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


Unruh-DeWitt Detector Differentiation of Black Holes and Exotic Compact Objects

Bob Holdom [email protected] Department of Physics, University of Toronto, 60 St. George St., Toronto, Ontario, Canada M5S 1A7    Robert B. Mann [email protected] Department of Physics and Astronomy, University of Waterloo, Waterloo, Ontario, Canada, N2L 3G1    Chen Zhang [email protected] Department of Physics, University of Toronto, 60 St. George St., Toronto, Ontario, Canada M5S 1A7 Department of Physics and Astronomy, University of Waterloo, Waterloo, Ontario, Canada, N2L 3G1
Abstract

We study the response of a static Unruh-DeWitt detector outside an exotic compact object (ECO) with a general reflective boundary condition in 3+1 dimensions. The horizonless ECO, whose boundary is extremely close to the would-be event horizon, acts as a black hole mimicker. We find that the response rate is notably distinct from the black hole case, even when the ECO boundary is perfectly absorbing. For a (partially) reflective ECO boundary, we find resonance structures in the response rate that depend on the different locations of the ECO boundary and those of the detector. We provide a detailed analysis in connection with the ECO’s vacuum mode structure and transfer function.

I Introduction

The discovery of gravitational waves from compact binary mergers have inspired many recent studies on exotic compact objects (ECOs), whose defining feature is their compactness: their radius is very close to that of a black hole with the same mass whilst lacking an event horizon. Explicit examples of such compact horizonless objects include gravastars Mazur:2001fv , boson stars Schunck:2003kk , wormholes, and 2-2 holes Holdom:2016nek . While some investigations are concerned with the consistency of semi-classical physics near horizons Baccetti:2018otf ; Terno:2020tsq , most studies are concerned with finding distinctive signatures from postmerger gravitational wave echoes Abedi:2016hgu ; Mark:2017dnq ; Conklin:2017lwb ; Cardoso:2019rvt ; Holdom:2019bdv , in which a wave that falls inside the gravitational potential barrier reflects off the ECO boundary and returns to the barrier after some time delay tdt_{d}.

In this paper, we study the response of an Unruh-DeWitt (UDW) detector unruh ; deWitt in this context. Assuming no angular momentum exchange, the light-matter interaction is well described as a local interaction between a scalar quantum field ϕ(x)\phi(x) and a local detector comprising two states: |0\left|0\right\rangle, with vanishing state energy, and |Ω\left|\Omega\right\rangle of energy Ω\Omega, where Ω\Omega may be a positive or negative real number Funai:2018wqq . If Ω>0\Omega>0 then |0\left|0\right\rangle is the ground state. Known as a UDW detector, the interaction Hamiltonian is

Hint=λχ(τ)μ(τ)ϕ(𝗑(τ)),H_{\text{int}}=\lambda\chi(\tau)\mu(\tau)\phi(\mathsf{x}(\tau))\,, (1)

where χ(τ)\chi(\tau) is the switching function encoding the time dependence of the interaction, whose coupling constant is λ\lambda, and 𝗑(τ)\mathsf{x}(\tau) is the detector’s trajectory with τ\tau its proper time. The quantity μ(τ)\mu({\tau}) is the monopole operator of the detector μ=eiΩτσ++eiΩτσ\mu=e^{i\Omega\tau}\sigma^{+}+e^{-i\Omega\tau}\sigma^{-} with σ+=|Ω0|\sigma^{+}=\left|\Omega\right\rangle\left\langle 0\right| and σ=|0Ω|\sigma^{-}=\left|0\right\rangle\left\langle\Omega\right| the respective raising and lowering operators.

Perturbing to first-order in λ\lambda, the transition probability is byd

P(Ω)=c2|0|μ(0)|Ω|2(Ω),P(\Omega)=c^{2}{|\langle 0|\mu(0)|\Omega\rangle|}^{2}\mathcal{F}\left(\Omega\right)\ , (2)

where (Ω)\mathcal{F}(\Omega) is the response function

(Ω)=dτdτ′′χ(τ)χ(τ′′)eiΩ(ττ′′)W(τ,τ′′).\mathcal{F}(\Omega)=\int_{-\infty}^{\infty}\mathrm{d}\tau^{\prime}\,\int_{-\infty}^{\infty}\mathrm{d}\tau^{\prime\prime}\,\chi(\tau^{\prime})\chi(\tau^{\prime\prime})\,\mathrm{e}^{-i\Omega(\tau^{\prime}-\tau^{\prime\prime})}\,W(\tau^{\prime},\tau^{\prime\prime})\,. (3)

W(τ,τ′′)=0|ϕ(𝗑(τ))ϕ(𝗑(τ′′))|0W(\tau^{\prime},\tau^{\prime\prime})=\langle 0|\phi(\mathsf{x}(\tau^{\prime}))\phi(\mathsf{x}(\tau^{\prime\prime}))|0\rangle is the pull-back of the Wightman distribution on the detector’s worldline. When the switching function is a Heaviside step function whose switch-on time is taken to be in the asymptotic past, we can drop the external τ\tau^{\prime}-integral of (3) by a change of variables and obtain the transition rate byd ; Louko:2007mu

˙(Ω)=dseiΩsW(s),\mathcal{\dot{F}}\left(\Omega\right)=\int^{\infty}_{-\infty}\,\mathrm{d}s\,\mathrm{e}^{-i\Omega s}\,W(s)\,, (4)

representing the number of particles detected per unit proper time.

The spacetime of a Schwarzschild black hole in 3+13+1 spacetime dimensions has the metric form (in geometric units where c=G=1c=G=1)

ds2=(12Mr)dt2+(12Mr)1dr2+r2(dθ2+sin2θdϕ2),ds^{2}=-\left(1-\frac{2M}{r}\right)dt^{2}+\left(1-\frac{2M}{r}\right)^{-1}dr^{2}+r^{2}\left(d\theta^{2}+\sin^{2}{\theta}d\phi^{2}\right), (5)

where MM is the black hole mass, with the event horizon at rH=2Mr_{H}=2M. Mode solutions of the Klein-Gordon equation for the massless scalar field ϕ\phi in this background have the form

ϕ=14πωρω(r)rYm(θ,ϕ)eiωt,\phi=\frac{1}{\sqrt{4\pi\omega}}\frac{\rho_{\omega\ell}(r)}{r}Y_{\ell m}\left(\theta,\phi\right)\operatorname{e}^{-i\omega t}, (6)

where YmY_{\ell m} is the spherical harmonic function. The radial function ρω\rho_{\omega\ell} satisfies

d2ρωdr2+{ω2V(r)}ρω=0,\frac{d^{2}\rho_{\omega\ell}}{dr^{*2}}+\left\{\omega^{2}-V(r)\right\}\rho_{\omega\ell}=0\,, (7)

where the tortoise coordinate rr^{*} satisfies dr/dr=(12M/r)1dr^{*}/dr=(1-2M/r)^{-1}. The gravitational potential is

V(r)=(12Mr)[(+1)r2+2Mr3]V(r)=\left(1-\frac{2M}{r}\right)\left[\frac{\ell(\ell+1)}{r^{2}}+\frac{2M}{r^{3}}\right] (8)

for the scalar field perturbation of angular momentum ll. V(r)V(r) has a peak at rpeakr_{\rm peak}, which is close to r=3Mr=3M. Note that we can rescale r¯=r/(2M),ω¯=2Mω\bar{r}=r/(2M),\,\bar{\omega}=2M\omega so that the dependence on MM in all equations is removed.

We model the ECO as an exterior Schwarzschild spacetime patched to a spherically symmetric interior metric at radius r=r0r=r_{0}, with r0/rH11r_{0}/r_{H}-1\ll 1, with a boundary condition at r0r_{0} parameterized by the reflection coefficient RWR^{W}. Typical quantum gravity arguments suggest that r0rHr_{0}-r_{H} is related to the Planck length, as for the brick wall model tHooft:1984kcu . The space between r0r_{0} and rpeakr_{\rm peak} acts as a cavity, and has an associated time delay111LIGO data favors the time delay td600Mt_{d}\approx 600\,M Holdom:2019bdv , which sets a preferred value for r0149.2r^{*}_{0}\approx-149.2.

td2(rpeakr0)t_{d}\equiv 2(r^{*}_{\rm peak}-r^{*}_{0}) (9)

expressed in tortoise coordinates. An ECO boundary closer to the would-be horizon (i.e., smaller r0r^{*}_{0}) maps to a larger time delay tdt_{d}. The tortoise coordinate can be chosen such that rpeak0r^{*}_{\rm peak}\approx 0.

II Formalism

In this section, we review the formalism for the calculation of the detector’s response rate in the black hole context Hodgkinson:2014iua , which we will then extend to the ECO scenario.

II.1 Black hole

For a black hole, the gravitational potential vanishes at r±r^{*}\to\pm\infty such that Eq. (7) has two simple asymptotic solutions: ρωlup(r)=eiωr\rho_{\omega l}^{\textrm{up}}(r^{*}\to\infty)=e^{i\omega r^{*}}, ρωlin(r)=eiωr\rho_{\omega l}^{\textrm{in}}(r^{*}\to-\infty)=e^{-i\omega r^{*}}. The normalized field modes

Φωin\displaystyle\operatorname{\Phi^{\text{in}}_{\omega\ell}} =Tωlinρωlinr,Φωup\displaystyle=T_{\omega l}^{\textrm{in}}\frac{\rho_{\omega l}^{\textrm{in}}}{r},\,\quad\operatorname{\Phi^{\text{up}}_{\omega\ell}} =Tωlupρωlupr,\displaystyle=T_{\omega l}^{\textrm{up}}\frac{\rho_{\omega l}^{\textrm{up}}}{r}\,, (10)

satisfy the respective boundary conditions

Φωlinr\displaystyle\Phi_{\omega l}^{\textrm{in}}r \displaystyle\to {Tωlineiωr,reiωr+Rωlineiωr,r\displaystyle\left\{\begin{array}[]{cc}\,T_{\omega l}^{\textrm{in}}e^{-i\omega r^{*}},&r^{*}\to-\infty\\ e^{-i\omega r^{*}}+R_{\omega l}^{\textrm{in}}e^{i\omega r^{*}},&r^{*}\to\infty\end{array}\right. (13)
Φωlupr\displaystyle\Phi_{\omega l}^{\textrm{up}}r \displaystyle\to {Rωlupeiωr+eiωr,rTωlupeiωr,r\displaystyle\left\{\begin{array}[]{ll}R_{\omega l}^{\textrm{up}}\,e^{-i\omega r^{*}}+e^{i\omega r^{*}},&r^{*}\to-\infty\\ T_{\omega l}^{\textrm{up}}\,e^{i\omega r^{*}},&r^{*}\to\infty\end{array}\right. (16)

such that Φωlin\Phi_{\omega l}^{\textrm{in}} represents a wave mode incident on the potential barrier from spatial infinity, giving rise to a reflected wave of amplitude RωlinR_{\omega l}^{\rm in} and to a transmitted wave of amplitude TωlinT_{\omega l}^{\textrm{in}}. Conversely, Φωlup\Phi_{\omega l}^{\textrm{up}} represents a wave incident from the past event horizon and is partially reflected by a factor of RωlupR_{\omega l}^{\rm up} and partially transmitted with an amplitude TωlupT_{\omega l}^{\textrm{up}}.

From the Wronskian relations, one can derive

Tωlin\displaystyle T_{\omega l}^{\textrm{in}} =2iωW[ρωin,ρωup],Rωin=W[ρωin,ρωup]W[ρωin,ρωup],\displaystyle=\frac{2i\omega}{W[\operatorname{\mathsf{\rho^{\text{in}}_{\omega\ell}}},\operatorname{\rho^{\text{up}}_{\omega\ell}}]},\,\,\,R^{\text{in}}_{\omega\ell}=-\frac{W[\operatorname{\mathsf{\rho^{\text{in}}_{\omega\ell}}},\rho^{\text{up}\,*}_{\omega\ell}]}{W[\operatorname{\mathsf{\rho^{\text{in}}_{\omega\ell}}},\operatorname{\rho^{\text{up}}_{\omega\ell}}]}, (17)

and222Note that some literature Mark:2017dnq ; Conklin:2017lwb ; Cardoso:2019rvt denote TωupT^{\text{up}}_{\omega\ell} as TBHT_{\rm BH}, and RωupR^{\text{up}}_{\omega\ell} as RBHR_{\rm BH}.

Tωup=Tωlin,,Rωup=TωinTωinRωinT^{\text{up}}_{\omega\ell}=T_{\omega l}^{\textrm{in}},\quad,\,\,R^{\text{up}}_{\omega\ell}=-\frac{T^{\text{in}}_{\omega\ell}}{T^{\text{in}*}_{\omega\ell}}R^{\text{in}*}_{\omega\ell} (18)

where WW denotes the Wronskian W[a,b]=abbaW[a,b]=ab^{\prime}-ba^{\prime}. Energy conservation furthermore implies |Tωup|2+|Rωup|2=1|T^{\text{up}}_{\omega\ell}|^{2}+|R^{\text{up}}_{\omega\ell}|^{2}=1, and |Tωin|2+|Rωin|2=1|T^{\text{in}}_{\omega\ell}|^{2}+|R^{\text{in}}_{\omega\ell}|^{2}=1 .

The Boulware state has positive-frequency modes with respect to the physical Schwarzschild time tt. It reduces to the Minkowski vacuum at spatial infinity but induces a diverging stress-energy near the black hole horizon. The quantum scalar field in this state can be expanded as

ψ==0m=+0dω(aωmupuωmup+aωminuωmin)+h.c.,\displaystyle\psi=\sum^{\infty}_{\ell=0}\sum^{+\ell}_{m=-\ell}\int^{\infty}_{0}\,\mathrm{d}\omega\,\Bigl{(}a^{\text{up}}_{\omega\ell m}u^{\text{up}}_{\omega\ell m}+a^{\text{in}}_{\omega\ell m}u^{\text{in}}_{\omega\ell m}\Bigr{)}+\text{h.c.}\,, (19)

where aωm(in, up)a^{\text{(in, up)}}_{\omega\ell m} are the respective annihilation operators, defining the Boulware state aωm(in, up)|0B=0a^{\text{(in, up)}}_{\omega\ell m}|0_{B}\rangle=0, and

uωm(in, up)(𝗑)\displaystyle u^{\text{(in, up)}}_{\omega\ell m}(\mathsf{x}) =14πωΦω(in, up)(r)Ym(θ,ϕ)eiωt\displaystyle=\frac{1}{\sqrt{4\pi\omega}}\Phi^{\text{(in, up)}}_{\omega\ell}(r)Y_{\ell m}(\theta,\phi)\operatorname{e}^{-i\omega t} (20)

are the respective field bases. Φωin\operatorname{\Phi^{\text{in}}_{\omega\ell}} and Φωup\operatorname{\Phi^{\text{up}}_{\omega\ell}} are the normalized field modes obtained from Eq. (10) with the normalization factor from Eq. (17) and Eq. (18).

Therefore, for a static detector at a coordinate distance RR away from the black hole, the Wightman function is333For a static detector, r=r=Rr=r^{\prime}=R, tt=Δt=Δτ/12M/Rt-t^{\prime}=\Delta t=\Delta\tau/\sqrt{1-2M/R}, and we can take θ=θ=ϕ=ϕ=0\theta=\theta^{\prime}=\phi=\phi^{\prime}=0 without loss of generality.

W(𝗑,𝗑)\displaystyle W(\mathsf{x},\mathsf{x}^{\prime}) =0B|ψ(𝗑)ψ(𝗑)|0B\displaystyle=\langle 0_{B}|\psi(\mathsf{x})\psi(\mathsf{x}^{\prime})|0_{B}\rangle
==00dω(2+1)16π2ωeiωΔτ/12M/R(|Φωup(R)|2+|Φωin(R)|2),\displaystyle=\sum^{\infty}_{\ell=0}\int^{\infty}_{0}\,\mathrm{d}\omega\,\frac{\left(2\ell+1\right)}{16\pi^{2}\omega}\operatorname{e}^{-i\omega\Delta\tau/\sqrt{1-2M/R}}\left(|\operatorname{\Phi^{\text{up}}_{\omega\ell}}(R)|^{2}+|\operatorname{\Phi^{\text{in}}_{\omega\ell}}(R)|^{2}\right)\,, (21)

with the field in the Boulware state. Substituting (21) into the detector response rate (4) and commuting the ss- and ω\omega-integrals, we obtain

˙Boulware(Ω)=Θ(Ω)8π|Ω|l=0(2+1)(|Φω~up(R)|2+|Φω~in(R)|2),\displaystyle\mathcal{\dot{F}}_{\rm Boulware}\left(\Omega\right)=\frac{\Theta(-\Omega)}{8\pi|\Omega|}\sum^{\infty}_{l=0}\left(2\ell+1\right)\left(|\Phi^{\text{up}}_{\tilde{\omega}\ell}(R)|^{2}+|\Phi^{\text{in}}_{\tilde{\omega}\ell}(R)|^{2}\right)\,, (22)

where ω~:=Ω12M/R\tilde{\omega}:=\Omega\sqrt{1-2M/R}. Note that for the Boulware state, the response rate vanishes for positive energies, meaning that the detector can only de-excite, or alternatively Ω\Omega can only take negative values. In the following study we choose the variable Ω/Tloc=8πMω~\Omega/T_{\rm loc}=8\pi M\tilde{\omega} as the effective gap parameter, where TlocT_{\text{loc}} is the local Hawking temperature, Tloc=1/(8πM12M/R)T_{\text{loc}}=1/(8\pi M\sqrt{1-2M/R}).

The other types of quantum vacuum fields for Schwarzchild spacetime, namely Hartle-Hawking state and Unruh state, have regular stress-energy across the future horizon (with Hartle-Hawking regular across the past horizon as well). As we explicitly show in Appendix A, the de-excitation rate of a static UDW detector with the field in the Hartle-Hawking (HH) and Unruh vacuum states overlap with that of the Boulware vacuum for energy gap Ω/Tloc5\Omega/T_{\rm loc}\lesssim-5, and only deviates to a small finite value at zero gap limit (˙HH|Ω/Tloc00.01\mathcal{\dot{F}}_{\rm HH}|_{\Omega/T_{\rm loc}\to 0}\lesssim 0.01), as shown in Figure 8b. Thus, we focus on the Boulware vacuum case in our following discussion.

II.2 ECO

An ECO has no event horizon and thus the vacuum must be of the Boulware type. Since only the in-modes satisfy the boundary condition at r0r^{*}_{0}, we must drop the up-mode contribution from the discussion above, yielding

ψ==0m=+0dω(aωminuωmin)+h.c.,\displaystyle\psi=\sum^{\infty}_{\ell=0}\sum^{+\ell}_{m=-\ell}\int^{\infty}_{0}\,\mathrm{d}\omega\,\Bigl{(}a^{\text{in}}_{\omega\ell m}u^{\text{in}}_{\omega\ell m}\Bigr{)}+\text{h.c.}\,, (23)

instead of Eq. (19), where

uωmin(𝗑)\displaystyle u^{\text{in}}_{\omega\ell m}(\mathsf{x}) =14πωΦωin(r)Ym(θ,ϕ)eiωt,\displaystyle=\frac{1}{\sqrt{4\pi\omega}}\Phi^{\text{in}}_{\omega\ell}(r)Y_{\ell m}(\theta,\phi)\operatorname{e}^{-i\omega t}\,, (24)

with

Φωin=Tωlinρωlinr.\operatorname{\Phi^{\text{in}}_{\omega\ell}}=T_{\omega l}^{\textrm{in}}\frac{\rho_{\omega l}^{\textrm{in}}}{r}. (25)

ρωlin\rho_{\omega l}^{\textrm{in}} is the in mode satisfying Eq. (7) with the boundary conditions parametrized by the reflectivity RWR^{W} of the ECO Conklin:2017lwb

Φωlinr\displaystyle\Phi_{\omega l}^{\textrm{in}}r \displaystyle\to {eiωr0Tωlin(eiω(rr0)+RWeiω(rr0)),rr0eiωr+Rωlineiωrr\displaystyle\left\{\begin{array}[]{cc}e^{-i\omega r^{*}_{0}}\,T_{\omega l}^{\textrm{in}}\left(e^{-i\omega(r^{*}-r^{*}_{0})}+R^{\rm W}\,e^{i\omega(r^{*}-r^{*}_{0})}\right),&r^{*}\to r^{*}_{0}\\ e^{-i\omega r^{*}}+R_{\omega l}^{\textrm{in}}e^{i\omega r^{*}}&r^{*}\to\infty\end{array}\right. (28)

in contrast to (13), where RW=1R^{\rm W}=1 (RW=1R^{\rm W}=-1) corresponds to the Neumann (Dirichlet) boundary condition, whereas RW=0R^{W}=0 means the ECO is perfectly absorbing. From the solution ρωin\operatorname{\mathsf{\rho^{\text{in}}_{\omega\ell}}} to Eq. (7) satisfying the appropriate boundary condition at r0r^{*}_{0}, TωinT^{\text{in}}_{\omega\ell} and RωinR^{\text{in}}_{\omega\ell} are obtained by insisting on the correct behaviour at large rr^{*}. We obtain

Tωin=2ωeiωrωρωin+iρωin|r,Rωin=e2iωr(ωρωiniρωin)ωρωin+iρωin|r,T^{\text{in}}_{\omega\ell}=\left.\frac{2\omega e^{-i\omega r^{*}}}{\omega\operatorname{\mathsf{\rho^{\text{in}}_{\omega\ell}}}+i\operatorname{\mathsf{\rho^{\text{in}}_{\omega\ell}}}^{\prime}}\right|_{r^{*}\to\infty},\quad R^{\text{in}}_{\omega\ell}=\left.\frac{e^{-2i\omega r^{*}}(\omega\operatorname{\mathsf{\rho^{\text{in}}_{\omega\ell}}}-i\operatorname{\mathsf{\rho^{\text{in}}_{\omega\ell}}}^{\prime})}{\omega\operatorname{\mathsf{\rho^{\text{in}}_{\omega\ell}}}+i\operatorname{\mathsf{\rho^{\text{in}}_{\omega\ell}}}^{\prime}}\right|_{r^{*}\to\infty}, (29)

where in practice ρωin\operatorname{\mathsf{\rho^{\text{in}}_{\omega\ell}}} and its derivatives ρωin\operatorname{\mathsf{\rho^{\text{in}}_{\omega\ell}}}^{\prime} are evaluated at a sufficiently large rr^{*}. As shown in Appendix C, we can introduce another set of ECO solutions similar to (16) as an analog of the up-mode ρup\rho_{\rm up} in the black hole setting, but only for mathematical use since only the in-modes satisfy the boundary condition at r0r^{*}_{0}. In this way, it is straightforward to show that the transmission and reflection coefficients of the ECO can be cast into the same form as Eq. (17). Energy conservation implies

|Tωin|2(1(RW)2)=1|Rωin|2,|T^{\text{in}}_{\omega\ell}|^{2}(1-(R^{W})^{2})=1-|R^{\text{in}}_{\omega\ell}|^{2}, (30)

so that |Rωin|=1|R^{\text{in}}_{\omega\ell}|=1 for RW=±1R^{W}=\pm 1.

Then, with the normalized Φωin\operatorname{\Phi^{\text{in}}_{\omega\ell}} field, we find

WECO(𝗑,𝗑)==00dω(2+1)16π2ωeiωΔτ/12M/R|Φωin(R)|2,W_{\rm ECO}(\mathsf{x},\mathsf{x}^{\prime})=\sum^{\infty}_{\ell=0}\int^{\infty}_{0}\,\mathrm{d}\omega\,\frac{\left(2\ell+1\right)}{16\pi^{2}\omega}\operatorname{e}^{-i\omega\Delta\tau/\sqrt{1-2M/R}}|\operatorname{\Phi^{\text{in}}_{\omega\ell}}(R)|^{2}\,, (31)

for the Wightman function, and in turn

˙ECO(Ω)=Θ(Ω)8π|Ω|l=0(2+1)|Φω~in(R)|2,\displaystyle\mathcal{\dot{F}_{\rm ECO}}\left(\Omega\right)=\frac{\Theta(-\Omega)}{8\pi|\Omega|}\sum^{\infty}_{l=0}\left(2\ell+1\right)|\Phi^{\text{in}}_{\tilde{\omega}\ell}(R)|^{2}\,, (32)

for the response rate.

III Results

The most distinctive feature in the comparison of an ECO to a black hole is the resonance spectrum of TωinT^{\text{in}}_{\omega\ell} (i.e., the transfer function). In Figure 1, we show the TωinT^{\text{in}}_{\omega\ell} spectrum of two benchmark models with small (td=66.5M)(t_{d}=66.5\,M) and large (td=160M)(t_{d}=160M) time delays.

Refer to caption
Refer to caption
Figure 1: log10|Tωin|\log_{10}|T^{\text{in}}_{\omega\ell}| for (a) small (td=66.5Mt_{d}=66.5M) and (b) large (td=160Mt_{d}=160M) time delays with RW=(0,1,0.5,1)R^{W}=(0,-1,0.5,1) for l=05l=0\sim 5.

The position and width of these resonances are determined from the real and imaginary parts of the complex poles of TωinT^{\text{in}}_{\omega\ell}, which are affected by both the ECO boundary condition and the gravitational potential. The spacings of the resonance spikes is related to the time delay such that444Note that the actual Δω\Delta\omega has a small frequency dependence resulting from the finite width of the gravitational potential around rpeakr^{*}_{\rm peak}.

Δω=2πtd,ΔΩ/Tloc=8πMΔω~=16π2Mtd.\Delta{\omega}=\frac{2\pi}{t_{d}},\quad\Delta{\Omega/T_{\rm loc}}=8\pi M\Delta{\tilde{\omega}}=\frac{16\pi^{2}M}{t_{d}}. (33)

Therefore, the resonance spacing is smaller for a larger time delay (i.e., when r0r_{0} is closer to the would-be horizon). For a zero RWR^{W}, TωinT^{\text{in}}_{\omega\ell} has no resonances at all since the boundary is perfectly absorbing, and thus is equivalent to the TωinT^{\text{in}}_{\omega\ell} of a black hole. Differing time delays in this case do not affect the result since different locations of a perfect-absorbing boundary are equivalent for the wave modes. As we will demonstrate, these resonance structures will also show up in the UDW detector’s response rate, yielding distinct structures in contrast to the black hole.

Using (32), the response rate can be calculated with the ll sum truncated to lmaxl_{\rm max} such that the contribution with l=lmax+1l=l_{\rm max}+1 is negligible. It turns out that lmaxl_{\rm max} increases for a larger radial distance RR and a larger |Ω/Tloc||\Omega/T_{\rm loc}|, and is insensitive to the boundary condition and time delay changes. For example, when |Ω/Tloc||\Omega/T_{\rm loc}| goes to 20, lmax5l_{\rm max}\approx 5 for R=4MR=4M, and lmax15l_{\rm max}\approx 15 for R=15MR=15M. We give an analytic approximation of lmaxl_{\rm max} in Appendix B. To illustrate some general features, we present in Figures 2 the response rate for the previous two benchmark examples with different (tdt_{d},RR).

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 2: M˙M\dot{\mathcal{F}} as a function of Ω/Tloc\Omega/T_{\text{loc}} for a static detector at R=2.5M,4M,15MR=2.5M,4M,15M (top to bottom) above an ECO of (left) small (td=66.5Mt_{d}=66.5\,M), and (right) large (td=160Mt_{d}=160M) time delays with RW=(0,1,0.5,1)R^{W}=(0,-1,0.5,1). The orange line is for the black hole case with Boulware vacuum.

Note that the first row of Figure 2 is for the case where the static detector is inside the cavity (R=2.5MR=2.5M), in contrast to the second and third rows where the detector is placed outside (R=4MR=4M and R=15MR=15M).

We easily see the striking resonance patterns of the response rate in the first row of Figure 2, which match those of |Tωin||T^{\text{in}}_{\omega\ell}| shown in Figure 1 that are independent of RR; cases with the smaller time delay have resonances more sparse in distribution, and vice versa. Comparing to those of the second and third rows, we see as the distance gets larger, the resonances become more suppressed in size. We give a detailed analysis in the next section.

In the second and thirds rows of Figure 2, we see a general trend for the detector’s response rate to grow roughly linearly with |Ω/Tloc||\Omega/T_{\rm loc}|. But in the second row, we see that an ECO with RW=0R^{W}=0 has a response rate that is relatively lower than that of the black hole, even though the boundaries of both objects are perfectly absorbing. Comparing Eq. (18) and Eq. (26), we see that this is because the ECO case has no up-mode contribution for any RWR^{W}. By comparison to the second row, the third row then shows that the extra up-mode contribution in the black hole response becomes relatively smaller for increasing RR. In the large RR limit all the response rates approach the Minkowski vacuum state result, which is proportional to Θ(Ω)Ω-\Theta(-\Omega)\Omega.

IV Analysis

From Eq. (28), we see the behaviour of Φωin\operatorname{\Phi^{\text{in}}_{\omega\ell}} and the resulting ˙ECO\mathcal{\dot{F}_{\rm ECO}} is dominated by TωinT^{\text{in}}_{\omega\ell} at small rr and by RωinR^{\text{in}}_{\omega\ell} at large rr up to some complex phase factors. To see this more explicitly, we can take absolute value of Φωlin\Phi_{\omega l}^{\textrm{in}}, obtaining

|Φωlin|r\displaystyle|\Phi_{\omega l}^{\textrm{in}}|r |Tωlin|1+RW2+2RWcos[2ω(rr0)], as rr0\displaystyle\to|T_{\omega l}^{\textrm{in}}|\sqrt{1+R^{W^{2}}+2R^{W}\cos[2\omega(r^{*}-r^{*}_{0})]},\text{ as }r^{*}\to r^{*}_{0} (34)
|Φωlin|r\displaystyle|\Phi_{\omega l}^{\textrm{in}}|r 1+|Rωin|2+2|Rωin|cos(2ωr+θ), as r.\displaystyle\to\sqrt{1+|R^{\text{in}}_{\omega\ell}|^{2}+2|R^{\text{in}}_{\omega\ell}|\cos(2\omega r^{*}+\theta)},\quad\,\,\,\,\text{ as }r^{*}\to\infty. (35)

from Eq. (28), where θ=arccos((Rωin)/|Rωin|)\theta=\arccos(\Re(R^{\text{in}}_{\omega\ell})/|R^{\text{in}}_{\omega\ell}|). The value of θ\theta can also have large fluctuations at the location of the TωinT^{\text{in}}_{\omega\ell} resonances, as the benchmark model in Figure 3a shows. We see that |Φωin|r|\operatorname{\Phi^{\text{in}}_{\omega\ell}}|r goes to |Tωlin|(1+RW)|T_{\omega l}^{\textrm{in}}|(1+R^{W}) when approaching the ECO boundary, and at large rr to some value between (1|Rωin|)(1-|R^{\text{in}}_{\omega\ell}|) and (1+|Rωin|)(1+|R^{\text{in}}_{\omega\ell}|), as the example in Figure 3b illustrates. From Eq. (30), we see that |Rωin|1|R^{\text{in}}_{\omega\ell}|\leq 1 for RW1R^{W}\leq 1, but |Tωin||T^{\text{in}}_{\omega\ell}| can display relatively large resonance peaks. This, in conjunction with (34) and (35), partially explains why |Φωin|r|\operatorname{\Phi^{\text{in}}_{\omega\ell}}|r and the corresponding contribution in ˙\mathcal{\dot{F}} at given ll exhibit smaller resonances at a larger distance.

Refer to caption
Refer to caption
Figure 3: (a) θ=arccos((Rωin)/|Rωin|)\theta=\arccos(\Re(R^{\text{in}}_{\omega\ell})/|R^{\text{in}}_{\omega\ell}|) and (b) |Φωin|R|\operatorname{\Phi^{\text{in}}_{\omega\ell}}|R (R=15MR=15M) for a time delay td=160Mt_{d}=160M with RW=1R^{W}=-1 for l=0l=0 (black) and l=1l=1 (red). Solid lines are the exact results from Eq. (25), while dashed lines denote the approximation from Eq. (35).

Note that Eqs. (34) and (35) give a good match to the exact result (25) for |rrpeak||r^{*}-r^{*}_{\rm peak}| sufficiently large, that is provided ω2\omega^{2} is larger than V(r)V(r), thus mapping to the parameter space of large ω\omega and small ll. These features can also be seen in the example shown in Fig (3)b, where the exact result of |Φωin|r|\operatorname{\Phi^{\text{in}}_{\omega\ell}}|r at l=0l=0 matches the approximate form well, with resonances fluctuating between 0 and +2, while all larger l1l\geq 1 show notable deviation at small |Ω/Tloc||\Omega/T_{\rm loc}| (=8πM|ω|)(=8\pi M\,|\omega|).

As Figure 1 shows, the resonance structures of TωinT^{\text{in}}_{\omega\ell} (and thus the similar structure of RωinR^{\text{in}}_{\omega\ell}) shift to a larger |Ω/Tloc||\Omega/T_{\rm loc}| for a larger ll. As rr increases, lmaxl_{\rm max} also becomes large (Appendix B), so that the ll-summed results have more dependence on large ll yielding a milder resonance structure at a given frequency. This also illustrates why the ˙\mathcal{\dot{F}} results in Figures 2 show milder resonances at a larger distance for a given Ω/Tloc\Omega/T_{\rm loc}. This further implies that we can identify a clear resonance structure at very large RR provided the time delay is large so that the resonances are located at small frequencies where lmaxl_{\rm max} is small. For illustration, we give a benchmark example of a very large time delay (td=600M)(t_{d}=600M) with the detector at a very large distance R=150MR=150M in Figure 4. We can see that the resonances of |Tωin||T^{\text{in}}_{\omega\ell}| at small ll in Figure 4a are clearly reflected in the response rate result shown in Figure 4b, even at this large RR.

Refer to caption
Refer to caption
Figure 4: (a) log10|Tωin|\log_{10}|T^{\text{in}}_{\omega\ell}| and (b) M˙M\dot{\mathcal{F}} for a large time delay td=600Mt_{d}=600M with RW=1R^{W}=-1. The M˙M\dot{\mathcal{F}} result has the static detector at a large distance R=150MR=150M with ll summed from zero to different lmaxl_{\rm max}.

V Case RW=0.98R^{W}=-0.98

For physical realizations of the ECO model, |RW||R^{W}| can be taken to be slightly less than one to account for a small amount of damping. Damping may be needed in the case of spinning ECOs to avoid an instability. In the case of a static 2-2-hole the damping may be explicitly calculated, and the resulting frequency dependent damping is largest for small frequencies Holdom:2020onl . For our static ECO study we shall illustrate the effect of damping by choosing a constant value RW=0.98R^{W}=-0.98. We show in Figure 5a the TωinT^{\text{in}}_{\omega\ell} function with large time delay td=160Mt_{d}=160M, with the related response rate at small distance (R=4MR=4M) depicted in Figure 5b. Comparing to the RW=1R^{W}=-1 case shown in Figure 1b and the right figure in the second row of Figure 2, we observe that the response rate for RW=0.98R^{W}=-0.98 is close to the RW=1R^{W}=-1 case, except that the resonance heights of TωinT^{\text{in}}_{\omega\ell} and ˙ECO\mathcal{\dot{F}_{\rm ECO}} are both noticeably milder than those for RW=1R^{W}=-1. It turns out that this feature is also present in all other examples we have examined, including the large-RR cases.

Refer to caption
Refer to caption
Figure 5: (a) log10|Tωin|\log_{10}|T^{\text{in}}_{\omega\ell}| and (b) M˙M\dot{\mathcal{F}} ( R=4MR=4M) as a function of Ω/Tloc\Omega/T_{\text{loc}} for a time delay td=160Mt_{d}=160M with RW=0.98R^{W}=-0.98.

For the RW=0.98R^{W}=-0.98 examples at large RR, the exact result of |Φωlin|r|\Phi_{\omega l}^{\textrm{in}}|r can be approximated by Eq. (35). From (30), when |RW||R^{W}| is close but not equal to unity, the resonances of TωinT^{\text{in}}_{\omega\ell} can show up as dips in |Rωin||R^{\text{in}}_{\omega\ell}|, as shown in Figure 6 for a benchmark example with a time delay td=160Mt_{d}=160M. These dips can potentially translate to large dips in M˙M\dot{\mathcal{F}} at large RR, considering Eq. (35).

Refer to caption
Refer to caption
Figure 6: (a) |Tωin||T^{\text{in}}_{\omega\ell}| and (b)|Rωin||R^{\text{in}}_{\omega\ell}| for a time delay td=160Mt_{d}=160M with RW=0.98R^{W}=-0.98.

However, our numerical results indicate that the resonance heights of ˙ECO\mathcal{\dot{F}_{\rm ECO}} for RW=0.98R^{W}=-0.98 are noticeably milder than those for RW=1R^{W}=-1 even at large RR. This relates to the fact that, despite the large dips in |Rωin||R^{\text{in}}_{\omega\ell}| for RW=0.98R^{W}=-0.98, it has either the same or smaller fluctuations in θ\theta than for RW=1R^{W}=-1, as a comparison of Figure 7a with Figure 3a shows.

Refer to caption
Refer to caption
Figure 7: (a) θ=arccos((Rωin)/|Rωin|)\theta=\arccos(\Re(R^{\text{in}}_{\omega\ell})/|R^{\text{in}}_{\omega\ell}|) and (b) |Φωin|R|\operatorname{\Phi^{\text{in}}_{\omega\ell}}|R (R=15MR=15M) for a time delay td=160Mt_{d}=160M with RW=0.98R^{W}=-0.98 at l=0l=0 (black) and l=1l=1 (red).

VI Summary

We have calculated the response rate of an Unruh-DeWitt detector placed away from a general exotic compact object (ECO) having an exterior spacetime identical to that of a Schwarzschild black hole. Due to the absence of an event horizon and the corresponding up-modes in the vacuum fields, a perfectly absorbing ECO (RW=0R^{W}=0) spacetime has a smaller detector response rate than that of the same-mass black hole, with a smaller ratio for a larger detector gap and a smaller distance. The cavity between an ECO’s reflective boundary (RW0R^{W}\neq 0) and the gravitational potential peak determines the time delay tdt_{d} and results in distinct resonance structures, connected to the poles of the transfer function TωinT^{\text{in}}_{\omega\ell}. Thus, the observed spacing of the resonance structures tells us about the fundamental length scale at which the ECO deviates from a black hole.

Finally, we presented a detailed analysis on the large-distance behaviour of the resonance structures and the damping effect in the response rate, and showed how a large cavity size can increase the chances of observing resonance structure at large distances.

Our study provides a new alternative way to differentiate a static black hole from an ECO with a static UDW detector. The obvious next generalization is to study what happens to a spinning ECO. Likewise, it will be interesting to see how the results change if we consider detectors in motion, such as in geodesic orbits. We leave these investigations for future work.

\appendixpage\addappheadtotoc

Appendix A Hartle-Hawking and Unruh vacuums

The Hartle-Hawking vacuum state has positive-frequency modes with respect to the horizon generators in Kruskal coordinates. It is regular across the horizon, and features outgoing and ingoing thermal fluxes from the past event horizon and the past null infinity, respectively. Accordingly, the response rate of a static detector above the black hole with fields in Hartle-Hawking (HH) state are Hodgkinson:2014iua :

˙HH(Ω)=18πΩ1eΩ/Tloc1l=0(2+1)(|Φω~up(R)|2+|Φω~in(R)|2),\displaystyle\mathcal{\dot{F}}_{\rm HH}\left(\Omega\right)=\frac{1}{8\pi\Omega}\frac{1}{\operatorname{e}^{\Omega/T_{\text{loc}}}-1}\sum^{\infty}_{l=0}\left(2\ell+1\right)\left(|\Phi^{\text{up}}_{\tilde{\omega}\ell}(R)|^{2}+|\Phi^{\text{in}}_{\tilde{\omega}\ell}(R)|^{2}\right)\,, (36)

Comparing to the Boulware vacuum case Eq. (22), we have

˙HH˙Boulware=1Θ(Ω)(1eΩ/Tloc),\frac{\mathcal{\dot{F}}_{\rm HH}}{\mathcal{\dot{F}}_{\rm Boulware}}=\frac{1}{\Theta(-\Omega)(1-\operatorname{e}^{\Omega/T_{\text{loc}}})}, (37)

which we plot in Figure 8a. We see that the de-excitation rate of a UDW detector in the Hartle-Hawking vacuum becomes coincident with that of the Boulware vacuum for a large energy gap, except for an enhancement at the small energy gap region, resulting a small finite response rate at zero frequency limit (˙HH|Ω/Tloc00.01\mathcal{\dot{F}}_{HH}|_{\Omega/T_{\rm loc}\to 0}\sim 0.01), as shown in Figure 8b for two benchmark examples.

Refer to caption
Refer to caption
Figure 8: (a)The ratio ˙HH/˙Boulware\mathcal{\dot{F}}_{HH}/\mathcal{\dot{F}}_{\rm Boulware} and (b) M˙BoulwareM\mathcal{\dot{F}}_{\rm Boulware} (solid orange) vs M˙HHM\mathcal{\dot{F}}_{\rm HH} (dashed blue) and M˙UnruhM\mathcal{\dot{F}}_{\rm Unruh} (dashed black) for a static detector at R=4MR=4M (top) and R=15MR=15M (bottom) outside a black hole.

The Unruh vacuum has positive-frequency up-modes with respect to the horizon generators in Kruskal coordinates, and positive-frequency in-modes with respect to the physical Schwarzschild time tt. The response rate for a static detector with fields in Unruh vacuum state thus is Hodgkinson:2014iua :

˙Unruh(Ω)=18πΩl=0(2+1)(1eΩ/Tloc1|Φω~up(R)|2Θ(Ω)|Φω~in(R)|2),\displaystyle\mathcal{\dot{F}}_{\rm Unruh}\left(\Omega\right)=\frac{1}{8\pi\Omega}\sum^{\infty}_{l=0}\left(2\ell+1\right)\left(\frac{1}{\operatorname{e}^{\Omega/T_{\text{loc}}}-1}|\Phi^{\text{up}}_{\tilde{\omega}\ell}(R)|^{2}-\Theta(-\Omega)|\Phi^{\text{in}}_{\tilde{\omega}\ell}(R)|^{2}\right)\,, (38)

Comparing Eq. (38) to Eq. (36) and Eq. (22), we see that the response rate for Unruh vacuum interpolates between those of Hartle-Hawking and Boulware cases. One can also see this in Figure 8b for two benchmark examples.

Appendix B Determination of lmaxl_{\rm max}

In section III, we calculated the response rate using Eq. (32), where the ll sum is taken from zero to lmaxl_{\rm max} such that the contribution with l=lmax+1l=l_{\rm max}+1 is negligible.

We find that the lmaxl_{\rm max} can be well approximated analytically from the relation

ω2V(lmax,R)0\omega^{2}-V(l_{\rm max},R)\approx 0 (39)

since the wave amplitude is heavily suppressed in the region of large (V(R)ω2)(V(R)-\omega^{2}). This gives

lmax12(1+18MR+4ω2R3R2M)l_{\rm max}\approx\left\lceil\frac{1}{2}\left(-1+\sqrt{1-\frac{8M}{R}+\frac{4\omega^{2}R^{3}}{R-2M}}\right)\right\rceil (40)

for given ω\omega and RR, where the outer bracket \lceil\cdots\rceil denotes the ceiling function. Then utilizing Ω/Tloc=8πMω~\Omega/T_{\rm loc}=8\pi M\tilde{\omega} we obtain lmax(Ω/Tloc,R)l_{\rm max}(\Omega/T_{\rm loc},R) accordingly, as shown in Figure 9. We observe the general feature that we only need to sum small ll values for small frequency or small RR region.

Refer to caption
Refer to caption
Figure 9: lmaxl_{\rm max} for a static detector at distance RR with (a) Ω/Tloc=0.5\Omega/T_{\rm loc}=-0.5 and (b) Ω/Tloc=20\Omega/T_{\rm loc}=-20.

Appendix C Wronskian relations for ECO

Here we derive the expressions of the transmission and reflection coefficients from the Wronskian relations. We can construct two sets of solutions, satisfying

Φωlinr\displaystyle\Phi_{\omega l}^{\textrm{in}}r \displaystyle\to {eiωr0Tωlin(eiω(rr0)+RWeiω(rr0)),rr0,eiωr+Rωlineiωr,r,\displaystyle\left\{\begin{array}[]{cc}e^{-i\omega r^{*}_{0}}\,T_{\omega l}^{\textrm{in}}\left(e^{-i\omega(r^{*}-r^{*}_{0})}+R^{\rm W}\,e^{i\omega(r^{*}-r^{*}_{0})}\right),&r^{*}\to r^{*}_{0}\,,\\ e^{-i\omega r^{*}}+R_{\omega l}^{\textrm{in}}e^{i\omega r^{*}},&r^{*}\to\infty\,,\end{array}\right. (43)
Φωlupr\displaystyle\Phi_{\omega l}^{\textrm{up}}r \displaystyle\to {Rωlupeiωr+eiωr,rr0,Tωlupeiωr,r,\displaystyle\left\{\begin{array}[]{ll}R_{\omega l}^{\textrm{up}}\,e^{-i\omega r^{*}}+e^{i\omega r^{*}},&r^{*}\to r^{*}_{0}\,,\\ T_{\omega l}^{\textrm{up}}\,e^{i\omega r^{*}},&r^{*}\to\infty\,,\end{array}\right. (46)

Note that Tωup=TBHT^{\text{up}}_{\omega\ell}=T_{\rm BH} and Rωup=RBHR^{\text{up}}_{\omega\ell}=R_{\rm BH} are defined in Eq. (18) for black hole with no RWR^{W} dependence.

Therefore, the Wronskian satisfies

W[Φωin,Φωup]{2iωTωin(1ei2ωr0RBHRW),rr0,2iωTωup,r.W[\operatorname{\Phi^{\text{in}}_{\omega\ell}},\operatorname{\Phi^{\text{up}}_{\omega\ell}}]\to\begin{cases}&2i\omega T^{\text{in}}_{\omega\ell}(1-e^{-i2\omega r^{*}_{0}}R_{\rm BH}R^{\rm W})\,,\quad\quad r^{*}\to r^{*}_{0}\,,\\ &2i\omega T^{\text{up}}_{\omega\ell}\,,\quad\quad r^{*}\to\infty\,.\end{cases} (47)

Since the Wronskian must be a constant over all rr^{*}, we have

Tωup=Tωin(1ei2ωr0RBHRW),T^{\text{up}}_{\omega\ell}=T^{\text{in}}_{\omega\ell}(1-e^{-i2\omega r^{*}_{0}}R_{\rm BH}R^{\rm W}), (48)

or

Tωin=TBH1ei2ωr0RBHRW.T^{\text{in}}_{\omega\ell}=\frac{T_{\rm BH}}{1-e^{-i2\omega r^{*}_{0}}{R_{\rm BH}}R^{\rm W}}. (49)

The resonances of |Tωin||T^{\text{in}}_{\omega\ell}| therefore result from the zeroes of 1ei2ωr0RBHRW=01-e^{-i2\omega r^{*}_{0}}{R_{\rm BH}}R^{\rm W}=0.

Since W[Φωin,Φωup]=TωinTωupW[ρωin,ρωup],W[\operatorname{\Phi^{\text{in}}_{\omega\ell}},\operatorname{\Phi^{\text{up}}_{\omega\ell}}]=T^{\text{in}}_{\omega\ell}T^{\text{up}}_{\omega\ell}W[\operatorname{\mathsf{\rho^{\text{in}}_{\omega\ell}}},\operatorname{\rho^{\text{up}}_{\omega\ell}}], from Eq. (47) we have

Tωin=2iωW[ρωin,ρωup].T^{\text{in}}_{\omega\ell}=\frac{2i\omega}{W[\operatorname{\mathsf{\rho^{\text{in}}_{\omega\ell}}},\operatorname{\rho^{\text{up}}_{\omega\ell}}]}. (50)

Note that the in-mode transmission coefficient TωinT^{\text{in}}_{\omega\ell} is equivalent to the transfer function calculated in ECO gravitational-wave studies555Note that ρωin2iωψωleft\operatorname{\mathsf{\rho^{\text{in}}_{\omega\ell}}}\leftrightarrow 2i\omega\operatorname{\mathsf{\psi^{\text{left}}_{\omega\ell}}}, ρωupψωright\operatorname{\rho^{\text{up}}_{\omega\ell}}\leftrightarrow\operatorname{\psi^{\text{right}}_{\omega\ell}} for the ψleft,right\psi^{\rm left,right} basis appearing in the gravitational wave literature Conklin:2017lwb . Correspondingly, Tωin=𝒦T^{\text{in}}_{\omega\ell}=\cal{K}, where 𝒦\cal{K} is the conventional transfer function defined as 𝒦=1/W[ψωleft,ψωright]{\cal K}=1/W[\operatorname{\mathsf{\psi^{\text{left}}_{\omega\ell}}},\operatorname{\psi^{\text{right}}_{\omega\ell}}]. Also since Φωin2iωTωinψωleft\operatorname{\Phi^{\text{in}}_{\omega\ell}}\leftrightarrow 2i\omega T^{\text{in}}_{\omega\ell}\operatorname{\mathsf{\psi^{\text{left}}_{\omega\ell}}}, ΦωupTωupψωright\operatorname{\Phi^{\text{up}}_{\omega\ell}}\leftrightarrow T^{\text{up}}_{\omega\ell}\operatorname{\psi^{\text{right}}_{\omega\ell}}, comparing Eq. (43) of this paper and Eq. (6) and Eq. (B1) of Ref. Conklin:2017lwb , the coefficients map as Tωin1/(2iωAin)=Atrans/Ain=TBHT^{\text{in}}_{\omega\ell}\leftrightarrow 1/(2i\omega A_{\rm in})=A_{\rm trans}/A_{\rm in}=T_{\rm BH}, RW=R=e2iωr0Aref/AtransR^{W}=R=e^{2i\omega r^{*}_{0}}\,A_{\rm ref}/A_{\rm trans}, RωinAout/AinR^{\text{in}}_{\omega\ell}\leftrightarrow A_{\rm out}/A_{\rm in}. And Rωup/Tωup=Dtrans, 1/Tωup=DrefR^{\text{up}}_{\omega\ell}/T^{\text{up}}_{\omega\ell}=D_{\rm trans},\,1/T^{\text{up}}_{\omega\ell}=D_{\rm ref}, or equivalently Tωup=1/Dref=TBHT^{\text{up}}_{\omega\ell}=1/D_{\rm ref}=T_{\rm BH}, Rωup=Dtrans/Dref=RBHR^{\text{up}}_{\omega\ell}=D_{\rm trans}/D_{\rm ref}=R_{\rm BH}. Mark:2017dnq ; Conklin:2017lwb . Furthermore, it matches the TωinT^{\text{in}}_{\omega\ell} calculation using Eq. (29) without involving any up mode.

Since

W[Φωin,Φωup]{2iωTωin(RBHe2iωr0RW)rr0,2iωTBHRωin,r,W[\operatorname{\Phi^{\text{in}}_{\omega\ell}},\Phi^{\text{up}*}_{\omega\ell}]\to\begin{cases}&2i\omega T^{\text{in}}_{\omega\ell}(R^{*}_{\rm BH}-e^{-2i\omega r^{*}_{0}}R^{\rm W})\,\quad\quad r^{*}\to r^{*}_{0}\,,\\ &-2i\omega T^{*}_{\rm BH}R^{\text{in}}_{\omega\ell}\,,\quad\quad\quad\quad\quad\quad\quad r^{*}\to\infty\,,\end{cases} (51)

we have

W[Φin,Φup]W[Φin,Φup]|r=TBHTBHW[ρin,ρup]W[ρin,ρup]=TBHTBHRωin,\frac{W[\Phi_{\rm in},\Phi^{*}_{\rm up}]}{W[\Phi_{\rm in},\Phi_{\rm up}]}\Bigg{|}_{r^{*}\to\infty}=\frac{T^{*}_{\rm BH}}{T_{\rm BH}}\frac{W[\rho_{\rm in},\rho^{*}_{\rm up}]}{W[\rho_{\rm in},\rho_{\rm up}]}=-\frac{T^{*}_{\rm BH}}{T_{\rm BH}}R^{\text{in}}_{\omega\ell}, (52)

so that

Rωin=W[ρin,ρup]W[ρin,ρup].R^{\text{in}}_{\omega\ell}=-\frac{W[\rho_{\rm in},\rho^{*}_{\rm up}]}{W[\rho_{\rm in},\rho_{\rm up}]}. (53)

Also, from

W[Φωin,Φωup]W[Φωin,Φωup]|rr0=TωinTωinW[ρωin,ρωup]W[ρωin,ρωup]=e2iωr0Tωin(RBH+RWe2iωr0)Tωin(RBHRW+e2iωr0)\frac{W[\operatorname{\Phi^{\text{in}}_{\omega\ell}},\Phi^{\text{up}*}_{\omega\ell}]^{*}}{W[\operatorname{\Phi^{\text{in}}_{\omega\ell}},\operatorname{\Phi^{\text{up}}_{\omega\ell}}]}\Bigg{|}_{r^{*}\to r^{*}_{0}}=\frac{T^{\text{in}*}_{\omega\ell}}{T^{\text{in}}_{\omega\ell}}\frac{W[\operatorname{\mathsf{\rho^{\text{in}}_{\omega\ell}}},\rho^{\text{up}\,*}_{\omega\ell}]^{*}}{W[\operatorname{\mathsf{\rho^{\text{in}}_{\omega\ell}}},\operatorname{\rho^{\text{up}}_{\omega\ell}}]}=\frac{e^{2i\omega r^{*}_{0}}T^{\text{in}*}_{\omega\ell}\left(-R_{\rm BH}+R^{\rm W}e^{2i\omega r^{*}_{0}}\right)}{T^{\text{in}}_{\omega\ell}\left(-R_{\rm BH}R^{\rm W}+e^{2i\omega r^{*}_{0}}\right)} (54)

we have

W[ρωin,ρωup]W[ρωin,ρωup]=e2iωr0(RBH+RWe2iωr0)(RBHRW+e2iωr0).\frac{W[\operatorname{\mathsf{\rho^{\text{in}}_{\omega\ell}}},\rho^{\text{up}\,*}_{\omega\ell}]^{*}}{W[\operatorname{\mathsf{\rho^{\text{in}}_{\omega\ell}}},\operatorname{\rho^{\text{up}}_{\omega\ell}}]}=\frac{e^{2i\omega r^{*}_{0}}\left(-R_{\rm BH}+R^{\rm W}e^{2i\omega r^{*}_{0}}\right)}{\left(-R_{\rm BH}R^{\rm W}+e^{2i\omega r^{*}_{0}}\right)}. (55)

Then, defining

W/=W[ρωin,ρωup]W[ρωin,ρωup],W^{/}=\frac{W[\operatorname{\mathsf{\rho^{\text{in}}_{\omega\ell}}},\rho^{\text{up}\,*}_{\omega\ell}]^{*}}{W[\operatorname{\mathsf{\rho^{\text{in}}_{\omega\ell}}},\operatorname{\rho^{\text{up}}_{\omega\ell}}]}, (56)

one has

RBH=e2iωr0RWW/1e2iωr0RWW/ for RW±1R_{\rm BH}=\frac{e^{2i\omega r^{*}_{0}}R^{\rm W}-W^{/}}{1-e^{-2i\omega r^{*}_{0}}R^{\rm W}W^{/}}\text{ for }R^{W}\neq\pm 1 (57)

and

W/=±e2iωr0 for RW=±1W^{/}=\pm e^{2i\omega r^{*}_{0}}\text{ for }R^{W}=\pm 1 (58)

from Eq. (55).

Referring to Eq. (50) and Eq. (53), and the definition of W/W^{/} (Eq. (56)), we obtain the relation

Rωin=TωinTωinW/R^{\text{in}}_{\omega\ell}=\frac{T^{\text{in}}_{\omega\ell}}{T^{\text{in}*}_{\omega\ell}}W^{/*} (59)

For the special case RW=±1R^{W}=\pm 1, inserting Eq. (58) into Eq. (59), we have

Rωin=±TωinTωine2iωr0 for RW=±1.R^{\text{in}}_{\omega\ell}=\pm\frac{T^{\text{in}}_{\omega\ell}}{T^{\text{in}*}_{\omega\ell}}e^{-2i\omega r^{*}_{0}}\text{ for }R^{W}=\pm 1. (60)

For other RWR^{W} value, inserting Eq. (57) into Eq. (59), we have

Rωin=TBH(RBHe2iωr0RW)TBH(1e2iωr0RBHRW) for RW±1R^{\text{in}}_{\omega\ell}=-\frac{T_{\rm BH}(R^{*}_{\rm BH}-e^{-2i\omega r^{*}_{0}}R^{W})}{T^{*}_{\rm BH}(1-e^{-2i\omega r^{*}_{0}}R_{\rm BH}R^{\rm W})}\text{ for }R^{W}\neq\pm 1 (61)

or equivalently

Rωin=TBHTBHRBH|RBH|2e2iωr0RBHRW1e2iωr0RBHRW.R^{\text{in}}_{\omega\ell}=-\frac{T_{\rm BH}}{T^{*}_{\rm BH}R_{\rm BH}}\cdot\frac{|R_{\rm BH}|^{2}-e^{-2i\omega r^{*}_{0}}R_{\rm BH}R^{W}}{1-e^{-2i\omega r^{*}_{0}}R_{\rm BH}R^{\rm W}}. (62)

Comparing Eq. (49) and Eq. (62), we see that the resonances of RωinR^{\text{in}}_{\omega\ell} occur at the same set of frequencies as those of TωinT^{\text{in}}_{\omega\ell}. However, when |RBH||R_{\rm BH}| goes to one (i.e., |TBH||T_{\rm BH}| to 0) at large ll for given ω\omega, the second multiplicative factor’s nominator cancel its denominator, making the resonance structures of RωinR^{\text{in}}_{\omega\ell} disappears. This helps to explain the diminishment of resonance structures for large-ll Φωin\operatorname{\Phi^{\text{in}}_{\omega\ell}} at large-RR.

From Eq. (61), we also see

Rωin=TBHRBHTBH for RW=0.R^{\text{in}}_{\omega\ell}=-\frac{T_{\rm BH}R^{*}_{\rm BH}}{T^{*}_{\rm BH}}\text{ for }R^{W}=0. (63)

Moreover, by matching the Wronskian W[Φωin,Φωin]W[\operatorname{\Phi^{\text{in}}_{\omega\ell}},\Phi^{\rm in*}_{\omega\ell}] at r0r^{*}_{0} and infinity we obtain the energy conservation relation |Tωin|2(1RW2)=1|Rωin|2|T^{\text{in}}_{\omega\ell}|^{2}(1-R^{W^{2}})=1-|R^{\text{in}}_{\omega\ell}|^{2}. Similarly, for the Wronskian W[Φωup,Φωup]W[\operatorname{\Phi^{\text{up}}_{\omega\ell}},\Phi^{\rm up*}_{\omega\ell}], we have |TBH|2+|RBH|2=1|T_{\rm BH}|^{2}+|R_{\rm BH}|^{2}=1.

Acknowledgements

This work was supported in part by the Natural Science and Engineering Research Council of Canada.

References

  • (1) P. O. Mazur and E. Mottola, [arXiv:gr-qc/0109035 [gr-qc]].
  • (2) F. E. Schunck and E. W. Mielke, Class. Quant. Grav. 20, R301-R356 (2003) [arXiv:0801.0307 [astro-ph]].
  • (3) B. Holdom and J. Ren, Phys. Rev. D 95, no.8, 084034 (2017) [arXiv:1612.04889 [gr-qc]].
  • (4) V. Baccetti, R. B. Mann, S. Murk and D. R. Terno, Phys. Rev. D 99, no.12, 124014 (2019) [arXiv:1811.04495 [gr-qc]].
  • (5) D. R. Terno, Phys. Rev. D 101, no.12, 124053 (2020) [arXiv:2003.12312 [gr-qc]].
  • (6) J. Abedi, H. Dykaar and N. Afshordi, Phys. Rev. D 96, no.8, 082004 (2017) doi:10.1103/PhysRevD.96.082004 [arXiv:1612.00266 [gr-qc]].
  • (7) Z. Mark, A. Zimmerman, S. M. Du and Y. Chen, Phys. Rev. D 96, no.8, 084002 (2017) doi:10.1103/PhysRevD.96.084002 [arXiv:1706.06155 [gr-qc]].
  • (8) R. S. Conklin, B. Holdom and J. Ren, Phys. Rev. D 98, no.4, 044021 (2018) doi:10.1103/PhysRevD.98.044021 [arXiv:1712.06517 [gr-qc]].
  • (9) V. Cardoso and P. Pani, Living Rev. Rel. 22, no.1, 4 (2019) doi:10.1007/s41114-019-0020-4 [arXiv:1904.05363 [gr-qc]].
  • (10) B. Holdom, Phys. Rev. D 101, no.6, 064063 (2020) doi:10.1103/PhysRevD.101.064063 [arXiv:1909.11801 [gr-qc]].
  • (11) W. G. Unruh, Phys. Rev.  D 14, 870 (1976).
  • (12) B. S. DeWitt, “Quantum gravity: the new synthesis”, in General Relativity; an Einstein centenary survey ed S. W. Hawking and W. Israel (Cambridge University Press, 1979) 680.
  • (13) N. Funai, J. Louko and E. Martín-Martínez, Phys. Rev. D 99, no.6, 065014 (2019) doi:10.1103/PhysRevD.99.065014 [arXiv:1807.08001 [quant-ph]].
  • (14) N. D. Birrell and P. C. W. Davies, Quantum Fields in Curved Space (Cambridge University Press 1982).
  • (15) J. Louko and A. Satz, Class. Quant. Grav. 25, 055012 (2008) doi:10.1088/0264-9381/25/5/055012 [arXiv:0710.5671 [gr-qc]].
  • (16) G. ’t Hooft, Nucl. Phys. B 256, 727-745 (1985) doi:10.1016/0550-3213(85)90418-3
  • (17) L. Hodgkinson, J. Louko and A. C. Ottewill, Phys. Rev. D 89, no.10, 104002 (2014) doi:10.1103/PhysRevD.89.104002 [arXiv:1401.2667 [gr-qc]].
  • (18) B. Holdom, [arXiv:2004.11285 [gr-qc]].