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

Jordan chains of hh-cyclic matrices, II

Andrew L. Nickerson [email protected] Pietro Paparella [email protected] [ Division of Engineering and Mathematics, University of Washington Bothell, Bothell, WA 98011-8246, USA
Abstract

McDonald and Paparella [Linear Algebra Appl. 498 (2016), 145–159] gave a necessary condition on the structure of Jordan chains of hh-cyclic matrices. In this work, that necessary condition is shown to be sufficient. As a consequence, we provide a spectral characterization of nonsingular, hh-cyclic matrices. In addition, we provide results for the Jordan chains corresponding to the eigenvalue zero of singular matrices. Along the way, a new characterization of circulant matrices is given.

keywords:
digraph , bipartite digraph , cyclically hh-partite digraph , circulant matrix
MSC:
[2020] 15A18 , 15A20 , 15B99
journal: Linear Algebra and its Applications

url]http://faculty.washington.edu/pietrop/

1 Introduction

A celebrated result in spectral graph theory states that a graph is bipartite if and only if the spectrum of its adjacency matrix is symmetric (i.e., λ-\lambda is an eigenvalue whenever λ\lambda is). However, as noted by Nikiforov [6, p. 3], a bipartite digraph can not, in general, be characterized by its spectrum alone and restrictions on the Jordan structure of the matrix must be considered.

Recall that a directed graph (or digraph, for brevity) Γ=(V,E)\Gamma=(V,E) consists of a finite, nonempty set VV of vertices, together with a set EV×VE\subseteq V\times V of arcs. If AA is an nn-by-nn matrix with entries over a field 𝔽\mathbb{F}, then the digraph of AA, denoted by Γ=ΓA\Gamma=\Gamma_{A}, has vertex set V={1,,n}V=\{1,\dots,n\} and arc set E={(i,j)V×Vaij0}E=\{(i,j)\in V\times V\mid a_{ij}\neq 0\}.

A digraph Γ=(V,E)\Gamma=(V,E) is called hh-partite if there is a partition P={V1,,Vh}P=\{V_{1},\ldots,V_{h}\} of VV such that, for each arc (i,j)E(i,j)\in E, there are positive integers k,{1,,h}k,\ell\in\{1,\dots,h\} such that iVi\in V_{\ell} and jVkj\in V_{k}. A digraph Γ=(V,E)\Gamma=(V,E) is called cyclically hh-partite if there is a partition P={V1,,Vh}P=\{V_{1},\dots,V_{h}\} of VV such that, for each arc (i,j)E(i,j)\in E, there is a positive integer {1,,h}\ell\in\{1,\dots,h\} such that iVi\in V_{\ell} and jV+1j\in V_{\ell+1} (where, for convenience, Vh+1:=V1V_{h+1}:=V_{1}). Notice that there is no distinction between a cyclically bipartite digraph and a bipartite digraph, but there is if h>2h>2. Furthermore, notwithstanding its more restrictive nature, characterizing the spectral properties of cyclically hh-partite digraphs subsumes the case of bipartite digraphs.

A matrix AA is called hh-cyclic or cyclically hh-partite if ΓA\Gamma_{A} is cyclically hh-partite. McDonald and Paparella [4, Lemma 4.1] showed that if AA is a nonsingular, hh-cyclic matrix with complex entries, then a given Jordan chain corresponding to an eigenvalue λ\lambda of AA determines a Jordan chain corresponding to λωk\lambda\omega^{k}, 1kh11\leq k\leq h-1, where ω:=exp(2πi/h)\omega:=\exp(2\pi\textup{i}/h) (for full details, see Lemma 4.10 below). As an immediate consequence, if the nonsingular Jordan block Jp(λ)J_{p}(\lambda) appears in any Jordan canonical form of AA, then the nonsingular Jordan block Jp(λωk)J_{p}(\lambda\omega^{k}) also appears in any Jordan canonical form of AA, k{1,,h1}\forall k\in\{1,\dots,h-1\} [4, Corollary 4.2].

The central aim of this work is to establish a converse of this result—i.e., if a given Jordan chain corresponding to an eigenvalue λ\lambda of a matrix A𝖬n()A\in\mathsf{M}_{n}(\mathbb{C}) determines a Jordan chain chain for λωk\lambda\omega^{k}, for every eigenvalue λ\lambda of AA, then AA is hh-cyclic (see Theorem 4.14 below). As a consequence, we provide a spectral characterization of nonsingular, hh-cyclic matrices. In addition, we provide results for Jordan chains corresponding to the eigenvalue zero of singular matrices and implicitly provide an algorithm on computing these chains. Along the way, we provide a new characterization of circulant matrices and use this characterization to provide a more rigorous proof of a result by McDonald and Paparella [4, Lemma 4.3].

2 Notation and Background

For nn\in\mathbb{N}, denote by: n\langle n\rangle the set {1,,n}\{1,\ldots,n\}; R(n)R(n) the set {0,1,,n1}\{0,1,\ldots,n-1\}; and ω=ωn\omega=\omega_{n} the complex number exp(2πi/n)\exp(2\pi\textup{i}/n).

The algebra of mm-by-nn matrices over a field 𝔽\mathbb{F} is denoted by 𝖬m×n(𝔽)\mathsf{M}_{m\times n}(\mathbb{F}); when m=nm=n, the set 𝖬m×n(𝔽)\mathsf{M}_{m\times n}(\mathbb{F}) is abbreviated to 𝖬n(𝔽)\mathsf{M}_{n}(\mathbb{F}). If A𝖬m×n(𝔽)A\in\mathsf{M}_{m\times n}(\mathbb{F}), then the (i,j)(i,j)-entry of AA is denoted by [A]ij[A]_{ij}, aija_{ij}, or ai,ja_{i,j}.

If A𝖬n()A\in\mathsf{M}_{n}(\mathbb{C}) and RR, C{1,,n}C\subseteq\{1,\dots,n\}, then A(R,C)A(R,C) denotes the submatrix of AA whose rows and columns are indexed by RR and CC, respectively.

If A,B𝖬m×n(𝔽)A,B\in\mathsf{M}_{m\times n}(\mathbb{F}), then the Hadamard product of AA and BB, denoted by ABA\circ B, is the m×nm\times n matrix such that [AB]ij=aijbij[A\circ B]_{ij}=a_{ij}b_{ij}.

If λ𝔽\lambda\in\mathbb{F}, then the nn-by-nn Jordan block with eigenvalue λ\lambda, denoted by Jn(λ)J_{n}(\lambda), is defined by

Jn(λ)=λi=1nEii+i=1n1Ei,i+1=[λ1λ1λ1λ]𝖬n(𝔽),J_{n}(\lambda)=\lambda\sum_{i=1}^{n}E_{ii}+\sum_{i=1}^{n-1}E_{i,i+1}=\begin{bmatrix}\lambda&1&\\ &\lambda&1\\ &&\ddots&\ddots\\ &&&\lambda&1\\ &&&&\lambda\end{bmatrix}\in\mathsf{M}_{n}(\mathbb{F}),

where EijE_{ij} is the nn-by-nn matrix whose (i,j)(i,j)-entry equals one and zero otherwise.

2.1 Cyclically hh-partite matrices

If ΓA\Gamma_{A} is cyclically hh-partite with partition PP, then AA is said to be h-cyclic with partition PP or that PP describes the hh-cyclic structure of A. The partition P={V1,,Vh}P=\{V_{1},\dots,V_{h}\} is consecutive if V1={1,,i1}V_{1}=\{1,\dots,i_{1}\}, V2={i1+1,,i2},,Vh={ih1+1,,n}V_{2}=\{i_{1}+1,\dots,i_{2}\},\ldots,V_{h}=\{i_{h-1}+1,\dots,n\}. If AA is hh-cyclic with consecutive partition PP, then AA is of the form

[\@arstrut12h\\1A12\\\\h-1A-h1,h\\hAh1] ,\displaystyle\hbox{}\vbox{\kern 0.86108pt\hbox{$\kern 0.0pt\kern 2.5pt\kern-5.0pt\left[\kern 0.0pt\kern-2.5pt\kern-5.55557pt\vbox{\kern-0.86108pt\vbox{\vbox{ \halign{\kern\arraycolsep\hfil\@arstrut$\kbcolstyle#$\hfil\kern\arraycolsep& \kern\arraycolsep\hfil$\@kbrowstyle#$\ifkbalignright\relax\else\hfil\fi\kern\arraycolsep&& \kern\arraycolsep\hfil$\@kbrowstyle#$\ifkbalignright\relax\else\hfil\fi\kern\arraycolsep\cr 5.0pt\hfil\@arstrut$\scriptstyle$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle 1$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle 2$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle\cdots$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle h\\1$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle A_{12}$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle\\\vdots$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle\ddots$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle\\h-1$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle A_{h-1,h}\\h$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle A_{h1}$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle$\hfil\kern 5.0pt&5.0pt\hfil$\scriptstyle$\hfil\kern 5.0pt\crcr}}}}\right]$}}, (2)

where Ai,i+1=A(Vi,Vi+1)𝖬|Vi|×|Vi+1|()A_{i,i+1}=A(V_{i},V_{i+1})\in\mathsf{M}_{|V_{i}|\times|V_{i+1}|}(\mathbb{C}), ih\forall i\in\langle h\rangle [1, p. 71]. If PP is not consecutive, then there is a permutation matrix QQ such that QAQQ^{\top}AQ is hh-cyclic with consecutive partition [1, p. 71]. Given that the eigenvalues of a matrix do not change with respect to permuatition similarity, hereinafter it is assumed, without loss generality, that every hh-cyclic matrix is of the form (2).

If xnx\in\mathbb{C}^{n} and

x=[x1xh],x=\begin{bmatrix}x_{1}\\ \vdots\\ x_{h}\end{bmatrix},

where xi|Vi|x_{i}\in\mathbb{C}^{|V_{i}|}, then xx is said to be conformably partitioned with AA (or PP).

Suppose that x1,,xpnx_{1},\dots,x_{p}\in\mathbb{C}^{n}. Recall that if x1x_{1} is an eigenvector corresponding to λ\lambda\in\mathbb{C} and Axi=λxi+xi1Ax_{i}=\lambda x_{i}+x_{i-1}, 1<ip1<i\leq p, then {x1,,xp}\{x_{1},\dots,x_{p}\} is called a Jordan chain of AA corresponding to λ\lambda. Furthermore, it can be shown that if {x1,,xp}\{x_{1},\ldots,x_{p}\} is a Jordan chain, then {x1,,xp}\{x_{1},\ldots,x_{p}\} is linearly independent and xk=(AλI)pkxpx_{k}=(A-\lambda I)^{p-k}x_{p}, 1kp1\leq k\leq p.

The following partial products of submatrices of an hh-cyclic matrix will be of use in the sequel.

Definition 2.1.

Let A𝖬n()A\in\mathsf{M}_{n}(\mathbb{C}) and suppose that AA is of the form (2). For ihi\in\langle h\rangle and pp\in\mathbb{N}, let

Bip:=j=h+1phAαj1(i),αj(i),B_{ip}:=\prod_{j=h+1-p}^{h}A_{\alpha^{j-1}(i),\alpha^{j}(i)},

where αSh\alpha\in S_{h} is the hh-cycle of order hh defined by α(i)=imodh+1\alpha(i)=i\bmod{h}+1. For ease of notation, the matrix BihB_{ih} is abbreviated to BiB_{i}. Notice that BiB_{i} is a square matrix of order |Vi||V_{i}|.

Remark 2.2.

Notice that

α1(i)={i1,1<ihh,i=1.\displaystyle\alpha^{-1}(i)=\begin{cases}i-1,&1<i\leq h\\ h,&i=1.\end{cases}

and αj\alpha^{j} is defined if j<0j<0 since α\alpha is an invertible map.

If AA is of the form (2), then

Ah=j=1hBj=[B100Bh]\displaystyle A^{h}=\bigoplus_{j=1}^{h}B_{j}=\begin{bmatrix}B_{1}&&\hbox{\multirowsetup\LARGE 0}\\ \hbox{\multirowsetup\LARGE 0}&\ddots&\\ &&B_{h}\end{bmatrix} (3)

(see, e.g., Brualdi and Ryser [1, p. 73]).

We note the following useful theorem due to Mirsky.

Theorem 2.3 (Mirsky [5, Theorem 1]).

Let A𝖬n()A\in\mathsf{M}_{n}(\mathbb{C}) and suppose that AA is of the form (2). If λ1,,λm\lambda_{1},\ldots,\lambda_{m} are the nonzero eigenvalues of B1B_{1}, then the spectrum of AA consists of nhmn-hm zeros and the hmhm hh-th roots of λ1,,λm\lambda_{1},\ldots,\lambda_{m}.

2.2 Circulant Matrices

We digress to briefly discuss circulant matrices (for a general reference, see Davis [2]).

Definition 2.4 ([2, p. 66]).

If r:=(r1,,rn)r:=\begin{pmatrix}r_{1},\ldots,r_{n}\end{pmatrix}, where rir_{i}\in\mathbb{C}, ini\in\langle n\rangle, and C𝖬n()C\in\mathsf{M}_{n}(\mathbb{C}), then CC is called a circulant or a circulant matrix with reference vector rr, denoted circ(r)\operatorname{circ}(r), if

cij=r((ji)modn)+1c_{ij}=r_{((j-i)\bmod{n})+1}

for every (i,j)n2(i,j)\in\langle n\rangle^{2}.

Definition 2.5.

Denote by eie_{i} the iith canonical basis vector of n\mathbb{C}^{n}. For nn\in\mathbb{N}, n2n\geq 2, the basic circulant of order nn, denoted by KnK_{n}, is the circulant matrix defined by Kn=circ(e2)K_{n}=\operatorname{circ}(e_{2}).

For example,

K2=[0110],K_{2}=\begin{bmatrix}0&1\\ 1&0\end{bmatrix},
K3=[010001100],K_{3}=\begin{bmatrix}0&1&0\\ 0&0&1\\ 1&0&0\end{bmatrix},

and, in general,

Kn=[In11].K_{n}=\begin{bmatrix}&I_{n-1}\\ 1&\end{bmatrix}.

The following result, although unsurprising, provides a characterization of circulant matrices that, to the best of our knowledge, has not appeared previously in the literature.

Theorem 2.6.

If C𝖬n()C\in\mathsf{M}_{n}(\mathbb{C}), then CC is a circulant matrix if and only if ci1,j1=ci2,j2c_{i_{1},j_{1}}=c_{i_{2},j_{2}} for all (i1,j1)(i2,j2)(i_{1},j_{1})\sim(i_{2},j_{2}), where \sim is the equivalence relation defined by

(i1,j1)(i2,j2)j1i1(j2i2)modn.\displaystyle(i_{1},j_{1})\sim(i_{2},j_{2})\iff j_{1}-i_{1}\equiv(j_{2}-i_{2})\bmod{n}.
Proof.

Suppose that C=circ(r)C=\operatorname{circ}(r). If (i1,j1)(i2,j2)(i_{1},j_{1})\sim(i_{2},j_{2}), then

ci1,j1=r((j1i1)modn)+1=r((j2i2)modn)+1=ci2,j2,c_{i_{1},j_{1}}=r_{((j_{1}-i_{1})\bmod{n})+1}=r_{((j_{2}-i_{2})\bmod{n})+1}=c_{i_{2},j_{2}},

as desired.

Conversely, suppose that ci1,j1=ci2,j2c_{i_{1},j_{1}}=c_{i_{2},j_{2}} for all (i1,j1)(i2,j2)(i_{1},j_{1})\sim(i_{2},j_{2}) and let

rk:=c1k,kn.r_{k}:=c_{1k},~{}\forall k\in\langle n\rangle.

If (i,j)n2(i,j)\in\langle n\rangle^{2} and m:=(ji)modnm:=(j-i)\bmod{n}, then

(ji)modn=mmodn=(m+11)modn,(j-i)\bmod{n}=m\bmod{n}=(m+1-1)\bmod{n},

i.e., (i,j)(1,m+1)(i,j)\sim(1,m+1). Thus,

cij=c1,m+1=rm+1=r((ji)modn)+1c_{ij}=c_{1,m+1}=r_{m+1}=r_{((j-i)\bmod{n})+1}

as desired. ∎

3 Nonsingular hh-cyclic Matrices

We briefly digress to address some properties of nonsingular, hh-cyclic matrices.

Theorem 3.7.

Let A𝖬n()A\in\mathsf{M}_{n}(\mathbb{C}) be hh-cyclic and nonsingular. If P={V1,,Vh}P=\{V_{1},\dots,V_{h}\} describes the hh-cyclic structure of AA, then |Vi|=|Vj||V_{i}|=|V_{j}|, i,jh\forall i,j\in\langle h\rangle.

Proof.

For contradiction, suppose that |Vi||Vj||V_{i}|\neq|V_{j}|. Without losing generality, it may be assumed that |Vj|>|Vi||V_{j}|>|V_{i}|. Notice that Bj𝖬|Vj|()B_{j}\in\mathsf{M}_{|V_{j}|}(\mathbb{C}) and

rank(Bj)\displaystyle\rank(B_{j}) rank(k=1hAαk1(j),αk(j))\displaystyle\leq\rank\left(\prod_{k=1}^{h}A_{\alpha^{k-1}(j),\alpha^{k}(j)}\right)
min{rank((A12)),,rank((Ah1,h)),rank((Ah,1))}\displaystyle\leq\min\left\{\rank{(A_{12})},\ldots,\rank{(A_{h-1,h})},\rank{(A_{h,1})}\right\}
|Vi|\displaystyle\leq|V_{i}|
<|Vj|,\displaystyle<|V_{j}|,

i.e., BjB_{j} is rank-deficient. Thus, 0σ(Bj)0\in\sigma(B_{j}), but

0σ(Bj)0σ(Ah)0σ(A)0\in\sigma(B_{j})\implies 0\in\sigma(A^{h})\iff 0\in\sigma(A)

which contradicts the assumption that AA is non-singular. ∎

Corollary 3.8.

If A𝖬n()A\in\mathsf{M}_{n}(\mathbb{C}) is hh-cyclic and nonsingular, then hh divides nn.

Remark 3.9.

The converse of Theorem 3.7 does not hold; indeed, consider the singular bipartite matrix

[0011001111001100].\begin{bmatrix}0&0&1&1\\ 0&0&1&1\\ 1&1&0&0\\ 1&1&0&0\end{bmatrix}.

4 Main Results

Given that it will be used in several of the subsequent results, hereinafter, αij:=(ij)modh\alpha_{ij}:=(i-j)\bmod{h}, i,j\forall i,j\in\mathbb{Z} and h\forall h\in\mathbb{N}.

The following result was established by McDonald and Paparella [4, Lemma 4.1] for nonsingular matrices; however, examining the proof reveals that the supposition is unnecessary.

Lemma 4.10 ([4, Lemma 4.1]).

Let A𝖬n()A\in\mathsf{M}_{n}(\mathbb{C}) and suppose that AA is of the form (2).

  1. 1.

    If {x0,j}j=1p\left\{x_{\langle 0,j\rangle}\right\}_{j=1}^{p} is a right Jordan chain corresponding to λσ(A)\lambda\in\sigma(A), where x0,jx_{\left<0,j\right>} is partitioned conformably with respect to AA as

    x0,j=[x1jxhj],jp,\displaystyle x_{\left<0,j\right>}=\begin{bmatrix}x_{1j}\\ \vdots\\ x_{hj}\end{bmatrix},~{}j\in\langle p\rangle,

    then, for kR(h)k\in R(h), the set

    {xk,j:=[(ωk)α1jx1j(ωk)αhjxhj]}j=1p\displaystyle\left\{x_{\left<k,j\right>}:=\begin{bmatrix}(\omega^{k})^{\alpha_{1j}}x_{1j}\\ \vdots\\ (\omega^{k})^{\alpha_{hj}}x_{hj}\end{bmatrix}\right\}_{j=1}^{p}

    is a right Jordan chain corresponding to λωk\lambda\omega^{k}.

  2. 2.

    If {yj,0}j=1p\left\{y_{\left<j,0\right>}\right\}_{j=1}^{p} is a left Jordan chain corresponding to λσ(A)\lambda\in\sigma(A), where yj,0y_{\left<j,0\right>} is partitioned conformably with respect to AA as

    yj,0=[yj1yjh],jp,\displaystyle y_{\left<j,0\right>}^{\top}=\begin{bmatrix}y_{j1}^{\top}&\cdots&y_{jh}^{\top}\end{bmatrix},~{}j\in\langle p\rangle,

    then, for kR(h)k\in R(h), the set

    {yj,k:=[(ωk)αj1yj1(ωk)αjhyjh]}j=1p\displaystyle\left\{y_{\left<j,k\right>}^{\top}:=\begin{bmatrix}(\omega^{k})^{\alpha_{j1}}y_{j1}^{\top}&\cdots&(\omega^{k})^{\alpha_{jh}}y_{jh}^{\top}\end{bmatrix}\right\}_{j=1}^{p}

    is a left Jordan chain corresponding to λωk\lambda\omega^{k}.

Remark 4.11.

Although McDonald and Paparella stated the aforementioned result for nonzero eigenvalues, the proof given is valid when λ=0\lambda=0. However, if any Jordan canonical form of AA has a singular Jordan block of size pp-by-pp, then the result does not yield another Jordan block (it does, at least, yield a different Jordan chain for that eigenvalue).

For example, consider the 33-cyclic matrix AA defined by

A=[011100000010000001000011100000100000].\displaystyle A=\left[\begin{array}[]{*{1}{c}|*{3}{c}|*{2}{c}}0&1&1&1&0&0\\ \hline\cr 0&0&0&0&1&0\\ 0&0&0&0&0&1\\ 0&0&0&0&1&1\\ \hline\cr 1&0&0&0&0&0\\ 1&0&0&0&0&0\end{array}\right].

Since B1=[4]B_{1}=\begin{bmatrix}4\end{bmatrix}, it follows that the nonzero eigenvalues of AA are 22/32^{2/3}, 22/3ω2^{2/3}\omega, and 22/3ω22^{2/3}\omega^{2}. Furthermore, since

[011100000010000001000011100000100000][000011]=[011000],\left[\begin{array}[]{*{1}{c}|*{3}{c}|*{2}{c}}0&1&1&1&0&0\\ \hline\cr 0&0&0&0&1&0\\ 0&0&0&0&0&1\\ 0&0&0&0&1&1\\ \hline\cr 1&0&0&0&0&0\\ 1&0&0&0&0&0\end{array}\right]\left[\begin{array}[]{r}0\\ \hline\cr 0\\ 0\\ 0\\ \hline\cr 1\\ -1\end{array}\right]=\left[\begin{array}[]{r}0\\ \hline\cr 1\\ -1\\ 0\\ \hline\cr 0\\ 0\end{array}\right],

and

[011100000010000001000011100000100000][011000],\left[\begin{array}[]{*{1}{c}|*{3}{c}|*{2}{c}}0&1&1&1&0&0\\ \hline\cr 0&0&0&0&1&0\\ 0&0&0&0&0&1\\ 0&0&0&0&1&1\\ \hline\cr 1&0&0&0&0&0\\ 1&0&0&0&0&0\end{array}\right]\left[\begin{array}[]{r}0\\ \hline\cr 1\\ -1\\ 0\\ \hline\cr 0\\ 0\end{array}\right],

it follows that

{x1:=[011000],x2:=[000011]}\left\{x_{1}:=\begin{bmatrix}0\\ 1\\ -1\\ 0\\ 0\\ 0\end{bmatrix},\ x_{2}:=\begin{bmatrix}0\\ 0\\ 0\\ 0\\ 1\\ -1\end{bmatrix}\right\}

is Jordan chain of length two associated with the eigenvalue zero. By Lemma 4.10,

{x^1=[(ω)α110(ω)α211(ω)α21(1)(ω)α210(ω)α310(ω)α310]=[0ωω000],x^2=[(ω)α120(ω)α220(ω)α220(ω)α220(ω)α321(ω)α32(1)]=[0000ωω]}\left\{\hat{x}_{1}=\begin{bmatrix}(\omega)^{\alpha_{11}}\cdot 0\\ (\omega)^{\alpha_{21}}\cdot 1\\ (\omega)^{\alpha_{21}}\cdot(-1)\\ (\omega)^{\alpha_{21}}\cdot 0\\ (\omega)^{\alpha_{31}}\cdot 0\\ (\omega)^{\alpha_{31}}\cdot 0\end{bmatrix}=\begin{bmatrix}0\\ \omega\\ -\omega\\ 0\\ 0\\ 0\end{bmatrix},\ \hat{x}_{2}=\begin{bmatrix}(\omega)^{\alpha_{12}}\cdot 0\\ (\omega)^{\alpha_{22}}\cdot 0\\ (\omega)^{\alpha_{22}}\cdot 0\\ (\omega)^{\alpha_{22}}\cdot 0\\ (\omega)^{\alpha_{32}}\cdot 1\\ (\omega)^{\alpha_{32}}\cdot(-1)\end{bmatrix}=\begin{bmatrix}0\\ 0\\ 0\\ 0\\ \omega\\ -\omega\end{bmatrix}\right\}

is also a Jordan chain corresponding to zero, but any Jordan canonical form of AA has one Jordan block of order two and one Jordan block of order one.

The following results were presented by McDonald and Paparella [4, Lemma 4.3 and Remark 4.4]. We repeat the results for the sake of completeness and to give more rigorous proofs that utilize the novel characterization of circulant matrices given in Theorem 2.6.

Lemma 4.12 (c.f. [4, Lemma 4.3]).

If kR(h)k\in R(h) and r\ell\in\langle r\rangle, then

Wk1\displaystyle W_{k\ell}^{1} :=ωk[(ωk)α1(ωk)αh][(ωk)α1(ωk)αh]\displaystyle:=\omega^{k}\begin{bmatrix}(\omega^{k})^{\alpha_{1\ell}}\\ \vdots\\ (\omega^{k})^{\alpha_{h\ell}}\end{bmatrix}\begin{bmatrix}(\omega^{k})^{\alpha_{\ell 1}}&\ldots&(\omega^{k})^{\alpha_{\ell h}}\end{bmatrix}
=circ(ωk,1,(ωk)h1,,(ωk)2)\displaystyle=\operatorname{circ}(\omega^{k},1,(\omega^{k})^{h-1},\ldots,(\omega^{k})^{2})

and

Wk2\displaystyle W_{k\ell}^{2} :=[(ωk)α1(ωk)αh][(ωk)α(+1)1(ωk)α(+1)h]\displaystyle:=\begin{bmatrix}(\omega^{k})^{\alpha_{1\ell}}\\ \vdots\\ (\omega^{k})^{\alpha_{h\ell}}\end{bmatrix}\begin{bmatrix}(\omega^{k})^{\alpha_{(\ell+1)1}}&\ldots&(\omega^{k})^{\alpha_{(\ell+1)h}}\end{bmatrix}
=circ(ωk,1,(ωk)h1,,(ωk)2).\displaystyle=\operatorname{circ}(\omega^{k},1,(\omega^{k})^{h-1},\ldots,(\omega^{k})^{2}).
Proof.

The following facts are easily established:

α\displaystyle\alpha βmodh(ωk)α=(ωk)β,k\displaystyle\equiv\beta\bmod{h}\Longrightarrow(\omega^{k})^{\alpha}=(\omega^{k})^{\beta},~{}k\in\mathbb{Z} (4)
αij\displaystyle\alpha_{ij} =(αi+αj)modh,i,j,\displaystyle=(\alpha_{i\ell}+\alpha_{\ell j})\bmod{h},~{}\forall i,j,\ell\in\mathbb{Z} (5)
αi+1,j\displaystyle\alpha_{i+1,j} =αi,j1=(αij+1)modh,i,j.\displaystyle=\alpha_{i,j-1}=(\alpha_{ij}+1)\bmod{h},~{}\forall i,j\in\mathbb{Z}. (6)

If kR(h)k\in R(h) and r\ell\in\langle r\rangle, then

[Wk1]ij\displaystyle\left[W_{k\ell}^{1}\right]_{ij} =ωk(ωk)αi(ωk)αj\displaystyle=\omega^{k}(\omega^{k})^{\alpha_{i\ell}}(\omega^{k})^{\alpha_{\ell j}}
=ωk(ωk)(αi+αj)modh\displaystyle=\omega^{k}(\omega^{k})^{(\alpha_{i\ell}+\alpha_{\ell j})\bmod{h}} (by (4))
=ωk(ωk)αij\displaystyle=\omega^{k}(\omega^{k})^{\alpha_{ij}} (by (5))
=(ωk)(αij+1)modh\displaystyle=(\omega^{k})^{(\alpha_{ij}+1)\bmod{h}} (by (4))
=(ωk)αi+1,j\displaystyle=(\omega^{k})^{\alpha_{i+1,j}} (by (6))

If (i,j)(p,q)(i,j)\sim(p,q), then

[Wk1]ij=(ωk)αi+1,j=(ωk)αp+1,q=[Wk1]pq,\left[W_{k\ell}^{1}\right]_{ij}=(\omega^{k})^{\alpha_{i+1,j}}=(\omega^{k})^{\alpha_{p+1,q}}=\left[W_{k\ell}^{1}\right]_{pq},

so that, by Theorem 2.6, Wk1W_{k\ell}^{1} is a circulant. Since

[Wk1]1j=(ωk)α2j=(ωk)(2j)modh,\left[W_{k\ell}^{1}\right]_{1j}=(\omega^{k})^{\alpha_{2j}}=(\omega^{k})^{(2-j)\bmod{h}},

it follows that

([Wk1]11,[Wk1]12,[Wk1]13,,[Wk1]1h)=(ωk,1,(ωk)h1,,(ωk)2),\displaystyle\begin{pmatrix}\left[W_{k\ell}^{1}\right]_{11},\left[W_{k\ell}^{1}\right]_{12},\left[W_{k\ell}^{1}\right]_{13},\ldots,\left[W_{k\ell}^{1}\right]_{1h}\end{pmatrix}=\begin{pmatrix}\omega^{k},1,(\omega^{k})^{h-1},\ldots,(\omega^{k})^{2}\end{pmatrix},

as desired.

For the second claim, notice that

[Wk2]ij\displaystyle\left[W_{k\ell}^{2}\right]_{ij} =(ωk)αi(ωk)α+1,j\displaystyle=(\omega^{k})^{\alpha_{i\ell}}(\omega^{k})^{\alpha_{\ell+1,j}}
=(ωk)αi(ωk)α,j1\displaystyle=(\omega^{k})^{\alpha_{i\ell}}(\omega^{k})^{\alpha_{\ell,j-1}} (by (6))
=(ωk)(αi+α,j1)modh\displaystyle=(\omega^{k})^{(\alpha_{i\ell}+\alpha_{\ell,j-1})\bmod{h}} (by (4))
=(ωk)αi,j1\displaystyle=(\omega^{k})^{\alpha_{i,j-1}} (by (5))
=(ωk)αi+1,j\displaystyle=(\omega^{k})^{\alpha_{i+1,j}} (by (6))
=[Wk1]ij,\displaystyle=\left[W_{k\ell}^{1}\right]_{ij},

as desired. ∎

Lemma 4.13 (c.f. [4, Remark 4.4]).

If kR(h)k\in R(h) and

Ck:=circ(ωk,1,(ωk)h1,,(ωk)2)𝖬h(),C_{k}:=\operatorname{circ}(\omega^{k},1,(\omega^{k})^{h-1},\ldots,(\omega^{k})^{2})\in\mathsf{M}_{h}(\mathbb{C}), (7)

then

k=0h1Ck=hKh=circ(0,h,0,,0).\sum_{k=0}^{h-1}C_{k}=hK_{h}=\operatorname{circ}(0,h,0,\ldots,0).
Proof.

Follows from the fact that circulant matrices are closed with respect to addition [2, Theorem 3.24] and the fact that

k=0h1(ωk)p=h,\sum_{k=0}^{h-1}(\omega^{k})^{p}=h,

if p=0p=0 and

k=0h1(ωk)p=k=0h1(ωp)k=(ωp)h1ωp1=(ωh)p1ωp1=0,\sum_{k=0}^{h-1}(\omega^{k})^{p}=\sum_{k=0}^{h-1}(\omega^{p})^{k}=\frac{(\omega^{p})^{h}-1}{\omega^{p}-1}=\frac{(\omega^{h})^{p}-1}{\omega^{p}-1}=0,

whenever p0p\neq 0. ∎

Theorem 4.14.

Let A𝖬n()A\in\mathsf{M}_{n}(\mathbb{C}) and let P={V1,,Vh}P=\{V_{1},\ldots,V_{h}\} be a partition of n\langle n\rangle.

  1. (i)

    Suppose that, for every eigenvalue λσ(A)\lambda\in\sigma(A) with corresponding right Jordan chain {x0,j}j=1p\left\{x_{\langle 0,j\rangle}\right\}_{j=1}^{p}, and whenever the vector x0,jx_{\left<0,j\right>} is partitioned conformably with respect to PP as

    x0,j=[x1jxhj],jp,\displaystyle x_{\left<0,j\right>}=\begin{bmatrix}x_{1j}\\ \vdots\\ x_{hj}\end{bmatrix},~{}j\in\langle p\rangle,

    the set

    {xk,j:=[(ωk)α1jx1j(ωk)αhjxhj]}j=1p\displaystyle\left\{x_{\left<k,j\right>}:=\begin{bmatrix}(\omega^{k})^{\alpha_{1j}}x_{1j}\\ \vdots\\ (\omega^{k})^{\alpha_{hj}}x_{hj}\end{bmatrix}\right\}_{j=1}^{p}

    is a right Jordan chain corresponding to λωk\lambda\omega^{k} for every kR(h)k\in R(h).

  2. (ii)

    Suppose that, for every eigenvalue λσ(A)\lambda\in\sigma(A) with corresponding left Jordan chain {yj,0}j=1p\left\{y_{\left<j,0\right>}\right\}_{j=1}^{p}, and whenever the vector yj,0y_{\left<j,0\right>} is partitioned conformably with respect to PP as

    yj,0=[yj1yjh],jp,\displaystyle y_{\left<j,0\right>}^{\top}=\begin{bmatrix}y_{j1}^{\top}&\cdots&y_{jh}^{\top}\end{bmatrix},~{}j\in\langle p\rangle,

    the set

    {yj,k:=[(ωk)αj1yj1(ωk)αjhyjh]}j=1p\displaystyle\left\{y_{\left<j,k\right>}^{\top}:=\begin{bmatrix}(\omega^{k})^{\alpha_{j1}}y_{j1}^{\top}&\cdots&(\omega^{k})^{\alpha_{jh}}y_{jh}^{\top}\end{bmatrix}\right\}_{j=1}^{p}

    is a left Jordan chain corresponding to λωk\lambda\omega^{k} for every kR(h)k\in R(h)

If the above hold, then AA is hh-cyclic with partition PP.

Proof.

By hypothesis, any Jordan canonical form of AA is of the form

S1AS=i=1m(k=0h1Jni(λiωk)),1m<n.S^{-1}AS=\bigoplus_{i=1}^{m}\left(\bigoplus_{k=0}^{h-1}J_{n_{i}}\left(\lambda_{i}\omega^{k}\right)\right),1\leq m<n.

For imi\in\langle m\rangle, let

Di:=diag(0,,0,j=0h1Jni(λi)𝑖,0,,0)D_{i}:=\operatorname{diag}\left(0,\ldots,0,\overset{i}{\overbrace{\bigoplus_{j=0}^{h-1}J_{n_{i}}(\lambda_{i})}},0,\ldots,0\right)

and Ai:=SDiS1𝖬n()A_{i}:=SD_{i}S^{-1}\in\mathsf{M}_{n}(\mathbb{C}), where diag(M1,,Mk)\operatorname{diag}(M_{1},\ldots,M_{k}) denotes the block-diagonal matrix with block-diagonal entries M1,,MkM_{1},\ldots,M_{k}.

For kR(h)k\in R(h), let CkC_{k} be defined as in (7). By Lemmas 4.12 and 4.13 and properties of the Hadamard product, notice that

