Topological susceptibility and axion properties in the presence of a strong magnetic field within the three-flavor NJL model
Abstract
We analyze the topological susceptibility and the axion properties in the presence of an external uniform magnetic field, considering a three flavor NJL model that includes strong CP violation through a ’t Hooft-like flavor mixing term. Both thermal and finite density effects are studied for magnetic fields up to 1 GeV2, and the corresponding phase transitions are analyzed. To capture the inverse magnetic catalysis effect at finite temperatures and densities, a magnetic field-dependent coupling constant is considered. Our analytical and numerical results are compared with those previously obtained from lattice QCD, chiral perturbation theory and other effective models.
I Introduction
As well known, Quantum Chromodynamics (QCD) contains gauge field configurations that carry topological charge [1]. These configurations interpolate between different vacua of the gluonic sector, giving rise to the so-called “ vacuum” [2]. At the level of the QCD Lagrangian, this nontrivial nature of the QCD vacuum implies the presence of a so-called -term of the form
(1) |
where is the strong coupling constant, and , stand for the gluon field tensor and its dual. In fact, in the context of the full Standard Model, the coefficient can be modified through a chiral rotation. Considering the weak interaction sector, by diagonalizing the Yukawa-generated quark mass matrix one has
(2) |
The parameter (which can be taken to run from 0 to , due to the periodicity of the action) can be in principle experimentally determined. Presently, measurements of the neutron electric dipole moment lead to the constraint . Since a nonzero value of would imply the existence of CP violation in the strong interaction sector, the measurement of such an unexpectedly low upper bound for the value of is known as the “strong CP problem”.
As proposed almost 50 years ago by Peccei and Quinn, a possible solution for this puzzle can be achieved by invoking the existence of an additional global U(1) chiral symmetry [3, 4]. The spontaneous breakdown of this symmetry leads to the presence of an associated Goldstone boson, the axion, which implies an extra anomalous contribution to the Lagrangian. Thus, the resulting effective potential gets minimized for a nonzero vacuum expectation value of the axion field, leading finally to a cancellation of the term and solving the strong CP problem. In practice, the above described mechanism can be basically implemented by promoting the parameter to a dynamical field , normalized by a dimensionful “decay constant” . The effective potential will be given by a periodic function of , becoming minimized for .
After the original introduction of the axion as a possible solution of the strong CP problem, its relevance has also been discussed in other contexts (for a recent review see e.g. Ref. [5]). For example, it has been proposed a candidate for cold dark matter. In this sense, the temperature dependence of the axion mass would play an important role in the estimation of its cosmic abundance [6, 7, 8]. Axions might also play a key role in astrophysics, in particular for the anomalous cooling of neutron stars (see e.g. Ref. [9, 10]). In this framework, the influence of nonzero density and background magnetic fields on the axion properties deserve particular interest.
The strength of the topological charge fluctuations in the QCD vacuum is quantified by the topological susceptibility , which is defined as a second derivative of the QCD partition function with respect to and is shown to be proportional to the axion mass . Different approaches to non-perturbative QCD have been used to determine this quantity. For example, within the framework of Nambu-Jona-Lasinio (NJL) type models, estimates of the topological susceptibility and its temperature dependence have been studied in Refs. [11, 12, 13]. On the other hand, lattice QCD (LQCD) results for at zero and finite temperature can be found in Refs. [14, 15, 16].
More recently, considerable attention has been paid to the modifications of and axion properties induced by the presence of strong magnetic fields. Besides its importance within the above mentioned astrophysical context, it has been realized that the effect of magnetic fields on the topological structure of the QCD vacuum can be significant for the study of heavy ion collisions. In fact, the existence of a chirality imbalance induced by topology, in the presence of strong magnetic fields produced in a noncentral heavy ion collision, can lead to the so-called chiral magnetic effect, according to which positive and negative charges get separated along the magnetic field direction [17, 18]. As discussed in Ref. [19], it is important to take into account the effects of both the temperature and the magnetic field on QCD vacuum fluctuations. These effects have been analyzed using Chiral Perturbation Theory (ChPT) [20, 21, 22], which should be adequate for relatively low values of the magnetic field. In addition, results have been obtained using two-flavor versions of the local [23] and nonlocal NJL model [24]. One of the aims of the present work is to extend those studies by considering a three-flavor version of the local NJL model. It is worth pointing out that in the SU(3) chiral NJL model the coupling that controls the flavor mixing effects related with the U(1)A anomaly can be better determined, taking into account the phenomenological values of meson masses. Moreover, three-flavor models allow for a more consistent comparison with very recent results obtained using LQCD simulations with 2+1 flavors [25]. Another important purpose of our work is to complement previous studies by analyzing nonzero density effects which, as already mentioned, might become relevant when axions are considered in astrophysical contexts.
