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

Instability of nonsingular black holes in nonlinear electrodynamics

Antonio De Felicea and Shinji Tsujikawab aCenter for Gravitational Physics and Quantum Information, Yukawa Institute for Theoretical Physics, Kyoto University, 606-8502, Kyoto, Japan
bDepartment of Physics, Waseda University, 3-4-1 Okubo, Shinjuku, Tokyo 169-8555, Japan
Abstract

We show that nonsingular black holes realized in nonlinear electrodynamics are always prone to Laplacian instability around the center because of a negative squared sound speed in the angular direction. This is the case for both electric and magnetic BHs, where the instability of one of the vector-field perturbations leads to enhancing a dynamical gravitational perturbation in the even-parity sector. Thus, the background regular metric is no longer maintained in a steady state. Our results suggest that the construction of stable, nonsingular black holes with regular centers, if they exist, requires theories beyond nonlinear electrodynamics.

preprint: YITP-24-124, WUCG-24-09

I Introduction

The vacuum black hole (BH) solutions predicted in General Relativity (GR) possess curvature singularities at their centers (r=0r=0). Under several physical assumptions of spacetime and matter, Penrose showed that such singularities arise as an endpoint of the gravitational collapse [1]. However, the existence of singularity-free BHs is not precluded by relaxing some of these assumptions. For example, the nonsingular BH proposed by Bardeen [2] has a regular center due to the absence of global hyperbolicity of spacetime postulated in Penrose’s theorem. Since quantum corrections to GR may manifest themselves in extreme gravity regimes, it is important to investigate whether curvature singularities can be eliminated in extended theories of gravity or matter.

It is known that nonlinear electrodynamics (NED) in the framework of GR allows the existence of spherically symmetric and static (SSS) BHs with regular centers [3, 4, 5, 6, 7, 8, 9, 10, 11, 12]. The Lagrangian {\cal L} of NED depends on F=(1/4)FμνFμνF=-(1/4)F_{\mu\nu}F^{\mu\nu} nonlinearly, where Fμν=μAννAμF_{\mu\nu}=\partial_{\mu}A_{\nu}-\partial_{\nu}A_{\mu} is the field strength of a covector field AμA_{\mu}. The action of Einstein-NED theory is given by

𝒮=d4xg[MPl22R+(F)],{\cal S}=\int{\rm d}^{4}x\sqrt{-g}\left[\frac{M_{\rm Pl}^{2}}{2}R+{\cal L}(F)\right]\,, (1)

where gg is the determinant of a metric tensor gμνg_{\mu\nu}, MPlM_{\rm Pl} is the reduced Planck mass, and RR is the Ricci scalar. For example, Euler-Heisenberg theory [13] in quantum electrodynamics has a low-energy effective Lagrangian =F+αF2{\cal L}=F+\alpha F^{2}, where αF2\alpha F^{2} is a correction to the Maxwell term FF. NED also accommodates Born-Infeld theory [14] with the Lagrangian =μ4[112F/μ4]{\cal L}=\mu^{4}[1-\sqrt{1-2F/\mu^{4}}], in which the electron’s self-energy is nondivergent by the finiteness of FF. These subclasses of NED, when coupled with GR, give rise to hairy BH solutions [15, 16, 17, 18], but there are in general curvature singularities at r=0r=0 unless the functional form of (F){\cal L}(F) is further extended.

The common procedure for realizing nonsingular BHs in NED is to assume the existence of regular metrics and reconstruct the Lagrangian {\cal L} from the field equations of motion [3]. In particular, for the magnetic BH, one can directly express {\cal L} as a function of FF [5]. Nonsingular electric BHs constructed in this manner have finite values of FF and the electric field everywhere. For magnetic BHs, there is the divergence of FF at r=0r=0, but the force exerted on a charged test particle is finite at any distance rr including r=0r=0 [6]. Nonsingular BHs can be designed to meet standard energy conditions, but not all regular solutions do [19]. For instance, violation of dominant energy conditions occurs for some regular BHs [2, 20].

Thus, at the background level, there are consistent regular BHs in NED evading Penrose’s theorem. To see whether these BHs do not suffer from theoretical pathologies, we need to address their linear stability by analyzing perturbations on the SSS background. The BH perturbations in NED were studied in Refs. [21, 22, 23, 24, 25] by focusing on the stability outside the outer horizon. It was found that there are viable parameter spaces in which the nonsingular BHs are plagued by neither ghosts nor Laplacian instabilities. However, the BH stability inside the horizon, especially around its center, is still unclear and has not been investigated to our best knowledge. In this letter, we will show that all nonsingular BHs in NED, including both electric and magnetic ones, are unstable due to the angular Laplacian instability around r=0r=0.

II Nonsingular BHs in NED

The line element on the SSS background is given by

ds2=f(r)dt2+h1(r)dr2+r2(dθ2+sin2θdφ2),{\rm d}s^{2}=-f(r){\rm d}t^{2}+h^{-1}(r){\rm d}r^{2}+r^{2}({\rm d}\theta^{2}+\sin^{2}\theta\,{\rm d}\varphi^{2})\,, (2)

where ff and hh are functions of the radial distance rr. We consider the following covector-field configuration

Aμ=[A0(r),0,0,qMcosθ],A_{\mu}=\left[A_{0}(r),0,0,-q_{M}\cos\theta\right]\,, (3)

where A0A_{0} is a function of rr, and qMq_{M} is a constant corresponding to a magnetic charge. Since the theory (1) has U(1)U(1) gauge invariance, AμAμ+μχgaugeA_{\mu}\to A_{\mu}+\partial_{\mu}\chi_{\rm gauge}, the longitudinal component A1(r)A_{1}(r) has been eliminated by choosing the gauge field as χgauge(r)=rA1(ρ)dρ\chi_{\rm gauge}(r)=-\int^{r}A_{1}(\rho){\rm d}\rho.

Varying the action (1) with respect to A0A_{0}, ff, and hh, it follows that

(h/fr2,FA0)\displaystyle\left(\sqrt{h/f}\,r^{2}{\cal L}_{,F}A_{0}^{\prime}\right)^{\prime} =\displaystyle= 0,\displaystyle 0\,, (4)
h1hr\displaystyle h^{\prime}-\frac{1-h}{r} =\displaystyle= rMPl2f(fhA02,F),\displaystyle\frac{r}{M_{\rm Pl}^{2}f}\left(f{\cal L}-hA_{0}^{\prime 2}{\cal L}_{,F}\right), (5)
ffhh\displaystyle\frac{f^{\prime}}{f}-\frac{h^{\prime}}{h} =\displaystyle= 0,\displaystyle 0\,, (6)

where a prime represents the derivative with respect to rr, and ,Fd/dF{\cal L}_{,F}\equiv{\rm d}{\cal L}/{\rm d}F. The explicit form of FF is given by

F=hA022fqM22r4.F=\frac{hA_{0}^{\prime 2}}{2f}-\frac{q_{M}^{2}}{2r^{4}}\,. (7)

From Eq. (6), we obtain f=Chf=Ch, where CC is a constant. Using time reparametrization invariance, we can impose f1f\to 1 as rr\to\infty, whereas the asymptotic flatness sets h1h\to 1 at spatial infinity. Then we have C=1C=1, so that

f=h.f=h\,. (8)

With this condition, Eqs. (4) and (5) give

A0\displaystyle A_{0}^{\prime} =\displaystyle= qEr2,F,\displaystyle\frac{q_{E}}{r^{2}{\cal L}_{,F}}\,, (9)
\displaystyle{\cal L} =\displaystyle= r2[qEA0+MPl2(rf+f1)],\displaystyle r^{-2}\left[q_{E}A_{0}^{\prime}+M_{\rm Pl}^{2}(rf^{\prime}+f-1)\right]\,, (10)

where qEq_{E} is a constant corresponding to an electric charge. Taking the rr-derivative of Eq. (10) and combining it with Eq. (9) to eliminate ,F{\cal L}_{,F}, we obtain

2qEr4A02+MPl2(2f2r2f′′)r4A0+2qEqM2=0,2q_{E}r^{4}A_{0}^{\prime 2}+M_{\rm Pl}^{2}\left(2f-2-r^{2}f^{\prime\prime}\right)r^{4}A_{0}^{\prime}+2q_{E}q_{M}^{2}=0\,, (11)

which algebraically determines A0A_{0}^{\prime} in terms of ff and its second radial derivative.

We are interested in nonsingular BHs with regular centers. To avoid the singularities of Ricci scalar RR, Ricci squared RμνRμνR_{\mu\nu}R^{\mu\nu}, Riemann squared RμνρσRμνρσR_{\mu\nu\rho\sigma}R^{\mu\nu\rho\sigma} at r=0r=0, we require that ff is expanded around r=0r=0 as [26]

f(r)=1+n=2fnrn,f(r)=1+\sum_{n=2}^{\infty}f_{n}r^{n}\,, (12)

where fnf_{n}’s are constants. The deviation of f(0)f(0) from 1 results in conical singularities. Any power of nn smaller than 2 leads to curvature singularities. Substituting Eq. (12) and its second radial derivative into Eq. (11), two branches of A0A_{0}^{\prime} have the leading-order terms ±(qEqM)2/(qEr2)\pm\sqrt{-(q_{E}q_{M})^{2}}/(q_{E}r^{2}). This means that there exist real solutions to A0A_{0}^{\prime} only if

qEqM=0.q_{E}q_{M}=0\,. (13)

Thus, the presence of dyon BHs with qE0q_{E}\neq 0 and qM0q_{M}\neq 0 is forbidden from the regularity of ff at r=0r=0. From Eq. (13), either qEq_{E} or qMq_{M} must be 0. This non-existence of regular dyon BH solutions breaks the electromagnetic duality present in linear electrodynamics, where the property of BHs is determined by their mass and total charge qT=qE2+qM2q_{T}=\sqrt{q_{E}^{2}+q_{M}^{2}} (see e.g., [27, 28]).

II.1 Purely electric BHs

For qE0q_{E}\neq 0 and qM=0q_{M}=0, the nonvanishing solution to A0A_{0}^{\prime} follows from Eq. (11), such that

A0=MPl2(r2f′′2f+2)2qE.A_{0}^{\prime}=\frac{M_{\rm Pl}^{2}(r^{2}f^{\prime\prime}-2f+2)}{2q_{E}}\,. (14)

Using the regular metric (12) around r=0r=0, we have

A0=2MPl2f3qEr3+5MPl2f4qEr4+𝒪(r5),A_{0}^{\prime}=\frac{2M_{\rm Pl}^{2}f_{3}}{q_{E}}\,r^{3}+\frac{5M_{\rm Pl}^{2}f_{4}}{q_{E}}r^{4}+{\cal O}(r^{5})\,, (15)

which approaches 0 as r0r\to 0. Substituting Eq. (14) into Eqs. (7) and (10), we obtain

F\displaystyle F =\displaystyle= MPl4(r2f′′2f+2)28qE2,\displaystyle\frac{M_{\rm Pl}^{4}(r^{2}f^{\prime\prime}-2f+2)^{2}}{8q_{E}^{2}}\,, (16)
\displaystyle{\cal L} =\displaystyle= MPl22(f′′+2fr).\displaystyle\frac{M_{\rm Pl}^{2}}{2}\left(f^{\prime\prime}+\frac{2f^{\prime}}{r}\right)\,. (17)

Around r=0r=0, these behave as F=2MPl4f32r6/qE2+𝒪(r7)F=2M_{\rm Pl}^{4}f_{3}^{2}r^{6}/q_{E}^{2}+{\cal O}(r^{7}) and =3MPl2(f2+2f3r)+𝒪(r2){\cal L}=3M_{\rm Pl}^{2}(f_{2}+2f_{3}r)+{\cal O}(r^{2}), which are both finite.

The nonsingular BH proposed by Ayon-Beato and Garcia [3] is characterized by the metric components

f=h=12Mr2(r2+r02)3/2+r02r2(r2+r02)2,f=h=1-\frac{2Mr^{2}}{(r^{2}+r_{0}^{2})^{3/2}}+\frac{r_{0}^{2}r^{2}}{(r^{2}+r_{0}^{2})^{2}}\,, (18)

where MM and r0r_{0} are constants. At large distances, Eq. (18) approaches the Reissner-Nordström (RN) metric components f=h=12M/r+qE2/(2MPl2r2)f=h=1-2M/r+q_{E}^{2}/(2M_{\rm Pl}^{2}r^{2}), with the correspondence qE=2MPlr0q_{E}=\sqrt{2}M_{\rm Pl}r_{0}. Around r=0r=0, the metric (18) is related to the coefficients in Eq. (12) as f2=(2Mr0)/r03f_{2}=-(2M-r_{0})/r_{0}^{3}, f3=0f_{3}=0, and f4=(3M2r0)/r05f_{4}=(3M-2r_{0})/r_{0}^{5}. So long as f2<0f_{2}<0, i.e., r0<2Mr_{0}<2M, the central region of BHs is approximately described by the de Sitter spacetime, which generates pressure against gravity.

For a given regular metric ff, we know both FF and {\cal L} as functions of rr from Eqs. (16) and (17). In this case, we can also express {\cal L} as a function of FF, provided that rr is explicitly written in terms of FF.

II.2 Purely magnetic BHs

For qM0q_{M}\neq 0 and qE=0q_{E}=0, Eqs. (9) and (10) give

A0\displaystyle A_{0}^{\prime} =\displaystyle= 0,\displaystyle 0\,, (19)
\displaystyle{\cal L} =\displaystyle= MPl2r2(rf+f1),\displaystyle M_{\rm Pl}^{2}r^{-2}(rf^{\prime}+f-1)\,, (20)

with F=qM2/(2r4)F=-q_{M}^{2}/(2r^{4}). Using the expansion (12), the Lagrangian is regular as =MPl2(3f2+4f3r)+𝒪(r2){\cal L}=M_{\rm Pl}^{2}(3f_{2}+4f_{3}r)+\mathcal{O}(r^{2}) around r=0r=0. For a given f(r)f(r), we can explicitly express {\cal L} as a function of FF by using Eq. (20). The nonsingular BH proposed by Dymnikova [7] corresponds to the metric components

f=h=14Mπr[arctan(rr0)r0rr2+r02],f=h=1-\frac{4M}{\pi r}\left[\arctan\left(\frac{r}{r_{0}}\right)-\frac{r_{0}r}{r^{2}+r_{0}^{2}}\right]\,, (21)

where r0=πqM2/(16MPl2M)r_{0}=\pi q_{M}^{2}/(16M_{\rm Pl}^{2}M) to recover the magnetic RN solution f=12M/r+qM2/(2MPl2r2)f=1-2M/r+q_{M}^{2}/(2M_{\rm Pl}^{2}r^{2}) at large distances. In this case, the Lagrangian is known as

=qM22(r2+r02)2=qM2(2r02+qM2/F)2.{\cal L}=-\frac{q_{M}^{2}}{2(r^{2}+r_{0}^{2})^{2}}=-\frac{q_{M}^{2}}{(\sqrt{2}r_{0}^{2}+\sqrt{-q_{M}^{2}/F})^{2}}\,. (22)

This recovers the standard Maxwell Lagrangian =F{\cal L}=F as F0F\to-0 (i.e., in the limit rr\to\infty).

III Angular Laplacian instabilities of nonsingular BHs

To study the linear stability of electric and magnetic BHs, we consider metric and vector-field perturbations on the SSS background (2) [29, 30, 31]. For the components of metric perturbations hμνh_{\mu\nu}, we choose

htt=f(r)H0(t,r)Yl(θ),htr=H1(t,r)Yl(θ),htθ=0,\displaystyle h_{tt}=f(r)H_{0}(t,r)Y_{l}(\theta),\quad h_{tr}=H_{1}(t,r)Y_{l}(\theta),\quad h_{t\theta}=0,
htφ=Q(t,r)(sinθ)Yl,θ(θ),hrr=f1(r)H2(t,r)Yl(θ),\displaystyle h_{t\varphi}=-Q(t,r)(\sin\theta)Y_{l,\theta}(\theta),\;\;h_{rr}=f^{-1}(r)H_{2}(t,r)Y_{l}(\theta),
hrθ=h1(t,r)Yl,θ(θ),hrφ=W(t,r)(sinθ)Yl,θ(θ),\displaystyle h_{r\theta}=h_{1}(t,r)Y_{l,\theta}(\theta),\quad h_{r\varphi}=-W(t,r)(\sin\theta)Y_{l,\theta}(\theta),
hθθ=0,hφφ=0,hθφ=0,\displaystyle h_{\theta\theta}=0,\quad h_{\varphi\varphi}=0,\quad h_{\theta\varphi}=0, (23)

where Yl(θ)Y_{l}(\theta) is the m=0m=0 component of spherical harmonics Ylm(θ,φ)Y_{lm}(\theta,\varphi). On the SSS background, we can focus on the axisymmetric modes (m=0m=0) without loss of generality. The covector-field perturbation δAμ\delta A_{\mu} has the following components

δAt=δA0(t,r)Yl(θ),δAr=δA1(t,r)Yl(θ),\displaystyle\delta A_{t}=\delta A_{0}(t,r)Y_{l}(\theta),\qquad\delta A_{r}=\delta A_{1}(t,r)Y_{l}(\theta),\qquad
δAθ=0,δAφ=δA(t,r)(sinθ)Yl,θ(θ),\displaystyle\delta A_{\theta}=0,\qquad\delta A_{\varphi}=-\delta A(t,r)(\sin\theta)Y_{l,\theta}(\theta)\,, (24)

where the choice δAθ=0\delta A_{\theta}=0 is an outcome of the presence of U(1)U(1) gauge symmetry.111Since there is U(1)U(1) gauge invariance under the transformation, AμAμ+μχgaugeA_{\mu}\to A_{\mu}+\partial_{\mu}\chi_{\rm gauge}, we can eliminate the even-parity mode δAθ=δA2(t,r)Yl,θ(θ)\delta A_{\theta}=\delta A_{2}(t,r)Y_{l,\theta}(\theta) by choosing the perturbed gauge field δχgauge=δA2(t,r)Yl(θ)\delta\chi_{\rm gauge}=-\delta A_{2}(t,r)Y_{l}(\theta) and making the field redefinitions δA0new=δA0δA˙2\delta A_{0}^{\rm new}=\delta A_{0}-\dot{\delta A}_{2}, δA1new=δA1δA2\delta A_{1}^{\rm new}=\delta A_{1}-\delta A_{2}^{\prime}. We note that the gauge choice (23) completely fixes the residual gauge degrees of freedom under the infinitesimal transformation xμxμ+ξμx^{\mu}\to x^{\mu}+\xi^{\mu}.

The three perturbations QQ, WW, δA\delta A belong to those in the odd-parity sector, while the six perturbations H0H_{0}, H1H_{1}, H2H_{2}, h1h_{1}, δA0\delta A_{0}, δA1\delta A_{1} correspond to those in the even-parity sector. We focus on the multiple modes l2l\geq 2 and expand the action (1) up to second order in perturbed fields. The total quadratic-order action can be expressed as 𝒮(2)=dtdr(1+2){\cal S}^{(2)}=\int{\rm d}t{\rm d}r\,({\cal L}_{1}+{\cal L}_{2}), where 1{\cal L}_{1} and 2{\cal L}_{2} are given in Appendix A. We introduce the following Lagrange multipliers

χ\displaystyle\chi =\displaystyle= rW˙rQ+2Q2,FrA0MPl2δA,\displaystyle r\dot{W}-rQ^{\prime}+2Q-\frac{2{\cal L}_{,F}rA_{0}^{\prime}}{M_{\rm Pl}^{2}}\delta A\,, (25)
V\displaystyle V =\displaystyle= δA0δA1˙+A02(H0H2),\displaystyle\delta A_{0}^{\prime}-\dot{\delta A_{1}}+\frac{A_{0}^{\prime}}{2}(H_{0}-H_{2})\,, (26)

where a dot represents the derivative with respect to tt. The dynamical fields χ\chi and VV correspond to the odd-parity gravitational perturbation and the even-parity electromagnetic perturbation, respectively. We also have the odd-parity electromagnetic mode δA\delta A and the even-parity gravitational mode ψ\psi defined by

ψ=rH2Lh1,whereL=l(l+1).\displaystyle\psi=rH_{2}-Lh_{1}\,,\quad{\rm where}\quad L=l(l+1)\,. (27)

Following the procedure explained in Appendix A, the second-order action, after the elimination of all nondynamical perturbations and the integration by parts, is expressed in the form

𝒮~(2)=dtdr(Ψ˙t𝑲Ψ˙+Ψt𝑮Ψ+Ψt𝑴Ψ+Ψt𝑺Ψ),\tilde{{\cal S}}^{(2)}=\int{\rm d}t{\rm d}r\left(\dot{\vec{\Psi}}^{t}{\bm{K}}\dot{\vec{\Psi}}+\vec{\Psi}^{\prime t}{\bm{G}}\vec{\Psi}^{\prime}+\vec{\Psi}^{t}{\bm{M}}\vec{\Psi}+\vec{\Psi}^{\prime t}{\bm{S}}\vec{\Psi}\right)\,, (28)

where 𝑲,𝑮,𝑴{\bm{K}},{\bm{G}},{\bm{M}} are 4×44\times 4 symmetric matrices with components like K11K_{11}, 𝑺{\bm{S}} is a 4×44\times 4 antisymmetric matrix, and

Ψt=(χ,δA,ψ,V).\vec{\Psi}^{t}=\left(\chi,\delta A,\psi,V\right)\,. (29)

In the eikonal limit (l1l\gg 1), we will derive the linear stability conditions for electric and magnetic BHs. Unlike past related works [21, 25], our results can be applied to the stability for both f>0f>0 and f<0f<0.

III.1 Purely electric BHs

For qE0q_{E}\neq 0 and qM=0q_{M}=0, the dynamical system of perturbations is decomposed into the odd-parity sector with ΨAt=(χ,δA)\vec{\Psi}_{\rm A}^{t}=(\chi,\delta A) and the even-parity sector with ΨBt=(ψ,V)\vec{\Psi}_{\rm B}^{t}=(\psi,V). When f>0f>0, the positivities of 𝑲A{\bm{K}}_{\rm A} and 𝑲B{\bm{K}}_{\rm B}, which are the 2×22\times 2 kinetic matrices of 𝑲{\bm{K}} associated with ΨAt\vec{\Psi}_{\rm A}^{t} and ΨBt\vec{\Psi}_{\rm B}^{t} respectively, determine the no-ghost conditions of four dynamical perturbations. So long as

,F>0,{\cal L}_{,F}>0\,, (30)

both 𝑲A{\bm{K}}_{\rm A} and 𝑲B{\bm{K}}_{\rm B} are positive definite. For f<0f<0, the no-ghost conditions are determined by the positivities of matrices 𝑮A{\bm{G}}_{\rm A} and 𝑮B{\bm{G}}_{\rm B} associated with ΨAt\vec{\Psi}_{\rm A}^{t} and ΨBt\vec{\Psi}_{\rm B}^{t} respectively. They are satisfied with the inequality (30).

For f>0f>0, the radial propagation speeds crc_{r} measured by a proper time τ=fdt\tau=\int f\,{\rm d}t are known by substituting the WKB-form solutions Ψt=Ψ0tei(ωtkr)\vec{\Psi}^{t}=\vec{\Psi}_{0}^{t}e^{-i(\omega t-kr)} into their perturbation equations, where Ψ0t\vec{\Psi}_{0}^{t} is a constant vector composed of (Ψ0t)A(\vec{\Psi}_{0}^{t})_{\rm A} and (Ψ0t)B(\vec{\Psi}_{0}^{t})_{\rm B}. This leads to the algebraic equations 𝑼A(Ψ0)A=0{\bm{U}}_{\rm A}(\vec{\Psi}_{0})_{\rm A}=0 and 𝑼B(Ψ0)B=0{\bm{U}}_{\rm B}(\vec{\Psi}_{0})_{\rm B}=0, where 𝑼A{\bm{U}}_{\rm A} and 𝑼B{\bm{U}}_{\rm B} are 2×22\times 2 matrices. The existence of nonvanishing solutions to (Ψ0)A(\vec{\Psi}_{0})_{\rm A} and (Ψ0)B(\vec{\Psi}_{0})_{\rm B} requires that det𝑼A=0{\rm det}\,{\bm{U}}_{\rm A}=0 and det𝑼B=0{\rm det}\,{\bm{U}}_{\rm B}=0. Taking the large ω\omega and kk limits, both equations lead to (ω2k2f2)2=0(\omega^{2}-k^{2}f^{2})^{2}=0. Substituting ω=kfcr\omega=kfc_{r} into this relation, we find

cr2=1,forallΨt=(χ,δA,ψ,V).c_{r}^{2}=1\,,\qquad{\rm for}~{}~{}~{}{\rm all}\quad\vec{\Psi}^{t}=\left(\chi,\delta A,\psi,V\right)\,. (31)

When f<0f<0, we exploit the WKB solution in the form Ψt=Ψ0tei(ωrkt)\vec{\Psi}^{t}=\vec{\Psi}_{0}^{t}e^{-i(\omega r-kt)}. This results in the dispersion relation (ω2f2k2)2=0(\omega^{2}f^{2}-k^{2})^{2}=0 for both ΨAt\vec{\Psi}_{\rm A}^{t} and ΨBt\vec{\Psi}_{\rm B}^{t}. Then, after the substitution of ω=kcr/(f)\omega=kc_{r}/(-f), we obtain the same squared radial propagation speeds as those in Eq. (31).

For f>0f>0, the angular propagation speeds cΩc_{\Omega} are derived by taking the large ω\omega and ll limits in det𝑼A=0{\rm det}\,{\bm{U}}_{\rm A}=0 and det𝑼B=0{\rm det}\,{\bm{U}}_{\rm B}=0. From det𝑼A=0{\rm det}\,{\bm{U}}_{\rm A}=0, we obtain the dispersion relation (r2ω2Lf)2=0(r^{2}\omega^{2}-Lf)^{2}=0. Substituting ω=cΩlf/r\omega=c_{\Omega}l\sqrt{f}/r into this relation and taking the limit l1l\gg 1, we find

cΩ2=1,forΨAt=(χ,δA).c_{\Omega}^{2}=1\,,\qquad{\rm for}\quad\vec{\Psi}_{\rm A}^{t}=\left(\chi,\delta A\right)\,. (32)

The two solutions following from det𝑼B=0{\rm det}\,{\bm{U}}_{\rm B}=0 are

cΩ2=1,forψ,\displaystyle c_{\Omega}^{2}=1\,,\qquad{\rm for}~{}~{}\psi\,, (33)
cΩ2=cE2,F,F+2F,FF,forV.\displaystyle c_{\Omega}^{2}=c_{E}^{2}\equiv\frac{{\cal L}_{,F}}{{\cal L}_{,F}+2F{\cal L}_{,FF}}\,,\qquad{\rm for}~{}~{}V\,. (34)

Since M44/K44=cE2(l2f/r2)M_{44}/K_{44}=-c_{E}^{2}(l^{2}f/r^{2}) for l1l\gg 1, we can identity cE2c_{E}^{2} as the squared angular propagation speed of VV. When f<0f<0, using the WKB-form solution Ψt=Ψ0tei(ωrkt)\vec{\Psi}^{t}=\vec{\Psi}_{0}^{t}e^{-i(\omega r-kt)} with ω=cΩl/(fr)\omega=c_{\Omega}l/(\sqrt{-f}\,r) results in the same values of cΩ2c_{\Omega}^{2} as those given in Eqs. (32)-(34).

III.2 Purely magnetic BHs

For qM0q_{M}\neq 0 and qE=0q_{E}=0, the system is separated into two sectors: type (C) with ΨCt=(χ,V)\vec{\Psi}_{\rm C}^{t}=(\chi,V) and type (D) with ΨDt=(δA,ψ)\vec{\Psi}_{\rm D}^{t}=(\delta A,\psi) [25, 32]. When f>0f>0, the positivities of 2×22\times 2 kinetic matrices 𝑲C{\bm{K}}_{\rm C} and 𝑲D{\bm{K}}_{\rm D} in each sector are ensured under the condition (30). This is also the case for f<0f<0, where the positivities of matrices 𝑮C{\bm{G}}_{\rm C} and 𝑮D{\bm{G}}_{\rm D} determine the no-ghost conditions.

Using the WKB-form solution Ψt=Ψ0tei(ωtkr)\vec{\Psi}^{t}=\vec{\Psi}_{0}^{t}e^{-i(\omega t-kr)} for f>0f>0, we obtain the two algebraic equations 𝑼C(Ψ0)C=0{\bm{U}}_{\rm C}(\vec{\Psi}_{0})_{\rm C}=0 and 𝑼D(Ψ0)D=0{\bm{U}}_{\rm D}(\vec{\Psi}_{0})_{\rm D}=0 in type C and D sectors, respectively. Taking the large ω\omega and kk limits for det𝑼C=0{\rm det}\,{\bm{U}}_{\rm C}=0 and det𝑼D=0{\rm det}\,{\bm{U}}_{\rm D}=0, we find that all four dynamical perturbations have the luminal squared radial propagation speeds cr2=1c_{r}^{2}=1. This property also holds for f<0f<0.

For the sector (C) with f>0f>0, taking the large ω\omega and ll limits for det𝑼C=0{\rm det}\,{\bm{U}}_{\rm C}=0 leads to the squared angular propagation speeds

cΩ2=1,forΨCt=(χ,V).c_{\Omega}^{2}=1\,,\qquad{\rm for}\quad\vec{\Psi}_{\rm C}^{t}=\left(\chi,V\right)\,. (35)

From the other equation det𝑼D=0{\rm det}~{}{\bm{U}}_{\rm D}=0, we obtain

cΩ2=cM2,F+2F,FF,F,forδA,\displaystyle c_{\Omega}^{2}=c_{M}^{2}\equiv\frac{{\cal L}_{,F}+2F{\cal L}_{,FF}}{{\cal L}_{,F}}\,,\qquad{\rm for}~{}~{}\delta A\,, (36)
cΩ2=1,forψ.\displaystyle c_{\Omega}^{2}=1\,,\qquad{\rm for}~{}~{}\psi\,. (37)

Since M22/K22=cM2l2f/r2M_{22}/K_{22}=-c_{M}^{2}l^{2}f/r^{2} for l1l\gg 1, we can identity cM2c_{M}^{2} as the squared angular propagation speed of δA\delta A. Unlike the electric BH, the odd-parity electromagnetic perturbation δA\delta A has a nontrivial propagation speed different from 1. Again, the results (35)-(37) are valid for f<0f<0.

IV Instability of nonsingular BHs

For the electric BH, we compute Eq. (34) by differentiating Eq. (9) and using the relation F=A02/2F=A_{0}^{\prime 2}/2. This gives ,FF=qE(rA0′′+2A0)/(r3A0′′A03){\cal L}_{,FF}=-q_{E}(rA_{0}^{\prime\prime}+2A_{0}^{\prime})/(r^{3}A_{0}^{\prime\prime}A_{0}^{\prime 3}) and hence cE2=rA0′′/(2A0)c_{E}^{2}=-rA_{0}^{\prime\prime}/(2A_{0}^{\prime}). By using Eq. (14), we obtain

cE2=cf2r(r2f′′′+2rf′′2f)2(r2f′′2f+2),c_{E}^{2}=c_{f}^{2}\equiv-\frac{r(r^{2}f^{\prime\prime\prime}+2rf^{\prime\prime}-2f^{\prime})}{2(r^{2}f^{\prime\prime}-2f+2)}\,, (38)

which depends on ff and its rr derivatives alone.

For the magnetic BH, we take the FF derivative of Eq. (20) and exploit the relation F=qM2/(2r4)F=-q_{M}^{2}/(2r^{4}). Then, we find that cM2c_{M}^{2} in Eq. (36) reduces to cf2c_{f}^{2} in Eq. (38). Thus, for a given metric function f(r)f(r), the squared angular propagation speeds cE2c_{E}^{2} and cM2c_{M}^{2} can be expressed in a unified manner. We have cf2=1c_{f}^{2}=1 at any distance rr for the RN metric f=12M/r+q2/(2MPl2r2)f=1-2M/r+q^{2}/(2M_{\rm Pl}^{2}r^{2}), but this property does not hold for nonsingular BHs.

Let us consider the nonsingular BH with the expansion (12) of ff around r=0r=0. Since f>0f>0 in this regime, the tt and rr coordinates play the timelike and spacelike roles, respectively. The expansion of cf2c_{f}^{2} leads to

cf2=325f44f3r+25f4236f3f58f32r2+𝒪(r3),c_{f}^{2}=-\frac{3}{2}-\frac{5f_{4}}{4f_{3}}r+\frac{25f_{4}^{2}-36f_{3}f_{5}}{8f_{3}^{2}}r^{2}+{\cal O}(r^{3})\,, (39)

which is valid for f30f_{3}\neq 0. Nonsingular BHs like (18) and (21) correspond to f3=0f_{3}=0, in which case we have

cf2=29f510f4r+81f52140f4f650f42r2+𝒪(r3).c_{f}^{2}=-2-\frac{9f_{5}}{10f_{4}}r+\frac{81f_{5}^{2}-140f_{4}f_{6}}{50f_{4}^{2}}r^{2}+{\cal O}(r^{3})\,. (40)

Thus, in both cases, the leading-order terms of cf2c_{f}^{2} are negative. For the metric function f=1+fnrn+𝒪(rn+1)f=1+f_{n}r^{n}+\mathcal{O}(r^{n+1}), we have cf2=n/2+𝒪(r)c_{f}^{2}=-n/2+{\cal O}(r) and hence cf21c_{f}^{2}\leq-1 for n2n\geq 2.

