12
Tidal effects and disruption in superradiant clouds: a numerical investigation Vitor Cardoso, Francisco Duque, and Taishi Ikeda CENTRA, Departamento de F´ ısica, Instituto Superior T´ ecnico – IST, Universidade de Lisboa – UL, Avenida Rovisco Pais 1, 1049 Lisboa, Portugal The existence of light, fundamental bosonic fields is an attractive possibility that can be tested via black hole observations. We study the effect of a tidal field – caused by a companion star or black hole – on the evolution of superradiant scalar-field states around spinning black holes. For small tidal fields, the superradiant “cloud” puffs up by transitioning to excited states and acquires a new spatial distribution through transitions to higher multipoles, establishing new equilibrium configurations. For large tidal fields the scalar condensates are disrupted; we determine numerically the critical tidal moments for this to happen and find good agreement with Newtonian estimates. We show that the impact of tides can be relevant for known black-hole systems such as the one at the center of our galaxy or the Cygnus X-1 system. The companion of Cygnus X-1, for example, will disrupt possible scalar structures around the BH for gravitational couplings as large as 2 × 10 -3 . I. INTRODUCTION The matter content of our universe is largely unknown, and has been the focus of an incredible effort in the last decades [1]. Most experimental searches are based on putative couplings between dark matter (DM) and stan- dard model fields. The null results of such experiments provide interesting constraints on the strength of such couplings, but are otherwise unable to shed light on new fundamental constituents of the universe. The universal nature of gravity suggests that new fields or particles behave in the same way as their standard model cousins when placed in gravitational fields. It is no surprise therefore that the only but solid evidence for DM interaction is so far of a purely gravitational nature. The advent of gravitational-wave (GW) astronomy provides a compelling case to understand further the behavior of DM in strong gravity situations [2–4]. Of particular rele- vance in this context are black hole (BH) spacetimes. In vacuum general relativity these are the simplest macro- scopic object one can conceive of, and ideal to be used as testing grounds for the presence of new fields or exten- sions of general relativity [1–6]. The DM density in our universe is measured to be small enough that its effects on the dynamics of compact objects – BHs in particular – are perturbatively small. The imprint that DM leaves on the GW is correspond- ingly small, but potentially measurable by future GW detectors [7–10]. However, should ultralight bosonic de- grees of freedom exist in nature [11, 12], superradiance will give rise to the development of massive structures (“clouds”) around spinning astrophysical BHs. This is a general mechanism that requires only minimal ingredi- ents. A simple minimally coupled massive field with no initial abundance suffices: a superradiant instability sets in, extracting rotational energy away from the BH and depositing it in any small boson fluctuation outside the BH. For the mechanism to be effective, the BH radius 2GM/c 2 needs to be of the order of the boson Comp- ton wavelength G/(c 2 μ) for a particle of mass m B (here μ = Gm B /(c~) is the mass parameter that will appear in all our equations). In other words, the mechanism is effective when 1. However, because BHs in our universe appear in a wide range of masses – that vary over 8 or more orders of magnitude, superradiance allows to effectively study or rule out boson masses varying by correspondingly large orders of magnitude [6, 13, 14]. The existence of superradiant clouds would lead to ob- servable signatures, such as peculiar holes in the mass- spin plane of BHs [13, 14], to monochromatic emission of GWs [14, 15] and to a significant stochastic background of GWs [16, 17]. The presence of such periodic, non- axisymmetric structures can leave imprints in planetary and stellar orbits, through Lindblad and co-rotation res- onances [18, 19], or – of interest for GW astronomy – through floating or sinking orbits [20–23]. There has been a significant progress in our understanding of the development of the superradiant instabilities [6]. There are two main factors that could alter, in a significant way, the formation of heavy boson clouds around BHs. In the presence of couplings between the ultralight boson and standard model fields, for example, the cloud growth can be suppressed, while stimulating bursts of light [24, 25]. Here, we focus instead on the effects that a com- panion star or BH have on the structure of the boson cloud. Tidal effects were studied recently, at an analyti- cal level, and using Newtonian dynamics for non relativis- tic fields [21–23, 26–28]. The motion of the binary can, at specific orbital frequencies, induce resonant transitions between growing and decaying modes of the boson, that enhance the cloud’s depletion and/or transfer energy and angular momentum to the companion through tidal ac- celeration [29]. This behaviour would leave distinctive imprints in the GW signal emitted by the binary, both as a monochromatic signal from the cloud or as modifica- tions in the GW waveform of the binary, due to finite-size effects (eg. variations on the spin-induced quadrupole or the tidal Love numbers) [27]. This situation could be of interest for eccentric BH binaries targeted by the space- interferometer LISA [28]. arXiv:2001.01729v3 [gr-qc] 21 Mar 2020

CENTRA, Departamento de F sica, Instituto Superior T ...Universidade de Lisboa { UL, Avenida Rovisco Pais 1, 1049 Lisboa, Portugal The existence of light, fundamental bosonic elds

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: CENTRA, Departamento de F sica, Instituto Superior T ...Universidade de Lisboa { UL, Avenida Rovisco Pais 1, 1049 Lisboa, Portugal The existence of light, fundamental bosonic elds

Tidal effects and disruption in superradiant clouds: a numerical investigation

Vitor Cardoso, Francisco Duque, and Taishi IkedaCENTRA, Departamento de Fısica, Instituto Superior Tecnico – IST,

Universidade de Lisboa – UL, Avenida Rovisco Pais 1, 1049 Lisboa, Portugal

The existence of light, fundamental bosonic fields is an attractive possibility that can be tested viablack hole observations. We study the effect of a tidal field – caused by a companion star or blackhole – on the evolution of superradiant scalar-field states around spinning black holes. For small tidalfields, the superradiant “cloud” puffs up by transitioning to excited states and acquires a new spatialdistribution through transitions to higher multipoles, establishing new equilibrium configurations.For large tidal fields the scalar condensates are disrupted; we determine numerically the criticaltidal moments for this to happen and find good agreement with Newtonian estimates. We showthat the impact of tides can be relevant for known black-hole systems such as the one at the centerof our galaxy or the Cygnus X-1 system. The companion of Cygnus X-1, for example, will disruptpossible scalar structures around the BH for gravitational couplings as large as Mµ ∼ 2 × 10−3.

I. INTRODUCTION

The matter content of our universe is largely unknown,and has been the focus of an incredible effort in the lastdecades [1]. Most experimental searches are based onputative couplings between dark matter (DM) and stan-dard model fields. The null results of such experimentsprovide interesting constraints on the strength of suchcouplings, but are otherwise unable to shed light on newfundamental constituents of the universe.

The universal nature of gravity suggests that new fieldsor particles behave in the same way as their standardmodel cousins when placed in gravitational fields. It is nosurprise therefore that the only but solid evidence for DMinteraction is so far of a purely gravitational nature. Theadvent of gravitational-wave (GW) astronomy providesa compelling case to understand further the behavior ofDM in strong gravity situations [2–4]. Of particular rele-vance in this context are black hole (BH) spacetimes. Invacuum general relativity these are the simplest macro-scopic object one can conceive of, and ideal to be used astesting grounds for the presence of new fields or exten-sions of general relativity [1–6].

The DM density in our universe is measured to besmall enough that its effects on the dynamics of compactobjects – BHs in particular – are perturbatively small.The imprint that DM leaves on the GW is correspond-ingly small, but potentially measurable by future GWdetectors [7–10]. However, should ultralight bosonic de-grees of freedom exist in nature [11, 12], superradiancewill give rise to the development of massive structures(“clouds”) around spinning astrophysical BHs. This isa general mechanism that requires only minimal ingredi-ents. A simple minimally coupled massive field with noinitial abundance suffices: a superradiant instability setsin, extracting rotational energy away from the BH anddepositing it in any small boson fluctuation outside theBH. For the mechanism to be effective, the BH radius2GM/c2 needs to be of the order of the boson Comp-ton wavelength G/(c2µ) for a particle of mass mB (hereµ = GmB/(c~) is the mass parameter that will appear

in all our equations). In other words, the mechanism iseffective when Mµ ∼ 1. However, because BHs in ouruniverse appear in a wide range of masses – that varyover 8 or more orders of magnitude, superradiance allowsto effectively study or rule out boson masses varying bycorrespondingly large orders of magnitude [6, 13, 14].

