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

Tidal response and near-horizon boundary conditions for spinning exotic compact objects

Baoyi Chen [email protected] Walter Burke Institute for Theoretical Physics and Theoretical Astrophysics, M/C 350-17, California Institute of Technology, Pasadena, CA 91125, US    Qingwen Wang [email protected] Perimeter Institute and the University of Waterloo, 31 Caroline St N, Waterloo, ON N2L 2Y5, Canada    Yanbei Chen [email protected] Walter Burke Institute for Theoretical Physics and Theoretical Astrophysics, M/C 350-17, California Institute of Technology, Pasadena, CA 91125, US
Abstract

Teukolsky equations for |s|=2|s|=2 provide efficient ways to solve for curvature perturbations around Kerr black holes. Imposing regularity conditions on these perturbations on the future (past) horizon corresponds to imposing an in-going (out-going) wave boundary condition. For exotic compact objects (ECOs) with external Kerr space time, however, it is not yet clear how to physically impose boundary conditions for curvature perturbations on their boundaries. We address this problem using the Membrane Paradigm, by considering a family of zero-angular-momentum fiducial observers (FIDOs) that float right above the horizon of a linearly perturbed Kerr black hole. From the reference frame of these observers, the ECO will experience tidal perturbations due to in-going gravitational waves, respond to these waves, and generate out-going waves. As it also turns out, if both in-going and out-going waves exist near the horizon, the Newman Penrose (NP) quantity ψ0\psi_{0} will be numerically dominated by the in-going wave, while the NP quantity ψ4\psi_{4} will be dominated by the out-going wave — even though both quantities contain full information regarding the wave field. In this way, we obtain the ECO boundary condition in the form of a relation between ψ0\psi_{0} and the complex conjugate of ψ4\psi_{4}, in a way that is determined by the ECO’s tidal response in the FIDO frame. We explore several ways to modify gravitational-wave dispersion in the FIDO frame, and deduce the corresponding ECO boundary condition for Teukolsky functions. Using the Starobinsky-Teukolsky identity, we subsequently obtain the boundary condition for ψ4\psi_{4} alone, as well as for the Sasaki-Nakamura and Detweiler’s functions. As it also turns out, reflection of spinning ECOs will generically mix between different \ell components of the perturbations fields, and be different for perturbations with different parities. It is straightforward to apply our boundary condition to computing gravitational-wave echoes from spinning ECOs, and to solve for the spinning ECOs’ quasi-normal modes.

I Introduction

A black hole (BH) is characterized by the event horizon, a boundary of the space-time region within which the future null infinity cannot be reached. The existence of a horizon has lead to the simplicity of black-hole solutions in general relativity and modified theories of gravity, although the notion of a horizon has also led to technical and conceptual problems. First of all, at the classical level, the even horizon has a teleological nature: its shape at a particular time-slice of a spacetime depends on what happens to the future of that slide. Even if we are provided with a full numerical solution of the Einstein’s equation (e.g., in the form of all metric components in a particular coordinate system), it is much harder to determine the location of the event horizon than trapped surfaces, whose definitions are more local.

In classical general relativity, it has been shown that a singularity (or singularities) should always exist inside the event horizon Penrose (1969), this requires that quantum gravity must be used to study the space-time geometry inside black holes. Naively, one expects corrections when space-time curvature is at the Planck scale. However, the unique causal structure of the horizon already leads to non-trivial quantum effects, e.g., Hawking radiation Hawking (1975, 1976). From considerations of quantum gravity, it has been proposed that space-time geometry near the horizon can be modified, even at scales larger than the Planck scale Lunin and Mathur (2002, 2002); Mathur (2005); Prescod-Weinstein et al. (2009); Braunstein et al. (2013); Almheiri et al. (2013); Bianchi et al. (2020a, b); Giddings (2017). It has also been conjectured that a phase transition might occur during the formation of black holes, leading to non-singular, yet extremely compact objects Mazur and Mottola (2004); Mathur (2008). All these considerations (or speculations) lead to a similar class of objects: their external space-time geometries mimic those of black holes except very close to the horizon. We shall refer to these objects as Exotic Compact Objects (ECOs).

Followed by the unprecedented discovery of gravitational waves from the binary BH merger event GW150914 Abbott et al. (2016), and follow-up observations of an order of 100\sim 100 binary black-hole merger events Abbott et al. (2019, 2020), we now know that dark compact objects do exist in our universe, and that their space-time geometry and dynamics are consistent with those of black holes in general relativity, better than order unity, and at scales comparable to the sizes of the black holes. Observations by the Event Horizon Telescope (EHT) provides yet another avenue toward near-horizon physics of black holes Akiyama et al. (2019a, b, c, d, e, f); Gralla et al. (2019); Giddings (2019).

Since the horizon is defined as the boundary of the unreachable region hence “absorbs” all radiation, instead of asking whether the horizon exists, a more testable question might be how absorptive the horizon is: any potential modifications to classical general relativity near the surface of an ECO, be it quantum or not, may impose a different physical boundary condition near the horizon. That is, for any incoming gravitational radiations, they not only can fall into the dark object, but may also get reflected, and then propagate to the infinity. In the context of a point particle orbiting a black-hole candidate, this was studied as a modified tidal interaction Fang and Lovelace (2005); Li and Lovelace (2008); Datta et al. (2020). Alternatively, a stronger probe of the reflectivity is provided by waves that propagate toward the horizon of the final (remnant) black hole after the merger of two black holes — in the form of repeated GW echoes at late times in the ringdown signal of a binary merger event Cardoso et al. (2016a, b); Cardoso and Pani (2017); Mark et al. (2017); Abedi et al. (2017); Micchi et al. (2020); Micchi and Chirenti (2020); Conklin (2020); Sago and Tanaka (2020); Burgess et al. (2018). Following this line of thought, the gravitational echoes has been extensively studied in different models of near-horizon structures Conklin et al. (2018); Oshita and Afshordi (2019); Wang et al. (2020); Cardoso et al. (2019); Buoninfante (2020); Bueno et al. (2018). Even though the idea of ECOs might be speculative, one can always regard the search for ECOs as one to quantify the darkness of the final objects in binary merger events, and in this way its importance cannot be overstated.

A major missing piece of the current echo program is how to apply boundary condition near the horizon for curvature perturbations obtained from the Teukolsky equation. This was discussed, by Nakano et al. Oshita et al. (2020) and Wang and Afshordi Wang et al. (2020), but for Kerr there are still more details to fill in — even though Kerr echoes have already been studied by several authors Micchi et al. (2020); Micchi and Chirenti (2020); Conklin (2020); Sago and Tanaka (2020); Burgess et al. (2018). This is the main problem we would like to address in this paper.

Imposing a near-horizon boundary condition is more straightforward in Schwarzschild spacetime. The Schwarzschild metric perturbations can be fully constructed from solutions of the Regge-Wheeler equation Regge and Wheeler (1957) and the Zerilli equation Zerilli (1970), both of which are wave equations that have regular asymptotic behaviors at horizon and infinity. These metric perturbations can then be used to connect to the response of the ECO to external perturbations. In the Kerr spacetime, perturbations are most efficiently described by the s=±2s=\pm 2 Teukolsky equations Teukolsky (1973) for curvature components that are projected along null directions, and therefore are less directly connected to tidal perturbations and responses of an ECO. Furthermore, the Teukolsky equations for the s=±2s=\pm 2 cases do not have short-range potentials, and result in solutions that do not have the standard form of incoming and out-going waves, leading to certain difficulties in finding numerical solutions.

To solve the second issue, the Teukolsky equations can be transformed into wave-like equations with short-ranged potentials, namely the Sasaki-Nakamura (SN) equations, via the Chandrasekhar-Sasaki-Nakamura (CSN) transformation Sasaki and Nakamura (1982a, b); Hughes (2000a). In order to define the near-horizon reflection of waves in the Kerr spacetime, it was proposed that the reflection should be applied to the SN functions — as has been widely used in many literatures regarding gravitational wave echoes Conklin et al. (2018); Nakano et al. (2017); Sago and Tanaka (2020); Li and Lovelace (2008); Du and Chen (2018); Micchi et al. (2020); Micchi and Chirenti (2020); Conklin (2020); Sago and Tanaka (2020). Despite the short-rangedness of the SN equation, the physical meaning of SN functions are less clear than the Teukolsky functions, especially in the Kerr case.

For the Kerr spacetime, Thorne, Price and MacDonald introduced the “Membrane Paradigm” (MP) Thorne et al. (1986), by considering a family of fiducial observers (FIDOs) with zero angular momentum. World lines of the collection of these observers form a “membrane”, which can be used as a proxy to think about the interaction between the black hole and the external universe. In order to recover the pure darkness of the black hole, the membrane must have the correct complex (in fact purely resistive) impedance for each type of flux/current, so that nothing is reflected. For example, the membrane must have the correct specific viscocity in order for gravitational waves not to be reflected, and the correct (electric) resistivity in order for electromagnetic waves not to be reflected. Extensive discussions were made on the physics viewed by the FIDOs, in particular tidal tensors measured by these observers in presence of gravitational waves. The picture was more recently used to visualize space-time geometry using Tendex and Vortex picture Nichols et al. (2011); Zhang et al. (2012); Nichols et al. (2012).

It has been proposed that reflectivity of ECOs can be modeled by altering the impedance of the ECO surface Maggio et al. (2020); Wang et al. (2020); Oshita et al. (2020). In this paper, we generalize this point of view to ECOs with nonzero spins. It is worth mentioning that the the membrane paradigm point of view has been taken by Datta et al. Datta et al. (2020); Datta (2020) to study the tidal heating of Kerr-like ECOs, although reflection of waves by the ECO was not described. In this paper we shall continue along with the membrane paradigm, and propose a physical definition of the ECO’s reflectivity.

In order to do so, we make a careful connection between Teukolsky functions, which efficiently describe wave propagation between the near-horizon region and infinity, and in-going and out-going tidal waves in the FIDO frame of the Membrane Paradigm. We then obtain boundary conditions for the Teukolsky equations in terms of tidal responses of the ECO in the FIDO frame. Here the fundamental assumption that we rely upon — as has also been made implicitly in previous ECO reflectivity literatures — is that the ECO has a simple structure in the FIDO frame — for example as a distribution of exotic matter that modifies dispersion relation of gravitational waves in the FIDO frame.

We organize the paper as follows. In Sec. II, by considering individual FIDOs, we introduce the modified boundary conditions in Teukolsky equations based on the tidal response of the ECO, and obtain input-output relations for Teukolsky equations in terms of that tidal response. In Sec. III, we more specifically consider a Rindler coordinate system near the horizon, and put our discussion into a more firm ground by relating the Teukolsky functions to Riemann tensor components in this coordinate system. We further consider modified gravitational-wave dispersion relations in the Rindler frame, and relate these relations to the ECO’s tidal response. In Sec. IV, we translate our reflection model into a model which fits most literatures on gravitational-wave echoes, in particular making connections to the SN formalism. In Sec. V, we apply our method to obtain the echo waveform as well as the quasi-normal modes (QNMs) of the ECOs. Showing that even- and odd-parity waves will generate different echoes, and generalize the breaking of QNM isospectrality found by Maggio et al. Maggio et al. (2020) to the spinning case. In Sec. VI, we summarize all results and propose possible future works.

Notation. We choose the natural units G=c=1G=c=1, and set the black hole mass M=1M=1. The following symbols are also used throughout the paper:

Δ\displaystyle\Delta =r22r+a2,\displaystyle=r^{2}-2r+a^{2}\,, (1)
Σ\displaystyle\Sigma =r2+a2cos2θ,\displaystyle=r^{2}+a^{2}\cos^{2}\theta\,, (2)
ρ\displaystyle\rho =(riacosθ)1.\displaystyle=-\left(r-ia\cos\theta\right)^{-1}\,. (3)

Here (t,r,θ,ϕ)(t,r,\theta,\phi) are the Boyer-Lindquist coordinates for Kerr black holes and aa is the black hole spin. The Kerr horizons are at the Boyer-Lindquist radius rH=1+1a2r_{H}=1+\sqrt{1-a^{2}}, while the inner horizons are at rC=11a2r_{C}=1-\sqrt{1-a^{2}}. The angular velocity of the horizon is given by ΩH=a/(2rH)\Omega_{H}=a/\left(2r_{H}\right). The tortoise coordinate is defined by

r=r+2rHrHrCln(rrH2)2rCrHrCln(rrC2).r_{*}=r+\frac{2r_{H}}{r_{H}-r_{C}}\ln\left(\frac{r-r_{H}}{2}\right)-\frac{2r_{C}}{r_{H}-r_{C}}\ln\left(\frac{r-r_{C}}{2}\right)\,. (4)

II The reflection boundary condition from tidal response

In a (3+1)(3+1)-splitting of the spacetime, the Weyl curvature tensor CabcdC_{abcd} naturally gets split into an “electric” part, which is responsible for the tidal effect, and a “magnetic” part, which is responsible for the frame-dragging effect. From now on, we will focus on the electric part, as it gives rise to the gravitational stretching and squeezing, i.e. the tidal force, which drives the geodesic deviations of particles that are slowly moving with respect to that slicing.

In MP, a relation is established between the Newman-Penrose quantity ψ0\psi_{0} near the future horizon and components of the tidal tensors in the FIDO frame. In this section, we will extend this to include waves “originating from past horizon”, which really are waves in the vicinity of the horizon but propagate toward the positive rr direction, see Fig. 2. More specifically, we seek to derive the relation among the tidal tensor components, the incoming waves, and the “reflected” (outgoing) waves due to the tidal response. This will establish our model of near-horizon reflection for the Teukolsky equations.

Refer to caption
Figure 1: Trajectory of the FIDO in a constant θ\theta slice of the Kerr spacetime in the (t,rcosϕ,rsinϕ)(t,r\cos\phi,r\sin\phi) coordinate system. Here the green surface indicates the ECO surface with r=br=b, while the black surface indicates the Kerr horizon. Each FIDO has r=br=b, but has (t,ϕ)=(t,ϕ0+ΩHt)(t,\phi)=(t,\phi_{0}+\Omega_{H}t).

II.1 FIDOs

Starting from the Boyer-Lindquist coordinate system (t,r,θ,ϕ(t,r,\theta,\phi), FIDOs in the MP are characterized by constant rr and θ\theta, but ϕ=const+ωϕt\phi={\rm const}+\omega_{\phi}t, with

ωϕ=2arΞ.\omega_{\phi}=\frac{2ar}{\Xi}\,. (5)

and

Ξ=(r2+a2)2a2Δsin2θ.\Xi=(r^{2}+a^{2})^{2}-a^{2}\Delta\sin^{2}\theta\,. (6)

Each FIDO carries an orthonormal tetrad of 111Note that MP uses different notations for the ρ\rho, Σ\Sigma.

er^\displaystyle\vec{e}_{\hat{r}} =ΔΣr,\displaystyle=\sqrt{\frac{\Delta}{\Sigma}}\vec{\partial}_{r}\,,\; eθ^\displaystyle\vec{e}_{\hat{\theta}} =θΣ,\displaystyle=\frac{\vec{\partial}_{\theta}}{\sqrt{\Sigma}}\,,\; (7)
eϕ^\displaystyle\vec{e}_{\hat{\phi}} =ΣΞϕsinθ,\displaystyle=\sqrt{\frac{\Sigma}{\Xi}}\frac{\vec{\partial}_{\phi}}{\sin\theta}\,,\; e0^\displaystyle\vec{e}_{\hat{0}} =1α(t+ωϕϕ),\displaystyle=\frac{1}{\alpha}\left(\vec{\partial}_{t}+\omega_{\phi}\vec{\partial}_{\phi}\right)\,,

with

α=ΣΔΞ.\alpha=\sqrt{\frac{\Sigma\Delta}{\Xi}}\,. (8)

Here e0^\vec{e}_{\hat{0}} is the four-velocity of the FIDO. The FIDOs have zero angular momentum (hence are also known as Zero Angular-Momentum Observers, or ZAMOs), since e0^\vec{e}_{\hat{0}} has zero inner product with ϕ\vec{\partial}_{\phi}. Here α\alpha is called the redshift factor, also known as the lapse function, since it relates the proper time of the FIDOs and the coordinate time tt.

Refer to caption
Figure 2: Waves that propagate toward the ECO surface can be approximated as propagating toward the future horizon, while those originate from the ECO surface can be approximated as originating from the past horizon.

Near the horizon, we have α0\alpha\rightarrow 0; FIDO’s tetrads are related to the Kinnersly tetrad Kinnersley (1969) via

lΣΔ(e0^+er^),nΔΣe0^er^2,meiβ(eθ^+ieϕ^)2,\displaystyle\vec{l}\approx\sqrt{\frac{\Sigma}{\Delta}}(\vec{e}_{\hat{0}}+\vec{e}_{\hat{r}})\,,\vec{n}\approx\sqrt{\frac{\Delta}{\Sigma}}\frac{\vec{e}_{\hat{0}}-\vec{e}_{\hat{r}}}{2}\,,\vec{m}\approx\frac{e^{i\beta}(\vec{e}_{\hat{\theta}}+i\vec{e}_{\hat{\phi}})}{\sqrt{2}}\,, (9)

where

β=tan1(acosθrH).\beta=-\tan^{-1}\left(\frac{a\cos\theta}{r_{H}}\right)\,. (10)

II.2 Tidal Tensor Components

Let us now introduce the electric-type tidal tensor \mathcal{E} as viewed by FIDOs, which can be formally defined as Zhang et al. (2012)

ij=hiahjcCabcdUbUd.\mathcal{E}_{ij}={h_{i}}^{a}{h_{j}}^{c}C_{abcd}U^{b}U^{d}\,. (11)

Here U=e0^U=\vec{e}_{\hat{0}} is the four-velocity of FIDOs as in Eq. (7), and hia=δia+UiUa{h_{i}}^{a}={\delta_{i}}^{a}+U_{i}U^{a} is the projection operator onto the spatial hypersurface orthogonal to UU. In particular, we look at the mmmm-component of the tidal tensor, as the gravitational-wave stretching and squeezing will be along these directions. Near the horizon, the tidal tensor component is then given by

mm=C0^m0^mΔ4Σψ0ΣΔψ4.\mathcal{E}_{mm}=C_{\hat{0}m\hat{0}m}\approx-\frac{\Delta}{4\Sigma}\psi_{0}-\frac{\Sigma}{\Delta}\psi_{4}^{*}\,. (12)

For convenience, let us define a new variable Υs{}_{s}\Upsilon which is the solution to the Teukolsky equation with spin weight ss. For s=±2s=\pm 2 we have

Υ2ρ4ψ4,\displaystyle{}_{-2}\Upsilon\equiv\rho^{-4}\psi_{4}\,, Υ+2ψ0\displaystyle{}_{+2}\Upsilon\equiv\psi_{0} .\displaystyle\,. (13)

We briefly review the Teukolsky formalism in Appendix. A. For perturbations that satisfy the linearized vacuum Einstein’s equation (in this case the Teukolsky equation), at rr_{*}\rightarrow-\infty, in general we can decompose Υs{}_{s}\Upsilon using the spin-weighted spheroidal harmonics Smωs(θ){}_{s}S_{\ell m\omega}(\theta), and write

Υ2(t,r,θ,ϕ)=\displaystyle{}_{-2}\Upsilon(t,r_{*},\theta,\phi)= mdω2πeiωtSmω2(θ)eimϕ\displaystyle\sum_{\ell m}\int\frac{d\omega}{2\pi}e^{-i\omega t}\,{}_{-2}S_{\ell m\omega}(\theta)\,e^{im\phi} (14)
×[ZmωholeΔ2eikr+Zmωrefleikr],\displaystyle\times\left[Z^{\rm hole}_{\ell m\omega}\Delta^{2}e^{-ikr_{*}}+Z^{\rm refl}_{\ell m\omega}\,e^{ikr_{*}}\right]\,,
Υ+2(t,r,θ,ϕ)=\displaystyle{}_{+2}\Upsilon(t,r_{*},\theta,\phi)= mdω2πeiωtSmω+2(θ)eimϕ\displaystyle\sum_{\ell m}\int\frac{d\omega}{2\pi}e^{-i\omega t}\,{}_{+2}S_{\ell m\omega}(\theta)\,e^{im\phi} (15)
×[YmωholeΔ2eikr+Ymωrefleikr],\displaystyle\times\left[Y^{\rm hole}_{\ell m\omega}\,\Delta^{-2}e^{-ikr_{*}}+Y^{\rm refl}_{\ell m\omega}\,e^{ikr_{*}}\right]\,,

where kωmΩHk\equiv\omega\,-\,m\,\Omega_{H}. We use the shorthand m=2m=\sum_{\ell m}\equiv\sum_{\ell=2}^{\infty}\sum_{m=-\ell}^{\ell}, in which \ell is the multipolar index, and mm is the azimuthal quantum number. Note this mm here should not be confused with the label mm in the Kinnersly tetrad basis. Here ZmωZ_{\ell m\omega} and YmωY_{\ell m\omega} are amplitudes for the radial modes, with “hole” labeling the left-propagation modes into the compact object (in this paper, left means toward direction with decreasing rr)  222Of course here we refer to ECOs instead of black holes, the label “hole” is for matching the notations from Teukolsky and Press (1974). and “refl” labeling the right-propagation (reflected) modes (in this paper, “right” means toward direction with increasing rr).

Note that for outgoing modes of either Υ+2{}_{+2}\Upsilon or Υ2{}_{-2}\Upsilon, we have

eiωteikreimϕ=eiω(tr)eim(ϕΩHr),e^{-i\omega t}\,e^{ikr_{*}}\,e^{im\phi}=e^{-i\omega(t-r_{*})}\,e^{im(\phi-\Omega_{H}r_{*})}\,, (16)

therefore the outgoing modes are functions of the retarded time u=tru=t-r_{*}, and the position-dependent angular coordinate ϕΩHr\phi-\Omega_{H}r_{*}. Similarly for ingoing modes, we have

eiωteikreimϕ=eiω(t+r)eim(ϕ+ΩHr),e^{-i\omega t}\,e^{-ikr_{*}}\,e^{im\phi}=e^{-i\omega(t+r_{*})}\,e^{im(\phi+\Omega_{H}r_{*})}\,, (17)

indicating that the ingoing modes are functions of the advanced time v=t+rv=t+r_{*}, and another position-dependent angular coordinate ϕ+ΩHr\phi+\Omega_{H}r_{*}. We can then write down one schematic expression for either Υ+2{}_{+2}\Upsilon or Υ2{}_{-2}\Upsilon, by decomposing both of them into left- and right- propagation components:

Υ+2(t,r,θ,ϕ)=\displaystyle{}_{+2}\Upsilon(t,r_{*},\theta,\phi)= ΥR+2(u,θ,φ)+1Δ2ΥL+2(v,θ,φ+),\displaystyle{}_{+2}\Upsilon^{R}(u,\theta,\varphi_{-})+\frac{1}{\Delta^{2}}{}_{+2}\Upsilon^{L}(v,\theta,\varphi_{+})\,, (18)
Υ2(t,r,θ,ϕ)=\displaystyle{}_{-2}\Upsilon(t,r_{*},\theta,\phi)= ΥR2(u,θ,φ)+Δ2ΥL2(v,θ,φ+),\displaystyle{}_{-2}\Upsilon^{R}(u,\theta,\varphi_{-})+\Delta^{2}{}_{-2}\Upsilon^{L}(v,\theta,\varphi_{+})\,, (19)

where we have defined

φ\displaystyle\varphi_{-} =ϕΩHr,\displaystyle=\phi-\Omega_{H}r_{*}\,, φ+\displaystyle\varphi_{+} =ϕ+ΩHr.\displaystyle=\phi+\Omega_{H}r_{*}\,. (20)

Here both the LL and RR components are finite, and the Δ\Delta represents the divergence/convergence behaviors of the components. As we can see here, once we specify these LL, and RR components on a constant tt slice, as functions of (r,θ,ϕ)(r_{*},\theta,\phi), we will be able to obtain their future, or past, values by inserting tt.

Here we also note that, while the vacuum/homogeneous perturbation of space-time geometry are encoded in both ψ0\psi_{0} and ψ4\psi_{4} — either of them suffices to describe the perturbation field Starobinskil and Churilov (1974); Teukolsky and Press (1974) 333One may imagine a very rough Electromagnetic analogy: for a vacuum EM wave (without electro- or magneto-static fields), both EE and BB fields contain the full information of the wave, since one can use Maxwell equations to convert one to the other. Nevertheless, when it comes to interacting with charges and currents, EE and BB play very different roles, and sometimes it is important to evaluate both EE and BB fields. . Near the horizon, the numerical value of ψ0\psi_{0} is dominated by left-propagating waves, while the numerical value of ψ4\psi_{4} is dominated by right-propagating waves. According to Eq. (12), we then have

mm14ΣΔΥL+2(v,θ,φ+)ρ4ΣΔ[ΥR2(u,θ,φ)].\mathcal{E}_{mm}\approx-\frac{1}{4\Sigma\Delta}{}_{+2}\Upsilon^{L}(v,\theta,\varphi_{+})-\frac{{\rho^{*}}^{4}\Sigma}{\Delta}\left[{{}_{-2}\Upsilon^{R}}(u,\theta,\varphi_{-})\right]^{*}\,. (21)

Note that both terms diverge toward rr_{*}\rightarrow-\infty and at the same order. This divergence correctly reveals the fact that the FIDOs will observe gravitational waves with the same fractional metric perturbation, but because the frequency of the wave gets increased, the curvature perturbation will diverge as α2\alpha^{-2}.

II.3 Linear Response Theory

Now, suppose we have a surface, 𝒮\mathcal{S} at a constant radius r=br_{*}=b_{*} (or in the Boyer-Lindquist coordinates r=br=b), with eκb1e^{\kappa b_{*}}\ll 1. Here κ=(rHrC)/2(rH2+a2)\kappa=(r_{H}-r_{C})/2(r^{2}_{H}+a^{2}) is the surface gravity of the Kerr black hole. To the right of the surface, for r>br_{*}>b_{*}, we have completely vacuum, and to the left of the surface, we have matter that are relatively at rest in the FIDO frame, we shall refer to this as the ECO region. The ECO is assumed to be extremely compact and 𝒮\mathcal{S} is close to the position, as viewed as part of its external Kerr spacetime.

For the moment, let us assume that linear perturbation theory holds throughout the external Kerr spacetime of the ECO. On 𝒮\mathcal{S} and to its right, mm\mathcal{E}_{mm} will be the sum of two pieces,

mm=mmsrc+mmresp,\mathcal{E}_{mm}=\mathcal{E}^{\rm src}_{mm}+\mathcal{E}^{\rm resp}_{mm}\,, (22)

with the first term

mmsrc=Δ4ΣΥsrc+2(v,θ,φ+)\mathcal{E}^{\rm src}_{mm}=-\frac{\Delta}{4\Sigma}{}_{+2}\Upsilon^{\rm src}(v,\theta,\varphi_{+}) (23)

a purely left-propagating wave that is sourced by processes away from the surface, e.g., an orbiting or a plunging a particle. The second term can be written as

mmresp=ρ4ΣΔ[Υrefl2(u,θ,φ)],\displaystyle\mathcal{E}^{\rm resp}_{mm}=-\frac{{\rho^{*}}^{4}\Sigma}{\Delta}\left[{}_{-2}\Upsilon^{\rm refl}(u,\theta,\varphi_{-})\right]^{*}\,, (24)

as the ECO’s response to the incoming gravitational wave.

Now we are prepared to discuss the reflecting boundary condition of the Teukolsky equations in terms of the tidal response of the ECO. According to the linear response theory, we can assume the linear tidal response of the ECO is proportional to the total tidal fields near the surface of the ECO. That is, we may introduce a new parameter η\eta, and write

mmresp=η(b,θ)mm.\mathcal{E}^{\rm resp}_{mm}=\eta(b,\theta)\,\mathcal{E}_{mm}\,. (25)

Here η\eta is analogous to the tidal love number. This leads to the following relation at r=br_{*}=b_{*}:

[Υrefl2(tb,θ,ϕΩHb)]Υsrc+2(t+b,θ,ϕ+ΩHb)=η1ηe4iβ4Δ2.\displaystyle\frac{\left[{}_{-2}\Upsilon^{\rm refl}(t-b_{*},\theta,\phi-\Omega_{H}b_{*})\right]^{*}}{{}_{+2}\Upsilon^{\rm src}(t+b_{*},\theta,\phi+\Omega_{H}b_{*})}=\frac{\eta}{1-\eta}\frac{e^{-4i\beta}}{4}\Delta^{2}\,. (26)

In particular, when η\eta\rightarrow\infty, we will have the Dirichlet boundary condition,

[Υrefl2(tb,θ,ϕΩHb)]Υsrc+2(t+b,θ,ϕ+ΩHb)=e4iβ4Δ2.\frac{\left[{}_{-2}\Upsilon^{\rm refl}(t-b_{*},\theta,\phi-\Omega_{H}b_{*})\right]^{*}}{{}_{+2}\Upsilon^{\rm src}(t+b_{*},\theta,\phi+\Omega_{H}b_{*})}=-\frac{e^{-4i\beta}}{4}\Delta^{2}\,. (27)

This then provides us with a prescription for obtaining the boundary condition at r=br_{*}=b_{*}. Once we know the left-propagating ψ0src\psi_{0}^{\rm src}, the reflected waves due to the tidal response are simply given by Eq. (26).

Let us now define a new parameter (b,θ)\mathcal{R}(b,\theta) as

(b,θ)η1η.\mathcal{R}(b,\theta)\equiv-\frac{\eta}{1-\eta}\,. (28)

This parameter has the physical meaning of being the reflectivity of the tidal fields on the ECO surface. This local response assumes that different angular elements of the ECO act independently, which is reasonable since on the ECO surface, and in the FIDO frame, the gravitational wavelength is blue shifted by α\alpha, hence, much less than the radius of the ECO.

However, the response may also depend on the history of the exerted tidal perturbation, and to account for this we can write a more general boundary condition on the ECO surface as

[Υrefl2(tb,θ,ϕΩHb)]=e4iβ4t𝑑t(b,θ;tt)Δ2Υsrc+2(t+b,θ,ϕ+ΩHbΩH(tt)).\displaystyle\!\!\left[{}_{-2}\Upsilon^{\rm refl}(t-b_{*},\theta,\phi-\Omega_{H}b_{*})\right]^{*}=-\frac{e^{-4i\beta}}{4}\int_{-\infty}^{t}dt^{\prime}\mathcal{R}(b,\theta;t-t^{\prime})\Delta^{2}{}_{+2}\Upsilon^{\rm src}(t^{\prime}+b_{*},\theta,\phi+\Omega_{H}b_{*}-\Omega_{H}(t-t^{\prime}))\,. (29)

Here the ΩH(tt)-\Omega_{H}(t-t^{\prime}) term has been inserted into the argument of Υsrc+2{}_{+2}\Upsilon^{\rm src} because the FIDO follows ϕ=ϕ0+ΩHt\phi=\phi_{0}+\Omega_{H}t (see Fig. 1). This is the key equation of our reflection model.

II.4 Mode Decomposition

We now have obtained the modified boundary condition (29) in terms of the Newman-Penrose quantities, and are ready to apply it to the Teukolsky formalism. The solutions to the s=2s=-2 Teukolsky equation, Υ2{}_{-2}\Upsilon, admits the near-horizon decomposition as in Eq. (14). In this equation, ZholeZ^{\rm hole} is the amplitude of the ingoing wave down to the ECO, which is contributed by the source, and ZreflZ^{\rm refl} is the amplitude of the reflected wave due to the tidal response. For s=+2s=+2, the corresponding amplitudes are YholeY^{\rm hole} and YreflY^{\rm refl}. We would like to derive a relation among the four amplitudes.

Near the ECO surface 𝒮\mathcal{S}, Υsrc+2{}_{+2}\Upsilon^{\rm src} is given by

Υsrc+2(v,θ,φ+)\displaystyle{}_{+2}\Upsilon^{\rm src}(v,\theta,\varphi_{+}) =mdω2πeiωvYmωholeΔ2Smω+2(θ,φ+),\displaystyle=\sum_{\ell m}\int\frac{d\omega}{2\pi}e^{-i\omega v}Y^{\rm hole}_{\ell m\omega}\,\Delta^{-2}\,{}_{+2}S_{\ell m\omega}(\theta,\varphi_{+})\,, (30)

where we have kept only the dominant piece—the left-propagating mode, and YmωholeY^{\rm hole}_{\ell m\omega} is the amplitude of that mode. The quantity Υrefl2{}_{-2}\Upsilon^{\rm refl} is given by

Υrefl2(u,θ,φ)\displaystyle{}_{-2}\Upsilon^{\rm refl}(u,\theta,\varphi_{-}) =mdω2πeiωuZmωreflSmω2(θ,φ),\displaystyle=\sum_{\ell m}\int\frac{d\omega}{2\pi}e^{-i\omega u}Z^{\rm refl}_{\ell m\omega}\,{}_{-2}S_{\ell m\omega}(\theta,\varphi_{-})\,, (31)

Inserting the above two equations into Eq. (29), we obtain

ZmωreflSmω2(θ,ϕ)\displaystyle\sum_{\ell}Z_{\ell m\omega}^{\rm refl}\,{}_{-2}S_{\ell m\omega}(\theta,\phi)
=\displaystyle= 14e4iβ2ikb(1)m+1kY-m-ωholeSmω2(θ,ϕ),\displaystyle\frac{1}{4}\sum_{\ell^{\prime}}e^{4i\beta-2ikb_{*}}(-1)^{m+1}\mathcal{R}^{*}_{-k^{*}}\,Y^{\rm hole^{*}}_{\ell^{\prime}\,\text{-}m\,\text{-}\omega^{*}}\,{}_{-2}S_{\ell^{\prime}m\omega^{*}}(\theta,\phi)\,, (32)

where ω(b,θ)\mathcal{R}_{\omega}(b,\theta) is the Fourier transform of (b,θ;tt)\mathcal{R}(b,\theta;t-t^{\prime}), and kωmΩHk\equiv\omega-m\Omega_{H}. During the derivation, we have used the fact that the spheroidal harmonic functions satisfy the relation

Smω2(θ,ϕ)=(1)mS-m-ω+2(θ,ϕ).{}_{-2}S^{*}_{\ell m\omega}(\theta,\phi)=(-1)^{m}{}_{+2}S_{\ell\,\text{-}m\,\text{-}\omega^{*}}(\theta,\phi)\,. (33)

Assuming the normalization that

0π|Smω2(θ)|2sinθdθ=1,\int_{0}^{\pi}|{}_{-2}S_{\ell m\omega}(\theta)|^{2}\sin\theta d\theta=1\,, (34)

from Eq. (II.4), we can write

Zmωrefl=(1)m+114e2ikbmωY-m-ωhole,Z_{\ell m\omega}^{\rm refl}=(-1)^{m+1}\frac{1}{4}e^{-2ikb_{*}}\sum_{\ell^{\prime}}\mathcal{M}_{\ell\ell^{\prime}m\omega}Y^{\rm hole^{*}}_{\ell^{\prime}\,\text{-}m\,\text{-}\omega^{*}}\,, (35)

with

mω\displaystyle\mathcal{M}_{\ell\ell^{\prime}m\omega} =0πω+mΩH(b,θ)e4iβ(θ)×\displaystyle=\int_{0}^{\pi}\mathcal{R}^{*}_{-\omega^{*}+m\Omega_{H}}(b,\theta)\,e^{4i\beta(\theta)}\times (36)
×Smω2(θ)Smω2(θ)sinθdθ.\displaystyle\times{}_{-2}S_{\ell^{\prime}m\omega}(\theta){}_{-2}S^{*}_{\ell m\omega}(\theta)\sin\theta\,d\theta\,.

In general the reflection will mix between modes with different \ell, but not different mm. Note that the mixing not only arises from the θ\theta dependence of (θ,b)\mathcal{R}(\theta,b), but also from the θ\theta dependence of β\beta. This mixing vanishes for the Schwarzschild case. For our calculation, it will be good to discard the phase term e4iβe^{4i\beta} and make the assumption that \mathcal{R} is independent of the angle θ\theta. But we should keep in mind that these assumptions only work well in the Schwarzschild limit a0a\rightarrow 0.

In the simplified scenario where mode mixing is ignored, we can write

Zmωrefl(1)m+114e2ikbω+mΩHY-m-ωhole,Z^{\rm refl}_{\ell m\omega}\approx\left(-1\right)^{m+1}\frac{1}{4}\,e^{-2ikb_{*}}\,\mathcal{R}^{*}_{-\omega^{*}+m\Omega_{H}}\,Y^{\rm hole^{*}}_{\ell\,\text{-}m\,\text{-}\omega^{*}}\,, (37)

In this way, the ω\omega frequency component of the ψ4\psi_{4} amplitude of each (l,m)(l,m) mode is related to the ω-\omega^{*} frequency component of ψ0\psi_{0} of the (l,m)(l,-m) mode. Here in a Fourier analysis, ω\omega is always real, but we have kept ω\omega^{*} so that our notation will directly apply to quasi-normal modes, where frequency can be complex.

III Wave Propagation in the vicinity of the horizon

In the previous section, we have obtained a new reflecting boundary condition (29) relating Newman-Penrose quantity ψ0\psi_{0} and ψ4\psi_{4} on a spherical surface near the Kerr horizon. This was further converted as a relation between frequency components of the in-coming ψ0\psi_{0} and the out-going ψ4\psi_{4}. Before moving on to the applications of these boundary conditions, in this section, we put the discussions of the previous section onto a more solid ground. We consider a concrete coordinate system associated with the FIDOs, and relate condition (29) to modified refractive indices or dispersion relations of gravitational waves in this coordinate system. This way of modeling the ECO can be thought of as a generalization of Refs. Wang et al. (2020); Oshita and Afshordi (2019); Oshita et al. (2020) to the spinning case.

III.1 Rindler approximations

Let us now study the propagation of waves near the horizon, and explore how emergent gravity might influence the boundary condition there.

Inside the ECO boundary 𝒮\mathcal{S}, we can consider propagation of metric perturbations in the near-horizon FIDO coordinate system. According to MP, the unperturbed metric takes the simple form Thorne et al. (1986):

ds2=α2dt¯2+dα2gH2+ΣHdθ¯2+4rH2ΣHsin2θ¯dϕ¯2,ds^{2}=-\alpha^{2}d\bar{t\vphantom{\theta}}^{2}+\frac{d\alpha^{2}}{g_{H}^{2}}+\Sigma_{H}d\bar{\theta}^{2}+\frac{4r_{H}^{2}}{\Sigma_{H}}\sin^{2}\bar{\theta}d\bar{\phi}^{2}\,, (38)

where

gH=rH12rH,\displaystyle g_{H}=\frac{r_{H}-1}{2r_{H}}\,, ΣH=rH2+a2cos2θ¯.\displaystyle\Sigma_{H}=r^{2}_{H}+a^{2}\cos^{2}\bar{\theta}\,. (39)

This metric, only valid for α1\alpha\ll 1, is a Rindler-like spacetime with spherical symmetry, with horizon located at α=0\alpha=0. According to the membrane paradigm Suen et al. (1988), the new radial coordinates (α,θ¯,ϕ¯)(\alpha,\,\bar{\theta},\,\bar{\phi}) are defined as

t¯\displaystyle\bar{t\vphantom{\theta}} =t,\displaystyle=t\,, (40)
α\displaystyle\alpha =(2gH2aΩHgHsin2θ)12(rrH)12,\displaystyle=\left(2g_{H}-2a\Omega_{H}g_{H}\sin^{2}\theta\right)^{\frac{1}{2}}\left(r-r_{H}\right)^{\frac{1}{2}}\,, (41)
θ¯\displaystyle\bar{\theta} =θΣH,θ4gH2ΣH2α2,\displaystyle=\theta-\frac{{\Sigma_{H}}_{,\theta}}{4g_{H}^{2}\Sigma_{H}^{2}}\alpha^{2}\,, (42)
ϕ¯\displaystyle\bar{\phi} =ϕΩHt.\displaystyle=\phi-\Omega_{H}t\,. (43)

The Kinnersley tetrad, near the horizon, can then be expressed in terms of the Rindler coordinates as

l\displaystyle\vec{l} =2rHΔ(t¯+gHαα),\displaystyle=\frac{2r_{H}}{\Delta}(\vec{\partial}_{\bar{t\vphantom{\theta}}}+g_{H}\alpha\vec{\partial}_{\alpha})\,, (44)
n\displaystyle\vec{n} =rHΣ(t¯gHαα),\displaystyle=\frac{r_{H}}{\Sigma}(\vec{\partial}_{\bar{t\vphantom{\theta}}}-g_{H}\alpha\vec{\partial}_{\alpha})\,, (45)
m\displaystyle\vec{m} =ρ2[iasinθ¯t¯aΩHsinθ¯cosθ¯1aΩHsin2θ¯αα\displaystyle=\frac{-\rho^{*}}{\sqrt{2}}\Big{[}ia\sin\bar{\theta}\vec{\partial}_{\bar{t\vphantom{\theta}}}-\frac{a\Omega_{H}\sin\bar{\theta}\cos\bar{\theta}}{1-a\Omega_{H}\sin^{2}\bar{\theta}}\alpha\vec{\partial}_{\alpha}
+θ¯+(isinθ¯iaΩHsinθ¯)ϕ¯],\displaystyle\qquad\quad+\vec{\partial}_{\bar{\theta}}+\left(\frac{i}{\sin\bar{\theta}}-ia\Omega_{H}\sin\bar{\theta}\right)\vec{\partial}_{\bar{\phi}}\Big{]}\,, (46)

where we have used the near-horizon approximations and discarded all 𝒪(α2)\mathcal{O}(\alpha^{2}) corrections.

For convenience, we introduce a new radial coordinate xx, which is related to the lapse function via

α=egHx.\alpha=e^{g_{H}x}\,. (47)

The regime xx\rightarrow-\infty is the horizon, where α0\alpha\rightarrow 0. In fact, (t,x)(t,x) is exactly the Cartesian coordinate of the Minkowski space in which this Rindler space is embedded. Now we consider metric perturbations of the trace-free form

hθ¯θ¯(t,x,θ¯,ϕ¯)\displaystyle h_{\bar{\theta}\bar{\theta}}(t,x,\bar{\theta},\bar{\phi})\ =ΣHH+(t,x,θ¯,ϕ¯),\displaystyle=\Sigma_{H}H_{+}(t,x,\bar{\theta},\bar{\phi})\,, (48)
hθ¯ϕ¯(t,x,θ¯,ϕ¯)\displaystyle h_{\bar{\theta}\bar{\phi}}(t,x,\bar{\theta},\bar{\phi}) =2rHsinθ¯H×(t,x,θ¯,ϕ¯),\displaystyle=2r_{H}\sin\bar{\theta}H_{\times}(t,x,\bar{\theta},\bar{\phi})\,, (49)
hϕ¯ϕ¯(t,x,θ¯,ϕ¯)\displaystyle h_{\bar{\phi}\bar{\phi}}(t,x,\bar{\theta},\bar{\phi}) =4rH2sinθ¯2H+(t,x,θ¯,ϕ¯)/ΣH.\displaystyle=-4r^{2}_{H}\sin\bar{\theta}^{2}H_{+}(t,x,\bar{\theta},\bar{\phi})/\Sigma_{H}\,. (50)

Note that H+,×H_{+,\times} are metric perturbations in the angular directions, measured in orthonormal bases. We first find that the Einstein’s equations reduce to

(t2+x2)Hp=0,p=+,×.(-\partial_{t}^{2}+\partial_{x}^{2})H_{p}=0\,,\quad p=+,\times. (51)

Again, to obtain the equations above we have only kept the leading terms in α\alpha-series. The absence of θ¯\bar{\theta} and ϕ¯\bar{\phi} derivatives in this equation supports the argument that tidal response of the ECO is local to each angular element on its surface, as we have made in Sec. II.3.

We can further decompose Hp(t,x)H_{p}(t,x) to the left- and right-propagating piece as

Hp(t,x,θ¯,ϕ¯)=HpL(t+x,θ¯,ϕ¯)+HpR(tx,θ¯,ϕ¯),H_{p}(t,x,\bar{\theta},\bar{\phi})=H_{p}^{L}(t+x,\bar{\theta},\bar{\phi})+H_{p}^{R}(t-x,\bar{\theta},\bar{\phi})\,, (52)

Using the Rindler approximations, we then find that the Weyl quantities ψ0\psi_{0} and ψ4\psi_{4} near horizon can be written as

ψ0(t,x,θ¯,ϕ¯)\displaystyle\psi_{0}(t,x,\bar{\theta},\bar{\phi}) =8rH2e2iβΔ2(t2gHx)L(t,x,θ¯,ϕ¯),\displaystyle=\frac{8r_{H}^{2}e^{2i\beta}}{\Delta^{2}}(\partial_{t}^{2}-g_{H}\partial_{x})\mathcal{H}^{L}(t,x,\bar{\theta},\bar{\phi})\,, (53)
[ρ4ψ4(t,x,θ¯,ϕ¯)]\displaystyle\Big{[}\rho^{-4}\psi_{4}(t,x,\bar{\theta},\bar{\phi})\Big{]}^{*} =2rH2e2iβ(t2gHx)R(t,x,θ¯,ϕ¯),\displaystyle=2r_{H}^{2}e^{-2i\beta}(\partial_{t}^{2}-g_{H}\partial_{x})\mathcal{H}^{R}(t,x,\bar{\theta},\bar{\phi})\,, (54)

where we have defined

L\displaystyle\mathcal{H}^{L} =H+L+iH×L,\displaystyle=H^{L}_{+}+iH^{L}_{\times}\,, R\displaystyle\mathcal{H}^{R} =H+R+iH×R.\displaystyle=H^{R}_{+}+iH^{R}_{\times}\,. (55)

Note that in this approximation, we only extract the leading behavior of ψ0\psi_{0} and ψ4\psi_{4} near the horizon, namely a left-going wave (rrH)2\sim(r-r_{H})^{-2} for ψ0\psi_{0}, and a right-going wave (rrH)0\sim(r-r_{H})^{0} for ψ4\psi_{4}. Here, we are considering wave propagation and reflection independently for each (θ¯,ϕ¯)(\bar{\theta},\bar{\phi}). Eq. (53) and Eq. (54) are consistent with our reflection model given in Eq. (29). For instance, in the case of total reflection, we have =1\mathcal{R}=-1, and all left-propagating modes HpLH^{L}_{p} become right-propagating modes HpRH^{R}_{p}.

Refer to caption
Figure 3: Illustration of the constant-xx contours in the (r,cosθ)(r_{*},\cos\theta) plane. Reflections from the same xx for different θ\theta will appear as being reflected at different rr_{*} for different θ\theta.

Let us now evaluate the Riemann tensor components in an orthonormal basis whose vectors point along the (t,x,θ¯,ϕ¯)(t,x,\bar{\theta},\bar{\phi}) coordinate axes. The results are

Rt¯^θ¯^t¯^θ¯^=Rt¯^ϕ¯^t¯^ϕ¯^\displaystyle R_{\hat{\bar{t\vphantom{\theta}}}\hat{\bar{\theta}}\hat{\bar{t\vphantom{\theta}}}\hat{\bar{\theta}}}=-R_{\hat{\bar{t\vphantom{\theta}}}\hat{\bar{\phi}}\hat{\bar{t\vphantom{\theta}}}\hat{\bar{\phi}}} =e2gHx2[t2+gHx]H+,\displaystyle=-\frac{e^{-2g_{H}x}}{2}\left[-\partial_{t}^{2}+g_{H}\partial_{x}\right]H_{+}\,, (56)
Rt¯^θ¯^t¯^ϕ¯^\displaystyle R_{\hat{\bar{t\vphantom{\theta}}}\hat{\bar{\theta}}\hat{\bar{t\vphantom{\theta}}}\hat{\bar{\phi}}} =e2gHx2[t2+gHx]H×.\displaystyle=-\frac{e^{-2g_{H}x}}{2}\left[-\partial_{t}^{2}+g_{H}\partial_{x}\right]H_{\times}\,. (57)

This also confirms the reflection model that we have obtained from the previous section.

We also point out that near the horizon, xx and rr_{*} differ by a additive constant for each (θ¯,ϕ¯)(\bar{\theta},\bar{\phi}). Let’s work out the θ¯\bar{\theta} dependence of the asymptotic shift between xx and rr_{*}. More specifically, near the horizon, the tortoise coordinate rr_{*} is approximately given by

r12gHln[2gH(rrH)]+,r_{*}\approx\frac{1}{2g_{H}}\ln[2g_{H}(r-r_{H})]+\mathcal{I}\,, (58)

with a constant

=rH+ln(rH2)12rHgHln(8rHgH2).\mathcal{I}=r_{H}+\ln\left(\frac{r_{H}}{2}\right)-\frac{1}{2r_{H}g_{H}}\ln\left(8r_{H}g^{2}_{H}\right)\,. (59)

Here we have neglected 𝒪(rrH)\mathcal{O}(r-r_{H}) terms. Note that rr_{*} is independent of θ¯\bar{\theta}. We define the difference between the two radial coordinates as

xrδ(θ¯),x-r_{*}\equiv\delta(\bar{\theta})-\mathcal{I}\,, (60)

where

δ(θ¯)=12gHln(1aΩHsin2θ¯).\delta(\bar{\theta})=\frac{1}{2g_{H}}\ln(1-a\Omega_{H}\sin^{2}\bar{\theta})\,. (61)

This may influence the mode mixing of reflected waves from an ECO whose surface has a constant redshift. In Fig. 3, we illustrate contant-xx contours in the (r,cosθ)(r_{*},\cos\theta) plane.

Finally let us derive the Teukolsky reflectivity \mathcal{R} using the Rindler approximation. Supposing for xx in certain regions we can write the wave solution as

(t,x,θ¯,ϕ¯)=L(t,x,θ¯,ϕ¯)+R(t,x,θ¯,ϕ¯),\displaystyle\mathcal{H}(t,x,\bar{\theta},\bar{\phi})=\mathcal{H}^{L}(t,x,\bar{\theta},\bar{\phi})+\mathcal{H}^{R}(t,x,\bar{\theta},\bar{\phi})\,, (62)

with

L(t,x,θ¯,ϕ¯)\displaystyle\mathcal{H}^{L}(t,x,\bar{\theta},\bar{\phi}) =mdk2πΘk(θ¯)eikxeikteimϕ¯,\displaystyle=\sum_{m}\int\frac{dk}{2\pi}\Theta_{k}(\bar{\theta})e^{-ikx}e^{-ikt}e^{im\bar{\phi}}\,, (63)
R(t,x,θ¯,ϕ¯)\displaystyle\mathcal{H}^{R}(t,x,\bar{\theta},\bar{\phi}) =mdk2πΘk(θ¯)ζkeikxeikteimϕ¯.\displaystyle=\sum_{m}\int\frac{dk}{2\pi}\Theta_{k}(\bar{\theta})\zeta_{k}e^{ikx}e^{-ikt}e^{im\bar{\phi}}\,. (64)

Here Θk(θ¯)\Theta_{k}(\bar{\theta}) gives the kk-dependent angular distribution. ζkζ(k)\zeta_{k}\equiv\zeta(k) is the reflection coefficient that converts left-propagating to right-propagating gravitational waves. Thus ψ0\psi_{0} and ψ4\psi_{4} are respectively given by

ψ\displaystyle\psi (t,x,θ¯,ϕ¯)0=8rH2e2iβ(θ¯)Δ2mdk2πΘk(θ¯)(k2+igHk)eikxeikteimϕ¯,{}_{0}(t,x,\bar{\theta},\bar{\phi})=\frac{8r^{2}_{H}e^{2i\beta(\bar{\theta})}}{\Delta^{2}}\sum_{m}\int\frac{dk}{2\pi}\Theta_{k}(\bar{\theta})(-k^{2}+ig_{H}k)e^{-ikx}e^{-ikt}e^{im\bar{\phi}}\,, (65)
(ρ4ψ\displaystyle(\rho^{-4}\psi )4(t,x,θ¯,ϕ¯)=2rH2e2iβ(θ¯)mdk2πΘk(θ¯)(k2igHk)ζkeikxeikteimϕ¯.{}_{4})^{*}(t,x,\bar{\theta},\bar{\phi})=2r_{H}^{2}e^{-2i\beta(\bar{\theta})}\sum_{m}\int\frac{dk}{2\pi}\Theta_{k}(\bar{\theta})(-k^{2}-ig_{H}k)\zeta_{k}e^{ikx}e^{-ikt}e^{im\bar{\phi}}\,. (66)

Now that we have obtained ψ0\psi_{0} and ψ4\psi_{4} using the Rindler approximations. We would like to relate ζ\zeta to the Teukolsky reflectivity \mathcal{R}. To accomplish this, recall that in Sec. II we have obtained a reflection relation (29) between ψ0\psi_{0} and ψ4\psi_{4} on the ECO surface. Since the relation is written in the Boyer-Lindquist coordinates, we first perform the coordinate transformations on ψ0\psi_{0} and ψ4\psi_{4} according to x=r+δ(θ)x=r_{*}+\delta(\theta)-\mathcal{I}, θ¯=θ\bar{\theta}=\theta, and ϕ¯=ϕΩHt\bar{\phi}=\phi-\Omega_{H}t. During the coordinate transformation, we have used the near-horizon approximations and discarded all 𝒪(α2)\mathcal{O}(\alpha^{2}) terms. The results are given by

ψ\displaystyle\psi (v,θ,φ+)0=8rH2e2iβ(θ)Δ2mdk2πΘk(θ)(k2+igHk)ei(k+mΩH)veikδ(θ)eikeimφ+,{}_{0}(v,\theta,\varphi_{+})=\frac{8r^{2}_{H}e^{2i\beta(\theta)}}{\Delta^{2}}\sum_{m}\int\frac{dk}{2\pi}\Theta_{k}(\theta)(-k^{2}+ig_{H}k)e^{-i(k+m\Omega_{H})v}e^{-ik\delta(\theta)}e^{ik\mathcal{I}}e^{im\varphi_{+}}\,, (67)
(ρ4ψ\displaystyle(\rho^{-4}\psi )4(u,θ,φ)=2rH2e2iβ(θ)mdk2πΘk(θ)(k2igHk)ζkei(k+mΩH)ueikδ(θ)eikeimφ.{}_{4})^{*}(u,\theta,\varphi_{-})=2r_{H}^{2}e^{-2i\beta(\theta)}\sum_{m}\int\frac{dk}{2\pi}\Theta_{k}(\theta)(-k^{2}-ig_{H}k)\zeta_{k}e^{-i(k+m\Omega_{H})u}e^{ik\delta(\theta)}e^{-ik\mathcal{I}}e^{im\varphi_{-}}\,. (68)

Using the reflection model (29), we obtain that

k=ζk(kigHk+igH)exp[2ikb+2ikδ(θ)2ik].\displaystyle\mathcal{R}_{k}=\zeta_{k}\left(\frac{-k-ig_{H}}{-k+ig_{H}}\right)\exp\left[2ikb_{*}+2ik\delta(\theta)-2ik\mathcal{I}\right]\,. (69)

Thus once we know ζ\zeta, the Teukolsky reflectivity \mathcal{R} can be readily obtained. Here we point out that the phase factor e2ikbe^{2ikb_{*}} here will cancel the e2ikbe^{-2ikb_{*}} factors in Sec. II.4. This is because in the previous section we chose bb_{*} as the location for the “surface of the ECO”, while in this section, the ECO is embedded into the xx coordinate system, therefore we no longer need to introduce a reference location bb_{*} as “surface of the ECO”. The information of ECO location will now be incorporated into ζk\zeta_{k}.

Before the end of this subsection, let us look at the factor mω\mathcal{M}_{\ell\ell^{\prime}m\omega} in Eq. (36), and see how the mode mixing show up in the reflected waves. We can pull out the angular dependence of this factor by defining

mω=(kigHk+igH)e2ikb2ik^mω,\displaystyle\mathcal{M}_{\ell\ell^{\prime}m\omega}=\left(\frac{-k-ig_{H}}{-k+ig_{H}}\right)e^{2ikb_{*}-2ik\mathcal{I}}\hat{\mathcal{M}}_{\ell\ell^{\prime}m\omega}\,, (70)

where

^mω=0πζω+mΩHeiΦmω(θ)Smω2(θ)Smω2(θ)sinθdθ,\displaystyle\hat{\mathcal{M}}_{\ell\ell^{\prime}m\omega}=\int_{0}^{\pi}\zeta^{*}_{-\omega^{*}+m\Omega_{H}}e^{i\Phi_{m\omega}(\theta)}{}_{-2}S_{\ell^{\prime}m\omega}(\theta){}_{-2}S^{*}_{\ell m\omega}(\theta)\sin\theta\,d\theta\,, (71)

with

Φmω(θ)=2(ωmΩH)δ(θ)+4β(θ).\Phi_{m\omega}(\theta)=2(\omega-m\Omega_{H})\delta(\theta)+4\beta(\theta)\,. (72)

As an example, we simply let ζ=1\zeta=1, thus ^mω\hat{\mathcal{M}}_{\ell\ell^{\prime}m\omega} directly shows the mixing of modes due to the phase. We plot the absolute value of ^mω\hat{\mathcal{M}}_{\ell\ell^{\prime}m\omega} for =2\ell=2, m=2m=2 and for various spin and \ell^{\prime} in Fig. 4. For a=0a=0, we have ^mω=1\hat{\mathcal{M}}_{\ell\ell^{\prime}m\omega}=1, indicating no mode mixing. As we raise the spin, modes get more mixed and the reflected waves attain more contributions from >2\ell^{\prime}>2 modes. This quantitatively shows the mixing of different \ell-modes is a significant feature for reflection of waves near horizon.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 4: The absolute values of the factor ^mω\hat{\mathcal{M}}_{\ell\ell^{\prime}m\omega} for various spin aa and \ell^{\prime}. This factor shows the mixing between different \ell-modes after an incoming single mode gets reflected on the surface of an exotic compact object. Here we have chosen =2\ell=2, m=2m=2 as an example. In general for higher spin the reflected waves gain more contributions from higher \ell^{\prime}-modes, thus the effect of mode mixing are not negligible for rapidly spinning ECOs.

III.2 Position-dependent damping of gravitational waves

We now calculate the reflectivity \mathcal{R} in a simplest setting — by adding dissipating terms in the linearized Einstein equation in the Rindler coordinate system, obtaining ζ\zeta, and then converting to \mathcal{R}. Wang et al. already introduced a model in which wave is damped, by introducing a complex “Young’s modulus” of space-time Wang et al. (2020). They name the reflection coefficient they found as the Boltzman reflectivity. As an alternative approach, let us introduce a position-dependent damping to gravitational waves, which increases as we approach the horizon. This model has the feature of being able to provide more well-posed differential equations.

To do so, we modify the linearized Einstein equation by adding an extra dissipation term, with coupling coefficient ϵ\epsilon, to the equation satisfied by the perturbation \mathcal{H} defined in Eq. (62):

t2ϵegHxt+x2=0.-\partial_{t}^{2}\mathcal{H}-\epsilon e^{-g_{H}x}\partial_{t}\mathcal{H}+\partial_{x}^{2}\mathcal{H}=0\,. (73)

Assuming harmonic time decomposition (x,t)=~(x)eikt\mathcal{H}(x,t)=\tilde{\mathcal{H}}(x)e^{-ikt}, we have

[d2dx2+k2+ikϵegHx]~(x)=0.\left[\frac{d^{2}}{dx^{2}}+k^{2}+ik\epsilon e^{-g_{H}x}\right]\tilde{\mathcal{H}}(x)=0\,. (74)

Here kk has the physical meaning of being the angular frequency of the perturbation measured by FIDOs, before blue shift. The modified Einstein equation then admits a general solution given by

~(x)=C1Γ(1iν)Jiν(1)(z)+C2Γ(1+iν)Jiν(1)(z),\tilde{\mathcal{H}}(x)=C_{1}\Gamma(1-i\nu)J^{(1)}_{-i\nu}(z)+C_{2}\Gamma(1+i\nu)J^{(1)}_{i\nu}(z)\,, (75)

where

ν\displaystyle\nu =2k/gH,\displaystyle=2k/g_{H}\,, z\displaystyle z =2eiπ4gHx2ϵk/gH,\displaystyle=2e^{\frac{i\pi}{4}-\frac{g_{H}x}{2}}\sqrt{\epsilon k}/g_{H}\,, (76)

and Jν(1)(z)J^{(1)}_{\nu}(z) is the Bessel function of the first kind. The appropriate solution which damps on the horizon is given by

C1C2=Γ(1+iν)Γ(1iν)eπν.\frac{C_{1}}{C_{2}}=-\frac{\Gamma(1+i\nu)}{\Gamma(1-i\nu)}e^{-\pi\nu}\,. (77)

Here we shall assume ϵ1\epsilon\ll 1. In this way, there is a region where x1x\ll-1, but still with ϵegHx1\epsilon e^{-g_{H}x}\ll 1. In other words, this is a region very close to the Kerr horizon, but here the damping has not yet turned on. In this region, the damping solution can be written as

~(x)eikx+ζDeikx,\tilde{\mathcal{H}}(x)\propto e^{-ikx}+\zeta_{\rm D}e^{ikx}\,, (78)

with

ζD(k)=Γ(1+2ik/gH)Γ(12ik/gH)e2ikgHlnkgHeπkgHe2ikgHlnϵ,k>0.\displaystyle\zeta_{D}(k)=-\frac{\Gamma(1+2ik/g_{H})}{\Gamma(1-2ik/g_{H})}e^{-\frac{2ik}{g_{H}}\ln\frac{k}{g_{H}}}e^{-\frac{\pi k}{g_{H}}}e^{-\frac{2ik}{g_{H}}\ln\epsilon}\,,\;k>0\,. (79)

Here we have imposed k>0k>0 because the sense of in-going and out-going waves changes for k<0k<0, where we need to write

ζD(k)=ζD(k).\zeta_{D}(-k)=\zeta_{D}^{*}(k)\,. (80)

This is the same form of reflectivity proposed by Wang et al. In Eq. (79), the first factor involving two Γ\Gamma functions is a pure phase factor that has a moderate variation at the scale kgHk\sim g_{H}, and the phase factor e2ikgHlnkgHe^{-\frac{2ik}{g_{H}}\ln\frac{k}{g_{H}}} is similar; the amplitude factor eπkgHe^{-\frac{\pi k}{g_{H}}} provides unity reflectivity for k0k\sim 0 and this reflectivity decreases as |k||k| increases. We plot |ζD(k)||\zeta_{D}(k)| for gH=1g_{H}=1 in Fig. 5.

The final phase factor in ζD\zeta_{D} can be written in the form of

e2ikgHlnϵ=e2ikxeff,xeff=1gHlnϵ.e^{-\frac{2ik}{g_{H}}\ln\epsilon}=e^{-2ikx_{\rm eff}}\,,\quad x_{\rm eff}=\frac{1}{g_{H}}\ln\epsilon\,. (81)

This provides an effective xx location around which most of the wave is reflected — as we can see, we no longer have a single location r=br=b for the ECO surface at which all the waves are reflected. To obtain the reflectivity \mathcal{R}, we simply insert ζD\zeta_{D} into Eq. (69), which adds an additional θ\theta-dependent phase factor.

III.3 GW Propagation in Matter

The damping term in the linearized Einstein equation causes reflection in the near-horizon region. In this subsection we consider another scenario where there exists some effective matter fields in the vicinity of the horizon. The effective stress-energy tensor is denoted as TABeffT_{AB}^{\rm eff}, and its existence may be related to the emergent nature of gravity. We now modify the linearized (1+1)(1+1)-Einstein equation (51) by adding the effective source, and get

t2+x2=16πe2gHxTABeff.-\partial_{t}^{2}\mathcal{H}+\partial_{x}^{2}\mathcal{H}=-16\pi e^{2g_{H}x}T_{AB}^{\rm eff}\,. (82)

In this equation, on the left hand side, we have a freely propagating GW in (1+1)(1+1)-Minkowski spacetime, while on the right hand side, we have the effect of emergent gravity.

III.3.1 Tidal response of matter

We now discuss how TABeffT_{\rm AB}^{\rm eff} should respond to \mathcal{H}. Suppose theses effective degrees of freedom act as matters that stay at rest in the FIDO frame. The ABAB component of the Riemann tensor is given by

RtAtB=12(t2+gHx).R_{tAtB}=\frac{1}{2}\left(-\partial_{t}^{2}+g_{H}\partial_{x}\right)\mathcal{H}\,. (83)

We postulate that the response of the effective matter is given by

TABeff=μ8πRτAτB,T^{\rm eff}_{AB}=\frac{\mu}{8\pi}R_{\tau A\tau B}\,, (84)

where τ\tau is the proper time for the Rindler metric (38), and μ\mu is a physical coupling constant measured in the local Lorentz frame of the FIDO, which can be dependent on the driving frequency felt by the FIDO. Thus we have

TABeff=μ8πα2RtAtB=μ16πα2(t2+gHx).T^{\rm eff}_{AB}=\frac{\mu}{8\pi\alpha^{2}}R_{tAtB}=\frac{\mu}{16\pi\alpha^{2}}\left(-\partial_{t}^{2}+g_{H}\partial_{x}\right)\mathcal{H}\,. (85)

Note that the Einstein’s equation is now modified into

GAB=μRτAτB.G_{AB}=\mu R_{\tau A\tau B}\,. (86)

With the effective stress-energy tensor, the metric equation of motion can now be written as

[(1+μ)t2+gHμx+x2]=0.\left[-(1+\mu)\partial_{t}^{2}+g_{H}\mu\partial_{x}+\partial_{x}^{2}\right]\mathcal{H}=0\,. (87)

Here (1+μ)(1+\mu) acts as the permeability of gravitational waves in matter, and decreases the speed of gravitational waves. Now let us consider two kinds of matter distributions for the exotic compact object.

III.3.2 Homogeneous star

For a simplest model, let us look at a homogeneous star with uniform μ\mu in the interior region. For a frequency independent μ\mu, we can write eikt+ik~x\mathcal{H}\propto e^{-ikt+i\tilde{k}x}, and the modified dispersion relation is given by k~=k~+\tilde{k}=\tilde{k}_{+} or k~\tilde{k}_{-}, with

k~±=igHμ2±(1+μ)k2gH2μ24.\tilde{k}_{\pm}=\frac{ig_{H}\mu}{2}\pm\sqrt{(1+\mu)k^{2}-\frac{g_{H}^{2}\mu^{2}}{4}}\,. (88)

We immediately note that gravitational waves become evanescent when

|k||kth|=|gHμ|21+μ.|k|\leq|k_{\rm th}|=\frac{|g_{H}\mu|}{2\sqrt{1+\mu}}\,. (89)

That is, we have a total reflection of all waves below ωth\omega_{\rm th}. Substantial reflection also takes place near the ωth\omega_{\rm th} frequency. For |k|>kth|k|>k_{\rm th} and positive μ\mu, waves will be amplified when propagating towards the xx\rightarrow-\infty direction, i.e., towards the horizon.

We may further postulate that μ\mu is of order unity inside a surface at which the surface gravity is blue-shifted to the Planck frequency ωP\omega_{P}:

μ={μ0,α1gH>ωP,0,otherwise.\mu=\left\{\begin{array}[]{ll}\mu_{0}\,,&\alpha^{-1}g_{H}>\omega_{P}\,,\\ \\ 0\,,&\mbox{otherwise}\,.\end{array}\right. (90)

The surface is then located at x=xPx=x_{P}, where

xP=1gHln(gHωP).x_{P}=\frac{1}{g_{H}}\ln\left(\frac{g_{H}}{\omega_{P}}\right)\,. (91)

As before, we write down the general solutions to Eq. (87) as (x,t)=~(x)eikt\mathcal{H}(x,t)=\tilde{\mathcal{H}}(x)e^{-ikt}. Outside the surface, we can write

~(x)eikx+ζMeikx.\tilde{\mathcal{H}}(x)\propto e^{-ikx}+\zeta_{M}e^{ikx}\,. (92)

Inside the surface we have

~(x)eik~x.\tilde{\mathcal{H}}(x)\propto e^{i\tilde{k}_{-}x}\,. (93)

Matching the solutions on the surface gives

ζM=(k+igHμ2(1+μ)k2gH2μ24kigHμ2+(1+μ)k2gH2μ24)e2ikxP,k>0.\zeta_{M}=\left(\frac{k+\frac{ig_{H}\mu}{2}-\sqrt{(1+\mu)k^{2}-\frac{g_{H}^{2}\mu^{2}}{4}}}{k-\frac{ig_{H}\mu}{2}+\sqrt{(1+\mu)k^{2}-\frac{g_{H}^{2}\mu^{2}}{4}}}\right)e^{-2ikx_{P}}\,,\quad k>0\,. (94)

Similarly to the previous section, ζM(k)=ζ(k)\zeta_{M}(-k)=\zeta^{*}(k). For |k|kth|k|\leq k_{\rm th}, we have |ζM|=1|\zeta_{M}|=1, indicating a total reflection of low frequency waves. For higher frequencies, |ζM||\zeta_{M}| approaches a constant

limk|ζM|=1+μ11+1+μ.\lim_{k\rightarrow\infty}|\zeta_{M}|=\frac{\sqrt{1+\mu}-1}{1+\sqrt{1+\mu}}\,. (95)

We plot |ζM||\zeta_{M}| for different μ\mus in Fig. 5. Since μ\mu is supposed to be a small number, high frequency waves have nearly zero reflection near the surface. This ζM\zeta_{M} is qualitatively similar to the Lorentzian reflectivity model adopted, e.g., by Ref. Du and Chen (2018).

Refer to caption
Figure 5: Absolute values of ζD\zeta_{D} and ζM\zeta_{M} as functions of the frequency kk. We have set gH=1g_{H}=1. Since ζ\zeta and the Teukolsky reflectivity \mathcal{R} only differ by a phase, |ζ||\zeta| is the same as |||\mathcal{R}|. The blue solid line represents |ζD||\zeta_{D}|. The yellow dotted line, the green dashed line, the red dot-dashed line, and the purple long dashed line give |ζM||\zeta_{M}| for μ=0.1,0.2,0.5,\mu=0.1,0.2,0.5,\infty respectively. The “Boltzman” reflectivity, i.e. ξD\xi_{D}, exponentially decays for higher frequencies. For our model of homogeneous stars, we have total reflection of waves on the ECO surface below a certain threshold frequency. Beyond the threshold frequency, the reflectivity gets decreased and converges to a constant. When μ\mu\rightarrow\infty, we have total reflection of waves for all the frequency range, which is equivalent to the case of inhomogeneous stars we have introduced.

III.3.3 Inhomogeneous star

Let us make μ\mu grow as a function of the location, with

μ=μ0eηx,\mu=\mu_{0}e^{-\eta x}\,, (96)

where μ0\mu_{0} and η\eta are positive constants. In this way, we successfully “revive” μ\mu near the horizon.

We write down the general solutions to Eq. (87) as (x,t)=~(x)eikt\mathcal{H}(x,t)=\tilde{\mathcal{H}}(x)e^{-ikt}, and obtain that

~(x)=\displaystyle\tilde{\mathcal{H}}(x)= A1eikxeikηln(μ0gHη)πkηM(a,b,z)+\displaystyle A_{1}e^{-ikx}e^{\frac{ik}{\eta}\ln\left(\frac{\mu_{0}g_{H}}{\eta}\right)-\frac{\pi k}{\eta}}M(a,b,z)+ (97)
A2eikxeikηln(μ0gHη)+πkηM(a,b,z),\displaystyle A_{2}e^{ikx}e^{-\frac{ik}{\eta}\ln\left(\frac{\mu_{0}g_{H}}{\eta}\right)+\frac{\pi k}{\eta}}M(a^{*},b^{*},z)\,,

where

a\displaystyle a =ikηk2gHη,\displaystyle=\frac{ik}{\eta}-\frac{k^{2}}{g_{H}\eta}\,, b\displaystyle b =1+2ikη,\displaystyle=1+\frac{2ik}{\eta}\,, z=μ0gHηeηx,\displaystyle z=\frac{\mu_{0}g_{H}}{\eta}e^{-\eta x}\,, (98)

A1A_{1}, A2A_{2} are some constants, and M(a,b,z)M(a,b,z) is the confluent hypergeometric function.

The hypergeometric function behaves asymptotically as

M(a,b,z)\displaystyle M(a,b,z) ezzabΓ(b)Γ(a),\displaystyle\sim e^{z}\,z^{a-b}\frac{\Gamma(b)}{\Gamma(a)}\,, z\displaystyle z ,\displaystyle\rightarrow\infty\,, (99)
M(a,b,z)\displaystyle M(a,b,z) 1,\displaystyle\sim 1\,, z\displaystyle z 0.\displaystyle\rightarrow 0\,. (100)

The solution that damps on the horizon is then given by

A2A1=e2πωηΓ(b)Γ(b)Γ(a)Γ(a).\frac{A_{2}}{A_{1}}=-e^{-\frac{2\pi\omega}{\eta}}\frac{\Gamma(b)}{\Gamma(b^{*})}\frac{\Gamma(a^{*})}{\Gamma(a)}\,. (101)

For xx in the region that μ0gHeηx1\mu_{0}g_{H}e^{-\eta x}\ll 1 and positive kk, this solution can then be written as

~(x)=eikx+ζNeikx,\tilde{\mathcal{H}}(x)=e^{-ikx}+\zeta_{N}e^{ikx}\,, (102)

where

ζN=\displaystyle\zeta_{N}= e[2ikηln(μ0gHη)]Γ(1+2ikη)Γ(12ikη)Γ(ikηk2gHη)Γ(ikηk2gHη).\displaystyle-e^{\left[-\frac{2ik}{\eta}\ln\left(\frac{\mu_{0}g_{H}}{\eta}\right)\right]}\frac{\Gamma\left(1+\frac{2ik}{\eta}\right)}{\Gamma\left(1-\frac{2ik}{\eta}\right)}\frac{\Gamma(-\frac{ik}{\eta}-\frac{k^{2}}{g_{H}\eta})}{\Gamma(\frac{ik}{\eta}-\frac{k^{2}}{g_{H}\eta})}\,. (103)

One immediately notes that |ζN|=1|\zeta_{N}|=1 for all real kk, indicating a total reflection of waves. This may due to the fact that our assumption of μ\mu in  (96) is equivalent to putting infinite numbers of reflecting surfaces near the horizon, i.e. the μ\mu\rightarrow\infty case in Fig. 5.

IV Boundary Condition in terms of Various Functions

In calculations for gravitational waveforms, one does not usually compute both ψ0\psi_{0} and ψ4\psi_{4}; the Sasaki-Nakamura formalism was also used to obtain faster numerical convergence. In this section, let us convert our boundary condition (35), which involves both ψ0\psi_{0} and ψ4\psi_{4} amplitudes, into those that only involve ψ4\psi_{4} amplitudes, and compare our reflectivity with the one defined using the Sasaki-Nakamura functions.

IV.1 Reflectivity for ψ4\psi_{4} mode amplitudes

The Newman-Penrose quantities ψ0\psi_{0} and ψ4\psi_{4} can be tranformed into each other using the Teukolsky-Starobinsky identities. The amplitude ZholeZ^{\rm hole} and YholeY^{\rm hole} are related by Teukolsky and Press (1974)

CmωYmωhole=DmωZmωhole,C_{\ell m\omega}Y_{\ell m\omega}^{\rm hole}=D_{\ell m\omega}Z^{\rm hole}_{\ell m\omega}\,, (104)

with

Dmω=64(2rH)4(ik)(k2+4ε2)(ik+4ε),D_{\ell m\omega}=64(2r_{H})^{4}(ik)\left(k^{2}+4\varepsilon^{2}\right)\left(-ik+4\varepsilon\right)\,, (105)

and CC is given by

|Cmω\displaystyle|C_{\ell m\omega} |2=((λ+2)2+4aωm4a2ω2)\displaystyle|^{2}=\left((\lambda+2)^{2}+4a\omega m-4a^{2}\omega^{2}\right) (106)
×[λ2+36aωm36a2ω2]\displaystyle\times\left[\lambda^{2}+36a\omega m-36a^{2}\omega^{2}\right]
+(2λ+3)(96a2ω248aωm)+144ω2(1a2),\displaystyle+(2\lambda+3)(96a^{2}\omega^{2}-48a\omega m)+144\omega^{2}(1-a^{2})\,,

with

ImC\displaystyle\text{Im}\,C =12ω,\displaystyle=12\omega\,, (107)
ReC\displaystyle\text{Re}\,C =+|C|2(ImC)2.\displaystyle=+\sqrt{|C|^{2}-(\text{Im}\,C)^{2}}\,. (108)

Here we have defined

ε=1a24rH,\varepsilon=\frac{\sqrt{1-a^{2}}}{4r_{H}}\,, (109)

and λλmω2\lambda\equiv{}_{-2}\lambda_{\ell m\omega} is the eigenvalue of the s=2s=-2 spin-weighted spheroidal harmonic. See Appendix. B for more discussions on the Teukolsky-Starobinsky identity.

Combining Eq. (35) with Eq. (104), we finally arrive at the relation between ZreflZ^{\rm refl} and ZinZ^{\rm in}, which is given by

Zmωrefl=𝒢mωZ-m-ωhole,Z^{\rm refl}_{\ell m\omega}=\sum_{\ell^{\prime}}\mathcal{G}_{\ell\ell^{\prime}m\omega}Z^{\rm hole^{*}}_{\ell^{\prime}\,\text{-}m\,\text{-}\omega^{*}}\,, (110)

where

𝒢mω=(1)m+114e2ikbmωDmωCmω.\mathcal{G}_{\ell\ell^{\prime}m\omega}=(-1)^{m+1}\frac{1}{4}e^{-2ikb_{*}}\mathcal{M}_{\ell\ell^{\prime}m\omega}\frac{D_{\ell^{\prime}m\omega}}{C_{\ell^{\prime}m\omega}}\,. (111)

We have used the relations Dmω=D-m-ωD_{\ell m\omega}=D^{*}_{\ell\,\text{-}m\,\text{-}\omega^{*}} and Cmω=C-m-ωC_{\ell m\omega}=C^{*}_{\ell\,\text{-}m\,\text{-}\omega^{*}} in order to obtain the above equation.

If we restrict ourselves to the simple case where \ell- and \ell^{\prime}- modes do not mix up, we may simply write Eq. (110) as

Zmωrefl\displaystyle Z^{\rm refl}_{\ell m\omega} =𝒢^mωZ-m-ωhole,\displaystyle=\hat{\mathcal{G}}_{\ell m\omega}\,Z^{\rm hole^{*}}_{\ell\,\text{-}m\,\text{-}\omega}\,, (112)

with

𝒢^mω\displaystyle\hat{\mathcal{G}}_{\ell m\omega} Dmω4Cmω-ω+mΩHeiφmωrefl,\displaystyle\equiv\frac{D_{\ell m\omega}}{4C_{\ell m\omega}}\,\mathcal{R}^{*}_{\text{-}\omega+m\Omega_{H}}\,e^{i\varphi^{\rm refl}_{\ell m\omega}}\,, (113)

and

φmωrefl=(m+1)π2kb.\varphi^{\rm refl}_{\ell m\omega}=(m+1)\pi-2kb_{*}\,. (114)

Eq. (112) says that, the (,m,ω)(\ell,\,m,\,\omega)-mode of gravitational-wave echoes are not induced by the reflection of the incoming (,m,ω)(\ell,\,m,\,\omega)-mode but the (,m,ω)(\ell,\,-m,\,-\omega^{*})-mode instead. The mixing of these two types of modes essentially indicates the breaking of isospectrality as pointed out by Ref. Maggio et al. (2020). We will get back to this point later. The other new result is the extra phase term φrefl\varphi^{\rm refl} for the reflected waves, which may be important for observations.

IV.2 Reflectivity for Sasaki-Nakamura Mode Amplitudes

Since most previous literatures on gravitational wave echoes base their models on the reflection of Sasaki-Nakamura (SN) functions, one may ask how the tidal reflectivity can be related to the SN reflectivity. (See Appendix A for a brief review of the SN formalism.) In the vicinity of the horizon, the s=2s=-2 SN function, i.e., the one associated with ψ4\psi_{4}, can be written as

XmωECO\displaystyle X^{\rm ECO}_{\ell m\omega} =ξmωholeeikr+ξmωrefleikr.\displaystyle=\xi^{\rm hole}_{\ell m\omega}\,e^{-ikr_{*}}+{\xi^{\rm refl}_{\ell m\omega}}\,e^{ikr_{*}}\,. rb.\displaystyle r_{*}\rightarrow b_{*}\,. (115)

Under the Chandrasekhar-Sasaki-Nakamura transformation, we have

ξmωhole=Zmωholedmω,ξmωrefl=Zmωreflfmω\xi^{\rm hole}_{\ell m\omega}=Z^{\rm hole}_{\ell m\omega}d_{\ell m\omega}\,,\quad{\xi^{\rm refl}_{\ell m\omega}}=\frac{Z^{\rm refl}_{\ell m\omega}}{f_{\ell m\omega}} (116)

with

dmω=2rH[\displaystyle d_{\ell m\omega}=\sqrt{2r_{H}}[ (824iω16ω2)rH2\displaystyle(8-24i\omega-16\omega^{2})r^{2}_{H}
+(12iam16+16amω+24iω)rH\displaystyle+(12iam-16+16am\omega+24i\omega)r_{H}
4a2m212iam+8],\displaystyle-4a^{2}m^{2}-12iam+8]\,, (117)

and

fmω=4k2rH[2krH+i(rH1)]η(rH).\displaystyle f_{\ell m\omega}=-\frac{4k\sqrt{2r_{H}}\left[2kr_{H}+i(r_{H}-1)\right]}{\eta(r_{H})}\,. (118)

Inserting Eqs. (116) into Eq. (110), we obtain boundary condition for the mω\ell m\omega components of the SN functions:

ξmωrefl=(1)m+1e2ikb4fmωmωDmωCmωdmωξ-m-ωhole,\xi_{\ell m\omega}^{\rm refl}=\frac{(-1)^{m+1}e^{-2ikb_{*}}}{4f_{\ell m\omega}}\sum_{\ell^{\prime}}\mathcal{M}_{\ell\ell^{\prime}m\omega}\frac{D_{\ell^{\prime}m\omega}}{C_{\ell^{\prime}m\omega}d_{\ell^{\prime}m\omega}}\xi^{\rm hole^{*}}_{\ell^{\prime}\,\text{-}m\,\text{-}\omega^{*}}\,, (119)

Here we have used the identity that

dmω=d-m-ω.d_{\ell m\omega}=d_{\ell\,\text{-}m\,\text{-}\omega^{*}}^{*}\,. (120)

As we will see later, the fact that reflection at the ECO surface turns the in-going (,m,ω)(\ell,-m,-\omega) SN components into out-going (l,m,ω)(l,m,\omega) SN components leads to the breaking of isospectrality, which has also been pointed out by Maggio et al. Maggio et al. (2020); here we take the further step of relating these coefficients to the tidal response of the ECO.

For the most simplified scenario where Z-m-ωhole=ZmωholeZ^{\rm hole^{*}}_{\ell\,\text{-}m\,\text{-}\omega^{*}}=Z^{\rm hole}_{\ell m\omega} and different \ell^{\prime}-modes do not mix, we may simply write

ξmωrefl=lmoSNξmωhole,\xi_{\ell m\omega}^{\rm refl}=\mathcal{R}_{lmo}^{\rm SN}\xi^{\rm hole}_{\ell m\omega}\,, (121)

where

mωSN=𝒦mωTSN-ω+mΩH,\mathcal{R}^{\rm SN}_{\ell m\omega}=\mathcal{K}_{\ell m\omega}^{\rm T\rightarrow SN}\mathcal{R}^{*}_{\text{-}\omega+m\Omega_{H}}\,, (122)

with

𝒦mωTSN=(1)m+1Dmω4Cmωfmωdmω.\mathcal{K}_{\ell m\omega}^{\rm T\rightarrow SN}=\frac{(-1)^{m+1}D_{\ell m\omega}}{4C_{\ell m\omega}f_{\ell m\omega}d_{\ell m\omega}}\,. (123)

This is a simple linear factor that converts \mathcal{R} into the SN\mathcal{R}^{\rm SN} that are used in SN calculations. In the Schwarzschild limit, we have

𝒦mωTSN=(1)m(4ωi)[12ω+iλ(λ+2)](4ω+i)[12ωiλ(λ+2)],a=0,\mathcal{K}_{\ell m\omega}^{\rm T\rightarrow SN}=\frac{(-1)^{m}(4\omega-i)\left[12\omega+i\lambda(\lambda+2)\right]}{(4\omega+i)\left[12\omega-i\lambda(\lambda+2)\right]}\,,\quad a=0\,, (124)

where λ=(1)(+2)\lambda=(\ell-1)(\ell+2). One immediately notes that |𝒦mωTSN|=1|\mathcal{K}_{\ell m\omega}^{\rm T\rightarrow SN}|=1 in the Schwarzschild limit. For spinning ECOs, we numerically investigate 𝒦mωTSN\mathcal{K}_{\ell m\omega}^{\rm T\rightarrow SN} for the (2,2)(2,2) mode for different spins in Fig. 6.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 6: Conversion factor 𝒦mωTSN\mathcal{K}_{\ell m\omega}^{\rm T\rightarrow SN} from the Teukolsky \mathcal{R} to the Sasaki-Nakamura SN\mathcal{R}_{\mathrm{SN}}. Here we have ignored the \ell-\ell^{\prime} mode mixing. We plot real and imaginary parts, as well as the modulus, of 𝒦mωTSN\mathcal{K}_{\ell m\omega}^{\rm T\rightarrow SN} for the (2,2)(2,2)-mode with a=0,0.3,0.7,1a=0,0.3,0.7,1 respectively. The gray dot-dashed line marks the horizon frequency mΩHm\Omega_{H}. In the Schwarzschild case, the two reflection coefficients only differ by a phase. For Kerr spacetimes we have |𝒦mωTSN|>1|\mathcal{K}_{\ell m\omega}^{\rm T\rightarrow SN}|>1 for both low and high frequencies, but |𝒦mωTSN||\mathcal{K}_{\ell m\omega}^{\rm T\rightarrow SN}| dips below 1 for some frequencies. Also note that |𝒦mωTSN|=1|\mathcal{K}_{\ell m\omega}^{\rm T\rightarrow SN}|=1 when ω\omega equals to the horizon frequency.

IV.3 Energy Contents of Incoming and Reflected Waves

The reflection coefficient we defined in last subsection is indeed the (square root of) power reflectivity of the gravitational waves on the ECO boundary. To see this, consider a solution to s=2s=-2 Teukolsky equation near the ECO surface. The energy flux down to the surface is given by Teukolsky and Press (1974)

dEholedω=mω64πk(k2+4ϵ2)(2rH)3|Ymωhole|2,\frac{dE_{\rm hole}}{d\omega}=\sum_{\ell m}\frac{\omega}{64\pi k(k^{2}+4\epsilon^{2})(2r_{H})^{3}}|Y^{\rm hole}_{\ell m\omega}|^{2}\,, (125)

while the energy propagating outward from the surface is given by

dErefldω=mω4πk(k2+4ϵ2)(2rH)3|Zmωrefl|2.\frac{dE^{\rm refl}}{d\omega}=\sum_{\ell m}\frac{\omega}{4\pi k(k^{2}+4\epsilon^{2})(2r_{H})^{3}}|Z^{\rm refl}_{\ell m\omega}|^{2}\,. (126)

Here ω\omega are all taken as real numbers. See Appendix B for detailed discussions on the energy flux and the energy conservation law. In the simple case of neglecting \ell-\ell^{\prime} mixing, incoming energy from the (-m-ω)(\ell\,\text{-}m\,\text{-}\omega)-mode will return from the (mω)(\ell m\omega)-mode, with

(dErefldω)mω=|-ω+mΩH|2(dEholedω)-m-ω\left(\frac{dE_{\rm refl}}{d\omega}\right)_{\ell m\omega}=|\mathcal{R}_{\text{-}\omega+m\Omega_{H}}|^{2}\left(\frac{dE_{\rm hole}}{d\omega}\right)_{\ell\,\text{-}m\,\text{-}\omega} (127)

This means our reflectivity \mathcal{R} indeed acts as an energy reflectivity.

V Waveforms and Quasi-Normal Modes of the ECO

In this section, we show how our ECO boundary conditions can be applied to echo computations and resonant conditions for quasi-normal modes. We shall also restrict ourselves to the case of

𝒢^mω=𝒢^-m-ω.\hat{\mathcal{G}}_{\ell m\omega}=\hat{\mathcal{G}}^{*}_{\ell\,\text{-}m\,\text{-}\omega^{*}}\,. (128)

This is satisfied by all the reflectivity models discussed in this paper, since in these cases the tidal response in the time-domain, (b,θ;t)\mathcal{R}(b,\theta;t) [Cf. (29)] is real-valued.

V.1 Even and Odd-Parity Echoes

In this subsection, we derive the gravitational-wave echo waveform based on our reflection model. Note that this echo can be the additional wave due to the reflection at the ECO surface during the inspiral phase — it does not necessarily has to be the echo that follows the ringdown phase of the coalescence wave.

Suppose now, we have some small perturbations towards the ECO spacetime. We assume that the source in the Teukolsky equation drives a Υ(0)2{}_{-2}\Upsilon^{(0)}, which has the following form at rr_{*}\rightarrow-\infty:

Υ(0)2=mdω2πZmωhole(0)Δ2eikrSmω2(θ,ϕ)eiωt.\displaystyle{}_{-2}\Upsilon^{\rm(0)}=\sum_{\ell m}\int\frac{d\omega}{2\pi}Z_{\ell m\omega}^{\rm hole\,(0)}\Delta^{2}e^{-ikr_{*}}{}_{-2}S_{\ell m\omega}(\theta,\phi)e^{-i\omega t}\,. (129)

This satisfies the Teukolsky equation with the appropriate source term away from the horizon, the out-going condition at infinity, but not the ECO boundary condition near the horizon. We will need to add an additional homogeneous solution, which satisfies the out-going boundary condition at infinity. Recall that for the radial part, we have

Rmω+={DmωinΔ2eikr+Dmωouteikr,rb,r3eiωr,r+.\displaystyle{R}^{+\infty}_{\ell m\omega}=\left\{\begin{array}[]{ll}D^{\rm in}_{\ell m\omega}\Delta^{2}e^{-ikr_{*}}+D^{\rm out}_{\ell m\omega}e^{ikr_{*}}\,,&r\rightarrow b\,,\\ \\ r^{3}e^{i\omega r_{*}}\,,&r\rightarrow+\infty\,.\end{array}\right.

Thus we add the following homogeneous solution to Υ(0)\Upsilon^{\rm(0)}:

Υecho2=mdω2πcmωRmω+Smω2(θ,ϕ)eiωt,{}_{-2}\Upsilon^{\rm echo}=\sum_{\ell m}\int\frac{d\omega}{2\pi}c_{\ell m\omega}R_{\ell m\omega}^{+\infty}{}_{-2}S_{\ell m\omega}(\theta,\phi)e^{-i\omega t}\,, (131)

so that Υ(0)2+Υecho2{}_{-2}\Upsilon^{(0)}+{}_{-2}\Upsilon^{\rm echo} is of the form (14), also satisfying (112). The asymptotic behavior of Υecho2{}_{-2}\Upsilon^{\rm echo} is given by

Υecho2={mdω2πcmωr3e+iωreiωtSmω2(θ,ϕ),r+,mdω2πcmω[DmωinΔ2eikr+Dmωouteikr]eiωtSmω2(θ,ϕ),rb.{}_{-2}\Upsilon^{\rm echo}=\left\{\begin{array}[]{ll}\displaystyle\sum_{\ell m}\int\frac{d\omega}{2\pi}c_{\ell m\omega}r^{3}e^{+i\omega r_{*}}e^{-i\omega t}{}_{-2}S_{\ell m\omega}(\theta,\phi)\,,&r_{*}\rightarrow+\infty\,,\\ \\ \displaystyle\sum_{\ell m}\int\frac{d\omega}{2\pi}c_{\ell m\omega}\left[D^{\rm in}_{\ell m\omega}\Delta^{2}e^{-ikr_{*}}+D^{\rm out}_{\ell m\omega}e^{ikr_{*}}\right]e^{-i\omega t}{}_{-2}S_{\ell m\omega}(\theta,\phi)\,,&r_{*}\rightarrow b_{*}\,.\\ \end{array}\right. (132)

Here we already see that the amplitudes clmωc_{lm\omega} directly give us the additional gravitational waves due to the reflecting surface. Identifying term by term between Υ(0)2+Υecho2{}_{-2}\Upsilon^{(0)}+{}_{-2}\Upsilon^{\rm echo} and Eq. (14), we find

Zmωhole=Zmωhole(0)+cmωDmωin,Zmωrefl=cmωDmωout.Z^{\rm hole}_{\ell m\omega}=Z^{\rm hole\,(0)}_{\ell m\omega}+c_{\ell m\omega}D^{\rm in}_{\ell m\omega}\,,\quad Z^{\rm refl}_{\ell m\omega}=c_{\ell m\omega}D^{\rm out}_{\ell m\omega}\,. (133)

Applying Eq. (110), we obtain

cmωDmωout\displaystyle c_{\ell m\omega}D^{\rm out}_{\ell m\omega} =𝒢mω[Z-m-ωhole(0)+c-m-ωD-m-ωin],\displaystyle=\sum_{\ell^{\prime}}\mathcal{G}_{\ell\ell^{\prime}m\omega}\left[Z^{\rm hole\,(0)*}_{\ell^{\prime}\,\text{-}m\,\text{-}\omega}+c^{*}_{\ell^{\prime}\,\text{-}m\,\text{-}\omega}D^{\rm in\,*}_{\ell^{\prime}\,\text{-}m\,\text{-}\omega}\right]\,,
c-m-ωD-m-ωout\displaystyle c_{\ell\,\text{-}m\,\text{-}\omega}^{*}D^{\rm out\,*}_{\ell\,\text{-}m\,\text{-}\omega} =𝒢-m-ω[Zmωhole(0)+cmωDmωin].\displaystyle=\sum_{\ell^{\prime}}\mathcal{G}^{*}_{\ell\ell^{\prime}\text{-}m\text{-}\omega}\left[Z^{\rm hole\,(0)}_{\ell^{\prime}m\omega}+c_{\ell^{\prime}m\omega}D^{\rm in}_{\ell^{\prime}m\omega}\right]\,. (134)

Here we restrict ourselves to real-valued ω\omega only. Using the symmetry of the Teukolsky equation, for real-valued ω\omega, it is straightforward to show that the homogeneous solutions have the symmetry that

Dmωin=D-m-ωin,Dmωout=D-m-ωout.D^{\rm in}_{\ell m\omega}=D^{\rm in\,*}_{\ell\,\text{-}m\,\text{-}\omega}\,,\quad D^{\rm out}_{\ell m\omega}=D^{\rm out\,*}_{\ell\,\text{-}m\,\text{-}\omega}\,. (135)

We can then write

(δDmωout𝒢mωDmωin𝒢¯mωDmωinδDmωout)\displaystyle\left(\begin{array}[]{cc}\delta_{\ell\ell^{\prime}}D^{\rm out}_{\ell m\omega}&-\mathcal{G}_{\ell\ell^{\prime}m\omega}D^{\rm in}_{\ell m\omega}\\ -\overline{\mathcal{G}}_{\ell\ell^{\prime}m\omega}D^{\rm in}_{\ell m\omega}&\delta_{\ell\ell^{\prime}}D^{\rm out}_{\ell m\omega}\end{array}\right) (cmωc-m-ω)=(𝒢mω00𝒢¯mω)(Z-m-ωin(0)Zmωin(0)),\displaystyle\left(\begin{array}[]{c}c_{\ell^{\prime}m\omega}\\ c_{\ell^{\prime}\,\text{-}m\,\text{-}\omega}^{*}\end{array}\right)=\left(\begin{array}[]{cc}\mathcal{G}_{\ell\ell^{\prime}m\omega}&0\\ 0&\overline{\mathcal{G}}_{\ell\ell^{\prime}m\omega}\end{array}\right)\left(\begin{array}[]{c}Z^{\rm in\,(0)*}_{\ell^{\prime}\,\text{-}m\,\text{-}\omega}\\ Z^{\rm in\,(0)}_{\ell^{\prime}m\omega}\end{array}\right)\,, (144)
𝒢¯mω𝒢-m-ω,\displaystyle\overline{\mathcal{G}}_{\ell\ell^{\prime}m\omega}\equiv\mathcal{G}_{\ell\ell^{\prime}\,\text{-}m\,\text{-}\omega}^{*}\,, (145)

where the components in all matrices are also block matrices with \ell and \ell^{\prime} representing sections of rows and columns. This will allow us to solve for cmωc_{\ell m\omega}, therefore leading to the additional out-going waves at infinity, i.e. the gravitational-wave echoes.

In the simple case where there is no \ell-\ell^{\prime} mixing for reflected waves (so that the relation between reflected waves and incoming waves is simply given by Eq. (112)), and that

𝒢^-m-ω𝒢^mω,\hat{\mathcal{G}^{*}}_{\ell\,\text{-}m\,\text{-}\omega}\equiv\hat{\mathcal{G}}_{\ell m\omega}\,, (146)

we can have simpler results. For each harmonic for the ZZ components (similar for the cc components), we can define symmetric and anti-symmetric quadrature amplitudes

Zmωhole(0),S\displaystyle Z^{\rm hole\,(0),S}_{\ell m\omega} Zmωhole(0)+Z-m-ωhole(0)2,\displaystyle\equiv\frac{Z^{\rm hole\,(0)}_{\ell m\omega}+Z^{\rm hole\,(0)\,*}_{\ell\,\text{-}m\,\text{-}\omega}}{\sqrt{2}}\,, (147)
Zmωhole(0),A\displaystyle Z^{\rm hole\,(0),A}_{\ell m\omega} Zmωhole(0)Z-m-ωhole(0)2i.\displaystyle\equiv\frac{Z^{\rm hole\,(0)}_{\ell m\omega}-Z^{\rm hole\,(0)\,*}_{\ell\,\text{-}m\,\text{-}\omega}}{\sqrt{2}i}\,. (148)

We then have

cmωS\displaystyle c_{\ell m\omega}^{\rm S} =𝒢^mωDmωout𝒢^mωDmωinZmωhole(0),S,\displaystyle=\frac{\hat{\mathcal{G}}_{\ell m\omega}}{D_{\ell m\omega}^{\rm out}-\hat{\mathcal{G}}_{\ell m\omega}D_{\ell m\omega}^{\rm in}}Z_{\ell m\omega}^{\rm hole\,(0),S}\,, (149)
cmωA\displaystyle c_{\ell m\omega}^{\rm A} =𝒢^mωDmωout+𝒢^mωDmωinZmωhole(0),A.\displaystyle=-\frac{\hat{\mathcal{G}}_{\ell m\omega}}{D_{\ell m\omega}^{\rm out}+\hat{\mathcal{G}}_{\ell m\omega}D_{\ell m\omega}^{\rm in}}Z_{\ell m\omega}^{\rm hole\,(0),A}\,. (150)

Here we see that the AA quadrature has a reflectivity of 𝒢^mω-\hat{\mathcal{G}}_{\ell m\omega}, compared with 𝒢^mω\hat{\mathcal{G}}_{\ell m\omega} for the SS quadrature. These quadratures correspond to electric- and magnetic-type perturbations.

As it turns out, non-spinning binaries, or those with spins aligned with the orbital angular momentum, only excite the SS quadrature — although generically both quadratures are excited — they will have different echoes. In the case when echoes are well-separated in the time domain, the first, third, and other odd echoes, the AA and SS will have transfer functions negative to each other, while for even echoes, they will have the same transfer function.

If we further simplify the problem by demanding cmω=c-m-ωc_{\ell m\omega}=c^{*}_{\ell\,\text{-}m\,\text{-}\omega}, Eq. (144) gives that

cmω=𝒢^mωDmωout𝒢^mωDmωinZmωhole(0).c_{\ell m\omega}=\frac{\hat{\mathcal{G}}_{\ell m\omega}}{D^{\rm out}_{\ell m\omega}-\hat{\mathcal{G}}_{\ell m\omega}\,D^{\rm in}_{\ell m\omega}}Z^{\rm hole\,(0)}_{\ell m\omega}\,. (151)

This expression coincides, for instance, with the one obtained in Mark et al. (2017) for the spherically-symmetric spacetime with a reflecting surface. Note that the phase factor e2ikbe^{-2ikb_{*}} has been absorbed into our definition of 𝒢^\hat{\mathcal{G}}.

V.2 Echoes driven by symmetric source terms

In our reflection model (112), as discussed in Ref. Hughes (2000b), the coefficients Z-m-ωholeZ^{\rm hole^{*}}_{\ell\,\text{-}m\,\text{-}\omega^{*}} and ZmωholeZ^{\rm hole}_{\ell m\omega} are related for quasi-circular orbits. For such orbits, one can define a series of frequencies as

ωmk=mΩϕ+kΩθ,\omega_{mk}=m\Omega_{\phi}+k\Omega_{\theta}\,, (152)

where Ωϕ\Omega_{\phi} and Ωθ\Omega_{\theta} are two fundamental frequencies defined for periodic motions in ϕ\phi and θ\theta. Then, for real frequencies, we can decompose the amplitude ZmωinZ^{\rm in}_{\ell m\omega} according to

Zmωhole=kZmkholeδ(ωωmk).Z^{\rm hole}_{\ell m\omega}=\sum_{k}Z^{\rm hole}_{\ell mk}\delta(\omega-\omega_{mk})\,. (153)

It is easy to check that for Kerr black holes,

Z-m-khole=(1)+kZmkhole.Z^{\rm hole^{*}}_{\ell\text{-}m\text{-}k}=(-1)^{\ell+k}Z^{\rm hole}_{\ell mk}\,. (154)

That is, if we consider a specific circular orbit, we have the symmetry that Z-m-ωholeZ^{\rm hole^{*}}_{\ell\,\text{-}m\,\text{-}\omega^{*}} is either equal to ZmωholeZ^{\rm hole}_{\ell m\omega}, or they differ by a minus sign. In this simple case, our reflection model (112) does not involve different modes, and the model becomes similar to those reflection models based on Sasaki-Nakamura functions like in Ref. Wang et al. (2020). However, if we consider the full quasi-circular motions, i.e. adding up all orbits, this symmetry no longer exists, and one has to consider the mixing of modes when dealing with the reflecting boundary. For general orbits that are not quasi-circular, the symmetry between Z-m-ωholeZ^{\rm hole^{*}}_{\ell\,\text{-}m\,\text{-}\omega^{*}} and ZmωholeZ^{\rm hole}_{\ell m\omega} may not exist.

Now for the symmetric source, where there is no mode mixing, let us consider a solution Υ(0)2{}_{-2}\Upsilon^{(0)} to the Teukolsky equation, which has the following form at rr_{*}\rightarrow-\infty:

Υ(0)2=mdω2πZmωhole(0)Δ2eikrSmω2(θ,ϕ)eiωt.\displaystyle{}_{-2}\Upsilon^{\rm(0)}=\sum_{\ell m}\int\frac{d\omega}{2\pi}Z_{\ell m\omega}^{\rm hole\,(0)}\Delta^{2}e^{-ikr_{*}}{}_{-2}S_{\ell m\omega}(\theta,\phi)e^{-i\omega t}\,. (155)

Following the same steps as in the last subsection, it is straightforward to show that the echo solution to the Teukolsky equation at infinity is given by

Υecho2=mdω2πZmωechoSmω2(θ,ϕ)eiωt,{}_{-2}\Upsilon^{\rm echo}=\sum_{\ell m}\int\frac{d\omega}{2\pi}Z^{\rm echo}_{\ell m\omega}{}_{-2}S_{\ell m\omega}(\theta,\phi)e^{-i\omega t}\,, (156)

with

Zmωecho=𝒢^mωDmωout𝒢^mωDmωinZmωhole(0),Z^{\rm echo}_{\ell m\omega}=\frac{\hat{\mathcal{G}}_{\ell m\omega}}{D^{\rm out}_{\ell m\omega}-\hat{\mathcal{G}}_{\ell m\omega}\,D^{\rm in}_{\ell m\omega}}Z^{\rm hole\,(0)}_{\ell m\omega}\,, (157)

where we have chosen the normalization Dmω=1D^{\infty}_{\ell m\omega}=1. The tidal reflectivity can also be directly related to the SN reflectivity as

mωSN=(1)m+1Dmω4Cmωfmωdmω-ω+mΩH.\mathcal{R}^{\rm SN}_{\ell m\omega}=(-1)^{m+1}\frac{D_{\ell m\omega}}{4C_{\ell m\omega}f_{\ell m\omega}d_{\ell m\omega}}\mathcal{R}^{*}_{\text{-}\omega+m\Omega_{H}}\,. (158)

In this simple scenario, the tidal reflectivity is exactly the energy reflectivity for each mode.

V.3 Quasi-Normal Modes and Breakdown of Isospectrality

For Quasi-Normal Modes, we set ZZ to zero, and analytically continue Eq. (144) to complex ω\omega. The QNM frequencies can be directly solved by setting the determinant of the lhs matrix of Eq. (144) to zero, i.e.

det(δDmωout𝒢mωDmωin𝒢-m-ωDmωinδDmωout)=0.\text{det}\left(\begin{array}[]{cc}\delta_{\ell\ell^{\prime}}D^{\rm out}_{\ell m\omega}&-\mathcal{G}_{\ell\ell^{\prime}m\omega}D^{\rm in}_{\ell m\omega}\\ -\mathcal{G}_{\ell\ell^{\prime}\,\text{-}m\,\text{-}\omega^{*}}^{*}D^{\rm in}_{\ell m\omega}&\delta_{\ell\ell^{\prime}}D^{\rm out}_{\ell m\omega}\end{array}\right)=0\,. (159)

This will in general cause a mixing between QNMs with different \ell, and break the isospectrality property of the Kerr spacetime and lead to two distinct QNMs for each (,m)(\ell,m).

Neglecting the \ell-\ell^{\prime} mixing, we can simply write

[Dmωout]2\displaystyle\left[D^{\rm out}_{\ell m\omega}\right]^{2} =𝒢^mω𝒢^¯mω[Dmωin]2,\displaystyle=\hat{\mathcal{G}}_{\ell m\omega}\bar{\hat{\mathcal{G}}}_{\ell m\omega}\left[D^{\rm in}_{\ell m\omega}\right]^{2}\,, 𝒢^¯mω𝒢^-m-ω.\displaystyle\bar{\hat{\mathcal{G}}}_{\ell m\omega}\equiv\hat{\mathcal{G}^{*}}_{\ell\,\text{-}m\,\text{-}\omega^{*}}\,. (160)

In the special case of 𝒢^mω=𝒢^¯mω\hat{\mathcal{G}}_{\ell m\omega}=\bar{\hat{\mathcal{G}}}_{\ell m\omega} (which is satisfied by all the reflectivity models discussed in this paper), we note that the ECO’s QNMs split into SS and AA modes, with ωnmS\omega_{n\ell m}^{S} and ωnmA\omega_{n\ell m}^{A} satisfying different equations:

DmωSout𝒢^mωSDmωSin\displaystyle D_{\ell m\omega_{S}}^{\rm out}-\hat{\mathcal{G}}_{\ell m\omega_{S}}D_{\ell m\omega_{S}}^{\rm in} =0,\displaystyle=0\,, (161)
DmωAout+𝒢^mωADmωAin\displaystyle D_{\ell m\omega_{A}}^{\rm out}+\hat{\mathcal{G}}_{\ell m\omega_{A}}D_{\ell m\omega_{A}}^{\rm in} =0.\displaystyle=0\,. (162)

This still breaks the isospectrality properties of Kerr spacetime. Note that this property has also been found and studied in Ref. Maggio et al. (2020) with their echo model which describes the ECO as a dissipative fluid. Since modes of the ECO are usually excited collectively, the main signature of the breakdown of isospectrality is still the fact that SS and AA echoes have alternating sign differences in even and odd echoes.

VI Conclusions

In this paper, we developed a more physical way to impose boundary conditions for Teukolsky functions near the surface of extremely compact objects. We adopted the Membrane Paradigm, and assumed that the ECO structure is well adapted to the coordinate system of the Fiducial Observers, which is an approximate Rindler coordinate system near the horizon. More specifically, assuming that the additional physics near an ECO can be viewed as modified propagation laws of gravitational waves in the Rindler coordinate system, we were able to obtain reflectivity models for spinning ECOs that are similar to those proposed by previous literature, when taking the Schwarzschild limit. In particular, the Boltzmann reflectivity of Oshita et al. was obtainable from a position-dependent damping of gravitational waves in the Rindler coordinate system, which might be thought of as due to the emergent nature of gravity.

As it has turned out, the most directly physical condition is between in-going components of ψ0\psi_{0} and out-going components of ψ4\psi_{4}, although relations between in-going and out-going components of ψ4\psi_{4}, as well as those of the Sasaki-Nakamura functions, can be obtained by using the Starobinsky-Teukolsky transformation, as well as the Chandrasekhar-Sasaki-Nakamura relations.

The deformation of space-time geometry due to the spin of the ECO causes a mixing between different \ell modes during reflection at the ECO surface; reflection at the ECO also takes (m,ω)(m,ω)(m,\omega)\rightarrow(-m,-\omega^{*}). This means an incoming (,m,ω)(\ell,m,\omega) mode is reflected into (,m,ω)(\ell^{\prime},-m,-\omega^{*}) modes. For moderately rapidly spinning holes, such \ell-\ell^{\prime} mixing is moderate, but non-negligible, which means accurately modeling echoes will indeed have to take such mixing into account. For incoming waves toward the ECO caused by a quasi-circular inspiral of a non-spinnining particle, the waveform has a definite partiy, and is invariant under the (m,ω)(m,ω)(m,\omega)\rightarrow(-m,-\omega^{*}) transformation. For more general waves, the (m,ω)(m,ω)(m,\omega)\rightarrow(-m,-\omega^{*}) map causes echoes from even- and odd-parity waves to differ from each other; it also causes the breakdown of quasi-normal mode isospectrality, as has been pointed out by Maggio et al. in the Schwarzschild case.

Acknowledgements.
The authors would like to thank Shuo Xin, Wenbiao Han, Ka-Lok R. Lo, Ling Sun and Niayesh Afshordi for useful conversations. BC and YC acknowledge the support from the Brinson Foundation, the Simons Foundation (Award Number 568762), and the National Science Foundation, Grants PHY-2011961, PHY-2011968. QW acknowledges the support from the University of Waterloo, Natural Sciences and Engineering Research Council of Canada (NSERC), and the Perimeter Institute for Theoretical Physics.

Appendix A The homogeneous Teukolsky and Sasaki-Nakamura equations

Perturbations of Kerr spacetime can be described by the Teukolsky equations Teukolsky (1973). In the vacuum case, one can decompose solutions to the homogeneous Teukolsky equation as

Υs=mdω2πeiωt+imϕRmωs(r)Smωs(θ),\displaystyle{}_{s}\Upsilon=\sum_{\ell m}\int\frac{d\omega}{2\pi}\,e^{-i\omega t+im\phi}\,{}_{s}R_{\ell m\omega}(r){}_{s}S_{\ell m\omega}(\theta)\,, (163)

where Smωs(θ){}_{s}S_{\ell m\omega}(\theta) is the spin-weighted spheroidal harmonic function, and ss is the spin weight. The Teukolsky equations are then separable, and the equations for RR and SS are respectively

[Δsddr(Δs+1ddr)+K22is(r1)KΔ+4isωrλmωs]Rmωs=0,\displaystyle\left[\Delta^{-s}\frac{d}{dr}\left(\Delta^{s+1}\frac{d}{dr}\right)+\frac{K^{2}-2is\left(r-1\right)K}{\Delta}+4is\omega r-{}_{s}\lambda_{\ell m\omega}\right]{}_{s}R_{\ell m\omega}=0\,, (164)
[1sinθddθ(sinθddθ)a2ω2sin2θ(m+scosθ)2sin2θ2aωscosθ+s+2maω+λmωs]Smωs=0,\displaystyle\left[\frac{1}{\sin\theta}\frac{d}{d\theta}\left(\sin\theta\frac{d}{d\theta}\right)-a^{2}\omega^{2}\sin^{2}\theta-\frac{(m+s\cos\theta)^{2}}{\sin^{2}\theta}-2a\omega s\cos\theta+s+2ma\omega+{}_{s}\lambda_{\ell m\omega}\right]{}_{s}S_{\ell m\omega}=0\,, (165)

where K=(r2+a2)ωmaK=(r^{2}+a^{2})\omega-ma, and λmωs{}_{s}\lambda_{\ell m\omega} is the eigenvalue of the spin-weighted spheroidal harmonic.

For s=2s=-2, the radial equation (164) admits two independent solutions, RmωH2{}_{-2}R^{\rm H}_{\ell m\omega} and Rmω2{}_{-2}R^{\infty}_{\ell m\omega}, which have the following asymptotic forms:

RmωH2\displaystyle{}_{-2}R^{\rm H}_{\ell m\omega} ={Bmωoutr3eiωr+Bmωinr1eiωr,r,BmωholeΔ2eikr,rrH;\displaystyle=\left\{\begin{array}[]{lr}B^{\rm out}_{\ell m\omega}r^{3}e^{i\omega r_{*}}+B^{\rm in}_{\ell m\omega}r^{-1}e^{-i\omega r_{*}}\,,&r\rightarrow\infty\,,\\ \\ B^{\rm hole}_{\ell m\omega}\Delta^{2}e^{-ikr_{*}}\,,&r\rightarrow r_{H}\,;\end{array}\right. (169)
Rmω2\displaystyle{}_{-2}R^{\rm\infty}_{\ell m\omega} ={Dmωr3eiωr,r,Dmωouteikr+DmωinΔ2eikr,rrH.\displaystyle=\left\{\begin{array}[]{lr}D^{\infty}_{\ell m\omega}r^{3}e^{i\omega r_{*}}\,,&r\rightarrow\infty\,,\\ \\ D^{\rm out}_{\ell m\omega}e^{ikr_{*}}+D^{\rm in}_{\ell m\omega}\Delta^{2}e^{-ikr_{*}}\,,&r\rightarrow r_{H}\,.\end{array}\right. (173)

The Sasaki-Nakamura-Chandrashekar transformation Sasaki and Nakamura (1982b) takes the Teukolsky radial function R2(r){}_{-2}R(r) to the Sasaki-Nakamura function X(r)X(r), and the Teukolsky equation becomes the Sasaki-Nakamura equation. The homogeneous SN equation is given by

d2Xmωdr2F(r)dXmωdrU(r)Xmω=0.\frac{d^{2}X_{\ell m\omega}}{dr^{2}_{*}}-F(r)\frac{dX_{\ell m\omega}}{dr_{*}}-U(r)X_{\ell m\omega}=0\,. (174)

The explicit expressions for F(r)F(r) and U(r)U(r) are given in Ref. Sasaki and Tagoshi (2003)’s Eqs. (51-58). The SN equation also admits two independent solutions, XmωHX^{\rm H}_{\ell m\omega} and XmωX^{\infty}_{\ell m\omega}, which have the asymptotic values:

XmωH\displaystyle X^{\rm H}_{\ell m\omega} ={Amωouteiωr+Amωineiωr,r,Amωholeeikr,rrH;\displaystyle=\left\{\begin{array}[]{lr}A^{\rm out}_{\ell m\omega}e^{i\omega r_{*}}+A^{\rm in}_{\ell m\omega}e^{-i\omega r_{*}}\,,&r\rightarrow\infty\,,\\ \\ A^{\rm hole}_{\ell m\omega}e^{-ikr_{*}}\,,&r\rightarrow r_{H}\,;\end{array}\right. (178)
Xmω\displaystyle X^{\rm\infty}_{\ell m\omega} ={Cmωeiωr,r,Cmωouteikr+Cmωineikr,rrH.\displaystyle=\left\{\begin{array}[]{lr}C^{\infty}_{\ell m\omega}e^{i\omega r_{*}}\,,&r\rightarrow\infty\,,\\ \\ C^{\rm out}_{\ell m\omega}e^{ikr_{*}}+C^{\rm in}_{\ell m\omega}e^{-ikr_{*}}\,,&r\rightarrow r_{H}\,.\end{array}\right. (182)

The amplitudes AA and CC can be related to the amplitudes BB and DD by matching the asymptotic solutions to the SN and the Teukolsky equation on the horizon and at infinity. The BB-coefficients and AA-coefficients are related by

Bmωin\displaystyle B^{\rm in}_{\ell m\omega} =14ω2Amωin,\displaystyle=-\frac{1}{4\omega^{2}}A^{\rm in}_{\ell m\omega}\,, (183)
Bmωout\displaystyle B^{\rm out}_{\ell m\omega} =4ω2c0Amωout,\displaystyle=-\frac{4\omega^{2}}{c_{0}}A^{\rm out}_{\ell m\omega}\,, (184)
Bmωhole\displaystyle B^{\rm hole}_{\ell m\omega} =1dmωAmωhole,\displaystyle=\frac{1}{d_{\ell m\omega}}A^{\rm hole}_{\ell m\omega}\,, (185)

and the DD-coefficients and CC-coefficients are related by

Dmωin\displaystyle D^{\rm in}_{\ell m\omega} =1dmωCmωin,\displaystyle=\frac{1}{d_{\ell m\omega}}C^{\rm in}_{\ell m\omega}\,, (186)
Dmωout\displaystyle D^{\rm out}_{\ell m\omega} =fmωCmωout,\displaystyle=f_{\ell m\omega}C^{\rm out}_{\ell m\omega}\,, (187)
Dmω\displaystyle D^{\infty}_{\ell m\omega} =4ω2c0Cmω,\displaystyle=-\frac{4\omega^{2}}{c_{0}}C^{\infty}_{\ell m\omega}\,, (188)

where

dmω=2rH[\displaystyle d_{\ell m\omega}=\sqrt{2r_{H}}[ (824iω16ω2)rH2\displaystyle(8-24i\omega-16\omega^{2})r^{2}_{H} (189)
+(12iam16+16amω+24iω)rH\displaystyle+(12iam-16+16am\omega+24i\omega)r_{H}
4a2m212iam+8],\displaystyle-4a^{2}m^{2}-12iam+8]\,,

and

fmω=4k2rH[2krH+i(rH1)]η(rH).\displaystyle f_{\ell m\omega}=-\frac{4k\sqrt{2r_{H}}\left[2kr_{H}+i(r_{H}-1)\right]}{\eta(r_{H})}\,. (191)

Here η(r)\eta(r) is defined by

η(r)=c0+c1r+c2r2+c3r3+c4r4,\displaystyle\eta(r)=c_{0}+\frac{c_{1}}{r}+\frac{c_{2}}{r^{2}}+\frac{c_{3}}{r^{3}}+\frac{c_{4}}{r^{4}}\,, (192)

with

c0\displaystyle c_{0} =12iω+λ(λ+2)12aω(aωm),\displaystyle=-12i\omega+\lambda(\lambda+2)-12a\omega(a\omega-m)\,, (193)
c1\displaystyle c_{1} =8ia[3aωλ(aωm)],\displaystyle=8ia[3a\omega-\lambda(a\omega-m)]\,,
c2\displaystyle c_{2} =24ia(aωm)+12a2[12(aωm)2],\displaystyle=-24ia(a\omega-m)+12a^{2}[1-2(a\omega-m)^{2}]\,, (194)
c3\displaystyle c_{3} =24ia3(aωm)24a2,\displaystyle=24ia^{3}(a\omega-m)-24a^{2}\,,
c4\displaystyle c_{4} =12a4,\displaystyle=12a^{4}\,,

where λλmω2\lambda\equiv{}_{-2}\lambda_{\ell m\omega} is the eigenvalue of the s=2s=-2 spin-weighted spheroidal harmonic.

Appendix B Conservation of energy for gravitational perturbations

In this section, we derive a new conservation relation among four energies, which corresponds to waves that are outgoing at infinity, ingoing at infinity, coming down to the “horizon”, and being reflected from the “horizon”, respectively. A derivation has been done by Teukolsky and Press in Teukolsky and Press (1974) for the relation among the first three energies. Here we extend their results to include the reflected one.

From the Newman-Penrose equations, one can derive the Teukolsky-Starobinsky identities for s=±2s=\pm 2, which can be written as

1012S2+12iωS2\displaystyle\mathscr{L}_{-1}\mathscr{L}_{0}\mathscr{L}_{1}\mathscr{L}_{2}\,{}_{2}S+12i\omega{}_{-2}S =CS2,\displaystyle=C{}_{-2}S\,, (195)
𝒟𝒟𝒟𝒟R2\displaystyle\mathscr{D}\mathscr{D}\mathscr{D}\mathscr{D}{}_{-2}R =14R2,\displaystyle=\frac{1}{4}\,{}_{2}R\,, (196)

where we have omitted (mω)(\ell m\omega)-indices in RR and SS for the sake of brevity. We will adopt these abbreviated notations throughout this section. The operators \mathscr{L} and 𝒟\mathscr{D} are defined by

n\displaystyle\mathscr{L}_{n} =θ+mcscθaωsinθ+ncotθ,\displaystyle=\partial_{\theta}+m\csc\theta-a\omega\sin\theta+n\cot\theta\,, (197)
𝒟\displaystyle\mathscr{D} =riK/Δ,\displaystyle=\partial_{r}-iK/\Delta\,, (198)

and CC is given by

|C\displaystyle|C |2=((λ+2)2+4aωm4a2ω2)\displaystyle|^{2}=\left((\lambda+2)^{2}+4a\omega m-4a^{2}\omega^{2}\right) (199)
×[λ2+36aωm36a2ω2]\displaystyle\times\left[\lambda^{2}+36a\omega m-36a^{2}\omega^{2}\right]
+(2λ+3)(96a2ω248aωm)+144ω2(1a2),\displaystyle+(2\lambda+3)(96a^{2}\omega^{2}-48a\omega m)+144\omega^{2}(1-a^{2})\,,

with

ImC\displaystyle\text{Im}\,C =12ω,\displaystyle=12\omega\,, (200)
ReC\displaystyle\text{Re}\,C =+|C|2(ImC)2.\displaystyle=+\sqrt{|C|^{2}-(\text{Im}\,C)^{2}}\,. (201)

Similarly we define

n\displaystyle\mathscr{L}^{\dagger}_{n} =n(ω,m),\displaystyle=\mathscr{L}_{n}(-\omega,-m)\,, (202)
𝒟\displaystyle\mathscr{D}^{\dagger} =𝒟(ω,m)=r+iK/Δ.\displaystyle=\mathscr{D}(-\omega,-m)=\partial_{r}+iK/\Delta\,. (203)

A complementary set of equations to Eqs. (195) and  (196) then gives

1012CS2\displaystyle\mathscr{L}^{\dagger}_{-1}\mathscr{L}^{\dagger}_{0}\mathscr{L}^{\dagger}_{1}\mathscr{L}^{\dagger}_{2}C{}_{-2}S +12iωCS2=C2S2,\displaystyle+12i\omega C^{*}{}_{2}S=C^{2}{}_{2}S\,, (204)
𝒟𝒟𝒟𝒟Δ2R2\displaystyle\mathscr{D}^{\dagger}\mathscr{D}^{\dagger}\mathscr{D}^{\dagger}\mathscr{D}^{\dagger}\Delta^{2}\,{}_{2}R =4|C|2Δ2R2.\displaystyle\!=\!4|C|^{2}\Delta^{-2}{}_{-2}R\,. (205)

Now let us derive the relation between ψ0\psi_{0} and ψ4\psi_{4} by using the Teukolsky-Starobinsky identities. Note that at large rr, the radial function Rs{}_{s}R has the following asymptotic behavior,

R2\displaystyle{}_{2}R =Yineiωrr+Youteiωrr5,\displaystyle=Y_{\rm in}\frac{e^{-i\omega r_{*}}}{r}+Y_{\rm out}\frac{e^{i\omega r_{*}}}{r^{5}}\,, (206)
R2\displaystyle{}_{-2}R =Zineiωrr+Zoutr3eiωr.\displaystyle=Z_{\rm in}\frac{e^{-i\omega r_{*}}}{r}+Z_{\rm out}r^{3}e^{i\omega r_{*}}\,. (207)

Plugging these asymptotic expressions into Eq. (196), and keep the terms leading in (1/r)(1/r)-expansions, we have

CYin=64ω4Zin.CY_{\rm in}=64\omega^{4}Z_{\rm in}\,. (208)

A set of useful identities can be used during the derivations are

Δ𝒟𝒟=\displaystyle\Delta\mathscr{D}\mathscr{D}=  2(iK+r1)𝒟+6iωr+λ,\displaystyle\,2(-iK+r-1)\mathscr{D}+6i\omega r+\lambda\,, (209)
Δ2𝒟𝒟𝒟=\displaystyle\Delta^{2}\mathscr{D}\mathscr{D}\mathscr{D}= [4iK(iKr+1)+(λ+2+2iωr)Δ]𝒟\displaystyle\left[4iK(iK-r+1)+(\lambda+2+2i\omega r)\Delta\right]\mathscr{D} (210)
2iK(λ+6iωr)+6iωΔ,\displaystyle-2iK(\lambda+6i\omega r)+6i\omega\Delta\,,
Δ3𝒟𝒟𝒟𝒟=\displaystyle\Delta^{3}\mathscr{D}\mathscr{D}\mathscr{D}\mathscr{D}= [Δ(4iK(λ+2)8iωr(r1))\displaystyle\,\left[\Delta\left(-4iK(\lambda+2)-8i\omega r(r-1)\right)\right. (211)
+8iK(K2+(r1)2)+8iωΔ2]𝒟\displaystyle+\left.8iK\left(K^{2}+(r-1)^{2}\right)+8i\omega\Delta^{2}\right]\mathscr{D}
+Δ[(λ+22iωr)(λ+6irω)\displaystyle+\Delta\left[(\lambda+2-2i\omega r)(\lambda+6ir\omega)\right.
12iω(iK+r1)]\displaystyle\left.-12i\omega(iK+r-1)\right]
+4iK(iK+r1)(λ+6irω).\displaystyle+4iK(iK+r-1)(\lambda+6ir\omega)\,.

Similarly, plugging the asymptotic expressions of the radial functions R±2{}_{\pm 2}R into Eq. (205), we obtain that

4ω4Yout=CZout.4\omega^{4}Y_{\rm out}=C^{*}Z_{\rm out}\,. (212)

On the horizon, the radial function Rs{}_{s}R is given by

R2\displaystyle{}_{2}R =YholeΔ2eikr+Yrefleikr,\displaystyle=Y_{\rm hole}\Delta^{-2}e^{-ikr_{*}}+Y_{\rm refl}e^{ikr_{*}}\,, (213)
R2\displaystyle{}_{-2}R =ZholeΔ+2eikr+Zrefleikr.\displaystyle=Z_{\rm hole}\Delta^{+2}e^{-ikr_{*}}+Z_{\rm refl}e^{ikr_{*}}\,. (214)

Plugging these expressions into Eq. (196) and (205), we obtain that

CYhole=64(2rH)4(ik)(ik+4ϵ)(k2+4ϵ2)Zhole,\displaystyle\!\!CY_{\rm hole}=64(2r_{H})^{4}(ik)(-ik+4\epsilon)(k^{2}+4\epsilon^{2})Z_{\rm hole}\,, (215)
4(2rH)4(ik)(ik4ϵ)(k2+4ϵ2)Yrefl=CZrefl.\displaystyle 4(2r_{H})^{4}(ik)(-ik-4\epsilon)(k^{2}+4\epsilon^{2})Y_{\rm refl}=C^{*}Z_{\rm refl}\,. (216)

In the Schwarzschild case, the energy conservation relations can be most easily seen from the Wronskian of two linearly independent homogeneous solutions to the perturbation equations such as the Regge-Wheeler equation. In the Teukolsky equation, due to the existence of the dR/drdR/dr_{*}-term, the Wronskian is then dependent on rr. To resolve this, one can rewrite the radial Teukolsky equation (164) in a form of

d2Y/dr2+VY=0,d^{2}Y/dr^{2}_{*}+VY=0\,, (217)

which is possible if one defines

Y\displaystyle Y =Δs/2(r2+a2)1/2R,\displaystyle=\Delta^{s/2}(r^{2}+a^{2})^{1/2}R\,, (218)
V\displaystyle V =[K22isK(r1)+Δ(4irωsλ2)s2(1a2)](r2+a2)2Δ(2r3+a2r24ra2+a4)(r2+a2)4.\displaystyle=\frac{\left[K^{2}-2isK(r-1)+\Delta(4ir\omega s-\lambda-2)-s^{2}(1-a^{2})\right]}{(r^{2}+a^{2})^{2}}-\frac{\Delta(2r^{3}+a^{2}r^{2}-4ra^{2}+a^{4})}{(r^{2}+a^{2})^{4}}\,. (219)

The Wronskian of any two solutions of Eq. (217) is then conserved. By equating the Wronskian evaluated at infinity and that on the horizon, we have

(dYsdrYsYsdYsdr)r=rH=(dYsdrYsYsdYsdr)r=.\left(\frac{d{}_{s}Y}{dr_{*}}{}_{-s}Y^{*}-{}_{s}Y\frac{d{}_{-s}Y^{*}}{dr_{*}}\right)_{r=r_{H}}=\left(\frac{d{}_{s}Y}{dr_{*}}{}_{-s}Y^{*}-{}_{s}Y\frac{d{}_{-s}Y^{*}}{dr_{*}}\right)_{r=\infty}\,. (220)

For s=2s=2, we substitute Eqs. (206), (207), (213) and (214) into the Wronskian equation, and we use Eqs. (208), (212), (215) and (216) to obtain that

iC|Yhole|232k(2rH)3(k2+4ϵ2)+256(ik)rH5(k2+4ϵ2)(k2+16ϵ2)|Yrefl|2C=iC|Yin|232ω3+8iω5|Yout|2C,\frac{-iC^{*}|Y_{\rm hole}|^{2}}{32k(2r_{H})^{3}(k^{2}+4\epsilon^{2})}+\frac{256(ik)r^{5}_{H}(k^{2}+4\epsilon^{2})(k^{2}+16\epsilon^{2})|Y_{\rm refl}|^{2}}{C}=\frac{-iC^{*}|Y_{\rm in}|^{2}}{32\omega^{3}}+\frac{8i\omega^{5}|Y_{\rm out}|^{2}}{C}\,, (221)

where ϵ\epsilon is defined in Eq. (109).

This is indeed the energy conservation law relating the ingoing energy at infinity EinE_{\rm in}, the outgoing energy at infinity EoutE_{\rm out}, the energy absorbed by the “horizon” EholeE_{\rm hole}, and the energy reflected from the horizon EreflE_{\rm refl}. The conservation law can be written as

dEindωdEoutdω=dEholedωdErefldω,\frac{dE_{\rm in}}{d\omega}-\frac{dE_{\rm out}}{d\omega}=\frac{dE_{\rm hole}}{d\omega}-\frac{dE_{\rm refl}}{d\omega}\,, (222)

in which the explicit expressions for the four energies are

dEindω\displaystyle\frac{dE_{\rm in}}{d\omega} =m164πω2|Yin|2=m64ω6π|C|2|Zin|2,\displaystyle=\sum_{\ell m}\frac{1}{64\pi\omega^{2}}|Y_{\rm in}|^{2}=\sum_{\ell m}\frac{64\omega^{6}}{\pi|C|^{2}}|Z_{\rm in}|^{2}\,, (223)
dEoutdω\displaystyle\frac{dE_{\rm out}}{d\omega} =m14πω2|Zout|2=m4ω6π|C|2|Yin|2,\displaystyle=\sum_{\ell m}\frac{1}{4\pi\omega^{2}}|Z_{\rm out}|^{2}=\sum_{\ell m}\frac{4\omega^{6}}{\pi|C|^{2}}|Y_{\rm in}|^{2}\,, (224)
dEholedω\displaystyle\frac{dE_{\rm hole}}{d\omega} =mω64πk(k2+4ϵ2)(2rH)3|Yhole|2\displaystyle=\sum_{\ell m}\frac{\omega}{64\pi k(k^{2}+4\epsilon^{2})(2r_{H})^{3}}|Y_{\rm hole}|^{2} (225)
=m64ωk(k2+4ϵ2)(k2+16ϵ2)(2rH)5π|C|2|Zhole|2,\displaystyle=\sum_{\ell m}\frac{64\omega k(k^{2}+4\epsilon^{2})(k^{2}+16\epsilon^{2})(2r_{H})^{5}}{\pi|C|^{2}}|Z_{\rm hole}|^{2}\,, (226)
dErefldω\displaystyle\frac{dE_{\rm refl}}{d\omega} =mω4πk(k2+4ϵ2)(2rH)3|Zrefl|2,\displaystyle=\sum_{\ell m}\frac{\omega}{4\pi k(k^{2}+4\epsilon^{2})(2r_{H})^{3}}|Z_{\rm refl}|^{2}\,, (227)
=m4ωk(k2+4ϵ2)(k2+16ϵ2)(2rH)5π|C|2|Yrefl|2.\displaystyle=\sum_{\ell m}\frac{4\omega k(k^{2}+4\epsilon^{2})(k^{2}+16\epsilon^{2})(2r_{H})^{5}}{\pi|C|^{2}}|Y_{\rm refl}|^{2}\,. (228)

References