Next Article in Journal
Exact Kerr–Schild Spacetimes from Linearized Kinetic Gravity Braiding
Previous Article in Journal
Constraining the Primordial Black Hole Mass Function by the Lensing Events of Fast Radio Bursts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Stability Analysis of Magnetized Quark Matter in Tsallis Statistics

Institute of Theoretical Physics, State Key Laboratory of Quantum Optics and Quantum Optics Devices, Shanxi University, Taiyuan 030006, China
*
Author to whom correspondence should be addressed.
Universe 2025, 11(9), 312; https://doi.org/10.3390/universe11090312
Submission received: 20 July 2025 / Revised: 5 September 2025 / Accepted: 10 September 2025 / Published: 12 September 2025

Abstract

In this work, we employ the nonextensive Nambu–Jona-Lasinio model to analyze the thermodynamic properties of magnetized quark matter. The nonequilibrium state is described in Tsallis distribution by a dimensionless parameter q. We find that within a reasonable temperature range, the system undergoes a crossover transition at the critical chemical potential, which is decreased by the increase of both the temperature and q value. In contrast to the enhanced stability by magnetic field in Boltzmann statistics, it is found that the stability of chiral restored matter in Tsallis statistics would be reduced by an increase of the magnetic field. Conversely, the increase of the q would enhance the stability of quark matter. Finally, we display the different magnetic effects on the stability in the chiral broken and restored regions.

1. Introduction

As a core component of the Standard Model of particle physics, quantum chromodynamics (QCD) not only explains the internal structure of hadrons but also provides a theoretical framework for the evolution of the early Universe and heavy ion collisions [1]. For many years, the theory has been used to predict and has confirmed the state of matter within a few microseconds after the Big Bang in the Relativistic Heavy Ion Collider (RHIC) and the Large Hadron Collider (LHC) experiments [2,3]. That is to say, quarks and gluons are liberated from confinement at high temperatures and high densities, forming quark gluon plasma (QGP) that only exists for a short time and exhibits the properties of a perfect fluid [4]. For the low energy region, in 1974, Kenneth G. Wilson proposed the lattice QCD (LQCD)-simulation method, which is one of the most effective methods in the nonperturbation theory. It would become the continuum field theory in the thermodynamic limit and the continuum limit [5]. At present, the QCD theory is utilized to achieve precise prediction of the matter construction and its progress has a profound impact on particle physics, nuclear physics, cosmology and even quantum information.
The phase diagram and phase transition of QCD are among the most interesting and challenging fields in modern physics. The critical temperature and chemical potential of transitions play a very important role in our understanding of the phase diagram and the theory itself [6,7]. Currently, the HotQCD Collaboration has provided precise pseudocritical temperature of the chiral phase transition by calculating different quantities, such as chiral order parameter and chiral susceptibility [8,9]. When the mass of a quark is infinite, the quark cannot be excited in vacuum. At this time, the phase transition of the system is the confinement–deconfinement transition, which is another independent phase transition in QCD [10]. Recently, the large N c argument indicates that there is an area in the phase diagram called quarkyonic matter, which could be discovered in heavy ion collisions and compact stars [11]. Due to the existence of strong magnetic fields, the thermodynamic properties of quark matter would be influenced, thereby causing the phase structures to become more abundant. The boundary line is confirmed to be changed by the magnetic fields. It is somewhat difficult to find the exact location of the critical end point (CEP), at which phase transitions of strong interacting matter occur. A mainstream view is that the CEP exists and there is only one, but in some of the literature the authors believe that there is no CEP or that there might be two CEPs, etc. [12,13,14,15,16].
Generally speaking, the most reliable source for the equation of state is the lattice gauge theory and its application to QCD, which provides relevant nonperturbation information from first principle calculations [17,18,19]. However, the lattice study at finite baryon chemical potential has been severely hindered due to the sign problem; other approaches such as effective theories need to be used for calculation, which have the properties and symmetries of QCD. In 1961, Yoichiro Nambu and G. Jona-Lasinio proposed the Nambu–Jona-Lasinio (NJL) model to describe the interaction between nucleons [20,21]. Later, the model was also employed to research questions such as the mass and decay constant of scalar mesons, and became a crucial theoretical model [22,23]. Nevertheless, the role of gluons is ignored in general NJL mean fields, resulting in the model’s inability to demonstrate the color confinement of quarks. In order to make up for the shortcoming, the polyakov loop, which is the dynamic gluon field, has been added on this basis [24,25].
In the conventional description of the QCD transition, equilibrium is typically based on Boltzmann–Gibbs (BG) statistics. However, in heavy ion collision experiments, the spatial configuration of the system is far from uniform, which prevents the establishment of global thermal equilibrium. In the development process of thermodynamic statistical models, the introduction of the Tsallis distribution better describes the nonequilibrium statistical phenomena in the high energy region. In 1988, Constantino Tsallis first proposed the concept of nonextensive entropy and formed a generalized distribution [26]. In the practice of the PHENIX and STAR at RHIC in BNL, the Tsallis distribution has developed into more general forms, including the Boltzmann, the Fermi–Dirac and the Bose–Einstein distribution [27,28]. As mentioned earlier, the Tsallis statistic is parameterized by a nonequilibrium factor q, and when q approaches 1, the distribution degenerates into the standard Boltzmann–Gibbs distribution [29]. In recent years, many studies have taken the Tsallis statistic into account to describe high energy physics [30,31,32,33,34]. Here, it should be emphasized that the strong magnetic fields existing in relativistic heavy ion collisions and neutron stars do play an important role in the thermodynamics of the system [35,36].
The paper is organized as follows. In Section 2, the thermodynamics of the nonextensive NJL model is reviewed in a strong magnetic field. In Section 3, the numerical results and discussion concentrate on the stability of quark matter affected by the magnetic field and nonextensive parameters. Finally we present our conclusions in Section 4.

2. Nonextensive NJL Model in the Strong Magnetic Field

We next present the main formulas necessary for the application of the nonextensive statistics to the NJL model. As mentioned above, the Tsallis entropy was proposed [30,37]
S q = k 1 p i q q 1 ,
where k is a conventional positive constant, p i represents the probability related to the event, q can be any real number. The concept is delicate, and the entropy satisfies all properties of Boltzmann–Gibbs entropy, except the additivity, so
S q ( A + B ) = S q ( A ) + S q ( B ) + ( 1 q ) S q ( A ) S q ( B ) ,
where A and B refer to two independent systems. The nonextensive phenomenon is considered to be related to the long-range interacting mechanic [38].
The function exp q ( x ) is usually defined as [26,37,39,40]
exp q ( x ) ( 1 + ( q 1 ) x ) 1 / ( q 1 ) x > 0 ( 1 + ( 1 q ) x ) 1 / ( 1 q ) x 0 ,
where x = ω i μ T is defined, ω i is the energy of the particle given by ω i = p 2 + M 2 and μ is the chemical potential. In the limit q 1 , the standard exponential lim q 1 exp q ( x ) exp ( x ) is recovered. The function of the q-logarithm as well given by
ln q ( x ) x q 1 1 q 1 x > 0 x 1 q 1 1 q x 0 .
By using the above formula, we can obtain the distribution function of fermions at the temperature T
f + ( x ) = 1 1 + ( 1 + ( q 1 ) x ) 1 q 1 x > 0 1 1 + ( 1 + ( 1 q ) x ) 1 1 q x 0 ,
where x = ω i μ T , ω i = p z 2 + M 2 + 2 k i | q i | B gives the effective energy and q i represents the charge of quark in strong magnetic fields [41,42].
In order to model dynamical chiral symmetry breaking, the Lagrangian density of the two-flavor and three-color NJL model in the presence of strong external magnetic fields is [43]
L N J L = ψ ¯ ( i / D m ) ψ + G [ ( ψ ¯ ψ ) 2 + ( ψ ¯ i γ 5 τ ψ ) 2 ] ,
where ψ represents the quark fields and m is the bare mass of fermions. The coupling of quarks to electromagnetic fields is given by / D = γ μ D μ and the covariant derivative D μ = μ + i Q A μ , which implies the sum of the flavor and color degrees of freedom. G is the four-fermion interaction coupling constant and τ is isospin Pauli matrice.
We all know that the thermodynamic potential is a function of natural variables and all other thermodynamic quantities are differentiated from it. In the finite temperature field theory and mean-field approximation, it can be expressed as
Ω = ( M m i ) 2 4 G + i = u , d Ω i ,
where the first term is the interaction term, and Ω i is defined as Ω i = Ω i vac + Ω i mag + Ω i med in the second term. The vacuum, the magnetic field, and the medium contributions at finite temperature and chemical potential are
Ω i vac   = N c 8 π 2 M 4 ln ( Λ + ϵ Λ M ) ϵ Λ Λ ( Λ 2 + ϵ Λ 2 ) ,
Ω i mag = N c ( | q i | B ) 2 2 π 2 ζ ( 1 , x i ) 1 2 ( x i 2 x i ) ln ( x i ) + x i 2 4 ,
Ω i med = k i = 0 a k i | q i | B N c T 4 π 2 d p ln q ( 1 + e ( ω i μ ) / T ) + ln q ( 1 + e ( ω i + μ ) / T ) = k i = 0 a k i | q i | B N c T 4 π 2 d p [ 1 + ( 1 φ + ) 1 1 q ] q 1 1 q 1 + [ 1 + ( 1 φ ) 1 1 q ] q 1 1 q 1 x > 0 k i = 0 a k i | q i | B N c T 4 π 2 d p [ 1 + ( 1 + φ + ) 1 q 1 ] q 1 1 q 1 + [ 1 + ( 1 + φ ) 1 q 1 ] q 1 1 q 1 x 0 ,
where the quantity ϵ Λ is defined as ϵ Λ = Λ 2 + M 2 . The ζ ( a , x ) = n = 0 1 ( a + n ) x is the Riemann–Hurwitz zeta function. The ζ ( 1 , x ) = d ζ ( z , x ) d z | z = 1 is defined in the magnetic field term [44]. Then, a k i = 2 δ k i 0 and k i are the degeneracy label and the Landau quantum number, respectively. The function φ ± in the medium term is expressed as φ ± = ( q 1 ) ( ω i μ ) T as a matter of convenience.
In the interaction term, M refers to the effective dynamical mass of the quark. The condensation that can be regarded as the order parameter is formed due to the pairing of quarks–antiquarks with the same chirality in the mean-field approximation [45]. So we have the gap equation
M i = m i 2 G ψ ¯ ψ ,
where the quark condensates include u and d quark contributions as ψ ¯ ψ = ϕ = i = u , d ϕ i . Here we use M = M u M d as the representation of the current mass. Similarly, the contributions from the quark condensates with flavor i include the vacuum, the magnetic field, and the medium term as [46,47,48]
ϕ i = ϕ i vac + ϕ i mag + ϕ i med ,
where each term reads
ϕ i vac   = M N c 2 π 2 Λ Λ 2 + M 2 M 2 ln ( Λ + Λ 2 + M 2 M ) ,
ϕ i mag = M | q i | B N c 2 π 2 ln [ Γ ( x i ) ] 1 2 ln ( 2 π ) + x i 1 2 ( 2 x i 1 ) ln ( x i ) ,
ϕ i med = k i = 0 a k i M | q i | B N c 4 π 2 f + ( ω i ) + f ( ω i ) ω i d p .
In order to obtain a good approximation, we choose the dimensionless quantity x i , defined as x i = M 2 / ( 2 | q i | B ) in the renormalized expression. In Equation (15), a k i = 2 δ k i 0 and k i are the degeneracy label and the Landau quantum number, respectively. In nonzero magnetic fields, the concepts of Landau quantization and magnetic catalysis, where the magnetic field is assumed to be oriented along z direction, should be implemented. So we can use the mapping d 3 p ( 2 π ) 3 | q i | B 2 π k d p z 2 π ( 2 δ k i 0 ) to compute the integral over the three momenta [49,50]. Here the distribution function of the system is in the form of nonextensive statistics because the global equilibrium is difficult to establish. It should be noted that the NJL model, as a non-renormalization model, needs to be regularized to remove the divergence of the vacuum. In the literature, there are many works that have discussed the renormalization of the free energy of charged fermions in a magnetic field. The appropriate scheme was proposed to regularize the integral. Schwinger’s proper time formalism regularized all integrals without separating the divergent (vacuum) piece from the convergent (thermomagnetic) contribution. The magnetic field independent regularization (MFIR) makes a full separation of the finite magnetic contributions and the divergencies.
Generally, the pressure can be obtained by P = Ω . In this context, we take care of the normalized pressure by subtracting vacuum pressure contribution at zero chemical potential from the total pressure given as
P eff = Ω P ( μ = 0 , T = 0 ) .
In terms of thermodynamic relations of the system, the free energy density reads
F = P eff + μ i n i ,
where n i is the number number of particles, it can be obtained through the relation
n i = Ω i μ | T .
The detailed formulas are as follows
n i = k i = 0 a k i | q i | B N c 4 π 2 d p [ 1 + ( 1 φ + ) 1 1 q ] q 2 ( 1 φ + ) q 1 q [ 1 + ( 1 φ ) 1 1 q ] q 2 ( 1 φ ) q 1 q x > 0 k i = 0 a k i | q i | B N c 4 π 2 d p [ 1 + ( 1 + φ + ) 1 q 1 ] q 2 ( 1 + φ + ) 2 q q 1 [ 1 + ( 1 + φ ) 1 q 1 ] q 2 ( 1 + φ ) 2 q q 1 x 0 .

3. Numerical Result and Discussion

In the framework of the SU(2) NJL model, the following parameters are adopted: m u = m d = 5.5 MeV for the up and down quarks, Λ = 587.9 MeV and G = 2.44 / Λ 2 for the momentum cutoff and the coupling constant [22]. It is worth noting that the maximum deviation of q should not exceed 20% [28,51,52,53,54]. Following this, we chose three different q values in the following Tsallis statistic for comparison.
In the literature, the dynamical mass of quarks serves as an order parameter to reflect the chiral crossover in the NJL model. Figure 1 depicts the dynamical mass of u and d quarks in the chiral phase transition at the magnetic field e B = 0.2 GeV2 and the nonextensive parameter q = 1.1 . Here we have chosen a typical value of the magnetic field, which is helpful for understanding the magnetic effects in heavy ion collisions and neutron stars [55]. Three different temperatures are denoted by black, red and blue curves, respectively. The dotted line corresponds to the dynamical mass of nonmagnetized quarks. It is obvious that the magnetic field and the temperature play competing roles to influence the quark mass at μ = 0 . For given finite temperatures, the curves exhibit a decline to lower values at high chemical potentials, indicating that the quarks undergo a chiral restoration. As the temperature increases, the crossover occurs at lower chemical potential, which means that high temperatures favor the chiral restoration at low chemical potentials. Moreover, compared with the zero magnetic field, a strong magnetic field leads to an increase of the critical chemical potential.
In a nonequilibrium system, the influence of the parameter q on quark matter is beyond doubt. At a fixed temperature and magnetic field, the variation of quark mass with chemical potential induced by q is shown in Figure 2. On the one hand, within a specific range, the increase in the q value reduces the critical chemical potential, which is determined by the peak of the derivative of the dynamical mass with respect to the chemical potential in the phase transition. On the other hand, the presence of strong magnetic fields shifts the critical chemical potential to higher values, reflecting the magnetic catalytic effect observed in the lattice QCD. Consequently, it can be concluded that the parameter q and the strong magnetic field exhibit opposite effects on the change of the critical chemical potential in nonextensive case.
In Figure 3, the normalized pressure at the finite chemical potential was calculated for different temperatures. The pressure is shown as a monotonically increasing function of quark chemical potential for all given temperatures. Furthermore, the lower the temperature, the smaller the effective pressure as expected. In addition, there are intersection points between the solid and dashed lines in each group. This implies that the effective pressure of the magnetized system exceeds that in the absence of a magnetic field at high chemical potentials. This similar phenomenon occurs for three sets of q values in Figure 4. The positions of the intersection points vary slightly by changing the q value. Roughly speaking, the magnetized pressure is larger than the nonmagnetized one in the chiral restoration. In the chiral broken phase at low chemical potential, one is in the opposite situation.
The free energy per baryon is defined according to the partition function, providing a good theoretical perspective for understanding thermodynamic stability. The lower free energy of quark matter corresponds to its more stable state. The free energy per baryon is shown at the temperature T = 0.08, 0.12, 0.16 GeV in Figure 5. With the increase of quark chemical potential, the free energy per baryon monotonically grows. On the contrary, the free energy per baryon is lower when the temperature rises. The zero magnetic field and strong magnetic field are marked by the dotted and solid lines, respectively. The two lines intersect at a moderate chemical potential value of up to 0.4 GeV. This indicates that the stability on both sides of the crossing point is determined by the presence or absence of the magnetic field. The inflection corresponds to the critical chemical potential at which the phase transition occurs. At high temperature T = 0.16 GeV, the system remains in the chiral restoration phase regardless of the magnitude of the chemical potential. Consequently, no inflection behavior is observed along the blue line. The magnetic field has weak influence on the free energy at low temperatures. At high temperatures, it is observed that the free energy without a magnetic field is lower than that of magnetized matter.
Similarly, as depicted in Figure 6, we investigated the influence of the nonextensive parameter q on the free energy per baryon. For a given q value, the free energy exhibits monotonic growth with the chemical potential, and the larger the q value is, the chemical potential of the inflection point shifts towards a lower value. In our previous calculation, we reached the conclusion that the existence of a strong magnetic field could potentially reduce the stability of quark matter in the lower chemical potential region. However, the increase of the parameter q contributes to enhancing its stability. The intertwining of the three groups of curves in the figure can be attributed to the fact that, within the phase transition region, the free energy of quark matter with a higher q value is the greatest.
Figure 7 explains the impact of different magnetic fields on the chiral phase transition and thermodynamic properties at finite temperature T = 0.12 GeV. In the left panel, the smooth change of the quark mass still exhibits the crossover in QCD community. The increase in the magnetic field leads to an increase in the pseudocritical chemical potential of the crossover, that is, the magnetic catalysis effect in the density direction. In the middle panel, the pressure function is slightly decreased by an increase of the magnetic field in the low chemical potential region. Inversely in the larger chemical potential region, the stronger magnetic field would result in larger pressure. Finally, the trend of free energy per baryon is displayed on the right panel. Similar to the behavior of pressure, the free energy per baryon exhibits different magnetic field effects at lower and higher chemical potential. It can be concluded that an increased magnetic field is favorable to the stability of chiral broken matter. Conversely, when the density increases, a strong magnetic field would reduce the stability of the chiral restored matter. The inflections on the curve correspond to the pseudocritical chemical potential of the phase transition in the left panel.

4. Summary

In this paper, we have investigated the chiral transition and the stability of quark matter in the nonextensive NJL model. Here, we focus on the influence of the finite chemical potential, nonextensive parameter q and magnetic fields on quark matter. At the fixed temperature, the chiral transition takes place at a critical chemical potential, which is decreased by an increase of the nonextensive parameter q and increased by an increase of the magnetic field. The stability is represented by the free energy per baryon. The increased temperature can result in lower free energy and therefore enhance the stability of the chiral broken matter. On the contrary, the chiral restored phase would have larger free energy at high temperature. Correspondingly, it is shown that the magnetic field also has different effects on the stability of quark matter. In particular, the stability of the chiral broken matter is enhanced by an increase of the magnetic field. Conversely, the chiral restored phase would have a larger free energy per baryon and result in a decrease in stability.
For heavy ion collision experiments, researchers investigate potential quark matter by generating extremely high temperature and high density conditions. The stability of matter directly influences the existence of new states of matter, such as quark gluon plasma, and affects the interpretation of experimental data, thereby verifying the theory of QCD phase diagrams, particularly in identifying their critical end points and phase boundaries. On the other hand, as stellar objects partially or entirely composed of strange quark matter (SQM), neutron stars may provide extreme environments that enable quark matter to exist stably [56]. According to the literature, if the strange quark matter is more stable than ordinary nuclear matter at high density, the conversion from two flavor to SQM occurs via a decay of the down quark to the strange quark ( u + d s + u ) [57]. The study of the stability of matter is directly related to characteristics of neutron stars, including the core structure, cooling mechanism and gravitational wave signals. Therefore, we hope that our conclusion will contribute to the investigation of nonequilibrium quark matter and serve as a bridge between experiments and astrophysical observations.

Author Contributions

Conceptualization, J.Z. and X.-J.W.; software, J.Z.; validation, X.-J.W.; formal analysis, J.Z. and X.-J.W.; writing—original draft preparation, J.Z.; writing—review and editing, J.Z.; supervision, X.-J.W.; funding acquisition, X.-J.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Natural Science Foundation of China for support through Grants No.11875181, No.12047571, and No.11705163. This work was also sponsored by the Fund for Shanxi “1331 Project” Key Subjects Construction.

Acknowledgments

The authors thank the referees and editors for their valuable suggestions.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Shuryak, E.V. Quantum Chromodynamics and the Theory of Superdense Matter. Phys. Rep. 1980, 61, 71–158. [Google Scholar] [CrossRef]
  2. Abelev, B.I.; Adams, J.; Aggarwal, M.M.; Ahammed, Z.; Amoett, J.; Anderson, S.D.; Aderson, M.; Arkhipkin, D.; Averichev, G.S.; Averichev, G.S.; et al. [STAR] Strange particle production in p + p collisions at s**(1/2) = 200-GeV. Phys. Rev. C 2007, 75, 064901. [Google Scholar] [CrossRef]
  3. Aamodt, K.; Abel, N.; Abeysekara, U.; Quintana, A.A.; Abramyan, A.; Adamová, D.; Aggarwal, M.M.; Aglieri Rinella, G.; Agocs, A.G.; Aguilar Salazar, S.; et al. [ALICE] Transverse momentum spectra of charged particles in proton-proton collisions at s = 900 GeV with ALICE at the LHC. Phys. Lett. B 2010, 693, 53–68. [Google Scholar] [CrossRef]
  4. Tannenbaum, M.J. Recent results in relativistic heavy ion collisions: From ‘a new state of matter’ to ‘the perfect fluid’. Rep. Prog. Phys. 2006, 69, 2005–2060. [Google Scholar] [CrossRef]
  5. Wilson, K.G. Confinement of Quarks. Phys. Rev. D 1974, 10, 2445–2459. [Google Scholar] [CrossRef]
  6. Asakawa, M.; Yazaki, K. Chiral Restoration at Finite Density and Temperature. Nucl. Phys. A 1989, 504, 668–684. [Google Scholar] [CrossRef]
  7. Costa, P.; Ruivo, M.C.; de Sousa, C.A. Thermodynamics and critical behavior in the Nambu-Jona-Lasinio model of QCD. Phys. Rev. D 2008, 77, 096001. [Google Scholar] [CrossRef]
  8. Steinbrecher, P. [HotQCD] The QCD crossover at zero and non-zero baryon densities from Lattice QCD. Nucl. Phys. A 2019, 982, 847–850. [Google Scholar] [CrossRef]
  9. Bazavov, A.; Ding, H.T.; Hegde, P.; Kaczmarek, O.; Karsch, F.; Karthik, N.; Laermann, E.; Lahiri, A.; Larsen, R.; Li, S.T.; et al. [HotQCD] Chiral crossover in QCD at zero and non-zero chemical potentials. Phys. Lett. B 2019, 795, 15–21. [Google Scholar] [CrossRef]
  10. Satz, H. Color deconfinement in nuclear collisions. Rep. Prog. Phys. 2000, 63, 1511. [Google Scholar] [CrossRef]
  11. McLerran, L.; Pisarski, R.D. Phases of cold, dense quarks at large N(c). Nucl. Phys. A 2007, 796, 83–100. [Google Scholar] [CrossRef]
  12. Kitazawa, M.; Koide, T.; Kunihiro, T.; Nemoto, Y. Chiral and color superconducting phase transitions with vector interaction in a simple model. Prog. Theor. Phys. 2002, 108, 929–951, Erratum in Prog. Theor. Phys. 2003, 110, 185–186. [Google Scholar] [CrossRef]
  13. Blaschke, D.; Volkov, M.K.; Yudichev, V.L. Coexistence of color superconductivity and chiral symmetry breaking within the NJL model. Eur. Phys. J. A 2003, 17, 103–110. [Google Scholar] [CrossRef]
  14. Hatsuda, T.; Tachibana, M.; Yamamoto, N.; Baym, G. New critical point induced by the axial anomaly in dense QCD. Phys. Rev. Lett. 2006, 97, 122001. [Google Scholar] [CrossRef] [PubMed]
  15. Carignano, S.; Nickel, D.; Buballa, M. Influence of vector interaction and Polyakov loop dynamics on inhomogeneous chiral symmetry breaking phases. Phys. Rev. D 2010, 82, 054009. [Google Scholar] [CrossRef]
  16. Bratovic, N.M.; Hatsuda, T.; Weise, W. Role of Vector Interaction and Axial Anomaly in the PNJL Modeling of the QCD Phase Diagram. Phys. Lett. B 2013, 719, 131–135. [Google Scholar] [CrossRef]
  17. Philipsen, O. The QCD equation of state from the lattice. Prog. Part. Nucl. Phys. 2013, 70, 55–107. [Google Scholar] [CrossRef]
  18. Ukawa, A. Kenneth Wilson and lattice QCD. J. Statist. Phys. 2015, 160, 1081. [Google Scholar] [CrossRef]
  19. Meyer, H.B. QCD at non-zero temperature from the lattice. In Proceedings of the 33rd International Symposium on Lattice Field Theory (LATTICE 2015), Kobe, Japan, 14–18 July 2015. [Google Scholar] [CrossRef]
  20. Nambu, Y.; Jona-Lasinio, G. Dynamical Model of Elementary Particles Based on an Analogy with Superconductivity. I. Phys. Rev. 1961, 122, 345–358. [Google Scholar] [CrossRef]
  21. Nambu, Y.; Jona-Lasinio, G. Dynamical model of elementary particles based on an analogy with superconductivity. II. Phys. Rev. 1961, 124, 246–254. [Google Scholar] [CrossRef]
  22. Hatsuda, T.; Kunihiro, T. QCD phenomenology based on a chiral effective Lagrangian. Phys. Rep. 1994, 247, 221–367. [Google Scholar] [CrossRef]
  23. Buballa, M. NJL model analysis of quark matter at large density. Phys. Rep. 2005, 407, 205–376. [Google Scholar] [CrossRef]
  24. Fukushima, K. Phase diagrams in the three-flavor Nambu-Jona-Lasinio model with the Polyakov loop. Phys. Rev. D 2008, 77, 114028, Erratum in Phys. Rev. D 2008, 78, 039902. [Google Scholar] [CrossRef]
  25. Lo, P.M. Polyakov loop susceptibilities in pure gauge system. J. Phys. Conf. Ser. 2014, 503, 012034. [Google Scholar] [CrossRef]
  26. Tsallis, C. Possible Generalization of Boltzmann-Gibbs Statistics. J. Statist. Phys. 1988, 52, 479–487. [Google Scholar] [CrossRef]
  27. Adare, A.; Afanasiev, S.; Aidala, C.; Ajitanand, N.N.; Akiba, Y.; Al-Bataineh, H.; Alexander, J.; Aoki, K.; Aphecetche, L.; Armendariz, R.; et al. [PHENIX] Identified charged hadron production in p + p collisions at s = 200 and 62.4 GeV. Phys. Rev. C 2011, 83, 064903. [Google Scholar] [CrossRef]
  28. Cleymans, J.; Worku, D. Relativistic Thermodynamics: Transverse Momentum Distributions in High-Energy Physics. Eur. Phys. J. A 2012, 48, 160. [Google Scholar] [CrossRef]
  29. Wilk, G.; Wlodarczyk, Z. On the interpretation of nonextensive parameter q in Tsallis statistics and Levy distributions. Phys. Rev. Lett. 2000, 84, 2770. [Google Scholar] [CrossRef]
  30. Wilk, G.; Wlodarczyk, Z. Power laws in elementary and heavy-ion collisions: A Story of fluctuations and nonextensivity? Eur. Phys. J. A 2009, 40, 299–312. [Google Scholar] [CrossRef]
  31. Marques, L.; Andrade-II, E.; Deppman, A. Nonextensivity of hadronic systems. Phys. Rev. D 2013, 87, 114022. [Google Scholar] [CrossRef]
  32. Wong, C.Y.; Wilk, G. Tsallis fits to pT spectra and multiple hard scattering in pp collisions at the LHC. Phys. Rev. D 2013, 87, 114007. [Google Scholar] [CrossRef]
  33. Marques, L.; Cleymans, J.; Deppman, A. Description of High-Energy pp Collisions Using Tsallis Thermodynamics: Transverse Momentum and Rapidity Distributions. Phys. Rev. D 2015, 91, 054025. [Google Scholar] [CrossRef]
  34. Deppman, A. Thermodynamics with fractal structure, Tsallis statistics and hadrons. Phys. Rev. D 2016, 93, 054001. [Google Scholar] [CrossRef]
  35. Kharzeev, D.E.; McLerran, L.D.; Warringa, H.J. The Effects of topological charge change in heavy ion collisions: ‘Event by event P and CP violation’. Nucl. Phys. A 2008, 803, 227–253. [Google Scholar] [CrossRef]
  36. Kadam, G.; Pal, S.; Bhattacharyya, A. Interacting hadron resonance gas model in magnetic field and the fluctuations of conserved charges. J. Phys. G 2020, 47, 125106. [Google Scholar] [CrossRef]
  37. Tsallis, C.; Mendes, R.S.; Plastino, A.R. The Role of constraints within generalized nonextensive statistics. Physica A 1998, 261, 534. [Google Scholar] [CrossRef]
  38. Tsallis, C. Introduction to Nonextensive Statistical Mechanics: Approaching a Complex World; Springer: New York, NY, USA, 2009; ISBN 978-0-387-85358-1/978-0-387-85359-8. [Google Scholar] [CrossRef]
  39. Cleymans, J.; Worku, D. The Tsallis Distribution in Proton-Proton Collisions at s = 0.9 TeV at the LHC. J. Phys. G 2012, 39, 025006. [Google Scholar] [CrossRef]
  40. Azmi, M.D.; Cleymans, J. The Tsallis Distribution at Large Transverse Momenta. Eur. Phys. J. C 2015, 75, 430. [Google Scholar] [CrossRef]
  41. Endrödi, G. QCD equation of state at nonzero magnetic fields in the Hadron Resonance Gas model. J. High Energy Phys. 2013, 2013, 23. [Google Scholar] [CrossRef]
  42. Bali, G.S.; Bruckmann, F.; Endrödi, G.; Katz, S.D.; Schäfer, A. The QCD equation of state in background magnetic fields. J. High Energy Phys. 2014, 2014, 177. [Google Scholar] [CrossRef]
  43. Menezes, D.P.; Benghi Pinto, M.; Avancini, S.S.; Perez Martinez, A.; Providencia, C. Quark matter under strong magnetic fields in the Nambu-Jona-Lasinio Model. Phys. Rev. C 2009, 79, 035807. [Google Scholar] [CrossRef]
  44. Boomsma, J.K.; Boer, D. The Influence of strong magnetic fields and instantons on the phase structure of the two-flavor NJL model. Phys. Rev. D 2010, 81, 074005. [Google Scholar] [CrossRef]
  45. Buballa, M.; Oertel, M. Strange quark matter with dynamically generated quark masses. Phys. Lett. B 1999, 457, 261–267. [Google Scholar] [CrossRef]
  46. Bali, G.S.; Bruckmann, F.; Endrodi, G.; Fodor, Z.; Katz, S.D.; Schafer, A. QCD quark condensate in external magnetic fields. Phys. Rev. D 2012, 86, 071502. [Google Scholar] [CrossRef]
  47. Avancini, S.S.; Menezes, D.P.; Providencia, C. Finite temperature quark matter under strong magnetic fields. Phys. Rev. C 2011, 83, 065805. [Google Scholar] [CrossRef]
  48. Rożynek, J.; Wilk, G. Nonextensive Nambu-Jona-Lasinio Model of QCD matter. Eur. Phys. J. A 2016, 52, 13, Erratum in Eur. Phys. J. A 2016, 52, 204. [Google Scholar] [CrossRef]
  49. Chakrabarty, S. Quark matter in strong magnetic field. Phys. Rev. D 1996, 54, 1306–1316. [Google Scholar] [CrossRef]
  50. Fraga, E.S.; Mizher, A.J. Chiral transition in a strong magnetic background. Phys. Rev. D 2008, 78, 025016. [Google Scholar] [CrossRef]
  51. Beck, C. Nonextensive statistical mechanics and particle spectra in elementary interactions. Physica A 2000, 286, 164–180. [Google Scholar] [CrossRef]
  52. Cleymans, J.; Hamar, G.; Levai, P.; Wheaton, S. Near-thermal equilibrium with Tsallis distributions in heavy ion collisions. J. Phys. G 2009, 36, 064018. [Google Scholar] [CrossRef]
  53. Beck, C. Superstatistics in high energy physics: Application to cosmic ray energy spectra and e+e annihilation. Eur. Phys. J. A 2009, 40, 267–273. [Google Scholar] [CrossRef]
  54. Aamodt, K.; Abel, N.; Abeysekara, U.; Abrahantes, A.Q.; Abramyan, A.; Adamova, D.; Aggarwal, M.M.; Aglieri Rinella, G.; Agocs, A.G.; Aguilar Salazar, S.; et al. [ALICE] Production of pions, kaons and protons in pp collisions at s = 900 GeV with ALICE at the LHC. Eur. Phys. J. C 2011, 71, 1655. [Google Scholar] [CrossRef]
  55. Tawfik, A.N.; Diab, A.M. Chiral magnetic properties of QCD phase-diagram. Eur. Phys. J. A 2021, 57, 200. [Google Scholar] [CrossRef]
  56. Benvenuto, O.G. White dwarf stars as strange quark matter detectors. J. Phys. G 2005, 31, L13–L17. [Google Scholar] [CrossRef]
  57. Bhattacharyya, A.; Ghosh, S.K.; Joardar, P.S.; Mallick, R.; Raha, S. The conversion of Neutron star to Strange star: A two step process. Phys. Rev. C 2006, 74, 065804. [Google Scholar] [CrossRef]
Figure 1. The dynamical mass of nonextensive quark matter as a function of chemical potential at different temperatures.
Figure 1. The dynamical mass of nonextensive quark matter as a function of chemical potential at different temperatures.
Universe 11 00312 g001
Figure 2. The dynamical mass of quark matter as a function of chemical potential for different nonextensive parameters at T = 0.12 GeV.
Figure 2. The dynamical mass of quark matter as a function of chemical potential for different nonextensive parameters at T = 0.12 GeV.
Universe 11 00312 g002
Figure 3. The behavior of the pressure for q = 1.1 as a function of chemical potential at different temperatures.
Figure 3. The behavior of the pressure for q = 1.1 as a function of chemical potential at different temperatures.
Universe 11 00312 g003
Figure 4. The behavior of the pressure for three sets of q value as a function of chemical potential at T = 0.12 GeV.
Figure 4. The behavior of the pressure for three sets of q value as a function of chemical potential at T = 0.12 GeV.
Universe 11 00312 g004
Figure 5. The variation of the free energy per baryon as a function of chemical potential at different temperatures.
Figure 5. The variation of the free energy per baryon as a function of chemical potential at different temperatures.
Universe 11 00312 g005
Figure 6. The variation of the free energy per baryon as a function of chemical potential for different nonextensive parameters.
Figure 6. The variation of the free energy per baryon as a function of chemical potential for different nonextensive parameters.
Universe 11 00312 g006
Figure 7. The thermodynamic quantities of nonextensive quark matter at finite temperatures are shown at different magnetic fields, from left to right as the dynamical mass, the normalized pressure and the free energy per baryon.
Figure 7. The thermodynamic quantities of nonextensive quark matter at finite temperatures are shown at different magnetic fields, from left to right as the dynamical mass, the normalized pressure and the free energy per baryon.
Universe 11 00312 g007
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, J.; Wen, X.-J. Stability Analysis of Magnetized Quark Matter in Tsallis Statistics. Universe 2025, 11, 312. https://doi.org/10.3390/universe11090312

AMA Style

Zhang J, Wen X-J. Stability Analysis of Magnetized Quark Matter in Tsallis Statistics. Universe. 2025; 11(9):312. https://doi.org/10.3390/universe11090312

Chicago/Turabian Style

Zhang, Jia, and Xin-Jian Wen. 2025. "Stability Analysis of Magnetized Quark Matter in Tsallis Statistics" Universe 11, no. 9: 312. https://doi.org/10.3390/universe11090312

APA Style

Zhang, J., & Wen, X.-J. (2025). Stability Analysis of Magnetized Quark Matter in Tsallis Statistics. Universe, 11(9), 312. https://doi.org/10.3390/universe11090312

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop