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

Search for continuous gravitational waves from neutron stars in five globular clusters with a phase-tracking hidden Markov model in the third LIGO observing run

L. Dunn School of Physics, University of Melbourne, Parkville, Victoria 3010, Australia OzGrav-Melbourne, Australian Research Council Centre of Excellence for Gravitational Wave Discovery, Parkville, Victoria 3010, Australia    A. Melatos School of Physics, University of Melbourne, Parkville, Victoria 3010, Australia OzGrav-Melbourne, Australian Research Council Centre of Excellence for Gravitational Wave Discovery, Parkville, Victoria 3010, Australia    P. Clearwater School of Physics, University of Melbourne, Parkville, Victoria 3010, Australia OzGrav-Melbourne, Australian Research Council Centre of Excellence for Gravitational Wave Discovery, Parkville, Victoria 3010, Australia    S. Suvorova School of Physics, University of Melbourne, Parkville, Victoria 3010, Australia OzGrav-Melbourne, Australian Research Council Centre of Excellence for Gravitational Wave Discovery, Parkville, Victoria 3010, Australia Department of Electrical and Electronic Engineering, University of Melbourne, Parkville, Victoria 3010, Australia    L. Sun OzGrav-ANU, Centre for Gravitational Astrophysics, College of Science, The Australian National University, Canberra ACT 2601, Australia    W. Moran Department of Electrical and Electronic Engineering, University of Melbourne, Parkville, Victoria 3010, Australia    R. J. Evans OzGrav-Melbourne, Australian Research Council Centre of Excellence for Gravitational Wave Discovery, Parkville, Victoria 3010, Australia Department of Electrical and Electronic Engineering, University of Melbourne, Parkville, Victoria 3010, Australia
(September 27, 2025)
Abstract

A search is performed for continuous gravitational waves emitted by unknown neutron stars in five nearby globular clusters using data from the third Laser Interferometer Gravitational-Wave Observatory (LIGO) observing run, over the frequency range 100100800Hz800\,\mathrm{Hz}. The search uses a hidden Markov model to track both the frequency and phase of the continuous wave signal from one coherent segment to the next. It represents the first time that a phase-tracking hidden Markov model has been used in a LIGO search. After applying vetoes to reject candidates consistent with non-Gaussian artifacts, no significant candidates are detected. Estimates of the strain sensitivity at 95% confidence h0,eff95%h_{0,\mathrm{eff}}^{95\%} and corresponding neutron star ellipticity ϵ95%\epsilon^{95\%} are presented. The best strain sensitivity, h0,eff95%=2.7×1026h_{0,\mathrm{eff}}^{95\%}=2.7\times 10^{-26} at 211Hz211\,\mathrm{Hz}, is achieved for the cluster NGC6544.

I Introduction

Continuous gravitational waves (CWs) are long-lasting and nearly monochromatic signals which may be emitted by a variety of sources, but which have so far remained undetected [1, 2]. In the frequency band covered by existing terrestrial interferometers, one promising class of CW sources is rapidly rotating neutron stars located within the Milky Way. There are many possible emission mechanisms through which neutron stars can emit gravitational waves, involving both the crust and the interior. These include deformations created by thermal [3, 4, 5, 6], magnetic [7, 8, 9, 10, 11, 12], and tectonic [13, 14, 15] effects, as well as fluid oscillations (particularly rr-modes) [16, 17, 18]. CW searches thus serve to probe the physics involved in these mechanisms.

Globular clusters are an interesting target for CW searches, as their dense cores provide a natural formation site for millisecond pulsars (MSPs), fast-spinning neutron stars which have been “recycled” by accretion from a binary companion [19, 20, 21, 22]. The gravitational wave strain scales as the square of the spin frequency for several important emission mechanisms [1]. Once recycled, MSPs may also experience further accretion episodes later in their life, as the stellar density in the cluster core enhances the likelihood of stellar encounters which disrupt the orbits of debris disks [23, 24, 25] or planets around the pulsar [26, 27, 28, 29, 30]. Two CW searches targeting globular clusters specifically have been carried out in the past [2]. Abbott et al. [24] carried out a search with 10 days of Initial Laser Interferometer Gravitational-Wave Observatory (LIGO) Science Run 6 data targeting unknown neutron stars in the globular cluster NGC 6544. Dergachev et al. [31] carried out a search with Advanced LIGO Observing Run 1 data targeting a small sky region containing both the Galactic Center and the globular cluster Terzan 5. Searches directed at the Galactic Center are broadly similar to globular cluster searches, being wide parameter space searches targeting an interesting sky region, and a number of these searches have been performed previously [32, 31, 33, 34]. Besides the two specific searches, globular cluster MSPs have been targeted in searches for known pulsars [35, 36, 37, 38, 39, 34].

This paper presents a search for CW emission from unknown neutron stars in five globular clusters in the Milky Way. We allow for the possibility that the phase model of the signal does not follow a simple Taylor expansion but includes some additional stochasticity (called “spin wandering”). The search uses a hidden Markov model (HMM) to track both the frequency and phase of the wandering signal [40, 41, 42]. HMMs have been used in a number of CW searches in the past [43, 44, 45, 46, 47, 48, 49, 50, 51, 52, 53, 54, 55] but previously tracked only the frequency, typically using the \mathcal{F}-statistic for isolated targets [56, 57] or the 𝒥\mathcal{J}-statistic for targets in binaries [58]. In contrast, the HMM implementation used in this paper, developed by Melatos et al. [42], employs a modified version of the \mathcal{B}-statistic [59, 42], which is a function of both frequency and phase. Phase tracking improves the sensitivity (by 10%\lesssim 10\% as it turns out) by downweighting noise features, that are not approximately phase-coherent over the search. We opt to use the phase-tracking HMM in this search both to take advantage of its enhanced sensitivity and as a technology demonstration in its own right. This is the first time that a phase-tracking HMM has been used in an astrophysical search, and its successful application here shows that the technique can be feasibly applied to searches over a wide parameter domain.

The structure of the paper is as follows. In Section II we describe the HMM framework in general, its implementation in this specific search, and the modified \mathcal{B}-statistic which serves as the coherent detection statistic. In Section III we discuss the setup of the search, including target selection and the parameter domain. In Section IV we discuss the candidates returned by the search, the veto procedures used to reject spurious candidates, and the results of applying the vetoes. In Section V we estimate the sensitivity of the search to CW emission from each cluster. In Section VI we summarise the results and look to future work.

II Algorithms

In this section we briefly recap the HMM formalism and its algorithmic components. We define the signal model (Section II.1), derive the form of the phase-sensitive \mathcal{B}-statistic which serves as the coherent detection statistic throughout this search (Section II.2), and explain how the \mathcal{B}-statistic is combined with phase-frequency transition probabilities as part of a HMM (Section II.3).

II.1 Signal model

Following Jaranowski et al. [56], the mass quadrupole gravitational wave signal from an isolated rotating neutron star emitting at twice the spin frequency can be written as a linear combination of four components

h(t)=Aμhμ(t),h(t)=A^{\mu}h_{\mu}(t), (1)

with μ{1,2,3,4}\mu\in\{1,2,3,4\}, where we adopt the Einstein summation convention. The AμA^{\mu} are the “amplitude parameters” given by Eqs. (32)–(35) in Jaranowski et al. [56]. They do not evolve and depend on the characteristic strain amplitude h0h_{0}, the cosine of the inclination angle cosι\cos\iota, the polarisation angle ψ\psi and the initial signal phase Φ0\Phi_{0}.111Strictly speaking the AμA^{\mu} also depend on the angle between the total angular momentum of the star and the axis of symmetry, θ\theta, as well as the angle between the interferometer arms, ζ\zeta [56]. The AμA^{\mu} are multiplied by a factor of sinζsin2θ\sin\zeta\sin^{2}\theta, but ζ=π/2\zeta=\pi/2 for the LIGO detectors, and we assume θ=π/2\theta=\pi/2 also, so this factor is unity. The hμ(t)h_{\mu}(t) are combinations of the antenna pattern functions [Eqs. (12) and (13) of Jaranowski et al. [56]] and the phase of the signal Φ(t)\Phi(t), whose evolution is expanded as a Taylor series, viz.

Φ(t)=2πk=0sf(k)tk+1(k+1)!+Φw(t)+ΦSSB(t).\Phi(t)=2\pi\sum_{k=0}^{s}f^{(k)}\frac{t^{k+1}}{(k+1)!}+\Phi_{\text{w}}(t)+\Phi_{\text{SSB}}(t). (2)

In Eq. (2), f(k)f^{(k)} is the kk-th time derivative of the frequency evaluated at t=0t=0, Φw(t)\Phi_{\text{w}}(t) is an additional stochastic wandering term, the nature of which is discussed in detail in Section II.3.2, and ΦSSB(t)\Phi_{\text{SSB}}(t) is the phase contribution from the transformation between the solar system barycentre (SSB) and the detector frame, cf. Eq. (14) of Jaranowski et al. [56]. For the form of hμ(t)h_{\mu}(t) see Eq. (36) in Ref. [56].

II.2 Phase-dependent \mathcal{B}-statistic

The \mathcal{B}-statistic was introduced by Prix and Krishnan [59] as a Bayesian alternative to the well-known \mathcal{F}-statistic [56]. The \mathcal{F}-statistic is obtained by analytically maximizing the likelihood function over AμA^{\mu} and hence h0h_{0}, cosι\cos\iota, ψ\psi and Φ0\Phi_{0}. The \mathcal{B}-statistic is obtained by instead marginalising over these parameters with a physically motivated prior which is uniform in h0h_{0}, cosι\cos\iota, ψ\psi, and Φ0\Phi_{0}. Melatos et al. [42] modified this approach to leave Φ0\Phi_{0} as a free parameter, so that we can compute the \mathcal{B}-statistic for many coherent segments and then track f(0)f^{(0)} and Φ0\Phi_{0} across these segments using a HMM as discussed in Section II.3. Here we briefly justify the form of the phase-dependent \mathcal{B}-statistic and discuss how to compute it in practice. The discussion draws on Refs. [56, 59, 42].

We begin with the log likelihood

lnΛ=x|h12h|h,\ln\Lambda=\innerproduct{x}{h}-\frac{1}{2}\innerproduct{h}{h}, (3)

where x(t)x(t) is the detector data and h(t)h(t) is the signal as defined in Eqs. (1) and (2). The inner product is defined as

x|y\displaystyle\innerproduct{x}{y} =4Re0dfx(f)y(f)Sh(f)\displaystyle=4\real\int_{0}^{\infty}\mathrm{d}f\,\frac{x(f)y^{*}(f)}{S_{h}(f)} (4)
2Sh(f0)0Tdtx(t)y(t),\displaystyle\approx\frac{2}{S_{h}(f_{0})}\int_{0}^{T}\mathrm{d}t\,x(t)y(t), (5)

where Sh(f)S_{h}(f) is the one-sided power spectral density of the noise at a frequency ff, TT is the total length of the data, and the approximate equality holds for finite-length signals localized spectrally to a narrow band around f0f_{0} in which Sh(f)Sh(f0)S_{h}(f)\approx S_{h}(f_{0}) can be taken as constant (which is a good approximation for the quasi-monochromatic signals of interest here [56]). Defining xμ=x|hμx_{\mu}=\innerproduct{x}{h_{\mu}} and hμ|hν=μν\innerproduct{h_{\mu}}{h_{\nu}}=\mathcal{M}_{\mu\nu} we write

lnΛ=Aμxμ12AμμνAν.\ln\Lambda=A^{\mu}x_{\mu}-\frac{1}{2}A^{\mu}\mathcal{M}_{\mu\nu}A^{\nu}. (6)

The phase-dependent \mathcal{B}-statistic is obtained by marginalising over cosι\cos\iota, ψ\psi, and h0h_{0}:

=11d(cosι)π/4π/4dψ0h0maxdh0p(cosι,ψ,h0)Λ.\mathcal{B}=\int_{-1}^{1}\mathrm{d}(\cos\iota)\int_{-\pi/4}^{\pi/4}\mathrm{d}\psi\int_{0}^{h_{0}^{\mathrm{max}}}\mathrm{d}h_{0}\,p(\cos\iota,\psi,h_{0})\Lambda. (7)

In Eq. (7), p(cosι,ψ,h0)=C/h0maxp(\cos\iota,\psi,h_{0})=C/h_{0}^{\mathrm{max}} is the prior, where h0maxh_{0}^{\mathrm{max}} is a large but arbitrary maximum h0h_{0} and CC is an irrelevant normalising constant. Since p(cosι,ψ,h0)p(\cos\iota,\psi,h_{0}) is a constant and can be taken outside the integral it is dropped going forward.

