1: %\documentclass[prd,showpacs,amsmath,amssymb,nofootinbib]{revtex4}
2: \documentclass[prd,twocolumn,showpacs,amsmath,amssymb,footinbib]{revtex4}
3: %\documentstyle[prl,multicol,aps,epsfig]{revtex}
4: \everymath{\displaystyle}
5: \usepackage[dvips]{graphicx}
6: \usepackage{longtable}
7: \newcommand \beq{\begin{eqnarray}}
8: \newcommand \eeq{\end{eqnarray}}
9: \newcommand{\nc}{\newcommand}
10: \def\DS {D\!\!\!\!/}
11: \def\vp{{\bf p}}
12: \def\qs{q\!\!\!/}
13: %\def\pfs{p\!\!\!/}
14: %\def\pfs{{p_F}\!\!\!\!\!\!\!/}
15: \def\ms{{\mu}\!\!\!/}
16: \def\pas{{\partial}\!\!\!/}
17: \def\simge{\mathrel{%
18: \rlap{\raise 0.511ex \hbox{$>$}}{\lower 0.511ex \hbox{$\sim$}}}}
19: \def\simle{\mathrel{
20: \rlap{\raise 0.511ex \hbox{$<$}}{\lower 0.511ex \hbox{$\sim$}}}}
21: \begin{document}
22:
23: \title{Thermal Phase Transitions and Gapless Quark Spectra\\
24: in Quark Matter at High Density}
25: \author{K. Iida,$^{1,2}$ T. Matsuura,$^3$
26: M. Tachibana,$^2$ and T. Hatsuda$^3$}
27: \affiliation{$^{1}$
28: RIKEN BNL Research Center, Brookhaven National
29: Laboratory, Upton, NY 11973\\
30: $^{2}$
31: The Institute of Physical and Chemical Research (RIKEN),
32: Wako, Saitama 351-0198, Japan\\
33: $^{3}$Department of Physics, University of Tokyo,
34: Tokyo 113-0033, Japan}
35:
36: \begin{abstract}
37: Thermal color superconducting phase transitions in three-flavor quark
38: matter at high baryon density are investigated in the Ginzburg-Landau (GL)
39: approach. We constructed the GL potential near the boundary with a normal
40: phase by taking into account nonzero quark masses, electric charge
41: neutrality, and color charge neutrality. We found that the density of states
42: averaged over paired quarks plays a crucial role in determining the phases
43: near the boundary. By performing a weak coupling calculation of the
44: parameters characterizing the GL potential terms of second order in the
45: pairing gap, we show that three successive second-order phase transitions
46: take place as the temperature increases: a modified color-flavor locked
47: phase ($ud$, $ds$, and $us$ pairings) $\to$ a ``dSC'' phase ($ud$ and
48: $ds$ pairings) $\to$ an isoscalar pairing phase ($ud$ pairing) $\to$ a normal
49: phase (no pairing). The Meissner masses of the gluons and the number of
50: gapless quark modes are also studied analytically in each of these phases.
51: \end{abstract}
52: \pacs{12.38.-t,12.38.Mh,26.60.+c}
53: \maketitle
54:
55: \section{Introduction}
56:
57: Over the past few years, the properties of color superconducting
58: quark matter and the phase structure at high baryon density, which were
59: originally studied in earlier publications \cite{BL}, have been
60: examined intensively (see, e.g., Ref.\ \cite{RWA} for reviews).
61: A color superconductor predicted to occur in weak coupling is a relativistic
62: system in which the long-range color magnetic interaction, which is
63: screened only dynamically by Landau-damping of exchanged gluons, is
64: responsible for the formation of the superconducting gap. The gap in weak
65: coupling has a non-BCS form,
66: $\Delta \propto \mu \exp( -3 \pi^2/\sqrt{2} g )$
67: with $g$ the strong coupling constant and $\mu$ the quark chemical potential
68: \cite{SON99}. Furthermore, the gap has a matrix structure with
69: $(N_c \times N_f)^2$ components due to various combinations of $N_c$ color
70: and $N_f$ flavor degrees of freedom.
71:
72: There is a good deal of evidence, obtained from various weak coupling
73: analyses \cite{Sch-NPB-575,Evans-NPB-581,IB,HH-PRD-68}, that
74: if all the quark masses are zero ($m_u=m_d=m_s=0$) and hence the quark
75: chemical potentials are equal ($\mu_u=\mu_d=\mu_s$), the quark matter is
76: in a color-flavor locked (CFL) phase \cite{CFL} at low temperature.
77: A transition from the CFL phase to the normal phase in mean-field theory is
78: of second order. In weak coupling, the transition temperature $T_c$ is
79: related to the zero temperature gap $\Delta_{T=0}$ as
80: $T_c= 2^{1/3} \times 0.57 \Delta_{T=0}$
81: \cite{SWR}, which is a BCS relation
82: except for a factor $2^{1/3}$.
83:
84: The massless situation thus described may be approximately realized at
85: asymptotically high densities where the quark masses are negligible compared
86: to the quark chemical potentials. However, the effect of nonzero quark
87: masses becomes important when the chemical potentials decrease. In the
88: presence of a quark mass difference characterized by $2m_s/(m_u+m_d) \sim 25$,
89: $\beta$ equilibrium, electric neutrality, and color neutrality combine to
90: produce non-trivial chemical potential differences between flavors and
91: and between colors, leading to new phases such as the gapless CFL (gCFL)
92: phase \cite{gCFL} in which two out of nine quark quasiparticles are gapless.
93:
94: In our recent Letter \cite{IMTH04}, we have studied what kind of phase
95: structure appears in $\beta$ equilibrated neutral quark matter near the
96: super-normal phase boundary when there are quark mass differences.
97: An advantage of studying the region near the phase boundary is that
98: in classifying the possible phase structures we can make use of the
99: model-independent Ginzburg-Landau (GL) analysis \cite{IB,IB3} in which
100: the thermodynamic potential difference between the superfluid and normal
101: phases is expanded in terms of the order parameter (the pairing gap).
102: Furthermore, by calculating the parameters controlling the thermodynamic
103: potential terms in weak coupling, one can {\rm prove} which phase is
104: realized below the super-normal boundary in the high density regime
105: where the weak coupling analysis is valid.
106:
107: A crucial observation found in Ref.\ \cite{IMTH04} is that the phase
108: structure near the critical temperature is essentially dictated by the
109: average density of states of different quarks. We also found that
110: the quark mass difference and the electric charge neutrality
111: (but not the color neutrality) play key roles in determining the phase
112: structure near the critical temperature. In weak coupling, it was shown that
113: the following successive phase transitions occur near the super-normal
114: phase boundary: a modified color-flavor locked (mCFL) phase ($ud$, $ds$,
115: and $us$ pairings) $\to$ a ``dSC'' phase ($ud$ and $ds$ pairings) $\to$ an
116: isoscalar two-flavor superconducting (2SC) phase ($ud$ pairing) $\to$ a
117: normal phase (no pairing).
118:
119: The purposes of this paper are (i) to give a detailed account of the
120: results given in Ref.\ \cite{IMTH04} and (ii) to investigate the elementary
121: excitation modes (gluons with the Meissner mass, massless gluons, gapped
122: quarks, and gapless quarks).
123:
124: The content of this paper is as follows. In Sec. II, we start with
125: a toy model that captures the essential features of the hierarchical
126: structure of the phase transitions. Then we construct a general form of the
127: GL potential with the quark masses in such a way that it fulfills symmetry
128: constraints. In Sec.\ III, we evaluate the parameters characterizing the GL
129: potential terms in the weak coupling region with unequal quark masses
130: ($m_{i}$, $i$=$u, d, s$) and unequal quark chemical potentials ($\mu_{i}$,
131: $i$=$u, d, s$) by utilizing the Cornwall-Jackiw-Tomboulis (CJT) effective
132: action \cite{CJT}. In Sec.\ IV, we consider a simplified situation
133: ($m_{u,d}=0$ and $m_s \neq 0$) and calculate the corrections to the GL
134: potential from nonzero $m_s$. In Sec.\ V, we introduce a parametrization of
135: the gap and analyze possible phases near the super-normal phase boundary at
136: finite temperature. For these phases, residual symmetries, the Meissner
137: masses of gluons, and the gapped and gapless quark modes are also
138: investigated. Section VI is devoted to summary and concluding remarks.
139:
140: \section{Ginzburg-Landau potential}
141:
142: In this section, as an introduction to later sections, we first study a
143: toy model that captures the essential features of such a hierarchical
144: structure of phase transitions as encountered in dense quark matter. Validity
145: of the various approximations adopted later will be also clarified by the
146: analysis of the toy model. We then construct a general form of the GL
147: potential with the quark masses such that it satisfies symmetry constraints.
148:
149: \subsection{A toy model}
150:
151: Let us assume two real order parameters, $X$ and $Y$, and consider a
152: GL potential of the following form:
153: \beq
154: \Omega &=& \alpha (X^2+Y^2) + \delta (X^2-Y^2) \nonumber \\
155: & & + \beta_1(X^2+Y^2)^2 + \beta_2(X^4+Y^4).
156: \label{eq:toy-GL}
157: \eeq
158: This is the most general form of the potential up to the quartic order with
159: the ``parity" symmetry, $X\leftrightarrow -X$ and $Y \leftrightarrow -Y$.
160: For the stability of the potential, $\beta_1+\beta_2 >0$ and
161: $2\beta_1+\beta_2>0$ are implicitly assumed.
162:
163: As we will see later, the coefficient $\alpha$ corresponds to
164: the reduced temperature in the chiral limit, while $\delta$ denotes the
165: degree of flavor symmetry breaking (in the present context, the breaking of
166: $X^2 \leftrightarrow Y^2$ symmetry) which is proportional to the quark mass
167: squared. The coefficients $\beta_1$ and $\beta_2$ are the quartic couplings
168: that do not change sign around the critical point, but more or less receive
169: corrections from the quark mass. Bearing in mind the results that will be
170: obtained in later sections, we parametrize these coefficients as
171: \beq
172: \label{eq:toy-alpha}
173: \alpha &=& \mu^2 t = \mu^2 \frac{T-T_c}{T_c}, \\
174: \label{eq:toy-delta}
175: \delta&=& \mu^2 \sigma = \mu^2 \frac{m^2}{g\mu^2} ,\\
176: \label{eq:toy-beta}
177: \beta_{1}=\beta_2 &\equiv& \beta= \frac{\mu^2}{T_c^2},
178: \eeq
179: where $\mu$ and $T_c$ are the quark chemical potential and the critical
180: temperature in the chiral limit. $m$ and $g$ are the quark mass and the QCD
181: coupling constant, respectively. In the last equalities of Eqs.\
182: (\ref{eq:toy-alpha})--(\ref{eq:toy-beta}), we mimic the actual
183: parametric forms of $t$, $\sigma$, and $\beta$ shown in Eqs.\ (\ref{b1b2}) and
184: (\ref{eq:full-sigma}) except for numerical factors. We do not have to
185: consider the quark mass corrections to $\beta_1$ and $\beta_2$ under the
186: condition to be shown shortly. $\delta$ is a key parameter which governs the
187: multiple structure of the phase transitions: Due to the presence of $\delta$,
188: the temperatures where the coefficients affixed to $X^2$ and $Y^2$ change sign
189: are no longer identical.
190:
191: The minimum of the potential can be simply obtained from
192: $\partial \Omega/\partial X=\partial \Omega/\partial Y=0$. Then one finds
193: three different phases:
194: \begin{enumerate}
195: \item a phase where both $X$ and $Y$ have non-vanishing
196: condensates,
197: \beq
198: \label{eq:XY-1}
199: X^2&=& -\frac{1}{6\beta}(\alpha+3\delta)=
200: - \frac{T_c^2}{6} (t+ 3 \sigma), \\
201: \label{eq:XY-2}
202: Y^2&=& -\frac{1}{6\beta}(\alpha-3\delta)=
203: - \frac{T_c^2}{6} (t- 3 \sigma).
204: \eeq
205: \item a phase where only $Y$ has non-vanishing
206: condensate,
207: \beq
208: X^2&=& 0, \\
209: Y^2&=& -\frac{1}{4\beta}(\alpha-\delta)=
210: - \frac{T_c^2}{4} (t- \sigma).
211: \eeq
212: \item a phase where there are no condensates,
213: \beq
214: X^2=Y^2=0.
215: \eeq
216: \end{enumerate}
217:
218: \begin{figure}[t]
219: \begin{center}
220: \includegraphics[scale=0.45]{toy-model.eps}
221: \end{center}
222: \vspace{-0.5cm}
223: \caption{ Condensates $X^2$ and $Y^2$ as a function of the
224: reduced temperature $t$ in a toy model. $\sigma$ is
225: the parameter that breaks ``flavor" symmetry $X^2 \rightarrow Y^2$.
226: }
227: \label{fig:toy-model}
228: \end{figure}
229:
230: The behavior of the condensates as a function of the reduced temperature
231: $t$ is illustrated in Fig.\ \ref{fig:toy-model}. The multiple structure of
232: the phase transitions induced by $\sigma$ shown in this figure is a general
233: feature which survives in a realistic situation that will be discussed in later
234: sections.
235:
236: The questions to be addressed here are (i) what is the region of
237: temperature where the GL potential Eq.\ (\ref{eq:toy-GL}) expanded up to the
238: quartic order is valid, and (ii) what is the justification of neglecting the
239: quark mass corrections to $\beta_1$ and $\beta_2$. These questions can be
240: answered if we restrict ourselves to the region
241: \beq
242: |t| < {\rm const} \cdot \sigma.
243: \label{eq:t-sig}
244: \eeq
245: Within this interval, both $X^2$ and $Y^2$ are of order or even smaller than
246: $\sigma T_c^2$. Therefore, the dimensionless expansion parameters
247: of the GL potential satisfy
248: \beq
249: \frac{(X^2,Y^2)}{T_c^2} < {\cal O}(\sigma),
250: \label{eq:t-sig2}
251: \eeq
252: which in turn guarantees that the higher order terms beyond the quartic order
253: are negligible. Furthermore, the additional quark mass corrections to
254: $\beta_1$ and $\beta_2$ give only higher order corrections to $X^2$ and $Y^2$
255: as can be seen from, e.g., Eqs.\ (\ref{eq:XY-1}) and (\ref{eq:XY-2});
256: $X^2$ and $Y^2$ are already proportional to $\sigma$ if Eq.\ (\ref{eq:t-sig})
257: is satisfied.
258:
259:
260: \subsection{General form of GL potential}
261:
262: In this subsection, we construct a general form of the GL potential on
263: the basis of the QCD symmetry under $G=SU(3)_C\times SU(3)_L\times SU(3)_R
264: \times U(1)_B$.
265:
266: For this purpose, let us first consider a basic variable, the gap
267: function $\phi_{bcjk}$. It is defined through the pairing gap $\Delta$ for a
268: quark (color $b$ and flavor $j$) and another quark (color $c$ and flavor $k$)
269: on the Fermi surface as will be shown in Sec.\ III [in particular, see
270: Eqs.\ (\ref{eq:Del-phi}) and (\ref{G})]. Focusing our attention on
271: spin-zero pairings of positive energy quarks in anti-symmetric combinations
272: of colors and of flavors, we may rewrite $\phi_{bcjk}$ as \cite{IB}
273: \begin{eqnarray}\label{phi-d}
274: [\phi_{bcjk}]_{L,R}=\epsilon_{abc} \epsilon_{ijk} [d_a^i]_{L,R} .
275: \end{eqnarray}
276: Here the 3$\times $ 3 matrix $[d_a^i]_{L}$ belongs to the
277: {\rm fundamental} representation of $SU(3)_C \times SU(3)_{L}$. Similar
278: properties hold for $[d_a^i]_{R}$ under the change $R \leftrightarrow L$.
279: In this subsection, we omit the color and flavor indices and write the gap
280: function as ${\rm d}_{L,R}$ for simplicity.
281:
282: Under the operation of $G_{L,R}=SU(3)_C \times SU(3)_{L,R}\times U(1)_B$,
283: the left- and right-handed quarks and the gap ${\rm d}_{L,R}$ transform as
284: \begin{eqnarray}
285: \psi_{L,R}&\rightarrow& e^{i\varphi}U_C U_{L,R}\psi_{L,R}, \\
286: {\rm d}_{L,R}&\rightarrow& e^{-2i\varphi}U_{L,R}{\rm d}_{L,R}U_C^{\rm T},
287: \label{transform1}
288: \end{eqnarray}
289: where $U_{L,R}$'s are $3\times 3$ unitary matrices corresponding to each
290: $SU(3)_{L,R}$ symmetry. The phase factor is associated with the
291: $U(1)_B$ rotation.
292:
293: We now try to incorporate small but nonzero quark masses and write down
294: possible terms of the GL potential allowed by the QCD symmetry $G$. In QCD
295: with quark masses, we have such a term as
296: \begin{equation}
297: {\cal L}=\bar{\psi}_L m \psi_R+h.c.,
298: \label{quarkmass}
299: \end{equation}
300: with $m$ being a $3\times 3$ mass matrix in flavor space. Just like the
301: standard procedure to construct the chiral Lagrangians, $m$ is assumed to
302: transform in the following way \cite{Georgi},
303: \begin{equation}
304: m\rightarrow U_L m U_R^{\dagger}.
305: \label{masstransform}
306: \end{equation}
307:
308: Following the transformation laws, Eqs.\ (\ref{transform1}) and
309: (\ref{masstransform}), we can write down a possible form of the GL potential
310: as an expansion in terms of the order parameter, which is, in the
311: present context, the on-shell gap function ${\rm d}_{L,R}$. Following the
312: discussion in Sec.\ II.A, we take into account the quark mass terms up to
313: ${\cal O}(m^2)$ only in the quadratic terms in ${\rm d}_{L,R}$ and obtain
314: \begin{eqnarray}\label{generalGL}
315: \Omega &=& (a_0)_L {\rm Tr} ({\rm d}_L^{\dagger}{\rm d}_L)
316: +(a_0)_R {\rm Tr} ({\rm d}_R^{\dagger}{\rm d}_R) \\ \nonumber
317: & &+a_1[{\rm Tr} ({\rm d}_L^{\dagger}m {\rm d}_R)+ h.c.] \\ \nonumber
318: & &+(a_2)_L {\rm Tr} ({\rm d}_L^{\dagger}mm^{\dagger}{\rm d}_L)
319: +(a_2)_R {\rm Tr} ({\rm d}_R^{\dagger}m^{\dagger}m {\rm d}_R) \\ \nonumber
320: & &+(\hat{a}_2)_L{\rm Tr} ({\rm d}_L^{\dagger} {\rm d}_L)
321: {\rm Tr} (m^{\dagger}m ) +(\hat{a}_2)_R
322: {\rm Tr} ({\rm d}_R^{\dagger} {\rm d}_R) {\rm Tr} (m^{\dagger}m ) \\
323: \nonumber
324: & & +c[{\rm det} m\cdot {\rm Tr}({\rm d}_R^{\dagger}m^{-1}{\rm d}_L)+ h.c.] \\
325: \nonumber
326: & &+(b_{1})_L [ {\rm Tr} ({\rm d}_L^{\dagger}{\rm d}_L)]^2
327: +(b_{1})_R [{\rm Tr} ({\rm d}_R^{\dagger}{\rm d}_R)]^2 \\ \nonumber
328: & &+ (b_{2})_L {\rm Tr} [({\rm d}_L^{\dagger}{\rm d}_L)^2]
329: +(b_{2})_R {\rm Tr} [({\rm d}_R^{\dagger}{\rm d}_R)^2] \\ \nonumber
330: & &+b_3 {\rm Tr} [({\rm d}_L^{\dagger}{\rm d}_L)
331: ({\rm d}_R^{\dagger}{\rm d}_R)]
332: +b_4 {\rm Tr} ({\rm d}_L^{\dagger}{\rm d}_L)
333: {\rm Tr} ({\rm d}_R^{\dagger}{\rm d}_R),
334: \end{eqnarray}
335: where the $a$'s, $b$'s, and $c$ are the expansion coefficients being
336: functions of temperature and chemical potential.
337:
338: Only the term proportional to $a_1$ in Eq.\ (\ref{generalGL}) breaks
339: $U(1)_A$ symmetry and thus originates from anomaly induced interactions.
340: It does not affect the phase structure near the super-normal phase boundary
341: at asymptotically high density as will be shown in Sec.\ IV.D. The term
342: proportional to $c$ also does not play an important role near the phase
343: boundary as will be discussed at the end of Sec.\ III.B.
344:
345: We can further take into account the effects of electric and color
346: neutralities and discuss possible terms to be included in
347: Eq.\ (\ref{generalGL}). In particular, the electric charge neutrality plays
348: a central role in this paper. Instead of writing down its general structure,
349: we will show its explicit form calculated in the weak coupling regime in
350: Sec.\ III. The effect enters into the quadratic terms of the GL potential.
351: As for the color neutrality, its effect is subdominant in the vicinity of the
352: phase boundary as has been already shown in Ref.\ \cite{IB}; the color
353: neutrality affects only the quartic part of the GL potential (see Sec.\ IV.C).
354:
355:
356: \section{Corrections to the GL potential in weak coupling}
357:
358: In this section, we calculate corrections to the GL potential by
359: unequal quark masses ($m_{i}$, $i$=$u, d, s$) and unequal quark chemical
360: potentials ($\mu_{i}$, $i$=$u, d, s$). To simplify the derivation, we make
361: several ansatze for the pairing gap:
362: \begin{enumerate}
363: \item
364: The pairing between quarks is assumed to be in the zero total angular
365: momentum ($J=0$) channel \cite{PR-PRD60}. Furthermore, it is assumed to be
366: in the $LL$ and $RR$ channel as in the massless case and in the
367: positive parity channel as in the presence of $U(1)_A$ breaking \cite{RWA}.
368: Then the gap can be written as
369: $\Delta(k)=\gamma^5 \Delta^{(1)}+ {\bf \gamma} \cdot {\bf {\hat k}}
370: \gamma^0 \gamma^5 \Delta^{(2)}$ \cite{BL}.
371:
372: \item
373: The pairing is projected onto positive energy states. In this case,
374: it is convenient to rewrite $\Delta(k)$ as
375: \begin{eqnarray}
376: \Delta(k)=\gamma^5 \phi(k_0, {\bf k}) \Lambda^+ ({\bf k}),
377: \label{eq:Del-phi}
378: \end{eqnarray}
379: where
380: $\Lambda^+ ({\bf k})=(1+\gamma^0 {\bf \gamma} \cdot {\bf {\hat k}}) /2$
381: is a projection operator onto the positive energy states of
382: massless quarks.
383:
384: \item
385: The pairing is assumed to take place in the color antisymmetric and
386: flavor antisymmetric channel. This color channel is indeed the most
387: attractive in weak coupling \cite{SON99,brown,PR}, while the flavor channel
388: is chosen in such a way as to satisfy the Pauli principle.
389: Then, the pairing gap of a quark (color $b$ and flavor $j$) and another quark
390: (color $c$ and flavor $k$) at the Fermi surface takes the form as given in
391: Eq.\ (\ref{phi-d}). Furthermore, the pairing in the positive parity channel
392: is represented by
393: \begin{eqnarray}\label{parity-even}
394: {\rm d}_L={\rm d}_R\equiv d.
395: \end{eqnarray}
396:
397: \item
398: We adopt the fact that the three-momentum dependence of the on-shell gap
399: $\Delta(q_0=\epsilon({\bf q}), {\bf q})$ or, equivalently,
400: $\phi({\bf q};T)\equiv
401: \phi(q_0=\epsilon({\bf q}), {\bf q})$
402: can be approximately factorized as \cite{PR}
403: \begin{equation}
404: \phi({\bf q};T) \simeq \phi(|{\bf q}|=\mu ; T) f({\bf q}),
405: \label{approx1}
406: \end{equation}
407: where
408: \begin{eqnarray}\label{f}
409: f({\bf q})&=& \sin {{\bar g}y}|_{T=0}, \\
410: {\bar g}&=& g/ 3 \sqrt{2} \pi, \\
411: b &\equiv& 256 \pi^4 (2/3g^2)^{5/2}, \label{b} \\
412: y({\bf q})&=&\ln \left[ \frac{2b \mu}{ ||{\bf q}|-\mu| + E({\bf q})} \right],
413: \\
414: E({\bf q})&=&\left[ ||{\bf q}|-\mu|^2+
415: |\phi({\bf q};T)|^2 \right]^{1/2}.
416: \end{eqnarray}
417:
418: \item
419: We will assume later in Sec.\ V that $d_a^i$ is diagonal in color-flavor
420: space even under the influence of nonzero quark masses and charge chemical
421: potentials. Then we will minimize the GL potential within the subspace.
422:
423: \end{enumerate}
424:
425: In the following, we start with a known result for massless quarks
426: (Sec.\ III.A) and then proceed to describe a more general situation with
427: unequal quark masses ($m_{i}$, $i$=$u, d, s$) and unequal quark chemical
428: potentials ($\mu_{i}$, $i$=$u, d, s$) in Sec.\ III.B.
429:
430: \subsection{Case with massless three flavors}
431:
432: For a homogeneous system composed of massless quarks $(m_{u,d,s}=0)$,
433: quarks have a common chemical potential $\mu$, and the GL potential near the
434: critical temperature $T_c$ expanded up to quartic order reads
435: \cite{IB,PIS00}
436: \begin{eqnarray}
437: \label{GL}
438: \Omega_0= \bar{\alpha} \sum_{a}|{\mathbf{d}_a}|^2
439: +\beta_1(\sum_{a}|{\mathbf{d}_a}|^2)^2
440: +\beta_2\sum_{ab}|{\mathbf{d}_a}^{\ast}
441: \cdot {\mathbf{d}}_b|^2, \nonumber \\
442: \end{eqnarray}
443: where $({\mathbf{d}_a})^i\equiv (d_a^u, d_a^d, d_a^s)$, and
444: the inner product is taken for flavor indices. Using Eq.\
445: (\ref{parity-even}), one can relate the coefficients in Eq.\ (\ref{generalGL})
446: to those in Eq.\ (\ref{GL}) as
447: \begin{eqnarray}
448: &&\beta_1=(b_{1})_{L}+(b_{1})_{R}+b_4, \nonumber \\
449: &&\beta_2=(b_{2})_{L}+(b_{2})_{R}+b_3, \nonumber \\
450: && \bar{\alpha}=(a_0)_L+(a_0)_R .
451: \end{eqnarray}
452: This potential is manifestly invariant under $SU(3)_C \times SU(3)_{L+R}
453: \times U(1)_{B}$ rotation. In the weak coupling approximation where the
454: one-gluon exchange force in the normal medium is responsible for the pairing,
455: the coefficients have been calculated as \cite{IB}
456: \begin{eqnarray}
457: \label{b1b2}
458: \beta_{1}=\beta_{2}
459: =\frac{7\zeta(3)}{8(\pi T_c)^{2}}N(\mu)
460: \equiv \beta, \
461: \bar{\alpha}=4 N(\mu) t \equiv \alpha_0 t.
462: \end{eqnarray}
463: Here $N(\mu) = \mu^2/2\pi^2$ is the density of states at the Fermi surface,
464: and $t=(T-T_c)/T_c$ is the reduced temperature. With the parameters
465: in Eq.\ (\ref{b1b2}), one finds a single second-order phase transition at
466: $T=T_c$ from the CFL phase ($ {d}_a^i \propto \delta_a^i$) to the
467: normal phase ($ {d}_a^i = 0$) in the mean-field theory \cite{IB}.
468:
469: \subsection{Case with unequal quark masses and chemical potentials}
470:
471: For a homogeneous system composed of unequal quark masses, there arise
472: differences in the chemical potential among flavors and among colors
473: due to charge and color neutrality conditions. The GL potential in this case
474: is obtained in a similar approach to that adopted in Ref.\ \cite{IB} where a
475: correction to Eq.\ (\ref{GL}) from color neutrality is calculated.
476:
477: First we start with the Nambu-Gor'kov two component field
478: ($\psi_{ai}, \bar{\psi}_{ai}^{C}$), where $\psi^C=C \bar{\psi}^{\rm T}$ is
479: the charge-conjugate spinor. The quark propagator of this field with
480: the Hartree-Fock contributions ignored in the diagonal part reads
481: \begin{eqnarray}
482: G(k) &\equiv&
483: \left(
484: \begin{array}{cc}
485: G^{(11)}(k)& G^{(12)}(k) \\
486: G^{(21)}(k)& G^{(22)}(k)
487: \end{array}
488: \right) \\
489: & =& \left(
490: \begin{array}{cc}
491: \gamma k+\gamma^{0}{\cal M} -m & {\tilde \Delta}(k) \\
492: \Delta(k) & \gamma k-\gamma^{0}{\cal M}^{\rm T}-m
493: \end{array}
494: \right)^{-1} .
495: \label{G}
496: \end{eqnarray}
497: Here ${\tilde \Delta}=\gamma^{0}\Delta^{\dagger}\gamma^{0}$, and
498: \begin{eqnarray}
499: {\cal M}_{abij}=\delta_{ab}\delta_{ij} \mu_{i} ,\ \ \
500: m_{abij}=\delta_{ab}\delta_{ij} m_{i},
501: \end{eqnarray}
502: are the quark chemical potentials and the quark masses in color ($a, b$) and
503: flavor ($i, j$) space, respectively. As in Sec.\ III.A, we again define
504: $\mu$ as a quark chemical potential in the chiral limit ($m_{u,d,s}=0$).
505: Once unequal quark masses are included under charge neutrality and $\beta$
506: equilibrium, $\mu_{i}$ differs from $\mu$. We define the deviation of
507: ${\cal M}_{abij}$ from the massless case as
508: $\delta {\cal M}_{abij}\equiv \delta_{ab}\delta_{ij} (\mu_{i}-\mu) $.
509:
510: The free quark propagator $G_0$ and the self-energy $\Sigma$ are defined
511: from the diagonal and off-diagonal components of $G^{-1}$ as usual:
512: \begin{eqnarray}
513: G^{-1}(k)=G_{0}^{-1}(k) - \Sigma(k)\ .
514: \label{GSig}
515: \end{eqnarray}
516:
517: Then the GL potential, which is a difference between the superfluid and
518: normal phases near the critical temperature, is written in the CJT form
519: \cite{CJT}
520: \begin{eqnarray}
521: &&\Omega = \Omega_{1} + \Omega_{2} \label{omegagl} \\
522: &&\Omega_{1} = \frac{T}{2}\sum_{n}
523: \int\frac{d^{3}q}{(2\pi)^{3}}
524: {\rm Tr}[-G(q)\Sigma(q)+\ln G_{0}^{-1}(q)G(q)] , \nonumber \\
525: \label{omega1} \\
526: &&\Omega_{2} = g^2 \frac{T^2}{4}
527: \sum_{m, n}
528: \int\frac{d^{3}k}{(2\pi)^{3}} \int\frac{d^{3}q}{(2\pi)^{3}} \nonumber \\
529: &&~~ \! \! \times {\rm Tr}\Bigl[
530: {\cal D}_{\mu\nu}^{\alpha\beta} (q-k)
531: \gamma^{\mu} \frac{\lambda^{\alpha}}{2} G^{(12)}(k)
532: \gamma^{\nu} {\left( \frac{\lambda^{\beta}}{2} \right)^{\rm T}} G^{(21)}(q)
533: \nonumber \\
534: &&\! \! +{\cal D}_{\mu\nu}^{\alpha\beta} (q-k)
535: \gamma^{\mu} {\left(\frac{\lambda^{\alpha}}{2}\right)^{\rm T}}G^{(21)}(k)
536: \gamma^{\nu} \frac{\lambda^{\beta}}{2}G^{(12)}(q)
537: \Bigl] .
538: \label{omega2}
539: \end{eqnarray}
540: Here $\Omega_2$ represents the two particle irreducible graphs in the
541: mean-field approximation with ${\cal D}(q)$ being the gluon propagator
542: in hard dense loop approximation, and $\lambda^{\alpha}$ the generators of
543: color $SU(3)$. The summations are taken over Matsubara frequencies of quarks.
544:
545: In the following, we will consider the corrections from
546: $ \delta{\cal M}_{abij}$ and $m_{abij}$ to the quadratic term in $\Delta$
547: assuming that the corrections are small. The corrections to the quartic terms
548: are negligible near the critical temperature. Figure 2 represents the
549: ${\cal O}(\Delta ^2)$ contributions to $\Omega_{1}$ and $\Omega_{2}$.
550:
551: First we consider $\Omega_{1}^{\Delta ^2}$,
552: \begin{eqnarray}
553: \Omega_{1}^{\Delta ^2}
554: &=&-\frac{T}{2}\sum_{n~ {\rm odd}}
555: \int\frac{d^{3}q}{(2\pi)^{3}}
556: {\rm Tr} \Biggl[ \frac{1}{\gamma q+\gamma^{0}{\cal M}-m}
557: {\tilde\Delta}(q) \nonumber\\
558: &&~~~~~~~~~\times \frac{1}{\gamma q-\gamma^{0}{\cal M}^{\rm T}-m}
559: \Delta(q) \Biggl] \nonumber \\
560: &=& -\sum_{abij} \frac{i}{4} \oint \frac{d^{4}q}{(2\pi)^{4}}
561: \tanh \left( \frac{q_0}{2T} \right) {\rm Tr} \Biggl[
562: \frac{1}{\gamma q+\gamma^{0}{\mu}_{i}-m_{i}} \nonumber \\ &&~~~~~~~~
563: \times {\tilde\Delta}(q)_{abij}
564: \frac{1}{\gamma q-\gamma^{0}{\mu}_{j}-m_{j}}
565: \Delta(q)_{baji} \Biggl].
566: \label{eq:O-1-D}
567: \end{eqnarray}
568: The first quark propagator in the right-hand side of Eq.\ (\ref{eq:O-1-D})
569: has particle and anti-particle poles,
570: \begin{eqnarray} \label{pp}
571: \epsilon^{p \pm}&=& \pm |{\bf q}| - p_F^{i} ,
572: \end{eqnarray}
573: while the second propagator has hole and anti-hole poles,
574: \begin{eqnarray} \label{hp}
575: \epsilon^{h \mp}&=& \mp |{\bf q}| + p_F^{j} ,
576: \end{eqnarray}
577: where $p_F^{i}=\sqrt{\mu_i^2-m_i^2}$ is the Fermi momentum of flavor $i$.
578: We define the shift of the chemical potential of flavor $i$ from $\mu$ as
579: \begin{eqnarray}
580: \delta \mu_i = \mu_i -\mu.
581: \end{eqnarray}
582: Then, up to ${\cal O}(m^2/\mu^2, \delta \mu/\mu)$, $p_F^{i}$ can be written as
583: \begin{eqnarray}\label{fermi-mom}
584: p_F^{i} \simeq \mu_{i}-m_{i}^2/2\mu.
585: \end{eqnarray}
586:
587: Of the poles in Eqs.\ (\ref{pp}) and (\ref{hp}), only two of them,
588: $q_0= \epsilon^{p +}$ (particles) and $q_0= \epsilon^{h -}$ (particle holes),
589: have relevant contributions to the potential. The other two poles contribute
590: to the gaps on antiparticle (antiparticle hole) mass shell, which are
591: irrelevant in our calculation. Let us define
592: \begin{eqnarray}
593: {\tilde q}&=&|{\bf q}|-\mu, \\
594: {\cal F} ({\tilde q}) &=&\frac{\tanh ({\tilde q}/2T)}{{\tilde q}} .
595: \end{eqnarray}
596: Then, combining the residues of the two poles, we obtain
597: \begin{eqnarray} \label{tree0}
598: \Omega_{1}^{\Delta ^2}=
599: \sum_{abij} \frac{1}{2}
600: \int \frac{d^3 q}{(2 \pi )^3}
601: \left(
602: {\cal F} ({\tilde q})
603: -\frac{\partial {\cal F} ({\tilde q})
604: }{\partial{\tilde q}} \delta p_{ij} \right)
605: |\phi_{abij}|^2 .
606: \end{eqnarray}
607: Here,
608: \begin{eqnarray}
609: \delta p_{ij} = (p_F^i + p_F^j ) /2 -\mu.
610: \end{eqnarray}
611: For later purpose, we also introduce an averaged shift of the chemical
612: potential as
613: \begin{eqnarray}
614: \delta \mu_{ij} = (\delta \mu_i + \delta \mu_j )/2.
615: \end{eqnarray}
616:
617:
618: In the diagrammatic language, the first term in the right-hand side of
619: Eq.\ (\ref{tree0}) corresponds to (a0) in Fig.\ \ref{fig:graphs}. The second
620: term corresponds to (a1) and (a2). (a3) vanishes because of the Lorentz
621: structure of the gaps together with the identity,
622: ${\rm Tr} (\Lambda^+ \Lambda^-) =0$.
623:
624:
625: To perform the momentum integration in Eq.\ (\ref{tree0}), it is
626: convenient to introduce a positive constant $\kappa \sim {\cal O}(g^{-1})$
627: which cancels at the end of the calculation \cite{PR,IB}. The cutoff of the
628: three-momentum $\Lambda$ is taken to be $b\mu$ for the sake of consistency
629: with the formula, Eq.\ (\ref{f}). We divide the region of the momentum
630: integral $-\mu < |{\bf q}|-\mu < b \mu $ into two parts,
631: $0< ||{\bf q}|- \mu| < \kappa \phi_0$ (region I), and
632: $-\mu < |{\bf q}|- \mu < -\kappa\phi_0$
633: and $\kappa\phi_0< |{\bf q}|- \mu< b\mu$ (region II) \cite{PR}. Here we
634: defined $\phi_0 \equiv [(1/9)\sum_{abij}|\phi_{abij}(\mu, T=0)|^2]^{1/2}$.
635: In each region, the following approximations can be safely taken: in region I,
636: $\sin {{\bar g}y}|_{T=0} \sim 1$, and in region II, $\tanh
637: \left[ (|{\bf q}| -\mu)/2T \right] \sim (|{\bf q}| -\mu) /||{\bf q}| -\mu|$.
638:
639: The first term in the right-hand side of Eq.\ (\ref{tree0}) is
640: calculated as
641: \begin{eqnarray}
642: {\rm (a0)}&=& \sum_{abij} \frac{1}{2}
643: \int \frac{d^3 q}{(2 \pi )^3}
644: \ {\cal F} ({\tilde q}) \
645: |\phi_{abij}|^2 \nonumber \\
646: &=&
647: \sum_{abij}
648: \frac{1}{2}N(\mu) \Biggl\{
649: 2 \int_{0}^{\kappa \phi_0}
650: \ d \tilde{q} \ {\cal F} ({\tilde q})
651: \nonumber\\
652: &-&
653: \Biggl( \int_{\ln(b \mu/ \kappa \phi_0)}^{\ln b}
654: +\int_{\ln(b \mu/ \kappa \phi_0)}^{0} \Biggl)
655: dy f^2(y)
656: \Biggl\}
657: |\phi_{abij}|^2 \nonumber \\
658: &=&
659: \sum_{abij}
660: N(\mu)
661: \left\{ \ln \left(\frac{T_c}{T} \right)+ \frac{\pi}{4 {\bar g}} \right\}
662: |\phi_{abij}|^2 . \label{omega--1}
663: \end{eqnarray}
664: In the first equality, we replaced the momentum in the measure by the density
665: of states at the Fermi surface. In the second equality, we used the weak
666: coupling approximation $g \ll 1$ and expanded $f(y)$ in terms of $g$.
667: We also used the explicit $g$ dependence of $T_c$ and $\phi_0$ in weak
668: coupling \cite{brown,PR},
669: $\ln(T_c/\mu) \sim \ln(\phi_0/\mu) \sim -3\pi^2/\sqrt2 g$,
670: to obtain the final form.
671:
672: The second term in the right-hand side of Eq.\ (\ref{tree0}) is more
673: involved:
674: \begin{eqnarray}
675: {\rm (a1)}&+& {\rm (a2)} \nonumber\\
676: &=& -\frac{1}{2}
677: \int \frac{d^3 q}{(2 \pi )^3}
678: \frac{\partial {\cal F} ( {\tilde q} )}{\partial {\tilde q} } \ \delta p \
679: |\phi|^2 \nonumber\\
680: &=&\frac{1}{4 \pi^2}
681: \int _{-\mu}^{b\mu} \ d{\tilde q} \ ({\tilde q}+\mu)^2
682: \{ {\cal F}({\tilde q})-{\cal F}({\tilde q}+ \delta p ) \}
683: f^2({\tilde q}) \ |\phi|^2 \ \nonumber \\
684: &\sim&
685: \frac{1}{4 \pi^2}
686: \Biggl[
687: \left\{ \int_{-\mu}^{b\mu}
688: -\int_{-\mu+ \delta p }^{b\mu+ \delta p } \right\}
689: \ d{\tilde q} \ {\cal F}({\tilde q}) f^2({\tilde q}) ({\tilde q}+\mu)^2
690: \nonumber\\
691: &&~~~~~~~~~+2 \ \delta p \
692: \int _{-\mu}^{b\mu} \ d{\tilde q} \ {\cal F}({\tilde q}) f^2({\tilde q})
693: ({\tilde q}+\mu)
694: \Biggl] \ |\phi|^2 \ \nonumber\\
695: &\sim&
696: \frac{1}{4 \pi^2}
697: \Biggl[
698: - {\cal F}(b\mu) f^2(b\mu ) (b\mu+\mu)^2 \nonumber\\
699: &&~~+2
700: \int _{-\mu}^{b\mu} \ d{\tilde q} \ {\cal F}({\tilde q}) f^2({\tilde q})
701: ({\tilde q}+\mu)
702: \Biggl] \ \delta p \ |\phi|^2, \
703: \label{inte} \
704: \end{eqnarray}
705: where we abbreviated internal indices for simplicity. $f(b\mu)$ appearing in
706: the first term of the last bracket in Eq.\ (\ref{inte}) is
707: ${\cal O}(g(\phi^2/(b \mu)^2))$ and can be neglected at high densities.
708: Then the second term leads to
709: \begin{eqnarray}
710: {\rm (a1)}&+&{\rm (a2)} \nonumber\\
711: &=&
712: \frac{1}{2 \pi^2} \mu\ \delta p \
713: \Biggl\{
714: 2 \int_{0}^{\kappa \phi_0}
715: \ d {\tilde q} \ {\cal F}({\tilde q}) \nonumber\\
716: &-&
717: \Biggl( \int_{\ln(b \mu/ \kappa \phi_0)}^{\ln b}
718: +\int_{\ln(b \mu/ \kappa \phi_0)}^{0}\Biggl)
719: dy f^2(y)
720: \Biggl\} |\phi|^2 \nonumber\\
721: &=&
722: -N(\mu)
723: \frac{\delta p}{\mu} \ln \left( \frac{T_c}{\mu} \right)
724: \ |\phi|^2. \label{omega--2}
725: \end{eqnarray}
726: Here we again expanded $f(y)$ in terms of $g$ and used the explicit $g$
727: dependence of $T_c$ in weak coupling to obtain the final form.
728:
729: Then the sum of Eqs.\ (\ref{omega--1}) and (\ref{omega--2}) becomes
730: \begin{eqnarray} \label{tree}
731: \Omega_{1}^{\Delta^2}&=&
732: {\rm (a0)}+{\rm (a1)}+{\rm (a2)}
733: \nonumber\\
734: &=&\sum_{abij}
735: N(\mu)
736: \Biggl[
737: \left\{ \ln \left(\frac{T_c}{T} \right)+ \frac{\pi}{4 {\bar g}} \right\}
738: \nonumber\\
739: &-&\frac{\delta p_{ij}}{\mu} \ln \left( \frac{T_c}{\mu} \right)\
740: \Biggl] |\phi_{abij}|^2 .
741: \end{eqnarray}
742:
743: \begin{figure}[b]
744: \begin{center}
745: \includegraphics[width=8cm,height=4.5cm]{graphs.eps}
746: \end{center}
747: \vspace{-0.5cm}
748: \caption{The quadratic terms of the GL potential. $\times$ and $+$
749: represent $m$ and $\delta \mu$, respectively. (a0)--(a3) represent
750: $\Omega_1^{\Delta^2}$, and (b0)--(b4) are the two particle irreducible graphs
751: $\Omega_2^{\Delta^2}$. (a0) and (b0) are the quadratic terms of the potential
752: in the $m=\delta \mu=0$ case. Among these graphs, (a3) and (b2) vanish
753: due to the Lorentz structure of the gap. (b3) vanishes because we consider
754: the gaps constructed by quarks having the same chirality. Only (a1), (a2),
755: (b1), and (b4) give finite correction to the original GL potential
756: [(a0) plus (b0)].}
757: \label{fig:graphs}
758: \end{figure}
759:
760: Let us next evaluate $\Omega_2^{\Delta ^2}$. In the diagrammatic
761: description, this corresponds to (b0)--(b4) in Fig.\ \ref{fig:graphs}
762: in which the quark propagator is expanded in powers of masses and
763: chemical potential differences. Of these graphs, (b2) vanishes because of the
764: Lorentz structure of the gaps, and (b3) vanishes because the gaps are
765: constructed by quarks having the same chirality. As a result, only (b0),
766: (b1), and (b4) have a finite value. We expand these diagrams by
767: $t \equiv |T-T_c|/T_c$ around $T_c$ up to ${\cal O}(t)$ for (b0) and
768: ${\cal O}(1)$ for (b1) and (b4).
769:
770: We use a relation between $T_c$ and $g$ when $m_i$ and $\delta \mu_i$
771: in the denominator of the quark propagator are taken to be zero \cite{IB},
772: \begin{eqnarray}\label{Tcg}
773: &&f({\bf q})=
774: -g^2 \frac{1}{2}\oint \frac{d^4 k}{ (2 \pi)^4} \tanh
775: \left( \frac{k_0}{2T_c}\right) {\rm Tr}\bigg[
776: {\cal D}_{\mu\nu}^{\alpha\beta} (q-k)
777: \gamma^{\mu} \nonumber \\
778: &&\times
779: {\left( \frac{\lambda^{\alpha}}{2} \right)^{\rm T}}
780: (\gamma k-\gamma^{0}{\cal M})^{-1}
781: (\gamma k+\gamma^{0}{\cal M})^{-1} \gamma^{\nu}
782: {\left( \frac{\lambda^{\beta}}{2} \right)} f({\bf k}) \bigg].\nonumber\\
783: \end{eqnarray}
784: This equation is derived by taking $T \rightarrow T_c$ in the gap equation
785: with the decomposition Eq.\ (\ref{approx1}).
786:
787: Using Eq.\ (\ref{Tcg}) which relates $T_c$ and $g$,
788: we can evaluate (b0), (b1), and (b4) in Fig.\ \ref{fig:graphs}. We derive (b0)
789: as follows. In leading order in $t$, (b0) becomes
790: \begin{eqnarray}
791: {\rm (b0)} & \rightarrow &
792: \frac{i}{4} \oint\frac{d^{4}q}{(2\pi)^{4}}
793: \tanh \left(\frac{q_0}{2T_c}\right) \nonumber\\
794: &&\times {\rm Tr}
795: \Biggl[
796: \biggl(
797: -g^2 T_c
798: \sum_{n~ {\rm odd}}
799: \int\frac{d^{3}k}{(2\pi)^{3}}
800: {\cal D}_{\mu\nu}^{\alpha\beta} (q-k)
801: \gamma^{\mu} {\left( \frac{\lambda^{\alpha}}{2} \right)^{\rm T}} \nonumber\\
802: &&~~~~~~~~~~~\times G^{(21)}(k)
803: \gamma^{\nu} {\left( \frac{\lambda^{\beta}}{2} \right)}
804: \biggl)
805: G^{(12)}(q)
806: \Biggl] \Bigg|_{T=T_c} \nonumber\\
807: &=&
808: \frac{i}{4}\oint \frac{d^4q}{(2\pi)^4} \tanh\left(\frac{q_0}{2T}\right)
809: \nonumber\\
810: &&~~~~~~~~~~~
811: \times {\rm Tr}\left [ f({\bf q}) \phi(T) G^{(12)}(q) \right]
812: \Bigg|_{T=T_c} \nonumber\\
813: &=&
814: \frac{i}{4}\oint \frac{d^4q}{(2\pi)^4} \tanh\left(\frac{q_0}{2T}\right)
815: {\rm Tr} [ \Delta(q) G^{(12)}(q)] \Bigg|_{T=T_c} \nonumber\\
816: &=&-\Omega_{1}^{\Delta^2}|_{T=T_c}.
817: \end{eqnarray}
818:
819: The ${\cal O}(t)$ term of (b0) becomes
820: \begin{eqnarray}
821: &&{\rm (b0)} \rightarrow \nonumber\\
822: && (T-T_c)\frac{i}{4} \frac{d}{dT} \biggl\{
823: \oint \frac{d^4k}{(2\pi)^4} \tanh\left(\frac{k_0}{2T}\right)
824: {\rm Tr} [G^{(21)}(k){\tilde\Delta(k)}] \nonumber\\
825: & &+
826: \oint \frac{d^4q}{(2\pi)^4} \tanh\left(\frac{q_0}{2T}\right)
827: {\rm Tr} [ \Delta(q) G^{(12)}(q)]
828: \biggl\} \Bigg|_{T=T_c}
829: \nonumber\\
830: &&=2(T-T_c)\frac{d}{dT}(-\Omega_{1}^{\Delta ^2})|_{T=T_c}.
831: \end{eqnarray}
832:
833: (b1) and (b4) can be calculated in a similar way, and they become
834: \begin{eqnarray}
835: {\rm (b1)}+{\rm (b4)}= N(\mu)\frac{2\delta p_{ij} }{\mu}
836: \ln \left( \frac{T_c}{\mu} \right)
837: \ |\phi_{abij}|^2.
838: \end{eqnarray}
839: Adding (b1) and (b4) to (b0), we finally obtain
840: \begin{eqnarray}
841: \Omega_{2}^{\Delta ^2}&=&{\rm (b0)}+{\rm (b1)}+{\rm (b4)} \nonumber\\
842: &=&-\Omega_{1}^{\Delta^2}|_{T=T_c}
843: +2 (T-T_c)\frac{d}{dT}(-\Omega_{1}^{\Delta ^2})|_{T=T_c} \nonumber\\
844: &&~~~+ N(\mu)\frac{2\delta p_{ij} }{\mu} \ln \left( \frac{T_c}{\mu} \right)
845: \ |\phi_{abij}|^2 \nonumber\\
846: &=&\sum_{abij}
847: N(\mu) \left\{
848: -\frac{\pi}{4 \bar{g}}+ 2\ln \left( \frac{T}{T_c} \right)
849: +\frac{2\delta p_{ij} }{\mu} \ln \left( \frac{T_c}{\mu} \right)
850: \ \right\} \nonumber\\
851: &&~~~~~~~~~~~~~~~~~~~~~~
852: \times |\phi_{abij}|^2 .\label{correction}
853: \end{eqnarray}
854:
855: Combining Eqs.\ (\ref{tree}) and (\ref{correction}), we obtain the final
856: result for the quadratic term of the GL free energy,
857: \begin{eqnarray}
858: \Omega_{GL}^{\Delta ^2}
859: &=&\sum_{abij} N(\mu)
860: \left\{
861: \ln \left( \frac{T}{T_c} \right)
862: - \frac{\delta p_{ij} }{\mu} \ln \left( \frac{\mu}{T_c} \right) \right\}
863: \nonumber \\
864: &&~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~
865: \times |\phi_{abij}|^2,
866: \label{eq:final-form-1}
867: \\
868: \delta p_{ij} &\simeq &
869: \delta \mu_{ij} - (m_i^2+m_j^2)/(4\mu).
870: \label{eq:final-form-2}
871: \end{eqnarray}
872:
873: It can be rewritten in a clearer form,
874: \begin{eqnarray}
875: \Omega_{GL}^{\Delta ^2}
876: &=&\sum_{abij} N(\mu)
877: \left( \frac{T- {T}_{c}^{ij}} {{T}_{c}} \right)| \phi_{abij}|^2,
878: \label{eq:final-form-3} \\
879: \frac{{T}_{c}^{ij}}{T_c}&=&
880: 1+ \frac{\delta p_{ij} }{\mu}
881: \ln \left( \frac{\mu}{T_c} \right).
882: \label{eq:final-form-4}
883: \end{eqnarray}
884:
885: Several comments are in order here.
886: \begin{itemize}
887:
888: \item The free energy correction depends only on the average shift of the
889: Fermi momenta of paired quarks, $\delta p_{ij}$, as can be seen from
890: Eq.\ (\ref{eq:final-form-1}). Furthermore, the correction may be decomposed
891: into a sum of the effects from the averaged chemical potential and the
892: averaged mass squared when $\delta p_{ij}$ is small, as shown in
893: Eq.\ (\ref{eq:final-form-2}).
894:
895: \item
896: Equations (\ref{eq:final-form-3}) and (\ref{eq:final-form-4}) indicate that
897: the shift $\delta p_{ij}$ affects the free energy correction through the
898: critical temperature $T_c^{ij}$ for the $(i,j)$ pairing. The larger the
899: shift $\delta p_{ij}$ is, the higher the melting temperature ${T}_{c}^{ij}$
900: becomes. In other words, one can compare the thermal stability of two
901: different pairs only by taking note of a difference in the average density
902: of states between them.
903:
904: \item
905: The unlocking at $T=0$ is different in mechanism from our unlocking near
906: $T_c$. At $T=0$, color-flavor unlocking due to the Fermi momentum
907: {\em mismatch} between paired quarks is expected at
908: $\mu \sim m_s^2 / \phi_0$ \cite{unlock}. In contrast, the Fermi momentum
909: {\em average} is important for our unlocking near $T_c$.
910:
911: \item
912: We can see that in the correction obtained above quark flavors do not mix
913: with each other. As long as we consider the pairing of positive energy quarks
914: and neglect that of antiquarks, the flavor structure of a gap always enters in
915: the form of $m_i^2$ instead of $m_i m_j (i \neq j)$.
916: \end{itemize}
917:
918: \section{GL potential in realistic quark matter}
919:
920: Up to now we constructed the GL potential for general quark masses
921: and chemical potential shifts assuming that the corrections are small.
922: In this section, we will focus on the case close to a realistic situation
923: by taking $m_{s} \neq 0$ with $m_{u,d}=0$ and by imposing $\beta$
924: equilibrium and charge neutrality conditions. This analysis clarifies the
925: role of the strange quark mass in possible phase structures near the
926: super-normal phase boundary at finite $T$.
927:
928: \subsection{Direct $m_s$ correction through the quark mass term}
929:
930: Let us first consider the correction directly proportional to
931: $m_s^2$ in Eqs.\ (\ref{eq:final-form-1}) and (\ref{eq:final-form-2}).
932: It is easy to see from the flavor structure of Eq.\ (\ref{eq:final-form-1})
933: that $m_s$ affects only $us$ and $ds$ pairings. This is also reasonable
934: from the physical point of view as will be shown below.
935:
936: By inserting $m_s\neq 0$ with $m_{u,d}=0$ and using Eq.\ (\ref{phi-d})
937: with Eq.\ (\ref{parity-even}), the direct $m_s$ correction in Eq.\
938: (\ref{eq:final-form-1}) becomes
939: \begin{eqnarray}\label{epsilon}
940: \epsilon
941: \sum_{a} ( |d_a^u|^2 + |d_a^d|^2 )
942: = \epsilon \sum_{a} (|{\mathbf{d}_a}|^2 -|d_a^s|^2).
943: \end{eqnarray}
944: Here the coefficient $\epsilon$ reads
945: \begin{eqnarray}
946: \label{epsilon-wc}
947: \epsilon \simeq
948: \alpha_0 \frac{m_s^2}{4 \mu^2}
949: \ln \left(\frac{\mu}{T_c}\right)
950: \sim 2 \alpha_0 \sigma,
951: \end{eqnarray}
952: where $\alpha_0=4N(\mu)$ is defined in Eq.\ (\ref{b1b2}), and
953: a dimensionless parameter $\sigma$ is introduced as
954: \begin{eqnarray}
955: \label{eq:full-sigma}
956: \sigma = \left( \frac{3 \pi^2} {8 {\sqrt 2}} \right) \frac{m_s^2} {g \mu^2}.
957: \end{eqnarray}
958: In Eq.\ (\ref{epsilon-wc}) we have used a weak coupling relation,
959: $\ln (T_c/\mu) \sim -{3\pi^2}/(\sqrt{2} g)$, which originates from the
960: long-range color magnetic interaction \cite{brown,PR}.
961:
962: Since the finite $m_s$ decreases the Fermi momentum of $s$ quarks,
963: $\delta p_{us}$ and $\delta p_{ds}$ are smaller than $\delta p_{ud}$, and
964: thus $\epsilon$ becomes positive such that $ud$ pairing is favored over $us$
965: and $ds$ pairings. Consequently, the CFL phase becomes asymmetric in flavor
966: space and its critical temperature is lowered, leading to the appearance of
967: the 2SC phase (${d}_a^i \propto \delta^{is}$) just below $T_c$
968: \cite{abuki}.
969:
970: \subsection{Indirect $m_s$ correction through the charge chemical potentials}
971:
972: We proceed to discuss the correction from the charge chemical
973: potentials, which is proportional to $\delta \mu_{ij}$ in Eqs.\
974: (\ref{eq:final-form-1}) and (\ref{eq:final-form-2}). Under $\beta$
975: equilibrium and charge neutrality, the electron chemical potential $\mu_e$ and
976: the shift $\delta\mu_i$ of the chemical potential of flavor $i$ are related as
977: \begin{eqnarray}\label{delmu}
978: \delta \mu_i =-q_i\mu_e
979: \end{eqnarray}
980: with $q_i$ being the electric charges of the quarks; $q_{u}=2/3$ and
981: $q_{d,s}=-1/3$. In weak coupling, where one may regard normal quark matter
982: and electrons as noninteracting Fermi gases, $\mu_e$ is related to $m_s$ as
983: \begin{eqnarray}\label{muemus}
984: \mu_e=m_s^2/4\mu .
985: \end{eqnarray}
986: This estimate is valid in the vicinity of $T_c$ where corrections to $\mu_e$
987: by a finite pairing gap affect only the quartic terms in the GL potential.
988:
989: By inserting Eqs.\ (\ref{delmu}) and (\ref{muemus}) and using
990: Eq.\ (\ref{phi-d}) with Eq.\ (\ref{parity-even}), the indirect $m_s$
991: correction in Eq.\ (\ref{eq:final-form-1}) becomes
992: \begin{eqnarray}\label{eta}
993: \eta~ (\frac{1}{3} \sum_{a}|{\mathbf{d}_a}|^2
994: -\sum_{a} |d_a^u|^2).
995: \end{eqnarray}
996: Here the coefficient $\eta$ reads
997: \begin{eqnarray}
998: \label{eta-wc}
999: \eta \simeq
1000: \alpha_0 \frac{m_s^2}{8\mu^2}
1001: \ln \left(\frac{\mu}{T_c}\right)
1002: \sim \alpha_0 \sigma .
1003: \end{eqnarray}
1004: Since $\delta \mu_{ds}$ is larger than $\delta \mu_{ud}$ and $\delta \mu_{us}$
1005: and it increases the average Fermi momentum of $ds$ quarks in Eq.\
1006: (\ref{eq:final-form-2}), the indirect $m_s$ correction favors $ds$ pairing
1007: than $ud$ and $us$ pairings.
1008:
1009: Note that we only consider the modification of $T_c$ due to nonzero
1010: $m_s$ through the properties of the superfluid phase. Actually, $T_c$ is
1011: modified not only by the properties of the superfluid phase but also of the
1012: normal phase. The modification by the normal phase enters through the Debye
1013: mass in the gluon propagator ${\cal D}$ in the relation between $T_c$ and $g$,
1014: Eq.\ (\ref{Tcg}). Fortunately the modification due to nonzero $m_s$ and
1015: charge neutrality in normal quark matter is like $T_c \rightarrow
1016: T_c (1+ {\cal O}(g \sigma))$. This modification is of higher order than
1017: that in superfluid quark matter, $T_c \rightarrow T_c (1+ {\cal O}(\sigma))$,
1018: to be derived in Sec.\ V.
1019:
1020: \subsection{Effect of color neutrality}
1021:
1022: We consider color neutrality of the system as well. In contrast to the
1023: case at $T=0$, however, it affects only the quartic terms in the GL potential
1024: because the possible chemical potential differences between colors are
1025: \cite{IB,RIS}
1026: \begin{eqnarray}
1027: \delta{\mu}_{ab}&=&
1028: -\frac{1}{9\mu}\ln\left(\frac{T_{c}}{\mu}\right)
1029: [3\sum_i (d_{a}^i d_{b}^{i*})
1030: - \delta_{ab} \sum_{c,i} |d_c^i|^{2}], \nonumber \\
1031: \label{muaGL}
1032: \end{eqnarray}
1033: which are already of quadratic order in the gap. Here $\delta \mu_{ab}=
1034: \mu_{ab} -\mu \delta_{ab}$, and ${\mu}_{ab}$ is the chemical potential
1035: conjugate to $n_{ab}=\sum_i \bar{\psi}_{bi} \gamma^0 \psi_{ai}$ \cite{IB}.
1036: In weak coupling the magnitude of the correction to the quartic terms is
1037: suppressed by ${\cal O}((T_c/g\mu)^2)$ compared to the leading quartic terms.
1038: Thus color neutrality has no essential consequence to the phase transitions
1039: considered in this paper. A major difference between the corrections from the
1040: charge neutrality and the color neutrality is that the former shifts the
1041: quark chemical potentials even in the normal phase, while the latter works
1042: only when the pairing occurs. This is why the former is more important
1043: than the latter near $T_c$.
1044:
1045: \subsection{Effect of instantons}
1046:
1047: In QCD with three flavors, instantons induce chirality-flipping
1048: six-fermion interactions. This interaction leads to a sextic term in the gap
1049: in the chiral limit, which is irrelevant near the super-normal phase boundary.
1050: If the strange quark mass enters, the direct instantons induce an effective
1051: four-fermion interaction between $u$ and $d$ quarks \cite{Sch}. This leads to
1052: a quadratic term in the GL potential, $\xi~\sum_a |d_a^s|^2$, which
1053: corresponds to the $a_1$ term in Eq.\ (\ref{generalGL}). An explicit
1054: calculation in weak coupling shows that $\xi\sim -\alpha_0 (m_s/\mu)
1055: (\Lambda_{\rm QCD}/\mu)^9 (1/g)^{14}$.
1056: The negative sign indicates that the instanton effect favors $ud$ pairing
1057: as does one-gluon exchange [see Eq.\ (\ref{epsilon})]. However, the magnitude
1058: of $\xi$ is highly suppressed at high densities. Therefore we will ignore
1059: this term in the following.
1060:
1061: In summary of Secs.\ IV.A--D, as far as we are close to the super-normal
1062: phase boundary at high density, we have only to consider the corrections
1063: from the strange quark mass and the electric charge neutrality, which favor
1064: $ud$ pairing and $ds$ pairing, respectively.
1065:
1066: \subsection{Validity of the approximations}
1067:
1068: Our analysis of the GL potential near $T_c$ is valid as long as
1069: (a) $g \ll 1$, (b) $\sigma \ll 1$, and (c) $T_c \ll g \mu $. The condition
1070: (a) is necessary for the dominance of the long-range magnetic interaction
1071: acting between quarks. The condition (b) allowed us to consider the
1072: corrections to the GL potential by the strange quark mass and differences
1073: in the chemical potential among flavors up to quadratic order in the gap.
1074: The condition (c) is imposed so that the correction to the quartic terms in
1075: the GL potential from the color neutrality can be neglected. These conditions
1076: are all satisfied at asymptotically high density.
1077:
1078: If we use the result for $T_c$ calculated in weak coupling
1079: \cite{brown,PR}, one finds that (c) is a consequence of (a). The conditions
1080: (a) and (b) can also be combined into
1081: \begin{eqnarray}
1082: \frac{m_s^2}{\mu^2} \ll g \ll 1.
1083: \end{eqnarray}
1084:
1085:
1086: \section{Melting pattern of diquark condensates}
1087:
1088: In this section, we clarify the phase structure near the transition
1089: temperature using the GL potential corrected by Eqs.\ (\ref{epsilon}) and
1090: (\ref{eta}). We start from assuming the form of the condensation and derive
1091: the temperature dependence of the gaps that minimize the potential. Then we
1092: discuss the properties of quark quasiparticles and transverse gluons in
1093: each of the phases near the transition temperature.
1094:
1095: \subsection{Phase structure with finite $m_s$ and charge neutrality}
1096:
1097: Since the two effects of nonzero $m_s$, characterized by Eqs.\
1098: (\ref{epsilon}) and (\ref{eta}), favor $ud$ pairing and $ds$ pairing,
1099: respectively, the finite temperature transition from the CFL to
1100: the normal phase at $m_s=0$ is significantly modified. In fact, as we show
1101: in detail below, successive color-flavor unlockings take place instead of a
1102: simultaneous unlocking of all color-flavor combinations. To describe this
1103: {\em hierarchical thermal unlocking}, it is convenient to
1104: introduce a parameterization
1105: \begin{eqnarray}
1106: \label{s}
1107: d_a^i
1108: =
1109: \left(
1110: \begin{array}{lll}
1111: \Delta_1&0&0 \\
1112: 0&\Delta_2&0 \\
1113: 0&0&\Delta_3 \\
1114: \end{array}
1115: \right).
1116: \end{eqnarray}
1117: We assume $\Delta_{1,2,3}$ to be real. We also name the phases for later
1118: convenience as \cite{IMTH04}
1119: \begin{eqnarray}
1120: \label{phase-def}
1121: \begin{array}{llcl}
1122: \Delta_{1,2,3} \neq 0 & & : & {\rm mCFL}, \\
1123: \Delta_1=0, & \Delta_{2,3}\neq 0 & :& {\rm uSC}, \\
1124: \Delta_2=0, & \Delta_{1,3} \neq 0 & :& {\rm dSC}, \\
1125: \Delta_3=0, & \Delta_{1,2} \neq 0 & :& {\rm sSC}, \\
1126: \Delta_{1,2}=0, & \Delta_3 \neq 0 &: & {\rm 2SC},
1127: \end{array}
1128: \end{eqnarray}
1129: where dSC (uSC, sSC) stands for superconductivity in which for $d$ ($u$, $s$)
1130: quarks all three colors are involved in the pairing.
1131:
1132: In terms of the parameterization (\ref{s}), the GL potential with
1133: corrections of ${\cal O}(m_s^2)$ to the quadratic term,
1134: Eqs.\ (\ref{epsilon}) and (\ref{eta}), reads
1135: \begin{eqnarray}
1136: \label{new-GL}
1137: \Omega &=&\bar{\alpha}' (\Delta_1^2+\Delta_2^2+\Delta_3^2)
1138: - \epsilon \Delta_3^2 - \eta \Delta_1^2\nonumber \\
1139: &+& \beta_1 (\Delta_1^2+\Delta_2^2+\Delta_3^2)^2
1140: + \beta_2 ( \Delta_1^4+\Delta_2^4+\Delta_3^4),
1141: \end{eqnarray}
1142: where $\bar{\alpha}' =\bar{\alpha}+\epsilon+\frac{\eta}{3}$.
1143:
1144: We proceed to analyze the phase structure dictated by Eq.\
1145: (\ref{new-GL}) with the weak coupling parameters (\ref{b1b2}),
1146: (\ref{epsilon-wc}), and (\ref{eta-wc}) up to leading order in $g$.
1147: A stable condensation must satisfy the gap equations
1148: \begin{eqnarray}
1149: \partial \Omega /\partial\Delta_{1,2,3}=0,
1150: \end{eqnarray}
1151: and it must minimize the potential Eq.\ (\ref{new-GL}). We compare the
1152: energies of all condensates in Eq.\ (\ref{phase-def}) and decide the most
1153: stable condensation at each temperature near $T_c$.
1154:
1155: In Figs.\ 3 and 4 the results thus obtained for the phase structure
1156: near $T_c$ are summarized.
1157:
1158: Figure 3(a) shows the second-order phase transition, CFL $\to$
1159: normal for $m_s=0$. Figures 3(b,c) represent how the phase transitions and
1160: their critical temperatures bifurcate as we introduce (b) effects of a nonzero
1161: $m_s$ in the quark propagator and then (c) effects of charge neutrality.
1162: In case (b), two second-order phase transitions arise:
1163: mCFL (with $\Delta_1=\Delta_2$) $\to$ 2SC at
1164: $T=T_c^s \equiv (1-4 \sigma)T_c$,
1165: and 2SC $\to$ normal at $T=T_c$. In case (c), there arise three
1166: successive second-order phase transitions, mCFL $\to$ dSC at $T=T_c^{\rm I}$,
1167: dSC $\to$ 2SC at $T=T_c^{\rm II}$, and 2SC $\to$ normal at $T=T_c^{\rm III}$.
1168: Shown in Fig.\ 4 is the $T$-dependence of the gaps $\Delta_{1,2,3}$ for
1169: the case (c). All the gaps are continuous functions of $T$, but their slopes
1170: are discontinuous at the critical points, which reflects the second order
1171: nature of the transitions in the mean-field treatment of Eq.\ (\ref{new-GL}).
1172:
1173: \begin{figure}[t]
1174: \begin{center}
1175: \includegraphics[width=7cm]{Tc.eps}
1176: \end{center}
1177: \vspace{-0.5cm}
1178: \caption{Transition temperatures of the
1179: three-flavor color superconductor in weak coupling:
1180: (a) all quarks are massless;
1181: (b) nonzero $m_s$ in the quark propagator is considered;
1182: (c) electric charge neutrality is further imposed.
1183: The numbers attached to the arrows are in units of $\sigma T_c$. From Ref.\
1184: \cite{IMTH04}.
1185: }
1186: \label{fig:Tc}
1187: \end{figure}
1188:
1189: \begin{figure}[t]
1190: \begin{center}
1191: \includegraphics[width=7cm]{conden.eps}
1192: \end{center}
1193: \vspace{-0.5cm}
1194: \caption{A schematic illustration of the gaps squared
1195: as a function of $T$. From Ref.\ \cite{IMTH04}.}
1196: \label{fig:conden}
1197: \end{figure}
1198:
1199: We may understand the bifurcation of the transition temperatures
1200: as follows. In the massless case (a), $T_c$ is degenerate
1201: between the CFL and 2SC phases, the chemical potential is common to all
1202: three flavors and colors, and the CFL phase is more favorable than
1203: the 2SC phase below $T_c$. As one goes from (a) to (b), the density of
1204: states of the $s$ quarks at the Fermi surface is reduced. Then the critical
1205: temperature for the CFL phase is lowered, and the 2SC phase is allowed
1206: to appear at temperatures between $T_c^s$ and $T_c$. As one goes from (b)
1207: to (c), the average chemical potential of $ds$ ($ud$ and $us$) quarks
1208: increases (decreases). Accordingly, the transition temperatures
1209: further change from $T_c$ to $T_c^{\rm III}$ and from $T_c^s$ to
1210: $T_c^{\rm I}$ and $T_c^{\rm II}$.
1211:
1212: Now we examine in more detail how the color-flavor unlocking in case
1213: (c) proceeds with increasing $T$ from the region below $T_c^{\rm I}$.
1214:
1215: \noindent
1216: (i)
1217: Just below $T_c^{\rm I}$, we have a CFL-like phase, but the three gaps
1218: take different values, with an order $\Delta_3 > \Delta_1 > \Delta_2 \neq 0$
1219: (the mCFL phase). The reason why this order is realized can be understood
1220: from the GL potential (\ref{new-GL}). The $\epsilon$-term and $\eta$-term in
1221: Eq.\ (\ref{new-GL}) tend to destabilize $us$ pairing ($\Delta_2$)
1222: relative to $ud$ pairing ($\Delta_3$) and $ds$ pairing ($\Delta_1$), and
1223: since $\epsilon > \eta (> 0)$, $ds$ pairing is destabilized more effectively
1224: than $ud$ pairing. The value of each gap in the mCFL phase reads
1225: \begin{eqnarray} \label{gap1}
1226: \Delta_i^2=\frac{\alpha_0}{8\beta} \left(
1227: \frac{T_c-T}{T_c} + c_i \sigma
1228: \right) ,
1229: \end{eqnarray}
1230: with $c_{1,2,3}=(-4/3, -16/3, 8/3)$.
1231: The mCFL phase has only a global symmetry $U(1)_{C+L+R} \times U(1)_{C+L+R}$
1232: in contrast to the global symmetry $SU(3)_{C+L+R}$ in the CFL phase with
1233: $m_{u,d,s}=0$. The remaining $U(1)$ generators are
1234: \begin{eqnarray}
1235: {\cal Q}_1&=&T_3- S_3, \nonumber\\
1236: {\cal Q}_2&=&T_8-S_8.
1237: \end{eqnarray}
1238: Here $T_{a}$ ($S_{i}$) are the part of the generators of $SU(3)_C$
1239: ($SU(3)_{L+R}$) with explicit forms $T_3=S_3=({1}/{2}){\rm diag}(1,-1,0)$ and
1240: $T_8=S_8=({1}/{2 {\sqrt 3}}){\rm diag}(1,1,-2)$. The free energy in this
1241: phase is
1242: \begin{eqnarray}
1243: \Omega_{mCFL}=-\frac{3\alpha_0^2}{16\beta}\left\{ \left(\frac{T-T_c}{T_c}
1244: +\frac{4}{3}\eta \right)^2
1245: +\frac{8}{3}\sigma^2 \right\}.
1246: \end{eqnarray}
1247: As $T$ increases, the first unlocking transition, the unlocking of
1248: $\Delta_2$ (the pairing between $Bu$ and $Rs$ quarks), takes place at the
1249: critical temperature,
1250: \begin{eqnarray}
1251: \label{Tc-1}
1252: T_c^{\rm I} = \left( 1 - \frac{16}{3} \sigma \right) T_c .
1253: \end{eqnarray}
1254:
1255: \noindent
1256: (ii) For $T_c^{\rm I} < T < T_c^{\rm II}$, $\Delta_2=0$ and
1257: \begin{eqnarray} \label{gap2}
1258: \Delta_i^2= \frac{\alpha_0}{6 \beta}
1259: \left(
1260: \frac{T_c-T}{T_c} + c_i \sigma
1261: \right) ,
1262: \end{eqnarray}
1263: with $c_{1,3}=(-7/3, 2/3)$.
1264: In this phase, we have only $ud$ and $ds$ pairings (the dSC phase), and there
1265: is a manifest symmetry, $U(1)_{C+L+R} \times U(1)_{C+L+R} \times
1266: U(1)_{C+V+B} \times U(1)_{C+V+B}$, where the corresponding $U(1)$ generators
1267: are
1268: \begin{eqnarray}
1269: {\cal Q}_1&=&T_3- S_3, \nonumber \\
1270: {\cal Q}_2&=&T_8-S_8, \nonumber \\
1271: {\cal Q}_3&=&Q+\frac{2}{\sqrt 3} T_8- 2 S_3, \nonumber \\
1272: {\cal Q}_4&=&Q+\frac{2}{\sqrt 3} S_8- 2 S_3, \label{qqq}
1273: \end{eqnarray}
1274: respectively.
1275: Here $Q={2}/{3}$ is the generator of baryon charge ($Q$) and acts on the gap
1276: as Eq.\ (\ref{transform1}).
1277: The free energy in this phase is
1278: \begin{eqnarray}
1279: \Omega_{dSC}=-\frac{\alpha_0^2}{6\beta}\left\{\left(\frac{T-T_c}{T_c}+\frac{5}{6}\eta \right)^2
1280: +\frac{3}{4}\sigma^2 \right\}.
1281: \end{eqnarray}
1282: At $T=T_c^{\rm II}$, the second unlocking transition, the unlocking of
1283: $\Delta_1$ (the pairing between $Gs$ and $Bd$ quarks), takes place at the
1284: critical temperature,
1285: \begin{eqnarray}
1286: \label{Tc-2}
1287: T_c^{\rm II} = \left( 1 -\frac{7}{3}\sigma \right) T_c.
1288: \end{eqnarray}
1289:
1290: \noindent
1291: (iii) For $T_c^{\rm II} < T < T_c^{\rm III}$, one finds the 2SC phase, which
1292: has only $ud$ pairing with
1293: \begin{eqnarray} \label{gap3}
1294: \Delta_3^2=\frac{\alpha_0}{4 \beta}
1295: \left(
1296: \frac{T_c-T}{T_c}
1297: -\frac{1}{3}\sigma
1298: \right) .
1299: \end{eqnarray}
1300: The 2SC phase has a symmetry $SU(2)_C \times SU(2)_{L+R} \times U(1)_{C+B}
1301: \times U(1)_{L+R+B}$, where the corresponding $U(1)$ generators are
1302: \begin{eqnarray}
1303: {\cal Q}_5&=&Q+{\sqrt 3} T_8, \nonumber \\
1304: {\cal Q}_6&=&Q+{\sqrt 3} S_8,
1305: \end{eqnarray}
1306: respectively.
1307: The free energy in this phase is
1308: \begin{eqnarray}
1309: \Omega_{2SC}=-\frac{\alpha_0^2}{8\beta}\left(\frac{T-T_c}{T_c}+\frac{1}{3}\sigma\right)^2 .
1310: \end{eqnarray}
1311: The final unlocking transition where
1312: $\Delta_3$ (the pairing between $Rd$ and $Gu$ quarks) vanishes occurs at
1313: \begin{eqnarray}
1314: \label{Tc-3}
1315: T_c^{\rm III} = \left( 1 -\frac{1}{3}\sigma \right) T_c .
1316: \end{eqnarray}
1317: Above $T_c^{\rm III}$, the system is in the normal phase. Note that in the
1318: temperature region $(T_c^{\rm I}-T)/T_c^{\rm I} < {\rm const} \cdot \sigma$,
1319: the gaps satisfy the condition ($\ref{eq:t-sig2}$), so the GL potential
1320: expanded up to quartic order in the gap is valid.
1321:
1322: So far, we have derived the parameters $\beta_{1,2}$, $\epsilon$, and
1323: $\eta$ in the weak coupling approximation and discussed the phase structure
1324: expected at asymptotically high density. Now we relax the weak coupling
1325: constraint and study possible phase structures expected from the GL potential
1326: of the form Eq.\ (\ref{new-GL}) with arbitrary coupling strengths. To
1327: simplify the argument, we assume $\beta_1=\beta_2 >0$ and draw the phase
1328: diagram in the space of the parameters $\epsilon$ and $\eta$ in Fig.\
1329: \ref{fig:epsilon-eta}.
1330:
1331: If we consider the case where the mCFL phase appears at temperature
1332: sufficiently below $T_c$, the relative magnitudes of the gaps behave
1333: as shown in Fig.\ \ref{fig:epsilon-eta}. If $\Delta_{2}$ is the smallest of
1334: the three gaps, we call the melting pattern ``dSC-type'' because the dSC phase
1335: would appear as $T$ increases. ``uSC-type'' and ``sSC-type'' are defined in
1336: a similar way. As in Fig.\ \ref{fig:epsilon-eta}, depending on the relative
1337: magnitude of $\epsilon$ and $\eta$, the dSC-type melting pattern is divided
1338: into two different classes in which the 2SC-type phase
1339: ($\Delta_1 \neq 0, \Delta_{2,3}=0$ or $\Delta_3 \neq 0, \Delta_{1,2}=0$)
1340: appears just below the transition temperature to the normal phase.
1341: The uSC-type and sSC-type melting patterns also have such a two-fold
1342: structure. The {\em hierarchical thermal unlocking} (mCFL $\rightarrow$
1343: dSC $\rightarrow$ 2SC $\rightarrow$ normal) is realized in weak coupling
1344: as shown by $\otimes$ in the figure. The figure indicates that
1345: it is a rather robust phenomenon near the super-normal phase boundary.
1346: In other words, the uSC or sSC phase appears only when $\epsilon$ or
1347: $\eta$ changes sign at least under the form of the GL potential in Eq.\
1348: (\ref{new-GL}) with an assumption $\beta_1=\beta_2>0$. We mention here
1349: that a recent analysis in Ref.\ \cite{Fukushima} using the Nambu-Jona-Lasinio
1350: (NJL) model shows that the signs and the ratios of the coupling strengths in
1351: the GL potential take similar values with our weak coupling values:
1352: $\beta_1/\beta_2=1$ (with $\beta_1>0$ and $\beta_2>0$)
1353: and $\zeta \equiv \epsilon / \eta \cong 2$--4 (with $\epsilon >0$ and
1354: $\eta >0$).
1355:
1356: \begin{figure}[t]
1357: \begin{center}
1358: \includegraphics[width=7cm]{epsilon-eta.eps}
1359: \end{center}
1360: \vspace{-0.5cm}
1361: \caption{
1362: The phase structure obtained from Eq.\ (\ref{new-GL}) in ($\epsilon, \eta$)
1363: plane for $\beta_1=\beta_2>0$. The mCFL phase is assumed to appear at
1364: sufficiently low temperature. $\otimes$ indicates the point calculated
1365: in the weak coupling approximation.
1366: }
1367: \label{fig:epsilon-eta}
1368: \end{figure}
1369:
1370: \subsection{Properties of quark and gluon modes}
1371:
1372: Next we resume considering about the weak coupling limit and discuss the
1373: properties of quark and gluon modes such as the excitation energies
1374: in each of the mCFL, dSC, and 2SC phases. The low energy excitations
1375: of these phases are massive or massless gauge bosons, gapped or gapless
1376: quark quasiparticles, and modes that are associated with fluctuations in
1377: the diquark fields. Here we concentrate on the gauge bosons and gapless
1378: fermions. As for gauge bosons, we only consider the transverse components
1379: of the gauge fields which, in the static limit, can become massive only
1380: in a superconductor. The longitudinal components are Debye screened
1381: in the vicinity of the transition temperature. Note that the
1382: "gapless fermions" in our previous paper \cite{IMTH04} have been defined only
1383: by diagonalizing the self-energy $\Sigma$ in Eq.\ (\ref{GSig}). In the
1384: present paper, we have diagonalized the full propagator $G^{-1}(k)$ to
1385: define the gapless fermions in a more precise manner as will be shown below.
1386:
1387:
1388: In Table I, we summarize the symmetries, the number of massive gauge
1389: bosons, and the number of gapless fermion modes in each phase. As we will
1390: show in detail below, the number of gapless quark modes depends not only on
1391: which phase we consider but also on the temperature itself. In each phase,
1392: in addition to the unpaired quark modes whose excitation energies are
1393: naturally zero, gapless modes appear if the temperature is in the region
1394: where the gaps satisfy the conditions listed in Table I.
1395:
1396: In each phase, the gauge symmetry is broken partially or even totally.
1397: The Meissner masses for the general condensate, Eq.\ (\ref{phase-def}),
1398: and the corresponding massive (or massless) gauge fields can be calculated
1399: using the GL theory coupled to gluon fields \cite{IB3} as
1400: \begin{eqnarray}
1401: &&m_{A_1}^2=m_{A_2}^2= \kappa_T g^2 (\Delta_1^2 +\Delta_2^2), \nonumber\\
1402: &&m_{A_4}^2=m_{A_5}^2=\kappa_T g^2 (\Delta_1^2 +\Delta_3^2), \nonumber\\
1403: &&m_{A_6}^2=m_{A_7}^2=\kappa_T g^2 (\Delta_2^2 +\Delta_3^2),
1404: \label{gluon-masss}\\
1405: &&m_{A_8}^2= \frac{4}{3}\kappa_T g^2
1406: \left(
1407: \frac{ \Delta_1^2 \Delta_2^2 +\Delta_2^2 \Delta_3^2 +\Delta_3^2 \Delta_1^2 }
1408: {\Delta_1^2+ \Delta_2^2 }
1409: \right), \nonumber\\
1410: &&m_{\tilde{A}}^2=\frac{4}{3}\kappa_T g^2
1411: \left(
1412: \frac{ \Delta_1^4 +\Delta_2^4 +\Delta_1^2 \Delta_2^2 }
1413: {\Delta_1^2+ \Delta_2^2 }
1414: \right) . \nonumber
1415: \end{eqnarray}
1416: Here $\tilde{A}$ is a mixed field of $A_3$ and $A_8$,
1417: \begin{eqnarray}
1418: \tilde{A}&=&
1419: \frac{\sqrt 3}{2}
1420: \frac{\Delta_1^2+ \Delta_2^2 }{\sqrt{ \Delta_1^4 +\Delta_2^4 +\Delta_1^2
1421: \Delta_2^2 }}A_3 \nonumber\\
1422: &&~~~~~~~~~+\frac{1}{2} \frac{\Delta_1^2-\Delta_2^2 }
1423: {\sqrt{ \Delta_1^4 +\Delta_2^4 +\Delta_1^2 \Delta_2^2 }}A_8.
1424: \end{eqnarray}
1425: Here we followed the notations in Refs.\ \cite{IB, IB3}. $\kappa_T$ is the
1426: stiffness parameter which controls the spatial variation of the gap. In
1427: weak coupling it is proportional to the quartic coupling $\beta$ as \cite{IB3}
1428: \begin{eqnarray}
1429: \kappa_T=\beta/3 = \frac{7\zeta(3)}{24(\pi T_c)^{2}}N(\mu).
1430: \end{eqnarray}
1431: We can reproduce the results in Table I in Ref.\ \cite{MIHB} by taking a limit
1432: $\Delta_{1,2} \rightarrow 0$ for the isoscalar 2SC phase and
1433: $\Delta_1=\Delta_2=\Delta_3$ for the CFL phase.
1434:
1435: As given in a full description in Ref.\ \cite{gapless-AKR}, we can count
1436: the gapless quark modes by examining the spectra of which the full propagator
1437: diverges:
1438: \begin{eqnarray} \label{fullD}
1439: {\rm det} G^{-1}(E, {\bf q})= {\cal G} _{Rd,Gu} {\cal G} _{Rs,Bu}
1440: {\cal G} _{Gs,Bd} {\cal G} _{Ru,Gd,Bs}= 0. \nonumber \\
1441: \end{eqnarray}
1442: The gap matrix $\Sigma (k)$ can be written in block-diagonal form if we
1443: adequately take the base whereas the noninteracting quark propagator $G_0 (k)$
1444: is diagonal. The determinant Eq.\ (\ref{fullD}) can be decomposed
1445: into three 4 $\times$ 4 determinants ${\cal G}_{Rd,Gu}$, ${\cal G}_{Rs,Bu}$,
1446: and ${\cal G}_{Gs,Bd}$, and one 6 $\times$ 6 determinant
1447: ${\cal G}_{Ru,Gd,Bs}$, which may effectively be written as
1448: \begin{eqnarray}
1449: \label{det1}
1450: {\cal G} _{Rd,Gu}
1451: &=& \det
1452: \left(
1453: \begin{array}{ll}
1454: \qs -\hat{p}_{F}^{d} & -\Delta_3\\
1455: -\tilde{\Delta}_3& ~\qs +\hat{p}_{F}^{u}
1456: \end{array}
1457: \right) \nonumber\\
1458: &\times &\det
1459: \left(
1460: \begin{array}{ll}
1461: \qs -\hat{p}_{F}^{u} & - \Delta_3 \\
1462: -\tilde{\Delta}_3& ~\qs +\hat{p}_{F}^{d}
1463: \end{array}
1464: \right) ,
1465: \end{eqnarray}
1466: \begin{eqnarray}
1467: \label{det2}
1468: {\cal G} _{Rs,Bu}
1469: &=& \det
1470: \left(
1471: \begin{array}{ll}
1472: \qs -\hat{p}_{F}^{s} & -\Delta_2\\
1473: -\tilde{\Delta}_2& ~\qs +\hat{p}_{F}^{u}
1474: \end{array}
1475: \right) \nonumber\\
1476: & \times &\det
1477: \left(
1478: \begin{array}{ll}
1479: \qs -\hat{p}_{F}^{u} & -{\Delta}_2\\
1480: -\tilde{\Delta}_2& ~\qs +\hat{p}_{F}^{s}
1481: \end{array}
1482: \right),
1483: \end{eqnarray}
1484: \begin{eqnarray}
1485: \label{det3}
1486: {\cal G} _{Gs,Bd}
1487: &=& \det
1488: \left(
1489: \begin{array}{ll}
1490: \qs -\hat{p}_{F}^{s} & -\Delta_1\\
1491: -\tilde{\Delta}_1 & ~\qs +\hat{p}_{F}^{d}
1492: \end{array}
1493: \right) \nonumber\\
1494: & \times &\det
1495: \left(
1496: \begin{array}{ll}
1497: \qs -\hat{p}_{F}^{d} & -\Delta_1\\
1498: -\tilde{\Delta}_1 & ~\qs +\hat{p}_{F}^{s}
1499: \end{array}
1500: \right),
1501: \end{eqnarray}
1502: \begin{eqnarray}
1503: {\cal G} _{Ru, Gd, Bs}
1504: =
1505: ~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~
1506: ~~~~~~~~~~~~~ &&\nonumber\\
1507: \det
1508: \left(
1509: \begin{array}{llllll}
1510: \qs -\hat{p}_{F}^{u} & 0& 0 & 0& \Delta_3 & \Delta_2 \\
1511: 0 & \qs -\hat{p}_{F}^{d}&0 & \Delta_3 & 0 &\Delta_1 \\
1512: 0& 0& \qs -\hat{p}_{F}^{s} & \Delta_2 & \Delta_1&0 \\
1513: 0& \tilde{\Delta}_3 & \tilde{\Delta}_2 & \qs +\hat{p}_{F}^{u} &0 & 0 \\
1514: \tilde{\Delta}_3 & 0 &\tilde{\Delta}_1 &0 & \qs +\hat{p}_{F}^{d}&0 \\
1515: \tilde{\Delta}_2 & \tilde{\Delta}_1&0 &0& 0& \qs +\hat{p}_{F}^{s}
1516: \end{array}
1517: \right). \nonumber \\
1518: \label{det4}
1519: \end{eqnarray}
1520: Here $\hat{p}_{F}^{i}=p_F^{i} \gamma_0$, and $p_F^{i}$ are the Fermi momenta
1521: of $i$ quarks. Only in Eqs.\ (\ref{det1})--(\ref{det4}) we abbreviate
1522: the Lorentz structure of the gap for simplicity; here $\Delta_i$ indicates
1523: $\gamma_5 \Lambda^{+} \Delta_i$, and $\tilde{\Delta}_i$ indicates
1524: $\gamma_0 (\gamma_5 \Lambda^{+} \Delta_i)\gamma_0$. Moreover, we include the
1525: effect of $m_s$ as a shift of the effective chemical potential $p_F^i$
1526: following Ref.\ \cite{gapless-AKR}. Using Eqs.\ (\ref{fermi-mom}),
1527: (\ref{delmu}), and (\ref{muemus}), $p_F^{i}$'s can be expressed as
1528: \begin{eqnarray}
1529: p_{F}^{u} & =&\mu -(2/3)\mu_e, \nonumber \\
1530: p_{F}^{d}&=&\mu +(1/3)\mu_e, \\
1531: p_{F}^{s} &=&\mu -(5/3)\mu_e. \nonumber
1532: \end{eqnarray}
1533: In the vicinity of the critical temperature, chemical potential differences
1534: between colors $\delta \mu_{ab}$ can be ignored as stated in the previous
1535: sections.
1536:
1537: As for the three 4 $\times$ 4 determinants, we can count the gapless
1538: modes as follows. First of all, ($Rs$,$Bu$) and ($Bu$,$Bd$,$Rs$,$Gs$)
1539: quarks are gapless in the dSC and 2SC phases, respectively, because they do
1540: not participate in the pairing. Even for quarks participating in the
1541: pairing, they can produce gapless excitations near the critical point.
1542: As an example, consider a Fermi system composed of two species having
1543: different Fermi momenta, $p_{F}^{(1)}$ and $p_{F}^{(2)}$, and a
1544: superconducting gap $\Delta$ of pairing with each other. Then the
1545: quasiparticle spectra obtained after the diagonalization of the matrix
1546: fermion propagator can be gapless when the gap is smaller than half of the
1547: difference of the Fermi momenta,
1548: \begin{eqnarray} \label{criterion}
1549: |p_{F}^{(1)}-p_{F}^{(2)}| > 2 \Delta.
1550: \end{eqnarray}
1551: In Ref.\ \cite{gapless-MS} the detailed analysis of the gapless
1552: superconductivity at finite temperature has been done; the chemical potentials
1553: and gaps have been calculated self-consistently at finite temperature.
1554:
1555: In our case, the difference of the chemical potentials depends on
1556: temperature only in higher orders so that the left hand side of Eq.\
1557: (\ref{criterion}) can be regarded as constant in temperature.
1558: Then, we can solve Eq.\ (\ref{criterion})
1559: analytically using Eqs.\ (\ref{gap1}), (\ref{gap2}), and (\ref{gap3}).
1560: When the conditions shown below for the mCFL, dSC and 2SC phases
1561: are satisfied, we obtain two more gapless modes in each phase.
1562: The conditions are
1563: \begin{itemize}
1564: \item{for the mCFL phase,
1565: $\Delta_2 < \frac{1}{2} \mu_e$. }
1566:
1567: \item {for the dSC phase,
1568: $\Delta_1 < \mu_e$. }
1569:
1570: \item {for the 2SC phase,
1571: $\Delta_3 < \frac{1}{2} \mu_e$}.
1572: \end{itemize}
1573: For example, let us consider the temperature satisfying $T_c^{\rm III} < T <
1574: T_c^{\rm II}$ (the dSC region). We basically have two unpaired quark modes
1575: ($Bu$ and $Rs$). Additionally, if the temperature satisfies $\Delta_1 <
1576: \mu_e $, then we obtain two more gapless modes.
1577:
1578: We move on to the 6 $\times$ 6 determinant. In this sector, there are
1579: basically no unpaired quark modes in the dSC and mCFL phases and one unpaired
1580: quark mode ($Bs$) in the 2SC phase. In order to find additional gapless modes,
1581: we examine the solutions which satisfy the equation
1582: \begin{eqnarray}
1583: 0&=&{\cal G} _{Ru, Gd, Bs}(E=0,{\bf q})^{1/2} \nonumber \\
1584: &=& (|{\bf q}|+ p_{F}^{u})^2(|{\bf q}|+ p_{F}^{d})^2
1585: (|{\bf q}|+ p_{F}^{s})^2 \nonumber \\
1586: &&[\{ x (x-\mu_e) (x+\mu_e) + x \Delta_1^2 +(x- \mu_e) \Delta_2^2 \nonumber \\
1587: &&~~~~~~~~~~+ (x+\mu_e) \Delta_3^2 \}^2
1588: +4\Delta_1^2 \Delta_2^2 \Delta_3^2].
1589: \end{eqnarray}
1590: Here $x=|{\bf q}|-\bar{\mu}$, where $\bar{\mu}=\mu-2\mu_e/3$ is the chemical
1591: potential averaged over $Ru$, $Gd$, and $Bs$ quarks. Then we find
1592: \begin{itemize}
1593: \item {for the mCFL phase, no more gapless modes.}
1594:
1595: \item {for the dSC phase, one more gapless mode corresponding to the
1596: mixture of $Ru$, $Gd$, and $Bs$ quarks. }
1597:
1598: \item{for the 2SC phase, if the gap satisfies $\Delta_3 < \frac{1}{2} \mu_e$,
1599: we obtain two more gapless modes corresponding to the mixtures of
1600: $Ru$ and $Gd$ quarks.}
1601:
1602: \end{itemize}
1603:
1604:
1605: In Fig.\ \ref{fig:gapless2}, we show the number of gapless quark modes
1606: together with the gaps as a function of temperature. We can see that the
1607: gapless modes always exist just below the critical temperatures
1608: $T^{\rm I}_c$, $T^{\rm II}_c$, and $T^{\rm III}_c$. The number of the
1609: gapless modes decreases as the temperature lowers at high density.
1610:
1611: \begin{figure}[t]
1612: \begin{center}
1613: \includegraphics[width=8.5cm]{gapless2.eps}
1614: \end{center}
1615: \vspace{-0.5cm}
1616: \caption{The number of gapless quark modes as a function of temperature,
1617: together with the gaps and the electron chemical potential.}
1618: \label{fig:gapless2}
1619: \end{figure}
1620:
1621: \section{Summary and concluding remarks}
1622:
1623: In this paper, we have investigated thermal phase transitions of the
1624: color superconducting quark matter in the presence of finite strange quark
1625: mass and under charge neutrality. We constructed a general form of the GL
1626: potential in such a way that it fulfills symmetry constraints and then
1627: obtained the weak coupling expression for the GL potential including
1628: the corrections from unequal quark masses ($m_{i}$, $i$=$u, d, s$) and
1629: unequal quark chemical potentials ($\mu_{i}$, $i$=$u, d, s$). The corrections
1630: to the quadratic term of the GL potential from $m_i$ and $\mu_i$ indicate
1631: that it is the {\em average} density of states of paired quarks which
1632: determines the critical temperatures of different pairs, as can be seen in
1633: Eqs.\ ($\ref{eq:final-form-3}$) and ($\ref{eq:final-form-4}$).
1634:
1635: We found that the effects of the strange quark mass ($m_s$) through the
1636: mass term in the quark propagator and through the electron charge chemical
1637: potential play a vital role near the super-normal phase boundary, while the
1638: color neutrality plays a minor role in contrast to the case at zero
1639: temperature.
1640:
1641: As shown in Figs.\ $\ref{fig:Tc}$ and $\ref{fig:conden}$, an interplay
1642: between these two effects of $m_s$ splits the single phase transition
1643: (CFL $\rightarrow$ normal) for the $u$-$d$-$s$ symmetric system to three
1644: successive second-order phase transitions
1645: (mCFL $\rightarrow$ dSC $\rightarrow$ 2SC $\rightarrow$ normal).
1646: The window of the "dSC" phase opened up between the 2SC and mCFL phases is
1647: a new phase where only the superconducting gaps with $d$ quarks are
1648: non-vanishing \cite{IMTH04}.
1649:
1650: The obtained phase structure, which is shown to be valid in the weak
1651: coupling (high density) region, may be applicable even in the strong coupling
1652: region as long as the successive transitions driven by the quadratic term of
1653: the GL potential take place in a small interval of temperature near the
1654: super-normal phase boundary. In connection with this, it was shown recently
1655: that the dSC phase discussed in the present paper may be replaced by the uSC
1656: phase in the low density region through a "doubly critical" point on the basis
1657: of the NJL model \cite{Fukushima} (see also Ref.\ \cite{uSC}). Moreover,
1658: if we take into account the chiral condensate at low density, some interplay
1659: between the broken chiral symmetry and the color superconductivity is expected
1660: \cite{Bub}. How to incorporate these aspects in our picture based on the
1661: GL potential is an interesting open problem.
1662:
1663:
1664: In the present paper, the properties of quark and gluon modes
1665: such as the excitation energies were also investigated. The excitation
1666: spectra of quarks indicate that in addition to unpaired quark modes, whose
1667: excitation energy is naturally zero, more than one gapless mode appear just
1668: below each critical temperature as summarized in Table I. Effects of the
1669: gapless modes on physical phenomena such as neutron star cooling is
1670: an interesting problem to be examined \cite{astro}.
1671:
1672:
1673: Throughout this paper, we have studied the phase transitions in the
1674: mean-field level. In weak coupling, thermal fluctuations of the gluon fields
1675: could change the second-order transition to the first-order one as described
1676: in Refs.\ \cite{MIHB,GH}. As far as the order of phase transition is
1677: concerned, the gluon fluctuations have the following effects (the detailed
1678: account will be given in our future publication): The second order transition,
1679: mCFL $\to$ dSC, remains second order. This is because all eight gluons are
1680: Meissner screened across $T=T_c^{\rm I}$ and thus cannot induce a cubic term
1681: in the order parameter in the GL potential. On the other hand, the
1682: transitions, dSC $\to$ 2SC and 2SC $\to$ normal, become weak first order
1683: since some of the gluons become massless at $T=T_c^{\rm I,II}$ (Table I).
1684: In order to obtain a final phase diagram we should study the competition
1685: between the shift of the critical temperature discussed in the present paper
1686: and that coming from the fluctuation effects.
1687:
1688: \begin{table*}[hbt]
1689: \caption{The symmetry, the number of Meissner screened gluons, and the number
1690: of gapless fermion modes in the mCFL, dSC, and 2SC phases. The results for
1691: the Meissner masses are given in Eq.\ (\ref{gluon-masss}). In each phase, if
1692: the temperature is in the region where the gaps satisfy the conditions listed
1693: in this table, more gapless modes appear in addition to unpaired quark modes.
1694: $(\cdots)$ shows the color and flavor of unpaired quarks. $\{ \cdots \}$
1695: shows which combination of quarks makes the gapless modes. }
1696: \begin{tabular}{|c|c|c|c|c|c|c|c|}
1697: \hline
1698: & Symmetry & Number of & \multicolumn{5}{|c|}{Number of gapless quark modes}\\
1699: \cline{4-8}
1700: & & massive gluons & unpaired quarks
1701: &\multicolumn{4}{|c|}{paired quarks} \\
1702: \cline{5-8}
1703: & & & & $\Delta_3 < \mu_e/2$
1704: &$\Delta_1 < \mu_e$ & $\Delta_2 < \mu_e/2$ & no condition\\
1705:
1706: \hline \hline
1707: mCFL &$[U(1)]^2 $ & 8 & 0 & - & - & 2 $\{ Rs,Bu\}$ &-\\
1708: \hline
1709: dSC & $[U(1)]^4 $ & 8 & 2 $(Rs,Bu)$ & - & 2$\{Gs,Bd\}$& - &1$\{Ru,Gd,Bs\}$\\
1710: \hline
1711: 2SC & $[SU(2)]^2 \times [U(1)]^2 $ & 5 &
1712: 5& 2 $\{ Rd, Gu\}$ & -&- &-\\
1713: & & & ($Bu$, $Bd$, $Bs$, $Rs$, $Gs$)
1714: & 2 $\{ Ru, Gd\}$ & & &\\
1715: \hline
1716: \end{tabular}
1717: \end{table*}
1718:
1719: \begin{acknowledgments}
1720: We are grateful to H. Abuki, G. Baym, K. Fukushima, and M. Huang
1721: for helpful discussions. K.I., M.T., and T.H. thank the Institute for
1722: Nuclear Theory, Univ.\ of Washington where a part of this work has been
1723: completed. They also thank K. Rajagopal, I. Shovkovy, T. Sch\"{a}fer, and
1724: C. Kouvaris for discussions. This work was supported in part by RIKEN Special
1725: Postdoctoral Researchers Grant No.\ A12-52010, and by the Grants-in-Aid of the
1726: Japanese Ministry of Education, Culture, Sports, Science, and Technology
1727: (No.~15540254).
1728: \end{acknowledgments}
1729:
1730: %\newpage
1731: %%%%%%%%%% REFERENCES %%%%%%%%%%
1732: \begin{thebibliography}{99}
1733: \bibitem{BL}
1734: D. Bailin and A. Love, Phys.\ Rep.\ {\bf 107}, 325 (1984);
1735: M. Iwasaki and T. Iwado, Phys.\ Lett.\ B {\bf 350}, 163 (1995);
1736: M. Alford, K. Rajagopal, and F. Wilczek, Phys.\ Lett.\ B {\bf 422}, 247 (1998);
1737: R. Rapp, T. Sch\"{a}fer, E.V. Shuryak, and M. Velkovsky,
1738: Phys.\ Rev.\ Lett.\ {\bf 81}, 53 (1998).
1739:
1740: \bibitem{RWA} For reviews, see
1741: K. Rajagopal and F. Wilczek, in {\em Handbook of QCD},
1742: edited by M. Shifman (World Scientific, Singapore, 2001);
1743: M.G. Alford, Ann.\ Rev.\ Nucl.\ Part.\ Sci.\ {\bf 51}, 131 (2001).
1744:
1745: \bibitem{SON99} D. T. Son, Phys.\ Rev.\ D {\bf 59}, 094019 (1999).
1746:
1747: \bibitem{Sch-NPB-575} T. Sch\"{a}fer, Nucl.\ Phys.\ {\bf B575}, 269 (2000).
1748:
1749: \bibitem{Evans-NPB-581}
1750: N.J. Evans, J. Hormuzdiar, S.D.H. Hsu, and M. Schwetz, Nucl.\ Phys.\
1751: {\bf B581}, 391 (2000).
1752:
1753: \bibitem{IB} K. Iida and G. Baym, Phys.\ Rev.\ D {\bf 63}, 074018
1754: (2001); {\em ibid.} {\bf 66}, 059903(E) (2002).
1755:
1756: \bibitem{HH-PRD-68}
1757: D.K. Hong and S.D.H. Hsu, Phys.\ Rev.\ D {\bf 68}, 034011 (2003).
1758:
1759: \bibitem{CFL} M. Alford, K. Rajagopal, and F. Wilczek, Nucl.\ Phys.\
1760: {\bf B537}, 443 (1999).
1761:
1762: \bibitem{SWR}
1763: A. Schmitt, Q. Wang, and D.H. Rischke, Phys.\ Rev.\ D {\bf 66}, 114010 (2002).
1764:
1765: \bibitem{gCFL}
1766: M. Alford, J. Berges and K. Rajagopal, Phys.\ Rev.\ Lett.\ {\bf 84},
1767: 598 (2000); M. Huang and I. Shovkovy, Nucl.\ Phys.\ {\bf A729}, 835 (2003);
1768: M. Alford, C. Kouvaris, and K. Rajagopal, Phys.\ Rev.\ Lett.\
1769: {\bf 92}, 222001 (2004).
1770:
1771: \bibitem{IMTH04}
1772: K. Iida, T. Matsuura, M. Tachibana, and T. Hatsuda, Phys.\ Rev.\ Lett.
1773: {\bf 93}, 132001 (2004).
1774:
1775: \bibitem{IB3} K. Iida and G. Baym, Phys.\ Rev.\ D {\bf 66}, 014015 (2002).
1776:
1777: \bibitem{CJT}
1778: J.M. Luttinger and J.C. Ward, Phys.\ Rev.\ {\bf 118}, 1417 (1960);
1779: G. Baym, Phys.\ Rev.\ {\bf 127}, 1391 (1962);
1780: J.M. Cornwall, R. Jackiw, and E. Tomboulis, Phys.\ Rev.\ D {\bf 10}, 2428
1781: (1974).
1782:
1783:
1784: \bibitem{Georgi} H. Georgi, {\em Weak Interactions and Modern Particle
1785: Theory}, (W.A.~Benjamin, Menlo Park, CA, 1984).
1786:
1787: \bibitem{PR-PRD60}
1788: R.D. Pisarski and D.H. Rischke, Phys.\ Rev.\ D {\bf 60}, 094013, (1999).
1789:
1790: \bibitem{brown}
1791: D.K. Hong, Nucl.\ Phys.\ {\bf B582}, 451 (2000);
1792: W.E. Brown, J.T. Liu, and H.C. Ren, Phys.\ Rev.\ D {\bf 62}, 054016 (2000).
1793:
1794: \bibitem{PR}
1795: R.D. Pisarski and D.H. Rischke, Phys.\ Rev.\ D {\bf 61}, 074017 (2000).
1796:
1797: \bibitem{PIS00} R.D. Pisarski, Phys.\ Rev.\ C {\bf 62}, 035202 (2000).
1798:
1799: \bibitem{unlock}
1800: M.G. Alford, J. Berges, and K. Rajagopal, Nucl.\ Phys.\ {\bf B558}, 219 (1999);
1801: T. Sch\"{a}fer and F. Wilczek, Phys.\ Rev.\ D {\bf 60}, 074014 (1999).
1802:
1803: \bibitem{abuki}
1804: H. Abuki, Prog.\ Theor.\ Phys.\ {\bf 110}, 937 (2003).
1805:
1806:
1807: \bibitem{RIS}
1808: D.D. Dietrich and D.H. Rischke, Prog.\ Part.\ Nucl. \ Phys.\ {\bf 53}, 305
1809: (2004).
1810:
1811:
1812: \bibitem{Sch}
1813: T. Sch\"{a}fer, Phys.\ Rev.\ D {\bf 65}, 094033 (2002).
1814:
1815:
1816: \bibitem{Fukushima}
1817: K. Fukushima, C. Kouvaris, and K. Rajagopal, \protect{hep-ph/0408322}.
1818:
1819:
1820: \bibitem{MIHB}
1821: T. Matsuura, K. Iida, T. Hatsuda, and G. Baym, Phys.\ Rev.\ D {\bf 69},
1822: 074012 (2004).
1823:
1824: \bibitem{gapless-AKR}
1825: M. Alford, C. Kouvaris, and K. Rajagopal, Phys.\ Rev.\ Lett.\ {\bf 92},
1826: 222001 (2004);
1827: M. Alford, C. Kouvaris, and K. Rajagopal, \protect{hep-ph/0406137}.
1828:
1829: \bibitem{gapless-MS}
1830: I. Shovkovy and M. Huang, Phys.\ Lett.\ B {\bf 564}, 205 (2003);
1831: M. Huang, and I. Shovkovy, Nucl.\ Phys.\ {\bf A729}, 835 (2003).
1832:
1833:
1834: \bibitem{uSC}
1835: S.B. Ruster, I.A. Shovkovy, and D.H. Rischke,
1836: Nucl.\ Phys.\ {\bf A743}, 127 (2004).
1837:
1838: \bibitem{Bub}
1839: See, e.g., M. Buballa, \protect{hep-ph/0402234} and
1840: references therein.
1841:
1842: \bibitem{astro}
1843: T. Sch\"{a}fer and K. Schwenzer, \protect{astro-ph/0410395};
1844: M. Alford, P. Jotwani, C. Kouvaris, J. Kundu, and K. Rajagopal
1845: \protect{astro-ph/0411560}.
1846:
1847: \bibitem{GH}
1848: I. Giannakis, D.f. Hou, H.c. Ren, and D.H. Rischke, \protect{hep-ph/0406031}.
1849:
1850: \end{thebibliography}
1851: \end{document}
1852: