arxiv: v1 [hep-lat] 13 Jul 2007

Similar documents
Analytic continuation from an imaginary chemical potential

The phase diagram of QCD from imaginary chemical potentials

arxiv:hep-lat/ v2 7 Sep 2004

arxiv:hep-ph/ v1 23 Jan 2003 QCD PHASE DIAGRAM AT SMALL DENSITIES FROM SIMULATIONS WITH IMAGINARY µ

The QCD phase diagram at low baryon density from lattice simulations

arxiv: v1 [hep-lat] 25 Feb 2012

The QCD phase diagram at real and imaginary chemical potential

Constraints on the QCD phase diagram from imaginary chemical potential

arxiv:hep-lat/ v1 25 Jun 2005

Massimo D Elia Dipartimento di Fisica and INFN Genova, Via Dodecaneso 33, I Genova, ITALY

Recent Progress in Lattice QCD at Finite Density

Taylor expansion in chemical potential for 2 flavour QCD with a = 1/4T. Rajiv Gavai and Sourendu Gupta TIFR, Mumbai. April 1, 2004

arxiv:hep-lat/ v1 7 Jan 2004

The QCD pseudocritical line from imaginary chemical potentials

Lattice QCD. QCD 2002, I. I. T. Kanpur, November 19, 2002 R. V. Gavai Top 1

from Taylor expansion at non-zero density From Lattices to Stars INT, University of Washington, Seattle, 28. April 2004

The High Density Region of QCD from an Effective Model

Constraints for the QCD phase diagram from imaginary chemical potential

Imaginary chemical potential and finite fermion density on the lattice

Critical end point of Nf=3 QCD at finite temperature and density

N f = 1. crossover. 2nd order Z(2) m, m

Exploring finite density QCD phase transition with canonical approach Power of multiple precision computation

Bulk Thermodynamics: What do we (want to) know?

Confining strings in representations with common n-ality

QCD Phase Diagram. M. Stephanov. U. of Illinois at Chicago. QCD Phase Diagram p. 1/13

QCD at Finite Density and the Sign Problem

Thermodynamics of (2+1)-flavor QCD from the lattice

A model for QCD at High Density and large Quark Mass

Chiral symmetry breaking, instantons, and monopoles

Finite Chemical Potential in N t = 6 QCD

arxiv:hep-lat/ v1 5 Oct 2006

Hopping parameter expansion to all orders using the Complex Langevin equation

arxiv: v1 [hep-lat] 18 Nov 2013

The chiral phase transition for two-flavour QCD at imaginary and zero chemical potential

arxiv: v2 [hep-lat] 13 Dec 2010

Properties of canonical fermion determinants with a fixed quark number

arxiv: v1 [hep-lat] 26 Dec 2009

PoS(LATTICE 2015)261. Scalar and vector form factors of D πlν and D Klν decays with N f = Twisted fermions

I would like to thank Colleagues in the World for

arxiv: v1 [hep-lat] 18 Dec 2013

EQUATION OF STATE AND FLUCTUATIONS FROM THE LATTICE Claudia Ratti University of Houston (USA)

arxiv: v1 [hep-lat] 19 Feb 2012

The critical point of QCD: what measurements can one make?

arxiv:hep-lat/ v2 13 Oct 1998

Edinburgh Research Explorer

The sign problem in Lattice QCD

The Equation of State for QCD with 2+1 Flavors of Quarks

Probing the QCD phase diagram with higher moments

Two-colour Lattice QCD with dynamical fermions at non-zero density versus Matrix Models

A Noisy Monte Carlo Algorithm

Baryon correlators containing different diquarks from lattice simulations

arxiv: v1 [hep-lat] 5 Nov 2018

Mass Components of Mesons from Lattice QCD

arxiv: v1 [hep-lat] 7 Oct 2007

Is the up-quark massless? Hartmut Wittig DESY

Nuclear Matter between Heaven and Earth: The QCD Phase Diagram

arxiv: v1 [hep-lat] 31 Oct 2014

Lattice QCD with Eight Degenerate Quark Flavors

Lattice QCD at nonzero baryon density

Gauge invariance of the Abelian dual Meissner effect in pure SU(2) QCD

Phase of the Fermion Determinant and the Phase Diagram of QCD

Catalytic effects of monopole in QCD

Maarten Golterman Department of Physics and Astronomy, San Francisco State University, San Francisco, CA 94132, USA

Phase diagram at finite temperature and quark density in the strong coupling limit of lattice QCD for color SU(3)

Can we locate the QCD critical endpoint with the Taylor expansion?

Canonical partition functions in lattice QCD

QCD Critical Point : Inching Towards Continuum

A Simple Idea for Lattice QCD at Finite Density

Exploring the QCD phase diagram with conserved charge fluctuations

arxiv: v1 [hep-lat] 5 Nov 2007

Orientifold planar equivalence.

Poisson statistics in the high temperature QCD Dirac spectrum

arxiv: v1 [hep-ph] 24 Oct 2012

Lattice QCD study for relation between quark-confinement and chiral symmetry breaking

Baryon number distribution in Lattice QCD

arxiv: v1 [hep-lat] 19 Jan 2018

Thermodynamics using p4-improved staggered fermion action on QCDOC

arxiv: v1 [hep-ph] 2 Nov 2009

Polyakov loop extended NJL model with imaginary chemical potential

The Polyakov Loop and the Eigenvalues of the Dirac Operator

arxiv:hep-lat/ v1 24 Jun 1998

Surprises in the Columbia plot

Multiple Critical Points in the QCD Phase Diagram

and B. Taglienti (b) (a): Dipartimento di Fisica and Infn, Universita di Cagliari (c): Dipartimento di Fisica and Infn, Universita di Roma La Sapienza

PoS(LATTICE 2013)500. Charmonium, D s and D s from overlap fermion on domain wall fermion configurations

Quark number density and susceptibility calculation under one-loop correction in the mean-field potential

The interplay of flavour- and Polyakov-loop- degrees of freedom

Continuous Insights in the Wilson Dirac Operator

Gapless Dirac Spectrum at High Temperature

Complex Langevin dynamics for nonabelian gauge theories

Phase diagram and EoS from a Taylor expansion of the pressure

The QCD phase diagram from the lattice

Magnetic properties of the QCD medium

First results from dynamical chirally improved fermions

MILC results and the convergence of the chiral expansion

Large-N c universality of phases in QCD and QCD-like theories

Monopole Condensation and Confinement in SU(2) QCD (1) Abstract

Quark Mass and Flavour Dependence of the QCD Phase Transition. F. Karsch, E. Laermann and A. Peikert ABSTRACT

Maria Paola Lombardo GGI Firenze March 2014

Transcription:

Imaginary chemical potentials and the phase of the fermionic determinant Simone Conradi and Massimo D Elia Dipartimento di Fisica dell Università di Genova and INFN, Sezione di Genova, Via Dodecaneso 33, I-646 Genova, Italy (Dated: February 9, 203) A numerical technique is proposed for an efficient numerical determination of the average phase factor of the fermionic determinant continued to imaginary values of the chemical potential. The method is tested in QCD with eight flavors of dynamical staggered fermions. A direct check of the validity of analytic continuation is made on small lattices and a study of the scaling with the lattice volume is performed. PACS numbers:.5.ha, 2.38 Gc, 2.38.Aw GEF-TH 8/07 arxiv:0707.987v [hep-lat] 3 Jul 2007 I. INTRODUCTION Lattice QCD simulations in presence of a finite density of baryonic matter are hindered by the well known sign problem. Consider for instance the QCD partition function Z(µ, µ) DUe SG[U] (detm[u, µ]) 2 = DUe SG[U] detm[u, µ] 2 e i2θ, () describing two flavors of quarks (or eight flavors in the case of staggered flavors) which are given an equal chemical potential µ: the determinant of the fermionic matrix M is in general complex (θ 0) for µ 0 and Monte Carlo simulations are not feasible. Various possibilities have been explored to circumvent the problem, like reweighting techniques [, 2, 3], the use of an imaginary chemical potential either for analytic continuation [4, 5, 6, 7, 8, 9, 0] or for reconstructing the canonical partition function [, 2, 3], Taylor expansion techniques [4, 5] and non-relativistic expansions [6, 7, 8]. The same is not true in the case of a finite isospin density, i.e. when quarks are given opposite chemical potentials. Indeed, due to the property detm[u, µ] = detm[u, µ], the partition function Z(µ, µ) = DUe SG[U] detm[u, µ] 2 (2) has a positive measure. That is also known as phase quenched QCD. The average value of the phase factor of the fermionic determinant, e i2θ (µ, µ), where the index indicates the partition function the expectation value refers to, gives a direct measurement of the severeness of the sign problem. e i2θ 0 will signal the stage at which the complex nature of the determinant will imply a significant difference between finite baryonic density and finite isospin density, as well a poor reliability of reweighting techniques (see Ref. [9] and references therein). It clearly follows from Eqs. () and (2) that the average phase factor is the expectation value of the ratio of two determinants; it can also be expressed as the ratio of two partition functions: e i2θ µ detm(µ) = detm( µ) (µ, µ) Z(µ, µ) Z(µ, µ). (3) Its direct numerical computation reveals a difficult numerical task as the lattice volume V increases, since it involves the numerical evaluation of fermionic determinants. It has been proposed recently [20, 2] to study the analytic continuation of the average phase factor to imaginary values of the chemical potential e i2θ iµ = detm(iµ) Z(iµ, iµ) = detm( iµ) (iµ, iµ) Z(iµ, iµ) DUe S G[U] det M[U, iµ] detm[u, iµ] DUe S G[U] detm[u, iµ] detm[u, iµ] (4) where Z(iµ, iµ) and Z(iµ, iµ) are the analytic continuation of the partition functions at finite baryonic and isospin chemical potentials respectively, which are both suitable for numerical simulations since det M[U, iµ] is always real. Numerical difficulties however are present also in this case: the observable to be averaged is still expressed in terms of fermionic determinants. Moreover in principle sampling problems deriving from a bad overlap between the two statistical distributions described by Z(iµ, iµ) and Z(iµ, iµ) may arise. In Ref. [2] the fermionic determinant has been estimated on the basis of the lowest lying eigenvalues of the fermionic matrix. In the present paper we propose a new technique which, making use of numerical strategies developed in different contexts, permits an exact evaluation of the average phase factor with a reasonable scaling of the required CPU time as the lattice volume is increased. In doing this we will fully exploit the possibility of performing numerical simulations of the partition function Z(iµ, iµ 2 ) for generic values of µ and µ 2. In Section II we illustrate two different possible methods, which are then numerically tested and compared in Section III for the theory with 8 staggered flavors.

2 II. THE METHOD The evaluation of the average phase factor, expressed like in Eq. (3) or Eq. (4) as the ratio of two different partition functions, resembles similar problems which are encountered in quite different contexts, like the evaluation of disorder parameters in statistical models and in lattice gauge theories. Explicit examples are given by monopole disorder parameters or by the t Hooft loop, which enters in various studies about color confinement. The major problem in those cases is the small overlap between the statistical distributions corresponding to two different partition functions, resulting in a poor sampling efficiency. Powerful techniques have been developed in both cases, consisting in either determining derivatives of the disorder parameters, from which the ratio of partition functions can then be reconstructed after integration [22, 23, 24], or in making use of various reweighting techniques, like that of rewriting the original ratio in terms of intermediate ratios which are more easily evaluable [25, 26, 27]. In the present case the major difficulty derives from a direct computation of the observables, which is expressed in terms of fermionic determinants, but sampling problems may in principle worsen the situation also in this case, especially in the large volume limit. In the following we will describe the application of both kind of techniques described above to the present case, and try to understand by numerical simulations which of them is best suited in this context. We describe at first how to reconstruct e i2θ in terms of derivatives. Consider the modified ratio R µ (ν) = Z(iµ, iν) Z(iµ, iµ) (5) where Z(iµ, iν) DUe SG[U] detm[u, iµ] detm[u, iν]. (6) It is clear that R µ ( µ) =, while R µ (µ) is the original ratio. It can be easily verified that ρ(ν) d dν lnr µ(ν) = d lnz(iµ, iν) ( dν ) = i Tr M d (iν) d(iν) M(iν) (iµ,iν). (7) Last quantity is nothing but i times the average number of quarks coupled to the chemical potential iν: it is purely imaginary for symmetry reasons, hence ρ(ν) is real, and can be computed using a noisy unbiased estimator. The average phase factor can then be obtained by integration ( µ ) e i2θ iµ = exp ρ(ν)dν (8) µ and no quark determinant must be explicitly computed. In practice, the derivative ρ(ν) will be computed for a discrete set of values of ν and then integrated numerically. The precision attained for e i2θ iµ will depend both on the statistical errors of the single determinations and on the systematic uncertainty linked to numerical integration; the last can be estimated for instance by varying the chosen interpolation procedure. In principle it is also possible to determine further derivatives of ρ in order to improve the integration accuracy. As a different method we consider rewriting e i2θ iµ as the product of N intermediate ratios: e i2θ iµ = Z(iµ, iµ) Z(iµ, iµ) = Z N Z N Z N Z N 2... Z Z 0 N r k (9) k= where Z N Z(iµ, iµ), Z 0 Z(iµ, iµ) while Z k DUe SG[U] detm[u, iµ]detm[u, i( µ + kδν)](0) with δν = 2µ/N. The idea is to compute each single ratio r k by a different Monte Carlo simulation. Apart from the increased overlap among each couple of partition functions, an improvement comes also from the simpler form in which the observable appearing in each ratio r k can be rewritten, for large enough N. Indeed we have: r k = detm(i(ν + δν))/ detm(iν) (iµ,iν) = exp (TrlnA(ν, δν)) (iµ,iν) () where ν = µ + (k )δν and A[U, ν, δν] M[U, iν] M[U, i(ν + δν)]. (2) If δν is small, the matrix A[U, ν, δν] is very close to the identity matrix Id for each configuration U. We can therefore expand the logarithm in Eq. () thus rewriting the following approximate expression for r k : r k exp ( Tr(A Id) 2 Tr(A Id)2 +... ) (3) Each trace in the exponential can be evaluted by a noisy estimator as follows: Tr(A[U] Id) n K K η (j) (A[U] Id) n η (j) (4) j= η (j) where η (j) is a random vector satisfying η (j) i i 2 η = δ i,i 2. The computation of each noise estimate in Eq. (4) can be made faster if, when applying the matrix A[U] = M[U, iν] M[U, i(ν + δν)] to the vector η (j) (or to (A[U] Id) n η (j) at higher orders), η (j) itself is taken as a starting tentative solution for the inverter giving M[U, iν] (M[U, i(ν + δν)]η (j) ): the guess is better and better as δν 0. The second method is not conceptually different from the first one: finite free energy differences are computed

3 400 0 µ = 0.025 µ = 0.05 µ = 0.075 µ = 0.0 µ = 0.20 300 200 00 µ = 0.025 µ = 0.05 ρ(ν) 0 ρ(ν) 0-00 -200-0 -300 - -0.8-0.6-0.4-0.2 0 0.2 0.4 0.6 0.8 ν/µ -400 - -0.8-0.6-0.4-0.2 0 0.2 0.4 0.6 0.8 ν/µ FIG. : ρ(ν) for various values of µ at β = 4.8 and L s = 4. FIG. 3: ρ(ν) for various values of µ at β = 4.8 and L s = 6. in this case instead of derivatives. However the numerical procedure is different and it is not clear apriori which approach is more convenient. In the second case no numerical integration must be performed, however one has the drawback that the exponential of a noisy unbiased estimator is biased, hence a large number K of vectors must be used and the final result must be checked to be independent of K. Moreover the systematic error involved in the truncation of the logarithm expansion, Eq. (3), must be properly estimated and kept under control. III. NUMERICAL RESULTS We have tested our methods for the theory with 8 staggered flavors of mass am = 0.. We will present results obtained on L 3 s L t lattices with L t = 4 and L s = 4, 8, 6. At zero chemical potential this theory presents a strong first order deconfinement/chiral transition, the critical coupling being β c 4.7 for L t = 4. We have performed simulations both in the deconfined region (β = 4.8) and in the low temperature confined region (β = 4.6). On the smallest lattice (L s = 4) we will compare our results directly with those obtained at real isospin chemical potential by a direct evaluation of the determinant phase. Numerical simulations have been performed mostly on the APEmille facility in Pisa. The INFN apenext facility in Rome has been used for the results on the largest lattice. The standard exact HMC algorithm [28] has been used with trajectories of length. A collection of the results obtained for imaginary chemical potentials is reported in Table I..06 0.8 0.6 0.4 0.2 µ = 0.025 µ = 0.05.04.02 ρ(ν) 0-0.2-0.4-0.6-0.8 0.98 0 5 0 5 20 25 30 K - - -0.8-0.6-0.4-0.2 0 0.2 0.4 0.6 0.8 ν/µ FIG. 2: ρ(ν) for various values of µ at β = 4.6 and L s = 4. FIG. 4: r + k (blank triangles) and the inverse of r k (blank circles) defined in Eq. (5), together with their geometric mean, i.e. r k (filled circles). Data are showed for L s = 4, µ = 0.05, ν = 0.045 and δν = 0.005.

4 A. Systematic errors and comparison of the methods In Figs., 2 and 3 we report various determinations of ρ(ν) (minus the imaginary part of the baryon number, Eq. (7)) obtained on discrete sets of points. It is apparent that ρ(ν) is a very smooth function of ν in all explored cases and independently of the lattice size and of the explored phase (confined or deconfined). In most cases it can even be approximated by a linear function; therefore numerical integration turns out to be an easy task. We have adopted a simple linear interpolation between consecutive points to obtain the results given in Table I, the reported errors derive from standard error propagation of the statistical errors of the single data points. We have verified, by changing the order of the interpolating polynomial, that the systematic error related to the interpolation-integration procedure is negligible with respect to the statistical one. Concerning the second method described in Section II, we have adopted a standard trick [25] in order to reduce systematic effects. Each partial ratio r k in Eq. (9) has been rewritten as detm(i(ν + δν)) r k = = r+ k detm(iν) iµ,iν r k detm(i(ν + δν))/ detm(i(ν + δν 2 )) iµ,i(ν+ δν 2 ) detm(i(ν))/ detm(i(ν + δν 2 )). (5) iµ,i(ν+ δν 2 ) r k can again be evaluated in a single simulation and a jackknife analysis has to be applied to obtain a correct error estimate. Two major benefits derive in this case. First, the reduced value δν/2 greatly improves the convergence of the logarithm expansion in Eq. (3). Second, the bias introduced by the finite number of noisy estimators, see Eqs. (3) and (4), gets largely cancelled in the ratio. That is apparent from Fig. 4, where we plot r + k and the inverse of r k defined in Eq. (5), and their geometric mean (i.e. r k ), as a function of the number K of noise vectors, in one particular sample case. It is clear that, while the single factors have a relatively slow convergence, their product is stable from K = 5 on. We have however always used K = 30 in our determinations. Regarding the logarithm expansion, Eq. (3), we have always adopted a third order approximation: in all cases the discrepancy with the result obtained at the second order is at least one order of magnitude smaller than the statistical uncertainty. The fact that the systematic error related to this expansion is well under control can be also appreciated from Table I, second and third row, showing Clearly one expects a non-smooth behaviour if Z(iµ, iµ) and Z 0 Z(iµ, iµ) belong to two different phases, so that some transition is met when ν goes from ν to ν, however in these cases analytic continuation itself is not applicable. L s β Im(µ) method e i2θ iµ HMC trajs 4 4.8 0.025 DER(0).00322(42) 700k 4 4.8 0.025 RAT(5).0030(8) 50k 4 4.8 0.025 RAT(0).0028() 300k 4 4.8 0.025 direct.0033() 40k 4 4.8 0.05 DER(20).008() 800k 4 4.8 0.05 RAT(0).022(6) 500k 4 4.8 0.075 DER(5).0266(7) 350k 4 4.8 0.0 DER(20).0454(6) 700k 4 4.8 0.20 DER(6).283(8) 700k 8 4.8 0.025 DER(0).064(9) 50k 8 4.8 0.025 RAT(5).0200(50) 50k 6 4.8 0.025 DER(0).0732(85) 60k 6 4.8 0.025 RAT(5).053(33) 40k 6 4.8 0.05 DER(0).368(30) 40k 4 4.6 0.025 DER(5).006(0) 200k 4 4.6 0.05 DER(0).0270(5) 350k TABLE I: Collection of determinations of the average phase factor continued to imaginary values of µ for various parameter sets and computation methods. In the fourth column the method used to obtain the determination is described: DER(N) stands for the integration of the first derivative ρ determined on a discrete set of (N+) points, Eq. (8); RAT(N) stands for the evaluation of N intermediate ratios r k, Eq. (9). Finally on the smallest lattices also a direct determination of the expectation value in Eq. (4) is reported for comparison. that the determination of e i2θ iµ is stable against the variation of the number of intermediate ratios. Let us now come to the comparison between the two methods. While they always give perfectly compatible results, thus confirming the absence of appreciable systematics, it is clear from Table I that, with a comparable numerical effort (in the last column we give the total number of Monte-Carlo trajectories used for each determination), the method described by Eq. (8) (integration of the derivative) furnishes more accurate determinations. We have therefore chosen this method in order to perform more extensive studies of e i2θ iµ. B. Test of analytic continuation The average phase factor computed at finite isospin chemical potential, at variance with that computed in the quenched theory, is expected [20, 2] to be an analytic function of µ 2 around µ 2 = 0 2. We can test directly analytic continuation by comparing our results with direct determinations of e i2θ iµ performed at real chemical potentials: this is done only for the smallest lattice (L s = 4), where the second determination is easily affordable. 2 It is even in µ for symmetry reasons.

5.2 quadratic fit, β = 4.8 quartic fit, β = 4.8 quadratic fit, β = 4.6 as shown in the figure. We obtain, at β = 4.8, A = 4.48(8), B = 5.7 ± 2.5 and χ 2 /d.o.f..3. Analyticity around µ 2 = 0 is therefore well verified. We stress that at β = 4.8 our largest value of the imaginary chemical potential is still below the first Roberge- Weiss phase transition at Im(µ) = π/(3l t ), hence within the expected range of validity of analytic continuation for µ 2 < 0 at high temperatures. C. Large volume scaling 0.8-0.04-0.02 0 0.02 0.04 µ 2 FIG. 5: e i2θ computed for different values of µ 2 at β = 4.8 and β = 4.6 on a 4 4 lattice. Best fit quadratic and quartic functions in µ 2 are displayed, showing good validity of analytic continuation. We plot in Fig. 5 results obtained at β = 4.6 and β = 4.8. The whole set of results obtained at real chemical potentials (µ 2 > 0) and imaginary chemical potentials (µ 2 < 0) can be described by a simple quadratic behaviour e i2θ = + Aµ 2 (6) in a range µ 2 0.0, with A = 4.4(9) and χ 2 /d.o.f..5 for β = 4.8 and A = 0.2(3) and χ 2 /d.o.f..8 for β = 4.6. If the range of values is extended a quartic term is necessary e i2θ = + Aµ 2 + Bµ 4 (7) We have performed numerical simulations at different values of L s in order to test both the behaviour of e i2θ and the efficiency of our method as the lattice volume is increased. In Fig. 6 we report determinations performed at fixed values of iµ and variable L s at β = 4.8. A behaviour e i2θ = + CL γ s (8) well describes the data with γ 2.5 for both values of iµ. Concerning the numerical efficiency, we notice that to obtain comparable uncertainties (of the order of 0 %) for e i2θ, on the largest lattice (6 3 4) we needed a CPU time which is less than one order of magnitude bigger than what needed on the smallest lattice (4 4 ). Considering that the two lattice volumes differ by a factor 64, we deduce that, at least for the quark mass considered in the present study, our method requires a numerical effort which scales in an affordable way with the lattice size..4 IV. CONCLUSIONS.3.2. β = 4.8, µ = 0.025 β = 4.8, µ = 0.050 0 5 0 5 20 L s FIG. 6: e i2θ iµ as a function of the spatial lattice size L s for two values of iµ. A best fit according to Eq. (8) is reported in both cases. We have presented two different techniques, described respectively by Eq. (8) and Eq. (9), for an efficient numerical determination of the average phase factor of the fermionic determinant continued to imaginary values of the chemical potential. We have applied both methods to QCD with 8 dynamical staggered flavors, verifying the absence of uncontrolled systematic effects and performing a comparison of the efficiencies, with the conclusion that the method based on the integration of the imaginary part of the baryon density, Eq. (8), is numerically more convenient. A fair good scaling of the efficiency is observed as the lattice volume is increased. We have also directly tested, on small lattices, the analiticity of the average phase factor around µ 2 = 0. The method proposed and tested in the present paper will be used in the future to perform more extensive studies, with more physical quark masses and number of flavors, of the average phase factor continued to imaginary chemical potentials.

6 [] I. M. Barbour, S. E. Morrison, E. G. Klepfish, J. B. Kogut and M. P. Lombardo, Nucl. Phys. Proc. Suppl. 60A, 220 (998). [2] Z. Fodor, S. D. Katz, Phys. Lett. B 534 (2002) 87; JHEP 0203, 04 (2002). [3] Z. Fodor, S. D. Katz and C. Schmidt, JHEP 0703, 2 (2007). [4] Ph. de Forcrand and O. Philipsen, Nucl. Phys. B 642, 290 (2002); Nucl. Phys. B 673, 70 (2003). [5] M. D Elia and M.P. Lombardo, Phys. Rev. D 67, 04505 (2003); Phys. Rev. D 70, 074509 (2004). [6] V. Azcoiti, G. Di Carlo, A. Galante and V. Laliena, Nucl. Phys. B 723, 77 (2005). [7] H. S. Chen and X. Q. Luo, Phys. Rev. D 72, 034504 (2005). [8] P. Giudice and A. Papa, Phys. Rev. D 69, 094509 (2004) [9] P. Cea, L. Cosmai, M. D Elia and A. Papa, JHEP 0702, 066 (2007). [0] M. D Elia, F. Di Renzo, M.P. Lombardo, arxiv:0705.384 [hep-lat]. [] A. Roberge and N. Weiss, Nucl. Phys. B 275, 734 (986). [2] S. Kratochvila and P. de Forcrand, PoS LAT2005, 67 (2006). [3] A. Alexandru, M. Faber, I. Horvath and K. F. Liu, Phys. Rev. D 72, 453 (2005). [4] C. R. Allton et al., Phys. Rev. D 66, 074507 (2002); Phys. Rev. D 7, 054508 (2005). [5] R. V. Gavai and S. Gupta, Phys. Rev. D 68, 034506 (2003); Phys. Rev. D 73, 04004 (2006). [6] T. C. Blum, J. E. Hetrick and D. Toussaint, Phys. Rev. Lett. 76, 09 (996). [7] J. Engels, O. Kaczmarek, F. Karsch and E. Laermann, Nucl. Phys. B 558, 307 (999). [8] R. De Pietri, A. Feo, E. Seiler and I. O. Stamatescu, arxiv:0705.3420 [hep-lat]. [9] K. Splittorff, PoS LAT2006, 023 (2006) [arxiv:hep-lat/060072]. [20] K. Splittorff and J. J. M. Verbaarschot, Phys. Rev. D 75, 6003 (2007). [2] K. Splittorff and B. Svetitsky, Phys. Rev. D 75, 4504 (2007). [22] L. Del Debbio, A. Di Giacomo and G. Paffuti, Phys. Lett. B 349, 53 (995). [23] P. Cea, L. Cosmai and A. D. Polosa, Phys. Lett. B 392, 77 (997) [24] L. Del Debbio, A. Di Giacomo, B. Lucini, Nucl. Phys. B594, 287 (200); Phys. Lett. B500, 326 (200). [25] Ph. de Forcrand, M. D Elia, M. Pepe, Phys. Rev. Lett. 86, 438 (200). [26] Ph. de Forcrand and M. Vettorazzo, Nucl. Phys. B686, 85 (2004). [27] M. D Elia and L. Tagliacozzo, Phys. Rev. D 74, 450 (2006). [28] S.A. Gottlieb, W. Liu, D. Toussaint, R.L. Renken and R.L. Sugar, Phys. Rev. D 35, 253 (987).