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: \def\qmm{$m^0=\mbox{diag}\,(m_u^0,
43: m_d^0, m_s^0)$}
44: \def\cond#1{\langle\bar #1\rangle}
45:
46:
47: \renewcommand{\theequation}%
48: {\arabic{section}.\arabic{equation}}
49: \makeatletter \@addtoreset{equation}{section} \makeatother
50:
51: \begin{document}
52:
53: % Use the \preprint command to place your local institutional report
54: % number in the upper righthand corner of the title page in preprint mode.
55: % Multiple \preprint commands are allowed.
56: % Use the 'preprintnumbers' class option to override journal defaults
57: % to display numbers if necessary
58: %\preprint{}
59:
60: %Title of paper
61: \title{Description of Deconfinement at Finite Matter Density in a Generalized Nambu--Jona-Lasinio Model}
62:
63:
64: \author{Hu Li}
65: \author{C.M. Shakin}
66: \email[email:]{casbc@cunyvm.cuny.edu}
67:
68: \affiliation{%
69: Department of Physics and Center for Nuclear Theory\\
70: Brooklyn College of the City University of New York\\
71: Brooklyn, New York 11210
72: }%
73:
74: \date{June, 2002}
75:
76: \begin{abstract}
77: Recent years have seen extensive applications of the
78: Nambu--Jona-Lasinio (NJL) model in the study of matter at high
79: density. There is a good deal of interest in the predictions of
80: diquark condensation and color superconductivity, with suggested
81: applications to the study the properties of neutron stars. As the
82: researchers in this field note, the NJL model does not describe
83: confinement, so that one is limited to the study of the deconfined
84: phase, which may set in at several times nuclear matter density.
85: Recently, we have extended the NJL model to include a covariant
86: confinement model. Our model may be used to study the properties
87: of the full range of light mesons, including their radial
88: excitations, in the 1-3 GeV energy domain. Most recently we have
89: used our extended model to provide an excellent fit to the
90: properties of the $\eta(547)$ and $\eta^\prime(958)$ mesons and
91: their radial excitations. The mixing angles and decay constants
92: are given successfully in our model. In the present work our goal
93: is to include a phenomenological model of deconfinement at finite
94: matter density, using some analogy to what is known concerning
95: ``string breaking" and deconfinement at finite temperature.
96: Various models may be used, but for this work we choose a specific
97: model for the density dependence of the parameters of our
98: confining interaction. We perform relativistic
99: random-phase-approximation (RPA) calculations of the properties of
100: the $\pi(138), K(495), f_0(980), a_0(980)$ and $K_0^*(1430)$
101: mesons and their radial excitations. In the model chosen for this
102: work, there are no mesonic states beyond about $2\rho_{NM}$, where
103: $\rho_{NM}$ is the density of nuclear matter. (The density for
104: deconfinement in our model may be moved to higher values by the
105: change of one of the parameters of the model.) This inability of
106: the model to support hadronic excitations at large values of the
107: density is taken as a signal of deconfinement. In addition to the
108: density dependence of the confining interaction, we use the
109: density-dependent quark mass values obtained in either the SU(2)
110: or SU(3)-flavor versions of the NJL model. We stress that other
111: assumptions for the density dependence of the confinement
112: potential, other than that used in this work, maybe considered in
113: future work, particularly if we are able to obtain further insight
114: in the dynamics of deconfinement at finite matter density.
115: \end{abstract}
116:
117: % insert suggested PACS numbers in braces on next line
118: \pacs{12.39.Fe, 12.38.Aw, 14.65.Bt}
119: % insert suggested keywords - APS authors don't need to do this
120: %\keywords{}
121:
122: %\maketitle must follow title, authors, abstract, \pacs, and \keywords
123: \maketitle
124:
125: % body of paper here - Use proper section commands
126: % References should be done using the \cite, \ref, and \label commands
127: \section{INTRODUCTION}
128:
129: In recent years we have developed a generalized
130: Nambu--Jona-Lasinio (NJL) model that incorporates a covariant
131: model of confinement [1-5]. The Lagrangian of the model is \be
132: {\cal L}=&&\bar q(i\slr
133: \partial-m^0)q +\frac{G_S}{2}\sum_{i=0}^8[
134: (\bar q\lambda^iq)^2+(\bar qi\gamma_5 \lambda^iq)^2]\nonumber\\
135: &&-\frac{G_V}{2}\sum_{i=0}^8[
136: (\bar q\lambda^i\gamma_\mu q)^2+(\bar q\lambda^i\gamma_5 \gamma_\mu q)^2]\nonumber\\
137: && +\frac{G_D}{2}\{\det[\bar q(1+\gamma_5)q]+\det[\bar
138: q(1-\gamma_5)q]\} \nonumber\\
139: &&+ {\cal L}_{conf}\,, \ee where the $\lambda^i(i=0,\cdots, 8)$
140: are the Gell-Mann matrices, with
141: $\lambda^0=\sqrt{2/3}\mathbf{\,1}$, $m^0=\mbox{diag}\,(m_u^0,
142: m_d^0, m_s^0)$ is a matrix of current quark masses and ${\cal
143: L}_{conf}$ denotes our model of confinement. Many applications
144: have been made in the study of light meson spectra, decay
145: constants, and mixing angles. In the present work we extend our
146: model to include a description of deconfinement at finite density.
147:
148: There has been extensive application of the NJL model in the study
149: of matter at high density, with particular interest in diquark
150: condensation and color superconductivity [6-9]. These studies find
151: application in the study of neutron stars. The NJL model is the
152: model of choice, since little insight into the properties of
153: matter at finite density can be obtained in lattice simulations of
154: QCD. This problem is associated with the introduction of a
155: chemical potential, which makes the Euclidean-space fermion
156: determinant complex.
157:
158: The use of the NJL model in the hadronic phase of matter is
159: limited, since the standard version of the model does not contain
160: a model of confinement [10-12]. It is clearly of value to extend
161: the NJL model so that one can study the full range of densities of
162: interest at this point in time. We are encouraged in this program
163: by recent results, obtained in lattice simulations of QCD with
164: dynamical quarks, that provide information on the temperature
165: dependence of the confining interaction [13]. It is generally
166: believed that the presence of matter will play a role similar to
167: that of finite temperature, with deconfinement taking place at
168: some finite density, which might be several times that of nuclear
169: matter. In the present work we make a specific assumption
170: concerning the density dependence of the confining field and then
171: calculate meson spectra in the presence of our density-dependent
172: confining interaction. We also take into account the density
173: dependence of the constituent quark masses, which is calculated in
174: the SU(2) or SU(3)-flavor version of the NJL model. As is well
175: known, the presence of matter leads to a reduction in the
176: magnitude of the quark vacuum condensates, which represents a
177: partial restoration of chiral symmetry in matter.
178:
179: Our calculations of the properties of mesons in matter is made
180: using a covariant random-phase-approximation (RPA) formalism,
181: which we have developed for the study of mesons in vacuum [1-5].
182: The organization of our work is as follows. In Section II we
183: provide a short review of our treatment of Lorentz-vector
184: confinement in our generalized NJL model. In Section III we
185: describe the variation of the up, down and strange quark
186: constituent quark masses in matter. In Section IV we discuss some
187: recent work concerning the temperature dependence of the confining
188: interaction, as obtained in lattice simulations of QCD with
189: dynamical quarks. We also specify the density dependence of the
190: confining field that we use in this work in Section IV. In Section
191: V we comment upon the phenomenon of pion condensation. (In our
192: work we introduce a small density dependence of the coupling
193: constants of the NJL model to simulate effects that prevent the
194: formation of a pion condensate in nuclear matter.) In Section VI
195: we discuss our covariant RPA calculations of meson properties in
196: vacuum and indicate how these calculations are modified in matter.
197: Results of our RPA calculations of the properties of pseudoscalar
198: mesons in matter are presented in Section VII, while Section VIII
199: contains similar results for scalar mesons. In the case of scalar
200: mesons, we study the $a_0(980), f_0(980)$, and $K_0^*(1430)$
201: mesons and their radial excitations. Finally, Section IX contains
202: some further discussion and conclusions.
203:
204:
205: \section{models of confinement}
206:
207: There are several models of confinement in use. One approach is
208: particularly suited to Euclidean-space calculations of hadron
209: properties. In that case one constructs a model of the quark
210: propagator by solving the Schwinger-Dyson equation. By appropriate
211: choice of the interaction one can construct a propagator that has
212: no on-mass-shell poles when the propagator is continued into
213: Minkowski space. Such calculations have recently been reviewed by
214: Roberts and Schmidt [14]. In the past, we have performed
215: calculations of the quark and gluon propagators in Euclidean space
216: and in Minkowski space. These calculations give rise to
217: propagators which did not have on-mass-shell poles [15-18].
218: However, for our studies of meson spectra, which included
219: descriptions of radial excitations, we found it useful to work in
220: Minkowski space.
221:
222: The construction of our covariant confinement model has been
223: described in a number of works [1-5]. We have made use of
224: Lorentz-vector confinement, so that the Lagrangian of our model
225: exhibits chiral symmetry. We begin with the form $V^C(r)=\kappa
226: r\mbox{exp}[-\mu r]$ and obtain the momentum-space potential via
227: Fourier transformation. Thus, \be V^C(\vec k-\vec
228: k\,^\prime)=-8\pi\kappa\left[\frac1{[(\vec k-\vec
229: k\,^\prime)^2+\mu^2]^2}-\frac{4\mu^2}{[(\vec k-\vec
230: k\,^\prime)^2+\mu^2]^3}\right]\,,\ee with the matrix form \be
231: \overline V{}\,^C(\vec k-\vec k\,^\prime)=\gamma^\mu(1)V^C(\vec
232: k-\vec k\,^\prime)\gamma_\mu(2)\,,\ee appropriate to
233: Lorentz-vector confinement. The potential of Eq. (2.1) is used in
234: the meson rest frame. We may write a covariant version of
235: $V^C(\vec k-\vec k^\prime)$ by introducing the four-vectors \be
236: \hat k^\mu=k^\mu-\frac{(k\cdot P)P^\mu}{P^2}\,, \ee and \be \hat
237: k^{\prime\,\mu}=k^{\prime\,\mu}-\frac{(k^\prime\cdot
238: P)P^\mu}{P^2}\,. \ee Thus, we have \be V^C(\hat k-\hat
239: k\,^\prime)=-8\pi\kappa\left[\frac1{[-(\hat k-\hat
240: k\,^\prime)^2+\mu^2]^2}-\frac{4\mu^2}{[-(\hat k-\hat
241: k\,^\prime)^2+\mu^2]^3}\right]\,.\ee Originally, the parameter
242: $\mu=0.010$ GeV was introduced to simplify our momentum-space
243: calculations. However, in the light of the following discussion,
244: we can remark that $\mu$ may be interpreted as describing
245: screening effects as they affect the confining potential [13].
246:
247: In our work, we found that the use of $\kappa=0.055$\gev2 gave
248: very good results for meson spectra. Here, $\kappa$ for the
249: Lorentz-vector potential is about one-fourth of the value of
250: $\kappa$ for Lorentz-scalar confinement. This difference arises
251: since the Dirac matrices $\gamma^\mu(1)\gamma_\mu(2)$ in Eq. (2.2)
252: give rise to a factor of 4 upon forming various Dirac trace
253: operations, so that the $\emph{effective}$ value of the string
254: tension is about the same in both Lorentz-scalar and
255: Lorentz-vector models of confinement.
256:
257: The potential $V^C(r)=\kappa r\mbox{exp}[-\mu r]$ has a maximum at
258: $r=1/\mu$, at which point the value is $V_{max}=\kappa/\mu e=
259: 2.023$ GeV. If we consider pseudoscalar mesons, which have $L=0$,
260: the continuum of the model starts at $E_{cont}=m_1+m_2+V_{max}$,
261: so that for $m_1=m_2=m_u=m_d=0.364$ GeV, $E_{cont}=2.751$ GeV. It
262: is also worth noting that the potential goes to zero for very
263: large r. Thus, there are scattering states whose lowest energy
264: would be $m_1+m_2$. However, barrier penetration plays no role in
265: our work. The bound states in the interior of the potential do not
266: communicate with these scattering states to any significant
267: degree. It is not difficult to construct a computer program that
268: picks out the bound states from all the states found upon
269: diagonalizing the random-phase-approximation Hamiltonian.
270:
271: Bound states in the confining field may be found by solving the
272: equation for the mesonic vertex function shown in Fig. 1a.
273: Inclusion of the short-range NJL interaction leads to an equation
274: for the vertex shown in Fig. 1b. We will return to a consideration
275: of Fig. 1b when we discuss our covariant RPA formalism in Section
276: VI.
277:
278: \begin{figure}
279: \includegraphics[bb=30 0 180 220, angle=90, scale=1.5]{fig1.eps}%
280: \caption{a) Bound states in the confining field (wavy line) may be found by solving
281: the equation for the vertex shown in this figure, b) Effects of both the confining
282: field and the short-range NJL interaction (filled circle) are included when solving
283: for the vertex shown in this figure.}
284: \end{figure}
285:
286: \section{calculation of constituent quark mass values}
287:
288: In this Section we report upon our calculation of the density
289: dependence of the constituent quark masses of the up (or down) and
290: strange quarks. The role of confinement in the calculation of the
291: constituent mass was studied in an earlier work in which
292: calculations were made in Euclidean space [19]. The results were
293: similar to those obtained in Minkowski-space calculations in which
294: confinement was neglected and it is the latter calculations which
295: we discuss here.
296:
297: The equations for the quark masses in the SU(3)-flavor NJL model
298: are [11] \be m_u=m_u^0-2G_S\langle\bar uu\rangle-G_D\langle\bar
299: dd\rangle\langle\bar ss\rangle\,,\\
300: m_d=m_d^0-2G_S\langle\bar dd\rangle-G_D\langle\bar
301: uu\rangle\langle\bar ss\rangle\,,\\
302: m_s=m_s^0-2G_S\langle\bar ss\rangle-G_D\langle\bar
303: uu\rangle\langle\bar dd\rangle\,,\ee where $\langle\bar uu\rangle,
304: \langle\bar dd\rangle$ and $\langle\bar ss\rangle$ are the quark
305: vacuum condensates. For example, with $N_c=3$, \be \langle\bar
306: uu\rangle=-4N_ci\myint k\frac{m_u}{\ksq-m_u^2+i\epsilon}\,.\ee If
307: this integral is evaluated in a Minkowski-space calculation, a
308: cutoff is used such that $|\vec k|\leq\Lambda_3$. Thus, \be
309: \langle\bar
310: uu\rangle=-4N_c\!\int^{\Lambda_3}\!\!\frac{d^3\!{k}}{(2\pi)^3}\,\frac{m_u}{2E_u(\vec
311: k)}\,,\ee etc. Here $E_u(\vec k)=\left[\vec
312: k{}^2+m_u^2\right]^{1/2}$.
313:
314: For studies at finite density, we consider the presence of two
315: Fermi seas of up and down quarks with Fermi momentum $k_F$. We
316: also take $m_u^0=m_d^0$ and obtain the density-dependent
317: equations, with $\langle\bar uu\rangle_\rho=\langle\bar
318: dd\rangle_\rho$, \be m_u(\rho)=m_u^0-2G_S\langle\bar
319: uu\rangle_\rho-G_D\langle\bar
320: dd\rangle_\rho\langle\bar ss\rangle_\rho\,,\\
321: m_s(\rho)=m_s^0-2G_S\langle\bar ss\rangle_\rho-G_D\langle\bar
322: uu\rangle_\rho\langle\bar dd\rangle_\rho\,.\ee Equation (3.5) is
323: now replaced by \be \langle\bar
324: uu\rangle_\rho=-4N_c\!\left[\int_0^{\Lambda_3}\!\!\frac{d^3\!{k}}{(2\pi)^3}\,\frac{m_u(\rho)}
325: {2E_u(\vec
326: k)}-\int_0^{k_F}\!\!\frac{d^3\!{k}}{(2\pi)^3}\,\frac{m_u(\rho)}
327: {2E_u(\vec k)}\right]\,,\ee with $E_u(\vec k)=\left[\vec
328: k{}^2+m_u^2(\rho)\right]^{1/2}$. On the other hand, since we do
329: not consider a background of strange matter, we have \be
330: \langle\bar
331: ss\rangle_\rho=-4N_c\!\int_0^{\Lambda_3}\!\!\frac{d^3\!{k}}{(2\pi)^3}\,\frac{m_s(\rho)}
332: {2E_s(\vec k)}\,,\ee with $E_s(\vec k)=\left[\vec
333: k{}^2+m_s^2(\rho)\right]^{1/2}$.
334:
335: We may argue that, with respect to our mean-field analysis, the
336: Fermi seas of up and down quarks yield contributions to the scalar
337: density that are similar to what would be obtained if the quarks
338: are organized into nucleons. One part of the argument is based
339: upon the well-known model-independent relation for the density
340: dependence of the condensate [20] \be \frac{\langle\bar
341: qq\rangle_\rho}{\langle\bar
342: qq\rangle_0}=\left(1-\frac{\sigma_N\rho}{f_\pi^2m_\pi^2}+\cdots\right)\,,\ee
343: where $\sigma_N$ is the pion-nucleon sigma term and $\rho$ is the
344: density of the matter. If we take $f_\pi=0.0942$ GeV,
345: $m_\pi=0.138$ GeV, $\rho_{NM}=(0.109\,\mbox{GeV})^3$ and
346: $\sigma_N=0.050$ GeV, we find a reduction of the condensate in
347: nuclear matter of 38\%, which is consistent with relativistic
348: models of nuclear matter [21, 22].
349:
350: We now consider the corresponding relation for a quark gas of up
351: and down quarks, \be \frac{\langle\bar qq\rangle_\rho}{\langle\bar
352: qq\rangle_0}=\left(1-\frac{\sigma_q\rho_q}{f_\pi^2m_\pi^2}+\cdots\right)\,,\ee
353: where $\rho_q$ is the density of quarks ($\rho_q=3\rho$) and
354: $\sigma_q$ is a ``quark sigma term". We have shown in earlier work
355: [23] that $\sigma_q$ is in the range of 15-17 MeV, so that Eq.
356: (3.10) and (3.11) imply that quite similar mean fields are
357: generated by the quark gas and by nuclear matter.
358:
359: In Table I and in Fig. 2, we show the results obtained when Eqs.
360: (3.8) and (3.9) are solved with $G_S=9.00$\gev{-2},
361: $G_D=-240$\gev{-5}, $\Lambda_3=0.631$ GeV, $m_u^0=0.0055$ GeV and
362: $m_s^0=0.130$ GeV. We note that the dependence of $m_u(\rho)$ on
363: density is approximately linear for $\rho/\rho_{NM}\leq2$, with a
364: 32\% reduction in the value of $m_u(\rho)$ when
365: $\rho/\rho_{NM}=1$. Another point to note is that $m_s(\rho)$ is
366: density-dependent for finite values of $G_D$, since the
367: $\langle\bar ss\rangle$ condensate is modified by the coupling to
368: the up and down quark condensates via the 't Hooft interaction.
369: This coupling becomes less important as the up and down quark
370: condensates are reduced at increasing density. [See Fig. 2.]
371:
372: \begin{figure}
373: \includegraphics[bb=0 0 400 220, angle=0, scale=1.5]{fig2.eps}%
374: \caption{The solution of Eqs. (3.6) and (3.7) for the density-dependent constituent
375: quark masses, $m_u(\rho)=m_d(\rho)$ and $m_s(\rho)$ are shown.
376: Here $G_S=9.00$\gev{-2}, $G_D=-240.0$\gev{-5}, $\Lambda_3=0.631$ GeV, $m_u^0=0.0055$ GeV
377: and $m_s^0=0.130$ GeV.}
378: \end{figure}
379:
380: We have also considered the solution for the SU(2) version of the
381: above equations \be m_u(\rho)=m_u^0-2G_S\langle\bar
382: uu\rangle_\rho\,,\ee and have used the parameters specified in the
383: Klevansky review article [10], $G_S=10.15$\gev{-2}, $m_u^0=0.0055$
384: GeV and $\Lambda_3=0.631$ GeV. The results for $m_u(\rho)$ are
385: similar to that seen in Fig. 2, except that $m_u(0)=0.336$ GeV.
386: [See Fig. 3.] In this case, $m_u(\rho)$ is reduced by about 32\%
387: when $\rho=\rho_{NM}$.
388:
389: \begin{table}%[H] add [H] placement to break table across pages
390: % \begin{ruledtabular}
391: \begin{tabular}{||@{\hspace{0.5cm}}
392: c@{\hspace{0.5cm}}|@{\hspace{0.5cm}}c@{\hspace{0.5cm}}
393: |@{\hspace{0.5cm}}c@{\hspace{0.5cm}}|@{\hspace{0.5cm}}c@{\hspace{0.5cm}}||}\hline\hline
394: $k_F^3$ &$\rho/\rho_{NM}$ &$m_u(\rho)$ &$m_s(\rho)$ \\
395: (\gev{3}) & &[GeV] &[GeV]\\\hline\hline
396: 0.00 &0.00 &0.358 &0.532\\\hline
397: 0.007 &0.364 &0.318 &0.515\\\hline
398: 0.010 &0.521 &0.300 &0.508\\\hline
399: 0.0140 &0.729 &0.276 &0.498\\\hline
400: 0.0192 &1.00 &0.242 &0.487\\\hline
401: 0.025 &1.302 &0.200 &0.475\\\hline
402: 0.030 &1.562 &0.162 &0.465\\\hline
403: 0.035 &1.823 &0.121 &0.457\\\hline
404: 0.040 &2.083 &0.0860 &0.452\\\hline
405: 0.045 &2.343 &0.0618 &0.449\\\hline
406: 0.050 &2.604 &0.0470 &0.448\\\hline
407: 0.055 &2.864 &0.0378 &0.448\\\hline
408: 0.060 &3.125 &0.0316 &0.448\\\hline
409: 0.065 &3.385 &0.0272 &0.447\\\hline\hline
410:
411: % Lines of table here ending with \\
412: \end{tabular}
413: \vspace{1.2cm}
414: \caption{Values of $m_u(\rho)$ and $m_s(\rho)$ obtained from the solution of Eqs. (3.6) and
415: (3.7) are given for various values of the ratio $\rho/\rho_{NM}$. (Here, $k_F^3=0.0192$\gev3
416: for nuclear matter, $m_u^0=0.0055$ GeV, $m_s^0=0.130$ GeV, $\Lambda_3=0.631$ GeV,
417: $G_S=9.00$\gev{-2}, $G_D=-240.0$\gev{-5}.)}
418: % \end{ruledtabular}
419: \end{table}
420:
421: \begin{figure}
422: \includegraphics[bb=0 0 400 220, angle=-0.5, scale=1.5]{fig3.eps}%
423: \caption{The solution of Eq. (3.12) for $m_u(\rho)$ is shown. Here $G_S=10.15$\gev{-2},
424: $m_u^0=0.0055$ GeV and $\Lambda_3=0.631$ GeV. (See Table V of Ref. [10].) The dashed line is a
425: linear approximation to the result which we use for $\rho\leq2\rho_{NM}$.
426: (Nuclear matter density
427: corresponds to $k_F^3=0.0192$ \gev3.}
428: \end{figure}
429:
430: \section{density and temperature dependence of the confining field}
431:
432: In part, our study has been stimulated by the results presented in
433: Ref. [13] for the temperature-dependent potential, $V(r)$, in the
434: case dynamical quarks are present. We reproduce some of the
435: results of that work in Fig. 4. There, the filled symbols
436: represent the results for $T/T_c=0.68, 0.80, 0.88$ and 0.94 when
437: dynamical quarks are present. This figure represents definite
438: evidence of ``string breaking", since the force between the quarks
439: appears to approach zero for $r > 1$ fm. This is not evidence for
440: deconfinement, which is found for $T=T_c$. Rather, it represents
441: the creation of a second $\bar qq$ pair, so that one has two
442: mesons after string breaking. Some clear evidence for string
443: breaking at zero temperature and finite density is reported in
444: Ref. [24].
445:
446: \begin{figure}
447: \includegraphics[bb=0 0 250 350, angle=-90, scale=1.2]{fig4.eps}%
448: \caption{A comparison of quenched (open symbols) and unquenched results (filled symbols) for
449: the interquark potential at finite temperature [13]. The dotted line is the zero temperature
450: quenched potential. Here, the symbols for $T=0.80T_c$ [open triangle], $T=0.88T_c$
451: [open circle], $T=0.80T_c$ [open square], represent the quenched
452: results. The results with dynamical fermions are given at $T=0.68T_c$ [solid downward-pointing
453: triangle], $T=0.80T_c$ [solid upward-pointing triangle], $T=0.88T_c$ [solid circle],
454: and $T=0.94T_c$ [solid square].}
455: \end{figure}
456:
457: In order to study deconfinement in our generalized NJL model, we
458: need to specify the interquark potential at finite density. We
459: start with our model that was described in Section II. In that
460: case we had $V^C(r)=\kappa r\mbox{exp}[-\mu r]$. For the model we
461: study in this work, we write \be V^C(r, \rho)=\kappa
462: r\mbox{exp}[-\mu(\rho) r]\ee and put \be
463: \mu(\rho)=\frac{\mu_0}{1-\left(\displaystyle\frac\rho{\rho_C}\right)^2}\,,\ee
464: with $\rho_C=2.25\rho_{NM}$ and $\mu_0=0.010$ GeV. With this
465: modification our results for meson spectra in the vacuum are
466: unchanged. Other forms than that given in Eqs. (4.1) and (4.2) may
467: be used. However, in this work we limit our analysis to the model
468: presented in these equations. The corresponding potentials for our
469: model of Lorentz-vector confinement are shown in Fig. 5 for
470: several values of $\rho/\rho_{NM}$.
471:
472: \begin{figure}
473: \includegraphics[bb=0 20 400 220, angle=0, scale=1.5]{fig5.eps}%
474: \caption{Values of $V(r,\rho)$ are shown, where $V(r,\rho)=\kappa r\exp[-\mu(\rho)r]$
475: and $\mu(\rho)=\mu_0/[1-(\rho/\rho_C)^2]$. Here $\rho_C=2.25\rho_{NM}$ and $\mu_0=0.010$ GeV.
476: The values of $\rho/\rho_{NM}$ are 0.0 [solid line], 0.50 [dotted line], 1.0 [dash line],
477: 1.50 [dash-dot line]. 1.75 [dash-dot-dot line], 2.0 [short-dash line],
478: and 2.1 [small dot line].}
479: \end{figure}
480:
481: We can see from Fig. 4 that, for $T=0.94T_c$, the use of dynamical
482: quarks leads to an approximately constant value of $V(r)=1000$ MeV
483: for larger $r$. If we perform a Fierz rearrangement of the
484: Lorentz-scalar potential to study pseudoscalar $q\bar q$ states,
485: one introduces a factor of 1/4, making the value at large $r$ to
486: be about 250 MeV. (See Eq. (B1) of Ref. [10].) However,
487: rearranging the Lorentz-vector potential to study pseudoscalar
488: $q\bar q$ states introduces a factor of 1. Now, let us consider
489: $\rho/\rho_{NM}=0.94(\rho_C/\rho_{NM})\simeq2.11$, and find the
490: maximum of our Lorentz-vector potential at that density from the
491: relation $V_{max}=\kappa/\mu(\rho)e$. Using our value for
492: $\mu(\rho)$ at $\rho/\rho_{NM}=2.11$, we obtain $V_{max}=0.227$
493: GeV. The value for the Lorentz-vector potential compares favorably
494: with the value of $V(r)$, for large $r$, quoted above. This result
495: suggests that, if the dynamics of chiral symmetry restoration and
496: deconfinement at finite temperature is somewhat analogous to the
497: deconfinement process at finite density, our use of
498: $\rho/\rho_{NM}=2.25$ may be a satisfactory choice.
499:
500: \section{pion condensation and the choice of the parameters of the interaction}
501:
502: It was suggested many years ago that the ground state of nuclear
503: matter might have an unusual structure due to presence of pionlike
504: excitations [25]. In finite nuclei such effects could imply
505: anomalous behavior in states with $J^\pi=0^-, 1^+, 2^-\ldots$,
506: etc. However, the nucleon-nucleon interaction is sufficiently
507: repulsive in the relevant channel so that pion condensation does
508: not take place at normal nuclear matter densities. That matter has
509: been discussed in Ref. [26]. A constant $g^\prime$ parametrizes
510: the strength of a nuclear force in the spin-isospin channel that
511: represents short-range correlation effects and exchange effects.
512: (See Eq. (5.11a) of Ref. [26].) The phenomenological value of
513: $g^\prime$, obtained from the study of nuclear excitations, is
514: sufficiently large so that pion condensation does not take place
515: until about three times nuclear matter density. (See Fig. 5.9 of
516: Ref. [26].)
517:
518: In our work we will model the effects that prevent pion
519: condensation by introducing a density-dependent interaction for
520: the pionic states calculated in the NJL model. We write \be
521: G_\pi(\rho)=G_\pi(0)[1-0.087\rho/\rho_{NM}]\,,\ee where the second
522: term in Eq. (5.1) represents medium effects that reduce the pion
523: self-energy in matter. Here $G_\pi(0)$ is the linear combination
524: of $G_S$ and $G_D$ given on page 269 of Ref. [12], \be
525: G_\pi=G_S+\frac{G_D}2\langle\bar ss\rangle\,.\ee Equation (5.1)
526: represents our scheme for parametrizing the nuclear matter effects
527: that prevent pion condensation. In our calculations of pionlike
528: excitations we put $G_\pi(0)=13.49$\gev{-2}, and used a constant
529: values of $G_V=11.46$\gev{-2}. We may check that our choice of
530: $G_\pi(0)$ is reasonable by using Eq. (5.2) with
531: $G_S=11.84$\gev{-2} and $-180$\gev{-5}$\leq G_D\leq 240$\gev{-5}.
532: These values of $G_S$ and $G_D$ were obtained in our extensive
533: study of the eta mesons [1]. Thus, if we take $\langle\bar
534: ss\rangle=-(0.258\,\mbox{GeV})^3$ and $G_D=-190$\gev{-5}, we find
535: $G_\pi(0)=13.47$\gev{-2}. This analysis suggests that, once we fix
536: our parameters in the study of the eta mesons, we can then infer
537: the parameters needed for our study of the pion in vacuum.
538:
539: For this work, in our study of the kaon, we use
540: $G_K(0)=13.07$\gev{-2} and $G_V=11.46$\gev{-2}. Note that [12] \be
541: G_K(0)=G_S+\frac{G_D}2\cond{dd}_0\,.\ee If we take
542: $G_S=11.84$\gev{-2}, $G_D=-190$\gev{-5} and
543: $\cond{uu}=-(0.240\,\mbox{GeV})^3$, we find
544: $G_K(0)=13.15$\gev{-2}, which is close to the value of
545: $G_K(0)=13.07$\gev{-2} used in our calculations. In our work we
546: have used \be G_K(\rho)=G_K(0)[1-0.087\rho/\rho_{NM}]\,.\ee In the
547: case of the kaon, about 40\% of the assumed density dependence of
548: $G_K(\rho)$ may be attributed to the density dependence of
549: $\cond{uu}_\rho$ or $\cond{dd}_\rho$. We may consider the relation
550: \be G_K(\rho)=G_S(\rho)+\frac{G_D}2\cond{dd}_\rho\,,\ee and use a
551: somewhat smaller reduction of $G_S(\rho)$ for the kaon than that
552: used for the pion in Eq. (5.5), since the reduction of
553: $\cond{uu}_\rho$ or $\cond{dd}_\rho$ in matter effectively reduces
554: the interaction strength.
555:
556: In the absence of $a_0-f_0$ coupling we have
557: $G_{33}^S=G_{a_0}=G_S-(G_D/2)\cond{ss}$ [12]. If we again put
558: $G_S=11.84$\gev{-2}, $G_D=-190$\gev{-5}, and
559: $\cond{ss}=-(0.258\,\mbox{GeV})^3$, we have
560: $G_{a_0}=10.21$\gev{-2}, which places the $a_0(980)$ at 1.13 GeV.
561: However, in the case of the scalar mesons there exist significant
562: contributions to the interaction from processes that describe the
563: scalar meson decay to two-meson channels. An extended discussion
564: of these effects was given in an early work on scalar mesons [27].
565: In the case of the $f_0(980)$ we presented a discussion of such
566: terms as they affect the energy predicted for the $f_0(980)$ in
567: Ref. [28].
568:
569: In order to take into account these effects, which are not
570: included in our RPA calculations, we increase the value of the
571: $a_0$ coupling constant to $G_{a_0}=13.10$\gev{-2}. That has the
572: effect of moving the $a_0(980)$ mass down to 980 MeV.
573:
574: We also introduce some density dependence of the interaction to
575: avoid an ``$a_0$ condensate", which would otherwise take place at
576: $\rho=1.75\rho_{NM}$, if we use
577: $m_u(\rho)=m_d(\rho)=0.0055+0.3585(1-0.4\rho/\rho_{NM})$. Thus, we
578: use $G_{a_0}(\rho)=G_{a_0}(0)[1-0.045\rho/\rho_{NM}]$ when we
579: allow for the rapid decrease in the value of $m_u(\rho)=m_d(\rho)$
580: given by the above expression. It is possible that the small
581: reduction of $G_{a_0}(\rho)$ in matter given above has it origin
582: in a somewhat smaller attraction generated at the larger densities
583: by the real part of the polarization operator that describes decay
584: to the two-meson channels [27, 28]. We will provide further
585: details of our treatment of the scalar mesons in Section VIII.
586:
587: \section{random phase approximation for mesonic excitations}
588:
589: In this work we report upon covariant random-phase-approximation
590: (RPA) calculations of meson spectra in vacuum and in dense matter.
591: Before writing the equations of our model, it is worth discussing
592: some properties of RPA calculations made for many-body systems
593: [29, 30]. For example, such calculations have been performed to
594: study excited states of nuclei. In the RPA one usually does not
595: attempt to construct the wave function of the ground state.
596: Rather, one considers amplitudes of particle-hole operators taken
597: between the excited state and the ground state. The dominant
598: amplitude usually involves the creation of a hole in the ground
599: state and the creation of a particle in what are predominantly
600: unoccupied states. Smaller amplitudes are found if one destroys a
601: hole in the ground state and destroys a particle in the
602: predominantly unoccupied states. These smaller amplitudes are only
603: nonzero, if one allows for correlations in the ground state.
604:
605: Such RPA calculations are particularly important for states that
606: are collective with respect to matrix elements of electromagnetic
607: transition operators, for example. In hadron physics the most
608: ``collective state" is the $\pi(138)$. In this case the ``large"
609: and ``small" components of the wave function, in the sense of the
610: RPA, are comparable in magnitude and approach equality in
611: magnitude as one approaches the chiral limit, when
612: $m_\pi\rightarrow0$.
613:
614: Another important feature of RPA calculations is that they may be
615: considered as an investigation of the properties of small
616: oscillations about the ground state. Thus, if one obtains an
617: imaginary energy value for the ground state, one infers that the
618: ground state is unstable. A new ground state must be constructed
619: that will yield real eigenvalues. (Note that imaginary eigenvalues
620: may be obtained, since the RPA Hamiltonian is not Hermitian.)
621:
622: There is a strong analogy that can be made between the
623: particle-hole RPA calculations described above and the calculation
624: of mesonic excitations. For example, a ``hole" in the ground state
625: (the vacuum) is an antiquark, while the particle state is the
626: quark. If we perform relativistic RPA calculations for the pion
627: and its radial excitations, an imaginary energy calculated for the
628: pion is a signal of pion condensation.
629:
630: Random-phase-approximation equations may be derived using the
631: vertex equation of Fig. 1b. The RPA equations for the study of the
632: pion, kaon, and eta mesons were derived in Ref. [1]. In the case
633: of the pion and kaon we include pseudoscalar---axial-vector
634: coupling. The most complex case is that of the eta mesons which,
635: in addition to pseudoscalar---axial-vector coupling, involves
636: singlet-octet coupling in the flavor sector.
637:
638: In this work we only record the equations in the simplest example,
639: that of RPA calculations for the $a_0$ mesons [31]. In this case
640: the large component is denoted as $\phi^+(k)$, while the small
641: component is $\phi^-(k)$. These functions are found to satisfy
642: coupled equations for mesons in vacuum: \be
643: 2E_u(k)\phi^+(k)+\int\!
644: d\kp\,[H_C(k,\kp)+H_{NJL}(k,\kp)]\phi^+(\kp)\\\nonumber+\int\!
645: d\kp \,H_{NJL}(k,\kp)\phi^-(\kp)=P^0\phi^+(k)\,,\ee \be
646: -2E_u(k)\phi^-(k)-\int\!
647: d\kp\,[H_C(k,\kp)+H_{NJL}(k,\kp)]\phi^-(\kp)\\\nonumber-\int\!
648: d\kp \,H_{NJL}(k,\kp)\phi^+(\kp)=P^0\phi^-(k)\,,\ee where
649: $E_u(k)=[\vec k\,{}^2+m_u^2]^{1/2}$, \be
650: H_C(k,\kp)=-\frac1{(2\pi)^2}\frac{[2V_0^C(k,\kp)k^2\kpsq+k\kp
651: V_1^C(k,\kp)]}{E_u(k)E_u(\kp)}\,,\ee and \be
652: H_{NJL}=\frac{8N_c}{(2\pi)^2}\frac{\ksq\kpsq
653: G_{a_0}e^{-\ksq/2\alpha^2}e^{-\kpsq/2\alpha^2}}{E_u(k)E_u(\kp)}\,.\ee
654: In Eq. (6.3) we have introduced \be
655: V_l^C(k,\kp)=\frac12\int_{-1}^1\!dx\, P_l(x)V^C(\vec k-\vec
656: k{}^\prime)\,.\ee Here, $x=\mbox{cos}\theta$ and $P_l(x)$ is a
657: Legendre function. The terms exp[$-\ksq/2\alpha^2$] and
658: exp[$-\kpsq/2\alpha^2$] are regulators with $\alpha=0.605$ GeV.
659:
660: In order to solve these equations in the presence of matter, we
661: replace $m_u$, $G_{a_0}$ and $\mu_0$ by $m_u(\rho)$,
662: $G_{a_0}(\rho)$ and $\mu(\rho)$. (Recall that
663: $\mu(\rho)=\mu_0/[1-(\rho/\rho_C)^2]$.) In our calculation for the
664: $a_0$ states we have taken
665: $m_u(\rho)=m_u^0+0.3585\,\mbox{GeV}\,[1-0.4\rho/\rho_{NM}]$ and
666: $G_{a_0}(\rho)=G_{a_0}(0)[1-0.045\rho/\rho_{NM}]$, with
667: $m_u^0=0.0055$ GeV. As an alternative, the mass values for
668: $m_u(\rho)=m_d(\rho)$ may be taken from Table I.
669:
670: \section{results of numerical calculations: pseudoscalar mesons}
671:
672: The choice of the parameters in the case of the pion and its
673: radial excitations was discussed in Section V. We use
674: $G_\pi(\rho)=G_\pi(0)[1-0.087\rho/\rho_{NM}]$ and
675: $m_u(\rho)=m_d(\rho)=0.0055+0.3585[1-0.4\rho/\rho_{NM}]$ with
676: $G_\pi(0)=13.49$\gev{-2} and $G_V=11.46$\gev{-2}. Also,
677: $\mu(\rho)=\mu_0/[1-(\rho/\rho_C)^2]$ with $\mu_0=0.010$ GeV and
678: $\rho_C=2.25\rho_{NM}$.
679:
680: The results of our calculations are shown in Fig. 6. At $\rho=0$,
681: the first radial excitation of the pion is found at 1.319 GeV. The
682: large number of states above 1.3 GeV have wave functions that are
683: dominated by either the $\gamma_5$ or $\gamma_0\gamma_5$ vertex.
684: The pion wave function has mainly a $\gamma_5$ vertex structure,
685: with a small admixture of the $\gamma_0\gamma_5$ vertex. (The
686: axial-vector part of the wave function makes a significant
687: contribution in the calculation of the pion decay constant,
688: $f_\pi$.)
689:
690: \begin{figure}
691: \includegraphics[bb=0 50 280 280, angle=0, scale=1.5]{fig6.eps}%
692: \caption{The mass values for the pion and its radial excitations are presented as a function of
693: the density of matter. Here, $G_\pi(\rho)=G_\pi(0)[1-0.087\rho/\rho_{NM}]$
694: and $m_u(\rho)=m_d(\rho)=m_u^0+0.3585\,\mbox{GeV}[1-0.4\rho/\rho_{NM}]$,
695: with $m_u^0=0.0055$ GeV. We use $G_\pi(0)=13.49$\gev{-2}, $G_V=11.46$\gev{-2}
696: and $\mu=\mu_0/[1-(\rho/\rho_C)^2]$, with $\mu_0=0.010$ GeV and $\rho_C=2.25\rho_{NM}$.}
697: \end{figure}
698:
699: It may be seen from the figure, that with the reduction of the
700: value of the constituent mass and of the confining field with
701: increasing values of $\rho/\rho_{NM}$, the radial excitations that
702: appear as bound states become fewer in number. Beyond
703: $\rho/\rho_{NM}=1.50$ only the nodeless pion wave function is
704: bound and that state is no longer supported beyond
705: $\rho/\rho_{NM}\simeq1.80$. That represents the beginning of the
706: deconfined phase in the case of the pion for the model introduced
707: in this work.
708:
709: Somewhat similar behavior is found for the kaon and its radial
710: excitations, as may be seen in Fig. 7. Here we have used the mass
711: values given in Table I and
712: $G_K(\rho)=G_K(0)[1-0.087\rho/\rho_{NM}]$ with
713: $G_K(0)=13.07$\gev{-2} and $G_V=11.46$\gev{-2}. Again we see only
714: a small increase of the mass of the nodeless state, the pseudo
715: Goldstone boson, as $\rho/\rho_{NM}$ is increased. We again find
716: deconfinement for $\rho/\rho_{NM}>1.8$. The density dependence of
717: $G_K(\rho)$ is taken to be the same as in the case of the pion.
718: However, in this case, we have noted previously that about 40\% of
719: the reduction of $G_K(\rho)$ with increasing density may be
720: ascribed to the density dependence of the up and down quark
721: condensates, $\cond {uu}_\rho$ and $\cond {dd}_\rho$. The
722: calculation of the density dependence of the coupling constants in
723: our model is a major undertaking and is beyond the scope of this
724: work.
725:
726: \begin{figure}
727: \includegraphics[bb=0 0 280 220, angle=0, scale=1.5]{fig7.eps}%
728: \caption{Mass values of the $K$ mesons are shown as a function of the density of matter.
729: Here we use $G_K(0)=13.07$\gev{-2}, $G_K(\rho)=G_K(0)[1-0.087\rho/\rho_{NM}]$,
730: $G_V=11.46$\gev{-2} and $\mu=\mu_0/[1-(\rho/\rho_C)^2]$, with $\mu_0=0.010$ GeV
731: and $\rho_C=2.25\rho_{NM}$. The mass values given in Table I are used.}
732: \end{figure}
733:
734: \section{results of numerical calculations: scalar mesons}
735:
736: We have recently discussed the properties of the $f_0(980)$,
737: giving particular attention to the role of the polarization
738: diagrams that describe the decay of the $f_0$ mesons to the
739: $\pi\pi$ or $K\bar K$ channels [28]. (See Fig. 2 of Ref. [28].)
740: However, when we diagonalize the RPA Hamiltonian we do not take
741: those terms into account. Calculations of such effects are more
742: easily made if we construct a quark-antiquark $T$ matrix. For a
743: single channel example we may write \be t(p^2)=-\frac
744: G{1-GJ(p^2)}\,,\ee where $G$ is the appropriate coupling constant
745: for that channel and $J(p^2)$ is the corresponding vacuum
746: polarization operator. In our model $J(p^2)$ is calculated with
747: the confining vertex function that appears in Fig. 1a as a
748: crosshatched region. (See Fig. 1 of Ref. [28].) The resulting
749: $J(p^2)$ is a real function, which is singular at the values of
750: $p^2$ for which there is a bound state in the confining field. If
751: we include polarization diagrams that describe coupling to
752: two-meson decay channels, Eq. (8.1) is modified to read \be
753: t(p^2)=-\frac
754: G{1-G[J(p^2)+\mbox{Re}K(p^2)+i\mbox{Im}K(p^2)]}\,.\ee The
755: calculation of $J(p^2)$ and $K(p^2)$ has been extensively
756: discussed in our earlier work. In the case of the scalar mesons,
757: inclusion of $\mbox{Re}K(p^2)$ can move the mass of the
758: lowest-energy state down by about 70-100 MeV [27, 28].
759:
760: In the case of the $a_0(980)$, the use of $G_S$ and $G_D$
761: determined in our study of the eta mesons places the $a_0(980)$ at
762: 1.13 GeV. In the present work we have increased the coupling
763: constant from $G_{a_0}=10.21$\gev{-2} to $G_{a_0}=13.10$\gev{-2}
764: to move the lowest $a_0$ state down to 980 MeV. That creates a
765: problem of ``$a_0$ condensation" which we avoid by taking
766: $G_{a_0}(\rho)=G_{a_0}(0)[1-0.045\rho/\rho_{NM}]$. One may
767: speculate that the effects that increase the effective coupling
768: strength from $G_{a_0}=10.21$\gev{-2} to $G_{a_0}=13.10$\gev{-2}
769: have some density dependence that reduces the induced attraction
770: at the higher densities.
771:
772: In Fig. 8 we show our results for the $a_0$ mesons. There we see
773: deconfinement at about $\rho=2.0\rho_{NM}$ which is a slightly
774: larger value of the density than that found for the other mesons
775: studied in this work. However, the behavior of the lowest $a_0$
776: state with increasing density is made somewhat uncertain because
777: of our lack of knowledge of the appropriate form for
778: $G_{a_0}(\rho)$.
779:
780: \begin{figure}
781: \includegraphics[bb=0 60 400 400, angle=0, scale=1]{fig8.eps}%
782: \caption{Mass values for the $a_0$ mesons are given as a function of the matter density.
783: Here, we have used $G_{a_0}(0)=13.10$\gev{-2} and
784: $G_{a_0}(\rho)=G_{a_0}(0)[1-0.045\rho/\rho_{NM}]$. We have
785: used $m_u=m_u^0+0.3585\,\mbox{GeV}[1-0.4\rho/\rho_{NM}]$
786: with $m_u^0=0.0055$ GeV. The dotted line results, if we put
787: $G_{a_0}(\rho)=G_{a_0}(0)[1-0.087\rho/\rho_{NM}]$ and use the mass values of
788: Table I. The dotted curve is similar to the curve for the
789: $a_0$ mass given in Ref. [31]. The curves representing the masses of
790: the radial excitations are changed very little when we use the second
791: form for $G_{a_0}(\rho)$ given above.}
792: \end{figure}
793:
794: For our study of the $f_0$ mesons we work in a singlet-octet
795: representation and use the coupling constants
796: $G_{00}^S=14.25$\gev{-2}, $G_{88}^S=10.65$\gev{-2} and
797: $G_{08}^S=G_{80}^S=0.4953$\gev{-2}. This choice yields 980 MeV for
798: the mass of the $f_0(980)$. The fact that $G_{00}^S>G_{88}^S$ is a
799: feature of the 't Hooft interaction and leads to the $f_0(980)$
800: being mainly a singlet state [28]. (For the $\eta(547)$ the
801: behavior of the 't Hooft interaction is such that
802: $G_{88}^S>G_{00}^S$ [12, 28] and, therefore, the $\eta(547)$ is
803: predominantly a flavor octet meson [1].)
804:
805: In our study of the $f_0$ mesons at finite density we use the mass
806: values of Table I and do not introduce any density dependence for
807: $G_{00}^S$, $G_{88}^S$ and $G_{08}^S$. The results of our
808: calculation are shown in Fig. 9. Since the $f_0(980)$ has a
809: significant $\bar ss$ component, the mass value only decreases
810: slowly, with a value of 700 MeV for the lowest $f_0$ state at
811: $\rho/\rho_{NM}=1.82$, where deconfinement sets in.
812:
813: \begin{figure}
814: \includegraphics[bb=0 50 280 270, angle=0, scale=1.5]{fig9.eps}%
815: \caption{The figure shows the mass values of the $f_0$ mesons as a
816: function of density. The mass values for the quarks are taken from
817: Table I. In a singlet-octet representation, we have used the
818: constants $G_{00}^S=14.25$\gev{-2}, $G_{08}^S=10.65$\gev{-2} and $G_{88}^S=0.4953$\gev{-2}.
819: Deconfinement takes place somewhat above $\rho=1.8\rho_{NM}$.
820: Here $\mu=\mu_0/[1-(\rho/\rho_C)^2]$
821: with $\mu_0=0.010$ GeV and $\rho_C=2.25\rho_{NM}$.}
822: \end{figure}
823:
824: In Ref. [28] we provide a discussion of the $T$ matrix for the
825: singlet-octet channels. There the role of $K_{00}^S(p^2)$,
826: $K_{08}^S(p^2)$ and $K_{88}^S(p^2)$ in lowering the energy
827: predicted for the $f_0(980)$ is discussed in some detail.
828:
829: Our results for the energy levels of the $K_0^*$ mesons are given
830: in Fig. 10. In this case we use a constant value for
831: $G_{K_0^*}=10.25$\gev{-2}. The results are hardly modified if we
832: allow for a small density dependence of $G_{K_0^*}$. Since the
833: $K_0^*$ mesons contain a strange quark, the density dependence of
834: their energies is not as marked as that of the $a_0$ mesons which
835: only contain up and down quarks in our model. In that regard, the
836: behavior of the $K_0^*$ mesons is more like that of the $f_0$
837: mesons, which have some strange quark content. Again we see
838: deconfinement for $\rho>1.8\rho_{NM}$.
839:
840: \begin{figure}
841: \includegraphics[bb=0 50 280 270, angle=0, scale=1.5]{fig10.eps}%
842: \caption{The figure shows the mass values obtained for the $K_0^*$ mesons as a
843: function of density. Here we use a constant $G_{K_0^*}=10.25$\gev{-2}.
844: Deconfinement takes place somewhat above $\rho=1.8\rho_{NM}$. The quark mass values
845: were taken from Table I.}
846: \end{figure}
847:
848: \section{discussion}
849:
850: We originally chose $\rho_C=2.25\rho_{NM}$, since the curve in
851: Fig. 2 that shows the values of $m_u(\rho)$ seemed to change its
852: behavior at about $k_F^3=0.045$\gev3, which corresponds to
853: $\rho\simeq2.3\rho_{NM}$. We can attempt to see if that is a
854: reasonable choice by noting that ``string breaking" should occur
855: when the energy of the extended string is equal to the energy of
856: the lowest two-meson state that can be formed when the string
857: breaks. Therefore, we may write $V_{max}=m_1+m_2$, where $m_1+m_2$
858: are the masses of the mesons in the final state. We then use
859: $V_{max}=\kappa/\mu(\rho)e$ to find a value $\mu(\rho)$ and obtain
860: $\rho/\rho_C$ from the expression
861: $\mu(\rho)=\mu_0/[1-(\rho/\rho_C)^2]$. We then put
862: $\rho_C=2.25\rho_{NM}$ and calculate the value of $\rho/\rho_{NM}$
863: where we might expect string breaking. We consider the final
864: states $\pi\pi$, $\pi K$, $\pi\eta$ and $K\bar K$. The
865: corresponding values of $\rho/\rho_{NM}$ are 2.09, 1.86, 1.83, and
866: 1.61 for $\rho_C=2.25\rho_{NM}$. Note that the $K(495)$ and
867: $K_0^*(1430)$ mesons can break up into the $\pi K$ system, while
868: the $a_0(980)$ is strongly coupled to the $\pi\eta$ channel. The
869: $f_0(980)$ is coupled both to the $\pi\pi$ and $K\bar K$ channels.
870: On the whole, the values of $\rho/\rho_{NM}$ calculated above are
871: generally consistent with the value of that quantity that leads to
872: deconfinement in our model. That result tends to suggest that, for
873: light mesons, the density that leads to string breaking may be
874: similar to the density for deconfinement. (In general, however,
875: these processes are distinct and further studies would be needed
876: to see if string breaking and deconfinement are related at finite
877: density.) We may suggest that, if the initial meson is of the same
878: type as the mesons that appear upon string breaking, it becomes
879: reasonable to suggest that the instability of the initial mesons
880: is also felt by the final state mesons, giving rise to the
881: relation of string breaking and deconfinement suggested above for
882: light mesons.
883:
884: A comprehensive discussion of meson properties at finite
885: temperature and density has been presented by Lutz, Klimt and
886: Weise [32]. Since those authors did not include a model of
887: confinement, they were able to calculate values of the meson
888: masses for large values of the density. Their Fig. 8 shows the
889: calculated masses of the nodeless pion, $f_0$ and $a_0$ mesons for
890: $0\leq\rho/\rho_{NM}\leq 3.5$. They also give the result for an
891: $f_0^\prime$ excitation. (The $f_0$ and $f_0^\prime$ exhibit
892: singlet-octet mixing.) Compared to our results, their value of the
893: $f_0$ mass falls more rapidly than ours, becoming degenerate with
894: the pion mass at about $\rho/\rho_{NM}=3$. On the other hand, the
895: mass of the $a_0$ in their work is about 600 MeV at
896: $\rho/\rho_{NM}=2$. They are able to derive systematic low-density
897: expansions for various quantities which provide important insight
898: into the results obtained in numerical studies. They also show
899: that effects due to finite quasiparticle size are important in
900: stabilizing the density and temperature dependence of the pion
901: mass. The main deficiency of their work is the absence of a model
902: of confinement. Therefore, we believe our work provides a natural
903: extension of the work reported in Ref. [32].
904:
905: It is worth noting that deconfinement takes place in our model at
906: about $\rho=1.8\rho_{NM}$, while the confining potential goes to
907: zero at $\rho=\rho_C=2.25\rho_{NM}$. That suggests that the
908: specific form we have chosen for the density dependence,
909: $\mu(\rho)=\mu_0/[1-(\rho/\rho_C)^2]$, is not particularly
910: important. What is more important is the behavior of our confining
911: potential, $V^C(r, \rho)$, shown in Fig. 5. There, we see that the
912: potential still has a substantial magnitude at
913: $\rho=1.75\rho_{NM}$ and $\rho=2.10\rho_{NM}$.
914:
915: Since the analysis of Ref. [32] is made in the absence of a model
916: of confinement, many analytic results can be obtained for the
917: behavior of various quantities when small changes in density and
918: temperature are considered. Indeed, the work of that reference
919: provides some support for our treatment of the pion and kaon. It
920: is shown that the Goldstone boson remains at zero mass in the
921: chiral limit as long as the system remains in the Goldstone-Nambu
922: mode of symmetry breaking. For finite current quark masses, we
923: quote the result given in Eq. (5.6) of Ref. [32] for $T=0$, \be
924: \frac{dm_\pi^2}{m_\pi^2}=\left(1-2m_u^2\langle
925: r_S^2\rangle\right)\frac{d\cond{uu}}{\cond{uu}}\,.\ee Here, $r_S$
926: is the quasiparticle radius. That quantity is defined in terms of
927: the form factor $F_S(\vec p-\vec p\,{}^\prime)$ that appears in
928: the matrix element of the u-quark scalar density \be \langle
929: u(\vec p\,{}^\prime)|\,\bar uu(0)|\,u(\vec p)\rangle=F_S(\vec
930: p-\vec p\,{}^\prime)\bar u(p\,{}^\prime)u(p)\,.\ee In Eq. (9.2)
931: $u(\vec p)$ denotes the Dirac spinor of a constituent u quark with
932: four-momentum $p$. The scalar mean-squared radius is then \be
933: \left.\langle r_S^2\rangle=6\frac d{dq^2}\ln
934: F_S(q^2)\right|_{\,q^2=0}\,.\ee (See Eq. (A.7) of Ref. [32] for an
935: explicit expression for $\langle r_S^2\rangle$ in terms of the
936: parameters of the NJL model.) With the well-known relation [20]
937: \be
938: \frac{d\cond{uu}}{\cond{uu}}=-\frac{\sigma_N\rho}{m_\pi^2f_\pi^2}\,,\ee
939: Eq. (9.1) becomes \be dm_\pi^2=-\left(1-2m_u^2\langle
940: r_S^2\rangle\right)\frac{\sigma_N\rho}{f_\pi^2}\,.\ee If one
941: ignores the quasiparticle size, one has
942: $dm_\pi^2=-(\sigma_N\rho/f_\pi^2)$ [33, 34], which implies pion
943: condensation at a critical density
944: $\rho_{crit}=f_\pi^2m_\pi^2/\sigma_N=(0.148\,\mbox{GeV})^3$, which
945: is about 2.5 $\rho_{NM}$.
946:
947: The second term in Eq. (9.5) works against condensation. With
948: $m_u=0.364$ GeV and $r_S=0.40$ fm [32] one finds that $\delta
949: m_\pi^2$ increases slowly with increasing density, as born out by
950: the calculations reported in Ref. [32]. Our choice of
951: $G_\pi(\rho)=G_\pi(0)[1-0.087\rho/\rho_{NM}]$ reproduces the
952: almost constant value of $m_\pi$. We see that the
953: density-dependent term in $G_\pi(\rho)$ plays a similar role in
954: our model as that played by the second term in Eq. (9.5).
955:
956: We have some confidence in our treatment of the pion and kaon at
957: finite density. We recall that we were able to find satisfactory
958: values of $G_\pi(0)$ and $G_K(0)$ using the values of $G_S$ and
959: $G_D$ obtained in our study of the eta mesons [1]. Therefore, our
960: work provides a unified approach for the nonet of pseudoscalar
961: mesons in the presence of a model of confinement.
962:
963: Since confinement is important for the $a_0(980)$ and $f_0(980)$
964: mesons, it is uncertain whether the results of Ref. [32] for the
965: properties of these mesons can be trusted. These mesons are in the
966: continuum of the NJL model without confinement and various
967: assumptions need to be made as to how the formalism is to be
968: applied. For a small increase in density, the mass of the $a_0$ in
969: our model and in Ref. [32] are similar. For the larger values of
970: density, the use of
971: $G_{a_0}(\rho)=G_{a_0}(0)[1-0.045\rho/\rho_{NM}]$ in model leads
972: to a rather small mass for the $a_0$ for $\rho\sim2\rho_{NM}$.
973: [See Fig. 8.]
974:
975: Our treatment of the $a_0$ mesons is less satisfactory than that
976: of $\pi$ and $K$ mesons, since coupled channel effects are
977: important in the case of the scalar mesons. Using the values of
978: $G_S$ and $G_D$ obtained in our study of the eta mesons [1], we
979: found the lowest $a_0$ state at 1.13 GeV. To place the $a_0$ at
980: 980 MeV, we increased the value of $G_{a_0}(0)$. That increase led
981: to the possibility of an $a_0$ condensation, which was removed by
982: reducing the coupling constant with increasing density. [See Fig.
983: 8.] However, it might be preferable to accept the value of 1.13
984: GeV for the mass of the $a_0$ and, therefore, avoid the problem of
985: $a_0$ condensation. Our difficulty in this case arises since we do
986: not know the density dependence of the processes that move the
987: $a_0$ mass from our predicted value to the experimental value of
988: 980 MeV.
989:
990: In our model we see some relation between the partial restoration
991: of chiral symmetry and deconfinement. With reference to Fig. 2, we
992: see that the up (or down) quark mass drops in a roughly linear
993: manner with increasing density up to about 2 or 2.5 times nuclear
994: matter density. With the reduction of the magnitude of the
995: confining field, as seen in Fig. 5, the combined effect of the
996: smaller confining field and reduced quark mass values leads to
997: deconfinement at about 1.8 $\rho_{NM}$. For the $f_0$, $K$ and
998: $K_0^*$ mesons, the reduction of the mass of the quarks is less
999: important, since these mesons have either one strange quark ($K$,
1000: $K_0^*$) or an $\bar ss$ component ($f_0$). However, deconfinement
1001: still takes place at about $\rho=1.8\rho_{NM}$ for the mesons.
1002:
1003: In future work we will study the dependence of the deconfinement
1004: process on both temperature and density. In addition, it would be
1005: desirable to have some understanding of the mechanism by which the
1006: increased matter density modifies the confining interaction.
1007:
1008:
1009: % If in two-column mode, this environment will change to single-column
1010: % format so that long equations can be displayed. Use
1011: % sparingly.
1012: %\begin{widetext}
1013: % put long equation here
1014: %\end{widetext}
1015:
1016: %\newpage
1017: \vspace{1.5cm}
1018: \noindent$\textbf{References}$\\[-2cm]
1019: \begin{thebibliography}{99}
1020: \bibitem{cms6}C. M. Shakin and Huangsheng Wang, Phys. Rev. D
1021: $\textbf{65}$\,, 094003 (2002).
1022: \bibitem{cms5}C. M. Shakin and Huangsheng Wang, Phys. Rev. D
1023: $\textbf{64}$\,, 094020 (2001).
1024: \bibitem{cms7}C. M. Shakin and Huangsheng Wang, Phys. Rev. D
1025: $\textbf{63}$\,, 074017 (2001).
1026: \bibitem{lhc}L. S. Celenza, Huangsheng Wang, and C. M. Shakin,
1027: Phys. Rev. C $\textbf{63}$\,, 025209 (2001).
1028: \bibitem{cms }C. M. Shakin and Huangsheng Wang, Phys. Rev. D
1029: $\textbf{63}$\,, 014019 (2000).
1030: \bibitem{kf}For reviews, see K. Rajagopal and F. Wilcek, in B. L. Ioffe Festscrift;
1031: \emph{At the Frontier of Particle Physics/Handbook of QCD}, M. Shifman ed.
1032: (World Scientific, Singapore 2001); M. Alford, hep-ph/0102047.
1033: \bibitem{mr}M. Alford, R. Rajagopal and F. Wilcek, Phys. Lett.
1034: B $\textbf{422}$\,, 247 (1998).
1035: \bibitem{rts}R. Rapp, T. Sch$\ddot a$fer, E. V. Shuryak and M. Velkovsky,
1036: Phys. Rev. Lett. $\textbf{81}$\,, 53 (1998).
1037: \bibitem{mjk}M. Alford, J. Berges and K. Rajagopal, Nucl.
1038: Phys. B $\textbf{558}$\,, 219 (1999).
1039: \bibitem{spk}S. P. Klevansky, Rev. Mod. Phys. $\textbf{64}$\,,
1040: 649 (1992).
1041: \bibitem{uvw}U. Vogl and W. Weise, Prog. Part. Nucl. Phys.
1042: $\textbf{27}$\,, 195 (1991).
1043: \bibitem{tt}T. Hatsuda and T. Kunihiro, Phys. Rep.
1044: $\textbf{247}$\,, 221 (1994).
1045: \bibitem{cofe}C. DeTar, O. Kaczmarek, F. Karsch, and E. Laermann, Phys.
1046: ReV. D $\textbf{59}$\,, 031501 (1998).
1047: \bibitem{cs}C. D. Roberts and S. M. Schmidt, Porg. Part. Nucl.
1048: Phys. $\textbf{45}$\,, S1 (2000).
1049: \bibitem{c1}C. Shakin, Ann. Phys. (NY) $\textbf{192}$\,, 254
1050: (1989).
1051: \bibitem{lhx}L. S. Celenza, Hui-Wen Wang and Xin-Hua Yang, Intl. J. Mod. Phys. A
1052: $\textbf{4}$\,, 3807 (1989).
1053: \bibitem{vlhsc}V. M. Bannur, L. S. Celenza, Huang-he Chen,
1054: Shun-fu Gao and C. M. Shakin, Intl. J. Mod. Phys. A $\textbf{5}$\,, 1479 (1990).
1055: \bibitem{vlch}V. M. Bannur, L. S. Celenza, C. M. Shakin, and Hui-Wen
1056: Wang, Description of the QCD Vacuum as a Random Medium,
1057: Brookly College Report: BCCNT89/032/189---unpublished.
1058: \bibitem{lbhqc}L. S. Celenza, Bing He, Hu Li, Qing Sun, and C. M. Shakin,
1059: nucl-th/0203010.
1060: \bibitem{ee}E. G. Drukarev and E. M. Levin, Prog. Part. Nucl,
1061: Phys. $\textbf{27}$\,, 77 (1991).
1062: \bibitem{bj}B. D. Serot and J. D. Walecka, in Advances in
1063: Nuclear Physics, Vol. 16, edited by J. W. Negele and E. Vogt
1064: (Plenum, New York, 1986).
1065: \bibitem{lc}L. S. Celenza and C. M. Shakin, \emph{Relativistic
1066: Nuclear Physics: Theories of Structure and Scattering} (World
1067: Scientific, Singapore, 1986).
1068: \bibitem{ncw}Nan-Wei Cao, C. M. Shakin and Wei-Dong Sun, Phys.
1069: Rev. C $\textbf{46}$\,, 2535 (1992).
1070: \bibitem{cb}C. Bernard \emph{et al.}, Phys. Rev. D $\textbf{64}$\,,
1071: 074509 (2001).
1072: \bibitem{ab}A. B. Migdal, Soviet Phys., JETP $\textbf{34}$\,,
1073: 1184 (1972).
1074: \bibitem{tw}T. Ericson and W. Weise, \emph{Pions and Nuclei}
1075: (Oxford Univ. Press, Oxford, 1988).
1076: \bibitem{lsbh}L. S. Celenza, Shun-fu Gao, Bo Huang, Huangsheng
1077: Wang, and C. M. Shakin, Phys. Rev. C $\textbf{61}$\,, 035201
1078: (2000).
1079: \bibitem{c2}C. M. Shakin, Phys. Rev. D $\textbf{65}$\,, 114011
1080: (2002).
1081: \bibitem{aj}A. L. Fetter and J. D. Walecka, \emph{Quantum Theory
1082: of Many-Particle Systems}, (Mc Graw-Hill, New York, 1971).
1083: \bibitem{dj}D. J. Rowe, \emph{Nuclear Collective Motion},
1084: (Methuen, London, 1970).
1085: \bibitem{cms8}C. M. Shakin and Huangsheng Wang, Phys. Rev. D
1086: $\textbf{63}$\,, 114007 (2001).
1087: \bibitem{msw}M. Lutz, S. Klimt, and W. Weise, Nucl. Phys. A
1088: \textbf{541}\,, 521 (1992).
1089: \bibitem{da}D. B. Kaplan and A. E. Nelson, Phys. Letter. B
1090: \textbf{175}\,, 57 (1986).
1091: \bibitem{ad}A. E. Nelson and D. B. Kaplan, Phys. Letter. B
1092: \textbf{192}\,, 193 (1987).
1093:
1094: \end{thebibliography}
1095: % Create the reference section using BibTeX:
1096: %\bibliography{basename of .bib file}
1097:
1098: \end{document}
1099: %
1100: % ****** End of file template.aps ******
1101: