This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

g-mode oscillations in hybrid stars:
A tale of two sounds

Prashanth Jaikumar [email protected] Department of Physics and Astronomy, California State University Long Beach, Long Beach, California 90840, USA    Alexandra Semposki [email protected]    Madappa Prakash [email protected] Department of Physics and Astronomy, Ohio University, Athens, Ohio 45701, USA    Constantinos Constantinou [email protected] INFN-TIFPA, Trento Institute of Fundamental Physics and Applications, Povo, 38123 Trento, Italy European Centre for Theoretical Studies in Nuclear Physics and Related Areas, Villazzano, 38123 Trento, Italy
Abstract

We study the principal core gg-mode oscillation in hybrid stars containing quark matter and find that they have an unusually large frequency range (\approx 200 - 600 Hz) compared to ordinary neutron stars or self-bound quark stars of the same mass. Theoretical arguments and numerical calculations that trace this effect to the difference in the behaviour of the equilibrium and adiabatic sound speeds in the mixed phase of quarks and nucleons are provided. We propose that the sensitivity of core gg-mode oscillations to non-nucleonic matter in neutron stars could be due to the presence of a mixed quark-nucleon phase. Based on our analysis, we conclude that for binary mergers where one or both components may be a hybrid star, the fraction of tidal energy pumped into resonant gg-modes in hybrid stars can exceed that of a normal neutron star by a factor of 2-3, although resonance occurs during the last stages of inspiral. A self-bound star, on the other hand, has a much weaker tidal overlap with the gg-mode. The cumulative tidal phase error in hybrid stars, Δϕ\Delta\phi\cong 0.5 rad, is comparable to that from tides in ordinary neutron stars, presenting a challenge in distinguishing between the two cases. However, should the principal gg-mode be excited to sufficient amplitude for detection in a postmerger remnant with quark matter in its interior, its frequency would be a possible indication for the existence of non-nucleonic matter in neutron stars.

preprint: APS/123-QED

I Introduction

The core of a neutron star (NS) can, in principle, support phases of dense deconfined quark matter (QM) [1]. Confirmation of the presence of quarks in NSs, however, has not been possible either through observations or from lattice-gauge calculations of finite baryon density matter starting from the Lagrangian of Quantum Chromodynamics (QCD). Although perturbative calculations of QM have been performed [2, 3, 4], their applicability is limited to baryon densities nB40nsn_{B}\gtrsim 40n_{s} [5], where ns0.16fm3n_{s}\simeq 0.16~{}\rm{fm^{-3}} is the nuclear matter saturation density. Such densities, however, lie well beyond the central densities ncn_{c} in the range 3-8nsn_{s} of observed NSs. In view of this conundrum, theoretical studies of QM in NSs have been exploratory in nature by positing either a sharp 1st-order or a smooth crossover hadron-to-quark phase transition. Depending on the treatment of the phase transition and the equations of state (EOSs) of hadronic and quark matter, either a phase of pure QM or a phase in which hadrons are admixed with quarks can be realized (for a detailed account, see Ref. [6] and an extensive list of references therein). In either case, stars with quarks are difficult to distinguish from normal NSs based on the knowledge of masses and radii alone as similar results can be obtained with both. While the long-term cooling of a NS can be affected by the presence of quarks, cooling data are relatively sparse and gathered over decades [7, 8]. Gravitational wave observations from compact binary mergers can be another probe of the EOS, but currently, constraints on tidal polarizability [9, 10, 11] from gravitational wave data [12] are consistent with both normal and quark-containing stars, depending on the theoretical assumptions made [6, 13, 14].

In this paper, we are particularly interested in how NS oscillations can shed light on the presence of QM in stars that contain an admixture of nucleons and quarks (termed hybrid stars). Andersson and Kokkotas [15] have proposed that NS oscillations (in particular, the f,pf,p modes) could be a “fingerprint” for the supra-nuclear EOS in gravitational wave data. A review of potential signatures of QM in NSs in the multi-messenger era, including the role of their oscillations, can be found in [16]. Along these lines, we offer in this work a new diagnostic of deconfined QM in NSs based on asteroseismology. We show that a steep rise in the frequency of the principal gg-mode (gravity mode) occurs as soon as QM appears in a mixed phase in the core, exceeding the typical core gg-mode frequency of a nucleonic star by a factor of two or more. This rise is essentially driven by a drop in the local equilibrium speed of sound at the onset of the phase transition, while the adiabatic sound speed changes only slightly. If this gg-mode becomes resonant with the tidal force during the late stages of a binary inspiral, the resulting energy transfer from the orbital motion to the star via tidal coupling can affect the phase of the gravitational waveform, and potentially signal a hybrid star.

NS oscillations are categorized by the nature of the dominant restoring force for the perturbation in question. Several types of modes can be supported by a star and it is desirable to investigate as many of them as possible in detail. These modes are typically excited and sustained in different regions of the star and their amplitudes and damping rates are subject to considerable uncertainty. Here, for reasons that will become apparent, we focus our attention on the gg-mode and its coupling to gravitational waves.

A gg-mode is a specific kind of non-radial fluid oscillation initiated when a parcel of fluid is displaced against the background of a stratified environment [17, 18]. While pressure equilibrium is rapidly restored via sound waves, chemical equilibrium can take longer causing buoyancy forces to oppose the displacement. Since cold NSs are not convective, the opposing force sets up stable oscillations throughout the core, with a typical frequency, called the (local) Brunt-Väisälä  frequency [19], which depends on the difference between the equilibrium and adiabatic sound speeds as well as the local metric coefficients. Convectively stable gg-modes exist for a wide range of models of the EOS [20]. Though the gg-mode in NSs has been studied before [21, 22, 23], with recent works incorporating additional features like hyperonic matter and superfluidity [24, 25, 26, 27, 28, 29, 30, 31, 32, 33, 34], the novel aspect of our work is the investigation of the gg-mode frequency in a phase transition from nuclear matter to a mixed phase of quarks and nucleons. We point out that a similar result was obtained in [30] for superfluid hyperonic stars. However, the calculations presented there were for a fixed NS mass of about 1.64MM_{\odot} with a radius of 13.6 km. While their chosen hyperonic EOS does support a maximum mass of 2.015M2.015M_{\odot} [35], the nuclear and quark EOSs chosen in our work satisfy additional observational and experimental constraints, and presents the effect for a wide range of masses up to the observed maximum.

This paper also extends the results of [36] by incorporating aspects of General Relativity in the fluid oscillation equations (while remaining within the Cowling approximation), updating the nuclear EOS to include consistency with radius constraints from tidal polarizability [37] and NICER data [38, 39] on the radius of 1.4M\simeq 1.4M_{\odot} NS. We also provide new analytical results for the two sound speeds in a mixed phase of quarks and nucleons. Our study can be of practical interest to investigations of the sound speed in NSs, which is attracting renewed attention [40, 41, 42, 43]. The matter of detecting gg-modes from the pre- or post-coalescence phase of binary NS mergers is not addressed in detail here, but we present an estimate of its impact on the accumulated tidal phase up to the merger.

It is pertinent to note that oscillation modes other than gg-modes can also potentially be affected by the presence of QM in NSs. Radial oscillation modes, which however do not couple to gravitational waves, were studied in [44]. Among non-radial modes, the ii-mode (interface mode) has been recently investigated for the special case of crystalline QM surrounded by a hadronic envelope in [45] and its frequency can range from (300-1500) Hz, which can be probed by current or third generation interferometric detectors. The rr-mode (Rossby mode) frequency and its damping rate for NSs containing QM also differs from a purely nucleonic one [46, 47, 48]. The ss-mode (shear mode) can be excited in a mixed phase of quarks and nucleons and is sensitive to the shear modulus of structures in the mixed phase [49], probing the surface tension of QM. The gg-mode oscillation in stars containing QM has been studied in the case of a sharp interface [50, 51] between hadronic and quark matter, yielding the spectrum of so-called discontinuity gg-modes, but these works assume a strong 1st-order phase transition and a large value of the surface tension for QM, while we study the case of a mixed phase of significant extent that would be favored if the same surface tension were small enough111Here, we do not explicitly study surface and curvature effects or the impact of a non-trivially structured mixed phase on the oscillation spectrum [52, 49, 53], but a more complete treatment of gg-modes in hybrid stars should address these issues.. The gg-mode for baryonic stars with superfluid effects was studied in [24, 25, 26, 27, 28, 29, 30, 31, 32, 33, 34] highlighting the subtle role of temperature and composition gradients in driving the gg-mode. In this work, we investigate the composition gradients induced by the mixed phase of quarks and nucleons, which supports unusually high-frequency gg-modes through its effect on the adiabatic and equilibrium sound speeds.

The organization of this paper is as follows. In Sec. II, we introduce the governing equations of the gg-mode and outline its relation to the two sound speeds. In Sec. III, we present the EOS models in the nucleonic and quark sectors chosen for our study of the gg-mode spectrum. The rationale for our parameter choices and the basic features of these models are highlighted here for orientation. We stress that our choices are representative, but not exhaustive. In Sec. IV, we derive expressions for the two sound speeds in the mixed phase of quarks and nucleons. Results for the sound speeds, the Brunt-Väisälä  frequency and the gg-mode frequency in nucleonic, quark and hybrid stars are gathered and interpreted in Sec. V. The tidal forcing and phase error due to gg-mode excitation are estimated in Sec. VI. Our summary, conclusions and outlook are contained in Sec. VII. The appendices contain details about the determination of parameters in the nuclear EOS model, the resulting NS structural properties and the two sound speeds.

II gg-mode Oscillations

In this section, we outline the equations for fluid oscillations and non-rotating NS structure that were used to determine the eigenfrequencies of the gg-mode. In general, the oscillatory displacement of a fluid element in a spherically symmetric star is represented by a vector field ξnlm(r)eiωt{\vec{\xi}}^{nlm}(\vec{r}){\rm e}^{-i\omega t} with n,ln,l and mm denoting the radial, azimuthal and magnetic mode indices. To be precise, the frequencies also carry subscripts nlmnlm implicitly understood, with degeneracies that are broken in more realistic cases such as with rotation or magnetic fields. For even-parity or spheroidal modes, separation into radial and tangential components yields ξrnlm(r)\xi_{r}^{nlm}(\vec{r}) = ηrnl(r)Ylm(θ,ϕ)\eta_{r}^{nl}(r)Y_{lm}(\theta,\phi) and ξnlm(r)\xi_{\perp}^{nlm}(\vec{r}) = rηnl(r)Ylm(θ,ϕ)r\eta_{\perp}^{nl}(r)\nabla_{\perp}Y_{lm}(\theta,\phi), respectively, where Ylm(θ,ϕ)Y_{lm}(\theta,\phi) are the spherical harmonics. From the perturbed continuity equation for the fluid, the tangential function η\eta_{\perp} can be traded for fluid variables as δp/ϵ\delta p/\epsilon = ω2rη(r)Ylm(θ,ϕ)eiωt\omega^{2}r\eta_{\perp}(r)Y_{lm}(\theta,\phi){\rm e}^{-i\omega t}, where δp\delta p is the corresponding local (Eulerian) pressure perturbation and ϵ\epsilon the local energy density. Within the relativistic Cowling approximation222The Cowling approximation neglects the back reaction of the gravitational potential and reduces the number of equations we have to solve. While this approximation is not strictly consistent with our fully general relativistic (GR) treatment of the equilibrium structure of the star, it does not change our conclusions qualitatively or even quantitatively, since this approximation is accurate for gg-mode frequencies at the few % level [54]., the equations of motion to be solved to determine the frequency of a particular mode are [55, 17, 29]

1eλ/2r2r[eλ/2r2ξr]+l(l+1)eνr2ω2δpp+ϵΔpγp=0\displaystyle-\frac{1}{\mathrm{e}^{\lambda/2}r^{2}}\frac{\partial}{\partial r}\left[\mathrm{e}^{\lambda/2}r^{2}\xi_{r}\right]+\frac{l(l+1)\mathrm{e}^{\nu}}{r^{2}\omega^{2}}\frac{\delta p}{p+\epsilon}-\frac{\Delta p}{\gamma p}=0
δpr+g(1+1cs2)δp+eλνh(N2ω2)ξr=0,\displaystyle\frac{\partial\delta p}{\partial r}+g\left(1+\frac{1}{c_{\mathrm{s}}^{2}}\right)\delta p+\mathrm{e}^{\lambda-\nu}h\left(N^{2}-\omega^{2}\right)\xi_{r}=0\,, (1)

where h=p+ϵh=p+\epsilon, and we have suppressed the indices on ω\omega and ξ\xi. Equation (1) involves thermodynamic quantities that follow from the specific EOS. Specifically, pp denotes pressure, ϵ\epsilon energy density, and γ\gamma the adiabatic index of the fluid. The Lagrangian variation of the pressure enters as Δp\Delta p, and is related to the Eulerian variation δp\delta p through the operator relation Δδ+ξ\Delta\equiv\delta+\xi\cdot\nabla. The symbol csc_{s} denotes the adiabatic sound speed, which is related to the adiabatic index as cs2=γp/(μnnB)c_{s}^{2}=\gamma p/(\mu_{n}n_{B}) where μn\mu_{n} is the neutron chemical potential333In beta-equilibrated charge neutral matter, the neutron chemical potential is sufficient to determine all other chemical potentials. and nBn_{B} the local baryon density. The equilibrium sound speed enters through the Brunt-Väisälä  frequency (NN) which is given by

N2g2(1ce21cs2)eνλ,\displaystyle N^{2}\equiv g^{2}\Big{(}\frac{1}{c_{e}^{2}}-\frac{1}{c_{s}^{2}}\Big{)}{\rm e}^{\nu-\lambda}\,, (2)

where g=ϕ=p/hg=-\nabla\phi=-\nabla p/h with h=ϵ+ph=\epsilon+p the enthalpy of the fluid. Finally, ν(r)\nu(r) and λ(r)\lambda(r) are metric functions of the unperturbed star which features in the Schwarzschild interior metric (r<Rr<R):

ds2gαβdxαdxβ=\displaystyle-\mathrm{d}s^{2}\equiv\,g_{\alpha\beta}\mathrm{d}x^{\alpha}\mathrm{d}x^{\beta}= eν(r)dt2+eλ(r)dr2\displaystyle-\mathrm{e}^{\nu(r)}\mathrm{d}t^{2}+\mathrm{e}^{\lambda(r)}\mathrm{d}r^{2} (3)
+r2(dθ2+sin2θdφ2).\displaystyle+r^{2}\left(\mathrm{d}\theta^{2}+\sin^{2}\theta\mathrm{d}\varphi^{2}\right).

Explicitly,

eλ(r)=11(2Gm(r)c2r)\displaystyle e^{\lambda(r)}=\frac{1}{1-\left(\frac{2Gm(r)}{c^{2}r}\right)} (4)

and

eν(r)=exp[2Gc20r((m(r)+4πp(r)r3c2)r(r2m(r)Gc2))𝑑r]eν0,\displaystyle e^{\nu(r)}=\exp\bigg{[}-\frac{2G}{c^{2}}\int_{0}^{r}\left(\frac{\left(m(r^{\prime})+\frac{4\pi p(r^{\prime})r^{\prime 3}}{c^{2}}\right)}{r^{\prime}\left(r^{\prime}-\frac{2m(r^{\prime})G}{c^{2}}\right)}\right)dr^{\prime}\bigg{]}{\rm e}^{\nu_{0}}, (5)

where m(r)m(r^{\prime}) is the enclosed mass of the star at rr^{\prime}. These metric functions must match to their exterior values at the surface r=Rr=R, hence the constant factor eν0{\rm e}^{\nu_{0}} [56].

In this work, we study the fundamental gg-mode with nn = 1 and fix the mode’s multipolarity at ll = 2. For the non-rotating stars we consider here, solutions are degenerate in mm. Note that our definition of the “fundamental” mode refers to the lowest radial order of the gg-mode which also has the highest frequency. This should not be confused with the qualitatively different ff-mode which is also referred to sometimes as the fundamental mode. Furthermore, overtones with lower frequency exist, but we do not perform any computations with them here, since the fundamental gg-mode has the highest frequency and will be excited during the final stage of the pre-merger phase when tidal forces are strongest. The system of equations in Eq. (1) cannot be solved analytically even with a simple model of a neutron star. Our aim will be to solve this numerically as an eigenvalue system for the gg-mode frequency ω\omega. Physically, the solution to this system of equations, under the boundary conditions Δp=0\Delta p=0 at the surface and ξr,δp/ϵ\xi_{r},\,\delta p/\epsilon regular at the center, only exists for discrete values of the mode frequency ω\omega. These values represent the gg-mode spectrum for a chosen stellar model. Because we have employed the Cowling approximation and ignored the perturbations of the metric that must accompany fluid perturbations, we cannot compute the imaginary part of the eigenfrequency (damping time) of the gg-mode444The damping time of gg-modes due to viscosity and gravitational wave emission, crudely estimated in [57, 36], suggests that the gg-mode can become secularly unstable for temperatures 108K<T<109K10^{8}~{}{\rm K}<T<10^{9}~{}{\rm K} for rotational speeds exceeding twice the gg-mode frequency of a static star.. We turn now to discuss the EOS models for nucleonic and quark matter employed in this work.

III Models for the Equation of State

The EOS models chosen in this work were predicated on the requirement that the squared sound speeds ce2c_{e}^{2} (see Eq.(25)) and cs2c_{s}^{2} could be calculated straightforwardly. In the nucleonic sector, we employ the model of Zhao and Lattimer (ZL) [43] which is consistent with nuclear systematics at and below the nuclear saturation density nsn_{s}. With suitable choices of the slope of the nuclear symmetry energy at nsn_{s} (see below), this EOS is also consistent with the recent chiral effective theory calculations of Drischler et al. [58] in which uncertainty estimates of the EOS up to 2ns2n_{s} were provided (see Fig. 2 of this reference). In addition, the ZL EOS is able to support 2M\simeq 2M_{\odot} stars required by mass measurements of heavy NSs [59], and is consistent with the recent radius measurements of 1.4M\sim 1.4M_{\odot} stars [38, 39] and the tidal deformability estimates from the binary neutron star merger GW170817 [37].

Among the many models and treatments available in the quark sector [6], we utilize the vMIT model of Gomes et al. [60] as a caricature of strongly interacting quarks at the densities attained within NSs. Such interactions between quarks are required to satisfy astrophysical data, particularly those of heavy mass NSs. For the treatment of the nucleon-to-quark transition at supra-nuclear densities, we employ the Gibbs construction [61] which renders the transition to be smooth. Alternative models and treatments that feature strong first- or second-order phase transitions will be undertaken in subsequent work.

III.1 The ZL EOS for Nucleonic Matter

For completeness and to set the stage for the calculation of the two sound speeds in the next section, relevant details of the ZL model are provided below. The total energy density of interacting nucleons in neutron star matter (NSM) is

ϵB=\displaystyle\epsilon_{B}= i=n,p1π20kFik2MB2+k2𝑑k\displaystyle\sum_{i=n,p}\frac{1}{\pi^{2}}\int_{0}^{k_{Fi}}k^{2}\sqrt{M_{B}^{2}+k^{2}}\,dk
+nBV(nn,np),\displaystyle+n_{B}V(n_{n},n_{p})\,, (6)

where the Fermi momenta kFi=(3π2ni)1/3k_{Fi}=(3\pi^{2}n_{i})^{1/3} with i=n,pi=n,p and nB=nn+npn_{B}=n_{n}+n_{p}, and MBM_{B} is the baryon mass in vacuum. In the ZL model, interactions between nucleons are written as

V(nn,np)V(u,x)=\displaystyle V(n_{n},n_{p})\equiv V(u,x)=  4x(1x)(a0u+b0uγ)\displaystyle\,4x(1-x)(a_{0}u+b_{0}u^{\gamma})
+(12x)2(a1u+b1uγ1),\displaystyle+(1-2x)^{2}(a_{1}u+b_{1}u^{\gamma_{1}}), (7)

where u=nB/nsu=n_{B}/n_{s} and the proton fraction x=np/nBx=n_{p}/n_{B}. Adding and subtracting a0u+b0uγa_{0}u+b_{0}u^{\gamma}, the above equation can be rewritten as

V(u,x)\displaystyle V(u,x) =V0+S2i(u)(12x)2\displaystyle=V_{0}+S_{2i}(u)(1-2x)^{2}\quad
with
V0\displaystyle V_{0} =a0u+b0uγ,\displaystyle=a_{0}u+b_{0}u^{\gamma},\quad
S2i(u)\displaystyle\quad S_{2i}(u) =(a1a0)u+b1uγ1b0uγ,\displaystyle=(a_{1}-a_{0})u+b_{1}u^{\gamma_{1}}-b_{0}u^{\gamma}\,, (8)

where the subscript “2i2i” in S2iS_{2i} refers to the interacting part of the total symmetry energy S2=S2k+S2iS_{2}=S_{2k}+S_{2i}, with S2kS_{2k} representing the kinetic part. Expanding the kinetic part in Eq. (6) to order (12x)2(1-2x)^{2}, we obtain the result555For the derivation of the kinetic part of the symmetry energy and its derivatives, see the Appendix.

S2k=18[1n2ϵBkx2]x=12=kF26EF,\displaystyle S_{2k}=\frac{1}{8}\left[\frac{1}{n}\frac{\partial^{2}\epsilon_{Bk}}{\partial x^{2}}\right]_{x=\frac{1}{2}}=\frac{k_{F}^{2}}{6E_{F}}\,, (9)

where kF=(3π2nB/2)1/3k_{F}=(3\pi^{2}n_{B}/2)^{1/3} is the Fermi wave number of symmetric nuclear matter (SNM) and EF=kF2+MB2E_{F}=\sqrt{k_{F}^{2}+M_{B}^{2}}. Collecting the results, the energy per baryon relative to MBM_{B} is given by

ϵBnBMB\displaystyle\frac{\epsilon_{B}}{n_{B}}-M_{B} =\displaystyle= E(u,x)=ESNM+S2(u)(12x)2\displaystyle E(u,x)=E_{{\rm SNM}}+S_{2}(u)(1-2x)^{2}
where
ESNM\displaystyle E_{{\rm SNM}} =\displaystyle= T1/2+V1/2=T1/2+(a0u+b0uγ),\displaystyle T_{1/2}+V_{1/2}=T_{1/2}+(a_{0}u+b_{0}u^{\gamma}),
S2(u)\displaystyle S_{2}(u) =\displaystyle= kF26EF+(a1a0)u+b1uγ1b0uγ.\displaystyle\frac{k_{F}^{2}}{6E_{F}}+(a_{1}-a_{0})u+b_{1}u^{\gamma_{1}}-b_{0}u^{\gamma}\,. (10)

The kinetic energy per baryon T1/2T_{1/2} in SNM (x=1/2x=1/2) is given by the expression

T1/2\displaystyle T_{1/2} =\displaystyle= ϵ1/2kinnBMBwith\displaystyle\frac{\epsilon_{1/2}^{\rm kin}}{n_{B}}-M_{B}\quad{\rm with}
ϵ1/2kin\displaystyle\epsilon_{1/2}^{\rm kin} =\displaystyle= 24π2[kFEF(kF2+MB22)\displaystyle\frac{2}{4\pi^{2}}\bigg{[}k_{F}E_{F}\left(k_{F}^{2}+\frac{M_{B}^{2}}{2}\right) (11)
\displaystyle- 12MB4ln(kF+EFMB)],\displaystyle\frac{1}{2}M_{B}^{4}\ln\left(\frac{k_{F}+E_{F}}{M_{B}}\right)\bigg{]}\,,

where nBn_{B}, kFk_{F} and EFE_{F} refer to those in SNM.

The baryon pressure pBp_{B} is

pB=nsu2dEBdu=pSNM+nsu2(12x)2dS2(u)du,\displaystyle p_{B}=n_{s}u^{2}\frac{dE_{B}}{du}=p_{\rm SNM}+n_{s}u^{2}(1-2x)^{2}\frac{dS_{2}(u)}{du}\,, (12)

where

pSNM\displaystyle p_{SNM} =\displaystyle= p1/2kin+ns(a0u2+γb0uγ+1),with\displaystyle p_{1/2}^{\rm kin}+n_{s}~{}(a_{0}u^{2}+\gamma b_{0}u^{\gamma+1}),\quad{\rm with}
p1/2kin\displaystyle p_{1/2}^{\rm kin} =\displaystyle= 212π2[kFEF(kF232MB2)\displaystyle\frac{2}{12\pi^{2}}\bigg{[}k_{F}E_{F}\left(k_{F}^{2}-\frac{3}{2}M_{B}^{2}\right)
+\displaystyle+ 32MB4ln(kF+EFMB)],and\displaystyle\frac{3}{2}M_{B}^{4}\ln\left(\frac{k_{F}+E_{F}}{M_{B}}\right)\bigg{]},\quad{\rm and}
udS2(u)du\displaystyle u\frac{dS_{2}(u)}{du} =\displaystyle= 23S2k[118(S2kkF)2]+(a1a0)u\displaystyle\frac{2}{3}S_{2k}\left[1-18\left(\frac{S_{2k}}{k_{F}}\right)^{2}\right]+(a_{1}-a_{0})u (13)
+\displaystyle+ b1γ1uγ1b0γuγ.\displaystyle b_{1}\gamma_{1}u^{\gamma_{1}}-b_{0}\gamma u^{\gamma}.

The incompressibility KBK_{B} in SNM is obtained from

KB\displaystyle K_{B} =\displaystyle= 9dpBdnB=9[2udEBdu+u2d2EBdu2]\displaystyle 9\frac{dp_{B}}{dn_{B}}=9\left[2u\frac{dE_{B}}{du}+u^{2}\frac{d^{2}E_{B}}{du^{2}}\right] (14)
=\displaystyle= 9{kF23EF+[2a0u+γ(γ+1)b0uγ]}\displaystyle 9\left\{\frac{k_{F}^{2}}{3E_{F}}+\left[2a_{0}u+\gamma(\gamma+1)b_{0}u^{\gamma}\right]\right\}

The energy per baryon in pure neutron matter (PNM in which x=0x=0) relative to the baryon mass is

EPNM\displaystyle E_{{\rm PNM}} =\displaystyle= T0+V0=T0+(a1u+b1u1γ)\displaystyle T_{0}+V_{0}=T_{0}+(a_{1}u+b_{1}u^{\gamma}_{1})
T0\displaystyle T_{0} =\displaystyle= ϵ0kinnBMBwith\displaystyle\frac{\epsilon_{0}^{\rm kin}}{n_{B}}-M_{B}\quad{\rm with}
ϵ0kin\displaystyle\epsilon_{0}^{\rm kin} =\displaystyle= 14π2[kFnEFn(kFn2+MB22)\displaystyle\frac{1}{4\pi^{2}}\bigg{[}k_{Fn}E_{Fn}\left(k_{Fn}^{2}+\frac{M_{B}^{2}}{2}\right) (15)
\displaystyle- 12MB4ln(kFn+EFnMB)],\displaystyle\frac{1}{2}M_{B}^{4}\ln\left(\frac{k_{Fn}+E_{Fn}}{M_{B}}\right)\bigg{]}\,,

where now nB=nnn_{B}=n_{n}, kFn=(3π2nn)1/3=21/3kFk_{Fn}=(3\pi^{2}n_{n})^{1/3}=2^{1/3}k_{F}, and EFn=kFn2+MB2E_{Fn}=\sqrt{k_{Fn}^{2}+M_{B}^{2}}.

The determination of the EOS constants in SNM and PNM, and relevant NS structural properties are summarized in Appendix A.

III.2 The vMIT Equation of State for Quark Matter

In recent years, variations of the original bag model [62] have been adopted [60, 63] to calculate the structure of NSs with quarks in their cores to account for 2M\geq 2\mathrm{M}_{\odot} maximum-mass stars. Termed as vMIT or vBag models, the QCD perturbative results are dropped and replaced by repulsive vector interactions between quarks in such works. We will provide some numerical examples of the vMIT model for contrast with other models as those of the vBag model turn out to be qualitatively similar.

The Lagrangian density of the vMIT bag model is

=i[ψ¯i(i∂̸miB)ψi+int]Θ,\mathcal{L}=\sum_{i}\left[\bar{\psi}_{i}\left(i\not{\partial}-m_{i}-B\right)\psi_{i}+\mathcal{L}_{\mathrm{int}}\right]\Theta, (16)

where int=pert+vec\mathcal{L}_{\mathrm{int}}=\mathcal{L}_{\mathrm{pert}}+\mathcal{L}_{\mathrm{vec}} describes quarks of mass mim_{i} confined within a bag as denoted by the Θ\Theta function. For three flavors i=u,d,si=u,d,s and three colors Nc=3N_{c}=3 of quarks, the number and baryon densities, energy density, pressure and chemical potentials in the bag model are given by

ni\displaystyle n_{i} =2NckFid3k(2π)3,nB=13ini\displaystyle=2N_{c}\int^{k_{Fi}}\frac{d^{3}k}{(2\pi)^{3}},\quad n_{B}=\frac{1}{3}\sum_{i}n_{i} (17)
ϵQ\displaystyle\epsilon_{Q} =2NcikFid3k(2π)3k2+mi2+ϵint+B\displaystyle=2N_{c}\sum_{i}\int^{k_{Fi}}\frac{d^{3}k}{(2\pi)^{3}}\sqrt{k^{2}+m_{i}^{2}}+\epsilon_{\mathrm{int}}+B (18)
pQ\displaystyle p_{Q} =2Nc3ikFid3k(2π)3k2k2+mi2+pintB\displaystyle=\frac{2N_{c}}{3}\sum_{i}\int^{k_{Fi}}\frac{d^{3}k}{(2\pi)^{3}}\frac{k^{2}}{\sqrt{k^{2}+m_{i}^{2}}}+p_{\mathrm{int}}-B (19)
μi\displaystyle\mu_{i} =kFi2+mi2+μint,i\displaystyle=\sqrt{k_{Fi}^{2}+m_{i}^{2}}+\mu_{\mathrm{int},i} (20)

The upper limit of the integrals kFik_{Fi} is the Fermi wave number for each species ii, which, at zero temperature, appropriately terminates the integration over kk. The first terms in ϵQ\epsilon_{Q} and in pQp_{Q} are free Fermi gas (FG) contributions, ϵFG\epsilon_{\mathrm{FG}} and pFGp_{\mathrm{FG}} respectively, the second terms are due to int \mathcal{L}_{\text{int }} and BB is the bag constant that accounts for the cost of confining the quarks inside a bag. The mim_{i} are quark masses, generally taken to be current quark masses. The uu and dd quark masses are commonly set to zero (because at high density, kFik_{Fi} in these cases far exceed mim_{i}), whereas that of the ss quark is taken at its Particle Data Group (PDG) value. The QCD perturbative calculations of ϵpert \epsilon_{\text{pert }} and ppert p_{\text{pert }}, and the ensuing results for the structure of NSs containing quarks within the cores as well as self-bound strange quark stars are discussed in [5]. At leading order of QCD corrections, the results are qualitatively similar to what one obtains by simply using the FG results with an appropriately chosen value of BB. As results of perturbative calculations are deemed to be valid only for nB40nsn_{B}\geq 40n_{s}, they are dropped in the vMIT model. The Lagrangian density from vector interactions

vec =Gviψ¯γμVμψ+(mV2/2)VμVμ,\quad\mathcal{L}_{\text{vec }}=-G_{v}\sum_{i}\bar{\psi}\gamma_{\mu}V^{\mu}\psi+\left(m_{V}^{2}/2\right)V_{\mu}V^{\mu}\,, (21)

where interactions among the quarks occur via the exchange of a vector-isoscalar meson VμV^{\mu} of mass mV,m_{V}, is chosen in Ref. [54]. Explicitly,

ϵQ\displaystyle\epsilon_{Q} =iϵFG,i+12(GvmV)2nQ2+B\displaystyle=\sum_{i}\epsilon_{\mathrm{FG},\mathrm{i}}+\frac{1}{2}\left(\frac{G_{v}}{m_{V}}\right)^{2}n_{Q}^{2}+B (22)
pQ\displaystyle p_{Q} =ipFG,i+12(GvmV)2nQ2B\displaystyle=\sum_{i}p_{\mathrm{FG},\mathrm{i}}+\frac{1}{2}\left(\frac{G_{v}}{m_{V}}\right)^{2}n_{Q}^{2}-B (23)
μi\displaystyle\mu_{i} =kFi2+mi2+(GvmV)2nQ,\displaystyle=\sqrt{k_{Fi}^{2}+m_{i}^{2}}+\left(\frac{G_{v}}{m_{V}}\right)^{2}n_{Q}\,, (24)

where nQ=ini,n_{Q}=\sum_{i}n_{i}, and the bag constant BB is chosen appropriately to enable a transition to matter containing quarks. Note that terms associated with the vector interaction above are similar to those in hadronic models. We studied model parameters in a wide range B1/4=(155180)MeVB^{1/4}=(155-180)~{}\mathrm{MeV} and a=(Gv/mV)2=(0.10.3)fm2a=\left(G_{v}/m_{V}\right)^{2}=(0.1-0.3)~{}\mathrm{fm}^{2} and report results for specific values within this range.

IV Sound Speeds in the Pure and Mixed Phases

As the difference of the adiabatic and equilibrium sound speeds drives the restoring force for gg-modes, it is instructive to collect some general expressions for these two sound speeds in the pure and mixed phases. For the pure phase of npenpe matter, these expressions are derived and applied in [20], but given their central role in this work, and the fact that we also extend the application to npeμnpe\mu and quark matter, we detail their derivation below for completeness. For the mixed phase, we derive expressions that have not, to our knowledge, been previously reported in the literature. First, a point of notation: the equilibrium squared sound speed is commonly defined in the literature [20, 64, 6] by the symbol cs2c_{s}^{2}, which we reserve here for the squared adiabatic sound speed, as in [29]. The equilibrium sound speed is defined by

ce2=dpdϵ,\displaystyle c_{e}^{2}=\frac{dp}{d\epsilon}\,, (25)

where pp and ϵ\epsilon are the total pressure and energy density in matter

ϵ=ϵB+l=e,μϵl,p=pB+l=e,μpl.\displaystyle\epsilon=\epsilon_{B}\,+\sum\limits_{l=e^{-},\,\mu^{-}}\epsilon_{l}\,,\quad p=p_{B}\,+\sum\limits_{l=e^{-},\,\mu^{-}}p_{l}\,. (26)

In Eq.(26), the leptonic energies are the TT=0 degenerate Fermi gas expressions for massive leptons. Being a total derivative, the derivative is taken along the curve satisfying both mechanical and chemical equilibrium, i.e., β\beta-equilibrium conditions hold. In NSM, when only npenpe are present in equilibrium, the composition at fixed baryon density (nn+npn_{n}+n_{p}) is completely fixed once the proton fraction xpx_{p} (=xex_{e} by charge neutrality) is determined. In this case, the squared adiabatic sound speed is defined as

cs2=(pϵ)x,\displaystyle c_{s}^{2}=\left(\frac{\partial p}{\partial\epsilon}\right)_{x}\,, (27)

where x=xp=xex=x_{p}=x_{e}. In the partial derivative, the composition is held fixed, i.e., β\beta-equilibrium conditions are imposed only after all derivatives have been evaluated. The resulting distinction between these two speeds plays an important role in determining the oscillation frequencies of non-radial oscillations such as gg-modes:

ω2(1ce21cs2)=(cs2ce2)ce2cs2.\displaystyle\omega^{2}\propto\left(\frac{1}{c_{e}^{2}}-\frac{1}{c_{s}^{2}}\right)=\frac{(c_{s}^{2}-c_{e}^{2})}{c_{e}^{2}c_{s}^{2}}\,. (28)

Note that both the above speeds are dependent on density which varies over a large range in NSM. Furthermore, an individual knowledge of both speeds is required. In what follows, we apply Eqs. (25) and (27) to the case of a pure and mixed phase.

IV.1 The Pure Phase

IV.1.1 Sound speeds in npenpe matter

It is useful to recast the general expressions Eqs. (25) and (27) in terms of derivatives of the individual chemical potentials with respect to density, since such expressions are amenable to both analytical and numerical checks. Without loss of generality, we have

ce2=dpdϵ=(dpdnB)/(dϵdnB),\displaystyle c_{e}^{2}=\frac{dp}{d\epsilon}=\left(\frac{dp}{dn_{B}}\right)\bigg{/}\left(\frac{d\epsilon}{dn_{B}}\right)\,, (29)

Considering npenpe matter as an example, differentiating the total energy energy density inclusive of electrons

ϵ(nB,x)=nB[MB+E(nB,x)]\displaystyle\epsilon(n_{B},x)=n_{B}[M_{B}+E(n_{B},x)] (30)

with respect to nBn_{B}, we have

(dϵdnB)=ϵnB+nB(dEdnB),\displaystyle\left(\frac{d\epsilon}{dn_{B}}\right)=\frac{\epsilon}{n_{B}}+n_{B}\left(\frac{dE}{dn_{B}}\right)\,, (31)

where E(nB,x)E(n_{B},x) is the energy per baryon. The second term on the right hand side of Eq. (31) becomes

(dEdnB)=(EnB)+(Ex)nB(dxdnB).\displaystyle\left(\frac{dE}{dn_{B}}\right)=\left(\frac{\partial E}{\partial n_{B}}\right)+\left(\frac{\partial E}{\partial x}\right)_{n_{B}}\left(\frac{dx}{dn_{B}}\right)\,. (32)

For the equilibrium sound speed, the β\beta-equilibrium condition (Ex)nB=0\left(\frac{\partial E}{\partial x}\right)_{n_{B}}=0 yields

(dEdnB)=(EnB).\displaystyle\left(\frac{dE}{dn_{B}}\right)=\left(\frac{\partial E}{\partial n_{B}}\right)\,. (33)

Thus,

(dϵdnB)=ϵnB+1nB[nB2(dEdnB)]=(ϵ+p)nB.\displaystyle\left(\frac{d\epsilon}{dn_{B}}\right)=\frac{\epsilon}{n_{B}}+\frac{1}{n_{B}}\left[n_{B}^{2}\left(\frac{dE}{dn_{B}}\right)\right]=\frac{(\epsilon+p)}{n_{B}}\,. (34)

From the thermodynamic identity, using charge neutrality (x=xex=x_{e}) and beta-equilibrium,

ϵ+p\displaystyle\epsilon+p =\displaystyle= μnnn+μpnp+μene\displaystyle\mu_{n}n_{n}+\mu_{p}n_{p}+\mu_{e}n_{e} (35)
=\displaystyle= μnnB(μnμpμe)nB=μnnB,\displaystyle\mu_{n}n_{B}-(\mu_{n}-\mu_{p}-\mu_{e})n_{B}=\mu_{n}n_{B}\,,

leading to the simple result

(dϵdnB)=(ϵnB)x=μn\displaystyle\left(\frac{d\epsilon}{dn_{B}}\right)=\left(\frac{\partial\epsilon}{\partial n_{B}}\right)_{x}=\mu_{n} (36)

This implies that

ce2\displaystyle c_{e}^{2} \displaystyle\equiv dpdϵ=1μn(dpdnB)=1μnnB(dμndnB)\displaystyle\frac{dp}{d\epsilon}=\frac{1}{\mu_{n}}\left(\frac{dp}{dn_{B}}\right)=\frac{1}{\mu_{n}}n_{B}\left(\frac{d\mu_{n}}{dn_{B}}\right)\, (37)
=\displaystyle= (dlnμndlnnB),\displaystyle\left(\frac{d\ln\mu_{n}}{d\ln n_{B}}\right)\,,

where we have again taken advantage of the thermodynamic identity to relate the required derivative of pp to that of μn\mu_{n}.

The adiabatic squared sound speed can be expressed as

cs2\displaystyle c_{s}^{2} =\displaystyle= (pϵ)x=(pnB)x/(ϵnB)x\displaystyle\left(\frac{\partial p}{\partial\epsilon}\right)_{x}=\left(\frac{\partial p}{\partial n_{B}}\right)_{x}\bigg{/}\left(\frac{\partial\epsilon}{\partial n_{B}}\right)_{x}\, (38)
=\displaystyle= nBϵ+p(pnB)x=1μavg(pnB)x,\displaystyle\frac{n_{B}}{\epsilon+p}\left(\frac{\partial p}{\partial n_{B}}\right)_{x}=\frac{1}{\mu_{\rm avg}}\left(\frac{\partial p}{\partial n_{B}}\right)_{x}\,, (39)

where we have used the equality in Eq. (34), which is valid at constant composition even in the absence of β\beta-equilibrium, and introduced an average chemical potential μavg=(iμini)/nB=(ϵ+p)/nB\mu_{\rm avg}=(\sum_{i}\mu_{i}n_{i})/n_{B}=(\epsilon+p)/n_{B}. Since p=pB+pep=p_{B}+p_{e},

cs2=1(ϵ+p)[(upBu)x+(upeu)x].c_{s}^{2}=\frac{1}{(\epsilon+p)}\left[\left(u\frac{\partial p_{B}}{\partial u}\right)_{x}+\left(u\frac{\partial p_{e}}{\partial u}\right)_{x}\right]\,. (40)

The required derivatives are analytic:

(upBu)x\displaystyle\left(u\frac{\partial p_{B}}{\partial u}\right)_{x} =\displaystyle= 29π2kF5EF+nsu[2a0u+b0γ(γ+1)uγ]\displaystyle\frac{2}{9\pi^{2}}\frac{k_{F}^{5}}{E_{F}}+n_{s}u\Big{[}2a_{0}u+b_{0}\gamma(\gamma+1)u^{\gamma}\Big{]} (41)
+\displaystyle+ nsu(12x)2{2kF227EF(19kF210EF2+3kF410EF4)\displaystyle n_{s}u(1-2x)^{2}\left\{\frac{2k_{F}^{2}}{27E_{F}}\left(1-\frac{9k_{F}^{2}}{10E_{F}^{2}}+\frac{3k_{F}^{4}}{10E_{F}^{4}}\right)\right.
+\displaystyle+ [2u(a1a0)+b1γ1(γ1+1)uγ1\displaystyle\left.\bigg{[}2u(a_{1}-a_{0})+b_{1}\gamma_{1}(\gamma_{1}+1)u^{\gamma_{1}}\right.
\displaystyle- b0γ(γ+1)uγ]}\displaystyle\left.b_{0}\gamma(\gamma+1)u^{\gamma}\bigg{]}\right\}
(upeu)x\displaystyle\left(u\frac{\partial p_{e}}{\partial u}\right)_{x} =\displaystyle= 13neμe.\displaystyle\frac{1}{3}n_{e}\mu_{e}\,. (42)

Thus, the difference of squared sound speeds becomes

cs2ce2=1μavg(pnB)x1μn(dpdnB).\displaystyle c_{s}^{2}-c_{e}^{2}=\frac{1}{\mu_{\rm avg}}\left(\frac{\partial p}{\partial n_{B}}\right)_{x}-\frac{1}{\mu_{n}}\left(\frac{dp}{dn_{B}}\right)\,. (43)

At this point all the necessary ingredients for the calculation of the speed-of-sound difference are present. It is instructive, however, to obtain a complementary expression in which its physical causes, namely β\beta-equilibrium and compositional gradients, are made explicit. To that end, we proceed as follows: Noting that

dpdnB=(pnB)x+(px)nBdxdnB,\displaystyle\frac{dp}{dn_{B}}=\left(\frac{\partial p}{\partial n_{B}}\right)_{x}+\left(\frac{\partial p}{\partial x}\right)_{n_{B}}\frac{dx}{dn_{B}}\,, (44)

Eq. (43) can be recast as

cs2ce2\displaystyle c_{s}^{2}-c_{e}^{2} =\displaystyle= (1μavg1μn)(pnB)x1μn(px)nBdxdnB\displaystyle\left(\frac{1}{\mu_{\rm avg}}-\frac{1}{\mu_{n}}\right)\left(\frac{\partial p}{\partial n_{B}}\right)_{x}-\frac{1}{\mu_{n}}\left(\frac{\partial p}{\partial x}\right)_{n_{B}}\frac{dx}{dn_{B}} (45)
=\displaystyle= xμ~μavgμn(pnB)x1μn(px)nBdxdnB\displaystyle-~{}\frac{x\tilde{\mu}}{\mu_{\rm avg}\mu_{n}}\left(\frac{\partial p}{\partial n_{B}}\right)_{x}-\frac{1}{\mu_{n}}\left(\frac{\partial p}{\partial x}\right)_{n_{B}}\frac{dx}{dn_{B}}\,

where μ~=μe+μpμn\tilde{\mu}=\mu_{e}+\mu_{p}-\mu_{n}. Anticipating that β\beta-equilibrium will be imposed at the end, we note that the first term above vanishes as μ~=0\tilde{\mu}=0, which leads to

cs2ce2=1μn(px)nBdxdnB.\displaystyle c_{s}^{2}-c_{e}^{2}=-\frac{1}{\mu_{n}}\left(\frac{\partial p}{\partial x}\right)_{n_{B}}\frac{dx}{dn_{B}}\,. (46)

Utilizing p=nB2EnBp=n_{B}^{2}\frac{\partial E}{\partial n_{B}} and interchanging the order of derivatives

cs2ce2=1μnnB2nB(Ex)nBdxdnB,\displaystyle c_{s}^{2}-c_{e}^{2}=-\frac{1}{\mu_{n}}n_{B}^{2}\frac{\partial}{\partial n_{B}}\left(\frac{\partial E}{\partial x}\right)_{n_{B}}\frac{dx}{dn_{B}}\,, (47)

which can be further rewritten as 666Observe the interesting relation px=nB2μ~nB\frac{\partial p}{\partial x}=n_{B}^{2}\frac{\partial\tilde{\mu}}{\partial n_{B}}, noted also in the context of bulk viscosity studies [65].

cs2ce2=1μnnB2(μ~nB)xdxdnB.\displaystyle c_{s}^{2}-c_{e}^{2}=-\frac{1}{\mu_{n}}n_{B}^{2}\left(\frac{\partial\tilde{\mu}}{\partial n_{B}}\right)_{x}\frac{dx}{dn_{B}}\,. (48)

It remains now to determine dxdnB\frac{dx}{dn_{B}}. As

dμ~=(μ~nB)xdnB+(μ~x)nBdx=0,\displaystyle d\tilde{\mu}=\left(\frac{\partial\tilde{\mu}}{\partial n_{B}}\right)_{x}dn_{B}+\left(\frac{\partial\tilde{\mu}}{\partial x}\right)_{n_{B}}dx=0\,, (49)
dxdnB=(μ~nB)x/(μ~x)nB.\displaystyle\frac{dx}{dn_{B}}=-\left(\frac{\partial\tilde{\mu}}{\partial n_{B}}\right)_{x}\bigg{/}\left(\frac{\partial\tilde{\mu}}{\partial x}\right)_{n_{B}}\,. (50)

With this relation, Eq. (48) becomes

cs2\displaystyle c_{s}^{2} =\displaystyle= ce2+[nB(μ~nB)x]2μn(μ~x)nB,\displaystyle c_{e}^{2}+\frac{\left[n_{B}\left(\frac{\partial\tilde{\mu}}{\partial n_{B}}\right)_{x}\right]^{2}}{\mu_{n}\left(\frac{\partial\tilde{\mu}}{\partial x}\right)_{n_{B}}}\,, (51)

which illustrates the influence of the density and compositional gradients of the two sound speeds. Thus far, we have simply retraced the steps originally given in [20] (Sec. 4.2 of this reference). In npenpe matter under the constraint of charge neutrality, the independent variables chosen are nBn_{B} and xx, and thus a partial derivative of μ~\tilde{\mu} with respect to nBn_{B} (xx) implies that xx (nBn_{B}) is fixed.

Casting the expressions for the sound speeds in terms of the chemical potentials is expedient, as illustrated below for the case of npenpe matter. Note that the average chemical potential μavg=μn\mu_{\rm avg}=\mu_{n} only in β\beta-equilibrium. At fixed xx, with the relation μ~=μeμ^\tilde{\mu}=\mu_{e}-\hat{\mu},

nBμ~nB\displaystyle n_{B}\frac{\partial\tilde{\mu}}{\partial n_{B}} =\displaystyle= u(μeμ^)u.\displaystyle u\frac{\partial(\mu_{e}-\hat{\mu})}{\partial u}\,. (52)

For the term in Eq. (52) involving electrons, we have

μe=c(3π2nsu)1/3x1/3anduμeu=μe3,\displaystyle\mu_{e}=\hbar c~{}(3\pi^{2}n_{s}u)^{1/3}x^{1/3}\quad{\rm and}\quad u\frac{\partial\mu_{e}}{\partial u}=\frac{\mu_{e}}{3}\,, (53)

while for baryons,777Details of the derivatives of the kinetic part of the symmetry energy are given in Appendix A.

μ^\displaystyle\hat{\mu} =\displaystyle= μnμp=4S2(u)(12x)with\displaystyle\mu_{n}-\mu_{p}=4S_{2}(u)(1-2x)\quad{\rm with}
S2(u)\displaystyle S_{2}(u) =\displaystyle= S2k+S2i=kF26EF\displaystyle S_{2k}+S_{2i}=\frac{k_{F}^{2}}{6E_{F}}
+\displaystyle+ (a1a0)u+b1uγ1b0uγ,\displaystyle(a_{1}-a_{0})u+b_{1}u^{\gamma_{1}}-b_{0}u^{\gamma}\,,
uS2k\displaystyle uS_{2k}^{\prime} =\displaystyle= 132S2k[118(S2kkF)2],\displaystyle\frac{1}{3}\cdot 2S_{2k}\left[1-18\left(\frac{S_{2k}}{k_{F}}\right)^{2}\right],
anduS2i\displaystyle{\rm and}\quad uS_{2i}^{\prime} =\displaystyle= (a1a0)u+γ1b1uγ1γb0uγ,\displaystyle(a_{1}-a_{0})u+\gamma_{1}b_{1}u^{\gamma_{1}}-\gamma b_{0}u^{\gamma}\,,
uS2\displaystyle uS_{2}^{\prime} =\displaystyle= uS2k+uS2i.\displaystyle uS_{2k}^{\prime}+uS_{2i}^{\prime}. (54)

Putting together the above results, we have

nBμ~nB=μe34(12x)uS2.\displaystyle n_{B}\frac{\partial\tilde{\mu}}{\partial n_{B}}=\frac{\mu_{e}}{3}-4(1-2x)~{}uS_{2}^{\prime}\,. (55)

Derivatives with respect to xx of μ~\tilde{\mu} at fixed density are also straightforward. For

μ~x=(μeμ^)x,\displaystyle\frac{\partial\tilde{\mu}}{\partial x}=\frac{\partial(\mu_{e}-\hat{\mu})}{\partial x}\,, (56)

we note that

μex\displaystyle\frac{\partial\mu_{e}}{\partial x} =\displaystyle= 13μexandμ^x=8S2(u)sothat\displaystyle\frac{1}{3}\frac{\mu_{e}}{x}\quad{\rm and}\quad\frac{\partial\hat{\mu}}{\partial x}=-8S_{2}(u)\quad{\rm so~{}that}
μ~x\displaystyle\frac{\partial\tilde{\mu}}{\partial x} =\displaystyle= 13μex+8S2(u).\displaystyle\frac{1}{3}\frac{\mu_{e}}{x}+8S_{2}(u)\,. (57)

The equivalence of Eqs. (40) and (51) is established analytically in Appendix A.5.

IV.1.2 Sound speeds in npeμnpe\mu matter

Going beyond the results in [20], one way to include muons is by choosing nBn_{B}, x=xe+xμx=x_{e}+x_{\mu} and xμyx_{\mu}\equiv y as the independent variables. The formal expression for the squared adiabatic sound speed remains the same as in npenpe matter, i.e., Eq. (40) but now (upe/u)x\left(u~{}\partial p_{e}/\partial u\right)_{x} [Eq. (42)] is replaced by

(uplepu)x,xμ=13neμe+13nμ(μμ2mμ2μμ).\left(u\frac{\partial p_{lep}}{\partial u}\right)_{x,x_{\mu}}=\frac{1}{3}n_{e}\mu_{e}+\frac{1}{3}n_{\mu}\left(\frac{\mu_{\mu}^{2}-m_{\mu}^{2}}{\mu_{\mu}}\right)\,. (58)

where lep=e,μlep=e^{-},\mu^{-}.

Furthermore, by retracing the steps leading to Eq. (51), its npeμnpe\mu equivalent is obtained as

cs2ce2=1μn(Px|nB,ydxdnB+Py|nB,xdydnB)c_{s}^{2}-c_{e}^{2}=-\frac{1}{\mu_{n}}\left(\left.\frac{\partial P}{\partial x}\right|_{n_{B},y}\frac{dx}{dn_{B}}+\left.\frac{\partial P}{\partial y}\right|_{n_{B},x}\frac{dy}{dn_{B}}\right) (59)

with 888The intermediate steps leading to Eqs. (60)-(61) are detailed in Appendix. A.6

dxdnB\displaystyle\frac{dx}{dn_{B}} =\displaystyle= μ~xy|nB,xμ~ynB|x,yμ~yy|nB,xμ~xnB|x,yμ~xy|nB,xμ~yx|nB,yμ~yy|nB,xμ~xx|nB,y\displaystyle\frac{\left.\frac{\partial\tilde{\mu}_{x}}{\partial y}\right|_{n_{B},x}\left.\frac{\partial\tilde{\mu}_{y}}{\partial n_{B}}\right|_{x,y}-\left.\frac{\partial\tilde{\mu}_{y}}{\partial y}\right|_{n_{B},x}\left.\frac{\partial\tilde{\mu}_{x}}{\partial n_{B}}\right|_{x,y}}{\left.\frac{\partial\tilde{\mu}_{x}}{\partial y}\right|_{n_{B},x}\left.\frac{\partial\tilde{\mu}_{y}}{\partial x}\right|_{n_{B},y}-\left.\frac{\partial\tilde{\mu}_{y}}{\partial y}\right|_{n_{B},x}\left.\frac{\partial\tilde{\mu}_{x}}{\partial x}\right|_{n_{B},y}} (60)
dydnB\displaystyle\frac{dy}{dn_{B}} =\displaystyle= μ~xx|nB,yμ~ynB|x,yμ~yx|nB,yμ~xnB|x,yμ~xx|nB,yμ~yy|nB,xμ~yx|nB,yμ~xy|nB,x.\displaystyle\frac{\left.\frac{\partial\tilde{\mu}_{x}}{\partial x}\right|_{n_{B},y}\left.\frac{\partial\tilde{\mu}_{y}}{\partial n_{B}}\right|_{x,y}-\left.\frac{\partial\tilde{\mu}_{y}}{\partial x}\right|_{n_{B},y}\left.\frac{\partial\tilde{\mu}_{x}}{\partial n_{B}}\right|_{x,y}}{\left.\frac{\partial\tilde{\mu}_{x}}{\partial x}\right|_{n_{B},y}\left.\frac{\partial\tilde{\mu}_{y}}{\partial y}\right|_{n_{B},x}-\left.\frac{\partial\tilde{\mu}_{y}}{\partial x}\right|_{n_{B},y}\left.\frac{\partial\tilde{\mu}_{x}}{\partial y}\right|_{n_{B},x}}~{}. (61)

The chemical potentials μ~x=μp+μeμn\tilde{\mu}_{x}=\mu_{p}+\mu_{e}-\mu_{n} and μ~y=μμμe\tilde{\mu}_{y}=\mu_{\mu}-\mu_{e} are zero in β\beta-equilibrated matter. Equations (59)-(61), while demonstrating that compositional gradients are at the core of g-mode oscillations, are lengthy and computationally more involved compared to the direct calculation of the adiabatic sound speed in npeμnpe\mu matter using Eqs. (41) and (58). For the sake of completeness, we provide here the explicit expressions for the adiabatic sound speed in npeμnpe\mu matter arising from  (59)-(61) which are in excellent numerical agreement with the more direct method:

cs2=ce2+1μn(T1+T2+T3+T4)\displaystyle c_{s}^{2}=c_{e}^{2}+\frac{1}{\mu_{n}}\Big{(}T_{1}+T_{2}+T_{3}+T_{4}\Big{)} (62)

where Tj=Nj/DT_{j}=N_{j}/D with

N1\displaystyle N_{1} =[μe34(12x)uS2]2\displaystyle=\left[\frac{\mu_{e}}{3}-4(1-2x)~{}uS_{2}^{\prime}\right]^{2} (63)
N2\displaystyle N_{2} =[μe34(12x)uS2]8S2xe[kFμkFexμxe]\displaystyle=\left[\frac{\mu_{e}}{3}-4(1-2x)~{}uS_{2}^{\prime}\right]8S_{2}x_{e}\left[\frac{k_{F_{\mu}}}{k_{F_{e}}}-\frac{x_{\mu}}{x_{e}}\right]
N3\displaystyle N_{3} =[kFμ23μe4(12x)uS2](μe+8S2xe)[xμxekFμkFe]\displaystyle=\left[\frac{k_{F_{\mu}}^{2}}{3\mu_{e}}-4(1-2x)~{}uS_{2}^{\prime}\right]\left(\mu_{e}+8S_{2}x_{e}\right)\left[\frac{x_{\mu}}{x_{e}}-\frac{k_{F_{\mu}}}{k_{F_{e}}}\right]
N4\displaystyle N_{4} =[kFμ23μe4(12x)uS2]kFμkFe[μe34(12x)uS2]\displaystyle=\left[\frac{k_{F_{\mu}}^{2}}{3\mu_{e}}-4(1-2x)~{}uS_{2}^{\prime}\right]\frac{k_{F_{\mu}}}{k_{F_{e}}}\left[\frac{\mu_{e}}{3}-4(1-2x)~{}uS_{2}^{\prime}\right]
D\displaystyle D =[μe3xe+8S2(1+kFμkFe)]\displaystyle=\left[\frac{\mu_{e}}{3x_{e}}+8S_{2}\left(1+\frac{k_{F_{\mu}}}{k_{Fe}}\right)\right]

and kFe=μek_{F_{e}}=\mu_{e} (massless electrons) and kFμk_{F_{\mu}}=μμ2mμ2\sqrt{{\mu_{\mu}}^{2}-m_{\mu}^{2}}. These equations explicitly display the connection to the nuclear symmetry energy S2S_{2} and its density derivative S2S_{2}^{\prime}.

At the muon threshold (kFμ,xμk_{F_{\mu}},x_{\mu}=0 \Rightarrow xx=xex_{e}), it is easy to see that N2,N3,N4N_{2},N_{3},N_{4}=0, while N1(N1npeN_{1}(\equiv N_{1}^{npe}) recovers Eq. (51) for npenpe matter. At extremely high baryon density, muons are ultra-relativistic (μμ\mu_{\mu}=kFμk_{F_{\mu}}=kFek_{F_{e}}, xμx_{\mu}=xex_{e}=x/2x/2) so that N2,N3N_{2},N_{3}=0, N1N_{1}=N4N_{4}=N1npe/2N_{1}^{npe}/2 and the total leptonic contribution to the sound speed is equally divided between electrons and muons.

IV.1.3 Sound speeds in Quark Matter

We now move to a discussion of sound speeds in dense quark matter at zero temperature. For the pure quark phase, the difference of the two sound speeds has been computed to leading order in the quark mass [66, 36] using the non-interacting 3-flavor FG model with massive quarks (see Sec.III.2). These expressions reveal that for the non-interacting FG model, a non-zero quark mass is necessary to support gg-modes. This is because a system of massless udsuds quarks is charge neutral with equal numbers of each flavor at any density; effectively, there is no change in composition with density to drive the gg-mode.

To leading order in the ss-quark’s mass msm_{s}, the Brunt-Väisälä frequency is [36]

Nq(g2πce)(ms2B),\displaystyle N_{q}\simeq\left(\frac{g}{2\pi c_{e}}\right)\left(\frac{m_{s}^{2}}{\sqrt{B}}\right)\,, (64)

where ce2=dpq/dϵqc_{e}^{2}=dp_{q}/d\epsilon_{q} is the equilibrium squared sound speed in QM999Numerically, Nq100N_{q}\approx 100 Hz for a current quark mass ms100m_{s}\approx 100 MeV, but the effect of interactions in addition to this yields significantly lower values for NqN_{q} [36].. It is possible to obtain an exact expression for ce2c_{e}^{2} and cs2c_{s}^{2} in QM for the FG model, and also for the vMIT model, as we show below.

The equilibrium sound speed may be simply calculated by numerically evaluating ce,vMIT2=dp/dϵc_{e,vMIT}^{2}=dp/d\epsilon in the pure quark phase. However, additional insight into its compositional structure is gained by expressing it in terms of the various chemical potentials involved. Starting from the relation (valid in β\beta-equilibrium)

μn=2μd+μu=(2μd+μu)+3anQ,\displaystyle\mu_{n}=2\mu_{d}+\mu_{u}=(2\mu_{d}^{\ast}+\mu_{u}^{\ast})+3an_{Q}\,, (65)

where μf=kFf2+mf2\mu_{f}^{\ast}=\sqrt{k_{F_{f}}^{2}+m_{f}^{2}} for a quark of flavor ff, and using ce2=dlnμn/lnnBc_{e}^{2}=d\ln\mu_{n}/\ln n_{B},101010In charge neutral and β\beta-equilibrated matter, μB=fxfμf+=e,μxμ=μn\mu_{B}=\sum_{f}x_{f}\mu_{f}+\sum\limits_{\ell=e,\mu}x_{\ell}\mu_{\ell}=\mu_{n} as in the nucleonic phase. we obtain

ce,vMIT2=1μn\displaystyle c_{e,vMIT}^{2}=\frac{1}{\mu_{n}} [\displaystyle\bigg{[} 13{2μd(1md2μd2)dlnnddlnnB\displaystyle\frac{1}{3}\left\{2\mu_{d}^{\ast}\left(1-\frac{m_{d}^{2}}{\mu_{d}^{\ast^{2}}}\right)\frac{d\ln n_{d}}{d\ln n_{B}}\right. (66)
+\displaystyle+ μu(1mu2μu2)dlnnudlnnB}\displaystyle\left.\mu_{u}^{\ast}\left(1-\frac{m_{u}^{2}}{\mu_{u}^{\ast^{2}}}\right)\frac{d\ln n_{u}}{d\ln n_{B}}\right\}
+\displaystyle+ 3anQ].\displaystyle 3an_{Q}\bigg{]}\,.

Contributions from the leptons are implicitly included in the above expression.

For the non-interacting FG model, the pressure p=fpFG(μf,μe)p=\sum_{f}p_{\rm FG}(\mu_{f},\mu_{e}). Introducing the partial fractions xf=nf/nBx_{f}=n_{f}/n_{B}, where nf=(μf2mf2)3/2/π2n_{f}=(\mu_{f}^{2}-m_{f}^{2})^{3/2}/\pi^{2} and ne=μe3/3π2n_{e}=\mu_{e}^{3}/3\pi^{2}, the partial derivative of pressure with respect to baryon density in the definition of the adiabatic sound speed in Eq. (39) can be re-expressed in terms of partial derivatives with respect to the various chemical potentials, yielding

cs,FG2=1μavg[f13μfxf(1mf2μf2)+13μexe],\displaystyle c_{s,{\rm FG}}^{2}=\frac{1}{\mu_{\rm avg}}\left[\sum_{f}\frac{1}{3}\mu_{f}x_{f}\left(1-\frac{m_{f}^{2}}{\mu_{f}^{2}}\right)+\frac{1}{3}\mu_{e}x_{e}\right]\,, (67)

where μavg=(f=u,d,s,enfμf)/nB\mu_{\rm avg}=(\sum\limits_{f=u,d,s,e}n_{f}\mu_{f})/n_{B}. Note that if all mfm_{f} = 0 (i.e, one is in a charge neutral phase with xex_{e} = 0), cs,FG2=ce,FG2=1/3c_{s,{\rm FG}}^{2}=c_{e,{\rm FG}}^{2}=1/3 and there can be no gg-modes. Inclusion of 𝒪(αs){\cal O}(\alpha_{s}) corrections to this model does not change the fact that a non-zero quark mass is necessary for gg-modes.

In the vMIT model given by Eqs. (24), μf=kFf2+mf2=μfanQ\mu_{f}^{\ast}=\sqrt{k_{F_{f}}^{2}+m_{f}^{2}}=\mu_{f}-an_{Q}, and as was done for the FG model, we compute partial derivatives with respect to μf\mu_{f}^{\ast}, noting that nf(μf)n_{f}(\mu_{f}) = nFG(μf)n_{\rm FG}(\mu_{f}^{\ast}). The resulting expression for the adiabatic sound speed in the vMIT model is

cs,vMIT2=1μavg\displaystyle c_{s,vMIT}^{2}=\frac{1}{\mu_{\rm avg}} [\displaystyle\bigg{[} f13μfxf(1mf2μf2)\displaystyle\sum_{f}\frac{1}{3}\mu_{f}^{\ast}x_{f}\left(1-\frac{m_{f}^{2}}{\mu_{f}^{\ast^{2}}}\right) (68)
+\displaystyle+ 13μexe+3anQ],\displaystyle\frac{1}{3}\mu_{e}x_{e}+3an_{Q}\bigg{]}\,,

where all quantities μf,μe,xe,xf\mu_{f},\mu_{e},x_{e},x_{f} are equilibrium values111111Inclusion of muons is straightforward and adds a term, 13μμxμ(1mμ2μμ2)\frac{1}{3}\mu_{\mu}x_{\mu}\left(1-\frac{m_{\mu}^{2}}{\mu_{\mu}^{\ast^{2}}}\right), on the right hand side of Eq. (67).. If we switch off interactions (aa\rightarrow 0), we recover results of the non-interacting FG model. Interestingly, if we retain the interaction term, but set all quark masses equal or to zero (implying that xex_{e} = 0), we find that cs,vMIT2=ce,vMIT2c_{s,vMIT}^{2}=c_{e,vMIT}^{2} so stable gg-modes are not supported in the pure quark phase. Therefore, while both sound speeds are modified by interactions, e.g.,

cs,vMIT2=1μavg[μq+3anQ]cs,FG2,\displaystyle c_{s,vMIT}^{2}=\frac{1}{\mu_{\rm avg}}\left[\mu_{q}^{*}+3an_{Q}\right]\neq c_{s,{\rm FG}}^{2}\,, (69)

at asymptotically high density where quark masses are negligible, there can be no gg-modes in quark matter in the vMIT model 121212gg-modes would still exist in a mixed phase of vMIT quark matter and nucleons as the electron fraction would vary from β\beta-processes involving nucleons.

Note that when all chemical potentials and partial fractions are set to their equilibrium values for cs,vMIT2c_{s,vMIT}^{2} in the pure phase, μavg=μn\mu_{\rm avg}=\mu_{n}. A comparison of Eqs. (66) and (68) reveals the differences between the two sound speeds. While effects of interactions enter in the same formal way for the two squared speeds, the occurrence of the logarithmic derivatives of the quark densities distinguishes ce,vMIT2c_{e,vMIT}^{2} from cs,vMIT2c_{s,vMIT}^{2} which features the partial fractions xfx_{f}. This difference is the principal reason for the latter to become larger than the former. In both cases, the dd-quark contributions are larger than those of uu and ss quarks.

IV.2 Sound Speeds in the Mixed Phase

Once we have expressions for the sound speed in a pure phase of quarks or nucleons, it is possible to compute the sound speed in the mixed phase of the two, obtained from a Gibbs construction. The only information required, other than the sound speeds in the pure phases, is the partial phase fraction of quarks χ\chi at any density. It is more convenient to begin with the reciprocal relation

1ce,mix2=(dϵmixdpmix).\displaystyle\frac{1}{c_{e,{\rm mix}}^{2}}=\left(\frac{d\epsilon_{\rm mix}}{dp_{\rm mix}}\right)\,. (70)

In a Gibbs mixed phase, the pressures in the two phases are equal, while the energy density is a proportional mix of the quark (qq) and nucleonic/hadronic (hh) phases: ϵmix=(1χ)ϵh+χϵq\epsilon_{\rm mix}=(1-\chi)\epsilon_{h}+\chi\epsilon_{q}. Substituting this in Eq.(70) gives

1ce,mix2=(1χ)ce,h2+χce,q2+(ϵqϵh)dχ/dnBdP/dnB.\displaystyle\frac{1}{c_{e,{\rm mix}}^{2}}=\frac{(1-\chi)}{c_{e,h}^{2}}+\frac{\chi}{c_{e,q}^{2}}+(\epsilon_{q}-\epsilon_{h})\frac{d\chi/dn_{B}}{dP/dn_{B}}\,. (71)

The derivatives in Eq.(71) must be computed numerically after solving for χ\chi, hence afford no particular advantage over a direct numerical computation of the sound speed from Eq.(70) itself. However, note that the last term in Eq.(71) is always positive in the mixed phase.

As before, the general definition of the adiabatic sound speed applies to the mixed phase

cs,mix2=(dpmixdϵmix)xi=const.=xi,eq.,\displaystyle c_{s,{\rm mix}}^{2}=\left(\frac{dp_{\rm mix}}{d\epsilon_{\rm mix}}\right)_{x_{i}={\rm const.}=x_{i,{\rm eq.}}}\,, (72)

and the thermodynamic identity becomes ϵmix+pmix\epsilon_{\rm mix}+p_{\rm mix} = iniμi\sum_{i}n_{i}\mu_{i} = nBμavgn_{B}\mu_{\rm avg}. Noting that the derivatives ϵh/nB,h\partial\epsilon_{h}/\partial_{n_{B,h}} and ϵq/nB,q\partial\epsilon_{q}/\partial_{n_{B,q}} are equal to the respective μavg\mu_{\rm avg}, it is once again more convenient to begin with the reciprocal relation

1cs,mix2=(dϵmixdpmix)xi=const.=xi,eq.,\displaystyle\frac{1}{c_{s,{\rm mix}}^{2}}=\left(\frac{d\epsilon_{\rm mix}}{dp_{\rm mix}}\right)_{x_{i}={\rm const.}=x_{i,{\rm eq.}}}\,, (73)

and use the chain rule to compute derivatives with respect to density. This leads to

(dϵmixdpmix)xi=xi,eq.=(1χ)μavg(phnB,h)+χμavg(pqnB,q),\displaystyle\left(\frac{d\epsilon_{\rm mix}}{dp_{\rm mix}}\right)_{x_{i}=x_{i,{\rm eq.}}}=\frac{(1-\chi)\mu_{\rm avg}}{\left(\frac{\partial p_{h}}{\partial n_{B},h}\right)}+\frac{\chi\mu_{\rm avg}}{\left(\frac{\partial p_{q}}{\partial n_{B},q}\right)}\,, (74)

which, using Eq. (39), becomes

1cs,mix2=(1χ)cs,h2+χcs,q2.\displaystyle\frac{1}{c_{s,{\rm mix}}^{2}}=\frac{(1-\chi)}{c_{s,h}^{2}}+\frac{\chi}{c_{s,q}^{2}}\,. (75)

Comparing Eqs. (71) and (75), and to the extent that the two sound speeds in the pure hadronic/quark phase are almost equal, we expect that the last term in Eq. (71) which tracks the rapidly changing composition in the mixed phase, is mainly responsible for cs,mix2>ce,mix2c_{s,\rm mix}^{2}>c_{e,\rm mix}^{2}. The more rapid the appearance of new chemical species and the softer the mixed phase, the larger the Brunt-Väisälä  frequency will be.

Furthermore, as will become evident from our results in Sec. V, the adiabatic sound speed is continuous across the transition to and from the mixed phase, while the equilibrium sound speed has a slight jump to accommodate the derivative of χ\chi. The reciprocal relation for the adiabatic sound speeds is reminiscent of the addition of resistors in a parallel circuit, with voltage as pressure and current as energy density. Such impedance analogies arise commonly in electrical engineering when modeling the behavior of transducers.

V Results

V.1 Structural properties, sound speeds and the Brunt-Väisälä frequency

Refer to caption
Refer to caption
Figure 1: Mass-radius curves for the ZL EOS without and with muons. Configurations with muons are slightly more compact, but both cases support Mmax2MM_{\rm max}\simeq 2M_{\odot}. Except for LL, the EOS parameters are K0=220K_{0}=220 MeV, Sv=31S_{v}=31 MeV, and γ1=1.6\gamma_{1}=1.6 for all curves.

Figure 1 shows MM-RR curves for ZL EOSs with and without muons for the indicated parameters in the caption. The radii of 1.4M\sim 1.4~{}M_{\odot} stars, R1.4R_{1.4}, for the different models shown lie within the bounds inferred from available data. For example, data from X-ray observations have yielded R1.4=9R_{1.4}=9-1414 km for canonical masses of 1.4M\sim 1.4~{}M_{\odot} [67, 68, 69]. Measured tidal deformations from gravitational wave data in the binary NS merger GW170817 give 8.9-13.2 km for binary masses in the range 1.36(1.17)-1.6(1.36) MM_{\odot} [70], whereas for the same masses Capano et al. [71] report 11±111\pm 1 km. X-ray pulse analysis of NICER data from PSR J0030+0451 by Miller et al. (2019) [39] finds 13.021.19+1.1413.02^{+1.14}_{-1.19} km for M=1.44±0.15MM=1.44\pm 0.15~{}M_{\odot}, whereas for the same star Riley et al. (2019) [38] obtain 12.711.19+1.1412.71^{+1.14}_{-1.19} km and M=1.340.16+0.15MM=1.34^{+0.15}_{-0.16}~{}M_{\odot}. The maximum masses (2M\simeq 2M_{\odot}) of these EOSs131313By adjusting the constants of the ZL EOS to make the EOS stiffer (yet causal) at supra-nuclear densities, masses larger than 2M2M_{\odot} can be obtained; an example will be shown later. are also within the uncertainties of high mass NSs which range from 1.908±0.016M1.908\pm 0.016~{}M_{\odot} to 2.270.15+0.17M2.27^{+0.17}_{-0.15}~{}M_{\odot} [72, 73, 74, 75, 76]. Although differences in R1.4R_{1.4} with and without muons for a given EOS are small, the appearance of muons in the star leads to distinct features in the Brunt-Väisälä frequency (see below).

Refer to caption
Figure 2: Difference between the adiabatic and equilibrium squared sound speeds (normalized to the squared speed of light) for the ZL EOS (K0K_{0}=220 MeV, Sv=31S_{v}=31 MeV, L=60L=60 MeV and γ1\gamma_{1}=1.6) without and with muons.

In Fig. 2, differences in the two squared sound speeds are shown as a function of nBn_{B} with and without muons for the ZL EOS with L=60L=60 MeV. The small jump at nB0.14fm3n_{B}\simeq 0.14~{}{\rm fm}^{-3}, the density at which muons appear, is caused by a sudden drop in the equilibrium sound speed. The differences at large densities are due to the increasing concentration of muons.

Refer to caption
Refer to caption
Figure 3: The Brunt-Väisälä  frequency in the NS for the ZL EOS (K0K_{0}=220 MeV, Sv=31S_{v}=31 MeV, L=60L=60 MeV, and γ1\gamma_{1}=1.6) without and with muons.
Refer to caption
Figure 4: EOS for the mixed phase of nucleons and quarks (middle curve) using the Gibbs construction. For the ZL EOS without muons, K0K_{0}=220 MeV, Sv=31S_{v}=31 MeV, L=60L=60 MeV, and γ1\gamma_{1}=1.6. Parameters for the vMIT EOS are: (mu,md,msm_{u},m_{d},m_{s})=(5,7,150) MeV, B1/4B^{1/4}=180 MeV and aa=0.1. The circle indicates the central pp and ϵ\epsilon of the maximum mass star (nc,max/nsn_{c,{\rm max}}/n_{s}=7.63, for Mmax/MM_{\rm max}/M_{\odot}=1.82).

Figure 3 shows the Brunt-Väisälä  frequency NN vs r/Rr/R in the star. In the results shown, the crust is assumed to be a homogeneous fluid for simplicity, hence NN vanishes there. The location where muons begin to appear is signaled by the small kink in the bottom panel. Overall, NN is slightly larger with muons in the density range in the core of a 1.4MM_{\odot} star, consistent with Fig. 2. This has a proportional impact on the gg-mode frequency as shown in the next section.

The EOS of the mixed phase following the Gibbs construction, and the ZL EOS for the nucleonic sector and the vMIT EOS for the quark sector is shown in Fig. 4. The ZL EOS does not include the small effect of muons. In the quark sector, muons have not been included since their impact relative to quarks is tiny.

Refer to caption
Figure 5: Quark fraction vs nBn_{B} corresponding to Fig. 4. The circle indicates the central density of the maximum mass star (nc,maxn_{c,{\rm max}}=1.22 fm3{\rm fm}^{-3} for Mmax/MM_{\rm max}/M_{\odot}=1.82).

The compositional change in the mixed phase is indicated by the quark fraction χ\chi in Fig. 5. The steep rise of χ\chi from the onset indicates the sort of rapid compositional change that can impact the gg-mode frequency. A similar effect has been reported [30] in the context of the appearance of strange baryons (e.g. hyperons), which is not a phase transition. Note, that for the EOSs considered, the central density of the maximum mass star, indicated by the filled circle on the curve, lies in the mixed phase so that the pure quark phase is not realized.

Refer to caption
Refer to caption
Figure 6: The two sound speeds (top panel) and their differences (bottom panel) in the mixed phase for the EOS parameters corresponding to Fig. 4. The pure quark phase is not achieved prior to the maximum mass in this case. The termination at nBn_{B}=0.08 fm-3 demarcates the core-crust boundary, since we assume csc_{s}=cec_{e} in the core. Both sound speeds take much smaller values in the crust than in the core. The circle indicates the central density of the maximum mass star (nc,maxn_{c,{\rm max}}=1.22 fm3{\rm fm}^{-3} for Mmax/MM_{\rm max}/M_{\odot}=1.82).

Figure 6 shows results for the individual sound speeds and their differences for the mixed phase. The two sound speeds in the mixed phase behave very differently. Specifically, the equilibrium sound speed suddenly drops (rises) at the onset (end) of the mixed phase, whereas the adiabatic sound speed varies smoothly.

Refer to caption
Figure 7: The Brunt-Väisälä  frequency in a hybrid star of mass 1.4M1.4~{}M_{\odot}. The ZL EOS does not include muons and parameters for the nuclear and quark EOS are as in Fig. 4. Quarks enter at nBn_{B}\simeq 0.42 fm-3 corresponding to r/Rr/R = 0.383, and the mixed phase extends beyond the central density. The value of NN decreases towards the core due to the decreasing value of gg, even as the sound speed difference does not change much.

The Brunt-Väisälä  frequency of a 1.4M1.4~{}M_{\odot} hybrid star is shown in Fig. 7. Note the broader width of the peak when quarks enter, and its location in denser regions of the star, as compared to the nucleonic stars depicted in Fig. 3. This explains why the gg-mode, which is a long-wavelength global oscillation, is strongly impacted by the mixed phase (see results in the next section).

Refer to caption
Refer to caption
Figure 8: The mass-radius curves for a hybrid star (Gibbs construction) with EOS parameters chosen such that the mixed phase supports Mmax=2.05MM_{\rm max}=2.05~{}M_{\odot}. In the left panel, muons are not included, whereas the right panel is with muons included.

Figure 8 shows MM-RR curves for a hybrid star whose Mmax=2.05MM_{\rm max}=2.05~{}M_{\odot}. This value is obtained by increasing the compressibility parameter of the ZL EOS from K0K_{0} = 220 to K0K_{0} = 240 MeV, and increasing γ1\gamma_{1} from 1.6 to 1.8 while maintaining causality. Including muons pushes the onset of the mixed phase to slightly higher densities, which causes the maximum mass of a hybrid star with muons to be higher than for a hybrid star without muons. This is in contrast to the effect of muons in an ordinary NS, where the softening results in a lower maximum mass. The leftmost curves in these figures refer to a self-bound quark star, and are shown here to provide contrast.

V.2 Boundary conditions for the gg-mode oscillation

Having established the equilibrium structure and computed the sound speeds, we have all the variables necessary to solve Eqs. (1) at hand, except for the boundary conditions that determine the (real) eigenfrequencies. The boundary conditions for Newtonian structure equations are obtained as a straightforward limiting case of Eqs. (1), and are discussed at length in [20]. To summarize those results, in the Newtonian non-relativistic case, regularity of ξr,δp/ρ\xi_{r},~{}\delta p/\rho can be checked by Taylor expansion around rr = 0. The resulting condition is:

r2ξr=lω2(Y0+ϕ0)rl+1,δpρ=Y0rl,\displaystyle r^{2}\xi_{r}=\frac{l}{\omega^{2}}(Y_{0}+\phi_{0})r^{l+1},\quad\frac{\delta p}{\rho}=Y_{0}r^{l}\,, (76)

where Y0,ϕ0Y_{0},~{}\phi_{0} are constants. For our purposes, ϕ0\phi_{0} = 0 since we ignore perturbations in the gravitational potential, as in [20]. Y0Y_{0} is an arbitrary normalization constant allowed by the linearity of these equations. Effectively, this means that the overall scale of the eigenfunctions is arbitrary. It must be determined by external factors, such as the strength of the force (tidal effects in a merger, for example). The normalization has no impact on the numerical value of the eigenfrequency. It is therefore conventional to choose Y0Y_{0} = 1. We will make, for simplicity, and without loss of generality, a slightly different choice:

lω2Y0=1\frac{l}{\omega^{2}}Y_{0}=1 (77)

so that (for ll=2), ξrr\xi_{r}\rightarrow r as r0r\rightarrow 0. In practice, we start the integration slight off-center, so ξr\xi_{r} will be small but non-zero. The other condition at the center becomes

δpρ=ω2lrleν0,\frac{\delta p}{\rho}=\frac{\omega^{2}}{l}r^{l}{\rm e}^{-\nu_{0}}\,, (78)

again, with ll = 2 for our case. For the relativistic form of the oscillation equations, the above conditions still apply with the change δpρδpϵ+p\frac{\delta p}{\rho}\rightarrow\frac{\delta p}{\epsilon+p}. The boundary condition at the surface is the vanishing of the Lagrangian pressure perturbation Δp=cs2Δϵ=0\Delta p=c_{s}^{2}\Delta\epsilon=0. This projects out the radial component of ξ\vec{\xi}. In the non-relativistic case, p=ρg\nabla p=-\rho g while in the relativistic case, p=gh\nabla p=-gh with h=(ϵ+p)h=(\epsilon+p) the enthalpy.

With some algebra, one can arrive at a simpler form of Eqs. (1):

dUdr\displaystyle\frac{dU}{dr} =\displaystyle= gcs2U+eλ/2[l(l+1)eνω2r2cs2]V\displaystyle\frac{g}{c_{s}^{2}}U+{\rm e}^{\lambda/2}\left[\frac{l(l+1){\rm e}^{\nu}}{\omega^{2}}-\frac{r^{2}}{c_{s}^{2}}\right]V
dVdr\displaystyle\frac{dV}{dr} =\displaystyle= eλ/2νω2N2r2U+gΔ(c2)V,\displaystyle{\rm e}^{\lambda/2-\nu}\frac{\omega^{2}-N^{2}}{r^{2}}U+g\Delta(c^{-2})V\,, (79)

where UU = r2eλ/2ξrr^{2}{\rm e}^{\lambda/2}\xi_{r}, VV = δp/(ϵ+p)\delta p/(\epsilon+p) and
Δ(c2)=ce2cs2\Delta(c^{-2})=c_{e}^{-2}-c_{s}^{-2}.

We employ a 4th-order Runge-Kutta scheme to find a global solution of the linear perturbation equations, Eqs. (V.2), subject to the boundary conditions for the relativistic case outlined above. Since the solution set comprises overtones, we selected the lowest order gg-mode (highest frequency) by checking that the radial eigenfunction ξr\xi_{r} has only one node inside the star. The corresponding eignefrequency is plotted in the figures that follow. We examine the trends of the gg-mode vs. mass for various parameter choices, for the pure nuclear, self-bound and hybrid stars.

Refer to caption
Refer to caption
Figure 9: Contrasts of the gg-mode frequencies vs mass of normal NSs for the ZL EOS without and with muons. The two curves with different LL’s in each panel are for EOSs with K0=220K_{0}=220 MeV, Sv=31S_{v}=31 MeV and γ1\gamma_{1}=1.6.

Figure 9 contrasts the influence of varying the density dependence of the symmetry energy, by changing the slope of the symmetry energy parameter LL at nsn_{s}, of the underlying ZL EOS for normal neutron stars with fixed K0=220K_{0}=220 MeV and Sv=31S_{v}=31 MeV. For LL = 60 MeV as well as LL=70 MeV, the softening effect of muons leads to a noticeable increase in the gg-mode frequency at a given mass. Comparing LL = 60 MeV with LL = 70 MeV for a fixed composition however, the gg-mode frequency for M>M\buildrel>\over{\sim}~{}0.5-0.6 M\mathrm{M}_{\odot}\;is higher for the stiffer EOS.

Refer to caption
Refer to caption
Figure 10: Contrasts of g-mode frequencies vs stellar mass in a hybrid star. Parameters of the EOSs are as in the insets. In the left panel, muons are not included, whereas the right panel is with muons included.

In Fig. 10, results contrasting the gg-mode frequencies in normal, hybrid, and self-bound stars are presented. The contents of this figure constitute the principal result of this work, viz., the abrupt rise in the scale of the gg-mode frequency at the onset of the mixed phase in the hybrid star. For the EOS parameters displayed in the figure, the jump occurs around 1.4 MM_{\odot}, so that a hybrid star in a merger would have a distinctly higher g-mode frequency than a normal NS. In the top panel, the ZL EOS does not include muons, whereas in the bottom panel the ZL EOS includes muons. The gg-mode frequency in the mixed phase is again higher than in a pure phase, but since the mixed phase appears at a higher density due to muons, the rise in the gg-mode is less dramatic compared to a hybrid star without muons. Results for the self-bound star are shown here for comparison, and to emphasize that its gg-mode frequency is comparatively small (10-50 Hz). Unlike the ff-mode frequency for the hybrid star, which gradually interpolates between those of the normal NS and self-bound star [50, 77] and shows no dramatic effects of compositional changes, the gg-mode frequency for the hybrid star is the highest of all and is sensitive to the onset of quarks - making it less subject to ambiguity. One does not need to know the mass of the star to ascertain if it can be a hybrid star if the gg-mode frequency can be precisely determined.

The unusually large gg-mode frequency for the hybrid star with a Gibbs mixed phase may be understood in a qualitative sense using general thermodynamic considerations without reference to details of the EOS. In general, the equilibrium sound speed ce,mixc_{e,\mathrm{mix}} in a system with two conserved charges (μB\mu_{B} and μQ\mu_{Q}) can be expressed as

ce,mix2\displaystyle c_{e,\mathrm{mix}}^{2} =\displaystyle= dpmix(μB,μQ)dϵmix\displaystyle\frac{dp_{\mathrm{mix}}\left(\mu_{B},\mu_{Q}\right)}{d\epsilon_{\mathrm{mix}}} (80)
=\displaystyle= pmixμB(dμBdϵmix)+pmixμQ(dμQdϵmix)\displaystyle\frac{\partial p_{\mathrm{mix}}}{\partial\mu_{B}}\left(\frac{d\mu_{B}}{d\epsilon_{\mathrm{mix}}}\right)+\frac{\partial p_{\mathrm{mix}}}{\partial\mu_{Q}}\left(\frac{d\mu_{Q}}{d\epsilon_{\mathrm{mix}}}\right)

where μQ\mu_{Q} is the charge chemical potential. Glendenning [61] showed that in such a situation, while μB\mu_{B} is smooth at the onset of the mixed phase, μQ\mu_{Q} is not, as there is freedom to rearrange charges between the two phases to achieve global charge neutrality and minimize the free energy. In fact, the steady rise with density of μQ\mu_{Q} in the pure nuclear phase changes abruptly to a decline in the mixed phase, tempering the equilibrium sound speed as shown by our numerical results presented in Fig. 6 and confirmed by other works [78] which use different EOS from ours for the nucleon and quark sector. On the other hand, the adiabatic sound speed cs,mixc_{s,\mathrm{mix}} is evaluated at fixed composition and shows no such effect, hence the difference of the two sound speeds (usually small in a pure phase) abruptly increases in the mixed phase. This is reflected as a positive jump in the Brunt-Väisälä frequency and therefore of the gg-mode  in the mixed phase.

VI gg-mode Energy and Tidal Forcing

Unlike the Sun, where convection from within can drive oscillations, any oscillations of an evolved NS likely require an external agent to excite and sustain the perturbation beyond its normal damping time. A violent event such as a NS merger is bound to produce oscillations in the pre-merger phase due to tidal forcing or in the postmerger (ringdown) phase as the hypermassive remnant relaxes to its stable rotating configuration. Here, we estimate the impact of the gg-mode on tidal phasing leading up to the merger, as the gg-mode spectrum in the postmerger remnant can be modified by thermal and convective effects which are beyond the scope of the current work. We follow [18] and assume spherically symmetric non-rotating stars, the Newtonian approximation to orbital dynamics and quadrupolar gravitational wave emission. These simplifying approximations allow for a first estimate of the excitation energy and amplitude of the gg-mode, as well as the phase difference due to dynamic tides associated to the gg-mode (not to be confused with the quasi-static tides due to global deformation). Our estimates can be systematically improved by going to the post-Newtonian approximation or numerical relativity.

The estimates are derived by modeling the NS as a forced simple harmonic oscillator with a mass MM_{\ast}, radius RR_{\ast} and a natural frequency ω\omega=ωg\omega_{g}, the angular frequency of the gg-mode. The forcing comes from the quadrupolar moment of the companion star’s gravitational force (mass MM), which couples to the gg-mode. By following the analysis of [18], we arrive at an expression for the accumulated phase error ΔΦ(t)\Delta\Phi(t) caused by the gg-mode:

ΔΦ(t)3πΓm[Ωe(t)Ω1](ΔEE),\displaystyle\Delta\Phi(t)\approx\frac{3\pi\Gamma}{m}\left[\frac{\Omega_{e}(t)}{\Omega}-1\right]\left(-\frac{\Delta E}{E}\right)\,, (81)

where E\triangle E is the energy pumped into the gg-mode, EE the total (kinetic plus potential) orbital energy of the system, Ωe(t)\Omega_{e}(t) the time-dependent orbital frequency of the binary, and Ω\Omega=ωg/m\omega_{g}/m. The quantity mm in Eq. (81) is the azimuthal mode index (mm=2 in this case). Finally, Γ\Gamma is a quantity that appears as a result of applying the stationary phase approximation to the evaluation of the time to resonance [18], and is quantified below. A ΔΦ(t)\Delta\Phi(t) of 𝒪(1){\cal O}(1) signifies a large deviation from the point particle approximation to the gravitational waveform from the merger. Explicitly, the quantity E\triangle E (for angular quantum number ll) is given by

ΔE\displaystyle\Delta E =\displaystyle= (5π384m)M/M[1+M/M](2l+1)/3(c2RGM)5/2\displaystyle\left(\frac{5\pi}{384m}\right)\frac{M/M_{\ast}}{[1+M/M_{\ast}]^{(2l+1)/3}}\left(\frac{c^{2}R_{\ast}}{GM_{\ast}}\right)^{5/2} (82)
×\displaystyle\times (ΩΩd)(4l7)/3(GM2R)Slm2.\displaystyle\left(\frac{\Omega}{\Omega_{d}}\right)^{(4l-7)/3}\left(\frac{GM_{\ast}^{2}}{R_{\ast}}\right)S_{lm}^{2}\,.

where Ωd\Omega_{d}=(GM/R3)1/2(GM_{\ast}/R_{\ast}^{3})^{1/2} is a natural frequency unit and SlmS_{lm} is proportional to the overlap integral between the mode eigenstate |lm|lm\rangle and the vector spherical harmonic |Plm=[rlYlm(θ,ϕ)]\left|P_{lm}\right\rangle=\nabla\left[r^{l}Y_{lm}(\theta,\phi)\right]. The total instantaneous orbital energy is E=GMM/2aE=-GMM_{\ast}/2a with a=a(t)a=a(t) the instantaneous orbital separation. The evolution of the orbital frequency for a circularized orbit using the formula for quadrupolar gravitational wave emission gives

Ωe(τ)=18(aMcc3)5/81τ3/8,\displaystyle\Omega_{e}(\tau)=\frac{1}{8}\left(\frac{aM_{c}}{c^{3}}\right)^{-5/8}\frac{1}{\tau^{3/8}}\,, (83)

where McM_{c} is the chirp mass of the binary system and τ\tau is the time to coalescence. All quantities on the right hand side in Eq. (81) can be calculated, once the parameters of the binary (M,M,RM,M_{\ast},R_{\ast}) and the resonant gg-mode frequency are fixed. We choose MM=MM_{\ast}=1.5 MM_{\odot} for neutron/hybrid stars and pure quark stars. The strongest tidal coupling is likely to the ll=mm=2 gg-mode whose characteristic frequency we choose as

ωg(NS)\displaystyle\omega_{g}(NS) \displaystyle\cong 2π(200)Hz,\displaystyle 2\pi(200)\,{\rm Hz}\,,~{}~{}
ωg(HS)\displaystyle\omega_{g}(HS) \displaystyle\cong 2π(300)Hz,and\displaystyle 2\pi(300)\,{\rm Hz}\,,~{}~{}\rm{and}~{}~{}
ωg(QS)\displaystyle\omega_{g}(QS) \displaystyle\cong 2π(40)Hz,\displaystyle 2\pi(40)\,{\rm Hz}\,,~{}~{} (84)

based on the gg-mode eigenfrequencies in the previous section. Even without computing SlmS_{lm}, one can estimate from Eq. (83) the time at which the gg-mode becomes resonant as

τ0(NS)\displaystyle\tau_{0}(NS) \displaystyle\cong 272ms,\displaystyle 272\,{\rm ms}\,,~{}~{}
τ0(HS)\displaystyle\tau_{0}(HS) \displaystyle\cong 103ms,and\displaystyle 103\,{\rm ms}\,,~{}~{}\rm{and}~{}~{}
τ0(QS)\displaystyle\tau_{0}(QS) \displaystyle\cong 22s,\displaystyle 22\,{\rm s}\,, (85)

where the the zero of time is the moment of coalescence. Assuming circularized orbits, standard equations of binary orbit evolution a(t)a(t) [79] give

a0(NS)\displaystyle a_{0}(NS) \displaystyle\cong 111km,\displaystyle 111~{}\mathrm{km}\,,~{}~{}
a0(HS)\displaystyle a_{0}(HS) \displaystyle\cong 85km,and\displaystyle 85~{}\mathrm{km}\,,~{}~{}{\rm and}~{}~{}
a0(QS)\displaystyle a_{0}(QS) \displaystyle\cong 326km.\displaystyle 326~{}\mathrm{km}\,. (86)

We note that the gg-mode for the hybrid star, which has a larger resonant frequency than neutron or quark stars, is excited later in the merger and is likely to be stronger in amplitude due to the close separation of the binary since the forcing term is 1/a3\propto 1/a^{3} for ll=2. Finally, from our calculations for the gg-mode eigenfunction and the associated density perturbation δϵ(r)\delta\epsilon(r), we estimate

SlmNS\displaystyle S_{lm}^{NS} \displaystyle\cong 4.5×103,\displaystyle\quad 4.5\times 10^{-3}\,,~{}~{}
SlmHS\displaystyle S_{{lm}}^{HS} \displaystyle\cong 6.2×103and\displaystyle\quad 6.2\times 10^{-3}\,~{}~{}{\rm and}~{}~{}
SlmQS\displaystyle S_{lm}^{QS} \displaystyle\cong 9.9×106\displaystyle\quad 9.9\times 10^{-6}\, (87)

using Eq. (40) for SlmS_{lm} from [32]. From these estimates, Eq. (82) can be utilized to yield the estimated fractional orbital energy pumped into the gg-mode:

|ΔEE|NS\displaystyle\left|\frac{\Delta E}{E}\right|^{NS} \displaystyle\cong 2.3×103,\displaystyle 2.3\times 10^{-3}\,,~{}~{}
|ΔEE|HS\displaystyle\left|\frac{\Delta E}{E}\right|^{HS} \displaystyle\cong 5.9×103,and\displaystyle 5.9\times 10^{-3}\,,~{}~{}~{}~{}{\rm and}~{}~{}
|ΔEE|QS\displaystyle\left|\frac{\Delta E}{E}\right|^{QS} \displaystyle\cong 2×109,\displaystyle\quad 2\times 10^{-9}\,, (88)

and finally from Eq. (81), we obtain the phase error due to the resonant excitation of the gg-mode to be

ϕNS\displaystyle\triangle\phi^{NS} \displaystyle\cong 0.8,and\displaystyle 0.8\,,~{}~{}{\rm and}~{}~{}
ϕHS\displaystyle\triangle\phi^{HS} \displaystyle\cong 0.45,\displaystyle 0.45\,,~{}~{}
ϕQS\displaystyle\triangle\phi^{QS} \displaystyle\cong 6×104.\displaystyle 6\times 10^{-4}\,. (89)

Note that ΔϕNS\Delta\phi^{NS} and ϕHS\triangle\phi^{HS} are comparable. Despite (ΔEE)\left(\frac{\Delta E}{E}\right) being larger for a hybrid star as expected, its higher gg-mode frequency means it is excited later in the merger, when there is less time left for accumulating a phase error. These results are very sensitive to the value of SlmS_{lm} (ΔΦSlm2\Delta\Phi\propto S_{lm}^{2}), which itself can vary by a factor of 2 or more depending on the EOS.

Comparison with other works

Table 1: Comparison of characteristic gg-mode frequencies (denoted by ωg\omega_{g} in the table) reported in a selection of the literature. As other works usually fix the stellar mass MM, we include this information. The symbol Λ\Lambda is used here as a shorthand to denote hyperonic degrees of freedom and SF denotes superfluidity in the nucleonic sector. Values of fgf_{g} that vary with the NS mass can be inferred from Figs. 9 and 10 of this work. The entries are representative, not exhaustive.
Authors [Ref.] Core MM fg=ωg/(2π)f_{g}=\omega_{g}/(2\pi)
Composition [MM_{\odot}] [kHz]
Reisenegger & Goldreich [17] npenpe 1.405 0.215
Lai [20] npenpe 1.4 0.073
Kantor & Gusakov [29] npenpe 1.4 0.13
Kantor & Gusakov [29] npeμnpe\mu 1.4 0.19
Kantor & Gusakov [29] npeμnpe\mu(SF) 1.4 0.46
Dommes & Gusakov [30] npeμΛnpe\mu\Lambda(SF) 1.634 0.742
Yu & Weinberg [33] npeμnpe\mu 1.4 0.13
Yu & Weinberg [33] npeμnpe\mu(SF) 2.0 0.45
Rau & Wasserman [34] npeμnpe\mu(SF) 2.0 0.45
Jaikumar  et al. [this work] npenpe 1.4 0.24
Jaikumar  et al. [this work] npeμnpe\mu 1.4 0.27
Jaikumar  et al. [this work] npeμqnpe\mu q 2.0 0.58

Table 1 compares our results for zero-temperature core gg-modes in the Gibbs mixed phase of hadrons and quarks with other works, some of which also find an enhancement of the frequency due to other compositional mixes or collective fluid effects like superfluidity, although values in the table do not include the effect of entrainment on the gg-mode, which has also been studied. Details about the different EOSs used, the effect of non-zero temperature and entrainment can be found by perusing the respective reference. We confirm the value of the gg-mode frequency for npenpe and npeμnpe\mu non-superfluid matter described by the Akmal-Pandharipande-Ravenhall (APR)-EOS as reported in [29], which also serves as a test of our numerics. In comparison to  [29] with the APR-EOS or  [33] with the SLy4 equation of state, the ZL-EOS yields a larger value of the gg-mode frequency as it is less stiff than either of those two. While the EOS and the treatment of gravitational perturbations differ between the cited works, the results for npeμnpe\mu matter with superfluidity appear to be in general agreement with each other. Note the considerably larger value of the gg-mode frequency for hyperonic stars with superfluidity compared to hybrid stars or superfluid neutron stars. All of these, in turn, are larger than non-superfluid neutron/hyperonic stars although the latter employ Newtonian gravity. A study of gg-mode frequencies and damping times in superfluid hybrid stars is a future objective that would make this comparison more complete.

VII Summary and Conclusions

The main objective of this work was to ascertain the characteristics of gg-mode oscillations of NSs containing QM in their cores. Toward this end, the nucleon-to-quark phase transition was treated using Gibbs construction which renders an otherwise sharp first order transition smooth. The cores of such hybrid stars accommodate admixtures of nucleonic and quark matter, the pure quark phase being never achieved. This feature, while allowing contrasts between the structural properties (e.g., MM vs RR) of normal and hybrid stars to be made also permits comparisons of observables that depend on their interior compositions, such as short- and long-term cooling, oscillation modes, etc. Determining the composition of the star is essential to break the degeneracy that exists in the masses and radii of normal and hybrid stars as one may be masquerading as the other.

The nucleonic part of the EOS used in this work tallies with nuclear systematics near and below nsn_{s} in addition to being consistent with results from modern chiral EFT calculations up to 2ns2n_{s} for which uncertainty quantifications have been made. The EOS employed in the quark sector is sufficiently stiff, hence non-perturbative, to support 2M\sim 2M_{\odot} NSs required by recent observations. Furthermore, the overall EOS gives radii of 1.4M\sim 1.4M_{\odot} stars that lie within the bounds of recent determinations. The EOS is also consistent with the tidal deformation inferred from gravitational wave detection in the event GW170817. Appendix A summarizes the structural properties for the EOSs used and provides mathematical details for the derivation of the sound speeds.

Unlike for MM-RR curves for which only the pressure vs density relation (EOS) is sufficient, the analysis of gg-mode oscillations requires simultaneous information about the equilibrium and adiabatic squared sound speeds, ce2=dp/dϵc_{e}^{2}=dp/d\epsilon and cs2=p/ϵ|xc_{s}^{2}=\partial p/\partial\epsilon|_{x}, where xx is the local proton fraction. The distinction between these two sound speeds plays a central role in determining the Brunt-Väisälä frequencies ω2ce2cs2\omega^{2}\propto c_{e}^{-2}-c_{s}^{-2} of non-radial gg-mode oscillations. Thus, a future detection of gg-modes would take gravitational wave astronomy beyond the current capability of MM-RR measurements to determine the composition of the star.

We find that the gg-mode is sensitive to the presence of QM in NSs, where quarks are part of a mixed phase with nuclear matter in the core. The equilibrium sound speed drops sharply at the boundary of the mixed phase (Fig. 5), raising the local Brunt-Väisälä  frequency and the fundamental gg-mode  frequency of the star (Fig. 6). Contrasts of gg-mode frequencies between normal and hybrid stars containing quark matter (Fig. 9) form the principal results of our work.

Our analysis leads to the conclusion that in binary mergers where one or both components may be a hybrid star, the fraction of tidal energy pumped into the resonant gg-mode in hybrid stars can exceed that of a NS by a factor of 2-3, although resonance occurs later in the inspiral. On the other hand, a self-bound star has a much weaker tidal overlap with the g-mode. The cumulative tidal phase error in hybrid stars Δϕ\Delta\phi\cong 0.5 is comparable to that from tides in ordinary NSs. While this happenstance may present a challenge in distinguishing between the two cases, should the g-mode be excited to sufficient amplitude in a postmerger remnant, its frequency spectrum would be a possible indication for the existence of non-nucleonic matter, including quarks. The detection of such gg-mode frequencies in binary mergers observed by current gravitational wave detectors seems challenging, but possible with next generation detectors.

The novel features of this work include (i) use of nucleonic EOSs that are consistent with constraints from modern chiral EFT calculations coupled with sufficiently stiff quark EOSs to calculate structural properties of hybrid stars that lie within the bounds of astrophysical measurements, (ii) a first calculation of the two sound speeds and the principal gg-mode frequency of hybrid stars employing Gibbs phase criteria, and (iii) a concomitant analysis of tidal phase effects in a binary merger due to gg-modes in hybrid stars. In future work, we aim to report on gg-mode frequencies in alternative treatments of quark matter in NSs such as a first order nucleon-to-quark phase transition and crossover transitions as in quarkyonic matter.

Acknowledgments.— We are grateful to the anonymous referee for meticulous review of the equations. We acknowledge discussions with Thomas Klähn on the non-perturbative EOS for quark matter. Thanks are due to Sophia Han for remarks on observational constraints on the EOS. P.J. is supported by the U.S. NSF Grant No. PHY-1608959 and PHY-1913693. The research of A.S. and M.P. was supported by the U.S. Department of Energy, Grant No. DE-FG02-93ER-40756. C.C. acknowledges support from the European Union’s Horizon 2020 research and innovation programme under the Marie Skłodowska-Curie grant agreement No. 754496 (H2020-MSCA-COFUND-2016 FELLINI).

Appendix A Determination of EOS constants in SNM and PNM for the ZL EOS

A.1 SNM

The constants a0,b0a_{0},b_{0} and γ\gamma in Eq. (10) for SNM are determined by utilizing the empirical properties of SNM at u=1u=1. Specifically, the values used are E1/2=B=16E_{1/2}=-B=-16 MeV at ns=0.16fm3n_{s}=0.16~{}{\rm fm^{-3}}, p1/2/ns=0p_{1/2}/n_{s}=0, and K1/2=220K_{1/2}=220 MeV. Manipulating the relations

B\displaystyle-B =\displaystyle= T1/2+a0+b0\displaystyle T_{1/2}+a_{0}+b_{0} (90)
0\displaystyle 0 =\displaystyle= T1/2+a0+γb0\displaystyle T_{1/2}^{\prime}+a_{0}+\gamma b_{0} (91)
K1/29\displaystyle\frac{K_{1/2}}{9} =\displaystyle= T1/2′′+2T1/2+2a0+γ(γ+1)b0,\displaystyle T_{1/2}^{\prime\prime}+2T_{1/2}^{\prime}+2a_{0}+\gamma(\gamma+1)b_{0}\,, (92)

the constants are given by

γ\displaystyle\gamma =\displaystyle= K1/2/9T1/2′′T1/2T1/2+B,\displaystyle\frac{K_{1/2}/9-T_{1/2}^{\prime\prime}}{T_{1/2}-T_{1/2}^{\prime}+B}\,,\quad
b0\displaystyle b_{0} =\displaystyle= K1/2/9T1/2′′γ(γ1)and\displaystyle\frac{K_{1/2}/9-T_{1/2}^{\prime\prime}}{\gamma(\gamma-1)}\quad{\rm and}
a0\displaystyle a_{0} =\displaystyle= BT1/2b0,\displaystyle-B-T_{1/2}-b_{0}\,, (93)

where T1/2=udT1/2duT_{1/2}^{\prime}=u\frac{dT_{1/2}}{du} and T1/2′′=u2d2T1/2du2T_{1/2}^{\prime\prime}=u^{2}\frac{d^{2}T_{1/2}}{du^{2}}. Explicit expressions for these derivatives are

T1/2\displaystyle T_{1/2}^{\prime} =\displaystyle= p1/2kinn|ns\displaystyle\left.\frac{p_{1/2}^{\rm kin}}{n}\right|_{n_{s}}
=\displaystyle= 1ns212π2[kFEF(kF232MB2)\displaystyle\frac{1}{n_{s}}\cdot\frac{2}{12\pi^{2}}\bigg{[}k_{F}E_{F}\left(k_{F}^{2}-\frac{3}{2}M_{B}^{2}\right)
+\displaystyle+ 32MB4ln(kF+EFMB)]kFs,\displaystyle\frac{3}{2}M_{B}^{4}\ln\left(\frac{k_{F}+E_{F}}{M_{B}}\right)\bigg{]}_{k_{Fs}}\,,
T1/2′′\displaystyle T_{1/2}^{\prime\prime} =\displaystyle= K1/2kin92T/12=kFs23EFs2T/12,\displaystyle\frac{K_{1/2}^{\rm kin}}{9}-2T_{/12}^{\prime}=\frac{k_{Fs}^{2}}{3E_{Fs}}-2T_{/12}^{\prime}, (94)

where kFs=(3π2ns/2)1/3k_{Fs}=(3\pi^{2}n_{s}/2)^{1/3}. To obtain the first term in the rightmost equality above, it is advantageous to use the thermodynamical identity p=nμϵp=n\mu-\epsilon for the kinetic parts, whence dpdn=ndμdn=dμdkFdkFdn\frac{dp}{dn}=n\frac{d\mu}{dn}=\frac{d\mu}{dk_{F}}\frac{dk_{F}}{dn}. The result quoted above ensues from the relations dkFdn=kF3n\frac{dk_{F}}{dn}=\frac{k_{F}}{3n} and μ=EF=kF2+MB2\mu=E_{F}=\sqrt{k_{F}^{2}+M_{B}^{2}} both evaluated at nsn_{s} and kFsk_{Fs}.

Numerical values of the derivatives and constants so derived are

T1/2\displaystyle T_{1/2} \displaystyle\simeq 21.79MeV,T1/214.34MeV,\displaystyle 21.79~{}{\rm MeV},~{}~{}~{}T_{1/2}^{\prime}\simeq 14.34~{}{\rm MeV,}
T1/2′′\displaystyle T_{1/2}^{\prime\prime} =\displaystyle= 5.030MeV,γ1.256,\displaystyle-5.030~{}{\rm MeV,}~{}~{}~{}\gamma\simeq 1.256,
a0\displaystyle~{}~{}~{}a_{0} \displaystyle\simeq 129.3MeV,andb091.49MeV.\displaystyle-129.3~{}{\rm MeV},\quad{\rm and}~{}~{}~{}b_{0}\simeq 91.49~{}{\rm MeV}. (95)

as in ZL. For other permissible values of K1/2K_{1/2} in the range 220±30220\pm 30 MeV, Eqs. (A.1, A.1) can be used to determine the corresponding constants.

A.2 PNM

In the PNM sector in which x=0x=0, the constants in Eq. (15) to determined are a1,b1a_{1},b_{1} and γ1\gamma_{1}. As in SNM, E0E_{0} and T0T_{0} are relative to the baryon mass MBM_{B}. Denoting the energy per baryon of PNM by E0E_{0}, its various terms and the associated pressure are

E0\displaystyle E_{0} =\displaystyle= T0+V0=T0+a1u+b1u1γ,\displaystyle T_{0}+V_{0}=T_{0}+a_{1}u+b_{1}u^{\gamma}_{1}\,,
p0\displaystyle p_{0} =\displaystyle= ns(u2dE0du)\displaystyle n_{s}\left(u^{2}\frac{dE_{0}}{du}\right) (96)
=\displaystyle= ns(u2T0+a1u2+γ1b1uγ1+1).\displaystyle n_{s}\left(u^{2}T_{0}^{\prime}+a_{1}u^{2}+\gamma_{1}b_{1}u^{\gamma_{1}+1}\right)\,.

Evaluating the above equations at u=1u=1 leads to

E0=SvB=T0+a1+b1,\displaystyle E_{0}=S_{v}-B=T_{0}+a_{1}+b_{1}\,, (97)
p0=ns(T0+a1+γ1b1),\displaystyle p_{0}=n_{s}(T_{0}^{\prime}+a_{1}+\gamma_{1}b_{1})\,, (98)

where Sv=(E0E1/2)S_{v}=(E_{0}-E_{1/2}) at u=1u=1. The last equation above is generally written as

p0ns\displaystyle\frac{p_{0}}{n_{s}} =\displaystyle= L3withL=3[ndSvdn]ns=3[uSv]u=1sothat\displaystyle\frac{L}{3}\quad{\rm with}\quad L=3\left[n\frac{dS_{v}}{dn}\right]_{n_{s}}=3~{}[uS_{v}^{\prime}]_{u=1}\quad{\rm so~{}that}
L3\displaystyle\frac{L}{3} =\displaystyle= T0+a1+γ1b1,\displaystyle T_{0}^{\prime}+a_{1}+\gamma_{1}b_{1}\,, (99)

where Sv=dSvduS_{v}^{\prime}=\frac{dS_{v}}{du}. Manipulating Eqs. (97) and (99) leads to the relations

b1\displaystyle b_{1} =\displaystyle= L3+BSv+T0T0γ11and\displaystyle\frac{\frac{L}{3}+B-S_{v}+T_{0}-T_{0}^{\prime}}{\gamma_{1}-1}\quad{\rm and}\quad
a1\displaystyle a_{1} =\displaystyle= SvBT0b1.\displaystyle S_{v}-B-T_{0}-b_{1}\,. (100)

Taking guidance from the empirical properties of isospin asymmetric nuclear matter, we choose Sv=31S_{v}=31 MeV, LL in the range (3030-7070) MeV, and γ1=5/3\gamma_{1}=5/3. The resulting values of the constants are

a1\displaystyle a_{1} \displaystyle\simeq (L2+14.72)MeVand\displaystyle-\left(\frac{L}{2}+14.72\right)~{}{\rm MeV}\quad{\rm and}\quad
b1\displaystyle b_{1} \displaystyle\simeq (L24.62)MeV.\displaystyle\left(\frac{L}{2}-4.62\right)~{}{\rm MeV}. (101)

A.3 Sensitivity of the EOS constants

The EOS constants above depend on the input values of B,ns,K0B,~{}n_{s},~{}K_{0} and Sv,L,γ1S_{v},~{}L,~{}\gamma_{1} in SNM and PNM, respectively. Although we have used representative values for these quantities at u=1u=1, nuclear data permits variations in them. Furthermore, one or more sets of these constants may be correlated, as for example, SVS_{V} and LL. Additional constraints are to support 2M\simeq 2M_{\odot} NSs and to maintain causality, at least within the star. These points must be borne in mind when varying the input values, particularly when correlated errors are present in theoretical evaluations of these quantities.

A.4 NS properties with the ZL EOSs

The various properties shown in Table 2 below are for beta-equilibrated normal NSs and correspond to variations in the characteristic properties of the ZL EOSs.

Table 2: Structural properties of nucleonic NSs with M=1.4MM=1.4\,M_{\odot} and MmaxM_{\rm max} for the ZL EOSs. For each mass, the compactness parameter β=(GM/c2R)(1.475/R)(M/M)\beta=(GM/c^{2}R)\simeq(1.475/R)(M/M_{\odot}), ncn_{c}, pcp_{c} and ycy_{c} are the central values of the density, pressure and proton fraction, respectively. The corresponding equilibrium squared speeds of sound are denoted by ce2c_{e}^{2}. The Λ\Lambda’s denote tidal deformabilities.
Property ZL-A ZL-B ZL-C Units
K0K_{0} 220 220 240 MeV
SvS_{v} 31 31 31 MeV
LL 50 70 60 MeV
γ1\gamma_{1} 1.6 1.6 1.8
R1.4R_{1.4} 11.77 12.69 12.61 km
β1.4\beta_{1.4} 0.175 0.163 0.164
nc,1.4/nsn_{c,1.4}/n_{s} 3.35 2.78 2.75
pc,1.4p_{c,1.4} 83.65 60.59 61.50 MeVfm3{\rm MeV~{}fm^{-3}}
(ce2)c,1.4(c_{e}^{2})_{c,1.4} 0.385 0.345 0.363 c2c^{2}
Λ1.4\Lambda_{1.4} 713.4 970.2 504.1
RmaxR_{\rm max} 10.01 10.68 10.8 km
MmaxM_{\rm max} 1.997 2.02 2.13 MM_{\odot}
βmax\beta_{\rm max} 0.294 0.279 0.291
nc,max/nsn_{\rm c,max}/n_{s} 7.71 6.96 6.67
pc,maxp_{\rm c,max} 798.89 602.04 646.7 MeVfm3{\rm MeV~{}fm^{-3}}
(ce2)c,max(c_{e}^{2})_{\rm c,max} 0.874 0.777 0.866 c2c^{2}
Λmax\Lambda_{\rm max} 9.11 9.6 6.39

Structural properties of hybrid stars are discussed and shown in various figures in the text.

A.5 Proof of equivalence of Eqs. (40) and (51)

Here we establish the equivalence of the direct approach to computing cs2c_{s}^{2} from Eq. (40) with that from Eq. (51) for a general EOS with a parabolic dependence of the proton fraction xx in the case of n,p,en,p,e matter. Both approaches yield identical results, which we have verified numerically as well.

The equilibrium squared sound speed is

ce2=(dPdϵ)eq=nBddnB(Pbar+Pe)(ϵ+P),c_{e}^{2}=\left(\frac{dP}{d\epsilon}\right)_{\rm eq}=\frac{n_{B}\frac{d}{dn_{B}}\left(P_{\rm bar}+P_{e}\right)}{\left(\epsilon+P\right)}\,, (102)

where the total pressure P=Pbar+PeP=P_{\rm bar}+P_{e} is comprised of the pressure from baryons and electrons. Writing

ddnB=nB+dxdnBx,\frac{d}{dn_{B}}=\frac{\partial}{\partial n_{B}}+\frac{dx}{dn_{B}}\frac{\partial}{\partial x}\,, (103)

we get

ce2=(nBPbarnB+nBPenB)ϵ+P+nB(Pbar xdxdnB+PexdxdnB)ϵ+P.c_{e}^{2}=\frac{\left(n_{B}\frac{\partial P_{bar}}{\partial n_{B}}+n_{B}\frac{\partial P_{e}}{\partial n_{B}}\right)}{\epsilon+P}+\frac{n_{B}\left(\frac{\partial P_{\text{bar }}}{\partial x}\frac{dx}{dn_{B}}+\frac{\partial P_{\text{e}}}{\partial x}\frac{dx}{dn_{B}}\right)}{\epsilon+P}\,. (104)

Comparing with Eq. (39), the first term on the right hand side is simply cs2c_{s}^{2}. For the specific case of an EOS with parabolic dependence in xx (of which the APR-EOS in Kantor & Gusakov [29] and the ZL-EOS  in Zhao and Lattimer [43] are examples), we have

E(u,x)=E0(u)+(12x)2S2(u);u=nB/n0.E(u,x)=E_{0}(u)+(1-2x)^{2}S_{2}(u);\quad u=n_{B}/n_{0}\,. (105)

where n0n_{0} is the saturation density, E0E_{0} the energy per baryons (neutrons and protons) and S2S_{2} the symmetry energy. Computing the pressure and its derivatives with respect to nBn_{B} and xx for the EOS in Eq. (105), we find

Pbarx=4nB(12x)uS2,\frac{\partial P_{\text{bar}}}{\partial x}=-4n_{B}(1-2x)uS_{2}^{\prime}\,, (106)

where the prime on S2S_{2} is with respect to uu. Since (assuming massless electrons) Pex=13nBμe\frac{\partial P_{\text{e}}}{\partial x}=\frac{1}{3}n_{B}\mu_{e}, it follows that

ce2=cs2+1μn(μe34(12x)uS2)nBdxdnB.c_{e}^{2}=c_{s}^{2}+\frac{1}{\mu_{n}}\left(\frac{\mu_{e}}{3}-4(1-2x)uS_{2}^{\prime}\right)n_{B}\frac{dx}{dn_{B}}\,. (107)

From the β\beta-equilibrium condition

μe=4S2(12x)μe/(4S2)=(12x),\mu_{e}=4S_{2}(1-2x)\Rightarrow\mu_{e}/(4S_{2})=(1-2x)\,, (108)

we get upon differentiation

14S2(dμednB)(μe4S22)(dS2dnB)=2(dxdnB).\frac{1}{4S_{2}}\left(\frac{d\mu_{e}}{dn_{B}}\right)-\left(\frac{\mu_{e}}{4S_{2}^{2}}\right)\left(\frac{dS_{2}}{dn_{B}}\right)=-2\left(\frac{dx}{dn_{B}}\right)\,. (109)

Using

(dμednB)=μe3x(dxdnB)+μe3nB,\left(\frac{d\mu_{e}}{dn_{B}}\right)=\frac{\mu_{e}}{3x}\left(\frac{dx}{dn_{B}}\right)+\frac{\mu_{e}}{3n_{B}}\,, (110)

and solving for dx/dnBdx/dn_{B} from Eq.(109), a minor rearrangement yields

nB(dxdnB)=(μe34(12x)uS2)(μe3x+8S2).n_{B}\left(\frac{dx}{dn_{B}}\right)=\frac{-\left(\frac{\mu_{e}}{3}-4(1-2x)uS_{2}^{\prime}\right)}{\left(\frac{\mu_{e}}{3x}+8S_{2}\right)}\,. (111)

Putting together Eq. (111) with Eq. (107), we get

ce2=cs21μn(μe34(12x)uS2)2(μe3x+8S2).\displaystyle c_{e}^{2}=c_{s}^{2}-\frac{1}{\mu_{n}}\frac{\left(\frac{\mu_{e}}{3}-4(1-2x)uS_{2}^{\prime}\right)^{2}}{\left(\frac{\mu_{e}}{3x}+8S_{2}\right)}\,. (112)

Comparing Eq.(112) with Eqs. (51), (55) and (57) in the text, namely,

cs2\displaystyle c_{s}^{2} =\displaystyle= ce2+[nB(μ~nB)x]2μn(μ~x)nB\displaystyle c_{e}^{2}+\frac{\left[n_{B}\left(\frac{\partial\tilde{\mu}}{\partial n_{B}}\right)_{x}\right]^{2}}{\mu_{n}\left(\frac{\partial\tilde{\mu}}{\partial x}\right)_{n_{B}}} (113)
nBμ~nB\displaystyle n_{B}\frac{\partial\tilde{\mu}}{\partial n_{B}} =\displaystyle= μe34(12x)uS2\displaystyle\frac{\mu_{e}}{3}-4(1-2x)~{}uS_{2}^{\prime} (114)
μ~x\displaystyle\frac{\partial\tilde{\mu}}{\partial x} =\displaystyle= 13μex+8S2(u),\displaystyle\frac{1}{3}\frac{\mu_{e}}{x}+8S_{2}(u)\,, (115)

we see that Eq. (51) is consistent with the direct definition of cs2c_{s}^{2} from Eq. (40) and that Eq. (51) applies in general for any form of the EOS with a parabolic dependence in xx, although in the text we arrived at Eqs. (55) and (57) in the context of the ZL-EOS.

A.6 Derivation of Eqs. (59)-(61)

In npeμnpe\mu matter, we choose the independent variables to be the baryon density nBn_{B}, the lepton fraction xx, and the muon fraction yxμy\equiv x_{\mu}. The electron fraction xex_{e} is the difference xyx-y.

The starting point for the speed-of-sound difference is

cs2ce2=1μavgPnB|x,y1μndPdnB.c_{s}^{2}-c_{e}^{2}=\frac{1}{\mu_{avg}}\left.\frac{\partial P}{\partial n_{B}}\right|_{x,y}-\frac{1}{\mu_{n}}\frac{dP}{dn_{B}}~{}. (116)

The total derivative of the pressure P(nB,x,y)P(n_{B},x,y) with respect to nBn_{B} is given by

dPdnB=PnB|x,y+Px|nB,ydxdnB+Py|nB,xdydnB\frac{dP}{dn_{B}}=\left.\frac{\partial P}{\partial n_{B}}\right|_{x,y}+\left.\frac{\partial P}{\partial x}\right|_{n_{B},y}\frac{dx}{dn_{B}}+\left.\frac{\partial P}{\partial y}\right|_{n_{B},x}\frac{dy}{dn_{B}} (117)

and therefore

cs2\displaystyle c_{s}^{2} \displaystyle- ce2=(1μavg1μn)PnB|x,y\displaystyle c_{e}^{2}=\left(\frac{1}{\mu_{avg}}-\frac{1}{\mu_{n}}\right)\left.\frac{\partial P}{\partial n_{B}}\right|_{x,y}
\displaystyle- 1μn(Px|nB,ydxdnB+Py|nB,xdydnB)\displaystyle\frac{1}{\mu_{n}}\left(\left.\frac{\partial P}{\partial x}\right|_{n_{B},y}\frac{dx}{dn_{B}}+\left.\frac{\partial P}{\partial y}\right|_{n_{B},x}\frac{dy}{dn_{B}}\right)
=\displaystyle= (μnμavgμavgμn)PnB|x,y\displaystyle\left(\frac{\mu_{n}-\mu_{avg}}{\mu_{avg}~{}\mu_{n}}\right)\left.\frac{\partial P}{\partial n_{B}}\right|_{x,y}
\displaystyle- 1μn(Px|nB,ydxdnB+Py|nB,xdydnB).\displaystyle\frac{1}{\mu_{n}}\left(\left.\frac{\partial P}{\partial x}\right|_{n_{B},y}\frac{dx}{dn_{B}}+\left.\frac{\partial P}{\partial y}\right|_{n_{B},x}\frac{dy}{dn_{B}}\right). (119)

The average chemical potential μavg\mu_{avg} is

μavg=(1x)μn+xμp+(xy)μe+yμy\mu_{avg}=(1-x)\mu_{n}+x\mu_{p}+(x-y)\mu_{e}+y\mu_{y} (120)

which means that

μnμavg\displaystyle\mu_{n}-\mu_{avg} =\displaystyle= x(μnμpμe)+y(μeμμ)\displaystyle x(\mu_{n}-\mu_{p}-\mu_{e})+y(\mu_{e}-\mu_{\mu}) (121)
\displaystyle\equiv xμ~xyμ~y\displaystyle-x\tilde{\mu}_{x}-y\tilde{\mu}_{y}

with the obvious definitions for μ~x\tilde{\mu}_{x} and μ~y\tilde{\mu}_{y}.

In β\beta-equilibrium μn=μavg\mu_{n}=\mu_{avg} (as well as μ~x=μ~y=0\tilde{\mu}_{x}=\tilde{\mu}_{y}=0) and, correspondingly,

cs2ce2=1μn(Px|nB,ydxdnB+Py|nB,xdydnB),c_{s}^{2}-c_{e}^{2}=-\frac{1}{\mu_{n}}\left(\left.\frac{\partial P}{\partial x}\right|_{n_{B},y}\frac{dx}{dn_{B}}+\left.\frac{\partial P}{\partial y}\right|_{n_{B},x}\frac{dy}{dn_{B}}\right)~{}, (122)

which is Eq. (59) in the main text. Using P=nB2EnB|x,yP=n_{B}^{2}\left.\frac{\partial E}{\partial n_{B}}\right|_{x,y}, the speed-of-sound difference can be expressed as

cs2ce2\displaystyle c_{s}^{2}-c_{e}^{2} =\displaystyle= nB2μnnB(Ex|nB,ydxdnB+Ey|nB,xdydnB)|x,y\displaystyle-\frac{n_{B}^{2}}{\mu_{n}}\frac{\partial}{\partial n_{B}}\left.\left(\left.\frac{\partial E}{\partial x}\right|_{n_{B},y}\frac{dx}{dn_{B}}+\left.\frac{\partial E}{\partial y}\right|_{n_{B},x}\frac{dy}{dn_{B}}\right)\right|_{x,y} (123)
=\displaystyle= nB2μn(μ~xnB|x,ydxdnB+μ~ynB|x,ydydnB).\displaystyle-\frac{n_{B}^{2}}{\mu_{n}}\left(\left.\frac{\partial\tilde{\mu}_{x}}{\partial n_{B}}\right|_{x,y}\frac{dx}{dn_{B}}+\left.\frac{\partial\tilde{\mu}_{y}}{\partial n_{B}}\right|_{x,y}\frac{dy}{dn_{B}}\right)~{}.

The calculation of dxdnB\frac{dx}{dn_{B}} and dydnB\frac{dy}{dn_{B}} begins from the total differentials of μ~x\tilde{\mu}_{x} and μ~y\tilde{\mu}_{y} which are

dμ~x=μ~xnB|x,ydnB+μ~xx|nB,ydx+μ~xy|nB,xdy=0d\tilde{\mu}_{x}=\left.\frac{\partial\tilde{\mu}_{x}}{\partial n_{B}}\right|_{x,y}dn_{B}+\left.\frac{\partial\tilde{\mu}_{x}}{\partial x}\right|_{n_{B},y}dx+\left.\frac{\partial\tilde{\mu}_{x}}{\partial y}\right|_{n_{B},x}dy=0 (124)

and

dμ~y=μ~ynB|x,ydnB+μ~yx|nB,ydx+μ~yy|nB,xdy=0.d\tilde{\mu}_{y}=\left.\frac{\partial\tilde{\mu}_{y}}{\partial n_{B}}\right|_{x,y}dn_{B}+\left.\frac{\partial\tilde{\mu}_{y}}{\partial x}\right|_{n_{B},y}dx+\left.\frac{\partial\tilde{\mu}_{y}}{\partial y}\right|_{n_{B},x}dy=0~{}. (125)

From the former differential, it follows that

dy=1μ~xy|nB,x(μ~xnB|x,ydnB+μ~xx|nB,ydx)dy=\frac{-1}{\left.\frac{\partial\tilde{\mu}_{x}}{\partial y}\right|_{n_{B},x}}\left(\left.\frac{\partial\tilde{\mu}_{x}}{\partial n_{B}}\right|_{x,y}dn_{B}+\left.\frac{\partial\tilde{\mu}_{x}}{\partial x}\right|_{n_{B},y}dx\right) (126)

which, when substituted in the latter, leads to

0\displaystyle 0 =\displaystyle= μ~ynB|x,ydnB+μ~yx|nB,ydx\displaystyle\left.\frac{\partial\tilde{\mu}_{y}}{\partial n_{B}}\right|_{x,y}dn_{B}+\left.\frac{\partial\tilde{\mu}_{y}}{\partial x}\right|_{n_{B},y}dx (127)
\displaystyle- μ~yy|nB,xμ~xy|nB,x(μ~xnB|x,ydnB+μ~xx|nB,ydx).\displaystyle\frac{\left.\frac{\partial\tilde{\mu}_{y}}{\partial y}\right|_{n_{B},x}}{\left.\frac{\partial\tilde{\mu}_{x}}{\partial y}\right|_{n_{B},x}}\left(\left.\frac{\partial\tilde{\mu}_{x}}{\partial n_{B}}\right|_{x,y}dn_{B}+\left.\frac{\partial\tilde{\mu}_{x}}{\partial x}\right|_{n_{B},y}dx\right)~{}.

One then collects terms proportional to dnBdn_{B} and dxdx

(μ~ynB|x,yμ~yy|nB,xμ~xy|nB,xμ~xnB|x,y)dnB\displaystyle\left(\left.\frac{\partial\tilde{\mu}_{y}}{\partial n_{B}}\right|_{x,y}-\frac{\left.\frac{\partial\tilde{\mu}_{y}}{\partial y}\right|_{n_{B},x}}{\left.\frac{\partial\tilde{\mu}_{x}}{\partial y}\right|_{n_{B},x}}\left.\frac{\partial\tilde{\mu}_{x}}{\partial n_{B}}\right|_{x,y}\right)dn_{B} (128)
=\displaystyle= (μ~yx|nB,yμ~yy|nB,xμ~xy|nB,xμ~xx|nB,y)dx\displaystyle-\left(\left.\frac{\partial\tilde{\mu}_{y}}{\partial x}\right|_{n_{B},y}-\frac{\left.\frac{\partial\tilde{\mu}_{y}}{\partial y}\right|_{n_{B},x}}{\left.\frac{\partial\tilde{\mu}_{x}}{\partial y}\right|_{n_{B},x}}\left.\frac{\partial\tilde{\mu}_{x}}{\partial x}\right|_{n_{B},y}\right)dx

or, equivalently,

dxdnB=μ~xy|nB,xμ~ynB|x,yμ~yy|nB,xμ~xnB|x,yμ~xy|nB,xμ~yx|nB,yμ~yy|nB,xμ~xx|nB,y.\frac{dx}{dn_{B}}=\frac{\left.\frac{\partial\tilde{\mu}_{x}}{\partial y}\right|_{n_{B},x}\left.\frac{\partial\tilde{\mu}_{y}}{\partial n_{B}}\right|_{x,y}-\left.\frac{\partial\tilde{\mu}_{y}}{\partial y}\right|_{n_{B},x}\left.\frac{\partial\tilde{\mu}_{x}}{\partial n_{B}}\right|_{x,y}}{\left.\frac{\partial\tilde{\mu}_{x}}{\partial y}\right|_{n_{B},x}\left.\frac{\partial\tilde{\mu}_{y}}{\partial x}\right|_{n_{B},y}-\left.\frac{\partial\tilde{\mu}_{y}}{\partial y}\right|_{n_{B},x}\left.\frac{\partial\tilde{\mu}_{x}}{\partial x}\right|_{n_{B},y}}~{}. (129)

Similarly,

dydnB=μ~xx|nB,yμ~ynB|x,yμ~yx|nB,yμ~xnB|x,yμ~xx|nB,yμ~yy|nB,xμ~yx|nB,yμ~xy|nB,x.\frac{dy}{dn_{B}}=\frac{\left.\frac{\partial\tilde{\mu}_{x}}{\partial x}\right|_{n_{B},y}\left.\frac{\partial\tilde{\mu}_{y}}{\partial n_{B}}\right|_{x,y}-\left.\frac{\partial\tilde{\mu}_{y}}{\partial x}\right|_{n_{B},y}\left.\frac{\partial\tilde{\mu}_{x}}{\partial n_{B}}\right|_{x,y}}{\left.\frac{\partial\tilde{\mu}_{x}}{\partial x}\right|_{n_{B},y}\left.\frac{\partial\tilde{\mu}_{y}}{\partial y}\right|_{n_{B},x}-\left.\frac{\partial\tilde{\mu}_{y}}{\partial x}\right|_{n_{B},y}\left.\frac{\partial\tilde{\mu}_{x}}{\partial y}\right|_{n_{B},x}}~{}. (130)

The speed-of-sound difference, as given by Eqs. (123), (129), and (130) is physically transparent because β\beta-equilibrium and compositional gradients are brought to the forefront via μ~x(y)\tilde{\mu}_{x(y)} and /x(y)\partial/\partial x(y), respectively (the latter two equations are Eqs. (60) and (61) in the main text). However, this intuitive picture comes at the expense of computational complexity.

References