Brought to you by:

The following article is Open access

Compressed Sensing Based RFI Mitigation and Restoration for Pulsar Signals

, , , and

Published 2022 August 22 © 2022. The Author(s). Published by the American Astronomical Society.
, , Citation Hao Shan et al 2022 ApJ 935 117 DOI 10.3847/1538-4357/ac8003

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/935/2/117

Abstract

In pulsar signal processing, two primary difficulties are (1) radio-frequency interference (RFI) mitigation and (2) information loss due to preprocessing and mitigation itself. Linear mitigation methods have a difficulty in RFI modeling, and accommodate a limited range of RFI morphologies. Thresholding methods suffer from manual factors and adaptability. There is also a distinct lack of methods dedicated to information loss. In this paper, a novel method "CS-Pulsar" is proposed. It carries out compressed sensing (CS) on time-frequency signals to accomplish RFI mitigation and signal restoration simultaneously. Curvelets allow an optimal sparse representation for multichannel pulsar signals containing the time-of-arrival dispersion relationship. CS-Pulsar mitigation is implemented using a regularized least-squares framework that does not require the statistics of RFI to be known beforehand. CS-Pulsar implements channel restoration, and useful signal contents are retrieved from the measurement error by a morphological component analysis aided by the root-mean-square envelope. These two steps allow CS-Pulsar to provide key signal details for special astrophysical purposes. Experiments of signal restoration for pulsar data from the Nanshan 26 m radio telescope reveal the advantage of CS-Pulsar. The method successfully removes false peaks due to on-pulse RFI in multipeaked pulsar profiles. CS-Pulsar also increases the timing accuracy and signal-to-noise ratio proving its feasibilities and prospects in astrophysical measurements.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Observed data of pulsar radio emission always suffer from various types of radio-frequency interference (RFI). The RFI is caused by, e.g., human communications technologies such as satellites, mobile base stations, or navigation radars. Linear mitigation methods such as principal component analysis (Li et al. 2006) and singular vector decomposition (Pen et al. 2009) can only remove limited types of RFI (Offringa et al. 2010), and have difficulty in RFI modeling (Zeng et al. 2021). RFI modeling can be simply understood as, e.g., RFI being an additive layer of the time-frequency signal profiles or being represented in a sparse transform domain. An alternative set of methods, called thresholding methods, include, e.g., the cumulative sum (Baan et al. 2004) and the combinatorial thresholding methods (VarThreshold and SumThreshold; Offringa et al. 2010). Some of them benefit from a combined effect of adjacent samples. However, this combined property cannot perform feature representation and approximation for multichannel signals containing the dispersion relationship. The SumThreshold also involves empirical factors and an additional rule to avoid excessive flagging.

The problem of information loss is also seemingly neglected in the field of pulsar signal processing, and few methods are dedicated to it. Offringa et al. (2015) presented detailed low-frequency RFI statistics and found that 1.1% of data were lost to RFI detection. Information loss is generally caused by two effects. The first is channel saturation. In channels containing both strong intermediate-frequency signals and excessive RFI, no useful information can be identified, and a common solution is to set all values as zero. This leads to complete and irreparable information loss in these channels. Thus, if zero-setting can be replaced with missing channel restoration, some of the lost information can be recovered. The second information-loss effect is smoothing of RFI mitigation itself, which requires additional signal retrieval methods after the channel restoration. In this paper, we consider information loss because it can impact high-precision astrophysical measurements, e.g., timing accuracy. To refine the current methods of dealing with RFI mitigation and information loss, a conceptual innovation solving the two problems in one framework is proposed here.

The basic principle of compressed sensing (CS; Donoho 2006; Candés et al. 2006b, 2006c; Candés & Tao 2005, 2006) is that a sparse or compressible signal can be exactly recovered from incomplete linear random or pseudorandom measurements through exploration of signal sparsity and optimization algorithms. The assumption of "sparsity" means that a signal has a limited number of nonzero entries, and "compressibility" is a relaxation of the assumption of "sparsity" for signals that are not strictly sparse (see Section 3.2). Unlike traditional measurements, which have to satisfy the Shannon/Nyquist sampling theorem (Nyquist 1928; Shannon 1949), CS measurements can attain sub-Nyquist rates without the limitation of Fourier bandwidth by measuring a linear combination of the signal components under a measurement matrix (Akçakaya & Tarokh 2010).

CS has already been applied to radio astronomy. In radio interferometry imaging (Wiaux et al. 2009), CS makes it possible that sparse or compressible astrophysical signals probed by incomplete and noisy Fourier measurements of interferometry can be accurately reconstructed based on a convex optimization. A CS-based deconvolution method (Li et al. 2011b) can reconstruct both point sources and extended sources using an isotropic undecimated wavelet transform. For another example, reconstruction of the Faraday dispersion function (Li et al. 2011a) from limited measurements of polarized radio emission is ill-conditioned, because these measurements are possible only for a limited range of wavelengths. A CS reconstruction model can be realized by reforming the Fourier relationship and projecting the Faraday dispersion function to the polarized emission. One can also refer to Jiang et al. (2015) for interferometric radio transient reconstruction. In radio RFI mitigation, CS has been applied especially to compressed sampled data of synthetic aperture radar (SAR). Cucho-Padin et al. (2019) proposed a compressive statistical sensing to estimate features of the cyclic spectrum of sub-Nyquist SAR data to mitigate quasiperiodic RFI. There are also CS methods exploring the sparse feature of RFI for SAR data, such as Mai et al. (2013) and Kong et al. (2016). It is worth mentioning that a technique belonging to optimization algorithms but non-CS is a dictionary-based nonconvex low-rank minimization (Tang et al. 2022), which can overcome improper punishments by singular value thresholding.

Despite these successes, applications of CS to radio pulsar signals remain underexplored. But exploration of sparsity in CS frameworks in these works inspired the application of curvelets (Candés et al. 2006a; Candés & Donoho 2000, 2004, 2005) and the attempt of signal restoration in this paper. For the second half of our problem, missing data restoration, we can draw from noteworthy results in seismic data processing. With 50% of sources missing, Herrmann et al. (2011) estimated the seismic data volumes by the curvelet coefficients and reconstructed conventionally sampled Nyquist data. Gan et al. (2016) used a CS approach based on a fast projection onto convex sets with sparsity constraint. For iterative thresholding, Chen et al. (2014) presented a data-driven percentile-half-thresholding for iterative shrinkage thresholding to reconstruct irregular seismic data. A toolbox of Sparco (van den Berg et al. 2007) can realize accurate seismic wavefield reconstruction. These advancements in, e.g., missing data reconstruction and iterative thresholding schemes have not yet been transferred to pulsar signal processing.

This paper proposes a novel CS-based model for simultaneous RFI mitigation and signal restoration. This work explores the sparsity of the dispersed pulses containing the time-of-arrival (TOA) feature for the purposes of special astrophysical measurements, e.g., high-precision pulsar timing. Throughout this paper, an assumption is made that pulsar signals are sparse or compressible. With this assumption, a nonlinear approximation of curvelet coefficients can be applied as a preliminary attempt to restore missing channels caused by the two effects mentioned above. This method can remove most types of RFI typically presented in pulsar signals, especially the on-pulse RFI, by the regularized least-squares (LS) framework. A morphological component analysis (MCA; Starck et al. 2005) based data retrieval is also introduced. Two advantages are restoration efficiency for missing channels and false peak removal in the mode of multiple peaks (Rankin & Rathnasree 1995; Rankin & Rathnasree 1997). The sparse operators will be given in Section 2. The method is introduced in Section 3. Section 4 shows the experimental results, and conclusions are drawn in Section 5.

2. Sparse Operators

The TOA–dispersion relationship is one of the most important features of pulsars, and also plays a decisive role in pulsar timing. Multichannel signals containing this relationship allow for feature representation and approximation via sparse transforms such as the fast Fourier transform, Haar wavelet, Daubechies wavelet (Daubechies 1992), discrete cosine transform (DCT), and the curvelet. The two-dimensional (2D) wavelets (Vetterli & Kovačević 1995; Strang & Nguyen 1996; Mallat 1997; Mertins 1999) are good at isolating point-like discontinuities, and can also be used for edge and curve detection (Li 2003; Mallat & Hwang 1992; Mallat & Zhong 1992). The curvelet is especially designed to deal with the second-order continuously differentiable singularities (i.e., C2-singularities) with its needle-shaped elements, which are particularly directionally sensitive (Candés & Donoho 2002). Signals containing the feature of the TOA–dispersion relationship are very suitable to be represented and approximated by curvelet coefficients. This paper will apply the second generation curvelet transform (Candés & Donoho 2004; Candés et al. 2006a).

The accuracy of feature approximation should be reviewed. Suppose a pulsar signal profile ${\boldsymbol{F}}\in {L}^{2}({{\mathbb{R}}}^{2})$ is uniformly regular, i.e., it belongs to Cβ . For the dispersion relationship, β = 2. The sparsity reveals that if F is singular along the C2 curve but otherwise smooth, the asymptotic approximation error obeys $\parallel {\boldsymbol{F}}-{{\boldsymbol{F}}}_{s}{\parallel }_{2}^{2}\leqslant K\,{{\rm{s}}}^{-2}$, where F s is the best s-term nonlinear approximation and K is a constant. The Fourier approximation achieved by the Fourier series fails as $\parallel {\boldsymbol{F}}-{{\boldsymbol{F}}}_{s}{\parallel }_{2}^{2}\leqslant K\,{{\rm{s}}}^{-1/2}$. The wavelet series is better than the Fourier series with only $\parallel {\boldsymbol{F}}-{{\boldsymbol{F}}}_{s}{\parallel }_{2}^{2}\leqslant K{{\rm{s}}}^{-1}$. Surprisingly, the curvelet achieves $\parallel {\boldsymbol{F}}-{{\boldsymbol{F}}}_{s}{\parallel }_{2}^{2}\leqslant K\cdot {\left(\mathrm{log}s\right)}^{3}\,{{\rm{s}}}^{-2}$. The optimality of the curvelet reveals that with the same s terms, no other sparse operators can yield a smaller asymptotic error. Compared with wavelets, the curvelet uses far fewer coefficients to represent anisotropic objects for a given accuracy. Sparsity is the core factor of CS, relying on the fact that most natural signals including pulsar signals are sparse or compressible in transform domains.

3. Methodology

3.1. Pulsar Signals and Preprocessing

Suppose a digital filter bank has P equal-width frequency channels. Upon saving the data to disk, the observation program divides the time samples in one pulse period into T phase bins. Let an observed 2D pulsar profile be a digital matrix F 0 within an observing bandwidth $[{\nu }_{1}^{1},{\nu }_{P}^{2}]$. Each frequency channel is tuned to receive a signal ${{\boldsymbol{f}}}_{p}^{\mathrm{ch}},1\leqslant p\leqslant P$ within a narrow bandwidth $[{\nu }_{p}^{1},{\nu }_{p}^{2}]$ with νp as the central frequency, where the vector ${{\boldsymbol{f}}}_{p}^{\mathrm{ch}}={({f}_{{pt}}^{\mathrm{ch}})}_{1\leqslant t\leqslant T}$. In this way, the broadband F 0 is split into adjacent narrowband signals. The dispersion frequency relationship (Lorimer & Kramer 2005) is calculated as

Equation (1)

where DC = 4.15 × 103 (MHz2 pc−1 cm3 s) is the dispersion constant, DM (parsecs per cubic centimeter) is the dispersion measure, and ${\nu }_{\mathrm{ref}}=({\nu }_{1}^{1}+{\nu }_{P}^{2})/2$ is the reference frequency and often chosen to be the central frequency of the broadband. Frequencies are all in megahertz. Let d be the distance to the pulsar, and ne be the electron number density. DM is defined as $\mathrm{DM}={\int }_{0}^{d}{n}_{{\rm{e}}}{\rm{d}}l$, where l is the line of sight. For the profiles collected by the Nanshan 26 m Radio Telescope, we set P = 1024 and T = 512 for calculation.

A necessary three-step preprocessing is carried out. Step 1: baseline removal by a channel mean subtraction. Step 2: zero-setting for saturated channels, resulting in an F 2. Step 3: a simple threshold-to-zero to remove the huge-amplitude RFI, resulting in an F . Let γ be the threshold of Step 3. The threshold-to-zero means that for an ${{\boldsymbol{f}}}_{p}^{\mathrm{ch}},1\leqslant p\leqslant P$, if $\max ({{\boldsymbol{f}}}_{p}^{\mathrm{ch}})-\min ({{\boldsymbol{f}}}_{p}^{\mathrm{ch}})\gt \gamma $, then ${{\boldsymbol{f}}}_{p}^{\mathrm{ch}}={\bf{0}}$. Suppose the median absolute deviation (MAD) of a matrix X is MAD( X ) = median∣ X − median( X )∣, then the standard deviation $\sigma =\tfrac{\mathrm{MAD}}{0.6745}$. For the case of Gaussian white noise, γ can be selected as 3σ (Starck et al. 2010). In Step 3, a preliminary removal for huge-amplitude RFI is carried out by setting $\gamma =6\sim 8\tfrac{\mathrm{MAD}\left({{\boldsymbol{F}}}_{2}\right)}{0.6745}$ or selecting a reasonable numerical number. Figure 1 takes PSR J1645-0317 (B1642-03), for example, and shows the process.

Figure 1.

Figure 1. The three-step preprocessing of PSR J1645-0317. (a) The observed original data. (b) Channel saturation and huge-amplitude RFI after baseline removal of Step 1. (c) Zooming-in details of the red rectangle in panel (b). Only an extremely weak signal portion can be identified by visual inspection. (d) Profile F with channel losses after Steps 2 and 3 with γ = 30. (e) Zooming-in details of panel (d).

Standard image High-resolution image

3.2. CS-based Mitigation and Restoration

Consider the standard CS problem with missing channels

Equation (2)

where ${\boldsymbol{\Phi }}\in {{\mathbb{R}}}^{L\times N}(L\leqslant N)$ is the CS measurement matrix and ${\boldsymbol{\epsilon }}\in {{\mathbb{R}}}^{L}$ is the measurement error. Signal ${\boldsymbol{u}}\in {{\mathbb{R}}}^{N}$ will be recovered via the noisy linear measurements ${\boldsymbol{f}}\in {{\mathbb{R}}}^{L}$, and f is the vector form of F and formed by concatenating every column f t , 1 ≤ tT of F . Let epsilon = σ ω , where ω is the vector of random Gaussian white noise and RFI, and σ controls the magnitude of epsilon . Unfortunately, there is no purely statistical way to decide the statistics of ω and the detailed information about σ. For the purpose of reconstruction, epsilon is a combined factor encompassing both Gaussian noise and RFI that can be executed with or without statistical characteristics. Φ has many more rows than columns. Suppose A is the number of the zero set or missing channels, then L = (PA)T and N = PT. Given Φ, the recovery of u from f is an underdetermined linear system leading to an ill-posed problem. To regularize it, assume that u has the property of "sparsity," i.e., u is s-sparse in ${\boldsymbol{\Theta }}={{\boldsymbol{\Psi }}}^{\dagger }\in {{\mathbb{R}}}^{M\times N}$ (M = N for a basis or M > N for a frame), such that u can be expressed by s nonzero coefficients in Θ, i.e., ∥Θ u 0 = ∥ α 0 = s, where ${\boldsymbol{\alpha }}\in {{\mathbb{R}}}^{M}$, MN, and † denotes the inverse transform. Then, u can be reconstructed with high accuracy from the incomplete measurements f . For most cases, signals of interest are not strictly sparse, but the assumption can be relaxed to "compressibility" if most of the energy in Θ u = α is contained in its largest s coefficients, i.e., amplitudes of the sorted scalar coefficients exhibit a polynomial decay: $| {\alpha }_{i}| \leqslant c\cdot {i}^{-\tfrac{1}{b}}$, 1 ≤ is, b ∈ [0, 1], where c is a constant. The smaller the b is, the faster the amplitudes decay, and the more compressible a signal is. Supposing Φ and Θ are two orthogonal basis or frames, and Φ is noise-like incoherent in the Θ domain, i.e., Φ Θ satisfies a sufficient condition named restricted isometry property (Candés & Tao 2005; Candés & Wakin 2008), CS tells us that the sparse coefficients α can be accurately recovered. Thus,

Equation (3)

Physically realizable random matrices that are incoherent to most transforms can be used as measurement matrices. One of the classical methods to solve Equation (3) is the generalized formulation of the unconstrained regularized LS based reconstruction, which reads

Equation (4)

where λ is a nonnegative regularization parameter balancing the data fidelity against the sparsity. $J:{{\mathbb{R}}}^{N}\to {\mathbb{R}}$ is a fixed nonnegative valued cost function, and it is also a penalization term of α to preserve the sparsity property. J could be various choices. For example, to solve the NP-hard problem of the 0 pseudo-norm, one can use its optimal convex approximation, i.e., a sparsity-promoting 1 minimization instead to assist the reconstruction. For another example, the total variation (TV) regularizer assumes that the pulse portions are sparse, and the usual discretizations of TV are assumed to be convex. The reason to apply the regularized LS framework of Equation (4) is that the statistics of α and w, and the detailed information about σ are not required (Vehkaperä et al. 2016). Thus, the measurement error is considered as a combination of the RFI and random Gaussian white noise. A suboptimal synthesis problem termed least absolute shrinkage and selection operator (Candés & Tao 2006; Tibshirani 1996; Chen et al. 1998) is considered:

Equation (5)

where $\parallel {\boldsymbol{\alpha }}{\parallel }_{1}={\sum }_{i=1}^{N}| {\alpha }_{i}| $. The fast iterative soft-thresholding algorithm (FISTA; Beck & Teboulle 2009) is derived from the proximal forward–backward iterative scheme (Bruck 1977; Passty 1979) in the general framework of splitting methods (Facchinei & Pang 2003). For solving Equation (5), the FISTA will be applied. Set $\lambda =\tfrac{a\cdot \mathrm{MAD}\left({\boldsymbol{F}}\right)}{0.6745}$, where a is a parameter that can be adjusted.

3.3. Signal Retrieval

The deleterious smoothing effect from RFI mitigation can cause precise astrophysical measurements (e.g., timing) to be negatively affected. It is necessary to perform signal retrieval after RFI mitigation to account for these effects. The signal retrieval will be carried out by a further MCA (Starck et al. 2005) decomposition of the measurement error epsilon . The MCA seeks the sparsest of all representations over the sparse dictionaries representing the morphological components. Suppose ${\boldsymbol{\epsilon }}\in {{\mathbb{R}}}^{L}$ contains three components of the removed signal details ${{\boldsymbol{u}}}^{{\rm{M}}}\in {{\mathbb{R}}}^{L}$, the RFI ${\boldsymbol{i}}\in {{\mathbb{R}}}^{L}$ and the Gaussian white noise ${\boldsymbol{w}}\in {{\mathbb{R}}}^{L}$, i.e., epsilon = u M + i + w . The MCA assumes that i can be considered as the texture-like features, which will be represented by, e.g., the local discrete cosine transform, and u M is regarded as the piecewise smooth component represented by the curvelet. After the default MCA decomposition, the components will also be extended to matrix forms ${{\boldsymbol{U}}}^{{\rm{M}}},{\boldsymbol{I}},{\boldsymbol{W}}\in {{\mathbb{R}}}^{P\times T}$ by supplementing zero vectors to the corresponding positions of their zero set channels. To retrieve finer details of the main peak, the upper root-mean-square envelope with a sliding window of length 128 points is applied, and U M is taken for example to interpret the retrieval strategy. Suppose the maximum value of the upper envelope of the signal after dedispersion of U M is eu, and phases of the signal samples above the threshold eu can be recorded. After a dispersion process, phases of samples in every channel can always be obtained by a simple phase shifting based on the recorded phases. In every channel, samples located on these phases should be recycled, and other samples will be set to zeros, resulting in a new profile U R with most entries being zeros. For components I and W , the thresholds are 3ei and 3ew, and the new profiles are I R and W R, respectively. Let R = U R + I R + W R and the matrix form of ${\boldsymbol{u}}\in {{\mathbb{R}}}^{N}$ be ${\boldsymbol{U}}\in {{\mathbb{R}}}^{P\times T}$; then a final profile P = U + R can be obtained. Figure 2 shows a flowchart for the algorithm of CS-Pulsar explaining the overall execution process from input of an observed signal profile F 0 to output of a final signal profile P . A normalization is carried out for two reasons. (1) The advantage of restoration efficiency should be clearly clarified. (2) Shape comparisons should be carried out between the reference signals and the restored ones. Several dedispersed signals are defined to avoid unnecessary confusions.

  • (1)  
    y 0: the standard reference signal.
  • (2)  
    y 2: dedispersion result of F 2 (after Steps 1 and 2).
  • (3)  
    y : dedispersion result of F (after Steps 1, 2, and 3).
  • (4)  
    y F, y H, y D, y C and y Cu: dedispersion results of U by the Fourier, Haar and Daubechies wavelets, DCT, and curvelet operators. m F, m H, m D, m C, and m Cu: dedispersion results of RFI mitigation by the five operators without considering the restored channels. p F, p H, p D, p C, and p Cu: dedispersion results of P by the five operators.

Figure 2.

Figure 2. Flowchart of overall execution process of the CS-Pulsar.

Standard image High-resolution image

Because there is no unified calibration standard at present, the reference signals are applied only for shape comparisons. Without loss of generality, the maximum value of y 0 will always be set as 1. Then, the maximum value of y is normalized as 1, and the other signals will be normalized according to their intensity ratios to y . The peak signal-to-noise ratio (PS/N: in dB) is used to measure the similarity of the restored signals. In order to show the restoration accuracy, a restoration ratio Br /B is applied as an assessing criterion, where B is the area of a restoration result, and Br is the area of restored signal contents.

4. Numerical Experiments

To test this new method, pulsar observations were made with the Nanshan 26 m Radio Telescope, which has a dual-channel cryogenic receiver with a center frequency of 1556 MHz and a simultaneous bandwidth of 320 MHz (Manchester et al. 2001). It applies a digital filter bank (DFB) built by the CSIRO Australia Telescope National Facility with a polyphase filter in field programmable gate array processors (Manchester et al. 2013). The DFB has a maximum bandwidth of 1 GHz and four primary modes of operation (i.e., folding, search, spectrometer, and baseband outputs). For practical timing observation, the DFB is configured to 1024 × 0.5 MHz channels and 8 bit sampling.

4.1. RFI Mitigation

Figure 3 shows PSR J1645-0317 as an example for mitigation and restoration. The number of the zero set channels is 74 with γ = 25, and a is set to 4. The maximum iteration number and the termination tolerance are set to 500 and 10−5, respectively. By visual inspection, after mitigation and channel restoration, flux intensity obtained by the curvelet operator shows a better continuity in frequency distribution, and the missing channels are sufficiently restored. Figure 4 shows comparisons between the different operators applied in the CS-Pulsar test in Figure 3 after mitigation, restoration, and signal retrieval. The standard signal of PSR J1645-0317 is the summation of all observed profiles between 2003 January 3 and 2009 December 31 with the total observing time 78,594 s. Signals y 0 (top left) and y (top right) are normalized to 1, and y 2 (top middle) and other signals are normalized according to their intensity ratios to y . Obviously, y 2 widens the signal and shows false peaks due to the on-pulse RFI. In this test, y can also be considered as a result of the traditional incoherent dedispersion. From middle left to bottom middle: dedispersion results of RFI mitigation (black), mitigation with restoration (blue), and restoration and signal retrieval (red) by the five operators. The PS/N values of the pulses after mitigation are 26.8353, 26.3932, 26.8876, 28.1907, and 30.1723, respectively. Bottom right: dedispersion results of all the restored 74 missing channels by the five operators, and the restoration ratios are 0.1321, 0.1088, 0.1135, 0.1329, and 0.1375, respectively. The purpose is to show the significant effect of the function of signal restoration by visual inspection.

Figure 3.

Figure 3. Mitigation and restoration results for PSR J1645–0317. The number of the zero set channels is 74 with γ = 25, and a is set to 4. The maximum iteration number and the termination tolerance are 500 and 10−5, respectively. Panels from left to right are profile F after preprocessing, profiles processed by CS-Pulsar with the Fourier, Haar, Daubechies, DCT, and curvelet operators before dedispersion, respectively. The restored zero set channels can be clearly seen. Flux in the final panel shows a better continuity in frequency distribution.

Standard image High-resolution image
Figure 4.

Figure 4. Dedispersion results of mitigation, restoration, and signal retrieval for PSR J1645–0317. Three panels of the top show y 0, y 2, and y , respectively. Five panels from the middle left to bottom middle: dedispersion results of only RFI mitigation (black), mitigation and restoration (blue), and restoration with MCA-based retrieval (red) by the Fourier, Haar and Daubechies wavelets, DCT, and curvelet operators. Bottom right: dedispersion results of all the restored 74 missing channels by the five operators.

Standard image High-resolution image

Figure 5 shows the restoration and retrieval results of PSR J0332 + 5434 with the mode of three peaks. The number of the zero set channels is 97 with γ = 80, and a is set to 7. The maximum iteration number and the termination tolerance are set to 1000 and 103, respectively. In cases of multipeaked pulse profiles, the on-pulse RFI will severely disturb the recognition of lower peaks after dedispersion (see Figure 5 left and right top). The main purpose is to show the mitigation efficiency in removing on-pulse RFI leading to false peaks. In the right bottom panel, the false peaks are successfully removed and the multipeaked signal shape is preserved in the process of mitigation and restoration. Another purpose is to demonstrate the retrieval result for the strongest peak. In the middle panel, the removed signal contents of the strongest peak can be reasonably retrieved with the aid of the upper root-mean-square envelope.

Figure 5.

Figure 5. Dedispersion results of channel restoration (by the curvelet operator) and signal retrieval of PSR J0332 + 5434. The number of the zero set channels is 97 with γ = 80, and a is set to 7. The maximum iteration number and the termination tolerance are set to 1000 and 10−3, respectively. The left and middle panels show y and p Cu, respectively. Right top and right bottom: zooming-in results of y with 3 false peaks in red rectangles, and p Cu with the false peaks removed.

Standard image High-resolution image

The restoration ratios with γ = 40 using the curvelet operators for PSRs J1935 + 1616 (B1933+16), J0332 + 5434 (B0329+54), J0814 + 7429 (B0809+74), J0826 + 2637 (B0823+26), and J1136 + 1551 (B1133+16) are 0.0995, 0.1243, 0.0362, 0.0278, and 0.0286, respectively. The purpose is to further reveal the efficiency of CS-based restoration. These values also imply considerable information loss that should have been considered in astrophysical measurements. For a profile, the larger the value of γ is, the less channels are removed and the lower the restoration ratio is, and vice versa. For PSR J0814 + 7429, the corresponding restoration ratios of γ = 30, 40, and 50 are 0.1683, 0.0362, and 0.0104, respectively.

4.2. Pulsar Timing

High-precision timing observations of millisecond-period pulsars have diverse applications, such as testing relativistic gravity (Kramer et al. 2006), investigating binary stellar evolution (Freire et al. 2011), and using pulsar timing arrays (Shannon et al. 2013) to detect gravitational waves emitted from supermassive black hole binaries. Astrophysical measurements require RFI mitigation as a necessary routine. If the CS-Pulsar method is applied to improve timing accuracy, it must not cause issues with timing-related analysis further down the pipeline. It must be proven to not cause systematic biases or underestimated uncertainties. To be an improvement on current methods, it should reduce the timing residuals and the estimated error.

Let v 1 and v 2 (in seconds) be residual vectors of the original profile and a processed one, and e 1 and e 2 (in seconds) be length vectors of their error bars, respectively. A residual vector reveals the goodness of fit of the rotation model of an estimated template to a standard one. The rms of a group of residuals can be applied to judge the respective timing accuracy. The standard templates corresponding to different a and operators are chosen as the summations of the observations without outliers, and the observations only contain residuals distributed like white noise. An error bar judges the estimated error or goodness of fit of the signal shape of a processed signal to a standard one, and its length is directly related to the S/R. The median error (ME) is defined as the median value of the lengths of a group of error bars to reveal the respective estimated error of TOA. In this subsection, the standard template-matching approach (Taylor 1992) is applied as the TOA measurement algorithm, and the toolboxes of PSRCHIVE and TEMPO2 are used.

The profiles of PSR J1935 + 1616 were collected from 2011 to 2014, and γ is set to $\tfrac{6\mathrm{MAD}\left({{\boldsymbol{F}}}_{2}\right)}{0.6745}$ in preprocessing. Each panel of Figure 6 shows a comparison of timing results of two groups of profiles, i.e., the profiles F 0 and P . Parameter a of the soft threshold is selected as 1, 2, 3, 5, 7, and 9, and only the results of a = 1, 3, and 9 are shown in the left, middle, and right panels, respectively. Figure 6 reveals a reasonable trend that the residuals of P are reduced with the increase of a, and the advantage of P over F 0 will be stabilized when a is large. Table 1 shows the values of the assessing criteria. When a is small, e.g., a < 3, the Gaussian noise and RFI still deteriorate the timing results. With the increase of a, the goodness of fit of the rotation model between the estimated and standard templates has been improved, and the processed signals have higher S/N values. In general, timing accuracy benefits from channel restoration and signal retrieval due to recovered key signal details, and the detrimental smoothing effect of mitigation can be largely avoided. Table 2 shows the timing performance of the profiles P by the five operators with a = 5.

Figure 6.

Figure 6. Comparisons of timing results for PSR J1935 + 1616 between the original profile F 0 (black) and P (red) with the curvelet operator. Parameter a of the soft threshold is selected as 1 (left), 3 (middle), and 9 (right), and γ is set to $\tfrac{6\mathrm{MAD}\left({{\boldsymbol{F}}}_{2}\right)}{0.6745}$ in preprocessing. The timing residuals of P will be reduced to a stable level with the increase of a.

Standard image High-resolution image

Table 1. Timing Results of the Criteria (rms in Seconds; ME in Seconds) by the Curvelet Operator with Different Thresholds

a rms( v 2)ME( e 2)
15.4562e–052.4812e–05
24.3316e–052.6609e–05
34.5773e–053.0942e–05
54.4773e–052.8006e–05
74.0210e–052.5388e–05
94.1391e–052.5990e–05

Download table as:  ASCIITypeset image

Table 2. Timing Criteria (rms in Seconds; ME in Seconds) by the Five Operators with a = 5

Operatorsrms( v 2)ME( e 2)
FFT6.6865e–52.3760e–5
Haar6.9740e–52.4354e–5
Daubechies6.6999e–52.5090e–5
DCT6.6912e–52.5090e–5
Curvelet6.6647e–52.4440e–5
rms( v 1) = 7.3655e–5, ME( e 1) = 3.2366

Download table as:  ASCIITypeset image

5. Conclusions

In this work, a compressed sensing method called CS-Pulsar is developed for simultaneous RFI mitigation and signal restoration, and is applied to pulsar timing. This method is novel because it is the first technique to leverage the sparsity of the signals containing the TOA feature for astrophysical purposes. The CS-Pulsar takes the advantage of the reconstruction framework of CS, and applies the regularized LS framework. It reconstructs missing data in removed channels, and then follows up with an MCA-based retrieval with the aid of the upper root-mean-square envelope to retrieve signal contents removed by RFI mitigation. The promising results of the application of CS-pulsar to real data imply that similar signal retrieval methods are highly effective for astrophysical tasks. CS-Pulsar has satisfying restoration efficiency, and shows improvements in model fitting accuracy and signal S/N. This proves that CS-Pulsar can be applied to follow-up timing-related research branches, which we will pursue in future work.

The authors thank the support of National Key R&D Program under grant No. 2018YFA0404602; NSFC under grant Nos. 11673056, 11873080, 12041304, and 11173042; China Scholarship Council with a Certificate No. (or student No.) 201409655001, and Natural Science Foundation of Xinjiang, China under grant No. 2018D01A24. The Nanshan 26 m Radio Telescope is partly supported by the Operation, Maintenance and Upgrading Fund for Astronomical Telescopes and Facility Instruments, budgeted from the Ministry of Finance of China (MOF) and administrated by the Chinese Academy of Sciences (CAS).

Please wait… references are loading.
10.3847/1538-4357/ac8003