hep-ph0302152/am.tex
1: \tolerance=10000
2: \newcommand{\be}{\begin{eqnarray}}
3: \newcommand{\ee}{\end{eqnarray}}
4: \documentclass[twocolumn,aps,showpacs]{revtex4}
5: %\documentstyle[preprint,aps]{revtex}
6: %\unitlength=1mm
7: \usepackage{graphics}
8: \begin{document}
9: %\tightenlines
10: %\draft
11: \title{$\rho$ propagation and dilepton production at finite pion density
12: and temperature} 
13: \author{Alejandro Ayala}  
14: \affiliation{Instituto de Ciencias Nucleares, Universidad Nacional 
15:          Aut\'onoma de M\'exico, Apartado Postal 70-543, 
16:          M\'exico Distrito Federal 04510, M\'exico.}
17: \author{Javier Magnin}
18: \affiliation{Centro Brasileiro de Pesquisas Fisicas, Rua Dr. Xavier
19:              Sigaud 150 - URCA CEP 22290-180 Rio de Janeiro Brazil.}
20: \begin{abstract}
21: 
22: We study the propagation properties of the $\rho$ vector in a dense
23: and hot pion medium. We introduce a finite value of the chemical
24: potential associated to a conserved pion number and argue that such
25: description is valid during the hadronic phase of a relativistic
26: heavy-ion collision, between chemical and thermal freeze-out, where the
27: strong interaction drives pion number to a fixed value. By invoking
28: vector dominance and $\rho$ saturation, we also study the finite pion
29: density effects into the low mass dilepton production rate. We find
30: that the distribution moderately widens and the position of the peak
31: shifts toward larger values of the pair invariant mass, at the same
32: time that the height of the peak decreases when the value of the
33: chemical potential grows. We conclude by arguing that for the
34: description of the dilepton spectra at ultra-relativistic energies,
35: such as those of RHIC and LHC, the proper treatment of the large pion
36: density might be a more important effect to consider than the
37: influence of a finite baryon density. 
38: 
39: \end{abstract}
40: 
41: \pacs{11.10.Wx, 11.30.Rd, 11.55.Fv, 25.75.-q }
42: 
43: \maketitle
44: 
45: \section{Introduction}
46: 
47: One of the most salient features of the low-mass dilepton spectra in
48: relativistic nucleus-nucleus collisions, from BEVALAC/SIS to SPS
49: energies, is the enhancement in the production yields for invariant
50: masses between 0.2 and 1 GeV, as compared to proton induced
51: reactions~\cite{experiments}. Since dilepton final states are mediated
52: by electromagnetic currents and these in turn are connected to vector
53: mesons, dilepton pairs represent a prime tool to study the evolution
54: of the dense and hot hadronic region formed in this kind of
55: collisions. For low invariant masses, the vector mesons involved are the
56: $\rho$, $\omega$ and $\phi$. Among these, $\rho$ plays an special
57: role given that its lifetime is smaller than the expected lifetime
58: of the interacting region and thus is able to probe different
59: stages during the collision of heavy systems.  
60: 
61: The favored explanations, able to account for a great
62: deal of the features of the measured low-mass dilepton spectra can be
63: divided in two categories: the {\it dropping} $\rho$ mass and the {\it
64: melting} of resonances scenarios~\cite{Rapp}. The first of these,
65: connected to the Brwon-Rho scaling conjecture~\cite{Brown} and the
66: decrease of the quiral quark condensate with temperature and baryon
67: density, states that the in-medium $\rho$ mass will sweep the entire
68: low invariant mass region as the system cools down from its initially
69: hot and dense state toward freeze-out. The second
70: scenario~\cite{Dominguez, Asakawa} states that the in-medium spectral
71: densities of the $\rho$ and its chiral partner, the $a_1$ (1260)
72: become broad and structureless, merging into a flat continuum as the
73: system approaches chiral symmetry restoration. 
74: 
75: An important ingredient for the success of the two above mentioned
76: scenarios is the presence of finite baryon density effects (see for
77: example Ref.~\cite{Chen}). In the
78: first case, baryons act as the source of strongly attractive scalar
79: fields. In the second case, interactions of the vector mesons with
80: baryonic resonances, characterized by large coupling constants,
81: overwhelm the relative strength of the interactions of $\rho$'s with
82: pions, despite the fact that at SPS energies, the relative abundance
83: of the latter is five times that of the former.
84: 
85: Nevertheless, at RHIC and moreover at LHC energies, the central
86: rapidity region is expected to become baryon free with the relative
87: abundance of pions being larger than at SPS energies. Consequently, at
88: ultra-relativistic energies, it is important to include in the
89: calculation of the dilepton spectrum the proper treatment of the
90: large pion density, particularly during the hadronic phase of
91: the collision. 
92: 
93: Recall that strictly speaking, pion number is not a conserved
94: quantity and that pion decay is driven by all the relevant
95: interactions, namely, strong, weak and electromagnetic.
96: However, the characteristic time for electromagnetic
97: and weak pion number-changing reactions, is very large compared to the
98: lifetime of the system created in relativistic heavy-ion collisions
99: and therefore, these processes are of no relevance for the propagation
100: properties of pions within the lifetime of the collision. As for the
101: case of strong processes, it is by now accepted that they drive pion
102: number toward chemical freeze-out at a temperature considerably higher
103: than the thermal freeze-out temperature and therefore, that
104: from chemical to thermal freeze-out, the pion system evolves with the
105: pion abundance held fixed~\cite{Bebie, Braun-Munzinger}. Under these
106: circumstances, it is possible to ascribe to the pion density a
107: chemical potential and consider the pion number as
108: conserved~\cite{Hung,Chungsik}. In this context, the role of a finite
109: pion chemical potential into a hadronic equation of state has been recently
110: investigated in Refs.~\cite{Teaney}. The effects of a finite isospin
111: chemical potential on the pion mass have also been recently studied in
112: Refs.~\cite{Loewe}. 
113: 
114: The description of hadronic degrees of freedom belongs to the realm of
115: nonperturbative phenomena and therefore has necessarily to rely on
116: effective approaches that implement the dynamical symmetries of
117: QCD. In a series of recent papers~\cite{Ayala, Ayala2} it has been
118: shown that the linear sigma model can be used as one of such effective
119: approaches to describe the pion propagation properties within a pion
120: medium at energies, temperatures and densities small compared to
121: the sigma mass. The sigma degree of freedom can be integrated out in a
122: systematic expansion to obtain an effective theory of like-isospin
123: pions interacting among themselves through an effective quartic
124: term with coupling $\alpha = 6(m_\pi^2/2f_\pi^2)$, where $m_\pi$ and
125: $f_\pi$ are the vacuum pion mass and decay constant, respectively.
126: 
127: In this paper we extend the use of such effective description to
128: study the interaction of pions with the $\rho$ vector, paying
129: particular attention to the effects that a finite pion density
130: introduce on the propagation properties of $\rho$ at finite
131: temperature. We find the finite density and temperature modifications
132: to the $\rho$ mass, width and dispersion relation. By invoking vector
133: dominance, we also study the effects that these modifications
134: introduce on the production of $e^+$ $e^-$ pairs near the $\rho$ peak.
135: 
136: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
137: \begin{figure}[t!] % fig 1
138: {\centering
139: {\includegraphics{amfig1.ps}}
140: \par}
141: \vspace{-3.0in}
142: \caption{Feynman diagrams representing the pion self-energy at one
143: loop. The wavy lines represent the $\rho$ whereas the
144: solid lines represent the pion. In the approximation where $m_\rho \gg
145: m_\pi, T$, only diagram $(a)$ contributes.} 
146: \end{figure}
147: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
148: 
149: The work is organized as follows: In Sec.~\ref{secII}, we introduce
150: the $\rho$ in the description by {\it gauging} the original effective
151: Lagrangian. The theory thus obtained closely resembles scalar
152: electrodynamics. We introduce a chemical potential associated to a
153: conserved number of pions and construct the modifications to the pion
154: propagator and $\pi$-$\rho$ vertex at finite density. In
155: Sec.~\ref{secIII}, we use this effective Lagrangian to compute
156: the pion self-energy and in Sec.~\ref{secIV} the $\rho$ self-energy and
157: from it, the modifications to its mass, width 
158: and dispersion curve at finite density and temperature. In
159: Sec.~\ref{secV}, we compute the dilepton rate assuming vector
160: dominance. We finally summarize and conclude in Sec.~\ref{secV}.
161: 
162: \section{Effective Lagrangian}\label{secII}
163: 
164: In Refs.~\cite{Ayala} it has been shown that starting from the linear
165: sigma model Lagrangian, including only meson degrees of freedom and
166: working in the kinematical regime where the pion momentum, mass and
167: temperature are small compared to the sigma mass, the two-loop pion
168: self energy can be formally obtained by means of the effective
169: Lagrangian  
170: \be
171:    {\mathcal L}=\frac{1}{2}\left(\partial^\mu{\mathbf{\phi}}\right)^2
172:    -\frac{1}{2}m_\pi^2{\mathbf{\phi}}^2 -\frac{\alpha}{4!}
173:    {\mathbf{\phi}}^4\, ,
174:    \label{effLag}
175: \ee
176: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
177: \begin{figure}[t!] % fig 2
178: {\centering
179: {\includegraphics{amfig2.ps}}
180: \par}
181: \vspace{-4.8in}
182: \caption{Feynman diagrams representing the $\rho$ self-energy at one
183: loop. The wavy lines represent the $\rho$ whereas the solid lines
184: represent the pion.} 
185: \end{figure}
186: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
187: where $\alpha=6(m_\pi^2/2f_\pi^2)$ and the factor $6$ comes from
188: considering the interaction of like-isospin pions in the vertex
189: \be
190:    i\Gamma_4^{ijkl}=-2i\left(\frac{m_\pi^2}{2f_\pi^2}\right)
191:    \left(\delta^{ij}\delta^{kl}+\delta^{ik}\delta^{jl}+\delta^{il}\delta^{jk}
192:    \right)\, .
193:    \label{vertmod}
194: \ee
195: In essence, the theory thus constructed and summarized by the
196: effective Lagrangian in Eq.~(\ref{effLag}) can be thought of as a theory
197: for the effective coupling $\alpha$ that encodes the dynamics of low
198: energy pion interactions. It can also be checked that
199: Eq.~(\ref{effLag}) reproduces the leading order modification to the
200: pion mass at finite temperature obtained from chiral perturbation
201: theory~\cite{Gasser}. 
202: 
203: In order to introduce the $\rho$ field, we {\it gauge} the
204: theory~\cite{Gale, Pisarski} described by the Lagrangian in
205: Eq.~(\ref{effLag}) replacing the derivative $\partial^\mu$ by the
206: covariant derivative $D^\mu$ given by 
207: \be
208:    \partial^\mu \phi\rightarrow D^\mu\phi = (\partial^\mu
209:    -ig\rho^\mu)\phi\, ,
210:    \label{covder}
211: \ee
212: where we have introduced the $\pi$-$\rho$ coupling constant $g$. Also,
213: by introducing the mass term and kinetic energy for the $\rho$ field,
214: the Lagrangian in Eq.~(\ref{effLag}) becomes
215: \be
216:    {\mathcal L}\rightarrow {\mathcal L}' &=& 
217:    \frac{1}{2}(D^\mu\phi)^2
218:    -\frac{1}{2}m_\pi^2\phi^2 -\frac{\alpha}{4!}
219:    \phi^4\nonumber\\
220:    &+&\frac{1}{2}m_\rho^2\rho^\mu\rho_\mu
221:    -\frac{1}{4}\rho_{\mu\nu}\rho^{\mu\nu}\, .
222:    \label{effLagprim}
223: \ee
224: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
225: \begin{figure}[t] % fig 3
226: {\centering
227: \resizebox*{0.4\textwidth}{0.2\textheight}{\includegraphics{amfig3a.eps}}
228: \par}
229: \vspace{0.7cm}
230: {\centering 
231: \resizebox*{0.4\textwidth}{0.2\textheight}{\includegraphics{amfig3b.eps}}
232: \par}
233: \caption{Thermal pion mass as a function of $(a)$ $T$ for
234: different values of $\mu$ ranging 
235: from $\mu=0$ to $\mu=130$ MeV and as a function of $(b)$ $\mu$ for
236: different values of $T$ ranging from $T=50$ to $T=150$ MeV.}
237: \end{figure}
238: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
239: Let us now pause briefly to describe the formalism that allows us to
240: introduce a finite chemical potential associated to a conserved pion
241: number. To this end, let us further modify the Lagrangian in
242: Eq.~(\ref{effLagprim}), writing it in terms of a complex scalar field
243: and regarding $\phi$ and $\phi^*$ as independent fields
244: \be
245:    {\mathcal L}'\rightarrow {\mathcal L}'' &=& 
246:    (D_\mu\phi)(D^\mu\phi^*)
247:    -m_\pi^2\phi\phi^* -\frac{\alpha}{4}
248:    (\phi\phi^*)^2\nonumber\\
249:    &+&\frac{1}{2}m_\rho^2\rho^\mu\rho_\mu
250:    -\frac{1}{4}\rho_{\mu\nu}\rho^{\mu\nu}\, .
251:    \label{effLagdoubleprim}
252: \ee
253: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
254: \begin{figure}[t] % fig 4
255: {\centering
256: \resizebox*{0.4\textwidth}{0.2\textheight}{\includegraphics{amfig4a.eps}}
257: \par}
258: \vspace{0.7cm}
259: {\centering 
260: \resizebox*{0.4\textwidth}{0.2\textheight}{\includegraphics{amfig4b.eps}}
261: \par}
262: \caption{Thermal $\rho$ mass as a function of $(a)$ $T$ for
263: different values of $\mu$ ranging 
264: from $\mu=0$ to $\mu=130$ MeV and as a function of $(b)$ $\mu$ for
265: different values of $T$ ranging from $T=50$ to $T=150$ MeV.} 
266: \end{figure}
267: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
268: The effective Lagrangian in Eq.~(\ref{effLagdoubleprim}) resembles
269: that of scalar electrodynamics with the photon field replaced by the
270: massive $\rho$ field. Invariance under global phase transformations 
271: \be
272:    \phi\rightarrow\phi' = e^{-i\lambda}\phi
273:    \label{globalphase}
274: \ee
275: leads to the conserved current
276: \be
277:    J^\mu = i(\phi^*\partial^\mu\phi - \phi\ \partial^\mu\phi^*) -
278:    2g\rho^\mu\phi^*\phi\, ,
279:    \label{conscurr}
280: \ee
281: and to the conserved charge $N$, that can be identified with the particle
282: number, given by
283: \be
284:    N=i\int d^3x(\phi^*\partial^0\phi - \phi\ \partial^0\phi^* +
285:    2ig\rho^0\phi^*\phi)\, .
286:    \label{conscharge}
287: \ee 
288: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
289: \begin{figure}[t] % fig 5
290: {\centering
291: \resizebox*{0.4\textwidth}{0.2\textheight}{\includegraphics{amfig5a.eps}}
292: \par}
293: \vspace{0.7cm}
294: {\centering 
295: \resizebox*{0.4\textwidth}{0.2\textheight}{\includegraphics{amfig5b.eps}}
296: \par}
297: \caption{Thermal $\rho$ half width as a function of $(a)$ $T$ for
298: different values of $\mu$ ranging 
299: from $\mu=0$ to $\mu=130$ MeV and as a function of $(b)$ $\mu$ for
300: different values of $T$ ranging from $T=50$ to $T=150$ MeV.} 
301: \end{figure}
302: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
303: Since $N$ is a conserved charge, there exists a chemical
304: potential $\mu$ conjugate to $N$ so that the grand partition function
305: is
306: \be
307:    Z(\beta ,\mu)={\mbox {Tr}}\ e^{-\beta(H-\mu N)}\, ,
308:    \label{partfunc}
309: \ee
310: where $\beta = 1/T$ is the inverse of the temperature $T$. From now
311: on, let us work explicitly in the imaginary-time formalism of thermal
312: field theory. It is straightforward to write a path integral
313: representation in Euclidean space of the grand partition function
314: $Z(\beta ,\mu)$. Care has to be taken by going through the Hamiltonian
315: form since $N$ depends on the time derivative of $\phi$ and
316: $\phi^*$. After integration over the conjugate fields
317: $\pi^*=\partial{\mathcal L}''/\partial(\partial^0\phi^*)$ and
318: $\pi=\partial{\mathcal L}''/\partial(\partial^0\phi)$ one can check
319: that in the exponent, there appear the combinations
320: \be
321:     -\phi^*\left(\frac{\partial^2}{\partial\tau^2} -
322:     2\mu\frac{\partial}{\partial\tau} + \mu^2 -
323:     m_\pi^2\right)\phi\, ,
324:     \label{comb1}
325: \ee
326: \be
327:     ig\rho^0\left(\phi^*\frac{\partial\phi}{\partial\tau} -
328:     \phi\frac{\partial\phi^*}{\partial\tau}\right) -
329:     2ig\mu\rho^0\phi^*\phi\, ,
330:     \label{comb2}
331: \ee
332: where $\tau$ is the Euclidean time. Going to frequency space the
333: Euclidean time derivatives get replaced by
334: \be
335:    \frac{\partial\phi}{\partial\tau}&\rightarrow&
336:    -i\omega_n\phi\nonumber\\ 
337:    \frac{\partial\phi^*}{\partial\tau}&\rightarrow&
338:    i{\omega}_{n'}\phi^*\, , 
339:    \label{replace}
340: \ee
341: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
342: \begin{figure}[t!] % fig 6
343: {\centering
344: \resizebox*{0.44\textwidth}{0.22\textheight}{\includegraphics{amfig6.eps}}
345: \par}
346: \caption{$\rho$ dispersion relation for $(a)$ longitudinal and $(b)$
347: transverse modes for $\mu=100$ MeV for
348: different values of $T$ ranging from $T=50$ to $T=150$ MeV.} 
349: \end{figure}
350: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
351: where the periodicity of the fields in the interval $ 0\leq \tau \leq
352: \beta$ makes the (Matsubara) frequencies $\omega_n$ and ${\omega}_{n'}$
353: be discrete and integer multiples of $2\pi T$. The combination in
354: Eq.~(\ref{comb1}) translates into a modification of the Matsubara pion
355: propagator which now reads as (hereafter capital letters are used to
356: denote four-vectors whereas lower case letters are used do denote the
357: components)
358: \be
359:    \Delta(i\omega_n,p;\mu)=\frac{1}{-(i\omega_n +\mu)^2 + p^2
360:    +m_\pi^2}\, ,
361:    \label{prop}
362: \ee
363: whereas the combination in Eq.~(\ref{comb2}) goes into a modification
364: of the $\pi$-$\rho$ vertex which now becomes
365: \be
366:    \Gamma_{\pi\rho}(P_\mu, {P'}_\mu;\mu)&=&
367:    -ig\{[-(i\omega_n + \mu),{\mathbf p}]\nonumber\\
368:    &+& [-(i{\omega}_{n'} + \mu),{\mathbf p'}]\}\, , 
369:    \label{vert}
370: \ee
371: The final outcome is that the introduction of a finite chemical
372: potential translates into the substitution
373: \be
374:    i\omega_n\rightarrow i\omega_n + \mu
375:    \label{subst}
376: \ee
377: both when this frequency appears in internal pion lines either in the
378: Matsubara propagator or in the the $\pi$-$\rho$ vertex. Notice that
379: these substitutions agree with the weel known result for the
380: periodicity of the Matsubara propagator in the mixed representation
381: given by
382: \be
383:    \Delta (\beta-\tau,E;\mu)=\Delta (\tau,E;-\mu)\, .
384:    \label{Matsbound}
385: \ee
386: 
387: \section{$\pi$ self-energy}\label{secIII}
388: 
389: The effective Lagrangian in Eq.~(\ref{effLagdoubleprim}) describes the
390: interactions between pions and $\rho$'s as well as among pions
391: themselves. The one-loop diagrams for the pion and $\rho$ 
392: self-energies are depicted in Figs.~1 and~2. For the pion self-energy, the
393: diagrams contain terms with internal $\rho$ propagators. Notice that
394: since $m_\rho \gg m_\pi,\ T$, these diagrams are strongly suppressed
395: at finite temperature compared to those made out exclusively of
396: internal pion lines and can thus be neglected. This approximation goes
397: along the lines of the reasoning used in Refs.~\cite{Ayala} where the
398: heavy internal sigma modes were systematically {\it pinched} to
399: construct the effective vertices and propagators leading to the
400: effective Lagrangian in Eq.~(\ref{effLag}). In this scheme, the pion
401: and $\rho$ self-energies decouple and the former gets modified only by
402: the diagram in Fig.~1a leading to a momentum independent correction of
403: the pion mass. We can thus take one step further going beyond naive
404: perturbation theory and adopt a resummation scheme for the pion
405: self-energy to look for the modification of the pion mass beyond
406: leading order. This can be implemented by writing the pion self-energy
407: explicitly as 
408: \be
409:    \Pi_0=\frac{\alpha}{2}T\sum_n\int\frac{d^3k}{(2\pi)^3}
410:    \frac{1}{K^2+m_\pi^2+\Pi_0}\, ,
411:    \label{pi0}
412: \ee
413: 
414: Equation~(\ref{pi0}) represents a self consistent relation for the
415: temperature and density dependent quantity $\Pi_0$. This is the well known
416: resummation for the {\it superdaisy} diagrams which constitute the dominant
417: contribution in the large-$N$ expansion~\cite{Dolan} of the Lagrangian in
418: Eq.~(\ref{effLag}). The solution to Eq.~(\ref{pi0}) is given by the
419: transcendental equation 
420: \be
421:    \Pi_0&=&\left(\frac{\alpha T}{4\pi^2}\right)\sqrt{m_\pi^2 + \Pi_0}
422:    \nonumber\\
423:    &\times&\sum_{n=1}^\infty K_1\left(\frac{n\sqrt{m_\pi^2 + \Pi_0}}{T}\right)
424:    \frac{\cosh (n\mu /T)}{n}\, .
425:    \label{solpi0}
426: \ee
427: Figure~3 shows the behavior of the pion thermal mass
428: $\tilde{m}_\pi=\sqrt{m_\pi^2 + \Pi_0}$ as a function of $(a)$ $T$ for
429: different values of $\mu$ and $(b)$ as a function of $\mu$ for
430: different values of $T$. We use the values $m_\pi=135$ MeV, $f_\pi=93$
431: MeV. From Fig.~3, we notice that $\tilde{m}_\pi$ grows monotonically
432: with both $T$ and $\mu$. 
433: 
434: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
435: \begin{figure}[t!] % fig 7
436: {\centering
437: \resizebox*{0.44\textwidth}{0.22\textheight}{\includegraphics{amfig7.eps}}
438: \par}
439: \caption{Dilepton production rates for $(a)$ longitudinal and $(b)$
440: transverse modes for $T=150$ MeV, $k=50$ MeV and $\mu=0, 100$ MeV.} 
441: \end{figure}
442: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
443: 
444: \section{$\rho$ self-energy}\label{secIV}
445: 
446: The explicit expression for the one-loop $\rho$ self-energy depicted in
447: Fig.~2 is given by
448: \be
449:    \Pi^{\mu\nu}&=&-g^2T\sum_n\int\frac{d^3p}{(2\pi)^3}
450:    \frac{(2P^\mu-K^\mu)(2P^\nu-K^\nu)}
451:    {(P^2+\tilde{m}_\pi^2)[(K-P)^2+\tilde{m}_\pi^2]}\nonumber\\
452:    &+&\delta^{\mu\nu}g^2T\sum_n\int\frac{d^3p}{(2\pi)^3}
453:    \frac{1}{P^2+\tilde{m}_\pi^2}\, ,
454:    \label{rhoself}
455: \ee
456: where, according to the discussion in Sec.~\ref{secII}, the internal
457: pion momentum $P^\mu$ and the $\rho$ momentum $K^\mu$ are 
458: \be
459:    P^\mu&=&[-(i\omega_n +\mu), {\mathbf p}]\nonumber\\
460:    K^\mu&=&(-i\omega, {\mathbf k})\, .
461:    \label{vecdef}
462: \ee  
463: Equation~(\ref{rhoself}) contains vacuum and matter contributions. It is
464: well known that the infinities coming from the vacuum pieces can be
465: reabsorbed into the redefinition of the bare masses and
466: couplings~\cite{Gale}. For what follows, we will concentrate on
467: the matter contributions.
468: 
469: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
470: \begin{figure}[t!] % fig 8
471: {\centering
472: \resizebox*{0.44\textwidth}{0.22\textheight}{\includegraphics{amfig8.eps}}
473: \par}
474: \caption{Dilepton production rates for $(a)$ longitudinal and $(b)$
475: transverse modes for $T=150$ MeV, $k=250$ MeV and $\mu=0, 100$ MeV.} 
476: \end{figure}
477: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
478: 
479: For a massive vector field, the tensor structure of its self-energy
480: can be written in terms of the longitudinal $P_L^{\mu\nu}$ and
481: transverse $P_T^{\mu\nu}$ projection tensors
482: \be
483:    \Pi^{\mu\nu}=F(K)P_L^{\mu\nu}+G(K)P_T^{\mu\nu}\, .
484:    \label{proj}
485: \ee
486: In Minkowski space, $P_{T,L}^{\mu\nu}$ are given by
487: \be
488:    P_T^{00}&=&P_T^{0i}=P_T^{i0}=0\nonumber\\
489:    P_T^{ij}&=&\delta^{ij}-k^ik^j/k^2\nonumber\\
490:    P_L^{\mu\nu}&=&-g^{\mu\nu}+K^\mu K^\nu/K^2 - P_T^{\mu\nu}\, .
491:    \label{projLT}
492: \ee
493: From the relation between the self-energy, and the full $D^{\mu\nu}$
494: and bare $D^{\mu\nu}_0$ $\rho$ propagators, we have, also in Minkowski
495: space 
496: \be
497:    -iD^{\mu\nu}=\frac{P_L^{\mu\nu}}{K^2-m_\rho^2-F}+
498:    \frac{P_T^{\mu\nu}}{K^2-m_\rho^2-G} + 
499:    \frac{K^\mu K^\nu}{m_\rho^2K^2}.
500:    \label{proprho}
501: \ee
502: In order to identify the coefficients $F$ and $G$ of the longitudinal
503: and transverse projectors, we take ${\mathbf k}$ along the
504: $z$-axis. Thus, in Minkowski space, their expressions are
505: \be
506:    F(K)&=&-\frac{K^2}{k_0k}\Pi^{03}\nonumber\\
507:    G(K)&=&\Pi^{11}\, ,
508:    \label{FandG}
509: \ee
510: where $\Pi^{03}$ and $\Pi^{11}$ are obtained from Eq.~(\ref{rhoself})
511: with the analytical continuation 
512: \be
513:    i\omega\rightarrow k_0 + i\epsilon\, ,
514:    \label{analcont}
515: \ee
516: that give the retarded functions and that can be performed after carrying
517: out the sum over the Matsubara frequencies. 
518: 
519: The sum over frequencies can be obtained by standard
520: techniques~\cite{Kapusta, LeBellac}. Considering the effects of a
521: finite chemical potential, the expressions of interest are
522: \begin{widetext}
523: \be
524:    T\sum_n \Delta(i\omega_n,p)\Delta(i(\omega-\omega_n),|{\mathbf
525:    k}-{\mathbf p}|)&=&-
526:    \sum_{s_1s_2=\pm}\frac{s_1s_2}{4E_pE_{|{\mathbf k}-{\mathbf p}|}}
527:    \frac{[1+n(s_1E_p-\mu)+n(s_2E_{|{\mathbf k}-{\mathbf p}|}+\mu)]}
528:    {(i\omega - s_1E_p - s_2E_{|{\mathbf k}-{\mathbf p}|})}\nonumber\\
529:    T\sum_n i\omega_n\Delta(i\omega_n,p)\Delta(i(\omega-\omega_n),|{\mathbf
530:    k}-{\mathbf p}|)&=&-
531:    \sum_{s_1s_2=\pm}\frac{s_1s_2(s_1E_p-\mu)}
532:    {4E_pE_{|{\mathbf k}-{\mathbf p}|}}
533:    \frac{[1+n(s_1E_p-\mu)+n(s_2E_{|{\mathbf k}-{\mathbf p}|}+\mu)]}
534:    {(i\omega - s_1E_p - s_2E_{|{\mathbf k}-{\mathbf p}|})}\, ,
535:    \label{sumsexp}
536: \ee
537: \end{widetext}
538: where 
539: \be
540:    n(x)=\frac{1}{e^{\beta |x|}-1}\, ,
541:    \label{be}
542: \ee
543: is the Bose-Einstein distribution,
544: $E_p=\sqrt{p^2+\tilde{m}_\pi^2}$, $E_{|{\mathbf k}-{\mathbf
545: p}|}=\sqrt{({\mathbf k}-{\mathbf p})^2+\tilde{m}_\pi^2}$ and the
546: function $\Delta$ is defined in Eq.~(\ref{prop}).
547: 
548: Using Eq.~(\ref{sumsexp}) into Eq.~(\ref{rhoself}) and by means of the
549: analytical continuation in Eq.~(\ref{analcont}), the real and
550: imaginary parts of $\Pi^{03}$ and $\Pi^{11}$ are given by
551: \begin{widetext}
552: \be
553:    {\mbox {Re}}\Pi^{03}&=&\frac{g^2}{8\pi^2}\int_0^\infty
554:    dp\left(\frac{p^2}{E_p}\right)\left[n(E_p-\mu)+n(E_p+\mu)\right]
555:    \left[k_0(A_-+A_+)-2E_p(A_--A_+)\right]\nonumber\\
556:    {\mbox {Im}}\Pi^{03}&=&\frac{g^2}{16\pi}\left(\frac{k_0}{k^2}\right)
557:    \Big\{\int_{\frac{k_0-ak}{2}}^{\frac{k_0+ak}{2}}
558:    dE_p(2E_p-k_0)^2[1+2n(E_p-k_0-\mu)]\theta (k_0)\nonumber\\
559:    &-&
560:    \int_{\frac{-k_0-ak}{2}}^{\frac{-k_0+ak}{2}}
561:    dE_p(2E_p+k_0)^2[1+2n(E_p+k_0+\mu)]\theta (-k_0)\Big\}\nonumber\\
562:    {\mbox {Re}}\Pi^{11}&=&\frac{g^2}{8\pi^2}\int_0^\infty
563:    dp\left(\frac{p^2}{E_p}\right)\left[n(E_p-\mu)+n(E_p+\mu)\right]
564:    \left[\frac{2(k_0^2+k^2)}{k^2}+(B_-+B_+)\right]\nonumber\\
565:    {\mbox {Im}}\Pi^{11}&=&-\frac{g^2}{32\pi}\left(\frac{1}{k^3}\right)
566:    \Big\{\int_{\frac{k_0-ak}{2}}^{\frac{k_0+ak}{2}}
567:    dE_p[4p^2k^2-(k^2-k_0^2+2k_0E_p)^2]
568:    [1+2n(E_p-k_0-\mu)]\theta (k_0)\nonumber\\
569:    &-&
570:    \int_{\frac{-k_0-ak}{2}}^{\frac{-k_0+ak}{2}}
571:    dE_p[4p^2k^2-(k^2-k_0^2-2k_0E_p)^2]
572:    [1+2n(E_p+k_0+\mu)]\theta (-k_0)\Big\}\nonumber\\
573:    A_\pm&=&\frac{1}{2k^2p}\left(4kp + [k_0^2\pm2k_0E_p]
574:    \ln\left|\frac{k_0^2-k^2\pm2k_0E_p-2kp}{k_0^2-k^2\pm2k_0E_p+2kp}
575:    \right|\right)\nonumber\\
576:    B_\pm&=&\left(\frac{(k_0^2-k^2\pm 2k_0E_p)^2-4k^2p^2}{4k^3p}\right)
577:    \ln\left|\frac{k_0^2-k^2\pm2k_0E_p-2kp}{k_0^2-k^2\pm2k_0E_p+2kp}\right|\, ,
578:    \label{ABpm}
579: \ee
580: \end{widetext}
581: where the function $a$ is defined as
582: \be
583:    a=\sqrt{1-\frac{4\tilde{m}_\pi^2}{(k_0^2-k^2)}}\, .
584:    \label{funca}
585: \ee
586: It is easy to check that in the limit $\mu\rightarrow 0$,
587: Eqs.~(\ref{ABpm}) reduce to the corresponding expressions found in
588: Ref.~\cite{Gale}. 
589: 
590: Figure~4 shows the behavior of the $\rho$ thermal mass obtained as the
591: solution for $k_0$ of either
592: \be
593:    k_0^2- m_\rho^2 - {\mbox {Re}}\left\{\begin{array}{c}
594:                                  F(k_0,k=0)\\
595:                                  G(k_0,k=0)
596:                                  \end{array}\right\}=0\, ,
597:    \label{disprho}
598: \ee
599: as a function of $(a)$ $T$ for different values of $\mu$ and $(b)$ as
600: a function of $\mu$ for different values of $T$. We use the value
601: $g^2/4\pi=2.93$ as determined by the $\rho$ width in vacuum. For
602: $k=0$ there is no distinction between transverse and longitudinal
603: modes and thus both 
604: Eqs.~(\ref{disprho}) lead to the same solution. For every value of
605: $(T,\mu)$, the solution is computed with the corresponding value for
606: $\tilde{m}_\pi (T,\mu)$. From Fig.~4, we see that
607: the thermal $\rho$ mass grows monotonically with both $T$ and $\mu$
608: and that the growth is larger for larger values of $\mu$.
609: 
610: Figure~5 shows the behavior of the $\rho$ decay rate (or half width)
611: obtained from either
612: \be
613:    \Gamma = -\frac{{\mbox {Im}}\left\{\begin{array}{c}
614:                              F(m_\rho(T,\mu),k=0)\\
615:                              G(m_\rho(T,\mu),k=0)
616:                             \end{array}\right\}}{2m_\rho(T,\mu)}
617:    \label{halfwidth}
618: \ee
619: as a function of $(a)$ $T$ for different values of $\mu$ and $(b)$ as
620: a function of $\mu$ for different values of $T$. Again, for $k=0$
621: there is no distinction between 
622: transverse and longitudinal modes. This can be shown analytically from
623: the explicit expressions of ${\mbox {Im}} F$ and ${\mbox {Im}} G$ in
624: Eqs.~(\ref{ABpm}) which for $k=0$ yield
625: \be
626:    {\mbox {Im}} F(k_0,k=0) &=& {\mbox {Im}} G(k_0,k=0)\nonumber\\
627:    &=&-\left(\frac{g^2}{48\pi}\right)k_0^2
628:    \left(1-\frac{4\tilde{m}_\pi^2}{k_0^2}\right)^{3/2}\nonumber\\
629:    &\times&\left[1+2n(k_0/2+\mu)\right]\, .
630:    \label{IMFGk0}
631: \ee
632: For $\mu=0$, Eq.~(\ref{IMFGk0}) coincides with the corresponding
633: expression in Ref.~\cite{Gale}.
634: 
635: From Fig.~5a we see that for a given value of $\mu$ the $\rho$ width
636: increases monotonically with temperature. From Fig.~5b we
637: observe that the width reaches a minimum at a finite value of $\mu$
638: and that this value increases as the temperature increases. This
639: behavior can be understood if we recall that when the density increases
640: so does the pion mass and thus the phase space available for the decay
641: products of $\rho$ narrows. The situation changes when the density is
642: large enough so that the increase in the mass of the $\rho$ becomes
643: steeper than the growth in the pion mass, widening the phase space
644: available for the decay process. 
645: 
646: Figure~6 shows the dispersion relation for $(a)$ longitudinal and
647: $(b)$ transverse $\rho$ modes for different values of $T$ and $\mu=100$
648: MeV. The main difference between the curves in each set can be
649: attributed to the increase of the $\rho$  mass with both $\mu$ and $T$.
650: 
651: \section{Dilepton rate}\label{secV}
652: 
653: The electromagnetic current can be identified with the underlying quark
654: structure of hadrons. For invariant masses below the charm threshold,
655: this current can be decomposed as
656: \be
657:    j_\mu^{\ \mbox {\small {em}}}=\frac{2}{3}\bar{u}\gamma_\mu u
658:    -\frac{1}{3}\bar{d}\gamma_\mu d -\frac{1}{3}\bar{s}\gamma_\mu s\, .
659:    \label{jemquarks}
660: \ee
661: Using the $SU(3)$ quark content of hadrons, this current, being
662: vectorial in nature, can be in turn identified with the current
663: constructed out of the vector mesons $\rho$, $\omega$ and $\phi$.
664: \be
665:    j_\mu^{\ \mbox {\small {em}}}=j_\mu^\rho + j_\mu^\omega + j_\mu^\phi\, ,
666:    \label{jemvectors}
667: \ee
668: This is the well known conjecture named {\it vector dominance model}
669: (VDM). Since the pion electromagnetic form factor is is almost
670: totally dominated by the $\rho$ for invariant masses below 1~GeV, 
671: a simplified picture to study the decay of this electromagnetic
672: current into low mass dilepton pairs is to consider that the current
673: in Eq.~(\ref{jemvectors}) is totally dominated by the $\rho$. A
674: further simplification stems from considering the spectral density of
675: $\rho$ as a simple pole located at its peak which in turn dictates
676: that the coupling of $\rho$ to the electromagnetic current is
677: $em_\rho^2/g$.
678: 
679: Under the assumption of VDM and $\rho$ saturation, the expressions for
680: the thermal dilepton rates from longitudinal and transverse $\rho$
681: modes are given, respectively, by~\cite{Gale} 
682: \begin{widetext}
683: \be
684:    E_+E_-\frac{dR}{d^3p_+d^3p_-}&=&\frac{1}{(2\pi)^6}
685:    \left(\frac{e^4}{g^2}\right)
686:    \left(\frac{m_\rho^4}{M^4}\right)
687:    \left[q^2-\frac{({\mathbf q}\cdot{\mathbf k})^2}{k^2}\right]
688:    \frac{-{\mbox {Im}}F}{(M^2-m_\rho^2-{\mbox {Re}}F)^2 + {\mbox
689:    {Im}}F^2}\left(\frac{1}{e^{\beta\omega_L}-1}\right),\nonumber\\
690:    E_+E_-\frac{dR}{d^3p_+d^3p_-}&=&\frac{1}{(2\pi)^6}
691:    \left(\frac{e^4}{g^2}\right)
692:    \left(\frac{m_\rho^4}{M^4}\right)
693:    \left[2M^2-q^2+\frac{({\mathbf q}\cdot{\mathbf k})^2}{k^2}\right]
694:    \frac{-{\mbox {Im}}G}{(M^2-m_\rho^2-{\mbox {Re}}G)^2 + {\mbox
695:    {Im}}G^2}\left(\frac{1}{e^{\beta\omega_T}-1}\right),
696:    \label{dileptonrate}
697: \ee
698: \end{widetext}
699: where $P^\mu_+=(E_+,{\mathbf p}_+)$ and $P^\mu_-=(E_-,{\mathbf p}_-)$
700: are the positron and electron four momenta, respectively,
701: $K^\mu=P^\mu_++P^\mu_-$, $Q^\mu=P^\mu_+-P^\mu_-$, $M^2=k_0^2-k^2$ is
702: the pair invariant mass, $\omega_{L,T}$ are the longitudinal and
703: transverse $\rho$ modes dispersion relations and we have neglected the
704: electron mass. 
705: 
706: Figures~7 and~8 show the dilepton production rates as functions of
707: the pair invariant mass $M$ for ${\mathbf q}\cdot{\mathbf k}=0$ for
708: fixed values of $T=150$ MeV and $k$ for two values of $\mu=0, 100$ MeV
709: for $(a)$ longitudinal and $(b)$ transverse modes. Figure~7 considers
710: an small value of $k=50$ MeV and Fig.~8 a larger value $k=250$ MeV. In
711: both cases we can see that the effect of the finite chemical potential is
712: to moderately widen the distribution and to (also moderately) displace
713: the position of the peak toward larger values of $M$. The most
714: significant effect is however the lowering of the peak for finite $\mu$
715: which is in agreement with the analysis of Ref.~\cite{Rapp2}.   
716:     
717: \section{Summary and conclusions}\label{secVI}
718: 
719: In this paper we have considered the effects of a finite pion density
720: on the propagation properties of $\rho$ mesons at finite
721: temperature. The $\rho$ has been introduced by gauging an effective
722: Lagrangian obtained from the linear sigma model in the kinematical
723: regime where the pion mass and temperature are small compared to the
724: sigma mass. The finite density is described in terms of a finite pion
725: chemical potential associated to a conserved pion number. We have argued
726: that such description is important for ultra relativistic heavy-ion
727: collisions at RHIC and LCH energies where the central rapidity region
728: is expected to become baryon free. In this situations, we expect that
729: the influence of the dropping $\rho$ mass or melting or resonances
730: scenarios to describe the dilepton spectra, which rely on the
731: effects of a finite baryon density, become less important than the
732: effects of the expected large pion density. This is so particularly
733: during the hadronic phase of the reaction, between chemical and
734: thermal freeze-out when the strong interaction drives pion-number to a
735: fixed value.  
736: 
737: We have found that the $\rho$ thermal mass increases monotonically
738: with both the temperature and the pion density. However, the $\rho$
739: width as a function of $\mu$ and for fixed $T$ starts off by decreasing,
740: reaching a minimum at a finite value of $\mu$. This behavior can be
741: understood by noticing that as the density increases, the pion mass
742: does too and the phase space available for the decay process
743: $\rho\rightarrow \pi \pi$ narrows up to a --temperature dependent-- value
744: of $\mu$. From this value, the increase in the thermal $\rho$ mass is
745: stronger than the increase in the thermal pion mass and this effect
746: produces a widening of the phase space for the process, thus producing
747: the $\rho$ width to rise.  
748: 
749: Under the assumption of VDM and $\rho$ saturation, we have also
750: computed the dilepton production rate as a function of the pair
751: invariant mass. We have found that the finite pion density produces
752: a moderate broadening of the distribution and a moderate increase of
753: the position of the peak. The finite pion density
754: also produces a decrease of the distribution at the position of the
755: peak compared to the $\mu=0$ case.
756: 
757: The problem posed by the renormalization of theories
758: involving resummation is by no means a simple one. In fact, in recent
759: years this subject has been much actively pursued in the
760: literature (see for example Refs.~\cite{Hees, Caldas, Blaizot}). The
761: consensus is that when the theory is vacuum renormalizable, it 
762: will still be renormalizable after resummation. The solution, which
763: can be formulated using different languages, has been shown to require
764: that the counterterms needed for vacuum renormalization in ordinary
765: perturbation theory need also to receive the benefits of resummation
766: and be considered self-consistently in the resummation procedure.
767: 
768: For the purposes of our work, let us stress that the resummation we
769: have implemented for the pion self-energy corresponds to the tadpole
770: approximation (see Sec III of Ref.~\cite{Hees}). In fact, the gap
771: equation that renders the finite temperature and density corrections
772: to the pion mass in our work, Eq.~(\ref{solpi0}), is 
773: identical to the thermal part of Eq. (24) of the above reference; we
774: have just gone one step further, expanding the integrand as a series
775: which allows us to analytically integrate each term. In Eq. (24) of
776: Ref.~\cite{Hees}, the renormalization of the vacuum self energy and
777: four-point function have been carried out on the physical mass-shell
778: condition and have taken into account the self-consistency. In the
779: language of this reference, this is so because in the tadpole
780: approximation, the renormalized four-point function is constant and
781: given just in terms of the original coupling constant on the
782: mass-shell condition. 
783: 
784: Had we gone beyond the tadpole approximation, the situation would
785: have not be that straightforward. In fact, it has also been shown in
786: Ref.~\cite{Hees} that the self-consistent resummation scheme requires
787: a more complicated renormalization of the four-point function to be
788: taken into account alongside the renormalization procedure for the
789: self-energy. The effects of this certainly more complete analysis
790: might be interesting with regard to further thermal modifications to
791: the pion mass in situations where the original coupling constant of
792: the $\phi^4$ theory is much larger that one but is certainly beyond the
793: scope of the present work where our coupling constant $\alpha$, as
794: determined by the scale of the interactions, set by the vacuum pion
795: mass, is of order 1 (in fact, what matters, as shown in
796: Eq.~(\ref{solpi0}) is the effective coupling constant given by
797: $\alpha/(4\pi^2)$ which is of order 0.1). 
798: 
799: Finally, in order to predict the final dilepton spectra at RHIC and LHC
800: energies, the results found in this work have to be placed into a
801: model for the evolution of the collision and also possibly to include
802: the effects of other vector or axial vector mesons. All this is for
803: future. 
804:  
805: \section*{Acknowledgments}
806: 
807: A. Ayala wishes to thank Centro Brasileiro de Pes\-qui\-sas Fi\-si\-cas for
808: their kind hospitality during the time when part of this work was done.
809: Support for this work has been received in part by DGAPA-UNAM under PAPIIT
810: grant number IN108001 and CONACyT under grant numbers 32279-E and 40025-F.
811: 
812: \begin{thebibliography} {20}
813: 
814: \bibitem{experiments}
815: G. Roche et al., (DLS Collaboration) Phys. Lett. {\bf B226}, 228
816: (1989); S. Bedoe et al., (DLS Collaboration) \prc {\bf 47}, 2840
817: (1993); G. Agakichiev et al., (CERES Col\-lab\-o\-ra\-tion), \prl {\bf 75},
818: 1272 (1995); D. Adamov\'a et al., (CERES/NA45 Collaboration), {\it Enhanced
819: production of low mass electron pairs in 40 AGeV Pb-Au collisions at
820: the CERN-SPS}, nucl-ex/0209024. 
821: 
822: \bibitem{Rapp}
823: For a comprehensive review of the theoretical and experimental status
824: on the subject see R. Rapp and J. Wambach, Adv. Nucl. Phys. {\bf 25},
825: 1 (2000).
826: 
827: \bibitem{Brown}
828: G.E. Brown and M. Rho, \prl {\bf 66}, 2720 (1991).
829: 
830: \bibitem{Dominguez}
831: C. Dominguez and M. Loewe, Z. Phys. C {\bf 49}, 423 (1991).
832: 
833: \bibitem{Asakawa}
834: M. Asakawa, C.M. Ko, P. L\'evai and X.J. Qiu, \prc {\bf 46}, R1159
835: (1992). 
836: 
837: \bibitem{Chen}
838: J. Chen, J. Li and P. Zhuang, JHEP, {\bf 0211}, 14 (2002). 
839: 
840: \bibitem{Bebie}
841: H. Bebie, P. Gerber, J.L. Goity and H. Leutwyler, Nucl. Phys. {\bf
842: B378}, 95 (1992). 
843: 
844: \bibitem{Braun-Munzinger}
845: P. Braun-Munzinger, J. Stachel, J. Wessels, and N. Xu, Phys. Lett. 
846: {\bf B344}, 43 (1995), {\it ibid} {\bf 365}, 1 (1996).
847: 
848: \bibitem{Hung}
849: C.M. Hung and E. Shuryak, Phys. Rev. C {\bf 57}, 1891 (1998).
850: 
851: \bibitem{Chungsik}
852: C. Song and V. Koch, Phys. Rev. C {\bf 55}, 3026 (1997). 
853: 
854: \bibitem{Teaney}
855: D. Teaney, {\it Chemical freezeout in heavy ion collisions},
856: nucl-th/0204023 (2002); T.Hirano and K. Tsuda, \prc {\bf 66}, 054905
857: (2002). 
858: 
859: \bibitem{Loewe}
860: M. Loewe and C. Villavicencio, {\it Thermal pions at finite density},
861: hep-ph/0206294, (2002); {\it Thermal pions at finite isospin chemical
862: potential}, hep-ph/0212275.
863: 
864: \bibitem{Ayala}
865: A. Ayala, S. Sahu and M. Napsuciale, Phys. Lett. {\bf B 479}, 156
866: (2000); A. Ayala and S. Sahu, Phys. Rev. D {\bf 62}, 056007 (2000). 
867: 
868: \bibitem{Ayala2}
869: A. Ayala, P. Amore and A. Aranda, \prc {\bf 66}, 045205 (2002).
870: 
871: \bibitem{Gasser}
872: J. Gasser and H. Leutwyler, Phys. Lett. {\bf B184}, 83 (1987).
873: 
874: \bibitem{Gale}
875: C. Gale and J. Kapusta, Nucl. Phys. {\bf B357}, 65 (1991).
876: 
877: \bibitem{Pisarski}
878: R.D. Pisarski, \prd {\bf 52}, R3773 (1995).
879: 
880: \bibitem{Dolan}
881: L. Dolan and R. Jackiw, Phys. Rev. D {\bf 9}, 3320 (1974);
882: S. Weinberg, {\it ibid} {\bf 9}, 3357 (1974).
883: 
884: \bibitem{Kapusta}
885: J.I. Kapusta, {\it Finite Temperature Field Theory} (Cambridge
886: University Press, Cambridge, 1989).
887: 
888: \bibitem{LeBellac}
889: M. Le Bellac, {\it Thermal Field Theory} (Cambridge University Press,
890: Cambridge, 1996).
891: 
892: \bibitem{Rapp2} 
893: R. Rapp and C. Gale, \prc {\bf 60}, 024903 (1999).
894: 
895: \bibitem{Hees}
896: H. van Hees and J. Knoll, \prd {\bf 65}, 105005 (2002).
897: 
898: \bibitem{Caldas}
899: H.C.G. Caldas, \prd {\bf 65}, 065005 (2002).
900: 
901: \bibitem{Blaizot}
902: J.-P. Blaizot, E. Iancu and U. Reinosa, {\it Renormalizability of
903: $\Phi$ derivable approximations in scalar $\phi^4$ theory},
904: hep-ph/0301201.  
905: 
906: 
907: \end{thebibliography}
908: 
909: \end{document}
910: 
911: 
912: 
913: 
914: 
915: