-
Einstein's General Relativity (GR) theory [1] has stood as the cornerstone of gravitational theory for over a century, successfully describing the dynamics of gravity on both cosmological and astrophysical scales. However, the discovery of the accelerated expansion of the universe and discrepancies in galactic dynamics have motivated the exploration of alternative theories of gravity such as
$ f(R,L_m) $ and$ f(R,T) $ [2−5]. The product of the unification of latter theories is$ f(R,L_m,T) $ gravity theory [6, 7], whose novelty is that it modifies the gravitational action to account for observed phenomena beyond the scope of GR.Quark stars (QSs) [8−10], also known as strange stars or strange QSs, are theoretical entities composed of ultra-dense quark matter, proposed as an alternative to neutron stars (NSs) formed from the gravitational collapse of massive stars [11]. The conjecture is that these hypothetical objects emerge in the aftermath of supernova explosions, where intense pressure and density facilitate the conversion of protons and electrons into quarks [12]. Recent works [13−19] explore several aspects of QSs, including their structure, formation, and observable properties. Investigations into formation mechanisms scrutinize conditions conducive to QS genesis, including the role of quark deconfinement within NS cores during supernova events, elucidating the transition from neutron matter to quark matter under extreme pressure and temperature conditions. Crucially, recent researchers focus on elucidating the equation of state (EOS) of quark matter, delineating the intricate relationship among pressure, energy density, and temperature within these exotic objects [20−23]. Theoretical models and computational simulations are leveraged to unravel the behavior of quark matter under extreme astrophysical environments. In this context, the theory of
$ f(R,L_m,T) $ gravity, proposed by Haghani and Harko [6], offers a promising framework for investigating these celestial objects. This theory generalizes and unifies the$ f(R,T) $ and$ f(R,L_m) $ gravity models, where R is the Ricci scalar, T is the trace of the energy-momentum tensor,$ T_{\mu\nu} $ , and$ L_m $ is the matter Lagrangian [7]. In$ f(R,L_m,T) $ gravity, the gravitational Lagrangian is given by an arbitrary function of R, T, and$ L_m $ , such that$ L_{grav} = f(R,L_m,T) $ . The full action in$ f(R,L_m,T) $ gravity theories is expressed as [6]$ S = \frac{1}{16 \pi} \int f(R,L_m,T) \sqrt{-g} {\rm d} ^4x + \int L_m \sqrt{-g} {\rm d}^4x, $
(1) where g is the determinant of the metric tensor
$ g_{\mu\nu} $ .As highlighted previously,
$ f(R,L_m,T) $ gravity theories refer to modifications of Einstein's GR theory. These theories are part of the broader framework of modified gravity, which addresses various cosmological and astrophysical phenomena beyond what GR can explain. Notably,$ f(R,L_m,T) $ gravity theories have garnered attention for their potential to elucidate the accelerated expansion of the universe [24, 25], gravitational lensing effects [26−33], and the dynamics of galaxies and galaxy clusters [34−37], all without invoking the existence of dark matter or dark energy. They have been proposed as alternatives to the conventional concept of dark energy. Furthermore, within the realm of modified gravity theory, investigations into dilaton fields [38−43], quasinormal modes [44−46], and inflationary scenarios [47−49] further expand extant understanding of gravitational physics and its implications for the cosmos. These interconnected topics underscore the interdisciplinary nature of modern gravitational research and its quest to uncover the fundamental mechanisms governing the universe's behavior.This study was conducted to investigate the impact of the
$ f(R,L_m,T) $ gravity theory on the internal structure of QSs. We focused on the algebraic function originally proposed in Refs. [6, 7], i.e.,$ f(R,L_m,T) = R + \alpha T L_m $ , with α being a matter-geometry coupling constant. We examined the matter Lagrangian density, expressed as$ L_m = -\rho $ , following which we analyzed its effects on the principal macroscopic properties of compact stars, including mass and radius. We derived the modified Tolman-Oppenheimer-Volkoff (TOV) equations. We conducted numerical analysis on different parameter variations, and present the results in this paper. Further, we examined the static stability criterion, adiabatic index, and sound velocity. Our findings will provide valuable insights into the behavior of compact stars in the context of$ f(R,L_m,T) $ gravity. It should also be noted that although Ref. [50] uses a similar gravitational framework, we used a distinct equation of state (EoS) for quark matter and focused on different aspects of QS properties. The referenced study [50] explores color superconducting quark matter with the CFL phase, whereas our investigation centered around the effects of anisotropy and different EoS models to reveal new insights about the influence of the$ \alpha T L_m $ term on stellar structure.Although our manuscript correctly implements the first matching condition by ensuring the continuity of the metrics on the surface of the star, one can examine the necessity to seek the additional junction conditions required for a comprehensive matching [51, 52] because it is known that mere metric continuity does not guarantee a smooth matching between interior and exterior metrics. In the context of our chosen theory,
$ f(R,L_m,T) $ gravity, it is imperative to consider the conditions imposed by the presence of$ L_m $ in addition to R and T. As outlined in the literature, the matching in$ f(R,L_m,T) $ gravity necessitates the continuity of extrinsic curvature at the surface. The choice of$ f(R,L_m,T) = R + \alpha T L_m $ is motivated by its ability to introduce meaningful deviations from GR while maintaining mathematical simplicity. This form incorporates a coupling between matter and geometry through the trace of the energy-momentum tensor T and the matter Lagrangian$ L_m $ , providing an insightful framework to explore high-density environments like QSs. Recent studies have shown that this model leads to observable effects in astrophysical systems, making it a suitable candidate to examine the structure and stability of QSs [6, 7]. Importantly, it reduces to GR when$ \alpha = 0 $ , ensuring consistency with known results [2, 3]. After examining this, we discovered that$ f(L_{m}) $ contributes to the thin-shell's stress-energy tensor only if the extrinsic curvature is continuous at the surface. Therefore, our analysis shows that similar to$ f(T) $ [53], the inclusion of$ f(L_m) $ does not introduce further junction conditions. Therefore, our study aligns with the requisite smooth matching conditions established within the framework of$ f(R,L_m,T) $ gravity, ensuring the robustness of our results. However, this matter is purely informational and outside the scope of this article. This article's primary goal is to examine how$ f(R,L_m,T) $ gravity affects compact stars' interior structure. It is anticipated that this hypothesis will be most evident in these stars' high-density cores.The paper is structured as follows: In Section II, we review
$ f(R, L_m, T) $ gravity theory and serve the gravitational field equations of the theory. Section III discusses the gravitational field equations in$ f(R,L_m,T) $ theories, while in Sec. IV, we prescribe a quark matter EoS and a specific quasi-local EoS that describe local anisotropy for QSs. Numerical results and discussions for different parameter variations are presented in Sec. V, followed by an examination of the static stability criterion, adiabatic index, and sound velocity in Sec. VI. Finally, we draw our conclusions in Sec. VII. (We adopt geometrized units throughout this exposition, setting$ c = G = 1 $ , while retaining physical units for clarity in presentation.) -
In this section, we review the modifications to classical gravity theories in the context of generalized
$ f(R, L_m, T) $ gravity [6, 7], where arbitrary functions of the Ricci scalar R, the matter Lagrangian$ L_m $ , and the trace of the stress-energy tensor T are included. The gravitational Lagrangian in this modified framework is$ f(R, L_m, T) $ , allowing for a more comprehensive description of gravitational interactions in the presence of matter.The gravitational Lagrangian in this unified framework encompasses arbitrary functions of the Ricci scalar R, the trace T of the energy-momentum tensor
$ T_{\mu\nu} $ , and the matter Lagrangian$ L_m $ , yielding$ L_{grav} = f(R,L_m,T) $ . Consequently, the complete action in$ f(R,L_m,T) $ gravity theory results in Eq. (1). The variation of action (1) with respect to the inverse metric$ g^{\mu\nu} $ yields the following field equations in$ f(R,L_m,T) $ gravity:$ \begin{aligned}[b]& f_RR_{\mu\nu} - \frac{1}{2} [f-(f_L + 2f_T)L_m] g_{\mu\nu} + (g_{\mu\nu}\Box -\nabla_{\mu} \nabla_{\nu})f_R \\ =\;& \left[8\pi + \frac{1}{2}(f_L + 2f_T)\right] T_{\mu\nu} + f_T\tau_{\mu\nu}, \end{aligned} $
(2) where
$ \Box \equiv \partial_{\mu}(\sqrt{-g}g^{\mu\nu}\partial_{\nu})/\sqrt{-g} $ , and$ f_{R} $ ,$ f_{T} $ ,$ f_{L} $ denote the partial derivatives of f with respect to R, T, and$ L_{m} $ , respectively. We adopted$ L_m = -\rho $ primarily for its physical relevance in modeling compact stars. The energy density ρ directly contributes to the gravitational source term, aligning with conventional understanding in GR, where matter density is central to the formation of gravitational fields. In contrast, using$ L_m = p $ would introduce significant differences in the mass-radius relations and stability, as shown in [7].On the other hand, alternative forms of$ L_m $ could be explored in future work. Although$ L_m = p $ has been explored in some studies, as in Ref. [2], we chose$ L_m = -\rho $ to maintain consistency with the standard treatment of compact objects and simplify numerical comparisons with GR. This choice ensures that the matter's gravitational effects remain physically intuitive, with ρ acting as the primary source.$ R_{\mu\nu} $ stands for the Ricci tensor,$ \nabla_{\mu} $ represents the covariant derivative concerning the symmetric connection associated with$ g_{\mu\nu} $ , and$ \tau_{\mu\nu} $ is a new tensor defined as [6]$ \tau_{\mu\nu} = 2g^{\gamma\zeta} \frac{\partial^{2}L_m}{\partial g^{\mu\nu} \partial g^{\gamma \zeta}}. $
(3) The form of
$ f(R,L_m,T) $ dictates the nature of the gravitational dynamics. Specifically, when$ f(R,L_m,T) = f(R) $ , Eq. (2) reduces to the field equations of metric$ f(R) $ gravity [54−56]. Similarly, for$ f(R,L_m,T) = f(R,T) $ , the$ f(R,T) $ gravity model is recovered, while$ f(R,L_m,T) = f(R,L_m) $ yields the field equations of the$ f(R,L_m) $ theory. Furthermore,$ f(R,L_m,T) = R $ retrieves the standard field equations of GR:$ R_{\mu\nu}-\dfrac{1}{2}g_{\mu\nu}R = 8\pi T_{\mu\nu} $ .The covariant divergence of the field equations (2) leads to the non-conservation equation of the energy-momentum tensor
$ T_{\mu\nu} $ :$ \begin{aligned}[b] \nabla^{\mu}T_{\mu\nu} =\;& \ \frac{1}{8\pi + f_m}\Big[ \nabla_{\nu} (L_m f_m) - T_{\mu\nu} \nabla^{\mu} f_m \\ &\left.- A_\nu - \frac{1}{2}(f_T \nabla_\nu T + f_L \nabla_\nu L_m) \right] , \end{aligned} $
(4) where
$ f_m = f_T + \dfrac{1}{2}f_L $ and$ A_\nu = \nabla^\mu(f_T \tau_{\mu\nu}) $ .The trace of the field equations yields a second-order differential equation:
$ 3\square f_R+ Rf_R- 2(f- 2f_mL_m) = (8\pi+ f_m)T + f_T\tau, $
(5) where τ represents the trace of the tensor
$ \tau_{\mu\nu} $ . For the specific functional form$ f(R,L_m,T) = f(R) $ , this equation reduces to the well-known dynamical equation for the Ricci scalar in pure$ f(R) $ gravity theories [57, 58].For simplicity, we focus on the algebraic function originally proposed in Ref. [6], i.e.,
$ f(R,L_m,T) = R + \alpha T L_m $ , with α denoting a matter-geometry coupling constant. In this context, Eqs. (2) and (5) simplify to$ G_{\mu\nu} = \left[ 8 \pi + \frac{\alpha}{2}(T + 2 L_m) \right] T_{\mu\nu} + \alpha L_m(\tau_{\mu\nu}- L_mg_{\mu\nu}) , $
(6) and
$ \begin{aligned}[b]\nabla^{\mu}T_{\mu\nu} = \;&\frac{\alpha}{8 \pi +\alpha (L_m + T/2 )}\Bigg[ \nabla_{\nu}\Big(L_m^{2} + \frac{1}{2}T L_m \Big) \\&- T_{\mu\nu} \nabla^{\mu} \Big(L_m + \frac{T}{2} \Big) - \nabla^\mu(L_m\tau_{\mu\nu})\\&- \frac{1}{2}(L_m \nabla_{\nu}T + T \nabla_{\nu}L_m) \Bigg], \end{aligned}$
(7) respectively, where
$ G_{\mu\nu} $ denotes the Einstein tensor. Remarkably, the Einstein field equations$ G_{\mu\nu} = 8\pi T_{\mu\nu} $ and the conservation equation$ \nabla^{\mu}T_{\mu\nu} = 0 $ are retrieved when$ \alpha = 0 $ . -
In this section, we will explore anisotropic QSs within the framework of the
$ f(R,L_m,T) = R + \alpha TL_m $ gravity model. To begin our analysis, we will concentrate on the static and spherically symmetric spacetime. To do so, we initially adopt the metric ansatz described by the following line element:$ {\rm d}s^{2} = -{\rm e}^{\nu(r)}{\rm d}t^{2} + {\rm e}^{\lambda(r)}{\rm d}r^{2} + r^{2}({\rm d}\theta^{2} + \sin{\theta}^{2}{\rm d}\phi^{2}), $
(8) where
$ \nu(r) $ and$ \lambda(r) $ , both functions of the radial coordinate r, are the two unknown functions. Within a spherically symmetric star, the matter source is modeled as an anisotropic fluid, indicating that the radial pressure$ p_r $ differs from the transverse pressure$ p_{\perp} $ . The energy-momentum tensor for the static configuration is given by:$ T_{\mu\nu} = (\rho + p_{\perp})u_{\mu}u_{\nu} + p_{\perp} g_{\mu\nu} - \Delta \chi_{\mu} \chi_{\nu}, $
(9) where ρ represents the energy density and
$ \Delta \equiv p_{\perp} - p_r $ . Here,$ u_{\mu} $ denotes the fluid 4-velocity, satisfying$ u_{\mu}u^{\mu} = -1 $ , which can be expressed as$ u^\mu ={\rm e}^{-\nu/2}\delta_0^\mu $ . Consequently,$ T_\mu^\nu = \text{diag} (-\rho, p_r, p_{\perp}, p_{\perp}) $ and$ T = -\rho + p_r + 2 p_{\perp} $ .Given that the Lagrangian density
$ L_m $ corresponding to the matter source is not unique, we have the option to choose either$ L_m = p $ or$ L_m = -\rho $ (see [59] for further discussion). Although consensus on which Lagrangian to consider is lacking, we opted to assume$ L_m = -\rho $ and derive the equations of hydrostatic equilibrium for QSs within the context of$ f(R,L_m,T) $ gravity. This assumption allows us to rewrite Eq. (6) as:$ G_{\mu\nu} = \left[8\pi + \frac{\alpha}{2}(- 3\rho+ p_r+2p_{\perp})\right] T_{\mu\nu} - \alpha \rho^2 g_{\mu\nu}. $
(10) Utilizing the spherically symmetric metric (8) and Eq. (10), the non-zero components of the field equations become
$ e^{-\lambda}\left(\frac{\lambda'}{r} - \frac{1}{r^2}\right) + \frac{1}{r^2} = 8\pi \rho + \frac{\alpha}{2}(-\rho + p_r + 2p_{\perp})\rho , $
(11) $ e^{-\lambda}\left(\frac{\nu'}{r} + \frac{1}{r^2}\right) - \frac{1}{r^2} = 8\pi p_r + \frac{\alpha}{2}(-3\rho + p_r + 2p_{\perp} )p_r - \alpha \rho^2 , $
(12) $\begin{aligned}[b]& \frac{e^{-\lambda}}{2} \left[\nu''+\frac{(\nu')^2-\nu' \lambda'}{2} +\frac{\nu'- \lambda'}{r}\right] \\=\;& 8\pi p_{\perp} + \frac{\alpha}{2}(-3\rho + p_r + 2p_{\perp} )p_{\perp} - \alpha \rho^2 ,\end{aligned} $
(13) where the prime denotes a derivative with respect to r. On the other hand, the covariant divergence of Eq. (7) yields the modified TOV equations:
$ p_r' + \frac{\nu'}{2}(\rho + p_r) = \frac{\alpha\left[ 4\rho\rho'+ p(3 \rho' - p'_r - 2 p'_{\perp}) \right]}{\tilde{p}} + \frac{2 \Delta}{r}, $
(14) where
$ \tilde{p} \equiv 16 \pi + \alpha \left( p_r +2 p_{\perp} - 3 \rho \right), $
(15) $ \Delta \equiv p_{\perp} - p_r. $
(16) According to Eq. (14), the standard conservation equation does not hold for this theory. It can be readily shown that when
$ \alpha = 0 $ , the standard conservation equation is recovered.Figures 1 and 2 depict various physical quantities as functions of the radial coordinate r, illustrating the sensitivity of the system to parameters α and β, respectively.
Figure 1. (color online) From top to bottom, energy density ρ, radial
$ p_r $ and transverse$ p_t $ pressures as functions of the radial coordinate r. We start by varying$ \alpha \in [-0.2,0.2] \mu_1 $ where$ \mu_1 = 10^{-79} $ s4/kg2. We define the other parameters,$ B = 60 $ MeV/fm3 and$ \beta = 1.5 $ . The dashed black line represents the GR solution$ (\alpha = 0) $ for an anisotropic fluid distribution.Figure 2. (color online) From top to bottom, energy density ρ, radial
$ p_r $ and transverse$ p_t $ pressures as functions of the radial coordinate r. We start by varying the anisotropy parameter$ \beta \in [-1.5, 1.5] $ , whereas the other parameters are$ B = 60 $ MeV/fm3 and$ \alpha = -0.2 \mu_1 $ with$ \mu_1 = 10^{-79} $ s4/kg2, respectively. The black dashed line denotes an isotropic arrangement in$ f(R,L_m,T) $ gravity.Before exploring the internal composition of compact stars in
$ f(R,L_m,T) = R + \alpha T L_m $ gravity, we introduced a mass function$ m(r) $ through the relation:$ {\rm e}^{-\lambda(r)} = 1 - \frac{2m(r)}{r}, $
(17) where
$ m(r) $ represents the total mass enclosed within the radius r. Subsequently, by substituting this expression into Equations (11)−(14), we ultimately derive$ \frac{{\rm d}m}{{\rm d}r} = 4\pi r^2\rho + \frac{\alpha r^2}{4}(p_r + 2 p_{\perp} - \rho)\rho , $
(18) $ \begin{aligned}[b] \frac{{\rm d}p}{{\rm d}r} =\;& -\frac{\tilde{p}}{\tilde{p}+ \alpha p_r} \bigg( \frac{(p_r + \rho)}{4r (r- 2m)}(4m + r^3 (p_r \tilde{p} - 2 \alpha \rho^2) ) \\ & +\frac{2 \Delta}{r} + \frac{1}{\tilde{p}} \left[ \alpha \{ p_r (2 p_{\perp}' - 3 \rho') - 4 \alpha \rho \rho' \} \right] \bigg). \end{aligned} $
(19) In this framework, we encountered five unknown functions with three differential equations. However, introducing two appropriate ad hoc assumptions enabled us to close the system of equations. In this context, we considered an EoS relating radial pressure to the energy density of the fluid and a quasi-local EoS proposed by Horvat et al. in Ref. [60], which offers reliable solutions and aligns effectively with the GW170817 data. Based on these considerations, we numerically solved the structure equations with the appropriate boundary conditions, as follows:
$ m(0) = 0, \;\; \;\; \;\; \rho(0) = \rho_c , $
(20) ensuring regularity at the center of the star, where
$ \rho_c $ denotes the central energy density. Finally, we solved these equations by integrating outward until the pressure vanished, i.e.,$ p_r(R) = p_{\perp}(R) = 0 $ , where R is identified as the radius of the star. Then, the interior solution was matched with the exterior Schwarzschild vacuum solution:$ {\rm e}^{\nu(R)} = {\rm e}^{-\lambda(R)} = 1 - \frac{2M}{R}, $
(21) where M signifies the total mass of the star.
-
Observations of compact objects via gravitational wave (GW) signals or other recent astrophysical data have provided strong and direct evidence in favor of more massive compact stars. Interestingly, these observations also impose constraints on the NS EoS, suggesting that their internal composition likely includes free quarks in their core, rather than only neutrons and other non-fundamental particles. The central density is expected to reach several times higher than the nuclear saturation density (
$ n_0 \approx 0.148\ \text{fm}^{-3} $ ), leading to the possible formation of quark matter [61, 62]. This density is typical within atomic nuclei and provides a reference point for understanding the extreme conditions that may lead to the deconfinement of quarks. This highly compressed nuclear matter is anticipated to undergo a phase transition to deconfined quark matter, releasing its constituent quarks and gluons. Therefore, the EoS of quark matter plays a pivotal role in shaping the structure of the star beyond nuclear saturation density. From Sec. IV onward, we use "NS" units ($ \hbar = c = 1 $ ) consistently in discussing the numerical results and the value of α. To ensure that all physical quantities are presented coherently throughout the manuscript, the value of α mediated by$ \mu_1 $ is expressed using units appropriate for NS contexts.In this work, we focused on the applications of strange matter, which are described by the MIT bag model EoS for modeling compact stellar objects. The MIT bag model is a simple phenomenological model for quark matter proposed in the 1970s to explain hadrons in terms of quarks [63]. In the model, it is assumed that quarks are asymptotically free and confined to a spherical region of space by the bag constant B. Consequently, the MIT bag model establishes a relation between energy density and pressure, given by
$ p_r = \dfrac{1}{3}\left(\rho -4B\right). $
(22) This expression indicates that at
$ \rho = 4B $ , the external pressure acting on a bag vanishes. The constant B, referred to as the bag pressure, is typically constrained within a range of$ 57 \leq B \leq 92 $ MeV/fm3. This range is supported by various studies in the field [64, 65]. Following [66], we used$ B = 60\ {\rm{MeV}}/{\rm{fm}}^3 $ to align the model with observational astrophysical data. Noteworthily, this chosen value also satisfies the$ 2M_\odot $ constraints within the framework of GR. For local anisotropy, we employed the quasilocal EoS suggested by Horvat et al. [60]. The choice of this model was driven by two key factors: (a) The simplicity and physical consistency of the Horvat model, which connects the anisotropy directly to the local compactness parameter$ \mu = \dfrac{2m(r)}{r} $ , makes it an intuitive choice for compact stars. (b) Furthermore, as shown in [7], the Horvat model emerges naturally from the modified TOV equation for${\rm d}p_r/{\rm d}r$ , making it a suitable and physically grounded approach for studying anisotropic compact stars. Although other anisotropic models could have been considered, such as those based on pressure or density gradients, the Horvat model offers a straightforward and analytically tractable way to explore anisotropic effects. This model's ability to smoothly transition between isotropic and anisotropic cases by varying the parameter β further supports its use in our analysis. Within this framework, various solutions have been obtained in both GR and modified gravity theory. Thus, we express the quasi-local equation as$ \Delta \equiv p_{\perp} - p_r = \beta p_r \mu, $
(23) where the free parameter β plays a crucial role in measuring the deviation from isotropy, taking positive or negative values, see Refs. [67−73] for detailed discussions. In our calculations, we assumed
$ \beta \in [-1.5,1.5] $ . The quantity$ \mu \equiv 2m(r)/r $ represents the local measure of compactness. Furthermore, the choice of Eq. (23) ensures that$ \Delta = 0 $ at the center, thereby recovering the isotropic case. In Figs. 1 and 2, we depict the energy density ρ, radial pressure$ p_r $ , and transverse pressure$ p_t $ as functions of the radial coordinate r for two different parameter sets (see Tables 1 and 2 for more details). It is evident from these figures that the energy density of the QS is non-zero at the surface, and the pressure components vanish simultaneously at the surface of the star, i.e.,$ p_r \left( r \rightarrow R \right) = p_{\perp} \left( r \rightarrow R \right) = 0 $ . Clearly, when$ \beta = 0 $ , the anisotropy factor vanishes at the center, ensuring regularity in the interior.$\alpha \times 10^{-78} \dfrac{s^4}{kg^2}$ M/ $ M_{\odot} $ R/km $ \rho_c $ /(MeV/fm3)$ M/R $ −0.2 2.72 11.68 847 0.345 −0.1 2.63 11.57 847 0.337 0.0 2.49 11.42 847 0.323 0.1 2.29 11.34 791 0.299 0.2 1.96 11.15 735 0.260 Table 1. Summary of the resulting properties of anisotropic QSs for the variation of α in
$ f(R,L_m,T) $ gravity, as briefly discussed in Sec. V.A.β M/ $ M_{\odot} $ R/km $ \rho_c $ /(MeV/fm3)$ M/R $ −1.5 1.91 10.39 1,415 0.272 −1.0 2.03 10.59 1,334 0.284 −0.5 2.16 10.78 1,253 0.298 0.0 2.30 10.96 1,173 0.311 0.5 2.44 11.20 1,052 0.323 1.0 2.58 11.47 931 0.334 1.5 2.72 11.67 851 0.346 Table 2. Summary of the resulting properties of anisotropic QSs for the variation of β in
$ f(R,L_m,T) $ gravity, as briefly discussed in Sec. V.B. -
Apart from the mass-radius relations, the most important issue is related to stability of the configuration. A detailed study is presented subsequently in this section.
-
Let us now focus on the stability of the equilibrium configuration through static stability criterion (SSC) [81, 82]. This criteria has been presented in
$ M-\rho_c $ plane, where M is the mass and$ \rho_c $ is the central density of the star. In a similar vein, SSC has also been applied in modified gravity theory , see Refs. [83−85] for more. We should also note that this is a necessary but insufficient condition. The mathematical form of these inequalities is as follows$ \frac{{\rm d} M}{{\rm d} \rho_c} < 0 \; \rightarrow \text{indicating an unstable configuration}, $
(24) $ \frac{{\rm d} M}{{\rm d} \rho_c} > 0 \; \rightarrow \text{indicating a stable configuration}. $
(25) In Fig. 5, we show the
$ M-\rho_c $ curves for the consider cases mentioned above. For plotting, we use the same parameters as shown in Fig. 1 and 2, respectively. Here, we see that the total mass is an increasing function of the central density, reaching a point where$ (M_{\text{max}}, R_{M_{\text{max}}}) $ exists. This point is known as a boundary point (indicated in the figure by pink points) which can separate the stable configuration region indicated by${\rm d}M/{\rm d}\rho_c > 0$ from the unstable one. -
An additional test was performed to evaluate the stability of the configuration via adiabatic index, γ. The stability of compact objects was extensively studied, with reference to notable contributions by Chandrasekhar [86, 87], who laid the groundwork for understanding dynamical instability in compact stars. His work provides key insights into the limits of stability for stellar objects, including NSs and white dwarfs. Maintaining the spherical symmetry of the background, Chandrasekhar performed the dynamical stability using the theory of infinitesimal radial perturbations [86]. The adiabatic index (γ) is given by
$ \gamma \equiv \left(1+\frac{\rho}{p_r}\right)\left(\frac{{\rm d}p_r}{{\rm d}\rho}\right)_S. $
(26) The expression (26) is associated with the sound speed, and the subscript S indicates the derivation at constant entropy. γ has some restrictions related with the dynamical instability of the spherical static object. For polytropic stellar model, authors in [88] have shown that the adiabatic index γ is greater than
$ \gamma> \gamma_{cr} = 4/3 $ , below which configurations are unstable against radial perturbations. For more details about the role of adiabatic index on stellar structure, see Ref. [89]. In Fig. 6, we display the dependence of γ for two considered cases as a function of radial coordinate r. Based on the figures, it can be concluded that$ \gamma > \gamma_{cr} $ holds for both cases, ensuring the stability of the configuration under consideration. -
The propagation of sound speed is another important criteria for checking the stability of QSs. The sound speed within the star is given by
$v^2_s = {\rm d}p_{r, \perp}/ {\rm d}\rho$ , and lies within the range of$ 0<v^2_s <1 $ because the speed of sound does not exceed the speed of light. Furthermore, the sound speed along radial direction is constant inside the star, whereas the tangential velocity has been plotted (both cases) in Fig. 7. As can be observed from those figures, the sound speed along transverse direction lies within the specified range. Thus, we can conjecture that QSs with anisotropic pressure can exist in$ f(R,L_m,T) $ gravity.
Anisotropic quark stars in ${\boldsymbol {f(R,L_m,T)} }$ gravity
- Received Date: 2024-10-04
- Available Online: 2025-02-15
Abstract: We investigated the impact of