We study the behavior of dynamical perturbations VV and ψ\psi around r=0r=0 for the electric BH. Expressing those fields as V=V~(t)eikrV=\tilde{V}(t)e^{ikr} and ψ=ψ~(t)eikr\psi=\tilde{\psi}(t)e^{ikr} for f>0f>0 and taking the limits l1l\gg 1 and lkrl\gg kr, the time-dependent parts approximately obey the differential equations

V~¨\displaystyle\ddot{\tilde{V}} +cf2fl2r2V~=f2Ωfcf4MPl2r3qEψ~,\displaystyle+c_{f}^{2}\,\frac{fl^{2}}{r^{2}}\tilde{V}=\frac{f^{2}\Omega_{f}c_{f}^{4}M_{\rm Pl}^{2}}{r^{3}q_{E}}\tilde{\psi}\,, (41)
ψ~¨\displaystyle\ddot{\tilde{\psi}} +fl2r2ψ~=l2qErcf2MPl2V~,\displaystyle+\frac{fl^{2}}{r^{2}}\tilde{\psi}=\frac{l^{2}q_{E}}{rc_{f}^{2}M_{\rm Pl}^{2}}\tilde{V}\,, (42)

where Ωfr2f′′2f+2\Omega_{f}\equiv r^{2}f^{\prime\prime}-2f+2. If we use the expansion f=1+fnrn+𝒪(rn+1)f=1+f_{n}r_{n}+{\cal O}(r^{n+1}) around r=0r=0, we have that Ωf=(n2n2)fnrn+𝒪(rn+1)\Omega_{f}=(n^{2}-n-2)\,f_{n}r^{n}+\mathcal{O}(r^{n+1}). It is possible to close Eqs. (41) and (42) for one single variable, say ψ~\tilde{\psi}, finding

ψ~˙˙˙˙+(1+cf2)fl2r2ψ~¨+cf2f2l4r4ψ~0,\ddddot{\tilde{\psi}}+\frac{(1+c_{f}^{2})\,fl^{2}}{r^{2}}\ddot{\tilde{\psi}}+\frac{c_{f}^{2}f^{2}l^{4}}{r^{4}}\tilde{\psi}\simeq 0\,, (43)

where we have neglected the term (f2cf2Ωfl2/r4)ψ~-(f^{2}c_{f}^{2}\Omega_{f}l^{2}/r^{4})\tilde{\psi}. Assuming the solution to Eq. (43) in the form ψ~(t)eiωt\tilde{\psi}(t)\propto e^{-i\omega t}, we obtain

ω2=cf2fl2r2,ω2=fl2r2.\omega^{2}=c_{f}^{2}\,\frac{fl^{2}}{r^{2}}\,,\qquad\omega^{2}=\frac{fl^{2}}{r^{2}}\,. (44)

Since cf2<0c_{f}^{2}<0 in a non-empty set centered around r=0r=0, there is always a growing-mode solution (ω2=cf2fl2/r2\omega^{2}=c_{f}^{2}fl^{2}/r^{2}) besides a stable one (ω2=fl2/r2\omega^{2}=fl^{2}/r^{2}). However, the presence of the former is enough to make the nonsingular BH unstable. We note that V~\tilde{V} obeys the same form of a fourth-order differential equation as Eq. (43), so that the two dynamical perturbations ψ\psi and VV in the even-parity sector (B) are subject to exponential growth.

The enhancement of ψ\psi and VV works as a backreaction to the background BH solution. Then, the background metric is no longer maintained as the steady forms like (18) and (21). For the magnetic BH, the same exponential growth of dynamical perturbations occurs for ψ\psi and δA\delta A in the sector (D). Such instability is generic for all nonsingular BHs constructed in the framework of NED–including those of Bardeen with metric f=12Mr2/(r2+r02)3/2f=1-2Mr^{2}/(r^{2}+r_{0}^{2})^{3/2} [2] and Hayward with metric f=12Mr2/(r3+2Mr02)f=1-2Mr^{2}/(r^{3}+2Mr_{0}^{2}) [20].

A typical time scale of instability arising from the negative value of cf2c_{f}^{2} around r=0r=0 is estimated as

tinsrcf2l.t_{\rm ins}\simeq\frac{r}{\sqrt{-c_{f}^{2}}l}\,. (45)

We recall that cf2-c_{f}^{2} is of order c2c^{2}, where we restored the speed of light cc. Since the distance rr associated with Laplacian instability is less than the outer horizon radius rhr_{h}, tinst_{\rm ins} is much shorter than rh/cr_{h}/c for l1l\gg 1. If we consider a BH with rh=10r_{h}=10 km, we have tins105/lt_{\rm ins}\lesssim 10^{-5}/l sec.

Refer to caption
Figure 1: We plot cf2c_{f}^{2} and ff versus r/r0r/r_{0} for the metric (21) with M=4r0M=4r_{0}. At the distance r<2r0r<\sqrt{2}r_{0}, we have cf2<0c_{f}^{2}<0. For the metric (18), cf2c_{f}^{2} exhibits similar behavior around r=0r=0.

The above results show that nonsingular BHs in NED are always plagued by angular Laplacian instability around r=0r=0. For example, the BH solution (21) has the following squared angular propagation speed

cf2=r22r02r2+r02.c_{f}^{2}=\frac{r^{2}-2r_{0}^{2}}{r^{2}+r_{0}^{2}}\,. (46)

As we estimated in Eq. (40), we have cf2=2c_{f}^{2}=-2 at r=0r=0. While cf2c_{f}^{2} approaches 1 as rr\to\infty, cf2c_{f}^{2} is negative in the region r<2r0r<\sqrt{2}r_{0}.

In Fig. 1, we plot cf2c_{f}^{2} and ff for the metric (21) with M=4r0M=4r_{0}, in which case there are two horizons at r1=0.69r0r_{1}=0.69r_{0} and r2=6.44r0r_{2}=6.44r_{0}. Since the expression (46) is valid at any distance r0r\geq 0, there is angular Laplacian instability for r<2r0r<\sqrt{2}r_{0} (including the region r1r<2r0r_{1}\leq r<\sqrt{2}r_{0} with f<0f<0). The crucial point is that nonsingular BHs always have a finite range of rr where ff is expanded as Eq. (12) around r=0r=0, in which regime cf2c_{f}^{2} is always negative. In Appendix B, we will confirm that the angular Laplacian instability is robust irrespective of the presence/absence of ghosts and the rescaling of dynamical perturbations.

V Conclusions

We have shown that nonsingular BHs in NED are inevitably subject to angular Laplacian instability around r=0r=0. This result holds for both electric and magnetic BHs, as the form (38) of cf2c_{f}^{2} is universal to both cases. The Laplacian instability we found is a physical one, in that the even-parity gravitational perturbation ψ\psi is subject to exponential growth through the angular instability of vector-field perturbations (VV for the electric BH and δA\delta A for the magnetic BH). The backreaction of enhanced perturbations to the background would not keep the regular metrics like (18) and (21) as they are.

Our no-go result for the absence of stable static nonsingular BHs is valid for NED, but this is not the case for more general theories. For example, it is of interest to study what happens by incorporating an additional scalar field ϕ\phi as the Lagrangian (ϕ,X,F){\cal L}(\phi,X,F) [33, 34, 35], where XX is a scalar kinetic term. If such theories with dynamical degrees of freedom still lead to the instability of regular BHs, nonlocal versions of the ultraviolet completion of gravity such as those proposed in Refs. [36, 37, 38, 39] may be the clue to the construction of stable nonsingular BHs.

Acknowledgements

We thank Valeri Frolov for useful discussions. The work of ADF was supported by the Japan Society for the Promotion of Science Grants-in-Aid for Scientific Research No. 20K03969. ST was supported by the Grant-in-Aid for Scientific Research Fund of the JSPS No. 22K03642 and Waseda University Special Research Project No. 2024C-474.

Appendix A: Second-order perturbed action

The second-order action of perturbations, which is obtained after the integration with respect to θ\theta and φ\varphi, can be written in the form 𝒮(2)=dtdr(1+2){\cal S}^{(2)}=\int{\rm d}t{\rm d}r\,({\cal L}_{1}+{\cal L}_{2}), where

1\displaystyle{\cal L}_{1} =\displaystyle= a0H02+H0[a1H2+La2h1+(a3+La4)H2+La5h1+La6δA]+Lb1H12+H1(b2H˙2+Lb3h˙1)\displaystyle a_{0}H_{0}^{2}+H_{0}\left[a_{1}H_{2}^{\prime}+La_{2}h_{1}^{\prime}+(a_{3}+La_{4})H_{2}+La_{5}h_{1}+La_{6}\delta A\right]+Lb_{1}H_{1}^{2}+H_{1}(b_{2}\dot{H}_{2}+Lb_{3}\dot{h}_{1}) (A.1)
+c0H22+LH2(c1h1+c2δA)+L(d0h˙12+d1h12)+Lh1(d2δA0+d3δA)\displaystyle+c_{0}H_{2}^{2}+LH_{2}(c_{1}h_{1}+c_{2}\delta A)+L(d_{0}\dot{h}_{1}^{2}+d_{1}h_{1}^{2})+Lh_{1}(d_{2}\delta A_{0}+d_{3}\delta A^{\prime})
+s1(δA0δA1˙)2+(s2H0+s3H2+Ls4δA)(δA0δA1˙)+L(s5δA02+s6δA12),\displaystyle+s_{1}(\delta A_{0}^{\prime}-\dot{\delta A_{1}})^{2}+(s_{2}H_{0}+s_{3}H_{2}+Ls_{4}\delta A)(\delta A_{0}^{\prime}-\dot{\delta A_{1}})+L(s_{5}\delta A_{0}^{2}+s_{6}\delta A_{1}^{2})\,,
2\displaystyle{\cal L}_{2} =\displaystyle= L[p1(rW˙rQ+2Q)2+p2δA(rW˙rQ+2Q)+p3δA2˙+p4δA2+Lp5δA2+(Lp6+p7)W2\displaystyle L[p_{1}(r\dot{W}-rQ^{\prime}+2Q)^{2}+p_{2}\delta A(r\dot{W}-rQ^{\prime}+2Q)+p_{3}\dot{\delta A^{2}}+p_{4}\delta A^{\prime 2}+Lp_{5}\delta A^{2}+(Lp_{6}+p_{7})W^{2} (A.2)
+(Lp8+p9)Q2+p10QδA0+p11Qh1+p12WδA1],\displaystyle~{}~{}+(Lp_{8}+p_{9})Q^{2}+p_{10}Q\delta A_{0}+p_{11}Qh_{1}+p_{12}W\delta A_{1}]\,,

where we used the condition h=fh=f, and

a0=r28A02(,F+A02,FF),a1=MPl2rf2,a2=MPl2f2,a3=MPl22r24(2A02,F+A04,FF),\displaystyle a_{0}=\frac{r^{2}}{8}A_{0}^{\prime 2}({\cal L}_{,F}+A_{0}^{\prime 2}{\cal L}_{,FF}),\qquad a_{1}=-\frac{M_{\rm Pl}^{2}rf}{2},\qquad a_{2}=\frac{M_{\rm Pl}^{2}f}{2},\qquad a_{3}=-\frac{M_{\rm Pl}^{2}}{2}-\frac{r^{2}}{4}(2{\cal L}-A_{0}^{\prime 2}{\cal L}_{,F}+A_{0}^{\prime 4}{\cal L}_{,FF}),
a4=MPl24,a5=MPl2(f+1)+r2(A02,F)4r,a6=qM2r2(,FA02,FF),b1=a4,\displaystyle a_{4}=-\frac{M_{\rm Pl}^{2}}{4},\qquad a_{5}=\frac{M_{\rm Pl}^{2}(f+1)+r^{2}({\cal L}-A_{0}^{\prime 2}{\cal L}_{,F})}{4r},\qquad a_{6}=\frac{q_{M}}{2r^{2}}({\cal L}_{,F}-A_{0}^{\prime 2}{\cal L}_{,FF}),\qquad b_{1}=-a_{4},
b2=MPl2r,b3=2a4,c0=a32,c1=a5,c2=a6,d0=a4,d1=f2r4(MPl2r2qM2,F),\displaystyle b_{2}=M_{\rm Pl}^{2}r,\qquad b_{3}=2a_{4}\,,\qquad c_{0}=-\frac{a_{3}}{2},\qquad c_{1}=-a_{5},\qquad c_{2}=-a_{6},\qquad d_{0}=-a_{4},\qquad d_{1}=\frac{f}{2r^{4}}(M_{\rm Pl}^{2}r^{2}-q_{M}^{2}{\cal L}_{,F}),
d2=A0,F,d3=qMf,Fr2,s1=r22(,F+A02,FF),s2=A0s1,s3=A0s1,\displaystyle d_{2}=-A_{0}^{\prime}{\cal L}_{,F},\qquad d_{3}=-\frac{q_{M}f{\cal L}_{,F}}{r^{2}},\qquad s_{1}=\frac{r^{2}}{2}({\cal L}_{,F}+A_{0}^{\prime 2}{\cal L}_{,FF}),\qquad s_{2}=A_{0}^{\prime}s_{1},\qquad s_{3}=-A_{0}^{\prime}s_{1},
s4=qMA0,FFr2,s5=,F2f,s6=f,F2,\displaystyle s_{4}=-\frac{q_{M}A_{0}^{\prime}{\cal L}_{,FF}}{r^{2}},\qquad s_{5}=\frac{{\cal L}_{,F}}{2f},\qquad s_{6}=-\frac{f{\cal L}_{,F}}{2}, (A.3)
p1=MPl24r2,p2=d2r,p3=s5,p4=s6,p5=qM2,FFr4,F2r6,p6=MPl2f4r2,\displaystyle p_{1}=\frac{M_{\rm Pl}^{2}}{4r^{2}},\qquad p_{2}=\frac{d_{2}}{r},\qquad p_{3}=s_{5},\qquad p_{4}=s_{6},\qquad p_{5}=\frac{q_{M}^{2}{\cal L}_{,FF}-r^{4}{\cal L}_{,F}}{2r^{6}},\qquad p_{6}=-\frac{M_{\rm Pl}^{2}f}{4r^{2}},
p7=d1,p8=MPl24r2f,p9=d1f2,p10=d3f2,p11=A0fp10,p12=f2p10,\displaystyle p_{7}=d_{1},\qquad p_{8}=\frac{M_{\rm Pl}^{2}}{4r^{2}f},\qquad p_{9}=-\frac{d_{1}}{f^{2}},\qquad p_{10}=-\frac{d_{3}}{f^{2}},\qquad p_{11}=-A_{0}^{\prime}fp_{10},\qquad p_{12}=-f^{2}p_{10}\,, (A.4)

where s4s_{4} vanishes for both the electric BH (qM=0q_{M}=0) and the magnetic BH (A0=0A_{0}^{\prime}=0). The second-order Lagrangians (A.1) and (A.2) with the coefficients (A.3) and (A.4) are valid both for f>0f>0 and f<0f<0.

We incorporate the dynamical fields VV and χ\chi as the forms of Lagrange multipliers

~1\displaystyle\tilde{{\cal L}}_{1} =\displaystyle= 1s1[δA0δA1˙+A02(H0H2)V]2,\displaystyle{\cal L}_{1}-s_{1}\left[\delta A_{0}^{\prime}-\dot{\delta A_{1}}+\frac{A_{0}^{\prime}}{2}(H_{0}-H_{2})-V\right]^{2}\,, (A.5)
~2\displaystyle\tilde{{\cal L}}_{2} =\displaystyle= 2Lp1[rW˙rQ+2Q2,FrA0MPl2δAχ]2.\displaystyle{\cal L}_{2}-Lp_{1}\left[r\dot{W}-rQ^{\prime}+2Q-\frac{2{\cal L}_{,F}rA_{0}^{\prime}}{M_{\rm Pl}^{2}}\delta A-\chi\right]^{2}\,.

Then, we consider the action 𝒮~(2)=dtdr(~1+~2)\tilde{{\cal S}}^{(2)}=\int{\rm d}t{\rm d}r\,(\tilde{{\cal L}}_{1}+\tilde{{\cal L}}_{2}) equivalent to 𝒮(2){\cal S}^{(2)}. We also introduce the dynamical perturbation ψ=rH2Lh1\psi=rH_{2}-Lh_{1} and express H2H_{2} in terms of ψ\psi and h1h_{1}. Since H02H_{0}^{2} vanishes in ~1\tilde{{\cal L}}_{1}, the variation of 𝒮~(2)\tilde{{\cal S}}^{(2)} with respect to H0H_{0} allows one to express h1h_{1} in terms of the other fields. After deriving the perturbation equations of motion for H1H_{1}, δA0\delta A_{0}, δA1\delta A_{1}, QQ, and WW, we can eliminate these fields from 𝒮~(2)\tilde{{\cal S}}^{(2)}. After the integration by parts, we finally obtain the second-order action of the form (28) containing four dynamical perturbations χ\chi, δA\delta A, ψ\psi, VV, and their t,rt,r derivatives.

Appendix B: No-ghost conditions

Let us discuss the no-ghost conditions in more detail. For the electric BH, we have ,F=qE/(r2A0)=2qE2/(MPl2r2Ωf){\cal L}_{,F}=q_{E}/(r^{2}A_{0}^{\prime})=2q_{E}^{2}/(M_{\rm Pl}^{2}r^{2}\Omega_{f}) from Eqs. (9) and (14). For the magnetic BH, we obtain ,F=MPl2r2Ωf/(2qM2){\cal L}_{,F}=M_{\rm Pl}^{2}r^{2}\Omega_{f}/(2q_{M}^{2}) from Eq. (20). Then, in both cases, the no-ghost condition (30) is equivalent to

Ωf=r2f′′2f+2>0.\Omega_{f}=r^{2}f^{\prime\prime}-2f+2>0\,. (A.1)

Using the expansion (12) around r=0r=0, this inequality translates to Ωf=4f3r3+10f4r4+𝒪(r5)>0\Omega_{f}=4f_{3}r^{3}+10f_{4}r^{4}+{\cal O}(r^{5})>0, which is always satisfied if f3>0f_{3}>0 (and if f4>0f_{4}>0 for the BH solution with f3=0f_{3}=0).

For the electric BH in the range f>0f>0, the second-order action of even-parity perturbations contains kinetic terms of VV and ψ\psi, as

𝒮~(2)=dtdr(qE2r2l2fΩfMPl2cf4V˙2+MPl2fl4ψ˙2),\tilde{{\cal S}}^{(2)}=\int{\rm d}t\,{\rm d}r\left(\frac{q_{E}^{2}r^{2}}{l^{2}f\Omega_{f}M_{\rm Pl}^{2}c_{f}^{4}}\,\dot{V}^{2}+\frac{M_{\rm Pl}^{2}f}{l^{4}}\dot{\psi}^{2}\cdots\right)\,, (A.2)

Thus, in the limit that r0r\to 0, there is no strong coupling associated with the vanishing kinetic terms. Under the no-ghost condition Ωf>0\Omega_{f}>0 together with the regular condition f>0f>0 around r=0r=0, the coefficients of V˙2\dot{V}^{2} and ψ˙2\dot{\psi}^{2} are both positive. One can perform the field definitions for VV and ψ\psi to make the kinetic terms in Eq. (A.2) canonical. However, this does not modify the squared angular propagation speed cf2c_{f}^{2}. Indeed, Eq. (43) shows the invariance under the field redefinition ψ~(r,l)ψ~\tilde{\psi}\to{\cal F}(r,l)\tilde{\psi}, where {\cal F} depends on rr and ll. For the magnetic BH, the same property for the invariance of cf2c_{f}^{2} also holds under the redefinition of δA\delta A and ψ\psi. Therefore, the angular instability around r=0r=0 is always present irrespective of no-ghost conditions and the rescaling of dynamical perturbations.

References