detection requires surveys which cover a wide area of the sky (see the vertical axis in Figures 24). The simplest strategy for identifying high-redshift afterglows is through all-sky surveys in the $`\gamma `$-ray or X-ray regimes. In particular, detection of high redshift sources will become feasible with the high trigger rate provided by the forthcoming Swift satellite, to be launched in 2003. Swift is expected to observe and localize $``$300 GRBs per year, and to repoint within 20-70 seconds its on-board x-ray and UV-optical instrumentation for continued afterglow studies. The high-resolution GRB coordinates obtained by Swift, will be transmitted to Earth within $``$50 seconds. Deep follow-up observations will then become feasible from the ground or using the highly-sensitive infrared instruments on board the Next Generation Space Telescope (NGST), scheduled for launch in 2008. Swift is sufficiently sensitive to trigger on the $`\gamma `$-ray emission from GRBs at redshifts $`z\stackrel{>}{}10`$ (Lamb & Reichart 1999). We note that even if some GRBs occur outside of galaxies, the intergalactic medium would be sufficiently dense to produce an afterglow at these high-redshifts since its mean density is $`\stackrel{>}{}10^4[(1+z)/10]^3\mathrm{cm}^3`$. The effect of beaming is expected to be minor for sources with millimeter and radio fluxes $`\stackrel{<}{}10^5`$ Jy, since the number counts at these fluxes are dominated by sources which are in the sub-relativistic phase of their expansion, and for which the energy output in the explosion is already isotropized (see Fig. 6b, 6c). For a characteristic energy output of GRBs of $`10^{52}`$ ergs, our model implies that at any time there should be $`15`$ GRBs from redshifts $`z\stackrel{>}{}5`$ across the sky, which are brighter than $`100`$ nJy at an observed wavelength of 2 $`\mu `$m. The Next Generation Space Telescope (NGST) will be able to measure the spectra of these sources. Prior to reionization, the spectrum of GRB afterglows might reveal the long sought-after Gunn-Peterson trough (Gunn & Peterson 1965), due to absorption by the neutral intergalactic medium. We thank F. Haardt for providing us the IGM opacity curves, and J. Bagla and R. Croft for useful discussions. BC thanks the CfA pre-doctoral fellowship program for support during the course of this work. This work was also supported in part by NASA grants NAG 5-7039 and NAG 5-7768, NSF grant 9900077, and by a grant from the Israel-US BSF.
warning/0002/hep-ph0002056.html
ar5iv
text
# How precisely can we determine the 𝜋NN coupling constant from the isovector GMO sum rule? ## INTRODUCTION: ROBUST FORM OF THE GMO RELATION The analysis to determine the $`\pi `$NN coupling constant should be clear and easily reproducible. One should do a detailed study for the statistical and systematic errors. The precise determination is an absolute statement, it could be erroneous and it should be improvable. In this perspective the Goldberger-Miyazawa-Oehme (GMO) sum rule might be a good candidate. It is a forward dispersion relation at zero energy for $`\pi `$N scattering. It assumes scattering amplitudes to be analytical functions satisfying crossing symmetry. At first isospin symmetry does not have to be assumed and it reads (for more details see e.g. ) with its numerical coefficients: $`g_c^2/4\pi =4.50J^{}+103.3[(a_{\pi ^{}p}a_{\pi ^+p})/2]`$, where $`J^{}`$ is in mb the weighted integral, $`J^{}=(1/4\pi ^2)_0^{\mathrm{}}(dk/\sqrt{k^2+m_\pi ^2})[\sigma _{\pi ^{}p}^{Total}(k)\sigma _{\pi ^+p}^{Total}(k)]`$ and $`a_{\pi ^\pm p}`$ are the $`\pi ^\pm p`$ scattering lengths in units of $`m_\pi ^1`$. All ingredients are physical observables but so far the lack of precision in $`a_{\pi ^\pm p}`$ (contribution of 2/3 to $`g^2/4\pi `$) led to applications of the GMO relation as consistency check or constraint . The 1s width of the $`\pi ^{}p`$ atom determines $`a_{\pi ^{}p\pi ^0n}=0.128(6)m_\pi ^1`$ and assuming isospin symmetry this gives $`a^{}=(a_{\pi ^{}p}a_{\pi ^+p})/2`$ and $`g_c^2/4\pi =14.2(4)`$ using $`J^{}`$= -1.077(47) mb. This is not accurate enough although improvements will come . We here report on a possible way to improve the precision on $`g_c^2/4\pi `$ . As $`a_{\pi ^{}p}`$ is precisely known ($`0.0883(8)m_\pi ^1`$) from energy shift in pionic hydrogen one can write: $$g_c^2/4\pi =4.50J^{}+103.3a_{\pi ^{}p}103.3(\frac{a_{\pi ^{}p}+a_{\pi ^+p}}{2}).$$ (1) and using the above cited $`J^{}`$ (to be calculated later), $`g_c^2=4.85(22)+9.12(8)103.3[(a_{\pi ^{}p}+a_{\pi ^+p})/2]`$. This (not our final result) shows that all the action is in the term $`1/2(a_{\pi ^{}p}+a_{\pi ^+p})`$,which, assuming isospin symmetry, is $`a^+`$. If this quantity is positive $`g_c^2/4\pi `$ is smaller than 14, if it is negative it is larger. One way to determine the small $`a^+`$ is to use the accurate $`\pi ^{}d`$ scattering length $`a_{\pi ^{}d}=0.0261(5)m_\pi ^1,`$ from the pionic deuterium 1s energy level. To leading order this is the coherent sum of the $`\pi ^{}`$ scattering lengths from the proton and neutron, which, assuming charge symmetry (viz, $`a_{\pi ^+p}=a_{\pi ^{}n}`$) is the term required in our ’robust’ relation (1) The strong cancellation between the two terms is then done by the physics. In order to match the precision using the width, we only need a theoretical precision in the description of the deuteron scattering length to about 30%. ## ZERO-ENERGY $`\pi ^{}`$-DEUTERON SCATTERING AND $`a^+`$ In multiple scattering theory of zero-energy s-wave pion scattering from point-like nucleons and in the fixed scattering-center approximation, the leading contribution is: $`a_{\pi ^{}d}^{static}=S+D\mathrm{}`$ with $`S=[(1+m_\pi /M)/(1+m_\pi /M_d)](a_{\pi ^{}p}+a_{\pi ^{}n}),`$ $`M`$ and $`M_d`$ being the nucleon and deuteron masses respectively. The double scattering term $`D`$ is: $$D=2\frac{(1+m_\pi /M)^2}{(1+m_\pi /M_d)}[(\frac{a_{\pi ^{}p}+a_{\pi ^{}n}}{2})^22(\frac{a_{\pi ^{}p}a_{\pi ^{}n}}{2})^2]<1/r>,$$ (2) and with our final scattering lengths $`D=0.0256m_\pi ^1`$ quite close to $`a_{\pi ^{}d}`$ experimental. Fig. 1. Our graphical determination of the $`\pi N`$ scattering lengths in excellent agreement with the central values of the experimental PSI group , $`a^+=22(43)10^4m_\pi ^1;a^{}=905(42)10^4m_\pi ^1`$. We shall here follow the recent theoretical multiple scattering investigation of Baru and Kudryatsev (B-K) . The comparison of typical contributions is listed in Table 1. 1)Fermi motion: the nucleons have a momentum distribution which produces an attractive contribution, calculable to leading order from $`<p^2>`$ of the nucleon momenta in the deuteron. The uncertainty of 7 comes from the D-state percentage in the deuteron, P<sub>D</sub> =4.3% vs. 5.7% for the Machleidt1 vs. the Paris wave functions. 2) Absorption correction: the absorption reaction $`\pi ^{}`$d $``$nn, using 3-body Faddeev approaches produces a repulsive ($``$20%) contribution (not included in B-K). These studies were done carefully but a modern reinvestigation of this term is highly desirable. 3) ”s-p’ interference: a $``$15% correction was obtained by B-K. We find that it is a model dependent contribution due nearly entirely to the Born term the contribution of which vanishes exactly. We have then not considered this contribution. 4)Form factor: this non-local effect enters mainly via the dominant isovector $`\pi `$N s-wave interaction, closely linked to $`\rho `$ exchange. It represents only a correction of about $``$10%. 5) Non-static effects: these produce only a rather small correction of about 4%. There are systematic cancellations between single and double scattering as was first demonstrated by Fäldt . It has been numerically investigated by B-K and we have adopted their value, the error of 6 reflects a lack of independent verification. 6) Isospin violation, higher order terms, p-wave double scattering,virtual pion scattering: these corrections are all small and controllable. The isospin violation in the $`\pi `$N interaction comes in part from the $`\pi ^\pm \pi ^0`$ mass difference where an additional check comes from the chiral approach . Based on this, we obtain the preliminary, though nearly final, values $`(a_{\pi ^{}p}+a_{\pi ^{}n})/2=(17\pm 3(\mathrm{statistic})\pm 9(\mathrm{systematic}))10^4m_\pi ^1`$ and $`(a_{\pi ^{}p}a_{\pi ^{}n})/2=(900\pm 12)10^4m_\pi ^1.`$ Our values represent a substantial improvement in accuracy as seen in Fig. 1. The contribution of the scattering lengths to g$`{}_{c}{}^{2}/4\pi `$ has here a precision of about 1%. ## TOTAL CROSS SECTION INTEGRAL $`J^{}`$ The cross section integral contributes only one third to the GMO relation. Total cross sections are inherently accurate and their contribution is calculated with accuracy, but for the high energy region. The possibility of systematic effects in the difference must be considered, particularly since Coulomb corrections have opposite sign for $`\pi ^\pm `$p. The only previous evaluation with a detailed discussion of errors is that by Koch . Later evaluations given in Table 2 find values within the errors, but the uncertainties are not stated and analyzed. In view of obtaining a clear picture of the origin of uncertainties we have reexamined this problem in spite of the consensus. We limit the discussion to the critical features. The typical shape of the integrands is shown in Fig. 2 up to 2 Gev/c. Fig. 2. The J<sup>-</sup> separate integrands for $`\pi ^\pm `$p as well as their difference as function of k<sub>lab</sub> together with the cumulative value of the integral J$`{}_{}{}^{}(k_{lab})`$ from the region $`0<k<k_{lab}`$. The integrands are in units of mb GeV/c. There are no total cross-section measurements below 160 MeV/c, but the accurately known $`a_{\pi ^\pm p}`$ give a strong constraint assuming isospin symmetry. The s- and p-wave contributions nearly cancels. A tiny correction occurs, since isospin is broken by the 3.3 MeV lower threshold for the $`\pi ^0`$n channel below the physical $`\pi ^{}`$p threshold. The main contributions come from the region of the $`\mathrm{\Delta }`$ resonance and just above. There are no strong cancellations in the difference between $`\pi ^\pm `$p cross sections in that region and the cross sections have been very carefully analyzed. We have first evaluated the hadronic cross sections up to 2 GeV/c based on the VPI phase-shift solution . In doing so Coulomb corrections and penetration factors have been taken into account in the adjustment to experimental data even if the treatment may not be optimal. It also allows for some isospin breaking, since the $`\mathrm{\Delta }`$ mass splitting is parameterized . In view of the not so high accuracy we aim for, this should be adequate. Bugg has emphasized that in the $`\pi ^+`$p scattering the total cross sections are systematically reduced at all energies by the Coulomb repulsion between the particles and, conversely, enhanced in $`\pi ^{}`$p scattering. One must correct for this effect, which gives a negative contribution to $`J^{}`$. Having made no correction for it at higher energies means that we will underestimate the coupling constant somewhat. In the region around 500 MeV/c there are long-standing problems with the experimental total cross section data . This uncertainty, larger than the Coulomb penetrability effects, should be resolved. So we have preferred to use the SM95 PWA solution as the best guide. The real uncertainty in J<sup>-</sup> comes from the high energy region and is linked to the relatively slow convergence of the integral. We have evaluated the different contributions (see Table 3) with no other Coulomb and penetration corrections than those introduced by the experimental authors above 2 GeV or by the theoretical analysis below 2 GeV. We find, based on (integration of hadronic cross section) the SM95 and Arndt 12/98 analysis below 2 GeV/c , and on the Regge pole PDG94 and PDG98 extrapolation beyond 240 GeV/c, the values $`J^{}=(1.075\pm 0.008)(30)`$ mb and $`(1.114\pm 0.008)(30)`$ mb respectively. We have adopted the mean value $`J^{}=(1.095\pm 0.008)(30)`$ mb. In our calculation we have added a systematic uncertainty from Coulomb penetration effect of $`\pm 0.017`$ from the region less than 2 GeV/c. ## RESULTS AND CONCLUSIONS We have derived first new values for the $`\pi `$N scattering lengths from the $`\pi ^{}`$d one: $$a^+\frac{a_{\pi ^{}p}+a_{\pi ^{}n}}{2}=(17\pm 3)(9))10^4m_\pi ^1,a^{}\frac{a_{\pi ^{}p}a_{\pi ^{}n}}{2}=900(12)10^4m_\pi ^1.$$ Our second conclusion concerns the charged $`\pi `$NN coupling constant using these new accurate values in (1) with $`J^{}=(1.095\pm 0.008)(30)`$ and charge symmetry: $$g_c^2/4\pi =(4.93\pm 0.04)(14)+(9.12\pm 0.08)+(0.18\pm 0.03)(9)=(14.23\pm 0.09)(17).$$ (3) The uncertainty comes mainly from $`J^{}`$. This coupling constant which agrees quite well with the text book value, 14.28(18) is intermediate between the low value deduced from the large data banks of NN and $`\pi `$N scattering data and the high value from np charge exchange cross sections . It is fully compatible with the latter, differing statistically by only about one standard deviation.
warning/0002/astro-ph0002182.html
ar5iv
text
# Stellar Evolution and Large Extra Dimensions ## I Introduction Recently there has been a revived interest in the physics of extra-spatial dimensions. In order to provide a framework of solving the hierarchy problem, in refs. , the fundamental Planck scale - where gravity becomes comparable in strength with the other interactions - was taken to be near the weak scale. The observed weakness of gravity at long distances is due to the presence of $`n`$ new spatial dimensions, with size $`R`$ which are large compared to the electroweak scale. The relation between the Planck mass in $`4`$ dimensions ($`M_{Pl}=\sqrt{\mathrm{}c/G_N}=\mathrm{1.2\hspace{0.17em}10}^{19}`$GeV/c<sup>2</sup>) and that in $`4+n`$ dimensions ($`M_s`$) is $$R^n=(\mathrm{}/c)^nM_{Pl}^2/(M_s^{n+2}\mathrm{\Omega }_n)$$ (1) where $`\mathrm{\Omega }_n`$ is the volume of the n-dimensional sphere with unit radius. Laboratory limits, essentially from LEP II give a lower bound on $`M_s`$ of about 1 TeV/c<sup>2</sup>. The choice $`M_s1`$ TeV/c<sup>2</sup> yields $`R10^{32/n17}`$ cm. The case $`n=1`$ gives $`R10^{15}`$cm which is excluded since it would modify newtonian gravitation at solar system distances. Already for $`n=2`$ one has $`R1`$mm which is the distance where our present experimental measurement of gravitational forces stops, and one needs information from different sources. In this context, one should remind that in last decades the improved knowledge of several physical mechanisms has allowed astrophysicists to produce stellar models with a significant degree of reliability. As a matter of fact, current stellar models nicely reproduce the large variety of stellar structures populating the sky, passing also some subtle tests as the ones recently provided by seismologic investigations of our sun. Such a success has already opened the way of using stellar structures as a natural laboratory to test the space allowed for new physics, i.e., to investigate the allowed modifications of the current physical scenario . This looks as a quite relevant opportunity, bearing in mind that a stellar structure is governed by the whole ensemble of physical laws investigated in terrestrial laboratories and that these stellar structures, in varying their mass and ages, experience a range of physical situations not yet reached in current laboratory experiments. On this basis, the “stellar laboratory” has already provided relevant constraints on several physical ingredient as, e.g., the existence of Weak Interacting Massive Particles or the neutrino magnetic moments . Astrophysical constraints on large extra dimensions have been discussed in and in . The main effect of large extra dimensions arises from the production of the Kaluza-Klein (KK) excitations of the graviton. The KK-graviton and matter interactions are of gravitational strength, so the KK states never become thermalized and always freely escape. The associated energy loss (through photon-photon annihilation, electron-positron annihilation, gravi-Compton-Primakoff scattering, gravi-bremsstrahlung, nucleon-nucleon bremsstrahlung) have been calculated in and observational constraints on $`M_s`$ have been derived from simple considerations on the energetics of sun, red giants and supernovae. In this paper we discuss in more detail the information on large extra dimensions which can be derived in the framework of stellar evolution theory and observation. The first part is devoted to the study of the sun, which represents a privileged laboratory in view of the richness and accuracy of available data. In particular we shall consider the following topics: i)As well known there is a remarkable agreement between the predictions of the Standard Solar Model (SSM) and the results of helioseismic observations, see e.g. . Production of KK gravitons provides a new energy loss, which will become incompatible with helioseismic constraints if the $`4+n`$ dimensional Planck mass $`M_s`$ is sufficiently low. In this way we shall determine lower limits on $`M_s`$ from helioseismic observations. ii)Despite its several successes, the standard solar model presents us with some puzzles, e.g. the deficit of solar neutrinos, see e.g. , the depletion of the photospheric lithium abundance, see e.g. , and - perhaps - an underestimate of the sound speed just below the convective envelope, see e.g. . Could it be that the new physics of KK-graviton production accounts for some of these anomalies? In addition, the efficiency of KK-graviton energy loss appears strongly dependent on the temperature. This suggests to consider stars experiencing internal temperatures much larger than in the sun and which, in turn, are particularly sensitive to the efficiency of cooling mechanisms. In this way the investigation will be extended to red giants structures, which will provide a much more stringent limit on $`M_s`$. In section II we give a first look at KK-graviton production in the sun, determining the order of magnitude of the acceptable $`M_s`$ and presenting the structure of solar models where the new energy loss is relevant. In section III we shall determine the helioseismic constraints on $`M_s`$, from data on the photospheric helium abundance and on sound speed in the energy production region. The effect of KK-graviton production on the “solar puzzles” mentioned above is discussed in sect. IV. Sect. V will be devoted to red giant stars. Our conclusions are summarized at the end of the paper, whereas in the appendix we collect the relevant formulas for the energy losses. ## II A first look at the effects of KK-graviton production in the sun It is interesting to compare the energy loss due to KK-graviton production with the energy production from the pp chain at the center of the sun. The values of density, temperature and chemical composition derived from the SSM of are presented in Table I. The results of other SSM calculations are similar, see e.g. . Energy loss and production rates, computed according to the results of Appendix A, are compared in Table II. The most important contribution always comes from the photon-photon annihilation. One expects that the solar solar structure would be drastically modified if the energy loss due to KK-gravitons becomes comparable with the nuclear energy production rate. In this way one can derive the following lower limits on $`M_s`$: $$n=2:M_s>140GeV/c^2$$ (2) $$n=3:M_s>3.5GeV/c^2$$ (3) This result which is essentially the same as that in ref. suggests that we concentrate on the $`n=2`$ case only. So far we assumed just a rough knowledge of the solar structure. One can expect that more detailed information, as that provided by helioseismology, provide more stringent constraints. To understand in more detail the effect of KK-graviton production we have built solar models which include this additional energy loss. The energy generation subroutine was modified so as to include the KK-graviton loss and the stellar evolution code FRANEC was run by varying the three free parameters of the model (initial helium abundance $`Y_{in}`$, initial metal abundance $`Z_{in}`$ and mixing length $`\alpha `$) until it provides a solar structure (i.e. it reproduces the observed solar luminosity, radius and photospheric metal abundance at the solar age). As an example, we present here the case $`M_s=0.2`$ TeV/c<sup>2</sup>. The main differences with respect to our SSM are depicted in Table III and Fig. 1. Several features can be easily understood by observing that the solar model with KK-graviton production has to produce, now and in the past, a higher amount of nuclear energy, in order to compensate for the additional energy loss. More hydrogen has been burnt into helium, and the initial helium abundance has to be reduced with respect to the SSM (otherwise one would get a stellar structure which, being too much helium rich, would be presently overluminous). Consequently, the present photospheric helium abundance $`Y_{ph}`$ is decreased with respect to the SSM prediction. Nevertheless, the central helium abundance is still somehow larger than in SSM and more energy is being produced in order to compensate for the KK-graviton losses. This is achieved with a somehow larger central temperature. In the solar core, both temperature and “mean molecular weight” are thus higher than in the SSM, so that one cannot a priori decide for the behaviour of the sound speed. In fact Fig. 1 shows a decrease near the center and a significant increase near $`R=0.2R_{}`$, i.e. in a region where helioseismic determinations are still very accurate. These observations will be useful for determining the relevant observables which are sensitive to $`M_s`$ and which can be constrained by means of helioseismology. In Fig. 2 we also compare the nuclear energy production rate with the losses due to KK-gravitons. The results are consistent with the qualitative energetics analysis discussed above. ## III Helioseismic constraints on $`M_s`$ Helioseismology provides detailed information on several solar properties. In particular, the sound speed profile and the photospheric helium abundance $`Y_{ph}`$ are determined with high accuracy. In ref. it was estimated that the isothermal sound speed squared, $`u=P/\rho `$ at distance $`R=0.2R_{}`$ is determined with an accuracy of about $`\mathrm{\Delta }u/u110^3`$, $$u_{0.2}^{}=(1.238\pm 0.001)10^{15}cm^2/s^2$$ (4) This uncertainty, defined as the “statistical” or “one sigma” error , was obtained by taking into account all possible contributions arising from: i) measurement errors, ii) the inversion method and iii) the choice of the reference model (the recent analysis of confirms the estimate of for each contribution to the uncertainty). These estimated errors were added in quadrature. With a similar attitude the uncertainty of $`Y_{ph}`$ was estimated: $$Y_{ph}^{}=0.249\pm 0.003$$ (5) Recent accurate standard solar model calculations are successful in reproducing sound speed in the energy production region as well as the photospheric helium abundance, their predictions being quite close to the central helioseismic estimates, see e.g. . On the other hand, as discussed in the previous section, both quantities are sensitive to the energy loss due to KK-graviton production For this reason, we concentrate here on $`Y_{ph}`$ and on the value of $`u`$ at $`R=0.2R_{}`$ hereafter $`u_{0.2}`$. We have built a series of solar models with $`M_s`$ in the range of few hundred GeV/c<sup>2</sup> in order to determine the dependence of both observables on $`M_s`$, see Figs. 3 and 4. For each observable $`Q`$ the results have been parametrized in the form $$Q(M_s)=Q_{SSM}(1+(m/M_s)^\alpha ),$$ (6) with the results for the parameters $`m`$ and $`\alpha `$ shown in Table IV. By requiring that the differences in the calculated observables do not exceed the helioseismic uncertainty, we get the following lower bounds on $`M_s`$: -from $`Y_{ph}`$ : $`M_s>0.23`$ TeV/c<sup>2</sup> -from $`u_{0.2}`$ : $`M_s>0.31`$ TeV/c<sup>2</sup> These bounds are stronger that that of Eq. (2), which was obtained by using crude energetical considerations. However, the accuracy of helioseismic method has yielded an improvement of just a factor of two. In fact, KK-graviton energy loss rate $`ϵ_{KK}`$ depends on high powers of $`M_s`$, so that drastic changements of $`ϵ_{KK}`$ result from just tiny modifications of $`M_s`$. ## IV Extra-dimensions and the puzzles of the SSM As well known, in front of its several successes, the standard solar model presents us with some puzzles: i) The signals measured by all solar neutrinos experiments are systematically lower than those predicted by SSMs, an effect which is now commonly ascribed to neutrino oscillations. ii) The observed photospheric lithium abundance is a factor of hundred smaller than the meteoritic value . Lithium is being continuosly mixed in the convective envelope, however -according to the SSM - it should not be destroyed by nuclear reactions since even at the bottom of the convective zone the temperature is not high enough to burn it. This signals some deficiency of the standard solar model, which is built in a one dimensional approximation and neglects rotation, see . iii) The helioseismically determined sound speed just below the convective envelope is somehow smaller (by 0.4%) with respect to the predictions of the most recent and accurate SSM calculations, see e.g. . It is thus natural to ask what is the effect of the hypothetical large extra dimensions on these items. <sup>*</sup><sup>*</sup>*We recall that in conversion of electron neutrinos to the light fermions propagating in the bulk of $`4+n`$ dimensions has been considered as a solution of the solar neutrino problem. Concerning solar neutrinos, the answer is already contained in the previous discussion. When KK-graviton production is effective, the central temperature increases and consequently the production of Beryllium and Boron neutrinos is increased, see Table III. KK-graviton production would thus make the neutrino puzzle even more serious. At the bottom of the convective zone the temperature would be even smaller than that predicted by SSM, see Table III, so that there is no help in lithium burning. Sound speed just below the convective enevelope is practically unchanged with respect the SSM, so that the disagreement cannot be affected. In short, KK graviton production would provide no cure to the SSM puzzles. ## V Red giants and KK-gravitons A glance at the current evolutionary scenario easily indicates low-mass Red Giant Branch (RGB) stars as good candidates for investigating the effects of KK-graviton production. As a matter of fact a RGB star reaches internal temperatures of the order of $`10^8`$ K. Moreover, the structure of RGB stars is quite sensitive to the cooling mechanisms which regulate the size of the He core at the He ignition. The size of He core in turn governs several observational quantities both in these RGB structures as well as in the subsequent phase of central He burning (Horizontal Branch, HB) stars. We will follow this approach discussing the effect of KK-graviton cooling on the evolution of suitable RGB structures. Comparison of theoretical predictions with available experimental (i.e. observational) data will allow to put more stringent constraints on the minimal $`4+n`$ dimensional Planck mass $`M_s`$. To perform our investigation we used our latest version of the FRANEC evolutionary code to predict the observational properties of stellar models with different metallicities but with a common age of the order of 10 Gyr, thus adequate for RGB stars actually evolving in galactic globular stellar clusters (GCs). In order to make more clear to the reader the following discussion, in Fig. 5 we show the typical Hertzprung-Russel diagram for a galactic GC (upper panel) and the corresponding theoretical one (lower panel) as obtained by using the prescriptions provided by our own computations. The most relevant evolutionary phases and observational features are clearly marked. The diagram represents the locus of stars for a given chemical composition and age but different masses. As the mass increases, the star moves from the Main Sequence location (H central burning phase) to the RGB (H shell burning phase) till reaching a maximum luminosity where the central He ignition occurs (RGB tip), driving the structure to the central He burning (Horizontal Branch) phase. Numerical experiments disclose that a stronger cooling has a little effect on the morphology of the diagram depicted in Fig. 5, but severe consequences on the internal structure of the star. Fig. 6 shows the predicted time dependence of the central temperature– density relation for selected values of $`M_s`$ and $`n=2`$. As expected, one finds that by increasing $`M_s`$ the efficiency of the extra-cooling decreases. Above $`M_s5`$ TeV/c<sup>2</sup> the effects on the evolutionary history of the stellar structure vanish. Even a quick inspection of data in Fig. 6 reveals that the assumption $`M_s1.5`$ TeV/c<sup>2</sup> (i.e. already above the current accelerator lower limit for $`M_s`$) is deeply affecting the structure so that one expects strong observational consequences. As a matter of fact, by exploring the case $`M_s`$=1 TeV/c<sup>2</sup> (the previous lower limit) one finds that RGB stars would fail to ignite Helium, running against the well-established evidence of Helium burning star in galactic GCs. Fig. 6 shows that increasing the cooling for each given central density the central temperature is lower, as expected. According to well known prescriptions of the stellar evolutionary theory, one can thus easily predict that the end of the RGB phase – i.e. the central He ignition – will be delayed and the mass of the He core at this stage will be larger. To discuss this point in some detail, we show in Fig. 7 the He core mass at the central He ignition (RGB tip) for selected assumptions about the value of $`M_s`$ and n=2. As shown in the same Fig. 7 in order to cover the range of metallicity (Z) spanned by the galactic GCs, computations have been performed for a $`0.8M_{}`$ star with Z=0.0002 and for $`1M_{}`$ star with Z=0.02. To constrain the value of $`M_s`$ one has now to discuss theoretical results in terms of observable quantities. In this context one finds that the extra-cooling is governing two main observational parameters: i) the luminosity of the RGB tip, ii) the luminosity of He burning HB stars. In both cases the stronger the extra-cooling the larger is the predicted luminosity. In this paper, we focus our attention on the first parameter only. Several papers have already remarked the good agreement between observations and standard model theoretical predictions . Such good agreement is shown also in Fig. 8 where we report luminosity (in bolometric absolute magnitudes) of the brightest observed RGB star in clusters with different metallicities as compared with theoretical predictions for the canonical scenario ($`M_s\mathrm{}`$). According to the discussions given in several papers (see e.g. ) the theoretical predictions should represent within about 0.1 mag. the upper envelope of the observed star luminosity, and this is precisely what one finds in Fig. 8. However, the same figure shows that for finite value of $`M_s`$ theoretical predictions move toward larger luminosity, in disagreement with observations. By inspection of data in Fig. 8 one can conclude conservatively that values of $`Ms`$ 3 TeV/c<sup>2</sup> are definitively ruled out by the observational tests, whereas a lower limit of 4 TeV/c<sup>2</sup> appears reasonably acceptable. Thus this detailed evolutionary investigation has improved the crude estimate of ref $`M_s>2`$ TeV/c<sup>2</sup>. ## VI Concluding remarks We summarize here the main points of this paper: i)Helioseismic constraints on the sound speed in the energy production region and on the photospheric helium abundance rule out values of the $`4+2`$ dimensional Planck mass below $`M_s=0.3`$ TeV/c<sup>2</sup>. ii)The introduction of additional energy loss due to KK-graviton production cannot be a cure to the puzzles posed by SSM calculations. In particular, the predicted neutrino signals would be even larger than those of the SSM. iii)Observational constraints for Red Giant stars evolving in galactic globulars imply $`M_s>34`$ TeV/c<sup>2</sup>. This bound is stronger than that provided by accelerator, thus indicating how useful it is, and hopefully it will be, the synergetic use of terrestrial and stellar laboratories. ###### Acknowledgements. We are extremely grateful to Z. Berezhiani, D. Comelli and F. Villante for discussions. This work is co-financed by the Ministero dell’Università e della Ricerca Scientifica e Tecnologica (MURST) within the “Astroparticle Physics” project. ## A Star energy-loss via KK-gravitons The energy loss rate per unit mass due to escaping KK gravitons has been calculated in . Three processes are important for KK-graviton production in the sun and in the red giants. The relevant formulas are collected below, in natural units as well as in units more useful for implementation in a stellar evolution code. For a comparison, the energy production rate per unit mass due to the pp-chain is also parametrized. ### a Photon photon annihilation to KK gravitons: $`\gamma +\gamma grav`$ When $`n`$ extra dimensions are effective, the Newtonian interaction potential \[ $`V1/(M_{pl}^2r)`$\]is modified to $`V1/(M_s^{n+2}r^{1+n})`$, so that the coupling of each particle to the gravitational field is proportional to $`1/M_s^{1+n/2}`$ and KK-graviton production cross sections are proportional to the square of this quantity. A thermal photon gas is uniquely specified by its temperature T and fundamental physical constants $`(\mathrm{},c`$ and $`K_B`$) so that dimensional considerations fix the dependence of the energy loss rate per unit volume $`Q_\gamma `$. In natural units, this has dimension of \[Energy\]<sup>5</sup>, so that one has $`Q_\gamma =A_\gamma (n)M_s^{n2}T^{n+7}`$ where $`A_\gamma (n)`$ are numerical coefficients given in eq. (7) of . The energy loss rate per unit mass is obtained by dividing Q by the mass density $`\rho `$. When temperature is in expressed in Kelvin degress, density in g/cm<sup>3</sup> and $`M_s`$ in TeV/c<sup>2</sup>, the energy loss $`ϵ_\gamma `$ in erg/g/s is thus: $$n=2ϵ_\gamma =7.2510^{66}\frac{T^9}{\rho M_s^4}$$ (A1) $$n=3ϵ_\gamma =4.4210^{82}\frac{T^{10}}{\rho M_s^5}$$ (A2) ### b Gravi-compton Primakoff scattering: $`\gamma +ee+grav`$ The expression for the energy loss is in this case: $$ϵ_{GCP}=B(n)\frac{\alpha }{m_e}\frac{n_e}{\rho }\frac{T^{n+5}}{M_s^{n+2}}$$ (A3) where the numerical coefficients $`B(n)`$ are found in eq. 15 of , $`m_e(n_e)`$ is the electron mass (numerical density). The dependence on $`M_s`$ is easily understood from the previous considerations, $`\alpha `$ comes in from electro-magnetic coupling of the electron and the factor $`n_e/\rho `$ clearly expresses the proportionality to the electron number per unit mass. Dimensional analysis is not sufficient to fully specify the dependence on temperature due to the presence of another mass scale, $`m_e`$, which is relevant for non-relativistic electrons, see . In the same units as in eq. (A1,A2): $$n=2ϵ_{GCP}=1.6910^{78}\frac{T^7n_e}{\rho M_s^4}$$ (A4) $$n=3ϵ_{GCP}=5.6010^{94}\frac{T^8n_e}{\rho M_s^5}$$ (A5) ### c Gravi-Bremsstrahlung: $`e+Ze+Z+grav`$ The energy loss is now: $$ϵ_{GB}=C(n)\alpha ^2\frac{n_e}{\rho }\frac{T^{n+1}}{M_s^{n+2}}\mathrm{\Sigma }_jn_jZ_j^2$$ (A6) where $`n_j`$ is the number density of nuclei with atomic number $`Z_j`$ and the numerical factors $`C(n)`$ are given in eq. (21) of . In the same units of eqs. (A1,A2) one has: $$n=2:ϵ_{GB}=5.8610^{75}\frac{T^3n_e}{\rho M_s^4}\mathrm{\Sigma }_jn_jZ_j^2$$ (A7) $$n=3:ϵ_{GB}=9.7410^{91}\frac{T^4n_e}{\rho M_s^5}\mathrm{\Sigma }_jn_jZ_j^2$$ (A8) The total energy loss due to KK-graviton production is $$ϵ_{KK}=ϵ_\gamma +ϵ_{GCP}+ϵ_{GB}$$ (A9) It is useful to compare the above energy losses with the e.m. energy production rate per unit mass from the pp-chain. The slowest reaction of the chain is the $`p+pd+e^++\nu _e`$ which in the temperature region of interest for the sun, has a rate $`<\sigma v>_{pp}=AT^{3.83}`$, with $`A=4.39810^{71}`$ cm<sup>3</sup>/s and $`T`$ is expressed in Kelvin. The energy production rate per unit mass through the ppI termination of the pp-chain is: $$ϵ_{ppI}=\frac{1}{4}\rho \frac{X^2}{m_H^2}Q_{em}<\sigma v>_{pp}$$ (A10) where $`Q_{em}=26.1`$ MeV is the average e.m. energy released in the $`4p+2e^{}^4He+2\nu _e`$, $`X`$ is the H-mass fraction and $`m_H`$ is the hydrogen mass. In the same units as in Eq. (A1) one has: $$ϵ_{ppI}=2.9710^{27}X^2\rho T^{3.83}$$ (A11) As well known the ppI branch is the main energy source in the Sun. Everywhere it gives the strongest contribution to the energy production rate: $$ϵ_{nuc}=ϵ_{ppI}+ϵ_{ppII}+ϵ_{ppIII}+ϵ_{CNO}.$$ (A12)
warning/0002/cond-mat0002224.html
ar5iv
text
# Ice XII in its second regime of metastability ## Abstract We present neutron powder diffraction results which give unambiguous evidence for the formation of the recently identified new crystalline ice phase , labeled ice XII, at completely different conditions. Ice XII is produced here by compressing hexagonal ice I<sub>h</sub> at T = 77, 100, 140 and 160 K up to 1.8 GPa. It can be maintained at ambient pressure in the temperature range 1.5 $`<`$ T $`<`$ 135 K. High resolution diffraction is carried out at T = 1.5 K and ambient pressure on ice XII and accurate structural properties are obtained from Rietveld refinement. At T = 140 and 160 K additionally ice III/IX is formed. The increasing amount of ice III/IX with increasing temperature gives an upper limit of T $``$ 150 K for the successful formation of ice XII with the presented procedure. Although, water has been extensively studied both experimentally and theoretically it still rewards us with new and unexpected properties. This has been demonstrated recently by (i) the detection of polyamorphism and (ii) the discovery of a new crystalline phase (ice XII) . Having been observed in different regions of water’s phase diagram the two phenomena were originally thought to be disconnected (Figure 1). The phenomenon of polyamorphism, i.e. the existence of two distinct amorphous phases, is still lacking a comprehensive understanding . Although polyamorphism is equally observed in other substances particularly interesting explanations have been put forward for water. These ideas are based on computer simulations and link polyamorphism to particularities of the supercooled liquid like phase segregation and a second critical point. Due to homogeneous crystallization supercooled water is experimentally inaccessible in the region where these phenomena are expected. The hypotheses must, therefore, be checked indirectly, e.g. by establishing the glassy character of the amorphous phases. The formation of the high–density amorphous ice (HDA) — achieved by compressing crystalline hexagonal ice I<sub>h</sub> at temperatures below 150 K to pressures exceeding 1 GPa (Fig. 1) — has received particular attention in this context . So far, the formation of HDA from ice I<sub>h</sub> has been reported as a well-defined transition channel. The contamination of the amorphous samples by crystalline impurities has been granted little attention . Only recently strong experimental indications have become available which imply that all these contaminations correspond to ice XII. As ice XII was originally observed in a completely different region of water’s phase diagram this shows that it is a rather prolific phase of water. Moreover, the co-production of ice XII has, as we will argue in the conclusions, far reaching implications for the I<sub>h</sub> to HDA transformation and, thus, on the origin of water’s polyamorphism. In this letter we show that the structure of ice XII produced at low temperatures is definitely identical with the phase characterized by Lobban et al. at higher temperatures. Furthermore, we find that no continuous connection between the two regions of apparent metastability exists. And finally, we identify the conditions which define whether the compression of I<sub>h</sub> results in ice XII or HDA. Our results are based on high–resolution neutron powder diffraction experiments on samples which are produced at various temperatures. All samples are treated in the following way. About 2.5 ml of D<sub>2</sub>O (purity 99.9, resistivity 1 M$`\mathrm{\Omega }`$cm) is frozen to common hexagonal ice I<sub>h</sub> and cooled to 77 K in a piston–cylinder apparatus. The samples are heated to the desired temperatures, namely 77, 100, 140 and 160 K and tempered for about 30 minutes. Each sample is then compressed to a maximum nominal pressure of 1.8 GPa. The rate of compression is 1 GPa/min at 77 K and 0.5 GPa/min at all other temperatures. Once the maximum pressure is attained the samples are cooled back to 77 K and finally recovered from the pressure device in liquid nitrogen, where they are powdered by using a mortar and a pestle. No pressure changes which could be interpreted as signs of phase transitions are observed in the course of cooling under pressure. The sample temperature as measured at the bottom of the pressure cylinder apparatus does not increase by more than 10 K during the compression phase. The diffraction experiments are carried out on the high–resolution diffractometer D2B at the Institute–Laue–Langevin in Grenoble, France . A Vanadium sample holder (diameter = 7 mm) and a standard cryostat are chosen as sample environment . Two different experimental setups are applied. The accurate structure determination of ice XII is carried out at 1.5 K on the sample produced at 77 K. For this measurement the horizontal incident beam divergence $`\alpha _1`$ is set to 10’ and the monochromator aperture (MA) to 10 mm. The detector is moved using $`\mathrm{\Delta }\mathrm{\Theta }=0.05^{}`$ steps. In this high resolution mode, the applied wavelength $`\lambda =1.59427\pm 0.00006`$ Å and the instrumental resolution are determined from an independent measurement of a silicon powder sample. For good statistics data are collected over 15 hours. All other measurements are performed at 110 K, with $`\alpha _1`$ = 30’ and MA = 50 mm, $`\mathrm{\Delta }\mathrm{\Theta }=0.05^{}`$ at the given wavelength. These lower resolution data are collected for 120 minutes (sample prepared at 77 K) and 30 minutes (all other samples) which is sufficient for a structural identification of the samples. Figure 2 displays the results of the high–resolution measurement at 1.5 K. A Rietveld refinement is performed without any constrains on the parameters using the program Fullprof . Our refinement confirms that the structure of the observed ice phase is unambiguously that of ice XII . This structure is characterized by twelve water molecules arranged in a tetragonal unit cell meeting the symmetry space group I4̄2d. The refined unit cell parameters are $`a=8.2816\pm 0.0002`$ Å and $`b=4.0361\pm 0.0001`$ Å which result in a calculated microscopic density of $`\rho =1.4397\pm 0.0003`$ g/cm<sup>3</sup> (D<sub>2</sub>O). We present in Table LABEL:tab1 the refined fractional coordinates and thermal factors and in Table LABEL:tab2 the calculated intra– and intermolecular distances and bond angles. Taking into account the different preparational conditions which are used here and in reference the cell constants, mass densities and atomic parameters are in good agreement with each other. Profile features in the diffraction pattern (Figure 2) which are not due to ice XII can be attributed to the sample environment and a slight contamination of the sample by some untransformed I<sub>h</sub> and simultaneously produced amorphous ice. This contamination is taken into account in the refinement by profile (I<sub>h</sub>, symmetry space group P$`6_3`$/mmc) and background (amorphous ice) matching . The peak which due to the sample environment is excluded from the calculation. The formation of amorphous ice shows the ambivalent nature of the applied pressure induced transition. The final residuals of the refinement are $`R_{wp}=3.48\%`$ and $`R_p=2.69\%`$. The diffraction results for samples prepared at different temperatures are compared in Figure 3. The presence of ice XII can be clearly identified in all of the data, although, its diffraction pattern is obscured progressively by an additional contribution at 140 and 160 K, respectively. To identify this additional contribution profile matching methods are used for ice phases expected in this region of water’s phase diagram. The most promising indexing scheme is given by the symmetry space group P$`4_12_12`$. This symmetry space group is uniquely inherent to ice III and ice IX, which differ in the degree of order in their proton sublattices . As in ice XII, twelve water molecules inhabit the tetragonal unit cell of ice III/IX. Table III gives the unit cell constants and the corresponding mass densities of ice XII and ice III/IX as determined by the profile matching. The profile shape function, which is used in the matching procedure, is fully accounted for by the instrument resolution function as determined from the samples prepared at 77 and 100 K (Figure 3a,b). As a consequence only the cell constants (Table III) are freely adjustable for the phase mixtures. Typical residuals of the matching for the ice XII and ice III/IX mixtures are $`R_{wp}6\%`$ and $`R_p8\%`$. The inset of Figure 3 shows a diffraction pattern taken at $`\lambda =3.00`$ Å on the time–of–flight spectrometer IN5 of a sample prepared at $`T165`$ K. The absence of the (220) peak of ice XII which should be well resolved at $`2\mathrm{\Theta }70^{}`$ indicates that the content of ice XII is negligibly small. Thus, above $`T>160`$ K solely ice III/IX is formed. The pressure induced formation of ice III/IX from I<sub>h</sub> at temperatures exceeding 140 K is in agreement with recent extensive studies . Formation of ice V at temperatures exceeding 180 K is not observed in any of our samples. This shows that despite the elevated compression rates used in our work the temperature of the sample during the compression stage stayed close to the cell temperature. Summarizing, we have demonstrated that ice I<sub>h</sub> can be successfully compacted either into the high–density amorphous state (HDA) or into the crystalline phase ice XII. Both phases are produced by application of pressure exceeding 1.0 GPa at temperatures below 150 K (Figure 1). Ice XII is a rather prolific feature in waters phase diagram. Its formation in two seemingly disconnected regions is unusual but not entirely surprising when we consider the fact that the formation is governed by dynamic variables like the compression rate, and, in addition, is in competition with other crystalline and amorphous phases. Ice XII can be recovered at ambient pressure and low temperature and can be stored at temperatures lower than 135 K. At higher temperatures it starts to transform apparently to the metastable cubic phase which itself is a precursor of the stable I<sub>h</sub> form. Production of different ice phases under seemingly identical conditions is not a new observation but related to the non-equilibrium character of the transitions. In the case of the metastable ice phases IV, XII and stable ice V (see Figure 1) phase discrimination is achieved via the cooling rates . In our example the compression rate seems to be the decisive control parameter. Having used higher compression rates than in previous experiments it was possible to obtain exclusive formation of ice XII. Given the higher density of recovered ice XII ($`\rho 1.44`$ g/cm<sup>3</sup>) in comparison to recovered HDA ($`\rho 1.30`$ g/cm<sup>3</sup>) and an anticipated kinetic character of the transitions such a dependence could be expected. The observation of explosive sound accompanied by abrupt loss of pressure indicates the development of shock waves during the compression which could play a major role in the transformation process. The here established competition between crystallization and amorphization under close experimental conditions has to be properly acknowledged by all theoretical attempts trying to explain amorphous ice formation under pressure. Crystallization implies a reorganization and not merely a deformation of water’s hydrogen bond network and, therefore, places more stringent conditions on the transition mechanism. A purely mechanical instability as recently proposed for the formation of HDA is, to our opinion, insufficient to explain the ice XII formation. This holds unless the mechanical collapse is accompanied by high molecular mobility as for example in the case of a thermodynamic mechanical instability or shock wave melting which allow for crystalline reassembly. The requirement for high mobility, necessary for ice XII formation, equally explains the threshold value of around 1.0 GPa which corresponds to water’s extrapolated melting line. ###### Acknowledgements. Helpful discussions with W.F. Kuhs are greatfully acknowledged. This work is financially supported by the German Bundesministerium für Bildung und Forschung project No. 03–FU4DOR–5.
warning/0002/math0002074.html
ar5iv
text
# 1.1. Proposition Cohomologie $`\text{L}^\text{2}`$ sur les revêtements d’une variété complexe compacte Frédéric Campana Université de Nancy I, Département de Mathématiques Jean-Pierre Demailly Université de Grenoble I, Institut Fourier Version du 27 janvier 2000, imprimée le 15 mars 2024, 15:43 Introduction. Andreotti-Vesentini \[AV\], Ohsawa \[Oh\], Gromov \[G\], Kollár \[K\], entre autres, ont montré que la théorie de Hodge d’une variété kählérienne compacte pouvait être définie avec les mêmes propriétés dans le cadre $`L^2`$ si cette variété était seulement complète. Par ailleurs, les théorèmes d’annulation de la géométrie kählérienne ou projective reposant sur la méthode de Kodaira-Bochner-Nakano admettent par nature des versions $`L^2`$ (voir Androtti-Vesentini \[AV\] et \[D\]). On se propose ici de définir une cohomologie $`L^2`$ naturelle sur tout revêtement étale d’un espace analytique complexe $`X`$, à valeurs dans le relèvement de tout faisceau analytique cohérent $``$ sur $`X`$. Cette cohomologie a toutes les propriétés habituelles de la cohomologie des faisceaux sur $`X`$ (suites exactes de cohomologie, suites spectrales, théorèmes d’annulation, en particulier), et ces propriétés sont obtenues en incorporant l’information issue des estimées $`L^2`$ dans les preuves standards des résultats correspondants. La cohomologie $`L^2`$ devrait offrir un cadre agréable pour étudier la géométrie des revêtements, en fournissant un formalisme fonctoriel jouissant des propriétés attendues. Lorsque l’espace $`X`$ de base est compact et que le revêtement est galoisien de groupe $`\mathrm{\Gamma }`$, on peut définir la $`\mathrm{\Gamma }`$-dimension des groupes de cohomologie $`L^2`$ associés à un faisceau cohérent sur la base. On établit en particulier leur finitude et on étend le théorème de l’indice $`L^2`$ de Atiyah dans ce cadre. Enfin, si $`X`$ est projective, on a des théorèmes d’annulation $`L^2`$ qui étendent naturellement les théorèmes d’annulation usuels (théorème de Kodaira-Serre, théorème de Kawamata-Viehweg $`\mathrm{}`$). Dans \[E\], P. Eyssidieux a annoncé la construction d’une telle cohomologie, en utilisant des procédés voisins de ceux présentés ici. Notons aussi qu’un théorème d’annulation $`L^2`$ en cohomologie $`L^2`$ similaire à 4.1 est énoncé par J. Kollár dans \[K\], 11.4. § 1. Norme $`\text{L}^\text{2}`$ sur les sections. 1.0. Soit $`X`$ une variété analytique complexe, $``$ un faisceau analytique cohérent sur $`X`$, et $`U`$ un ouvert relativement compact de $`X`$. On dira que $`U`$ est $``$-admissible s’il existe un ouvert de Stein $`V`$ contenant $`U`$ et tel que $`U`$ soit relativement compact dans $`V`$, ainsi qu’un morphisme surjectif $`f:𝒪_V^r_{|V}`$ de faisceaux de $`𝒪_V`$-modules sur $`V`$. Un tel morphisme sera appelé une $`0`$-présentation de $``$. Si le fibré vectoriel trivial $`V\times C^r`$ est muni d’une métrique hermitienne $`h`$, on définira, pour $`sH^0(U,𝒪_U^r)`$ la norme $`s`$ par : $`s^2={\displaystyle _U}h(s,s)𝑑\mu `$, où $`\mu `$ est la forme volume d’une métrique fixée sur $`X`$. On notera $`H_{(2)}^0(U,𝒪^r)`$ l’espace vectoriel (de Hilbert) des $`s`$ tels que $`s<+\mathrm{}`$, et $`H_{(2)}^0(U,):=f_{}H_{(2)}^0(U,𝒪^r)H^0(U,)`$. On notera que cet espace est indépendant des choix $`(h,\mu `$, et même $`f`$ – voir ci-dessous) faits. On le munit de la norme $`L^2`$ quotient : pour $`\sigma =f_{}(s)H_{(2)}^0(U,)`$, on pose $$\sigma :=inf\{sf_{}(s)=:fs=\sigma ,sH_{(2)}^0(U,𝒪^r)\}.$$ ###### 1.1. Proposition Si $`\sigma =0`$, alors $`\sigma =0`$. Autrement dit, la semi-norme ainsi définie sur $`H_{(2)}^0(U,)`$ est une norme. De plus $`H_{(2)}^0(U,)`$ équipé de cette norme est un espace de Hilbert isométrique à l’orthogonal $`(\text{Ker}f_{})^{}`$ de $`\text{Ker}f_{}:H_{(2)}^0(U,𝒪^r)H_{(2)}^0(U,)`$ dans $`H_{(2)}^0(U,𝒪^r)`$, et $`(\text{Ker}f_{})`$ est fermé dans $`H_{(2)}^0(U,𝒪^r)`$. Démonstration. Il suffit de montrer que $`Kerf_{}`$ est fermé dans $`H_{(2)}^0(U,𝒪^r)`$. Or ceci résulte du fait que la topologie $`L^2`$ est plus forte que la topologie de la convergence uniforme sur les compacts de $`U`$ (\[W\], III.7), et du fait bien connu que le noyau $`Kerf_{}:H^0(U,𝒪^r)H^0(U,)`$ est fermé pour la topologie de la convergence uniforme sur les compacts (c’est le cas pour les sections à valeurs dans un sous-faisceau quelconque, \[H\], 6.3.5 et chap. 7). ###### 1.2. Définition Deux espaces de Hilbert $`(E,h_i)`$ $`(i=1,2)`$ sur le même espace sous-jacent $`E`$ sont dits équivalents si les normes $`h_1`$ et $`h_2`$ définissent la même topologie (ou encore : s’il existe $`0<A<B`$ tels que : $`Ah_1<h_2<Bh_1`$). ###### 1.3. Corollaire A équivalence près, l’espace de Hilbert $`(H_{(2)}^0(U,),)`$ est indépendant des choix $`(h;\mu ;f)`$ faits. Démonstration. Seule l’indépendance vis à vis de $`f`$ mérite d’être vérifiée : soient $`f_i:𝒪_V^{r_i}_{|V}0`$ $`(i=1,2)`$ deux $`0`$-présentations de $`_{|V}`$ sur un $`V`$ commun. Puisque $`V`$ est Stein, il existe $`\phi :𝒪_V^{r_i}𝒪_V^{r_j}`$, avec $`ij`$, tel que $`f_j\phi =f_i`$. Cette application fournit la continuité de l’application identique de $`H_{(2)}^0(U,)`$ muni de la norme déduite de $`f_i`$ dans lui-même muni de la norme déduite de $`f_j`$. 1.4. Soit $`U^{}U`$ ; l’application naturelle de restriction : $`\text{res}:H_{(2)}^0(U,)H_{(2)}^0(U^{},)`$ est continue, et compacte si $`U^{}U`$. L’affirmation est en effet claire dans le cas de faisceaux localement libres, et le cas général s’en déduit. 1.5. Soit $`u:𝒢`$ un morphisme de faisceaux, et $`U`$ un ouvert qui soit à la fois $``$-admissible et $`𝒢`$-admissible relativement à un même ouvert de Stein $`V`$. On a alors un morphisme induit $`u_{(2)}:H_{(2)}^0(U,)H_{(2)}^0(U,𝒢)`$ continu. En effet, soit $`f:𝒪_V^r_{|V}`$ une $`0`$-présentation de $`_{|V}`$ et $`g:𝒪_V^{r+s}𝒢_{|V}`$ une $`0`$-présentation de $`𝒢_{|V}`$ choisie en sorte que $`gi=uf`$, où $`i:𝒪_V^r𝒪_V^{r+s}`$ est l’injection des $`r`$-premières composantes. Alors le noyau du morphisme $`f_{}:H_{(2)}^0(U,𝒪_V^r)H^0(U,)`$ s’envoie par $`i`$ dans le noyau de $`g_{}:H_{(2)}^0(U,𝒪_V^{r+s})H^0(U,𝒢)`$, et on en déduit le morphisme $`u_{(2)}`$ voulu par passage au quotient. De plus $`u_{(2)}`$ est injectif si $`u`$ est injectif, et $`u_{(2)}`$ est surjectif si $`u`$ est surjectif. 1.6. Le foncteur $`H_{(2)}^0(U,)`$ n’est en général pas exact. Pour le voir, on peut considérer par exemple le morphisme injectif $`u:𝒪_{C^2}𝒪_{C^2}`$, $`sz_2s`$. Alors le morphisme induit $$u_{(2)}:H_{(2)}^0(U,𝒪)H_{(2)}^0(U,𝒪)$$ n’est pas d’image fermée sur la boule unité $`U=B(0,1)C^2`$ (on peut vérifier que la section $`(1z_1)^{3/2}`$ n’est pas dans $`L^2(U)`$, tandis que $`z_2(1z_1)^{3/2}`$ est dans $`L^2(U)`$, et par suite $`z_2(1z_1)^{3/2}`$ est seulement dans l’adhérence de l’image). Ceci montre qu’on ne peut pas avoir une suite exacte $$H_{(2)}^0(U,𝒪_{C^2})\stackrel{u_{\left(2\right)}}{}H_{(2)}^0(U,𝒪_{C^2})H_{(2)}^0(U,𝒪_{C\times \{0\}}).$$ 1.7. Le défaut d’exactitude du foncteur sections $`L^2`$ pourra être pallié par l’observation suivante: soit $$\stackrel{u}{}𝒢\stackrel{v}{}$$ une suite exacte de faisceaux admettant des $`0`$-présentations sur un ouvert de Stein $`V`$, et soient $`U^{}UV`$ des ouverts de Stein. Il existe une constante $`C>0`$ telle que pour tout élément $`g`$ dans le noyau de $`v_{(2)}:H_{(2)}^0(U,𝒢)H_{(2)}^0(U,)`$, on puisse trouver un élément $`fH_{(2)}^0(U^{},)`$ tel que $`u_{(2)}(f)=g_{|U^{}}`$ et $`f_{L^2(U^{})}Cg_{L^2(U)}`$. En effet, la topologie de $`H_{(2)}^0(U,𝒢)`$ est plus forte que la topologie de la convergence uniforme sur les compacts de $`U`$ (induite par passage au quotient à partir d’une présentation $`𝒪^N𝒢`$ et de la topologie d’espace de Fréchet sur $`H^0(U,𝒪^N)`$). On conclut à partir de la suite exacte d’espaces de Fréchet $$H^0(U,)H^0(U,𝒢)H^0(U,)$$ et du fait que le morphisme de restriction $`H^0(U,)H_{(2)}^0(U^{},)`$ est continu. 1.8. Remarque. Les notions d’espaces de sections $`L^2`$ peuvent également être définies de manière analogue pour un espace analytique $`X`$ arbitraire (réduit ou non), en plongeant localement l’ouvert de Stein $`VX`$ dans un espace ambiant lisse $`C^N`$, et en considérant l’extension triviale $`𝒢`$ du faisceau $`_{|V}`$ à $`C^N`$ (telle que $`𝒢_{|C^NV}=0`$). Les normes $`L^2`$ pour un ouvert $`UV`$ sont alors calculées en travaillant sur un ouvert de Stein $`U^{}C^N`$ tel que $`U^{}V=U`$. § 2. Image directe $`\text{L}^\text{2}`$. 2.0. Conventions. Soit $`X`$ un espace analytique complexe et $`p:\stackrel{~}{X}X`$ un revêtement étale de $`X`$. Soit $``$ un faisceau analytique cohérent sur $`X`$ et $`\stackrel{~}{}:=p^{}`$ son relèvement à $`\stackrel{~}{X}`$. Si $`UV`$ sont des ouverts de $`X`$ avec $`V`$ Stein et $`f:𝒪_V^r_V`$ une 0-résolution de $``$ sur $`V`$, on notera : $`\stackrel{~}{U}=:p^1(U);\stackrel{~}{V}:=p^1(V)`$ ; et $`\stackrel{~}{f}:𝒪_{\stackrel{~}{V}}^r\stackrel{~}{}_{\stackrel{~}{V}}`$ les relèvements correspondants à $`\stackrel{~}{X}`$. On dira que $`V`$ est $`p`$-simple si chaque composante connexe de $`\stackrel{~}{V}`$ est appliquée par $`p`$ sur une composante connexe de $`V`$, bijectivement. On supposera cette condition satisfaite. Soit $`h`$ une métrique hermitienne sur le fibré trivial $`V\times C^r`$ associé à $`𝒪_V^r`$, et $`\stackrel{~}{h}`$ son relèvement à $`\stackrel{~}{V}\times C^r`$, associé à $`𝒪_{\stackrel{~}{V}}^r`$. Ceci permet de définir la notion de norme $`L^2`$ pour $`\stackrel{~}{s}H^0(\stackrel{~}{U},\stackrel{~}{})`$, grâce à la définition de 1.0, et aussi $`H_{(2)}^0(\stackrel{~}{U},\stackrel{~}{})`$ qui, muni de cette norme, est un espace de Hilbert. De plus, les arguments du § 1 montrent que l’espace de Hilbert $`(H_{(2)}^0(\stackrel{~}{U},\stackrel{~}{}),)`$ est indépendant des choix $`(V,f,h,\mu )`$ faits, à équivalence près. ###### 2.1. Définition Soit $`W`$ un ouvert de $`X`$, et $`\stackrel{~}{W}=:p^1(W)`$. Soit $`\stackrel{~}{s}H^0(\stackrel{~}{W},\stackrel{~}{})`$. On dit que $`\stackrel{~}{s}`$ est localement $`L^2`$ sur $`X`$ si, pour chaque $`xW`$, il existe des voisinage ouverts $`UV`$ de $`x`$ dans $`X`$, avec $`V`$ Stein et $`p`$-simple, tels que la restriction de $`\stackrel{~}{s}`$ à $`\stackrel{~}{U}`$ soit dans $`H_{(2)}^0(\stackrel{~}{U},\stackrel{~}{})`$. 2.2. L’ensemble, noté : $`H_{(2),\mathrm{loc}}^0(\stackrel{~}{W},\stackrel{~}{})`$ des $`\stackrel{~}{s}`$ de $`H^0(\stackrel{~}{W},\stackrel{~}{})`$ qui sont localement $`L^2`$ sur $`X`$ forme clairement un espace vectoriel complexe. On a de plus, des applications naturelles de restriction res$`{}_{W,W^{}}{}^{}:H_{(2),\mathrm{loc}}^0(\stackrel{~}{W},\stackrel{~}{})H_{(2)}^0(\stackrel{~}{W}^{},\stackrel{~}{})`$ pour $`WW^{}`$, ouverts de $`X`$, et donc un préfaisceau à valeurs dans la catégorie des espaces vectoriels complexes : $$WH_{(2),\mathrm{loc}}^0(\stackrel{~}{W},\stackrel{~}{}).$$ Il est immédiat de vérifier que ce préfaisceau est un faisceau d’espaces vectoriels complexes sur $`X`$. ###### 2.3. Définition Le faisceau ainsi défini : $`WH_{(2),\mathrm{loc}}^0(\stackrel{~}{W},\stackrel{~}{})`$ sur $`X`$ est noté $`p_{(2)}\stackrel{~}{}`$ ; il est appelé le faisceau image directe $`L^2`$ de $`\stackrel{~}{}`$ par $`p`$. 2.4. On munit maintenant naturellement $`\left(p_{(2)}\stackrel{~}{}\right)`$ d’une structure de $`𝒪_X`$-module comme suit : si $`\stackrel{~}{s}H_{(2),\mathrm{loc}}^0(\stackrel{~}{W},\stackrel{~}{})`$ représente un germe de section de $`\left(p_{(2)}\stackrel{~}{}\right)`$ en $`xW`$, et si $`\phi H^0(W,𝒪_W)`$, alors $`(p^{}\phi \stackrel{~}{s})H_{(2),\mathrm{loc}}^0(\stackrel{~}{W},\stackrel{~}{})`$, qui est donc un $`H^0(W,𝒪_W)`$-module. 2.5. Remarque. Le faisceau $`p_{(2)}(\stackrel{~}{})`$ n’est en général cohérent que si $`p`$ est fini, auquel cas il se réduit à l’image directe $`p_{}\stackrel{~}{}`$ usuelle. Si $`p`$ est infini et $`0`$, alors $`p_{(2)}(\stackrel{~}{})`$ n’est jamais cohérent. ###### 2.6. Proposition Soit $`𝒢`$ une suite exacte de faisceaux cohérents analytiques sur $`X`$ ; alors la suite naturelle d’images directes $`L^2:p_{(2)}\stackrel{~}{}p_{(2)}\stackrel{~}{𝒢}p_{(2)}\stackrel{~}{}`$ est exacte. $`(`$Autrement dit : le foncteur d’image directe $`L^2`$ par $`p`$ est exact$`)`$. Démonstration. Soit $`V`$ un ouvert $`p`$-simple sur lequel $``$, $`𝒢`$ et $``$ admettent des $`0`$-présentations. et $`U^{}UV`$ des ouverts de Stein connexes comme dans 1.7. Toutes les composantes connexes $`U_j^{}U_jV_j`$ de $`\stackrel{~}{U}^{}`$, $`\stackrel{~}{U}`$, $`\stackrel{~}{V}`$ sont alors en isomorphisme avec $`U^{}UV`$. Il existe par conséquent une constante $`C`$ indépendante de $`j`$ telles que les sections $`g_j`$ du noyau de $`H_{(2)}^0(U_j,\stackrel{~}{𝒢})H_{(2)}^0(U_j,\stackrel{~}{})`$ se relèvent en des sections $`f_j`$ de $`H_{(2)}^0(U_j^{},)`$, avec $`f_j_{L^2(U_j^{})}Cg_j_{L^2(U_j)}`$. Toute section $`g=g_j`$ dans le noyau de $`H_{(2)}^0(\stackrel{~}{U},\stackrel{~}{𝒢})H_{(2)}^0(\stackrel{~}{U},\stackrel{~}{})`$ se relève donc en une section $`f=f_jH_{(2)}^0(\stackrel{~}{U}^{},\stackrel{~}{})`$ dans $`L^2(\stackrel{~}{U}^{})`$. ###### 2.7. Corollaire On a un isomorphisme naturel de faisceaux de $`𝒪_X`$-modules :$`\eta :p_{(2)}\stackrel{~}{}\stackrel{}{}\left(p_{(2)}𝒪_{\stackrel{~}{X}}\right)_{𝒪_X}`$. Démonstration. C’est immédiat à partir des définitions lorsque $``$ est localement libre. En général, on conclut à partir de l’exactitude du foncteur image directe $`L^2`$, par récurrence sur la longueur d’une résolution libre (locale) de $``$ sur $`X`$. 2.8. Exemple : Supposons que $``$ admette une résolution finie localement libre (si $`X`$ est projective, c’est toujours le cas, par un résultat de J.-P. Serre) : $`0_n_{n1}\mathrm{}_00`$. Alors : $`p_{(2)}\stackrel{~}{}`$ admet une résolution finie de même longueur par des faisceaux localement de la forme $`\left(p_{(2)}𝒪_{\stackrel{~}{X}}\right)^r`$ : $$0p_{(2)}\stackrel{~}{}_n\mathrm{}p_{(2)}\stackrel{~}{}_0p_{(2)}0.$$ Lorsque $`X`$ est projective, les $`_k`$ peuvent être pris de la forme : $$=\underset{j=1}{\overset{r}{}}𝒪_X(a_j),a_jZ$$ et on a donc : $$p_{(2)}=\underset{j=1}{\overset{r}{}}p_{(2)}𝒪_{\stackrel{~}{X}}_{𝒪_X}𝒪_X(a_j).$$ ###### 2.9. Proposition Soit $`Y`$ un sous-espace de $`X`$, $`^{}`$ un faisceau analytique cohérent sur $`Y`$, et $`=i_{}^{}`$ son extension par $`0`$ sur $`X𝑟Y`$ (image directe par l’injection $`i:YX`$). Alors, si $`p^{}:\stackrel{~}{Y}:=p^1(Y)Y`$ est le revêtement étale de $`Y`$ induit par $`p:\stackrel{~}{X}X`$, on a $`p_{(2)}\stackrel{~}{}=\stackrel{~}{i}_{}(p_{(2)}^{}\stackrel{~}{}^{})`$$`\stackrel{~}{i}:\stackrel{~}{Y}\stackrel{~}{X}`$ est l’injection naturelle. Démonstration. On observe d’abord que si $`V`$ est un ouvert de Stein simple dans $`X`$ et si $`UV`$ est un voisinage dans $`X`$ d’un point $`xY`$ (resp. $`U^{}YU`$ un voisinage de $`x`$ dans $`Y`$), on a un morphisme de restriction continu $`H_{(2)}^0(U,)H_{(2)}^0(U^{},^{})`$. En considérant les sections sur les revêtements $`\stackrel{~}{U}`$ et $`\stackrel{~}{U}^{}`$ puis en passant à la limite inductive sur $`U`$ et $`U^{}`$, on en déduit qu’on a un morphisme de restriction $`p_{(2)}\stackrel{~}{i}_{}(p_{(2)}^{}^{})`$ continu. Il s’agit en fait d’un isomorphisme. En effet, prenons des voisinages ouverts de Stein $`U`$, $`U_1`$ de $`x`$ dans $`X`$ (resp. $`U^{}`$ de $`x`$ dans $`Y`$), tels que $`UU_1V`$ et $`YU_1U^{}`$. On a un homomorphisme surjectif d’espaces de Fréchet $`H^0(U_1,)H^0(YU_1,^{})`$, qui est par suite un morphisme ouvert. Pour toute section $`s^{}`$ de $`^{}`$ sur $`U^{}`$, on peut alors trouver une section $`s`$ de $``$ sur $`U_1`$ telle que $`s_{|YU_1}=s_{|YU_1}^{}`$ et $`s_{L^{\mathrm{}}(U)}Cs^{}_{L^{\mathrm{}}(U_1^{})}`$ pour un certain $`U_1^{}YU_1`$. Au niveau des normes $`L^2`$, ceci implique $`s_{L^2(U)}Cs^{}_{L^2(U^{})}`$. On conclut en passant aux sections sur les revêtements $`\stackrel{~}{U}`$ et $`\stackrel{~}{U}^{}`$. §3. Cohomologie $`\text{L}^\text{2}`$. 3.0. Les conventions sont celles de 2.0. ###### 3.1. Définition Soit $`W`$ un ouvert de $`X`$. On définit la cohomologie $`L^2`$ de $`\stackrel{~}{W}`$ à valeurs dans $`\stackrel{~}{}`$ comme étant celle du faisceau $`\left(p_{(2)}\stackrel{~}{}\right)`$ sur $`W`$. On note alors $`H_{(2)}^q(\stackrel{~}{W},\stackrel{~}{})`$ $`(q0)`$ le $`q`$-ième groupe de cohomologie ainsi défini. ###### 3.2. Théorème Soit $`0𝒢0`$ une suite exacte de faisceaux analytiques cohérents sur $`X`$. On peut lui associer une suite exacte longue de cohomologie $`L^2`$, fonctorielle en $``$ $$0H_{(2)}^0(\stackrel{~}{X},\stackrel{~}{})\mathrm{}H_{(2)}^q(\stackrel{~}{X},\stackrel{~}{})H_{(2)}^q(\stackrel{~}{X},\stackrel{~}{𝒢})H_{(2)}^q(\stackrel{~}{X},\stackrel{~}{})H_{(2)}^{q+1}(\stackrel{~}{X},\stackrel{~}{})\mathrm{}$$ Démonstration. Cette suite exacte résulte immédiatement de l’exactitude du foncteur $`p_{(2)}`$ et des propriétés usuelles de la cohomologie. ###### 3.3. Proposition Soit un diagramme commutatif $$\begin{array}{ccc}\stackrel{~}{Y}& \stackrel{\stackrel{~}{f}}{}& \stackrel{~}{X}\\ \\ p^{}& & p\\ Y& \stackrel{f}{}& X\end{array}$$ où les flèches verticales sont des revêtements, $`f:YX`$ un morphisme analytique et $`p^{}:\stackrel{~}{Y}Y`$ le revêtement image inverse de $`p`$ par $`f`$. Soit $``$ un faisceau analytique cohérent sur $`Y`$. Alors pour tout $`q0`$ on a la formule de commutation $$p_{(2)}(R^qf_{})^\stackrel{~}{}=R^q\stackrel{~}{f}_{}(p_{(2)}^{}\stackrel{~}{}).$$ Démonstration. Les deux faisceaux en question sont les faisceaux associés au préfaisceau $$UH^q(f^1(U),p_{(2)}^{}\stackrel{~}{}),UH_{(2)}^q(\stackrel{~}{f}^1(\stackrel{~}{U}),\stackrel{~}{}),$$ et ces deux préfaisceaux coïncident par définition de la cohomologie $`L^2`$. 3.4. Isomorphisme de Dolbeault. On suppose que $`X`$ est une variété lisse, que $``$ est localement libre, et on note $`F`$ le fibré vectoriel (supposé muni d’une métrique hermitienne) associé sur $`X`$. Soit $`\stackrel{~}{}^{r,q}`$ le faisceau des formes différentielles $`v`$ de type $`(r,q)`$ à valeurs dans $`\stackrel{~}{F}`$ et à coefficients $`L_{\mathrm{loc}}^2`$ sur $`\stackrel{~}{X}`$, telles que $`\overline{}v`$ soient aussi $`L_{\mathrm{loc}}^2`$. Soit $`p_{(2)}\stackrel{~}{}^{r,q}`$ le (pré)faisceau sur $`X`$ défini par $`UH_{(2),\mathrm{loc}}^0(\stackrel{~}{U},\stackrel{~}{}^{r,q})`$, image directe $`L^2`$ de $`\stackrel{~}{}^{r,q}`$ sur $`X`$, à savoir le faisceau des formes différentielles qui sont $`L^2`$ localement au dessus de $`X`$, sur les ouverts de la forme $`\stackrel{~}{U}=p^1(U)`$. L’opérateur $`\overline{}`$ fournit un complexe de faisceaux sur $`X`$ $$0p_{(2)}\stackrel{~}{}p_{(2)}\stackrel{~}{}^{0,0}\overline{}\mathrm{}\overline{}p_{(2)}\stackrel{~}{}^{0,q}p_{(2)}\stackrel{~}{}^{0,q+1}\mathrm{}p_{(2)}\stackrel{~}{}^{0,n}0.$$ Ce complexe est exact, en vertu du théorème d’existence de Hörmander-Andreotti-Vesentini pour les solutions $`L^2`$ de l’opérateur $`\overline{}`$: en effet, on va appliquer ce théorème sur des ouverts $`\stackrel{~}{U}`$ qui revêtent un ouvert de Stein $`UX`$ quelconque, en prenant une métrique kählérienne $`\omega `$ sur $`X`$ et des poids de la forme $`\stackrel{~}{\phi }=\phi p`$, où $`\phi `$ est choisi en sorte que $`i\overline{}\phi +\mathrm{Ricci}_\omega +\mathrm{Courbure}_F\omega `$. Pour toute forme $`w`$ telle que $`\overline{}w=0`$, on voit alors que l’équation $`\overline{}v=w`$ admet une solution $`v`$ telle que $`vw`$ en norme $`L^2`$. On obtient ainsi une résolution fine de $`p_{(2)}\stackrel{~}{}`$ (par des faisceaux de modules sur le faisceau d’anneaux des fonctions $`C^{\mathrm{}}`$ sur $`X`$), de sorte que le complexe de Dolbeault $`L^2`$ calcule bien la cohomologie $`L^2`$ définie en 3.1. Nous pouvons énoncer: ###### 3.5. Proposition (Isomorphisme de Dolbeault) Soit $`X`$ une variété analytique complexe, $``$ un $`𝒪_X`$-module localement libre sur $`X`$, et $`(L_{(2),\mathrm{loc}}^{0,q}(\stackrel{~}{X},\stackrel{~}{}),\overline{})`$ le complexe des $`(0,q)`$-formes sur $`\stackrel{~}{X}`$, à valeurs dans $`\stackrel{~}{}`$ et localement $`L^2`$ sur $`X`$ ainsi que leur $`\overline{}`$. Alors la cohomologie de ce complexe s’identifie à $`H_{(2)}^{}(\stackrel{~}{X},\stackrel{~}{})`$. Comme dans la situation classique, l’isomorphisme de Dolbeault fournit un moyen commode pour prouver des théorèmes d’annulations. ###### 3.6. Théorème Soit $``$ un faisceau cohérent sur un espace complexe $`X`$. a) Si $`U`$ est un ouvert de Stein, alors $`H_{(2)}^q(\stackrel{~}{U},\stackrel{~}{})=H^q(U,p_{(2)}\stackrel{~}{})=0`$ pour $`q>0`$. b) Soit $`WU`$ une paire de Runge d’ouverts de Stein, alors le morphisme de restriction $`H_{(2)}^0(\stackrel{~}{U},\stackrel{~}{})H_{(2)}^0(\stackrel{~}{W},\stackrel{~}{})`$ est d’image dense pour la topologie de la convergence $`L^2`$ au dessus des compacts de $`W`$. Démonstration. Rappelons qu’une paire d’ouverts de Stein $`WU`$ est dite de Runge si l’enveloppe holomorphe convexe de toute partie compacte de $`W`$ relativement à l’algèbre des fonctions holomorphes sur $`U`$ est compacte dans $`W`$. On sait alors que l’image du morphisme de restriction $`𝒪(U)𝒪(W)`$ est dense, et que pour tout compact $`K`$ de $`W`$ il existe une fonction d’exhaustion strictement plurisousharmonique $`\phi _K`$ sur $`X`$ telle que l’ensemble de niveau $`X_c=\{\phi _K<c\}`$ vérifie $`KX_cW`$. Le théorème 3.5 se prouve en 3 étapes. Étape 1. $`X`$ est lisse et $``$ est localement libre. Dans ce cas, il suffit d’utiliser l’isomorphisme de Dolbeault 3.4, et d’appliquer le théorème de Hörmander-Andreotti-Vesentini (\[H\], \[AV\]) avec des poids plurisousharmoniques $`\phi `$ sur $`U`$ à croissance arbitrairement grande lorsqu’on s’approche du bord de $`U`$, de manière à faire converger les normes $`L^2`$. L’assertion sur les paires de Runge se démontre comme les théorèmes 4.3.2 et 5.2.10 de Hörmander \[H\], en utilisant des poids de la forme $`e^{N\phi _K}`$, $`N0`$, pour assurer la convergence uniforme des approximations au voisinage de $`K`$. Bien entendu, cette étape permet aussi de couvrir le cas où $`U`$ est un ouvert de Stein dans un espace $`X`$ quelconque, il suffit de plonger $`V`$ dans un espace de Stein ambiant lisse et de prolonger le faisceau $``$ par $`0`$ en dehors de $`X`$. Étape 2. L’ouvert $`U`$ est contenu dans un ouvert de Stein simple $`V`$ sur lequel $``$ admet une résolution libre. Soit $$0_n_{n1}_n\mathrm{}_0$$ une résolution libre de $``$. On raisonne par récurrence sur la longueur $`n`$ de la résolution. Si $`n=0`$, alors $``$ est libre et on applique l’étape 1. En général, soit $`𝒢`$ le noyau de $`_0`$. Alors $`𝒢`$ admet une résolution libre de longueur $`n1`$, et par hypothèse de récurrence on a $`H_{(2)}^q(\stackrel{~}{U},\stackrel{~}{𝒢})=0`$ pour $`q>0`$. La suite exacte $$0𝒢_00$$ fournit une suite exacte longue de cohomologie $$0=H_{(2)}^q(\stackrel{~}{U},\stackrel{~}{}_0)H_{(2)}^q(\stackrel{~}{U},\stackrel{~}{})H_{(2)}^{q+1}(\stackrel{~}{U},\stackrel{~}{𝒢})=0,q>0,$$ ce qui conclut la récurrence. Le fait que le morphisme $`H_{(2)}^0(\stackrel{~}{W},\stackrel{~}{}_0)H_{(2)}^0(\stackrel{~}{W},\stackrel{~}{})`$ soit surjectif ramène l’assertion sur les paires de Runge au cas d’un faisceau localement libre. Étape 3. Cas général. On utilise la classique “méthode des bosses” d’Andreotti-Grauert. Pour cela, on choisit un recouvrement localement fini $`U=_{jN}U_j`$ assez fin de $`U`$, par des ouverts $`U_j`$ ayant les propriétés suivantes: $``$$`U_j`$ est un ouvert de Stein relativement compact dans $`U`$, et $`(U_j,U)`$ est une paire de Runge ; $``$$`V_j=U_0U_1\mathrm{}U_j`$ est un ouvert de Stein et $`(V_j,U)`$ est une paire de Runge. On choisit le recouvrement $`(U_j)`$ assez fin pour que chaque $`U_j`$ soit contenu dans un ouvert de Stein simple sur lequel $``$ admet une résolution libre. On démontre maintenant par récurrence sur $`j`$ que $`(\text{a}_j)`$$`H_{(2)}^q(\stackrel{~}{V}_j,\stackrel{~}{})=0`$ pour tout $`q>0`$, $`(\text{b}_j)`$Si $`WV_j`$ est une paire de Runge, alors la restriction $`H_{(2)}^0(\stackrel{~}{V}_j,\stackrel{~}{})H_{(2)}^0(\stackrel{~}{W},\stackrel{~}{})`$ est d’image dense. Pour $`j=0`$ on a $`V_0=U_0`$ et $`(\text{a}_0)`$, $`(\text{b}_0)`$ résultent de l’étape 2. Pour passer de l’étape $`j`$ à l’étape $`j+1`$, on utilise la suite exacte $$\begin{array}{cc}\hfill \mathrm{}H_{(2)}^{q1}(\stackrel{~}{V}_j,\stackrel{~}{})& H_{(2)}^{q1}(\stackrel{~}{U}_{j+1},\stackrel{~}{})H_{(2)}^{q1}(\stackrel{~}{V}_j\stackrel{~}{U}_{j+1},\stackrel{~}{})\hfill \\ \hfill H_{(2)}^q(\stackrel{~}{V}_{j+1},\stackrel{~}{})H_{(2)}^q(\stackrel{~}{V}_j,\stackrel{~}{})& H_{(2)}^q(\stackrel{~}{U}_{j+1},\stackrel{~}{})H_{(2)}^q(\stackrel{~}{V}_j\stackrel{~}{U}_{j+1},\stackrel{~}{})\mathrm{}\hfill \end{array}$$ qui résulte de l’application de la suite exacte de Mayer-Vietoris au faisceau $`(p_2)_{}\stackrel{~}{}`$. Pour $`q2`$, l’étape 2 et l’hypothèse de récurrence entraînent $$H_{(2)}^{q1}(\stackrel{~}{V}_j\stackrel{~}{U}_{j+1},\stackrel{~}{})=H_{(2)}^q(\stackrel{~}{U}_{j+1},\stackrel{~}{})=0,\text{resp.}H_{(2)}^q(\stackrel{~}{V}_j,\stackrel{~}{})=0,$$ d’où $`H_{(2)}^q(\stackrel{~}{V}_{j+1},\stackrel{~}{})=0`$. Si $`q=1`$, on utilise de plus le fait que la restriction $`H_{(2)}^0(\stackrel{~}{V}_j,\stackrel{~}{})H_{(2)}^0(\stackrel{~}{V}_j\stackrel{~}{U}_{j+1},\stackrel{~}{})`$ est d’image dense pour voir que le morphisme continu $$H_{(2)}^0(\stackrel{~}{V}_j\stackrel{~}{U}_{j+1},\stackrel{~}{})H_{(2)}^1(\stackrel{~}{V}_{j+1},\stackrel{~}{})$$ est nécessairement nul. Ceci implique alors $`H_{(2)}^1(\stackrel{~}{V}_{j+1},\stackrel{~}{})=0`$ et l’assertion $`(\text{a}_{j+1})`$ est démontrée. L’assertion $`(\text{b}_{j+1})`$, quant à elle, s’obtient comme suit. Soit $`WV_{j+1}`$ une paire de Runge. Alors $`WV_jV_j`$ et $`WU_{j+1}U_{j+1}`$ sont des paires de Runge pour lesquelles on peut appliquer l’hypothèse de récurrence $`(\text{b}_j)`$, resp. l’étape 2. Si $`s`$ est une section de $`H_{(2)}^0(\stackrel{~}{W},\stackrel{~}{})`$, on peut approximer $`s`$ en topologie $`L^2`$ au dessus de tout compact de $`WV_j`$, resp. de $`WU_{j+1}`$, par des sections $`s_jH_{(2)}^0(\stackrel{~}{V}_j,\stackrel{~}{})`$, resp. $`t_{j+1}H_{(2)}^0(\stackrel{~}{U}_{j+1},\stackrel{~}{})`$. La différence $`s_jt_{j+1}`$ définit un $`1`$-cocycle de Čech sur $`V_{j+1}`$ relativement au recouvrement $`(V_j,U_{j+1})`$. Comme $`H_{(2)}^1(\stackrel{~}{V}_{j+1},\stackrel{~}{})=0`$, on a un morphisme surjectif d’espaces de Fréchet $$H_{(2)}^0(\stackrel{~}{V}_j,\stackrel{~}{})H_{(2)}^0(\stackrel{~}{U}_{j+1},\stackrel{~}{})H_{(2)}^0(\stackrel{~}{V}_{j+1},\stackrel{~}{}).$$ Or, si $`s_j`$ et $`t_{j+1}`$ sont des approximations suffisamment bonnes de $`s`$, la différence $`s_jt_{j+1}`$ peut être choisie arbitrairement petite dans la topologie de l’espace de Fréchet but. D’après le théorème de l’application ouverte, on peut trouver des sections $`\sigma _jH_{(2)}^0(\stackrel{~}{V}_j,\stackrel{~}{})`$ et $`\tau _{j+1}H_{(2)}^0(\stackrel{~}{U}_{j+1},\stackrel{~}{})`$ arbitrairement petites telles que $`\sigma _j\tau _{j+1}=s_jt_{j+1}`$ sur $`\stackrel{~}{V}_j\stackrel{~}{U}_{j+1}`$. Alors $`s_j\sigma _j`$ et $`t_{j+1}\tau _{j+1}`$ se recollent en une section sur $`V_{j+1}=V_jU_{j+1}`$ qui approxime $`s`$ d’aussi près qu’on veut sur $`W`$. Un raisonnement standard de passage à la limite à la Mittag-Leffler permet d’atteindre la nullité de la cohomologie sur $`U=V_j`$ et le théorème de Runge pour la paire $`WU`$ à partir du théorème de Runge sur les paires $`WV_jV_{j+1}`$, $`V_{j+1}V_{j+2}`$, $`\mathrm{}`$, etc. ###### 3.7. Corollaire Soit $`𝒰:=(U_\lambda )_{\lambda \mathrm{\Lambda }}`$ un recouvrement ouvert localement fini de $`X`$ par des ouverts de Stein $`U_\lambda `$. Ce recouvrement est alors de Leray pour $`p_{(2)}`$ et on a un isomorphisme naturel : $$H_{(2)}^{}(\stackrel{~}{𝒰},\stackrel{~}{})\stackrel{}{}H_{(2)}^{}(\stackrel{~}{X},\stackrel{~}{}),$$ $`H_{(2)}^{}(\stackrel{~}{𝒰},\stackrel{~}{}):=H^{}(𝒰,p_{(2)}\stackrel{~}{})`$ est la cohomologie de Čech de $`p_{(2)}\stackrel{~}{}`$ relative au recouvrement $`𝒰`$ de $`X`$. 3.8. Explicitons l’assertion de 3.7 : soit $`N_q(𝒰)`$ le $`q`$-nerf du recouvrement $`𝒰`$, constitué des intersections non vides de $`(q+1)`$ des éléments de $`𝒰`$, et soit $`C_{(2),\mathrm{loc}}^q(\stackrel{~}{𝒰},\stackrel{~}{})`$ le groupe des $`q`$-cochaînes à valeurs dans $`\stackrel{~}{}`$, définies sur les ouverts $`\stackrel{~}{𝒰}_{(q)}=p^1(𝒰_{(q)})`$ et localement $`L^2`$ au dessus de $`𝒰_{(q)}N_q(𝒰)`$, avec les applications de cobord $$\delta _q:C_{(2),\mathrm{loc}}^q(\stackrel{~}{𝒰},\stackrel{~}{})C_{(2),\mathrm{loc}}^{q+1}(\stackrel{~}{𝒰},\stackrel{~}{})$$ usuelles. Alors $`H_{(2)}^{}(\stackrel{~}{𝒰},\stackrel{~}{})`$ est la cohomologie du complexe ainsi défini. Les espaces $`C_{(2),\mathrm{loc}}^q(\stackrel{~}{𝒰},\stackrel{~}{})`$ sont naturellement munis de la topologie de la convergence en norme $`L^2`$ au dessus des compacts $`p`$-simples contenus dans les $`𝒰_{(q)}`$ (on ne prend bien entendu en compte simultanément qu’un nombre fini de ces intersections), ce qui en fait des espaces de Fréchet. On munit $`H_{(2)}^{}(\stackrel{~}{𝒰},\stackrel{~}{})`$ de la topologie quotient correspondante (qui n’est pas nécessairement séparée). 3.9. Désignons par $`B_{(2)}^q(\stackrel{~}{𝒰},\stackrel{~}{})`$ et $`Z_{(2)}^q(\stackrel{~}{𝒰},\stackrel{~}{})`$ respectivement l’image de $`\delta _{q1}`$ et le noyau de $`\delta _q`$ $`(`$avec $`B_{(2)}^0(\stackrel{~}{𝒰},\stackrel{~}{})=0)`$. On a alors une suite exacte $$0\underset{¯}{H}_{(2)}^q(\stackrel{~}{𝒰},\stackrel{~}{})H_{(2)}^q(\stackrel{~}{𝒰},\stackrel{~}{})\overline{H}_{(2)}^q(\stackrel{~}{𝒰},\stackrel{~}{})0,$$ $`\overline{B}_{(2)}^q(\stackrel{~}{𝒰},\stackrel{~}{})`$ est l’adhérence dans $`Z_{(2)}^q(\stackrel{~}{𝒰},\stackrel{~}{})`$ de $`B_{(2)}^q(\stackrel{~}{𝒰},\stackrel{~}{})`$ et $$\underset{¯}{H}_{(2)}^q(\stackrel{~}{𝒰},\stackrel{~}{})=\overline{B}_{(2)}^q(\stackrel{~}{𝒰},\stackrel{~}{})/B_{(2)}^q(\stackrel{~}{𝒰},\stackrel{~}{}),\overline{H}_{(2)}^q(\stackrel{~}{𝒰},\stackrel{~}{})=Z_{(2)}^q(\stackrel{~}{𝒰},\stackrel{~}{})/\overline{B}_{(2)}^q(\stackrel{~}{𝒰},\stackrel{~}{}).$$ L’espace $`\overline{H}_{(2)}^q(\stackrel{~}{𝒰},\stackrel{~}{})`$ est par définition un espace de Fréchet, mais $`\underset{¯}{H}_{(2)}^q(\stackrel{~}{𝒰},\stackrel{~}{})`$ est muni de la topologie grossière et, s’il est non nul, la cohomologie $`L^2`$ n’est pas séparée. On va voir, cependant, que la topologie de $`H_{(2)}^q(\stackrel{~}{𝒰},\stackrel{~}{})`$ est essentiellement indépendante du choix du recouvrement. Soient en effet $`𝒰^{}`$, $`𝒰`$ des recouvrements de Stein de $`X`$. On suppose $`𝒰^{}`$ plus fin que $`𝒰`$ et muni d’une application de raffinement $`\rho `$ vers $`𝒰`$. Alors, dans le diagramme commutatif associé $`(q0)`$, $$\begin{array}{ccccccccc}0& & \underset{¯}{H}_{(2)}^q(\stackrel{~}{𝒰},\stackrel{~}{})& & H_{(2)}^q(\stackrel{~}{𝒰},\stackrel{~}{})& & \overline{H}_{(2)}^q(\stackrel{~}{𝒰},\stackrel{~}{})& & 0\\ \\ & & \underset{¯}{\rho }& & \rho & & \overline{\rho }\\ \\ 0& & \underset{¯}{H}_{(2)}^q(\stackrel{~}{𝒰}^{},\stackrel{~}{})& & H_{(2)}^q(\stackrel{~}{𝒰}^{},\stackrel{~}{})& & \overline{H}_{(2)}^q(\stackrel{~}{𝒰}^{},\stackrel{~}{})& & 0\end{array}$$ les applications verticales $`\rho `$, $`\underset{¯}{\rho }`$, $`\overline{\rho }`$ sont des isomorphismes topologiques (d’espaces de Fréchet en ce qui concerne $`\overline{\rho }`$). En effet, si (pour simplifier) on désigne par $`(C,\delta )`$ et $`(C^{},\delta ^{})`$ les complexes de Fréchet impliqués, l’isomorphisme de Leray implique que $`\rho `$ est un isomorphisme algébrique, et il est clair par ailleurs que l’application de restriction $`\rho :CC^{}`$ est continue. On a alors une application surjective $`\delta ^{}r:C^{}ZZ^{}`$ entre espaces de Fréchet. Le théorème de l’application ouverte montre que cette application est ouverte, et il en est donc de même pour l’application induite $`\rho :H=Z/\delta CH^{}=Z^{}/\delta ^{}C^{}`$. Ceci montre déjà que $`\rho `$ est un isomorphisme topologique. L’assertion pour $`\underset{¯}{\rho }`$ et $`\overline{\rho }`$ s’en déduit immédiatement, puisque $`\underset{¯}{H}`$ est la partie grossière et $`\overline{H}`$ le quotient séparé de la cohomologie. Ceci nous mène à la définition suivante. ###### 3.10. Définition L’espace de cohomologie $`H_{(2)}^q(\stackrel{~}{X},\stackrel{~}{})`$ $`(q0)`$ est muni d’une topologie naturelle pour laquelle, si $`\underset{¯}{H}_{(2)}^q(\stackrel{~}{X},\stackrel{~}{})`$ est l’adhérence de $`0`$, alors $$\overline{H}_{(2)}^q(\stackrel{~}{X},\stackrel{~}{})=H_{(2)}^q(\stackrel{~}{X},\stackrel{~}{})/\underset{¯}{H}_{(2)}^q(\stackrel{~}{X},\stackrel{~}{})$$ est un espace de Fréchet. On appellera $`\overline{H}_{(2)}^q(\stackrel{~}{X},\stackrel{~}{})`$ $`(`$resp. $`\underset{¯}{H}_{(2)}^q(\stackrel{~}{X},\stackrel{~}{}))`$ la cohomologie $`L^2`$ séparée de $`X`$ à valeurs dans $`\stackrel{~}{}`$ $`(`$resp. le noyau de la cohomologie $`L^2`$ de $`X`$ à valeurs dans $`\stackrel{~}{})`$. 3.11. Remarque. Par construction, les objects $`\underset{¯}{H}_{(2)}^q(\stackrel{~}{X},\stackrel{~}{})`$ et $`\overline{H}_{(2)}^q(\stackrel{~}{X},\stackrel{~}{})`$ sont fonctoriels en $``$ et $`X`$. 3.12. Remarque. Il résulte immédiatement de 2.9 que si $``$ est supporté par la sous-variété $`Y`$ de $`X`$, alors : $`H_{(2)}^{}(\stackrel{~}{X},\stackrel{~}{})H_{(2)}^{}(\stackrel{~}{Y},\stackrel{~}{})`$ est un isomorphisme topologique. 3.13. Remarque. Si $`X`$ est un espace compact, on peut choisir un recouvrement ouvert fini $`𝒰=(U_j)`$ fini par des ouverts de Stein $`p`$-simples, puis rétrécir un peu chacun des ouverts $`U_j`$ en des ouverts $`U_j^{\prime \prime }U_j^{}U_j`$ tels que $`U_j^{}`$ et $`U_j^{\prime \prime }`$ soient de Runge dans $`U_j`$. Les morphismes de restriction donnent lieu à des flèches $$C_{(2),\mathrm{loc}}^q(\stackrel{~}{𝒰},\stackrel{~}{})C_{(2)}^q(\stackrel{~}{𝒰^{}},\stackrel{~}{})C_{(2),\mathrm{loc}}^q(\stackrel{~}{𝒰^{\prime \prime }},\stackrel{~}{})$$ où les termes extrêmes sont des espaces de Fréchet et le terme central un espace de Hilbert (on prend sur ce terme la topologie $`L^2`$ globale sur $`𝒰^{}`$). En cohomologie, on un isomorphisme entre les termes extrêmes, ce qui prouve que la cohomologie du terme central se surjecte sur cette cohomologie. La cohomologie séparée $`\overline{H}_{(2)}^q(\stackrel{~}{X},\stackrel{~}{})`$ possède donc alors une topologie d’espace de Hilbert. 3.14. Remarque. Supposons maintenant que $`X`$ soit une variété compacte lisse et que $``$ soit un faisceau analytique localement libre sur $`X`$. Les arguments de \[G\] s’appliquent encore dans ce contexte et montrent que si $`Z_{(2)}^{0,q}(\stackrel{~}{X},\stackrel{~}{})C_{(2)}^{0,q}(\stackrel{~}{X},\stackrel{~}{})`$ est le noyau du $`\overline{}`$, alors : $`\left(Z_{(2)}^{0,q}(\stackrel{~}{X},\stackrel{~}{})/(\overline{\text{Im}\overline{}})\right)`$ s’identifie à l’espace de Hilbert (par ellipticité de $`\overline{}`$) des formes $`\mathrm{\Delta }_\overline{}`$-harmoniques de type $`(0,q)`$ et $`L^2`$ sur $`\stackrel{~}{X}`$ à valeurs dans $`\stackrel{~}{}`$. On voit, de plus, que $`_{(2)}^{0,q}(\stackrel{~}{X},\stackrel{~}{})`$ s’identifie canoniquement à la cohomologie réduite $`\overline{H}_{(2)}^q(\stackrel{~}{X},\stackrel{~}{})`$ définie en 3.10. ###### 3.15. Corollaire (dualité de Serre) Soit $`X`$ une variété complexe compacte lisse, et $``$ un faisceau analytique cohérent localement libre sur $`X`$. Soit $`p:\stackrel{~}{X}X`$ un revêtement étale. Il existe une isométrie antilinéaire $`\sigma :\overline{H}_{(2)}^q(\stackrel{~}{X},\mathrm{\Omega }_{\stackrel{~}{X}}^r\stackrel{~}{})\overline{H}_{(2)}^{nq}(\stackrel{~}{X},\mathrm{\Omega }_{\stackrel{~}{X}}^{nr}\stackrel{~}{}^{})`$ si $`n`$ est la dimension $`(`$pure$`)`$ de $`X`$. 3.16. Cas galoisien. Dans le cas particulier où le revêtement $`p:\stackrel{~}{X}X`$ de la variété complexe compacte $`X`$ est galoisien, de groupe $`\mathrm{\Gamma }`$, on a une opération naturelle du groupe $`\mathrm{\Gamma }`$ sur tous les objets définis précédemment : $`\stackrel{~}{}`$, $`p_{(2)}\stackrel{~}{}`$, $`C_{(2),\mathrm{loc}}^q(\stackrel{~}{𝒰},\stackrel{~}{})`$, $`H_{(2)}^q(\stackrel{~}{𝒰},\stackrel{~}{})`$, $`\mathrm{}`$, et ce pour tout $`q0`$ et tout recouvrement de Stein $`p`$-simple $`𝒰`$. On a donc aussi une action de $`\mathrm{\Gamma }`$ sur les espaces de cohomologie $`H_{(2)}^q(\stackrel{~}{X},\stackrel{~}{})`$, $`\underset{¯}{H}_{(2)}^q(\stackrel{~}{X},\stackrel{~}{})`$, $`\overline{H}_{(2)}^q(\stackrel{~}{X},\stackrel{~}{})`$. Dans le cas où $`X`$ est compacte, cette opération induit une action unitaire sur l’espace de Hilbert $`C_{(2)}^q(\stackrel{~}{𝒰},\stackrel{~}{})`$. ###### 3.17. Proposition Soit $`p:\stackrel{~}{X}X`$ un revêtement galoisien de groupe $`\mathrm{\Gamma }`$ de la variété complexe compacte $`X`$ ; soit $``$ un faisceau cohérent sur $`X`$, et $`𝒰`$ un recouvrement de Stein ouvert, fini et $`p`$-simple de $`X`$. Cette action définit une action de $`\mathrm{\Gamma }`$ sur $`H_{(2)}^{}(\stackrel{~}{X},\stackrel{~}{})`$ qui préserve chacune des semi-normes préhilbertiennes (équivalentes entre elles) dont cet espace peut être muni. En particulier, $`\mathrm{\Gamma }`$ agit sur $`\underset{¯}{H}_{(2)}^{}(\stackrel{~}{X},\stackrel{~}{})`$ et $`\overline{H}_{(2)}^{}(\stackrel{~}{X},\stackrel{~}{})`$, de manière unitaire sur ce dernier espace $`(`$qui est de Hilbert$`)`$. § 4. Théorèmes d’annulation. 4.0. Les nombreux résultats d’annulation accessibles par les techniques $`L^2`$ usuelles vont en général se transcrire mot pour mot pour donner des versions s’appliquant en cohomologie $`L^2`$. Nous présentons ici quelques énoncés parmi les plus fondamentaux. ###### 4.1. Théorème de Kodaira-Serre $`L^2`$ Soit $`X`$ une variété projective lisse, et $``$ un faisceau analytique cohérent sur $`X`$. Soit $``$ un fibré en droites ample sur $`X`$ et $`p:\stackrel{~}{X}X`$ un revêtement étale de $`X`$. Il existe $`m_0=m_0(,)`$, indépendant du revêtement $`p`$, tel que $`H_{(2)}^q(X,\stackrel{~}{}(m))=0`$ si $`q>0`$ et $`mm_0`$ $`(`$on pose ici comme d’habitude $`(m):=^m)`$. Démonstration. $``$ admet une résolution localement libre de longueur $`0rn=dim_C(X)`$. Si $`r=0`$, $``$ est localement libre, et le résultat est conséquence directe de 3.5 ci-dessus et de l’existence de solutions $`L^2`$ à l’équation : $`\overline{}\stackrel{~}{v}=\stackrel{~}{w}`$, avec $`\overline{}\stackrel{~}{w}=0`$ et $`\stackrel{~}{w}`$ section $`L^2`$ de $`\stackrel{~}{}^{0,1}`$ (\[D\], théorème 5.1). Sinon, on procède à nouveau par récurrence sur $`r`$, supposant le résultat vrai pour $`r10`$. Dans ce cas, l’assertion résulte immédiatement de la suite exacte longue de cohomologie $`L^2`$ (théorème 3.2) associée à la suite exacte de faisceaux : $`0𝒢0`$, où $``$ est localement libre et où $`𝒢`$ admet une résolution localement libre de longueur $`(r1)`$. 4.2. Exemple (Cet exemple a partiellement motivé la construction présentée ici). Soit $`X`$ une variété projective, $``$ et $``$ des faisceaux analytiques cohérents sur $`X`$, avec $``$ fibré en droites ample. Soit $`p:\stackrel{~}{X}X`$ un revêtement étale de $`X`$ et $`Y`$ un sous-schéma (non nécessairement réduit de $`X`$). Il existe $`m_0:=m_0(,,X,Y)`$ tel que, pour $`mm_0`$ le morphisme de restriction naturel $$H_{(2)}^0(\stackrel{~}{X},\stackrel{~}{}\stackrel{~}{}^m)H_{(2)}^0(\stackrel{~}{Y},\stackrel{~}{}_{|Y}\stackrel{~}{}_{|Y}^m)$$ soit surjectif. On utilise en effet la suite-exacte $$0_Y_{|Y}0$$ $`_{|Y}=/_Y`$ est la restriction de $``$ au sous-schéma $`Y`$. Le théorème de Kodaira-Serre implique l’annulation du groupe $$H_{(2)}^1(\stackrel{~}{X},\stackrel{~}{}_{\stackrel{~}{Y}}\stackrel{~}{}\stackrel{~}{}^m)$$ pour $`mm_0`$ assez grand, d’où le résultat. Ceci s’applique entre autres au cas où $`Y`$ est le schéma ponctuel associé à l’anneau $`𝒪_X/_a^{k+1}`$ des jets d’ordre $`k`$ de fonctions en un point $`a`$ de $`X`$. On voit alors que les $`k`$-jets de $`p_{(2)}\left(\stackrel{~}{}\stackrel{~}{}^m\right)`$ sont engendrés pour $`mm_0`$ assez grand par les sections globales $`L^2`$ du faisceau $`\stackrel{~}{}\stackrel{~}{}^m`$ sur $`\stackrel{~}{X}`$ (il faut voir que $`m_0`$ peut être choisi indépendant de $`a`$, mais c’est immédiat en contrôlant un tant soit peu les résolutions libres globales des anneaux $`𝒪_X/_a^{k+1}`$). Les résultats suivants sont des transcriptions immédiates des résultats $`L^2`$ classiques pour les variétés kählériennes complètes (\[AV\], \[D\]), et nous les énonçons donc sans commentaires (le corollaire 4.5 étant par exemple déjà mentionné dans \[K\], 11.4). ###### 4.3. Théorème de Akizuki-Kodaira-Nakano $`L^2`$ Soit $`X`$ une variété projective lisse de dimension complexe $`n`$, $`p:\stackrel{~}{X}X`$ un revêtement étale de $`X`$ et $``$ un fibré en droites ample sur $`X`$. Alors on a $$H_{(2)}^q(\stackrel{~}{X},\mathrm{\Omega }_{\stackrel{~}{X}}^r\stackrel{~}{})=0\text{si }q+rn+1\text{.}$$ ###### 4.4. Théorème de Nadel $`L^2`$ Soit $`X`$ une variété compacte $`(`$projective ou de Moishezon$`)`$, lisse, $`p:\stackrel{~}{X}X`$ un revêtement étale de $`X`$ et $``$ un fibré en droites sur $`X`$. On suppose que $``$ possède une métrique hermitienne singulière $`h`$ dont la $`(1,1)`$-forme de courbure $`\mathrm{\Theta }_h()`$ est positive au sens des courants, minorée par une $`(1,1)`$ forme de classe $`C^{\mathrm{}}`$ définie positive. Alors on a $$H_{(2)}^q(\stackrel{~}{X},K_{\stackrel{~}{X}}\stackrel{~}{}\stackrel{~}{}(h))=0\text{pour tout }q1\text{,}$$ $`(h)𝒪_X`$ désigne l’idéal multiplicateur des germes de fonctions holomorphes $`f`$ telles que $`|f|^2e^\phi <+\mathrm{}`$ $`(e^\phi `$ désignant le poids qui représente localement la métrique $`h)`$. ###### 4.5. Corollaire (Théorème de Kawamata-Viehweg $`L^2`$) Si $`X`$ est une variété de Moishezon lisse et $`p:\stackrel{~}{X}X`$ un revêtement étale de $`X`$. On suppose donné un fibré en droites $``$ numériquement équivalent à la somme d’un $`Q`$-diviseur $`D`$ nef $`(`$numériquement effectif$`)`$, et d’un $`Q`$-diviseur effectif $`E`$. (i) Si $`D`$ est gros, alors $`H_{(2)}^q(\stackrel{~}{X},K_{\stackrel{~}{X}}\stackrel{~}{}\stackrel{~}{}(E))=0`$ pour tout $`q1`$. (ii) Si $`D`$ est de dimension numérique $`\nu n=dimX`$, l’annulation a lieu pour $`q>n\nu `$. (iii) Si $`D`$ est de dimension numérique $`\nu n`$ et si l’idéal $`(E)`$ est un faisceau inversible $`(`$i.e. l’idéal d’un diviseur effectif$`)`$, la cohomologie séparée $`\overline{H}_{(2)}^q(\stackrel{~}{X},\stackrel{~}{}^1(\stackrel{~}{}(E))^1)`$ est nulle pour tout $`q<\nu `$. Démonstration. Rappelons que $`(E)`$ est le faisceau associé à $`\phi =\frac{1}{k}\mathrm{log}|g|`$, où $`g`$ est un générateur de $`𝒪(kE)`$. La preuve de 4.5 consiste en une réduction à 4.4, à peu près identique à celle effectuée dans le cas classique. (i) Si $`D`$ est ample, le résultat résulte directement de 4.4, en munissant $`𝒪(D)`$ d’une métrique lisse à courbure positive et $`𝒪(E)`$ de la métrique associée au poids $`e^\phi `$ (dont la courbure est le courant d’intégration $`[E]`$). En général, si $`D`$ est seulement nef et gros, on peut écrire $`D=D^{}+F`$ avec $`D^{}`$ ample et $`F`$ un $`Q`$-diviseur effectif aussi petit que l’on veut. On peut en particulier supposer que $`(E+F)=(E)`$. (ii) On se ramène au cas où la dimension numérique est maximale par un argument classique de sections hyperplanes et un raisonnement par récurrence sur la dimension. De façon précise, on choisit un diviseur lisse $`Y`$ très ample dans $`X`$ et on considère la suite exacte courte $$0K_XK_X𝒪(Y)K_Y_{|Y}0.$$ L’annulation de la cohomologie du terme central est obtenue par Kodaira-Serre en prenant $`Y`$ assez grand, tandis que l’annulation de la cohomologie en degré $`q1`$ du terme de droite résulte de l’hypothèse de récurrence. (iii) C’est un cas particulier de (ii), si on utilise la dualité de Serre. Il serait intéressant de savoir si la cohomologie non séparée $`H_{(2)}^q(\stackrel{~}{X},\stackrel{~}{}^1(\stackrel{~}{}(E))^1)`$ est nulle elle aussi. La difficulté est que c’est une cohomologie “duale” d’une cohomologie $`L^2`$, qui ne s’obtient pas directement par application d’estimations $`L^2`$ globales. § 5. Théorème de finitude et théorème de l’indice. 5.0. Notre objectif est ici d’étendre au cas de faisceaux analytiques cohérents quelconques le théorème de l’indice $`L^2`$ de Atiyah \[A\]. La preuve en est purement formelle à partir des résultats des sections précédentes. ###### 5.1. Théorème Soit $`X`$ un espace analytique compact et $``$ un faisceau analytique cohérent sur $`X`$. Soit $`p:\stackrel{~}{X}X`$ un revêtement étale galoisien de groupe $`\mathrm{\Gamma }`$. Pour tout $`q0`$, le groupe $`H_{(2)}^q(\stackrel{~}{X},\stackrel{~}{})`$ est un $`\mathrm{\Gamma }`$-module $`L^2`$ de présentation finie. En particulier, la $`\mathrm{\Gamma }`$-dimension de $`\overline{H}_{(2)}^q(\stackrel{~}{X},\stackrel{~}{})`$, notée $`h_{(2)}^q(\stackrel{~}{X},\stackrel{~}{})`$ est (un nombre réel) fini. De plus, la caractéristique d’Euler $`L^2`$ $$\chi _{(2)}(\stackrel{~}{X},\stackrel{~}{}):=\underset{q=0}{\overset{n}{}}(1)^qh_{(2)}^q(\stackrel{~}{X},\stackrel{~}{})$$ sur $`\stackrel{~}{X}`$ est égale à la caractéristique d’Euler ordinaire: $$\chi _{(2)}(\stackrel{~}{X},\stackrel{~}{})=\chi (X,):=\underset{q=0}{\overset{n}{}}(1)^qh^q(X,).$$ (Voir l’appendice pour les notions hilbertiennes requises, en particulier 6.4 et 6.5). Démonstration. Si $`X`$ est lisse et $``$ localement libre, c’est le théorème de l’indice $`L^2`$ d’Atiyah (\[A\]). En général, on raisonne par récurrence sur la dimension $`n=dimX`$, en utilisant un dévissage de $``$ et une résolution des singularités. Supposons le théorème déjà démontré en dimension $`n1`$; les résultats sont triviaux en dimension $`0`$, car si $`X=\{p\}`$, on a $$\chi (X,)=h^0(\{p\},)=dim_p$$ tandis que $`H_{(2)}^0(\stackrel{~}{X},\stackrel{~}{})=\mathrm{}^2(\mathrm{\Gamma })_p`$, d’où $`\chi _{(2)}(\stackrel{~}{X},\stackrel{~}{})=dim_p`$. En général, si $`X`$ n’est pas réduit, on peut considérer sa réduction $`X_{\mathrm{red}}`$ et la filtration de $``$ par les $`𝒩^p`$, où $`𝒩`$ désigne l’idéal des éléments nilpotents de $`𝒪_X`$. Le gradué $`𝒩^p/𝒩^{p+1}`$ de cette filtration est constitué de faisceaux cohérents sur $`𝒪_{X_{\mathrm{red}}}`$. Par additivité de la caractéristique d’Euler (ordinaire ou $`L^2`$, grâce à 3.2), on est ramené au cas où $`X`$ est réduit. Si $`X`$ n’est pas lisse, on utilise le théorème de Hironaka pour trouver une désingularisation $`f:YX`$. Soit $`p^{}:\stackrel{~}{Y}`$ le revêtement image réciproque de $`p:\stackrel{~}{X}X`$ par $`f`$ et $`𝒢=f^{}`$, qui est un faisceau cohérent sur $`Y`$. On a un morphisme naturel injectif $`f_{}𝒢`$, et le faisceau quotient $`f_{}𝒢/`$ est à support dans le lieu singulier $`X_{\mathrm{sing}}`$. D’après l’hypothèse de récurrence, les résultats sont vrais pour le faisceau quotient $`f_{}𝒢/`$, et on est donc ramené à traiter le cas du faisceau image directe $`f_{}𝒢`$. On utilise alors les suites spectrales de Leray $$\begin{array}{cc}\hfill H^{}(X,R^{}f_{}𝒢)& H^{}(Y,𝒢),\hfill \\ \hfill H_{(2)}^{}(\stackrel{~}{X},R^{}\stackrel{~}{f}_{(2)}\stackrel{~}{𝒢})& H_{(2)}^{}(\stackrel{~}{Y},\stackrel{~}{𝒢})\hfill \end{array}$$ (l’existence de la deuxième suite spectrale est une conséquence du théorème de Leray et de la Proposition 3.3). Supposons que le résultat soit déjà démontré dans le cas d’une variété lisse. Alors $`\chi (Y,𝒢)=\chi _{(2)}(\stackrel{~}{Y},\stackrel{~}{𝒢})`$ puisque $`Y`$ est lisse, et de même pour $`q>0`$ on a $`\chi (X,R^qf_{}𝒢)=\chi _{(2)}(\stackrel{~}{X},R^q\stackrel{~}{f}_{(2)}\stackrel{~}{𝒢})`$ pour $`q>0`$ par hypothèse de récurrence sur la dimension (les faisceaux $`R^qf_{}𝒢`$, $`q>0`$, sont supportés par $`X_{\mathrm{sing}}`$ qui est de dimension$`n1`$). Comme il y a préservation de la caractéristique d’Euler dans les différents niveaux d’une suite spectrale, il y a équivalence à prouver le résultat pour la paire $`(X,R^0f_{}𝒢)`$ ou pour la paire $`(Y,𝒢)`$, ce qui fait qu’on est ramené au cas où l’espace ambiant $`X`$ est lisse de dimension $`n`$. Dans ce cas, soit $`_{\mathrm{tors}}`$ la partie de torsion de $``$. Cette partie est à support en codimension $`n1`$, donc l’hypothèse de récurrence s’y applique. La suite exacte $$0_{\mathrm{tors}}/_{\mathrm{tors}}0$$ et l’additivité de la caractéristique d’Euler montre que l’on peut supposer $``$ sans torsion. En appliquant de nouveau le théorème de Hironaka, il existe une modification analytique propre $`f:YX`$, tel que $`𝒢=f^{}`$ soit localement libre sur $`Y`$. Des arguments identiques à ceux qui précèdent ramènent la preuve du cas de la paire $`(X,)`$ au cas de la paire $`(Y,𝒢)`$. On conclut en appliquant cette fois le théorème de l’indice $`L^2`$ de Atiyah \[A\] à $`(Y,𝒢)`$. § 6. Appendice : présentation hilbertienne et $`\mathrm{\Gamma }`$-dimension. ###### 6.1. Définition Soit $`H`$ un espace vectoriel complexe. Une présentation hilbertienne de $`H`$ est la donnée d’une application linéaire continue $`\delta :CZ`$ entre deux espaces de Hilbert et d’un isomorphisme (algébrique) $`(Z/\delta C)\stackrel{}{}H`$. On note $`(\delta C)`$ (resp. $`(\overline{\delta C})`$) l’image dans $`Z`$ de $`C`$ (resp. son adhérence). Associée à une telle présentation est définie une suite exacte : $$0\underset{¯}{H}=:(\overline{\delta C}/\delta C)H=:(Z/\delta C)\overline{H}:=(Z/\overline{\delta C})0.$$ On appelle $`\underset{¯}{H}`$ (resp. $`\overline{H}`$) le noyau (resp. la réduction) de $`H`$ relative à cette présentation hilbertienne. (Ces notions ne dépendent que de la classe d’équivalence des espaces de Hilbert). Une application entre deux présentations hilbertiennes $`\delta :CZ`$ et $`\delta ^{}:C^{}Z^{}`$ est un diagramme commutatif $`s:CC^{}`$ et $`r:ZZ^{}`$ d’applications linéaires continues telles que $`r\delta =\delta ^{}s`$. Une telle application induit un diagramme commutatif : $$\begin{array}{ccccccccc}0& & \underset{¯}{H}& & H& & \overline{H}& & 0\\ \\ & & \underset{¯}{r}& & \left[r\right]& & \overline{r}\\ \\ 0& & \underset{¯}{H}^{}& & H^{}& & \overline{H}^{}& & 0\end{array}$$ $`(6.1^{})`$ Deux présentations hilbertiennes sont dites compatibles s’il existe une application entre elles. ###### 6.2. Proposition La situation étant celle décrite en 6.1, alors : i. Si $`[r]`$ est surjective, $`\overline{r}`$ est surjective. Si $`[r]`$ est injective, $`\underset{¯}{r}`$ est injective. ii. Si $`[r]`$ est bijective, $`\overline{r}`$ et $`\underset{¯}{r}`$ sont bijectives. Donc : $`\overline{r}`$ est une équivalence d’espaces de Hilbert. (En particulier, deux présentations hilbertiennes de $`H`$ fournissent le même noyau et la même réduction de $`H`$). Démonstration (de 6.2). La première assertion est évidente. Pour établir la seconde et montrer que $`\overline{r}:(Z/\overline{\delta C})(Z^{}/\overline{\delta ^{}C^{}})`$ est injective, if faut vérifier la propriété suivante : si $`r(z)(\overline{\delta ^{}C^{}})`$, alors $`z(\overline{\delta C})`$. Remarquons que, puisque $`[r]`$ est surjective, l’application $$\delta ^{}r:C^{}ZZ^{}$$ est surjective. Soit $`K`$ son noyau. Alors $`(\delta ^{}r)`$ admet un relèvement continu $`\phi :ZK^{}`$ qui est un isomorphisme d’espaces de Hilbert (où $`K^{}`$ est l’orthogonal de $`K`$) $$K=\{(c^{},\xi )r(\xi )=\delta ^{}(c^{})\}.$$ Soit alors $`(z+\overline{\delta C})\text{Ker}(\overline{r})`$ ; on a donc : $`r(z)=z^{}(\overline{\delta ^{}C^{}})`$, et $`z^{}=lim\delta ^{}(c_n^{})`$. Soit $`(\gamma _n^{},z_n):=(c_n^{},z)\phi \left(z^{}\delta ^{}(c_n^{})\right)K`$, on a $`\phi \left(z^{}\delta ^{}(c_n^{})\right)0`$ et $`z_nz`$, en particulier. Or, $$\delta ^{}(\gamma _n^{})+r(z_n)=(\delta ^{}r)(\gamma _n^{},z_n)=\delta ^{}(c_n^{})+r(z)\left(z^{}\delta ^{}(c_n^{})\right)=r(z)z^{}=0,$$ donc $`r(z_n)=\delta ^{}\gamma _n^{}`$, et $`z_n\delta C`$ par injectivité de $`[r]`$. On a donc bien $`z(\overline{\delta C})`$ comme annoncé. 6.3. $`\mathrm{\Gamma }`$-présentation. Soit $`\mathrm{\Gamma }`$ un groupe discret agissant sur l’espace vectoriel complexe $`H`$. Une $`\mathrm{\Gamma }`$-présentation hilbertienne $`\delta :CZ`$ de $`H`$ est une présentation hilbertienne telle ue $`\mathrm{\Gamma }`$ agisse de manière unitaire et équivariante sur $`C`$ et $`Z`$, l’action sur le quotient $`(Z/\delta C)`$ étant celle sur $`H`$. Une telle présentation munit $`\underset{¯}{H}`$ et $`\overline{H}`$ d’une action de $`\mathrm{\Gamma }`$, qui est unitaire sur $`\overline{H}`$. 6.4. $`\mathrm{\Gamma }`$-dimension. Soit $`V`$ un espace de Hilbert muni d’une action unitaire de $`\mathrm{\Gamma }`$. On dit que $`V`$ est de $`\mathrm{\Gamma }`$-dimension finie s’il existe un sous-ensemble fini $`\{v_1,\mathrm{},v_m\}`$ de $`V`$ tel que le sous-espace vectoriel engendré par les $`(\gamma .v_i)`$ $`(\gamma \mathrm{\Gamma },1im)`$ soit dense dans $`V`$. De manière équivalente, $`V`$ est un quotient ou un sous-espace de $`\left(\mathrm{}^2(\mathrm{\Gamma })_CC^m\right)`$, ceci de manière compatible avec les actions naturelles de $`\mathrm{\Gamma }`$ (triviale sur $`C^m`$). Dans cette situation, on définit la $`\mathrm{\Gamma }`$-dimension $`dim_\mathrm{\Gamma }(V)`$, qui est un nombre réel $`(`$inférieur ou égal à $`m`$, ici, et indépendant du choix des $`(v_i))`$. Voir \[P\] pour cette notion. Les propriétés fondamentales (utilisées ici) de cette notion sont les suivantes : (6.4.1) $`dim_\mathrm{\Gamma }(V)0`$, avec égalité si et seulement si $`V=0`$. (6.4.2) Si $`V`$ est isomorphe à un sous-espace dense de $`W`$, alors $$dim_\mathrm{\Gamma }V=dim_\mathrm{\Gamma }W.$$ (6.4.3) Si $`V=WW^{}`$ (somme directe orthogonale), alors : $$dim_\mathrm{\Gamma }V=dim_\mathrm{\Gamma }W+dim_\mathrm{\Gamma }W^{}.$$ (6.4.4) $`dim_\mathrm{\Gamma }\left(\mathrm{}^2(\mathrm{\Gamma })\right)=1`$. (6.4.5) Si $`\mathrm{\Gamma }`$ est un groupe fini de cardinal $`|\mathrm{\Gamma }|`$, alors $`dim_\mathrm{\Gamma }V=|\mathrm{\Gamma }|^1dim_CV`$. 6.5. $`\mathrm{\Gamma }`$-présentation finie. Soit $`\gamma :CZ`$ une $`\mathrm{\Gamma }`$-présentation de l’espace vectoriel complexe $`H`$, muni d’une $`\mathrm{\Gamma }`$-action. On dit que $`(\delta :CZ)`$ est une $`\mathrm{\Gamma }`$-présentation finie de $`H`$ si $`C`$ et $`Z`$ sont de $`\mathrm{\Gamma }`$-dimensions finies. Alors $`(\overline{\delta C})`$ est de $`\mathrm{\Gamma }`$-dimension finie au plus égale à celle de $`C`$, et $`\overline{H}`$ est aussi de $`\mathrm{\Gamma }`$-dimension finie. En fait, $`dim_\mathrm{\Gamma }\overline{H}=dim_\mathrm{\Gamma }Zdim_\mathrm{\Gamma }\overline{\delta C}`$. On pose alors $`dim_\mathrm{\Gamma }H=dim_\mathrm{\Gamma }\overline{H}`$. Remarquons que cette définition peut être donnée même si $`C`$ n’est pas supposé être de $`\mathrm{\Gamma }`$-dimension finie. Bibliographie \[AG\] A. Andreotti, H. Grauert, Théorèmes de finitude pour la cohomologie des espaces complexes. Bull. Soc. Math. France 90 (1962), 193–259. \[AV\] A. Andreotti, E. Vesentini, Carleman estimates for the Laplace-Beltrami equation in complex manifolds. Publ. Math. I.H.E.S. 25 (1965), 81–130. \[A\] M. Atiyah, Elliptic operators, discrete groups and Von Neumann algebras. Astérisque 32-33 (1976), 43-72. \[D\] J.-P. Demailly, Estimations $`L^2`$ pour l’opérateur $`\overline{}`$ d’un fibré vectoriel hermitien semi-positif. Ann. Sc. ENS 15 (1982), 457-511. \[E\] P. Eyssidieux, Théorie de l’adjonction $`L^2`$ sur le revêtement universel. Preprint (1997). \[G\] M. Gromov, Kähler hyperbolicity and $`L^2`$-Hodge theory. J. Diff. Geom. 33 (1991), 263-291. \[H\] L. Hörmander, An introduction to Complex Analysis in several variables. 1st edition, Elsevier Science Pub., New York, 1966, 3rd revised edition, North-Holland Math. library, Vol 7, Amsterdam (1990). \[K\] J. Kollár, Shafarevitch maps and automorphic forms. Princeton University Press (1995). \[Oh\] T. Ohsawa, A reduction theorem for cohomology groups of very strongly $`q`$-convex Kähler manifolds. Invent. Math. 63 (1981) 335-354 ; 66 (1982) 391-393. \[P\] P. Pansu, An introduction to $`L^2`$ Betti numbers. Preprint (1994). \[W\] A. Weil, Introduction à l’étude des variétés kählériennes. Hermann (1958). Frédéric Campana Université de Nancy I, Faculté des Sciences, Département de Mathématiques BP 239, 54506 Vandoeuvre les Nancy, France E-mail: campana@iecn.u-nancy.fr Jean-Pierre Demailly Université de Grenoble I, Institut Fourier, UMR 5582 du CNRS BP74, 100 rue des Maths, 38402 Saint-Martin d’Hères, France E-mail: demailly@ujf-grenoble.fr
warning/0002/hep-th0002101.html
ar5iv
text
# Wilson lines on noncommutative tori ## I Introduction Gauge theories on noncommutative spaces arise as low energy effective theories on $`D`$-brane world volumes in the presence of the background $`B`$-field in the theory of open strings . The simplest noncommutative space where a gauge theory can be constructed is the space $`𝐑_\theta ^d`$ with coordinates $`x^i`$ obeying the Heisenberg commutation relations, $$[x^i,x^j]=i\theta ^{ij},$$ (1) where $`\theta ^{ij}=\theta ^{ji}`$ is a constant anti-symmetric matrix. In string theory, $`\theta `$ is given by the following formula $$\theta =(2\pi \alpha ^{})^2(g+2\pi \alpha ^{}B)^1B(g2\pi \alpha ^{}B)^1,$$ where $`\alpha ^{}`$ is the inverse to the string tension, $`g`$ is the metric and $`B`$ is the $`B`$-field on the brane world-volume. Functions on the space $`𝐑_\theta ^d`$ can be identified with ordinary functions on $`𝐑^d`$ with the noncommutative product given by the Moyal formula, $$(uv)(x)=\left(e^{\frac{i}{2}\theta ^{ij}_i^x_j^y}u(x)v(y)\right)_{y=x}.$$ (2) Along with the space $`𝐑_\theta ^d`$ it is natural to consider its compactification $`T_\theta ^d`$. We can choose the compactification radii $`R_i`$ such that the functions on $`T_\theta ^d`$ are invariant with respect to the shifts, $$f(x^1+2\pi n_1R^1,\mathrm{},x^d+2\pi n_dR^d)=f(x^1,\mathrm{},x^d),$$ where $`n_i`$ are arbitrary integers. Since $`\theta ^{ij}`$ is a constant tensor, the Moyal product is invariant with respect to such shifts. Therefore, the $``$-product restricts to periodic functions. An important ingredient in the ordinary gauge theory is the notion of a Wilson line, $$W_\gamma (A)=\mathrm{Tr}P\mathrm{exp}(i_\gamma A),$$ (3) where $`A`$ is a gauge field and $`\gamma `$ is a closed contour. The expression $`W_\gamma (A)`$ is gauge invariant and, hence, defines an observable in the gauge theory. One reason why $`W_\gamma (A)`$ is an important object is given by the ’t Hooft’s criterion of confinement . Another reason is as follows. Suppose that the curvature of $`A`$, $$F_{ij}=_iA_j_jA_i+i[A_j,A_i],$$ vanishes. Locally, the condition $`F=0`$ implies that $`A`$ is a pure gauge. Globally, flat connections (gauge fields with vanishing curvature) may have moduli. One can distinguish a nontrivial gauge field with $`F=0`$ from a pure gauge by looking at the Wilson lines along noncontractible contours. The standard application of this principle is the Aharonov-Bohm effect where the phase of the Wilson line for electromagnetic vector potential influences the interference pattern. It is the goal of this paper to study the notion of a Wilson line in a gauge theory on the noncommutative torus $`T_\theta ^d`$. In contrast to the previous suggestions , we only consider the case of flat connections. In our view, there is a conceptual difficulty in the case of $`F_{ij}0`$: the definition of a Wilson line contains a contour $`\gamma `$. It might be a nontrivial task to find noncommutative subspaces of a noncommutative space. Technically speaking, the requirement is that the algebra of functions vanishing on the subspace be an ideal in the noncommutative algebra of functions. For instance, if $`\theta `$ is nondegenerate and generic, the noncommutative torus $`T_\theta ^d`$ has no subspaces different from $`T_\theta ^d`$ itself. In particular, there is not a single ‘noncommutative contour’ on $`T_\theta ^d`$. In the case of $`F_{ij}=0`$, the Wilson line only depends on the homotopy class of the contour $`\gamma `$. We propose a natural gauge invariant object which generalizes $`W_\gamma (A)`$ in the case of flat noncommutative gauge fields. It has been shown in that one can construct a transformation (the Seiberg-Witten map) which identifies the noncommutative gauge fields on $`𝐑_\theta ^d`$ (and on $`T_\theta ^d`$) for different values of $`\theta `$. The Seiberg-Witten map on general noncommutative spaces was recently studied in . We show that our noncommutative Wilson line is invariant with respect to the Seiberg-Witten map. In other words, we prove that the Seiberg-Witten map for flat connections is isomonodromic. It would be very interesting to see whether the noncommutative Wilson lines have some natural interpretation in the $`D`$-brane Physics. The following elementary example can be used as a starting point. The energy spectrum of a non-relativistic particle of mass $`m`$ and charge $`q`$ confined to a circle of length $`L`$ in the constant vector potential $`A`$ has the form, $$E_n=\frac{1}{2m}\frac{4\pi ^2\mathrm{}^2}{L^2}\left(n+i\mathrm{ln}(M)\right)^2,$$ where $$M=\mathrm{exp}\left(i\frac{qAL}{2\pi \mathrm{}}\right)$$ is the monodromy of the Wilson line winding around the circle. In the case of noncommutative torus, the charged particle is replaced by the end-point of an open string. It is plausible that the noncommutative Wilson line shifts the spectrum of an open string in a similar fashion. ## II Noncommutative Wilson lines In ordinary gauge theory, there are two ways to define a Wilson line for a gauge field $`A`$ with vanishing curvature $`F`$. For the purpose of this paper we restrict our attention to the Wilson lines on the torus $`T^d`$. Since a gauge field with vanishing curvature is locally a pure gauge, one can solve the equation, $$_ig(x)=iA_i(x)g(x),$$ (4) where $`g`$ is not necessarily periodic. Different solutions of (4) can be obtained from each other by multiplication by a constant from the right, $`g(x)g(x)h`$. Under gauge transformations solutions of equation (4) get multiplied from the left, $`g(x)h(x)g(x)`$. One can now introduce left and right monodromy matrices, $`M_i=g^1g_i`$ and $`\stackrel{~}{M}_i=g_ig^1`$, where $`gg(x^1,\mathrm{},x^d)`$ and $`g_ig(x^1,\mathrm{},x^i+2\pi R^i,\mathrm{},x^d)`$ are solutions of (4). It is easy to see that the monodromies $`M_i`$ are gauge-invariant and do not depend on the point $`x`$. If one replaces the solution $`g(x)`$ by some other solution of equation (4), $`g(x)h`$, the monodromy $`M_i`$ gets conjugated, $`M_ih^1M_ih`$. Therefore, the trace of $`M_i`$ is a gauge invariant object which is independent of the choice of a solution of (4) and can be used as the definition of the Wilson line, $`W_i=\mathrm{Tr}M_i`$. The monodromies $`\stackrel{~}{M}_i`$ are the usual ordered exponentials, $$\stackrel{~}{M}_i=P\mathrm{exp}(i_{x^i}^{x^i+2\pi R^i}A).$$ They are independent of the choice of the solution $`g(x)`$ but have explicit dependence on the point $`x`$ and transform covariantly under the gauge transformations, $`\stackrel{~}{M}_ih(x)\stackrel{~}{M}_ih(x)^1`$. Again, taking a trace yields a gauge invariant object independent of all choices which coincides with $`W_i`$. In the noncommutative gauge theory, only the prescription which uses monodromies $`M_i`$ survives. It is still possible to construct objects similar to $`\stackrel{~}{M}_i`$ (see e.g. ) which transform covariantly under the gauge transformations, $`\stackrel{~}{M}_ih(x)\stackrel{~}{M}_ih(x)^1`$. However, the matrix trace fails to be cyclic under the Moyal product, that is $`\mathrm{Tr}(uv)\mathrm{Tr}(vu)`$. This implies that $`\mathrm{Tr}\stackrel{~}{M}_i`$ is not gauge-invariant. In what follows we present the definition of a Wilson line in noncommutative gauge theory which uses the monodromy $`M_i`$. In the noncommutative case, the curvature of the gauge field is defined using the $``$-product (see e. g. ), $$F_{ij}=_iA_j_jA_iiA_iA_j+iA_jA_i.$$ (5) The infinitesimal gauge transformations are of the form, $$\delta _\lambda A_i=_i\lambda +i\lambda A_iiA_i\lambda .$$ (6) Let us mention that not every compact Lie group can be used as a gauge group on $`T_\theta ^d`$. For instance, $`G=U(N)`$ works for any $`\theta `$ whereas $`G=SU(N)`$ in general does not give rise to a gauge group on the noncommutative torus. More precisely, the $``$-product $`gh`$ of two unitary matrices is always unitary, but in general $`\mathrm{det}(gh)\mathrm{det}(g)\mathrm{det}(h)`$. Equations (5) and (6) are obtained from the standard formulas in ordinary (commutative) gauge theory by introducing $``$-products instead of the ordinary products. It is slightly more complicated to find a counterpart of the general formula for gauge transformations, $$A_i^h=i_ihh_{}^1+hA_ih_{}^1.$$ (7) Here $`h(x)`$ is an element of the gauge group, and $`h_{}^1`$ is the inverse of $`h`$ with respect to the $``$-product, $`hh_{}^1=h_{}^1h=1`$. Note that for $`\theta 0`$ the element $`h_{}^1`$ differs from the ordinary inverse $`h^1`$. More explicitly, $$h_{}^1=h^1+\frac{i}{2}\theta ^{kl}h^1(_kh)h^1(_lh)h^1+𝒪(\theta ^2).$$ We now turn to the case of a noncommutative gauge field $`A`$ with vanishing curvature $`F=0`$. Then, one may expect that at least locally $`A`$ is a pure gauge. That is, there is an element of the gauge group $`g`$ such that, $$_ig=iA_ig.$$ (8) At this point it is convenient to view gauge fields on $`T_\theta ^d`$ as periodic gauge fields on $`𝐑_\theta ^d`$. Then, we expect that equation (8) admits global solutions. For now we assume that such solutions exist and return to this issue in the next section. Our next task is to show that the ratio $`m=g_{}^1g^{}`$ of two solutions $`g`$ and $`g^{}`$ of equation (8) is a constant. Indeed, $`_im=g_{}^1_igg_{}^1g^{}+g_{}^1_ig^{}`$ (9) $`=ig_{}^1(A_iA_i)g^{}=0.`$ (10) Let us mention that in the derivation of this equation we have used that the $``$-product on $`𝐑_\theta ^d`$ satisfies the Leibniz rule, $`_i(fg)=_ifg+f_ig`$. In general, this property does not hold on noncommutative spaces. If $`g(x^1,\mathrm{},x^d)`$ is a solution of (8) then so is $`g_i(x^1,\mathrm{},x^d)=g(x^1,\mathrm{},x^i+2\pi R^i,\mathrm{},x^d)`$. By equation (9), the ratio $`M_i=g_{}^1g_i`$ is independent of the point on $`T_\theta ^d`$. We call the group elements $`M_i`$ noncommutative monodromies of the equation (8). As usual, the value of the monodromy depends on the choice of the solution $`g(x)`$. For some other solution, $`g^{}(x)=g(x)h=g(x)h`$ one gets, $`M_i^{}=h^1g_{}^1g_ih=h^1M_ih`$. Here we used that the $``$-product with a constant coincides with the ordinary product. Finally, we can define the Wilson lines which are independent of the choice of the solution $`g(x)`$, $$W_i(A)=\mathrm{Tr}M_i=\mathrm{Tr}(g_{}^1g_i).$$ (11) Let us stress, that, since $`M_i`$ are coordinate independent, we use the ordinary matrix trace here. Similar to their commutative counterparts (3), the Wilson lines $`W_i(A)`$ are gauge invariant, $$W_i(A^h)=\mathrm{Tr}(g_{}^1h_{}^1hg_i)=W_i(A),$$ where $`g(x)h(x)g(x)`$ is the composition of two gauge transformations. We put forward the definition (11) of noncommutative Wilson lines $`W_i(A)`$. In the next Section we apply the Seiberg-Witten map to the noncommutative gauge theory at hand and show that our Wilson lines $`W_i(A)`$ are invariant with respect to this transformation. Let us remark that although the considerations of this Section are restricted to the case of gauge fields with vanishing curvature, they may be viewed as the general definition of a noncommutative Wilson line. Indeed, in the commutative case the Wilson lines are assigned to closed contours. When restricted to a 1-dimensional contour, any gauge field has vanishing curvature and one can use the monodromy $`M`$ instead of the ordered exponential $`\stackrel{~}{M}`$ to construct the Wilson line. In the noncommutative context it is more difficult to construct submanifolds. In particular, as we pointed out in Introduction, the noncommutative tori with generic $`\theta `$ have no nontrivial submanifolds. However, if one can find a subtorus $`\mathrm{\Gamma }`$ of the noncommutative torus $`T_\theta ^d`$ such that the restriction of the gauge field $`A`$ to $`\mathrm{\Gamma }`$ has vanishing curvature, one can define the Wilson lines of $`A`$ along $`\mathrm{\Gamma }`$. If $`\mathrm{\Gamma }`$ is 1-dimensional, it defines one Wilson line observable, similar to commutative gauge theory. ## III The Seiberg-Witten map It was established in that the gauge theories on noncommutative tori $`T_\theta ^d`$ with different values of the deformation parameter $`\theta `$ are equivalent to each other. More explicitly, let $`\theta `$ and $`\theta +\delta \theta `$ be two infinitesimally close values of the deformation parameter. Then, there exists the Seiberg-Witten map $`A\widehat{A}(A)`$ and $`\lambda \widehat{\lambda }(\lambda ,A)`$ of the gauge fields and infinitesimal gauge parameters on $`𝐑_\theta ^d`$ to those on $`𝐑_{\theta +\delta \theta }^d`$ such that $$\widehat{A}(A+\delta _\lambda A)=\widehat{A}(A)+\delta _{\widehat{\lambda }(\lambda ,A)}\widehat{A}(A).$$ (12) This map is given by explicit formulas (see equation (3.8) in ). In this paper we only need the form of the Seiberg-Witten transformation for $`A`$ with $`F=0`$, $$\delta _\theta A_i=\frac{1}{4}\delta \theta ^{kl}(A_k_lA_i+_lA_iA_k).$$ (13) We show in the Appendix that under the Seiberg-Witten map solutions of the equation (8) transform according to formula, $$\delta _\theta g=\frac{1}{4i}\delta \theta ^{kl}A_kA_lg.$$ (14) It is convenient to introduce a special notation for the ‘gauge parameter’, $$\lambda (\delta \theta ,A)=\frac{1}{4}\delta \theta ^{kl}A_kA_l.$$ (15) Note, however, that $`\delta _\theta A_i`$ is not a gauge transformation. More exactly, $$\delta _{\lambda (\delta \theta ,A)}A_i=\frac{1}{4}\delta \theta ^{kl}(A_k_lA_i_lA_iA_k).$$ The reason is that in contrast to partial derivatives $`_i`$, the transformation $`\delta _\theta `$ violates the Leibniz rule for the Moyal product. Indeed, the exponential form of (2) implies, $$\delta _\theta (uv)=(\delta _\theta u)v+u(\delta _\theta v)+\frac{i}{2}\delta \theta ^{ij}_iu_jv.$$ (16) In particular, putting together equations (13)-(16) we obtain, $$\delta _\theta (g_{}^1)=g_{}^1\delta _\theta gg_{}^1.$$ (17) Surprisingly, there is no minus sign on the right hand side. This is due to the contribution of the last term in (16). We are now prepared to address the question of variation of the Wilson lines under the Seiberg-Witten map. Let us apply the transformation $`\delta _\theta `$ to equation $`g_i(x)=g(x)M_i`$. The result reads, $$i\lambda (\delta \theta ,A)g_i=i\lambda (\delta \theta ,A)gM_i+g\delta _\theta M_i.$$ Thus, we conclude that $`\delta _\theta M_i=0`$ and monodromies $`M_i`$ are $`\theta `$-independent, and so are the Wilson lines $`W_i(A)`$. Finally, we return to the question of existence of solution $`g(x)`$ of the equation (8). Clearly, such solutions exist for $`\theta =0`$. We can use these solutions as initial conditions in the pair of differential equations (13) and (14). Choosing a path between $`\theta _0=0`$ and some other value $`\theta _1`$, one can, at least formally, obtain solutions of equation (8) for $`\theta =\theta _1`$ by solving equations (13) and (14). In general, these solutions depend on the path between $`\theta _0`$ and $`\theta _1`$. In other words, two infinitesimal transformations $`\delta _\theta ^1`$ and $`\delta _\theta ^2`$ may have a nonvanishing commutator. This commutator applied to gauge fields $`A`$ was first computed in . We obtain an elegant formula for the variation of $`g`$, $$[\delta _\theta ^1,\delta _\theta ^2]g=\frac{1}{16}\delta \theta _1^{kl}\delta \theta _2^{mn}(_lA_n_mA_k_mA_k_lA_n).$$ There is no obvious reason for the right hand side to vanish even in the abelian case for $`\theta 0`$. ## Appendix Our goal is to show that transformation (14) is consistent with formula (8) provided that eq. (13) holds. Applying the partial derivative $`_i`$ to equation(14) and then using twice that $`F=0`$, we obtain $`\delta _\theta (_ig)=`$ $`={\displaystyle \frac{1}{4i}}\delta \theta ^{kl}(_iA_kA_l+A_k_iA_l+iA_kA_lA_i)g`$ $`={\displaystyle \frac{1}{4i}}\delta \theta ^{kl}(_kA_iA_l+A_k_lA_i+iA_iA_kA_l)g.`$ Taking into account the antisymmetry of $`\theta ^{ij}`$, we can rewrite this expression as follows: $`\delta _\theta (_ig)={\displaystyle \frac{1}{4i}}\delta \theta ^{kl}(2_kA_iA_l+_lA_iA_k`$ $`+A_k_lA_i+iA_iA_kA_l)g`$ $`={\displaystyle \frac{1}{2}}\delta \theta ^{kl}_kA_i_lg+i(\delta _\theta A_i)g+iA_i\delta _\theta g`$ $`=i\delta _\theta (A_ig).`$ Here we used formulas (13) and (14) in the third line, and (8) and (16) in the last line. Acknowledgments: We would like to thank K. Okuyama, V. Roubtsov and V. Schomerus for useful discussions. A.A. thanks the organizers and the participants of the conference on Noncommutative Gauge Theory (Leiden, November 1999) for inspiring discussions. The visit of A.B. to the Institute for Theoretical Physics, Uppsala University was supported by the grant INTAS 96-196 and by the grant VISBY-380 of the Swedish Institute.
warning/0002/hep-ex0002007.html
ar5iv
text
# Partial Wave Analysis of 𝐽/𝜓→𝛾⁢(𝐾^±⁢𝐾_𝑆⁰⁢𝜋^∓) ## Abstract BES data on $`J/\psi \gamma (K^\pm K_S^0\pi ^{})`$ are presented. There is a strong peak due to $`\eta (1440)/\iota `$, which is fitted with a Breit-Wigner amplitude with $`s`$-dependent widths for decays to $`K^{}K`$, $`\kappa K`$, $`\eta \pi \pi `$ and $`\rho \rho `$; $`\kappa `$ refers to the $`K\pi `$ S-wave. At a $`K\overline{K}\pi `$ mass of $`2040`$ MeV, there is a second peak with width $`400`$ MeV; $`J^P=0^{}`$ is preferred over $`1^+`$ and $`2^{}`$ respectively by 5.2 and 6.8 standard deviations. It is a possible candidate for a $`0^{}`$ $`s\overline{s}g`$ hybrid partner of $`\pi (1800)`$. preprint: BIHEP-EP1-2000-02 There have been earlier data from Mark III and DM2 for $`J/\psi `$ radiative decays to $`K^\pm K_S^0\pi ^{}`$, as well as $`K^+K^{}\pi ^0`$. Recently, the BES group has published data on the latter channel . Here we present BES data on decays to $`K^\pm K_S^0\pi ^{}`$. These data have lower backgrounds than for $`K^+K^{}\pi ^0`$, because of the identification of $`K_S^0\pi ^+\pi ^{}`$. Consequently, the partial wave analysis may be extended up to a $`K\overline{K}\pi `$ mass of 2300 MeV, covering an interesting structure at $`2040`$ MeV. The Beijing Spectrometer(BES) has collected $`7.8\times 10^6J/\psi `$ triggers, used here. Details of the detector are given in Ref. . We describe briefly those detector elements playing a crucial role in the present measurement. Tracking is provided by a 10 superlayer main drift chamber (MDC). Each superlayer contains four layers of sense wires measuring both the position and the ionization energy loss ($`d`$E$`/dx`$) of charged particles. The momentum resolution is $`\sigma _P/P=1.7\%\sqrt{1+P^2}`$, where $`P`$ is the momentum of charged tracks in GeV/$`c`$. The resolution of the $`d`$E$`/dx`$ measurement is $`\pm 9\%`$, providing good $`\pi /K`$ separation and proton identification for momenta up to 600 MeV/c. An array of 48 scintillation counters surrounding the MDC measures the time-of-flight (TOF) of charged tracks with a resolution of 330$`ps`$ for hadrons. Outside the TOF system is an electromagnetic calorimeter made of lead sheets and streamer tubes and having a $`z`$ positional resolution of 4 cm. The energy resolution scales as $`\sigma _E/E=22\%/\sqrt{E}`$, where $`E`$ is the energy in GeV. Outside the shower counter is a solenoidal magnet producing a 0.4 Tesla magnetic field. Each candidate event is required to have four charged tracks. Each track must have a good helix fit in the polar angle range $`0.8<\mathrm{cos}\theta <0.8`$ and a transverse momentum $`>60`$ MeV/c. A vertex is required within an interaction region $`\pm 30`$ cm longitudinally and 3 cm radially. A positive identification of just one $`K^\pm `$ is required using time of flight and/or $`dE/dx`$. Events are fitted kinematically to the 4C hypothesis $`J/\psi \gamma (K^\pm \pi ^{}\pi ^+\pi ^{})`$, requiring a confidence level $`>5\%`$. Backgrounds arise mainly from $`\pi ^0K^\pm \pi ^{}\pi ^+\pi ^{}`$ and $`K^\pm \pi ^{}\pi ^+\pi ^{}`$. Those events giving a better fit to these channels are rejected. Next, we require $`U_{miss}=E_{miss}P_{miss}<0.15`$ GeV/c<sup>2</sup>, so as to reject the events with multi-photons or more or less than one charged kaon; here, $`E_{miss}`$ and $`P_{miss}`$ are, respectively, the missing energy and missing momentum of all charged particles. The momentum of the $`K^\pm \pi ^{}\pi ^+\pi ^{}`$ system transverse to the photon $`P_{t\gamma }^2=4P_{miss}{}_{}{}^{2}\mathrm{sin}_{}^{2}(\theta _{m\gamma }/2)<0.005`$ (GeV/c)<sup>2</sup> is required in order to remove the background $`J/\psi \pi ^0K^\pm \pi ^{}\pi ^+\pi ^{}`$; here $`\theta _{m\gamma }`$ is the angle between the missing momentum and the photon direction. Finally, $`K_s^0`$ are selected with a cut on the $`\pi ^+\pi ^{}`$ invariant mass, $`M_{\pi ^+\pi ^{}}M_{K_s^0}<25`$ MeV. Fig. 1(a) shows the $`\pi ^+\pi ^{}`$ invariant mass closest to the $`K_S^0`$ mass before the $`K_s^0`$ are selected; a very strong signal $`K_S^0`$ is seen. The number of surviving events is 1095 with 57 $`\pm `$ 5 non-$`K_S^0`$ background under the $`K_S^0`$. For our final fit, we use 683 events below a $`K\overline{K}\pi `$ mass of 2.3 GeV. A constraint to the $`K_S^0`$ vertex does not improve the signal/background ratio further, but loses some events. The effects of the various selection cuts on the data is simulated with a full Monte Carlo of the BES detector including the decay path of the $`K_s^0`$; 250,000 Monte Carlo events are successfully fitted to $`J/\psi \gamma (K^\pm K_S^0\pi ^{})`$. All background reactions are similarly fitted to this channel. The estimated background is $`29\pm 7\%`$, mostly from $`J/\psi \pi ^0(K^\pm K_S^0\pi ^{})`$, some from non-$`K_S^0`$ events. It peaks at about 2.3 GeV, and follows phase space closely. We have included it in the amplitude analysis, but it has little effect, since all genuine signals have a characteristic dependence on either or both of production and decay angles. Fig. 1(b) shows the $`K^\pm K_S^0\pi ^{}`$ mass spectrum; the dark histogram shows the estimated background in the analysis region. There is a conspicuous and somewhat asymmetric peak due to $`\eta (1440)/\iota `$, similar to the earlier data from Mark III, DM2 and BES. At high mass, there is a distinct peak at 2040 MeV. Fig. 2 shows Dalitz plots for three mass ranges: (a) 1360-1560 MeV, (b) 1600–1750 MeV, and (c) 1800–2200 MeV; fits are shown in (d), (e) and (f). There is a conspicuous $`K^{}K`$ decay mode in the first region of the $`\eta (1440)`$. At higher masses, it disappears rapidly, and the mass projections shown in Fig. 3 are consistent with decays to $`\kappa K`$ only; above 1560 MeV, there is no significant evidence for $`a_0(980)K`$, $`a_0K\overline{K}`$. We have carried out a partial wave analysis using amplitudes constructed from Lorentz-invariant combinations of the 4-vectors and the photon polarization for $`J/\psi `$ initial states with helicity $`\pm 1`$. Cross sections are summed over photon polarisations. The relative magnitudes and phases of the amplitudes are determined by a maximum likelihood fit. We include $`K\overline{K}\pi `$ states with quantum numbers $`0^{}`$, $`1^+`$, $`2^{}`$ and $`2^+`$. There are two helicity amplitudes for $`1^+`$, three for $`2^{}`$ and three for $`2^+`$. Because production is via an electromagnetic transition, the same phase is used for different helicity amplitudes to the same final state. Different phases are allowed for different decay channels, e.g $`K^{}K`$ and $`\kappa K`$, because of strong interaction effects due to rescattering. The analysis is discussed separately for the mass region of $`\eta (1440)`$ and the 2040 MeV peak. The $`\eta (1440)`$ has been fitted using a Breit-Wigner amplitude with $`s`$-dependent width: $$f=\frac{\mathrm{\Lambda }}{M^2siM[\mathrm{\Gamma }_{K^{}K}(s)+\mathrm{\Gamma }_{\eta \sigma }(s)+\mathrm{\Gamma }_{\rho \rho }(s)+\mathrm{\Gamma }_{\kappa K}(s)]}.$$ (1) The numerator $`\mathrm{\Lambda }`$ is a complex coupling constant. The $`\mathrm{\Gamma }(s)`$ are taken to be proportional to the available phase space for each channel, evaluated numerically . The $`\eta \pi \pi `$ phase space is taken from $`\eta \sigma `$, the dominant channel, but $`a_0(980)\pi `$ phase space is similar and both are slowly varying over this mass region. The magnitude of each $`\mathrm{\Gamma }`$ is adjusted iteratively so that cross sections integrated over the resonance agree with the branching ratios determined experimentally. The magnitude of $`\mathrm{\Gamma }_{\eta \sigma }`$ has been obtained from BES data on radiative decays to $`\eta \pi \pi `$ . That for $`\mathrm{\Gamma }_{\rho \rho }`$ has been obtained by fitting BES data on radiative decays to $`4\pi `$ , including in the fit $`\eta (1440)`$ and the broad $`\eta (1800)`$. Here $`\eta (1800)`$ refers to the very broad $`0^{}`$ signal ($`\mathrm{\Gamma }1`$ GeV) derived by Bugg and Zou from an analysis of several channels of $`J/\psi `$ radiative decay. Values of $`\mathrm{\Gamma }_{K^{}K}`$ and $`\mathrm{\Gamma }_{\kappa K}`$ are obtained from the present data. In the mass region of the $`\eta (1440)`$, half the $`\kappa K`$ signal comes from the low mass tail of $`\eta (1800)`$ and its constructive interference with $`\eta (1440)`$. That is, if $`\eta (1800)`$ is removed from the fit, the $`\kappa K`$ width of $`\eta (1440)`$ needs to be doubled. Removing the $`\eta (1800)`$ has a significant, but not dramatic, effect on log likelihood, which changes by 4.8 for 2 extra parameters. The $`f_1(1420)`$ is also included in the amplitude analysis, and a small component due to $`f_1(1285)`$. Both optimise close to the masses and widths quoted by the Particle Data Group (PDG) , so we fix them at PDG values. The amplitude analysis distinguishes cleanly between quantum numbers $`1^+`$ and $`0^{}`$ for $`K^{}K`$ decays. If the whole $`\eta (1440)`$ signal is fitted with $`J^P=1^+`$ (optimising its mass and width), log likelihood is worse by 11.4, a significant amount. (Our definition of log likelihood is such that it increases by 0.5 for a one standard deviation change in one parameter). In the earlier analysis of BES data on the $`K^+K^{}\pi ^0`$ final state , a fairly large amplitude was fitted for $`\eta (1800)`$. The smaller background in present data and the wider mass range allow us to show that this component should in fact be rather small. Its effects on the $`\eta (1440)`$ may be replaced with some increase in the total width of that resonance and an increase in its width for decays to $`K^{}K`$. Present results for the fitted widths are shown in Fig. 4 and branching fractions in Table 1. The fit is compared with the $`K\overline{K}\pi `$ mass spectrum by the histogram in Fig. 5. A free fit to the mass gives 1440 MeV. However, $`\eta \pi \pi `$ data give a resonance mass of $`1405\pm 5`$ MeV, according to the summary by the Particle Data Group . The $`s`$-dependent width we use for $`\eta (1440)`$ explains naturally a mass difference of 20 MeV between $`\eta \pi \pi `$ and $`K\overline{K}\pi `$ data; the rapidly increasing phase space for $`K^{}K`$ makes the $`K\overline{K}\pi `$ channel peak higher and explains also the asymmetric shape of the peak, which rises rapidly on the lower side of the peak and falls more slowly on the upper side. A small ($`15`$ MeV) discrepancy remains between the peaks fitted to $`\eta \pi \pi `$ and $`K\overline{K}\pi `$. We adopt a compromise between fitting these data and $`\eta \pi \pi `$ by using a mass of 1432 MeV, but the effect on other conclusions is negligible. Interferences between $`\eta (1440)`$ and the broad $`\eta (1800)`$ depend on their relative phases and can shift the peak in different data sets; so we do not regard this small discrepancy as a matter for concern. Around 1650 MeV, there is some indication for a narrow $`K\overline{K}\pi `$ peak. However, fitting it requires an unreasonably narrow width $`30`$ MeV. An $`s\overline{s}`$ state at this mass has no obvious non-strange partners. If fitted, it is only a two standard deviation effect. Therefore we discard it as a statistical flutuation. Including it has negligible effects on parameters fitted to $`\eta (1440)`$ and the peak at 2040 MeV. We now turn to the latter peak. It cannot be explained by the very broad $`\eta (1800)`$, which has a completely different and much flatter shape, illustrated by the shaded area in Figure 5(b) below. We fit it with a simple Breit-Wigner amplitude of constant width. Its mass and width optimise at $`M=2040\pm 50`$ MeV, $`\mathrm{\Gamma }=400\pm 90`$ MeV. We have tried fits to this peak with resonances having quantum numbers $`0^{}`$, $`1^+`$ and $`2^{}`$; for standard $`q\overline{q}`$ states, one does not expect $`3^+`$ in kaonic channels until 2300 MeV. We find that log likelihood is better for $`0^{}`$ than $`1^+`$ by 16.4. The latter has one additional parameter, so it is a poorer fit by 5.2 standard deviations. If a combination of $`0^{}`$ and $`1^+`$ amplitudes is used, log likelihood improves only by 0.6, and the fitted $`1^+`$ component is very small: 4.4% of $`0^{}`$ in cross section. These results are not sensitive to the $`\eta (1800)`$ contribution: removing it, the distinction between quantum numbers $`0^{}`$ and $`1^+`$ for the 2040 MeV peak remains at a log likelihood difference of 12.9. We have also tried adding or substituting $`2^{}`$. Alone it gives a poor fit, worse in log likelihood than $`0^{}`$ by 27.9. This demonstrates that $`2^{}`$ and $`0^{}`$ are well separated by their distinctively different angular distributions. If it is added freely to the fit, it improves log likelihood by 3.2 for three extra parameters; this cannot be considered significant. Fig. 5 shows magnitudes of components fitted in the amplitude analysis when the 2040 MeV peak is fitted as $`0^{}`$. The slight differences between Figs.5(b) and (d) is due to interferences of $`\eta (1440)`$ and $`\eta (2040)`$ with the broad $`\eta (1800)`$ in Fig. 5(b). Branching fractions for production and decay, including the dominant interferences, are given in Table 2. Values are integrated up to a $`K\overline{K}\pi `$ mass of 2.3 GeV. Decays to $`K^\pm K_S\pi ^{}`$ have a branching ratio 1/3 of all $`K\overline{K}\pi `$ decays. We correct all measured branching ratios by this factor 3, so as to quote branching fractions for all $`K\overline{K}\pi `$ charge states. The overall branching fraction, summed over all final states is $`(6.0\pm 0.4\pm 2.1)\times 10^3`$. We now discuss possible interpretations for the 2040 MeV peak. Our data for $`J/\psi `$ radiative decays to $`\eta \pi ^+\pi ^{}`$ were fitted using an $`\eta (1760)`$ with a width of 250 MeV and an $`\eta _2(1840)`$. The $`\eta (1760)`$ is entirely distinct from $`\eta (1800)`$, which has a much larger width. A possible interpretation is that it is the $`n=3q\overline{q}`$ state. Then the $`\eta (2040)`$ observed here could be its $`s\overline{s}`$ partner. However, the VES collaboration has identified a $`\pi (1800)`$ with curious decay modes to $`f_0(1300)\pi `$, $`f_0(980)\pi `$ and $`K_0(1430)K`$, but not $`\rho \pi `$. There has been speculation that this is an $`I=1`$ hybrid . The $`\eta (1760)`$ would make a natural partner; its decays to $`\eta \sigma `$ and $`a_0(980)\pi `$ are to be expected for a hybrid. It is natural to expect a corresponding $`s\overline{s}g`$ state decaying to $`\kappa K`$ in the $`K\overline{K}\pi `$ channel roughly 200–250 MeV above the peak in $`\eta \pi \pi `$. In $`J/\psi `$ radiative decays, the amplitude for production of $`q\overline{q}`$ states is suppressed by two powers of $`\alpha _s`$, required to couple intermediate gluons to quarks; at 2040 MeV, $`\alpha _s0.41`$. Production of a hybrid will only be suppressed by one power of $`\alpha _s`$ in amplitude. We therefore examine the possible interpretation of $`\eta (1760)`$ as a $`q\overline{q}g`$ hybrid. For a hybrid, the branching fraction expected in the $`K\overline{K}\pi `$ channel is half that for $`\eta \pi \pi `$, since in $`J/\psi `$ decays intermediate gluons couple equally to $`u\overline{u}`$, $`d\overline{d}`$ and $`s\overline{s}`$. If fitted as $`0^{}`$, the branching ratio for the 2040 MeV peak in $`K\overline{K}\pi `$ is $`(2.1\pm 0.1\pm 0.7)\times 10^3`$; this value is obtained after allowing for interferences with $`\eta (1800)`$ and includes the error in the overall normalisation. It is to be compared with the branching ratio for $`\eta (1760)`$ in $`\eta \pi \pi `$ of $`(1.8\pm 0.75)\times 10^3`$ . These values are consistent within the sizable errors with the expectation for hybrids. The magnitude of branching ratio we now fit to $`\eta (1800)\kappa K`$ is $`(0.58\pm 0.03\pm 0.20)\times 10^3`$. Again, the error includes the overall normalisation uncertainty. It compares with $`(1.08\pm 0.45)\times 10^3`$ fitted to $`\eta \pi \pi `$ decays . Within the errors, these values are now consistent with flavour-blind decays of a glueball. In summary, present data contain less background than earlier data on $`J/\psi \gamma (K^+K^{}\pi ^0)`$ and allow a somewhat improved determination of the properties of $`\eta (1440)`$. Its dominant decay mode is to $`K^{}K`$. This suggests it is the first radial excitation of $`\eta (958)`$, probably mixed with the broad $`\eta (1800)`$, in order to account for its strong production in $`J/\psi `$ radiative decays. We now find a small component of $`\eta (1800)`$ decaying to $`\kappa K`$. We observe a peak at 2040 MeV which may be fitted with a $`0^{}`$ resonance of width 400 MeV. $`J^P=0^{}`$ is preferred over $`1^+`$ and $`2^{}`$ respectively by 5.2 and 6.8 standard deviations. Its branching fraction, when compared with the $`\eta \pi \pi `$ channel, would be consistent with interpretation as a $`0^{}`$ $`s\overline{s}g`$ hybrid. The BES group thanks the staff of IHEP for technical support in running the experiment. This work is supported in part by China Postdoctoral Science Foundation and National Natural Science Foundation of China under contract Nos. 19991480, 19825116 and 19605007; and by the Chinese Academy of Sciences under contract No. KJ 95T-03(IHEP). We also acknowledge financial support from the Royal Society for collaboration between Chinese and UK groups.
warning/0002/astro-ph0002136.html
ar5iv
text
# Cosmic Star Formation History Required from Infrared Galaxy Number Count : Future Prospect for Infrared Imaging Surveyor (IRIS) ## 1 Introduction Recent infrared and sub-mm surveys revealed a very steep slope of galaxy number count compared with that expected from no evolution model, and provided a new impetus to the related field. Such excess of galaxy number count is generally understood as a consequence of strong galaxy evolution, i.e. rapid change of the star formation rate in galaxies. Now Japanese infrared satellite project Astro-F (Infrared Imaging Surveyor: IRIS) is in progress, and we calculated the expected number count by a simple empirical method (Takeuchi et al. 1999; Hirashita et al. 1999). The applied model was based on the IRAS surveys, and we need more realistic one to study the detailed observational plans and follow-up strategies. In order to construct the advanced model, we first compiled the infrared/sub-mm SEDs of galaxies obtained by ISO, SCUBA, and other facilities, and derived average SEDs for various classes of galaxies. Then, using the local luminosity function, we studied the required galaxy evolutionary history statistically. ## 2 Model Description We construct the SEDs of galaxies based on the IRAS color–luminosity relation (Smith et al. 1987; Soifer & Neugebauer 1991). For the infrared – sub-mm component, we consider PAH (polycyclic aromatic hydrocarbon), graphite and silicate dust spectra (Dwek et al. 1996). Detailed PAH band emission parameters are taken from Allamandola et al. (1989). For the longer wavelength regime, power-law continuum produced by synchrotron radiation ($`\nu ^\alpha `$) dominates. We set $`\alpha =0.7`$ according to Condon (1992). The SEDs are presented in Fig. 1. We applied the double power-law form (Soifer et al. 1987) for the local luminosity function, and assumed pure luminosity evolution in this study. This is depicted in Fig. 2. The faint-end slope does not affect the result, because the number count is an integrated value along with redshift and volume, and its contribution is small. ## 3 Results and Discussions ### 3.1 Evolutionary history We treat the evolutionary change of galaxy luminosities as a stepwise nonparametric form, in order to explore the most suitable evolutionary history which reproduces the present observational results. Furthermore the constraint from Cosmic Infrared Background (CIRB) should be considered as another observational constraint to the models of evolutionary history. #### 3.1.1 CIRB First, we searched the evolutionary pattern which satisfy the constraint required from CIRB and we found three patterns (Evolution $`13`$) shown in Fig. 3. These three evolutionary patterns satisfy the constraint from CIRB (Fig. 4). In order to satisfy the high background intensity at $`150\mu `$m, the high evolutionary factor at $`z0.8`$ is a mandatory. We note that too large evolutionary factor at high $`z`$ would produce the excess around $`1000\mu `$m. We obtain IR $`60\mu `$m luminosity density along with redshift (Fig. 5) from Fig. 3. Figure 5 show the rapid evolution in $`\rho _\mathrm{L}(60\mu \mathrm{m})`$. The increase is well described by $`(1+z)^5`$ ! We need such a sudden rise of $`\rho _\mathrm{L}(60\mu \mathrm{m})`$ to reproduce the very high CIRB intensity at $`150\mu `$m mentioned before. The peak of the IR luminosity density is located at $`z1`$. #### 3.1.2 number count Recently, new observational results of the galaxy number counts at the mid-infrared and far-infrared (mainly by ISO), and submillimeter (SCUBA and others). We are able to compile these data as well as previously obtained IRAS data and radio data. It is obviously important to calculate multiband number count predictions and compare the multiband observations, because the response of the results to the galaxy evolutionary form varies with different wavelengths. In principle, the galaxy number count is an integrated value along with redshift $`z`$, and the information of the redshift is not available. But the redshift degeneracy can be solved to some extent, by treating the multiband observational results at the same time. We check whether the three evolutionary patterns found in the previous section also satisfy the constraints from observations of number counts. We compare our number counts with observations in Fig. 6 (infrared), and Fig. 7 (sub-mm – radio). Every pattern satisfys the constraints. When we especially focus on the sub-mm number counts, the evolution 3 is the most desirable. ### 3.2 Infrared Imaging Surveyor (IRIS) Infrared Imaging Surveyor (IRIS) is a satellite which will be launched in 2003, by the M-V rocket of the ISAS (the Institute of Space and Astronautical Science in Japan). One of the main purposes of the IRIS mission is an all-sky survey at far-infrared (FIR) with a flux limit much deeper than that of IRAS. Detailed information of IRIS is available at http://koala.astro.isas.ac.jp/Astro-F/index-e.html. In order to examine the performance of the survey, we estimated the FIR galaxy counts in two narrow bands (i.e. N60 and N170) based on models described in the previous section. We found that a large number of galaxies ( $`10^7`$ in the whole sky) will be detected in this survey (Fig. 8). ###### Acknowledgements. We wish to thank Dr. Hiroshi Shibai, Dr. Izumi Murakami and Dr. Hideo Matsuhara for helpful discussions. Dr. Kimiaki Kawara also deserves our thanks for providing us their number count in the Lockman Hole for reference. TTI is grateful to Dr. Hiroki Kurokawa for continuous encouragement. TTT, HH, and KY acknowledge the Research Fellowships of the Japan Society for the Promotion of Science for Young Scientists. We are grateful to all the participants who gave us useful suggestions and comments at the Symposium.
warning/0002/nucl-th0002043.html
ar5iv
text
# A hybrid version of the tilted axis cranking model and its application to 128Ba ## I Introduction The transitional nuclei in the region $`A=130140`$ show regular $`\mathrm{\Delta }I=1`$ regular bands, characterized by large $`B(M1)/B(E2)`$ ratios, the lack of signature splitting and relatively low dynamical moments of inertia . These bands have an intermediate character. They stand between the collective high-$`K`$ bands of well deformed nuclei, for which collective rotation is the dominant mechanism of generating the angular momentum , and the magnetic rotation of near spherical nuclei, for which few high-$`j`$ particles and holes generate most of the angular momentum by means of the shears mechanism (see, for example ). The question of how changes magnetic into collective rotation has not been studied yet. So far, only the magnetic dipole band in <sup>128</sup>Ba has been investigated from this point of view. Another intriguing question is the possibility of a chiral character of rotation . Their softness with respect to triaxial deformations makes the nuclei in the $`A=130`$ region particularly good candidates for identifying this new symmetry type. The Tilted Axis Cranking (TAC) model has turned out to be an appropriate theoretical tool for the description of the magnetic dipole bands. This model is a natural generalization of the cranking model for situations where the axis of rotation does not coincide with a principal axis of the density distribution of the rotating nucleus, and thus the signature is not a good quantum number. Since introduced, TAC has proven to be a reliable approximation for the energies and intraband transitions in both normally and weakly deformed nuclei . However, in the case of the four quasiparticle magnetic dipole band in <sup>128</sup>Ba, the TAC calculations predicted the wrong parity and a too early termination of the band and provided only fair description of the electro magnetic transition data . <sup>128</sup>Ba is one of the best studied nuclei in this mass range and the above mentioned $`\mathrm{\Delta }I=1`$ rotational band is a good test case for the TAC model. The purpose of the present work is to try to identify the origin of the discrepancies and remove them. The version of TAC used in was based on the Nilsson Hamiltonian. The parameters of this well-tried potential were carefully adjusted for various mass regions, where it was successfully used in standard cranking calculations as well as in TAC calculations . However, the $`A=130`$ region is known to be problematic for the model. In fact no really satisfactory set of Nilsson parameters for this region is available so far. Thus we attribute the discrepancies of the TAC calculations for <sup>128</sup>Ba with the later measurements to the general problems of the Nilsson potential in this region. On the other hand, the Woods-Saxon potential works very well around $`A=130`$ . Encouraged by this, we adapt the Nilsson potential as close as possible to the Woods-Saxon one. We call this the hybrid potential, which is the basis for new TAC calculations for <sup>128</sup>Ba. Instead of the parameterizing the single particle levels of the spherical modified oscillator in the standard way by means of an $`ls`$\- and an $`l^2`$-term, the hybrid model directly takes the energies of the spherical Woods-Saxon potential. The deformed part of the hybrid potential is an anisotropic harmonic oscillator. This compromise keeps the simplicity of the Nilsson potential, because coupling between the oscillator shell can be approximately taken into account by means of stretched coordinates , and it amounts to a minor modification of the existing TAC code. On the other hand, it has turned out to be a quite good approximation of the realistic flat bottom potential as long as the deformation is moderate. The hybrid was used to calculate the triaxial shapes of liquid sodium clusters . The results agree very well with later calculations using the correct radial profile of the deformed part of the potential. ## II TAC implementation The TAC model is discussed in more detail in . The brief presentation in this paper focuses at the suggested improvement of this approach. The starting point is the mean field Routhian<sup>*</sup><sup>*</sup>*For simplicity, only one type of particles is spelled out. The extension to both types is obvious. $$h^{}=h_{sph}+V_{def}(\epsilon _2,\gamma ,\epsilon _4)\mathrm{\Delta }\left(P+P^+\right)\lambda N\omega \left(\mathrm{sin}\vartheta j_1+\mathrm{cos}\vartheta j_3\right)$$ (1) where $`h_{sph}`$ denotes the spherical part including the spin-orbit term and $`V_{def}(\epsilon _2,\gamma ,\epsilon _4)`$ the deformed part of the Nilsson single-particle Hamiltonian. (see, e.g. ). The pairing field in eq.(1) is determined by the gap parameter $`\mathrm{\Delta }`$ and the monopole pairing operator $`P=_kc_kc_{\overline{k}}`$ while the chemical potential $`\lambda N`$ is needed to satisfy on average the particle number conservation. The two-dimensional cranking term $`\omega \left(\mathrm{sin}\vartheta j_1+\mathrm{cos}\vartheta j_3\right)`$ is the new element of the TAC, as compared to the standard cranking model, which is recovered for $`\vartheta `$ = 0 or 90. The angle $`\vartheta `$ fixes the tilt of the cranking axis with respect to the intrinsic 3-axis within the principal (1-3) plane of the deformed potential. By diagonalization in stretched coordinates, neglecting $`\mathrm{\Delta }N=2`$ shell mixing, this TAC Routhian yields quasi-particle energies and quasi-particle states, from which the many-body configuration $`|\omega ,\epsilon _2,\gamma ,\epsilon _4,\vartheta >`$ of interest is constructed. The mean field is found for a given frequency $`\omega `$ and fixed configuration by minimizing the total Routhian $$E^{}(\omega ,\epsilon _2,\gamma ,\epsilon _4,\vartheta )=<\omega ,\epsilon _2,\gamma ,\epsilon _4,\vartheta |h^{}|\omega ,\epsilon _2,\gamma ,\epsilon _4,\vartheta >$$ (2) with respect to the deformation parameters $`(\epsilon _2,\gamma ,\epsilon _4)`$ and the tilt angle $`\vartheta `$. The value of $`\mathrm{\Delta }`$ is kept fixed at two values: 80% of the experimental odd-even mass difference and zero. At the equilibrium angle $`\vartheta =\vartheta _{}`$ (minimum) the cranking axis is parallel to the direction of the angular momentum vector $`\stackrel{}{J}=(<j_1>,<j_3>)`$. After the minimum is found the various electro magnetic observables of interest are obtained by means of the following semiclassical expressions $`<I2I2|_2(E2)|II>=<_2(E2)>=`$ (3) $`=\sqrt{{\displaystyle \frac{5}{4\pi }}}\left({\displaystyle \frac{eZ}{A}}\right)\left[\sqrt{{\displaystyle \frac{3}{8}}}<Q_0^{}>(\mathrm{sin}\vartheta )^2+{\displaystyle \frac{1}{4}}<Q_2^{}+Q_2^{}>\left(1+(\mathrm{cos}\vartheta )^2\right)\right],`$ (4) $`<I1I1|_1(E2)|II>=<_1(E2)>=`$ (5) $`=\sqrt{{\displaystyle \frac{5}{4\pi }}}\left({\displaystyle \frac{eZ}{A}}\right)\left[\mathrm{sin}\vartheta \mathrm{cos}\vartheta (\sqrt{{\displaystyle \frac{3}{2}}}<Q_0^{}>{\displaystyle \frac{1}{2}}<Q_2^{}+Q_2^{}>)\right].`$ (6) $`<I1I1|_1(M1)|II>=<_1(M1)>=`$ (7) $`=\sqrt{{\displaystyle \frac{3}{8\pi }}}\left[\mu _3\mathrm{sin}\vartheta \mu _1\mathrm{cos}\vartheta \right],`$ (8) where $`<_\nu >`$ is the expectation value of the transition operator with the TAC configuration $`|>`$, the components of which refer to the lab system. The intrinsic quadrupole moments $`Q_\mu ^{}`$ are calculated with respect to the principle axes (1, 2, 3). The same holds for the magnetic moments $$\mu _1=\mu _N(J_{1,p}+(\eta 5.581)S_{1,p}\eta 3.82S_{1,n}),\mu _3=\mu _N(J_{3,p}+(\eta 5.581)S_{3,p}\eta 3.82S_{3,n}),$$ (9) where the free nucleonic magnetic moments are attenuated by a factor of $`\eta =0.7`$. The reduced transition probabilities are $`B(M1,\mathrm{\Delta }I`$ $`=`$ $`1)=<_1(M1)>^2`$ $`B(E2,\mathrm{\Delta }I`$ $`=`$ $`2)=<_2(E2)>^2`$ The mixing ratio is $$\delta =\frac{<_1(E2)>}{<_1(M1)>}.$$ (10) We apply the Strutinsky renormalization procedure to calculate the total Routhian $$E^{}(\omega )=E_{LD}(\omega =0)E_{smooth}+<\omega |h^{}|\omega >,$$ (11) where $`E_{LD}=E_{LD}(\epsilon _2,\gamma ,\epsilon _4,\epsilon _4)`$ means the liquid drop energy and $`E_{smooth}`$ is the the smooth part of the mean-field energy calculated from the single-particle energies at $`\omega =0`$. This version of the TAC has turned out to be quite successful for well deformed nuclei (see and references therein). The new element of the present paper consists in the hybrid potential, which approximates the well-established deformed Woods-Saxon potential, yet preserving the existing convenient TAC environment. For this purpose the spherical part $`h_{sph}`$ in Eq.(1) is replaced the spherical Woods-Saxon Hamiltonian for the nucleus of interest. In the present work the universal Woods-Saxon parameters are used (see e.g. ). Technically, the replacement is rather simple, because the existing TAC code uses states of good $`l,j,m`$ as a basis. The spherical Nilsson energies $`e_{N,l,j}^{(nil)}`$ are replaced by the spherical Woods - Saxon energies $`e_{N,l,j}^{(ws)}`$. It turns out to be unproblematic to associate the quantum numbers of the two different potentials. For a given combination $`l,j`$ the third quantum number $`N`$ is found by counting from the state with the lowest energy. The fact that the spherical Woods-Saxon code uses a harmonic oscillator basis permitted a check of the algorithm. The major component of the Woods-Saxon wavefunction agrees with the state found by our counting algorithm. In the high-lying part of the single-particle spectrum (three shells above the valence shell or higher) there are occasional ambiguities in assigning the states. Small errors of this kind are not expected to have any consequences at moderate or small deformation. The states do not couple strongly to the states near the Fermi surface. They contribute only to the smooth level density used in the Strutinsky renormalization, which will not be affected by small shifts of the levels. Such a replacement of the spherical single-particle energies is a common practice in large scale shell-model configuration mixing and similar calculations, where they are often used as adjustable parameters . The effect of the replacement is illustrated on Fig.1, which shows the deformation dependence of the proton single-particle levels of <sup>128</sup>Ba for the Nilsson, the Woods-Saxon and the hybrid Hamiltonians. The close similarity of the levels of the Woods-Saxon and the hybrid models is obvious. The hybrid has a somewhat later and sharper crossing between the positive parity levels, which also show stronger tendency to arrange into pairs of pseudo spin doublets. These treats are inherited from the Nilsson Hamiltonian, which controls the change with deformation. The main difference between the Nilsson Hamiltonian and the other two is the lower energy of the negative parity levels originating from $`h_{11/2}`$. It seems to be the reason for the discrepancies between the previous TAC predictions and the experiment, as will be demonstrated in the next section. ## III The M1 band in <sup>128</sup>Ba In the previous TAC calculations for the $`\mathrm{\Delta }I=1`$ band in <sup>128</sup>Ba the four quasi-particle configurations $`[\pi (h_{11/2})^2\nu (h_{11/2}(d_{5/2}g_{7/2}))]`$ Indicating the major components, we denote the mixed Nilsson state by $`(d_{5/2}g_{7/2})`$. for the negative parity and $`[\pi (h_{11/2})^2\nu (h_{11/2})^2]`$ for the positive parity were found to be the lowest ones in energy at the oblate deformation of $`\epsilon =0.26`$, $`\gamma =60^o`$. This deformation was determined at $`\omega =0.2`$ MeV by minimizing the total Routhian calculated from the Quadrupole-Quadrupole interaction. With the present version of TAC we find a significantly smaller prolate equilibrium deformation of $`\epsilon =0.205`$ and $`\gamma =0^o`$. The lowest four quasi-particle configuration now turns out to be $`[\pi (h_{11/2}(d_{5/2}g_{7/2}))\nu (h_{11/2}(d_{5/2}g_{7/2}))]`$. Fig.2 shows the quasi proton and quasi neutron levels for various cranking frequencies at tilt angle $`\vartheta =90^o`$ and for various angles at $`\omega =0.2`$ MeV. The equilibrium value of $`\vartheta `$, which minimizes $`E^{}`$, is found to be $`\vartheta _{}=52.5^o`$ for $`\omega =0.2`$ MeV. As seen, the different position of the $`h_{11/2}`$ orbitals in the hybrid TAC has drastic consequences. The deformation changes from oblate to prolate, resulting in a different configuration of the M1 band. This is not surprising in a region where the energy difference between oblate and prolate shape is small. The calculations of the $`\mathrm{\Delta }I=1`$ band in the present work are built on this new configuration $`[\pi (h_{11/2}(d_{5/2}g_{7/2}))(\nu h_{11/2}(d_{5/2}g_{7/2}))]`$, which is the lowest four-quasi particle TAC solution. In agreement with the experiment, it has positive parity. The excitation of four quasi-particles significantly reduces the pairing gaps. In order to better grasp the influence of this blocking on characteristics of the band, we did two calculations: one with $`\mathrm{\Delta }_\nu =0.88`$ MeV and $`\mathrm{\Delta }_\pi =1.04`$ MeV corresponding to 80% of the experimental even odd mass difference and one with zero pairing. Electro magnetic transition properties present a stringent test of the nuclear models. The experimental information about lifetimes and mixing ratios of the magnetic dipole band was accumulated in several experiments . Fig.3 compares the experimental $`B(M1)`$ and $`B(E2)`$ values and mixing ratios with the results of TAC model obtained for the configuration $`[\pi (h_{11/2}(d_{5/2}g_{7/2}))\nu (h_{11/2}(d_{5/2}g_{7/2}))]`$. All electro-magnetic characteristics of the band are well reproduced by both the paired and unpaired calculations. Curiously, the quenching of pairing influences to some extent the $`B(M1)`$ values but leaves almost unchanged the B(E2) values up to spin 18$`\mathrm{}`$. The TAC calculation seems to slightly overestimate the deformation. In contrast with the previous TAC calculation, it is possible to follow the band all the way up to spin 26$`\mathrm{}`$. Fig.4 shows the measured and the calculated function $`J(\omega )`$ of the spin on the angular frequency. It is more sensitive to the changes of the pair correlations. While the unpaired calculation gives a nearly linear function with the moment of inertia $`𝒥^{(2)}=dJ/d\omega `$ close to the measured one, the paired calculation exhibits a substantially lower moment of inertia for low rotational frequencies and an upbend at higher ones. The experiment is in between. This seems to indicate that the pair field is weak in this nucleus and a more refined treatment of pairing is needed. It is noted that in the calculations the band extends down to $`\omega =0.1MeV`$ and $`J=11\mathrm{}`$. In experiment there is an irregularity around $`I=12,13\mathrm{}`$. It may be caused by mixing of the $`13^+`$ state with the another $`13^+`$ state, which lies nearby in energy ($`\mathrm{\Delta }E=0.045MeV`$) and into which the $`14^+`$ also decays . In , the $`\mathrm{\Delta }I=1`$ band was analyzed in terms of a pure high-K band using the familiar expressions for the $`B(M1)`$ and $`B(E2)`$ values for the axial symmetric rotor . Adjusting three free parameters, the K-value, the intrinsic quadrupole moment $`Q_{}`$, and the gyromagnetic factor $`|g_Kg_R|`$, a good fit of the electro magnetic decay data was obtained. The quality is practically the same as in our calculation without parameters in Fig.4. The TAC calculation contains much more physical information as e.g. the specific configuration on which the band is built and the band energies. The knowledge of the intrinsic state can be used to derive further structure information, e. g. the geometrical coupling scheme shown in Fig.5, which enables one to see how the total spin is formed from the quasiparticle orbitals and how it changes with the rotational frequency. Apparently, most of the angular momentum gain along the band is of collective nature, however the high-j quasiparticles from proton and neutron $`(d_{5/2}g_{7/2})`$ and $`h_{11/2}`$ orbitals do also substantially contribute by means of the shears mechanism. With increasing frequency, the 3-component of the spin vector $`\stackrel{}{J}`$ stays practically at $`<J_3>`$ 9$`\mathrm{}`$. However, this does not mean that the corresponding TAC configuration behaves like a structureless the high-K rotor. Fig.6 shows that the calculated dependence $`\vartheta (\omega )`$ of the tilt angle on the rotational frequency by no means follows curve expected for the strong coupling limit. Thus, the present case lies in between a good shears band and a good high-K band. The $`\mathrm{\Delta }I=1`$ band in <sup>128</sup>Ba is an example for a rotational band of intermediate nature. ## IV Summary The strength of the TAC model is that it can predict the appearance of $`\mathrm{\Delta }I=1`$ rotational bands and is able to describe microscopically their electro magnetic decay properties. This is achieved by taking into consideration the orientation of the rotational axis with respect to the deformed potential, which is fixed along a principal axis in the conventional cranking model. The intrinsic TAC configuration of a rotational band is found by searching for a local minimum on the multi-parameter surface of the total Routhian. Therefore, it is crucial to calculate these surfaces as reliable as possible. In the present work the Strutinsky renormalization and a hybrid single-particle potential were implemented in order to improve the total Routhian. The hybrid potential combines the spherical Woods-Saxon single-particle energies with the deformed part of the Nilsson potential. For moderately deformed nuclei it gives deformed single-particle levels that are quite close to the deformed Woods-Saxon levels. We applied the hybrid TAC to the previously investigated four quasiparticle magnetic dipole band in <sup>128</sup>Ba. The lowest equilibrium configuration is found to be $`[\pi (h_{11/2}(d_{5/2}g_{7/2}))\nu (h_{11/2}(d_{5/2}g_{7/2}))]`$. It has positive parity and a prolate axial deformation of $`\epsilon _2=0.20`$. The microscopic TAC calculations describes rather well the experimental energies, $`B(M1)`$ and $`B(E2)`$ values, as well as the branching and mixing ratios. The dipole band in <sup>128</sup>Ba has an intermediate structure. A comparable amount of angular momentum is generated by the shears mechanism, active for the $`h_{11/2}`$ and $`(d_{5/2}g_{7/2})`$ quasiparticles, and by collective rotation. It is an example for the transition from collective to magnetic rotation. The hybrid potential turned out to be crucial for the good agreement between the calculation and the data. It substantially improves the Nilsson potential, which has problems in the region around mass 130. It seems to be a promising starting point for studying the intriguing interplay between triaxial deformation and the orientation of the rotational axis. ## V Acknowledgments V.D. wishes to thank the Foundation ”Bulgarian Science and Culture” for its support. We should like to thank Prof. P. von Brentano and Dr. I. Wiedenhöver, who kept us up with the data. The work was partially carried out under the Grant DE-FG02-95ER40934.
warning/0002/quant-ph0002047.html
ar5iv
text
# Continuous pumping and control of mesoscopic superposition state in a lossy QED cavity ## I Introduction Along the last two decades a consensus has been established about the importance of the effects of the environment on a macroscopic system to explain the non-observation of superposition of quantum states . The formal treatment of a non-isolated macroscopic quantum system of interest assumes a unitary evolution of the whole system composed by system of interest \+ environment \+ (possibly) measurement apparatus. In the dynamical process called decoherence, the environment drives the macroscopic superposition state into a statistical mixture in a very short time, as compared to the relaxation time . As a matter of fact, even microscopic systems suffer from the effects of the environment since they are not perfectly isolated, however less drastically. The construction of mesoscopic superposition states of the electromagnetic field (EMF) in a cavity (cat states) has attracted attention due to the experimental observation of its very short lifetime as a superposition. The most recent proposals for preparing mesoscopic superposition states relies on strategies whose aim consists in keeping them in a high degree of purity, by precluding the noise coming from the reservoir, in order to delay the decoherence process. One proposal for creating and sustaining cat states is based on the confinement of an EMF in a superconducting cavity . In the authors proposed and achieved an experiment consisting of the preparation of a cat state (a superposition of coherent states in the microwave region) in a Fabry-Perot open superconducting cavity of high quality factor then measuring its decoherence time using the interaction with a beam of two-level Rydberg atoms with an electromagnetic field going through the cavity . In such an experiment energy and information loss are unavoidable, not permitting the existence of the cat state for a sufficiently long time, so constituting a drawback for its use for technological purposes; the cavity field ends up in a thermal state . Other proposals for suppressing the action of the environment on the coherence of mesoscopic superposition field states have been presented in the last years. One of them considers a stroboscopic feedback scheme : a stream of two-level atoms, all prepared in the same state, cross a cavity where each atom interacts dispersively with the field. If the delay time between sequential atoms is short enough, all atoms are to be detected in the same state, e.g., the excited state. This is an indirect measure of the coherence and parity of the initial field superposition state. However, when a cavity photon is lost, the following atom will be detected in the ground state. When this event occurs a subsequent atom prepared in the excited state should be sent to interact in resonance with the field in order to compensate the lost energy and phase, so hopping to restore its original state by absorbing a photon from the atom. This procedure needs a full mastering of the atom-field interactions by the experimenter. However, due to the poor efficiency of the atomic detectors, the stroboscopic feedback scheme looses its full reliability. In order to reduce the velocity of the decoherence process of a cat state in a lossy QED cavity (hereby referred as C), in this paper another practical strategy is proposed. It considers the continuous action of a classical pumping field - a single mode microwave signal - in C, during the running time of the experiment. We begin by showing that under the action of pumping and at temperature $`T=0`$K, an arbitrary initial state of the field in C goes asymptotically to a coherent state. The pumping action compensates the energy lost to the environment, but not the initial available information about the state (interference of probability amplitudes or coherence) as it is not sensitive to the phase information of the field state. This can only be achieved with the combined action of pumping together with the injection of Rydberg atoms through the cavity, which permits sustaining the energy of the field and reconstructing the initial coherence. This process can also use feedback atoms, as proposed in , to guarantee a full efficiency for maintaining the initial cat state. Another important question raised in the present paper: For an open system constantly fed by an external source how does evolve the decoherence and relaxation process? This paper is organized as follows: In Section II we review the mechanism for generating superposition field states in superconducting cavities. Section III is devoted to obtain of the Heisenberg equations for the field operators that govern the evolution of the continuously pumped quantum state. In Section IV we discuss how a cat state is generated in a cavity and how it evolves when the action of the combined pumping field plus environment proceeds. Section V is dedicated to the study of the decoherence process of the cavity state. In Section VI we propose a strategy which combines the action of atoms and pumping to restore the initial superposition of the field state and finally in Section VII we present a summary of this work. ## II Generation of Schrödinger ‘cat’ states The experimental apparatus for the generation of field superposition states consist of a beam of Rydberg atoms crossing three cavities, R<sub>1</sub>, R<sub>2</sub> and C. R<sub>1</sub> and R<sub>2</sub> are low quality cavities (Ramsey zones), but C is a high-Q superconducting cavity, where a coherent state was previously injected by a microwave source. The atoms are initially prepared in circular states of quantum principal number of the order of 50 which are well designed for these experiments since their life time is over $`3\times 10^2`$s . The usual method of Ramsey interferometry consists in injecting classical fields in the Ramsey zones R<sub>1</sub> and R<sub>2</sub> during the interaction time with the atoms . The transition between two nearly orthogonal atomic states, $`|e`$ (excited) and $`|g`$ (ground), is resonant with the R<sub>1</sub> and R<sub>2</sub> fields, and the transition strength is set by selecting the velocity of the atom, which suffers a rotation in the space spanned by state vectors $`\{|e,|g\}`$. The experiment begins by preparing the Rydberg atom in state $`|e`$, which is then rotated in R<sub>1</sub> to the superposition state $$|\mathrm{\Psi }_a=\frac{1}{\sqrt{2}}\left(|e+|g\right).$$ (1) Subsequently the atom interacts with the field in $`C`$, whose dynamics is described by the Jaynes-Cummings Hamiltonian $$H=\mathrm{}\omega a^{}a+\frac{1}{2}\mathrm{}\omega _0\sigma _z+\mathrm{}\kappa \left(a\sigma ^++a^{}\sigma ^{}\right)$$ (2) where $`\sigma _z|ee||gg|`$, $`\sigma ^+|eg|`$ and $`\sigma ^{}|ge|`$ are atomic pseudo-spin operators, $`a`$ ($`a^{}`$) is the annihilation (creation) operator for the field mode of frequency $`\omega `$ in $`C`$, $`\kappa `$ is the atom-field coupling constant and $`\omega _0`$ is the atomic transition frequency. The cavity $`C`$ is tuned near resonance with the atomic transition frequency $`\omega _i`$, between states $`|e`$ and $`|i`$, where $`|i`$ is a reference state with energy level above that of $`|e`$. The transition frequency $`\omega _i`$ is distinct of any other one involving the state $`|g`$. The mode geometry inside the cavity is such that the intensity of the field increases and decreases smoothly along the atomic trajectory inside $`C`$. For sufficiently slow atoms and for sufficiently large detuning between $`\omega `$ and $`\omega _i`$, the atom-field evolution is adiabatic and no photonic absorption or emission occurs . However, dispersive effects are very important \- the atom crossing $`C`$ in the state $`|e`$ induces a phase shift in the cavity field which can be adjusted by a suited selection of the atomic velocity ($`100`$ m/s). For a phase shift $`\pi `$, a coherent field $`|\alpha `$ in $`C`$ is turned into $`|\alpha `$. On the other hand, an atom in state $`|g`$ does not introduce any phase shift on the cavity field. Therefore, an atom in state (1) crossing $`C`$ will lead the system C \+ atom in the correlated state $$\frac{1}{\sqrt{2}}\left(|e+|g\right)|\alpha \frac{1}{\sqrt{2}}\left(|e|\alpha +|g|\alpha \right).$$ (3) The atom crosses the cavity in a time of order of 10<sup>-4</sup> s, which is well bellow the relaxation time of the field inside $`C`$ (typically 10<sup>-3</sup>-10<sup>-2</sup>s for Niobium superconducting cavities) and bellow the atomic spontaneous emission time (3$`\times 10^2`$s) . When the atom is submitted to a second $`\pi /2`$ pulse, in R<sub>2</sub>, the total state will be transformed as $$\frac{1}{\sqrt{2}}\left(|e|\alpha +|g|\alpha \right)\frac{1}{\sqrt{2}}\left[|e\frac{1}{\sqrt{2}}\left(|\alpha |\alpha \right)+|g\frac{1}{\sqrt{2}}\left(|\alpha +|\alpha \right)\right].$$ (4) Therefore, if the atom is detected in the state $`|g`$ or $`|e`$ the field in $`C`$ will be projected to the state $$|\mathrm{\Psi }_c=\frac{1}{N}\left(|\alpha +\mathrm{cos}\phi |\alpha \right),$$ (5) with $`\phi =0`$ ($`\pi `$) if the atom is detected in the state $`|g`$ ($`|e`$), $`N=\sqrt{2\left(1+\mathrm{cos}\phi \text{e}^{2\left|\alpha \right|^2}\right)}`$ is thenormalization constant and the density operator for the superposition state (5) is given by $$\rho _C=|\mathrm{\Psi }_c\mathrm{\Psi }_c|=\frac{1}{N^2}\left\{|\alpha ><\alpha |+|\alpha ><\alpha |+\mathrm{cos}\phi \left(|\alpha \alpha |+|\alpha \alpha |\right)\right\}.$$ (6) When such a state is produced inside the cavity, the presence of dissipative effects alters its free evolution, introducing an amplitude damping as well as a coherence loss term. At temperature $`T=0`$K the density operator (6) evolves according to $`\rho _C(t)`$ $`=`$ $`{\displaystyle \frac{1}{N^2}}\{|\alpha \text{e}^{\gamma t/2}><\alpha \text{e}^{\gamma t/2}|+|\alpha \text{e}^{\gamma t/2}><\alpha \text{e}^{\gamma t/2}|+\mathrm{cos}\phi \text{e}^{2|\alpha |^2(1\text{e}^{\gamma t})}`$ (8) $`\times [|\alpha \text{e}^{\gamma t/2}><\alpha \text{e}^{\gamma t/2}|+|\alpha \text{e}^{\gamma t/2}><\alpha \text{e}^{\gamma t/2}|]\},`$ where two characteristic times are involved. The first one, the decoherence time is the time in which the pure state Eq.(8) is turned into a statistical mixture $$\rho _C(t)\frac{1}{2}\left\{|\alpha \text{e}^{\gamma t/2}><\alpha \text{e}^{\gamma t/2}|+|\alpha \text{e}^{\gamma t/2}><\alpha \text{e}^{\gamma t/2}|\right\};$$ (9) the other one is the damping or relaxation time of the field, $`t_c`$ =$`\gamma ^1`$, being the characteristic time when the energy dissipation becomes important, driving the field asymptotically to a vacuum state. The decoherence phenomenon is characterized by the factor $`\mathrm{exp}\left[2\left|\alpha \right|^2\left(1\text{e}^{\gamma t}\right)\right]`$, and for short times, $`\gamma t1`$, it turns to be $`\mathrm{exp}\left[2\left|\alpha \right|\gamma t\right]`$. The decoherence time $`t_d=\left(2\gamma \left|\alpha \right|^2\right)^1`$ will be called for future reference, the free decoherence time. ## III Theory of classical pumping of lossy cavities We are going to show how a stationary coherent field state is generated in cavities by the action of continuous pumping and how this can change the decoherence process due to the energy loss. In the experimental apparatus discussed in the last section the pumping consists in maintaining the microwave radiation in C during the experimental running time. A single EM mode in C interacts with the reservoir modes, represented by a vast number of harmonic oscillators , accounting for the energy dissipation of the field in C. In the rotating wave approximation the total Hamiltonian is $$H=\mathrm{}\omega _0a^{}a+\underset{k}{}\mathrm{}\omega _kb_k^{}b_k+\mathrm{}\underset{k}{}\left(\lambda _ka^{}b_k+\lambda _k^{}ab_k^{}\right)+\mathrm{}\left[F\text{e}^{i\omega t}a^{}+F^{}\text{e}^{i\omega t}a\right]$$ (10) where $`\omega _0`$ is the mode frequency of the cavity, $`\omega _k`$ is the frequency of the $`k`$-th mode of the reservoir, $`\lambda _k`$ is the field-reservoir coupling constant and $`F`$ is the coupling constant between the cavity and pumping fields, proportional to the pumping field amplitude. Operators $`a^{}`$ ($`a`$) and $`b_k^{}`$ ($`b_k`$) are the bosonic creation (annihilation) operators of the field mode and the reservoir, respectively. Let us suppose that initially the quantum field and reservoir are uncoupled, $$|\mathrm{\Psi }_T;t=0|\psi _F|\varphi _R,$$ (11) where $`|\psi _F`$ is the state of the field and $`|\varphi _R`$ is the state of the reservoir. The Heisenberg equations for $`a`$ and $`b_k`$ are given by $`\dot{a}`$ $`=`$ $`{\displaystyle \frac{1}{i\mathrm{}}}[a,H]=i\omega _0ai{\displaystyle _k}\lambda _kb_kiF\text{e}^{i\omega t},`$ (12) $`\dot{b}_k`$ $`=`$ $`{\displaystyle \frac{1}{i\mathrm{}}}[b_k,H]=i\omega _kb_ki\lambda _k^{}a,`$ (13) and the formal solution to Eq. (13) is $$b_k(t)=\text{e}^{i\omega _kt}b_k(0)i\lambda _k^{}_0^ta(t^{})\text{e}^{i\omega _k(t^{}t)}𝑑t^{}.$$ (14) The rapid oscillation of the free field evolution can be eliminated by introducing in Eq. (12) the operator of slow variation in time $`A\text{e}^{i\omega _0t}a,`$ whose equation of motion is $$\dot{A}=i_k\lambda _kb_k\text{e}^{i\omega _0t}iF\text{e}^{i\omega _0t}\text{e}^{i\omega t}.$$ (15) Substituting Eq. (14) into Eq. (15) we get an equation for $`A`$ only, $$\dot{A}=i_k\lambda _kb_k(0)\text{e}^{i(\omega _k\omega _0)t}\underset{k}{}\left|\lambda _k\right|^2_0^tA(t^{})\text{e}^{i\left(\omega _k\omega _0\right)(t^{}t)}𝑑t^{}iF\text{e}^{i\omega _0t}\text{e}^{i\omega t}.$$ (16) Using the Wigner-Weisskopf approximation into the above equation (see details of calculations in Ap. A) and after some algebraic manipulation the solution of the Heisenberg equation for the operator $`a`$ writes as $$a(t)=u(t)a(0)+\underset{k}{}v_k(t)b_k(0)+w(t),$$ (17) where $$u(t)=\text{e}^{\frac{\gamma }{2}t}\text{e}^{i\omega _0t},$$ (18) $$v_k(t)=\lambda _k\text{e}^{i\omega _kt}\frac{\left[1\text{e}^{\frac{\gamma }{2}t}\text{e}^{i\left(\omega _k\omega _0\right)t}\right]}{\omega _0\omega _ki\frac{\gamma }{2}},$$ (19) and $$w(t)=F\text{e}^{i\omega t}\frac{\left[1\text{e}^{\frac{\gamma }{2}t}\text{e}^{i\left(\omega \omega _0\right)t}\right]}{\omega \omega _0+i\frac{\gamma }{2}},$$ (20) $`\gamma `$ (defined in Ap. A) is the damping constant. ### A Characteristic function and field state representation Any density operator can be spanned by the overcomplete basis of the coherent states having associated a Glauber-Sudarshan $`P`$-distribution, $$\rho (t)=d^2\gamma P(\gamma ;t)|\gamma \gamma |.$$ (21) The normal ordered characteristic function (CF) associated to $`\rho (t)`$ is given by $$\chi _N(\eta ,t)=\mathrm{Tr}\left[\rho (t)\text{e}^{\eta a^{}}\text{e}^{\eta ^{}a}\right]=\mathrm{Tr}\left[\rho (0)\text{e}^{\eta a^{}(t)}\text{e}^{\eta ^{}a(t)}\right],$$ (22) where the term in the middle is written in the Schrödinger picture and the last one is in the Heisenberg picture. The $`P`$ Glauber-Sudarshan distribution is related to the normal ordered CF by a double Fourier transform (FT) $$P(\gamma ;t)=\frac{1}{\pi ^2}d^2\eta \text{e}^{\gamma \eta ^{}\gamma ^{}\eta }\chi _N(\eta ,t),$$ (23) whereas the Wigner function is defined as a double Fourier transform of the symmetric ordered CF by $$W(\zeta ;t)=\frac{1}{\pi ^2}d^2\eta \text{e}^{\zeta \eta ^{}\zeta ^{}\eta }\chi _S(\eta ,t).$$ (24) Both CF’s are related through $$\chi _S(\eta ,t)\mathrm{Tr}\left[\rho (t)\text{e}^{\eta a^{}\eta ^{}a}\right]=\text{e}^{\frac{1}{2}\left|\eta \right|^2}\chi _N(\eta ,t).$$ (25) Substituting Eq. (25) into Eq. (24) and using the inverse FT of Eq. (23), we relate the Wigner function to the P-distribution as $$W(\zeta ;t)=\frac{2}{\pi }d^2\gamma \text{e}^{2\left|\zeta \gamma \right|^2}P(\gamma ,t).$$ (26) The symmetric ordered CF associated to the $`\omega _0`$ mode in cavity $`C`$ is given, in the Heisenberg picture, by $$\chi _S^F(\eta ,t)=\mathrm{Tr}_{F+R}\left[\rho _{F+R}(0)\text{e}^{\eta a^{}(t)\eta ^{}a(t)}\right],$$ (27) where the trace operation runs over the field and reservoir coordinates and the subsystems are assumed initially uncorrelated, $$\rho _{F+R}(0)=\rho _F(0)\rho _R(0).$$ (28) Inserting operator (17) and its Hermitian conjugate into Eq. (27), the CF for the field writes $`\chi _S^F(\eta ,t)`$ $`=`$ $`\mathrm{Tr}_{F+R}\left\{\rho _{F+R}(0)\mathrm{exp}[\eta (u^{}(t)a^{}+{\displaystyle \underset{k}{}}v_k^{}(t)b_k^{}+w^{}(t))h.c.]\right\}`$ (29) $`=`$ $`\mathrm{Tr}_{F+R}\left\{\rho _{F+R}(0)\text{e}^{\eta w^{}(t)n^{}w(t)}\mathrm{exp}\left[\eta u^{}(t)a^{}\eta ^{}u(t)a\right]\mathrm{exp}\left[{\displaystyle \underset{k}{}}\left(\eta v_k^{}(t)b_k^{}\eta ^{}v_k(t)b_k\right)\right]\right\}`$ (30) $`=`$ $`\text{e}^{\eta w^{}(t)\eta ^{}w(t)}\mathrm{Tr}_F\left\{\rho _F(0)\mathrm{exp}\left[\eta u^{}(t)a^{}\eta ^{}u(t)a\right]\right\}`$ (32) $`\times \mathrm{Tr}_R\left\{\rho _R(0)\mathrm{exp}\left[{\displaystyle \underset{k}{}}\left(\eta v_k^{}(t)b_k^{}\eta ^{}v_k(t)b_k\right)\right]\right\}`$ $`=`$ $`\text{e}^{\eta w^{}(t)\eta ^{}w(t)}\chi _S^F(\eta u^{}(t),0)\mathrm{Tr}_R\left\{\rho _R(0)\mathrm{exp}\left[{\displaystyle \underset{k}{}}\left(\eta v_k^{}(t)b_k^{}\eta ^{}v_k(t)b_k\right)\right]\right\},`$ (33) with $$\chi _S^F(\eta u^{}(t),0)\mathrm{Tr}_F\left[\rho _F(0)\text{e}^{\eta u^{}(t)a^{}\eta ^{}u(t)a}\right].$$ (34) For a thermalized reservoir the state is given by $$\rho _R(0)=_kd^2\beta _k\frac{1}{\pi n_k}\text{e}^{\frac{\left|\beta _k\right|^2}{n_k}}|\beta _k\beta _k|,$$ (35) where $`n_k`$ is the mean occupation number of the $`k`$-th oscillator mode. So, Eq. (34) can be written as $$\chi _S^F(\eta ,t)=\text{e}^{\eta w^{}(t)\eta ^{}w(t)}\chi _S^F(\eta u^{}(t),0)_k\text{e}^{\frac{1}{2}\left|\eta \right|^2\left|v_k(t)\right|^2}d^2\beta _k\frac{1}{\pi n_k}\text{e}^{\frac{\left|\beta _k\right|^2}{n_k}}\mathrm{exp}\left[\eta v_k^{}(t)\beta _k^{}\eta ^{}v_k(t)\beta _k\right].$$ (36) The integral in Eq. (36) is easily solved whit the help of the identity $$\frac{1}{\pi }d^2\eta \text{e}^{\mu \left|\eta \right|^2+\lambda \eta +\nu \eta ^{}}=\frac{1}{\mu }\text{e}^{\frac{\lambda \nu }{\mu }},\left(\text{Re}\left(\mu \right)>0\right)$$ (37) and the CF writes $`\chi _S^F(\eta ,t)`$ $`=`$ $`\text{e}^{\eta w^{}(t)\eta ^{}w(t)}\chi _S^F(\eta u^{}(t),0){\displaystyle _k}\text{e}^{\left|\eta \right|^2\left|v_k(t)\right|^2\left(\frac{1}{2}+n_k\right)}`$ (38) $`=`$ $`\text{e}^{\eta w^{}(t)\eta ^{}w(t)}\chi _S^F(\eta u^{}(t),0)\text{e}^{\left|\eta \right|^2_k\left|v_k(t)\right|^2\left(\frac{1}{2}+n_k\right)}.`$ (39) Substituting the discrete sum by an integration we obtain after a minor algebra $$\underset{k}{}\left|v_k(t)\right|^2\left(\frac{1}{2}+n_k\right)=\left(1\text{e}^{\gamma t}\right)\left(\frac{1}{2}+\overline{n}\right)$$ (40) with $`\overline{n}\left(\text{e}^{\beta \omega _0}1\right)^1`$, $`\beta =\left(k_BT\right)^1`$, where $`k_B`$ is the usual Boltzmann constant and $`T`$ is the reservoir temperature. Substituting Eq. (40) into Eq. (39) we obtain $$\chi _S^F(\eta ,t)=\chi _S^F(\eta u^{}(t),0)\text{e}^{\eta w^{}(t)\eta ^{}w(t)}\text{e}^{\left|\eta \right|^2\left(1\text{e}^{\gamma t}\right)\left(\frac{1}{2}+\overline{n}\right)}.$$ (41) For a reservoir at $`T=0`$K, $`\overline{n}=0`$, the symmetrically ordered CF becomes $`\chi _S^F(\eta ,t)`$ $`=`$ $`\mathrm{Tr}_F\left[\rho _F(0)\text{e}^{\eta u^{}(t)a^{}\eta ^{}u(t)a}\right]\text{e}^{\eta w^{}(t)\eta ^{}w(t)}\text{e}^{\frac{1}{2}\left|\eta \right|^2\left(1\text{e}^{\gamma t}\right)}`$ (42) $`=`$ $`\mathrm{Tr}_F\left[\rho _F(0)\text{e}^{\eta u^{}(t)a^{}}\text{e}^{\eta ^{}u(t)a}\right]\text{e}^{\eta w^{}(t)\eta ^{}w(t)}\text{e}^{\frac{1}{2}\left|\eta \right|^2};`$ (43) comparing the RHS of the second equality with the normal ordered CF, Eq. (25), we identify the following relation $$\chi _N^F(\eta ,t)=\chi _N^F(\eta u^{}(t),0)\text{e}^{\eta w^{}(t)\eta ^{}w(t)}.$$ (44) At this point it is important to emphasize that we have not mentioned yet the initial state of the field inside the cavity. Eq. (44) allows one to obtain the evolved density operator for an arbitrary initial state. The dynamics of the system cavity field \+ reservoir correlates the initial states of the subsystems entailing energy dissipation and loss of coherence during evolution. In the next section we show that when the reservoir is at $`T=0`$K, both, the cavity field and the reservoir states, remain uncorrelated in the course of the evolution. In the absence of pumping, the field state is called dissipative coherent state. ## IV Generation of states in the dissipative cavity ### A Coherent states Let us first consider the situation when the initial state of the field in the cavity C is $$\rho _C(0)=|\alpha \alpha |;$$ (45) introducing it into the normal ordered CF Eq. (22) we have $`\chi _N^F(\eta u^{}(t),0)`$ $`=`$ $`\mathrm{Tr}_F\left[|\alpha \alpha |\text{e}^{\eta u^{}(t)a^{}}\text{e}^{\eta ^{}u(t)a}\right]`$ (46) $`=`$ $`\alpha \left|\text{e}^{\eta u^{}(t)a^{}}\text{e}^{\eta ^{}u(t)a}\right|\alpha `$ (47) $`=`$ $`\text{e}^{\eta u^{}(t)\alpha ^{}\eta ^{}u(t)\alpha },`$ (48) and substituting into Eq. (44) one gets $$\chi _N^F(\eta ,t)=\text{e}^{\eta \left[u^{}(t)\alpha ^{}+w^{}(t)\right]\eta ^{}\left[u(t)\alpha +w(t)\right]}.$$ (49) However, since this result must be identical to the normal ordered CF obtained in the Schrödinger picture, it follows that $$\chi _N^F(\eta ,t)=\mathrm{Tr}_F\left[\rho _F(t)\text{e}^{\eta a^{}}\text{e}^{\eta ^{}a}\right]=\psi _F(t)\left|\text{e}^{\eta a^{}}\text{e}^{\eta ^{}a}\right|\psi _F(t),$$ (50) with $`\rho _F(t)=|\psi _F(t)\psi _F(t)|`$. Then, if we compare the term at the LHS of the second equality of Eq. (50) to Eq. (49), it can be directly verified that one gets $$|\psi _F(t)=|u(t)\alpha +w(t),$$ (51) as a consequence of the disentanglement between the field and the reservoir states, only at $`T=0`$K. Thus the density operator for the continuously pumped field is given by $`\rho _F(t)`$ $`=`$ $`|u(t)\alpha +w(t)u(t)\alpha +w(t)|`$ (52) $`=`$ $`|\text{e}^{\frac{\gamma }{2}t}\text{e}^{i\omega _0t}\alpha +w(t)\text{e}^{\frac{\gamma }{2}t}\text{e}^{i\omega _0t}\alpha +w(t)|,`$ (53) where $$w(t)=F\text{e}^{i\omega t}\frac{\left[1\text{e}^{\frac{\gamma }{2}t}\text{e}^{i\left(\omega \omega _0\right)t}\right]}{\omega \omega _0+i\frac{\gamma }{2}}.$$ (54) By adjusting the pumping field in resonance with the cavity field ($`\omega =\omega _0`$), we have $$w(t)=i\frac{2F}{\gamma }\text{e}^{i\omega _0t}\left(1\text{e}^{\frac{\gamma }{2}t}\right),$$ (55) and the density operator becomes $$\rho _F(t)=|\text{e}^{i\omega _0t}\left[\text{e}^{\frac{\gamma }{2}t}\alpha i\frac{2F}{\gamma }\left(1\text{e}^{\frac{\gamma }{2}t}\right)\right]\text{e}^{i\omega _0t}\left[\text{e}^{\frac{\gamma }{2}t}\alpha i\frac{2F}{\gamma }\left(1\text{e}^{\frac{\gamma }{2}t}\right)\right]|.$$ (56) Setting the relation between the system parameters, $`F=i\alpha \gamma /2`$, all the terms multiplying $`\mathrm{exp}(\gamma t/2)`$ cancel and the field in the cavity remains coherent, oscillating at frequency $`\omega _0`$, $$\rho _C(t)=|\text{e}^{i\omega _0t}\alpha \text{e}^{i\omega _0t}\alpha |.$$ (57) In this way despite the dissipative effect, the pumping action compensates the lost energy, establishing the stationary coherent field state in the cavity. This result is independent of the cavity quality factor $`Q\omega _0/\gamma `$, showing that coherent fields are quite stable. For the generation of another coherent state it is sufficient to adjust the pumping field amplitude. Asymptotically the field state is stationary, $$\underset{t\mathrm{}}{lim}\rho _F(t)|\text{e}^{i\left(\omega _0t+\frac{\pi }{2}\right)}\frac{2F}{\gamma }\text{e}^{i\left(\omega _0t+\frac{\pi }{2}\right)}\frac{2F}{\gamma }|,$$ (58) even if the field in the cavity is initially in the vacuum state $`\alpha =0`$. This result shows how a lossy cavity fills up coherently when pumped by a classical source of EM radiation . ### B Superposition state Let us consider now that the state (57) is sustained in the cavity and that the experiment described in Sec. 2 is going on. With the pumping field acting continuously we consider the adiabatic passage of a Rydberg atom across the cavity $`C`$, i.e., the time of flight of the atom is very small compared to the relaxation time of the field. It is worth noting that if the detuning between the atomic transition frequency $`\omega _i`$ and the cavity field is sufficiently large, the atomic presence inside the cavity do not changes considerably its frequency mode distribution. For an atom prepared initially in the state $`|e`$ the density operator of the system atom+field is written as $$\rho _{F+A}=\rho _A\rho _F=|ee|\rho _F.$$ (59) The resonant interaction of the atom with the field in $`R_1`$, rotates the atomic state by a $`\pi /2`$, $$|e\frac{1}{\sqrt{2}}\left(|e+|g\right),$$ (60) and the joint density operator writes as $$\rho _{F+A}=\frac{1}{2}\left(|e+|g\right)\left(e|+g|\right)\rho _F.$$ (61) Due to the dispersive interaction of the atom with the field in $`C`$ the joint state is given by $$\rho _{F+A}=\frac{1}{2}\left(|ee\left|\text{e}^{i\pi a^{}a}\rho _F\text{e}^{i\pi a^{}a}+\right|gg\left|\rho _F+\right|eg\left|\text{e}^{i\pi a^{}a}\rho _F+\right|ge|\rho _F\text{e}^{i\pi a^{}a}\right),$$ (62) once the state $`|e`$ is always associated with the phase shift operator $`\mathrm{exp}(i\pi a^{}a)`$ in this experiment. Then the outgoing atom passing through $`R_2`$ is submitted to a new $`\pi /2`$ rotation and the joint state becomes $`\rho _{F+A}`$ $`=`$ $`{\displaystyle \frac{1}{4}}\left(|e+|g\right)\left(e|+g|\right)\text{e}^{i\pi a^{}a}\rho _F\text{e}^{i\pi a^{}a}+\left(|e+|g\right)\left(e|+g|\right)\rho _F`$ (64) $`+\left(|e+|g\right)\left(e|+g|\right)\text{e}^{i\pi a^{}a}\rho _F+\left(|e+|g\right)\left(e|+g|\right)\rho _F\text{e}^{i\pi a^{}a}.`$ If the atom is detected in the state $`|g`$ or $`|e`$, the field state collapses instantaneously to $$\rho _F^{\left(\genfrac{}{}{0pt}{}{g}{e}\right)}=\frac{1}{4}\left[\text{e}^{i\pi a^{}a}\rho _F\text{e}^{i\pi a^{}a}+\rho _F+\mathrm{cos}\phi \left(\text{e}^{i\pi a^{}a}\rho _F+\rho _F\text{e}^{i\pi a^{}a}\right)\right],$$ (65) where $$\rho _F^g=g\left|\rho _{F+A}\right|g,\rho _F^e=e\left|\rho _{F+A}\right|e$$ (66) and $`\phi =0`$ or $`\pi `$ depending on the atom being detected in state $`|g`$ or $`|e`$, respectively. The final state can be obtained from Eq. (65) when the initial state is known; for example, if $`\rho _F=|\alpha \alpha |`$ is the initial state of the field in $`C`$, we have from Eq. (57) ($`\alpha `$ containing the time-dependent phase $`e^{i\omega _0t}`$) $$\rho _F^{\left(\genfrac{}{}{0pt}{}{g}{e}\right)}=\frac{1}{N^2}\left[|\alpha \alpha |+|\alpha \alpha |+\mathrm{cos}\phi \left(|\alpha \alpha |+|\alpha \alpha |\right)\right],$$ (67) with $`N=\sqrt{2\left(1+\mathrm{cos}\phi \text{e}^{2\left|\alpha \right|^2}\right)}`$. Then, immediately after the atomic detection, the collapsed state of the field decoheres due to the effects of pumping and energy dissipation. Its explicit time dependence is obtained by first constructing the CF by substituting Eq. (67) into Eq. (44), $`\chi _N(\eta ,t)`$ $`=`$ $`{\displaystyle \frac{1}{N^2}}\{\text{e}^{\eta \left(u^{}(t)\alpha ^{}+w^{}(t)\right)\eta ^{}\left(u(t)\alpha +w(t)\right)}+\text{e}^{\eta \left(u^{}(t)\alpha ^{}w^{}(t)\right)+\eta ^{}\left(u(t)\alpha w(t)\right)}`$ (69) $`\mathrm{cos}\phi \text{e}^{2\left|\alpha \right|^2}[\text{e}^{\eta \left(u^{}(t)\alpha ^{}+w^{}(t)\right)+\eta ^{}\left(u(t)\alpha w(t)\right)}+\text{e}^{\eta \left(u^{}(t)\alpha ^{}w^{}(t)\right)\eta ^{}\left(u(t)\alpha +w(t)\right)}]\},`$ then by comparing, again, the expressions in both, the Schrödinger and Heisenberg pictures, we obtain the density operator for the field state $`\rho _F(t)`$ $`=`$ $`{\displaystyle \frac{1}{N^2}}\left\{\right|\text{e}^{\frac{\gamma }{2}t}\alpha +w(t)\text{e}^{\frac{\gamma }{2}t}\alpha +w(t)|+|(\text{e}^{\frac{\gamma }{2}t}\alpha w(t))(\text{e}^{\frac{\gamma }{2}t}\alpha w(t))|`$ (72) $`+\mathrm{cos}\phi \text{e}^{2\left|\alpha \right|^2(1\text{e}^{\gamma t})}\left[\text{e}^{\left[\text{e}^{\gamma t}\left(\alpha ^{}w(t)\alpha w^{}(t)\right)\right]}\right|(\text{e}^{\frac{\gamma }{2}t}\alpha w(t))\text{e}^{\frac{\gamma }{2}t}\alpha +w(t)|`$ $`+\text{e}^{\left[\text{e}^{\gamma t}\left(\alpha ^{}w(t)\alpha w^{}(t)\right)\right]}|\text{e}^{\frac{\gamma }{2}t}\alpha +w(t)(\text{e}^{\frac{\gamma }{2}t}\alpha w(t))\left|\right]\}.`$ When the amplitude of the field is adjusted to $`F=i\alpha \gamma /2`$ we have $$w(t)=\alpha \left(1\text{e}^{\frac{\gamma }{2}t}\right),$$ (73) and $`\rho _F(t)`$ $`=`$ $`{\displaystyle \frac{1}{N^2}}\left\{\right|\alpha \alpha |+|\alpha (12\text{e}^{\frac{\gamma }{2}t})\alpha (12\text{e}^{\frac{\gamma }{2}t})|`$ (75) $`+\mathrm{cos}\phi \text{e}^{2\left|\alpha \right|^2(1\text{e}^{\gamma t})}\left[|\alpha \left(12\text{e}^{\frac{\gamma }{2}t}\right)\alpha |+|\alpha \alpha \left(12\text{e}^{\frac{\gamma }{2}t}\right)|\right],`$ which shows the time evolution of the quantum state. Asymptotically this state goes to the coherent state $`\rho _F(t)=|\alpha \alpha |`$, which acts as an attractor for other initial quantum states. Independently of the initial amplitude $`\alpha `$, the pumping field supplies energy continuously to the cavity, sustaining the field in a pure coherent state, Eq. (58). ## V Decoherence of a continuously pumped field Now we analyze the evolution of the density operator at times far from the asymptotic regime, in which case we can observe the effects of the classical pumping on the evolution of the state. From Eq. (72) we observe that the coherence terms (non-diagonal) are modified, in comparison to the pumping-free decoherence, by a factor $`\mathrm{exp}\left[\pm \text{e}^{\gamma t}\left(\alpha ^{}w(t)\alpha w^{}(t)\right)\right]`$. However, due to the oscillatory character of these factors, the non-diagonal terms do not sustain the coherence of the field which is continuously attenuated by the damping factor $`\mathrm{exp}\left[2\left|\alpha \right|^2(1\text{e}^{\gamma t})\right],`$ as shown in Sec. 2. The quantum characteristic of a field state can be visualized when represented by a Wigner function , obtained from the Fourier transform of the symmetrically ordered CF Eq. (24). The state given by Eq. (72) has as Wigner function $`W(\zeta ,t)`$ $`=`$ $`{\displaystyle \frac{2}{N^2\pi }}\{\mathrm{exp}[2|\zeta \text{e}^{\gamma t/2}\alpha w(t)|^2]+\mathrm{exp}[2|\zeta +\text{e}^{\gamma t/2}\alpha w(t)|^2]`$ (77) $`2\mathrm{cos}\phi \text{e}^{2\left|\zeta w(t)\right|^2}\mathrm{exp}[2|\alpha |^2(1\text{e}^{\gamma t})]\mathrm{cos}\left[4\text{e}^{\gamma t/2}\text{Im}\left[(\zeta w(t))\alpha ^{}\right]\right]\},`$ where the first two exponential functions are Gaussians centered at $`\left[\text{e}^{\gamma t/2}\alpha +w(t)\right]`$ and $`\left[\text{e}^{\gamma t/2}\alpha w(t)\right]`$ respectively, representing the two distinct states, $`|\text{e}^{\frac{\gamma }{2}t}\alpha +w(t)`$ and $`|\text{e}^{\frac{\gamma }{2}t}\alpha +w(t)`$. The (third) coherence term is composed by three factors, a Gaussian centered at $`w(t)`$, a sinoid modulation, $`\mathrm{cos}\left[4\text{e}^{\gamma t/2}\text{Im}\left[\left(\zeta w(t)\right)\alpha ^{}\right]\right]`$ and the factor responsible for the decoherence, $`\mathrm{exp}\left[2\left|\alpha \right|^2\left(1\text{e}^{\gamma t}\right)\right]`$. The modulation and the time of decoherence given by the last factor depend on the intensity of the state. The larger is $`\left|\alpha \right|^2`$ the faster is the decoherence. In Figs. 1-3 three configurations of the Wigner function (77) for $`\left|\alpha \right|^2=5`$ are shown at three distinct times, $`t=0`$, $`t=\gamma ^1`$ and $`t\mathrm{}`$. For $`F=1`$ we observe the progressive evolution of the superposition state, driven continuously to a stationary coherent state, Fig. 3, representing $$\underset{t\mathrm{}}{lim}W(\zeta ,t)=\frac{2}{\pi }\mathrm{exp}\left[2\left|\zeta \text{e}^{i\left(\omega _0t+\frac{\pi }{2}\right)}\frac{2F}{\gamma }\right|^2\right].$$ (78) The coherence term is suppressed in a time shorter than the time of relaxation of the state, still given by the free decoherence time $`t_d=\left(2\left|\alpha \right|^2\gamma \right)^1`$, being null the effect of the pumping on the coherence terms. The evolution of the superposition state shows the decoherence and relaxation processes (loss of purity) as analyzed through the linear entropy, $`S`$ $`=`$ $`\mathrm{Tr}_\mathrm{F}\left[\rho _\mathrm{F}(\mathrm{t})\rho _\mathrm{F}^2(\mathrm{t})\right]`$ (79) $`=`$ $`1{\displaystyle \frac{2}{N^4}}\{1+4\text{e}^{2\left|\alpha \right|^2}+\text{e}^{4\left|\alpha \right|^2\text{e}^{\gamma t}}+\text{e}^{4\left|\alpha \right|^2\left(1\text{e}^{\gamma t}\right)}`$ (81) $`+\text{e}^{4\left|\alpha \right|^2}\text{e}^{2\left|\alpha \right|^2\text{e}^{\gamma t}}\mathrm{cos}\left[2\text{e}^{\gamma t/2}\text{Im}\left(w(t)\alpha ^{}\right)\right]\}.`$ In Fig. 4 we plotted $`S`$ against $`\gamma t`$ for $`\left|\alpha \right|^2=5`$, where the state is initially pure. As the decoherence goes on the state evolves into a mixture, $`\mathrm{Tr}\rho ^2<1`$, and the entropy increases meaning that there is a flux of information to the reservoir. Although the pumping field is able to restore the energy lost by the cavity field, it is not able to establish back the information of the original superposition state encoded by the coherence terms. Certainly here the process of decoherence is tied to the loss of energy of the field to the reservoir; however there are situations where the information transfer does not occur necessarily together with an energy transfer . The information transfer strongly depends on the phase relation of the superposition of quantum states . Despite that reversible subsystems can exhibit the decoherence and recoherence at constant mean energy , this characteristic ceases to be true for irreversible subsystems. Decoherence still occurs in a characteristic time, which is dependent on the field relaxation time and the field energy, as shown in Sec. (2). An open question remains: Is the information flow (decoherence) always accompanied with an energy flow, or this is only valid for open irreversible systems? We emphasize that the time irreversible character of these models of reservoirs follows from the introduction of approximations as the Wigner-Weisskopf and Markov. ## VI Atoms and pumping The attempt to sustain the field, against decoherence, in a superposition of coherent states by using a classical pumping field is not effective because the insertion of photons for compensating those lost to the reservoir is not phase sensitive. The pumping is only sufficient to re-establish the energy lost to the reservoir and not the original superposition state. Asymptotically only a stationary coherent state is established in the cavity. However, the maintenance of the superposition state could be possible if an additional process accounting for re-establishing the original coherence is considered. In the experiment proposed in , once the superposition is created in $`C`$, the field interacts with atoms sent sequentially through $`C`$. The authors argue that this procedure refreshes the initial coherence. Here we analyze the same process of sending atoms through the cavity, but with the pumping field included. At time $`T,`$ after the detection of the first atom, the state of the field in $`C`$ is given by $`\rho _F(T)`$, Eq. (72); then a second atom is released, going through the same interaction process as the former. After crossing $`R_1`$, the second atom + $`C`$-field joint state is given by $$\rho _{F+A_2}(T)=\frac{1}{2}\left(|e+|g\right)_2\left(e|+g|\right)_2\rho _F(T),$$ (82) and the dispersive interaction in the $`C`$-field produces the entangled joint state $`\rho _{F+A_2}(T)`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left(\right|ee|_2\text{e}^{i\pi a^{}a}\rho _F(T)\text{e}^{i\pi a^{}a}+|gg|_2\rho _F(T)+|eg|_2\text{e}^{i\pi a^{}a}\rho _F(T)`$ (84) $`+|ge|_2\rho _F(T)\text{e}^{i\pi a^{}a}).`$ After crossing the cavity $`R_2`$ the joint state suffers a new transformation, becoming $`\rho _{F+A_2}(T)`$ $`=`$ $`{\displaystyle \frac{1}{4}}[\left(\right|e+|g\left)_2\right(e|+g\left|\right)_2\text{e}^{i\pi a^{}a}\rho _F(T)\text{e}^{i\pi a^{}a}+(|e+|g\left)_2\right(e|+g\left|\right)_2\rho _F(T)`$ (86) $`+\left(\right|e+|g\left)_2\right(e|+g\left|\right)_2\text{e}^{i\pi a^{}a}\rho _F(T)+(|e+|g\left)_2\right(e|+g\left|\right)_2\rho _F(T)\text{e}^{i\pi a^{}a}].`$ If the atom is detected in the $`|g`$ or $`|e`$ state the field will collapse instantaneously to $$\rho _F^{\left(\genfrac{}{}{0pt}{}{g}{e}\right)}(T)=\frac{1}{4}\left(\text{e}^{i\pi a^{}a}\rho _F(T)\text{e}^{i\pi a^{}a}+\rho _F(T)\pm \text{e}^{i\pi a^{}a}\rho _F(T)\pm \rho _F(T)\text{e}^{i\pi a^{}a}\right),$$ (88) with the signal $`+`$ ($``$) standing for $`|g`$ ($`|e`$). Substituting Eq. (72) for $`\rho _F(T)`$ in Eq. (88) we obtain the conditional expression for $`\rho _F^{\left(\genfrac{}{}{0pt}{}{g}{e}\right)}(T)`$. In short, the probability for the second atom be detected in either state $`|g`$ or $`|e`$ is given by $`P_{\left(\genfrac{}{}{0pt}{}{g}{e}\right)}(T)`$ $`=`$ $`\mathrm{Tr}_F\left[\rho _F^{\left(\genfrac{}{}{0pt}{}{g}{e}\right)}(T)\right]`$ (89) $`=`$ $`{\displaystyle \frac{1}{2}}\left(1\pm \text{Re}\left\{\mathrm{Tr}\left[\text{e}^{i\pi a^{}a}\rho _C(T)\right]\right\}\right)`$ (90) $`=`$ $`{\displaystyle \frac{1}{2}}\{1\pm {\displaystyle \frac{\text{e}^{2\left|w(T)\right|^2}}{1+\mathrm{cos}\phi \text{e}^{2\left|\alpha \right|^2}}}[\text{e}^{2\left|\alpha \right|^2\text{e}^{\gamma t}}\mathrm{cosh}\left(4\text{e}^{\frac{\gamma T}{2}}\text{Re}\left[\alpha w^{}(T)\right]\right)`$ (92) $`+\mathrm{cos}\phi \text{e}^{2\left|\alpha \right|^2\left(1\text{e}^{\gamma T}\right)}\mathrm{cos}\left(4\text{e}^{\frac{\gamma T}{2}}\text{Im}\left[\alpha w^{}(T)\right]\right)]\},`$ where $`\phi =0`$ ($`\pi `$) for the first atom detected in the state $`|g_1`$ ($`|e_1`$) and the signal $`+()`$ for the second atom detected in the state $`|g_2`$ ($`|e_2`$). Analyzing Eq. (92) one verifies that if the second atom is detected instantaneously after the first one, $`\gamma T1,`$ one gets $$P_{\left(\genfrac{}{}{0pt}{}{g}{e}\right)}=\frac{1}{2}\left[1\pm \frac{\text{e}^{2\left|\alpha \right|^2}+\mathrm{cos}\phi }{1+\mathrm{cos}\phi \text{e}^{2\left|\alpha \right|^2}}\right],$$ (93) and for $`\left|\alpha \right|1,`$ $$P_{\left(\genfrac{}{}{0pt}{}{g}{e}\right)}=\frac{1}{2}\left[1\pm \mathrm{cos}\phi \right],$$ (94) which is the result obtained in without pumping: If the first atom is detected in $`|g`$ or $`|e`$, the field in $`C`$ collapses to an even or odd cat field state, $`|\mathrm{\Psi }_C=\frac{1}{N}\left(|\alpha +\mathrm{cos}\phi |\alpha \right)`$, $`\phi =0`$ or $`\pi `$ respectively. Now let us suppose that the first atom is detected in the state $`|e`$, then the odd cat state is generated in $`C`$. For $`Tt_d`$ (the time interval between sequentially emitted atoms being quite small) the second atom can be detected either in the state $`|g`$, with conditional probability $`P(e,g,T)0`$, or in the state $`|e`$, with conditional probability $`P(e,e,T)1`$, and so on for the subsequent atoms crossing the apparatus. In this manner the atoms crossing the apparatus sustain (approximately) the superposition state. The measurement of the field state in $`C`$ by the atoms refresh its superposition character, so turning the environment induced decoherence almost ineffective. Thus if an experiment can be done where $`Tt_d`$, a kind of Zeno effect takes place in a continuous measurement process. When the second atom is detected in a different state from the former, the original cat state changes its parity. If one wishes to maintain the parity of the original cat state a resonant interaction could be used to restore the state of the field to its initial state. Such a process, could be outlined as the feedback process reported in , once the resonant interaction time can be controlled to produce a single photon exchange between the atom and field. When the cavity field looses a photon the state of the field flips from odd to even cat state and vice-versa. As the initial field state (prepared by the first atom) is an odd cat state and the second atom is detected at $`|g`$, then a conditional feedback process must be activated and the field flips to the even cat state. The case $`Tt_d`$ is better understood by observing the behavior of the conditional probabilities $`P(g,e,T)`$ and $`P(e,e,T)`$ in Figs. 5 and 6, where these quantities are plotted as function of $`\gamma T`$ for several values of pumping field intensity, $`\left|F\right|^2`$, and for $`\left|\alpha \right|^2=5`$. In the absence of pumping both conditional probabilities go to zero asymptotically because the field in $`C`$ ends in a vacuum state. However, the pumping modifies this trend: the higher the pumping field intensity the faster $`P(g,e,T)`$ and $`P(e,e,T)`$ will attain $`1/2`$, an upper limit that does not depend on the intensity of the pumping. This limit means that the pumping action drives the cavity field to a coherent state and with the next atomic interaction another superposition state is generated. It has 50% of probability to be an even or odd cat state depending on the state in which the second atom is measured. This process that guarantee a 50% efficiency for the detection of the same initial superposition is not very useful if we do not introduce a supplementary process. The efficiency can be increased if a conditional measurement is used for assuring that for each ‘wrong’ result (the atom not being detected in the required state) a resonant feedback atom is sent through the cavity to flip the parity of the field state. It is worth to mention that this process which guarantees an efficiency for the generation of the same superposition state up to 93 % was proposed in for controlling the parity of a field cat state in a quantum logic gate encoding. It is important to note that the classical pumping acts on the cavity-field relaxation time. The stronger the pumping intensity, $`|F|^2`$, the faster will be the relaxation of any initial state to a coherent state. For $`|F|^2=1`$, the time delay between sequentially emitted atoms should be about $`\gamma T3`$, defining a minimum time interval for state reconstruction. While the feedback process is fully dependent on the atomic detectors efficiency, the proposed process for delaying the cavity-field decoherence does not depend. Thus, this process is feasible as soon as each atom of the sequence is prepared in the required state and time, as discussed above. Actually, nowadays it is not an easy task to achieve an efficient control of atomic injection for sending exactly one atom at a time in the cavity . For instance, sending a single atom into a cavity means to send an atomic pulse with an average number of $`0.2`$ atoms, making negligible the chance of finding simultaneously two atoms in the cavity . However, the required technology for energy supply - feeding the cavity continuously with a classical source - is already available since it is employed in current experiments . ## VII Summary and discussion The proposed scheme of the paper shows how a classical pumping field drives any initial state prepared in a lossy cavity into a stationary coherent state. The pumping compensates the lost energy due to the cavity damping mechanism;however, due to the phase insensitivity, this energy feeding does not re-establish the initial superposition of two coherent states, destroyed during the decoherence process. The pumping does not change the time of decoherence of an initial cat state, which remains the same as in the free decoherence case, showing that the information flows from the cavity field to the environment at the same rate independently from the amount of supplied energy. However, the combined action of pumping together with a sequential injection of atoms interacting dispersively with the cavity field (atomic quantum non-demolition measurement) can be used for partially conserving an initial cat state in the cavity. This state can be partially conserved by an atom ‘measuring’ the cavity field state, thus re-establishing partially its original coherence.This result is to be compared with that in , where the mechanism of atomic quantum non-demolition measurement is used without pumping the cavity. In Figs. 5 and 6 we show that for large enough delay times between sequentially injected atoms the action of pumping $`(F0)`$ contributes to reset the initial cat state. This may be important in a practical implementation of quantum processors. The importance for seeking a process that may sustain the coherence of a superposition state is based on the possibility of encoding information in the field state. We expect that even and odd cat states could be used for this purpose because they constitute an orthogonal basis, which should be a sufficient condition to encode qubits. As reported in , we can consider the even cat state as being the $`0`$ qubit and the odd cat state as the $`1`$ qubit, $`|0_L{\displaystyle \frac{1}{N_+}}\left(|\alpha +|\alpha \right)\mathrm{and}|1_L={\displaystyle \frac{1}{N_{}}}\left(|\alpha |\alpha \right).`$ These states can only be used to encode qubits while as pure states, however, dissipation precludes their existence as such. In conclusion, drawing strategies to suppress or at least to delay the decoherence time is therefore extremely important for technological purposes and worth to be pursued. ###### Acknowledgements. MCO acknowledges the financial support from FAPESP (Brazil). MHYM and SSM acknowledge partial support from CNPq (Brazil). ## A Solution of the Heisenberg equation The solution to Eq. (16), goes closely along the lines of Louisell , its Laplace transform is $`(\dot{A})`$ $``$ $`{\displaystyle _0^{\mathrm{}}}\text{e}^{st}\dot{A}𝑑t=i{\displaystyle \underset{k}{}}\lambda _kb_k(0){\displaystyle _0^{\mathrm{}}}\text{e}^{st}\text{e}^{i(\omega _k\omega _0)t}𝑑t`$ (A2) $`{\displaystyle \underset{k}{}}\left|\lambda _k\right|^2{\displaystyle _0^{\mathrm{}}}\text{e}^{st}𝑑t{\displaystyle _0^t}A(t^{})\text{e}^{i(\omega _k\omega _0)(t^{}t)}𝑑t^{}iF{\displaystyle _0^{\mathrm{}}}\text{e}^{st}\text{e}^{i(\omega \omega _0)t}𝑑t.`$ The integrals give $$_0^{\mathrm{}}\text{e}^{st}𝑑t_0^tA(t^{})\text{e}^{i(\omega _k\omega _0)(t^{}t)}𝑑t^{}=\frac{\stackrel{~}{A}(s)}{s+i(\omega _k\omega _0)},$$ (A3) $$_0^{\mathrm{}}\text{e}^{st}\text{e}^{i(\omega _k\omega _0)t}𝑑t=\frac{1}{s+i(\omega _k\omega _0)},$$ (A4) $$_0^{\mathrm{}}\text{e}^{st}\text{e}^{i(\omega \omega _0)t}𝑑t=\frac{1}{s+i(\omega \omega _0)},$$ (A5) and $$_0^{\mathrm{}}\text{e}^{st}\frac{d}{dt}\left[A(t)\right]𝑑t=s\stackrel{~}{A}(s)A(0),$$ (A6) with $`\stackrel{~}{A}(s)(A(t))`$. Substituting these in Eq. (A2), after a little algebra one gets $$\stackrel{~}{A}(s)=\frac{A(0)i\frac{F}{s+i(\omega \omega _0)}i_k\frac{\lambda _kb_k(0)}{s+i(\omega _k\omega _0)}}{\left[s+_k\frac{\left|\lambda _k\right|^2}{s+i(\omega _k\omega _0)}\right]}.$$ (A7) The Wigner-Weisskopf approximation assumes that in the denominator of the LHS in the above equation the frequency spectrum of the reservoir is densely distributed around the cavity characteristic frequency $`\omega _0`$, such that one can replace the discrete sum by an integration over the reservoir frequencies having a distribution $`g(\omega )`$ and do the so-called ‘pole approximation’, $`{\displaystyle \underset{k}{}}{\displaystyle \frac{\left|\lambda _k\right|^2}{s+i\left(\omega _k\omega _0\right)}}`$ $`=`$ $`i{\displaystyle \underset{k}{}}{\displaystyle \frac{\left|\lambda _k\right|^2}{\left(\omega _k\omega _0\right)is}}`$ (A8) $`=`$ $`\underset{s0}{lim}\left\{i{\displaystyle _0^{\mathrm{}}}𝑑\omega ^{}{\displaystyle \frac{g(\omega ^{})\left|\lambda (\omega ^{})\right|^2}{\left(\omega ^{}\omega _0\right)is}}\right\}.`$ (A9) Considering only the first order shift in the simple pole in $`\omega _0`$ in the above integral we have the Wigner-Weisskopf approximation for $`s0`$ $`{\displaystyle \underset{k}{}}{\displaystyle \frac{\left|\lambda _k\right|^2}{s+i\left(\omega _k\omega _0\right)}}`$ $`=`$ $`i{\displaystyle 𝑑\omega ^{}g(\omega ^{})\left|\lambda (\omega ^{})\right|^2\left[\frac{1}{\left(\omega ^{}\omega _0\right)}+i\pi \delta (\omega ^{}\omega _0)\right]}`$ (A10) $`=`$ $`{\displaystyle \frac{\gamma }{2}}+i\mathrm{\Delta }\omega ,`$ (A11) where $$\gamma 2\pi g(\omega _0)\left|\lambda (\omega _0)\right|^2,$$ (A12) is the damping constant and $$\mathrm{\Delta }\omega 𝑑\omega ^{}\frac{g(\omega ^{})\left|\lambda (\omega ^{})\right|^2}{\omega ^{}\omega _0},$$ (A13) is the frequency shift. So Eq.(A7) can be written as $`\stackrel{~}{A}(s)`$ $`=`$ $`{\displaystyle \frac{1}{s+\frac{\gamma }{2}+i\mathrm{\Delta }\omega }}A(0)i{\displaystyle \underset{k}{}}{\displaystyle \frac{\lambda _k}{\left[s+i(\omega _k\omega _0)\right]\left(s+\frac{\gamma }{2}+i\mathrm{\Delta }\omega \right)}}b_k(0)`$ (A15) $`i{\displaystyle \frac{F}{\left[s+i\left(\omega \omega _0\right)\right]\left(s+\frac{\gamma }{2}+i\mathrm{\Delta }\omega \right)}}.`$ Now the calculation of the inverse Laplace transform $`A(t)`$ $`=`$ $`{\displaystyle \frac{A(0)}{2\pi i}}{\displaystyle \text{e}^{st}\frac{1}{s+\frac{\gamma }{2}+i\mathrm{\Delta }\omega }𝑑s}`$ (A18) $`{\displaystyle \frac{1}{2\pi }}{\displaystyle \underset{k}{}}\lambda _kb_k(0){\displaystyle \text{e}^{st}\frac{1}{\left[s+i(\omega _k\omega _0)\right]}\frac{1}{\left(s+\frac{\gamma }{2}+i\mathrm{\Delta }\omega \right)}𝑑s}`$ $`{\displaystyle \frac{1}{2\pi }}F{\displaystyle \text{e}^{st}\frac{1}{\left[s+i\left(\omega \omega _0\right)\right]}\frac{1}{\left(s+\frac{\gamma }{2}+i\mathrm{\Delta }\omega \right)}𝑑s},`$ where $`A(t)=e^{i\omega _0t}a(t)`$ and disregarding the small frequency shift $`\mathrm{\Delta }\omega `$, gives after a little algebra the solution to the Heisenberg equation (12), $$a(t)=u(t)a(0)+\underset{k}{}v_k(t)b_k(0)+w(t),$$ (A19) where $$u(t)=\text{e}^{\frac{\gamma }{2}t}\text{e}^{i\omega _0t},$$ (A20) $$v_k(t)=\lambda _k\text{e}^{i\omega _kt}\frac{\left[1\text{e}^{\frac{\gamma }{2}t}\text{e}^{i\left(\omega _k\omega _0\right)t}\right]}{\omega _0\omega _ki\frac{\gamma }{2}},$$ (A21) and $$w(t)=F\text{e}^{i\omega t}\frac{\left[1\text{e}^{\frac{\gamma }{2}t}\text{e}^{i\left(\omega \omega _0\right)t}\right]}{\omega \omega _0+i\frac{\gamma }{2}}.$$ (A22) Figure Captions Fig. 1 Wigner distribution function for the superposition state $`\left(|\alpha +|\alpha \right)/\sqrt{2}`$ with $`\left|\alpha \right|^2=5`$. The central structure represents the coherence of the quantum state. Fig. 2 Wigner distribution function for the state of Fig. 1 evolved to $`\gamma t=1`$. The original coherence was suppressed by the environment action and the state suffers a continuous displacement due to the pumping field. Fig.3 Asymptotic Wigner distribution function for the state of Fig. 1. The original superposition state evolved asymptotically to a coherent state due to the classical pumping. Fig. 4 The evolution in time (in units of $`\gamma ^1`$) of the linear entropy for the continuously pumped initial superposition state. The pumping does not affect the coherence terms, the state evolves from a pure state to a mixture and then to a pure state again, as in the absence of the pumping, but the final state is a coherent state instead of a vacuum state. Fig. 5 The conditional probability $`P(g,e,T)`$ (first atom in $`|e`$ and second in $`|g`$) increases with the interaction time $`T`$ (in units of $`\gamma ^1`$) as the pumping field intensity increases, saturating at $`0.5`$. Fig. 6 As like as $`P(g,e,T)`$ in Fig. 5, the conditional probability $`P(e,e,T)`$ (first and second atoms in $`|e`$ ) increases with the interaction time $`T`$ as the pumping field intensity increases, saturating at $`0.5`$. This means that the cavity field state has $`50\%`$ chance to be left in an even or odd cat state.