Brought to you by:
Paper

Bayesian inference on EMRI signals using low frequency approximations

, , and

Published 25 June 2012 © 2012 IOP Publishing Ltd
, , Citation Asad Ali et al 2012 Class. Quantum Grav. 29 145014 DOI 10.1088/0264-9381/29/14/145014

0264-9381/29/14/145014

Abstract

Extreme mass ratio inspirals (EMRIs) are thought to be one of the most exciting gravitational wave sources to be detected with LISA. Due to their complicated nature and weak amplitudes the detection and parameter estimation of such sources is a challenging task. In this paper we present a statistical methodology based on Bayesian inference in which the estimation of parameters is carried out by advanced Markov chain Monte Carlo (MCMC) algorithms such as parallel tempering MCMC. We analysed high and medium mass EMRI systems that fall well inside the low frequency range of LISA. In the context of the Mock LISA Data Challenges, our investigation and results are also the first instance in which a fully Markovian algorithm is applied for EMRI searches. Results show that our algorithm worked well in recovering EMRI signals from different (simulated) LISA data sets having single and multiple EMRI sources and holds great promise for posterior computation under more realistic conditions. The search and estimation methods presented in this paper are general in their nature, and can be applied in any other scenario such as AdLIGO, AdVIRGO and Einstein Telescope with their respective response functions.

Export citation and abstract BibTeX RIS

1. Introduction

It is likely that most of the galaxies, including our own milky way, host super massive black holes (SMBHs) with masses of order 106M (M = solar mass) or larger in their centres. These SMBHs are surrounded by a large population of stellar mass compact objects (COs) such as neutron stars, white dwarfs and small black holes with masses ∼10 M. Due to multi-body interactions most of these COs occasionally end up being trapped in an orbit which passes too close to the central mass. Once captured in the strong gravitational field of an SMBH, a CO then starts orbiting around the central mass in an eccentric orbit which decays with time and the CO eventually spirals into the central mass. These inspirals are called extreme mass ratio inspirals (EMRIs) because of the large difference in the masses of the two bodies. EMRIs are considered to be one of the most important potential sources of gravitational waves (GW's), to be detected with the laser interferometer space antenna6 (LISA) [1]. These inspirals may encode information about the structure of the central black hole, its evolution and other features such as the Lense–Thirring effects [2] and spin–orbit coupling [3].

Bayesian approaches along with MCMC methods have been in use by different groups working on the GW source detection and parameter estimation problems (see e.g. [47] and many others). In the context of EMRI signals the reversible jump MCMC algorithm was employed in [8]. Another Monte Carlo, but not rigorously Markovian, approach in which the successive states are chosen from directed proposal distributions, was used in [911].

In this work we present results of the applications of the Bayesian approach using the parallel tempering MCMC (PTMCMC) algorithm [12] for the detection and parameter estimation of EMRI signals in LISA data. The noise spectrum is assumed to be unknown and for this reason we used Whittle's approximation to the Gaussian likelihood [13]. The Whittle likelihood uses the approximate properties of discrete Fourier transform (DFT) and assumes that the DFTs are approximately independent normally distributed with mean zero and variance proportional to the power spectral density.

This paper is organized as follows: in section 2 we present the Bayesian statistical model, comprising the waveform model, our assumptions for the error distribution with unknown power spectrum, and prior distributions of the parameters. Section 3 briefly explains the Bayesian approach to detection and parameter estimation and details the MCMC search algorithm. The implementation of the algorithm is described in detail in section 4. Conclusions are given in section 5.

2. Bayesian statistical model

In this section, we outline all components of our statistical model that encompasses the waveform model and detector response, the observation error assumptions that lead to the Whittle likelihood and the prior distribution of the parameters.

2.1. The waveform model and detector response

The EMRI sources given in MLDC [14] LISA data are generated by Barack and Cutler's analytic kludge waveform (AKW) approximation [15]. The waveform model is described by 14-dimensional parameter set θ = (ν0, M, μ, e0, $\tilde{\gamma _0}$, Φ0, α0, θS, ϕS, λ, χ, θK, ϕK, DL), where ν0 is the initial orbital frequency, μ and M are the masses of CO and SMBH, respectively, e0 is the initial eccentricity, $\tilde{\gamma _0}$, Φ0 and α0 are the initial orbital phase angles, θS and ϕS are the ecliptic latitude and longitude, respectively, λ is the orbital inclination, χ is SMBH's spin with θK, ϕK its orientation angles and DL is luminosity distance. In the Barack and Cutler parametrization, the log-transformed values of the parameters ν0, μ, M and DL are used. For the MCMC searches a suitable alternative is to use the truncated AKWs [10]. These waveforms can be computed ∼3 times faster than full AKWs and there is a typical overlap of ∼90–95% between the two. The multiple harmonics and the complex structure of EMRIs pose a challenge as far as their computation and the statistical estimation of their source parameters is concerned. The likelihood surface for EMRIs contains multiple local peaks corresponding to different harmonics, of which some are as high as 85% of the peak of the dominant harmonic. Furthermore, the orbital evolution (over time) introduces much more uncertainty to the likelihood surface. This means that the orbital parameters keep changing throughout the signal life. To overcome these issues, sophisticated algorithms are needed that can deal with multi-modality due to multiple harmonics and the increased uncertainty due to the time-varying nature of orbital parameters. One such algorithm will be presented in the subsequent sections.

The full descriptions of both the full AKW and truncated AKW models are extensive so we give only the final expressions of the model which is a pair of two polarization signals defined as

Equation (1)

Equation (2)

where ψ(t) is the polarization angle defined as

Equation (3)

where θS and ϕS are ecliptic latitude and longitude, respectively, and θL(t) and ϕL(t) are the time-varying angles specifying the instantaneous direction of the angular momentum.

Unlike ground-based interferometers the LISA arms are at 60° angles, and the interferometers are susceptible to '+' and '×' polarizations. Due to its orbital dynamics, the detector response is complicated. The detector output may be derived numerically [1618], while in the limit of long wavelength/low frequency approximation (LFA) the mapping simplifies [15, 19], which is the regime we will be concerned with here. The three spacecraft's output may be re-combined into two variables with stochastically independent noise components, namely the 'A' and 'E' variables [20]; data analysis in the following is going to be based on these two time series.

Although the LFA is simpler and faster, the amplitude of the final LISA response is not exactly the same as that of [17, 18], which we will refer to as the full LISA response throughout this paper. While working on MLDC 4 blind data, we found that the amplitudes of the LFA response are by a factor of ∼3 larger than those derived by full LISA response. We compared the two responses for several (noise-less) EMRI signals posted on the MLDC webpage [21] including those given in the training data of MLDC 4 and saw the same phenomena. This difference in amplitudes results in an incorrect estimation of luminosity distance (DL) and sky location angles (θS, ϕS) as we have found in several test MCMC searches on MLDC 1B EMRI data sets. Particularly, during MCMC searches the chains for sky location angles used to get locked at wrong positions and then nothing could move them from those positions. The same phenomenon (wrong estimation of distance and sky location) was observed by another group too [10, 22]. As an ad hoc solution we divided the overall amplitude in our signal model by 3 and used this option throughout the MLDC 4 blind searches, more investigation regarding the exact origin is required. When the overall amplitude in LFA was divided by 3, the sampler was able to converge to the correct values of the luminosity distance and sky location angles. It is of interest to note that in both cases (i.e. the adjusted and unadjusted amplitudes) the other key parameters, e.g. ν, μ, M, e and χ are almost unaffected and remain the same. Later on, near the submission of MLDC 4 blind entries, we further found that for high mass EMRIs the difference factor is generally rather larger than 3, it is generally ∼6.

2.2. Prior information

In order to implement the a priori information about the parameters to be inferred, we followed the specifications given in the MLDC Task Force's documentation [14, 21], and for different parameters we employed the following prior distributions: μ ∈ Uniform[9.5, 10.5]M, e0 ∈ Uniform[0.15, 0.25], χ ∈ Uniform[0.5, 0.7]. The range of the prior distribution of e0 was changed to [0.10, 0.40] because of its initial values given for different MLDC EMRI sources being beyond the range [0.15, 0.25]. The prior for SMBH mass are different for different sources: M ∈ Uniform[0.95, 1.05] 107M (high mass), M ∈ Uniform[4.75, 5.25] 106M (medium mass) and M ∈ Uniform[0.95, 1.05] 106M (low mass). For initial azimuthal orbital frequency ν0, the prior distribution was set to be uniform over the range [5.0 × 10−5, 0.01]. For the polar angles we assumed isotropic priors. For some parameters such as ν0, μ, M and DL, their logarithmic values were used; as in [15], for ease of specification of proposal distributions, the prior densities for these parameters were transformed accordingly. For the noise power spectrum we used the conjugate Inv-χ2 prior distribution, where the prior scale parameters were estimated based on a disjoint stretch of the same data set [23]. In the following subsections we present some results obtained with the application of the above search and estimation algorithm on different MLDC data sets containing single and multiple EMRI sources.

3. Parameter estimation

In this section, we briefly describe the Bayesian approach to statistical inference and posterior computation via MCMC. Furthermore, we give an overview of the tempering methods and the parallel tempering strategy that we employed.

3.1. Posterior inference

We are interested in inferring a priori unknown parameters θ from measured data y. To this end we derive the parameters' posterior probability distribution, which expresses the information on the actual parameter values after consideration of the observed data by assigning probabilities to regions of the parameter space. Given the relationship between parameters and data through the likelihood function p(y|θ), the posterior density function is given by Bayes' theorem as

Equation (4)

[24]. Here p(θ) is the prior probability density, expressing any information we have about the parameters before accounting for the measured data. The posterior distribution (4) is essentially given by the product of prior and likelihood, while the evidence p(y) = ∫p(y|θ) p(θ) dθ is commonly of minor concern for parameter estimation purposes, as it only constitutes a normalizing constant.

3.2. Monte Carlo integration

Bayes theorem supplies the posterior probability distribution in terms of its density function. In order to extract information relevant for any particular purpose, one may be interested in marginal posterior distributions, posterior expectations, quantiles, etc; what is commonly required is the evaluation of integrals with respect to the posterior distribution. As these integrals are rarely analytically tractable, stochastic integration methods are commonly used to approach posterior inference. Such Monte Carlo methods aim at approximating the desired integrals via random sampling; instead of computing an expectation value analytically, random draws from the distribution of interest are generated, and the expectation is then approximated by an average. Analogous procedures are applied for other figures of interest, like quantiles or marginal densities for example, and an arbitrary accuracy may be achieved by increasing the sample size. A popular variety of such Monte Carlo procedures is the Metropolis(–Hastings) algorithm [25]; what makes this particular algorithm attractive for Bayesian inference is the fact that it only requires the unnormalized probability density function of the distribution of interest as an input, which in this case is supplied by Bayes' theorem (4). Instead of providing stochastically independent samples from the distribution, the Metropolis–Hastings algorithm will generate a Markov chain of subsequently dependent draws whose stationary distribution is the distribution of interest. At each step in the generation of the random sequence, only ratios of probability density values need to be considered, so that an overall normalization constant (like p(y) in (4)) does not need to be known. To approximate a target distribution with density function p(θ|y), the Metropolis(–Hastings) algorithm employs an acceptance–rejection rule to construct a Markov chain from an auxiliary density q( · | · ), which is referred to as proposal density or transition probability function. Being at current state θ(t) the acceptance probability for moving to a new state θ' is defined as

Equation (5)

Taking q( · | · ) to be symmetric, i.e. q(θ|θ') = q(θ'|θ), leads to the basic Metropolis algorithm [26] with acceptance probability given by

Equation (6)

3.3. Tempering methods

In most multi-dimensional cases, the density surface of complicated target distributions turns out to have multiple secondaries or local modes that are well separated by deep valleys of low probability regions. The simple Metropolis(–Hastings) algorithm tends to get stuck at some of these local modes for a prohibitively long time before reaching the global mode. Analogous to an annealing process, temperature is used to scale the target density in order to flatten the local modes so that the MCMC sampler moves freely towards the global maximum without being trapped in local modes. At a given temperature T ⩾ 1, samples are generated from a tempered version of the target density p(θ|y) defined as

Equation (7)

where T = 1 yields the actual target distribution. This heating is equivalent to increasing the standard deviation of the target density by a factor $\sqrt{T}$, therefore as T increases the heated distribution becomes flatter and gets closer to the uniform distribution, which enables the Markov chain to move more freely and hence faster towards higher probability regions.

Some of the most popular tempering methods are simulated annealing [27], simulated tempering [28] and PTMCMC aka Metropolis coupled-MCMC. PTMCMC introduced in [12], is a powerful optimization of the simple Metropolis(–Hastings) algorithm which is very effective in improving the mixing of MCMC chains and in particular in escaping the local modes. The algorithm works by running multiple MCMC chains in parallel, each simulating a separate target density characterized by a different temperature and occasionally attempting swaps of its current states. As stated above, in principle, a high temperature chain sees the density of the target distribution as more flattened relative to a low temperature chain, which means that the high temperature chain can move more freely across the valleys of low probability regions in between modes. In order to make the low temperature chain benefit from the high temperature chain that may be sampling near another mode, an exchange of information about the current states is attempted by proposing a swap of the current states using an additional Metropolis acceptance–rejection step. Detailed descriptions of this algorithm can be found in [7, 29, 30].

3.4. The likelihood function

The data, which is sampled at discrete time steps, can be represented by the following signal plus noise model,

Equation (8)

where the deterministic component s(t, θ) is the true signal model depending on parameters θ and epsilon(t) is a zero-mean stationary time series with unknown spectral density [23].

When the data are in (discrete) frequency domain for an unknown noise spectral density the likelihood function is defined as

Equation (9)

where ν = ⌊(N − 1)/2⌋ is the greatest integer less than or equal to (N − 1)/2, $\tilde{y}(f_{j})$ and $\tilde{s}(f_j,\theta )$ are Fourier transformed observables and model signal, respectively, S(fj) is the one-sided power spectral density and K is the normalizing constant.

When the noise spectrum is assumed known then one can omit the constant term, log (Sn(fj)), from equation (9) to obtain

Equation (10)

In the literature, equation (9) is known as Whittle's approximation to the Gaussian likelihood or simply Whittle likelihood [31]. The Whittle likelihood assumes that the DFTed residuals are approximately independent complex normally distributed with mean zero and power spectrum S(fj). The complete description of the Bayesian estimation of the noise spectral density is explained in [23]. The likelihood computation is commonly simplified by restricting the summation to the limited frequency range relevant to the signals of interest. Furthermore, as mentioned in section 2, the two TDI observables A and E are independent, their joint likelihood is just the product of their individual likelihoods.

4. Implementation

Although LFA is much more time efficient compared to full LISA response, EMRI waveforms are still computationally expensive and hence we had to use only shorter data segments ranging in length from one to four weeks. We used heavy-tailed Student-t proposals which are very useful for good mixing. The parallel computation on multiple processors or cores is accomplished by including the message passing interface library [32] with the relevant additional programming scripts in the main analysis code. In all of our searches we used eight or ten chain PTMCMC.

4.1. Single EMRI sources: example searches

In the simplest scenario the algorithm was applied to two EMRI data sets taken from earlier MLDC rounds; 1B.3.2 and 1C.3.1. 1B.3.2 contains a single medium mass EMRI source and 1C.3.1 contains a high mass EMRI source buried in the LISA instrument noise only. In these attempts the adjusted LFA was used and there was no problem in recovering the signal parameters using the true values as the starting points for the MCMC search, thus we used completely blind searches on two weeks long stretches of the data, to test the performance of our algorithm. In both cases the blind searches were conducted in multiple stages, i.e. first several shorter MCMC chains were run in parallel without a swapping step from random starting points and those chains were chosen that showed stability and for which the SNR and likelihood values were larger than others. The modes of those chains were then set as the starting points in the next MCMC run. These steps were repeated for a few times to narrow down the search range. This approach is somewhat similar to that used in [9]. In the following, summaries of the posterior distribution of the parameters of the EMRI sources given in MLDC 1B.3.2 (medium mass) and MLDC 1C.3.1 (high mass) are given. Figure 1 display a typical shorter MCMC run demonstrating how the chains for different parameters find their true modes. Table 1 and figure 2 presents the posterior summary of the parameters for source 1B.3.2. Similarly, table 2 and figure 3 show the results for source 1C.3.1. From these results it is clear that both signals were recovered with a great accuracy. The widths of different marginal posterior densities show that almost the whole prior range was searched before convergence.

Figure 1.

Figure 1. A typical (initial) shorter MCMC run. The trace plots of marginal posterior MCMC samples for some key parameters for the EMRI training source MLDC 1B.3.2. The grey dashed lines indicate the true values.

Standard image
Figure 2.

Figure 2. Kernel density estimates of the marginal posterior densities for all 14 parameters for the EMRI source MLDC 1B.3.2. The solid lines indicate the true values.

Standard image
Figure 3.

Figure 3. Kernel density estimates of the marginal posterior densities for all 14 parameters for the EMRI source MLDC 1C.3.1 (high mass source). The solid lines indicate the true values.

Standard image

Table 1. Posterior results and true parameter values for the recovered medium mass EMRI signal given in actual MLDC 1B.3.2 training data set.

Parameters Mean Std. dev. Mode 95% BCI True values Prior range
log(ν0)  −7.961 883 0.001 556  −7.962 237 (−7.962 645, −7.959 406)  −7.962 205 log [5.0 × 10−5, 0.01]Hz
log(μ)  2.108 143 0.137 913  2.122 627 (1.853 307, 2.323 680)  2.290 258 log [9.5, 10.5]M
log(M) 15.429 324 0.014 565 15.432 732 (15.401 371, 15.460 317) 15.431 952 log [4.75, 5.25] 106M
e0  0.208 397 0.015 804  0.210 424 (0.183 391, 0.229 404)  0.215 401 [0.10, 0.40] units
$\tilde{\gamma _{0}}$  4.384 615 1.229 447  4.845 658 (1.232 442, 5.543 105)  2.033 2973 [0, 2π] rad
Φ0  4.682 341 0.695 824  5.011 981 (3.201 36, 5.340 058)  5.999 900 [0, 2π] rad
θS  0.862 168 0.531 047  0.649 613 (0.492 233, 2.114 493)  1.222 330 [0, π] rad
ϕS  2.710 713 1.905 070  1.580 461 (1.092 997, 6.237 064)  2.934 625 [0, 2π] rad
λ  2.331 862 0.068 092  2.302 875 (2.251 945, 2.463 038)  2.289 951 [0, π] rad
α0  2.884 914 2.493 591  0.649 972 (0.087 803, 6.210 357)  1.609 215 [0, 2π] rad
χ  0.589 667 0.039 206  0.578 192 (0.515 752, 0.659 08)  0.574 818 [0.5, 0.7] M2
θK  2.478 239 0.773 029  2.870 962 (0.597 917, 3.096 335)  1.403 416 [0, π] rad
ϕK  3.373 622 2.155 152  5.050 124 (0.223 237, 6.042 101)  6.223 129 [0, 2π] rad
log(DL)  −0.538 350 1.128 170  −0.135 892 (−2.721 751, 0.832 533)  −0.584 778 log [0, ] log(GPC)

Table 2. Posterior results and true parameter values for the recovered high mass EMRI signal given in actual MLDC 1C.3.1 training data set.

Parameters Mean Std. dev. Mode 95% BCI True values Prior range
log(ν0) −8.592 307 0.005 407 −8.591 044 (−8.604 332, −8.584 98) −8.591 472 log [5.0 × 10−5, 0.01]Hz
log(μ) 2.026 831 0.185 767 2.209 483 (1.741 593, 2.322 630) 2.320 877 log [9.5, 10.5]M
log(M) 16.110 918 0.001 331 16.114 347 (16.090 741, 16.132 173) 16.119 304 log [0.95, 1.05] 107M
e0 0.205 405 0.014 601 0.204 826 (0.181 549, 0.226 693) 0.195 337 [0.10, 0.40] units
$\tilde{\gamma _{0}}$ 3.425 854 2.011 777 2.116 807 (0.189 037, 6.143 949) 4.381 526 [0, 2π] rad
Φ0 3.948 827 1.273 087 3.366 273 (0.966 292, 5.721 986) 3.441 184 [0, 2π] rad
θS 0.616 343 0.537 823 0.406 881 (0.063 561, 1.823 876) 1.235 677 [0, π] rad
ϕS 2.944 769 1.580 446 2.915 200 (0.817 490, 6.224 024) 4.054 785 [0, 2π] rad
λ 1.652 216 0.605 062 2.291 762 (0.719 750, 2.352 924) 2.358 963 [0, π] rad
α0 2.610 697 1.849 872 1.261 552 (0.361 536, 5.797 795) 2.158 356 [0, 2π] rad
χ 0.612 676 0.051 827 0.650 991 (0.502 682, 0.662 555) 0.636 644 [0.5, 0.7] M2
θK 1.887 538 0.871 457 2.464 749 (0.472 588, 2.974 532) 2.036 360 [0, π] rad
ϕK 2.827 086 2.093 720 1.124 890 (0.152 617, 6.081 997) 1.260 128 [0, 2π] rad
log(DL) −1.943 526 0.763 724 −1.795 853 (−3.843 257, −0.939 327) −1.518 042 log [0, ] log(GPC)

4.2. Multiple EMRIs

4.2.1. MLDC 4 results

The approach was applied to detect signals generated by EMRI sources given in both training and blind data sets issued in the revised MLDC round 4. Looking at the amount of noise in these data, we attempted to recover signals from high mass EMRI systems only. Moreover, there were no medium mass sources in the training data set. Some preliminary results were presented at GWPAW (January 26–29, 2011, Milwaukee, WI, USA). The MLDC notation for a high mass EMRI sources are EMRI-1-0, EMRI-1-1 and so on, and for a medium mass EMRI sources are EMRI-2-0, EMRI-2-1 and so on, whereas we denote the corresponding estimated sources (in the blind data) simply by High-0, High-1 and Med-0.

4.2.2. Training data

The training data contains three high mass EMRI sources which are somewhat similar to each other, therefore a joint MCMC search was conducted to recover them. In an eight chain MCMC search on two weeks long data segments, three chains were started from the true parameter values of the three signals while the rest of the chains were started from the values in the vicinity of the true parameters. Figure 4 displays the results of this joint search. In the plots of kernel density estimates, different types of vertical lines denote the true values of the parameters of the three different high mass EMRI signals, named EMRI-1-0, EMRI-1-1 and EMRI-1-2 in MLDC 4 training keys, that can be found at the challenge webpage [33]. After running for a sufficiently large number of iterations (∼ 4 × 106) it was observed that the third signal (dashed-dot-dashed vertical lines) was dominating the other two as can be seen in figure 4, even though the overall mean swap acceptance rate between chains was ∼35%. Thus the code was restarted with the starting values of different chains somewhat similar to the true parameter values of the third signal. These results are given in table 3 and figure 5. We can see that all the parameters, except the luminosity distance, DL, and some of the angles, are estimated with great accuracy and most of the parameters' chains show stability. From the MCMC trace plots the distance parameter seemed to be over estimated and the sky location (θS, ϕS) seemed to have been locked at a different position. The wrong sky location was attributed to the fact that we were using LFA and are experiencing the same problem as was observed in [10], which also used the LFA [15]. At this stage, the problem of amplitude differences between the full LISA response and LFA was not known to us.

Figure 4.

Figure 4. Kernel density estimates of the marginal posterior densities for all 14 parameters of the first three EMRI sources in MLDC 4 training set. In these plots the solid lines indicate the true parameter values of the source EMRI-1-0, whereas the dashed and dashed-dot-dashed lines indicate the true parameter values for sources EMRI-1-1 and EMRI-1-2, respectively, in MLDC 4 training data.

Standard image
Figure 5.

Figure 5. Kernel density estimates of the marginal posterior densities for all 14 parameters of the third EMRI (EMRI-1-2) signal in MLDC 4 training data. The solid lines indicate the true parameter values.

Standard image

Table 3. Posterior results for the third EMRI (EMRI-1-2) signal given in MLDC 4 training data set.

Parameters Mean Std. dev. Mode 95% BCI True values Prior range
log(ν0) −8.615 057 0.008 435 −8.618 036 (−8.624 964, −8.593 739) −8.616 096 log [5.0 × 10−5, 0.01]Hz
log(μ) 2.247 796 0.519 295 2.185 567 (1.433 541, 3.218 01) 2.350 281 log [9.5, 10.5]M
log(M) 16.152 486 0.002 285 16.152 319 (16.148 756, 16.156 314) 16.153 307 log [0.95, 1.05] 107M
e0 0.184 634 0.019 464 0.195 090 (0.145 334, 0.203 263) 0.186 712 [0.10, 0.40] units
$\tilde{\gamma _{0}}$ 3.830 427 1.407 267 4.734 815 (1.089 336, 5.408 43) 5.138 873 [0, 2π] rad
Φ0 3.770 239 1.079 385 3.244 351 (2.128 266, 5.591 350) 2.084 779 [0, 2π] rad
θS 2.455 871 0.434 043 2.611 267 (1.311 937, 2.755 807) 2.305 033 [0, π] rad
ϕS 3.901 775 0.485 260 3.780 766 (3.583 514, 4.948 959) 4.707 928 [0, 2π] rad
λ 1.115 781 0.224 159 1.180 035 (0.572 802, 1.395 177) 1.155 677 [0, π] rad
α0 1.965 982 1.351 007 1.372 157 (0.521 520, 5.008 045) 1.708 861 [0, 2π] rad
χ 0.573 573 0.024 460 0.576 902 (0.540 204, 0.616 836) 0.579 723 [0.5, 0.7] M2
θK 1.235 264 0.685 402 1.127 951 (0.096 748, 2.711 275) 1.429 748 [0, π] rad
ϕK 1.719 245 2.280 232 0.380 733 (0.088 012, 5.899 353) 0.745 268 [0, 2π] rad
log(DL) −1.262 741 0.596 673 −1.216 699 (−2.187 479, −0.196 473) −2.299 291 log [0, ] log(GPC)

It was also evident that after some 800 000 iterations one of the neighbouring chains of the true (T = 1) chain found some other mode; however, overall the true chain was unaffected. This second mode could either correspond to, most probably, a low strength harmonic of this same EMRI signal as such a close harmonic usually corresponds to higher frequencies than the true one, or to a harmonic of another EMRI signal. This sort of overlapping and sharing of characteristics between different signals will be quite common in such complicated cases and will result in confusions among different EMRI signals.

4.2.3. Blind data

For the MLDC 4 blind data as a first step an eight chain MCMC search was conducted on the first two weeks in which all the eight chains were started from random values corresponding to a high mass EMRI source. Table 4 shows a summary of the posterior estimates and figure 6 shows the kernel density estimates of the marginal posterior densities. The MCMC chains for all parameters showed great stability except for the CO's mass μ and distance log DL the chains were showing somewhat oscillatory and correlated behaviour.

Figure 6.

Figure 6. Kernel density estimates of the marginal posterior densities for all 14 parameters for the detected EMRI source in MLDC 4 blind data. The solid lines indicate the true solutions of high mass source EMRI-1-1 in MLDC 4 blind data.

Standard image

Table 4. Posterior results for the detected high mass EMRI signal given in MLDC 4 blind data set in the first step.

Parameters Mean Std. dev. Mode 95% BCI Prior range
log(ν0) −8.573 457 0.003 015 −8.573 259 (−8.574 035, −8.572 342) log [5.0 × 10−5, 0.01]Hz
log(μ) 1.409 599 0.623 259 1.179 280 (0.589 129, 2.263 096) log [9.5, 10.5]M
log(M) 16.133 081 0.008 141 16.134 940 (16.132 126, 16.136 896) log [0.95, 1.05] 107M
e0 0.154 880 0.014 667 0.152 115 (0.139 163, 0.166 299) [0.10, 0.40] units
$\tilde{\gamma _{0}}$ 1.584 359 0.708 484 1.486 492 (0.859 918, 2.269 012) [0, 2π] rad
Φ0 4.197 498 0.336 251 4.147 553 (3.834 165, 4.592 633) [0, 2π] rad
θS 2.534 345 0.098 177 2.540 198 (2.507 845, 2.579 459) [0, π] rad
ϕS 3.822 423 0.097 805 3.809 877 (3.744 967, 3.886 976) [0, 2π] rad
λ 0.298 106 0.365 414 0.249 978 (0.150 734, 0.347 221) [0, π] rad
α0 3.072 884 2.491 800 5.761 360 (0.150 223, 6.189 003) [0, 2π] rad
χ 0.499 902 0.009 253 0.501 648 (0.488 186, 0.508 558) [0.5, 0.7] M2
θK 2.218 992 0.075 129 2.207 951 (2.161 368, 2.274 100) [0, π] rad
ϕK 1.046 394 0.370 040 0.986 633 (0.870 475, 1.121 522) [0, 2π] rad
log(DL) −0.928 824 0.803 5100 −0.986 087 (−1.974 076, −0.018 514) log [0, ] log(GPC)

The empirical correlation between these two parameters was ∼0.92, which is quite high. It could not be deciphered why this (high correlation) happened as such a phenomenon was found neither in earlier nor in subsequent searches. The chains for all the angles were also stable except for the parameter α0, which was vibrating between two different modes. An interesting result which was observed in our earlier searches, in which the LFA response was not adjusted (dividing by 3), is that for all time regions the joint plots of the sky location angles indicated a similar behaviour, though in the plots of the kernel density estimates of these two angles obtained for different time regions the (strongest) modes were different. For different time regions the joint plots of the two sky location angles are shown in figure 8. These sky locations remained the same despite using different proposal distributions. At this stage we realized that amplitudes of LFA response are in general by a factor of ∼3 larger than full LISA response amplitudes. As an ad hoc solution to this problem, due to time limitation, the LFA amplitudes were divided by 3, which, in single source searches such as MLDC 1B (as shown in section 4.1), resulted not only in the correct estimation of distance and sky location parameters but also, in MLDC 4 searches, the sky location angles never got locked on wrong positions.

In order to search the entire two years of the blind data several more MCMC searches were conducted on different time regions (data segments). In these searches we used a week long data segments chosen at regular gaps (generally six weeks) covering the entire two years duration of the data or till the anticipated plunge time of a typical EMRI signal (for most of the EMRI sources given in earlier rounds of MLDC the plunge generally occurs during [$1\frac{1}{2} - 1\frac{3}{4}$] years, although in some cases the plunges are occurred in the last month of the two years duration). Each time region was searched individually by running eight or ten chain MCMCs. The search strategy was the same as explained in section 4.1. Attempts were made to recover both types of EMRI signals (high mass and medium mass) in these searches. In the end, for each type of EMRI source four possible best fits were chosen on the basis of high SNRs, to be submitted to MLDC 4. Among the submitted entries, the posterior results for a medium mass source 'Med-2' and a high mass source 'High-3' are presented in tables 5 and 6 and figure 7. The rest of our entries and solutions can be found on MLDC 4 webpage [33]. The kernel density estimates of the marginal posterior densities for most of the parameters show multi-modality. However, in most cases the strongest modes can be clearly recognized. In all these searches the average rates of the regular Metropolis acceptance were in 20–40% while the average swap acceptance rates were in 16%–30%. The recovered medium EMRI source 'Med-2', had a very high SNR but the estimated parameters are somewhat away from the true ones, demonstrating that these are probably the secondaries of the true densities. The 'High-3' has a lower SNR but the estimated parameters are quite close to the true ones.

Figure 7.

Figure 7. Kernel density estimates of the marginal posterior densities for all 14 parameters for the high mass EMRI source 'High-3' in MLDC 4 blind data. The dashed and solid lines corresponds to the true solutions of two high mass sources EMRI-1-0 and EMRI-1-1, respectively, in MLDC 4 blind data.

Standard image
Figure 8.

Figure 8. The joint plots of sky location angles for different time regions demonstrate that there are four most probable sky positions either of the same source or there are two or more EMRI sources located in different sky regions. The dashed lines indicate the maximum a posteriori values.

Standard image

Table 5. Posterior results for the detected EMRI signal 'Med-2' given in MLDC 4 blind data set.

Source Parameters Med-2 Mean SNR-A = 62.921 Mode Std. dev. SNR-E = 70.950 Mean ± 2StdDev SNR-AE = 103.708 Prior range
ν0  0.000 341 1537  0.000 340 9980 9.80e−7 (0.000 339 0515, 0.000 342 9558) [5.0 × 10−5, 0.01] Hz
μ 13.036 900 58 10.505 164 49 3.21e+0 (6.603943913, 16.710 996 09) [9.5, 10.5]M
M 4840 967.8525 4838 561.9866 6.93e+4 (4706 125.2326, 4974 725.6905) [0.95, 1.05] 107M
e0  0.218 266  0.217 780 2.23e−2 (0.173 1910, 0.262 3679) [0.10, 0.40] units
$\tilde{\gamma _{0}}$  4.691 678  5.305 615 1.23e+0 (2.840 7054, 7.770 524) [0, 2π] rad
Φ0  5.176 204  4.961 252 8.86e−1 (3.189 7754, 6.732 728) [0, 2π] rad
θS  1.382 196  1.408 894 2.88e−1 (0.833 3279, 1.984 459) [0, 2π] rad
ϕS  4.412 142  4.646 086 9.16e−1 (2.814 0819, 6.478 091) [0, 2π] rad
λ  0.694 621  0.681 557 2.78e−1 (0.126 5363, 1.236 579) [0, 2π] rad
α0  2.624 212  0.261 676 2.68e+0 (−5.101 531, 5.624 883) [0, 2π] rad
χ  0.581 468  0.578 720 2.12e−2 (0.536 2277, 0.621 211) [0.5, 0.7] M2
θK  1.588 288  1.562 271 3.12e−1 (0.938 9395, 2.185 603) [0, 2π] rad
ϕK  2.572 490  2.615 519 5.01e−1 (1.612 6596, 3.618 377) [0, 2π] rad
DL(GPC)  0.079 2323  0.090 8401 3.62e−2 (0.016 161 34, 0.510 5964) [0, ] GPC

Table 6. Posterior results for the detected EMRI signal 'High-3' given in MLDC 4 blind data set.

Source Parameters High-3 Mean SNR-A = 19.777 Mode Std. dev. SNR-E = 17.406 Mean ± 2SthDev SNR-AE = 26.346 Prior range
ν0 0.000 192 9449 0.000 192 7831 8.81e−7 (0.000 191 0318, 0.000 194 5503) [5.0 × 10−5, 0.01] Hz
μ 8.816 421 59 10.126 919 09 1.24e+0 (7.479 619 72, 13.711 190 41) [9.5, 10.5]M
M 9362 002.4476 9462 096.5945 2.86e+5 (8901 852.0130, 10 057 600.5800) [0.95, 1.05] 107M
e0 0.184 721 0.198 859 2.89e−2 (0.148 464, 0.249 254) [0.10, 0.40] units
$\tilde{\gamma _{0}}$ 2.247 122 1.133 451 1.59e+0 (−2.049 031, 4.315 933) [0, 2π] rad
Φ0 3.681 986 4.570 893 1.38e+0 (1.791 032, 7.350 754) [0, 2π] rad
θS 1.864 284 2.062 0089 5.55e−1 (0.952 001, 3.172 016) [0, π] rad
ϕS 3.088 018 0.309 6088 2.44e+0 (−4.581 664, 5.200 881 77) [0, 2π] rad
λ 2.147 808 2.472 5720 8.12e−1 (0.848 402, 4.096 742 41) [0, π] rad
α0 2.908 103 0.541 1894 2.13e+0 (−3.719 176, 4.801 554 57) [0, 2π] rad
χ 0.626 321 0.573 2350 4.31e−2 (0.486 933, 0.659 537 07) [0.5, 0.7]
θK 1.466 041 1.656 5712 6.97e−1 (0.262 243, 3.050 899 78) [0, π] rad
ϕK 2.621 246 1.278 5729 1.58e+0 (−1.894 009, 4.451 155) [0, 2π] rad
DL (GPC) 0.112 901 98 0.094 923 33 3.15e−2 (0.066 4814, 0.135 5331) [0, ] GPC

4.3. Overlaps and SNRs

4.3.1. True signals: LFA versus full LISA response

When the true solutions for the MLDC 4 blind EMRIs were released, we tried to assess the performance of LFA and, in particular, the performance of the full AKW and TAKW in LFA regime by comparing the overlaps of their A (≡ hI), E (≡ hII) channels with those obtained with full LISA response. For this purpose all the true signals (noiseless) were generated using both lisatools package and our model codes. To calculate the overlap the following formula was used

Equation (11)

where $\langle a, b\rangle = \int _{-\infty }^{\infty } \frac{a(f) b^*(f)}{S(f)}\, {\rm d}f$, the noise weighted inner product of two functions a, b. All these quantities are in frequency domain. The lisatools package uses the full AKW model to generate the polarization signals (h+, h×) of EMRIs and then employs the full LISA response function to generate TDI variables X, Y and Z, that can then be transformed to A, E channels. Similarly, our codes can generate h+ and h× polarizations using AKW or TAKW and then compute hI, hII using LFA. The overlap statistics are given in tables 7 and 8. These results demonstrate that in the LFA regime the TAKW model gives better results than the full AKW. Therefore, with the LFA, it makes sense to use TAKW model for estimating the EMRI signals buried in LISA data.

Table 7. True signals: overlaps of the AKW and TAKW in LFA against AKW in full LISA response.

  AKW+LFA versus AKW+LISA TAKW+LFA versus AKW+LISA
Signals EMRI-1-0 EMRI-1-1 EMRI-2-0 EMRI-1-0 EMRI-1-1 EMRI-2-0
Overlap A 0.238 0.750 0.893 0.978 0.886 0.986
Overlap E 0.191 0.844 0.913 0.952 0.874 0.977
Combined overlap 0.214 0.795 0.903 0.965 0.880 0.981

Table 8. Overlaps and SNRs of the detected signals for different sources.

    Med-0 Med-1 Med-2 Med-3
EMRI-2-0 Overlap A 0.501 0.102 0.219 0.185
  Overlap E 0.664 0.161 0.134 0.115
  Combined overlap 0.577 0.128 0.171 0.146
  SNR A 38.133 1.647 65.307 62.921
  SNR E 42.003 1.190 80.562 70.950
  Combined SNR 56.731 2.032 103.708 94.831
    High-0 High-1 High-2 High-3
EMRI-1-0 Overlap A 0.021 0.114 0.017 0.055
  Overlap E 0.018 0.084 0.041 0.125
  Combined overlap 0.019 0.098 0.027 0.083
EMRI-1-1 Overlap A 0.018 0.199 0.018 0.031
  Overlap E 0.007 0.286 0.024 0.057
  Combined overlap 0.012 0.238 0.021 0.042
  SNR A 11.173 11.262 2.688 19.777
  SNR E 15.589 10.227 3.303 17.406
  Combined SNR 19.180 15.213 4.258 26.346

4.3.2. True versus detected signals

Overlaps of the estimated signals (TAKW + LFA) with true signals (AKW + full LISA response) and their SNRs are shown in table 8. Different symbols indicate the correspondence between the best matches of estimated signals with true signals. The estimated medium mass signal 'Med-2' is believed to have a best match with the only medium mass source in MLDC 4 blind data. The low degree of overlaps and SNRs of the estimated high mass signals is due to the fact that for these sources the difference factor between the LFA and full LISA response is ∼6 rather than ∼3.

5. Conclusion

In this paper we report on the application of a Bayesian approach to statistical inference on the highly challenging problem of detecting and estimating parameters of GW signals produced by EMRI's in the data from a future space-based interferometric mission such as LISA. The LFA is used to model the LISA response and the Whittle likelihood as a realistic approximation to the unknown error distribution. Differences in the amplitudes of LFA and the LISA response function were found and temporarily fixed. Further investigations are required to overcome these discrepancies. We assume the noise power spectrum to be unknown and formulate an Inv-χ2 posterior with a conjugate prior distribution for the unknown noise parameters. We have implemented a parallel tempering MCMC algorithm to sample from the posterior distribution of all 14 unknown parameters, the first instance of a fully Markovian algorithm for EMRI detection and characterization in the context of MLDC LISA data analysis challenges. This algorithm has the potential to sample from multi-modal distributions and thus to avoid getting trapped in local modes. Our simulation results using single as well as multiple EMRI sources in MLDC training and blind data should be seen as a snapshot of an investigation still in progress. However, they show that this strategy holds great promise for posterior computation under more realistic conditions and hints at areas where the algorithm could be refined and tuned.

Acknowledgments

AA acknowledges support from the Higher Education Commission of Pakistan. The work of NC was partially supported by NSF grant PHY-0854790. We also thank the Max-Planck-Institut für Gravitationsphysik (Albert-Einstein-Institut), Hannover, Germany, for access to their computing facilities.

Footnotes

  • We realize that the state of LISA is in limbo after NASA's decision to leave the project, but the European Space Agency is investigating carrying on with a scaled down version of the project, and the conclusions we reach in this paper will certainly be applicable to that mission if it moves forward.

Please wait… references are loading.
10.1088/0264-9381/29/14/145014