1: \documentclass[12pt,twoside,fleqn]{article}
2: \usepackage{epsfig}
3: \usepackage{cite}
4: %\usepackage[dvips]{rotating}
5: %\usepackage{multirow}
6: %\usepackage{exscale}
7: %\usepackage{amsbsy}
8: %\usepackage{amssymb}
9: %\usepackage[dvips]{color}
10: % put your own definitions here:
11: \def\lsim{\raise0.3ex\hbox{$<$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}}
12: \def\gsim{\raise0.3ex\hbox{$>$\kern-0.75em\raise-1.1ex\hbox{$\sim$}}}
13: %-----------------------------------------------------------------------
14: % The lines below are necessary in order to enumerate the equations
15: % according to the sections where they are.
16: \makeatletter
17: \@addtoreset{equation}{section}
18: \@addtoreset{figure}{section}
19: \makeatother
20: %\renewcommand{\theequation}{\arabic{equation}}
21: %\renewcommand{\thetable}{\arabic{table}}
22: %\renewcommand{\thefigure}{\arabic{figure}}
23: \renewcommand{\theequation}{\thesection.\arabic{equation}}
24: \renewcommand{\thetable}{\thesection.\arabic{table}}
25: \renewcommand{\thefigure}{\thesection.\arabic{figure}}
26: %-----------------------------------------------------------------------
27: \setlength{\parskip}{2ex}
28: \setlength{\textwidth}{16cm}
29: \setlength{\textheight}{22.5cm}
30: \setlength{\oddsidemargin}{0.0cm}
31: \setlength{\evensidemargin}{0.0cm}
32: \setlength{\topmargin}{-0.5cm}
33: % this causes footnotes to be numbered by letters rather than numbers
34: \renewcommand{\thefootnote}{\alph{footnote}}
35: %
36: \renewcommand{\arraystretch}{1.25}
37: \arraycolsep3mm
38: %
39: \newcommand{\bq}{\begin{equation}}
40: \newcommand{\eq}{\end{equation}}
41: \newcommand{\bqa}{\begin{eqnarray}}
42: \newcommand{\eqa}{\end{eqnarray}}
43: \newcommand{\bqas}{\begin{eqnarray*}}
44: \newcommand{\eqas}{\end{eqnarray*}}
45: \newcommand{\bdm}{\begin{displaymath}}
46: \newcommand{\edm}{\end{displaymath}}
47: \newcommand{\Ns}{N_{\sigma}}
48: \newcommand{\Nt}{N_{\tau}}
49: \newcommand{\Tr}{ {}{\rm Tr}{} }
50: \newcommand{\tr}{\mbox{Tr~}}
51: \newcommand{\re}{\mbox{Re~}}
52: \newcommand{\nn}{\nonumber}
53: \newcommand{\bfmu}{\hbox{\boldmath $\mu$\unboldmath}}
54: \newcommand{\bfmut}{\hbox{\boldmath $\mu^2$\unboldmath}}
55: %
56: \unitlength0.6cm
57: %
58: \begin{document}
59: \thispagestyle{empty}
60: %
61: \mbox{} \hfill BI-TP 2005/02\\[0mm]
62: \mbox{} \hfill SWAT/05/422\\
63: \begin{center}
64: {\Huge \bf
65: Thermodynamics of Two Flavor \\
66: QCD to Sixth Order in \\[1mm]
67: Quark Chemical Potential}
68:
69: %
70: \vspace*{1.0cm}
71: {\large
72: \bf C.R.~Allton\rlap,$^{\rm a}$ M. D\"oring\rlap,$^{\rm b}$
73: S.~Ejiri\rlap,$^{\rm b}$
74: S.J.~Hands\rlap,$^{\rm a}$ O.~Kaczmarek\rlap,$^{\rm b}$
75: F.~Karsch\rlap,$^{\rm b}$
76: E.~Laermann\rlap,$^{\rm b}$ and
77: K. Redlich$^{\rm b,c}$
78: }
79: %
80: \vspace*{0.6cm}
81:
82: {\normalsize
83: $^{\rm a}$ Department of Physics, University of Wales Swansea,
84: Singleton Park,\\ Swansea SA2 8PP, U.K. \\
85: $^{\rm b}$ Fakult\"at f\"ur Physik,
86: Universit\"at Bielefeld, D-33615 Bielefeld, Germany. \\
87: $^{\rm c}$ Institute of Theoretical Physics, University of Wroclaw,\\
88: PL-50204 Wroclaw, Poland.\\
89: }
90: %
91: \end{center}
92: %\setlength{\baselineskip}{1.3\baselineskip}
93: \vspace*{0.6cm}
94: \centerline{\large ABSTRACT}
95: \baselineskip 20pt
96: \noindent
97: We present results of a simulation of two flavor QCD on a $16^3\times4$
98: lattice using p4-improved staggered fermions with bare quark mass
99: $m/T=0.4$. Derivatives of the thermodynamic grand canonical partition
100: function $Z(V,T,\mu_u,\mu_d)$ with respect to
101: chemical potentials $\mu_{u,d}$ for different quark flavors are calculated
102: up to sixth order, enabling estimates of
103: the pressure and the quark number density as well as the chiral condensate
104: and various
105: susceptibilities as functions of $\mu_q = (\mu_u +\mu_d)/2$
106: via Taylor series expansion. Furthermore, we analyze baryon as well as
107: isospin fluctuations and discuss the
108: relation between the radius of convergence of the Taylor series
109: and the chiral critical point in the QCD phase diagram. We argue that
110: bulk thermodynamic observables do not, at present, provide direct
111: evidence for the existence of a chiral critical point in the QCD
112: phase diagram.
113: Results are compared to high temperature perturbation theory as well
114: as a hadron resonance gas model.
115:
116: \vfill
117: \noindent
118: %\mbox{}January 2005\\
119: \mbox{}PACS numbers: 12.38.Gc, 12.38.Mh\\
120: %\eject
121: \baselineskip 15pt
122:
123:
124: \section{Introduction}
125: \label{sec:intro}
126:
127: The thermodynamics of strongly interacting matter has been studied
128: extensively in lattice calculations at vanishing quark chemical
129: potential, $\mu_q$ (or baryon $\mu_B\equiv 3\mu_q$) \cite{review}.
130: Current lattice calculations strongly suggest that the transition from
131: the hadronic low temperature phase to the high temperature phase is a
132: continuous, non-singular but rapid transition happening in a narrow temperature
133: interval around the transition temperature $T_0\simeq 170$~MeV.
134: Recent advances in the development of techniques for lattice calculations
135: at non-zero quark chemical potential $\mu_q$
136: \cite{Fodor1,us1,Gavai1,Lombardo1,Philipsen1} have also enabled the first
137: exploratory studies of the QCD phase diagram and bulk thermodynamics in a regime
138: of small $\mu_q$, {\it i.e.} for $\mu_q/T\ \lsim \ 1$ and $T\ \gsim \ 0.8\ T_0$.
139:
140: Guided by phenomenological models which suggest that
141: at low temperature and non-zero quark chemical potential
142: the low and high density
143: regions will be separated by a first order phase transition
144: it has been speculated \cite{Stephanov}
145: that a $2^{nd}$ order phase transition point,
146: the chiral critical point, exists in the interior of the QCD phase diagram,
147: at which the line of first order transitions ends. For smaller
148: values of $\mu_q/T$ the low and high temperature regime will then be only
149: separated by a crossover transition. The first exploratory studies at
150: non-zero quark chemical potential indeed gave evidence for such a chiral
151: critical point \cite{Fodor1}, although subsequent investigations made clear that at
152: present any quantitative statement about the location \cite{Where,Fodor2}
153: and maybe even about the existence of such a chiral critical point is premature.
154: The first calculations of the baryonic contribution to the pressure in
155: strongly interacting matter
156: \cite{us1,FKS,us2,Fodoreos,Gavaieos,Lombardoeos}
157: also suggested that the transition and the
158: thermodynamic behaviour in the high-T phase resemble the picture found
159: previously at zero net baryon density. In this case ($\mu_q=0$) the
160: transition occurs over a narrow temperature interval. Beyond a region
161: $T \in [T_0,1.5 T_0]$ where there are large deviations from ideal gas
162: behaviour, thermodynamic observables rapidly approach the high-T ideal
163: gas
164: limit; for instance, the pressure agrees with the Stefan-Boltzmann
165: prediction to within $\sim 20\%$.
166: In the low temperature hadronic phase it has been found that a hadron resonance
167: gas model provides an astonishingly good description of basic features of the
168: $T$ and $\mu_q$-dependence of thermodynamic observables \cite{KRT}.
169:
170: Maybe one of the largest changes compared to the thermodynamics at
171: $\mu_q\equiv 0$ has been observed in the temperature dependence of the
172: quark number and isovector susceptibilities \cite{us2}. For $\mu_q=0$
173: these observables show a similar temperature dependence. They have been found
174: to change rapidly at the transition temperature but continue to increase
175: monotonically at larger temperatures
176: \cite{Gottlieb,suscept,Gupta02,suscept2,Gavai04}.
177: For $\mu_q > 0$, however,
178: the quark number susceptibility develops a pronounced peak at the transition
179: temperature while the isovector susceptibility continues to show a temperature
180: dependence similar to that found at $\mu_q = 0$. Such a behavior, indeed, is
181: expected to occur in QCD in the vicinity of a $2^{nd}$ order phase transition
182: point \cite{Kunihiro}. Susceptibilities thus may provide the most direct
183: evidence for the existence of a $2^{nd}$ order phase transition in the QCD
184: phase diagram. Finding the characteristic volume and/or quark mass dependent
185: universal scaling behavior\footnote{The $2^{nd}$ order critical point in the
186: QCD phase diagram is expected to belong to the universality class of the
187: 3-dimensional Ising model.} of susceptibilities would undoubtedly establish the
188: existence of a chiral critical point.
189:
190: In this paper we want to extend our previous study of thermodynamics at
191: non-zero quark chemical potential \cite{us1,us2}, which is based on a
192: Taylor series expansion around $\mu_q=0$, to the $6^{th}$
193: order\footnote{Some results on the radius of convergence
194: have been reported recently in an $8^{th}$ order Taylor expansion for two flavor
195: QCD \cite{Gavai04}.}. Going to higher orders in the expansion is of
196: particular importance for the analysis of higher derivatives of the
197: density of the grand potential expressed in units of the
198: temperature\footnote{We will call this in the following the grand potential
199: although conventionally the extensive quantity $VT^4\Omega$ is called the
200: grand potential.}, $\Omega (V,T,\mu)\equiv (V T^3)^{-1}\ln Z$, {\it i.e.}
201: for an analysis of generalized susceptibilities. Accordingly our main
202: emphasis will be to further analyze the properties of quark number and
203: isovector susceptibilities, relate them to diagonal and non-diagonal
204: flavor susceptibilities, calculate the chiral susceptibility and discuss to
205: what extent the pronounced peaks
206: found in some of these susceptibilities give evidence for the existence of a
207: chiral critical point in the QCD phase diagram.
208:
209: This paper is organized as follows. We start in the next section by summarizing
210: basic results on QCD thermodynamics at non-zero quark chemical potential
211: obtained in high temperature perturbation theory \cite{Blaizot,Vuorinen}
212: and expectations based
213: on properties of the hadron resonance gas model at low temperature
214: \cite{Hagedorn,pbm}.
215: In section 3 we present results on the calculation of various
216: thermodynamic observables obtained from a Taylor expansion up to $6^{th}$
217: order in $\mu_q/T$. In section 4 we analyze the convergence properties
218: of the Taylor series. Section 5 is devoted to a discussion of the reweighting
219: approach to QCD thermodynamics at $\mu_q \ne 0$ and its comparison to the
220: Taylor expansion approach. In section 6 we give our conclusions. An Appendix
221: contains details on the Taylor expansion of various observables studied here.
222:
223:
224: \section{Thermodynamics at low and high temperature}
225: \label{sec:hrgm}
226:
227: In this section we want to briefly discuss the
228: density or chemical potential dependence of thermodynamic
229: observables in the asymptotic high temperature regime as well as at
230: low temperatures in the hadronic phase of QCD. On the former we have information
231: from high temperature perturbation theory which
232: recently has been extended to ${\cal O}(g^6\ln g)$ \cite{Vuorinen}
233: also for non-vanishing quark chemical potentials $\mu_f$, $f=u,~d,..$, and
234: as for $\mu_f\equiv 0$ \cite{Kajantie} is thus now known to all
235: perturbatively calculable orders. Although a comparison
236: of the perturbative expansion with lattice calculations at $\mu_f =0$
237: suggests that quantitative agreement cannot be expected at temperatures
238: close to the transition temperature, $T_0$, we can gain
239: useful insight into the structure of the Taylor expansion used in
240: lattice calculations to study thermodynamics at $\mu_f \ne 0$.
241:
242: In the low temperature hadronic phase a systematic QCD based analysis
243: is difficult and one generally has to rely on calculations
244: within the framework of effective low energy theories or to
245: phenomenological approaches.
246: Here we will focus on a discussion of the properties of a hadron
247: resonance gas model, which recently has been compared to lattice results
248: at non-zero quark chemical potential quite successfully \cite{KRT}
249: and which also is known to describe experimental results on the
250: chemical freeze-out of particle ratios observed in heavy ion collisions
251: rather well \cite{pbm}.
252:
253: \subsection{High temperature perturbation theory}
254: \label{subsec:pert}
255:
256: In the infinite temperature limit the grand potential of QCD\footnote{We
257: suppress here the volume dependence of $\Omega$. Perturbative
258: calculations are performed in the thermodynamic limit. The volume
259: dependence, however, has to be analyzed more carefully in lattice
260: calculations.},
261: $\Omega (T,\mu) \equiv \ln {\cal Z}(V,T,\mu) /VT^3$, which is equivalent
262: to the pressure in units of $T^4$, approaches that of a free
263: quark-gluon gas (Stefan-Boltzmann (SB) gas),
264: \begin{eqnarray}
265: \frac{p_{SB}}{T^4} = \Omega^{(0)} (T,\mu) = \frac{8 \pi^2}{45} +
266: \sum_{f=u,d,..} \left[\frac{7 \pi^2}{60} +
267: \frac{1}{2} \left(\frac{\mu_f}{T}\right)^2
268: + \frac{1}{4 \pi^2} \left(\frac{\mu_f}{T}\right)^4
269: \right] \quad ,
270: \label{eq:free}
271: \end{eqnarray}
272: where the first term gives the contribution of the gluon sector and
273: the sum over the fermion sector extends over $n_f$ different flavors.
274: In Eq.~(\ref{eq:free}) we only gave the result for massless quarks and
275: gluons. Also in the following we will restrict our discussion of
276: perturbative results to the case of QCD with massless quarks.
277: In this section we also use $\mu$ to denote the entire set of $n_f$ different
278: chemical potentials $(\mu_u,~\mu_d,...)$. We also introduce the shorthand
279: notation $\bfmut \equiv \sum_{f} \left(\mu_f / T\right)^2$.
280:
281: The additive structure of contributions arising from gluons
282: and the different fermion flavor sectors persists at ${\cal O}(g^2)$.
283: Only starting at ${\cal O}(g^3)$ is a coupling among the different partonic
284: sectors introduced. At this order it is induced through the non-vanishing
285: electric (Debye) mass term $m_E$
286: \cite{Vuorinen},
287: \begin{eqnarray}
288: \Omega (T,\mu) = \Omega^{(0)} (T,\mu) + g^2\; \Omega^{(2)} (T,\mu)
289: + g^3\; \Omega^{(3)} (T,\mu) + {\cal O}(g^4)\quad ,
290: \end{eqnarray}
291: with
292: \begin{eqnarray}
293: \Omega^{(2)} (T,\mu)
294: \hspace*{-0.2cm}&=&\hspace*{-0.2cm}
295: -\left( \frac{1}{6} +
296: \frac{5 n_f}{72} + \frac{1}{4 \pi^2} \bfmut
297: +\frac{1}{8 \pi^4}\sum_{f=u,d,..} \left(\frac{\mu_f}{T}\right)^4
298: \right) \quad , \nonumber \\
299: \Omega^{(3)} (T,\mu)
300: \hspace*{-0.2cm}&=&\hspace*{-0.2cm} \frac{1}{6 \pi} \left(
301: \frac{m_E}{gT}\right)^3 =
302: \frac{1}{6 \pi} \left( 1 +\frac{n_f}{6} + \frac{1}{2 \pi^2}
303: \bfmut \right)^{3/2} \; .
304: \label{pg3}
305: \end{eqnarray}
306: The electric mass term introduces
307: a dependence of a given quark flavor sector on changes in another sector,
308: {\it i.e.} the quark number density in a flavor sector $\ell$,
309: \begin{equation}
310: \frac{n_\ell}{T^3} = \frac{\partial \Omega (T,\mu)}{\partial \mu_\ell/T} \quad ,
311: \end{equation}
312: depends on the other quark chemical potentials only at ${\cal O}(g^3)$.
313: This is also reflected
314: in the structure of diagonal and non-diagonal susceptibilities
315: \cite{Gavai1,Gottlieb,suscept,Gupta02,suscept2,Gavai04}.
316: \begin{equation}
317: \frac{\chi_{ff}(T,\mu)}{T^2} =
318: \frac{\partial^{2} \Omega (T,\mu)}{\partial (\mu_f/T)^2} \quad ,\quad
319: \frac{\chi_{fk}(T,\mu)}{T^2} =
320: \frac{\partial^{2} \Omega (T,\mu)}{\partial (\mu_f/T)
321: \partial (\mu_k/T)} \quad .
322: \label{sus}
323: \end{equation}
324: The diagonal susceptibilities are non-zero in the ideal gas
325: limit and, moreover, the leading order perturbative term stays non-zero
326: also in the limit of vanishing quark chemical potential,
327: \begin{eqnarray}
328: \frac{\chi_{ff}(T,\mu)}{T^2} \hspace{-0.1cm}&=&\hspace{-0.1cm}
329: 1 + \frac{3}{\pi^2}\left(\frac{\mu_f}{T}\right)^2
330: + {\cal O}(g^2) \quad .
331: \label{sus_lead_d}
332: \end{eqnarray}
333: The non-diagonal susceptibilities, however, receive non-zero contributions
334: only at ${\cal O}(g^3)$. The leading perturbative
335: contribution is positive and inversely proportional to the electric
336: screening mass. However, it vanishes in the limit of vanishing
337: chemical potentials. In this case the first non-zero contribution
338: arises at ${\cal O}(g^6 \ln 1/g)$ \cite{Blaizot},
339: \begin{eqnarray}
340: \frac{\chi_{fk}(T,\mu)}{T^2} \hspace{-0.1cm}&=&\hspace{-0.1cm}
341: \frac{g^3}{2 \pi^5} \left( 1 +\frac{n_f}{6} + \frac{1}{2 \pi^2}
342: \bfmut \right)^{-1/2}
343: \frac{\mu_\ell}{T} \frac{\mu_k}{T}
344: + {\cal O}(g^4)\quad , \\
345: \frac{\chi_{fk}(T,0)}{T^2} \hspace{-0.1cm}&\simeq&\hspace{-0.1cm}
346: - \frac{5}{144 \pi^6}\; g^6 \ln 1/g \quad .
347: \label{sus_lead_nd}
348: \end{eqnarray}
349:
350: As we are going to discuss lattice calculations at non-zero chemical
351: potential which are based on a Taylor expansion of the grand potential
352: $\Omega$ in terms of $\mu_f/T$ it also is instructive to consider a
353: Taylor expansion of the perturbative series for $\Omega$.
354: While perturbative terms up to ${\cal O}(g^2) $ only contribute
355: to the series up to ${\cal O}(\mu^4)$ higher order terms in the
356: expansion start receiving non-vanishing contributions at ${\cal O}(g^3)$.
357: These again arise from the electric mass term, {\it i.e.} from an expansion of
358: $m_E/gT$ in powers of $\bfmut$;
359: \begin{eqnarray}
360: \Omega^{(3)} &=& \frac{1}{6\pi}\left( 1 +\frac{n_f}{6}\right)^{3/2}
361: + \frac{1}{8\pi^3}\left( 1 +\frac{n_f}{6}\right)^{1/2}
362: \bfmut
363: + \frac{1}{64\pi^5}\left( 1 +\frac{n_f}{6}\right)^{-1/2}
364: \left( \bfmut \right)^2 \nonumber \\[2mm]
365: &~&- \frac{1}{768\pi^7}\left( 1 +\frac{n_f}{6}\right)^{-3/2}
366: \left( \bfmut \right)^3
367: +{\cal O}(\mu^8) \quad .
368: \end{eqnarray}
369: Note that the expansion coefficients up to and including ${\cal O}(\mu^4)$
370: are positive. Starting at ${\cal O}(\mu^6)$ they alternate in sign.
371: From the sign of the $\mu^6$-term it follows via Eq.~(\ref{sus}) that the
372: leading perturbative contribution to the expansion coefficient
373: of diagonal as well as non-diagonal susceptibilities at ${\cal O}(\mu^4)$
374: is negative, with the latter being an order of magnitude smaller\footnote{It
375: has also recently
376: been pointed out in the context of a large-$N_c$ expansion that
377: flavor non-diagonal contributions to the free energy are suppressed by
378: $O(1/N_c^2)$ \cite{Nc}.}.
379:
380: Rather than discussing the thermodynamics of QCD at non vanishing
381: quark chemical potential in terms of chemical potentials related to
382: quark numbers in different flavor channels, it is convenient
383: to introduce chemical potentials which are related to
384: conserved quantum numbers considered at low energy,
385: {\it i.e.} quark or baryon number and isospin. In the case
386: of two flavor QCD, which we are going to analyze in our lattice
387: calculations, we thus introduce also the quark chemical potential,
388: $\mu_q = (\mu_u+\mu_d)/2$ and the isovector chemical potential\footnote{This
389: definition of the isovector chemical potential differs from that used in
390: \cite{us2} by a factor of 2.}
391: $\mu_I = (\mu_u-\mu_d)/2$. In analogy to Eq.~(\ref{sus}) we then can
392: introduce quark number and isovector susceptibilities,
393: \begin{eqnarray}
394: \frac{\chi_q}{T^2}
395: = \frac{\partial^{2} \Omega}{\partial (\mu_q/T)^2}
396: \equiv 2 \left( \chi_{uu} + \chi_{ud} \right)
397: \quad , \quad
398: \frac{\chi_I}{T^2}
399: = \frac{\partial^{2} \Omega}{\partial (\mu_I/T)^2}
400: \equiv 2 \left( \chi_{uu} - \chi_{ud} \right) \quad ,
401: \label{eq:chiqi}
402: \end{eqnarray}
403: where in the second equality we have assumed degenerate $(u,d)$-quark
404: masses.
405:
406: \subsection{Low temperature hadron resonance gas}
407:
408: The success in describing particle abundance ratios observed in
409: heavy ion experiments at varying beam energies in terms of equilibrium
410: properties of a hadron resonance gas model \cite{pbm} begs a
411: comparison of this model for the low temperature hadronic phase with
412: lattice QCD calculations. Indeed this led to astonishingly good
413: agreement \cite{KRT}.
414:
415: In the hadron resonance gas (HRG) model it is assumed that for $T<T_0$ the
416: QCD partition function can be approximated by that of a non-interacting
417: gas of hadron resonances, either bosonic mesons or fermionic
418: baryons. This, however, does not mean that interactions in dense
419: hadronic matter have been ignored;
420: in the spirit of Hagedorn's bootstrap model \cite{Hagedorn}
421: the inclusion of
422: heavy resonances as stable particles also takes care of the interaction
423: among the hadrons in the dense gas at low temperature.
424:
425: The partition
426: function of the hadron resonance gas may be split into mesonic and
427: baryonic contributions,
428: \begin{eqnarray}
429: \ln{\cal Z}_{HRG}(T,V,\mu_q,\mu_I)
430: =\sum_{i\in\;mesons}\hspace{-3mm} \ln{\cal Z}^{M}_{m_i}(T,V,\mu_q,\mu_I)
431: +\hspace{-3mm}
432: \sum_{i\in\;baryons}\hspace{-3mm} \ln{\cal Z}^{B}_{m_i}(T,V,\mu_q,\mu_I)\; ,
433: \label{eq:ZHRG}
434: \end{eqnarray}
435: where
436: \begin{equation}
437: \ln{\cal Z}^{M/B}_{m_i}(T,V\mu_q,\mu_I)
438: =\mp{V\over{2\pi^2}}\int_0^\infty dk k^2
439: \ln(1\mp z_ie^{-\varepsilon_i/T}) \quad ,
440: \label{eq:ZMB}
441: \end{equation}
442: with energies $\varepsilon_i^2=k^2+m_i^2$ and fugacities
443: \begin{equation}
444: z_i=\exp\left((3B_i\mu_q+2I_{3i}\mu_I)/T\right) \quad .
445: \label{eq:fuga}
446: \end{equation}
447: Here $B_i$ is the baryon number and $I_{3i}$ denotes the third component of
448: the isospin of the species in question. The upper sign in Eq.~(\ref{eq:ZMB})
449: refers to bosons, and the lower sign to fermions. Note that with this
450: convention anti-particles must be counted separately in Eq.~(\ref{eq:ZHRG})
451: with fugacity $z_i^{-1}$, and self-conjugate species such as
452: $\pi^0, \eta^\prime$ have $z=1$. If the logarithms are expanded in powers
453: of fugacity, the integral over momenta, $k$, can be performed. This yields
454: \begin{equation}
455: \ln{\cal Z}^{M/B}_{m_i}=\frac{VTm_i^2}{2\pi^2} \sum_{\ell=1}^\infty
456: \left\{\matrix{1\cr(-1)^{\ell+1}}\right\}\ell^{-2} K_2\left(\frac{\ell m_i}{
457: T}\right) z_i^{\ell} \quad ,
458: \label{eq:Zmi}
459: \end{equation}
460: where the upper factor in braces applies to bosons and the lower to fermions,
461: and $K_2$ is a modified Bessel function. For large argument, {\it i.e.} for
462: $m_i \gg T$ the Bessel function can be approximated by
463: $K_2(x)\sim\sqrt{\pi/2x}\; {\rm e}^{-x} (1+15/8x +{\cal O}(x^{-2}))$.
464: Terms with $\ell\geq2$ in the series given in Eq.~(\ref{eq:Zmi}) thus are
465: exponentially suppressed. For temperatures and quark chemical
466: potentials less than a typical scale of about $200~$MeV it generally
467: suffices to keep the first term in the sum appearing in Eq.~(\ref{eq:Zmi}).
468: The only species for which this step would need further justification is
469: the pion; clearly for realistic pion masses more care must be taken
470: when evaluating the sum over $\ell$. This, however, does not influence the
471: $\mu_q$-dependence of the hadron resonance gas; $B_i=0$ for mesons
472: and their contribution thus is independent of $\mu_q$. At
473: $\mu_I=0$ one readily derives,
474: \begin{equation}
475: {p(T,\mu_q)\over T^4}= \frac{1}{VT^3} \ln{\cal Z}_{HRG}\biggr|_{\mu_I=0}
476: \simeq G(T)+F(T)\cosh \left( \frac{3\mu_q}{T} \right) \quad ,
477: \label{eq:pHRG}
478: \end{equation}
479: with $G(T)=(VT^3)^{-1}\sum_{i\in mesons} \ln {\cal Z}^M_{m_i}$ and
480: $F(T)$ given by the Boltzmann approximation to the fermion partition
481: function of baryons,
482: \begin{eqnarray}
483: F(T) &=& {1\over\pi^2}\sum_{i\in\;baryons}\left({m_i\over T}\right)^2
484: K_2\left({m_i\over T}\right)\quad .
485: \label{eq:Phip}
486: \end{eqnarray}
487: Note that each term in the sum for $F$ now counts both baryon and
488: anti-baryon.
489: As the meson sector of the partition functions is independent of $\mu_q$
490: the mesonic component does not contribute to $\chi_q$,
491: \begin{equation}
492: {\chi_q\over T^2}=9F(T)\cosh{{3\mu_q}\over T} \quad .
493: \label{eq:chiHRG}
494: \end{equation}
495: Similarly, at $\mu_I = 0$ we find for the isovector susceptibility
496: \begin{equation}
497: {\chi_I\over T^2}
498: %\biggr|_{\mu_I=0}
499: =G^I(T) +F^I(T)\cosh{{3\mu_q}\over T}
500: \label{eq:chiHRGI}
501: \end{equation}
502: with
503: \begin{eqnarray}
504: G^I(T) &=& {1\over2\pi^2}\sum_{i\in\;mesons}
505: \sum_{\ell=1}^\infty
506: (2I_{3i})^2
507: \left({m_i\over T}\right)^2 K_2\left({\ell m_i\over T}\right) \quad ,\nonumber \\
508: F^I(T) &=& {1\over \pi^2}\sum_{i\in\;baryons}(2I_{3i})^2
509: \left({m_i\over T}\right)^2 K_2\left({m_i\over T}\right) \quad .
510: \label{eq:Phichi}
511: \end{eqnarray}
512: where again only in the baryon sector $\ell > 1$ terms have been
513: neglected.
514:
515: \noindent
516: The resonance gas model in the (partial)
517: Boltzmann approximation given by
518: Eqs.~(\ref{eq:pHRG})-(\ref{eq:Phichi}) leads to simple predictions
519: for the dependence of thermodynamic observables on the quark chemical
520: potential $\mu_q$. In particular, it predicts that ratios of the
521: density dependent part of thermodynamic observables are insensitive
522: to details of the hadronic mass spectrum.
523: Nor do they depend explicitly on temperature, but instead only on
524: the ratio $\mu_q/T$. For instance, one finds
525: \begin{equation}
526: \frac{p(T,\mu_q)-p(T,0)}{\chi_q T^2} = \frac{\cosh(3 \mu_q/T) -1}{9
527: \cosh(3 \mu_q/T)} \quad , \quad
528: \frac{n_q}{\mu_q \chi_q}
529: =\frac{T}{3\mu_q} \tanh \left( \frac{3\mu_q}{T} \right) \quad .
530: \label{eq:npchiq}
531: \end{equation}
532: Similarly ratios of Taylor expansion coefficients of thermodynamic
533: quantities, $X$, are temperature and spectrum independent. For an
534: observable $X$, of generic form $X=G^X +F^X \cosh(3 \mu_q/T)$, the
535: expansion in $\mu_q/T$ is given by
536: \begin{equation}
537: X=\sum_{n=0}^\infty c_n^X (T) (\mu_q/T)^{n}, \hspace{5mm}
538: \end{equation}
539: with
540: \begin{equation}
541: c_0^X=G^X+F^X, \hspace{5mm} c_{2n}^X=\frac{9^n}{(2n)!} F^X
542: \end{equation}
543: and $c_{2n+1}^X=0$.
544: Hence, ratios are given by
545: \begin{equation}
546: \frac{c_{2n+2}^X}{c_{2n}^X}=\frac{9}{(2n+2)(2n+1)} \hspace{5mm}
547: {\rm for} \ n \geq 1 \quad .
548: \label{eq:HRGratios}
549: \end{equation}
550: These ratios as well as ratios of physical observables calculated within
551: the resonance gas approximation will be compared to corresponding lattice results
552: in the following.
553:
554:
555:
556: \section{Taylor expansion for two flavor QCD}
557: \label{sec:eos}
558:
559: The basic concepts of our Taylor expansion approach to QCD
560: thermodynamics at non-zero quark chemical potential have been introduced
561: in \cite{us2} where observables have been analyzed
562: up to ${\cal O}(\mu_q^4)$. Here we extend the analysis up to the sixth
563: order in $\mu_q/T$ with significantly improved statistics.
564:
565: Our calculations have been performed for two flavor
566: QCD on an $16^3\times4$ lattice with bare quark mass $ma=0.1$
567: using Symanzik-improved gauge and p4-improved
568: staggered fermion actions. These parameters are identical to those used
569: for our analysis of thermodynamic observables
570: at non-zero chemical potential up to ${\cal O}(\mu_q^4)$
571: \cite{us1,us2}.
572: The simulation uses the hybrid R molecular dynamics algorithm,
573: and measurements were performed on equilibrated configurations separated
574: by 5 units of molecular dynamics time $\tau$. The gauge couplings,
575: $\beta=6/g^2$, used for our calculations
576: cover the interval [3.52,4.0], which corresponds to a temperature range
577: $T/T_{0}\in[0.76,1.98]$, where $T_{0}$ is the pseudocritical temperature
578: at $\mu=0$ for which we use\footnote{In \cite{us1} we determined as
579: critical coupling $\beta_c=3.649(2)$. Our current analysis favors a
580: slightly larger value, $\beta_c=3.655(5)$.} $\beta_c=3.65$. This value is
581: used to define the temperature scale $T/T_0$. The number of
582: configurations generated at each $\beta$-value is given in
583: Table~\ref{tab:configs}.
584: The third and seventh columns give the sample sizes used in \cite{us2},
585: where coefficients up to $n=4$ were calculated. The numbers of additional
586: configurations generated for the present study of expansion coefficients
587: up to $n=6$ are listed in the fourth and eighth columns.
588: It can be seen that we have increased our
589: statistics in the hadronic phase by a factor 4-5 and in the plasma phase
590: by a factor 3.
591:
592: For the calculation of various operator traces we use the
593: method of noisy estimators. We generally found that expectation values
594: involving odd derivatives of $\ln {\rm det}M$ with respect to $\mu$ are
595: noisier and require averages over more random vectors than needed to
596: estimate expectation values involving only even
597: derivatives. Odd derivatives have to appear in even numbers in an
598: expectation value in order for this to be non-zero. Such
599: expectation values behave very much like susceptibilities and receive
600: their largest contributions in the vicinity of $T_0$. Still their total
601: contribution to the expansion of e.g. the pressure is found to be small
602: up to ${\cal O}(\mu^4)$. It only becomes sizeable at ${\cal O}(\mu^6)$.
603: For these reasons we used 100 stochastic noise vectors to estimate operator
604: traces on each configuration for $\beta\in[3.60,3.68]$. For other $\beta$
605: values 50 noise vectors were found to suffice.
606:
607: \begin{table}[t]
608: \centering
609: \setlength{\tabcolsep}{1.0pc}
610: \begin{tabular}{|llll|llll|}
611: \hline
612: $\beta$ & $T/T_c$ & \#(2,4) & $\!\!\!\!$\#(2,4,6) &
613: $\beta$ & $T/T_c$ & \#(2,4) & $\!\!\!\!$\#(2,4,6) \\
614: \hline
615: 3.52 & 0.76 & 1000 & 3500 &
616: 3.70 & 1.11 & 800 & 2000 \\
617: 3.55 & 0.81 & 1000 & 3500 &
618: 3.72 & 1.16 & 500 & 2000 \\
619: 3.58 & 0.87 & 1000 & 3500 &
620: 3.75 & 1.23 & 500 & 1000 \\
621: 3.60 & 0.90 & 1000 & 3800 &
622: 3.80 & 1.36 & 500 & 1000 \\
623: 3.63 & 0.96 & 1000 & 3500 &
624: 3.85 & 1.50 & 500 & 1000 \\
625: 3.65 & 1.00 & 1000 & 4000 &
626: 3.90 & 1.65 & 500 & 1000 \\
627: 3.66 & 1.02 & 1000 & 4000 &
628: 3.95 & 1.81 & 500 & 1000 \\
629: 3.68 & 1.07 & $\;\,$800 & 3600 &
630: 4.00 & 1.98 & 500 & 1000 \\
631: \hline
632: \end{tabular}
633: \caption{Sample size at each $\beta$ value.}
634: \smallskip
635: \label{tab:configs}
636: \end{table}
637:
638: \subsection{Pressure, quark number density and susceptibilities}
639: \label{subsec:pressure}
640:
641: To start the discussion of lattice results on thermodynamics
642: of two flavor QCD for small values of the quark chemical potential we
643: will present results obtained from a Taylor expansion of the
644: grand potential $\Omega(T,\mu_u,\mu_d)\equiv
645: \Omega(T,\mu_q+\mu_I,\mu_q-\mu_I)$ and some of its derivatives.
646: We will consider expansions in terms of $\mu_q/T$ at fixed, vanishing
647: $\mu_I$. The pressure is then given by
648:
649: \begin{table}[h]
650: %\centering
651: \setlength{\tabcolsep}{0.8pc}
652: \begin{tabular}{|l|lll|lll|}
653: \hline
654: $T/T_c$ & $c_2$ & $c_4\times10$ & $c_6\times10^2$
655: & $c^I_2$ & $c^I_4\times10$ & $c^I_6\times10^2$\\
656: \hline
657: 0.76 & 0.0243(19) & 0.238(61) & $\!\!$-1.12(121)
658: & 0.0649(6) & 0.098(5) & 0.23(9) \\
659: 0.81 & 0.0450(20) & 0.377(64) & 1.98(141)
660: & 0.0874(8) & 0.140(6) & 0.44(10) \\
661: 0.87 & 0.0735(23) & 0.506(68) & 1.69(155)
662: & 0.1206(11) & 0.216(8) & 0.60(13) \\
663: 0.90 & 0.1015(24) & 0.765(72) & 2.06(159)
664: & 0.1551(14) & 0.302(12) & 0.83(18) \\
665: 0.96 & 0.2160(31) & 1.491(135) & 4.96(260)
666: & 0.2619(21) & 0.564(23) & 1.47(37) \\
667: 1.00 & 0.3501(32) & 2.133(121) & $\!\!$-5.00(359)
668: & 0.3822(26) & 0.839(28) & 0.26(49) \\
669: 1.02 & 0.4228(33) & 2.258(118) & $\!\!$-4.49(312)
670: & 0.4501(27) & 0.909(28) & 0.02(44) \\
671: 1.07 & 0.5824(23) & 1.417(62) & $\!\!$-5.73(158)
672: & 0.5972(21) & 0.741(17) & $\!\!$-0.75(26) \\
673: 1.11 & 0.6581(20) & 0.951(39) & $\!\!$-1.65(62)
674: & 0.6662(18) & 0.618(11) & $\!\!$-0.18(10) \\
675: 1.16 & 0.7091(15) & 0.763(24) & $\!\!$-0.31(26)
676: & 0.7156(14) & 0.564(6) & $\!\!$-0.03(4) \\
677: 1.23 & 0.7517(16) & 0.667(23) & $\!\!$-0.44(23)
678: & 0.7573(13) & 0.527(5) & $\!\!$-0.06(3) \\
679: 1.36 & 0.7880(11) & 0.572(12) & $\!\!$-0.09(11)
680: & 0.7906(9) & 0.495(3) & $\!\!$-0.03(1) \\
681: 1.50 & 0.8059(10) & 0.539(10) & $\!\!$-0.17(7)
682: & 0.8076(7) & 0.477(2) & $\!\!$-0.05(1) \\
683: 1.65 & 0.8157(8) & 0.499(7) & $\!\!$-0.13(8)
684: & 0.8169(7) & 0.461(2) & $\!\!$-0.05(1) \\
685: 1.81 & 0.8203(8) & 0.497(7) & $\!\!$-0.11(6)
686: & 0.8218(6) & 0.452(1) & $\!\!$-0.05(1) \\
687: 1.98 & 0.8230(7) & 0.473(6) & 0.03(4)
688: & 0.8250(6) & 0.441(1) & $\!\!$-0.03(1) \\
689: \hline
690: \end{tabular}
691: \caption{Taylor expansion coefficients $c_n(T)$ and $c^I_n(T)$.}
692: \smallskip
693: \label{tab:c_n}
694: \end{table}
695:
696: \begin{equation}
697: {p\over T^4}\equiv\Omega(T,\mu_q,\mu_q)=
698: \sum_{n=0}^\infty c_n(T) \left({\mu_q\over T}\right)^n.\label{eq:p}
699: \end{equation}
700: CP symmetry implies that the series is even in $\mu_q$, so
701: the coefficients $c_n$ are non-zero only for $n$ even, and are defined as
702: \begin{equation}
703: c_n (T)={1\over n!}{{\partial^n \Omega}
704: \over{{\partial(\mu_q/T)^n}}}\biggr\vert_{\mu_q=0}=
705: {1\over n!}{N_\tau^{3}\over N_\sigma^3}{{\partial^n\ln{\cal Z}}\over
706: {\partial(\mu N_\tau)^n}}\biggr\vert_{\mu=0} \quad ,
707: \label{eq:cn}
708: \end{equation}
709: where in the second equality we have explicitly specified that the calculations
710: have been performed on a $N_\sigma^3\times N_\tau$ lattice with dimensionless
711: quark chemical potential $\mu\equiv\mu_q a$. ${\cal Z}$ denotes the
712: lattice regularized partition function for two flavor QCD.
713: Similarly we calculate the quark number density
714: \begin{equation}
715: \frac{n_q(T,\mu_q)}{T^3}=\frac{\partial \Omega(T,\mu_q,\mu_q)}{\partial \mu_q/T}
716: =2c_2\left(\frac{\mu_q}{T}\right) +4c_4\left(\frac{\mu_q}{T}\right)^3
717: +6c_6\left(\frac{\mu_q}{T}\right)^5 + \cdots \quad ,
718: \label{eq:nq}
719: \end{equation}
720: as well as the quark and isovector susceptibilities using
721: Eq.~(\ref{eq:chiqi}),
722: \begin{eqnarray}
723: {\chi_q(T,\mu_q)\over T^2}
724: &=&2c_2+12c_4\left({\mu_q\over T}\right)^2+30c_6\left({\mu_q\over
725: T}\right)^4+\cdots\label{eq:chiq}\\
726: {\chi_I(T,\mu_q)\over T^2}
727: &=&2c^I_2+12c^I_4\left({\mu_q\over T}\right)^2+30c^I_6\left({\mu_q\over
728: T}\right)^4+\cdots\label{eq:chiI}
729: \end{eqnarray}
730: where
731: \begin{equation}
732: c^I_n={1\over n!}{\partial^n\; \Omega
733: (T,\mu_q+\mu_I,\mu_q-\mu_I)
734: \over{{\partial(\mu_I/T)^2}\partial(\mu_q/T)^{n-2}}}
735: \biggr\vert_{\mu_q=0,\mu_I=0} \quad.
736: \end{equation}
737: Explicit expressions for $c_2$, $c_4$, $c_6$
738: and $c^I_2$, $c^I_4$ and $c^I_6$ are given in the Appendix.
739: Note that the expansion for the quark number susceptibility
740: $\chi_q$ given in Eq.~(\ref{eq:chiq}) is a derivative of the grand
741: potential at $\mu_I\equiv 0$ and thus
742: has the same radius of convergence as that of the pressure and quark
743: number density given by Eqs.~(\ref{eq:p}) and (\ref{eq:nq}).
744: The expansion coefficient $c_2^I$ also defines the first term in an
745: expansion of the pressure at non-zero isospin. This series
746: may have a different radius of convergence \cite{Kogut}; indeed, since
747: the lightest particle carrying non-zero
748: isospin $I_3$ in the hadronic phase is the pion, we might expect the
749: expansion to break down in the chiral limit for arbitrarily small $\mu_I$.
750:
751: \begin{figure}[tb]
752: \begin{center}
753: \begin{minipage}[c][5.2cm][c]{5.0cm}
754: \begin{center}
755: \epsfig{file=c2c2IvT.eps, width=5.0cm}\\[-1mm]
756: %(a)
757: \end{center}
758: \end{minipage}
759: \begin{minipage}[c][5.2cm][c]{5.0cm}
760: \begin{center}
761: \epsfig{file=c4c4IvT.eps, width=5.0cm}\\[-1mm]
762: %(b)
763: \end{center}
764: \end{minipage}
765: \begin{minipage}[c][5.2cm][c]{5.0cm}
766: \begin{center}
767: \epsfig{file=c6c6IvT.eps, width=5.0cm}\\[-1mm]
768: %(c)
769: \end{center}
770: \end{minipage}
771: \caption{The Taylor expansion coefficients $c_n$ and $c^I_n$ for $n=2, 4$ and 6
772: as functions of $T/T_{0}$.}
773: \label{fig:c2c4}
774: \end{center}
775: \end{figure}
776:
777: The coefficient $c_0 (T)$ gives
778: the pressure in units of $T^4$ at vanishing baryon density and can be
779: calculated using the integral method \cite{KLP}. It is the only expansion
780: coefficient which also requires lattice calculations at zero temperature.
781: Higher order terms can be calculated directly from gauge field
782: configurations generated on finite temperature lattices. They, however,
783: require additional derivatives of $\ln\mbox{det}M$,
784: where $M$ is the quark matrix. They are evaluated at fixed
785: temperature, {\it i.e.} fixed gauge coupling $\beta$, by calculating
786: combinations of traces of products of
787: $\partial^m M/\partial\mu^m$ and $M^{-1}$ (see Appendix).
788:
789:
790: Results for the Taylor expansion coefficients are listed in
791: Table~\ref{tab:c_n}. In Fig.~(\ref{fig:c2c4}) we plot $c_n$ and
792: $c^I_n$ for $n=2,~4$ and 6 as functions of
793: $T$. A comparison with Figs.~3 and 8 of Ref.~\cite{us2} reveals the improvement
794: in statistics of the current study. The same features are apparent:
795: namely $c_2$ and $c^I_2$ both rise steeply across $T_0$
796: with $c_2^I> c_2$ as is obvious from the explicit expressions given for
797: these coefficients in the appendix; they
798: reach a plateau at approximately 80\% of the value $n_{\rm f}/2$
799: predicted in the Stefan-Boltzmann (SB) limit, {\it i.e.} for free
800: massless quarks;
801: $c_4$ rises steeply to peak at $T\simeq T_{0}$ before approaching its
802: SB limit value $n_{\rm f}/4\pi^2$ from above, whereas the peak in
803: $c^I_4$ is much less marked\footnote{The difference is largely due to the
804: dominance of the disconnected term
805: $\langle(\partial^2\ln\mbox{det}M/\partial\mu^2)^2\rangle -
806: \langle \partial^2\ln\mbox{det}M/\partial\mu^2 \rangle^2$
807: which contributes to
808: $c_4$ with a coefficient three times that of its contribution to $c^I_4$.}.
809:
810: \begin{figure}[tb]
811: \begin{center}
812: \begin{minipage}[c][7.8cm][c]{7.4cm}
813: \begin{center}
814: \epsfig{file=dpO4O6vT.eps, width=7.4cm}\\[-1mm]
815: (a)
816: \end{center}
817: \end{minipage}
818: \begin{minipage}[c][7.8cm][c]{7.4cm}
819: \begin{center}
820: \epsfig{file=nqO4O6vT.eps, width=7.4cm}\\[-1mm]
821: (b)
822: \end{center}
823: \end{minipage}
824: \caption{The $\mu_q$ dependent contribution to the pressure (left)
825: and the quark number density (right) as functions of $T/T_{0}$
826: for various values of the quark chemical potential calculated
827: from a Taylor series in $6^{th}$ order. Also shown as dashed lines
828: are results from a $4^{th}$ order expansion in $\mu_q/T$.}
829: \label{fig:Deltap}
830: \end{center}
831: \end{figure}
832:
833: As can be seen in Table~\ref{tab:c_n} in the high temperature
834: phase the $6^{th}$ order expansion coefficients generally are an
835: order of magnitude smaller than the $4^{th}$ order coefficients. In the
836: low temperature phase they are still a factor 3-5 smaller. As a consequence
837: the $6^{th}$ order contributions to the pressure and quark number density
838: are small for $\mu_q/T \le 1$. This is seen in
839: Fig.~(\ref{fig:Deltap}) which shows $\Delta p/T^4 \equiv
840: (p(T,\mu_q)-p(T,0))/T^4$ and $n_q/T^3$
841: %using results for $c_n(T)$
842: in the range $0\le \mu_q/T \leq 1$.
843: Here we also
844: show as dashed lines results obtained from a Taylor expansion
845: which includes only terms up to $4^{th}$ order in $\mu_q/T$. This
846: suggests that the expansion for the pressure and quark number density
847: is converging rapidly for $\mu_q/T < 1$. Even for $\mu_q/T=1$ the
848: differences between the $4^{th}$ and $6^{th}$-order results are small
849: and partly influenced by statistics. We also note that in the high
850: temperature regime, $T\gsim 1.5 T_0$, our results are compatible with
851: the continuum extrapolated
852: (quenched) results obtained with an unimproved staggered
853: fermion action \cite{Gavaieos}. This supports the expectation that
854: deviations from the continuum limit are strongly suppressed with our
855: improved action.
856:
857:
858: \begin{figure}[tb]
859: \begin{center}
860: \begin{minipage}[c][7.8cm][c]{7.4cm}
861: \begin{center}
862: \epsfig{file=chiqO4O6vT.eps, width=7.4cm}\\[-1mm]
863: (a)
864: \end{center}
865: \end{minipage}
866: \begin{minipage}[c][7.8cm][c]{7.4cm}
867: \begin{center}
868: \epsfig{file=chiIO4O6vT.eps, width=7.4cm}\\[-1mm]
869: (b)
870: \end{center}
871: \end{minipage}
872: \caption{The quark number susceptibility $\chi_q/T^2$ (left) and
873: isovector susceptibility $\chi_I/T^2$ (right) as functions of $T/T_0$ for
874: various $\mu_q/T$ ranging from $\mu_q/T=0$ (lowest curve) rising in
875: steps of 0.2 to $\mu_q/T=1$, calculated
876: from a Taylor series in $6^{th}$ order. Also shown as dashed lines
877: are results from a $4^{th}$ order expansion in $\mu_q/T$.
878: }
879: \label{fig:chiIq}
880: \end{center}
881: \end{figure}
882:
883: Next we turn to a discussion of quark number and isovector susceptibilities
884: which are shown in Fig.~(\ref{fig:chiIq}). They have been obtained
885: using Eqs.~(\ref{eq:chiq}) and (\ref{eq:chiI}). Again we show the corresponding
886: $4^{th}$-order results as dashed lines in these figures.
887: These lines agree with our old results shown as Fig. 9 of \cite{us2}.
888: The effect of the new term proportional to $c_6(T)$ is to shift the apparent
889: maximum in $\chi_q(T)$, arising from the sharply peaked $\mu_q^4$-contribution
890: proportional to $c_4(T)$, to lower temperature.
891: This suggests that the transition temperature at non-zero $\mu_q$ determined
892: from the peak position of susceptibilities indeed moves to temperatures
893: smaller than the transition temperature $T_0$ determined at $\mu_q=0$. The
894: figure, however, also shows that at least for $T<T_0$ the $6^{th}$-order
895: contribution can be sizeable and still suffers from statistical errors.
896: Better statistics and the contribution from higher orders in the Taylor
897: expansion thus will be needed to get good quantitative results for
898: susceptibilities in the hadronic phase.
899:
900: There is also a pronounced dip in $\chi_q(T)$ for $T/T_{0}\simeq1.05$ which,
901: together with the increased error bars makes the presence of a peak in $\chi_q$
902: less convincing then it is without the inclusion of the $\mu_q^6$-contribution.
903: However, the error bars also reflect the problem we have at present in determining
904: this additional contribution with sufficient accuracy to include it in the
905: calculation of higher order derivatives of the partition function.
906: On the other side, Fig.~(\ref{fig:chiIq}) confirms that a significant peak
907: is not present in the isovector channel.
908: In fact, if a critical endpoint exists in the ($T,\mu$)-plane of the
909: QCD phase diagram, this is expected to belong to
910: the Ising universality class, implying
911: that exactly one 3$d$ scalar degree of freedom becomes massless at this point.
912: Since both $\bar\psi\psi$ and $\bar\psi\gamma_0\psi$ are isoscalar and Galilean
913: scalars, both are candidates to interpolate this massless field, and hence
914: we can expect divergent fluctuations in both quark number and chiral
915: susceptibilities at this point. The latter will be discussed in section 3.3.
916:
917: The difference in the temperature dependence of $\chi_q$ and $\chi_I$
918: also reflects the strong correlation between
919: fluctuations in different flavor components. This will become clear from our
920: discussion of flavor diagonal and non-diagonal susceptibilities
921: in the next section.
922:
923:
924: \subsection{Flavor diagonal and non-diagonal susceptibilities}
925:
926:
927: Using the relation between quark number and isovector susceptibilities
928: on the one hand
929: and diagonal and non-diagonal susceptibilities on the other hand
930: (Eq.~(\ref{eq:chiqi}))
931: we also can define expansions for the latter,
932: \begin{eqnarray}
933: {\chi_{uu}(T, \mu_q)\over T^2} =
934: &=&2c^{uu}_2+12c^{uu}_4\left({\mu_q\over T}\right)^2+
935: 30c^{uu}_6\left({\mu_q\over T}\right)^4+\cdots\label{eq:chiuu} \quad ,
936: \\
937: {\chi_{ud}(T, \mu_q)\over T^2} =
938: &=&2c^{ud}_2+12c^{ud}_4\left({\mu_q\over T}\right)^2+
939: 30c^{ud}_6\left({\mu_q\over T}\right)^4+\cdots\label{eq:chiud} \quad ,
940: \end{eqnarray}
941: with $c_n^{uu}=(c_n+c_n^I)/4$ and $c_n^{ud}=(c_n-c_n^I)/4$.
942:
943: As discussed in the previous section the expansion coefficients $c_n$ and $c^I_n$
944: become quite similar at high temperature. This was to be expected from the
945: discussion of the structure
946: of the high temperature perturbative expansion given in
947: section~\ref{subsec:pert} as $c_n$ and $c_n^I$ differ only by
948: contributions coming from non-diagonal susceptibilities, which enter
949: with opposite sign in these two coefficients. It thus is instructive
950: to analyze directly the expansion coefficients of $\chi_{uu}$ and
951: $\chi_{ud}$. These are listed in Table~\ref{tab:cuu_n} and plotted in
952: Fig.~(\ref{fig:uu}). We note that the errors on these quantities have been
953: obtained from an independent jackknife analysis and
954: thus are not simply obtained by adding errors for $c_n$ and $c_n^I$.
955:
956: \begin{table}[h]
957: %\centering
958: \setlength{\tabcolsep}{0.8pc}
959: \begin{tabular}{|l|lll|lll|}
960: \hline
961: $T/T_c$ &
962: $c_2^{uu} \times 10^2$ & $c_4^{uu} \times 10^2$ & $c_6^{uu} \times 10^2$ &
963: $c_2^{ud} \times 10^2$ & $c_4^{ud} \times 10^2$ & $c_6^{ud} \times 10^2$ \\
964: \hline
965: 0.76 & $\,\,$2.23(6) & 0.84(16) & $\!\!$-0.22(32)
966: & $\!\!$-1.015(42) & 0.35(14) & $\!\!$-0.34(29)\\
967: 0.81 & $\,\,$3.31(6) & 1.29(17) & 0.60(37)
968: & $\!\!$-1.060(45) & 0.59(15) & 0.38(33)\\
969: 0.87 & $\,\,$4.85(8) & 1.81(18) & 0.57(42)
970: & $\!\!$-1.177(46) & 0.72(16) & 0.27(36)\\
971: 0.90 & $\,\,$6.41(9) & 2.67(20) & 0.72(44)
972: & $\!\!$-1.339(45) & 1.16(16) & 0.31(36)\\
973: 0.96 & 11.95(12) & 5.14(39) & 1.61(74)
974: & $\!\!$-1.148(45) & 2.32(29) & 0.87(57)\\
975: 1.00 & 18.31(14) & 7.43(37) & $\!\!$-1.19(101)
976: & $\!\!$-0.802(32) & 3.23(24) & $\!\!$-1.31(78)\\
977: 1.02 & 21.82(15) & 7.92(36) & $\!\!$-1.12(88)
978: & $\!\!$-0.681(27) & 3.37(23) & $\!\!$-1.13(69)\\
979: 1.07 & 29.49(11) & 5.39(20) & $\!\!$-1.62(46)
980: & $\!\!$-0.369(17) & 1.69(11) & $\!\!$-1.25(33)\\
981: 1.11 & 33.11(9) & 3.92(12) & $\!\!$-0.46(18)
982: & $\!\!$-0.205(20) & 0.83(7) & $\!\!$-0.37(13) \\
983: 1.16 & 35.62(7) & 3.32(7) & $\!\!$-0.08(8)
984: & $\!\!$-0.162(17) & 0.50(5) & $\!\!$-0.07(6) \\
985: 1.23 & 37.73(7) & 2.98(7) & $\!\!$-0.13(6)
986: & $\!\!$-0.140(22) & 0.35(5) & $\!\!$-0.09(5) \\
987: 1.36 & 39.47(5) & 2.67(4) & $\!\!$-0.03(3)
988: & $\!\!$-0.063(18) & 0.19(3) & $\!\!$-0.01(2) \\
989: 1.50 & 40.34(4) & 2.54(3) & $\!\!$-0.06(2)
990: & $\!\!$-0.043(16) & 0.15(2) & $\!\!$-0.03(2) \\
991: 1.65 & 40.81(4) & 2.40(2) & $\!\!$-0.04(2)
992: & $\!\!$-0.029(14) & 0.10(1) & $\!\!$-0.02(2) \\
993: 1.81 & 41.05(3) & 2.37(2) & $\!\!$-0.04(2)
994: & $\!\!$-0.040(14) & 0.11(2) & $\!\!$-0.01(1) \\
995: 1.98 & 41.20(3) & 2.29(2) & $\!\!$-0.00(1)
996: & $\!\!$-0.051(13) & 0.08(1) & 0.02(1) \\
997: \hline
998: \end{tabular}
999: \caption{Taylor expansion coefficients $c_n^{uu}(T)$ and $c_n^{ud}(T)$.}
1000: \smallskip
1001: \label{tab:cuu_n}
1002: \end{table}
1003:
1004: Fig.~(\ref{fig:uu}) clearly shows that for $T>T_0$
1005: the various expansion coefficients
1006: rapidly approach the corresponding ideal gas values, which is zero for
1007: all non-diagonal expansion coefficients, $c^{ud}_n$. In fact, as discussed
1008: in section~\ref{subsec:pert} the latter receive contributions
1009: only at ${\cal O}(g^6\ln 1/g^2)$ for $n=2$ and ${\cal O}(g^3)$ for $n>2$.
1010: Moreover, it
1011: is interesting to note that despite the small magnitude of these
1012: contributions the leading order perturbative results correctly
1013: predict the sign of all expansion coefficients
1014: for $T>T_0$, {\it i.e.} $c_{2,6}^{ud}<0$, $c_4^{ud} >0$, and
1015: $c_{2,4}^{uu} > 0$, $c_6^{uu} < 0$. Furthermore, the
1016: order of magnitude for $c_2^{ud}$, {\it i.e.} $|c_2^{ud} | \simeq 5\cdot 10^{-4}$
1017: at $T\simeq 2 T_0$,
1018: agrees with the perturbative estimate\footnote{A previous analysis
1019: \cite{Gupta02} reported values for $c^{ud}_2$ consistent
1020: with zero for temperatures $T/T_0\ge 1.25$ within an error of $10^{-6}$.
1021: This has been found subsequently to be incorrect. The corrected
1022: values \cite{Gavai04} are qualitatively
1023: consistent with our findings. (The second panel
1024: of Fig.~10 in \cite{Gavai04} should be compared to twice the value of
1025: $c_2^{ud}$ shown in our Fig.~(\ref{fig:uu}).)
1026: Similar agreement is found with the results of
1027: \cite{suscept2}.
1028: }
1029: \cite{Blaizot}.
1030:
1031: \begin{figure}[tb]
1032: \begin{center}
1033: \begin{minipage}[c][9.5cm][c]{5.0cm}
1034: \begin{center}
1035: \epsfig{file=cuu2vT.eps, width=5.0cm}\\[-1mm]
1036: \epsfig{file=cud2vT.eps, width=5.0cm}\\[-1mm]
1037: %(a)
1038: \end{center}
1039: \end{minipage}
1040: \begin{minipage}[c][9.5cm][c]{5.0cm}
1041: \begin{center}
1042: \epsfig{file=cuu4vT.eps, width=5.0cm}\\[-1mm]
1043: \epsfig{file=cud4vT.eps, width=5.0cm}\\[-1mm]
1044: %(b)
1045: \end{center}
1046: \end{minipage}
1047: \begin{minipage}[c][9.5cm][c]{5.0cm}
1048: \begin{center}
1049: \epsfig{file=cuu6vT.eps, width=5.0cm}\\[-1mm]
1050: \epsfig{file=cud6vT.eps, width=5.0cm}\\[-1mm]
1051: %(c)
1052: \end{center}
1053: \end{minipage}
1054: \caption{The Taylor expansion coefficients $c_n^{uu}$ (upper row) of diagonal and
1055: $c^{ud}_n$ (lower row) of non-diagonal susceptibilities
1056: for $n=2,~4$ and 6 as functions of $T/T_{0}$.}
1057: \label{fig:uu}
1058: \end{center}
1059: \end{figure}
1060:
1061: A striking feature of the expansion coefficients $c_n^{uu}$ and
1062: $c_n^{ud}$ is that for $n>2$ they become similar in magnitude close to
1063: $T_0$. In fact, $c_4^{uu}$ and $c_4^{ud}$ both have pronounced peaks at
1064: $T_0$ with $(c_4^{ud}/c_4^{uu})_{\rm peak} \simeq 0.45$ and the expansion
1065: coefficients for $n=6$ are identical within errors. This suggests that
1066: any divergent piece in $\chi^{uu}$, which could occur when $\mu_q/T$
1067: approaches the radius of convergence of the Taylor expansion, will
1068: show up with identical strength also in $\chi^{ud}$. This, in turn,
1069: implies that the singular behavior will add up constructively in the
1070: quark number susceptibility whereas it can cancel in the isovector
1071: susceptibility giving rise to finite values for $\chi_I$ at such a
1072: critical point. Even for smaller, non-critical values of $\mu_q/T$, however,
1073: the rapid
1074: rise of $c_4^{ud}(T)$ for $T\simeq T_0$ is important. It shows that
1075: non-diagonal susceptibilities will become large at non-zero chemical
1076: potential in the transition region from the low to the high temperature
1077: phase, {\it i.e.} fluctuations in different flavor channels, which are
1078: uncorrelated at high temperature, become strongly correlated in the
1079: transition region. This correlation is also reflected in the errors of the
1080: various expansion coefficients, which are of similar size for $c_n^{uu}$ and
1081: $c_n^{ud}$ but much reduced in the difference, $\chi_I$.
1082:
1083: The above considerations also suggest that the electric charge susceptibility,
1084: \begin{eqnarray}
1085: \chi_C=\left({2\over3}{{\partial\;}\over{\partial\mu_u}}
1086: -{1\over3}{{\partial\;}\over{\partial\mu_d}}\right)
1087: \left({2\over3}{{\partial p}\over{\partial\mu_u}}
1088: -{1\over3}{{\partial p}\over{\partial\mu_d}}\right)
1089: &=&\frac{1}{9}\left( 4\chi_{uu}+\chi_{dd} -4\chi_{ud}\right) \nonumber \\
1090: ~&=& \frac{1}{4}\left( \chi_I + {1\over 9}\chi_q \right) \quad ,
1091: \end{eqnarray}
1092: will be singular whenever the diagonal and non-diagonal susceptibilities
1093: are singular as the cancellation between the corresponding singular parts
1094: will be incomplete. We show the charge susceptibility in
1095: Fig.~(\ref{fig:charge}).
1096: As it is dominated by the contribution from the isovector susceptibility
1097: any possible singular contribution arising from $\chi_q$ will be weak. It thus
1098: may not be too surprising that a peak does not yet show up in $\chi_C$.
1099:
1100: \begin{figure}[tb]
1101: \begin{center}
1102: \begin{minipage}[c][7.8cm][c]{7.4cm}
1103: \begin{center}
1104: \epsfig{file=chiCHvT.eps, width=7.4cm}\\[-1mm]
1105: \end{center}
1106: \end{minipage}
1107: \caption{The charge susceptibility $\chi_C/T^2$ as a function of $T/T_0$ for
1108: various $\mu_q/T$ ranging from $\mu_q/T=0$ (lowest curve) rising in
1109: steps of 0.2 to $\mu_q/T=1$, calculated
1110: from a Taylor series in $6^{th}$ order.}
1111: \label{fig:charge}
1112: \end{center}
1113: \end{figure}
1114:
1115:
1116:
1117: \subsection{Mass derivatives and chiral condensate}
1118: \label{subsec:chiral}
1119:
1120: The transition between low and high temperature phases of strongly
1121: interacting matter is expected to be closely related to chiral symmetry
1122: restoration. It is therefore also of interest to analyze the
1123: dependence of the chiral condensate on the quark chemical potential.
1124: We will do so in the framework of a Taylor expansion of the grand
1125: potential,
1126: \begin{equation}
1127: \frac{\langle \bar{\psi} \psi \rangle}{T^3}
1128: =\left( \frac{N_\tau}{N_{\sigma}}\right)^3
1129: \frac{\partial \ln{\cal Z}}{\partial m/T}
1130: = \sum_{n=0}^{\infty} c_n^{\bar{\psi}\psi}(T) \left(
1131: \frac{\mu_q}{T}\right)^n \quad ,
1132: \label{eq:pbp}
1133: \end{equation}
1134: with
1135: \begin{equation}
1136: c_n^{\bar{\psi} \psi} = \frac{1}{n!} \left.
1137: \frac{\partial^n \langle \bar{\psi} \psi \rangle/T^3}{\partial (\mu_q/T)^n}
1138: \right|_{\mu_q=\mu_I=0} = \frac{1}{n! } \left.
1139: \frac{\partial^{n+1} \Omega}{\partial (\mu_q/T)^n \partial m/T}
1140: \right|_{\mu_q=\mu_I=0}
1141: %= \frac{\partial c_n}{\partial m/T}
1142: \quad .
1143: \label{eq:pbpcoef}
1144: \end{equation}
1145: Here we expressed the bare lattice quark masses, $ma$, in units of the
1146: temperature by using $m/T\equiv ma N_\tau$.
1147: For $n>0$ the expansion coefficients of the chiral condensate,
1148: are directly related to derivatives of the
1149: expansion coefficients of the grand potential $\Omega$ with respect
1150: to the quark mass, {\it i.e.} $c_n^{\bar{\psi} \psi} = \partial
1151: c_n/\partial (m/T)$. For $n=0$ this holds true up to a contribution
1152: arising from the normalization of the pressure at $(T=0,\mu_q=0)$.
1153: As such the coefficients $c_n^{\bar{\psi} \psi}$ also provide information on
1154: the quark mass
1155: dependence of other thermodynamic observables like pressure, number density
1156: or susceptibilities. For instance, the change of the quark number
1157: susceptibility with quark mass is given by
1158: \begin{equation}
1159: \frac{\partial \chi_q/T^2}{\partial m/T}
1160: = 2c_2^{\bar{\psi} \psi}+12c_4^{\bar{\psi} \psi}\left({\mu_q\over T}\right)^2+
1161: {\cal O}(\mu_q^4) \quad .
1162: \label{eq:chiqm}
1163: \end{equation}
1164: \begin{figure}[tb]
1165: \begin{center}
1166: \begin{minipage}[c][5.2cm][c]{5.0cm}
1167: \begin{center}
1168: \epsfig{file=dc0dmt.eps, width=5.0cm}\\[-1mm]
1169: %(a)
1170: \end{center}
1171: \end{minipage}
1172: \begin{minipage}[c][5.2cm][c]{5.0cm}
1173: \begin{center}
1174: \epsfig{file=dc2dmt.eps, width=5.0cm}\\[-1mm]
1175: %(b)
1176: \end{center}
1177: \end{minipage}
1178: \begin{minipage}[c][5.2cm][c]{5.0cm}
1179: \begin{center}
1180: \epsfig{file=dc4dmt.eps, width=5.0cm}\\[-1mm]
1181: %(c)
1182: \end{center}
1183: \end{minipage}
1184: \caption{The Taylor expansion coefficients $c_n^{\bar{\psi} \psi}$
1185: of the chiral condensate for $n=0, 2$ and 4 as functions of $T/T_{0}$.
1186: Also shown are the coefficients $c_n^{I,\bar{\psi} \psi}$ for $n=0, 2$
1187: which define the quark mass derivatives of the isovector susceptibility
1188: in analogy to Eq.~(\ref{eq:chiqm}).}
1189: \label{fig:dcc}
1190: \end{center}
1191: \end{figure}
1192: We have calculated the derivatives of $c_n$ with respect to the quark mass
1193: for $n=0$, 2 and 4. These derivatives are shown in Fig.~(\ref{fig:dcc})
1194: together with the corresponding derivatives for the expansion coefficients
1195: of the isovector susceptibility, which have a similar temperature dependence,
1196: \begin{equation}
1197: \frac{\partial \chi_I/T^2}{\partial m/T}
1198: = 2c_2^{I,\bar{\psi} \psi}+
1199: 12c_4^{I,\bar{\psi} \psi}\left({\mu_q\over T}\right)^2+
1200: {\cal O}(\mu_q^4) \quad .
1201: \label{eq:chiIm}
1202: \end{equation}
1203: We note that the expansion coefficients $c_n^{\bar{\psi} \psi}$ are negative for
1204: $n>0$ and $T\le 0.96 T_0$. The chiral condensate thus will drop at fixed
1205: temperature
1206: with increasing $\mu_q/T$ and the chiral susceptibilities will increase
1207: in the hadronic phase with decreasing quark mass. This, together with
1208: the change of sign in $c_4^{\bar{\psi} \psi}$ at $T\simeq T_0$, will shift
1209: the transition point at non-zero $\mu_q/T$ to lower temperatures. In
1210: Fig.~(\ref{fig:ccn}) we show the chiral condensate and the related chiral
1211: susceptibility\footnote{The chiral susceptibility introduced here is not
1212: the complete derivative of the chiral condensate with respect to the quark
1213: mass.
1214: As frequently done also at $\mu_q = 0$ we define the chiral susceptibility
1215: by ignoring a contribution from the connected part which would arise
1216: in the derivative $\partial \langle \bar{\psi} \psi \rangle / \partial m$.
1217: Nonetheless $\chi_{\bar{\psi} \psi}$ seems to capture the leading singular
1218: behavior that should show up at a $2^{nd}$ order critical point
1219: \cite{FKEL}.}
1220: obtained from a Taylor expansion up to and including ${\cal O}(\mu_q^4)$,
1221: \begin{eqnarray}
1222: \frac{\chi_{\bar{\psi} \psi}}{T^2}
1223: &=& \frac{N_\tau}{N_\sigma^3} \left( \frac{n_{\rm f}}{4}\right)^2
1224: \left[ \left\langle \left( {\rm tr} M^{-1} \right)^2 \right\rangle
1225: - \left\langle {\rm tr} M^{-1} \right\rangle^2
1226: \right] \nonumber \\
1227: &=& c_0^{\chi}+c_2^{\chi} \left( \frac{\mu_q}{T}\right)^2
1228: +c_4^{\chi} \left( \frac{\mu_q}{T}\right)^4 + {\cal O}\left(
1229: \left( \mu_q/T\right)^6 \right)
1230: \quad .
1231: \label{eq:chisus}
1232: \end{eqnarray}
1233: Obviously $\chi_{\bar{\psi} \psi}$ develops a much
1234: more pronounced peak for $\mu_q/T >0$ than at vanishing chemical potential
1235: which, moreover, is shifted to smaller temperatures.
1236: However, as will become clear from the discussion in the next section
1237: the peaks found in $\chi_{\bar{\psi}\psi}$ and also in other susceptibilities
1238: have to be analyzed and interpreted carefully. They reflect the abrupt
1239: transition from the
1240: hadronic regime to the high temperature phase in which fluctuations of the
1241: chiral condensate are suppressed, but do not signal the presence of
1242: a $2^{nd}$ order phase transition unambiguously. The rapid rise of
1243: susceptibilities in the hadronic phase
1244: is strongly correlated to the increase in the pressure and is also
1245: present in a hadron gas which does not show any singular behavior at the
1246: transition temperature.
1247: \begin{figure}
1248: \begin{center}
1249: \epsfig{file=ccndt.eps, width=7.4cm}
1250: \epsfig{file=csudt.eps, width=7.4cm}
1251: \smallskip
1252: \caption{The chiral condensate $\langle \bar{\psi} \psi \rangle$ (left)
1253: and chiral susceptibility $\chi_{\bar{\psi} \psi}$ (right)
1254: as a function of $T/T_0$ for $\mu_q/T=0,~0.4$ and 0.8. The chiral
1255: condensate drops with increasing $\mu_q/T$ and the peak in
1256: $\chi_{\bar{\psi} \psi}$ becomes more pronounced.}
1257: \label{fig:ccn}
1258: \end{center}
1259: \end{figure}
1260:
1261:
1262: Also the expansion of the chiral condensate and related observables are
1263: compatible with the HRG model. In fact, a comparison of the
1264: temperature dependence of $c_n^{\bar{\psi} \psi}$ shown in Fig.~(\ref{fig:dcc})
1265: with that of the expansion coefficients $c_n$ of the grand potential shown
1266: in Fig.~(\ref{fig:c2c4}) suggests a strong similarity between
1267: $-c_n^{\bar{\psi} \psi}$ and $c_{n+2}$. On the other
1268: hand, for $T<T_0$ the ratio $c_4^{\bar{\psi} \psi} /c_2^{\bar{\psi} \psi}$
1269: agrees within errors with the ratio $c_4/c_2$ shown in
1270: Fig.~(\ref{fig:rc2c4c6}). All this is consistent with the HRG
1271: model where the quark mass (or spectrum) dependence only enters through
1272: the functions $F(T)$ and $G(T)$ and does not modify the dependence on
1273: $\mu_q/T$.
1274:
1275:
1276:
1277:
1278: \section{Radius of convergence and the hadron resonance gas}
1279: \label{sec:radius}
1280:
1281: So far we have not discussed the range of validity of the Taylor expansion.
1282: In general the Taylor series will only converge for
1283: $\mu_q/T < \rho$ (or $\mu_q/T \le \rho$) where the radius of convergence,
1284: $\rho$,
1285: is determined by the zero of ${\cal Z} (T,\mu_q,\mu_q)$ closest to the origin
1286: of the complex $\mu_q$ plane. If this zero happens to lie on the
1287: real axis the radius of convergence coincides with a critical point of the
1288: QCD partition function. A sufficient condition for this is that
1289: all expansion coefficients are positive \cite{Gaunt}.
1290: Apparently this is the case for
1291: all coefficients $c_n(T)$ with $T/T_0 < 0.96$ that have been calculated so
1292: far by us\footnote{In \cite{Gavai04} it is reported that $c_8(T)$ is
1293: negative for $T<0.95T_0$, however, the statistical significance of this
1294: result unfortunately is not given.}.
1295: Above $T_0$, however, we find from the calculation
1296: of $c_6(T)$ that the expansion coefficients do not stay strictly positive.
1297: This is in accordance with our expectation to find a chiral critical point
1298: at some temperature $T < T_0$.
1299:
1300: \begin{figure}[tb]
1301: \begin{center}
1302: \begin{minipage}[c][5.2cm][c]{5.0cm}
1303: \begin{center}
1304: \epsfig{file=rc2c4c6.eps, width=5.0cm}\\[-1mm]
1305: (a)
1306: \end{center}
1307: \end{minipage}
1308: \begin{minipage}[c][5.2cm][c]{5.0cm}
1309: \begin{center}
1310: \epsfig{file=rc2c4c6IV.eps, width=5.0cm}\\[-1mm]
1311: (b)
1312: \end{center}
1313: \end{minipage}
1314: \begin{minipage}[c][5.2cm][c]{5.0cm}
1315: \begin{center}
1316: \epsfig{file=ccrc0c2c4.eps, width=5.0cm}\\[-1mm]
1317: (c)
1318: \end{center}
1319: \end{minipage}
1320: \caption{(a) The ratios $c_4/c_2$ and $c_6/c_4$, and
1321: (b) $c^I_4/c^I_2$ and $c^I_6/c^I_4$
1322: as well as (c) $c_2^{\bar{\psi}\psi}/c_0^{\bar{\psi}\psi}$ and
1323: $c_4^{\bar{\psi}\psi}/c_2^{\bar{\psi}\psi}$ as functions of $T$.
1324: Horizontal lines indicate the HRG prediction for $T<T_{0}$, and
1325: the SB prediction for $T>T_{0}$. Note the difference in vertical scale
1326: between the plots.}
1327: \label{fig:rc2c4c6}
1328: \end{center}
1329: \end{figure}
1330:
1331: The radius of convergence of the Taylor series
1332: for $\Omega (T,\mu_q,\mu_q)$ can be estimated by inspecting ratios of subsequent
1333: expansion coefficients,
1334: \begin{equation}
1335: \rho=\lim_{n\to\infty}\rho_{2n}\equiv\lim_{n\to\infty}\sqrt{\biggl\vert
1336: {c_{2n}\over c_{2n+2}}\biggr\vert} \quad ,
1337: \end{equation}
1338: where the square root arises because the Taylor expansion of the grand
1339: potential $\Omega$ is an even series in $\mu_q/T$. The ratios $c_{2n+2}/c_{2n}$
1340: are shown in Fig.~(\ref{fig:rc2c4c6}) together with ratios of the expansion
1341: coefficients $c_n^I$ of the isovector susceptibility
1342: and $c_n^{\bar{\psi}\psi}$ of the chiral condensate.
1343: It is obvious that these ratios rapidly change across $T_0$ and approach the value
1344: of corresponding ratios obtained in the high temperature ideal gas limit.
1345: Another remarkable feature, however, is that below $T_0$ the ratios
1346: involving expansion coefficients of the $\mu_q$-dependent parts of
1347: $\Omega$, $\chi_I$ and $\chi_{\bar{\psi}\psi}$ are almost temperature independent.
1348: In fact, these ratios are consistent with the corresponding ratios deduced
1349: from the grand potential of a hadron resonance gas (Eq.~(\ref{eq:HRGratios})),
1350: {\it i.e.} ${c_4/ c_2}=c_4^{\bar{\psi}\psi}/c_2^{\bar{\psi}\psi}={3/4}$
1351: and ${c_6/ c_4}={c^I_6/ c^I_4}={3/10}$. In
1352: ratios that contain the lowest order expansion coefficients, {\it i.e.}
1353: $c_0$, $c_0^{\bar{\psi}\psi}$ and $c_2^I$, the spectrum dependence
1354: does not cancel because the lowest order expansion coefficients also
1355: depend on the meson sector which is not the case for higher order
1356: coefficients. These ratios thus show a significant temperature
1357: dependence as can be seen for $c^I_4/ c^I_2$ and
1358: $c_2^{\bar{\psi}\psi}/ c_0^{\bar{\psi}\psi}$ shown in Fig.~(\ref{fig:rc2c4c6}).
1359:
1360: \begin{figure}
1361: \begin{center}
1362: \epsfig{file=rdpchimu6.eps, width=7.4cm}
1363: \epsfig{file=rnqchimu.eps, width=7.4cm}
1364: \smallskip
1365: \caption{The baryonic part of the pressure
1366: divided by the quark number susceptibility, $\Delta p / \chi_q$ (left),
1367: and the normalized derivative of pressure with respect to quark number
1368: density, $n_q/\chi_q$ (right), as a function of $\mu_q/T$ for various $T/T_0$.
1369: Horizontal lines show the infinite temperature ideal gas values and
1370: the HRG model prediction ($T\le T_0$) (solid lines) and expanded to
1371: $6^th$-order in $\mu_q/T$ (dashed lines). The difference is visible
1372: only for $\Delta p / \chi_q$ at $\mu_q/T=0.8$.}
1373: \label{fig:rnqchiq}
1374: \end{center}
1375: \end{figure}
1376:
1377: Similar information is contained in the ratios of
1378: physical observables, e.g. the quark number density or
1379: pressure over the quark number susceptibility, introduced in
1380: Eq.~(\ref{eq:npchiq}).
1381: These ratios are shown in Fig.~(\ref{fig:rnqchiq}). Here
1382: $n_q/T^3$ and $\chi_q/T^2$ have been calculated
1383: using Eqs.~(\ref{eq:nq}) and (\ref{eq:chiq}) up to ${\cal O}(\mu_q^6)$.
1384: The ratio, $n_q/\chi_q= (\partial p/ \partial \mu_q)/(\partial n_q/ \partial
1385: \mu_q)= \partial p/ \partial n_q$ is related
1386: to the isothermal compressibility, $\kappa_T = \chi_q/n_q^2$ which diverges
1387: at a $2^{nd}$ order phase transition point, {\it i.e.} at a
1388: point at which $\partial p/ \partial n_q =0$, the number
1389: of particles is unstable under small changes in the pressure
1390: (mechanical instability) and large density fluctuations occur. This instability
1391: leads to a divergence in the quark number susceptibility \cite{Kunihiro}.
1392: A $2^{nd}$ order phase transition is
1393: thus expected to be signalled by a zero in both ratios shown in
1394: Fig.~(\ref{fig:rnqchiq}). On the other hand, for $\mu_I=0$
1395: these ratios are expected to be
1396: constant in an ideal quark-gluon plasma as well as in a hadron resonance gas,
1397: \begin{equation}
1398: \frac{n_q^{SB}}{\mu_q \chi_q^{SB}}
1399: =\frac{1+\frac{1}{\pi^2}\left( \frac{\mu_q}{T}\right)^2}{1+
1400: \frac{3}{\pi^2}\left( \frac{\mu_q}{T}\right)^2} \quad , \quad
1401: \frac{n_q^{HRG}}{\mu_q \chi_q^{HRG}}
1402: =\frac{T}{3\mu_q} \tanh \left( \frac{3\mu_q}{T} \right) \quad .
1403: \end{equation}
1404: The corresponding values are indicated in Fig.~(\ref{fig:rnqchiq})
1405: by horizontal lines.
1406:
1407: As far as the determination of a possible $2^{nd}$ order
1408: critical point at non-vanishing quark chemical potential (chiral
1409: critical point) is concerned
1410: Fig.~(\ref{fig:rc2c4c6}) and Fig.~(\ref{fig:rnqchiq}) contain identical
1411: information.
1412: For $T\le 0.96 T_0$ bulk thermodynamic observables agree with predictions
1413: based on an HRG model, which in itself does not show any
1414: critical behavior as function of $\mu_q/T$ at fixed $T$.
1415: In particular, there is no hint for a dip in $\Delta p/\chi_q$ or
1416: $n_q/\chi_q$ which could
1417: signal the presence of a second order transition point.
1418: The same observation, albeit with
1419: larger statistical errors, holds for ratios involving
1420: the chiral susceptibility $\chi_{\bar{\psi}\psi}$. Nonetheless, all
1421: these quantities change rapidly in the transition from the low temperature
1422: to the high temperature regime and, moreover, at $T=T_0$ the $6^{th}$ order
1423: expansion
1424: coefficients clearly cannot be described within the HRG
1425: model. Due to the good agreement with the HRG model and its
1426: Taylor expansion at lower temperature we cannot, however, present
1427: an upper limit for the radius of convergence below $T_0$;
1428: the ratios shown in
1429: Fig.~(\ref{fig:rc2c4c6}) suggest that a lower limit is given by
1430: $(\mu_q/T)_c \gsim 1$.
1431: Also from the analysis of the temperature dependence
1432: of bulk thermodynamic observables we get, at present, no unambiguous
1433: evidence for the existence of a phase transition. At present, therefore,
1434: we cannot
1435: rule out that in the temperature range covered by our analysis
1436: ($T\gsim 0.8T_0$) the transition to the high temperature phase is a rapid
1437: crossover transition rather than a phase transition. This situation then would be
1438: similar to that at $\mu_q=0$.
1439: In order to exclude this possibility we would need, in the future:
1440: to consider even
1441: higher orders in the Taylor series; to scan in more detail the small
1442: temperature interval $[0.95T_0,T_0]$; and to explore systematically
1443: the quark mass and volume dependence of our results.
1444: These issues are partially
1445: addressed already in the next section where we discuss the use of
1446: reweighting techniques to calculate some thermodynamic observables and compare
1447: results obtained within this approach with results from the Taylor expansion.
1448:
1449: The good agreement found here for different ratios of Taylor expansion
1450: coefficients calculated on the lattice and within the HRG model suggests
1451: that we may use this information for a more detailed analysis of the
1452: composition of hadronic matter at temperatures below $T_0$.
1453: In Ref.~\cite{KRT} also the temperature dependence of thermodynamic
1454: observables like the pressure or the quark number susceptibility have
1455: been compared to the HRG model.\footnote{For these observables
1456: a good functional agreement between lattice data and the leading
1457: $\mu_q$ dependent term in the HRG model has also been noted
1458: in \cite{Lombardoeos} within the imaginary chemical potential approach.}
1459: In order to do so the hadron spectrum
1460: has been adjusted to the conditions realized in the lattice calculations,
1461: {\it i.e.} all masses have been shifted to larger values as the lattice
1462: calculations have been performed with unphysically large quark masses.
1463: This approach can also be turned around. The HRG model for $T< T_0$ can
1464: be used as an ansatz to determine the contributions
1465: of the mesonic and baryonic parts of the spectrum without making assumptions
1466: on the distortion of the spectrum due to the unphysical quark mass values.
1467:
1468: As outlined in section 2,
1469: within the Boltzmann approximation the HRG model yields a simple
1470: dependence of the pressure on the quark chemical potential.
1471: The relation given in Eq.~(\ref{eq:pHRG}) can easily be extended to also
1472: include a non-vanishing isovector chemical potential.
1473: Neglecting the mass difference among isospin partners the pressure can be
1474: written as,
1475: \begin{eqnarray}
1476: \frac{p(T,\mu_q,\mu_I)}{T^4} &\simeq& G^{(1)}(T)
1477: +G^{(3)}(T) \frac{1}{3} \left(2\cosh\left(\frac{2\mu_I}{T}\right)+1 \right)
1478: \nonumber \\
1479: && +F^{(2)}(T) \cosh \left( \frac{3\mu_q}{T} \right)
1480: \cosh \left( \frac{\mu_I}{T} \right) \\
1481: &&
1482: +F^{(4)}(T) \frac{1}{2}\cosh \left( \frac{3\mu_q}{T} \right) \left[
1483: \cosh \left( \frac{\mu_I}{T} \right) +\cosh \left( \frac{3\mu_I}{T} \right)
1484: \right]~~, \nonumber
1485: \label{eq:pHRGmu}
1486: \end{eqnarray}
1487: where $G^{(1)}, G^{(3)}, F^{(2)}$ and $F^{(4)}$ are the contributions to
1488: the pressure at $\mu_q=\mu_I=0$ arising from isosinglet mesons
1489: $(\eta, \ldots,~ [B_i=0, I_{3i}=0])$,
1490: isotriplet mesons $(\pi, \ldots,~[B_i=0, I_{3i}=\{0, \pm1\}])$,
1491: isodoublet baryons $(n, p, \ldots,~[B_i=\pm1, I_{3i}=\{\pm1/2\}])$ and
1492: isoquartet baryons $(\Delta, \ldots,~[B_i=\pm1, I_{3i}=\{\pm1/2, \pm3/2\}])$,
1493: respectively. These functions contain all the information on the
1494: hadron spectrum in different quantum number channels. Performing the
1495: Taylor expansion of the pressure as well as quark number and isovector
1496: susceptibilities allows to relate these functions to combinations of the
1497: various Taylor expansion coefficients. This way one finds
1498: %\begin{eqnarray}
1499: \begin{equation}
1500: G^{(3)}(T)=\frac{3}{4}c_2^I-c_4^I, \hspace{8mm}
1501: F^{(2)}(T)=\frac{5}{18}c_2 -\frac{2}{3}c_4^I, \hspace{8mm}
1502: F^{(4)}(T)= -\frac{1}{18}c_2 +\frac{2}{3}c_4^I~~.
1503: \label{eq:gf}
1504: %\end{eqnarray}
1505: \end{equation}
1506: The various contributions to the pressure are shown
1507: in Fig.~(\ref{fig:pHRGmu})a for $\mu_q =0$. With increasing quark chemical
1508: potential the relative weight of hadrons in different quantum number
1509: channels changes. As expected the baryonic component becomes more
1510: important with increasing $\mu_q$ (Fig.~(\ref{fig:pHRGmu})b and c). We find
1511: that for $\mu_q/T \gsim 0.6$ the baryonic sector gives the dominant
1512: contribution.
1513:
1514: \begin{figure}[tb]
1515: \begin{center}
1516: \begin{minipage}[c][5.2cm][c]{5.0cm}
1517: \begin{center}
1518: \epsfig{file=prsvTHRGsim.eps, width=5.0cm}\\[-1mm]
1519: (a)
1520: \end{center}
1521: \end{minipage}
1522: \begin{minipage}[c][5.2cm][c]{5.0cm}
1523: \begin{center}
1524: \epsfig{file=prsvTbyHRGmu04.eps, width=5.0cm}\\[-1mm]
1525: (b)
1526: \end{center}
1527: \end{minipage}
1528: \begin{minipage}[c][5.2cm][c]{5.0cm}
1529: \begin{center}
1530: \epsfig{file=prsvTbyHRGmu08.eps, width=5.0cm}\\[-1mm]
1531: (c)
1532: \end{center}
1533: \end{minipage}
1534: \caption{Contribution of different hadronic channels to the total
1535: pressure $p/T^4$ obtained by using the HRG ansatz. Shown are results
1536: for (a) $\mu_q/T=0$, (b) $\mu_q/T=0.4$ and (c) $\mu_q/T=0.8$.}
1537: \label{fig:pHRGmu}
1538: \end{center}
1539: \end{figure}
1540:
1541:
1542: \section{Reweighting Approach}
1543: \label{sec:rew}
1544:
1545: \begin{figure}
1546: \begin{center}
1547: \epsfig{file=phflvT.eps, width=8.0cm}
1548: \smallskip
1549: \caption{Contour plot of the variance of the phase of the quark
1550: determinant, $\sigma (\theta)$ calculated for
1551: $\theta^{(3)}$ in the $(T/T_0, \mu_q/T)$ plane.
1552: Contour lines for $\sigma ( \theta^{(3)} )$ are
1553: given in steps of $\pi/4$ ranging from $\pi/4$ (lowest curve) to
1554: $2 \pi$.}
1555: \label{fig:phasefl}
1556: \end{center}
1557: \end{figure}
1558:
1559: \begin{figure}[tb]
1560: \begin{center}
1561: \begin{minipage}[c][7.8cm][c]{7.4cm}
1562: \begin{center}
1563: \epsfig{file=chiqmurew.eps, width=7.4cm}\\[-1mm]
1564: (a)
1565: \end{center}
1566: \end{minipage}
1567: \begin{minipage}[c][7.8cm][c]{7.4cm}
1568: \begin{center}
1569: \epsfig{file=chiImurew.eps, width=7.4cm}\\[-1mm]
1570: (b)
1571: \end{center}
1572: \end{minipage}
1573: \caption{Susceptibilities $\chi_q/T^2$ (left) and $\chi_I/T^2$ (right) for
1574: various $\mu_q/T$ ranging from $\mu_q/T=0$ (lowest curve) rising in
1575: steps of 0.2 to $\mu_q/T=1$. Results are obtained
1576: by a combination of reweighting (solid lines and data points) and from
1577: a $6^{th}$-order Taylor expansion (dashed lines).}
1578: \label{fig:chirew}
1579: \end{center}
1580: \end{figure}
1581:
1582: An alternative to a strict Taylor expansion of thermodynamic observables
1583: in terms of $\mu_q/T$ is the reweighting approach. Here the
1584: dependence of the grand potential on the quark chemical potential
1585: is included in the calculation of observables, $X$, by
1586: shifting the $\mu_q$-dependent piece of the QCD action into the
1587: calculation of expectation values rather than taking it into account
1588: in the statistical
1589: weights used for the generation of gauge field configurations.
1590: This reweighting approach has been used to analyze the
1591: thermodynamics of QCD at
1592: non-zero chemical potential \cite{FKS,Fodoreos}. Within this approach
1593: thermodynamic observables $X(\beta,\mu)$ are estimated via the expression
1594: \begin{equation}
1595: \langle X\rangle_{(\beta,\mu)}=
1596: {{\langle X\; e^{{n_{\rm f}\over4}\Delta\ln {\rm det}M}
1597: e^{-\Delta S_g}\rangle_{(\beta_0,0)}}\over
1598: {\langle e^{{n_{\rm f}\over4}\Delta\ln {\rm det}M}
1599: e^{-\Delta S_g} \rangle_{(\beta_0,0)}}} \quad ,
1600: \label{eq:reweight}
1601: \end{equation}
1602: where $\Delta\ln\mbox{det}M\equiv \ln{\rm det}M(\mu) -\ln{\rm det}M(0)$
1603: and $\Delta S_g \equiv S_g(\beta)-S_g(\beta_0)$ is the difference of the
1604: gluonic part of the QCD action.
1605: The expectation values on the RHS of Eq.~(\ref{eq:reweight}) are obtained
1606: in simulations at $(\beta_0,0)$. In \cite{us1} we implemented a version of
1607: Eq.~(\ref{eq:reweight}) in which the reweighting factor $\Delta\ln {\rm det}M$
1608: as well as the operator $X$ itself have been replaced by a Taylor series
1609: about $\mu=0$. The advantage over an exact evaluation of $\mbox{det} M$
1610: \cite{Fodor1} clearly
1611: is that the required expressions are calculable with relatively little
1612: computational effort even on large lattices. In our initial study we
1613: performed the expansion consistently up to and including ${\cal O}(\mu^2)$.
1614: Here we extend
1615: this analysis by expanding $\ln\mbox{det}M$ up to and including
1616: terms of ${\cal O}(\mu^6)$. Unlike the direct
1617: evaluation of thermodynamic observables in terms of a Taylor expansion
1618: up to a certain order the reweighting approach with a Taylor expanded weight
1619: factor also includes effects of higher orders in $\mu_q/T$ which
1620: are partially resummed in the exponentiated observables
1621: $\exp \{ {n_{\rm f}\over4}\Delta\ln {\rm det}M \}$.
1622:
1623: The effectiveness of any reweighting approach strongly depends on the overlap
1624: between the ensemble simulated at $(\beta_0, \mu_0=0)$ and that corresponding
1625: to the true equilibrium state at $(\beta, \mu)$ which one wants to analyze.
1626: This can be judged by inspecting the average phase factor of the
1627: complex valued quark determinant, $\langle
1628: e^{i\theta}\rangle_{(\beta_0,0)}$, where the quark determinant is written
1629: as $\mbox{det}M=\vert\mbox{det}M\vert e^{i\theta}$.
1630: Reweighting loses its reliability once $\langle e^{i\theta}\rangle\ll1$
1631: as both expectation values appearing in the numerator and denominator
1632: of Eq.~(\ref{eq:reweight}) then become difficult to control \cite{Shinji}.
1633: In our approach we estimate the phase factor via the variance
1634: of the phase $\theta$,
1635: $\sigma(\theta) = \sqrt{\langle \theta^2 \rangle - \langle \theta\rangle^2}$,
1636: where we approximate the phase by its Taylor expansion up to
1637: ${\cal O}(\mu^{2n-1})$,
1638: \begin{equation}
1639: \theta^{(n)}=
1640: {{n_{\rm f}\over4}}\mbox{Im}\sum_{j=1}^n{\mu^{2j-1}\over{(2j-1)!}}
1641: {{\partial^{2j-1}\ln\mbox{det}M}\over{\partial\mu^{2j-1}}} \quad .
1642: \end{equation}
1643: As discussed and shown in Fig.~6 of \cite{Shinji}, the value of $\mu_q/T$
1644: for which
1645: the standard deviation of $\theta^{(n)}$ exceeds $\pi / 2$ is a reasonable
1646: criterion for judging the applicability of reweighting in our
1647: simulated systems. In Fig.~(\ref{fig:phasefl}) we show contour lines for
1648: the variance of $\theta^{(3)}$. All contour lines
1649: drop dramatically in the vicinity of $T_0$;
1650: the contour corresponding to $\sigma(\theta^{(3)}) = \pi/2$ yields
1651: $\mu_q/T\approx 1.5$ at $T\simeq1.2T_{0}$ and
1652: reaches a minimum value of about $0.6$ at $T\simeq0.9T_{0}$. Reweighting
1653: is thus much easier to control in
1654: the high temperature phase than in the hadronic phase.
1655:
1656: In Fig.~(\ref{fig:chirew}) we plot quark number and isovector
1657: susceptibilities for various $\mu_q/T$
1658: calculated by using reweighting where possible.
1659: The operators needed for the susceptibilities are calculated with an error
1660: of $O(\mu^6)$,
1661: the reweighting factor $\Delta\ln\mbox{det}M$ with error $O(\mu^8)$.
1662: Dashed lines show a comparison
1663: with the results obtained from direct Taylor expansion of the susceptibilities
1664: up to and including ${\cal O}(\mu_q^4$) (see Fig.~(\ref{fig:chiIq})).
1665:
1666: As anticipated above, for $T\lsim T_{0}$ reweighting becomes difficult
1667: to control for $\mu_q/T>0.6$. However, where reweighting appears to be
1668: statistically under control,
1669: it agrees well with the direct Taylor expansion and even shows similar
1670: features in the statistical errors; {\it i.e.} the signal
1671: is much noisier in the vicinity of $T_{0}$ for $\chi_q$ than for $\chi_I$.
1672:
1673:
1674: \section{Conclusions}
1675:
1676: We have extended our analysis of the thermodynamics of two flavor QCD at
1677: non-zero quark chemical potential to the $6^{th}$ order in a Taylor
1678: expansion around $\mu_q/T = 0$. We find clear evidence for a rapid
1679: transition from a low temperature hadronic phase to a high temperature
1680: quark-gluon plasma phase which is signalled by large fluctuations in the
1681: quark number density and the chiral condensate. The transition temperature
1682: shifts to smaller values with increasing quark chemical
1683: potential. Above $T_0$ the Taylor expansion coefficients and bulk thermodynamic
1684: observables agree with qualitative features of the perturbative high
1685: temperature expansion and approach the ideal gas limit to within
1686: $\sim 20\%$ for $T\gsim 1.5 T_0$. Thermodynamics in the low temperature
1687: phase agrees well with predictions based on a hadron resonance gas for
1688: temperatures $T\lsim 0.96 T_0$ and $\mu_q/T\lsim 1$.
1689:
1690: From the analysis of bulk thermodynamic observables alone we cannot
1691: provide strong evidence for the existence of a $2^{nd}$ order phase
1692: transition point in the QCD phase diagram. At present we cannot rule out
1693: the transition being a rapid crossover in the entire parameter space
1694: covered by our analysis. In particular, we have
1695: shown that large fluctuations in the quark number and the chiral
1696: condensate are consistent with expectations based on a hadron resonance
1697: gas. The current estimates on the radius of convergence of the Taylor
1698: expansion favor a critical value of the quark chemical potential close
1699: to $\mu_q \approx T_0$. The good agreement of the expansion coefficients
1700: with those of a hadron gas, however, prohibit any firm conclusion
1701: on the location and even on the existence of the chiral critical point.
1702:
1703: Likewise, we cannot rule out that a $2^{nd}$ order transition occurs at
1704: temperatures closer to $T_0$ than the largest value in the hadronic phase,
1705: $T=0.96 T_0$ which we have analyzed here.
1706: In order to improve on the current analysis it would be important to
1707: perform calculations at smaller quark masses in a narrower temperature
1708: interval around $T_0$. An improvement over the current statistical
1709: errors on the $6^{th}$ order coefficient as well as the analysis of
1710: higher order expansion coefficients with high statistics is needed.
1711:
1712:
1713: \vfill
1714: \newpage
1715:
1716:
1717: \section*{Acknowledgments}
1718: Numerical work was performed using APEmille computers in Swansea,
1719: supported by PPARC grant PPA/G/S/1999/00026, and Bielefeld.
1720: SJH acknowledges support from PPARC.
1721: Work of the Bielefeld group has been supported partly through the DFG
1722: under grant FOR 339/2-1, the DFG funded Graduate School GRK 881 and a grant
1723: of the BMBF under contract no. 06BI102. KR acknowledges partial support by
1724: KBN under contract 2PO3 (06925).
1725:
1726: \appendix
1727: \section{Appendix: Taylor expansion coefficients}
1728: \label{sec:appA}
1729:
1730: In this appendix, we derive some equations which are used in the
1731: calculation of the various thermodynamic quantities and expansion
1732: coefficients of the Taylor series presented in this study.
1733: The partition function is given by
1734: \begin{eqnarray}
1735: {\cal Z} = \int
1736: {\cal D}U (\det M)^{n_{\rm f}/4} e^{-S_g} \quad ,
1737: \label{eq:partition}
1738: \end{eqnarray}
1739: with $U\in SU(3)$. The expectation value of a physical quantity,
1740: $\left\langle {\cal O} \right\rangle$ is then obtained as
1741: \begin{eqnarray}
1742: \left\langle {\cal O} \right\rangle
1743: \hspace{-4mm} &=& \hspace{-4mm}
1744: \frac{1}{\cal Z} \int {\cal D}U {\cal O}
1745: (\det M)^{n_{\rm f}/4} e^{-S_g} \quad ,
1746: \end{eqnarray}
1747: and its derivatives with respect to quark chemical potential and quark
1748: mass are given by,
1749: \begin{eqnarray}
1750: \frac{\partial \langle {\cal O} \rangle}{\partial \mu}
1751: \hspace{-4mm} &=& \hspace{-4mm}
1752: \left\langle
1753: \frac{\partial {\cal O}}{\partial \mu} \right\rangle
1754: + \frac{n_{\rm f}}{4} \left(
1755: \left\langle {\cal O} \;
1756: \frac{\partial (\ln \det M)}{\partial \mu} \right\rangle
1757: -\left\langle {\cal O} \right\rangle
1758: \left\langle
1759: \frac{\partial (\ln \det M)}{\partial \mu} \right\rangle \right) \quad ,
1760: \\ \nonumber \\
1761: \frac{\partial \langle {\cal O} \rangle}{\partial m}
1762: \hspace{-4mm} &=& \hspace{-4mm}
1763: \left\langle
1764: \frac{\partial {\cal O}}{\partial m} \right\rangle
1765: + \frac{n_{\rm f}}{4} \left(
1766: \left\langle {\cal O} \;
1767: \frac{\partial (\ln \det M)}{\partial m} \right\rangle
1768: -\left\langle {\cal O} \right\rangle
1769: \left\langle
1770: \frac{\partial (\ln \det M)}{\partial m} \right\rangle \right) \nonumber \\
1771: \hspace{-4mm} &=& \hspace{-4mm}
1772: \left\langle
1773: \frac{\partial {\cal O}}{\partial m} \right\rangle
1774: + \frac{n_{\rm f}}{4} \left(
1775: \left\langle {\cal O} \;
1776: {\rm tr} M^{-1} \right\rangle
1777: -\left\langle {\cal O} \right\rangle
1778: \left\langle {\rm tr} M^{-1} \right\rangle \right) \quad .
1779: \end{eqnarray}
1780: Here we use $m$ as the dimensionless quark mass value instead of $ma$,
1781: and also $\mu = \mu_q a$ for the dimensionless quark chemical potential.
1782: The temperature is $T=(N_{\tau} a)^{-1}$ and the volume is
1783: $V=(N_{\sigma} a)^3$.
1784: Moreover, we introduce for simplification,
1785: \begin{eqnarray}
1786: {\cal C}_n = \frac{n_{\rm f}}{4} \frac{\partial^n {\rm tr} M^{-1}}
1787: {\partial \mu^n}
1788: = \frac{n_{\rm f}}{4} \frac{\partial^{n+1} \ln \det M}
1789: {\partial \mu^n \partial m} \quad , \quad
1790: {\cal D}_n = \frac{n_{\rm f}}{4} \frac{\partial^n \ln \det M}
1791: {\partial \mu^n} \quad .
1792: \label{eq:basic}
1793: \end{eqnarray}
1794: All Taylor expansion coefficients used in this paper can be expressed
1795: in terms of expectation values of certain combinations of ${\cal C}_n$
1796: and ${\cal D}_n$. The required derivatives of $\ln \det M$ and
1797: ${\rm tr}M^{-1}$ are explicitly given in the following.
1798:
1799: \paragraph{Derivatives of \boldmath $\ln \det M$: }
1800: \begin{eqnarray}
1801: \frac{\partial \ln \det M}{\partial \mu}
1802: \hspace{-4mm} &=& \hspace{-4mm}
1803: {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu} \right) ,
1804: \label{eq:dermu1} \\
1805: \frac{\partial^2 \ln \det M}{\partial \mu^2}
1806: \hspace{-4mm} &=& \hspace{-4mm}
1807: {\rm tr} \left( M^{-1} \frac{\partial^2 M}{\partial \mu^2} \right)
1808: - {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1809: M^{-1} \frac{\partial M}{\partial \mu} \right) ,
1810: \label{eq:dermu2} \\
1811: %
1812: \frac{\partial^3 \ln \det M}{\partial \mu^3}
1813: \hspace{-4mm} &=& \hspace{-4mm}
1814: {\rm tr} \left( M^{-1} \frac{\partial^3 M}{\partial \mu^3} \right)
1815: -3 {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1816: M^{-1} \frac{\partial^2 M}{\partial \mu^2} \right)
1817: \nonumber \\ && \hspace{-4mm}
1818: +2 {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1819: M^{-1} \frac{\partial M}{\partial \mu}
1820: M^{-1} \frac{\partial M}{\partial \mu} \right) ,
1821: \label{eq:dermu3} \\
1822: %
1823: \frac{\partial^4 \ln \det M}{\partial \mu^4}
1824: \hspace{-4mm} &=& \hspace{-4mm}
1825: {\rm tr} \left( M^{-1} \frac{\partial^4 M}{\partial \mu^4} \right)
1826: -4 {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1827: M^{-1} \frac{\partial^3 M}{\partial \mu^3} \right) \nonumber \\
1828: && \hspace{-2cm}
1829: -3 {\rm tr} \left( M^{-1} \frac{\partial^2 M}{\partial \mu^2}
1830: M^{-1} \frac{\partial^2 M}{\partial \mu^2} \right)
1831: +12 {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1832: M^{-1} \frac{\partial M}{\partial \mu}
1833: M^{-1} \frac{\partial^2 M}{\partial \mu^2} \right) \nonumber \\
1834: && \hspace{-2cm}
1835: -6 {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1836: M^{-1} \frac{\partial M}{\partial \mu}
1837: M^{-1} \frac{\partial M}{\partial \mu}
1838: M^{-1} \frac{\partial M}{\partial \mu} \right) ,
1839: \label{eq:dermu4} \\
1840: %
1841: \frac{\partial^5 \ln \det M}{\partial \mu^5}
1842: \hspace{-4mm} &=& \hspace{-4mm}
1843: {\rm tr} \left( M^{-1} \frac{\partial^5 M}{\partial \mu^5} \right)
1844: -5 {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1845: M^{-1} \frac{\partial^4 M}{\partial \mu^4} \right) \nonumber \\
1846: && \hspace{-25mm}
1847: -10 {\rm tr} \left( M^{-1} \frac{\partial^2 M}{\partial \mu^2}
1848: M^{-1} \frac{\partial^3 M}{\partial \mu^3} \right)
1849: +20 {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1850: M^{-1} \frac{\partial M}{\partial \mu}
1851: M^{-1} \frac{\partial^3 M}{\partial \mu^3} \right) \nonumber \\
1852: && \hspace{-25mm}
1853: +30 {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1854: M^{-1} \frac{\partial^2 M}{\partial \mu^2}
1855: M^{-1} \frac{\partial^2 M}{\partial \mu^2} \right) \nonumber \\
1856: && \hspace{-25mm}
1857: -60 {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1858: M^{-1} \frac{\partial M}{\partial \mu}
1859: M^{-1} \frac{\partial M}{\partial \mu}
1860: M^{-1} \frac{\partial^2 M}{\partial \mu^2} \right) \nonumber \\
1861: && \hspace{-25mm}
1862: +24 {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1863: M^{-1} \frac{\partial M}{\partial \mu}
1864: M^{-1} \frac{\partial M}{\partial \mu}
1865: M^{-1} \frac{\partial M}{\partial \mu}
1866: M^{-1} \frac{\partial M}{\partial \mu} \right) ,
1867: \label{eq:dermu5} \\
1868: %
1869: \frac{\partial^6 \ln \det M}{\partial \mu^6}
1870: \hspace{-4mm} &=& \hspace{-4mm}
1871: {\rm tr} \left( M^{-1} \frac{\partial^6 M}{\partial \mu^6} \right)
1872: -6 {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1873: M^{-1} \frac{\partial^5 M}{\partial \mu^5} \right) \nonumber \\
1874: && \hspace{-25mm}
1875: -15 {\rm tr} \left( M^{-1} \frac{\partial^2 M}{\partial \mu^2}
1876: M^{-1} \frac{\partial^4 M}{\partial \mu^4} \right)
1877: -10 {\rm tr} \left( M^{-1} \frac{\partial^3 M}{\partial \mu^3}
1878: M^{-1} \frac{\partial^3 M}{\partial \mu^3} \right) \nonumber \\
1879: && \hspace{-25mm}
1880: +30 {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1881: M^{-1} \frac{\partial M}{\partial \mu}
1882: M^{-1} \frac{\partial^4 M}{\partial \mu^4} \right)
1883: +60 {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1884: M^{-1} \frac{\partial^2 M}{\partial \mu^2}
1885: M^{-1} \frac{\partial^3 M}{\partial \mu^3} \right) \nonumber \\
1886: && \hspace{-25mm}
1887: +60 {\rm tr} \left( M^{-1} \frac{\partial^2 M}{\partial \mu^2}
1888: M^{-1} \frac{\partial M}{\partial \mu}
1889: M^{-1} \frac{\partial^3 M}{\partial \mu^3} \right)
1890: +30 {\rm tr} \left( M^{-1} \frac{\partial^2 M}{\partial \mu^2}
1891: M^{-1} \frac{\partial^2 M}{\partial \mu^2}
1892: M^{-1} \frac{\partial^2 M}{\partial \mu^2} \right) \nonumber \\
1893: && \hspace{-25mm}
1894: -120 {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1895: M^{-1} \frac{\partial M}{\partial \mu}
1896: M^{-1} \frac{\partial M}{\partial \mu}
1897: M^{-1} \frac{\partial^3 M}{\partial \mu^3} \right) \nonumber \\
1898: && \hspace{-25mm}
1899: -180 {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1900: M^{-1} \frac{\partial M}{\partial \mu}
1901: M^{-1} \frac{\partial^2 M}{\partial \mu^2}
1902: M^{-1} \frac{\partial^2 M}{\partial \mu^2} \right) \nonumber \\
1903: && \hspace{-25mm}
1904: -90 {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1905: M^{-1} \frac{\partial^2 M}{\partial \mu^2}
1906: M^{-1} \frac{\partial M}{\partial \mu}
1907: M^{-1} \frac{\partial^2 M}{\partial \mu^2} \right) \nonumber \\
1908: && \hspace{-25mm}
1909: +360 {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1910: M^{-1} \frac{\partial M}{\partial \mu}
1911: M^{-1} \frac{\partial M}{\partial \mu}
1912: M^{-1} \frac{\partial M}{\partial \mu}
1913: M^{-1} \frac{\partial^2 M}{\partial \mu^2} \right) \nonumber \\
1914: && \hspace{-25mm}
1915: -120 {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1916: M^{-1} \frac{\partial M}{\partial \mu}
1917: M^{-1} \frac{\partial M}{\partial \mu}
1918: M^{-1} \frac{\partial M}{\partial \mu}
1919: M^{-1} \frac{\partial M}{\partial \mu}
1920: M^{-1} \frac{\partial M}{\partial \mu} \right) .
1921: \label{eq:dermu6}
1922: \end{eqnarray}
1923:
1924: \paragraph{Derivatives of \boldmath ${\rm tr} M^{-1}$: }
1925: \begin{eqnarray}
1926: \frac{\partial {\rm tr} M^{-1}}{\partial \mu}
1927: \hspace{-4mm} &=& \hspace{-4mm}
1928: - {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1929: M^{-1} \right) \\
1930: \frac{\partial^2 {\rm tr} M^{-1}}{\partial \mu^2}
1931: \hspace{-4mm} &=& \hspace{-4mm}
1932: - {\rm tr} \left( M^{-1} \frac{\partial^2 M}{\partial \mu^2}
1933: M^{-1} \right)
1934: + 2 {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1935: M^{-1} \frac{\partial M}{\partial \mu} M^{-1} \right) \\
1936: \frac{\partial^3 {\rm tr} M^{-1}}{\partial \mu^3}
1937: \hspace{-4mm} &=& \hspace{-4mm}
1938: - {\rm tr} \left( M^{-1} \frac{\partial^3 M}{\partial \mu^3}
1939: M^{-1} \right)
1940: +3 {\rm tr} \left( M^{-1} \frac{\partial^2 M}{\partial \mu^2}
1941: M^{-1} \frac{\partial M}{\partial \mu} M^{-1} \right) \\
1942: && \hspace{-2cm}
1943: +3 {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1944: M^{-1} \frac{\partial^2 M}{\partial \mu^2} M^{-1} \right)
1945: -6 {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1946: M^{-1} \frac{\partial M}{\partial \mu} M^{-1}
1947: \frac{\partial M}{\partial \mu} M^{-1} \right) \nonumber \\
1948: \frac{\partial^4 {\rm tr} M^{-1}}{\partial \mu^4}
1949: \hspace{-4mm} &=& \hspace{-4mm}
1950: - {\rm tr} \left( M^{-1} \frac{\partial^4 M}{\partial \mu^4}
1951: M^{-1} \right)
1952: +4 {\rm tr} \left( M^{-1} \frac{\partial^3 M}{\partial \mu^3}
1953: M^{-1} \frac{\partial M}{\partial \mu} M^{-1} \right) \\
1954: && \hspace{-2cm}
1955: +6 {\rm tr} \left( M^{-1} \frac{\partial^2 M}{\partial \mu^2}
1956: M^{-1} \frac{\partial^2 M}{\partial \mu^2} M^{-1} \right)
1957: +4 {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1958: M^{-1} \frac{\partial^3 M}{\partial \mu^3} M^{-1} \right) \nonumber \\
1959: && \hspace{-2cm}
1960: -12 {\rm tr} \left( M^{-1} \frac{\partial^2 M}{\partial \mu^2}
1961: M^{-1} \frac{\partial M}{\partial \mu} M^{-1}
1962: \frac{\partial M}{\partial \mu} M^{-1} \right) \nonumber \\
1963: && \hspace{-2cm}
1964: -12 {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1965: M^{-1} \frac{\partial^2 M}{\partial \mu^2} M^{-1}
1966: \frac{\partial M}{\partial \mu} M^{-1} \right) \nonumber \\
1967: && \hspace{-2cm}
1968: -12 {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1969: M^{-1} \frac{\partial M}{\partial \mu} M^{-1}
1970: \frac{\partial^2 M}{\partial \mu^2} M^{-1} \right) \nonumber \\
1971: && \hspace{-2cm}
1972: +24 {\rm tr} \left( M^{-1} \frac{\partial M}{\partial \mu}
1973: M^{-1} \frac{\partial M}{\partial \mu} M^{-1}
1974: \frac{\partial M}{\partial \mu} M^{-1}
1975: \frac{\partial M}{\partial \mu} M^{-1} \right) \nonumber
1976: \end{eqnarray}
1977:
1978: \noindent
1979: Having defined the explicit representation of ${\cal C}_n$ and ${\cal D}_n$
1980: we now can proceed to define the expansion coefficients for various
1981: thermodynamic quantities discussed in this paper.
1982:
1983: \paragraph{Pressure \boldmath $(p)$ \unboldmath :}
1984: The pressure is obtained from the logarithm of the QCD partition function.
1985: Its expansion is defined in Eqs.~(\ref{eq:p}) and (\ref{eq:cn}). The leading
1986: expansion coefficient $c_0$ is given by the pressure calculated at $\mu_q=0$.
1987: All higher order expansion coefficients are given in terms of
1988: derivatives of $\ln {\cal Z}$.
1989: \begin{eqnarray}
1990: \frac{p}{T^4} \equiv \Omega =
1991: \frac{1}{VT^3} \ln {\cal Z} = \sum_{n=0}^{\infty} c_n
1992: \left( \frac{\mu_q}{T} \right)^n \quad ,
1993: \end{eqnarray}
1994: with
1995: \begin{equation}
1996: c_n = \frac{1}{n! VT^3} \frac{\partial^n \ln {\cal Z}}{\partial (\mu_q/T)^n}
1997: \biggr|_{\mu=0}
1998: \end{equation}
1999: To generate the expansion we first consider derivatives of $\ln {\cal Z}$
2000: for $\mu \ne 0$. For the first derivative we find
2001: \begin{equation}
2002: \frac{\partial \ln {\cal Z}}{\partial \mu} \equiv {\cal A}_1 =
2003: \left\langle {\cal D}_1 \right\rangle \quad .
2004: \end{equation}
2005: Higher order derivatives are generated using the relation
2006: \begin{eqnarray}
2007: \frac{\partial {\cal A}_n}{\partial \mu}
2008: = {\cal A}_{n+1} - {\cal A}_n {\cal A}_1 \quad ,
2009: \label{eq:An1}
2010: \end{eqnarray}
2011: where ${\cal A}_n$ is defined as
2012: \begin{equation}
2013: {\cal A}_n \equiv \left\langle
2014: \exp\{ -{\cal D}_0 \} \frac{\partial^n \exp\{ {\cal D}_0 \} }{\partial \mu^n}
2015: \right\rangle \quad .
2016: \label{eq:an}
2017: \end{equation}
2018: With this we can generate higher order derivatives of $\ln {\cal Z}$
2019: iteratively using
2020: \begin{equation}
2021: \partial^{n+1} \ln {\cal Z}/\partial \mu^{n+1} =
2022: \partial^{n} {\cal A}_1 /\partial \mu^{n} \quad .
2023: \label{eq:Zn1}
2024: \end{equation}
2025: Explicitly we find from Eq.~(\ref{eq:an})
2026: \begin{eqnarray}
2027: %{\cal A}_1 \hspace{-4mm} &=& \hspace{-4mm}
2028: %\left\langle {\cal D}_1 \right\rangle,
2029: %\\
2030: {\cal A}_2 \hspace{-4mm} &=& \hspace{-4mm}
2031: \left\langle {\cal D}_2 \right\rangle
2032: +\left\langle {\cal D}_1^2 \right\rangle,
2033: \\
2034: {\cal A}_3 \hspace{-4mm} &=& \hspace{-4mm}
2035: \left\langle {\cal D}_3 \right\rangle
2036: +3\left\langle {\cal D}_2 {\cal D}_1 \right\rangle
2037: +\left\langle {\cal D}_1^3 \right\rangle,
2038: \\
2039: {\cal A}_4 \hspace{-4mm} &=& \hspace{-4mm}
2040: \left\langle {\cal D}_4 \right\rangle
2041: +4\left\langle {\cal D}_3 {\cal D}_1 \right\rangle
2042: +3\left\langle {\cal D}_2^2 \right\rangle
2043: +6\left\langle {\cal D}_2 {\cal D}_1^2 \right\rangle
2044: +\left\langle {\cal D}_1^4 \right\rangle,
2045: \\
2046: {\cal A}_5 \hspace{-4mm} &=& \hspace{-4mm}
2047: \left\langle {\cal D}_5 \right\rangle
2048: +5\left\langle {\cal D}_4 {\cal D}_1 \right\rangle
2049: +10\left\langle {\cal D}_3 {\cal D}_2 \right\rangle
2050: +10\left\langle {\cal D}_3 {\cal D}_1^2 \right\rangle
2051: +15\left\langle {\cal D}_2^2 {\cal D}_1 \right\rangle
2052: +10\left\langle {\cal D}_2 {\cal D}_1^3 \right\rangle
2053: \nonumber \\ && \hspace{-10mm}
2054: +\left\langle {\cal D}_1^5 \right\rangle,
2055: \\
2056: {\cal A}_6 \hspace{-4mm} &=& \hspace{-4mm}
2057: \left\langle {\cal D}_6 \right\rangle
2058: +6\left\langle {\cal D}_5 {\cal D}_1 \right\rangle
2059: +15\left\langle {\cal D}_4 {\cal D}_2 \right\rangle
2060: +10\left\langle {\cal D}_3^2 \right\rangle
2061: +15\left\langle {\cal D}_4 {\cal D}_1^2 \right\rangle
2062: +60\left\langle {\cal D}_3 {\cal D}_2 {\cal D}_1 \right\rangle
2063: \nonumber \\ && \hspace{-5mm}
2064: +15\left\langle {\cal D}_2^3 \right\rangle
2065: +20\left\langle {\cal D}_3 {\cal D}_1^3 \right\rangle
2066: +45\left\langle {\cal D}_2^2 {\cal D}_1^2 \right\rangle
2067: +15\left\langle {\cal D}_2 {\cal D}_1^4 \right\rangle
2068: +\left\langle {\cal D}_1^6 \right\rangle \quad .
2069: \end{eqnarray}
2070: From Eq.~(\ref{eq:Zn1}) we then obtain through repeated
2071: application of Eq.~(\ref{eq:An1}),
2072: \begin{eqnarray}
2073: \frac{\partial \ln {\cal Z}}{\partial \mu}
2074: \hspace{-4mm} &=& \hspace{-4mm}
2075: {\cal A}_1 ,
2076: \\
2077: \frac{\partial^2 \ln {\cal Z}}{\partial \mu^2}
2078: \hspace{-4mm} &=& \hspace{-4mm}
2079: {\cal A}_2 - {\cal A}_1^2 ,
2080: \\
2081: \frac{\partial^3 \ln {\cal Z}}{\partial \mu^3}
2082: \hspace{-4mm} &=& \hspace{-4mm}
2083: {\cal A}_3 -3 {\cal A}_2 {\cal A}_1 +2 {\cal A}_1^3 ,
2084: \\
2085: \frac{\partial^4 \ln {\cal Z}}{\partial \mu^4}
2086: \hspace{-4mm} &=& \hspace{-4mm}
2087: {\cal A}_4 -4 {\cal A}_3 {\cal A}_1 -3 {\cal A}_2^2
2088: +12 {\cal A}_2 {\cal A}_1^2 -6 {\cal A}_1^4 ,
2089: \\
2090: \frac{\partial^5 \ln {\cal Z}}{\partial \mu^5}
2091: \hspace{-4mm} &=& \hspace{-4mm}
2092: {\cal A}_5 -5 {\cal A}_4 {\cal A}_1 -10 {\cal A}_3 {\cal A}_2
2093: +20 {\cal A}_3 {\cal A}_1^2 +30 {\cal A}_2^2 {\cal A}_1
2094: -60 {\cal A}_2 {\cal A}_1^3 +24 {\cal A}_1^5 ,
2095: \\
2096: \frac{\partial^6 \ln {\cal Z}}{\partial \mu^6}
2097: \hspace{-4mm} &=& \hspace{-4mm}
2098: {\cal A}_6 -6 {\cal A}_5 {\cal A}_1 -15 {\cal A}_4 {\cal A}_2
2099: -10 {\cal A}_3^2 +30 {\cal A}_4 {\cal A}_1^2
2100: \nonumber \\ && \hspace{-9mm}
2101: +120 {\cal A}_3 {\cal A}_2 {\cal A}_1
2102: +30 {\cal A}_2^3 -120 {\cal A}_3 {\cal A}_1^3
2103: -270 {\cal A}_2^2 {\cal A}_1^2
2104: +360 {\cal A}_2 {\cal A}_1^4 -120 {\cal A}_1^6 \; .
2105: \end{eqnarray}
2106: These relations simplify considerably for $\mu=0$ as all odd
2107: expectation values vanish, {\it i.e.} ${\cal A}_n =0$ for $n$ odd. In fact,
2108: $\partial^n (\ln \det M)/\partial \mu^n$
2109: is strictly real for $n$ even and pure imaginary for $n$ odd.
2110: Using this property, the odd derivatives of the pressure vanish
2111: and also the even derivatives become rather simple. This defines
2112: the expansion coefficients $c_n$ introduced in Eqs.~(\ref{eq:p}) and
2113: (\ref{eq:cn}),
2114: \begin{eqnarray}
2115: c_2 &\equiv& \frac{1}{2}\frac{\partial^2 (p/T^4)}{\partial (\mu_{q}/T)^2}
2116: \biggr|_{\mu_q=0} =
2117: \frac{1}{2} \frac{N_{\tau}}{N_{\sigma}^3}\; {\cal A}_2 \quad , \nonumber \\
2118: c_4 &\equiv& \frac{1}{4!} \frac{\partial^4 (p/T^4)}{\partial (\mu_{q}/T)^4}
2119: \biggr|_{\mu_q=0} =
2120: \frac{1}{4!} \frac{1}{N_{\sigma}^3 N_{\tau}} ({\cal A}_4 -3 {\cal A}_2^2) \quad ,
2121: \nonumber \\
2122: c_6 &\equiv& \frac{1}{6!} \frac{\partial^6 (p/T^4)}{\partial (\mu_{q}/T)^6}
2123: \biggr|_{\mu_q=0} =
2124: \frac{1}{6!} \frac{1}{N_{\sigma}^3 N_{\tau}^3}
2125: ({\cal A}_6 -15 {\cal A}_4 {\cal A}_2 +30 {\cal A}_2^3) \quad .
2126: \label{eq:dpmu0}
2127: \end{eqnarray}
2128: Here all expectation values ${\cal A}_n$ are now meant to be evaluated at
2129: $\mu =0$.
2130: \paragraph{
2131: %Quark number susceptibility \boldmath $(\chi_q)$ and
2132: Isovector susceptibility $(\chi_I)$\unboldmath :}
2133: While the Taylor expansion of the quark number susceptibility is easily
2134: obtained from that of the pressure we need to introduce the expansion
2135: of the isovector susceptibility. This has been done in Eq.~(\ref{eq:chiI}).
2136: More explicitly the isovector susceptibility is given by
2137: \begin{eqnarray}
2138: \frac{\chi_{I}}{T^2}
2139: \hspace{-4mm} &=& \hspace{-4mm}
2140: \frac{N_{\tau}}{N_{\sigma}^3} \left(
2141: \frac{\partial^2 \ln {\cal Z}}{\partial \bar{\mu}_u^2}
2142: -\frac{\partial^2 \ln {\cal Z}}{\partial \bar{\mu}_u
2143: \partial \bar{\mu}_d}
2144: - \frac{\partial^2 \ln {\cal Z}}{\partial \bar{\mu}_u
2145: \partial \bar{\mu}_d}
2146: + \frac{\partial^2 \ln {\cal Z}}{\partial \bar{\mu}_d^2}
2147: \right) \nonumber \\
2148: && \hspace{-2cm}
2149: = \frac{N_{\tau}}{N_{\sigma}^3} \left[ \sum_{f=u,d} \left\langle
2150: \frac{1}{4}
2151: \frac{\partial^2 (\ln \det M_{f})}{\partial \bar{\mu}_{f}^2}
2152: \right\rangle
2153: + \left\langle \left( \frac{1}{4}
2154: \frac{\partial (\ln \det M_u)}{\partial \bar{\mu}_u}
2155: - \frac{1}{4} \frac{\partial (\ln \det M_d)}
2156: {\partial \bar{\mu}_d} \right)^2 \right\rangle \right.
2157: \nonumber \\
2158: && \hspace{-5mm}
2159: - \left. \left\langle \frac{1}{4}
2160: \frac{\partial (\ln \det M_u)}{\partial \bar{\mu}_u}
2161: - \frac{1}{4}
2162: \frac{\partial (\ln \det M_d)}{\partial \bar{\mu}_d}
2163: \right\rangle^2 \right] \quad .
2164: \end{eqnarray}
2165: Throughout this paper we have considered the case of
2166: $\bar{\mu}_u = \bar{\mu}_d \equiv \mu_q a \equiv \mu$. The isovector
2167: chemical potential $\mu_I$ has been set equal to zero after appropriate
2168: derivatives have been taken.
2169: In the Taylor expansions of $\chi_I$, which is performed
2170: at $\mu=0$ in terms of $\mu_q/T$, the derivatives with respect to $\mu_u$
2171: and $\mu_d$ then become identical, {\it i.e.}
2172: $[\partial^n (\ln \det M_{u(d)})/\partial \bar{\mu}_{u(d)}^n]
2173: (\bar{\mu}_{u(d)}) \equiv [\partial^n (\ln \det M)/\partial \mu^n] (\mu)$.
2174:
2175: The calculation of the isovector susceptibility thus reduces to the
2176: calculation of
2177: \begin{equation}
2178: \frac{\chi_{I}}{T^2}
2179: %\hspace{-4mm} = \hspace{-4mm}
2180: = \frac{N_{\tau}}{2N_{\sigma}^3} \left\langle
2181: \frac{\partial^2 (\ln \det M)}{\partial \mu^2}
2182: \right\rangle \quad .
2183: \end{equation}
2184: To define the expansion of the isovector susceptibility,
2185: $\chi_I$ at fixed $\mu_I=0$ around $\mu_q=0$ we set up an iterative
2186: scheme similar to that introduced for the pressure.
2187: We introduce the additional kernel ${\cal D}_2$ in Eq.~(\ref{eq:an})
2188: to generate ${\cal B}_n$ for arbitrary $\mu_q \ne 0$,
2189: \begin{equation}
2190: {\cal B}_{n+2} \equiv \left\langle
2191: \exp\{ -{\cal D}_0 \}
2192: \frac{\partial^n {\cal D}_2\exp\{ {\cal D}_0 \} }{\partial \mu^n}
2193: \right\rangle \quad .
2194: \label{eq:bn}
2195: \end{equation}
2196: This yields,
2197: \begin{eqnarray}
2198: {\cal B}_2
2199: \hspace{-4mm} &=& \hspace{-4mm}
2200: \left\langle {\cal D}_2 \right\rangle ,
2201: \\
2202: {\cal B}_3
2203: \hspace{-4mm} &=& \hspace{-4mm}
2204: \left\langle {\cal D}_3 \right\rangle
2205: + \left\langle {\cal D}_2 {\cal D}_1 \right\rangle ,
2206: \\
2207: {\cal B}_4
2208: \hspace{-4mm} &=& \hspace{-4mm}
2209: \left\langle {\cal D}_4 \right\rangle
2210: +2 \left\langle {\cal D}_3 {\cal D}_1 \right\rangle
2211: + \left\langle {\cal D}_2^2 \right\rangle
2212: + \left\langle {\cal D}_2 {\cal D}_1^2 \right\rangle ,
2213: \\
2214: {\cal B}_5
2215: \hspace{-4mm} &=& \hspace{-4mm}
2216: \left\langle {\cal D}_5 \right\rangle
2217: +3 \left\langle {\cal D}_4 {\cal D}_1 \right\rangle
2218: +4 \left\langle {\cal D}_3 {\cal D}_2 \right\rangle
2219: +3 \left\langle {\cal D}_3 {\cal D}_1^2 \right\rangle
2220: +3 \left\langle {\cal D}_2^2 {\cal D}_1 \right\rangle
2221: + \left\langle {\cal D}_2 {\cal D}_1^3 \right\rangle
2222: \\
2223: {\cal B}_6
2224: \hspace{-4mm} &=& \hspace{-4mm}
2225: \left\langle {\cal D}_6 \right\rangle
2226: +4 \left\langle {\cal D}_5 {\cal D}_1 \right\rangle
2227: +7 \left\langle {\cal D}_4 {\cal D}_2 \right\rangle
2228: +4 \left\langle {\cal D}_3^2 \right\rangle
2229: +6 \left\langle {\cal D}_4 {\cal D}_1^2 \right\rangle
2230: +16 \left\langle {\cal D}_3 {\cal D}_2 {\cal D}_1 \right\rangle
2231: \nonumber \\ &&
2232: +3 \left\langle {\cal D}_2^3 \right\rangle
2233: +4 \left\langle {\cal D}_3 {\cal D}_1^3 \right\rangle
2234: +6 \left\langle {\cal D}_2^2 {\cal D}_1^2 \right\rangle
2235: + \left\langle {\cal D}_2 {\cal D}_1^4 \right\rangle .
2236: \end{eqnarray}
2237: The derivative of ${\cal B}_n$ with respect to the chemical potential
2238: satisfies,
2239: \begin{equation}
2240: \frac{\partial {\cal B}_n}{\partial \mu}
2241: = {\cal B}_{n+1} - {\cal B}_n {\cal A}_1 \quad .
2242: \end{equation}
2243: Starting with $(N_{\tau}/N_{\sigma}^3) {\cal B}_2 = \chi_I/T^2$,
2244: we obtain the equations for
2245: $\partial^n (\chi_I/T^2) / \partial (\mu_q/T)^n$ iteratively,
2246: and then use again the CP symmetry for $\mu=0$,
2247: i.e. ${\cal A}_n$ and ${\cal B}_n$ are zero for $n$ odd.
2248: Therefore the odd derivatives in the expansion of $\chi_I/T^2$
2249: at $\mu=0$ are zero and
2250: the even derivatives define the expansion coefficients $c_n^I$ used in
2251: Eq.~(\ref{eq:chiI}),
2252: \begin{eqnarray}
2253: c_2^I &\equiv& \frac{1}{2} \frac{\chi_I}{T^2}
2254: \biggr|_{\mu_q=0} =
2255: \frac{1}{2} \frac{N_{\tau}}{N_{\sigma}^3} {\cal B}_2 \quad , \nonumber \\
2256: c_4^I &\equiv& \frac{1}{4!} \frac{\partial^2 (\chi_I/T^2)}{\partial (\mu_{q}/T)^2}
2257: \biggr|_{\mu_q=0} =
2258: \frac{1}{4!}\frac{1}{N_{\sigma}^3 N_{\tau}} ({\cal B}_4 - {\cal B}_2 {\cal A}_2) ,
2259: \nonumber \\
2260: c_6^I &\equiv& \frac{1}{6!} \frac{\partial^4 (\chi_I/T^2)}{\partial (\mu_{q}/T)^4}
2261: \biggr|_{\mu_q=0} =
2262: \frac{1}{6!} \frac{1}{N_{\sigma}^3 N_{\tau}^3}
2263: ({\cal B}_6 -6 {\cal B}_4 {\cal A}_2 - {\cal B}_2 {\cal A}_4
2264: +6 {\cal B}_2 {\cal A}_2^2 ) \quad .
2265: \end{eqnarray}
2266:
2267: \paragraph{ Chiral condensate \boldmath $(\langle \bar{\psi} \psi \rangle)$
2268: and disconnected chiral susceptibility $(\chi_{\bar{\psi} \psi})$\unboldmath :}
2269: We also discuss the Taylor expansion of the chiral condensate and the related
2270: chiral susceptibility,
2271: \begin{eqnarray}
2272: \frac{\langle \bar{\psi} \psi \rangle}{T^3}
2273: \hspace{-4mm} &=& \hspace{-4mm}
2274: \frac{N_\tau^2}{N_{\sigma}^3 } \frac{\partial \ln {\cal Z}}{\partial m}
2275: = \frac{N_\tau^2}{N_{\sigma}^3 } \frac{n_{\rm f}}{4} \left\langle
2276: \frac{\partial \ln \det M}{\partial m} \right\rangle
2277: = \frac{N_\tau^2}{N_{\sigma}^3 } \frac{n_{\rm f}}{4} \left\langle
2278: {\rm tr} M^{-1} \right\rangle \quad ,
2279: \label{eq:chicon} \\
2280: \frac{\chi_{\bar{\psi} \psi}}{T^2}
2281: \hspace{-4mm} &=& \hspace{-4mm}
2282: \frac{N_\tau}{N_{\sigma}^3 } \left( \frac{n_{\rm f}}{4} \right)^2
2283: \left[ \left\langle \left(
2284: {\rm tr} M^{-1} \right)^2 \right\rangle
2285: - \left\langle {\rm tr} M^{-1} \right\rangle^2
2286: \right] \quad .
2287: \label{eq:chisusapp}
2288: \end{eqnarray}
2289: The iterative scheme is similar to that introduced for the isovector
2290: susceptibility but with the generating kernel ${\cal D}_2$ in Eq.~(\ref{eq:bn})
2291: replaced by ${\cal C}_0$. This yields
2292: \begin{eqnarray}
2293: {\cal F}_0
2294: \hspace{-4mm} &=& \hspace{-4mm}
2295: \left\langle {\cal C}_0 \right\rangle, \\
2296: {\cal F}_1
2297: \hspace{-4mm} &=& \hspace{-4mm}
2298: \left\langle {\cal C}_1 \right\rangle
2299: + \left\langle {\cal C}_0 {\cal D}_1 \right\rangle, \\
2300: {\cal F}_2
2301: \hspace{-4mm} &=& \hspace{-4mm}
2302: \left\langle {\cal C}_2 \right\rangle
2303: + 2 \left\langle {\cal C}_1 {\cal D}_1 \right\rangle
2304: + \left\langle {\cal C}_0 {\cal D}_2 \right\rangle
2305: + \left\langle {\cal C}_0 {\cal D}_1^2 \right\rangle, \\
2306: {\cal F}_3
2307: \hspace{-4mm} &=& \hspace{-4mm}
2308: \left\langle {\cal C}_3 \right\rangle
2309: + 3 \left\langle {\cal C}_2 {\cal D}_1 \right\rangle
2310: + 3 \left\langle {\cal C}_1 {\cal D}_2 \right\rangle
2311: + \left\langle {\cal C}_0 {\cal D}_3 \right\rangle
2312: + 3 \left\langle {\cal C}_1 {\cal D}_1^2 \right\rangle
2313: + 3 \left\langle {\cal C}_0 {\cal D}_2 {\cal D}_1 \right\rangle
2314: \nonumber \\ && \hspace{-5mm}
2315: + \left\langle {\cal C}_0 {\cal D}_1^3 \right\rangle \quad , \\
2316: {\cal F}_4
2317: \hspace{-4mm} &=& \hspace{-4mm}
2318: \left\langle {\cal C}_4 \right\rangle
2319: + 4 \left\langle {\cal C}_3 {\cal D}_1 \right\rangle
2320: + 6 \left\langle {\cal C}_2 {\cal D}_2 \right\rangle
2321: + 4 \left\langle {\cal C}_1 {\cal D}_3 \right\rangle
2322: + \left\langle {\cal C}_0 {\cal D}_4 \right\rangle
2323: + 6 \left\langle {\cal C}_2 {\cal D}_1^2 \right\rangle
2324: + 12 \left\langle {\cal C}_1 {\cal D}_2 {\cal D}_1 \right\rangle
2325: \nonumber \\ && \hspace{-5mm}
2326: + 4 \left\langle {\cal C}_0 {\cal D}_3 {\cal D}_1 \right\rangle
2327: + 3 \left\langle {\cal C}_0 {\cal D}_2^2 \right\rangle
2328: + 4 \left\langle {\cal C}_1 {\cal D}_1^3 \right\rangle
2329: + 6 \left\langle {\cal C}_0 {\cal D}_2 {\cal D}_1^2 \right\rangle
2330: + \left\langle {\cal C}_0 {\cal D}_1^4 \right\rangle \quad .
2331: \end{eqnarray}
2332: These expectation values again satisfy
2333: \begin{equation}
2334: \frac{\partial {\cal F}_n}{\partial \mu}
2335: = {\cal F}_{n+1} - {\cal F}_n {\cal A}_1 \quad .
2336: \end{equation}
2337: This gives the expansion coefficients $c_n^{\bar{\psi}\psi}$ for the
2338: chiral condensate defined in Eqs.~(\ref{eq:pbp}) and (\ref{eq:pbpcoef}),
2339: \begin{eqnarray}
2340: c_0^{\bar{\psi}\psi} &\equiv&
2341: \frac{\left\langle \bar{\psi} \psi \right\rangle}{T^3}
2342: \biggr|_{\mu_q =0}
2343: = \frac{N_\tau^2}{N_{\sigma}^3 } {\cal F}_0 \quad , \\
2344: c_2^{\bar{\psi}\psi} &\equiv& \frac{1}{2}
2345: \frac{\partial^2 \left\langle \bar{\psi} \psi \right\rangle /T^3
2346: }{\partial (\mu_{q}/T)^2} \biggr|_{\mu_q=0} =
2347: \frac{1}{2} \frac{1}{N_{\tau}^2} \frac{\partial^2
2348: \left\langle \bar{\psi} \psi \right\rangle/T^3}{\partial \mu^2} \biggr|_{\mu=0}
2349: \nonumber \\
2350: &=& \frac{1}{2}\frac{1}{N_{\sigma}^3 }
2351: \left( {\cal F}_2 - {\cal F}_0 {\cal A}_2 \right) ,
2352: \\
2353: c_4^{\bar{\psi}\psi} &\equiv& \frac{1}{4!}
2354: \frac{\partial^4 \left\langle \bar{\psi} \psi \right\rangle /T^3
2355: }{\partial (\mu_{q}/T)^4} \biggr|_{\mu_q=0} =
2356: \frac{1}{4!} \frac{1}{N_{\tau}^4} \frac{\partial^4
2357: \left\langle \bar{\psi} \psi \right\rangle/T^3}{\partial \mu^4} \biggr|_{\mu=0}
2358: \nonumber \\
2359: &=& \frac{1}{4!} \frac{1}{N_{\sigma}^3 N_{\tau}^2}
2360: \left( {\cal F}_4 -6 {\cal F}_2 {\cal A}_2 - {\cal F}_0 {\cal A}_4
2361: +6 {\cal F}_0 {\cal A}_2^2 \right) ,
2362: \end{eqnarray}
2363: where we used again that
2364: ${\cal A}_n$ and ${\cal F}_n$ are zero for $n$ odd ant $\mu =0$.
2365: Hence the odd derivatives in the expansion vanish. Note that these
2366: expansion coefficients also control the quark mass dependence of the
2367: quark number susceptibility given in Eq.~(\ref{eq:chiqm}). The first
2368: derivative of the isovector susceptibility with respect to the quark mass,
2369: \begin{eqnarray}
2370: \frac{\partial (\chi_I/T^2)}{\partial m/T}
2371: = \frac{1}{N_{\sigma}^3} \left[ \left\langle
2372: \frac{n_{\rm f}}{4}
2373: \frac{\partial^3 (\ln \det M)}{\partial \mu^2 \partial m}
2374: \right\rangle
2375: + \left\langle \left( \frac{n_{\rm f}}{4} \right)^2
2376: \frac{\partial^2 (\ln \det M)}{\partial \mu^2}
2377: {\rm tr} M^{-1} \right\rangle \right.
2378: \nonumber \\ \left.
2379: - \left\langle \frac{n_{\rm f}}{4}
2380: \frac{\partial^2 (\ln \det M)}{\partial \mu^2} \right\rangle
2381: \left\langle \frac{n_{\rm f}}{4} {\rm tr} M^{-1} \right\rangle \right]
2382: \quad ,
2383: \end{eqnarray}
2384: for which we have introduced the Taylor expansion in Eq.~(\ref{eq:chiIm})
2385: requires the calculation of further expectation values ${\cal I}_n$,
2386: \begin{eqnarray}
2387: {\cal I}_2
2388: \hspace{-4mm} &=& \hspace{-4mm}
2389: \left\langle {\cal C}_2 \right\rangle
2390: + \left\langle {\cal C}_0 {\cal D}_2 \right\rangle ,
2391: \\
2392: {\cal I}_3
2393: \hspace{-4mm} &=& \hspace{-4mm}
2394: \left\langle {\cal C}_3 \right\rangle
2395: + \left\langle {\cal C}_2 {\cal D}_1 \right\rangle
2396: + \left\langle {\cal C}_1 {\cal D}_2 \right\rangle
2397: + \left\langle {\cal C}_0 {\cal D}_3 \right\rangle
2398: + \left\langle {\cal C}_0 {\cal D}_2 {\cal D}_1 \right\rangle ,
2399: \\
2400: {\cal I}_4
2401: \hspace{-4mm} &=& \hspace{-4mm}
2402: \left\langle {\cal C}_4 \right\rangle
2403: +2 \left\langle {\cal C}_3 {\cal D}_1 \right\rangle
2404: +2 \left\langle {\cal C}_1 {\cal D}_3 \right\rangle
2405: +2 \left\langle {\cal C}_2 {\cal D}_2 \right\rangle
2406: + \left\langle {\cal C}_2 {\cal D}_1^2 \right\rangle
2407: +2 \left\langle {\cal C}_1 {\cal D}_2 {\cal D}_1 \right\rangle
2408: \nonumber \\ && \hspace{-4mm}
2409: +\left\langle {\cal C}_0 {\cal D}_4 \right\rangle
2410: +2 \left\langle {\cal C}_0 {\cal D}_3 {\cal D}_1 \right\rangle
2411: + \left\langle {\cal C}_0 {\cal D}_2^2 \right\rangle
2412: + \left\langle {\cal C}_0 {\cal D}_2 {\cal D}_1^2 \right\rangle \quad .
2413: \end{eqnarray}
2414: With these and the coefficients ${\cal B}_n$ and ${\cal F}_n$
2415: one finds for the expansion coefficients,
2416: \begin{eqnarray}
2417: c_2^{I,\bar{\psi}\psi} &=& \frac{1}{2}
2418: \frac{\partial (\chi_I/T^2)}{\partial m/T} \biggr|_{\mu=0}
2419: = \frac{1}{2} \frac{1}{N_{\sigma}^3} ({\cal I}_2 - {\cal B}_2 {\cal F}_0 )
2420: \quad ,
2421: \nonumber \\
2422: c_4^{I,\bar{\psi}\psi} &=& \frac{1}{4!}
2423: \frac{\partial^3 (\chi_I/T^2)}{\partial (\mu_{q}/T)^2 \partial m/T}
2424: \biggr|_{\mu=0
2425: }
2426: \nonumber \\
2427: &=& \frac{1}{4!}
2428: \frac{1}{N_{\sigma}^3 N_{\tau}^2}
2429: ({\cal I}_4 - {\cal I}_2 {\cal A}_2
2430: - {\cal B}_4 {\cal F}_0 - {\cal B}_2 {\cal F}_2
2431: +2 {\cal B}_2 {\cal F}_0 {\cal A}_2 ) \quad .
2432: \end{eqnarray}
2433: \noindent
2434: Finally we present the expansion of $\chi_{\bar{\psi} \psi}$ which is
2435: generated in analogy to the isovector susceptibility by
2436: using ${\cal C}_0^2$ as a kernel in Eq.~(\ref{eq:an}). We find
2437: \begin{eqnarray}
2438: {\cal G}_0
2439: \hspace{-4mm} &=& \hspace{-4mm}
2440: \left\langle {\cal C}_0^2 \right\rangle, \\
2441: {\cal G}_1
2442: \hspace{-5mm} &=& \hspace{-5mm}
2443: 2 \left\langle {\cal C}_1 {\cal C}_0 \right\rangle
2444: + \left\langle {\cal C}_0^2 {\cal D}_1 \right\rangle, \\
2445: {\cal G}_2
2446: \hspace{-4mm} &=& \hspace{-4mm}
2447: 2 \left\langle {\cal C}_2 {\cal C}_0 \right\rangle
2448: + 2 \left\langle {\cal C}_1^2 \right\rangle
2449: + 4 \left\langle {\cal C}_1 {\cal C}_0 {\cal D}_1 \right\rangle
2450: + \left\langle {\cal C}_0^2 {\cal D}_2 \right\rangle
2451: + \left\langle {\cal C}_0^2 {\cal D}_1^2 \right\rangle, \\
2452: {\cal G}_3
2453: \hspace{-4mm} &=& \hspace{-4mm}
2454: 2 \left\langle {\cal C}_3 {\cal C}_0 \right\rangle
2455: + 6 \left\langle {\cal C}_2 {\cal C}_1 \right\rangle
2456: + 6 \left\langle {\cal C}_2 {\cal C}_0 {\cal D}_1 \right\rangle
2457: + 6 \left\langle {\cal C}_1^2 {\cal D}_1 \right\rangle
2458: + 6 \left\langle {\cal C}_1 {\cal C}_0 {\cal D}_2 \right\rangle
2459: + \left\langle {\cal C}_0^2 {\cal D}_3 \right\rangle
2460: \nonumber \\ && \hspace{-5mm}
2461: + 6 \left\langle {\cal C}_1 {\cal C}_0 {\cal D}_1^2 \right\rangle
2462: + 3 \left\langle {\cal C}_0^2 {\cal D}_2 {\cal D}_1 \right\rangle
2463: + \left\langle {\cal C}_0^2 {\cal D}_1^3 \right\rangle, \\
2464: {\cal G}_4
2465: \hspace{-4mm} &=& \hspace{-4mm}
2466: 2 \left\langle {\cal C}_4 {\cal C}_0 \right\rangle
2467: + 8 \left\langle {\cal C}_3 {\cal C}_1 \right\rangle
2468: + 6 \left\langle {\cal C}_2^2 \right\rangle
2469: + 8 \left\langle {\cal C}_3 {\cal C}_0 {\cal D}_1 \right\rangle
2470: + 24 \left\langle {\cal C}_2 {\cal C}_1 {\cal D}_1 \right\rangle
2471: \nonumber \\ && \hspace{-5mm}
2472: + 12 \left\langle {\cal C}_2 {\cal C}_0 {\cal D}_2 \right\rangle
2473: + 12 \left\langle {\cal C}_1^2 {\cal D}_2 \right\rangle
2474: + 8 \left\langle {\cal C}_1 {\cal C}_0 {\cal D}_3 \right\rangle
2475: + \left\langle {\cal C}_0^2 {\cal D}_4 \right\rangle
2476: + 12 \left\langle {\cal C}_2 {\cal C}_0 {\cal D}_1^2 \right\rangle
2477: \nonumber \\ && \hspace{-5mm}
2478: + 12 \left\langle {\cal C}_1^2 {\cal D}_1^2 \right\rangle
2479: + 24 \left\langle {\cal C}_1 {\cal C}_0 {\cal D}_2 {\cal D}_1 \right\rangle
2480: + 4 \left\langle {\cal C}_0^2 {\cal D}_3 {\cal D}_1 \right\rangle
2481: + 3 \left\langle {\cal C}_0^2 {\cal D}_2^2 \right\rangle
2482: + 8 \left\langle {\cal C}_1 {\cal C}_0 {\cal D}_1^3 \right\rangle
2483: \nonumber \\ && \hspace{-5mm}
2484: + 6 \left\langle {\cal C}_0^2 {\cal D}_2 {\cal D}_1^2 \right\rangle
2485: + \left\langle {\cal C}_0^2 {\cal D}_1^4 \right\rangle.
2486: \end{eqnarray}
2487: These expectation values satisfy,
2488: \begin{eqnarray}
2489: \frac{\partial {\cal G}_n}{\partial \mu}
2490: = {\cal G}_{n+1} - {\cal G}_n {\cal A}_1 \quad .
2491: \end{eqnarray}
2492: The expansion of the disconnected chiral susceptibility defined in
2493: Eq.~(\ref{eq:chisus}) is then given by,
2494: \begin{eqnarray}
2495: c_0^{\chi} &\equiv& \frac{\chi_{\bar{\psi} \psi}}{T^2} \biggr|_{\mu=0}
2496: = \frac{N_\tau}{N_{\sigma}^3} \left({\cal G}_0 - {\cal F}_0^2 \right),
2497: \\
2498: c_2^{\chi} &\equiv& \frac{1}{2}
2499: \frac{\partial^2 \chi_{\bar{\psi} \psi}/T^2}{\partial (\mu_{q}/T)^2}
2500: \biggr|_{\mu=0}
2501: %= \frac{1}{2}
2502: %\frac{\partial^2 \chi_{\bar{\psi} \psi}}{\partial \mu^2} \biggr|_{\mu=0}
2503: %\nonumber \\
2504: = \frac{1}{2N_{\sigma}^3 N_{\tau}} \left[
2505: {\cal G}_2 - {\cal G}_0 {\cal A}_2
2506: -2 \left( {\cal F}_2 - {\cal F}_0 {\cal A}_2 \right) {\cal F}_0
2507: \right] ,
2508: \\
2509: c_4^{\chi} &\equiv& \frac{1}{4!}
2510: \frac{\partial^4 \chi_{\bar{\psi} \psi}/T^2}{\partial (\mu_{q}/T)^4}
2511: \biggr|_{\mu=0}
2512: %= \frac{1}{4!\;N_{\tau}^2}
2513: %\frac{\partial^4 \chi_{\bar{\psi} \psi}}{\partial \mu^4} \biggr|_{\mu=0}
2514: \nonumber \\
2515: &=& \frac{1}{4!\; N_{\sigma}^3 N_{\tau}^3} \left[
2516: {\cal G}_4 -6 {\cal G}_2 {\cal A}_2 - {\cal G}_0 {\cal A}_4
2517: +6 {\cal G}_0 {\cal A}_2^2
2518: -6 \left( {\cal F}_2 - {\cal F}_0 {\cal A}_2 \right)^2
2519: \right.
2520: \nonumber \\
2521: && \hspace{15mm} \left.
2522: -2 \left({\cal F}_4 -6 {\cal F}_2 {\cal A}_2 - {\cal F}_0 {\cal A}_4
2523: +6 {\cal F}_0 {\cal A}_2^2 \right) {\cal F}_0
2524: \right] .
2525: \end{eqnarray}
2526:
2527:
2528: \clearpage
2529: \begin{thebibliography}{99}
2530: \baselineskip 10pt
2531: %
2532: \bibitem{review}
2533: see for instance: F. Karsch, Lect. Notes Phys. {\bf 583}
2534: (2002) 209.
2535: %
2536: \bibitem{Fodor1}
2537: Z. Fodor and S. Katz, JHEP {\bf 0203} (2002) 014.
2538: %
2539: \bibitem{us1}
2540: C.R. Allton, S. Ejiri, S.J. Hands, O. Kaczmarek, F. Karsch, E. Laermann,
2541: C. Schmidt and L. Scorzato, Phys. Rev. D {\bf 66} (2002) 074507.
2542: %
2543: \bibitem{Gavai1}
2544: R.V. Gavai and S. Gupta, Phys. Rev. D {\bf 64} (2001) 074506.
2545: %
2546: \bibitem{Lombardo1}
2547: M. D'Elia and M. P. Lombardo, Phys. Rev. D {\bf 67} (2003) 014505.
2548: %
2549: \bibitem{Philipsen1}
2550: P. de Forcrand and O. Philipsen, Nucl. Phys. {\bf B642} (2002) 290.
2551: %
2552: \bibitem{Stephanov}
2553: M. Stephanov, K. Rajagopal and E.V. Shuryak,
2554: Phys. Rev. Lett. 81 (1998) 4816.
2555: %
2556: \bibitem{Where}
2557: C.R. Allton, S. Ejiri, S.J. Hands, O. Kaczmarek, F. Karsch, E. Laermann,
2558: C. Schmidt, Prog. Theor. Phys. Suppl. 153 (2004) 118.
2559: %
2560: \bibitem{Fodor2}
2561: Z. Fodor and S. Katz, JHEP 0404 (2004) 050.
2562: %
2563: \bibitem{FKS}
2564: Z.~Fodor, S.D.~Katz and K.K.~Szabo,
2565: %``The QCD equation of state at nonzero densities: Lattice result,''
2566: Phys.\ Lett.\ B {\bf 568} (2003) 73.
2567: %[arXiv:hep-lat/0208078].
2568: %%CITATION = HEP-LAT 0208078;%%
2569: %
2570: \bibitem{us2}
2571: C.R.~Allton, S.~Ejiri, S.J.~Hands, O.~Kaczmarek, F.~Karsch, E.~Laermann and
2572: C.~Schmidt,
2573: %``The equation of state for two flavor QCD at non-zero chemical potential,''
2574: Phys.\ Rev.\ D {\bf 68} (2003) 014507.
2575: %[arXiv:hep-lat/0305007].
2576: %
2577: \bibitem{Fodoreos}
2578: F. Csikor, G. I. Egri, Z. Fodor, S. D. Katz, K. K. Szabo and A. I. Toth,
2579: JHEP {\bf 0405} (2004) 046.
2580: %
2581: \bibitem{Gavaieos}
2582: R.V. Gavai and S. Gupta, Phys. Rev. D {\bf 68} (2003) 034506;\\
2583: R.V. Gavai, S. Gupta and R. Roy, Prog.\ Theor.\ Phys.\ Suppl.\ {\bf
2584: 153} (2004) 270.
2585: %
2586: \bibitem{Lombardoeos}
2587: M. D'Elia and M.P. Lombardo, Phys. Rev. D {\bf 70} (2004) 074509.
2588: %
2589: \bibitem{KRT}
2590: F.~Karsch, K.~Redlich and A.~Tawfik,
2591: %``Hadron resonance mass spectrum and lattice QCD thermodynamics,''
2592: Eur.\ Phys.\ J.\ C {\bf 29} (2003) 549
2593: %[arXiv:hep-ph/0303108]
2594: %%CITATION = HEP-PH 0303108;%%
2595: and
2596: %``Thermodynamics at non-zero baryon number density:
2597: %A comparison of lattice
2598: %and hadron resonance gas model calculations,''
2599: Phys.\ Lett.\ B {\bf 571} (2003) 67.
2600: %[arXiv:hep-ph/0306208].
2601: %%CITATION = HEP-PH 0306208;%%
2602: %
2603: \bibitem{Gottlieb}
2604: S. Gottlieb, W. Liu, R.L. Renken, R.L. Sugar and D. Toussaint,
2605: Phys.\ Rev.\ Lett.\ {\bf 59} (1987) 2247 and
2606: Phys. Rev. D {\bf 38} (1988) 2888.
2607: %%CITATION = PHRVA,D38,2888;%%
2608: %
2609: \bibitem{suscept}
2610: R.V. Gavai, J. Potvin and S. Sanielevici, Phys.\ Rev.\ D {\bf 40} (1989) R2743;\\
2611: C. Bernard et al., Phys.\ Rev.\ D {\bf 54} (1996) 4585;\\
2612: S. Gottlieb et al., Phys.\ Rev.\ D {\bf 55} (1997) 6852.
2613: %
2614: \bibitem{Gupta02}
2615: R.V. Gavai, S. Gupta and P. Majumdar, Phys. Rev. D {\bf 65} (2002) 054506.
2616: %
2617: \bibitem{suscept2}
2618: C. Bernard et al.,Nucl.\ Phys.\ B (Proc. Suppl.) 119 (2003) 523,
2619: {\it QCD Thermodynamics with Three Flavors of Improved Staggered Quarks},
2620: hep-lat/0405029.
2621: %
2622: \bibitem{Gavai04}
2623: R.V. Gavai and S. Gupta, {\it The critical end point of QCD}, hep-lat/0412035.
2624: %
2625: \bibitem{Kunihiro}
2626: T. Kunihiro, Phys. Lett. B {\bf 271} (1991) 395.
2627: %%CITATION = HEP-LAT 0204010;
2628: %
2629: \bibitem{Blaizot}
2630: J.-P. Blaizot, E. Iancu and A. Rebhan, Phys. Lett. B {\bf 523} (2001) 143.
2631: %
2632: \bibitem{Vuorinen}
2633: A. Vuorinen, Phys.\ Rev.\ D {\bf 68} (2003) 054017.
2634: %
2635: \bibitem{Hagedorn}
2636: R. Hagedorn, Nuovo Cimento {\bf 35} (1965) 395.
2637: %
2638: \bibitem{pbm}
2639: P. Braun-Munzinger, K. Redlich and J. Stachel, in:
2640: Quark Gluon Plasma 3
2641: (Edts. R.C. Hwa and Xin-Nian Wang), World-Scientific 2004.
2642: %
2643: \bibitem{Kajantie}
2644: K. Kajantie, M. Laine, K. Rummukainen and Y. Schr\"oder,
2645: Phys.\ Rev.\ D {\bf 67} (2003) 105008.
2646: %
2647: \bibitem{Nc}
2648: T.D. Cohen, {\it Large $N_c$ QCD at non-zero chemical potential},
2649: hep-ph/0410156; \\
2650: D. Toublan, {\it A large $N_c$ perspective on the QCD phase diagram},
2651: hep-th/0501069.
2652: %
2653: \bibitem{Kogut}
2654: J.B. Kogut and D.K. Sinclair, Phys. Rev. D {\bf 66} (2002) 034505 and
2655: Phys. Rev. D {\bf 70} (2004) 094501.
2656: %
2657: \bibitem{KLP}
2658: F. Karsch, E. Laermann and A. Peikert, Phys. Lett. B {\bf 478} (2000) 447.
2659: %%CITATION = HEP-LAT 0002003;%%
2660: %
2661: \bibitem{FKEL}
2662: F.~Karsch and E.~Laermann, Phys.\ Rev.\ D {\bf 50} (1994) 6954.
2663: %
2664: \bibitem{Gaunt}
2665: see for instance: S. Gaunt and A.J. Gutmann in {\it Phase Transitions
2666: and Critical Phenomena 3} (Edts. C. Domb and M.S. Green), Academic Press
2667: 1974.
2668: %
2669: \bibitem{Shinji}
2670: S.~Ejiri,
2671: %``Remarks on the multi-parameter reweighting method for the study of lattice
2672: %QCD at non-zero temperature and density,''
2673: Phys.\ Rev.\ D {\bf 69} (2004) 094506.
2674: %[arXiv:hep-lat/0401012].
2675: %%CITATION = HEP-LAT 0401012;%%
2676: \end{thebibliography}
2677:
2678: \end{document}
2679:
2680: