This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

From Signal Space To STP-CS

Daizhan Cheng This work is supported partly by the National Natural Science Foundation of China (NSFC) under Grants 62073315.
Abstract

Under the assumption that a finite signal with different sampling lengths or different sampling frequencies is considered as equivalent signals, the signal space is considered as the quotient space of {\mathbb{R}}^{\infty} over equivalence. The topological structure and the properties of signal space are investigated. Using them some characteristics of semi-tensor product based compressed sensing (STP-CS) are revealed. Finally, a systematic analysis of the construction of sensing matrix based on balanced incomplete block design (BIBD) is presented.

Index Terms:
Semi-tensor product (STP), compressed sensing (CS), sensing matrix, lossless (lossy) compression/dicompression.

I Introduction

The compressed sensing (CS), as a new technique for signal processing, was initially proposed by Donoho, Candes, Tao et al. in 2006 [18, 4]. The basic idea of CS is to compress a signal xx of dimension nn through a matrix Am×nA\in{\mathcal{M}}_{m\times n}, m<<nm<<n, called the sensing matrix, to get a sampled data yRmy\in R^{m}. That is,

y=Ax.\displaystyle y=Ax. (1)

The sensing matrix is obtained as follows: Assume the original signal is θn\theta\in{\mathbb{R}}^{n}, and the sampled data is y=Φθmy=\Phi\theta\in{\mathbb{R}}^{m}. Then CS tries to design Ψ\Psi such that θ=Ψx\theta=\Psi x. where xx is with limited number nonzero entries. Then the sensing matrix is obtained as A=ΦΨA=\Phi\Psi. Since xx is constructed in such a way, it is possible to recover xx from sampled date yy [5, 6].

A fundamental characteristic is the spark of AA, which is the smallest number of the dependent columns of AA. Then the follows is a criterion to justify if the original signal can be recovered from sampled data.

Denote by w(x)w(x) the number of nonzero entries in vector xx, and Σkn\Sigma^{n}_{k} the set of xnx\in{\mathbb{R}}^{n} with w(x)kw(x)\leq k. Then we have the following result.

Proposition I.1

[17] Consider equation (1). Assume spark(A)>2kspark(A)>2k, then the equation has at most one solution xΣknx\in\Sigma^{n}_{k}.

Hence the CS broken the Nyquist sampling theorem, which says that to avoid aliasing the sample frequency should be greater than twice of signal highest frequency [21].

Construct the sensing matrix is one of the fundamental issues in CS. It has been investigated by many authors. For instance, [16, 15] proposed a technique to construct a structurally random matrix for the sensing matrix. Many deterministic sensing matrices have been constructed [1, 33, 36]. To reduce the storage space of the sensing matrices, many methods have been developed, including orthogonal bases [20], low-rank matrices [26, 3], etc.

Semi-tensor product (STP) of matrices/vectors is a generalization of conventional matrix/vector product, which removes the dimension restriction on factor matrices/vectors of the conventional one[10]. The STP has firstly been used to CS by the pioneer work of [32], where the technique is named as STP-CS. The basic idea for STP-CS is to replace fundamental CS-model (1) by

y=Ax(=(A0Is)x),\displaystyle y=A\ltimes x(=(A_{0}\otimes I_{s})x), (2)

where A0m0×n0A_{0}\in{\mathcal{M}}_{m_{0}\times n_{0}}, m=m0sm=m_{0}s and n=n0sn=n_{0}s.

There are two obvious advantages of STP-CS: (i) It can reduce the storage of sensing matrix; (ii) it can be used for all nn dimensional signals as long as m|nm|n.

There are some fundamental characteristics which are used to evaluate the quality of sensing matrices.

Definition I.2

[19] (RIP(restricted isometry property)) Matrix AA is said to satisfy the (k,δ)(k,\delta)-RIP, if there exists δ=δkA(0,1)\delta=\delta_{k}^{A}\in(0,1), such that

(1δ)x22Ax22(1+δ)x22,xΣk.\displaystyle(1-\delta)\|x\|_{2}^{2}\leq\|Ax\|_{2}^{2}\leq(1+\delta)\|x\|_{2}^{2},\quad\forall x\in\Sigma_{k}. (3)

When k<m<nk<m<n, RIP can ensure a lossless recovering of the signal.

Unfortunately, in general, verifying RIP is difficult. An alternative criterion is the coherence.

Definition I.3

[19]

Assume Am×nA\in{\mathcal{M}}_{m\times n}, m<nm<n. the coherence of AA, denoted by μ(A)\mu(A), is defined as follows.

μ(A)=max1ijn<ai,aj>aiaj,\displaystyle\mu(A)=\max_{1\leq i\neq j\leq n}\frac{<a_{i},a_{j}>}{\|a_{i}\|\|a_{j}\|}, (4)

where, ai=Coli(A)a_{i}=\operatorname{Col}_{i}(A).

Remark I.4
  • (i)

    [32]

    μ(A)[nmm(n1),1].\displaystyle\mu(A)\in\left[\sqrt{\frac{n-m}{m(n-1)}},1\right]. (5)
  • (ii)

    A sufficient condition for a lossless recovering the signal is [19]

    k<12(1+1μ(A)).\displaystyle k<\frac{1}{2}\left(1+\frac{1}{\mu(A)}\right). (6)

    Hence the smaller the μ(A)\mu(A) the larger the kk is allowed, which means more signals can be treated.

It was proved in [32] that the sensing matrix AIsA\otimes I_{s} has exactly the same CS-related properties as AA.

Proposition I.5

[32] Consider AA and AIsA\otimes I_{s}, where Am×nA\in{\mathcal{M}}_{m\times n}, m<nm<n, and s+s\in{\mathbb{Z}}^{+}.

  • (i)
    spark(AIs)=spark(A).\displaystyle spark(A\otimes I_{s})=spark(A). (7)
  • (ii)

    AA satisfies (k,δ)(k,\delta)-RIP, if and only, AIsA\otimes I_{s} satisfies (k,δ)(k,\delta)-RIP with the same kk and δ\delta. (The “only if” part is claimed and proved in [32], and the “if” part is straightforward verifiable.)

  • (iii)
    μ(AIs)=μ(A).\displaystyle\mu(A\otimes I_{s})=\mu(A). (8)

Proposition I.5 provides a theoretical foundation for STP-CS.

Since then, there have been considerable researches for STP-CS. For instance, a further development of STP-CS called the PTPCSPTP-CS is presented by [27], [25] considered how to construct a random sensing matrix for STP-CS. Some applications shown the efficiency of STP-CS. For instance, the design of one-bit compressed sensing by STP-CS was proposed by [22], [7] uses STP-CS for secure medical image transmission; the problem of reducing storage space of sensing matrix using STP-CS was investigated by [30]; a secure image encryption scheme by using STP-CS is proposed by [31]; [23] considered the storage constrained smart meter sensing using STP-CS; [35] proposed a reversible image hiding algorithm using STP-CS, etc.

The purpose of this paper is threefold: (1) Propose the structure of signal space. under the assumption that a finite signal with different sampling lengths or different sampling frequencies is considered as equivalent signals, the signal space is considered as the quotient space of {\mathbb{R}}^{\infty} over equivalence. The structure of signal space provides a solid framework for design of STP-CS. (2) Reveal some fundamental properties of STP-CS. (3) Propose a new technique for constructing sensing matrix based on BIBD.

The rest of this paper is organized as follows. Section 2 is a systematic review on both left and right STPs, their quotient spaces, and their vector space structure. Section 3 investigates the topological structure of the signal space, which is the quotient space of cross-dimensional Euclidian space, and its topology is deduced from distance. In Section 4 the signal space is embedded into various functional spaces, including Frėchet space, Balach space, and Hilbert space. These functional spaces provide different structures for signal space, which reveals related properties of signal space. Section 5 provides a generalized dimension free STP-CS, which contains the STP-CS in current literature as its spacial case. An precise lossless recovering condition for STP-CS is obtained. Section 6 proposes a new constructing technique for balanced incomplete block design (BIBD) based design of sensing matrix. Finally, Section 7 is a brief conclusion.

Before ending this section we give a list of notations.

  • {\mathbb{R}}: The set of real numbers.

  • {\mathbb{Q}} (Q+Q^{+}) : The set of (positive) rational numbers.

  • {\mathbb{Z}} (+{\mathbb{Z}}^{+}): Set of (positive) integers.

  • t|rt|r (trt\nmid r): t,r+t,r\in{\mathbb{Z}}^{+} and r/t+r/t\in{\mathbb{Z}}^{+} (r/t+r/t\not\in{\mathbb{Z}}^{+}).

  • {\mathbb{R}}^{\infty}: =n=1n{\mathbb{R}}^{\infty}=\bigcup_{n=1}^{\infty}{\mathbb{R}}^{n}.

  • m×n{\mathcal{M}}_{m\times n}: set of m×nm\times n real matrices.

  • μ:={Am×n|m/n=μ}{\mathcal{M}}_{\mu}:=\left\{A_{m\times n}\;|\;m/n=\mu\right\} where μ+\mu\in{\mathbb{Q}}^{+}.

  • :=m=1n=1m×n{\mathcal{M}}:=\bigcup_{m=1}^{\infty}\bigcup_{n=1}^{\infty}{\mathcal{M}}_{m\times n}

  • [a,b][a,b]: the set of integers {a,a+1,,b}\{a,a+1,\cdots,b\}, where aba\leq b.

  • x\lceil x\rceil: the smallest integer upper bound of xx, i.e., the smallest integer sxs\geq x.

  • lcm(n,p)\operatorname{lcm}(n,p) : the least common multiple of nn and pp.

  • gcd(n,p)\gcd(n,p) : the greatest common divisor of nn and pp.

  • δni\delta_{n}^{i}: the ii-th column of the identity matrix InI_{n}.

  • Δn:={δni|i=1,,n}\Delta_{n}:=\left\{\delta_{n}^{i}|i=1,\cdots,n\right\}.

  • 𝟏k:=(1,,1k)T{\bf 1}_{k}:=(\underbrace{1,\cdots,1}_{k})^{T}.

  • Span()\operatorname{Span}(\cdot): subspace or dual subspace generated by \cdot.

  • +¯\bar{+}: semi-tensor addition.

  • \ltimes: (left) matrix-matrix semi-tensor product, which is the default one.

  • \rtimes: right semi-tensor product of matrices.

  • \vec{\ltimes} (\vec{\rtimes}): left (right) matrix-vector semi-tensor product.

  • Σkn\Sigma^{n}_{k}: set of nn dimensional vector with less than or equal to kk non-zero entries.

  • Σk/sn\Sigma^{n}_{k/s}: set of nn dimensional vector with less than or equal to kk non-zero entries per ss length.

  • w(x)=0(x)w(x)=\ell_{0}(x): nonzero entries of vector xx(called the degree of xx).

  • spark(A)spark(A): the smallest number of dependent columns of AA.

  • μ(A)\mu(A): coherence of AA.

II Left and Right STP - Preliminaries

Definition II.1

[10] Assume Am×nA\in{\mathcal{M}}_{m\times n}, Bp×qB\in{\mathcal{M}}_{p\times q}, xrx\in{\mathbb{R}}^{r}, t=lcm(n,p)t=\operatorname{lcm}(n,p), and s=lcm(n,r)s=\operatorname{lcm}(n,r).

  • (i)

    The left matrix-matrix (MM-) STP is defined by

    AB:=(AIt/n)(BIt/p)mt/n×qt/p.\displaystyle A\ltimes B:=\left(A\otimes I_{t/n}\right)\left(B\otimes I_{t/p}\right)\in{\mathcal{M}}_{mt/n\times qt/p}. (9)
  • (ii)

    The right MM-STP is defined by

    AB:=(It/nA)(It/pB)mt/n×qt/p.\displaystyle A\rtimes B:=\left(I_{t/n}\otimes A\right)\left(I_{t/p}\otimes B\right)\in{\mathcal{M}}_{mt/n\times qt/p}. (10)
  • (iii)

    The left matrix-vector (MV-) STP is defined by

    Ax:=(AIs/n)(x𝟏s/r)s.\displaystyle A\vec{\ltimes}x:=\left(A\otimes I_{s/n}\right)\left(x\otimes{\bf 1}_{s/r}\right)\in{\mathbb{R}}^{s}. (11)
  • (iv)

    The right MV-STP is defined by

    Ax:=(Is/nA)(𝟏s/rx)s.\displaystyle A\vec{\rtimes}x:=\left(I_{s/n}\otimes A\right)\left({\bf 1}_{s/r}\otimes x\right)\in{\mathbb{R}}^{s}. (12)

Denote the set of finite-dimensional matrices by

=m=1n=1m×n.{\mathcal{M}}=\bigcup_{m=1}^{\infty}\bigcup_{n=1}^{\infty}{\mathcal{M}}_{m\times n}.

And the set of finite-dimensional vectors by

=n=1n.{\mathbb{R}}^{\infty}=\bigcup_{n=1}^{\infty}{\mathbb{R}}^{n}.

The basic properties of STPs are listed as follows:

Proposition II.2

[10] Let A,B,CA,B,C\in{\mathcal{M}}, x,yx,y\in{\mathbb{R}}^{\infty}.

  1. (1)

    (Consistency)

    • (i)

      When n=pn=p, the MM-STPs are degenerated to conventional MM-product. That is,

      AB=AB=AB.\displaystyle A\ltimes B=A\rtimes B=AB. (13)
    • (ii)

      When n=rn=r, the MV-STPs are degenerated to conventional MV-product. That is,

      Ax=Ax=Ax.\displaystyle A\vec{\ltimes}x=A\vec{\rtimes}x=Ax. (14)
  2. (2)

    (Associativity)

    • (i)
      (AB)C=A(BC).\displaystyle(A\ltimes B)\ltimes C=A\ltimes(B\ltimes C). (15)
    • (ii)
      (AB)C=A(BC).\displaystyle(A\rtimes B)\rtimes C=A\rtimes(B\rtimes C). (16)
    • (iii)
      (AB)x:=A(Bx).\displaystyle(A\ltimes B)\vec{\ltimes}x:=A\vec{\ltimes}(B\vec{\ltimes}x). (17)
    • (iv)
      (AB)x:=A(Bx).\displaystyle(A\rtimes B)\vec{\rtimes}x:=A\vec{\rtimes}(B\vec{\rtimes}x). (18)
  3. (3)

    (Distributivity)

    • (i)
      (A+B)C=AC+BC,C(A+B)=CA+CB.\displaystyle\begin{array}[]{l}(A+B)\ltimes C=A\ltimes C+B\ltimes C,\\ C\ltimes(A+B)=C\ltimes A+C\ltimes B.\end{array} (21)
    • (ii)
      (A+B)C=AC+BC,C(A+B)=CA+CB.\displaystyle\begin{array}[]{l}(A+B)\rtimes C=A\rtimes C+B\rtimes C,\\ C\rtimes(A+B)=C\rtimes A+C\rtimes B.\end{array} (24)
    • (iii)
      (A+B)x=Ax+Bx,A(x+y)=Ax+Ay.\displaystyle\begin{array}[]{l}(A+B)\vec{\ltimes}x=A\vec{\ltimes}x+B\vec{\ltimes}x,\\ A\vec{\ltimes}(x+y)=A\vec{\ltimes}x+A\vec{\ltimes}y.\\ \end{array} (27)
    • (iv)
      (A+B)x=Ax+Bx,A(x+y)=Ax+Ay.\displaystyle\begin{array}[]{l}(A+B)\vec{\rtimes}x=A\vec{\rtimes}x+B\vec{\rtimes}x,\\ A\vec{\rtimes}(x+y)=A\vec{\rtimes}x+A\vec{\rtimes}y.\\ \end{array} (30)

Note that when the addition of two matrices (vectors) appears to the equalities in Proposition II.2, the dimensions of two adding members must be the same.

Definition II.3
  • (i)

    Two matrices AA and BB are said to be left equivalent, denoted by ABA\sim_{\ell}B, if there exist two identities IαI_{\alpha} and IβI_{\beta} such that

    AIα=BIβ.\displaystyle A\otimes I_{\alpha}=B\otimes I_{\beta}. (31)

    The left equivalence class of AA is denoted by

    A:={B|BA}.\langle A\rangle_{\ell}:=\{B\;|\;B\sim_{\ell}A\}.
  • (ii)

    Two matrices AA and BB are said to be right equivalent, denoted by ArBA\sim_{r}B, it there exists two identities IαI_{\alpha} and IβI_{\beta} such that

    IαA=IβB.\displaystyle I_{\alpha}\otimes A=I_{\beta}\otimes B. (32)

    The right equivalence class of AA is denoted by

    Ar:={B|BrA}.\langle A\rangle_{r}:=\{B\;|\;B\sim_{r}A\}.
  • (iii)

    Two vectors xx and yy are said to be left equivalent, denoted by xyx\leftrightarrow_{\ell}y, if there exist two 1-vectors 𝟏α{\bf 1}_{\alpha} and 𝟏β{\bf 1}_{\beta} such that

    x𝟏α=y𝟏β.\displaystyle x\otimes{\bf 1}_{\alpha}=y\otimes{\bf 1}_{\beta}. (33)

    The left equivalence class of xx is denoted by

    x¯:={y|yx}.\bar{x}_{\ell}:=\{y\;|\;y\leftrightarrow_{\ell}x\}.
  • (iv)

    Two vectors xx and yy are said to be right equivalent, denoted by xryx\leftrightarrow_{r}y, if there exist two 1-vectors 𝟏α{\bf 1}_{\alpha} and 𝟏β{\bf 1}_{\beta} such that

    𝟏αx=𝟏βy.\displaystyle{\bf 1}_{\alpha}\otimes x={\bf 1}_{\beta}\otimes y. (34)

    The right equivalence class of xx is denoted by

    x¯r:={y|yrx}.\bar{x}_{r}:=\{y\;|\;y\leftrightarrow_{r}x\}.

It is easy to verify that the equivalences defined in Definition II.3 are equivalence relations.

Proposition II.4

[11]

  • (i)

    If ABA\sim_{\ell}B, then there exists a CC\in{\mathcal{M}} such that

    A=CIb,B=CIa.\displaystyle A=C\otimes I_{b},\quad B=C\otimes I_{a}. (35)

    w.l.g. (without lose of generality) we can assume in (31) the gcd(α,β)=1\gcd(\alpha,\beta)=1, then in (35)

    b=β,anda=α.b=\beta,~{}\mbox{and}~{}a=\alpha.
  • (ii)

    If ArBA\sim_{r}B, then there exists a CC\in{\mathcal{M}} such that

    A=IbC,B=IaC.\displaystyle A=I_{b}\otimes C,\quad B=I_{a}\otimes C. (36)

    Assume in (32) the gcd(α,β)=1\gcd(\alpha,\beta)=1, then in (36)

    b=β,anda=α.b=\beta,~{}\mbox{and}~{}a=\alpha.
  • (iii)

    If xyx\leftrightarrow_{\ell}y, then there exists a zz\in{\mathbb{R}}^{\infty} such that

    x=z𝟏b,y=z𝟏a.\displaystyle x=z\otimes{\bf 1}_{b},\quad y=z\otimes{\bf 1}_{a}. (37)

    Assume in (33) the gcd(α,β)=1\gcd(\alpha,\beta)=1, then in (37)

    b=β,anda=α.b=\beta,~{}\mbox{and}~{}a=\alpha.
  • (iv)

    If xryx\leftrightarrow_{r}y, then there exists a zz\in{\mathbb{R}}^{\infty} such that

    x=𝟏bz,y=𝟏az.\displaystyle x={\bf 1}_{b}\otimes z,\quad y={\bf 1}_{a}\otimes z. (38)

    Assume in (34) the gcd(α,β)=1\gcd(\alpha,\beta)=1, then in (38)

    b=β,anda=α.b=\beta,~{}\mbox{and}~{}a=\alpha.
Definition II.5
  • (i)

    AA\in{\mathcal{M}} is said to be left (right) reducible, if there exist an identity matrix IsI_{s}, s>1s>1 and an A0A_{0}\in{\mathcal{M}}, such that

    A=A0Is,(correspondingly,A=IsA0).\displaystyle A=A_{0}\otimes I_{s},(\mbox{correspondingly,}~{}A=I_{s}\otimes A_{0}). (39)

    Otherwise, it is called left (right) irreducible.

  • (ii)

    AA\in{\mathbb{R}}^{\infty} is said to be left (right) reducible, if there exist an 1-vector 𝟏s{\bf 1}_{s}, s>1s>1 and an x0x_{0}\in{\mathbb{R}}^{\infty}, such that

    x=x0𝟏s,(correspondingly,x=𝟏sx0).\displaystyle x=x_{0}\otimes{\bf 1}_{s},(\mbox{correspondingly,}~{}x={\bf 1}_{s}\otimes x_{0}). (40)

    Otherwise, it is called left (right) irreducible.

Proposition II.6

[11]

  • (i)

    Consider A\langle A\rangle_{\ell} (Ar\langle A\rangle_{r}). There exists a unique left irreducible A0A_{0}^{\ell} (correspondingly, right irreducible A0rA_{0}^{r}), such that

    A={An=A0In|n+}.(correspondingly,Ar={An=InA0r|n+}.)\displaystyle\begin{array}[]{l}\langle A\rangle_{\ell}=\{A_{n}=A^{\ell}_{0}\otimes I_{n}\;|\;n\in{\mathbb{Z}}^{+}\}.\\ (\mbox{correspondingly,}~{}\langle A\rangle_{r}=\{A_{n}=I_{n}\otimes A^{r}_{0}\;|\;n\in{\mathbb{Z}}^{+}\}.)\\ \end{array} (43)
  • (ii)

    Consider x¯\bar{x}_{\ell} (x¯r\bar{x}_{r}). There exists a unique left irreducible x0x_{0}^{\ell} (correspondingly, right irreducible x0rx_{0}^{r}), such that

    x¯={xn=x0𝟏n|n+}.(correspondingly,x¯r={xn=𝟏nx0r|n+}.)\displaystyle\begin{array}[]{l}\bar{x}_{\ell}=\{x_{n}=x^{\ell}_{0}\otimes{\bf 1}_{n}\;|\;n\in{\mathbb{Z}}^{+}\}.\\ (\mbox{correspondingly,}~{}\bar{x}_{r}=\{x_{n}={\bf 1}_{n}\otimes x^{r}_{0}\;|\;n\in{\mathbb{Z}}^{+}\}.)\\ \end{array} (46)
Definition II.7
  • (i)

    Consider {\mathcal{M}}. The left equivalent classes form the left quotient space as

    Σ:=/={A|A}.\displaystyle\Sigma_{\ell}:={\mathcal{M}}/\sim_{\ell}=\{\langle A\rangle_{\ell}\;|\;A\in{\mathcal{M}}\}. (47)
  • (ii)

    Consider {\mathcal{M}}. The right equivalent classes form the right quotient space as

    Σr:=/r={Ar|A}.\displaystyle\Sigma_{r}:={\mathcal{M}}/\sim_{r}=\{\langle A\rangle_{r}\;|\;A\in{\mathcal{M}}\}. (48)
  • (iii)

    Consider {\mathbb{R}}^{\infty}. The left equivalent classes form the left quotient space as

    Ω:=/={x¯|x}.\displaystyle\Omega_{\ell}:={\mathbb{R}}^{\infty}/\leftrightarrow_{\ell}=\{\bar{x}_{\ell}\;|\;x\in{\mathbb{R}}^{\infty}\}. (49)
  • (iv)

    Consider {\mathbb{R}}^{\infty}. The right equivalent classes form the right quotient space as

    Ωr:=/r={x¯r|x}.\displaystyle\Omega_{r}:={\mathbb{R}}^{\infty}/\leftrightarrow_{r}=\{\bar{x}_{r}\;|\;x\in{\mathbb{R}}^{\infty}\}. (50)
Remark II.8

In this section we carefully reviewed the left(right) STP, left(right) equivalence, and left(right) quotient space for both matrices and vectors. It is easy to see that the left (STP) system, including the left STP, left equivalence, and left quotient space, is “mirror symmetric” to the right (STP) system, including the right STP, right equivalence, and right quotient space. Up to this paper, all the researches are concentrated mainly on the left system. In the rest of this paper, we are only interested on left system too. We claim that all the results obtained for the left system are also available for the right system. Unless elsewhere stated, the corresponding verifications are straightforward.

Hereafter we assume the default system is the left one. That is,

{:=,:=,A:=A,x¯:=x¯,Ω:=Ω,Σ:=Σ.\displaystyle\begin{cases}\sim:=\sim_{\ell},\\ \leftrightarrow:=\leftrightarrow_{\ell},\\ \langle A\rangle:=\langle A\rangle_{\ell},\\ \bar{x}:=\bar{x}_{\ell},\\ \Omega:=\Omega_{\ell},\\ \Sigma:=\Sigma_{\ell}.\\ \end{cases} (51)

Unless elsewhere stated.

III Signal Space

A signal length nn can be considered as a vector in Euclidian space n{\mathbb{R}}^{n} [19].

Definition III.1

Two signals xmx\in{\mathbb{R}}^{m} and yny\in{\mathbb{R}}^{n} are said to be equivalent if xyx\leftrightarrow y.

Remark III.2
  • (i)

    According to Proposition II.4, xryx\leftrightarrow_{r}y means there exists a zz such that x=𝟏αzx={\bf 1}_{\alpha}\otimes z and y=𝟏βzy={\bf 1}_{\beta}\otimes z. Hence we have

    x=(zT,zT,,zT)Tα,y=(zT,zT,,zT)Tβ.\begin{array}[]{l}x=\underbrace{(z^{T},z^{T},\cdots,z^{T})^{T}}_{\alpha},\\ y=\underbrace{(z^{T},z^{T},\cdots,z^{T})^{T}}_{\beta}.\\ \end{array}

    That is, xx and yy are obtained from same signal zz with different sampling time lengths. This fact makes the physical meaning of equivalence clear.

    If we consider xx and yy from frequency domain, they are exactly the same.

  • (ii)

    Similarly, xyx\leftrightarrow_{\ell}y means there exists a zz such that x=z𝟏αx=z\otimes{\bf 1}_{\alpha} and y=z𝟏βy=z\otimes{\bf 1}_{\beta}. Assume z=(z1,z2,,zk)Tz=(z_{1},z_{2},\cdots,z_{k})^{T}, then we have

    x=(z1,z1,,z1α,z2,z2,,z2α,,zk,zk,,zkα)T,y=(z1,z1,,z1β,z2,z2,,z2β,,zk,zk,,zkβ)T.\begin{array}[]{l}x=(\underbrace{z_{1},z_{1},\cdots,z_{1}}_{\alpha},\underbrace{z_{2},z_{2},\cdots,z_{2}}_{\alpha},\cdots,\underbrace{z_{k},z_{k},\cdots,z_{k}}_{\alpha})^{T},\\ y=(\underbrace{z_{1},z_{1},\cdots,z_{1}}_{\beta},\underbrace{z_{2},z_{2},\cdots,z_{2}}_{\beta},\cdots,\underbrace{z_{k},z_{k},\cdots,z_{k}}_{\beta})^{T}.\\ \end{array}

    It can be considered as both xx and yy are obtained from same signal zz with different sampling frequencies (with certain approximation).

Hence, from finite signal point of view, the right (or left) equivalence is physically meaningful.

It is clear that under the left equivalence the signal space becomes

Ω=Ω.\Omega=\Omega_{\ell}.

And a signal xx becomes x¯\bar{x}. Let x0x¯x_{0}\in\bar{x} be irreducible, then x0x_{0} is called the atom signal, and we define

dim(x¯):=dim(x0).\displaystyle\dim(\bar{x}):=\dim(x_{0}). (52)

III-A Vector Space Structure on Signal Space

Definition III.3

[11] Let x,yx,y\in{\mathbb{R}}^{\infty} with xmx\in{\mathbb{R}}^{m} and yny\in{\mathbb{R}}^{n}, t=lcm(m,n)t=\operatorname{lcm}(m,n). Then the left (right) semi-tensor addition (STA) of xx and yy is defied by

x±y:=(x𝟏t/m)±(y𝟏t/n)t;x±ry:=(𝟏t/mx)±(𝟏t/ny)t.\displaystyle\begin{array}[]{l}x\vec{\pm}_{\ell}y:=\left(x\otimes{\bf 1}_{t/m}\right)\pm\left(y\otimes{\bf 1}_{t/n}\right)\in{\mathbb{R}}^{t};\\ x\vec{\pm}_{r}y:=\left({\bf 1}_{t/m}\otimes x\right)\pm\left({\bf 1}_{t/n}\otimes y\right)\in{\mathbb{R}}^{t}.\end{array} (55)

The physical meaning of these additions are very clear: Assume xx is a signal with period mm and yy is a signal with period nn, then the addition of these two signals, x+ryx\vec{+}_{r}y, is a signal with period tt. Similarly, assume xx is a signal with frequency 1/m1/m and yy is a signal with period 1/n1/n, then the addition of these two signals, x+yx\vec{+}_{\ell}y, is a signal with frequency 1/t1/t.

Figure 1 shows that the black signal has period 22 and the red signal has period 33, then their addition x+ryx\vec{+}_{r}y, (blue line) is a signal with period lcm(2,3)=6\operatorname{lcm}(2,3)=6; their addition x+yx\vec{+}_{\ell}y, (green line) is a signal with frequency 1/61/6.

012623xxyyx+ryx\vec{+}_{r}yx+yx\vec{+}_{\ell}y
Figure 1: Signal Addition
Proposition III.4

[11] The addition defined by (55) ls consistent with the equivalence. That is, if x1x2x_{1}\leftrightarrow x_{2} and y1y2y_{1}\leftrightarrow y_{2}, then

x1±y1x2±y2.\displaystyle x_{1}\vec{\pm}y_{1}\leftrightarrow x_{2}\vec{\pm}y_{2}. (56)

Using Proposition III.4, the following results are obvious.

Proposition III.5

[11]

  • (i)

    +\vec{+} is properly defined on Ω\Omega by

    x¯±y¯=x±y¯,x¯,y¯Ω.\displaystyle\bar{x}\vec{\pm}\bar{y}=\overline{x\vec{\pm}y},\quad\bar{x},\bar{y}\in\Omega. (57)
  • (ii)

    Define the scalar product on Ω\Omega by

    ax¯:=ax¯,a,x¯Ω.\displaystyle a\cdot\bar{x}:=\overline{ax},\quad a\in{\mathbb{R}},\bar{x}\in\Omega. (58)

    With +\vec{+} and scalar product \cdot, Ω\Omega is a vector space.

III-B Topological Structure on Ω\Omega

Definition III.6

Let x,yx,y\in{\mathbb{R}}^{\infty} with xmx\in{\mathbb{R}}^{m} and yny\in{\mathbb{R}}^{n}, t=lcm(m,n)t=\operatorname{lcm}(m,n).

  • (i)

    The inner product of xx and yy is defined by

    x,y𝒱:=1t(x𝟏t/m,y𝟏t/n,\displaystyle\langle x,y\rangle_{{\mathcal{V}}}:=\frac{1}{t}\langle(x\otimes{\bf 1}_{t/m},y\otimes{\bf 1}_{t/n}\rangle, (59)

    where

    (x𝟏t/m,y𝟏t/n=(x𝟏t/m)T(y𝟏t/n)\langle(x\otimes{\bf 1}_{t/m},y\otimes{\bf 1}_{t/n}\rangle=(x\otimes{\bf 1}_{t/m})^{T}(y\otimes{\bf 1}_{t/n})

    is the conventional inner product on t{\mathbb{R}}^{t}.

  • (ii)

    The norm of xx is defined by

    x𝒱:=x,x𝒱=1nxTx.\displaystyle\|x\|_{{\mathcal{V}}}:=\sqrt{\langle x,x\rangle_{{\mathcal{V}}}}=\frac{1}{\sqrt{n}}\sqrt{x^{T}x}. (60)
  • (iii)

    The distance of xx and yy is defined by

    d𝒱(x,y):=xy𝒱.\displaystyle d_{{\mathcal{V}}}(x,y):=\|x\vec{-}y\|_{{\mathcal{V}}}. (61)

Using this inner product, the angle between two elements x,yx,y\in{\mathbb{R}}^{\infty}, denoted by θ=θx,y\theta=\theta_{x,y}, is defined by

cos(θ)=x,y𝒱x𝒱y𝒱.\displaystyle\cos(\theta)=\frac{\langle x,y\rangle_{{\mathcal{V}}}}{\|x\|_{{\mathcal{V}}}\|y\|_{{\mathcal{V}}}}. (62)

Using this distance, the distance deduced topology can be obtained, which is denoted by 𝒯d{\mathcal{T}}_{d}. Apart from this 𝒯d{\mathcal{T}}_{d}, there is another topology on {\mathbb{R}}^{\infty}. Naturally, each n{\mathbb{R}}^{n} can be considered as a component (i.e., a clopen set) of {\mathbb{R}}^{\infty}. And within each n{\mathbb{R}}^{n} the conventional Euclidian space topology is used. Then overall, such a topology is called the natural topology, denoted by 𝒯n{\mathcal{T}}_{n}. (,𝒯n)({\mathbb{R}}^{\infty},{\mathcal{T}}_{n}) is disconnected. Precisely speaking, its fundamental group is +{\mathbb{Z}}^{+}.

It is easy to verify the following result.

Proposition III.7

[11] Consider {\mathbb{R}}^{\infty}. Let x,yx,y\in{\mathbb{R}}^{\infty}. Then d(x,y)=0d(x,y)=0, if and only if, xyx\leftrightarrow y.

Proposition III.8

[11] The inner product defined by (59) is consistent with the equivalence. That is, if x1x2x_{1}\leftrightarrow x_{2} and y1y2y_{1}\leftrightarrow y_{2}, then

x1,y1𝒱=x2,y2𝒱.\displaystyle\langle x_{1},y_{1}\rangle_{{\mathcal{V}}}=\langle x_{2},y_{2}\rangle_{{\mathcal{V}}}. (63)

According to Proposition III.8, the inner product, norm, distance defined by Definition III.6 for {\mathbb{R}}^{\infty} can be extended to Ω\Omega. Hence the following definitions for Ω\Omega are all properly defined.

Definition III.9
  • (i)

    The inner product of x¯\bar{x} and y¯\bar{y} is defined by

    x¯,y¯𝒱:=x,y𝒱,x¯,y¯Ω.\displaystyle\langle\bar{x},\bar{y}\rangle_{{\mathcal{V}}}:=\langle x,y\rangle_{{\mathcal{V}}},\quad\bar{x},\bar{y}\in\Omega. (64)
  • (ii)

    The norm of x¯\bar{x} is defined by

    x¯𝒱:=x𝒱,x¯Ω.\displaystyle\|\bar{x}\|_{{\mathcal{V}}}:=\|x\|_{{\mathcal{V}}},\bar{x}\in\Omega. (65)
  • (iii)

    The distance of x¯\bar{x} and y¯\bar{y} is defined by

    d𝒱(x¯,y¯):=d𝒱(x,y),x¯,y¯Ω.\displaystyle d_{{\mathcal{V}}}(\bar{x},\bar{y}):=d_{{\mathcal{V}}}(x,y),\bar{x},\bar{y}\in\Omega. (66)

With the topology deduced by (66) Ω\Omega is a topological space. Because of Proposition (61), this distance deduced topology is homeomorphic to quotient topology. That is,

Ω=[(,𝒯n)/](,𝒯d).\Omega=\left[\left({\mathbb{R}}^{\infty},{\mathcal{T}}_{n}\right)/\leftrightarrow\right]\cong\left({\mathbb{R}}^{\infty},{\mathcal{T}}_{d}\right).

In the following proposition we summarize some known basic properties of Ω\Omega.

Proposition III.10

[11, 12] Consider the signal space Ω\Omega.

  • (i)] Ω\Omega is an infinite dimensional vector space, while each element has finite dimension. That is,

    dim(x¯)<,x¯Ω.\dim(\bar{x})<\infty,\quad\bar{x}\in\Omega.
  • (ii)

    As a topological space, Ω\Omega is second countable, separable, Hausdorff.

  • (iii)

    Ω\Omega is an inner product space, but not Hilbert.

Since Ω\Omega is an inner product space, then the following geometric structure is obvious.

Corollary III.11
  • (i)

    For any two points x¯,y¯Ω\bar{x},\bar{y}\in\Omega, the angle between them (the same as the angle between xx and yy) , denoted by θx¯,y¯\theta_{\bar{x},\bar{y}}, is determined by

    cos(θx¯,y¯)=cos(θx,y)=x,y𝒱x𝒱y𝒱.\displaystyle\cos(\theta_{\bar{x},\bar{y}})=\cos(\theta_{x,y})=\frac{\langle x,y\rangle_{{\mathcal{V}}}}{\|x\|_{{\mathcal{V}}}\|y\|_{{\mathcal{V}}}}. (67)
  • (ii)

    x¯\bar{x} and y¯\bar{y} (as well as xx and yy) are said to be parallel (orthogonal), if cos(θx,y)=1\cos(\theta_{x,y})=1 (correspondingly, cos(θx,y)=0\cos(\theta_{x,y})=0 ).

Let xmx\in{\mathbb{R}}^{m}. The projection of xx onto n{\mathbb{R}}^{n}, denoted by πnm(x)\pi^{m}_{n}(x), is defined by

πnm(x)=argminy(d𝒱(y,x)).\displaystyle\pi^{m}_{n}(x)=argmin_{y}(d_{{\mathcal{V}}}(y,x)). (68)
Proposition III.12

[12]

  • (i)

    The projection of xmx\in{\mathbb{R}}^{m} onto n{\mathbb{R}}^{n} is

    y0:=πnm(x)=Πnmx,\displaystyle y_{0}:=\pi^{m}_{n}(x)=\Pi^{m}_{n}\ltimes x, (69)

    where (t=lcm(m,n)t=\operatorname{lcm}(m,n))

    Πnm=nt(In𝟏t/nT)(Im𝟏t/m).\Pi^{m}_{n}=\frac{n}{t}\left(I_{n}\otimes{\bf 1}^{T}_{t/n}\right)\left(I_{m}\otimes{\bf 1}_{t/m}\right).
  • (ii)

    xy0x\vec{-}y_{0} is perpendicular to y0y_{0}, ((xy0)y0(x\vec{-}y_{0})\perp y_{0}). That is, the angle between them, denoted by θ\theta satisfies

    cos(θ)=0.\displaystyle\cos(\theta)=0. (70)
  • (iii)

    For any yny\in{\mathbb{R}}^{n}, (xy0)y(x\vec{-}y_{0})\perp y.

(See Figure 2 for the projection.)

n{\mathbb{R}}^{n}xxxy0x\vec{-}y_{0}y0y_{0}yy0
Figure 2: Projection
Remark III.13

The projection πnm\pi^{m}_{n} can be extended naturally to a projection between subspaces of Ω\Omega.

III-C Linear Mapping on Ω\Omega

Consider the set of matrices {\mathcal{M}} and the equivalence =\sim=\sim_{\ell}. Similarly to vectors, we define the reducibility.

Definition III.14

Let AA\in{\mathcal{M}}. AA is said to be (left) reducible, if there exists an identity matrix IsI_{s}, s>1s>1 and a matrix A0A_{0} such that

A=A0Is.\displaystyle A=A_{0}\otimes I_{s}. (71)

Otherwise, it is irreducible.

Using Proposition II.4, one sees easily that in A\langle A\rangle there exists a unique irreducible A0A_{0} such that

A={An=A0In|n+}\langle A\rangle=\{A_{n}=A_{0}\otimes I_{n}\;|\;n\in{\mathbb{Z}}^{+}\}

Next, consider the linear mapping on {\mathbb{R}}^{\infty}. We have the following result:

Proposition III.15

[11] Assume A,BA,B\in{\mathcal{M}} with ABA\sim B, and x,yx,y\in{\mathbb{R}}^{\infty} with xyx\leftrightarrow y. Then the MV-STP \vec{\ltimes} is consistent with both equivalences. That is,

AxBy.\displaystyle A\vec{\ltimes}x\leftrightarrow B\vec{\ltimes}y. (72)

Recall the quotient space of matrices, Σ=Σ\Sigma=\Sigma_{\ell}, using Proposition III.15, the following result is obvious.

Corollary III.16
  • (i)

    A linear mapping on image space Ω\Omega is a mapping π:Σ×ΩΩ\pi:\Sigma\times\Omega\rightarrow\Omega, which can be expressed by

    Ax¯:=Ax¯.\displaystyle\langle A\rangle\vec{\ltimes}\bar{x}:=\overline{A\vec{\ltimes}x}. (73)
  • (ii)

    For two linear mappings A\langle A\rangle and B\langle B\rangle,

    A(Bx¯)=(AB)x,A,BΣ,x¯Ω.\displaystyle\langle A\rangle\vec{\ltimes}(B\vec{\ltimes}\bar{x})=(\langle A\rangle\ltimes\langle B\rangle)\vec{\ltimes}x,\quad\langle A\rangle,\langle B\rangle\in\Sigma,\bar{x}\in\Omega. (74)

IV Structure of Signal Space

IV-A Orthonormal Basis of Signal Space

First, we search a basis of Ω=Ω\Omega=\Omega_{\ell}.

It was claimed by [37] that

𝒱:={1,δij|j<i,gcd(i,j)=1;i+}\displaystyle{\mathcal{B}}_{{\mathcal{V}}}:=\left\{1,\delta_{i}^{j}\;|\;j<i,\gcd(i,j)=1;\;i\in{\mathbb{Z}}^{+}\right\} (75)

is a basis of Ω\Omega. Unfortunately, 𝒱{\mathcal{B}}_{{\mathcal{V}}} is only a generating set of Ω\Omega, which was proved by [37]. But there are some linearly depending terms. Say,

δ41+δ43δ21=0.\delta_{4}^{1}\vec{+}\delta_{4}^{3}\vec{-}\delta_{2}^{1}=0.

Searching a basis of Ω\Omega is a fundamental task for understanding and using the signal space.

Denote

𝒱n:=𝒱Δn,𝒱m:=nm𝒱n,Ω1=Span{1},Ωn=Span{δnj|j<n,gcd(j,n)=1},n2,Ωm=nmΩn.\begin{array}[]{l}{\mathcal{B}}_{{\mathcal{V}}_{n}}:={\mathcal{B}}_{{\mathcal{V}}}\bigcap\Delta_{n},\\ {\mathcal{B}}^{m}_{{\mathcal{V}}}:=\bigcup_{n\leq m}{\mathcal{B}}_{{\mathcal{V}}_{n}},\\ \Omega_{1}=\operatorname{Span}\{1\},\\ \Omega_{n}=\operatorname{Span}\{\delta_{n}^{j}\;|\;j<n,\;\gcd(j,n)=1\},\quad n\geq 2,\\ \Omega^{m}=\bigcup_{n\leq m}\Omega_{n}.\\ \end{array}

1<s+1<s\in{\mathbb{Z}}^{+} is said to be a multi-fold divisor of n+n\in{\mathbb{Z}}^{+} if s2|ns^{2}|n.

Note that since Ω=Span(𝒱)\Omega=\operatorname{Span}({\mathcal{B}}_{{\mathcal{V}}}), we can gradually choose

𝒟𝒱n𝒱n,n=1,2,,{\mathcal{D}}_{{\mathcal{V}}_{n}}\subset{\mathcal{B}}_{{\mathcal{V}}_{n}},\quad n=1,2,\cdots,

and denote

𝒟𝒱m:=nm𝒟𝒱n,{\mathcal{D}}^{m}_{{\mathcal{V}}}:=\bigcup_{n\leq m}{\mathcal{D}}_{{\mathcal{V}}_{n}},

such that the elements of 𝒟m𝒱{\mathcal{D}}^{m}{{\mathcal{V}}} are linearly independent, and

Ωm=Span(𝒟m𝒱).\displaystyle\Omega^{m}=\operatorname{Span}({\mathcal{D}}^{m}{{\mathcal{V}}}). (76)

That is, 𝒟𝒱m{\mathcal{D}}^{m}_{{\mathcal{V}}} is a basis of Ωm\Omega^{m}.

Let mm\rightarrow\infty. The basis of Ω\Omega is obtained.

To describe this process clearly, the following theory shows the relationship between 𝒟𝒱n{\mathcal{D}}_{{\mathcal{V}}_{n}} and 𝒱n{\mathcal{B}}_{{\mathcal{V}}_{n}}.

Theorem IV.1
Ωn1+Span(𝒟𝒱n)=Ωn,\displaystyle\Omega^{n-1}+\operatorname{Span}({\mathcal{D}}_{{\mathcal{V}}_{n}})=\Omega^{n}, (77)

if and only if, nn has no multi-fold divisor.

Proof.

Since

Ωn=Ωn1+Span(𝒱n),\Omega^{n}=\Omega^{n-1}+\operatorname{Span}({\mathcal{B}}_{{\mathcal{V}}_{n}}),

(77) tells us how to choose the elements from 𝒱n{\mathcal{B}}_{{\mathcal{V}}_{n}}, which are linearly independent with elements in Ωn1\Omega^{n-1}.

(Sufficiency:)

Assume nn has no multi-fold divisor. We have to show that

{Ωn1,𝒱n}\displaystyle\left\{\Omega^{n-1},{\mathcal{B}}_{{\mathcal{V}}_{n}}\right\} (78)

are linearly independent, which implies that 𝒟𝒱n=𝒱n{\mathcal{D}}_{{\mathcal{V}}_{n}}={\mathcal{B}}_{{\mathcal{V}}_{n}}.

Let {δni1,,δnis}\{\delta_{n}^{i_{1}},\cdots,\delta_{n}^{i_{s}}\} be linearly dependent with Ωn1\Omega^{n-1}. Then there exists a xΩn1x\in\Omega^{n-1} such that it can be expressed by

x¯=c1δni1++csδnis¯.\displaystyle\bar{x}=\overline{c_{1}\delta_{n}^{i_{1}}\vec{+}\cdots\vec{+}c_{s}\delta_{n}^{i_{s}}}. (79)

Since

c1δni1++csδnisSpan(𝒱n),c_{1}\delta_{n}^{i_{1}}\vec{+}\cdots\vec{+}c_{s}\delta_{n}^{i_{s}}\in\operatorname{Span}({\mathcal{B}}_{{\mathcal{V}}_{n}}),

say x0x¯x_{0}\in\bar{x} is irreducible. then x0Span(𝒱n)x_{0}\in\operatorname{Span}({\mathcal{B}}_{{\mathcal{V}}_{n}}). But dim(x0)<n\dim(x_{0})<n, hence back to Euclidian space, we should have

𝟏tx0=c1δni1++csδnis,\displaystyle{\bf 1}_{t}\otimes x_{0}=c_{1}\delta_{n}^{i_{1}}\vec{+}\cdots\vec{+}c_{s}\delta_{n}^{i_{s}}, (80)

where t|nt|n, say n=tjn=tj.

It is obvious that x00x_{0}\neq 0, because the elements in 𝒱n{\mathcal{B}}_{{\mathcal{V}}_{n}} are linearly independent. Say, the ii-th component of x0x_{0} denoted by x0i0x_{0}^{i}\neq 0. Then

𝟏tx0=[δjiδjiδji]+[ξjiξjiiξji],{\bf 1}_{t}\otimes x_{0}=\begin{bmatrix}\delta_{j}^{i}\\ \delta_{j}^{i}\\ \vdots\\ \delta_{j}^{i}\end{bmatrix}+\begin{bmatrix}\xi_{j}^{-i}\\ \xi_{j}^{-i}i\\ \vdots\\ \xi_{j}^{-i}\end{bmatrix},

where ξjij\xi_{j}^{-i}\in{\mathbb{R}}^{j} with ii-th component equals 0. To meet (80), we need

δni+(k1)t,k[1,j].\displaystyle\delta_{n}^{i+(k-1)t},\quad\forall k\in[1,j]. (81)

We claim that there exists at least one kk such that

i+(k1)j(modt)=0.\displaystyle i+(k-1)j(\mod~{}t)=0. (82)

Since nn has no multi-fold divisor, we have gcd(j,t)=1\gcd(j,t)=1. (Otherwise, say gcd(j,t)=r\gcd(j,t)=r, then n=jtn=jt has rr as its multi-fold divisor. )

Then for any i<ti<t as kk runs from 11 to tt i+(k1)j(modt)i+(k-1)j(\mod~{}t) takes all values from [0,t1][0,t-1]. That is, there exists a k0[1,m]k_{0}\in[1,m] such that i+(k01)j=μti+(k_{0}-1)j=\mu t. That is, gcd(i+(k01)j,n)>0\gcd(i+(k_{0}-1)j,n)>0. Hence

δni+(k0=1)t𝒱n.\delta_{n}^{i+(k_{0}=1)t}\not\in{\mathcal{B}}_{{\mathcal{V}}_{n}}.

That is, (81) does not satisfied, and hence (80 is not true. Hance, (78) is not a linearly independent set.

(Necessity) Assume n=k2jn=k^{2}j. where k>1k>1. We have to show (78) is a linearly dependent set. We can, w.l.g., assume kk is a prime number. Then

𝟏kδkji=δni+δni+j++δni+(k1)ji[1,k1].\begin{array}[]{l}{\bf 1}_{k}\otimes\delta^{i}_{kj}=\delta_{n^{i}}+\delta_{n}^{i+j}+\cdots+\delta_{n}^{i+(k-1)j}\\ ~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}i\in[1,k-1].\end{array}

Note that

gcd(i+sj,n)=gcd(i,k)=1,i[1,k1],s[0,k1],\gcd(i+sj,n)=\gcd(i,k)=1,\quad i\in[1,k-1],\;s\in[0,k-1],

Then we have

δni+δni+j++δni+(k1)jSpan(𝒱n).\delta_{n^{i}}+\delta_{n}^{i+j}+\cdots+\delta_{n}^{i+(k-1)j}\in\operatorname{Span}({\mathcal{B}}_{{\mathcal{V}}_{n}}).\\

That is,

Ωn1Span(𝒱n){0},\Omega^{n-1}\bigcap\operatorname{Span}({\mathcal{B}}_{{\mathcal{V}}_{n}})\neq\{0\},

which implies that (78) is a linearly dependent set.

\Box

From the proof of Theorem IV.1, one sees easily how to choose elements of 𝒱n{\mathcal{B}}_{{\mathcal{V}}_{n}} to form 𝒟𝒱n{\mathcal{D}}_{{\mathcal{V}}_{n}}, and hence the basis of Ωn\Omega^{n}, which is 𝒟𝒱n=1sn𝒟𝒱s{\mathcal{D}}^{n}_{{\mathcal{V}}}=\bigcup_{1\leq s\leq n}{\mathcal{D}}_{{\mathcal{V}}_{s}}.

Corollary IV.2
  • (i)

    Assume nn has no multi-divisor, then

    𝒟𝒱n=𝒱n.\displaystyle{\mathcal{D}}_{{\mathcal{V}}_{n}}={\mathcal{B}}_{{\mathcal{V}}_{n}}. (83)
  • (ii)

    Assume nn has multi-divisor. Express n=s2jn=s^{2}\cdot j, where jj has no multi-divisor. (Such an expression is unique.) Then

    𝒟𝒱n={δnj𝒱n|j(s1)sj}\displaystyle{\mathcal{D}}_{{\mathcal{V}}_{n}}=\left\{\delta_{n}^{j}\in{\mathcal{B}}_{{\mathcal{V}}_{n}}\;|\;j\leq(s-1)s\cdot j\right\} (84)
Example IV.3
  • (i)

    Consider n=12:=s2jn=12:=s^{2}\cdot j. Then s=2s=2 and j=3j=3, and

    𝒱n={δ121,δ125,δ127,δ1211}.{\mathcal{B}}_{{\mathcal{V}}_{n}}=\{\delta_{12}^{1},\delta_{12}^{5},\delta_{12}^{7},\delta_{12}^{11}\}.

    Now we calculate (s1)sj=6(s-1)s\cdot j=6. Hence,

    𝒟𝒱n={δ121,δ125}.{\mathcal{D}}_{{\mathcal{V}}_{n}}=\{\delta_{12}^{1},\delta_{12}^{5}\}.

    Note that

    δ121+δ127=𝟏2δ61,\delta_{12}^{1}+\delta_{12}^{7}={\bf 1}_{2}\otimes\delta_{6}^{1},

    and

    δ125+δ1211=𝟏2δ65,\delta_{12}^{5}+\delta_{12}^{11}={\bf 1}_{2}\otimes\delta_{6}^{5},

    Hence, remove δ125\delta_{12}^{5} and δ1211\delta_{12}^{11} is reasonable.

  • (ii)

    Consider n=27:=s2jn=27:=s^{2}\cdot j. Then s=3s=3 and j=3j=3, and

    𝒱n={δ271,δ272,δ274,δ275,δ277,δ278,δ2710,δ2711,δ2713,δ2714,δ2716,δ2717,δ2719,δ2720,δ2722,δ2723,δ2725,δ2726}.\begin{array}[]{l}{\mathcal{B}}_{{\mathcal{V}}_{n}}=\{\delta_{27}^{1},\delta_{27}^{2},\delta_{27}^{4},\delta_{27}^{5},\delta_{27}^{7},\delta_{27}^{8},\delta_{27}^{10},\delta_{27}^{11},\\ \delta_{27}^{13},\delta_{27}^{14},\delta_{27}^{16},\delta_{27}^{17},\delta_{27}^{19},\delta_{27}^{20},\delta_{27}^{22},\delta_{27}^{23},\delta_{27}^{25},\delta_{27}^{26}\}.\\ \end{array}

    Calculate (s1)sj=18(s-1)s\cdot j=18. Hence,

    𝒟𝒱n={δ271,δ272,δ274,δ275,δ277,δ278,δ2710,δ2711,δ2713,δ2714,δ2716,δ2717}.\begin{array}[]{l}{\mathcal{D}}_{{\mathcal{V}}_{n}}=\{\delta_{27}^{1},\delta_{27}^{2},\delta_{27}^{4},\delta_{27}^{5},\delta_{27}^{7},\delta_{27}^{8},\delta_{27}^{10},\delta_{27}^{11},\\ \delta_{27}^{13},\delta_{27}^{14},\delta_{27}^{16},\delta_{27}^{17}\}.\\ \end{array}

    To see the rest elements have to be deleted, we have

    𝟏3δ9k=δ27k+δ27k+9+δ27k+18,k[1,8].{\bf 1}_{3}\otimes\delta_{9}^{k}=\delta_{27}^{k}+\delta_{27}^{k+9}+\delta_{27}^{k+18},\quad k\in[1,8].

    As kk goes from 11 to 88, one sees easily that δ2719,δ2720,,δ2726\delta_{27}^{19},\delta_{27}^{20},\cdots,\delta_{27}^{26} have to be removed.

Now we are able to construct a basis of Ω\Omega. Using Corollary IV.2, the basis elements can be obtained. A few leading elements are listed as follows.

d1=1,d2=δ21,d3=δ31,d4=δ32,d5=δ41,d6=δ51,d7=δ52,d8=δ53,d9=δ54,d10=δ61,d11=δ65,d12=δ71,d13=δ72,d14=δ73,d15=δ74,d16=δ75,d17=δ76,d18=δ81,d19=δ83,d20=δ91,d21=δ92,d22=δ94,d23=δ95,d24=δ101,d25=δ103,d26=δ107,d27=δ109,\displaystyle\begin{array}[]{ccccc}d_{1}=1,&d_{2}=\delta_{2}^{1},&d_{3}=\delta_{3}^{1},&d_{4}=\delta_{3}^{2},&d_{5}=\delta_{4}^{1},\\ d_{6}=\delta_{5}^{1},&d_{7}=\delta_{5}^{2},&d_{8}=\delta_{5}^{3},&d_{9}=\delta_{5}^{4},&d_{10}=\delta_{6}^{1},\\ d_{11}=\delta_{6}^{5},&d_{12}=\delta_{7}^{1},&d_{13}=\delta_{7}^{2},&d_{14}=\delta_{7}^{3},&d_{15}=\delta_{7}^{4},\\ d_{16}=\delta_{7}^{5},&d_{17}=\delta_{7}^{6},&d_{18}=\delta_{8}^{1},&d_{19}=\delta_{8}^{3},&d_{20}=\delta_{9}^{1},\\ d_{21}=\delta_{9}^{2},&d_{22}=\delta_{9}^{4},&d_{23}=\delta_{9}^{5},&d_{24}=\delta_{10}^{1},&d_{25}=\delta_{10}^{3},\\ d_{26}=\delta_{10}^{7},&d_{27}=\delta_{10}^{9},&\cdots&{}\hfil&{}\hfil\\ \end{array} (91)

Since Ω\Omega is an inner product, using Gram-Schmidt orthonormalization algorithm, we can get orthonormal basis as

e1=1,e2=(1,1)T,e3=1/2(2,1,1)T,e4=3/2(0,1,1)T,e5=2(1,0,1,0)T,e6=12(4,1,1,1,1)T,ϵ7=5/12(0,3,1,1,1)T,ϵ8=5/6(0,0,2,1,1)Tϵ9=5/2(0,0,0,1,1)T,\displaystyle\begin{array}[]{l}e_{1}=1,~{}~{}e_{2}=(1,-1)^{T},\\ e_{3}=\sqrt{1/2}(2,-1,-1)^{T},~{}~{}e_{4}=\sqrt{3/2}(0,1,-1)^{T},\\ e_{5}=\sqrt{2}(1,0,-1,0)^{T},~{}~{}e_{6}=\frac{1}{2}(4,-1,-1,-1,-1)^{T},\\ \epsilon_{7}=\sqrt{5/12}(0,3,-1,-1,-1)^{T},\\ \epsilon_{8}=\sqrt{5/6}(0,0,2,-1,-1)^{T}\\ \epsilon_{9}=\sqrt{5/2}(0,0,0,1,-1)^{T},~{}~{}\cdots\\ \end{array} (98)
Remark IV.4
  • (i)

    The orthonormal basis can easily be obtained up to finite terms via computer numerically.

  • (ii)

    Using orthonormal basis, Ω\Omega can be imbedded isometrically into Hilbert space 2\ell_{2} [34].

  • (iii)

    Inspired that the discussion of this subsection is about Ω=Ω\Omega=\Omega_{\ell}, since the elements in the basis are neither left reducible no right reducible, it is easy to see that the basis of Ω\Omega_{\ell} is also the basis of Ωr\Omega_{r}.

IV-B Norms of Signal Space

Consider x=(x1,x2,)x=(x_{1},x_{2},\cdots), which is a sequence of real numbers, denote it by EE^{\infty}, which is a vector space. We can define a set of norms on EE^{\infty} as

  • 0\ell_{0} norm:

    x0:=w(x).\displaystyle\|x\|_{0}:=w(x). (99)

    Then

    0:={xE|x0<}.\displaystyle\ell_{0}:=\{x\in E^{\infty}\;|\;\|x\|_{0}<\infty\}. (100)
  • p\ell_{p} norm:

    xp:=(i=1|xi|p)1/p,p=1,2,.\displaystyle\|x\|_{p}:=\left(\mathop{\sum}\limits_{i=1}^{\infty}|x_{i}|^{p}\right)^{1/p},\quad p=1,2,\cdots. (101)

    Then

    p:={xE|xp<},quadp=1,2,.\displaystyle\ell_{p}:=\{x\in E^{\infty}\;|\;\|x\|_{p}<\infty\},quadp=1,2,\cdots. (102)

Then the following facts are well known [28].

Proposition IV.5
  • (i)

    0\ell_{0} is a Frèchet space ***On a vector space VV, a mapping \|\cdot\| is called a pseudo-norm, if x0,andx=0impliesx=0.\|x\|\geq 0,~{}\mbox{and}~{}\|x\|=0~{}\mbox{implies}~{}x=0. x+yx+y,x,yV,\|x+y\|\leq\|x\|+\|y\|,\quad x,y\in V, x=x.\|-x\|=\|x\|. A complete pseudo-normed space is called a Frèchet space.

  • (ii)

    p\ell_{p}, p1p\geq 1 are Banach spaces.

  • (iii)

    2\ell_{2} is an Hilbert space.

Now go back to the set of signals. If a signal xnx\in{\mathbb{R}}^{n}. Then it is easy to be embedded into EE^{\infty} as

ψ:x(x1,,xn,0,0).\psi:x\mapsto(x_{1},\cdots,x_{n},0,0\cdots).

In this way, xpx\in\ell_{p}, p0\forall p\geq 0. This ψ\psi is commonly used in signal processing community. But if we consider infinite-dimensional signal or the infinite union of finite dimensional signals [19], we must be very careful.

Note that EE^{\infty} is a vector space, but it is not a topological space. All p\ell_{p} are its subspaces. Using the distance deduced by the norm (pseudo-norm), all p\ell_{p} become topological spaces. The topologies deduced by the distances are denoted by 𝒯p{\mathcal{T}}_{\ell_{p}}.

Consider {\mathbb{R}}^{\infty}. As aforementioned on {\mathbb{R}}^{\infty} there are two commonly used topologies: natural topology (𝒯n{\mathcal{T}}_{n}) and distance deduced topology (calTd{calT}_{d}).

Consider the natural topology first. Under this topology, any two points p,qp,q\in{\mathbb{R}}^{\infty}, pqp\neq q, are separable (under T2T_{2}, or Hausdorff sense). Hence,

ψ()p,p0.\displaystyle\psi\left({\mathbb{R}}^{\infty}\right)\subset\ell_{p},\quad\forall p\geq 0. (103)

Because each element in {\mathbb{R}}^{\infty} has only finite terms.

The mapping ψ\psi, defined by (103) is one-to-one, hence ψ\psi can be considered as an imbedding. And then we can pose the subspace topology of p\ell_{p} to {\mathbb{R}}^{\infty}. For instance, let p=2p=2. Then {\mathbb{R}}^{\infty} is an inner product space. But it is not a Hilbert space, because it is not complete.

All spaces

(,𝒯p),p0,({\mathbb{R}}^{\infty},{\mathcal{T}}_{\ell_{p}}),\quad p\geq 0,

are topological spaces. Some properties follow from the definition immediately.

Proposition IV.6
  • (i)

    (,𝒯p)({\mathbb{R}}^{\infty},{\mathcal{T}}_{\ell_{p}}) homeomorphic to neither (,𝒯n)({\mathbb{R}}^{\infty},{\mathcal{T}}_{n}) no (,𝒯d)({\mathbb{R}}^{\infty},{\mathcal{T}}_{d}).

  • (ii)

    If pqp\neq q, then (,𝒯p)({\mathbb{R}}^{\infty},{\mathcal{T}}_{\ell_{p}}) does not homeomorphic to (,𝒯q)({\mathbb{R}}^{\infty},{\mathcal{T}}_{\ell_{q}}).

  • (iv)

    Consider RnR^{n} , which is considered as the space of signals with fixed length. Assume it has the subspace topology of {\mathbb{R}}^{\infty}, then

    (n,𝒯p|n)(n,𝒯n|n)(n,𝒯d|n).\displaystyle({\mathbb{R}}^{n},{\mathcal{T}}_{\ell_{p}}|_{{\mathbb{R}}^{n}})\cong({\mathbb{R}}^{n},{\mathcal{T}}_{n}|_{{\mathbb{R}}^{n}})\cong({\mathbb{R}}^{n},{\mathcal{T}}_{d}|_{{\mathbb{R}}^{n}}). (104)
Remark IV.7

(104) shows that when the signals have a fixed dimension, all the vector space topologies are equivalent. Only the dimension varying signals or infinity dimensional signals are considered, the different topologies become meaningful.

Next, we consider the distance deduced topology (calTd{calT}_{d}).

Assume {ei|i=1,2,}\{e_{i}\;|\;i=1,2,\cdots\} is an orthonormal basis of Ω=/\Omega={\mathbb{R}}^{\infty}/\leftrightarrow . Consider x=(x1,,xn)Tnx=(x_{1},\cdots,x_{n})^{T}\in{\mathbb{R}}^{n}. Then

x¯=i=1mnξiei,\bar{x}=\mathop{\sum}\limits_{i=1}^{m_{n}}\xi_{i}e_{i},

where,

mn=|𝒟𝒱n|n.m_{n}=|{\mathcal{D}}^{n}_{{\mathcal{V}}}|\leq n.

Define ϕ:E\phi:{\mathbb{R}}^{\infty}\rightarrow E^{\infty} as

ϕ(x):=(ξ1,ξ2,,ξmn,0,0,)\displaystyle\phi(x):=(\xi_{1},\xi_{2},\cdots,\xi_{m_{n}},0,0,\cdots) (105)

Now consider {\mathbb{R}}^{\infty}. It is obvious that

ψ()p,p0.\displaystyle\psi\left({\mathbb{R}}^{\infty}\right)\subset\ell_{p},\quad\forall p\geq 0. (106)

Because each element is {\mathbb{R}}^{\infty} has only finite terms.

Since the ψ\psi in (106) is one=to-one, then ψ\psi can be considered as an embedding. Hence, we can pose {\mathbb{R}}^{\infty} the subspace topology of p\ell_{p}. For instance, we assume p=2p=2. Then {\mathbb{R}}^{\infty} becomes an inner product space. Note that

(,𝒯p),p0,({\mathbb{R}}^{\infty},{\mathcal{T}}_{\ell_{p}}),\quad p\geq 0,

are all topological spaces. According to the definitions, the following proposition is easily verifiable.

Proposition IV.8
  • (i)

    (,𝒯p)({\mathbb{R}}^{\infty},{\mathcal{T}}_{\ell_{p}}) is homeomorphic to neither (,𝒯n)({\mathbb{R}}^{\infty},{\mathcal{T}}_{n}) no (,𝒯d)({\mathbb{R}}^{\infty},{\mathcal{T}}_{d}).

  • (ii)

    If pqp\neq q, then (,𝒯p)({\mathbb{R}}^{\infty},{\mathcal{T}}_{\ell_{p}}) is not homeomorphic to (,𝒯q)({\mathbb{R}}^{\infty},{\mathcal{T}}_{\ell_{q}}).

  • (iv)

    Consider RnR^{n} with the inherited (subspace) topology of {\mathbb{R}}^{\infty}. Then

    (n,𝒯p|n)(n,𝒯n|n)(n,𝒯d|n).\displaystyle({\mathbb{R}}^{n},{\mathcal{T}}_{\ell_{p}}|_{{\mathbb{R}}^{n}})\cong({\mathbb{R}}^{n},{\mathcal{T}}_{n}|_{{\mathbb{R}}^{n}})\cong({\mathbb{R}}^{n},{\mathcal{T}}_{d}|_{{\mathbb{R}}^{n}}). (107)
Remark IV.9

(107) shows that when signals of fixed dimension are considered, any topologies are the same. Only when the dimension-varying signals are investigated, the different topologies may cause different results.

V Dimension-free STP-CS

According to Corollary III.16, if all signals are expressed as elements in signal space, the left STP-CS can be formally expressed

y¯=Ax¯,\displaystyle\bar{y}=\langle A\rangle_{\ell}\vec{\ltimes}\bar{x}, (108)

where x¯Ω\bar{x}\in\Omega_{\ell} is the original signal, y¯Ω\bar{y}\in\Omega_{\ell} is the sampled data, and AΣ\langle A\rangle_{\ell}\in\Sigma_{\ell} is the sensing matrix.

We call (108) a dimension-free STP-CS, because it can be used to treat signals of arbitrary dimensions.

Correspondingly, if we take Ωr\Omega_{r} as the signal space, then the right STP-CS can be expressed as

y¯=Arx¯,\displaystyle\bar{y}=\langle A\rangle_{r}\vec{\ltimes}\bar{x}, (109)

where x¯Ωr\bar{x}\in\Omega_{r} is the original signal, y¯Ωr\bar{y}\in\Omega_{r} is the sampled data, and ArΣr\langle A\rangle_{r}\in\Sigma_{r} is the sensing matrix.

Both the expressions (108) and (109) are dimension-free. That is, there is no dimension restriction on the original or compressed signal. But in practical use, dimension depending expression is more convenient. Let us consider a particular (matrix-vector) form for STP-CS: We can w.l.g., assume that AA is left irreducible. Otherwise, we can use irreducible A0AA_{0}\in\langle A\rangle to replace AA.

Proposition V.1
  • (i)

    Assume xpx\in{\mathbb{R}}^{p}, Am×nA\in{\mathcal{M}}_{m\times n}, and p=snp=sn. Then (108) becomes

    y=(AIs)x,\displaystyle y=(A\otimes I_{s})x, (110)

    which is essentially the same as (2).

  • (ii)

    Assume xpx\in{\mathbb{R}}^{p}, Am×nA\in{\mathcal{M}}_{m\times n}, and lcm(n,p)=t\operatorname{lcm}(n,p)=t, where t=sn=rpt=sn=rp, r>1r>1. Then (108) becomes

    y=(AIs)(x𝟏r).\displaystyle y=(A\otimes I_{s})(x\otimes{\bf 1}_{r}). (111)
  • (iii)

    Dimension-varying signal: Assume D={dii[1,]}+}D=\{d_{i}\>\>i\in[1,\ell]\}\subset{\mathbb{Z}}^{+}\} is a finite set, and xpx\in{\mathbb{R}}^{p}, pDp\in D. Then there exists a switching signal σ(t)D\sigma(t)\in D, t0t\geq 0, such that

    y(t)=(AIs(σ(t)))(x(t)𝟏r(σ(t))).\displaystyle y(t)=(A\otimes I_{s(\sigma(t))})(x(t)\otimes{\bf 1}_{r(\sigma(t))}). (112)

Correspondingly, using right system, we have

Proposition V.2
  • (i)

    Assume xpx\in{\mathbb{R}}^{p}, Am×nA\in{\mathcal{M}}_{m\times n}, and p=snp=sn. Then (109) becomes

    y=(IsA)x.\displaystyle y=(I_{s}\otimes A)x. (113)
  • (ii)

    Assume xpx\in{\mathbb{R}}^{p}, Am×nA\in{\mathcal{M}}_{m\times n}, and lcm(n,p)=t\operatorname{lcm}(n,p)=t, where t=sn=rpt=sn=rp, r>1r>1. Then (109) becomes

    y=(IsA)(𝟏rx).\displaystyle y=(I_{s}\otimes A)({\bf 1}_{r}\otimes x). (114)
  • (iii)

    Dimension-varying signal: Assume D={dii[1,]}+}D=\{d_{i}\>\>i\in[1,\ell]\}\subset{\mathbb{Z}}^{+}\} is a finite set, and xpx\in{\mathbb{R}}^{p}, pDp\in D. Then there exists a switching signal σ(t)D\sigma(t)\in D, t0t\geq 0, such that

    y(t)=(Is(σ(t))A)(𝟏r(σ(t))x(t)).\displaystyle y(t)=(I_{s(\sigma(t))}\otimes A)({\bf 1}_{r(\sigma(t))}\otimes x(t)). (115)

In practical use, the conventional matrix expression is necessary.

  • (i)

    Case 1:

    Assume n|pn|p with p=nsp=ns.

This case is commonly assumed in current literature.

We consider the right STP-CS first. Then we have (113).

Proposition V.3

Consider (113). Then

spark(IsA)=spark(A).\displaystyle spark(I_{s}\otimes A)=spark(A). (116)

Proof. Since

IsA=[A000A000A].\displaystyle I_{s}\otimes A=\begin{bmatrix}A&0&\cdots&0\\ 0&A&\cdots&0\\ \vdots&~{}&~{}&~{}\\ 0&0&\cdots&A\\ \end{bmatrix}. (117)

Then it is clear that the smallest set of linearly dependent columns can only be found within each AA. This fact leads to (116).

\Box

Using Proposition I.1, we have the following result.

Corollary V.4

Consider (113). If spark(A)>2kspark(A)>2k, then for each ysmy\in{\mathbb{R}}^{sm} there is at most one solution xΣkpx\in\Sigma^{p}_{k}.

In fact, we can get a better result for STP-CS.

Let x=(x1,x2,,xs)Tx=(x^{1},x^{2},\cdots,x^{s})^{T}, where xinx^{i}\in{\mathbb{R}}^{n}, i[1,s]i\in[1,s]. Define

Σk/np:={x=(x1,x2,,xs)Tp|xiΣkn}.\displaystyle\Sigma^{p}_{k/n}:=\{x=(x^{1},x^{2},\cdots,x^{s})^{T}\in{\mathbb{R}}^{p}\;|\;\forall x^{i}\in\Sigma^{n}_{k}\}. (118)
Proposition V.5

Consider (113). If spark(A)>2kspark(A)>2k, then for each ysmy\in{\mathbb{R}}^{sm} there is at most one solution xΣk/npx\in\Sigma^{p}_{k/n}.

Proof. Using (117) we can get the following decomposed system as

{y1=Ax1,y2=Ax2,,ys=Ax2.\displaystyle\begin{cases}y^{1}=Ax^{1},\\ y^{2}=Ax^{2},\\ \vdots,\\ y^{s}=Ax^{2}.\end{cases} (119)

Using Proposition I.1 to each sub-equations, the conclusion follows.

\Box

Next, we consider left STP-CS. Then we have (110).

Lemma V.6

[10] Let Am×nA\in{\mathcal{M}}_{m\times n}, Bp×qB\in{\mathcal{M}}_{p\times q}. Then

W[m,p](AB)Wq,n=BA.\displaystyle W_{[m,p]}(A\otimes B)W_{q,n}=B\otimes A. (120)

Define

z(x)=W[n,s]x:=(z1,z2,,zs)T.z(x)=W_{[n,s]}x:=(z^{1},z^{2},\cdots,z^{s})^{T}.

It is well known that zz is obtained from xx by an element permutation. Then similarly to Proposition V.5, we have the following result.

Proposition V.7

Consider (110). If spark(A)>2kspark(A)>2k, then for each ysmy\in{\mathbb{R}}^{sm} there is at most one solution xx with z(x)Σk/npz(x)\in\Sigma^{p}_{k/n}.

Proof. Using Lemma V.6, we have

(AIs)x=W[s,m](IsA)W[n,s]x=W[s,m](IsA)z=y.\displaystyle\begin{array}[]{l}(A\otimes I_{s})x\\ =W_{[s,m]}(I_{s}\otimes A)W_{[n,s]}x\\ =W_{[s,m]}(I_{s}\otimes A)z\\ =y.\end{array} (125)

(125) is equivalent to

(IsA)z=W[s,m]1y=W[m,s]y.\displaystyle(I_{s}\otimes A)z=W^{-1}_{[s,m]}y=W_{[m,s]}y. (126)

The conclusion follows from Proposition V.5 immediately.

\Box

  • (ii)

    Case 2: Assume npn\nmid p with lcm(n,p)=t\operatorname{lcm}(n,p)=t.

Then the (108) becomes

(AIt/n)x~=y~,\displaystyle(A\otimes I_{t/n})\tilde{x}=\tilde{y}, (127)

where x~=x𝟏t/p\tilde{x}=x\otimes{\bf 1}_{t/p}, y~=y𝟏tm/n\tilde{y}=y\otimes{\bf 1}_{tm/n}. (127) has exactly the same form as (113).

the (109) becomes

(It/nA)x~=y~,\displaystyle(I_{t/n}\otimes A)\tilde{x}=\tilde{y}, (128)

where x~=𝟏t/px\tilde{x}={\bf 1}_{t/p}\otimes x, y~=𝟏tm/ny\tilde{y}={\bf 1}_{tm/n}\otimes y. (128) has exactly the same form as (110).

We conclude that Case 2 can be converted into Case 1. Then the results for Case 1 are available for Case 2.

Definition V.8

Let AΩ\langle A\rangle\in\Omega_{\ell} (or AΩr\langle A\rangle\in\Omega_{r} ). Then

spark(A):=spark(A0),(orspark(Ar):=spark(A0),\displaystyle spark(\langle A\rangle_{\ell}):=spark(A_{0}),\quad(\mbox{or}~{}spark(\langle A\rangle_{r}):=spark(A_{0}), (129)

where A0AA_{0}\in\langle A\rangle_{\ell} is left irreducible (A0ArA_{0}\in\langle A\rangle_{r} is right irreducible) .

Next, we consider some further properties.

Definition V.9

(Coherence)

Assume A=A0IsA=A_{0}\otimes I_{s}, where A0A_{0} is left irreducible (or A=ItA0A=I_{t}\otimes A_{0}, where A0A_{0} is right irreducible). Then the coherence of A\langle A\rangle is defined as

μ(A):=μ(A0).\displaystyle\mu(\langle A\rangle):=\mu(A_{0}). (130)

Finally, we consider the RIP. Denote by

Σk/n=pnΣk/np.\displaystyle\Sigma_{k/n}=\bigcup_{p\geq n}\Sigma^{p}_{k/n}. (131)

Then we can define dimension-free RIP as follows.

Definition V.10

A\langle A\rangle with A0m×nA_{0}\in{\mathcal{M}}_{m\times n} is said to satisfy the (k,δ)(k,\delta)-RIP, if there exists δ=δkA(0,1)\delta=\delta_{k}^{A}\in(0,1), such that

(1δ)x𝒱2A0x𝒱2(1+δ)x𝒱2,xΣk/n.\displaystyle(1-\delta)\|x\|_{{\mathcal{V}}}^{2}\leq\|A_{0}\vec{\rtimes}x\|_{{\mathcal{V}}}^{2}\leq(1+\delta)\|x\|_{{\mathcal{V}}}^{2},\quad\forall x\in\Sigma_{k/n}. (132)
Remark V.11
  • (i)

    It is easy to verify that for any AAA\in\langle A\rangle, the coherence of AA is the same. Hence the Definition V.9 is reasonable. In fact, A0A_{0} can be replaced by any AAA\in\langle A\rangle.

  • (ii)

    The inner product and norm can also be replaced by ,𝒱\langle\cdot,\cdot\rangle_{{\mathcal{V}}} and 𝒱\|\cdot\|_{{\mathcal{V}}}. which are related to the topological structure of Ω\Omega.

  • (iii)

    Definition V.9 seems more restrictive than the original definition for fixed dimensional case. In fact, it is not. Note that if ABA\sim B and xyx\leftrightarrow y, then

    Ax𝒱=By𝒱.\|A\ltimes x\|_{{\mathcal{V}}}=\|B\ltimes y\|_{{\mathcal{V}}}.

    Hence if A0A_{0} satisfies the (k,δ)(k,\delta)-RIP as defined in Definition I.2, then A\langle A\rangle satisfies the (k,δ)(k,\delta)-RIP as defined in Definition V.10.Hence, the fixed dimension (k,δ)(k,\delta)-RIP implies the dimension-free (k,δ)(k,\delta)-RIP, which ensures the precise reconstruction for much more signals.

VI BIBD-Based Sensing Matrix

VI-A BIBD-Based Construction

This subsection gives a briefly review for the construction of sensing matrix based on balanced incomplete block design (BIBD), proposed in [25].

Definition VI.1

Consider a matrix Aα×βA\in{\mathcal{M}}_{\alpha\times\beta}.

  • (i)

    AA is called a sign matrix if

    ai,j{1,1},i[1,α],j[1,β].a_{i,j}\in\{1,-1\},\forall i\in[1,\alpha],j\in[1,\beta].
  • (ii)

    AA is called a Boolean matrix if

    ai,j{1,0},i[1,α],j[1,β].a_{i,j}\in\{1,0\},\forall i\in[1,\alpha],j\in[1,\beta].

    The column degree of AA is denoted by

    dc(Colj(A))=i=1αai,j,j[1,β].d_{c}(\operatorname{Col}_{j}(A))=\mathop{\sum}\limits_{i=1}^{\alpha}a_{i,j},\quad j\in[1,\beta].

    The row degree of AA is denoted by

    dr(Rowi(A))=j=1βai,j,i[1,α].d_{r}(\operatorname{Row}_{i}(A))=\mathop{\sum}\limits_{j=1}^{\beta}a_{i,j},\quad i\in[1,\alpha].
Definition VI.2

[25] Let X={x1,x2,,xα}X=\{x_{1},x_{2},\cdots,x_{\alpha}\} and ={P1,P2,,Pβ}2X\mathbb{P}=\{P_{1},P_{2},\cdots,P_{\beta}\}\subset 2^{X}, i.e., each block Pj2XP_{j}\in 2^{X}, j[1,β]j\in[1,\beta].

  • (i)

    Hα×βH\in{\mathcal{B}}_{\alpha\times\beta} is defined by

    hi,j={1,xiPj,0,Otherwise.h_{i,j}=\begin{cases}1,\quad x_{i}\in P_{j},\\ 0,\quad Otherwise.\end{cases}

    HH is called an incidence matrix.

  • (ii)

    \mathbb{P} is said to have BIBD, if its index matrix satisfies the following conditions.

    • Each element xix_{i} appears exactly in rr blocks.

    • Each block contains exactly 2k<α2\leq k<\alpha elements, i.e., |Pj|=k|P_{j}|=k, j\forall j.

    • Each pair of elements in X appears exactly in λ\lambda blocks.

    Precisely, \mathbb{P} is said to have a (α,β,r,k,λ)(\alpha,\beta,r,k,\lambda)-BIBD.

Assume X={x1,x2,,xα}X=\{x_{1},x_{2},\cdots,x_{\alpha}\} and ={P1,P2,,Pα}\mathbb{P}=\{P_{1},P_{2},\cdots,P_{\alpha}\}. [25] proposed a way to construct a deterministic sensing matrix as follows.

Assume \mathbb{P} has (α,α,α1,α1,α2)(\alpha,\alpha,\alpha-1,\alpha-1,\alpha-2)-BIBD, then its incidence matrix can be expressed as

H=[1110110110110111]\displaystyle H=\begin{bmatrix}1&1&\cdots&1&0\\ 1&1&\cdots&0&1\\ \vdots&~{}&~{}&~{}&~{}\\ 1&0&\cdots&1&1\\ 0&1&\cdots&1&1\\ \end{bmatrix} (133)

The sensing matrix AA is constructed through two expansions.

  • Vertical Expansion:

Algorithm VI.3

Assume an incidence matrix HH is given.

  • Step 1: Keep first column unchanged. Start from second column. Keep the first 11 in column 2 unchanged. If the second 11 meets 0 at the same row of the first column, keep this 11 unchanged. Otherwise, move all remaining elements (including this 11) down, till its same row element in first column being 0. Then do the same thing for third 11 and keep doing for all other 11 elements one by one in the order.

  • Step 2: For third column. Allow it has one 11, which is on the same row of the 11 for first and second columns. Otherwise, move this 11 and all the below elements down. Until the above requirement satisfies, which keeps the inner product of the first or second column with third one is 11.

  • Step k-1: For kk-th column, similarly to third column, as long as the inner product of kk-th column with one of the first k1k-1 columns is greater than 11, move the corresponding 11 with all below elements down, until the inner product requirement is satisfied.

We give a simple example to depict this algorithm.

Example VI.4

[25] Assume α=4\alpha=4, the incidence matrix is

H=[1110110110110111]H=\begin{bmatrix}1&1&1&0\\ 1&1&0&1\\ 1&0&1&1\\ 0&1&1&1\\ \end{bmatrix}

Its vertical expanded matrix becomes

Hv=[1110100110000101010000110010]H_{v}=\begin{bmatrix}1&1&1&0\\ 1&0&0&1\\ 1&0&0&0\\ 0&1&0&1\\ 0&1&0&0\\ 0&0&1&1\\ 0&0&1&0\\ \end{bmatrix}

It was proved in [25] that Hvm×αH_{v}\in{\mathcal{M}}_{m\times\alpha}, where

m=α23α+3.\displaystyle m=\alpha^{2}-3\alpha+3. (134)

Moreover, the coherence of HvH_{v} is

μ(Hv)=1α1.\displaystyle\mu(H_{v})=\frac{1}{\alpha-1}. (135)
  • Horizontal Expansion:

Definition VI.5

Let A(α1)×rA\in{\mathcal{M}}_{(\alpha-1)\times r} be a sign matrix, and Dr×rD\in{\mathcal{M}}_{r\times r} be a nonsingular diagonal matrix, say D=diag(d1,d2,,dr)D=diag(d_{1},d_{2},\cdots,d_{r}), where ds>0d_{s}>0, s[1,r]s\in[1,r] and didjd_{i}\neq d_{j}, iji\neq j. Then B:=ADB:=AD is called an embedding matrix.

Algorithm VI.6

Assume the vertically expanded incident matrix HvH_{v} and an embedding matrix BB are given.

For each column of HvH_{v}, say, Colj(Hv)\operatorname{Col}_{j}(H_{v}), its first 11 is replaced by Row1(B)\operatorname{Row}_{1}(B), second 11 is replaced by Row2(B)\operatorname{Row}_{2}(B), \cdots, till last 11 is replaced by Rowα1(B)\operatorname{Row}_{\alpha-1}(B). 0 is replaced by 0,0,,0r\underbrace{0,0,\cdots,0}_{r}. The resulting matrix is expressed as

A:=HvB.\displaystyle A:=H_{v}\odot B. (136)

The following result and the estimation (6) are fundamental for this design.

Theorem VI.7

[2] Let Hα×βH\in{\mathcal{M}}_{\alpha\times\beta} be an incidence matrix with column degree

dc(Colj(H))=d,j[1,β],d_{c}(\operatorname{Col}_{j}(H))=d,\quad j\in[1,\beta],

Bd×sB\in{\mathcal{M}}_{d\times s} be an embedding matrix, and Φ=HB\Phi=H\odot B. Then

μ(Φ)=max{μ(H),μ(B)}.\displaystyle\mu(\Phi)=\max\{\mu(H),\mu(B)\}. (137)
***The original requirement for BB is [2] “the elements of BB have the same absolute values in the same column, but the elements have different absolute values in different columns.” Such matrices can be constructed as in the Definition 137 for embedding matrix.
Corollary VI.8

Using BIBD-based design, the best solution is

μ(Φ)=μ(Hv)=1α1,\displaystyle\mu(\Phi)=\mu(H_{v})=\frac{1}{\alpha-1}, (138)

which can be reached as long as

μ(B)1α1.\displaystyle\mu(B)\leq\frac{1}{\alpha-1}. (139)
Example VI.9

Recall Example VI.4. It is easy to find an embedding matrix as

B=[111111111111]\displaystyle B=\begin{bmatrix}1&1&1&-1\\ 1&-1&1&1\\ 1&1&-1&1\end{bmatrix} (140)

. Then the sensing matrix can be constructed as

Φ:=HvB=[1111111111110000111100000000111100001111000000000000000011110000000011110000000000001111000000001111111111110000].\displaystyle\begin{array}[]{l}\Phi:=H_{v}\odot B\\ =\left[\begin{array}[]{cccccccc}1&1&1&-1&1&1&1&-1\\ 1&-1&1&1&0&0&0&0\\ 1&1&-1&1&0&0&0&0\\ 0&0&0&0&1&-1&1&1\\ 0&0&0&0&1&1&-1&1\\ 0&0&0&0&0&0&0&0\\ 0&0&0&0&0&0&0&0\\ \end{array}\right.\\ \left.\begin{array}[]{cccccccc}1&1&1&-1&0&0&0&0\\ 0&0&0&0&1&1&1&-1\\ 0&0&0&0&0&0&0&0\\ 0&0&0&0&1&-1&1&1\\ 0&0&0&0&0&0&0&0\\ 1&-1&1&1&1&1&-1&1\\ 1&1&-1&1&0&0&0&0\\ \end{array}\right].\\ \end{array} (158)

.

It is easy to verify that

μ(B)=1α1=13.\mu(B)=\frac{1}{\alpha-1}=\frac{1}{3}.

Hence,

μ(Φ)=μ(Hv)=13.\mu(\Phi)=\mu(H_{v})=\frac{1}{3}.

[29] provides some results for constructing matrix BB, which meets (139). Some other examples are Hadamard matrix or DFT matrix and certain their generalizations. Unfortunately, the latter are not embedding matrix. We will look for embedding matrices which meets (139).

VI-B A modified BIBD-Based Sensing Matrix

This subsection aims on improving the BIBD-based sensing matrix.

First, we improve the vertical expansion.

Definition VI.10

Let Hα×αH\in{\mathcal{M}}_{\alpha\times\alpha} be an incident matrix with column degree α1\alpha-1. A new vertical expansion, denoted by HH_{*}, is defined as follows.

H=[H1H2Hα1]m×α,\displaystyle H_{*}=\begin{bmatrix}H_{1}\\ H_{2}\\ \vdots\\ H_{\alpha-1}\end{bmatrix}\in{\mathcal{M}}_{m\times\alpha}, (159)

where

H1=[𝟏α1,Iα1],H2=[0α2,𝟏α2,Iα2],Hα2=[02,,02α3,𝟏2,I2],Hα1=[0,,0α2,1,1].\begin{array}[]{l}H_{1}=\left[{\bf 1}_{\alpha-1},I_{\alpha-1}\right],\\ H_{2}=\left[0_{\alpha-2},{\bf 1}_{\alpha-2},I_{\alpha-2}\right],\\ \vdots\\ H_{\alpha-2}=\left[\underbrace{0_{2},\cdots,0_{2}}_{\alpha-3},{\bf 1}_{2},I_{2}\right],\\ H_{\alpha-1}=\left[\underbrace{0,\cdots,0}_{\alpha-2},1,1\right].\end{array}
Example VI.11

Recall the HH in Example VI.4. Then

H=[110010101001011001010011]H_{*}=\begin{bmatrix}1&1&0&0\\ 1&0&1&0\\ 1&0&0&1\\ 0&1&1&0\\ 0&1&0&1\\ 0&0&1&1\end{bmatrix}

The following proposition comes from the construction immediately.

Proposition VI.12

Let Hα×αH\in{\mathcal{M}}_{\alpha\times\alpha} be an incident matrix with column degree α1\alpha-1. Then the vertically expanded matrix HH_{*} satisfies the following conditions.

  • (i)

    Hm×αH_{*}\in{\mathcal{M}}_{m\times\alpha}, where

    m=α(α1)2.\displaystyle m=\frac{\alpha(\alpha-1)}{2}. (160)
  • (ii)
    dc(Colj(H)=α1,j[1,α];dr(Rowi(H)=2,i[1,m].\displaystyle\begin{array}[]{l}d_{c}(\operatorname{Col}_{j}(H_{*})=\alpha-1,\quad j\in[1,\alpha];\\ d_{r}(\operatorname{Row}_{i}(H_{*})=2,\quad i\in[1,m].\\ \end{array} (163)
  • (iii)

    The coherence

    μ(H)=1α1.\mu(H_{*})=\frac{1}{\alpha-1}.

Comparing with HvH_{v}, it is obvious that HH_{*} has the same column degree and the same coherence with HvH_{v}. Moreover, HH_{*} has less rows (α(α1)2\frac{\alpha(\alpha-1)}{2}) comparing with HvH_{v} (α23α+3\alpha^{2}-3\alpha+3). The smaller the mm is, the higher the compress rate is obtained.

Next, we consider how to construct an embedding matrix to meet (139).

Lemma VI.13

Assume B=ADB=AD is an embedding matrix, where Aα×βA\in{\mathcal{M}}_{\alpha\times\beta} is a sign matrix, and D=diag(d1,,dβ)D=diag(d_{1},\cdots,d_{\beta}) is a nonsingular diagonal matrix with di>0d_{i}>0, i[1,β]i\in[1,\beta] and didjd_{i}\neq d_{j}, iji\neq j. Then

μ(B)=μ(A).\displaystyle\mu(B)=\mu(A). (164)

Proof. Denote Ai:=Coli(A)A_{i}:=\operatorname{Col}_{i}(A).

μ(Bi,Bj)=diAi,djAjdiAidjAj|=|di||dj|Ai,Aj|di||dj|AiAj|=μ(Ai,Aj),ij.\begin{array}[]{ccl}\mu(B_{i},B_{j})&=&\frac{\langle d_{i}A_{i},d_{j}A_{j}\rangle}{\|d_{i}A_{i}\|\|d_{j}A_{j}|}\\ {}\hfil&=&\frac{|d_{i}||d_{j}|\langle A_{i},A_{j}\rangle}{|d_{i}||d_{j}|\|A_{i}\|\|A_{j}|}\\ {}\hfil&=&\mu(A_{i},A_{j}),\quad i\neq j.\end{array}

The conclusion follows. \Box

It is clear now to construct an embedding matrix, we need only to construct a sign matrix, satisfied (139).

Proposition VI.14

Let Aα×βA\in{\mathcal{M}}_{\alpha\times\beta} be a sign matrix.

  • (i)

    Assume α\alpha is an even number. Then AA satisfies (139), if and only if,

    Ai,Aj=0,ij.\displaystyle\langle A_{i},A_{j}\rangle=0,\quad i\neq j. (165)
  • (ii)

    Assume α\alpha is an odd number. Then AA satisfies (139), if and only if,

    Ai,Aj=±1,ij.\displaystyle\langle A_{i},A_{j}\rangle=\pm 1,\quad i\neq j. (166)

Proof.

  • (i)

    Assume tt is even. Set

    N+=|{k|ak,iak,j=1}|,N=|{k|ak,iak,j=1}|.\begin{array}[]{l}N_{+}=|\{k\;|\;a_{k,i}a_{k,j}=1\}|,\\ N_{-}=|\{k\;|\;a_{k,i}a_{k,j}=-1\}|.\end{array}

    Then

    N+N=0,±2,±4,.N_{+}-N_{-}=0,\pm 2,\pm 4,\cdots.

    Then the corresponding coherence

    μ(Ai,Aj)=0,2/α,4/α,.\mu(A_{i},A_{j})=0,2/\alpha,4/\alpha,\cdots.

    Hence only when N+=NN_{+}=N_{-}, (139) is satisfied, which leads to (165).

  • (ii)

    Assume α\alpha is odd. Then

    N+N=±1,±3,±5,.N_{+}-N_{-}=\pm 1,\pm 3,\pm 5,\cdots.

    The corresponding coherence

    μ(Ai,Aj)=1/α,3/α,5/α,.\mu(A_{i},A_{j})=1/\alpha,3/\alpha,5/\alpha,\cdots.

    Hence only when N+N=±1N_{+}N_{-}=\pm 1, (139) is satisfied, which leads to (166).

\Box

Definition VI.15
  • (i)

    A sign matrix satisfying (165) is called an orthogonal column matrix (OCM).

  • (ii)

    A sign matrix satisfying (166) is called an almost orthogonal column matrix (AOCM).

  • (iii)

    A sign matrix 𝕆αα×β\mathbb{O}_{\alpha}\in{\mathcal{M}}_{\alpha\times\beta} is called a largest OCM (LCOM), if it is an OCM with maximum number of columns.

  • (iv)

    A sign matrix 𝕌αα×β\mathbb{U}_{\alpha}\in{\mathcal{M}}_{\alpha\times\beta} is called a largest AOCM(LACOM), if it is an AOCM with maximum number of columns.

VI-C Constructing 𝕆\mathbb{O} and 𝕌\mathbb{U}

Assume the vertically expanded matrix Hm×αH_{*}\in{\mathcal{M}}_{m\times\alpha} is obtained as in (159). Then we consider how to construct 𝕆\mathbb{O} (or 𝕌\mathbb{U}), where

𝕆t×s,(or𝕆t×s),\mathbb{O}\in{\mathcal{M}}_{t\times s},\quad(\mbox{or}~{}\mathbb{O}\in{\mathcal{M}}_{t\times s}),

where

t=α1.t=\alpha-1.

And we wish s/ts/t to be as large as possible.

Lemma VI.16

Assume AA is an OCM (or AOCM).

  • (i)

    Replacing Colj(A)\operatorname{Col}_{j}(A) by Colj(A)-\operatorname{Col}_{j}(A), the resulting matrix is still an OCM (or AOCM).

  • (ii)

    Replacing Rowi(A)\operatorname{Row}_{i}(A) by Rowi(A)-\operatorname{Row}_{i}(A), the resulting matrix is still an OCM(or AOCM).

  • (iii)

    Doing a row (or column) permutation, the resulting matrix is still an OCM(or AOCM).

Proof. A simple computation shows that all the aforementioned operations do not change the coherence μ(A)\mu(A). The conclusion follows.

\Box

In this subsection, denote by ABA\sim B, if BB is obtained from AA by the transformations defined in Lemma VI.16.

Assume Ξ={ξ1,ξ2,,ξs}\Xi=\{\xi^{1},\xi^{2},\cdots,\xi^{s}\} is a set of tt-dimensional orthogonal vectors. Using Lemma VI.16, we can, w.l.g., assume

ξ1:=𝟏tΞ.\xi^{1}:={\bf 1}_{t}\in\Xi.

Assume tt is even, say t=2t2t=2t_{2}, where t2+t_{2}\in{\mathbb{Z}}_{+} . then, w.l.g., we can assume

ξ2=[11]𝟏t2Ξ.\xi^{2}=\begin{bmatrix}1\\ -1\end{bmatrix}\otimes{\bf 1}_{t_{2}}\in\Xi.

Now assume

ξ3=[η1η2]Ξ.\xi^{3}=\begin{bmatrix}\eta_{1}\\ \eta_{2}\end{bmatrix}\in\Xi.

Since ξ3\xi^{3} is orthogonal with ξ1\xi^{1}, we have

N+(η1)=N(η2),N(η1)=N+(η2).N_{+}(\eta_{1})=N_{-}(\eta_{2}),\quad N_{-}(\eta_{1})=N_{+}(\eta_{2}).

Since ξ3\xi^{3} is orthogonal with ξ2\xi^{2}, we have

N+(η1)=N+(η2),N(η1)=N(η2).N_{+}(\eta_{1})=N_{+}(\eta_{2}),\quad N_{-}(\eta_{1})=N_{-}(\eta_{2}).

We conclude that

N+(η1)=N(η1),N+(η2)=N(η2).\displaystyle N_{+}(\eta_{1})=N_{-}(\eta_{1}),\quad N_{+}(\eta_{2})=N_{-}(\eta_{2}). (167)

(167) implies the following two facts:

  • (i)

    If there exists ξ3\xi^{3}, then t2t_{2} is an even number, and denoted by t2=2t3t_{2}=2t_{3}.

  • (ii)
    η1,η2{[11]𝟏t3,[11]𝟏t3}.\eta_{1},\eta_{2}\in\left\{\begin{bmatrix}1\\ 1\end{bmatrix}\otimes{\bf 1}_{t_{3}},\begin{bmatrix}1\\ -1\end{bmatrix}\otimes{\bf 1}_{t_{3}}\right\}.

Hence, we can exactly have two more elements for Ξ\Xi, which are

ξ3=[(11)𝟏t3(11)𝟏t3];ξ4=[(11)𝟏t3(11)𝟏t3]\xi^{3}=\begin{bmatrix}\begin{pmatrix}1\\ -1\end{pmatrix}\otimes{\bf 1}_{t_{3}}\\ \begin{pmatrix}1\\ -1\end{pmatrix}\otimes{\bf 1}_{t_{3}}\\ \end{bmatrix};\quad\xi^{4}=\begin{bmatrix}\begin{pmatrix}1\\ -1\end{pmatrix}\otimes{\bf 1}_{t_{3}}\\ \begin{pmatrix}-1\\ 1\end{pmatrix}\otimes{\bf 1}_{t_{3}}\\ \end{bmatrix}

Continuing this procedure, the following result about 𝕆\mathbb{O} can be obtained.

As for 𝕌t\mathbb{U}_{t}, since we are not able to find a sign vector xx such that the coherence of xx and 𝟏t{\bf 1}_{t} is ±1\pm 1. 𝕌t=𝕆t\mathbb{U}_{t}=\mathbb{O}_{t}. We, therefore, have the following result.

Theorem VI.17

Assume t=2pqt=2^{p}q, where qq is an odd number.

  • (i)

    𝕆t\mathbb{O}_{t} can be obtained as

    𝕆t=𝕆2𝕆2𝕆2p𝟏qα×2p.\displaystyle\mathbb{O}_{t}=\underbrace{\mathbb{O}_{2}\otimes\mathbb{O}_{2}\otimes\cdots\otimes\mathbb{O}_{2}}_{p}\otimes{\bf 1}_{q}\in{\mathcal{M}}_{\alpha\times 2^{p}}. (168)

    where

    𝕆2=[1111].\mathbb{O}_{2}=\begin{bmatrix}1&1\\ 1&-1\end{bmatrix}.

    Moreover, 𝕆t\mathbb{O}_{t} is unique up to column (or row) sign changes and column (or row) permutations.

  • (ii)
    𝕌t=𝕆t.\displaystyle\mathbb{U}_{t}=\mathbb{O}_{t}. (169)

We conclude that

Corollary VI.18

Consider even tt. When t=2pt=2^{p}, the largest ratio s/ts/t can be obtained as

maxt(s/t)=1.\displaystyle\max_{t}(s/t)=1. (170)

Next, we consider odd case, i.e., set t=2k+1t=2k+1. Then 𝕆t=\mathbb{O}_{t}=\emptyset. We construct 𝕌t\mathbb{U}_{t}.

Motivated by the case of even tt, we may choose t=2p1t=2^{p}-1.

Proposition VI.19

Consider odd tt. When t=2p1t=2^{p}-1, the ratio s/ts/t can be obtained as

maxt(s/t)=2p2p1,\displaystyle\max_{t}(s/t)=\frac{2^{p}}{2^{p}-1}, (171)

which is slightly larger than 11.

Proof. Consider 𝕆2p\mathbb{O}_{2^{p}}. Deleting the first row, the remainder becomes a 𝕌2p1\mathbb{U}_{2^{p}-1}.

\Box.

Proposition VI.20
Proposition VI.21

The U𝕌2p1U\in\mathbb{U}_{2^{p}-1} generated from 𝕆2p\mathbb{O}_{2^{p}} is a maximum almost orthogonal matrix.

Proof. Assume

U𝕌2p1U\in\mathbb{U}_{2^{p}-1}

is generated from 𝕆2p\mathbb{O}_{2^{p}} by delete its first row. Then

𝕆2p=[𝟏2pTU].\mathbb{O}_{2^{p}}=\begin{bmatrix}{\bf 1}^{T}_{2^{p}}\\ U\end{bmatrix}.

Hence,

Coli(U),Colj(U)=1,ij.\langle\operatorname{Col}_{i}(U),\operatorname{Col}_{j}(U)\rangle=-1,\quad i\neq j.

Note that

Col1(U)=𝟏2p1,\operatorname{Col}_{1}(U)={\bf 1}_{2^{p}-1},

then

N+(Colj(U))=2p11,N(Colj(U))=2p1;j[2,2p].\displaystyle\begin{array}[]{l}N_{+}(\operatorname{Col}_{j}(U))=2^{p-1}-1,\\ N_{-}(\operatorname{Col}_{j}(U))=2^{p-1};\quad j\in[2,2^{p}].\\ \end{array} (174)

We use contradiction to prove UU is the maximum almost orthogonal matrix. To this end, assume x2p1x\in{\mathbb{R}}^{2^{p}-1} is a sign vector, satisfying x,u=±1\langle x,u\rangle=\pm 1, uCol(U)u\in\operatorname{Col}(U). i.e. [U,x]𝕌2p1[U,x]\in\mathbb{U}_{2^{p}-1}.

According to Lemma LABEL:lcs.3.3.1, we can assume

x,𝟏2p1=1.\langle x,{\bf 1}_{2^{p}-1}\rangle=-1.

Then xx satisfies (174). First,

x,u=1,uCol(U)\langle x,u\rangle=-1,\quad\forall u\in\operatorname{Col}(U)

is impossible. Otherwise,

[𝟏2p1Ux]𝕆2p.\begin{bmatrix}{\bf 1}_{2^{p}}&1\\ U&x\end{bmatrix}\in\mathbb{O}_{2^{p}}.

This contradicts the structure of maximum 𝕆2p\mathbb{O}_{2^{p}}. Hence, there is at least one y=Colj(U)y=\operatorname{Col}_{j}(U), 2j2p2\leq j\leq 2^{p}, such that

x,y=1.\displaystyle\langle x,y\rangle=1. (175)

Since yy satisfies (174), we have

|{i|yi=1,xi=1}|:=s,|{i|yi=1,xi=1}|=2p11s,|{i|yi=1,xi=1}|:=t,|{i|yi=1,xi=1}|=2p1t.\begin{array}[]{l}|\{i\;|\;y_{i}=1,x_{i}=1\}|:=s,\\ |\{i\;|\;y_{i}=1,x_{i}=-1\}|=2^{p-1}-1-s,\\ |\{i\;|\;y_{i}=-1,x_{i}=1\}|:=t,\\ |\{i\;|\;y_{i}=-1,x_{i}=-1\}|=2^{p-1}-t.\\ \end{array}

xx also satisfies (174), hence

s+t=2p11.\displaystyle s+t=2^{p-1}-1. (176)

According to (175), we have

s+2p1t=2p1.s+2^{p-1}-t=2^{p-1}.

That is

s=t.\displaystyle s=t. (177)

Using (176) and (177) yields

s=t=2p112,s=t=\frac{2^{p-1}-1}{2},

this is a contradiction because ss and tt are integers.

\Box

Example VI.22
  • (i)

    Consider t=3t=3. Note that

    𝕆4=[1111111111111111].\mathbb{O}_{4}=\begin{bmatrix}1&1&1&1\\ 1&-1&1&-1\\ 1&1&-1&-1\\ 1&-1&-1&1\\ \end{bmatrix}.

    Then we have

    𝕌3=[111111111111][111111111111].\mathbb{U}_{3}=\begin{bmatrix}1&-1&1&-1\\ 1&1&-1&-1\\ 1&-1&-1&1\\ \end{bmatrix}\sim\begin{bmatrix}1&1&1&1\\ 1&-1&-1&1\\ 1&1&-1&-1\\ \end{bmatrix}.
  • (ii)

    Consider t=7t=7. Using 𝕆8\mathbb{O}_{8}, we have

    𝕌7=[11111111111111111111111111111111111111111111111111111111][11111111111111111111111111111111111111111111111111111111].\begin{array}[]{l}\mathbb{U}_{7}=\begin{bmatrix}1&-1&1&-1&1&-1&1&-1\\ 1&1&-1&-1&1&1&-1&-1\\ 1&-1&-1&1&1&-1&-1&1\\ 1&1&1&1&-1&-1&-1&-1\\ 1&-1&1&-1&-1&1&-1&1\\ 1&1&-1&-1&-1&-1&1&1\\ 1&-1&-1&1&-1&1&1&-1\\ \end{bmatrix}\\ \sim\begin{bmatrix}1&1&1&1&1&1&1&1\\ 1&-1&-1&1&1&-1&-1&1\\ 1&1&-1&-1&1&1&-1&-1\\ 1&-1&1&-1&-1&1&-1&1\\ 1&1&1&1&-1&-1&-1&-1\\ 1&-1&-1&1&-1&1&1&-1\\ 1&1&-1&-1&-1&-1&1&1\\ \end{bmatrix}.\end{array}
Remark VI.23
  • (i)

    If t=2p+1t=2^{p}+1, we may one arbitrary row to 𝕆2p\mathbb{O}_{2^{p}} to get a 𝕌2p+12p+1×2p\mathbb{U}_{2^{p}+1}\in{\mathcal{M}}_{2^{p}+1\times 2^{p}}. But this one might not be the one with maximum ratio s/ts/t. For instance, we have

    𝕌5=[1111111111111111111111111]5×5.\mathbb{U}_{5}=\begin{bmatrix}1&1&1&1&1\\ 1&1&1&-1&-1\\ 1&1&-1&1&-1\\ 1&-1&-1&-1&1\\ 1&-1&1&1&-1\\ \end{bmatrix}\in{\mathcal{M}}_{5\times 5}.
  • (ii)

    Unlike even case, so far we don’t know how to construct 𝕆t\mathbb{O}_{t} for general odd tt.

  • (iii)

    Our conjecture is (171) is the best ratio for odd tt.

  • (iv)

    Using (160) and (170), we have that when t=α1t=\alpha-1 is even the best compression rate is 22. If the above conjecture is correct, when tt is odd, the best compression rate is slightly higher than 22.

VII Conclusion

The purpose of this paper is to reveal some mathematical foundations for the STP-CS. First, the signal space is presented. The STP based equivalence leads to a quotient space, which is the signal space. Second, a coordinate-free STP-CS is presented. It is also revealed that STP-CS has an advantage in 0\ell_{0}. Third, a systematic construction of BIBD-based sensing matrix is obtained.

References

  • [1] A. Amini, F. Marvasti, Deterministic construction of binary, bipolar, and ternary compressed sensing matrices. IEEE Trans. Inform. Theory, Vol. 57, No. 4, 2360– 2370, 2011.
  • [2] A. Amini, V. Montazerhodjat, F. Marvasti, Matrices with small coherence using p-ary block codes, IEEE Sign. Proc., 60, 1 , 172-181, 2012.
  • [3] T.T. Cai, A. Zhang, Sparse representation of a polytope and recovery of sparse signals and low-rank matrices. IEEE Trans. Inform. Theory, Vol. 60, No. 1, 122–132, 2014.
  • [4] E.J. Candes, J. Romberg, T. Tao, Robust uncertainty principles: signal reconstruction from highly incomplete frequencyinformation, IEEE Trans. Inf. Theory, Vol. 52, No. 2, 489-509, 2006.
  • [5] E. Candes, J.K. Romberg, T. Tao, Stable signal recovery from incomplete and inaccurate measurements, Commun. Pure Appl. Math., Vol. 59, No. 8, 1207-1223, 2006.
  • [6] E. Candes, J.K. Romberg, Sparsity and incoherence in compressive sampling, Inverse Prob., Vol. 23, No. 3, 969-985, 2007.
  • [7] X. Chai, J. Fu, Z. Gan, et al., Exploiting semi-tensor product compressed sensing and hybrid cloud for secure medical image trasmissian, IEEE Int. Thing. J., Vol. 10, No. 8, 7380-7392, 2023.
  • [8] D. Cheng, Matrix and Polynomial Approach to Dynamic Control Systems,
  • [9] D. Cheng, H. Qi, Z. Li, Analysis and Control of Boolean Networks - A Semi-tensor Product Approach, Springer, London, 2011.
  • [10] D. Cheng, H. Qi, Y. Zhao, An Introduction to Semi-tensor Product of Matrices and Its Applications, World Scientific, Singapore, 2012.
  • [11] D. Cheng, On equivalence of Matrices, Asian J. Math., Vol. 23, No. 2, 257-348, 2019.
  • [12] D. Cheng, From Dimension-Free Matrix Theory to Cross-Dimensional Dynamic Systems, Elsevier, London, 2019.
  • [13] D. Cheng, Z. Ji, From dimension-free manofolds to dimension-varying control systems, Commun. Inform. Sys., Vol. 23, No. 1, 85-150, 2023.
  • [14] D. Cheng, Z. Ji, Finite and Dimension-free dynamic systems, Science Press, Beijing, 2023 (in Chinese).
  • [15] N. Cleju, Optimized projections for compressed sensing via rank-constrained nearest correlation matrix, Appl. Comput. Harmon. Anal., Vol. 36, No. 3, 495–507, 2014.
  • [16] T.T. Do, L. Gan, N.H. Nguyen, et al., Fast and efficient compressive sensing using structurally random matrices, IEEE Trans. Signal Proc., Vol. 60, No. 1, 139–154, 2012.
  • [17] D.L. Donoho, M. Elad, Optimally sparse representation in general (nonorthogonal) dictionaries via l1l_{1} minimization, Proc. Natl. Acad. Sci., Vol. 100, No. 5, 2197-2202, 2003.
  • [18] D.L. Donoho, Compressed sensing, IEEE Trans. Inf. Theory, Vol. 52, No. 4, 1289-1306, 2006.
  • [19] M.F. Duarte, Y.C. Eldar, Structured compressed sensing: From theory to applications, IEEE Trans. Sign. Proc., Vol. 59, No. 9, 4053-4085, 2011.
  • [20] M.F. Duarte, R.G. Baraniuk, Kronecker compressive sensing. IEEE Trans. Image Proc., Vol. 21, No. 2, 494–504, 2012.
  • [21] R.W. Hamming, Digital Filters, Prentice-Hall, New Jersey, 1977.
  • [22] J. Hou, X. Liu, Semi-tensor product-based one-bit compressed sensing. Eurasip J. Adv. Sign. Proc., 2023(1): 112.
  • [23] A. Joshi, A. Yerudkar, C. D. Vecchio, L. Glielmo, Storage Constrained Smart Meter Sensing using Semi-Tensor Product, Proc. 2019 IEEE Int. Conf. Sys. Man Cyb., Bari, Italy, 51-56, 2019.
  • [24] J.L. Kelley, I. Namioka, Linear Topological Spaces, Springer-Verlag, New York, 1963.
  • [25] J. Liang, H. Peng, L. Li, F. Tong, Construction of structured random measurement matrices in semi-tensor product compressed sensing based on combinatorial designs. Sensors, Vol. 22, No. 21, 8260, 2022.
  • [26] K. Lee, Y. Wu, Y. Bresler, Near optimal compressed sensing of sparse rankone matrices via sparse power factorization. Comput. Scie., Vol. 92, No. 4, 621– 624, 2013.
  • [27] H. Peng, Y. Mi, L. Li, H.E. Stanley, and Y. Yang, P-tensor product in compressed sensing, IEEE Intern. Thing. J., Vol. 6, No. 3, 3492-3510, 2-19.
  • [28] A.E. Taylor, D.C. Lay, Introduction to Functional Analysis, 2nd Ed., John Wiley & Sons, Inc., New York, 1980.
  • [29] F. Tong, L. Li, J. Pemg, Y. Yang, Flexible construction of compressed sensing matrices with low storage space and low coherence, Signal Processing, 182 (2021) 107951.
  • [30] J. Wang, S. Ye, R. Yue, C. Chen, Lower storage space for compressive sensing: Semi-tensor product approach, EURASIP J. Image Video Proc., Vol. 2017, No. 51, DOI. 10.1186/s-13640-017-0199-9, 2017.
  • [31] W. Wen, Y. Hong, Y. Fang, Me. Li, Mi. Li, A visually secure image encryption scheme based on semi-tensor product compressed sensing, Signal Processing, Vol. 173, 107580, 2020.
  • [32] D. Xie, H. Peng, L. Li, Y. Yang, Semi-tensor compressed sensing, Dig. Sign. Proc., Vol. 58, 85-92, 2016.
  • [33] Y. Xu, W. Yin, and S. Osher. Learning circulant sensing kernels, Inv. Prob. Imag., Vol/ 8, No. 3, 901-923, 2014.
  • [34] S. Xue, L. Zhang, Z. Xie, W. Yan, K. Zhang, Embedding cross-dimensional vector space into l2l^{2}, J. Syst. Sci. Compl., Vol. 36, 2309-2324, 2023.
  • [35] G. Ye, M. Liu, W. Yap, B. Goi, Reversible image hiding algorithm based on compressive sensing and deep learning. Nonlinear Dynamics, Vol. 111, No. 14, 13535-13560.
  • [36] H. Yuan, H. Song, X. Sun, et al., Compressive sensing measurement matrix construction based on improved size compatible array LDPC code. IET Image Proc., Vol. 9, No. 11, 993–1001, 2015.
  • [37] , K. Zhang, K.H. Johansson, Long-term behavior of cross-dimensional linear dynamical systems, Proc. CCC 2018, 158-163, 2018.
  • [38] Q. Zhang, B. Wang, J. Feng, Solution and stability of continuous-time cross-dimensional linear systems, Fron. Inform. Tech., Vol. 22, No. 2, 210-221, 2021.