We now substitute Eq. (6) into Eq. (7) and perform the triple integral. Writing 𝒜μ=h0𝒜μ\mathcal{A}^{\mu}=h_{0}\mathcal{A}^{\prime\mu}, with 𝒜μ\mathcal{A}^{\prime\mu} independent of h0h_{0}, and collapsing the triple integral []\int[\ldots] for notational convenience, we obtain

=[]exp(h0𝒜μxμh022𝒜μμν𝒜ν).\mathcal{B}=\int[\ldots]\exp\left(h_{0}\mathcal{A}^{\prime\mu}x_{\mu}-\frac{h_{0}^{2}}{2}\mathcal{A}^{\prime\mu}\mathcal{M}_{\mu\nu}\mathcal{A}^{\prime\nu}\right). (8)

We use the following identity to perform the integral over h0h_{0}:

0Xdxexp(AxBx2)=(π4B)1/2exp(A24B)[erf(A2B)+erf(2BXA2B)].\int_{0}^{X}\mathrm{d}x\,\exp(Ax-Bx^{2})=\left(\frac{\pi}{4B}\right)^{1/2}\exp\left(\frac{A^{2}}{4B}\right)\left[\erf\left(\frac{A}{2\sqrt{B}}\right)+\erf\left(\frac{2BX-A}{2\sqrt{B}}\right)\right]. (9)

We also take h0maxh_{0}^{\mathrm{max}}, which is arbitrary, to be sufficiently large that the second erf()\erf(\ldots) term approaches unity. The result is

=11d(cosι)π/4π/4dψ(π2𝒜μμν𝒜ν)1/2exp[(𝒜μxμ)22𝒜μμν𝒜ν][erf(𝒜μxμ2𝒜μμν𝒜ν)+1].\mathcal{B}=\int_{-1}^{-1}\,\mathrm{d}(\cos\iota)\int_{-\pi/4}^{\pi/4}\,\mathrm{d}\psi\left(\frac{\pi}{2\mathcal{A}^{\prime\mu}\mathcal{M}_{\mu\nu}\mathcal{A}^{\prime\nu}}\right)^{1/2}\exp\left[\frac{(\mathcal{A}^{\prime\mu}x_{\mu})^{2}}{2\mathcal{A}^{\prime\mu}\mathcal{M}_{\mu\nu}\mathcal{A}^{\prime\nu}}\right]\left[\erf\left(\frac{\mathcal{A}^{\prime\mu}x_{\mu}}{\sqrt{2\mathcal{A}^{\prime\mu}\mathcal{M}_{\mu\nu}\mathcal{A}^{\prime\nu}}}\right)+1\right]. (10)

The ingredients of the integrand — 𝒜μ\mathcal{A}^{\prime\mu}, μν\mathcal{M}_{\mu\nu} and xμx_{\mu} — are all straightforward to compute using existing functionality in lalsuite, the standard software suite for analysis of LIGO, Virgo, and KAGRA data [60]. We use the existing graphics processing unit (GPU) framework [61] to compute the \mathcal{F}-statistic and hence xμx_{\mu}, the most expensive step. We also integrate Eq. (10) using GPUs, running a simple Simpson integrator on a 11×1111\times 11 grid of cosι\cos\iota and ψ\psi values. We verify using synthetic data that the relative error in the value of the \mathcal{B}-statistic when computed using an 11×1111\times 11 grid compared to a finer 501×501501\times 501 grid is typically small, with a mean fractional error of 1.9% and a fractional error exceeding 10% occurring in 1.2% of the 1.7×1071.7\times 10^{7} \mathcal{B}-statistic values (corresponding to five 0.5Hz0.5\,\mathrm{Hz} bands with a coherence time of five days and eight phase bins) computed for the purposes of this test.

When computing \mathcal{B} we ignore Φw(t)\Phi_{\text{w}}(t) in Eq. (2). By assumption this term is small compared to the other terms in Eq. (2) over each coherent segment during which \mathcal{B} is computed, amounting to 1rad\lesssim 1\,\mathrm{rad}. The stochastic wandering is tracked instead by the HMM, as discussed in Section II.3.2.

II.3 HMMs and the Viterbi algorithm

The number of signal templates needed to cover an astrophysically relevant parameter domain grows quickly for a phase-coherent search, as the observational timespan TobsT_{\text{obs}} increases. Hence a number of “semicoherent” techniques for CW searches have been developed, which perform phase-coherent computations on subdivided segments of length TcohT_{\text{coh}} and combine the segments without enforcing phase coherence between them. Most semicoherent methods completely discard the phase information between segments (see e.g. Refs. [40, 62, 63, 64]). Some, termed “loosely coherent”, retain some phase information but do not enforce perfect phase continuity [65, 66, 67]. While there is some loss of sensitivity inherent in this approach — the sensitivity of semicoherent searches scales as (TcohTobs)1/4(T_{\mathrm{coh}}T_{\mathrm{obs}})^{-1/4} [62, 68], compared to Tobs1/2T_{\mathrm{obs}}^{-1/2} for coherent searches — there is a large computational saving. Some semicoherent methods also offer flexibility in the face of the unexpected. For example, although fully coherent methods can accommodate deviations away from an assumed signal model through explicit searches over the parameters determining the phase evolution [69, 70, 71], if there is a stochastic component to the phase evolution which operates on timescales shorter than the observing timespan then a semicoherent approach may be needed. A HMM solved by the Viterbi algorithm to find the most likely sequence of signal frequencies f(t)f(t) is one such approach, which has been used in a number of CW searches in the past [43, 44, 45, 47, 46, 48, 49, 50, 51, 52, 53, 54, 55]. This approach was extended to include tracking of Φ(t)\Phi(t) by Melatos et al. [42], and is the subject of the present paper. In this section we review the basic framework of HMMs and the way it is put into practice for the present, loosely coherent search.

II.3.1 HMMs

HMMs model the evolution and observation of systems fulfilling two important criteria: a) the evolution of the internal state of the system is Markovian, and b) the internal state is hidden from the observer but may be probed probabilistically through measurement. This section briefly describes the formalism involved; for a more detailed review of HMMs see Rabiner [72].

The hidden state of the system is assumed here to be discretised and bounded, so that each state is in the set 𝒬={q1,q2,,qNQ}\mathcal{Q}=\{q_{1},q_{2},\ldots,q_{N_{Q}}\}, where NQN_{Q} is the total number of hidden states. Time is likewise discretised, with the HMM occupying the hidden state q(ti)𝒬q(t_{i})\in\mathcal{Q} at each timestep ti{t0,t1,t2,,tNT}t_{i}\in\{t_{0},t_{1},t_{2},\ldots,t_{N_{T}}\} ,where t0t_{0} is an “initial timestep” before any observations have been made. A sequence of possible hidden states occupied during a particular realisation of the HMM is denoted by Q=[q(t0),q(t1),q(t2),,q(tNT)]Q=[q(t_{0}),q(t_{1}),q(t_{2}),\ldots,q(t_{N_{T}})]. The dynamics of the system are captured by the transition matrix A(qj,qi)A(q_{j},q_{i}), whose entries are defined as

A(qj,qi)=Pr[q(tn+1)=qjq(tn)=qi].A(q_{j},q_{i})=\Pr[q(t_{n+1})=q_{j}\mid q(t_{n})=q_{i}]. (11)

At each timestep we assume that the system is observed. For a given realisation we write the sequence of observations as O=[o(t1),o(t2),,o(tNT)]O=[o(t_{1}),o(t_{2}),\ldots,o(t_{N_{T}})]. The set of possible observations need not be finite, discretised, or bounded. All we require is that there exists an emission probability function L(o,qi)L(o,q_{i}) defined as

L(o,qi)=Pr[o(tn)=oq(tn)=qi],L(o,q_{i})=\Pr[o(t_{n})=o\mid q(t_{n})=q_{i}], (12)

i.e. the probability of observing oo at the timestep tnt_{n}, if the hidden state qiq_{i} is occupied at tnt_{n}.

Finally we specify a belief about the state of the system at the beginning of the analysis, i.e.

Π(qi)=Pr[q(t0)=qi].\Pi(q_{i})=\Pr[q(t_{0})=q_{i}]. (13)

In this paper we have no reason to impose anything other than a uniform prior, viz. Π(qi)=NQ1\Pi(q_{i})=N_{Q}^{-1}.

Given the above ingredients, it is straightforward to compute the probability of a hidden state sequence QQ, given a sequence of observations OO:

Pr(QO)=Π[q(t0)]n=1NTL[o(tn),q(tn)]A[q(tn),q(tn1)].\Pr(Q\mid O)=\Pi[q(t_{0})]\prod_{n=1}^{N_{T}}L[o(t_{n}),q(t_{n})]A[q(t_{n}),q(t_{n-1})]. (14)

The goal is to compute argmaxQPr(QO)\operatorname*{arg\,max}_{Q}\Pr(Q\mid O).

II.3.2 Phase tracking

In this paper, the full observation is divided into segments of length Tcoh=Tobs/NTT_{\text{coh}}=T_{\mathrm{obs}}/N_{T}, which are analysed coherently. Each coherent segment is a timestep in the HMM. The hidden states are pairs of frequencies and phases at the start of a segment, q(tn)=[f(tn),Φ(tn)]q(t_{n})=[f(t_{n}),\Phi(t_{n})]. The phase-dependent \mathcal{B}-statistic, defined in Section II.2, plays the role of lnL(o,qi)\ln L(o,q_{i}). It then remains to specify the transition matrix. We follow Melatos et al. [42] and adopt the following model for the evolution of the signal frequency ff and phase Φ\Phi:

dfdt\displaystyle\derivative{f}{t} =σξ(t),\displaystyle=\sigma\xi(t), (15)
dΦdt\displaystyle\derivative{\Phi}{t} =f,\displaystyle=f, (16)

where ξ(t)\xi(t) is a zero-mean white noise term, i.e.

ξ(t)\displaystyle\langle\xi(t)\rangle =0\displaystyle=0 (17)
ξ(t)ξ(t)\displaystyle\langle\xi(t)\xi(t^{\prime})\rangle =δ(tt)\displaystyle=\delta(t-t^{\prime}) (18)

and σ\sigma parametrises the noise strength. The forward transition matrix A[(fj,Φj),(fi,Φi)]A[(f_{j},\Phi_{j}),(f_{i},\Phi_{i})] is readily obtained by solving the Fokker-Planck equation corresponding to Eqs. (15)–(18). For practical reasons it is useful to also know the backward transition matrix

Ab(qi,qj)=Pr[q(tn)=qiq(tn+1)=qj],A^{\mathrm{b}}(q_{i},q_{j})=\Pr[q(t_{n})=q_{i}\mid q(t_{n+1})=q_{j}], (19)

which is similarly obtained by solving the backward Fokker-Planck equation for this system.

Details of the transition matrices are given in Appendix A, and an illustrative figure is shown in Fig. 1.

Refer to caption
Refer to caption
Figure 1: Illustrative plots of the transition matrix A(qj,qi)=Pr[q(tn+1)=qjq(tn)=qi]A(q_{j},q_{i})=\Pr[q(t_{n+1})=q_{j}\mid q(t_{n})=q_{i}] according to the model described in Section II.3.2. The top panel shows a heatmap of A(qj,qi)A(q_{j},q_{i}) where qi=[Φ(tn),f(tn)]q_{i}=[\Phi(t_{n}),f(t_{n})] is an arbitrary initial state and qj=[Φ(tn+1),f(tn+1)]q_{j}=[\Phi(t_{n+1}),f(t_{n+1})] is the final state. The heatmap axes are ΔΦ=Φ(tn+1)Φ(tn)\Delta\Phi=\Phi(t_{n+1})-\Phi(t_{n}) and Δf=f(tn+1)f(tn)\Delta f=f(t_{n+1})-f(t_{n}) in units of turns (cycles of 2π2\pi radians) and bins [of width (2Tcoh)1(2T_{\mathrm{coh}})^{-1}] respectively. The bottom panel shows slices of A(qj,qi)A(q_{j},q_{i}) as a function of ΔΦ\Delta\Phi for Δf=1, 0,+1\Delta f=-1,\,0,\,+1 bins. Parameters: Tcoh=5dT_{\text{coh}}=5\,\mathrm{d}, σ=1.76×109s3/2\sigma=1.76\times 10^{-9}\,\mathrm{s}^{-3/2}.

The top panel of Fig. 1 shows a heatmap of A(qj,qi)A(q_{j},q_{i}) as a function of ΔΦ=Φ(tn+1)Φ(tn)\Delta\Phi=\Phi(t_{n+1})-\Phi(t_{n}) and Δf=f(tn+1)f(tn)\Delta f=f(t_{n+1})-f(t_{n}). As shown in Melatos et al. [42] and Appendix A, it is approximately a truncated Gaussian centred on (ΔΦ,Δf)=(0,0)(\Delta\Phi,\Delta f)=(0,0), with positive covariance between ΔΦ\Delta\Phi and Δf\Delta f. The bottom panel shows slices through the distribution at Δf=1,0,+1bins\Delta f=-1\,,0\,,+1\,\mathrm{bins}. These slices are the distributions used to compute the transition probabilities in the HMM. The Δf=±1\Delta f=\pm 1 bin distributions are identical up to the location of the peak and each contain approximately 20% of the total probability, while the Δf=0\Delta f=0 bin contains the other 60%. The transitions probabilities are broad in phase. All three distributions have similar widths. From this figure we see that the probability is concentrated in the same bin as the previous timestep, but some leakage into neighboring bins does occur.

Melatos et al. [42] included a damping term γf-\gamma f in Eq. (15). This approximates the tendency of the spin frequency of the neutron star to approach zero in the long term, i.e. it represents the long-term spin down process. In this search we explicitly search over the spin-down rate f(1)f^{(1)} as a parameter in the deterministic part of the signal model (see Sections III.5.1 and III.6.1), and assume that once the deterministic spin down is taken into account, the mean drift of the signal frequency is negligible. As such, we impose the condition γ(2fTcoh2)1\gamma\ll(2fT_{\text{coh}}^{2})^{-1} which implies |f(tn+1)f(tn)|(2Tcoh)1|f(t_{n+1})-f(t_{n})|\ll(2T_{\mathrm{coh}})^{-1} [42]. This renders the effect of γf-\gamma f in Eq. (15) negligible, and in this work we do not include it. The approximation is justified in more detail in Appendix A.

II.3.3 Optimal Viterbi path

With the HMM fully specified, we turn to the question of to use it in a CW search. To find signal candidates we use the Viterbi algorithm [72, 40] to find the most likely sequence of hidden states,

Q=argmaxQPr(QO).Q^{*}=\operatorname*{arg\,max}_{Q}\Pr(Q\mid O). (20)

More precisely, for each possible terminating frequency bin fif_{i} we find the most likely sequence of hidden states which terminates in that bin,

Q(fi)=argmaxQPr[QO,f(tNT)=fi].Q^{*}(f_{i})=\operatorname*{arg\,max}_{Q}\Pr[Q\mid O,f(t_{N_{T}})=f_{i}]. (21)

These paths are sorted according to their probability Pr[Q(fi)O]\Pr[Q^{*}(f_{i})\mid O]. Paths exceeding a predetermined threshold are marked for follow-up analysis. The determination of the threshold as a function of the desired false alarm probability is discussed in Section IV.1. To maximize numerical accuracy, we work with log probabilities so that the products in Eq. (14) become sums. Hence we threshold candidates according to their log likelihood =lnPr[Q(fi)O]\mathcal{L}=\ln\Pr[Q^{*}(f_{i})\mid O].

To maximize computational performance the HMM tracking step is performed on GPUs, as with the computation of the \mathcal{B}-statistic described in Section II.2.

III Setting up a globular cluster search

In this section we describe the astronomical targets and data inputs (Sections III.1 and III.2 respectively) for the globular cluster search performed in this paper, the choices of certain control parameters which dictate the behaviour of the HMM (Sections III.3 and III.4 respectively), and the astrophysical and computational motivations for the parameter domain and its associated grid spacing (Sections III.5 and III.6 respectively).

III.1 Target selection

Globular clusters are dense stellar environments, with correspondingly high rates of dynamical interactions. They are prodigious manufacturers of low-mass X-ray binaries (LMXBs) [19, 20, 22], which are the progenitors of MSPs, neutron stars which have been “recycled” to millisecond spin periods via accretion [21]. Although accretion from a binary companion offers several mechanisms to increase the mass quadrupole moment of a neutron star [73, 3, 74], the search in this paper is not tuned to sources in binaries, as the \mathcal{B}-statistic in Section II.2 does not include the effects of binary motion on the frequency modulation of the signal (doing so would require a blind search over the binary parameters, adding greatly to the cost of the search). Hence the ideal target in this paper is a neutron star that has spent some time in a binary system historically, so as to be recycled to spin frequencies in the LIGO band above 100Hz100\,\mathrm{Hz} (see Section III.5.1), but was disrupted since by a secondary encounter and became an isolated source.

The catalog of Milky Way globular clusters compiled by Harris [75] (2010 edition) contains parameters for 157 clusters. With limited resources, we are unable to search every known globular cluster for CW emission. Therefore we develop a simple ranking statistic to guide target selection. The rest of this section describes that statistic and the resulting target list.

Verbunt and Freire [76] introduced a parameter222Verbunt and Freire [76] labelled this parameter γ\gamma, here we relabel it as η\eta to avoid confusion with the γ\gamma introduced in Section II.3.2. η\eta which they refer to as the “single-binary encounter rate”. Roughly speaking it equals the reciprocal of the lifetime of a binary, before it is disrupted by a secondary encounter. It scales with globular cluster parameters as

ηρcrc,\eta\propto\frac{\sqrt{\rho_{\text{c}}}}{r_{\text{c}}}, (22)

where ρc\rho_{\text{c}} is the core luminosity density and rcr_{\text{c}} is the core radius. Larger values of η\eta are correlated with properties which are encouraging for the present search [76]. First and foremost, isolated MSPs are more prevalent in high-η\eta clusters, perhaps due to recycling followed by ejection from the binary by secondary encounters. Second, high-η\eta clusters contain pulsars far from the core. A pulsar kicked out of the core by a dynamical interaction should sink back to the core due to dynamical friction, as long as it remains bound [77]. Hence detections of MSPs far from the core suggests that the secondary encounter which ejected the pulsar from the binary happened relatively recently so the overall rate of such encounters is indeed enhanced in high-η\eta clusters.

We follow Abbott et al. [24] and prioritize targets according to the cluster figure of merit η1/2D1\eta^{1/2}D^{-1}, where DD is the distance to the cluster. The figure of merit follows from the age-based limit on strain amplitude previously used in searches for CWs from supernova remnants [78, 48, 79], viz.

h0max=1.26×1024(3.3kpcD)300yra,h_{0}^{\mathrm{max}}=1.26\times 10^{-24}\left(\frac{3.3\,\mathrm{kpc}}{D}\right)\sqrt{\frac{300\,\mathrm{yr}}{a}}, (23)

where aa is the age of the object. Recalling that η\eta is the reciprocal of the binary disruption timescale, and that shorter timescales are preferred for our application, we replace aa by η1\eta^{-1}, following Ref. [24]. This approach yields the figure of merit η1/2D1=ρc1/4rc1/2D1\eta^{1/2}D^{-1}=\rho_{\mathrm{c}}^{1/4}r_{\mathrm{c}}^{-1/2}D^{-1} (up to a multiplicative constant which we ignore, as we are only interested in the ranking of the clusters).

Ranking the clusters listed in the Harris catalog [75] according to this figure of merit and taking the top five, we arrive at the list in Table 1.

Name Right ascension Declination rcr_{\mathrm{c}} [pc] log10(ρc/Lpc3)\log_{10}(\rho_{\mathrm{c}}/L_{\odot}\,\mathrm{pc}^{-3}) DD [kpc] η1/2D1\eta^{1/2}D^{-1}
NGC 6325 17h17m59.21s17^{\mathrm{h}}17^{\mathrm{m}}59.21^{\mathrm{s}} 234557.6′′-23^{\circ}45^{\prime}57.6^{\prime\prime} 6.8×1026.8\times 10^{-2} 5.52 7.8 11.8
Terzan 6 17h50m46.38s17^{\mathrm{h}}50^{\mathrm{m}}46.38^{\mathrm{s}} 311631.4′′-31^{\circ}16^{\prime}31.4^{\prime\prime} 9.9×1029.9\times 10^{-2} 5.83 6.8 13.4
NGC 6540 18h06m08.60s18^{\mathrm{h}}06^{\mathrm{m}}08.60^{\mathrm{s}} 274555.0′′-27^{\circ}45^{\prime}55.0^{\prime\prime} 4.6×1024.6\times 10^{-2} 5.85 5.3 25.4
NGC 6544 18h07m20.58s18^{\mathrm{h}}07^{\mathrm{m}}20.58^{\mathrm{s}} 245950.4′′-24^{\circ}59^{\prime}50.4^{\prime\prime} 4.4×1024.4\times 10^{-2} 6.06 3.0 52.2
NGC 6397 17h40m42.09s17^{\mathrm{h}}40^{\mathrm{m}}42.09^{\mathrm{s}} 534027.6′′-53^{\circ}40^{\prime}27.6^{\prime\prime} 3.3×1023.3\times 10^{-2} 5.76 2.3 65.5
Table 1: Cluster parameters for the five clusters targeted in this paper, from Harris [75]. The sky position (columns two and three) refers to the centroid of the light distribution. The core radius is denoted by rcr_{\mathrm{c}}, the central luminosity density is denoted by ρc\rho_{\mathrm{c}}, and the distance is denoted by DD. The final column is the figure of merit used to rank the clusters, as discussed in Section III.1.

The cluster with the second highest figure of merit is NGC 6544, which was previously targeted in the 10-day coherent search using data from the sixth LIGO science run performed by Abbott et al. [24]. We caution that the figure of merit is not directly related to an expected strain amplitude, despite its origin in the age-based limit for supernova remnants. It should instead be regarded as a convenient encapsulation of the two primary criteria for this search: isolated sources (higher η\eta) are better, as are closer sources (lower DD). The target selection strategy employed here is relatively simple and by no means optimal. A comprehensive study of how CW sources might form in globular clusters, and how observable cluster parameters correlate with these formation processes, is certainly worthwhile but beyond the scope of this work.

III.2 Data

This paper analyzes data from the third observing run (O3) of the Advanced LIGO detectors [80], two dual-recycled Fabry-Pérot Michelson interferometers with 4km4\,\mathrm{km} arms located in Hanford, Washington (H1) and Livingston, Louisiana (L1) in the United States of America [81]. The data cover the time period 1st April 2019 15:00 UTC (GPS 1238166018) to 27th March 2020 17:00 UTC (GPS 1269363618), with a commissioning break between 1st October 2019 and 1st November 2019 [82]. We analyze a calibrated data stream [83, 84], and so are working with measurements of the dimensionless strain hh.

The O3 LIGO data contain noise transients. A self-gating procedure mitigates their impact on continuous wave searches [85, 86]. The detector data are packaged into “Short Fourier Transforms” (SFTs) generated from 1800s1800\,\mathrm{s} of data. Multiple SFTs are combined to construct the \mathcal{B}-statistic in each coherent segment of length TcohT_{\mathrm{coh}} (see Section II.2). In the event that there are not enough data available to compute the \mathcal{B}-statistic in a given coherent segment, due to detector downtime, the value of lnL(o,qi)\ln L(o,q_{i}) in that segment is set to zero for all qiq_{i}. In this paper the only data gaps significant enough to prevent the computation of the \mathcal{B}-statistic for a coherent segment occur during the commissioning break mentioned above.

III.3 Coherence time

The duration of each coherent step TcohT_{\mathrm{coh}} is an important parameter. It controls the sensitivity of the search, although the dependence is fairly weak, scaling roughly as (TobsTcoh)1/4(T_{\mathrm{obs}}T_{\mathrm{coh}})^{-1/4} [62, 68]. It also plays a role in determining the HMM’s tolerance to deviations away from a deterministic signal model, either via Φw(t)\Phi_{\mathrm{w}}(t) in Eq. (2), or via errors in the parameters which determine the first and third terms of Eq. (2). In this paper we cover a wide parameter space, so it makes sense to set TcohT_{\mathrm{coh}} relatively low to reduce the number of templates and hence the computational cost In this search we adopt Tcoh=5dT_{\mathrm{coh}}=5\,\mathrm{d} , c.f. Tcoh=10dT_{\mathrm{coh}}=10\,\mathrm{d} in some HMM searches published previously [44, 46, 53, 52].

Tracking intrinsic spin wandering is a secondary priority in this paper. Electromagnetic measurements of timing noise in non-accreting pulsars, which are the objects primarily targeted in this search, suggest that Tcoh5dT_{\mathrm{coh}}\sim 5\,\mathrm{d} with the transition matrix described in Section II.3 allows conservatively for more spin wandering than is observed, assuming that the spin wandering in the CW signal tracks the electromagnetic observations. The timescale of decoherence due to spin wandering is typically 1yr\gtrsim 1\,\mathrm{yr} [87]. Hence any plausible choice of TcohT_{\mathrm{coh}} with regard to computational cost (Tcoh30dT_{\mathrm{coh}}\lesssim 30\,\mathrm{d}) allows for more spin wandering than what is expected astrophysically. As a precaution, we inject spin wandering above the astrophysically expected levels when assessing the performance of the HMM method in Sections III.4 and V. This allows us to verify that we remain sensitive to a wide range of signal models, including some that are unlikely a priori.

III.4 HMM control parameters

III.4.1 Timing noise strength

The phase-tracking implementation of the HMM assumes that the evolution of the rotational phase is driven by a random walk in the frequency, with a strength parametrised by a control parameter σ\sigma [see Equations (15)–(18)]. Our approach here is to fix TcohT_{\mathrm{coh}} based on the particular details of the search, whether that involves prior belief about the nature of the spin wandering, or a desire to keep the number of sky position/spindown templates under control (see Section (III.3). With TcohT_{\mathrm{coh}} fixed, we then set σ\sigma to be consistent with the signal model implied by our choice of TcohT_{\mathrm{coh}}. Setting σ\sigma appropriately is a subtle task. If σ\sigma is too large, the transition probabilities become uniform in both frequency and phase, the dynamical constraints imposed by the noise model are lost, and the method reduces approximately to the phase-maximized \mathcal{F}-statistic [56]. If σ\sigma is too small then the HMM struggles to track stochastic spin wandering or other unexpected phase evolution. Melatos et al. [42] recommended

σ=σ=(4Tcoh3)1/2\sigma=\sigma^{*}=(4T_{\text{coh}}^{3})^{-1/2} (24)

but did not validate Eq. (24) empirically.

There is no unique prescription for choosing σ\sigma. The uncertainty about the physical origin of the time-varying mass-quadrupole moment makes it difficult to argue in general that one choice of σ\sigma (reflecting the assumed nature of the underlying noise process) is optimal for a given search. However, σ\sigma endows the signal model with flexibility and allows the HMM to absorb small errors in parameters determining the secular spin evolution, e.g. sky position and spin-down rate. This in turn reduces the number of templates and hence the computational cost. As the HMM is restricted by construction to moving by at most one frequency bin per timestep, there is a limit to the parameter error which can be absorbed.333For sufficiently large mismatches in sky position or spindown a significant fraction of the total power is also lost in each coherent step, whether or not the HMM is allowed to track the frequency evolution of the signal across coherent steps. The aim, then, is to check whether Eq. (24) strikes an acceptable balance, allowing for both secular and stochastic deviations away from a Taylor-expanded signal model, as reflected in our choice of TcohT_{\mathrm{coh}}, without discarding the advantage gained by phase tracking.

We test the suitability of Eq. (24) via synthetic data injections into both Gaussian noise and the LIGO O3 data. For a range of h0h_{0} values we compute the detection probability PdP_{\text{d}} for σ=0.2σ\sigma=0.2\sigma^{*}, σ\sigma^{*}, and 5σ5\sigma^{*} as well as for no phase tracking, at a fixed false alarm probability of 101310^{-13} per frequency bin. We inject moderate frequency wandering via stepwise changes in f¨\ddot{f}; in each of 52 7-day chunks of injected data f¨\ddot{f} is randomly chosen from a uniform distribution spanning 1.13×1018f¨/(1Hzs2)1.13×1018-1.13\times 10^{-18}\leq\ddot{f}/(1\,\mathrm{Hz}\,\mathrm{s}^{-2})\leq 1.13\times 10^{-18}. The values of ϕ\phi, ff, and f˙\dot{f} are updated self-consistently as the signal evolves. With the chosen f¨\ddot{f} distribution, f(t)f(t) wanders by approximately 28 frequency bins on average, after correcting for the mean f˙=Tobs1[f(Tobs)f(0)]\dot{f}=T_{\mathrm{obs}}^{-1}[f(T_{\mathrm{obs}})-f(0)]. Note that the timescale of f¨\ddot{f} evolution is deliberately chosen not to coincide with Tcoh=5dT_{\text{coh}}=5\,\mathrm{d} to challenge the algorithm. The sky position (right ascension α\alpha, declination δ\delta) and initial value of f˙\dot{f} is chosen at random. The value of ff is chosen uniformly over the full search range when injecting into synthetic Gaussian noise, but fixed at 645.1442Hz645.1442\,\mathrm{Hz} when injecting into LIGO O3 data to avoid intersecting with non-Gaussian artefacts in the data. Signal recovery is performed at the injected sky position, with a 104Hz10^{-4}\,\mathrm{Hz} frequency band centred on f=f(0)f=f(0) and f˙=[f(Tobs)f(0)]/Tobs\dot{f}=\left[f(T_{\text{obs}})-f(0)\right]/T_{\text{obs}}. The injection and recovery setup is specified in Table 2.

Parameter Value
α\alpha 𝒰(0,2π)rad\mathcal{U}(0,2\pi)\,\mathrm{rad}
cosδ\cos\delta 𝒰(1,1)\mathcal{U}(-1,1)
f(0)f(0) 𝒰(100,700)Hz\mathcal{U}(100,700)\,\mathrm{Hz}
f˙(0)\dot{f}(0) 𝒰(5×1010,0)Hzs1\mathcal{U}(-5\times 10^{-10},0)\,\mathrm{Hz}\,\mathrm{s}^{-1}
cosι\cos\iota 𝒰(1,1)\mathcal{U}(-1,1)
ψ\psi 0.49rad0.49\,\mathrm{rad}
ϕ0\phi_{0} 0rad0\,\mathrm{rad}
Detectors H1, L1
Sh(f)1/2S_{h}(f)^{1/2} 4×1024Hz1/24\times 10^{-24}\,\mathrm{Hz}^{-1/2}
Spin wandering timescale 7d7\,\mathrm{d}
TcohT_{\mathrm{coh}} 5d5\,\mathrm{d}
TobsT_{\mathrm{obs}} 364d364\,\mathrm{d}
Table 2: Injection and recovery parameters for the synthetic data tests used to motivate the choices of HMM control parameters as described in Section III.4. 𝒰(a,b)\mathcal{U}(a,b) denotes a uniform distribution between aa and bb.

In order to find the detection probability PdP_{\text{d}} at a given probability of false alarm PfaP_{\text{fa}} we require an estimate of the noise-only distribution of the log likelihoods returned by the HMM. To obtain this we fit an exponential tail to the distribution of log likelihoods returned from a large number of simulations, using the procedure outlined in Appendix B and adopted in previous searches [53, 88]. The injected amplitude of the signal is characterised by h0,effh_{0,\mathrm{eff}}, a combination of h0h_{0} and cosι\cos\iota which determines approximately the signal-to-noise ratio for many CW search methods and is defined as [89]

(h0,eff)2=h02[(1+cos2ι)/2]2+(cosι)22.\left(h_{0,\text{eff}}\right)^{2}=h_{0}^{2}\frac{[(1+\cos^{2}\iota)/2]^{2}+(\cos\iota)^{2}}{2}. (25)

For cosι=1\cos\iota=1 we have h0,eff=h0h_{0,\text{eff}}=h_{0}, for cosι=0\cos\iota=0 we have h0,eff=h0/8h_{0,\text{eff}}=h_{0}/\sqrt{8}, and averaging over a uniform distribution in cosι\cos\iota gives h0,eff=(2/5)1/2h0h_{0,\text{eff}}=(2/5)^{1/2}h_{0}. Figure 2 graphs PdP_{\text{d}} as a function of h0,effh_{0,\mathrm{eff}} for the choices of σ\sigma outlined above, showing both the case where the signals are injected into simulated Gaussian noise and the case where the signals are injected into the LIGO O3 data.

Refer to caption
Refer to caption
Figure 2: Detection probability PdP_{\text{d}} versus effective wave strain h0,effh_{0,\text{eff}} for three choices of the HMM control parameter σ\sigma as well as a control experiment with zero phase tracking (see legend), as described in Section III.4.1. The top panels shows the results for injections into synthetic Gaussian noise, while the bottom panel shows the results for injections into the real LIGO O3 data. The h0,effh_{0,\mathrm{eff}} scales differ, because the noise floor is not the same for the two cases.

In both cases the choice σ=σ\sigma=\sigma^{*} performs as well as or better than the alternatives as well as no phase tracking across the full h0,effh_{0,\mathrm{eff}} range. We adopt σ=σ\sigma=\sigma^{*} throughout the rest of this paper.

III.4.2 Number of phase bins

Another important tunable parameter in the HMM is the number of bins used to track the signal phase, NΦN_{\Phi}. Melatos et al. [42] fixed NΦ=32N_{\Phi}=32. It is worth checking how reducing NΦN_{\Phi} affects performance. We seek to reduce NΦN_{\Phi} without sacrificing sensitivity, because the computational cost is proportional to NΦN_{\Phi}.

To this end we test NΦ=4N_{\Phi}=4, 88, and 3232 (c.f. Section III.4.1). Sensitivity curves showing PdP_{\text{d}} as a function of h0,effh_{0,\mathrm{eff}} are displayed in Figure 3.

Refer to caption
Figure 3: Detection probability PdP_{\text{d}} versus effective wave strain h0,effh_{0,\text{eff}} for three choices of the number of phase bins NΦN_{\Phi} as well as a control experiment with zero phase tracking, as described in Section III.4.2. The three choices of NΦN_{\Phi} perform similarly.

The differences are relatively minor, so we proceed with NΦ=8N_{\Phi}=8 in what follows.

III.5 Parameter ranges

Ranges must be set for four parameters: the frequency and frequency derivative of the CW signal, and the sky regions covered for each cluster. Here we discuss each in turn, and justify the choices made in the implementation of the search.

III.5.1 Frequency and frequency derivative

In principle the possible frequency of CW emission from neutron stars spans a wide range. Observed pulsar spin frequencies range from less than 1Hz1\,\mathrm{Hz} [90, 91] up to 716Hz716\,\mathrm{Hz} [92]. Although the relationship between the rotation frequency frotf_{\text{rot}} and the CW signal frequency ff depends on the precise nature of the emission mechanism, one normally assumes frotf2frotf_{\text{rot}}\leq f\leq 2f_{\text{rot}} [1]. We remind the reader that the notional target for this search is a recycled neutron star, whose rotation frequency is likely to fall between 0.1kHz0.1\,\mathrm{kHz} and 0.7kHz0.7\,\mathrm{kHz}, as for observed millisecond pulsars. Note that 80%{\sim}80\% of pulsars above 0.1kHz0.1\,\mathrm{kHz} fall in the range 0.1frot/1kHz0.40.1\leq f_{\text{rot}}/1\,\mathrm{kHz}\leq 0.4 [93]. There are also some practical considerations when it comes to selecting the frequency range. The LIGO interferometers are polluted by Newtonian noise below 20Hz{\sim}20\,\mathrm{Hz}, and narrowband spectral features (instrumental lines) below 0.1kHz\sim 0.1\,\mathrm{kHz} [94, 95].

The spin-down rate of a CW signal is weakly constrained a priori. Under the optimistic assumption that CW emission is responsible for all of the spindown of the emitting body, the strain amplitude is equal to

h0sd=\displaystyle h_{0}^{\mathrm{sd}}= 2.6×1025(1kpcD)(Izz1038kgm2)1/2\displaystyle 2.6\times 10^{-25}\left(\frac{1\,\mathrm{kpc}}{D}\right)\left(\frac{I_{zz}}{10^{38}\,\mathrm{kg}\,\mathrm{m}^{2}}\right)^{1/2}
×(100Hzf)1/2(|f˙|1011Hzs1)1/2\displaystyle\times\left(\frac{100\,\mathrm{Hz}}{f}\right)^{1/2}\left(\frac{|\dot{f}|}{10^{-11}\,\mathrm{Hz}\,\mathrm{s}^{-1}}\right)^{1/2} (26)

where IzzI_{zz} is the component of the moment of inertia tensor about the rotation axis. The expected sensitivity of this search is roughly [56, 78, 96]

h0sens35Sh1/2(TobsTcoh)1/4,h_{0}^{\mathrm{sens}}\approx 35S_{h}^{1/2}(T_{\mathrm{obs}}T_{\mathrm{coh}})^{-1/4}, (27)

which implies f˙1010Hzs1\dot{f}\sim-10^{-10}\,\mathrm{Hz}\,\mathrm{s}^{-1} for h0sens=h0sdh_{0}^{\mathrm{sens}}=h_{0}^{\mathrm{sd}}. To beat the spin-down limit we require h0sens<h0sdh_{0}^{\mathrm{sens}}<h_{0}^{\mathrm{sd}}. We therefore wish to pick a bound on f˙\dot{f} such that the cross-over point h0sens=h0sdh_{0}^{\mathrm{sens}}=h_{0}^{\mathrm{sd}} occurs at a high enough frequency that we beat the spin-down limit for a significant fraction of plausible signal frequencies. Figure 4 shows the h0sensh_{0}^{\mathrm{sens}} curve with three h0sdh_{0}^{\mathrm{sd}} curves overplotted with f˙=1011Hzs1-\dot{f}=10^{-11}\,\mathrm{Hz}\,\mathrm{s}^{-1}, f˙=1×1010Hzs1-\dot{f}=1\times 10^{-10}\,\mathrm{Hz}\,\mathrm{s}^{-1}, and f˙=5×1010Hzs-\dot{f}=5\times 10^{-10}\,\mathrm{Hz}\,\mathrm{s}. We set D=5kpcD=5\,\mathrm{kpc} in all cases.

Refer to caption
Figure 4: Comparison of h0sensh_{0}^{\mathrm{sens}} [Eq. (27)] with h0sdh_{0}^{\mathrm{sd}} [Eq. (26)], with h0sdh_{0}^{\mathrm{sd}} computed for three values of f˙\dot{f} (D=5kpcD=5\,\mathrm{kpc} in all cases). For the f˙=5×1010Hzs1\dot{f}=-5\times 10^{-10}\,\mathrm{Hz}\,\mathrm{s}^{-1} case, corresponding to the bound adopted in this paper (see Section III.5.1), we have h0sens<h0sdh_{0}^{\mathrm{sens}}<h_{0}^{\mathrm{sd}} for f0.8kHzf\lesssim 0.8\,\mathrm{kHz}.

The f˙=5×1010Hzs1\dot{f}=-5\times 10^{-10}\,\mathrm{Hz}\,\mathrm{s}^{-1} curve crosses over h0sensh_{0}^{\mathrm{sens}} at approximately f=0.8kHzf=0.8\,\mathrm{kHz}. Hence for 0.1f/1kHz0.80.1\leq f/1\,\mathrm{kHz}\leq 0.8 and 5f˙/1010Hzs10-5\leq\dot{f}/10^{-10}\,\mathrm{Hz}\,\mathrm{s}^{-1}\leq 0 we expect to approach or beat the spin-down limit across the full frequency range for each cluster. The range includes the bulk of the astrophysical prior on the signal frequency, recalling that the majority of MSPs have frot<0.4kHzf_{\mathrm{rot}}<0.4\,\mathrm{kHz} and thus f<0.8kHzf<0.8\,\mathrm{kHz}. We therefore adopt 0.1f/1kHz0.80.1\leq f/1\,\mathrm{kHz}\leq 0.8 and 5×1010f˙/1Hzs10-5\times 10^{-10}\leq\dot{f}/1\,\mathrm{Hz}\,\mathrm{s}^{-1}\leq 0 in the present paper. To facilitate data management, the total frequency range is divided into 0.5Hz0.5\,\mathrm{Hz} sub-bands which are searched independently.

We note that electromagnetically timed MSPs have spin-down rates orders of magnitude lower than our sensitivity limits according to Fig. 4 [93]. Their spin down is typically attributed to magnetic dipole braking, with the recycling process suppressing the magnetic dipole moment [97, 98, 9, 10, 99, 100]. Hence the signals to which this search is sensitive are most likely to come from a new class of objects with significantly higher spin-down torques, possibly dominated by gravitational radiation reaction [10, 101].

III.5.2 Sky position

Four of the clusters targeted in this search (NGC 6325, NGC 6397, NGC 6544, and Terzan 6) are core-collapsed; their density profiles peak sharply at the core. At the same time, our cluster selection criterion favors clusters which produce disrupted binaries where the neutron star may be ejected from the core. Indeed clusters similar to those selected are observed to contain millisecond pulsars many core radii away from the centre of the cluster [76]. Therefore, as a precaution, we extend the search to cover sky positions outside the core. Specifically, we extend the search out to the tidal radius as recorded in Harris [75], ensuring that every point within this radius is covered by at least one sky position template.

III.6 Parameter spacing

In this paper we search a fairly wide parameter domain (precisely how wide is discussed in Section III.5). In general, multiple templates are needed to cover the domain. In this section we discuss how closely spaced the templates must be. Sophisticated work has been done on this problem in the context of CW searches (e.g. [102, 103, 104, 105]), but we proceed empirically here, because the Viterbi tracking step makes it difficult to apply previous results.

III.6.1 Frequency and frequency derivative

We follow previous HMM-based searches and take the frequency resolution to be (2Tcoh)1(2T_{\mathrm{coh}})^{-1}. Other choices are possible. In principle any desired frequency resolution may be achieved by padding or decimating the data, e.g. δf=(10Tcoh)1\delta f=(10T_{\mathrm{coh}})^{-1} [48, 34], but δf=(2Tcoh)1\delta f=(2T_{\mathrm{coh}})^{-1} has been successfully used many times previously (see e.g. Refs. [40, 43, 48, 53, 52]) and is assumed in Eq. (24).

The HMM accommodates some displacement between the f˙\dot{f} value used to compute the \mathcal{B}-statistic and the true f˙\dot{f} by letting the signal wander into adjacent frequency bins at the boundaries between coherent segments. To determine the f˙\dot{f} resolution we empirically estimate the mismatch, i.e. the fractional loss of signal power which is incurred as the value of f˙\dot{f} used in the search is displaced away from the true value, f˙inj\dot{f}_{\mathrm{inj}}, with

m(f˙)=(f˙)(f˙inj)(f˙inj),m(\dot{f})=\frac{\mathcal{L}^{\prime}(\dot{f})-\mathcal{L}^{\prime}(\dot{f}_{\mathrm{inj}})}{\mathcal{L}^{\prime}(\dot{f}_{\mathrm{inj}})}, (28)

where we define =noise\mathcal{L}^{\prime}=\mathcal{L}-\mathcal{L}_{\text{noise}} and noise\mathcal{L}_{\text{noise}} is the mean likelihood returned in noise over a 0.5Hz0.5\,\mathrm{Hz} sub-band. Note that (f˙)\mathcal{L}^{\prime}(\dot{f}) is the maximum \mathcal{L} value over a 0.5Hz0.5\,\mathrm{Hz} sub-band centred on the injected frequency, and so m(f˙)m(\dot{f}) does not approach 11 but plateaus at approximately 0.80.8, due to the maximisation over noise-only frequency paths (whereas noise\mathcal{L}_{\mathrm{noise}} is the mean noise-only \mathcal{L} value in a 0.5Hz0.5\,\mathrm{Hz} sub-band, not the maximum). We compute m(f˙)m(\dot{f}) by injecting moderately loud signals into Gaussian noise, with h0=6×1026h_{0}=6\times 10^{-26} and Sh1/2=4×1024Hz1/2S_{h}^{1/2}=4\times 10^{-24}\,\mathrm{Hz}^{-1/2}. An example of m(f˙)m(\dot{f}) is displayed in Fig. 5, generated from a single realisation of noise and signal. All mismatch curves are smoothed using a Savitzky-Golay filter.

Refer to caption
Figure 5: Example mismatch profile m(f˙)m(\dot{f}) versus f˙\dot{f}, as defined by Equation (28). The profile is generated from a single random realisation of both noise and signal, with all parameters other than f˙\dot{f} fixed to their injected values. The injected f˙\dot{f} value is denoted by f˙inj\dot{f}_{\mathrm{inj}}.

The shape of m(f˙)m(\dot{f}) is consistent across noise realisations. Hence we take the f˙\dot{f} resolution to be the median difference between the two f˙\dot{f} values satisfying m(f˙)=0.2m(\dot{f})=0.2, drawn from 100 noise realisations. This gives an f˙\dot{f} resolution of δf˙=3.7×1012Hzs1\delta\dot{f}=3.7\times 10^{-12}\,\mathrm{Hz}\,\mathrm{s}^{-1}. We verify that the resolution is independent of the injected ff and f˙\dot{f} values and adopt it throughout the search.

III.6.2 Sky location

A mismatch in sky location induces a small apparent variation in the signal frequency due to a residual Doppler shift. Although the locations of globular clusters are well-known, they are extended objects on the sky. In this search we take a conservative approach and search out to the tidal radius of each cluster, as discussed in Section III.5.2. In general a single sky pointing does not suffice to cover the full sky area of each cluster. As with a mismatch in f˙\dot{f}, we expect the transition matrix in Equation (15) to track some of the residual Doppler variation. Nevertheless it is important to check how much variation can be accommodated before a significant amount of signal power is lost.

We specify the sky location by right ascension (α\alpha) and declination (δ\delta), referenced to the J2000 epoch. We empirically estimate the mismatch profiles m(α)m(\alpha) and m(δ)m(\delta) and thence derive resolutions in α\alpha and δ\delta. As with f˙\dot{f}, we smooth the mismatch profiles using a Savitzky-Golay filter. The treatment is somewhat more involved than in Section III.6.1, as m(α)m(\alpha) and m(δ)m(\delta) depend on the injected values αinj\alpha_{\mathrm{inj}} and δinj\delta_{\mathrm{inj}} as well as ff, i.e. their shape is not a universal function of sky position offset, because the Doppler modulation varies across the sky. In particular, sky positions at declinations near the ecliptic poles experience relatively little Doppler modulation, and hence proportionally smaller residual Doppler modulations [24]. In this paper, we target five relatively small regions of the sky so we do not produce a full skymap of m(α,δ)m(\alpha,\delta). Instead we measure the α\alpha and δ\delta resolutions independently for each cluster. The clusters themselves are fairly small (diameter 10arcmin\lesssim 10\,\mathrm{arcmin}), so we approximate the α\alpha and δ\delta resolutions to be uniform within each cluster.

Each cluster is covered by ellipses arranged in a hexagonal grid, with their semimajor and semiminor axes given by the resolutions in δ\delta and α\alpha corresponding to m(δ)0.2m(\delta)\leq 0.2 and m(α)0.2m(\alpha)\leq 0.2. The centre of one ellipse coincides with the cluster core, and the whole cluster out to the tidal radius is covered by at least one ellipse. The centres of the ellipses are taken as the sky position templates in the search. For simplicity we do not consider degeneracy between sky position and f˙\dot{f}, which may reduce modestly the number of templates, but adds to the complexity of implementing the search.

Examples of m(α)m(\alpha) and m(δ)m(\delta) are shown in the top two panels of Fig. 6, calculated for Terzan 6 at a frequency of 600Hz600\,\mathrm{Hz}. They are qualitatively similar, with well-defined troughs, but the resolutions, again defined as the distances between the two abcissae where m=0.2m=0.2, differ significantly — the resolution inferred from the m(α)m(\alpha) profile is 1.9arcmin1.9\,\mathrm{arcmin}, compared to 15arcmin15\,\mathrm{arcmin} for m(δ)m(\delta). The bottom panel of 6 shows the measured resolution in α\alpha and δ\delta as a function of signal frequency for a series of injections at the sky location of Terzan 6.

Refer to caption
Refer to caption
Refer to caption
Figure 6: Example sky resolution computations. The top and middle panels show example mismatch profiles m(α)m(\alpha) and m(δ)m(\delta) respectively, calculated for the sky location of Terzan 6 at a frequency of 600Hz600\,\mathrm{Hz} for a single random realisation of noise and signal. The resolutions are 1.9arcmin1.9\,\mathrm{arcmin} and 15arcmin15\,\mathrm{arcmin} — note the difference in scale along the horizontal axes. The bottom panel shows the resolution in right ascension (blue curve) and declination (orange curve) as a function of frequency for the sky location of Terzan 6. Each point is derived by averaging the mismatch measured from 100 random realisations, and the curves are interpolated from the injection results by fitting a power law function, as described in the text.

Clearly there is a strong frequency dependence, and it would be wasteful to simply apply the finest resolution across a wide band. Consequently, we produce curves like those in the bottom panel of Fig. 6 for each target. We interpolate by fitting a power-law function Afk+BAf^{k}+B for AA, kk, and BB using the lmfit package [106] to find the resolution in each 0.5Hz0.5\,\mathrm{Hz} sub-band.

IV Candidates

In this section we present the results of the search. We set the log likelihood threshold in terms of the desired false alarm probaility PfaP_{\mathrm{fa}} in Section IV.1. We study the candidates produced by a first-pass analysis in Section IV.2 and apply vetoes to reject those not of astrophysical origin. Finally we study the veto survivors in more depth in Section IV.3.

IV.1 Thresholds

We set a threshold log likelihood th\mathcal{L}_{\text{th}} above which candidates are kept for further analysis. The threshold is determined by the desired PfaP_{\text{fa}}, and is computed according to the procedure outlined in Appendix B. In short, we obtain a sample of \mathcal{L} values calculated in the absence of a signal, and fit an exponential distribution to the tail of this sample. The exponential distribution is used to calculate the value of th\mathcal{L}_{\mathrm{th}} that gives the desired PfaP_{\mathrm{fa}}. Empirical noise-only distributions of \mathcal{L} are obtained from 10410^{4} random sky pointings in each of three 0.5Hz0.5\,\mathrm{Hz} bands beginning at 263.5Hz263.5\,\mathrm{Hz}, 322Hz322\,\mathrm{Hz}, and 703Hz703\,\mathrm{Hz}, chosen arbitrarily but checked to ensure that the detector amplitude spectral densities (ASDs) show no obvious spectral artifacts. We find that th\mathcal{L}_{\mathrm{th}} is similar across the three bands (within ±1\pm 1 of the mean value per cluster), so we take the mean value over the three bands to be th\mathcal{L}_{\mathrm{th}} as a good approximation. We do not obtain separate noise-only distributions for each cluster.

The choice of PfaP_{\text{fa}} is subjective. Historically different searches have used a range of values [24, 39, 33, 48, 53, 52]. In this paper, we take a relatively aggressive approach (i.e. PfaP_{\text{fa}} is relatively high), setting th\mathcal{L}_{\mathrm{th}} on a per-cluster basis such that across all templates in that cluster, we expect one false alarm on average, and so expect the full search to produce five false alarms. This total number of expected false alarms is broadly comparable to other published HMM-based searches, which have adopted PfaP_{\mathrm{fa}} values giving false alarm rates ranging from 0.15 per search [48] to 18 per search [53]. The search for CWs from NGC 6544 carried out by Abbott et al. [24] adopted a false alarm rate of 184 per search, while the search targeting Terzan 5 and the galactic center carried out by Dergachev et al. [31] did not set a threshold explicitly in terms of a false alarm rate.

The total number of sky templates and hence th\mathcal{L}_{\mathrm{th}} differs between clusters. The per-cluster sky template counts and values of th\mathcal{L}_{\text{th}} corresponding to a false alarm rate of one per cluster are quoted in Table 3.

Cluster Sky templates th\mathcal{L}_{\mathrm{th}}
NGC 6544 3022 278.9278.9
NGC 6325 8416 283.9283.9
NGC 6540 8957 284.2284.2
Terzan 6 28462 289.8289.8
NGC 6397 43909 291.9291.9
Table 3: Number of sky templates and th\mathcal{L}_{\mathrm{th}} values for each target cluster as discussed in Section IV.1, chosen to give a false alarm rate of one per cluster on average. Targets are ordered by the number of sky templates.

IV.2 First pass candidates and vetoes

The number of above-threshold candidates in each cluster is listed in the pre-veto column of Table 4. The total across all clusters is 6.4×1076.4\times 10^{7}, with the candidates distributed approximately evenly among the five clusters444For comparison, the total number of Viterbi paths evaluated in this search is approximately 1.1×10131.1\times 10^{13}.. As discussed in Section IV.2.1, these pre-veto numbers are lower bounds, as most candidates are from heavily disturbed sub-bands and are not saved. The candidate counts in the subsequent columns are complete.

While 6.4×1076.4\times 10^{7} is far greater than the five expected based on the arguments in Section IV.1, we emphasise that those arguments are based on the assumption that the noise is approximately Gaussian. All but a handful at most of the candidates in the pre-veto column of Table 4 are due to non-Gaussian features in the data. In this section and the next we apply a sequence of veto procedures designed to reject candidates that are consistent with non-Gaussian noise.

IV.2.1 Disturbed sub-band veto

Some of the 0.5Hz0.5\,\mathrm{Hz} sub-bands are so disturbed by non-Gaussian artifacts that they are unusable in the absence of noise cancellation [107, 108, 109, 110]. In view of this, we impose a blanket veto on sub-bands which produce output files larger than 100MB100\,\mathrm{MB} (corresponding to more than 1.67×1051.67\times 10^{5} candidates per file). Between 3.93.9% and 5.45.4% of the full search band is removed thus, with the smallest number of removed sub-bands in NGC 6544 (55 out of 1400), and the highest in NGC 6397 (75 out of 1400). These percentages are roughly in line with the all-sky O3a search reported by Abbott et al. [111], for which 13% of the observing band was removed due to an unmanageable number of candidates (noting that the band in that search extended to 2000Hz2000\,\mathrm{Hz}). After this veto, approximately 4.9×1054.9\times 10^{5} candidates remain.

IV.2.2 Known-lines veto

The known-lines veto is simple and efficient, rejecting many candidates with relatively little computational cost. The LIGO Scientific Collaboration maintains a list of known instrumental lines [112, 95]. When assessing coincidence we must take into account that the line frequencies and bandwidths in Ref. [112] are quoted in the detector frame, while the HMM frequency path of a candidate is recorded in the Solar System barycentre frame. The resulting Doppler shift amounts to approximately ±104f\pm 10^{-4}f [113]. For each candidate we check whether any portion of the HMM frequency path overlaps with a known line in either detector, after correcting for the Doppler shift. In the event of any overlap the candidate is vetoed. After this veto, 1825218252 candidates remain.

IV.3 Second pass survivors and vetoes

The 1825218252 survivors from the first pass vetoes are still daunting to follow up manually. We apply three more vetos to further exclude candidates that are consistent with non-Gaussian noise artifacts.

IV.3.1 Cross-cluster veto

The cross-cluster veto excludes candidates which have frequencies overlapping (up to Doppler modulation) with candidates in another cluster. Candidates which appear at the same frequency in multiple clusters are consistent with non-Gaussian noise artifacts.

Given the targeted false alarm rate of one per cluster, we expect five candidate which are due to Gaussian noise fluctuations, and a candidate which overlaps with one of these will be falsely dismissed under this veto. Here we briefly show that the probability of a collision between a Gaussian false alarm and an astrophysical candidate is low. Our criterion for coincidence between two candidates is |f1f2|<104f1|f_{1}-f_{2}|<10^{-4}f_{1}, where f1f_{1} and f2f_{2} are the terminal frequencies of the two candidate frequency paths, and 10410^{-4} is the Doppler broadening factor. To evaluate the probability of a collision, we simulate a search by picking a candidate frequency uniformly over the search range 100Hzf800Hz100\,\mathrm{Hz}\leq f\leq 800\,\mathrm{Hz}, which is designated as the frequency of the astrophysical candidate, and then pick five more frequencies, again uniformly, which represent the frequencies of the Gaussian false alarm candidates. We then check the coincidence criterion for each Gaussian false alarm against the astrophysical candidate. Simulating 10510^{5} searches in this way, we find 6767 collisions, i.e. the probability of falsely dismissing an astrophysical candidate due to coincidence with a Gaussian false alarm is approximately 7×1047\times 10^{-4}, which we find to be acceptably low. We therefore veto any candidate for which there is a candidate in another cluster with respective frequencies f1f_{1} and f2f_{2} satisfying |f1f2|<104f1|f_{1}-f_{2}|<10^{-4}f_{1}.

1797717977 candidates are vetoed this way. We find that all of the vetoed candidates have at least 4 cross-cluster counterparts, i.e. distinct combinations of α\alpha, δ\delta, and f˙\dot{f} with at least one frequency path with >th\mathcal{L}>\mathcal{L}_{\mathrm{th}} overlapping with the vetoed candidate — multiple cross-cluster counterparts may be associated with the same cluster. After this veto, 275275 candidates remain.

IV.3.2 Single-interferometer veto

If a candidate is due to a non-Gaussian feature which is present in the data from only one interferometer, we expect the significance of the candidate to increase when data from only that interferometer is used. We thus veto candidates satisfying X>\mathcal{L}_{X}>\mathcal{L}_{\cup}, where X\mathcal{L}_{X} is the log likelihood of the candidate template using only data from detector XX and \mathcal{L}_{\cup} is the log likelihood computed using data from both detectors. Before carrying out this veto we group candidates that have frequencies within 102Hz10^{-2}\,\mathrm{Hz} of each other, in order to reduce the computation required and enable manual verification of the results of the veto procedure. There are 10 candidate groups in total, appearing in two clusters, NGC 6397 and NGC 6544. To compute X\mathcal{L}_{X} over the relevant frequency band we use the loudest sky position and f˙\dot{f} template in each candidate group, and \mathcal{L}_{\cup} is taken to be the loudest log likelihood value in the candidate group. Plots showing X\mathcal{L}_{X} in comparison to \mathcal{L}_{\cup} for each candidate are collected in Appendix C. In all but one case, one X\mathcal{L}_{X} value peaks above \mathcal{L}_{\cup}, indicating that the candidate group is likely to be due to a disturbance in the corresponding interferometer and is not astrophysical. These candidate groups are vetoed.

After the single-interferometer veto, two candidates remain.

IV.3.3 Unknown lines veto

The line list used for the known-lines veto [112] includes only lines for which there is good evidence of a non-astrophysical origin. Other line-like features are visible by eye in plots of the detector ASDs but have not yet been confidently associated with a terrestrial source.

As a further check on the final remaining candidate group, we manually inspect the ASD in each detector to check whether the candidate group overlaps with a loud non-Gaussian feature in the data, taking into account the Doppler modulation. Candidates which overlap with a clear feature in an ASD are vetoed, on the grounds that we expect any astrophysical signal to be at most marginally visible in the detector ASDs at distances 2kpc\gtrsim 2\,\mathrm{kpc} [114]. A plot showing the detector ASDs and corresponding \mathcal{L} values for the loudest candidate in the remaining candidate group, in the cluster NGC 63976397 at f652.756Hzf\approx 652.756\,\mathrm{Hz}, is shown in Figure 7.

Refer to caption
Figure 7: Individual detector ASDs (top panel) and joint-detector log likelihoods (bottom panel) for the candidate group which survived the single-interferometer veto. The vertical dotted lines in the top panel indicate the extent of the Doppler modulation, and the horizontal dashed line in the bottom panel indicates th\mathcal{L}_{\mathrm{th}}. The cluster name and approximate ending frequency of the loudest candidate within the group are shown at the top of the figure. There is a clear disturbance in the H1 data within the Doppler modulation window, and no corresponding feature in the L1 data. This candidate group is vetoed.

There is a clear, broad peak in the H1 ASD within the Doppler modulation range of the candidate, with no corresponding feature visible in the L1 ASD. We therefore veto this candidate group.

After the unknown lines veto, no candidates remain.

For completeness, in Appendix D we present plots analogous to Fig. 7 for all ten candidate groups which survive the cross-cluster veto.

IV.4 Summary of vetoes

After applying the veto procedures described in Sections IV.2 and IV.3, we are left with no surviving candidates. A summary of the results of the veto procedures is given in Table 4.

Cluster Pre-veto DS KL CC 1IFO UL
NGC 6544 9.3×1069.3\times 10^{6} 162509162509 29182918 8686 0 0
NGC 6325 1.1×1071.1\times 10^{7} 6833868338 33543354 0 0 0
NGC 6540 1.1×1071.1\times 10^{7} 6865668656 235235 0 0 0
Terzan 6 9.9×1069.9\times 10^{6} 145782145782 99529952 0 0 0
NGC 6397 1.2×1071.2\times 10^{7} 4356643566 17931793 189189 22 0
Total 6.4×1076.4\times 10^{7} 488851488851 1825218252 275275 22 0
Table 4: Results of applying the veto procdures described in Sections IV.2 and IV.3. Each column gives the number of candidates surviving after the corresponding veto. The abbreviated column headings are: disturbed sub-band (DS), known-lines (KL), cross-cluster (CC), single-interferometer (1IFO) and unknown lines (UL). Note that the number of pre-veto candidates is a lower bound, as the candidate list from any sub-band which produces more than 1.67×1051.67\times 10^{5} candidates is not saved and is instead marked as vetoed by the disturbed sub-band veto (see Section IV.2.1).

The disturbed sub-band and known-lines vetos reject most of the candidates but leave 1825218252, still a significant number. The cross-cluster veto further rejects all but 275275 candidates, which are grouped into ten narrow frequency bands. We find that for nine of the ten groups, one of the single-detector log likelihoods peaks at a higher value than the joint-detector peak, indicating that these candidates are due to a non-Gaussian feature which is peculiar to one detector. Hence these groups are vetoed. The remaining group is in the vicinity of a broad, loud disturbance in the H1 detector and is therefore also vetoed.

V Sensitivity

In this section we estimate the sensitivity of the search to CW emission from each targeted globular cluster. Specifically, for each cluster we estimate h0,eff95%h_{0,\text{eff}}^{95\%}, the effective strain amplitude [defined in Eq. (25)] at which we would expect to detect a signal 95% of the time. We draw the distinction between the search sensitivity, defined as above, and an upper limit. An upper limit is a limit on the loudest astrophysical signal actually present in the data, and is necessarily estimated with reference to the results of the search (i.e. estimation of the upper limit in a given sub-band depends on the loudest candidate in that sub-band). By contrast the search sensitivity is not a function of the results of the search, only the search configuration. We prefer to estimate the search sensitivity here as it is straightforward to interpret and requires less computation. Computing upper limits requires estimating a relationship between the noise ASD in a given sub-band, Sh(f)1/2S_{h}(f)^{1/2}, the log likelihood of the loudest candidate in that sub-band, max\mathcal{L}_{\mathrm{max}}, and the effective strain amplitude which produces a candidate at least that loud 95% of the time, h0,eff95%h_{0,\text{eff}}^{95\%}. This can be done, of course, but it is more involved and requires costly signal injection studies. Here we estimate the search sensitivity in a small number of sub-bands and subsequently interpolate across the full search band, as described below.

In order to estimate the search sensitivity we inject synthetic signals into the real O3 data. We inject at sky positions slightly offset in right ascension from the true cluster position, so that the sky templates do not overlap with the search, so as to avoid the possibility of contamination by an astrophysical signal from the cluster.In a given 0.5Hz0.5\,\mathrm{Hz} sub-band we inject signals at ten evenly spaced h0,effh_{0,\text{eff}} levels between 1.7×10261.7\times 10^{-26} and 3.5×10263.5\times 10^{-26} and determine the detection rate as a function of h0,effh_{0,\text{eff}}. We then fit these rates as a function of h0,effh_{0,\text{eff}} to a sigmoid curve, following Abbott et al. [48]. We use the best-fit parameters to determine the h0,effh_{0,\text{eff}} value yielding Pd=0.95P_{\text{d}}=0.95, in that sub-band.

The off-target injections and recoveries are treated much the same as in the actual search. The key difference is that the centre of the injection-containing mock cluster is offset from the true cluster centre by five tidal radii in right ascension. The placement of α\alpha, δ\delta, and f˙\dot{f} templates mimics that of the real search, but to save computational resources we only search f˙\dot{f} templates within ±3δf˙\pm 3\delta\dot{f} of the injected f˙\dot{f}. We also search a reduced frequency range ±5×104Hz\pm 5\times 10^{-4}\,\mathrm{Hz} around the injection frequency. The value of th\mathcal{L}_{\text{th}} is the same as in the real search555In the searches around these mock injections the number of templates is smaller than the number used in real searches, implying a smaller th\mathcal{L}_{\mathrm{th}} for the targeted false alarm rate. Nonetheless the aim is to ascertain the sensitivity of the real searches, and so we use the th\mathcal{L}_{\mathrm{th}} values quoted in Table 3., and we claim a successful recovery if the maximum recovered value of \mathcal{L} exceeds th\mathcal{L}_{\text{th}} regardless of which template returns the maximum. As we set th\mathcal{L}_{\text{th}} on a per-cluster basis (see Section IV.1), we expect different sensitivity curves for each cluster, with clusters requiring fewer sky position templates (and hence smaller th\mathcal{L}_{\text{th}} values) reaching deeper sensitivities.

We determine the value of h0,eff95%h_{0,\text{eff}}^{95\%} directly following the above recipe in a small number of 0.5Hz0.5\,\mathrm{Hz} bands (beginning at 168.2Hz168.2\,\mathrm{Hz}, 242.6Hz242.6\,\mathrm{Hz}, 424.3Hz424.3\,\mathrm{Hz}, 562.8Hz562.8\,\mathrm{Hz}, 622.2Hz622.2\,\mathrm{Hz}, and 702.0Hz702.0\,\mathrm{Hz}) and then interpolate to cover the full observing band. The value of h0,eff95%h_{0,\mathrm{eff}}^{95\%} is a strong function of frequency, but to a good approximation it is directly proportional to the detector-averaged ASD Sh(f)1/2S_{h}(f)^{1/2}. Therefore in order to interpolate the estimated h0,eff95%h_{0,\text{eff}}^{95\%} values in the six sub-bands to a h0,eff95%h_{0,\mathrm{eff}}^{95\%} curve which covers the entire observing band we estimate the sensitivity depth 𝒟95%\mathcal{D}^{95\%}, which is the constant of proportionality between h0,eff95%h_{0,\mathrm{eff}}^{95\%} and Sh(f)1/2S_{h}(f)^{1/2} and is defined as

𝒟95%=Sh(f)1/2h0,eff95%.\mathcal{D}^{95\%}=\frac{S_{h}(f)^{1/2}}{h_{0,\text{eff}}^{95\%}}. (29)

We compute Sh(f)S_{h}(f) using lalpulsar_ComputePSD, averaging over all data used in the search from both detectors. There is some scatter in the computed 𝒟95%\mathcal{D}^{95\%} values in the six 0.5Hz0.5\mathrm{Hz} sub-bands, with 0.96<𝒟95%/𝒟¯95%<1.040.96<\mathcal{D}^{95\%}/\overline{\mathcal{D}}^{95\%}<1.04 for each cluster, where 𝒟¯95%\overline{\mathcal{D}}^{95\%} is the mean value of 𝒟95%\mathcal{D}^{95\%} across the six sub-bands. There is no clear trend in 𝒟95%\mathcal{D}^{95\%} as a function of frequency. As 𝒟95%\mathcal{D}^{95\%} is defined to be the constant of proportionality between Sh(f)1/2S_{h}(f)^{1/2} and h0,eff95%h_{0,\mathrm{eff}}^{95\%}, to obtain a curve of h0,eff95%h_{0,\mathrm{eff}}^{95\%} for each cluster over the full band we divide Sh(f)1/2S_{h}(f)^{1/2} by the 𝒟¯95%\overline{\mathcal{D}}^{95\%} value for that cluster.

The resulting h0,eff95%h_{0,\text{eff}}^{95\%} sensitivity curves are shown in the top panel of Figure 8.

Refer to caption
Refer to caption
Figure 8: Sensitivity estimates for the five targeted clusters. Top panel: characteristic wave strain at 95% detection probability, h0,eff95%h_{0,\mathrm{eff}}^{95\%}, as a function of signal frequency, ff (in Hz), interpolated from the average estimated sensitivity depth defined in Eq. (29) over six 0.5Hz0.5\,\mathrm{Hz} bands. The curves for NGC 6397 and Terzan 6 are too close to be distinguished, as are the curves for NGC 6540 and NGC 6325. Bottom panel: minimum ellipticity ϵ95%\epsilon^{95\%} as a function of ff for the values of DD listed in Table 1 and the fiducial value of IzzI_{zz} in Eq. (30). The clusters are color-coded according to the legend.

The 𝒟95%\mathcal{D}^{95\%} values used to produce the h0,eff95%h_{0,\mathrm{eff}}^{95\%} curves are listed in Table 5.

Cluster 𝒟¯95%\overline{\mathcal{D}}^{95\%}
Terzan 6 116.3
NGC 6397 117.0
NGC 6325 119.3
NGC 6540 120.2
NGC 6544 125.9
Table 5: Estimated sensitivity depths 𝒟¯95%\overline{\mathcal{D}}^{95\%} for the five targeted clusters. The value of 𝒟¯95%\overline{\mathcal{D}}^{95\%} is calculated by taking the mean of the 𝒟95%\mathcal{D}^{95\%} values [see Eq. (29)] computed in six sub-bands.

The differences in sensitivity between the five clusters are due to the varying th\mathcal{L}_{\mathrm{th}} values for each cluster (see Table 3), which follow from the number of sky position templates (see Sections III.5.2 and III.6.2). The variation between clusters is fairly small: the least sensitive cluster (Terzan 6) is approximately 10% worse than the most sensitive (NGC 6544), achieving h0,eff95%=2.9×1026h_{0,\mathrm{eff}}^{95\%}=2.9\times 10^{-26} and 2.7×10262.7\times 10^{-26} respectively at the most sensitive frequency of f=211Hzf=211\,\mathrm{Hz}.

Out of astrophysical interest, we also convert h0,eff95%h_{0,\mathrm{eff}}^{95\%} in Figure 8 into estimates of the minimum ellipticity of a source which would have been detected by this search assuming emission at twice the rotational frequency, according to the relation [51]

ϵ95%=\displaystyle\epsilon^{95\%}= 9.5×106(h0,eff95%1025)(Izz1038kgm2)1\displaystyle 9.5\times 10^{-6}\left(\frac{h_{0,\text{eff}}^{95\%}}{10^{-25}}\right)\left(\frac{I_{zz}}{10^{38}\,\mathrm{kg}\,\mathrm{m}^{2}}\right)^{-1}
×(f100Hz)2(D1kpc).\displaystyle\times\left(\frac{f}{100\,\mathrm{Hz}}\right)^{-2}\left(\frac{D}{1\,\mathrm{kpc}}\right). (30)

Curves of ϵ95%\epsilon^{95\%} versus signal frequency for the five clusters are shown in the bottom panel of Figure 8. The smallest minimum ellipticity estimates are achieved at the top of the observing band, 800Hz800\,\mathrm{Hz}, due to the f2f^{-2} factor in Eq. (30). The factor of DD introduces larger variation between the clusters compared to the variation in h095%h_{0}^{95\%} values, and modifies the relative ordering — the smallest minimum ellipticity estimate is 1.6×1071.6\times 10^{-7}, for the nearest cluster NGC 6397 (D=2.3kpcD=2.3\,\mathrm{kpc}), while the largest is 5.2×1075.2\times 10^{-7} for the farthest cluster NGC 6325 (D=7.8kpcD=7.8\,\mathrm{kpc}).

VI Conclusion

We carry out a search for continuous gravitational radiation from unknown neutron stars in five globular clusters. This is the first time that all but one of the chosen clusters have been targeted. For NGC 6544 which was previously searched using LIGO S6 data, the sensitivity is improved approximately one order of magnitude [24] (although note that the signal models and parameter domain differ, so a direct, like-for-like comparison is impossible). This is also the first search to use a phase-tracking HMM in conjunction with a phase-dependent version of the \mathcal{B}-statistic [59, 42]; previous HMM searches tracked frequency only, not phase. Phase tracking brings with it an increase in sensitivity, but also increases the computational cost. For searches aiming to optimize sensitivity at fixed computing cost, a careful analysis of the tradeoff between sensitivity and cost is crucial. We do not undertake this exercise in this paper, as a key aim is to demonstrate the use of the phase-tracking HMM in an astrophysical search for the first time.

We find no significant candidates. We estimate the 95%95\% effective strain amplitude sensitivity for each cluster, along with a corresponding estimate of the 95%95\% ellipticity sensitivity. The minimum per-cluster values of h0,eff95%h_{0,\text{eff}}^{95\%} lie in the range 2.5h0,eff95%/10263.02.5\leq h_{0,\mathrm{eff}}^{95\%}/10^{-26}\leq 3.0, achieved at approximately 211Hz211\,\mathrm{Hz}, while the minimum values of ϵ95%\epsilon^{95\%} are achieved at the top end of the search band (800Hz800\,\mathrm{Hz}), and lie in the range 1ϵ95%/10761\leq\epsilon^{95\%}/10^{-7}\leq 6. The lowest h0,eff95%h_{0,\text{eff}}^{95\%} values are obtained for the cluster NGC 6544, while the lowest ϵ95%\epsilon^{95\%} values are obtained for NGC 6397.

We surpass the spin-down limit for |f˙|1011Hzs1|\dot{f}|\gtrsim 10^{-11}\,\mathrm{Hz}\,\mathrm{s}^{-1}, with variation between clusters due primarily to the difference in distance. As discussed in Section III.5, electromagnetically visible MSPs have spin-down rates well below this limit, so this search is primarily sensitive to “gravitars” — neutron stars whose spin-down behaviour is dominated by gravitational rather than electromagnetic torques [101]. The inferred limits on the ellipticity of the star probe up to an order of magnitude below theoretical upper limits, viz. ϵ106\epsilon\lesssim 10^{-6} [115, 116].

Acknowledgments

Parts of this research are supported by the Australian Research Council (ARC) Centre of Excellence for Gravitational Wave Discovery (OzGrav) (project numbers CE170100004 and CE230100016) and ARC Discovery Project DP170103625. This material is based upon work supported by NSF’s LIGO Laboratory which is a major facility fully funded by the National Science Foundation. LD is supported by an Australian Government Research Training Program Scholarship. LS is supported by the ARC Discovery Early Career Researcher Award, Project Number DE240100206. This work was performed in part on the OzSTAR national facility at Swinburne University of Technology. The OzSTAR program receives funding in part from the Astronomy National Collaborative Research Infrastructure Strategy allocation provided by the Australian Government.

References

Appendix A Transition probabilities

The transition probabilities used in this work are essentially the same as those derived by Melatos et al. [42]. They are are derived from the system of equations (15) and (16) except that we discard the damping term γf-\gamma f included in Eq. (15) by Melatos et al. [42]. In this appendix we reproduce the form for Ab(qi,qj)=Pr[q(tn)=qiq(tn+1)=qj]=pB(tn,qi)A^{\mathrm{b}}(q_{i},q_{j})=\Pr[q(t_{n})=q_{i}\mid q(t_{n+1})=q_{j}]=p^{\mathrm{B}}(t_{n},q_{i}) which appears in Melatos et al. [42], and subsequently justify the choice to drop the γf-\gamma f term. We then write down the transition probabilities in the γ0\gamma\to 0 limit, which we adopt in the search.

With Δq=(Δf,ΔΦ)\Delta q=(\Delta f,\Delta\Phi) specifying the difference between the final and initial states, the distribution pB(tn,Δq)p^{\mathrm{B}}(t_{n},\Delta q) is a wrapped Gaussian [41, 42],

pB(tn,Δq)=\displaystyle p^{\mathrm{B}}(t_{n},\Delta q)= (2π)1(det𝚺)1/2\displaystyle(2\pi)^{-1}(\det\mathbf{\Sigma})^{-1/2}
×m=exp[(ΔqQm)𝚺1(ΔqQm)T],\displaystyle\times\sum_{m=-\infty}^{\infty}\exp\left[-(\Delta q-Q_{m})\mathbf{\Sigma}^{-1}(\Delta q-Q_{m})^{\mathrm{T}}\right], (31)

where we have

(Qm)1\displaystyle(Q_{m})_{1} =0\displaystyle=0 (32)
(Qm)2\displaystyle(Q_{m})_{2} =m,\displaystyle=-m, (33)
𝚺11\displaystyle\mathbf{\Sigma}_{11} =σ22γ[1exp(2γτ)],\displaystyle=\frac{\sigma^{2}}{2\gamma}\left[1-\exp(-2\gamma\tau)\right], (34)
𝚺12=𝚺21\displaystyle\mathbf{\Sigma}_{12}=\mathbf{\Sigma}_{21} =σ22γ2[1exp(γτ)]2,\displaystyle=\frac{\sigma^{2}}{2\gamma^{2}}\left[1-\exp(-\gamma\tau)\right]^{2}, (35)
𝚺22\displaystyle\mathbf{\Sigma}_{22} =σ22γ3{1+2γτ[2exp(γτ)]2},\displaystyle=\frac{\sigma^{2}}{2\gamma^{3}}\left\{1+2\gamma\tau-\left[2-\exp(-\gamma\tau)\right]^{2}\right\}, (36)

with τ=tn+1tn=Tcoh\tau=t_{n+1}-t_{n}=T_{\text{coh}}. The condition γ<(2fTcoh2)1\gamma<(2fT_{\text{coh}}^{2})^{-1} in Ref. [42] ensures that the argument of the exponent, γτ(2fTcoh)1\gamma\tau\sim(2fT_{\text{coh}})^{-1}, is O(108)O(10^{-8}).

Taylor expanding the exponentials in Eqs. (34)–(36) about γτ=0\gamma\tau=0 allows us to write these expressions as a prefactor multiplied by a power series in γτ\gamma\tau with a constant term that is O(1)O(1), e.g.

𝚺12=σ2τ22(1γτ+).\mathbf{\Sigma}_{12}=\frac{\sigma^{2}\tau^{2}}{2}\left(1-\gamma\tau+\ldots\right). (37)

Given γτ108\gamma\tau\sim 10^{-8} we conclude by inspecting (37) and its counterparts that the effect of the damping term on the moments of the transition probability distribution is negligible in this case. We therefore take the γ0\gamma\to 0 limit and write

𝚺11\displaystyle\mathbf{\Sigma}_{11} =σ2τ,\displaystyle=\sigma^{2}\tau, (38)
𝚺12=𝚺21\displaystyle\mathbf{\Sigma}_{12}=\mathbf{\Sigma}_{21} =σ2τ22,\displaystyle=\frac{\sigma^{2}\tau^{2}}{2}, (39)
𝚺22\displaystyle\mathbf{\Sigma}_{22} =σ2τ32.\displaystyle=\frac{\sigma^{2}\tau^{3}}{2}. (40)

The transition probabilities are obtained by substituting (38)–(40) into (31).

Appendix B Detection thresholds

In order to reduce the number of candidates from a search to a manageable level, we set thresholds on the log likelihood of the paths returned from the search and do not retain any paths which fall below the threshold for followup analysis. The thresholds are set to target a desired false alarm probability PfaP_{\text{fa}}, i.e. one in every Pfa1P_{\text{fa}}^{-1} paths has >th\mathcal{L}>\mathcal{L}_{\text{th}} in sub-bands where the data are not significantly polluted by non-Gaussian features.

Since in general we want PfaP_{\text{fa}} to be small, setting th\mathcal{L}_{\text{th}} requires knowledge of the behaviour of the distribution of path log likelihoods in the tail. In this work we assume that the tail of the \mathcal{L} distribution is exponential, as confirmed empirically in previous work [53, 88]. We use off-target searches in the real data to infer the parameters of the exponential distribution. We follow the method described in Appendix A 1 of Abbott et al. [53], which we briefly recapitulate here.

We assume that the likelihood distribution p()p(\mathcal{L}) has the form

p()=Aλexp[λ(tail)]\displaystyle p(\mathcal{L})=A\lambda\exp[-\lambda(\mathcal{L}-\mathcal{L}_{\text{tail}})] for >tail,\displaystyle\text{for }\mathcal{L}>\mathcal{L}_{\text{tail}}, (41)

where AA and λ\lambda are constants and tail\mathcal{L}_{\text{tail}} is an empirically determined cut-off above which the distribution is approximately exponential. We have a sample of NN paths, of which NtailN_{\text{tail}} have >tail\mathcal{L}>\mathcal{L}_{\text{tail}}. The sample is generated by searching random sky positions and f˙\dot{f} values in 0.5Hz0.5\,\mathrm{Hz} sub-bands that appear free from disturbance based on the detector noise ASDs. Then AA is simply Ntail/NN_{\text{tail}}/N, and the maximum-likelihood estimate of λ\lambda is

λ^=Ntaili=1Ntailitail.\hat{\lambda}=\frac{N_{\text{tail}}}{\sum_{i=1}^{N_{\text{tail}}}\mathcal{L}_{i}-\mathcal{L}_{\text{tail}}}. (42)

For a desired PfaP_{\text{fa}} we then estimate th\mathcal{L}_{\text{th}} to be

th(Pfa)=1λ^ln(NPfaNtail)+tail.\mathcal{L}_{\text{th}}(P_{\text{fa}})=-\frac{1}{\hat{\lambda}}\ln\left(\frac{NP_{\text{fa}}}{N_{\text{tail}}}\right)+\mathcal{L}_{\text{tail}}. (43)

Appendix C Single-interferometer veto

This appendix collects the plots of the single-detector log likelihood values X\mathcal{L}_{X} in the vicinity of the ten candidate groups which pass through the single-interferometer veto described in Section IV.3.2. The plots are displayed in Figure 9. The log likelihoods are computed using the template of the loudest candidate in each group. The frequency of each candidate is indicated by a vertical dashed line, and the joint-detector log likelihood \mathcal{L}_{\cup} is indicated by a horizontal dashed line. A candidate group is vetoed if X>\mathcal{L}_{X}>\mathcal{L}_{\cup} for either detector. Nine of the ten candidate groups are vetoed this way.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 9: Single-detector log likelihoods X\mathcal{L}_{X} versus frequency for the ten candidate groups surviving the cross-cluster veto. The vertical dashed lines indicate the frequency of the loudest candidate in each group. The horizontal dashed lines indicate the loudest joint-detector log likelihood in each group, \mathcal{L}_{\cup}. Nine of the ten candidate groups satisfy X>\mathcal{L}_{X}>\mathcal{L}_{\cup} and are therefore vetoed. The cluster name and approximate terminal frequency of the loudest candidate in each group are recorded above each panel.
Refer to caption
Refer to caption
Figure 9: Single-detector log likelihoods X\mathcal{L}_{X} versus frequency for the ten candidate groups surviving the cross-cluster veto. The vertical dashed lines indicate the frequency of the loudest candidate in each group. The horizontal dashed lines indicate the loudest joint-detector log likelihood in each group, \mathcal{L}_{\cup}. Nine of the ten candidate groups satisfy X>\mathcal{L}_{X}>\mathcal{L}_{\cup} and are therefore vetoed. The cluster name and approximate eding frequency of the loudest candidate in each group are recorded above each panel. (cont.)

Appendix D Unknown lines

This appendix collects plots of the single-detector ASDs in the vicinity of the ten candidate groups which survive the cross-cluster veto (Section IV.3.1). The plots are displayed in Figure 10. Although all but one of these candidate groups are vetoed by the single-interferometer veto (Section IV.3.2), we present the single-detector ASDs for completeness. Seven of the ten candidate groups show a disturbance in one ASD within the Doppler modulation window of the candidate group.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 10: Single-detector ASDs (units: Hz1/2\mathrm{Hz}^{-1/2}) and log likelihoods versus frequency (units: Hz) for the ten candidate groups surviving the cross-cluster veto. The vertical, dotted lines in the ASD plots indicate the extent of the Doppler modulation of the candidate group. The horizontal, dashed lines in the log likelihood plots indicate the value of th\mathcal{L}_{\mathrm{th}} for the candidate’s cluster. Seven of the ten candidate groups show a clear disturbance in one of the detector ASDs. The cluster name and approximate ending frequency of the loudest path in each candidate group are recorded above each panel.