1: %\documentstyle[aps,multicol,epsf]{revtex}
2: \documentstyle[aps,epsf]{revtex}
3: %
4: \newcommand{\figsize}{10 cm}
5:
6: \input epsf
7: \begin{document}
8:
9: \title{Random matrix model for chiral symmetry breaking and color
10: superconductivity in QCD at finite density.\thanks{An animated gif movie
11: showing the evolution of the phase diagram with the chiral and diquark
12: coupling parameters can be viewed at
13: http://www.nbi.dk/\~{}vdheyden/QCDpd.html}}
14:
15: \author{Beno\^\i t Vanderheyden and A. D. Jackson \\
16: The Niels Bohr Institute, Blegdamsvej 17, DK-2100 Copenhagen \O, Denmark.}
17:
18: \date{\today}
19:
20: \maketitle
21:
22: \begin{abstract}
23:
24: We consider a random matrix model which describes the competition between
25: chiral symmetry breaking and the formation of quark Cooper pairs in QCD at
26: finite density. We study the evolution of the phase structure in
27: temperature and chemical potential with variations of the strength of the
28: interaction in the quark-quark channel and demonstrate that the phase
29: diagram can realize a total of six different topologies. A vector
30: interaction representing single-gluon exchange reproduces a topology
31: commonly encountered in previous QCD models, in which a low-density chiral
32: broken phase is separated from a high-density diquark phase by a
33: first-order line. The other five topologies either do not possess a
34: diquark phase or display a new phase and new critical points. Since these
35: five cases require large variations of the coupling constants away from the
36: values expected for a vector interaction, we conclude that the phase
37: diagram of finite density QCD has the topology suggested by single-gluon
38: exchange and that this topology is robust.
39:
40: \end{abstract}
41:
42: \section{Introduction}
43:
44: A number of early and recent model studies of finite density
45: QCD~\cite{Bar77,BaiLov84,AlfRaj98,RapSch98} have suggested that quark Cooper
46: pairs may form above some critical density and lead to `color
47: superconducting' matter. Although perturbation theory performed on
48: single-gluon exchange suggests pairing gaps of a few MeV~\cite{BaiLov84},
49: some recent calculations including non-perturbative interactions, either in
50: the form of the Nambu Jona-Lasinio (NJL)
51: model~\cite{AlfRaj98,BerRaj99,SchKle99,low} or those induced by
52: instantons~\cite{RapSch98,CarDia99}, indicate gaps as large as $100$ MeV.
53: Theses values imply that color superconductivity may be relevant to the
54: physics of heavy-ion collisions and neutron stars.
55:
56: Quark pairing singles out one color direction and thus spontaneously breaks
57: $SU(3)_{\rm color}$ to $SU(2)_{\rm color}$. The pattern of symmetry breaking
58: may however be richer, since the formation of condensates in one $\langle qq
59: \rangle$ channel competes with the breaking of chiral symmetry in the
60: orthogonal $\langle \bar q q\rangle$ channel. In an earlier
61: paper~\cite{VanJac99} we formulated random matrix models for both chiral and
62: diquark condensations in the limit of two quarks flavors and zero chemical
63: potential. Our aim was to understand the phase structures which result from
64: the competition between the two forms of order solely on the basis of the
65: underlying symmetries. In this spirit, we constructed random matrix
66: interactions for which the single quark Hamiltonian satisfies two basic
67: requirements: (1) its block structure reflects the color $SU(N_c)$ and chiral
68: $SU(2)_{\rm L} \times SU(2)_{\rm R}$ symmetries of QCD and (2) the single
69: quark Hamiltonian is Hermitian. This last condition ensures the existence of
70: well-defined relationships between the order parameters and the spectral
71: properties of the interactions. In particular, condition (2) is obeyed by
72: single-gluon exchange, which is generally regarded as the relevant
73: description of QCD. Aside from conditions (1) and (2), the dynamics of the
74: interactions does not contain any particular structure. In order to solve
75: the model exactly, we described this ``dynamics'' by independent Gaussian
76: distributions of matrix elements.
77:
78: In practice, what distinguishes the interactions from one another is their
79: respective coupling constants, $A$ and $B$, in the $\langle qq \rangle$ and
80: $\langle \bar q q \rangle$ channels. The ratio $B/A$ measures the balance
81: between chiral and diquark condensation forces. The absolute magnitudes of
82: $A$ and $B$ play a secondary role. They introduce a scale in the condensation
83: fields but do not affect the phase structure. We have shown
84: in~\cite{VanJac99} that the condition (2) of Hermiticity forces $B/A$ to be
85: smaller or equal to $N_c/2$. This constraint results in the absence of a
86: stable diquark phase in the limit of zero chemical potential.
87:
88: The purpose of this paper is to extend our previous analysis to non-zero
89: chemical potentials, for which the interactions cease to be Hermitian.
90: Non-Hermitian interactions lead to considerable difficulties in numerical
91: calculations in both lattice and random matrix theories. Standard Monte
92: Carlo techniques fail because the fermion determinant in the action is
93: complex and the sampling weights are no longer positive definite. Obtaining
94: reliable results thus requires a proper treatment of cancellations between a
95: large number of terms; a problem which has not yet found a satisfactory
96: solution~\cite{Biel,Ste96,KogMat83,BarBeh86,KogLom95,BarMor97}. The
97: present random matrix models possess exact solutions. The saddle-point
98: methods used in~\cite{VanJac99} to derive the free-energy as a
99: function of the condensation fields remain valid at finite densities and give
100: exact results in the thermodynamic limit of matrices of infinite dimensions.
101: Here, we will use these methods to calculate the thermodynamics quantities as
102: a function of the condensation fields and deduce the phase diagrams from the
103: field configurations which maximize the pressure.
104:
105: We will discover that the pressure function has a simple analytic dependence
106: which leads to polynomial gap equations. This situation reminds us of the gap
107: equations obtained in a Landau-Ginzburg theory near criticality, and the
108: random matrix approach is analogous in many respects. Both theories
109: associate a change of symmetry with a change of state in the system. They
110: are both mean-field and describe the dynamics of a reduced number of degrees
111: of freedom $N$, where $N$ scales with the volume of the system. In random
112: matrix models, these degrees of freedom can be related to the low-lying quark
113: excitation modes in the gluon background, i.e. the zero modes in an instanton
114: approximation of that background~\footnote{This relation is explicit in the
115: random matrix models where only chiral symmetry is
116: considered~\cite{ShuVer93,Ver94,Ver99}; it is expected to remain true once
117: color symmetry is also taken into account~\cite{VanJac99}.}. In the
118: Landau-Ginzburg formulation, the degrees of freedom correspond to the
119: long-wavelength modes which remain after coarse-graining. There is, however,
120: an essential difference in the construction of the two theories. A
121: Landau-Ginzburg theory starts with the specification of an effective
122: potential for the relevant degrees of freedom. A random matrix model starts
123: at a more microscopic level with the construction of an interaction which,
124: once integrated over, produces an effective potential. The integration can
125: carry along dynamical constraints and thus restricts the allowed range of
126: coupling constants. An example of such restriction is the Hermiticity
127: condition in~\cite{VanJac99}, which implies that $B/A \le N_c/2$. This
128: condition is characteristic of the dynamics of the interactions which we
129: consider and remains true at finite chemical potential. We will therefore
130: take it into account in the following.
131:
132: In this paper, we consider random matrix models in which both chiral and
133: color condensations can take place and study the resulting phase diagrams in
134: temperature and chemical potential. We review the model of
135: Ref.~\cite{VanJac99} and discuss the general form of the pressure as a
136: function of the condensation fields in Sec.~\ref{s:models}. We then solve the
137: gap equations and analyze the six topologies that the phase diagram can
138: assume in Sec.~\ref{s:pd}. We compare our results with those of QCD
139: effective models, discuss the possibility of other symmetry breaking patterns
140: and the extension to the case of non-zero quark current masses in
141: Sec.~\ref{s:discuss}. Section~\ref{s:conclusions} presents our conclusions.
142:
143: \section{Formulation of the model}
144: \label{s:models}
145:
146: \subsection{The partition function}
147: \label{ss:models}
148:
149: The generalization of the matrix models introduced in Ref.~\cite{VanJac99} to
150: finite quark densities is straightforward. We represent the quark fields for
151: each of the two flavors by $\psi_1^{\phantom \dagger}$ and $\psi_2^{\phantom
152: \dagger}$. The partition function at temperature $T$ and chemical potential
153: $\mu$ is then
154: \begin{eqnarray}
155: && Z(\mu,T) = \int {\cal D}H\,{\cal D}\psi_1^\dagger\, {\cal
156: D}\psi_1^{\phantom\dagger}
157: \,{\cal D}\psi_2^T\,{\cal D}\psi_2^* \nonumber \\
158: && \times \exp{\left[i \left( \begin{array}{c}
159: \psi_1^\dagger \\
160: \rule{0pt}{1ex} \\
161: \psi_2^T
162: \end{array}
163: \right)^T
164: \left( \begin{array}{cc}
165: {\cal H} + \left(\pi T+ i \mu\right) \gamma_0 + i m &
166: {\eta P_\Delta}\\
167: \rule{0pt}{1ex} & \rule{0pt}{1ex} \\
168: -\eta^* P_\Delta&
169: -{\cal H}^T + \left(\pi T-i \mu\right) \gamma_0^T - i m
170: \end{array}
171: \right)
172: \left( \begin{array}{c}
173: \psi_1^{\phantom\dagger} \\
174: \rule{0pt}{1ex} \\
175: \psi_2^*
176: \end{array}
177: \right)
178: \right].}
179: \label{partfunc}
180: \end{eqnarray}
181: Here, ${\cal H}$ is a matrix of dimension $4 \times N_c \times N$ which
182: represents the interaction of a single quark with a gluon background. Its
183: measure, ${\cal D} H$, will be discussed below. The parameters $m$ and
184: $\eta$ select a particular direction for chiral and color symmetry breaking
185: and are to be taken to zero in the appropriate order at the end of the
186: calculations. The current quark mass $m$ is to be associated with the chiral
187: order parameters $\langle \psi^\dagger_1 \psi_1^{\phantom \dagger} \rangle$
188: and $\langle \psi^T_2 \psi_2^* \rangle$. The complex parameter $\eta$ is to
189: be associated with the order parameter for Cooper pairing, $\langle \psi_2^T
190: P_\Delta \psi_1^{\phantom \dagger} \rangle$, in which $P_\Delta \equiv i C
191: \gamma_5 \lambda_2$ ($C$ is the charge conjugation operator) selects the
192: quark-quark combinations which are antisymmetric in spin and color, i.e. a
193: chiral isosinglet, Lorentz scalar, and color $\bar 3$ state. Note that in
194: order to permit the construction of correlations between the fields $\psi_1$
195: and $\psi_2^T$, we have transposed the single quark propagator in the second
196: flavor, hence the upperscript $T$.
197:
198: The interaction ${\cal H}$ is intended to mimic the effects of gluon
199: fields and thus explicitly includes the desired chiral and color
200: symmetries. Of central interest is single-gluon exchange, which has the
201: chiral block-structure
202: \begin{eqnarray}
203: {\cal H_{\rm sge}}& = & \left(
204: \begin{array}{cc}
205: 0 & W_{\rm sge} \\
206: W^\dagger_{\rm sge} & 0
207: \end{array}
208: \right),
209: \label{chiralblock}
210: \end{eqnarray}
211: where $W_{\rm sge}$ has the spin and color block-structure of a vector
212: interaction,
213: \begin{eqnarray}
214: W_{\rm sge} & = & \sum_{\mu =1}^4 \sum_{a=1}^8 \sigma_\mu^+ \otimes \lambda^a \otimes
215: A^{\mu a}.
216: \label{W}
217: \end{eqnarray}
218: Here, $\sigma_\mu^+=(1,i\vec \sigma)_\mu$ are the $2\times2$ spin matrices,
219: and $\lambda_a$ denote the $N_c \times N_c$ Gell-Mann matrices. The $A^{\mu
220: a}$ are real $N \times N$ matrices which represent the gluon fields.
221:
222: Since we want to explore the evolution of the phase structure as the balance
223: between chiral and diquark condensations is changed, we consider the larger
224: class of Hermitian interactions to which single-gluon exchange belongs.
225: As noted in the introduction, this
226: choice is motivated by the fact that Hermitian interactions
227: have a clear relationship between the order parameters and the
228: spectral properties, in the form of Banks-Casher formulae~\cite{BanCas80}.
229: We write an Hermitian interaction ${\cal H}$ as an expansion into a direct
230: product of the sixteen Dirac matrices $\Gamma_C$ times the $N_c^2$ color
231: matrices. The matrix elements are given by
232: \begin{eqnarray}
233: {\cal H}_{\lambda i \alpha k;\kappa j \beta l} & = &
234: \sum_{C=1}^{16} \left( \Gamma_C^{\phantom T} \right)_{\lambda i;\kappa j}
235: \sum_{a=1}^{N_c^2} \Lambda^a_{\alpha\beta}
236: \left(A^{C a}_{\lambda\kappa}\right)_{k l},
237: \label{calH}
238: \end{eqnarray}
239: where the indices ($\lambda$,$\kappa$), $(i,j)$, and $(\alpha,\beta)$
240: respectively denote chiral, spin, and color quantum numbers, while $(k,l)$
241: are matrix indices running from $1$ to $N$. The $\Lambda^a$ represent the
242: color matrices $\lambda^a$ when $a \le N_c^2 -1$ and the diagonal matrix
243: $(\delta_c)_{\alpha\beta}=\delta_{\alpha\beta}$ when $a = N_c^2$. The
244: normalization for color matrices is ${\rm Tr}[\lambda^a \lambda^b]=2
245: \delta_{ab}$ and ${\rm Tr}[\delta_c^2]=N_c$; the normalization of the Dirac
246: matrices is ${\rm Tr}[\Gamma_C \Gamma_{C'}]=4 \delta_{C C'}$.
247:
248: The random matrices $A^{Ca}$ are real when $C$ is vector or axial
249: vector ($C=V,A$) and real symmetric when $C$ is scalar, pseudoscalar, or
250: tensor $(C=S,P,T)$~\cite{VanJac99}. Their measure is
251: \begin{eqnarray}
252: {\cal D} H &=& \left\{\prod_{C a}\prod_{\lambda \kappa} {\cal
253: D}A^{Ca}_{\lambda\kappa}\right\}
254: \exp\left[ - N \sum_{C a}
255: \sum_{\lambda\kappa}\,\beta_C\Sigma_{Ca}^2 \,\,{\rm Tr}
256: [A^{Ca}_{\lambda\kappa} (A^{Ca}_{\lambda\kappa})^T] \right],
257: \label{measure}
258: \end{eqnarray}
259: where ${\cal D}A^{Ca}_{\lambda \kappa}$ are Haar measures. Here, $\beta_C=1$
260: for $C=V,A$ and $\beta_C=1/2$ for $C=S,P,T$. We want to mimic interactions
261: which in a four-dimensional field theory would respect color $SU(N_c)$ and
262: Lorentz invariance in the vacuum. Therefore, we choose a single variance
263: $\Sigma_{Ca}$ for all channels which transform equally under color and space
264: rotations.
265:
266: The temperature and chemical potential enter the model in Eq.~(\ref{Omega})
267: through the inclusion of the first Matsubara frequency in the single quark
268: propagator. Such $T$ and $\mu$ dependence is certainly oversimplified but none
269: the less sufficient to produce the desired physics. Our purpose is to
270: understand the general topology of the phase diagram and not to provide
271: explicit numbers. More refined treatments including, for instance, all
272: Matsubara frequencies would modify the details of the phase diagram and map
273: every $(\mu,T)$ coordinate to a new one. However, any such mapping will
274: necessarily be monotonic and will conserve the topology. We note in
275: particular that we do not assume a temperature dependence of the variances,
276: i.e. we neglect the $T$-dependence of the gluon background. A non-analytic
277: behavior in $T$ should only arise in the contribution to the thermodynamics
278: from the degrees of freedom related to chiral and diquark condensations. We
279: thus expect a realistic $T$-dependence of the gluon background (and of the
280: variances) to be smooth and not to affect the overall phase topology.
281:
282: We choose the signs of the $T$- and $\mu$-dependencies to mimic a diquark
283: condensate which is uniform in time, i.e. which does not contain a proper
284: pairing frequency. In a microscopic theory formulated in four-momentum space,
285: the absence of a proper frequency leads to pairing between particle and hole
286: excitations with energies which are symmetric around the Fermi surface. We
287: simulate this effect by selecting opposite $T$-dependences for the fields
288: $\psi_1^{\phantom T}$ and $\psi^T_2$, while maintaining the same
289: $\mu$-dependence.
290:
291: \subsection{The pressure function}
292: \label{s:pressure}
293:
294: The integration over the random matrix interactions is Gaussian and can thus
295: be performed exactly. Following the procedure of Ref.~\cite{VanJac99}, we use
296: a Hubbard-Stratonovitch transformation to introduce two auxiliary variables
297: $\sigma$ and $\Delta$, to be associated with the chiral and pairing order
298: parameters respectively
299: \begin{eqnarray}
300: \sigma &\leftrightarrow&
301: \langle \psi^\dagger\psi \rangle \equiv
302: \langle \psi^\dagger_1 \psi^{\phantom \dagger}_1 \rangle =
303: - \langle \psi^T_2 \psi^{*}_2 \rangle, \nonumber \\
304: \Delta &\leftrightarrow&
305: \langle \psi \psi \rangle \equiv
306: \langle \psi^T_2 P_\Delta \psi^{\phantom T}_1 \rangle =
307: \left(\langle \psi^\dagger_1 P_\Delta \psi^*_2 \rangle\right)^*.
308: \label{HS}
309: \end{eqnarray}
310: An integration over the fermion fields then reduces the
311: partition function to
312: \begin{eqnarray}
313: Z(\mu,T) &=& \int d\sigma d\Delta \exp\left[-4 N \Omega(\sigma,\Delta)\right],
314: \label{partfunceff}
315: \end{eqnarray}
316: where $\Omega(\sigma,\Delta)$ is the negative of the pressure,
317: $P(\sigma,\Delta)$, per degree of freedom and per unit spin and flavor,
318: \begin{eqnarray}
319: \Omega(\sigma,\Delta) & = & - P(\sigma,\Delta) \nonumber \\
320: & = & A \Delta^2 + B \sigma^2 -{1\over 2} \bigg\{
321: (N_c-2) \log\Big[ \Big( (\sigma + m - \mu)^2 + T^2 \Big)
322: \Big( (\sigma + m + \mu)^2+ T^2 \Big) \Big] \nonumber \\
323: && +
324: 2 \log\Big[ \Big( (\sigma + m - \mu)^2 + T^2 + |\Delta+\eta| ^2 \Big)
325: \Big( (\sigma + m + \mu)^2 + T^2 + |\Delta+\eta|^2 \Big) \Big]
326: \bigg\},
327: \label{Omega}
328: \end{eqnarray}
329: where we have dropped the prefactors $\pi$ in the temperature dependence for
330: simplicity. Here, the coupling constants $A$ and $B$ are weighted averages
331: of the Fierz coefficients $f^{Ca}_\Delta$ and $f^{Ca}_\chi$ obtained
332: respectively by projecting the interaction $\Gamma_C \otimes \Lambda_a$ onto
333: chiral and diquark channels,
334: \begin{eqnarray}
335: A = 2 \left(\sum_{Ca} \Sigma_{Ca}^{-2} f_\Delta^{Ca}\right)^{-1}, \,\,
336: B = 2 \left(\sum_{Ca} \Sigma_{Ca}^{-2} f_\chi^{Ca}\right)^{-1}.
337: \label{AB}
338: \end{eqnarray}
339: To make contact to microscopic theories, we note that a small coupling
340: limit in either channel corresponds to a small Fierz constant and hence to
341: large parameters $A$ or $B$. This limit favors small fields $\Delta$ or
342: $\sigma$. Because we can always rescale the condensation fields by either
343: $\sqrt{A}$ or $\sqrt{B}$, the only independent parameter in Eq.~(\ref{Omega})
344: is the ratio of $B/A$, which by virtue of Eq.~(\ref{AB}) is a measure of the
345: balance between the condensation forces. Again, Hermitian matrices ${\cal H}$
346: satisfy $B/A \le N_c/2$.
347:
348: The mass $m$ and the parameter $\eta$ explicitly break chiral and color
349: symmetries. They act as external fields which select a particular direction
350: for the condensation pattern, and should be taken to zero at the end of the
351: calculations. (They can also be kept constant to study the effect of a small
352: external field, a point which we take in Sec.~\ref{s:discuss}.) They are
353: useful for obtaining the order parameters from derivatives of the partition
354: function. In the thermodynamic limit $N \to \infty$, $Z(\mu,T)$ in
355: Eq.~(\ref{partfunceff}) obeys
356: \begin{eqnarray}
357: \lim_{N \to \infty} \log Z(\mu,T) = - \lim_{N \to \infty} 4 N \,\,
358: {\min}_{\sigma,\Delta} \left\{\Omega(\sigma,\Delta)\right\},
359: \end{eqnarray}
360: where the right side represents the global minimum of $\Omega$ in
361: Eq.~(\ref{Omega}) for fixed $\mu$ and $T$. The order parameters for chiral and
362: diquark condensations are given as
363: \begin{eqnarray}
364: \langle \psi^\dagger \psi \rangle = \lim_{m \to 0} \lim_{N \to \infty}
365: \left. \frac{1}{4 N_f N} \frac{\partial \log Z}{\partial m}
366: \right|_{\eta,\eta^*=0}, \quad
367: \langle \psi^T \psi \rangle = - \lim_{\eta,\eta^* \to 0} \lim_{N\to \infty}
368: \left. \frac{1}{4 N} \frac{\partial \log Z}{\partial \eta^*}\right|_{m=0},
369: \label{orderparameters}
370: \end{eqnarray}
371: where the number of flavor is $N_f = 2$. Note that the thermodynamic limit $N
372: \to \infty$ must be taken first before the small field limit $m,\eta \to 0$,
373: see~\cite{JacVer96}. Given the $m$- and $\eta$-dependences of the log terms
374: in Eq.~(\ref{Omega}), we have
375: \begin{eqnarray}
376: \langle \psi^\dagger \psi \rangle = B\, \sigma(\mu, T), \quad
377: \langle \psi^T \psi \rangle = A\, \Delta(\mu,T),
378: \end{eqnarray}
379: where $\sigma(\mu,T)$ and $\Delta(\mu,T)$ are the condensation fields
380: which minimize $\Omega(\sigma,\Delta)$ for fixed $\mu$ and $T$.
381:
382: Equation~(\ref{Omega}) is the main expression from which we will deduce the
383: phase structure as a function of the ratio $B/A$. Its form is very simple to
384: understand. The quadratic terms correspond to the energy cost for creating
385: static field configurations with finite $\sigma$ and $\Delta$. The log terms
386: represent the energy of interaction between the condensation fields and the
387: quark degrees of freedom. They can be written in a compact form as ${\rm
388: Tr}[\log S(\sigma,\Delta)] = \log {\rm det}~S(\sigma, \Delta)$, where ${\rm
389: Tr}$ is a trace in flavor, spin, and color, and $S$ is the single quark
390: propagator in a background of $\sigma$ and $\Delta$ fields. Substituting the
391: Mastubara frequency $T \to i p_4$, it becomes clear that the poles of $S$ in
392: $p_4$ correspond to the excitation energies of the system. From
393: Eq.~(\ref{Omega}), we see that two colors develop gapped excitations with
394: \begin{eqnarray}
395: p_4 & = & \pm \sqrt{(\sigma \mp \mu)^2+\Delta^2},
396: \label{gapped}
397: \end{eqnarray}
398: where the plus and minus signs respectively correspond to particle and
399: antiparticle modes. The $N_c-2$ remaining colors have ungapped excitation
400: with
401: \begin{eqnarray}
402: p_4 & = & \pm |\sigma \mp \mu|.
403: \label{ungapped}
404: \end{eqnarray}
405:
406: It is important to recognize that the potential $\Omega$ in Eq.~(\ref{Omega})
407: contains the contribution to the thermodynamics from only the low energy
408: modes of Eqs.~(\ref{gapped}) and (\ref{ungapped}). The right side of
409: Eq.~(\ref{Omega}) thus corresponds to the non-analytic piece in the
410: thermodynamic potential which describes the critical physics related to
411: chiral and color symmetry breaking. In a microscopic model, the right side of
412: Eq.~(8) would also contain a smooth analytic component $\Omega_{\rm
413: reg}(\mu,T)$ which arises from all other, non-critical degrees of freedom
414: in the system. Our model does not contain this contribution and thus cannot
415: be taken as a quantitative description of bulk thermodynamics properties; the
416: model is constructed to describe the critical properties and should be used
417: as such.
418:
419: We show in the next section that the forms of the potential
420: $\Omega(\sigma,\Delta)$ in Eq.~(\ref{Omega}) and of the associated excitation
421: energies in Eqs.~(\ref{gapped}) and (\ref{ungapped}) are sufficient to
422: produce a rich variety of phase diagrams which illustrate in a clear way the
423: interplay between chiral and color symmetries. In particular, we will find
424: that single-gluon exchange reproduces the topology obtained in many
425: microscopic models of finite density QCD, see for
426: instance~\cite{RapSch98,CarDia99,BerRaj99,SchKle99}.
427:
428: \section{Exploring the phase diagrams}
429: \label{s:pd}
430:
431: We noted earlier that the potential $\Omega$ in Eq.~(\ref{Omega}) is very
432: similar to a Landau-Ginzburg functional. Although $\Omega(\sigma,\Delta)$ is
433: not an algebraic function, it is equivalent to a polynomial of order
434: $\sigma^6$ and $\Delta^4$: The gap equations reduce to coupled polynomial
435: equations of fifth order in $\sigma$ and third order in $\Delta$.
436: Generically, four types of solutions exist:
437: \begin{itemize}
438: \item (i) the $0$-phase, the trivial phase in which both
439: $\sigma$ and $\Delta$ vanish,
440: \item (ii) the $\chi$-phase, in which chiral symmetry
441: is spontaneously broken but $\Delta=0$,
442: \item (iii) the $\Delta$-phase, in which
443: color symmetry is spontaneously broken but $\sigma=0$,
444: \item (iv) the $\chi\Delta$
445: phase, a mixed broken symmetry phase in which both fields are
446: non-vanishing, $\sigma \neq 0$ and $\Delta \neq 0$.
447: \end{itemize}
448: The $\chi\Delta$-phase is thermodynamically distinct from the $\chi$- and
449: $\Delta$-phases and is not a ``mixture'' of these two phases.
450:
451: At a given $\mu$, $T$, and $B/A$, each of the solutions (i) - (iv) can either
452: be a minimum or a saddle-point of $\Omega$. The complex flow of these
453: solutions with the variation of $B/A$ leads to large variety of phase
454: diagrams. The phase structures can however be grouped according to their
455: topologies as shown in Figs.~\ref{f:panel1}-~\ref{f:panel6}.~\footnote{We
456: have rescaled in each figure the units of $T$ and $\mu$ for clarity.}
457: Figure~\ref{f:panel1} shows the case of smallest values $B/A$, which favor
458: chiral over diquark condensation. $B/A$ then increases continuously from
459: Fig.~\ref{f:panel2} to~\ref{f:panel6}. We now discuss Figs.~1-6 in the
460: six following subsections. We indicate in parentheses the corresponding
461: ranges of $B/A$ and discuss their limits in the text.
462:
463: {\bf Pure chiral condensation:} ($0 \le B/A \le 0.139 N_c$ or $0 \le B/A \le
464: 0.418$ for $N_c=3$; see Fig.~1). Generically, the global minimum of $\Omega$
465: in Eq.~(\ref{Omega}) is realized by field configurations which maximize the
466: log terms while keeping reasonably low values of the quadratic terms $A
467: \Delta^2 + B \sigma^2$. We consider first the limit $A \gg B$. The minimum
468: of $\Omega$ must then correspond to $\Delta = 0$ in order to avoid the large
469: energy penalty $\sim A \Delta^2$. No diquark condensation occurs in this
470: case, and color enters only as a prefactor $N_c$ in the number of degrees of
471: freedom. This becomes clear if we absorb $N_c$ into $B$ and set $B$ to $1$;
472: we then recover the potential $\Omega$ studied in the chiral random matrix
473: models~\cite{HalJac98} which neglect color altogether.
474:
475: We briefly recall the phase structure in this case. The gap equation for the
476: chiral field, $\partial \Omega(\sigma,0)/\partial \sigma = 0$, has a trivial
477: root $\sigma =0$ and four other roots which satisfy the following quadratic
478: equation for $\sigma^2$
479: \begin{eqnarray}
480: N_c (\mu^2-T^2)+B (\mu^2+T^2)^2 +
481: (2 B (T^2-\mu^2)-N_c) \sigma^2 & + & B \sigma^4 = 0.
482: \label{sigma4}
483: \end{eqnarray}
484: In the high temperature limit $T \gg \mu$, both roots in $\sigma^2$ are
485: negative. The only real solution of the gap equation is $\sigma=0$, and the
486: system is in the symmetric phase. Decreasing $T$ for fixed $\mu$, one
487: encounters a line of second-order phase transitions
488: \begin{eqnarray}
489: L_{\chi,2} \equiv N_c (\mu^2-T^2)+B(\mu^2+T^2)^2 & = & 0,
490: \label{sigma_second}
491: \end{eqnarray}
492: where one of the roots $\sigma^2$ in Eq.~(\ref{sigma4}) vanishes. Below
493: $L_{\chi,2}$, the trivial root becomes a local maximum of $\Omega(\sigma,0)$,
494: and we have a pair of local minima at the real roots $\sigma = \pm \sigma_0$,
495: where
496: \begin{eqnarray}
497: \sigma_0 = \left(\frac{N_c}{2 B}- T^2+\mu^2 +
498: \frac{\sqrt{N_c^2-16 B^2 \mu^2 T^2}}{2 B}\right)^{1/2}.
499: \label{sigma0}
500: \end{eqnarray}
501: These roots correspond to a chiral broken phase. They become
502: degenerate with $\sigma = 0$ on the second-order line $L_{\chi,2}$, where the
503: potential $\Omega$ scales as $\Omega(\sigma,0) - \Omega(0,0) \approx
504: \sigma^4$. Thus, the critical exponents near $L_{\chi,2}$ are those of a
505: mean-field $\phi^4$ theory.
506:
507: The second-order line ends at a tricritical point $(\mu_3,T_3)$ at which all
508: five roots of the gap equation vanish and where
509: $\Omega(\sigma,0)-\Omega(0,0)\approx \sigma^6$, giving now critical exponents
510: of a mean-field $\phi^6$ theory. From Eq.~(\ref{sigma4}), this happens when
511: $2 B (T_3^2-\mu_3^2)-N_c=0$, which with the use of Eq.~(\ref{sigma_second})
512: gives
513: \begin{eqnarray}
514: \mu_3 = \sqrt{\frac{N_c}{4 B}} \sqrt{\sqrt{2}-1}, \quad \quad
515: T_3 = \sqrt{\frac{N_c}{4 B}} \sqrt{\sqrt{2}+1}.
516: \label{tricritical}
517: \end{eqnarray}
518: For $\mu > \mu_3$, the transition between the chiral and trivial phases
519: is first-order and takes place along the line of equal pressure
520: \begin{eqnarray}
521: L_{\chi,1} \equiv && \frac{N_c}{2} \left(1+\sqrt{1- 16 \mu^2 T^2
522: \frac{B^2}{N_c^2} } -\log \left[\frac{N_c^2}{2 B^2} \left( 1+ \sqrt{1-16
523: \mu^2 T^2 \frac{B^2}{N_c^2}} \right) \right] \right)
524: \nonumber \\
525: && + N_c \log\left[\mu^2+T^2\right] + B (\mu^2-T^2) = 0.
526: \label{Lchi1}
527: \end{eqnarray}
528: This line intercepts the $T=0$ axis at
529: $\mu=\mu_1$ which obeys
530: \begin{eqnarray}
531: 1+{B \mu_1^2\over N_c}+\log({B \mu_1^2 \over N_c})=0.
532: \label{mu1}
533: \end{eqnarray}
534: This gives $\mu_1 = 0.528 \sqrt{N_c/B}$, or $\mu_1 = 0.914/\sqrt{B}$ for
535: $N_c=3$.
536:
537:
538: $L_{\rm \chi,1}$ in Eq.~(\ref{Lchi1}) is a triple line. To see this and
539: clarify the character of $(\mu_3,T_3)$, it is useful to consider the effect
540: of a non-zero quark current mass $m$~\cite{HalJac98}. A mass $m \neq 0$
541: selects a particular direction for chiral condensation. If we now consider
542: the three-dimensional parameter space $(\mu,T,m)$, the region delimited by
543: $L_{\chi,2}$ and $L_{\chi,1}$ in the plane $m=0$ thus appears to be a surface
544: of coexistence of the two ordered phases with the chiral fields $\pm
545: \sigma_0$ of Eq.~(\ref{sigma0}). Along $L_{\chi,1}$, this surface meets two
546: other `wing' surfaces which extend symmetrically into the regions $m >0$ and
547: $m <0$. Each of the wings is a coexistence surface between one of the
548: ordered phases whose chiral field continues to $\sigma = \pm \sigma_0$ as
549: $m\to 0$ and the high temperature phase. Hence, $L_{\chi,1}$ marks the
550: coexistence of three phases and is a triple line. The three phases become
551: identical at $(\mu_3,T_3)$, which is thus a tricritical point. The second-
552: and first-order lines $L_{\chi,1}$ and $L_{\chi,2}$ join tangentially at
553: $(\mu_3,T_3)$, see~\cite{DomGre84}.
554:
555: The onset of diquark condensation modifies the topology that we have just
556: described. This takes place for coupling ratios $B/A \ge 0.139 N_c$ to which
557: we now turn.
558:
559: {\bf The QCD case:} ($0.139 N_c \le B/A \le \alpha_1(N_c)$ or $0.418 \le B/A
560: \le 1.05$ for $N_c=3$; see Fig.~2). To understand the conditions for the
561: onset of diquark condensation, consider a pure diquark phase by setting
562: $\sigma=0$ in Eq.~(\ref{Omega}). The gap equation $\partial \Omega/\partial
563: \Delta=0$ then has three solutions: $\Delta = 0$, and $\Delta=\pm\Delta_0$
564: where $\Delta_0 = \sqrt{2/A -\mu^2-T^2}$. In the high $T$ and $\mu$ phase,
565: only the trivial root is real and the system is in the symmetric phase.
566: Inside the semicircle
567: \begin{eqnarray}
568: L_{\rm \Delta,2} \equiv 2/A-\mu^2-T^2=0,
569: \end{eqnarray}
570: the trivial root becomes a maximum of $\Omega(0,\Delta)$, and we have a pair
571: of two real minima $\Delta = \pm \Delta_0$ for which color symmetry is
572: spontaneously broken. The roots $\pm \Delta_0$ go continuously to zero as
573: one approaches $L_{\Delta,2}$, which is thus a second-order line provided no
574: other phase develops. The fate of other ordering forms depends on the
575: pressure in the diquark phase,
576: \begin{eqnarray}
577: P_\Delta=-\Omega_\Delta & = & A (\mu^2+T^2) - 2 + (N_c-2) \log\left( \mu^2 +
578: T^2 \right)+ 2 \log\left(\frac{2}{A}\right).
579: \end{eqnarray}
580: The maximum pressure $P_\Delta$ is reached on $L_{\Delta,2}$, where it
581: also equals the pressure of the trivial phase. The condition for the
582: onset of diquark condensation is then clear: The semicircle must lie
583: in part outside the region occupied by the chiral phase. Then, the
584: maximum pressure in the diquark phase is higher than that of the
585: chiral phase, and the diquark phase is stable near the semicircle.
586:
587: To find the minimum ratio $B/A$ for which this condition holds, we compare
588: the dimensions of the chiral phase to the radius of the semicircle, $\mu_{\rm
589: semi}\equiv \sqrt{2/A}$. When $B/A \ll 1$, the line $L_{\Delta,2}$ lies
590: well inside the boundaries of the chiral phase, whose linear dimensions are
591: $T=\sqrt{N_c/B}$ (along $\mu=0$) and $\mu_1=0.528 \sqrt{N_c/B}$ (along
592: $T=0$). The onset of diquark condensation requires therefore $B/A$ to be
593: large enough that the semicircle crosses the first-order line between chiral
594: and trivial phases. This takes place along $T=0$ when $\mu_{\rm
595: semi}=\mu_{1}$, or
596: \begin{eqnarray}
597: \sqrt{2/A} = 0.528 \sqrt{N_c/B},
598: \end{eqnarray}
599: which gives $B/A = 0.139 N_c$, or $B/A = 0.418$ for $N_c =3$.
600:
601: For larger values of $B/A$, the diquark phase exists in a region delimited by
602: $L_{\Delta,2}$, on which it coexists with the symmetric phase, and by a
603: first-order line of equal pressure with the chiral phase. This phase diagram
604: is realized for ratios including that corresponding to single-gluon exchange,
605: $B/A = N_c/ (2 (N_c-1))$ (or $3/4$ for $N_c=3$), which is the ratio taken in
606: Fig.~\ref{f:panel2}. We have verified that the two segments of first-order
607: lines, between the chiral and the diquark phases on the one hand and the
608: chiral and trivial phases on the other hand, join tangentially.
609: Figure~\ref{f:fields} shows the chiral and diquark fields as a function of
610: $\mu$ and $T$. It is worth noting that the chiral field vanishes with a
611: square root law near the second-order line $L_{\rm \chi,2}$ and is
612: discontinuous along the first-order line $L_{\chi,1}$. The diquark field
613: vanishes with a square root law all along the semicircle line.
614:
615: The topology in Fig.~\ref{f:panel2} can also be summarized by a simple
616: counting argument. Consider the difference between the pressures in the
617: diquark and the chiral phases along the $T=0$ axis,
618: \begin{eqnarray}
619: \Delta \Omega = \Omega_\chi-\Omega_\Delta
620: = N_c-2 +(A+B)\mu^2- 2\log \frac{A \mu^2}{2} + N_c \log\frac{B \mu^2}{N_c}.
621: \label{DeltaOmega}
622: \end{eqnarray}
623: This difference varies from $\Delta \Omega = N_c (1 + (2B)/(N_c A)
624: +\log[(2B)/(N_c A)])$ as $\mu \to \sqrt{2/A}$, to $\Delta \Omega = (N_c-2)
625: \log \mu^2$ as $\mu \to 0$. We have just argued that $\Delta\Omega$ can be
626: negative in the former limit if $B/A> 0.139 N_c$. In the latter limit,
627: however, $\Delta\Omega$ is always negative for $N_c > 2$ and the system is
628: necessarily in the chiral phase. That the diquark phase must be metastable
629: for small $\mu$ is clearly a consequence of the fact that chiral condensation
630: uses all $N_c$ colors while diquark condensation uses only colors $1$ and
631: $2$. This counting argument was mentioned in early models of color
632: superconductivity~\cite{AlfRaj98} and is here completely manifest. For $N_c
633: > 2$, the diquark phase must appear at densities higher than those
634: appropriate for the chiral phase.
635:
636: {\bf Onset of the mixed broken symmetry phase:} ($\alpha_1(N_c) \le B/A \le
637: N_c/\sqrt{8}$ or $1.05 \le B/A \le 1.06$ for $N_c=3$; see Fig.~3). The chiral
638: solution $\sigma_0$, Eq.~(\ref{sigma0}), ceases to be a local minimum of
639: $\Omega(\sigma,0)$ for sufficiently large $\mu$. It turns into a saddle
640: point along a line where the second derivative $\partial^2
641: \Omega/\partial \Delta^2$ vanishes,
642: \begin{eqnarray}
643: L_{\rm mix} \equiv
644: \left(N_c A - 2 B\right) \left( N_c+\sqrt{N_c^2-16 B^2 \mu^2 T^2}\right)
645: - 8 B^2 \mu^2 = 0.
646: \label{mixline}
647: \end{eqnarray}
648: A new minimum of $\Omega(\sigma,\Delta)$ with both $\sigma \neq 0$ and
649: $\Delta \neq 0$ develops in the region to the right of $L_{\rm mix}$.
650: This new minimum corresponds to a new phase, the $\chi\Delta$-phase, which
651: competes with the diquark phase. When the $\chi\Delta$-phase first appears
652: along $L_{\rm mix}$, it has the same pressure as the chiral phase.
653: Therefore, the condition for this new phase to realize the largest pressure
654: is that the instability line $L_{\rm mix}$ lies in the region spanned by the
655: chiral phase. Then, along $L_{\rm mix}$, the new phase necessarily has a
656: pressure that exceeds that in the diquark phase and is favored.
657:
658: To see what ratios $B/A$ are needed for the mixed broken symmetry phase,
659: consider the point where $L_{\rm mix}$ meets the $T=0$ axis,
660: \begin{eqnarray}
661: \mu_{\rm mix}&=& \sqrt{\frac{N_c}{2 B} \left({N_c A\over 2 B}-1\right)}.
662: \label{mumix}
663: \end{eqnarray}
664: When $B/A \ll 1$, $\mu_{\rm mix} \sim \sqrt{N_c/B}/ \sqrt{ 4B/A} \gg
665: \sqrt{N_c/B}$ and $L_{\rm mix}$ lies well outside the chiral phase. $L_{\rm
666: mix}$ then moves towards the $\chi$-phase as $B/A$ increases. It first
667: crosses the first-order line between chiral and diquark phases when $\mu_{\rm
668: mix}=\mu_1$, where $\mu_1$ is the point of equal pressure which obeys
669: $\Delta\Omega=0$ in Eq.~(\ref{DeltaOmega}). This condition gives a ratio $B/A
670: = \alpha_1(N_c)$, where $\alpha_1$ is a non-trivial function of $N_c$ which
671: we study in the appendix. Here, we note only that $\alpha_1(3)=1.05$ for
672: three colors. When $B/A > \alpha_1(N_c)$, we find that the mixed broken
673: symmetry phase develops in the wedge bordered by the second-order line
674: $L_{\rm mix}$ of Eq.~(\ref{mixline}), and a first-order line on which it
675: coexists with the diquark phase. The phase diagram is shown in
676: Fig.~\ref{f:panel3}, while an expanded view of the wedge of mixed broken
677: symmetry and of the first-order line near $(\mu_3,T_3)$ are
678: respectively shown in Figs.~\ref{f:panel3b} and~\ref{f:panel3c}.
679:
680: {\bf A new critical point:} ($N_c/\sqrt{8} \le B/A \le \alpha_2(N_c)$ or
681: $1.06 \le B/A \le 1.163$ for $N_c = 3$; see Fig.~4). As $B/A$ increases above
682: $\alpha_1(N_c)$, the semicircle $2/A = \mu^2+T^2$ grows relative to the
683: chiral phase, and eventually reaches the tricritical point $(\mu_3,T_3)$ of
684: Eq.~(\ref{tricritical}) when $B/A = N_c / \sqrt{8}$. At this stage, the
685: segment of first-order line between chiral and trivial phases disappears. For
686: $B/A > N_c/\sqrt{8}$, the semicircle meets the second-order line between
687: chiral and trivial phases, $L_{\chi,2}$ in Eq.~(\ref{sigma_second}), at a new
688: critical point
689: \begin{eqnarray}
690: \mu_4 = \sqrt{{1\over A}\left(1 - {2 B \over N_c A}\right)}, \quad
691: T_4 = \sqrt{{1\over A}\left(1 + {2 B \over N_c A}\right)},
692: \label{tetracritical}
693: \end{eqnarray}
694: which now separates three phases as shown in Fig. \ref{f:panel4}.
695:
696: {\bf Coexistence of four phases:} ($\alpha_2(N_c) \le B/A \le N_c/2$ or
697: $1.163 \le B/A \le 1.5$ for $N_c=3$; see Fig.~5). With still higher coupling
698: ratios $B/A$, the wedge of mixed broken symmetry in Fig.~\ref{f:panel4} grows
699: in size relative to the other phases. Its tip reaches the new critical point
700: $(\mu_4,T_4)$ when $B/A$ satisfies the condition
701: \begin{eqnarray}
702: {B \over A} = \alpha_2(N_c) \equiv \frac{4 N_c - N_c^{3/2}
703: \sqrt{2 N_c-4}}{4 (4 - N_c)},
704: \label{alpha2}
705: \end{eqnarray}
706: the derivation of which we detail in the appendix. We just note here that
707: $\alpha_2( 3 ) = 1.163$. This coupling ratio also marks the appearance of
708: two new critical points, as illustrated in Fig.~\ref{f:panel5}. First, when
709: $B/A > \alpha_2(N_c)$, the point $(\mu_4,T_4)$ characterizes the coexistence
710: of all four phases, and it has thus become a {tetracritical} point. The
711: pressure of the system at $(\mu_4,T_4)$ has the form
712: $\Omega(\sigma,\Delta)-\Omega(0,0) \approx a \sigma^4 + b \Delta^4+ c
713: \sigma^2 \Delta^2$, where $a$, $b$, and $c$ are constants detailed in the
714: appendix. We note in particular that none of the four second-order lines in
715: Fig.~5 join tangentially at the tetracritical point.
716:
717: The second point appearing above $B/A=\alpha_2(N_c)$ is a tricritical
718: point $(\mu_{\rm 3m},T_{\rm 3m})$ which lies on the boundary between the
719: $\chi\Delta$- and the $\Delta$-phases. Its origin can be understood from the
720: similarity between the characters of the mixed broken symmetry phase and the
721: chiral phase. The diquark field in the $\chi\Delta$-phase is
722: \begin{eqnarray}
723: \Delta(\sigma) & =& \left({1+\sqrt{1+4 A^2 \mu^2 \sigma^2}\over A}
724: -\mu^2-T^2-\sigma^2 \right)^{1/2}.
725: \end{eqnarray}
726: Substituting $\Delta$ in $\Omega(\sigma,\Delta)$ for this expression, we find
727: that the gap equation for the chiral field, $\partial
728: \Omega(\sigma,\Delta(\sigma))/\partial \sigma = 0$, has five roots. Despite
729: the fact that $\Delta(\sigma) \neq 0$, the dynamics of these roots as a
730: function of $\mu$ and $T$ is identical to that of the roots in the pure
731: chiral condensation case discussed at the beginning of this section. In
732: particular, the transition to the $\Delta$-phase starts out second-order near
733: $(\mu_4,T_4)$ and takes place along a line where three of the five roots
734: vanish. This second-order line is thus determined by the condition
735: \begin{eqnarray}
736: \left.\frac{d^2 \Omega}{d \sigma^2}(\sigma,\Delta(\sigma))\right|_{\sigma \to
737: 0}& = & 0,
738: \label{partial2Om}
739: \end{eqnarray}
740: which is the analog of Eq.~(\ref{sigma_second}) and gives
741: \begin{eqnarray}
742: L_{\chi\Delta \to \Delta,2} \equiv
743: (N_c-2) (\mu^2-T^2) + \left(B- A(1-A \mu^2)\right)\left(\mu^2+T^2\right)^2=0.
744: \label{mixtocol}
745: \end{eqnarray}
746: In analogy with the pure chiral case, the transition becomes first order at
747: high $\mu$; second- and first-order segments join tangentially at a
748: tricritical point $(\mu_{3\rm m},T_{3\rm m})$.
749:
750: We determine the location of this point as follows. We observe that, as in
751: the pure chiral broken case, the potential along the second-order line
752: $L_{\chi\Delta \to \Delta,2}$ has the form $\Omega(\sigma,\Delta(\sigma)) -
753: \Omega(0,0) \approx \sigma^4$. (See Eq.~(\ref{partial2Om}).)
754: This scaling form becomes
755: $\Omega(\sigma,\Delta(\sigma)) -\Omega(0,0) \approx \sigma^6$ at $(\mu_{3\rm
756: m},T_{3\rm m})$ where all five roots vanish.
757: Thus, the tricritical point is located on the line $L_{\chi\Delta \to
758: \Delta,2}$ at the point where
759: \begin{eqnarray}
760: \frac{d^4\Omega}{d\sigma^4}\left(\sigma,\Delta(\sigma)\right) & = & 0,
761: \label{trimix}
762: \end{eqnarray}
763: which is the analog of the condition $2 B(T^2-\mu^2)-N_c=0$ for the pure
764: chiral case. We have solved Eq.~(\ref{trimix}) numerically to determine
765: $(\mu_{3\rm m},T_{3 \rm m})$ in Figs.~\ref{f:panel5} and~\ref{f:panel6}.
766:
767: {\bf Disappearance of the chiral phase.} ($\alpha_2(N_c) \le B/A$; see
768: Figs.~6 and~10). The wedge of mixed broken symmetry gains space in the
769: $(\mu,T)$ plane at the expense of the chiral phase as $B/A$ increases above
770: $\alpha_2(N_c)$. The chiral phase eventually shrinks to a vertical line along
771: the $\mu=0$ axis when $B/A=N_c/2$, as shown in Fig.~\ref{f:panel5b}. That the
772: chiral phase is still present for $\mu=0$ can be understood as follows. When
773: $B/A =N_c/2$ and $\mu = 0$, both condensation fields appear in
774: Eq.~(\ref{Omega}) in the combination $\sigma^2+\Delta^2$. A pure diquark
775: solution $(\sigma,\Delta)=(0,\Delta_0)$ can thus always be rotated into a
776: pure chiral solution $(\sigma,\Delta)=(\Delta_0,0)$~\cite{VanJac99}, and thus
777: does not represent an independent phase. This symmetry is however no longer
778: present as soon as $\mu \neq 0$, in which case the diquark phase becomes
779: thermodynamically independent from the chiral phase.
780:
781: Recall that the ratio $B/A = N_c/2$ is the maximum ratio that our model can
782: realize~\cite{VanJac99}. It is instructive however to explore higher coupling
783: ratios, although they do not necessarily describe physical situations, by
784: setting $B/A > N_c/2$ by hand in Eq.~(\ref{Omega}). We find that higher $B/A$
785: force the diquark phase to grow in size at the expense of the mixed broken
786: symmetry phase. Figure~\ref{f:panel6} shows one example with $B/A =1.8$ and
787: $N_c = 3$. The similarities between the critical properties of the mixed
788: broken symmetry phase and those of the pure chiral broken phase are now
789: clear: Compare for instance the critical lines between Figs.~\ref{f:panel1}
790: and~\ref{f:panel6}.
791:
792: \section{Discussion}
793: \label{s:discuss}
794:
795: { \bf Baryon density discontinuity.} We have argued that the potential
796: $\Omega$ in Eq.~(8) represents the non-analytic contribution to the
797: thermodynamics which is directly associated with the breaking of chiral and
798: color symmetry. The potential $\Omega$ should therefore not be used too
799: literally in computations of the bulk properties of the phases we have
800: encountered. It may however give reasonable estimates of discontinuities near
801: a phase transition. For instance in the pure chiral case, the random matrix
802: models of Ref.~\cite{HalJac98} estimate that the baryon density $n_B = -(1/3)
803: \partial \Omega(\mu,T)/\partial \mu$ changes discontinuously at the first
804: order point along $T=0$ by an amount $\Delta n_B \sim 2.5\, n_0$. Here, $n_0
805: = 0.17\, {\rm fm}^{-3}$ is the density of normal nuclear matter. The result
806: for $\Delta n_B$ relies on an evaluation of the number of degrees of freedom
807: $N$ from instanton models and seems to be a reasonable estimate of the
808: baryon density discontinuity~\cite{HalJac98}.
809:
810: $\Delta n_B$ is modified by the presence of diquark condensation. We
811: consider single-gluon exchange with $N_c=3$, which realizes $B/A=3/4$, and
812: work in the limit $T=0$. Taking the derivative of Eq.~(\ref{DeltaOmega}) with
813: respect to $\mu$, we find the discontinuity in baryon density at the point
814: $\mu_1$ of equal pressure between chiral and color phases to be
815: \begin{eqnarray}
816: N_\Delta-N_\chi & = & \frac{2}{\mu_1} \left(1+\frac{7}{3} B \mu_1^2\right).
817: \label{deltaNcol}
818: \end{eqnarray}
819: Here, $\mu_1$ obeys $\Omega_\chi=\Omega_\Delta$ in Eq.~(\ref{DeltaOmega}); we
820: have $\mu_1 = 0.87 /\sqrt{B}$. Hence, $N_\Delta - N_\chi \sim 6.4
821: \,\sqrt{B}$ in the appropriate unit of inverse chemical potential. Were no
822: diquark condensation to occur, we would have found a discontinuity
823: \begin{eqnarray}
824: N_0-N_\chi & = & \frac{6}{\mu_1} \left( 1+ \frac{B \mu_1^2}{3} \right),
825: \end{eqnarray}
826: where $\mu_1$ is now the point of equal pressure between chiral and trivial
827: phases. We want to keep the same $B$ as in Eq.~(\ref{deltaNcol}) so as to
828: compare two situations which realize the same vacuum chiral field $\sigma_0 =
829: \sqrt{3/B}$. The condition of equal pressure in Eq.~(\ref{mu1}) gives then
830: $\mu_1=0.914/\sqrt{B}$ and we have $N_0-N_\chi \sim 8.4 \,\sqrt{B}$. Thus, we
831: find that diquark condensation reduces the discontinuity in baryon density by
832: roughly twenty five percent.
833:
834: {\bf Away from the chiral limit.} We now turn to study the effects of a small
835: quark current mass $m$ in Eq.~(\ref{Omega}). For $m \neq 0$, chiral symmetry
836: is explicitely broken, and the chiral condensate $\langle \psi^\dagger
837: \psi\rangle$ ceases to be a good order parameter. This affects the phase
838: diagrams in Figs. 1-6 in a number of ways. We consider the effect of a
839: small mass $m$ chosen so that $m \sim 10$ MeV in units for which the vacuum
840: chiral field is $\sigma \approx \sqrt{3/B} \sim 400$ MeV and illustrate a
841: few cases in Figs. 11-14.
842:
843: Figure 11 shows the limit of small ratios $B/A$ which favor chiral over
844: diquark condensation. Since $\langle \psi^\dagger \psi \rangle $ is no longer
845: a good order parameter, any two given points in the phase diagram can be
846: connected by a trajectory along which no thermodynamic discontinuity occurs.
847: It results that the second-order line $L_{\chi,2}$ in
848: Eq.~(\ref{sigma_second}) is no longer present when $m \neq 0$. There remains,
849: however, a first-order line, which ends at a regular critical point
850: $(\mu_c,T_c)$. This point can be located as follows. Along the first-order
851: line, the potential $\Omega(\sigma,0)$ has two minima of equal depth
852: separated by a single maximum. All three extrema become degenerate at the
853: critical point $(\mu_c,T_c)$, past which $\Omega(\sigma,0)$ possesses only
854: one minimum. The location of $(\mu_c,T_c)$ can thus be determined from the
855: condition that $\Omega(\sigma,0)$ scales as
856: \begin{eqnarray}
857: \Omega(\sigma,0) - \Omega(0,0) \sim (\sigma - \sigma_0)^4,
858: \end{eqnarray}
859: at $(\mu_c,T_c)$ and for small deviations $|\sigma-\sigma_0|$. The location
860: of $(\mu_c, T_c)$, as well as $\sigma_0$, are then determined by requiring
861: the first three derivatives of $\Omega(\sigma,0)$ to vanish at the critical
862: point and for $\sigma = \sigma_0$. A small mass tends to increase the
863: pressure in the low density `phase' with respect to that in the high density
864: `phase'. Its results that a mass $m$ displaces the first-order line
865: $L_{\chi,1}$ of Eq.~(\ref{Lchi1}) to higher $\mu$ by an amount linear in $m$.
866: The pressure increase also delays the onset of diquark condensation; the
867: ratio $B/A$ for which the diquark appears first increases linearly with $m$.
868:
869: Figure 12 shows the case of QCD: there is now a second-order line
870: $L_{\Delta,2}$ which separates the high temperature phase from a mixed broken
871: symmetry phase; the dominant effect of $m$ on the diquark phase in Fig.~2 is
872: to produce a small chiral field $\sigma \sim m$. The effect on the
873: thermodynamics in both the chiral and the trivial phases is, however,
874: second order in $m$, and so is the displacement of the second-order line
875: $L_{\Delta,2}$ from Fig.~2 to Fig.~12.
876:
877: The evolution of the phase diagram for higher $B/A$ parallels the evolution
878: we have outlined for $m=0$ in Figs. 3-6. The phase with a finite diquark
879: field grows until it eventually reaches the critical point $(\mu_c,T_c)$.
880: Higher ratios lead to a phase structure in which the first- and second-order
881: lines intersect as shown in Fig. 13. A wedge of another mixed broken symmetry
882: phase, initially with a large chiral field and a small diquark field, appears
883: at still higher $B/A$ on the left of the first-order line. This wedge grows
884: in size until it encounters the boundary of the other mixed broken symmetry
885: phase. The second-order lines then merge into a single continuous line as
886: shown in Fig.~14. This line is the locus of point for which
887: \begin{eqnarray}
888: \left. \frac{\partial^2 \Omega(\sigma,\Delta)}{\partial
889: \Delta^2}\right|_{\Delta = 0} = 0,
890: \end{eqnarray}
891: and on which a chiral solution with a vanishing diquark field turns into a
892: saddle-point of $\Omega(\sigma,\Delta)$. Inside this boundary, the global
893: minimum of $\Omega(\sigma, \Delta)$ describes a single mixed broken symmetry
894: phase. This phase again exhibits properties similar to those of a chiral
895: phase. It contains in particular a first-order line which ends at a critical
896: point $(\mu_{\rm cm},T_{\rm cm})$ at which the potential $\Omega$ scales as
897: $\Omega(\sigma, \Delta(\sigma)) - \Omega(0,0) \sim (\sigma-\sigma_0)^4$,
898: where $\Delta(\sigma)$ is the diquark field for fixed $\sigma$.
899:
900: To summarize, the effects of a small mass is linear for the chiral and mixed
901: broken symmetry phases and quadratic for the diquark and trivial phases. The
902: chiral field no longer represents a good order parameter, and the second-order
903: lines of vanishing second derivatives with respect to $\sigma$, i.e.
904: $L_{\chi,2}$ in Eq.~(\ref{sigma_second}) and $L_{\rm mix}$ in
905: Eq.~(\ref{mixline}), disappear. The tricritical points become regular critical
906: points while the tetracritical point in Fig.~5 disappears.
907:
908: {\bf The $N_c=2$ and $N_c \to \infty$ limits.} In order to make connection
909: with known results and to obtain some insight on the dependence of the phase
910: structure on the number of colors, it is interesting to consider the limits
911: $N_c = 2$ and $N_c \to \infty$. First of all when $N_c=2$, the Hermitian
912: matrix models realize only a single coupling ratio $B/A = 1$~\cite{VanJac99}. We
913: focus on single-gluon exchange for which the potential in
914: Eq.~(\ref{Omega}) becomes
915: \begin{eqnarray}
916: \Omega_{2}(\sigma,\Delta) & = & A (\sigma^2 +\Delta^2) -
917: \log[(\sigma-\mu)^2+\Delta^2+T^2] - \log[(\sigma+\mu)^2+\Delta^2+T^2].
918: \label{Om2}
919: \end{eqnarray}
920: A rotational symmetry appears at $\mu =0 $ as $\Omega_2$ depends on chiral
921: and diquark fields via the combination $\sigma^2+\Delta^2$. Diquark and
922: chiral phases do not in this case represent independent states. We find for
923: the combined fields that
924: \begin{eqnarray}
925: \sigma^2 + \Delta^2 & = & {2 \over A} - T^2,
926: \end{eqnarray}
927: below a critical temperature $T_c = \sqrt{2/A}$, while symmetry is restored
928: above $T_c$. This rotational symmetry is a consequence of the pseudo-reality
929: of $SU(2)$-QCD, a property by which the Dirac operator
930: $D = i \sum_{\mu a} \gamma^\mu \lambda^a A_{\mu a} + m$ commutes with $\tau_2
931: C\gamma^5 K$ where $\tau_2$ is the antisymmetric $2\times 2$ color matrix and
932: $K$ the complex conjugation operator. For $\mu = 0$, this property permits
933: one to arrange color and flavor symmetries into a higher $SU(4)$ symmetry,
934: see for instance~\cite{DiaPet92,SmiVer95,KogSte99}.
935:
936: At finite $\mu$, however, the $SU(4)$ symmetry is explicitly broken.
937: The global minimum of $\Omega(\sigma,\Delta)$ always has $\sigma = 0$, and
938: the system prefers diquark condensation over chiral symmetry breaking.
939: This results agrees with instanton models~\cite{CarDia99,DiaFor96}, and
940: lattice calculations~\cite{DagKar86}. We now have a second-order phase
941: transition from an ordered state with $\Delta^2 = 2/A -\mu^2 - T^2$ in the
942: low $T$ and $\mu$ region to a symmetric phase at high $T$ and $\mu$.
943:
944: The opposite limit $N_c \to \infty$ is more subtle. In microscopic models, it
945: is expected on general grounds that the quark-quark interaction is suppressed
946: with respect to the $\bar qq$ channel by powers of
947: $1/N_c$~\cite{DiaFor96,tHo73}. In the present model, we observe that diquark
948: condensation disappears as $N_c \to \infty$ if the interaction is
949: single-gluon exchange. Its coupling ratio $B/A = N_c/(2 N_c - 2) \to 1/2$ as
950: $N_c \to \infty$. Thus, we have $B/A \ll 0.139 N_c$ and the only possible
951: topology for the phase diagram is that in Fig.~1. Hence, no diquark
952: condensate forms. Other Hermitean interactions with $B/A \sim O(N_c)$ can
953: however explore the full range of phase diagrams which display diquark
954: condensation. The actual number of possible toplogies is however reduced to
955: five, as $\alpha_2(N_c)\to N_c/\sqrt{8}$ for $N_c \to \infty$ (see Appendix)
956: and Fig.~4 can no longer be realized.
957:
958: We summarize the variation with $N_c$ of the coupling ratios characterizing a
959: change in topology and of the ratio realized by single-gluon exchange in
960: Table 1.
961:
962: {\bf Comparison with a microscopic model:} In contrast to microscopic models,
963: the random matrix interaction does not lead to a logarithmic instability of
964: the gap equation near the Fermi surface. To clarify this effect, we consider
965: diquark condensation in a chiral symmetric phase at $T=0$ and
966: compare the random matrix approach to a microscopic model. We choose
967: here the NJL study of Berges and Rajagopal~\cite{BerRaj99}. In the random
968: matrix formulation the gap equation
969: \begin{eqnarray}
970: 2 A \Delta & = & {\Delta \over \Delta^2 + \mu^2}
971: \label{BCSgapRMM}
972: \end{eqnarray}
973: has a non-trivial root $\Delta \sim \sqrt{\mu_c - \mu}$ with
974: $\mu_c = \sqrt{2/A}$. The model of~\cite{BerRaj99} gives an equation of the
975: form
976: \begin{eqnarray}
977: 2 A' \Delta & = & \Delta \int_0^\infty dq q^2 F^4(q) \left({1\over
978: \sqrt{(q-\mu)^2+\Delta^2 F^4(q)}}+
979: {1\over \sqrt{(q+\mu)^2+\Delta^2 F^4(q)}}\right),
980: \label{BCSgap}
981: \end{eqnarray}
982: where $F(q)$ is an appropriate form factor which falls off over $q \sim
983: \Lambda \sim O(\Lambda_{\rm QCD})$. Clearly, the right side of
984: Eq.~(\ref{BCSgap}) contains a singularity for $q \sim \mu$ and $\Delta \to 0$
985: which is absent in Eq.~(\ref{BCSgapRMM}). This singularity has two
986: consequences. First, the behavior of the diquark field for large $\mu$
987: depends sensitively on the form factor. As $\Delta \to 0$, the singularity at
988: $q \sim \mu$ in the right side of Eq.~(\ref{BCSgap}) gives $ 2 A' \approx
989: \mu^2 F^4(\mu) \log \Lambda \mu /\Delta^2$ to logarithmic order. Thus,
990: instead of the square root behavior $\Delta(\mu)\sim (\mu_c - \mu)^{1/2}$ of
991: the random matrix approach, $\Delta$ now vanishes as $\mu \to \infty$ as
992: $\Delta(\mu) \propto \exp(- c/(\mu^2 F^4(\mu)))$, where $c$ is a
993: constant.~\footnote{This result is only true for a smooth cutoff $F(q)$. For
994: a sharp cutoff $F(q)=\Theta(\Lambda - \mu)$, where $\Theta(x)$ is the
995: Heavyside function, the diquark field exhibits the square root
996: singularity.} This tail is exponentially sensitive to form factors; with
997: regards to establishing the general topology of the phase diagram, such
998: behavior is not qualitatively different from $\Delta = 0$. The second
999: consequence of the logarithmic singularity in Eq.~(\ref{BCSgap}) is that
1000: $\Delta(\mu)$ must be non-monotonic for intermediate $\mu$. This profile can
1001: be seen by drawing in the plane $(\Delta,\mu)$ the lines of constant height
1002: for the right side of Eq.~(\ref{BCSgap}): these lines must go around the
1003: point $(\mu_0,0)$ where the logarithmic singularity is the strongest as
1004: $\Delta \to 0$, and $\Delta(\mu)$ reaches a maximum at $\mu_0$.
1005:
1006: {\bf Color dependency in the chiral fields.} The patterns of symmetry
1007: breaking which we have described can become even richer if we allow the
1008: chiral fields to depend on color. This possibility arises in the instanton
1009: model of Carter and Diakonov~\cite{CarDia99}, who remarked that the
1010: Dyson-Gorkov equations close in color space on the condition that gapped and
1011: ungapped quarks can develop different masses. We studied the effects of this
1012: additional degree of freedom in the limit of zero chemical
1013: potential~\cite{VanJac99} and found no change in the phase structure for
1014: ratios $B/A \le N_c/2$. The situation is different for finite $\mu$.
1015: Choosing different masses for the two gapped and the $N_c-2$ ungapped quarks
1016: leads in some limits to an increase in the pressure of phases with finite
1017: diquark fields. The main consequence is an increase of both the parameter
1018: range and the region in the $(\mu,T)$ plane for which the mixed broken
1019: symmetry phase exists. This result is obvious since the mixed broken symmetry
1020: phase is the only phase which can exploit this additional degree of freedom.
1021: However, there is also a general mechanism at work by which a change of phase
1022: structure always takes place above a certain threshold for the coupling
1023: constants.
1024:
1025: We wish to illustrate these effects in a few cases and concentrate on the
1026: limit $N_c =3$ and $m = 0$ for clarity. We denote the chiral fields for color
1027: $1$ and $2$ by $\sigma_1$ and that for color $3$ by $\sigma_3$. In order to
1028: permit these fields to be different, we include the projection of the
1029: interaction onto a chiral-$\lambda_8$ channel as described
1030: in~\cite{VanJac99}. For $\mu\neq 0$, the thermodynamical potential becomes
1031: \begin{eqnarray}
1032: \Omega (\sigma_1,\sigma_3,\Delta) &=& A \Delta^2 + B(\beta_1
1033: \sigma_1^2+\beta_2 \sigma_1 \sigma_3 + \beta_3 \sigma_3^2)
1034: - \log [(\sigma_1 - \mu)^2 + \Delta^2 + T^2]
1035: - \log [(\sigma_1 + \mu)^2 + \Delta^2 + T^2] \nonumber \\
1036: && - {1 \over 2} \log [ (\sigma_3 -\mu)^2 + T^2 ] -{1\over 2} \log[(\sigma_3 +
1037: \mu)^2 + T^2],
1038: \label{Omegasplit}
1039: \end{eqnarray}
1040: where
1041: \begin{eqnarray}
1042: \beta_1 = {4\over 9} + {C \over 3 B}, \quad
1043: \beta_2 = {4 \over 9} - {2 C \over 3 B}, \quad
1044: \beta_3 = {1 \over 9} + {C \over 3 B},
1045: \label{beta123}
1046: \end{eqnarray}
1047: describe the coupling between $\sigma_1$ and $\sigma_3$. $C$ is a coupling
1048: constant associated with the chiral-$\lambda_8$ channel~\cite{VanJac99}. By
1049: fine tuning the various Lorentz and color channels which compose the random
1050: matrix interactions, we can realize a range of ratios $B/C$ for any fixed
1051: $B/A$.~\footnote{This range is $-3/16 \le B/C \le 3/2$ for $B/A < 3/4$ and
1052: then reduces linearly to $B/C = 3/2$ for $B/A=3/2$.} New patterns of
1053: symmetry breaking only appears when $C > 0$ and the chiral-$\lambda_8$
1054: channel is attractive~\cite{VanJac99}.
1055:
1056: Consider first single-gluon exchange. The coupling ratios are $B/A = 3/4$ and
1057: $C/B = -3/16$. The chiral-$\lambda_8$ channel is repulsive, and the phases
1058: with the largest pressure always satisfy $\sigma_1 = \sigma_3$, which is the
1059: case shown in Fig.~2. If we now fine tune the interactions so as to increase
1060: $B/C$ to $B/C \sim 3/2$ while keeping $B/A = 3/4$, the phase diagram changes
1061: substantially. The first-order line between the chiral and trivail phases
1062: splits at the point where it meets the diquark transition line into two
1063: first-order lines. Together with the $T=0$ axis, they delimitate a wedge of
1064: mixed broken symmetry phase with $\sigma_1 \neq \sigma_3$ and $\Delta \neq
1065: 0$. Thus, with an attractive channel $B/C \sim 3/2$ the mixed broken
1066: symmetry phase appears much earlier than previously discussed. This is an
1067: extreme case. For fixed $B/A$, the mixed broken symmetry phase does not
1068: appear immediately as $B/C$ increases from $0$: there is a threshold value
1069: above which the new phase appears (this value is $B/C \sim 1.29$ for
1070: $B/A=0.75$). Therefore, it takes large variations of the coupling constants
1071: $B/A$ and $B/C$ away from the values expected for single-gluon exchange to
1072: modify the phase structure of Fig.~2.
1073:
1074: To complete the picture of the effects of an attractive channel $C>0$,
1075: consider next $B/A = 3/2$. This ratio can only be realized by a color
1076: diagonal interaction for which $B/C = 3/2$~\cite{VanJac99}. In this case,
1077: there is no freedom in fine tuning $B/C$ to other values. The effect of
1078: splitting masses in colors is now maximal. $\beta_2$ in Eq.~(\ref{beta123})
1079: vanishes and $\sigma_3$ decouples from the other two fields. Its gap equation
1080: leads to the same solutions as a pure chiral phase with $\sigma \neq 0$ and
1081: $\Delta =0$ and the phase diagram for the third color is that of Fig.~1. For
1082: $B/A = B/C =3/2$, the partial pressure for colors $1$ and $2$ has the same
1083: form as in the limit $N_c = 2$, see Eq.~(\ref{Om2}). There is then a
1084: rotational symmetry between $\sigma_1$ and $\Delta$ for $\mu = 0$ and
1085: $\sigma_1$ vanishes for $\mu \neq 0$ while $\Delta = 2/A -\mu^2 -T^2$ for
1086: $\mu^2+T^2 \le 2/A$. The overall phase diagram is obtained by superposing the
1087: two pictures; we have a mixed broken symmetry phase with $\sigma_1=0$,
1088: $\sigma_3 \neq 0$, and $\Delta \neq 0$ on the left of the first-order line
1089: $L_{\chi,1}$ of Eq.~(\ref{Lchi1}). This phase is contiguous to a pure diquark
1090: phase which develops as in Fig.~2 on the right of $L_{\chi,1}$ and below the
1091: semicircle. With respect to Fig.~5, the mixed broken symmetry phase thus
1092: occupies a wider area in the $(\mu,T)$ plane.
1093:
1094: {\bf A color-$6$ condensate.} It has been suggested that an energy gain may
1095: result if the third color condenses in a spin-$1$ color symmetric
1096: state~\cite{AlfRaj98}. We find no such condensation for single-gluon
1097: exchange. We can however fine tune the interactions so as to keep $B/A =3/2$
1098: and allow this channel to develop. The Fierz constant for projecting on a
1099: color $6$ is half that for a $\bar 3$ color state. Thus, we can
1100: understand qualitatively how a color-$6$ behaves by repeating our previous
1101: analysis with an additional phase which now may develop inside a semicircle
1102: of radius $\sqrt{2}$ smaller than the radius of the diquark phase. This
1103: semicircle crosses the first-order line beween chiral and diquark phases at
1104: $B/A \sim 0.754$ and a color-$6$ phase can, in favorable cases, increase the
1105: pressure of the system. However, as far as QCD is concerned, a color-$6$
1106: condensate should be very small since its threshold ratio is very close to
1107: $B/A = 3/4$. This is an example of a result which cannot be regarded as
1108: robust: in a microscopic theory, the fate of the color-$6$ phase will
1109: inevitably depend on the details of the interaction and on whether these give
1110: rise to exponential tails for the associated condensation field. By contrast,
1111: the chiral and diquark phases are fully developed for $B/A =3/4$ and the
1112: phase diagram in Fig.~2 should thus be considered robust.
1113:
1114:
1115: \section{Conclusions}
1116: \label{s:conclusions}
1117:
1118: Our random matrix model leads to a thermodynamic potential which has a very
1119: simple form. Yet, it contains enough physics to illustrate in a clear way the
1120: interplay between chiral symmetry breaking and the formation of quark Cooper
1121: pairs. We have found that this interplay results in a variety of phase
1122: diagrams which can be characterized by a total of only six different
1123: topologies. Single-gluon exchange leads to the topology shown in Fig.~2, a
1124: phase diagram familiar from microscopic models. We have considered the
1125: chiral and scalar diquark channels, which seem the most promising ones. We
1126: have found that chiral and diquark phases are fully developped for the ratio
1127: realized by single-gluon exchange and that it takes large variations in the
1128: coupling ratios $B/A$ and $C/B$ to depart from that result. On the other
1129: hand, other less attractive condensation channels seem sensitive to coupling
1130: constant ratios and are expected in microscopic models to depend on the
1131: details of the interactions. Furthermore, these channels develop
1132: weak condensates at best. It is worth keeping in mind that the present picture
1133: is mean-field; we expect that quantum fluctuations will inevitably have large
1134: effects on the weak channels, which should thus not be considered robust. We
1135: conclude that QCD with two light flavors should realize the topology
1136: suggested by single-gluon exchange and that this topology is stable against
1137: variations in the detailed form of the microscopic interactions.
1138:
1139: Many effects lie in the mismatch between the number of colors involved in
1140: each order parameter: all $N_c$ colors contribute to the chiral field, while
1141: a Cooper pair involves two colors. A result of this for $N_c = 3$ is that the
1142: chiral broken phase prevails at low densities. Furthermore, our model
1143: reproduces to a reasonable extent the expected limits at small and large
1144: $N_c$. We expect such counting arguments to be valid in both microscopic
1145: models and lattice calculations. More generally, we believe that random
1146: matrix models can provide insight into calculations of QCD at finite density
1147: by providing simple illustrations for many of the mechanisms which are at
1148: work.
1149:
1150: Not all the phase diagrams that we have studied are directly relevant to QCD.
1151: However, many of their characteristics such as the presence of tricritical
1152: and tetracritical points are generic to systems in which two forms of order
1153: compete. Our model could naturally be extended to the study of nuclear or
1154: condensed matter systems in which such competion takes place. The
1155: construction of a genuine theory seems technically involved at first
1156: glance~\cite{VanJac99} but in fact contains only three basic ingredients: the
1157: identification of the symmetries at play and their associated order
1158: parameters, the knowledge of the elementary excitations in a background of
1159: condensed fields, and the calculation of the range of coupling constants
1160: realized by the random matrix interactions. These three components are
1161: sufficient to build a thermodynamical potential and determine the resulting
1162: phase structures in parameter space.
1163:
1164: \section*{Acknowledgements}
1165:
1166: We are grateful for stimulating discussions with G. Carter, D. Diakonov, H.
1167: Heiselberg, and K. Splittorff.
1168:
1169: \appendix
1170: \section{Calculation of $\alpha_1(N_c)$ and $\alpha_2(N_c)$}
1171: \label{a:alpha1}
1172:
1173: In this appendix, we calculate the ratios $\alpha_1(N_c)$ and
1174: $\alpha_2(N_c)$, which respectively characterize the onset of the
1175: $\chi\Delta$-phase and the appearance of the tetracritical point.
1176:
1177: \subsection{The ratio $\alpha_1(N_c)$.}
1178:
1179: The mixed broken symmetry phase appears first for the coupling ratio $B/A$
1180: for which the instability line $L_{\rm mix}$ of Eq.~(\ref{mixline}) crosses
1181: the first-order line between chiral and diquark phases. This crossing occurs
1182: on the $T=0$ axis. The line $L_{\rm mix}$ meets the $T=0$ axis at $\mu =
1183: \mu_{\rm mix}$, Eq.~(\ref{mumix}),
1184: \begin{eqnarray}
1185: \mu_{\rm mix}^2 = \frac{N_c^2 A}{4 B^2} - \frac{N_c}{2 B}, \nonumber
1186: \end{eqnarray}
1187: while the condition of equal pressure at that point gives
1188: \begin{eqnarray}
1189: \Omega_\Delta - \Omega_\chi = (N_c - 2)+(A+B) \mu_{\rm mix}^2-2 \log \frac{A
1190: \mu_{\rm mix}^2}{2}+ N_c \log \frac{B \mu_{\rm mix}^2}{N_c} = 0.
1191: \end{eqnarray}
1192: Combining these two equations gives the determining equation for
1193: $\alpha_1 = B/A$,
1194: \begin{eqnarray}
1195: (N_c-2) & + & \frac{N_c}{2 \alpha_1^2} \left(\frac{N_c}{2}-\alpha_1\right)
1196: \left(1+\alpha_1\right)
1197: - 2 \log \left[\frac{N_c}{4 \alpha_1} \left(\frac{N_c}{2 \alpha_1}-1\right)
1198: \right]
1199: + N_c \log \left[\frac{1}{2}\left(\frac{N_c}{2 \alpha_1} -1\right)\right] = 0.
1200: \end{eqnarray}
1201: This is a transcendental relation which needs to be solved numerically for
1202: each $N_c$. In particular, we find
1203: $\alpha_1(2)=1$, $\alpha_1(3)=1.05$,
1204: and $\alpha_1(N_c \to \infty)\sim 0.321 N_c$.
1205:
1206: \subsection{The ratio $\alpha_2(N_c)$.}
1207:
1208: Two conditions determine $\alpha_2(N_c)$. Coming from small $B/A$,
1209: $\alpha_2(N_c)$ corresponds to the ratio at which the wedge of the
1210: $\chi\Delta$-phase reaches the critical point $(\mu_4,T_4)$ of
1211: Eq.~(\ref{tetracritical}). Decreasing $B/A$ from large values,
1212: $\alpha_2(N_c)$ marks the coexistence of the tetracritical point
1213: $(\mu_4,T_4)$ and the tricritical point $(\mu_{3\rm m},T_{3\rm m})$. We now
1214: show that these two conditions are equivalent, and thus that the transition
1215: from Fig.~4 to Fig.~5 is continuous.
1216:
1217: To proceed, we concentrate on the region
1218: near $(\mu_4,T_4)$ where we perform a small
1219: field expansion of the potential $\Omega(\sigma,\Delta)$ in
1220: Eq.~(\ref{Omega}),
1221: \begin{eqnarray}
1222: \Omega(\sigma,\Delta) & \approx & \Omega(0,0) + a_0\, \sigma^2 + b_0\, \Delta^2+
1223: \frac{a_1^2}{2}\, \sigma^4 + \frac{b_1^2}{2}\, \Delta^4 +c_1 \,\sigma^2
1224: \, \Delta^2 + {\cal O} ({\rm min}^6(\Delta,\sigma)).
1225: \label{Omegaexpanded}
1226: \end{eqnarray}
1227: Here, the coefficients of the quadratic terms are linear in $\delta \mu = \mu
1228: -\mu_4$ and $\delta T = T-T_4$ and vanish at $(\mu_4,T_4)$, while those of
1229: the quartic terms are
1230: \begin{eqnarray}
1231: a_1^2 = \frac{2 B^2}{N_c}-\frac{A^2 N_c}{4},\quad
1232: b_1^2 = \frac{A^2}{2}, \quad
1233: c_1 =\frac{2 A B}{N_c} -\frac{A^2}{2},
1234: \label{a1b1c1}
1235: \end{eqnarray}
1236: to leading order in $\delta\mu$ and $\delta T$.
1237:
1238: The condition that the $\chi\Delta$-phase includes $(\mu_4,T_4)$ can be
1239: determined as follows. While $B/A$ increases from below $\alpha_2(N_c)$, the
1240: tip of the wedge of the $\chi\Delta$-phase slides along the the line of equal
1241: pressures between chiral and diquark phases, which we denote by $L_{\chi \to
1242: \Delta,1}$. Minimizing $\Omega(\sigma,\Delta)$ in
1243: Eq.~(\ref{Omegaexpanded}) to find the respective pressure in the $\chi$- and
1244: $\Delta$-phases, this line is determined near $(\mu_4,T_4)$ by
1245: \begin{eqnarray}
1246: L_{\chi\to \Delta,1} \approx \frac{b_0^2}{b_1^2} - \frac{a_0^2}{a_1^2} = 0.
1247: \label{firstnear4}
1248: \end{eqnarray}
1249: The left boundary of the $\chi\Delta$-phase is the instability line
1250: \begin{eqnarray}
1251: L_{\rm mix} \equiv
1252: \left.
1253: \frac{\partial^2 \Omega}{\partial \Delta^2}
1254: \right|_{\Delta = 0,\sigma = \sigma_0} = 0,
1255: \end{eqnarray}
1256: where $\sigma_0$ is the $\sigma$ field in the chiral phase. From Eq.~(\ref{Omegaexpanded}), this gives near $(\mu_4,T_4)$
1257: \begin{eqnarray}
1258: L_{\rm mix} \approx b_0 - \frac{a_0}{a_1^2}{c_1}=0.
1259: \label{Lmixnear4}
1260: \end{eqnarray}
1261: The tip $P_{\rm mix}$ of the $\chi\Delta$-phase is the intercept of $L_{\rm
1262: mix}$ and $L_{\chi\to\Delta,1}$. From Eqs.~(\ref{firstnear4}) and
1263: (\ref{Lmixnear4}) $P_{\rm mix}$ thus obeys
1264: \begin{eqnarray}
1265: a_0^2 \left(c_1^2 - a_1^2 b_1^2 \right) = 0.
1266: \label{alpha2eq}
1267: \end{eqnarray}
1268: Now, imagine that $B/A$ is strictly smaller but near
1269: $\alpha_2(N_c)$. Then, $P_{\rm mix} \neq (\mu_4,T_4)$, and we have $a_0 \neq
1270: 0$. The location of $P_{\rm mix}$ is therefore determined by the condition
1271: $a_1^2 b_1^2 = c_1^2$, where the coefficients $a_1$,
1272: $b_1$ and $c_1$ are those of Eq.~(\ref{a1b1c1}) augmented by linear
1273: corrections in $\delta \mu$ and $\delta T$. If we now let $B/A \to
1274: \alpha_2(N_c)$, the tip $P_{\rm mix}$ reaches $(\mu_4,T_4)$ and the linear
1275: corrections in $a_1$, $b_1$, and $c_1$ vanish. Using Eq.~(\ref{a1b1c1}), and
1276: solving $a_1^2 b_1^2 = c_1^2$ for $B/A$ gives
1277: \begin{eqnarray}
1278: \alpha_2(N_c)={B \over A} \equiv \frac{4 N_c - N_c^{3/2}
1279: \sqrt{2 N_c-4}}{4 (4 - N_c)}, \nonumber
1280: \end{eqnarray}
1281: which is the result stated in Eq.~(\ref{alpha2}).
1282:
1283: Coming from large ratios $B/A$, the determining condition for $B/A =
1284: \alpha_2(N_c)$ requires that $(\mu_4,T_4)$ obeys Eq.~(\ref{trimix}),
1285: \begin{eqnarray}
1286: \frac{d^4\Omega}{d\sigma^4}(\sigma,\Delta(\sigma)) & = & 0, \nonumber
1287: \end{eqnarray}
1288: where $\Delta(\sigma)$ is the diquark field in the $\chi\Delta$-phase. Taking
1289: the derivative of Eq.~(\ref{Omegaexpanded}) with respect to $\Delta$, we find
1290: $\Delta^2(\sigma) \approx - (c_1 \sigma^2 +b_0)/b_1^2$ which when inserted
1291: back in Eq.~(\ref{Omegaexpanded}) gives
1292: \begin{eqnarray}
1293: \frac{d^4\Omega}{d\sigma^4}(\sigma,\Delta(\sigma)) & \approx & a_1^2 -
1294: \frac{c_1^2}{b_1^2},
1295: \end{eqnarray}
1296: in the neighborhood of $(\mu_4,T_4)$. Setting the right side to zero, we
1297: recover the previous condition of Eq.~(\ref{alpha2eq}).
1298:
1299:
1300: \newpage
1301:
1302: \begin{figure}[htb]
1303: \epsfxsize=\figsize
1304: \centerline{
1305: \epsffile{pd_fig1.eps}}
1306: \vspace{.5cm}
1307: \caption{
1308: Phase diagram for $B/A \le 0.418$ and $N_c=3$, as a function of the quark
1309: chemical potential $\mu$ and the temperature $T$. For this and all the
1310: following figures, continuous curves represent second-order lines while
1311: first-order lines are plotted with dots. Here, $\chi$ is the chiral phase
1312: and $\Delta$ is the diquark phase. First- and second-order lines join
1313: tangentially at the tricritical point $(\mu_3,T_3)$.}
1314: \label{f:panel1}
1315: \end{figure}
1316:
1317: \vspace{2cm}
1318:
1319: \begin{figure}[hb]
1320: \epsfxsize=\figsize
1321: \centerline{
1322: \epsffile{pd_fig2.eps}}
1323: \vspace{0.5cm}
1324: \caption{Phase diagram for the coupling ratio realized by single-gluon
1325: exchange, $B/A=0.75$. The transition from the chiral phase $\chi$ to the
1326: diquark phase $\Delta$ is first-order while that from the $\Delta$ to the
1327: trivial phase is second-order.}
1328: \label{f:panel2}
1329: \end{figure}
1330:
1331:
1332:
1333: \begin{figure}[ht]
1334: \epsfxsize=\figsize
1335: \centerline{
1336: \epsffile{pd_fig3.eps}}
1337: \vspace{.5cm}
1338: \caption{ Phase diagram for a ratio $B/A=1.054$. The mixed broken symmetry
1339: phase $\chi\Delta$ emerges out of the chiral phase via a second-order
1340: transition at $\mu_{\rm mix}$, and undergoes a first-order transition
1341: towards the diquark phase at $\mu_{1}$.}
1342: \label{f:panel3}
1343: \end{figure}
1344:
1345: \vspace{2cm}
1346:
1347: \begin{figure}[hb]
1348: \epsfxsize=\figsize
1349: \centerline{
1350: \epsffile{pd_fig4.eps}}
1351: \vspace{.5cm}
1352: \caption{
1353: Phase diagram for a ratio $B/A=1.1$. Compared to
1354: Fig.~3,
1355: the tricritical point $(\mu_3,T_3)$ no longer exists.
1356: The trivial, chiral, and diquark phases now meet at the new critical point
1357: $(\mu_4,T_4)$. We note in particular that the first and second-order lines
1358: bordering the chiral phase do not meet tangentially at $(\mu_4,T_4)$.}
1359: \label{f:panel4}
1360: \end{figure}
1361:
1362:
1363:
1364: \begin{figure}[ht]
1365: \epsfxsize=\figsize
1366: \centerline{
1367: \epsffile{pd_fig5.eps}}
1368: \vspace{.5cm}
1369: \caption{Phase diagram for a ratio $B/A=1.4$. Compared to
1370: Fig.~4, $(\mu_4,T_4)$ has become a tetracritical point, at the
1371: intersection of the four phases. There is also a new tricritical point
1372: $(\mu_{3\rm m},T_{3\rm m})$ where the first- and second-order line
1373: separating the $\chi\Delta$- and the $\Delta$-phases join tangentially.}
1374: \label{f:panel5}
1375: \end{figure}
1376:
1377: \vspace{2cm}
1378:
1379: \begin{figure}[hb]
1380: \epsfxsize=\figsize
1381: \centerline{
1382: \epsffile{pd_fig6.eps}}
1383: \vspace{.5cm}
1384: \caption{Phase diagram for a ratio $B/A=1.8$. This topology is characteristic
1385: of large ratios $B/A$, which favor diquark over chiral condensation. The
1386: $\Delta$-phase occupies a large part of the phase diagram, at the expense
1387: of the mixed broken symmetry phase. The chiral phase has completely
1388: vanished. We note that since Hermitian interactions realize ratios $B/A <
1389: 1.5$, this last topology actually describes a case out of reach for our
1390: model.}
1391: \label{f:panel6}
1392: \end{figure}
1393:
1394:
1395:
1396: \begin{figure}[hb]
1397: \epsfxsize=\figsize
1398: \centerline{
1399: \epsffile{fields.eps}}
1400: \caption{Order parameters for the coupling ratio of single-gluon exchange,
1401: $B/A=0.75$. The corresponding phase diagram is shown in Fig.~2. Chiral and
1402: diquark fields vanish continuously along the second-order lines.}
1403: \label{f:fields}
1404: \end{figure}
1405:
1406: \vspace{2cm}
1407:
1408: \begin{figure}[ht]
1409: \epsfxsize=\figsize
1410: \centerline{
1411: \epsffile{pd_fig3c.eps}
1412: }
1413: \caption{Expanded view of Fig.~3 near the wedge of mixed broken symmetry.
1414: We note that the second-order line between the $\chi$- to the $\chi\Delta$
1415: phase does not meet tangentially with the first-order line between the
1416: $\chi\Delta$- and the $\Delta$-phases, as indicated by the short dashed
1417: line.}
1418: \label{f:panel3b}
1419: \end{figure}
1420:
1421:
1422:
1423: \begin{figure}[hb]
1424: \epsfxsize=\figsize
1425: \centerline{
1426: \epsffile{pd_fig3b.eps}
1427: }
1428: \caption{
1429: Expanded view of Fig.~3 near the tricritical point $(\mu_3,T_3)$, showing
1430: that there is a very short first-order segment between the chiral and
1431: trivial phases. }
1432: \label{f:panel3c}
1433: \end{figure}
1434:
1435: \vspace{2cm}
1436:
1437: \begin{figure}[ht]
1438: \epsfxsize=\figsize
1439: \centerline{
1440: \epsffile{pd_fig5b.eps}}
1441: \vspace{.5cm}
1442: \caption{Phase diagram for the maximal ratio $B/A=1.5$ that can be
1443: realized by the random matrix interactions. With respect to Fig.~5, the
1444: thermodynamic competition favors the mixed broken symmetry phase
1445: at the expense of the chiral phase.}
1446: \label{f:panel5b}
1447: \end{figure}
1448:
1449:
1450:
1451: \begin{figure}[ht]
1452: \epsfxsize=\figsize
1453: \centerline{
1454: \epsffile{pd_fig1m.eps}}
1455: \vspace{.5cm}
1456: \caption{Phase diagram for a ratio $B/A =0.2$ and a small quark mass
1457: $m$.}
1458: \label{f:fig1m}
1459: \end{figure}
1460:
1461: \vspace{2cm}
1462:
1463: \begin{figure}[ht]
1464: \epsfxsize=\figsize
1465: \centerline{
1466: \epsffile{pd_fig2m.eps}}
1467: \vspace{.5cm}
1468: \caption{ Phase diagram for the coupling ratio of single-gluon
1469: exchange, $B/A = 0.75$, and a small quark mass $m \sim 10$ {MeV}.
1470: The first-order line ends at a critical point $(\mu_c,T_c)$.}
1471: \label{f:fig2m}
1472: \end{figure}
1473:
1474:
1475:
1476: \begin{figure}[ht]
1477: \epsfxsize=\figsize
1478: \centerline{
1479: \epsffile{pd_fig3m.eps}}
1480: \vspace{.5cm}
1481: \caption{Phase diagram for $B/A=1.0$ and a small quark mass $m \sim 10$
1482: MeV.}
1483: \label{f:fig3m}
1484: \end{figure}
1485:
1486: \vspace{2cm}
1487:
1488:
1489: \begin{figure}[ht]
1490: \epsfxsize=\figsize
1491: \centerline{
1492: \epsffile{pd_fig4m.eps}}
1493: \vspace{.5cm}
1494: \caption{Phase diagram for $B/A = 1.4$ and a small quark mass $m$. The
1495: first-order line ends at a critical point $(\mu_{\rm cm},T_{\rm cm})$.}
1496: \label{f:fig4m}
1497: \end{figure}
1498:
1499:
1500: \newpage
1501:
1502: \begin{tabular}{l p{1em} r @{.} l p{1em} r @{.} l p{1em} r @{.} l l }
1503: \multicolumn{11}{c}{Table 1. A few characteristic coupling ratios.} \\
1504: \hline
1505: \hline
1506: Ratio $B/A$ & &
1507: \multicolumn{2}{c}{$N_c = 2$} &
1508: &
1509: \multicolumn{2}{c}{$N_c = 3$} &
1510: &
1511: \multicolumn{3}{l}{$N_c \to \infty$} \\
1512: \hline
1513: Onset of diquark condensation & & \multicolumn{2}{l}{} &
1514: &
1515: 0 & 418 &
1516: &
1517: 0 & 139 & $N_c$ \\
1518: $\alpha_1(N_c)$ & & 1 & 0 &
1519: &
1520: 1 & 050 &
1521: &
1522: 0 & 321 & $N_c$ \\
1523: Disappearance of the tricritical point. & & \multicolumn{2}{l}{} &
1524: &
1525: 1 & 061 &
1526: &
1527: 0 & 354 & $N_c$ \\
1528: $\alpha_2(N_c)$ & & 1 & 0 &
1529: &
1530: 1 & 163 &
1531: &
1532: 0 & 354 & $N_c$ \\
1533: Maximum ratio $N_c/2$ & & 1 & 0 &
1534: &
1535: 1 & 5 &
1536: &
1537: 0 & 5 & $N_c$ \\
1538: Single-gluon exchange & & 1 & 0 &
1539: &
1540: 0 & 75 &
1541: &
1542: 0 & 5 \\
1543: \hline
1544: \hline
1545: \end{tabular}
1546:
1547:
1548:
1549:
1550: \begin{thebibliography}{99}
1551: \bibitem{Bar77}
1552: B. Barrois, Nucl. Phys. {\bf B129}, 390 (1977);
1553:
1554: \bibitem{BaiLov84}
1555: D. Bailin and A. Love, Phys. Rept. {\bf 107}, 325 (1984).
1556:
1557: \bibitem{AlfRaj98}
1558: M. Alford, K. Rajagopal, and F. Wilczek, Phys. Lett. B {\bf 422}, 247 (1998).
1559:
1560:
1561: \bibitem{RapSch98}
1562: R. Rapp, T. Sch{\"a}fer, E. V. Shuryak, and M. Velkovsky,
1563: Phys. Rev. Lett. {\bf 81}, 53 (1998); hep-ph/9904353.
1564:
1565:
1566:
1567: \bibitem{CarDia99}
1568: G. W. Carter and D. Diakonov, Phys. Rev. D {\bf 60}, 016004 (1999).
1569:
1570: \bibitem{BerRaj99}
1571: J. Berges and K. Rajagopal, Nucl. Phys. {\bf B538}, 215 (1999).
1572:
1573: \bibitem{SchKle99}
1574: T. M. Schwarz, S. P. Klevansky, and G. Papp, Phys. Rev. C {\bf 60},
1575: 055205 (1999).
1576:
1577:
1578: \bibitem{low}
1579: For studies of diquark condensation at low densities, see also S. Ying,
1580: Phys. Lett B {\bf 285}, 341 (1992); Ann. Phys. (N. Y.) {\bf 250}, 69 (1996);
1581: S. P\'epin, M. C. Birse, J. A. McGovern, and N. R. Walet, hep-ph/9912475.
1582:
1583: \bibitem{VanJac99}
1584: B. Vanderheyden and A. D. Jackson, Phys. Rev. D {\bf 61}, 076004 (2000).
1585:
1586: \bibitem{Biel}
1587: See the procceedings of {\it QCD at finite baryon densities}, Bielfeld,
1588: April 1998, F. Karsh and M. P. Lombardo (Eds.), Nucl. Phys. {\bf A642} (1998).
1589:
1590:
1591: \bibitem{Ste96}
1592: M. A. Stephanov, Phys. Rev. Lett. {\bf 76}, 4471 (1996);
1593: Nucl. Phys. Proc. Suppl. {\bf 53}, 469 (1997).
1594:
1595: \bibitem{KogMat83}
1596: J. Kogut, H. Matsuoka, M. Stone, H.W. Wyld, S. Shenker, J. Shigemitsu, and
1597: D. K. Sinclair, Nucl. Phys. {\bf B225} [FS9], 93 (1983).
1598:
1599: \bibitem{BarBeh86}
1600: I. M. Barbour, N. Behlil, E. Dagotta, F. Karsch, A. Moreao, M. Stone, and
1601: H. W. Wyld, Nucl. Phys. {\bf B275} [FS17] 296, (1986).
1602:
1603: \bibitem{KogLom95}
1604: J. B. Kogut, M. P. Lombardo, and D. K. Sinclair, Phys. Rev D {\bf 51}, 1282
1605: (1995).
1606:
1607: \bibitem{BarMor97}
1608: I. M. Barbour, S. E. Morrison, E. G. Klepfish, J. B. Kogut, and
1609: M. P. Lombardo, Phys. Rev. D {\bf 56}, 7063 (1997).
1610:
1611: \bibitem{ShuVer93}
1612: E. V. Shuryak and J. J. M. Verbaarschot, Nucl. Phys.
1613: {\bf A560}, 306 (1993).
1614:
1615: \bibitem{Ver94}
1616: J. J. M. Verbaarschot, Phys. Rev. Lett. {\bf 72} 2531
1617: (1994); Phys. Lett. B {\bf 329}, 351 (1994).
1618:
1619: \bibitem{Ver99}
1620: For a review, see J. J. M. Verbaarschot, hep-ph/9902394.
1621:
1622: \bibitem{BanCas80}
1623: T. Banks and A. Casher, Nucl. Phys. {\bf B169}, 103 (1980).
1624:
1625: \bibitem{HalJac98}
1626: M.A. Halasz, A.D. Jackson, R.E. Shrock, M.A. Stephanov, J.J.M. Verbaarschot,
1627: Phys.Rev. D {\bf 58}, 096007 (1998).
1628:
1629: \bibitem{JacVer96}
1630: A. D. Jackson and J. J. M. Verbaarschot, Phys. Rev. {\bf D53}, 7223 (1996).
1631:
1632: \bibitem{DomGre84}
1633: I. D. Lawrie and S. Larbach, in {\it Phase Transitions and
1634: Critical Phenomena}, Edited by C. Domb and J. L.
1635: Lebowitz (Academic Press, London, 1984), vol. {\bf 9}, p. 1.
1636:
1637:
1638: \bibitem{DiaPet92}
1639: D. Diakonov and V. Petrov, in {\it Quark Cluster Dynamics},
1640: Lecture Notes in Physics, edited by K. Goeke, P. Kroll, and H. Petry
1641: (Springer-Verlag, Berlin, 1992) p. 288.
1642:
1643: \bibitem{SmiVer95}
1644: A. Smilga and J. J. M. Verbaarschot, Phys. Rev. D {\bf 51}, 829 (1995).
1645:
1646: \bibitem{KogSte99}
1647: J. B. Kogut, M. A. Stephanov, and D. Toublan,
1648: Phys. Lett. B {\bf 464}, 183 (1999).
1649:
1650: \bibitem{DiaFor96}
1651: D. Diakonov, H. Forkel, and M. Lutz, Phys. Lett. {\bf B373}, 147 (1996).
1652:
1653: \bibitem{DagKar86}
1654: E. Dagotta, F. Karsh, and A. Moreo, Phys. Lett. B {\bf 169}, 421 (1986).
1655:
1656: \bibitem{tHo73}
1657: G. 't Hooft, Nucl. Phys. {\bf B72}, 461 (1972).
1658:
1659:
1660:
1661:
1662:
1663: \end{thebibliography}
1664: \widetext
1665: \end{document}
1666:
1667:
1668:
1669:
1670:
1671:
1672:
1673:
1674:
1675: