Open access peer-reviewed chapter

Sparsity in Bayesian Signal Estimation

Written By

Ishan Wickramasingha, Michael Sobhy and Sherif S. Sherif

Submitted: 29 November 2016 Reviewed: 07 August 2017 Published: 02 November 2017

DOI: 10.5772/intechopen.70529

From the Edited Volume

Bayesian Inference

Edited by Javier Prieto Tejedor

Chapter metrics overview

1,645 Chapter Downloads

View Full Metrics

Abstract

In this chapter, we describe different methods to estimate an unknown signal from its linear measurements. We focus on the underdetermined case where the number of measurements is less than the dimension of the unknown signal. We introduce the concept of signal sparsity and describe how it could be used as prior information for either regularized least squares or Bayesian signal estimation. We discuss compressed sensing and sparse signal representation as examples where these sparse signal estimation methods could be applied.

Keywords

  • inverse problems
  • signal estimation
  • regularization
  • Bayesian methods
  • signal sparsity

1. Introduction

In engineering and science, a system typically refers to a physical process whose outputs are generated due to some inputs [1, 2]. Examples of systems include measuring instruments, imaging devices, mechanical and biomedical devices, chemical reactors and others. A system could be abstracted as a block diagram,

where x and y represent the inputs and outputs of the system, respectively. The block, A, formalizes the relation between these inputs and the outputs using mathematical equations [2, 3]. Depending on the nature of the system, the relation between its inputs and outputs could be either linear or nonlinear. For a linear relation, the system is called a linear system and it would be represented by a set of linear equations [3, 4]

y=Ax.E1

In this chapter, we will restrict our attention to linear systems, as they could adequately represent many actual systems in a mathematically tractable way.

When dealing with systems, two typical types of problems arise, forward and inverse problems.

1.1. Forward problems

In a forward problem, one would be interested in obtaining the output of a system due to a particular input [5, 6]. For linear systems, this output is the result of a simple matrix-vector product, Ax. Forward problems usually become more difficult as the number of equations increases or as uncertainties about the inputs, or the behavior of the system, are present [6].

1.2. Inverse problems

In an inverse problem, one would be interested in inferring the inputs to a system x that resulted in observed outputs, i.e., measured y [5, 6]. Another formulation of an inverse problem is to identify the behavior of the system, i.e., construct A, from knowledge of different input and output values. This problem formulation is known as system identification [1, 7, 8]. In this chapter, we will only consider the input inference problem. The nature of the input x to be inferred further leads to two broad categories of this problem: estimation, and classification. In input estimation, the input could assume an infinite number of possible values [4, 9], while in input classification the input could assume only a finite number (usually small) of possible values [4, 9]. Accordingly, in input classification, one would like to only assign an input to a predetermined signal class. In this chapter, we will only focus on estimation problems, particularly on restoring an input signal x from noisy data y that is obtained using a linear measuring system represented by a matrix A.

Advertisement

2. Signal restoration as example of an inverse problem

To solve the above signal restoration problem, we need to estimate input signal x through the inversion of matrix A. This could be a hard problem because in many cases the inverse of A might not exist, or the measurement data, y, might be corrupted by noise. The existence of the inverse of A depends on the number of acquired independent measurements relative to the dimension of the unknown signal [5, 10]. The conditions for the existence of a stable solution of any inverse problem, i.e., for an inverse problem to be well-posed, have been addressed by Hadamard as:

  • Existence: for measured output y there exists at least one corresponding input x.

  • Uniqueness: for measured output y there exists only one corresponding input x.

  • Continuity: as the input x changes slightly, the output y changes slightly, i.e., the relation between x and y is continuous.

These conditions could be applied to linear systems as conditions on the matrix A. Let the matrix ARn×m, such that Rn×m denotes the set of matrices of dimension n × m with its elements being real values. The matrix equation, yn × 1 = An × m xm × 1, is equivalent to n linear equations with m unknowns. The matrix A is a linear transformation that maps input signals from its domain DA=Rm to its range RA=Rn [4, 5, 10]. For any measured output signal yRn, we could identify three cases based on the values of n and m.

2.1. Underdetermined linear systems

In this case, n < m, i.e., the number of equations is less than the number of unknowns,

A=a11a12a1man1an2anm.E2

If these equations are consistent, Hadamard’s Existence condition will be satisfied. However, Hadamard’s Uniqueness condition is not satisfied because the Null Space(A) ≠ {0}, i.e., there exist z ≠ 0 ∈ Null Space(A) such that,

Ax+z=y.E3

This linear system is called under-determined because its equations, i.e., system constraints, are not enough to uniquely determine x [4, 5]. Thus, the inverse of A does not exist.

2.2. Overdetermined linear systems

In this case, m > n, the number of equations is more than the number of unknowns,

A=a11a21a1ma2man1anm.E4

If these equations are consistent, Hadamard’s Existence condition will not be satisfied. However, Hadamard’s Uniqueness condition will be satisfied, if A has full rank. In this case, Null Space(A) = {0}, i.e.,

Ax+0=Ax=y.E5

This linear system is called over-determined, because its equations, i.e., system constraints, are too many for x to exist [4, 5]. Also, the inverse of A does not exist.

2.3. Square linear systems

The case where m = n, the number of equations is equal to the number of unknowns,

A=a11a1nan1ann.E6

If A has full rank, its Null Space(A) = {0} and both Hadamard’s Existence and Uniqueness conditions will be satisfied. In addition, if A has a small condition number, the relation between x, y will be continuous, and Hadamard’s Continuity condition will be satisfied [4, 5, 10]. In this case, the inverse problem formulated by this system of linear equations is well-posed.

Advertisement

3. Methods for signal estimation

In this section, we will focus on the estimation of an input signal x from a noisy measurement y of the output of a linear system A.

The linear system shown in Figure 1, could be modeled as,

y=Ax+v.E7

Figure 1.

Linear system with noisy output measurement.

where v is additive Gaussian noise. As a consequence of the Central Limit Theorem, this assumption of Gaussian distributed noise is valid for many output measurement setups.

Statistical Estimation Theory allows one to obtain an estimate x̂ of a signal x that is input to a known system A from measurement y (see Figure 2) [11, 12]. However, this estimate x̂ is not unique, as it depends on the choice of the used estimator from the different ones available. In addition to measurement y, if other information about the input signal is available, it could be used as prior information to constrain the estimator to produce a better estimate of x. Signal estimation for overdetermined systems could be achieved without any prior information about the input signal. However, for underdetermined systems, prior information is necessary to ensure a unique estimate.

Figure 2.

Signal estimation using prior information.

3.1. Least squares estimation

If there is no information available about the statistics of the measured data,

y=Ax+v,E8

least squares estimation could be used. The least squares estimate is obtained by minimizing the square of the L2 norm of the error between the measurement and the linear model, v = y − Ax. It is given by

x̂=arg minxy-Ax22.E9

The L2 norm is a special case of the p-norm of a vector, where p = 2, that is defined as xp=i=1mxip1p. In Eq. (9), the unknown x is considered deterministic, so its statistics are not required. The noise v in this formulation is implicitly assumed to be white noise with variance σ2 [13, 14]. Least squares estimation is typically used to estimate input signals x in overdetermined problems. Since x̂ is unique in this case, no prior information, additional constraints, for x is necessary.

3.2. Weighted least squares estimation

If the noise v in Eq. (8) is not necessarily white and its second order statistics, i.e., mean and covariance matrix, are known, then weighted least squares estimation could be used to further improve the least squares estimate. In this estimation method, measurement errors are not weighted equally, but a weighting matrix C explicitly specifies such the weights. The weighted least squares estimate is given by

x̂=arg minxC-1/2y-Ax22.E10

We note that the least squares problem, Eq. (9), is a special case of the weighted least squares problem, Eq. (10), when C = σ2I.

3.3. Regularized least squares estimation

In underdetermined problems, the introduction of additional constraints on x, also known as regularization, could ensure the uniqueness of the obtained solution. Standard least squares estimation could be extended, through regularization, to solve underdetermined estimation problems. The regularized least squares estimate is given by

arg minxy-Ax22+λLx2,E11

where L is a matrix specifying the additional constraints and λ is a regularization parameter whose value determines the relative weights of the two terms in the objective function. If the combined matrix AL has full rank, the regularized least squares estimate x̂ is unique [4]. In this optimization problem, the unknown x is once again considered deterministic, so its statistics are not required. It is worthwhile noting that while regularization is necessary to solve underdetermined inverse problems, it could also be used to improve numerical properties, e.g., condition number, of either linear overdetermined or linear square inverse problems.

3.4. Maximum likelihood estimation

If the probability distribution function (pdf) of the measurement y, parameterized by an unknown deterministic input signal x, is available, then the maximum likelihood estimate of x is given by,

x̂=arg maxxfy|x.E12

This maximum likelihood estimate x̂ is obtained by assuming that measurement y is the most likely measurement to occur given the input signal x. This corresponds to choosing the value of x for which the probability of the observed measurement y is maximized. In maximum likelihood estimation, the negative log of the likelihood function, f(y|x), is typically used to transform Eq. (12) into a simpler minimization problem. When, f(yx) is a Gaussian distribution, N(Ax, C), minimizing the negative log of the likelihood function is equivalent to solving the weighted least squares estimation problem.

3.5. Bayesian estimation

If the conditional pdf of the measurement y, given an unknown random input signal x, is known, in addition to the marginal pdf of x, representing prior information about x, is given, then a Bayesian estimation method would be possible. The first step to obtain one of the many possible Bayesian estimates of x is to use Bayes rule to obtain the a posteriori pdf,

fxy=fy|x)fxfy|x)fx.E13

Once this a posteriori pdf is known, different Bayesian estimates x̂ could be obtained. For example, the minimum mean square error estimate is given by,

x̂MMSE=Exfxy=Exfy|x)fxfy|x)fx,E14

while the maximum a priori (MAP) estimate is given by,

x̂MAP=argmaxxfxy=argmaxxfy|x)fx.E15

We note that the maximum likelihood estimate, Eq. (12), is a special case of the MAP estimate, when f(x) is a uniform pdf over the entire domain of x. The use of prior information is essential to solve underdetermined inverse problems, but it also improves the numerical properties, e.g., condition number, of either linear overdetermined or linear square inverse problems.

3.5.1. Bayesian least squares estimation

In least squares estimation, the vector x is assumed to be an unknown deterministic variable. However, in Bayesian least squares estimation, it is considered a vector of scalar random variables that satisfies statistical properties given by an a priori probability distribution function [5]. In addition, in least squares estimation, the L2 norm of the measurement error is minimized, while in Bayesian least squares estimation, it is the estimation error, e=x̂-x, not measurement error, that is used [5]. Since x is assumed to be a random vector, the estimation error e will also be a random vector. Therefore, the Bayesian least squares estimate could be obtained by minimizing the condtional mean of the square of the estimation error, given measurement, y,

x^=argminxE[ (x^x)T(x^x)y ].E16

When x has a Gaussian distribution and A represents a linear system, then measurement y will also have a Gaussian distribution. In this case, the Bayesian least squares estimate given by Eq. (16) could be reinterpreted as a regularized least squares estimate given by,

x̂=argminxy-Ax22+μ-x,E17

where μ is the mean of the a priori distribution of x [5]. Therefore, a least squares Bayesian estimate is analogous to a regularized least squares estimate, where a priori information about x is expressed as additional constraints on x in the form of a regularization term.

3.5.2. Advantages of Bayesian estimation over other estimation methods

Bayesian estimation techniques could be used, given that a reliable a priori distribution is known, to obtain an accurate estimate of a signal x, even if the number available measurements is smaller than the dimension of the signal to estimated. In this underdetermined case, Bayesian estimation could accurately estimate a signal while un-regularized least squares estimation or maximum likelihood estimation could not. The use of prior information in Bayesian estimation could also improves the numerical properties, e.g., condition number, of either linear overdetermined or linear square inverse problems. This could be understood by keeping in mind the mathematical equivalence between obtaining one scalar measurement related to x, and specifying one constraint that x has to satisfy. Therefore, as the number of available measurements significantly increases, both Bayesian and maximum likelihood estimates would converge to the same estimate.

Bayesian estimation also could be easily adapted to estimate dynamic signals that change over time. This is achieved by sequentially using past estimates of a signal, e.g., xt − 1, as prior information to estimate its current value xt. More generally, Bayesian estimation could be easily adapted for data fusion, i.e., combination of multiple partial measurements to estimate a complete signal in remote sensing, stereo vision and tomographic imaging, e.g., Positron emission tomography (PET), Magnetic resonance imaging (MRI), computed tomography (CT) and optical coherence tomography (OCT). Bayesian methods could also easily fuse all available prior information to provide an estimate based on measurements, in addition to all known information about a signal.

Bayesian estimation techniques could be extended in straight forward ways to estimate output signals of nonlinear systems or signals that have complicated probability distributions. In these cases, numerical Bayesian estimates are typically obtained using Monte Carlo methods.

3.5.3. Sparsity as prior information for underdetermined Bayesian signal estimation

Sparse signal representation means the representation of a signal in a domain where most of its coefficients are zero. Depending on the nature of the signal, one could find an appropriate domain where it would be sparse. This notion could be useful in signal estimation because assuming that the unknown signal x is sparse could be used as prior information to obtain an accurate estimate of it, even if only a small number of measurements are available. The rest of this chapter will focus on using signal sparsity as prior information for underdetermined Bayesian signal estimation.

Advertisement

4. Sparse signal representation

As shown in Figure 3, a sinusoid is a dense signal in the time domain. However, it could be represented by a single value, i.e., it has a sparse representation, in the frequency domain.

Figure 3.

A sinusoid in time and frequency domains.

We note that any signal could have a sparse representation in a suitable domain [15]. A sparse signal representation means a representation of the signal in a domain where most of its coefficients are zero. Sparse signal representations have many advantages including:

  1. A sparse signal representation requires less memory for its storage. Therefore, it is a fundamental concept for signal compression.

  2. A sparse signal representation could lead to simpler signal processing algorithms. For example, signal denoising could be achieved by simple thresholding operations in a domain where the signal is known to be sparse.

  3. Sparse signal representations have fewer coefficients than dense signal representations. Therefore, the computational cost for sparse representations would be lower than for dense representations.

4.1. Signal representation using a dictionary

A dictionary D is a collection of vectors {φn}Γ, indexed by a parameter n ϵ Γ equal to the dimension of a signal f, where we could represent f as a linear combination [16],

f=nϵΓcnφn.E18

If the vectors {φn}Γ are linearly independent, then such dictionary is called a basis. Representing a signal as a linear combination of sinusoids, i.e., using a Fourier dictionary, is very common. Wavelet dictionaries and Chirplet dictionaries are also common dictionaries for signal representation. Dictionaries could be combined together to obtain a larger dictionary, where n ϵ Γ is larger than the dimension the signal f, that is called an overcomplete dictionary or a frame.

4.1.1. Signal representation using a basis

A set of vectors form a basis for Rn if they span Rn and are linearly independent. A basis in a vector space V is a set X of linearly independent vectors such that every vector in V is a linear combination of elements in X. A vector space V is finite dimensional if it has a finite number of basis vectors [17].

Depending on the properties of {φn}Γ, bases could be classified into different types, e.g., orthogonal basis, orthonormal basis, biorthogonal basis, global basis and local basis. For an orthogonal basis, its basis vectors in the vector space V are mutually orthogonal,

φmφn=0formn.E19

For an orthonormal basis, its basis vectors in the vector space V are mutually orthogonal and have unit length,

φmφn=δm-n,E20

where δ(m − n) is the Kronecker delta function. For a biorthogonal basis, its basis vectors are not orthogonal to each other, but they are orthogonal to vectors in another basis, φ̃nnϵΓ, such that

φmφ̃n=δm-n.E21

In addition, depending on the domain (support) on which these basis vectors are defined, we could also classify a basis as either global or local. Sinusoidal basis vectors used for the discrete Fourier transform are defined on the entire domain (support) of f, so they are considered a global basis. Many wavelet basis vectors used for the discrete wavelet transform are defined on only part of the domain (support) of f, so they are considered a local basis.

4.1.2. Signal representation using a frame

A frame is a set of vectors {φn}Γ that spans Rn and could be used to represent a signal f from the inner products {⟨fφn⟩}Γ. A frame allows the representation of a signal as a set of frame coefficients, and its reconstruction from these coefficients in a numerically stable way

f=nϵΓfφnφn.E22

Frame theory analyzes the completeness, stability, and redundancy of linear discrete signal representations [18]. A frame is not necessarily a basis, but it shares many properties with bases. The most important distinction between a frame and a basis is that the vectors that comprise a basis are linearly independent, while those comprising frame could be linearly dependent. Frames are also called overcomplete dictionaries. The redundancy in the representation of a signal using frames could be used to obtain sparse signal representations.

4.2. Sparse signal representation as a regularized least squares estimation problem

If designed to concentrate the energy of a signal in a small number of dimensions, an orthogonal basis would be the minimum-size dictionary that could yield a sparse representation of this signal [15]. However, finding an orthogonal basis that yields a highly sparse representation for a given signal is usually difficult or impractical. To allow more flexibility, the orthogonality constraint is usually dropped, and overcomplete dictionaries (frames) are usually used. This idea is well explained in the following quote by Stephane Mallat:

“In natural languages, a richer dictionary helps to build shorter and more precise sentences. Similarly, dictionaries of vectors that are larger than bases are needed to build sparse representations of complex signals. Sparse representations in redundant dictionaries can improve pattern recognition, compression, and noise reduction but also the resolution of new inverse problems. This includes super resolution, source separation, and compressed sensing” [15].

Thus representing a signal using a particular overcomplete dictionary has the following goals [16]

  • Sparsity—this representation should be more sparse than other representations.

  • Super resolution—the resolution of the signal when represented using this dictionary should be higher than when represented in any other dictionary.

  • Speed—this representation should be computed in O(n) or O(n log(n)) time.

A simple way to obtain an overcomplete dictionary A is to use a union of basis Ai that would result in the following representation of a signal y,

(y)=([ A1 ][ A2 ][ A3 ][ A4 ][ A5 ])A(x)y=Ax,E23

where A is a n × m matrix representing the dictionary and x are the coefficients representing y in the domain defined by A. Since A represents an overcomplete dictionary, the number of its rows will be less than the number of its columns. Eq. (23) is a formulation of the signal representation problem as an underdetermined inverse problem.

To obtain a sparse solution for Eq. (23) one needs to find an m × 1 coefficient vector x̂, such that,

x̂=argminxy-Ax22+λx0,E24

where ‖x0 is the cardinality of vector x, i.e., its number of nonzero elements, and λ > 0 is a regularization parameter that quantifies the tradeoff between the signal representation error,y-Ax22, and its sparsity level, ‖x0 [19]. The cardinality of vector x is sometimes referred to as the L0 norm of x, even though ‖x0 is actually a pseudo norm that does not satisfy the requirements of a norm in Rm. This sparse signal representation problem, Eq. (24), has a form similar to the regularized least squares estimation problem, Eq. (11), that would be underdetermined in the case of an overcomplete dictionary. Because of the correspondence between regularized least squares estimation and Bayesian estimation, the problem of finding a sparse representation of a signal could be formulated as a Bayesian estimation problem.

Advertisement

5. Compressed sensing

Compressed sensing involves the estimation of a signal using a number of measurements that are significantly less than its dimension [20]. By assuming that the unknown signal is sparse in the domain where the measurements were acquired, one could use this sparsity constraint as prior information to obtain an accurate estimate of the signal from relatively few measurements.

Compressed sensing is closely related to signal compression that is routinely used for efficient storage or transmission of signals. Compressed sensing was inspired by this question: instead of the typical signal acquisition followed by signal compression, is there a way to acquire (sense) the compressed signal in the first place? If possible, it would significantly reduce the number of measurements and the computation cost [20]. In addition, this possibility would allow acquisition of signals that require extremely high, hence impractical, sampling rates [21]. As an affirmative answer to this question, compressed sensing was developed to combine signal compression with signal acquisition [20]. This is achieved by designing the measurement setup to acquire signals in the domain where the unknown signal is assumed to be sparse.

In compressed sensing, we consider the estimation of an input signal xRn from m linear measurements, where m ≪ n. As discussed above, this problem could be written as an underdetermined linear system,

y=Ax,E25

where yRm and ARm×n represent the measurements and measurement (sensing) matrix, respectively.

Assuming that the unknown signal x is s-sparse, i.e., x ∈ ∑s has only s nonzero elements, in the domain specified by the measurement (sensing) matrix A, and assuming that A satisfies the restricted isometry property (RIP) of order 2s, i.e., there exists a constant δ2s ∈ (0, 1) such that,

1-δ2sz22Az221+δ2sz22,E26

for all z ∈ ∑2s, then x could be reconstructed from m ≥ s measurements by different optimization algorithms [20]. When the measurements y are noiseless, x could be exactly estimated from,

minxx0subjecttoAx=y.E27

However, when the measurements y are contaminated by noise,x could be obtained as the regularized least squares estimate,

x̂=argminxAx-y22+λx0.E28

This minimization problem could also be mathematically reformulated and solved as a Bayesian estimation problem.

Advertisement

6. Obtaining sparse solutions for signal representation and signal estimation problems

From Sections 4 and 5 we note that the problem of obtaining a sparse signal representation, Eq. (24) and the problem of sparse signal estimation in compressed sensing, Eq. (28), both have the same mathematical form [11, 22],

x̂=argminxy-Ax22+λx0.E29

In this section, we describe different approaches to solving this minimization problem. From Eq. (29), we note that the first term of its RHS, y-Ax22, represents either signal reconstruction error (sparse signal representation problem) or measurement fitting error (sparse signal estimation in compressed sensing problem), while the second term of its RHS,‖x0, represents the cardinality (number of nonzero coefficients) of the unknown signal. The regularization parameter λ specifies the tradeoff between these two terms in the objective function. The selection of an appropriate value of λ to balance the reconstruction, or fitting error, and signal sparsity is very important. Regularization theory and Bayesian approaches could provide ways to determine optimal values of λ [2326].

Convex optimization problems is a class of optimization problems that are significantly easier to solve compared to nonconvex problems [34]. Another advantage of convex optimization problems is that any local solution, e.g., a local minimum, is guaranteed to be a global solution. We note that obtaining an exact solution for the minimization problem in Eq. (29) is difficult because it is nonconvex. Therefore, one could either seek an approximate solution to this nonconvex problem or approximate this problem by a convex optimization whose exact solution could be obtained easily.

Considering the general regularized least squares estimation problem,

x̂=argminxy-Ax22+λxp,E30

we note that it is a nonconvex optimization problem for 0 ≤ p < 1 and a convex optimization problem for p ≥ 1. One alternative to approximate Eq. (29) by a convex optimization problem, one could relax the strict condition of minimizing the cardinality of the signal, ‖x0, by replacing by it by the sparsity-promoting condition of minimizing the L1 norm of the signal, ‖x1. Another alternative to approximate Eq. (29) by another nonconvex optimization problem that is easier to solve than the original problem using a Bayesian formulation, is to replace ‖x0 by ‖xp, 0 < p < 1. The minimization of Eq. (30) using ‖xp, 0 < p < 1 would result in a higher degree of signal sparsity compared to when ‖x1 is used. This could be understood visually by examining Figure 4, that shows the shapes of two-dimensional unit balls using (pseudo)norms with different values of p.

Figure 4.

Two-dimensional unit ball using different (pseudo)norms. (a) L0, (b) L0−1, and (c) L1.

We explain further details in the following subsections.

6.1. Obtaining a sparse signal solution using L0 minimization

The sparsest solution of the regularized least squares estimation problem, Eq. (29) would be obtained when p = 0 in ‖xp. As shown in Figure 5, the solution of the regularized least squares problem, x̂, is given by the intersection of the circles, possibly ellipses, representing the solution of the unconstrained least squares estimation problem and the unit ball using L0 representing the constraint of minimizing L0. In this case of minimizing L0, the unconstrained least squares solution will always intersect the unit ball at an axis, this yielding the most possible sparse solution. However, as mentioned earlier, this L0 minimization problem is difficult to solve because it is nonconvex. Approximate solutions for this problem could be obtained using greedy optimization algorithms, e.g., Matching Pursuits [27] and Least Angle Regression (LARS) [28].

Figure 5.

Regularized least squares using L0.

6.2. Obtaining a sparse signal solution using L1 minimization

On relaxing the nonconvex regularized least squares using L0 minimization problem, by setting p = 1, we obtain the convex L1 minimization problem. As shown in Figure 4(c), the unit ball using the L1 norm covers a larger area than the unit ball using the L0 pseudo norm, shown in Figure 4(a). Therefore, as shown in Figure 6, the solution for the regularized least squares problem using the L1 minimization would be sparse, but it should not be expected to be as sparse as the L0 minimization problem.

Figure 6.

Regularized least squares using L1.

This L1 minimization problem could be solved easily using various algorithms, e.g., Basis Pursuits [16], Method of frames (MOF) [29], Lasso [30, 31], and Best Basis Selection [32, 33]. A Bayesian formulation of this L1 minimization problem is also possible by assuming that the a priori probability distribution of x is Laplacian, x ~ e− |x|.

6.3. Obtaining a sparse signal solution using L0 − 1 minimization

As discussed above, solving the regularized least squares problem with L0 minimization should yield the sparsest signal solution. However, only approximate solutions are available for this difficult nonconvex problem. Alternatively, solving the regularized least squares problem with L1 minimization should yield an exact sparse solution that would be less sparse than in the L0 case, but it is considerably easier to obtain.

The regularized least squares problem could also be formulated as an L0 − 1 minimization problem. As ‖xp, 0 < p < 1,that we abbreviate as L0 − 1, is not an actual norm, this optimization problem would be nonconvex [34]. The advantage of using L0 − 1 minimization is that, as shown in Figure 4(b), compared to unit ball using the L1 norm, the unit ball using the L0 − 1 pseudo norm has a narrower area that is concentrated around the axes. Therefore, as shown in Figure 7, the L0 − 1 minimization problem should yield a sparser solution compared to the L1 minimization problem.

Figure 7.

Regularized least squares using L0−1.

Another advantage of using L0 − 1 minimization is that this nonconvex optimization problem could be easily formulated as a Bayesian estimation problem that could be solved using Markov Chain Monte Carlo (MCMC) methods. As shown in Figure 8, the product of student-t probability distributions has a shape similar to the unit ball using the L0 − 1 pseudo norm, so student-t distributions could be used as a priori distributions to approximate the L0 − 1 pseudo norm.

Figure 8.

Product of two student-t probability distributions.

6.4. Bayesian method to obtain a sparse signal solution using L0 − 1 minimization

As mentioned in Section 3.5, the first step to obtaining one of the many possible Bayesian estimates of x is to use Bayes rule to obtain the a posteriori pdf,

fxy=fy|x)fxfy|x)fx.E31

Using this a posteriori distribution, one could obtain a sparse signal solution using L0 − 1 minimization, as the maximum a posteriori (MAP) estimate given by Eq. (15). Compared to other Bayesian estimates, the MAP estimate could be easier to obtain because the calculation of the normalizing constant, ∫f(y|x)f(x),would not be needed. The maximization of the product of conditional probability distribution of y given x and the a priori distribution of x is equivalent to the minimizing of the sum of their negative logarithms,

x̂MAP=arg minx-logpy|x-logpx.E32

In the case of white Gaussian measurement noise, p(y|x) ~ Nx (Axσ2I) where -logpy|xy-Ax22, which the first term of the RHS of Eq. (30). As discussed in the previous section, the a priori probability p(x) corresponding to L0 − 1 minimization could be represented as a product of univariate student-t probability distribution functions [14],

px=i=1Mstudxi01ϑ=i=1MΓϑ+12ϑπΓϑ21+xi2ϑ-ϑ+12,E33

where Γ is the Gamma function, and ϑ is the number of degrees of freedom of the student-t distribution. Since this a priori distribution function is not an exponential function, we would use Eq. (15) instead of Eq. (32) to obtain the MAP estimate.

Because the prior is not a Gaussian distribution, there is no simple closed form expression for the posterior, p(x|y) with a student-t a priori probability distribution. However, we could express each student-t distribution as an infinite weighted sum of Gaussian distributions, where the hidden variables hi determine their variances [14].

px=i=1MNxi0,1/hiGamhiϑ/2,ϑ/2dhi=Nx0H-1i=1MGamhiϑ/2,ϑ/2dH,E34

where the matrix H contains the hidden variables hii=1M on its diagonal and has zeros elsewhere, and Gamhiϑ/2,ϑ/2 is the gamma probability distribution function with parameters (ϑ/2, ϑ/2). Using this approximation, the a posteriori pdf could be written as

px|ypy|xpx=NxAx,σ2INx0H-1i=1MGamhiϑ2ϑ2dH=NxAx,σ2INx0H-1i=1MGamhiϑ2ϑ2dH.E35

The product of two Gaussian distributions is also a Gaussian distribution [35],

Nxμ1Σ1Nxμ2Σ2=k.NxμΣ,E36

where the mean and covariance (μ, Σ) of the new Gaussian distribution in Eq. (36) is given by,

μ=Σ1-1+Σ2-1-1Σ1-1μ1+Σ2-1μ2andΣ=Σ1-1+Σ2-1-1,E37

and k is a constant. Therefore, we could simplify the product of two the Gaussian distributions given in Eq. (35) as,

NxAx,σ2I.Nx0H-1=k.Nxσ-2I+H-1σ-2Ax,σ-2I+H-1.E38

From Eqs. (35) and (38) we could write p(x|y) as,

px|y=kNxσ-2I+H-1σ-2Ax,σ-2I+H-1i=1MGamhiϑ2ϑ2dH.E39

We still could not compute the integral in Eq. (39) in closed form. However, we could maximize the RHS of Eq. (39) over the hidden variables H to obtain an approximation for the a posteriori probability distribution function

px|yarg maxHNxσ-2I+H-1σ-2Ax,σ-2I+H-1i=1MGamhiϑ2ϑ2.E40

Eq. (40) would be a good approximation of p(x|y), if the actual distribution over the hidden variables is concentrated tightly around its mode [14]. When hi has a large value, its corresponding ith component of the a priori probability distribution function p(x) would have a small variance, 1hi, so that this ith component of p(x) could be set to zero. Therefore, this ith dimension of the prior p(x) would not contribute to the solution of Eq. (30), thus increasing its sparsity.

Since both Gaussian and gamma pdfs in Eq. (40) are members of the exponential family of probability distributions, we could obtain x̂MAP by maximizing the sum of their logarithms. Section 3.5 in [11] and Section 8.6 in [14] describe an iterative optimization method to obtain x̂MAP from the approximate a posteriori probability distribution function given by Eq. (40).

Advertisement

7. Conclusion

In this chapter, we described different methods to estimate an unknown signal from its linear measurements. We focused on the underdetermined case where the number of measurements is less than the dimension of the unknown signal. We introduced the concept of signal sparsity and described how it could be used as prior information for either regularized least squares or Bayesian signal estimation. We discussed compressed sensing and sparse signal representation as examples where these sparse signal estimation methods could be applied.

References

  1. 1. Keesman KJ. System Identification: An Introduction. London: Springer Science & Business Media; 2011
  2. 2. Von Bertalanffy L. General system theory. New York. 1968;41973(1968):40
  3. 3. Chen C-T. Linear System Theory and Design. New York, NY: Oxford University Press, Inc.; 1999
  4. 4. Moon TK, Stirling WC. Mathematical Methods and Algorithms for Signal Processing. Upper Saddle River, NJ: Prentice Hall; 2000
  5. 5. Fieguth P. Statistical Image Processing and Multidimensional Modeling. New York, NY: Springer Science+Business Media, LLC; 2011
  6. 6. Tarantola A. Inverse problem theory and methods for model parameter estimation. Philadelphia, PA: Society for Industrial and Applied Mathematics; 2005. p. 1-37
  7. 7. Ljung L. Perspectives on system identification. Annual Reviews in Control. 2010 Apr 30;34(1):1-2
  8. 8. Wellstead PE. Non-parametric methods of system identification. Automatica. 1981 Jan 1;17(1):55-69
  9. 9. Shanmugan KS, Breipohl AM. Random Signals: Detection, Estimation, and Data Analysis. New York, NY: Wiley; 1997
  10. 10. Saad Y. Iterative Methods for Sparse Linear Systems. Philadelphia, PA: Society for Industrial and Applied Mathematics; 2003.
  11. 11. Bishop CM. Pattern Recognition and Machine Learning. New York, NY: Springer; 2006
  12. 12. Mendel JM. Lessons in Estimation Theory for Signal Processing, Communications, and Control. Englewood Cliffs, N.J.: Prentice-Hall; 1995
  13. 13. Sorenson HW. Least-squares estimation: From Gauss to Kalman. IEEE Spectrum. 1970 Jul;7(7):63-68
  14. 14. Prince SJ. Computer Vision: Models, Learning, and Inference. Cambridge: Cambridge University Press; 2012
  15. 15. Mallat S. A Wavelet Tour of Signal Processing: The Sparse Way. Amsterdam: Academic Press; 2009
  16. 16. Chen SS, Donoho DL, Saunders MA. Atomic decomposition by basis pursuit. SIAM Review. 2001;43(1):129-159
  17. 17. Paul R. Halmos. Finite-Dimensional Vector Spaces. Mineola, UNITED STATES: Dover Publications; 2017
  18. 18. Mallat S. A Wavelet Tour of Signal Processing. San Diego: Academic Press; 1999
  19. 19. Shannon CE. A mathematical theory of communication. ACM SIGMOBILE Mobile Computing and Communications Review. 2001;5(1):3-55
  20. 20. Eldar YC, Kutyniok G, editors. Compressed Sensing: Theory and Applications. Cambridge: Cambridge University Press; 2012
  21. 21. Asif MS. Dynamic compressive sensing: Sparse recovery algorithms for streaming signals and video [Doctoral dissertation]. Georgia Institute of Technology
  22. 22. Huang K, Aviyente S. Sparse representation for signal classification. In: NIPS. Vol. 19; 2006. pp. 609-616
  23. 23. Poggio T, Torre V, Koch C. Computational vision and regularization theory. Nature. 1985 Sep 26;317(6035):314-319
  24. 24. Tikhonov AN, Arsenin VI. Solutions of Ill-posed Problems. Washington, DC: Winston; 1977 Jan
  25. 25. Wahba G, Wendelberger J. Some new mathematical methods for variational objective analysis using splines and cross validation. Monthly Weather Review. 1980 Aug;108(8):1122-1143
  26. 26. Lin Y, Lee DD. Bayesian L1-Norm Sparse Learning. In: 2006 IEEE International Conference on Acoustics Speech and Signal Processing Proceedings. Toulouse, France: vol. 5; 2006. p. V?V.
  27. 27. Mallat SG, Zhang Z. Matching pursuits with time-frequency dictionaries. IEEE Transactions on Signal Processing. 1993 Dec;41(12):3397-3415
  28. 28. Efron B, Hastie T, Johnstone I, Tibshirani R. Least angle regression. The Annals of Statistics. 2004 Apr;32(2):407-499
  29. 29. Daubechies I. Time-frequency localization operators: A geometric phase space approach. IEEE Transactions on Information Theory. 1988 Jul;34(4):605-612
  30. 30. Tibshirani R. Regression shrinkage and selection via the lasso. Journal of the Royal Statistical Society. Series B (Methodological). 1996 Jan 1;58(1):267?288
  31. 31. Yuan M, Lin Y. Model selection and estimation in regression with grouped variables. Journal of the Royal Statistical Society: Series B (Statistical Methodology). 2006 Feb 1;68(1):49-67
  32. 32. Coifman RR, Wickerhauser MV. Entropy-based algorithms for best basis selection. IEEE Transactions on Information Theory. 1992 Mar;38(2):713-718
  33. 33. Rao BD, Kreutz-Delgado K. An affine scaling methodology for best basis selection. IEEE Transactions on Signal Processing. 1999 Jan;47(1):187-200
  34. 34. Boyd S, Vandenberghe L. Convex Optimization. New York: Cambridge University Press; 2004.
  35. 35. Bromiley P. Products and convolutions of Gaussian probability density functions. Tina-Vision Memo. 2003;3(4):1?13.

Written By

Ishan Wickramasingha, Michael Sobhy and Sherif S. Sherif

Submitted: 29 November 2016 Reviewed: 07 August 2017 Published: 02 November 2017