Brought to you by:
Paper

Application of a Hough search for continuous gravitational waves on data from the fifth LIGO science run

, , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , and

Published 3 April 2014 © 2014 IOP Publishing Ltd
, , Citation J Aasi et al 2014 Class. Quantum Grav. 31 085014 DOI 10.1088/0264-9381/31/8/085014

0264-9381/31/8/085014

Abstract

We report on an all-sky search for periodic gravitational waves in the frequency range 50–1000 Hz with the first derivative of frequency in the range −8.9 × 10−10 Hz s−1 to zero in two years of data collected during LIGO's fifth science run. Our results employ a Hough transform technique, introducing a χ2 test and analysis of coincidences between the signal levels in years 1 and 2 of observations that offers a significant improvement in the product of strain sensitivity with compute cycles per data sample compared to previously published searches. Since our search yields no surviving candidates, we present results taking the form of frequency dependent, 95% confidence upper limits on the strain amplitude h0. The most stringent upper limit from year 1 is 1.0 × 10−24 in the 158.00–158.25 Hz band. In year 2, the most stringent upper limit is 8.9 × 10−25 in the 146.50–146.75 Hz band. This improved detection pipeline, which is computationally efficient by at least two orders of magnitude better than our flagship Einstein@Home search, will be important for 'quick-look' searches in the Advanced LIGO and Virgo detector era.

Export citation and abstract BibTeX RIS

1. Introduction

The focus of this article is the search for evidence of continuous gravitational waves (GWs), as might be radiated by nearby, rapidly spinning neutron stars, in data from the Laser interferometer gravitational-wave observatory (LIGO) [1]. The data used in this paper were produced during LIGO's fifth science run (S5) that started on November 4, 2005 and ended on October 1, 2007.

Spinning neutron stars are promising sources of GW signals in the LIGO frequency band. These objects may generate continuous GWs through a variety of mechanisms including non-axisymmetric distortions of the neutron star, unstable oscillation modes in the fluid part of the star and free precession [26]. Independently of the specific mechanism, the emitted signal is a quasi-periodic wave whose frequency changes slowly during the observation time due to energy loss through GW emission, and possibly other mechanisms. At an Earth-based detector the signal exhibits amplitude and phase modulations due to the motion of the Earth with respect to the source.

A number of searches have been carried out previously in LIGO data [718] including: targeted searches in which precise pulsar ephemerides from radio, x-ray or γ-ray observations can be used in a coherent integration over the full observation span; directed searches in which the direction of the source is known precisely, but for which little or no frequency information is known; and all-sky searches in which there is no information about location or frequency.

All-sky searches for unknown neutron stars must cope with a very large parameter space volume. Optimal methods based on coherent integration over the full observation time are completely unfeasible since the template bank spacing decreases dramatically with observation time, and even for a coherent time baseline of just few days, a wide-frequency-band all-sky search is computationally extremely challenging. Therefore hierarchical approaches have been proposed [1924] which incorporate semi-coherent methods into the analysis. These techniques are less sensitive for the same observation time but are computationally inexpensive. The Hough transform [7, 10, 21, 25, 26] is an example of such a method and has been used in previous wide-parameter-space searches published by the LIGO and Virgo Collaborations. Moreover it has also been used in the hierarchical approach for Einstein@Home searches, as the incoherent method to combine the information from coherently analyzed segments [14, 18].

In this paper we report the results of an all-sky search making use of the 'weighted Hough' method [10, 25, 26]. The 'weighted Hough' was developed to improve the sensitivity of the 'standard Hough' search [7, 21] and allows us to analyze data from multiple detectors, taking into account the different sensitivities.

The work presented here achieves improved sensitivity compared to previous Hough searches [7, 10] by splitting the run into two year-long portions and requiring consistency between signal levels in the two separate years for each candidate event, in addition to incorporating a χ2-test [27]. This new pipeline is efficient at rejecting background, allowing us to lower the event threshold and achieve improved sensitivity. The parameter space searched in our analysis covers the frequency range 50 < f < 1000 Hz and the frequency time-derivative range $-8.9\times 10^{-10}<\dot{f}<0$ Hz s−1. We detect no signals, so our results are presented as strain amplitudes h0 excluded at 95% confidence, marginalized over the above $\dot{f}$ interval.

Through the use of significant distributed computing resources [28], another search [18] has achieved better sensitivity on the same data as the search described here. But the Einstein@Home production run on the second year of S5 LIGO data required about 9.5 months, used a total of approximately 25 000 CPU (central processing unit) years [18], and required five weeks for the post-processing on a cluster with 6720 CPU cores. The search presented in this paper used only 500 CPU months to process each of the two years of data, representing a computational cost more than two orders of magnitude smaller. This is also an order of magnitude smaller than the computational cost of the semi-coherent 'PowerFlux' search reported in a previous paper [17].

The significance of our analysis is through offering an independent analysis to cross-check these results, and a method that allows the attainment of sensitivity close to that of the Einstein@Home search at substantially reduced computational burden. This technique will be particularly important in the advanced LIGO and Virgo detector when applied to 'quick-look' searches for nearby sources that may have detectable electromagnetic counterparts. Moreover, the Hough transform is more robust than other computationally efficient semi-coherent methods with respect to noise spectral disturbances [10] and phase modeling of the signal. In particular, it is also more robust than the Einstein@Home search to the non inclusion of second order frequency derivatives.

An important feature to note is that the sensitivity of the Hough search is proportional to $1/( N^{1/4} \sqrt{T_{{{{\rm coh}}}}})$ or $N^{1/4}/ \sqrt{T_{{{{\rm obs}}}}}$, assuming Tobs = NTcoh, being N the number of data segments coherently integrated over a time baseline Tcoh and combined using the Hough transform over the whole observation time Tobs, while for a coherent search over the whole observation time, the sensitivity is proportional to $1/ \sqrt{T_{{{{\rm obs}}}}}$. This illustrates the lost of sensitivity introduced combining the different data segments incoherently but, of course, this is compensated by the lesser computational requirements of the semi-coherent method.

For sufficiently short segments (Tcoh of the order of 30 min or less), the signal remains within a single Fourier frequency bin in each segment. In this case a simple Fourier transform can be applied as a coherent integration method. As the segment duration Tcoh is increased, it becomes necessary to account for signal modulations within each segment by computing the so-called $\mathcal{F}$-statistic [29] over a grid in the space of phase evolution parameters, whose spacing decreases dramatically with time baseline Tcoh. This results in a significant increase in the computational requirements of the search and also limits the significant thresholds for data points selection and the ultimate sensitivity of the search.

The search presented here is based on 30 min long coherent integration times, being this the reason for the significant reduction of the computational time compared to the Einstein@Home search [18] in which the span of each segment was set equal to 25 h. For an in-depth discussion on how to estimate and optimize the sensitivity of wide area searches for spinning neutron stars at a given computational cost, we refer the reader to [22, 23, 30].

This paper is organized as follows: section 2 briefly describes the LIGO interferometers and the data from LIGO's S5 run. Section 3 defines the waveforms we seek and the associated assumptions we have made. In section 4 we briefly review the Hough-transform method. Section 5 describes the χ2 test implemented for the analysis of the full S5 data. Section 6 gives a detailed description of the search pipeline and results. Upper limit computations are provided in section 7. The study of some features related to the χ2-veto is presented in section 8. Section 9 discusses variations, further improvements and capabilities of alternative searches. Section 10 concludes with a summary of the results.

2. Data from the LIGO's fifth science run

During LIGO's S5 run the LIGO detector network consisted of a 4-km interferometer in Livingston, Louisiana (called L1) and two interferometers in Hanford, Washington, one a 4-km and another 2-km (H1 and H2, respectively). The S5 run spanned a nearly two-year period of data acquisition. This run started at 16:00 UTC on November 4, 2005 at Hanford and at 16:00 UTC on November 14, 2005 at Livingston Observatory; the run ended at 00:00 UTC on October 1, 2007. During this run, all three LIGO detectors had displacement spectral amplitudes very near their design goals of 1.1 × 10−19 m ⋅ Hz−1/2 in their most sensitive frequency band near 150 Hz for the 4-km detectors and, in terms of GW strain, the H2 interferometer was roughly a factor of two less sensitive than the other two over most of the relevant band.

The data were acquired and digitized at a rate of 16384 Hz. Data acquisition was periodically interrupted by disturbances such as seismic transients (natural or anthropogenic), reducing the net running time of the interferometers. In addition, there were 1–2 week commissioning breaks to repair equipment and address newly identified noise sources. The resulting duty factors for the interferometers, defined as the fraction of the total run time when the interferometer was locked (i.e., all the interferometer control servos operating in their linear regime) and in its low configuration, were approximately 69% for H1, 77% for H2, and 57% for L1 during the first eight months. A nearby construction project degraded the L1 duty factor significantly during this early period of the S5 run. By the end of the S5 run, the cumulative duty factors had improved to 78% for H1, 79% for H2, and 66% for L1.

In the paper the data from each of the three LIGO detectors is used to search for continuous GW signals. In table 1 we provide the reference GPS initial and final times for the data collected for each detector, together with the number of hours of data used for the analysis, where each data segment used was required to contain at least 30 min of continuous interferometer operation.

Table 1. The reference GPS initial and final time for the data collected during the LIGO's S5 run, together with the number of hours of data used for the analysis.

  First year Second year
Detector Start End Hours Start End Hours
H1 815 410 991 846 338 742 5710 846 375 384 877 610 329 6295
H2 815 201 292 846 340 583 6097.5 846 376 386 877 630 716 6089
L1 816 070 323 846 334 700 4349 846 387 978 877 760 976 5316.5

3. The waveform model

Spinning neutron stars may generate continuous GWs through a variety of mechanisms. Independently of the specific mechanism, the emitted signal is a quasi-periodic wave whose frequency changes slowly during the observation time due to energy loss through GW emission, and possibly other mechanisms. The form of the received signal at the detector is

Equation (1)

where t is time in the detector frame, ψ is the polarization angle of the wave and F+, × characterize the detector responses for the two orthogonal polarizations [29, 31]. For an isolated quadrupolar GW emitter, characterized by a rotating triaxial-ellipsoid mass distribution, the individual components h+, × have the form

Equation (2)

where ι describes the inclination of the source's rotation axis to the line of sight, h0 is the wave amplitude and Φ(t) is the phase evolution of the signal.

For such a star, the GW frequency, f, is twice the rotation frequency and the amplitude h0 is given by

Equation (3)

where d is the distance to the star, Izz is the principal moment of inertia with respect to its spin axis, ε the equatorial ellipticity of the star, G is Newton's constant and c is the speed of light.

Note that the search method used in this paper is sensitive to periodic signals from any type of isolated GW source, though we present upper limits in terms of h0. Because we use the Hough method, only the instantaneous signal frequency in the detector frame, 2πf(t) = dΦ(t)/dt, needs to be calculated. This is given, to a very good approximation, by the non-relativistic Doppler expression:

Equation (4)

where $\hat{f}(t)$ is the instantaneous signal frequency in the solar system barycenter (SSB), v(t) is the detector velocity with respect to the SSB frame and n is the unit-vector corresponding to the sky location of the source. In this analysis, we search for $\hat{f}(t)$ signals well described by a nominal frequency f0 at the start time of the S5 run t0 and a constant first time derivative $\dot{f}$, such that

Equation (5)

These equations ignore corrections to the time interval tt0 at the detector compared with that at the SSB and relativistic corrections. These corrections are negligible for the search described here.

4. The Hough transform

The Hough transform is a well known method for pattern recognition that has been applied to the search for continuous GWs. In this case the Hough transform is used to find hypothetical signals whose time-frequency evolution fits the pattern produced by the Doppler modulation of the detected frequency, due to the Earth's rotational and orbital motion with respect to the SSB, and the time derivative of the frequency intrinsic to the source. Further details can be found in [7, 21, 25, 26]; here we only give a brief summary.

The starting point for the Hough transform are N short Fourier transforms (SFTs). Each of these SFTs is digitized by setting a threshold ρth on the normalized power

Equation (6)

Here ${\tilde{x}_k}$ is the discrete Fourier transform of the data, the frequency index k corresponds to a physical frequency of fk = k/Tcoh, Sn(fk) is the single sided power spectral density of the detector noise and Tcoh is the time baseline of the SFT. The kth frequency bin is selected if ρk ⩾ ρth, and rejected otherwise. In this way, each SFT is replaced by a collection of zeros and ones called a peak-gram. This is the simplest method of selecting frequency bins, for which the optimal choice of the threshold ρth is 1.6 [21]. Alternative conditions could be imposed [3234], that might be more robust against spectral disturbances.

For our choice, the probability that a frequency bin is selected is $q = {\rm e}^{-\rho _{{{\rm th}}}}$ for Gaussian noise and η, given by

Equation (7)

is the corresponding probability in the presence of a signal. λk is the signal to noise ratio (SNR) within a single SFT, and for the case when there is no mismatch between the signal and the template:

Equation (8)

with $\tilde{h}(f)$ being the Fourier transform of the signal h(t).

Several flavors of the Hough transform have been developed [21, 25, 35] and used for different searches [7, 10, 14]. The Hough transform is used to map points from the time-frequency plane of our data (understood as a sequence of peak-grams) into the space of the source parameters. Each point in parameter space corresponds to a pattern in the time-frequency plane, and the Hough number count n is the weighted sum of the ones and zeros, $n_k^{(i)}$, of the different peak-grams along this curve. For the 'weighted Hough' this sum is computed as

Equation (9)

where the choice of weights is optimal, in the sense of [25], if defined as

Equation (10)

where $F_{+1/2}^{(i)}$ and $F_{\times 1/2}^{(i)}$ are the values of the beam pattern functions at the mid point of the ith SFT and are normalized according to

Equation (11)

The natural detection statistic is the significance (or critical ratio) defined as:

Equation (12)

where 〈n〉 and σ are the expected mean and standard deviation for pure noise. Furthermore, the relation between the significance and the false alarm probability α, in the Gaussian approximation [21], is given by

Equation (13)

5. The χ2 veto

χ2 time-frequency discriminators are commonly used for GW detection. Originally, they were designed for broadband signals with a known waveform in a data stream [36]. But they can be adapted for narrowband continuous signals, as those expected from rapidly rotating neutron stars. The essence of these tests is to 'break up' the data (in time or frequency domain) and to see if the response in each chunk is consistent with what would be expected from the purported signal.

In this paper, a chi-square test is implemented as a veto, in order to reduce the number of candidates in the analysis of the full S5 data. The idea for this χ2 discriminator is to split the data into p non-overlapping chunks, each of them containing a certain number of SFTs {N1, N2, ..., Np}, such that

Equation (14)

and analyze them separately, obtaining the Hough number-count nj which, for the same pattern across the different chunks, would then satisfy

Equation (15)

where n is the total number-count for a given point in parameter space. The χ2 statistic will look along the different chunks to see if the number count accumulates in a way that is consistent with the properties of the signal and the detector noise. Small values of χ2 are consistent with the hypothesis that the observed significance arose from a detector output which was a linear combination of Gaussian noise and the continuous wave signal. Large values of χ2 indicate either the signal did not match the template or that the detector noise was non-Gaussian.

In the following subsections we derive a χ2 discriminator for the different implementations of the Hough transform and show how the veto curve was derived for LIGO S5 data.

5.1. The standard Hough

In the simplest case in which all weights are set to unity, the expected value and variance of the number count are

Equation (16)

Equation (17)

Consider the p quantities defined by

Equation (18)

With this definition, it holds true that

Equation (19)

and the expectation value of the square of Δnj is

Equation (20)

Therefore we can define the χ2 discriminator statistic by

Equation (21)

This corresponds to a χ2-distribution with p − 1 degrees of freedom. To implement this discriminator, we need to measure, for each point in parameter space, the total number-count n, the partial number-counts nj and assume a constant value of η = n/N.

5.2. The weighted Hough

In the case of the weighted Hough the result given by equation (21) can be generalized. Let Ij be the set of SFT indices for each different p chunks, thus the mean and variance of the number-count become

Equation (22)

and we can define

Equation (23)

so that 〈Δnj〉 = 0, $\sum _{j=1}^p \Delta n_j=0$. Hence, the χ2 discriminator would now be:

Equation (24)

In a given search, we can compute the $\sum _{i\in I_j} w_i$, $\sum _{i\in I_j} w_i^2$ for each of the p chunks, but the different ηi values cannot be measured from the data itself because they depend on the exact SNR for each single SFT as defined in equations (7) and (8). For this reason, the discriminator we proposed is constructed by replacing ηi → η*, where η* = n/N. In this way, from equation (24) we get

Equation (25)

In principle, one is free to choose the different p chunks of data as one prefers, but it is reasonable to split the data into segments in such a way that they would contribute a similar relative contribution to the total number count. Therefore we split the SFT data in such a way that the sum of the weights in each block satisfies

Equation (26)

Further details and applications of this χ2 on LIGO S4 data can be found in [27].

5.3. The S5 χ2 veto curve

We study the behavior of this χ2 discriminator in order to characterize the χ2-significance plane in the presence of signals and derive empirically the veto curve.

For this purpose we use the full LIGO S5 SFT data, split into both years, in the same way that is done in the analysis, and inject a large number of Monte Carlo simulated continuous GW signals into the data, varying the amplitude, frequency, frequency derivative, sky location, as well as the nuisance parameters cos ι, ψ and ϕ0 of the signals. Those injections are analyzed with the multi-interferometer Hough code using the same grid resolution in parameter space as is used in the search.

To characterize the veto curve, nine 0.25 Hz bands, spread in frequency and free of known large spectral disturbances have been selected. These are: 102.5, 151, 190, 252.25, 314.1, 448.5, 504.1, 610.25 and 710.25 Hz. Monte Carlo injections in those bands have been performed separately in the data from both years. Since the results were comparable for both years a single veto curved is derived.

In total 177 834 injections are considered with a significance value lower than 70. The results of these injections in terms of (s, χ2) are presented in figure 1. The χ2 values obtained correspond to those by splitting the data in p = 16 segments.

Figure 1.

Figure 1. Top left: mean value of the significance versus mean of the χ2 and the fitted power law curve for 177 834 simulated injected signals. Top right: Mean value of the significance versus mean χ2 standard deviation and the fitted power law curve. Bottom: significance-χ2 plane for the injections, together with the fitted mean curve (dot-dashed line) and the veto curve (dashed line) corresponding to the mean χ2 plus five times its standard deviation.

Standard image High-resolution image

Then we proceed as follows: first we sort the points with respect to the significance, and we group them in sets containing 1000 points. For each set we compute the mean value of the significance, the mean of the χ2 and its standard deviation. With these reduced set of points we fit two power laws p − 1 + asc and $\sqrt{2p-2}+b \, s^d$ to the (mean s, mean χ2) and (mean s, std χ2) respectively, obtaining the following coefficients (with 95% confidence bounds):

The veto curve we will use in this analysis corresponds to the mean curve plus five times the standard deviation

Equation (27)

This curve vetoes 25 of the 177 834 injections considered with significance lower than 70, that could translate into a false dismissal rate of 0.014. In figure 1 we show the fitted curves and the $ \bar{\chi }^2$ veto curve compared to the result of the injections.

6. Description of the all-sky search

In this paper, we use a new pipeline to analyze the data from the S5 run of the LIGO detectors to search for evidence of continuous GWs, that might be radiated by nearby unknown rapidly spinning isolated neutron stars. Data from each of the three LIGO interferometers is used to perform the all-sky search. The key difference from previous searches is that, starting from 30 min SFTs, we perform a multi-interferometer search analyzing separately the two years of the S5 run, and we study coincidences among the source candidates produced by the first and second years of data. Furthermore, we use a χ2 test adapted to the Hough transform searches to veto potential candidates. The pipeline is shown schematically in figure 2.

Figure 2.

Figure 2. Pipeline of the Hough search.

Standard image High-resolution image

A separate search was run for each successive 0.25 Hz band within the frequency range 50–1000 Hz and covering frequency time derivatives in the range −8.9 × 10−10 Hz s−1 to zero. We use a uniform grid spacing equal to the size of a SFT frequency bin, δf = 1/Tcoh = 5.556 × 10−4 Hz. The resolution $\delta \dot{f}$ is given by the smallest value of $\dot{f}$ for which the intrinsic signal frequency does not drift by more than a frequency bin during the observation time Tobs in the first year: $\delta \dot{f} = \delta f / T_{{{{\rm obs}}}}\sim 1.8\times 10^{-11} \,\mathrm{Hz} \mathrm{\hbox{ }s}^{-1}$. This yields 51 spin-down values for each frequency. $\delta \dot{f}$ is fixed to the same value for the search on the first and the second year of S5 data. The sky resolution, δθ, is frequency dependent, as given by equation (4.14) of [21], that we increase by a factor 2. As explained in detail in section V.B.1 of [10], the sky-grid spacing can be increased with a negligible loss in SNR, and for previous PowerFlux searches [10, 13, 17] a factor 5 of increase was used in some frequency ranges to analyze LIGO S4 and S5 data.

The set of SFTs are generated directly from the calibrated data stream, using 30-min intervals of data for which the interferometer is operating in what is known as science mode. With this requirement, we search 32295 SFTs from the first year of S5 (11402 from H1, 12195 from H2 and 8698 from L1) and 35401 SFTs from the second year (12590 from H1, 12178 from H2 and 10633 from L1).

6.1. A two-step hierarchical Hough search

The approach used to analyze each year of data is based on a two-step hierarchical search for continuous signals from isolated neutron stars. In both steps, the weighted Hough transform is used to find signals whose frequency evolution fits the pattern produced by the Doppler shift and the spin-down in the time-frequency plane of the data. The search is done by splitting the frequency range in 0.25 Hz bands and using the SFTs from multiple interferometers.

In the first stage, and for each 0.25 Hz band, we break up the sky into smaller patches with frequency dependent size in order to use the look up table approach to compute the Hough transform, which greatly reduces the computational cost. The look up table approach benefits from the fact that, according to the Doppler expression (4), the set of sky positions consistent with a given frequency bin fk at a given time correspond to annuli on the celestial sphere centered on the velocity vector v(t). In the look up table approach, we precompute all the annuli for a given time and a given search frequency mapped on the sky search grid. Moreover, it turns out that the mapped annuli are relatively insensitive to changes in frequency and can therefore be reused a large number of times. The Hough map is then constructed by selecting the appropriate annuli out of all the ones that have been found and adding them using the corresponding weights. A detailed description of the look up table approach with further details of implementation choices can be found in [21].

But limitations on the memory of the computers constrain the volume of data (i.e., the number of SFTs) that can be analyzed at once and the parameter space (e.g., size and resolution of the sky-patches and number of spin-down values) we can search over. For this reason, in this first stage, we select the best 15000 SFTs (according to the noise floor and the beam pattern functions) for each frequency band and each sky-patch and apply the Hough transform on the selected data. The size of the sky-patches ranges from ∼0.4 rad × 0.4 rad at 50 Hz to ∼0.07 rad × 0.07 rad at 1 kHz and we calculate the weights only for the center of each sky-patch. This was set in order to ensure that the memory usage will never exceed the 0.8 GB and this search could run on the Merlin/Morgane dual compute cluster at the Albert Einstein Institute128. A top-list keeping the best 1000 candidates is produced for each 0.25 Hz band for the all-sky search.

Figure 3 shows the histograms of the percentage of SFTs that each detector contributes for the different sky locations for a band at 420 Hz for the first year of S5 data. At this particular frequency, the detector that contributes the most is H1 between 44–64%, giving the maximum contribution near the poles, L1 contributes between 28.1–45.5% with its maximum around the equator, and H2 contributes at most 21.7% of the SFTs. As shown in figure 1 in [37], the maximum contribution of H2 corresponds to those sky regions where L1 contributes the least. If SFT selection had been based only upon the weights due to the noise floor, the H2 detector would not have contributed at all in this first stage.

Figure 3.

Figure 3. Histograms of the percentage of SFTs that each detector has contributed in the first stage to the all-sky search. These figures correspond to a 0.25 Hz band at 420 Hz for the first year of S5 data. The vertical axes are the number of sky-patches.

Standard image High-resolution image

In a second stage, we compute the χ2 value for all the candidates in the top-list in each 0.25 Hz band. This is done by dividing the data into 16 chunks and summing weighted binary zeros or ones along the expected path of the frequency evolution of a hypothetical periodic GW signal in the digitized time-frequency plane of our data. Since there are no computational limitations, we use the complete set of available SFTs from all three interferometers, and we also get a new value of the significance using all the data. In this way we reduce the mismatch of the template, since the number count is obtained without the roundings introduced by the look up table approach and the weights are computed for the precise sky location and not for the center of the corresponding patch. All these refinements contribute also to a potential improvement of sensitivity when a threshold is subsequently applied to the recomputed significance (described below).

Figure 4 shows the maximum-significance value in each 0.25 Hz band obtained for the first and second years of S5 data.

Figure 4.

Figure 4. Maximum value of the significance for each 0.25 Hz band for both years of LIGO S5 data.

Standard image High-resolution image
Figure 5.

Figure 5. Percentage of the number of candidates vetoed due to a large χ2 value for each 0.25 Hz band for both years of LIGO S5 data.

Standard image High-resolution image
Figure 6.

Figure 6. Left: Number of templates analyzed in each 0.25 Hz band as a function of frequency. Right: Significance threshold for a false alarm level of 1/(number of templates) (solid line), compared to 10/(number of templates) (dashed line) and 0.5/(number of templates) (dot-dashed line) in each band.

Standard image High-resolution image
Figure 7.

Figure 7. Surviving candidates from both years after applying the χ2 veto and setting a threshold in the significance.

Standard image High-resolution image
Figure 8.

Figure 8. Significance of the coincidence candidates from the two years of LIGO S5 data. The upper and lower plots correspond to the first and second year respectively.

Standard image High-resolution image

6.2. The post-processing

After the multi-interferometer Hough search is performed on each year of S5 data between 50 and 1000 Hz, a top list keeping the best 1000 candidates is produced for each 0.25 Hz band. This step yields 3.8 × 106 candidates for each year. The post-processing of these results has the following steps:

  • (i)  
    Remove those 0.25 Hz bands that are affected by power lines or violin modes.A total of 96 bands are removed. These bands are given in table 2.
  • (ii)  
    Remove all the 0.25 Hz bands for which the χ2 vetoes more than a 95% of the elements in the top list.Figure 5 shows the χ2 veto level for all the frequency bands for both years. With this criterion, 144 and 131 0.25 Hz bands would be vetoed for the first and second year of data respectively. These first two steps leave a total 3548 bands in which we search for coincidence candidates and set upper limits; a total of 252 bands were discarded.
  • (iii)  
    Set a threshold on the significance.Given the relation of the Hough significance and the Hough false alarm probability (see equation (13)), we set a threshold on the candidate's significance that corresponds to a false alarm of 1/(number of templates) for each 0.25 Hz band. Figure 6 shows the value of this threshold at different frequencies.
  • (iv)  
    Apply the χ2 veto.From the initial 3.8 × 106 elements in the top list for each year, after excluding the noisy bands, applying the χ2 veto and setting a threshold on the significance, the number of candidates remaining in the 3548 'clean' bands are 31 427 for the first year and 50 832 for the second year. Those are shown in figure 7.
  • (v)  
    Selection of coincident candidates.For each of the four parameters: frequency, spin-down and sky location, we set the coincidence window with a size equal to five times the grid spacing used in the search and centered on the values of the candidates parameters. Therefore the coincidence window always contains 625 cells in parameter space, with frequency-dependent size, according to the search grid. This window is computed for each of the candidates selected from the first year of data and then we look for coincidences among the candidates of the second year, making sure to translate their frequency to the reference time of the starting time of the run, taking into account their spin-down values. Extensive analysis of software injected signals, in different frequency bands, have been used to determine the size of this coincidence window. This was done by comparing the parameters of the most significant candidates of the search, using the same pipeline, in both years of data.With this procedure, we obtain 135 728 coincidence pairs, corresponding to 5823 different candidates of the first year that have coincidences with 7234 different ones of the second year. Those are displayed in figure 8. All those candidates cluster in frequency in 34 groups. The most significant outlier at 108.857 Hz corresponds to a simulated pulsar signal injected into the instrument as a test signal. The most significant events in each cluster are shown in tables 3 and 4. Notice how with this coincidence step the overall number of candidates has been reduced by a factor 5.4 for the first year and a factor 7.0 for the second year. Furthermore, without the coincidence step, the candidates are spread over all frequencies, whereas the surviving coincident candidates are clustered in a few small regions, illustrating the power of this procedure on real data.

Table 2. Initial frequency of the 0.25 Hz bands excluded from the search.

Excluded Bands (Hz) Description
[n60 − 0.25, n60 + 0.25] n = 1 to 16 Power lines
[343.0, 344.75] Violin modes
[346.5, 347.75] Violin modes
[348.75, 349.25] Violin modes
[685.75, 689.75] Violin mode harmonics
[693.0, 695.5] Violin mode harmonics
[697.5, 698.75] Violin mode harmonics

Table 3. Summary of first year coincidence candidates, including the frequency band, the number of candidates in each cluster and showing the details of the most significant candidate in each of the 34 clusters. Shown are the significance s, the χ2 value, the detected frequency at the start of the run (SSB frame) f0, the spin-down $\dot{f}$, and the sky position (RA, dec).

  Band (Hz) Num. s χ2 f0 (Hz) $\dot{f}$ (Hz s−1)  RA (rad)  dec (rad)
 1   50.001–50.003   12  7.103  64.115  50.0022  0 −1.84  0.69
 2  50.997–51.004    7  5.431  57.322  51.0028  0 −2.73  0.94
 3  52.000–52.016   44  6.886  54.108  52.0139 −21.4e-11 −1.99 −0.08
 4  52.786–52.793    6  6.094  40.089  52.7911  0  0  1.33
 5  53.996–54.011 1136 11.880  60.583  54.0039 −10.7e-11  2.68 −1.21
 6  54.996–55.011   82  6.995  48.289  55.0067  0  3.00 −0.47
 7  55.749–55.749    7  5.940  37.902  55.7489  0 −0.23  1.12
 8  56.000–56.016  167  8.085  66.130  56.0056 −7.1e-11  0.20  0.22
 9  56.997–57.011 1370 24.751 175.459  57.0028 −12.5e-11 −2.10  1.40
10  58.000–58.015   89  7.901  61.360  58.0128 −35.7e-11 −2.32  0.43
11  61.994–62.000  195  6.597  26.445  61.9989  0  0.98  0.21
12  62.996–63.009 1324 23.188 131.619  63.0028 −8.9e-11  0 −1.40
13  64.996–65.000  101  6.252  52.374  64.9994 −5.3e-11 −0.17  0.03
14  65.378–65.381    3  5.586  29.000  65.3806 −16.1e-11 −2.17  1.15
15  65.994–66.013  919 11.662  35.587  65.9994 −7.1e-11  0.15  1.38
16  67.006–67.006    1  5.801  11.950  67.0056 −17.8e-11 −1.13 −0.20
17  67.993–68.009   39  6.061  44.319  68.0017 −14.3e-11 −1.55  0.66
18  72.000–72.000    4  5.604  17.668  72.0000 −1.8e-11  1.57 −1.11
19  86.002–86.024   14  6.786  47.054  86.0150 −17.8e-11  1.78  1.09
20  90.000–90.000    2  5.554  58.338  90.0000 −3.6e-11  1.54 −1.05
21 108.857–108.860   50 64.850 552.161 108.8570  0  3.10 −0.60
22 111.998–111.998    1  5.673  46.493 111.9980 −7.1e-11 −0.54  1.18
23 118.589–118.613   18  7.072  52.181 118.5990 −57.1e-11  2.90 −0.46
24 160.000–160.000    1  5.571  18.847 160.0000  0  1.56 −1.15
25 178.983–179.026   21  7.483  43.380 179.0010 −3.6e-11 −1.59  1.17
26 181.000–181.038    8  6.309  13.174 181.0170 −8.9e-11 −1.12 −0.89
27 192.000–192.002    5  7.976  41.042 192.0000 −1.8e-11 −1.51  1.17
28 341.763–341.765    3  6.292  39.340 341.7630 −1.8e-11 −1.63  1.21
29 342.680–342.684    5  6.113  36.893 342.6800 −3.6e-11  1.47 −1.14
30 345.721–345.724   17  6.835  38.540 345.7230 −12.5e-11 −1.07  1.44
31 346.306–346.316    9  6.973  30.298 346.3070 −3.6e-11  1.32 −1.13
32 394.099–394.100    3 10.617  95.063 394.1000  0 −1.58  1.16
33 575.163–575.167   57 26.058 146.929 575.1640 −1.8e-11 −2.53  0.06
34 671.728–671.733  101 13.878 132.197 671.7290  0  1.54 −1.17

Table 4. Summary of 2nd year coincidence candidates, showing the details of the most significant candidate in each of the 34 clusters.

  Band (Hz) Num. s χ2 f0 (Hz) $\dot{f}$ (Hz s−1) RA (rad) dec (rad)
 1  50.001–50.004   10  9.345   77.005  50.0006 −5.3e-11  1.25 −1.05
 2  50.993–51.003   16  7.106   17.670  51.0011 −1.8e-11 −1.95  1.32
 3  52.000–52.012   38  8.800   30.888  52.0094 −17.8e-11 −2.17 −0.08
 4  52.784–52.792   14 13.269  128.801  52.7911 −7.1e-11  0.77  1.31
 5  53.996–54.006 1380 23.431   69.473  54.0033 −14.3e-11 −1.92  1.31
 6  54.995–55.008  405 11.554   73.521  55.0056 −5.3e-11 −2.98  0.13
 7  55.748–55.749   46  7.102   34.151  55.7489 −1.8e-11 −0.10  0.13
 8  56.000–56.008  161 11.478   64.077  56.0006 −1.8e-11  1.58 −1.13
 9  56.996–57.004 1478 47.422  412.817  56.9989  0 −0.31  1.43
10  58.000–58.008  166 11.675   65.764  58.0000  0  1.59 −1.13
11  61.993–62.001  400  9.361   56.108  61.9989  0 −1.16  1.02
12  62.997–63.004 1138 42.432  360.615  63.0039 −5.3e-11 −1.51  1.41
13  64.994–65.000  288 12.557  119.361  64.9978  0  0.61 −1.14
14  65.377–65.378    2  5.777   34.580  65.3772 −14.3e-11 −1.84  1.03
15  65.995–66.006 1162 20.864   67.221  66.0022 −5.3e-11 −1.51  1.40
16  66.999–66.999    5  6.207   19.546  66.9989 −17.8e-11 −1.08  0.03
17  67.994–68.007  199 11.682   82.662  68.0006 −1.8e-11  1.65 −1.14
18  71.999–72.000   10  6.802   45.727  72.0000 −1.8e-11  1.41 −1.14
19  86.002–86.013   15  7.368   54.707  86.0094 −10.7e-11  2.10  0.75
20  89.999–90.000    4  9.008   86.486  90.0000  0  1.54 −1.19
21 108.857–108.858   27 77.157 1613.230 108.8580 −5.3e-11  2.99 −0.71
22 111.996–111.996    1  5.775   59.202 111.9960 −7.1e-11 −0.37  1.00
23 118.579–118.589   19  7.398   69.372 118.5820 −41.0e-11  2.79 −0.68
24 160.000–160.000    2  7.898   15.079 160.0000  0  1.60 −1.17
25 178.984–179.014   24  7.089   33.229 179.0010 −10.7e-11 −1.79  1.35
26 180.998–181.019    8  6.587   28.605 181.0180 −28.5e-11 −0.39  1.07
27 191.999–192.001    8  6.582   45.301 192.0000 −3.6e-11  1.51 −1.19
28 341.762–341.764   19  7.859   71.736 341.7630 −3.6e-11 −1.63  1.17
29 342.677–342.680   19  7.569   45.255 342.6790 −14.3e-11  1.47 −1.11
30 345.718–345.721   12  7.890   44.570 345.7200 −12.5e-11 −0.91  1.38
31 346.303–346.309   14  7.803   45.756 346.3070  0  1.46 −1.21
32 394.099–394.100    2  9.877   95.232 394.1000  0 −1.58  1.16
33 575.163–575.165   53 40.576  415.830 575.1640 −1.8e-11 −2.53  0.06
34 671.729–671.732   87 13.600  116.261 671.7320 −1.8e-11  1.59 −1.15

Noise lines were identified by previously performed searches [13, 14, 17, 18, 38] as well as the search described in this paper. Several techniques were used to identify the causes of outliers, including the calculation of the coherence between the interferometer output channel and physical environment monitoring channels and the computation of high resolution spectra. A dedicated analysis code 'FScan' [39] was also created specifically for identification of instrumental artifacts. Problematic noise lines were recorded and monitored throughout S5.

In addition, a number of particular checks were performed on the coincidence outliers, including: a detailed study of the full top-list results, for those 0.25 Hz bands where the candidates were found—in order to check if candidates are more dominant in a given year, or if they cluster in certain regions of parameter space; and a second search using the data of the two most sensitive detectors, H1 and L1 separately—in order to see if artifacts could be associated to a given detector, consistent with the observed spectra.

All of the 34 outliers were investigated and were all traced to instrumental artifacts or hardware injections (see details in table 5). Hence the search did not reveal any true continuous GW signals.

Table 5. Description of the coincidence outliers, together with the maximum value of the significance in both years.

  Bands (Hz) s 1y s 2y Comment
 1   50.001–50.004  7.103  9.345 L1 1 Hz Harmonic from control/data acquisition system
 2  50.993–51.004  5.431  7.106 L1 1 Hz Harmonic from control/data acquisition system
 3  52.000–52.016  6.886  8.800 L1 1 Hz Harmonic from control/data acquisition system
 4  52.784–52.793  6.094 13.269 Instrumental line in H1
 5  53.996–54.011 11.880 23.431 Pulsed heating sideband on 60 Hz mains
 6  54.995–55.011  6.995 11.554 1 Hz Harmonic from control/data acquisition system
 7  55.748–55.749  5.940  7.102 Instrumental line in L1
 8  56.000–56.016  8.085 11.478 L1 1 Hz Harmonic from control/data acquisition system
 9  56.996–57.011 24.751 47.422 Pulsed heating sideband on 60 Hz mains
10  58.000–58.015  7.901 11.675 L1 1 Hz Harmonic from control/data acquisition system
11  61.993–62.001  6.597  9.361 L1 1 Hz Harmonic from control/data acquisition system
12  62.996–63.009 23.188 42.432 Pulsed heating sideband on 60 Hz mains
13  64.994–65.000  6.252 12.557 L1 1 Hz Harmonic from control/data acquisition system
14  65.377–65.381  5.586  5.777 Instrumental line in L1—member of offset 1 Hz comb
15  65.994–66.013 11.662 20.864 Pulsed sideband on 60 Hz mains
16  66.999–67.006  5.801  6.207 L1 1 Hz Harmonic from control/data acquisition system
17  67.993–68.009  6.061 11.682 1 Hz Harmonic from control/data acquisition system
18  71.999–72.000  5.604  6.802 1 Hz Harmonic from control/data acquisition system
19  86.002–86.024  6.786  7.368 Instrumental line in H1
20  89.999–90.000  5.554  9.008 Instrumental line in H1
21 108.857–108.860 64.850 77.157 Hardware injection of simulated signal (ip3)
22 111.996–111.998  5.673  5.775 16 Hz harmonic from data acquisition system
23 118.579–118.613  7.072  7.398 Sideband of mains at 120 Hz
24 160.000–160.000  5.571  7.898 16 Hz harmonic from data acquisition system
25 178.983–179.026  7.483  7.089 Sideband of mains at 180 Hz
26 180.998–181.038  6.309  6.587 Sideband of mains at 180 Hz
27 191.999–192.002  7.976  6.582 16 Hz harmonic from data acquisition system
28 341.762–341.765  6.292  7.859 Sideband of suspension wire resonance in H1
29 342.677–342.684  6.113  7.569 Sideband of suspension wire resonance in H1
30 345.718–345.724  6.835  7.890 Sideband of suspension wire resonance in H1
31 346.303–346.316  6.973  7.803 Sideband of suspension wire resonance in H1
32 394.099–394.100 10.617  9.877 Sideband of calibration line at 393.1 Hz in H1
33 575.163–575.167 26.058 40.576 Hardware injection of simulated signal (ip2)
34 671.728–671.733 13.878 13.600 Instrumental line in H1

7. Upper limits estimation and astrophysical reach

The analysis of the Hough search presented here has not identified any convincing continuous GW signal. Hence, we proceed to set upper limits on the maximum intrinsic GW strain h0 that is consistent with our observations for a population of signals described by an isolated triaxial rotating neutron star.

As in the previous S2 and S4 searches [7, 10], we set a population-based frequentist upper limit, assuming random positions in the sky, in the GW frequency range [50, 1000] Hz and with spin-down values in the range −8.9 × 10−10 Hz s−1 to zero. The rest of the nuisance parameters, cos ι, ψ and ϕ0, are assumed to be uniformly distributed. As commonly done in all-sky, all-frequency searches, the upper limits are given in different frequency sub-bands, here chosen to be 0.25 Hz wide. Each upper limit is based on the most significant event from each year in its 0.25 Hz band. Our goal is to find the value of h0 (denoted $h_0^{95\%}$) such that 95% of the signal injections at this amplitude would be recovered by our search and are more significant than the most significant candidate from the actual search in that band, thus providing the 95% confidence all-sky upper limit on h0.

Our procedure for setting upper-limits uses partial Monte Carlo signal injection studies, using the same search pipeline as described above, together with an analytical sensitivity estimation. As in the previous S4 Hough search [10], upper limits can be computed accurately without extensive Monte Carlo simulations. Up to a constant factor C, that depends on the grid resolution in parameter space, they are given by

Equation (28)

where

Equation (29)

Si is the average value of the single sided power spectral noise density of the ith SFT in the corresponding frequency sub-band, αH is the false alarm and βH the false dismissal probability.

The utility of this fit is that having determined the value of C in a small frequency range, it can be extrapolated to cover the full bandwidth without performing any further Monte Carlo simulations. Figure 9 shows the value of the constant C for a number of 0.1 Hz frequency bands. More precisely, this is the ratio of the upper limits measured by means of Monte-Carlo injections in the multi-interferometer Hough search to the quantity $h_0^{95\%}/C$ as defined in equation (28). The value of $\mathcal{S}$ is computed using the false alarm αH corresponding to the observed loudest event, in a given frequency band, and a false dismissal rate βH = 0.05, in correspondence to the desired confidence level of 95%, i.e., $\mathcal{S}\rightarrow s^*/ \sqrt{2} + \textrm {erfc}^{-1}(0.1)$, where s* is the highest significance value in the frequency band. This yields a scale factor C of 8.32 ± 0.19 for the first year and 8.25 ± 0.16 for the second year of S5. With these values we proceed to set the upper limits for all the frequency bands. The validity of equation (28) was studied in [10] using LIGO S4 data. In that paper upper limits were measured for each 0.25 Hz frequency band from 100 to 1000 Hz using Monte Carlo injections and compared with those prescribed by this analytical approximation. Such comparison study showed that the values obtained using equation (28) have an error smaller than 5% for bands free of large instrumental disturbances. For an in-depth study of how to analytically estimate the sensitivity of wide parameter searches for GW pulsars, we refer the reader to [30].

Figure 9.

Figure 9. Ratio of the upper limits measured by means of Monte-Carlo injections in the multi-interferometer Hough search to the quantity $h_0^{95\%}/C$ as defined in Equation (28). The top figure corresponds to the first year of S5 data and the bottom one to the second year. The comparison is performed by doing 500 Monte-Carlo injections for 10 different amplitude in several small frequency bands. 153 and 144 frequency bands have been used for the first and second year respectively.

Standard image High-resolution image

The 95% confidence all-sky upper limits on h0 from this multi-interferometer search for each year of S5 data are shown in figure 10. The best upper limits correspond to 1.0 × 10−24 for the first year of S5 in the 158–158.25 Hz band, and 8.9 × 10−25 for the second year in the 146.5–146.75 Hz band. There is an overall 15% calibration uncertainty on these upper limits. No upper limits are provided in the 252 vetoed bands, that were excluded from the coincidence analysis, since the analytical approximation would not be accurate enough. These excluded frequency bands are marked in the figure.

Figure 10.

Figure 10. The 95% confidence all-sky upper limits on h0 from the hierarchical Hough multi-interferometer search together with excluded frequency bands. The best upper limits correspond to 1.0 × 10−24 for the S5 first year in the 158–158.25 Hz band, and 8.9 × 10−25 for the S5 second year in the 146.5–146.75 Hz band.

Standard image High-resolution image

Figure 11 provides the maximum astrophysical reach of our search for each year of the S5 run. The top panel shows the maximum distance to which we could have detected a source emitting a continuous wave signal with strain amplitude $h_0^{95\%}$. The bottom panel does not depend on any result from the search. It shows the corresponding ellipticity values as a function of frequency. For both plots the source is assumed to be spinning down at the maximum rate considered in the search −8.9 × 10−10 Hz s−1, and emitting in GWs all the energy lost. This follows formulas in paper [10] and assumes the canonical value of 1038 kg m2 for Izz in equation (3).

Figure 11.

Figure 11. These plots represent the distance range (in kpc) and the maximum ellipticity, respectively, as a function of frequency. Both plots are valid for neutron stars spinning down solely due to gravitational radiation and assuming a spin-down value of −8.9 × 10−10 Hz s−1. In the upper plot, the excluded frequency bands for which no upper limits are provided have not been considered.

Standard image High-resolution image

Around the frequencies of greatest sensitivity, we are sensitive to objects as far away as 1.9 and 2.2 kpc for the first and second year of S5 and with an ellipticity ε around 10−4. Normal neutron stars are expected to have ε less than 10−5 [40, 41]. Such plausible value of ε could be detectable by a search like this if the object were emitting at 350 Hz and at a distance no further than 750 pc. For a source of fixed ellipticity and frequency, this search had a bit less range than the Einstein@Home search on the same data [18].

8. Applications of the χ2 veto and hardware-injected signals

A novel feature of the search presented here is the implementation of the χ2 veto. It is worth mentioning that this discriminator has been able to veto all the violin modes present in the data and many other narrow instrumental artifacts. Figure 12 demonstrates how well the χ2 veto used works on those frequency bands affected by violin modes.

Figure 12.

Figure 12. Significance and χ2 values obtained for all the elements in the top list for the second year of S5 data for the frequency bands 325–355 Hz and 685–699 Hz. Those two frequency bands include violin modes. Marked in dark red appear all the elements vetoed by the χ2 test. The solid line corresponds to the veto curve.

Standard image High-resolution image

As part of the testing and validation of search pipelines and analysis code, simulated signals are added into the interferometer length control system to produce mirror motions similar to what would be generated if a GW signal were present. These are the so-called hardware-injected pulsars. During the S5 run ten artificial pulsars were injected. Four of these pulsars: P2, P3, P5 and P8, at frequencies 575.16, 108.85, 52.81 and 193.4 Hz respectively, were strong enough to be detected by the multi-interferometer Hough search (see table III in [18] for the detailed parameters). The hardware injections were not active all the time, having a duty factor of about 60%.

The fact that these signals were not continuously present in the data caused the χ2 test to veto most of the templates associated with them, since they did not behave like the signals we were looking for. In particular, for the second year of S5, the elements of the top-list in frequency band containing P8 were vetoed by the χ2 test at the 99.4% level, and therefore that band was excluded from the analysis. The bands containing injected pulsars P2 and P3 were vetoed at the 87.7% and 94.5% level respectively, including the most significant events. Figure 13 shows the behavior of the χ2 veto for the 0.25 Hz band starting at 108.75 Hz that contains pulsar P3.

Figure 13.

Figure 13. Significance and χ2 values obtained for all the elements in the top list for the second year of S5 data for 0.25 Hz band starting at 108.75 Hz that contains a hardware injected simulated pulsar signal. Marked in dark red appear all the elements vetoed by the χ2 test. The solid line corresponds to the veto curve.

Standard image High-resolution image

In the frequency band 52.75–53.0 Hz, the candidates in the top-list were all produced by the 52.79 Hz instrumental artifact present in H1 and consequently the search failed to detect P5. This suggests that, in future analysis, smaller frequency intervals should be used to produce the top list of candidates, to prevent missing GW signals due to the presence of instrumental line-noise closeby.

9. Alternative strategies and future improvements

The search presented in this paper is more robust than but not as sensitive as the hierarchical all-sky search performed by the Einstein@Home distributed computing project on the same data [18], which for example, in the 0.5 Hz-wide band at 152.5 Hz, excluded the presence of signals with a h0 greater than 7.6 × 10−25 at a 90% confidence level. This later run used the Hough transform method to combine the information from coherent searches on a time scale of about a day and it was very computationally intensive. At the same time, the Einstein@Home search, due to its larger coherent baseline, is more sensitive to the fact that the second spin-down is not included in the search. The Hough transform method has also proven to be more robust against transient spectral disturbances than the StackSlide or PowerFlux semi-coherent methods [10].

Other strategies can be applied to perform all-sky multi-interferometer searches using the Hough transform operating on successive SFTs. In this section we estimate the sensitivity of the semi-coherent Hough search for two hypothetical searches to illustrate its capabilities, by either varying the duration of the total amount of data used in the multi-interferometer search, or by lowering the threshold for candidate selection.

In the first case we consider a search over the full S5 data with the same criteria of selecting candidates as presented here, i.e. setting a threshold in the significance for a false alarm level equivalent to one candidate per 0.25 Hz band, but using the full data. This first search would be more sensitive since we increase the number of SFTs to search over. In this case the sensitivity can be estimated from equation (28) and using the desired significance threshold as the 'loudest' event. In this case we should take into account that the number of templates for a two years search is double that for a single year, because of the increase of spin-down values that are resolvable. This corresponds to the dot-dashed line in the significance threshold in figure 6. Different sensitivity confidence levels can also be provided by modifying the false dismissal rate in equation (29) accordingly:

In the second case, we consider only the data from second year of S5 data but lower the threshold in the significance such that the false alarm would be ten candidates per 0.25 Hz band (see dashed line in figure 6).

In figure 14 we show the projected sensitivities for these two searches for different confidence levels. The best sensitivity would be for the search performed on the combined full two years of S5 data. For example in the frequency interval 159.75–160.0 Hz the estimated sensitivity levels are of 8.1 × 10−25, 7.9 × 10−25 and 7.1 × 10−25 at the 95%, 90% and 50% confidence level respectively. In the second case, corresponding to an increase in false alarm rate but with a reduced amount of data, the best sensitivities are of 8.8 × 10−25, 8.5 × 10−25 and 7.6 × 10−25 at the 95%, 90% and 50% confidence level respectively.

Figure 14.

Figure 14. Projected sensitivities at different confidence levels for (top) a combined search over the full S5 data using the same false alarm and (bottom) sensitivity of the second year of S5 but increasing the false alarm rate.

Standard image High-resolution image

Notice that although both strategies are, in principle, more sensitive than the search presented in this paper, they would produce many more candidates. These would need eliminating either by demanding coincidence between two searches of comparable sensitivity, or by follow-up using a more sensitive, computationally intensive, search. Coincidence analysis will be explored in future searches owing to the efficiency with which background noise is removed. Follow up studies will always be computationally limited. Therefore the follow up capacity will actually limit the event threshold for candidate selection. Without the inclusion of a coincidence analysis, the event threshold will have to be set higher, therefore compromising the potential sensitivity of the search itself. Moreover, for the first hypothetical case, in order to achieve the projected sensitivity, one would need to perform the search over the entire 67696 SFTs available at once. If one wanted to use the 'look up table' approach to compute the Hough transform over the two years of data, the computational cost would increase by a factor of nine with respect to the one year search presented in this paper, and the memory usage would increase from 0.8 to 7.2 GB, for the same sky-patch size. The memory needs could be reduced by decreasing the sky-patch size, but at additional computational cost. Another consequence of analyzing both years together is that the spin-down step size in the production search would have had to be reduced significantly.

There are a number of areas where further refinements could improve the sensitivity of the Hough search. In particular, one could decrease the grid spacing in parameter space in order to reduce the maximum mismatch allowed, increase the duration of the SFTs to increase the SNR within a single SFT, the development of further veto strategies to increase the overall efficiency of the analysis, as well as the tracking and establishing of appropriate data-cleaning strategies to remove narrow-band disturbances present in the peak-grams [39, 42, 43]. Several of these ideas are being addressed and will be implemented in the 'Frequency Hough all-sky search' using data from the Virgo second and fourth science runs to analyze data between 20 and 128 Hz.

10. Conclusions

In summary, we have reported the results of an all-sky search for continuous, nearly monochromatic gravitational waves on data from LIGO's fifth science (S5) run, using a new detection pipeline based on the Hough transform. The search covered the frequency range 50–1000 Hz and with the frequency's time derivative in the range −8.9 × 10−10 Hz s−1 to zero. Since no evidence for gravitational waves has been observed, we have derived upper limits on the intrinsic gravitational wave strain amplitude using a standard population-based method. The best upper limits correspond to 1.0 × 10−24 for the first year of S5 in the 158–158.25 Hz band, and 8.9 × 10−25 for the second year in the 146.5–146.75 Hz band (see figure 10).

This new search pipeline has allowed to process outliers down to significance from 5.10 at 50 Hz to 6.13 at 1000 Hz permitting deeper searchers than in previous Hough all-sky searches [10]. A set of new features have been included into the multi-detector Hough search code to be able to cope with large amounts of data and the memory limitations on the machines. In addition, a χ2 veto has been applied for the first time for continuous gravitational wave searches. This veto might be very useful for the analysis of the most recent set of data produced by the LIGO and Virgo interferometers (science runs S6, VSR2 and VSR4) whose data at lower frequencies are characterized by larger contamination of non-Gaussian noise than for S5.

Although the search presented here is not the most sensitive one on the same S5 data, this paper shows the potential of the new pipeline given the advantage of the lower computational cost of the Hough search and its robustness compared to other methods, and suggests further improvements to increase the sensitivity and overall efficiency of the analysis.

Acknowledgments

The authors gratefully acknowledge the support of the United States National Science Foundation for the construction and operation of the LIGO Laboratory, the Science and Technology Facilities Council of the United Kingdom, the Max-Planck-Society, and the State of Niedersachsen/Germany for support of the construction and operation of the GEO600 detector, and the Italian Istituto Nazionale di Fisica Nucleare and the French Centre National de la Recherche Scientifique for the construction and operation of the Virgo detector. The authors also gratefully acknowledge the support of the research by these agencies and by the Australian Research Council, the International Science Linkages program of the Commonwealth of Australia, the Council of Scientific and Industrial Research of India, the Istituto Nazionale di Fisica Nucleare of Italy, the Spanish Ministerio de Economía y Competitividad, the Conselleria d'Economia Hisenda i Innovació of the Govern de les Illes Balears, the Foundation for Fundamental Research on Matter supported by the Netherlands Organisation for Scientific Research, the Polish Ministry of Science and Higher Education, the FOCUS Programme of Foundation for Polish Science, the Royal Society, the Scottish Funding Council, the Scottish Universities Physics Alliance, The National Aeronautics and Space Administration, the Carnegie Trust, the Leverhulme Trust, the David and Lucile Packard Foundation, the Research Corporation, and the Alfred P Sloan Foundation. This document has been assigned LIGO Laboratory document number LIGO-P1300071.

Footnotes

Please wait… references are loading.
10.1088/0264-9381/31/8/085014