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

Horizon Surface Gravity in Corotating Black Hole Binaries

Alexandre Le Tiec LUTH, Observatoire de Paris, PSL Research University, CNRS, Université Paris Diderot, Sorbonne Paris Cité, 92190 Meudon, France    Philippe Grandclément LUTH, Observatoire de Paris, PSL Research University, CNRS, Université Paris Diderot, Sorbonne Paris Cité, 92190 Meudon, France
(August 18, 2025)
Abstract

For binary systems of corotating black holes, the zeroth law of black hole mechanics states that the surface gravity is constant over each component of the horizon. Using the approximation of a conformally flat spatial metric, we compute sequences of quasi-equilibrium initial data for corotating black hole binaries with irreducible mass ratios in the range 10:11:110:1-1:1. For orbits outside the innermost stable one, the surface gravity is found to be constant on each component of the apparent horizon at the sub-percent level. We compare those numerical results to the analytical predictions from post-Newtonian theory at the fourth (4PN) order and from black hole perturbation theory to linear order in the mass ratio. We find a remarkably good agreement for all mass ratios considered, even in the strong-field regime. In particular, our findings confirm that the domain of validity of black hole perturbative calculations appears to extend well beyond the extreme mass-ratio limit.

pacs:
04.25.dg, 04.25.Nx, 04.30.Db, 97.60.Lf

I Introduction and summary

The first detections of gravitational waves from coalescing black hole binaries has herald a new era in astronomy Ab.al2.16; Ab.al3.16; Ab.al.17; Ab.al2.17. By the end of this decade, an increasingly sensitive network of terrestrial laser interferometers (LIGO, Virgo, KAGRA) will detect up to several hundreds of such events per year Ab.al.18. Those gravitational-wave observations will not only tell us about the underlying population of binary black holes, but also allow testing the general theory of relativity (GR), and improve our understanding of the dynamics of compact binary systems in a highly dynamical, strong-field regime Ab.al4.16.

In this paper, we focus on corotating black hole binaries, which correspond to the particular case where the black holes are rotating synchronously with the orbital motion. Heuristically, these can be viewed as stationary in a corotating frame. A great deal of attention has already been devoted to such systems Go.al.02; Gr.al.02; Fr.al.02; Co2.02; Bl.02; Da.al.02; CoPf.04; Kl.04; Ca.al2.06, because they represent the only configuration of two black holes that can possess a true helical Killing field, and for which the rotation state of each black hole is fully constrained. Interestingly, the exact helical symmetry allows defining the horizon surface gravity of each black hole in the standard way Poi. Moreover, according to the zeroth law of binary black hole mechanics Fr.al.02, the surface gravity must be constant over each component of the horizon.

While sequences of quasi-equilibrium initial data for corotating black hole binaries have been constructed long ago, the main interest at the time was to extract the binding energy of the system, as a function of the circular-orbit frequency, in order to locate the innermost circular orbit. Here we revisit this problem, but focus instead on the horizon surface gravity of each black hole, which, rather surprisingly, has never been investigated in this context. Using the standard approximation of a conformally flat spatial metric, we compute sequences of quasi-equilibrium initial data for corotating black hole binaries with irreducible mass ratios 10:110:1, 2:12:1 and 1:11:1, and extract the surface gravity of each black hole.

On the other hand, if the orbital separation is large, then such systems can be investigated in the context of the post-Newtonian (PN) approximation to GR. Each black hole can then be modelled as a spinning point particle. Various arguments indicate that the point-particle analogue of the horizon surface gravity of a black hole is the “redshift variable,” a conserved quantity associated with the helical symmetry Le.al.12; Bl.al.13; Po3.15; Zi.al.16. Building upon the recent derivation of the 4PN equations of motion for binary systems of nonspinning point particles Da.al.14; Be.al.16; Ma.al.18, we add the relevant spin-orbit contributions and compute the redshifts in corotating point-particle binaries up to 4PN order,111As usual we refer to nnPN as the order corresponding to terms 𝒪(c2n)\mathcal{O}(c^{-2n}) in the equations of motion. for comparison to the numerical results.

Corotating black hole binaries have also been studied in the limit of extreme mass ratios, in the context of linear black hole perturbation theory. The authors of Ref. GrLe.13 showed that, for a Kerr black hole perturbed by a small corotating moon, the perturbed event horizon is a Killing horizon of the helical Killing field. The surface gravity is constant over the horizon, and the perturbation in surface gravity was computed analytically, together with the perturbation in the horizon angular velocity GrLe.13. Combining those results, we compute in closed form the surface gravity of the larger body, to linear order in the mass ratio. We also compute the redshift of the corotating point particle, i.e., the analogue of the horizon surface gravity of the smaller body, still to linear order in the mass ratio.

A growing body of work FiDe.84; An.al.95; Fa.al.04; Sp.al2.11; Le.al.11; Le.al2.12; Le.al.13; Na.13; vdM.17 suggests that the domain of validity of black hole perturbative calculations can be extended well beyond the extreme mass-ratio limit, by means of a simple but physically-motived rescaling of the component masses. Here, we perform such a rescaling and show that the rescaled perturbative predictions are in remarkable agreement with the numerical relativity (NR) results obtained by solving (part of) the Einstein field equations. As pointed out in Ref. Le2.14, this opens the exciting prospect of using black hole perturbation theory to model the gravitational-wave emission from intermediate mass-ratio inspirals or even compact binaries with comparable masses.

The remainder of this paper is organized as follows. In Sec. II we give a summary of the known results on the laws of mechanics for single and binary black hole systems. Section III details our computations of quasi-equilibrium initial data for corotating black hole binaries with mass ratios in the range 10:11:110:1-1:1, from which we extract the surface gravity. We then compute the (analogue of the) surface gravity for corotating point particles in Sec. IV, in the context of the PN approximation, and for a Kerr black hole perturbed by a corotating moon in Sec. V, using linear perturbation theory. Finally, we compare those numerical and analytical results in Sec. VI.

Throughout this paper our conventions are those of Wal. The metric signature is (+++)(-+++) and we use “geometrized units” where G=c=1G=c=1. Greek indices α,β,\alpha,\beta,\dots are abstract, while Latin indices i,j,i,j,\dots are used for spatial coordinate components in a given coordinate system.

II Mechanics of single and binary black holes

In this section, we provide a summary of known results on the laws of mechanics for single (isolated) and binary black hole systems. For the convenience of the reader, we collect in one place many results that were previously established in several papers. We discuss the various zeroth and first laws of mechanics that have been derived for isolated, stationary black holes in Sec. II.1, for binary systems of corotating black holes in Sec. II.2, for a spinning black hole perturbed by a corotating moon in Sec. II.3, for two spinning point particles in Sec. II.4, and for two corotating point particles in Sec. II.5. An analogy between the horizon surface gravity of a black hole and the redshift of a point particle is discussed in Sec. II.6.

II.1 Single stationary black hole

First, we consider an isolated, stationary black hole of mass MM, angular momentum SS, and horizon surface area AA. Hereafter, we will frequently use the irreducible mass μ=A/(16π)\mu=\sqrt{A/(16\pi)} instead of the horizon surface area AA in various formulas.

II.1.1 Zeroth law

The event horizon of a stationary black hole is a Killing horizon, i.e., a null hypersurface \mathcal{H} whose geodesic generators are tangent to a Killing vector field kαk^{\alpha}. For any Killing horizon, the surface gravity κ\kappa is the scalar field defined on \mathcal{H} by

κ212αkβαkβ|.\kappa^{2}\equiv-\frac{1}{2}\nabla^{\alpha}k^{\beta}\,\nabla_{\alpha}k_{\beta}\big{|}_{\mathcal{H}}\,. (1)

Assuming the dominant energy condition, Bardeen, Carter and Hawking Ba.al.73 established that the scalar field (1) must be constant over \mathcal{H}. The zeroth law of black hole mechanics then states that the surface gravity κ\kappa of a stationary black hole is constant over its event horizon \mathcal{H}. The constant κ\kappa is also known as the non-affinity coefficient, because it is directly related to the failure of the null geodesic generators to be affinely parameterized:

kααkβ|=κkβ|.k^{\alpha}\nabla_{\alpha}k^{\beta}|_{\mathcal{H}}=\kappa\,k^{\beta}|_{\mathcal{H}}\,. (2)

Specializing to a Kerr black hole of mass MM and angular momentum (or spin) SS, we denote by tαt^{\alpha} the timelike Killing field normalized to 1-1 at infinity, and by ϕα\phi^{\alpha} the axial Killing field with integral curves of parameter length 2π2\pi. The horizon-generating Killing field kαk^{\alpha} must be a linear combination of tαt^{\alpha} and ϕα\phi^{\alpha}, the only two generators of isometries of the Kerr metric. Choosing the normalization of kαk^{\alpha} such that kαtα=1k^{\alpha}t_{\alpha}=-1 at spatial infinity, we thus have

kα=tα+ωϕα,k^{\alpha}=t^{\alpha}+\omega\,\phi^{\alpha}\,, (3)

where the constant ω\omega can be interpreted as the angular velocity of the horizon’s generators, namely the apparent rotation rate of the black hole, as measured by a static, inertial observer at infinity. With this choice of normalization, the condition kαkα|=0k^{\alpha}k_{\alpha}|_{\mathcal{H}}=0 and the definition (1) yield the following expressions for the constant surface gravity and angular velocity of a Kerr black hole Poi:

κ\displaystyle\kappa =1χ22M(1+1χ2),\displaystyle=\frac{\sqrt{1-\chi^{2}}}{2M\bigl{(}1+\sqrt{1-\chi^{2}}\bigr{)}}\,, (4a)
ω\displaystyle\omega =χ2M(1+1χ2),\displaystyle=\frac{\chi}{2M\bigl{(}1+\sqrt{1-\chi^{2}}\bigr{)}}\,, (4b)

where χS/M2\chi\equiv S/M^{2} is the dimensionless Kerr parameter, such that 0χ<10\leqslant\chi<1. For a nonspinning black hole, ω=0\omega=0 and κ=1/(4M)\kappa=1/(4M), as expected from a (naive) Newtonian calculation of the surface gravity of a spherically symmetric body of mass MM and typical radius 2M2M.

II.1.2 First law

On the other hand, the celebrated first law of black hole mechanics of Bardeen et al. Ba.al.73 relates the changes in the mass, spin and surface area of nearby stationary and axisymmetric black holes solutions according to222Recalling the relation A=16πμ2A=16\pi\mu^{2} between the surface area AA and the irreducible mass μ\mu, the last term in the right-hand side of Eq. (5) could also be written as 4κμδμ4\kappa\mu\,\delta\mu.

δM=ωδS+κ8πδA.\delta M=\omega\,\delta S+\frac{\kappa}{8\pi}\,\delta A\,. (5)

Iyer and Wald IyWa.94 gave a strengthened form of this result, proving that it holds for arbitrary nonsingular, vacuum perturbations of a stationary black hole that are asymptotically flat at spatial infinity and regular on the event horizon \mathcal{H}. Moreover, a “physical process version” of the first law (5) was derived GaWa.01, and extensions to electrovac and non-vacuum spacetimes are known as well Ca.10. By noticing that the mass MM is a homogeneous function of degree one in the variables S1/2S^{1/2} and A1/2A^{1/2}, it can be shown that there exists a “first integral” associated with the variational formula (5), namely Sm.73

M=2ωS+κA4π.M=2\omega S+\frac{\kappa A}{4\pi}\,. (6)

Hereafter, it will prove convenient to have the expressions for the mass MM and spin SS as functions of the irreducible mass μ\mu and the horizon angular velocity ω\omega. Combining the first law (5) with the Christodoulou mass formula M2=μ2+S2/(4μ2)M^{2}\!=\!\mu^{2}+S^{2}/(4\mu^{2}) for Kerr black holes Ch.70, it can easily be shown that

M\displaystyle M =μ14(μω)2,\displaystyle=\frac{\mu}{\sqrt{1-4(\mu\omega)^{2}}}\,, (7a)
S\displaystyle S =4μ3ω14(μω)2.\displaystyle=\frac{4\mu^{3}\omega}{\sqrt{1-4(\mu\omega)^{2}}}\,. (7b)

In the slow rotation limit μω1\mu\omega\ll 1, one finds the expression S4μ3ωS\simeq 4\mu^{3}\omega for the black hole spin, as expected for a body of mass MμM\simeq\mu with typical radius 2M2M and rotational velocity 2Mω2M\omega. Then, substituting for Eqs. (7) into the formula (4a), we obtain the following expression for the normalized surface gravity 4μκ4\mu\kappa of a Kerr black hole as a function of μ\mu and ω\omega:

4μκ=18(μω)214(μω)2.4\mu\kappa=\frac{1-8(\mu\omega)^{2}}{\sqrt{1-4(\mu\omega)^{2}}}\,. (8)

II.2 Two corotating black holes

In a seminal work, Friedman, Uryū and Shibata Fr.al.02 generalized the zeroth and first laws discussed above to spacetimes containing multiple black holes and/or a generic distribution of perfect fluid matter sources with compact support, having in mind the application of their general results to binary systems of black holes and/or neutron stars.333See Ref. Ur.al.10 for the generalization of their work to the magnetohydrodynamic case.

II.2.1 Zeroth law

We consider vacuum, binary black hole spacetimes endowed with a global helical Killing vector kαk^{\alpha}.​ Following Refs.​ Go.al.02; Fr.al.02, a spacetime is said to have a helical Killing symmetry if the generator of the isometry can be written in the form

kα=tα+Ωϕα,k^{\alpha}=t^{\alpha}+\Omega\,\phi^{\alpha}\,, (9)

where Ω>0\Omega>0 is a constant, tαt^{\alpha} is timelike and ϕα\phi^{\alpha} is spacelike with circular orbits of parameter length 2π2\pi. In general, neither tαt^{\alpha} nor ϕα\phi^{\alpha} is a Killing vector, but the combination (9) is a Killing vector for a specific value of the constant Ω\Omega, which can be interpreted as the angular frequency of the binary black hole system.

Physically, the only allowed configurations correspond to a state of corotation, for which the null geodesic generators of each component of the horizon are tangent to the Killing vector field (9). Indeed, if corotation were not realized, the resulting non-vanishing shear would result in the growth of the horizon’s surface areas, in contradiction with the hypothesis of helical symmetry. For such corotating binaries, each component a\mathcal{H}_{a} of the event horizon is a Killing horizon with horizon-generating Killing field kα|ak^{\alpha}|_{\mathcal{H}_{a}}. The authors of Ref. Fr.al.02 could then show that the surface gravity κa\kappa_{a}, as defined by (1), is indeed constant on a\mathcal{H}_{a}:

κa212αkβαkβ|a=const.\kappa_{a}^{2}\equiv-\frac{1}{2}\nabla^{\alpha}k^{\beta}\,\nabla_{\alpha}k_{\beta}\big{|}_{\mathcal{H}_{a}}=\text{const.} (10)

However, in GR it is known that helically symmetric spacetimes cannot be asymptotically flat GiSt.84; De.89; Kl.04. This fact can easily be understood from a heuristic point of view: in order to maintain the binary on a fixed circular orbit, the energy radiated in gravitational waves needs to be compensated by an equal amount of incoming radiation. Far away from the source, the resulting system of standing waves ends up dominating the energy content of the spacetime, such that the falloff conditions necessary to ensure asymptotic flatness cannot be satisfied. This lack of asymptotic flatness implies that there is no natural normalization of the helical Killing field kαk^{\alpha}, and hence the numerical value of the surface gravity κa\kappa_{a} of each black hole is entirely free.

Nevertheless, asymptotic flatness can be recovered if, loosely speaking, the gravitational radiation can be “turned off.” This can be achieved, in particular, using the Isenberg, Wilson and Mathews approximation to GR, also known as the conformal flatness condition (CFC) approximation IsNe.80; WiMa.89; Is.08. In Sec. III, we will rely on this approximation to compute sequences of quasi-equilibrium initial data for corotating black hole binaries, and to properly normalize the surface gravity of each black hole.

II.2.2 First law

Consider a vacuum spacetime containing multiple black holes, and endowed with a global Killing vector field kαk^{\alpha}. The Noether current associated with kαk^{\alpha} assigns to each spacetime a conserved charge QQ. The main result established in Ref. Fr.al.02 relates the variation δQ\delta Q of the conserved charge to the variations δAa\delta A_{a} of the black hole’s horizon surface areas, namely

δQ=aκa8πδAa,\delta Q=\sum_{a}\frac{\kappa_{a}}{8\pi}\,\delta A_{a}\,, (11)

where κa\kappa_{a} is the constant horizon surface gravity of the aath black hole. When matter sources are present, additional terms appear in the right-hand side of this formula. We are interested in applying the general result (11) to the particular case of a binary system of corotating black holes moving along circular orbits.

For such spacetimes, the geometry is invariant along the integral curve of a helical Killing vector field (9). In general, neither tαt^{\alpha} nor ϕα\phi^{\alpha} are Killing vectors. However, for asymptotically flat spacetimes tαt^{\alpha} and ϕα\phi^{\alpha} are asymptotically Killing, and the variation of the conserved charge QQ takes the form δQ=δMADMΩδJ\delta Q=\delta M_{\text{ADM}}-\Omega\,\delta J Fr.al.02, where MADMM_{\text{ADM}} is the Arnowitt-Deser-Misner (ADM) mass and JJ is the total angular momentum, both defined as surface integrals at spatial infinity. Then, the general formula (11) reduces to the first law of binary black hole mechanics:

δMADM=ΩδJ+aκa8πδAa.\delta M_{\text{ADM}}=\Omega\,\delta J+\sum_{a}\frac{\kappa_{a}}{8\pi}\,\delta A_{a}\,. (12)

Notice that, in the binary black hole case, the horizon angular velocity ω\omega of a single rotating black hole is replaced by the orbital frequency Ω\Omega of the binary system.

Just like the first law (5) for a single black hole admits a “first integral,” namely Smarr’s formula (6), in the binary black hole case there is also a first integral associated with the variational formula (12), namely Le.al.12

MADM=2ΩJ+aκaAa4π.M_{\text{ADM}}=2\Omega J+\sum_{a}\frac{\kappa_{a}A_{a}}{4\pi}\,. (13)

This algebraic relationship will play a key role in Sec. III, by allowing us to test the numerical accuracy of our sequences of quasi-equilibrium initial data for corotating binary black holes.

II.3 Black hole and corotating moon

Gralla and Le Tiec GrLe.13 have shown how the zeroth and first laws of black hole mechanics can be extended to the case where a Kerr black hole is perturbed by a small orbiting compact object (a “moon”) on the corotating orbit, the unique circular equatorial orbit whose orbital frequency matches the angular velocity of the Kerr horizon. The natural framework for such a problem is that of black hole perturbation theory. Therefore, in this section we consider a Kerr black hole of mass MM and spin SS, perturbed by a corotating point particle of mass mMm\ll M, and we work to linear order in the small mass ratio qm/Mq\equiv m/M.

II.3.1 Zeroth law

The metric perturbation generated by the corotating moon is helically symmetric, asymptotically flat at future null infinity,444By working to linear order in the mass ratio, the metric perturbation is (in a sense) taken to be “infinitesimally small,” such than an helically symmetric spacetime can be asymptotically flat as well Ke.al2.10. and regular on the black hole’s perturbed event horizon \mathcal{H}. Combining those properties, it can be shown that the expansion and shear of \mathcal{H} must vanish, thus ensuring (through rigidity arguments) that the perturbed event horizon is a Killing horizon GrLe.13. The horizon-generating Killing field, say kαk^{\alpha}, must then be a linear combination of (i) the helical Killing field α\ell^{\alpha} of the perturbed spacetime, simply given by α=tα+ω¯ϕα\ell^{\alpha}=t^{\alpha}+\bar{\omega}\,\phi^{\alpha} in adapted coordinates, where ω¯\bar{\omega} is the angular velocity of the background Kerr black hole [recall Eq. (4b)], and (ii) perturbations of the background Killing fields inherited from the isometries of the Kerr geometry. Since the perturbed spacetime is asymptotically Minkowskian at future null infinity, the horizon-generating Killing field kαk^{\alpha} may be normalized there according to

kα=tα+(ω¯+𝒟ω)ϕα,k^{\alpha}=t^{\alpha}+\bigl{(}\bar{\omega}+\mathcal{D}\omega\bigr{)}\,\phi^{\alpha}\,, (14)

where the constant 𝒟ω=𝒪(q)\mathcal{D}\omega=\mathcal{O}(q) is the perturbation in angular velocity induced by the moon. By analogy with the binary black hole case, the constant Ωω¯+𝒟ω\Omega\equiv\bar{\omega}+\mathcal{D}\omega will be referred to as the circular-orbit frequency of the binary system.

Since \mathcal{H} is a Killing horizon, the surface gravity κ\kappa of the perturbed black hole, as defined by (1), is constant over \mathcal{H}. It can be written in the form κ=κ¯+𝒟κ\kappa=\bar{\kappa}+\mathcal{D}\kappa, where κ¯\bar{\kappa} is the surface gravity of the background Kerr black hole [recall Eq. (4a)], and 𝒟κ=𝒪(q)\mathcal{D}\kappa=\mathcal{O}(q) is the constant perturbation induced by the corotating moon. Remarkably, simple closed-form expressions for the perturbations 𝒟κ\mathcal{D}\kappa and 𝒟ω\mathcal{D}\omega in surface gravity and angular velocity can be established, namely GrLe.13

𝒟κ\displaystyle\mathcal{D}\kappa =zκ¯HM,\displaystyle=z\,\bar{\kappa}\,\frac{\partial H}{\partial M}\,, (15a)
𝒟ω\displaystyle\mathcal{D}\omega =zω¯HM+zHS,\displaystyle=z\,\bar{\omega}\,\frac{\partial H}{\partial M}+z\,\frac{\partial H}{\partial S}\,, (15b)

where zkαuαz\equiv-k^{\alpha}u_{\alpha} is the “redshift” of the particle, the conserved orbital quantity associated with the helical Killing symmetry of the perturbed spacetime, with uαu^{\alpha} the particle’s four-velocity. The formulas (15) involve the canonical Hamiltonian HH that controls the timelike geodesic motion of a test particle of mass mm in the Kerr geometry.

An explicit calculation reveals that the relative changes in horizon surface gravity and angular velocity induced by the orbiting moon are given by the expressions Le.12; GrLe.13

𝒟κκ¯\displaystyle\frac{\mathcal{D}\kappa}{\bar{\kappa}} =qv21+χv31+2χv3χ2v413v2+2χv3,\displaystyle=-q\,\,\frac{v^{2}}{1+\chi v^{3}}\,\frac{1+2\chi v^{3}-\chi^{2}v^{4}}{\sqrt{1-3v^{2}+2\chi v^{3}}}\,, (16a)
𝒟ωω¯\displaystyle\frac{\mathcal{D}\omega}{\bar{\omega}} =qv21+χv32sv3χ(1+sv4)χ13v2+2χv3,\displaystyle=q\,\,\frac{v^{2}}{1+\chi v^{3}}\,\frac{2sv^{3}-\chi(1+sv^{4})}{\chi\sqrt{1-3v^{2}+2\chi v^{3}}}\,, (16b)

where we introduced the shorthands v3MΩ/(1χMΩ)v^{3}\!\equiv\!M\Omega/(1-\chi M\Omega) and s2+21χ2χ2s\!\equiv\!2+2\sqrt{1-\chi^{2}}-\chi^{2}. At this order of approximation, we may substitute Ω\Omega by ω¯\bar{\omega} and use the relation (4b) between ω¯\bar{\omega} and χ\chi, so that the relative changes (16) are functions of χ\chi (or ω¯\bar{\omega}) only. Notice that the change in surface gravity induced by the tidal field of the moon is negative, while the change in horizon angular velocity is positive. Hence, from the point of view of a distant inertial observer, the apparent rotation rate of the black hole is increased by the corotating moon.

II.3.2 First law

Adapting Iyer’s and Wald’s formalism IyWa.94 to nonvacuum perturbations of a nonstationary black hole spacetime that are asymptotically flat at future null infinity, the authors of GrLe.13 could generalize the first law (5) to binary systems of spinning black holes with corotating moons. Their first law relates variations in the Bondi mass MBM_{\text{B}} and angular momentum JBJ_{\text{B}} of the perturbed spacetime, as measured at future null infinity, to variations in the perturbed black hole’s surface area AA and the moon’s mass mm, and reads555The symbol δ\delta, used to denote a variation within the three-parameter family of black hole with corotating moon spacetimes, should not be confused with the symbol 𝒟\mathcal{D}, used to denote a small change induced by the nonvanishing mass mm of the orbiting moon.

δMB=ΩδJB+κ8πδA+zδm,\delta M_{\text{B}}=\Omega\,\delta J_{\text{B}}+\frac{\kappa}{8\pi}\,\delta A+z\,\delta m\,, (17)

where we recall that zz is the particle’s “redshift.” Here, κ=κ¯+𝒟κ\kappa=\bar{\kappa}+\mathcal{D}\kappa is the constant horizon surface gravity of the perturbed black hole, while Ω=ω¯+𝒟ω\Omega=\bar{\omega}+\mathcal{D}\omega is the circular-orbit frequency of the binary [recall Eq. (14)]. The derivation of the first law (17) is analogous to that outlined in Sec. II.2. In particular, the combination δMBΩδJB\delta M_{\text{B}}-\Omega\,\delta J_{\text{B}} appears when evaluating the variation of the conserved Noether charge associated with the helical Killing vector field kα=tα+Ωϕαk^{\alpha}=t^{\alpha}+\Omega\,\phi^{\alpha}, except that the surface integral is performed on a two-sphere at future null infinity, and not at spatial infinity.

Moreover, just like the first laws (5) and (12) for isolated black holes and corotating black hole binaries admit first integral relations, there is a first integral associated with the first law (17) for a black hole with a corotating moon, namely GrLe.13

MB=2ΩJB+κA4π+zm.M_{\text{B}}=2\Omega J_{\text{B}}+\frac{\kappa A}{4\pi}+zm\,. (18)

This can be established either by applying Stokes’ theorem to the Noether charge associated with kαk^{\alpha}, or by combining Euler’s theorem for homogeneous functions with Eq. (17). Note that the first law (17) and the first integral (18) were established up to relative 𝒪(q)\mathcal{O}(q).

II.4 Two spinning point particles

Next, we consider binary systems of compact objects, modelled as (spinning) point particles with constant masses mam_{a}. Although the concept of a point particle does not make sense in exact GR GeTr.87—the closest thing to a point particle in GR is a black hole—, in the context of approximation schemes such as PN theory Bl.14, Einstein’s equation can be coupled to a distributional matter source in a meaningful manner, provided that a regularization scheme (e.g. dimensional regularization) is used to handle the divergent self-field of each particle.

Since a point particle has no “spatial” extension, and in particular no horizon, a putative zeroth law of mechanics is meaningless for binary systems of point particles. Still, a first law can be derived by assuming the existence of a global helical Killing field kαk^{\alpha}, and by taking a formal point-particle limit in the generalized first law of Fr.al.02, in the case of two self-gravitating balls of perfect fluid. The resulting first law of binary point-particle mechanics reads Le.al.12

δMADM=ΩδJ+azaδma,\delta M_{\text{ADM}}=\Omega\,\delta J+\sum_{a}z_{a}\,\delta m_{a}\,, (19)

where zauaαkαz_{a}\equiv-u_{a}^{\alpha}k_{\alpha} is the “redshift” of particle aa, a conserved orbital quantity associated with the helical Killing symmetry, with uaαu_{a}^{\alpha} the four-velocity of particle aa. Choosing coordinates adapted to that isometry, such that kα=(t)α+Ω(ϕ)αk^{\alpha}={(\partial_{t})}^{\alpha}+\Omega\,{(\partial_{\phi})}^{\alpha} holds everywhere, and not merely in a neighbourhood of infinity, the redshift of particle aa is simply za=dτa/dtz_{a}=\mathrm{d}\tau_{a}/\mathrm{d}t, the ratio of the proper times elapsed along the worldline of the particle and along that of a distant static observer. As expected, the first integral associated with the variational formula (19) is

MADM=2ΩJ+azama.M_{\text{ADM}}=2\Omega J+\sum_{a}z_{a}m_{a}\,. (20)

The first law (19) was later recovered, and extended to spinning point particles, using the canonical ADM Hamiltonian framework Bl.al.13. Assuming that the conservative dynamics of a binary system of spinning point particles with masses mam_{a} and canonical spins 𝐒a\mathbf{S}_{a} derives from an autonomous Fokker-type two-body Hamiltonian HF(𝐱a,𝐩a;ma,𝐒a)H_{\text{F}}(\mathbf{x}_{a},\mathbf{p}_{a};m_{a},\mathbf{S}_{a}), with 𝐱a\mathbf{x}_{a} and 𝐩a\mathbf{p}_{a} the canonical positions and momenta, the first law for circular orbits and spins aligned or anti-aligned with the orbital angular momentum reads666Strictly speaking, the mass appearing in the left-hand side of Eq. (21) is the on-shell value of the local, autonomous Fokker Hamiltonian HF(𝐱a,𝐩a;ma,𝐒a)H_{\text{F}}(\mathbf{x}_{a},\mathbf{p}_{a};m_{a},\mathbf{S}_{a}).

δMADM=ΩδJ+a[zaδma+(ΩaΩ)δSa].\delta M_{\text{ADM}}=\Omega\,\delta J+\sum_{a}\left[z_{a}\,\delta m_{a}+(\Omega_{a}-\Omega)\,\delta S_{a}\right]. (21)

Interestingly, while the precession frequencies Ωa\Omega_{a} of the spins are given by the partial derivatives of the two-body Hamiltonian with respect to the particle’s spins Da.al.08, it was shown in Ref. Bl.al.13 that (at least to linear order in the spins) the redshifts zaz_{a} are given by the partial derivatives of the Hamiltonian with respect to the particle’s masses:

Ωa=HFSa,za=HFma.\Omega_{a}=\frac{\partial H_{\text{F}}}{\partial S_{a}}\,,\qquad z_{a}=\frac{\partial H_{\text{F}}}{\partial m_{a}}\,. (22)

Using the canonical ADM framework, the first law (19) for circular orbits was also extended to generic bound (i.e. eccentric) orbits in Refs. Le.15; BlLe.17. These first laws for binary systems of (spinning) point masses, as well as the associated first integrals, have been explictly checked to hold true up to 3PN order included, as well as the logarithmic contributions at the 4PN and 5PN orders.

II.5 Two corotating point particles

Motivated by the first law (12) for binary systems of corotating black holes, the authors of Ref. Bl.al.13 suggested modelling each spinning point particle as an equilibrium black hole in a tidal environment, endowed with an “irreducible mass” μa\mu_{a} and a “proper rotation frequency” ωa\omega_{a}. Then, by analogy with the first law (5) for a single, isolated black hole, each spinning point particle with mass mam_{a} and spin SaS_{a} was assumed to obey the variational formula

δma=caδμa+ωaδSa.\delta m_{a}=c_{a}\,\delta\mu_{a}+\omega_{a}\,\delta S_{a}\,. (23)

To obtain explicit expressions for the coefficients ca=(ma/μa)Sac_{a}=(\partial m_{a}/\partial\mu_{a})_{S_{a}} and ωa=(ma/Sa)μa\omega_{a}=(\partial m_{a}/\partial S_{a})_{\mu_{a}}, it was further assumed that each spinning point particle obeys a Christodoulou mass formula of the type ma2=μa2+Sa2/(4μa2)m_{a}^{2}=\mu_{a}^{2}+S_{a}^{2}/(4\mu_{a}^{2}). Then, for each particle the “response coefficient” cac_{a}, which is the point-particle analogue of the normalized surface gravity 4μκ4\mu\kappa of an isolated black hole, can be expressed as a function of e.g. μa\mu_{a} and ωa\omega_{a}, and similarly for the mass mam_{a} and spin SaS_{a}. From Eqs. (7) and (8), we immediately find

ma\displaystyle m_{a} =μa14(μaωa)2,\displaystyle=\frac{\mu_{a}}{\sqrt{1-4(\mu_{a}\omega_{a})^{2}}}\,, (24a)
Sa\displaystyle S_{a} =4μa3ωa14(μaωa)2,\displaystyle=\frac{4\mu_{a}^{3}\omega_{a}}{\sqrt{1-4(\mu_{a}\omega_{a})^{2}}}\,, (24b)
ca\displaystyle c_{a} =18(μaωa)214(μaωa)24μaκ¯a.\displaystyle=\frac{1-8(\mu_{a}\omega_{a})^{2}}{\sqrt{1-4(\mu_{a}\omega_{a})^{2}}}\equiv 4\mu_{a}\bar{\kappa}_{a}\,. (24c)

Substituting for Eq. (23) into the first law (21) for binary systems of spinning point particles yields a formula that involves the variations δμa\delta\mu_{a} and δSa\delta S_{a} of the irreducible masses and the spins, namely δMADM=ΩδJ+a[cazaδμa+(ΩaΩ+zaωa)δSa]\delta M_{\text{ADM}}=\Omega\,\delta J+\sum_{a}\left[c_{a}z_{a}\,\delta\mu_{a}+(\Omega_{a}-\Omega+z_{a}\omega_{a})\,\delta S_{a}\right]. By analogy with the first law (12) for binary systems of corotating black holes, the proper rotation state of each spinning particle is constrained to a corotation state if, and only if, the coefficients in front of the variations δSa\delta S_{a} vanish, yielding the corotation conditions (for a=1,2a=1,2)

zaωa=ΩΩa.z_{a}\,\omega_{a}=\Omega-\Omega_{a}\,. (25)

Physically, this means that the redshifted proper rotation frequency of each spinning point particle, zaωaz_{a}\omega_{a}, must be equal to the circular orbit frequency Ω\Omega, as seen in a frame rotating at the angular rate Ωa\Omega_{a} with respect to an inertial frame of reference.

When the corotating condition (25) is imposed, the first law (21) for binary systems of spinning point particles reduces to

δMADM=ΩδJ+acazaδμa.\delta M_{\text{ADM}}=\Omega\,\delta J+\sum_{a}c_{a}z_{a}\,\delta\mu_{a}\,. (26)

Because the first law (12) for corotating black hole binaries can be written in the equivalent form δMADM=ΩδJ+a4μaκaδμa\delta M_{\text{ADM}}=\Omega\,\delta J+\sum_{a}4\mu_{a}\kappa_{a}\,\delta\mu_{a}, it appears that the renormalized redshifts cazac_{a}z_{a} in corotating point particle binaries are analogous to the normalized surface gravities 4μaκa4\mu_{a}\kappa_{a} in corotating black hole binaries. This will be further discussed in the next Sec. II.6. To conclude this section, we point out that the first integral relation associated with the variational first law (26) reads, as expected,

MADM=2ΩJ+acazaμa.M_{\text{ADM}}=2\Omega J+\sum_{a}c_{a}z_{a}\mu_{a}\,. (27)

II.6 Surface gravity and redshift

We can now compare the first law (12) for corotating black hole binaries to the first law (26) for corotating point particles binaries, as well as the first integrals (13) and (27). Clearly, if the irreducible masses μa\mu_{a} are identified for each body, then the surface gravity κa\kappa_{a} of each corotating black hole is analogous to the redshift zaz_{a} of each corotating point particle, and more precisely

4μaκacaza.4\mu_{a}\kappa_{a}\longleftrightarrow c_{a}z_{a}\,. (28)

Recall that the coefficient cac_{a} introduced in Eq. (23) is nothing but the normalized horizon surface gravity of an isolated black hole of irreducible mass μa\mu_{a} and horizon angular velocity ωa\omega_{a}, say 4μaκ¯a4\mu_{a}\bar{\kappa}_{a} [see Eqs. (8) and (24c)]. Therefore, the analogy (28) can equivalently be written as

κaκ¯aza.\frac{\kappa_{a}}{\bar{\kappa}_{a}}\longleftrightarrow z_{a}\,. (29)

Since the redshift of a particle satisfies za1z_{a}\leqslant 1, the analogy (29) suggests that in corotating black hole binary systems the tidal interaction decreases the surface gravity of each black hole (with respect to its value in isolation). This agrees with the prediction of the perturbative calculation discussed in Sec. II.3.

In the large separation limit, i.e. when Ω0\Omega\to 0, we know that za1z_{a}\to 1 for each corotating particle, while κaκ¯a=1/(4μa)\kappa_{a}\to\bar{\kappa}_{a}=1/(4\mu_{a}) for each corotating black hole. Indeed, in that limit, the spin and proper rotation frequency of each body must vanish to maintain the corotation. To test the soundness of the analogy (28) beyond the large separation limit, we will compute the normalized black hole surface gravities 4κaμa4\kappa_{a}\mu_{a} in Sec. III, as well as the 4PN expansions of the particle’s renormalized redshifts cazac_{a}z_{a} in Sec. IV, and compare those results, expressed a functions of the circular-orbit frequency Ω\Omega, in Sec. VI.

Interestingly, for black hole binaries with large mass ratios, the analogy (29) is supported by a comparison, at the Hamiltonian level, of the expression for the surface gravity of a black hole tidally perturbed by a corotating moon, and that for the redshift of the most massive particle (say body 22). Indeed, recalling Eqs. (15a) and (22), we find the analogy

κκ¯=H~Mz2=HFm2,\frac{\kappa}{\bar{\kappa}}=\frac{\partial\tilde{H}}{\partial M}\quad\longleftrightarrow\quad z_{2}=\frac{\partial H_{\text{F}}}{\partial m_{2}}\,, (30)

where we introduced the Hamiltonian H~M+zH\tilde{H}\equiv M+zH, including the mass MM of the background hole. The occurence of the moon’s redshift zz in H~\tilde{H} comes from the fact that the Hamiltonian HH of a test mass in Kerr is parameterized by the proper time τ\tau, while the Fokker-type two- body Hamiltonian HFH_{\text{F}} is parameterized by the coordinate time tt.

Similarly, by comparing the expression (15b) for the perturbed angular velocity of the black hole to the corotation condition (25) with the formulas (22) for the point particle’s proper rotation frequency, we note a formal analogy between ω¯\bar{\omega} and ω2\omega_{2}, namely

ω¯=(H~M)1(ΩH~S)ω2=(HFm2)1(ΩHFS2).\bar{\omega}=\biggl{(}\frac{\partial\tilde{H}}{\partial M}\biggr{)}^{-1}\biggl{(}\Omega-\frac{\partial\tilde{H}}{\partial S}\biggr{)}\quad\longleftrightarrow\quad\omega_{2}=\biggl{(}\frac{\partial H_{\text{F}}}{\partial m_{2}}\biggr{)}^{-1}\biggl{(}\Omega-\frac{\partial H_{\text{F}}}{\partial S_{2}}\biggr{)}\,. (31)

In Secs. IV and V below, we will find that the large mass-ratio limits of the PN expansions of z2(Ω)z_{2}(\Omega) and ω2(Ω)\omega_{2}(\Omega) are in full agreement with the PN expansions of the perturbative expressions for (κ/κ¯)(Ω)(\kappa/\bar{\kappa})(\Omega) and ω¯(Ω)\bar{\omega}(\Omega), thus supporting the analogies (30) and (31).

On the other hand, still for black hole binaries with large mass ratios, the analogy (29) can be made rigorous for the smaller compact object (say body 11). Indeed, using a method of matched asymptotic expansions, the author of Ref. Po3.15 recently proved that the horizon surface gravity κ1\kappa_{1} of a small black hole of mass mm in the tidal environment of a large black hole of mass MmM\gg m is closely related to the redshift zz1z\equiv z_{1} of a point mass mm moving like a test particle in a certain effective metric satisfying the linearized vacuum Einstein equation. More precisely, to linear order in the small mass ratio q=m/Mq=m/M, it can be shown that Po3.15

κ1κ¯1=z.\frac{\kappa_{1}}{\bar{\kappa}_{1}}=z\,. (32)

At this order of approximation, the smaller black hole can be taken to be nonrotating, such that m=μ1m=\mu_{1}, κ¯1=1/(4μ1)\bar{\kappa}_{1}=1/(4\mu_{1}) and c1=1c_{1}=1. Hence the result (32) proves that, at least to linear order in the mass ratio and for the smaller body 11, Eq. (28) is an identity, and not a mere analogy. See also the discussion in Ref. Zi.al.16.

III Numerical relativity

In this section, we present our calculations of sequences of quasi-equilibrium initial data for corotating black hole binaries with mass ratios in the range 10:11:110:1-1:1. In particular, we explain how the horizon surface gravity of each black hole is computed. The formulation used to construct the initial data is outlined in Sec. III.1, the dependence of our numerical data on the resolution is discussed in Sec. III.2, and our results are presented in Sec. III.3.

III.1 Formulation

The numerical results presented here are similar to those previosly published in Gr.al.02; Gr.10; Ur.al.12. The basic properties of such data are as follows. We employ the 3+1 formalism of GR so that the four-dimensional line element reads

ds2=(N2BiBi)dt2+2Bidxidt+γijdxidxj,\mathrm{d}s^{2}=-\bigl{(}N^{2}-B_{i}B^{i}\bigr{)}\,\mathrm{d}t^{2}+2B_{i}\,\mathrm{d}x^{i}\mathrm{d}t+\gamma_{ij}\,\mathrm{d}x^{i}\mathrm{d}x^{j}\,, (33)

where NN is the lapse, BiB^{i} the shift vector and γij\gamma_{ij} the three-dimensional spatial metric. Spatial indices are lowered and raised by means of γij\gamma_{ij} and its inverse γij\gamma^{ij}, so that e.g. BiγijBjB_{i}\equiv\gamma_{ij}B^{j}. Circularity of the orbit is enforced by imposing the existence of a helical Killing vector kαk^{\alpha}; recall Eq. (9). From a practical point of view, the equations are written in the corotating frame, such that kα=(t)αk^{\alpha}=\left(\partial_{t}\right)^{\alpha} and the time derivatives of the various fields vanish.

Our corotation setup allows us to consider binary black hole systems with an exact helical Killing symmetry, such that the apparent horizon is a Killing horizon, and the surface gravity is well defined over each component of the horizon. This should be contrasted to the recent work of Ref.Zi.al.16, where the authors consider more generic and astrophysically realistic binary configurations, but do not have a well-defined (geometrical and unambiguous) notion of black hole surface gravity.

Far from the system, it is demanded that the spacetime be asymptotically flat. However, as discussed in Sec. II.2, this is not strictly compatible with the existence of a global helical symmetry. Indeed, if a true helical Killing vector existed, the system would have an infinite lifetime. This fact, coupled with the emission of gravitational waves would lead to diverging metric quantities, being the signature of this accumulated emission. There are two possible approaches to cure that problem. Either one can (to some extend) relax the Killing condition, or one needs to find a way to suppress the gravitational radiation. In this work, we opt for the latter solution. There are various ways to proceed. The most widely used is to impose that the solution be spatially conformally flat. The spatial metric γij\gamma_{ij} is then simply related to the three-dimensional flat metric fijf_{ij} through a conformal factor Ψ\Psi, namely

γij=Ψ4fij.\gamma_{ij}=\Psi^{4}f_{ij}\,. (34)

Of course, the price to pay is that not all of Einstein’s equations can be exactly satisfied.

With conformal flatness, the spacetime depends on five fields: the lapse NN, the conformal factor Ψ\Psi and the shift vector BiB^{i}. Those fields obey a subset of Einstein’s equations, namely the trace of the evolution equation (35a), the Hamiltonian constraint (35b) and the momentum constraints (35c). The slicing (i.e. the choice of time) is done via maximal slicing, for which the trace of the extrinsic curvature tensor vanishes. The spacelike coordinates are fixed by the choice of corotating coordinates and the conformal flatness approximation. The resulting system of equations is then

DiDiN\displaystyle D_{i}D^{i}N =2DiNDiΨΨ+NΨ4AijAij,\displaystyle=-2D^{i}N\frac{D_{i}\Psi}{\Psi}+N\Psi^{4}A_{ij}A^{ij}\,, (35a)
DiDiΨ\displaystyle D_{i}D^{i}\Psi =Ψ58AijAij,\displaystyle=-\frac{\Psi^{5}}{8}A_{ij}A^{ij}\,, (35b)
DjDjBi+13DiDjBj\displaystyle D_{j}D^{j}B^{i}+\frac{1}{3}D^{i}D_{j}B^{j} =2Aij(DjN6NDjΨΨ),\displaystyle=2A^{ij}\biggl{(}D_{j}N-6N\frac{D_{j}\Psi}{\Psi}\biggr{)}\,, (35c)

where DD denotes the covariant derivative compatible with the three-dimensional Euclidean metric fijf_{ij}, while AijDiBj/NA^{ij}\equiv D^{\langle i}B^{j\rangle}/N is the conformal extrinsic curvature tensor.

The presence of two corotating black holes is enforced by imposing the following boundary conditions on two spheres corresponding to the apparent horizons:

N\displaystyle N =n0,\displaystyle=n_{0}\,, (36a)
rΨ+Ψ2r\displaystyle\partial_{r}\Psi+\frac{\Psi}{2r} =Ψ34Aijsisj,\displaystyle=-\frac{\Psi^{3}}{4}A_{ij}s^{i}s^{j}\,, (36b)
Bi\displaystyle B^{i} =n0Ψ2si,\displaystyle=\frac{n_{0}}{\Psi^{2}}\,s^{i}\,, (36c)

where sis^{i} is the unit normal to each hole (with respect to the metric fijf_{ij}). Equation (36a) is a free choice of the lapse on the boundaries of the physical domain; typically one uses n0=0.1n_{0}=0.1. This basically comes from the fact that maximal slicing is a differential gauge choice, defined uniquely up to the boundaries. This was discussed is some length in Refs. Ca.al2.06; GoJa.06. Equation (36b) enforces that the two spheres are apparent horizons, and (36c) that the black holes are corotating. Asymptotic flatness is explicitly recovered by requesting that N=1N=1, Ψ=1\Psi=1 and Bi=Ω(φ)iB^{i}=\Omega\left(\partial_{\varphi}\right)^{i} at infinity. (Recall that we work in the corotating frame).

For each separation, the constant angular velocity Ω\Omega is determined by the numerical code to ensure the equality of the ADM and Komar masses:

MADM=MK.M_{\text{ADM}}=M_{\text{K}}\,. (37)

This condition is closely related to relativistic generalizations of the virial theorem GoBo.94; Fr.al.02, and it has been used extensively to compute binary black hole configurations, e.g. in Gr.al.02; Gr.10; Ur.al.12. The five elliptic equations (35) are solved by spectral methods implemented via the Kadath library. Bispherical coordinates are used and the algorithm is the same as the one already described in Refs. Gr.10; UrTs.11. The version of the code used for this work is more general in the sense that it can deal with black holes of different masses. When doing so, one needs to find the location of the rotation axis of the binary. This is done numerically and it is obtained from the fact that the total (ADM) linear momentum of the binary must vanish. The linear momentum can be computed as a surface integral at spatial infinity, and the condition can be rewritten as

limrAij(ey)idSj=0,\lim_{r\to\infty}\int A^{ij}\left(e_{y}\right)_{i}{\rm d}S_{j}=0\,, (38)

where ey\vec{e}_{y} is the unit vector (with respect to the flat metric) along the yy-axis and dSi\mathrm{d}S_{i} is the flat surface element at infinity. The various axes are defined as follows : the xx-direction joins the center of the holes and the zz-direction is the one of the rotation axis, so that the orbital plane is z=0z=0. The yy-direction is then in the orbital plane, perpendicular to the line joining the holes. Due to the symmetries of the problem, only the yy component of the momentum is not trivially zero. It vanishes only when the axis is at the right location on the xx-axis.

The coordinate radii of the holes are also unknowns. The sequences of corotating binaries are obtained by requesting that the irreducible mass of each hole is held fixed (see Sec. III.3.2 below). The size of the numerical domain is allowed to vary and the numerical code finds the appropriate radii so that the masses have the desired values. More precisely, the irreducible masses are defined by μaAa/(16π)\mu_{a}\equiv\sqrt{A_{a}/(16\pi)}, where AaA_{a} is the surface area of the aath hole, given by the following surface integral over a cross section SaS_{a} of the corresponding horizon:

Aa=SaΨ4dS.A_{a}=\int_{S_{a}}\!\Psi^{4}\,{\rm d}S\,. (39)

Since, by construction, the horizons have a vanishing expansion, their areas AaA_{a} do not depend on a particular choice of slicing As.al.00. We computed quasi-equilibrium initial data for irreducible mass ratios qμμ1/μ2{1,1/2,1/10}q_{\mu}\equiv\mu_{1}/\mu_{2}\in\{1,1/2,1/10\}.

In order to compute the horizon surface gravity (10) of each hole, we employ a formula given in terms of the 3+1 quantities. A suitable expression can be found, e.g., in Eq. (10.10) of Ref. GoJa.06. There, the normal vector is normalized with respect to the spatial metric γij\gamma_{ij}, whereas the vector sis^{i} appearing in our Eqs. (36b) and (36c) is normalized with respect to the flat metric fijf_{ij}. It follows that the two normal vectors differ by a factor of Ψ2\Psi^{2}. Moreover, the term proportional to ααlnN\ell^{\alpha}\nabla_{\alpha}\ln N in Eq. (10.10) of Ref. GoJa.06 is related to the evolution of the lapse function, and it vanishes for our quasi-stationary configurations. Taking all of that into account, the 3+1 expression for the surface gravity reduces to

κa=siDiNΨ2NAijsisj|Sa.\kappa_{a}=\frac{s^{i}D_{i}N}{\Psi^{2}}-NA_{ij}s^{i}s^{j}\Big{|}_{S_{a}}. (40)

III.2 Effects of resolution

Next, we explore the variations of some quantities of physical interest with respect to the numerical resolution. The three-dimensional physical space is split into several (typically a dozen) computational domains. In each domain the number of spectral coefficients is the same with respect to the three space dimensions. We denote by NN this number and compute binary configuration for N=9N=9, 1111, 1313 and 1515 coefficients.

Refer to caption
Figure 1: Normalized root-mean-square variation (41) of the surface gravity of each black hole, as a function of the number of points NN in each spatial dimension. Results for irreducible mass ratios qμμ1/μ2{1,1/2,1/10}q_{\mu}\equiv\mu_{1}/\mu_{2}\in\{1,1/2,1/10\} are shown. For each unequal-mass configuration, two curves are plotted, one for each black hole, labelled by the fractional masses μ1/μ\mu_{1}/\mu and μ2/μ\mu_{2}/\mu, with μμ1+μ2\mu\equiv\mu_{1}+\mu_{2} the total irreducible mass. Two different separations are displayed for qμ=1/10q_{\mu}=1/10.

The first test has to do with the zeroth law of binary black hole mechanics, being the fact that the horizon surface gravity (10) of each hole is constant. To measure this we define the normalized root-mean-square variation (or relative standard deviation) of the surface gravity κa\kappa_{a} of body aa by

σ¯(κa)κa2κa2/κa.\bar{\sigma}(\kappa_{a})\equiv\sqrt{\langle\kappa_{a}^{2}\rangle-\langle\kappa_{a}\rangle^{2}}/\langle\kappa_{a}\rangle\,. (41)

At each point one can compute κa\kappa_{a} by Eq. (40), and the various averages appearing in (41) are taken using all the collocation points located on the horizon of each black hole. Figure 1 shows σ¯(κa)\bar{\sigma}(\kappa_{a}) as a function of NN for one configuration with equal masses, one with a mass ratio qμ=1/2q_{\mu}=1/2 and two with qμ=1/10q_{\mu}=1/10. All these configurations correspond to separations close to the innermost circular orbit (ICO), namely the configuration of minimum binding energy; see Sec. III.3.2 below. For a mass ratio qμ=1/10q_{\mu}=1/10 we show an additional configuration with a smaller separation, corresponding to an orbital frequency μΩ=0.13\mu\Omega=0.13, where μμ1+μ2\mu\equiv\mu_{1}+\mu_{2} is the total irreducible mass. (In that case the ICO is located at μΩ=0.025\mu\Omega=0.025; see Tab. 1). All the curves exhibit the same behavior, being the convergence of σ¯(κa)\bar{\sigma}(\kappa_{a}) to a finite value as the resolution increases. It means that the error on the zeroth law is not coming from a numerical error, since that would depend on the resolution. Rather, it arises from the physical assumptions used when computing the binary configurations, such as our use of the CFC approximation, which implies that only a subset of Einstein’s equations are solved.

Another test is provided by the first integral relation (13) associated with the variational first law (12). For each configuration, we compute the global quantities MADMM_{\text{ADM}}, JJ and Ω\Omega, as well as the quasi-local quantities μa\mu_{a} and κa\langle\kappa_{a}\rangle. Figure 2 shows the violation of this algebraic relation, as measured by the difference

εMADM2ΩJa4μa2κa.\varepsilon\equiv M_{\text{ADM}}-2\Omega J-\sum_{a}4\mu_{a}^{2}\langle\kappa_{a}\rangle\,.\vskip-8.5359pt (42)

More precisely, Fig. 2 displays the variation of the dimensionless quantity ε/μ\varepsilon/\mu as a function of resolution, for the same configurations as the ones of Fig. 1. The data for qμ=1q_{\mu}=1 and 1/21/2 exhibit a similar behavior, where the error saturates. The convergence with resolution is slightly less clean than for σ¯(κa)\bar{\sigma}(\kappa_{a}), as the curves show a slight increase of the error for N=15N=15. This increase is probably due to the difficulty in computing very accurately the global quantities MADMM_{\text{ADM}} and JJ. At this stage it is unclear how the error (42) would behave at even higher resolution, which is unfortunately out of reach of our current computational resources. Nevertheless, from Fig. 2, it is clear that the deviation from the first integral does not converge to zero as the resolution increases, but instead to a small but finite value. From this point of view it is very similar to what is observed for the zeroth law.

Refer to caption
Figure 2: Error (42) on the first integral (13) as a function of the numerical resolution, still for irreducible mass ratios qμ{1,1/2,1/10}q_{\mu}\in\{1,1/2,1/10\}. Two different separations are displayed for qμ=1/10q_{\mu}=1/10.

The data for qμ=1/10q_{\mu}=1/10 at a separation close to the ICO (red triangles) exhibits a different behavior where no saturation is visible. The plausible explanation is that, in this case, the deviation from the first integral is so small that the error is still dominated by the effects of a finite resolution. This is consistent with results from Fig. 1, where the values of σ¯(κa)\bar{\sigma}(\kappa_{a}) for qμ=1/10q_{\mu}=1/10 at the ICO are significantly smaller that in all the other cases depicted. This explanation is also supported by data for a smaller separation (green diamonds). The black holes being closer, the errors induced by the CFC approximation are larger, as confirmed by Fig. 1. In that case, the error on the first law is no longer dominated by resolution effects, such that the curve shows a plateau.

III.3 Results

Unless stated otherwise, all the results shown below correspond to the highest resolution, namely N=15N=15 for irreducible mass ratios qμ=1q_{\mu}=1 and 1/21/2. For qμ=1/10q_{\mu}=1/10, most of the results were computed while using N=13N=13. A few configurations with N=15N=15 were also computed to confirm that they were almost indistinguishable from the N=13N=13 ones.

III.3.1 Surface gravity and first integral

The top panel of Fig. 3 shows the normalized horizon surface gravity 4μaκa4\mu_{a}\kappa_{a} of two corotating black holes with mass ratio qμ=1/2q_{\mu}=1/2, for a circular-orbit frequency μΩ=0.195\mu\Omega=0.195. The horizons are depicted as spheres of constant coordinate radius. The bottom panel of Fig. 3 displays the relative variations in surface gravity,

Δ(κa)κaκaκa,\Delta(\kappa_{a})\equiv\frac{\kappa_{a}-\langle\kappa_{a}\rangle}{\langle\kappa_{a}\rangle}\,, (43)

for the same configuration. Those reach at most a few percents, which shows that the zeroth law is satisfied to a high degree of accuracy, even for such a strong-field orbit. Indeed, for this configuration the ICO corresponds to an orbital frequency μΩICO=0.083\mu\Omega_{\text{ICO}}=0.083.

Refer to caption
Figure 3: Top panel: Normalized surface gravity 4μaκa4\mu_{a}\kappa_{a} of each black hole for a corotating binary with mass ratio qμ=μ1/μ2=1/2q_{\mu}=\mu_{1}/\mu_{2}=1/2 and a circular-orbit frequency μΩ=0.195\mu\Omega=0.195. Bottom panel: Relative variations Δ(κa)=(κaκa)/κa\Delta(\kappa_{a})=(\kappa_{a}-\langle\kappa_{a}\rangle)/\langle\kappa_{a}\rangle in horizon surface gravity for the same binary configuration.

Figure 4 illustrates the deviation from the zeroth law. More precisely, it shows the value of σ¯(κa)\bar{\sigma}(\kappa_{a}) as a function of the orbital frequency μΩ\mu\Omega, for mass ratios qμ{1,1/2,1/10}q_{\mu}\in\{1,1/2,1/10\}. The vertical gray lines show the location of the ICO in each case. For corotating orbits outside the ICO, the overall variation in surface gravity is always small, i.e., less than 0.3%0.3\%, 0.4%0.4\% and 0.5%0.5\% for mass ratios qμ=1q_{\mu}=1, 1/21/2 and 1/101/10, respectively. However, for highly relativistic orbits inside the ICO, those variations can reach several percents, especially for the heaviest body.

Moreover, the variations displayed in Fig. 4 are well above the numerical error level, and they show a monotonic increase as a function of frequency, for all configurations. This indicates that such variations are physical, in the sense that the zeroth law (10) is not exactly satisfied. As discussed in Sec. III.2, this is to be expected because of the CFC approximation implemented in our initial data, which prevents us from having exact, helically symmetric solutions to the full set of Einstein field equations.

Refer to caption
Figure 4: Normalized root-mean-square variation (41) of the horizon surface gravity of each black hole, as a function of the circular-orbit frequency μΩ\mu\Omega, for mass ratios qμ{1,1/2,1/10}q_{\mu}\in\{1,1/2,1/10\}. The gray vertical lines show the location of the innermost circular orbit for each mass ratio.

Figure 5 depicts the error ε/μ\varepsilon/\mu on the first integral, defined by (42), as a function of the orbital frequency μΩ\mu\Omega, for mass ratios qμ=1q_{\mu}=1 and 1/21/2. The first integral relation is satisfied at the 0.3%0.3\% and 0.1%0.1\% level at the ICO for qμ=1q_{\mu}=1 and 1/21/2, and the dimensionless differences (ε/μ)(μΩ)(\varepsilon/\mu)(\mu\Omega) show a trend similar to that observed in Fig. 4 for the normalized variations (41) in surface gravity. The data for qμ=1/10q_{\mu}=1/10 are too noisy to be displayed in Fig. 5, because in that case it gets challenging to extract the global quantities MADMM_{\text{ADM}} and JJ. The quantities (41) and (42) provide meaningful measures of the error (with respect to an exact solution of the full set of field equations) introduced by the CFC approximation.

Refer to caption
Figure 5: Violation (42) of the first integral relation (13), as a function of the orbital frequency μΩ\mu\Omega, for mass ratios qμ=1q_{\mu}=1 and qμ=1/2q_{\mu}=1/2.

III.3.2 Innermost circular orbit

The validity of the condition δAa=0\delta A_{a}=0, which we impose to compute our sequences of quasi-equilibrium initial data, is supported by the results from time evolutions of inspiralling black hole binaries. For instance, in the simplest case of nonspinning binaries, the total irreducible mass (defined as the sum of the irreducible masses of the apparent horizons) deviates from its initial value by only a few parts in 10610^{6} over the entire inspiral phase Bo.al2.07; Bu.al.12. This observed change is smaller than the numerical errors in the simulations, and analytical estimates Al.01; Po2.04; TaPo.08 suggest that the increase in irreducible mass due to tidal heating should be even smaller in this case. For equal-mass binaries with equal spins χ=+0.97\chi=+0.97, however, the variation in total irreducible mass is noticeable over the inspiral, but the relative changes only increase from 10610^{-6} to 1%1\% as the orbital velocity v(MΩ)1/3v\equiv(M\Omega)^{1/3} ramps up from 0.30.3 to 0.70.7 Lo.al.12. (Similar results hold for even larger spin values Sc.al.15.) Since our corotating binaries have much smaller (dimensionless) spins, no larger than 0.20.2 over our orbital frequency range, the approximation of constant irreducible masses is excellent.

Combining the condition δAa=0\delta A_{a}=0 with the first law of binary black hole mechanics (12) implies that, along a given sequence,

dEdΩ=ΩdJdΩ,\frac{\mathrm{d}E}{\mathrm{d}\Omega}=\Omega\,\frac{\mathrm{d}J}{\mathrm{d}\Omega}\,, (44)

where EMADM(μ1+μ2)E\equiv M_{\text{ADM}}-(\mu_{1}+\mu_{2}) is the binding energy. While in principle we could determine the functions E(Ω)E(\Omega) and J(Ω)J(\Omega) and compute their slopes, in practice Eq. (44) only provides a weak test of the first law because of the sparseness of our data, and because of the limited accuracy of the measurements of the global quantities MADMM_{\text{ADM}} and JJ for a given Ω\Omega.

Along a given sequence of circular orbits, the orbit for which the binding energy E(Ω)E(\Omega) is minimized is referred to as the minimum energy circular orbit (MECO), or more often as the innermost circular orbit (ICO). So according to (44) the ICO frequency ΩICO\Omega_{\text{ICO}} satisfies

dEdΩ|ICO=dJdΩ|ICO=0.\frac{\mathrm{d}E}{\mathrm{d}\Omega}\bigg{|}_{\text{ICO}}=\frac{\mathrm{d}J}{\mathrm{d}\Omega}\bigg{|}_{\text{ICO}}=0\,. (45)

The notion of ICO is closely related to that of innermost stable circular orbit (ISCO), which is defined by the vanishing of radial frequency associated with small eccentricity perturbations away from a given circular orbit. Indeed, for a test particle on a bound orbit around a Kerr black hole, the notions of ICO and ISCO are equivalent. More generally, when a meaningful conservative/dissipative split of the dynamics of a compact binary system can be defined,777This is known to be the case up to at least fourth order in the PN approximation Da.al.14; Be.al.16; Da.al.16; Be.al.17; Be.al2.17, as well as for linear black hole perturbations HiFl.08. if the conservative dynamics derives from a Hamiltonian, then the notions of ICO and ISCO are equivalent Bu.al.03; ViFl.15; Fu.al.17. When dissipative effects are included, however, the transition from inspiral to plunge does not occure at a particular, sharp frequency (the ICO frequency), but is replaced instead by a smooth transition over a frequency range OrTh.00; BuDa.00; BaSa.09. The notion of ICO has thus been used extensively as a useful reference point for comparisons between different analytical approximation methods and numerical techniques, e.g. in Da.al3.00; Pf.al.00; Gr.al.02; Bl.02; Da.al.02; CoPf.04; Ca.al2.06; Fa.11; Fa2.11.

Table 1 shows the numerical values of the circular-orbit frequency μΩ\mu\Omega, the reduced binding energy E/μE/\mu and the dimensionles angular momentum J/μ2J/\mu^{2} of the ICO for corotating binaries with mass ratios qμ{1,1/2,1/10}q_{\mu}\in\{1,1/2,1/10\}, as computed using the CFC and PN approximations. The PN predictions were computed from the 4PN-accurate expressions for the binding energy E(Ω)E(\Omega) and angular momentum J(Ω)J(\Omega) of binary systems of corotating points particles, as given by the formulas (69)–(70) in App. A. The accuracy on the CFC quantities is asserted from convergence studies for the same resolutions as in Figs. 1 and 2. Convergence gets slower as the mass ratio decreases. For qμ=1/10q_{\mu}=1/10 our highest available resolution is not high enough to extract the binding energy in a satisfactory manner. While for all mass ratios the agreement between the PN and NR predictions improves with increasing order, the PN series does not appear to converge precisely towards the CFC results, with systematic relative differences of a few percents, consistent with previous results (see for instance Ref. Da.al.02). This is most likely due to the fact that those two approximations are of a different nature.

     qμ=1q_{\mu}=1         qμ=1/2q_{\mu}=1/2         qμ=1/10q_{\mu}=1/10
μΩICO\mu\Omega_{\text{ICO}} EICO/μE_{\text{ICO}}/\mu JICO/μ2J_{\text{ICO}}/\mu^{2} μΩICO\mu\Omega_{\text{ICO}} EICO/μE_{\text{ICO}}/\mu JICO/μ2J_{\text{ICO}}/\mu^{2} μΩICO\mu\Omega_{\text{ICO}} EICO/μE_{\text{ICO}}/\mu JICO/μ2J_{\text{ICO}}/\mu^{2}
1PN 0.5220.522 0.04054-0.04054 0.62080.6208 0.5250.525 0.03614-0.03614 0.55100.5510 0.5370.537 0.01365-0.01365\;\, 0.20340.2034
2PN 0.08090.0809 0.01446-0.01446 0.88190.8819 0.06840.0684 0.01159-0.01159 0.81760.8176 0.02500.0250 0.002282-0.002282 0.40100.4010
3PN 0.09840.0984 0.01557-0.01557 0.85890.8589 0.07810.0781 0.01230-0.01230 0.80200.8020 0.02550.0255 0.002301-0.002301 0.39980.3998
4PN 0.09980.0998 0.01572-0.01572 0.85970.8597 0.07870.0787 0.01230-0.01230 0.80230.8023 0.02550.0255 0.002303-0.002303 0.39970.3997
CFC 0.1050.105 0.0164-0.0164 0.8430.843 0.0830.083 0.0126-0.0126 0.7890.789 0.0250.025 --- 0.4(5)0.4(5)
Table 1: Numerical values of the innermost circular orbit (ICO) frequency μΩICO\mu\Omega_{\text{ICO}}, reduced binding energy EICO/μE_{\text{ICO}}/\mu and dimensionless angular momentum JICO/μ2J_{\text{ICO}}/\mu^{2}, for corotating binaries with mass ratios qμ{1,1/2,1/10}q_{\mu}\in\{1,1/2,1/10\}, as computed using the CFC and PN approximations. The numerical data for qμ=1/10q_{\mu}=1/10 are difficult to compute due to a limited resolution (see also Sec. III.3.1).

IV Post-Newtonian theory

In this section we shall derive, in the context of the PN approximation, the 4PN-accurate expressions of the redshifts for binary systems of corotating spinning point masses. This requires the knowledge of the redshifts up to 4PN order for nonspinning systems (Sec. IV.1), that of the spin-orbit contributions up to next-to-leading order (Sec. IV.2), and the application of the corotation condition (25) to express the spins as functions of the circular-orbit frequency (Sec. IV.3), from which the renormalized redshifts can be derived (Sec. IV.4).

IV.1 Redshifts for nonspinning binaries

For nonspinning (NS) binaries, the relations zans(Ω)z^{\text{ns}}_{a}(\Omega) were computed up to 2PN and later to 3PN order in Refs. De.08; Bl.al.10. The leading and next-to-leading order logarithmic contributions at 4PN and 5PN orders, which originate from gravitational-wave tails, were then obtained in Refs. Bl.al2.10; Da.10; Le.al.12. Those results were all derived starting from the definition of the redshift in terms of the spacetime metric gαβ(x)g_{\alpha\beta}(x) evaluated at the coordinate location xα=yaαx^{\alpha}=y_{a}^{\alpha} of the particle aa, namely

za=dτadt=[gαβ(ya)vaαvaβ]1/2.z_{a}=\frac{\mathrm{d}\tau_{a}}{\mathrm{d}t}=\left[-g_{\alpha\beta}(y_{a})v_{a}^{\alpha}v_{a}^{\beta}\right]^{1/2}. (46)

Moreover, the nonlogarithmic 4PN terms were also computed, from the known 4PN contribution to the circular-orbit binding energy JaSc.13; BiDa.13 combined with the first law of mechanics (19) for binary systems of nonspinning point masses Le.al.12; Le.15; BlLe.17. All these results are valid for arbitrary mass ratios. Up to 4PN order, we have

z1ns\displaystyle z_{1}^{\text{ns}} =1+(3434Δ+ν2)x+(916916Δν218Δν+524ν2)x2\displaystyle=1+\left(-\frac{3}{4}-\frac{3}{4}\Delta+\frac{\nu}{2}\right)x+\left(-\frac{9}{16}-\frac{9}{16}\Delta-\frac{\nu}{2}-\frac{1}{8}\Delta\,\nu+\frac{5}{24}\nu^{2}\right)x^{2}
+(27322732Δν2+1916Δν3932ν2132Δν2+ν316)x3\displaystyle\qquad\!+\left(-\frac{27}{32}-\frac{27}{32}\Delta-\frac{\nu}{2}+\frac{19}{16}\Delta\,\nu-\frac{39}{32}\nu^{2}-\frac{1}{32}\Delta\,\nu^{2}+\frac{\nu^{3}}{16}\right)x^{3}
+(405256405256Δ+[3834164π2]ν+[68893844164π2]Δν\displaystyle\qquad\!+\left(-\,\frac{405}{256}-\frac{405}{256}\Delta+\left[\frac{38}{3}-\frac{41}{64}\pi^{2}\right]\nu+\left[\frac{6889}{384}-\frac{41}{64}\pi^{2}\right]\Delta\,\nu\right.
+[3863576+41192π2]ν293128Δν2+973864ν371728Δν3+9110368ν4)x4\displaystyle\qquad\qquad\!\!\left.+\left[-\frac{3863}{576}+\frac{41}{192}\pi^{2}\right]\nu^{2}-\frac{93}{128}\Delta\,\nu^{2}+\frac{973}{864}\nu^{3}-\frac{7}{1728}\Delta\,\nu^{3}+\frac{91}{10368}\nu^{4}\right)x^{4}
+(17015121701512Δ+[32915+12911024π2+645γE+325ln(16x)]ν\displaystyle\qquad\!+\left(-\,\frac{1701}{512}-\,\frac{1701}{512}\Delta+\left[-\frac{329}{15}+\frac{1291}{1024}\pi^{2}+\frac{64}{5}\gamma_{\text{E}}+\frac{32}{5}\ln{(16x)}\right]\nu\right.
+[246893840+12911024π2+645γE+325ln(16x)]Δν+[712071536+451256π2]Δν2\displaystyle\qquad\qquad\!\!+\left[-\frac{24689}{3840}+\frac{1291}{1024}\pi^{2}+\frac{64}{5}\gamma_{\text{E}}+\frac{32}{5}\ln{(16x)}\right]\Delta\,\nu+\left[-\frac{71207}{1536}+\frac{451}{256}\pi^{2}\right]\Delta\,\nu^{2}
+[101917923040+67033072π2+6415γE+3215ln(16x)]ν2+[356551691222551152π2]ν3\displaystyle\qquad\qquad\!\!+\left[-\frac{1019179}{23040}+\frac{6703}{3072}\pi^{2}+\frac{64}{15}\gamma_{\text{E}}+\frac{32}{15}\ln{(16x)}\right]\nu^{2}+\left[\frac{356551}{6912}-\frac{2255}{1152}\pi^{2}\right]\nu^{3}
+43576Δν3562141472ν4+5541472Δν418762208ν5)x5+o(x5),\displaystyle\qquad\qquad\!\!\left.+\,\,\frac{43}{576}\,\Delta\,\nu^{3}-\frac{5621}{41472}\,\nu^{4}+\frac{55}{41472}\,\Delta\,\nu^{4}-\frac{187}{62208}\,\nu^{5}\right)x^{5}+o(x^{5})\,, (47)

where x(mΩ)2/3x\equiv(m\Omega)^{2/3} is the small frequency-related PN parameter, with mm1+m2m\equiv m_{1}+m_{2} the total mass, while νm1m2/m2\nu\equiv m_{1}m_{2}/m^{2} is the symmetric mass ratio and Δ(m2m1)/m=14ν\Delta\equiv(m_{2}-m_{1})/m=\sqrt{1-4\nu} is the reduced mass difference. (We assume that m1m2m_{1}\leqslant m_{2}.) The expression for z2nsz_{2}^{\text{ns}} is easily deduced by setting ΔΔ\Delta\to-\Delta in Eq. (IV.1). Note the occurence of a logarithmic running and of Euler’s constant γE\gamma_{\text{E}} at 4PN order.

IV.2 Spin-orbit contributions to the redshifts and spin precession frequencies

Since our goal is to compute the redshifts for corotating spinning point-particle binaries, we need to account for the spin contributions to the redshifts. At the relative 4PN accuracy, however, it will prove sufficient to include the leading order 1.5PN and the next-to-leading order 2.5PN spin-orbit (SO) contributions. Up to next-to-leading order, the SO contributions to the redshift of body 1 read Bl.al.13

z1so\displaystyle z_{1}^{\text{so}} =[(13+Δ3+23ν)νχ1+(1+Δ176ν56Δν+23ν2)χ2]x5/2\displaystyle=\left[\left(-\frac{1}{3}+\frac{\Delta}{3}+\frac{2}{3}\nu\right)\nu\,\chi_{1}+\left(1+\Delta-\frac{17}{6}\nu-\frac{5}{6}\Delta\,\nu+\frac{2}{3}\nu^{2}\right)\chi_{2}\right]x^{5/2}
+[(12Δ2+1918ν1918Δνν29)νχ1+(32+32Δ173ν83Δν\displaystyle+\biggl{[}\left(\frac{1}{2}-\frac{\Delta}{2}+\frac{19}{18}\nu-\frac{19}{18}\Delta\,\nu-\frac{\nu^{2}}{9}\right)\nu\,\chi_{1}+\biggl{(}\frac{3}{2}+\frac{3}{2}\Delta-\frac{17}{3}\nu-\frac{8}{3}\Delta\,\nu
+17936ν2+4136Δν2ν39)χ2]x7/2+𝒪(x9/2),\displaystyle\qquad+\frac{179}{36}\nu^{2}+\frac{41}{36}\Delta\,\nu^{2}-\frac{\nu^{3}}{9}\biggr{)}\,\chi_{2}\biggr{]}\,x^{7/2}+\mathcal{O}(x^{9/2})\,, (48)

where χaSa/ma2\chi_{a}\equiv S_{a}/m_{a}^{2} are the dimensionless spins of the two compact objects. The expression for z2soz_{2}^{\text{so}} is easily deduced by setting ΔΔ\Delta\to-\Delta and χ1χ2\chi_{1}\leftrightarrow\chi_{2} in Eq. (IV.2). We will also need the expressions for the spin precession frequencies Ωa\Omega_{a} as functions of the circular-orbit frequency Ω\Omega. These were computed up to next-to-leading order in Bl.al.13, from the spin-orbit part of the binary canonical Hamiltonian, and up to next-to-next-to-leading order in Refs. Bo.al.13 and Do.al.14. For body 11, we have

Ω1\displaystyle\Omega_{1} =Ω{(34+34Δ+ν2)x+(916+916Δ+54ν58Δνν224)x2\displaystyle=\Omega\,\biggl{\{}\left(\frac{3}{4}+\frac{3}{4}\Delta+\frac{\nu}{2}\right)x+\left(\frac{9}{16}+\frac{9}{16}\Delta+\frac{5}{4}\nu-\frac{5}{8}\Delta\,\nu-\frac{\nu^{2}}{24}\right)x^{2}
+(2732+2732Δ+316ν398Δν10532ν2+532Δν2ν348)x3+𝒪(x4)}.\displaystyle\qquad\quad+\left(\frac{27}{32}+\frac{27}{32}\Delta+\frac{3}{16}\nu-\frac{39}{8}\Delta\,\nu-\frac{105}{32}\nu^{2}+\frac{5}{32}\Delta\,\nu^{2}-\frac{\nu^{3}}{48}\right)x^{3}+\mathcal{O}(x^{4})\biggr{\}}\,. (49)

Now, when reducing to corotating binaries, we will consider the sum za=zans+zaso+𝒪(χ2)z_{a}=z_{a}^{\text{ns}}+z_{a}^{\text{so}}+\mathcal{O}(\chi^{2}) and impose the corotation condition (25) to express the proper rotation frequencies ωa\omega_{a}, and then the dimensionless spins χa\chi_{a}, as functions of the circular-orbit frequency Ω\Omega. Importantly, while doing so, it is not necessary to account for spin-spin (or any higher order in the spins) contributions to the redshifts zaz_{a}. Indeed, spin-spin effects occure at the leading 2PN order, but since the spins themselves contribute at the leading 1.5PN order for corotating binaries [see Eq. (52) below], these spin-spin interactions are pushed to 5PN order, which is beyond the accuracy of our calculations.

Beware, however, that the leading spin-spin couplings in the Hamiltonian dynamics contribute terms at the relative 0.5PN order beyond the leading term in the expansion (IV.2) of the spin precession frequencies, which then contribute at the relative 2PN order for corotating binaries. Since the first law (21) has only been established to linear order in the spins, we do not include such quadratic-in-spin effects here. Including these would not impact our final result for the renormalized redshifts, Eq. (IV.4) below.

IV.3 Proper rotation frequencies and spins for corotating binaries

We now consider binary systems of corotating spinning point particles. Such systems are fully characterized by the component masses mam_{a}, or equivalently by the irreducible masses μa\mu_{a}, and the angular frequency Ω\Omega of the orbit. In particular, the redshifts zaz_{a}, the spins SaS_{a}, the spin precession frequencies Ωa\Omega_{a} and the proper rotation frequencies ωa\omega_{a} should all be functions of these three parameters. As discussed in Sec. II above, the condition for corotation is given by Eq. (25), which is equivalent to

ωa(Ω)=za1(Ω)[ΩΩa(Ω)],\omega_{a}(\Omega)=z^{-1}_{a}(\Omega)\left[\Omega-\Omega_{a}(\Omega)\right], (50)

where the redshifts zaz_{a} and spin precession frequencies Ωa\Omega_{a} are functions of the binary’s orbital frequency Ω\Omega (and the component masses mam_{a}) only. Substituting for Eqs. (IV.1) and (IV.2) into (50) yields the following 3PN-accurate expression for the proper rotation frequency of the largest body:

ω2=Ω{1νxν(32ν3)x2ν(118+2Δ5ν)x3+𝒪(x4)}.\omega_{2}=\Omega\left\{1-\nu\,x-\nu\left(\frac{3}{2}-\frac{\nu}{3}\right)x^{2}-\nu\left(\frac{11}{8}+2\Delta-5\nu\right)x^{3}+\mathcal{O}(x^{4})\right\}. (51)

Here, it is sufficient to use the expression (IV.1) for nonspinning bodies because the spin-orbit terms (IV.2) contribute at leading relative 4PN order in Eq. (51). Interestingly, even though z1z2z_{1}\neq z_{2} and Ω1Ω2\Omega_{1}\neq\Omega_{2}, we notice that ω1=ω2\omega_{1}=\omega_{2} up to relative 2PN order Bl.al.13. In the Newtonian limit x0x\to 0 and in the extreme mass-ratio limit ν0\nu\to 0, we recover ωa=Ω\omega_{a}=\Omega, in agreement with physical intuition. In Sec. VI.2, we will demonstrate how using the 3PN-accurate result (51) for equal-mass binaries (for which ν=14\nu=\tfrac{1}{4} and Δ=0\Delta=0), instead of the Newtonian result ωa=Ω\omega_{a}=\Omega, improves significantly the measurements of quasi-local spins performed in Ref. Ca.al2.06.

We may then compute the dimensionless spins χa=Sa/ma2\chi_{a}=S_{a}/m_{a}^{2} as functions of the circular-orbit frequency Ω\Omega by combining Eqs. (24a) and (24b), in which we substitute the formula (51) for the proper rotation frequencies. Up to next-to-leading order, which is sufficient for our purposes, we obtain

χa=4μaωa14(μaωa)2=2y3/2(1Γ){1ηy+𝒪(y2)},\chi_{a}=4\mu_{a}\omega_{a}\sqrt{1-4(\mu_{a}\omega_{a})^{2}}=2y^{3/2}\left(1\mp\Gamma\right)\left\{1-\eta\,y+\mathcal{O}(y^{2})\right\}, (52)

where y(μΩ)2/3y\equiv(\mu\Omega)^{2/3} is a frequency-related PN parameter, with the frequency adimensionalized using the total irreducible mass μμ1+μ2\mu\equiv\mu_{1}+\mu_{2}, while ημ1μ2/μ2\eta\equiv\mu_{1}\mu_{2}/\mu^{2} is the symmetric irreducible mass ratio and Γ(μ2μ1)/μ=14η\Gamma\equiv(\mu_{2}-\mu_{1})/\mu=\sqrt{1-4\eta} the irreducible mass difference. The parameters (μ,η,Γ,y)(\mu,\eta,\Gamma,y) are analogous to the parameters (m,ν,Δ,x)(m,\nu,\Delta,x) used above, but are defined in terms of the irreducible masses μa\mu_{a}, rather than the ordinary masses mam_{a}. For corotating systems, Eqs. (24a) and (51) can be combined to give

m\displaystyle m =μ{1+(26η)(12ηy)y3+η(6+643η10η2)y5+o(y5)},\displaystyle=\mu\left\{1+\left(2-6\eta\right)\left(1-2\eta\,y\right)y^{3}+\eta\biggl{(}-6+\frac{64}{3}\eta-10\eta^{2}\biggr{)}y^{5}+o(y^{5})\right\}, (53a)
x\displaystyle x =y{1+(434η)(12ηy)y3+o(y4)},\displaystyle=y\left\{1+\biggl{(}\frac{4}{3}-4\eta\biggr{)}\!\left(1-2\eta\,y\right)y^{3}+o(y^{4})\right\}, (53b)
ν\displaystyle\nu =η{1(28η)(12ηy)y3+o(y4)},\displaystyle=\eta\,\Bigl{\{}1-\left(2-8\eta\right)\left(1-2\eta\,y\right)y^{3}+o(y^{4})\Bigr{\}}\,, (53c)
Δ\displaystyle\Delta =Γ{1+4ηy38η2y4+o(y4)}.\displaystyle=\Gamma\,\Bigl{\{}1+4\eta\,y^{3}-8\eta^{2}y^{4}+o(y^{4})\Bigr{\}}\,. (53d)

As anticipated in Sec. IV.2 above, for corotating binaries the spins contribute at the leading 1.5PN order. The expressions (52) for the dimensionless spins χa\chi_{a} are to be substituted into the leading and next-to-leading SO contributions (IV.2) to the redshift, which then contribute at the 3PN and 4PN orders, respectively.

IV.4 Renormalized redshifts for corotating binaries

Next, to compute the renormalized redshifts cazac_{a}z_{a} for corotating spinning point particles, we also need to control the 4PN expansions of the response coefficients cac_{a} as functions of the circular-orbit frequency Ω\Omega. By substituting Eq. (51) into (24c), we obtain the following 4PN result for body 11:

c1\displaystyle c_{1} =1+(3+3Γ+6η)y3+η(66Γ12η)y4\displaystyle=1+\left(-3+3\Gamma+6\eta\right)y^{3}+\eta\left(6-6\Gamma-12\eta\right)y^{4}
+η(99Γ23η+5Γη+10η2)y5+o(y5).\displaystyle\qquad+\eta\left(9-9\Gamma-23\eta+5\Gamma\,\eta+10\eta^{2}\right)y^{5}+o(y^{5})\,. (54)

Notice the absence of contributions at 𝒪(y)\mathcal{O}(y) and at 𝒪(y2)\mathcal{O}(y^{2}). In the test-particle limit μ10\mu_{1}\to 0, Γ1\Gamma\to 1 and η0\eta\to 0, such that c1=4μ1κ¯1=1+𝒪(q2)c_{1}=4\mu_{1}\bar{\kappa}_{1}=1+\mathcal{O}(q^{2}) and c2=4μ2κ¯2=16y3+o(y5)+𝒪(q)c_{2}=4\mu_{2}\bar{\kappa}_{2}=1-6y^{3}+o(y^{5})+\mathcal{O}(q).

Finally, by adding the nonspinning and spin-orbit contributions to the redshift, Eqs. (IV.1) and (IV.2), and using the expression (52) for the corotating spins, we obtain after multiplying by Eq. (IV.4) and expanding in powers of the small PN parameter:

c1z1\displaystyle c_{1}z_{1} =1+(3434Γ+η2)y+(916916Γη218Γη+524η2)y2\displaystyle=1+\left(-\frac{3}{4}-\frac{3}{4}\Gamma+\frac{\eta}{2}\right)y+\left(-\frac{9}{16}-\frac{9}{16}\Gamma-\frac{\eta}{2}-\frac{1}{8}\Gamma\,\eta+\frac{5}{24}\eta^{2}\right)y^{2}
+(12332+6932Γ+112η+1916Γη3932η2132Γη2+η316)y3\displaystyle\qquad\!+\left(-\frac{123}{32}+\frac{69}{32}\Gamma+\frac{11}{2}\eta+\frac{19}{16}\Gamma\,\eta-\frac{39}{32}\eta^{2}-\frac{1}{32}\Gamma\,\eta^{2}+\frac{\eta^{3}}{16}\right)y^{3}
+(363256+363256Γ+[2334164π2]η+[11293844164π2]Γη\displaystyle\qquad\!+\left(\frac{363}{256}+\frac{363}{256}\Gamma+\left[\frac{23}{3}-\frac{41}{64}\pi^{2}\right]\eta+\left[\frac{1129}{384}-\frac{41}{64}\pi^{2}\right]\Gamma\,\eta\right.
+[983576+41192π2]η293128Γη2+973864η371728Γη3+9110368η4)y4\displaystyle\qquad\qquad\!\!\left.+\left[-\frac{983}{576}+\frac{41}{192}\pi^{2}\right]\eta^{2}-\frac{93}{128}\Gamma\,\eta^{2}+\frac{973}{864}\eta^{3}-\frac{7}{1728}\Gamma\,\eta^{3}+\frac{91}{10368}\eta^{4}\right)y^{4}
+(603512+603512Γ+[49415+12911024π2+645γE+325ln(16y)]η\displaystyle\qquad\!+\left(\frac{603}{512}+\frac{603}{512}\Gamma+\left[-\frac{494}{15}+\frac{1291}{1024}\pi^{2}+\frac{64}{5}\gamma_{\text{E}}+\frac{32}{5}\ln{(16y)}\right]\eta\right.
+[1475693840+12911024π2+645γE+325ln(16y)]Γη+[337671536+451256π2]Γη2\displaystyle\qquad\qquad\!\!+\left[-\frac{147569}{3840}+\frac{1291}{1024}\pi^{2}+\frac{64}{5}\gamma_{\text{E}}+\frac{32}{5}\ln{(16y)}\right]\Gamma\,\eta+\left[-\frac{33767}{1536}+\frac{451}{256}\pi^{2}\right]\Gamma\,\eta^{2}
+[70333923040+67033072π2+6415γE+3215ln(16y)]η2+[169351691222551152π2]η3\displaystyle\qquad\qquad\!\!+\left[-\frac{703339}{23040}+\frac{6703}{3072}\pi^{2}+\frac{64}{15}\gamma_{\text{E}}+\frac{32}{15}\ln{(16y)}\right]\eta^{2}+\left[\frac{169351}{6912}-\frac{2255}{1152}\pi^{2}\right]\eta^{3}
+43576Γη3562141472η4+5541472Γη418762208η5)y5+o(y5).\displaystyle\qquad\qquad\!\!\left.+\,\,\frac{43}{576}\,\Gamma\,\eta^{3}-\frac{5621}{41472}\,\eta^{4}+\frac{55}{41472}\,\Gamma\,\eta^{4}-\frac{187}{62208}\,\eta^{5}\right)y^{5}+o(y^{5})\,. (55)

The result for body 22, such that μ2μ1\mu_{2}\geqslant\mu_{1}, is straightforwardly deduced by setting ΓΓ\Gamma\to-\Gamma.

Anticipating a comparison to the perturbative results to be derived in Sec. V, we consider the PN expansion (IV.4), as well as that for body 22, in the limit qμμ1/μ21q_{\mu}\equiv\mu_{1}/\mu_{2}\ll 1 of large mass ratios. To do so, it is convenient to express the final results in terms of the frequency-related parameter u(μ2Ω)2/3u\equiv(\mu_{2}\Omega)^{2/3}, rather than y=(μΩ)2/3=u(1+qμ)2/3y=(\mu\Omega)^{2/3}=u\,(1+q_{\mu})^{2/3}. We find

c1z1\displaystyle c_{1}z_{1} =132u98u22716u3+363128u4+603256u5+qμ(uu2u3+[4634132π2]u4\displaystyle=1-\frac{3}{2}u-\frac{9}{8}u^{2}-\frac{27}{16}u^{3}+\frac{363}{128}u^{4}+\frac{603}{256}u^{5}+q_{\mu}\,\biggl{(}u-u^{2}-u^{3}+\biggl{[}\frac{46}{3}-\frac{41}{32}\pi^{2}\biggr{]}u^{4}
+[98815+1291512π2+1285γE+645ln(16u)]u5)+o(u5,qμ),\displaystyle\qquad+\biggl{[}-\frac{988}{15}+\frac{1291}{512}\pi^{2}+\frac{128}{5}\gamma_{\text{E}}+\frac{64}{5}\ln{(16u)}\biggr{]}u^{5}\biggr{)}+o(u^{5},q_{\mu})\,, (56a)
c2z2\displaystyle c_{2}z_{2} =16u3qμ(u+32u2+278u312116u41005128u5)+o(u5,qμ).\displaystyle=1-6u^{3}-q_{\mu}\left(u+\frac{3}{2}u^{2}+\frac{27}{8}u^{3}-\frac{121}{16}u^{4}-\frac{1005}{128}u^{5}\right)+o(u^{5},q_{\mu})\,. (56b)

V Black hole perturbation theory

In this section, we use the results of Ref. GrLe.13 to express the surface gravity of a spinning black hole perturbed by a corotating moon in terms of the circular-orbit frequency (Sec. V.1). We then use the result (32) in Sec. II.6, together with recent gravitational self-force results, to compute the surface gravity of the orbiting moon itself (Sec. V.2). Finally, we rely on the symmetry properties of the PN expression (IV.4) for the renormalized redshifts to introduce a physically-motivated “mass-ratio rescaling” of the perturbative expressions (Sec. V.3).

V.1 Surface gravity of a black hole with a corotating moon

The formulas (4a) and (16a) are expressed in terms of the spin χ\chi and angular velocity ω¯\bar{\omega} of the background Kerr black hole. Instead, we wish to express the surface gravity κ2κ=κ¯+𝒟κ+𝒪(q2)\kappa_{2}\equiv\kappa=\bar{\kappa}+\mathcal{D}\kappa+\mathcal{O}(q^{2}) of the perturbed event horizon as a function of the circular-orbit frequency Ω\Omega of the binary. To do so, we simply need to invert the expression for Ω=ω¯+𝒟ω(ω¯)+𝒪(q2)\Omega=\bar{\omega}+\mathcal{D}\omega(\bar{\omega})+\mathcal{O}(q^{2}), as given by Eq. (16b) with χ=4μω¯[14(μω¯)2]1/2\chi\!=\!4\mu\bar{\omega}[1-4(\mu\bar{\omega})^{2}]^{1/2} and v3=μω¯[14(μω¯)2]3/2+𝒪(q)v^{3}\!=\!\mu\bar{\omega}[1-4(\mu\bar{\omega})^{2}]^{-3/2}+\mathcal{O}(q). We readily find

ω¯=Ω{1qu(1+2u)14u31+u+𝒪(q2)},\bar{\omega}=\Omega\left\{1-q\,u\left(1+2u\right)\sqrt{\frac{1-4u^{3}}{1+u}}+\mathcal{O}(q^{2})\right\}, (57)

where we recall that u=(μΩ)2/3u=(\mu\Omega)^{2/3}, with μ=μ2\mu=\mu_{2} the irreducible mass of the background Kerr black hole.888Beware that in Sec. V.1 and V.2, the symbol μ\mu denotes the irreducible mass of the background Kerr black hole, i.e. μ=μ2\mu=\mu_{2}, and not the total irreducible mass like everywhere else. Importantly, the 3PN expansion of this perturbative result is in full agreement with the large mass-ratio limit of the 3PN result (51), namely

ω¯=ω2=Ω{1q(u+32u2+278u3)+𝒪(q2,u4)}.\bar{\omega}=\omega_{2}=\Omega\left\{1-q\,\biggl{(}u+\frac{3}{2}u^{2}+\frac{27}{8}u^{3}\biggr{)}+\mathcal{O}(q^{2},u^{4})\right\}. (58)

This provides significant support to both the post-Newtonian and the black hole perturbative calculations. Substituting for (57) into Eqs. (4a) and (16a), and expanding to first order in the mass ratio, we find

4μ2κ2=18u314u3qμu(1+2u+4u28u316u4)(1+u)(14u3)+𝒪(qμ2).4\mu_{2}\kappa_{2}=\frac{1-8u^{3}}{\sqrt{1-4u^{3}}}-q_{\mu}\,\frac{u\left(1+2u+4u^{2}-8u^{3}-16u^{4}\right)}{\sqrt{(1+u)(1-4u^{3})}}+\mathcal{O}(q_{\mu}^{2})\,. (59)

The expression for an isolated black hole, i.e. the first term in the right-hand side of Eq. (59), is simply given by (8) with ω=ω¯=Ω\omega=\bar{\omega}=\Omega at that order of approximation. Notice also that in the mass ratio qq we have substituted the Kerr black hole mass M=m2M=m_{2} by the irreducible mass μ=μ2\mu=\mu_{2} through

q=qμ14(μω¯)2=qμ14u3+𝒪(qμ2).q=q_{\mu}\sqrt{1-4(\mu\bar{\omega})^{2}}=q_{\mu}\sqrt{1-4u^{3}}+\mathcal{O}(q_{\mu}^{2})\,. (60)

As is clear from (59), the 𝒪(qμ)\mathcal{O}(q_{\mu}) correction to the horizon surface gravity of the background Kerr black hole is negative. In the weak-field limit u0u\to 0, the perturbative result (59) gives

4μ2κ2=16u3qμ(u+32u2+278u312116u41005128u5)+𝒪(u6,qμ2),4\mu_{2}\kappa_{2}=1-6u^{3}-q_{\mu}\left(u+\frac{3}{2}u^{2}+\frac{27}{8}u^{3}-\frac{121}{16}u^{4}-\frac{1005}{128}u^{5}\right)+\mathcal{O}(u^{6},q_{\mu}^{2})\,, (61)

in full agreement with the large mass-ratio limit (56b) of the 4PN result (IV.4). This check of consistency gives additional support to the analogy (28) between point particle redshift and black hole horizon surface gravity.

V.2 Surface gravity of the corotating moon

We are now going to derive an explicit expression for the normalized surface gravity 4μ1κ14\mu_{1}\kappa_{1} of the orbiting moon (the lightest black hole), to linear order in the mass ratio. In Sec. II.6 we mentionned a proof, given in Ref. Po3.15, that the normalized surface gravity 4μ1κ14\mu_{1}\kappa_{1} is in fact equal to the redshift zz1z\equiv z_{1} of a massive point particle undergoing a gravitational self-force effect, that follows a geodesic motion in a certain smooth effective metric. Hence, Eq. (32) implies

4μ1κ1=z=z(0)(Ω,χ)+qz(1)(Ω,χ)+𝒪(q2),4\mu_{1}\kappa_{1}=z=z_{(0)}(\Omega,\chi)+q\,z_{(1)}(\Omega,\chi)+\mathcal{O}(q^{2})\,, (62)

where z(0)=[13v2+2χv3]1/2/(1+χv3)z_{(0)}=[1-3v^{2}+2\chi v^{3}]^{1/2}/(1+\chi v^{3}) is the redshift of a test mass on a circular equatorial orbit in a Kerr background of dimensionless spin χ\chi, while qz(1)(Ω,χ)q\,z_{(1)}(\Omega,\chi) is the (conservative) gravitational self-force correction at a given circular-orbit frequency Ω\Omega. Substituting for the expressions χ=4μω¯[14(μω¯)2]1/2\chi=4\mu\bar{\omega}[1-4(\mu\bar{\omega})^{2}]^{1/2} and v3=MΩ/(1χMΩ)=μΩ/([14(μω¯)2]1/2χμΩ)v^{3}=M\Omega/(1-\chi M\Omega)=\mu\Omega/([1-4(\mu\bar{\omega})^{2}]^{1/2}-\chi\mu\Omega), and specifying to the corotating case by imposing Eq. (57), we readily obtain

4μ1κ1=(12u)(1+u)(14u3)+q{z(1)(u)4u5(1+2u)21+u}+𝒪(q2).4\mu_{1}\kappa_{1}=(1-2u)\sqrt{(1+u)(1-4u^{3})}+q\left\{z_{(1)}(u)-4u^{5}\frac{(1+2u)^{2}}{1+u}\right\}+\mathcal{O}(q^{2})\,. (63)

On the other hand, numerical data for z(1)(Ω;χ)z_{(1)}(\Omega;\chi) has been computed in Refs. Sh.al.12; vdMSh.15 for a sample of background Kerr black hole spins and circular-orbit frequencies. However, those data are too sparse for our purposes. Instead, we shall use the recent analytical gravitational self-force results of Ka.al.16, which provide the 8.5PN expansion of z(1)(Ω;χ)z_{(1)}(\Omega;\chi) for any Kerr black hole spin 0χ<10\leqslant\chi<1. Specifying to the corotating case, i.e., substituting χ=4u3/2[14u3]1/2\chi=4u^{3/2}[1-4u^{3}]^{1/2} in their Eqs. (4.2)–(4.3), as well as M=μ[14u3]1/2M=\mu[1-4u^{3}]^{-1/2}, we find999Actually, the authors of Ref. Ka.al.16 computed the 8.5PN expansion of the gravitational self-force contribution to the quantity U1/zU\equiv 1/z, which they denoted ΔU\Delta U. We simply have z(1)=z(0)2ΔUz_{(1)}=-z_{(0)}^{2}\Delta U.

z(1)\displaystyle z_{(1)} (u)=uu2u3+(5234132π2)u4+(95815+1291512π2+1285γE+645ln(16u))u5\displaystyle(u)=u-u^{2}-u^{3}+\biggl{(}\frac{52}{3}-\frac{41}{32}\pi^{2}\biggr{)}u^{4}+\biggl{(}-\frac{958}{15}+\frac{1291}{512}\pi^{2}+\frac{128}{5}\gamma_{\text{E}}+\frac{64}{5}\ln{(16u)}\biggr{)}u^{5}
+(25877873150+1267791536π29976105γE2367215ln2+2437ln34988105lnu)u6+o(u6).\displaystyle+\biggl{(}-\frac{2587787}{3150}+\frac{126779}{1536}\pi^{2}-\frac{9976}{105}\gamma_{\text{E}}-\frac{23672}{15}\ln{2}+\frac{243}{7}\ln{3}-\frac{4988}{105}\ln{u}\biggr{)}u^{6}+o(u^{6})\,. (64)

To avoid an exceedingly lengthy expression, we only display here the result up to 5PN order. The analytical values of the higher order coefficients are given in Eqs. (74)–(75) of App. B, up to 8.5PN order. In Sec. VI below, we shall use this 8.5PN-accurate expansion to evaluate the normalized surface gravity (63) of the smaller black hole.

Importantly, we can verify that the 4PN expansion of the perturbative result (63)–(V.2), together with Eq. (60), is in full agreement with the perturbative expansion (56a) of the 4PN result (IV.4), namely

4μ1\displaystyle 4\mu_{1} κ1=132u98u22716u3+363128u4+603256u5\displaystyle\kappa_{1}=1-\frac{3}{2}u-\frac{9}{8}u^{2}-\frac{27}{16}u^{3}+\frac{363}{128}u^{4}+\frac{603}{256}u^{5}
+qμ[uu2u3+(4634132π2)u4+(98815+1291512π2+1285γE+645ln(16u))u5]\displaystyle+q_{\mu}\,\biggl{[}u-u^{2}-u^{3}+\biggl{(}\frac{46}{3}-\frac{41}{32}\pi^{2}\biggr{)}u^{4}+\biggl{(}-\frac{988}{15}+\frac{1291}{512}\pi^{2}+\frac{128}{5}\gamma_{\text{E}}+\frac{64}{5}\ln{(16u)}\biggr{)}u^{5}\biggr{]}
+o(u5,qμ).\displaystyle+o(u^{5},q_{\mu})\,. (65)

V.3 Rescaling of the perturbative results

The 4PN formula (IV.4) is naturally expressed in terms of the PN parameter y=(μΩ)2/3y=(\mu\Omega)^{2/3}, where the orbital frequency is normalized by the total irreducible mass μ=μ1+μ2\mu=\mu_{1}+\mu_{2}. Hence, to compare our perturbative formulas (59) and (63) to the 4PN and NR results, we first use the relation u=y(1+qμ)2/3u=y\,(1+q_{\mu})^{-2/3}, expand to linear order in the irreducible mass ratio qμq_{\mu}, and obtain

4μ1κ1\displaystyle 4\mu_{1}\kappa_{1} =(12y)(1+y)(14y3)+𝒪(qμ2)\displaystyle=(1-2y)\sqrt{(1+y)(1-4y^{3})}+\mathcal{O}(q_{\mu}^{2})
+qμ[y(1+2y+4y28y316y4)(1+y)(14y3)4y514y3(1+2y)21+y+14y3z(1)(y)],\displaystyle\quad+q_{\mu}\biggl{[}\frac{y\,(1+2y+4y^{2}-8y^{3}-16y^{4})}{\sqrt{(1+y)(1-4y^{3})}}-4y^{5}\sqrt{1-4y^{3}}\,\frac{(1+2y)^{2}}{1+y}+\sqrt{1-4y^{3}}\,z_{(1)}(y)\biggr{]}\,, (66a)
4μ2κ2\displaystyle 4\mu_{2}\kappa_{2} =18y314y3qμ[y(1+2y+4y28y316y4)(1+y)(14y3)4y3(38y3)(14y3)3/2]+𝒪(qμ2).\displaystyle=\frac{1-8y^{3}}{\sqrt{1-4y^{3}}}-q_{\mu}\biggl{[}\frac{y\,(1+2y+4y^{2}-8y^{3}-16y^{4})}{\sqrt{(1+y)(1-4y^{3})}}-\frac{4y^{3}\left(3-8y^{3}\right)}{(1-4y^{3})^{3/2}}\biggr{]}+\mathcal{O}(q_{\mu}^{2})\,. (66b)

Remarkably, we observe that the first terms in the 𝒪(qμ)\mathcal{O}(q_{\mu}) contributions therein are identical. While the former comes from the substitution uyu\to y in the 𝒪(qμ0)\mathcal{O}(q_{\mu}^{0}) contribution to (66a), the latter comes from combining the same substitution in the 𝒪(qμ0)\mathcal{O}(q_{\mu}^{0}) contribution to (66b) with the 𝒪(q)\mathcal{O}(q) correction to the horizon angular velocity (57).

Motivated by a large body of work FiDe.84; An.al.95; Fa.al.04; Sp.al2.11; Le.al.11; Le.al2.12; Le.al.13; Na.13; vdM.17 suggesting that the scaling qμη=qμ/(1+qμ)2q_{\mu}\to\eta=q_{\mu}/(1+q_{\mu})^{2} considerably extends the domain of validity of black hole perturbative calculations, we wish to restore in the expansions (66) some of the missing information regarding the symmetry property by exchange 121\leftrightarrow 2 of the two bodies. As argued in App. A, the general form (73) of the renormalized redshifts cazac_{a}z_{a}, together with the analogy (28), suggest that the surface gravity of each black hole should naturally be written in the form

4μaκa=a(y)±b(y)Γ+c(y)η±d(y)Γη+𝒪(η2).4\mu_{a}\kappa_{a}=a(y)\pm b(y)\,\Gamma+c(y)\,\eta\pm d(y)\,\Gamma\eta+\mathcal{O}(\eta^{2})\,. (67)

The key point is that the a priori unknown functions a(y)a(y), b(y)b(y), c(y)c(y) and d(y)d(y) can be uniquely determined by expanding Eqs. (67) up to linear in the mass ratio, using Γ=12qμ+𝒪(qμ2)\Gamma=1-2q_{\mu}+\mathcal{O}(q_{\mu}^{2}), and equating the resulting expressions to the formulas (66). Doing so, we readily obtain

a(y)\displaystyle a(y) =12(12y)(1+y)(14y3)+18y3214y3,\displaystyle=\frac{1}{2}(1-2y)\sqrt{(1+y)(1-4y^{3})}+\frac{1-8y^{3}}{2\sqrt{1-4y^{3}}}\,, (68a)
b(y)\displaystyle b(y) =12(12y)(1+y)(14y3)18y3214y3,\displaystyle=\frac{1}{2}(1-2y)\sqrt{(1+y)(1-4y^{3})}-\frac{1-8y^{3}}{2\sqrt{1-4y^{3}}}\,, (68b)
c(y)\displaystyle c(y) =14y3{z(1)(y)22y5(1+2y)21+y}+2y3(38y3)(14y3)3/2,\displaystyle=\sqrt{1-4y^{3}}\,\biggl{\{}\frac{z_{(1)}(y)}{2}-2y^{5}\,\frac{(1+2y)^{2}}{1+y}\biggr{\}}+\frac{2y^{3}(3-8y^{3})}{{(1-4y^{3})}^{3/2}}\,, (68c)
d(y)\displaystyle d(y) =14y3{z(1)(y)22y5(1+2y)21+y}+14y48y5(1+y)(14y3)16y3+16y6(14y3)3/2.\displaystyle=\sqrt{1-4y^{3}}\,\biggl{\{}\frac{z_{(1)}(y)}{2}-2y^{5}\,\frac{(1+2y)^{2}}{1+y}\biggr{\}}+\frac{1-4y^{4}-8y^{5}}{\sqrt{(1+y)(1-4y^{3})}}-\frac{1-6y^{3}+16y^{6}}{{(1-4y^{3})}^{3/2}}\,. (68d)

In particular, the coefficients aa and bb are simply given by the half-sum and the half-difference of the 𝒪(qμ0)\mathcal{O}(q_{\mu}^{0}) contributions to Eq. (66), respectively, while the coefficient cc is given by the half-sum of the 𝒪(qμ)\mathcal{O}(q_{\mu}) contributions. Expanding (67)–(68) in powers of qμq_{\mu}, one recovers the original expressions (66), by construction. One the other hand, expanding in powers of yy, one recovers the 4PN formula (IV.4), up to uncontrolled terms 𝒪(η2)\mathcal{O}(\eta^{2}) and higher. For equal-mass binaries, Γ=0\Gamma=0 and the expressions (67)–(68) for 4μ1κ14\mu_{1}\kappa_{1} and 4μ2κ24\mu_{2}\kappa_{2} coincide.

VI Comparisons

In this section, we shall compare the predictions from our numerical relativity initial data, the post-Newtonian approximation, and black hole perturbation theory for corotating black hole binaries. In Sec. VI.1 we compare the computed surface gravities for irreducible mass ratios qμ{1,1/2,1/10}q_{\mu}\in\{1,1/2,1/10\}. We then revisit the problem of diagnostics of spin measurements in quasi-equilibrium initial data of corotating black hole binaries in Sec. VI.2. Throughout this section, we use the total irreducible mass μ=μ1+μ2\mu=\mu_{1}+\mu_{2} to adimensionalize all frequencies.

VI.1 Surface gravity

Motivated by the analogy pointed out in Refs. Le.al.12; Bl.al.13 between black hole surface gravity and point particle redshift, the authors of Ref. Zi.al.16 introduced a concept of “redshift” for a dynamical black hole, from a properly normalized and averaged notion of surface gravity, in the context of NR simulations of nonspinning black hole binaries. They compared their numerical results for these “black hole redshifts” to the known 3PN prediction Bl.al.10 for the redshifts of point particle binaries for circular orbits, and found excellent agreement during the entire inspiral, with relative differences 𝒪(103)\mathcal{O}(10^{-3}).

In this work, we adopt a strategy that, in a sense, is the opposite to that used in Ref. Zi.al.16. In Sec. III we have computed the surface gravity 4μaκa4\mu_{a}\kappa_{a} of each corotating black hole. Then, motivated by the analogy (28), we are going to compare those numerical results to the 4PN-accurate prediction (IV.4) for the redshift cazac_{a}z_{a} of each corotating point particle. Moreover, we will compare our nonlinear result for the surface gravity of each black hole to the predictions (66) and (67)–(68) from linear black hole perturbation theory.

VI.1.1 Mass ratio 1:1

We start with the simplest case of equal-mass binaries, for which κ1=κ2\kappa_{1}=\kappa_{2} by symmetry. Figure 6 shows the normalized horizon surface gravity 4μaκa4\mu_{a}\kappa_{a} of both black holes as a function of the orbital frequency μΩ\mu\Omega, as computed in Sec. III using the CFC approximation, and in Sec. IV using the PN approximation, up to 4PN order. More precisely, the data points with error bars display 4μaκa[1±σ¯(κa)]4\mu_{a}\langle\kappa_{a}\rangle[1\pm\bar{\sigma}(\kappa_{a})], and the various colored curves show the successive PN approximations to cazac_{a}z_{a}, up to 4PN order, computed from (IV.4) with η=14\eta=\frac{1}{4} and Γ=0\Gamma=0. The vertical gray line shows the location of the ICO for this particular configuration, at μΩICO0.105\mu\Omega_{\text{ICO}}\simeq 0.105, as computed from the binding energy E(Ω)E(\Omega) of the sequence of quasi-equilibrium initial data (see Tab. 1).

The NR data shows that the surface gravity decreases fairly steeply with increasing orbital frequency, as the effect of the tidal field of the companion gets stronger. In the Newtonian limit where μΩ0\mu\Omega\to 0, we recover the expected behavior 4μaκa14\mu_{a}\kappa_{a}\to 1, as appropriate for isolated nonspinning black holes. While the Newtonian and 1PN approximations to 4μaκa4\mu_{a}\kappa_{a} compare quite poorly to the NR results, except at low frequencies (i.e., at large orbital separations), the 2PN, 3PN and 4PN approximations perform very well, even for highly relativistic orbits beyond the ICO. In particular, the 3PN and 4PN predictions—which cannot be distinguished on the graph—are both within the NR error bars over the entire orbital-frequency range for which we have data.

Refer to caption
Figure 6: Normalized horizon surface gravity 4μaκa4\mu_{a}\kappa_{a} in equal-mass corotating binaries, as a function of the circular-orbit frequency μΩ\mu\Omega, computed using the CFC and the PN approximations. The data points with error bars display 4μaκa[1±σ¯(κa)]4\mu_{a}\langle\kappa_{a}\rangle[1\pm\bar{\sigma}(\kappa_{a})]. The various colored curves show the successive PN approximations to cazac_{a}z_{a}, up to 4PN order, computed from Eq. (IV.4) with η=14\eta=\frac{1}{4} and Γ=0\Gamma=0. The vertical gray line displays the location of the innermost circular orbit, at μΩICO0.105\mu\Omega_{\text{ICO}}\simeq 0.105.

VI.1.2 Mass ratios 2:1 and 10:1

Next, we consider the case of a binary with mass ratio qμ=1/2q_{\mu}=1/2, for which μΩICO0.083\mu\Omega_{\text{ICO}}\simeq 0.083. Figure 7 displays the NR results for each black hole, which can be distinguished by the size of the filled black circle in the figure’s legend. First, we notice that the (normalized) surface gravity of the heaviest black hole is larger than that of the lightest black hole. This indicates that the tidal field of the largest body has a bigger effect on the smaller body than the other way around, in agreement with physical intuition. Figure 7 also shows the 3PN and 4PN predictions for each black hole. We observe a slight improvement from 3PN to 4PN order, for both black holes. Overall, the agreement with the NR results is very good, all the way to the ICO, and even beyond. The relative differences reach 2%2\% for large orbital frequencies, say at μΩ0.22.4μΩICO\mu\Omega\sim 0.2\sim 2.4\,\mu\Omega_{\text{ICO}}, corresponding to an orbital separation rΩμ/(μΩ)2/33μr_{\Omega}\equiv\mu/(\mu\Omega)^{2/3}\sim 3\mu.

Our third and last case corresponds to a mass ratio qμ=1/10q_{\mu}=1/10, for which μΩICO0.025\mu\Omega_{\text{ICO}}\simeq 0.025. Figure 8 shows the NR results, as well as the analytic predictions from the PN approximation and from linear black hole perturbation theory. Just like for the cases qμ=1q_{\mu}=1 and 1/21/2, the 4PN predictions perform very well over the full range of circular-orbit frequencies considered. More remarkably, the predictions from linear black hole perturbation theory are in excellent agreement with the NR data, despite the moderate mass ratio, all the way to separations rΩ3μr_{\Omega}\sim 3\mu (i.e. well inside the ICO), for both black holes. In fact, the perturbative prediction for the smaller body is in better agreement with the NR data than the 4PN prediction.

Refer to caption
Figure 7: Same as in Fig. 6 but for a mass ratio 2:12:1, for which μΩICO0.083\mu\Omega_{\text{ICO}}\simeq 0.083. The black holes can be distinguished by the size of the filled black circles in the legend. By convention μ1μ2\mu_{1}\leqslant\mu_{2} so black hole 1 (BH1) is lighter than black hole 2 (BH2). For all but the rightmost pale-blue data point the CFC error bars are smaller than the size of the point symbol.
Refer to caption
Figure 8: Same as in Fig 7 but for a mass ratio 10:110:1, for which μΩICO0.025\mu\Omega_{\text{ICO}}\simeq 0.025. The dashed curves labelled “BHP” show the predictions from linear black hole perturbation theory, as given in (66). For the lighter black hole (BH1), the error bars on the CFC data points are all smaller than the size of the point symbol.

VI.1.3 Mass ratio rescaling

Finally, in Fig. 9 we collect all the NR results discussed previously, together with the predictions from black hole perturbation theory in each case, following the mass-ratio rescaling introduced in Sec. V.3. Remarkably, the rescaled perturbative predictions (67)–(68) are in excellent agreement with the NR data in all cases, for both black holes. The agreement in the equal-mass case (black curve and symbols) is especially stricking. Indeed, the rescaled linear perturbative predictions are compatible with the nonlinear results (agreement within the NR error bars) over most of the frequency range. The largest disagreement occures for the smallest black hole, for a mass ratio qμ=1/10q_{\mu}=1/10, for which the relative difference grows up to 3%\sim 3\% at rΩ3μr_{\Omega}\sim 3\mu. Restricting ourselves to physically relevant orbits outside the ICO, the agreement reaches the 0.3%0.3\% level or better for all mass ratios considered.

Refer to caption
Figure 9: Same as in Figs. 6, 7 and 8, all binary configurations combined. The solid curves show the predictions from linear black hole perturbation theory, following the mass-ratio rescaling discussed in Sec. V.3, as given in Eqs. (67)–(68). The vertical gray lines show the location of the innermost circular orbit for each irreducible mass ratio μ2/μ1\mu_{2}/\mu_{1}.

This illustrates, once again, that the domain of validity of black hole perturbative calculations appears to extend well beyond the extreme mass-ratio limit. As argued in Ref. Le2.14, this opens the exciting prospect of using black hole perturbation theory and the gravitational self-force framework to model the gravitational-wave emission from intermediate mass-ratio inspirals, or even compact binaries with comparable masses.

This observation is also supported by recent works showing that (i) the Einstein equation is, in a sense, not so nonlinear Ha2.14 and that (ii) the relativistic gravitational dynamics of an arbitrary-mass-ratio two-spinning-black-hole system is equivalent, at first post-Minkowskian order, to that of a spinning test black hole in a background Kerr spacetime Da.16; Vi.18.

Refer to caption
Figure 10: Relative differences ΔS/S\Delta S/S between several quasi-local measurements of spins for equal-mass corotating black hole binaries, as functions of the circular-orbit frequency μΩ\mu\Omega. The 2PN and 3PN curves are almost on top of each other.

VI.2 Quasi-local spin

In binary black hole spacetimes, there is no canonical definition of spin for each black hole. However, several quasi-local notions of spins have been introduced in the litterature BrYo.93; As.al.01; CoWh.07 (see Sz.04; AsKr.04 for reviews). In particular, the authors of Ref. Ca.al2.06 used two methods to estimate the spins of equal-mass, corotating black holes binaries: one based on an approximate axial Killing vector, SKS_{K}, and one based on a flat space Killing vector constructed from Cartesian-like coordinates, SzS_{z}. In the case of a sequence of quasi-equilibrium initial data for equal-mass, corotating black hole binaries, those spin diagnostics are listed in Table IV of Ref. Ca.al2.06, for a particular choice of lapse boundary condition on the excision surfaces. The relative difference between SKS_{K} and SzS_{z} remains below 0.6%0.6\% over the orbital frequency range 0.01μΩ0.110.01\lesssim\mu\Omega\lesssim 0.11.

On the other hand, the discussion in Sec. II.5 suggests that each black hole in a corotating binary can, to some extent, be modelled as an isolated hole immersed in a tidal environment. Then, a third natural spin measurement is provided by the Kerr-inspired formula (24b), in which the proper rotation frequency ωa\omega_{a} must be expressed as a function of the circular-orbit frequency Ω\Omega of the binary. Such a relationship was precisely determined, up to 3PN order, in Eq. (51). We denote by SS the spin measurement obtained by inserting Eq. (51) into the formula (24b), without performing any subsequent PN expansion.

Figure 10 shows the relative differences ΔS/SSK/S1\Delta S/S\equiv S_{K}/S-1 and Sz/S1S_{z}/S-1 while using the Newtonian (labelled 0PN), 1PN, 2PN and 3PN approximations for the black hole’s proper rotation frequencies (51) in the expression (24b) for the spin SS of a black hole of irreducible mass μa\mu_{a} and proper rotation frequency ωa\omega_{a}. This figure updates Fig. 4 of Ref. Ca.al2.06. We make the three following observations:

  • At a given PN order, the SKS_{K} spin diagnostic is always in better agreement with the SS spin diagnostic than the SzS_{z} spin diagnostic;

  • Increasing the PN order systematically reduces the disagreements between the multiple spin diagnostics, from up to 8%8\% at Newtonian order to less than 0.3%0.3\% at 3PN order;

  • Given the small numerical value of the 3PN coefficient for equal masses in Eq. (51), the difference between the 2PN and 3PN-accurate predictions for the SS spin diagnostic is small over the orbital frequency range considered, such that the 2PN and 3PN curves are almost on top of each other.

In particular, this suggests that the corotation condition (25) for spinning point masses captures most of the physics of corotating black hole binaries, which may (in part) explain the remarkable agreement between the NR and PN results for the horizon surface gravity, as discussed in Sec. VI.1. That being said, despite a trend that shows better agreement at increasingly higher PN orders, one should not expect such comparisons to reach an arbitrarily high level of agreement, because each of those spins diagnostics comes with caveats.

Acknowledgements.
ALT acknowledges financial support through a Marie Curie FP7 Integration Grant within the 7th European Union Framework Programme (PCIG13-GA-2013-630210).

Appendix A Energy and angular momentum of corotating binaries

In this appendix, we give the expressions for the binding energy EE and angular momentum JJ of corotating compact binaries, as functions of the circular-orbit frequency Ω\Omega, up to 4PN order. First, recall that for nonspinning binaries, the 4PN-accurate expressions of the binding energy and angular momentum are known and read Da.al.14; Be.al.17

Ens\displaystyle\!\!\!E^{\text{ns}} =μηy2{1+(34η12)y+(278+198ηη224)y2\displaystyle=-\frac{\mu\eta\,y}{2}\biggl{\{}1+\left(-\frac{3}{4}-\frac{\eta}{12}\right)y+\left(-\frac{27}{8}+\frac{19}{8}\eta-\frac{\eta^{2}}{24}\right)y^{2}
+(67564+[3444557620596π2]η15596η2355184η3)y3\displaystyle\qquad\qquad\quad\,\;+\left(-\frac{675}{64}+\biggl{[}\frac{34445}{576}-\frac{205}{96}\pi^{2}\biggr{]}\eta-\frac{155}{96}\eta^{2}-\frac{35}{5184}\eta^{3}\right)y^{3}
+(3969128+[1236715760+90371536π2+89615γE+44815ln(16y)]η\displaystyle\qquad\qquad\quad\,\;+\left(-\frac{3969}{128}+\left[-\frac{123671}{5760}+\frac{9037}{1536}\pi^{2}+\frac{896}{15}\gamma_{\text{E}}+\frac{448}{15}\ln(16y)\right]\eta\right.
+[4984493456+3157576π2]η2+3011728η3+7731104η4)y4+o(y4)},\displaystyle\left.\qquad\qquad\qquad\quad+\left[-\frac{498449}{3456}+\frac{3157}{576}\pi^{2}\right]\eta^{2}+\frac{301}{1728}\eta^{3}+\frac{77}{31104}\eta^{4}\right)y^{4}+o(y^{4})\biggr{\}}\,, (69a)
ΩJns\displaystyle\!\!\!\Omega J^{\text{ns}} =μηy{1+(32+η6)y+(278198η+η224)y2\displaystyle=\mu\eta\,y\,\biggl{\{}1+\left(\frac{3}{2}+\frac{\eta}{6}\right)y+\left(\frac{27}{8}-\frac{19}{8}\eta+\frac{\eta^{2}}{24}\right)y^{2}
+(13516+[6889144+4124π2]η+3124η2+71296η3)y3\displaystyle\qquad\qquad\;\;+\left(\frac{135}{16}+\biggl{[}-\frac{6889}{144}+\frac{41}{24}\pi^{2}\biggr{]}\eta+\frac{31}{24}\eta^{2}+\frac{7}{1296}\eta^{3}\right)y^{3}
+(2835128+[98869576064551536π21283γE643ln(16y)]η\displaystyle\qquad\qquad\;\;+\left(\frac{2835}{128}+\left[\frac{98869}{5760}-\frac{6455}{1536}\pi^{2}-\frac{128}{3}\gamma_{\text{E}}-\frac{64}{3}\ln(16y)\right]\eta\right.
+[35603534562255576π2]η22151728η35531104η4)y4+o(y4)},\displaystyle\left.\qquad\qquad\qquad\;+\left[\frac{356035}{3456}-\frac{2255}{576}\pi^{2}\right]\eta^{2}-\frac{215}{1728}\eta^{3}-\frac{55}{31104}\eta^{4}\right)y^{4}+o(y^{4})\biggr{\}}\,, (69b)

where we recall that y=(μΩ)2/3y=(\mu\Omega)^{2/3}, with μ=μ1+μ2\mu=\mu_{1}+\mu_{2} the total irreducible mass and η=μ1μ2/μ2\eta\!=\!\mu_{1}\mu_{2}/\mu^{2}. Since for nonspinning binaries ma=μam_{a}=\mu_{a}, the sets of variables (μ,η,y)(\mu,\eta,y) and (m,ν,x)(m,\nu,x) coincide. Notice the logarithmic running at 4PN order, related to the occurence of gravitational-wave tails at that order.

On the other hand, by combining the leading order and next-to-leading order spin-orbit contributions to the binding energy and angular momentum with the contributions coming from the replacement of the masses mam_{a} by the irreducible masses μa\mu_{a} and the spins χa\chi_{a} by their expressions as functions of the orbital frequency Ω\Omega (see Ref. Bl.al.13), we obtain the additional spin-related contributions EcorE^{\text{cor}} and JcorJ^{\text{cor}} in the corotating case. Those explicitly read

Ecor\displaystyle\!\!\!E^{\text{cor}} =μ(26η)y3+μη(10+25η)y4+μη(21+3203η3256η2)y5+o(y5),\displaystyle=\mu\left(2-6\eta\right)y^{3}+\mu\eta\left(-10+25\eta\right)y^{4}+\mu\eta\left(-21+\frac{320}{3}\eta-\frac{325}{6}\eta^{2}\right)y^{5}+o(y^{5}), (70a)
ΩJcor\displaystyle\!\!\!\Omega J^{\text{cor}} =μ(412η)y3+μη(16+40η)y4+μη(30+2243η45512η2)y5+o(y5).\displaystyle=\mu\left(4-12\eta\right)y^{3}+\mu\eta\left(-16+40\eta\right)y^{4}+\mu\eta\left(-30+\frac{224}{3}\eta-\frac{455}{12}\eta^{2}\right)y^{5}+o(y^{5}). (70b)

These 2PN, 3PN and 4PN spin-related contributions have to be added to the 4PN-accurate expressions (69) for spinless binaries, such that for corotating binaries the binding energy and angular momentum are given by E=Ens+EcorE=E^{\text{ns}}+E^{\text{cor}} and J=Jns+JcorJ=J^{\text{ns}}+J^{\text{cor}}. The 3PN results were first (incorrectly) computed in Ref. Bl.02, by using the leading-order formula ωa=Ω\omega_{a}=\Omega for the proper rotation frequencies of the spinning particles, and were later corrected in Ref. Bl.al.13 using Eq. (51). Here we extended these calculations to 4PN order.

One can then check that the formulas (69) and (70), together with the expressions (IV.4) for the particles’ (renormalized) redshifts, obey the first law (26) for corotating binaries, which is equivalent to the partial differential equations

EΩ|μa\displaystyle\frac{\partial E}{\partial\Omega}\bigg{|}_{\mu_{a}} =ΩJΩ|μa,\displaystyle=\Omega\,\frac{\partial J}{\partial\Omega}\bigg{|}_{\mu_{a}}\,, (71a)
Eμa|Ω\displaystyle\frac{\partial E}{\partial\mu_{a}}\bigg{|}_{\Omega} =ΩJμa|Ω+caza1.\displaystyle=\Omega\,\frac{\partial J}{\partial\mu_{a}}\bigg{|}_{\Omega}+c_{a}z_{a}-1\,. (71b)

Similarly, one can easily check that the first integral (27) is also satisfied.

The partial differential equations (71) can also be combined to establish that the redshifts cazac_{a}z_{a} take a simple form when expressed in terms of the variables yy and η\eta. Indeed, performing the change of variables (Ω,μ1,μ2)(y,μ,η)(\Omega,\mu_{1},\mu_{2})\to(y,\mu,\eta), we have (see also Sec. IV B in Ref. Le.al.12)

caza1=EμaΩJμa=^+2y3^y+12(1±Γ4η)^η,c_{a}z_{a}-1=\frac{\partial E}{\partial\mu_{a}}-\Omega\,\frac{\partial J}{\partial\mu_{a}}=\hat{\mathcal{E}}+\frac{2y}{3}\,\frac{\partial\hat{\mathcal{E}}}{\partial y}+\frac{1}{2}(1\pm\Gamma-4\eta)\,\frac{\partial\hat{\mathcal{E}}}{\partial\eta}\,, (72)

where we introduced the convenient combination ^(EΩJ)/μ\mathcal{\hat{E}}\equiv(E-\Omega J)/\mu, heuristically the binary’s reduced binding energy in a corotating frame. Now, on dimensional grounds alone, we know that ^(y,η)=k0ηke(k)(y)\mathcal{\hat{E}}(y,\eta)=\sum_{k\geqslant 0}\eta^{k}e_{(k)}(y), with e(k)e_{(k)} dimensionless functions of the dimensionless variable yy. This can be checked explicitly, for instance up to 4PN order, using the expressions (69) and (70). Hence, Eq. (72) implies for the renormalized redshifts

caza=f(y,η)±Γg(y,η),c_{a}z_{a}=f(y,\eta)\pm\Gamma\,g(y,\eta)\,, (73)

for some functions ff and gg that can (at least formally) be written as f(y,η)=k0ηkf(k)(y)f(y,\eta)=\sum_{k\geqslant 0}\eta^{k}f_{(k)}(y) and g(y,η)=k0ηkg(k)(y)g(y,\eta)\!=\!\sum_{k\geqslant 0}\eta^{k}g_{(k)}(y). This can also be checked explicitly, up to 4PN order, from the formula (IV.4). Recalling the analogy (28), or following a similar argument starting from the first law (12) for corotating black hole binaries, the general result (73) motivated us in Sec. V.3 to rewrite the perturbative expansions (66) in the form (67)–(68).

Appendix B Redshift of a corotating point particle in a Kerr background

In this appendix, we consider the PN expansion of the gravitational self-force contribution z(1)(Ω)z_{(1)}(\Omega) to the redshift of a point particle on the unique circular equatorial corotating orbit around a Kerr black hole of given spin. To do so, we use the results of Ka.al.16, who computed the 8.5PN expansion of the quantity ΔU=z(1)/z(0)2\Delta U=-z_{(1)}/z_{(0)}^{2}, where z(0)(Ω)z_{(0)}(\Omega) is given in the first line below (62), for a generic circular equatorial orbit, i.e., for any value of the Kerr black hole spin. Here, we specify to the corotating case by substituting the formula χ=4u3/2[14u3]1/2\chi=4u^{3/2}[1-4u^{3}]^{1/2} into their Eqs. (4.1)–(4.2). Moreover, since we normalize the orbital frequency Ω\Omega using the irreducible mass μ\mu, rather than the black hole mass MM, we also substitute M=μ[14u3]1/2M=\mu[1-4u^{3}]^{-1/2}. Up to 8.5PN order, we readily obtain

z(1)(u)=k=118/2(ak+bklnu+ckln2u)uk+o(u18/2),z_{(1)}(u)=\sum_{k=1}^{18/2}\left(a_{k}+b_{k}\ln{u}+c_{k}\ln^{2}{\!u}\right)u^{k}+o(u^{18/2})\,, (74)

where the index kk runs over integer and half-integer values, with the non-zero coefficients

a1\displaystyle a_{1} =a2=a3=1,\displaystyle=-a_{2}=-a_{3}=1\,, (75a)
a4\displaystyle a_{4} =5234132π2,\displaystyle=\frac{52}{3}-\frac{41}{32}\pi^{2}\,, (75b)
a5\displaystyle a_{5} =95815+1291512π2+1285γE+2565ln2,\displaystyle=-\frac{958}{15}+\frac{1291}{512}\pi^{2}+\frac{128}{5}\gamma_{\text{E}}+\frac{256}{5}\ln{2}\,, (75c)
a6\displaystyle a_{6} =25877873150+1267791536π29976105γE23672105ln2+2437ln3,\displaystyle=-\frac{2587787}{3150}+\frac{126779}{1536}\pi^{2}-\frac{9976}{105}\gamma_{\text{E}}-\frac{23672}{105}\ln{2}+\frac{243}{7}\ln{3}\,, (75d)
a6.5\displaystyle a_{6.5} =13696525π,\displaystyle=\frac{13696}{525}\pi\,, (75e)
a7\displaystyle a_{7} =3922767414175+8104603671769472π22800873262144π43576882835γE\displaystyle=-\frac{39227674}{14175}+\frac{810460367}{1769472}\pi^{2}-\frac{2800873}{262144}\pi^{4}-\frac{357688}{2835}\gamma_{\text{E}}
+680881ln219447ln3,\displaystyle\quad\,+\frac{6808}{81}\ln{2}-\frac{1944}{7}\ln{3}\,, (75f)
a7.5\displaystyle a_{7.5} =3686933675π,\displaystyle=-\frac{368693}{3675}\pi\,, (75g)
a8\displaystyle a_{8} =4792193805194191008125+48090893574972477260800π2+56161864116777216π4\displaystyle=-\frac{4792193805194}{191008125}+\frac{4809089357497}{2477260800}\pi^{2}+\frac{561618641}{16777216}\pi^{4}
+126820742385457375γE109568525γE2438272525γEln2\displaystyle\quad\,+\frac{12682074238}{5457375}\gamma_{\text{E}}-\frac{109568}{525}\gamma_{\text{E}}^{2}-\frac{438272}{525}\gamma_{\text{E}}\ln{2}
+179721761265457375ln2438272525ln22+1570436124640ln3\displaystyle\quad\,+\frac{17972176126}{5457375}\ln{2}-\frac{438272}{525}\ln^{2}{2}+\frac{15704361}{24640}\ln{3}
+195312519008ln5+20485ζ(3),\displaystyle\quad\,+\frac{1953125}{19008}\ln{5}+\frac{2048}{5}\zeta(3)\,, (75h)
a8.5\displaystyle a_{8.5} =3612092923274425π,\displaystyle=-\frac{361209292}{3274425}\pi\,, (75i)
a9\displaystyle a_{9} =359148730284808317381739375+124200083594298192484403200π2227872841284736442450944π4\displaystyle=\frac{3591487302848083}{17381739375}+\frac{1242000835942981}{92484403200}\pi^{2}-\frac{22787284128473}{6442450944}\pi^{4}
1857416227468496621125γE+744310411025γE2+2555584735γEln23790849γEln3\displaystyle\quad\,-\frac{1857416227468}{496621125}\gamma_{\text{E}}+\frac{7443104}{11025}\gamma_{\text{E}}^{2}+\frac{2555584}{735}\gamma_{\text{E}}\ln{2}-\frac{37908}{49}\gamma_{\text{E}}\ln{3}
4281723492044496621125ln2+60877761575ln223790849ln2ln31895449ln23\displaystyle\quad\,-\frac{4281723492044}{496621125}\ln{2}+\frac{6087776}{1575}\ln^{2}{2}-\frac{37908}{49}\ln{2}\ln{3}-\frac{18954}{49}\ln^{2}{3}
+977811113907196196000ln3513671875370656ln587616105ζ(3),\displaystyle\quad\,+\frac{977811113907}{196196000}\ln{3}-\frac{513671875}{370656}\ln{5}-\frac{87616}{105}\zeta(3)\,, (75j)
a9.5\displaystyle a_{9.5} =25601004551766831048863816000π2344755255125πγE4689510455125πln2+2191361575π3,\displaystyle=-\frac{2560100455176683}{1048863816000}\pi-\frac{23447552}{55125}\pi\gamma_{\text{E}}-\frac{46895104}{55125}\pi\ln{2}+\frac{219136}{1575}\pi^{3}\,, (75k)
b5\displaystyle b_{5} =645,b6=4988105,b7=1788442835,\displaystyle=\frac{64}{5}\,,\quad b_{6}=-\frac{4988}{105}\,,\quad b_{7}=-\frac{178844}{2835}\,, (75l)
b8\displaystyle b_{8} =63410371195457375109568525γE219136525ln2,\displaystyle=\frac{6341037119}{5457375}-\frac{109568}{525}\gamma_{\text{E}}-\frac{219136}{525}\ln{2}\,, (75m)
b9\displaystyle b_{9} =935064864134496621125+744310411025γE+1277792735ln21895449ln3,\displaystyle=-\frac{935064864134}{496621125}+\frac{7443104}{11025}\gamma_{\text{E}}+\frac{1277792}{735}\ln{2}-\frac{18954}{49}\ln{3}\,, (75n)
b9.5\displaystyle b_{9.5} =1172377655125π,\displaystyle=-\frac{11723776}{55125}\pi\,, (75o)
c8\displaystyle c_{8} =27392525,c9=186077611025,\displaystyle=-\frac{27392}{525}\,,\quad c_{9}=\frac{1860776}{11025}\,, (75p)

where ζ(s)\zeta(s) is the Riemann zeta function. The increase in “transcendentality” of the numbers entering the PN expansion coefficients as the order increases is discussed in BiDa2.14. Note also the half-integral contributions occuring at the leading 5.5PN order, first pointed out in Sh.al.14. These were later shown to be closely related to gravitational-wave tails-of-tails Bl.al.14; Bl.al2.14.

In Sec. V.1 we used the 8.5PN-accurate formula (74)–(75) in order to approximate the surface gravity (66a) of the smaller body, to linear order in the (irreducible) mass ratio qμq_{\mu}, and in Eqs. (68), following the mass-ratio rescaling. Ideally, one should use numerical data for z(1)(u)z_{(1)}(u) that does not rely on a PN expansion. However, we checked that the perturbative prediction (66a) with the PN expansion (74)–(75) is fairly insensitive to the PN order at which we truncate the 𝒪(qμ)\mathcal{O}(q_{\mu}) contribution from z(1)(u)z_{(1)}(u). More precisely, we checked that for qμ=1/10q_{\mu}=1/10 the relative differences between the 6.5PN, 7.5PN and 8.5PN approximations remain below 1%1\% for μΩ0.16\mu\Omega\lesssim 0.16. For larger orbital frequencies, however, the PN expansion becomes unreliable, and exact numerical gravitational self-force data would be needed.