Extremal Black Holes in Dynamical Chern-Simons Gravity

Robert McNees,1 Leo C. Stein,2 and Nicolás Yunes3 1Loyola University Chicago, Department of Physics, Chicago, IL 60660, USA. 2TAPIR, Walter Burke Institute for Theoretical Physics, California Institute of Technology, Pasadena, CA 91125, USA 3Department of Physics, Montana State University, Bozeman, MT 59717, USA. (Dated: December 18, 2015) Rapidly rotating black hole solutions in theories beyond play a key role in experimental gravity, as they allow us to compute observables in extreme that deviate from the predictions of general relativity (GR). Such solutions are often difficult to find in beyond- GR theories due to the inclusion of additional fields that couple to the metric non-linearly and non-minimally. In this paper, we consider rotating black hole solutions in one such theory, dynamical Chern-Simons gravity, where the Einstein-Hilbert action is modified by the introduction of a dynamical scalar field that couples to the metric through the Pontryagin density. We treat dynamical Chern- Simons gravity as an effective field theory and thus work in the decoupling limit, where corrections are treated as small perturbations from general relativity. We perturb about the maximally-rotating Kerr solution, the so-called extremal limit, and develop mathematical insight into the analysis techniques needed to construct solutions for generic spin. First we find closed-form, analytic expressions for the extremal scalar field, and then determine the trace of the metric perturbation, giving both in terms of Legendre decompositions. Retaining only the first three and four modes in the Legendre representation of the scalar field and the trace respectively suffices to ensure a fidelity of over 99% relative to full numerical solutions. The leading-order mode in the Legendre expansion of the trace of the metric perturbation contains a logarithmic divergence at the extremal Kerr horizon, which is likely to be unimportant as it occurs inside the perturbed dynamical Chern-Simons horizon. The techniques employed here should enable the construction of analytic, closed-form expressions for the scalar field and metric perturbations on a background with arbitrary rotation.

PACS numbers: 04.30.-w,04.50.Kd,04.25.-g,04.25.Nx

I. INTRODUCTION ing BH solutions have not yet been found is dynamical Chern-Simons (dCS) gravity [3]. This theory modifies Einstein’s theory of general relativity has passed a the Einstein-Hilbert action by introducing a dynamical plethora of Solar System and binary pulsar tests [1], but (pseudo) scalar field that couples non-minimally to the it has not been tested in depth in the extreme gravity metric through the Pontryagin density. This interaction regime [2] where the gravitational interaction is non-linear leads to a scalar field evolution equation that is sourced and dynamical. A number of new observations will al- by the Pontryagin density, and modified metric field equa- low us to test this regime of Einstein’s theory: in the tions with third derivatives. The latter have cast doubt gravitational wave spectrum through advanced LIGO and on whether full dCS is well-posed as an initial value prob- its partners, when compact objects collide; in the radio lem [4], and also on whether stable BH solutions exist. spectrum with the Event Horizon Telescope, when an We take the point of view that dCS must be treated as an accretion disk illuminates its host black hole and creates effective field theory, since it is motivated from the low- a ‘shadow’; and in the X-ray spectrum with the Chandra energy limit of compactified heterotic string theory [5, 6] Telescope, when gas heats up and glows as it accretes in (for a review see [3]), from effective field theories of in- arXiv:1512.05453v1 [gr-qc] 17 Dec 2015 the black hole . Such future observations will flation [7], and from loop quantum gravity [8]. Thus the either confirm Einstein’s theory at unprecedented levels theory is treated in the decoupling limit, where deforma- or reveal new phenomena in the extreme gravity regime. tions from GR are treated perturbatively, order-reducing Solutions that represent rotating black holes (BHs) in the field equations. theories of gravity beyond general relativity (GR) are an When treated as an effective theory, BH solutions in essential ingredient of tests in the extreme gravity regime. dCS have been found in certain limits. Jackiw and Pi [9] Constraining these theories requires a metric with which showed that the Schwarzschild metric is also a solution to calculate observables. Once a metric is available, one in dCS that represents a non-rotating BH. Linear sta- can investigate the linear and non-linear stability of the bility of high-frequency waves about the Schwarzschild solution through a mode analysis, calculate the gravita- background was suggested by [10]. The axionic hair on tional waves emitted as two BHs inspiral, compute the a slowly-rotating BH was first found in [11]. The metric ‘shadow’ cast by a BH when illuminated by an accretion solution to linear order in spin was found contempora- disk, and determine the energy spectrum of the radiation neously and independently by Yunes and Pretorius [12], emitted by gas accreting into the BH. and Konno, et al. [13]. This was expanded to quadratic One beyond-GR gravity theory in which generic rotat- order by Yagi, Yunes and Tanaka [14]. The first rapidly- 2 rotating dCS BH studies were carried out by Konno and decomposition technique we employed in the extremal Takahashi [15] and Stein [16] who investigated the behav- limit is also applicable to sub-extremal BHs [16], and if ior of the dynamical scalar field about a rapidly rotating the tensor equations admit decoupling and separation of Kerr background. Stein also investigated the trace of the variables, could be used to find the full metric deforma- metric perturbation, and found that the extremal limit tion. may be singular, which partly motivated the present work. The remainder of this paper presents details of the At present, nobody has succeeded in constructing the full techniques developed, the solutions obtained and their metric of generic rotating BHs in dCS gravity, despite properties. Henceforth, we use the following conventions. over two decades of work in that direction [9, 11–19]. Latin letters (a, b, c, . . .) in index lists stand for abstract As a first step toward finding full rotating BH solutions indices. Parentheses and square brackets in index lists in effective dCS gravity, we study the extremal limit, stand for symmetrization and anti-symmetrization respec- i.e. the limit in which the BH spin is close to the maximal tively. The metric signature will be (−, +, +, +) and we Kerr value. The extremal limit is of interest not only choose units in which c = 1. However, we do not set G because the mathematics simplify significantly, but also or h to unity. All other conventions follow the standard because of the Kerr/CFT conjecture that posits a dual treatment of [26, 27]. holographic description in terms of a two-dimensional conformal field theory [20–25]. Working in the extremal limit, we have found a general, closed-form expression for II. THE ABC OF DCS the Legendre mode-decomposed scalar field. The radial structure of the scalar field is more complicated than that Dynamical Chern-Simons gravity [3, 9] is a four- of the slowly-rotating case. Whereas the slow-rotation dimensional theory defined by the action case only requires a finite polynomial expansion, the rapid- rotation case is characterized by natural logarithms and I = IEH + ICS + Iϑ + IMat . (1) arctangents. The angular structure of the scalar field is still predominantly dipolar. We find that retaining only The first term is the Einstein-Hilbert action the first 3 non-vanishing modes of the scalar field suffices Z   4 √ 1 to achieve a fidelity above 99% in the entire domain IEH = d x −g R , (2) 2κ2 relative to the full numerical solution. With analytic, closed-form expressions for the extremal where κ2 = 8πG, R is the Ricci scalar associated with the scalar field in hand, we then focus on the BH metric, treat- gµν , and g is the metric determinant. The ing the dCS correction as a small perturbation of the Kerr last term in Eq. (1) is the action for all matter degrees background. Using a convenient (harmonic) gauge, we of freedom, which couple minimally to the metric tensor show that the modified field equations for the trace of the and do not couple to ϑ. metric perturbation can be solved in quadrature in terms The Chern-Simons correction is mediated by a of another Legendre mode decomposition. As in the scalar canonically-normalized scalar field ϑ, whose kinetic term field case, the radial structure of the metric perturbation in the action is is quite different from that of slowly-rotating BHs, with Z √  1  a logarithmic divergence at the extremal Kerr horizon for I = d4x −g − (∂ ϑ)(∂aϑ) . (3) the dominant monopole mode. This divergence confirms ϑ 2 a the one which was conjectured in [16] by one of the present authors. This divergence, however, may be unphysical This scalar field couples non-minimally to the metric since the Kerr horizon is likely “inside” the perturbed dCS through the potential term in the action horizon. The angular structure of the trace of the metric Z   4 √ 1 α ∗ perturbation is still predominantly monopolar. We find ICS = d x −g − ϑ RR , (4) that retaining only the first 4 non-vanishing modes of 4 κ the trace of the metric perturbation suffices to achieve a where the Pontryagin density is defined via fidelity above 99% in the entire domain relative to the full numerical solution. 1 ∗RR := ∗RabcdR = abef R cd R , (5) The results we have obtained have various consequences abcd 2 ef abcd for the study of BHs in beyond-GR gravity theories. The logarithmic divergence of the trace of the metric suggests and abcd is the Levi-Civita tensor. Notice that the defini- a conjecture: that the dCS-corrected horizon is “outside” tion of the Pontryagin density here differs from that of [3] of the Kerr horizon for all possible values of angular mo- by a minus sign, which is compensated by an additional mentum, protecting against a naked singularity. Further, minus sign in ICS. our results suggest that generic rotating BH solutions Variation of the action with respect to the metric yields in beyond-GR theories will not have the simple rational- the field equations polynomial form that the Kerr metric enjoys, and may 2 require more complicated functional forms. The mode- Gab + 2 ακ Cab = κ Tab , (6) 3 where Gab is the , and the traceless ‘C- the observation and its uncertainties. This constraint can tensor’ is defined as then be mapped to a constraint on α to find p|α| < δ1/4L. ab cde(a b) ∗ d(ab)c Currently, the best constraint on the dCS coupling param- C = (∇cϑ)  ∇eR d + (∇c∇dϑ) R . (7) p 8 eter is |α| . 10 km and it comes from observations of The stress-energy tensor decomposes linearly into a term Lense-Thirring precession from satellites in orbit around that depends only on the matter degrees of freedom and Earth [30]. Such a weak constraint makes sense when one a term that depends only on the scalar field, i.e. Tab = realizes that for these kind of experiments the character- Mat ϑ 3 1/2 8 Tab + Tab, where the latter is istic length scale L = [R⊕/(GM⊕)] ≈ 2 × 10 km. The aforementioned theoretical motivations suggest ϑ 1 c Tab := (∇aϑ)(∇bϑ) − gab (∇cϑ)(∇ ϑ) . (8) that one treat dCS as an effective theory valid up to 2 some cut-off scale, i.e., the scale above which higher-order Variation of the action with respect to the scalar field curvature terms in the action cannot be neglected [2]. We yields its evolution equation will here restrict attention to physical scenarios in which

α ∗ the effective theory is valid, and since we are interested ϑ = RR, (9) 4κ only in black holes, this means√ we restrict attention to those with masses GM  α. When this is the case, we where stands for the d’Alembertian operator. Notice  can work in the decoupling limit of the theory, i.e. we that there is no potential associated with the scalar field, perform a perturbative expansion of the field equations which implies it is a long-ranged field. This vanishing and their solutions in powers of ζ. Henceforth, dCS is (or flat) potential means that ϑ retains a global shift exclusively treated in the decoupling limit. symmetry, ϑ → ϑ+const., because ∗RR is related to a The decoupling limit can be implemented in practice topological invariant [28]. Retaining this shift symmetry by expanding the metric tensor and the scalar field in may be important to protect against certain quantum powers of ζ. In this paper, we will expand the metric and corrections. the scalar field as follows: The theoretical motivation to study dCS is varied. From a string-theory standpoint, non-minimal scalar couplings (0) (1/2) (1) g = g + ζ1/2 g + ζ g + O(ζ3/2) , (10) of the form of Eq. (4) arise in the low-energy limit of ab ab ab ab 1 1 1 heterotic string theory upon four-dimensional compactifi- ϑ = ϑ˜(0) + ζ1/2 ϑ˜(1/2) + ζ ϑ˜(1) + O(ζ3/2) , (11) cation [5, 6] (for a review see [3]). From a loop quantum κ κ κ gravity standpoint, dCS arises when the Barbero-Immirzi where the superscript denotes the order in ζ of each parameter is promoted to a scalar field in the presence term. Notice that a factor of κ−1 in the expansion for of fermions [8, 29]. From a cosmology standpoint, the the scalar field ensures that ϑ˜(n) is dimensionless. As interaction in Eq. (4) arises as one of three terms that we are perturbing about ζ = 0, our background solution remain in an effective field theory treatment of single-field (g(0), ϑ˜(0)) must solve the field equations for GR and a free inflation [7]. ˜(0) The choice of conventions made here differs from that massless scalar field. Choosing trivial initial data for ϑ gives ϑ˜(0) = 0 at all times, so we find (g(0), ϑ˜(0)) = (gGR, 0) of [3]. The mapping between the two sets is κ = 1/(2κ2), ab ab AY at zeroth order, where gGR is some known GR solution. β = 1 and α = α/κ. Moreover, we retain all factors ab AY AY If we next examine the system at order ζ1/2, we find that of G, or equivalently of κ, since we do not set G to (1/2) unity. Without requiring the action to have any spe- gab satisfies a homogeneous linear equation due to the ˜(0) cific sets of units, demanding consistency between IEH, vanishing of ϑ . Therefore, again, trivial initial data −1 2 gives g(1/2) = 0 at all times. ICS and Iϑ implies [ϑ] = [κ] and [α] = L , where ab L stands for units of length. Given some GR solution Thus to the order we are working, our expansion is with characteristic length scale L, corrections are then 2 4 GR (1) 3/2 controlled by the dimensionless parameter ζ := α /L . gab = gab + ζ gab + O(ζ ) , (12) −4 One can see this by noting that |∂abϑ| ∝ (α/κ)L from 1 1/2 (1/2) 1 −6 ϑ = 0 + ζ ϑ˜ + O(ζ ) . (13) Eq. (9), which implies that |Cab| ∝ (α/κ)L from Eq. (7). κ Then the fractional corrections to GR are proportional 2 −4 0 to (ακ|Cab|/|Gab|) ∝ α L = ζ. Henceforth, we will focus on BH solutions, with the O(ζ ) GR Current constraints on dCS are rather weak because term in the metric, gab , being simply the Kerr metric. dCS corrections are relevant only in scenarios where the The O(ζ1/2) term in the scalar field, ϑ(1/2), is sourced spacetime curvature is large. One can see this by noting by the Kerr metric and, in turn, this sources the O(ζ) (1) that dCS corrections to the gravitational field are sourced correction to the metric, gab . To be within the regime of by the scalar field, which in turn is only sourced by the validity of the perturbative expansion, we require ζ  1, spacetime curvature. In fact, one can easily show through and since for the Kerr black hole the typical curvature the argument given in the previous paragraph that con- length scale is L = GM, we take straints on the α parameter of dCS will be roughly pro- portional to a power of L. Let us assume that some obser- α2 ζ =  1 . (14) vation places the constraints |ζ| < δ, where δ is related to (GM)4 4

Notice that this definition differs from others in the liter- proportional to Σ(∗RR(0)) and given explicitly by ature [3] in that we do not include a factor of 1/κ2 in ζ, 2 2 2 2 2 2 but rather we factor it out in the scalar field directly. (1/2) χrψ˜ (3˜r − χ ψ )(˜r − 3χ ψ ) s (˜r, ψ) = 24 . (21) In this paper we are concerned with solutions that Σ˜ 5 represent rotating BHs spinning near extremality, so in addition to the decoupling expansion we will also perform Equation (20) admits a solution via separation of vari- a near-extremal expansion. Letting the BH spin angular ables, by expanding the solution momentum be |~J|, we can define the BH dimensionless ∞ ~ 2 ˜(1/2) X ˜(1/2) spin parameter χ := |J|/(GM) . We can then expand ϑ = ϑ` (˜r)P`(ψ) , (22) all fields in the problem in a bivariate expansion, i.e. a `=0 simultaneous expansion in both ζ  1 and χ ∼ 1, namely where P`(·) are Legendre functions of the first kind. The (n) (n,0) (n,1) 2 (n,2) 3 radial modes ϑ˜(1/2)(˜r) then satisfy the equation gab = gab + ε gab + ε gab + O(ε ) , (15) ` ϑ˜(n) = ϑ˜(n,0) + ε ϑ˜(n,1) + ε2 ϑ˜(n,2) + O(ε3) , (16)   ˜ ˜(1/2) ˜(1/2) (1/2) ∂r˜ ∆∂r˜ϑ` − `(` + 1)ϑ` = s` (˜r) , (23) where ε := (1 − χ2)1/2 is a near-extremality parameter, (1/2) i.e. ε  1 for near-extremal BHs. with source functions s` (r˜) given by the modes in the Legendre decomposition of Eq. (21)

Z 1 III. : SOLUTION (1/2) 2` + 1 (1/2) s` (˜r) = dψ P`(ψ)s (˜r, ψ) . (24) 2 −1 We wish to solve the evolution equation for the scalar field [Eq. (9)] to leading order in ζ. To this order, the Note that the source function in Eq. (21) is odd in the Pontryagin density on the right-hand side of Eq. (9) is variable ψ, so its Legendre expansion (as well as that of evaluated on the unmodified Kerr spacetime. The wave the scalar field) will only contain odd modes: ` = 2n + 1 operator on the left-hand side can also be evaluated on for all n ∈ N. the Kerr spacetime, since corrections will be of O(ζ). In The integral in Eq. (24) can be evaluated in closed form polynomial Boyer-Lindquist coordinates, the scalar field in terms of known functions: evolution equation is evaluated on the line element [31] 1 ` `+1 (1/2) Γ( 2 )Γ(` + 4) χ s (˜r) = (−1) 2 ` ` 1 `+4 ∆ 2 Σ 2 Γ(` + ) r˜ g(0)dxadxb = − dt − a Γ dφ + dr2 (17) 2 ab h  2  Σ ∆ × 3 F `+4 , `+5 ; ` + 3 ; − χ 2 2 1 2 2 2 r˜2 Σ 2 Γ  2 2  + dψ + (r + a )dφ − a dt ,  2 i Γ Σ `+4 `+7 3 χ −(` + 5) 2F1 2 , 2 ; ` + 2 ; − r˜2 , (25) where the usual polar angle θ has been replaced with a coordinate ψ = cos θ, and Γ := 1 − ψ2 = sin2 θ. The where 2F1(·, ·; ·; ·) is the ordinary hypergeometric function mass of the black hole is M and it rotates with angular and ` is odd. One can show, via identities for hypergeo- momentum per unit mass a = |~J|/(GM), where −GM ≤ metric functions, that this expression is equivalent to one a ≤ GM. The functions Σ and ∆ are given previously in [32]. The solution of Eq. (23) can be obtained through the Σ = r2 + a2ψ2 (18) method of variation of parameters (see AppendixA). p 2 ∆ = r2 − 2GMr + a2 , (19) Defining a new variable η = (r˜−1)/ 1 − χ , the solution ˜(1/2) of Eq. (23) [see also Eq. (A3)] for the mode function ϑ` so that the background horizons, where ∆ = (r − r+)(r − is p 2 2 r−) = 0, are located at r± = GM ± (GM) − a . Z η It will be convenient to replace all quantities with ˜(1/2) 0 (1/2) 0p 2 0 ϑ` (˜r) = P`(η) dη s` (1 + η 1 − χ ) Q`(η ) dimensionless variables by scaling out factors of GM: ∞ Z η r˜ = r/(GM) and χ = a/(GM), so that the rescaled 0 (1/2) 0p 2 0 ˜ 2 2 2 − Q`(η) dη s` (1 + η 1 − χ ) P`(η ) , (26) functions ∆ = ∆/(GM) = (r˜ − 1) − (1 − χ ) and 1 Σ˜ = Σ/(GM)2 = r˜2 + χ2ψ2. Assuming a stationary and axisymmetric solution for the scalar field, the O(α) term where Q`(·) are Legendre functions of the second kind. in Eq. (9) then takes the form This solution is regular at r˜+, and approaches zero as r˜ → ∞. ˜(1/2) ˜(1/2) (1/2) ∂r˜ ∆˜ ∂r˜ϑ + ∂ψ Γ∂ψϑ = s (˜r, ψ) (20) Our eventual goal is to evaluate Eq. (26) in closed form for the full range of the rotation parameter, −1 ≤ χ ≤ 1. where factors of (α/κ) and (GM) have canceled from The slow rotation limit of the field, i.e. the solution in both sides of the equation. The source s(1/2)(r,˜ ψ) is a |χ|  1 expansion, is already well-understood; it was 5

first derived in [12], verified in [13], and extended to The overall `-dependent factor comes from the definite 0 0 second order in rotation in [14]. Similarly, it is also integration of Q`(η ) in the range η ∈ (1, ∞), while (1/2, 0) possible to systematically solve the scalar field equation s` (1) is Eq. (25) evaluated at χ = 1 and r˜ = 1. of motion in the near-extremal expansion introduced in With these results, we can now give a general expression Sec.II. Expanding the source functions of Eq. (25) for for the radial modes of the scalar field. They take the ε  1, we find form

(1/2) (1/2, 0) 2 (1/2, 2) 4 (1/2, 4) 6 ˜(1/2, 0) s` (˜r) = s` (˜r) + ε s` (˜r) + ε s` (˜r) + O(ε ) . ϑ` (˜r) = A`(˜r) + B`(˜r) arccot(˜r) (31)   (27) r˜ − 1 + C`(˜r) log √ , 2 Recall that the second superscript in each of these terms r˜ + 1 represents the order in ε at which it enters the near- where the functions A`(˜r), B`(˜r), and C`(˜r) are extremal expansion. Because the χ → 1 limit is regular for (1/2) (1/2, 0) (1/2) ` s` (˜r), s` (˜r) is simply s` (˜r) evaluated at χ = 1. `−1 `(` − 1) X k A (˜r) = (−1) 2 γ (˜r − 1) (32) In this paper we will only consider the extremal limit, ` (˜r − 1)`+1 k ε → 0, which is the leading term in the near-extremal k=0 2` + 1 (2` + 1)(4˜r − `(` + 1)) expansion. This corresponds to the limit χ → ±1 of − + the dimensionless spin parameter χ. The homogeneous (˜r2 + 1)2 4 (˜r2 + 1) solutions are regular in this limit [see Eqs. (A8)-(A9)], `−1 X k and the solution for the scalar field [see Eq. (A3)] at ε = 0 + α`,k r˜ is k=0

" r˜ (1/2, 0) ` Z 0 `+1 `(` − 1) (1/2, 0) 1 ` 0 s` (˜r ) X k ϑ˜ (˜r) = (˜r − 1) dr˜ B`(˜r) = (−1) 2 + β`,k r˜ (33) ` 0 `+1 (˜r − 1)`+1 2` + 1 ∞ (˜r − 1) k=0 # Z r˜ `+1 (` + 1)(` + 2) ` 1 0 0 ` (1/2, 0) 0 C`(˜r) =(−1) 2 (˜r − 1) . (34) − `+1 dr˜ (˜r − 1) s` (˜r ) . (28) 2 (˜r − 1) 1 The constants γk appearing in A`(r˜) are the first ` + 1 (1/2, 0) The source functions at leading-order in ε, s` (r˜), are terms in the Taylor expansion of arccot(˜r) around r˜ = 1 given by Eq. (24) or Eq. (25) evaluated at |χ| = 1. The boundary conditions are the same as before: each mode 1  ∂  (1/2, 0) γ = arccot(x) . (35) ϑ˜ is regular at r˜ = 1, and goes to zero as r˜ → ∞. k k ` + k! ∂x The integrals in Eq. (28) can be readily evaluated for x=1 specific values of `. For example, the ` = 1 radial mode The remaining 2`+1 coefficients α`,k and β`,k are fixed by is given by imposing the boundary conditions: each mode falls off as −(`+1)  r˜ − 1  r˜ at large r˜, and takes the value Eq. (30) at r˜ = 1. ϑ˜(1/2, 0)(˜r) = 3(˜r − 1) log √ + 3(˜r − 1) arccotr ˜ Alternately, the condition at r˜ = 1 can be replaced with 1 2 r˜ + 1 the requirement that the leading asymptotic behavior of 3(˜r − 1)(2˜r2 +r ˜ + 3) the mode is given by Eq. (C25). The coefficients α`,k + 2 2 . (29) 2(˜r + 1) and β`,k for the first several modes are given explicitly in With more work, we can also give an expression for general AppendixB. values of ` in terms of finite-order rational polynomials, arccot(r˜), and the log which appears above. But before IV. SCALAR FIELD: PROPERTIES we can give the general form, we first have to establish two results for the behavior of the modes at large r˜ and at r˜ = 1. The far-field behavior of the modes is dominated by Let us now discuss some properties of the scalar field the second integral in Eq. (28), since the first one decays solution obtained in the previous section. We begin by with a higher power of r. This second integral converges plotting the first five (odd) modes in Fig.1. Observe ˜(1/2, 0) −(`+1) that the integrated norm of ϑ˜(1/2, 0) decays exponentially in this limit, and thus, ϑ` ∼ r˜ as r˜  1. The ` near-horizon behavior of the modes is dominated by the with `. This is because this function is a spectral solution first integral in Eq. (28), but its asymptotic behavior to a differential equation with a C∞ source, so it must as r˜ ∼ 1 cannot be easily discerned from that equation. converge exponentially with mode number. Observe also Instead, it is easier to return to Eq. (26) and set χ = 1, from Fig.1 that the ` = 1 mode of the field vanishes at remembering that the horizon limit r˜ → 1 is equivalent to the horizon. Modes with ` > 1 are finite but non-zero√ at η → 1 (this is the case for all values of χ). In this limit, the horizon, with values that scale like `5/2(1 + 2)−(`+1) only the first line of Eq. (26) contributes, leading to for `  1. By including contributions from a sufficient number of 1 ˜(1/2, 0) (1/2, 0) modes, we can construct an arbitrarily accurate approxi- ϑ` (1) = − s` (1) . (30) `(` + 1) mation of the full, extremal scalar field. An approximation 6

0.4

0.2

0

FIG. 1. The first five radial mode functions of the scalar field. The vertical dashed line indicates the location of the event -0.2 horizon of an extremal black hole. The ` = 1 mode vanishes at the horizon, while modes with ` > 1 are all non-zero at r˜ = 1. -0.4 using the first five modes is shown in Fig.2, as a function of both radius and polar angle. The accuracy of such an approximation can be characterized using a slicing- independent measure of the scalar field energy through the ADM energy. Let ua be a timelike unit vector normal to a hypersurface S, with γab the induced metric on S. Then the scalar field’s contribution to the energy is Z 3 √ a ϑ b E = d x γ u Tabt (36) S FIG. 2. The behavior of the scalar field on the extremal back- where tb is the Killing vector ∂/∂t. This energy can be ground, approximated by its first five Legendre modes. The perturbatively expanded in powers of ζ, coordinates rψ˜ and r˜p1 − ψ2 correspond to r˜cos θ and r˜sin θ, respectively, in conventional Boyer-Lindquist coordinates. E = ζ E(1) + ζ2 E(2) + ... (37)

The scalar field’s ADM energy at leading order can further 4 be computed via the spectral decomposition, than one part in 10 . Observe that the accuracy increases exponentially with N. Observe also that if we wish to ∞ capture 99% of the energy in the field, it suffices to keep (1) X ˜(1) E = M Ek , (38) only up to the first three odd modes, i.e. N = 5. Finally, k=1 note that the energy in the scalar field is dominated by the behavior of the scalar field close to the horizon. If with the dimensionless E˜(1) functions given by k one is interested in regimes of spacetime outside some 1 1 Z ∞ h i two-sphere with radius r  M, then the full scalar field ˜(1) ˜ ˜(1/2) 2 ˜(1/2) 2 E = dr˜ ∆(∂r˜ϑ ) + k(k + 1)(ϑ ) . can be accurately modeled using just the dipole (` = 1) k 4 2k + 1 k k r˜+ and octupole (` = 3) modes. (39) The fractional difference between the total energy in the scalar field and the energy in the first N modes is then V. TRACE OF THE METRIC PERTURBATION: N SOLUTION M X (1) δN = 1 − E˜ . (40) E(1) k k=1 The leading correction to the metric in Eq. (12) is Figure3 shows the fractional difference δN for the first determined by the O(ζ) term in the metric equation of seven nonvanishing modes. The contribution from the first motion [Eq. (6)]. We work in a gauge where the covariant (1) five – up to ` = 9 – differs from the total energy by less divergence of gab is proportional to the derivative of its 7

Then, the equation of motion [Eq. (43)] again separates, giving the radial equation ● ● h i ˜ (1) ∂r˜∆∂r˜ − `(` + 1) g` (˜r) = S`(˜r) . (45) ●

● The source functions S`(r˜) are the Legendre modes of the right-hand side of Eq. (42), i.e. ● Z 1 ● 2` + 1 S`(˜r) = dψP`(ψ)Sg(˜r, ψ) , (46) 2 −1 ● where the source function Sg is simply

p (0) ˜(1/2) 2 FIG. 3. The fractional difference between the scalar field’s Sg(˜r, ψ) := −2 −g (∇ϑ ) . (47) (1) contribution to the ADM energy, and the contribution from The solution for the mode functions g` (r˜) is then given the field’s first N modes. by Eq. (A3), which in this case becomes

(1) ab (1) Z r˜ trace g = g(0)gab with respect to the background metric: (1) 1 + 0 − 0 0 g` = H` (˜r) dr˜ H` (˜r )S`(˜r ) W` ∞ a (1) 1 (1) ∇ gab = ∇bg . (41) Z r˜ ! 2 − 0 + 0 0 −H` (˜r) dr˜ H` (˜r )S`(˜r ) . (48) This gauge leads to simplifications in the O(ζ) term of r˜+ Eq. (6), but its structure is still too complicated to allow for a simple solution. As a first step towards determining Note that the source is quadratic in the scalar field, which (1) has odd Legendre modes. Thus, both the trace of the gab , we take the trace of the O(ζ) correction to Eq. (6) to find metric perturbation and its source function have even Legendre modes: ` = 2n for all n ∈ N. ∇2g(1) = −2(∇ϑ(1/2))2 . (42) One approach to evaluating the integrals in Eq. (48) is to express the Legendre modes of the source function in Assuming a stationary and axisymmetric solution and terms of the scalar field modes and their radial derivatives. transforming to dimensionless variables, this reduces to The resulting integrals are significantly more complicated h i than the ones we encountered in Sec.III, so we will opt for ˜ (1) ˜ ˜(1/2) 2 ˜(1/2) 2 ∂r˜∆∂r˜ + ∂ψΓ∂ψ g = −2∆(∂r˜ϑ ) − 2Γ(∂ψϑ ) . a different approach. One can express Eq. (48) in terms (43) of a simpler set of integrals through multiple integrations- by-parts (noting that the source S` depends on Sg, which As with the scalar field, we can express g(1) in a Legen- in turn is proportional to the squared derivative of the dre decomposition as scalar field) and application of the scalar field evolution equation [Eq. (20)]. Doing so, the modes of the trace of (1) X (1) g = g` (˜r) P`(ψ) . (44) the metric perturbation are given by `

2` + 1 Z r˜ Z 1 (1) + 0 − 0 ˜(1/2) 0 0 g` (˜r) = H` (˜r) dr˜ dψ H` (˜r )P`(ψ)ϑ (˜r , ψ)s(˜r , ψ) W` ∞ −1 Z r˜ Z 1 − 0 + 0 ˜(1/2) 0 0 − H` (˜r) dr˜ dψ H` (˜r )P`(ψ)ϑ (˜r , ψ)s(˜r , ψ) r˜+ −1 Z r˜ Z 1 Z r˜ Z 1 ! + 0 µ 0 − 0 µ 0 + H` (˜r) dr˜ dψ ∂µV− (˜r , ψ) − H` (˜r) dr˜ dψ ∂µV+ (˜r , ψ) , (49) ∞ −1 r˜+ −1

where s(r˜0, ψ) is the scalar field source given in Eq. (21), and we have defined

µ 1 (1/2) p µν ± V := (ϑ˜ )2 −g(0) g ∂ H P  ± 2 (0) ν ` `

1 ± p µν (1/2) − H P −g(0) g ∂ (ϑ˜ )2 . (50) 2 ` ` (0) ν 8

The integrals of total derivatives in Eq. (49) can be sim- For general `, the integration over ψ can be expressed in √ ψψ (0) terms of the standard 3j-symbols. The resulting expres- plified by noting that (i) −g g(0) = Γ, which vanishes when evaluated at the limits of integration ψ = ±1, and sion is (ii) contributions at spatial infinity and at the horizon 2 ` k j vanish due to the behavior of the scalar field modes ϑ˜(1/2), (1, 0) X √ g` (˜r) = 2 × ± (0) rr ˜ 0 0 0 the homogeneous solutions H` , and −g g(0) = ∆. The k,j modes of the trace of the metric perturbation are then " Z r˜ ϑ˜(1/2, 0)(˜r0)s(1/2, 0)(˜r0) ` 0 k j Z r˜ Z 1 × (˜r − 1) dr˜ 0 `+1 (1) 2` + 1 (˜r − 1) g (˜r) = H+(˜r) dr˜0 dψ H−(˜r0)P (ψ)ϑ˜(1/2)s(1/2) ∞ ` W ` ` ` ` ∞ −1 1 Z r˜ ! − dr˜0(˜r0 − 1)` ϑ˜(1/2, 0)(˜r0)s(1/2, 0)(˜r0) Z r˜ Z 1 (˜r − 1)`+1 k j − 0 + 0 ˜(1/2) (1/2) ∞ − H` (˜r) dr˜ dψ H` (˜r )P`(ψ)ϑ s # r˜+ −1 2` + 1 − ϑ˜(1/2, 0)(˜r) ϑ˜(1/2, 0)(˜r) (54) 2` + 1 Z 1   2 k j ˜(1/2) 2 − dψ P`(ψ) ϑ (˜r, ψ) , (51) 2 −1 The radial integrals in Eqs. (53) and (54), though still where in the first and second lines ϑ˜(1/2) and s(1/2) are complicated, are more tractable than the integrals that both functions of r˜0 and ψ. We have simplified the last result from expressing the source function in terms of the line by extracting a factor of W`, defined in Eq. (A4), modes of the scalar field in Eq. (48). which is a constant. We have not yet obtained a closed-form expression Let us now focus on the extremal limit. With the nor- for the trace of the metric perturbation on the extremal malizations defined in AppendixA, the factor W` = 2`+1 background. The main difficulty, apparent in Eqs. (52)- and the homogeneous solutions are given by Eqs. (A8)- (54), is that the source term depends on the full tower of (A9). We can then write Legendre modes of the scalar field. Using our expressions for the modes of the scalar field and its source, Eq. (31) Z r˜ Z 1 P (ψ)ϑ˜(1/2, 0)s(1/2, 0) (1, 0) ` 0 ` and Eq. (25), it is possible to evaluate individual terms in g` (˜r) = (˜r − 1) dr˜ dψ `+1 ∞ −1 (˜r − 1) these sums. However, we have not been able to perform 1 Z r˜ Z 1 the sums themselves. Indeed, the analytic results for − dr˜0 dψ (˜r − 1)`P (ψ)ϑ˜(1/2, 0)s(1/2, 0) (˜r − 1)`+1 ` the individual terms are sufficiently complicated that we 1 −1 turn to approximations and numerical analysis, which we (2` + 1) Z 1 ˜(1/2, 0) 2 discuss in the next section. − dψP`(ψ)(ϑ ) . (52) 2 −1 This completes the formal solution for the modes of the VI. TRACE OF THE METRIC trace of the metric perturbation in the extremal limit in PERTURBATION: PROPERTIES integral form. The angular integrals in Eq. (52) can be evaluated in closed form using the Legendre decomposition of the The results of Sec.IV suggest that the first three or scalar field and the source function. From Eq. (22) and four modes of the scalar field capture most of its physics, Eq. (24) we have and should be sufficient for analyzing the behavior of the trace of the metric perturbation. But first, let us consider ∞ a few important properties of the modes g(1, 0) that can ˜(1/2, 0) X ˜(1/2, 0) ` ϑ (˜r, ψ) = ϑk (˜r)Pk(ψ) be extracted from the integral form of the solution. k=1 At large radius, r˜  1, the second line of Eq. (52) ∞ (1/2, 0) X (1/2, 0) dominates and the leading behavior of the mode is s (˜r, ψ) = sj (˜r)Pj(ψ) , (1, 0) −(`+1) g` ∼ r˜ . This is because the first line of Eq. (52) j=1 decays with a higher power of r˜, while the third line is pro- where the sums are over odd integers in both cases. Using portional to (ϑ˜(1/2, 0))2 and therefore decays as r˜−2(`+1). the orthonormality of Legendre functions, the ` = 0 mode Near the horizon the first and third line of Eq. (52) dom- is given by inate. The asymptotic behavior as r˜ → 1 is most easily extracted by first evaluating Eq. (48) at r˜ = r˜ = 1 + ε, ∞ " Z r˜ ˜(1/2, 0) 0 (1/2, 0) 0 + (1, 0) X 2 ϑ (˜r )s (˜r ) 0 g (˜r) = dr˜0 k k changing the integration variable to η = (r˜ − 1)/ε, and 0 2k + 1 r˜0 − 1 k=1 ∞ then taking the ε → 0 limit, which gives Z r˜ 1 1 0 (1/2, 0) 0 (1/2, 0) 0 Z − dr˜ ϑ˜ (˜r )s (˜r ) (1, 0) (1, 0) k k g (1) = S (1) dη Q`(η) . (55) r˜ − 1 1 ` ` # ∞ 1 − ϑ˜(1/2, 0)(˜r)2 . (53) The overall factor of S(1, 0)(1), the source function evalu- 2 k ` ated at the extremal Kerr horizon, can be expressed in 9

FIG. 4. The (absolute value of) Legendre modes of the trace of the metric perturbation evaluated at the horizon of the FIG. 5. The first four radial modes of the trace of the metric extremal background, on a logarithmic scale, as a function of perturbation as a function of r˜. The dashed vertical line harmonic number `. indicates the horizon of the extremal background. The ` = 0 mode exhibits a logarithmic divergence as r˜ → 1. terms of the source functions for the scalar field. Evalu- ating Eq. (46) at r˜ = 1, using lim ∆˜ (∂ ϑ˜)2 = 0, and r˜→1 r˜ This precisely matches the log behavior of the solution performing the angular integral yields near r˜ = 1, which can be extracted from both analytical (1, 0) ∞   and numerical results for g0 (˜r). (1, 0) X 2(2` + 1) ` j k S (1) = − As explained at the end of the previous section, analyt- ` p 0 0 0 (1, 0) k,j=1 j(j + 1)k(k + 1) ical results for the r˜-dependence of the modes g` are sufficiently complicated that a numerical analysis is called ` j k  × s(1/2, 0)(1)s(1/2, 0)(1) , (56) for. The first four modes of the trace of the metric pertur- 0 1 −1 j k bation are shown in Fig.5. These are numerical solutions, where again the result is expressed in terms of 3j-symbols. obtained with closed form expressions for the source that For ` ≥ 2 (recall that ` is even) the integral in Eq. (55) include contributions from modes of the scalar field with converges to ` ≤ 21. It is immediately apparent that the ` = 0 mode dominates the trace of the metric perturbation, even away 1 from the log-divergent behavior near r˜ = 1. g(1, 0)(1) = − S(1, 0)(1) . (57) ` `(` + 1) ` We expect, based on the behavior shown in Fig.5 that g(1, 0) is well-approximated by its first few Legendre modes. In this case the mode is finite at the horizon of the ex- In the case of the scalar field, a similar conclusion was tremal background, just like the modes of the scalar field. justified by examining mode-by-mode contributions to (1, 0) The values g` (1) are plotted against ` in Fig.4, where the ADM energy. There is not an obvious analog for the observe that rather than falling off monotonically with `, trace of the metric perturbation, so instead we consider the ` = 4 mode is suppressed relative to the ` = 6 and the fractional difference, as a function of r˜, between g(1, 0) ` = 8 modes. and its approximation by the first N modes For the ` = 0 mode the integral in Eq. (55) does not converge, and the mode has a logarithmic divergence as N ! 1 X (1, 0) r˜ → 1: δN (˜r) = 1 − g (˜r) . (60) g(1, 0)(˜r) ` `=0 (1, 0) (1, 0) lim g0 (˜r) ∼ S0 (1) log(˜r − 1) . (58) r˜→1 The fractional difference for N = 0, 2, 4, 6 is shown in The meaning of this divergence will be discussed in Fig.6. As expected, the log-divergence of the ` = 0 Sec.VII; suffice it to say here that this logarithmic diver- modes means that the fractional difference δ0(r˜) → 0 as gence does not imply the existence of a naked singularity r˜ → 1. Figure6 shows that if one wishes an accuracy at the perturbed horizon. For ` = 0, Eq. (56) reduces to of no more than about 10%, then retaining only the a single sum which can be evaluated numerically ` = 0 mode suffices. To obtain a higher accuracy, more modes are needed. In particular, since the ` = 4 mode is ∞ suppressed at r˜ = 1 relative to the ` = 6 and 8 modes, (1, 0) X 2 (1/2, 0) S (1) = − (s (1))2 (59) 0 k(k + 1)(2k + 1) k one must include modes up to ` = 6 to obtain uniform k=1 accuracy of at least one percent. Observe, however, that ' − 3.52572 . if one is interested in the trace of the metric perturbation 10

2.0

1.5

FIG. 6. The fractional difference between the trace of the metric perturbation at θ = 0, and its approximation including 1.0 only modes with ` ≤ N.

0.5 outside a larger radius, such as for r˜ & 2, then g(1, 0) is approximated at better than percent precision with only the first two or three modes. (1, 0) We conclude that Legendre modes g` with ` ≤ 4 capture almost all of the physics of g(1, 0), except near r˜ = 1 where the ` = 6 and ` = 8 modes may be required. Note that the fractional error defined in Eq. (60) puts a bound on the fidelity of our approximation of g(1, 0) at ψ = 1 (θ = 0), where P`(ψ) = 1 for all `. However, since the ` = 0 mode dominates, and −1 ≤ P`(ψ) ≤ 1 for ` ≥ 2, the fractional error δN (r˜) gives an upper bound on the fidelity of the approximation in the full (r,˜ ψ) plane. An FIG. 7. The trace of the metric perturbation on the extremal approximation of g(1, 0) by its first four Legendre modes background, approximated by its first four Legendre modes. is shown in Fig.7. Near the extremal horizon r˜ = 1 (the solid line) the logarithmic The analysis above uses modes of the scalar field with divergence of the monopole term dominates. ` ≤ 21 to approximate the source for the trace of the met- ric perturbation. However, as we saw in Sec.IV, the first three modes of the scalar field account for most of its con- in the slow-rotation limit. We then solved for the Legendre tribution to the ADM energy. Indeed, the behavior of the decomposition of the trace of the metric perturbation. first few modes of g(1, 0) is largely unchanged if we include We discovered that retaining 3 (4) terms in the Legendre fewer modes of the scalar field. In particular, we achieve expansion suffices to ensure a fidelity of at least 99% in (1, 0) comparable results for the mode g` by approximating the scalar field (trace of the metric perturbation) relative the scalar field by its first N = ` + 1 modes. to numerical solutions. The trace of the metric perturbation in harmonic gauge exhibits a logarithmic divergence, but this is probably VII. DISCUSSION not a problem. The divergence occurs at the location of the extremal Kerr horizon, which need not coincide with This paper explored rotating black holes in dCS. Using the location of the dCS corrected horizon. Indeed, in an effective field theory treatment of dCS, we worked the slow-rotation expansion, the horizon was seen to be 2 in the decoupling limit where dCS corrections are small shifted outward to rhor = rhor,Kerr + (915/28672)ζMχ + perturbations from GR solutions. We have further focused O(χ4) [14]. Thus, one may expect that in the extremal on BHs that spin at the maximal Kerr rate, the so-called limit the Kerr and the dCS horizons do not coincide either. extremal limit. With these assumptions in hand, we then It may also be the case that the extremality condition in solved for the dynamical scalar field in closed analytic dCS is shifted away from J = M. form, through a Legendre decomposition that we found The techniques found above rely heavily on Legendre was dominated by the dipole term. The radial structure expansions, but our work suggests that these can be of this decomposition includes natural logarithms and truncated at a finite mode number without losing much arctangents, unlike the simple polynomial results obtained of the overall behavior of the function. In particular, if 11 one wishes to carry out astrophysical tests of GR, certain Appendix A: Solution of the Scalar Equation of observables may be sensitive to only certain regions of Motion spacetime that need not include the horizon. For example, BH shadow observations are most sensitive to the location The equations of motion for the scalar field and the of the light-ring, while astrophysical observations of the trace of the metric perturbation have the same general energy spectrum of radiation emitted by accretion disks form, so let us briefly establish some conventions for the are most sensitive to the location of the innermost stable solutions of such equations. First, consider an equation circular orbit. For such observations, it may suffice to of the form keep only the first few modes in a Legendre expansion provided the BH is not rotating maximally. The reason ∂r˜(∆˜ ∂r˜I`) − `(` + 1) I` = K` (A1) here is two-fold. First, the light-ring and ISCO are both pushed away from the horizon as the spin decreases, and + − with some source K`. Denote by H` and H` the solutions the approximation by a finite number of Legendre modes of the homogeneous equation improves away from the immediate vicinity of the horizon. ˜ ± ± Second, our studies indicate that, in general, the fidelity ∂r˜(∆∂r˜H` ) − `(` + 1) H` = 0 , (A2) of an approximation at a fixed number of modes improves + − away from extremality, as shown in [16]. with H` regular at r˜+ and H` → 0 at r˜ → ∞. We use the method of variation of parameters to find the general The results obtained here open the door for new inves- solution to the inhomogeneous equation. The solution tigations of rotating BHs in dCS gravity. For example, of (A1) that is both regular at r˜ and goes to zero as the methods we employed could be extended to the next- + r˜ → ∞ can be expressed in terms of the homogeneous order term in a near-extremal expansion, or better yet, solutions and the source as for BHs that rotate with arbitrary spins. We have already obtained partial results for the latter, with closed-form Z r˜ results for the first two modes of the scalar field (and nu- 1 + 0 − 0 0 I` = × H` (˜r) dr˜ H` (˜r )K`(˜r ) merical results for all other modes). Ultimately, of course, W` ∞ one would like to solve for the full metric perturbation of Z r˜ ! dCS BHs, and not just the trace. − 0 + 0 0 − H` (˜r) dr˜ H` (˜r )K`(˜r ) . (A3) r˜+

Here a factor of ∆˜ has canceled inside each integral, allow- ing us to pull out a constant W`; this constant depends on the Wronskian of the homogeneous solutions, ACKNOWLEDGMENTS ˜ − + W` ≡ ∆ × W [H` ,H` ] (A4) = ∆˜ × H−∂ H+ − H+∂ H− . We would like to thank the Kavli Institute for Theo- ` r˜ ` ` r˜ ` retical Physics for their hospitality during the completion It is straightforward to verify that this is constant using of this work. We would also like to thank Kent Yagi, Eq. (A2). Frans Pretorius, and Albion Lawrence for useful discus- For the Kerr background, the homogeneous solutions sions. RM acknowledges support from a Loyola University can be written as Chicago Summer Research Stipend. LCS acknowledges ! that support for this work was provided by NASA through ` + + 2 r˜ − 1 Einstein Postdoctoral Fellowship Award Number PF2- H (˜r) = c (1 − χ ) 2 P` (A5) ` ` p 2 130101 issued by the Chandra X-ray Observatory Center, 1 − χ ! which is operated by the Smithsonian Astrophysical Ob- `+1 − − 2 − r˜ − 1 H (˜r) = c (1 − χ ) 2 Q` . (A6) servatory for and on behalf of the NASA under contract ` ` p 2 NAS8-03060, and further acknowledges support from the 1 − χ NSF grant PHY-1404569. NY acknowledges support from NSF CAREER Award PHY-1250636. The research of RM where P`(·) and Q`(·) are Legendre functions of the first and NY was supported in part by the National Science and second kind, respectively. A standard identity for Legendre functions then gives the factor W = c+ c−. Foundation under Grant No. NSF PHY11-25915. Some ` ` ` p 2 calculations used the computer algebra-system Maple, in The factors of 1 − χ in Eq. (A5)-(A6) have been combination with the GRTensorII package [33]. Other chosen so that the extremal limit, χ → ±1, is regular. ± calculations used the computer algebra-system Mathe- Otherwise, the overall normalization factors c` are arbi- matica, in combination with the xTensor package [34– trary. A convenient choice is to set 36]. Finally, RM (@mcnees) and LCS (@duetosymmetry) would like to thank Twitter for facilitating discussion + `! − (2` + 1)!! c` = , c` = . (A7) during the early stages of this collaboration. (2` − 1)!! `! 12

Then W` = 2` + 1, and in the extremal limit the homoge- We adopt this normalization throughout Secs.III andV. neous solutions are simply

Appendix B: Expressions for Radial Modes

+ ` lim H` = (˜r − 1) (A8) The radial mode function for general ` is given in |χ|→1 Eq. (31). In this form, each mode depends on 2` + 1 − 1 coefficients α and β . The coefficients for the modes lim H` = . (A9) `,k `,k |χ|→1 (˜r − 1)`+1 up to ` = 9 are given below.

k α1,k β1,k α3,k β3,k α5,k β5,k α7,k β7,k α9,k β9,k 70 15 68 85 31719 189 49019 1377 0 0 -3 3 2 − 5 − 8 70 16 1260 − 128 1015 2955 22317 4347 351417 315405 1 - 3 -25 −75 8 8 − 80 − 4 896 128 8645 735 90819 8127 13832005 34155 2 -- 15 75 − 8 − 2 20 8 − 896 − 16 4263 4935 44079 33705 3351161 1065405 3 -- - 25 8 4 − 16 − 4 384 32 819 4935 21009 73395 9069467 1183545 4 -- -- − 8 − 8 2 16 − 128 − 64 987 14067 44793 3456585 6416487 5 -- -- - 8 − 4 − 4 128 64 1989 14931 30089455 292215 6 ------4 4 − 384 − 8 2133 2523125 2579445 7 ------− 4 128 32 279565 2579445 8 ------− 128 − 128 286605 9 ------128

Appendix C: Representations of the Scalar Field of Eq. (28). From here on, different representations take different routes to arrive at a solution to Eq. (23) in the Equation (31) provides one representation of the solu- extremal limit, so we tackle each of them separately below. tion to Eq. (23) for arbitrary harmonic number ` after a Legendre decomposition and in an expansion to leading order in ζ (i.e. in the GR deformation) and in ε (i.e., in 1. Incomplete Beta Function Representation the extremal limit). This form of the solution depends on 2` + 1 coefficients (α`,k and β`,k) that are fixed by Introducing expansion Eq. (C1) for the scalar source imposing appropriate boundary conditions on the mode. into the solution Eq. (28) for the scalar field: In this appendix, we present two additional representa- ∞ " Z r˜ 0 (1/2, 0) X dr˜ tions of the solution to the scalar field evolution equation ϑ˜ (˜r) = α (˜r − 1)` that may be preferable in some applications. As in the ` `,n (˜r0 − 1)`+1 r˜0`+4+2n n=0 ∞ case of Eq. (31), these representations will have both ad- (C3) vantages and disadvantages that we will describe in detail. Z r˜ # The solutions start by representing the source function in 1 0 0 ` 1 − `+1 dr˜ (˜r − 1) 0`+4+2n , the extremal limit in terms of a series: (˜r − 1) 1 r˜ ∞ (1/2, 0) X 2` + 1 s (˜r) = α , (C1) where we have already imposed appropriate boundary ` `,n r˜`+4+2n n=0 conditions. The integrals can be evaluated in closed-form where we have introduced the constants to obtain 1 ˜(1/2, 0) `−1 n (` + 2n + 2) Γ(` + 4 + 2n) Γ( 2 ) ϑ` (˜r) = β1(˜r) + β2(˜r) + β3(˜r) (C4) α = (−1) 2 (−1) , `,n 2`+1+2n Γ(n + 1) Γ(` + n + 3 ) 2 where we have defined (C2) ∞   X ` in terms of the Gamma function Γ(·). The factor of 2` + 1 β1(˜r) = − α`,n(˜r − 1) B1/r˜ 2` + 4 + 2n, −` , in Eq. (C1) has been introduced to simplify some expres- n=0 sions, by canceling a similar factor in the denominator (C5) 13

∞   X α`,n in terms of the incomplete Beta function Bx(a, b) (see β (˜r) = B 2n + 3, ` + 1 , (C6) 2 (˜r − 1)`+1 1/r˜ e.g. Sec. 8.17 of [37]), n=0 ∞ Z x X α`,n Γ(` + 1) Γ(2n + 3) β (˜r) = − , (C7) B (a, b) ≡ ta−1(1 − t)b−1dt . (C8) 3 (˜r − 1)`+1 Γ(` + 4 + 2n) x n=0 0 Evaluating at x = 1 gives the ordinary Beta function, B(a, b) = B1(a, b).

The sums over n can be evaluated in closed form for β2 and β3. The latter can be summed into `+1 1 " # (−1) 2 ` Γ(` + 1)Γ( ) 1 β (˜r) = 2 ` (2` + 1) − 2(` − 1)(` + 1) F − 1 , 1; ` + 3 ; −1 , (C9) 3 `+2 3 `+1 2 1 2 2 2 Γ(` + 2 ) (˜r − 1) ˜(1,0) which gives the leading behavior of ϑ` at large r˜. The sum over n for β2 can be evaluated using the series representation of the incomplete Beta function appropriate for Eq. (C6), n−1 X (−1)j Γ(n) B m, n = xm+j . (C10) x m + j Γ(j + 1) Γ(n − j) j=0 Permuting the order of the sums over j and n yields ` (−1)(`+1)/2 Γ(` + 1)Γ(1/2)Γ(4 + `) X 1 1 1 β (˜r) = − (−1)j 2 (˜r − 1)`+1 22+`Γ(5/2 + `) (3 + j)(5 + j) Γ(1 + j)Γ(1 + ` − j) r˜5+j j=0  3 + j 4 + ` 5 + ` 5 + j 3 + 2` 1  × (5 + j)(2 + `)(3 + 2`)r ˜2 F , , ; , ; − 3 2 2 2 2 2 2 r˜2 5 + j 6 + ` 7 + ` 7 + j 5 + 2` 1  −(3 + j)(4 + `)(5 + `) F , , ; , ; − , (C11) 3 2 2 2 2 2 2 r˜2 where P FQ(·; ·; ·) is the generalized hypergeometric function. We have not succeeded in finding a closed-form expression for the above sum over j, but the sum can be performed explicitly given any value of `. One is then only left with the sum over n for β1. To obtain an expression for this sum, we start with the following representation of the incomplete Beta function relevant for Eq. (C5):

" `+3+2n !# (−1)`+1(` + 4 + 2n)Γ(2` + 4 + 2n) r˜ − 1 1 X 1 1 B (2` + 4 + 2n, −`) = ln + 1 − 1/r˜ Γ(` + 1)Γ(` + 5 + 2n) r˜ r˜ − 1 k(k + 1) r˜k k=1 `−2 Γ(2` + 4 + 2n) 1 X (−1)k+1Γ(` − k) − (˜r − 1)k−` . (C12) Γ(` + 1) r˜`+3+2n Γ(2` + 4 + 2n − k) k=0

This allows us to write β1(˜r) := β4(˜r) + β5(˜r), where we have defined ∞ " `+3+2n !# X (−1)`+1(` + 4 + 2n)Γ(2` + 4 + 2n) r˜ − 1 1 X 1 1 β (˜r) = −(˜r − 1)` α ln + 1 − 4 `,n Γ(` + 1)Γ(` + 5 + 2n) r˜ r˜ − 1 k(k + 1) r˜k n=0 k=1 (C13) ∞ `−2 1 1 X 1 X (−1)k Γ(` − k) β (˜r) = − α Γ(2` + 4 + 2n) (˜r − 1)k . (C14) 5 Γ(` + 1) r˜`+3 `,n r˜2n Γ(2` + 4 + 2n − k) n=0 k=0

The function β5(˜r) can be simplified further by performing the sums to obtain `−2 1 X Γ(` − k) β (˜r) = (−1)(`+1)/221+`(1 + `)(2 + `)Γ(4 + `) (−1)k (˜r − 1)k 5 r˜5+` Γ(6 − k + 2`) k=0  4 + ` 5 + ` 5 + 2` 3 + 2` 4 − k + 2` 5 − k + 2` 1  × (k − 2` − 5) (3 + 2`)(k − 4 − 2`)r ˜2 F , , 2 + `, ; , , ; − 4 3 2 2 2 2 2 2 r˜2 6 + ` 7 + ` 7 + 2` 5 + 2` 6 − k + 2` 7 − k + 2` 1  −2 (4 + `) (5 + `) (5 + 2`) F , , 3 + `, ; , , ; − . (C15) 4 3 2 2 2 2 2 2 r˜2 14

The function β4(˜r) can also be simplified by performing some of the sums in closed form to obtain ∞ " `+3+2n !# (−1)` X (` + 4 + 2n)Γ(2` + 4 + 2n) r˜ − 1 1 X 1 1 β (˜r) = (˜r − 1)` α ln + 1 − 4 Γ(` + 1) `,n Γ(` + 5 + 2n) r˜ r˜ − 1 k(k + 1) r˜k n=0 k=1 ∞ `+3+2n (−1)(3`+1)/2 Γ(` + 3)  r˜ − 1 (−1)`+1 X X 1 1 = (˜r − 1)`−1 1 + (˜r − 1) ln + (˜r − 1)`−1 γ , 2 Γ(` + 1) r˜ Γ(` + 1) `,n k(k + 1) r˜k n=0 k=1 (C16) where we have defined (` + 4 + 2n)Γ(2` + 4 + 2n) γ := α . (C17) `,n `,n Γ(` + 5 + 2n) The last term in Eq. (C16) can also be represented as follows:

∞ `+3+2n ∞  ∞  X X 1 1 (−1)(`+1)/2 1  X X  1 1 γ = Γ(` + 3) G , ` + 3 γ + `,n k(k + 1) r˜k 2 r˜  `,n (` + 4 + 2j)(` + 5 + 2j) r˜`+4+2j n=0 k=1 j=0 n=j+1 1 1  + , (C18) (` + 5 + 2j)(` + 6 + 2j) r˜`+5+2j

where we have defined the new function with

∞ N k X 1 X x σ1(˜r) := a`,n , (C21) G(x, N) := , (C19) r˜`+1+n k(k + 1) n=0 k=1 ∞ X 1 σ (˜r) := b , (C22) 2 `,n r˜`+4+2n for some x ∈ < and N ∈ N. This function is the first N n=0 terms of the Taylor series for 1−log(1−x)+x−1 log(1−x) about x = 0. Notice that the sum in this new function is into Eq. (23) and find recursion relations for the a`,n and finite, and thus G(1/r,˜ N) is simply a polynomial in 1/r˜. b`,n coefficients. Given a particular value of `, the remaining sum over j The recursion relations for the a`,n can be solved to can be performed explicitly. obtain (` + n)! a = a , (C23) `,n `! n! `,0 2. Radial Series Representation which then leads to a`,0 σ1(˜r) = . (C24) Instead of using variation of parameters to solve (˜r − 1)`+1 Eq. (23), we will search for a series solution. We thus insert the ansatz Since this is the leading behavior of the scalar field at large r˜, we can determine the coefficient a`,0 by comparing it with the incomplete Beta function representation of the ˜(1,0) ϑ` (˜r) = σ1(˜r) + σ2(˜r) , (C20) previous subsection:

∞ `+1 √ X (−1) 2 π `!  (` + 2) 6  a = − α B(2n + 3, ` + 1) = F 3 , 2; ` + 3 ; −1 − F 5 , 3; ` + 5 ; −1 . `,0 `,n 2` Γ(` + 3 ) 2 1 2 2 Γ(` + 5 ) 2 1 2 2 n=0 2 2 (C25)

Resumming the coefficients b`,n is more complicated. We can solve the recursion relations to express the b`,n coefficient as finite sums that depend on the coefficients α`,n in the series expansion of the source:

jmax jmax Γ(` + 4 + n) X Γ(3 + 2j) Γ(` + 4 + n) X Γ(2` + 4 + 2j) b = α − α , (C26) `,n Γ(4 + n) Γ(` + 4 + 2j) `,j Γ(2` + 5 + n) Γ(` + 4 + 2j) `,j j=0 j=0 15 where jmax = n/2 if n is even, and jmax = (n + 1)/2 if n is odd. When one tries to perform the full infinite sum of the b`,n coefficients over n to find σ2(r˜), one finds a familiar problem: the coefficients of Eq. (C26) are finite sums with an upper limit that depends on n, which must then be summed to infinity. To get around this problem, we can rewrite each finite sum as the difference of two infinite sums:

Γ(` + 4 + n) Γ(` + 4 + n) b = c − d , (C27) `,n Γ(4 + n) `,jmax Γ(2` + 5 + n) `,jmax where √ " `−1 π (` + 1)(4` − 7) 2 (−1)k(2k + 3)Γ(k + 5 ) (`4 − 2`2 + 9` + 10) 2 2 1 5 c`,k = (−1) `+1 × 3 + √ 3 + 5 2F1(− 2 , 1; ` + 2 ; −1) 2 2 Γ(` + 2 ) π Γ(` + 2 + k) 2 Γ(` + 2 ) 4 3 k 5 (` + 5` + ` + 5) 1 5 2 (−1) (` − 1) Γ(k + 2 ) 5 5 − 5 2F1( 2 , 1; ` + 2 ; −1) + √ 5 2F1(1, k + 2 ; ` + k + 2 ; −1) 2 Γ(` + 2 ) π Γ(` + 2 + k) `−1  2 (` + 2) 1 3 + 3 2F1(` − 2 , `; ` + 2 ; −1) , (C28) Γ(` + 2 )

`−1  1 (−1)k2`+2(k + 1)(` + k + 2)Γ(` + k + 3) d = (−1) 2 × − Γ(` + 3) + (C29) `,k 2 Γ(k + 2) (−1)k2`+1(` + 2)Γ(` + k + 3)  + F (1, ` + k + 3; k + 2; −1) . Γ(k + 2) 2 1

With the b`,n coefficients expressed in this form, the second sum for the scalar field becomes

∞ X Γ(` + 4 + n) Γ(` + 4 + n)  1 σ (˜r) = c − d . (C30) 2 Γ(4 + n) `,jmax Γ(2` + 5 + n) `,jmax r˜`+4+2n n=0 We have not succeeded in finding closed-form expressions for the sum over n given a generic ` value, but the sum can be performed for a given value of `.

Appendix D: Series Solutions for the Trace of the Note that the coefficient of the log(r˜ − 1) term is (D1) Metric Perturbation evaluated at r˜ = 1, as in Eq. (58). The modes with ` ≥ 2 can be expressed as a series involving incomplete Beta Instead of truncating the Legendre expansion of the functions: scalar field, it is also possible to construct series approxi- mations of the modes g(1, 0). We first note that the source ∞ ` (1, 0) X h ` term (46) can be expanded in powers of 1/r˜. For the g` (˜r) = e`,n × − (˜r − 1) B1/r˜ (2` + 2 + n, −`) ` = 0 mode this series takes the form n=0 ∞ 1 i (1, 0) X 1 − B1−1/r˜ (` + 1, n + 1) . (D4) S (˜r) = e , (D1) (˜r − 1)`+1 0 0,n r˜4+n n=0 while for ` ≥ 2 it is Since these solutions are obtained directly from Eq. (48) ∞ they already satisfy the correct boundary conditions at (1, 0) X 1 S (˜r) = e . (D2) r˜ → ∞ and r˜ = 1. ` `,n r˜`+2+n n=0 Given the expansion of the source functions, one can obtain a series solution of the equation of motion Eq. (45) In terms of the series coefficients for the source, the ` = 0 directly. For the ` = 0 mode this solution takes the form mode is

∞ " n+3 r˜ − 1 1 ∞ n (1, 0) X X (1, 0) f0,0 X 1 X n + 1 − j g0 (˜r) = e0,n × log + g (˜r) = + e r˜ j r˜j 0 (˜r − 1) r˜4+n (n + 4)(j + 3) 0,j n=0 j=1 n=0 j=0 # 1 1  1  (D5) + − 1 . (D3) n + 3 r˜ − 1 r˜n+3 while for ` ≥ 2 it is 16

∞ n   (1, 0) f`,0 X (` + n)! X j! (2` + j + 1)! g (˜r) = + e − . (D6) ` (˜r − 1)`+1 r˜`+2+n `,j (n + 1)!(` + j + 1)! (` + j + 1)!(2` + n + 2)! n=0 j=0

In both cases the coefficient f`,0 of the leading term can be expressed in terms of one or more integrals of the source; for ` = 0 it is Z 1 (1, 0) f0,0 = dr˜ S0 (˜r) . (D7) ∞

[1] C. M. Will, Living Reviews in Relativity 17 (2014), [21] A. M. Ghezelbash, JHEP 08, 045 (2009), arXiv:0901.1670 10.1007/lrr-2014-4. [hep-th]. [2] N. Yunes and X. Siemens, Living Rev.Rel. 16, 9 (2013), [22] Y. Matsuo and T. Nishioka, JHEP 12, 073 (2010), arXiv:1304.3473 [gr-qc]. arXiv:1010.4549 [hep-th]. [3] S. Alexander and N. Yunes, Phys.Rept. 480, 1 (2009), [23] S. Carlip, JHEP 04, 076 (2011), [Erratum: arXiv:0907.2562 [hep-th]. JHEP01,008(2012)], arXiv:1101.5136 [gr-qc]. [4] T. Delsate, D. Hilditch, and H. Witek, Phys. Rev. D91, [24] G. Compere, Living Rev. Rel. 15, 11 (2012), 024027 (2015), arXiv:1407.6727 [gr-qc]. arXiv:1203.3561 [hep-th]. [5] M. B. Green and J. H. Schwarz, Phys. Lett. B149, 117 [25] O. J. C. Dias, J. E. Santos, and M. Stein, JHEP 10, 182 (1984). (2012), arXiv:1208.3322 [hep-th]. [6] M. B. Green, J. H. Schwarz, and E. Witten, Superstring [26] C. W. Misner, K. S. Thorne, and J. A. Wheeler, Gravi- Theory. Vol. 2: Loop Amplitudes, Anomalies And Phe- tation (W. H. Freeman, San Francisco, 1973). nomenology (Cambridge: Cambridge University Press, [27] R. M. Wald, General Relativity (University of Chicago 1988). Press, 1984). [7] S. Weinberg, Phys.Rev. D77, 123541 (2008), [28] K. Yagi, L. C. Stein, and N. Yunes, (2015), arXiv:0804.4291 [hep-th]. arXiv:1510.02152 [gr-qc]. [8] V. Taveras and N. Yunes, Phys.Rev. D78, 064070 (2008), [29] S. Mercuri and V. Taveras, Phys.Rev. D80, 104007 (2009), arXiv:0807.2652 [gr-qc]. arXiv:0903.4407 [gr-qc]. [9] R. Jackiw and S. Pi, Phys.Rev. D68, 104012 (2003), [30] Y. Ali-Haimoud and Y. Chen, Phys.Rev. D84, 124033 arXiv:gr-qc/0308071 [gr-qc]. (2011), arXiv:1110.5329 [astro-ph.HE]. [10] D. Ayzenberg, K. Yagi, and N. Yunes, Phys. Rev. D89, [31] R. P. Kerr, Phys.Rev.Lett. 11, 237 (1963). 044023 (2014), arXiv:1310.6392 [gr-qc]. [32] L. C. Stein, (2014), arXiv:1407.0744 [gr-qc]. [11] B. A. Campbell, M. J. Duncan, N. Kaloper, and K. A. [33] “GRTensorII,” This is a package which runs within Maple Olive, Phys. Lett. B251, 34 (1990). but distinct from packages distributed with Maple. It [12] N. Yunes and F. Pretorius, Phys.Rev. D79, 084043 (2009), is distributed freely on the World-Wide-Web from the arXiv:0902.4669 [gr-qc]. address: http://grtensor.org. [13] K. Konno, T. Matsuyama, and S. Tanda, [34] J. M. Martin-Garcia, R. Portugal, and L. R. U. Manssur, Prog.Theor.Phys. 122, 561 (2009), arXiv:0902.4767 Comput.Phys.Commun. 177, 640 (2007), arXiv:0704.1756 [gr-qc]. [cs.SC]. [14] K. Yagi, N. Yunes, and T. Tanaka, Phys.Rev. D86, [35] J. M. Martin-Garcia, D. Yllanes, and R. Portugal, Com- 044037 (2012), arXiv:1206.6130 [gr-qc]. put.Phys.Commun. 179, 586 (2008), arXiv:0802.1274 [15] K. Konno and R. Takahashi, Phys. Rev. D90, 064011 [cs.SC]. (2014), arXiv:1406.0957 [gr-qc]. [36] D. Brizuela, J. M. Martin-Garcia, and G. A. Mena Maru- [16] L. C. Stein, Phys. Rev. D90, 044061 (2014), gan, Gen.Rel.Grav. 41, 2415 (2009), arXiv:0807.0824 [gr- arXiv:1407.2350 [gr-qc]. qc]. [17] D. Grumiller and N. Yunes, Phys. Rev. D77, 044015 [37] DLMF, “NIST Digital Library of Mathematical Functions,” (2008), arXiv:0711.1868 [gr-qc]. http://dlmf.nist.gov/, Release 1.0.10 of 2015-08-07, online [18] N. Yunes and C. F. Sopuerta, Phys. Rev. D77, 064007 companion to [38]. (2008), arXiv:0712.1028 [gr-qc]. [38] F. W. J. Olver, D. W. Lozier, R. F. Boisvert, and C. W. [19] C. F. Sopuerta and N. Yunes, Phys. Rev. D80, 064006 Clark, eds., NIST Handbook of Mathematical Functions (2009), arXiv:0904.4501 [gr-qc]. (Cambridge University Press, New York, NY, 2010) print [20] M. Guica, T. Hartman, W. Song, and A. Strominger, companion to [37]. Phys. Rev. D80, 124008 (2009), arXiv:0809.4266 [hep-th].