microscopic origin of ua qcd - arxiv · 2018-10-02 · arxiv:1502.06190v1 [hep-lat] 22 feb 2015...

15
arXiv:1502.06190v1 [hep-lat] 22 Feb 2015 Microscopic Origin of U A (1) Symmetry Violation in the High Temperature Phase of QCD Viktor Dick, 1 Frithjof Karsch, 1, 2 Edwin Laermann, 1 Swagato Mukherjee, 2 and Sayantan Sharma 2 1 Fakult¨ at f¨ ur Physik, Universit¨ at Bielefeld, D-33615 Bielefeld, Germany 2 Brookhaven National Laboratory, Upton, NY 11973, USA We investigate the low-lying eigenmodes of the Dirac matrix with the aim to gain more insight into the temperature dependence of the anomalous UA(1) symmetry. We use the overlap operator to probe dynamical QCD configurations generated with (2+1)-flavors of highly improved staggered quarks. We find no evidence of a gap opening up in the infrared region of the eigenvalue spectrum even at 1.5 Tc, Tc being the chiral crossover temperature. Instead, we observe an accumulation of near-zero eigenmodes. We argue that these near-zero eigenmodes are primarily responsible for the anomalous breaking of the axial symmetry still being effective. At 1.5 Tc, these near-zero eigenmodes remain localized and their distribution is consistent with the dilute instanton gas picture. At this temperature, the average size of the instantons is 0.223(8) fm and their density is 0.147(7) fm -4 . PACS numbers: 12.38.Gc, 11.15.Ha, 11.30.Rd, 11.15.Kc I. INTRODUCTION Owing to the near-degeneracy and smallness of the up and down quark masses, the Quantum Chromo- dynamics (QCD) Lagrangian possesses an approximate U L (2) × U R (2) SU L (2) × SU R (2) × U V (1) × U A (1) symmetry. The fact that we do not see parity doublet hadrons in our world implies that the SU L (2) × SU R (2) chiral symmetry is spontaneously broken down to the SU V (2) isospin symmetry of the vacuum. It is well known from first principle lattice QCD studies that above the chiral crossover [1–4] temperature of T c = 154(9) MeV [5] the chiral symmetry of QCD gets restored. On the other hand, the axial U A (1) symmetry of the QCD Lagrangian is always broken due to the presence of quantum fluctuations. This gives rise to the well-known anomalous non-conservation of the axial current [6, 7]. The explicit violation of the global U A (1) symmetry is due to the presence of topologically nontrivial gauge field configurations [8]. Although U A (1) is not an exact sym- metry of QCD, the magnitude of its breaking near T c is expected to influence the nature of the chiral phase tran- sition in the limit of two vanishingly small light quark masses. Perturbative renormalization group studies of model quantum field theories with the same global sym- metries as QCD suggest that if U A (1) is not effectively restored at T c , the chiral phase transition is of second order, belonging to the 3-dimensional O(4) universality class [9–12]. If the axial symmetry gets effectively re- stored for T T c , the chiral phase transition can be ei- ther of first order [9, 10] or of second order with the sym- metry breaking pattern U L (2) × U R (2) U V (2) [11, 12]. In order to resolve the nature of the phase transition of QCD with two light quark flavors it is thus important to understand the significance of the anomalous U A (1) in the high temperature phase. At low temperatures, U A (1) is also broken explicitly by the presence of a non vanishing vacuum chiral con- densate. In the chirally symmetric phase, the vacuum condensate vanishes and the mechanism of global U A (1) breaking can be studied directly. The microscopic mech- anism for U A (1) breaking in the chirally symmetric phase of QCD presents an intriguing puzzle. The chiral conden- sate, which is the order parameter related to the restora- tion of chiral symmetry in QCD with massless quarks, can be expressed in terms of the eigenvalues λ of the Dirac operator as ¯ ψψV →∞ 0 dλ 2(λ,m) λ 2 + m 2 , (1) where ρ(λ,m) is the eigenvalue density. On the other hand, U A (1) is not a global symmetry, so one cannot define a corresponding order parameter. For two light quark flavors, an approximate restoration of U A (1) would result in the degeneracy of the correlation functions of the pion and the scalar iso-triplet delta meson [13]. Specifi- cally, the difference of the integrated correlation functions of these mesons in terms of the eigenvalues of the Dirac operator is χ π χ δ = d 4 x + (x)(0)〉−〈δ + (x)δ (0) V →∞ 0 dλ 4m 2 ρ(λ,m) (λ 2 + m 2 ) 2 . (2) In the limit of vanishingly small quark mass m and infi- nite volume V , the chiral condensate is proportional to the density of near-zero eigenvalues in accordance with the Banks-Casher relation [14], ¯ ψψ= πρ(0, 0). For T T c , chiral symmetry gets restored and the chiral or- der parameter ¯ ψψvanishes implying that ρ(0, 0) must also vanish. Motivated by the free theory limit at fi- nite temperature, where the spectral density ρ(λ, 0) has a gap up to the lowest fermion Matsubara frequency, 0 λ<πT , one possibility by which chiral symmetry restoration may occur in the chiral limit is through the generation of a gap in the infrared part of the eigenvalue spectrum. Such a scenario, however, would also lead to

Upload: others

Post on 22-Apr-2020

9 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Microscopic Origin of UA QCD - arXiv · 2018-10-02 · arXiv:1502.06190v1 [hep-lat] 22 Feb 2015 Microscopic Origin ofUA(1) Symmetry Violation in theHigh Temperature Phase of QCD Viktor

arX

iv:1

502.

0619

0v1

[he

p-la

t] 2

2 Fe

b 20

15

Microscopic Origin of UA(1) Symmetry Violation in the High Temperature Phase ofQCD

Viktor Dick,1 Frithjof Karsch,1, 2 Edwin Laermann,1 Swagato Mukherjee,2 and Sayantan Sharma2

1Fakultat fur Physik, Universitat Bielefeld, D-33615 Bielefeld, Germany2Brookhaven National Laboratory, Upton, NY 11973, USA

We investigate the low-lying eigenmodes of the Dirac matrix with the aim to gain more insightinto the temperature dependence of the anomalous UA(1) symmetry. We use the overlap operatorto probe dynamical QCD configurations generated with (2+1)-flavors of highly improved staggeredquarks. We find no evidence of a gap opening up in the infrared region of the eigenvalue spectrumeven at 1.5 Tc, Tc being the chiral crossover temperature. Instead, we observe an accumulation ofnear-zero eigenmodes. We argue that these near-zero eigenmodes are primarily responsible for theanomalous breaking of the axial symmetry still being effective. At 1.5 Tc, these near-zero eigenmodesremain localized and their distribution is consistent with the dilute instanton gas picture. At thistemperature, the average size of the instantons is 0.223(8) fm and their density is 0.147(7) fm−4.

PACS numbers: 12.38.Gc, 11.15.Ha, 11.30.Rd, 11.15.Kc

I. INTRODUCTION

Owing to the near-degeneracy and smallness of theup and down quark masses, the Quantum Chromo-dynamics (QCD) Lagrangian possesses an approximateUL(2) × UR(2) ≡ SUL(2) × SUR(2) × UV (1) × UA(1)symmetry. The fact that we do not see parity doublethadrons in our world implies that the SUL(2)× SUR(2)chiral symmetry is spontaneously broken down to theSUV (2) isospin symmetry of the vacuum. It is well knownfrom first principle lattice QCD studies that above thechiral crossover [1–4] temperature of Tc = 154(9)MeV[5] the chiral symmetry of QCD gets restored.

On the other hand, the axial UA(1) symmetry of theQCD Lagrangian is always broken due to the presence ofquantum fluctuations. This gives rise to the well-knownanomalous non-conservation of the axial current [6, 7].The explicit violation of the global UA(1) symmetry isdue to the presence of topologically nontrivial gauge fieldconfigurations [8]. Although UA(1) is not an exact sym-metry of QCD, the magnitude of its breaking near Tc isexpected to influence the nature of the chiral phase tran-sition in the limit of two vanishingly small light quarkmasses. Perturbative renormalization group studies ofmodel quantum field theories with the same global sym-metries as QCD suggest that if UA(1) is not effectivelyrestored at Tc, the chiral phase transition is of secondorder, belonging to the 3-dimensional O(4) universalityclass [9–12]. If the axial symmetry gets effectively re-stored for T ∼ Tc, the chiral phase transition can be ei-ther of first order [9, 10] or of second order with the sym-metry breaking pattern UL(2)×UR(2) → UV (2) [11, 12].In order to resolve the nature of the phase transition ofQCD with two light quark flavors it is thus important tounderstand the significance of the anomalous UA(1) inthe high temperature phase.

At low temperatures, UA(1) is also broken explicitlyby the presence of a non vanishing vacuum chiral con-densate. In the chirally symmetric phase, the vacuum

condensate vanishes and the mechanism of global UA(1)breaking can be studied directly. The microscopic mech-anism for UA(1) breaking in the chirally symmetric phaseof QCD presents an intriguing puzzle. The chiral conden-sate, which is the order parameter related to the restora-tion of chiral symmetry in QCD with massless quarks,can be expressed in terms of the eigenvalues λ of theDirac operator as

〈ψψ〉 V →∞→∫ ∞

0

dλ2m ρ(λ,m)

λ2 +m2, (1)

where ρ(λ,m) is the eigenvalue density. On the otherhand, UA(1) is not a global symmetry, so one cannotdefine a corresponding order parameter. For two lightquark flavors, an approximate restoration of UA(1) wouldresult in the degeneracy of the correlation functions of thepion and the scalar iso-triplet delta meson [13]. Specifi-cally, the difference of the integrated correlation functionsof these mesons in terms of the eigenvalues of the Diracoperator is

χπ − χδ =

d4x[

〈iπ+(x)iπ−(0)〉 − 〈δ+(x)δ−(0)〉]

V →∞→∫ ∞

0

dλ4m2 ρ(λ,m)

(λ2 +m2)2. (2)

In the limit of vanishingly small quark mass m and infi-nite volume V , the chiral condensate is proportional tothe density of near-zero eigenvalues in accordance withthe Banks-Casher relation [14], 〈ψψ〉 = πρ(0, 0). ForT & Tc, chiral symmetry gets restored and the chiral or-der parameter 〈ψψ〉 vanishes implying that ρ(0, 0) mustalso vanish. Motivated by the free theory limit at fi-nite temperature, where the spectral density ρ(λ, 0) hasa gap up to the lowest fermion Matsubara frequency,0 ≤ λ < πT , one possibility by which chiral symmetryrestoration may occur in the chiral limit is through thegeneration of a gap in the infrared part of the eigenvaluespectrum. Such a scenario, however, would also lead to

Page 2: Microscopic Origin of UA QCD - arXiv · 2018-10-02 · arXiv:1502.06190v1 [hep-lat] 22 Feb 2015 Microscopic Origin ofUA(1) Symmetry Violation in theHigh Temperature Phase of QCD Viktor

2

the vanishing of χπ − χδ, i.e. to the effective restorationof both chiral and UA(1) symmetry.

In fact, more rigorous calculations based on chiralWard identities for up to 4-point correlation functionsshow [15] that if the eigenvalue density for QCD withtwo light quark flavors is an analytic function in m2,it must have the form limm→0 ρ(λ,m) ∼ λ3 + O(λ4) inthe chirally symmetric phase, similar to that for the freetheory at T = 0. It was further shown [15] that, inthis case all correlation functions up to 6-point whichare related through UA(1) symmetry will be degener-ate, making the anomalous breaking of UA(1) invisible inthese correlation functions. Thus, if UA(1) breaking is fi-nite through the nondegeneracy of the 2-point correlationfunctions such as χπ − χδ, the eigenvalue density mustbe nonanalytic in m2. Two such possible forms of theinfrared eigenvalue spectrum, compatible with 〈ψψ〉 = 0but χπ − χδ 6= 0 for m → 0, have been speculated in[16], namely ρ(λ,m) ∼ m2δ(λ) and ρ(λ,m) ∼ |m|. Thefunctional form of the infrared eigenvalue density in thechirally symmetric phase of QCD remains an open andinteresting theoretical question.

The global UA(1) breaking at T = 0 is intimately con-nected to the presence of topologically nontrivial configu-rations of the QCD gauge fields [8]. It is well known thatlocalized topological structures like instantons give riseto zero modes of the Dirac operator and the correspond-ing wavefunctions remain localized [17]. The occurrenceof near-zero modes can possibly also be traced back tothe underlying topology of the gauge field configurations.For example, the particular form of the eigenvalue den-sity of the Dirac operator ρ(λ,m) ∼ m2δ(λ) in the in-frared can be motivated from the fact that a small shiftfrom zero of the near-zero modes resulting from the weakinteractions among widely separated instantons and an-tiinstantons can be neglected, leading to a δ(λ) behavior.The m2 factor naturally arises from the two light fermiondeterminants. At high enough temperatures, it has beenshown that the dilute gas of instantons is a reasonabledescription of pure gauge theory [18, 19]. Within thisapproximation the instanton density is suppressed withdecreasing value of the gauge coupling and eventuallyvanishes when T → ∞ [18]. This dilute gas model is ex-pected to be a good description of the high temperaturephase of QCD as well [18]. The UA(1) breaking can be ex-plained within such a model only for small enough valuesof the gauge coupling, i.e. at sufficiently high tempera-ture and for small sizes of the instantons. It is howeverunclear whether such a mechanism can explain the UA(1)breaking for the more relevant temperature range from Tcto a few times Tc. Near Tc the instantons and antiinstan-tons may not be widely separated and weakly interacting,as described by the Instanton Liquid Model (ILM) [20].In this model, chiral symmetry breaking arises due tothe fermion modes associated with strongly interactingand overlapping instantons. As the temperature is in-creased, it was proposed that there is a transition from aliquid phase of disordered instantons and antiinstantons

to a phase of instanton-antiinstanton molecules [21, 22].The chiral symmetry restoration at finite temperaturemay not necessarily be due to the suppression of the in-stantons [23] but rather due to the temperature depen-dence of the fermion determinant, which favors polarizedinstanton-antiinstanton molecules. Thus, it is entirelypossible that for T . Tc . 2Tc other nonperturbativemechanisms responsible for UA(1) breaking may gener-ate an accumulation of near-zero modes leading to a morecomplex form of the infrared eigenvalue spectrum.

Since topological structures in QCD are inherentlynonperturbative, lattice QCD techniques are ideallysuited to address issues related to the UA(1). Anoma-lous UA(1) breaking at high temperature was studied onthe lattice by looking at the nondegeneracy of the 2 pointcorrelation functions χπ−χδ using staggered fermion for-mulation, which preserves a remnant of the continuumchiral symmetry on the lattice, in [24] and more recentlywith an improved staggered fermion formulation in [25].In both cases, it was observed that UA(1) was not effec-tively restored at T & Tc. Recently, this issue was re-visited in a study of the infrared eigenvalue spectrum ofhighly improved staggered quarks (HISQ) [26] and a sim-ilar conclusion was reached [27]. However, for staggeredfermions the connection between topology and fermionzero modes and hence the reproduction of the index the-orem is a very subtle issue [28]. Attempts to study thisproblem with fermions with exact chiral symmetry onthe lattice have produced juxtaposing results. Studieswith domain wall fermions [16, 29] but with a heavierthan physical pion mass of 200 MeV support the sce-nario that UA(1) remains broken for T & Tc. The eigen-value density of the Dirac operator for T . 1.2Tc has asmall peak structure in the infrared favoring the form ofρ(λ,m) ∼ m2δ(λ), which can largely account for the ori-gin of the axial anomaly. However, even in these studies aclear separation of the zero and near-zero modes was notpossible due to moderate residual chiral symmetry break-ing effects induced by the mixing of the left and righthanded fermions along the finite fifth dimension. Onthe other hand, another independent preliminary studywith so-called optimal domain wall fermions and physicalpion mass reports the restoration of UA(1) above Tc [30].A study using overlap fermions restricted to the trivialtopological sector of QCD on relatively small volumesalso suggests that UA(1) is effectively restored at T ∼ Tc[31]. However, it is well known that simulations withfixed topology are more sensitive to finite volume effectsand, at present, it is difficult to perform calculations withlarger lattice volumes or for fluctuating topology in thecase of overlap fermions due to prohibitively large com-putational costs.

Measuring the underlying topology of the gauge fieldson the lattice requires careful analysis. The gauge fieldsare defined as links connecting the adjacent lattice sites,which can be continuously deformed to unity. To studylocalized topological structures one has to remove theultraviolet fluctuations of the fields or approach succes-

Page 3: Microscopic Origin of UA QCD - arXiv · 2018-10-02 · arXiv:1502.06190v1 [hep-lat] 22 Feb 2015 Microscopic Origin ofUA(1) Symmetry Violation in theHigh Temperature Phase of QCD Viktor

3

sively to the minimum of the classical action. The latteris done by cooling the gauge configurations [32]. The ul-traviolet fluctuations can be reduced using smearing [33],which involves replacing each gauge link by an averageover the neighboring links. It is then possible to mea-sure the topological charge on the lattice using the dis-cretized version of the integrated FF operator. The mostpopularly used smearing technique, known as the hyper-cubic (HYP) smearing [34], has been shown to providea good estimate of the topological susceptibility. How-ever, successive smearing may lead to small instantonsbeing undetected and a change in the large scale struc-ture of the gauge fields. Alternatively, one can make useof the index theorem [35], which relates the difference be-tween the number of right and left-handed fermion zeromodes to the topological charge of the gauge fields. Theadvantage of this method is that it naturally connectstopological structures and the eigenmodes of the Diracoperator, which are global quantities and depend on thegauge links of the entire lattice.

In the present work we address the temperature depen-dence of UA(1) and probable microscopic mechanisms re-sponsible for its breaking in the high temperature phaseof QCD by studying the infrared eigenmodes of over-lap fermions [38] on the background of dynamical (2+1)flavors of HISQ gauge field configurations with nearlyphysical fermion masses and large volumes. This HISQdiscretization scheme has been used for extensive studieson QCD thermodynamics [5, 36] and has small discretiza-tion errors, resulting in the least taste symmetry breakingamong all commonly used staggered fermion discretiza-tions. Preliminary studies with HISQ fermions [37] alsoprovide hints that in the continuum limit for two vanish-ingly small light quark masses, the QCD chiral transitionmay belong to the 3-dimensional O(4) universality class.The issues about the lack of an index theorem of theHISQ is overcome by using the Overlap Dirac fermionsto probe the topology of the HISQ gauge configurations.Overlap Dirac fermions circumvent the Nielsen-NinomiyaNo-go theorem [39] by sacrificing the ultra-locality crite-rion, preserve an exact chiral symmetry [38] and an exactindex theorem [40] and reproduce the correct anomaly[41] even at nonzero lattice spacing. Employing the indextheorem for the overlap fermions, the topological struc-tures in SU(2) [19] as well as SU(3) pure gauge theories[19, 42] have been studied earlier. However, the rela-tionship between chiral and UA(1) symmetry also cannotbe addressed within the framework of pure gauge theoryand the presence of light dynamical fermions is necessaryto address this question. Furthermore, it is a priori notevident whether the same dilute instanton gas picturealso applies for QCD with near-physical, light dynami-cal fermions as the presence of light fermions would leadto interactions between instantons and induce anomalyeffects.

This work is structured as follows: In Sec. II we pro-vide all necessary computational details pertaining tothis work. In III A we check the distribution of the topo-

logical charge measured by using the exact index theoremof overlap fermions. In III B we present and discuss ourresults on the eigenvalue distribution of overlap fermionson the dynamical HISQ configurations. The contribu-tion of the low-lying eigenmodes towards UA(1) break-ing is discussed in III C, while IIID contains our resultsabout the functional form of the eigenvalue density andits implications for UA(1) breaking. In III E we verify therobustness of the occurrence of near-zero modes. FromIII F to III G we discuss various different properties ofthe zero and near-zero eigenmodes. Finally, in IV wesummarize and conclude this work. Preliminary resultsof this work were previously presented in [43].

II. COMPUTATIONAL DETAILS

The set of (2 + 1)-flavor HISQ configurations used inthis work was generated by the HotQCD collaboration[5]. Two lattice sizes were used in this study, 243× 6 and323 × 8. The strange quark mass ms is set to its phys-ical value and the light quark mass in all these sets ofconfigurations are chosen to be ml = ms/20, which cor-responds to a Goldstone pion mass of mπ = 160MeV inthe continuum. We studied 5 sets of configurations, twoat T ∼ Tc, two at T ∼ 1.2Tc and one at T ∼ 1.5Tc. Here,Tc = 154(9) MeV [5] is the chiral crossover temperaturein the continuum limit. Near Tc, in addition, we studiedconfigurations generated by the Bielefeld-BNL collabora-tion [37] with lattice size 323 × 6 and a light quark massof ml = ms/40, which corresponds to mπ = 110 MeV.This was to study whether the UA(1) breaking survivesas the chiral limit is approached. We considered 90–160configurations of each set, typically separated by 100 tra-jectories, and computed the eigenvalues of the overlapDirac operator on them. The lattice sizes, strange tolight quark mass ratio, temperatures and relevant statis-tics are shown in Tab. I.

N3

σ ×Nτ ml/ms T [MeV] N Nλ

243 × 6 1/20 162.3 120 200

323 × 6 1/40 162.3 90 400

323 × 8 1/20 165.6 120 200

243 × 6 1/20 199.0 100 100

323 × 8 1/20 196.0 100 100

323 × 8 1/20 237.1 160 50

Tab. I: Lattice size (N3

σ × Nτ ), mass ratio (ml/ms), tem-perature (T ), number of configurations (N) and number ofeigenvalues that were computed per configuration (Nλ) foreach ensemble.

We probe the low-lying eigenmodes of these HISQgauge ensembles through the use of the massless over-lap Dirac fermion operator

Dov =M [1 + γ5sgn [γ5DW (−M)]] , (3)

Page 4: Microscopic Origin of UA QCD - arXiv · 2018-10-02 · arXiv:1502.06190v1 [hep-lat] 22 Feb 2015 Microscopic Origin ofUA(1) Symmetry Violation in theHigh Temperature Phase of QCD Viktor

4

where DW is the standard Wilson Dirac operator withthe parameter 0 < M < 2.

For the implementation of the sign function in theoverlap operator, we computed the lowest 20 eigenvec-

tors of D†WDW using the Kalkreuter-Simma (KS) Ritz

algorithm [44]. The sign function was computed forthese low modes explicitly, while for the higher modesit was approximated by a Zolotarev rational function.The number of terms in the Zolotarev function was keptto be 15. The overlap operator satisfies the Ginsparg-Wilson(GW) relation with a deviation of no more than10−7 at low temperatures and 10−10 at high tempera-tures. The square of the sign function deviated fromidentity by about 10−7–10−9.

For each temperature, 50 lowest eigenvalues of D†ovDov

were computed using the KS algorithm. The zero modesof D†

ovDov come with chiralities ±1. The nonzero eigen-values come in degenerate pairs with chiralities havingopposite signs but equal magnitudes, which is usuallydifferent from unity. These features of the spectrum al-low us to distinguish between the near and exact zeromodes within a few iterations of the KS algorithm. TheKS algorithm was run until the relative error on thenonzero eigenvalues ofD†

ovDov were estimated to be lowerthan 10−4 on average and the separation between zeroand nonzero modes was clearly seen. In most cases, thenumber of eigenvalues was later increased by computingthe eigenvalues of PDovP , where P is the projection torighthanded or lefthanded modes. This projected opera-tor has the advantage that it only takes half the time tobe applied and each nonzero eigenvector can be related toa pair of eigenvectors of D†

ovDov, further reducing com-putation time and memory requirement. However, thealgorithm becomes quite unstable if the subspace thatP projects onto contains zero modes and can only beused on the opposite chiralities after identifying the zeromodes.

For some applications, the eigenvectors of Dov wererequired and not only those of D†

ovDov. Each degener-ate pair of nonzero eigenvectors of the squared opera-tor spans a two-dimensional space that also contains twoeigenvectors of Dov, which are related to each other byan application of γ5 and have eigenvalues that are com-plex conjugates of each other. These eigenvectors couldbe obtained by applying an appropriate unitary transfor-mation to each of the original pairs.

We also checked the optimal value of the parameterMused in the construction of the overlap operator. Frompartially quenched studies it is known that for certain

choices of M , the corresponding D†WDW can have very

small eigenvalues, leading to the presence of spurious zeromodes in the overlap operator [45]. We verified that forconfigurations without zero modes the choice of M didnot affect the eigenvalues significantly within our preci-sion. Moreover, for configurations with zero modes wechose M such that the sign function and the GW rela-tion were determined with highest accuracy, ensuring thebest implementation of the overlap operator. Except for

a few cases, especially near Tc, we chose M = 1.8.

For the 323 × 6 lattice with a light quark mass ofml = ms/40, the gauge configurations were rough and

the convergence of the KS for D†WDW was slow, leading

to imprecise estimates of the GW relation and the signfunction. In this case, we did two levels of HYP smear-ing to smoothen out the ultraviolet fluctuations. Thisenabled us to achieve a more precise estimation of theoverlap sign function, similar to the precision achievedfor the other ensembles. The effects of the smearing arefurther discussed in Section III E.

III. RESULTS

A. Topological charge distributions

The topological charge Q was measured by countingthe number of zero-modes of the overlap operator anddetermining their chiralities,

Q = n+ − n− , (4)

where n+ (n−) is the number of zero modes with chiral-ity +1 (−1). Since the underlying HISQ gauge configu-rations were generated at zero strong CP violating angleθ, at any temperature all the topological sectors shouldbe spanned. However, the configurations may have beentrapped in one topological sector and the autocorrelationtimes in such cases may be large. To avoid autocorrela-tion effects we usually chose the configurations to be sep-arated by 100 Rational Hybrid Monte-Carlo trajectories.The time histories of the topological charge are shown inFig. 1. It is evident that the autocorrelation effects areunder control. The distribution is ergodic enough andon average 〈Q〉 ≃ 0, for all temperatures. This gives usconfidence that the statistics is sufficient in our presentstudy. Moreover, we also observe many configurationswith |Q| ≥ 1 in all studied ensembles. In fact, even at1.5Tc more than a third of the total number of configu-rations have |Q| = 1, confirming the importance of theQ 6= 0 configurations.

B. Eigenvalue spectra

In this section we show the eigenvalue density of theoverlap operator at three temperatures – near Tc, at1.2Tc and at a yet higher temperature of 1.5Tc. Theoverlap fermion matrix in Eq. (3) is a normal matrix. In

the complex plane, its dimensionless eigenvalues, λ, lieon a circle centered at M and with a radius M , obeying|λ−M |2 = M2. The eigenvalues measured in our studylie on the circle very close to the origin with a very smallreal part. Hence, we always plot the eigenvalue densityas a function of λ, where aλ = Imλ. On the lattice, the

Page 5: Microscopic Origin of UA QCD - arXiv · 2018-10-02 · arXiv:1502.06190v1 [hep-lat] 22 Feb 2015 Microscopic Origin ofUA(1) Symmetry Violation in theHigh Temperature Phase of QCD Viktor

5

-20

-10

0

10

20 Q

-20

-10

0

10

20 Q

-20

-10

0

10

20 Q

-15-10-505

1015

Q

-15-10-505

1015

Q

-15-10-505

1015

Q

∼ Tc, Nτ = 8,mπ = 160MeV

∼ Tc, Nτ = 6,mπ = 160MeV

∼ Tc, Nτ = 6,mπ = 110MeV

1.2Tc, Nτ = 8,mπ = 160MeV

1.2Tc, Nτ = 6,mπ = 160MeV

1.5Tc, Nτ = 8,mπ = 160MeV

Fig. 1: Time histories of the topological charge, calculatedfrom the zero modes of the overlap Dirac operator, for theHISQ configurations at T ∼ Tc, T ∼ 1.2 Tc and T ∼ 1.5 Tc.Configurations belonging to the same production stream areconnected with lines.

eigenvalue density is defined as

a3ρ(λ) =1

N3σNτ

i

δ (aλ− aλi) , (5)

where the sum only includes values on the left part ofthe semicircle with Reλ < M , and excludes values nearλ = 2M even though their imaginary part would also besmall.The eigenvalue distribution at three different temper-

atures is shown in Figs. 2, 3 and 4. The spectrum istruncated at some large eigenvalue since we measure onlya finite number of them. We indicate the point beyondwhich the spectrum is not trustworthy anymore by a ver-tical line in red in each of the plots. It is estimated byfirst taking the highest computed eigenvalue of each con-figuration and then taking the minimum of these valuesover all the analyzed configurations.As we emphasized earlier, the KS algorithm allowed

us to distinguish the zero modes from the near-zeromodes using the chirality properties of the correspondingeigenvectors. In general, the eigenvalue distribution has

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8

λ/T

ρ(λ)/T 3

∼ Tc 323 × 8mπ = 160MeV

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8

λ/T

ρ(λ)/T 3

∼ Tc 243 × 6mπ = 160MeV

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8

λ/T

ρ(λ)/T 3

∼ Tc 323 × 6mπ = 110MeV

Fig. 2: The eigenvalue density of the overlap operator on323 × 8, 243 × 6 and 323 × 6 HISQ configurations near Tc.Only nonzero modes are included. The vertical line denotesthe range of validity due to the finite number of computedeigenvalues.

three distinct features – the zero mode peak, a near-zeromode accumulation and the bulk eigenvalue region. NearTc, the first bin contains a large contribution from zeromodes which are omitted in Fig. 2 to focus on the infraredphysics of only the near-zero eigenvalues. At this tem-perature, we do not observe any gap in the infrared partof the eigenvalue spectrum. The near-zero modes andthe bulk modes appear to overlap significantly and thenear-zero modes tend to develop a peak towards the in-frared region. This peak becomes sharper as the light seaquark mass is lowered from ml = ms/20 to ml = ms/40at fixed lattice spacing 1/6T . It also becomes sharperwhen we go to a finer lattice, from Nτ = 6 to Nτ = 8

Page 6: Microscopic Origin of UA QCD - arXiv · 2018-10-02 · arXiv:1502.06190v1 [hep-lat] 22 Feb 2015 Microscopic Origin ofUA(1) Symmetry Violation in theHigh Temperature Phase of QCD Viktor

6

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8

λ/T

ρ(λ)/T 3

1.2Tc 323 × 8

λ = 0

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8

λ/T

ρ(λ)/T 3

1.2Tc 243 × 6

λ = 0

Fig. 3: The eigenvalue density for HISQ configurations usingthe overlap operator at 1.2 Tc. The lattice sizes are 323 × 8and 243 × 6, respectively. The red line marks the range ofvalidity.

at a fixed pion mass of 160MeV. This trend suggeststhat the near-zero mode accumulations will remain asthe chiral and the continuum limits are approached.

At temperatures 1.2Tc and 1.5Tc, both the zero modesdenoted by the red bar and the near-zero and bulk modesare shown in Fig. 3 and Fig. 4. The separation betweenthe near-zero mode accumulation and the bulk eigenvalueregion becomes even more evident with increasing tem-perature. At 1.2Tc, we study the eigenvalue spectrumat two different lattice spacings to estimate whether theinfrared part of the spectrum is strongly affected by thelattice cutoff effects at higher temperatures. Keeping thephysical bin size the same in units of λ/T for compari-son, we observe that the infrared region of the eigenvaluedensity remains practically unchanged when the latticespacing goes from 1/6T to 1/8T at a fixed temperature T .This gives us confidence that the near-zero modes are notdue to dislocations of the gauge fields. A more detailedstudy about the lattice artifacts is given in Sec. III D.

The number of zero and near-zero modes both de-crease as the temperature is increased to 1.5Tc as shownin Fig. 4. There is a small peak of near-zero modes,while the number of bulk eigenvalues starts to rise veryslowly and only gives a significant contribution beyondλ0 ≃ 0.4T . This is reminiscent of some kind of bandedge separating the two different regimes of eigenvalues,

0

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6

λ/T

ρ(λ)/T 3

1.5Tc 323 × 8

λ = 0

0

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6

λ/T

ρ(λ)/T 3

1.5Tc 323 × 8only Q = 0

Fig. 4: The eigenvalue density for 323×8 HISQ configurationsusing the overlap operator for 1.5 Tc for all values of Q andalso separately for the Q = 0 sector. The red line marks therange of validity.

which is studied in detail in Sec. IIIG. Even at this tem-perature we do not observe a gap in the infrared sector ofthe eigenvalue spectrum. The presence of these near-zeromodes is not due to the fact that we are also samplingconfigurations belonging to nonzero topological sectors inour study. This is evident from the lower panel of Fig. 4,where we show the eigenvalue distribution of only thoseconfigurations with topological charge Q = 0 at 1.5Tc.The presence of near-zero modes is also observed for thisparticular subset of configurations.

C. Near-zero eigenmodes and axial symmetry

breaking

As introduced in Sec. I, ω ≡ χπ − χδ as defined inEq. (2) is a measure that quantifies UA(1) breaking.Through a chiral Ward identity it can also be obtained

ω =〈ψψ〉m

− χconn (6)

from the chiral condensate⟨

ψψ⟩

= TV

tr(

D−1m ∂mDm

)⟩

and the connected chiral susceptibility χconn =TV

∂mtr(

D−1m ∂mDm

)⟩

, where Dm = Dov(1−am/2M)+am is the Dirac operator for overlap quarks with a (va-lence) quark mass m.

Page 7: Microscopic Origin of UA QCD - arXiv · 2018-10-02 · arXiv:1502.06190v1 [hep-lat] 22 Feb 2015 Microscopic Origin ofUA(1) Symmetry Violation in theHigh Temperature Phase of QCD Viktor

7

Thus, in terms of the eigenvalues of the overlap oper-ator

a2ω =1

N3σNτ

[ 〈|Q|〉(am)2

(7)

+

λ6=0

2(am)2(4M2 − |λ|2)2[

|λ|2(4M2 − (am)2) + 4(am)2M2

]2

.

The first term is the contribution from the zero modeswhich vanishes in the thermodynamic limit.Thus, having determined the low-lying eigenvalues of

the overlap operator, ω can be computed from them.However, to explore the physics of the underlying HISQconfigurations the overlap valence quark mass that en-ters Eq. (7) has to be tuned against some physical quan-tity measured on the same gauge configurations. In thepresent work we adopt a very simple strategy: we roughlytune the strange valence quark mass, ms, and then studyω as a function of the light valence quark mass within arange of ml = ms/20 to ml = ms/2. To tune the strangevalence quark mass we use the renormalized difference ofthe pseudo-scalar (ηss) and the (connected) scalar sus-ceptibilities in the strange quark sector, m2

sωs/T4. We

calculated this renormalized quantity using strange va-lence overlap fermions and matched it with the samequantity calculated independently for the strange HISQsea quarks. For the overlap fermions this quantity wasestimated in two parts. We first calculated this quan-tity from the predetermined low-lying eigenvalues usingEq. (7), and then added the contribution of higher eigen-values by performing inversions on the eigenspace orthog-onal to the low-lying eigenmodes using random sourcevectors. The result of the strange valence quark masstuning near Tc is shown in Fig. 5. In order to monitorfinite volume effects, we considered the exact zero modesseparately and do not observe a significant contributionfrom them to this quantity. The tuned masses obtainedwith and without the zero modes differ by about 5%.In Fig. 6, we show a renormalized measure of UA(1)

breaking, namely mlmsω/T4, for a range of the light va-

lence quark mass between ms/20 and ms/2. Assuminga Breit-Wigner distribution for the near-zero mode peakto model the δ(λ) like distribution discussed in the intro-duction, i.e. ρ(λ)/T 3 = ρ0A/(A

2+λ2), the contributionsof the near-zero modes in this renormalized measure ofUA(1) breaking can be characterized as

mlms

T 4ω ∝ ρ0ms

A+ 2ml

(A+ml)2. (8)

In our partially quenched setup, where only the valencelight quark mass ml is varied, this quantity has a finitevalue in the chiral limit ml → 0. In a full dynamicalsetup, its behavior will be governed by the dependenceof A and ρ0 on the light sea quark mass. The simplestcase of both A and ρ0 being proportional to the lightsea quark mass, which is compatible with the trends

1.8

2

2.2

2.4

2.6

2.8

3

3.2

3.4

3.6

3.8

0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2

ams

m2sωs/T 4

323× 8

243× 6

Fig. 5: The tuning of the strange valence quark mass for over-lap fermions on Nτ = 6 (blue) and Nτ = 8 (red) lattices nearTc. The horizontal lines mark the results for m2

sωs/T4 (cf.

Eq. (6)) independently obtained for the HISQ sea fermions.Filled (empty) points denote the overlap result with (without)the zero mode contribution.

discussed in Sec. III D, will also give a finite value andshould be approximated in a partially quenched study.For light valence quark masses near to or smaller thanthe smallest near-zero eigenvalues, the computed quan-tity will approach zero, which is visible for the lowestmasses near Tc. This is a finite volume effect in the sensethat a larger volume will sample more eigenvalues in thenear-zero mode region, which reduces the magnitude ofthe smallest eigenvalue and pushes this effect towardszero. Beyond that, we observe a smooth dependence onml/ms that is compatible with the Breit-Wigner ansatzand which is independent of the lattice spacing. It is alsoevident from Fig. 6 that the contribution of the near-zero modes to ω is substantially larger than that fromthe bulk modes. When going from T ∼ Tc to T ∼ 1.2Tc,this quantity does not decrease significantly, supportingour conclusion that UA(1) is not effectively restored si-multaneously with the chiral symmetry.

D. The functional form of the eigenspectra

In order to understand the general functional form ofthe eigenvalue density and in particular the near-zero re-gion, we make a fit ansatz consisting of a Breit-Wignerpeak for the near-zero modes and a polynomial behaviorfor the bulk part of the spectrum of the form,

ρ(λ)

T 3=

ρ0A

A2 + λ2+ cλα . (9)

We address three issues in this section. Firstly, a gen-eral idea about the dependence of the near-zero modepeak on the sea quark mass is necessary to understandwhat happens in the chiral limit. Secondly, it is impor-tant to check the dependence of the near-zero modes onthe lattice cutoff to establish that these are physical andnot mere lattice artifacts. Finally, the leading exponent

Page 8: Microscopic Origin of UA QCD - arXiv · 2018-10-02 · arXiv:1502.06190v1 [hep-lat] 22 Feb 2015 Microscopic Origin ofUA(1) Symmetry Violation in theHigh Temperature Phase of QCD Viktor

8

0

0.5

1

1.5

2

2.5

3

3.5

0 0.1 0.2 0.3 0.4 0.5

ml/ms

mlms (χπ − χδ) /T4

∼ Tc

Nτ = 8, λ0 = 0.35TNτ = 6, λ0 = 0.50T

0

0.5

1

1.5

2

2.5

0 0.1 0.2 0.3 0.4 0.5ml/ms

mlms (χπ − χδ) /T4 1.2Tc

Nτ = 8, λ0 = 0.3TNτ = 6, λ0 = 0.4T

Fig. 6: A renormalized measure of UA(1) breaking for a rangeof valence light quark masses, ms/20 ≤ ml ≤ ms/2 for en-sembles with different Nτ at T ∼ Tc and T ∼ 1.2 Tc. Thefilled points denote the contribution from near-zero modes(λ < λ0), while the empty points were calculated only fromthe bulk modes (λ > λ0).

that characterizes the rise of the bulk also provides in-formation regarding the restoration of UA(1), hence itsdependence on the temperature and lattice cutoff needsto be studied.

The fit to the eigenvalue spectrum near Tc for differ-ent sea quark masses is shown in Fig. 7. The error barsfor each bin have been determined by a jackknife proce-dure over the set of gauge configurations. The param-eter A, characterizing the width of the near-zero modepeak, falls from 0.35(4)T to 0.151(7)T when going fromml = ms/20 to ms/40 and the prefactor ρ0, indirectlycontrolling the height of the peak, also goes down from0.28(3)T to 0.191(7)T . This generally supports the pic-ture of a delta function like peak forming in the chirallimit, with a decreasing peak height. However, as far asone can tell from two data points, the dependence of ρ0does not seem to be quadratic in the light sea quark massnear Tc, arguing against the dilute instanton gas pictureas a good description of QCD at this temperature.

To examine cutoff effects one needs to compare a renor-malized version of the eigenvalue density at different lat-tice spacings. One way to renormalize the eigenvaluesis to scale them by the previously tuned strange quarkmass. The corresponding renormalized eigenvalue den-sity is then msρ(λ) since it leaves the quantity ms〈ψψ〉

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8

ρ(λ)/T3

λ/T

~ Tc

mπ=160 MeVmπ=110 MeV

Fig. 7: Eigenvalue distribution at T ∼ Tc for Nτ = 6 and twodifferent sea quark masses compared to fits with Eq. (9).

unchanged under renormalization. We therefore take thesame ansatz as in Eq. (9) with λ→ λ/ms and the densityreplaced by its renormalized definition msρ(λ)/T

4. Fitsto the renormalized spectrum in dimensionless units attwo different temperatures, Tc and 1.2 Tc, are shown inFig. 8. Taking a closer look at the near-zero mode peak,it is evident that the accumulation of near-zero modes isalmost independent of the lattice spacing and is unlikelyto be just a lattice artifact.

Finally, we discuss the exponent α of the characteristicλα rise of the bulk eigenvalues. As mentioned in Sec. I,under the assumption of analyticity of the eigenvalue den-sity in m2, detailed analytical calculations based on upto 4-point chiral Ward identities show [15] that in thechiral symmetric phase of QCD the leading λ depen-dence should be similar to that for the free theory, i.e.limm→0 ρ(λ,m) ∼ λ3. In such a case, the effect of UA(1)breaking should be invisible in at least up to 6-point cor-relation functions. In light of this, it is interesting tocharacterize the rise of the bulk eigenvalues. As shownin Fig. 7, near Tc the rise of the bulk eigenvalues for ourtwo Nτ = 6 lattices with quark masses ml = ms/20 andml = ms/40 is described by the exponents α = 0.92(5)and α = 0.98(4), respectively. A fit to the renormal-ized eigenvalue spectrum near Tc further yields 0.86(2)for Nτ = 8, see Fig. 8. Thus, near Tc a linear rise of bulkeigenvalues is favored for both lattice spacings and quarkmasses. Interestingly, chiral perturbation theory [46] andthe ILM [47] show that a contribution to ρ(λ) linear inλ, which results in a non vanishing connected susceptibil-ity χconn, is absent for two light flavors and only presentfor Nf > 2. On the other hand, for staggered fermionsaway from the continuum limit, taste violations lead toχconn 6= 0 also for Nf = 2 [48].

At 1.2Tc the rise of the bulk eigenvalues has a differentexponent. From Fig. 8 it is evident that α ≃ 2, indepen-dent of the lattice spacing. Note that a bulk eigenvaluedensity rising quadratically with λ does not contribute to

Page 9: Microscopic Origin of UA QCD - arXiv · 2018-10-02 · arXiv:1502.06190v1 [hep-lat] 22 Feb 2015 Microscopic Origin ofUA(1) Symmetry Violation in theHigh Temperature Phase of QCD Viktor

9

0

0.2

0.4

0.6

0.8

1

1.2

1.4

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8

ρ(λ)ms/T4

λ/ms

~ Tc

Nτ=8

Nτ=6

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8

λ/ms

ρ(λ)ms/T4 1.2 Tc

Nτ=8Nτ=6

Fig. 8: The renormalized eigenvalue spectra near Tc (top) andat 1.2 Tc (bottom), both for two different lattice spacings. Thelines indicate fits to Eq. (9) with λ → λ/ms and ρ(λ)/T 3 →msρ(λ)/T

4.

ω. A similar linear behavior for T ∼ Tc and a quadraticrise for T ∼ 1.2Tc of the bulk eigenvalues was also ob-served in the previous study with domain wall fermions[29] with a heavier pion mass, corroborating that the bulkrise is independent of the sea quark mass. On the otherhand, the characteristic free theory like cubic rise of thebulk eigenvalue density was only observed at 1.5Tc asshown in Fig. 9. At 1.5Tc, the near-zero mode peak re-duces significantly and the separation between the bulkand the near-zero modes is distinctly visible. Our fitansatz for the bulk is modified accordingly as (λ − λ0)

α

to represent this feature, which gives a smaller χ2 per de-grees of freedom than the original ansatz in Eq. (9). Theparameter α and the goodness of fit at different temper-atures are compiled in Tab. II.

E. Robustness of the zero and near-zero modes

Detecting topological objects with fermion zero modeshas the advantage that by construction the zero modesdepend on all the gauge links distributed on a lat-

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 0.2 0.4 0.6 0.8 1 1.2 1.4

ρ(λ)/T3 1.5 Tc

λ/T

Fig. 9: Eigenvalue distribution at 1.5 Tc together witha (λ− λ0)

α fit to the bulk eigenmodes.

N3

σ ×Nτ ml/ms T [MeV] α χ2/dof

243 × 6 1/20 162.3 0.92(5) 1.32

323 × 6 1/40 162.3 0.98(4) 1.84

323 × 8 1/20 165.6 0.86(2) 0.92

243 × 6 1/20 199.0 1.9(2) 1.16

323 × 8 1/20 196.0 1.9(1) 1.21

323 × 8 1/20 237.1 3.0(4) 1.30

Tab. II: Lattice size (N3

σ ×Nτ ), mass ratio (ml/ms), temper-ature (T ), the exponent α characterizing the λα rise of thebulk eigenvalues λ and the goodness of the fits performed onthe eigenvalue distribution.

tice. Still, if the underlying gauge fields are not smoothenough, the method might be hampered by the pres-ence of unphysical fermion modes localized on structurescalled dislocations, which typically have a smaller clas-sical action than instantons. These are lattice artifacts,i.e. effects of finite lattice spacing, and should disappearas the continuum limit is approached. It is therefore im-portant to make sure that the observed infrared fermionmodes are physical and do not solely arise as lattice ar-tifacts. If the zero and near-zero eigenvalues are entirelydue to the presence of dislocations, they are expected todisappear as the gauge fields are smoothed using smear-ing techniques. To check this we performed HYP smear-ing on the 323 × 8 configurations with ml/ms = 1/20 at1.5Tc.Such smoothing methods are mandatory if one wants

to compute the topological charge by means of a dis-cretized version of its field theoretic definition

Q =1

32π2

d4xF aµν F

aµν . (10)

The optimal number of smearing levels is usually chosensuch that the topological charge measured this way on thesmeared configurations has a value close to an integer.

Page 10: Microscopic Origin of UA QCD - arXiv · 2018-10-02 · arXiv:1502.06190v1 [hep-lat] 22 Feb 2015 Microscopic Origin ofUA(1) Symmetry Violation in theHigh Temperature Phase of QCD Viktor

10

0

0.1

0.2

0.3

0.4

-6 -4 -2 0 2 4 6Q

P(Q)

1.5 TcOV-unsmeared

HYP smeared

Fig. 10: The distribution of the topological charge, Q, at1.5 Tc for 323 × 8 HISQ configurations measured from thezero modes of the overlap operator and also from the gluonicoperator FF on the same gauge configurations after 10 levelsof HYP smearing.

On our present lattices it turned out that 10 levels ofHYP smearing were sufficient to give an integer value ofthe FF operator summed over the whole lattice at 1.5Tc.On all of these smoothed configurations we could verifythat the value of Q obtained from FF matched exactlywith the topological charge obtained by counting the zeromodes of the overlap operator after smearing.

In Fig. 10, we compare the histogram for the topologi-cal charge measured by counting the fermion zero modesof unsmeared configurations with that measured usingthe purely gluonic observable FF on the same config-urations after smearing. Apparently, some zero modesdisappear in the course of the smearing process; yet, thecomparison indicates that the fermionic zero modes forthese configurations do not arise only due to gauge fielddislocations and are reflecting continuum physics.

A similar behavior is observed for the near-zero modes.In Fig. 11, the comparison of the eigenvalue spectrum ofthe overlap operator on the original unsmeared config-urations with the smeared ones reveals that the num-ber of near-zero modes is somewhat reduced by smearingbut that the near-zero mode accumulation is still presenteven after a substantial amount of smearing, indicatingthat these are not mere lattice artifacts like dislocations.We also found that the typical eigenvectors associatedwith the near-zero modes on a smeared configuration ap-pear to be slightly less localized compared to the un-smeared case, suggesting that the reduction of the near-zero modes may be caused by the loss of small instantonsdue to smearing.

Finally, as already shown in Fig. 8, near Tc as well asat 1.2Tc the comparison of the renormalized eigenvaluespectra at two different lattice spacings, 1/6T and 1/8T ,indicates that the near-zero mode accumulation remainsnearly unchanged as the lattice spacing is reduced. Thus,it is unlikely that the near-zero mode accumulation arisesprimarily due to lattice artifacts.

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0 0.5 1 1.5 2λ/T

ρ(λ)/T 3

before smearing

after smearing

Fig. 11: The eigenvalue density at 1.5 Tc and Nτ = 8 beforeand after 10 steps of HYP smearing. The empty boxes are ob-tained when including zero modes and the dashed lines markthe ranges of validity.

F. Profiles of the zero and near-zero modes at high

temperature

The fermion zero mode associated with an instanton is,at T = 0, localized in the region occupied by the instan-ton. At nonzero temperature however, the compactifica-tion of Euclidean time leads to periodic copies of instan-tons. Such classical finite action solutions of the gaugefields on the manifold R3 × S1 are known as calorons.When the instanton size is much smaller than 1/T , thecopies do not feel the effect of the neighbors and behavelike zero temperature instantons. However, the overlapbetween the instanton copies may be larger and theirsizes become comparable to 1/T . Explicit solutions areknown for trivial [49] as well as nontrivial [50, 51] holon-omy. In the case of trivial holonomy it has been observedthat the caloron turns into a magnetic monopole [52] inpure gauge theory. Calorons with nontrivial holonomyhave more interesting features with monopole substruc-tures [53]. However, in our studies we have not attemptedto take a detailed look at the monopoles.

In order to gain more inside into the structure of theinfrared modes in QCD, we looked at the profiles of thezero and the near-zero modes at 1.5Tc. Examples rep-resenting the majority of configurations with topologicalcharge |Q| = 1 are shown in Figs. 12 and 13. In thesefigures the density of the wavefunctions, ψ†(x)ψ(x), ismeasured along two spacetime directions, summing overthe other two directions and the internal degrees of free-dom (color and spin). We observe that the zero modes arelocalized along the spatial as well as the temporal direc-tions. In a few cases, the width of the zero modes in thecompact temporal direction is somewhat larger and thedensity profile represents the overlap between the nearestcopies.

The near-zero modes typically exhibit a two-peakstructure, in the density as well as in the chirality(ψ†(x)γ5ψ(x)) profile, cf. Fig. 14. In the latter case, the

Page 11: Microscopic Origin of UA QCD - arXiv · 2018-10-02 · arXiv:1502.06190v1 [hep-lat] 22 Feb 2015 Microscopic Origin ofUA(1) Symmetry Violation in theHigh Temperature Phase of QCD Viktor

11

0 1

2 3

4y (fm) 0

1

2

3

4

t (fm)

0

0.04

0.08

0.12

0.16

ψ+ψ

0

0.04

0.08

0.12

0.16

0 1

2 3

4

x (fm) 0 1

2 3

4 y (fm)

0

0.04

0.08

0.12

0.16

ψ+ψ

0

0.04

0.08

0.12

0.16

Fig. 12: Space-time profile of a zero mode at 1.5 Tc for atypical gauge configuration with Q = 1.

chirality contained in the two peaks is of opposite sign.These profiles provide a strong hint toward a picturewhere two zero modes with equal and opposite chirali-ties interact weakly, becoming a pair of near-zero modeswhen properly superposed.The fermion zero mode ψ0(x) associated with an in-

stanton is known analytically [17], giving a density ofthe form

ψ†0(x)ψ0(x) =

2ρ2

π2(x2 + ρ2)3, (11)

where ρ is the radius of the instanton. When three of thespacetime coordinates are integrated over, the densityalong the remaining fourth coordinate, say y, becomes

fy(y) =ρ2

2(y2 + ρ2)3/2. (12)

An estimate of the instanton size ρ can be obtained byeither finding the distance where this integrated den-sity falls below 1/

√8 of its maximal value or by fit-

ting Eq. (12) to the measured density. When estimat-ing the size in the temporal direction, the second ap-proach has the advantage that it can also accommodatecases where the size of the instanton ∼ 1/T and the pe-riodic copies in the temporal direction have noticeableoverlap. This is achieved by replacing the fit function by∑n

k=−n fτ (τ + k/T ).

0 1

2 3

4y (fm) 0

1

2

3

4

t (fm)

0

0.02

0.04

0.06

ψ+ψ

0

0.02

0.04

0.06

0 1

2 3

4

x (fm) 0 1

2 3

4 y (fm)

0

0.02

0.04

0.06

ψ+ψ

0

0.02

0.04

0.06

Fig. 13: Space-time profile of a near-zero mode at 1.5 Tc forthe same gauge configuration depicted in Fig. 12.

At 1.5Tc, we measured the wavefunction density of thezero modes along each coordinate direction for all config-urations with |Q| = 1 by summing over the other threedirections. The radii of the profiles along the x, y andz directions, ρx, ρy and ρz, were averaged over to givea spatial radius of ρσ = 0.223(8) fm, essentially indepen-dent of which method was used. Along the temporal di-rection, using the adjusted fit ansatz we obtained a radiusthat was only slightly larger, namely ρτ = 0.24(1) fm.In order to study the distribution of the distance be-

tween the instanton and antiinstanton forming a pair,the lattice points with the lowest and highest chiral den-sity of the corresponding near-zero modes were identi-fied and the distance between them was measured. Theresult is shown in Fig. 15. We compare this with theexpected distribution of the separation if an instantonand an antiinstanton were to be distributed on the lat-tice randomly, independent of each other. It can be seenthat the measured distribution compares quite well withthe random distribution with a slight excess at small sep-arations, further supporting the picture that instantonsand antiinstantons are weakly interacting.In a weakly interacting random ensemble of such topo-

logical objects, the assumption of independent occur-rence results in a Poisson distribution for the total num-ber n of instantons and antiinstantons ,

Pκ(n) = e−κκn/n! , (13)

Page 12: Microscopic Origin of UA QCD - arXiv · 2018-10-02 · arXiv:1502.06190v1 [hep-lat] 22 Feb 2015 Microscopic Origin ofUA(1) Symmetry Violation in theHigh Temperature Phase of QCD Viktor

12

ψ+γ5ψ

0 1

2 3

4y (fm) 0

1

2

3

4

t (fm)

-0.06-0.04-0.02

0 0.02 0.04 0.06

-0.06

-0.04

-0.02

0

0.02

0.04

0.06

0 1

2 3

4

x (fm) 0 1

2 3

4 y (fm)

-0.06

-0.04

-0.02

0

0.02

0.04

0.06

ψ+γ5ψ

-0.06

-0.04

-0.02

0

0.02

0.04

0.06

Fig. 14: Space-time profile of chirality of the same near-zeromode depicted in Fig. 13 at 1.5 Tc.

0

0.02

0.04

0.06

0.08

0.1

0.12

0 5 10 15 20 25 30

0.5fm 1fm 1.5fm 2fm 2.5fm 3fm

d/a

p(d/a) measuredrandom

Fig. 15: Distribution of distances between the instanton andantiinstanton that couple to give rise to a near-zero mode.The blue points show the expected distribution if the instan-ton and antiinstanton were to be distributed randomly andindependently of each other on a 323 × 8 lattice.

where κ is a parameter that is equal to the ensemble av-erage 〈n〉 as well as the variance σ2 and n is obtainedby counting the number of eigenvalues below a cutoffλ0. To fix the cutoff, we fitted Eq. (13) to the distri-bution of n for different cutoffs and compared the χ2

per degree of freedom. The value that was closest to 1,namely 1.03, was obtained for λ0 = 0.44T . The result-ing distribution is shown in Fig. 16 together with the

0

0.05

0.1

0.15

0.2

0.25

0 2 4 6 8 10

n

P (n) λ < 0.44TPoisson fit

Fig. 16: Configuration distribution of the total number n ofzero and near-zero modes at 1.5 Tc and Poisson fit to the data.

Poisson fit which gives κ = 4.5(2). At this cutoff, weobtained 〈n〉 = 4.50(14) and σ2 = 4.2(4) from averagingover the configurations, confirming that instantons andantiinstantons indeed occur almost independently.Taking 4.5(2) as the average total number of instan-

tons and antiinstantons per configuration, we obtain adensity of 0.147(7) fm−4, which is much lower than thevalue predicted from the strongly interacting dense ILM,1 fm−4 [20]. However, since we are at a temperature muchhigher than the chiral crossover temperature Tc, it is notsurprising that we get values expected rather from a di-lute and weakly interacting ensemble.Altogether, the findings described in this section are

giving support to a picture in which for sufficiently hightemperatures of T & 1.5Tc, the infrared behavior of QCDcan be described as that of a dilute gas of instantons andantiinstantons which are interacting weakly with eachother.

G. The localization properties of the eigenmodes

The profiles of the near-zero modes have suggested thatthese are localized structures. To further quantify theirlocalization properties and those of the bulk modes, westudy the so-called participation ratio (PR), defined fora normalized eigenvector ψ(x) of the Dirac operator as

PR =1

N3σNτ

[

x

(

ψ†(x)ψ(x))2

]−1

. (14)

It is the fraction of the total lattice volume occupied bythe eigenmode. If the eigenvector is distributed equallyon the entire four-volume, this quantity is unity.First we use the PR to corroborate our observation

made in Sec. III F. If indeed the near-zero modes repre-sent weakly interacting instanton-antiinstanton pairs, thePR of a typical near-zero mode should be about twice aslarge as that of a zero mode. The comparison of the PRof the near-zero modes and of the zero modes of |Q| = 1configurations is shown in Fig. 17. The ratio of the aver-age PR of a near-zero and that of a zero mode indeed is

Page 13: Microscopic Origin of UA QCD - arXiv · 2018-10-02 · arXiv:1502.06190v1 [hep-lat] 22 Feb 2015 Microscopic Origin ofUA(1) Symmetry Violation in theHigh Temperature Phase of QCD Viktor

13

0

0.01

0.02

0.03

0.04

0.05

0.06

1e-05 0.0001 0.001 0.01 0.1 1

λ/T

PR zero modesnear-zero modes

Fig. 17: Participation ratio of zero modes of |Q| = 1 config-urations and near-zero modes (λ/T < 0.4) as well as theiraverage values as function of the eigenvalue λ at T ∼ 1.5 Tc.

1.85. The PR values of near-zero modes fluctuate aboutthe mean value, so not all of them support this picture,but there is a significant fraction that does.

At 1.2Tc, the PR histograms are measured for two val-ues of the lattice spacing and compiled in Fig. 18. Thelow-lying eigenvalues are more localized than the bulkmodes; however, the PR gradually increases as one goestowards the bulk. At 1.5Tc, the low-lying modes belowλ ≤ 0.4T average to a PR value of about 0.01. As oneenters the bulk eigenvalue region for λ > 0.4T , see forexample Fig. 9, there appears to be a rise of the PR. Thevalue of λ ≈ 0.4T may thus be considered as a mobilityedge separating the localized near-zero eigenstates fromthe delocalized bulk states. The presence of localized aswell as delocalized states is observed in disordered semi-conductors whose dynamics is described by the AndersonHamiltonian. In the Anderson model, the electron statesat the band edge are localized whereas the states at theband center remain delocalized within the lattice. Thecorresponding eigenvalues of the Anderson Hamiltonianchange from Poisson statistics at the band edge to ran-dom matrix theory statistics at the band center.

A comparison with random matrix model predictionscan be achieved by looking at the distribution of the dis-tance s between two consecutive eigenvalues, i.e. the levelspacing distribution. In order to understand its univer-sal properties, it is necessary to map the eigenvalues ontonew values using an unfolding procedure [54], thereby re-moving the non-universal global scale of the system. Byconstruction, the unfolded eigenvalues that are obtainedby this method have a mean level spacing of unity.

While the unfolded level spacings of the localizednear-zero modes should follow a Poisson distribution be-cause of their mostly independent occurrence, the bulkmodes should be strongly mixed. If the bulk is highlydisordered, the corresponding level spacing distributionshould follow the same distribution as the eigenvaluesof a random matrix of an appropriate symmetry group.The Dirac operator for QCD with matter fields in thefundamental representation falls into the same symme-

0

0.05

0.1

0.15

0.2

0.25

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6λ/ms

PR 1.2 TcNτ=6Nτ=8

0

0.05

0.1

0.15

0.2

0.25

0.3

0 0.2 0.4 0.6 0.8 1 1.2 1.4

λ /T

T=1.5 TcPR

Fig. 18: The PR for eigenvectors at 1.2 and 1.5 Tc. The 1.2 Tc

data are for two different lattice spacings to study the cutoffdependence of our results.

try group as Gaussian Unitary ensembles (GUE) whereasfor the case of two colors it is in the same universal-ity class as Gaussian Orthogonal ensembles (GOE). Thelevel spacing distribution for the unfolded bulk eigenval-ues of the overlap operator on HISQ configurations at1.5Tc is shown in Fig. 19. It shows an agreement witha random matrix theory with Gaussian unitary matri-ces. This is in general agreement with similar studiesof the localization of low-lying modes done before forthe quenched theory [19, 56–58] and also with dynamicalstaggered fermions [55, 59, 60] on smaller lattice sizes.

IV. CONCLUSIONS

In this work we have investigated the temperature de-pendence of the anomalous UA(1) symmetry breakingin the high temperature phase of QCD with two lightquark flavors. To this end we have employed the over-lap Dirac operator exploiting its property of preservingthe index theorem even at nonvanishing lattice spacing.We have applied the overlap operator on large volumeHISQ gauge field configurations and computed its low-lying eigenmodes. We observe the presence of zero aswell as near-zero modes in the investigated temperaturerange of Tc . T . 1.5Tc. By comparing the low-lyingeigenmodes from two lattice spacings and studying the

Page 14: Microscopic Origin of UA QCD - arXiv · 2018-10-02 · arXiv:1502.06190v1 [hep-lat] 22 Feb 2015 Microscopic Origin ofUA(1) Symmetry Violation in theHigh Temperature Phase of QCD Viktor

14

0

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2

P(s

)

s

GOEGUE

Fig. 19: The universal level spacing distribution for λ > 0.4 Tat 1.5 Tc compared to the Random Matrix theory GaussianUnitary ensemble (GUE) and the Gaussian Orthogonal en-semble (GOE).

effects of smearing we have shown that these infraredmodes are not mere lattice cutoff effects.We mainly analyzed configurations which have been

obtained at a light sea quark mass corresponding to apion mass of 160MeV. However, within the set of con-figurations at our disposal, at a temperature near Tc wecould confirm the accumulation of the near zero eigenval-ues also at a quark mass considerably below its physicalvalue.By quantifying the contribution of the near-zero eigen-

modes to a specific combination of two point correla-tion functions, χπ − χδ, we conclude that these modesare primarily responsible for the anomalous breakingof the axial symmetry in QCD still being visible forTc . T . 1.5Tc. Through detailed studies of their

spacetime profiles, localization properties and distribu-tions over gauge configurations we have shown that forT ∼ 1.5Tc the near-zero modes follow the behavior asexpected of a gas of widely separated, weakly interactinginstantons and antiinstantons. At 1.5Tc we find the den-sity of (anti)instantons to be 0.147(7) fm−4, with a typi-cal radius of 0.223(8) fm. At this temperature, the spatialvolume of our lattice was∼ (3.3 fm)3 with 1/T ∼ 0.83 fm,suggesting that the instanton gas is indeed dilute, theinstanton size is smaller than 1/T and our chosen vol-ume being large enough to accommodate more that oneinstanton–antiinstanton pair on average. In conclusion,our study suggests that at T ∼ 1.5Tc the origin of globalUA(1) breaking in QCD is due to the dilute gas of weaklyinteracting instantons and antiinstantons.

For an independent confirmation of our results, itwould clearly be desirable to carry out a similar anal-ysis with dynamical chiral fermions. While the latticespacing effects in this work appear to be small, it willfurther be necessary to control the subtle extrapolationsto the continuum as well as the chiral limit in such afuture investigation.

Acknowledgments

This work has been supported in part through con-tract de-sc0012704 with the U.S. Department of Energy,the BMBF under grant 05P12PBCTA, EU under grants238353 and 283286 and the GSI BILAER grant. Numeri-cal calculations have been performed using GPU clustersat Bielefeld University. The GPU codes used in our workwere in part based on some publicly available QUDA li-braries [61]. SS would like to thank Gernot Akemannand Mario Kieburg for discussions and their helpful sug-gestions.

[1] C. Bernard et. al., Phys. Rev. D 71, 034504 (2005).[2] M. Cheng et. al., Phys. Rev. D 74, 054507 (2006).[3] Y. Aoki et. al., Nature 443, 675678 (2006).[4] T. Bhattacharya et. al., Phys. Rev. Lett. 113, 082001

(2014).[5] A. Bazavov et. al., Phys. Rev. D 85, 054503 (2012).[6] S. L. Adler, Phys. Rev. 177, 2426 (1969).

J. Bell and R. Jackiw, Nuovo. Cim. A 60, 47 (1969).[7] K. Fujikawa, Phys. Rev. Lett. 42, 1195 (1979).[8] G. ’t Hooft, Phys. Rev. Lett. 37, 8 (1976).[9] R. D. Pisarski and F. Wilczek, Phys. Rev. D 29, 338

(1984).[10] A. Butti, A. Pelissetto, E. Vicari, JHEP 0308, 029 (2003).[11] A. Pelissetto and E. Vicari, Phys. Rev. D 88, 105018

(2013).[12] M. Grahl and D. H. Rischke, Phys. Rev. D 88, 056014

(2013).[13] E. Shuryak, Comments Nucl. Part. Phys. 21, 235 (1994).[14] T. Banks and A. Casher, Nucl. Phys. B 169, 103 (1980).[15] S. Aoki, H. Fukaya, Y. Taniguchi, Phys. Rev. D 86,

114512 (2012).

[16] A. Bazavov et. al. Phys. Rev. D 86, 094503 (2012).[17] G. t’Hooft, Phys. Rev. D 14, 3432 (1976).[18] D. J. Gross, R. D. Pisarski and L. G. Yaffe, Rev. Mod.

Phys. 53, 43 (1981).[19] R. G. Edwards, U. M. Heller, J. Kiskis and R. Narayanan,

Phys. Rev. D 61, 074504 (2000).[20] E. V. Shuryak, Nucl. Phys. B 203, 93 (1982).[21] E. -M. Ilgenfritz and E. V. Shuryak Nucl. Phys. B 319,

511 (1989).[22] T. Schaefer, E. Shuryak and J. J. M. Verbaarschot, Phys.

Rev. D 51, 1267 (1995).[23] E. -M. Ilgenfritz and E. V. Shuryak, Phys. Lett. B 325,

263 (1994).[24] S. Chandrasekharan and N. H. Christ, Nucl. Phys. Proc.

Suppl. 47, 527 (1996).[25] M. Cheng et. al., Eur. Phys. J. C 71, 1564 (2011).[26] E. Follana et al. (HPQCD Collaboration and UKQCD

Collaboration), Phys. Rev. D 75, 054502 (2007).[27] H. Ohno, U. M. Heller, F. Karsch and S. Mukherjee, PoS

LATTICE 2011, 210 (2011); PoS LATTICE 2012, 095(2012).

Page 15: Microscopic Origin of UA QCD - arXiv · 2018-10-02 · arXiv:1502.06190v1 [hep-lat] 22 Feb 2015 Microscopic Origin ofUA(1) Symmetry Violation in theHigh Temperature Phase of QCD Viktor

15

[28] D. H. Adams, Phys. Rev. Lett. 104, 141602 (2010).[29] M. I. Buchoff et. al., Phys. Rev. D 89, 054514 (2014).[30] T. W. Chiu et. al., PoS LATTICE2013 165 (2014).[31] G. Cossu et. al., Phys. Rev. D 87, 114514 (2013).[32] Ph. de Forcrand, M. Garcia Perez and I.-O. Stamatescu,

Nucl. Phys. Proc. Suppl. 53, 557 (1997).[33] T. DeGrand, A. Hasenfratz, and T. G. Kovacs, Nucl.

Phys. B 520, 301 (1998).[34] A. Hasenfratz and F. Knechtli, Phys. Rev. D 64, 034504

(2002).[35] M. Atiyah and I. Singer, Bull. Amer. Math. Soc, 69 (3)

422 (1963).[36] A. Bazavov et. al. [HotQCD Collaboration], Phys. Rev.

D 90, 094503 (2014).[37] H.-T. Ding et. al., PoS LATTICE 2013, 157 (2014).[38] R. Narayanan and H. Neuberger, Phys. Rev. Lett. 71,

3251 (1993);H. Neuberger, Phys. Lett. B 417, 141 (1998).

[39] S. B. Nielsen and M. Ninomiya, Nucl. Phys. B 185, 20(1981).

[40] P. Hasenfratz, V. Laliena and F. Niedermeyer, Phys.Lett. B 427, 125 (1998).

[41] D. H. Adams, Annals Phys. 296, 131 (2002).[42] R. V. Gavai, S. Gupta and R. Lacaze, Phys. Rev. D 65,

094504 (2002).[43] S. Sharma et. al., PoS LATTICE 2013, 164 (2014).[44] T. Kalkreuter and H. Simma, Comput. Phys. Commun.

93, 33 (1996).[45] A. Hasenfratz, R. Hoffmann, S. Schaefer, JHEP 0711,

071 (2007).[46] A. V. Smilga and J. Stern, Phys. Lett. B 318, 531 (1993).[47] J. J. M. Verbaarschot, Nucl. Phys. B 427, 534 (1994).[48] C. Bernard, C. E. DeTar, Z. Fu and S. Prelovsek, Phys.

Rev. D 76, 094504 (2007).[49] B. J. Harrington and H. K. Shepard, Phys. Rev. D 17,

2122 (1978).[50] T. Kraan and P. van Baal, Nucl. Phys. B 533, 627 (1998).[51] K. -M. Lee and C. -H. Lu, Phys. Rev. D 58, 025011

(1998).[52] P. Rossi, Nucl. Phys. B 149, 170 (1979).[53] V. Bornyakov, E. -M. Ilgenfritz, B. Martemyanov, V.

Mitrjushkin and M. Mueller-Preussker, Phys. Rev. D 87,114508 (2013).

[54] T. Guhr, A. Mueller-Groeling and H. A. Weidenmueller,Phys. Rept. 299, 189 (1998).

[55] A. M. Garcia-Garcia and J. Osborn, Phys. Rev. D 75,034503 (2007).

[56] R. Gavai and S. Gupta, Phys. Rev. D 65, 094504 (2002).[57] T. G. Kovacs, Phys. Rev. Lett. 104, 031601 (2010).[58] T. G. Kovacs and F. Pittler, Phys. Rev. Lett. 105, 192001

(2010).[59] R. Gavai and S. Gupta, Phys. Rev. D 77, 114506 (2008).[60] M. Giordano, T. G. Kovacs, F. Pittler, Phys. Rev. Lett.

112, 102002 (2014).[61] M. A. Clark et. al., Comput. Phys. Commun. 181, 1517

(2010).