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

Stability of relativistic stars with scalar hairs

Ryotaro Kase1, Rampei Kimura2, Seiga Sato3, and Shinji Tsujikawa3 1Department of Physics, Faculty of Science, Tokyo University of Science, 1-3, Kagurazaka, Shinjuku-ku, Tokyo 162-8601, Japan
2Waseda Institute for Advanced Study, Waseda University 19th building, 1-21-1 Nishiwaseda, Shinjuku-ku, Tokyo 169-0051, Japan
3Department of Physics, Waseda University, 3-4-1 Okubo, Shinjuku, Tokyo 169-8555, Japan
Abstract

We study the stability of relativistic stars in scalar-tensor theories with a nonminimal coupling of the form F(ϕ)RF(\phi)R, where FF depends on a scalar field ϕ\phi and RR is the Ricci scalar. On a spherically symmetric and static background, we incorporate a perfect fluid minimally coupled to gravity as a form of the Schutz-Sorkin action. The odd-parity perturbation for the multipoles l2l\geq 2 is ghost-free under the condition F(ϕ)>0F(\phi)>0, with the speed of gravity equivalent to that of light. For even-parity perturbations with l2l\geq 2, there are three propagating degrees of freedom arising from the perfect-fluid, scalar-field, and gravity sectors. For l=0,1l=0,1, the dynamical degrees of freedom reduce to two modes. We derive no-ghost conditions and the propagation speeds of these perturbations and apply them to concrete theories of hairy relativistic stars with F(ϕ)>0F(\phi)>0. As long as the perfect fluid satisfies a weak energy condition with a positive propagation speed squared cm2c_{m}^{2}, there are neither ghost nor Laplacian instabilities for theories of spontaneous scalarization and Brans-Dicke (BD) theories with a BD parameter ωBD>3/2\omega_{\rm BD}>-3/2 (including f(R)f(R) gravity). In these theories, provided 0<cm210<c_{m}^{2}\leq 1, we show that all the propagation speeds of even-parity perturbations are sub-luminal inside the star, while the speeds of gravity outside the star are equivalent to that of light.

pacs:
04.50.Kd, 95.36.+x, 98.80.-k

I Introduction

The dawn of gravitational-wave (GW) astronomies Abbott2016 shed new light on the physics around compact objects such as black holes and neutron stars (NSs). For example, the GW170817 event GW170817 arising from binary NSs placed constraints on the mass-radius relation of NSs by their tidal deformations. The accumulation of GW events in the future will allow us to probe the accuracy of General Relativity (GR) in the strong gravitational regime Berti ; Barack . In particular, whether or not some extra degrees of freedom are present around compact objects are a great concern, along with the problem of dark energy and dark matter.

The simplest candidate for such an extra degree of freedom is a scalar field ϕ\phi. Theories in which the scalar field is coupled to the gravitational sector are called scalar-tensor theories Fujii . The nonminimal coupling F(ϕ)RF(\phi)R, where FF is a function of ϕ\phi and RR is the Ricci scalar, is a typical example of the direct interaction between the scalar field and gravity. In the presence of baryonic fluids, the matter sector indirectly feels an interaction with the scalar field through the coupling to gravity. This matter coupling manifests itself after performing a so-called conformal transformation to the Einstein frame in which the Ricci scalar does not have a direct coupling to ϕ\phi DeFelice:2010aj . In scalar-tensor theories, such matter couplings can modify the internal structure of relativistic stars.

In scalar-tensor theories with the nonminimal coupling F(ϕ)RF(\phi)R, it is known that there are some NS solutions with scalar hairs on a spherically symmetric and static background. For the theories in which the coupling F(ϕ)F(\phi) contains an even power-law function ϕn\phi^{n}, there exists a field profile of nonvanishing ϕ\phi, while satisfying F,ϕ(0)=0F_{,\phi}(0)=0 as in GR (where F,ϕ=dF/dϕF_{,\phi}={\rm d}F/{\rm d}\phi) Damour ; Damour2 . If the second derivative F,ϕϕF_{,\phi\phi} is positive at ϕ=0\phi=0, then the GR branch (ϕ=0\phi=0 everywhere) can be unstable due to a negative mass squared proportional to F,ϕϕ(0)-F_{,\phi\phi}(0). This allows a possibility for triggering tachyonic growth of the scalar field toward a scalarized branch with nonvanishing ϕ\phi, whose phenomenon is dubbed spontaneous scalarization.

A typical example of the nonminimal coupling triggering spontaneous scalarization is of the form F(ϕ)=eβϕ2/(2Mpl2)F(\phi)=e^{-\beta\phi^{2}/(2M_{\rm pl}^{2})} Damour ; Damour2 , where β\beta is a constant and MplM_{\rm pl} is the reduced Planck mass. Provided that β<0\beta<0, the two conditions F,ϕ(0)=0F_{,\phi}(0)=0 and F,ϕϕ(0)>0F_{,\phi\phi}(0)>0 are satisfied. More precisely, spontaneous scalarization can occur for the coupling β<4.35\beta<-4.35 Harada:1998ge ; Novak:1998rk ; Silva:2014fca ; Freire:2012mg , whose upper bound is insensitive to the change of equations of state inside the Ns. The properties of hairy solutions and its observational consequences were investigated in several contexts, e.g., GW asteroseismology Sotani1 ; Sotani2 , rotating NSs Sotani:2012eb ; Doneva:2013qva ; Doneva:2014faa ; Pani:2014jra , and the influence on particle geodesics around NSs Doneva:2014uma ; Pappas:2015npa . Recent studies showed that black holes can also exhibit spontaneous scalarization, in the presence of couplings to a Gauss-Bonnet term Kleihaus:2015aje ; Doneva:2017bvd ; Silva:2017uqg ; Antoniou:2017acq ; Antoniou:2017hxj ; Minamitsuji:2018xde ; Cunha and to an electromagnetic field Stefanov ; Herdeiro1 ; Herdeiro2 ; Herdeiro3 ; Ikeda .

There are also other nonminimally coupled theories admitting hairy NS solutions, e.g., Brans-Dicke (BD) theories Brans with a scalar potential. The nonminimal coupling in BD theories can be expressed in the form F(ϕ)=e2Qϕ/MplF(\phi)=e^{-2Q\phi/M_{\rm pl}}, where QQ is a constant related to the so-called BD parameter ωBD\omega_{\rm BD} as 2Q2=1/(3+2ωBD)2Q^{2}=1/(3+2\omega_{\rm BD}) Yoko . Since this coupling does not satisfy conditions for the occurrence of spontaneous scalarization, there exist only hairy solutions with nonvanishing scalar-field profiles Cooney:2009rr ; Arapoglu:2010rz ; Orellana:2013gn ; Astashenok:2013vza ; Yazadjiev:2014cza ; Resco:2016upv . For example, f(R)f(R) gravity belongs to a class of BD theories with ωBD=0\omega_{\rm BD}=0 Ohanlon ; Chiba03 , in the presence of a scalar potential arising from the deviation from GR. For the Starobinsky model f(R)=R+R2/(6m2)f(R)=R+R^{2}/(6m^{2}) Staro , the existence of a positive constant mass mm gives rise to an exponential growing mode of ϕ\phi outside the star Ganguly ; Kase:2019dqc . This is not the case for the models f(R)=R+aRpf(R)=R+aR^{p} with 1<p<21<p<2 Kase:2019dqc ; Dohi:2020bfs , in which the effective mass of ϕ\phi can approach 0 toward spatial infinity. In the latter case, there are hairy NS solutions with the mass-radius relation modified from that in GR.

The hairy NS solutions in theories of spontaneous scalarization and BD theories have an additional pressure induced by a matter coupling with the scalar field. Under the occurrence of spontaneous scalarization, for example, the additional pressure works to be repulsive against gravity, in which case the radius and mass of star tend to be increased relative to those in GR (see Refs. Maselli ; Minami ; Chagoya ; Kase:2017egk ; Chagoya2 ; Ogawa ; Kase:2020yhw for related papers). Naively, one might expect that such hairy solutions can be unstable against perturbations. To study the stability of relativistic stars with scalar hairs, it is necessary to properly incorporate perturbations of the matter sector besides those of gravity and the scalar field. In this paper, we will address this issue by dealing with baryonic matter as a perfect fluid.

For a k-essence scalar field with the Lagrangian P(X)P(X) Scherrer , where PP is a function of the field kinetic energy X=μϕμϕ/2X=-\partial_{\mu}\phi\partial^{\mu}\phi/2, it is known that the corresponding energy-momentum tensor reduces to that of a perfect fluid for a time-like scalar field, i.e., X>0X>0 Hu05 ; Arroja . This is the case for a time-dependent cosmological background, so that the k-essence Lagrangian was extensively used to describe the perfect-fluid dynamics especially in the context of late-time cosmic acceleration DMT10 ; KT14b ; Heisenberg:2016eld ; Gum . On the spherically symmetric and static background, however, the scalar field is space-like (X<0X<0) and hence the perfect fluid cannot be described by the k-essence Lagrangian. Instead, we will employ the matter action advocated by Schutz and Sorkin Sorkin , which allows one to describe the perfect fluid in any curved background (see also Refs. Brown ; DGS ) .

To accommodate both theories of spontaneous scalarization and BD theories presented in Sec. II, we will consider scalar-tensor theories given by the Lagrangian =G4(ϕ)R+G2(ϕ,X){\cal L}=G_{4}(\phi)R+G_{2}(\phi,X) in the presence of a perfect fluid. This belongs to a subclass of Horndeski theories with second-order field equations of motion Horndeski ; Horn1 ; Horn2 ; Horn3 . We carry out all the analysis in the Jordan frame, in which the matter sector is minimally coupled to gravity. After deriving covariant equations of motion and applying them to the spherically symmetric and static background in Sec. III, we proceed to the discussion about the separation of perturbations into odd- and even-parity modes in Sec. IV. In Horndeski theories without matter, the stabilities of hairy black hole solutions against odd- and even-parity perturbations were studied in Refs. DeFelice:2011ka ; Kobayashi:2012kh ; Kobayashi:2014wsa ; Kobayashi:2014eva ; Ogawa:2015pea ; Babichev:2016rlq ; Khoury:2020aya .

In Sec. V, we expand the full action up to second order in odd-parity perturbations with the multipoles l2l\geq 2 and show that the ghost is absent for G4>0G_{4}>0 with the propagation speed of gravity equivalent to that of light. In Sec. VI, we derive conditions for the absence of ghosts and Laplacian instabilities in the even-parity sector. The propagating degrees of freedom are different depending on the multipoles ll, so we separately discuss the three cases l2l\geq 2, l=1l=1, and l=0l=0. The analysis of odd- and even-parity perturbations was also performed in Ref. Sotani:2005qx for the nonminimal coupling of Refs. Damour ; Damour2 by using an energy-momentum tensor of the perfect fluid. Since this approach is based on the equations of motion rather than the action principle, it is not straightforward to identify the dynamical degrees of freedom and their stability conditions. In this paper, we address this problem by dealing with the perfect fluid as the Schutz-Sorkin action and integrate out all the nondynamical perturbations from the action. Indeed, the dynamical degree of freedom in the matter sector is of a nontrivial form, which affects the propagation of scalar GWs.

In Sec. VII, we apply our general stability conditions to hairy relativistic stars present in theories of spontaneous scalarization and BD theories. We show that these hairy solutions are stable under the conditions G4>0G_{4}>0, ρ+P>0\rho+P>0, and cm2>0c_{m}^{2}>0, where ρ\rho, PP, and cm2c_{m}^{2} are the density, pressure and sound speed squared of matter respectively. In particular, as long as 0<cm210<c_{m}^{2}\leq 1, all the speeds of propagation associated with even-parity perturbations in the gravity sector are subluminal inside the star, while they are equivalent to the speed of light outside the star. This fact is consistent with the speed of gravity constrained from the GW170817 event GW170817 .

Throughout the paper, we adopt the natural units for which the speed of light cc, the reduced Planck constant \hbar, and the Boltzmann constant kBk_{B} are set to unity.

II Theories of scalarized relativistic stars

In this section, we briefly review several theories of relativistic stars with scalar hairs. We consider a scalar field ϕ\phi with a nonminimal coupling of the form F(ϕ)RF(\phi)R. The scalar field can have a kinetic term of the form ω(ϕ)X\omega(\phi)X, where ω\omega is a function of ϕ\phi and X=(1/2)gμνμϕνϕX=-(1/2)g^{\mu\nu}\partial_{\mu}\phi\partial_{\nu}\phi is the kinetic energy (with metric tensor gμνg^{\mu\nu}). We also allow for the existence of a field potential V(ϕ)V(\phi). Then, the action of such scalar-tensor theories is given by

𝒮=d4xg[Mpl22F(ϕ)R+ω(ϕ)XV(ϕ)]+𝒮m(gμν,Ψm),{\cal S}=\int{\rm d}^{4}x\sqrt{-g}\left[\frac{M_{\rm pl}^{2}}{2}F(\phi)R+\omega(\phi)X-V(\phi)\right]+{\cal S}_{m}(g_{\mu\nu},\Psi_{m})\,, (1)

where gg is a determinant of metric tensor gμνg_{\mu\nu}, and MplM_{\rm pl} is the reduced Planck mass. The action 𝒮m{\cal S}_{m} corresponds to that of the matter field Ψm\Psi_{m}. We assume that the matter field is minimally coupled to gravity in the Jordan frame given by the metric gμνg_{\mu\nu}. In Sec. III, we specify the action 𝒮m{\cal S}_{m} to be of the form of a perfect fluid.

Performing a so-called conformal transformation (gμν)E=F(ϕ)gμν(g_{\mu\nu})_{E}=F(\phi)g_{\mu\nu} of the metric, we obtain the following Einstein-frame action DeFelice:2010aj ,

𝒮=d4xgE[Mpl22RE12(gμν)EμφνφVE(φ)]+𝒮m(F1(ϕ)(gμν)E,Ψm),{\cal S}=\int{\rm d}^{4}x\sqrt{-g_{E}}\left[\frac{M_{\rm pl}^{2}}{2}R_{E}-\frac{1}{2}(g^{\mu\nu})_{E}\partial_{\mu}\varphi\partial_{\nu}\varphi-V_{E}(\varphi)\right]+{\cal S}_{m}\left(F^{-1}(\phi)(g_{\mu\nu})_{E},\Psi_{m}\right)\,, (2)

where the subscript “EE” represents quantities in the Einstein frame, and

dφdϕ=32(MplF,ϕF)2+ωF,VE(φ)=VF2.\frac{{\rm d}\varphi}{{\rm d}\phi}=\sqrt{\frac{3}{2}\left(\frac{M_{\rm pl}F_{,\phi}}{F}\right)^{2}+\frac{\omega}{F}}\,,\qquad V_{E}(\varphi)=\frac{V}{F^{2}}\,. (3)

In the Einstein frame the field φ\varphi does not have a direct coupling with the Ricci scalar RER_{E}, but the matter sector has an interaction with the scalar field through the metric (gμν)E=F(ϕ)gμν(g_{\mu\nu})_{E}=F(\phi)g_{\mu\nu}.

In what follows, we will present two classes of theories which belong to the action (1).

II.1 Theories of spontaneous scalarization

When the nonminimal coupling F(ϕ)F(\phi) is present, spontaneous scalarization can occur inside relativistic stars for a massless scalar field ϕ\phi. The canonical scalar field φ\varphi in the Einstein frame can be chosen to be equivalent to ϕ\phi. Since dφ/dϕ=1{\rm d}\varphi/{\rm d}\phi=1 in this case, it follows that ω=[13Mpl2F,ϕ2/(2F2)]F\omega=[1-3M_{\rm pl}^{2}F_{,\phi}^{2}/(2F^{2})]F. In the absence of the potential V(ϕ)V(\phi), the Jordan-frame action (1) is expressed in the form,

𝒮=d4xg[Mpl22F(ϕ)R+(13Mpl2F,ϕ22F2)F(ϕ)X]+𝒮m(gμν,Ψm).{\cal S}=\int{\rm d}^{4}x\sqrt{-g}\left[\frac{M_{\rm pl}^{2}}{2}F(\phi)R+\left(1-\frac{3M_{\rm pl}^{2}F_{,\phi}^{2}}{2F^{2}}\right)F(\phi)X\right]+{\cal S}_{m}(g_{\mu\nu},\Psi_{m})\,. (4)

For the theories (4) with F,ϕ(0)=0F_{,\phi}(0)=0, there is a GR branch of spherically symmetric and static NS solutions characterized by ϕ=0\phi=0 everywhere. About the NS solution in GR, the effective mass squared for small perturbations is given by meff2=(Mpl2/2)[F,ϕϕ(0)/ω(0)]Rm_{\rm eff}^{2}=-(M_{\rm pl}^{2}/2)[F_{,\phi\phi}(0)/\omega(0)]R. As long as the conditions F,ϕϕ(0)>0F_{,\phi\phi}(0)>0 and ω(0)>0\omega(0)>0 are satisfied with a positive RR, there is a tachyonic instability of the GR branch. Then, the spontaneous growth of ϕ\phi can occur toward the other nontrivial branch with ϕ0\phi\neq 0. The conditions for the occurrence of spontaneous scalarization correspond to F,ϕ(0)=0F_{,\phi}(0)=0, F,ϕϕ(0)>0F_{,\phi\phi}(0)>0, and ω(0)>0\omega(0)>0, so it is necessary to have an even power-law dependence of ϕ\phi for the coupling F(ϕ)F(\phi).

The nonminimal coupling chosen by Damour and Esposito-Farese Damour ; Damour2 is given by

F(ϕ)=eβϕ2/(2Mpl2),F(\phi)=e^{-\beta\phi^{2}/(2M_{\rm pl}^{2})}\,, (5)

where β\beta is a constant. In this case, the function ω(ϕ)\omega(\phi) in front of the kinetic term XX in Eq. (4) reads

ω(ϕ)=(13β2ϕ22Mpl2)eβϕ2/(2Mpl2).\omega(\phi)=\left(1-\frac{3\beta^{2}\phi^{2}}{2M_{\rm pl}^{2}}\right)e^{-\beta\phi^{2}/(2M_{\rm pl}^{2})}\,. (6)

Then, the conditions F,ϕ(0)=0F_{,\phi}(0)=0, F,ϕϕ(0)>0F_{,\phi\phi}(0)>0, and ω(0)>0\omega(0)>0 are satisfied for β<0\beta<0. Hence, for negative β\beta, the NS can have a nontrivial branch with a modified internal structure by the presence of nonminimal coupling with the scalar field.

A common procedure for the analysis of spontaneous scalarization is to study the solutions in the Einstein frame first and then transform back to the Jordan frame to compute physical quantities such as the mass and radius of relativistic stars Damour ; Damour2 ; Harada:1998ge ; Novak:1998rk ; Silva:2014fca ; Freire:2012mg ; Chen ; Morisaki . However, all the analysis can be performed in the Jordan-frame action (4) without any reference to the Einstein frame. In the Jordan frame the matter sector is minimally coupled to gravity, so it is also straightforward to incorporate it as a perfect fluid described by a Schutz-Sokin action (see Sec. III).

II.2 Brans-Dicke theories

The action in BD theories Brans with a scalar potential V(ϕ)V(\phi) can be expressed in the form Yoko ,

𝒮=d4xg[Mpl22F(ϕ)R+(16Q2)F(ϕ)XV(ϕ)]+𝒮m(gμν,Ψm),{\cal S}=\int{\rm d}^{4}x\sqrt{-g}\left[\frac{M_{\rm pl}^{2}}{2}F(\phi)R+\left(1-6Q^{2}\right)F(\phi)X-V(\phi)\right]+{\cal S}_{m}(g_{\mu\nu},\Psi_{m})\,, (7)

where the nonminimal coupling is given by

F(ϕ)=e2Qϕ/Mpl.F(\phi)=e^{-2Q\phi/M_{\rm pl}}\,. (8)

The constant QQ characterizes the coupling strength between the scalar field ϕ\phi and gravity, which is related to the BD parameter ωBD\omega_{\rm BD} as

2Q2=13+2ωBD.2Q^{2}=\frac{1}{3+2\omega_{\rm BD}}\,. (9)

The action (7) belongs to a sub-class of scalar-tensor theories (1). The minimally coupled scalar field in GR corresponds to the limit Q0Q\to 0, i.e., ωBD\omega_{\rm BD}\to\infty. In the absence of matter the ghost is absent for ωBD>3/2\omega_{\rm BD}>-3/2 Fujii , which is consistent with the positivity on the left-hand-side of Eq. (9).

The metric f(R)f(R) gravity is accommodated by the action (7) with the correspondence DeFelice:2010aj ,

Q=16,V(ϕ)=Mpl22(FRf),F=fR=e2Qϕ/Mpl.Q=-\frac{1}{\sqrt{6}}\,,\qquad V(\phi)=\frac{M_{\rm pl}^{2}}{2}\left(FR-f\right)\,,\qquad F=\frac{\partial f}{\partial R}=e^{-2Q\phi/M_{\rm pl}}\,. (10)

In this case, the action (7) reduces to 𝒮=d4xg(Mpl2/2)f(R)+𝒮m(gμν,Ψm){\cal S}=\int{\rm d}^{4}x\sqrt{-g}\,(M_{\rm pl}^{2}/2)f(R)+{\cal S}_{m}(g_{\mu\nu},\Psi_{m}), with the BD parameter ωBD=0\omega_{\rm BD}=0 Ohanlon ; Chiba03 . Provided that f(R)f(R) contains nonlinear functions of RR, the gravitational sector propagates one scalar degree of freedom ϕ\phi. This scalar field, which is related to RR through the last relation of Eq. (10), has a potential V(ϕ)V(\phi) of the gravitational origin.

If the potential has a constant mass mm like the Starobinsky model f(R)=R+R2/(6m2)f(R)=R+R^{2}/(6m^{2}), it is difficult to realize a stable field profile satisfying the boundary condition ϕ0\phi\to 0 at spatial infinity due to the existence of an exponentially growing mode outside the star Ganguly ; Kase:2019dqc . If the effective mass of ϕ\phi approaches 0 toward spatial infinity, there exist regular NS solutions without the exponential growth of ϕ\phi. An explicit example of the latter is the model f(R)=R+aRpf(R)=R+aR^{p}, where aa and pp are constants in the ranges a>0a>0 and 1<p<21<p<2 Kase:2019dqc ; Dohi:2020bfs . In this case, the scalar potential is approximately given by V(ϕ)ϕp/(p1)V(\phi)\propto\phi^{p/(p-1)}. This includes the self-coupling potential V(ϕ)ϕ4V(\phi)\propto\phi^{4} (for p=4/3p=4/3).

III Scalar-tensor theories with matter

In this paper, we focus on scalar-tensor theories given by the action,

𝒮=d4xg[G4(ϕ)R+G2(ϕ,X)]+𝒮m(gμν,Ψm),{\cal S}=\int{\rm d}^{4}x\sqrt{-g}\left[G_{4}(\phi)R+G_{2}(\phi,X)\right]+{\cal S}_{m}(g_{\mu\nu},\Psi_{m})\,, (11)

where G4G_{4} is a function of the scalar field ϕ\phi, and G2G_{2} depends on both ϕ\phi and XX. The action (11) accommodates scalar-tensor theories with Eq. (1) as a special case. A perfect fluid minimally coupled to gravity can be described by a Schutz-Sorkin action of the form Sorkin ; Brown ; DGS

𝒮m=d4x[gρ(n)+Jμ(μ+𝒜iμi)].{\cal S}_{m}=-\int{\rm d}^{4}x\left[\sqrt{-g}\,\rho(n)+J^{\mu}(\partial_{\mu}\ell+{\cal A}_{i}\partial_{\mu}{\cal B}^{i})\right]\,. (12)

The matter density ρ\rho depends on the fluid number density nn alone. The vector field JμJ^{\mu} corresponds to a current, while the scalar quantity \ell is a Lagrange multiplier. In terms of JμJ^{\mu}, the number density can be expressed as

n=gμνJμJνg.n=\sqrt{\frac{g_{\mu\nu}J^{\mu}J^{\nu}}{g}}\,. (13)

The fluid four-velocity uμu_{\mu} is related to JμJ_{\mu}, as

uμ=Jμng.u_{\mu}=\frac{J_{\mu}}{n\sqrt{-g}}\,. (14)

From Eq. (13), there is the relation uμuμ=1u^{\mu}u_{\mu}=-1. The quantities 𝒜i{\cal A}_{i} and i{\cal B}^{i} are spatial vectors characterizing intrinsic vector modes.

III.1 Covariant equations of motion

Varying the action (11) with respect to \ell and JμJ^{\mu} respectively, we obtain

μJμ\displaystyle\partial_{\mu}J^{\mu} =\displaystyle= 0,\displaystyle 0\,, (15)
μ\displaystyle\partial_{\mu}\ell =\displaystyle= ρ,nuμ𝒜iμi,\displaystyle\rho_{,n}u_{\mu}-{\cal A}_{i}\partial_{\mu}{\cal B}^{i}\,, (16)

where we used the property n/Jμ=uμ/g\partial n/\partial J^{\mu}=-u_{\mu}/\sqrt{-g}. Here and in the following, we use the notation ρ,n=ρ/n\rho_{,n}=\partial\rho/\partial n. Variations of the action (12) with respect to 𝒜i{\cal A}_{i} and i{\cal B}^{i} lead, respectively, to

Jμμi=0,\displaystyle J^{\mu}\partial_{\mu}{\cal B}^{i}=0\,, (17)
Jμμ𝒜i=0,\displaystyle J^{\mu}\partial_{\mu}{\cal A}_{i}=0\,, (18)

where we used Eq. (15). Taking note of the relations Jμ=nguμJ^{\mu}=n\sqrt{-g}\,u^{\mu} and μ(guμ)=gμuμ\partial_{\mu}(\sqrt{-g}\,u^{\mu})=\sqrt{-g}\,\nabla_{\mu}u^{\mu}, where μ\nabla_{\mu} is the covariant derivative operator, the current conservation (15) can be expressed in the form,

uμμρ+(ρ+P)μuμ=0,u^{\mu}\nabla_{\mu}\rho+(\rho+P)\nabla_{\mu}u^{\mu}=0\,, (19)

where PP is the matter pressure defined by

Pnρ,nρ.P\equiv n\rho_{,n}-\rho\,. (20)

Variation of the matter Lagrangian Lm=[gρ(n)+Jμ(μ+𝒜iμi)]L_{m}=-[\sqrt{-g}\,\rho(n)+J^{\mu}(\partial_{\mu}\ell+{\cal A}_{i}\partial_{\mu}{\cal B}^{i})] with respect to gμνg^{\mu\nu} gives

δLm=δgρ(n)gρ,nδnJμ(ν+𝒜iνi)δgμν.\delta L_{m}=-\delta\sqrt{-g}\rho(n)-\sqrt{-g}\rho_{,n}\delta n-J_{\mu}\left(\partial_{\nu}\ell+{\cal A}_{i}\partial_{\nu}{\cal B}^{i}\right)\delta g^{\mu\nu}\,. (21)

By exploiting Eq. (16) as well as the relations,

δg=12ggμνδgμν,δn=n2(gμνuμuν)δgμν,\delta\sqrt{-g}=-\frac{1}{2}\sqrt{-g}\,g_{\mu\nu}\delta g^{\mu\nu}\,,\qquad\delta n=\frac{n}{2}\left(g_{\mu\nu}-u_{\mu}u_{\nu}\right)\delta g^{\mu\nu}\,, (22)

it follows that

2gδLmδgμν=(ρ+P)uμuν+PgμνTμν,-\frac{2}{\sqrt{-g}}\frac{\delta L_{m}}{\delta g^{\mu\nu}}=\left(\rho+P\right)u_{\mu}u_{\nu}+Pg_{\mu\nu}\equiv T_{\mu\nu}\,, (23)

where TμνT_{\mu\nu} corresponds to an energy-momentum tensor of the perfect fluid.

Varying the total action (11) with respect to gμνg^{\mu\nu}, the resulting gravitational equations of motion are given by

2G4GμνG2gμνG2,Xμϕνϕ2G4,ϕ(μνϕgμνϕ)2G4,ϕϕ(μϕνϕ+2Xgμν)=Tμν,2G_{4}G_{\mu\nu}-G_{2}g_{\mu\nu}-G_{2,X}\nabla_{\mu}\phi\nabla_{\nu}\phi-2G_{4,\phi}\left(\nabla_{\mu}\nabla_{\nu}\phi-g_{\mu\nu}\Box\phi\right)-2G_{4,\phi\phi}\left(\nabla_{\mu}\phi\nabla_{\nu}\phi+2Xg_{\mu\nu}\right)=T_{\mu\nu}\,, (24)

where =gμνμν\square=g^{\mu\nu}\nabla_{\mu}\nabla_{\nu}. Since the perfect fluid is minimally coupled to gravity, the corresponding energy-momentum tensor is divergence-free, i.e.,

μTμν=0.\nabla^{\mu}T_{\mu\nu}=0\,. (25)

This property is consistent with the left-hand-side of Eq. (24). Multiplying μν\mu^{\nu} for Eq. (25), we obtain

uνμTμν=[uμμρ+(ρ+P)μuμ]=0,u^{\nu}\nabla^{\mu}T_{\mu\nu}=-\left[u_{\mu}\nabla^{\mu}\rho+(\rho+P)\nabla^{\mu}u_{\mu}\right]=0\,, (26)

which also follows from Eq. (19). Since Eq. (19) arises from Eq. (15), Eq. (26) corresponds to the current conservation.

Let us also introduce a unit vector nνn^{\nu} orthogonal to uνu_{\nu}, such that

nνuν=0,nνnν=1.n^{\nu}u_{\nu}=0\,,\qquad n^{\nu}n_{\nu}=1\,. (27)

Multiplying nνn^{\nu} for Eq. (25), we find

nνμTμν=(ρ+P)nνuμμuν+nμμP=0,n^{\nu}\nabla^{\mu}T_{\mu\nu}=(\rho+P)n^{\nu}u_{\mu}\nabla^{\mu}u_{\nu}+n_{\mu}\nabla^{\mu}P=0\,, (28)

and hence

nμμP=(ρ+P)nνuμμuν.n_{\mu}\nabla^{\mu}P=-(\rho+P)n^{\nu}u_{\mu}\nabla^{\mu}u_{\nu}\,. (29)

Inside a compact object, this can be interpreted as a balance between the pressure and gravity.

III.2 Background equations

Let us consider a spherically symmetric and static background given by the line element,

ds2=f(r)dt2+h(r)1dr2+r2(dθ2+sin2θdφ2),{\rm d}s^{2}=-f(r){\rm d}t^{2}+h(r)^{-1}{\rm d}r^{2}+r^{2}\left({\rm d}\theta^{2}+\sin^{2}\theta\,{\rm d}\varphi^{2}\right)\,, (30)

where f(r)f(r) and h(r)h(r) are functions of rr. For the matter sector, we take the following configuration,

Jμ=(gN(r),0,0,0),𝒜i=0,J^{\mu}=\left(\sqrt{-g}\,N(r),0,0,0\right)\,,\qquad{\cal A}_{i}=0\,, (31)

where g=f1/2h1/2r2sinθ\sqrt{-g}=f^{1/2}h^{-1/2}r^{2}\sin{\theta} and NN is a function of the radial coordinate rr. We note that JμJ^{\mu} is not a four vector since the second term on the right-hand-side of Eq. (12) is not multiplied by the volume factor g\sqrt{-g}. If we alternatively define J~μ=Jμ/g\tilde{J}^{\mu}=J^{\mu}/\sqrt{-g}, this can be regarded as a four vector whose component depends on rr alone, i.e., J~μ=(N(r),0,0,0)\tilde{J}^{\mu}=(N(r),0,0,0). This is the reason why JμJ^{\mu} contains the θ\theta-dependent term arising from g\sqrt{-g}.

Assuming that N(r)N(r) is positive, the number density (13) reads

n(r)=f(r)1/2N(r),n(r)=f(r)^{1/2}N(r)\,, (32)

which depends on rr. Substituting Eqs. (31) and (32) into Eq. (14), it follows that

uμ=(f(r)1/2,0,0,0).u^{\mu}=\left(f(r)^{-1/2},0,0,0\right)\,. (33)

From the definition (23) of the fluid energy-momentum tensor, we have

Tνμ=diag(ρ(r),P(r),P(r),P(r)).T^{\mu}_{\nu}={\rm diag}\left(-\rho(r),P(r),P(r),P(r)\right)\,. (34)

From the tttt, rrrr, θθ\theta\theta components Eq. (24), we obtain

(2G4r+G4,ϕϕ)h+2(G4,ϕϕ′′+2G4,ϕϕr+G4,ϕϕϕ2)h+2G4(h1)r2G2=ρ,\displaystyle\left(\frac{2G_{4}}{r}+G_{4,\phi}\phi^{\prime}\right)h^{\prime}+2\left(G_{4,\phi}\phi^{\prime\prime}+\frac{2G_{4,\phi}\phi^{\prime}}{r}+G_{4,\phi\phi}\phi^{\prime 2}\right)h+\frac{2G_{4}(h-1)}{r^{2}}-G_{2}=-\rho\,, (35)
(2G4r+G4,ϕϕ)hff+4G4,ϕϕhr+2G4(h1)r2G2G2,Xϕ2h=P,\displaystyle\left(\frac{2G_{4}}{r}+G_{4,\phi}\phi^{\prime}\right)h\frac{f^{\prime}}{f}+\frac{4G_{4,\phi}\phi^{\prime}h}{r}+\frac{2G_{4}(h-1)}{r^{2}}-G_{2}-G_{2,X}\phi^{\prime 2}h=P\,, (36)
G4h2f2(2ff′′f2)+(12G4h+G4hr+G4,ϕϕh)ff+(G4r+G4,ϕϕ)h\displaystyle\frac{G_{4}h}{2f^{2}}\left(2ff^{\prime\prime}-f^{\prime 2}\right)+\left(\frac{1}{2}G_{4}h^{\prime}+\frac{G_{4}h}{r}+G_{4,\phi}\phi^{\prime}h\right)\frac{f^{\prime}}{f}+\left(\frac{G_{4}}{r}+G_{4,\phi}\phi^{\prime}\right)h^{\prime}
+2(G4,ϕϕϕ2+G4,ϕϕ′′+G4,ϕϕr)hG2=P,\displaystyle+2\left(G_{4,\phi\phi}\phi^{\prime 2}+G_{4,\phi}\phi^{\prime\prime}+\frac{G_{4,\phi}\phi^{\prime}}{r}\right)h-G_{2}=P\,, (37)

where a prime represents the derivative with respect to rr. The φφ\varphi\varphi component of Eq. (24) gives the same equation as (37). We note that Eq. (26) is trivially satisfied on the background (30). For the unit vector nνn_{\nu} obeying the property (27), we can choose nμ=(0,h1/2,0,0)n_{\mu}=(0,h^{-1/2},0,0). Then, Eq. (29) reduces to

P+f2f(ρ+P)=0.P^{\prime}+\frac{f^{\prime}}{2f}\left(\rho+P\right)=0\,. (38)

The equation of motion for the scalar field follows by varying the action (11) with respect to ϕ\phi. This amounts to substituting the rr derivative of Eq. (36) as well as Eqs. (35) and (36) into Eq. (38), with the elimination of PP from Eqs. (36) and (37) to solve for G4G_{4}. Then, the scalar field obeys the following equation,

ϕ\displaystyle{\cal E}_{\phi} \displaystyle\equiv G2,ϕ+G2,Xh[ϕ′′+h2hϕ+(2r+f2f)ϕ]+G2,ϕXhϕ2G2,XX(ϕ′′+h2hϕ)h2ϕ2\displaystyle G_{2,\phi}+G_{2,X}h\left[\phi^{\prime\prime}+{h^{\prime}\over 2h}\phi^{\prime}+\left({2\over r}+{f^{\prime}\over 2f}\right)\phi^{\prime}\right]+G_{2,\phi X}h\phi^{\prime 2}-G_{2,XX}\left(\phi^{\prime\prime}+{h^{\prime}\over 2h}\phi^{\prime}\right)h^{2}\phi^{\prime 2} (39)
+G4,ϕ[2r2(1hrh)(h+4hr)f2f+f22f2hf′′fh]=0.\displaystyle+G_{4,\phi}\left[{2\over r^{2}}(1-h-rh^{\prime})-\left(h^{\prime}+{4h\over r}\right){f^{\prime}\over 2f}+{f^{\prime 2}\over 2f^{2}}h-{f^{\prime\prime}\over f}h\right]=0\,.

If the equation of state,

P=P(ρ),P=P(\rho)\,, (40)

is given inside a star, we can integrate Eqs. (35)-(38) with Eq. (40) to solve for ff, hh, ϕ\phi, PP, and ρ\rho. The integration is performed up to the radius of star (characterized by P=0P=0) with the regular boundary conditions f=h=ϕ=P=ρ=0f^{\prime}=h^{\prime}=\phi^{\prime}=P^{\prime}=\rho^{\prime}=0 at r=0r=0.

IV Perturbations on the spherically symmetric and static background

To study the stability of relativistic stars around the spherically symmetric and static space-time (30), we consider metric perturbations hμνh_{\mu\nu} on the background metric g¯μν\bar{g}_{\mu\nu}, such that

gμν=g¯μν+hμν.g_{\mu\nu}=\bar{g}_{\mu\nu}+h_{\mu\nu}\,. (41)

There are also perturbations of the scalar field ϕ\phi and quantities appearing in the Schutz-Sorkin action (12). In terms of the spherical harmonics Ylm(θ,φ)Y_{lm}(\theta,\varphi), the scalar field can be expressed in the form,

ϕ=ϕ¯(r)+l,mδϕlm(t,r)Ylm(θ,φ),\phi=\bar{\phi}(r)+\sum_{l,m}\delta\phi_{lm}(t,r)Y_{lm}(\theta,\varphi)\,, (42)

where ϕ¯(r)\bar{\phi}(r) is the background value, and δϕlm\delta\phi_{lm} is the perturbed part being a function of tt and rr. In the following, we omit the subscripts l,ml,m from the perturbation δϕlm(t,r)\delta\phi_{lm}(t,r) and also apply the same rule to other perturbed quantities appearing below. Under the rotation in two-dimensional plane (θ,φ\theta,\varphi), any scalar perturbation has the parity ()l(-)^{l}, which is called the even mode Regge:1957td ; Zerilli:1970se . The odd mode corresponds to a perturbation with the parity ()l+1(-)^{l+1}. The metric components htt,htr,hrrh_{tt},h_{tr},h_{rr} transform as scalars under a two-dimensional rotation in the (θ,φ\theta,\varphi) plane, so they only possess even-parity modes.

The θ\theta and φ\varphi components of any vector field VμV_{\mu} contain both even- and odd-modes, whereas the temporal and radial components VtV_{t} and VrV_{r} possess the even mode alone. To accommodate the odd-parity contribution to VμV_{\mu}, we introduce the tensor Eab=γεabE_{ab}=\sqrt{\gamma}\,\varepsilon_{ab}, where γ\gamma is the determinant of two dimensional metric γab\gamma_{ab} (with a,ba,b either θ\theta or φ\varphi), and εab\varepsilon_{ab} is the anti-symmetric symbol with εθφ=1\varepsilon_{\theta\varphi}=1. The θ,φ\theta,\varphi components of fluid four velocity uμu_{\mu}, for example, can be expressed as

ua\displaystyle u_{a} =\displaystyle= av+Eabbv~=l,mv(t,r)aYlm(θ,φ)+l,mv~(t,r)EabbYlm(θ,φ),\displaystyle\nabla_{a}v+E_{ab}\nabla^{b}\tilde{v}=\sum_{l,m}v(t,r)\nabla_{a}Y_{lm}(\theta,\varphi)+\sum_{l,m}\tilde{v}(t,r)E_{ab}\nabla^{b}Y_{lm}(\theta,\varphi)\,, (43)

where, in the second line, we expressed two scalars vv and v~\tilde{v} in terms of the expansion of spherical harmonics. The first and second terms of Eq. (43) correspond to even- and odd-parity perturbations, respectively. The metric components htθ,htφ,hrθ,hrφh_{t\theta},h_{t\varphi},h_{r\theta},h_{r\varphi} transform as vectors under the two-dimensional rotation in the (θ,φ\theta,\varphi) plane, so they can be also expressed in terms of the sum of even- and odd-modes analogous to Eq. (43).

The components TabT_{ab} of any symmetric tensor TμνT_{\mu\nu} contain both even- and odd-modes. For instance, the metric components habh_{ab} are written in the form,

hab\displaystyle h_{ab} =\displaystyle= Kgab+abG+12(EaccbU+EbccaU)\displaystyle Kg_{ab}+\nabla_{a}\nabla_{b}G+\frac{1}{2}\left({E_{a}}^{c}\nabla_{c}\nabla_{b}U+{E_{b}}^{c}\nabla_{c}\nabla_{a}U\right) (44)
=\displaystyle= l,m[K(t,r)gabYlm+G(t,r)abYlm]+12l,mU(t,r)[EaccbYlm+EbccaYlm],\displaystyle\sum_{l,m}\left[K(t,r)g_{ab}Y_{lm}+G(t,r)\nabla_{a}\nabla_{b}Y_{lm}\right]+\frac{1}{2}\sum_{l,m}U(t,r)\left[E_{a}{}^{c}\nabla_{c}\nabla_{b}Y_{lm}+E_{b}{}^{c}\nabla_{c}\nabla_{a}Y_{lm}\right]\,,

where, in the second line, we expressed the three scalars KK, GG, and UU in terms of the expansion of spherical harmonics.

In summary, metric perturbations hμνh_{\mu\nu} and perturbed quantities present in the action (11) can be decomposed into even- and odd-modes in the following way.

IV.1 Odd-parity perturbations

The components of odd-mode metric perturbations are written as

htt=htr=hrr=0,\displaystyle h_{tt}=h_{tr}=h_{rr}=0\,, (45)
hta=l,mQ(t,r)EabbYlm(θ,φ),hra=l,mW(t,r)EabbYlm(θ,φ),\displaystyle h_{ta}=\sum_{l,m}Q(t,r)E_{ab}\nabla^{b}Y_{lm}(\theta,\varphi)\,,\qquad h_{ra}=\sum_{l,m}W(t,r)E_{ab}\nabla^{b}Y_{lm}(\theta,\varphi)\,, (46)
hab=12l,mU(t,r)[EaccbYlm(θ,φ)+EbccaYlm(θ,φ)].\displaystyle h_{ab}=\frac{1}{2}\sum_{l,m}U(t,r)\left[E_{a}{}^{c}\nabla_{c}\nabla_{b}Y_{lm}(\theta,\varphi)+E_{b}{}^{c}\nabla_{c}\nabla_{a}Y_{lm}(\theta,\varphi)\right]\,. (47)

The vector field JμJ_{\mu} in the Schutz-Sorkin action (12) has the following components associated with the odd-parity sector,

Jt=J¯t,Jr=0,Ja=l,mg¯δj(t,r)EabbYlm(θ,φ),J_{t}=\bar{J}_{t}\,,\qquad J_{r}=0\,,\qquad J_{a}=\sum_{l,m}\sqrt{-\bar{g}}\,\delta j(t,r)E_{ab}\nabla^{b}Y_{lm}(\theta,\varphi)\,, (48)

where

J¯t=n(r)f(r)g¯,\bar{J}_{t}=-n(r)\sqrt{f(r)}\sqrt{-\bar{g}}\,, (49)

with the background value g¯=f(r)/h(r)r2sinθ\sqrt{-\bar{g}}=\sqrt{f(r)/h(r)}r^{2}\sin\theta. The intrinsic vectors 𝒜i{\cal A}_{i} and i{\cal B}^{i} are expressed in the form,

𝒜i=δ𝒜i,i=xi+δi,{\cal A}_{i}=\delta{\cal A}_{i}\,,\qquad{\cal B}^{i}=x^{i}+\delta{\cal B}^{i}\,, (50)

with the odd-parity perturbed components,

δ𝒜r=0,δ𝒜a=l,mδ𝒜(t,r)EabbYlm(θ,φ),\displaystyle\delta{\cal A}_{r}=0\,,\qquad\delta{\cal A}_{a}=\sum_{l,m}\delta{\cal A}(t,r)E_{ab}\nabla^{b}Y_{lm}(\theta,\varphi)\,, (51)
δr=0,δa=l,mδ(t,r)EabbYlm(θ,φ),\displaystyle\delta{\cal B}^{r}=0\,,\qquad\delta{\cal B}^{a}=\sum_{l,m}\delta{\cal B}(t,r){E^{a}}_{b}\nabla^{b}Y_{lm}(\theta,\varphi)\,, (52)

where the term xix^{i} in i{\cal B}^{i} is normalized such that the background contribution to ji\partial_{j}{\cal B}^{i} reduces to δji\delta^{i}_{j}.

Let us consider the infinitesimal gauge transformation xμxμ+ξμx_{\mu}\to x_{\mu}+\xi_{\mu}, where

ξt=ξr=0,ξa=l,mΛ(t,r)EabbYlm(θ,φ).\xi_{t}=\xi_{r}=0,\qquad\xi_{a}=\sum_{l,m}\Lambda(t,r)E_{ab}\nabla^{b}Y_{lm}(\theta,\varphi)\,. (53)

Then, the metric perturbations QQ, WW, and UU transform, respectively, as

QQ+Λ˙,WW+Λ2rΛ,UU+2Λ,Q\to Q+\dot{\Lambda}\,,\qquad W\to W+\Lambda^{\prime}-\frac{2}{r}\Lambda\,,\qquad U\to U+2\Lambda\,, (54)

where a dot represents a derivative with respect to tt. For the multipoles l2l\geq 2, we choose the Regge-Wheeler gauge Regge:1957td characterized by

U=0.U=0\,. (55)

For the dipole (l=1l=1), the perturbation habh_{ab} vanishes identically, so we need to handle this case separately.

IV.2 Even-parity perturbations

The components of metric perturbations in the even-parity sector are given by

htt=f(r)l,mH0(t,r)Ylm(θ,φ),htr=hrt=l,mH1(t,r)Ylm(θ,φ),hrr=h(r)1l,mH2(t,r)Ylm(θ,φ),\displaystyle h_{tt}=f(r)\sum_{l,m}H_{0}(t,r)Y_{lm}(\theta,\varphi)\,,\qquad h_{tr}=h_{rt}=\sum_{l,m}H_{1}(t,r)Y_{lm}(\theta,\varphi)\,,\qquad h_{rr}=h(r)^{-1}\,\sum_{l,m}H_{2}(t,r)Y_{lm}(\theta,\varphi)\,,
hta=hat=l,mβ(t,r)aYlm(θ,φ),hra=har=l,mα(t,r)aYlm(θ,φ),\displaystyle h_{ta}=h_{at}=\sum_{l,m}\beta(t,r)\nabla_{a}Y_{lm}(\theta,\varphi)\,,\qquad h_{ra}=h_{ar}=\sum_{l,m}\alpha(t,r)\nabla_{a}Y_{lm}(\theta,\varphi)\,,
hab=l,m[K(t,r)gabYlm(θ,φ)+G(t,r)abYlm(θ,φ)].\displaystyle h_{ab}=\sum_{l,m}\left[K(t,r)g_{ab}Y_{lm}(\theta,\varphi)+G(t,r)\nabla_{a}\nabla_{b}Y_{lm}(\theta,\varphi)\right]\,. (56)

The scalar field ϕ\phi is expressed in the form (42). The components of JμJ_{\mu} containing even-parity perturbations are of the forms,

Jt=J¯t+l,mg¯δJt(t,r)Ylm(θ,φ),Jr=l,mg¯δJr(t,r)Ylm(θ,φ),Ja=l,mg¯δJ(t,r)aYlm(θ,φ).J_{t}=\bar{J}_{t}+\sum_{l,m}\sqrt{-\bar{g}}\,\delta J_{t}(t,r)Y_{lm}(\theta,\varphi)\,,\qquad J_{r}=\sum_{l,m}\sqrt{-\bar{g}}\,\delta J_{r}(t,r)Y_{lm}(\theta,\varphi)\,,\qquad J_{a}=\sum_{l,m}\sqrt{-\bar{g}}\,\delta J(t,r)\nabla_{a}Y_{lm}(\theta,\varphi)\,. (57)

The intrinsic spatial vector fields 𝒜i{\cal A}_{i} and i{\cal B}_{i} are given by Eq. (50) with the perturbed components,

δ𝒜r=l,mδ𝒜1(t,r)Ylm(θ,φ),δ𝒜a=l,mδ𝒜2(t,r)aYlm(θ,φ),\displaystyle\delta{\cal A}_{r}=\sum_{l,m}\delta{\cal A}_{1}(t,r)Y_{lm}(\theta,\varphi)\,,\qquad\delta{\cal A}_{a}=\sum_{l,m}\delta{\cal A}_{2}(t,r)\nabla_{a}Y_{lm}(\theta,\varphi)\,, (58)
δr=l,mδ1(t,r)Ylm(θ,φ),δa=l,mδ2(t,r)aYlm(θ,φ).\displaystyle\delta{\cal B}_{r}=\sum_{l,m}\delta{\cal B}_{1}(t,r)Y_{lm}(\theta,\varphi)\,,\qquad\delta{\cal B}_{a}=\sum_{l,m}\delta{\cal B}_{2}(t,r)\nabla_{a}Y_{lm}(\theta,\varphi)\,. (59)

Let us consider the infinitesimal gauge transformation xμxμ+ξμx_{\mu}\to x_{\mu}+\xi_{\mu}, where

ξt=l,m𝒯(t,r)Ylm(θ,φ),ξr=l,m(t,r)Ylm(θ,φ),ξa=l,mΘ(t,r)aYlm(θ,φ).\xi_{t}=\sum_{l,m}{\cal T}(t,r)Y_{lm}(\theta,\varphi)\,,\qquad\xi_{r}=\sum_{l,m}{\cal R}(t,r)Y_{lm}(\theta,\varphi)\,,\qquad\xi_{a}=\sum_{l,m}\Theta(t,r)\nabla_{a}Y_{lm}(\theta,\varphi)\,. (60)

Then, the perturbations H0H_{0}, H1H_{1}, H2H_{2}, β\beta, α\alpha, KK, GG, and δϕ\delta\phi transform, respectively, as Kobayashi:2014wsa ; Motohashi:2011pw

H0H0+2f𝒯˙fhf,H1H1+˙+𝒯ff𝒯,H2H2+2h+h,\displaystyle H_{0}\to H_{0}+\frac{2}{f}\dot{{\cal T}}-\frac{f^{\prime}h}{f}{\cal R}\,,\qquad H_{1}\to H_{1}+\dot{{\cal R}}+{\cal T}^{\prime}-\frac{f^{\prime}}{f}{\cal T}\,,\qquad H_{2}\to H_{2}+2h{\cal R}^{\prime}+h^{\prime}{\cal R}\,, (61)
ββ+𝒯+Θ˙ff𝒯,αα++Θ2rΘ,KK+2rh,GG+2r2Θ,\displaystyle\beta\to\beta+{\cal T}+\dot{\Theta}-\frac{f^{\prime}}{f}{\cal T}\,,\qquad\alpha\to\alpha+{\cal R}+\Theta^{\prime}-\frac{2}{r}\Theta\,,\qquad K\to K+\frac{2}{r}h{\cal R}\,,\qquad G\to G+\frac{2}{r^{2}}\Theta\,, (62)
δϕδϕϕ¯h,\displaystyle\delta\phi\to\delta\phi-\bar{\phi}^{\prime}h{\cal R}\,, (63)

where a dot represents the derivative with respect to tt. For the multipoles l2l\geq 2, the transformation scalars 𝒯{\cal T} and Θ\Theta can be fixed by choosing the gauge:

β=0,G=0.\beta=0\,,\qquad G=0\,. (64)

To fix the other transformation scalar {\cal R}, there are several different gauge choices listed below.

(i)Uniformcurvaturegauge:K=0,\displaystyle{\rm(i)~{}Uniform~{}curvature~{}gauge}:~{}K=0\,, (65)
(ii)Unitarygauge:δϕ=0,\displaystyle{\rm(ii)~{}Unitary~{}gauge}:~{}\delta\phi=0\,, (66)
(iii)Spatiallydiagonalgauge:α=0.\displaystyle{\rm(iii)~{}Spatially~{}diagonal~{}gauge}:~{}\alpha=0\,. (67)

The physics is not affected by different choices of gauges. As in Refs. DeFelice:2011ka ; Motohashi:2011pw ; Kobayashi:2014wsa , we will choose the uniform curvature gauge (i) to compute the second-order action of even-parity perturbations.

The above argument of gauge fixings is valid for the multipoles l2l\geq 2. For the monopole (l=0l=0), the perturbations β\beta, α\alpha, and GG vanish identically. For the dipole (l=1l=1), the perturbations in habh_{ab} appear only as the combination KGK-G, so the decomposition into the two components KK and GG is redundant. We will separately study these cases in Sec. VI.

IV.3 Matter density perturbation and velocity potential

We discuss the structure of perturbations in the Schutz-Sorkin action in more detail. The matter density perturbation δρ\delta\rho is related to the perturbation δn\delta n of fluid number density given by Eq. (13). The components (48) of JμJ_{\mu} in the odd-parity sector do not give rise to the first-order perturbation of nn. On the other hand, the components (57) in the even-parity sector generate the first-order perturbation of nn. From this first-order perturbation δn\delta n, we define the matter density perturbation δρ(t,r)\delta\rho(t,r), as

δn=l,mδρ(t,r)ρ,n(r)Ylm(θ,φ).\delta n=\sum_{l,m}\frac{\delta\rho(t,r)}{\rho_{,n}(r)}Y_{lm}(\theta,\varphi)\,. (68)

After the expansion of nn in terms of even-mode perturbations, we will convert δn\delta n to δρ(t,r)\delta\rho(t,r) for computing the second-order action.

For the quantity \ell appearing in the action (12), we will derive its explicit form by using the constraint (16). Up to first order in perturbations, the partial derivatives of \ell with respect to a=θ,φa=\theta,\varphi are given by

a=ρ,nuaδ𝒜a,\partial_{a}\ell=\rho_{,n}u_{a}-\delta{\cal A}_{a}\,, (69)

where we used Eq. (50). Since we are considering the derivative of the scalar quantity \ell, we only need to consider the even-mode contribution to uau_{a}, i.e., the first term in Eq. (43). On using the second of Eq. (58) and integrating Eq. (69) with respect to aa, it follows that

=(t,r)+l,m[ρ,n(r)v(t,r)δ𝒜2(t,r)]Ylm(θ,φ),\ell={\cal F}(t,r)+\sum_{l,m}\left[\rho_{,n}(r)v(t,r)-\delta{\cal A}_{2}(t,r)\right]Y_{lm}(\theta,\varphi)\,, (70)

where {\cal F} is a function of tt and rr. The term ρ,n(r)v(t,r)\rho_{,n}(r)v(t,r) in Eq. (70) corresponds to the first-order quantity, so that ρ,n\rho_{,n} should be evaluated on the background. The time derivative t\partial_{t}\ell is equivalent to ρ,n(r)u¯t(r)=ρ,n(r)f(r)\rho_{,n}(r)\bar{u}_{t}(r)=-\rho_{,n}(r)\sqrt{f(r)} at the background level. The integration of the relation t=ρ,n(r)f(r)\partial_{t}{\cal F}=-\rho_{,n}(r)\sqrt{f(r)} gives (t,r)=ρ,n(r)f(r)t{\cal F}(t,r)=-\rho_{,n}(r)\sqrt{f(r)}\,t, so that

=ρ,n(r)f(r)t+l,m[ρ,n(r)v(t,r)δ𝒜2(t,r)]Ylm(θ,φ).\ell=-\rho_{,n}(r)\sqrt{f(r)}\,t+\sum_{l,m}\left[\rho_{,n}(r)v(t,r)-\delta{\cal A}_{2}(t,r)\right]Y_{lm}(\theta,\varphi)\,. (71)

The temporal and radial components of four velocity uμu_{\mu} can be expressed, respectively, as

ut=f(r)+l,mδut(t,r)Ylm(θ,φ),ur=l,mδur(t,r)Ylm(θ,φ),u_{t}=-\sqrt{f(r)}+\sum_{l,m}\delta u_{t}(t,r)Y_{lm}(\theta,\varphi)\,,\qquad u_{r}=\sum_{l,m}\delta u_{r}(t,r)Y_{lm}(\theta,\varphi)\,, (72)

where δut\delta u_{t} and δur\delta u_{r} are functions of tt and rr. Since t=ρ,nut\partial_{t}\ell=\rho_{,n}u_{t} up to first order in perturbations, we obtain the correspondence,

δut=f(r)cm2δρρ+P+v˙,\delta u_{t}=\sqrt{f(r)}c_{m}^{2}\frac{\delta\rho}{\rho+P}+\dot{v}\,, (73)

where cm2c_{m}^{2} is the matter sound speed squared defined by

cm2nρ,nnρ,n.c_{m}^{2}\equiv\frac{n\rho_{,nn}}{\rho_{,n}}\,. (74)

Similarly, we take the rr derivative of Eq. (71) and compare it with the relation r=ρ,nurδ𝒜r\partial_{r}\ell=\rho_{,n}u_{r}-\delta{\cal A}_{r}. In doing so, we exploit the property,

ddr(ρ,n(r)f(r))=f(r)(ρ,n+f2fρ,n)=0,\frac{{\rm d}}{{\rm d}r}\left(\rho_{,n}(r)\sqrt{f(r)}\right)=\sqrt{f(r)}\left(\rho_{,n}^{\prime}+\frac{f^{\prime}}{2f}\rho_{,n}\right)=0\,, (75)

where the second equality follows from Eq. (38) with ρ,n=P/n\rho_{,n}^{\prime}=P^{\prime}/n. Then, the perturbation δur\delta u_{r} can be expressed as

δur=vf2fv+δ𝒜1δ𝒜2ρ,n.\delta u_{r}=v^{\prime}-\frac{f^{\prime}}{2f}v+\frac{\delta{\cal A}_{1}-\delta{\cal A}_{2}^{\prime}}{\rho_{,n}}\,. (76)

Equations (73) and (76) show the correspondence between the perturbations δρ\delta\rho, δ𝒜1\delta{\cal A}_{1}, δ𝒜2\delta{\cal A}_{2} and the components of four velocity uμu_{\mu}. Since uμ=Jμ/(ng)u_{\mu}=J_{\mu}/(n\sqrt{-g}), there are also relations between the perturbations δJr\delta J_{r}, δJ\delta J in Eq. (57) and vv, δ𝒜1\delta{\cal A}_{1}, δ𝒜2\delta{\cal A}_{2} appeared above. In Sec. VI, we will address this issue.

V Odd-parity perturbations

We expand the action (11) up to second-order in odd-parity perturbations. In doing so, we choose the Regge-Wheeler gauge (55) for l2l\geq 2. For the dipole (l=1l=1) the condition U=0U=0 automatically holds, so we study this case separately at the end of this section.

The scalar field ϕ\phi does not possess the odd-parity perturbation, so we can use the background value of G4(ϕ)G_{4}(\phi) for the expansion of G4(ϕ)RG_{4}(\phi)R. The field kinetic energy XX is expanded as

X=12hϕ2h22r2ϕ2W2[(θYlm)2+(φYlm)2sin2θ]+𝒪(ε4),X=-\frac{1}{2}h\phi^{\prime 2}-\frac{h^{2}}{2r^{2}}\phi^{\prime 2}W^{2}\left[(\partial_{\theta}Y_{lm})^{2}+\frac{(\partial_{\varphi}Y_{lm})^{2}}{\sin^{2}\theta}\right]+{\cal O}(\varepsilon^{4})\,, (77)

where εn\varepsilon^{n} represents the nn-th order of perturbations. We expand the k-essence Lagrangian as G2(ϕ,X)=G2(r)+G2,XδXG_{2}(\phi,X)=G_{2}(r)+G_{2,X}\delta X, where δX\delta X is the second-order perturbation in Eq. (77).

The fluid number density (13) can be decomposed into the background part n(r)n(r) and the second-order perturbed part δn\delta n given by

δn=n2r2(hW22fQ2+2nfQδjδj2n2)[(θYlm)2+(φYlm)2sin2θ]+𝒪(ε4).\delta n=\frac{n}{2r^{2}}\left(hW^{2}-\frac{2}{f}Q^{2}+\frac{2}{n\sqrt{f}}Q\delta j-\frac{\delta j^{2}}{n^{2}}\right)\left[(\partial_{\theta}Y_{lm})^{2}+\frac{(\partial_{\varphi}Y_{lm})^{2}}{\sin^{2}\theta}\right]+{\cal O}(\varepsilon^{4})\,. (78)

Then, the matter density ρ\rho in the action (12) contains the second-order perturbation ρ,n(r)δn\rho_{,n}(r)\delta n. As we derived in Eq. (71), the quantity \ell has only even-mode perturbations and hence =ρ,n(r)f(r)t\ell=-\rho_{,n}(r)\sqrt{f(r)}\,t for odd modes. We caution that the vector component JtJ^{t} contains the second-order odd-parity perturbation δJt\delta J^{t} besides the background value J¯t=n(r)r2sinθ/h\bar{J}^{t}=n(r)r^{2}\sin\theta/\sqrt{h}. The perturbations δJθ\delta J^{\theta} and δJφ\delta J^{\varphi} correspond to the first-order perturbation. Hence the terms δJt˙+a=θ,φ(J¯tδ𝒜aδa˙+δJaδ𝒜a)\delta J^{t}\dot{\ell}+\sum_{a=\theta,\varphi}(\bar{J}^{t}\delta{\cal A}_{a}\dot{\delta{\cal B}^{a}}+\delta J^{a}\delta{\cal A}_{a}) give rise to the second-order contribution to Eq. (12).

After expanding Eq. (11) up to quadratic order in odd-parity perturbations, the resulting second-order action contains terms multiplied by δ𝒜(t,r)\delta{\cal A}(t,r). Varying this action with respect to δ𝒜\delta{\cal A}, it follows that

δ˙=nQfδjnr2.\dot{\delta{\cal B}}=\frac{nQ-\sqrt{f}\,\delta j}{nr^{2}}\,. (79)

After substituting this relation into the second-order action, the terms related to δj\delta j appear as the quadratic dependence δj2\delta j^{2}. Hence the variation of the action with respect to δj\delta j gives

δj=0.\delta j=0\,. (80)

In this way, the perturbations δ𝒜\delta{\cal A}, δ\delta{\cal B}, δj\delta j are integrated out from the second-order action.

The next step is to perform the integral with respect to θ\theta and φ\varphi. In this procedure, it is sufficient to set m=0m=0 and multiply the action by 2π2\pi. As for the integration with respect to θ\theta, we use the formulas of integrals of Yl0Y_{l0} and their θ\theta derivatives given in Appendix B of Ref. Kase:2018voo . The resulting quadratic-order action contains the tt and rr derivatives of perturbations WW and QQ, so we integrate some of them by parts. By using the background Eqs. (35) and (36), the second-order action of odd-parity perturbations reduces to

𝒮odd(2)=l,mdtdr[L2hfG4(W˙Q+2Qr)2L(L2)fhG42r2W2+L(L2)G42fhr2Q2],{\cal S}^{(2)}_{\rm odd}=\sum_{l,m}\int{\rm d}t{\rm d}r\left[\frac{L}{2}\sqrt{\frac{h}{f}}G_{4}\left(\dot{W}-Q^{\prime}+\frac{2Q}{r}\right)^{2}-\frac{L(L-2)\sqrt{fh}\,G_{4}}{2r^{2}}W^{2}+\frac{L(L-2)G_{4}}{2\sqrt{fh}\,r^{2}}Q^{2}\right]\,, (81)

where

L=l(l+1).L=l(l+1)\,. (82)

From Eq. (81), we observe that the presence of the perfect fluid does not affect the evolution of odd-mode perturbations. In the following, we will study the two different cases: (i) l2l\geq 2 and (ii) l=1l=1, in turn.

V.1 l2l\geq 2

The variation of Eq. (81) with respect to the nondynamical variable QQ leads to a constraint equation for QQ. However, the presence of the Q2Q^{\prime 2} term does not allow one to solve explicitly for QQ. To overcome this problem, we introduce the Lagrange multiplier χ(t,r)\chi(t,r) and express the action (81) in the form,

𝒮odd(2)=l,mdtdr{L2hfG4[2χ(W˙Q+2Qr)χ2]L(L2)fhG42r2W2+L(L2)G42fhr2Q2}.{\cal S}^{(2)}_{\rm odd}=\sum_{l,m}\int{\rm d}t{\rm d}r\left\{\frac{L}{2}\sqrt{\frac{h}{f}}G_{4}\left[2\chi\left(\dot{W}-Q^{\prime}+\frac{2Q}{r}\right)-\chi^{2}\right]-\frac{L(L-2)\sqrt{fh}\,G_{4}}{2r^{2}}W^{2}+\frac{L(L-2)G_{4}}{2\sqrt{fh}\,r^{2}}Q^{2}\right\}\,. (83)

Varying the action (83) with respect to WW and QQ, respectively, we obtain

W\displaystyle W =\displaystyle= r2f(L2)χ˙,\displaystyle-\frac{r^{2}}{f(L-2)}\dot{\chi}\,, (84)
Q\displaystyle Q =\displaystyle= [(hχ+2hχ)r+4hχ]G4ffhrχG4+2fhrϕχG4,ϕ2(L2)fG4r.\displaystyle-\frac{[(h^{\prime}\chi+2h\chi^{\prime})r+4h\chi]G_{4}f-f^{\prime}hr\chi G_{4}+2fhr\phi^{\prime}\chi G_{4,\phi}}{2(L-2)fG_{4}}r\,. (85)

Substituting these relations and their tt and rr derivatives into Eq. (83) and integrating it by parts, the second-order action is expressed in the form,

𝒮odd(2)=l,mdtdr(𝒦χ˙2+𝒢χ2+χ2),{\cal S}^{(2)}_{\rm odd}=\sum_{l,m}\int{\rm d}t{\rm d}r\left({\cal K}\dot{\chi}^{2}+{\cal G}\chi^{\prime 2}+{\cal M}\chi^{2}\right)\,, (86)

where

𝒦\displaystyle{\cal K} =\displaystyle= Lhr2G42f3/2(L2),𝒢=Lh3/2r2G42f(L2),\displaystyle\frac{L\sqrt{h}r^{2}G_{4}}{2f^{3/2}(L-2)}\,,\qquad{\cal G}=-\frac{Lh^{3/2}r^{2}G_{4}}{2\sqrt{f}(L-2)}\,, (87)
\displaystyle{\cal M} =\displaystyle= Lh[{(2L+4h4h′′r24hr)f2+(f′′h+fh)fr2f2hr2}G42\displaystyle-L\sqrt{h}[\{(2L+4h-4-h^{\prime\prime}r^{2}-4h^{\prime}r)f^{2}+(f^{\prime\prime}h+f^{\prime}h^{\prime})fr^{2}-f^{\prime 2}hr^{2}\}G_{4}^{2} (88)
2{(G4,ϕϕϕ2+G4,ϕϕ′′)h+G4,ϕhϕ}f2r2G4+2f2hr2ϕ2G4,ϕ2]/[4f5/2(L2)G4].\displaystyle-2\{(G_{4,\phi\phi}\phi^{\prime 2}+G_{4,\phi}\phi^{\prime\prime})h+G_{4,\phi}h^{\prime}\phi^{\prime}\}f^{2}r^{2}G_{4}+2f^{2}hr^{2}\phi^{\prime 2}G_{4,\phi}^{2}]/[4f^{5/2}(L-2)G_{4}]\,.

There is one propagating degree of freedom χ\chi arising from the gravitational sector. Since L6L\geq 6 for l2l\geq 2, the condition for the absence of ghosts corresponds to 𝒦>0{\cal K}>0, i.e.,

2G4>0.{\cal F}\equiv 2G_{4}>0\,. (89)

Let us derive the propagation speed of χ\chi along the radial direction. Assuming the solution of the form χ=ei(ωtkr)\chi=e^{i(\omega t-kr)} and taking the limits of large ω\omega and kk, the dispersion relation following from Eq. (86) reads

ω2𝒦+k2𝒢=0.\omega^{2}{\cal K}+k^{2}{\cal G}=0\,. (90)

The propagation speed squared in terms of the coordinates tt and rr is c^r2=ω2/k2=𝒢/𝒦=fh\hat{c}_{r}^{2}=\omega^{2}/k^{2}=-{\cal G}/{\cal K}=fh. The propagation speed crc_{r} along the radial direction in proper time τ=fdt\tau=\int\sqrt{f}{\rm d}t is given by cr=dr/dτc_{r}={\rm d}r_{*}/{\rm d}\tau, where dr=dr/h{\rm d}r_{*}={\rm d}r/\sqrt{h}. Since crc_{r} is related to c^r=dr/dt\hat{c}_{r}={\rm d}r/{\rm d}t as cr=c^r/fhc_{r}=\hat{c}_{r}/\sqrt{fh}, it follows that

cr2=1.c_{r}^{2}=1\,. (91)

Hence there is no Laplacian instability of the odd-mode perturbation χ\chi in the radial direction.

Along the angular direction, we employ the solution of the form χ=ei(ωtlθ)\chi=e^{i(\omega t-l\theta)}. For large multipoles (l1l\gg 1), the term χ2{\cal M}\chi^{2} in Eq. (86) contributes to the dispersion relation besides the term 𝒦χ˙2{\cal K}\dot{\chi}^{2}, so that

ω2𝒦+=0,\omega^{2}{\cal K}+{\cal M}=0\,, (92)

where

𝒦=hr2G42f3/2+𝒪(l2),=hG42fl2+𝒪(l).{\cal K}=\frac{\sqrt{h}r^{2}G_{4}}{2f^{3/2}}+{\cal O}(l^{-2})\,,\qquad{\cal M}=-\frac{\sqrt{h}G_{4}}{2\sqrt{f}}\,l^{2}+{\cal O}(l)\,. (93)

In terms of the time coordinate tt, the propagation speed squared along the angular direction is c^Ω2=ω2r2/l2=(/𝒦)r2/l2=f\hat{c}_{\Omega}^{2}=\omega^{2}r^{2}/l^{2}=-({\cal M}/{\cal K})r^{2}/l^{2}=f, where we have taken the limit ll\to\infty in the last equality. In proper time, the propagation speed is given by cΩ=rdθ/dτ=c^Ω/fc_{\Omega}=r{\rm d}\theta/{\rm d}\tau=\hat{c}_{\Omega}/\sqrt{f}. Hence, in the limit l1l\gg 1, we obtain

cΩ2=1.c_{\Omega}^{2}=1\,. (94)

This means that there is no Laplacian instability along the angular direction either. We have thus shown that the stability of the odd-mode perturbation χ\chi is ensured under the no-ghost condition G4>0G_{4}>0. For the theories (11), the speed of odd-parity GWs is equivalent to that of light.

V.2 l=1l=1

For the dipole there is the relation U=0U=0, so we have a residual gauge degree of freedom. Since L=2L=2 in this case, the last two terms in the square bracket of Eq. (81) vanish. Then, the second-order action reduces to

𝒮odd(2)=l,mdtdrhfG4(W˙Q+2Qr)2.{\cal S}^{(2)}_{\rm odd}=\sum_{l,m}\int{\rm d}t{\rm d}r\sqrt{\frac{h}{f}}G_{4}\left(\dot{W}-Q^{\prime}+\frac{2Q}{r}\right)^{2}\,. (95)

For the gauge choice W=0W=0, the transformation scalar Λ\Lambda in Eq. (54) is given by

Λ(t,r)=r2dr~W(t,r~)r~2+r2𝒞1(t),\Lambda(t,r)=-r^{2}\int{\rm d}\tilde{r}\frac{W(t,\tilde{r})}{\tilde{r}^{2}}+r^{2}{\cal C}_{1}(t)\,, (96)

where 𝒞1(t){\cal C}_{1}(t) is an arbitrary function of tt corresponding to a gauge mode. We vary the action (95) with respect to WW and QQ, and set W=0W=0 in the end. This leads to

𝒰˙\displaystyle\dot{\cal U} =\displaystyle= 0,\displaystyle 0\,, (97)
(r2𝒰)\displaystyle\left(r^{2}{\cal U}\right)^{\prime} =\displaystyle= 0,\displaystyle 0\,, (98)

where

𝒰(t,r)hfG4(Q2Qr).{\cal U}(t,r)\equiv\sqrt{\frac{h}{f}}G_{4}\left(Q^{\prime}-\frac{2Q}{r}\right)\,. (99)

The integrated solutions to Eqs. (97) and (98) are given by

𝒰=Cr2,{\cal U}=\frac{C}{r^{2}}\,, (100)

where CC is a constant. Substituting Eq. (100) into Eq. (99) and solving it for QQ, it follows that

Q(t,r)=r2dr~Cr~4fh1G4+r2𝒞2(t),Q(t,r)=r^{2}\int{\rm d}\tilde{r}\frac{C}{\tilde{r}^{4}}\sqrt{\frac{f}{h}}\frac{1}{G_{4}}+r^{2}{\cal C}_{2}(t)\,, (101)

where 𝒞2(t){\cal C}_{2}(t) is an arbitrary function of tt corresponding to the gauge mode. The gauge modes appearing in Eqs. (96) and (101) can be eliminated by choosing

𝒞˙1(t)=𝒞2(t).\dot{\cal C}_{1}(t)={\cal C}_{2}(t)\,. (102)

Substituting Eq. (100) into Eq. (95) with the gauge choice W=0W=0, we obtain

𝒮odd(2)=l,mdtdrfhC2r4G4,{\cal S}^{(2)}_{\rm odd}=\sum_{l,m}\int{\rm d}t{\rm d}r\sqrt{\frac{f}{h}}\frac{C^{2}}{r^{4}G_{4}}\,, (103)

which means that there is no dynamical propagating degree of freedom for l=1l=1.

We note that the perturbation W˙Q+2Q/r\dot{W}-Q^{\prime}+2Q/r appearing in the action (95) is gauge-invariant. In terms of the field χ\chi introduced in Eq. (83), the variation of the action with respect to WW gives χ˙=0\dot{\chi}=0. Since χ\chi depends on rr alone, the gauge-invariant perturbation χ=W˙Q+2Q/r\chi=\dot{W}-Q^{\prime}+2Q/r does not work as a dynamical perturbation. This is consistent with the argument given above.

VI Even-parity perturbations

We proceed to the derivation of stability conditions in the even-parity sector by expanding the action up to second order in perturbations. Since the second-order action is different depending on the multipoles ll, we will discuss the three cases: (A) l2l\geq 2, (B) l=0l=0, and (C) l=1l=1, in turn.

VI.1 l2l\geq 2

For l2l\geq 2, we choose the uniform curvature gauge given by

K=0,β=0,G=0,K=0\,,\qquad\beta=0\,,\qquad G=0\,, (104)

under which 𝒯{\cal T}, {\cal R}, and Θ\Theta in the gauge transformation (62) are fixed. In the gravity sector, we are left with four metric perturbations H0H_{0}, H1H_{1}, H2H_{2}, and α\alpha. For the perfect fluid, we consider the vector field JμJ_{\mu} in the form (57) and adopt the configuration (50) with the perturbations δ𝒜i\delta{\cal A}_{i} and δi\delta{\cal B}_{i} given by Eqs. (58) and (59). From Eq. (71), the Lagrange multiplier \ell contains the velocity potential v(t,r)v(t,r) and the perturbation δ𝒜2(t,r)\delta{\cal A}_{2}(t,r). This expression of \ell is used for expanding the Schutz-Sorkin action. The matter perturbation δρ\delta\rho is related to the perturbation of number density δn\delta n, as Eq. (68). The density ρ\rho in the Schutz-Sorkin action is expanded in the form,

ρ=ρ(r)+ρ,nδn+ρ,n2ncm2δn2+𝒪(ε3),\rho=\rho(r)+\rho_{,n}\delta n+\frac{\rho_{,n}}{2n}c_{m}^{2}\delta n^{2}+{\cal O}(\varepsilon^{3})\,, (105)

where cm2c_{m}^{2} is defined by Eq. (74).

In the scalar-field sector, we perform the expansions,

G2(ϕ,X)\displaystyle G_{2}(\phi,X) =\displaystyle= G2(r)+G2,ϕδϕ+G2,XδX+12G2,ϕϕδϕ2+12G2,XXδX2+G2,ϕXδϕδX+𝒪(ε3),\displaystyle G_{2}(r)+G_{2,\phi}\delta\phi+G_{2,X}\delta X+\frac{1}{2}G_{2,\phi\phi}\delta\phi^{2}+\frac{1}{2}G_{2,XX}\delta X^{2}+G_{2,\phi X}\delta\phi\delta X+{\cal O}(\varepsilon^{3})\,, (106)
G4(ϕ)\displaystyle G_{4}(\phi) =\displaystyle= G4(r)+G4,ϕδϕ+12G4,ϕϕδϕ2+𝒪(ε3),\displaystyle G_{4}(r)+G_{4,\phi}\delta\phi+\frac{1}{2}G_{4,\phi\phi}\delta\phi^{2}+{\cal O}(\varepsilon^{3})\,, (107)

where δϕ=l,mδϕ(t,r)Ylm\delta\phi=\sum_{l,m}\delta\phi(t,r)Y_{lm}, and

δX\displaystyle\delta X =\displaystyle= l,m12hϕ(ϕH22δϕ)Ylm+l,m12f[δϕ˙2+h2ϕ2H12h{fδϕ2+2ϕ(H1δϕ˙fH2δϕ)+fϕ2H22}]Ylm2\displaystyle\sum_{l,m}\frac{1}{2}h\phi^{\prime}\left(\phi^{\prime}H_{2}-2\delta\phi^{\prime}\right)Y_{lm}+\sum_{l,m}\frac{1}{2f}\left[\dot{\delta\phi}^{2}+h^{2}\phi^{\prime 2}H_{1}^{2}-h\left\{f\delta\phi^{\prime 2}+2\phi^{\prime}(H_{1}\dot{\delta\phi}-fH_{2}\delta\phi^{\prime})+f\phi^{\prime 2}H_{2}^{2}\right\}\right]Y_{lm}^{2} (108)
l,m12r2(δϕhϕα)2[(θYlm)2+(φYlm)2sin2θ]+𝒪(ε3).\displaystyle-\sum_{l,m}\frac{1}{2r^{2}}\left(\delta\phi-h\phi^{\prime}\alpha\right)^{2}\left[(\partial_{\theta}Y_{lm})^{2}+\frac{(\partial_{\varphi}Y_{lm})^{2}}{\sin^{2}\theta}\right]+{\cal O}(\varepsilon^{3})\,.

Since it is sufficient to consider the mode m=0m=0, we will do so in the following discussion. We expand the total action (11) up to second order in even-mode perturbations. Varying the resulting second-order action with respect to δ1\delta{\cal B}_{1}, δ2\delta{\cal B}_{2}, δJ\delta J, and δJr\delta J_{r}, respectively, it follows that

δ𝒜˙1=0,\displaystyle\dot{\delta{\cal A}}_{1}=0\,, (109)
δ𝒜˙2=0,\displaystyle\dot{\delta{\cal A}}_{2}=0\,, (110)
nvδJ=0,\displaystyle nv-\delta J=0\,, (111)
2(ρ+P)δJr2n2(δ𝒜1δ𝒜2)n(ρ+P)(2vffv)=0.\displaystyle 2\left(\rho+P\right)\delta J_{r}-2n^{2}\left(\delta{\cal A}_{1}-\delta{\cal A}_{2}^{\prime}\right)-n\left(\rho+P\right)\left(2v^{\prime}-\frac{f^{\prime}}{f}v\right)=0\,. (112)

We solve Eqs. (111) and (112) for δJ\delta J and δJr\delta J_{r}, respectively, and substitute them into the total second-order action. After this procedure, there exist the terms proportional to δ𝒜1δ˙1\delta{\cal A}_{1}\dot{\delta{\cal B}}_{1} and δ𝒜2δ˙2\delta{\cal A}_{2}\dot{\delta{\cal B}}_{2} with time-independent coefficients, but they can be integrated out on account of Eqs. (109) and (110). The second-order action containing the perturbations δ𝒜1\delta{\cal A}_{1} and δ𝒜2\delta{\cal A}_{2} reduces to 𝒮δ𝒜1,2=l,mdtdrδ𝒜1,2{\cal S}_{\delta{\cal A}_{1,2}}=\sum_{l,m}\int{\rm d}t{\rm d}r{\cal L}_{\delta{\cal A}_{1,2}}, where

δ𝒜1,2=nr22(ρ+P)hf[(ρ+P)(2fH1+fv2fv)(δ𝒜1δ𝒜2)nf(δ𝒜1δ𝒜2)2].{\cal L}_{\delta{\cal A}_{1,2}}=\frac{nr^{2}}{2(\rho+P)}\sqrt{\frac{h}{f}}\left[(\rho+P)\left(2\sqrt{f}H_{1}+f^{\prime}v-2fv^{\prime}\right)(\delta{\cal A}_{1}-\delta{\cal A}_{2}^{\prime})-nf\left(\delta{\cal A}_{1}-\delta{\cal A}_{2}^{\prime}\right)^{2}\right]\,. (113)

The combination δ𝒜1δ𝒜2\delta{\cal A}_{1}-\delta{\cal A}_{2}^{\prime} can be replaced with the perturbation δur\delta u_{r} given by Eq. (76). Then, varying the action (113) with respect to δur\delta u_{r}, we obtain

δur=H1f.\delta u_{r}=\frac{H_{1}}{\sqrt{f}}\,. (114)

On using this relation, the Lagrangian (113) can be expressed in terms of vv, its rr derivative, and H1H_{1}, as

δ𝒜1,2=(ρ+P)r2h8f3/2(2fH1+fv2fv)2.{\cal L}_{\delta{\cal A}_{1,2}}=\frac{(\rho+P)r^{2}\sqrt{h}}{8f^{3/2}}\left(2\sqrt{f}H_{1}+f^{\prime}v-2fv^{\prime}\right)^{2}\,. (115)

In the full quadratic-order action, there are also terms containing the perturbations vv and δρ\delta\rho, which arise from JμμJ^{\mu}\partial_{\mu}\ell and ρ(n)\rho(n) in Eq. (12).

On using the background Eqs. (35)-(37) and performing the integration by parts, the second-order action of even-mode perturbations can be expressed in the form 𝒮even(2)=l,mdtdr{\cal S}_{\rm even}^{(2)}=\sum_{l,m}\int{\rm d}t{\rm d}r{\cal L}, where

\displaystyle{\cal L} =\displaystyle= H0[a1δϕ′′+a2δϕ+a3H2+La4α+(a5+La6)δϕ+(a7+La8)H2+La9α+a10δρ]\displaystyle H_{0}\left[a_{1}\delta\phi^{\prime\prime}+a_{2}\delta\phi^{\prime}+a_{3}H_{2}^{\prime}+La_{4}\alpha^{\prime}+\left(a_{5}+La_{6}\right)\delta\phi+\left(a_{7}+La_{8}\right)H_{2}+La_{9}\alpha+a_{10}\delta\rho\right] (116)
+Lb1H12+H1(b2δϕ˙+b3δϕ˙+b4H˙2+Lb5α˙)+c1δϕ˙H˙2+H2[c2δϕ+(c3+Lc4)δϕ+Lc5α+c~5v˙]+c6H22\displaystyle+Lb_{1}H_{1}^{2}+H_{1}\left(b_{2}\dot{\delta\phi}^{\prime}+b_{3}\dot{\delta\phi}+b_{4}\dot{H}_{2}+Lb_{5}\dot{\alpha}\right)+c_{1}\dot{\delta\phi}\dot{H}_{2}+H_{2}\left[c_{2}\delta\phi^{\prime}+(c_{3}+Lc_{4})\delta\phi+Lc_{5}\alpha+\tilde{c}_{5}\dot{v}\right]+c_{6}H_{2}^{2}
+Ld1α˙2+Lα(d2δϕ+d3δϕ)+Ld4α2+e1δϕ˙2+e2δϕ2+(e3+Le4)δϕ2+Lf1v2+f2δρ2+f3δρv˙.\displaystyle+Ld_{1}\dot{\alpha}^{2}+L\alpha\left(d_{2}\delta\phi^{\prime}+d_{3}\delta\phi\right)+Ld_{4}\alpha^{2}+e_{1}\dot{\delta\phi}^{2}+e_{2}\delta\phi^{\prime 2}+\left(e_{3}+Le_{4}\right)\delta\phi^{2}+Lf_{1}v^{2}+f_{2}\delta\rho^{2}+f_{3}\delta\rho\,\dot{v}\,.

The background-dependent coefficients a1,a2,a_{1},a_{2},... are explicitly given in Appendix. In comparison to the paper by Kobayashi, Motohashi, Suyama (KMS) Kobayashi:2014wsa without the perfect fluid, there is the notational difference of a factor 1/21/2, i.e., a1=a1KMS/2,,e4=e4KMS/2a_{1}=a_{1}^{\rm KMS}/2,...,e_{4}=e_{4}^{\rm KMS}/2, due to the different normalization of YlmY_{lm}. Each second equality in Eq. (A.1) among coefficients (e.g., a5=a2a1′′a_{5}=a_{2}^{\prime}-a_{1}^{\prime\prime}) is valid even in full Horndeski theories containing the dependence of G3(ϕ,X)G_{3}(\phi,X), G4(ϕ,X)G_{4}(\phi,X), and G5(ϕ,X)G_{5}(\phi,X) Kobayashi:2014wsa .

When the perfect fluid is absent, there are two propagating degrees of freedom in the even-parity sector. One of them is the field perturbation δϕ\delta\phi, and the other is the following combination DeFelice:2011ka ; Kobayashi:2014wsa ,

ψa3H2+La4α+a1δϕ,\psi\equiv a_{3}H_{2}+La_{4}\alpha+a_{1}\delta\phi^{\prime}\,, (117)

which corresponds to the dynamical perturbation in the gravity sector. The variable ψ\psi is analogous to the dynamical perturbation taken by Moncrief Moncrief and Zerilli Zerilli:1970se in the Regge-Wheeler gauge (α=β=G=0\alpha=\beta=G=0). While the Moncrief-Zerilli variable Lousto:1996sx is the combination of H2H_{2} and KK, the perturbation (117) contains H2H_{2} and α\alpha. Due to the existence of the term a1δϕa_{1}\delta\phi^{\prime} in Eq. (117), the derivatives a1δϕ′′a_{1}\delta\phi^{\prime\prime} and a3H2a_{3}H_{2}^{\prime} in Eq. (116) can be simultaneously replaced with ψ\psi^{\prime} Kobayashi:2014wsa .

In the perfect-fluid sector, we introduce the following dynamical matter perturbation,

δρmδρ+2fhr3f1f3[ha3(2ffr)+Lfra4]ψhf1r2[a1(2ffr)+f2rb3]ff3[ha3(2ffr)+Lfra4]δϕ.\delta\rho_{m}\equiv\delta\rho+\frac{2\sqrt{f}\,hr^{3}f_{1}}{f_{3}[ha_{3}(2f-f^{\prime}r)+Lfra_{4}]}\psi^{\prime}-\frac{hf_{1}r^{2}[a_{1}(2f-f^{\prime}r)+f^{2}rb_{3}]}{\sqrt{f}f_{3}[ha_{3}(2f-f^{\prime}r)+Lfra_{4}]}\delta\phi^{\prime}\,. (118)

If we try to obtain the second-order action of dynamical perturbations in terms of δρ\delta\rho, this gives rise to the apparent dynamical terms ψ˙2\dot{\psi}^{\prime 2} and δϕ˙2\dot{\delta\phi}^{\prime 2}. However, they can be eliminated by introducing the second and third terms on the right-hand-side of Eq. (118).

Varying the Lagrangian (116) with respect to the nondynamical perturbations H0H_{0}, H1H_{1}, and vv, respectively, we obtain

a1δϕ′′+a2δϕ+a3H2+La4α+(a5+La6)δϕ+(a7+La8)H2+La9α+a10δρ=0,\displaystyle a_{1}\delta\phi^{\prime\prime}+a_{2}\delta\phi^{\prime}+a_{3}H_{2}^{\prime}+La_{4}\alpha^{\prime}+\left(a_{5}+La_{6}\right)\delta\phi+\left(a_{7}+La_{8}\right)H_{2}+La_{9}\alpha+a_{10}\delta\rho=0\,, (119)
H1=12Lb1(b2δϕ˙+b3δϕ˙+b4H˙2+Lb5α˙),\displaystyle H_{1}=-\frac{1}{2Lb_{1}}\left(b_{2}\dot{\delta\phi}^{\prime}+b_{3}\dot{\delta\phi}+b_{4}\dot{H}_{2}+Lb_{5}\dot{\alpha}\right)\,, (120)
v=12Lf1(f3δρ˙+c~5H˙2).\displaystyle v=\frac{1}{2Lf_{1}}\left(f_{3}\dot{\delta\rho}+\tilde{c}_{5}\dot{H}_{2}\right)\,. (121)

Taking the rr derivative of Eq. (117) and substituting it into Eq. (119), the nondynamical perturbation α\alpha can be expressed in terms of ψ\psi, δϕ\delta\phi, and δρm\delta\rho_{m} and their first radial derivatives. From Eq. (117), the variable H2H_{2} and its time derivative are written in terms of ψ\psi, α\alpha, δϕ\delta\phi^{\prime} and their first time derivatives. We plug these relations and Eqs. (120)-(121) into Eq. (116). After the integration by parts, the resulting second-order action is expressed in the form,

𝒮even(2)=l,mdtdr(𝒳˙t𝑲𝒳˙+𝒳t𝑮𝒳+𝒳t𝑸𝒳+𝒳t𝑴𝒳),{\cal S}_{\rm even}^{(2)}=\sum_{l,m}\int{\rm d}t{\rm d}r\left(\dot{\vec{\mathcal{X}}}^{t}{\bm{K}}\dot{\vec{\mathcal{X}}}+\vec{\mathcal{X}}^{{}^{\prime}t}{\bm{G}}\vec{\mathcal{X}}^{{}^{\prime}}+\vec{\mathcal{X}}^{t}{\bm{Q}}\vec{\mathcal{X}}^{{}^{\prime}}+\vec{\mathcal{X}}^{t}{\bm{M}}\vec{\mathcal{X}}\right)\,, (122)

where 𝑲{\bm{K}}, 𝑮{\bm{G}}, 𝑸{\bm{Q}}, 𝑴{\bm{M}} are 3×33\times 3 matrices, with

𝒳t=(δρm,ψ,δϕ).\vec{\mathcal{X}}^{t}=\left(\delta\rho_{m},\psi,\delta\phi\right)\,. (123)

To derive no-ghost conditions, we only resort to relations among the coefficients presented in the second equalities of Eq. (A.1) in the Appendix and define the following quantities,

μ4a3fh,2a4fh,𝒫1hμ2fr22(fr44μ2h),𝒫2h(2rff)μ,\mu\equiv-\frac{4a_{3}}{\sqrt{fh}}\,,\qquad{\cal H}\equiv\frac{2a_{4}}{\sqrt{fh}}\,,\qquad{\cal P}_{1}\equiv\frac{h\mu}{2fr^{2}{\cal H}^{2}}\left(\frac{fr^{4}{\cal H}^{4}}{\mu^{2}h}\right)^{\prime}\,,\qquad{\cal P}_{2}\equiv-h\left(2-\frac{rf^{\prime}}{f}\right)\mu\,, (124)

where {\cal H}, 𝒫1{\cal P}_{1}, and 𝒫2{\cal P}_{2} are the same as those introduced in Ref. Kobayashi:2014wsa . It is convenient to notice the following relation,

𝒫2=r(r+)𝒫22rμμ𝒫1𝒫2r224μrhff1,{\cal P}_{2}^{\prime}=\frac{r(r{\cal H}^{\prime}+{\cal H}){\cal H}{\cal P}_{2}-2r\mu{\cal H}{\cal F}-\mu{\cal P}_{1}{\cal P}_{2}}{r^{2}{\cal H}^{2}}-\frac{4\mu r}{{\cal H}}\sqrt{\frac{h}{f}}f_{1}\,, (125)

where we recall that =2G4{\cal F}=2G_{4} for the theories (11).

The ghost is absent under the following three conditions,

K11>0,K11K22K12K21>0,det𝑲>0.K_{11}>0\,,\qquad K_{11}K_{22}-K_{12}K_{21}>0\,,\qquad{\rm det}\,{\bm{K}}>0\,. (126)

The first condition corresponds to

K11=(2rL+𝒫2)2r42fh(ρ+P)L[2rL+𝒫22(ρ+P)r3]2>0,K_{11}=\frac{(2r{\cal H}L+{\cal P}_{2})^{2}r^{4}}{2\sqrt{fh}\,(\rho+P)L[2r{\cal H}L+{\cal P}_{2}-2(\rho+P)r^{3}]^{2}}>0\,, (127)

which is satisfied for ρ+P>0\rho+P>0. The second translates to

K11K22K12K21=4(L𝒫1)μ2r4f22L2(ρ+P)[2rL+𝒫22(ρ+P)r3]2>0.K_{11}K_{22}-K_{12}K_{21}=\frac{4(L{\cal P}_{1}-{\cal F})\mu^{2}r^{4}}{f^{2}{\cal H}^{2}L^{2}(\rho+P)[2r{\cal H}L+{\cal P}_{2}-2(\rho+P)r^{3}]^{2}}>0\,. (128)

Provided that ρ+P>0\rho+P>0, the condition (128) holds for L𝒫1>0L{\cal P}_{1}-{\cal F}>0. Finally, the third condition is given by

det𝑲=2r4[h2μ2(L2)(2𝒫1)r2(ρ+P){2r4L(ρ+P)+2hrμ(L2)hμ2(L𝒫1)}](fh)5/2(ρ+P)L22ϕ2[2rL+𝒫22(ρ+P)r3]2>0.{\rm det}\,{\bm{K}}=\frac{2r^{4}[h^{2}\mu^{2}(L-2){\cal F}(2{\cal P}_{1}-{\cal F})-r^{2}(\rho+P)\{{\cal H}^{2}r^{4}L(\rho+P)+2{\cal H}{\cal F}hr\mu(L-2)-h\mu^{2}(L{\cal P}_{1}-{\cal F})\}]}{(fh)^{5/2}(\rho+P)L^{2}{\cal H}^{2}\phi^{\prime 2}[2r{\cal H}L+{\cal P}_{2}-2(\rho+P)r^{3}]^{2}}>0\,. (129)

As long as ρ+P>0\rho+P>0, the condition (129) is satisfied for a positive numerator. When the perfect fluid is absent, the no-ghost condition corresponds to (2𝒫1)>0{\cal F}(2{\cal P}_{1}-{\cal F})>0 Kobayashi:2014wsa . Indeed, this condition can be recovered by taking the limit ρ+P+0\rho+P\to+0 in Eq. (129). Adding the perfect fluid modifies the third no-ghost condition. In the above derivation of no-ghost conditions, we only used the relations among coefficients in the second-order action (116), so the results (127)-(129) are valid even in full Horndeski theories with more general coefficients a1a_{1} etc given in Ref. Kobayashi:2014wsa .

In the limit of large wave number kk, the three propagation speeds crc_{r} along the radial direction in proper time can be obtained by solving

det|fhcr2𝑲+𝑮|=0.{\rm det}\left|fhc_{r}^{2}{\bm{K}}+{\bm{G}}\right|=0\,. (130)

The matrix components G11G_{11}, G12G_{12}, and G13G_{13} of symmetric matrix 𝑮{\bm{G}}, which are related to the matter perturbation δρm\delta\rho_{m}, vanish identically. This is attributed to the fact that the velocity potential vv in Eq. (43) arises from the θ\theta and φ\varphi components of the four velocity uμu_{\mu}. There is no propagation of the matter perturbation in the radial direction, so the corresponding value of cr2c_{r}^{2} yields

cr12=0.c_{r1}^{2}=0\,. (131)

As for the other two radial speeds of propagation, the derivation of their general expressions applicable to full Horndeski theories is not straightforward due to a mixture of the gravitational propagation speed with the perfect-fluid sector. Hence we focus on scalar-tensor theories given by the action (11) in the following. Then, the two propagation speed squares read

cr±2=C22C1[1±14C1C3C22],c_{r\pm}^{2}=\frac{C_{2}}{2C_{1}}\left[-1\pm\sqrt{1-\frac{4C_{1}C_{3}}{C_{2}^{2}}}\right]\,, (132)

where

C1\displaystyle C_{1} =\displaystyle= G2,XG4[4(L2)G4h+Lr2(ρ+P)]+2G4,ϕ2[6(L2)G4h+(2L1)r2(ρ+P)],\displaystyle G_{2,X}G_{4}\left[4(L-2)G_{4}h+Lr^{2}(\rho+P)\right]+2G_{4,\phi}^{2}\left[6(L-2)G_{4}h+(2L-1)r^{2}(\rho+P)\right]\,, (133)
C2\displaystyle C_{2} =\displaystyle= G2,XG4[8(L2)G4h+Lr2(1+cm2)(ρ+P)]2G4,ϕ2[12(L2)G4h+r2(2L1)(1+cm2)(ρ+P)]\displaystyle-G_{2,X}G_{4}\left[8(L-2)G_{4}h+Lr^{2}(1+c_{m}^{2})(\rho+P)\right]-2G_{4,\phi}^{2}\left[12(L-2)G_{4}h+r^{2}(2L-1)(1+c_{m}^{2})(\rho+P)\right] (134)
+G2,XXG4hϕ2[4(L2)G4h+Lr2(ρ+P)],\displaystyle+G_{2,XX}G_{4}h\phi^{\prime 2}\left[4(L-2)G_{4}h+Lr^{2}(\rho+P)\right]\,,
C3\displaystyle C_{3} =\displaystyle= G4[4(L2)G4h+cm2Lr2(ρ+P)](G2,XG2,XXhϕ2)+G4,ϕ2[12(L2)G4h+2(2L1)cm2r2(ρ+P)].\displaystyle G_{4}\left[4(L-2)G_{4}h+c_{m}^{2}Lr^{2}(\rho+P)\right](G_{2,X}-G_{2,XX}h\phi^{\prime 2})+G_{4,\phi}^{2}\left[12(L-2)G_{4}h+2(2L-1)c_{m}^{2}r^{2}(\rho+P)\right]\,.

If we consider the theories containing only a linear function of XX in G2G_{2}, i.e.,

G2,XX=0,G_{2,XX}=0\,, (136)

then the propagation speed squares (132) reduce to

cr22\displaystyle c_{r2}^{2} =\displaystyle= 1(1cm2)r2[(L2)G4,ϕ2+L(G2,XG4+3G4,ϕ2)](ρ+P)4(L2)G4h(G2,XG4+3G4,ϕ2)+r2[(L2)G4,ϕ2+L(G2,XG4+3G4,ϕ2)](ρ+P),\displaystyle 1-\frac{(1-c_{m}^{2})r^{2}[(L-2)G_{4,\phi}^{2}+L(G_{2,X}G_{4}+3G_{4,\phi}^{2})](\rho+P)}{4(L-2)G_{4}h(G_{2,X}G_{4}+3G_{4,\phi}^{2})+r^{2}[(L-2)G_{4,\phi}^{2}+L(G_{2,X}G_{4}+3G_{4,\phi}^{2})](\rho+P)}\,, (137)
cr32\displaystyle c_{r3}^{2} =\displaystyle= 1.\displaystyle 1\,. (138)

In theories containing nonlinear functions of XX in G2G_{2}, there is the deviation of cr32c_{r3}^{2} from 1 analogous to that of k-essence scalar in Minkowski space-time. Then, cr32c_{r3}^{2} corresponds to the speed of propagation for δϕ\delta\phi, whereas cr22c_{r2}^{2} to that for ψ\psi. The results (137) and (138) are valid for scalar-tensor theories given by the action (1). For ρ+P>0\rho+P>0 and cm21c_{m}^{2}\neq 1, there is the deviation of cr22c_{r2}^{2} from 1. This property is different from that in Horndeski theories without the perfect fluid, in which case the speed of even-parity gravitational perturbation ψ\psi is the same as that of the odd-parity sector Kobayashi:2014wsa .

The propagation speed cΩc_{\Omega} in the angular direction is known by solving

det|l2fcΩ2𝑲+r2𝑴|=0.{\rm det}\left|l^{2}fc_{\Omega}^{2}{\bm{K}}+r^{2}{\bm{M}}\right|=0\,. (139)

The mass matrix 𝑴{\bm{M}} is important only for large ll, so we will take the limit L=l(l+1)1L=l(l+1)\gg 1 in the following discussion. For the theories given by the action (11), the propagation speed squares cΩ12c_{\Omega 1}^{2}, cΩ22c_{\Omega 2}^{2}, and cΩ32c_{\Omega 3}^{2} of the perturbations δρm\delta\rho_{m}, ψ\psi, and δϕ\delta\phi are given, respectively, by

cΩ12\displaystyle c_{\Omega 1}^{2} =\displaystyle= cm2,\displaystyle c_{m}^{2}\,, (140)
cΩ22\displaystyle c_{\Omega 2}^{2} =\displaystyle= 1r2(G2,XG4+4G4,ϕ2)(ρ+P)4hG2,XG42+[G2,Xr2(ρ+P)+12hG4,ϕ2]G4+4G4,ϕ2r2(ρ+P),\displaystyle 1-\frac{r^{2}(G_{2,X}G_{4}+4G_{4,\phi}^{2})(\rho+P)}{4hG_{2,X}G_{4}^{2}+[G_{2,X}r^{2}(\rho+P)+12hG_{4,\phi}^{2}]G_{4}+4G_{4,\phi}^{2}r^{2}(\rho+P)}\,, (141)
cΩ32\displaystyle c_{\Omega 3}^{2} =\displaystyle= 1.\displaystyle 1\,. (142)

The speed of propagation in the gravity sector is affected by the perfect fluid.

From the above discussions, the Laplacian stabilities of even-mode perturbations along the radial and angular directions are absent under the conditions cr±20c_{r\pm}^{2}\geq 0 and cΩ220c_{\Omega 2}^{2}\geq 0 with cm20c_{m}^{2}\geq 0.

VI.2 l=0l=0

For the monopole mode l=0l=0, the perturbations β\beta, α\alpha, and GG vanish identically. We choose the gauge K=0K=0 to fix the radial transformation scalar {\cal R} in ξr\xi_{r}. The second-order Lagrangian for l=0l=0 can be derived by setting L=0L=0 and α=0\alpha=0 in Eq. (116), such that

\displaystyle{\cal L} =\displaystyle= H0[(a1δϕf2b3δϕf2b4H2)r22f1H2+a10δρ]+b2a1H1(a1δϕf2b3δϕf2b4H2)\displaystyle H_{0}\left[\left(a_{1}\delta\phi^{\prime}-\frac{f}{2}b_{3}\delta\phi-\frac{f}{2}b_{4}H_{2}\right)^{\prime}-\frac{r^{2}}{2}f_{1}H_{2}+a_{10}\delta\rho\right]+\frac{b_{2}}{a_{1}}H_{1}\left(a_{1}\delta\phi^{\prime}-\frac{f}{2}b_{3}\delta\phi-\frac{f}{2}b_{4}H_{2}\right)^{\cdot} (143)
+c1δϕ˙H˙2+H2(c2δϕ+c3δϕ+c~5v˙)+c6H22+e1δϕ˙2+e2δϕ2+e3δϕ2+f2δρ2+f3δρv˙.\displaystyle+c_{1}\dot{\delta\phi}\dot{H}_{2}+H_{2}\left(c_{2}\delta\phi^{\prime}+c_{3}\delta\phi+\tilde{c}_{5}\dot{v}\right)+c_{6}H_{2}^{2}+e_{1}\dot{\delta\phi}^{2}+e_{2}\delta\phi^{\prime 2}+e_{3}\delta\phi^{2}+f_{2}\delta\rho^{2}+f_{3}\delta\rho\,\dot{v}\,.

In the following we choose the gauge H0=0H_{0}=0. In this case, there is a gauge mode 𝒞(r){\cal C}(r) in the temporal transformation scalar 𝒯{\cal T}. This appears as the gauge mode 𝒞(f/f)𝒞{\cal C}^{\prime}-(f^{\prime}/f){\cal C} in the gauge transformation of H1H_{1}, see Eq. (61). Varying the action (143) with respect to H1H_{1}, we have

a1δϕf2b3δϕf2b4H2=𝒟(r),a_{1}\delta\phi^{\prime}-\frac{f}{2}b_{3}\delta\phi-\frac{f}{2}b_{4}H_{2}={\cal D}(r)\,, (144)

where 𝒟(r){\cal D}(r) is an arbitrary function of rr. The gauge mode 𝒞(f/f)𝒞{\cal C}^{\prime}-(f^{\prime}/f){\cal C} in H1H_{1} can be eliminated by properly choosing the rr-dependent function 𝒟(r){\cal D}(r). This rr-dependent function depends on the background alone, so it does not affect the dynamics of perturbations Motohashi:2011pw ; Kobayashi:2014wsa . Hence we drop such contributions to the second-order action of perturbations in the following. We solve Eq. (144) for H2H_{2} and take the time derivative of H2H_{2}. Substituting H2H_{2} and H˙2\dot{H}_{2} into Eq. (143) and integrating it by parts, the action contains the two dynamical fields,

𝒳t=(v,δϕ).\vec{\mathcal{X}}^{t}=\left(v,\delta\phi\right)\,. (145)

After the integration by parts, the reduced action is expressed in the form,

𝒮even(2)=l,mdtdr(𝒳˙t𝑲𝒳˙+𝒳t𝑮𝒳+𝒳t𝑴𝒳+𝒳˙t𝑹𝒳+𝒳˙t𝑻𝒳),{\cal S}_{\rm even}^{(2)}=\sum_{l,m}\int{\rm d}t{\rm d}r\left(\dot{\vec{\mathcal{X}}}^{t}{\bm{K}}\dot{\vec{\mathcal{X}}}+\vec{\mathcal{X}}^{\prime t}{\bm{G}}\vec{\mathcal{X}}^{\prime}+\vec{\mathcal{X}}^{t}{\bm{M}}\vec{\mathcal{X}}+\dot{\vec{\mathcal{X}}}^{t}{\bm{R}}\vec{\mathcal{X}}^{\prime}+\dot{\vec{\mathcal{X}}}^{t}{\bm{T}}\vec{\mathcal{X}}\right)\,, (146)

where the nonvanishing components of 2×22\times 2 matrices 𝑲{\bm{K}}, 𝑮{\bm{G}}, 𝑴{\bm{M}}, 𝑹{\bm{R}}, and 𝑻{\bm{T}} are

K11\displaystyle K_{11} =\displaystyle= (ρ+P)r22fhcm2,\displaystyle\frac{(\rho+P)r^{2}}{2\sqrt{fh}\,c_{m}^{2}}\,, (147)
K22\displaystyle K_{22} =\displaystyle= 1fhϕ2[2𝒫1+r2(ρ+P)(μ4r)2hμ],\displaystyle\frac{1}{\sqrt{fh}\phi^{\prime 2}}\left[2{\cal P}_{1}-{\cal F}+\frac{r^{2}(\rho+P)(\mu-4r{\cal H})}{2h\mu}\right]\,, (148)
G22\displaystyle G_{22} =\displaystyle= a32e2a1a3c2+a12c6a32,\displaystyle\frac{a_{3}^{2}e_{2}-a_{1}a_{3}c_{2}+a_{1}^{2}c_{6}}{a_{3}^{2}}\,, (149)
M22\displaystyle M_{22} =\displaystyle= a32e3+(a1a2)[(a1a2)c6+a3c3]a32+[a1a3c3+(a1a2)(2a1c6a3c2)2a32],\displaystyle\frac{a_{3}^{2}e_{3}+(a_{1}^{\prime}-a_{2})[(a_{1}^{\prime}-a_{2})c_{6}+a_{3}c_{3}]}{a_{3}^{2}}+\left[\frac{a_{1}a_{3}c_{3}+(a_{1}^{\prime}-a_{2})(2a_{1}c_{6}-a_{3}c_{2})}{2a_{3}^{2}}\right]^{\prime}\,, (150)
R12\displaystyle R_{12} =\displaystyle= R21=f1a1r22fa3,\displaystyle R_{21}=-\frac{f_{1}a_{1}r^{2}}{2\sqrt{f}\,a_{3}}\,, (151)
T12\displaystyle T_{12} =\displaystyle= T21=r8f3/2a32[a1(2ff1ra34a3ff12a3frf1+a3frf1)+2a3ff1r(2a23a1)].\displaystyle-T_{21}=-\frac{r}{8f^{3/2}a_{3}^{2}}\left[a_{1}(2ff_{1}ra_{3}^{\prime}-4a_{3}ff_{1}-2a_{3}frf_{1}^{\prime}+a_{3}f^{\prime}rf_{1})+2a_{3}ff_{1}r(2a_{2}-3a_{1}^{\prime})\right]\,. (152)

We note that 𝑹{\bm{R}} and 𝑻{\bm{T}} are symmetric and anti-symmetric matrices, respectively.

The ghosts are absent under the two conditions,

K11>0,\displaystyle K_{11}>0\,, (153)
K22>0.\displaystyle K_{22}>0\,. (154)

where the former is satisfied for ρ+P>0\rho+P>0 and cm2>0c_{m}^{2}>0. On using 𝒳t=ei(ωtkr)\vec{\mathcal{X}}^{t}=e^{i(\omega t-kr)} as a solution in the radial direction, the dispersion relation for large ω\omega and kk yields

det(ω2𝑲ωk𝑹+k2𝑮)=0.{\rm det}(\omega^{2}{\bm{K}}-\omega k{\bm{R}}+k^{2}{\bm{G}})=0\,. (155)

The propagation speed crc_{r} in proper time can be obtained by substituting ω=fhcrk\omega=\sqrt{fh}c_{r}k into Eq. (155). The resulting two solutions are given by

(cr12)l=0=0,\displaystyle\left(c_{r1}^{2}\right)_{l=0}=0\,, (156)
(cr22)l=0=1fhK22(G22R122K11)=14h2ϕ2G2,XXG42+(1cm2)(ρ+P)r2G4,ϕ24hG2,XG42+[12hG4+(ρ+P)r2]G4,ϕ2.\displaystyle\left(c_{r2}^{2}\right)_{l=0}=-\frac{1}{fhK_{22}}\left(G_{22}-\frac{R_{12}^{2}}{K_{11}}\right)=1-\frac{4h^{2}\phi^{\prime 2}G_{2,XX}G_{4}^{2}+(1-c_{m}^{2})(\rho+P)r^{2}G_{4,\phi}^{2}}{4hG_{2,X}G_{4}^{2}+[12hG_{4}+(\rho+P)r^{2}]G_{4,\phi}^{2}}\,. (157)

There is no radial propagation in the perfect-fluid sector, but the scalar perturbation δϕ\delta\phi propagates with the speed (cr2)l=0\left(c_{r2}\right)_{l=0}. The Laplacian instability can be avoided for (cr22)l=00\left(c_{r2}^{2}\right)_{l=0}\geq 0. Unless l1l\gg 1, the kinetic term 𝒳˙t𝑲𝒳˙\dot{\vec{\mathcal{X}}}^{t}{\bm{K}}\dot{\vec{\mathcal{X}}} dominates over the mass term 𝒳t𝑴𝒳\vec{\mathcal{X}}^{t}{\bm{M}}\vec{\mathcal{X}} for large ω\omega, so we do not consider the propagation along the angular direction.

VI.3 l=1l=1

For the dipole mode l=1l=1, the dependence of metric perturbations habh_{ab} occurs through the combination KGK-G. After fixing the gauge to be G=0G=0 and β=0\beta=0, we can set K=0K=0. Since the latter does not correspond to the gauge fixing, we will choose the gauge δϕ=0\delta\phi=0 to fix {\cal R}. Then, we can simply set δϕ=0\delta\phi=0 and L=2L=2 in the second-order action (116) and define the dynamical variables,

ψa3H2+La4α,δρmδρ+2fhr3f1f3[ha3(2ffr)+Lfra4]ψ.\psi\equiv a_{3}H_{2}+La_{4}\alpha\,,\qquad\delta\rho_{m}\equiv\delta\rho+\frac{2\sqrt{f}\,hr^{3}f_{1}}{f_{3}[ha_{3}(2f-f^{\prime}r)+Lfra_{4}]}\psi^{\prime}\,. (158)

We follow the similar procedure to that taken in Eqs. (119)-(121) and eliminate the nondynamical variables H0H_{0}, α\alpha, H1H_{1}, and vv. After the integration by parts, the resulting second-order action is of the form (122) with the two dynamical perturbations,

𝒳t=(δρm,ψ).\vec{\mathcal{X}}^{t}=\left(\delta\rho_{m},\psi\right)\,. (159)

The no-ghost conditions, which are determined by the 2×22\times 2 matrix 𝑲{\bm{K}}, are

K11\displaystyle K_{11} =\displaystyle= (4r+𝒫2)2r44fh(ρ+P)[4r+𝒫22(ρ+P)r3]2>0,\displaystyle\frac{(4r{\cal H}+{\cal P}_{2})^{2}r^{4}}{4\sqrt{fh}\,(\rho+P)[4r{\cal H}+{\cal P}_{2}-2(\rho+P)r^{3}]^{2}}>0\,, (160)
det𝑲\displaystyle{\rm det}\,{\bm{K}} =\displaystyle= (2𝒫1)μ2r4f22(ρ+P)[4r+𝒫22(ρ+P)r3]2>0.\displaystyle\frac{(2{\cal P}_{1}-{\cal F})\mu^{2}r^{4}}{f^{2}{\cal H}^{2}(\rho+P)[4r{\cal H}+{\cal P}_{2}-2(\rho+P)r^{3}]^{2}}>0\,. (161)

Taking the limit L2L\to 2, these results coincide with Eqs. (127) and (128), respectively.

The propagation speeds along the radial direction are known by solving Eq. (130) for the 2×22\times 2 matrices 𝑲{\bm{K}} and 𝑮{\bm{G}}. Since the nonvanishing component of 𝑮{\bm{G}} is G11G_{11} alone, the propagation speed squared associated with δρm\delta\rho_{m} is

(cr12)l=1=0.\left(c_{r1}^{2}\right)_{l=1}=0\,. (162)

The other solution, which corresponds to the propagation of ψ\psi, is given by

(cr22)l=1=2r2[fr22cm2(ρ+P)h(4μc5+82c6+μ2d4)]fh(2𝒫1)μ2.\left(c_{r2}^{2}\right)_{l=1}=\frac{2r^{2}[\sqrt{f}\,r^{2}{\cal H}^{2}c_{m}^{2}(\rho+P)-\sqrt{h}(4{\cal H}\mu c_{5}+8{\cal H}^{2}c_{6}+\mu^{2}d_{4})]}{\sqrt{f}h(2{\cal P}_{1}-{\cal F})\mu^{2}}\,. (163)

For the theories given by the action (11), the Laplacian instability is absent under the condition,

(cr22)l=1=1G4[(1cm2)(ρ+P)+h2ϕ4G2,XX]G4(ρ+P+hϕ2G2,X)+3hϕ2G4,ϕ20.\left(c_{r2}^{2}\right)_{l=1}=1-\frac{G_{4}[(1-c_{m}^{2})(\rho+P)+h^{2}\phi^{\prime 4}G_{2,XX}]}{G_{4}(\rho+P+h\phi^{\prime 2}G_{2,X})+3h\phi^{\prime 2}G_{4,\phi}^{2}}\geq 0\,. (164)

This is not recovered by taking the limit L2L\to 2 in Eq. (132), so it gives the additional stability condition to that for l2l\geq 2.

VII Stability of relativistic stars in concrete theories

We study the stability of relativistic stars in scalar-tensor theories by using the results derived in Secs. V and VI. In doing so, we first summarize conditions for the absence of ghosts and Laplacian instabilities in the general theories (11) and apply them to specific theories discussed in Sec. II.

First of all, the ghost in the odd-parity sector is absent under the condition (89), i.e.,

G4>0.G_{4}>0\,. (165)

For even-mode perturbations, the no-ghost conditions (127), (153), and (160), which correspond to stabilities in the matter sector for the modes l2l\geq 2, l=0l=0, and l=1l=1 respectively, are satisfied for

ρ+P>0,cm2>0.\rho+P>0\,,\qquad c_{m}^{2}>0\,. (166)

In the following, we will consider relativistic stars composed by baryonic matter obeying the inequalities (166). The other no-ghost condition for l=1l=1, i.e., Eq. (161), is the special case of Eq. (128), so we do not need to consider the former. Then, the remaining no-ghost conditions are given by Eqs. (128), (129), and (154). Under the inequalities (165) and (166), they translate, respectively, to

𝒦1\displaystyle{\cal K}_{1} \displaystyle\equiv (L2)h(2G4+G4,ϕrϕ)2+Lr2[G4(ρ+P)+κhϕ2]>0,\displaystyle(L-2)h(2G_{4}+G_{4,\phi}r\phi^{\prime})^{2}+Lr^{2}[G_{4}(\rho+P)+\kappa h\phi^{\prime 2}]>0\,, (167)
𝒦2\displaystyle{\cal K}_{2} \displaystyle\equiv (L2)[4κG4h+G4,ϕ2r2(ρ+P)]+Lκr2(ρ+P)>0,\displaystyle(L-2)[4\kappa G_{4}h+G_{4,\phi}^{2}r^{2}(\rho+P)]+L\kappa r^{2}(\rho+P)>0\,, (168)
𝒦3\displaystyle{\cal K}_{3} \displaystyle\equiv 4κG4h+G4,ϕ2r2(ρ+P)>0,\displaystyle 4\kappa G_{4}h+G_{4,\phi}^{2}r^{2}(\rho+P)>0\,, (169)

where L>2L>2 in Eqs. (167) and (168), and we defined

κG2,XG4+3G4,ϕ2.\kappa\equiv G_{2,X}G_{4}+3G_{4,\phi}^{2}\,. (170)

For l2l\geq 2, the propagation speeds of odd-mode perturbations are equivalent to 1. The stabilities of even-mode perturbations are ensured as long as the speeds of propagation given in Eqs. (137), (141), (157), and (164) are nonnegative. In the case with G2,XX=0G_{2,XX}=0, these conditions translate to

cr22\displaystyle c_{r2}^{2} =\displaystyle= 1(1cm2)r2[(L2)G4,ϕ2+Lκ](ρ+P)4(L2)G4hκ+r2[(L2)G4,ϕ2+Lκ](ρ+P)\displaystyle 1-\frac{(1-c_{m}^{2})r^{2}[(L-2)G_{4,\phi}^{2}+L\kappa](\rho+P)}{4(L-2)G_{4}h\kappa+r^{2}[(L-2)G_{4,\phi}^{2}+L\kappa](\rho+P)} (171)
=\displaystyle= 4(L2)G4hκ+cm2r2[(L2)G4,ϕ2+Lκ](ρ+P)4(L2)G4hκ+r2[(L2)G4,ϕ2+Lκ](ρ+P)0,\displaystyle\frac{4(L-2)G_{4}h\kappa+c_{m}^{2}r^{2}[(L-2)G_{4,\phi}^{2}+L\kappa](\rho+P)}{4(L-2)G_{4}h\kappa+r^{2}[(L-2)G_{4,\phi}^{2}+L\kappa](\rho+P)}\geq 0\,,
cΩ22\displaystyle c_{\Omega 2}^{2} =\displaystyle= 1r2(G4,ϕ2+κ)(ρ+P)4G4hκ+r2(G4,ϕ2+κ)(ρ+P)=4G4hκ4G4hκ+r2(G4,ϕ2+κ)(ρ+P)0,\displaystyle 1-\frac{r^{2}(G_{4,\phi}^{2}+\kappa)(\rho+P)}{4G_{4}h\kappa+r^{2}(G_{4,\phi}^{2}+\kappa)(\rho+P)}=\frac{4G_{4}h\kappa}{4G_{4}h\kappa+r^{2}(G_{4,\phi}^{2}+\kappa)(\rho+P)}\geq 0\,, (172)
(cr22)l=0\displaystyle(c_{r2}^{2})_{l=0} =\displaystyle= 1(1cm2)G4,ϕ2r2(ρ+P)4G4hκ+G4,ϕ2r2(ρ+P)=4G4hκ+cm2G4,ϕ2r2(ρ+P)4G4hκ+G4,ϕ2r2(ρ+P)0,\displaystyle 1-\frac{(1-c_{m}^{2})G_{4,\phi}^{2}r^{2}(\rho+P)}{4G_{4}h\kappa+G_{4,\phi}^{2}r^{2}(\rho+P)}=\frac{4G_{4}h\kappa+c_{m}^{2}G_{4,\phi}^{2}r^{2}(\rho+P)}{4G_{4}h\kappa+G_{4,\phi}^{2}r^{2}(\rho+P)}\geq 0\,, (173)
(cr22)l=1\displaystyle(c_{r2}^{2})_{l=1} =\displaystyle= 1(1cm2)G4(ρ+P)G4(ρ+P)+hκϕ2=cm2G4(ρ+P)+hκϕ2G4(ρ+P)+hκϕ20.\displaystyle 1-\frac{(1-c_{m}^{2})G_{4}(\rho+P)}{G_{4}(\rho+P)+h\kappa\phi^{\prime 2}}=\frac{c_{m}^{2}G_{4}(\rho+P)+h\kappa\phi^{\prime 2}}{G_{4}(\rho+P)+h\kappa\phi^{\prime 2}}\geq 0\,. (174)

Now, we discuss stability conditions in two concrete theories presented in Sec. II.

VII.1 Theories of spontaneous scalarization

The action in theories of spontaneous scalarization is given by Eq. (4), i.e.,

G4(ϕ)=Mpl22F(ϕ),G2(ϕ,X)=(13Mpl2F,ϕ22F2)F(ϕ)X.G_{4}(\phi)=\frac{M_{\rm pl}^{2}}{2}F(\phi)\,,\qquad G_{2}(\phi,X)=\left(1-\frac{3M_{\rm pl}^{2}F_{,\phi}^{2}}{2F^{2}}\right)F(\phi)X\,. (175)

In this case, the quantity κ\kappa reduces to

κ=Mpl22F2(ϕ)>0.\kappa=\frac{M_{\rm pl}^{2}}{2}F^{2}(\phi)>0\,. (176)

The positivity of κ\kappa means that, along with the conditions (165) and (166), the absence of ghost and Laplacian instabilities is manifestly guaranteed. This is the case for the coupling (5) chosen by Damour and Esposito-Farese, where G4G_{4} is positive. In addition, as long as cm2c_{m}^{2} is in the range 0<cm210<c_{m}^{2}\leq 1, all the propagation speed squares computed above are subluminal. Since all the conditions (167)-(169) and (171)-(174) are irrelevant to the potential, a massive scalar field coupled to matter with positive coupling F>0F>0 investigated in Refs. Chen ; Morisaki has neither ghost nor Laplacian instabilities either.

Refer to caption
Refer to caption
Figure 1: (Left) The mass MM of NS normalized by the solar mass MM_{\odot} versus the star radius rsr_{s} in theories of spontaneous scalarization. We choose the Damour and Esposito-Farese nonminimal coupling (5) with β=6\beta=-6. We consider the SLy equation of state inside the star. For increasing central density ρc\rho_{c}, the mass-radius relation moves in the direction shown as an arrow. The mass-radius relation in GR is plotted as a thin dashed line. (Right) The radial propagation speed squared cr22c_{r2}^{2} for l=2l=2 versus the distance rr normalized by the star radius rsr_{s}. The plots represented as (a), (b), (c) correspond to the cases shown as the same labels in the left panel. Spontaneous scalarization occurs in all these three cases with the subluminal values of cr22c_{r2}^{2} inside the star.

In the left panel of Fig. 1, we plot the mass-radius relation of relativistic star for the nonminimal coupling (5) with β=6\beta=-6. We choose the SLy equation of state inside the star, whose analytic representation is given in Ref. Haensel:2004nu . For each central density ρc\rho_{c}, the boundary condition of ϕ\phi at r=0r=0 is iteratively searched to realize the asymptotic behavior ϕ(r)0\phi(r)\to 0 as rr\to\infty. For the central density 4.3ρ0<ρc<14.4ρ04.3\rho_{0}<\rho_{c}<14.4\rho_{0}, where ρ0=1.6749×1014gcm3\rho_{0}=1.6749\times 10^{14}~{}{\rm g}\cdot{\rm cm}^{-3}, we find that there exists the scalarized branch with ϕ(0)0\phi(0)\neq 0 besides the GR branch with ϕ(r)=0\phi(r)=0 everywhere. As the central density increases, the mass-radius relation shifts toward the direction of an arrow depicted in Fig. 1.

Three cases (a), (b), (c) shown in the left panel of Fig. 1 are in the region where spontaneous scalarization takes place. In the right panel, we plot the propagation speed squared (171) versus r/rsr/r_{s} for l=2l=2 in three different cases (a), (b), (c). For β=6\beta=-6, we find that the matter sound speed squared is subluminal (0<cm210<c_{m}^{2}\leq 1) in the region where spontaneous scalarization occurs. In cases (a), (b), (c) of Fig. 1, we have cm2=0.56,0.74,0.88c_{m}^{2}=0.56,0.74,0.88 at the center of NS, respectively. The superluminal propagation of cm2c_{m}^{2} arises for the high central density ρc18ρ0\rho_{c}\gtrsim 18\rho_{0}, but this is the region in which only the GR branch is present. In the right panel of Fig. 1, we can confirm that cr22c_{r2}^{2} is subluminal inside the star for the scalarized branch. From Eq. (171) the value of cr22c_{r2}^{2} at r=0r=0 is equivalent to 1. For increasing rr from the center, cr22c_{r2}^{2} decreases from 11. Around the surface of star, both ρ\rho and PP rapidly drop down toward 0, so cr22c_{r2}^{2} begins to increase toward the value 1. Besides cases (a), (b), (c), we numerically confirmed that the above subluminal property generally holds throughout the region in which spontaneous scalarization takes place.

From Eq. (174), we have (cr22)l=1=cm2(c_{r2}^{2})_{l=1}=c_{m}^{2} at r=0r=0 due to the boundary condition ϕ(0)=0\phi^{\prime}(0)=0. Provided that 0<cm210<c_{m}^{2}\leq 1, (cr22)l=1(c_{r2}^{2})_{l=1} also remains subluminal inside the star and approaches the value 1 toward the surface. Taking the limit ρ,P0\rho,P\to 0 in Eqs. (171)-(174), the propagation speed squares cr22c_{r2}^{2}, cΩ22c_{\Omega 2}^{2}, (cr22)l=0\left(c_{r2}^{2}\right)_{l=0}, and (cr22)l=1\left(c_{r2}^{2}\right)_{l=1} are all equivalent to 1 outside the star. This fact is consistent with the observations of speed of GWs GW170817 .

VII.2 Brans-Dicke theories

Let us proceed to the BD theories given by the action (7), i.e.,

G4(ϕ)=Mpl22e2Qϕ/Mpl,G2(ϕ,X)=(16Q2)e2Qϕ/MplXV(ϕ).G_{4}(\phi)=\frac{M_{\rm pl}^{2}}{2}e^{-2Q\phi/M_{\rm pl}}\,,\qquad G_{2}(\phi,X)=\left(1-6Q^{2}\right)e^{-2Q\phi/M_{\rm pl}}X-V(\phi)\,. (177)

We are considering the coupling Q2>0Q^{2}>0, i.e., the BD parameter in the range ωBD>3/2\omega_{\rm BD}>-3/2. Then, it follows that

κ=Mpl22e4Qϕ/Mpl>0.\kappa={M_{\rm pl}^{2}\over 2}e^{-4Q\phi/M_{\rm pl}}>0\,. (178)

Therefore, together with the conditions (166), the absence of ghost and Laplacian instabilities is automatically guaranteed. As long as 0<cm210<c_{m}^{2}\leq 1, the propagation speed squares (171)-(174) are subluminal inside the star.

VIII Conclusions

In this paper, we studied the stability of relativistic stars against odd- and even-parity perturbations in scalar-tensor theories given by the action (11). Our interest is the application to hairy NS solutions which are known to exist in theories of spontaneous scalarization and BD theories (see Sec. II). For this purpose, we need to properly deal with the matter sector as a form of the perfect fluid. The Schutz-Sorkin action (12) is suitable for describing the perfect fluid in any background space-time. In Sec. III, we derived covariant equations of motion in scalar-tensor theories (11) with the matter action (12) and applied them to the spherically symmetric and static background.

In Sec. IV, we decomposed the perturbations of gravity, scalar-field, and perfect-fluid sectors into the odd- and even-parity modes. To our knowledge, this type of decomposition including the perfect fluid as a form of the Schutz-Sorkin action was not addressed in the literature. We also defined the matter density perturbation δρ\delta\rho and velocity potential vv as the quantities related to the number density nn and four velocity uμ=Jμ/(ng)u_{\mu}=J_{\mu}/(n\sqrt{-g}), respectively. The radial direction is singled out on the spherically symmetric and static background, in which case the velocity potential is associated with the θ\theta and φ\varphi components of uμu_{\mu}.

In Sec. V, we expanded the action (11) up to quadratic order in odd-parity perturbations and obtained the second-order action of the form (81). The perfect fluid does not affect the evolution of GWs in the odd-parity sector. For the multipoles l2l\geq 2, there is one dynamical degree of freedom with the propagation speed equivalent to that of light. In this case, the ghost is absent under the condition G4>0G_{4}>0. For l=1l=1, there is no dynamical propagation of odd-parity perturbations.

In Sec. VI, we obtained the second-order action of even-parity perturbations with the multipoles l2l\geq 2 in the form (116) after eliminating some nondynamical perturbations appearing in the Schutz-Sorkin action. We found that there are three propagating degrees of freedom characterized by δρm\delta\rho_{m}, ψ\psi, and δϕ\delta\phi, where δρm\delta\rho_{m} and ψ\psi are given, respectively, by Eqs. (118) and (117). After integrating out all the other nondynamical perturbations, the final second-order action reduces to the form (122) with (123). From the kinetic matrix 𝑲{\bm{K}}, we showed that the ghosts are absent under the three conditions (127), (128), and (129). Since we only exploited the relations among coefficients in Eq. (116) applicable to full Horndeski theories, our no-ghost conditions are also valid for Horndeski theories by modifying the coefficients (A.1) in the Appendix to those presented in Ref. Kobayashi:2014wsa .

The matter perturbation δρm\delta\rho_{m} associated with the even-parity sector does not propagate along the radial direction by reflecting the property of velocity potential mentioned above. In scalar-tensor theories given by the action (11), the other propagation speed squares are given by Eq. (132). If G2,XX=0G_{2,XX}=0, which is the case for theories of scalarized relativistic stars discussed in Sec. II, the radial speeds of propagation associated with the perturbations ψ\psi and δϕ\delta\phi reduce, respectively, to Eqs. (137) and (138). For cm21c_{m}^{2}\neq 1, the speed cr2c_{r2} of scalar GWs is different from 1 inside relativistic stars. In the limit l1l\gg 1, we also derived the three propagation speeds along the angular direction as Eqs. (140)-(142). Again, the perfect fluid affects the angular propagation of scalar GWs.

For the monopole mode (l=0l=0) in the even-parity sector, we showed that there are two dynamical perturbations vv and δϕ\delta\phi with the reduced action of the form (146). In this case, the no-ghost conditions are given by Eqs. (153) and (154) with the two radial propagation speeds (156) and (157), so that the latter propagation of scalar-field perturbation δϕ\delta\phi is modified by the presence of perfect fluid. For the dipole mode (l=1l=1), the dynamical perturbations correspond to δρm\delta\rho_{m} and ψ\psi defined by Eq. (158). In this case, the no-ghost conditions correspond to the l1l\to 1 limit of those derived for l2l\geq 2. However, the speed of scalar GWs cannot be recovered in the same limit, so it gives an additional stability condition to those obtained for l2l\geq 2.

In Sec. VII, we summarized the stability conditions for the absence of ghost and Laplacian instabilities and applied them to the theories with G2,XX=0G_{2,XX}=0. Provided that the ghost is absent in the odd-parity sector (G4>0G_{4}>0) and that the perfect fluid satisfies the properties ρ+P>0\rho+P>0 and cm2>0c_{m}^{2}>0, the sign of κ\kappa defined by Eq. (170) is crucial for the stability of even-parity perturbations. In theories of spontaneous scalarization and BD theories with ωBD>3/2\omega_{\rm BD}>-3/2, we showed that κ\kappa is positive, under which there are neither ghosts nor Laplacian instabilities. Moreover, as long as 0<cm210<c_{m}^{2}\leq 1, the propagation speeds are subluminal inside the star. Indeed, we confirmed this property in theories of spontaneous scalarization with the nonminimal coupling taken by Damour and Esposito-Farese by numerically computing the radial propagation speed squared cr22c_{r2}^{2} inside the NS. In such theories, the propagation speeds of GWs in both odd- and even-parity sectors are equivalent to that of light outside the star, so they are consistent with observations of the GW170817 event.

We have thus shown that hairy relativistic stars in scalar-tensor theories given by the action (1) are stable against odd- and even-parity perturbations under mild conditions. The next step is to probe the signature of scalar hairs from observations. In addition to the oscillation of scalar GWs discussed in Ref. Sotani:2005qx , the tidal deformations of NS binaries Flanagan:2007ix ; Damour:2009vw ; Binnington:2009bb ; Hinderer:2009ca may allow one to distinguish between NSs in scalar-tensor theories and in GR. Our general formulation of perturbations around relativistic stars will provide a useful framework for dealing with such problems.

Acknowledgements

R. Kase is supported by the Grant-in-Aid for Young Scientists B of the JSPS No. 17K14297. ST is supported by the Grant-in-Aid for Scientific Research Fund of the JSPS No. 19K03854.

Appendix: Coefficients in the second-order action of even-parity perturbation

In this appendix, we summarize the coefficients of the reduced action (116). They are given by

a1\displaystyle a_{1} =\displaystyle= r2fhG4,ϕ,\displaystyle r^{2}\sqrt{fh}\,G_{4,\phi}\,,
a2\displaystyle a_{2} =\displaystyle= r2fh[hrϕ(G2,X+4G4,ϕϕ)+G4,ϕ(rh+4h)]=fh(a1fh)(ϕ′′ϕf2f)a1+rϕ(ffhh)a4+r2ϕf1,\displaystyle\frac{r}{2}\sqrt{\frac{f}{h}}\left[hr\phi^{\prime}\left(G_{2,X}+4G_{4,\phi\phi}\right)+G_{4,\phi}(rh^{\prime}+4h)\right]=\sqrt{fh}\left(\frac{a_{1}}{\sqrt{fh}}\right)^{\prime}-\left(\frac{\phi^{\prime\prime}}{\phi^{\prime}}-\frac{f^{\prime}}{2f}\right)a_{1}+\frac{r}{\phi^{\prime}}\left(\frac{f^{\prime}}{f}-\frac{h^{\prime}}{h}\right)a_{4}+\frac{r^{2}}{\phi^{\prime}}f_{1},
a3\displaystyle a_{3} =\displaystyle= rfh2(rϕG4,ϕ+2G4)=ϕ2a1ra4,a4=fhG4,\displaystyle-\frac{r\sqrt{fh}}{2}\left(r\phi^{\prime}G_{4,\phi}+2G_{4}\right)=-\frac{\phi^{\prime}}{2}a_{1}-ra_{4}\,,\qquad a_{4}=\sqrt{fh}\,G_{4}\,,
a5\displaystyle a_{5} =\displaystyle= 12fh[2G4,ϕ(h1)+2r(hG4,ϕ+2hϕG4,ϕϕ)+r2{2(ϕ2G4,ϕϕϕ+ϕ′′G4,ϕϕ)h+hϕG4,ϕϕG2,ϕ}]\displaystyle\frac{1}{2}\sqrt{\frac{f}{h}}\left[2G_{4,\phi}(h-1)+2r\left(h^{\prime}G_{4,\phi}+2h\phi^{\prime}G_{4,\phi\phi}\right)+r^{2}\{2(\phi^{\prime 2}G_{4,\phi\phi\phi}+\phi^{\prime\prime}G_{4,\phi\phi})h+h^{\prime}\phi^{\prime}G_{4,\phi\phi}-G_{2,\phi}\}\right]
=\displaystyle= a2a1′′,\displaystyle a_{2}^{\prime}-a_{1}^{\prime\prime}\,,
a6\displaystyle a_{6} =\displaystyle= fhG4,ϕ=12rϕfh[r(2a4fh)+2a4fh],\displaystyle-\sqrt{\frac{f}{h}}G_{4,\phi}=-\frac{1}{2r\phi^{\prime}}\sqrt{\frac{f}{h}}\left[r\left(\frac{2a_{4}}{\sqrt{fh}}\right)^{\prime}+\frac{2a_{4}}{\sqrt{fh}}-{\cal F}\right]\,,
a7\displaystyle a_{7} =\displaystyle= 14fh[4hG4+4r(hG4+2hϕG4,ϕ)+r2{hϕ2(G2,X+4G4,ϕϕ)+2(2hϕ′′+hϕ)G4,ϕ}]=a3r22f1,\displaystyle-\frac{1}{4}\sqrt{\frac{f}{h}}\left[4hG_{4}+4r(h^{\prime}G_{4}+2h\phi^{\prime}G_{4,\phi})+r^{2}\{h\phi^{\prime 2}(G_{2,X}+4G_{4,\phi\phi})+2(2h\phi^{\prime\prime}+h^{\prime}\phi^{\prime})G_{4,\phi}\}\right]=a_{3}^{\prime}-\frac{r^{2}}{2}f_{1},
a8\displaystyle a_{8} =\displaystyle= fhG42=a42h,\displaystyle-\sqrt{\frac{f}{h}}\frac{G_{4}}{2}=-\frac{a_{4}}{2h}\,,
a9\displaystyle a_{9} =\displaystyle= 12rfh[(rh+2h)G4+2hrϕG4,ϕ]=a4+(1rf2f)a4,a10=r22fh,\displaystyle\frac{1}{2r}\sqrt{\frac{f}{h}}\left[(rh^{\prime}+2h)G_{4}+2hr\phi^{\prime}G_{4,\phi}\right]=a_{4}^{\prime}+\left(\frac{1}{r}-\frac{f^{\prime}}{2f}\right)a_{4}\,,\qquad a_{10}=\frac{r^{2}}{2}\sqrt{\frac{f}{h}}\,,
b1\displaystyle b_{1} =\displaystyle= hfG42=a42f,b2=2r2hfG4,ϕ=2fa1,\displaystyle\sqrt{\frac{h}{f}}\frac{G_{4}}{2}=\frac{a_{4}}{2f}\,,\qquad b_{2}=-2r^{2}\sqrt{\frac{h}{f}}\,G_{4,\phi}=-\frac{2}{f}a_{1}\,,
b3\displaystyle b_{3} =\displaystyle= r2hf3/2[ϕf(G2,X+2G4,ϕϕ)fG4,ϕ]=2f(a1a2),b4=rhf(rϕG4,ϕ+2G4)=2fa3,\displaystyle-r^{2}\frac{\sqrt{h}}{f^{3/2}}\left[\phi^{\prime}f\left(G_{2,X}+2G_{4,\phi\phi}\right)-f^{\prime}G_{4,\phi}\right]=\frac{2}{f}\left(a_{1}^{\prime}-a_{2}\right)\,,\qquad b_{4}=r\sqrt{\frac{h}{f}}\left(r\phi^{\prime}G_{4,\phi}+2G_{4}\right)=-\frac{2}{f}a_{3}\,,
b5\displaystyle b_{5} =\displaystyle= hfG4=2b1,\displaystyle-\sqrt{\frac{h}{f}}G_{4}=-2b_{1}\,,
c1\displaystyle c_{1} =\displaystyle= r2fhG4,ϕ=a1fh,c2=r2hf[rϕf(hϕ2G2,XXG2,X)+(rf+4f)G4,ϕ],\displaystyle-\frac{r^{2}}{\sqrt{fh}}G_{4,\phi}=-\frac{a_{1}}{fh}\,,\qquad c_{2}=-\frac{r}{2}\sqrt{\frac{h}{f}}\left[r\phi^{\prime}f(h\phi^{\prime 2}G_{2,XX}-G_{2,X})+(rf^{\prime}+4f)G_{4,\phi}\right]\,,
c3\displaystyle c_{3} =\displaystyle= 12fh[fh(r2ϕ2G2,ϕX4rϕG4,ϕϕ2G4,ϕ)+fr2G2,ϕ+2fG4,ϕhrf(rϕG4,ϕϕ+2G4,ϕ)],\displaystyle\frac{1}{2\sqrt{fh}}\left[fh(r^{2}\phi^{\prime 2}G_{2,\phi X}-4r\phi^{\prime}G_{4,\phi\phi}-2G_{4,\phi})+fr^{2}G_{2,\phi}+2fG_{4,\phi}-hrf^{\prime}(r\phi^{\prime}G_{4,\phi\phi}+2G_{4,\phi})\right]\,,
c4\displaystyle c_{4} =\displaystyle= fhG4,ϕ,c5=12rhf[2frϕG4,ϕ+G4(rf+2f)],c~5=r2(ρ+P)2h,\displaystyle\sqrt{\frac{f}{h}}G_{4,\phi}\,,\qquad c_{5}=-\frac{1}{2r}\sqrt{\frac{h}{f}}\left[2fr\phi^{\prime}G_{4,\phi}+G_{4}(rf^{\prime}+2f)\right]\,,\qquad\tilde{c}_{5}=-\frac{r^{2}(\rho+P)}{2\sqrt{h}}\,,
c6\displaystyle c_{6} =\displaystyle= 18hf[4fG4+4r(fG4+2fϕG4,ϕ)+r2ϕ(fhϕ3G2,XXfϕG2,X+2fG4,ϕ)],\displaystyle\frac{1}{8}\sqrt{\frac{h}{f}}\left[4fG_{4}+4r(f^{\prime}G_{4}+2f\phi^{\prime}G_{4,\phi})+r^{2}\phi^{\prime}(fh\phi^{\prime 3}G_{2,XX}-f\phi^{\prime}G_{2,X}+2f^{\prime}G_{4,\phi})\right]\,,
d1\displaystyle d_{1} =\displaystyle= hfG42=a42f,d2=2fhG4,ϕ=2hc4,d3=fhr[rϕ(G2,X+2G4,ϕϕ)2G4,ϕ],d4=fhr2G4,\displaystyle\sqrt{\frac{h}{f}}\frac{G_{4}}{2}=\frac{a_{4}}{2f}\,,\qquad d_{2}=2\sqrt{fh}\,G_{4,\phi}=2hc_{4}\,,\qquad d_{3}=\frac{\sqrt{fh}}{r}\left[r\phi^{\prime}\left(G_{2,X}+2G_{4,\phi\phi}\right)-2G_{4,\phi}\right]\,,\qquad d_{4}=\frac{\sqrt{fh}}{r^{2}}G_{4},
e1\displaystyle e_{1} =\displaystyle= r22fhG2,X=1ϕfh[(ff+h2h)a1+a22a12rha6],e2=r2fh2(G2,Xhϕ2G2,XX),\displaystyle\frac{r^{2}}{2\sqrt{fh}}G_{2,X}=\frac{1}{\phi^{\prime}fh}\left[\left(\frac{f^{\prime}}{f}+\frac{h^{\prime}}{2h}\right)a_{1}+a_{2}-2a_{1}^{\prime}-2rha_{6}\right]\,,\qquad e_{2}=-\frac{r^{2}\sqrt{fh}}{2}\left(G_{2,X}-h\phi^{\prime 2}G_{2,XX}\right)\,,
e3\displaystyle e_{3} =\displaystyle= r22fhϕϕ,e4=12fhG2,X,\displaystyle\frac{r^{2}}{2}\sqrt{\frac{f}{h}}\frac{\partial{\cal E}_{\phi}}{\partial\phi}\,,\qquad e_{4}=-\frac{1}{2}\sqrt{\frac{f}{h}}G_{2,X}\,,
f1\displaystyle f_{1} =\displaystyle= ρ+P2fh,f2=cm2r22(ρ+P)fh,f3=r2h,\displaystyle-\frac{\rho+P}{2}\sqrt{\frac{f}{h}}\,,\qquad f_{2}=-\frac{c_{m}^{2}r^{2}}{2(\rho+P)}\sqrt{\frac{f}{h}}\,,\qquad f_{3}=-\frac{r^{2}}{\sqrt{h}}\,, (A.1)

where {\cal F} and ϕ{\cal E}_{\phi} are defined by Eqs. (89) and (39) respectively.

References

  • (1) B. P. Abbott et al. [LIGO Scientific and Virgo Collaborations], Phys. Rev. Lett.  116, 061102 (2016) [arXiv:1602.03837 [gr-qc]].
  • (2) B. P. Abbott et al. [LIGO Scientific and Virgo Collaborations], Phys. Rev. Lett.  119, 161101 (2017) [arXiv:1710.05832 [gr-qc]].
  • (3) E. Berti et al., Class. Quant. Grav. 32, 243001 (2015) [arXiv:1501.07274 [gr-qc]].
  • (4) L. Barack et al., Class. Quant. Grav. 36, 143001 (2019) [arXiv:1806.05195 [gr-qc]].
  • (5) Y. Fujii and K. Maeda, “The scalar-tensor theory of gravitation”, Cambridge University Press (2003).
  • (6) A. De Felice and S. Tsujikawa, Living Rev. Rel. 13, 3 (2010) [arXiv:1002.4928 [gr-qc]].
  • (7) T. Damour and G. Esposito-Farese, Phys. Rev. Lett.  70, 2220 (1993).
  • (8) T. Damour and G. Esposito-Farese, Phys. Rev. D 54, 1474 (1996) [gr-qc/9602056].
  • (9) T. Harada, Phys. Rev. D 57, 4802 (1998) [gr-qc/9801049].
  • (10) J. Novak, Phys. Rev. D 58, 064019 (1998) [gr-qc/9806022].
  • (11) H. O. Silva, C. F. B. Macedo, E. Berti and L. C. B. Crispino, Class. Quant. Grav.  32, 145008 (2015) [arXiv:1411.6286 [gr-qc]].
  • (12) P. C. C. Freire et al., Mon. Not. Roy. Astron. Soc.  423, 3328 (2012) [arXiv:1205.1450 [astro-ph.GA]].
  • (13) H. Sotani and K. D. Kokkotas, Phys. Rev. D 70, 084026 (2004) [arXiv:gr-qc/0409066 [gr-qc]].
  • (14) H. Sotani, Phys. Rev. D 89, 064031 (2014) [arXiv:1402.5699 [astro-ph.HE]].
  • (15) H. Sotani, Phys. Rev. D 86, 124036 (2012) [arXiv:1211.6986 [astro-ph.HE]].
  • (16) D. D. Doneva, S. S. Yazadjiev, N. Stergioulas and K. D. Kokkotas, Phys. Rev. D 88, 084060 (2013) [arXiv:1309.0605 [gr-qc]].
  • (17) D. D. Doneva, S. S. Yazadjiev, K. V. Staykov and K. D. Kokkotas, Phys. Rev. D 90, 104021 (2014) [arXiv:1408.1641 [gr-qc]].
  • (18) P. Pani and E. Berti, Phys. Rev. D 90, 024025 (2014) [arXiv:1405.4547 [gr-qc]].
  • (19) D. D. Doneva, S. S. Yazadjiev, N. Stergioulas, K. D. Kokkotas and T. M. Athanasiadis, Phys. Rev. D 90, 044004 (2014) [arXiv:1405.6976 [astro-ph.HE]].
  • (20) G. Pappas and T. P. Sotiriou, Mon. Not. Roy. Astron. Soc. 453, 2862-2876 (2015) [arXiv:1505.02882 [gr-qc]].
  • (21) B. Kleihaus, J. Kunz, S. Mojica and E. Radu, Phys. Rev. D 93, 044047 (2016) [arXiv:1511.05513 [gr-qc]].
  • (22) D. D. Doneva and S. S. Yazadjiev, Phys. Rev. Lett.  120, 131103 (2018) [arXiv:1711.01187 [gr-qc]].
  • (23) H. O. Silva, J. Sakstein, L. Gualtieri, T. P. Sotiriou and E. Berti, Phys. Rev. Lett.  120, 131104 (2018) [arXiv:1711.02080 [gr-qc]].
  • (24) G. Antoniou, A. Bakopoulos and P. Kanti, Phys. Rev. Lett.  120, 131102 (2018) [arXiv:1711.03390 [hep-th]].
  • (25) G. Antoniou, A. Bakopoulos and P. Kanti, Phys. Rev. D 97, 084037 (2018) [arXiv:1711.07431 [hep-th]].
  • (26) M. Minamitsuji and T. Ikeda, Phys. Rev. D 99, 044017 (2019) [arXiv:1812.03551 [gr-qc]].
  • (27) P. V. P. Cunha, C. A. R. Herdeiro and E. Radu, Phys. Rev. Lett.  123, 011101 (2019) [arXiv:1904.09997 [gr-qc]].
  • (28) I. Z. Stefanov, S. S. Yazadjiev and M. D. Todorov, Mod. Phys. Lett. A 23, 2915 (2008) [arXiv:0708.4141 [gr-qc]].
  • (29) C. A. R. Herdeiro, E. Radu, N. Sanchis-Gual and J. A. Font, Phys. Rev. Lett.  121, 101102 (2018) [arXiv:1806.05190 [gr-qc]].
  • (30) P. G. S. Fernandes, C. A. R. Herdeiro, A. M. Pombo, E. Radu and N. Sanchis-Gual, Class. Quant. Grav.  36, no. 13, 134002 (2019) [arXiv:1902.05079 [gr-qc]].
  • (31) P. G. S. Fernandes, C. A. R. Herdeiro, A. M. Pombo, E. Radu and N. Sanchis-Gual, Phys. Rev. D 100, 084045 (2019) [arXiv:1908.00037 [gr-qc]].
  • (32) T. Ikeda, T. Nakamura and M. Minamitsuji, Phys. Rev. D 100, 104014 (2019) [arXiv:1908.09394 [gr-qc]].
  • (33) C. Brans and R. H. Dicke, Phys. Rev.  124, 925 (1961).
  • (34) S. Tsujikawa, K. Uddin, S. Mizuno, R. Tavakol and J. Yokoyama, Phys. Rev. D 77, 103009 (2008) [arXiv:0803.1106 [astro-ph]].
  • (35) A. Cooney, S. DeDeo and D. Psaltis, Phys. Rev. D 82, 064033 (2010) [arXiv:0910.5480 [astro-ph.HE]].
  • (36) A. S. Arapoglu, C. Deliduman and K. Y. Eksi, JCAP 1107, 020 (2011) [arXiv:1003.3179 [gr-qc]].
  • (37) M. Orellana, F. Garcia, F. A. Teppa Pannia and G. E. Romero, Gen. Rel. Grav.  45, 771 (2013) [arXiv:1301.5189 [astro-ph.CO]].
  • (38) A. V. Astashenok, S. Capozziello and S. D. Odintsov, JCAP 1312, 040 (2013) [arXiv:1309.1978 [gr-qc]].
  • (39) S. S. Yazadjiev, D. D. Doneva, K. D. Kokkotas and K. V. Staykov, JCAP 1406, 003 (2014) [arXiv:1402.4469 [gr-qc]].
  • (40) M. Aparicio Resco, A. de la Cruz-Dombriz, F. J. Llanes Estrada and V. Zapatero Castrillo, Phys. Dark Univ.  13, 147 (2016) [arXiv:1602.03880 [gr-qc]].
  • (41) J. O‘Hanlon, Phys. Rev. Lett.  29, 137 (1972).
  • (42) T. Chiba, Phys. Lett. B 575, 1 (2003) [astro-ph/0307338].
  • (43) A. A. Starobinsky, Phys. Lett. B 91, 99 (1980).
  • (44) A. Ganguly, R. Gannouji, R. Goswami and S. Ray, Phys. Rev. D 89, 064019 (2014) [arXiv:1309.3279 [gr-qc]].
  • (45) R. Kase and S. Tsujikawa, JCAP 09, 054 (2019) [arXiv:1906.08954 [gr-qc]].
  • (46) A. Dohi, R. Kase, R. Kimura, K. Yamamoto and M. a. Hashimoto, arXiv:2003.12571 [gr-qc].
  • (47) A. Maselli, H. O. Silva, M. Minamitsuji and E. Berti, Phys. Rev. D 93, 124056 (2016) [arXiv:1603.04876 [gr-qc]].
  • (48) M. Minamitsuji and H. O. Silva, Phys. Rev. D 93, 124041 (2016) [arXiv:1604.07742 [gr-qc]].
  • (49) J. Chagoya, G. Niz and G. Tasinato, Class. Quant. Grav. 34, 165002 (2017) [arXiv:1703.09555 [gr-qc]].
  • (50) R. Kase, M. Minamitsuji and S. Tsujikawa, Phys. Rev. D 97, 084009 (2018) [arXiv:1711.08713 [gr-qc]].
  • (51) J. Chagoya and G. Tasinato, JCAP 08, 006 (2018) [arXiv:1803.07476 [gr-qc]].
  • (52) H. Ogawa, T. Kobayashi and K. Koyama, Phys. Rev. D 101, 024026 (2020) [arXiv:1911.01669 [gr-qc]].
  • (53) R. Kase, M. Minamitsuji and S. Tsujikawa, arXiv:2001.10701 [gr-qc] (Physical Review D to appear).
  • (54) R. J. Scherrer, Phys. Rev. Lett. 93, 011301 (2004) [arXiv:astro-ph/0402316 [astro-ph]].
  • (55) D. Giannakis and W. Hu, Phys. Rev. D 72, 063502 (2005) [astro-ph/0501423].
  • (56) F. Arroja and M. Sasaki, Phys. Rev. D 81, 107301 (2010) [arXiv:1002.1376 [astro-ph.CO]].
  • (57) A. De Felice, S. Mukohyama and S. Tsujikawa, Phys. Rev. D 82, 023524 (2010) [arXiv:1006.0281 [astro-ph.CO]].
  • (58) R. Kase and S. Tsujikawa, Phys. Rev. D 90, 044073 (2014) [arXiv:1407.0794 [hep-th]].
  • (59) L. Heisenberg, R. Kase and S. Tsujikawa, Phys. Lett. B 760, 617-626 (2016) [arXiv:1605.05565 [hep-th]].
  • (60) A. E. Gumrukcuoglu, R. Kimura and K. Koyama, Phys. Rev. D 101, 124021 (2020) [arXiv:2003.11831 [gr-qc]].
  • (61) B. F. Schutz and R. Sorkin, Annals Phys.  107, 1 (1977).
  • (62) J. D. Brown, Class. Quant. Grav.  10, 1579 (1993) [gr-qc/9304026].
  • (63) A. De Felice, J. M. Gerard and T. Suyama, Phys. Rev. D 81, 063527 (2010) [arXiv:0908.3439 [gr-qc]].
  • (64) G. W. Horndeski, Int. J. Theor. Phys.  10, 363 (1974).
  • (65) C. Deffayet, X. Gao, D. A. Steer and G. Zahariade, Phys. Rev. D 84, 064039 (2011) [arXiv:1103.3260 [hep-th]].
  • (66) T. Kobayashi, M. Yamaguchi and J. ’i. Yokoyama, Prog. Theor. Phys.  126, 511 (2011) [arXiv:1105.5723 [hep-th]].
  • (67) C. Charmousis, E. J. Copeland, A. Padilla and P. M. Saffin, Phys. Rev. Lett.  108, 051101 (2012) [arXiv:1106.2000 [hep-th]].
  • (68) A. De Felice, T. Suyama and T. Tanaka, Phys. Rev. D 83, 104035 (2011) [arXiv:1102.1521 [gr-qc]].
  • (69) T. Kobayashi, H. Motohashi and T. Suyama, Phys. Rev. D 85, 084025 (2012) [arXiv:1202.4893 [gr-qc]].
  • (70) T. Kobayashi, H. Motohashi and T. Suyama, Phys. Rev. D 89, 084042 (2014) [arXiv:1402.6740 [gr-qc]].
  • (71) T. Kobayashi and N. Tanahashi, PTEP 2014, 073E02 (2014) [arXiv:1403.4364 [gr-qc]].
  • (72) H. Ogawa, T. Kobayashi and T. Suyama, Phys. Rev. D 93, no.6, 064078 (2016) [arXiv:1510.07400 [gr-qc]].
  • (73) E. Babichev, C. Charmousis and A. Lehebel, Class. Quant. Grav. 33, no.15, 154002 (2016) [arXiv:1604.06402 [gr-qc]].
  • (74) J. Khoury, M. Trodden and S. S. C. Wong, arXiv:2007.01320 [astro-ph.CO].
  • (75) H. Sotani and K. D. Kokkotas, Phys. Rev. D 71, 124038 (2005) [arXiv:gr-qc/0506060 [gr-qc]].
  • (76) P. Chen, T. Suyama and J. Yokoyama, Phys. Rev. D 92, 124016 (2015) [arXiv:1508.01384 [gr-qc]].
  • (77) S. Morisaki and T. Suyama, Phys. Rev. D 96, 084026 (2017) [arXiv:1707.02809 [gr-qc]].
  • (78) T. Regge and J. A. Wheeler, Phys. Rev.  108, 1063 (1957).
  • (79) F. J. Zerilli, Phys. Rev. Lett.  24, 737 (1970).
  • (80) H. Motohashi and T. Suyama, Phys. Rev. D 84, 084041 (2011) [arXiv:1107.3705 [gr-qc]].
  • (81) R. Kase, M. Minamitsuji, S. Tsujikawa and Y. L. Zhang, JCAP 02, 048 (2018) [arXiv:1801.01787 [gr-qc]].
  • (82) V. Moncrief, Annals Phys. 88, 323-342 (1974).
  • (83) C. O. Lousto and R. H. Price, Phys. Rev. D 55, 2124-2138 (1997) [arXiv:gr-qc/9609012 [gr-qc]].
  • (84) P. Haensel and A. Y. Potekhin, Astron. Astrophys.  428, 191 (2004) [astro-ph/0408324].
  • (85) E. E. Flanagan and T. Hinderer, Phys. Rev. D 77, 021502 (2008) [arXiv:0709.1915 [astro-ph]].
  • (86) T. Damour and A. Nagar, Phys. Rev. D 80 (2009) 084035 [arXiv:0906.0096 [gr-qc]].
  • (87) T. Binnington and E. Poisson, Phys. Rev. D 80 (2009) 084018 [arXiv:0906.1366 [gr-qc]].
  • (88) T. Hinderer, B. D. Lackey, R. N. Lang and J. S. Read, Phys. Rev. D 81, 123016 (2010) [arXiv:0911.3535 [astro-ph.HE]].