Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
License: arXiv.org perpetual non-exclusive license
arXiv:2404.05364v1 [astro-ph.HE] 08 Apr 2024
thanks: E-mail: kimsanginn@gmail.comthanks: E-mail: cyhui@cnu.ac.kr, huichungyue@gmail.com

Autoregressive Search of Gravitational Waves: Denoising

Sangin Kim11{}^{1}start_FLOATSUPERSCRIPT 1 end_FLOATSUPERSCRIPT    C. Y. Hui11{}^{1}start_FLOATSUPERSCRIPT 1 end_FLOATSUPERSCRIPT    Jianqi Yan22{}^{2}start_FLOATSUPERSCRIPT 2 end_FLOATSUPERSCRIPT    Alex P. Leung33{}^{3}start_FLOATSUPERSCRIPT 3 end_FLOATSUPERSCRIPT    Kwangmin Oh44{}^{4}start_FLOATSUPERSCRIPT 4 end_FLOATSUPERSCRIPT    A. K. H. Kong55{}^{5}start_FLOATSUPERSCRIPT 5 end_FLOATSUPERSCRIPT    L.C.-C. Lin66{}^{6}start_FLOATSUPERSCRIPT 6 end_FLOATSUPERSCRIPT    Kwan-Lok Li66{}^{6}start_FLOATSUPERSCRIPT 6 end_FLOATSUPERSCRIPT 11{}^{1}start_FLOATSUPERSCRIPT 1 end_FLOATSUPERSCRIPTDepartment of Astronomy and Space Science, Chungnam National University, 9 Daehak-ro, Yuseong-gu, Daejeon 34134, Republic of Korea 22{}^{2}start_FLOATSUPERSCRIPT 2 end_FLOATSUPERSCRIPTFaculty of Innovation Engineering, Macau University of Science and Technology, Avenida Wai Long, Taipa 999078, Macau 33{}^{3}start_FLOATSUPERSCRIPT 3 end_FLOATSUPERSCRIPTDepartment of Physics, The University of Hong Kong, 999077, Hong Kong 44{}^{4}start_FLOATSUPERSCRIPT 4 end_FLOATSUPERSCRIPTDepartment of Physics and Astronomy, Michigan State University, East Lansing, MI 48824, USA 55{}^{5}start_FLOATSUPERSCRIPT 5 end_FLOATSUPERSCRIPTInstitute of Astronomy, National Tsing Hua University, Hsinchu 30013, Taiwan 66{}^{6}start_FLOATSUPERSCRIPT 6 end_FLOATSUPERSCRIPTDepartment of Physics, National Cheng Kung University, No.1, University Road, Tainan City 701, Taiwan
(April 8, 2024)
Abstract

Because of the small strain amplitudes of gravitational-wave (GW) signals, unveiling them in the presence of detector/environmental noise is challenging. For visualizing the signals and extracting its waveform for a comparison with theoretical prediction, a frequency-domain whitening process is commonly adopted for filtering the data. In this work, we propose an alternative template-free framework based on autoregressive modeling for denoising the GW data and extracting the waveform. We have tested our framework on extracting the injected signals from the simulated data as well as a series of known compact binary coalescence (CBC) events from the LIGO data. Comparing with the conventional whitening procedure, our methodology generally yields improved cross-correlation and reduced root mean square errors with respect to the signal model.

preprint: APS/123-QED

I Introduction

The existence of gravitational wave (GW) is one of the most remarkable predictions of general relativity (GR)[1, 2]. GW is a tidal acceleration that propagates in spacetime at the speed of light. According to the Einstein field equation, it requires stress at an order of c2/8πG1043similar-tosuperscript𝑐28𝜋𝐺superscript1043c^{2}/8\pi G\sim 10^{43}italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 8 italic_π italic_G ∼ 10 start_POSTSUPERSCRIPT 43 end_POSTSUPERSCRIPT N m22{}^{-2}start_FLOATSUPERSCRIPT - 2 end_FLOATSUPERSCRIPT to produce a unit of curvature. Therefore, the amplitude of GW is expected to be very small. And it requires catastrophic phenomena involving compact objects to produce such tiny ripples in the spacetime (e.g. binary black hole mergers).

It is the small amplitude of GW that makes the detection challenging. The first compelling evidence for the existence of GW came indirectly from the long-term pulsar timing of the Hulse-Taylor binary PSR B1913+16 [3]. The behavior of this binary (e.g. decay of orbital period) is fully consistent with the prediction by GR as the system loses its orbital energy in GW.

Thanks to the improved sensitivity, on 14 September 2015, the advanced Laser Interferometer Gravitational-wave Observatory (LIGO) has directly detected a GW event, GW150914, from a binary black hole (BBH) coalescence for the first time [4]. This has opened the possibility of exploring our Universe without limiting to the window of electromagnetic radiation. Two years later, the era of multi-messenger astronomy was highlighted by the discovery of GW event GW170817 resulted from the merger of two neutron stars (NSs) [5, 6, 7], which was found to be associated with the γlimit-from𝛾\gamma-italic_γ -ray burst GRB 170817A [8]. This marks the first case that both GW and electromagnetic radiation were detected from the same astrophysical object.

Currently, in the Gravitational Wave Transient Catalog (GWTC) maintained by LIGO/Virgo/KAGRA collaboration 111https://gwosc.org/eventapi/html/allevents/ [10, 11, 12, 13], there are 93 GW transient events so far have been confidently detected (i.e. probability of origin from an astrophysical source pastro>0.5subscript𝑝astro0.5p_{\rm astro}>0.5italic_p start_POSTSUBSCRIPT roman_astro end_POSTSUBSCRIPT > 0.5). These include 89 from BBH coalescence, 2 from NS-NS mergers, and 2 from BH-NS mergers. Apart from these confident events, there are >20absent20>20> 20 marginal candidates.

For further advancing GW astronomy, while enhancing the instrumental sensitivity is vitally important [e.g. 14], progress can also be achieved by improving the methodology of data processing and analysis. Currently, the standard search method for CBC events in the GW community is matched filtering [cf. 15], which is done by cross-correlating a template of known waveform and the interferometer output at different time delays to produce a filtered output. With the signal-to-noise ratio (SNR) as the ratio of the value of the filtered output to the corresponding value root mean square value for the noise, it can be proved that a matched filter comprises the ratio of the template of the actual waveform to the spectral noise density of the interferometer can optimize the SNR under several assumptions [16].

Although the technique of matched filtering has unveiled a considerable population of GW events as aforementioned, it has a number of limitations. For the matched filter to have optimal performance, the data have to fulfill the assumptions of wide-sense stationarity (WSS) and zero-means, which are generally not satisfied in the raw interferometric data. And most importantly, the construction of matched filter requires knowledge of waveform for the expected signal. However, the forms of GW signal from many possible sources are poorly modeled (e.g. highly eccentric BH binaries) or even unknown (e.g. fast radio bursts). In such cases, matched filtering technique cannot be employed. Even for the cases that the waveform can be determined such as circular BH binaries, this technique still requires the construction of a large template bank to cover a sufficiently large parameter space. A search over this extensive template bank by brute-force is computationally expensive.

Furthermore, even though the technique of matched filtering is capable to detect the GW signals from CBC events with known waveform, it does not enable one to visualize the signal directly. For visualizing the GW signal from the data and extracting its waveform for a comparison with the prediction by numerical relativity [e.g. Fig. 1 in 4], one must filter the raw time series with a bandpass filter for removing the data out of the detectors’ most sensitive frequency band as well as apply the frequency-domain whitening process for suppressing the colored noises at low frequencies and the spectral lines resulted from instrumental/background effects. Frequency-domain whitening is a procedure to equalize the spectrum through dividing the Fourier coefficients by the estimate of the amplitude spectral density of the noise. While this is considered as a standard procedure and has been adopted for noise suppression in many works [e.g. 17, 18], it is still important to explore alternative techniques for de-noising and compare their performance with that of conventional whitening filter. For example, Tsukada et al. have proposed a time-domain whitening filter for optimizing latency in the CBC data analysis pipeline [19].

In this paper, we explore the feasibility of a template-free method based on autoregressive modeling in filtering GW data. In Section II, we provide an overview of the methodology. In Section III, we will demonstrate the feasibility of our framework by a series of experiments. And we will summarize our results and provide an outlook for further development in Section IV.

II Methodology

II.1 Autoregression

Time series data that we acquire in nature can be affected by various random processes and exhibit stochastic behaviors. Due to the high sensitivity of GW detectors, the raw data are typically corrupted with the various kinds of noise [e.g. 20], in which the assumptions for the matched filter to attain the optimal performance such as stationarity are generally not fulfilled. In our proposed framework, we adopted an autoregressive (AR) approach in developing an efficient time-domain noise filtering scheme without any a priori knowledge on the noise.

In a recent astronomical application of AR modeling, Caceres et al. have developed a methodology of the autoregressive planet search (ARPS) for treating a wide variety of stochastic processes so as to improve the search of transit signals by exoplanets in the residuals after noise reduction [21]. In exoplanet search, people are looking for small dips in the light curve resulting from a transit submerged by the much larger brightness variability of the parent stars. And the aperiodic colored noise in the photometric data is notoriously difficult to treat [22]. With a procedure based on AR, Caceres et al. have demonstrated the stellar variability can be identified and removed.

We notice that the aforementioned challenge is shared by the GW astronomy, namely searching for the small strain amplitude of GW signals in the presence of instrumental/environmental noise with amplitude orders of magnitude larger. This comparison has motivated us to explore whether the AR technique can also be applied in extracting GW signals.

AR modeling can be applied to any dynamical system whose status in the present time has a dependence on its past status (i.e. autocorrelated behavior). The simplest model AR(p𝑝pitalic_p) can be built by regression with the estimate at time t𝑡titalic_t, x^tsubscript^𝑥𝑡\hat{x}_{t}over^ start_ARG italic_x end_ARG start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT, being modeled by the linear combination of past values xtisubscript𝑥𝑡𝑖x_{t-i}italic_x start_POSTSUBSCRIPT italic_t - italic_i end_POSTSUBSCRIPT plus a random noise term:

x^t=i=1paixti+ϵtsubscript^𝑥𝑡subscriptsuperscript𝑝𝑖1subscript𝑎𝑖subscript𝑥𝑡𝑖subscriptitalic-ϵ𝑡\hat{x}_{t}=\sum^{p}_{i=1}a_{i}x_{t-i}+\epsilon_{t}over^ start_ARG italic_x end_ARG start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = ∑ start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_t - italic_i end_POSTSUBSCRIPT + italic_ϵ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT (1)

where p𝑝pitalic_p is the order of AR model (i.e. the number of lags in the model), aisubscript𝑎𝑖a_{i}italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are the model parameters, xtisubscript𝑥𝑡𝑖x_{t-i}italic_x start_POSTSUBSCRIPT italic_t - italic_i end_POSTSUBSCRIPT is the i𝑖iitalic_i-th past data, and ϵtsubscriptitalic-ϵ𝑡\epsilon_{t}italic_ϵ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT is the noise term distributed as a Gaussian with zero mean and unknown variance.

In the application of ARPS, the best-fit AR model can be treated as a good estimator of stochastic noise. By subtracting the model from the raw data, the residuals can be obtained as follow:

xr=xx^subscript𝑥𝑟𝑥^𝑥x_{r}=x-\hat{x}italic_x start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = italic_x - over^ start_ARG italic_x end_ARG (2)

where x^^𝑥\hat{x}over^ start_ARG italic_x end_ARG is the best-fit AR model on data x𝑥xitalic_x. From xrsubscript𝑥𝑟x_{r}italic_x start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT, we can investigate whether there is any astrophysically-interesting signals can be extracted [see 21, 23].

II.2 Autoregressive integrated moving-average model (ARIMA)

However, for an AR model to provide a legitimate description of the data, the time series is assumed to be stationary. For the time series with systematic trends, such data cannot be treated by the AR model. For converting a non-stationary time series into a stationary one, the differencing operation is found to be efficient (e.g. xt=xtxt1superscriptsubscript𝑥𝑡subscript𝑥𝑡subscript𝑥𝑡1x_{t}^{{}^{\prime}}=x_{t}-x_{t-1}italic_x start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT = italic_x start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT - italic_x start_POSTSUBSCRIPT italic_t - 1 end_POSTSUBSCRIPT where xtsuperscriptsubscript𝑥𝑡x_{t}^{{}^{\prime}}italic_x start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT end_POSTSUPERSCRIPT is the differenced series obtained from the change between consecutive values in the original time series). With the backshift operator B𝐵Bitalic_B defined as Bxt=xt1𝐵subscript𝑥𝑡subscript𝑥𝑡1Bx_{t}=x_{t-1}italic_B italic_x start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = italic_x start_POSTSUBSCRIPT italic_t - 1 end_POSTSUBSCRIPT, the aforementioned process can be described as (1B)xt1𝐵subscript𝑥𝑡(1-B)x_{t}( 1 - italic_B ) italic_x start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT. This is known as first-order differencing. To generalize the process to a higher orders, the operation can be modified as

xd=(1B)dxtsubscript𝑥𝑑superscript1𝐵𝑑subscript𝑥𝑡x_{d}=(1-B)^{d}x_{t}italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = ( 1 - italic_B ) start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT (3)

which is commonly referred to as Integrated process (I) of dthsuperscript𝑑thd^{\rm th}italic_d start_POSTSUPERSCRIPT roman_th end_POSTSUPERSCRIPT order. The output xdsubscript𝑥𝑑x_{d}italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT can be modeled as a stationary time series.

While AR model uses past values in a time series to predict the current value, a Moving Average model of order q𝑞qitalic_q, MA(q𝑞qitalic_q), predicts the current value by a linear combination of past error terms

x^t=j=1qbiϵtj+ϵtsubscript^𝑥𝑡subscriptsuperscript𝑞𝑗1subscript𝑏𝑖subscriptitalic-ϵ𝑡𝑗subscriptitalic-ϵ𝑡\hat{x}_{t}=\sum^{q}_{j=1}b_{i}\epsilon_{t-j}+\epsilon_{t}over^ start_ARG italic_x end_ARG start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = ∑ start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_ϵ start_POSTSUBSCRIPT italic_t - italic_j end_POSTSUBSCRIPT + italic_ϵ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT (4)

where ϵtjsubscriptitalic-ϵ𝑡𝑗\epsilon_{t-j}italic_ϵ start_POSTSUBSCRIPT italic_t - italic_j end_POSTSUBSCRIPT is the error term for the jlimit-from𝑗j-italic_j -th time step in the time series xtsubscript𝑥𝑡x_{t}italic_x start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT and bisubscript𝑏𝑖b_{i}italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are the model parameter.

Combining Equation 1, Equation 3 and, Equation 4, an AutoRegressive Integrated Moving-Average model (ARIMA) can be constructed as:

(1B)dxt=i=1paixti+j=1qbiϵtj+ϵtsuperscript1𝐵𝑑subscript𝑥𝑡subscriptsuperscript𝑝𝑖1subscript𝑎𝑖subscript𝑥𝑡𝑖subscriptsuperscript𝑞𝑗1subscript𝑏𝑖subscriptitalic-ϵ𝑡𝑗subscriptitalic-ϵ𝑡(1-B)^{d}x_{t}=\sum^{p}_{i=1}a_{i}x_{t-i}+\sum^{q}_{j=1}b_{i}\epsilon_{t-j}+% \epsilon_{t}( 1 - italic_B ) start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = ∑ start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_t - italic_i end_POSTSUBSCRIPT + ∑ start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_ϵ start_POSTSUBSCRIPT italic_t - italic_j end_POSTSUBSCRIPT + italic_ϵ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT (5)

with aisubscript𝑎𝑖a_{i}italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and bisubscript𝑏𝑖b_{i}italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT estimated simultaneously for the whole time series by any optimization method (e.g. maximum likelihood). On the other hand, the orders of the model (e.g. p𝑝pitalic_p, q𝑞qitalic_q, d𝑑ditalic_d) can be determined by the procedures of model selection.

By applying ARIMA modeling to the light curves of 156717 stars as observed by NASA’s Kepler satellite, Caceres et al. shows that the brightness fluctuations of the parent stars can be effectively reduced. Subsequent searches of transit-like signals from the ARIMA residuals resulted in a recovery of a significant fraction of confirmed exoplanets [23].

We started our experiment by testing whether the existing code auto.arima from the R package forecast222https://www.r-project.org is capable to fit the ARIMA model to GW data with the orders and the parameters of the model estimated automatically. auto.arima was used in constructing ARPS pipeline [23].

We have tested whether auto.arima can extract the waveform of GW150914 from LIGO data obtained by both detectors in Hanford (H) and Livingston (L) with a sampling frequency of 4 kHz. The data were downloaded from the event portal managed by the GW Open Science Center (GWOSC)333https://www.gw-openscience.org/eventapi/html/allevents/. However, since the strain amplitude is at an order of h1018less-than-or-similar-tosuperscript1018h\lesssim 10^{-18}italic_h ≲ 10 start_POSTSUPERSCRIPT - 18 end_POSTSUPERSCRIPT, the modeling apparently suffered from an underflow problem.

In order to circumvent the underflow, we have normalized the data to the order of unity. In this case, auto.arima yields the orders of (p,q,d)=(4,0,1). By subtracting the resultant model from the original data, we have obtained the residuals. However, even with the aid of a low-pass filter, we are unable to identify any waveform which resembles that of GW150914 from the residuals. In view of such an undesirable behavior, instead of employing auto.arima as in [21, 23], we are going to develop our own algorithm which is more suitable for reducing noise in GW data.

Refer to caption
Figure 1: The structure of our proposed framework seqARIMA.
Refer to caption
Figure 2: Illustrations for the effects of each stage in seqARIMA by using the LIGO-H data of GW150914.

In our proposed framework, we break the noise reduction process into a sequence of procedures as shown in Figure 1. Hereafter we refer it as sequential ARIMA model (seqARIMA), which consists of four stages: integrated process, autoregressive process, moving-average process, and bandpass filtering.

In this section, we take the LIGO data of GW150914 (Hanford, 32 s, 4 kHz sampling) to demonstrate the performance of seqARIMA. The effects of each stage in our procedure are illustrated in Figure 2 and described in the following subsections (i.e. subsection II.3-subsection II.6).

II.3 Integrated process

As a first step of our proposed framework, integrated process plays an essential role for ensuring the stationarity of a given GW time series.

For demonstrating the procedure, we start with the raw LIGO-H data of GW150914 (top panel in Figure 2) and with the parameter corresponding to the order of differencing initialized as d=0. We employ the Kwiatkowski–Phillips–Schmidt–Shin (KPSS) test [26], which is a standard test for stationarity [27], to examine whether the raw data exhibits trend and non-zero mean level. Figure 2 clearly shows that the raw data is non-stationary.

Instead of ensuring global stationarity on the entire input time series, we consider local or segmented stationarity within time windows which are comparable to the signal duration from the typical BBH coalescence. A segment within a time series refers to a sequence of data points collected or recorded at a given time interval. Most time series segmentation algorithms can be classified into three primary categories: sliding windows, top-down, and bottom-up approaches [28]. Lovrić et al. have demonstrated that the process of segmenting time series into a limited number of homogeneous segments aids in the extraction of time segments with similar observations [29]. These techniques process the input time series and return a piecewise linear representation (PLR). Our method, however, divides time series into non-overlapping times series of equal length for the best performance. Since the total length of the time series is much longer (32 s) than the duration of the typical BBH coalescence signal (i.e. less-than-or-similar-to\lesssim 0.5 seconds), the whole time series is divided into 64 segments and the KPSS test is applied on each segment with the length 0.5 s (i.e. ts=0.5subscript𝑡s0.5t_{\rm s}=0.5italic_t start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT = 0.5). If the p𝑝pitalic_p-values from KPSS test on all those segments are greater than or equal to our predefined threshold (p𝑝pitalic_p-value = 0.1), the given data is determined as a stationary time series. The threshold is chosen to be larger than the conventional value of p𝑝pitalic_p-value = 0.05 so as to reduce the false negatives.

Otherwise, if there is any segment exhibits non-stationary behavior, we modify the parameter d𝑑ditalic_d as d+1𝑑1d+1italic_d + 1 and apply differencing by Equation 3. Such process will be iterated until xdsubscript𝑥𝑑x_{d}italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT satisfies the stationary condition by passing the KPSS test (See algorithm 1). In the second panel of Figure 2, we show an optimal differencing model for our test data with d=2𝑑2d=2italic_d = 2.

Since the result of hypothesis test can be influenced by the volume of data used, we further test the robustness of algorithm 1 by running KPSS tess on different tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. In the aforementioned experiment, we took ts=0.5subscript𝑡𝑠0.5t_{s}=0.5italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 0.5 s which gives 2000similar-toabsent2000\sim 2000∼ 2000 data points for a sampling rate of 4 kHz. For investigating the possible impact of non-stationarity detection by the length of data segment, we have re-run algorithm 1 on the same data by varying tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT from 0.25 s to 0.75 s. And we found that all cases yield the same optimal differencing model with d=2𝑑2d=2italic_d = 2. In view of this, we conclude that the results from KPSS test and hence algorithm 1 is robust and our adopted segment length of ts=0.5subscript𝑡𝑠0.5t_{s}=0.5italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 0.5 s is sufficient.

Input : xtsubscript𝑥𝑡x_{t}italic_x start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT, the input time series,
ttsubscript𝑡tt_{\rm t}italic_t start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT, the time length of xtsubscript𝑥𝑡x_{t}italic_x start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT,
ts=0.5subscript𝑡s0.5t_{\rm s}=0.5italic_t start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT = 0.5, the length of each segment,
cp=0.1subscript𝑐𝑝0.1c_{p}=0.1italic_c start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = 0.1, the threshold for the acceptance of p-values
Output : xdsubscript𝑥𝑑x_{d}italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT
Initialization : ns=tt/tssubscript𝑛ssubscript𝑡tsubscript𝑡sn_{\rm s}=\lceil t_{\rm t}/t_{\rm s}\rceilitalic_n start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT = ⌈ italic_t start_POSTSUBSCRIPT roman_t end_POSTSUBSCRIPT / italic_t start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT ⌉, the number of segments,
𝐒={s1,,sns}𝐒subscript𝑠1subscript𝑠subscript𝑛s\mathbf{S}=\{s_{1},...,s_{n_{\rm s}}\}bold_S = { italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_s start_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT end_POSTSUBSCRIPT }, the consecutive segments
1 d=0𝑑0d=0italic_d = 0
2 while true do
3       xd=(1B)dxtsubscript𝑥𝑑superscript1𝐵𝑑subscript𝑥𝑡x_{d}=(1-B)^{d}x_{t}italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = ( 1 - italic_B ) start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT for i=1𝑖1i=1italic_i = 1 to nssubscript𝑛normal-sn_{\rm s}italic_n start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT do
4             Perform KPSS test on sisubscript𝑠𝑖s_{i}italic_s start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT
5             Compute the p𝑝pitalic_p-values for the level stationarity pLsubscript𝑝𝐿p_{L}italic_p start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT-value and the trend stationarity pTsubscript𝑝𝑇p_{T}italic_p start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT-value with sisubscript𝑠𝑖s_{i}italic_s start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT
6            
7      if pLcpsubscript𝑝𝐿subscript𝑐𝑝p_{L}\geq c_{p}italic_p start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ≥ italic_c start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT and pTcpsubscript𝑝𝑇subscript𝑐𝑝p_{T}\geq c_{p}italic_p start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT ≥ italic_c start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT si{1,2,,ns}for-allsubscript𝑠𝑖12normal-…subscript𝑛𝑠\forall s_{i\in\{1,2,...,n_{s}\}}∀ italic_s start_POSTSUBSCRIPT italic_i ∈ { 1 , 2 , … , italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT } end_POSTSUBSCRIPT then
8             return xdsubscript𝑥𝑑x_{d}italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT
9            
10      else
11             d=d+1𝑑𝑑1d=d+1italic_d = italic_d + 1
12            
13      
Algorithm 1 Integrated process (Sec. II C)

II.4 Autoregressive process

Once the stationary data set is obtained as an output xdsubscript𝑥𝑑x_{d}italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT of the integrated process, we build the AR model of xdsubscript𝑥𝑑x_{d}italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT. A set of AR models can be produced by:

x^d,p=i=1paixd,ti+ϵtsubscript^𝑥𝑑𝑝subscriptsuperscript𝑝𝑖1subscript𝑎𝑖subscript𝑥𝑑𝑡𝑖subscriptitalic-ϵ𝑡\hat{x}_{d,p}=\sum^{p}_{i=1}a_{i}x_{d,t-i}+\epsilon_{t}over^ start_ARG italic_x end_ARG start_POSTSUBSCRIPT italic_d , italic_p end_POSTSUBSCRIPT = ∑ start_POSTSUPERSCRIPT italic_p end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_d , italic_t - italic_i end_POSTSUBSCRIPT + italic_ϵ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT (6)

where p{1,,pmax}𝑝1subscript𝑝maxp\in\{1,...,p_{\rm max}\}italic_p ∈ { 1 , … , italic_p start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT }.

For constructing a set of candidate models, we need to fix the upper-bound of p𝑝pitalic_p which is set by the hyper-parameter pmaxsubscript𝑝maxp_{\rm max}italic_p start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT. The optimal AR model is determined by model selection based on Akaike Information Criteria [30, AIC], which is defined as AIC=2p+nln(σp^2)𝐴𝐼𝐶2𝑝𝑛superscript^subscript𝜎𝑝2AIC=2p+n\ln(\hat{\sigma_{p}}^{2})italic_A italic_I italic_C = 2 italic_p + italic_n roman_ln ( over^ start_ARG italic_σ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) where n𝑛nitalic_n is the sample size and σp^2superscript^subscript𝜎𝑝2\hat{\sigma_{p}}^{2}over^ start_ARG italic_σ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is the maximum likelihood estimator for the variance of the noise term. For each model in the set, AIC is calculated. And the model that attains the optimal AIC will be selected.

We have examined the effect with different values of pmaxsubscript𝑝maxp_{\rm max}italic_p start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT by subtracting the corresponding optimal AR model from the data (cf. Equation 2). The results are shown in Figure 4 which show that pmax=8192subscript𝑝max8192p_{\rm max}=8192italic_p start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 8192 (corresponding to similar-to\sim 2 seconds for 4 kHz sampling frequency) can lead to recognizable waveform in the residuals. We have examined whether the result can be further improved by setting pmaxsubscript𝑝maxp_{\rm max}italic_p start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT at higher values. However, among all the experiments presented in this work, we found that the optimal p𝑝pitalic_p selected by AIC all converged below 8192819281928192 even with pmaxsubscript𝑝maxp_{\rm max}italic_p start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT set at higher values. In view of this, we fixed pmax=8192subscript𝑝max8192p_{\rm max}=8192italic_p start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 8192 as the hyper-parameter throughout this work for an efficient computation.

In our framework, the model parameters (i.e. AR coefficients {ai}subscript𝑎𝑖\{a_{i}\}{ italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT }) are estimated by Burg method, which fits the model to xdsubscript𝑥𝑑x_{d}italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT for minimizing the sums of squares of forward and backward linear prediction errors [31, 32]. The function ar.burg from the R package stats is adopted in our experiment.

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Refer to caption
(d)
Refer to caption
(e)
Refer to caption
(f)
Figure 3: (upper panels) PSDs of the pure noise data, AR residuals and the whitened data. (lower panels) Q-Q plots for comparing the distributions of the raw/processed data (red lines) and with a Gaussian distribution (black lines).

Since AR is the major component of our procedure, before we apply it to the data with a CBC signal embeded, we have first investigated its performance on the pure noise data and examined whether the processed data can satisfy the requirements of stationarity and normality [cf. 33]. For this test, we have used both simulated and real noise data. We started by generating 100 simulated noise data of 32 s from sampling the updated Advanced LIGO sensitivity design curve 444https://dcc.ligo.org/LIGO-T1800044/public. And we have processed them with algorithm 1 and algorithm 2. For comparison, we have also separately processed the simulated noise with the standard whitening. To quantify the difference between the distribution of the data from normality, we have run the Anderson-Darling (A-D) test [35]. Taking the plimit-from𝑝p-italic_p -value of 0.05 as the benchmark for rejecting the null hypothesis, all the simulated data fail to pass the A-D test which yield a mean plimit-from𝑝p-italic_p -value of 1016similar-toabsentsuperscript1016\sim 10^{-16}∼ 10 start_POSTSUPERSCRIPT - 16 end_POSTSUPERSCRIPT. This suggests they are all significantly different from a Gaussian distribution. After subtracting the noise data from the AR models, 70%similar-toabsentpercent70\sim 70\%∼ 70 % of these samples become conform with normality (yield a plimit-from𝑝p-italic_p -value >0.05absent0.05>0.05> 0.05). And we found that whitening results in a similar fraction that pass A-D test.

On the other hand, all these simulated noise data are found to pass KPSS test and do not demonstrate any non-stationarity. In order to search for the non-stationary noise data for the experiment, we have searched over the LIGO data. We have chosen 24 time segments which do not encompass any confirmed GW events and yield a plimit-from𝑝p-italic_p -value smaller than 0.05 in the KPSS test. Also all these segments do not conform with normality which yield a mean plimit-from𝑝p-italic_p -value of 1014similar-toabsentsuperscript1014\sim 10^{-14}∼ 10 start_POSTSUPERSCRIPT - 14 end_POSTSUPERSCRIPT in the A-D test. After whitening (or processing with our method), all 24 processed pure noise data pass the KPSS test (all yield plimit-from𝑝p-italic_p -value >0.1absent0.1>0.1> 0.1). Also, both methods result in similar fraction (65%similar-toabsentpercent65\sim 65\%∼ 65 %) for passing the normality test. Therefore, we conclude that both our method and whitening have a comparable performance on the pure noise in attaining normality and stationarity. As an example, we compare the power spectral density (PSD) for one of our real noise sample with those of AR residuals and whitened data in Figure 3. These plots also demonstrate the capability of a AR model in line removal. In the low panels, we have also constructed the quantile-quantile (Q-Q) plots for comparing the distribution of the data with the normal distribution. They clearly show that both AR residuals and the whitened data distribute as a Gaussian.

For our test data which encompasses the transient signal from GW150914, an optimal AR order of p=7931𝑝7931p=7931italic_p = 7931 is obtained. Before passing to the next stage of processing, we obtain the residual time series xrsubscript𝑥𝑟x_{r}italic_x start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT by subtracting the optimal model from xdsubscript𝑥𝑑x_{d}italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT (cf. Equation 2, algorithm 2). The waveform of AR residual data is shown in the third panel in Figure 2, in which the modulation resulting from the BBH coalescence starts emerging. To further suppress the random fluctuation in xrsubscript𝑥𝑟x_{r}italic_x start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT, we proceed to the next stage (see below).

Input : xdsubscript𝑥𝑑x_{d}italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT, the input time series,
pmax=8192subscript𝑝max8192p_{\rm max}=8192italic_p start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 8192: the maximum number of the order of the autoregressive process
Output : xrsubscript𝑥𝑟x_{r}italic_x start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT
1 Estimate the parameters ϕpsubscriptitalic-ϕ𝑝\phi_{p}italic_ϕ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT of an AR(p𝑝pitalic_p) model on xdsubscript𝑥𝑑x_{d}italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT with the variance σp2subscriptsuperscript𝜎2𝑝\sigma^{2}_{p}italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT calculated using Burg’s method and
2 {ϕp,popt}=argminϕp,poptAICsubscriptitalic-ϕ𝑝subscript𝑝optsubscriptsubscriptitalic-ϕ𝑝subscript𝑝opt𝐴𝐼𝐶\{\phi_{p},p_{\rm opt}\}=\mathop{\arg\min}\limits_{\phi_{p},p_{\rm opt}}{AIC}{ italic_ϕ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT , italic_p start_POSTSUBSCRIPT roman_opt end_POSTSUBSCRIPT } = start_BIGOP roman_arg roman_min end_BIGOP start_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT , italic_p start_POSTSUBSCRIPT roman_opt end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_A italic_I italic_C for p{1,,pmax}𝑝1subscript𝑝maxp\in\{1,...,p_{\rm max}\}italic_p ∈ { 1 , … , italic_p start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT }
3 with AIC=2p+nln(σp^2)𝐴𝐼𝐶2𝑝𝑛superscript^subscript𝜎𝑝2AIC=2p+n\ln(\hat{\sigma_{p}}^{2})italic_A italic_I italic_C = 2 italic_p + italic_n roman_ln ( over^ start_ARG italic_σ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT )
4 Obtain time series x^poptsubscript^𝑥subscript𝑝opt\hat{x}_{p_{\rm opt}}over^ start_ARG italic_x end_ARG start_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT roman_opt end_POSTSUBSCRIPT end_POSTSUBSCRIPT from the optimal model {ϕp,popt}subscriptitalic-ϕ𝑝subscript𝑝opt\{\phi_{p},p_{\rm opt}\}{ italic_ϕ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT , italic_p start_POSTSUBSCRIPT roman_opt end_POSTSUBSCRIPT }
5 return the residual xr=xdx^poptsubscript𝑥𝑟subscript𝑥𝑑subscript^𝑥subscript𝑝optx_{r}=x_{d}-\hat{x}_{p_{\rm opt}}italic_x start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT - over^ start_ARG italic_x end_ARG start_POSTSUBSCRIPT italic_p start_POSTSUBSCRIPT roman_opt end_POSTSUBSCRIPT end_POSTSUBSCRIPT
Algorithm 2 Autoregressive process (Sec. II D)
Refer to caption
Figure 4: Effects of AR with different pmaxsubscript𝑝maxp_{\rm max}italic_p start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT on the difference data of d=2𝑑2d=2italic_d = 2. These time series are the residuals obtained by subtracting the fitted AR model from the data (i.e. xrsubscript𝑥𝑟x_{r}italic_x start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT outputted by algorithm 2).

II.5 Moving-average process

Different from the conventional ARIMA model (Equation 5) in which MA performs the regression with the past forecast errors ϵisubscriptitalic-ϵ𝑖\epsilon_{i}italic_ϵ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. In our framework, MA refers to the method of estimating the trend in the residuals which is taken as a form of low-pass finite impulse response filter. The process is expressed as follows:

xma=1qj=kkxt+jsubscript𝑥ma1𝑞superscriptsubscript𝑗𝑘𝑘subscript𝑥𝑡𝑗x_{\rm ma}=\frac{1}{q}\sum_{j=-k}^{k}x_{t+j}italic_x start_POSTSUBSCRIPT roman_ma end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_q end_ARG ∑ start_POSTSUBSCRIPT italic_j = - italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT italic_t + italic_j end_POSTSUBSCRIPT (7)

where q𝑞qitalic_q is the order of MA and k=(q1)/2𝑘𝑞12k=(q-1)/2italic_k = ( italic_q - 1 ) / 2. Since we consider a two-sided (centered) MA, if an even value of order q𝑞qitalic_q is specified, two MAs of k𝑘kitalic_k rounded-down and rounded-up will be averaged [cf. 36].

For choosing the value of q𝑞qitalic_q, a model with small q𝑞qitalic_q might have the signal remain buried by the random fluctuations in xrsubscript𝑥𝑟x_{r}italic_x start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT. On the other hand, a large q𝑞qitalic_q can smear out the signal. Since the GW from the BBH coalescence has its frequency varying, a MA model with a fixed q𝑞qitalic_q can suffer from the aforementioned trade-off. Another problem of a single MA model is found when large values of q𝑞qitalic_q are adopted. In the Figure 5, we show the power spectral density (PSD) of the output from the MA model with different q𝑞qitalic_q annotated as blue lines in each panel. Within our concerned frequency band (32--512 Hz), We found that power spectral leakage starts appearing with q7𝑞7q\geq 7italic_q ≥ 7, which can be possibly resulted from over-smoothing.

To overcome the aforementioned problems, rather than using only a single MA(q𝑞qitalic_q), we adopted the method of Ensemble of Averages (EoA) [cf. 37] which combines MAs from a range of q𝑞qitalic_q (q{q\in\{italic_q ∈ {1…qmax}q_{\rm max}\}italic_q start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT }). It aggregates a number of MAs with an ensemble of moving averages. In our work, we utilize EoA and demonstrate that using median as the collector function can be a very effective filter (algorithm 3).

Figure 6 shows the outputs of EoA for different choices of qmaxsubscript𝑞maxq_{\rm max}italic_q start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT. Empirically, we found that qmax=20subscript𝑞max20q_{\rm max}=20italic_q start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 20 with a median collector function gives a desirable result. It can eliminate most random fluctuations and retains the signal fidelity as all three stages of coalescence (i.e., inspiral, merger, and ringdown) can be clearly visualized. Furthermore, with EoA, the problem of spectral power leakage in PSD is resolved (annotated as red lines in Figure 5).

Input : xrsubscript𝑥𝑟x_{r}italic_x start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT,
q{qmin,,qmax}𝑞subscript𝑞minsubscript𝑞maxq\in\{q_{\rm min},...,q_{\rm max}\}italic_q ∈ { italic_q start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT , … , italic_q start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT }, the order of the moving average
where qmin=1subscript𝑞min1q_{\rm min}=1italic_q start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT = 1 and qmax=20subscript𝑞max20q_{\rm max}=20italic_q start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 20 by default
Output : xEoAsubscript𝑥EoAx_{\rm EoA}italic_x start_POSTSUBSCRIPT roman_EoA end_POSTSUBSCRIPT
1 for q=qmin𝑞subscript𝑞normal-minq=q_{\rm min}italic_q = italic_q start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT to qmaxsubscript𝑞normal-maxq_{\rm max}italic_q start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT do
2       xq=1qj=kkxr+jsubscript𝑥𝑞1𝑞subscriptsuperscript𝑘𝑗𝑘subscript𝑥𝑟𝑗x_{q}=\frac{1}{q}\sum^{k}_{j=-k}x_{r+j}italic_x start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_q end_ARG ∑ start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j = - italic_k end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_r + italic_j end_POSTSUBSCRIPT
3x^=median({xqmin,,xqmax})^𝑥𝑚𝑒𝑑𝑖𝑎𝑛subscript𝑥subscript𝑞minsubscript𝑥subscript𝑞max\hat{x}=median(\{x_{q_{\rm min}},...,x_{q_{\rm max}}\})over^ start_ARG italic_x end_ARG = italic_m italic_e italic_d italic_i italic_a italic_n ( { italic_x start_POSTSUBSCRIPT italic_q start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT end_POSTSUBSCRIPT , … , italic_x start_POSTSUBSCRIPT italic_q start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT end_POSTSUBSCRIPT } )
4 return x^^𝑥\hat{x}over^ start_ARG italic_x end_ARG
Algorithm 3 Moving-average process (Sec. II E)
Refer to caption
Figure 5: PSDs of the outputs from EoA with different qmaxsubscript𝑞maxq_{\rm max}italic_q start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT (red lines). These are compared with the corresponding result from MA with single value of q𝑞qitalic_q (blue lines). Cases of different qmaxsubscript𝑞maxq_{\rm max}italic_q start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT and q𝑞qitalic_q are shown.
Refer to caption
Figure 6: EoA with a median collector on the output of AR stage. Collector function aggregates MAs from q=1𝑞1q=1italic_q = 1 to q=qmax𝑞subscript𝑞maxq=q_{\rm max}italic_q = italic_q start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT.

II.6 Bandpass filtering

In the last step of our framework, we have xEoAsubscript𝑥EoAx_{\rm EoA}italic_x start_POSTSUBSCRIPT roman_EoA end_POSTSUBSCRIPT bandpass filtered in the frequency range of 325123251232-51232 - 512 Hz for removing noise out of this band (e.g. seismic noise at low frequencies and photon shot noise at high frequencies). We adopt the Finite Impulse Response (FIR) filter by using the functions filtfilt and fir1 from the R package signal555https://cran.r-project.org/web/packages/signal/index.html. The bandpass filtered signal of GW150914 is shown at the bottom panel in Figure 2.

In comparison with the output from the previous step (i.e. xEoAsubscript𝑥EoAx_{\rm EoA}italic_x start_POSTSUBSCRIPT roman_EoA end_POSTSUBSCRIPT), no significant improvement can be found as a result of bandpass, which indicates that seqARIMA has already efficiently suppressed the noise in the raw data.

Although the bandpass does not appear to be necessary in the case of GW150914, we keep it in our framework to ensure all unwanted modulations outside this band are removed for the sake of comparing with the whitening results.

III Experimental results

III.1 Simulated data

For comparing the performance between the frequency-domain whitening filter and seqARIMA in noise reduction as well as waveform visualization, we have carried out a series of experiments. We started by simulating clean waveforms of a BBH coalescence at different luminosity distance dLsubscript𝑑𝐿d_{L}italic_d start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT by the code get_td_waveform from pycbc 666https://pycbc.org with the model SEOBNRv4_opt.

We have considered dLsubscript𝑑𝐿d_{L}italic_d start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT in a range from 200-4000 Mpc with a step size of ΔdLΔsubscript𝑑𝐿\Delta d_{L}roman_Δ italic_d start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT=200 Mpc. In order to analyse the denoising performance for a variety of waveform, for each dLsubscript𝑑𝐿d_{L}italic_d start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT, we have generated 100 waveform of randomly sampled individual component masses m1subscript𝑚1m_{1}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and m2subscript𝑚2m_{2}italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. For the other parameters such as dimensionless spin and eccentricity, the default values of get_td_waveform are adopted (See 777https://pycbc.org/pycbc/latest/html/pycbc.waveform.html). These waveform are defined as the signals s𝑠sitalic_s.

For the sampling of waveform parameters, we have firstly fitted the distributions of m1subscript𝑚1m_{1}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and m2subscript𝑚2m_{2}italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT from all the 81 confirmed BBH CBC events with the R package gamlss888https://www.gamlss.com. Among all the distribution functions available in gamlss999https://search.r-project.org/CRAN/refmans/gamlss.dist/html/gamlss.family.html, generalized Beta distributions of second kind provides the best description in accordance with AIC. And we sampled m1subscript𝑚1m_{1}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and m2subscript𝑚2m_{2}italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT from these best-fitted distributions.

For dLsubscript𝑑𝐿d_{L}italic_d start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT, we have generated 100 noise data of 32 s, which is defined as n𝑛nitalic_n. They are sampled from a PSD simulated by aLIGOZeroDetHighPower in pycbc from LALsimulation with low_freq_cutoff of 15 Hz. Each of them are generated with different random seed. The preparation of the simulated data was finished by injecting s𝑠sitalic_s in a random time location of n𝑛nitalic_n.

This simulated dataset allows us to compare the performance of seqARIMA and whitening filter in extracting the injected signal at varying dLsubscript𝑑𝐿d_{L}italic_d start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT. In both methods of seqARIMA and whitening, the same bandpass filter of 325123251232-51232 - 512 Hz were applied. For whitening, we have adopted a segment length of 4 s and the an overlap percentage of 50%percent5050\%50 % in all experiments. Such choices of whitening parameters follow the standards given by the pycbc documentation 101010https://pycbc.org

For quantifying the fidelity of the extracted signal, we computed the cross-correlation functions (CCFs) defined as,

CCF(t)=t=s(t)s^(tt)CCFsuperscript𝑡superscriptsubscript𝑡𝑠𝑡^𝑠𝑡superscript𝑡{\rm CCF}(t^{\prime})=\sum_{t=-\infty}^{\infty}s(t)\hat{s}(t-t^{\prime})roman_CCF ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) = ∑ start_POSTSUBSCRIPT italic_t = - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_s ( italic_t ) over^ start_ARG italic_s end_ARG ( italic_t - italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) (8)

where s𝑠sitalic_s is the simulated waveform and s^^𝑠\hat{s}over^ start_ARG italic_s end_ARG is the denoised data. Then we obtained the maximum values of |CCF|CCF|{\rm CCF}|| roman_CCF |, CCFmaxsubscriptCCFmax{\rm CCF_{max}}roman_CCF start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT, as the metric of measuring the similarity between the s𝑠sitalic_s and s^^𝑠\hat{s}over^ start_ARG italic_s end_ARG. In order to evaluate the noise reduction performance, we also computed the root-mean-square errors (RMSEs) defined as,

RMSE=N(ss^)2NRMSEsuperscript𝑁superscript𝑠^𝑠2𝑁{\rm RMSE}=\sqrt{\frac{\sum^{N}(s-\hat{s})^{2}}{N}}roman_RMSE = square-root start_ARG divide start_ARG ∑ start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ( italic_s - over^ start_ARG italic_s end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_N end_ARG end_ARG (9)

where N𝑁Nitalic_N is the length of data, which reflects how the noise is suppressed in the whole time series. For each dLsubscript𝑑𝐿d_{L}italic_d start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT, we have re-sampled n𝑛nitalic_n with 100 different random seeds and computed the median and the 95% confidence interval of CCFmaxsubscriptCCFmax{\rm CCF}_{\rm max}roman_CCF start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT and RMSE from this sample.

In Figure 7, the results are shown for dL=subscript𝑑𝐿absentd_{L}=italic_d start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT =400, 2000, 3000, and 4000 Mpc. For a visual comparison of the similarity of the extracted signal and the injected waveform s𝑠sitalic_s, we have also overlaid s𝑠sitalic_s in all the panels of Figure 7 as the red solid curves.

In the left panels of Figure 8, we compare how CCFmaxsubscriptCCFmax{\rm CCF}_{\rm max}roman_CCF start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT and RMSE vary with dLsubscript𝑑𝐿d_{L}italic_d start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT in both schemes. The error bars represent 95% confidence intervals calculated from 100 simulated waveform with randomly sampled m1subscript𝑚1m_{1}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and m2subscript𝑚2m_{2}italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT as well as different random seeds for generating the noise. Comparing the extracted signals by these two methods, we found that those obtained by seqARIMA generally have a larger degree of similarity with s𝑠sitalic_s and lower level of noise. Although whitening process attains better results for small distance (dL600subscript𝑑𝐿600d_{L}\leq 600italic_d start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ≤ 600 Mpc), seqARIMA has shown advantage in de-noising for increasing dLsubscript𝑑𝐿d_{L}italic_d start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT (i.e. larger CCFmaxsubscriptCCFmax{\rm CCF}_{\rm max}roman_CCF start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT and reduced RMSE).

In the right panels of Figure 8, we show the fractional improvements in both metrics as yielded by seqARIMA at different dLsubscript𝑑𝐿d_{L}italic_d start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT. Comparing with the whitening results at dL=4000subscript𝑑𝐿4000d_{L}=4000italic_d start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT = 4000 Mpc, seqARIMA has improved CCFmaxsubscriptCCFmax{\rm CCF}_{\rm max}roman_CCF start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT by 42similar-toabsent42\sim 42~{}∼ 42% and suppressed RMSE by 23similar-toabsent23\sim 23~{}∼ 23%.

Refer to caption
Figure 7: The comparison of extracted waveforms from the simulated data (top panel) by whitening (middle panel) and seqARIMA denoising (bottom panel) for dL=subscript𝑑𝐿absentd_{L}=italic_d start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT =400, 2000, 3000, and 4000 Mpc, from top to bottom, respectively. In each panel, we have overlaid the injected signal (red lines) on the data (black lines). For the sake of comparison, the strain amplitudes are normalized in each panel.
Refer to caption
Figure 8: (left panel) The comparison of CCFmaxsubscriptCCFmax{\rm CCF_{max}}roman_CCF start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT and RMSE between the injected signal and the waveform extracted from seqARIMA (red triangles) and the whitening process (blue circles) with varying dLsubscript𝑑𝐿d_{L}italic_d start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT. The error bars represent 95% confidence errors calculated from 100 sampled parameters and 100 different random seeds. (right panel) Fractional difference of CCFmaxsubscriptCCFmax{\rm CCF_{max}}roman_CCF start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT and RMSE resulted from seqARIMA denoising with respect to the corresponding metrics resulted from whitening as a function of dLsubscript𝑑𝐿d_{L}italic_d start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT.

III.2 LIGO data

To demonstrate the capability of seqARIMA in handling real data, we have attempted to extract the signals from a number of known GW events from the LIGO data. In this test, we have chosen all the events in GWTC-1 as observed during the first and second observation runs (O1 and O2) in 2015-2017 [44] plus two additional interesting events. All the data with a length of 4096 s with 4 kHz sampling frequency are obtained from GWOSC. Except for the NS-NS merger GW170817, we windowed the 4096 s data with a frame of 32 s for all the events. For GW170817, because of its much longer timescale, we apply a window of 50 s instead.

Event Name Dataset RMSE CCFmaxmax{}_{\rm max}start_FLOATSUBSCRIPT roman_max end_FLOATSUBSCRIPT
H1 L1 H1 L1
GW150914 Whiten 0.034 0.0288 0.628 0.488
seqARIMA 0.024 0.0321 0.646 0.607
-29.4 % +11.4 % +2.83 % +24.4 %
GW151012 Whiten 0.0318 0.0326 0.127 0.0922
seqARIMA 0.0304 0.0311 0.201 0.177
-4.39 % -4.71 % +57.9 % +92.1 %
GW151226 Whiten 0.0119 0.0119 0.0763 0.0736
seqARIMA 0.0114 0.0117 0.145 0.109
-4.05 % -1.81 % +90.6 % +47.6 %
GW170104 Whiten 0.0289 0.0283 0.212 0.212
seqARIMA 0.0262 0.0264 0.298 0.293
-9.27 % -6.71 % +40.3 % +38.2 %
GW170608 Whiten 0.0145 0.0144 0.0926 0.105
seqARIMA 0.0138 0.0143 0.18 0.143
-4.92 % -1.23 % +94 % +37 %
GW170729 Whiten 0.0375 0.0391 0.278 0.379
seqARIMA 0.033 0.0309 0.443 0.51
-12.2 % -20.9 % +59.7 % +34.5 %
GW170809 Whiten 0.0328 0.0309 0.246 0.305
seqARIMA 0.0315 0.0304 0.367 0.35
-3.75% -1.5 % +49.4 % +15 %
GW170814 Whiten 0.0309 0.027 0.303 0.402
seqARIMA 0.0255 0.0276 0.467 0.546
-17.6 % +2.0 % +54 % +35.8 %
GW170817 Whiten 0.00836 0.00828 0.0719 0.106
seqARIMA 0.00816 0.00793 0.113 0.163
-2.38 % -4.16 % +57.1 % +54.1 %
GW170818 Whiten 0.035 0.0335 0.119 0.273
seqARIMA 0.0325 0.0319 0.218 0.359
-7.19 % -4.99 % +82.9 % +31.2 %
GW170823 Whiten 0.0336 0.0349 0.306 0.3
seqARIMA 0.0352 0.0337 0.434 0.362
+4.78 % -3.51 % +41.7 % +20.7 %
GW190814 Whiten 0.0118 0.0117 0.0951 0.109
seqARIMA 0.0113 0.0112 0.171 0.182
-4.68 % -4.62 % +79.7 % +67.2 %
GW200105 Whiten - 0.0121 - 0.0316
seqARIMA - 0.0115 - 0.174
- -5.59 % - +449 %
Table 1: The comparison of RMSE and CCFmaxmax{}_{\rm max}start_FLOATSUBSCRIPT roman_max end_FLOATSUBSCRIPT yielded by whitening and seqARIMA on 13 confirmed CBC events as observed by LIGO with reference to the model waveform as specified in the literature. The third row of each event show the percentage change resulted from seqARIMA with respect to whitening. For GW200105, LIGO-H was not operational during this event as hence there is no data available [45].

III.2.1 GWTC-1 events

All 11 events in GWTC-1 can be well extracted by seqARIMA. In Figure 9, we show the spectrograms/oscillograms of the extracted signals from three representative cases, GW150914, GW151012, and GW170817, as detected by both observatories in Hanford (H: left panels) and Livingston (L: right panels). For the results of other GWTC-1 events, we have put them in the Appendix (Figure 11).

GW150914 is the first case that a GW signal was directly detected [4]. Its high signal-to-noise (SNR) of 26 has put it among the strongest signals of BBH merger detected so far. In Section II, we have already used this case for illustrating the feasibility of seqARIMA, in which we demonstrate that the signal of GW150914 can be clearly recovered. In the top row of Figure 9, we have produced the spectrograms of this event with Q-transform for visualizing how the frequency of the signal varies over the entire process. The characteristic sweeping chirp can be clearly seen in the spectrograms.

GW151012 is the BBH merger detected with a SNR of 10, which puts it as the weakest signal in GWTC-1 [44]. Its low significance as found from the initial discovery in O1 did not make it as a confirmed detection. And hence it was firstly considered as a candidate which was named as LVT151012 [46]. With a more detailed analysis, it was found to meet the criteria of a confident detection and was subsequently re-named as GW151012. In the second row of Figure 9, we show the spectrograms of the signals of GW151012 as extracted by seqARIMA. The chirp-like feature can be seen from the denoised data though it is not as clear as in the case of GW150914 because of its low significance.

The GW signal from the event GW170817 is resulted from a merging NS-NS binary, which is the first GW event that has the counterpart detected across the whole electromagnetic spectrum [5, 6, 7, 8]. It is associated with a short γlimit-from𝛾\gamma-italic_γ -ray burst GRB170817A, detected by Fermi Gamma-ray Burst Monitor (GBM) 1.7 s after the coalescence [8]. It has provided a long-sought evidence for the link between NS-NS mergers and short γlimit-from𝛾\gamma-italic_γ -ray bursts. Unlike BBHs, the inspiral time of GW170817 is much longer. Therefore, we take this event as a test for the capability of our framework in handling a signal with a longer timescale. Apart from adopting a wider window in the analysis, since LIGO-L data of GW170817 suffered from the transient noise (or glitch) at the GPS time of 1187008881.389 (around 1.1 s before the coalescence), we have used the data after noise subtraction following the glitch model described in [5]111111https://www.gw-openscience.org/events/GW170817. The spectrograms of GW170817 resulted from seqARIMA denoising are shown in the bottom panels of Figure 9. The inspiral and the merging process over 30similar-toabsent30\sim 30∼ 30 s can be clearly visualized in both data.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 9: The spectrograms and oscillograms of seqARIMA-denoised LIGO data of GW150914, GW151012, and GW170817 as selected from GWTC-1 (Results for the other GWTC-1 events are shown in Figure 11). The color scale for the normalized power of the spectrogram is given at the upper-left corner of each panel. The reference epochs t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT for each case are the event time reported in GWOSC. For GW170817, the glitch-removed LIGO-L data is used.

III.2.2 GW190814 & GW200105_162426

Apart from reproducing the GWTC-1 events, we have further tested our framework on two additional sources: GW190814 & GW200105_162426. These events were chosen because their inferred properties are somewhat different from those 11 events in GWTC-1.

GW190814 was detected in the third observing run (O3) with a SNR of 25 [48]. Parameter estimation suggests that the masses of the compact objects in their progenitor binary are highly unequal. While one component has its mass estimated as 23Msimilar-toabsent23subscript𝑀direct-product\sim 23M_{\odot}∼ 23 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT which is consistent with a stellar BH, the mass of the other one is likely lying in a range of 2.53Msimilar-toabsent2.53subscript𝑀direct-product\sim 2.5-3M_{\odot}∼ 2.5 - 3 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT which put it in a mass gap of being either a very massive NS or a low-mass BH. In the Fig. 1 of [48], we notice that the timescale of GW190814 is 23similar-toabsent23\sim 2-3∼ 2 - 3 s long which is different from those of GWTC-1 sources. Therefore, we have included GW190814 in our test.

For the same reason, we have also included GW200105_162426 (hereafter GW200105) in our experiment. It was detected by a single detector (LIGO-L) during O3 with an SNR of 14similar-toabsent14\sim 14∼ 14 [45]. It is estimated to have component masses of 8.9Msimilar-toabsent8.9subscript𝑀direct-product\sim 8.9M_{\odot}∼ 8.9 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and 1.9Msimilar-toabsent1.9subscript𝑀direct-product\sim 1.9M_{\odot}∼ 1.9 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT which makes it likely an NS-BH binary. The signal of GW200105 shows a track of excess power with increasing frequency over 3similar-toabsent3\sim 3∼ 3 s in the spectrogram [see Fig. 1 in 45].

In Figure 10, we show the spectrograms of these two sources produced in our framework. The tracks of the signals in both cases are clearly visible. In comparing the spectrogram of GW200105 resulted from seqARIMA and the one obtained from spectral whitening as shown in Fig. 1 of [45], we found that our result can attain a higher clarity which shows the inspiraling stage has a duration up to 6similar-toabsent6\sim 6∼ 6 s.

In order to compare the performance of signal extraction by whitening and seqARIMA, we computed the CCFmaxsubscriptCCFmax{\rm CCF}_{\rm max}roman_CCF start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT and RMSE resulted from both schemes with reference to the waveforms generated by pycbc with the model of SEOBNRv4_opt for BBHs and IMRPhenomPv2 for GW170817 (BNS), GW190814 (Mass-gap), and GW200105 (NSBH) according to the parameters given in the corresponding literature. The results are summarized in Table 1. For comparing RMSE between seqARIMA and whitening, we have seen general improvement in most cases. However, there are a few cases that the noise reduction resulted from seqARIMA are worse than that from whitening. The most notable one is from the LIGO-L data of GW150914. This might suggest that for the events with SNR sufficiently large as in the case of GW150914, seqARIMA may not have the advantage over the conventional whitening. This is also reflected by the non-monotonic behavior for the small values of dLsubscript𝑑𝐿d_{L}italic_d start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT in Figure 8. On the other hand, in terms of CCFmaxsubscriptCCFmax{\rm CCF}_{\rm max}roman_CCF start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT (i.e. the similarity between the extracted signals and the model), seqARIMA has shown improvement in all our tested cases.

Refer to caption
Refer to caption
Refer to caption
Figure 10: The spectrograms and oscillograms of seqARIMA-denoised data of GW190814 (LIGO-H, L) and GW200105_162426 (LIGO-L).

IV Summary & Future Prospects

In this work, we have proposed a novel de-noising technique in processing GW data, which is based on autoregressive modeling. By coupling with other techniques (i.e. integrated process, EoA), we have developed a framework we refer as seqARIMA pipeline (cf. Figure 1). The effects of each component in the pipeline have been investigated (see Figure 2 and subsection II.3-subsection II.6 for details). We have tested the performance of our proposed framework with a series of experiments.

We have examined the ability of seqARIMA pipeline in extracting the simulated GW signal with varying waveform and distance. By comparing the noise-subtracted time series and the injected signal (Figure 7), we have computed CCFmaxmax{}_{\rm max}start_FLOATSUBSCRIPT roman_max end_FLOATSUBSCRIPT and RMSE resulted from both seqARIMA and whitening process. At larger distance, we found that seqARIMA can attain a higher CCFmaxmax{}_{\rm max}start_FLOATSUBSCRIPT roman_max end_FLOATSUBSCRIPT and lower RMSE than those resulted from whitening (Figure 8).

We have also applied our method in extracting a number of known GW events from the LIGO data. All 11 events cataloged in GWTC-1 can be well recovered by seqARIMA (Figure 9 & Figure 11). We have further tested the method in two additional sources GW190814 (mass-gap object) and GW200105 (NS-BH merger), which have the timescale of their GW signals different from those in GWTC-1. We showed their signals can also be successfully extracted (Figure 10).

We have further compared the CCFmaxmax{}_{\rm max}start_FLOATSUBSCRIPT roman_max end_FLOATSUBSCRIPT and RMSE resulted from both seqARIMA and whitening by comparing the noise-subtracted time series of these events with the model waveforms generated in accordance with the parameters specified in the corresponding literature (see Table 1). We found that seqARIMA generally yields improvement over whitening in terms of these performance metrics.

We have demonstrated that seqARIMA can enhance the noise suppression and therefore it is capable to provide an alternative to the conventional frequency-domain whitening process. For further improving the denoising performance, seqARIMA can be coupled with deep learning. Many recent studies have investigated the feasibility of denoising the GW data with deep neural network and showed that this can significantly suppress the noise and recover the signal [e.g. 49, 50, 51, 52]. We notice that these recent studies remain using whitening as a preprocessing procedure. Therefore, it will be encouraging to explore whether combining seqARIMA with these machine-learning based architectures can boost the denoising performance to a further extent. By substituting whitening with our proposed method, dedicated studies can also explore whether parameter estimation can also be benefited from seqARIMA.

We can also consider the feasibility of incorporating seqARIMA into a template-free low-latency detection pipeline. Since whitening can be a dominant source for the latency, it is desirable to reduce the computational cost in this stage [e.g. 19]. However, the conventional frequency-domain whitening process do not have many degree of freedom for improving the computational efficiency. On the other hand, the complexity of seqARIMA can be controlled by the hyper-parameters pmaxsubscript𝑝maxp_{\rm max}italic_p start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT and qmaxsubscript𝑞maxq_{\rm max}italic_q start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT, which gives the flexibility of this process. For example, in trading off the fidelity of the extracted signal, a low pmaxsubscript𝑝maxp_{\rm max}italic_p start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT can result in a more efficient modeling. Therefore, one can examine whether seqARIMA can be adopted in a candidate identification pipeline. With the improved noise subtraction, the signal from a CBC or burst event can possibly be identified as a cluster of bright pixels in the spectrograms [e.g. 53, 54] which allows a GW event candidate to be detected without a priori knowledge of its waveform. A quantitative analysis on the execution speed of our proposed framework will be important for examining the capability of rapid real-time processing.

Acknowledgements.
The authors would like to thank Dr. Wang He for his valuable comments for improving the quality of this work. S.K. is supported by the National Research Foundation of Korea grant 2022R1F1A1073952. C.Y.H. is supported by the research fund of Chungnam National University and by the National Research Foundation of Korea grant 2022R1F1A1073952. A.K.H.K. is supported by the National Science and Technology Council of Taiwan through grants 111-2112-M-007-020 and 112-2112-M-007-042. L.C.C.L is supported by NSTC of Taiwan through grant Nos. 110-2112-M-006-006-MY3 and 112-2811-M-006-019. K.L.L. is supported by the National Science and Technology Council of the Republic of China (Taiwan) through grant 111-2636-M-006-024, and he is also a Yushan Young Fellow supported by the Ministry of Education of the Republic of China (Taiwan). J.Y. and A.P.L. is supported by the Science and Technology Development Fund, Macau SAR (No. 0079/2019/A2).

References

Appendix A Spectrograms and oscillograms of GWTC-1 events produced by our framework

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 11: The spectrograms and oscillograms of seqARIMA-denoised LIGO data of GWTC-1 sources which are not shown in Figure 9.
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 12: The spectrograms and oscillograms of seqARIMA-denoised LIGO data of GWTC-1 sources which are not shown in Figure 9.