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

cmlargesymbols0 cmlargesymbols1 cmlargesymbols0 cmlargesymbols1 cmlargesymbols2 cmlargesymbols3

Gravitational bremsstrahlung from spinning binaries
in the post-Minkowskian expansion

Massimiliano Maria Riva    Filippo Vernizzi    Leong Khim Wong Université Paris-Saclay, CNRS, CEA, Institut de Physique Théorique, 91191 Gif-sur-Yvette, France
Abstract

We present a novel calculation of the four-momentum that is radiated into gravitational waves during the scattering of two arbitrarily spinning bodies. Our result, which is accurate to leading order in GG, to quadratic order in the spins, and to all orders in the velocity, is derived by using a Routhian-based worldline effective field theory formalism in concert with a battery of analytic techniques for evaluating loop integrals. While nonspinning binaries radiate momentum only along the direction of their relative velocity, we show that the inclusion of spins generically allows for momentum loss in all three spatial directions. We also verify that our expression for the radiated energy agrees with the overlapping terms from state-of-the-art calculations in post-Newtonian theory.

. ntroduction

The burgeoning field of gravitational-wave astronomy [1, 2] will offer new opportunities to explore questions in fundamental physics, test the nature of strong-field gravity, and constrain various binary formation and evolution channels [3, 4, 5, 6, 7, 8, 9]. As binary systems with spinning black holes constitute one of the primary sources of gravitational waves, modeling precisely how spin influences a binary’s inspiral is essential for making robust detections and performing accurate parameter estimation studies [10, 11, 12].

In the traditional approach to the two-body problem, one makes the so-called post-Newtonian (PN) expansion [13]: the equations of motion for the binary and the gravitational field are solved order by order simultaneously in powers of GG and v2v^{2}; respectively, the gravitational constant and the square of the relative velocity between the two bodies. Since the two parameters are related by the virial theorem, this perturbative scheme is ideally suited to the study of bound orbits.

An alternative approach, which lends itself more naturally to the study of unbounded orbits (i.e., scattering encounters), is the post-Minkowskian (PM) approximation [14]. Here, one also expands in powers of GG, but keeps vv fully relativistic. While the study of unbounded orbits may, at face value, seem far removed from the coalescing binaries that gravitational-wave detectors observe, quantities computed in one scenario can be linked to the other via, e.g., analytic continuation [15, 16, 17, 18]. Alternatively, PM calculations could also be used as inputs to improve the accuracy of the effective-one-body approach [14, 19, 20, 21, 22, 23, 24]—a popular semianalytic method for constructing waveform templates.

In recent years, rapid advancements in the PM program have been driven by the scattering-amplitudes community [25, 26, 27, 28, 29, 30, 31, 32, 33, 34, 35, 36, 37, 38, 39, 40], who (at present) have pushed out calculations in the conservative sector up to 4PM; i.e., up to O(G4)O(G^{4}) [41, 42]. Analogous results were also obtained independently through various worldline effective field theory (EFT) approaches [43, 44, 45, 46, 47, 48]. These results were later extended to include tidal deformation [49, 50, 51, 52, 53, 54, 55, 56] and spin effects [57, 58, 59, 60, 61, 62, 63, 64, 65, 66, 67, 68, 69].

Developments in the radiative sector are more recent. The four-momentum emitted into gravitational waves by a nonspinning binary was first computed at leading (3PM) order in Refs. [70, 71] via the “KMOC” approach [29], independently in Ref. [72] via the eikonal approach, and then in Ref. [73] via the worldline EFT approach. Tidal contributions were later included in Ref. [74]. (See also Refs. [75, 76, 77, 78, 79, 80, 81, 82, 83, 84, 85, 67, 82, 86, 87, 88, 89, 68] for related works on radiative effects.) Notably absent from the literature, however, is the inclusion of spins in the radiated observables at 3PM.

To be precise, the outgoing waveform from a spinning binary has been computed up to 2PM in Ref. [85]. Using this to compute the radiated four-momentum at 3PM is challenging, however, because of the multiscale nature of the resulting integrals, which have so far proven to be intractable unless one also performs a low-velocity expansion [83, 84, 85]. Fortunately, this is not the only option, and indeed our goal in this paper is to compute the four-momentum radiated at 3PM up to quadratic order in the spins and to all orders in the velocity.

We bypass the aforementioned complications with the waveform by formulating the problem as an integral of the outgoing graviton momentum over phase space, weighted by (what is essentially) the square of the binary’s (pseudo) stress-energy tensor. The latter we construct by using the worldline EFT formalism, while the loop integrals that arise are computed by appropriating powerful techniques from high-energy physics [90]; namely, reverse unitarity [91, 92, 93, 94], a reduction to master integrals via integration by parts [95, 96, 97], and differential equation methods [98, 99, 100, 101, 102, 103], as previously used in Refs. [70, 71, 72, 73, 74] for the nonspinning case.

These techniques are described in more detail in Sec. II. In Sec. III, we discuss the key features of our main result, but owing to its length, we present the full expression only in the Supplemental Material [104], a computer-readable version of which is available in the ancillary files attached to the arXiv submission of this paper. Also included in the Supplemental Material are explicit expressions for some of the intermediate quantities that we calculate, like the stress-energy tensor. We conclude in Sec. IV.

I. ethods

 Worldline effective field theory

Consider the case of two spinning bodies approaching one another from infinity, and suppose that their distance of closest approach remains much larger than their individual radii. In this scenario, the details of their scattering encounter are well described by an EFT in which the two bodies are treated as point particles traveling along the worldlines of their respective centers of energy. Their dynamics are conveniently described by a Routhian [65, 105, 106, 107], which for each body of mass mm reads

=\displaystyle\mathcal{R}= 12(mgμν𝒰μ𝒰ν+ωμab𝒮ab𝒰μ\displaystyle-\frac{1}{2}\bigg{(}mg_{\mu\nu\,}\mathcal{U}^{\mu}\mathcal{U}^{\mathstrut\nu}+\omega_{\mu}^{ab}\mathcal{S}_{ab}\mathcal{U}^{\mu}
1m𝒰a𝒰eRebcd𝒮ab𝒮cd1mCEEab𝒮a𝒮c)cb.\displaystyle-\frac{1}{m}\mathcal{U}_{a}\,\mathcal{U}^{e}R_{ebcd}\mathcal{S}^{ab}\mathcal{S}^{cd}-\frac{1}{m}C_{E}E_{ab}\mathcal{S}^{a}{}_{c}\mathcal{S}{}^{cb}\bigg{)}. (1)