This article is organized as follows. Following Ref. [26], in Sec. II.A we review the general formalism corresponding to a three-flavor NJL model at finite temperature and chemical potential in the presence of a constant magnetic field, including the mentioned field. Then, in Sec. II.B, we derive the expressions needed to obtain the quantities of interest related with the topological susceptibility and the axion properties. Numerical results for these quantities are discussed in Sec. III. Finally, in Sec. IV we summarize our results and present our main conclusions.
II Theoretical formalism
II.1 Effective NJL lagrangian and mean field equations at zero temperature
We consider a three-flavor NJL Lagrangian that includes scalar and pseudoscalar chiral quark couplings as well as a ’t Hooft six-fermion interaction term, in the presence of an external electromagnetic field . Moreover, we take into account the coupling to a field of the form given by Eq. (1). This can be effectively done through the ’t Hooft term (which accounts for the chiral U(1)A anomaly) by performing a chiral rotation of quarks fields by an angle . In this way, the effective Euclidean action is given by [27, 26]
(3) |
where and are coupling constants, stands for a quark three-flavor vector, is the corresponding current quark mass matrix, and . In addition, denote the Gell-Mann matrices, with , where is the unit matrix in the three flavor space. The coupling of quarks to the electromagnetic field is implemented through the covariant derivative , where represents the quark electric charge matrix, i.e. , being the proton electric charge. In the present work we consider a static and constant magnetic field in the -direction.
We proceed by bosonizing the action in terms of scalar and pseudoscalar fields, introducing also corresponding auxiliary and fields. Following a standard procedure, we start from the partition function
(4) |
By introducing functional delta functions, the scalar () and pseudoscalar () currents in can be replaced by and , and one can perform the functional integration on the fermionic fields and . Then, to carry out the integration over the auxiliary fields we use the stationary phase approximation (SPA), keeping the functions and that minimize the integrand of the partition function. This yields a set of coupled equations among the bosonic fields, from which one can take and to be implicit functions of and . Finally, we consider the mean field (MF) approximation, expanding the bosonized action in powers of field fluctuations around the corresponding translationally invariant mean field values and . Thus, we write and , where, due to charge conservation, only , with can be nonzero. For convenience, we introduce the notation and .
At the mean field level, the Euclidean action per unit volume reads
(5) | |||||
where stands for trace in Dirac space, while
(6) |
is the inverse mean field quark propagator for each flavor. Here, we have used the definitions and . Moreover, in Eq. (5) are the values of the auxiliary fields at the mean field level within the SPA approximation, i.e. . They satisfy the conditions
(7) |
where the values for indices , and are equivalent to labels . From the conditions and one can get now the “gap equations”
(8) |
where the quark-antiquark condensates are given by
(9) |
As is well known, the quark propagator can be written in different ways. For convenience we use [28, 29]
(10) |
where is the so-called Schwinger phase, which depends on the gauge choice, and vanishes for . We express in the Landau level form
(11) | |||||
Here, are generalized Laguerre polynomials, with the convention . We have also introduced the definitions and , together with . “Perpendicular” and “parallel” gamma matrices have been collected into vectors and , and, in the same way, we have defined vectors and . Note that in our convention .
The resulting expressions for and are divergent and have to be properly regularized. We use here the magnetic field independent regularization (MFIR) scheme, in which one subtracts from the unregulated integral the limit and then adds it in a regulated form. In this way, we obtain
(12) |
where
(13) |
For the quantity we use here a 3D cutoff regularization. Thus, we have
(14) |
where . On the other hand, the “magnetic piece” is found to be given by
(15) |
where .
The associated Euclidean regularized action per unit volume is given by
(16) | |||||
where
(17) |
with
(18) | |||||
(19) |
Here stands for the derivative of the Hurwitz zeta function.
Let us now consider the case of a system in equilibrium at a finite temperature and quark chemical potential . We follow here a similar analysis as the one carried out in Ref. [26]. The expressions for the mean field values and can be written as in Eqs. (12), replacing the function by
(20) |
where
(21) |
with , .
II.2 Mean field topological susceptibility, axion mass and axion self-coupling
As stated, the topological susceptibility is useful to analyze how the manifestations of the chiral anomaly are affected by the temperature and the external magnetic field. This quantity is given by
(24) |
where is the topological charge
(25) |
In the framework of the above introduced effective NJL model, can be simply calculated by taking the second derivative of the effective action with respect to , viz.
(26) |
The evaluation of the first derivative of with respect to can be obtained taking into account the SPA equations
(27) |
together with the MF conditions
(28) |
Then, from Eq. (5) one has
(29) |
Using Eqs. (7), and noting that (see Eqs. (12)), one gets
(30) |
where are the current quark masses. Moreover, it can be shown that the terms in the above sum over flavors are equal to each other. Indeed, from Eqs. (7) one has
(31) | |||||
where the last expression does not depend on . To derive this relation we have used the property
(32) |
Thus, we have
(33) |
where can be either , or .
Notice that if any of the current quark masses (e.g. ) is taken to be equal to zero, one immediately obtains , i.e., the Lagrangian in Eq. (3) becomes independent of . This is due to the existence of an additional U(1) global symmetry. In this limit the value of becomes unobservable, and, as is well known, the strong CP problem vanishes.
From the above equations it is easy to see that, as required by the Peccei-Quinn mechanism, the minimization condition leads to the mean field values , . On the other hand, at the mean field level, the second derivative of the action with respect to is nothing but the axion mass squared. Thus, the topological susceptibility and the axion mass are simply related by
(34) |
To evaluate the second derivative in the above expression we need to determine . We have
(35) |
The expressions for and can be readily obtained from Eqs. (12) and the subsequent expressions of the functions for zero and nonzero temperature. The partial derivatives of and with respect to can be calculated by solving the coupled equations
(36) |
where we have introduced the shorthand notation
(37) |
In the limit the equations get simplified, leading to the result
(38) |
We recall that is equal to twice the scalar condensate .
The expression in Eq. (38) can be compared to previous results obtained in the approximate chiral limit, where current masses are taken to be relatively small. At the lowest order in one has
(39) |
Moreover, assuming one can approximate
(40) |
where . This in agreement with the expressions found in Refs. [30, 22] and [31] in the contexts of ChPT and a linear sigma model, respectively. In the limit , the above expressions for reduce to the lowest-order ChPT Leutwyler-Smilga relation [32]
(41) |
Finally, it is also interesting to study the axion self-coupling arising from the -dependent effective potential. It is usual to focus on the (i.e. ) term in the effective action, defining the coupling parameter as
(42) |
III Results
III.1 Model parameters
Before presenting the numerical results for the topological susceptibility, we introduce the parameterization that has been chosen for the above discussed three-flavor version of the NJL model. Following Refs. [33, 34], we adopt the parameter set MeV, MeV, MeV, and . This set has been determined in such a way that for one obtains meson masses MeV, MeV and MeV, together with a pion decay constant MeV.
It is well known that local NJL-like models fail to reproduce the inverse magnetic catalysis (IMC) effect at finite temperature. To address this issue, the possibility of allowing the coupling constant to depend on the magnetic field has been considered [35, 36, 37]. Taking into account the analysis carried out in Ref. [35], we consider a functional form
(43) |
where the parameters , , and are determined by fitting the dependence of the pseudocritical chiral transition temperatures to those obtained through LQCD calculations [38]. The values of the parameters are [35] , , and , with MeV. The effect on the pseudocritical chiral restoration temperatures (at zero baryon chemical potential) is shown in Fig. 1, where we plot —normalized to — as a function of the magnetic field, considering the case of a constant coupling and the case in which one has a -dependent coupling as in Eq. (43) [35]. For comparison, the dependence of the normalized pseudocritical temperatures obtained from Lattice QCD calculations is also shown (gray band in the figure) [38]. The normalization temperature in our model is found to be MeV, somewhat larger than the critical value obtained from LQCD calculations, MeV [39].

III.2 Zero chemical potential
Let us start by quoting our numerical results for both zero quark chemical potential and zero temperature. In Fig. 2 we show the values of the topological susceptibility (upper panel) and the axion self-coupling parameter (lower panel) as functions of the magnetic field, normalized by the corresponding values at , namely MeV and GeV. Black dashed and solid lines correspond to and , respectively. It can be seen that in both cases and show an enhancement with the magnetic field. Within errors, our results for the topological susceptibility are shown to be in agreement with those obtained from LQCD calculations (blue squares) [25] for a temperature MeV (which is well below the critical temperature, therefore the value of should be rather close to the one at ). In addition, we include for comparison the curves for corresponding to a two-flavor NJL model, taken from Ref. [23], both for constant and -dependent couplings (red dashed and solid lines, respectively). The above mentioned value for at obtained within our model can be compared with the results obtained from ChPT [40] and LQCD [14] analyses, which lead to MeV. In the case of the axion self-coupling, from ChPT the estimation GeV is obtained [41].

The behavior of the above quantities for nonzero temperatures is shown in Fig. 3, where we show the numerical results for and as functions of for three representative values of the magnetic field. The pseudocritical chiral transition temperatures have been defined taking the maximum values of the slopes , where , for each value of the magnetic field. Left and right panels correspond to the results for constant and -dependent couplings, respectively. As expected, one finds a sudden drop of both and at , signalling the restoration of chiral symmetry in the light quark sector. Notice that the curves for tend to show a peak located at .


In Fig. 4 we compare the results for obtained within our model, Eq. (38) (solid lines) with those arising from the approximate expressions in Eqs. (39) and (40) (dashed and dotted lines, respectively). The curves correspond to the case , for , 0.5 GeV2 and 1 GeV2. It can be seen that up to all expressions are approximately equivalent for the considered values of the magnetic field. Beyond the chiral restoration transition, the curves corresponding to the approximate expressions in Eqs. (39) and (40) show some deviation with respect to the full result in Eq. (38). This difference is mainly due to the fact that the first term into the brackets, on the right hand side of Eq. (38), becomes nonnegligible in this region.

Our numerical results for the temperature dependence of can also be compared with those recently obtained from LQCD calculations, see Ref. [25]. To perform the comparison we consider the normalized quantity introduced in that work. In Fig. 5 we show our results (black curves) together with those quoted in Ref. [25] (shaded bands) for and 0.8 GeV2. We include in the figure just the curves that correspond to the case of the -dependent coupling , which, as stated, is the one consistent with LQCD results for IMC. Once again, it is seen that the predictions of the NJL model show qualitative agreement with LQCD calculations.

In Fig. 6 we go back to our results for and as functions of the temperature, this time normalizing to the corresponding value at and taking the magnetic field values and GeV2, in order to compare our results (black solid lines) with those obtained within the two-flavor NJL model studied in Ref. [23] (red dashed lines). As in Fig. 5, our results correspond to , given by Eq. (43). The values of the temperature have been normalized to the critical temperatures , which are somewhat different for both models. This is in part due to the fact that in Ref. [23] an explicit dependence on both and has been assumed for the coupling . It is seen that for the two-flavor model the peak of at is slightly higher, while the fall of both and for is less pronounced than in the case of the three-flavor NJL model. Nonetheless, it could be said that the behavior of and is found to be qualitatively similar for both models.


III.3 Finite chemical potential
We turn now to discuss our numerical results for systems at nonzero quark chemical potential . For a better comprehension, we start by briefly reviewing the phase diagrams in the plane, shown in Fig. 7. As expected, for low temperatures the system undergoes a first order chiral restoration transition at given critical chemical potentials . In the figure we show the corresponding transition lines (solid lines in the figure) for some representative values of the magnetic field. These first order transition lines finish at some critical end points (CEPs), whose positions depend on the external magnetic field. For higher values of the temperature, the transitions turn into smooth crossovers (dashed lines in the figure), as discussed for the systems at .
Left and right panels of Fig. 7 correspond to constant and -dependent , respectively. As stated, the assumption of a dependence as the one in Eq. (43) leads to IMC, implying a significant change in the phase diagrams with respect to those obtained for the case. On the other hand, the behavior of the CEP with the magnetic field is known to be strongly model-dependent. It is seen that our results for the case are found to be qualitatively similar to those obtained in Ref. [42], where a three-flavor Polyakov-NJL model with a -dependent coupling is studied; however, the behavior of the CEP is shown to be different from the one found in Ref. [43], where a nonlocal NJL-like model is considered (notice that in this type of model IMC is naturally obtained [44, 45]).


In Fig. 8 we show the behavior of as a function of the temperature, taking some representative values of the chemical potential and the magnetic field. The curves show clearly the first order and crossover transitions, both for the cases of (left panels) and (right panels). We have also verified that for nonzero the expressions in Eqs. (39) and (40) still approximate with good accuracy (%) the exact result in Eq. (38), for temperatures that lie below the chiral transition. Then, in Fig. 9 we show the behavior of the axion self-coupling (times the dimensionful scale ), for the same values of and . It is seen that the peak in —which at was found to occur at the pseudocritical temperature— goes to infinity when the transition reaches the CEP, and turns into a first order transition jump beyond the CEP. This shows that at the transition the coupling behaves like the derivative of an order parameter, and the position of the peak could be used to define the pseudocritical transition temperature related with the restoration of the U(1)A symmetry.




To conclude this section, we discuss the numerical results obtained for zero temperature and finite quark chemical potential. In Fig. 10 we show the behavior of the critical chemical potential as a function of the magnetic field, both for constant and -dependent couplings. The values are normalized to . It is interesting to notice that for GeV2 the models show again an IMC-like effect, i.e., a decrease of the critical chemical potential for increasing values of the magnetic field. This effect has been observed in various effective approaches to low energy QCD, including local and nonlocal NJL-like models [46, 47, 48, 49, 50]. For it is seen that the IMC extends to larger values of the magnetic field, while for constant the values of reach a minimum and then get increased. This growth is consistent with the results in Ref. [48], where a constant value of is assumed, while the persistent IMC behavior is similar to the one obtained in nonlocal NJL models [50], where nonlocality leads to an effective dependence of the couplings on the magnetic field [45].

In Fig. 11 we show the behavior of and as functions of the chemical potential, for three representative values of . Left and right panels correspond to constant and -dependent couplings, respectively. The first order chiral transitions can be clearly observed. Moreover, it can be seen that beyond these transitions there is a second discontinuity, which corresponds to the partial chiral symmetry restoration related to the quark-antiquark condensate. Finally, in Fig. 12 we show the behavior of and as functions of the magnetic field, for some selected values of . Here we use a logarithmic scale, in order to focus on the region of low , and we just include the results for (the curves for are similar for values of up to about 0.3 GeV2). As one can see from Fig. 10, for low values of the system lies in the chiral symmetry broken phase for all considered values of . Then, for MeV a first order transition is found at some intermediate value of , while for values of beyond MeV the system lies in the partially restored chiral symmetry phase. Typically, in this region one finds for a series of magnetic oscillations related to the van Alphen-de Haas effect [51]. In fact, as shown in the figure, one finds a sequence of first order transitions that correspond to the values of that satisfy the relation () for integer . The value of is found to be relatively large for right above at low values of the magnetic field, and becomes significantly reduced when is increased up to GeV2.



IV Summary and conclusions
In the present work we have analyzed the topological susceptibility and the axion properties in the presence of a strong magnetic field, considering a three flavor NJL model that includes strong CP violation through a ’t Hooft-like flavor mixing term. The behavior of the relevant quantities for systems at finite temperature and quark chemical potential have been studied.
As is well known, when the scalar/pseudoscalar coupling is kept constant (i.e., when it does not depend on the magnetic field) local NJL models are not able to reproduce the inverse magnetic catalysis effect at finite temperature. Therefore, we have considered both the case of a constant and the one in which one assumes a dependent coupling , chosen in such a way that the model is able to adequately reproduce the dependence of the critical chiral transition temperatures obtained in lattice QCD.
We have shown that within these three flavor NJL models the topological susceptibility has a rather simple expression (see Eq. (38)) in terms of the quark condensates, the current quark masses and the strength of the flavor mixing term. In addition, we have shown that close to the chiral limit this expression reduces to the one obtained in other approaches to non-perturbative QCD such as chiral perturbation theory [30, 22] and the linear sigma model [31].
At , using standard values of model parameters we obtain, for vanishing external magnetic field, MeV and GeV, in reasonable agreement with values obtained from LQCD and/or ChPT [52]. For nonzero magnetic field, in agreement with previous analyses we find that the topological susceptibility gets increased with . Clearly, this can be understood by noticing that, according to the previously mentioned theoretical expressions, is approximately proportional to the quark condensates, which exhibit the well known magnetic catalysis effect. Moreover, we find that the axion self-coupling also increases with the magnetic field.
For the case of both nonzero temperature and magnetic field, we find that, as expected, and remain approximately constant as functions of , up to the critical temperatures . Beyond these values we find for both quantities a sudden drop, signalling the restoration of the U(1)A symmetry in the light quark sector. We also note that the curves for tend to show a peak located at . When comparing our results with those of the SU(2) NJL model analyzed in Ref. [23], it is seen that for the two-flavor model the peak of at is slightly higher, while the fall of both and observed for is less pronounced than in the case of the three-flavor model. In any case, it could be said that the behavior of and is found to be qualitatively similar for both models. Regarding the comparison with finite temperature LQCD, our predictions for in the case of the -dependent coupling show qualitative agreement with those obtained from LQCD calculations.
As stated, we have completed our analysis by considering systems at nonzero quark chemical potential. Curves showing the behavior of and as functions of the temperature are given for various values of the chemical potential, showing both crossover and first order transitions. It is found that the approximate expressions for in Eqs. (39) and (40) work very well (within %) in the chirally broken phase (i.e., at low temperatures and/or chemical potentials), whereas they show some deviation in the restored phase. In the case of crossover transitions, it is seen that behaves like the derivative of an order parameter; thus, the position of the corresponding peak can be used to define the pseudocritical transition temperature associated with the restoration of symmetry. At zero temperature and finite , it is seen that for the critical chemical potentials exhibit inverse magnetic catalysis, showing a similar behavior as the one observed in nonlocal NJL-like models. In the chirally restored phase, for and low magnetic fields we observe a pattern of magnetic oscillations in the values of both and , related to the well-known van Alphen-de Haas effect.
In general, it is seen that the case in which , which (by construction) exhibits inverse magnetic catalysis at finite temperature, provides the best agreement with results from LQCD at zero chemical potential and yields a phase diagram that is consistent with the expected phenomenology at finite chemical potential. We find that the behavior of and is predominantly determined by the light quark condensates, i.e., by the chiral SU(2) symmetry restoration, and shows reasonable agreement with other effective models, including approximate results from chiral perturbation theory.
Acknowledgements
This work has been supported in part by Consejo Nacional de Investigaciones Científicas y Técnicas and Agencia Nacional de Promoción Científica y Tecnológica (Argentina), under Grants No. PIP2022-GI-11220210100150CO, No. PICT20-01847, No. PICT22-03-00799, and by the National University of La Plata (Argentina), Project No. X960.
References
- [1] A. A. Belavin, A. M. Polyakov, A. S. Schwartz and Y. S. Tyupkin, Phys. Lett. B 59, 85-87 (1975).
- [2] C. G. Callan, Jr., R. F. Dashen and D. J. Gross, Phys. Lett. B 63, 334-340 (1976).
- [3] R. D. Peccei and H. R. Quinn, Phys. Rev. Lett. 38, 1440-1443 (1977).
- [4] R. D. Peccei, Lect. Notes Phys. 741, 3-17 (2008) [arXiv:hep-ph/0607268 [hep-ph]].
- [5] L. Di Luzio, M. Giannotti, E. Nardi and L. Visinelli, Phys. Rept. 870, 1-117 (2020) [arXiv:2003.01100 [hep-ph]].
- [6] J. Preskill, M. B. Wise and F. Wilczek, Phys. Lett. B 120, 127-132 (1983).
- [7] L. F. Abbott and P. Sikivie, Phys. Lett. B 120, 133-136 (1983).
- [8] M. Dine and W. Fischler, Phys. Lett. B 120, 137-141 (1983).
- [9] G. G. Raffelt, Lect. Notes Phys. 741, 51-71 (2008) [arXiv:hep-ph/0611350 [hep-ph]].
- [10] A. Sedrakian, Phys. Rev. D 93, no.6, 065044 (2016) [arXiv:1512.07828 [astro-ph.HE]].
- [11] K. Fukushima, K. Ohnishi and K. Ohta, Phys. Rev. C 63, 045203 (2001) [arXiv:nucl-th/0101062 [nucl-th]].
- [12] P. Costa, M. C. Ruivo, C. A. de Sousa, H. Hansen and W. M. Alberico, Phys. Rev. D 79, 116003 (2009) [arXiv:0807.2134 [hep-ph]].
- [13] G. A. Contrera, D. G. Dumm and N. N. Scoccola, Phys. Rev. D 81, 054005 (2010) [arXiv:0911.3848 [hep-ph]].
- [14] S. Borsanyi, Z. Fodor, J. Guenther, K. H. Kampert, S. D. Katz, T. Kawanai, T. G. Kovacs, S. W. Mages, A. Pasztor and F. Pittler, et al. Nature 539, no.7627, 69-71 (2016) [arXiv:1606.07494 [hep-lat]].
- [15] P. Petreczky, H. P. Schadler and S. Sharma, Phys. Lett. B 762, 498-505 (2016) [arXiv:1606.03145 [hep-lat]].
- [16] Y. Taniguchi, K. Kanaya, H. Suzuki and T. Umeda, Phys. Rev. D 95, no.5, 054502 (2017) [arXiv:1611.02411 [hep-lat]].
- [17] K. Fukushima, D. E. Kharzeev and H. J. Warringa, Phys. Rev. D 78, 074033 (2008) [arXiv:0808.3382 [hep-ph]].
- [18] D. E. Kharzeev, J. Liao and P. Tribedy, [arXiv:2405.05427 [nucl-th]].
- [19] M. Asakawa, A. Majumder and B. Muller, Phys. Rev. C 81, 064912 (2010) [arXiv:1003.2436 [hep-ph]].
- [20] P. Adhikari, Phys. Lett. B 825, 136826 (2022) [arXiv:2103.05048 [hep-ph]].
- [21] P. Adhikari, Nucl. Phys. B 974, 115627 (2022) [arXiv:2111.06196 [hep-ph]].
- [22] P. Adhikari, Nucl. Phys. B 982, 115853 (2022) [arXiv:2203.00200 [hep-ph]].
- [23] A. Bandyopadhyay, R. L. S. Farias, B. S. Lopes and R. O. Ramos, Phys. Rev. D 100, no.7, 076021 (2019) arXiv:1906.09250 [hep-ph]].
- [24] M. S. Ali, C. A. Islam and R. Sharma, Phys. Rev. D 104, no.11, 114026 (2021) [arXiv:2009.13563 [hep-ph]].
- [25] B. B. Brandt, G. Endrődi, J. J. H. Hernández and G. Markó, [arXiv:2409.00796 [hep-lat]].
- [26] B. Chatterjee, H. Mishra and A. Mishra, Phys. Rev. D 91, no.3, 034031 (2015) [arXiv:1409.3454 [hep-ph]].
- [27] J. K. Boomsma and D. Boer, Phys. Rev. D 80, 034019 (2009) [arXiv:0905.4660 [hep-ph]].
- [28] V. A. Miransky and I. A. Shovkovy, Phys. Rept. 576, 1-209 (2015) [arXiv:1503.00732 [hep-ph]].
- [29] J. O. Andersen, W. R. Naylor and A. Tranberg, Rev. Mod. Phys. 88, 025001 (2016) [arXiv:1411.7176 [hep-ph]].
- [30] Y. Y. Mao et al. [TWQCD], Phys. Rev. D 80, 034502 (2009) [arXiv:0903.2146 [hep-lat]].
- [31] M. Kawaguchi, S. Matsuzaki and A. Tomiya, Phys. Rev. D 103, no.5, 054034 (2021) [arXiv:2005.07003 [hep-ph]].
- [32] H. Leutwyler and A. V. Smilga, Phys. Rev. D 46, 5607-5632 (1992).
- [33] P. Rehberg, S. P. Klevansky and J. Hufner, Phys. Rev. C 53 (1996), 410-429 [arXiv:hep-ph/9506436 [hep-ph]].
- [34] M. Coppola, W. R. Tavares, S. S. Avancini, J. C. Sodré and N. N. Scoccola, [arXiv:2410.05568 [hep-ph]].
- [35] M. Ferreira, P. Costa, O. Lourenço, T. Frederico and C. Providência, Phys. Rev. D 89 (2014) no.11, 116011 [arXiv:1404.5577 [hep-ph]].
- [36] R. L. S. Farias, K. P. Gomes, G. I. Krein and M. B. Pinto, Phys. Rev. C 90 (2014) no.2, 025203 [arXiv:1404.3931 [hep-ph]].
- [37] R. L. S. Farias, V. S. Timoteo, S. S. Avancini, M. B. Pinto and G. Krein, Eur. Phys. J. A 53 (2017) no.5, 101 [arXiv:1603.03847 [hep-ph]].
- [38] G. S. Bali, F. Bruckmann, G. Endrodi, Z. Fodor, S. D. Katz, S. Krieg, A. Schafer and K. K. Szabo, JHEP 02 (2012), 044 [arXiv:1111.4956 [hep-lat]].
- [39] A. Bazavov et al. [HotQCD], Phys. Lett. B 795, 15-21 (2019) [arXiv:1812.08235 [hep-lat]].
- [40] M. Gorghetto and G. Villadoro, JHEP 03, 033 (2019) [arXiv:1812.01008 [hep-ph]].
- [41] G. Grilli di Cortona, E. Hardy, J. Pardo Vega and G. Villadoro, JHEP 01, 034 (2016) [arXiv:1511.02867 [hep-ph]].
- [42] P. Costa, M. Ferreira, D. P. Menezes, J. Moreira and C. Providência, Phys. Rev. D 92, no.3, 036012 (2015) [arXiv:1508.07870 [hep-ph]].
- [43] J. P. Carlomagno, S. A. Ferraris, D. Gomez Dumm and A. G. Grunfeld, Phys. Rev. D 108, no.5, 056029 (2023) [arXiv:2305.15540 [hep-ph]].
- [44] V. P. Pagura, D. Gomez Dumm, S. Noguera and N. N. Scoccola, Phys. Rev. D 95, no.3, 034013 (2017) [arXiv:1609.02025 [hep-ph]].
- [45] D. Gómez Dumm, M. F. Izzo Villafañe, S. Noguera, V. P. Pagura and N. N. Scoccola, Phys. Rev. D 96, no.11, 114012 (2017) [arXiv:1709.04742 [hep-ph]].
- [46] F. Preis, A. Rebhan and A. Schmitt, JHEP 03, 033 (2011) [arXiv:1012.4785 [hep-th]].
- [47] F. Preis, A. Rebhan and A. Schmitt, Lect. Notes Phys. 871, 51-86 (2013) [arXiv:1208.0536 [hep-ph]].
- [48] P. G. Allen and N. N. Scoccola, Phys. Rev. D 88, 094005 (2013) [arXiv:1309.2258 [hep-ph]].
- [49] P. G. Allen, A. G. Grunfeld and N. N. Scoccola, Phys. Rev. D 92, no.7, 074041 (2015) [arXiv:1508.04724 [hep-ph]].
- [50] S. A. Ferraris, D. G. Dumm, A. G. Grunfeld and N. N. Scoccola, Eur. Phys. J. A 57, no.4, 141 (2021) [arXiv:2103.00982 [hep-ph]].
- [51] D. Ebert, K. G. Klimenko, M. A. Vdovichenko and A. S. Vshivtsev, Phys. Rev. D 61, 025005 (2000) [arXiv:hep-ph/9905253 [hep-ph]].
- [52] Z. Y. Lu, M. L. Du, F. K. Guo, U. G. Meißner and T. Vonk, JHEP 05 (2020), 001 doi:10.1007/JHEP05(2020)001 [arXiv:2003.01625 [hep-ph]].