Izvestiya: Mathematics
RUS  ENG    JOURNALS   PEOPLE   ORGANISATIONS   CONFERENCES   SEMINARS   VIDEO LIBRARY   PACKAGE AMSBIB  
General information
Latest issue
Forthcoming papers
Archive
Impact factor
Guidelines for authors
Submit a manuscript

Search papers
Search references

RSS
Latest issue
Current issues
Archive issues
What is RSS



Izv. RAN. Ser. Mat.:
Year:
Volume:
Issue:
Page:
Find






Personal entry:
Login:
Password:
Save password
Enter
Forgotten password?
Register


Izvestiya: Mathematics, 2022, Volume 86, Issue 4, Pages 770–781
DOI: https://doi.org/10.4213/im9137e
(Mi im9137)
 

This article is cited in 1 scientific paper (total in 1 paper)

Maltsev equal-norm tight frames

S. Ya. Novikov, V. V. Sevost'yanova

Samara National Research University
References:
Abstract: A frame in $\mathbb{R}^d$ is a set of $n\geqslant d$ vectors whose linear span coincides with $\mathbb{R}^d$. A frame is said to be equal-norm if the norms of all its vectors are equal. Tight frames enable one to represent vectors in $\mathbb{R}^d$ in the form closest to the representation in an orthonormal basis. Every equal-norm tight frame is a useful tool for constructing efficient computational algorithms. The construction of such frames in $\mathbb{C}^d$ uses the matrix of the discrete Fourier transform, and the first constructions of equal-norm tight frames in $\mathbb{R}^d$ appeared only at the beginning of the 21st century. The present paper shows that Maltsev's note of 1947 was decades ahead of its time and turned out to be missed by the experts in frame theory, and Maltsev should be credited for the world's first design of an equal-norm tight frame in $\mathbb{R}^d$. Our main purpose is to show the historical significance of Maltsev's discovery. We consider his paper from the point of view of the modern theory of frames in finite-dimensional spaces. Using the Naimark projectors and other operator methods, we study important frame-theoretic properties of the Maltsev construction, such as the equality of moduli of pairwise scalar products (equiangularity) and the presence of full spark, that is, the linear independence of any subset of $d$ vectors in the frame.
Keywords: matrix, equal-norm tight frame, analysis operator, synthesis operator, orthogonal rows of a matrix, equiangular frame, full spark.
Funding agency Grant number
Ministry of Science and Higher Education of the Russian Federation 075-02-2022-878
This paper has been written under the development program of Volga Region Mathematical Center (contract no. 075-02-2022-878).
Received: 29.12.2020
Revised: 25.05.2021
Russian version:
Izvestiya Rossiiskoi Akademii Nauk. Seriya Matematicheskaya, 2022, Volume 86, Issue 4, Pages 162–174
DOI: https://doi.org/10.4213/im9137
Bibliographic databases:
Document Type: Article
UDC: 517.982.254
MSC: 42C15, 46C05, 15B99
Language: English
Original paper language: Russian

§ 1. Introduction

Let $n$ and $d$ be integers, $n\geqslant d$. A frame in $\mathbb{R}^d$ is a complete family of $n$ vectors in $d$-dimensional Euclidean space, generalizing the concept of a basis by omitting the requirement of linear independence of all vectors. More precisely, a family of vectors $\{\varphi_j\}_{j=1}^n$ is called a frame of $\mathbb{R}^d$ if there are numbers $a,b$, $0<a\leqslant b< \infty$, such that for all $ \mathbf{x}\in \mathbb{R}^d$ we have

$$ \begin{equation*} a\|\mathbf{x}\|^2\leqslant \sum_{j=1}^n|\langle \mathbf{x},\varphi_j\rangle |^2\leqslant b\|\mathbf{x}\|^2. \end{equation*} \notag $$

The numbers $a$ and $b$ are referred to as frame bounds. They are not uniquely defined. Usually, they are understood as the optimal bounds, that is, the minimum $b$ and the maximum $a$.

In finite-dimensional spaces, the concept of a frame is equivalent to that of a complete system, that is, $\operatorname{span}\{\varphi_j\}_{j=1}^n=\mathbb{R}^d$. This and other well-known facts about frames can be found in [1].

The synthesis operator for a finite system $\{\varphi_j\}_{j=1}^n$ of vectors in $\mathbb{R}^d$ is defined as $\mathbf{\Phi} \colon \mathbb{R}^n \to \mathbb{R}^d$, $\mathbf{\Phi} \mathbf{x} :=\sum_{j=1}^n \mathbf{x}(j)\varphi_j$, where $\mathbf{x}(j)$ denotes the $j$th coordinate of $\mathbf{x}\in \mathbb{R}^n$. The conjugate of the synthesis operator is the analysis operator $\mathbf{\Phi}^* \colon \mathbb{R}^d \to \mathbb{R}^n$, defined by $(\mathbf{\Phi}^*\mathbf{y})(j) = \langle \varphi_j,\mathbf{y}\rangle$ for $j = 1,\dots,n$.

The composite $\mathbf{\Phi}^*\mathbf{\Phi}\colon\mathbb{R}^n\to \mathbb{R}^n$ of the analysis and synthesis operators determines the Gram matrix, that is, the $(n\times n)$-matrix whose $(j,j')$th entry is equal to $(\mathbf{\Phi}^*\mathbf{\Phi})(j,j') = \langle\varphi_j,\varphi_{j'}\rangle$. The composite $\mathbf{\Phi}\mathbf{\Phi}^* \colon \mathbb{R}^d\to \mathbb{R}^d$ in the inverse order is the frame operator $\mathbf{\Phi}\mathbf{\Phi}^*\mathbf{y} = \sum_{j=1}^n\langle\varphi_j,\mathbf{y}\rangle\varphi_j$. It is well known that sequences $\{\varphi_j\}_{j=1}^n$ and $\{\widehat{\varphi}_j\}_{j=1}^n$ in $\mathbb{R}^d$ have equal Gram matrices if and only if there is a unitary operator $\mathbf{U}$ such that $\mathbf{U}\varphi_j=\widehat{\varphi}_j$ for all $j = 1,\dots,n$.

For a single vector $\{\varphi_j\}$, the synthesis and analysis operators take the form

$$ \begin{equation*} \varphi_j\colon \mathbb{R}\to\mathbb{R}^d,\quad \varphi_j x = x\varphi_j;\qquad \varphi_j^*\colon \mathbb{R}^d\to\mathbb{R},\quad \varphi_j^*\mathbf{y} = \langle\varphi_j, \mathbf{y} \rangle. \end{equation*} \notag $$
The frame operator may be written in this notation as $\mathbf{\Phi}\mathbf{\Phi}^*=\sum_{j=1}^n\varphi_j\varphi_j^*$.

Hence the matrix representing the synthesis operator $\mathbf{\Phi}$ is the $(d\times n)$-matrix whose columns are the coordinates of the frame vectors $\{\varphi_j\}_{j=1}^n$. In particular, if the frame bounds are equal to each other, then the frame operator is represented by the diagonal matrix $\mathbf{\Phi}\mathbf{\Phi^*}{=}\, a\mathbf{I}_{\mathbb{R}^d}$ with $a>0$, and this formula gives a simple frame representation of vectors or signals:

$$ \begin{equation*} \mathbf{x}=\frac 1a\mathbf{\Phi}\mathbf{\Phi^*}\mathbf{x} =\frac1{a}\sum_{j=1}^n\langle\varphi_j,\mathbf{x}\rangle\varphi_j, \qquad \mathbf{x}\in \mathbb{R}^d. \end{equation*} \notag $$
These frames are said to be tight or $a$-tight, and $1$-tight frames are called Parseval frames.

Tight frames all of whose vectors have equal norms are of particular interest. These frames are called equal-norm tight frames. In general, a frame $\{\varphi_j\}_{j=1}^n$ is said to be equal-norm if there is a $c>0$ such that $\|\varphi_{j}\|^2=c$, $j=1,\dots,n$.

Given an equal-norm $a$-tight frame, we have a connection between $d$, $c$ and $a$:

$$ \begin{equation*} da=\operatorname{trace}(a\,\mathbf{I}_{\mathbb{R}^d}) =\operatorname{trace}(\mathbf{\Phi}\mathbf{\Phi^*}) =\operatorname{trace}(\mathbf{\Phi^*}\mathbf{\Phi})= \sum_{j=1}^n\|\varphi_j\|^2=nc. \end{equation*} \notag $$
In particular, $d=nc$ for Parseval frames. Hence equal-norm Parseval frames have norms $\leqslant 1$.

§ 2. Naimark’s theorem and its matrix consequences

The following result of Naimark [2] turned out to be fundamental for numerical constructions of frames. It is repeatedly cited in the papers on frames and their applications.

Theorem 1. If $\mathbf{\Phi}= \{\varphi_j\}_{j=1}^n$ is a Parseval frame in $\mathbb{R}^d$, then there is an orthonormal basis $\{\mathbf{b}_j\}_{j=1}^n$ in $\mathbb{R}^n$ such that $\varphi_j=\mathbf{P}_{\mathbb{R}^d} \mathbf{b}_j$ for all $j$, where $\mathbf{P}_{\mathbb{R}^d}$ is the orthogonal projection of $\mathbb{R}^n$ to $\mathbb{R}^d$.

The converse also holds and is easily provable.

Theorem 2. If $\{\mathbf{b}_j\}_{j=1}^n$ is an orthonormal basis in $\mathbb{R}^n$, then $\mathbf{\Phi}= \{\mathbf{P}_{\mathbb{R}^d} \mathbf{b}_j\}_{j=1}^n$ is a Parseval frame in $\mathbb{R}^d$, where $\mathbf{P}_{\mathbb{R}^d}$ is the orthogonal projection of $\mathbb{R}^n$ to $\mathbb{R}^d$.

Proof. If $\varphi_j=\mathbf{P}_{\mathbb{R}^d} \mathbf{b}_j$, $j=1,\dots, n,$ and $\mathbf{x}\in \mathbb{R}^d$, then
$$ \begin{equation*} \|\mathbf{x}\|^2=\|\mathbf{P}_{\mathbb{R}^d} \mathbf{x}\|^2=\sum_{j=1}^n|\langle \mathbf{P}_{\mathbb{R}^d} \mathbf{x},\mathbf{b}_j\rangle |^2 =\sum_{j=1}^n|\langle \mathbf{x},\mathbf{P}_{\mathbb{R}^d}\mathbf{b}_j\rangle|^2 =\sum_{j=1}^n|\langle \mathbf{x},\varphi_j\rangle |^2. \end{equation*} \notag $$
Hence $\mathbf{\Phi}$ is a Parseval frame. $\Box$

Theorems 1 and 2 were extended in [3] to frames of general form, which turned out to be projections of Riesz bases. Projections of general orthogonal systems (not necessarily complete) were considered in [4].

Let

be the matrix of the synthesis operator, that is, the $(d\times n)$-matrix whose columns are the coordinates of the frame vectors $\{\varphi_j\}_{j=1}^n$ in the canonical basis.

Theorem 3. Let $\{\varphi_j\}_{j=1}^n$ be a frame in $\mathbb{R}^d$. Then the coordinates of the vectors $\varphi_j$ may be regarded as the first $d$ coordinates of some vectors in $\mathbb{R}^n$ which form a basis in $\mathbb{R}^n$. If $\{\varphi_j\}_{j=1}^n$ is a tight frame, then the coordinates of the vectors $\varphi_j$ are the first $d$ coordinates of some vectors which form an orthogonal basis in $\mathbb{R}^n$.

Proof. Let $\{\varphi_j\}_{j=1}^n$ be an arbitrary frame in $\mathbb{R}^d$. The analysis operator $\Phi^*\colon \mathbb{R}^d\to \mathbb{R}^n$ has matrix

If $\Phi^*\mathbf{x}=\mathbf{0}$, then

$$ \begin{equation*} 0=\|\Phi^*\mathbf{x}\|^2=\sum_{j=1}^n|\langle \varphi_j,\mathbf{x}\rangle |^2. \end{equation*} \notag $$
Since $\operatorname{span}\{\varphi_j\}_{j=1}^n=\mathbb{R}^d$, we have $\mathbf{x}=\mathbf{0}$. Hence, $\Phi^*$ is an injective mapping. We can therefore extend $\Phi^*$ to a bijection $\widetilde{\Phi}^*\colon \mathbb{R}^n\to \mathbb{R}^n$, for example, in the following way. Let $\{\mathbf{e}_j\}_{j=1}^n$ be the canonical basis in $\mathbb{R}^n$. Then we choose a basis $\{\mathbf{b}_j\}_{j=d+1}^n$ for the orthogonal complement of the range $\mathcal{R}_{\Phi^*}$ in $\mathbb{R}^n$ and set by definition $\widetilde{\Phi}^*\mathbf{e}_j:=\mathbf{b}_j$, $j= d+1,\dots, n$. The matrix $\widetilde{\Phi}^*$ is an $(n\times n)$-matrix whose first $d$ columns coincide with the matrix $\Phi^*$:

Since $\widetilde{\Phi}^*$ is surjective by construction, the columns of this matrix are complete in $\mathbb{R}^n$. Hence the rows of the matrix $\widetilde{\Phi}^*$ are complete in $\mathbb{R}^n$, that is, they are linearly independent and constitute a basis for $\mathbb{R}^n$.

If $\{\varphi_j\}_{j=1}^n$ is an $a$-tight frame for $\mathbb{R}^d$, then, as noted above, $\Phi\Phi^*{=}\,a I$. Hence the rows of the matrix of $\Phi$ are orthogonal vectors in $\mathbb{R}^n$. By adding $n\,{-}\,d$ rows to this matrix in such a way that the extended matrix has orthogonal rows, we obtain an $(n\,{\times}\,n)$-matrix with orthogonal rows. Hence the columns of the extended matrix are also orthogonal. $\Box$

§ 3. Maltsev’s matrices. Equiangular frames. Full-spark frames

Maltsev published the paper [5] in 1947. He answered a question in [6] by constructing a matrix with orthogonal rows of unit length and columns of equal length. We call such matrices Maltsev’s matrices. In fact it was the first construction of an equal-norm Parseval frame.

While Naimark’s paper mentioned in § 2 is frequently cited in the literature about frames, Maltsev’s note went unnoticed. Modern constructions of equal-norm tight frames for $\mathbb{R}^d$ appeared only at the beginning of the 21st century [7].

Consider the construction of Maltsev’s matrices in more detail. We first justify the possibility of considering only the case when $d<n/2$.

Lemma 1. Suppose that $\{\varphi_j\}_{j=1}^{n}$ is an $a$-tight frame for $\mathbb{R}^d$ and $d<n$. Then the $(n\times n)$-matrix $(1/a)\mathbf{\Phi}^* \mathbf{\Phi}$ is the matrix of orthogonal projection to a $d$-dimensional subspace of $\mathbb{R}^n$. The Gram matrix $\mathbf{\Phi}^* \mathbf{\Phi}$ has eigenvalues $a$ of multiplicity $d$ and $0$ of multiplicity $n-d$.

Proof. Since $\{\varphi_j\}_{j=1}^{n}$ is an $a$-tight frame, we have
$$ \begin{equation*} \varphi_l=\frac{1}{a}\sum_{j=1}^n\langle\varphi_j, \varphi_l\rangle\varphi_j, \qquad l=1,\dots, n. \end{equation*} \notag $$
Hence, for any $k$, $l$ we obtain
$$ \begin{equation*} \langle\varphi_k,\varphi_l\rangle=\biggl\langle\varphi_k,\frac{1}{a}\sum_{j=1}^n \langle\varphi_j,\varphi_l\rangle\varphi_j\biggr\rangle= \frac{1}{a}\sum_{j=1}^n\langle\varphi_k,\varphi_j\rangle \langle\varphi_j, \varphi_l\rangle \end{equation*} \notag $$
or
$$ \begin{equation*} \mathbf{\Phi}^*\mathbf{\Phi}=\frac{1}{a}(\mathbf{\Phi}^*\mathbf{\Phi})^2. \end{equation*} \notag $$

If $\mathbf{P}:=(1/a)\mathbf{\Phi}^* \mathbf{\Phi}$, then $\mathbf{P}=\mathbf{P}^2$. Since $\mathbf{P}$ is self-adjoint, we see that $\mathbf{P}$ is the orthogonal projection to a $d$-dimensional subspace. The matrix of this orthogonal projection has $d$ eigenvalues equal to $1$, and $n-d$ zero eigenvalues, $\operatorname{trace} \mathbf{P} =\operatorname{rank} \mathbf{P} = d$. Hence $\mathbf{\Phi}^* \mathbf{\Phi}= a\mathbf{P}$ has $d$ eigenvalues equal to $a$, and $n-d$ zero eigenvalues,

$$ \begin{equation*} \operatorname{trace}(\mathbf{\Phi}^* \mathbf{\Phi}) = ad = \sum_{j=1}^n\|\varphi_j\|^2,\qquad \operatorname{rank} (\mathbf{\Phi}^* \mathbf{\Phi}) = d.\qquad\Box \end{equation*} \notag $$

Theorem 4. Every $a$-tight frame $\{\varphi_j\}_{j=1}^{n}$ for $\mathbb{R}^d$ has a Naimark complement frame $\{\psi_j\}_{j=1}^{n}$ with the following properties:

(i) $\{\psi_j\}_{j=1}^{n}$ is an $a$-tight frame for $\mathbb{R}^{(n-d)}$;

(ii) the Gram matrix $\mathbf{\Psi}^* \mathbf{\Psi}=a \mathbf{I}_{\mathbb{R}^n} - \mathbf{\Phi}^* \mathbf{\Phi}$;

(iii) $\mathbf{\Phi}\mathbf{\Psi}^*=\mathbf{0}$.

Proof. (i) A direct verification shows that the matrix $\mathbf{Q}:=\mathbf{I}_{\mathbb{R}^n}-(1/a)\mathbf{\Phi}^*\mathbf{\Phi}$ satisfies the equality $\mathbf{Q}=\mathbf{Q}^*=\mathbf{Q}^2,$ and its eigenvalues are $1$ and $0$ with multiplicities $n-d$ and $d$ respectively. Hence the columns $\widetilde{\psi}_j:=\mathbf{Q}\mathbf{e}_j$, $j=1,\dots, n$, of the matrix $\mathbf{Q}$, where $\{\mathbf{e}_j\}_{j=1}^{n}$ is an orthonormal basis for $\mathbb{R}^n$, determine $\operatorname{span}\bigl(\{\widetilde{\psi}_j\}_{j=1}^{n}\bigr)=\mathbb{R}^{(n-d)}$.

If $\mathbf{f}\in \operatorname{span}\bigl(\{\widetilde{\psi}_j\}_{j=1}^{n}\bigr)$, then $\mathbf{Q}\mathbf{f}=\mathbf{f}$,

$$ \begin{equation*} \mathbf{f}=\mathbf{Q}\biggl(\sum_{j=1}^n\langle\mathbf{Q} \mathbf{f}, \mathbf{e}_j\rangle \mathbf{e}_j\biggr)=\sum_{j=1}^n\langle \mathbf{f}, \mathbf{Q}\mathbf{e}_j\rangle \mathbf{Q} \mathbf{e}_j=\sum_{j=1}^n\langle \mathbf{f}, \widetilde{\psi}_j\rangle \widetilde{\psi}_j, \end{equation*} \notag $$
that is, the system of vectors $\{\widetilde{\psi}_j\}_{j=1}^{n}$ is a Parseval frame for $\operatorname{span}\bigl(\{\widetilde{\psi}_j\}_{j=1}^{n}\bigr)$.

(ii) The vectors $\psi_j= \sqrt{a} \widetilde{\psi}_j$, $j=1,\dots, n$, form an $a$-tight frame for the space $\operatorname{span}(\{\psi_j\}_{j=1}^{n}) =\operatorname{span}\bigl(\{\widetilde{\psi}_j\}_{j=1}^{n}\bigr)$ and $\mathbf{\Psi}^*\mathbf{\Psi}=a\mathbf{I}_{\mathbb{R}^n}-\mathbf{\Phi}^*\mathbf{\Phi}$.

(iii) By the lemma, we have

$$ \begin{equation*} (\mathbf{\Psi}\mathbf{\Phi}^*)^*(\mathbf{\Psi}\mathbf{\Phi}^*) =\mathbf{\Phi}\mathbf{\Psi}^*\mathbf{\Psi}\mathbf{\Phi}^* =\mathbf{\Phi}(a\mathbf{I}-\mathbf{\Phi}^*\mathbf{\Phi})\mathbf{\Phi}^* =a\mathbf{\Phi}\mathbf{\Phi}^*-(\mathbf{\Phi}\mathbf{\Phi}^*)^2=\mathbf{0}. \end{equation*} \notag $$
The Frobenius norm of the matrix
$$ \begin{equation*} \|\mathbf{\Psi}\mathbf{\Phi}^*\|^2_{\mathrm{Fro}}= \operatorname{trace}\bigl((\mathbf{\Psi}\mathbf{\Phi}^*)^*(\mathbf{\Psi}\mathbf{\Phi}^*)\bigr) =\operatorname{trace}\mathbf{0}=0. \end{equation*} \notag $$
Hence $\mathbf{\Psi}\mathbf{\Phi}^*=\mathbf{0}$, and its adjoint is $\mathbf{\Phi}\mathbf{\Psi}^*=\mathbf{0}$. $\Box$

Theorem 4 asserts that we can add $n-d$ rows to the $(d\times n)$-matrix $\mathbf{\Phi}$ of the synthesis operator for an equal-norm tight frame of vectors in $\mathbb{R}^d$ in such a way that the extended $(n\times n)$-matrix is orthogonal. Moreover, the added rows form the $((n-d)\times n)$-matrix of the synthesis operator $\mathbf{\Psi}$ for an equal-norm tight frame for the space $\mathbb{R}^{(n-d)}$.

When $d=n/2$, we can take for $\mathbf{\Phi}$ two orthogonal $(d\times d)$-matrices multiplied by $1/\sqrt{2}$ and placed side by side.

Consider Maltsev’s design for $d<n/2$.

For completeness, we recall a construction (following [5]) of Sylvester–Walsh matrices, which form a subset of Hadamard matrices.

Remark 1. For every positive integer $m$ there is a $(2^m\times 2^m)$-matrix $\mathbf{H}_{2^m}$ with entries $\pm 1$ and with mutually orthogonal rows and columns.

Indeed, for $m=1$ we have a matrix

$$ \begin{equation*} \mathbf{H}_2= \begin{pmatrix} 1 & 1 \\ 1 & -1 \end{pmatrix}. \end{equation*} \notag $$
We define higher-order matrices by putting
$$ \begin{equation*} \mathbf{H}_{2^m}=\left( \begin{array}{c|c} \rule{0pt}{14pt}\mathbf{H}_{2^{m-1}}&\mathbf{H}_{2^{m-1}}\\ \hline \rule{0pt}{14pt}\mathbf{H}_{2^{m-1}}&-\mathbf{H}_{2^{m-1}}\\ \end{array}\right)=\mathbf{H}_2\otimes \mathbf{H}_{2^{m-1}}, \end{equation*} \notag $$
$2\leqslant m\in \mathbb{N}$, where $\otimes$ denotes the Kronecker product.

The Sylvester–Walsh matrices are closely connected with discrete Walsh functions, which have many useful applications [8].

By the remark, for any positive integers $m$ and $k\leqslant 2^m$ there are $k$ mutually orthogonal rows of length $2^m$ with entries $\pm 1$.

The number $n$ of columns of the required matrix $\mathbf{\Phi}$ is represented by

$$ \begin{equation*} n= 2^{m_1}+\dots + 2^{m_t},\qquad 0\leqslant m_1<\dots< m_t. \end{equation*} \notag $$
There is a number $s$ such that $1\leqslant s\leqslant t-1$ and $2^{m_s}< d\leqslant 2^{m_{s+1}}$.

We construct the matrix $\mathbf{\Phi}$ of the synthesis operator for an equal-norm tight frame in two steps. At the first step we obtain a matrix with orthogonal rows and with entries $\pm 1$ and $0$. At the second step we normalize it without losing the orthogonality. The matrix of the first step is as follows:

All non-zero blocks are filled with $\pm 1$. The block $A_{1,1}$ is a $(2^{m_1}\times 2^{m_1})$-matrix with orthogonal rows, the block $A_{2,2}$ is a $((2^{m_2}-2^{m_1})\times 2^{m_2})$-matrix with orthogonal rows, $\dots$, the block $A_{s,s}$ is a $((2^{m_s}- 2^{m_{s-1}})\times 2^{m_s})$-matrix with orthogonal rows.

Thus the left part of $\mathbf{\Phi}$ consists of $2^{m_s}$ rows and $2^{m_1}+\dots + 2^{m_s}$ columns.

All blocks with second subscript $s+1$ consist of $p:= 2^{m_{s+1}}+\dots + 2^{m_t}$ columns. The number of rows in any block $A_{i,s+1}$, $i=1, \dots, s$, is equal to that of $A_{i,i}$, $i=1, \dots, s$, and the block $A_{s+1,s+1}$ consists of $d-2^{m_s}$ rows and $p$ columns.

Each block $A_{i,i}$, $i=1, \dots, s$, contains $2^{m_i}-2^{m_{i-1}}\leqslant 2^{m_i}$ orthogonal rows of length $2^{m_i}$. The existence of such rows is ensured by the remark.

We take a closer look at the right side of the matrix.

Split the block

into $t-s$ blocks $B_j$ of size $d\times 2^{m_j}$, $j=s+1,\dots,t$. Each block $B_j$ has $d\leqslant 2^{m_j}$ rows. Hence the blocks $B_j$ can be filled by $\pm1$ in such a way that the rows of length $2^{m_j}$ are mutually orthogonal. We obtain orthogonal rows of length $p$ by combining the rows of the blocks $B_j$. This yields a $(d\times n)$-matrix with entries $\pm 1$ and $0$ and with mutually orthogonal rows.

We proceed to normalize the entries of this matrix. All entries of the blocks $A_{i,i}$, $i=1,\dots,s$, are multiplied by $\alpha_i$, $i=1,\dots,s$, and all entries of the blocks $A_{j,s+1}$, $j=1,\dots,s+1$, are multiplied by $\beta_j$, $j=1,\dots,s+1$, in order to make the length of any row equal to $1$ and the length of any column equal to $\sqrt{d/n}$. The rows of the matrix remain orthogonal when we multiply the rows by any numbers. To specify the numbers $\alpha_i$ and $\beta_j$, we calculate the sums of squares of the entries in each column of $\Phi$:

$$ \begin{equation} \alpha_1^2\cdot2^{m_1}=\frac{d}{n},\quad \alpha_2^2\cdot(2^{m_2}-2^{m_1})=\frac{d}{n},\quad \dots,\quad \alpha_s^2\cdot(2^{m_s}-2^{m_{s-1}})=\frac{d}{n}, \end{equation} \tag{3.1} $$
$$ \begin{equation} \beta_1^2\cdot2^{m_1}+\beta_2^2\cdot(2^{m_2}-2^{m_1})+\dots+ \beta_s^2\cdot(2^{m_s}-2^{m_{s-1}})+ \beta_{s+1}^2\cdot(d-2^{m_{s}})=\frac{d}{n}. \end{equation} \tag{3.2} $$
Using (3.1), we obtain the values
$$ \begin{equation*} \alpha_1^2=\frac{d}{n\cdot2^{m_1}},\qquad \alpha_i^2=\frac{d}{n\cdot(2^{m_i}-2^{m_{i-1}})},\quad i=2,\dots,s. \end{equation*} \notag $$
By the condition of normalization of rows,
$$ \begin{equation} \begin{aligned} \, &\alpha_1^2\cdot2^{m_1}+\beta_1^2\cdot(2^{m_{s+1}}+\dots+2^{m_t})=1, \\ &\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots\dots \\ &\alpha_s^2\cdot2^{m_s}+\beta_s^2\cdot(2^{m_{s+1}}+\dots+2^{m_t})=1, \\ &\beta_{s+1}^2\cdot(2^{m_{s+1}}+\dots+2^{m_t})=1. \end{aligned} \end{equation} \tag{3.3} $$
Substituting the known values of $\alpha_i^2$ in (3.3), we have
$$ \begin{equation} \begin{aligned} \, \beta_1^2 &=\frac{1}{2^{m_{s+1}}+\dots+2^{m_t}}\biggl(1-\frac{d}{n}\biggr), \\ \beta_j^2 &=\frac{1}{2^{m_{s+1}}+\dots+2^{m_t}} \biggl(1-\frac{d\cdot2^{m_j}}{n(2^{m_j}-2^{m_{j-1}})}\biggr),\qquad j=2,\dots,s, \\ \beta_{s+1}^2 &=\frac{1}{2^{m_{s+1}}+ \dots+2^{m_t}}. \end{aligned} \end{equation} \tag{3.4} $$
It follows from the inequality $d<n/2$ that
$$ \begin{equation*} \begin{aligned} \, &1-\frac{d\cdot2^{m_j}}{n(2^{m_j}-2^{m_{j-1}})}> 1-\frac{\frac{n}{2}\cdot2^{m_j}}{n(2^{m_j}-2^{m_{j-1}})} \\ &\qquad=\frac{(2^{m_j}-2^{m_{j-1}})-2^{m_j-1}}{2^{m_j}-2^{m_{j-1}}} =\frac{2^{m_j-1}-2^{m_{j-1}}}{2^{m_j}-2^{m_{j-1}}}\geqslant0. \end{aligned} \end{equation*} \notag $$
Hence the right-hand sides of the equalities (3.4) are positive. This guarantees the existence of the numbers $\beta_j$.

The resulting numbers $\alpha_i$ and $\beta_j$ satisfy the equality (3.2). The systems of vectors thus obtained will be called Maltsev equal-norm tight frames.

We now consider examples.

Example. Consider the case when $d=3$, $n=7$. The number $n$ is represented as $n=2^0+2^1+2^2$. Then the synthesis matrix is

We find the values of $\alpha_i$ and $\beta_j$ from the normalization conditions:

Example. Consider the case when $d=7$, $n=15$. Since $15=2^0+2^1+2^2+2^3$, we obtain

where
$$ \begin{equation*} \begin{aligned} \, A_{1,1} &=\begin{pmatrix} \sqrt{\dfrac{7}{15}} \end{pmatrix},\qquad A_{2,2}=\begin{pmatrix} \sqrt{\dfrac{7}{15}}&\sqrt{\dfrac{7}{15}}\end{pmatrix}, \\ A_{3,3} &=\begin{pmatrix} \sqrt{\dfrac{7}{30}}&\sqrt{\dfrac{7}{30}}&\sqrt{\dfrac{7}{30}}&\sqrt{\dfrac{7}{30}} \\ \sqrt{\dfrac{7}{30}}&-\sqrt{\dfrac{7}{30}}&\sqrt{\dfrac{7}{30}}&-\sqrt{\dfrac{7}{30}} \end{pmatrix}, \\ A_{1,4} &=\begin{pmatrix} \sqrt{\dfrac{1}{15}}&\sqrt{\dfrac{1}{15}}&\sqrt{\dfrac{1}{15}}&\sqrt{\dfrac{1}{15}} &\sqrt{\dfrac{1}{15}}&\sqrt{\dfrac{1}{15}}&\sqrt{\dfrac{1}{15}}&\sqrt{\dfrac{1}{15}} \end{pmatrix}, \\ A_{2,4} &=\begin{pmatrix} \dfrac{1}{2\sqrt{30}}&-\dfrac{1}{2\sqrt{30}}&\dfrac{1}{2\sqrt{30}}&-\dfrac{1}{2\sqrt{30}} &\dfrac{1}{2\sqrt{30}}&-\dfrac{1}{2\sqrt{30}}&\dfrac{1}{2\sqrt{30}}&-\dfrac{1}{2\sqrt{30}} \end{pmatrix}, \\ A_{3,4} &= \begin{pmatrix} \dfrac{1}{2\sqrt{30}}&\dfrac{1}{2\sqrt{30}}&-\dfrac{1}{2\sqrt{30}}&-\dfrac{1}{2\sqrt{30}} &\dfrac{1}{2\sqrt{30}}&\dfrac{1}{2\sqrt{30}}&-\dfrac{1}{2\sqrt{30}}&\! -\dfrac{1}{2\sqrt{30}} \\ \dfrac{1}{2\sqrt{30}}&-\dfrac{1}{2\sqrt{30}}&-\dfrac{1}{2\sqrt{30}}&\dfrac{1}{2\sqrt{30}} &\dfrac{1}{2\sqrt{30}}&-\dfrac{1}{2\sqrt{30}}&-\dfrac{1}{2\sqrt{30}}&\dfrac{1}{2\sqrt{30}} \end{pmatrix}, \\ A_{4,4} &=\begin{pmatrix} \dfrac{1}{2\sqrt{2}}&\dfrac{1}{2\sqrt{2}}&\dfrac{1}{2\sqrt{2}}&\dfrac{1}{2\sqrt{2}} &-\dfrac{1}{2\sqrt{2}}&-\dfrac{1}{2\sqrt{2}}&-\dfrac{1}{2\sqrt{2}}&-\dfrac{1}{2\sqrt{2}} \\ \dfrac{1}{2\sqrt{2}}&-\dfrac{1}{2\sqrt{2}}&\dfrac{1}{2\sqrt{2}}&-\dfrac{1}{2\sqrt{2}} &-\dfrac{1}{2\sqrt{2}}&\dfrac{1}{2\sqrt{2}} &-\dfrac{1}{2\sqrt{2}}&\dfrac{1}{2\sqrt{2}} \\ \dfrac{1}{2\sqrt{2}}&\dfrac{1}{2\sqrt{2}}&-\dfrac{1}{2\sqrt{2}}&-\dfrac{1}{2\sqrt{2}} &-\dfrac{1}{2\sqrt{2}}&-\dfrac{1}{2\sqrt{2}}&\dfrac{1}{2\sqrt{2}}&\dfrac{1}{2\sqrt{2}} \end{pmatrix}. \end{aligned} \end{equation*} \notag $$

Example. Here is a Hadamard matrix of order $16$ constructed by Maltsev’s algorithm:

$$ \begin{equation} \left(\begin{array}{cccccccccccccccc} +&+&+&+&+&+&+&+&+&+&+&+&+&+&+&+\\ +&-&+&-&+&-&+&-&+&-&+&-&+&-&+&-\\ +&+&-&-&+&+&-&-&+&+&-&-&+&+&-&-\\ +&-&-&+&+&-&-&+&+&-&-&+&+&-&-&+\\ +&+&+&+&-&-&-&-&+&+&+&+&-&-&-&-\\ +&-&+&-&-&+&-&+&+&-&+&-&-&+&-&+\\ +&+&-&-&-&-&+&+&+&+&-&-&-&-&+&+\\ +&-&-&+&-&+&+&-&+&-&-&+&+&-&+&-\\ +&+&+&+&+&+&+&+&-&-&-&-&-&-&-&-\\ +&-&+&-&+&-&+&-&-&+&-&+&-&+&-&+\\ +&+&-&-&+&+&-&-&-&-&+&+&-&-&+&+\\ +&-&-&+&+&-&-&+&-&+&+&-&-&+&+&-\\ +&+&+&+&-&-&-&-&-&-&-&-&+&+&+&+\\ +&-&+&-&-&+&-&+&-&+&-&+&+&-&+&-\\ +&+&-&-&-&-&+&+&-&-&+&+&+&+&-&-\\ +&-&-&+&-&+&+&-&-&+&+&-&+&-&-&+ \end{array}\right). \end{equation} \tag{3.5} $$
Since the matrix consists of the numbers $\pm 1$, we show only their signs. When $d=6$ and $n=16$, we can choose any $6$ rows in the Hadamard $(16\times 16)$-matrix (3.5) and obtain equal-norm tight frames for $\mathbb{R}^6$. However, additional properties of the frame can be achieved by a special selection of rows.

We recall that an equal-norm frame $\{\varphi_j\}_{j=1}^n$ is said to be equiangular if there is a number $\gamma\geqslant 0$ such that $|\langle\varphi_{j},\varphi_{j^{'}}\rangle|=\gamma$ for all $j\neq j'$.

The matrix formed by the rows $1$, $9$, $5$, $3$, $2$, $16$ of the matrix (3.5) is the synthesis matrix of an equiangular frame. Another way of obtaining an equiangular tight frame was given in [9]. It uses the rows with numbers $2$, $3$, $4$, $6$, $11$, $16$. We do not know whether other choices of rows can give an equiangular frame.

Maltsev frames for $n\neq 2^k$ cannot be equiangular because the columns in the left part of the matrix are orthogonal to each other while those in the right part form at least one non-orthogonal pair.

In the literature on sparse representations, much attention is paid to the spark of the system (see [10]).

Definition 1. The spark of a $(d\times n)$-matrix $\mathbf{\Phi}$ is the minimal number of linearly dependent columns of $\mathbf{\Phi}$.

If $\operatorname{spark}(\Phi)=d+1$, then any subset of $d$ columns consists of linearly independent vectors. In this case, the system is called a full-spark frame.

Proposition 1. There are $4$ linearly dependent vectors in any Maltsev equal-norm tight frame of $n\geqslant 4$ vectors.

Proof. If $\{\varphi_i\}_{i=1}^n$ is a Maltsev tight frame with $n\geqslant4$, then the binary representation of $n$ contains powers $2^t$ with $t\geqslant2$. Let $t$ be the maximal number such that $2^t\leqslant n$. Then the last block $B_t$ of the $(d\times n)$-matrix $\Phi$ is of size $d\times 2^t$. We claim that there are $4$ linearly dependent columns in the block $B_t.$

Indeed, relabel the frame vectors in such a way that $\varphi_1,\dots,\varphi_{2^t}$ are the last $2^t$ columns of the matrix $\Phi$ (hence they form the block $B_t$). By the construction of Maltsev frames, $B_t$ is obtained by deleting any $2^t-d$ rows from the matrix

$$ \begin{equation*} \mathbf{H}_{2^t}=\left( \begin{array}{c|c} \rule{0pt}{14pt}\mathbf{H}_{2^{t-1}}&\mathbf{H}_{2^{t-1}}\\ \hline \rule{0pt}{14pt}\mathbf{H}_{2^{t-1}}&-\mathbf{H}_{2^{t-1}} \end{array}\right). \end{equation*} \notag $$
Let $B$ (resp. $B'$) be the result of deleting $2^{t-1}-p$ rows (resp. $2^{t-1}-d+p$ rows) from the matrix $\mathbf{O}_{t-1}$. Then the size of $B$ is $p\times2^{t-1}$, the size of $B'$ is $(d-p)\times2^{t-1}$ and we have
$$ \begin{equation*} B_t=\left( \begin{array}{c|c} \rule{0pt}{14pt}B & B\\ \hline \rule{0pt}{14pt}B' & -B' \end{array}\right). \end{equation*} \notag $$
For every $i=1,\dots,2^{t-1}$, the sum of the columns of the block $B_t$ with numbers $i$ and $k_i=2^{t-1}+i$ is equal to
$$ \begin{equation*} \varphi_{i}+\varphi_{k_i}= \left(\begin{array}{c} 2\beta_1\\ \dots\\ 2\beta_p\\0\\ \dots\\0\\ \end{array}\right) \begin{array}{l} \hspace{-6pt}\left.\begin{array}{c} \\\\\\ \end{array}\right\}p\\ \hspace{-6pt}\left.\begin{array}{c} \\\\\\ \end{array}\right\}d-p \end{array}, \end{equation*} \notag $$
where the numbers $\beta_j$ are those corresponding to the rows of the block $B_t$ in the construction of the Maltsev frame. Hence for any $i$, $j$ we have
$$ \begin{equation*} \varphi_{i}+\varphi_{k_i}=\varphi_{j}+\varphi_{k_j}. \end{equation*} \notag $$
Since the block $B_t$ has at least four columns, we conclude that among the frame vectors there are four linearly dependent vectors. $\Box$

Hence Maltsev tight frames can be full-spark only when $\operatorname{spark}(\Phi)=d+1\leqslant4$.

Let $\{\varphi_i\}_{i=1}^n$ be a full-spark Maltsev tight frame. We have seen that $d\leqslant3$. Since multiplication of the blocks $A_{i,j}$ by the non-zero numbers $\alpha_i$ and $\beta_j$ in the construction of $\Phi$ preserves the orthogonality of rows and the linear dependence or independence of columns, we can assume that all non-zero entries of the matrix $\Phi$ are equal to $\pm1$.

To find the possible sizes of the block $B_t$, we can always multiply the rows and columns of $\Phi$ by $\pm 1$ in such a way that all the entries in the first row and the first column of $B_t$ are equal to $1$.

Suppose that $d=2$ and $B_t$ has more than two columns. Since $\varphi_1=\begin{pmatrix}1\\ 1\\\end{pmatrix}$ and the first coordinates of $\varphi_2$ and $\varphi_3$ are equal to 1, we obtain two equal vectors in the frame for any choice of the second coordinates. This contradicts the equality $\operatorname{spark}(\Phi)=3$. Hence a full-spark Maltsev frame for $\mathbb{R}^2$ is possible only for $n=3$. This frame is uniquely determined by the algorithm above:

When $d=3$, a simple search yields a unique choice for the block $B_t$ (up to permutation of columns):

$$ \begin{equation*} B_t=\left(\begin{array}{cccc} 1&1&1&1\\ 1&1&-1&-1\\ 1&-1&-1&1\\ \end{array}\right). \end{equation*} \notag $$
Thus we obtain the following restrictions on the number of columns in $B_t$ and the number of vectors in the frame: $n\leqslant2^0+2^1+2^2=7$. Putting $n=5,6,7$, we see that none of the Maltsev frames is full-spark. Hence the matrix $B_2$ gives the unique possible full-spark Maltsev frame for $\mathbb{R}^3$.

Corollary 1. There are exactly two full-spark equal-norm tight Maltsev frames:


1. O. Christensen, An introduction to frames and Riesz bases, Appl. Numer. Harmon. Anal., Birkhäuser Boston, Inc., Boston, MA, 2003  crossref  mathscinet  zmath
2. M. A. Naimark, “Spectral functions of a symmetric operator”, Izv. Akad. Nauk SSSR Ser. Mat., 4:3 (1940), 277–318 (Russian)  mathnet  mathscinet  zmath
3. B. S. Kashin and T. Yu. Kulikova, “A note on the description of frames of general form”, Mat. Zametki, 72:6 (2002), 941–945  mathnet  crossref  mathscinet  zmath; English transl. Math. Notes, 72:6 (2002), 863–867  crossref
4. S. Ya. Novikov, “Bessel sequences as projections of orthogonal systems”, Mat. Zametki, 81:6 (2007), 893–903  mathnet  crossref  mathscinet  zmath; English transl. Math. Notes, 81:6 (2007), 800–809  crossref
5. A. I. Maltsev, “Remark on the paper 'A formula of Gauss in the theory of least squares' by A. N. Kolmogorov, A. A. Petrov, and Yu. M. Smirnov”, Izv. Akad. Nauk SSSR. Ser. Mat., 11:6 (1947), 567–568 (Russian)  mathnet  mathscinet  zmath
6. A. N. Kolmogorov, A. A. Petrov, and Yu. M. Smirnov, “A formula of Gauss in the theory of least squares”, Izv. Akad. Nauk SSSR. Ser. Mat., 11:6 (1947), 561–566 (Russian)  mathnet  mathscinet  zmath
7. P. G. Casazza and M. T. Leon, “Existence and construction of finite tight frames”, J. Concr. Appl. Math., 4:3 (2006), 277–289  mathscinet  zmath
8. M. S. Bespalov, “Eigenspaces of the discrete Walsh transform”, Probl. peredachi inform., 46:3 (2010), 60–79  mathnet  mathscinet  zmath; English transl. Problems Inform. Transmission, 46:3 (2010), 253–271  crossref
9. M. Fickus, J. Jasper, D. G. Mixon, and J. Peterson, “Hadamard equiangular tight frames”, Appl. Comput. Harmon. Anal., 50:1 (2021), 281–302  crossref  mathscinet  zmath
10. M. Elad, Sparse and redundant representations. From theory to applications in signal and image processing, Springer, New York, 2010  mathscinet  zmath

Citation: S. Ya. Novikov, V. V. Sevost'yanova, “Maltsev equal-norm tight frames”, Izv. RAN. Ser. Mat., 86:4 (2022), 162–174; Izv. Math., 86:4 (2022), 770–781
Citation in format AMSBIB
\Bibitem{NovSev22}
\by S.~Ya.~Novikov, V.~V.~Sevost'yanova
\paper Maltsev equal-norm tight frames
\jour Izv. RAN. Ser. Mat.
\yr 2022
\vol 86
\issue 4
\pages 162--174
\mathnet{http://mi.mathnet.ru/im9137}
\crossref{https://doi.org/10.4213/im9137}
\mathscinet{http://mathscinet.ams.org/mathscinet-getitem?mr=4461245}
\adsnasa{https://adsabs.harvard.edu/cgi-bin/bib_query?2022IzMat..86..770N}
\transl
\jour Izv. Math.
\yr 2022
\vol 86
\issue 4
\pages 770--781
\crossref{https://doi.org/10.4213/im9137e}
\isi{https://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=Publons&SrcAuth=Publons_CEL&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=000992245100006}
\scopus{https://www.scopus.com/record/display.url?origin=inward&eid=2-s2.0-85165645608}
Linking options:
  • https://www.mathnet.ru/eng/im9137
  • https://doi.org/10.4213/im9137e
  • https://www.mathnet.ru/eng/im/v86/i4/p162
  • This publication is cited in the following 1 articles:
    Citing articles in Google Scholar: Russian citations, English citations
    Related articles in Google Scholar: Russian articles, English articles
    Известия Российской академии наук. Серия математическая Izvestiya: Mathematics
    Statistics & downloads:
    Abstract page:354
    Russian version PDF:58
    English version PDF:69
    Russian version HTML:201
    English version HTML:53
    References:59
    First page:8
     
      Contact us:
     Terms of Use  Registration to the website  Logotypes © Steklov Mathematical Institute RAS, 2024