<<

KCL-PH-TH/2021-04 CERN-TH-2021-016 Nonlinear gravitational-wave memory from cusps and kinks on cosmic strings

1, 1, 2, Alexander C. Jenkins ∗ and Mairi Sakellariadou † 1Theoretical Particle and Cosmology Group, Physics Department, King’s College London, University of London, Strand, London WC2R 2LS, UK 2Theoretical Physics Department, CERN, Geneva, (Dated: 23 July 2021) The nonlinear memory effect is a fascinating prediction of (GR), in which oscillatory gravitational-wave (GW) signals are generically accompanied by a monotonically-increasing strain which persists in the detector long after the signal has passed. This effect is directly accessible to GW observatories, and presents a unique opportunity to test GR in the dynamical and nonlinear regime. In this article we calculate, for the first time, the nonlinear memory signal associated with GW bursts from cusps and kinks on cosmic string loops, which are an important target for current and future GW observatories. We obtain analytical waveforms for the GW memory from cusps and kinks, and use these to calculate the ‘memory of the memory’ and other higher-order memory effects. These are among the first memory observables computed for a cosmological source of GWs, with previous literature having focused almost entirely on astrophysical sources. Surprisingly, we find that the cusp GW signal diverges for sufficiently large loops, and argue that the most plausible explanation for this divergence is a breakdown in the weak-field treatment of GW emission from the cusp. This shows that previously-neglected strong gravity effects must play an important rôle near cusps, although the exact mechanism by which they cure the divergence is not currently understood. We show that one possible resolution is for these cusps to collapse to form primordial black holes (PBHs); the kink memory signal does not diverge, in agreement with the fact that kinks are not predicted to form PBHs. Finally, we investigate the prospects for detecting memory from cusps and kinks with current and future GW observatories, considering both individual memory bursts and the contribution of many such bursts to the stochastic GW background. We find that in the scenario where the cusp memory divergence is cured by PBH formation, the memory signal is strongly suppressed and is not likely to be detected. However, alternative resolutions of the cusp divergence may in principle lead to much more favourable observational prospects.

CONTENTS B. Late-time memory 15 C. Full waveform 15 I. Introduction2 D. Time-domain waveform near the arrival time 15 E. Radiated energy 16 II. Nonlinear GW memory3 F. Higher-order memory 16 A. Late-time memory4 G. Caveats of our approach 17 B. Frequency-domain memory waveforms4 VI. Detection prospects 17 III. Cosmic string burst waveforms5 A. Burst searches 17 IV. Memory from cusps5 B. Stochastic background searches 18 A. Beaming effects5 C. Consequences for the loop distribution B. Late-time memory6 function 19 C. Full waveform6 VII. Summary and conclusion 19 arXiv:2102.12487v2 [gr-qc] 22 Jul 2021 D. Time-domain waveform near the arrival time7 E. Radiated energy7 F. Ultraviolet divergence of the radiated energy8 Acknowledgments 20 G. Second-order memory9 H. Higher-order memory, and another A. Understanding the origin of the higher-order divergence 10 memory divergence 20 I. Memory from cusp collapse 11 1. Gaussian pulse as a toy model 20 2. Application to cosmic strings 21 V. Memory from kinks 14 3. Application to compact binaries 21 A. Beaming effects 15 B. Angular integrals for higher-order cusp memory 22

C. Angular integrals for kink memory 23 ∗ [email protected][email protected] References 25 2

I. INTRODUCTION to the appeal of nonlinear GW memory as an observational probe of the underlying gravitational theory. The advent of gravitational-wave (GW) astronomy has With these motivations in mind, there has been a sig- given us unprecedented observational access to gravity in nificant effort in recent years to calculate [6, 34, 45–51] the dynamical, nonlinear regime, allowing us to test the and search for [52–68] memory waveforms associated with predictions of Einstein’s General Relativity (GR) in this CBCs, as these are the primary target of GW observat- regime as never before [1–3]. One such prediction is the ories like LIGO/Virgo and pulsar timing arrays (PTAs), GW memory effect [4–6], in which the passage of a GW and are predicted to be abundant sources of nonlinear signal causes a permanent offset in the distance between memory. two nearby freely-falling test masses; i.e., the system In this article, we focus instead on the nonlinear memory ‘remembers’ the passage of the GW signal, rather than generated by another key GW source: cosmic strings [69– just oscillating and returning to its initial configuration. 72]—linelike topological defects which may have formed in Effects like this are a generic prediction for any GW a cosmological phase transition in the early Universe due signal whose source has unbound components that es- to a spontaneously-broken U(1) symmetry, whose produc- cape to infinity—for example, hyperbolic binary encoun- tion is a generic prediction of many theories beyond the ters [7–12], core-collapse supernovae [13–17], gamma-ray Standard Model [73]. These strings self-intersect to form bursts [18–21], and particle decay [22–24]. The resulting closed loops [74, 75], which oscillate at relativistic speeds displacement is then called the linear memory, as it is and emit strong GW bursts through sharp features called already present in the GW signal obtained by solving cusps and kinks [76], making them important targets for the linearised Einstein equation. Gravitationally-bound GW searches. The amplitude of these GW signals is set by the dimensionless string tension Gµ, which is related to systems, such as the compact binary coalescences (CBCs) 2 the symmetry breaking scale η by Gµ (η/mPl) , with observed by LIGO/Virgo [25–28], do not source significant ∼ amounts of linear GW memory.1 mPl the Planck mass. Searches for cosmic strings with LIGO/Virgo [77–81] and pulsar timing arrays (PTAs) [82– In a now-classic paper, Christodoulou [31] used rig- 84] have so far returned only null results,2 allowing us orous asymptotic methods to show that there is also a to place a conservative upper limit on the string tension nonlinear memory effect, in which essentially any GW of 11 (there is some uncertainty due to the signal is accompanied by a permanent strain offset (the Gµ . 10− modelling of the cosmic string loop network—the most same effect was also discovered soon after by Blanchet stringent constraints are at the level of 15 [81]). and Damour [32] in the very different context of post- Gµ . 10− There are at least two reasons to expect a priori that Minkowskian theory). Intuitively, we can understand this cosmic strings could be an important source of nonlinear by appreciating that linearised GWs carry energy and memory.3 momentum, and their effective stress-energy can there- fore source further perturbations to the spacetime metric 1. GW memory waveforms tend to be associated with beyond linear order. By heuristically treating the grav- lower frequencies than the primary GW emission itons sourced at linear order as GW sources in their own that sources them [6]. This is because the memory right, Thorne [33] showed soon after Christodoulou’s pa- grows monotonically over a period of order the dur- per that the nonlinear memory can be interpreted as the ation of the primary signal, which is typically much cumulative linear memory associated with these grav- itons escaping to null infinity. By searching for nonlinear memory with GW observatories, we can therefore directly 2 The NANOGrav Collaboration have recently found evidence for a probe the ‘ability of gravity to gravitate’ (in the words of stochastic process in the pulsar timing residuals of their 12.5-year reference [34]). dataset, with a common spectrum across all pulsars [85]. Several While originally derived in the context of classical grav- authors have pointed out that this is consistent with a stochastic itational physics, it has since been recognised that effects GW background from cosmic strings [86–90]. While this is very exciting, we note that the NANOGrav data do not yet provide analogous to the nonlinear GW memory are a generic conclusive evidence for the quadrupolar cross-correlations between feature of field theories with massless degrees of free- pulsars that one would expect from a GW signal, and further data dom [22, 23, 35–37], and that these memory effects are are thus needed to confidently identify the cause of the common intimately related to both the asymptotic symmetry group stochastic process. If the observed signal is indeed due to GWs, of spacetime and so-called ‘soft theorems’ which govern it is perfectly consistent with an astrophysical background from inspiralling supermassive black hole binaries, meaning that more the production of low-energy massless particles in the exotic interpretations must be explored with caution. quantum theory [38–44]. These deep theoretical links add 3 We note that cosmic string loops can also source significant amounts of linear memory by emitting radiation in the under- lying matter fields. These effects are ignored by the Nambu- Goto approximation we adopt here, and can only be resolved by field-theory simulations. In reality, we expect loops to gener- 1 CBCs do generate some linear memory if the final black hole has ate a combination of linear and nonlinear memory, by radiating a non-zero recoil velocity, but this linear memory signal is likely both matter and GWs. This has recently been demonstrated for far too weak to be detected [29], even for systems with maximal collapsing circular loops, using numerical-relativity/field-theory recoil velocities (the so-called ‘super-kick’ configuration) [30]. simulations [91]. 3

longer than the oscillation period of the primary our analysis. In sectionIV, we calculate the nonlinear signal. This means that, e.g., massive stellar binary GW memory signal associated with cusps, obtaining the black holes which merge near the bottom end of the simple frequency-domain waveform (21); we show that the LIGO/Virgo frequency band produce memory sig- total GW energy radiated by this memory signal diverges nals which are shifted to lower frequencies, and are for Nambu-Goto strings, and regularise this divergence thus challenging to detect with LIGO/Virgo. Cos- by imposing a cutoff at the scale of the string width δ; mic string signals, on the other hand, have durations we then go on to consider higher-order memory effects which are comparable to their oscillation period (i.e. (the nonlinear GW memory sourced by the memory GWs they have a ‘burst-like’ morphology, which is very themselves) and show that accounting for all such con- different to an inspiralling compact binary), mean- tributions leads to a divergence for large loops, which ing that we should expect the resulting memory persists even after applying the string-width regularisa- signals to have power at similar frequencies to the tion; finally, we show that this divergence is cured if the original signal. What’s more, the cosmic string sig- cusp collapses to form a primordial black hole (PBH), as nals we consider also have significant power at very we recently proposed in reference [93]. In sectionV we high frequencies, and it is possible that the heredit- repeat the memory calculation for kinks, obtaining the ary nature of the memory effect could transfer some leading-order waveform (70); unlike in the cusp case, the of this power, enhancing the amplitude of the signal memory signal is strongly suppressed at high frequencies at observable frequencies. (A similar idea of ‘orphan’ due to interference effects, and no divergence occurs. In signals, where the memory emission is detectable sectionVI we study the observable consequences of our even though the primary signal is not, was studied memory waveforms for GW searches, under the assump- in reference [92].) tion that the cusp memory divergence is cured by PBH formation; we find that in this scenario the memory is 2. The angular pattern of the GW memory signal on strongly suppressed, and is beyond the reach of current or the sky is typically different to that of the primary planned GW searches. Finally, we summarise our results GW emission that sources it. Cosmic string cusps in sectionVII. We discuss the cause of the higher-order and kinks emit GWs in narrow beams, meaning that memory divergence in appendixA, and give some tech- only a very small fraction of all cusps and kinks nical details of the angular distribution of the memory are oriented such that their GWs can be observed. GWs from cusps and kinks in appendicesB andC. We However, if the associated GW memory signal is set c = 1 throughout, but keep G and ~ explicit. more broadly distributed on the sphere, this might allow us to observe some of the many cusps and kinks whose beams are not oriented towards us. II. NONLINEAR GW MEMORY We find that both expectations are borne out by our cal- We work in terms of the complex GW strain, culations below: the GW memory from cusps and kinks is indeed emitted in a much broader range of directions 1 + h(t, r) h+ ih = (eij ieij×)hij, (1) than the initial beam, and the memory signal does indeed ≡ − × 2 − have a similar frequency profile to the primary signal, with the high-frequency behaviour of the primary GWs where t is the retarded time, r is a 3-vector point- A playing an important rôle in determining the strength ing from the source to the observer, and eij are the of the memory effect. In fact, we show that these two transverse-traceless (TT) polarisation tensors. We dis- ingredients lead to a divergence in the memory signal tinguish between the primary, oscillatory strain signal from cusps on sufficiently large loops, potentially signi- (sourced at linear order) and the additional strain due (0) (1) fying a breakdown of the standard weak-field approach to the memory effect by writing these as h and h respectively (reserving h(n) with n 2 for the ‘memory for calculating the GW signal from cusps. We attempt to ≥ clarify the root cause of this breakdown, and suggest one of the memory’ and other higher-order memory contribu- possible resolution, based on the cusp-collapse scenario tions, which we discuss in sectionsIVG andVF). The of reference [93]. Other resolutions are possible however, leading nonlinear memory correction term can be written and ultimately a fully general-relativistic treatment will in gauge-invariant form as [6, 33] be required to understand the true behaviour of cusps. Z t Z + 3 (0) (1) 2G 2 (eij ieij×)ˆri0rˆj0 d Egw The remainder of this article is structured as follows. r rˆ0 − h (t, ) = dt0 d 2 . In sectionII we introduce the standard expression for r S2 1 rˆ · rˆ0 dt0 d rˆ0 −∞ − the nonlinear GW memory, equation (2), and develop (2) from it some useful formulae for calculating the late-time We see immediately that since the integral is over the (non- 3 (0) 2 memory (i.e. the total strain offset after the GW sig- negative) GW energy flux from the source, d Egw / dt d rˆ, nal has passed), equation (9), and the frequency-domain we generically obtain a nonzero memory correction which memory waveform, equation (12). In sectionIII we briefly grows monotonically with time while the source is ‘on’. recap the standard cusp and kink waveforms as derived By projecting onto the GW polarisation tensors we ensure in reference [76], which are the other key ingredient of that only the TT part of the angular integral contributes; 4 this TT part vanishes if the emitted flux is exactly iso- This problem disappears when we consider the late-time tropic, but is generically non-vanishing for anisotropic memory, i.e. the total strain offset caused by the cusp, emission. Note that equation (2) corresponds to just one of many ∆h(1) lim h(1)(t), (8) t quadratic terms that appear on the right-hand side of the ≡ →∞ relaxed Einstein field equation [46]. While this term has as the time integral in equation (6) is then over the entire received particular attention in the literature due to its real line, and we can thus use Parseval’s theorem to write hereditary nature and its pleasing intuitive interpretation as the nonlinear gravitational counterpart to the linear 2 Z Z (1) 2π (0) 2 GW memory effect, the other quadratic terms will also ∆h (r) = df rfh˜ (f, r0) , (9) r 0 | | give nonlinear corrections to the GWs emitted by cosmic R rˆ strings. Our results here therefore represent just one ˜(0) particular nonlinear contribution to these GW signals. where h is the Fourier transform of the primary strain Indeed, since we argue below that nonlinear effects are signal, and the extra factor of frequency comes from the important near cusps on large cosmic string loops, it would time derivative on the strain. be interesting to explore these additional contributions further in future work. Returning to equation (2), we write the polarisation B. Frequency-domain memory waveforms tensors as [94] The late-time memory (9) is useful for giving a sense of + e = θˆiθˆj φˆiφˆj, e× = θˆiφˆj + φˆiθˆj, (3) the total size of the memory effect, but in many cases is ij − ij not directly observable. For example, the test masses in where θˆ and φˆ are the standard spherical polar unit vec- ground-based GW interferometers like LIGO and Virgo tors orthogonal to rˆ. It is then straightforward to show are not freely-falling in the plane of the interferometer that arms; they are acted upon by feedback control systems at low frequencies to mitigate seismic noise [97]. These + ˆ ˆ 2 (e ie×)ˆr0rˆ0 = [(θ iφ) · rˆ0] . (4) low-frequency forces mean that the test masses cannot sus- ij − ij i j − tain a permanent displacement after the GW has passed. We can also rewrite the energy flux (2) in terms of the However, the ‘ramping up’ of the memory signal from linear GW strain of the source, using the Isaacson for- zero at early times to ∆h(1) at late times can be measured mula [95, 96] if it contains power in the sensitive frequency band of the interferometer. 3 2 d Egw rh˙ = h| | i, (5) We are therefore interested in calculating the full dt d2rˆ 16πG memory signal in the frequency domain. The simplest way of doing this is to use the result4 where the dot denotes a time derivative. The leading GW memory term then becomes Z t  1 i dt0 g(t0) = g˜(0)δ(f) g˜(f) (10) Z t Z 2 2πf (1) dt0 (0) 2 F − h (t, r) = rh˙ (t0, r0) , (6) −∞ 2r rˆ0 | | −∞ for a general function g(t), where [ ] denotes the Four- F ··· where we have introduced the shorthand ier transform, and g˜ = [g]. Applying this to equation (6), we obtain F 2 2 Z Z d rˆ0 [(θˆ iφˆ) · rˆ0] [ ] [ ], (7) Z Z − (1) 1 (1) i dt 2πift (0) 2 rˆ0 ··· ≡ S2 4π 1 rˆ · rˆ0 ··· h˜ (f) = ∆h δ(f) e− rh˙ . − π 0 2 − 2 f R 2r rˆ | | for brevity. The temporal average is necessary in (11) equation (5) to give a gauge-invarianth· · ·i result, but is re- Note that we can neglect the term proportional to δ(f), as moved due to the time integral in equation (6). Through- this only contributes at f = 0, and we are interested here out we refer to the oscillatory strain h(0) on the RHS that in frequencies which are accessible to GW experiments sources the memory as the ‘primary’ GW emission. [this zero-frequency term is still captured in equation (9)]. By replacing each factor of h˙ (0) with its Fourier transform

A. Late-time memory

One drawback of equation (6) is that it is written in 4 R t 0 0 One can show this by noting that −∞ dt g(t ) is just the con- terms of the time-domain primary strain, whereas the volution of g(t) with the Heaviside step function Θ(t). Since cosmic string waveforms that we want to consider are F[Θ] = (1/2)δ(f) − i/(2πf), the result follows from the convolu- much more naturally expressed in the frequency domain. tion theorem, F[Θ ∗ g] = F[Θ]F[g]. 5

and massaging the resulting expression, we find The waveforms are nonzero only for frequencies above the base mode of the loop, 2/`, and only if the observer lies h˜(1)(f) = inside a small beam with opening angle Z Z iπ 2 (0) (0) df 0 f 0(f 0 f)r h˜ (f 0, rˆ0)h˜ ∗(f 0 f, rˆ0). 22/3 0 1/3 − rf R rˆ − − θb(f) ( f `)− . (15) (12) ' 31/6 | | For cusps, there is a single well-defined beaming direction This is the simplest way of calculating the frequency- rˆ , whereas for kinks the beam is over a one-dimensional domain memory using only the primary frequency-domain c ˜(0) ‘fan’ of directions on the sphere, and rˆk represents the signal h . direction in this fan which is closest to the observer’s line of sight rˆ. III. COSMIC STRING BURST WAVEFORMS IV. MEMORY FROM CUSPS We consider cosmic string loops in the Nambu-Goto (i.e. zero-width) approximation (although at some points We begin by calculating the nonlinear memory from below it is necessary to reintroduce a finite string width δ). cusps, inserting the primary waveform (13a) into equa- The Nambu-Goto equations of motion imply that these tions (9) and (12) to obtain the late-time memory and loops should generically develop sharp features called frequency-domain waveform, and then iterating this pro- ‘cusps’, where the tangent vector of a point on the loop cess to obtain higher-order memory corrections. instantaneously becomes null, generating a strong burst of GWs. String intercommutations also generically lead to discontinuities in the loop’s tangent vector called ‘kinks’. A. Beaming effects Nambu-Goto loops oscillate at a frequency 2/` inversely related to their invariant length `, with their motion sourcing GWs at integer multiples of this base frequency. The anisotropic beaming of the GWs from cusps is what Since this frequency is typically many orders of magnitude gives rise to a nonzero memory effect (due to its nonzero below the frequency band relevant to ground-based inter- TT projection), and is captured in the spherical integral ferometers like LIGO/Virgo, we are typically interested Z in very high-order harmonics of the loops. In this high- Θ(rˆc · rˆ0 cos θb), (16) frequency regime, the GW emission of cosmic string loops rˆ0 − is dominated by cusps and kinks. (High-frequency GWs where we are using the shorthand (7). To compute this can also be produced through kink-kink collisions [98], integral, it is convenient to define polar coordinates rˆ0 = but this emission mode is exactly isotropic and so does (θ0, φ0) such that the North pole θ0 = 0 coincides with not contribute to the memory effect.) The asymptotic the centre of the beam rˆ . The integrand then only has high-frequency waveforms for cusps and kinks on a loop c support for θ0 [0, θb], so that equation (16) becomes with dimensionless tension Gµ and invariant length ` can ∈ be approximated by [76] Z Θ(rˆc · rˆ0 cos θb) 2/3 rˆ0 − (0) Gµ` ˜ rˆ rˆ · rˆ Z θb Z 2π ˆ ˆ 2 hc (f, ) Ac 4/3 Θ( c cos θb)Θ( f 2/`), sin θ0 dφ0 [(θ iφ) · rˆ0] ' r f − | | − = dθ0 − (17) | | (13a) 0 2 0 2π 1 rˆ · rˆ0 4 ι  −  1/3 cos 1 2 (0) Gµ` 2 ˜ = 2 2 (1 cos θb) cos θb cos ι sin θb , hk (f, rˆ) Ak Θ(rˆ · rˆk cos θb)Θ( f 2/`), 2 ' r f 5/3 − | | − sin ι − − | | (13b) 1 where ι cos− rˆc · rˆ is the inclination of the beam to ≡ where the dimensionless pre-factors are the observer’s line of sight. In the high-frequency regime where the primary cusp and kink waveforms are valid, the 2/3 beam angle is very small, so we expand equation (17) to 8 2 leading order in θ to obtain Ac 2 0.8507, b ≡ Γ (1/3) 3 ≈ (14) Z 2 √Ac θb Ak 0.2936. Θ(rˆc · rˆ0 cos θb) (1 + cos ι). (18) ≡ π ≈ rˆ0 − ' 4 The fact that the frequency-domain waveforms (13a) There are a few remarks worth making about equa- and (13b) are real and even around f = 0 implies that tion (18). First, we note that it assumes ι θb; when the corresponding time-domain waveforms are real, and instead the inclination is much smaller than the beaming therefore that the cusps and kinks are linearly polarised. angle, ι θb, the integral (16) drops to zero. The fact  6

f 0 < f 0 f f 0 > f 0 f | | | − | | | | − | + −∞ ∞

0 2/` f/2 f 2/` f 0 ± ±

(1) 0 Figure 1. Schematic illustration of the different contributions to h˜c (f) from the integral over f in equation (12). By introducing a dimensionless dummy variable u ≡ f 0/|f| we obtain the two integrals shown in equation (20), one corresponding to the finite interior region 2/` < f 0 < f − 2/`, and the other corresponding to the two semi-infinite exterior regions f 0 < −2/` and f 0 > f + 2/`. that the result (18) is purely real, despite the integrand into equation (12). In doing so, we must be careful to being complex, shows that the memory strain is linearly correctly account for the behaviour of the two frequency polarised, much like the primary cusp and kink waveforms. arguments f 0 and f 0 f; in particular, the integrand is 2 − Geometrically, the θb/4 factor represents the fraction of only nonzero when both of these arguments have mag- the sphere taken up by the beam, while the (1 + cos ι) nitude greater than 2/`, and the size of the beam angle factor shows how the strength of the memory effect varies θb must always be set by whichever of the two arguments with inclination. In particular, we notice that the memory has greater magnitude (as this corresponds to a smaller, strain is nonzero when the observer lies outside of the more restrictive beam). These different contributions are beam, ι > θb. In fact, the observed memory strain van- illustrated in figure1 for the case where f is positive. ishes only when the beam is face-on (ι θb) or face-off Assuming that f > 4/`, we find (ι = π), as in both cases the angular pattern of the primary | | (1) GW emission is isotropic around the line of sight. i∆hc h˜(1)(f) = It is interesting to note that this angular pattern—a c −22/33π`1/3f f 1/3 broad distribution, except at very small | | (1 + cos ι) "Z Z 1/2 # inclinations∼ where the memory signal drops to zero—is ∞ du du 1/3 1/3 , exactly the same as that of the linear GW memory gener- × 2/ f ` u (1 + u) − 2/ f ` u (1 u) | | | | − ated by the ejection of an ultrarelativistic ‘blob’ of matter (20) from a massive object along a fixed axis [18]. Intuitively, this makes complete sense: the setup here is essentially where u is just a dimensionless dummy variable. We the same, except that the ‘blob’ is replaced by a burst of can simplify this by taking f 4/`, as this is the | |  GWs. regime where the primary waveforms are valid; in this high-frequency limit, both integrals can be evaluated ana- lytically. Inserting equation (19), the final result is B. Late-time memory 2 2/3 ˜(1) (Gµ) ` hc (f) iBc (1 + cos ι)Θ( f 2/`), (21) Now that we have computed the angular integral (16), it ' − rf f 1/3 | | − | | is straightforward to obtain the late-time memory from the with a numerical prefactor, cusp. Inserting equations (13a) and (15) into equation (9) and integrating over frequency, we find   2 2/3 4π 2 2 5 Bc πA (3/2) 2F1( , ; ; 1) ≡ c √ − 3 3 3 − (22) (1) 2/3 2 ` 3 3 ∆h = 2 3 (πAcGµ) (1 + cos ι) . (19) c × r 4.764, ≈ Inserting the numerical factors, this has a maximum value where 2F1 is a hypergeometric function. While we as- (1) 2 of ∆hc 59.43 (Gµ) `/r for nearly-face-on cusps ι & 0, sumed f > 4/` in order to obtain equation (20), we and smoothly≈ tapers× to zero for face-off cusps ι = π. We have checked| | that our final expression (21) underestim- note that equation (19) should be taken with a pinch of ates the true memory signal in the region 2/` < f < 4/` salt, as it is sensitive to the low-frequency regime where (assuming the primary signal is accurate at these| frequen-| the primary waveform is less accurate; nonetheless, we cies, which we expect to be the case to within an order expect this to give a reasonable estimate of the magnitude of magnitude if one ignores the low-frequency motion of the memory effect. not associated with the cusp), so we can safely leave the low-frequency cutoff at 2/` as in the primary waveform. The simple expression (21) thus gives a conservative but C. Full waveform generally accurate model of the time-varying part of the cusp memory waveform at all frequencies greater that the We now calculate the full frequency-domain cusp fundamental mode of the loop, while the zero-frequency memory waveform by inserting equations (13a) and (15) offset is described by equation (19). 7

equation (21) to obtain a simple closed-form expression in the time domain,

2 (1) (1) 5/3 (Gµ) h (t) h (t0) 2 3πBc (1 + cos ι) c − c ' − r  Γ (2/3)  (t t0) 1 2/3 , × − − (4π t t0 /`) | − | (23)

where Γ (z) is Euler’s Gamma function, and Γ (2/3) 1.354. (Here we have re-introduced the time of arrival of≈ the primary cusp signal, t0, rather than setting it to zero.) This time-domain waveform is shown in figure3. Note that this is real, which means that the signal is linearly +-polarised, just like the primary waveform h(0). It is important to note that since equation (21) is only valid for high frequencies f 2/`, equation (23) must  only be valid for a short duration t t0 ` around the arrival time. However, for typical| − loop|  sizes this ‘short duration’ is actually much longer than the relevant observational timescale (typically a few seconds), so the approximation in equation (23) may be a useful one. Figure 2. A cartoon illustration of the angular distribution of the energy radiated by a cusp. The primary emission (blue) is −1/3 concentrated in a narrow beam of width θb ∼ (f`) , while E. Radiated energy the first-order memory emission (violet) is proportional to (1 + cos ι)2, and the second-order memory emission (red) is ˜ proportional to sin4 ι. Given a frequency-domain waveform h(f), we can cal- culate the corresponding (one-sided, logarithmic) GW energy flux spectrum, Comparing equation (21) with the primary cusp wave- d3E πr2f 3   form (13a), we see that they are remarkably similar to gw = h˜(f) 2 + h˜( f) 2 . (24) each other, with both being given by the same simple d(ln f) d2rˆ 4G | | | − | frequency power law f 4/3 and the same dependence − In the context of radiation from a cosmic string loop, it is on the loop length ` (the∼ latter being required by dimen- helpful to normalise this with respect to the total energy of sional arguments). There are, however, some important the loop, µ`. We therefore define a dimensionless energy differences: spectrum,

1. The memory GW emission has broad support on 3 the sphere, while the primary waveform only has 1 d Egw (f, rˆ) 2 . (25) support inside a narrow beam (see figure2). ≡ µ` d(ln f) d rˆ

2. The memory waveform is suppressed by an addi- For the primary cusp waveform (13a), this is given by tional power of Gµ (this makes intuitive sense, given (0) π 2 1/3 that it is a nonlinear effect sourced by the primary  (f, rˆ) A Gµ(f`) Θ(θb ι)Θ(f 2/`), (26) c ' 2 c − − GW emission). while for the cusp memory (21), we find 3. The numerical constant in front of the memory wave- form, Bc, is an order of magnitude larger than that (1) π 2 3 1/3 2 c (f, rˆ) Bc (Gµ) (f`) (1+cos ι) Θ(f 2/`). (27) in front of the primary waveform, Ac. ' 2 − (0) (1) The first point is particularly crucial, as it lies at the The fact that h˜c is purely real while h˜c is purely imagin- heart of the divergent behaviour that we investigate in ary means that there is no coherent cross-energy between (0) (1) sectionsIVF andIVG. the two contributions, so the total energy is just c +c . For observers in the beaming direction the energy spec- tra of the primary and memory signals scale in the exact D. Time-domain waveform near the arrival time same way with frequency, but with a much smaller coeffi- cient for the memory signal. However, the picture changes While we have focused on deriving the memory signal in drastically when integrating the spectra over the sphere the frequency domain, we can inverse-Fourier-transform to compute the total emission, as the primary waveform 8

F. Ultraviolet divergence of the radiated energy Primary cusp signal Memory signal The total fraction of the loop’s energy radiated by the Cusp with memory cusp can be found by integrating equation (29),

Z ∞ df ¯(f). (31) E ≡ 0 f ) t

( For the primary cusp signal, this gives + h (0) 2/3 2 3 (πAc) Gµ 1. (32) Ec '  For the memory signal, however, the integral (31) diverges. This is clearly unphysical, and shows a breakdown in the validity of equation (21). Note however that this break- down is not in the low-frequency regime where we know that equation (21) is inaccurate; rather, the integral has an ultraviolet divergence that goes like f 1/3 at high 2 1 0 1 2 frequencies . Instead, we can understand∼ this diver- − − t/s f gence as a breakdown→ ∞ of the Nambu-Goto approximation for the loop dynamics. Figure 3. The time-domain GW strain hc(t) from a cusp, We recall here the justification for the Nambu-Goto with and without the leading-order memory contribution (23). action, following section 6.1 of reference [72], so that we The memory is exaggerated by a factor of ∼ 1/Gµ here to can clarify how this fails for cusps at high frequencies. make it visible. For small inclinations ι < θb the observer lies within the cusp’s beam, and sees the memory superimposed Assuming there are no non-gravitational long-range in- on the primary cusp signal (solid red line). If the inclination teractions between widely-separated string segments, a is very small, ι  θb, then the memory vanishes and only generic cosmic string can be described by a local world- the primary cusp signal is observable (solid blue line). In sheet Lagrangian of the schematic form most cases however, the observer lies outside of the beam, and only the memory is observable (red dashed line). Note ~2κ2 = µ + α~κ + β + , (33) that the higher-order memory contributions (order n ≥ 2) are L − µ ··· not shown here; these would look like step functions in the time domain, with height that either diverges rapidly with where µ is the string tension as before, κ represents space- 3 n [‘large’ loops, ` & δ/(Gµ) ] or converges so rapidly that time curvature (with indices suppressed), and α, β are the contribution to the total signal is negligible [‘small’ loops, 3 numerical constants. In most situations, the curvature ` δ/(Gµ) ]. 2 . associated with a loop is of the order κ `− , where ` is the length of the loop. This is many orders∼ of magnitude 2 2/3 larger than the width of the loop, is then suppressed by a factor of θ (f`)− . We write b ∼ these isotropically-averaged spectra as 1/2 p δ (µ/~)− = `Pl/ Gµ, (34) Z ∼ 2 ¯(f) d rˆ(f, rˆ), (28) where `Pl is the Planck length. This implies that ≡ S2 2 ~κ/µ (δ/`) 1, (35) such that ∼   1/3 so that we can drop all but the first term in equation (33), (0) 2 2 1/3 ¯c (f) (πAc) Gµ(f`)− Θ(f 2/`), leaving the Nambu-Goto action. This ‘zero-width’ ap- ' 3 − (29) proximation for the loop is equivalent to taking ~ 0 in (1) 8 2 3 1/3 → ¯c (f) (πBc) (Gµ) (f`) Θ(f 2/`). equation (33), effacing the microphysics of the underlying ' 3 − field-theory model and keeping only the classical part of We see that the isotropic energy spectrum due to the the action. This is valid for loops larger than a critical memory emission is blue-tilted (i.e. grows with frequency), scale, while the primary spectrum is red-tilted. This means that 3/2 ` δ/(Gµ) `Pl/(Gµ) , (36) the memory emission dominates at very high frequencies, & ∼  3 as field-theory calculations show that loops smaller than 3 Ac f > this lose much of their energy through matter radiation [71, 16` B Gµ c (30) 99] or through topological unwinding and dispersion [91,   1  3 22 ` − Gµ − 100], such that the Nambu-Goto approximation breaks 10 Hz 11 . ≈ × pc 10− down. Note, however, that these effects cannot resolve 9

Nambu-Goto cusp and continuous if a finite width is introduced. [In this Finite-width cusp heuristic setup, it is not immediately clear whether or not the smoothing afforded by a finite string width is strong enough to prevent higher-curvature terms in the Lagrangian (33) from becoming important near the cusp; we return to this point later.] We therefore consider only GWs with frequency f < 1/δ. This has no impact on the primary waveform, since ) t

( 1/2 c   ˙ h 38 Gµ 1/δ 10 Hz 11 , (38) ≈ × 10− which is beyond the reach of any current or planned GW experiments. However, the cutoff does impact the GW memory. For example, setting an upper limit of f = 1/δ in the integral (31), we find a finite value for the energy radiated by the cusp memory,

(1) 2 3 1/3 t c 8(πBc) (Gµ) (`/δ) E ' (39) 2 19/6 1/3 8(πBc) (Gµ) (`/`Pl) . Figure 4. The time derivative of the cusp strain signal close ∼ to the peak, |t − t0|  `, for an observer in the beaming Numerically, this gives direction, ι = 0. This diverges for the standard Nambu- Goto waveform (13a), ruining the validity of the Nambu-Goto  Gµ 19/6 ` 1/3 approximation, and causing a divergence in the energy radiated (1) 15 c 3 10− 11 , (40) by the first-order memory. By introducing a frequency cutoff E ≈ × × 10− pc due to the finite string width, f < 1/δ, we see that the derivative becomes finite and continuous, and the first-order which shows that the energy radiated due to the first- memory divergence is regularised. order memory effect is smaller than that from the primary emission for observationally-allowed values of the string 11 1/3 tension, Gµ 10− . However, the factor of (`/δ) 1 .  the divergence identified above, as this occurs for much in equation (39) is concerning. Based on equation (39), 6 larger loops ` δ/(Gµ), for which matter effects should cusps on GUT-scale strings (Gµ = 10− ) would radiate far be negligible.  more energy through the memory effect than through the 4 The Nambu-Goto dynamics of loops generically predicts primary GWs; for ` & 10− pc they would radiate more the formation of cusps, where the GW strain looks locally than the entire energy of the loop. Such a situation would 1/3 like h(t) t t0 . As was already recognised in be unphysical, and would indicate the breakdown of the reference [∼76], | the− curvature| associated with this scales as validity of the primary cusp waveform. One might argue that this is not an issue, since GUT-scale Nambu-Goto 5/3 strings are already ruled out by observations; however, κ h¨ t t0 − , (37) ∼ ∼ | − | we show below that accounting for higher-order memory which diverges at the peak, clearly ruining the validity effects exacerbates the problem, leading to similar un- of the Nambu-Goto approximation (35). This problem physical results even for much lower string tensions. does not manifest itself in the primary cusp waveform, We should note that the cutoff imposed here is rather since the beaming angle θb decreases fast enough with ad-hoc, and is done in a way that is agnostic to the frequency to ensure the total radiated energy is finite, cf. underlying microphysics of the string. In reality, the equation (32). However, the energy flux in the centre of divergence will be resolved in a way which may depend on the beam is still divergent, and as we have shown here, this the microphysics, and the GW memory observables may sources further divergences due to the nonlinear nature be sensitive to this; indeed, this is implied by the fact of gravity. Since the memory effect is not beamed, there that equation (32) depends directly on the string width is nothing suppressing it at high frequencies. δ. The simplest way to regularise this divergence is to impose an ultraviolet cutoff at the string width scale. By truncating the frequency-domain cusp waveform at f G. Second-order memory ∼ 1/δ, the cusp is smoothed out on timescales t t0 δ, and the curvature reaches a finite maximum| value− | that ∼ The memory waveform (21) describes the spacetime 5/3 scales like κ δ− . This smoothing is shown explicitly curvature generated by the energy-momentum of the in figure4, where∼ we see that the time derivative of the primary GWs from the cusp. However, these memory strain diverges in the Nambu-Goto case, but is finite GWs themselves carry energy-momentum, and will in 10 turn act as a source of their own GW memory (see refer- The waveform (43) can be used to calculate a second- ence [50] for a discussion of this effect in the context of order correction to the energy radiated by the cusp, binary black hole coalescences). We refer to this ‘memory 2 3   of the memory’ here as the second-order memory effect. (2) πr f ˜(0) ˜(1) ˜(2) 2 ˜(0) ˜(1) 2 c = hc + hc + hc hc + hc The calculation of this second-order memory contri- 2Gµ` | | − | | 2 3 bution is straightforward for cusps, simply substituting πr f  (2) 2 (1) (2)  = h˜ + 2h˜ h˜ ∗ equation (21) into the RHS of equation (12) and following 2Gµ` | c | c c all of the same steps as before. The key difference is in  π3 the angular integral, = B4(Gµ)7(`/δ)2/3f` sin4 ι 2 c Z 2 1 2  (1 + cos ι0) = sin ι, (41) 2 3 5 1/3 2/3 2 + π Bc (Gµ) (`/δ) (f`) sin ι(1 + cos ι) rˆ0 6 which, due to the broader emission of the first-order Θ(1/δ f )Θ( f 2/`). 2 × − | | | | − memory signal, is not suppressed by a factor of θb; cf. (44) equation (18). The resulting expressions for the late-time We see that, unlike for the first-order correction, there memory and frequency-domain waveform are (1) (2) are now nonzero cross-terms due to h˜c and h˜c being 2 Z (2) 2(πBc) 4 4/3 2 ∞ df exactly in phase with each other, resulting in two new ∆hc = (Gµ) ` sin ι 2/3 , 3r 2/` f contributions to the energy. Integrating over frequency 2  and emission direction, the total energy is given by iπB Z ∞ df h˜(2)(f) = c (Gµ)4`4/3 sin2 ι 0 c 1/3 1/3 16 − 3rf 2/` f 0 (f 0 + f ) (2) 4 7 5/3 3 5 | | c = (πBc) (Gµ) (`/δ) + 4(πBc) (Gµ) (`/δ). Z f /2  E 15 | | df 0 (45) + , f 1/3( f f )1/3 Comparing with the corresponding first-order result (39) 2/` 0 | | − 0 (42) we see a clear pattern emerging, where the second term in equation (45) is multiplied by a factor of 2 2/3 ∼ both of which contain integrals which diverge due to πBc(Gµ) (`/δ) compared to the first-order energy, and high-frequency contributions from the first-order memory. multiplying by the same factor again gives the first term Introducing the f < 1/δ cutoff from sectionIVF once in equation (45). This factor is greater than unity so long again, we find to leading order in δ, as the loop length ` is larger than 2 (2) 2π ` 2 4 1/3 2 δ ∆hc Bc (Gµ) (`/δ) sin ι, ` 3/2 3 ' r ∗ ≡ (πBc) (Gµ) (43) (46) (2)   7/2 ˜(2) i∆hc Gµ − hc (f) Θ(1/δ f )Θ( f 2/`). 90 m , π 11 ' − 2 f − | | | | − ≈ × 10− 4/3 This is interesting in that it departs from the the f − which applies to all macroscopically-large loops. scaling of both the primary waveform (13a) and the∼ first- One would naively expect successive memory correc- order memory waveform (21); the second-order memory tions for a generic GW source to become less and less instead has a slower decay with frequency, due to the important at higher order; the fact that they become lack of beaming in the first-order memory. In fact, for more important for such a large class of cosmologically- intermediate frequencies 2/` < f < 1/δ the second- relevant cosmic string loops is surprising, and suggests order memory is identical to the| Fourier| transform of that something unphysical is happening. Indeed, if we (2) a Heaviside step function of height ∆hc . This means plug numerical values into equation (45), we find that 14 11 47/10 that on timescales much shorter than the loop oscillation loops larger than about 10 m (Gµ/10− )− (i.e. × (2) period `/2 and much longer than the light-crossing time not much larger than the solar system) have c greater of the loop width δ, the second-order memory signal looks than unity, meaning that they radiate awayE more than like a step function in the time domain; this is because the total energy of the loop. This represents a signific- the signal is dominated by high frequencies f . 1/δ, ant worsening of the issue we identified at the end of and therefore ‘switches on’ in a very short time interval sectionIVF. The problem only gets worse when we 5 t t0 . δ. go beyond second-order corrections, as we demonstrate | − | below.

5 This picture is closely related to the zero-frequency limit (ZFL) H. Higher-order memory, and another divergence method for calculating radiation from high-energy gravitational scattering [8, 12, 14, 101], which leverages the fact that the scatter- ing is effectively instantaneous compared to the GW frequencies Iterating the procedure described above to calculate considered. the third-order GW memory, it is straightforward to show 11 that it obeys the same step-function-like relation that we where Ln(ι) and Sn(ι) are polynomials in cos ι which de- found at second order, scribe the angular pattern of the memory for large and (3) small loops, respectively—these are described in detail ˜(3) i∆hc in appendixB. For small loops, all memory effects are hc (f) = Θ(1/δ f )Θ( f 2/`), (47) − 2πf − | | | | − subdominant compared to the primary emission, and equa- with the late-time memory given by tion (53) gives a convergent geometric series. For large loops, on the other hand, the memory emission becomes 2 Z Z   (3) 2π 2 2 ˜(2) 2 ˜(1)˜(2) stronger at each order, and equation (53) gives a lacun- ∆hc = df r f hc + 2hc hc ∗ , (48) 0 r R rˆ | | ary series which diverges extremely quickly. One might where we have made sure to include both of the second- hope that the polynomials Ln(ι) decrease in magnitude order energy contributions as source terms for the third- fast enough to counteract the divergence, but we find order memory. Upon integration, this becomes empirically in appendixB that

  n−2 (3) 4δ 8/3 2 1 2 2 2 (54) ∆hc = (`/` ) sin ι 1 cos ι Ln(ι) 5(2/5) sin ι, − 5r ∗ − 3 | | ≈   (49) 3 2δ 2 2 3 so the series diverges as long as ` & (5/2) ` 1 km (`/` ) sin ι 1 + cos ι , 11 7/2 ∗ ≈ × − r ∗ 5 (Gµ/10− )− , as shown in figure5. We discuss the cause of this divergence in appendixA, and argue that it where ` is the Gµ-dependent lengthscale defined in equa- ∗ is caused by a super-Planckian GW energy flux from the tion (46). cusp. The third-order memory clearly scales differently for loops with ` ` (which we call ‘large loops’) compared to those with` ∗ ` (which we call ‘small loops’). When going to fourth order ∗ and beyond, we obtain an increasing I. Memory from cusp collapse number of cross-terms in the energy at each order (starting (n) 2 (n 1) (n) (n 2) (n) The results of the previous section show that, even with with h˜c , then 2h˜c − h˜c ∗, 2h˜c − h˜c ∗, and so on, | | (1) (n) an ultraviolet cutoff in place at the scale of the string down to 2h˜ h˜ ∗), resulting in a proliferation of terms c c width, the standard cusp waveform (13a) leads to a di- in the resulting memory expressions, each with a different vergence for all ‘large’ loops with length ` δ/(Gµ)3. power of `/` . It is therefore much simpler to treat small & The fact that the divergence appears in observable, gauge- and large loops∗ separately, and focus on the leading power independent quantities (the memory strain and, therefore, of `/` in each case. This results in a very economical the radiated energy) means that something unphysical formula∗ for iterating the memory calculation, valid for all must be going on. Since the only inputs to our calcula- n 2, ≥ tion are the cusp waveform (13a) and the GW memory  Z r (n 1) 2 formula (2), at least one of these two ingredients must  ∆hc − , for ` ` δ 0 | |  ∗ break down for cusps on loops of this size. ∆h(n) = rˆ (50) c 6πr Z In order to track down the cause of the divergence, let us  ∆ (n 1) ˜(1) for  2 hc − hc (1/δ) , ` ` list the assumptions that go into equations (2) and (13a): δ rˆ0 | |  ∗ where we have evaluated the frequency integral in each 1. The GW frequency is assumed to be much greater case, leaving just the angular integral. The frequency- than the fundamental mode of the loop, f 2/`. domain waveform is then given by the same quasi-step- This is because the waveform (13a) is derived using function form as before, the universal behaviour of the loop on scales ` (n) near the cusp.  ˜(n) i∆hc hc = Θ(1/δ f )Θ( f 2/`), (51) − 2πf − | | | | − 2. The loop’s dynamics are assumed to follow the and the corresponding energy spectra are given by Nambu-Goto action (33) on lengthscales larger than the loop width δ. (We have imposed a cutoff that  r2f  (n) 2 for effaces scales below this, in order to regularise the  ∆hc , ` ` (n) 8πGµ`| |  ∗ first divergence we encountered in sectionIVF.) c = 2 2 (52)  r f (n) ˜(1)  ∆hc hc , for ` ` 2Gµ` | |  ∗ 3. The loop is assumed to evolve according to the flat-space equations of motion derived from (33); Solving equation (50) iteratively with equation (43) as i.e. gravitational backreaction is assumed to be an input, we find for all n 2, ≥ negligible. δ n (`/` )2 /3L (ι), for ` `  n 4. The GWs generated by the loop are assumed to (n) r ∗  ∗ (53) ∆hc = be well-described by linear perturbations on a flat δ 2n/3  (`/` ) Sn(ι), for ` ` background. r ∗  ∗ 12

100 primary emission 1st-order memory 8 . 10− . . 8th-order memory 16 10 limn nth-order memory − →∞

24 10− ) `

( 32

c 10− E

40 10−

48 10−

56 10−

64 10− 12 9 6 3 0 3 6 10− 10− 10− 10− 10 10 10 `/m 100 primary emission 1st-order memory 8 . 10− . . 8th-order memory 10 16 ` = ` − ∗

24 10− ) `

( 32

c 10− E

40 10−

48 10−

56 10−

64 10− 12 9 6 3 0 3 6 10− 10− 10− 10− 10 10 10 `/m

Figure 5. The fractional energy radiated by a cusp at different orders in the memory expansion as a function of loop length, −11 with Gµ = 10 . The top panel shows the standard cusp case, clearly illustrating the divergence at ` & `∗ ≈ 90 m. The bottom panel shows the cusp collapse case, for which the radiated energy at each order drops to a small, `-independent value for ` & `∗, curing the divergence. 13

As mentioned earlier, the first assumption cannot be ated with cusp collapse, but there are two main qualitative the source of the problem, as the divergence is associated differences it introduces compared to the standard cusp with very high frequencies near the string width scale. waveform: (i) the Fourier transform of the primary strain (0) The second assumption is robust so long as (i) the signal, h˜c , is reduced by a factor 1/2, as it only re- higher-order curvature terms in the worldsheet Lag- ceives contributions from the half of≈ the signal at times rangian (33) are negligible, and (ii) the strings are created before the peak; (ii) more importantly, there is a loss of through the breaking of a local gauge symmetry, so that power at frequencies f & 1/(Gµ`), due to the truncation the underlying field theory does not give rise to long- immediately before the peak. Both of these effects in- range interactions (i.e. we are not considering global fluence the corresponding GW memory signal. We can strings, which would instead be described by the Kalb- account for (i) by multiplying the strain at each order Ramond action [72]). As mentioned in sectionIVF, while in the memory expansion by the appropriate power of introducing a finite string width prevents the curvature 1/2, and can approximate the effect of (ii) by introducing from diverging, it does not necessarily guarantee that the a sharp cutoff at frequency f = 1/(Gµ`), which is equi- higher-order curvature terms are negligible. In principle valent to replacing δ Gµ` in all previous expressions. the memory divergence identified here could be cured by The critical lengthscale→(46) which previously marked the departures from the Nambu-Goto action near the cusp. onset of the divergence then becomes However, these departures would have to take place on ` scales much greater than the string width, which seems (55) ` 3/2 2 `, very difficult to achieve. ∗ → (πBc) (Gµ)  By elimination, it seems that the problem is mostly which means that we are always in the ‘small loop’ regime, likely due to assumptions 3 and 4: i.e., that the flat-space ` ` ; higher-order memory corrections are suppressed description of the cusp’s dynamics and GW generation is by powers∗ of Gµ, and the divergence is avoided completely, inconsistent. Indeed, we can trace the divergence back to as shown in figure5. Note that the cusp collapse process 3 the fact that the integrated GW energy flux diverges in is only predicted to take place for loops with ` & δ/(Gµ) , R (0) the centre of the cusp’s beam, d(ln f) c (f, rˆc) , and that the GW memory from smaller loops is still which already suggests that a flat-space description→ ∞ is described by the results given above, with ` given by insufficient. This divergent flux is hidden somewhat by equation (46). ∗ 1/3 the narrowness of the beam, θb (f`)− , which ensures More explicitly, if the divergence is indeed cured by (0) ∼ that c is finite, but we have shown that the GW memory invoking cusp collapse, then the GW observables from inheritsE and amplifies the divergence. the cusp are as follows: the primary waveform is One possible resolution is that the gravitational backre- 2/3 (0) Gµ` action of the loop on itself could smooth out the cusp on h˜ Ac Θ(rˆ · rˆc cos θb)Θ( f 2/`) c r f 4/3 scales much larger than δ. However, there is large body of ' − | | − |(| (56) literature on cosmic string backreaction [102–112] which (1/2)Θ(1/Gµ` f ), for ` ` − | |  ∗ indicates that the backreaction timescale is `/(Gµ), × Θ(1/δ f ), for ` ` much longer than the loop’s oscillation period.∼ While it − | |  ∗ is true that these studies all assume linearised gravity, with ` given by (46); the first-order memory waveform is ∗ and are thus likely to underestimate the magnitude of 2 2/3 the effect near cusps, it is still hard to see how the loop ˜(1) (Gµ) ` hc iBc (1 + cos ι)Θ( f 2/`) could backreact fast enough to smooth out the cusp on ' − rf f 1/3 | | − ( | | (57) relatively large scales before the peak of the signal. (1/4)Θ(1/Gµ` f ), for ` ` All of this suggests that we need some strong-gravity − | |  ∗ × Θ(1/δ f ), for ` ` mechanism which acts on a very short timescale while the − | |  ∗ cusp is forming, and suppresses the cusp’s GW emission with the corresponding late-time memory given by at frequencies far below the cutoff, f 1/δ. In refer- ence [93] we proposed exactly such a mechanism. We (1) 2/3 2 ` ∆hc 2 3 (πAcGµ) (1 + cos ι) argued that when cusps form on sufficiently large cos- ≈ × r ( (58) mic string loops, they source such extreme spacetime 1/4, for ` `  ∗ curvature that a small portion of the loop could collapse × 1, for ` ` to form a black hole at a time Gµ` before the peak of  ∗ the cusp emission. Remarkably,∼ the loops for which this and the nth-order memory waveforms for n 2 are all ‘cusp collapse’ process is predicted to occur are those with step-function-like, ≥ 3 length ` & δ/(Gµ) —exactly the same loops for which (n) i∆hc the higher-order memory divergence occurs. We suggest h˜(n) Θ( f 2/`) that this is no coincidence, and that one plausible solution c ≈ − 2πf | | − ( (59) to the divergence we have identified here is given by the Θ(1/Gµ` f ), for ` ` results of reference [93]. − | |  ∗ × Θ(1/δ f ), for ` ` It is difficult to calculate the precise GW signal associ- − | |  ∗ 14 with height given by total energy emission is ( (0) 2/3 2 1/4, for ` ` 3 (πAc) Gµ  ∗ (61) Ec ≈ × 1, for ` `  ∗  2n−1   n+1  ` πBc 1+ 2  (Gµ) 3 Ln(ι), for ` ` with only cusps above the collapse threshold being subject (n) ∗ ∆hc r 4  to the 1/4 reduction in GW power due to the truncation ≈ δ 2n/3  (`/` ) Sn(ι), for ` ` of the signal. However, the difference becomes more r ∗  ∗ significant in the memory emission, (60)

1 2 8/3  (πBc) (Gµ) , for ` ` (1) ∗ c 2  (62) E ≈ 3/2 2 8(πBc) (Gµ) (`/` ), for ` ` ∗  ∗ It is interesting to note that loops with length just below where there is an extra power of (Gµ)2/3 for cusps above the cusp-collapse threshold, ` . ` , emit more energy the collapse threshold, compared to those just below it. through GW memory than those above∗ the threshold. We This pattern continues for higher-order memory, where can see this already in the primary emission, where the the total radiated energy for all n 2 is given by ≥

n   2 2 n+2 Z rˆ 1 πBc 2 /3 d 2  (Gµ) Ln(ι) , for ` ` (n) 2 4 S2 4π | |  ∗ (63) Ec ≈ Z 2rˆ 1 3/2 2 (2n 1)/3 d  (πBc) (Gµ) (`/` ) − 6(1 + cos ι)Sn(ι), for ` ` 2 ∗ S2 4π  ∗

We note briefly that reference [93] also discussed a We consider here the simplest case, where the beam further GW signature associated with cusp collapse: the of the kink traces out a great circle on the sphere at a high-frequency quasi-normal ringing of the newly-formed constant rate. We choose our polar coordinates such that black hole after the collapse. We neglect this effect here this circle lies in the equatorial plane, θk = π/2. The however, as it is hard to say anything concrete about the azimuthal direction of the kink’s beam is then given by phase evolution and angular pattern of this additional φk = 4πσt/` with σ = 1, where the two different signs GW emission, and this prevents us from calculating the correspond to left- and± right-moving kinks respectively.6 associated memory signal. It would be interesting to From equation (13b) we then have revisit this contribution to the memory if and when more 1/3 detailed phase-coherent cusp collapse waveform models (0) Gµ` iσφf`/2 h˜ (f, rˆ) Ak e− Θ(θb ι )Θ( f 2/`), become available. k ' r f 5/3 − | | | | − | | (64) where the inclination ι π/2 θ describes the angle between the observer’s line≡ of sight− and the closest point V. MEMORY FROM KINKS on the equatorial plane, and takes values ι [ π/2, π/2] (with positive/negative values corresponding∈ − to the ob- Unlike cusps (which are transient), kinks are persistent server being above/below the plane). Note that we have iσφf`/2 features of cosmic string loops: they propagate around picked up a phase factor e− to account for the time the loop at the speed of light, continuously emitting GWs at which the kink passes closest to the line of sight. As in a beam which traces out a one-dimensional ‘fan’ of we show below, this direction-dependent phase ultimately directions, like a lighthouse. This fact is usually unim- leads to a strong suppression of the kink memory signal portant when computing the primary GW emission from compared to the cusp case. kinks, since the beam only overlaps with the observer’s line of sight for a small fraction of the loop oscillation time, meaning that kinks are effectively transient sources 6 for a given observer. However, in order to calculate a The definition of whether a given kink is left- or right-moving is somewhat arbitrary here. For concreteness, we call kinks with GW memory signal, we need to know the primary GW σ = +1 ‘left-moving’; these move anti-clockwise around the loop flux in all directions over the entire history of the source, when viewed from the North pole θ = 0. Conversely, kinks with which for kinks means specifying the beaming direction σ = −1 are called ‘right-moving’; these move clockwise when as a function of time. viewed from θ = 0. 15

A. Beaming effects account for the two competing beam angles θb(f 0) and θb(f 0 f), we obtain − As in the cusp case, the first step in calculating the i25/3π(A Gµ)2`1/3 GW memory signal is to compute the angular integral h˜(1)(f) k K ( σι) k ' − 31/6rf f 2/3 f`/2 − that captures the beaming effects of the kink, | | " 1/2 # Z ∞ du Z du Z 0 iσφ f`/2 2/3 2/3 e− Θ(θb ι ). (65) × 2/ f ` u (1 + u) − 2/ f ` u (1 u) 0 − | | | | | | − rˆ (69)

Here the small circle of radius θb around the North pole that we considered for cusps has been replaced with a for f > 4/`. (This is analogous to equation (20) from the cusp| | case.) Taking the limit , we can evaluate narrow band of half-width θ around the equator, and we f 2/` b the integrals analytically to find| |  have included the φ-dependent phase factor from equa- tion (64). It is straightforward to integrate out the zenith 2 1/3 (1) (Gµ) ` angle θ if we keep only the leading-order term in θ ; this h˜ (f) iBk K ( σι)Θ( f 4/`), (70) 0 b k ' − rf f 2/3 f`/2 − | | − gives | | Z where the numerical constant is iσφ0f`/2 e− Θ(θb ι ) θbKf`/2( σι), (66) 2  5/3  rˆ0 − | | ' − πAk 2 2π 1 1 4 Bk 2F1( , ; ; 1) ≡ √3 3 3√3 − 3 3 3 − (71) where we have defined a family of azimuthal integrals, 0.5830, ≈ Z 2π 2 dφ inφ (sin ι cos φ i sin φ) and where we have set the frequency cutoff at double the Kn(ι) = e − . (67) 2π 1 cos ι cos φ fundamental mode f = 2/` due to the different behaviour 0 − 0 of the angular integral Kn(ι) for n = 1 compared to n 2. Notice that the argument n is always an integer, as the This waveform (70) shares many features with both≥ GW frequency is always an integer multiple of the loop’s the cusp memory waveform (21) and the primary kink fundamental mode 2/` (although we often ignore this waveform (64). The most important difference from both when taking the continuum limit at high frequencies). of those waveforms is the dependence on inclination, which Computing the integral equation (67) for general n there- here is a function of frequency. Using the results derived fore corresponds to finding the Fourier spectrum of a in appendixC, we can rewrite equation (70) as complicated nonlinear function of φ; we perform this cal- 2 1/3 culation explicitly in appendixC. (1) (Gµ) ` 4 sin ι h˜ (f) iBk | | k ' − rf f 2/3 cos2 ι | |  f`/2  cos ι B. Late-time memory  Θ(f 4/`), for σι < 0  1 + sin ι −  | | f`/2 ×  cos ι − Using equation (9), along with the expression for the  Θ( f 4/`), for σι > 0  1 + sin ι − − angular integral K0(ι) from equation (C5), we find that | | the late-time memory from the kink is given by (72)

5/6 meaning that the memory signal from left-moving kinks (1) 3 ` 2 4 sin ι + cos 2ι 3 contains only negative frequencies above the equatorial ∆hk = (πAkGµ) | | 2 − . (68) r cos ι plane and only positive frequencies below the plane, and

(1) vice versa for right-moving kinks. Inserting numerical values, the strongest effect is ∆ hk For high frequencies n 1 the angular integral Kn(ι) 4.250 (Gµ)2`/r when the observer lies in the plane≈ | |  − × has a maximum value of 4/(e n ) at inclination ι 1/n. of the kink ι = 0, an order of magnitude smaller than This means that the kink' memory| | signal is only observable' the maximum cusp memory. The late-time kink memory very close to the plane of the kink (but not in the plane, decreases smoothly to zero as one approaches the poles where it vanishes; see figure6), and is suppressed by an ι π/2. → ± extra power of frequency compared to the primary signal, (1) 8/3 h˜ f − . k ∼ C. Full waveform D. Time-domain waveform near the arrival time The calculation here is very similar to the cusp case in sectionIVC. Inserting equation (64) into equation (12), As with the cusp case, we can reverse-Fourier-transform and taking care to enforce the frequency cutoffs and to equation (72) to find the time-domain memory strain 16 around time of arrival of the primary kink signal, t t0 `. Unlike the cusp case, we obtain a signal which| − is| not  linearly polarised, but contains both + and polarisation content, ×

2 (1) (1) 16πBk(Gµ) h (t) h (t0) (t t0) k,+ − k,+ ' 21/3r − sin ι  1 + sin ι  | |E 2 ln | | , × cos2 ι 2/3 cos ι 2 2 (1) (1) 64π Bk(Gµ) 2 hk, (t) hk, (t0) 1/3 (t t0) × − × ' 2 `r − sin σι  1 + sin ι  E 1/3 2 ln | | , × cos2 ι − cos ι (73) where we have used the generalised exponential integral R zx n function, En(z) ∞ dx e− x− . The -polarised com- ≡ 1 × ponent is suppressed by an additional factor of (t t0)/` 1, meaning that the signal is still approximately− +- polarised. Figure 6. A cartoon illustration of the angular distribution of An important difference with respect to the cusp case the energy radiated by a kink. The primary emission (blue) −1/3 is that since kinks are long-lived rather than transient, is concentrated in a narrow fan of half-width θb ∼ (f`) their memory signal is not concentrated around a partic- around the plane of the kink, while the memory emission (red) ular arrival time, but can in principle be observed at all is concentrated in two lobes either side of this plane, which are exponentially suppressed as one approaches either of the times. One can easily obtain expressions analogous to directions normal to the plane. equation (73) at any point in the kink’s periodic motion by substituting in the appropriate time when evaluating the reverse Fourier transform; in general this leads to a so that at high frequencies, the isotropic energy spectrum mixing between the +- and -polarisation modes. × is approximately

3 (1) 2 (Gµ) E. Radiated energy ¯ (f) 32(πBk) Θ(f 4/`). (77) k ' (f`)10/3 −

As in the cusp case, we are interested in the total en- We see that the angular pattern of the kink memory ergy radiated by the kink, and how the memory emission strongly suppresses the isotropic spectrum at high fre- adds to this. Using equation (25), we find that the dimen- quencies, meaning that unlike in the cusp case, there is sionless energy spectra for the primary emission and the no frequency range where the memory contribution dom- first-order memory are inates, and the total radiated energy converges without imposing an ultraviolet cutoff, (0) π 2 1/3  (f, rˆ) A Gµ(f`)− Θ(θb ι )Θ(f 2/`), k ' 2 k − | | − (0) 3 2 f` 4 5/6 2 (Gµ) sin ι cos ι k 3 (πAk) Gµ, (1) rˆ 2 − E ' k (f, ) 4πBk 1/3 f` Θ(f 4/`). 2 (78) ' (f`) (1 + sin ι ) − (1) 3(πBk) | | (Gµ)3. (74) Ek ' 28/3 5 × Integrating over the sphere, the primary emission is sup- 1/3 pressed by a factor of the beaming angle θb (f`)− , giving ∼ F. Higher-order memory

5/3 (0) 2 2 2/3 As with the cusp case, we can iterate the memory ¯ (f) (πAk) Gµ(f`)− Θ(f 2/`). (75) k ' 31/6 − calculation with our 1st-order memory waveform (70) as an input to calculate the 2nd-order memory effect (the The spherical integral for the memory contribution can ‘memory of the memory’). This involves calculating the be evaluated explicitly to give angular integral

Z 2 f` 4 Z 2 sin ι cos − ι 8π iσφ0f`/2 d rˆ f` = , (76) e− Kf 0`/2( σι0)K(f 0 f)`/2( σι0), (79) 2 (1 + sin ι ) f`(f` 2)(f` + 2) 0 − − − S | | − rˆ 17 which in the high-frequency regime f 0 , f 0 f 2/` is which would typically contain structure on scales smaller well-approximated by | | | − |  than the loop length `, causing the path of the beam to vary on those scales. We expect that such structure is  16Kf`/2( σι) only likely to make a significant qualitative difference to − 3 Θ(f 0 4/`)Θ(f 0 f 4/`) ' (f 0` f`/2) − − − our results if it is on a scale corresponding to the GW −  frequency of interest. We note that small-scale structure Θ( f 0 4/`)Θ( f 0 + f 4/`) . − − − − − on loops is expected to be damped over time through (80) gravitational backreaction [103], so our simple treatment here is not likely to be too unreasonable. In any case, With this result to hand, the remaining steps are very sim- we do not expect such considerations to change the main ilar to the calculations for the 1st-order memory described conclusion of this section: that memory from kinks is above, yielding highly suppressed compared to the cusp case. We have also neglected the fact that kinks always appear 512πB2(Gµ)4 h˜(2)(f) i k K ( σι)Θ( f 4/`) in pairs on loops, with one left-mover for every right-mover. k ' − rf f 10/3`7/3 f`/2 − | | − A realistic loop is likely to have several pairs of kinks, | | Z ∞ du with each of these sourcing a GW memory signal like 2/3 2/3 3 the one calculated here. Since the kinks travel around × 4/ f ` u (u + 1) (u + 1/2) | | the loop at the same average rate, it is possible that 6664 (Gµ)4 i × K ( σι)Θ( f 4/`). the superposition of their memory signals could give rise ≈ − rf f 10/3`7/3 f`/2 − | | − to interesting coherent effects. However, regardless of | | (81) whether the kink memory signals are coherent or not, the GW energy flux will still be of the same order of The fact that this has the exact same dependence on the magnitude, so our main conclusions are unaffected. inclination ι as the 1st-order memory makes it straightfor- ward to iterate the process to higher orders. Doing this, we find that for all n 2, the kink GW memory is given schematically by ≥ VI. DETECTION PROSPECTS

2n 3 (n) i(Gµ) f` Having calculated the nonlinear GW memory wave- h˜ (f) − Kf`/2( σι)Θ( f 4/`), (82) k ∼ r( f `)8n/3 − | | − forms associated with cusps and kinks, it is natural to ask | | whether these signals are detectable with current or future multiplied by some numerical constant (for which there GW observatories. Clearly the divergent behaviour dia- does not seem to be a simple expression for all n). gnosed for cusps in sectionIVG could, in principle, have The situation here is drastically different from the cusp important observational implications, depending on how case. For cusps we saw that each successive order in the the divergence is regulated. For the purposes of this sec- memory was suppressed by larger powers of Gµ 1, but tion we assume the divergence is resolved along the lines also enhanced by larger powers of `/δ 1, and that in  of the cusp-collapse scenario described in sectionIVI. certain situations the latter would dominate, causing a We show below that, under this assumption, the cusp divergence. For kinks, we see instead that each order in and kink memory signals are suppressed so strongly that the memory is not only suppressed by powers of Gµ, but 8/3 they are well beyond the reach of GW observatories, even is further suppressed by a factor of ( f `)− 1 each | |  futuristic third-generation interferometers like Einstein time. The signal falls off quickly enough with frequency Telescope (ET) [122] and Cosmic Explorer (CE) [123]. that higher-order memory contributions are not sensitive to the string-width scale, and no factors of `/δ appear. This means that there is no situation where the kink memory diverges, and that the higher-order contributions A. Burst searches are negligible in any observational scenario. This lack of divergence is in agreement with the cusp-collapse mechan- We start by calculating the expected detection horizons ism we have invoked as a possible resolution for the cusp for individual bursts of GW memory from cusps and divergence, as kinks are not predicted to form PBHs [93]. kinks—i.e. the maximum distance at which a burst can be detected, on average. We assume a matched-filter search, such that the optimal root-mean-square SNR (averaging G. Caveats of our approach over sky location and polarisation angle) for a frequency- domain waveform h˜(f) which is isotropically averaged We have only considered the simplest case where kink’s over the source inclination is given by [94] beam traverses a fixed plane at a constant rate, in order to make detailed analytical calculations feasible. This 1/2 " Z ˜ 2 # X 4 2 ∞ h(f) situation is highly idealised, and is not representative of ρrms = sin αI df | | . (83) 5 PI (f) the loops one would find in a cosmological loop network, I 0 18

100 where r(z) is now the comoving distance to the source, and fs = (1 + z)f is the source-frame frequency, with z 7 10− the redshift. We assume that any cosmic string signal with ρrms 12 can be confidently detected; in reality, this 14 ≥ 10− threshold depends on the distribution of non-Gaussian noise transients (‘glitches’) in the network, and how closely 21 these are able to mimic the waveforms of interest, but we 10−

) ignore these details here. f

( 28 Even assuming an optimistic third-generation GW de- 10− gw tector network consisting of Einstein Telescope plus two Ω 7 10 35 Cosmic Explorers, we find that given current constraints − on the string tension8 the detection horizon for a cusp memory signal is at most 2 pc. This corresponds to a 10 42 primary GWs − leading-order memory negligibly small detection rate,≈ as very few cosmic strings higher-order memory 49 LIGO/Virgo O3 are expected within a volume of this size. For kink memory 10− LISA the result is even more pessimistic, with a detection ho- PPTA rizon of (a few times larger than the Earth- 56 0.008 AU 10− 14 8 2 4 10 16 ≈ 10− 10− 10− 10 10 10 Moon distance). Larger values of the string tension Gµ f/Hz would boost the detectability of these memory bursts, but would be in conflict with existing observational results.

Figure 7. Contributions to the SGWB energy spectrum Ωgw(f) from cosmic strings at different orders in the memory expan- sion, assuming that the memory divergence is resolved through B. Stochastic background searches the cusp-collapse scenario described in sectionIVI. We see that the memory effect is negligible compared to the primary The combined GW emission from many loops through- emission. Here we assume ‘model 2’ [83, 113] of the string −11 out cosmic history gives rise to a stochastic GW back- network, and a string tension of Gµ = 8.1 × 10 (this is ground (SGWB) [70–72, 76, 78, 81, 128–134]. The intens- the largest string tension allowed by current constraints in ity of this SGWB as a function of frequency is usually the cusp-collapse scenario [93]; smaller tensions suppress the expressed as a fraction of the cosmological critical density memory effect even further). The dashed/dotted curves show 2 the contributions from the matter/radiation era, respectively, ρcrit = 3H0 /(8πG) in terms of the density parameter, with the solid curves showing the combined spectra. A distinct 1 dρgw change in all three of the radiation-era spectra is visible at Ωgw(f) , (85) around f ≈ 10 Hz; at frequencies above this, the spectra are ≡ ρcrit d(ln f) increasingly dominated by small loops which do not undergo cusp-collapse, and this is why the memory effect becomes more which for cosmic string loops can be written as prominent (though still undetectable). The magenta curve Z Z 16πGµ X 4 shows the power-law-integrated (PI) sensitivity curve [114] Ωgw(f) = Ni dt d` a n(`, t)¯i(f/a, `), 3H2 from the LIGO/Virgo O3 isotropic stochastic search [80], which 0 i is publicly available at reference [115]. The green curve shows (86) the Parkes Pulsar Timing Array (PPTA) PI curve [116, 117], where t is cosmic time, a(t) is the FLRW scale factor, and calculating using the code from reference [114], which is pub- n(`, t) d` is the comoving number density of loops with licly available at reference [118]. The cyan curve shows the length between and . The sum is over different projected LISA [119] power-law-integrated sensitivity curve, ` ` + d` as described in references [120, 121]. GW signals (cusps, kinks, and kink-kink collisions), with Ni the mean number of signals per loop per oscillation period, and ¯i the isotropic energy spectrum of each signal, as defined in equation (28). The sum here is over different GW detectors, with PI (f) In figure7 we show the SGWB spectrum from a particu- representing the noise power spectral density (PSD) of lar model of the loop network, including the contributions detector I, and αI the opening angle between the two interferometer arms (this angle enters through the de- tector’s response function; LIGO, Virgo, and CE have an opening angle of π/2, while each of ET’s three interfero- 7 For ET we use the ‘ET-D’ noise PSD developed in reference [124], meters has an opening angle of π/3). It is convenient to while for CE we use the ‘CE-2’ noise PSD developed in refer- rewrite this in terms of the fractional energy spectrum, ence [125]. 8 Stochastic GW background constraints from PTAs [82, 83] imply using equation (25) to give −11 that Gµ . 8.1×10 in the cusp-collapse scenario [93] for ‘model 2’ [113] of the loop network. The other network model considered " #1/2 most frequently in the literature (‘model 3’ [126, 127]) gives an 2Gµ` Z ∞ ¯(f ) −15 X 2 s even more stringent constraint of Gµ . 4.0 × 10 [81]. (See ρrms = 2 2 sin αI df 3 , (84) 5π r f PI (f) references [78, 81, 128] for an overview of these models.) I 0 s 19

from first-order and higher-order GW memory from cusps strings, deriving detailed analytical waveform models for and kinks. We see that while the primary SGWB reaches the memory GWs, including the ‘memory of the memory’ a plateau at high frequencies, the memory contributions and other higher-order memory effects. These are among to the SGWB grow with frequency above 10 Hz; this the first memory observables computed for a cosmological makes sense given the slower fall-off of the cusp≈ memory source of GWs, with previous literature having focused emission at high frequency compared to the primary cusp almost entirely on astrophysical sources. and kink signals. However, each order in memory is sup- The leading-order cusp memory waveform (21) that pressed by an additional factor of (Gµ)2, and this suppres- we have found is strikingly similar to the primary GW sion is strong enough to render the memory contribution signal from the cusp (13a), with the same characteristic 4/3 ∼ unobservable at all frequencies. f − frequency power-law. However, one very important difference is that this memory signal is emitted in all directions, unlike the primary signal which is confined to C. Consequences for the loop distribution function 1/3 a narrow beam of width θb f − . As a result, we find that the total GW energy radiated∼ by the cusp memory Additional energy loss due to GW memory emission diverges for a Nambu-Goto (i.e., zero-width) loop. This means that loops decay more rapidly, modifying the dis- divergence can be regularised by introducing a cutoff at tribution of loop sizes. This could in principle leave dis- the scale of the string width δ `Pl/√Gµ, but this then tinguishable imprints on the two observational probes introduces powers of `/δ 1 to∼ the higher-order memory mentioned in the previous sections. However, we show be- terms, causing the sum of all memory contributions to 3 low that such imprints are suppressed by powers of Gµ in diverge for loops of length ` & δ/(Gµ) . the cusp-collapse scenario, rendering them unobservable. In sectionIVI we have argued that the most plaus- Following reference [135], we write the loop decay rate ible explanation for this unphysical memory divergence as is the assumption that the spacetime containing the cos- d` mic string loop is well-described by a flat background = Γ Gµ (`), (87) dt − J with linear perturbations. We have suggested that some strong-gravity mechanism must kick in on loops of length where Γ 50 is a numerical constant, and (`) is an 3 to suppress the high-frequency GW emission arbitrary≈ function of the loop length which isJ equal to ` & δ/(Gµ) from cusps, thereby curing the divergence. Our earlier unity in the standard Nambu-Goto case. If differs work reference [93] provides exactly such a mechanism: from unity then the faster/slower loop decay rateJ leads to the collapse of a small cosmic string segment near the cusp a smaller/greater number of loops once the equilibrium to form a PBH. Remarkably, this cusp-collapse process ‘scaling’ distribution is reached; reference [135] shows that is predicted to occur for all loops of length ` δ/(Gµ)3— the modified loop distribution function n(`, t) can then & exactly the same loops for which we have diagnosed the be obtained from the standard Nambu-Goto distribution memory divergence. We have shown explicitly that if by inserting appropriate factors of . cusp collapse does indeed occur for these loops, the cor- We account for enhanced energyJ loss through GW responding GW memory is strongly suppressed, and the memory radiation by writing divergence is cured. X∞ (n) (1) We have also calculated the memory emission associated (`) = E 1 + E , (88) J (0) ' (0) with kinks, and have shown that this is suppressed due to n=0 E E interference between GWs emitted by the kink at different where (n) is the dimensionless GW energy radi- points in its history. There is thus no situation in which ated throughE nth-order memory, as given by equa- the kink memory signal diverges; this accords with the tions (61), (62), and (63) for cusps and equation (78) cusp-collapse description, as kinks are not predicted to for kinks. We find that (`) 1 is at most Gµ for both form PBHs in that scenario. cusps and kinks, meaningJ that− any modifications∼ to the In sectionVI we have investigated the detection pro- loop distribution function are suppressed by powers of Gµ, spects for these cusp and kink memory signals, calcu- and are thus unobservable. (This is assuming the memory lating the expected detection rate of individual memory divergence is resolved through cusp collapse; of course, it signals for matched-filter searches with third-generation is still possible that the memory emission could be much GW observatories like Einstein Telescope and Cosmic larger, and thereby lead to significant imprints on the Explorer, as well as the contribution of memory GWs loop distribution, and indeed on the other observational to the SGWB spectrum, and possible imprints in the probes we have discussed.) cosmic string loop distribution function. We find that by requiring the string tension to agree with existing 10 observational bounds (Gµ . 10− ) and invoking cusp VII. SUMMARY AND CONCLUSION collapse to prevent the memory from diverging, the res- ulting memory signal is very strongly suppressed, and is We have thoroughly explored the nonlinear GW memory not likely to be detected by any current or upcoming GW associated with cusps and kinks on Nambu-Goto cosmic observatories. Of course, if cusp collapse does not occur, 20 then it is possible that large loops could source much a toy model in which the primary GW is a Gaussian pulse stronger memory signals; however, one would then need which reaches our detector at retarded time t = 0, an alternative means of resolving the cusp divergence. Our work demonstrates the importance of considering A  t2  h(0)(t) = exp . (A1) the nonlinear memory associated with a broader class of r −2σ2 GW sources than just compact binaries; we have shown that the memory effect is interesting not just from an ob- This has amplitude A and width σ, both of which have servational point of view, but also as a tool for sharpening dimensions of length. We ignore the polarisation and our theoretical understanding and modelling of said GW angular pattern of the pulse, though these can play an sources. For cosmic strings in particular, as an unexpected important role (e.g., for the case of kinks, as discussed by-product of our analysis we have shown that the stand- above), and focus on the pulse’s behaviour as a function ard Nambu-Goto description of cusps is unphysical, and of time. The Fourier transform of equation (A1) is that strong gravity effects (possibly including PBH form- ation) could play an important rôle in a more complete A  (2πfσ)2  understanding of their dynamics. This motivates further h˜(0)(f) = √2πσ exp , (A2) work to better understand cusps in full GR, including r − 2 nonlinear gravitational effects beyond those considered here, and thereby compute reliable waveform predictions so we see that the pulse has a characteristic frequency of 1/σ. The pulse is smooth and infinitely differentiable for GW observatories. It would also be very interesting ∼ to study the linear memory associated with cusps, and to for all finite σ; however, by making σ small, we can make understand how this contributes to the total GW emission; the pulse arbitrarily ‘sharp’ (equivalently, we can make the however, such a study would require us to go beyond the characteristic frequency arbitrarily large). This sharpness Nambu-Goto approach adopted here. More generally, our can be quantified by the maximum value of the time results motivate a broader examination of the nonlinear derivative of the strain, memory effect in GW astronomy and cosmology, to see (0) 1/2 what other surprises may be in store. max rh˙ (t) = e− A/σ t | | [amplitude] [characteristic frequency], ∼ × ACKNOWLEDGMENTS (A3)

We thank Josu Aurrekoetxea and Eric Thrane for their i.e. if the dimensionless ratio A/σ is much greater than valuable feedback on this work. A. C. J. is supported unity the pulse is ‘sharp’, and if A/σ is much less than by King’s College London through a Graduate Teaching unity the pulse is ‘soft’. Scholarship. M. S. is supported in part by the Science and The pulse’s first-order memory can be written as Technology Facility Council (STFC), United Kingdom, under research grant No. ST/P000258/1. This is LIGO At2  t2  rh˙ (1) = C rh˙ (0) 2 = C exp , (A4) document number P2100040. | | σ2 −σ2

where C is some constant arising from the angular integral, Appendix A: Understanding the origin of the which we assume to be (1) for all orders in the memory higher-order memory divergence expansion. (This assumptionO can easily be violated: e.g., if the integrand has vanishing TT component then C = 0.) Here we derive a condition (A5) on a generic GW signal We immediately see that the memory signal is larger than h(t) which, we argue heuristically, is necessary for the the primary GW signal near the arrival time ( t σ) if associated nonlinear memory to diverge. We find this and only if A/σ 1, or equivalently, | | ∼ condition by considering a simple toy model in which  we can tune the ‘sharpness’ of the signal to find where max rh˙ (t) 1. (A5) the divergence sets in. We then show that this condition t | |  predicts the divergence for cusps on ‘large’ cosmic string loops, while also providing an explanation for why the Using the Isaacson formula (5), we see that this is equi- memory from compact binaries never diverges. valent to

dEgw 1 mPl max = , (A6) 1. Gaussian pulse as a toy model t dt  G tPl

Our goal here is to derive a condition on how ‘sharp’ a i.e. if the GW energy flux is super-Planckian. putative GW signal must be to give rise to a divergent Moving on to the second-order memory, there are two nonlinear memory expansion. To investigate this, we use contributions: the self-energy of the first-order memory, 21 and its cross-energy with the primary signal, We can circumvent this issue by considering the first- order memory signal from the cusp as the source of the rh˙ (2) = Cr2( h˙ (1) + h˙ (0) 2 h˙ (0) 2) divergence: this should give C = (1) as it has a broad | | − | | O = C rh˙ (1) 2 + 2Cr2h˙ (1)h˙ (0) emission pattern, and is not concentrated into a narrow | | beam. Indeed, using equation (23) we find that the max- 4 At  2t2  imum time derivative for the first-order cusp memory = C3 exp (A7) σ2 − σ2 signal is  3  2  ˙ (1) 2 2/3 2 At 3t max rhc (Gµ) (`/δ) . (A12) 2C exp . t | | ∼ − σ2 −2σ2 Applying the condition (A5) again, we find that the cusp The general pattern is straightforward from here. Treating memory signal should diverge if the ‘sharp’ regime and the ‘soft’ regime A/σ 1 A/σ 1 3 separately (analogous to the separation between the ‘large ` & δ/(Gµ) , (A13) loop’ and ‘small loop’ regimes for cusps), we find for all which is exactly what we find from the careful analysis in n 1, ≥ sectionIVH. This supports the idea that equation (A5)  2n gives a necessary condition for the memory divergence, 1 CAt  t2   exp , A/σ 1 which is generally applicable if the factor C arising from C σ2 −2σ2  the angular integral is not too small. rh˙ (n) n+1 '  1  2CAt  t2    2 exp 2 , A/σ 1 4C − σ −2σ  3. Application to compact binaries (A8) which shows that the memory expansion diverges if equa- Here we use a very simple heuristic analysis to show that tion (A5) holds. ˙ is at most for compact binary coalescences, so Of course, there are many ways in which this argument rh(t) (1) that| the| memoryO divergence condition (A5) never holds. could fail, some of which we have already mentioned (in We neglect numerical factors throughout. particular, if C 1). However, the general takeaway is The GW signal from a CBC can be written as that if the time derivative of a GW strain signal is large, ˙ iφ(t) then a memory divergence might occur. For maxt rh 1 h(t) = (t)e , (A14) on the other hand, it seems very likely that the| memory|  A expansion must always converge. If this is indeed the case, where the leading-order Newtonian contributions to the then we could interpret equation (A5) as a necessary, but amplitude and phase are [94] not sufficient, condition for the memory divergence. (G )5/4  τ 5/8 M , φ , (A15) A ∼ rτ 1/4 ∼ G M 2. Application to cosmic strings with τ the time until coalescence and η3/5M the chirp mass, where η 1/4 is the dimensionlessM ≡ mass ratio. How does this fit into our results for cosmic string The time derivative≤ of the strain (A14) is therefore cusps? Neglecting numerical constants, the time-domain s cusp waveform looks like q G 5/2 G 5/4 rh˙ (t) = r ˙ 2 + rφ˙ 2 M + M . Gµ | | | A| | A| ∼ τ τ h(0)(t) `2/3 t 1/3 + constant, (A9) (A16) c r ∼ − | | Formally, this diverges in the Newtonian analysis as τ → so the time derivative diverges at t = 0 due to the absolute 0. However, introducing a cutoff at the ISCO radius value function. If we introduce a cutoff at the string width truncates the signal at a finite value of τ, scale δ, we find that 8/5 τmin η− G , (A17) ∼ M ˙ (0) 2/3 max rhc Gµ(`/δ) . (A10) t | | ∼ so that p max rh˙ (t) η2 + η4. (A18) Naively applying the condition (A5), we would thus expect t | | ∼ the cusp memory signal to diverge for loops of length Since η is at most (1), we see that the condition (A5) 3/2 O ` & δ/(Gµ) . (A11) is never met. This makes complete sense, given that we know that the memory from CBC signals cannot diverge. However, this does not happen, for the simple reason As a by-product, equation (A18) leads us to conjecture that there is a small factor C 1 arising from the an- that equal-mass binaries (η = 1/4) should give rise to gular integral, and as mentioned above this causes the stronger higher-order memory effects than extreme mass- condition (A5) to fail. ratio inspirals (η 1).  22

1.2 n = 2 . . 2 . 1.0 n = 6

1 5

/ 0.8 | ) ) ι ι ( ( n n S L n |

2 0.6 0 − 1) n 2 − ( 2) /

(5 0.4 1 −

n = 2 0.2 . . 2 . − n = 12 0.0 1.0 0.5 0.0 0.5 1.0 1.0 0.5 0.0 0.5 1.0 − − cos ι − − cos ι

Figure 8. Functions describing the strength of the nth-order cusp memory signal as a function of inclination, with Ln(ι) representing ‘large’ loops `  `∗, and Sn(ι) representing ‘small’ loops `  `∗.

Appendix B: Angular integrals for higher-order cusp and equation (7) becomes memory Z Z 2 d rˆ0 2iφ0 = (1 + cos θ0)e− , (B5) Here we derive the angular patterns of the higher-order rˆ0 S2 4π cusp memory in the ‘large-loop’ (` ` ) and ‘small-loop’ n  ∗ so that we can use a binomial expansion of cos ι0 = (` ` ) limits, which are described by the functions Ln(ι) n  ∗ (cos θ0 cos ι cos φ0 sin θ0 sin ι) to write and Sn(ι) respectively, with ι the inclination between the − cusp beaming direction and the observer’s line of sight. n   X n k n k These are defined by inserting the nth-order memory Cn(ι) = ( sin ι) cos − ι k − formula (53) into the iterative relation (50), which gives k=0 Z Z +1 dx 2 n k 2 k/2 (B6) Ln+1(ι) = Ln(ι0) , (1 + x)x − (1 x ) 0 × 1 2 − rˆ | | − Z (B1) Z 2π dφ0 2iφ0 k Sn+1(ι) = 6(1 + cos ι0)Sn(ι0), e− cos φ0, rˆ0 × 0 2π for all n 2, with where we have set x = cos θ0. Using the Beta function ≥ identity 2 L2(ι) = S2(ι) = 2 sin ι. (B2) Z 1 Γ (a)Γ (b) R a 1 b 1 (B7) (The integral symbol rˆ0 is defined in equation (7).) B(a, b) dx x − (1 x) − = , ≡ 0 − Γ (a + b) We find that Ln and Sn can be written as polynomials n 1 in cos ι (of order 2 − and n, respectively). The iterative we then obtain, for all n 0, process (B1) can therefore be carried out by evaluating a ≥ n 2k 2(n k) simpler family of integrals, X √π(n k)(2n)! cos ι sin − ι C2n(ι) = − , Z 22n+1k!(n k + 1)!Γ (n + 3 ) n k=0 2 Cn(ι) cos ι0. (B3) − n 2k+1 2(n k) ≡ rˆ0 X √π(n k)(2n + 1)! cos ι sin − ι C2n+1(ι) = − . 22n+2k!(n k + 1)!Γ (n + 5 ) To do so, we choose our rˆ0 = (θ0, φ0) coordinates such that k=0 − 2 θ0 is zero along the line of sight, so that cos θ0 rˆ · rˆ0, (B8) with the cusp beaming direction defined relative≡ to the line of sight by rˆc · rˆ = cos ι. We then have Calculating each iteration of equation (B1) is then reduced to writing down the appropriate linear combination of cos ι0 rˆc · rˆ0 = cos θ0 cos ι cos φ0 sin θ0 sin ι, (B4) Cn(ι) for different n. ≡ − 23

1.00 series and the binomial theorem, n = 0

0.75 n = 1 1 X∞ k k = cos ι cos φ n = 2 1 cos ι cos φ . k=0 0.50 . − (C3) k ! k n = 20 X∞ X k cos ι i(k 2m)φ = e− − . 2k 0.25 k=0 m=0 m ) ι (

n 0.00 This converges everywhere on ι [ π/2, +π/2] except for

K ι = 0, in which case we use equation∈ −(C2) instead. We can 0.25 then integrate term-by-term to extract the contribution − from each order in the series. The result, valid for all 0.50 n 0, is − ≥ K (ι) = 0.75 n ! 2 − X∞ 2k + n 2 cos ι2k+n 21 + sin ι − − 1.00 k 2 2 − 1.0 0.5 0.0 0.5 1.0 k=max(0,2 n) − − sin ι − ! X∞ 2k + n cos ι2k+n+2 2 − 2 Figure 9. The angular integral Kn(ι) for non-negative n, as k=0 k given by equation (C5). The corresponding curves for negative ! 2 X∞ 2k + n + 2 cos ι2k+n+21 sin ι n are obtained by reflecting around ι = 0. + − . 2 2 k=0 k (C4) The resulting expressions for Sn(ι) and Ln(ι) for the first few n 2 are shown in TablesII andI respectively. The first sum must be carried out separately for the three We find empirically≥ that the large-loop angular functions cases n = 0, n = 1, and n > 1, yielding the following can be approximated by expressions:

2n−2 2 4 sin ι + cos 2ι 3 Ln(ι) 5(2/5) sin ι (B9) K0(ι) = | | − , | | ≈ 2 cos2 ι sin ι 1 with an accuracy of 10%, as is illustrated in the left K1(ι) = | | − cos ι ∼ 2 sin ι panel of figure8. The small-loop angular functions Sn(ι) | | do not seem to follow such a simple pattern. 1 sin ι + sin ι(1 + 2 sin ι + sin ι ) + (sin ι 1) − | | | | , − cos3 ι 4 sin ι cos ι n  Appendix C: Angular integrals for kink memory 2 , ι > 0 Kn(ι) = cos ι 1 + sin ι , for n > 1. 0, ι 0 ≤ Here we compute the integral Kn(ι) defined in equa- (C5) tion (67) for all angular modes n Z. Note that while ∈ the integrand is complex, the integral itself is always real, These clearly agree with equation (C2) for ι = π/2. Note ± as the imaginary part of the integrand is an odd function that for all n, the integral Kn(ι) is not differentiable at ι = of φ. Note also that we can focus on non-negative n 0 0; this is related to the fact that the series in equation (C3) ≥ by exploiting the symmetry property diverges at ι = 0. However, despite this formal divergence, we see that equation (C5) agrees with equation (C2) in K n(ι) = Kn( ι). (C1) the limit ι 0, whether this limit is taken from above − − or from below.→ We therefore use equation (C5) over the For some special values of ι we can obtain the full full domain ι [ π/2, +π/2]. The resulting curves for spectrum immediately from equation (67), the first few non-negative∈ − n are illustrated in figure9. For n 2, the integral only has support for ι > 0, and 1 peaks at≥ an inclination ι given by Kn(0) = δn,0 (δn,1 + δn, 1), ∗ − − 2 − (C2) 1 p 2 Kn( π/2) = δn, 2. sin ι = (n n 4). (C6) ± ± ∗ 2 − − For general values ι [ π/2, +π/2], we proceed by expand- When calculating the kink memory signal we are interested ing the denominator∈ of− equation (67) using the geometric in the high-frequency regime, which corresponds to large 24

Table I. The first few large-loop angular functions, as defined by equations (B1) and (B2). Note that Ln(ι) is a polynomial in cos ι of order 2n−1, such that the number of terms grows exponentially with n. Despite this apparent complexity, we find that these formulae can be approximated by the simple expression (B9).

n Ln(ι) 2 2 sin2 ι 2 sin2 ι 3 − (5 − cos 2ι) 15 sin2 ι 4 [3737 cos 2ι − 262 cos 4ι + 7(−2070 + cos 6ι)] 141750 sin2 ι  5 91170878299511 cos 2ι − 8844505203863 cos 4ι + 627703754313 cos 6ι − 31676232034 cos 8ι 123092512042500000  +1106354755 cos 10ι + 1001(−326445712830 − 24241 cos 12ι + 245 cos 14ι) sin2 ι  6 2337436144549410554623113364688535605014856 cos 2ι 4375937836474054406594550954572130000000000000000 −252383762462012251730735377642370477964415 cos 4ι + 22676810284676710047197528844734128748260 cos 6ι −1739064548768558706903275538411759890046 cos 8ι + 114707816611411474747133898122922244740 cos 10ι −6538676998608577029462451915641765109 cos 12ι + 322110616165743776447552901776475040 cos 14ι −13638911209793535304662793111510580 cos 16ι + 490882191972236519729403447127344 cos 18ι +1463(−10081167364753857595829662265 cos 20ι + 246246861424248720285801892 cos 22ι +1495(−3145080481324598891074 cos 24ι + 44031911264492955804 cos 26ι  +6525(−569906751862925826959924173050076 − 61436527080929 cos 28ι + 271199768840 cos 30ι)))

Table II. The first few small-loop angular functions, as defined by equations (B1) and (B2). Note that Sn(ι) is a polynomial in cos ι of order n.

n Sn(ι) 2 2 sin2 ι 2 sin2 ι 3 − (5 + 3 cos ι) 5 2 sin2 ι 4 (25 + 24 cos ι + 3 cos 2ι) 25 2 sin2 ι 5 − (154 + 201 cos ι + 42 cos 2ι + 3 cos 3ι) 175 sin2 ι 6 (15575 + 28032 cos ι + 7932 cos 2ι + 960 cos 3ι + 45 cos 4ι) 12250 sin2 ι 7 − [466422 cos ι + 172824 cos 2ι + 5(29702 + 5619 cos 3ι + 450 cos 4ι + 15 cos 5ι)] 245000 sin2 ι 8 (−30194 + 299024 cos ι + 157771 cos 2ι + 32552 cos 3ι + 3490 cos 4ι + 200 cos 5ι + 5 cos 6ι) 245000 sin2 ι 9 − [4934669 cos ι + 6432382 cos 2ι + 5(−2137388 + 347047 cos 3ι + 46508 cos 4ι + 3565 cos 5ι + 154 cos 6ι + 3 cos 7ι)] 13475000 sin2 ι  10 − 5257311104 cos ι + 2472686192 cos 2ι + 5(214599584 cos 3ι + 36187180 cos 4ι 10375750000  +7(−383512899 + 491360 cos 5ι + 28400 cos 6ι + 960 cos 7ι + 15 cos 8ι)) sin2 ι  11 − − 2359095296230 cos ι − 55089058544 cos 2ι + 116827655852 cos 3ι + 27832319320 cos 4ι 1888386500000  +3261899900 cos 5ι + 49(−59685302066 + 4749680 cos 6ι + 214995 cos 7ι + 5850 cos 8ι + 75 cos 9ι) sin2 ι  12 − 213800187213104 cos ι − 34153675430026 cos 2ι + 5(−37491558286234 + 292405153376 cos 3ι 122745122500000 +238091184664 cos 4ι + 37145938400 cos 5ι + 3229010267 cos 6ι  +179962104 cos 7ι + 6563970 cos 8ι + 147000 cos 9ι + 1575 cos 10ι) 25 n. In the limit n , we have and is only observable very close to (but strictly outside → ∞ of) the plane of the kink. ι 1/n, Kn(ι ) 4/(en), (C7) ∗ ' ∗ ' meaning that the memory signal is strongly suppressed,

[1] B. P. Abbott et al. (LIGO Scientific Collaboration, 69, 971–1144 (2006), arXiv:astro-ph/0509456. Virgo Collaboration), “Tests of general relativity with [17] Christian D. Ott, “The Gravitational Wave Signature GW150914,” Phys. Rev. Lett. 116, 221101 (2016), of Core-Collapse Supernovae,” Class. Quant. Grav. 26, [Erratum: Phys. Rev. Lett. 121, 129902 (2018)], 063001 (2009), arXiv:0809.0695 [astro-ph]. arXiv:1602.03841 [gr-qc]. [18] Ehud B. Segalis and Amos Ori, “Emission of gravit- [2] Nicolas Yunes, Kent Yagi, and Frans Pretorius, “The- ational radiation from ultrarelativistic sources,” Phys. oretical Physics Implications of the Binary Black-Hole Rev. D 64, 064018 (2001), arXiv:gr-qc/0101117. Mergers GW150914 and GW151226,” Phys. Rev. D 94, [19] Norichika Sago, Kunihito Ioka, Takashi Nakamura, and 084002 (2016), arXiv:1603.08955 [gr-qc]. Ryo Yamazaki, “Gravitational wave memory of gamma- [3] R. Abbott et al. (LIGO Scientific Collaboration, Virgo ray burst jets,” Phys. Rev. D 70, 104012 (2004), Collaboration), “Tests of general relativity with binary arXiv:gr-qc/0405067. black holes from the second LIGO-Virgo gravitational- [20] Ofek Birnholtz and Tsvi Piran, “Gravitational wave wave transient catalog,” Phys. Rev. D 103, 122002 memory from gamma ray bursts’ jets,” Phys. Rev. D (2021), arXiv:2010.14529 [gr-qc]. 87, 123007 (2013), arXiv:1302.5713 [astro-ph.HE]. [4] V. B. Braginsky and L. P. Grishchuk, “Kinematic Res- [21] Shota Akiba, Megumi Nakada, Chiyo Yamaguchi, and onance and Memory Effect in Free Mass Gravitational Koichi Iwamoto, “Gravitational Wave Memory from the Antennas,” Sov. Phys. JETP 62, 427–430 (1985). Relativistic Jet of Gamma-Ray Bursts,” Publ. Astron. [5] Vladimir B. Braginsky and Kip S. Thorne, Soc. Jap. 65, 59 (2013), arXiv:1303.6566 [astro-ph.HE]. “Gravitational-wave bursts with memory and experi- [22] Alexander Tolish and Robert M. Wald, “Retarded Fields mental prospects,” Nature 327, 123–125 (1987). of Null Particles and the Memory Effect,” Phys. Rev. D [6] Marc Favata, “The gravitational-wave memory effect,” 89, 064008 (2014), arXiv:1401.5831 [gr-qc]. Class. Quant. Grav. 27, 084036 (2010), arXiv:1003.3486 [23] Alexander Tolish, Lydia Bieri, David Garfinkle, and [gr-qc]. Robert M. Wald, “Examination of a simple example of [7] Y. B. Zel’dovich and A. G. Polnarev, “Radiation of gravitational wave memory,” Phys. Rev. D 90, 044060 gravitational waves by a cluster of superdense stars,” (2014), arXiv:1405.6396 [gr-qc]. Sov. Astron. 18, 17 (1974). [24] Bruce Allen, “Gravitational wave stochastic background [8] Larry Smarr, “Gravitational Radiation from Distant En- from cosmological particle decay,” Phys. Rev. Res. 2, counters and from Headon Collisions of Black Holes: The 012034 (2020), arXiv:1910.08213 [gr-qc]. Zero Frequency Limit,” Phys. Rev. D 15, 2069–2077 [25] Gregory M. Harry (LIGO Scientific Collaboration), “Ad- (1977). vanced LIGO: The next generation of gravitational wave [9] Michael Turner, “Gravitational radiation from point- detectors,” Class. Quant. Grav. 27, 084006 (2010). masses in unbound orbits: Newtonian results,” Astro- [26] J. Aasi et al. (LIGO Scientific Collaboration), “Ad- phys. J. 216, 610–619 (1977). vanced LIGO,” Class. Quant. Grav. 32, 074001 (2015), [10] Michael Turner and Clifford M. Will, “Post-Newtonian arXiv:1411.4547 [gr-qc]. gravitational bremsstrahlung,” Astrophys. J. 220, 1107– [27] F. Acernese et al. (Virgo Collaboration), “Advanced 1124 (1978). Virgo: a second-generation interferometric gravitational [11] Sándor J. Kovács and Kip S. Thorne, “The Generation wave detector,” Class. Quant. Grav. 32, 024001 (2015), of Gravitational Waves. 4. Bremsstrahlung,” Astrophys. arXiv:1408.3978 [gr-qc]. J. 224, 62–85 (1978). [28] R. Abbott et al. (LIGO Scientific Collaboration, Virgo [12] Robert J. Bontz and Richard H. Price, “The spectrum Collaboration), “GWTC-2: Compact Binary Coales- of radiation at low frequencies,” Astrophys. J. 228, cences Observed by LIGO and Virgo During the First 560–575 (1979). Half of the Third Observing Run,” Phys. Rev. X 11, [13] Reuben Epstein, “The Generation of Gravitational Radi- 021053 (2021), arXiv:2010.14527 [gr-qc]. ation by Escaping Supernova Neutrinos,” Astrophys. J. [29] Marc Favata, “Gravitational-wave memory revisited: 223, 1037–1045 (1978). memory from the merger and recoil of binary black holes,” [14] Michael S. Turner, “Gravitational Radiation from Su- J. Phys. Conf. Ser. 154, 012043 (2009), arXiv:0811.3451 pernova Neutrino Bursts,” Nature 274, 565–566 (1978). [astro-ph]. [15] Adam Burrows and John Hayes, “Pulsar recoil and grav- [30] Manuela Campanelli, Carlos O. Lousto, Yosef Zlochower, itational radiation due to asymmetrical stellar collapse and David Merritt, “Maximum gravitational recoil,” Phys. and explosion,” Phys. Rev. Lett. 76, 352–355 (1996), Rev. Lett. 98, 231102 (2007), arXiv:gr-qc/0702133. arXiv:astro-ph/9511106. [31] Demetrios Christodoulou, “Nonlinear nature of gravit- [16] Kei Kotake, Katsuhiko Sato, and Keitaro Takahashi, ation and gravitational wave experiments,” Phys. Rev. “Explosion mechanism, neutrino burst, and gravitational Lett. 67, 1486–1489 (1991). wave in core-collapse supernovae,” Rept. Prog. Phys. [32] Luc Blanchet and Thibault Damour, “Hereditary effects 26

in gravitational radiation,” Phys. Rev. D 46, 4304–4319 memory for binary coalescence events,” Phys. Rev. D (1992). 103, 044012 (2021), arXiv:2009.06351 [gr-qc]. [33] Kip S. Thorne, “Gravitational-wave bursts with memory: [52] D. Kennefick, “Prospects for detecting the Christodoulou The Christodoulou effect,” Phys. Rev. D 45, 520–524 memory of gravitational waves from a coalescing compact (1992). binary and using it to measure neutron star radii,” Phys. [34] Marc Favata, “Nonlinear gravitational-wave memory Rev. D 50, 3587–3595 (1994). from binary black hole mergers,” Astrophys. J. Lett. [53] Rutger van Haasteren and Yuri Levin, “Gravitational- 696, L159–L162 (2009), arXiv:0902.3660 [astro-ph.SR]. wave memory and pulsar timing arrays,” Mon. Not. [35] Lydia Bieri and David Garfinkle, “An electromagnetic Roy. Astron. Soc. 401, 2372 (2010), arXiv:0909.0954 analogue of gravitational wave memory,” Class. Quant. [astro-ph.IM]. Grav. 30, 195009 (2013), arXiv:1307.5098 [gr-qc]. [54] Naoki Seto, “Search for Memory and Inspiral Gravita- [36] Leonard Susskind, “Electromagnetic Memory,” (2015), tional Waves from Super-Massive Binary Black Holes arXiv:1507.02584 [hep-th]. with Pulsar Timing Arrays,” Mon. Not. Roy. Astron. [37] Monica Pate, Ana-Maria Raclariu, and Andrew Soc. 400, L38 (2009), arXiv:0909.1379 [astro-ph.CO]. Strominger, “Color Memory: A Yang-Mills Analog of [55] J. M. Cordes and F. A. Jenet, “Detecting gravitational Gravitational Wave Memory,” Phys. Rev. Lett. 119, wave memory with pulsar timing,” Astrophys. J. 752, 261602 (2017), arXiv:1707.08016 [hep-th]. 54 (2012). [38] Temple He, Vyacheslav Lysov, Prahar Mitra, and An- [56] D. R. Madison, J. M. Cordes, and S. Chatterjee, “As- drew Strominger, “BMS supertranslations and Wein- sessing Pulsar Timing Array Sensitivity to Gravitational berg’s soft graviton theorem,” JHEP 05, 151 (2015), Wave Bursts with Memory,” Astrophys. J. 788, 141 arXiv:1401.7026 [hep-th]. (2014), arXiv:1404.5682 [astro-ph.HE]. [39] Andrew Strominger and Alexander Zhiboedov, “Gravita- [57] J. B. Wang, G. Hobbs, W. Coles, R. M. Shannon, tional Memory, BMS Supertranslations and Soft Theor- X. J. Zhu, D. R. Madison, M. Kerr, V. Ravi, M. J. ems,” JHEP 01, 086 (2016), arXiv:1411.5745 [hep-th]. Keith, R. N. Manchester, Y. Levin, M. Bailes, N. D. R. [40] Sabrina Pasterski, Andrew Strominger, and Alexander Bhat, S. Burke-Spolaor, S. Dai, S. Osłowski, W. van Zhiboedov, “New Gravitational Memories,” JHEP 12, Straten, L. Toomey, N. Wang, and L. Wen, “Searching 053 (2016), arXiv:1502.06120 [hep-th]. for gravitational wave memory bursts with the Parkes [41] Sabrina Pasterski, “Asymptotic Symmetries and Pulsar Timing Array,” Mon. Not. Roy. Astron. Soc. Electromagnetic Memory,” JHEP 09, 154 (2017), 446, 1657–1671 (2015), arXiv:1410.3323 [astro-ph.GA]. arXiv:1505.00716 [hep-th]. [58] Z. Arzoumanian, A. Brazier, S. Burke-Spolaor, S. J. [42] Daniel Kapec, Vyacheslav Lysov, Sabrina Pasterski, and Chamberlin, S. Chatterjee, B. Christy, J. M. Cordes, Andrew Strominger, “Higher-dimensional supertransla- N. J. Cornish, P. B. Demorest, X. Deng, T. Dolch, J. A. tions and Weinberg’s soft graviton theorem,” Ann. Math. Ellis, R. D. Ferdman, E. Fonseca, N. Garver-Daniels, Sci. Appl. 02, 69–94 (2017), arXiv:1502.07644 [gr-qc]. F. Jenet, G. Jones, V. M. Kaspi, M. Koop, M. T. Lam, [43] Éanna É. Flanagan and David A. Nichols, “Conserved T. J. W. Lazio, L. Levin, A. N. Lommen, D. R. Lorimer, charges of the extended Bondi-Metzner-Sachs algebra,” J. Luo, R. S. Lynch, D. R. Madison, M. A. McLaughlin, Phys. Rev. D 95, 044002 (2017), arXiv:1510.03386 S. T. McWilliams, D. J. Nice, N. Palliyaguru, T. T. [hep-th]. Pennucci, S. M. Ransom, X. Siemens, I. H. Stairs, D. R. [44] Andrew Strominger, Lectures on the Infrared Structure of Stinebring, K. Stovall, J. Swiggum, M. Vallisneri, R. van Gravity and Gauge Theory ( Press, Haasteren, Y. Wang, and W. W. Zhu (NANOGrav Col- 2018) arXiv:1703.05448 [hep-th]. laboration), “NANOGrav Constraints on Gravitational [45] Alan G. Wiseman and Clifford M. Will, “Christo- Wave Bursts with Memory,” Astrophys. J. 810, 150 doulou’s nonlinear gravitational wave memory: Eval- (2015), arXiv:1501.05343 [astro-ph.GA]. uation in the quadrupole approximation,” Phys. Rev. D [59] Paul D. Lasky, Eric Thrane, Yuri Levin, Jonathan 44, 2945–2949 (1991). Blackman, and Yanbei Chen, “Detecting gravitational- [46] Marc Favata, “Post-Newtonian corrections to the wave memory with LIGO: implications of GW150914,” gravitational-wave memory for quasi-circular, inspiralling Phys. Rev. Lett. 117, 061102 (2016), arXiv:1605.01415 compact binaries,” Phys. Rev. D 80, 024002 (2009), [astro-ph.HE]. arXiv:0812.0069 [gr-qc]. [60] Huan Yang and Denis Martynov, “Testing Gravitational [47] Denis Pollney and Christian Reisswig, “Gravitational Memory Generation with Compact Binary Mergers,” memory in binary black hole mergers,” Astrophys. J. Phys. Rev. Lett. 121, 071102 (2018), arXiv:1803.02429 Lett. 732, L13 (2011), arXiv:1004.4209 [gr-qc]. [gr-qc]. [48] Marc Favata, “The Gravitational-wave memory from [61] Aaron D. Johnson, Shasvath J. Kapadia, Andrew Os- eccentric binaries,” Phys. Rev. D 84, 124013 (2011), borne, Alex Hixon, and Daniel Kennefick, “Prospects arXiv:1108.3121 [gr-qc]. of detecting the nonlinear gravitational wave memory,” [49] David A. Nichols, “Spin memory effect for compact Phys. Rev. D 99, 044045 (2019), arXiv:1810.09563 binaries in the post-Newtonian approximation,” Phys. [gr-qc]. Rev. D 95, 084048 (2017), arXiv:1702.03300 [gr-qc]. [62] Kristina Islo, Joseph Simon, Sarah Burke-Spolaor, and [50] Colm Talbot, Eric Thrane, Paul D. Lasky, and Xavier Siemens, “Prospects for Memory Detection with Fuhui Lin, “Gravitational-wave memory: waveforms and Low-Frequency Gravitational Wave Detectors,” (2019), phenomenology,” Phys. Rev. D 98, 064031 (2018), arXiv:1906.11936 [astro-ph.HE]. arXiv:1807.00990 [astro-ph.HE]. [63] Atul K. Divakarla, Eric Thrane, Paul D. Lasky, [51] Neev Khera, Badri Krishnan, Abhay Ashtekar, and and Bernard F. Whiting, “Memory Effect or Cosmic Tommaso De Lorenzo, “Inferring the gravitational wave String? Classifying Gravitational-Wave Bursts with 27

Bayesian Inference,” Phys. Rev. D 102, 023010 (2020), [78] B. P. Abbott et al. (LIGO Scientific Collaboration, Virgo arXiv:1911.07998 [gr-qc]. Collaboration), “Constraints on cosmic strings using data [64] K. Aggarwal, Z. Arzoumanian, P. T. Baker, A. Brazier, from the first Advanced LIGO observing run,” Phys. Rev. P. R. Brook, S. Burke-Spolaor, S. Chatterjee, J. M. D 97, 102002 (2018), arXiv:1712.01168 [gr-qc]. Cordes, N. J. Cornish, F. Crawford, H. T. Cromartie, [79] B. P. Abbott et al. (LIGO Scientific Collaboration, K. Crowter, M. DeCesar, P. B. Demorest, T. Dolch, Virgo Collaboration), “Search for the isotropic stochastic J. A. Ellis, R. D. Ferdman, E. C. Ferrara, E. Fonseca, background using data from Advanced LIGO’s second N. Garver-Daniels, P. Gentile, D. Good, J. S. Hazboun, observing run,” Phys. Rev. D 100, 061101 (2019), A. M. Holgado, E. A. Huerta, K. Islo, R. Jennings, arXiv:1903.02886 [gr-qc]. G. Jones, M. L. Jones, D. L. Kaplan, L. Z. Kelley, [80] R. Abbott et al. (LIGO Scientific Collaboration, Virgo J. S. Key, M. T. Lam, T. J. W. Lazio, L. Levin, D. R. Collaboration, KAGRA Collaboration), “Upper Limits Lorimer, J. Luo, R. S. Lynch, D. R. Madison, M. A. on the Isotropic Gravitational-Wave Background from McLaughlin, S. T. McWilliams, C. M. F. Mingarelli, Advanced LIGO’s and Advanced Virgo’s Third Observing C. Ng, D. J. Nice, T. T. Pennucci, N. S. Pol, S. M. Run,” (2021), arXiv:2101.12130 [gr-qc]. Ransom, P. S. Ray, X. Siemens, J. Simon, R. Spiewak, [81] R. Abbott et al. (LIGO Scientific Collaboration, Virgo I. H. Stairs, D. R. Stinebring, K. Stovall, J. K. Swiggum, Collaboration, KAGRA Collaboration), “Constraints on S. R. Taylor, M. Vallisneri, R. van Haasteren, S. J. Cosmic Strings Using Data from the Third Advanced Vigeland, C. A. Witt, and W. W. Zhu (NANOGrav LIGO–Virgo Observing Run,” Phys. Rev. Lett. 126, Collaboration), “The NANOGrav 11 yr Data Set: Limits 241102 (2021), arXiv:2101.12248 [gr-qc]. on Gravitational Wave Memory,” Astrophys. J. 889, 38 [82] Paul D. Lasky, Chiara M. F. Mingarelli, Tristan L. (2020), arXiv:1911.08488 [astro-ph.HE]. Smith, John T. Giblin, Eric Thrane, Daniel J. Rear- [65] Moritz Hübner, Colm Talbot, Paul D. Lasky, and Eric don, Robert Caldwell, Matthew Bailes, N. D. Ramesh Thrane, “Measuring gravitational-wave memory in the Bhat, Sarah Burke-Spolaor, Shi Dai, James Dempsey, first LIGO/Virgo gravitational-wave transient catalog,” George Hobbs, Matthew Kerr, Yuri Levin, Richard N. Phys. Rev. D 101, 023011 (2020), arXiv:1911.12496 Manchester, Stefan Osłowski, Vikram Ravi, Pablo A. [astro-ph.HE]. Rosado, Ryan M. Shannon, Renée Spiewak, Willem van [66] Oliver M. Boersma, David A. Nichols, and Patricia Straten, Lawrence Toomey, Jingbo Wang, Linqing Wen, Schmidt, “Forecasts for detecting the gravitational-wave Xiaopeng You, and Xingjiang Zhu, “Gravitational-wave memory effect with Advanced LIGO and Virgo,” Phys. cosmology across 29 decades in frequency,” Phys. Rev. Rev. D 101, 083026 (2020), arXiv:2002.01821 [astro- X 6, 011035 (2016), arXiv:1511.05994 [astro-ph.CO]. ph.HE]. [83] Jose J. Blanco-Pillado, Ken D. Olum, and Xavier [67] Michael Ebersold and Shubhanshu Tiwari, “Search for Siemens, “New limits on cosmic strings from gravitational nonlinear memory from subsolar mass compact bin- wave observation,” Phys. Lett. B 778, 392–396 (2018), ary mergers,” Phys. Rev. D 101, 104041 (2020), arXiv:1709.02434 [astro-ph.CO]. arXiv:2005.03306 [gr-qc]. [84] N. Yonemaru, S. Kuroyanagi, G. Hobbs, K. Taka- [68] Lior M. Burko and Gaurav Khanna, “Climbing up the hashi, X.-J. Zhu, W. A. Coles, S. Dai, E. Howard, memory staircase: Equatorial zoom-whirl orbits,” Phys. R. Manchester, D. Reardon, C. Russell, R. M. Shannon, Rev. D 102, 084035 (2020), arXiv:2007.12545 [gr-qc]. N. Thyagarajan, R. Spiewak, and J.-B. Wang, “Search- [69] T. W. B. Kibble, “Topology of Cosmic Domains and ing for gravitational wave bursts from cosmic string cusps Strings,” J. Phys. A 9, 1387–1398 (1976). with the Parkes Pulsar Timing Array,” Mon. Not. Roy. [70] Alexander Vilenkin, “Cosmic Strings and Domain Walls,” Astron. Soc. 501, 701–712 (2021), arXiv:2011.13490 Phys. Rept. 121, 263–315 (1985). [gr-qc]. [71] M. B. Hindmarsh and T. W. B. Kibble, “Cosmic strings,” [85] Zaven Arzoumanian, Paul T. Baker, Harsha Blumer, Rept. Prog. Phys. 58, 477–562 (1995), arXiv:hep- Bence Bécsy, Adam Brazier, Paul R. Brook, ph/9411342. Sarah Burke-Spolaor, Shami Chatterjee, Siyuan Chen, [72] A. Vilenkin and E. P. S. Shellard, Cosmic Strings James M. Cordes, Neil J. Cornish, Fronefield Craw- and Other Topological Defects, Cambridge Monographs ford, H. Thankful Cromartie, Megan E. DeCesar, on Mathematical Physics (Cambridge University Press, Paul B. Demorest, Timothy Dolch, Justin A. Ellis, 2000). Elizabeth C. Ferrara, William Fiore, Emmanuel Fonseca, [73] Rachel Jeannerot, Jonathan Rocher, and Mairi Sakellari- Nathan Garver-Daniels, Peter A. Gentile, Deborah C. adou, “How generic is cosmic string formation in SUSY Good, Jeffrey S. Hazboun, A. Miguel Holgado, Kristina GUTs,” Phys. Rev. D 68, 103514 (2003), arXiv:hep- Islo, Ross J. Jennings, Megan L. Jones, Andrew R. ph/0308134. Kaiser, David L. Kaplan, Luke Zoltan Kelley, Joey Sha- [74] T. W. B. Kibble and Neil Turok, “Selfintersection of piro Key, Nima Laal, Michael T. Lam, T. Joseph W. Cosmic Strings,” Phys. Lett. B 116, 141–143 (1982). Lazio, Duncan R. Lorimer, Jing Luo, Ryan S. Lynch, [75] Neil Turok, “Grand Unified Strings and Galaxy Forma- Dustin R. Madison, Maura A. McLaughlin, Chiara tion,” Nucl. Phys. B 242, 520–541 (1984). M. F. Mingarelli, Cherry Ng, David J. Nice, Timothy T. [76] Thibault Damour and Alexander Vilenkin, “Gravitational Pennucci, Nihan S. Pol, Scott M. Ransom, Paul S. Ray, wave bursts from cusps and kinks on cosmic strings,” Brent J. Shapiro-Albert, Xavier Siemens, Joseph Simon, Phys. Rev. D 64, 064008 (2001), arXiv:gr-qc/0104026. Renée Spiewak, Ingrid H. Stairs, Daniel R. Stinebring, [77] B. P. Abbott et al. (LIGO Scientific Collaboration), Kevin Stovall, Jerry P. Sun, Joseph K. Swiggum, “First LIGO search for gravitational wave bursts from Stephen R. Taylor, Jacob E. Turner, Michele Vallis- cosmic (super)strings,” Phys. Rev. D 80, 062002 (2009), neri, Sarah J. Vigeland, and Caitlin A. Witt (NANO- arXiv:0904.4718 [astro-ph.CO]. Grav Collaboration), “The NANOGrav 12.5 yr Data Set: 28

Search for an Isotropic Stochastic Gravitational-wave 42, 2505–2520 (1990). Background,” Astrophys. J. Lett. 905, L34 (2020), [104] Edmund J. Copeland, D. Haws, and M. Hindmarsh, arXiv:2009.04496 [astro-ph.HE]. “Classical theory of radiating strings,” Phys. Rev. D 42, [86] John Ellis and Marek Lewicki, “Cosmic String Inter- 726–730 (1990). pretation of NANOGrav Pulsar Timing Data,” Phys. [105] R. A. Battye and E. P. S. Shellard, “String radiative Rev. Lett. 126, 041304 (2021), arXiv:2009.06555 [astro- back reaction,” Phys. Rev. Lett. 75, 4354–4357 (1995), ph.CO]. arXiv:astro-ph/9408078. [87] Simone Blasi, Vedran Brdar, and Kai Schmitz, “Has [106] Alessandra Buonanno and Thibault Damour, “Gravita- NANOGrav found first evidence for cosmic strings?” tional, dilatonic and axionic radiative damping of cosmic Phys. Rev. Lett. 126, 041305 (2021), arXiv:2009.06607 strings,” Phys. Rev. D 60, 023517 (1999), arXiv:gr- [astro-ph.CO]. qc/9801105. [88] Wilfried Buchmuller, Valerie Domcke, and Kai Schmitz, [107] Brandon Carter and Richard A. Battye, “Nondivergence “From NANOGrav to LIGO with metastable cos- of gravitational selfinteractions for Goto-Nambu strings,” mic strings,” Phys. Lett. B 811, 135914 (2020), Phys. Lett. B 430, 49–53 (1998), arXiv:hep-th/9803012. arXiv:2009.10649 [astro-ph.CO]. [108] Jeremy M. Wachter and Ken D. Olum, “Gravitational [89] Ligong Bian, Rong-Gen Cai, Jing Liu, Xing-Yu Yang, smoothing of kinks on cosmic string loops,” Phys. Rev. and Ruiyu Zhou, “Evidence for different gravitational- Lett. 118, 051301 (2017), [Erratum: Phys.Rev.Lett. wave sources in the NANOGrav dataset,” Phys. Rev. D 121, 149901 (2018)], arXiv:1609.01153 [gr-qc]. 103, L081301 (2021), arXiv:2009.13893 [astro-ph.CO]. [109] Jeremy M. Wachter and Ken D. Olum, “Gravitational [90] Jose J. Blanco-Pillado, Ken D. Olum, and Jeremy M. backreaction on piecewise linear cosmic string loops,” Wachter, “Comparison of cosmic string and superstring Phys. Rev. D 95, 023519 (2017), arXiv:1609.01685 models to NANOGrav 12.5-year results,” Phys. Rev. D [gr-qc]. 103, 103512 (2021), arXiv:2102.08194 [astro-ph.CO]. [110] Jose J. Blanco-Pillado, Ken D. Olum, and Jeremy M. [91] Josu C. Aurrekoetxea, Thomas Helfer, and Eugene A. Wachter, “Gravitational backreaction near cosmic string Lim, “Coherent Gravitational Waveforms and Memory kinks and cusps,” Phys. Rev. D 98, 123507 (2018), from Cosmic String Loops,” Class. Quant. Grav. 37, arXiv:1808.08254 [gr-qc]. 204001 (2020), arXiv:2002.05177 [gr-qc]. [111] David F. Chernoff, Éanna É. Flanagan, and [92] Lucy O. McNeill, Eric Thrane, and Paul D. Lasky, Barry Wardell, “Gravitational backreaction on a cosmic “Detecting Gravitational Wave Memory without Par- string: Formalism,” Phys. Rev. D 99, 084036 (2019), ent Signals,” Phys. Rev. Lett. 118, 181103 (2017), arXiv:1808.08631 [gr-qc]. arXiv:1702.01759 [astro-ph.IM]. [112] Jose J. Blanco-Pillado, Ken D. Olum, and Jeremy M. [93] Alexander C. Jenkins and Mairi Sakellariadou, “Prim- Wachter, “Gravitational backreaction simulations of ordial black holes from cusp collapse on cosmic strings,” simple cosmic string loops,” Phys. Rev. D 100, 023535 (2020), arXiv:2006.16249 [astro-ph.CO]. (2019), arXiv:1903.06079 [gr-qc]. [94] Michele Maggiore, Gravitational Waves. Vol. 1: The- [113] Jose J. Blanco-Pillado, Ken D. Olum, and Benjamin ory and Experiments, Oxford Master Series in Physics Shlaer, “The number of cosmic string loops,” Phys. Rev. (Oxford University Press, 2007). D 89, 023512 (2014), arXiv:1309.6637 [astro-ph.CO]. [95] Richard A. Isaacson, “Gravitational Radiation in the [114] Eric Thrane and Joseph D. Romano, “Sensitivity curves Limit of High Frequency. I. The Linear Approximation for searches for gravitational-wave backgrounds,” Phys. and Geometrical Optics,” Phys. Rev. 166, 1263–1271 Rev. D 88, 124032 (2013), arXiv:1310.5300 [astro- (1968). ph.IM]. [96] Richard A. Isaacson, “Gravitational Radiation in the [115] R. Abbott et al. (LIGO Scientific Collaboration, Virgo Limit of High Frequency. II. Nonlinear Terms and the Collaboration, KAGRA Collaboration), “Data for ‘Upper Ef fective Stress Tensor,” Phys. Rev. 166, 1272–1279 Limits on the Isotropic Gravitational-Wave Background (1968). from Advanced LIGO’s and Advanced Virgo’s Third [97] Peter R. Saulson, Fundamentals of interferometric grav- Observing Run’,” https://dcc.ligo.org/G2001287/ itational wave detectors (World Scientific, 1995). public (2021). [98] Christophe Ringeval and Teruaki Suyama, “Stochastic [116] R. M. Shannon, V. Ravi, L. T. Lentati, P. D. Lasky, gravitational waves from cosmic string loops in scaling,” G. Hobbs, M. Kerr, R. N. Manchester, W. A. Coles, JCAP 12, 027 (2017), arXiv:1709.03845 [astro-ph.CO]. Y. Levin, M. Bailes, N. D. R. Bhat, S. Burke-Spolaor, [99] Mark Srednicki and Stefan Theisen, “Nongravitational S. Dai, M. J. Keith, S. Osłowski, D. J. Reardon, W. van Decay of Cosmic Strings,” Phys. Lett. B 189, 397 Straten, L. Toomey, J.-B. Wang, L. Wen, J. S. B. (1987). Wyithe, and X.-J. Zhu, “Gravitational waves from bin- [100] Thomas Helfer, Josu C. Aurrekoetxea, and Eu- ary supermassive black holes missing in pulsar observa- gene A. Lim, “Cosmic String Loop Collapse in Full tions,” Science 349, 1522–1525 (2015), arXiv:1509.07320 General Relativity,” Phys. Rev. D 99, 104028 (2019), [astro-ph.CO]. arXiv:1808.06678 [gr-qc]. [117] J. P. W. Verbiest, L. Lentati, G. Hobbs, R. van [101] Robert V. Wagoner, “Low Frequency Gravitational Rad- Haasteren, P. B. Demorest, G. H. Janssen, J.-B. Wang, tiation From Collapsing Systems,” Phys. Rev. D 19, G. Desvignes, R. N. Caballero, M. J. Keith, D. J. Cham- 2897–2901 (1979). pion, Z. Arzoumanian, S. Babak, C. G. Bassa, N. D. R. [102] Arthur Christopher Thompson, “Dynamics of Cosmic Bhat, A. Brazier, P. Brem, M. Burgay, S. Burke-Spolaor, String,” Phys. Rev. D 37, 283–297 (1988). S. J. Chamberlin, S. Chatterjee, B. Christy, I. Cognard, [103] Jean M. Quashnock and David N. Spergel, “Gravita- J. M. Cordes, S. Dai, T. Dolch, J. A. Ellis, R. D. Ferd- tional Selfinteractions of Cosmic Strings,” Phys. Rev. D man, E. Fonseca, J. R. Gair, N. E. Garver-Daniels, 29

P. Gentile, M. E. Gonzalez, E. Graikou, L. Guille- Gravitational Wave Observatories,” Class. Quant. Grav. mot, J. W. T. Hessels, G. Jones, R. Karuppusamy, 28, 094013 (2011), arXiv:1012.0908 [gr-qc]. M. Kerr, M. Kramer, M. T. Lam, P. D. Lasky, A. Las- [125] Evan D. Hall, Kevin Kuns, Joshua R. Smith, Yuntao sus, P. Lazarus, T. J. W. Lazio, K. J. Lee, L. Levin, Bai, Christopher Wipf, Sebastien Biscans, Rana X Ad- K. Liu, R. S. Lynch, A. G. Lyne, J. Mckee, M. A. hikari, Koji Arai, Stefan Ballmer, Lisa Barsotti, Yanbei McLaughlin, S. T. McWilliams, D. R. Madison, R. N. Chen, Matthew Evans, Peter Fritschel, Jan Harms, Brit- Manchester, C. M. F. Mingarelli, D. J. Nice, S. Os- tany Kamai, Jameson Graef Rollins, David Shoemaker, łowski, N. T. Palliyaguru, T. T. Pennucci, B. B. P. Bram Slagmolen, , and Hiro Yamamoto, Perera, D. Perrodin, A. Possenti, A. Petiteau, S. M. “Gravitational-wave physics with Cosmic Explorer: Lim- Ransom, D. Reardon, P. A. Rosado, S. A. Sanidas, its to low-frequency sensitivity,” Phys. Rev. D 103, A. Sesana, G. Shaifullah, R. M. Shannon, X. Siemens, 122004 (2021), arXiv:2012.03608 [gr-qc]. J. Simon, R. Smits, R. Spiewak, I. H. Stairs, B. W. [126] Christophe Ringeval, Mairi Sakellariadou, and Francois Stappers, D. R. Stinebring, K. Stovall, J. K. Swig- Bouchet, “Cosmological evolution of cosmic string loops,” gum, S. R. Taylor, G. Theureau, C. Tiburzi, L. Toomey, JCAP 02, 023 (2007), arXiv:astro-ph/0511646. M. Vallisneri, W. van Straten, A. Vecchio, Y. Wang, [127] Larissa Lorenz, Christophe Ringeval, and Mairi L. Wen, X. P. You, W. W. Zhu, and X.-J. Zhu, “The Sakellariadou, “Cosmic string loop distribution on all International Pulsar Timing Array: First Data Release,” length scales and at any redshift,” JCAP 10, 003 (2010), Mon. Not. Roy. Astron. Soc. 458, 1267–1288 (2016), arXiv:1006.0931 [astro-ph.CO]. arXiv:1602.03640 [astro-ph.IM]. [128] Pierre Auclair, Jose J. Blanco-Pillado, Daniel G. [118] Eric Thrane and Joseph D. Romano, “Sensitivity curves Figueroa, Alexander C. Jenkins, Marek Lewicki, Mairi for searches for gravitational-wave backgrounds,” https: Sakellariadou, Sotiris Sanidas, Lara Sousa, Danièle A. //dcc.ligo.org/LIGO-P1300115/public (2013). Steer, Jeremy M. Wachter, and Sachiko Kuroy- [119] Pau Amaro-Seoane et al. (LISA), “Laser Interfero- anagi, “Probing the gravitational wave background meter Space Antenna,” (2017), arXiv:1702.00786 [astro- from cosmic strings with LISA,” JCAP 04, 034 (2020), ph.IM]. arXiv:1909.00819 [astro-ph.CO]. [120] Chiara Caprini, Daniel G. Figueroa, Raphael Flauger, [129] A. Vilenkin, “Gravitational radiation from cosmic Germano Nardini, Marco Peloso, Mauro Pieroni, An- strings,” Phys. Lett. B 107, 47–50 (1981). gelo Ricciardone, and Gianmassimo Tasinato, “Recon- [130] Bruce Allen, “The Stochastic gravity wave background: structing the spectral shape of a stochastic gravitational Sources and detection,” in Les Houches School of Physics: wave background with LISA,” JCAP 11, 017 (2019), Astrophysical Sources of Gravitational Radiation (1996) arXiv:1906.09244 [astro-ph.CO]. pp. 373–417, arXiv:gr-qc/9604033. [121] Tristan L. Smith and Robert Caldwell, “LISA for Cos- [131] Michele Maggiore, “Gravitational wave experiments and mologists: Calculating the Signal-to-Noise Ratio for early universe cosmology,” Phys. Rept. 331, 283–367 Stochastic and Deterministic Sources,” Phys. Rev. D (2000), arXiv:gr-qc/9909001. 100, 104055 (2019), arXiv:1908.00546 [astro-ph.CO]. [132] Xavier Siemens, Vuk Mandic, and Jolien Creighton, [122] M. Punturo et al., “The Einstein Telescope: A “Gravitational wave stochastic background from cosmic third-generation gravitational wave observatory,” Class. (super)strings,” Phys. Rev. Lett. 98, 111101 (2007), Quant. Grav. 27, 194002 (2010). arXiv:astro-ph/0610920. [123] David Reitze, Rana X. Adhikari, Stefan Ballmer, Barry [133] Chiara Caprini and Daniel G. Figueroa, “Cosmological Barish, Lisa Barsotti, GariLynn Billingsley, Duncan A. Backgrounds of Gravitational Waves,” Class. Quant. Brown, Yanbei Chen, Dennis Coyne, Robert Eisenstein, Grav. 35, 163001 (2018), arXiv:1801.04268 [astro- Matthew Evans, Peter Fritschel, Evan D. Hall, Al- ph.CO]. bert Lazzarini, Geoffrey Lovelace, Jocelyn Read, B. S. [134] Alexander C. Jenkins and Mairi Sakellariadou, “Aniso- Sathyaprakash, David Shoemaker, Joshua Smith, Calum tropies in the stochastic gravitational-wave background: Torrie, Salvatore Vitale, Rainer Weiss, Christopher Formalism and the cosmic string case,” Phys. Rev. D Wipf, and Michael Zucker, “Cosmic Explorer: The 98, 063509 (2018), arXiv:1802.06046 [astro-ph.CO]. U.S. Contribution to Gravitational-Wave Astronomy [135] Pierre Auclair, Danièle A. Steer, and Tanmay beyond LIGO,” Bull. Am. Astron. Soc. 51, 035 (2019), Vachaspati, “Particle emission and gravitational radi- arXiv:1907.04833 [astro-ph.IM]. ation from cosmic strings: observational constraints,” [124] S. Hild et al., “Sensitivity Studies for Third-Generation Phys. Rev. D 101, 083511 (2020), arXiv:1911.12066 [hep-ph].