-
Relativistic heavy-ion collision experiments offer a distinct opportunity to reproduce the extreme state of matter that existed at the universe's beginning, known as quark-gluon plasma (QGP). Experiments conducted at facilities like the Large Hadron Collider (LHC) at the European Organization for Nuclear Research (CERN) [1−3] and the Relativistic Heavy Ion Collider (RHIC) at Brookhaven National Laboratory (BNL) [4−7] have provided substantial evidence for the creation of QGP, which has significantly advanced the exploration and understanding of the strongly interacting matter properties under extreme temperature conditions. Additionally, the Beam Energy Scan (BES) initiative at RHIC [8, 9], along with current research programs at the Facility for Antiproton and Ion Research (FAIR) [10, 11] and the Nuclotron-based Ion Collider Facility (NICA) [12, 13], are working to explore the characteristics of baryon-rich nuclear matter.
The role of transport coefficients is crucial for describing the evolution of the bulk matter created in relativistic heavy-ion collisions. These coefficients provide insights into how much a system deviates from ideal hydrodynamics and reveal important information about fluid dynamics and critical phenomena [14−18]. Extracting these coefficients accurately from experimental data and evaluating them using various theoretical approaches is currently a prominent area of research. The shear viscosity (η) of hot quantum chromodynamics (QCD) matter has garnered significant interest, primarily because of the surprisingly small value of the ratio of shear viscosity to entropy density,
$ \eta/s_q = 1/4\pi $ , which may resemble a nearly perfect fluid [19]. This finding has led to speculation about the existence of a strongly interacting quark-gluon plasma (sQGP). However, the bulk viscosity to entropy density ratio,$ \zeta_b/s_q $ , has been suggested to increase near the critical temperature$ T_c $ [20−22]. The decreasing value of$ \eta/s_q $ and increasing value of$ \zeta_b/s_q $ near$ T_c $ was found to be consistent with the lattice calculations [23, 24]. The electrical conductivity$ \sigma_{el} $ is important for explaining the enhancement of low-mass dimuons [25] and serves as an essential input for magnetohydrodynamic simulations [26, 27]. Another key transport coefficient for hydrodynamic evolution at finite chemical potential is thermal conductivity κ, which was explored in Refs. [28−31]. These collective findings emphasize that transport coefficients play an important role in measuring the properties of strongly interacting relativistic QCD matter and understanding its phase transitions [15].In recent years, transport coefficients have been extensively studied using a variety of effective QCD [32−37] and hadronic models [38−45]. However, these models employ a Boltzmann-Gibbs (BG) statistics-based approach, which is valid only for systems with strong dynamical correlations, a homogeneous and infinite heat bath, long-range interactions, and microscopic memory effects [46−50]. However, in the initial stages of heavy ion collision experiments, these conditions are rarely met. Hence, some quantities develop power-law-tailed distributions and become nonextensive. To address these issues, Tsallis proposed nonextensive statistics as a generalization of the BG statistics by introducing a dimensionless parameter q to account for all potential variables that violate the assumptions of the standard BG statistics [46]. In this framework, Tsallis proposed a generalized non-additive entropy
$ s_q = \frac{1-\sum_{i = 1}^{w}P_i^q}{q-1}, $
(1) where w is the number of microstates,
$ P_i $ is the probability distribution with$ \sum_{i = 1}^{w}P_i = 1 $ , and q is a positive real number called the nonextensive q-parameter. Assuming equiprobability ($ P_i = 1/w $ ), the Tsallis entropy in Eq. (1) reduces to [46, 51]$ s_q = \frac{w^{1-q}-1}{1-q} = \ln_qw, $
(2) where the q-logarithm is defined as [52]
$ \ln_{q}(x) \equiv \frac{x^{1-q}-1}{1-q}, $
(3) with corresponding q-exponential expressed as
$ \exp_{q}(x) = [1+(1-q)x]^{1/(1-q)}. $
(4) The non-additivity of the entropy
$ s_q $ follows from the non-additivity of the q-logarithm [46]; considering two independent systems A and B with$ P_{(A+B)} = P_{A}P_{B} $ , the generalized entropy of the system takes the form$ s_q(A+B) = s_q(A)+s_q(B)+(1-q)s_q(A)s_q(B), $
(5) where
$ |1-q| $ quantifies the degree of non-equilibration, i.e., how far the system is away from equilibrium. For$ q>1 $ , it describes intrinsic fluctuations of temperature in the system [53, 54]. In Ref. [55], it was observed that temperature fluctuations are measured by the divergence of q from unity, while the Boltzmann limit (q = 1) does not show any temperature fluctuation. Thus, utilizing the nonextensive Tsallis statistics within the dynamical model is highly advantageous for examining the transport coefficients, as they are not well-defined during the initial stage of heavy-ion collisions when the system is in non-equilibrium. In the present work, we attempt to see possible deviations from standard BG statistics for the values of$ q>1 $ , as these values have been seen in numerous phenomenological investigations of high-energy heavy ion collisions [55−57]. In the limit$ q\rightarrow1 $ , the nonextensive entropy reduces to the usual BG entropy, i.e.,$ s_{q = 1} = s_{\rm BG} $ . The q-parameter is incorporated into the specific dynamical formulas of the model and enables a straightforward phenomenological test against possible deviations from the BG framework.Tsallis nonextensive statistics has become increasingly important in recent years due to its ability to fit transverse momentum distributions across a broad range of collision energies, as demonstrated by the STAR [58], PHENIX [59], ALICE [60], and CMS [61] collaborations. In light of this, nonextensivity has been incorporated into many theoretical models to study the properties of QCD matter. These include a generalized Quantum Hydrodynamics (
$ q- $ QHD) model [62], nonextensive version of the Nambu-Jona-Lasinio ($ q- $ NJL) model [63], nonextensive version of the MIT bag ($ q- $ MIT) model [64], generalized linear sigma model ($ q- $ LSM) [65], and nonextensive Polyakov chiral$S U$ (3) quark mean field ($ q- $ PCQMF) model [66]. Recently, the$ q- $ PNJL model was employed to study the transport coefficients [67], critical exponents [68], and fluctuations in the baryon number [69]. Nonextensivity has also been incorporated within the relaxation time approximation of kinetic theory to study the viscous [70] and conductive coefficients [71] of hot and dense magnetized QCD matter. Furthermore, the bulk properties of protoneutron stars [72] and the thermodynamics of a black hole [73] have also been explored using nonextensive statistics.In the present study, we aim to utilize the q-PCQMF to investigate the transport coefficients of strongly interacting QCD matter at finite temperatures and chemical potentials. We study the temperature variations of the transport coefficients using the expressions obtained from kinetic theory and relaxation time approximation at zero and finite values of chemical potentials. Additionally, we include the presence of quark back reaction by replacing the usual Polyakov loop potential with the QCD glue potential [74, 75]. This is done by substituting the pure gauge temperature
$ T_{\rm YM} $ with the glue potential temperature$ T_{\rm glue} $ . The remainder of this paper is organized as follows: In Sec. II, we give a brief introduction of the nonextensive version of the q-PCQMF model. Sec. III discusses the expressions of the transport coefficients. The impact of the q-parameter on the transport coefficients of strongly interacting QCD matter is discussed in Sec. IV. Finally, a brief summary and our conclusions are presented in Sec. V. -
The thermodynamic potential density of the q-extended Polyakov chiral
$S U$ (3) quark mean field model in the mean field approximation is defined as [66]$ \begin{aligned}[b]\Omega_{q} =\;& \mathcal{U}(\Phi,\bar{\Phi},T) - {\cal{L}}_M- {\cal{V}}_{vac} \\&+ \sum\limits_{i = u,d,s}\frac{-\gamma_i k_BT}{(2\pi)^3}\int_0^\infty {\rm d}^3k\left\{ {\rm ln_{q}} F_{q}^-+ {\rm ln_{q}} F^+_{q}\right\}, \end{aligned}$
(6) where
$ {\cal{L}}_M = {\cal{L}}_{\Sigma\Sigma} +{\cal{L}}_{VV} +{\cal{L}}_{SB} $ is the meson interaction term. In this model, the attractive part of the interactions between quarks is represented by scalar meson fields$ \sigma, \zeta, $ and δ, while the repulsive part is represented by vector fields$ \rho, \omega $ , and ϕ. The Polyakov loop fields Φ and$ \bar{\Phi} $ are included in the model to study deconfinement phase transition. Additionally, the model incorporates broken scale invariance by introducing a scalar dilaton field χ [76−78]. For the scalar meson, the self-interaction term$ {\cal{L}}_{\Sigma\Sigma} $ is expressed in terms of the scalar fields as$\begin{aligned}[b] {\cal{L}}_{\Sigma\Sigma} =\;& -\frac{1}{2} \, k_0\chi^2 \left(\sigma^2+\zeta^2+\delta^2\right)+k_1 \left(\sigma^2+\zeta^2+\delta^2\right)^2 \\&+k_2\left(\frac{\sigma^4}{2} +\frac{\delta^4}{2}+3\sigma^2\delta^2+\zeta^4\right) +k_3\chi\left(\sigma^2-\delta^2\right)\zeta -k_4\chi^4\\&-\frac14\chi^4 {\rm ln}\frac{\chi^4}{\chi_0^4} + \frac{\rm d} 3\chi^4 {\rm ln}\left(\left(\frac{\left(\sigma^2-\delta^2\right)\zeta}{\sigma_0^2\zeta_0}\right)\left(\frac{\chi^3}{\chi_0^3}\right)\right), \end{aligned} $
(7) where
$ \sigma_0 = - f_\pi $ and$ \zeta_0 = \dfrac{1}{\sqrt{2}} ( f_\pi - 2 f_K) $ correspond to the vacuum values of the σ and ζ fields, where$ f_{\pi} = $ 93 MeV and$ f_K = $ 115 MeV are the pion and kaon decay constants, respectively. The value of$ d = 6/33 $ is chosen to produce the correct trace anomaly for three flavours and three colours of quarks [79]. The vector meson self-interaction term is given by$ \begin{aligned}[b]{\cal{L}}_{VV} =\;& \frac{1}{2} \, \frac{\chi^2}{\chi_0^2} \left( m_\omega^2\omega^2+m_\rho^2\rho^2+m_\phi^2\phi^2\right)\\&+g_4\left(\omega^4+6\omega^2\rho^2+\rho^4+2\phi^4\right),\end{aligned}$
(8) where
$ m_{\phi} = $ 1020 MeV is the ϕ meson mass, and$ m_{\omega} = m_{\rho} = $ 783 MeV is the mass of ω and ρ mesons. Finally, the spontaneous symmetry-breaking term$ {\cal{L}}_{SB} $ is written as [80]$ {\cal{L}}_{SB} = -\frac{\chi^2}{\chi_0^2}\left[m_\pi^2f_\pi\sigma + \left(\sqrt{2}m_K^2f_K-\frac{m_\pi^2}{\sqrt{2}}f_\pi\right)\zeta\right]. $
(9) The term
$ \mathcal{U}(\Phi,\bar{\Phi}, T) $ in Eq. (6) is the Polyakov loop effective potential, which is given in logarithmic form by [81, 82]$ \begin{aligned}[b]\frac{{\cal{U}}(\Phi,\bar{\Phi})}{T^4} =\;& -\frac{a(T)}{2}\bar{\Phi}\Phi+b(T)\mathrm{ln}\big[1-6\bar{\Phi}\Phi\\&+4(\bar{\Phi}^3+\Phi^3)-3(\bar{\Phi}\Phi)^2\big],\end{aligned}$
(10) where
$ a(T) = a_0+a_1\bigg(\frac{T_0}{T}\bigg)+a_2\bigg(\frac{T_0}{T}\bigg)^2,\ \ b(T) = b_3\bigg(\frac{T_0}{T}\bigg)^3, $
(11) with
$ a_0 = 1.81 $ ,$ a_1 = -2.47 $ ,$ a_2 = 15.2 $ , and$ b_3 = -1.75 $ [83]. Incorporating the effects of the backreaction of quarks leads to substituting the Polyakov loop potential with the QCD glue potential [74]. Denoting the Polyakov loop potential in Eq. (10) as$ {\cal{U}}_{\rm YM} $ , the improved glue Polyakov loop potential${\cal{U}}_{\rm glue}$ can be written as [75]$ \frac{{\cal{U}}_{\rm glue}(\Phi,\bar{\Phi},T_{\rm glue})}{T_{\rm glue}} = \frac{{\cal{U}}_{\rm YM}(\Phi,\bar{\Phi},T_{\rm YM})}{T_{\rm YM}}, $
(12) with
$T = T_{0}^{\rm YM}\left[1 + 0.57\left(\dfrac{T_{\rm glue}}{T_{0}^{\rm glue}-1}\right)\right]$ and$T_0 = T_{0}^{\rm YM}$ on the right-hand side (RHS) of the Eq. (10). In the present work, we take$T_{0}^{\rm YM} = T_{0}^{\rm glue} = 200$ MeV. In the last term of Eq. (6),$ \gamma_i = 2 $ is the spin degeneracy factor, and$ F_{q}^- = 1+\exp_{q}(-3E^-)+3\Phi \exp_{q}(-E^-)+3\bar{\Phi}\exp_{q}(-2E^-), $
(13) $ F_{q}^+ = 1+\exp_{q}(-3E^+)+3\bar{\Phi} \exp_{q}(-E^+)+3\Phi \exp_{q}(-2E^+), $
(14) where
$ E^- = (E_i^*(k)-{\mu_i}^{*})/k_BT $ and$ E^+ = (E_i^*(k)+{\mu_i}^{*})/k_BT $ , and$ E_i^*(k) = \sqrt{m_i^{*2}+k^2} $ represents the effective energy of quarks. The in-medium quark chemical potential$ \mu_i^* $ can be written in terms of quark chemical potential$ \mu_i $ as$ {\mu_i}^{*} = \mu_i-g_\omega^i\omega-g_\phi^i\phi-g_\rho^i\rho $ . Here,$ g^i_{\omega} $ ,$ g^i_{\phi} $ , and$ g^i_{\rho} $ are the coupling coefficients between vector meson fields and various quarks. The in-medium mass of quarks$ {m_i}^{*} = -g_{\sigma}^i\sigma - g_{\zeta}^i\zeta - g_{\delta}^i\delta + \Delta m_i $ , where$ g_{\sigma}^i $ ,$ g_{\zeta}^i $ , and$ g_{\delta}^i $ represent the coupling constants between scalar meson fields and various quarks,$ \Delta m_{u,d} = 0 $ , and$ \Delta m_s = 29 $ MeV. The term$ {\cal{V}}_{vac} $ in Eq. (6) is subtracted to get zero vacuum energy. The temperature dependence of the scalar and vector fields is obtained by minimizing the thermodynamic potential density in Eq. (6) with respect to these fields, i.e.,$\begin{aligned}[b]& \frac{\partial\Omega_q} {\partial\sigma} = \frac{\partial\Omega_q}{\partial\zeta} = \frac{\partial\Omega_q}{\partial\delta} = \frac{\partial\Omega_q}{\partial\chi} = \frac{\partial\Omega_q}{\partial\omega}\\ =\;& \frac{\partial\Omega_q}{\partial\rho} = \frac{\partial\Omega_q}{\partial\phi} = \frac{\partial\Omega_q}{\partial\Phi} = \frac{\partial\Omega_q}{\partial\bar\Phi} = 0. \end{aligned}$
(15) The resulting coupled equations are provided in the Appendix. The vector density
$ \rho_{q,i} $ and scalar density$ \rho_{q,i}^s $ in the$ q- $ PCQMF model are defined as$ \rho_{q,i} = \gamma_{i}N_c\int\frac{{\rm d}^{3}k}{(2\pi)^{3}} \Big(f_{q,i}(k)-\bar{f}_{q,i}(k) \Big), $
(16) and
$ \rho_{q,i}^{s} = \gamma_{i}N_c\int\frac{{\rm d}^{3}k}{(2\pi)^{3}} \frac{m_{i}^{*}}{E^{\ast}_i(k)} \Big(f_{q,i}(k)+\bar{f}_{q,i}(k) \Big), $
(17) respectively, with q-modified Fermi-distribution functions for quarks and antiquarks
$ f_{q,i}(k) = \frac{\Phi \exp^{q}_{q}(-E^-)+2\bar{\Phi} \exp^{q}_{q}(-2E^-)+\exp^{q}_{q}(-3E^-)} {[1+3\Phi \exp_{q}(-E^-)+3\bar{\Phi} \exp_{q}(-2E^-)+\exp_{q}(-3E^-)]^{q}} , $
(18) $ \bar{f}_{q,i}(k) = \frac{\bar{\Phi} \exp^{q}_{q}(-E^+)+2\Phi \exp^{q}_{q}(-2E^+)+\exp^{q}_{q}(-3E^+)} {[1+3\bar{\Phi} \exp_{q}(-E^+)+3\Phi \exp_{q}(-2E^+)+\exp_{q}(-3E^+)]^{q}}, $
(19) respectively.
Note that as q approaches 1, the standard Fermi-distribution functions are restored, leading us back to the conventional (extensive) PCQMF model. Additionally, as temperature,
$ T\rightarrow 0 $ , the q-dependent thermodynamic potential density$ \Omega_q $ defined in Eq. (6) also returns to its standard (extensive) form defined in Ref. [83], as long as$ q>1 $ . This implies that the nonextensive effects are more prominent in heavy-ion collision experiments where the temperature reaches a few orders of MeV and the value of q remains greater than 1 [84−86].The parameters of the model used in the present study are summarized in Table 1. These are adjusted to accurately reproduce the vacuum masses of π, K, σ, ζ, and χ and the average masses of η and
$ \eta^{'} $ [79]. The relations of the baryon, isospin, and strangeness chemical potential are defined as$ k_0 $ $ k_1 $ $ k_2 $ $ k_3 $ $ k_4 $ $ g_s $ $ \rm{g_v} $ $ \rm{g_4} $ d $ \rho_0 $ /fm$ ^{-3} $ 4.94 2.12 −10.16 −5.38 −0.06 4.76 6 37.5 0.18 0.15 $ \sigma_0 $ /MeV$ \zeta_0 $ /MeV$ \chi_0 $ /MeV$ m_\pi $ /MeV$ f_\pi $ /MeV$ m_K $ /MeV$ f_K $ /MeV$ m_\omega $ /MeV$ m_\phi $ /MeV$ m_\rho $ /MeV−93 −96.87 254.6 139 93 496 115 783 1020 783 $ g_{\sigma}^u $ $ g_{\sigma}^d $ $ g_{\sigma}^s $ $ g_{\zeta}^u $ $ g_{\zeta}^d $ $ g_{\zeta}^s $ $ g_{\delta}^u $ $ g_{\delta}^d $ $ g_{\delta}^s $ $ g^u_{\omega} $ 3.36 3.36 0 0 0 4.76 3.36 −3.36 0 3.86 $ g^d_{\omega} $ $ g^s_{\omega} $ $ g^u_{\phi} $ $ g^d_{\phi} $ $ g^s_{\phi} $ $ g^u_{\rho} $ $ g^d_{\rho} $ $ g^s_{\rho} $ 3.86 0 0 0 5.46 3.86 −3.86 0 Table 1. Parameters used in the present work [79].
$ \mu_B = \frac{3}{2}(\mu_u+\mu_d), $
(20) $ \mu_I = \frac{1}{2}(\mu_u-\mu_d), $
(21) $ \mu_S = \frac{1}{2}(\mu_u+\mu_d-2\mu_s), $
(22) respectively. Here,
$ \mu_u $ ,$ \mu_d $ , and$ \mu_s $ represent the chemical potentials of up, down, and strange quarks, respectively. -
Transport coefficients for a system in the hydrodynamic regime can be determined using the Kubo formalism [87, 88], assuming that the relaxation time is shorter than the system's lifetime. The expressions for the transport coefficients obtained using this formalism are identical to those derived within a quasiparticle approach in kinetic theory using the relaxation time approximation (RTA) [89, 90]. In kinetic theory, the transport coefficients are derived using the Boltzmann transport equation, which can be written in the relaxation time approximation (RTA) as
$ k^{\mu}\partial_{\mu}f_{i} = C[f], $
(23) where
$ C[f] $ is the collision integral. To study the transport coefficients, we are interested in small departures of the distribution function from the equilibrium,$ \delta f_{i}(\vec x,\vec k,t) = f_{i}^{'}(\vec x,\vec k,t) - f_{i}(\vec x,\vec k,t), $
(24) where
$ f_i $ is the local equilibrium distribution of quarks, and$ f^{'}_{i} $ is the non-equilibrium distribution function. Under nonextensive statistics, the equilibrium distributions are modified to their$ q- $ modified versions. This results in a$ q- $ generalized transport equation known as the nonextensive Boltzmann transport equation (NEBE) [91],$ k^{\mu}\partial_{\mu}f_{q,i} = C_q[f], $
(25) where
$ C_q[f] $ is the$ q- $ deformed collision term. In Ref. [92], the authors demonstrated that it is valid to employ conventional methods for calculating transport coefficients, beginning with NEBE. These computations yield relations for all transport coefficients that are formally analogous to those derived from the conventional Boltzmann-Gibbs distributions. The expressions of various transport coefficients used in the present work are presented below [88, 93, 94]:$ \eta = \frac{2N_{c}}{15T}\sum\limits_{i = u,d,s}\int\frac{{\rm d}^3k}{(2\pi)^3}\tau\left(\frac{k^{2}}{E_i^{*}}\right)^{2}[f_{q,i}(1-f_{q,i})+\bar{f}_{q,i}(1-\bar{f}_{q,i})], $
(26) $\begin{aligned}[b] \zeta_b =\;& \frac{2N_{c}}{T}\sum\limits_{i = u,d,s}\int\frac{{\rm d}^3k}{(2\pi)^3}\tau\frac{1}{E_i^{*2}}\Bigg[\left(\frac{1}{3}-c_{sq}^{2}\right)k^{2}-c_{sq}^{2}m_i^{*2}\\&+c_{sq}^{2}m_i^{*}T\frac{{\rm d}m_i^{*}}{{\rm d}T}\Bigg]^{2} \left[f_{q,i}(1-f_{q,i})+\bar{f}_{q,i}(1-\bar{f}_{q,i})\right], \end{aligned} $
(27) $ \begin{aligned}[b]\sigma_{el} =\;& \frac{2N_{c}}{3T}\sum\limits_{i = u,d,s}e_i^{2}\int\frac{{\rm d}^3k}{(2\pi)^3}\tau\left(\frac{k}{E_i^{*}}\right)^{2}\\&\times[f_{q,i}(1-f_{q,i})+\bar{f}_{q,i}(1-\bar{f}_{q,i})], \end{aligned}$
(28) $ \begin{aligned}[b] \kappa =\;& \frac{2N_c}{3T^2}\sum\limits_{i = u,d,s}\int\frac{{\rm d}^3k}{(2\pi)^3}\tau\left(\frac{k}{E_i^*}\right)^2[(E_i^*-h_q)^2f_{q,i}(1-f_{q,i})\\&+(E_i^*+h)^2 \bar{f}_{q,i}(1-\bar{f}_{q,i})]. \end{aligned} $
(29) Notably,
$ f_{q, i} $ and$ \bar{f}_{q, i} $ are not the standard Fermi-distribution functions but rather the q-version of the Fermi-distribution functions, given by Eqs. (18) and (19), respectively.$ c^2_{sq} $ is the speed of sound at constant entropy defined as$ c_{sq}^{2} = \left(\partial p_q/\partial \epsilon_q\right)_{s_{q}} = s_{q}/c_{vq} $ , and$ c_{vq} $ is the specific heat at constant volume defined as$ c_{vq} = \left(\partial\epsilon_q/\partial T\right)_V $ . The pressure is given by$ p_q = -\Omega_q $ , while the energy density and entropy density are defined as$\epsilon_q = \Omega_q+ \sum_{i = u,d,s} \mu_i^* \rho_i+ Ts_q$ and$ s_q = -\partial \Omega_q/\partial T $ , respectively. The heat function$ h_q = (\epsilon_q + p_q)/\rho_q $ diverges at$ \mu = 0 $ , where$ \rho_q $ diverges. The relaxation time τ is a measure of the timescale over which the distribution function relaxes back to equilibrium and is defined as [95]$ \tau = \frac{1}{5.1T\alpha_S^2\log(\frac{1}{\alpha_S})[1+0.12(2N_f+1)]}, $
(30) where
$ \alpha_s $ is the temperature and chemical potential-dependent strong coupling constant given by [96, 97]$\begin{aligned}[b]& \alpha_S(T,\mu) = \frac{6\pi}{(33-2N_f)\log\left[\frac{T}{\Lambda_T}\sqrt{1+(\frac{\mu}{\pi T})^2}\right]}\\&\times\left[1-\frac{3(153-19N_f)}{(33-2N_f)^2}\frac{\log\left(2\log\frac{T}{\Lambda_T}\sqrt{1+(\frac{\mu}{\pi T})^2}\right)}{\log\left(\frac{T}{\Lambda_T}\sqrt{1+(\frac{\mu}{\pi T})^2}\right)}\right],\end{aligned} $
(31) with
$ \Lambda_T = 70 $ MeV [97]. -
The coupled equations are obtained by minimizing thermodynamic potential density,
$ \Omega_q $ , with respect to the various fields of the q-PCQMF model and are given as$ \begin{aligned}[b] \frac{\partial \Omega_q}{\partial \sigma} =\;& k_{0}\chi^{2}\sigma-4k_{1}\left( \sigma^{2}+\zeta^{2} +\delta^{2}\right)\sigma-2k_{2}\left( \sigma^{3}+3\sigma\delta^{2}\right) -2k_{3}\chi\sigma\zeta - \frac{\rm d}{3} \chi^{4} \bigg (\frac{2\sigma}{\sigma^{2}-\delta^{2}}\bigg ) +\left( \frac{\chi}{\chi_{0}}\right) ^{2}m_{\pi}^{2}f_{\pi}\\ &- \left(\frac{\chi}{\chi_0}\right)^2m_\omega\omega^2 \frac{\partial m_\omega}{\partial\sigma} - \left(\frac{\chi}{\chi_0}\right)^2m_\rho\rho^2 \frac{\partial m_\rho}{\partial\sigma} -\sum\limits_{i = u,d} g_{\sigma}^i\rho_{q,i}^{s} = 0 , \end{aligned} $
(A1) $ \begin{aligned}[b] \frac{\partial \Omega_q}{\partial \zeta} =\;& k_{0}\chi^{2}\zeta-4k_{1}\left( \sigma^{2}+\zeta^{2}+\delta^{2}\right) \zeta-4k_{2}\zeta^{3}-k_{3}\chi\left( \sigma^{2}-\delta^{2}\right)-\frac{\rm d}{3}\frac{\chi^{4}}{{\zeta}} + \left(\frac{\chi}{\chi_{0}} \right) ^{2}\left[ \sqrt{2}m_{K}^{2}f_{K}-\frac{1}{\sqrt{2}} m_{\pi}^{2}f_{\pi}\right]\\&-\left(\frac{\chi}{\chi_0}\right)^2m_\phi\phi^2 \frac{\partial m_\phi}{\partial\zeta} -\sum\limits_{i = s} g_{\zeta}^i\rho_{q,i}^{s} = 0 , \end{aligned} $
(A2) $ \frac{\partial \Omega_q}{\partial \delta} = k_{0}\chi^{2}\delta-4k_{1}\left( \sigma^{2}+\zeta^{2}+\delta^{2}\right) \delta-2k_{2}\left( \delta^{3}+3\sigma^{2}\delta\right) +\mathrm{2k_{3}\chi\delta \zeta} + \frac{2}{3} {\rm d} \chi^4 \left( \frac{\delta}{\sigma^{2}-\delta^{2}}\right) -\sum\limits_{i = u,d} g_{\delta}^i\rho_{q,i}^{s} = 0 , $
(A3) $ \begin{aligned}[b] \frac{\partial \Omega_q}{\partial \chi} =\;& \mathrm{k_{0}\chi} \left( \sigma^{2}+\zeta^{2}+\delta^{2}\right)-k_{3} \left( \sigma^{2}-\delta^{2}\right)\zeta + \chi^{3}\left[1 +{\rm {ln}}\left( \frac{\chi^{4}}{\chi_{0}^{4}}\right) \right] +(4k_{4}-d)\chi^{3} - \frac{4}{3} {\rm d} \chi^{3} {\rm {ln}} \Bigg ( \bigg (\frac{\left( \sigma^{2} -\delta^{2}\right) \zeta}{\sigma_{0}^{2}\zeta_{0}} \bigg ) \bigg (\frac{\chi}{\mathrm{\chi_0}}\bigg)^3 \Bigg )+\\& \frac{2\chi}{\chi_{0}^{2}}\left[ m_{\pi}^{2} f_{\pi}\sigma +\left(\sqrt{2}m_{K}^{2}f_{K}-\frac{1}{\sqrt{2}} m_{\pi}^{2}f_{\pi} \right) \zeta\right] - \frac{\chi}{{\chi^2_0}}({m_{\omega}}^2 \omega^2+{m_{\rho}}^2\rho^2) = 0, \end{aligned} $
(A4) $ \frac{\partial \Omega_q}{\partial \omega} = \frac{\chi^2}{\chi_0^2}m_\omega^2\omega+4g_4\omega^3+12g_4\omega\rho^2 - \sum\limits_{i = u,d}g_\omega^i\rho_{q,i} = 0, $
(A5) $ \frac{\partial \Omega_q}{\partial \rho} = \frac{\chi^2}{\chi_0^2}m_\rho^2\rho+4g_4\rho^3+12g_4\omega^2\rho - \sum\limits_{i = u,d}g_\rho^i\rho_{q,i} = 0, $
(A6) $ \frac{\partial \Omega_q}{\partial \phi} = \frac{\chi^2}{\chi_0^2}m_\phi^2\phi+8g_4\phi^3 - \sum\limits_{i = s}g_\phi^i\rho_{q,i} = 0, $
(A7) $ \begin{aligned}[b] \frac{\partial \Omega_q}{\partial \Phi} =\;&\bigg[\frac{-a(T)\bar{\Phi}}{2}-\frac{6b(T) (\bar{\Phi}-2{\Phi}^2+{\bar{\Phi}}^2\Phi) }{1-6\bar{\Phi}\Phi+4(\bar{\Phi}^3+\Phi^3)-3(\bar{\Phi}\Phi)^2}\bigg]T^4 -\sum\limits_{i = u,d,s}\frac{2k_BTN_C}{(2\pi)^3} \\ & \int_0^\infty d^3k \bigg[\frac{\exp_q\left(\dfrac{-(E_i^*(k)-{\mu_i}^{*})}{k_BT}\right)}{\left(1+\exp_q\left(\dfrac{-3(E_i^*(k)-{\mu_i}^{*})}{k_BT}\right)+3\Phi \exp_q\left(\dfrac{-(E_i^*(k)-{\mu_i}^{*})}{k_BT}\right) +3\bar{\Phi}\exp_q\left(\dfrac{-2(E_i^*(k)-{\mu_i}^{*})}{k_BT}\right)\right)^q} \\& +\frac{\exp_q\left(\dfrac{-2(E_i^*(k)+{\mu_i}^{*})}{k_BT}\right)}{\left(1+\exp_q\left(\dfrac{-3(E_i^*(k)+{\mu_i}^{*})}{k_BT}\right) +3\bar{\Phi} \exp_q\left(\dfrac{-(E_i^*(k)+{\mu_i}^{*})}{k_BT}\right)+3\Phi \exp_q\left(\dfrac{-2(E_i^*(k)+{\mu_i}^{*})}{k_BT}\right)\right)^q}\bigg] = 0, \end{aligned}$
(A8) $ \begin{aligned}[b] \frac{\partial \Omega_q}{\partial \bar{\Phi}} =\;& \bigg[\frac{-a(T)\Phi}{2}-\frac{6b(T) (\Phi-2{\bar{\Phi}}^2+{\Phi}^2\bar{\Phi}) }{\mathrm{1-6\bar{\Phi}\Phi+4(\bar{\Phi}^3+\Phi^3)-3(\bar{\Phi}\Phi)^2}}\bigg]T^4 -\sum\limits_{i = u,d,s}\frac{2k_BTN_C}{(2\pi)^3} \\& \int_0^\infty d^3k \bigg[\frac{\exp_q\left(\dfrac{-2(E_i^*(k)-{\mu_i}^{*})}{k_BT}\right)}{\left(1+\exp_q\left(\dfrac{-3(E_i^*(k)-{\mu_i}^{*})}{k_BT}\right)+3\Phi \exp_q\left(\dfrac{-(E_i^*(k)-{\mu_i}^{*})}{k_BT}\right) +3\bar{\Phi}\exp_q\left(\dfrac{-2(E_i^*(k)-{\mu_i}^{*})}{k_BT}\right)\right)^q} \\& +\frac{\exp_q\left(\dfrac{-(E_i^*(k)+{\mu_i}^{*})}{k_BT}\right)}{\left(1+\exp_q\left(\dfrac{-3(E_i^*(k)+{\mu_i}^{*})}{k_BT}\right) +3\bar{\Phi} \exp_q\left(\dfrac{-(E_i^*(k)+{\mu_i}^{*})}{k_BT}\right)+3\Phi \exp_q\left(\dfrac{-2(E_i^*(k)+{\mu_i}^{*})}{k_BT}\right)\right)^q}\bigg] = 0. \end{aligned} $
(A9)
Impact of nonextensivity on the transport coefficients of strongly interacting QCD matter
- Received Date: 2024-09-19
- Available Online: 2025-02-15
Abstract: Tsallis nonextensive statistics is applied to study the transport coefficients of strongly interacting matter within the Polyakov chiral