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

Termination of Superradiance from a Binary Companion

Xi Tong xtongac@connect.ust.hk    Yi Wang phyw@ust.hk    Hui-Yu Zhu hzhuav@connect.ust.hk Department of Physics, The Hong Kong University of Science and Technology, Clear Water Bay, Kowloon, Hong Kong, P.R.China The HKUST Jockey Club Institute for Advanced Study, The Hong Kong University of Science and Technology, Clear Water Bay, Kowloon, Hong Kong, P.R.China
Abstract

We study the impact of a binary companion on black hole superradiance at orbital frequencies away from the gravitational-collider-physics (GCP) resonance bands. A superradiant state can couple to a strongly absorptive state via the tidal perturbation of the companion, thereby acquiring a suppressed superradiance rate. Below a critical binary separation, this superradiance rate becomes negative, and the boson cloud gets absorbed by the black hole. This critical binary separation leads to tight constraints on GCP. Especially, a companion with mass ratio q>103q>10^{-3} invalidates all GCP fine structure transitions, as well as almost all Bohr transitions except those from the |ψ211|\psi_{211}\rangle state. Meanwhile, the backreaction on the companion manifests itself as a torque acting on the binary, producing floating/sinking orbits that can be verified via pulsar timing. In addition, the possible termination of cloud growth may help to alleviate the current bounds on the ultralight boson mass from various null detections.

I Introduction

Superradiance instability of ultralight bosons near a spinning black hole is a well-studied topic. The dissipative ergoregion of a Kerr black hole amplifies incoming waves and radiates out particles classically, a phenomenon widely known as superradiance [1, 2, 3]. More interestingly, if the bosonic particle carries a non-zero mass, its mass barrier located at a Compton wavelength away from the ergoregion reflects the amplified wave back, thereby generating more particles [4, 5, 6]. These copiously produced bosons then condensate into clouds surrounding the black hole, with structures similar to those of an electron in the hydrogen atom [6, 7].

Such a gravitational atom enjoys very rich phenomenology. For an isolated gravitational atom, the boson cloud quickly extracts the spin of black hole up to a saturation value. Observing black holes with spin higher than the saturation value then constrains superradiance and the boson properties [8, 9, 10]. The boson cloud also emits monochromatic gravitational waves via pair annihilations [7, 11], which are potential targets for gravitational wave detectors such as LIGO and LISA. Null detection then puts bounds on the boson mass [12, 13, 14, 15]. On the other hand, if the gravitational atom is in a binary system, more exciting physics kick in. For instance, a gravitational atom in a binary exhibits resonant transitions triggered by the orbital motion [16, 17, 18], observable via gravitational wave probes and pulsar timing [19, 20]. These so-called Gravitational Collider Physics (GCP) [17] transitions contain valuable information about the structure of the cloud as well as the properties of the boson. In addition, the mass quadrupole of the cloud induces orbital precession in an eccentric binary [21], observable from the gravitational wave or pulsar-timing signatures. At orbital separations close to the cloud radius, molecular structures emerge with distinctive observational signatures [22, 23, 24].

However, we point out that, hidden in many of the phenomena above, is the crucial assumption of the existence of a boson cloud with total mass not far away from (e.g.e.g., 11-33 orders of magnitude smaller than) that of the black hole itself. In the case of a truly isolated black hole, superradiance guarantees the saturation of cloud growth, and the cloud can only deplete slowly via gravitational wave emission. In the presence of a binary companion, however, the stability of cloud is far from obvious. For example, it is known that the resonant GCP transitions may deplete the cloud efficiently [16, 25, 26], while ionization effects may evaporate the cloud [27, 18, 28].

In this work, we highlight the fact that the superradiant modes (e.g.,|ψ322e.g.,~{}|\psi_{322}\rangle) can overlap with the dangerous absorptive modes (e.g.,|ψ300e.g.,~{}|\psi_{300}\rangle) via the tidal perturbation of the binary companion. Depending on the binary separation, the superradiance growth rate may be suppressed, and even become negative if the companion is too close (see FIG. 1 for a cartoon illustration). Consequently, the cloud cannot exist when the binary separation is below a critical distance. This dramatically affects the gravitational atom phenomenology. First, the parameter space of certain GCP transitions becomes constrained. Second, we may still observe orbital period derivative changes due to the backreaction of cloud absorption, which produces floating/sinking orbits away from the GCP resonance bands. Third, the bound on boson masses from various null detections may be alleviated, due to unknown companion objects depleting the cloud.

Refer to caption
Figure 1: The life cycle of a boson near a rotating black hole in binary. I: The boson is produced via black hole superradiance instability. II: The tidal perturbation of the binary companion (gray) turns the boson from a superradiant state (|ψ322|\psi_{322}\rangle) to an absorptive state (|ψ300|\psi_{300}\rangle) in an off-resonance fashion. III: The boson is quickly reabsorbed by the black hole. This whole process contributes negatively to the growth rate of the boson cloud, thereby suppressing superradiance.

This paper is organized as follows. In Sect. II, we first review the basic ingredients of black hole superradiance and tidal perturbation theory. Then in Sect. III, we compute the suppressed superradiance rate due to the gravitational perturbation of the binary, and determine the critical distances below which superradiance is turned off. We discuss the implications for the gravitational atom phenomenology in Sect. IV. At last, we conclude in Sect. V. We set G==c=1G=\hbar=c=1 throughout the paper and our conventions and notations largely follow from [17, 19].

II The gravitational atom and tidal perturbations

In this section, we begin with a lightning review of black hole superradiance instability and tidal perturbation theory. Consider an ultralight (scalar) boson field with mass μ\mu around a Kerr black hole with mass MM and dimensionless spin parameter a~\tilde{a}. The bosons generated by the Kerr black hole condensate into clouds and form a bound state together with the black hole. This spectrum of this gravitational atom is governed by the Klein-Gordon equation in Kerr spacetime:

(gαβαβμ2)Φ=0.\left(g^{\alpha\beta}\,\nabla_{\alpha}\,\nabla_{\beta}-\mu^{2}\right)\Phi=0~{}. (1)

If the gravitational fine structure constant αGMμ1\alpha\equiv GM\mu\ll 1 is perturbatively small, the cloud is mostly non-relativistic and we can factor out a rest-mass dynamical phase from the boson field,

Φ12μeiμtψ+c.c..\Phi\equiv\frac{1}{\sqrt{2\mu}}e^{-i\mu t}\psi+\text{c.c.}~{}. (2)

Using the Boyer-Lindquist coordinates, we insert (2) into the Klein-Gordon equation and expand in powers of α\alpha, the resulting equation takes a Schrödinger form at leading order,

itψ(t,𝐫)=H0ψ(t,𝐫),H012μ𝐫2+V(r),i\partial_{t}\psi(t,\mathbf{r})=H_{0}\psi(t,\mathbf{r})~{},~{}H_{0}\equiv-\frac{1}{2\mu}\partial_{\mathbf{r}}^{2}+V(r)~{}, (3)

where the potential V(r)=αr+𝒪(α2)V(r)=-\frac{\alpha}{r}+\mathcal{O}(\alpha^{2}) resembles that of a hydrogen atom. Solving the system with in-going boundary condition at the black hole outer horizon yields the quasi-stationary bound states, which at large distance (rMr\gg M) approximate to

ψnlm(𝒓)=Rnl(r)Ylm(θ,ϕ)ei(ωnlmμ)t,\psi_{nlm}(\bm{r})=R_{nl}(r)\,Y_{lm}(\theta,\phi)\,e^{-i\,(\omega_{nlm}-\mu)t}~{}, (4)

where RnlR_{nl} is the hydrogen radial function and n,l,mn,l,m are the usual principle, angular and magnetic quantum numbers. The key difference from that of hydrogen atom is that the eigenvalues of boson clouds are in general complex (hence the instability),

ωnlm=Enlm+iΓnlm,\omega_{nlm}=E_{nlm}+i\,\Gamma_{nlm}~{}, (5)

due to an absorptive horizon and dissipative ergoregion. The detailed form of the energy levels and decay widths can be found from inspecting the higher-order corrections of the Kerr geometry. Up to hyperfine splittings, the energy levels are given by [16, 29]

Enlm\displaystyle E_{nlm} =μ(1α22n2α48n4(3n2l1)α4n4(l+1/2))\displaystyle=\mu\left(1-\frac{\alpha^{2}}{2\,n^{2}}-\frac{\alpha^{4}}{8\,n^{4}}-\frac{(3\,n-2\,l-1)\,\alpha^{4}}{n^{4}\,(l+1/2)}\right) (6)
+2a~mα5n3l(l+1/2)(l+1)+𝒪(α6).\displaystyle+\frac{2\,\tilde{a}\,m\,\alpha^{5}}{n^{3}\,l\,(l+1/2)\,(l+1)}+\mathcal{O}(\alpha^{6})~{}.

The decay widths are given at α0.4\alpha\lesssim 0.4 via the Detweiler approximation111Note that numerics shows a factor-of-two uncertainty in the Detweiler approximation, although this is not important for our purpose. See a recent refinement in [30] [31],

Γn00\displaystyle\Gamma_{n00} =4n3(1+1a~2)μα5\displaystyle=-\frac{4}{n^{3}}(1+\sqrt{1-\tilde{a}^{2}})\,\mu\,\alpha^{5} (7)
Γnlm\displaystyle\Gamma_{nlm} =2r~+Cnlglm(a~,α,ω)(mΩHωnlm)α4l+5.\displaystyle=2\tilde{r}_{+}\,C_{nl}\,g_{lm}(\tilde{a},\alpha,\omega)\,(m\,\Omega_{H}-\omega_{nlm})\,\alpha^{4l+5}.

Here r~r/M\tilde{r}\equiv r/M, r~+1+1a~2\tilde{r}_{+}\equiv 1+\sqrt{1-\tilde{a}^{2}} and ΩHa~/[2M2(1+1+a~2)]\Omega_{H}\equiv\tilde{a}/[2\,M^{2}\,(1+\sqrt{1+\tilde{a}^{2}})] is the angular velocity of the outer horizon. The definition of CnlC_{nl} and glmg_{lm} are given as.

Cnl24l+1(n+l)!n2l+4(nl1)![l!(2l)!(2l+1!)]2,C_{nl}\equiv\frac{2^{4l+1}(n+l)!}{n^{2l+4}(n-l-1)!}\,\left[\frac{l!}{(2l)!\,(2l+1!)}\right]^{2}, (8)
glm(a~,α,ω)k=1l(k2(1a~2)+(a~m2r~+Mω)2).g_{lm}(\tilde{a},\alpha,\omega)\equiv\prod_{k=1}^{l}\left(k^{2}\,(1-\tilde{a}^{2})+(\tilde{a}m-2\tilde{r}_{+}M\omega)^{2}\right). (9)

For an isolated black hole, the occupation number of superradiant states (Γnlm>0\Gamma_{nlm}>0) grows at a timescale Tnlm(grow)=Γnlm1α(4l+5)T_{nlm}^{(\text{grow})}=\Gamma_{nlm}^{-1}\propto\alpha^{-(4l+5)} until saturation. In contrast, states with Γnlm<0\Gamma_{nlm}<0 will be absorbed into the black hole at a timescale Tnlm(absorb)=Γnlm1α(4l+5)T_{nlm}^{(\text{absorb})}=\Gamma_{nlm}^{-1}\propto\alpha^{-(4l+5)}. After saturation, the superradiant states constantly emit monochromatic gravitational waves and deplete at a much longer time scale T(deplete)α(4l+10)T^{(\text{deplete})}\propto\alpha^{-(4l+10)}.

Now consider inducing a binary companion of mass MM_{*}. At leading order, we can understand the influence of a binary companion as generating a time-periodic tidal field around the gravitational atom. In the Fermi normal coordinates, the Newtonian potential of the binary companion can be expanded using spherical harmonics as

V=\displaystyle V_{*}= αql2|m|llm(ι,φ)Ylm(θ,ϕ)\displaystyle-\alpha\,q\,\sum_{l_{*}\geqslant 2}\sum_{|m_{*}|\leqslant l_{*}}\,\mathcal{E}_{l_{*}m_{*}}(\iota_{*},\varphi_{*})\,Y_{l_{*}m_{*}}(\theta,\phi)
×(rlRl+1Θ(Rr)+Rlrl+1Θ(rR)).\displaystyle\times\left(\frac{r^{l_{*}}}{R_{*}^{l_{*}+1}}\,\Theta(R_{*}-r)+\frac{R_{*}^{l_{*}}}{r_{*}^{l_{*}+1}}\,\Theta(r-R_{*})\right)~{}. (10)

Here q=M/Mq=M_{*}/M is the mass ratio of the companion and the gravitational atom, RR_{*} is the binary separation, and ι,φ\iota_{*},\varphi_{*} are the inclination angle and true anomaly, respectively. Θ(x)\Theta(x) is the Heaviside step function. The tidal moment \mathcal{E} can be computed from the geometry of the binary. For equatorial orbits, the quadrupole contribution simplifies considerably to

22=126π5e±2iφ,21=0,20=π5.\displaystyle\mathcal{E}_{2\mp 2}=\frac{1}{2}\sqrt{\frac{6\pi}{5}}e^{\pm 2i\varphi_{*}},\mathcal{E}_{2\mp 1}=0,\mathcal{E}_{20}=-\sqrt{\frac{\pi}{5}}~{}. (11)

The tidal field of the binary companion decays as Rl+1R_{*}^{l_{*}+1} for the ll_{*}-th multipole moment. Although VV_{*} is small for large separations, it introduces weak mixings between energy eigenstates of H0H_{0}. This overlap of two H0H_{0} eigenstates is calculated as

ψnlm|V(t)|ψnlm\displaystyle\langle\psi_{n^{\prime}l^{\prime}m^{\prime}}|V_{*}(t)|\psi_{nlm}\rangle\equiv (1)m+1αq\displaystyle(-1)^{m^{\prime}+1}\alpha\,q
l,mlm(t)𝒢mmmlllIr,\displaystyle\sum_{l_{*},m_{*}}\mathcal{E}_{l_{*}m_{*}}(t)\mathcal{G}^{l^{\prime}l_{*}l}_{-m^{\prime}m_{*}m}\,I_{r}, (12)

Here 𝒢mmmlll\mathcal{G}^{l^{\prime}l_{*}l}_{-m^{\prime}m_{*}m} is the Gaunt integral,

𝒢mmmlll=dΩYlm(θ,ϕ)Ylm(θ,ϕ)Ylm(θ,ϕ),\mathcal{G}_{-m^{\prime}m_{*}m}^{l^{\prime}l_{*}l}=\int{\rm{d}}\Omega Y_{l^{\prime}-m^{\prime}}(\theta,\phi)Y_{l_{*}m_{*}}(\theta,\phi)Y_{lm}(\theta,\phi)~{}, (13)

which implicitly imposes a set of selection rules,

m+m+m=0\displaystyle-m^{\prime}+m_{*}+m=0 (14)
l+l+l=2p, for p\displaystyle l+l_{*}+l^{\prime}=2p,\text{ for }p\in\mathbb{Z}
|ll|ll+l.\displaystyle|l-l^{\prime}|\leqslant l_{*}\leqslant l+l^{\prime}~{}.

And IrI_{r} is the radial integral,

Ir\displaystyle I_{r} =0Rr2drRnl(r)Rnl(r)rlRl+1\displaystyle=\int_{0}^{R^{*}}r^{2}{\rm{d}}rR^{*}_{n^{\prime}l^{\prime}}(r)R_{nl}(r)\frac{r^{l_{*}}}{R_{*}^{l_{*}+1}} (15)
+Rr2drRnl(r)Rnl(r)Rlrl+1.\displaystyle+\int_{R^{*}}^{\infty}r^{2}{\rm{d}}rR^{*}_{n^{\prime}l^{\prime}}(r)R_{nl}(r)\frac{R_{*}^{l_{*}}}{r_{*}^{l_{*}+1}}~{}.

The radial integral is dominated by the first term if the companion is far outside the cloud, i.e.i.e., Rn2r1R_{*}\gg n^{2}r_{1}, where r1(μα)1=M/α2r_{1}\equiv(\mu\alpha)^{-1}=M/\alpha^{2} is the Bohr radius.

The dynamics of the system is then determined from solving the full Hamiltonian HH0+VH\equiv H_{0}+V_{*} with orbital backreaction. It is widely known that the periodic tidal perturbation VV_{*} can trigger atomic resonances when the orbital frequency matches the energy difference of two H0H_{0}-eigenstates. These GCP resonances exhibit distinctive floating/sinking features in their orbital evolutions, which can be observed via multiple messengers [20].

III Termination of superradiance

One might think that when the binary separation is much larger than the cloud radius, i.e.i.e., Rr1(μα)1R_{*}\gg r_{1}\sim(\mu\alpha)^{-1}, the gravitational perturbation of the binary companion (Vαr12/R3μα2\langle V_{*}\rangle\sim\alpha r_{1}^{2}/R_{*}^{3}\ll\mu\alpha^{2}) may be too small to affect the cloud dynamics, which is typically of order H0μα2\langle H_{0}\rangle\sim\mu\alpha^{2}. However, one must recall that the superradiance rate, which governs cloud formation, is also small. For instance, the |ψ322|\psi_{322}\rangle state has a superradiance rate Γ322μα12\Gamma_{322}\sim\mu\alpha^{12}. Therefore, it might be the case that the cloud formation process can be dramatically influenced by the presence of a binary companion. Indeed, we will see that the tidal perturbation of the binary companion typically suppresses the superradiance rate, and may even terminate superradiance and prevent the cloud from ever growing up.

III.1 The adiabatic case

We start with the perturbed Hamiltonian H=H0+VH=H_{0}+V_{*} of the boson. Its matrix element is given by

ψnlm|H|ψnlm=ωnlmδnnδllδmm+ψnlm|V|ψnlm.\langle\psi_{n^{\prime}l^{\prime}m^{\prime}}|H|\psi_{nlm}\rangle=\omega_{nlm}\delta_{n^{\prime}n}\delta_{l^{\prime}l}\delta_{m^{\prime}m}+\langle\psi_{n^{\prime}l^{\prime}m^{\prime}}|V_{*}|\psi_{nlm}\rangle~{}. (16)

The diagonal terms are dominated by the eigenvalues of the free Hamiltonian while the off-diagonal terms are led by the gravitational level mixings due to the companion. The level mixings couple two states together whenever the selection rule allows. In particular, the superradiant states can be coupled to the absorptive states, bringing negative contributions to the imaginary part of the frequencies. Most dangerous of all absorptive states are the spherically symmetric |ψn00|\psi_{n00}\rangle states, since they decay the fastest. Hence we expect the leading order correction to the superradiant rate of |ψnlm|\psi_{nlm}\rangle to come from its mixing with |ψn00|\psi_{n00}\rangle (if the selection rule allows). Therefore, we will focus on a two-state subspace {|1,|2}\{|1\rangle,|2\rangle\}, where |1|1\rangle denotes a superradiant state |ψnlm|\psi_{nlm}\rangle and |2|2\rangle denotes a highly absorptive state such as |ψn00|\psi_{n00}\rangle. In matrix form, the perturbed Hamiltonian reads

H=(ω1+V11V12V21ω2+V22)(E¯1+iΓ1ηηE¯2+iΓ2),\displaystyle H=\left(\begin{matrix}\omega_{1}+V_{11}&V_{12}\\ V_{21}&\omega_{2}+V_{22}\end{matrix}\right)\equiv\left(\begin{matrix}\bar{E}_{1}+i\Gamma_{1}&\eta^{*}\\ \eta&\bar{E}_{2}+i\Gamma_{2}\end{matrix}\right)~{}, (17)

where we have denoted ηV21\eta\equiv V_{21}, ω¯iωi+Vii\bar{\omega}_{i}\equiv\omega_{i}+V_{ii} and E¯iEi+Vii\bar{E}_{i}\equiv E_{i}+V_{ii}, i=1,2i=1,2. Notice that both E¯i(t)\bar{E}_{i}(t) and η(t)\eta(t) implicitly depend on time through the binary orbital motion. Thus in the adiabatic limit |η˙E¯i2|,|E¯˙iE¯i2|1|\frac{\dot{\eta}}{\bar{E}_{i}^{2}}|,|\frac{\dot{\bar{E}}_{i}}{\bar{E}_{i}^{2}}|\ll 1, the state of the cloud ci(t)ψi|ψ(t)c_{i}(t)\equiv\langle\psi_{i}|\psi(t)\rangle can be solved using the WKB approximation,

ci(t)=Ci+eiλ+𝑑t+Cieiλ𝑑t,i=1,2,c_{i}(t)=C_{i+}e^{-i\int\lambda_{+}dt}+C_{i-}e^{-i\int\lambda_{-}dt}~{},~{}i=1,2~{}, (18)

where λ±\lambda_{\pm} are the two instantaneous eigenvalues of the perturbed Hamiltonian (17),

λ±\displaystyle\lambda_{\pm}\equiv ω¯1+ω¯22±|η|2+(ω¯1ω¯22)2\displaystyle\frac{\bar{\omega}_{1}+\bar{\omega}_{2}}{2}\pm\sqrt{|\eta|^{2}+\left(\frac{\bar{\omega}_{1}-\bar{\omega}_{2}}{2}\right)^{2}}
\displaystyle\simeq {E¯1+|η|2E¯1E¯2+i[Γ1Γ1Γ2(E¯1E¯2)2|η|2]+E¯2+|η|2E¯2E¯1+i[Γ2Γ2Γ1(E¯1E¯2)2|η|2]\displaystyle\left\{\begin{array}[]{lr}\bar{E}_{1}+\frac{|\eta|^{2}}{\bar{E}_{1}-\bar{E}_{2}}+i\left[\Gamma_{1}-\frac{\Gamma_{1}-\Gamma_{2}}{(\bar{E}_{1}-\bar{E}_{2})^{2}}|\eta|^{2}\right]\text{, }+&\\ \\ \bar{E}_{2}+\frac{|\eta|^{2}}{\bar{E}_{2}-\bar{E}_{1}}+i\left[\Gamma_{2}-\frac{\Gamma_{2}-\Gamma_{1}}{(\bar{E}_{1}-\bar{E}_{2})^{2}}|\eta|^{2}\right]\text{, }-&\end{array}\right. (22)

The approximation is made under the limit |E¯1,2||η|,|Γ1,2||\bar{E}_{1,2}|\gg|\eta|,|\Gamma_{1,2}|, which is always valid given the superradiance context. Thus we see that λ+\lambda_{+} (λ\lambda_{-}) represents the corrected frequency of the quasi-stationary state |1|1\rangle (|2|2\rangle). In particular, their imaginary parts are modified. Since we have assumed Γ10\Gamma_{1}\geqslant 0 and Γ2<0\Gamma_{2}<0, the effective superradiance rate is now suppressed:

Γ~1Γ1+ΔΓ1,ΔΓ1Γ1Γ2(E¯1E¯2)2|η(R)|2<0.\tilde{\Gamma}_{1}\equiv\Gamma_{1}+\Delta\Gamma_{1},~{}\Delta\Gamma_{1}\simeq-\frac{\Gamma_{1}-\Gamma_{2}}{(\bar{E}_{1}-\bar{E}_{2})^{2}}|\eta(R_{*})|^{2}<0~{}. (23)

This suppression term in the effective superradiance rate mainly depends on the distance RR_{*} between the binary. If the absorptive state |2|2\rangle happens to be a highly dangerous state such as |ψn00|\psi_{n00}\rangle, the corresponding suppression can be significant if RR_{*} is not large. For instance, consider the case n=3,l=m=2n=3,l=m=2. The fine-structure splitting is

E322E3000.04s1(α0.1)5(M10M)1.E_{322}-E_{300}\simeq 0.04~{}\text{s}^{-1}\left(\frac{\alpha}{0.1}\right)^{5}\left(\frac{M}{10M_{\odot}}\right)^{-1}~{}. (24)

The superradiance/absorption rates at maximal black hole spin (a~=1\tilde{a}=1) are

Γ322\displaystyle\Gamma_{322} 3×1013s1(α0.1)13(M10M)1,\displaystyle\simeq 3\times 10^{-13}~{}\text{s}^{-1}\left(\frac{\alpha}{0.1}\right)^{13}\left(\frac{M}{10M_{\odot}}\right)^{-1}~{}, (25)
Γ300\displaystyle\Gamma_{300} 3×103s1(α0.1)5(M10M)1.\displaystyle\simeq-3\times 10^{-3}~{}\text{s}^{-1}\left(\frac{\alpha}{0.1}\right)^{5}\left(\frac{M}{10M_{\odot}}\right)^{-1}~{}. (26)

Then at a binary separation R=105MR_{*}=10^{5}M, the size of the level mixing is

|η(R)|2×107s1(α0.1)3q0.2(M10M)1.|\eta(R_{*})|\simeq 2\times 10^{-7}~{}\text{s}^{-1}\left(\frac{\alpha}{0.1}\right)^{-3}\frac{q}{0.2}\left(\frac{M}{10M_{\odot}}\right)^{-1}~{}. (27)

This results in a correction to the superradiance rate

ΔΓ322\displaystyle\Delta\Gamma_{322}\simeq 7×103q2α10M5R6\displaystyle-7\times 10^{3}\frac{q^{2}}{\alpha^{10}}\frac{M^{5}}{R_{*}^{6}}
\displaystyle\simeq 0.6×1013s1(α0.1)10(q0.2)2(M10M)1,\displaystyle-0.6\times 10^{-13}~{}\text{s}^{-1}\left(\frac{\alpha}{0.1}\right)^{-10}\left(\frac{q}{0.2}\right)^{2}\left(\frac{M}{10M_{\odot}}\right)^{-1}~{}, (28)

which already gives Γ~32245Γ322\tilde{\Gamma}_{322}\sim\frac{4}{5}\Gamma_{322}, i.e.i.e., a reduction of 20%20\%. As one expects from the quadrupole tidal perturbation, ΔΓ322\Delta\Gamma_{322} grows as R6R_{*}^{-6} as the binary separation decreases. Therefore, inside a critical distance R<R,cR_{*}<R_{*,c} defined from the condition

Γ~nlm(R,c)=Γnlm+ΔΓnlm(R,c)0,\tilde{\Gamma}_{nlm}(R_{*,c})=\Gamma_{nlm}+\Delta\Gamma_{nlm}(R_{*,c})\equiv 0~{}, (29)

the effective superradiance rate becomes negative, and the mode |ψnlm|\psi_{nlm}\rangle can no longer grow even at maximal black hole spin. For the |ψ322|\psi_{322}\rangle example, the critical distance is easily read out from (25) and (28):

R,c(322)106km(α0.1)23/6(q0.2)1/3M10M.R_{*,c}(322)\simeq 10^{6}~{}\text{km}\left(\frac{\alpha}{0.1}\right)^{-23/6}\left(\frac{q}{0.2}\right)^{1/3}\frac{M}{10M_{\odot}}~{}. (30)

Before moving on to further discussions, we would like to make a few comments on suppressed superradiance and the critical distance.

  1. \bullet

    First, it is interesting to see the interplay of UV and IR effects here. The imaginary part of the eigenvalues of the free Hamiltonian H0H_{0} can be understood as an UV effect, which is relevant near the black hole horizon scale rUVMr_{UV}\sim M. Yet its correction involves IR effects which has nothing to do with the black hole geometry. These co-rotating bosons produced via superradiant scattering leak to an IR scale rIR(μα)1r_{IR}\sim(\mu\alpha)^{-1}, and are turned into non-/counter-rotating bosons by the tidal perturbation of the binary companion, which then get absorbed by the black hole again in the UV (see again FIG. 1).

  2. \bullet

    Second, if the binary separation is shorter than the critical distance R,c(nlm)R_{*,c}(nlm), there can be two scenarios. If the cloud has not been formed, yet the black hole spin appears to be capable of superradiance (Γnlm>0\Gamma_{nlm}>0), superradiance will be shut off. No |ψnlm|\psi_{nlm}\rangle-cloud will be produced, and black hole spin cannot drop below the saturation value for modes with magnetic quantum number mm. This is because, unlike Hawking radiation, superradiance is not spontaneous and can only amplify a given cloud state. But if such a cloud state is easily depleted, there is no seed particles left for amplification, hence no extraction of black hole spin. On the other hand, if the cloud has already been formed, it will decay slowly at the rate |Γ~nlm||\tilde{\Gamma}_{nlm}|. Like GCP resonances, this process generates backreaction to the binary orbit and can be observed. More discussions on this possibility will be elaborated in Sect. IV.2.

  3. \bullet

    Third, the superradiance termination we are dealing with here is similar but not identical to resonant depletion scenarios considered in the literature [16, 25, 26]. Specifically, we need no requirement on the orbital frequency or orbit direction (co-rotating/counter-rotating). In fact, we will mostly stay away from the GCP resonance bands and analyze the viability of GCP resonances considering superradiance termination in Sect. IV.1.

III.2 Beyond adiabaticity and multiple states

The above calculations are performed under the assumption of adiabaticity. However, when the orbital frequency is high, the naive WKB approximation (18) breaks down and we need a better treatment of the perturbed Hamiltonian HH. For simplicity, we will specialize to the case of equatorial circular orbits. The tidal moments thus reduce to

lm=4π2l+1Ylm(π/2,φ(t))=elmeimφ(t),\mathcal{E}_{l_{*}m_{*}}=\frac{4\pi}{2l_{*}+1}Y_{l_{*}m_{*}}(\pi/2,\varphi_{*}(t))=e_{l_{*}m_{*}}e^{im_{*}\varphi_{*}(t)}~{}, (31)

where

elm4π2l+1(lm)!(l+m)!Plm(0)e_{l_{*}m_{*}}\equiv\sqrt{\frac{4\pi}{2l_{*}+1}\frac{(l_{*}-m_{*})!}{(l_{*}+m_{*})!}}P_{l_{*}}^{m_{*}}(0) (32)

and Plm(x)P_{l_{*}}^{m_{*}}(x) is the associated Legendre polynomial. This eimφ(t)e^{im_{*}\varphi_{*}(t)} time dependence is inherited by the off-diagonal term η(t)\eta(t). Therefore, we follow the standard procedure and perform a time-dependent unitary transformation to go into the co-rotating frame [32, 17, 22],

HD=U(t)(H(t)it)U(t),\displaystyle H_{D}=U(t)^{\dagger}(H(t)-i\partial_{t})U(t),
with U(t)eiφ(t)Lz,\displaystyle\text{with }U(t)\equiv e^{-i\varphi_{*}(t)L_{z}}~{}, (33)

where LzL_{z} is the cloud angular momentum operator along the spin axis. In component form, the Hamiltonian in the co-rotating frame reads

HD=(E¯1+iΓ1m1φ˙|η||η|E¯2+iΓ2m2φ˙).H_{D}=\left(\begin{matrix}\bar{E}_{1}+i\Gamma_{1}-m_{1}\dot{\varphi}_{*}&|\eta|\\ |\eta|&\bar{E}_{2}+i\Gamma_{2}-m_{2}\dot{\varphi}_{*}\end{matrix}\right)~{}. (34)

Notice that we have applied the selection rule (14) and set m=m2m1m_{*}=m_{2}-m_{1}. Now the fast oscillations in HH are eliminated and we are left with a co-rotating frame Hamiltonian HDH_{D} that varies slowly in time only through R(t)R_{*}(t). Going through the same procedure as before, we obtain the correction to the superradiance rate of |1|1\rangle:

ΔΓ1Γ1Γ2[E¯1E¯2(m1m2)φ˙(R)]2|η(R)|2.\Delta\Gamma_{1}\simeq-\frac{\Gamma_{1}-\Gamma_{2}}{\left[\bar{E}_{1}-\bar{E}_{2}-(m_{1}-m_{2})\dot{\varphi}_{*}(R_{*})\right]^{2}}|\eta(R_{*})|^{2}~{}. (35)

Clearly, in the adiabatic limit, |φ˙Ei|1|\frac{\dot{\varphi}}{E_{i}}|\ll 1, the above expression reduces to (23). The impact on superradiance becomes severe when the denominator vanishes. This corresponds to the resonant depletion scenario [16, 25, 26], where the cloud state |1|1\rangle largely mixes into |2|2\rangle and is quickly absorbed into the black hole. Since we focus more on the off-resonance scenario, the denominator will typically be dominated either by E¯1E¯2\bar{E}_{1}-\bar{E}_{2} (the adiabatic case) or by (m1m2)φ˙(m_{1}-m_{2})\dot{\varphi} (the diabatic case).

Since both the tidal perturbations and the instability rates are small compared to the energy levels, the two-state result (35) can be readily generalized to that of multiple states. To leading order in |ηij||\eta_{ij}|, the impact on superradiance is just a simple summation,

ΔΓ1i=nlmΓ1Γi[E¯1E¯i(m1mi)φ˙(R)]2|η1i(R)|2,\Delta\Gamma_{1}\simeq-\sum_{i=n^{\prime}l^{\prime}m^{\prime}}\frac{\Gamma_{1}-\Gamma_{i}}{\left[\bar{E}_{1}-\bar{E}_{i}-(m_{1}-m_{i})\dot{\varphi}_{*}(R_{*})\right]^{2}}|\eta_{1i}(R_{*})|^{2}~{}, (36)

where η1i\eta_{1i} contains a sum of tidal moments l,mim1\mathcal{E}_{l_{*},m_{i}-m_{1}}. The critical distance is again determined by requiring Γ~1=Γ1+ΔΓ1=0\tilde{\Gamma}_{1}=\Gamma_{1}+\Delta\Gamma_{1}=0. Due to the RR_{*} dependence in the denominator, there may be multiple solutions of R,cR_{*,c} when entering/exiting a resonance. We are more interested in the off-resonance solutions.

Refer to caption
Figure 2: The effective superradiance rate Γ~nlm\tilde{\Gamma}_{nlm} as a function of the binary separation RR_{*} for co-rotating orbits (upper panel) and counter-rotating orbits (lower panel). The solid lines correspond to the maximal black hole spin a~=1\tilde{a}=1, while the dashed lines correspond to the previous saturation black hole spin 4(m1)α(m1)2+4α2\frac{4(m-1)\alpha}{(m-1)^{2}+4\alpha^{2}} for the mm-th state (e.g.e.g., a~=2α1+α2\tilde{a}=\frac{2\alpha}{1+\alpha^{2}} when considering the m=3m=3 state). The other parameters are chosen to be α=0.1\alpha=0.1, M=10MM=10M_{\odot} and q=0.2q=0.2. It is clear that the effective superradiance rate reduces to the Detweiler value at large binary separations, while dropping below zero generically at a finite critical distance.

In FIG. 2, we have plotted the binary-separation-dependence of the effective superradiance rate Γ~nlm(R)\tilde{\Gamma}_{nlm}(R_{*}) for different states and black hole spins. It can be seen that the effective superradiance rate Γ~nlm\tilde{\Gamma}_{nlm} tends to the Detweiler value Γnlm\Gamma_{nlm} at large binary separations, since the tidal perturbations become negligible in this limit. As we decrease the binary separation, Γ~nlm\tilde{\Gamma}_{nlm} quickly drops to zero near a critical distance, where superradiance is terminated due to mixing into absorptive states. The sharp peaks and valleys are caused by GCP resonances, where the mixing effect is non-perturbatively large. Interestingly, when the black hole spin is high (e.g.e.g., a~=1\tilde{a}=1), the high-ll states such as |ψ433|\psi_{433}\rangle and |ψ544|\psi_{544}\rangle receive positive enhancement to the superradiance rate first, before negative suppression terms take over and terminate superradiance as RR_{*} decreases. This is because of their mixings into low-ll states such as |ψ411|\psi_{411}\rangle and |ψ522|\psi_{522}\rangle, which possess larger positive growth rates. Yet as the black hole spin drops below the threshold for superradiating these low-ll states, the superradiance rate Γ~nlm\tilde{\Gamma}_{nlm} for high-ll states can no longer benefit from the mixings, and thus monotonically drop to zero as RR_{*} decreases, as indicated by the dashed lines in FIG. 2.

The critical distance can be solved numerically from (29) in the case of multiple states beyond adiabaticity. As mentioned before, we focus on systems with an orbital frequency away from the GCP resonance bands, because such a configuration occupies the most proportion of the binary lifetime and are more typical statistically. Henceforth, it will be convenient to perform a “Wick” rotation and replace the denominator in (36) by

1[E¯1E¯i(m1mi)φ˙(R)]2\displaystyle\frac{1}{\left[\bar{E}_{1}-\bar{E}_{i}-(m_{1}-m_{i})\dot{\varphi}_{*}(R_{*})\right]^{2}}
1(E¯1E¯i)2+[(m1mi)φ˙(R)]2.\displaystyle\to\frac{1}{(\bar{E}_{1}-\bar{E}_{i})^{2}+[(m_{1}-m_{i})\dot{\varphi}_{*}(R_{*})]^{2}}~{}. (37)

This modification removes the resonance poles, but it keeps the off-resonance physics mostly unaltered. Since the tidal-perturbation-theory calculations are under the assumption of R>rnR_{*}>r_{n}, the resulting critical distance must be greater than the cloud radius,

R,c(nlm)>rn=n2r1 for consistency.R_{*,c}(nlm)>r_{n}=n^{2}r_{1}\text{ for consistency}~{}. (38)

Under this constraint, we solve (29) for the critical distance and plot its dependence on α\alpha and qq in FIG. 3. Because we have chosen a maximal black hole spin, the resulting critical distance is the minimal requirement for cloud stability. This means superradiance of a given state |ψnlm|\psi_{nlm}\rangle will be terminated completely below R<R,c(nlm)R_{*}<R_{*,c}(nlm), and an existing cloud will decay quickly at a rate comparable to its original growth rate. From FIG. 3, we see that for fixed α\alpha and MM (hence fixed boson mass), the critical distance decreases with a smaller mass ratio. This is in agreement with intuition, since a lighter (heavier) binary companion is expected to have weaker (stronger) gravitational perturbations on the cloud dynamics, resulting in a broader (narrower) safe region. Cloud states with higher ll appear to be less stable in the presence of a binary companion, as their critical distances are larger. At last, we stress that as an off-resonance quantity, the critical distance solved in this way does not depend on the orbit orientation, which matters only for the triggering of a GCP resonance.

Refer to caption
Figure 3: The critical distance R,cR_{*,c} (in units of the horizon size) as a function of α\alpha and qq for different cloud states. The gray region is excluded by the consistency constraint (38). The black hole spin is chosen to be maximal, a~=1\tilde{a}=1. Therefore, below the critical distance, no net superradiance is present and no existing boson cloud can last for a period longer than its original typical growth time.

IV Observational consequences

Given the potential threat of a binary companion to the boson cloud, how does this superradiance termination effect influence the phenomenology of the gravitational atom? In this section, we will briefly examine three of its main consequences.

IV.1 Implications for GCP transitions

First, the termination of superradiance poses a constraint on observable GCP transitions. This is natural since GCP is based on resonant cloud transitions. If there is no cloud when the binary enters the resonance band, there is certainly no cloud transition, and no observable signal. Therefore, roughly speaking, in order to have successful GCP transitions, we must require that the binary separation at resonance to be greater than the critical distance of the initial cloud state222Notice that both sides of (39) depend on α\alpha and qq. Thus this is effectively a constraint on the relative sizes of three mass parameters (namely, μ,M,M\mu,M,M_{*}) in the system.,

R,r(nlmnlm)>R,c(nlm).R_{*,r}(nlm\to n^{\prime}l^{\prime}m^{\prime})>R_{*,c}(nlm)~{}. (39)

Otherwise, the cloud either cannot form, or would have been depleted via Γ~nlm<0\tilde{\Gamma}_{nlm}<0 long before entering the resonance band.

Scanning through the parameter space, we plot the “safe” region for various GCP transitions in FIG. 4. It is clear from the plot that superradiance termination poses a tight bound on the parameters of the binary system. Bohr transitions with floating orbits are most severely constrained, because they happen at relatively high orbital frequencies. This means the binary separation at resonance is short, and the cloud is vulnerable to absorption. Hyperfine transitions, on the contrary, are influenced the least. In particular, we note that the |ψ211|ψ211|\psi_{211}\rangle\to|\psi_{21~{}-1}\rangle and |ψ322|ψ300|\psi_{322}\rangle\to|\psi_{300}\rangle transitions are safe to occur throughout the parameter space. This is because they happen at low orbital frequencies, where the overlap between the superradiant mode and absorptive modes is still small. Overall, except the hyperfine ones and Bohr transitions starting from the state |ψ211|\psi_{211}\rangle, most transitions are forbidden for mass ratio q103q\gtrsim 10^{-3}. They are allowed only below a certain mass ratio, corresponding to a small mass of the binary companion.

Refer to caption
Figure 4: The “safe” region for having various hyperfine, fine and Bohr GCP transitions. The solid lines represent the boundary where the resonance distance is equal to the critical distance, and its shaded side satisfies (39) and is allowed. Note that the hyperfine transitions |ψ211|ψ211|\psi_{211}\rangle\to|\psi_{21~{}-1}\rangle and |ψ322|ψ300|\psi_{322}\rangle\to|\psi_{300}\rangle are allowed throughout the parameter space. For Bohr transitions, we also included the constraint R,r(nlmnlm)>max{n2r1,n2r1}R_{*,r}(nlm\to n^{\prime}l^{\prime}m^{\prime})>\max\left\{n^{2}r_{1},n^{\prime 2}r_{1}\right\} to keep the validity of tidal perturbation theory. The black hole spin is chosen to be the saturation value of the initial cloud state.

IV.2 Observing cloud absorption

Before entering the critical distance R,cR_{*,c}, the cloud may have already grown up. Then the negative effective superradiance rate within R,cR_{*,c} will gradually deplete the cloud. The loss of angular-momentum-carrying particles backreacts on the orbit and produces an effective torque on the binary companion, leading to floating/sinking orbits in a manner similar to GCP transitions. The difference here is that such orbital backreaction does not occur at a fixed resonance band.

The total angular momentum of the cloud is

Sc(t)=Sc,0(m1|c1(t)|2+i1mi|ci(t)|2),S_{c}(t)=S_{c,0}\left(m_{1}|c_{1}(t)|^{2}+\sum_{i\neq 1}m_{i}|c_{i}(t)|^{2}\right)~{}, (40)

where Sc,0S_{c,0} is the cloud angular momentum at the saturation of |ψ1|\psi_{1}\rangle, and 0|c1|210\leqslant|c_{1}|^{2}\leqslant 1 is the percentage of cloud occupation. It effectively describes the boson particle number, i.e.i.e., the larger |c1|2|c_{1}|^{2} is, there are more particles in the cloud. Since except the superradiant mode |ψ1|\psi_{1}\rangle with m1>0m_{1}>0, most other modes are usually unoccupied (ci0c_{i}\simeq 0), we can approximate Sc(t)Sc,0m1|c1(t)|2S_{c}(t)\simeq S_{c,0}m_{1}|c_{1}(t)|^{2}. In the case of a planar circular orbit, considering backreaction of the cloud evolution yields an orbital period derivative

P˙=(P˙)GR+(P˙)C,\dot{P}=(\dot{P})_{\text{GR}}+(\dot{P})_{\text{C}}~{}, (41)

where (P˙)GR(\dot{P})_{\text{GR}} represents the usual orbital decay in General Relativity (GR),

(P˙)GR=965(2π)8/3q(1+q)1/3M5/3P5/3,(\dot{P})_{\text{GR}}=-\frac{96}{5}(2\pi)^{8/3}\frac{q}{(1+q)^{1/3}}M^{5/3}P^{-5/3}~{}, (42)

and (P˙)C(\dot{P})_{\text{C}} is the extra contribution due to cloud backreaction [17, 19],

(P˙)C=3(2π)1/3(1+q)2/3Sc,0m1M2d|c1(t)|2dtM1/3P2/3.(\dot{P})_{\text{C}}=-3(2\pi)^{1/3}(1+q)^{-2/3}\frac{S_{c,0}m_{1}}{M^{2}}\frac{d|c_{1}(t)|^{2}}{dt}M^{1/3}P^{2/3}~{}. (43)

The cloud evolution would have been unitary but for the absorption, which means c1(t)eΓ~1tc_{1}(t)\propto e^{-\tilde{\Gamma}_{1}t} and

d|c1(t)|2dt=2Γ~1(R)|c1(t)|2.\frac{d|c_{1}(t)|^{2}}{dt}=-2\tilde{\Gamma}_{1}(R_{*})|c_{1}(t)|^{2}~{}. (44)

For instance, using Kepler’s law to rewrite RR_{*} in terms of PP, we can obtain the fractional correction to the period derivative for |ψ322|\psi_{322}\rangle,

(P˙)C(P˙)GR\displaystyle\frac{(\dot{P})_{\text{C}}}{(\dot{P})_{\text{GR}}}
\displaystyle\simeq 15|c322|2(α0.1)9q(1+q)7/3(M10M)5/3(P1 hr)5/3.\displaystyle-15|c_{322}|^{2}\left(\frac{\alpha}{0.1}\right)^{-9}\frac{q}{(1+q)^{7/3}}\left(\frac{M}{10M_{\odot}}\right)^{5/3}\left(\frac{P}{1\text{ hr}}\right)^{-5/3}. (45)

The minus sign shows that this is a floating orbit. Thus the correction to the orbital period derivative can be significant for certain parameter choices333Since the cloud has been decaying for some time after entering the critical distance, we generally expect |c1|2|c_{1}|^{2} to be a small number. In principle, assuming natural evolution, one can solve the whole history and determine |c1|2|c_{1}|^{2}. Here, for simplicity, we will treat it as a free parameter.. Such corrections should be detectable via multi-messengers such as gravitational waves and pulsar timing [33, 34]. For instance, consider the gravitational atom in a pulsar-black hole binary, the periastron time shift is calculated by

ΔPtP(0)0tdtP(t)12P˙Pt2,\Delta_{P}\equiv t-P(0)\int_{0}^{t}\frac{dt^{\prime}}{P(t^{\prime})}\approx\frac{1}{2}\frac{\dot{P}}{P}t^{2}~{}, (46)

where we have Taylor expanded the orbital period to the linear order. In order to observe the extra periastron time shift caused by the backreaction of cloud absorption, we must require the deviation from the GR result to be larger than the timing error of pulse counting,

|ΔP(ΔP)GR|>σΔP.|\Delta_{P}-(\Delta_{P})_{\text{GR}}|>\sigma_{\Delta_{P}}~{}. (47)

For a pulsar with rotation period τ\tau and pulse width wτw\lesssim\tau, the timing error σΔP\sigma_{\Delta_{P}} can be roughly estimated as [19]

σΔP1t/1daywtobs/P,\sigma_{\Delta_{P}}\sim\frac{1}{\sqrt{t/1\text{day}}}\frac{w}{t_{\text{obs}}/P}~{}, (48)

where tobst_{obs} is the duration of a continuous observation window every day. In FIG. 5, we plot the timing accuracy for a pulsar of width w=0.01w=0.01s and total observation time TobsT_{\text{obs}}. We see that with enough observational time (e.g.e.g., TobsT_{\text{obs}}\sim 1 decade), much parameter region can be uncovered, even when the cloud is extremely dilute (e.g.e.g., |c1|2106|c_{1}|^{2}\sim 10^{-6}).

Refer to caption
Figure 5: The parameter region reachable for detecting cloud absorption using a pulsar with mass M=1.4MM_{*}=1.4M_{\odot} and width w=0.01w=0.01s, for different cloud occupation fraction |c1|2|c_{1}|^{2}. Here we have taken the state |ψ322|\psi_{322}\rangle as an example. The orbital period is chosen to be near the critical distance, P=0.95P(R,c(322))P=0.95P(R_{*,c}(322)). Different contours represent different total observational time TobsT_{\text{obs}}. The daily observation time is chosen to be tobs=5t_{\text{obs}}=5 hr. In addition to the timing accuracy, requirements on cloud formation and depletion are also included: T(grow)>106T^{(\text{grow})}>10^{6} yr, T(deplete)<108T^{(\text{deplete})}<10^{8} yr [17, 19].

IV.3 Toward relaxing the bound on boson mass

As mentioned before, boson clouds in isolated black holes only deplete by emitting gravitational waves whose frequency is peaked around the boson rest mass μ\mu. Thus the null detection of such near-monochromatic gravitational waves poses constraints on the mass of the ultralight boson, under certain assumptions on the initial black hole spin, cloud ages and the distance from earth. For instance, the LIGO search for continuous gravitational waves from the Milky Way center has reported constraints for the boson mass rage 101310^{-13}eV<μ<1012<\mu<10^{-12}eV [14] (see also [12, 35, 13, 15]). However, it is likely that spinning black holes are not isolated in the galactic center, where the density of objects is high. Therefore, in such a high-density and highly dynamical region, black hole superradiance may be influenced by the presence of a binary companion or other objects, possibly even suppressed to the extent of terminated cloud growth. Henceforth, the gravitational wave null detection bound could be relaxed. We have also seen in Sect. III.2 that states with higher ll are more severely affected by the tidal perturbation of nearby objects. Thus we expect the bound should be state-dependent, with cloud states of higher ll being constrained less.

Another boson mass bound comes from measuring the mass-spin distribution of black holes (the so-called Regge plane). Assuming efficient superradiance growth of the cloud, the black hole angular momentum will quickly be extracted, and its spin parameter is cut off at a value less than 1. This effect manifests itself as a gap on the black hole Regge plane, which can be tested statistically with LIGO and LISA [8, 9, 10]. Considering superradiance termination, such constraints are also subjected to relaxation if the black hole is in a binary. As mentioned in Sect. III.2, if the superradiance of a certain state is shut off by a close companion, the black hole spin cannot further decrease by producing that state. If, in addition, this superradiant state is the one with the lowest ll that the black hole spin is capable of producing, all other states will be shut off. For example, if R<R,c(322)R_{*}<R_{*,c}(322), and a~<4α1+4α2\tilde{a}<\frac{4\alpha}{1+4\alpha^{2}}, both |ψ322|\psi_{322}\rangle and |ψ211|\psi_{211}\rangle cannot grow, and neither can the higher-ll modes (since they have even larger R,cR_{*,c}). Then the black hole spin will not be cut off at the saturation value α1+α2\frac{\alpha}{1+\alpha^{2}} of the |ψ322|\psi_{322}\rangle state. Such considerations must be taken into account in the statistical analysis of the black hole Regge plane.

Other bounds come from, for example, testing the birefringence of photons [36] and the stellar kinematics [37] near supermassive black holes, with the help of powerful telescopes such as EHT [38, 39] and the Keck Observatory [40]. These bounds also assume the presence of cloud, and are subjected to a similar relaxation as mentioned here, since supermassive black holes are known to be surrounded by numerous stars. A detailed analysis of the impact of superradiance termination on various boson mass bounds is beyond the scoop of this paper, but it no doubt deserves further investigations in future works.

V Conclusion

Superradiance instability of ultralight bosons near a rotating black hole leads to lots of interesting phenomena, many of which are based on the existence of a hydrogen-atom-like cloud. In this work, however, we raise the question on the robustness of black hole superradiance in the presence of a binary companion. We have found that through the tidal perturbation of the companion, superradiant states can be coupled to dangerous absorptive states and receive negative corrections to its growth rate, thereby suppressing superradiance. We have found that for a given cloud state, there exists a critical binary distance below which its superradiance is terminated. This fact leads to several important consequences. On one hand, it poses tight constraints on possible GCP transitions. For instance, except hyperfine transitions and Bohr transitions starting from |ψ211|\psi_{211}\rangle, almost all other transitions must have a mass ratio q1q\ll 1. On the other hand, after entering the critical distance, an existing cloud is absorbed back into the black hole. This process can be observed via multiple messengers such as pulsar timing. In addition, the termination of superradiance implies the absence of certain cloud states. This effect is expected to relax the constraints on the ultralight boson mass using various methods that rely on the existence of such clouds.

We have followed a simplistic route in the current work. And there are certainly many improvements to make and prospects to explore in the future. To name a few, first, the analysis of superradiance termination itself can be generalized to more realistic scenarios with generic orbits. Second, we need to understand in detail how superradiance termination can affect the current boson mass bounds. Third, aside from bounded binary orbits considered here, external objects with unbounded orbits may also influence the cloud. It is then interesting to study the stability of the gravitational atom for more complex dynamics such as three-body evolutions and scattering with stellar objects.

Acknowledgments

We thank Chon Man Sou for helpful discussions. XT and YW were supported in part by the National Key R&D Program of China (2021YFC2203100), the NSFC Excellent Young Scientist Scheme (Hong Kong and Macau) Grant No. 12022516, and by the RGC of Hong Kong SAR, China (Project No. CRF C6017-20GF). HYZ was supported in part by a grant from the RGC of the Hong Kong SAR, China (Project No. 16303220).

References