The existence of superradiant clouds would lead to ob-servable signatures, such as peculiar holes in the mass-spin plane of BHs [13, 14], to monochromatic emission ofGWs [14, 15] and to a significant stochastic backgroundof GWs [16, 17]. The presence of such periodic, non-axisymmetric structures can leave imprints in planetaryand stellar orbits, through Lindblad and co-rotation res-onances [18, 19], or – of interest for GW astronomy –through floating or sinking orbits [20–23]. There hasbeen a significant progress in our understanding of thedevelopment of the superradiant instabilities [6]. Thereare two main factors that could alter, in a significant way,the formation of heavy boson clouds around BHs. In thepresence of couplings between the ultralight boson andstandard model fields, for example, the cloud growth canbe suppressed, while stimulating bursts of light [24, 25].

Here, we focus instead on the effects that a com-panion star or BH have on the structure of the bosoncloud. Tidal effects were studied recently, at an analyti-cal level, and using Newtonian dynamics for non relativis-tic fields [21–23, 26–28]. The motion of the binary can,at specific orbital frequencies, induce resonant transitionsbetween growing and decaying modes of the boson, thatenhance the cloud’s depletion and/or transfer energy andangular momentum to the companion through tidal ac-celeration [29]. This behaviour would leave distinctiveimprints in the GW signal emitted by the binary, bothas a monochromatic signal from the cloud or as modifica-tions in the GW waveform of the binary, due to finite-sizeeffects (eg. variations on the spin-induced quadrupole orthe tidal Love numbers) [27]. This situation could be ofinterest for eccentric BH binaries targeted by the space-interferometer LISA [28].

arX

iv:2

001.

0172

9v3

[gr

-qc]

21

Mar

202

0

Page 2: CENTRA, Departamento de F sica, Instituto Superior T ...Universidade de Lisboa { UL, Avenida Rovisco Pais 1, 1049 Lisboa, Portugal The existence of light, fundamental bosonic elds

2

II. SETUP

Our starting point is that of a Kerr BH spacetime per-turbed by a distant companion. The BH is surroundedby a superradiant cloud, which is assumed to cause neg-ligible backreaction in the spacetime. The geometry willalways be kept fixed in this work, in the sense that thescalar field never backreacts back. This working hypoth-esis holds true for most of the situations of interest [14],and is specially appropriate here: as we explain belowthe timescales that we can probe are much shorter thanany superradiant-growth timescales. A companion ofmass Mc is now present, at a distance R, and locatedat θ = θc, φ = φc in the BH sky. The companion inducesa change δds2tidal in the geometry. Thus, our spacetimegeometry is described by

ds2 = ds2Kerr + δds2tidal . (1)

For the tidal perturbation induced by the companion, weconsider the non-spinning approximation, and we findin Regge-Wheeler gauge that the dominant quadrupoletidal contribution is [30–32]

δds2 =∑m

r2E2mY2m(θ, φ)(f2dt2 + dr2 + (r2 − 2M2)dΩ2)

E2m =8πε

5M2Y ∗2m(θc, φc) , (2)

where f = 1 − 2M/r and we neglect subdominantmagnetic-type contributions and multipoles higher thanthe quadrupole. For more details see Appendix A. Weintroduce a dimensionless tidal parameter

ε =McM

2

R3, (3)

which measures the strength of the tidal moment. Co-ordinates are Boyer-Lindquist at large distance. Thisapproximation is not accurate close to the BH horizon,where spin effects change the tidal description. However,for all the parameters considered here the cloud is local-ized sufficiently far-away that these effects ought to bevery small. We will focus exclusively on static tides (or inother words, we consider large separations R). We stressthat we are using coordinates adapted to the BH: thecompanion position should in general be time-dependent,but we focus exclusively on slowly moving companions.

We consider a massive, minimally coupled scalar fieldΦ evolving on the above fixed geometry. The scalar isdescribed by the Klein-Gordon equation(

∇µ∇µ − µ2)

Φ = 0 . (4)

Note that, for zero rotation, distances r M and non-relativistic fields, the Klein-Gordon equation can be ex-pressed as in Eq. (3.4) of Ref. [27], amenable to a pertur-bation treatment. Some of the implications are summa-rized in the Appendix. However, we consider the KleinGordon in full generality, by evolving it numerically. To

express Eq. (4) as a Cauchy problem, we use the standard3+1 decomposition of the metric,

ds2 = −α2dt2 + γij(dxi + βidt)(dxj + βjdt) , (5)

where α is the lapse function, βi is a shift vector, andγij is the 3-metric on spacial hypersurface. We also in-troduce the scalar momentum Π,

Π = −nµ∇µΦ . (6)

The evolution equation for the axion field is written as

∂tΦ = −αΠ + LβΦ ,

∂tΠ = α(−D2Φ + µ2Φ +KΠ)−DiαDiΦ + LβΠ .

We use Cartesian Kerr-Schild coordinates (t, x, y, z) [33].The quantities extracted from our numerical simula-

tion are multipolar components of the scalar Φ,

Φl,m(t, r) =

∫dΩΦ(t, r, θ, φ)Yl,m(θ, φ) . (7)

We use as initial data the following profile, adequate todescribing quasi-stationary states around a BH [14, 34]

Φ(t, r, θ, φ) = A0rMµ2e−rMµ2/2 cos(φ− ωRt) sin θ . (8)

We will also show below that these are indeed good de-scription of stationary states for small couplings Mµ .0.2. Here A0 is an arbitrary amplitude related to themass in the axion cloud, and ωR ∼ µ is the bound-statefrequency.

The spacetime of a real astrophysical binary is asymp-totically flat. However, because we are using only anapproximation to the full problem, where the companionis supposed to be a large distance away, the geometry (2)is no longer asymptotically flat. To avoid unphysical be-havior at large distances, we force the geometry to beasymptotically flat, by replacing the far region with

ds2 = ds2Kerr + (1−W) δds2tidal , (9)

where W =W(r) is a following piecewise function

W(r) =

1 (r > 1)

W5 (0 < r < 1)

0 (r < 0).

(10)

Here, r = (r − rth)/w and W5(r) is chosen to matchsmoothly with the required asymptotic behavior, sowe choose a 5th-order polynomial satisfying W5(1) =1,W5(0) = W ′5(0) = W ′′5 (0) = W ′5(1) = W ′′5 (1) = 0.The transition region has a width w = 500M and is lo-cated at rth/M '

√0.9× 5/(8πε). These parameters

were chosen to ensure that the bosonic cloud sits entirelyin a region described by Eq. (2).

The evolution equations were integrated using fourth-order spatial discretization and a Runge-Kutta method.Accuracy requirements, finite size of the numerical gridand computational power all contribute to limit the

Page 3: CENTRA, Departamento de F sica, Instituto Superior T ...Universidade de Lisboa { UL, Avenida Rovisco Pais 1, 1049 Lisboa, Portugal The existence of light, fundamental bosonic elds

3

timescales that one is able to access. Here, we evolvethese systems for timescales ∼ 7000M .

Although we have results for general BH spin param-eter, we focus mostly on states around a non-spinningBH. These states are not superradiant in origin, andarise due to the fine-tuned initial data. However, theyare extremely long-lived (the decay timescale is of theorder of the superradiant growth timescale if the BH wasspinning), as we show below, with a lifetime that far ex-ceeds that of all the tidally-induced transitions studiedhere. Thus, BH spin is important to generate the scalarclouds, but has little impact on some of the physics oftides. In addition, the tidal field in Eq. (2) is adapted toa non-spinning BH. Our numerical results indeed showonly a very mild dependence on BH spin. With the ex-ception of Ref. [28], all previous results on tidal effectsin superradiant clouds focus on the small Mµ couplingparameter, consider a flat background on which the su-perradiant states evolve, and have only used linearizedanalysis for small tidal fields. Our framework can go be-yond all these limitations.

0 1000 2000 3000 4000 5000 6000 7000t/M

0.8

0.6

0.4

0.2

0.0

0.2

0.4

0.6

0.8

1,1

FIG. 1. A dipolar scalar cloud around a Schwarzschild BH.This figure shows the time evolution of initial conditions (8)for a dipole with gravitational coupling Mµ = 0.1 around anon-spinning BH, and in absence of a companion (ε = 0). Thefield is extracted at r = 60M .

III. RESULTS

A. Weak tides: transitions to new stationary states

We start by evolving the initial data described abovearound an isolated, non-spinning BH (ε = 0). The non-vanishing multipolar component of the field is shown inFig. 1. The amplitude of the field varies by a few percentover the time interval of ∼ 7000M . This time interval(7000 dynamical timescales) also corresponds to ∼ 100scalar field periods of oscillation. The scalar field and en-ergy density along an equatorial slice are shown in Fig. 2

FIG. 2. Field (left) and energy density (right) distributionalong the equatorial plane for the same initial data as Fig. 1.The field is dipolar, as expected, whereas the energy densityat the equator is almost – but not exactly – symmetric alongthe rotation axis. The lengthscale of these images is of order100M .

2000 1000 0 1000 2000x/M

0.8

0.6

0.4

0.2

0.0

0.2

0.4

0.6

0.8= 0, t/M = 0= 0, t/M = 2000= 10 8, t/M = 2000

FIG. 3. Dependence of the field Φ along the x− axis at differ-ent instants for a coupling Mµ = 0.1. In the absence of a com-panion (ε = 0) and despite a slight change in the profile, thefield has no nodes. It has a local extremum at ∼ r = 200M aspredicted by a small Mµ expansion for the fundamental mode(in the convention adopted in this paper n = l+1 = 2). In thepresence of a weak tidal field (ε = 10−8) the field develops adifferent radial profile with one node, pointing to a significantcomponent of overtones. Our results indicate a sizeable exci-tation of the second excited state n = 4, which has extremaat r/M ∼ 170, 850.

at t = 7000M . The density is almost (but not exactly)symmetric along this slice.

We now turn on a weak tidal field by letting ε = 10−8,produced by a companion star on the x-axis. We callthis a weak tide since no nonperturbative feature is seenon timescales of ∼ 6000M . As we will argue below, itis possible that new features appear at very late times,which we are unable to probe currently. Previous, ana-lytical studies focused on transitions between the over-tones [27, 28]. We do see transitions between overtoneswith the same angular index, leading to an expansion ofthe cloud: overtones are localized at rBohr ∼ n2/(Mµ2).

Page 4: CENTRA, Departamento de F sica, Instituto Superior T ...Universidade de Lisboa { UL, Avenida Rovisco Pais 1, 1049 Lisboa, Portugal The existence of light, fundamental bosonic elds

4

The appearance of higher overtones is apparent in Fig. 3,showing the x− dependence of the field initially and att = 2000M . It is clear than the companion triggers exci-tation of overtones, which manifest themselves via nodesin the field. Although this profile also includes the oc-tupolar l = 3 component, it is two orders of magnitudesmaller than the dipolar term, as we discuss below, andunable to explain all the structure in Fig. 3.

The data in Figure 3 indicates that on timescales. 2000M the second excited state n = 4 (in our conven-tion, states are labeled by an integer n = l + 1, l + 2, ...)dominates the transitions. In fact, a small coupling Mµexpansion (see Appendix B) shows that the n = 3 statehas extrema at r/M = 175, 1024 which are not appar-ent in Fig. 3. The second excited state n = 4 however,is predicted to have extrema at r/M = 170, 875, 2155 inagreement with our numerical results (but note that thelast point is challenging to confirm numerically, as thegrid size and spurious reflections affect a proper evalua-tion of eigenfunctions at large distances).

One can quantify the relative excitation using the or-thonormality between eigenstates corresponding to dif-ferent overtones, which enable us to extract the ampli-tude cn of a specific overtone from the numerical datavia

cn =

∫ ∞0

dr r2R∗n1 (r) Φ (r) , (11)

where Rnl (r) are the radial “hydrogenic” functions dis-cussed in the Appendix B and Φ (r) corresponds to thenumerical data at a given radial direction (e.g. θ = π/2and ϕ = 0). We are implicitly taking the numericaldata to be only composed by l = 1 modes, which is areasonable assumption considering our previous discus-sion. This expression is actually only valid in the non-relativistic (and far-region) limit, but we expect it to pro-vide reasonable estimates at small Mµ when the cloudis localized far away from the BH. As we explained, ourgrid size is limited and for these parameters (Mµ = 0.1),it captures a couple of nodes but not more. Thus, therelative amplitude excitation determined in this way isaffected by some numerical error, and is expected to bemore accurate at early times when the signal is dom-inated by the fundamental mode. To avoid reflectionsfrom the outer boundary of our numerical grid and tran-sition radius of our metric, we use time-domain data fort . 1000M .

Our results are shown in Table I, and include estimatesfor the transition timescale from time-dependent pertur-bation theory. Our numerical results are within a factortwo from the prediction from perturbation theory. Thisdisagreement can be explained by (at least) two factors:i. perturbation theory uses a small Mµ expansion whichis inaccurate at Mµ & 0.1 (a glance at Fig. 1 in Ref. [28]shows how factors of two can easily arise from such an ap-proximation); ii. transitions between intermediate statescomplicate substantially the calculation of mode excita-tion. Bearing this in mind, our results along with pertur-bation theory explain why the first excited state is not yet

TABLE I. Timescales ttrans and relative amplitudes predictedby time-independent perturbation theory and those obtainedfrom numerical data (at t = 1000M), for the most relevant1st order transitions from the initial state l = 1 state, withMµ = 0.1. The second column shows the timescale to tran-sition from the initial to the (nlm) state, as obtained fromtime-dependent perturbation theory. The third column showsthe relative amplitude of overtones, relative to the fundamen-tal mode, from time-independent perturbation theory (and inparenthesis the corresponding ratio of the field componentsat r = 60M). Finally, the last column shows the relativeamplitude of overtones as obtained from our numerical data.The entries in the third and fourth column agree to within afactor two, with the exception of the l = m = 1, n = 3 mode,for which the timescale needed for excitation is larger thanthe instant at which the coefficients were extracted.

(n lm) ttrans/Mcnlmc211

(φnlmφ211

)cNumn

cNum2

3 1 1 1888 1.03 (0.85) 0.221

4 1 1 458 0.236 (0.13) 0.094

5 1 1 173 0.113 (0.046) 0.058

dominant: the timescale for its excitation is the largestamong those in the table. In fact, our results are consis-tent with transition occurring on the timescales predictedfrom the table for the n = 3, 4, 5 modes.

The most apparent feature of our simulations, however,are transitions to octupolar and higher poles, inducedby the external tide. This is depicted in Fig. 4, wherewe show the evolution of the dipolar l = m = 1 andoctupolar l = m = 3 mode as time progresses. It isapparent that the magnitude of the dipolar mode is nowdecreasing, and that a fraction of this energy is going intohigher modes, specifically the octupolar l = 3,m = 1, 3.Such migration changes the spatial distribution of energydensity, apparent in Fig. 5.

Our results in the small external tide regime are con-sistent with perturbation theory prediction, in particularthat the amplitude of the l = 3 mode scales with theexternal tide ε [27, 28]. Perhaps one of the cleanest in-dications of the validity of the perturbative frameworkis the excitation of the l = 3,m = 1 mode. Perturba-tion theory predicts that the relative amplitude of thel = m = 3 mode is

√5/3 ∼ 1.29 larger than that of the

l = 3,m = 1 mode, and this depends exclusively on anangular matrix (no radial dependence). Our numericalresults show a relative amplitude across all times and ex-traction radii consistent with such prediction, as shownin the figure. The l = 3 mode seems to saturate, but ascan be seen from the figure, the l = 1 is still decreasing.Some energy is most likely flowing down the horizon, butwe do not fully understand this behavior.

B. Strong tides: tidal disruption of clouds

For large tidal fields, one expects the scalar configura-tion to be disrupted. A star of mass M∗, radius R∗, in

Page 5: CENTRA, Departamento de F sica, Instituto Superior T ...Universidade de Lisboa { UL, Avenida Rovisco Pais 1, 1049 Lisboa, Portugal The existence of light, fundamental bosonic elds

5

0 1000 2000 3000 4000 5000 6000 7000t/M

0.8

0.6

0.4

0.2

0.0

0.2

0.4

0.6

0.8

1,1

0 1000 2000 3000 4000 5000 6000t/M

0.0015

0.0010

0.0005

0.0000

0.0005

0.0010

0.0015

l,m

3, 3

3, 1 * 1.29

FIG. 4. Dipolar (l = m = 1, left) and octupolar (l = 3, m = 1, 3 right) component of the scalar cloud when in the presence ofa weak tidal field, for the same initial conditions as in Fig. 1 (non-spinning BH and gravitational coupling Mµ = 0.1), but nowin the presence of a companion of mass ε = 10−8. The l = 3,m = 1 mode amplitude relative to the l = m = 3 was re-scaled bythe perturbation theory prediction (

√5/3 ∼ 1.29). The agreement is very good throughout the evolution.

FIG. 5. Snapshot of a tidally deformed scalar cloud. Thesnapshot depicts the energy density along the equator of ascalar cloud which was set initially around a non-spinningBH. In the absence of a companion mass, the energy densityis almost spherical and remains so for thousands of dynamicaltimescales. Here, the simulation starts with one symmetricinitial scalar energy distribution, but in the presence of a star,such that the tidal parameter ε = 10−8. The gravitationalcoupling Mµ = 0.1. The snapshot is taken after 7000M bywhich the system settled to a new stationary configuration.

the presence of a companion of mass Mc at distance R ison the verge of disruption if – up to numerical factors oforder unity,

M∗/R2∗ = 2McR∗/R

3 . (12)

For configurations where the mass in the scalar cloud isa fraction of that of the BH, M∗ = M and its radius isof the order of R∗ & 5/(Mµ2) (see Appendix). As such,we find the critical moment

εcrit ≈(Mµ)6

250. (13)

Our simulations are consistent with this behavior. Fig-ures 6–7 summarize our findings at large companionmasses. We find that the initial dipolar mode quicklytransfers energy to the octupole, which then drains intohigher and higher multipoles. This signals a transfer ofenergy to lower and lower angular scales, and in our caseit is telling us that the cloud is being disrupted, loosingmass to asymptotic regions. A snapshot of the energydensity in Fig. 7 shows precisely this.

Extracting precise values for the critical tide from ournumerical simulations is complicated by two facts: theseare extended scalar configurations, and to understandwhether there is mass being lost to large distances re-quires large numerical grids. In addition, a seemingly sta-ble cloud on some timescale can eventually be disruptedwhen evolved on longer timescales. Typically, our numer-ical simulations last for ∼ 6000M . With this in mind, weestimate a threshold εcrit ∼ 2 × 10−7 for Mµ = 0.2, forwhich disruption is clearly seen after 4000M . This crit-ical value agrees remarkably well with Eq. (13). On theother hand, for Mµ = 0.1 we see disruption on the sim-ulated timescales only for ε & 2 × 10−8, a factor fourdifference from the prediction of Eq. (13). It is possi-ble that disruption does happen for smaller tides, but ontimescales that we are currently unable to probe. Notealso that disruption can be stimulated by transitions toovertones, an intermediate process which occurs as we’vejust discussed, and which “puffs up” the cloud, increasingits size to a few times the estimate 1/(Mµ2) and therefore

Page 6: CENTRA, Departamento de F sica, Instituto Superior T ...Universidade de Lisboa { UL, Avenida Rovisco Pais 1, 1049 Lisboa, Portugal The existence of light, fundamental bosonic elds

6

0 1000 2000 3000 4000 5000 6000 7000t/M

0.6

0.4

0.2

0.0

0.2

0.4

0.61,

1

0 1000 2000 3000 4000 5000 6000 7000t/M

0.0075

0.0050

0.0025

0.0000

0.0025

0.0050

0.0075

3,3

FIG. 6. Tidally disrupting cloud, and cascading to lower scales. This figure shows the time evolution of the dipolar andoctupolar components of the scalar field (with gravitational coupling Mµ = 0.1) for a companion with ε = 10−7 such that thecloud is disrupted. We observe that the cloud torn apart, loosing energy to asymptotically large distances, away from the BH.Thus, energy is cascading to higher and higher multipoles as time progresses. The field is extracted at r = 60M .

FIG. 7. Snapshot of a tidally disrupting cloud. The snap-shot depicts the energy density along the equator of a scalarcloud which was set initially around a non-spinning BH. Inthe absence of a companion mass, the energy density is al-most spherical and remains so for thousands of dynamicaltimescales. Here, the simulation starts with one symmetricinitial scalar energy distribution, but in the presence of a starfor which ε = 10−7. The gravitational coupling Mµ = 0.1.The snapshot is taken after 7000M and is leading to disrup-tion of the cloud.

reducing the critical tide. However, such transitions canoccur on large timescales. To summarize, our results areconsistent with the behavior of Eq. (13), though clearlyevolutions lasting for one order of magnitude more could

zoom in better on the prefactor.

IV. APPLICATION TO ASTROPHYSICALSYSTEMS

Known BHs with companions include the Cygnus X-1 system and the center of our galaxy. Cygnus X-1 isa binary system composed of a BH of mass MBH ∼15M, a companion with Mc ∼ 20M at a distanceR ∼ 0.2 AU ∼ 3 × 1010 m [35]. With these parameters,we find ε ∼ 5 × 10−19. For it to sit at the critical tide,Mµ ∼ 2×10−3. The timescale τ for growth of clouds viasuperradiance is of order τ ∼ (Mµ)−9M [6], too large tobe meaningful for scalar fields, but potentially affectingvectors (τ ∼ (Mµ)−7M) [33, 36, 37]. The tide is alsosmall enough that it should not be affecting any of theconstraints derived from the possible non-observation ofGWs from the system [38, 39].

On the other hand, at the center of our galaxy thereis a supermassive BH of mass ∼ 4× 106M with knowncompanions [40, 41]. For the closest known star, S2, witha pericenter distance of∼ 1400MBH we find ε ∼ 2×10−15,or a critical coupling Mµ ∼ 9 × 10−3 (we assume Mc ∼20M but the result above is only mildly dependent onthe unknown mass of S2). This is now a potential sourceof tidal disruption for interesting coupling parameters,and will certainly affect the estimates using pure dipo-lar modes to estimate GW emission. However, note thatour approximations always require that the companionsits outside the cloud (R > R∗, or the approximationin Eq. (2) would break down). Using Eq. (12), dis-ruption together with such a condition always requiresMc > M/2.

Note that, at the verge of tidal disruption by a com-

Page 7: CENTRA, Departamento de F sica, Instituto Superior T ...Universidade de Lisboa { UL, Avenida Rovisco Pais 1, 1049 Lisboa, Portugal The existence of light, fundamental bosonic elds

7

panion, the binary itself is emitting GWs at a rate

Ebinary =32

5

M2cM

3

R5, (14)

where we assume the companion to be much lighter thanthe BH. The GW flux emitted by the cloud-BH systemscales as [14, 16, 34]

Ecloud ∼1

50

(MS

M

)2

(Mµ)14. (15)

Thus, GW emission by the binary dominates the signalwhenever

Mc

MS& (MS/M)5(5Mµ/2)12 , (16)

with MS the mass in the scalar cloud [14, 16, 34]. There-fore, in the context of GW emission and detection, for allpractical purposes, disruption will not affect our abilityto probe the system: if it was visible via monochromaticemission by the cloud before disruption, it will be seenafter disruption as a binary.

Note that tidal disruption of the cloud is a relevantpossibility for these systems, since the cloud is genericallynot depleted due to mode mixing by the time the systemreaches the Roche radius; In fact, for cloud depletiondue to mode mixing to be effective, the system needsto be in a resonant epoch for a long time [27]. Thisrequires a particular combination of the mass ratio andgravitational coupling Mµ which can only be realized ina small region in the possible parameter space (see Figs. 7and 8 in Ref. [27]).

V. CONCLUSIONS

Massive, spinning BHs provide us with the tantalizingpossibility to test fundamental fields on scales which areotherwise inacessible. These fields can be all or a fractionof DM and may or not couple to standard model fields. Inother words, BHs are ideal detectors of ultralight fields [6,13, 23].

The mechanism behind this extraordinary ability is su-perradiance, which works very much like tidal accelera-tion in the Earth-moon system [6, 29]. In the presence ofa light field, a spinning BH may transfer a large fractionof its rotational energy to a “cloud” of bosons orbiting theBH. Such effect leads to BH spindown, emission of nearlymonochromatic radiation, etc. We started here a numer-ical study of the impact of a possible companion star orBH on the development of such superradiant cloud. Wesee transitions to higher overtones and to higher multi-poles, stretching and deforming the cloud. Weak tidalfields (i.e., light or far-away companions) slightly deformthe cloud, affecting GW emission by the system. Thechanges induced by tidal fields have not been computedyet. For tidal fields larger than the threshold of Eq. (13),the companion simply breaks the cloud apart. Since such

structures are usually much larger than the BH – asshown by Eq. (B7) – superradiant clouds are typicallyeasier to disrupt than stars. In fact, BH systems such asthe one at the center of our galaxy or the Cygnus X-1binary system may easily disrupt scalar clouds.

Our results generalize to a number of situations. Al-though we have discussed only tides acting along theequator, we have performed evolutions for polar tides(along the z−axis), and found the same phenomenol-ogy. This includes overtone and transitions between mul-tipoles and tidal disruption, even if quantitatively differ-ent. Our setup is that of a real scalar field, but the resultsgeneralize to complex scalars. These are interesting froma BH uniqueness perspective since they can lead to trulystationary hairy solutions (as opposed to real fields whichlead to long-lived states which are not of the Kerr family,but which eventually must decay to Kerr) [6, 42].

Simulations of binaries are challenging. To overcomeissues with long-term simulations of two bodies, we re-place the companion with its lowest order tidal moment.Such approximation has problems of its own, and requirescareful handling of boundary conditions, grid sizes etc.Currently, we are unable to probe tidal fields which varyon short timescales: these would lead to “superluminal”motion on the outerpart of our computational domain; assuch, we are unable to probe resonances arising from tidaleffects. Resonances are interesting on their own [23, 27]and might lead to floating or sinking orbits which lead toclear imprints in GW signals [18, 20, 21, 23].

ACKNOWLEDGEMENTS

We are grateful to Hirotaka Yoshino for useful com-ments and suggestions, and to Miguel Zilhao for usefulsuggestions and advice on the numerical simulations. V.C. was partially funded by the Van der Waals Professo-rial Chair. V. C. would like to thank Waseda Universityfor warm hospitality and support while this work was fi-nalized. V.C. acknowledges financial support providedunder the European Union’s H2020 ERC ConsolidatorGrant “Matter and strong-field gravity: New frontiersin Einstein’s theory” grant agreement no. MaGRaTh–646597. F.D. acknowledges financial support provided byFCT/Portugal through grant SFRH/BD/143657/2019.This project has received funding from the EuropeanUnion’s Horizon 2020 research and innovation pro-gramme under the Marie Sklodowska-Curie grant agree-ment No 690904. We thank FCT for financial sup-port through Project No. UIDB/00099/2020. We ac-knowledge financial support provided by FCT/Portugalthrough grant PTDC/MAT-APL/30043/2017. The au-thors would like to acknowledge networking support bythe GWverse COST Action CA16104, “Black holes, grav-itational waves and fundamental physics.”

Page 8: CENTRA, Departamento de F sica, Instituto Superior T ...Universidade de Lisboa { UL, Avenida Rovisco Pais 1, 1049 Lisboa, Portugal The existence of light, fundamental bosonic elds

8

Appendix A: Tides in General Relativity

The general theory of tidally deformed compact ob-jects in General Relativity lays its foundations on linearperturbations around a background spacetime describingthe compact object [30, 43, 44]

gµν = g(0)µν + hµν , (A1)

where g(0)µν is the background spacetime metric and hµν

is a small perturbation. For a body perturbed by anexternal field, we expect hµν to encapsule the direct con-tribution of that external field and the corresponding lin-ear response of the perturbed object due to gravitationalinteraction.

The standard strategy to compute hµν is to pick aspecific gauge and solve the linearized field equations fora chosen background. The most practical situation iswhen the background is spherically-symmetric and static,in which case the line element reads

ds2 = −F (r) dt2+G (r) dr2+r2dθ2+r2 sin2 θdϕ2 . (A2)

In this scenario, the perturbation hµν is expanded inspherical harmonics

Y lm(θ, ϕ) ≡

√2l + 1

(l −m)!

(l +m)!Pml (cos θ) eimϕ . (A3)

and due to axisymmetry, decomposed in even and oddparts. In the Regge-Wheeler gauge, these read

hevenµν =

F (r) H lm

0 (t, r) Y lm H lm1 (t, r) Y lm 0 0

H lm1 (t, r) Y lm G (r) H lm

2 (t, r) Y lm 0 0

0 0 r2Klm (t, r) Y lm 0

0 0 0 r2 sin2 θKlm (t, r) Y lm

, (A4)

hoddµν =

0 0 hlm0 (t, r) Slmθ hlm0 (t, r) Slmϕ0 0 hlm1 (t, r) Slmθ hlm1 (t, r) Slmϕ

hlm0 (t, r) Slmθ hlm1 (t, r) Slmθ 0 0

hlm0 (t, r) Slmϕ hlm1 (t, r) Slmϕ 0 0

, (A5)

where (Slmθ , Slmϕ

)≡(−Y lm,ϕ / sin θ, sin θ Y lm,θ

). (A6)

The aforementioned separation of hµν into the exter-nal field and respective tidal response can be made ex-plicit by means of an asymptotic expansion in multi-pole moments (check Eqs. (B9) and (B10) of Ref. [44]).The even-parity sector is controled by the polar tidalmoments EL, where the subscript L ≡ a1a2 . . . al isa multi-index labeling the (2l + 1) components of thissymmetric-trace-free tensor. One can further decom-posed these in spherical-harmonics through EL (t)xL =rl∑m Elm (t)Y lm (θ, ϕ). The same procedure applies to

the axial sector, but here they are controlled by the axialtidal moments BL, which follow the same decomposition,BL (t)xL =

∑m r

lBlm (t)Y lm (θ, ϕ).

For a specific tidal field, the tidal moments can be de-termined by performing an asymptotic matching with apost-Newtonian environment, in a domain much largerthan the typical lenghtscale of the deformed compact ob-ject [32, 45]. We are interested in a binary system whereone of bodies is a BH. Centering ourselves in it, and treat-ing the corresponding companion as a post-Newtonianmonopole of mass Mc at distance R, the lowest contribu-tion to the tidal field is given by the l = 2 quadrupolar

moments [32]

Eab = −3Mc

R3n〈ab〉 +O

(c−2), (A7)

Bab = −6Mc

R3[n× v](a nb) +O

(c−2), (A8)

where R is the position of the companion in the BHframe centered, n ≡ R/R, the brackets 〈. . . 〉 denote sym-metrization and trace removal, and v is the relative ve-locity of the binary. We stress that the two terms dis-played are not of the same PN order. While the firstterm in Eab corresponds to the Newtonian limit and theommited term is a 1PN correction, the lowest order termin Bab is already considered to be a 1PN contribution,despite not being directly surpressed by c−2 (see sectionII of Ref. [32]). Higher multipole moments are also sub-leading with respect to above quadrupoles.

Finally, a further simplification is introduced by as-suming that hµν is independent of time. This is theso called regime of static tides, when the binary evolu-tion happens in much larger timescales than the internaldynamics of each body, and the whole system evolvesadiabatically. This is what happens in the inspirallingphase of the binary. The corresponding perturbationsfor a Schwarzschild background are those presented inthe main text in Eq. (2), where we took the Newtonian

Page 9: CENTRA, Departamento de F sica, Instituto Superior T ...Universidade de Lisboa { UL, Avenida Rovisco Pais 1, 1049 Lisboa, Portugal The existence of light, fundamental bosonic elds

9

limit of the tidal quadrupole moments expanded in spher-ical harmonics, as explained above.

Appendix B: Perturbation Theory in QuantumMechanics

In the non-relativistic limit, the scalar cloud obeys anequation which is formally equivalent to Schrodinger’sequation with a Coulomb potential, governed by a singleparameter,

α ≡Mµ . (B1)

This can be seen by making the standard ansatz for thedynamical evolution of Φ [27, 46, 47]

Φ (t, r) =1√2µ

(ψ (t, r) e−iµt + ψ∗ (t, r) eiµt

), (B2)

where ψ is a complex field which varies on timescalesmuch larger than 1/µ. Then, one can re-write Eq. (4) as

i∂

∂tψ =

(− 1

2µ∇2 − α

r

)ψ , (B3)

where we kept only terms of order O(r−1)

and linear inα.

The normalized eigenstates of the system arehydrogenic-like, with an adapted “fine structure con-stant” α and “reduced Bohr radius” a0 [6, 48],

ψnlm = e−i(ωnlm−µ)tRnl (r)Ylm (θ, φ) , (B4)

Rnl (r) = C

(2r

na0

)lL2l+1n−l−1

(2r

na0

)e−

rna0 ,

a0 = 1/µα , C =

√(2

na0

)3(n− l − 1)!

2n (n+ l)!,(B5)

where L2l+1n−l−1 is a generalized Laguerre polynomial 1. We

are adopting the convention for the quantum numbersused in Refs. [27, 28]. The eigenvalue is, up to terms oforder α5 [49]

ωnlm = µ

(1− α2

2n2− α4

8n4+

(2l − 3n+ 1)α4

n4 (l + 1/2)

). (B6)

We can estimate the size of the axion cloud by computingthe expectation value of the radius on a given state

〈r〉 =

∫ ∞0

dr r3R2nl (r) =

a02

(3n2 − l(l + 1)

). (B7)

When the binary companion is included, the tidal per-turbation can be treated in the framework of perturba-tion theory in Quantum Mechanics. The tidal potential

1 We adopt the same normalization as the built-in function ofMathematica.

δV entering in Schrodinger’s equation due to δds2tidal (2)is represented by a step function

δV = −θ (t− t0)Mc µ

R

∑|m|≤2

5

( rR

)2Y ∗lm (θc, φc)Ylm (θ, φ) ,

(B8)

where t0 is the instant when we turn it on and θ (t) isthe Heaviside function. Therefore, though there is animplicit time dependence, if one lets the system evolvefor sufficient time, it will end in a final stationary state(ignoring the loss of energy at the horizon). To describethe final picture, time-independent perturbation theoryis enough.

Let us recall the standard procedure of time-independent perturbation theory. We work in theSchrodinger’s picture (though there are no ambiguitiesfor the time-independent problem), and are trying tosolve Schrodinger’s equation

H |ψi〉 = ωi |ψi〉 , (B9)

H = H0 + λ δV , (B10)

whereH0 is the Hamiltonian of the unperturbed problem,δV is the potential corresponding to the perturbation,and λ is a dimensionless expansion parameter varyingbetween 0 (no perturbation) and 1 (full perturbation).Since we are now referring to a generic problem, we havedropped the triple indices of the “hydrogenic” spectrumand instead label different eigenstates |ψi〉 of the Hamil-tonian (and the respective eigenvalue frequencies ωi) bya single index 2.

When the system is non-degenerate, the eigenstates∣∣∣ψ(0)k

⟩of the unperturbed problem (which are assumed

to be known and in our case are given by Eq. (B5)), are

in one-to-one correspondence with the eingenvalues, ω(0)k ,

H0

∣∣∣ψ(0)i

⟩= ω

(0)i

∣∣∣ψ(0)i

⟩, (B11)

and ψ(0)n form a complete orthonormal basis⟨

ψ(0)m |ψ(0)

n

⟩= δmn , (B12)

Now, we expand the eigenstates of the perturbed sys-

tem, ψi, in terms of the basis ψ(0)k

|ψi〉 =∑k

cki

∣∣∣ψ(0)k

⟩, (B13)

and plugging in this ansatz in (B9), the coefficients ckiand the eigenvalues ωi can be obtained as a power series

2 In Quantum Mechanics literature, it is common to use E forthe (energy) eigenvalues, but since we are working in naturalunits, ~ = 1, and there is no distinction between these and the(frequency) eigenvalues being used

Page 10: CENTRA, Departamento de F sica, Instituto Superior T ...Universidade de Lisboa { UL, Avenida Rovisco Pais 1, 1049 Lisboa, Portugal The existence of light, fundamental bosonic elds

10

in λ. If the perturbation is small enough, we expect thefirst order expansions to be a good approximation [50]

ωi = ω(0)i + λω

(1)i , (B14)

cki = c(0)ki + λ c

(1)ki , (B15)

ω(1)i =

⟨ψ(0)i

∣∣∣ δV ∣∣∣ψ(0)i

⟩, (B16)

c(1)ki =

⟨ψ(0)k

∣∣∣ δV ∣∣∣ψ(0)i

⟩ω(0)i − ω

(0)k

, k 6= i , (B17)

where we omitted terms of order O(λ2). In the end, we

set λ = 1, which is the same as reabsorbing it in δV .The timescales for the transitions between two modes

can be estimated using time-dependent perturbation the-ory. This involves introducing the interaction picture andperform a Dyson series on the time-evolution operator.Since the eigenstates remain the same as in the time-independent unperturbed case, we will skip details onthis procedure and directly import the result for the first-order correction on the coefficients cki for a step-functionperturbation [51]

c(1)ki =

〈ψk| δV |ψi〉ωi − ωk

(1− e−i(ωi−ωk)t

). (B18)

Both the states |ψi〉 and frequencies ωi should be under-stood as the ones for the unperturbed system, but weommit subscripts to avoid clustering, Then, the proba-bility of the transition |i〉 → |k〉 is∣∣∣c(1)ki ∣∣∣2 = 4

∣∣∣∣ 〈ψk| δV |ψi〉ωi − ωk

∣∣∣∣2 sin2

((ωk − ωi)t

2

).(B19)

Although we do not have a continuum spectrum, forlarge timescales we can take this limit. Then, at fixed t,

we can treat the probabilities∣∣∣c(1)ki ∣∣∣2 as functions of

∆ωki = |ωk − ωi| . (B20)

Plotting it for different instants of time, one can ver-ify this function becomes increasingly peaked around∆ωki = 0 as t increases (check Fig. 5.8 of Ref. [51]).This central peak scales with t2 and has a typical widthof 1/t. If we wait enough time ∆t since the perturba-tion is introduced, the only transitions with appreciableprobability are those satisfying

∆t = 2π/∆ωki . (B21)

The final conclusion is that the typical timescale ∆tfor the transition |i〉 → |k〉 to happen is

∆ωki ∆t ∼ 1 , (B22)

which, if we momentarily insert factors of ~, can be seenas a manifestation of the energy-time uncertainty princi-ple [51]

∆E∆t ∼ ~ . (B23)

Returning to our problem, the initial data (8) corre-sponds to the stationary state (reintroducing the triple“hydrogenic”” indices)

|i〉 ∝(∣∣∣ψ(0)

211

⟩−∣∣∣ψ(0)

21−1

⟩), (B24)

up to a proportionality constant reflecting the renormal-ization done for numerical purposes. The final stateshould correspond to a stationary state |f〉 which we cancompute using the machinery developed before There isstill a caveat, which is the degeneracy between states withthe same quantum number m (B5). Though a rotatingBH will lift this degeneracy, the energy shifts due to theperturbations considered are orders of magnitude higherthan the energy scale associated with the rotation. Thus,for non-degenerate perturbation theory to be controlled,we would have to perform it at higher orders then whatwe presented.

In the degenerate scenario, the equations presented are

invalid (for example (B17) diverges when ω(0)k = ω

(0)i ).

Instead, we use the freedom in making a linear combi-nation of unperturbed degenerate eigenstates, so that inevery degenerate subspace, we pick a basis of the Hilbertspace that diagonalizes the full Hamiltonian H (B10).After this step, we can apply non-degenerate perturba-tion theory, namely Eqs. (B16) and (B17), using the new“good” basis.

Finally, the numerical data we present in the main textcorresponds to multipole expansions of the field Φ andnot to the coefficients cki (B13) describing the mix of theunperturbed states. To obtain these multipoles we haveto select them from the space representation of the finalstate. The amplitude coefficients of the mode |ψnlm〉 areobtained via

cnlm ∝〈ψnlm |δV | i〉ω(0)21 − ω

(0)nl

, (B25)

φnlm (r) ∝ cnlmRnl (r) . (B26)

In the end, we are interested in the ratio between am-plitudes so the constant of proportionality is irrelevant.The matrix elements appearing here are explicitly pre-sented in Eqs.(3.7)-(3.9) of Ref. [27]. Notice that therelative amplitude between modes with the same quan-tum numbers n and l is completely determined by theangular integrals, and since these are (quasi)degenerate,they will also follow similar time evolutions. As a con-sequence, their relative amplitude is independent of timeand the value of α, even at higher orders in perturbationtheory. This is illustrated in Fig. 4 for φn33/φn31.

A summary of the time-independent perturbation the-ory for transitions between overtones is shown in Ta-ble I for Mµ = 0.1, ε = 10−8. The relative ampli-tudes cnlm/c211 indicate that the perturbation is not thatsmall. This is even more obvious if we compute thefirst order corrections to the frequency eigenvalues (B16)which, for this configuration, are of O

(10−3

)for over-

tones n > 3, as illustrated in Table II. For this reason,

Page 11: CENTRA, Departamento de F sica, Instituto Superior T ...Universidade de Lisboa { UL, Avenida Rovisco Pais 1, 1049 Lisboa, Portugal The existence of light, fundamental bosonic elds

11

when computing the timescales of the transitions (B22),we used the first order corrected ωnlm.

TABLE II. First order corrected frequencies ωnlm predictedby time-independent theory, for a non-rotating BH and acompanion with the configuration Mµ = 0.1, ε = 10−8.A spinning BH would break the degeneracy between stateswith the same l but different m quantum number. How-ever, these corrections enter the frequency spectrum (B6) onlyat order α5. For the above configuration, these would yieldωn33−ωn31 ∼ 10−6a/M n3, where a is the angular momentumparameter a = J/M .

(n l) ωnlm × 102

2 1 9.9754

3 1 9.9224

4 1 9.7570

5 1 9.3980

4 3 9.9001

[1] G. Bertone and M. P. Tait, Tim, “A new era in thesearch for dark matter,” Nature 562 no. 7725, (2018)51–56, arXiv:1810.01668 [astro-ph.CO].

[2] L. Barack et al., “Black holes, gravitational waves andfundamental physics: a roadmap,” Class. Quant. Grav.36 no. 14, (2019) 143001, arXiv:1806.05195 [gr-qc].

[3] V. Baibhav et al., “Probing the Nature of Black Holes:Deep in the mHz Gravitational-Wave Sky,”arXiv:1908.11390 [astro-ph.HE].

[4] M. Maggiore et al., “Science Case for the EinsteinTelescope,” arXiv:1912.02622 [astro-ph.CO].

[5] V. Cardoso and L. Gualtieri, “Testing the black holeno-hair hypothesis,” Class. Quant. Grav. 33 no. 17,(2016) 174001, arXiv:1607.03133 [gr-qc].

[6] R. Brito, V. Cardoso, and P. Pani, “Superradiance,”Lect. Notes Phys. 906 (2015) pp.1–237,arXiv:1501.06570 [gr-qc].

[7] K. Eda, Y. Itoh, S. Kuroyanagi, and J. Silk,“Gravitational waves as a probe of dark matterminispikes,” Phys. Rev. D91 no. 4, (2015) 044045,arXiv:1408.3534 [gr-qc].

[8] C. F. B. Macedo, P. Pani, V. Cardoso, and L. C. B.Crispino, “Into the lair: gravitational-wave signatures ofdark matter,” Astrophys. J. 774 (2013) 48,arXiv:1302.2646 [gr-qc].

[9] E. Barausse, V. Cardoso, and P. Pani, “Canenvironmental effects spoil precision gravitational-waveastrophysics?,” Phys. Rev. D89 no. 10, (2014) 104059,arXiv:1404.7149 [gr-qc].

[10] V. Cardoso and A. Maselli, “Constraints on theastrophysical environment of binaries withgravitational-wave observations,” arXiv:1909.05870

[astro-ph.HE].[11] A. Arvanitaki, S. Dimopoulos, S. Dubovsky, N. Kaloper,

and J. March-Russell, “String Axiverse,” Phys. Rev.D81 (2010) 123530, arXiv:0905.4720 [hep-th].

[12] D. J. E. Marsh, “Axion Cosmology,” Phys. Rept. 643

(2016) 1–79, arXiv:1510.07633 [astro-ph.CO].[13] A. Arvanitaki and S. Dubovsky, “Exploring the String

Axiverse with Precision Black Hole Physics,” Phys. Rev.D83 (2011) 044026, arXiv:1004.3558 [hep-th].

[14] R. Brito, V. Cardoso, and P. Pani, “Black holes asparticle detectors: evolution of superradiantinstabilities,” Class. Quant. Grav. 32 no. 13, (2015)134001, arXiv:1411.0686 [gr-qc].

[15] A. Arvanitaki, M. Baryakhtar, S. Dimopoulos,S. Dubovsky, and R. Lasenby, “Black Hole Mergers andthe QCD Axion at Advanced LIGO,” Phys. Rev. D95no. 4, (2017) 043001, arXiv:1604.03958 [hep-ph].

[16] R. Brito, S. Ghosh, E. Barausse, E. Berti, V. Cardoso,I. Dvorkin, A. Klein, and P. Pani, “Gravitational wavesearches for ultralight bosons with LIGO and LISA,”Phys. Rev. D96 no. 6, (2017) 064050,arXiv:1706.06311 [gr-qc].

[17] R. Brito, S. Ghosh, E. Barausse, E. Berti, V. Cardoso,I. Dvorkin, A. Klein, and P. Pani, “Stochastic andresolvable gravitational waves from ultralight bosons,”Phys. Rev. Lett. 119 no. 13, (2017) 131101,arXiv:1706.05097 [gr-qc].

[18] M. C. Ferreira, C. F. B. Macedo, and V. Cardoso,“Orbital fingerprints of ultralight scalar fields aroundblack holes,” Phys. Rev. D96 no. 8, (2017) 083017,arXiv:1710.00830 [gr-qc].

[19] M. Boskovic, F. Duque, M. C. Ferreira, F. S. Miguel,and V. Cardoso, “Motion in time-periodic backgroundswith applications to ultralight dark matter haloes atgalactic centers,” Phys. Rev. D98 (2018) 024037,arXiv:1806.07331 [gr-qc].

[20] V. Cardoso, S. Chakrabarti, P. Pani, E. Berti, andL. Gualtieri, “Floating and sinking: The Imprint ofmassive scalars around rotating black holes,” Phys. Rev.Lett. 107 (2011) 241101, arXiv:1109.6021 [gr-qc].

[21] J. Zhang and H. Yang, “Gravitational floating orbitsaround hairy black holes,” Phys. Rev. D99 no. 6,

Page 12: CENTRA, Departamento de F sica, Instituto Superior T ...Universidade de Lisboa { UL, Avenida Rovisco Pais 1, 1049 Lisboa, Portugal The existence of light, fundamental bosonic elds

12

(2019) 064018, arXiv:1808.02905 [gr-qc].[22] J. Zhang and H. Yang, “Dynamic Signatures of Black

Hole Binaries with Superradiant Clouds,”arXiv:1907.13582 [gr-qc].

[23] D. Baumann, H. S. Chia, R. A. Porto, and J. Stout,“Gravitational Collider Physics,” arXiv:1912.04932

[gr-qc].[24] T. Ikeda, R. Brito, and V. Cardoso, “Blasts of Light

from Axions,” Phys. Rev. Lett. 122 no. 8, (2019)081101, arXiv:1811.04950 [gr-qc].

[25] M. Boskovic, R. Brito, V. Cardoso, T. Ikeda, andH. Witek, “Axionic instabilities and new black holesolutions,” Phys. Rev. D99 no. 3, (2019) 035006,arXiv:1811.04945 [gr-qc].

[26] A. Arvanitaki, M. Baryakhtar, and X. Huang,“Discovering the QCD Axion with Black Holes andGravitational Waves,” Phys. Rev. D91 no. 8, (2015)084011, arXiv:1411.2263 [hep-ph].

[27] D. Baumann, H. S. Chia, and R. A. Porto, “ProbingUltralight Bosons with Binary Black Holes,” Phys. Rev.D99 no. 4, (2019) 044001, arXiv:1804.03208 [gr-qc].

[28] E. Berti, R. Brito, C. F. B. Macedo, G. Raposo, andJ. L. Rosa, “Ultralight boson cloud depletion in binarysystems,” Phys. Rev. D99 no. 10, (2019) 104039,arXiv:1904.03131 [gr-qc].

[29] V. Cardoso and P. Pani, “Tidal acceleration of blackholes and superradiance,” Class. Quant. Grav. 30(2013) 045011, arXiv:1205.3184 [gr-qc].

[30] V. Cardoso, E. Franzin, A. Maselli, P. Pani, andG. Raposo, “Testing strong-field gravity with tidal Lovenumbers,” Phys. Rev. D95 no. 8, (2017) 084014,arXiv:1701.01116 [gr-qc].

[31] V. Cardoso and F. Duque, “Environmental effects inGW physics: tidal deformability of black holesimmersed in matter,” arXiv:1912.07616 [gr-qc].

[32] S. Taylor and E. Poisson, “Nonrotating black hole in apost-Newtonian tidal environment,” Phys. Rev. D78(2008) 084016, arXiv:0806.3052 [gr-qc].

[33] H. Witek, V. Cardoso, A. Ishibashi, and U. Sperhake,“Superradiant instabilities in astrophysical systems,”Phys. Rev. D87 no. 4, (2013) 043513, arXiv:1212.0551[gr-qc].

[34] H. Yoshino and H. Kodama, “Gravitational radiationfrom an axion cloud around a black hole: Superradiantphase,” PTEP 2014 (2014) 043E02, arXiv:1312.2326[gr-qc].

[35] J. A. Orosz, J. E. McClintock, J. P. Aufdenberg, R. A.Remillard, M. J. Reid, R. Narayan, and L. Gou, “Themass of the black hole in cygnus x-1,” The AstrophysicalJournal 742 no. 2, (Nov, 2011) 84.http://dx.doi.org/10.1088/0004-637X/742/2/84.

[36] P. Pani, V. Cardoso, L. Gualtieri, E. Berti, andA. Ishibashi, “Black hole bombs and photon massbounds,” Phys. Rev. Lett. 109 (2012) 131102,arXiv:1209.0465 [gr-qc].

[37] V. Cardoso, O. J. C. Dias, G. S. Hartnett,M. Middleton, P. Pani, and J. E. Santos, “Constrainingthe mass of dark photons and axion-like particlesthrough black-hole superradiance,” JCAP 1803 no. 03,(2018) 043, arXiv:1801.01420 [gr-qc].

[38] H. Yoshino and H. Kodama, “Probing the stringaxiverse by gravitational waves from Cygnus X-1,”PTEP 2015 no. 6, (2015) 061E01, arXiv:1407.2030[gr-qc].

[39] L. Sun, R. Brito, and M. Isi, “Search for ultralightbosons in Cygnus X-1 with Advanced LIGO,”arXiv:1909.11267 [gr-qc].

[40] GRAVITY Collaboration, R. Abuter et al., “Detectionof the gravitational redshift in the orbit of the star S2near the Galactic centre massive black hole,” Astron.Astrophys. 615 (2018) L15, arXiv:1807.09409[astro-ph.GA].

[41] S. Naoz, C. M. Will, E. Ramirez-Ruiz, A. Hees, A. M.Ghez, and T. Do, “A hidden friend for the galacticcenter black hole, Sgr A*,” arXiv:1912.04910

[astro-ph.GA].[42] C. A. R. Herdeiro and E. Radu, “Kerr black holes with

scalar hair,” Phys. Rev. Lett. 112 (2014) 221101,arXiv:1403.2757 [gr-qc].

[43] T. Binnington and E. Poisson, “Relativistic theory oftidal Love numbers,” Phys. Rev. D80 (2009) 084018,arXiv:0906.1366 [gr-qc].

[44] T. Damour and A. Nagar, “Relativistic tidal propertiesof neutron stars,” Phys. Rev. D80 (2009) 084035,arXiv:0906.0096 [gr-qc].

[45] E. Poisson and C. M. Will, Gravity: Newtonian,Post-Newtonian, Relativistic. Cambridge UniversityPress, 2014.

[46] D. N. Page, “Classical and quantum decay ofoscillatons: Oscillating selfgravitating real scalar fieldsolitons,” Phys. Rev. D70 (2004) 023002,arXiv:gr-qc/0310006 [gr-qc].

[47] R. F. P. Mendes and H. Yang, “Tidal deformability ofboson stars and dark matter clumps,” Class. Quant.Grav. 34 no. 18, (2017) 185001, arXiv:1606.03035[astro-ph.CO].

[48] S. L. Detweiler, “KLEIN-GORDON EQUATION ANDROTATING BLACK HOLES,” Phys. Rev. D22 (1980)2323–2326.

[49] D. Baumann, H. S. Chia, J. Stout, and L. ter Haar,“The Spectra of Gravitational Atoms,”arXiv:1908.10370 [gr-qc].

[50] D. Griffiths, Introduction to Quantum Mechanics.Cambridge University Press, 2017.https://books.google.pt/books?id=0h-nDAAAQBAJ.

[51] J. Sakurai and J. Napolitano, Modern QuantumMechanics. Addison Wesley, 2011.https://books.google.pt/books?id=N4I-AQAACAAJ.