nucl-th0203010/P2.tex
1: %% ****** Start of file template.aps ****** %
2: %%
3: %%
4: %%   This file is part of the APS files in the REVTeX 4 distribution.
5: %%   Version 4.0 of REVTeX, August 2001
6: %%
7: %%
8: %%   Copyright (c) 2001 The American Physical Society.
9: %%
10: %%   See the REVTeX 4 README file for restrictions and more information.
11: %%
12: 
13: \documentclass[prc,preprint,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
14: 
15: % You should use BibTeX and apsrev.bst for references
16: % Choosing a journal automatically selects the correct APS
17: % BibTeX style file (bst file), so only uncomment the line
18: % below if necessary.
19: %\bibliographystyle{apsrev}
20: \usepackage[mathscr]{eucal}
21: \usepackage{graphicx}
22: \def\slr#1{\setbox0=\hbox{$#1$}           % set a box for #1
23:    \dimen0=\wd0                                 % and get its size
24:    \setbox1=\hbox{/} \dimen1=\wd1               % get size of /
25:    \ifdim\dimen0>\dimen1                        % #1 is bigger
26:       \rlap{\hbox to \dimen0{\hfil/\hfil}}      % so center / in box
27:       #1                                        % and print #1
28:    \else                                        % / is bigger
29:       \rlap{\hbox to \dimen1{\hfil$#1$\hfil}}   % so center #1
30:       /                                         % and print /
31:    \fi}
32: 
33: \def\kp{k^{\,\prime}}
34: \def\kpsq{k^{\,\prime\,2}}
35: \def\ksq{k^2}
36: \def\kdp{k^{\,\prime\prime}}
37: \def\myint#1{\!\int\!\!\frac{d^4\!{#1}}{(2\pi)^4}\,}
38: \def\gev#1{ GeV${}^{#1}$}
39: \def\be{\begin{eqnarray}}
40: \def\ee{\end{eqnarray}}
41: 
42: 
43: \renewcommand{\theequation}%
44:     {\arabic{section}.\arabic{equation}}
45: \makeatletter \@addtoreset{equation}{section} \makeatother
46: 
47: \begin{document}
48: 
49: % Use the \preprint command to place your local institutional report
50: % number in the upper righthand corner of the title page in preprint mode.
51: % Multiple \preprint commands are allowed.
52: % Use the 'preprintnumbers' class option to override journal defaults
53: % to display numbers if necessary
54: %\preprint{}
55: 
56: %Title of paper
57: \title{Quark Condensates and Momentum-Dependent Quark Masses
58:  in a Nonlocal Nambu--Jona-Lasinio Model}
59: 
60: \author{Bing He}
61: \author{Hu Li}
62: \author{Qing Sun}
63: \author{C.M. Shakin}
64:  \email[email:]{casbc@cunyvm.cuny.edu}
65: \affiliation{%
66: Department of Physics and Center for Nuclear Theory\\
67: Brooklyn College of the City University of New York\\
68: Brooklyn, New York 11210
69: }%
70: 
71: \date{January, 2002}
72: 
73: \begin{abstract}
74: % insert abstract here
75: The Nambu--Jona-Lasinio (NJL) model has been extensively studied
76: by many researchers. In previous work we have generalized the NJL
77: model to include a covariant model of confinement. In the present
78: work we consider further modification of the model so as to
79: reproduce the type of Euclidean-space momentum-dependent quark
80: mass values obtained in lattice simulations of QCD. This may be
81: done by introducing a nonlocal interaction, while preserving the
82: chiral symmetry of the Lagrangian. In other work on nonlocal
83: models, by other researchers, the momentum dependence of the quark
84: self-energy is directly related to the regularization scheme. In
85: contrast, in our work, the regularization is independent of the
86: nonlocality we introduce. It is of interest to note that the value
87: of the condensate ratio,
88: $\langle\bar{s}s\rangle/\langle\bar{u}d\rangle$, is about 1.7 when
89: evaluated using chiral perturbation theory and is only about 1.1
90: in standard applications of the NJL model. We find that our
91: nonlocal model can reproduce the larger value of the condensate
92: ratio when reasonable values are used for the strength of the 't
93: Hooft interaction. (In an earlier study of the $\eta$(547) and
94: $\eta$'(958) mesons, we found that use of the larger value of the
95: condensate ratio led to a very good fit to the mixing angles and
96: decay constants of these mesons.) We also study the density
97: dependence of both the quark condensate and the momentum-dependent
98: quark mass values. Without the addition of new parameters, we
99: reproduce the density dependence of the condensate given by a
100: well-known model-independent expression valid for small baryon
101: density. The generalization of our model to include a model of
102: confinement required the introduction of an additional parameter.
103: The further generalization to obtain a nonlocal model also
104: requires additional parameters. However, we believe our results
105: are of sufficient interest so as to compensate for the
106: introduction of the additional parameters in our formalism.
107: \end{abstract}
108: 
109: % insert suggested PACS numbers in braces on next line
110: \pacs{12.39.Fe, 12.38.Aw, 14.65.Bt}
111: % insert suggested keywords - APS authors don't need to do this
112: %\keywords{}
113: 
114: %\maketitle must follow title, authors, abstract, \pacs, and \keywords
115: \maketitle
116: 
117: % body of paper here - Use proper section commands
118: % References should be done using the \cite, \ref, and \label commands
119: \section{INTRODUCTION}
120: 
121: The Nambu--Jona-Lasinio model has been extensively studied for
122: several decades [1-3]. In recent years there has been strong
123: interest in the study of quark matter at high densities, using the
124: NJL model and related models. In particular, one finds color
125: superconductivity under certain conditions and it has been
126: suggested that some compact stars might be made of superconducting
127: quark matter [4-10]. It is our belief that in such studies one
128: should use a model which reproduces, as well as possible, know
129: features of QCD. In this work we wish to generalize the
130: SU(3)-flavor version of the NJL model to be consistent with the
131: type of momentum-dependent quark masses found in lattice
132: simulations of QCD [11]. For example, in Figs. 1 and 2 we show
133: some of the results obtained in Ref. [11]. It may be seen from
134: these figures that, in Euclidean space, the quark mass goes over
135: to the current mass for Euclidean momentum $k \gtrsim$ 2 GeV. On
136: the other hand, the NJL model, in the standard analysis [1], gives
137: rise to a constant value for the constituent quark mass. As we
138: will see, the form of the nonlocality used here is different from
139: that used in Refs. [6, 12-16]. For example, in reference [12] the
140: $q\bar{q}$ vertex is modified by a form factor which depends on
141: the relative momentum of the quark and antiquark, while, in
142: another scheme, a form factor is associated with each quark line
143: appearing in a diagram. In the latter procedure no further
144: regularization is needed. However, for the problem considered in
145: this work, the nonlocal models that appear in the literature are
146: of limited applicability, since, in the limit of zero current
147: quark mass, the quark self-energy is proportional to the form
148: factors used to define the nonlocality [12]. In contrast, in the
149: current work, we calculate the form of the quark self-energy after
150: introducing a regulator and a nonlocal quark interaction. We
151: stress that the procedure used here differs from any that appears
152: in the literature.
153: 
154: The organization of our work is as follows. In Section II we
155: review the standard analysis for the condensates and ``gap
156: equation" of the SU(3)-flavor NJL model [1]. We then go on to
157: review a procedure for including the contribution to the quark
158: self-energy due to the addition of a confining interaction for the
159: case of the SU(2)-flavor model. In Section III we introduce a
160: nonlocal interaction in the SU(3)-flavor NJL model while
161: maintaining the separable nature of the interaction. We also
162: describe the approximation used for the 't Hooft interaction in
163: the nonlocal model. In Section IV and V we present some of the
164: results of our numerical calculations. In Section VI we discuss
165: the density dependence of the quark condensate and the momentum
166: dependent quark mass. Finally, Section VII contains some
167: additional discussion and conclusions.
168: 
169:  \begin{figure}
170:  \includegraphics[
171:  bb=240 0 250 450,
172:  angle=-0.5
173:  , scale=0.5
174:  ]{fig1.eps}%
175:  \caption{Quark mass values obtained in Ref. [11]
176:  for various current quark masses: $m^0=91$ MeV [
177:  circles], $m^0=54$ MeV [crosses] and $m^0=35$ MeV
178:  [diamonds].}
179:  \end{figure}
180: 
181:  \begin{figure}
182:  \includegraphics[bb=280 20 290 480, scale=0.6]{fig2.eps}%
183:  \caption{Quark mass values obtained in Ref. [11]
184:  using an extrapolation of the current quark mass to zero.
185:  (The small dip at $k\sim 1.6$ GeV is not statistically significant
186:  [11].}
187:  \end{figure}
188: 
189: \section{The quark self-energy in the NJL model}
190: 
191: In order to best introduce the nonlocal model, we will first
192: review the calculation of the quark self-energy in the local
193: SU(3)-flavor NJL model and then proceed to add a confinement
194: interaction, as was done in an earlier study of the quark
195: self-energy [17]. We consider the generalization to a model with
196: nonlocal short-range and 't Hooft interactions in the next
197: section. The Lagrangian of the model is
198: \begin{eqnarray}
199: {\cal L}=&&\bar q(i\slr
200: \partial-m^0)q +\frac{G_S}{2}\sum_{i=0}^8[
201: (\bar q\lambda^iq)^2+(\bar qi\gamma_5 \lambda^iq)^2]\nonumber\\
202: && +\frac{G_D}{2}\{\det[\bar q(1+\gamma_5)q]+\det[\bar
203: q(1-\gamma_5)q]\}\,.
204: \end{eqnarray}
205: Here $m^0$ is the matrix of quark current masses, $m^0 =
206: \mbox{diag\,}(m_u^0, m_d^0, m_s^0)$ and the $\lambda^i
207: (i=1,\cdots,8)$ are the Gell-Mann matrices. Further, $\lambda_0 =
208: \sqrt{2\slash 3}\,\openone$, with \openone being the unit matrix
209: in the flavor space.
210: 
211: The quark propagator is written as
212: \begin{equation}
213: iS(k)=\frac{i}{\slr k-\Sigma(k)+i\epsilon}\,,
214: \end{equation}
215: with
216: \begin{equation}
217: \Sigma(k)=A(k^2)+B(k^2)\slr k\,.
218: \end{equation}
219: We may define
220: \begin{equation}
221: M_u(k^2)=\frac{A_u(k^2)}{1-B_u(k^2)}\,,
222: \end{equation}
223: and
224: \begin{equation}
225: Z_u(k^2)=\frac{1}{1-B_u(k^2)}\,,
226: \end{equation}
227: with similar definitions for $M_d(k^2),M_s(k^2),Z_d(k^2)$ and
228: $Z_s(k^2)$.
229: 
230: In the absence of a confinement model, we have $B(k^2)=0,
231: A_u(k^2)=A_d(k^2)=m_u$ and $A_s(k^2)=m_s$, where $m_u$ and and
232: $m_s$ are constants. (Here, we have take $m_u^0=m_d^0$.) In this
233: case we have [2]
234: \begin{equation}
235: m_u=m_u^0-2G_S\langle\bar uu\rangle-G_D\langle\bar
236: dd\rangle\langle\bar ss\rangle\,,
237: \end{equation}
238: \begin{equation}
239: m_d=m_d^0-2G_S\langle\bar dd\rangle-G_D\langle\bar
240: uu\rangle\langle\bar ss\rangle\,,
241: \end{equation}
242: \begin{equation}
243: m_s=m_s^0-2G_S\langle\bar ss\rangle-G_D\langle\bar
244: uu\rangle\langle\bar dd\rangle\,.
245: \end{equation}
246: These equations are depicted in Fig. 3a, where the last term
247: represents the 't Hooft interaction. The up quark vacuum
248: condensate is given by
249: \begin{eqnarray}
250: \langle\bar uu\rangle&=&-N_ci\myint{k}
251: \texttt{Tr}\frac{C(k^2)}{\slr
252: k-m_u+i\epsilon}\,,\\\nonumber\\
253: &=&-4N_ci\myint{k}\frac{m_uC(k^2)}{k^2-m_u^2+i\epsilon}\,.
254: \end{eqnarray}
255: Here, $C(k^2)$ is a function needed to regulate the integral. In
256: this work we will use the Pauli-Villars procedure and evaluate the
257: integral in Euclidean space, as was done in Ref. [17]. In the
258: general case we may write
259: \begin{equation}
260: \langle\bar uu\rangle=\;-4N_ci\myint{k}
261: \frac{Z_u(k^2)M_u(k^2)C(k^2)}{k^2-M_u^2(k^2)+i\epsilon}\,,
262: \end{equation}
263: where $Z_u(k^2)$ and $M_u^2(k^2)$ were defined in Eqs. (2.4) and
264: (2.5).
265: 
266: In the appendix of Ref. [17] we considered Lorentz-vector
267: confinement, with
268: \begin{equation}
269: \overline
270: {V^c}(k_E-k_E^{\,\prime})=\gamma^\mu(1)\gamma_\mu(2)V^c(k_E-k_E^{\,\prime})
271: \end{equation}
272: and
273: \begin{equation}
274: {V^c}(k_E-k_E^{\,\prime})=\;-8\pi\kappa\left\{\frac{1}{[(k_E-k_E^{\,\prime})^2+\mu^2]^2}
275: -\frac{4\mu^2}{[(k_E-k_E^{\,\prime})^2+\mu^2]^3}\right\}\,,
276: \end{equation}
277: where $k_E^\mu-k_E^{\prime\mu}$ denotes the Euclidean-space
278: momentum transfer. Here, $\mu$ is a small parameter introduced to
279: soften the momentum-space singularities. Note that the form
280: \begin{equation}
281: {V^c}(\vec k-\vec
282: k^{\,\prime})=\;-8\pi\kappa\left\{\frac{1}{[(\vec k-\vec
283: k^{\,\prime})^2+\mu^2]^2} -\frac{4\mu^2}{[(\vec k-\vec
284: k^{\,\prime})^2+\mu^2]^3}\right\}\,,
285: \end{equation}
286: represents the Fourier transform of $V^c(r)=\kappa re^{-\mu r}$,
287: so that for small $\mu, V^c(r)$ approximates a linear potential
288: over the relevant range of \emph r. In Fig. 3b we show the
289: equation for the self-energy when the confining field is included.
290: 
291: In Ref. [17] we obtained the following coupled equations in the
292: case of the SU(2)-flavor model
293: \begin{equation}
294: A(k^2)=\;i\myint{\kp}
295: \frac{[-4V^c(k-k^{\,\prime})+4N_cn_fG_S]A(\kpsq)}{k^{\,\prime
296: \,2}[1-B(k^{\,\prime \,2})]^2-A^2(k^{\,\prime \,2})+i\epsilon}\,,
297: \end{equation}
298: 
299: 
300: \begin{equation}
301: k^2B(k^2)=\;i\myint{\kp}\frac{2(k\cdot
302: k^{\,\prime})[1-B(k^{\,\prime
303: \,2})]V^c(k-k^{\,\prime})}{k^{\,\prime \,2}[1-B(k^{\,\prime
304: \,2})]^2-A^2(k^{\,\prime \,2})+i\epsilon}\,.
305: \end{equation}
306: These equations were solved after passing to Euclidean space and
307: including a Pauli-Villars regulator of the form
308: \begin{equation}
309: C(k_E^{\,2})=
310: \frac{2\Lambda^4}{[k_E^{\,2}+A^2(k_E^{\,2})+\Lambda^2][k_E^{\,2}+A^2(k_E^{\,2})+2\Lambda^2]}
311: \end{equation}
312: in Euclidean space. Note that the form
313: \begin{equation}
314: \widetilde C(k_E^{\,2})=
315: \frac{2\Lambda^4}{[k_E^{\,2}\left(1-B(k_E^{\,2})\right)^2
316: +A^2(k_E^{\,2})+\Lambda^2][k_E^{\,2}\left(1-B(k_E^{\,2})
317: \right)^2+A^2(k_E^{\,2})+2\Lambda^2]}
318: \end{equation}
319: may also be used. (In Eqs. (2.15) and (2.16) $V^c$ appears with a
320: sign opposite to that given in Ref. [17], since, in that work, we
321: used a negative value of $\kappa$. For the present work we use a
322: positive value of $\kappa$ to be consistent with all of our other
323: publications.)
324: 
325:  \begin{figure}
326:  \includegraphics[bb=180 25 190 250, angle=-0.2, scale=0.8]{fig3.eps}%
327:  \caption{a)A diagrammatic representation of the equation
328:  for the self-energy for a quark of momentum $k$. The first
329:  term on the right is the contribution of the current quark mass $m^0$.
330:  The second term corresponds to the term proportional to $G_S$ in Eqs. (3.6)--(3.8)
331:  and the last term represents the 't Hooft interaction.
332:  b)The self-energy equation in the presence of a confining interaction (wavy line.)
333:  Without the 't Hooft interaction, we have a representation of SU(2)-flavor
334:  model studied in Ref. [17]. (See Eqs. (2.15) and (2.16).)}
335:  \end{figure}
336: 
337: \section{A nonlocal NJL model}
338: 
339: In this section we describe the procedures we use to create a
340: nonlocal version of our generalized NJL model. We consider the
341: second term of Eq. (2.1) and make the replacement
342: \begin{eqnarray}
343: \frac{G_S}{2}&&\sum_{i=0}^8[(\bar q(x)\lambda_iq(x))^2+(\bar
344: q(x)i\gamma_5\lambda_iq(x))^2]\\\nonumber
345: &&\longrightarrow\frac{G_S}{2}\sum_{i=0}^8\{[\bar
346: q(x)\lambda^if(x)q(x)\cdot \bar q(y)\lambda^if(y)q(y)\\[0.4cm]\nonumber
347: &&\hspace{0.9cm}+[\bar q(x)i\gamma_5\lambda^if(x)q(x)\cdot\bar
348: q(y)i\gamma_5\lambda^if(y)q(y)]\}\,.
349: \end{eqnarray}
350: This replacement corresponds to the use of a separable interaction
351: $V(x-y)=G_Sf(x)f(y)$.
352: 
353: A related modification may be made for the 't Hooft interaction.
354: It is useful, however, to describe these modifications as they
355: affect momentum-space calculations. With reference to Fig. 4, we
356: replace $G_S$ by $f(k_1-k_2)G_Sf(k_3-k_4).$ In the evaluation of
357: the second term of Fig. 2 we need
358: $G_S(k-k^{\,\prime})=f(k-k^{\,\prime})G_Sf(k-k^{\,\prime})$ and
359: choose to write
360: \begin{eqnarray}
361: G_S(k-k^{\,\prime})&=&\:\exp[-(k-k^{\,\prime})^{\,2n}/2\beta]
362: G_S\exp[-(k-k^{\,\prime})^{\,2n}/2\beta]\\\nonumber
363: &=&\:G_S\exp[-(k-k^{\,\prime})^{\,2n}/\beta]\,.
364: \end{eqnarray}
365: In this work we take $n=4$ and $\beta=20$ GeV${}^{8}$. In Fig. 5
366: we exhibit the function $F(k^2)=\exp[-k^{\,2n}/\beta]$. It is
367: clear that many other functions may be chosen.
368: 
369: We now rewrite Eqs. (2.15) and (2.16) for the up, down and strange
370: quarks. For example, for the SU(3)-flavor case
371: \begin{eqnarray}
372: A_u(k^2)&=&m_u^0+i\myint{\kp}
373: \frac{[-4V^c(k-k^{\,\prime})+8N_cG_S(k-k^{\,\prime})]A_u(k^{\,\prime\,2})}
374: {k^{\,\prime\,2}[1-B_u(k^{\,\prime\,2})]^2-A_u^2(k^{\,\prime\,2})+i\epsilon}\,,
375: \\[0.35cm]
376: k^2B_u(k^2)&=&i\myint{\kp}
377: \frac{2(k\cdot\kp)[1-B_u(\kpsq)]V^c(k-\kp)}
378: {k^{\,\prime\,2}[1-B_u(k^{\,\prime\,2})]^2-A_u^2(k^{\,\prime\,2})+i\epsilon}\,,
379: \end{eqnarray}
380: with similar equations for $A_d(k^2), B_d(k^2),$ etc. Again, these
381: equations are solved after passing to Euclidean space and
382: introducing regulator functions: $C_u(\ksq),\, C_d(\ksq) \texttt{
383: and }C_s(\ksq)$. In Euclidean space we write \be
384: C_u(\ksq)=\frac{2\Lambda^4}{[\ksq+A_u^2(\ksq)+\Lambda^2][\ksq+A_u^2(\ksq)+2\Lambda^2]}\,,
385: \ee etc. Note that without the 't Hooft interaction the equations
386: for the up, down and strange quarks are uncoupled.
387: 
388:  \begin{figure}
389:  \includegraphics[bb=280 75 290 250, scale=0.6]{fig4.eps}%
390:  \caption{The figure indicates the replacement of the local quark interaction, $iG_S$,
391:  by the nonlocal (separable) term, $iG_S(k_1-k_2, k_3-k_4)=iG_Sf(k_1-k_2)f(k_3-k_4)$.
392:  The distinction between our separable model and that of Ref. [12], for example, is that
393:  in Ref. [12] the alternative replacement $iG_S\longrightarrow iG_S(k_2+k_4, k_1+k_3)
394:  =iG_Sf(k_2+k_4)f(k_1+k_3)$ was used.}
395:  \end{figure}
396: 
397:  \begin{figure}
398:  \includegraphics{fig5.eps}%
399:  \caption{The correlation function $F(k)=\exp[-k^{\,2n}/\beta]$ is
400:  shown for $n=4$ and $\beta=20$\gev{8}.}
401:  \end{figure}
402: 
403: Our treatment of the 't Hooft interaction is based upon a
404: generalization of the last term in Eqs. (2.6)-(2.8). With
405: reference to the third term on the right in Fig. 6, we introduce a
406: correlation between the quark of momentum \emph{k} and the quark
407: of momentum $\kp$. We also include a correlation between the quark
408: of momentum \emph{k} and that of momentum $\kp{}^\prime$.
409: (Therefore, our procedure does not introduce a correlation between
410: the two quarks in the separate condensates. At this stage of the
411: development of our model that seems to be a reasonable
412: approximation and avoids having to define a three-quark
413: correlation function, $f(k-\kp, \kp-\kdp, k-\kdp).$) For example,
414: we generalize the term $-G_D\langle\bar dd\rangle\langle\bar
415: ss\rangle$ to obtain the contribution to $A_u(k)$:\be A_u^t(k)=
416: -G_D\left[-4N_ci\myint{\kp}\frac{C_d(\kpsq)Z_u(\kpsq)M_u(\kpsq)f^2(k-\kp)}
417: {\kpsq[1-B_u(\kpsq)]^2-A_u^2(\kpsq)}\right]
418: \\[0.35cm]\nonumber \hspace{0.6cm}\times\left[-4N_ci\myint{\kdp}
419: \frac{C_s(\kdp{}^2)Z_s(\kdp{}^2)M_s(\kdp{}^2)f^2(k-\kdp)}
420: {\kdp{}^2[1-B_s(\kdp{}^2)]^2-A_s^2(\kdp{}^2)}\right] \ee in
421: Minkowski space. This expression is then evaluated in Euclidean
422: space and added to the right-hand side of Eq. (3.3). In a similar
423: fashion, we calculate $A_d^t(k)$ and $A_s^t(k)$.
424: 
425:  \begin{figure}
426:  \begin{center}
427:  \includegraphics[bb=200 50 210 250, angle=-2, scale=0.8]{fig6.eps}%
428:  \caption{The quark self-energy equation is depicted, with nonlocal terms replacing
429:  $G_S$ and $G_D$. [See Fig. 4.]}
430:  \end{center}
431:  \end{figure}
432: 
433: \section{Numerical Results: Condensates and Constituent Mass Values}
434: 
435: There is a good deal of flexibility in choosing the regulators
436: $C_u(\ksq)$, $C_d(\ksq)$, and $C_s(\ksq)$. Also, various forms
437: could be chosen for the correlation functions, $f(k-\kp)$.
438: Previously, in our Minkowski-space studies of the $\eta$ mesons we
439: used $G_S = 11.84$ GeV${}^{-2}$ and $G_D\simeq -200$\gev{-5} [18].
440: However, in that work we used a Gaussian regulator in Minkowski
441: space so that a direct comparison with the present study can not
442: be made. On the other hand, we do not expect to find radically
443: different parameters, if the constituent masses in the two
444: calculations are similar. For example, in our earlier work, in
445: which $m_u$ and $m_s$ were parameters, we used $m_u=0.364$\gev{},
446: which can be compared to the value of $M_u(0)$ calculated here. We
447: have also used either $m_s=0.565$\gev{} [19-23] or
448: $m_s=0.585$\gev{} [18], values which may be compared to $M_s(0)$.
449: 
450: To proceed, we take $\Lambda=1.0$\gev{}, $G_S=13.30$\gev{-2},
451: $\kappa=0.055$\gev{2}, $m_u^0=0.0055$\gev{}, $m_s^0=0.130$\gev{},
452: $\mu=0.010$\gev{} and $\beta=20.0$\gev{8}. We then consider
453: values of $G_D=0$, $G_D=-20G_S$, $G_D=-30G_S$ and $G_D=-40G_S$.
454: The results of our calculations are given in Table I. Recall that
455: the function $F(k)$ does not appear in our expression for the
456: condensates. The calculation of the condensates includes the
457: Pauli-Villars regulators, $C_u(\ksq)$, $C_d(\ksq)$ and
458: $C_s(\ksq)$, however. [See Eqs. (2.9)-(2.11).] In our calculation
459: of the properties of the $\eta$ mesons [18] we had $-G_D/G_S\simeq
460: 15-18$, since we used $G_D$ values in the range
461: $-180$\gev{-5}$\leq G_S\leq -220$\gev{-5} in that work.
462: 
463: In order to specify a value of $G_D$ for this work, we note that a
464: calculation based upon chiral perturbation theory yields
465: $\langle\bar ss\rangle/\langle\bar uu\rangle = 1.689$ [24].
466: Inspection of Table I suggests that the values of $G_D$, other
467: than $G_D=0$, given in Table I are acceptable. For
468: $G_D=-266$\gev{-5} we have $M_u(0)= 0.377$\gev{} and $M_s(0)=
469: 0.555$\gev{}, which are reasonably close to the phenomenological
470: parameters $m_u= 0.364$\gev{} and $m_s= 0.565$\gev{} used in our
471: earlier work [19-23].
472: 
473: It is worth noting that, in standard application of the
474: SU(3)-flavor NJL model, one finds $\langle\bar
475: ss\rangle/\langle\bar uu\rangle \sim 1.1$ [1], so that the results
476: shown in Table I are encouraging, given that the value for
477: $\langle\bar ss\rangle/\langle\bar uu\rangle$ obtained using
478: chiral perturbation theory is about 1.7 [24], as noted above.
479: 
480: \begin{table}%[H] add [H] placement to break table across pages
481: % \begin{ruledtabular}
482:  \begin{tabular}{||@{\hspace{0.2cm}}l@{\hspace{0.2cm}}||@{\hspace{0.5cm}}
483:  l@{\hspace{0.5cm}}|@{\hspace{0.5cm}}l@{\hspace{0.5cm}}
484:  |@{\hspace{0.5cm}}l@{\hspace{0.5cm}}|@{\hspace{0.5cm}}l@{\hspace{0.2cm}}||}\hline\hline
485:  $G_D$ [\gev{-5}]    &0.0     &-266.0 &-399.0 &-532.0\\\hline
486:  $M_u(0)$ [GeV]      &0.334   &0.377  &0.396  &0.416\\\hline
487:  $M_s(0)$ [GeV]      &0.538   &0.555  &0.564  &0.575\\\hline
488:  $\langle\bar uu\rangle^\frac{1}{3}$ [GeV]
489:                       &-0.207  &-0.215 &-0.217 &-0.220\\\hline
490:  $\langle\bar ss\rangle^\frac{1}{3}$ [GeV]
491:                       &-0.2605 &-0.261 &-0.261 &-0.261\\\hline
492:  $\frac{\langle\bar ss\rangle}{\langle\bar uu\rangle}$
493:                       &2.00    &1.80   &1.73   &1.68  \\\hline
494:  $A_u(0)$ [GeV]      &0.447   &0.481  &0.496  &0.512 \\\hline
495:  $B_u(0)$            &-0.335  &-0.276 &-0.253 &-0.233\\\hline
496:  $A_s(0)$ [GeV]      &0.614   &0.628  &0.636  &0.645  \\\hline
497:  $B_s(0)$            &-0.139  &-0.131 &-0.127 &-0.122\\\hline\hline
498: % Lines of table here ending with \\
499:  \end{tabular}
500:  \vspace{1.2cm}
501:  \caption{Calculated values for the condensates and for $A(0),
502: B(0), \mbox{and } M(0)$ are given for the up and strange quarks
503: for four values of $G_D$. The parameters $m_u^0=0.0055$ GeV,
504: $m_s^0=0.130$ GeV, $\kappa=0.055$\gev2, $\beta=20$\gev8,
505: $\mu=0.010$ GeV, $\Lambda=1.0$ GeV, $G_S=13.30$\gev{-2} were used.
506: The values of $\kappa$ and $\mu$ were fixed in earlier work
507: [18-23]. Values of $\langle\bar uu\rangle\simeq\langle\bar
508: dd\rangle\simeq -(0.240\pm 0.025\mbox{ GeV}){}^3$ have been
509: suggested [28], so we see that our calculated values are at, or
510: near, the lower limit for that quantity.}
511: % \end{ruledtabular}
512:  \end{table}
513: 
514: \section{numerical results:\\momentum dependence of the constituent quark masses}
515: 
516: In Fig. 7 we show $M_u(k)$, where \emph k is the magnitude of the
517: Euclidean momentum. The dashed line exhibits the result without
518: the confining interaction ($\kappa=0$). It is interesting to see
519: that inclusion of confinement improves the shape of the curve when
520: we compare our results to the lattice results shown in Figs. 1 and
521: 2. We note that $M_u(k)$ goes over to $m_u^0=0.0055$\gev{} for
522: large \emph k. In Fig. 8 we show $B_u(k)$. (Recall that
523: $Z_u(k)=[1-B_u(k)]^{-1}$.) We remark that $B_u(k)=0$ when
524: $\kappa=0$. In Figs. 9 and 10 we show $M_s(k)$ and $B_s(k)$,
525: respectively. As expected, we find that $M_s(k)$ goes over to
526: $m_s^0=0.130$\gev{} when \emph k is large.
527: 
528:  \begin{figure}
529:  \includegraphics{fig7.eps}%
530:  \caption{Values of $M_u(k)$ are shown for the parameters $m_u^0=0.0055$ GeV,
531:  $m_s^0=0.130$ GeV, $\kappa=0.055$\gev{2}, $\mu=0.010$ GeV, $\Lambda=1.0$ GeV,
532:  $\beta=20$\gev{8}, $G_S=13.3$\gev{-2}, and $G_D=-266$\gev{-5}. The dashed line
533:  shows the result without confinement ($\kappa=0$).}
534:  \end{figure}
535: 
536:  \begin{figure}
537:  \includegraphics{fig8.eps}%
538:  \caption{Values of $B_u(k)$ are shown. (See caption to Fig. 7.)}
539:  \end{figure}
540: 
541:  \begin{figure}
542:  \includegraphics{fig9.eps}%
543:  \caption{Values of $M_s(k)$ are shown. (See caption to Fig. 7.)}
544:  \end{figure}
545: 
546:  \begin{figure}
547:  \includegraphics{fig10.eps}%
548:  \caption{Values of $B_s(k)$ are shown. (See caption to Fig. 7.)}
549:  \end{figure}
550: 
551: \section{density dependence of the quark mass and quark
552: condensate}
553: 
554: As stated earlier, the behavior of the NJL model for finite values
555: of the baryon density is an extensively explored topic [1, 25,
556: 26], with particular recent emphasis on color superconductivity
557: [4-10]. In this section we explore the behavior of our model at
558: finite baryon density. (It should be noted that the nature of the
559: phase transition describing chiral symmetry restoration at finite
560: density is quite model dependent. For example, the inclusion of
561: current quark masses can change a strong first-order transition to
562: a smooth second-order transition [26].)
563: 
564: A comprehensive study of the thermodynamics of the three-flavor
565: NJL model has been reported in Ref. [27]. There it is found that
566: the up, down and strange quark masses are essentially constant up
567: to the density where a first-order phase transition appears. At
568: that point, the up and down quark masses drop from a value of
569: about 380 MeV to about 30 MeV. That behavior differs from the
570: behavior expected at low density. For example, we have the
571: well-known relation between the value of the condensate and the
572: baryon density of nuclear matter \be \frac{\langle\bar
573: qq\rangle_\rho}{\langle\bar
574: qq\rangle{}_0}=1-\frac{\sigma_N\rho_B}{f_\pi^2m_\pi^2}\,, \ee
575: where $\sigma_N$ is the pion-nucleon sigma term. This relation is
576: valid to first-order in the density. It may be derived, in the
577: case of nuclear matter, by writing \be \langle\bar qq\rangle_\rho
578: = \langle\bar qq\rangle_0+\langle N|\bar qq|N\rangle \rho_B \ee
579: and making use of the definition of the pion-nucleon sigma term,
580: $\sigma_N$, and the Gell-Mann--Oakes--Renner relation. If we put
581: $\sigma_N=0.045$ GeV, we have $\langle\bar
582: qq\rangle_\rho/\langle\bar qq\rangle_0=1-0.273\rho_B$, where
583: $\rho_B$ is in\gev{3} units. For nuclear matter $\rho_B=$(\,0.109
584: GeV)${}^3$, so we see that the condensate is reduced by about
585: 35\%, if we evaluate Eq. (6.1) at nuclear matter density. We can
586: check whether the density dependence given by Eq. (6.1) is
587: reproduced in our model, since it should not matter whether the
588: scalar density of the background matter is generated by quarks in
589: nucleons or by the presence of free quarks. In the former case, we
590: may write, for the baryon density, \be
591: \rho_B=4\!\int^{\,k_F}\hspace{-0.3cm}\frac{d^3\!k}{(2\pi)^3} \ee
592: where the factor of 4 is arises from the product of the spin and
593: isospin factors. In the case of quarks, we have \be
594: \rho_B=4N_c\left(\frac{1}{3}\right)
595: \!\int^{\,k_F}\hspace{-0.3cm}\frac{d^3\!k}{(2\pi)^3} \ee where, in
596: this case, the factor of 4 again arises from the spin and isospin
597: factor. (Both up and down quarks are present in equal numbers.)
598: The color factor, $N_c=3$, is cancelled by the baryon number of
599: 1/3 of each quark.
600: 
601: We need to modify the equations for the quark self-energy to take
602: into account the presence of the Fermi seas of up and down quarks
603: whose Fermi momentum is $k_F$. We take one Fermi sea to be
604: composed of on-mass-shell up quarks with constituent mass
605: $M_u(0)$. The following term is then added to the equation for
606: $A_u(k)$. \be A_u^{(\rho)}(k)=-(2G_S)N_c2\!\int^{\,k_F}
607: \hspace{-0.3cm}\frac{d^3\!\kp}{(2\pi)^3}\,\frac{M_u(0)}{E_u(k)}f^2(k-\kp)
608: \ee where $E_u(k)=\left[\vec k^2+M_u^2(0)\right]^\frac{1}{2}$. The
609: second factor of 2 in Eq. (6.5) reflects the spin degeneracy. We
610: note that $M_u(0)$ is density-dependent and could be written as
611: $M_u(0,\rho)$ in keeping with the labelling of Fig. 11, where
612: $M_u(k,\rho)$ is used. Also, if $f(k-\kp)=1$, $A_u^{(\rho)}(k)$
613: would then represent $-2G_S\,\rho_S$, where $\rho_S$ is the scalar
614: density associated with the up-quark Fermi sea.
615: 
616: In Fig. 11 we show $M(k,\rho)$ calculated for four values of
617: $\rho$ and in Fig. 12 we show \emph M(0) as a function of $k_F^3$.
618: Since the quarks in the Fermi sea are taken to be on-mass-shell,
619: we would, in principle, require $M_u(k,\rho)$ in Minkowski space.
620: However, since $k_F$ = 0.268 GeV for the case of nuclear matter,
621: only a very modest extrapolation of the curves shown in Fig. 11 is
622: needed for the densities considered in this work. In Fig. 13 we
623: show the value of the up quark condensate as a function of
624: $k_F^3$. (Note that $k_F^3 = 19.2\times 10^{-3}$\gev{3} represents
625: the density of nuclear matter.) It is seen that, for small values
626: of the density, the density dependence of the condensate
627: reproduces what is expected from Eq. (6.1). (If we extrapolate the
628: curve using a linear approximation, the condensate is reduced by
629: about 30\% at nuclear matter density.)
630: 
631:  \begin{figure}
632:  \includegraphics{fig11.eps}%
633:  \caption{The values of $M(k,\rho)$ are shown for $\rho/\rho_{NM}=0.26$ [dotted line],
634:  $\rho/\rho_{NM}=0.52$ [dashed line], and $\rho/\rho_{NM}=0.78$ [dash-dot line].}
635:  \end{figure}
636: 
637:  \begin{figure}
638:  \includegraphics{fig12.eps}%
639:  \caption{Values of $M(0)$ are shown as a function of $k_F^3$. Note that $\rho_B
640:  =(2/3\pi^2)k_F^3.$}
641:  \end{figure}
642: 
643:  \begin{figure}
644:  \includegraphics{fig13.eps}%
645:  \caption{The values of the up quark condensate are
646:  given as a function of $k_F^3$. Note that $\rho_B
647:  =(2/3\pi^2)k_F^3.$}
648:  \end{figure}
649: 
650: \section{discussion}
651: 
652: We have remarked earlier in this work that the large values of the
653: condensate ratio $\langle\bar ss\rangle/\langle\bar uu\rangle$
654: seen in Table I play a role in obtaining a good fit to the mixing
655: angles of the $\eta$(947) and $\eta\prime$(958) mesons [18]. To
656: understand this remark we note that the effective singlet-octet
657: coupling constants for pseudoscalar states are [3] \be
658: G_{00}^P&=&G_S-\frac{2}{3}(\alpha+\beta+\gamma)\frac{G_D}{2}\,,\\
659: G_{88}^P&=&G_S-\frac{1}{3}(\gamma-2\alpha-2\beta)\frac{G_D}{2}\,,
660: \ee and \be
661: \hspace{-0.5cm}G_{08}^P=-\frac{\sqrt2}{6}(2\gamma-\alpha-\beta)\frac{G_D}{2}\,,
662: \ee where $\alpha=\langle\bar uu\rangle$, $\beta=\langle\bar
663: dd\rangle$ and $\gamma=\langle\bar ss\rangle$. We take
664: $\alpha=\beta$, so that \be
665: G_{08}^P=-\frac{\sqrt2}{3}(\gamma-\alpha)\frac{G_D}{2}\,.\ee If
666: $\gamma=1.7\alpha$, the result for $G_{08}^P$ is six times larger
667: than when $\gamma=1.1\alpha$.
668: 
669: In addition to the effects of $G_{08}^P$, singlet-octet mixing is
670: induced by the quantity [18] \be
671: E_{08}(k)=\frac{2\sqrt2}{3}[E_u(k)-E_s(k)]\,,\ee where
672: $E_u(k)=\left[\vec k^2+m_u^2\right]^\frac{1}{2}$, etc. It is found
673: that, since $G_{08}^P$ and $E_{08}(k)$ tend to cancel in our
674: formalism, the significant singlet-octet mixing generated by
675: $E_{08}(k)$ is reduced by the values of $G_{08}^P$ obtained for
676: the larger value of the ratio $\langle\bar ss\rangle/\langle\bar
677: uu\rangle$, with the result that we reproduce the values of the
678: mixing angles found in other studies that make use of experimental
679: data to obtain values for the mixing angles [18].
680: 
681: In our earlier work, which was carried out in Minkowski space, the
682: values of $m_u=m_d$ and $m_s$ were taken as parameters. Inspection
683: of our figures which exhibit values of $M_u(k)$ and $M_s(k)$
684: suggests that an extrapolation into Minkowski space may be made if
685: $\ksq$ is not too large. The fact that $M_u(0)$ and $M_s(0)$ are
686: close to our phenomenological parameters for $G_D=-266$\gev{-5} is
687: encouraging and suggests that some support for our choice of quark
688: mass parameters may be found in our Euclidean-space analysis.
689: 
690: The full consequences of separating the specification of the
691: nonlocality of the quark interaction from the choice of the
692: regulator of the theory should be explored more fully. Although
693: that feature of our model introduces greater flexibility, that
694: comes with the disadvantage of having to introduce other
695: parameters in the model. We have made only limited variation of
696: the form of the nonlocality and the regulator. For further
697: applications it may be of interest to explore a more comprehensive
698: parameter variation. It is also necessary to extend the
699: calculations reported in Figs. 11--13 to larger values of the
700: density than those considered here. That step will require more
701: complex methods for solving our nonlinear equations for the
702: self-energy than the simple iteration scheme we have used thus
703: far.
704: 
705: Our work may be compared to that of Alkofer, Watson and Weigel
706: [29] who have solved the Schwinger-Dyson equation using a gluon
707: propagator whose low-momentum behavior is enhanced by a Gaussian
708: function. (That modification requires the introduction of two
709: phenomenological parameters [30].) The behavior found for $A(k)$
710: and $B(k)$ in Euclidean space is similar to that obtained in this
711: work. (See Fig. 1 of Ref. [29].) Those authors also solve the
712: Bethe-Salpeter equation to obtain the properties of various $q\bar
713: q$ mesons with generally satisfactory results. It is of interest
714: to note that the Minkowski space solution for $A(k)$ and $B(k)$ is
715: such that the quark can go on-mass-shell. That feature may be
716: related to our work [18-23] in which we use on-mass-shell quarks
717: with masses $m_u=m_d=0.364$ GeV and $m_s=0.565$ GeV (or 0.585 GeV
718: [18]) when solving the Bethe-Salpeter equation in our study of
719: $q\bar q$ mesons.
720: 
721: % If in two-column mode, this environment will change to single-column
722: % format so that long equations can be displayed. Use
723: % sparingly.
724: %\begin{widetext}
725: % put long equation here
726: %\end{widetext}
727: 
728: %\newpage
729: \vspace{1.5cm}
730: \noindent$\textbf{References}$\\[-2cm]
731: \begin{thebibliography}{99}
732:     \bibitem{spk}S. P. Klevansky, Rev. Mod. Phys. $\textbf{64}$\,,
733:     649 (1992).
734:     \bibitem{uvw}U. Vogl and W. Weise, Prog. Part. Nucl. Phys.
735:     $\textbf{27}$\,, 195 (1991).
736:     \bibitem{tt}T. Hatsuda and T. Kunihiro, Phys. Rep.
737:     $\textbf{247}$\,, 221 (1994).
738:     \bibitem{kf}For reviews, see K. Rajagopal and F. Wilcek, in B.
739:     L. Ioffe Festscrift, At the Frontier of Particle
740:     Physics/Handbook of QCD, M.Shifman ed. (World Scientific,
741:     Singapore, 2001);\\
742:     M. Alford, hep-ph/0102047.
743:     \bibitem{mjk}M. Alford, J. Berges and K. Rajagopal, Nucl.
744:     Phys. B $\textbf{558}$\,, 219 (1999);\\
745:     J. Kundu and K. Rajagopal, hep-ph/0112206 (2002).
746:     \bibitem{cd}C. Gocke, D. Blaschke, A. Khalatyan and H.
747:     Grigoria, hep-ph/0104183-v2 (2002).
748:     \bibitem{ia}I. A. Shovkovy, hep-ph/0110352 (2002).
749:     \bibitem{mr}M. Alford, R. Rajagopal and F. Wilcek, Phys. Lett.
750:     B $\textbf{422}$\,, 247 (1998).
751:     \bibitem{dt}D. T. Son, Phys. Rev. D $\textbf{59}$\,, 094019
752:     (1999).
753:     \bibitem{ts}T. Sh$\ddot a$fer, E. V. Shuryak and M. Velkovsky,
754:     Phys. Rev. Lett. $\textbf{81}$\,, 53 (1998).
755:     \bibitem{js}J. Skullerud, D. B. Leinweber, and A. G. Williams,
756:     Phys. Rev. D $\textbf{64}$\,, 074508 (2001).
757:     \bibitem{hi}H. Ito, W. W. Buck and F. Gross, Phys. Rev. C
758:     $\textbf{43}$\,, 2483 (1991); C $\textbf{45}$\,, 1918
759:     (1992).
760:     \bibitem{ss}S. Schmidt, D. Blaschke, Y. L. Kalinovsky, Phys.
761:     Rev. C $\textbf{50}$\,, 435 (1994).
762:     \bibitem{rs}R. S. Plant and M. C. Birse, Nucl. Phys. A
763:     $\textbf{628}$\,, 607 (1998).
764:     \bibitem{rd}R. D. Bowler and M. C. Birse, Nucl. Phys. A
765:     $\textbf{582}$\,, 655 (1995).
766:     \bibitem{rm}R. S. Plant and M. C. Birse, hep-ph/0007340 and
767:     references therein.
768:     \bibitem{ls}L. S. Celenza, Xiang-Dong Li, and C. M. Shakin,
769:     Phys. Rev. C $\textbf{55}$\,, 1492 (1997).\\
770:     In this reference a negative value of $\kappa$ was used.
771:     Therefore, the sign on the right-hand sides of Eqs. (2.4) and
772:     (2.5), appearing in this reference, should be corrected to be
773:     positive.
774:     \bibitem{cm}C. M. Shakin and Huangsheng Wang, Role of the 't
775:     Hooft interaction in the calculation of the mixing angles of
776:     the $\eta$(547) and $\eta^\prime$(958) mesons, Brooklyn
777:     College\\
778:     Report No. BCCNT: 01/082/307 (2001).
779:     To be published in Physical Rev. C [DH8173].
780:     \bibitem{cms}C. M. Shakin and Huangsheng Wang, Phys. Rev. D
781:     $\textbf{63}$\,, 014019 (2000).
782:     \bibitem{lsc}L. S. Celenza, Huangsheng Wang, and C. M. Shakin,
783:     Phys. Rev. C $\textbf{63}$\,, 025209 (2001).
784:     \bibitem{cmsh}C. M. Shakin and Huangsheng Wang, Phys. Rev. D
785:     $\textbf{63}$\,, 074017 (2001).
786:     \bibitem{cms4}C. M. Shakin and Huangsheng Wang, Phys. Rev. D
787:     $\textbf{63}$\,, 114007 (2001).
788:     \bibitem{cms5}C. M. Shakin and Huangsheng Wang, Phys. Rev. D
789:     $\textbf{64}$\,, 094020 (2001).
790:     \bibitem{ga}G. Amoros, J. Bijnens, and P. Talavera, Nucl.
791:     Phys. B $\textbf{602}$\,, 87 (2001).
792:     \bibitem{ma}M. Asakawa and K. Yazaki, Nucl. Phys. A
793:     $\textbf{504}$\,, 668 (1989).
794:     \bibitem{vb}V. Bernard, Ulf-G. Meissner, and I. Zahed, Phys.
795:     Rev. D $\textbf{36}$\,, 819 (1987).
796:     \bibitem{fg}F. Gastineau, R. Nebauer, and J. Aichelin,
797:     hep-ph/0101289 (2001).
798:     \bibitem{msa}M. Shifman, A. Vainstein and V. Zakarov, Nucl.
799:     Phys. B $\textbf{147}$\,, 385 (1979); 448 (1979);\\
800:     L. Reinders, H. Rubinstein and Y. Yazaki, Phys. Rep.
801:     $\textbf{127}$\,, 2 (1985).
802:     \bibitem{ra}R. Alkofer, P. Watson and H. Weigel,
803:     hep-ph/0202053 (2002).
804:     \bibitem{pm}P. Maris and P. C. Tandy, Phys. Rev. C
805:     $\textbf{60}$\,, 055214 (1999).
806: 
807: 
808: \end{thebibliography}
809: % Create the reference section using BibTeX:
810: %\bibliography{basename of .bib file}
811: 
812: \end{document}
813: %
814: % ****** End of file template.aps ******
815: