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\mytint#1{\!\int\!\!\frac{d^3\!{#1}}{(2\pi)^3}\,}
39: \def\gev#1{ GeV${}^{#1}$}
40: \def\be{\begin{eqnarray}}
41: \def\ee{\end{eqnarray}}
42:
43:
44: \renewcommand{\theequation}%
45: {\arabic{section}.\arabic{equation}}
46: \makeatletter \@addtoreset{equation}{section} \makeatother
47:
48: \begin{document}
49:
50: % Use the \preprint command to place your local institutional report
51: % number in the upper righthand corner of the title page in preprint mode.
52: % Multiple \preprint commands are allowed.
53: % Use the 'preprintnumbers' class option to override journal defaults
54: % to display numbers if necessary
55: %\preprint{}
56:
57: %Title of paper
58: \title{Quark and Nucleon Self-Energy in Dense Matter}
59:
60: \author{L.S.Celenza}
61: \author{Hu Li}
62: \author{C.M. Shakin}
63: \email[email:]{casbc@cunyvm.cuny.edu}
64: \author{Qing Sun}
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{April, 2002}
72:
73: \begin{abstract}
74: In a recent work we introduced a nonlocal version of the
75: Nambu--Jona-Lasinio(NJL) model that was designed to generate a
76: quark self-energy in Euclidean space that was similar to that
77: obtained in lattice simulations of QCD. In the present work we
78: carry out related calculations in Minkowski space, so that we can
79: study the effects of the significant vector and axial-vector
80: interactions that appear in extended NJL models and which play an
81: important role in the study of the $\rho$, $\omega$ and $a_1$
82: mesons. We study the modification of the quark self-energy in the
83: presence of matter and find that our model reproduces the behavior
84: of the quark condensate predicted by the model-independent
85: relation $\langle\bar qq\rangle_{\rho} = \langle\bar
86: qq\rangle_0(1-\sigma_N\rho_N/f_{\pi}^2m_{\pi}^2+\cdots)$, where
87: $\sigma_N$ is the pion-nucleon sigma term and $\rho_N$ is the
88: density of nuclear matter. (Since we do not include a model of
89: confinement, our study is restricted to the analysis of quark
90: matter. We provide some discussion of the modification of the
91: above formula for quark matter.) The inclusion of a quark current
92: mass leads to a second-order phase transition for the restoration
93: of chiral symmetry. That restoration is about 80\% at twice
94: nuclear matter density for the model considered in this work. We
95: also find that the part of the quark self-energy that is
96: explicitly dependent upon density has a strong negative
97: Lorentz-scalar term and a strong positive Lorentz-vector term,
98: which is analogous to the self-energy found for the nucleon in
99: nuclear matter when one makes use of the Dirac equation for the
100: nucleon. In this work we calculate the nucleon self-energy in
101: nuclear matter using our model of the quark self-energy and obtain
102: satisfactory results in agreement with the values of the scalar
103: and vector nucleon potentials in matter found in either
104: theoretical or phenomenological studies.
105: \end{abstract}
106:
107: % insert suggested PACS numbers in braces on next line
108: \pacs{12.39.Fe, 12.38.Aw, 14.65.Bt}
109: % insert suggested keywords - APS authors don't need to do this
110: %\keywords{}
111:
112: %\maketitle must follow title, authors, abstract, \pacs, and \keywords
113: \maketitle
114:
115: % body of paper here - Use proper section commands
116: % References should be done using the \cite, \ref, and \label commands
117: \section{INTRODUCTION}
118:
119: In recent years there has been a great deal of interest in
120: understanding the properties of quark matter at relatively low
121: temperature and high density [1-5]. Since it is difficult to study
122: QCD at finite chemical potential using lattice simulations, the
123: model of choice for such studies has been the Nambu--Jona-Lasinio
124: model [6-8]. Of particular interest is the prediction of diquark
125: condensates and color superconductivity at high densities. (Such
126: studies may be relevant to the properties of neutron stars.)
127: Calculations made for dense matter using the NJL model are carried
128: out in Minkowski space. These calculations are limited in that
129: they do not include a model of confinement and are, therefore,
130: unable to provide a comprehensive description of the hadronic
131: phase present at low density and temperature. However, it is
132: generally believed that a good deal of information may be gained
133: by studying quark matter, with a proper study of the hadronic
134: phase deferred until some future time.
135:
136: As is well known, the NJL model provides a microscopic dynamical
137: description of chiral symmetry breaking with the generation of
138: associated quark vacuum condensates and constituent masses. In the
139: standard versions of the NJL model [6-8], the constituent quark
140: mass that is generated in the model is a constant. However, it is
141: known from lattice simulations of QCD that the constituent mass
142: goes over to the current quark mass when the quark momentum,
143: $p^2$, is less than about -2\gev{2} [9]. It is our belief that, if
144: we are to use the NJL model to study dense matter, it is desirable
145: to make the model as realistic as possible. To that end, we have
146: introduced a nonlocal version of the NJL model [10] that is able
147: to reproduce the Euclidean-space behavior of the quark mass seen
148: in lattice simulations of QCD [9]. To carry out that program we
149: have introduced a momentum-dependent $q\bar q$ interaction in the
150: calculation of the quark self-energy and have separated the
151: regularization of the model from the specification of that
152: interaction. (This procedure requires the introduction of
153: additional parameters into the model.)
154:
155: Recently, we have seen an attempt to obtain the quark self-energy
156: in Minkowski space by analytic continuation of a Euclidean-space
157: form based upon gluon exchange enhanced at small momentum
158: transfers to simulate confinement [11]. The resulting
159: Minkowski-space values exhibit resonant-like structures and very
160: large values of the constituent mass. Such results do not appear
161: to have any natural physical interpretation.
162:
163: We note that the quark self-energy may be written as \be
164: \Sigma(\ksq)= A(\ksq)+ B(\ksq)\slr{k} \ee in vacuum. In matter
165: there is another four-vector, $\eta^\mu$, that describes the
166: motion of the matter rest frame. It is useful to put $\eta^2=1$
167: and to note that, if we work in the matter rest frame, we can put
168: $\eta^\mu=[1,0,0,0]$. In matter, we have \be \Sigma(\ksq,\eta\cdot
169: \!k)=A(\ksq,\eta\cdot\!k)+B(\ksq,\eta\cdot\!k)\overline{\slr
170: k}+C(\ksq,\eta\cdot\!k)\slr\eta\,, \ee where we have found it
171: useful to introduce the four-vector \be
172: \overline{k^\mu}=k^\mu-(k\cdot\!\eta)\eta^\mu\,. \ee Note that
173: $\overline{k^0}=0$ in the matter rest frame. In that frame, we may
174: write \be \Sigma(k^0, \vec k)=A(k^0, \vec k)-B(k^0, \vec
175: k)\vec\gamma\cdot\!\vec k+\gamma^0C(k^0, \vec k)\,. \ee As we will
176: see, $\textit{A}$ and $\textit{B}$ satisfy coupled nonlinear
177: equations, while $\textit{C}$ may be calculated independently.
178: (Note that $\textit{A}$ and $\textit{C}$ have the same dimension,
179: while $\textit{B}$ is dimensionless.)
180:
181: We first discuss our results for the case in which we neglect the
182: dependence of $\textit{A}$, $\textit{B}$, and $\textit{C}$ on
183: $k^0$. We anticipate that the dependence on $k^0$ will be
184: relatively weak and justify that assumption later in this work. As
185: a further simplification, we will at first neglect $\textit{B}$
186: and study the behavior of $A(\vec k,\rho)$, where $\rho$ is the
187: density of quark matter which is taken to contain equal numbers of
188: up and down quarks. In this case, the maximum value of $A(\vec
189: k,\rho)$ is found at $\vec k=0$, leading to a simpler presentation
190: of our results. We then go on to the consideration of the coupled
191: equations for $A(\vec k,\rho)$ and $B(\vec k,\rho)$. As usual, we
192: may introduce a density and momentum-dependent mass defined by \be
193: M(\vec k,\rho)= \frac{A(\vec k,\rho)}{1-B(\vec k,\rho)}\,. \ee We
194: provide values of $A(\vec k,\rho)$, $M(\vec k,\rho)$ and $C(\vec
195: k,\rho)$ in the following sections.
196:
197: We note that the inclusion of $C(\vec k,\rho)$ precludes the
198: passage to Euclidean space. The analogous problem arises when one
199: introduces a finite chemical potential. As we will see, $C(\vec
200: k,\rho)$ is quite large at finite density and can only be
201: neglected at very small densities. In vacuum, Lorentz invariance
202: leads to a relation between $C(\ksq)$ and $B(\ksq)$. However, our
203: formalism does not maintain Lorentz invariance. (For example, our
204: regulator depends only upon $|\vec k|$.) Therefore, in the
205: following we will only calculate the contribution to $C(\vec
206: k,\rho)$ from the matter, which takes the form of two Fermi seas
207: of up and down positive-energy quarks with Fermi momentum $k_{F}$.
208:
209: The Lagrangian of our model is \be {\cal L}=&&\bar q(i\slr
210: \partial-m^0)q +\frac{G_S}{2}\sum_{i=0}^8[
211: (\bar q\lambda^iq)^2+(\bar qi\gamma_5 \lambda^iq)^2]\nonumber\\
212: &&-\frac{G_V}{2}\sum_{i=0}^8[
213: (\bar q\lambda^i\gamma_\mu q)^2+(\bar q\lambda^i\gamma_5 \gamma_\mu q)^2]\nonumber\\
214: && +\frac{G_D}{2}\{\det[\bar q(1+\gamma_5)q]+\det[\bar
215: q(1-\gamma_5)q]\} \nonumber\\
216: &&+ {\cal L}_{\mbox{\tiny{conf}}}\,. \ee This Lagrangian was used
217: in Ref. [10] with $G_V=0$. There, both the condensate values and
218: self-energies were given for the up (down) and strange quarks. For
219: the present work we neglect the 't Hooft interaction $(G_D=0)$ and
220: our model of confinement. We also drop the strange quark from
221: consideration, so that the quark current mass matrix is
222: $m^0=\mbox{\,diag}(m_u^0, m_d^0)$ with $m_u^0=m_d^0$. Since the 't
223: Hooft interaction contributes to the self-energy [10], its neglect
224: leads us to use a larger value of $G_S$ than that we would use if
225: the 't Hooft interaction were included in our analysis.
226:
227: The organization of our work is as follows. In Section II we
228: present the coupled nonlinear equations that determine $A(\vec
229: k,\rho)$ and $B(\vec k,\rho)$ and also provide an expression for
230: $C(\vec k,\rho)$, and the density-dependent condensate
231: $\langle\bar uu\rangle_\rho=\langle\bar dd\rangle_\rho$. We
232: present values of $A(\vec k,\rho)$ and the condensate, when we
233: neglect $B(\vec k,\rho)$. In Section III we consider finite values
234: of $B(\vec k,\rho)$ and provide results for $A(\vec k,\rho)$,
235: $M(\vec k,\rho)$, $C(\vec k,\rho)$ and the condensate. In Section
236: IV, we consider the dependence of $\textit{A}$, $\textit{B}$ and
237: $\textit{C}$ on $k^0$ and provide some results for $A(k_0, \vec
238: k)$ in vacuum. In Section V we use the quark self-energy
239: calculated here to obtain the nucleon self-energy in nuclear
240: matter in a simple model. Finally, Section VI contains some
241: further discussion and conclusions.
242:
243:
244: \section{the quark self-energy}
245:
246: \begin{figure}
247: \includegraphics[bb=0 0 400 400, angle=-2, scale=0.7]{fig1.eps}%
248: \caption{The equation for the quark self-energy that was solved in Euclidean
249: space in Ref. [10] is shown. Here, $m^0$ is the current quark mass. The
250: open circle represents the momentum-dependent $q\bar q$ interaction
251: of the nonlocal model. The third term on the right-hand side of the figure
252: represents the 't Hooft interaction and the fourth term arises from our
253: model of confinement. The heavy lines are quark propagators which include the
254: self-energy, $\Sigma(k)$, in their definition. [See Eq. (2.1).]}
255: \end{figure}
256:
257:
258: In our earlier work we obtained the quark self-energy from the
259: solution of the equation depicted in Fig. 1 [10]. There, the open
260: circle is a momentum-dependent quark interaction, obtained by the
261: replacement \be G_S\longrightarrow f(k-\kp)\,G_Sf(k-\kp)\,,\ee
262: where $k$ and $\kp$ are the quark momenta entering (or leaving)
263: the interaction. We have used \be
264: f(k-\kp)=\mbox{exp}[-(k-\kp)\,^{2n}/2\beta]\ee with $n=4$ and
265: $\beta = 20$\gev{8} in Ref. [10]. The corresponding nonlocal
266: Lagrangian is given in Ref. [10]. On the right-hand side of Fig.1,
267: the 't Hooft interaction (third term) and the confinement
268: interaction (fourth term) are neglected for the purposes of this
269: work. In the second term we have contributions from the
270: negative-energy states in vacuum as well as the positive-energy
271: states at finite density, specified by the quark Fermi momentum,
272: $k_{F}$, of the up and down quark Fermi seas.
273:
274:
275: In general, the quark propagator is \be iS(k,
276: k\cdot\!\eta)=\frac{i}{\slr k-\Sigma(\ksq,
277: k\cdot\!\eta)+i\epsilon}\,,\ee which we will write as \be iS(k^0,
278: \vec k)=\frac{i}{(k^0-C(k^0, \vec k))\gamma^0-(1-B(k^0,\vec
279: k))\vec \gamma\cdot\!\vec k-A(k^0, \vec k)+i\epsilon}\,,\ee in the
280: matter rest frame. In a first approximation we write \be iS(k^0,
281: \vec k)=\frac{i}{(k^0-C(\vec k))\gamma^0-(1-B(\vec k))\vec
282: \gamma\cdot\!\vec k-A(\vec k)+i\epsilon}\,,\ee with $A(\vec k)$,
283: $B(\vec k)$ and $C(\vec k)$ density-dependent, in general. (On
284: occasion we will write $A(\vec k, \rho)$, etc.) We see that the
285: presence of $C(\vec k)$ precludes the passage to Euclidean space
286: that was made in Ref.[10]. That is analogous to the problem
287: created by the introduction of a chemical potential in the
288: formalism. Note that \be (k^0-C(\vec k))^2-(1-B(\vec k))^2\vec
289: k^2-A^2(\vec k)+i\epsilon \nonumber\\
290: =[k^0-E^+(\vec k)+i\epsilon][k^0-E^-(\vec k)-i\epsilon]\ee with
291: \be E^+(\vec k)=C(\vec k)+\sqrt{\vec k^2(1-B(\vec k))^2+A^2(\vec
292: k^2)}\ee and \be E^-(\vec k)=C(\vec k)-\sqrt{\vec k^2(1-B(\vec
293: k))^2+A^2(\vec k^2)}\,.\ee Here $E^+(\vec k)$ may be interpreted
294: as the (on-mass-shell) energy of the positive-energy states, while
295: $E^-(\vec k)$ refers to the negative-energy states.
296:
297: In order to simplify the notation somewhat, we write \be iS(k^0,
298: \vec k)=\frac{i}{\slr\Pi(k^0, \vec k)-A( \vec k)+i\epsilon}\,,\ee
299: where we have defined a four-component quantity \be \Pi^\mu(k^0,
300: \vec k)=[k^0-C(\vec k),(1-B(\vec k))\vec k]\,.\ee
301:
302: We also introduce a scalar quantity \be \rho_S(\vec
303: k)=iN_c(-1)\myint\kp\frac{4A(\vec
304: \kp)f^2(k-\kp)}{\Pi^2(k^{\,\prime\,0},\vec\kp)-A^2(\vec\kp)+i\epsilon}\,,\ee
305: where the minus sign is due to the closed Fermion loop in Fig. 1
306: and the factor of 4 comes from forming the trace associated with
307: the closed loop. We see that $\rho_S(\vec k)$ does not depend upon
308: $k^0$, since we are making an on-shell approximation in the
309: calculation of $f(k-\kp)$. In vacuum we have \be -iA(\vec
310: k)=-im^0+(2G_Si)\rho_S(\vec k)\,,\ee or \be A(\vec
311: k)=m^0-2G_S\rho_S(\vec k)\,,\ee where $\rho_S(\vec k)$ is real and
312: negative. We remark when $f(k-\kp)=1$, $A(\vec k)\rightarrow m$
313: and $\rho_S(\vec k)\rightarrow\langle\bar uu\rangle$, so that we
314: regain the usual result [6-8] \be m_u=m_u^0-2G_S\langle\bar
315: uu\rangle \,,\ee \be m_d=m_d^0-2G_S\langle\bar dd\rangle \,.\ee
316: (Note that our $G_S$ is one-half of the $G_S$ defined in Ref.
317: [7].)
318:
319: The evaluation of $\rho_S(\vec k)$ proceeds by closing to contour
320: in the complex $k^0$ plane. In the vacuum we find \be \rho_S(\vec
321: k)=-2N_c\mytint\kp\frac{f^2(k-\kp)A(\vec\kp)} {\sqrt{\vec
322: k^{\,\prime2}(1-B(\vec\kp))^2+A^2(\vec\kp)}}\,.\ee We define \be
323: E(\vec k)=\sqrt{\vec k^2(1-B(\vec k))^2+A^2(\vec k)}\,,\ee and \be
324: (k-\kp)^2=(E(\vec k)-E(\vec \kp))^2-(\vec k-\vec \kp)^2\,,\ee so
325: that $f(k-\kp)$ depends only upon $|\vec k|$, $|\vec\kp|$ and the
326: angle between $\vec k$ and $\vec \kp$.
327:
328: Integrals such as that in Eq. (2.16) require regularization. For
329: our calculations we insert a factor $\mbox{exp}[-\vec
330: k^{\,\prime2}\!/\,\alpha^2]$ with $\alpha=0.60$ GeV. (We have used
331: the same Gaussian regulator in our calculations of meson spectra,
332: where we have used $\alpha=0.605$ GeV [12-14]. However, those
333: calculations included a model of confinement, so that we can not
334: directly take over the parameters $G_S$, $G_V$ and $G_D$ used in
335: those works.) We remark that an expression for the
336: density-dependent condensate $\langle \bar uu\rangle_\rho$ may be
337: obtained by using Eq. (2.16) with $f(k-\kp)=1$.
338:
339: In the presence of matter it is useful to separate the propagator
340: into two parts, one of which will give rise to the explicitly
341: density-dependent terms. In this regard, it is useful to
342: generalize Eq. (5.8) of Ref. [6]. We define \be \Lambda^{(+)}(\vec
343: k)=\frac{E(\vec k)\gamma^0-\vec \gamma\cdot\vec
344: k(1-B(\vec k))+A(\vec k)}{2A(\vec k)}\,,\\
345: \Lambda^{(-)}(-\vec k)=\frac{-E(\vec k)\gamma^0-\vec
346: \gamma\cdot\vec k(1-B(\vec k))+A(\vec k)}{2A(\vec k)}\,,\ee where
347: $E(\vec k)=[\vec k^2(1-B(\vec k))^2+A^2(\vec k)]^{1/2}$. Then \be
348: S(k)=&&\frac{A(\vec k)}{E(\vec k)}\left[\frac{\Lambda^{(+)}(\vec
349: k)}{k^0-E^+(\vec k)} - \frac{\Lambda^{(-)}(-\vec k)}{k^0-E^-(\vec
350: k)}\right]\\ \nonumber &+&2\pi i\frac{A(\vec k)}{E(\vec
351: k)}\Lambda^{(+)}(\vec k)\,\theta (k_F-|\vec
352: k|)\,\delta(k^0-E^+(\vec k))\ee In the limit that $A(\vec
353: k)\rightarrow m^*$ and $B(\vec k)=C(\vec k)=0$, we have, with \be
354: E^*(\vec k)=\sqrt{\vec k^2+m^{*2}}\,,\ee \be
355: S(k)&=&\frac{m^*}{E^*(\vec k)}\left[\frac{\Lambda^{(+)}(\vec
356: k)}{k^0-E^*(\vec k)}
357: - \frac{\Lambda^{(-)}(-\vec k)}{k^0+E^*(\vec k)}\right]\\
358: \nonumber &+&2\pi i\frac{m^*}{E^*(\vec k)}\Lambda^{(+)}(\vec
359: k)\,\theta (k_F-|\vec k|)\,\delta(k^0-E^*(\vec k))\\ &=&
360: \frac{\slr k+m^*}{k^2-m^{*2}}+\frac{i\pi}{E^*(\vec k)}(\slr
361: k+m^*)\,\theta(k_F-|\vec k|)\,\delta(k^0-E^*(\vec k))\,,\ee which
362: agrees with Eq. (5.8) of Ref. [6].
363:
364: In the presence of matter, we have \be A(\vec k,
365: \rho)=m^0-2G_S[\rho_S^{\mbox{\tiny{vac}}}(\vec
366: k)-\rho_S^{\mbox{\tiny{mat}}}(\vec k)]\,,\ee where
367: $\rho_S^{\mbox{\tiny{mat}}}(\vec k)$ is calculated in the same
368: manner as $\rho_S^{\mbox{\tiny{vac}}}(\vec k)$, except that the
369: upper limit of the integral over $|\vec \kp|$ is $k_{F}$. Equation
370: (2.21) is a generalization of a corresponding equation that may be
371: found in Klevansky's review. (See Eqs. (5.18) of Ref. [6].)
372:
373: If we neglect $B(\vec k)$, Eqs. (2.11) and (2.16) provide a
374: nonlinear equation for $A(\vec k)$ which may be solved by
375: iteration. The results of such a calculation are reported in Table
376: I where, for nuclear matter, we put $k_{{F}}=0.268$ GeV. In Figs.
377: 2 and 3 we show values of the condensate $\langle \bar uu\rangle$
378: and $A(0,\rho)$ as a function of the density. These results may be
379: usefully discussed in terms of the relation [15]\be\langle \bar
380: uu\rangle_\rho=\langle \bar
381: uu\rangle_0\left(1-\frac{\sigma_N\rho_N}{f_\pi^2m_\pi^2}+\cdots\right)\,.\ee
382: Here $\sigma_N$ is the pion-nucleon sigma term and $\rho_N$ is the
383: density of nucleons. If we put $\sigma_N=0.050$ GeV,
384: $\rho_N=(0.109\,\mbox{GeV})^3$, $f_\pi=0.0942$ GeV and
385: $m_\pi=0.138$ GeV, we find a 38\% reduction of the condensate at
386: nuclear matter density, which agrees with our results given in
387: Table I and Fig. 2. It is of interest to note that the linear
388: dependence on the density implied by Eq. (2.26) appears to be
389: valid up to about twice nuclear matter density. However, one may
390: be concerned that, since we study quark matter rather than nuclear
391: matter, Eq. (2.26) may not be appropriate. Consider, however, the
392: relation \be\langle \bar uu\rangle_\rho=\langle \bar
393: uu\rangle_0\left(1-\frac{\sigma_q\rho_q}{f_\pi^2m_\pi^2}+\cdots\right)\,,\ee
394: where $\sigma_q$ is the quark sigma term and $\rho_q$ is the
395: number density of the quarks, which we may put equal to $3\rho_N$.
396: Thus, if $3\sigma_q=\sigma_N$, we may use Eq. (2.26). The fact
397: that $\sigma_q\simeq15$ MeV has been discussed by Vogl and Weise
398: [7]. We have also discussed this matter in great detail in Ref.
399: [16], where we calculated similar values of $\sigma_q$ using the
400: standard version of the NJL model. We conclude that the use of Eq.
401: (2.26), or Eq. (2.27), with an appropriate value of $\sigma_q$, is
402: satisfactory. For example, if $\sigma_q=\sigma_N/3$, as suggested
403: in Ref. [7], the two relations imply the same density dependence
404: of the condensate.
405:
406: In Fig. 3 we show $A(0, \rho)$ which is the density-dependent mass
407: parameter of the theory when $B(\vec k,\rho)=0$. We see that $A(0,
408: \rho)$ follows the trend seen in Fig.2 for the density dependence
409: of the condensate.
410:
411: \begin{figure}
412: \includegraphics[bb=0 0 400 200, angle=-0.5, scale=1]{fig2.eps}%
413: \caption{Values of the condensate $\langle\bar uu\rangle$ are given as a
414: function of $10^3k_{F}^3$. For nuclear matter $10^3k_{F}^3=19.2$\gev 3. [See Table I.]
415: Here $G_S=13.0$\gev{-2} and $B(\vec k,\rho)$ is put equal to zero.}
416: \end{figure}
417:
418: \begin{figure}
419: \includegraphics[bb=0 0 400 250, angle=-0.5, scale=1]{fig3.eps}%
420: \caption{Values of $A(0,\rho)$ are given as a function of $10^3k_{F}^3$.
421: [See Table I and the caption of Fig. 2.]}
422: \end{figure}
423:
424: \begin{table}%[H] add [H] placement to break table across pages
425: % \begin{ruledtabular}
426: \begin{tabular}{||@{\hspace{0.5cm}}
427: c@{\hspace{0.5cm}}|@{\hspace{0.5cm}}c@{\hspace{0.5cm}}
428: |@{\hspace{0.5cm}}c@{\hspace{0.5cm}}|@{\hspace{0.5cm}}c@{\hspace{0.5cm}}||}\hline\hline
429: $10^3k_{F}^3$ &$-\langle\bar uu\rangle^{1/3}$ &$-10^3\langle\bar uu\rangle$ &$A(0,\rho)$ \\
430: (\gev{3}) &(GeV) &(\gev{3}) &(GeV)\\\hline\hline
431: 0 &0.241 &14.0 &0.371\\\hline
432: 10 &0.225 &11.4 &0.298\\\hline
433: 19.2(nm) &0.205 &8.67 &0.226\\\hline
434: 25 &0.189 &6.75 &0.177\\\hline
435: 30 &0.171 &5.00 &0.133\\\hline
436: 40 &0.134 &2.46 &0.077\\\hline
437: 50 &0.113 &1.44 &0.041\\\hline
438: 60 &0.0990 &0.967 &0.030\\\hline
439: 70 &0.0915 &0.898 &0.024\\\hline
440: 80 &0.0853 &0.620 &0.021\\\hline
441: 90 &0.0805 &0.522 &0.018\\\hline\hline
442: % Lines of table here ending with \\
443: \end{tabular}
444: \vspace{1.2cm}
445: \caption{Values of the condensate and $A(0,\rho)$ are
446: given for $G_S=G_V=13.0$\gev{-2} and $m_u^0=m_d^0=0.005$ GeV. Here
447: $k_{F}=0.268$ GeV for nuclear matter. We note the reduction of
448: the condensate of 38\% and a 40\% reduction of $A(0,\rho)$ at nuclear matter
449: density [$10^3k_{F}^3=19.2$\gev 3]. Here $\alpha=0.60$ GeV is used in the
450: Gaussian regulator $\mbox{exp}[-\vec k^2/\alpha^2]$.}
451: % \end{ruledtabular}
452: \end{table}
453:
454:
455: \section{lorentz-vector terms of the quark self-energy}
456:
457: We now write $\Sigma(\vec k)=\Sigma_S(\vec k)+\Sigma_V(\vec k)$
458: where \be \Sigma_V(\vec k)=-\vec\gamma\cdot\vec kB(\vec
459: k)+\gamma^0C(\vec k)\,.\ee For the calculation of $C(\vec k)$ we
460: obtain the contribution from the last term in Eq. (2.21), with the
461: result that \be C(\vec k)=2G_V\rho_2^V(\vec k)\ee with \be
462: \rho_2^V(\vec
463: k)=2N_c\!\int^{k_{F}}\!\!\!\frac{d^3\!{\kp}}{(2\pi)^3}\,f^{\,2}(k-\kp)\,.\ee
464: An expression for $B(\vec k)$ may be found from the relation \be
465: -i[-\vec\gamma\cdot\vec kB(\vec k)]=(-2G_Vi)(-1)i\myint \kp
466: S(\kp)f^{\,2}(k-\kp)\ee if we only keep the term proportional to
467: $\vec \gamma\cdot\vec\kp$ in the expression for the quark
468: propagator. In vacuum we may compare corresponding terms in Eq.
469: (3.4) \be -\vec\gamma\cdot\vec kB(\vec
470: k)=-2G_V\myint\kp\frac{-\vec\gamma\cdot\vec \kp[1-B(\vec
471: \kp)]f^{\,2}(k-\kp)}{[\kp_0-E^+(\vec\kp)+i\epsilon][\kp_0-E^-(\vec\kp)-i\epsilon]}\,.\ee
472: Thus, \be B^{\mbox{\tiny{vac}}}(\vec
473: k)=2G_V\rho_1^{\mbox{\tiny{vac}}}(\vec k)\,,\ee with \be |\vec
474: k|\rho_1^{\mbox{\tiny{vac}}}(\vec k)=2N_c\mytint\kp\frac{|\vec
475: \kp|(\hat k\cdot\hat\kp)f^{\,2}(k-\kp)[1-B(\vec\kp)]}{\sqrt{\vec
476: k^{\,\prime2}[1-B(\vec\kp)]^2+A^2(\vec \kp)}}\,.\ee Here, $\hat k$
477: and $\hat\kp$ are unit vectors. As in the calculation of $A(\vec
478: k)$, Eq. (3.6) is generalized to read \be B(\vec
479: k)=2G_V[\rho_1^{\mbox{\tiny{vac}}}(\vec
480: k)-\rho_1^{\mbox{\tiny{mat}}}(\vec k)]\,,\ee where
481: $\rho_1^{\mbox{\tiny{mat}}}(\vec k)$ is calculated using Eq. (3.7)
482: with an upper limit on $|\vec\kp|$ of $k_{F}$.
483:
484: In Table II we present results of our calculation of the
485: condensate, $A(0,\rho)$, $A(0,\rho)/[1-B(0,\rho)]$, $C(0,\rho)$,
486: $B(0,\rho)$ and $A(0, \rho)-A(0,0)$. We also define \be U_S(\vec
487: k, \rho)=A(\vec k, \rho)-A(\vec k, 0)\,,\ee where $ U_S(\vec k,
488: \rho)$ is the density-dependent modification of $A(\vec k, 0)$ in
489: matter.
490:
491: It may be seen from the values given in Table II that there is a
492: thirty percent reduction of the condensate at the density of
493: nuclear matter, while the value of $A(0)$ is reduced by
494: thirty-nine percent. We would obtain a thirty percent reduction of
495: the condensate if $\sigma_N=39$ MeV.
496:
497: In Figs. 4 and 5 we exhibit values of $A(\vec k,\rho)$ and $A(\vec
498: k,\rho)/[1-B(\vec k, \rho)]$ for various densities and in Fig. 6
499: we present values of $C(\vec k,\rho)$. Figure 7 shows the values
500: of $U_S(\vec k,\rho_{NM})$ and $C(\vec k,\rho_{NM})$.
501:
502: \begin{table}%[H] add [H] placement to break table across pages
503: % \begin{ruledtabular}
504: \begin{tabular}{||@{\hspace{0.2cm}}c@{\hspace{0.2cm}}|@{\hspace{0.2cm}}c@{\hspace{0.2cm}}
505: |@{\hspace{0.2cm}}c@{\hspace{0.2cm}}|@{\hspace{0.2cm}}c@{\hspace{0.2cm}}|@{\hspace{0.2cm}}
506: c@{\hspace{0.2cm}}|@{\hspace{0.2cm}}c@{\hspace{0.2cm}}|@{\hspace{0.2cm}}c@{\hspace{0.2cm}}
507: |@{\hspace{0.2cm}}c@{\hspace{0.2cm}}||}\hline\hline
508: $10^3k_{F}^3$ &$-\langle\bar uu\rangle^{1/3}$ &$-10^3\langle\bar uu\rangle$ &$A(0,\rho)$
509: &$\frac{A(0,\rho)}{1-B(0,\rho)}$ &$U_S(0,\rho)$ &$C(0,\rho)$ &$B(0,\rho)$\\
510: (\gev{3}) &(GeV) &(\gev{3}) &(GeV) &(GeV) &(GeV) &(GeV) & \\\hline\hline
511: 0 &0.2401 &13.85 &0.365 &0.347 &0 &0 &-0.0544\\\hline
512: 10 &0.2279 &11.84 &0.292 &0.272 &-0.073 &0.0794 &-0.0736\\\hline
513: 19.2 &0.2128 &9.64 &0.223 &0.203 &-0.142 &0.0988 &-0.1019\\\hline
514: 30 &0.1755 &5.36 &0.114 &0.0942 &-0.252 &0.115 &-0.210 \\\hline
515: 40 &0.1441 &2.99 &0.0606 &0.0435 &-0.304 &0.126 &-0.394 \\\hline
516: 50 &0.1291 &2.15 &0.0432 &0.0279 &-0.322 &0.136 &-0.545 \\\hline\hline
517: % Lines of table here ending with \\
518: \end{tabular}
519: \vspace{1.2cm}
520: \caption{Various values are given for the case $G_S=13.5$\gev{-2}, $G_V=10.0$\gev{-2},
521: $m_u^0=m_d^0=0.005$ GeV and $\alpha=0.60$ GeV. Note a reduction of 30\% for the condensate
522: and 39\% for $A(0,\rho)$ at nuclear matter density, where $k_{F}=0.268$ GeV and $10^3k_{F}^3
523: =19.2$\gev 3.}
524: % \end{ruledtabular}
525: \end{table}
526:
527:
528: \begin{figure}
529: \includegraphics[bb=0 0 400 200, angle=-0.5, scale=1]{fig4.eps}%
530: \caption{Values of $A(\vec k, \rho)$ are given as a function of $|\vec k|$
531: for various densities: a) $10^3k_{{F}}^3=0$ [solid line]; b) $10^3k_{{F}}^3=10.0$\gev 3
532: [dashed line]; c) $10^3k_{{F}}^3=19.2$\gev 3 [dotted line]; d) $10^3k_{{F}}^3=30.0$\gev 3
533: [dot-dash line] and e) $10^3k_{{F}}^3=40.0$\gev 3 [short dash]. Here $G_S=13.5$\gev{-2},
534: $G_V=10.0$\gev{-2}, $m^0=0.005$ GeV and $\alpha=0.60$ GeV. [See Table II.]}
535: \end{figure}
536:
537: \begin{figure}
538: \includegraphics[bb=0 0 400 250, angle=-0.5, scale=1]{fig5.eps}%
539: \caption{The quantity $A(\vec k,\rho)/[1-B(\vec k,\rho)]$, which plays
540: the role of a momentum- and density-dependent mass parameter, is shown. [See Table
541: II and the caption of Fig. 4.]}
542: \end{figure}
543:
544: \begin{figure}
545: \includegraphics[bb=0 0 400 250, angle=-0.5, scale=1]{fig6.eps}%
546: \caption{Values of $C(\vec k, \rho)$ are shown for various densities:
547: a) $10^3k_{F}^3=10.0$\gev 3 [dashed line]; b) $10^3k_{F}^3=19.2$\gev 3 [solid line];
548: c) $10^3k_{F}^3=30.0$\gev 3 [dotted line] and d) $10^3k_{F}^3=40.0$\gev 3
549: [dash-dot line]. Here $G_S=13.5$\gev{-2}, $G_V=10.0$\gev{-2},
550: $m^0=0.005$ GeV and $\alpha=0.60$ GeV. }
551: \end{figure}
552:
553: \begin{figure}
554: \includegraphics[bb=0 0 400 250, angle=-0.5, scale=1]{fig7.eps}%
555: \caption{The values of $U(\vec k, \rho_{NM})=A(\vec k, \rho_{NM})-A(\vec k, 0)$
556: [solid line] and $C(\vec k, \rho_{NM})$ [dashed line] are shown. $U(\vec k, \rho_{NM})$
557: represents the density-dependent correction to the vacuum value of the
558: scalar term of the quark self-energy.}
559: \end{figure}
560:
561:
562: \section{off-mass-shell effects}
563:
564: In our calculations we have neglected the $k_0$ dependence of
565: $\textit{A}$, $\textit{B}$ and $\textit{C}$. In this Section we
566: provide some justification of that approximation. We can combine
567: Eqs. (2.9) and (2.11) to yield a nonlinear equation for
568: $A(k^0,\vec k)$ in vacuum \be A(k^0,\vec
569: k)=-(2G_Si)N_c(-1)\myint\kp
570: \frac{4A(k^{\,\prime0},\vec\kp)f^{\,2}(k-\kp)}{\slr k^{\,\prime}
571: -A(k^{\,\prime0},\vec\kp)+i\epsilon}\,.\ee Here we have neglected
572: $m^0$ and $B(k^0,\vec k)$. We see that $k^0$ dependence arises
573: from $f^{\,2}(k-\kp)$ which is given by \be
574: f^{\,2}(k-\kp)=\mbox{exp}[-(k-\kp)^{\,2n}/\beta]\,,\ee with \be
575: (k-\kp)^2=(k^0-k^{\,\prime0})^2-(\vec k-\vec \kp)^2\,,\ee which
576: differs from the on-mass-shell version given in Eq. (2.18).
577:
578: The solution of Eq. (4.1) for $A(k^0, \vec k)$ is shown in Fig. 8
579: for various values of $k^0$. It is seen that the dependence on
580: $k^0$ is weak as was assumed in this work.
581:
582: \begin{figure}
583: \includegraphics[bb=0 0 400 200, angle=-0.5, scale=1.5]{fig8.eps}%
584: \caption{Values of $A(k^0, \vec k)$ are shown as a function of $|\vec k|$ for
585: various values of $k^0$.}
586: \end{figure}
587:
588:
589: \section{the nucleon self-energy in matter}
590:
591: If one uses the Dirac equation to describe the interaction of a
592: nucleon with a nucleus, or with nuclear matter, it is found that a
593: strong scalar attraction is needed as well as a strong vector
594: repulsion [17, 18]. The scalar field is of the order of -400 MeV
595: and the vector field is about 300 MeV. It is of interest to see if
596: the nucleon self-energy, $\Sigma_N=V_S+\gamma^0V_V$, can be
597: calculated in terms of the quark self-energy obtained in this
598: work. To carry out this program we use a simple model of the
599: nucleon in which a quark is coupled to a scalar diquark. The full
600: complexity of the wave function, including vector diquarks and
601: various relativistic effects, is discussed in Ref. [19].
602:
603: We can calculate the nucleon self-energy in nuclear matter using a
604: triangle diagram in which one of the lower two vertices of the
605: triangle represents a vertex function for a zero-momentum nucleon
606: to emit a quark of momentum $\vec k$ leaving a spectator
607: (on-mass-shell) diquark of momentum $-\vec k$. The other lower
608: vertex represents the inverse process. At the upper vertex we
609: insert the quark self-energy, $\Sigma(\vec k)=U_S(\vec
610: k)+\gamma^0C(\vec k)$ calculated in this work and integrate over
611: $\vec k$. We make use of Fig. 4 of Ref. [19] and parametrize the
612: product of the vertex function and the quark Greens function by
613: the quark-diquark wave function \be \psi(\vec k)=\frac{1}{\sqrt
614: N}\, e^{-\vec k^2/\lambda^2}u(\vec k,s)\,,\ee where $u(\vec k, s)$
615: is a spinor for a quark of momentum $\vec k$ and spin projection
616: $s$ [19].
617:
618: The normalization for a single quark is obtained from the relation
619: \be \frac{1}{N}\mytint ke^{-2\vec
620: k^2/\lambda^2}\left(1+\frac{{\vec k}^2}{[E_q(\vec
621: k)+m_q]^2}\right)=1\,,\ee where we put $\lambda=0.18$ GeV to
622: correspond to the results of Ref. [19]. We may then relate the
623: nucleon self-energy to the quark self-energy. For a nucleon of
624: momentum $\vec P=0$, we have, with $E_q(\vec k)=[\vec
625: k^2+m_q^2]^{1/2}$, and $m_q=0.364$ GeV, \be V_V=\frac{3}N\mytint k
626: C(\vec k)\,e^{-2\vec k^2/\lambda^2}\left(1+\frac{{\vec
627: k}^2}{[E_q(\vec k)+m_q]^2}\right)\,,\ee and \be V_S=\frac3N\mytint
628: k U_S(\vec k)\,e^{-2\vec k^2/\lambda^2}\left(1-\frac{{\vec
629: k}^2}{[E_q(\vec k)+m_q]^2}\right)\,.\ee The difference of sign in
630: the brackets appearing in Eqs. (5.3) and (5.4) is due to the
631: different behavior of the Dirac matrices, $\mathbf {1}$ and
632: $\gamma^0$, at the upper vertex of the triangle.
633:
634: Since the momentum content of the quark-diquark wave function is
635: small [19], we expect that $V_S\simeq 3U_S(0)$ and $V_V\simeq
636: 3C(0)$, so that $V_V\simeq 296$ MeV and $V_S\simeq -426$ MeV. A
637: more careful evaluation of the integrals in Eqs. (5.3) and (5.4)
638: yields $V_V=295$ MeV and $V_S=-392$ MeV, which is in general
639: accord with the values given in Refs. [17] and [18].
640:
641: \section{discussion}
642:
643: The behavior of matter at high density and low temperature has
644: received a good deal of attention in the last few years [1-5]. A
645: large part of the work in this area has made use of the NJL model.
646: In our work we have attempted to modify the NJL model so that its
647: predictions are in greater accord with QCD. As a first step, in
648: Ref. [10] we introduced a momentum-dependent $q\bar q$ interaction
649: which allowed us to reproduce the Euclidean-space behavior for the
650: constituent quark mass obtained in lattice simulation of QCD [9].
651: In the present study we have considered the important vector
652: interactions of an extended NJL model. Of particular interest is
653: the behavior of the quark condensate at finite density. We find
654: that the linear behavior in the density exhibited in Eq. (2.26)
655: holds in our model up to about twice nuclear matter density. Also,
656: we note that the use of a nonzero current quark mass is important
657: for that result, since in the absence of an explicit chiral
658: symmetry breaking term, the model exhibits a first-order phase
659: transition at about 1.25 times nuclear matter density.
660:
661: Another point of interest are the results shown in Fig. 7. The
662: quark self-energy is similar to that found in relativistic nuclear
663: physics, with strong scalar attraction and strong vector
664: repulsion. The simple calculation reported in Section V suggests
665: that the quark self-energy, when multiplied by 3, provides a
666: satisfactory estimate of the nucleon self-energy which is in
667: accord with results of relativistic nuclear physics [17, 18].
668:
669: There is a body of work based upon the solutions of the
670: Schwinger-Dyson and Bethe-Salpeter equations with a
671: phenomenological form for the gluon propagator [20-23]. This body
672: of work is reviewed in Ref. [24]. In contrast to the results of
673: our work, it is found that the quark condensate
674: $\textit{incre{as}es}$ with increasing chemical potential. The
675: authors of Ref. [20] argue that the baryon density is zero up to
676: the critical chemical potential, $\mu_c$, for deconfinement. They
677: state that ``This result is an expected consequence of confinement
678: which entails that each additional quark must be locally paired
679: with an antiquark thereby increasing the density of condensate
680: pairs as $\mu$ is increased. For this reason, as long as
681: $\mu<\mu_c$, there is no excess of particles over antiparticles in
682: the vacuum and hence the baryon number density remains zero." We
683: note, however, that the baryon density usually considered in such
684: calculations is due to the presence of $\textit{nucleons}$, which
685: have a finite baryon density. As noted earlier, in our work the
686: baryon density is due to the presence of Fermi seas of up and down
687: quarks, which serve to provide a model of the baryon density that
688: would arise due to the presence of nucleons. We suggest that the
689: conclusions presented in Ref. [20] do not refer to the situation
690: of physical interest in which one studies the value of the quark
691: condensate in nuclei or in nuclear matter.
692:
693:
694: % If in two-column mode, this environment will change to single-column
695: % format so that long equations can be displayed. Use
696: % sparingly.
697: %\begin{widetext}
698: % put long equation here
699: %\end{widetext}
700:
701: \newpage
702: \vspace{1.5cm}
703: \noindent$\textbf{References}$\\[-2cm]
704: \begin{thebibliography}{99}
705: \bibitem{kf}For reviews, see K. Rajagopal and F. Wilcek, in B. L. Ioffe Festscrift;
706: $\textit{At the Frontier of Particle Physics/Handbook of QCD}$, M. Shifman ed. (World Scientific,
707: Singapore, 2001); M. Alford, hep-ph/0102047.
708: \bibitem{mr}M. Alford, R. Rajagopal and F. Wilcek, Phys. Lett.
709: B $\textbf{422}$\,, 247 (1998).
710: \bibitem{rts}R. Rapp, T. Schafer, E. V. Shuryak and M. Velkovsky,
711: Phys. Rev. Lett. $\textbf{81}$\,, 53 (1998).
712: \bibitem{mjk}M. Alford, J. Berges and K. Rajagopal, Nucl.
713: Phys. B $\textbf{558}$\,, 219 (1999).
714: \bibitem{jk}J. Kundu and K. Rajagopal, hep-ph/0112206 (2002).
715: \bibitem{spk}S. P. Klevansky, Rev. Mod. Phys. $\textbf{64}$\,,
716: 649 (1992).
717: \bibitem{uvw}U. Vogl and W. Weise, Prog. Part. Nucl. Phys.
718: $\textbf{27}$\,, 195 (1991).
719: \bibitem{tt}T. Hatsuda and T. Kunihiro, Phys. Rep.
720: $\textbf{247}$\,, 221 (1994).
721: \bibitem{js}J. Skullerud, D. B. Leinweber, and A. G. Williams,
722: Phys. Rev. D $\textbf{64}$\,, 074508 (2001).
723: \bibitem{bhqc}Bing He, Hu Li, Qing Sun, and C. M. Shakin,
724: nucl-th/0203010.
725: \bibitem{ra}R. Alkofer, P. Watson and H. Weigel,
726: hep-ph/0202053.
727: \bibitem{cm}C. M. Shakin and Huangsheng Wang, Phys. Rev. D $\textbf{65}$\,,
728: 094003 (2002).
729: \bibitem{cms5}C. M. Shakin and Huangsheng Wang, Phys. Rev. D
730: $\textbf{64}$\,, 094020 (2001).
731: \bibitem{lsbh}L. S. Celenza, Shun-fu Gao, Bo Huang, Huangsheng
732: Wang, and C. M. Shakin, Phys. Rev. C $\textbf{61}$\,, 035201
733: (2000).
734: \bibitem{ee}E. G. Drukarev and E. M. Levin, Prog. Part. Nucl,
735: Phys. $\textbf{27}$\,, 77 (1991).
736: \bibitem{ncw}Nan-Wei Cao, C. M. Shakin and Wei-Dong Sun, Phys.
737: Rev. C $\textbf{46}$\,, 2535 (1992).
738: \bibitem{bj}B. D. Serot and J. D. Walecka, in Advances in
739: Nuclear Physics, Vol. 16, edited by J. W. Negele and E. Vogt
740: (Plenum, New York, 1986).
741: \bibitem{lc}L. S. Celenza and C. M. Shakin, $\textit{Relativistic
742: Nuclear Physics: Theories of Structure and Scattering}$ (World
743: Scientific, Singapore, 1986).
744: \bibitem{jlcw}J. Szweda, L. S. Celenza, C. M. Shakin and W.-D.
745: Sun, Few-Body Systems $\textbf{20}$\,, 93 (1996).
746: \bibitem{agcsa}A. Bender, G. I. Poulis, C. D. Roberts, S.
747: Schmidt and A. W. Thomas, Phys. Lett. B $\textbf{431}$, 263
748: (1998).
749: \bibitem{dcs}D. Blaschke, C. D. Roberts, and S. Schmidt, Phys.
750: Lett. B $\textbf{425}$, 232 (1998).
751: \bibitem{pcsp}P. Maris, C. D. Roberts, S. M. Schmidt, and P.
752: C. Tandy, Phys. Rev. C $\textbf{63}$, 025202 (2001).
753: \bibitem{pcs}P. Maris, C. D. Roberts and S. Schmidt,
754: Phys. Rev. C $\textbf{57}$, R2821 (1998).
755: \bibitem{cs}C. D. Roberts and S. M. Schmidt, Prog. Part. Nucl.
756: Phys. $\textbf{45}$, S1(2000).
757:
758: \end{thebibliography}
759: % Create the reference section using BibTeX:
760: %\bibliography{basename of .bib file}
761:
762: \end{document}
763: %
764: % ****** End of file template.aps ******
765: