You are on page 1of 29

Perturbations of slowly rotating black holes:

massive vector fields in the Kerr metric


Paolo Pani,1, ∗ Vitor Cardoso,1, 2, † Leonardo Gualtieri,3, ‡ Emanuele Berti,2, 4, § and Akihiro Ishibashi5, 6, ¶
1
CENTRA, Departamento de Fı́sica, Instituto Superior Técnico,
Universidade Técnica de Lisboa - UTL, Av. Rovisco Pais 1, 1049 Lisboa, Portugal.
2
Department of Physics and Astronomy, The University of Mississippi, University, MS 38677, USA.
3
Dipartimento di Fisica, Università di Roma “La Sapienza” & Sezione INFN Roma1, P.A. Moro 5, 00185, Roma, Italy.
4
California Institute of Technology, Pasadena, CA 91109, USA
5
Theory Center, Institute of Particle and Nuclear Studies,
High Energy Accelerator Research Organization (KEK), Tsukuba, 305-0801, Japan
6
Department of Physics, Kinki University, Higashi-Osaka 577-8502, Japan
We discuss a general method to study linear perturbations of slowly rotating black holes which
arXiv:1209.0773v2 [gr-qc] 22 Oct 2012

is valid for any perturbation field, and particularly advantageous when the field equations are not
separable. As an illustration of the method we investigate massive vector (Proca) perturbations
in the Kerr metric, which do not appear to be separable in the standard Teukolsky formalism.
Working in a perturbative scheme, we discuss two important effects induced by rotation: a Zeeman-
like shift of nonaxisymmetric quasinormal modes and bound states with different azimuthal number
m, and the coupling between axial and polar modes with different multipolar index ℓ. We explicitly
compute the perturbation equations up to second order in rotation, but in principle the method can
be extended to any order. Working at first order in rotation we show that polar and axial Proca
modes can be computed by solving two decoupled sets of equations, and we derive a single master
equation describing axial perturbations of spin s = 0 and s = ±1. By extending the calculation
to second order we can study the superradiant regime of Proca perturbations in a self-consistent
way. For the first time we show that Proca fields around Kerr black holes exhibit a superradiant
instability, which is significantly stronger than for massive scalar fields. Because of this instability,
astrophysical observations of spinning black holes provide the tightest upper limit on the mass of
the photon: mγ . 4 × 10−20 eV under our most conservative assumptions. Spin measurements for
the largest black holes could reduce this bound to mγ . 10−22 eV or lower.

PACS numbers: 04.40.Dg, 04.62.+v, 95.30.Sf

I. INTRODUCTION in the context of high-energy physics [11]. Perturbation


theory can shed light on several open issues, such as the
Linear perturbations of black holes (BHs) play a ma- stability properties of BH spacetimes in higher dimen-
jor role in physics. Many astrophysical processes can be sions and in asymptotically anti-de Sitter spacetimes.
modeled as small deviations from analytically known BH Within the gauge-gravity duality, some of the correla-
backgrounds: for instance, perturbative calculations of tion functions and transport coefficients are related to
quasinormal modes (QNMs) are useful to describe the the lowest order BH QNMs. In a semiclassical treatment
late stages of compact binary mergers or gravitational of BH evaporation, the calculation of greybody factors
collapse [1–5], and BH perturbation theory provides an (which may be of direct interest for ongoing experiments)
accurate description of the general relativistic dynam- relies heavily on our ability to understand wave scatter-
ics of extreme mass-ratio inspirals [6–8]. Even when the ing in rotating BH spacetimes.
understanding of complex astrophysical phenomena re- BH perturbation theory is a useful tool to investi-
quires numerical relativity, the set up of the numerical gate issues in astrophysics and high-energy physics as
simulations and the interpretation of the results are eas- long as the radial and angular parts of the perturbation
ier when we can rely on prior perturbative knowledge of equations are separable. This usually happens when the
the problem (see e.g. [9, 10]). background spacetime has special symmetries. If sep-
Besides their interest in modeling gravitational-wave arable, the perturbation equations in the frequency do-
sources for present and future detectors in Einstein’s the- main reduce to a system of ordinary differential equations
ory and in various proposed extensions of general relativ- (ODEs) [12]. Separability is the norm if the background
ity, perturbative studies of BH dynamics are also relevant spacetime is spherically symmetric. Teukolsky [13] dis-
covered that a large class of perturbation equations is ex-
ceptionally separable in the Kerr metric, the underlying
∗ paolo.pani@ist.utl.pt reason being that the Kerr spacetime is of type D in the
† vitor.cardoso@ist.utl.pt Petrov classification. Separability is much more difficult
‡ Leonardo.Gualtieri@roma1.infn.it to achieve in the Kerr-Newman metric [12] and in higher
§ berti@phy.olemiss.edu spacetime dimensions [14]. In the Kerr background, per-
¶ akihiro.ishibashi@kek.jp turbations induced by massless fields with integer and
2

half-integer spins (including Dirac and Rarita-Schwinger to higher spin values. For all these reasons we are con-
fields [12, 13, 15]) are all separable, but this does not seem fident that perturbative studies of slowly rotating BHs
to be possible for some classes of perturbations, such as will further our understanding of several interesting open
massive vector (Proca) perturbations [16]. problems in astrophysics and high-energy physics.
Here we discuss a general method to study linear per- As an interesting testing ground of the slow-rotation
turbations of slowly rotating BH backgrounds that is par- approximation, here we focus on massive vector (Proca)
ticularly useful when the perturbation variables are not perturbations of slowly rotating Kerr BHs. It is well
separable. The method is an extension of Kojima’s work known that massive bosonic fields in rotating BH space-
on perturbations of slowly rotating neutron stars [17–19]. times can trigger superradiant instabilities [31–37]. For
Slowly rotating backgrounds are “close enough” to spher- scalar fields the instability is well studied. It is regulated
ical symmetry that an approximate separation of the per- by the dimensionless parameter M µ (in units G = c = 1),
turbation equations in radial and angular parts becomes where M is the BH mass and ms = µ~ is the bosonic
possible. The field equations can be Fourier transformed field mass1 , and it is strongest for maximally spinning
in time and expanded in spherical harmonics and they BHs, when M µ ∼ 1. For a solar mass BH and a field
reduce, in general, to a coupled system of ODEs. This of mass ms ∼ 1 eV the parameter M µ ∼ 1010 . In
approach can be seen as a two-parameter perturbative this case the instability is exponentially suppressed [38]
expansion [20], where the small parameters are the am- and in many cases of astrophysical interest the instability
plitude of the perturbation and the angular velocity of timescale would be larger than the age of the universe.
the background. The method we present can in princi- However, strong superradiant instabilities (M µ ∼ 1) can
ple be extended to any order in the rotation parameter; occur either for light primordial BHs which may have
here we derive the perturbation equations explicitly up been produced in the early universe [39–41] or for ul-
to second order. tralight exotic particles found in some extensions of the
To first order in rotation, the final system of coupled standard model, such as the “string axiverse” scenario
ODEs can be simplified by dropping terms that couple [42, 43]. In this scenario, massive scalar fields with
perturbations with different values of the harmonic in- 10−33 eV < ms < 10−18 eV could play a key role in
dices. This is a surprisingly good approximation: extend- cosmological models. Superradiant instabilities may al-
ing previous arguments by Kojima [19], we show that the low us to probe the existence of such ultralight bosonic
neglected coupling terms can affect the frequencies and fields by producing gaps in the mass-spin BH Regge spec-
damping times of the QNMs only at second or higher trum [42, 43], by modifying the inspiral dynamics of com-
order in the rotation rate. We support the analytical pact binaries [37, 44, 45] or by inducing a “bosenova”, i.e.
argument by numerical results, showing that the modes collapse of the axion cloud (see e.g. [46–48]).
computed with and without the coupling terms coincide Similar instabilities are expected to occur for massive
to first order in the rotation rate. At second order in ro- hidden U (1) vector fields, which are also a generic feature
tation, the coupling of perturbations with different har- of extensions of the standard model [49–52]. Massive vec-
monic indices cannot be neglected. However, a notion of tor perturbations of rotating BHs are expected to induce
“conserved quantum number” ℓ is preserved: perturba- a superradiant instability, but an explicit demonstration
tions with given parity and harmonic index ℓ are coupled of this effect has been lacking. While this problem has
with perturbations with opposite parity and harmonic in- been widely studied for massive scalar fields [31–37], the
dices ℓ ± 1, and with perturbations with the same parity case of massive vector fields is still uncharted territory;
and harmonic indices ℓ ± 2 [21]. incursions in the topic seem to be restricted to nonro-
tating backgrounds [16, 53–55]. The main reason is that
The main limitation of the present approach is the the Proca equation, which describes massive vector fields,
slow-rotation approximation. Currently this is not a
does not seem to be separable in the Kerr background. In
restriction in extensions of general relativity including the slow-rotation approximation we can reduce the prob-
quadratic curvature corrections, where rotating BH so-
lem of finding QNMs to a tractable system of coupled
lutions are only known in the slow-rotation limit [22– ODEs, where polar perturbations with angular index ℓ
26]. Furthermore, as we show in this paper, a slow-
are generically coupled to axial perturbations with index
rotation approximation is sufficient to describe important ℓ ± 1 (and viceversa).
effects, like the superradiant instability of massive fields
In this paper we derive the Proca perturbation equa-
around rotating BHs. While a second-order expansion is tions up to second order in rotation, and for the first time
needed to describe superradiance consistently, even the
we show that rotating BHs are indeed unstable to massive
first-order approximation provides accurate results well vector perturbations in the superradiant regime. At first
beyond the nominal region of validity of the approxima-
order in rotation, perturbations with a given parity and
tion. Similar extrapolations have been used in the past
to predict the existence and timescale of r-mode instabil-
ities in the relativistic theory of stellar perturbations (see
e.g. [27–30]), and it is not unreasonable to expect that 1 In this paper, with a slight abuse of notation, we will use µ for
some quantities computed in the small-rotation regime both the scalar field mass (µ = ms /~) and the vector field mass
(e.g. reflection coefficients) can be safely extrapolated (µ = mv /~). The meaning should be clear from the context.
3

5 6 7 8 9 10
10
1.0
10 10 10 10 10 In addition we compute stable and unstable bound-state
modes (which are spatially localized within the vicinity
0.8
of the BH and decay exponentially at spatial infinity [35])
0.6
2

of massive scalar perturbations up to second order in ro-


J/M

0.4 tation, and we compare our results to the numerical cal-


0.2 culations by Dolan [35]. We find good quantitative agree-
polar S=-1, fit II
0.0 ment with numerical results for nonsuperradiant frequen-
0.8 cies, and we reproduce the imaginary part of superradi-
0.6 antly unstable modes within a factor 3 for ã . 0.7, i.e.
2
J/M

0.4 even for moderately large spin.


0.2 Extrapolations of our perturbative results indicate
0.0
polar S=-1, fit I
that, because of the strong superradiant instability of
0.8
Proca perturbations, astrophysical BH spin measure-
mv=10
-18
eV ments may set the most stringent bounds on the masses
0.6
2

-19
of vector fields [59]. This is shown in Fig. 1, where we
J/M

mv=10 eV -21
mv=2x10 eV
0.4 mv=10
-20
eV summarize the astrophysical implications of our results
-21 -18
0.2 mv=2x10 eV mv=10 eV
(cf. Sect. VII for details). To be more specific, in Fig. 1
axial
0.0 5 6 7 8 9 10 we show exclusion regions in the “BH Regge plane” (cf.
10 10 10 10 10 10
M/MO. Fig. 3 of [43]), i.e. we plot contours corresponding to an
instability timescale of the order of a typical accretion
FIG. 1. (color online) Contour plots in the BH Regge
timescale (the Salpeter time τSalpeter = 4.5 × 107 yr) for
plane [43] corresponding to an instability timescale shorter four different masses of the Proca field (mv = 10−18 eV,
than a typical accretion timescale, τSalpeter = 4.5 × 107 yr, for 10−19 eV, 10−20 eV and 2 × 10−21 eV) and for axial
different values of the vector field mass mv = µ~ (from left modes (bottom panel) and polar modes (top and middle
to right: mv = 10−18 eV, 10−19 eV, 10−20 eV, 2 × 10−21 eV). panels). For polar modes we consider only the S = −1
For polar modes we consider the S = −1 polarization, which polarization, which provides the strongest instability (cf.
provides the strongest instability, and we use both Eq. (100) Sec. (VI D) for details). While our numerical results for
(fit II, top panel) and Eq. (95) (fit I, middle panel), and for the axial modes are supported by an analytical formula,
axial modes we use Eq. (95) (bottom panel). Dashed lines in the polar case we have used two different functions to
bracket our estimated numerical errors. The experimental fit the numerical data at second order in the BH spin.
points (with error bars) refer to the mass and spin estimates
The plot shows that essentially any spin measurement
of supermassive BHs listed in Table 2 of [56]; the rightmost
point corresponds to the supermassive BH in Fairall 9 [57]. for supermassive BHs with 106 M⊙ . M . 109 M⊙ would
Supermassive BHs lying above each of these curves would be exclude a wide range of vector field masses.
unstable on an observable timescale, and therefore they ex- The rest of the paper is organized as follows. In Sec. II
clude a whole range of Proca field masses. we outline the general method, which is valid for any sta-
tionary and axisymmetric background and for any kind of
perturbation. In Sec. III we study massive Klein-Gordon
perturbations of slowly rotating Kerr BHs up to second
angular index ℓ are coupled to perturbations with ℓ ± 1 order in rotation and outline how the method can be ex-
and opposite parity (cf. Eqs. (59), (60) and (61) below). tended, at least in principle, to higher orders. In Sec. IV
However we show both analytically and numerically that we specialize to Proca perturbations of slowly rotating
this coupling does not affect the eigenfrequencies of the Kerr BHs. In Sec. V we set up the eigenvalue problem
BH. In addition, the axial sector is described by a sin- for quasinormal modes and bound states, and in Sec. VI
gle master equation [Eq. (66) below], which is valid for we present our numerical results. Section VII deals with
both massive scalar and axial Proca perturbations. At the implications of the superradiant instability for astro-
second order in rotation the structure of the equations physics and fundamental physics. In Sec. VIII we draw
remain the same, but the ℓ ± 1 couplings do affect the some conclusions and we discuss possible future applica-
eigenfrequency spectrum. tions of the method. To improve readability we relegate
Quasinormal frequencies, damping times of stable some technical material to the appendices. Appendix A
modes and instability timescales of unstable modes in lists the coefficients appearing in the scalar and Proca
the slow-rotation limit can be computed using standard perturbation equations. Appendix B shows explicitly the
methods [4, 16, 35, 58]. As a test of our approxima- perturbation equations for a Proca field on a slowly ro-
tion scheme we revisit two problems for which “exact” tating Kerr background to first order in rotation, and
solutions are known in the Kerr background. We study appendix C generalizes Detweiler’s analytical calculation
QNMs of massless vector perturbations to first order in of the unstable massive scalar modes of a Kerr BH [60] to
rotation and we find that our slow-rotation method re- axial perturbations of a massive vector field up to linear
produces known results [4] with accuracy better than 1% order in rotation.
when the dimensionless Kerr parameter ã = J/M 2 . 0.3. To reduce the risk of typographical errors and facilitate
4

comparison with our results, we made many of our calcu- transform the perturbations, and expand them in tensor
lations available online as Mathematica notebooks [61]. spherical harmonics:
(i)
δXµ1 ... (t, r, ϑ, ϕ) = δXℓm (r)Yµℓm (i) −i ωt
1 ...
e , (3)
II. PERTURBATIONS OF A SLOWLY ℓm (i)
ROTATING BLACK HOLE: GENERAL METHOD where Yµ1 ... is a basis of scalar, vector or tensor harmon-
ics (depending on the tensorial nature of the perturbation
δX) and the frequency ω is, in general, complex. The
In this section we outline the strategy to study generic (i)
(scalar, vector, tensor, etcetera) perturbations of any sta- perturbation variables δXℓm (r) can be classified as “po-
tionary, axisymmetric background up to second order in lar” or “axial” depending on their behavior under parity
the rotation parameter, but in principle the formalism transformations (ϑ → π − ϑ, ϕ → ϕ + π): polar and
can be extended to any order. We consider the most axial perturbations are multiplied by (−1)ℓ and (−1)ℓ+1 ,
general stationary axisymmetric spacetime respectively.
With the introduction of the expansion (3), the linear
ds20 = −H 2 dt2 +Q2 dr2 +r2 K 2 dϑ2 + sin2 ϑ(dϕ − Ldt)2 , response of the system is fully characterized by the quan-
 
(i)
tities δXℓm (r). The perturbation equations, expanded
where H, Q, K and L are functions of r and ϑ only. To in spherical harmonics and Fourier transformed in time,
second order, the metric above can be expanded as [62] yield a coupled system of ODEs in the perturbation func-
(i)
  tions δXℓm (r). In the case of a spherically symmetric
2 2 2B2 background, perturbations with different values of (ℓ, m),
ds0 = − F (r) [1 + F2 ] dt + B(r)−1
1+ dr2
r − 2M as well as perturbations with opposite parity, are decou-
+ r2 (1 + k2 ) dϑ2 + sin2 ϑ(dϕ − ̟dt)2 , pled. In a rotating, axially symmetric background, per-
 
(1)
turbations with different values of m are decoupled2 but
where M is the mass of the spacetime, ̟ is a function perturbations with different values of ℓ are not. However,
of r linear in the rotation parameter, and F2 , B2 and k2 in the limit of slow rotation there is a Laporte-like “se-
are functions of r and ϑ quadratic in the rotation pa- lection rule” [21]: at first order in ã, perturbations with
rameter. Because H, K and Q all transform like scalars a given value of ℓ are only coupled to those with ℓ ± 1 and
under rotation, they can be expanded in scalar spher- opposite parity, similarly to the case of rotating stars. At
ical harmonics which, due to axisymmetry, reduce to second order, perturbations with a given value of ℓ are
the Legendre polynomials Pℓ (ϑ). As shown in Ref. [62], also coupled to those with ℓ ± 2 and same parity, and so
only ℓ = 0 and ℓ = 2 polynomials contribute at sec- on.
ond order in rotation; therefore F2 can be expressed as In general, the perturbation equations can be written
F2 (r, ϑ) = Fr (r) + Fϑ (r)P2 (ϑ), and the same applies to in the following form:
B2 (r, ϑ) and k2 (r, ϑ). 0 = Aℓ + ãmĀℓ + ã2 Âℓ
At first order in rotation the background (1) can be
written in the simpler form + ã(Qℓ P̃ℓ−1 + Qℓ+1 P̃ℓ+1 )
h i
+ ã2 Qℓ−1 Qℓ Ăℓ−2 + Qℓ+2 Qℓ+1 Ăℓ+2 + O(ã3 ) ,
ds20 = − F (r)dt2 + B(r)−1 dr2 + r2 d2 Ω
(4)
− 2̟(r) sin2 ϑdϕdt . (2)
2
0 = Pℓ + ãmP̄ℓ + ã P̂ℓ
Note that, given a nonrotating metric, the gyromagnetic + ã(Qℓ Ãℓ−1 + Qℓ+1 Ãℓ+1 )
function ̟(r) can be computed using the approach orig- h i
inally developed by Hartle [62]. For example, slowly + ã2 Qℓ−1 Qℓ P̆ℓ−2 + Qℓ+2 Qℓ+1 P̆ℓ+2 + O(ã3 ) .
rotating Kerr BHs [cf. Eq. (18) below] correspond to (5)
F (r) = B(r) = 1 − 2M/r and ̟ = 2M 2 ã/r, where M
and J = M 2 ã are the mass and the angular momentum Here
of the BH. Furthermore, the general metrics (1) and (2) r
ℓ 2 − m2
encompass (among others) slowly rotating Kerr-Newman Qℓ = ; (6)
4ℓ2 − 1
BHs and BH solutions in modified gravity theories; the
latter have been derived analytically only up to first or- Aℓ , Āℓ , Ãℓ , Âℓ , Ăℓ are linear combinations of the axial
der [22, 25] (and more recently to second order [26]) in perturbations and of their derivatives, with multipolar
the BH spin.
Scalar, vector and tensor field equations in the back-
ground metric (1) can be linearized in the field pertur- 2 From now on we will append the relevant multipolar index ℓ to
bations (that we shall schematically denote by δX) as any perturbation variable but we will omit the index m, because
well as in the BH angular momentum. We neglect terms in an axisymmetric background it is possible to decouple the
of second order in the perturbation amplitude δX and perturbation equations so that all quantities have the same value
terms of third order in the rotation parameter ã, Fourier of m.
5

index ℓ; similarly, Pℓ , P̄ℓ , P̃ℓ , P̂ℓ , P̆ℓ are linear combina- where ω0 is the eigenfrequency of the nonrotating space-
tions of the polar perturbations and of their derivatives, time and ωn is the n-th order correction (note that ω1
with index ℓ. The dimensionless parameter ã keeps track and ω2 are generically polynomials in m but, due to the
of the order of the various terms in the slow-rotation ex- above symmetry, ω1 is an even polynomial). Crucially,
pansion. only the terms (P̄ℓ , Āℓ ) in Eqs. (4) and (5) can contribute
The structure of Eqs. (4)–(5) is the following. Pertur- to ω1 . Indeed, due to the factor ã in front of all terms
bations with a given parity and index ℓ are coupled to: (P̄ℓ , Āℓ , P̃ℓ , Ãℓ ) and to their linearity, at first order in ã
(i) perturbations with opposite parity and index ℓ ± 1 at we can simply take the zeroth order (in rotation) expan-
order ã; (ii) perturbations with same parity and same in- sion of these terms. That is, to our level of approxima-
dex ℓ up to order ã2 ; (iii) perturbations with same parity tion the terms (P̄ℓ , Āℓ , P̃ℓ , Ãℓ ) in Eqs. (4) and (5) only
and index ℓ ± 2 at order ã2 . contain the perturbations of the nonrotating, spherically
From Eq. (6) it follows that Q±m = 0, and therefore symmetric background. Since the latter do not explicitly
if |m| = ℓ the coupling of perturbations with index ℓ to depend on m, the m dependence in Eq. (8) can only arise
perturbations with indices ℓ − 1 and ℓ − 2 is suppressed. from the terms (P̄ℓ , Āℓ ) to zeroth order (recall that the
Since the contribution |m| = ℓ dominates the linear re- Qℓ ’s are even functions of m, so they do not contribute).
sponse of the system, this general property of Eqs. (4)– In the case of slowly rotating stars, this argument is re-
(5) is reminiscent of a “propensity rule” in atomic theory, inforced by numerical simulations for the m = 0 axisym-
which states that transitions ℓ → ℓ + 1 are strongly fa- metric modes [19] and in the general nonaxisymmetric
vored over transitions ℓ → ℓ − 1 (cf. [21] and references case [63], which show that the coupling terms do not af-
therein). fect the first order corrections to the QNMs. In this work
Due to the coupling between different multipolar in- we shall verify that the same result holds also for mass-
dices, the spectrum of the solutions of Eqs. (4)–(5) is less vector (e.g. electromagnetic) modes and for Proca
extremely rich. However, if we are interested in the char- modes of a Kerr BH. If we are interested in the modes of
acteristic modes of the slowly rotating background to first the rotating background to O(ã) the coupling terms can
or to second order in ã, the perturbation equations can therefore be neglected, and the eigenvalue problem can
be considerably simplified. We discuss the truncation to be written in the form
first and to second order in the next two sections, respec-
Aℓ + ãmĀℓ = 0 , (9)
tively.
Pℓ + ãmP̄ℓ = 0 . (10)
In these equations the polar and axial perturbations (as
A. Eigenvalue spectrum to first order well as perturbations with different values of the har-
monic indices) are decoupled from each other, and can
Let us consider the first order expansion of Eqs. (4) and be studied independently.
(5). We expand all quantities to first order and we ignore
the terms Âℓ , Ăℓ , P̂ℓ and P̆ℓ , which are multiplied by ã2 .
Furthermore the terms (P̃ℓ , Ãℓ ) in Eqs. (4) and (5) do B. Eigenvalue spectrum to second order
not contribute to the eigenfrequencies at first order in
ã. This was first shown by Kojima [19] using symmetry For what concerns the calculation of the eigenfrequen-
arguments for the axisymmetric case m = 0. Here we cies to second order in ã, the equations above can be
extend his argument to any value of m, for the generic simplified as follows. To begin with, we remark that all
set of equations (4) and (5). Let us start by noting that, of the coefficients in Eqs. (4) and (5) are linear com-
at first order, Eqs. (4) and (5) are invariant under the binations of axial or polar perturbation functions, aℓm ,
simultaneous transformations pℓm , respectively. These functions can be expanded in
the rotation parameter as
aℓm → ∓aℓ−m , pℓm → ±pℓ−m , (7a)
(0) (1) (2)
ã → −ã , m → −m , (7b) aℓm = aℓm + ã aℓm + ã2 aℓm
(0) (1) (2)
pℓm = pℓm + ã pℓm + ã2 pℓm . (11)
where aℓm (pℓm ) schematically denotes all of the axial
(polar) perturbation variables with indices (ℓ, m). The The terms Ăℓ±2 and P̆ℓ±2 are multiplied by factors ã2 ,
invariance follows from the linearity of the terms in so they only depend on the zeroth-order perturbation
Eqs. (4) and (5) and from the fact that the Qℓ ’s are even (0) (0)
functions, aℓm , pℓm . The terms Ãℓ±1 and P̃ℓ±1 are mul-
functions of m. The boundary conditions that define the tiplied by factors ã, so they only depend on zeroth- and
characteristic modes of the BH are also invariant under (0) (0) (1) (1)
first-order perturbation functions aℓm , pℓm , aℓm , pℓm .
the transformation above (cf. Eqs. (72) and (77) below).
Since in the nonrotating limit axial and polar perturba-
Therefore in the slow-rotation limit the eigenfrequencies
tions are decoupled, a possible consistent set of solutions
can be expanded as (0)
of the system (4)–(5) has aℓ±2m ≡ 0; another consis-
(0)
ω = ω0 + m ω1 ã + ω2 ã2 + O(ã3 ) , (8) tent set of solutions of the same system has pℓ±2m ≡ 0.
6

Such solutions, which we can call, following Refs. [29, 30], this section by noting that this procedure can be applied
“axial-led” and “polar-led” perturbations respectively, to any slowly rotating spacetime and to any kind of per-
can be found by solving the following subsets of the equa- turbation. For concreteness, in the next section we shall
tions (4)–(5): specialize the method to the case of massive scalar and
vector perturbations of the Kerr metric.
Aℓ + ãmĀℓ + ã2 Âℓ + ã(Qℓ P̃ℓ−1 + Qℓ+1 P̃ℓ+1 ) = 0 ,
(12)
III. MASSIVE SCALAR PERTURBATIONS OF
Pℓ+1 + ãmP̄ℓ+1 + ãQℓ+1 Ãℓ = 0 , (13)
SLOWLY ROTATING KERR BLACK HOLES
Pℓ−1 + ãmP̄ℓ−1 + ãQℓ Ãℓ = 0 , (14)
and In order to illustrate how the slow-rotation expansion
works, in this section we begin by working out the sim-
Pℓ + ãmP̄ℓ + ã2 P̂ℓ + ã(Qℓ Ãℓ−1 + Qℓ+1 Ãℓ+1 ) = 0 , plest case. We consider the massive Klein-Gordon equa-
(15) tion for a scalar field perturbation φ around a rotating
Aℓ+1 + ãmĀℓ+1 + ãQℓ+1 P̃ℓ = 0 , (16) BH, and we work out the perturbation equations up to
second order in rotation. The formalism applies to a
Aℓ−1 + ãmĀℓ−1 + ãQℓ P̃ℓ = 0 . (17) generic stationary and axisymmetric background, but it
In the second and third equations of the two systems is natural to focus on the case of interest in general rel-
above we have dropped the Ãℓ±2 and P̃ℓ±2 terms, be- ativity, i.e. the Kerr metric in Boyer-Lindquist coordi-
cause these only enter at zeroth order, and we have set nates:
(0) (0)
4rM 2
 
aℓ±2m ≡ 0 and pℓ±2m ≡ 0. 2M r Σ
Interestingly, within this perturbative scheme a notion
2
dsKerr = − 1 − dt2 + dr2 − ã sin2 ϑdϕdt
Σ ∆ Σ
of “conserved quantum number” ℓ is still meaningful,
2rM 3 2 4
 
2 2 2 2 2
even though, for any given ℓ, rotation couples terms with +Σdϑ + (r + M ã ) sin ϑ + ã sin ϑ dϕ2 ,
opposite parity and different multipolar index. In the Σ
most relevant case from the point of view of superradiant (18)
instabilities, i.e. the case ℓ = m, the last equations (14)
and (17) are automatically satisfied, because Qm = 0 and where Σ = r2 + 2 2 2
√M ã cos ϑ, ∆ = (r − r+ )(r − r− ) and
2
r± = M (1 ± 1 − ã ). In what follows we shall ex-
perturbations with ℓ < m automatically vanish.
It is worth stressing that, even though in principle pand the metric and all other quantities of interest to
there may be modes which do not belong to the classes of second order in ã. At this order, the event horizon r+ ,
“axial-led” or “polar-led” perturbations, all solutions be- the Cauchy horizon r− and the outer ergosphere rS + can
longing to one of these classes which fulfill the appropri- be written in the form
ate boundary conditions defining QNMs or bound states ã2 M ã2
 
are also solutions of the full system (4)–(5) and belong r+ = 2M 1 − , r− = ,
4 2
to the eigenspectrum (up to second order in the rota-
ã2
 
tion rate). This is particularly relevant since, as we shall rS + = 2M 1 − cos2 ϑ . (19)
show, these classes contain also superradiantly unstable 4
modes.
In particular, note that second-order corrections are nec-
In the case of massive scalar perturbations, as we show
essary for the ergoregion to be located outside the event
in the next section, the coupling to ℓ±2 can be eliminated
horizon.
by defining a suitable linear combination of the eigen-
The massive Klein-Gordon equation reads
functions with different multipolar indices [cf. Eq. (37)],
so that our procedure allows us to compute the entire φ = µ2 φ , (20)
BH spectrum up to second order. By a similar, suitable
“rotation” in eigenfunction space it may be possible to where ms = µ~ is the mass of the scalar field. We de-
recast Eqs. (4)–(5) in the forms (12)–(14) and (15)–(17), compose the field in spherical harmonics:
but we did not find a general proof of this conjecture. If
the conjecture is correct, studying the “axial-led” and Ψ (r)
√ ℓ
X
φ= e−i ωt Y ℓ (ϑ, ϕ) , (21)
“polar-led” systems may be sufficient to describe the en- r + ã2 M 2
2
ℓm
tire BH spectrum at second order in rotation.
To summarize, the eigenfrequencies (or at least a sub- and expand the square root above to second order in ã.
set of the eigenfrequencies) of the general system (4)–(5) Schematically, we obtain the following equation:
can be found, at first order in ã, by solving the two de-
coupled sets (9) and (10) for axial and polar perturba- Aℓ Y ℓ + Dℓ cos2 ϑY ℓ = 0 , (22)
tions, respectively. At second order in ã we must solve
either the set (12)–(14) or the set (15)–(17) for “axial- where a sum over (ℓ, m) is implicit, and the explicit form
led” and “polar-led” modes, respectively. We conclude of Aℓ and Dℓ is given in Appendix A 1. The crucial point
7

is that Dℓ is proportional to ã2 , so the second term in form (cos ϑ)n Y ℓ will appear. As we discuss in the next
the equation above is zero to first order in rotation. If section this is also true for spin-1 or spin-2 perturbations,
we consider a first-order expansion in rotation the scalar except that now the perturbation equations will contain
equation is already decoupled, and it can be cast in the combinations of vector and tensor spherical harmonics,
form and this introduces terms such as (sin ϑ)n ∂ϑ Y ℓ , which
can be decoupled in a similar fashion by repeated appli-
4mM 2 ãω
 
2M cation of the identities listed above. This procedure is
D̂2 Ψℓ − + F 3 Ψℓ = 0 , (23)
r3 r well known in quantum mechanics, and the coefficients
Qℓ are related to the usual Clebsch-Gordan coefficients.
where F = 1 − 2M/r, dr/dr∗ = F and we defined the Using the explicit form of the coefficients given in Ap-
operator [16] pendix A 1, the field equations (31) schematically read
d2
 
ℓ(ℓ + 1)
"
2
D̂2 = 2 + ω − F 2
+µ . (24) d2 Ψℓ
dr∗ r2 + Vℓ Ψℓ + ã2 Uℓ+2 Ψℓ+2 + Uℓ−2 Ψℓ−2
dr∗2
Equation (23) coincides with Teukolsky’s master equa-
#
d2 Ψℓ+2 d2 Ψℓ−2
tion [13] for spin s = 0 perturbations expanded to first or- +Wℓ+2 + Wℓ−2 = 0,
der in ã. The coupling to perturbations with indices ℓ ± 1 dr∗2 dr∗2
vanishes for a simple reason: Klein-Gordon perturbations (32)
are polar quantities, and at first order the Laporte-like
selection rule implies that polar perturbations with in- where we have defined the tortoise coordinate via
dex ℓ should couple to axial perturbations with ℓ ± 1, dr/dr∗ ≡ f = ∆/(r2 + a2 ) (expanded at second order)
but the latter are absent in the spin-0 case. At second and V , U and W are some potentials, whose explicit form
order, perturbations with harmonic index ℓ are coupled is not needed here.
to perturbations with the same parity and ℓ ± 2, but this Note that the coupling to the ℓ±2 terms is proportional
coupling does not contribute to the eigenfrequencies for to ã2 . For a calculation accurate to second order in ã the
the reasons discussed in the previous section. Therefore terms in parenthesis can be evaluated at zeroth order,
(0)
we must solve a single scalar equation for given values of and therefore the functions Ψℓ±2 must be solutions of
ℓ and m, that we write schematically as
(0)
d2 Ψℓ±2 (0) (0)
2
Pℓ + ãmP̄ℓ + ã P̂ℓ = 0 . (25) + Vℓ±2 Ψℓ±2 = 0 . (33)
dr∗2
In order to confirm this general result, let us separate By substituting these relations in Eq. (32) we get
the angular part of Eq. (22). This can be achieved by
using the identities [17] d2 Ψℓ 
(0) (0) (0)

(0)
2
+ Vℓ Ψℓ +ã2 Uℓ+2 − Vℓ+2 Wℓ+2 Ψℓ+2
dr∗
cos ϑY ℓ = Qℓ+1 Y ℓ+1 + Qℓ Y ℓ−1 , (26)  
(0) (0) (0) (0)
sin ϑ∂ϑ Y ℓ = Qℓ+1 ℓY ℓ+1 − Qℓ (ℓ + 1)Y ℓ−1 , (27) +ã2 Uℓ−2 − Vℓ−2 Wℓ−2 Ψℓ−2 = 0 .(34)
cos2 ϑY ℓ = Q2ℓ+1 + Q2ℓ Y ℓ

Finally, making use of the expressions for V , U and W ,
+Qℓ+1 Qℓ+2 Y ℓ+2 + Qℓ Qℓ−1 Y ℓ−2 , the field equations can be reduced to
(28) d2 Ψℓ ã2 M 2 (r − 2M ) µ2 − ω 2

cos ϑ sin ϑ∂ϑ Y ℓ = ℓQ2ℓ+1 − (ℓ + 1)Q2ℓ Y ℓ
 + Vℓ Ψℓ =
dr∗2 r3
+Qℓ+1 Qℓ+2 ℓY ℓ+2
h i
(0) (0)
× Qℓ+1 Qℓ+2 Ψℓ+2 + Qℓ−1 Qℓ Ψℓ−2 , (35)
−Qℓ Qℓ−1 (ℓ + 1)Y ℓ−2 , (29)
where the potential is given by
as well as the orthogonality property of scalar spherical   
harmonics: 2 2M ℓ(ℓ + 1) 2M 2
Vℓ = ω − 1 − + 3 +µ
Z
′ ′
r r2 r
Y ℓ Y ∗ ℓ dΩ = δ ℓℓ . (30) 4ãmωM 2

r3
The result reads, schematically, ã M 2 
2
−24M 2 − 4M r ℓ(ℓ + 1) − 3 + r2 µ2

+ 6
Aℓ + (Q2ℓ+1 + Q2ℓ )Dℓ r
+2M r3 ω 2 + r2 ℓ(ℓ + 1) + m2 + r2 (µ2 − ω 2 ) − 1

+Qℓ−1 Qℓ Dℓ−2 + Qℓ+2 Qℓ+1 Dℓ+2 = 0 . (31)
−r3 (r − 2M )(µ2 − ω 2 ) Q2ℓ + Q2ℓ+1 .

(36)
By repeated use of the identity (26) we can separate
the perturbation equations at any order in ã. Indeed, As we previously discussed, the couplings to terms with
because of the expansion in ã, only combinations of the indices ℓ ± 2 can be neglected in the calculation of the
8

modes. In the scalar case this can be shown explicitly as where mv is the mass of the vector field. Note that, as a
follows. If we define consequence of Eq. (41), the Lorenz condition ∇µ Aµ = 0
is automatically satisfied, i.e. in the massive case there is
Zℓ = ψℓ − ã2 [cℓ+2 ψℓ+2 − cℓ ψℓ−2 ] , (37) no gauge freedom and the field Aµ propagates 2s + 1 = 3
degrees of freedom [16].
where In the µ = 0 case, Teukolsky showed that the equa-
M 2 µ2 − ω 2 Qℓ−1 Qℓ
 tions for spin-1 perturbations around a Kerr BH are sep-
cℓ = , (38) arable [13], the angular part being described by spin-1
2(2ℓ − 1) spheroidal harmonics. In the case of a massive field the
then, at second order in rotation, Eq. (35) can be written separation does not appear to be possible, and one is left
as a single equation for Zℓ : with a set of coupled partial differential equations. To
avoid these difficulties we shall consider the slow-rotation
d2 Zℓ limit of the Kerr metric and apply the approach described
+ Vℓ Zℓ = 0 , (39) in Sec. II. The procedure is equivalent in spirit to that
dr∗2
described for scalar perturbations, but it is more involved
which can be solved by standard methods. This equation due to the spin-1 nature of the vector field.
coincides with Teukolsky’s master equation [13] for spin
s = 0 perturbations expanded at second order in ã. This
is a nontrivial consistency check for the slow-rotation ex- A. Harmonic expansion of the Proca equation on a
slowly rotating background
pansion. In particular, the coefficients Qℓ in Eq. (36)
agree with an expansion of Teukolsky’s spheroidal eigen-
values to second order in ã [64]. In fact, by extending Following the notation of [63, 65] we set xµ = (t, r, xb )
our procedure to arbitrary order in ã we can reconstruct with xb = (ϑ, ϕ). We also introduce the metric of the
the Teukolsky scalar potential order by order. This can two-sphere γab = diag(1, sin2 ϑ).
be viewed as an independent check of the standard pro- Any vector field can be decomposed in a set of vector
cedure, which consists of expanding the angular equation spherical harmonics [65]
to obtain the angular eigenfrequencies (see [64] and ref- Ybℓ = ∂ϑ Y ℓ , ∂ϕ Y ℓ ,

erences therein).  
In addition, as discussed in the previous section, the 1
Sℓb = ∂ϕ Y ℓ , − sin ϑ∂ϑ Y ℓ , (42)
fact that neglecting the ℓ ± 2 couplings is equivalent to a sin ϑ
field redefinition implies that the entire BH spectrum of where Y ℓ (ϑ, ϕ) are the scalar spherical harmonics. We
(massive scalar) QNMs and bound states can be found expand the electromagnetic potential as follows [16]:
by neglecting those couplings.  
ℓ ℓ
Finally, the near-horizon behavior of Eq. (39) reads X  u(1) Y /r 
 
X 0
0 ℓ ℓ
δAµ (t, r, ϑ, ϕ) =  +  u(2) Y /(rf )  ,
Zℓ ∼ e−i kH r∗ , (40) ℓ ℓ
u(4) Sb /Λ
ℓ,m ℓ,m uℓ(3) Ybℓ /Λ
where kH = ω − mΩH and ΩH ∼ ã/(4M ) + O(ã3 ). Note (43)
that, by virtue of the second order expansion, we get where Λ = ℓ(ℓ + 1). Because of their transformation
precisely Vℓ ∼ kH 2
close to the horizon. properties under parity, the functions uℓ(i) belong to the
The possibility to obtain a single equation for any given polar sector when i = 1, 2, 3, and to the axial sector
ℓ and m is a special feature of scalar perturbations. The when i = 4. In the nonrotating case the two sectors are
underlying reason is that scalar perturbations have defi- decoupled [16]. The Proca equation (41), linearized in
nite parity, so the mixing between perturbations of differ- the perturbations uℓ(i) (i = 1, 2, 3, 4), can be written in
ent parity cannot occur. As we show in the next section, the following form:
this property does not necessarily hold for perturbations 
(I) (I) (I)

of higher spin. δΠI ≡ Aℓ + Ãℓ cos ϑ + Dℓ cos2 ϑ Y ℓ
 
(I) (I)
+ Bℓ + B̃ℓ cos ϑ sin ϑ∂ϑ Y ℓ = 0 , (44)
IV. MASSIVE VECTOR PERTURBATIONS OF
Yℓ
δΠϑ ≡ αℓ + ρℓ sin2 ϑ ∂ϑ Y ℓ − i mβℓ

SLOWLY ROTATING KERR BLACK HOLES
sin ϑ

The field equations of a massive vector field (also + (ηℓ + σℓ cos ϑ) sin ϑY = 0 , (45)
known as Proca’s equation) read δΠϕ Yℓ
≡ βℓ + γℓ sin2 ϑ ∂ϑ Y ℓ + i mαℓ

sin ϑ sin ϑ
Πν ≡ ∇σ F σν − µ2 Aν = 0 , (41) + (ζℓ + λℓ cos ϑ) sin ϑY ℓ = 0 , (46)
where Fµν = ∂µ Aν −∂ν Aµ and Aµ is the vector potential. where a sum over (ℓ, m) is implicit and I denotes either
Maxwell’s equations are recovered when µ = mv /~ = 0, the t component or the r component. The various radial
9

coefficients in the equations above are given in terms of radial equations:


the perturbation functions uℓ(i) in Appendix A 2. h i h i
(I) (I) (I) (I) (I)
Aℓ + Q2ℓ+1 Dℓ + ℓB̃ℓ + Q2ℓ Dℓ − (ℓ + 1)B̃ℓ +
The perturbation equations (44)–(46) can be simpli- h i h i
(I) (I) (I) (I)
fied by using the Lorenz identity ∇µ Aµ = 0. To second Qℓ Ãℓ−1 + (ℓ − 1)Bℓ−1 + Qℓ+1 Ãℓ+1 − (ℓ + 2)Bℓ+1 +
order in rotation, and for the background metric (18), h i
(I) (I)
this condition reads Qℓ−1 Qℓ Dℓ−2 + (ℓ − 2)B̃ℓ−2
h i
(I) (I)
+Qℓ+2 Qℓ+1 Dℓ+2 − (ℓ + 3)B̃ℓ+2 = 0 , (50)

(2) (2) (2)
 Λαℓ − i mζℓ + Q2ℓ+1 ℓ [ℓρℓ + σℓ ] + Q2ℓ (ℓ + 1) [(ℓ + 1)ρℓ − σℓ ]
δΠL ≡ Aℓ + Ãℓ cos ϑ + Dℓ cos2 ϑ Y ℓ
  +Qℓ [−(ℓ + 1)ηℓ−1 − i m((ℓ − 1)γℓ−1 + λℓ−1 )]
(2) (2)
+ Bℓ + B̃ℓ cos ϑ sin ϑ∂ϑ Y ℓ = 0 , (47) +Qℓ+1 [ℓηℓ+1 + i m((ℓ + 2)γℓ+1 − λℓ+1 )]
−Qℓ−1 Qℓ (ℓ + 1) [(ℓ − 2)ρℓ−2 + σℓ−2 ]
+Qℓ+2 Qℓ+1 ℓ [−(ℓ + 3)ρℓ+2 + σℓ+2 ] = 0 , (51)
Λβℓ + i mηℓ + Q2ℓ+1 ℓ [ℓγℓ + λℓ ] + Q2ℓ (ℓ + 1) [(ℓ + 1)γℓ − λℓ ]
where the various coefficients are again listed in Ap-
pendix A 2. +Qℓ [−(ℓ + 1)ζℓ−1 + i m((ℓ − 1)ρℓ−1 + σℓ−1 )]
+Qℓ+1 [ℓζℓ+1 − i m((ℓ + 2)ρℓ+1 − σℓ+1 )]
Each of the coefficients in Eqs. (44)–(47) is a linear −Qℓ−1 Qℓ (ℓ + 1) [(ℓ − 2)γℓ−2 + λℓ−2 ]
combination of perturbation functions with either polar
or axial parity (cf. Appendix A 2). Therefore we can +Qℓ+2 Qℓ+1 ℓ [−(ℓ + 3)γℓ+2 + λℓ+2 ] = 0 . (52)
divide them into two sets:
Note that Eqs. (50)–(52) have exactly the same structure
as Eqs. (4)–(5).

(j)
Polar: Aℓ , αℓ , ζℓ , B̃ℓ , Dℓ , ρℓ , σℓ ,
B. Proca perturbation equations at first order
(j) (j)
Axial: Ãℓ , Bℓ , βℓ , ηℓ , λℓ , γℓ ,
In order to make the equations more tractable, in this
section we focus on the first-order corrections only. The
second-order analysis is presented in Sec. IV C below. At
where j = 0, 1, 2. In order to separate the angular vari- first order, Eqs. (50)–(52) simplify to
ables in Eqs. (44)–(47) we compute the following inte-
grals:
h i
(I) (I) (I)
Aℓ + Qℓ Ãℓ−1 + (ℓ − 1)Bℓ−1
h i
(I) (I)
Z + Qℓ+1 Ãℓ+1 − (ℓ + 2)Bℓ+1 = 0 , (53)
∗ℓ
δΠI Y dΩ , (I = t, r, L) ; (48a) Λαℓ − i mζℓ
Z − Qℓ (ℓ + 1)ηℓ−1 + Qℓ+1 ℓηℓ+1 = 0 , (54)
δΠa Yb∗ ℓ γ ab dΩ , (a , b = ϑ, ϕ) ; (48b) Λβℓ + i mηℓ
− Qℓ (ℓ + 1)ζℓ−1 + Qℓ+1 ℓζℓ+1 = 0 , (55)
Z
δΠa S∗b ℓ γ ab dΩ , (a , b = ϑ, ϕ) . (48c)
where now the coefficients are linear in ã.
To begin with, let us focus on the equations for
monopole perturbations, ℓ = m = 0. The longitudi-
Using the orthogonality properties of scalar and vector nal mode of a massive vector field (unlike the massless
harmonics, Eq. (30), the relations case) is dynamical. Since m = 0, the monopole only ex-
cites axisymmetric modes. In the nonrotating case these
are described by a single equation belonging to the polar
sector [55] (see also [16]); for ℓ = 0, only the first two
Z Z
′ ′ ′
Ybℓ Yb∗ ℓ γ ab dΩ = Sℓb S∗b ℓ γ ab dΩ = Λδ ℓℓ , components u0(1) and u0(2) are defined. However, in the
Z
′ slowly rotating case these components are coupled to the
Ybℓ S∗b ℓ γ ab dΩ = 0 , (49) ℓ = 1 axial component u1(4) through Eq. (53).
When ℓ = m = 0, Q0 = 0 and Eq. (53) reduces to
h i
(I) (I) (I)
as well as the identities (26)–(29), we find the following A0 + Q1 Ã1 − 2B1 = 0 , (56)
10

where I = 0, 1, 2. This is an extreme example of the On the other hand, the axial sector leads to
“propensity rule” discussed in the general case: at first
order the monopole is only coupled with axial perturba- 4ãM 2 mω ℓ
D̂2 uℓ(4) − u(4)
tions with ℓ = 1. Using the explicit form of the coeffi- r3
2
cients A(I) , Ã(I) , B (I) given in Appendix A 2 (truncated 6i ãM F 
(ℓ + 1)Qℓm ψ ℓ−1 − ℓQℓ+1m ψ ℓ+1 (61)

=−
at first order), the equations above can be written as a 5
r ω
single equation for u0(2) :
(see Appendix B for details), where we have defined the
polar function
d2
  
2(r − 3M )
+ ω2 − F + µ2 u0(2)
dr∗2 r3 ℓ
ψ ℓ = Λ + r2 µ2 uℓ(2) − (r − 2M )u′ (3) .

(62)

2i 3ãM 2 ωF 1
= u(4) , (57) Note that this term is similar to Eq.(25) in Ref. [16]. As
r3
expected, the axial perturbation uℓ(4) is coupled to the
where at first order the tortoise coordinate r∗ is the same polar functions with ℓ ± 1. Equations (59), (60) and (61)
as in the Schwarzschild case, and it is defined by dr/dr∗ = describe the massive vector perturbations of a Kerr BH
F . The source term u1(4) is the solution of Eq. (53) with to first order in ã for ℓ > 0. These equations reduce to
ℓ = 1. To first order in ã we can approximate u1(4) (that those in Ref. [16] when ã = 0.
is multiplied by ã in the source term) by its zeroth-order Since the right-hand side of Eq. (61) is proportional to
expansion in ã, which is a solution of ã, in our perturbative framework we can first solve for
uℓ±1 ℓ±1
(2) and u(3) to zeroth order in the rotation rate, and
d2
  
2 2 2 then use these solutions as a source term in Eq. (61). As
+ω −F 2 +µ u1(4) = 0 . (58)
dr∗2 r we proved in Sec. II, the coupling on the right-hand side
of Eq. (61) does not affect the QNMs to first order in rota-
Note that Eq. (58) is precisely the axial perturbation tion. In Sec. VI we shall verify this property numerically.
equation in the nonrotating case for ℓ = 1 [16]. Eqs. (57) Therefore, for the purpose of computing eigenfrequencies
and (58) fully describe the dynamics of the polar ℓ = 0 to first order in ã we can simply consider the following
perturbations in the slow-rotation approximation. As we polar equations:
proved in Sec. II, the coupling on the right-hand side of  
Eq. (57) does not affect the QNMs to first order in rota- ℓ 2F 3M h ℓ i
D̂2 u(2) − 2 1 − u(2) − uℓ(3)
tion. In addition, since for the monopole m = 0, to this r r
order the frequency is the same as in the Schwarzschild 2ãM m2 h
Λ 2r2 ω 2 + 3F 2 uℓ(2)

case, which is extensively discussed in Ref. [16]. = 5ω
Λr
Let us now turn to modes with ℓ > 0. The equations 

i
+3F rΛF u′ (2) − r2 ω 2 + ΛF uℓ(3) ,

for ℓ > 0 at first order in ã are derived in Appendix B (63)
by using the Lorenz condition (B2) in order to eliminate 2F Λ
uℓ(1) . The polar sector is fully described by the system D̂2 uℓ(3) + 2 uℓ(2) =
r
2ãM 2 m h 2 2 ℓ 2 ′ℓ 2 2
 ℓ
i
2r ω u + 3rF u − 3 Λ + r µ F u ,
 
2F 3M h ℓ i
(3) (3) (2)
D̂2 uℓ(2) − 2 1− u(2) − uℓ(3) r5 ω
r r (64)
2
2ãM m h
Λ 2r2 ω 2 + 3F 2 uℓ(2)

= 5 as well as the decoupled axial equation
Λr
 ω i

+3F rΛF u′ (2) − r2 ω 2 + ΛF uℓ(3)

4ãM 2 mω ℓ
D̂2 uℓ(4) − u(4) = 0 . (65)
6i ãM 2 F ω h i r3
ℓ−1 ℓ+1
− (ℓ + 1)Q ℓ u (4) − ℓQ ℓ+1 u (4) , (59)
Λr3 Within our perturbative scheme, any eigenfrequency of
2F Λ Eq. (65) is also a solution of the coupled Eq. (61) as
D̂2 uℓ(3) + 2 uℓ(2) =
r long as uℓ±1 ℓ±1
(2) = u(3) = 0, which is a trivial solution of
2
2ãM m h
2 2 ℓ 2 ′ℓ 2 2
 ℓ
i
Eqs. (59) and (60) for ℓ ± 1 at zeroth order. Further-
2r ω u (3) + 3rF u (3) − 3 Λ + r µ F u (2) ,
r5 ω more, consistently with our argument in Sec. II, we have
(60) checked that there are no other modes to order O(ã) (cf.
Sec. VI).
where here and in the following a prime denotes deriva- Note the similarity between Eq. (65), describing axial
tion with respect to r and the operator D̂2 is defined Proca modes, and Eq. (23) for massive scalar pertur-
in Eq. (24). Note that while Eq. (60) only involves po- bations. Indeed, one can show that the generalization
lar perturbations, in Eq. (59) we also have a coupling to of Eq. (65) to the background metric (2) (i.e. without
uℓ±1
(4) . specializing to the slowly rotating Kerr metric) can be
11

written in a form that includes also massive scalar per- consistent subset of Eqs. (50)–(52) reads
turbations. This “master equation” reads h i
(i) (i) (i)
0 = Aℓ + Q2ℓ+1 Dℓ + ℓB̃ℓ
h i h i
(i) (i) (i) (i)
1

2m̟(r)ω +Q2ℓ Dℓ − (ℓ + 1)B̃ℓ + Qℓ Ãℓ−1 + (ℓ − 1)Bℓ−1
F BΨ′′ℓ + [B ′ F + F ′ B] Ψ′ℓ + ω 2 −
r2
h i
2 (i)
+Qℓ+1 Ãℓ+1 − (ℓ + 2)Bℓ+1 ,
(i)
  ′ ′

Λ B BF
−F 2
+ µ2 + (1 − s2 ) + Ψℓ = 0 , 0 = Λαℓ − i mζℓ + Q2ℓ+1 ℓ [ℓρℓ + σℓ ]
r 2r 2rF
(66) +Q2ℓ (ℓ + 1) [(ℓ + 1)ρℓ + σℓ ]
+Qℓ [−(ℓ + 1)ηℓ−1 − i m((ℓ − 1)γℓ−1 + λℓ−1 )]
+Qℓ+1 [ℓηℓ+1 + i m((ℓ + 2)γℓ+1 − λℓ+1 )] ,
and it can be simplified by introducing a generalized
√ tor- 0 = Λβℓ + i mηℓ + Q2ℓ+1 ℓ [ℓγℓ + λℓ ]
toise coordinate y(r) such that dr/dy = F B. In the
+Q2ℓ (ℓ + 1) [(ℓ + 1)γℓ + λℓ ]
equation above s is the spin of the perturbation (s = 0
for scalar perturbations and s = ±1 for vector perturba- +Qℓ [−(ℓ + 1)ζℓ−1 + i m((ℓ − 1)ρℓ−1 + σℓ−1 )]
tions with axial parity). In the nonrotating case Eq. (66) +Qℓ+1 [ℓζℓ+1 − i m((ℓ + 2)ρℓ+1 − σℓ+1 )] . (69)
is exact; it also includes gravitational perturbations of a
Schwarzschild BH if s = ±2 and F = B = 1 − 2M/r. In As in the scalar case, the second-order coefficients gener-
the slowly rotating case Eq. (66) is a complete descrip- ally contain second derivatives of the perturbation func-
ℓ ℓ±1 ℓ±2
tion of massive scalar perturbations for s = 0, whereas for tions, i.e. u′′ (i) , u′′ (i) and u′′ (i) . Since the coefficients
s = ±1 it describes the axial sector without the ℓ → ℓ ± 1 are already of second order, we can use the perturbation
couplings, i.e. it is a generalization of Eq. (65). equations in the nonrotating limit in order to eliminate
these second derivatives. After some manipulation we get
the final two sets of equations, that can be schematically
written as

DA ΥA ℓ + V′ A ΥA ℓ = 0 , (70)
′ ℓ ′ ℓ
C. Proca perturbation equations at second order D P ΥP + V P ΥP = 0 , (71)
where D′ A,P are second order differential operators,
In this section we briefly present the derivation of the ΥA ℓ = (uℓ(4) , uℓ±1 ℓ±1 ℓ ℓ ℓ ℓ±1
(2) , u(3) ), ΥP = (u(2) , u(3) , u(4) ), and
Proca eigenvalue problem to second order in ã. ′
V A,P are matrices.
By using the Lorenz condition to eliminate the spuri- We remark that for a given value of ℓ and m, ΥA and
ous uℓ(1) mode, Eqs. (50)–(52) can be written as ΥP are five- and four-dimensional vectors, respectively,
while ΨA and ΨP in Eqs. (67) and (68) are seven- and
eight-dimensional vectors, respectively. This is a very
convenient simplification, because large systems of equa-
DA ΨA ℓ + VA ΨA ℓ = 0 , (67) tions are computationally more demanding. Furthermore

DP ΨP + VP ΨP = 0 ,ℓ
(68) the couplings to perturbations with index ℓ − 1 vanish if
ℓ = m, because as usual the factors Qm = 0, and in this
case we are left with two subsystems of dimension three.

where DA,P are second order differential operators, VA,P


are matrices, ΨA ℓ = (uℓ(4) , uℓ±1 ℓ±1 ℓ±2 ℓ
(2) , u(3) , u(4) ) and ΨP = V. THE EIGENVALUE PROBLEM FOR
QUASINORMAL MODES AND BOUND STATES
(uℓ(2) , uℓ(3) , uℓ±1 ℓ±2 ℓ±2
(4) , u(2) , u(3) ).
The explicit form of the
equations above is quite lengthy; therefore we do not
show it in this article, but we make it available online [61]. One of the key advantages of the slow-rotation approx-
It can be obtained using the procedure explained above imation is that the perturbation equations can be solved
and the coefficients listed in Appendix A 2. The func- using well-known numerical approaches. The integration
tion uℓ(1) can be obtained from the Lorenz condition once proceeds exactly in the same way as in the nonrotating
case. For example, for Proca perturbations of a slowly ro-
the three dynamical degrees of freedom are known. Note
tating Kerr BH we can compute the characteristic modes
that Eqs. (70)-(71) are particular cases of Eqs. (12)-(14)
by following a procedure similar to the Schwarzschild case
and Eqs. (15)-(17), respectively.
discussed in Ref. [16], as follows.
If we are interested in the eigenfrequencies up to sec- The perturbation equations (70)–(71), together with
ond order in ã we can drop the couplings to ℓ ± 2 per- appropriate boundary conditions at the horizon and at
turbations, for reasons discussed in Sec. II. Therefore, a infinity, form an eigenvalue problem for the frequency
12
p
spectrum. In the near-horizon limit, DA,P → d2 /dr∗2 where k∞ = µ2 − ω 2 , so that Re[k∞ ] > 0. The bound-
and VA,P → (ω − mΩH )2 , so that at the horizon we ary conditions B(i) = 0 yield purely outgoing waves at
impose purely ingoing-wave boundary conditions: infinity, i.e. QNMs [4]. If instead we impose C(i) = 0 we
get states that are spatially localized within the vicinity
uℓ(i) ∼ uℓ(i)H e−i kH r∗ , (72) of the BH and decay exponentially at spatial infinity, i.e.
bound states (see e.g. [16, 35]). Stable QNMs are more
where uℓ(i)H is a constant and challenging to compute than bound states. In the former
case the boundary conditions above imply that purely
mã outgoing waves blow up as r∗ → ∞, whereas purely in-
kH = ω − mΩH = ω − + O(ã3 ) . (73) going waves are exponentially suppressed at infinity. For
4M
bound states, direct integration of the equations com-
We have introduced the horizon angular frequency ΩH = bined with a shooting method is sufficient, but QNMs
a/(2M r+ ) and expanded it to second order. are more efficiently computed by other means, for exam-
In the massless case, due to the boundary condi- ple via continued fraction methods [4].
tion (72), superradiant scattering for scalar, electromag-
netic and gravitational perturbations is possible when
ωR < mΩH [66], i.e. (to second order in rotation) when
A. On the superradiant regime and the second
order expansion
4M ωR
ã > , (74)
m
Here we comment on some important features that
where ωR is the real part of the mode frequency, i.e. would be missed in a first order treatment. First, at
ω = ωR + i ωI . Superradiance is possible because the second order the structure of the background metric is re-
energy flux at the horizon markably different, because all metric coefficients acquire
O(ã2 ) corrections. This affects the location of the event

Z
Ėr+ ≡ lim dϑdϕ −gTtr (75) horizon, of the ergosphere and of the inner Cauchy hori-
r→r+ zon [cf. Eq. (19)]. Second-order corrections are known to
approximate the Kerr solution much better than a first-
is negative when kH < 0. In the equation above Tµν is order expansion (see e.g. [67]).
the stress-energy tensor of the perturbation. The same
From a dynamical point of view, axisymmetric modes
argument can be applied to massive scalar perturbations.
acquire second-order corrections that break the m = 0
The case of massive vector perturbations is more in-
degeneracy of the first-order case. Most importantly, at
volved. For purely axial perturbations close to the hori-
second order the superradiant regime of vector and scalar
zon, by using Eq. (75) and the stress-energy tensor of a
fields can be described within our perturbative approach
Proca field we get Ėr+ = ℓm Ėrℓ+ with
P
in a self-consistent way. The superradiant condition (74)
may look like a first-order effect. However, at the onset
ωkH of superradiance ωM ∼ ã (at least in the most relevant
Ėrℓ+ = |uℓ |2 + O(ã3 ) , (76)
4ℓ(ℓ + 1)M 6 (4)H cases, i.e. when m is of order unity). In this case, terms
like ω 2 in the field equations (which are crucial, in partic-
which shows that the energy flux across the horizon is ular in the study of the mode spectrum) are of the same
negative when kH < 0. The general case involves terms order of magnitude as second order quantities.
proportional to each component uℓ(i)H , as well as terms
The field equations are of second differential order, so
proportional to µ2 . the linearized field equations (at first order in ã) contain
As we discuss in Sec. VI below, our results turn out at most terms of order
to be very accurate for moderate values of ã and they
are reliable even in the superradiant regime, defined by ωM , (ωM )2 , ã , ãωM . (78)
Eq. (74), as long as ωR M ≪ 1. Superradiant scat-
tering leads to instabilities for massive scalar perturba- In principle, terms of order ã(ωM )2 would also be al-
tions [35, 60]. In the Proca case, the numerical results lowed, but those do not appear in the linearized Proca
discussed in Sec. VI below show that, when the super- equations (50)–(52). In the massless case this is consis-
radiant condition (74) is met, the imaginary part of the tent with a first-order expansion of the Teukolsky equa-
modes crosses zero. Thus our numerical data show hard tion, which does not contain any ã(ωM )2 term for scalar,
evidence, for the first time, that massive vector fields vector and tensor perturbations.
trigger (as expected) a superradiant instability. Thus, if ωM ∼ ã the second and fourth terms listed
The asymptotic behavior of the solution at infinity above would be as large as second-order terms in ã, which
reads are neglected in a first-order expansion. In a second-order
M (µ2 −2ω2 ) M (µ2 −2ω2 ) expansion, instead, one also keeps terms proportional to
uℓ(i) ∼ B(i) e−k∞ r r− k∞ + C(i) ek∞ r r k∞ , ã2 . This would be enough to consistently describe the
(77) superradiant regime up to first order in ã, i.e. up to terms
13

(2) (3)
∼ ãω and ∼ ω 2 . For a related discussion in the case of where Un = (an , an ) is a two-dimensional vectorial
neutron star r-modes, we refer the reader to Ref. [68]. coefficient and αn , βn , γn , δn , ρn and σn are 2 × 2
matrices, whose explicit form we do not present here for
brevity but it is available online [61]. By using a matrix-
B. Continued fraction method for quasinormal valued Gaussian elimination [4, 69, 70] the system above
modes and bound states can be reduced to a three-term matrix-valued recurrence
relation, which can be solved with the method discussed
From a conceptual point of view, numerical calcula- in Ref. [16].
tions in the slow-rotation approximation are not more The continued fraction method works very well for
complicated than in the nonrotating case. For example, both QNMs and bound-state modes. We computed
in order to apply the continued fraction method we can QNMs and checked that they yield the correct limit in
write down a recurrence relation similar to Eqs. (31) and the massless case (see Sec. VI B below), but in the fol-
(32) of Ref. [16], starting from Eqs. (59), (60) and (65), lowing we will focus mainly on bound states, that can
by imposing the ansatz become unstable in the superradiant regime.
X
uℓ(i) = (r − r+ )−2ikH rν e−qr an(i) (r − r+ )n , (79)
n C. Direct integration and Breit-Wigner resonance
2
method for bound states
where ν = −q + ω /q + 2ikH and q = ±k∞ for bound
states and for QNMs, respectively. For the axial equation
this ansatz leads to a three-term recurrence relation of To compute bound state frequencies it is possible to use
the form either a direct integration method or the Breit-Wigner
resonance method. We start with a series expansion of
(4) (4)
α0 a1 + β0 a0 = 0 , the solution close to the horizon:
(4) (4)
αn an+1 + βn a(4)
n + γn an−1 = 0 , n > 0,
X
uℓ(i) ∼ e−i kH r∗ b(i) n
n (r − r+ ) , (83)
whose coefficients (to first order in ã, and setting M = 1 n

in these equations only) read where r+ is expanded to second order and the coefficients
(i) (i)
αn = −4(1 + n)q 2 (1 + n − 4i ω) − 4i ãm(1 + n)q 2 (, 80) bn (n > 0) can be computed in terms of b0 by solving
βn = 4q q(1 + ℓ(ℓ + 1) + 2n2 + q(3 + 4q) the near-horizon equations order by order. In the direct
integration method, the field equations are integrated
+n(2 + 6q) − s2 − 4i q(1 + 2n + 3q)ω

outwards up to infinity, where the condition C(i) = 0
−(1 + 2n + 12q)ω 2 + 4i ω 3 ) in Eq. (77) is imposed (see Ref. [16] for details).
+4i ãmq q + 2nq + 3q 2 − 2i qω − ω 2 ,

(81) The Breit-Wigner resonance method, also known as
the standing-wave approach [63, 71–73], is well suited
γn = −4 (q(n + q) − 2i qω − ω 2 )2 − s2 q 2
 
to computing QNMs and bound states of the system of
−4i ãmq q(n + q) − 2i qω − ω 2 , equations (59)–(61) in the case of slowly damped modes,


(82) i.e. those with ωI ≪ ωR . In this case the eigenvalue


problem can be solved by looking for minima of a real-
for s = ±1. If we set s = 0, the equations above are valued function of a real variable [72]. We briefly explain
also valid for massive scalar perturbations of a Kerr BH the procedure below, where we extend it to deal with a
in the slowly rotating limit because, as we have shown system of coupled equations. For clarity we only consider
above, there exists a single master equation describing first-order corrections, but our argument applies also to
both massive scalar and axial vector modes [cf. Eq. (66)]. the second-order case described by Eqs. (70)–(71) and,
We have investigated the massive scalar case to test the in principle, to any order.
robustness of our results, because in this case the per- Since ℓ = 0, 1, 2, .., the full system (59), (60) and (61)
turbation equations on a generic Kerr background are formally contains an infinite number of equations. In
separable [60] and the eigenvalues can be computed by a practice, we can truncate it at some given value of ℓ,
direct solution of the Teukolsky equation [4, 35]. compute the modes as explained below, and finally check
In the slowly rotating case the polar sector leads to a convergence by increasing the truncation order. Let us
six-term, matrix-valued [16] recurrence relation suppose we truncate the axial sector at ℓ = L and the
polar sector at ℓ = L + 1, i.e. for a given m we assume
α0 U1 + β0 U0 = 0 ,
α1 U2 + β1 U1 + γ1 U0 = 0 , n > 0, uℓ(4) ≡ 0 , uℓ+1
(j) ≡ 0 when ℓ ≥ L, (84)
α2 U3 + β2 U2 + γ2 U1 + δ2 U0 = 0 , n > 1,
with j = 1, 2, 3 denoting the polar perturbations. All
α3 U4 + β3 U3 + γ3 U2 + δ3 U1 + ρ3 U0 = 0 , n > 2,
perturbations vanish identically for ℓ < |m|.
αn Un+1 + βn Un + γn Un−1 + δn Un−2 When m = 0, the truncation above reduces the sys-
+ρn Un−3 + σn Un−4 = 0 , n > 3, tem to N = 3L coupled second-order ODEs for L − 1
14

axial functions and 2L + 1 polar functions, including the Therefore, on the real–ω axis, close to the real part of
monopole, described by Eq. (57). When |m| > 0 the the mode,
truncated system contains N = 3L − 3|m| + 2 second- 2
order ODEs (for L−|m| axial functions and 2L−2|m|+2 | det Sm (ω)|2 ∝ (ω − ωR ) + ωI2 . (90)
polar functions). In all cases we are left with a system
of N second-order ODEs for N perturbation functions, To summarize, to find the slowly damped modes it is suf-
which we collectively denote by y(p) (p = 1, ..., N ). ficient to integrate the truncated system N times for real
At the horizon each function is described by ingoing values of the frequency ω, construct the matrix Sm (ω)
and outgoing waves. We impose a purely ingoing wave and find the minima of the function | det Sm |2 , which rep-
boundary condition analogous to Eq. (83), resent the real part of the modes. Then the imaginary
part (in modulus) of the mode can be extracted through
a quadratic fit, as in Eq. (90). We note that the same
X
y(p) ∼ e−i kH r∗ c(p) n
n (r − r+ ) , (85)
n
method can be straightforwardly extended to compute
the slowly damped modes of the general systems (12)–
(p) (14) and (15)–(17).
where again the coefficients cn (n > 0) can be com-
(p)
puted in terms of c0 . A family of solutions at infinity
M ωR M ωI
is then characterized by N parameters, corresponding to
the N -dimensional vector of the near-horizon coefficients, No coupling, ℓ = 1 (DI) 0.099484532 1.1 · 10−7
(p) Full system, L = 1, 2, 3, 4 (BW) 0.099484563 1.1 · 10−7
c0 = {c0 }. At infinity we look for exponentially decay-
ing solutions, which correspond to bound, slowly damped No coupling, ℓ = 1 (DI) 0.09987320753 4 · 10−11
modes. The spectrum of these modes can be obtained as Full system, L = 2, 3, 4 (BW) 0.09987183250 5 · 10−11
follows. We first choose a suitable orthogonal basis for
(p) TABLE I. Examples of polar (top rows) and axial (bottom
the N -dimensional space of the initial coefficients c0 rows) bound-state modes for m = 1, M µ = 0.1 and ã = 0.1
(see also Ref. [16]). We perform N integrations from the computed at first order. Modes were computed via direct
horizon to infinity and construct the N × N matrix integration (DI) of the system (59)–(60) without couplings
 (1) (2) (N )  ℓ → ℓ ± 1, as well as with the Breit-Wigner (BW) method
y(1) y(1) ... y(1) applied to the full system (59)–(60) and (61), for different
 (1) (2) truncation orders L. The modes are insensitive to the trun-
 y(2) y(2) ... ... 

Sm (ω) = lim  ...

... ... ... 

, (86) cation order within the quoted numerical accuracy.
r→∞ 
 ... ... ... ... 

(1) (N ) By applying the Breit-Wigner method we have numer-
y(N ) ... ... y(N )
ically verified the argument given in Sec. II, i.e. that the
where the superscripts denote a particular vector of the ℓ → ℓ ± 1 couplings do not affect the eigenfrequencies
(1) (2)
basis, i.e. y(p) corresponds to c0 = {1, 0, 0, ..., 0}, y(p) in the slow-rotation limit. As an example, in Table I
(N )
we compare two modes (m = 1, ã = 0.1, M µ = 0.1)
corresponds to c0 = {0, 1, 0, ..., 0} and y(p) corresponds of the full system computed with the Breit-Wigner res-
to c0 = {0, 0, 0, ..., 1}. Finally, the bound-state frequency onance method for several truncation orders L and the
ω0 = ωR + i ωI is obtained by imposing same modes computed with a direct integration of the
system of equations without the ℓ → ℓ ± 1 couplings. Up
det Sm (ω0 ) = 0 . (87) to numerical accuracy the mode is insensitive to the trun-
So far we have not imposed the Breit-Wigner assumption cation order and, most importantly, it agrees very well
ωI ≪ ωR . In fact, the procedure discussed above can be with that computed for the system without couplings.
used to perform a general direct integration in the case Note that L = 1 corresponds to the uncoupled polar sys-
of coupled systems. tem with ℓ = 1. Therefore the small discrepancy is not
By expanding Eq. (87) about ωR and assuming ωI ≪ due to the coupling terms, but to some inherent numer-
ωR we get [63] ical error of the resonance method, which becomes less
accurate when the imaginary part of the mode is tiny.
d [det Sm (ω)]
det Sm (ω0 ) ≃ det Sm (ωR ) + i ωI = 0,
dω ωR VI. RESULTS
(88)
which gives a relation between d [det Sm (ω)]/dω and
det Sm at ω = ωR . We consider the function det Sm A. Numerical procedure
restricted to real values of ω. A Taylor expansion for
(real) ω close to ωR yields (using the relation above): In principle, by integrating the full systems (70) and
  (71) for a given truncation order L and a given value of m
ω − ωR one can obtain the full spectrum of quasinormal modes
det Sm (ω) ≃ det Sm (ωR ) 1 − ∝ ω − ωR − i ωI .
i ωI and bound modes for both the axial and polar sectors
(89) and for any ℓ < L. We have computed bound states
15

1
and QNMs via the Breit-Wigner procedure of Sec. V C. 10
l=1
We double-checked the results using two additional, inde- ωR, m=1
0 1% error
pendent techniques, also described above: the continued 10 ωI, m=1
fraction method and direct integration of the reduced ωR, m=0
equations (70) and (71). The results agree within nu- 10
-1
ωI, m=0
merical accuracy. The Breit-Wigner procedure and the ωR, m=-1

∆ [%]
continued fraction method have been implemented up to -2 ωI, m=-1
10
first order in the rotation parameter; the direct integra-
tion, instead, has been extended up to second order in -3
ã, in order to validate the results of the first-order inte- 10
gration, as we shall discuss below. Furthermore, when -4
ωI ≪ ωR we found very good agreement with the Breit- 10
Wigner method (cf. Table I).
-5
Exploring the full parameter space of the eigenfrequen- 10 0.01 0.1
cies at second order in ã is numerically demanding. For J/M
2

this reason, in Sec. VI D below we will present a more


extensive survey of results at first order, and compare FIG. 2. (color online) Percentage difference between QNMs
them to the second-order calculation in Sec. VI E in se- of massless vector perturbations in the slow-rotation limit at
lected cases. first order and the “full” numerical solution of the Teukolsky
equation in the Kerr metric [4]. The deviation scales like ã2
as ã → 0.
B. A consistency check: massless vector
perturbations
QNMs computed with our approach deviate by less than
As a preliminary test of our method, we have com- 1% from their exact values if ã . 0.3.
puted the QNMs of massless vector (i.e., electromagnetic)
perturbations of a Kerr BH to first order in ã. These re- -11
sults can be compared with those obtained by solving the 10
Teukolsky equation without imposing the slow-rotation
approximation (see e.g. [4]). In the massless limit the
axial equation (65) reduces to -12
10
|MωI|

d2 u(4) 4ãM 2 mω ℓ(ℓ + 1)


 
2
+ ω − − F u(4) = 0 . (91)
dr∗2 r3 r2 l=m=1 Mµ=0.1
-13 Exact, Teukolsky
10
For the polar sector in the massless limit, we can define 1st order
exactly the same master variable as in the nonrotating 2nd order
3rd order
case [74]. The polar sector in the massless limit is de- 4th order
scribed by a fourth-order equation. As in the nonrotat- -14
10 0
ing case [16], one solution of this equation is a pure-gauge 0.1 0.2 0.3 0.4 0.5 0.6 0.7
2
mode and can be eliminated. In the slow-rotation ap- J/M
proximation we can recast the other solution in terms of
FIG. 3. (color online) Comparison between the exact,
a master function, which also satisfies Eq. (91). There-
Teukolsky-based result and the results obtained by our slowly
fore axial and polar perturbations have the same spectra. rotating approximation at first and second order for the imag-
This is consistent with the fact that electromagnetic per- inary part of the scalar fundamental bound-state mode with
turbations of the full Kerr geometry are described by a ℓ = m = 1 and µM = 0.1. For comparison, we have also com-
single master equation in the Teukolsky formalism [4]. puted the same mode as obtained by expanding the Teukolsky
Due to isospectrality, in the massless limit we only need equation at third and fourth order.
to solve Eq. (91). The modes can be computed via the
continued fraction method introduced above, where the
coefficients of the recurrence relation can be obtained by
setting µ = 0 and s = ±1 in Eqs. (80)–(82).
We compared our results with the exact massless vec- C. A second test: bound state modes for scalar
tor modes of a Kerr BH, computed by solving the Teukol- perturbations
sky equation, for ℓ = 1 and different values of m (see
Fig. 2). The first order approximation performs very We can also investigate the accuracy of the slow-
well, even for relatively large values of the BH rotation rotation approximation for massive scalar perturbations,
rate. Remarkably, the real and imaginary parts of the another case in which the Teukolsky equation can be
16

solved exactly (see e.g. [35]). In Fig. 3 we show the imag- ℓ = 0, S = 1, in agreement with the rules for addition of
inary part of the bound state modes with ℓ = m = 1 angular momenta [16]. In the limit ã = 0 we recover this
and M µ = 0.1, computed by solving Eq. (39) at first scaling. We have also analyzed the ã-dependence of the
and second order via direct integration, against the ex- modes for small ã. When M µ ≪ 1 we find the expected
act Teukolsky-based result obtained with the continued hydrogen-like behavior, ωR ∼ µ, which remains valid also
fraction method [4, 35]. For comparison, we also show in the slowly rotating case. When M µ . 0.1 the real part
the results obtained by solving Teukolsky’s equation at of the modes is roughly independent of ã, and it is very
third and fourth order. As shown in Fig. 3, the imaginary well approximated by the relation
part crosses the axis when the superradiant condition is "  2 #
satisfied. In the stable branch (ã . 0.4) even first-order 2 2 Mµ
+ O µ4 ,

results are in good quantitative agreement with the “ex- ωR = µ 1 − (93)
ℓ+n+S +1
act” calculation. In the unstable branch (ã & 0.4) the
first-order approximation is in only qualitative agreement where n ≥ 0 is the overtone number (cf. [35]). For the
with the exact result, but the second-order approxima- axial case (S = 0) this relation is validated by the ana-
tion is in quantitative agreement with the numerics at lytical results presented in Appendix C [see in particular
the onset of the instability. It is still quite remarkable Eq. (C9)], where we solve the axial equations in the limit
that even first-order results can correctly reproduce the M µ ≪ 1. Equation (93) is also supported by the nonro-
onset of the instability. We shall return to this consider- tating result given in Eq. (49) of [16] (see also [53]).
ation below, when we will deal with the massive vector Equation (93) predicts a degeneracy for modes with
case. the same value of ℓ + n + S when M µ ≪ 1. In the Breit-
Wigner method, the mode frequencies can be identified
4
l+n+S=1
as minima of the real-valued function | det Sm |2 . The de-
30 3
generacy predicted by Eq. (93) is not exact for M µ = 0.1,
2 as illustrated in the inset of Fig. 4, where we display the
20 l+n+S=2 1 minima of | det Sm |2 when ã = 0.1. The first minimum
log10(|det Sm|)

0 on the right corresponds to ℓ + n + S = 0, which can


10 -1 only be achieved for the fundamental polar mode with
1.23×10
-4 -4
1.32×10
(ℓ, n, S) = (1, 0, −1). When ℓ+n+S = 1 we have a three-
0 fold degeneracy, corresponding to (ℓ, n, S) = (1, 0, 0),
l+n+S=1 (2, 0, −1), (1, 1, −1). The approximate nature of the de-
generacy is shown in the inset, where three distinct (al-
-10
l+n+S=0
beit very close) minima appear. For ℓ + n + S = 2 there
is a five-fold degeneracy, which can be resolved with high
-20 -4 -3 enough resolution. Note that according to Eq. (95) the
10 10
M(µ- Re[ω]) imaginary part of the modes is tiny when ℓ+n+S is large.
This makes it difficult to numerically resolve higher over-
FIG. 4. (color online) Bound state modes obtained with a tones and modes with large ℓ.
first-order Breit-Wigner method applied to the full system. The imaginary part of nonaxisymmetric modes shows
We show the determinant | det Sm | as a function of the real the typical Zeeman-like splitting for different values of
part of the frequency for M µ = 0.1 and ã = 0.1. According m when ã 6= 0, as shown in Fig. 5 for the axial modes
to Eq. (93), the real part of modes with the same ℓ + n + S is and for the polar mode with S = −1. For m > 0 the
approximately degenerate for M µ ≪ 1. imaginary part of the frequency decreases (in modulus)
as ã increases, and it has a zero crossing when

ωR ∼ µ = mΩH ∼ m + O(ã3 ) , (94)
4M
D. Proca modes to first order
which according to Eq. (74) corresponds to the onset of
the superradiant regime. Recall that a second-order cal-
Unlike the massless case, axial and polar modes for
culation is needed to describe the superradiant regime in
massive vector perturbations are not isospectral. Fur-
a self-consistent way, because the latter is well beyond
thermore there exist two classes of polar modes, which
the nominal regime of validity of the first-order approx-
can be distinguished by their “polarization” [16]. For
imation. It is quite remarkable that even the first-order
small M µ, Rosa and Dolan found that the imaginary
approximation predicts that the instability should “turn
part of the bound states of a Schwarzschild BH scales as
on” at the right point, as shown in Fig. 6 for ℓ = m = 1
ωI ∼ µ(M µ)4ℓ+5+2S , (92) and several values of µ. The quantitative accuracy of the
first-order approximation is questionable in the superra-
where S = 0 for axial modes and S = ±1 for the two diant regime. However, we will now show that second-
classes of polar modes. The monopole corresponds to order results indicate that we are correctly capturing the
17

-13
10

-9
10
-14
10
|MωI|

|MωI|
-10
-15 10
10 Mµ=0.05 l=1
m=1 Mµ=0.05 l=1
m=0 stable unstable
stable unstable m=1
m=-1 m=0
m=1, analytical m=-1
-16 -11
10 0 0.05 0.1 0.15 0.2 0.25 10 0 0.05 0.1 0.15 0.2 0.25
2 2
J/M J/M

FIG. 5. (color online) Absolute value of the imaginary part of the axial (left panel) and of the polar S = −1 (right panel)
vector modes as a function of the BH rotation rate ã for ℓ = 1, M µ = 0.05 and different values of m, computed at first order.
For m = 1, the modes cross the axis and become unstable when the superradiance condition (94) is met. In the left panel, the
red dot-dashed line denotes the analytical result (C12).

order of magnitude of the instability even for moderately the superradiant regime as long as M ω ≪ 1 and ã ≪ 1.
large rotation. Due to the hydrogen-like spectrum of the bound states,
this means that we are limited to considering small values
of µM . Again, this expectation is validated by Fig. 7.
E. Proca modes to second order At second order, all curves in Fig. 7 show a maximum
at large values of the spin. This maximum is an arti-
As discussed in Sec. IV C, the perturbation equations fact of the second-order approximation: a similar fea-
for massive vector fields at second order in rotation, ture appears also in the scalar case in which the exact,
Eqs. (70) and (71), are obtained from Eqs. (69), using Teukolsky-based result is monotonic (cf. Fig. 3). In what
the perturbation equations in the nonrotating limit to follows we have discarded numerical data beyond such
eliminate the second derivatives. The explicit form of maxima, and we have taken this truncation into account
the equations is available online [61]. when estimating numerical errors.
Solving the eigenvalue problem for these equations is The bottom line of this discussion is that the first or-
numerically more demanding than in the case of pertur- der calculation provides a reliable estimate of the order
bations at first order in ã, especially when the imaginary of magnitude of the instability. The second-order calcu-
part of the modes is tiny (as in the axial case and in the lation should be quantitatively reliable even in the unsta-
small-mass limit). For this reason we present only a se- ble regime, at least up to moderately large values of ã.
lection of results that allow us to validate the accuracy of Roughly speaking, the reason why the first-order calcu-
the first-order approximation, and (more in general) to lation captures the correct qualitative properties of the
estimate the errors due the slow-rotation approximation. instability can be traced back to the fact that the super-
The imaginary part of the fundamental S = 0 axial and radiant instability condition ω < mΩH is a “first order
S = −1 polar modes as a function of ã at first and second effect”. Because ΩH is an odd function of ã, a first-order
order are shown in Fig. 7. The slow-rotation method cor- approximation of this relation can be expected to be valid
rectly predicts the onset of the Proca instability already modulo third-order corrections.
at first order, where (a priori) large deviations could be
expected. Furthermore, the value of ã which corresponds
to the onset of the instability is slightly smaller at second F. The minimum instability growth scale
order, consistently with the fact that the horizon location
gets negative second-order contributions [cf. Eq. (19)]. The agreement between first- and second-order calcu-
Fig. 7 actually displays a remarkably good quantitative lations can be used to estimate the order of magnitude
agreement between the first- and second-order calcula- of the instability by extrapolation. We will fit our data
tions. The main difference is that in the superradiant at the onset of the instability, and then extrapolate to
region the instability predicted by the first-order expan- the superradiant region. This procedure should give us
sion is weaker (by a factor of a few) than the instability at least a correct order of magnitude for the instability
computed at second order. Another important point is timescale. We expect the extrapolation to work better
that our perturbative approach can consistently describe in the axial case, where bound-state frequencies have a
18

-10
10 -7
Mµ=0.10 10 Mµ=0.10
Mµ=0.09 Mµ=0.09
Mµ=0.08
-12 -8 Mµ=0.08
10 Mµ=0.07 10
Mµ=0.07
Mµ=0.06
-9 Mµ=0.06
10
|MωI|

|MωI|
-14 Mµ=0.05
10
Mµ=0.05
Mµ=0.04 -10
10
Mµ=0.04
-16
10 Mµ=0.03
-11
10
Mµ=0.03
l=m=1 l=m=1
-18 -12
10 0 0.1 0.2 0.3 0.4 0.5 10 0 0.1 0.2 0.3 0.4 0.5
2 2
J/M J/M

FIG. 6. (color online) Imaginary part of the axial (left panel) and of the polar S = −1 (right panel) vector modes as a function
of the BH rotation rate ã = J/M 2 for ℓ = m = 1 and several values of µ. Although beyond the regime of validity of the first
order approximation, the modes cross the axis and become unstable precisely when the superradiance condition (94) is met.
As shown in Fig. 7, the first-order results are also in qualitatively good agreement with those obtained at second order.

-10
10
-7
Mµ=0.10 10 Mµ=0.10
-11
10
Mµ=0.08
-8 Mµ=0.08
-12 10
10
|MωI|

|MωI|

-13 Mµ=0.06
10 -9 Mµ=0.06
10
-14
10
-10
10
-15
10
l=m=1 l=m=1
-16 -11
10 0 0.1 0.2 0.3 0.4 0.5 10 0 0.1 0.2 0.3 0.4 0.5
2 2
J/M J/M

FIG. 7. (color online) Imaginary part of the axial S = 0 vector modes (left panel) and for polar S = −1 vector modes (right
panel) computed at second order in the rotation for ℓ = m = 1 and several values of µ. The dotted thinner lines refer to the
first order calculation. The modes cross the axis and become unstable when the superradiance condition is met.

monotonic behavior as functions of ã (cf. the left panel Here and in the following, numerical errors are esti-
of Fig. 6). We estimate that close to the onset of the mated by comparing the results at first and second order,
superradiant instability the imaginary part of the modes and by taking the maximum deviation between the fit
should scale as follows: and the data. We generate numerical data with increas-
ing accuracy until convergence is reached. In particular,
M ωI ∼ γSℓ (ãm − 2r+ µ) (M µ)4ℓ+5+2S , (95)
in the direct-integration method it is important to retain
where γSℓ is a coefficient that depends on S and ℓ. For a sufficiently large number of terms in the series expan-
axial modes with ℓ = 1 our numerical data yield γ01 ≈ sion of the boundary condition at infinity, Eq. (77), in
0.09±0.03, in good agreement with the analytical formula order to achieve sufficient accuracy. This is expected be-
in the limit M µ ≪ 1, cause bound-state modes decay exponentially, and there-
fore it is challenging to perform a numerical integration
1
(ℓ=1 , axial)
M ωI (ãm − 2r+ µ) (M µ)9 ,
= (96) at large distances.
12
which is derived in Appendix C. From the equation above We have applied the same method to compute unstable
we get γ01 = 1/12 ∼ 0.0833. Expression (96) is compared massive scalar modes [60], whose imaginary part scales
to the numerical results in Fig. 5. as in the case of axial vector modes with S = 0. In this
19

scalar
case we find γ01 ≈ 0.03 ± 0.01, again in reasonable be surprising. Actually, Eq. (95) performs remarkably
agreement with analytical expectations (by setting s = 0 well even for values of ã close to extremality [43]: it over-
scalar
in Eq. (96) we get γ01 = 1/48 ∼ 0.0208). Comput- estimates the results in Table III in Ref. [35] by only 3%
ing the modes in the limit M µ ≪ 1 is numerically chal- when ã = 0.7 and by less than 70% when ã = 0.99. In
lenging because of the steep scaling of ωI with M µ [cf. our view, this agreement is quite impressive.
Eq. (95)] and this affects the precision of the fit. Despite Let us apply the same argument to vector fields. From
these numerical difficulties, we are able to recover the cor- Eq. (95) we expect the stronger instability when ℓ = m =
rect order of magnitude for the instability timescale when 1. For ã = 0.7 this corresponds to
M µ ≪ 1. In any case, the observational bounds that we
will address in the next section depend only mildly on M µmin ∼ 0.187 , M ωImax ∼ 5.8γ11 × 10−10 ,
the precise value of γSℓ . For the polar modes we find M µmin ∼ 0.184 , M ωImax ∼ 1.7γ01 × 10−8 ,
γ−11 ≈ 20 ± 10, while extracting the coefficient γ11 with
sufficient precision is numerically challenging even for the M µmin ∼ 0.179 , M ωImax ∼ 5.1γ−11 × 10−7 ,
fundamental mode with ℓ = 1, since ωI ∼ (M µ)11 . for S = 1, 0, −1, respectively. Our data at second or-
Obviously our method becomes inaccurate when ã → der are compatible with γ01 ≈ 0.09 and γ−11 ≈ 20. As
1, and the fit (95) can provide at most an order of we discussed, we could not extract the coefficient γ11
magnitude for the instability timescale in the extremal with sufficient precision. Nevertheless it is not difficult
limit. However, in the massive scalar case Eq. (95) with to show that the modes with S = 1 are indeed those
S = 0 is exact in the M µ ≪ 1 limit, for any value with the smallest (in modulus) imaginary part, i.e. they
of ã [60]. Furthermore, in the massive vector case, the are the least interesting for what concerns the instability.
good agreement between the first- and second-order ex- This confirms the expectation that polar perturbations
pansions suggests that the slow-rotation approximation with S = −1 should have the strongest instability in the
should be quite accurate even for moderately large spins, rapidly rotating limit, as conjectured in Ref. [16]. In the
ã . 0.7. Therefore it is reasonable to expect that extrap- Proca case the strongest instability should occur on a
olations of Eq. (95) from the slow-rotation limit should at timescale
least provide the correct order of magnitude (and possi-
bly a reasonable quantitative estimate) of the instability M (M µ)−7
timescale. τvector = ωI−1 ∼ . (98)
γ−11 (ã − 2µr+ )
It is tempting to extrapolate our predictions to generic
values of ã and to discuss possible astrophysical impli- The timescale above must be compared with that for the
cations of the superradiant instability for massive vec- massive scalar case [60] for ℓ = m = 1,
tor fields around Kerr BHs. Our conclusions rely on
extrapolation, so they should be confirmed by indepen- 48M (M µ)−9
τscalar ∼ . (99)
dent means. Let us assume that the imaginary part of ã − 2µr+
the modes is generically described by Eq. (95), where
Roughly speaking, the instability timescale against
we left the coefficient γSℓ undetermined, since its precise
value (whose order of magnitude approximately ranges vector polar perturbations is of order τvector ∼
−1
10−2 γ−11 (M/M⊙ ) s.
between 0.1 and 10, depending on S and ℓ) is not crucial
for the following discussion. From Eq. (95) we find that In the next section we shall use the results obtained
from our numerics at second order to discuss some im-
the instability growth timescale has a minimum (i.e., ωI
portant astrophysical consequences of the Proca instabil-
has a maximum) for
ity. In this context it is crucial to have a robust estimate
5 + 4ℓ + 2S of the instability timescale in the M µ ≪ 1 limit. In
µmin = mã , (97) the axial case our numerical results are supported by the
4r+ (3 + 2ℓ + S)
analytical formula (96), which provides strong support
and the corresponding value of the timescale τ = ωI−1 is that the fit (95) represents the correct behavior when
given by Eq. (95) evaluated at µ = µmin . From Eq. (95) M µ ≪ 1. Unfortunately in the polar case we do not
we also see that (excluding the case ãm ≃ 2r+ µ) this is have the same analytical insight and, as discussed above,
a typical value of the instability timescale. the small-mass regime is challenging to investigate nu-
If we apply the same procedure to scalar fields (ℓ = 1, merically. In order to verify the reliability of Eq. (95) for
−1
S = 0, γ01 = 48), we find that the minimum instabil- polar perturbations we have tried several choices of the
ity corresponds to ã = 1, M µmin = 0.45 and M ωImax = fitting functions in the most relevant case, ℓ = m = 1
1.6 × 10−6 . Notice that this simple argument relies on an and S = −1. It turns out that Eq. (95) is a conservative
(a priori unjustified) extrapolation of analytical expres- choice in the polar case, as other fits generically predict
sions valid for M µ ≪ 1 to the regime M µ ∼ 1. In the a stronger instability. Furthermore, by comparing the re-
scalar case this prediction overestimates the numerical sults in Figs. 6 and 7 for the axial and for the polar case,
results by an order of magnitude: in fact, Refs. [34, 35] it is clear that the polar case exhibits a more complex
found a minimum instability M ωI ∼ 1.5 × 10−7 at behavior, which is hard to reproduce without some ana-
M µ ∼ 0.42 for ã ∼ 0.99. This disagreement should not lytical insight. In fact our numerical data are also very
20

well reproduced by the following fitting function: momentum from the BH, leaving behind a Kerr metric
with dimensionless spin parameter below the superradi-
M ωI ∼ (ã − ãSR ) [η0 (M µ)κ0 + η1 ã(M µ)κ1 ] , (100) ant threshold: in other words, the angular frequency of
where the BH horizon must be such that ΩH . µ. By this
p argument, even a single reliable supermassive BH spin
−m + m2 + 16(M µ)2 4M µ measurement can impose a stringent constraint on the
ãSR = ∼ + O(µ3 ) , (101)
2µM m allowed mass range of massive vector fields. Bounds on
the mass of the vector field follow from the requirement
corresponding to the superradiance threshold [Eq. (94)]
that astrophysical spinning BHs should be stable, in the
when ωR ∼ µ. The fit (100) has been obtained by im-
sense that the timescale (98) should be larger than some
posing a second-order (in ã) functional form and ωI = 0
observational threshold.
when ã = ãSR , consistently with our data. Finally, the
functional dependence on µ of the two remaining terms 10
-17

has been obtained by fitting the data in the instability


region and for 0.03 . M µ . 0.1. We found η0 ≈ −6.5±2,
µ=Ω H
η1 ≈ 2.1 ± 1, κ0 ≈ 6.0 ± 0.1 and κ1 ≈ 5.0 ± 0.3. While Stable
the fit (95) is physically more appealing [16], Eq. (100) -18
10
reproduces our numerical data in the whole instability 4

mv [eV]
region within a factor of two. It would be interesting to τ =10 yr
better understand the behavior of the polar instability Unstable
in the limit M µ ≪ 1 using different approaches, such τ = τSalpeter
-19
as a time-evolution analysis or a full numerical evolution 10
τ = τHubble
of the Proca equation. In the following we shall discuss
how the choice of either Eq. (95) or Eq. (100) affects the τ > τHubble
astrophysical implications of our results. polar S=-1, fit I
-20
10 0 0.2 0.4 0.6 0.8 1
2
J/M
VII. ASTROPHYSICAL IMPLICATIONS OF
THE PROCA INSTABILITY
FIG. 8. (color online) Bounds on the photon mass mv = ~µ
obtained by extrapolating the instability timescale (98) for
Our numerical results in the previous section indicate S = −1 polar Proca modes of a Kerr BHs up to ã = 1.
that polar perturbations with S = −1 have the shortest The region delimited by the curves corresponds to an insta-
instability timescale, as conjectured in [16]. For fixed val- bility timescale τ < τHubble (continuous red line), τ < τSalpeter
ues of ã and µ, the instability timescale for polar pertur- (dashed green line) and τ < 104 yr (dot-dashed blue line), re-
bations is smaller by two-three orders of magnitude than spectively. We set γ−11 ≈ 20 ± 10 in Eq. (98) and consider
in the axial case. While this conclusion relies on an ex- a Kerr BH with M = 107 M⊙ , but the results have a simple
scaling with γ−11 and with the BH mass (cf. Eqs. (102) and
trapolation of calculations that are valid (strictly speak-
(103)). Thin dashed lines bracket our estimated numerical
ing) only in the slow-rotation limit, we have shown that errors.
a similar extrapolation in the scalar case is in remark-
able quantitative agreement with numerical results that
do not rely on the slow rotation approximation. Further- For isolated BHs we can consider the observational
more, in the Proca case a second-order calculation gives threshold to be the age of the Universe, τHubble =
solid evidence that our extrapolation is reliable, both 1.38 × 1010 yr. A more conservative assumption is to
qualitatively and quantitatively. Therefore it is reason- include possible spin growth due to mergers with other
able to expect that extrapolations from the slow-rotation BHs and/or accretion. These effects are in competition
limit should at least provide the correct order of magni- with superradiant angular momentum extraction.
tude (and possibly a reliable quantitative estimate) of the The most likely mechanism to produce fast-spinning
polar instability timescale. Here we explore the tantaliz- BHs3 is prolonged accretion [78]. In this case we can
ing astrophysical implications of the Proca superradiant compare the superradiance timescale to the shortest
instability. Unless otherwise stated, we shall focus on
the most relevant modes, those with ℓ = m = 1, which
correspond to the stronger instability.
3 Supermassive BHs with moderate spin (say ã ∼ 0.7, as in Fairall
9 [57]) could also be produced by comparable-mass BH merg-
A. Implications from the existence of supermassive ers [9]. The timescale for mergers depends on details of the
black holes “final parsec problem”, which are poorly known [75]. The mini-
mum timescale necessary for BHs to merge is the gravitational-
radiation timescale, which is of the same order as the Salpeter
The most conservative assumption on the final state of timescale (cf. Eq. (10) in [75] or Eq. (26) in [76]), but it is likely
the instability is that the radiation will extract angular that the timescale for binary BHs to merge will be dominated by
21

10 10 10
10 10 10
9 9
9 10 10
10 8 τ 8 τSalpeter
10 Salpeter
10
10
8 τSalpeter 7 7
10 10
7 6
10 10 10
6
5
6 10 5
10 4 10
τ [yr]

τ [yr]

τ [yr]
10 4
5 3 10
10 10 3
4
Axial (S=0) 2 Polar (S=-1, fit I) 10 Polar (fit II)
10 10 2
MCG-6-30-15
10
1 MCG-6-30-15 10 MCG-6-30-15
3 Fairall 9 Fairall 9 1 Fairall 9
10 SWIFT J2127.4+5654
10
0 SWIFT J2127.4+5654 10 SWIFT J2127.4+5654
2 1H0707-495 -1 1H0707-495 0 1H0707-495
10 Mrk 79 10 Mrk 79 10 Mrk 79
Mrk 335 -2 Mrk 335 -1 Mrk 335
1 NGC 7469 10 NGC 7469
10 NGC 7469
10 -3 -2
0
NGC 3783 10 NGC 3783 10 NGC 3783
10 -4 -3
-21 -20 -19 -18 -17 10 -21 -20 -19 -18 -17 10 -21 -20 -19 -18 -17
10 10 10 10 10 10 10 10 10 10 10 10 10 10 10
mv [eV] mv [eV] mv [eV]

FIG. 9. (color online) Instability timescale (in years) as a function of the vector mass mv (in eV) for the eight supermassive
BHs listed in Table 2 of [56]. The left panel refers to the axial case (S = 0, ℓ = m = 1); the middle panel to the polar case,
Eq. (95) with S = −1, ℓ = m = 1 and γ−11 ≈ 20; the right panel to the polar case, but using the alternative fitting function
given in Eq. (100). The horizontal line corresponds to τ = τSalpeter = 4.5 × 107 yr.

timescale over which accretion could spin up the BH. can be well approximated by the formula
Thin-disk accretion can increase the BH spin from ã = 0  1/(5+4ℓ+2S)
√ ã ≈ 1 with a corresponding mass increase by a factor
to ~ M
m∗v = ~µ∗ = , (102)
6 [79]. If we assume that mass growth occurs via accre- M mγSℓ τc
tion at the Eddington limit, so that the BH mass grows
exponentially with e-folding time given by the Salpeter where τc is the threshold time. If τc = τHubble and in the
timescale τSalpeter = 4.5 × 107 yr, then the minimum S = −1 polar case, we get
timescale for the BH spin to grow from ã = 0 to ã ≈ 1 via 6/7
7 × 10−20 107 M⊙

thin-disk accretion is comparable to τSalpeter. Note that ∗
mv ∼ eV . (103)
1/7 M
accretion at the Eddington limit is a rather conservative γ −11
assumption, since more typically we would expect ac-
cretion to be sub-Eddington. Furthermore fast-spinning Note that, although extracting the values of the fitting
BHs are hard to produce in the presence of radiative ef- parameters can be numerically challenging, our results
fects [80] or magnetohydrodynamics [81], and “chaotic” depend very mildly on the precise value of γ−11 . From
accretion [82] would make large spin parameters even less Eq. (98) we see that the instability timescale is inversely
likely. proportional to γ−11 . Therefore the near-horizontal “ob-
servability threshold” in Fig. 9 that would correspond to
For illustration, in Fig. 8 we consider a Kerr BH with choosing τ = τSalpeter rather than τ = τHubble can be
M = 107 M⊙ , so that mv = 10−17 (M µ) eV. We assume obtained multiplying γ−11 in Eq. (103) by the ratio of
γ−11 ≈ 20 ± 10 in Eq. (98) (consistently with our numer- the two timescales, τSalpeter/τHubble ≃ 3 × 10−3 . Then
ical results) and we plot: (i) the superradiant threshold the vector mass bound (103) would increase by a mod-
(dashed black line), corresponding to the points where est factor of 2.3. What is crucial in obtaining strong
µ ≈ ΩH and the instability timescale (in the perturba- astrophysical bounds is having a large BH mass, because
tive treatment) becomes infinite; (ii) the contours cor- 6/7
responding to an instability timescale τ = τHubble (con- bounds scale like 107 M⊙ /M .
tinuous red line), τ = τSalpeter (dashed green line) and Brenneman et al. [56] have recently presented a list
τ = 104 yr (dash-dotted blue line), respectively. Each of eight supermassive BH mass and spin estimates. In
curve is shown with the corresponding error bars. Fig. 9 we show the superradiant instability timescale τ
(in years) as a function of mv (in eV) for these supermas-
In the region delimited by the continuous red line the sive BHs. The dashed horizontal line denotes τSalpeter,
instability timescale is shorter than the Hubble time which we (conservatively) assume as the threshold for
(of course, similar considerations apply to the dashed the timescale to be observationally significant. For each
green and dash-dotted blue lines). For large ã the lower system we compute instability timescales using the BH
branches of the “critical” curves are roughly horizontal. mass and spin estimates given in Table 2 of [56] (we se-
By solving ωI−1 = τc in the limit M µ ≪ 1 this regime lect either the average spin value quoted in the table, or
the lower limit on the spin when the observations im-
ply ã > 0.98). The left panel refers to axial modes with
S = 0 and ℓ = m = 1, for which we can compute the
instability timescale analytically: cf. Eq. (96) and Ap-
the dynamical friction timescale, which is typically of the order pendix C. The middle and right panels shows that quali-
of a few Gyrs for satellites in a Milky Way-type halo [77] and tatively similar results hold for polar modes with S = −1
hence larger than the Salpeter timescale. and ℓ = m = 1, which exhibit the strongest instability.
22

In the polar case, we computed the instability timescale is whether the coupling of accreting matter to massive
using both Eq. (95) (fit I) and Eq. (100) (fit II). The bosons can quench the instability. In principle massive
results are qualitatively similar and show that, indepen- photons (unlike hidden U (1) fields, for which the inter-
dently of the fitting function, the axial bounds on mv action with matter is very small) can couple strongly to
are typically one-two orders of magnitude less stringent matter. However it is unlikely that this will significantly
than the polar bounds. Fig. 9 shows that existing mea- affect the superradiant instability discussed here, for two
surements of supermassive BH spins rule out Proca field reasons. The first is that unstable modes are large-scale
masses in the whole range 10−20 eV. mv . 10−17 eV. coherent modes whose Compton wavelength is compara-
Note that arguments based on the superradiant insta- ble to (or larger than) the size of the BH’s event hori-
bility can only be used to set “exclusion windows” on the zon, and accretion disks are typically charge-neutral over
boson mass, rather then upper bounds [42, 43]. How- these lengthscales, so any possible coupling with ordinary
ever, observations of BHs in different mass ranges can neutral matter is incoherent and most likely inefficient.
exclude different mass windows. For example, stellar- The second is that accretion disks are expected to be
size BHs, M ≃ (5 − 20)M⊙ , can set an exclusion region localized on the equatorial plane, and therefore can af-
of 10−13 eV. mv . 10−9 eV. This mass range is relevant fect at most some (but not all) of the unstable modes.
for so-called “dark photons” coupled to sterile neutri- The investigation of the superradiant instability in the
nos [83]. If the identified intermediate-mass BHs were presence of matter requires further work, but the above
confirmed to exist and if the largest known supermas- considerations suggest that vacuum estimates should be
sive BHs with M ≃ 2 × 1010 M⊙ [84, 85] were confirmed reliable. Spin estimates for slowly accreting BHs (such
to have sufficiently large spin, combined observations of as the BH at the center of our own galaxy) are the most
spinning BHs can potentially exclude the entire mass reliable, in that they should be less sensitive to details of
windows 10−21 eV. mv . 10−9 eV [42, 43]. The lat- the interaction of vector fields with matter.
ter is complementary to the bounds set by several other
experiments and observations that exclude light bosonic
particle with larger mass [52, 83, 86]. VIII. CONCLUSIONS AND EXTENSIONS
Using current data on supermassive BHs [56], the best
bound comes from Fairall 9 [57], for which the polar in- BH perturbation theory is a powerful tool, with impor-
stability implies a conservative bound (including mea- tant applications in astrophysics and high-energy physics,
surement errors) mv . 10−20 eV or mv . 10−21 eV, but its applicability is limited when the perturbation
depending on the fitting function [59]. Although the in- equations cannot be separated. We have discussed a gen-
stability for axial modes is weaker, in this case our re- eral method to circumvent these difficulties in the case of
sults are more precise and imply a bound as stringent slowly rotating BH spacetimes. The method was origi-
as mv . 4 × 10−20 eV if we consider τ < τSalpeter , or nally developed to study the modes of rotating stars but,
mv ×10−20 eV if we consider τ < τHubble . Even under our as we showed, it applies to any slowly rotating BH back-
very conservative assumptions these results are of great ground and to any kind of perturbation. Within this gen-
significance, since the current best bound on the photon eral framework, we discussed two effects induced by ro-
mass is mγ < 10−18 eV [52, 86]. Combining the exclusion tation: (i) the Zeeman-like splitting of the eigenfrequen-
window derived above, 10−20 . mv . 10−17 eV, with ex- cies, that breaks the degeneracy of modes with different
isting bounds [52, 86], we find the most stringent upper azimuthal number m, and (ii) a Laporte-like selection
bound on the photon mass, mγ . 10−20 eV. rule, in which modes with multipolar index ℓ are coupled
We remind the reader that our results are also sum- to those with multipolar index ℓ ± 1 and opposite parity
marized in Fig. 1, where we show instability windows for and to those with ℓ ± 2 and same parity. We have ex-
axial and polar modes together with supermassive BH tended Kojima’s arguments to show that these couplings
mass and spin estimates from Ref. [56]. do not affect the QNM frequencies of the spacetime to
first order in the rotation rate, and we verified this claim
by numerical integration of the perturbation equations.
B. Nonlinear effects, other couplings As a first relevant application of this method, we stud-
ied massive vector (Proca) perturbations on a slowly ro-
An important ingredient that was not taken into ac- tating Kerr background, i.e. the only interesting class
count in our study is the nonlinear evolution of the insta- of perturbations of Kerr BHs in four dimensions that
bility, that can modify the background geometry. Photon does not seem to be separable. Using our approach we
self-interactions are very weak, being suppressed by the reduced the equations to a set of coupled ODEs, that
mass of the electron. Therefore, it is quite likely that can be solved by a straightforward extension of methods
the outcome of the instability will be a slow and gradual valid in the nonrotating case. We have computed the
drainage of the hole’s rotational energy. However, exotic corrections to the Proca eigenfrequencies of a Kerr BH
massive vector fields may show nonlinearities as those to first and to second order in the rotation rate. For the
studied in Ref. [46, 47] for the scalar case. first time we showed that the Proca bound-state modes
Another important aspect that should be investigated become unstable when the superradiant condition (74)
23

is met. We proved this numerically for the polar modes, ACKNOWLEDGMENTS


and both numerically and analytically for the axial modes
in the limit M µ ≪ 1. Unfortunately the method becomes We wish to thank the Axiverse Project members
intrinsically inaccurate for large values of the rotation (especially Hideo Kodama), Antonino Flachi, Stefano
rate. This prevents us from accurately computing the Liberati and João Rosa for valuable discussions, and
instability timescale close to extremality. By fitting and the referees for their constructive comments. This
extrapolating our data we confirmed the scaling with mv work was supported by the DyBHo–256667 ERC Start-
conjectured in Ref. [16], and we also estimated the order ing Grant, the NRHEP–295189 FP7–PEOPLE–2011–
of magnitude of the instability timescale. We expect our IRSES Grant, and by FCT - Portugal through PTDC
results to be reliable up to ã ∼ 0.7. Our estimates should projects FIS/098025/2008, FIS/098032/2008, CTE-
be confirmed by more accurate methods, but they pro- ST/098034/2008, CERN/FP/123593/2011. A.I. was
vide strong evidence that measurements of rotating su- supported by JSPS Grant-in-Aid for Scientific Research
permassive BH spins set the most stringent upper bounds Fund No. 22540299 and No. 22244030. E.B. was sup-
on the photon mass (cf. Fig. 1). ported by NSF Grant No. PHY-0900735 and NSF CA-
REER Grant No. PHY-1055103. P.P. acknowledges
Besides their astrophysical implications, our results are financial support provided by the European Commu-
relevant also for some classes of gravitational theories nity through the Intra-European Marie Curie contract
with higher-order curvature terms in the action. For aStronGR-2011-298297 and the kind hospitality of the
example,√as proved by Buchdahl [87], a theory defined Department of Physics, University of Rome “Sapienza”
by L = −g R + αR[ab] R[ab] in the Palatini approach
 and of the International School for Advanced Studies
is dynamically equivalent to the Einstein-Proca system (SISSA) in Trieste. V.C. thanks the Yukawa Institute
when the vector mass µ2 = 3/|α| (see also Ref. [88]). for Theoretical Physics at Kyoto University for hospital-
The Proca instability of Kerr BHs in this theory can put ity during the YITP-T-11-08 workshop on “Recent ad-
constraints on the parameter α, which is related to the vances in numerical and analytical methods for black hole
nonmetricity of the connection. dynamics”.

While limited by the slow-rotation assumption, the Appendix A: Coefficients of the perturbation
power of the present method consists in its generality. equations
The slow-rotation approximation allows us to study lin-
ear perturbations of any slowly rotating spacetime even In this Appendix we list the explicit form of the co-
when the equations are nonseparable. For instance it efficients appearing in the perturbation equations. We
(i) (i) (i)
can be used to study gravitational-electromagnetic per- remark that the coefficients Aℓ , Dℓ , Aℓ , Bℓ , Dℓ incor-
turbations of Kerr-Newman BHs in four dimensions [12], porate terms at zeroth, first and second order in the ro-
QNMs of Myers-Perry BHs [14] and BH greybody fac- tation parameter ã. Therefore, there is no direct corre-
tors in higher-dimensional rotating spacetimes [11, 54]. spondence between these coefficients and the quantities
Another interesting extension is the stability analysis of at fixed order in ã, such as Aℓ , Āℓ , Âℓ , etc., which have
rotating BHs in asymptotically anti-de Sitter spacetimes. been introduced in Sec. II. In the following, a prime de-
This is relevant in the context of the gauge/gravity du- notes derivative with respect to the radial coordinate r.
ality, as the QNMs of anti-de Sitter BHs in five dimen-
sions are holographically related to specific correlation
functions and transport coefficients of the dual bound- 1. Coefficients of the scalar equation
ary theory [4].
The explicit form of the coefficients appearing in
We note that perturbations of rotating stars have been Eq. (22) is
studied up to second order in the Cowling approximation Aℓ = −2 M 2 r2 4M 2 + 2M r Λ − 1 + r2 µ2 + r4 ω 2
 
(see e.g. Refs. [89–91]), where metric perturbations are
−r2 Λ + r2 µ2 Ψℓ + r(r − 2M ) (2M Ψ′ℓ

neglected and the equations have a structure reminiscent
+r(r − 2M )Ψ′′ℓ ))] + 8ãmM 4 r3 ωΨℓ + ã2 M 4 36M 2

of our general Eqs. (4) and (5).
+2M r Λ − 11 + r2 µ2 − r2 Λ − 2 + 2m2 + r2 µ2
 

Finally we remark that the techniques discussed in this r3 ω 2



2 2

paper are not limited to general relativity. They are also + 8M − 2M r + r Ψℓ
r − 2M
useful to study gravitational perturbations and the lin-
ear stability of slowly rotating BH metrics in alterna- +r(r − 2M ) (10M Ψ′ℓ − r(2M + r)Ψ′′ℓ )] , (A1)
2 4 2 2 2 2
 
tive theories of gravity, for example in quadratic grav- Dℓ = 2ã M 4M − Λr + 2M r Λ − 1 + r ω Ψℓ
ity [22, 25, 26]. We hope to report on these interesting +r(r − 2M ) (2M Ψ′ℓ + r(r − 2M )Ψ′′ℓ )] . (A2)
extensions in the near future.
24

2. Coefficients of the Proca perturbation equations

In this Appendix we list the coefficients appearing in Eqs. (44)–(47). The coefficients read

uℓ(1) −r2 Λ + 2m2 + µ2 r2 − 2 + 32M 2 + 2M r Λ + µ2 r2 − 10


"  
(0) ã2 M 2
Aℓ = +
2r6 2M − r
r   
ℓ ℓ ℓ
 
2
(2M − r) ℓ(ℓ + 1)(2M − r) −12M u′(1) + r(2M + r)u′′ (1) + ir2 ωu′ (2) − 2im2 r2 ωuℓ(3)
ℓ(ℓ + 1)(r − 2M )
  
i 2ãmM 2 Λrωuℓ(1) − i (2M − r)u′ ℓ(3) − r2 ω 2 uℓ(3) + Λuℓ(2)
−iΛrω −8M 2 − 2M r + r2 uℓ(2) +

Λr4 (2M − r)
   
ℓ ℓ
(2M − r) Λ + µ2 r2 uℓ(1) + r(r − 2M )2 u′′ (1) + irω (2M − r) uℓ(3) − ru′ (2) + (r − 4M )uℓ(2)

− , (A3)
r3 (2M − r)
(1) ã2 M 2 h
Λuℓ(2) (2M − r) ℓ2 (2M + r) + ℓ(2M + r) + r 2m2 + µ2 r(2M + r) + r2 ω 2 8M 2 + 2M r + r2
 
Aℓ = 4 3
2Λr (2M − r)
i
ℓ  ℓ
+2m2 r(r − 2M )2 u′ (3) + iΛrω(2M − r) 8M 2 − 2M r + r2 u′ (1) − iℓ(ℓ + 1)ω(2M − r) 8M 2 − 6M r + 3r2 uℓ(1)

   
ℓ ℓ
2ãmM 2 r2 ω(r − 2M )u′ (3) + iΛ (2M − r)uℓ(1) + r (r − 2M )u′ (1) + 2irωuℓ(2)
+
ℓ(ℓ + 1)r4 (r − 2M )2
 
ℓ ℓ
(2M − r) (2M − r)u′ (3) + ir2 ωu′ (1) + uℓ(2) (2M − r) Λ + µ2 r2 + r3 ω 2 − irω(2M − r)uℓ(1)
 
+ , (A4)
r2 (r − 2M )2
   

ã2 M 2 (2M − r) −2m2 ruℓ(3) − Λr(r − 2M )u′ (2) − Λ(2M − r)uℓ(2) − iΛrω 8M 2 − 2M r + r2 uℓ(1)

(2)
Aℓ =
ℓ(ℓ + 1)(r − 2M )2
   
2 ′ℓ ℓ ℓ 2 ℓ
2 ℓ
4ãmM r(rωu(3) + iΛu(1) ) ℓ 2r (2M − r) ru (2) + u (2) − u (3) (r) − ir ωu (1)
+ + , (A5)
Λ(2M − r) 2M − r
(0)
4ãM 2 uℓ(4) 4ã2 mM 3 ωuℓ(4)
Ãℓ = − , (A6)
r5 Λr5

(1) 2iã2 mM 2 u′ (4)
Ãℓ =− , (A7)
Λr4
(2)
Ãℓ = 0, (A8)
 

ã2 mM 2 ωuℓ(4) 2ãM 2 (r − 2M )u′ (4) + r2 ω 2 uℓ(4)
(0)
Bℓ = − , (A9)
ℓ(ℓ + 1)r3 (2M − r) Λr4 (2M − r)
ℓ ℓ
(1) iã2 mM 2 u′ (4) 2iãM 2 ωu′ (4)
Bℓ = − , (A10)
ℓ(ℓ + 1)r3 (2M − r) Λr2 (2M − r)
(2)
2iã2 mM 2 ruℓ(4) 4iãM 2 r2 ωuℓ(4)
Bℓ = − , (A11)
Λ(r − 2M ) Λ(r − 2M )
(0)
4ã2 M 3 (Λuℓ(1) + irωuℓ(3) )
B̃ℓ = , (A12)
ℓ(ℓ + 1)r6
 

2ã2 M 2 (r − 2M )u′ (3) − Λuℓ(2)
(1)
B̃ℓ = , (A13)
Λr4 (2M − r)
(2)
B̃ℓ = 0, (A14)
25

(0) ã2 M 2 h

 
′ℓ ′′ ℓ

ℓ ′ℓ

Dℓ = −(2M − r)(6M − Λr)u (1) + r (2M − r) 6M u (1) − r(r − 2M )u (1) + irω u (3) − ru (2)
r6 (2M − r)
i
−irω(10M − r)uℓ(2) , (A15)

(1) ã2 M 2 h
ℓ 2 2
 
′ℓ ′ℓ


i
Dℓ = 4 u (2) −ℓ(ℓ + 1)(2M − r) − 2M r ω + (2M − r) (r − 2M )u (3) − 2iM rωu (1) + 2iM ω(2M − r)u (1) ,
r (r − 2M )2
(A16)
2 2 h
(2) 2ã M 

 i
Dℓ = −(2M − r) ru′ (2) + uℓ(2) − u(3) ℓ (r) + 2iM rωuℓ(1) , (A17)
2M − r
and
ã2 M 2 h
ℓ ℓ ℓ ℓ
αℓ = 4 2
−48M 3u′ (3) + 4M 2 rℓ2 u′ (2) + 4ℓM 2 ru′ (2) + 48M 2 ru′ (3) − 4M r3 ω 2 uℓ(3) + iΛr2 ω(2M + r)uℓ(1)
2Λr (r − 2M )
i
ℓ ℓ ℓ ℓ ℓ ℓ
−4M r2 ℓ2 u′ (2) − 4ℓM r2 u′ (2) − 12M r2 u′ (3) − 3Λ(r − 2M )2 uℓ(2) + 4M r(r − 2M )2 u′′ (3) + r3 ℓ2 u′ (2) + ℓr3 u′ (2)
2ãmM 2 (rωuℓ(3) − iΛuℓ(1) ) 1 h
ℓ ℓ ℓ
+ + 4M 2 u′ (3) − 2µ2 M r2 uℓ(3) − 2M rℓ2 u′ (2) − 2ℓM ru′ (2) + Λ(2M − r)uℓ(2)
ℓ(ℓ + 1)r3 (2M − r) Λr2 (2M − r)
i
ℓ ℓ ℓ ℓ
−2M ru′ (3) − r(r − 2M )2 u′′ (3) + µ2 r3 uℓ(3) − r3 ω 2 uℓ(3) + iΛr2 ωuℓ(1) + r2 ℓ2 u′ (2) + ℓr2 u′ (2) , (A18)
  
ℓ ℓ
ã2 M 2 −uℓ(4) ℓ(ℓ + 1)(r − 2M )2 − 2M r3 ω 2 − 2M (r − 2M )2 ru′′ (4) − 3u′ (4)

2ãmM 2 ωuℓ(4)
βℓ = −
Λr4 (r − 2M )2 Λr2 (2M − r)
 
ℓ ℓ
uℓ(4) (2M − r) Λ + µ2 r2 + r3 ω 2 + (2M − r) r(2M − r)u′′ (4) − 2M u′(4)
 
+ , (A19)
Λr2 (2M − r)
    
ℓ ℓ ℓ
ã2 mM 2 2ΛM rωuℓ(1) − i −2M r2 ω 2 uℓ(3) + (r − 2M ) −8M u′(3) − r(r − 2M )u′′ (3) + Λru′ (2) + Λ(8M − r)uℓ(2)
ζℓ =−
Λr4 (2M − r)
  

6ãM 2 (2M − r)uℓ(1) + r (r − 2M )u′ (1) + irωuℓ(2)
− , (A20)
r4 (2M − r)
  ′ℓ 
2 2 2 2 2 2 2 ′′ ℓ
 ℓ
2iãM 2 ωuℓ(4) iã mM −2 6M − 5M r + r u (4) + 2M Λ + r ω u(4) + r(r − 2M ) u (4)
ηℓ = − , (A21)
2M r2 − r3 Λr4 (2M − r)
ã2 M 2 h
ℓ ℓ ℓ
ρℓ = 4
12M 2 u′ (3) − 2M r2 ω 2 uℓ(3) + 2iΛM rωuℓ(1) − 2M rℓ2 u′ (2) − 2ℓM ru′ (2) + 3Λ(2M − r)uℓ(2)
Λr (2M − r)
i
ℓ ℓ ℓ ℓ ℓ
−10M ru′(3) − r(r − 2M )2 u′′ (3) + r2 ℓ2 u′ (2) + ℓr2 u′ (2) + 2r2 u′ (3) , (A22)
4ã2 M 3 uℓ(4)
λℓ = − , (A23)
r5
σℓ = 0 , (A24)
2 2
ã M h 
ℓ ℓ
i
uℓ(4) ℓ(ℓ + 1)(2M − r) + 2M r2 ω 2 + (2M − r) r(2M − r)u′′ (4) − 8M u′ (4) .

γℓ = (A25)
Λr4 (2M − r)
(A26)

Appendix B: Proca equation for a slowly rotating first order in ã. Equation (55) reads
Kerr BH
4ãM 2 mω ℓ
D̂2 uℓ(4) − u(4)
r3
6ãM 2 h 
ℓ−1 ℓ−1 ′ ℓ−1

= (ℓ + 1)Q ℓ F u (1) − i rωu (2) − F ru (1)
r4 
In this Appendix we list the perturbation equations ℓ+1 ℓ+1 ′ ℓ+1
i
for a Proca field on a slowly rotating Kerr background to +ℓQℓ+1 i rωu(2) − F u(1) + F ru (1) , (B1)
26

where the operator D̂2 is defined in Eq. (24). The Lorenz R(r) = Ψ/r we get
condition (Eq. (53) with I = 2) becomes  
d dR
r2 F r2 F + r4 ω 2 − 4ãmrω

dr dr
 

i rωuℓ(1) + F uℓ(2) − uℓ(3) + ru′ (2)
2M s2
 
2ãM 2 m  ℓ rω ℓ  −r2 F ℓ(ℓ + 1) + µ2 r2 − R = 0. (C1)
− i u (1) + u r
r2 Λ (3)
2 h
2i ãM ω i For s = 0, the equation above is equivalent to Eq. (7)
= (ℓ + 1)Qℓ uℓ−1
(4) − ℓQ ℓ+1 u ℓ+1
(4) . (B2) in Ref. [60] at first order. For s = ±1 and ã = 0 it is

equivalent to Eq. (51) in Ref. [16].
Using Eq. (B2), Eqs. (53) with I = 0, 1 and Eq. (54) read Following Starobinski’s method of matching asymp-
totics [92], we first expand Eq. (C1) for r ≫ M :
2F ℓ(ℓ + 1) ℓ 2ãM 2 m h
D̂2 uℓ(3) + u + rω(3uℓ(2) − 2uℓ(3) ) d2 (rR) 2 2M µ2 ℓ(ℓ + 1)
 
2 (2) 4
r r ω − µ 2
+ − rR = 0 . (C2)


i dr2 r r2
+3i F uℓ(1) − ru′ (1) = 0 , (B3)
2F

3M
h i The solution of this equation with the correct boundary
D̂2 uℓ(2) − 2 1 − uℓ(2) − uℓ(3) condition at infinity reads
r r
2ãM m 2 h
ℓ ℓ ℓ
i R∞ (x) ∝ xℓ e−x/2 U (ℓ + 1 − ν, 2ℓ + 2, x) , (C3)
− ℓ(ℓ + 1)(2rωu (2) − 3i F u (1) ) − 3rωF u (3)
ℓ(ℓ + 1)r4
where U (p, q, x) is one of the confluent hypergeomet-
6i ãM 2 F ω h ℓ−1 ℓ+1
i
ric functions [93], x = 2k∞ r, k∞ 2
= µ2 − ω 2 and
=− (ℓ + 1)Q ℓ u (4) − ℓQ ℓ+1 u (4) , (B4)
ℓ(ℓ + 1)r3 2
ν = M µ /k∞ . The small distance (x ≪ 1) behavior
2M  ℓ
 of the solution above reads
D̂2 uℓ(1) − 2 i ωuℓ(2) + F u′ (1)
r (2ℓ + 1 + n)!
2ãM 2 m h R∞ (r) ∼ (−1)n (2k∞ r)ℓ
(2ℓ + 1)!
i
ℓ ℓ ′ℓ
− ℓ(ℓ + 1)(2rωu (1) − i u (2) ) + i rF u (3)
ℓ(ℓ + 1)r4 + (−1)n+1 δν(2ℓ)!(2k∞ r)−ℓ−1 , (C4)
2
2ãM F h 
ℓ−1 ′ ℓ−1

=− (ℓ + 1)Q ℓ 2ℓu (4) + ru (4) where we have defined δν = ν − ℓ − 1 − n and assumed
ℓ(ℓ + 1)r4
 i δν ∼ x2ℓ+1 ≪ 1 [60].
′ ℓ+1
+ℓQℓ+1 2(ℓ + 1)uℓ+1 (4) − ru (4) . (B5) On the other hand, Eq. (C1) can be solved analytically
also when r ≪ max(ℓ/ω, ℓ/µ). In this limit, by intro-
The system (B1), (B3)–(B5) can be greatly simplified by ducing a new coordinate z = (r − r+ )/r+ , the equation
a systematic use of Eq. (B2) to eliminate uℓ(1) and uℓ±1 reduces to
(1) .
Solving Eq. (B2) for uℓ(1) and substituting in Eq. (B3)–
 
d dR
+ P 2 − ℓ(ℓ + 1)Z + s2 z R = 0 . (C5)
 
Z Z
(B4) we get, to first order in ã, the equations listed in dz dz
the main text, namely Eqs.(59) and (60).
In order to simplify the axial equation (B1), we con- We defined Z = z(z + 1) and
sider Eq. (B2) with ℓ − 1 and ℓ + 1 and solve it for uℓ−1
(1) P = −2M kH = −2M (ω − mΩH ) , (C6)
and uℓ+1
(1) , respectively. To first order in ã, and making
also use of Eq. (59) and (60), we get Eq. (61). and we neglected O(ã2 ) terms in P 2 . The general so-
lution of Eq. (C5) is a combination of hypergeometric
functions. By imposing ingoing waves at the horizon, i.e.
Appendix C: Analytical results for the axial modes R ∼ z iP at z ∼ 0, we get the general solution with the
correct boundary condition:
In this appendix we generalize Detweiler’s calcula- RH (r)∝ e−2P π (−1)2iP z iP ×
tion [60] of the unstable massive scalar modes of a Kerr 2 F1 [−ℓ + i P + σ, 1 + ℓ + i P + σ, 1 + 2iP, −z] ,
BH to the case of a massive vector field. We focus on
the axial sector and, unlike Detweiler, we work in the where 2 F√1 (a, b, c, z) is the hypergeometric function [93]
slow-rotation limit up to first order in ã. and σ = s2 − P 2 . The large-distance limit of the equa-
As discussed in the main text, the axial vector equa- tion above reads
tion (65) and the equation for scalar perturbations (66) 
(2M )−ℓ Γ[1 + 2ℓ]

can be written in a unified form, Eq. (66). Thus, in RH (r) ∼ rℓ
order to include both scalar and axial vector perturba- Γ [1 + ℓ + i P − σ] Γ [1 + ℓ + i P + σ]
(2M )1+ℓ Γ[−1 − 2ℓ]
 
tions, we shall solve Eq. (66) in the Kerr case, where
+ r−ℓ−1 . (C7)
F = B = 1 − 2M/r and ̟ = 2ãM 2 /r. By defining Γ [−ℓ + i P − σ] Γ [−ℓ + i P + σ]
27

The key point of the method of matching asymptotics prefactor. We find


is that, when |M ω| ≪ ℓ and M µ ≪ ℓ, there exists an
(s=±1)
overlapping region where Eqs. (C7) and (C4) are both ωI (1 + ℓ)2
valid. By equating the coefficients of rℓ and r−ℓ−1 one = , (C13)
(s=0)
ωI ℓ2
can fix the ratio of the overall factors and the frequency
of the mode. Finally, we get which is independent of µ, ã and n. The instability
timescale of the ℓ = 1 axial vector mode is 4 times shorter
(4k∞ M )2ℓ+1 Γ[−1 − 2ℓ]Γ[2 + n + 2ℓ] than that of the scalar mode. Equation (C12) is com-
δν = × pared to the numerical results in Fig. 5.
Γ[1 + n]Γ[1 + 2ℓ]2 Γ[2 + 2ℓ]
Γ [1 + iP − σ + ℓ] Γ [1 + i P + σ + ℓ]
. (C8) On the analytical result in the massive scalar case
Γ [iP − σ − ℓ] Γ [iP + σ − ℓ]

Although this is not directly related to the Proca prob-


As in the case of massive scalar perturbations, the real
lem, we wish to remark that when s = 0 our equa-
and imaginary parts of the mode are related to µ and δν
tion (C11) differs from Eq. (28) in Ref. [60] (expanded
by [60]
at first order) by a factor 2 for any ℓ and any n. We be-
2 lieve this is either due to an inconsistent limit (explained
below) or to a typo in Ref. [60] which propagates from


µ2 − ω R
2
= µ2 , (C9) Eqs. (21)-(24) to Eq. (25), where a factor 1/2 is missing.
ℓ+n+1
 3 Thus, Eq. (29) in Detweiler should read γ = µã(µM )8 /48
δν Mµ for consistency with our Eq. (C12).
i ωI = . (C10)
M ℓ+n+1 In order to understand the source of this discrep-
ancy we refer the reader to the work by Furuhashi and
In the slow-rotation limit we can further simplify Nambu [94], who extended Detweiler’s calculation to the
Eq. (C8). To first order in P we get case of Kerr-Newman BHs, reproducing the results of
Ref. [60] in the zero-charge limit. Ref. [94] presents a
detailed derivation of the instability timescale, facilitat-
iP (4k∞ M )2ℓ+1 Γ[2 + 2ℓ + n] ing intermediate comparisons with our own calculation.
δν∼ ×
Γ[1 + 2ℓ]2 Γ[2 + 2ℓ]2 Γ[1 + n] We note that the results in Ref. [94] seem to be par-
s   tially flawed because the limit defined in Eq. (23) in
Y j−1−s−ℓ
Γ[1 + ℓ − s]Γ[1 + ℓ + s]Γ[1 + ℓ]2 , Ref. [94], Γ(−n)/Γ(−m) = (−1)n−m m!/n!, is valid only
s−j −ℓ
j=1 when n and m are integers. However, in a rotating back-
(C11) ground the angular eigenvalues are not (in general) in-
tegers. For example, in the scalar case the separation
constant λ = ℓ(ℓ + 1) + O(ãM ω) [60]. In the M ω ≪ 1
When s = 0 we recover the result for the scalar case [60]
limit, this can be accounted for by considering ℓ → ℓ + ǫ,
to first order in P (but see the subsection below for a
with ǫ ≪ 1. When s = 0, the evaluation of Eq. (C8)
discussion). In addition, our results correctly reduce to
above involves limits of the form
those found in Ref. [16] in the nonrotating limit. We can
now evaluate the instability timescale. The fundamental Γ(−2ℓ − 2ǫ − 1) (−1)ℓ+1 ℓ!
mode reads lim = , (C14)
ǫ→0 Γ(−ℓ − ǫ) 2 (2ℓ + 1)!

(ℓ=1) (s + 1)!(s + 1) which differs by a factor 2 from Eq. (23) in Ref. [94] (see
M ωI = (ãm − 2r+ µ) (M µ)9 , also the discussion in Ref. [95]). We believe that this is
48
(C12) the reason for our factor 2 discrepancy with respect to
2 Furuhashi and Nambu (and presumably also with respect
(ℓ=2) [(2 − s)!(2 + s)!]
M ωI = (ãm − 2r+ µ) (M µ)13 , to Detweiler’s calculation). Despite numerical difficulties
885735 in the M µ ≪ 1 limit, our numerical calculations for ℓ =
m = 1 are in better agreement with a prefactor of 1/48,
for ℓ = 1 and ℓ = 2, respectively. Note that ωI ∼ µ4ℓ+5 as computed in Eq. (C12): see also Sec. VI F and Fig. 6
for any value of s, so the only difference arises in the in Ref. [35].

[1] H.-P. Nollert, Class.Quant.Grav. 16, R159 (1999). [3] V. Ferrari and L. Gualtieri, Gen.Rel.Grav. 40, 945
[2] K. D. Kokkotas and B. G. Schmidt, Living Rev.Rel. 2, 2 (2008), arXiv:0709.0657 [gr-qc].
(1999), arXiv:gr-qc/9909058 [gr-qc]. [4] E. Berti, V. Cardoso, and A. O. Starinets,
28

Class.Quant.Grav. 26, 163001 (2009), arXiv:0905.2975 arXiv:hep-th/0502206 [hep-th].


[gr-qc]. [35] S. R. Dolan, Phys.Rev. D76, 084001 (2007),
[5] R. Konoplya and A. Zhidenko, Rev.Mod.Phys. 83, 793 arXiv:0705.2880 [gr-qc].
(2011), arXiv:1102.4014 [gr-qc]. [36] J. Rosa, JHEP 1006, 015 (2010), arXiv:0912.1780 [hep-
[6] P. Amaro-Seoane, J. R. Gair, M. Freitag, M. Cole- -th].
man Miller, I. Mandel, et al., Class.Quant.Grav. 24, [37] V. Cardoso, S. Chakrabarti, P. Pani, E. Berti,
R113 (2007), arXiv:astro-ph/0703495 [ASTRO-PH]. and L. Gualtieri, Phys.Rev.Lett. 107, 241101 (2011),
[7] L. Barack, Class.Quant.Grav. 26, 213001 (2009), arXiv:1109.6021 [gr-qc].
arXiv:0908.1664 [gr-qc]. [38] T. Zouros and D. Eardley, Annals Phys. 118, 139 (1979).
[8] E. Poisson, A. Pound, and I. Vega, Living Rev.Rel. 14, [39] S. Hawking, Mon.Not.Roy.Astron.Soc. 152, 75 (1971).
7 (2011), arXiv:1102.0529 [gr-qc]. [40] Y. B. Zel’Dovich and I. D. Novikov, Astron.Zh. 43, 758
[9] E. Berti, V. Cardoso, J. A. Gonzalez, U. Sperhake, (1966).
M. Hannam, et al., Phys.Rev. D76, 064034 (2007), [41] B. J. Carr and S. W. Hawking, Mon.Not.Roy.Astron.Soc.
arXiv:gr-qc/0703053 [GR-QC]. 168, 399 (1974).
[10] E. Berti, V. Cardoso, T. Hinderer, M. Lemos, [42] A. Arvanitaki, S. Dimopoulos, S. Dubovsky, N. Kaloper,
F. Pretorius, et al., Phys.Rev. D81, 104048 (2010), and J. March-Russell, Phys.Rev. D81, 123530 (2010),
arXiv:1003.0812 [gr-qc]. arXiv:0905.4720 [hep-th].
[11] V. Cardoso, L. Gualtieri, C. Herdeiro, U. Sperhake, P. M. [43] A. Arvanitaki and S. Dubovsky, Phys.Rev. D83, 044026
Chesler, et al., (2012), arXiv:1201.5118 [hep-th]. (2011), arXiv:1004.3558 [hep-th].
[12] S. Chandrasekhar, The mathematical theory of black [44] N. Yunes, P. Pani, and V. Cardoso, Phys.Rev. D85,
holes (Oxford University Press, New York, 1992). 102003 (2012), arXiv:1112.3351 [gr-qc].
[13] S. A. Teukolsky, Astrophys.J. 185, 635 (1973). [45] J. Alsing, E. Berti, C. M. Will, and H. Zaglauer,
[14] M. Durkee and H. S. Reall, Class.Quant.Grav. 28, 035011 Phys.Rev. D85, 064041 (2012), arXiv:1112.4903 [gr-qc].
(2011), arXiv:1009.0015 [gr-qc]. [46] H. Kodama and H. Yoshino, (2011), arXiv:1108.1365
[15] G. Torres del Castillo and G. Silva-Ortigoza, Phys.Rev. [hep-th].
D42, 4082 (1990). [47] H. Yoshino and H. Kodama, Prog.Theor.Phys. 128, 153
[16] J. G. Rosa and S. R. Dolan, Phys.Rev. D85, 044043 (2012), arXiv:1203.5070 [gr-qc].
(2012), arXiv:1110.4494 [hep-th]. [48] G. Mocanu and D. Grumiller, Phys.Rev. D85, 105022
[17] Y. Kojima, Phys.Rev. D46, 4289 (1992). (2012), arXiv:1203.4681 [astro-ph.CO].
[18] Y. Kojima, Astrophys.J. 414, 247 (1993). [49] M. Goodsell, J. Jaeckel, J. Redondo, and A. Ringwald,
[19] Y. Kojima, Progress of Theoretical Physics 90, 977 JHEP 0911, 027 (2009), arXiv:0909.0515 [hep-ph].
(1993). [50] J. Jaeckel and A. Ringwald, Ann.Rev.Nucl.Part.Sci. 60,
[20] M. Bruni, L. Gualtieri, and C. F. Sopuerta, 405 (2010), arXiv:1002.0329 [hep-ph].
Class.Quant.Grav. 20, 535 (2003), arXiv:gr-qc/0207105 [51] P. G. Camara, L. E. Ibanez, and F. Marchesano, JHEP
[gr-qc]. 1109, 110 (2011), arXiv:1106.0060 [hep-th].
[21] S. Chandrasekhar and V. Ferrari, Proc.Roy.Soc.Lond. [52] A. S. Goldhaber and M. M. Nieto, Rev.Mod.Phys. 82,
A433, 423 (1991). 939 (2010), arXiv:0809.1003 [hep-ph].
[22] N. Yunes and F. Pretorius, Phys.Rev. D79, 084043 [53] D. Gal’tsov, G. Pomerantseva, and G. Chizhov,
(2009), arXiv:0902.4669 [gr-qc]. Sov.Phys.J. 27, 697 (1984).
[23] K. Konno, T. Matsuyama, and S. Tanda, [54] C. Herdeiro, M. O. Sampaio, and M. Wang, Phys.Rev.
Prog.Theor.Phys. 122, 561 (2009), arXiv:0902.4767 D85, 024005 (2012), arXiv:1110.2485 [gr-qc].
[gr-qc]. [55] R. Konoplya, Phys.Rev. D73, 024009 (2006),
[24] P. Pani and V. Cardoso, Phys.Rev. D79, 084031 (2009), arXiv:gr-qc/0509026 [gr-qc].
arXiv:0902.1569 [gr-qc]. [56] L. Brenneman, C. Reynolds, M. Nowak, R. Reis,
[25] P. Pani, C. F. Macedo, L. C. Crispino, and V. Cardoso, M. Trippe, et al., Astrophys.J. 736, 103 (2011),
Phys.Rev. D84, 087501 (2011), arXiv:1109.3996 [gr-qc]. arXiv:1104.1172 [astro-ph.HE].
[26] K. Yagi, N. Yunes, and T. Tanaka, Phys.Rev. D86, [57] S. Schmoll, J. Miller, M. Volonteri, E. Cackett,
044037 (2012), arXiv:1206.6130 [gr-qc]. C. Reynolds, et al., Astrophys.J. 703, 2171 (2009),
[27] N. Andersson, Astrophys.J. 502, 708 (1998), arXiv:0908.0013 [astro-ph.HE].
arXiv:gr-qc/9706075 [gr-qc]. [58] E. Leaver, Proc.Roy.Soc.Lond. A402, 285 (1985).
[28] Y. Kojima, Mon.Not.Roy.Astron.Soc. 293, 49 (1998), [59] P. Pani, V. Cardoso, L. Gualtieri, E. Berti, and
arXiv:gr-qc/9709003 [gr-qc]. A. Ishibashi, (2012), to appear in Physical Review Let-
[29] K. H. Lockitch and J. L. Friedman, Astrophys.J. 521, ters.
764 (1999), arXiv:gr-qc/9812019 [gr-qc]. [60] S. L. Detweiler, Phys. Rev. D22, 2323 (1980).
[30] K. H. Lockitch, N. Andersson, and J. L. Friedman, [61] http://blackholes.ist.utl.pt/?page=Files,
Phys.Rev. D63, 024019 (2001), arXiv:gr-qc/0008019 http://www.phy.olemiss.edu/~ berti/qnms.html.
[gr-qc]. [62] J. B. Hartle, Astrophys.J. 150, 1005 (1967).
[31] W. H. Press and S. A. Teukolsky, Nature 238, 211 (1972). [63] V. Ferrari, L. Gualtieri, and S. Marassi, Phys.Rev. D76,
[32] T. Damour, N. Deruelle, and R. Ruffini, Lett.Nuovo 104033 (2007), arXiv:0709.2925 [gr-qc].
Cim. 15, 257 (1976). [64] E. Berti, V. Cardoso, and M. Casals, Phys.Rev. D73,
[33] V. Cardoso, O. J. Dias, J. P. Lemos, and S. Yoshida, 024013 (2006), arXiv:gr-qc/0511111 [gr-qc].
Phys.Rev. D70, 044039 (2004), arXiv:hep-th/0404096 [65] U. Gerlach and U. Sengupta, Phys.Rev. D22, 1300
[hep-th]. (1980).
[34] V. Cardoso and S. Yoshida, JHEP 0507, 009 (2005), [66] S. Teukolsky and W. Press, Astrophys.J. 193, 443 (1974).
29

[67] E. Berti, F. White, A. Maniopoulou, and [84] N. J. McConnell, C.-P. Ma, K. Gebhardt, S. A. Wright,
M. Bruni, Mon.Not.Roy.Astron.Soc. 358, 923 (2005), J. D. Murphy, T. R. Lauer, J. R. Graham, and
arXiv:gr-qc/0405146 [gr-qc]. D. O. Richstone, Nature (London) 480, 215 (2011),
[68] K. H. Lockitch and N. Andersson, Class.Quant.Grav. 21, arXiv:1112.1078 [astro-ph.CO].
4661 (2004), arXiv:gr-qc/0106088 [gr-qc]. [85] N. J. McConnell, C.-P. Ma, J. D. Murphy, K. Gebhardt,
[69] E. W. Leaver, Phys.Rev. D41, 2986 (1990). T. R. Lauer, J. R. Graham, S. A. Wright, and D. O.
[70] C. Simmendinger, A. Wunderlin, and A. Pelster, Phys. Richstone, ArXiv e-prints (2012), arXiv:1203.1620 [as-
Rev. E 59, 5344 (1999), arXiv:math-ph/0106011. tro-ph.CO].
[71] K. S. Thorne, Astrophys. J. 158, 1 (1969). [86] J. Beringer et al. (Particle Data Group), Phys.Rev. D86,
[72] S. Chandrasekhar and V. Ferrari, Proc.Roy.Soc.Lond. 010001 (2012).
A437, 133 (1992). [87] H. A. Buchdahl, J.Phys.A A12, 1235 (1979).
[73] E. Berti, V. Cardoso, and P. Pani, Phys.Rev. D79, [88] V. Vitagliano, T. P. Sotiriou, and S. Liberati, Phys.Rev.
101501 (2009), arXiv:0903.5311 [gr-qc]. D82, 084007 (2010), arXiv:1007.3937 [gr-qc].
[74] R. Ruffini, Black Holes: les Astres Occlus (Gordon and [89] H. Sotani, K. Kokkotas, and N. Stergioulas,
Breach Science Publishers, 1973). Mon.Not.Roy.Astron.Soc. 375, 261 (2007),
[75] D. Merritt and M. Milosavljevic, Living Rev.Rel. 8, 8 arXiv:astro-ph/0608626 [astro-ph].
(2005), arXiv:astro-ph/0410364 [astro-ph]. [90] A. Stavridis, A. Passamonti, and K. Kokkotas,
[76] A. Sesana, F. Haardt, and P. Madau, Astrophys.J. 660, Phys.Rev. D75, 064019 (2007), arXiv:gr-qc/0701122
546 (2007), arXiv:astro-ph/0612265 [astro-ph]. [gr-qc].
[77] E. Barausse, Mon.Not.Roy.Astron.Soc. 423, 2533 (2012), [91] A. Passamonti, A. Stavridis, and K. Kokkotas,
arXiv:1201.5888 [astro-ph.CO]. Phys.Rev. D77, 024029 (2008), arXiv:0706.0991 [gr-qc].
[78] E. Berti and M. Volonteri, Astrophys.J. 684, 822 (2008), [92] A. A. Starobinsky, Zh. Eksp. Teor. Fiz 64, 48 (1973), sov.
arXiv:0802.0025 [astro-ph]. Phys. JETP 37, 28 (1973).
[79] J. M. Bardeen, Nature 226, 64 (1970). [93] M. Abramowitz and I. A. Stegun, Handbook of Mathe-
[80] K. S. Thorne, Astrophys.J. 191, 507 (1974). matical Functions with Formulas, Graphs, and Mathe-
[81] C. F. Gammie, S. L. Shapiro, and J. C. McKinney, As- matical Tables, ninth dover printing, tenth gpo printing
trophys.J. 602, 312 (2004), arXiv:astro-ph/0310886 [as- ed. (Dover, New York, 1964).
tro-ph]. [94] H. Furuhashi and Y. Nambu, Prog.Theor.Phys. 112, 983
[82] A. R. King and J. Pringle, Mon.Not.Roy.Astron.Soc.Lett. (2004), arXiv:gr-qc/0402037 [gr-qc].
373, L93 (2006), arXiv:astro-ph/0609598 [astro-ph]. [95] J. G. Rosa, (2012), arXiv:1209.4211 [hep-th].
[83] R. Harnik, J. Kopp, and P. A. Machado, JCAP 1207,
026 (2012), arXiv:1202.6073 [hep-ph].

You might also like