Joint Sparse Recovery Using Signal Space Matching Pursuit
Abstract
In this paper, we put forth a new joint sparse recovery algorithm called signal space matching pursuit (SSMP). The key idea of the proposed SSMP algorithm is to sequentially investigate the support of jointly sparse vectors to minimize the subspace distance to the residual space. Our performance guarantee analysis indicates that SSMP accurately reconstructs any row -sparse matrix of rank in the full row rank scenario if the sampling matrix satisfies , which meets the fundamental minimum requirement on to ensure exact recovery. We also show that SSMP guarantees exact reconstruction in at most iterations, provided that satisfies the restricted isometry property (RIP) of order with
where is the number of indices chosen in each iteration. This implies that the requirement on the RIP constant becomes less restrictive when increases. Such behavior seems to be natural but has not been reported for most of conventional methods. We also show that if , then by running more than iterations, the performance guarantee of SSMP can be improved to . Furthermore, we show that under a suitable RIP condition, the reconstruction error of SSMP is upper bounded by a constant multiple of the noise power, which demonstrates the robustness of SSMP to measurement noise. Finally, from extensive numerical experiments, we show that SSMP outperforms conventional joint sparse recovery algorithms both in noiseless and noisy scenarios.
Index Terms:
Joint sparse recovery, multiple measurement vectors (MMV), subspace distance, rank aware order recursive matching pursuit (RA-ORMP), orthogonal least squares (OLS), restricted isometry property (RIP)I Introduction
In recent years, sparse signal recovery has received considerable attention in image processing, seismology, data compression, source localization, wireless communication, machine learning, to name just a few [3, 2, 4, 5]. The main goal of sparse signal recovery is to reconstruct a high dimensional -sparse vector ( where denotes the number of nonzero elements in ) from its compressed linear measurements
(1) |
where () is the sampling (sensing) matrix. In various applications, such as wireless channel estimation [6, 5], sub-Nyquist sampling of multiband signals [7, 8], angles of departure and arrival (AoD and AoA) estimation in mmWave communication systems [9], and brain imaging [10], we encounter a situation where multiple measurement vectors (MMV) of a group of jointly sparse vectors are available. By the jointly sparse vectors, we mean multiple sparse vectors having a common support (the index set of nonzero entries). In this situation, one can dramatically improve reconstruction accuracy by recovering all the desired sparse vectors simultaneously [11, 12]. The problem to reconstruct a group of jointly -sparse vectors111In the sequel, we assume that are linearly independent. is often referred to as the joint sparse recovery problem [17]. Let be the measurement vector of acquired through the sampling matrix . Then the system model describing the MMV can be expressed as
(2) |
where and .
The main task of joint sparse recovery problems is to identify the common support shared by the unknown sparse signals. Once the support is determined accurately, then the system model in (2) can be reduced to independent overdetermined systems and thus the solutions can be found via the conventional least squares (LS) approach. The problem to identify the support can be formulated as
(3) |
where is the submatrix of that contains the columns indexed by and is the subspace spanned by the columns of . A naive way to solve (3) is to search over all possible subspaces spanned by . Such a combinatorial search, however, is exhaustive and thus infeasible for most practical scenarios. Over the years, various approaches to address this problem have been proposed [11, 12, 13, 14, 15, 17, 16]. Roughly speaking, these approaches can be classified into two categories: 1) those based on greedy search principles (e.g., simultaneous orthogonal matching pursuit (SOMP) [13] and rank aware order recursive matching pursuit (RA-ORMP) [16]) and 2) those relying on convex optimization (e.g., mixed norm minimization [14]). Hybrid approaches combining greedy search techniques and conventional methods such as multiple signal classification (MUSIC) have also been proposed (e.g., compressive MUSIC (CS-MUSIC) [15] and subspace-augmented MUSIC (SA-MUSIC) [17]).
In this paper, we put forth a new algorithm, called signal space matching pursuit (SSMP), to improve the quality of joint sparse recovery. Basically, the SSMP algorithm identifies columns of that span the measurement space. By the measurement space, we mean the subspace spanned by the measurement vectors . Towards this end, we recast (3) as
(4) |
In solving this problem, SSMP exploits the notion of subspace distance which measures the closeness between two vector spaces (see Definition 1 in Section II-A) [18]. Specifically, SSMP chooses multiple, say , columns of that minimize the subspace distance to the measurement space in each iteration.
The contributions of this paper are as follows:
-
1)
We propose a new joint sparse recovery algorithm called SSMP (Section II). From the simulation results, we show that SSMP outperforms conventional techniques both in noiseless and noisy scenarios (Section VI). Specifically, in the noiseless scenario, the critical sparsity (the maximum sparsity level at which exact reconstruction is ensured [22]) of SSMP is about 1.5 times higher than those obtained by conventional techniques. In the noisy scenario, SSMP performs close to the oracle least squares (Oracle-LS) estimator222Oracle-LS is the estimator that provides the best achievable bound using prior knowledge on the common support. in terms of the mean square error (MSE) when the signal-to-noise ratio (SNR) is high.
-
2)
We analyze a condition under which the SSMP algorithm exactly reconstructs any group of jointly -sparse vectors in at most iterations in the noiseless scenario (Section III).
-
–
In the full row rank scenario (),333Note that has at most nonzero rows so that . In this sense, we refer to the case where as the full row rank scenario. we show that SSMP accurately recovers any group of jointly -sparse vectors in iterations if the sampling matrix satisfies (Theorem 1(i))
where is the maximum number of columns in that are linearly independent. This implies that under a mild condition on (any columns of are linearly independent), SSMP guarantees exact reconstruction with measurements, which meets the fundamental minimum requirement on the number of measurements to ensure perfect recovery of [16].
-
–
In the rank deficient scenario (), we show that SSMP reconstructs accurately in at most iterations if satisfies the restricted isometry property (RIP [27]) of order with (Theorem 1(ii))
(5) Using the monotonicity of the RIP constant (see Lemma 3), one can notice that the requirement on the RIP constant becomes less restrictive when the number of (linearly independent) measurement vectors increases. This behavior seems to be natural but has not been reported for conventional methods such as SOMP [13] and mixed norm minimization [14]. In particular, if is on the order of , e.g., , then (5) is satisfied under
which implies that SSMP ensures exact recovery with overwhelming probability as long as the number of random measurements scales linearly with [27, 31].
-
–
-
3)
We analyze the performance of SSMP in the scenario where the observation matrix is contaminated by noise (Section IV). Specifically, we show that under a suitable RIP condition, the reconstruction error of SSMP is upper bounded by a constant multiple of the noise power (Theorems 2 and 3), which demonstrates the robustness of SSMP to measurement noise.
-
4)
As a special case, when , we establish the performance guarantee of SSMP running more than iterations (Section V). Specifically, we show that SSMP exactly recovers any -sparse vector in iterations under (Theorem 5)
In contrast to (5), this bound is a constant and unrelated to the sparsity . This implies that even when is not on the order of , SSMP guarantees exact reconstruction with random measurements by running slightly more than iterations.
We briefly summarize the notations used in this paper.
-
•
Let ;
-
•
For , is the cardinality of and is the set of all indices in but not in ;
-
•
For a vector , is the restriction of to the elements indexed by ;
-
•
We refer to having at most nonzero rows as a row -sparse signal and define its support as the index set of its nonzero rows;444For example, if and , then is a row -sparse matrix with support .
-
•
The submatrix of containing the rows indexed by is denoted by ;
We use the following notations for a general matrix .
-
•
The -th column of is denoted by ;
-
•
The -th row of is denoted by ;
-
•
The submatrix of containing the columns indexed by is denoted by ;
-
•
If has full column rank, is the pseudoinverse of where is the transpose of ;
-
•
We define as the column space of ;
-
•
The Frobenius norm and the spectral norm of are denoted by and , respectively;
-
•
The mixed -norm of is denoted by , i.e., ;
-
•
The minimum, maximum, and -th largest singular values of are denoted by , , and , respectively;
-
•
The orthogonal projections onto a subspace and its orthogonal complement are denoted by and , respectively. For simplicity, we write and instead of and , respectively.
II The Proposed SSMP Algorithm
As mentioned, we use the notion of subspace distance in solving our main problem (4). In this section, we briefly introduce the definition of subspace distance and its properties, and then describe the proposed SSMP algorithm.
II-A Preliminaries
We begin with the definition of subspace distance. In a nutshell, the subspace distance is minimized if two subspaces coincide with each other and maximized if two subspaces are perpendicular to each other.
Definition 1 (Subspace distance [18]).
Let and be subspaces in , and let and be orthonormal bases of and , respectively. Then the subspace distance between and is
(6) |
As a special case, suppose and are one-dimensional subspaces in (i.e., and are two lines in ). Further, let and be orthonormal bases of and , respectively, and be the angle between and . Then the subspace distance between and is (see Fig. 1)
(7) |
where (a) is because and are one-dimensional (i.e., in (6)). One can easily see that is maximal when (i.e., ) and if and only if (i.e., ). In fact, is maximized when and are orthogonal and if and only if and coincide with each other [18]. Also, note that the subspace distance is a proper metric between subspaces since it satisfies the following three properties of a metric [18, 19]:
-
(i)
for any subspaces , and the equality holds if and only if .
-
(ii)
for any subspaces .
-
(iii)
for any subspaces .
Proof.
It suffices to show that two constraints and are equivalent. If , then by the property (i) of the subspace distance. Conversely, if , then and thus . ∎
It is worth mentioning that problem (8) can be extended to the noisy scenario by relaxing the constraint to for some properly chosen threshold .
II-B Algorithm Description
The proposed SSMP algorithm solves problem (8) using the greedy principle. Note that the greedy principle has been popularly used in sparse signal recovery for its computational simplicity and competitive performance [20, 21, 22, 23]. In a nutshell, SSMP sequentially investigates the support to minimize the subspace distance to the residual space.
In the first iteration, SSMP chooses multiple, say , columns of the sampling matrix that minimize the subspace distance to the measurement space . Towards this end, SSMP computes the subspace distance
between and its orthogonal projection onto the subspace spanned by each column of . Let , then SSMP chooses (the indices corresponding to the smallest subspace distances). In other words, the estimated support is given by
(9) |
For example, let , , and be the angle between and (see Fig. 2). Also, suppose SSMP picks up two indices in each iteration (i.e., ). Then, by (7), so that (see in Fig. 2) and . After updating the support set , SSMP computes the estimate of the desired signal by solving the LS problem:
Note that and , where is the -dimensional zero matrix. Finally, SSMP generates the residual matrix
which will be used as an observation matrix for the next iteration.
The general iteration of SSMP is similar to the first iteration except for the fact that the residual space is used instead of the measurement space in the support identification. In other words, SSMP identifies the columns of that minimize the subspace distance to the residual space in the -th iteration. Let be the set of indices newly chosen in the -th iteration, then
(10) |
These set of operations are repeated until the iteration number reaches the maximum value 555In the estimation step, SSMP computes . In order to perform this task, should have full column rank and thus . or the pre-defined stopping criterion
is satisfied ( is a stopping threshold).
Suppose SSMP runs iterations in total and more than indices are chosen (i.e., ). Then, by identifying the rows of with largest -norms, the estimated support is pruned to its subset consisting of elements, i.e.,
Finally, SSMP computes the row -sparse estimate of by solving the LS problem on . This pruning step, also known as debiasing, helps to reduce the reconstruction error in the noisy scenario [34].
II-C Refined Identification Rule
As shown in (10), in the -th iteration, SSMP computes the subspace distance between the residual space and its projection space for all remaining indices . Since this operation requires projection operators , it would clearly be computationally prohibitive, especially for large . In view of this, it is of importance to come up with a computationally efficient way to perform the support identification. In the following proposition, we introduce a simplified selection rule under which the number of projections required in each identification step is independent of the signal dimension .
Proposition 1.
Consider the system model in (2). Suppose the SSMP algorithm chooses indices in each iteration. Then the set of indices chosen in the -th iteration satisfies
(11) |
Proof.
See Appendix A. ∎
Note that the selection rule in (11) requires only two projection operators and in each identification step. One can also observe from (11) that SSMP performs the identification step by one simple matrix multiplication followed by the selection of columns with largest -norms. Specifically, if is the -normalized counterpart of , i.e.,
then SSMP picks column indices of with largest -norms. In Algorithm 1, we summarize the proposed SSMP algorithm.
III Exact Joint Sparse Recovery via SSMP
In this section, we analyze a sufficient condition under which SSMP recovers any row -sparse matrix accurately in the noiseless scenario. In this case, we set the stopping threshold to zero. Also, we assume that the number of indices chosen in each iteration satisfies . Then SSMP performs at most iterations before stopping (see Algorithm 1). Finally, we assume that the sampling matrix has unit -norm columns since the performance of SSMP is not affected by the -normalization of columns in .666Note that the term in (11) is not affected by the -normalization of columns in .
III-A Definition and Lemmas
In our analysis, we employ the RIP framework, a widely used tool to analyze sparse recovery algorithms [25, 21, 22, 26, 23, 24].
Definition 2 (RIP [27]).
A matrix is said to satisfy the RIP of order if there exists a constant such that
(12) |
for any -sparse vector . In particular, the minimum value of satisfying (12) is called the RIP constant and denoted by .
We next introduce several lemmas useful in our analysis.
Lemma 2 ([17, Lemma A.2]).
Let and with , then
Lemma 2 implies that if has full column rank, so does the projected matrix . The next lemma describes the monotonicity of the RIP constant.
Lemma 3 ([22, Lemma 1]).
If a matrix satisfies the RIP of orders and , then .
Lemma 4 ([28, Lemma 1]).
Let and . If satisfies the RIP of order , then for any ,
Lemma 4 implies that if satisfies the RIP of order , then the projected matrix obeys the RIP of order and the corresponding RIP constant satisfies .
Recall that the SSMP algorithm picks the indices of the largest elements in in the -th iteration, where is the -normalized counterpart of (see Algorithm 1). This implies that SSMP chooses at least one support element in the -th iteration if and only if the largest element in is larger than the -th largest element in . The following lemma provides a lower bound of .
Lemma 5.
Consider the system model in (2). Let be the estimated support and be the residual generated in the k-th iteration of the SSMP algorithm. Also, let be the -normalized counterpart of . If has full column rank and , then
(13) |
where .
Proof.
See Appendix C. ∎
The following lemmas play an important role in bounding the -th largest element in .
Lemma 6 ([29, Lemma 3]).
Let and . If satisfies the RIP of order , then for any ,
Lemma 7 ([28, Lemma 2]).
Let and with . If satisfies the RIP of order , then for any ,
III-B Performance Guarantee of SSMP
We now analyze a condition of SSMP to guarantee exact reconstruction of any row -sparse signal in iterations. In the sequel, we say that SSMP is successful in the -th iteration if at least one support element is chosen (i.e., ). First, we present a condition under which SSMP chooses at least support elements in the first iterations (i.e., ).
Proposition 2.
Consider the system model in (2), where has unit -norm columns and any nonzero rows of are linearly independent. Let be the number of indices chosen in each iteration of the SSMP algorithm. If satisfies the RIP of order with
(14) |
then SSMP picks at least support elements in the first iterations.
Proof.
We show that for each of . First, we consider the case where . This case is trivial since and thus
Next, we assume that for some integer . In other words, we assume that the SSMP algorithm chooses at least support elements in the first iterations. In particular, we consider the case where , since otherwise . Under this assumption, we show that SSMP picks at least one support element in the -th iteration. As mentioned, SSMP is successful in the -th iteration if and only if the largest element in is larger than the -th largest element in . In our proof, we build a lower bound of and an upper bound of and then show that the former is larger than the latter under (14).
Lower bound of :
Note that . Then, since any nonzero rows of are linearly independent. Also, note that since consists of columns. As a result, we have
(15) |
In addition, since
(16) |
satisfies the RIP of order . Then, by Lemma 5, we have
(17) |
Let , then
(18) |
where (a) is because for each of . Note that satisfies the RIP of order . Then, by Lemma 4, the projected matrix obeys the RIP of order and the corresponding RIP constant satisfies
(19) |
Also, by the definition of the RIP (see Definition 2), we have
(20) |
Finally, by combining (17)-(20), we obtain
(21) |
where the last inequality follows from Lemma 3.
Upper bound of :
We build an upper bound of in two different ways and combine the results.
First, let be the index corresponding to the -th largest element in and . Since , we have
(22) |
Also, since by (16), satisfies the RIP of order and thus
(23) | ||||
(24) |
where (a) and (b) follow from Lemmas 6 and 3, respectively. By combining (22) and (24), we obtain
(25) |
We next obtain an upper bound of in a different way. Since is the -th largest element, we have
(26) |
where (a) is because by Lemma 6, is an orthonormal basis of , and (b) follows from Lemma 7. Also, since is an orthonormal basis of and ,777Note that , where the last equality follows from (15). we have
(27) |
where (a) follows from Lemma 4. Using this together with (26), we have
(28) |
where the last inequality follows from Lemma 3.
When is ?
Thus far, we have shown that SSMP picks at least support elements in the first iterations under (14). We next analyze the performance of SSMP when at least support elements are chosen.
Proposition 3.
Consider the system model in (2), where has unit -norm columns and any nonzero rows of are linearly independent. Let be the number of indices chosen in each iteration of the SSMP algorithm. Suppose SSMP picks at least support elements in the first iterations (i.e., ). If satisfies
(31) |
then SSMP chooses support elements in the -th iteration.
Proof.
In a nutshell, we will show that
(32a) | |||
(32b) |
If this argument holds, then support elements are chosen since SSMP picks the indices of the largest elements in in the -th iteration (see Algorithm 1).
Proof of (32a):
Since and any nonzero rows of are linearly independent, we have
Then, by Lemma 5, we have
(33) |
Also, since for each of , we have
(34) |
By combining (33) and (34), we obtain
which in turn implies (32a).
Proof of (32b):
If for some incorrect index , then
which implies that the matrix does not have full column rank. This is a contradiction since
where (a) and (b) follow from Lemma 2 and (31), respectively.
Therefore, for all of . ∎
We are now ready to establish a sufficient condition for SSMP to guarantee exact reconstruction of any row -sparse matrix.
Theorem 1.
Consider the system model in (2), where has unit -norm columns and any nonzero rows of are linearly independent. Let be the number of indices chosen in each iteration of the SSMP algorithm. Then, SSMP exactly reconstructs from in at most iterations if one of the following conditions is satisfied:
-
(i)
and satisfies
(35) -
(ii)
and satisfies the RIP of order with (14).
Proof.
We show that SSMP picks all support elements in at most iterations under (i) or (ii). It is worth pointing out that even if several incorrect indices are added, SSMP still reconstructs accurately as long as all the support elements are chosen [23, eq. (11)].
Case 1: and satisfies (35).
In this case, it suffices to show that support elements are chosen in the -th iteration for each of .
First, if , then and thus
Therefore, SSMP picks support elements in the first iteration by Proposition 3.
Next, we assume that the argument holds up to (). Then, since
for each of , support elements are chosen in each of the first iterations. In other words, SSMP does not choose an incorrect index until the -th iteration (i.e., ). Thus, we have
and then support elements are chosen in the ()-th iteration by Proposition 3.
Case 2: and satisfies the RIP with (14).
In this case, we have by Proposition 2. In other words, SSMP picks at least support elements during the first iterations. Then, since
,888Note that since satisfies the RIP of order , any columns of are linearly independent. and one can deduce (in a similar way to the case 1) that SSMP picks the rest of support elements by running additional iterations. As a result, SSMP picks all support elements and reconstructs accurately in at most iterations. ∎
Remark 1.
The assumption that any nonzero rows of are linearly independent is fairly mild since it applies to many naturally acquired signals. For example, any random matrix whose entries are drawn i.i.d. from a continuous probability distribution (e.g., Gaussian, uniform, exponential, and chi-square) obeys this assumption [17, 30].
Theorem 1 indicates that SSMP does not require any RIP condition to guarantee exact reconstruction in the full row rank scenario . In [16, Theorem 2], it has been shown that
(36) |
is the fundamental minimum requirement on to ensure exact joint sparse recovery. Combining this with Theorem 1, one can see that SSMP guarantees exact reconstruction with the minimum requirement on in the full row rank scenario. This is in contrast to conventional joint sparse recovery algorithms such as SOMP [13], M-ORMP [11], and mixed norm minimization [14], which require additional conditions on (e.g., null space property) to guarantee exact reconstruction (see [16, Theorems 4 and 5]). In addition, if the sampling matrix satisfies , then SSMP recovers any row -sparse signal accurately with measurements in the full row rank scenario, which meets the fundamental minimum number of measurements to ensure exact joint sparse recovery [16].
Furthermore, one can see that the requirement (14) on the RIP constant becomes less restrictive when the number of (linearly independent) measurement vectors increases, since decreases with and the upper bound in (14) increases with . Such behavior seems to be natural but has not been reported for conventional methods such as SOMP and mixed norm minimization.
In Table I, we summarize performance guarantees of various joint sparse recovery algorithms including SSMP. One can observe that SSMP is very competitive both in full row rank and rank deficient scenarios.
Full Row Rank () | Rank Deficient () | ||
---|---|---|---|
Is sufficient? | Is there any known guarantee that improves with ? | ||
SOMP [13] | No | No | |
M-ORMP [11] | No | No | |
-norm minimization [14] | No | No | |
CS-MUSIC [15] | Yes | No | |
SA-MUSIC [17] | Yes | Yes ( [17]) | |
SSMP | Yes | Yes (see (14)) |
It is well-known that a random matrix whose entries are drawn i.i.d. from a Gaussian distribution satisfies the RIP of order with with overwhelming probability , provided that
(37) |
where is the constant depending on [27, 31]. When combined with Theorem 1, one can notice that SSMP requires a smaller number of (random Gaussian) measurements for exact joint sparse recovery as increases. In particular, if is on the order of , e.g., , then (14) is satisfied under
This implies that SSMP accurately recovers any row -sparse matrix in at most iterations with overwhelming probability as long as the number of random measurements scales linearly with .
We would like to mention that when analyzing the number of measurements ensuring exact joint sparse recovery, probabilistic approaches have been popularly used. For example, it has been shown in [30, Theorem 9], [32], [33, Table I] that if measurement vectors are available, then measurements are sufficient for exact joint sparse recovery. Main benefit of our result, when compared to the previous results, is that it holds uniformly for all sampling matrices and row sparse signals. For example, the result in [30] holds only for a Gaussian or Bernoulli sampling matrix and a fixed row sparse signal that is independent of the sampling matrix. Also, results in [32, 33] are based on an asymptotic analysis where the dimension and sparsity level of a desired sparse signal go to infinity. In contrast, we put our emphasis on the finite-size problem model () so that our result is more realistic and comprehensive.
III-C Connection With Previous Efforts
First, we consider the SSMP algorithm in the single measurement vector (SMV) scenario (i.e., ). By Proposition 1, the set of indices chosen in the -th iteration is
(38) |
where is the residual defined as . One can see that the selection rule of SSMP simplifies to the support identification rule of the multiple orthogonal least squares (MOLS) algorithm when [34, Proposition 1]. In view of this, SSMP can also be considered as an extension of MOLS to the MMV scenario. Moreover, since MOLS reduces to the conventional OLS algorithm when it chooses one index in each iteration [35, 36, 34], SSMP includes OLS as a special case when . Using these connections with Theorem 1, one can establish the performance guarantees of MOLS and OLS, respectively, as follows:
(39a) | |||
(39b) |
In [34], it has been shown that MOLS accurately recovers any -sparse vector in at most iterations under
(40a) | |||
(40b) |
Clearly, the proposed guarantees (39a) and (39b) are less restrictive than (40a) and (40b), respectively. Furthermore, we would like to mention that there exists a -sparse vector that cannot be recovered by OLS running iterations under [37, Example 2], which implies that a sufficient condition of OLS running iterations cannot be less restrictive than
(41) |
One can see that the gap between (39b) and (41) is very small and vanishes for large , which demonstrates the near-optimality of (39b) for OLS running iterations.
Next, we consider the case where the SSMP algorithm picks one index in each iteration (i.e., ). Then the selection rule in (11) simplifies to
(42) |
In this case, SSMP reduces to the RA-ORMP algorithm [16]. Exploiting the relationship between SSMP and RA-ORMP, one can deduce from Theorem 1 that RA-ORMP exactly recovers any row -sparse matrix of rank in iterations under
(43) |
which is consistent with the best known guarantee for RA-ORMP [29]. It is worth mentioning that (43) is a near-optimal recovery condition of RA-ORMP running iterations, since there exists a row -sparse matrix of rank that cannot be recovered by RA-ORMP running iterations under [29, Theorem 2]. In Table II, we summarize the relationship between the proposed SSMP, MOLS, OLS, and RA-ORMP algorithms and their performance guarantees.
IV Robustness of SSMP to Measurement Noise
Thus far, we have focused on the performance guarantee of SSMP in the noiseless scenario. In this section, we analyze the performance of SSMP in the more realistic scenario where the observation matrix is contaminated by noise :
(44) |
Here, the noise matrix is assumed to be bounded (i.e., for some ) or Gaussian. In this paper, we exclusively consider the bounded scenario, but our analysis can be easily extended to the Gaussian noise scenario after small modifications (see [38, Lemma 3]).
In our analysis, we employ the Frobenius norm of the reconstruction error as a performance measure since exact recovery of is not possible in the noisy scenario. Also, we assume that the number of indices chosen in each iteration satisfies . Then SSMP continues to perform an iteration until the iteration number reaches or for some (see Algorithm 1). The following theorem presents an upper bound of when SSMP is terminated by the condition .
Theorem 2.
Consider the system model in (44) where is bounded. Suppose SSMP picks indices in each iteration and for some . Also, suppose satisfies the RIP of order . Then the output of SSMP satisfies
In particular, when , satisfies
(45) |
Proof.
Recall that if SSMP chooses more than indices (i.e., ), then by identifying the rows of with largest -norms, is pruned to its subset consisting of elements (see Algorithm 1). Let be the row -sparse matrix defined as and . Then, one can show that (see Appendix D)
(46) |
and
(47) |
Also, since , we have
(48) |
where (a) follows from (112) in Appendix A. By combining (46)-(48), we obtain the desired result. ∎
Theorem 2 implies that if the SSMP algorithm is terminated by the condition
then is upper bounded by a constant multiple of the noise power , which demonstrates the robustness of SSMP to the measurement noise. One can also deduce from (45) that is recovered accurately (i.e., ) if SSMP finishes before running iterations in the noiseless scenario.
We next consider the case where SSMP finishes after running iterations. In our analysis, we first establish a condition of SSMP choosing all support elements (i.e., ) and then derive an upper bound of the reconstruction error under the obtained condition. The following proposition presents a condition under which SSMP picks at least one support element in the -th iteration.
Proposition 4.
Consider the system model in (44), where has unit -norm columns, any nonzero rows of are linearly independent, and is bounded. Let be the estimated support generated in the -th iteration of the SSMP algorithm and be the number of indices chosen in each iteration. Suppose there exists at least one remaining support element after the -th iteration (i.e., ). Also, suppose satisfies the RIP of order and obeys
(49) |
Then, the following statements hold:
-
(i)
If and
(50) then SSMP chooses at least one support element in the -th iteration.
-
(ii)
If and
(51) then SSMP picks support elements in the -th iteration.
Proof.
We consider the following two cases: 1) and 2) .
-
1)
:
Recall that SSMP chooses at least one support element in the -th iteration if the largest element in is larger than the -th largest element in , where is the -normalized counterpart of (see Algorithm 1). In our proof, we construct a lower bound of and an upper bound of and then show that the former is larger than the latter under (49) and (50).
Lower bound of :
Note that for each of ,
(52) |
where (a) is from the triangle inequality. Thus, satisfies
(53) |
where the last inequality follows from (21).
Upper bound of :
We construct an upper bound of in two different ways and then combine the results.
First, we note that for each of , satisfies
(54) | ||||
(55) |
where (a) follows from the triangle inequality and (b) is from (22) and (23). Therefore, the -th largest element in also satisfies
(56) |
We next derive an upper bound of in a different way. Let be the index corresponding to the -th largest element in , then satisfies
(57) |
where (a) is due to (54), (b) follows from the Cauchy-Schwarz inequality, and (c) is from (26) and (27).
When is ?
(59) |
where . Also, it has been shown in [17, Proposition 7.6] that if the condition number of the matrix obeys999In our case, (60) is satisfied since , where (a) is due to (49) and (b) follows from the definition of the RIP.
(60) |
then satisfies101010In [17, Proposition 7.6], it has been shown that (61) holds when . After small modifications, the proof can readily be extended to the case where .
(61) | ||||
(62) |
where (a) follows from the definition of the RIP. Then, by combining (59) and (62), we have
(63) |
One can easily check that under (50), the right-hand side of (63) is strictly larger than zero. Therefore, , and hence SSMP picks at least one support element (the index corresponding to ) in the -th iteration.
-
2)
:
Proposition 4 indicates that if more than support elements remain after the -th iteration, then SSMP picks at least one support element in the -th iteration under (50). This in turn implies that SSMP chooses at least support elements in the first iterations, provided that the sampling matrix obeys the RIP of order and the corresponding RIP constant satisfies
(64) |
In particular, in the noiseless case (), so that (64) is satisfied under (14). Thus, SSMP picks at least support elements in the first iterations under (14), which coincides with the result in Proposition 2.
The next proposition presents a relationship between and noise.
Proposition 5.
Consider the system model in (44) where is bounded. If , then satisfies
(65) |
Proof.
?
Since
it suffices to show that for any unit vector in . Let be an arbitrary unit vector in , then
(66) |
where (a) follows from the triangle inequality. Note that for any . In particular, when , we have
where the last inequality follows from (66).
?
Let be an arbitrary unit vector in , then . Also, let , then
Since is an arbitrary unit vector in , we have , which is the desired result. ∎
Since , it is clear from Proposition 5 that
One can observe that the upper bound increases with the noise power . In particular, if , then , which in turn implies that the measurement space coincides with the signal space .
Having the results of Propositions 4 and 5 in hand, we are now ready to establish a condition under which SSMP picks all support elements.
Theorem 3.
Consider the system model in (44), where has unit -norm columns, any nonzero rows of are linearly independent, and is bounded. Suppose SSMP chooses indices in each iteration. Also, suppose obeys the RIP of order and the corresponding RIP constant satisfies
(67) |
Then SSMP picks all support elements in at most iterations if one of the following conditions is satisfied:
-
(i)
and satisfies
(68) -
(ii)
and satisfies
(69)
Proof.
By Propositions 4 and 5, the SSMP algorithm chooses at least support elements in the first iterations under (69). Furthermore, similar to the proof of Theorem 1, one can show that if SSMP picks at least support elements in the first iterations, then SSMP chooses the remaining support elements by running additional iterations under (68). Also, using , one can easily show that
and thus (68) is satisfied under (69). By combining these results, we can conclude that SSMP picks all support elements in at most iterations if (i) or (ii) holds. ∎
In the noiseless scenario (), so that (69) is satisfied under (14). Combining this with Theorem 3, one can see that SSMP chooses all support elements and recovers accurately in at most iterations under (14), which is consistent with the result in Theorem 1. One can also infer from Theorem 3 that all support elements are chosen if
(70) |
where
Note that is a decreasing function of and thus the upper bound in (70) also decreases with . Then, since decreases with , the RIP condition in (70) becomes less restrictive when increases. Furthermore, note that
increases with the number of (linearly independent) measurement vectors so that the upper bound in (70) also increases with . Thus, when increases, the requirement on the RIP constant becomes less restrictive (see Fig. 3). This behavior seems to be natural but has not been reported for conventional joint sparse recovery algorithms such as SOMP, M-ORMP, and mixed norm minimization techniques [13, 11, 14, 16]. Moreover, we note that if all support elements are chosen, then the output of SSMP satisfies111111We note that the result in (71) is the extension of the result in [34, eq. (19)], which is related to the SMV version of SSMP, to the MMV scenario. This extension can be easily done by taking similar steps to the proofs of (46) and (47) in Appendix D.
(71) |
This means that the reconstruction error of SSMP is upper bounded by a constant multiple of the noise power , which clearly demonstrates the robustness of the SSMP algorithm to measurement noise.
Finally, it is worth mentioning that our analysis can readily be extended to the scenario where the input signal is approximately row -sparse (a.k.a. row compressible), meaning that for some small .121212 is the matrix obtained from by maintaining the rows with largest -norms and setting all the other rows to the zero vector. In this case, one can establish a condition ensuring the robustness of the SSMP algorithm by 1) partitioning the observation matrix as
(72) |
2) considering as modified measurement noise satisfying (see Appendix E)
(73) |
and then 3) applying Theorems 2 and 3 to the system model in (72).
V SSMP Running More Than Iterations
Thus far, we have analyzed the performance of SSMP running at most iterations. Our result in Theorem 1 implies that if the number of measurement vectors is on the order of (e.g., ), then SSMP recovers any row -sparse matrix accurately with overwhelming probability as long as the number of random measurements scales linearly with . However, if is not on the order of (e.g., ), then the upper bound of the proposed guarantee (14) is inversely proportional to , which requires that should scale with (see (37)).
In the compressed sensing literature, there have been some efforts to improve the performance guarantee by running an algorithm more than iterations [41, 42, 45, 43, 44]. For example, it has been shown in [42, Theorem 6.25] that the optimal performance guarantee of the conventional OMP algorithm running iterations [40] can be relaxed to if it runs iterations. In this section, we show that if , then by running more than iterations, SSMP ensures exact reconstruction with random measurements.
V-A Main Results
In this subsection, we provide our main results. In the next theorem, we demonstrate that if support elements remain after some iterations, then under a suitable RIP condition, SSMP picks the remaining support elements by running a specified number of additional iterations.
Theorem 4.
Consider the SSMP algorithm and the system model in (1) where has unit -norm columns. Let be the number of indices chosen in each iteration and be the number of remaining support elements after the -th iteration. Let be an integer such that . If obeys the RIP of order and the corresponding RIP constant satisfies
(74a) | |||
(74b) | |||
(74c) |
then
(75) |
Proof.
See Section V-B. ∎
Theorem 4 implies that if support elements remain, then under (74a)-(74c), SSMP chooses all these elements by running additional iterations. In particular, when , we obtain the following result.
Theorem 5.
Consider the system model in (1) where has unit -norm columns. Let be the number of indices chosen in each iteration of the SSMP algorithm and be an integer such that . Suppose obeys the RIP of order and the corresponding RIP constant satisfies (74a)-(74c). Then the SSMP algorithm accurately recovers from in iterations.
Note that if , then (74a)-(74c) are satisfied under , , and , respectively. Combining this with Theorem 5, one can see that SSMP ensures exact recovery of any -sparse vector in iterations under
(76) |
In particular, our result indicates that the OLS algorithm, a special case of SSMP for (see Table II), recovers any -sparse vector accurately in iterations under (76). In Table III, we summarize the performance guarantee of SSMP for different choices of .
The beneficial point of (76) is that the upper bound is a constant and unrelated to the sparsity . This implies that by running slightly more than iterations, SSMP accurately recovers any -sparse vector with overwhelming probability with random measurements (see (37)). This is in contrast to the guarantees (39a) and (39b) of SSMP running iterations, which require that the number of random measurements should scale with .
Number of Iterations | Performance Guarantee | ||
---|---|---|---|
V-B Proof of Theorem 4
The proof of Theorem 4 is based on induction on the number of remaining support elements. We note that this proof technique is similar in spirit to the works of Zhang [41] and Foucart and Rauhut [42].
First, we consider the case where . In this case, and thus
Next, we assume that the argument holds up to an integer . In other words, we assume that if the number of remaining support elements is less than , then SSMP chooses all these elements by running additional iterations. Under this assumption, we show that if support elements remain after the -th iteration, then
(77) |
-
1)
Preliminaries
Before we proceed, we define notations used in our analysis. Without loss of generality, let and . We define the subset of as
(78) |
Let
(79) |
then by (74a). Also, let be the integer such that
(80a) | ||||
(80b) | ||||
(80c) | ||||
(80d) |
If (80d) holds for , then we take . We note that always exists, since when so that (80d) holds at least for . From (80a)-(80d), one can easily see that
(81) |
for each of . Also, if , then we have
(82) |
where (a) follows from [45, eq. (21)] and (b) is because by (79). We next provide lemmas useful in our proof.
Lemma 8.
For any integer satisfying and , the residual of the SSMP algorithm satisfies
(83) |
Proof.
See Appendix F. ∎
Lemma 9.
For any integer , , and , the residual generated in the -th iteration of the SSMP algorithm satisfies
(84) |
where
(85) |
Proof.
See Appendix G. ∎
-
2)
Sketch of Proof
We now proceed to the proof of (77). In our proof, we consider the following two cases: i) and ii) .
i) : In this case, one can show that (see justifications in Section V-C)
(86) |
where and
(87) |
This implies that , since otherwise
which is a contradiction to (86). In other words, at most support elements remain after the -th iteration. Then, by the induction hypothesis, SSMP picks the remaining support elements by running additional iterations, i.e.,
(88) |
Note that
(89) |
where (a) is because
Then, by combining (88) and (89), we have
(90) |
Also, note that131313If , then it is trivial. If , then so that .
Using this together with (90), we obtain (77), which completes the proof.
ii) : In this case, one can show that (see justifications in Section V-D)
(91) |
which in turn implies that . In other words, at most support elements remain after the -th iteration. Then, by the induction hypothesis, SSMP picks the remaining support elements by running iterations, i.e.,
(92) |
Also, one can show that141414If , then it is trivial. If , then so that . As a result, .
(93) |
By combining (92) and (93), we obtain (77), which is the desired result.
V-C Proof of (86)
In our proof, we build an upper bound of and a lower bound of and then show that the former is smaller than the latter under (74b).
Upper bound of :
Note that
(94) |
where (a) is from Lemma 4. Also, one can show that (see justifications in Appendix H)
(95) |
and thus the RIP constant of order satisfies
(96) |
by Lemma 3. Using this together with (94), we obtain
(97) |
Lower bound of :
Let and for each of . Then, by applying Lemma 9 with , , and , we have
(98) |
for each of . Note that
(99) |
where (a), (b), and (c) follow from (85), Lemma 3, and (96) respectively. By combining (98) and (99), we have
(100) |
for each of .151515If , then (100) holds since due to the orthogonal projection at each iteration of SSMP and . Now, after taking similar steps to [43, p. 4201], one can show that
which is equivalent to
(101) |
When is ?
V-D Proof of (91)
In our proof, we build an upper bound of and a lower bound of and then show that the former is smaller than the latter under (74c).
Upper bound of :
By taking similar steps to the proof of (97), we can show that
(103) |
Lower bound of :
From Lemma 8, we have
(104) |
where (a) is because and (b) follows from Lemma 3.161616Again, if , then (104) holds since by the orthogonal projection at each iteration. After re-arranging terms, we have
(105) |
Note that
(106) |
where (a) and (b) follow from the RIP and Lemma 3, respectively. Also, note that
(107) |
where (a), (b), and (c) follow from Lemma 3, (81), and (79), respectively. By combining (105)-(107), we have
(108) |
When is ?
VI Simulation Results
In this section, we study the performance of the proposed SSMP algorithm through empirical simulations. In our simulations, we use an sampling matrix () whose entries are drawn i.i.d. from a Gaussian distribution . For each , we generate a row -sparse matrix whose support is uniformly chosen at random. Nonzero entries of are drawn i.i.d. from a standard Gaussian distribution or binary () random distribution. We refer to these two types of signals as the Gaussian signal and the 2-ary pulse amplitude modulation (2-PAM) signal, respectively. In our simulations, the following joint sparse recovery algorithms are considered:
VI-A Noiseless Scenario
In this subsection, we study the empirical performance of SSMP in the noiseless scenario. In this case, the observation matrix follows the system model in (2). We perform independent trials for each point of the algorithm and compute the exact reconstruction ratio (ERR) defined as [22, 23]
By comparing the critical sparsity (the maximum sparsity level at which exact reconstruction is ensured [22]), recovery accuracy of different algorithms can be evaluated.
(a) Gaussian signal,
(b) 2-PAM signal,
First, we study the recovery performance in the full row rank case. In Fig. 4, we plot the ERR performance as a function of sparsity . We can observe that the proposed SSMP algorithm shows a perfect performance (i.e., ERR=1) regardless of the sparsity and the type of a row sparse signal. We also observe that RA-ORMP, which can be viewed as a special case of SSMP for (see Table II), achieves an excellent performance. This is because the simulations are performed in the scenario where , and SSMP guarantees exact reconstruction in this scenario if (see Theorem 1). On the other hand, conventional algorithms such as SOMP, M-ORMP, -norm minimization, and RA-OMP are imperfect when is large, which agrees with the result that these algorithms do not uniformly guarantee exact recovery under (see Table I).
(a) Gaussian signal,
(b) 2-PAM signal,
(c) Gaussian signal,
(d) 2-PAM signal,
(a) Gaussian signal, ,
(b) Gaussian signal, ,
Next, we investigate the recovery performance of SSMP in the rank deficient case (). In Fig. 5, we plot the ERR performance as a function of . In general, we observe that the ERR performance improves with the number of measurement vectors. We also observe that SSMP outperforms other joint sparse recovery algorithms in terms of the critical sparsity for both Gaussian and 2-PAM signals. For example, when and the desired signal is Gaussian, the critical sparsity of SSMP is times higher than those obtained by conventional recovery algorithms (see Fig. 5(a)). In Fig. 6, we plot the ERR performance as a function of the number of measurements. In these simulations, we set the sparsity level to , for which none of recovery algorithms uniformly guarantees exact recovery. Overall, we observe that ERR improves with and the number of measurements required for exact reconstruction decreases with . From Fig. 6(a), we observe that SSMP recovers any row -sparse signal accurately when , while other algorithms require more than measurements. Interestingly, from Fig. 6(b), we observe that SSMP ensures perfect recovery with measurements, which meets the fundamental minimum number of measurements () required for exact joint sparse recovery [16].
(a) Gaussian signal, ,
(b) Gaussian signal, ,
Finally, we study the empirical performance of SSMP in the scenario where the desired signal is approximately row sparse. Recall that is approximately row -sparse if for some small . For an approximately row -sparse signal, we define the support as the index set of the rows with largest -norms. For each , we generate an approximately row -sparse matrix whose support is uniformly chosen at random. The elements of and are drawn i.i.d. from Gaussian distributions and , respectively. In our simulations, is set to so that
As a performance measure for this scenario, we employ the exact support recovery ratio (ESRR):
In Fig. 7, we plot the ESRR performance as a function of when . In general, we observe that the ESRR performance degrades with . In particular, one can see that the ESRR performance of CS-MUSIC degrades severely with . For example, if , then the critical sparsity of CS-MUSIC is 20 (see Fig. 5(c)). However, if , then the ESRR of CS-MUSIC is less than one even when (see Fig. 7(a)). On the other hand, critical sparsities of other algorithms remain the same when increases from to . We also observe that the proposed SSMP algorithm is very competitive in recovering approximately row sparse signals. Specifically, the critical sparsity of SSMP is , which is about times higher than critical sparsities of other approaches.
VI-B Noisy Scenario
In this subsection, we study the empirical performance of SSMP in the noisy scenario. In this case, the observation matrix follows the system model in (44), and we employ the MSE as a performance measure where
For each simulation point of the algorithm, we perform 10,000 independent trials and average the MSE. In our simulations, we set the stopping threshold of SSMP to as in Theorem 2.
(a) Gaussian signal, ,
(b) Gaussian signal, ,
In Fig. 8, we plot the MSE performance as a function of SNR (in dB) which is defined as . In these simulations, the entries of are generated at random from Gaussian distribution .171717One can easily check that , since each component of is generated independently from and is a row -sparse matrix whose nonzero entries are drawn independently from . Then, from the definition of SNR, we have . As a benchmark, we also plot the MSE performance of the Oracle-LS estimator, the best possible estimator using prior knowledge on the support. From the figures, we observe that the MSE performance of SSMP improves linearly with SNR, but the performance improvement of conventional algorithms diminishes with SNR. In particular, SSMP performs close to the Oracle-LS estimator in the high SNR regime ( dB).
VII Conclusion and Discussion
In this paper, we proposed a new joint sparse recovery algorithm called signal space matching pursuit (SSMP) that greatly improves the reconstruction accuracy over conventional techniques. The key idea behind SSMP is to sequentially investigate the support of a row sparse matrix to minimize the subspace distance to the residual space. Our theoretical analysis indicates that under a mild condition on the sampling matrix, SSMP exactly reconstructs any row -sparse matrix of rank using measurements, which meets the fundamental minimum number of measurements to ensure perfect recovery of . We also showed that SSMP guarantees exact reconstruction in at most iterations, provided that satisfies the RIP of order with
(110) |
This implies that the requirement on the RIP constant becomes less restrictive when the number of measurement vectors increases. Such behavior seems to be natural but has not been reported for most of conventional methods. The result in (110) also implies that if is on the order of , then SSMP ensures perfect recovery with overwhelming probability as long as the number of random measurements scales linearly with . We further showed that if , then by running iterations, the guarantee (110) can be significantly improved to . This implies that although is not on the order of , SSMP guarantees exact reconstruction with random measurements by running slightly more than iterations. Moreover, we showed that under a proper RIP condition, the reconstruction error of SSMP is upper bounded by a constant multiple of the noise power, which demonstrates the robustness of SSMP to measurement noise. Finally, from numerical experiments, we confirmed that SSMP outperforms conventional joint sparse recovery algorithms both in noiseless and noisy scenarios.
Finally, we would like to mention some future directions. Firstly, in this work, the number of indices chosen in each iteration is fixed. It would be more flexible and useful if this constraint is relaxed to the variable length. To achieve this goal, a deliberately designed thresholding operation is needed. Secondly, in analyzing a condition under which SSMP chooses support elements in the first iterations, we only considered the scenario where SSMP picks at least one support element in each iteration. One can see that although SSMP fails to choose a support element in some iterations, it can still choose support elements by identifying multiple support elements at a time. It would be an interesting future work to improve the proposed guarantee (110) for this more realistic scenario. Lastly, our result in Theorem 1 implies that if is not on the order of , then SSMP running iterations requires that the number of random measurements should scale with , which is worse than the conventional linear scaling of (). When , we can overcome this limitation by running more than iterations. Extension of our analysis for general to obtain an improved performance guarantee of SSMP is also an interesting research direction.
Appendix A Proof of Proposition 1
Proof.
From (10), the set of indices chosen in the ()-th iteration of SSMP satisfies
(111) |
From the definition of subspace distance, one can show that (see justifications in Appendix B)
(112) |
where is an orthonormal basis of . By combining (111) and (112), we have
(113) |
Also, note that can be decomposed as [36, eq. (12)]
Using this together with (113), we obtain
which is the desired result. ∎
Appendix B Proof of (112)
Appendix C Proof of Lemma 5
Proof.
Note that
(116) |
Let . Then, by Eckart-Young theorem [46], we have
(117) |
By combining (116) and (117), we obtain
(118) |
We now take a look at . Since where , we have and thus
As a result, we have
(119) |
Note that since has full column rank, the projected matrix also has full column rank by Lemma 4, and so does its -normalized counterpart (i.e. ). Also, since has full column rank, . Combining these together with (118) and (119), we obtain the desired result (13). ∎
Appendix D Proofs of (46) and (47)
D-A Proof of (46)
From the triangle inequality, we have
(120) |
Note that since is the set of indices corresponding to the rows of with largest -norms (see Algorithm 1) and is the row -sparse matrix satisfying and , is the best row -sparse approximation of , i.e.,
(121) |
Then, we have . Using this together with (120), we obtain
(122) |
where the last inequality is because obeys the RIP of order and and are row - and -sparse, respectively. Also, since , we have
(123) |
where the last inequality follows from the triangle inequality. By combining (122) and (123), we obtain the desired result in (46).
D-B Proof of (47)
Note that is row -sparse since and are row -sparse matrices. Then, we have
(124) |
where the last inequality follows from the triangle inequality. Also, by noting that (see Algorithm 1) and , we have
(125) |
where (a) follows from the triangle inequality and (b) is because and are row -sparse and thus is a row -sparse matrix. Finally, by combining (124) and (125), we obtain the desired result in (47).
Appendix E Proof of (73)
Proof.
Since by the triangle inequality, it suffices to show that
(126) |
Note that if satisfies the RIP of order , then for any (not necessarily sparse) vector , we have [21, Proposition 3.5]
(127) |
Then, we have
(128) |
where (a) and (b) are from (127) and Minkowski inequality, respectively. Let , then we have
(129) |
where (a) follows from Cauchy-Schwarz inequality. By combining (128) and (129), we obtain (126), which completes the proof. ∎
Appendix F Proof of Lemma 8
Proof.
For any integer such that , we have [45, eq. (C.10)]
(130) |
where is the set of indices chosen in the -th iteration of the SSMP algorithm and (a) is because . Also, note that
(131) |
where (a) and (b) follow from Lemma 6 and (38), respectively, and (c) is because for each of . Furthermore, it has been shown in [45, eq. (C.5)] that
(132) |
By combining (130)-(132), we obtain (83), which is the desired result. ∎
Appendix G Proof of Lemma 9
Proof.
For each of , we have
(133) |
where (a) is from Lemma 8, (b) is because for , and (c) is from Lemma 3.181818If , then (133) clearly holds true since due to the orthogonal projection at each iteration of SSMP. By subtracting both sides of (133) by , we obtain
(134) |
By plugging into (134), we have
(135a) | |||
(135b) | |||
(135c) |
Finally, by combining (135a)-(135c), we obtain (84), which is the desired result. ∎
Appendix H Proof of (95)
References
- [1] J. Kim and B. Shim, “Multiple orthogonal least squares for joint sparse recovery,” in Proc. IEEE Int. Symp. Inform. Theory, Vail, CO, USA, Jun. 2018, pp. 61–65.
- [2] E. J. Candès and M. B. Wakin, “An introduction to compressive sampling,” IEEE Signal Process. Mag., vol. 25, no. 2, pp. 21–30, Mar. 2008.
- [3] D. L. Donoho, “Compressed sensing,” IEEE Trans. Inform. Theory, vol. 52, no. 4, pp. 1289–1306, Apr. 2006.
- [4] V. Papyan, Y. Romano, J. Sulam, and M. Elad, “Theoretical foundations of deep learning via sparse representations: A multilayer sparse model and its connection to convolutional neural networks,” IEEE Signal Process. Mag., vol. 35, no. 4, pp. 72–89, Jul. 2018.
- [5] J. W. Choi, B. Shim, Y. Ding, B. Rao, and D. I. Kim, “Compressed sensing for wireless communications: Useful tips and tricks,” IEEE Commun. Surveys Tuts., vol. 19, no. 3, pp. 1527–1550, 3rd Quart., 2017.
- [6] J. W. Choi and B. Shim, “Statistical recovery of simultaneously sparse time-varying signals from multiple measurement vectors,” IEEE Trans. Signal Process., vol. 63, no. 22, pp. 6136–6148, Nov. 2015.
- [7] P. Feng and Y. Bresler, “Spectrum-blind minimum-rate sampling and reconstruction of multiband signals,” in Proc. IEEE Int. Conf. Acoust., Speech, Signal Process., Atlanta, GA, USA, May 1996, pp. 1668–1691.
- [8] M. Mishali and Y. C. Eldar, “From theory to practice: Sub-Nyquist sampling of sparse wideband analog signals,” IEEE J. Sel. Topics Signal. Process., vol. 4, no. 2, pp. 375–391, Apr. 2010.
- [9] Y. Y. Lee, C. H. Wang, and Y. H. Huang, “A hybrid RF/baseband precoding processor based on parallel-index-selection matrix-inversion-bypass simultaneous orthogonal matching pursuit for millimeter wave MIMO systems,” IEEE Trans. Signal Process., vol. 63, no. 2, pp. 305–317, Jan. 2015.
- [10] O. Lee, J. M. Kim, Y. Bresler, and J. C. Ye, “Compressive diffuse optical tomography: Noniterative exact reconstruction using joint sparsity,” IEEE Trans. Med. Imag., vol. 30, no. 5, pp. 1129–1142, May 2011.
- [11] S. F. Cotter, B. D. Rao, K. Engan, and K. Kreutz-Delgado, “Sparse solutions to linear inverse problems with multiple measurement vectors,” IEEE Trans. Signal Process., vol. 53, no. 7, pp. 2477–2488, Jul. 2005.
- [12] J. Chen and X. Huo, “Theoretical results on sparse representations of multiple-measurement vectors,” IEEE Trans. Signal Process., vol. 54, no. 12, pp. 4634–4643, Dec. 2006.
- [13] J. A. Tropp, A. C. Gilbert, and M. J. Strauss, “Algorithms for simultaneous sparse approximation. part I: Greedy pursuit,” Signal Process., vol. 86, no. 3, pp. 572–588, Mar. 2006.
- [14] J. A. Tropp, “Algorithms for simultaneous sparse approximation. part II: Convex relaxation,” Signal Process., vol. 86, no. 3, pp. 589–602, Mar. 2006.
- [15] J. M. Kim, O. K. Lee, and J. C. Ye, “Compressive MUSIC: Revisiting the link between compressive sensing and array signal processing,” IEEE Trans. Inform. Theory, vol. 58, no. 1, pp. 278–301, Jan. 2012.
- [16] M. E. Davies and Y. C. Eldar, “Rank awareness in joint sparse recovery,” IEEE Trans. Inform. Theory, vol. 58, no. 2, pp. 1135–1146, Feb. 2012.
- [17] K. Lee, Y. Bresler, and M. Junge, “Subspace methods for joint sparse recovery,” IEEE Trans. Inform. Theory, vol. 58, no. 6, pp. 3613–3641, Jun. 2012.
- [18] L. Wang, X. Wang, and J. Feng, “Subspace distance analysis with application to adaptive Bayesian algorithm for face recognition,” Pattern Recognition, vol. 39, no. 3, pp. 456–464, Mar. 2006.
- [19] X. Sun, L. Wang, and J. Feng, “Further results on the subspace distance,” Pattern Recognition, vol. 40, no. 1, pp. 328–329, Jan. 2007.
- [20] Y. C. Pati, R. Rezaiifar, and P. S. Krishnaprasad, “Orthogonal matching pursuit: Recursive function approximation with applications to wavelet decomposition,” in Proc. 27th Annu. Asilomar Conf. Signals, Syst., and Comput., Pacific Grove, CA, USA, Nov. 1993, pp. 40–44.
- [21] D. Needell and J. A. Tropp, “CoSaMP: Iterative signal recovery from incomplete and inaccurate samples,” Appl. Comput. Harmon. Anal., vol. 26, no. 3, pp. 301–321, May 2009.
- [22] W. Dai and O. Milenkovic, “Subspace pursuit for compressive sensing signal reconstruction,” IEEE Trans. Inform. Theory, vol. 55, no. 5, pp. 2230–2249, May 2009.
- [23] J. Wang, S. Kwon, and B. Shim, “Generalized orthogonal matching pursuit,” IEEE Trans. Signal Process., vol. 60, no. 12, pp. 6202–6216, Dec. 2012.
- [24] S. Kwon, J. Wang, and B. Shim, “Multipath matching pursuit,” IEEE Trans. Inform. Theory, vol. 60, no. 5, pp. 2986–3001, May 2014.
- [25] E. J. Candès, “The restricted isometry property and its implications for compressed sensing,” Comptes Rendus Mathematique, vol. 346, no. 9-10, pp. 589–592, May 2008.
- [26] M. A. Davenport and M. B. Wakin, “Analysis of orthogonal matching pursuit using the restricted isometry property,” IEEE Trans. Inform. Theory, vol. 56, no. 9, pp. 4395–4401, Sep. 2010.
- [27] E. J. Candès and T. Tao, “Decoding by linear programming,” IEEE Trans. Inform. Theory, vol. 51, no. 12, pp. 4203–4215, Dec. 2005.
- [28] B. Li, Y. Shen, Z. Wu, and J. Li, “Sufficient conditions for generalized orthogonal matching pursuit in noisy case,” Signal Process., vol. 108, pp. 111–123, Mar. 2015.
- [29] J. Kim and B. Shim, “Nearly sharp restricted isometry condition of rank aware order recursive matching pursuit,” to appear in Proc. IEEE Int. Symp. Inform. Theory, Paris, France, Jul. 2019.
- [30] J. D. Blanchard and M. E. Davies, “Recovery guarantees for rank aware pursuits,” IEEE Signal Process. Lett., vol. 19, no. 7, pp. 427–430, Jul. 2012.
- [31] R. Baraniuk, M. Davenport, R. DeVore, and M. Wakin, “A simple proof of the restricted isometry property for random matrices,” Construct. Approx., vol. 28, no. 3, pp. 253–263, Dec. 2008.
- [32] Y. Jin and B. D. Rao, “Support recovery of sparse signals in the presence of multiple measurement vectors,” IEEE Trans. Inform. Theory, vol. 59, no. 5, pp. 3139–3157, May 2013.
- [33] S. Park, N. Y. Yu, and H. N. Lee, “An information-theoretic study for joint sparsity pattern recovery with different sensing matrices,” IEEE Trans. Inform. Theory, vol. 63, no. 9, pp. 5559–5571, Sept. 2017.
- [34] J. Wang and P. Li, “Recovery of sparse signals using multiple orthogonal least squares,” IEEE Trans. Signal Process., vol. 65, no. 8, pp. 2049–2062, Apr. 2017.
- [35] S. Chen, S. A. Billings, and W. Luo, “Orthogonal least squares methods and their application to non-linear system identification,” Int. J. Control, vol. 50, no. 5, pp. 1873–1896, 1989.
- [36] L. Rebollo-Neira and D. Lowe, “Optimized orthogonal matching pursuit approach,” IEEE Signal Process. Lett., vol. 9, no. 4, pp. 137–140, Apr. 2002.
- [37] J. Wen, J. Wang, and Q. Zhang, “Nearly optimal bounds for orthogonal least squares,” IEEE Trans. Signal Process., vol. 65, no. 20, pp. 5347–5356, Oct. 2017.
- [38] T. T. Cai and L. Wang, “Orthogonal matching pursuit for sparse signal recovery with noise,” IEEE Trans. Inform. Theory, vol. 57, no. 7, pp. 4680–4688, Jul. 2011.
- [39] P. A. Wedin, “On angles between subspaces of a finite dimensional inner product space,” Matrix Pencils, pp. 263–285, 1983.
- [40] Q. Mo, “A sharp restricted isometry constant bound of orthogonal matching pursuit,” arXiv:1501.01708, 2015.
- [41] T. Zhang, “Sparse recovery with orthogonal matching pursuit under RIP,” IEEE Trans. Inform. Theory, vol. 57, no. 9, pp. 6215–6221, Sep. 2011.
- [42] S. Foucart and H. Rauhut, A mathematical introduction to compressive sensing, in Applied and Numerical Harmonic Analysis, Birkhauser, 2013.
- [43] J. Wang and B. Shim, “Exact recovery of sparse signals using orthogonal matching pursuit: How many iterations do we need?,” IEEE Trans. Signal Process., vol. 64, no. 16, pp. 4194–4202, Aug. 2016.
- [44] A. Cohen, W. Dahmen, and R. DeVore, “Orthogonal matching pursuit under the restricted isometry property,” Construct. Approx., vol. 45, no. 1, pp. 113–127, Feb. 2017.
- [45] J. Wang, S. Kwon, P. Li, and B. Shim, “Recovery of sparse signals via generalized orthogonal matching pursuit: A new analysis,” IEEE Trans. Signal Process., vol. 64, no. 4, pp. 1076–1089, Feb. 2016.
- [46] C. Eckart and G. Young, “The approximation of one matrix by another of lower rank,” Psychometrika, vol. 1, no. 3, pp. 211–218, 1936.