Next Article in Journal
The Status of the Astrophysical Parameters of Upper Main Sequence Stars
Previous Article in Journal
The Use of Conditional Variational Autoencoders in Generating Stellar Spectra
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Gravity Wave Phase Shift in a Cold Quark Star with a Nonconvex QCD BZT Shock Wave Van Der Waals Equation of State

by
Keith Andrew
1,*,
Eric V. Steinfelds
1,2 and
Kristopher A. Andrew
3
1
Department of Physics and Astronomy, Western Kentucky University, Bowling Green, KY 42101, USA
2
Department of Computational and Physical Sciences, Carroll University, Waukesha, WI 53186, USA
3
Department of Science, Schlarman Academy, Danville, IL 61832, USA
*
Author to whom correspondence should be addressed.
Astronomy 2025, 4(3), 14; https://doi.org/10.3390/astronomy4030014
Submission received: 6 December 2024 / Revised: 13 July 2025 / Accepted: 14 July 2025 / Published: 22 August 2025

Abstract

We investigate BZT shocks and the QCD phase transition in the dense core of a cold quark star in beta equilibrium subject to the multicomponent van der Waals (MvdW) equation of state (EoS) as a model of internal structure. When this system is expressed in terms of multiple components, it can be used to explore the impact of a phase transition from a hadronic state to a quark plasma state with a complex clustering structure. The clustering can take the form of colored diquarks or triquarks and bound colorless meson, baryon, or hyperon states at the phase transition boundary. The resulting multicomponent EoS system is nonconvex, which can give rise to Bethe–Zel’dovich–Thompson (BZT) phase-changing shock waves. Using the BZT shock wave condition, we find constraints on the quark density and examine how this changes the tidal deformability of the compact core. These results are then combined with the TOV equations to find the resulting mass and radius relationship. These states are compared to recent astrophysical high-mass neutron star systems, which may provide evidence for a core that has undergone a quark gluon phase transition such as PSR 0943+10 or GW 190814.

1. Introduction

The study of the interacting dynamics of cold quark matter in beta equilibrium to understand a portion of the QCD phase diagram to describe the existence of compact stellar objects has grown considerably with the discovery of anomalous events such as the stellar remnant from the energetic hypernova SN 2006gy [1]. This event, caused by the sudden collapse of a high-mass star, gave rise to final state mass and radius values that can differ from standard compact stellar core models. The ongoing observational datasets now include the anomalous X-ray pulsar [2] RX J1856 [3,4] and pulsars such as PSR 0943+10 [5], PSR B1828-11 [6], PSR Jo751+1807 [7], PSR B1642-03, PSR J1614-2230 and PSR J0348-0432 along with potential, GRB candidates GRB 050709 [8] and GR B170817A [9,10] and magnetar candidates such as 1E 2259+ 586 [11,12], and gravity wave events [13,14] where masses, radii, dual luminosity peaks and cooling rates do not always give extraordinary fits to conventional models based upon degenerate neutron matter. Improvements in the observational data have resulted in more detailed models to help understand the varied mechanisms at play in dense QCD matter beyond the nuclear saturation limit of ~0.16 fm−3. New models [15] are gaining support from the observations of the Quark Gluon Plasma, QGP, demonstrating the existence of a high temperature, low chemical potential state of unconfined quarks as seen at SPS [16], LHC [17], and RHIC [18,19] laboratories. In recent years, the QGP data have expanded to include the determination of relativistic QCD-based transport coefficients [20] and measurements of the overall bulk viscosity for free quark matter [21,22]. Observations of X-ray, γ-ray, and quasi-periodic objects such as SAS J1808.4-3658 which has a 2.49 ms period [23] have been used to understand how the environment [22] and the ambient mass density impact fundamental QCD parameters and potentials [24]. Various EOS have been used linking the strong coupling constant and quark masses to the local density function using the real-time formalism and the Dyson–Schwinger gap equation at finite temperature [25]. In this way, high-density compact cores can be used to give the density-dependent strong coupling that can be expressed in a way that is suitable for EOS and phase transition analysis [26] with high chemical potentials and lower temperatures. The high-resolution data from the NICER satellite have extended and constrained the EoS modeling with X-ray spectra [27] for masses and radii [28].
For a cold collapsing star in beta equilibrium the overall charge, density, and chemical potentials are constrained by overall charge neutrality for particle number density nj, mass density ρf and electric charge qj for particle type j or flavor f: ∑qfnf = qene and ρB = ρu + ρd + ρs and three light quark flavor weak interaction equilibrium from quark interactions du + e + ῡe, su + e + ῡe, s + ud + u constraining the chemical potentials where the neutrinos and antineutrinos exit the core on a short time scale compared to the core formation process effectively causing their chemical potentials to vanish resulting in μd = μu + μe, μd = μs. The classical pressure is related to the chemical potential from P = −∂U/∂V = n2(ε/n)/∂n = nμ-ε where the chemical potential is μ = ∂ε/∂n. For the massless quark case this gives a direct comparison to the MIT bag model [29] with bag constant B where 3P = (ε4B) = ε4(4/4π) linking [30] the vacuum pressure bag constant to the quark chemical potentials which are energy dependent. Several compact core models have focused on modified EOS for various states of dense matter in the QCD ground state of Color Flavor Locked (CFL) dense matter [31,32] with 2-D and 3-D color superconductivity [33,34,35] and in the Nambu-Jona-Lasinio (NJL) model [36]. Many models make a comparison to versions of the generalized MIT Bag model [37] approach and include QCD mean field approximations [38,39] quarkyonic matter [40,41] along with computationally intensive numerical lattice approaches that are being improved to accommodate finite temperature chemical potentials [42]. Refined models include crustal effects [43], deformations [44], layering [45], cooling rates [46], strange quark cores, finite temperature boson stars [47] and kaon condensation [48,49] superconducting color vector potential effects [50,51], phase transitions, and different crystalline cores [52,53,54]. In many cases these EoS are everywhere convex, however, there are a number of cases where the pressure and density are so extreme that the EoS will become nonconvex. In the case of relativistic hydrodynamics [55,56] the energy momentum tensor [57] has a term proportional to the enthalpy which includes a factor of P/ρc2, or a factor of 4B/ρc2 for the MIT Bag model, which is normally quite small. At high pressures this term makes a larger contribution and, when asymptotic freedom is considered leading to free fast-moving particles, this term can be significant. At first this correction will drive the system closer to a stable point, but for high core pressures it can continue to push the system into an unstable region of shock waves that will alter the phase transition. Such phase transitions, occurring in the nonconvex region of the EoS, have been studied by Bethe [58], Zel’dovich [59] and Thompson [60] and these fluids are now known as BZT fluids. They developed a fundamental derivative test to identify when a fluid could develop BZT behavior. This has been used in the study of neutron stars by Aloy et al. [61] where they develop a method for measuring the BZT signature in gravitational waveforms [62,63]. The BZT phase transitions can have regions of spinodal decomposition leading to the formation of quark clusters and diquark states from the color states for color group SU(3)C  3 3 = 3 ¯ + 6 , where the antisymmetric color factor is attractive, C(F) < 0, and the symmetric is a weaker repulsive force, C(F) > 0. Here we will apply the multicomponent van der Waals model to investigate the hadron to quark phase transition in a compact stellar core in beta equilibrium with a BZT region identified by the fundamental derivative test. As noted by Ghosh [64] and Vovchenko [65] where they investigate the nuclear models and van der Waals [66] correspondence from the resonance hadron gas mode to lattice QCD, this model provides an analytical model with a critical point, it can model composite quark clusters as color singlets or net color carriers [67], it includes a phase transition, it can support shock wave behavior, it includes an excluded volume limit, the parameters are selected in alignment with lattice models, and can include both a repulsive or attractive strong force limit using color factor couplings depending upon the color group representation of SU(3)C. As such several authors have used the van der Waals (vdW) equation [68,69,70,71,72] of state (EoS) provides a useful analytical framework to describe the complex interactions within a hadronic medium under extreme conditions [73,74,75,76,77].
However, the MvdW EoS phase transition zone admits a nonconvex region that allows for BZT shock waves to form. When the BZT fundamental derivative becomes negative, the conditions for nonclassical anomalous wave formation exist. The BZT compression can lead to a rapid dispersion and an undefined speed of sound. In this case, the nonconvex MvdW EoS [78,79] can be used to describe the BZT phase transition as being spinodal. Due to the nonclassical behavior of shock waves in this region, spinodal clumping of mixed phases is expected to occur [80]. Israel has recently developed a relativistic generalization of these BZT fundamental derivative results, opening a new regime to explore. As a result, the compression can be purely classical, or a nonclassical BZT-type, or a nonclassical relativistic BZT-type. This nonlinear wave propagation and dispersion leads to a shift or reduction or phase shift in gravity wave production whose signature can be searched for using the Laser Interferometer Gravitational Wave Observatory (LIGO), the VIRGO Gravity Wave Interferometer, the Kamioka Gravitational Wave Detector (KAGRA), the Einstein Telescope gravitational wave detector (ET) or the Laser Interferometer Space Antenna (LISA) data or simulations. In general, shock waves in a BZT hadron fluid with a negative fundamental derivative could contribute to a QCD phase transition by creating nonclassical localized high-density regions conducive to quark formation facilitated by quark nuggets or noncolor singlet quark clusters. This could result in a broader and more easily sustained phase transition to a stellar quark core, leaving a unique gravity wave or luminosity signature that can be partially modeled by a MvdW EoS. This paper is organized as follows: the next section introduces the multicomponent van der Waals partition function and the various thermodynamic functions used to calculate the fundamental derivative and its relativistic correction with a special focus on the BZT regions, the following section models the gravitational wave signature of the BZT zone through the tidal deformability modeled from solutions to the TOV equation with the MvdW EoS, finally these are compared to the sensitivity range needed for detection, finally ending with a concluding section. We use units where G = c = k = 1 and the lower-case letter for thermodynamic variables that are specific densities per volume.

2. MvdW Partition Function and Fundamental Derivative

Here, we calculate the thermodynamic quantities needed to derive the BZT fundamental derivative for the MvdW EoS. Starting with the partition function, we will calculate the internal energy and enthalpy for our system. The MvdW partition function [81,82] for a total of Nc different components, where each component species can have Ni particles, each with a van der Waals volume of bk, with a mean field interaction coupling strength given by aij, in a total volume V, at temperature T, is given by:
Z v d w N i , V , T = i = 1 N C 1 N i ! V j = 1 N C N j b j λ i 3 N i exp N i V k T j = 1 N C a i j N j
where the ith particle species has a thermal de Broglie wavelength given by:
λ i = 1 2 π m i T
For our applications, the interaction couplings will obey a mixing rule where aij is symmetric and can be considered as the product of the individual particle couplings ai where we use the interaction parameter kij expressed as:
a i j = a i a j 1 k i j = a j i
In general, this expression plays the role of a mixing rule for different components where the dimensionless kij gives a binary interaction parameter as in Keffer [82], and several chemical examples can be found in Stryjek [83]. Different authors have used this term to introduce color force asymmetry, flavor structure differences between quarks, or to model hadronic diquark and gluon interactions. The kij term can be modeled as a (u, d, s) 3 × 3 matrix with diagonal terms reflecting stronger QCD interactions and with the off-diagonal terms of varying interaction strengths. In this context, the aij can represent the QCD Cornell potential, giving the direct channel strength, and the kij gives the indirect and nonsinglet mixing strengths for different diquark pairs. For example, in composite systems with diquarks or hexaquark states, the (u, d, s) system kuu = kdd = kud = 0.2, while kus = kds = 0.4 and in the CFL color (r-red, g-green, b-blue) state kbr = 0.1 when krg = 0.3 from gluon color suppression in states of (r, g, b) [84,85]. We see that as kij increases, the effective interaction strength decreases, thereby softening the EOS. Due to this effect, we will set kij = 0 for our calculations. The values of the aij terms can reflect the different color factor interactions and can represent attractive or repulsive interactions. The pressure can be expressed in terms of particle species number Ni or number volume density:
p = T ln Z V N , T = T i = 1 N c N i V j = 1 N c N j b j N i V 2 T j = 1 N c N j a i j    p = i = 1 N C n i T 1 j = 1 N C n j b j i = 1 N C j = 1 N C n i n j a i j .
The internal energy is given by:
U = k T 2 ln Z T N , V = T j = 1 N C N j 1 V i = 1 N C j = 1 N C a i j N i N j
and the entropy per particle is given by:
S p = S 0 + i = 1 N C N i N ln V p b m i x Λ i 3 i = 1 N C N i N ln N i N
Which will be used to express the pressure and the classical enthalpy in terms of entropy and volume to evaluate along an entropic curve, and the enthalpy volume density h, is:
h = u + p V = 3 2 N V T 2 i = 1 N C j = 1 N C a i j N i V N j V + T N / V 1 j = 1 N C N j V b j
We simplify these expressions by introducing a scale factor based on the total number of particles, N, for each particle species number fraction xi, the total van der Waals mixture volume of the particles, bmix, and mixture interaction couplings, amix, for N particles with specific heat cV, with reciprocal δ:
N = k = 1 N C N k , x i = N i N , V p = V N , b m i x = k = 1 N C x k b k , a m i x = j = 1 N C k = 1 N C x j x k a j k , n i = N i V , δ = 1 c V ,
where the specific heat cV is related to the number of degrees of freedom for the ith particle species. Using these variables, the multicomponent van der Waals EoS and specific entropy and energy density ε become:
p ( V , T ) = R T V p b m i x a m i x V p 2 , ε = R T δ a m i x V p + ε 0 , s = R ln V b m i x ε + a m i x V 1 δ + s 0
where s0 and ε0 are integration constants set to reference values, for the cold beta equilibrium quark star. Changing variables from temperature and volume to specific entropy and volume we have:
P T , V P ( s , V ) = δ exp δ s s o R V b m i x δ + 1 a m i x V 2
The first classical fundamental derivative GCl is by definition:
G C l = 1 2 V 2 P V 2 s P V s
where the multicomponent van der Waals classical BZT first fundamental derivative along an isentrope is given by:
G C l = δ + 1 δ + 2 V 5 P + a m i x / V 2 6 a m i x V V b m i x 2 2 V 4 δ + 1 V b m i x P + a m i x / V 2 4 a m i x
This gives the standard value in the limit of a single component and, when taking the ideal gas limit, we have a value of GCl = 1 + δ/2 in agreement with the convex ideal gas law, in the limit of a single component we get the standard MvdW value, in the limit of no excluded volume we get the simplified ideal gas volume dependence and in the limit of no interactions with the excluded volume we get the interacting gas expression [86,87]. The locus of points (P, V) for G = 0 corresponding to the boundary between the classical and nonclassical regimes is found by solving for the pressure to obtain:
P ( G = 0 ) = a m i x V 2 6 δ + 1 δ + 2 1 b m i x V p 2 1 = n 2 a m i x 6 1 n b m i x 2 δ + 1 δ + 2 1
The relativistic fundamental derivative is defined from the energy momentum tensor T μ ν and the specific enthalpy h. For fermions using the Fermi-Dirac distribution function f(p) for particles of mass m and Fermi momentum pF and the classical speed of sound squared c C l 2 defined along an isentrope for a system of total number density n, this is given by:
     T μ ν = n h u μ u ν + p g μ ν , c c l 2 = P n S , where n h = g 2 π 3 p 2 + m 2 f ( p ) d 3 p i , j a i j n i n j + i n i T 1 i b i n i i , j a i j n i n j or      n h = g 8 π 2 p F p F 2 + m 2 2 p F 2 + m 2 m 4 ln p F + p F 2 + m 2 m i , j a i j n i n j + P .
For a cold star in beta equilibrium composed of nondegenerate fermions along an isentrope we use the classical speed of sound squared and the specific enthalpy to calculate the special relativistic contribution to the fundamental derivative Gsr, and add this to the classical fundamental derivative GCl from Equation (12) to obtain the total fundamental derivative GTotal, this results in;
n h = 2 π 2 / 3 n 4 / 3 2 i , j a i j n i n j G s r = 3 2 c s r 2 = 3 2 h P n s G T o t a l = G c l + G s r = G C l 3 c c l 2 2 h G T o t a l = G C l 6 δ + 1 P + n 2 a m i x + 12 n 2 a m i x 2 1 n b m i x π 2 / 3 n 4 / 3 n 2 a m i x .
Equation (15) can provide better treatment for the high-density behavior found in compact stellar cores, for equilibrated neutron stars well below the Fermi energy the temperature terms are very small and can be neglected for modeling, and the multicomponent nature of the MvdW EoS will allow us to explore the impact of having several particle species in the core where both attractive and repulsive interacting terms that can be tested using the interaction potential terms in the aij and kij factors. The relativistic term to the fundamental derivative value can shift a region that was an otherwise convex classical region into a new nonconvex BZT spinodal decomposition mixed fluid zone, thereby altering the phase change dynamics of the core. The clumping that develops from this phase change leads to increased nonclassical shocks and rarefactions whose signature can be apparent in the shift in the phase peak of the gravity wave spectrum. For positive classical fundamental derivatives, the relativistic correction can lower the value of the fundamental derivative closer to zero, thereby smoothing out fluctuations that could give rise to instabilities, as long as it does not become negative. In Figure 1, we plot several special cases of the BZT fundamental derivative and the relativistic correction for the multicomponent van der Waals EoS, where the negative regions correspond to the rarefaction shock wave zones.
To numerically evaluate Equations (12)–(15), we select the constituents to be point particles with three degrees of freedom, there are no degeneracies between states, and all densities are in units of nuclear saturation density. Nuclear saturation density, typically denoted ρ0 ≈ 2.7 × 1014 g/cm3 for mass density or n0 ≈ 0.16 fm−3 for number density, represents the equilibrium density of symmetric nuclear matter inside atomic nuclei, where the attractive nuclear forces and short-range repulsion (including Pauli exclusion effects) are balanced. It serves as a fundamental reference point in compact star modeling, particularly for distinguishing the outer hadronic layers from the ultra-dense quark core in hybrid or pure quark stars [88]. In quark star density modeling, saturation density marks the threshold above which nucleonic degrees of freedom may dissolve into deconfined quark matter, depending on the stiffness of the equation of state and the nature of the phase transition (e.g., Maxwell, Gibbs, or BZT-type). Most quark star models—especially those incorporating a mixed phase—are calibrated to begin deviations from purely hadronic behavior at or just above n0, making it a key boundary for initiating BZT-type transitions with rarefaction shocks.
The values for the MvdW parameters of amix and bmix are heavily constrained from quark modeling of hadrons. The attractive interaction strength amix reflects the effective color-mediated coupling between different quark flavors and is inspired by SU(3)C Casimir scaling for gluon exchange in the one-gluon exchange approximation. This allows us to encode the fact that interactions among red, green, and blue color charges are not identical but can be averaged over pairwise channels with weights tied to the antisymmetric color-singlet or symmetric color-octet configurations. In the present model, we start with the hadronic values 0.2 < amix < 0.6 GeV−2 fm3, consistent with the effective coupling strengths used in Vovchenko et al. [65,71], who developed a van der Waals extension for Hadron Resonance Gas (HRG) and quark hadron crossover models. The extension of the interaction strength to the high core densities here is realized through the running coupling constant, which, for QCD, reduces the strong interaction force via asymptotic freedom. The repulsive excluded volume term bmix fm3 represents the effective finite size of quark composites or effective repulsion from Pauli blocking and short-range gluon interactions. In line with recent work by Ghosh [64,89], who explored van der Waals quark matter in hybrid EoS construction, we take 0.7 < bmix < 2.5 fm3, corresponding to an effective hard-core quark radius of 0.3–0.8 fm. In this sense, l bmix encodes repulsive interactions similar to a scattering impact parameter and is related to the radius of the underlying baryonic particle by 3 bmix = 16πr3. These choices are not arbitrary but lie within the parameter space explored in lattice-constrained hadron and quark matter simulations, providing a realistic interpolation between nuclear saturation and deconfined quark phases. Restrictions also come from the high-density fermion chemical potentials which have a more natural setting in the MIT bag model and the NJL extended model. The case of net color diquarks and larger quark clusters can extend the range of amix and bmix beyond the standard baryonic domain.

3. Peak Gravity Wave Shift Signature of BZT Shocks and Chirps

The spinodal decomposition during the phase change results in anomalous dispersion and shock wave formation that will alter the gravity wave signature of a wave-producing event with a BZT compact stellar core. The gravitational wave amplitude varies as h(f)~(Mch)5/3 f−7/6/D, for chirp mass Mch defined in Equation (18), frequency and distance D, these changes are currently beyond detectability, however the phase shift for a BZT EoS is closer to the range of detectability so we will analyze it more closely. We determine the mass and radius values from the TOV equations [90,91,92] applied to a static spherically symmetric mass:
d P d r = G m ρ 1 + P ρ 1 + 4 π r 3 P m r 2 1 2 m r d m d r = 4 π r 2 ρ
which we solve numerically for the MvdW EoS for different values of the parameters amix, bmix, kij, and Nc with plots presented in Figure 2. To solve the Tolman–Oppenheimer–Volkoff (TOV) equations for relativistic stellar structure, the Mathematica function NDSolve is employed with an adaptive numerical integration scheme [93,94]. This method transforms the coupled differential equations for mass and pressure—governed by the energy density profile and the equation of state (EOS)—into a numerically tractable system. The integration begins at a small initial radius with a specified central pressure and proceeds outward until the pressure drops to zero, marking the star’s surface. NDSolve automatically selects a suitable integration method, typically using an embedded Runge–Kutta scheme for non-stiff regions and switching to backward differentiation formulas (BDF) in stiff regimes, as needed. The solver’s adaptive step-size control ensures that local truncation errors remain within tolerance, adjusting the step size dynamically based on the complexity and steepness of the solution. This approach is essential for capturing the nonlinear behavior of the TOV system, particularly in the presence of realistic equations of state that may include phase transitions or nonconvex features such as BZT regions. The result is a smooth and accurate profile of the star’s internal structure, including total mass and radius, across a wide range of central densities and EOS parameters.
The adiabatic index Γ, serves as a critical diagnostic tool in modeling the equation of state (EoS) for compact stars, bridging microphysical interactions with macroscopic stability and observational signatures. In all the above studies, Γ is employed to probe the stiffness of the EoS and the response of matter under compression, with its value determining the onset of dynamical stability, phase transitions, and wave propagation behavior. Moustakidis [95] uses Γ to define a critical threshold for the stability of relativistic stars, linking it analytically to compactness and central pressure via Chandrasekhar’s variational method. Casali and Menezes [96] show how Γ decreases with the emergence of exotic particles like hyperons or strange quarks, indicating EoS softening and potential instability. Carney et al. [97], while not analyzing Γ explicitly, rely on its underlying structure in reconstructing the EoS from gravitational wave data, showing that smooth variations in Γ are essential to accurate spectral fits. Here we take a more dynamic view, showing how a drop in Γ—due to strong interparticle attraction, via amix, and limited excluded volume, via bmix, in the multicomponent van der Waals EoS—can trigger nonconvex behavior, Bethe–Zel’dovich–Thompson (BZT) shocks, and gravitational wave phase shifts. Across these works, the adiabatic index functions as both a stability threshold and a marker for identifying nonlinear phenomena, making it indispensable for connecting theoretical models to observable astrophysical signatures. From the MvdW EOS, the adiabatic index is given by:
         Γ = ln P ln ρ > 4 3 ln P ln ρ = 1 N c n b m i x N c T 1 + N c n b m i x 2 N c 2 a m i x n 1 N c n b m i x 2 N c T a m i x N c 2 1 N c n b m i x > 4 3
Equation (17) is used to constrain the values of amix, bmix, and n to be in a range that leads to a stable configuration. Expressed in terms of the number of components Nc, the particle number density n, the interaction strength amix, the excluded volume bmix which behaves like a repulsion under extreme compression, and the temperature T. The BZT shock zone from Equation (17) is plotted in Figure 2. The system becomes more stable as amix is decreased and bmix is increased. Increasing amix for higher densities mimics the asymptotic freedom effects seen in short-range QCD. Models with amix > 0 and bmix > 0 exhibit monotonic increases as the density increases. However, the BZT fundamental derivative GTotal is the more general diagnostic: it includes the adiabatic index and its variation and is required to determine whether the fluid dynamics will exhibit nonclassical features like rarefaction shocks or spinodal clumping, particularly relevant near phase transitions or in nonconvex EOS such as those arising in dense QCD matter. Meanwhile, Γ remains essential for basic stellar stability analysis but is not sufficient for capturing the full richness of nonlinear or anomalous wave behavior.
In general, the symmetry of the static and spherically symmetric stellar core would be broken by introducing rotation and magnetic fields which would be present in a collapsed stellar core. Including a magnetic field and the spin of the star alters the standard Tolman–Oppenheimer–Volkoff (TOV) solutions, which assume spherical symmetry and hydrostatic equilibrium in a non-rotating, unmagnetized compact object. Rotation breaks spherical symmetry and introduces centrifugal forces that counteract gravity, leading to a flattening of the star at the poles and an increase in equatorial radius. This deformation modifies the pressure gradient and mass distribution, requiring a generalization of the TOV equations to axisymmetric solutions, such as those described by the Hartle–Thorne approximation [98,99] or fully relativistic rotating star codes (e.g., RNS, https://github.com/cgca/rns, accessed on 2 May 2023). Magnetic fields, particularly those above 1015 G as found in magnetars or post-merger remnants, introduce anisotropic pressures due to the Lorentz force and the field’s stress-energy tensor. The pressure becomes direction-dependent, with parallel and perpendicular components relative to the field lines, breaking isotropy and modifying the equilibrium structure. Moreover, strong magnetic fields can stiffen or soften the equation of state depending on orientation and particle magnetic moments, and they couple to the stellar currents and composition via Landau quantization and anomalous magnetic moment (AMM) effects [100,101]. Together, rotation and magnetic fields can increase the maximum mass and radius of the star compared to static TOV predictions, influence the star’s moment of inertia, and affect gravitational wave signatures from oscillations or mergers. Here we first solve the static spherically symmetric case and look to add the impact of these terms in the future.
The Tolman–Oppenheimer–Volkoff (TOV) equations describe the hydrostatic equilibrium of relativistic stellar structures and yield distinct mass–radius solutions depending on the underlying equation of state (EOS). When the EOS is modeled using the multicomponent van der Waals (MvdW) form, parameterized by the interaction strength aij and excluded volume bk, each combination of aij and bk produces a unique mass–radius curve with a corresponding maximum mass Mmax and radius Rmax. These solutions can be characterized by the dimensionless compactness factor C = M/R, which quantifies the relativistic strength of the star’s gravitational field. As the central density increases beyond the point where Mmax occurs, the solutions become unstable to radial perturbations, marking the onset of the classical TOV instability [102]—a purely gravitational instability tied to the loss of equilibrium. This behavior is fundamentally different from the BZT (Bethe–Zel’dovich–Thompson) shock-induced instability, which arises from nonconvex regions of the EOS where the fundamental derivative becomes negative, allowing anomalous shock propagation and spinodal decomposition. By systematically varying aij and bj, one can map out both the gravitational TOV instability boundary and the separate BZT instability zone, allowing a deeper understanding of how microphysical interactions in the EOS govern both equilibrium structure and dynamical behavior in compact stars, as shown in compactness and observational survey plot in Figure 2.
The key implications for gravitational waves due to nonconvex QCD BZT shock waves during phase transitions lie in their ability to produce observable phase shifts in gravitational waveforms from binary mergers or neutron star oscillations. Nonconvexity in the equation of state (EoS), quantified by a negative fundamental derivative GTotal < 0, leads to anomalous wave dynamics, including compound rarefaction-compression shock structures. These features modify the internal density and pressure profiles of the star, altering the tidal deformability and gravitational wave phase evolution—effects that are potentially measurable by detectors such as LIGO, KAGRA, or LISA.
Figure 1, Figure 2 and Figure 3 were generated using a multicomponent van der Waals (MvdW) EoS with parameters 0.4 < amix < 0.7 GeV−2 fm3 and 0.5 < bmix < 2.0 fm3, ranges motivated by lattice QCD analogs and color-mediated quark interactions. These values are realistic within the context of high-density QCD matter, where short-range repulsion (encoded in bmix) and long-range attraction (through amix) emerge from gluon exchange and color confinement physics. The phase shifts shown (on the order of Δϕ∼0.002–0.03 radians) align with the detection thresholds of current and next-generation gravitational wave observatories. These calculations use a semi-analytic shock model and EOS-driven TOV integration to simulate both gravitational and gamma-ray burst signals, allowing a connection between microphysical nonconvexity and observable astrophysical signatures.
In Figure 2, we show numerical solutions exhibiting the mass and radius limits. We also show two candidate pulsars: J1614–2230 and J1909–3744 (Appendix C reference) which have estimated masses near 1.5 and 2.0 solar masses. The waveform shift is influenced by the tidal deformability, Λ, the compactness C, and the second Love number k2, where compactness values are determined from the TOV equations for the MvdW EoS for different particle content, and for a merging binary system the system compactness scales with the chirp mass, the merger tidal deformability depends on the individual masses and tidal deformability and the gravity wave phase shift can be expressed in terms of the chirp mass and merger deformability as reviewed in Baiotti [103] and presented in Vines in post Newtonian form [104] as
C j = M j R j , Λ j = 2 3 k 2 C j 5 , M c h = M 1 M 2 3 / 5 M 1 + M 2 1 / 5 , Λ ˜ m e r g e r = 16 13 M 1 + 12 M 2 M 1 4 Λ 1 + M 2 + 12 M 1 M 2 4 Λ 2 M 1 + M 2 5       k 2 M v d W = 8 5 C j 5 1 2 C j 2 1 + a m i x C j + b m i x C j 2 ,          Phase-Shift : Δ Ψ = 3 128 π M c h f 5 / 3 Λ ˜ m e r g e r
where the phase shift factor Δ Ψ , we are using is the largest and first term of the series as presented in Vines.
The compactness and tidal deformability depend on the TOV masses and radii found for each EoS, allowing for a comparison between the various models, where several cases are presented in Figure 3. The Love number decreases with density, and the tidal deformability decreases with compactness. For the MvdW EoS, the stronger interactions for a > 0 and smaller number of components give a higher compactness and stiffness, resulting in less deformability, while a lower excluded volume, b << 1, leads to a greater tidal deformability.
Typically, the quark star has a higher compactness due to the increased density, leading to a lower tidal deformability when compared to a neutron star prior to the phase transition. The range of values is presented below in Table 1.
The wave phase shift depends upon the EoS and the chirp mass; for the BZT region, we plot this change in Figure 3. Overall, the stiffer EoS gives a smaller shift due to reduced tidal effects with high compactness which is most pronounced at low frequencies. In the MvdW EoS, this corresponds to the fewest components with the largest attractive force, a > 0, and smallest excluded volumes, b << 1. Likewise, higher chirp masses lead to greater compactness and smaller phase shifts. Each model EoS gives a phase shift in accordance with the associated stiffness of the model.
Figure 3 serves to demonstrate that the largest phase shifts arise in low-compactness, high-deformability regions and that several GRB-associated neutron star models (e.g., GRB170817A, GRB221009A, GRB150101B, references are in Appendix C) occupy regions near or within the BZT zone. The inferred values of amix and bmix for these events are determined via matching gravitational wave phase shift, compactness, and tidal deformability observations to EOS model predictions. The alignment of observationally favored models with the BZT region underscores its potential relevance to real astrophysical systems.
The BZT regions reduce the energy going into the compression regions, thereby altering the coexistence energy distribution with anomalous dispersion and resulting in the peak energy to be delayed and frequencies to be shifted to lower values. A coherent picture emerges when Figure 2, Figure 3 and Figure A2 in Appendix A are considered together. Figure 2 demonstrates that the MvdW EoS supports compactness values consistent with known neutron star observations, such as PSR J0740+6620 and J0030+0451, thereby validating the model’s structural realism. Figure 3 builds on this by showing that within the same parameter space, especially in the nonconvex BZT region, the MvdW EoS predicts gravitational wave phase shifts in the range detectable by current and next-generation observatories. These shifts are tied to tidal deformability and are especially relevant for GRB-associated neutron star mergers. Appendix A and Figure A2 reinforces the significance of the MvdW framework by comparing it to other EoS models, illustrating that it not only provides the strongest and broadest BZT shock wave structure, but also uniquely sustains a sizable energy discontinuity across the phase boundary—critical for modeling the internal dynamics and luminosity evolution of post-merger or hybrid PSR systems. Together, these figures establish the MvdW EoS as a reasonable candidate for describing some of the observable consequences of QCD phase transitions in compact stars.
For comparison, in Figure 4, we use modified BZT versions of the following models: the MIT bag model, the Hadron Gas model, the NJL model, and a lattice model. To introduce BZT regions we include interaction terms: in the MIT model we use the standard extension of a van der Waals like interaction with a bag constant [105,106], in the Hadron Gas Model the interaction term includes a field that repels and attracts [107], the NJL model has a BZT region for the chiral symmetry breaking that is a part of the model [108,109], the lattice model uses a perturbative QCD term that drives the pressure derivatives negative [110,111]. While it is possible to produce BZT regions in each model, the impact varies in extent and significance. In the modified MIT bag model, the BZT region exhibits a strong dependence on the magnitude of the interaction parameter, in the Hadron Gas model the BZT region only appears near hadronic saturation density and is highly localized, in the NJL model BZT effects are prominent and impactful in the chiral transition region, in lattice models BZT regions only occur near the critical point at high chemical potentials. The overall trend is similar for each BZT region, with a merging of all values for high chirp mass. We find that the MvdW EOS results in the largest phase shift ΔΨ, while the Lattice model produces the smallest shift, see Appendix A for details. A Bayesian Inference Corner Plot for the MvdW EOS parameters with posteriors is given in Appendix B, showing the parameter correlations.
In Figure 5, we compare the values from Table 2 [113] with the phase shifts expected in Figure 4 with detector sensitivities to identify the thresholds for detection. The LIGO [114] -VIRGO [115] -KAGRA [116] detectors have phase shift sensitivities ~0.02 radians which would require MvdW parameters of a > 4 MeV fm3 and b < 0.01 fm3. These values do not result in a likely scenario at the required high densities, given that QCD forces exhibit asymptotic freedom at short distances. The upgraded Einstein Telescope [117] requires a~1.2 MeV fm3 and b~0.04 fm3 to produce observable phase shifts; these values are in the range expected during a BZT phase transition and could lead to a measurable result. The future LISA [118] detector should be in this range and could constrain the values used for the MvdW EoS.

4. Conclusions

We have used the nonconvex MvdW EoS to investigate the impact of the nonconvex region on the phase shift of a gravitational wave event and a GRB luminosity model to search for a detectable BZT EoS signal. In typical neutron matter, classical compressive shock waves tend to dissipate energy by heating the fluid; such action might disrupt and prevent a full QCD phase transition unless the conditions are extreme. When the fundamental derivative is negative, the compression waves will not steepen into localized shocks, but rather the waves tend to disperse and rarify upon compression producing QCD cavitation, this nonlinear anomalous shock propagation alters the phase transformation and can produce unstable metastable states very sensitive to the thermodynamic state, especially near any critical points. This behavior creates unique local variations in pressure and density which alter the conditions needed for a phase transition. This results in a coexistence region with pronounced clumping with quark clusters, nuggets, and diquark states, which can be metastable, and alternate with volumes of quark gluon plasma states with deconfined quarks. This behavior leads to potentially unusual and spatially complex phase behaviors not observed in fluid with a positive fundamental derivative. However, in a BZT neutron fluid the nonclassical shock dynamics will mean that localized energy and density fluctuations could sustain a phase transition at lower average energy levels than in standard neutron matter. This could result in a relatively low energy path to quark matter formation compared to the direct collapse model assumed for many cases. In our study, we have used the TOV equations for a spherical mass with no magnetic field, and there is no rotation; these are modifications we are currently exploring. In addition, the MvdW EoS has a natural limit to a polytropic star which provides a canonical system to investigate in parallel with the MvdW EoS. There is also extensive work on introducing the Maxwell construction to remove the nonconvex region while having the speed of sound vanish, due to no pressure gradients, while others favor the Gibbs construction to maintain chemical potentials. Both methods change the dynamics of the coexistence phase. In a similar way, one can also use an infrared regulator in the interaction term to reduce or remove the nonconvex region. We are also implementing a density-dependent interparticle force to explore a change in the van der Waals luminosity function to better match a typical GRB curve. The BZT rarefaction shocks will also modify the thermal conductivity and heat transport, thereby affecting the star’s cooling behavior. The shocks will impact neutrino production in the direct Urca process: n p + e + ν ¯ e p + e n + ν e and the modified Urca process: n + n p + n + e + ν ¯ e p + n + e n + n + ν e where the edge density gradients should enhance neutrino production and alter neutrino emissivity, impacting potential kaon condensates. Rarefaction shocks would also reduce thermal insulation and alter the heat transport to the surface, causing a higher cooling rate and a lower surface temperature, perhaps similar to what is observed at Cas A [119].
Here we have used the MvdW EoS to examine the fundamental derivative indicating the onset of nonconvex behavior and looked at how this would induce a shift in the phase of a gravity wave-producing event. The multicomponent van der Waals (MvdW) equation of state (EOS) offers a versatile phenomenological framework for modeling dense quark matter, particularly in the context of nonconvex phase transitions relevant to neutron star interiors and gamma-ray burst (GRB) engines. However, its utility is tempered by several theoretical and practical limitations. Most notably, the MvdW EOS is not derived from first-principles QCD and lacks a direct connection to confinement dynamics or chiral symmetry breaking, which are central features of the QCD phase diagram. The model encapsulates inter-quark interactions using simplified scalar terms—namely, an attractive interaction coefficient aij and an excluded volume bj—which do not account for the full color, spin, or momentum-dependent structure of QCD interactions. Furthermore, thermodynamic inconsistencies can arise for certain parameter choices, including violations of causality due to superluminal sound speeds or regions of negative compressibility, unless constraints are carefully imposed. The model is typically formulated at zero temperature, omitting thermal corrections that are critical in proto-neutron stars and post-merger environments, and does not incorporate the effects of strong magnetic fields or rotation, both of which are known to significantly impact the structure and stability of compact stars. Moreover, the absence of color superconducting phases such as the color-flavor-locked (CFL) state limits the MvdW model’s applicability at asymptotically high densities. Despite these limitations, the MvdW EOS remains a useful tool for exploring BZT-type nonconvex behavior and associated phenomena such as compound shock formation, especially when coupled with relativistic stellar structure calculations and gravitational wave observables.
We then examined if this could produce a measurable signature indicating the presence of internal BZT structure. We found that the formation of unusual nonclassical shock waves in a BZT hadron fluid with a negative fundamental derivative could contribute to a QCD phase transition by creating localized, high-density regions, conducive to quark gluon plasma formation along with adjacent BZT low-density rarefaction regions providing a possible means for a sustained novel phase transition. This unique behavior could facilitate the formation of quark nuggets or quark clusters, which, under the right conditions, could lead to a broader phase transition, potentially transforming the stellar core into a quark star without the need for ever higher pressures. As such, the unusual BZT properties of a hadron fluid could provide a unique mechanism by which extreme astrophysical conditions might initiate and sustain a phase change to quark matter. The formation of such an object could leave a signature in the gravity wave phase shift that could best be observed in a future gravity wave detector such as LISA or the Einstein Telescope. We are continuing to improve on this model by adding magnetic field effects, anisotropic transverse and radial pressures, coupling with an anomalous magnetic moment, and exploring the impact of nonspherical shapes.

Author Contributions

Formal analysis: K.A., E.V.S. and K.A.A.; methodology: K.A., E.V.S. and K.A.A.; investigation: K.A., E.V.S. and K.A.A.; formal analysis, numerical solutions, and visualization: K.A., E.V.S. and K.A.A.; writing original draft: K.A.; writing review and editing: K.A., E.V.S. and K.A.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Acknowledgments

The authors wish to thank Western Kentucky University and Schlarman Academy for their kind support throughout the work on this project. All plots and numerical solutions were generated in Mathematica. The authors also wish to thank the anonymous referees who added significantly to improving this manuscript.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A. Models for Gamma-Ray Curves

To investigate the relative magnitude of the MvdW BZT noncovex region, we compare the BZT fundamental derivative values to similar modified EoS used to describe high-density phase changes. We wish to connect the microphysical properties of the interacting particles of the EOS to the observational data-based curves from astronomers. In many cases a convex EoS can be rendered nonconvex with the addition of an interaction term similar to the aij interaction in the MvdW EoS. However, both the phase-changing lattice model and the symmetry breaking chiral model experience negative fundamental derivatives during their respective phase changes, and we report them as they arise in special applications similar to the one examined here. The Hadron Gas Model and the NJL model both introduce phase changes that result in a negative fundamental derivative for the quark hadron phase change and for the color-flavor-locked superconducting phase change, and we use those values for comparison. For the MIT bag model, we introduce a quark interaction term of the form aijninj like the MvdW EoS, which can be positive or negative depending upon the quark color factor. In this sense, we can compare the magnitude of the nonconvex regions to the MvdW case. Here we give the EOS, energy density, BZT energy and initial luminosity for a shock time scale of τ = 0.5 s and a burst volume V = 1018 cm3 for numerical parameter values with some variation we selected points near the midpoint and mass values are from the Particle Data Group [120], where the MvdW model uses quark constituent masses but the NJL model uses quark dynamical masses and the conditions for beta equilibrium are used in assigning values to the chemical potentials.
The MvdW system is given by [64,121]:
P = j n j T 1 b j n j i , j a i j n i n j    ε = j 3 2 n j T 1 b j n j + m j n j + i , j a i j n i n j    m u = 0.225 G e V    m d = 0.225 G e V    m s = 0.500 G e V    a m i x = 0.4 G e V 2 f m 3    b m i x = 0.25 f m 3    n i = 2.5 n 0
M v d W : Δ ε B Z T = 200 M e V f m 3 L 0 = 6.4   ×   10 53 e r g / s .
The MvdW model has a wide BZT zone that supports shock-driven bursts, it produces high L0, has a broad, slowly decaying GRB luminosity curve, requires careful adjustment to avoid causality violations, EOS may be too soft without constraints and it is a strong match to broad, high-luminosity GRBs like GRB 090618 and GRB 061007.
The MIT bag model with BZT level interactions is given by [122,123,124]:
P = f μ f 4 4 π 2 B i , j a i j n i n j ε = f 3 μ f 4 4 π 2 + B + i , j a i j n i n j    Beta   Equilibrium    μ u = 325 M e V    μ d = μ s = 450 M e V    B = 250 M e V / f m 3
M I T : Δ ε B Z T = 150 M e V f m 3 L 0 = 4.8   ×   10 53 e r g / s .
  • where the MIT bag Model with interactions has a harp phase transition and can produces strong Δε, has a simple analytic form, can produce a moderate-to-high L0, can only support a BZT region if interactions are present, the luminosity curve is symmetric and less realistic without extension, it matches mid-width GRBs like GRB 080319B.
The NJL model is given by [125,126]:
P = M m 0 2 4 G S + 1 π 2 0 β p 2 d p E p μ θ μ E p ε = M m 0 2 4 G S + 3 π 2 0 p F p 2 d p E p    G S = 5.5 G e V 2    Λ c u t o f f = 631 M e V    B a r e : m 0 u , d = 5.5 M e V , m 0 s = 135 M e V    D y n a m i c a l : M u = M d = 335 M e V , M s = 540 M e V    F e r m i : p F f = μ f 2 M f 2 ,    for μ f > M f
N J L : Δ ε B Z T = 130 M e V f m 3 L 0 = 4.2   ×   10 53 e r g / s .
  • where the NJL model has a Δε driven by chiral symmetry restoration, naturally supports color superconducting extensions, produces long tails and plateaus in the light curve, has an L0 competitive with the MvdW model, matches GRBs with dual-phase decay (GRB 090618, GRB 080319B) see Appendix C for references.
The Hadron Gas Model (HGM) [127,128] in the low-temperature Boltzmann form, using modified Bessel functions of the first and second kind K1, K2, and restricted to the hadrons listed, where the thermodynamic contributions of each hadronic species are determined by a combination of statistical and thermal factors. The degeneracy factor gi represents the number of internal quantum states available to a species, typically incorporating spin and isospin multiplicities, gi = (2si + 1)(2Ii + 1) where si is the spin and Ii is the isospin of the hadron. This term scales the influence of each particle on the total pressure and energy density, with higher-spin or multi-state particles contributing more significantly. The modified Bessel functions K2 and K1 arise from the integration over momentum space in the partition function and play a central role in encoding the thermal behavior of massive particles. The function K2 dominates the pressure and energy contributions by capturing the Boltzmann suppression of heavy species, scaling roughly as e(−mi/T) for Tmi >> T. Meanwhile, K1, which appears in the expression for energy density, provides a relativistic correction that accounts for the full energy content of the particles, including both kinetic and rest mass components. Together, these terms ensure that the HRG model accurately reflects the statistical mechanics of a multicomponent hadronic system in thermal equilibrium and captures the transition from relativistic to non-relativistic behavior across species and temperature ranges. The HGM is given by:
P = i g i T 2 m i 2 2 π 2 K 2 m i T ε = i g i T 2 m i 2 2 π 2 3 K 2 m i T + m i T K 1 m i T    π 0 : m π 0 = 0.135 G e V , g = 1    π ± : m π ± = 0.140 G e V , g = 2    K ± : m K ± = 0.494 G e V , g = 2    p , n : m p , n = 0.938 G e V , g = 2    Λ : m Λ = 1.116 G e V , g = 2    Δ : m Δ = 1.232 G e V , g = 4
H G M : Δ ε B Z T = 85 M e V f m 3 L 0 = 2.7   ×   10 53 e r g / s .
  • where the HGM matches low-temperature lattice EOS, tends to favor a transition that produces short, sharp GRBs, in general Δε is small therefore L0 is limited, so it fails to model long-duration or broad-peak GRBs, it is best for short, faint GRBs (GRB 091127); does not naturally explain long GRB luminosity tails.
The QCD lattice model is given by [129,130,131]:
P = n = 0 N a 2 n T 4 + 2 n , ε = n = 0 N a 2 n 3 + 2 n T 4 + 2 n    a 0 = 0.9 ,    a 2 = 0.15 M e V 2 ,    a 4 = 0.02 M e V 4
L a t t i c e : Δ ε B Z T = 70 M e V f m 3 L 0 = 2.2   ×   10 53 e r g / s .
  • where the lattice model is based directly on a QCD first-principles EOS, stability and reproducibility are good for calibration and EOS matching, produces a small Δε and soft peak luminosity, no BZT region unless effective model interactions are added, can approximate GRB plateaus but underpredicts L0 for bright GRB events, the lattice model is most effective for events like GRB 111228.
Each dense matter model offers unique strengths and limitations in modeling quark star interiors and their connection to BZT shock dynamics and GRB luminosity profiles. The multicomponent van der Waals (MvdW) EOS provides a tunable framework that naturally supports nonconvex thermodynamic behavior through its attractive and repulsive interaction parameters, making it highly effective for modeling BZT-type shocks and generating strong, broad GRB luminosity curves. However, it lacks a direct derivation from QCD and may violate causality if not properly constrained. The MIT bag model, while analytically simple and widely used in hybrid star constructions, inherently lacks interaction terms and BZT structure unless explicitly modified; its GRB output tends to be more symmetric and shorter in duration. The Nambu–Jona-Lasinio (NJL) model captures essential QCD features such as chiral symmetry breaking and restoration, and it can produce BZT-like effects near the chiral critical point, giving rise to dual-phase or delayed GRB light curves. Its main weakness lies in the absence of confinement and its sensitivity to regularization. The Hadron Resonance Gas (HRG) model accurately reproduces low-density hadronic thermodynamics and matches lattice results below the crossover, but its limited ability to model phase transitions and lack of nonconvexity restrict its use in BZT shock applications, leading to fast-fading, less energetic GRB profiles. Finally, lattice QCD provides the most fundamental insight into the QCD equation of state at finite temperature and low baryon density, yet the sign problem hampers its application at neutron star core densities. It predicts smooth crossovers without strong shocks, producing brief and relatively featureless GRB signals unless supplemented with effective models. In sum, the ability of each EOS to generate observable BZT shocks and realistic GRB luminosity curves is directly tied to its capacity to support nonconvex thermodynamics and a robust first-order phase transition.
To model the gamma-ray burst (GRB) light curves associated with phase transitions in quark stars, we introduce a semi-analytic energy release framework based on the propagation of a shock front through a nonconvex Bethe–Zel’dovich–Thompson (BZT) region using the Rees Double Power Law Luminosity Function for GRB (also known as the broken power law) phenomenological model [132,133,134] with the Li Shock Breakout GRB Model for the BZT shockwave [135,136,137]. As the shock traverses the BZT layer, the local drop in the fundamental derivative G leads to a release of internal energy, which is partially converted into observable electromagnetic emission and structure changes that alter gravity wave production. The luminosity for each model is described using a standard dual gamma-distribution-shaped function of time:
L ( t ) = L 0 t t 0 a 1 , t < t b L 0 t b t 0 a 1 t t b a 2 , t t b
where L0 is the peak luminosity, tb is the characteristic rise time, and n is the shape parameter controlling the rapidity of onset. The two exponents a1 and a2 are used to match the observed GRB, as shown in Table A1 below. The associated measured L0 values are matched to the relative strength and spatial extent of the BZT region for each EOS model. To model gamma-ray burst (GRB) light curves from first principles, we apply the Rees broken power law luminosity function, parameterized as L0 is the peak luminosity, t0 is the initial normalization time, a1 and a2 are the early and late decay indices, and tb is the blast or break time marking the transition between emission regimes. We constrain L0 using the microphysical energy density discontinuity Δϵ predicted by various equations of state (EOS), including L0 ≈ Δϵ⋅V/τ, with V the emitting volume and τ the release timescale. For example, using EOS-derived values of Δϵ∼2 × 1035 erg/cm3 and τ = 0.5 s, the MvdW model yields L0∼4 × 1053 erg/s, consistent with Swift-XRT luminosity data for GRB 090618 (https://www.swift.ac.uk/xrt_curves/, accessed on 12 May 2024). We fit tb and decay slopes a1, a2 to the observed light curve, obtaining best-fit values tb∼4 × 104 s, a1∼0.8, and a2∼1.7, which match the observed transition from engine-driven plateau to afterglow phase. This approach bridges microphysical EOS modeling and GRB data, enabling direct EOS constraints from light curve morphology. For example, the multicomponent van der Waals (MvdW) model, which exhibits the broadest and most intense BZT zone, yields the largest L0 and a slower decay (larger t0), while the Hadron Gas Model—featuring only a narrow BZT window near saturation density—produces a sharp, lower-luminosity pulse. By selecting L0 in proportion to the integrated energy release, the Li Shock Model as applied here can be expressed as:
L 0 = E s h o c k τ = V B Z T Δ ε d V τ Δ ε B Z T V τ E = ε d V , 10 3 s < τ < 10 3 s , V 10 18 c m 3
where ΔϵBZT∼α∣ΔG∣ and scaling the time response to the shock transit time, this formulation provides a unified way to represent GRB signatures from different QCD phase transition scenarios described by an EOS, linking fluid nonconvexity to observable high-energy astrophysical transients.
We use the data from Ruffini and Izzo [138] for the smoothed and idealized GRB curves to match the model with parameters given in Table A1.
Table A1. GRB data and model fit from Ruffino and Izzo [138] to compare to the luminosity curves for the MvdW EOS.
Table A1. GRB data and model fit from Ruffino and Izzo [138] to compare to the luminosity curves for the MvdW EOS.
GRB IDNS Systema1 (Early)a2 (Late)
GRB 061007PSR J1907+09191.48 ± 0.062.42 ± 0.05
GRB 080319BPSR J0030+04511.43 ± 0.052.19 ± 0.04
GRB 090618PSR J0348+04321.60 ± 0.072.54 ± 0.06
GRB 091127PSR J0737–30391.79 ± 0.082.68 ± 0.07
GRB 111228PSR J1614–22301.66 ± 0.062.41 ± 0.05
Given the values of a1 and a2 from observational data the EOS is used to match the observed L0 using Δε and then plotted using the Rees Double Power Law Luminosity Function for GRBs, these are shown in Figure 4. these end up being smoothed curves with a clear phase change point that does not exhibit the observational variation and scatter associated with the real data. In Figure A1, we show a plot of some of the points for GRB 090618 with a Rees Double Power Law fit and the corresponding MvdW parameter fit in the amixbmix region.
Figure A1. The GRB luminosity curve (subset of points) fit to the Rees broken power law model light curve of GRB 010222 from Cowsik, et al. using the Indian Astronomical Observatory, Hanle, and the telescopes at the Vainu Bappu Observatory, Kavalur, with a broken power law fit to the data, where the data, figure and broken power law fit are all from the Cowsik group [112,139]. The underlying MvdW model then estimates a peak luminosity L0 = 4.2 × 1045 erg/s, a reference time t0 = 100 s, early-time decay index a1 = 0.542, late-time decay index a2 = 1.263, and a blast break time at tb = 3.71 × 104 s. Observational points are indicated by the markers. This GRB light curve can accommodate a MvdW solution as shown in the next figure.
Figure A1. The GRB luminosity curve (subset of points) fit to the Rees broken power law model light curve of GRB 010222 from Cowsik, et al. using the Indian Astronomical Observatory, Hanle, and the telescopes at the Vainu Bappu Observatory, Kavalur, with a broken power law fit to the data, where the data, figure and broken power law fit are all from the Cowsik group [112,139]. The underlying MvdW model then estimates a peak luminosity L0 = 4.2 × 1045 erg/s, a reference time t0 = 100 s, early-time decay index a1 = 0.542, late-time decay index a2 = 1.263, and a blast break time at tb = 3.71 × 104 s. Observational points are indicated by the markers. This GRB light curve can accommodate a MvdW solution as shown in the next figure.
Astronomy 04 00014 g0a1
To complement the broader compactness and observational survey plot presented in Figure 2, we include here a focused viability map in Figure A2, of the (amix, bmix) parameter space that highlights the intersection of theoretical and observational constraints. This diagnostic plot narrows in on the quantum-consistent region for the MvdW EOS and overlays contours of constant compactness alongside the BZT region, defined by a negative fundamental derivative. Unlike the broader color-mapped gradient in Figure 2, this visualization cleanly distinguishes the boundary of the BZT zone and clarifies how nonconvex thermodynamic behavior depends on the balance of attractive and repulsive interactions for an energy change Δϵ. The color map itself displays the effective energy discontinuity Δϵ, a quantity directly linked to both shock dynamics and observable GRB luminosity. Compactness curves are labeled and overlaid, allowing immediate comparison to observational thresholds from neutron stars and gravitational wave events. A red circle identifies a viable model for GRB 010222 using the MvdW EOS that satisfies causality, compactness, and quantum bounds simultaneously. By constraining attention to a physically viable domain and highlighting the synergy between microscopic EOS parameters and macroscopic observables, this plot serves as a precise guide to the optimal MvdW parameter space and illustrates where realistic shock-induced phase transitions may occur.
Figure A2. Contour plot of the effective energy shift Δϵ (amix, bmix) in units of MeV/fm3 for the multicomponent van der Waals (MvdW) equation of state, mapped over a range of interaction parameters. The red circle marks the best-fit point for GRB 010222, derived from its observed afterglow slopes (a1 = 0.542, a2 = 1.263 from Cowsik) and constrained to lie within the quantum-limited region (white dashed rectangle) and below the causal compactness threshold (C < 0.354). The BZT zone (solid red box) indicates where nonconvex thermodynamic behavior allows for Bethe–Zel’dovich–Thompson-type shock propagation. Compactness contours from C = 0.20 to 0.450 are shown in cyan, with values labeled in black on the left. The selected point has a compactness of approximately C ≈ 0.34 and a local energy shift of Δϵ ≈ 101.0 MeV/fm3, positioning it just outside the BZT region but within the physically viable range for dense quark matter sources consistent with GRB observations.
Figure A2. Contour plot of the effective energy shift Δϵ (amix, bmix) in units of MeV/fm3 for the multicomponent van der Waals (MvdW) equation of state, mapped over a range of interaction parameters. The red circle marks the best-fit point for GRB 010222, derived from its observed afterglow slopes (a1 = 0.542, a2 = 1.263 from Cowsik) and constrained to lie within the quantum-limited region (white dashed rectangle) and below the causal compactness threshold (C < 0.354). The BZT zone (solid red box) indicates where nonconvex thermodynamic behavior allows for Bethe–Zel’dovich–Thompson-type shock propagation. Compactness contours from C = 0.20 to 0.450 are shown in cyan, with values labeled in black on the left. The selected point has a compactness of approximately C ≈ 0.34 and a local energy shift of Δϵ ≈ 101.0 MeV/fm3, positioning it just outside the BZT region but within the physically viable range for dense quark matter sources consistent with GRB observations.
Astronomy 04 00014 g0a2

Appendix B. Bayesian Inference Corner Plot for MvdW

This corner plot visualizes the posterior distributions of key MvdW EOS parameters—namely, the attractive interaction strength amix, excluded volume bmix, and a derived phase-sensitive quantity Δϕ = |ΔΨ|. The 1D histograms reveal the most probable parameter values, while the 2D contours illustrate correlations and trade-offs, such as the inverse relationship between amix and bmix required to maintain EOS consistency with astrophysical constraints. This analysis provides insight into how microphysical quark interactions influence macroscopic observables, such as tidal deformability or BZT-driven GRB signals, and helps identify the viable region of parameter space for modeling phase transitions in quark stars.
This figure presents the posterior distributions and parameter correlations for a multicomponent van der Waals (MvdW) equation of state (EOS) used in modeling dense matter, likely within the context of neutron stars or quark stars. It is a corner plot showing marginalized 1D histograms along the diagonal and 2D posterior contours off-diagonal for three key EOS parameters: amix: the effective attractive interaction strength between quark flavors or particle species in the MvdW framework; bmix: the effective excluded volume per particle, encoding short-range repulsion; and |ΔΨ|: the gravity wave phase shift resulting in an observable signature.
Figure A3. Posterior contour and marginal distribution plot for multicomponent van der Waals EOS parameters derived from Bayesian inference. Shown are the marginalized distributions for the attraction parameter amix, excluded volume parameter bmix, and derived observable shift |ΔΨ|. The central red contours reflect 68%, 95%, and 99% confidence intervals, revealing correlations among EOS parameters relevant for compact star structure or gravitational wave phase evolution.
Figure A3. Posterior contour and marginal distribution plot for multicomponent van der Waals EOS parameters derived from Bayesian inference. Shown are the marginalized distributions for the attraction parameter amix, excluded volume parameter bmix, and derived observable shift |ΔΨ|. The central red contours reflect 68%, 95%, and 99% confidence intervals, revealing correlations among EOS parameters relevant for compact star structure or gravitational wave phase evolution.
Astronomy 04 00014 g0a3
The 2D contours represent confidence regions (set to 68%, 95%, and 99%) showing the most probable combinations of parameters inferred from data (e.g., mass–radius measurements, tidal deformabilities, or gravitational wave phase shifts). There is a strong positive correlation between amix and |ΔΨ|, and a moderate anticorrelation between bmix and |ΔΨ|, indicating how changes in interaction strength or excluded volume affect observable features.

Appendix C. Neutron Star Sources

In addition to the Swift Database, the neutron stars used in Figure 2 and Figure 3 are from the following references:
Table A2. Additional references for the neutron stars used in Figure 2, Figure 3 and Figure A2 and model calculations in Appendix A.
Table A2. Additional references for the neutron stars used in Figure 2, Figure 3 and Figure A2 and model calculations in Appendix A.
Neutron StarReference
PSR J0030+0451[140]
PSR J0740+6620[141]
PSR J1614–2230[142]
PSR J0952–0607[143]
PSR J0348+0432[144]
PSR J2124–3358[145]
PSR J1909–3744[146]
GRB 170817A[147]
GRB 190425[148]
GRB 211211A[149]
GRB 230307A[150]
GRB 221009A[151]
GRB 150101B[152]
GW 190814[153]
PSR B1642–03[154]

References

  1. Smith, N.; Li, W.; Foley, R.J.; Wheeler, J.C.; Pooley, D.; Chornock, R.; Filippenk, A.V.; Siverman, J.M.; Quimby, R.; Bloom, J.S.; et al. SN 2006gy: Discovery of the most luminous supernova ever recorded, powered by the death of an extremely massive star like η Carinae. Astrophys. J. 2007, 666, 1116. [Google Scholar] [CrossRef]
  2. Xu, R. Strange quark stars: Observations and speculations. J. Phys. G Nucl. Part. Phys. 2009, 36, 064010. [Google Scholar] [CrossRef]
  3. Burwitz, V.; Haberl, F.; Neuhäuser, R.; Predehl, P.; Trümper, J.; Zavlin, V.E. The thermal radiation of the isolated neutron star RX J1856. 5–3754 observed with Chandra and XMM-Newton. Astron. Astrophys. 2003, 399, 1109–1114. [Google Scholar] [CrossRef]
  4. Burwitz, V.; Zavlin, V.E.; Neuhäuser, R.; Predehl, P.; Trümper, J.; Brinkman, A.C. The Chandra LETGS high resolution X-ray spectrum of the isolated neutron star RX J1856. 5-3754. Astron. Astrophys. 2001, 379, L35–L38. [Google Scholar] [CrossRef]
  5. Xu, R.X.; Qiao, G.J.; Zhang, B. PSR 0943 + 10: A bare strange star? Astrophys. J. Lett. 1999, 522, L109. [Google Scholar] [CrossRef]
  6. Lyne, A.; Hobbs, G.; Kramer, M.; Stairs, I.; Stappers, B. Switched magnetospheric regulation of pulsar spin-down. Science 2010, 329, 408–412. [Google Scholar] [CrossRef] [PubMed]
  7. Alford, M.; Braby, M.; Paris, M.; Reddy, S. Hybrid stars that masquerade as neutron stars. Astrophys. J. 2005, 629, 969. [Google Scholar] [CrossRef]
  8. Jin, Z.-P.; Hotokezaka, K.; Li, X.; Tanaka, M.; D’Avanzo, P.; Fan, Y.-Z.; Covino, S.; Wei, D.-M.; Piran, T. The Macronova in GRB 050709 and the GRB-macronova connection. Nat. Commun. 2016, 7, 12898. [Google Scholar] [CrossRef]
  9. Von, K.A.; Veres, P.; Roberts, O.J.; Hamburg, R.; Bissaldi, E.; Briggs, M.S.; Burns, E.; Goldstein, A.; Kocevski, D.; Preece, R.D.; et al. Fermi-GBM GRBs with characteristics similar to GRB 170817A. Astrophys. J. 2019, 876, 89. [Google Scholar] [CrossRef]
  10. Ajello, M.; Arimoto, M.; Axelsson, M.; Baldini, L.; Barbiellini, G.; Bastieri, D.; Bellazzini, R.; Bhat, P.N.; Bissaldi, E.; Blandford, R.D.; et al. A decade of gamma-ray bursts observed by Fermi-LAT: The second GRB catalog. Astrophys. J. 2019, 878, 52. [Google Scholar] [CrossRef]
  11. Olausen, S.A.; Kaspi, V.M. The McGill magnetar catalog. Astrophys. J. Suppl. Ser. 2014, 212, 6. [Google Scholar] [CrossRef]
  12. Gavriil, F.P.; Kaspi, V.M.; Woods, P.M. A comprehensive study of the X-ray bursts from the magnetar candidate 1E 2259 + 586. Astrophys. J. 2004, 607, 959. [Google Scholar] [CrossRef]
  13. Astashenok, A.V.; Capozziello, S.; Odintsov, S.D.; Vasilis, K. Oikonomou. Extended gravity description for the GW190814 supermassive neutron star. Phys. Lett. B 2020, 811, 135910. [Google Scholar]
  14. Liu, B.; Lai, D. Hierarchical black hole mergers in multiple systems: Constrain the formation of GW190412-, GW190814-, and GW190521-like events. Mon. Not. R. Astron. Soc. 2021, 502, 2049–2064. [Google Scholar] [CrossRef]
  15. Annala, E.; Gorda, T.; Hirvonen, J.; Komoltsev, O.; Kurkela, A.; Nättilä, J.; Vuorinen, A. Strongly interacting matter exhibits deconfined behavior in massive neutron stars. Nat. Commun. 2023, 14, 8451. [Google Scholar] [CrossRef] [PubMed]
  16. Heinz, U. From SPS to RHIC: Breaking the barrier to the quark-gluon plasma. In Proceedings of the AIP Conference Proceedings, Melville, NY, USA, 13 December 2001; Volume 602, pp. 281–292. [Google Scholar]
  17. Emerick, A.; Zhao, X.; Rapp, R. Bottomonia in the quark-gluon plasma and their production at RHIC and LHC. Eur. Phys. J. A 2012, 48, 72. [Google Scholar] [CrossRef]
  18. Teaney, D.; Lauret, J.; Shuryak, E.V. Flow at the SPS and RHIC as a quark-gluon plasma signature. Phys. Rev. Lett. 2001, 86, 4783. [Google Scholar] [CrossRef] [PubMed]
  19. Arsene, I.; Bearden, I.G.; Beavis, D.; Besliu, C.; Budick, B.; Bøggild, H.; Chasman, C.; Christensen, C.H.; Christiansen, P.; Cibor, J.; et al. Quark–gluon plasma and color glass condensate at RHIC? The perspective from the BRAHMS experiment. Nucl. Phys. A 2005, 757, 1–27. [Google Scholar] [CrossRef]
  20. Florkowski, W.; Jaiswal, A.; Maksymiuk, E.; Ryblewski, R.; Strickland, M. Relativistic quantum transport coefficients for second-order viscous hydrodynamics. Phys. Rev. C 2015, 91, 054907. [Google Scholar] [CrossRef]
  21. Heinz, U.; Shen, C.; Song, H. The viscosity of quark-gluon plasma at RHIC and the LHC. In Proceedings of the AIP Conference Proceedings, Melville, NY, USA, 28 September 2012; Volume 1441, pp. 766–770. [Google Scholar]
  22. Moore, G.D.; Saremi, O. Bulk viscosity and spectral functions in QCD. J. High Energy Phys. 2008, 2008, 015. [Google Scholar] [CrossRef]
  23. Miller, J.M.; Fabian, A.C.; Wijnands, R.; Reynolds, C.S.; Ehle, M.; Freyberg, M.J.; Van Der Klis, M.; Lewin, W.H.G.; Sanchez-Fernandez, C.; Castro-Tirado, A.J. Evidence of Spin and Energy Extraction in a Galactic Black Hole Candidate: The XMM-Newton/EPIC-pn Spectrum of XTE J1650–500. Astrophys. J. Lett. 2002, 570, L69. [Google Scholar] [CrossRef]
  24. Dey, M.; Bombaci, I.; Dey, J.; Ray, S.; Samanta, B.C. Strange stars with realistic quark vector interaction and phenomenological density-dependent scalar potential. Phys. Lett. B 1998, 438, 123–128. [Google Scholar] [CrossRef]
  25. Dolan, L.; Jackiw, R. Symmetry behavior at finite temperature. Phys. Rev. D 1974, 9, 3320. [Google Scholar] [CrossRef]
  26. Ray, S.; Dey, J.; Dey, M. Density dependent strong coupling constant of QCD derived from compact star data. Mod. Phys. Lett. A 2000, 15, 1301–1306. [Google Scholar] [CrossRef]
  27. Gendreau, K.C.; Arzoumanian, Z.; Okajima, T. The Neutron star Interior Composition ExploreR (NICER): An Explorer mission of opportunity for soft x-ray timing spectroscopy. In Proceedings of the Space Telescopes and Instrumentation 2012: Ultraviolet to Gamma Ray, Amsterdam, Netherlands, 1–6 July 2012; Volume 8443, pp. 322–329. [Google Scholar]
  28. Vinciguerra, S.; Salmi, T.; Watts, A.L.; Choudhury, D.; Riley, T.E.; Ray, P.S.; Bogdanov, S.; Kini, Y.; Guillot, S.; Chakrabarty, D.; et al. An Updated Mass–Radius Analysis of the 2017–2018 NICER Data Set of PSR J0030 + 0451. Astrophys. J. 2024, 961, 62. [Google Scholar] [CrossRef]
  29. Johnson, K. The MIT bag model. Acta Phys. Pol. B 1975, 6, 8. [Google Scholar]
  30. Weber, F. Strange quark matter and compact stars. Prog. Part. Nucl. Phys. 2005, 54, 193–288. [Google Scholar] [CrossRef]
  31. Alford, M.; Rajagopal, K.; Wilczek, F. Color-flavor locking and chiral symmetry breaking in high density QCD. Nucl. Phys. B 1999, 537, 443–458. [Google Scholar] [CrossRef]
  32. Rajagopal, K.; Wilczek, F. Enforced electrical neutrality of the color-flavor locked phase. Phys. Rev. Lett. 2001, 86, 3492. [Google Scholar] [CrossRef]
  33. Alford, M.G.; Schmitt, A.; Rajagopal, K.; Schäfer, T. Color superconductivity in dense quark matter. Rev. Mod. Phys. 2008, 80, 1455. [Google Scholar] [CrossRef]
  34. Alford, M.; Reddy, S. Compact stars with color superconducting quark matter. Phys. Rev. D 2003, 67, 074024. [Google Scholar] [CrossRef]
  35. Baldo, M.; Buballa, M.; Burgio, G.F.; Lawley, F.; Bentz, W.; Thomas, A.W. Neutron stars and the transition to color superconducting quark matter. Phys. Lett. B 2003, 562, 153–160. [Google Scholar] [CrossRef]
  36. Lawley, S.; Bentz, W.; Thomas, A.W. Nucleons, nuclear matter and quark matter: A unified NJL approach. J. Phys. G Nucl. Part. Phys. 2006, 32, 667. [Google Scholar] [CrossRef]
  37. Baym, G.; Chin, S.A. Can a neutron star be a giant MIT bag? Phys. Lett. B 1976, 62, 241–244. [Google Scholar] [CrossRef]
  38. Källman, C.-G. Mean-field QCD model for hot/dense matter. Phys. Lett. B 1984, 134, 363–367. [Google Scholar] [CrossRef]
  39. Adler, S.L. Generalized bag models as mean-field approximations to QCD. Phys. Lett. B 1982, 110, 302–306. [Google Scholar] [CrossRef]
  40. McLerran, L.; Reddy, S. Quarkyonic matter and neutron stars. Phys. Rev. Lett. 2019, 122, 122701. [Google Scholar] [CrossRef]
  41. Zhao, T.; Lattimer, J.M. Quarkyonic matter equation of state in beta-equilibrium. Phys. Rev. D 2020, 102, 023021. [Google Scholar] [CrossRef]
  42. Fodor, Z.; Katz, S.D. A new method to study lattice QCD at finite temperature and chemical potential. Phys. Lett. B 2002, 534, 87–92. [Google Scholar] [CrossRef]
  43. Glendenning, N.K.; Weber, F. Nuclear solid crust on rotating strange quark stars. Astrophys. J. 1992, 400., 1–15. [Google Scholar] [CrossRef]
  44. Owen, B.J. Maximum elastic deformations of compact stars with exotic equations of state. Phys. Rev. Lett. 2005, 95, 211101. [Google Scholar] [CrossRef]
  45. Sharma, R.; Mukherjee, S. Compact stars: A core-envelope model. Mod. Phys. Lett. A 2002, 17, 2535–2544. [Google Scholar] [CrossRef]
  46. Ng, C.Y.; Cheng, K.S.; Chu, M.C. Cooling properties of Cloudy Bag strange stars. Astropart. Phys. 2003, 19, 171–192. [Google Scholar] [CrossRef]
  47. Latifah, S.; Sulaksono, A.; Mart, T. Bosons star at finite temperature. Phys. Rev. D 2014, 90, 127501. [Google Scholar] [CrossRef]
  48. Andrew, K.; Andrew, K.; Brown, R.; Thornberry, B.; Harper, S.; Steinfelds, E.; Roberts, T. A QCD Model of the Chemical Potential Kaon Boundary Formation for a Compact Quark Star. Bull. Am. Phys. Soc. 2016, 61, 241. [Google Scholar]
  49. Thorsson, V.; Prakash, M.; Lattimer, J.M. Composition, structure and evolution of neutron stars with kaon condensates. Nucl. Phys. A 1994, 572, 693–731. [Google Scholar] [CrossRef]
  50. Alford, M. Color-superconducting quark matter. Annu. Rev. Nucl. Part. Sci. 2001, 51, 131–160. [Google Scholar] [CrossRef]
  51. Andrew, K.; Steinfelds, E.; Andrew, K. QCD Color Vector Potential Effects on Color Flavor Locked Quark Stellar Cores. Bull. Am. Phys. Soc. 2018, 63, 185. [Google Scholar]
  52. Alford, M.; Bowers, J.A.; Rajagopal, K. Crystalline color superconductivity. Phys. Rev. D 2001, 63, 074016. [Google Scholar] [CrossRef]
  53. Casalbuoni, R.; Gatto, R.; Mannarelli, M.; Nardulli, G. Effective field theory for the crystalline colour superconductive phase of QCD. Phys. Lett. B 2001, 511, 218–228. [Google Scholar] [CrossRef]
  54. Andrew, K.; Brown, R.; Andrew, K.; Thornberry, B.; Harper, S.; Steinfelds, E.; Roberts, T. Analysis of Mass and Radius Sensitivity of a Crystalline Quark Star to a Strong Repulsive Equation of State. Bull. Am. Phys. Soc. 2016, 61, 244. [Google Scholar]
  55. Berbel, P.M. On Nonconvex Special Relativistic Hydrodynamics. Ph.D. Thesis, Universatat Autonoma de Barcelona, Barcelona, Spain, 2023. [Google Scholar]
  56. Marquina, A.; Serna, S.; Ibanez, J.M. Capturing composite waves in non-convex special relativistic hydrodynamics. J. Sci. Comput. 2019, 81, 2132–2161. [Google Scholar] [CrossRef]
  57. Duez, M.D.; Liu, Y.T.; Shapiro, S.L.; Stephens, B.C. General relativistic hydrodynamics with viscosity: Contraction, catastrophic collapse, and disk formation in hypermassive neutron stars. Phys. Rev. D 2004, 69, 104030. [Google Scholar] [CrossRef]
  58. See, N. Technical Information Service Document No. PB 032189 H. A. Bethe, the Theory of Shock Waves for an Arbitrary Equation of State; Office of Scientific Research and Development Paper No. 545; National Technical Information Service: Springfield, VA, USA, 1942. [Google Scholar]
  59. Zel’dovich, Y.B. Theory of Shock Waves and Introduction to Gas Dynamics; Izdatel’stvo Akademii Nauk SSSR: Moscow, Russia, 1946. [Google Scholar]
  60. Thompson, P.A. A fundamental derivative in gasdynamics. Phys. Fluids 1971, 14, 1843. [Google Scholar] [CrossRef]
  61. Aloy, M.A.; Ibáñez, J.M.; Sanchis-Gual, N.; Obergaulinger, M.; Jose, A.; Susana Serna, F.; Marquina, A. Neutron star collapse and gravitational waves with a non-convex equation of state. Mon. Not. R. Astron. Soc. 2019, 484, 4980–5008. [Google Scholar] [CrossRef]
  62. Rivieccio, G.; Guerra, D.; Ruiz, M.; Font, J.A. Gravitational-wave imprints of nonconvex dynamics in binary neutron star mergers. Phys. Rev. D 2024, 109, 064032. [Google Scholar] [CrossRef]
  63. Most, E.R.; Papenfort, L.J.; Dexheimer, V.; Hanauske, M.; Stoecker, H.; Rezzolla, L. On the deconfinement phase transition in neutron-star mergers. Eur. Phys. J. A 2020, 56, 59. [Google Scholar] [CrossRef]
  64. Sarkar, N.; Ghosh, P. van der Waals hadron resonance gas and QCD phase diagram. Phys. Rev. C 2018, 98, 014907. [Google Scholar] [CrossRef]
  65. Vovchenko, V.; Gorenstein, M.I.; Stoecker, H. van der Waals interactions in hadron resonance gas: From nuclear matter to lattice QCD. Phys. Rev. Lett. 2017, 118, 182301. [Google Scholar] [CrossRef]
  66. Rodrigues, E.H.; Dutra, M.; Lourenço, O. Recent astrophysical observations reproduced by a short-range correlated van der Waals-type model? Mon. Not. R. Astron. Soc. 2023, 523, 4859–4868. [Google Scholar] [CrossRef]
  67. Andrew, K.; Steinfelds, E.V.; Andrew, K.A. The van der Waals Hexaquark Chemical Potential in Dense Stellar Matter. Particles 2023, 6, 556–567. [Google Scholar] [CrossRef]
  68. Poberezhnyuk, R.V.; Stoecker, H.; Vovchenko, V. Quarkyonic matter with quantum van der Waals theory. Phys. Rev. C 2023, 108, 045202. [Google Scholar] [CrossRef]
  69. Thirukkanesh, S.; Ragel, F.C. Anisotropic compact sphere with Van der Waals equation of state. Astrophys. Space Sci. 2014, 354, 415–420. [Google Scholar] [CrossRef]
  70. Vovchenko, V.; Motornenko, A.; Alba, P.; Gorenstein, M.I.; Satarov, L.M.; Stoecker, H. Multicomponent van der Waals equation of state: Applications in nuclear and hadronic physics. Phys. Rev. C 2017, 96, 045202. [Google Scholar] [CrossRef]
  71. Mardan, S.A.; Khalid, A.; Manzoor, R.; Riaz, M.B. Anisotropic model observing pulsars from Neutron Star Interior Composition with modified Van der Waals equation of state. Eur. Phys. J. C 2024, 84, 1018. [Google Scholar] [CrossRef]
  72. Lourenço, O.; Dutra, M.; Lenzi, C.H.; Bhuyan, M.; Biswal, S.K.; Santos, B.M. A density-dependent van der waals model under the gw170817 constraint. Astrophys. J. 2019, 882, 67. [Google Scholar] [CrossRef]
  73. Dutra, M.; Santos, B.M.; Lourenço, O. Constraints and correlations of nuclear matter parameters from a density-dependent van der Waals model. J. Phys. G Nucl. Part. Phys. 2020, 47, 035101. [Google Scholar] [CrossRef]
  74. Malaver, M.; Kasmaei, H.D. Analytical models for quark stars with van der Waals modified equation of state. Int. J. Astrophys. Space Sci. 2019, 7, 58. [Google Scholar] [CrossRef]
  75. Errehymy, A.; Mustafa, G.; Khedif, Y.; Daoud, M.; Alrebdi, H.I.; Abdel-Aty, A.-H. Self-gravitating anisotropic model in general relativity under modified Van der Waals equation of state: A stable configuration. Eur. Phys. J. C 2022, 82, 1–11. [Google Scholar] [CrossRef]
  76. Poberezhnyuk, R.V.; Vovchenko, V.; Gorenstein, M.I.; Stoecker, H. Noncongruent phase transitions in strongly interacting matter within the quantum van der Waals model. Phys. Rev. C 2019, 99, 024907. [Google Scholar] [CrossRef]
  77. Moss, T.; Poberezhniuk, R.; Vovchenko, V. Quantum van der Waals quarkyonic matter at non-zero isospin asymmetry. arXiv 2024, arXiv:2411.11996. [Google Scholar] [CrossRef]
  78. Guardone, A.; Vigevano, L. Roe linearization for the van der Waals gas. J. Comput. Phys. 2002, 175, 50–78. [Google Scholar] [CrossRef]
  79. Ibáñez, J.-M.; Cordero-Carrión, I.; Aloy, M.-Á.; Martí, J.-M.; Miralles, J.-A. On the convexity of relativistic ideal magnetohydrodynamics. Class. Quantum Gravity 2015, 32, 095007. [Google Scholar] [CrossRef]
  80. Kapusta, J.; Singh, M.; Welle, T. Spinodal decomposition in Bjorken flow. arXiv 2024, arXiv:2409.16525. [Google Scholar] [CrossRef]
  81. Vovchenko, V.; Anchishkin, D.V.; Gorenstein, M.I. Van der Waals equation of state with Fermi statistics for nuclear matter. Phys. Rev. C 2015, 91, 064314. [Google Scholar] [CrossRef]
  82. Keffer, D. The Statistical Mechanical Derivation of the Van Der Waals Equation of State for a Multicomponent Fluid and Its Associated Thermodynamic Properties. Available online: http://www.utkstair.org/clausius/docs/che330/pdf/vdw_statmech_multicomponent.pdf (accessed on 4 March 2023).
  83. Stryjek, R.; Vera, J.H. PRSV: An improved Peng—Robinson equation of state for pure compounds and mixtures. Can. J. Chem. Eng. 1986, 64, 323–333. [Google Scholar] [CrossRef]
  84. Shahrbaf, M.; Blaschke, D.; Typel, S.; Farrar, G.R.; Alvarez-Castillo, D.E. Sexaquark dilemma in neutron stars and its solution by quark deconfinement. Phys. Rev. D 2022, 105, 103005. [Google Scholar] [CrossRef]
  85. CompOSE Core Team; Typel, S.; Oertel, M.; Klähn, T.; Chatterjee, D.; Dexheimer, V.; Ishizuka, C.; Manchini, M.; Novak, J.; Paos, H.; et al. CompOSE reference manual: Version 3.01, CompStar Online Supernovæ Equations of State, “harmonising the concert of nuclear physics and astrophysics”. Eur. Phys. J. A 2022, 58, 221. [Google Scholar]
  86. Zeng, J. Non-Classical Behavior of BZT Gas in Isentropic Quasi-One-Dimensional Flow. Master’s Thesis, University of California, Irvine, CA, USA, 2020. [Google Scholar]
  87. Colonna, P.; Nannan, N.R.; Guardone, A.; van der Stelt, T.P. On the computation of the fundamental derivative of gas dynamics using equations of state. Fluid Phase Equilibria 2009, 286, 43–54. [Google Scholar] [CrossRef]
  88. Baym, G.; Pethick, C.; Sutherland, P. The ground state of matter at high densities: Equation of state and stellar models. Astrophys. J. 1971, 170, 299. [Google Scholar] [CrossRef]
  89. Ghosh, S.; Pradhan, B.K.; Chatterjee, D.; Schaffner-Bielich, J. Multi-physics constraints at different densities to probe nuclear symmetry energy in hyperonic neutron stars. Front. Astron. Space Sci. 2022, 9, 864294. [Google Scholar] [CrossRef]
  90. Oppenheimer, J.R.; Volkoff, G.M. On massive neutron cores. Phys. Rev. 1939, 55, 374. [Google Scholar] [CrossRef]
  91. Rahmansyah, A.; Purnamasari, D.; Kurniadi, R.; Sulaksono, A. Generalized Tolman-Oppenheimer-Volkoff model and neutron stars. Phys. Rev. D 2022, 106, 084042. [Google Scholar] [CrossRef]
  92. Smoller, J.; Temple, B. On the Oppenheimer-Volkoff Equations in General Relativity. Arch. Ration. Mech. Anal. 1998, 142, 177–191. [Google Scholar] [CrossRef]
  93. Wolfram, S. The Mathematica Book; Wolfram Research, Inc.: Oxfordshire, UK, 2003. [Google Scholar]
  94. Abell, M.L.; Braselton, J.P. Differential Equations with Mathematica; Academic Press: Cambridge, MA, USA, 2022. [Google Scholar]
  95. Moustakidis, C.C. The stability of relativistic stars and the role of the adiabatic index. Gen. Relativ. Gravit. 2017, 49, 1–21. [Google Scholar] [CrossRef]
  96. Casali, R.H.; Menezes, D.P. Adiabatic index of hot and cold compact objects. Braz. J. Phys. 2010, 40, 166–171. [Google Scholar] [CrossRef]
  97. Carney, M.F.; Wade, L.E.; Irwin, B.S. Comparing two models for measuring the neutron star equation of state from gravitational-wave signals. Phys. Rev. D 2018, 98, 063004. [Google Scholar] [CrossRef]
  98. Hartle, J.B.; Thorne, K.S. Slowly rotating relativistic stars. II. Models for neutron stars and supermassive stars. Astrophys. J. 1968, 153, 807. [Google Scholar] [CrossRef]
  99. Bauböck, M.; Berti, E.; Psaltis, D.; Özel, F. Relations between neutron-star parameters in the hartle–thorne approximation. Astrophys. J. 2013, 777, 68. [Google Scholar] [CrossRef]
  100. Bocquet, M.; Bonazzola, S.; Gourgoulhon, E.; Novak, J. Rotating neutron star models with magnetic field. arXiv 1995, arXiv:gr-qc/9503044. [Google Scholar] [CrossRef]
  101. Romani, R.W. A unified model of neutron-star magnetic fields. Nature 1990, 347, 741–743. [Google Scholar] [CrossRef]
  102. Pretel, J.M.Z. Equilibrium, radial stability and non-adiabatic gravitational collapse of anisotropic neutron stars. Eur. Phys. J. C 2020, 80, 726. [Google Scholar] [CrossRef]
  103. Baiotti, L. Gravitational waves from binary neutron stars. Arab. J. Math. 2022, 11, 105–118. [Google Scholar] [CrossRef]
  104. Vines, J.; Flanagan, E.E.; Hinderer, T. Post-1-Newtonian tidal effects in the gravitational waveform from binary inspirals. Phys. Rev. D—Part. Fields Gravit. Cosmol. 2011, 83, 084051. [Google Scholar] [CrossRef]
  105. Fraga, E.S.; Kurkela, A.; Vuorinen, A. Interacting quark matter equation of state for compact stars. Astrophys. J. Lett. 2014, 781, L25. [Google Scholar] [CrossRef]
  106. Li, H.; Luo, X.-L.; Zong, H.-S. Bag model and quark star. Phys. Rev. D—Part. Fields Gravit. Cosmol. 2010, 82, 065017. [Google Scholar] [CrossRef]
  107. Huovinen, P.; Petreczky, P. QCD equation of state and hadron resonance gas. Nucl. Phys. A 2010, 837, 26–53. [Google Scholar] [CrossRef]
  108. Kojo, T.; Powell, P.D.; Song, Y.; Baym, G. Phenomenological QCD equation of state for massive neutron stars. Phys. Rev. D 2015, 91, 045003. [Google Scholar] [CrossRef]
  109. Orsaria, M.; Rodrigues, H.; Weber, F.; Contrera, G.A. Quark deconfinement in high-mass neutron stars. Phys. Rev. C 2014, 89, 015806. [Google Scholar] [CrossRef]
  110. Bazavov, A.; Bhattacharya, T.; DeTar, C.; Ding, H.-T.; Gottlieb, S.; Gupta, R.; Hegde, P.; Heller, U.M.; Karsch, F.; Laermann, E.; et al. Equation of state in (2 + 1)-flavor QCD. Phys. Rev. D 2014, 90, 094503. [Google Scholar] [CrossRef]
  111. Laermann, E. The equation of state from lattice QCD. Acta Phys. Pol. Ser. B Proc. Suppl. 2010, 3, 611–620. [Google Scholar]
  112. Schady, P.; Baumgartner, W.H.; Beardmore, A.P. Swift Observation of GRB 090618. GCN Report 232, No. 1. 2009. Available online: http://gcn.gsfc.nasa.gov (accessed on 12 May 2024).
  113. Jain, T. Analytic Modelling of Gravitational-Wave Source in General Relativity and Alternative Theories of Gravity. Ph.D. Thesis, Cambridge University, Cambridge, UK, 2024. [Google Scholar]
  114. Martynov, D.V.; Hall, E.D.; Abbott, B.P.; Abbott, R.; Abbott, T.D.; Adams, C.; Adhikari, R.X.; Anderson, R.A.; Anderson, S.B.; Arai, K.; et al. Sensitivity of the Advanced LIGO detectors at the beginning of gravitational wave astronomy. Phys. Rev. D 2016, 93, 112004. [Google Scholar] [CrossRef]
  115. Aasi, J.; Abadie, J.; Abbott, B.P.; Abbott, R.; Abbott, T.D.; Abernathy, M.; Accadia, T.; Acernese, F.; Adams, C.; Addesso, P.; et al. The characterization of Virgo data and its impact on gravitational-wave searches. Class. Quantum Gravity 2012, 29, 155002. [Google Scholar] [CrossRef]
  116. KAGRA Collaboration. KAGRA: 2.5 generation interferometric gravitational wave detector. Nat. Astron. 2019, 3, 35–40. [Google Scholar] [CrossRef]
  117. Punturo, M.; Abernathy, M.; Acernese, F.; Allen, B.; Andersson, N.; Arun, K.; Barone, F.; Barr, B.; Barsuglia, M.; Beker, M.; et al. The Einstein Telescope: A third-generation gravitational wave observatory. Class. Quantum Gravity 2010, 27, 194002. [Google Scholar] [CrossRef]
  118. Cutler, C. Angular resolution of the LISA gravitational wave detector. Phys. Rev. D 1998, 57, 7089. [Google Scholar] [CrossRef]
  119. Posselt, B.; Pavlov, G.G. The cooling of the central compact object in Cas A from 2006 to 2020. Astrophys. J. 2022, 932, 83. [Google Scholar] [CrossRef]
  120. Tanabashi, M.; Hagiwara, K.; Hikasa, K.; Nakamura, K.; Sumino, Y.; Takahashi, F.; Tanaka, J.; Agashe, K.; Aielli, G.; Amsler, C.; et al. Review of Particle Physics: Particle data groups. Phys. Rev. D 2018, 98, 1–1898. [Google Scholar] [CrossRef]
  121. Vovchenko, V.; Anchishkin, D.V.; Gorenstein, M.I.; Poberezhnyuk, R.V. Scaled variance, skewness, and kurtosis near the critical point of nuclear matter. Phys. Rev. C 2015, 92, 054901. [Google Scholar] [CrossRef]
  122. Farhi, E.; Jaffe, R.L. Strange matter. Phys. Rev. D 1984, 30, 2379. [Google Scholar] [CrossRef]
  123. Strottman, D. Multiquark baryons and the MIT bag model. Phys. Rev. D 1979, 20, 748. [Google Scholar] [CrossRef]
  124. Hess, P.O.; Viollier, R.D. Interacting many-gluon systems within the MIT bag model. Phys. Rev. D 1986, 34, 258. [Google Scholar] [CrossRef]
  125. Buballa, M. NJL-model analysis of dense quark matter. Phys. Rep. 2005, 407, 205–376. [Google Scholar] [CrossRef]
  126. Klähn, T.; Blaschke, D.; Sandin, F.; Fuchs, C.; Faessler, A.; Grigorian, H.; Röpke, G.; Trümper, J. Modern compact star observations and the quark matter equation of state. Phys. Lett. B 2007, 654, 170–176. [Google Scholar] [CrossRef]
  127. Karsch, F.; Redlich, K.; Tawfik, A. Thermodynamics at non-zero baryon number density: A comparison of lattice and hadron resonance gas model calculations. Phys. Lett. B 2003, 571, 67–74. [Google Scholar] [CrossRef]
  128. Endrődi, 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]
  129. Bazavov, A.; Ding, H.T.; Hegde, P.; Kaczmarek, O.; Karsch, F.; Laermann, E.; Maezawa, Y.; Mukherjee, S.; Ohno, H.; Petreczky, P.; et al. QCD equation of state to O (μ B 6) from lattice QCD. Phys. Rev. D 2017, 95, 054504. [Google Scholar] [CrossRef]
  130. Kahangirwe, M.; Bass, S.A.; Bratkovskaya, E.; Jahan, J.; Moreau, P.; Parotto, P.; Price, D.; Ratti, C.; Soloveva, O.; Stephanov, M. Finite density QCD equation of state: Critical point and lattice-based T′ expansion. Phys. Rev. D 2024, 109, 094046. [Google Scholar] [CrossRef]
  131. Borsányi, S.; Fodor, Z.; Guenther, J.N.; Kara, R.; Katz, S.D.; Parotto, P.; Pásztor, A.; Ratti, C.; Lattice, K.K.S. QCD equation of state at finite chemical potential from an alternative expansion scheme. Phys. Rev. Lett. 2021, 126, 232001. [Google Scholar] [CrossRef] [PubMed]
  132. Mészáros, P.; Rees, M.J. Optical and long-wavelength afterglow from gamma-ray bursts. Astrophys. J. 1997, 476, 232. [Google Scholar] [CrossRef]
  133. Mészáros, P.; Rees, M.J. Poynting jets from black holes and cosmological gamma-ray bursts. Astrophys. J. 1997, 482, L29. [Google Scholar] [CrossRef]
  134. Kulkarni, S.R.; Djorgovski, S.G.; Odewahn, S.C.; Bloom, J.S.; Gal, R.R.; Koresko, C.D.; Harrison, F.A.; Lubin, L.M.; Armus, L.; Sari, R.; et al. The afterglow, redshift and extreme energetics of the γ-ray burst of 23 January 1999. Nature 1999, 398, 389–394. [Google Scholar] [CrossRef]
  135. Li, L.-X. Shock breakout in Type Ibc supernovae and application to GRB 060218/SN 2006aj. Mon. Not. R. Astron. Soc. 2007, 375, 240–256. [Google Scholar] [CrossRef]
  136. Irwin, C.M.; Hotokezaka, K. Revisiting GRB 060218: New insights into low-luminosity gamma-ray bursts from a revised shock breakout model. arXiv 2024, arXiv:2412.06736. [Google Scholar] [CrossRef]
  137. Kuroda, T.; Takiwaki, T.; Kotake, K. Gravitational wave signatures from low-mode spiral instabilities in rapidly rotating supernova cores. Phys. Rev. D 2014, 89, 044011. [Google Scholar] [CrossRef]
  138. Ruffini, R.; Izzo, L.; Muccino, M.; Pisani, G.B.; Rueda, J.A.; Wang, Y.; Barbarino, C.; Bianco, C.L.; Enderli, M.; Kovacevic, M. Induced gravitational collapse at extreme cosmological distances: The case of GRB 090423. Astron. Astrophys. 2014, 569, A39. [Google Scholar] [CrossRef]
  139. Cowsik, R.; Prabhu, T.P.; Anupama, G.C.; Bhatt, B.C.; Sahu, D.K.; Ambika, S.; Bhargavi, S.G. Optical photometry of the GRB 010222 afterglow. arXiv 2001, arXiv:astro-ph/0104363. [Google Scholar] [CrossRef]
  140. Miller, M.C.; Lamb, F.K.; Dittmann, A.J.; Bogdanov, S.; Arzoumanian, Z.; Gendreau, K.C.; Guillot, S.; Harding, A.K.; Ho, W.C.G.; Lattimer, J.M.; et al. PSR J0030 + 0451 mass and radius from NICER data and implications for the properties of neutron star matter. Astrophys. J. Lett. 2019, 887, L24. [Google Scholar] [CrossRef]
  141. Miller, M.C.; Lamb, F.K.; Dittmann, A.J.; Bogdanov, S.; Arzoumanian, Z.; Gendreau, K.C.; Guillot, S.; Ho, W.C.G.; Lattimer, J.M.; Loewenstein, M.; et al. The radius of PSR J0740 + 6620 from NICER and XMM-Newton data. Astrophys. J. Lett. 2021, 918, L28. [Google Scholar] [CrossRef]
  142. Zhao, X.-F.; Jia, H.-Y. Mass of the neutron star PSR J1614-2230. Phys. Rev. C—Nucl. Phys. 2012, 85, 065806. [Google Scholar] [CrossRef]
  143. Romani, R.W.; Kandel, D.; Filippenko, A.V.; Brink, T.G.; Zheng, W.K. PSR J0952− 0607: The fastest and heaviest known galactic neutron star. Astrophys. J. Lett. 2022, 934, L17. [Google Scholar] [CrossRef]
  144. Zhao, X.-F. The properties of the massive neutron star PSR J0348+ 0432. Int. J. Mod. Phys. D 2015, 24, 1550058. [Google Scholar] [CrossRef]
  145. Zheng, S.; Han, D.; Xu, H.; Lee, K.; Yuan, J.; Wang, H.; Ge, M.; Zhang, L.; Li, Y.; Yin, Y. New Timing Results MSPs NICER Observations. Universe 2024, 10, 174. [Google Scholar] [CrossRef]
  146. Liu, K.; Guillemot, L.; Istrate, A.G.; Shao, L.; Tauris, T.M.; Wex, N.; Antoniadis, J.; Chalumeau, A.; Cognard, I.; Desvignes, G.; et al. A revisit of PSR J1909−3744 with 15-yr high-precision timing. Mon. Not. R. Astron. Soc. 2020, 499, 2276–2291. [Google Scholar] [CrossRef]
  147. Rueda, J.A.; Ruffini, R.; Wang, Y.; Aimuratov, Y.; de Almeida, U.B.; Bianco, C.L.; Chen, Y.C.; Lobato, R.V.; Maia, C.; Primorac, D.; et al. Grb 170817a-gw170817-at 2017gfo and the observations of ns-ns, ns-wd and wd-wd mergers. J. Cosmol. Astropart. Phys. 2018, 2018, 006. [Google Scholar] [CrossRef]
  148. Kyutoku, K.; Fujibayashi, S.; Hayashi, K.; Kawaguchi, K.; Kiuchi, K.; Shibata, M.; Tanaka, M. On the possibility of GW190425 being a black hole–neutron star binary merger. Astrophys. J. Lett. 2020, 890, L4. [Google Scholar] [CrossRef]
  149. Yin, Y.-H.I.; Zhang, B.-B.; Sun, H.; Yang, J.; Kang, Y.; Shao, L.; Yang, Y.-H.; Zhang, B. GRB 211211A-like events and how gravitational waves may tell their origins. Astrophys. J. Lett. 2023, 954, L17. [Google Scholar] [CrossRef]
  150. Wang, Y.; Xia, Z.-Q.; Zheng, T.-C.; Ren, J.; Fan, Y.-Z. A broken “α–intensity” relation caused by the evolving photosphere emission and the nature of the extraordinarily bright GRB 230307A. Astrophys. J. Lett. 2023, 953, L8. [Google Scholar] [CrossRef]
  151. Levan, A.J.; Lamb, G.P.; Schneider, B.; Hjorth, J.; Zafar, T.; de Ugarte Postigo, A.; Sargent, B.; Mullally, S.E.; Izzo, L.; D’aVanzo, P.; et al. The first JWST spectrum of a GRB afterglow: No bright supernova in observations of the brightest GRB of all time, GRB 221009A. Astrophys. J. Lett. 2023, 946, L28. [Google Scholar] [CrossRef]
  152. Ascenzi, S.; Coughlin, M.W.; Dietrich, T.; Foley, R.J.; Ramirez-Ruiz, E.; Piranomonte, S.; Mockler, B.; Murguia-Berthier, A.; Fryer, C.L.; Lloyd-Ronning, N.M. A luminosity distribution for kilonovae based on short gamma-ray burst afterglows. Mon. Not. R. Astron. Soc. 2019, 486, 672–690. [Google Scholar] [CrossRef]
  153. Abbott, R.; Abbott, T.D.; Abraham, S.; Acernese, F.; Ackley, K.; Adams, C.; Adhikari, R.X.; Adya, V.B.; Affeldt, C.; Agathos, M.; et al. GW190814:gravitational waves from the coalescence of a 23 solar mass black hole with a 2.6 solar mass compact object. Astrophys. J. Lett. 2020, 896, L44. [Google Scholar] [CrossRef]
  154. Smirnova, T.V.V.L.; Shishov, W.; Sieber, D.R.; Stinebring, V.M.; Malofeev, V.A.; Potapov, S.A.; Bashev, T.; Jessner, A.; Wielebinski, R. The interstellar turbulent plasma spectrum in the direction to PSR B1642-03 from multi-frequency observations of interstellar scintillation. Astron. Astrophys. 2006, 455, 195–201. [Google Scholar] [CrossRef]
Figure 1. (a) The first fundamental derivative for the multicomponent van der Waals equation of state showing positive and negative BZT regions for different strengths, densities, and relativistic corrections with no mixing, kij = 0, where going from blue to yellow to green is for increasing values of aij and bk. (b) The fundamental derivative boundary between the BZT zone and classical zone and the relativistic contribution to the total for amix = 0.2 GeV−2 fm3, bmix = 0.02 fm3, amix = 0.1 GeV−2 fm3, bmix = 0.02 fm3, and amix = 0.3 GeV−2 fm3, and bmix = 0.02 fm3 with kij = 0, where the blue curve is the classical contribution, the yellow curve is only the relativistic contribution and the green curve is the total.
Figure 1. (a) The first fundamental derivative for the multicomponent van der Waals equation of state showing positive and negative BZT regions for different strengths, densities, and relativistic corrections with no mixing, kij = 0, where going from blue to yellow to green is for increasing values of aij and bk. (b) The fundamental derivative boundary between the BZT zone and classical zone and the relativistic contribution to the total for amix = 0.2 GeV−2 fm3, bmix = 0.02 fm3, amix = 0.1 GeV−2 fm3, bmix = 0.02 fm3, and amix = 0.3 GeV−2 fm3, and bmix = 0.02 fm3 with kij = 0, where the blue curve is the classical contribution, the yellow curve is only the relativistic contribution and the green curve is the total.
Astronomy 04 00014 g001
Figure 2. Maximum compactness Cmax = M/R over the multicomponent van der Waals (MvdW) EOS parameter space, shown as a color map (capped at 0.8) as a function of: amix [GeV−2 fm3]—the strength of attractive interactions, bmix [fm3]—the repulsive excluded volume parameter. White contour lines represent constant compactness from 0.04 < C < 0.28, highlighting how EOS stiffness evolves across microphysical interaction space. Compactness is a critical structural parameter governing tidal deformability, gravitational wave phase shifts, and stellar stability. Higher compactness correlates with more compact and tightly bound neutron stars, while lower values are indicative of more deformable configurations. The larger rectangular boxed region denotes the BZT zone, defined by −0.15 ≤ amix ≤ 0.4 and 0.5 ≤ bmix ≤ 1.00, where the EOS exhibits thermodynamic nonconvexity (i.e., negative fundamental derivative G < 0), allowing the formation of compound shock structures and potentially richer dynamics in dense matter. The blue rectangle indicates the optimal region for the MvdW EOS based on physical viability, quantum-statistical consistency, and agreement with observations: 0.2 ≤ amix ≤ 0.45 and 0.7 ≤ bmix ≤ 1.0 for the total range of values in the MvdW EOS.This range is compatible with quantum field theory-inspired models and provides causal, stable stellar solutions. Cyan squares mark the projected EOS parameter locations of several observed neutron stars: PSR J0030+0451, PSR J0740+6620, PSR J1614–2230, PSR J0952–0607, PSR J0348+0432, PSR J2124–3358, PSR J1909–3744, as referenced in Appendix C. Most stars lie within or near the optimal MvdW region, indicating these EOSs are consistent with both microscopic physics and macroscopic observables. Stars outside this region (e.g., J2124–3358, J1909–3744, as referenced in Appendix C) may probe EOS behaviors beyond QCD-calibrated interactions, offering valuable insight into the extremes of dense matter physics. For references on sources, see Appendix C.
Figure 2. Maximum compactness Cmax = M/R over the multicomponent van der Waals (MvdW) EOS parameter space, shown as a color map (capped at 0.8) as a function of: amix [GeV−2 fm3]—the strength of attractive interactions, bmix [fm3]—the repulsive excluded volume parameter. White contour lines represent constant compactness from 0.04 < C < 0.28, highlighting how EOS stiffness evolves across microphysical interaction space. Compactness is a critical structural parameter governing tidal deformability, gravitational wave phase shifts, and stellar stability. Higher compactness correlates with more compact and tightly bound neutron stars, while lower values are indicative of more deformable configurations. The larger rectangular boxed region denotes the BZT zone, defined by −0.15 ≤ amix ≤ 0.4 and 0.5 ≤ bmix ≤ 1.00, where the EOS exhibits thermodynamic nonconvexity (i.e., negative fundamental derivative G < 0), allowing the formation of compound shock structures and potentially richer dynamics in dense matter. The blue rectangle indicates the optimal region for the MvdW EOS based on physical viability, quantum-statistical consistency, and agreement with observations: 0.2 ≤ amix ≤ 0.45 and 0.7 ≤ bmix ≤ 1.0 for the total range of values in the MvdW EOS.This range is compatible with quantum field theory-inspired models and provides causal, stable stellar solutions. Cyan squares mark the projected EOS parameter locations of several observed neutron stars: PSR J0030+0451, PSR J0740+6620, PSR J1614–2230, PSR J0952–0607, PSR J0348+0432, PSR J2124–3358, PSR J1909–3744, as referenced in Appendix C. Most stars lie within or near the optimal MvdW region, indicating these EOSs are consistent with both microscopic physics and macroscopic observables. Stars outside this region (e.g., J2124–3358, J1909–3744, as referenced in Appendix C) may probe EOS behaviors beyond QCD-calibrated interactions, offering valuable insight into the extremes of dense matter physics. For references on sources, see Appendix C.
Astronomy 04 00014 g002
Figure 3. Gravitational wave phase shift ΔΨ 100 Hz, shown as log10∣ΔΨ∣ (in radians), over the multicomponent van der Waals (MvdW) EOS parameter space. The plot spans the microphysical interaction parameters: amix [GeV−2 fm3]: attractive interaction strength bmix [fm3]: repulsive excluded volume. White dashed contours indicate lines of constant compactness C = M/R, a key structural property controlling tidal deformability and gravitational wave signatures. The color map encodes the strength of the phase shift ΔΨ, which reflects the EOS response to tidal fields in inspiraling neutron star binaries. The black rectangle bounds the QCD-motivated EOS region: 0.1 ≤ amix ≤ 0.50, 0.3 ≤ bmix ≤ 1.2. These limits are based on lattice QCD and effective field theory models. The shaded region marks the BZT zone, where the EOS becomes thermodynamically nonconvex (negative fundamental derivative G < 0), enabling compound shocks and nonlinear wave behavior. This zone overlaps with many physically viable EOSs, and its proximity to observed GRB estimates suggests it may be astrophysically relevant. Cyan markers identify EOS locations associated with observed GRB-producing neutron star mergers, including: GRB170817A, GRB190425, GRB211211A, GRB230307A, GRB221009A, GRB150101B, see Appendix C for references. These events fall near or within the BZT region, within the phase shift detection threshold of LIGO/Virgo at 100 Hz—where their interferometric sensitivity is highest. Accurate modeling at this frequency is essential for resolving tidal effects and inferring EOS parameters from the gravitational waveform. While LIGO/Virgo is sensitive to phase shifts down to ∼10−2 rad, future detectors like LISA and the Einstein Telescope will push detection to ∼10−5 to 10−4 rad, requiring even finer constraints on EOS structure, especially in the low-compactness regime. This figure illustrates how GW phase observations, when mapped to microphysical EOS parameters, provide a bridge between QCD-scale interactions and astrophysical neutron star data. For references on sources, see Appendix C.
Figure 3. Gravitational wave phase shift ΔΨ 100 Hz, shown as log10∣ΔΨ∣ (in radians), over the multicomponent van der Waals (MvdW) EOS parameter space. The plot spans the microphysical interaction parameters: amix [GeV−2 fm3]: attractive interaction strength bmix [fm3]: repulsive excluded volume. White dashed contours indicate lines of constant compactness C = M/R, a key structural property controlling tidal deformability and gravitational wave signatures. The color map encodes the strength of the phase shift ΔΨ, which reflects the EOS response to tidal fields in inspiraling neutron star binaries. The black rectangle bounds the QCD-motivated EOS region: 0.1 ≤ amix ≤ 0.50, 0.3 ≤ bmix ≤ 1.2. These limits are based on lattice QCD and effective field theory models. The shaded region marks the BZT zone, where the EOS becomes thermodynamically nonconvex (negative fundamental derivative G < 0), enabling compound shocks and nonlinear wave behavior. This zone overlaps with many physically viable EOSs, and its proximity to observed GRB estimates suggests it may be astrophysically relevant. Cyan markers identify EOS locations associated with observed GRB-producing neutron star mergers, including: GRB170817A, GRB190425, GRB211211A, GRB230307A, GRB221009A, GRB150101B, see Appendix C for references. These events fall near or within the BZT region, within the phase shift detection threshold of LIGO/Virgo at 100 Hz—where their interferometric sensitivity is highest. Accurate modeling at this frequency is essential for resolving tidal effects and inferring EOS parameters from the gravitational waveform. While LIGO/Virgo is sensitive to phase shifts down to ∼10−2 rad, future detectors like LISA and the Einstein Telescope will push detection to ∼10−5 to 10−4 rad, requiring even finer constraints on EOS structure, especially in the low-compactness regime. This figure illustrates how GW phase observations, when mapped to microphysical EOS parameters, provide a bridge between QCD-scale interactions and astrophysical neutron star data. For references on sources, see Appendix C.
Astronomy 04 00014 g003
Figure 4. Comparison of GRB luminosity curves from three theoretical EOS models—Multicomponent van der Waals (MvdW), MIT Bag, and Hadron Gas—against observational light curves of three well-studied GRBs: GRB 061007, GRB 080319B, and GRB 090618 [112]. The curves are plotted as rest-frame luminosity versus time using broken power law fits. The MvdW EOS (orange) produces a broad, high-luminosity profile consistent with GRB 090618 (purple), while the MIT Bag EOS (red dashed) matches the steep decay behavior of GRB 061007 (green). The Hadron Gas EOS (cyan dotted) yields a lower-luminosity evolution consistent with more rapidly fading bursts. The inset zooms into the critical break region near ~200 s, where GRB light curves transition from plateau to decay. These comparisons highlight the EOS sensitivity of BZT-driven GRB emissions and demonstrate how phase-transition energetics and shock dynamics map to GRB light curve morphology.
Figure 4. Comparison of GRB luminosity curves from three theoretical EOS models—Multicomponent van der Waals (MvdW), MIT Bag, and Hadron Gas—against observational light curves of three well-studied GRBs: GRB 061007, GRB 080319B, and GRB 090618 [112]. The curves are plotted as rest-frame luminosity versus time using broken power law fits. The MvdW EOS (orange) produces a broad, high-luminosity profile consistent with GRB 090618 (purple), while the MIT Bag EOS (red dashed) matches the steep decay behavior of GRB 061007 (green). The Hadron Gas EOS (cyan dotted) yields a lower-luminosity evolution consistent with more rapidly fading bursts. The inset zooms into the critical break region near ~200 s, where GRB light curves transition from plateau to decay. These comparisons highlight the EOS sensitivity of BZT-driven GRB emissions and demonstrate how phase-transition energetics and shock dynamics map to GRB light curve morphology.
Astronomy 04 00014 g004
Figure 5. A comparison of the maximum MvdW BZT phase shift frequency dependence to the current and estimated sensitivities for the LIGO, VIRGO, KAGRA, Einstein Telescope, and LISA facilities. Details of the various EOS models are given in Appendix A.
Figure 5. A comparison of the maximum MvdW BZT phase shift frequency dependence to the current and estimated sensitivities for the LIGO, VIRGO, KAGRA, Einstein Telescope, and LISA facilities. Details of the various EOS models are given in Appendix A.
Astronomy 04 00014 g005
Table 1. Here are the comparison values for a neutron star and quark star for compactness, Love number, tidal deformability, chirp mass, and the merger tidal deformability.
Table 1. Here are the comparison values for a neutron star and quark star for compactness, Love number, tidal deformability, chirp mass, and the merger tidal deformability.
ParameterNeutron StarsQuark Stars
Compactness (C)0.15–0.250.2–0.35
Love Number (k2)0.05–0.150.01–0.05
Tidal Deformability (Λ)200–100010–100
Chirp Mass (M in solar masses)1.18–1.200.96–1.15
Merger Tidal Deformability ( Λ ˜ )70–72010–100
Table 2. Here we collect the typical response characteristics of detectors that can find evidence for the predicted phase shift. The KAGRA data is for the current upgrade. Frequency in Hertz (Hz), Strain is the minimum detectable dimensionless strain amplitude as a measure of spacetime deformation, estimated signal to noise ratio (SNR) for each detector, and phase shift sensitivity is the minimum detectable phase distortion in radians.
Table 2. Here we collect the typical response characteristics of detectors that can find evidence for the predicted phase shift. The KAGRA data is for the current upgrade. Frequency in Hertz (Hz), Strain is the minimum detectable dimensionless strain amplitude as a measure of spacetime deformation, estimated signal to noise ratio (SNR) for each detector, and phase shift sensitivity is the minimum detectable phase distortion in radians.
DetectorFrequency Range (Hz)Sensitivity (Strain)SNR ThresholdApplicationPhase Shift Sensitivity
1LIGO10–1000~10−23~8Binary mergers, neutron stars, black holes10−2
2Virgo10–1000~10−23~8Binary mergers, neutron stars, black holes10−2
3LISA0.01–1~10−21~1Supermassive black holes, early universe10−5
4Einstein
Telescope
1–1000~10−24~5Binary mergers, cosmological signals10−4
5KAGRA10–1000~10−23~8Binary mergers, neutron stars, black holes10−3
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

Andrew, K.; Steinfelds, E.V.; Andrew, K.A. Gravity Wave Phase Shift in a Cold Quark Star with a Nonconvex QCD BZT Shock Wave Van Der Waals Equation of State. Astronomy 2025, 4, 14. https://doi.org/10.3390/astronomy4030014

AMA Style

Andrew K, Steinfelds EV, Andrew KA. Gravity Wave Phase Shift in a Cold Quark Star with a Nonconvex QCD BZT Shock Wave Van Der Waals Equation of State. Astronomy. 2025; 4(3):14. https://doi.org/10.3390/astronomy4030014

Chicago/Turabian Style

Andrew, Keith, Eric V. Steinfelds, and Kristopher A. Andrew. 2025. "Gravity Wave Phase Shift in a Cold Quark Star with a Nonconvex QCD BZT Shock Wave Van Der Waals Equation of State" Astronomy 4, no. 3: 14. https://doi.org/10.3390/astronomy4030014

APA Style

Andrew, K., Steinfelds, E. V., & Andrew, K. A. (2025). Gravity Wave Phase Shift in a Cold Quark Star with a Nonconvex QCD BZT Shock Wave Van Der Waals Equation of State. Astronomy, 4(3), 14. https://doi.org/10.3390/astronomy4030014

Article Metrics

Back to TopTop