The translational degrees of freedom (d.o.f.s) of this body are encoded in four worldline coordinates xμ(τ)x^{\mu}(\tau), which chart the integral curve of the four-velocity 𝒰μdxμ/dτ{\mathcal{U}^{\mu}\equiv\text{d}x^{\mu}/\text{d}\tau}. Only three of these are needed to specify the position of the body uniquely, however, and so we remove the remaining unphysical d.o.f. by imposing the constraint 𝒰μ𝒰μ=1{\mathcal{U}_{\mu}\mathcal{U}^{\mu}=1} [43, 65]. Meanwhile, the body’s rotational d.o.f.s are encoded in the antisymmetric spin tensor 𝒮ab\mathcal{S}_{ab}, which notably is defined in a locally flat frame with coordinates {ya}\{y^{a}\}. Tensors defined in the general coordinate frame {xμ}\{x^{\mu}\} are transformed into the former by way of the vielbein eμaya/xμ{e_{\mu}^{a}\equiv\partial y^{a}/\partial x^{\mu}} (e.g., 𝒰aeμa𝒰μ{\mathcal{U}^{a}\equiv e^{a}_{\mu}\mathcal{U}^{\mu}}), which also defines for us the spin connection ωμab:-gρσeσbμeρa{\omega_{\mu}^{ab}\coloneq g^{\rho\sigma}\!e^{b}_{\sigma}\nabla_{\!\mu}^{\vphantom{a}}e^{a}_{\rho}}. Only three of the six components in 𝒮ab\mathcal{S}_{ab} are needed to specify the spin of the body uniquely; hence, the other three d.o.f.s are to be removed by imposing a spin supplementary condition (SSC) [108, 109, 110, 111]. Given the relativistic nature of the problem, the covariant SSC, 𝒰a𝒮ab=0+O(𝒮3){\,\mathcal{U}^{a}\mathcal{S}_{ab}=0+O(\mathcal{S}^{3})} [111], proves to be the most convenient choice.

The first three terms in Eq. (1) are universal, in the sense that they apply to any body with a mass monopole and spin dipole. Higher-order multipole moments, however, are sensitive to the body’s internal structure, and this is why the final term in Eq. (1), which describes the self-induced quadrupole moment of the rotating body (EμνRμρνσ𝒰ρ𝒰σ{E_{\mu\nu}\equiv R_{\mu\rho\nu\sigma}\mathcal{U}^{\rho}\mathcal{U}^{\sigma}}), is accompanied by the Wilson coefficient CEC_{E}. Kerr black holes have CE=1{C_{E}=1} [106], although this value can be larger for objects like neutron stars [112, 113]. Infinitely many more terms can be appended to Eq. (1) should we wish to include even higher-order multipoles [114], or other finite-size effects like tidal deformations [115, 74], but these all come with higher powers of either the curvature tensors or the spin, and so are irrelevant to our purposes here.

As the Routhian behaves like a Lagrangian from the point of view of the translational d.o.f.s, but like a Hamiltonian with regards to the rotational d.o.f.s, the equations of motion for each body follow from a mixture of Euler-Lagrange and Hamilton equations [116]; namely,

δδxμdτ=0andddτ𝒮ab={𝒮ab,}.\frac{\delta}{\delta x^{\mu}}\int\text{d}\tau\,\mathcal{R}=0\quad\text{and}\quad\frac{\text{d}}{\text{d}\tau}\mathcal{S}^{ab}=\{\mathcal{S}^{ab},\mathcal{R}\}. (2)

The only nontrivial Poisson bracket we require is [105]

{𝒮ab,𝒮cd}=ηac𝒮bd+ηbd𝒮acηad𝒮bcηbc𝒮ad,\{\mathcal{S}^{ab},\mathcal{S}^{cd}\}=\eta^{ac}\mathcal{S}^{bd}+\eta^{bd}\mathcal{S}^{ac}-\eta^{ad}\mathcal{S}^{bc}-\eta^{bc}\mathcal{S}^{ad}, (3)

where ηab\eta^{ab} is the Minkowski metric with a mostly minus signature. Note that the aforementioned constraints on 𝒰μ\mathcal{U}^{\mu} and 𝒮ab\mathcal{S}_{ab} should be imposed only after all functional derivatives and Poisson brackets have been evaluated.

To fully specify our EFT, we must also endow the metric gμνg_{\mu\nu} with its own dynamics; hence, we take the full effective action to be S=SEH+SFP+A=12dτAA{S=S_{\text{EH}}+S_{\text{FP}}+\sum_{A=1}^{2}\int\text{d}\tau_{A}\mathcal{R}_{A}}, where SEHS_{\text{EH}} is the Einstein-Hilbert action, SFPS_{\text{FP}} is a Faddeev-Popov term that enforces the de Donder gauge, and the label A{1,2}{A\in\{1,2\}} is used to distinguish between the binary’s two constituents.

 Stress-energy tensor

As a precursor to computing the radiated four-momentum, we first determine the stress-energy tensor TμνT^{\mu\nu} for the binary as a whole. This object is sourced by the multipolar moments of the two bodies, as well as by the energy stored in nonlinear interactions of the gravitational field, and can be obtained perturbatively from our EFT with the help of Feynman diagrams once we expand the action in powers of κhμν:-gμνημν{\kappa h_{\mu\nu}\coloneq g_{\mu\nu}-\eta_{\mu\nu}}, with κ32πG{\kappa\equiv\sqrt{32\pi G}}. Crucially, since

eμa=ηaν(ημν+12κhμν18κ2hμρhρ+νO(κ3h3))e_{\mu}^{a}=\eta^{a\nu}\bigg{(}\eta_{\mu\nu}+\frac{1}{2}\kappa h_{\mu\nu}-\frac{1}{8}\kappa^{2}h_{\mu\rho}h^{\rho}{}_{\nu}+O(\kappa^{3}h^{3})\bigg{)} (4)

in this expansion, Greek and Latin indices are now indistinct. The spin tensors are, nonetheless, still defined in their respective locally flat frames [65].

Refer to caption
Figure 1: Feynman rules relevant to our calculation.

Only the Feynman rules shown in Fig. 1 are required at the order in GG to which we are working. In dd-dimensional momentum space (we work in dd dimensions for added generality), the graviton propagator in Fig. 1(a) is iPμνρσ/k2iP_{\mu\nu\rho\sigma}/k^{2}, with Pμνρσ=ημ(ρησ)ν+ημνηρσ/(d2){P_{\mu\nu\rho\sigma}=\eta_{\mu(\rho}\eta_{\sigma)\nu}+\eta_{\mu\nu}\eta_{\rho\sigma}/(d-2)} (an iϵi\epsilon prescription is unnecessary at this order because the loop integrals never hit the poles at k2=0{k^{2}=0} [83]), while the worldline vertex for single-graviton emission is

Fig. 1(b)=12iκdτAeikxA[mA𝒰A𝒰Aμ+νikρ𝒮A𝒰Aρ(μν)\displaystyle\text{Fig.~\ref*{fig:feynman_rules}(b)}=-\frac{1}{2}i\kappa\!\int\text{d}\tau_{A}\,e^{ik\cdot x_{A}}\bigg{[}m_{A\,}\mathcal{U}_{A}{}^{\mu\,}\mathcal{U}_{A}{}^{\nu}+ik_{\rho}\mathcal{S}_{A}{}^{\rho(\mu}\mathcal{U}_{A}{}^{\nu)}\allowdisplaybreaks
+1mAkρkσ𝒰Aα(𝒰A𝒮A(μ𝒮Aν)ρ+σα𝒰A𝒮Aρ𝒮Aσ(μ)ν)α\displaystyle\;\;+\frac{1}{m_{A}}k_{\rho}k_{\sigma}\mathcal{U}_{A\,\alpha}\big{(}\mathcal{U}_{A}{}^{(\mu}\mathcal{S}_{A}{}^{\nu)\rho}\mathcal{S}_{A}{}^{\sigma\alpha}+\mathcal{U}_{A}{}^{\rho}\mathcal{S}_{A}{}^{\sigma(\mu}\mathcal{S}_{A}{}^{\nu)\alpha}\big{)}\allowdisplaybreaks
+12mACEAkρkσ(𝒮A𝒮Aρα𝒰Aσα𝒰Aμν\displaystyle\;\;+\frac{1}{2m_{A}}C_{E_{A}}k_{\rho}k_{\sigma}\big{(}\mathcal{S}_{A}{}^{\rho\alpha}\mathcal{S}_{A}{}^{\sigma}{}_{\alpha}\,\mathcal{U}_{A}{}^{\mu}\mathcal{U}_{A}{}^{\nu}\allowdisplaybreaks
+2𝒰A𝒮Aρ𝒮Aασα𝒰A(μ+ν)𝒮A𝒮A(μα𝒰Aν)α𝒰Aρ)σ].\displaystyle\;\;+2\,\mathcal{U}_{A}{}^{\rho}\mathcal{S}_{A}{}^{\sigma\alpha}\mathcal{S}_{A\,\alpha}{}^{(\mu}\mathcal{U}_{A}{}^{\nu)}+\mathcal{S}_{A}{}^{(\mu}{}_{\alpha}\mathcal{S}_{A}{}^{\nu)\alpha}\mathcal{U}_{A}{}^{\rho}\mathcal{U}_{A}{}^{\sigma}\big{)}\bigg{]}. (5)

Expressions for the two remaining vertices, which are much lengthier, are presented in the Supplemental Material [104].

In addition to making the weak-field expansion above, we must also expand the body variables XA(xA,𝒰A,𝒮A){X_{A}\equiv(x_{A},\mathcal{U}_{A},\mathcal{S}_{A})} about their initial straight-line trajectories in order to achieve manifest power counting in GG. We therefore write [43, 65]

XA(τA)=X¯A(τA)+n=1δ(n)XA(τA),X_{A}(\tau_{A})=\overline{X}_{A}(\tau_{A})+\sum_{n=1}^{\infty}\delta^{(n)}X_{A}(\tau_{A}), (6)

where δ(n)XA\delta^{(n)}X_{A} is the O(Gn)O(G^{n}) deflection away from the initial trajectory X¯A\overline{X}_{A} due to the gravitational pull of the other body. The 1PM deflections δ(1)XA\delta^{(1)}X_{A}, which we will need later in our calculation, were previously computed using Eq. (2) in Refs. [43, 65]. (They are reproduced in the Supplemental Material [104] for completeness.) As for X¯A\overline{X}_{A}, we write

x¯=AμbAμ+uAμτA,𝒰¯=AμuAμ,and𝒮¯=AμνmAsAμν,\overline{x}{}^{\mu}_{A}=b_{A}^{\mu}+u_{A}^{\mu}\tau_{A}^{\mathstrut},\quad\overline{\mathcal{U}}{}^{\mu}_{A}=u_{A}^{\mu},\;\;\text{and}\;\;\overline{\mathcal{S}}{}_{A}^{\mu\nu}=m_{A}^{\mathstrut}s_{A}^{\mu\nu}, (7)

where the constant vectors uAμu_{A}^{\mu} and bAμb_{A}^{\mu} are the initial velocity and orthogonal displacement of the AAth body, respectively, while the constant tensor sAμνs_{A}^{\mu\nu} describes its initial spin per unit mass. Note that sAμνuAν=0{s_{A}^{\mu\nu}u_{A\,\nu}=0} as per the covariant SSC, while the impact parameter bμ:-b1μb2μ{b^{\mu}_{\mathstrut}\coloneq b_{1}^{\mu}-b_{2}^{\mu}} satisfies buA=0{b\cdot u_{A}=0} [43].

Refer to caption
Figure 2: Feynman diagrams contributing to the stress-energy tensor up to next-to-leading order in GG. While not drawn explicitly, our calculation includes the mirror inverses of (a) and (b), which are obtained by interchanging the body labels 12{1\leftrightarrow 2} and redefining the loop momentum qkq{q\mapsto k-q}.

We now use these rules to compute the (tree-level) expectation value hμν(k)κPμνρσTρσ(k)/(2k2){{\langle h_{\mu\nu}(k)\rangle}\equiv\kappa P_{\mu\nu\rho\sigma}T^{\rho\sigma}(k)/(2k^{2})}, from which the (classical) stress-energy tensor TμνT^{\mu\nu} may be extracted. At leading order in GG, only the diagram in Fig. 2(a), with XAX_{A} replaced by X¯A\overline{X}_{A}, contributes. The result is

TLOμν(k)=A=12δ(kuA)mAeikbA[uAuAμ+νikρsAρuA(μν)\displaystyle T^{\mu\nu}_{\text{LO}}(k)=\sum_{A=1}^{2}\text{\smash{\raisebox{2.58334pt}{--}}\kern-4.73611pt$\delta$}(k\cdot u_{A})m_{A}e^{ik\cdot b_{A}}\Big{[}u_{A}{}^{\mu}u_{A}{}^{\nu\mathstrut}+ik^{\rho}s_{A\,\rho}{}^{(\mu}u_{A}{}^{\nu)}
12CEA(kρsAsAσρσkαα)uAuAμ]ν,\displaystyle-\frac{1}{2}C_{E_{A}}(k_{\rho}s_{A}{}^{\rho\sigma}{s}_{A\,\sigma}{}^{\alpha}k_{\alpha})u_{A}{}^{\mu}u_{A}{}^{\nu}\Big{]}, (8)

where the delta function δ(x)2πδ(x){\text{\smash{\raisebox{2.58334pt}{--}}\kern-4.73611pt$\delta$}(x)\equiv 2\pi\delta(x)} comes from having performed the integral over τA\tau_{A} in Eq. (5).

All three diagrams in Fig. 2 contribute at next-to-leading order in GG. From Fig. 2(a), we extract the O(G)O(G) part of the diagram by expanding XAX_{A} up to 1PM, whereas for Figs. 2(b) and 2(c), it suffices to replace XAX_{A} by X¯A\overline{X}_{A}. The total result is

TNLOμν(k)=κ2M2ν4qΔ12(q,k)q2(kq)2tμν(q,k)eiqbeikb2,T^{\mu\nu}_{\text{NLO}}(k)=\frac{\kappa^{2}M^{2}\nu}{4}\!\int_{q}\frac{\Delta_{12}(q,k)}{q^{2}(k-q)^{2}}t^{\mu\nu}(q,k)\,e^{iq\cdot b}e^{ik\cdot b_{2}}, (9)

where M=m1+m2{M=m_{1}+m_{2}} is the binary’s total mass, ν=m1m2/M2{\nu=m_{1}m_{2}/M^{2}} is its symmetric mass ratio, qddq/(2π)d{\int_{q}\!\equiv\!\int\text{d}^{d}q/(2\pi)^{d}}, and Δ12(q,k):-δ(qu1)δ((kq)u2)\Delta_{12}(q,k)\coloneq\,\text{\smash{\raisebox{2.58334pt}{--}}\kern-4.73611pt$\delta$}(q\cdot u_{1})\,\text{\smash{\raisebox{2.58334pt}{--}}\kern-4.73611pt$\delta$}((k-q)\cdot u_{2}). An explicit expression for the object tμνt^{\mu\nu}, accurate to O(s2)O(s^{2}), is presented in the Supplemental Material [104].

 Loop integrals

Given the above, we may now compute the radiated four-momentum via the definition [117]

Pradμ=κ24kδ+(k2)kμ[Tαν(k)PανρσTρσ(k)],P^{\mu}_{\text{rad}}=\frac{\kappa^{2}}{4}\!\int_{k}\text{\smash{\raisebox{2.58334pt}{--}}\kern-4.73611pt{\hbox{\delta}}}_{+}(k^{2})k^{\mu}\big{[}T^{\alpha\nu}(k)P_{\alpha\nu\rho\sigma}T^{*\rho\sigma}(k)\big{]}, (10)

where δ+(k2)ddk/(2π)d{\text{\smash{\raisebox{2.58334pt}{--}}\kern-4.73611pt$\delta$}_{+}(k^{2})\,\text{d}^{d}k/(2\pi)^{d}} is the Lorentz-invariant phase-space measure for the emission of on shell gravitons. Observe that because the delta functions in Eq. (8) have compact support away from k2=0{k^{2}=0}, TLOμνT^{\mu\nu}_{\text{LO}} is nonradiative and so does not contribute to PradμP^{\mu}_{\text{rad}}. It therefore suffices to substitute Eq. (9) into Eq. (10) when working to leading order in GG. We then find

Pradμ=\displaystyle P^{\mu}_{\text{rad}}=\> κ6M4ν264k,q1,q2tαν(q1,k)Pανρσtρσ(q2,k)q12q22(kq1)2(kq2)2kμ\displaystyle\frac{\kappa^{6}M^{4}\nu^{2}}{64}\int_{k,q_{1},q_{2}}\!\!\frac{t^{\alpha\nu}(q_{1},k)\,P_{\alpha\nu\rho\sigma}\,t^{*\rho\sigma}(q_{2},k)}{q_{1}^{2}q_{2}^{2}(k-q_{1})^{2}(k-q_{2})^{2}}k^{\mu}
×ei(q1q2)bδ+(k2)Δ12(q1,k)Δ12(q2,k).\displaystyle\times e^{i(q_{1}-q_{2})\cdot b}\,\text{\smash{\raisebox{2.58334pt}{--}}\kern-4.73611pt$\delta$}_{+}(k^{2})\,\Delta_{12}(q_{1},k)\Delta_{12}(q_{2},k). (11)

Next, we define new momentum variables q=q1q2{q=q_{1}-q_{2}}, 1=q2{\ell_{1}=-q_{2}}, and 2=q1k{\ell_{2}=q_{1}-k} [73] in order to write

Pradμ=κ6M4ν264qδ(qu1)δ(qu2)eiqbQμ(q).P^{\mu}_{\text{rad}}=\frac{\kappa^{6}M^{4}\nu^{2}}{64}\!\int_{q}\text{\smash{\raisebox{2.58334pt}{--}}\kern-4.73611pt{\hbox{\delta}}}(q\cdot{u}_{1})\text{\smash{\raisebox{2.58334pt}{--}}\kern-4.73611pt{\hbox{\delta}}}(q\cdot{u}_{2})\,e^{iq\cdot b}Q^{\mu}(q). (12)

The radiated four-momentum may thus be viewed as the inverse Fourier transform of some object QμQ^{\mu}, which is expressible as a sum of terms in which qμq^{\mu}, uAμu_{A}^{\mu}, and sAμνs_{A}^{\mu\nu} are contracted amongst themselves and with the two-loop integrals

Gn1n9μ1μiν1νj:-1,21μ11μi2ν12νjρ¯1n1ρ2n2ρ3n3ρ¯4n4ρ5n5ρ6n6ρ¯7n7ρ8n8ρ9n9.G^{\mu_{1}\cdots\,\mu_{i}\nu_{1}\cdots\,\nu_{j}}_{n_{1}\cdots\,n_{9}}\coloneq\int_{\ell_{1},\ell_{2}}\frac{\ell_{1}^{\mathstrut\mu_{1}}\cdots\,\ell_{1}^{\mathstrut\mu_{i}}\,\ell_{2}^{\mathstrut\nu_{1}}\cdots\,\ell_{2}^{\mathstrut\nu_{j}}}{\underline{\rho}_{1}^{n_{1}}\rho_{2}^{n_{2}}\rho_{3}^{n_{3}}\underline{\rho}_{4}^{n_{4}}\rho_{5}^{n_{5}}\rho_{6}^{n_{6}}\underline{\rho}_{7}^{n_{7}}\rho_{8}^{n_{8}}\rho_{9}^{n_{9}}}. (13)

Following Ref. [90], we define ρ2=21u2\rho_{2}={-2\ell_{1}\cdot u_{2}}, ρ3=22u1\rho_{3}={-2\ell_{2}\cdot u_{1}}, ρ5=12{\rho_{5}=\ell_{1}^{2}}, ρ6=22{\rho_{6}=\ell_{2}^{2}}, ρ8=(1q)2{\rho_{8}=(\ell_{1}-q)^{2}}, and ρ9=(2q)2{\rho_{9}=(\ell_{2}-q)^{2}}, while the underlined variables are used to denote the presence of delta functions; i.e., 2/ρ¯=1δ(1u1){2/\underline{\rho}{}_{1}=\text{\smash{\raisebox{2.58334pt}{--}}\kern-4.73611pt$\delta$}(\ell_{1}\cdot u_{1})}, 2/ρ¯=4δ(2u2){2/\underline{\rho}{}_{4}=\text{\smash{\raisebox{2.58334pt}{--}}\kern-4.73611pt$\delta$}(\ell_{2}\cdot u_{2})}, and 1/ρ¯=7δ+(k2){1/\underline{\rho}{}_{7}=\text{\smash{\raisebox{2.58334pt}{--}}\kern-4.73611pt$\delta$}_{+}(k^{2})} with kq12{k\equiv q-\ell_{1}-\ell_{2}}. All of the integrals in QμQ^{\mu} have n1=n4=n7=1n_{1}=n_{4}={n_{7}=1}. Reverse unitarity then allows us to treat these delta functions as cut propagators [91, 92, 93, 94].

The tensor-valued nature of these integrals make them cumbersome to evaluate as is, but fortunately they can all be reduced to scalar-valued ones via a suitable basis decomposition [85]. Specifically, we expand each of the loop momenta Aμ\ell_{A}^{\mu} (A{1,2}{A\in\{1,2\}}) in the numerator as

Aμ=(Au1)uˇ1+μ(Au2)uˇ2+μ(Aq)q2qμ+Aμ,\ell_{A}^{\mu}=(\ell_{A}\cdot u_{1})\,\check{u}_{1}{}^{\mu}+(\ell_{A}\cdot{u}_{2})\,\check{u}_{2}{}^{\mu}+\frac{(\ell_{A}\cdot q)}{q^{2}}q^{\mu}_{\mathstrut}+\ell_{A\bot}^{\mu}, (14)

where uˇ1μ:-(u1μγu2μ)/(1γ2){\check{u}_{1}^{\mu}\coloneq(u_{1}^{\mu}-\gamma u_{2}^{\mu})/(1-\gamma^{2})} and uˇ2μ:-uˇ1μ|12{\check{u}_{2}^{\mu}\!\coloneq\check{u}_{1}^{\mu}|_{1\leftrightarrow 2}} are the dual vectors to the two initial velocities (uˇAuB=δAB{\check{u}_{A}\cdot u_{B}=\delta_{AB}}), γu1u2{\gamma\equiv u_{1}\cdot u_{2}} is the Lorentz factor for the relative velocity vv, and Aμ\ell_{A\bot}^{\mu} is the part of Aμ\ell_{A}^{\mu} that is orthogonal to u1u_{1}, u2u_{2}, and qq [the delta functions in Eq. (12) guarantee that quA=0{q\cdot u_{A}=0}]. The three inner products in Eq. (14) are then easily rewritten in terms of the variables ρi\rho_{i} and q2q^{2} only.

As for Aμ\ell_{A\bot}^{\mu}, the fact that the denominator of Eq. (13) is invariant under (1,2)(1,2)(\ell_{1\bot},\ell_{2\bot})\mapsto-(\ell_{1\bot},\ell_{2\bot}) implies that any term in the numerator with ii powers of 1\ell_{1\bot} and jj powers of 2\ell_{2\bot} will integrate to zero if i+j{i+j} is odd. If instead i+j=2{i+j=2}, then rotational invariance on the hypersurface orthogonal to u1u_{1}, u2u_{2}, and qq allows us to replace

AμBν(AρρσBσ)d3μν\ell_{A\bot}^{\mathstrut\mu}\ell_{B\bot}^{\mathstrut\nu}\mapsto\frac{(\ell_{A}^{\mathstrut\rho}\bot_{\rho\sigma}\ell_{B}^{\mathstrut\sigma})}{d-3}\bot^{\mu\nu} (15)

under the integral, where the metric on this hypersurface is μν=ημνuˇ1u1μνuˇ2u2μνqμqν/q2{\bot^{\mu\nu}=\eta^{\mu\nu}-\check{u}_{1}{}^{\mu}u_{1}{}^{\nu}-\check{u}_{2}{}^{\mu}u_{2}{}^{\nu}-q^{\mu}q^{\nu}/q^{2}}, and note that the inner product (AρρσBσ){(\ell_{A}^{\mathstrut\rho}\bot_{\rho\sigma}\ell_{B}^{\mathstrut\sigma})} is easily rewritten solely in terms of the variables ρi\rho_{i}, q2q^{2}, and γ\gamma. Analogous replacement rules can be derived for the i+j=4{i+j=4} case by positing the ansatz AμBνCρDσc1μνρσ+c2μρνσ+c3μσρν\ell_{A\bot}^{\mathstrut\mu}\ell_{B\bot}^{\mathstrut\nu}\ell_{C\bot}^{\mathstrut\rho}\ell_{D\bot}^{\mathstrut\sigma}\mapsto{c_{1}\bot^{\mu\nu}\bot^{\rho\sigma}}+{c_{2}\bot^{\mu\rho}\bot^{\nu\sigma}}+{c_{3}\bot^{\mu\sigma}\bot^{\rho\nu}} and then solving for the coefficients {c1,c2,c3}{\{c_{1},c_{2},c_{3}\}} by taking appropriate contractions. The same can be done for all i+j2{i+j\in 2\mathbb{Z}}, although in practice we encounter only integrals with i+j5{i+j\leq 5}.

The object QμQ^{\mu} is now a sum of terms in which different combinations of qμq^{\mu}, uAμu_{A}^{\mu}, and sAμνs_{A}^{\mu\nu} are contracted with one another and multiplied by one of the scalar-valued integrals Gn1n9G_{n_{1}\cdots\,n_{9}}. At this stage, 31003100 different scalar integrals enter into QμQ^{\mu}, but not all of them are independent. After using the LiteRed software package [118, 119] to identify nontrivial integration by parts relations between the different integrals [95, 96, 97], we find that they reduce to a set of only four master integrals; the same four as in Eqs. (4.13)–(4.16) of Ref. [73]. These are solved via differential equation methods [98, 99, 100, 101, 102, 103]; see Ref. [73] for details.

All that remains is to compute the Fourier transform in Eq. (12). We do so by using another family of master integrals,

Inμ1μj:-qδ(qu1)δ(qu2)qμ1qμj(q2)neiqb,I^{\mu_{1}\cdots\,\mu_{j}}_{n}\coloneq\int_{q}\text{\smash{\raisebox{2.58334pt}{--}}\kern-4.73611pt{\hbox{\delta}}}(q\cdot u_{1})\text{\smash{\raisebox{2.58334pt}{--}}\kern-4.73611pt{\hbox{\delta}}}(q\cdot u_{2})\frac{q^{\mu_{1}}\cdots q^{\mu_{j}}}{(-q^{2})^{n}}e^{iq\cdot b}, (16)

whose scalar-valued member evaluates to

In=Γ(d/2n1)Γ(n)(bμ12μνbν)n+1d/24nπ(d2)/2(γ21)1/2,I_{n}=\frac{\Gamma(d/2-n-1)}{\Gamma(n)}\frac{(-b_{\mu}\bot_{12}^{\mu\nu}b_{\nu})^{n+1-d/2}}{4^{n}\pi^{(d-2)/2}(\gamma^{2}-1)^{1/2}}, (17)

with 12μν=ημνuˇ1μu1νuˇ2μu2ν{\bot_{12}^{\mu\nu}=\eta^{\mu\nu}_{\mathstrut}-\check{u}_{1}^{\mathstrut\mu}u_{1}^{\mathstrut\nu}-\check{u}_{2}^{\mathstrut\mu}u_{2}^{\mathstrut\nu}} [67]. Its tensor-valued cousins follow from differentiation; e.g., Inμ=iIn/bμ{I^{\mu}_{n}=-i\partial I_{n}/\partial b_{\mu}}.

II. esults

 Radiated four-momentum

Now specializing to four dimensions, it becomes convenient to decompose our final result into components along the basis vectors {u1,u2,b^,l^}\{u_{1},u_{2},\hat{b},\hat{l}\}, where b^μ:-bμ/b{\hat{b}^{\mu}\coloneq b^{\mu}/b} (bbμbμ){(b\equiv\sqrt{-b_{\mu}b^{\mu}})} and l^μ:-ϵμνρσu1νu2ρb^σ/γ21{\hat{l}_{\mu}\coloneq\epsilon_{\mu\nu\rho\sigma}u_{1}^{\mathstrut\nu}u_{2}^{\mathstrut\rho}\hat{b}^{\mathstrut\sigma}_{\mathstrut}/\sqrt{\gamma^{2}-1}} are the unit vectors pointing along the impact parameter and orbital angular momentum, respectively. After also eliminating the spin tensors in favor of the Pauli-Lubanski spin vectors sAμ:-ϵμuAννρσsAρσ/2{s_{A}^{\mu}\coloneq\epsilon^{\mu}_{\mathstrut}{}_{\nu\rho\sigma}u_{A}^{\mathstrut\nu}s_{A}^{\mathstrut\rho\sigma}/2}, we find that we can write

Pradμ=G3M4πν2b3(𝒞u1uˇ1μ+𝒞u2uˇ2μ𝒞b^b^μ𝒞l^l^μ).P^{\mu}_{\text{rad}}=\frac{G^{3}M^{4}\pi\nu^{2}}{b^{3}}\big{(}\mathcal{C}_{u_{1}}\check{u}_{1}^{\mu}+\mathcal{C}_{u_{2}}\check{u}_{2}^{\mu}-\mathcal{C}_{\hat{b}}\hat{b}^{\mu}-\mathcal{C}_{\hat{l}}\hat{l}^{\mu}\big{)}. (18)

The components 𝒞V\mathcal{C}_{V}, with V{u1,u2,b^,l^}{V\in\{u_{1},u_{2},\hat{b},\hat{l}\}}, are dimensionless functions of only the Lorentz factor γ\gamma, the two Wilson coefficients CEAC_{E_{A}}, and the six inner products (sAV)/b(s_{A}\cdot V)/b. (There are only six because sAuA=0{s_{A}\cdot u_{A}=0} by definition.)

Table 1: Functions of the Lorentz factor γ\gamma appearing in Eq. (19).
fIf_{\text{I}} 210γ6552γ5+339γ4912γ3+3148γ23336γ+115148(γ21)3/2\displaystyle\frac{210\gamma^{6}-552\gamma^{5}+339\gamma^{4}-912\gamma^{3}+3148\gamma^{2}-3336\gamma+1151}{48(\gamma^{2}-1)^{3/2}}
35γ4+60γ3150γ2+76γ58γ21log(1+γ2)\displaystyle-\frac{35\gamma^{4}+60\gamma^{3}-150\gamma^{2}+76\gamma-5}{8\sqrt{\gamma^{2}-1}}\log\!\left(\!\frac{1+\gamma}{2}\!\right)
+70γ7165γ5+112γ333γ16(γ21)2cosh1γ\displaystyle+\frac{70\gamma^{7}-165\gamma^{5}+112\gamma^{3}-33\gamma}{16(\gamma^{2}-1)^{2}}\cosh^{-1}\!\gamma
fIIf_{\text{II}} 210γ6356γ5111γ41627γ3+5393γ24741γ+135216(γ+1)(γ21)\displaystyle\frac{210\gamma^{6}-356\gamma^{5}-111\gamma^{4}-1627\gamma^{3}+5393\gamma^{2}-4741\gamma+1352}{16(\gamma+1)(\gamma^{2}-1)}
105γ4+345γ3405γ2+147γ488(γ+1)log(1+γ2)\displaystyle-\frac{105\gamma^{4}+345\gamma^{3}-405\gamma^{2}+147\gamma-48}{8(\gamma+1)}\log\!\left(\!\frac{1+\gamma}{2}\!\right)
+210γ6405γ4+135γ216(γ21)3/2cosh1γ\displaystyle+\frac{210\gamma^{6}-405\gamma^{4}+135\gamma^{2}}{16(\gamma^{2}-1)^{3/2}}\cosh^{-1}\!\gamma
fIIIf_{\text{III}} 210γ6279γ5219γ41350γ3+4732γ24243γ+124516(γ+1)(γ21)\displaystyle\frac{210\gamma^{6}-279\gamma^{5}-219\gamma^{4}-1350\gamma^{3}+4732\gamma^{2}-4243\gamma+1245}{16(\gamma+1)(\gamma^{2}-1)}
21γ4+66γ384γ2+30γ92(γ+1)log(1+γ2)\displaystyle-\frac{21\gamma^{4}+66\gamma^{3}-84\gamma^{2}+30\gamma-9}{2(\gamma+1)}\log\!\left(\!\frac{1+\gamma}{2}\!\right)
+42γ681γ4+27γ24(γ21)3/2cosh1γ\displaystyle+\frac{42\gamma^{6}-81\gamma^{4}+27\gamma^{2}}{4(\gamma^{2}-1)^{3/2}}\cosh^{-1}\!\gamma
fIVf_{\text{IV}} 425γ51215γ4+2491γ33957γ2+2992γ76016(γ+1)(γ21)2\displaystyle-\frac{425\gamma^{5}-1215\gamma^{4}+2491\gamma^{3}-3957\gamma^{2}+2992\gamma-760}{16(\gamma+1)(\gamma^{2}-1)^{2}}
84γ6+459γ5825γ4138γ3+666γ2321γ+758(γ+1)(γ21)2log(1+γ2)\displaystyle-\frac{84\gamma^{6}+459\gamma^{5}-825\gamma^{4}-138\gamma^{3}+666\gamma^{2}-321\gamma+75}{8(\gamma+1)(\gamma^{2}-1)^{2}}\log\!\left(\!\frac{1+\gamma}{2}\!\right)
+168γ7+78γ6414γ5171γ4+261γ3+81γ227γ16(γ+1)(γ21)5/2cosh1γ\displaystyle+\frac{168\gamma^{7}+78\gamma^{6}-414\gamma^{5}-171\gamma^{4}+261\gamma^{3}+81\gamma^{2}-27\gamma}{16(\gamma+1)(\gamma^{2}-1)^{5/2}}\cosh^{-1}\!\gamma

The fact that PradμP^{\mu}_{\text{rad}} is a polar vector strongly constrains which inner products can appear at any given order, and in which combinations. For instance, because 𝒞u1\mathcal{C}_{u_{1}}, 𝒞u2\mathcal{C}_{u_{2}}, and 𝒞b^\mathcal{C}_{\hat{b}} must all be even under parity, they can only depend on (sAl^)/b(s_{A}\cdot\hat{l})/b at linear order in the spins. Indeed, we find explicitly that

𝒞u1\displaystyle\mathcal{C}_{u_{1}} =fI(γ)+1b[(s1l^)fII(γ)+(s2l^)fIII(γ)]+O(s2),\displaystyle=f_{\text{I}}(\gamma)+\frac{1}{b}\big{[}(s_{1}\cdot\hat{l})f_{\text{II}}(\gamma)+(s_{2}\cdot\hat{l})f_{\text{III}}(\gamma)\big{]}+O(s^{2}),
𝒞l^\displaystyle\mathcal{C}_{\hat{l}} =1b[(s1u2)+(s2u1)]fIV(γ)+O(s2),\displaystyle=\frac{1}{b}\big{[}(s_{1}\cdot u_{2})+(s_{2}\cdot u_{1})\big{]}f_{\text{IV}}(\gamma)+O(s^{2}), (19)

while 𝒞b^=0+O(s2){\mathcal{C}_{\hat{b}}=0+O(s^{2})}. The remaining component 𝒞u2\mathcal{C}_{u_{2}} can be obtained from 𝒞u1\mathcal{C}_{u_{1}} by swapping the body labels 12{1\leftrightarrow 2}, since PradμP^{\mu}_{\text{rad}} must be symmetric under this interchange. As an added consequence, 𝒞b^\mathcal{C}_{\hat{b}} and 𝒞l^\mathcal{C}_{\hat{l}} must be odd and even under this interchange, respectively.

The four functions {fI,,fIV}{\{f_{\text{I}},\,\dots,f_{\text{IV}}\}} in Eq. (19) depend purely on γ\gamma and are presented in Table 1 [fIf_{\text{I}}, which appears at O(s0)O(s^{0}), was previously determined in Refs. [70, 71, 72, 73], but is reproduced here for completeness]. An additional 21 functions of γ\gamma, with similar analytic structures, appear at O(s2)O(s^{2}). These are presented in the Supplemental Material [104]. Notice that 𝒞b^\mathcal{C}_{\hat{b}} and 𝒞l^\mathcal{C}_{\hat{l}} both vanish when the spins are aligned along l^\hat{l} (or, indeed, when they are zero); hence, for so-called “aligned-spin” configurations, for which the binary’s motion is confined to a plane, we see that momentum is lost only in the direction of the relative velocity.

 Consistency checks

To validate Eq. (18) against the existing literature, we compare results for the energy ΔE\Delta E radiated in the center-of-mass frame. This is computed in our approach as ΔE=(Pradptot)/E{\Delta E=(P_{\text{rad}}\cdot p_{\text{tot}})/E}, where ptotμ=m1u1μ+m2u2μ{p^{\mu}_{\text{tot}}=m_{1}^{\mathstrut}u_{1}^{\mu}+m_{2}^{\mathstrut}u_{2}^{\mu}} and E2=ptot2=M2[1+2ν(γ1)]{E^{2}=p_{\text{tot}}^{2}=M^{2}[1+2\nu(\gamma-1)]}. Since b^μ\hat{b}^{\mu} and l^μ\hat{l}^{\mu} are purely spatial in this frame, sAb^{s_{A}\cdot\hat{b}} and sAl^{s_{A}\cdot\hat{l}} are equivalent to the three-dimensional dot products 𝐬A^𝐛{-\,\mathbf{s}_{A}\cdot\hat{\vphantom{b}}\kern-1.4pt\mathbf{b}} and 𝐬A𝐥^{-\,\mathbf{s}_{A}\cdot\hat{\mathbf{l}}}, respectively, and note that s1u2𝐬1𝐯{s_{1}\cdot u_{2}\simeq\mathbf{s}_{1}\cdot\mathbf{v}} while s2u1𝐬2𝐯{s_{2}\cdot u_{1}\simeq-\,\mathbf{s}_{2}\cdot\mathbf{v}} after expanding to first order in the relative 3-velocity 𝐯\mathbf{v}. Having done so, our result for ΔE\Delta E agrees with that of Ref. [85], which is accurate to leading order in 𝐯\mathbf{v} and to quadratic order in the spins, once we also replace (𝐛,𝐬A,CEA)(𝐛,𝐬A,1CEA)(\mathbf{b},\mathbf{s}_{A},C_{E_{A}})\mapsto(-\mathbf{b},-\mathbf{s}_{A},1-C_{E_{A}}) to account for differing conventions.

As a second consistency check, we use analytic continuation by way of the boundary-to-bound (B2B) map [15, 16, 17] to convert our result for ΔE\Delta E into the energy ΔEell\Delta E_{\text{ell}} radiated during one period of ellipticlike motion. This is accomplished in three steps. Owing to current limitations of the B2B map, we first specialize to aligned-spin configurations. Next, we must transform from the covariant SSC to the canonical (Newton-Wigner) SSC [110] for the map to work. This generally entails transforming (𝐛,𝐬A)(\mathbf{b},\mathbf{s}_{A}) to new canonical variables (𝐛c,𝐬Ac)(\mathbf{b}_{\text{c}},\mathbf{s}_{A\,\text{c}}) [59, 60], but 𝐬A𝐬Ac{\mathbf{s}_{A}\equiv\mathbf{s}_{A\,\text{c}}} in the aligned-spin case; hence, only the magnitude of the impact parameter must be transformed. The rule is

bp=bcpEM2E[Ea+(m1m2)a],bp_{\infty}=b_{\text{c}}p_{\infty}-\frac{E-M}{2E}\big{[}Ea_{+}-(m_{1}-m_{2})a_{-}\big{]}, (20)

where p=M2νγ21/E{p_{\infty}=M^{2}\nu\sqrt{\gamma^{2}-1}/E} is the initial momentum of either body in the center-of-mass frame, a±=(𝐬1±𝐬2)𝐥^{a_{\pm}=(\mathbf{s}_{1}\!\pm\mathbf{s}_{2})\cdot\hat{\mathbf{l}}}, and we may define Lc=bcp{L_{\text{c}}=b_{\text{c}}p_{\infty}} as the canonical orbital angular momentum. Finally, we obtain ΔEell\Delta E_{\text{ell}} from ΔE\Delta E via [17]

ΔEell(,Lc,a±)=ΔE(,Lc,a±)ΔE(,Lc,a±),\Delta E_{\text{ell}}(\mathcal{E},L_{\text{c}},a_{\pm})=\Delta E(\mathcal{E},L_{\text{c}},a_{\pm})-\Delta E(\mathcal{E},-L_{\text{c}},-a_{\pm}), (21)

having eliminated γ\gamma in favor of (EM)/(Mν){\mathcal{E}\equiv(E-M)/(M\nu)}. The left-hand side follows after analytic continuation from positive to negative values of \mathcal{E}. Expanded in powers of \mathcal{E}, we find that our result, which is valid in the large-angular-momentum limit [17], agrees with the overlapping terms from PN theory up to 3PN in Ref. [17], and up to 4PN in Refs. [120, 121].

V. onclusion

The worldline EFT approach has proven to be an efficient way of obtaining classical radiated observables from the scattering encounter of two compact objects, be they Schwarzschild or Kerr black holes, or even neutron stars. This work closes an important gap in the PM literature by computing the radiated four-momentum at 3PM up to quadratic order in the spins and to all orders in the velocity. Remarkably, integrating over the loop momenta required knowledge of only four master integrals—the same four as in the nonspinning case—which explains why the analytic structure of our result is similar to that of Refs. [70, 71, 72, 73, 74], despite the inclusion of spins. At low velocities, our radiated energy is consistent with the existing literature, including the case of the energy loss from a bound system during a single orbit, which we derived via analytic continuation.

These results should prove invaluable for making further consistency checks in the future. We expect, for example, that our 3PM result for the radiated energy should reemerge in the tail term of the (as yet unknown) conservative potential for spinning binaries at 4PM, analogously to how the tail term [41, 42, 47] was found to match the radiated energy [70, 71, 72, 73] in the nonspinning case. In the future, it would also be interesting to reconstruct the radiated flux for bound systems (via the approach in Ref. [17]) from our calculation of the energy loss, so as to make a more direct comparison with PN results, since it is the former that directly impacts the binary’s inspiral via the balance equation [13].

Acknowledgements.
It is a pleasure to thank Gihyuk Cho, Gregor Kälin, and Rafael Porto for providing us with their result for the radiated energy up to 4PN. We acknowledge use of the xAct package [122] for Mathematica in our calculations. This work was partially supported by the Centre National d’Études Spatiales (CNES).
Note added.—While this manuscript was undergoing peer review, we became aware of similar calculations being undertaken by Gustav Jakobsen and Gustav Mogull. We thank them for verifying that their result for the radiated four-momentum is in agreement with ours.

References