Ai\displaystyle A_{i} =k=0h1(j=1piλiωkxk,jyj,k+j=1pi1xk,jyj+1,k)\displaystyle=\sum_{k=0}^{h-1}\left(\sum_{j=1}^{p_{i}}\lambda_{i}\omega^{k}x_{\left<k,j\right>}y_{\left<j,k\right>}^{\top}+\sum_{j=1}^{p_{i}-1}x_{\left<k,j\right>}y_{\left<j+1,k\right>}^{\top}\right)
=k=0h1(j=1piλiωk[(ωk)α1jx1j(ωk)αhjxhj][(ωk)αj1yj1(ωk)αjhyjh]+\displaystyle=\sum_{k=0}^{h-1}\left(\sum_{j=1}^{p_{i}}\lambda_{i}\omega^{k}\begin{bmatrix}(\omega^{k})^{\alpha_{1j}}x_{1j}\\ \vdots\\ (\omega^{k})^{\alpha_{hj}}x_{hj}\end{bmatrix}\begin{bmatrix}(\omega^{k})^{\alpha_{j1}}y_{j1}^{\top}&\cdots&(\omega^{k})^{\alpha_{jh}}y_{jh}^{\top}\end{bmatrix}\right.+
j=1pi1[(ωk)α1jx1j(ωk)αhjxhj][(ωk)αj+1,1yj+1,1(ωk)αj+1,hyj+1,h])\displaystyle\qquad\left.\sum_{j=1}^{p_{i}-1}\begin{bmatrix}(\omega^{k})^{\alpha_{1j}}x_{1j}\\ \vdots\\ (\omega^{k})^{\alpha_{hj}}x_{hj}\end{bmatrix}\begin{bmatrix}(\omega^{k})^{\alpha_{j+1,1}}y_{j+1,1}^{\top}&\cdots&(\omega^{k})^{\alpha_{j+1,h}}y_{j+1,h}^{\top}\end{bmatrix}\right)
=λik=0h1j=1piWkj1x0,jyj,0+k=0h1j=1pi1Wkj2x0,jyj+1,0\displaystyle=\lambda_{i}\sum_{k=0}^{h-1}\sum_{j=1}^{p_{i}}W_{kj}^{1}\circ x_{\langle 0,j\rangle}y_{\langle j,0\rangle}^{\top}+\sum_{k=0}^{h-1}\sum_{j=1}^{p_{i}-1}W_{kj}^{2}\circ x_{\langle 0,j\rangle}y_{\langle j+1,0\rangle}^{\top}
=λij=1pi(k=0h1Ckx0,jyj,0)+j=1pi1(k=0h1Ckx0,jyj+1,0)\displaystyle=\lambda_{i}\sum_{j=1}^{p_{i}}\left(\sum_{k=0}^{h-1}C_{k}\circ x_{\langle 0,j\rangle}y_{\langle j,0\rangle}^{\top}\right)+\sum_{j=1}^{p_{i}-1}\left(\sum_{k=0}^{h-1}C_{k}\circ x_{\langle 0,j\rangle}y_{\langle j+1,0\rangle}^{\top}\right)
=λij=1pi[(k=0h1Ck)x0,jyj,0]+j=1pi1[(k=0h1Ck)x0,jyj+1,0]\displaystyle=\lambda_{i}\sum_{j=1}^{p_{i}}\left[\left(\sum_{k=0}^{h-1}C_{k}\right)\circ x_{\langle 0,j\rangle}y_{\langle j,0\rangle}^{\top}\right]+\sum_{j=1}^{p_{i}-1}\left[\left(\sum_{k=0}^{h-1}C_{k}\right)\circ x_{\langle 0,j\rangle}y_{\langle j+1,0\rangle}^{\top}\right]
=λihj=1piKhx0,jyj,0+hj=1pi1Khx0,jyj+1,0\displaystyle=\lambda_{i}h\sum_{j=1}^{p_{i}}K_{h}\circ x_{\langle 0,j\rangle}y_{\langle j,0\rangle}^{\top}+h\sum_{j=1}^{p_{i}-1}K_{h}\circ x_{\langle 0,j\rangle}y_{\langle j+1,0\rangle}^{\top}
=λihj=1pi[x1jyj2x(h1)jyjhxhjyj1]+\displaystyle=\lambda_{i}h\sum_{j=1}^{p_{i}}\begin{bmatrix}&x_{1j}y_{j2}^{\top}&&\\ &&\ddots&\\ &&&x_{(h-1)j}y_{jh}^{\top}\\ x_{hj}y_{j1}^{\top}&&&\end{bmatrix}+
hj=1pi1[x1jyj+1,2xh1,jyj+1,hxhjyj+1,1]\displaystyle\qquad h\sum_{j=1}^{p_{i}-1}\begin{bmatrix}&x_{1j}y_{j+1,2}^{\top}&&\\ &&\ddots&\\ &&&x_{h-1,j}y_{j+1,h}^{\top}\\ x_{hj}y_{j+1,1}^{\top}&&&\end{bmatrix}
=[A12(i)Ah1,h(i)Ah1(i)],\displaystyle=\begin{bmatrix}&A_{12}^{(i)}&&\\ &&\ddots&\\ &&&A_{h-1,h}^{(i)}\\ A_{h1}^{(i)}&&&\end{bmatrix},

where

Ak,k+1(i):=λihj=1pixkjyj,k+1+hj=1pi1xkjyj+1,k+1𝖬|Vk|×|Vk+1|(),\displaystyle A_{k,k+1}^{(i)}:=\lambda_{i}h\sum_{j=1}^{p_{i}}x_{kj}y_{j,k+1}^{\top}+h\sum_{j=1}^{p_{i}-1}x_{kj}y_{j+1,k+1}^{\top}\in\mathsf{M}_{|V_{k}|\times|V_{k+1}|}(\mathbb{C}),

khk\in\langle h\rangle, and, for convenience, h+1=1h+1=1.

Clearly, the matrices A1,,AmA_{1},\dots,A_{m} are hh-cyclic with partition PP and since A=i=1mAiA=\sum_{i=1}^{m}A_{i}, it follows that AA is hh-cyclic with partition PP. ∎

Example 4.15.

Consider the matrices

J=[0100000000010000]J=\left[\begin{array}[]{*{2}{c}|*{2}{c}}0&1&0&0\\ 0&0&0&0\\ \hline\cr 0&0&0&1\\ 0&0&0&0\end{array}\right]

and

S=[xaxaybybzczcwdwd]S=\begin{bmatrix}x&a&x&-a\\ y&b&y&-b\\ z&c&-z&c\\ w&d&-w&d\end{bmatrix}

where det(S)=4(aybx)(cwdz)0\det(S)=-4(ay-bx)(cw-dz)\neq 0.

If A=SJS1A=SJS^{-1} and P={{1,2},{3,4}}P=\{\{1,2\},\{3,4\}\}, then AA satisfies the hypotheses of Theorem 4.14, so AA is bipartite. Indeed, a calculation via a computer algebra system reveals that

A=[00wxcwdzxzcwdz00wycwdzyzcwdzyzaybxxzaybx00wyaybxxwaybx00].\displaystyle A=\begin{bmatrix}0&0&\frac{wx}{cw-dz}&\frac{-xz}{cw-dz}\\ 0&0&\frac{wy}{cw-dz}&\frac{-yz}{cw-dz}\\ \frac{yz}{ay-bx}&\frac{-xz}{ay-bx}&0&0\\ \frac{wy}{ay-bx}&\frac{-xw}{ay-bx}&0&0\end{bmatrix}.

The following result yields a complete characterization of the Jordan structure for invertible hh-cyclic matrices.

Theorem 4.16.

If A𝖬n()A\in\mathsf{M}_{n}(\mathbb{C}) is nonsingular and P={V1,,Vh}P=\{V_{1},\ldots,V_{h}\} is a partition of n\langle n\rangle, then AA is hh-cyclic with partition PP if and only if the Jordan chains of AA satisfy conditions (i) and (ii) of Theorem 4.14.

Proof.

Follows by Lemma 4.10 and Thereom 4.14. ∎

5 Jordan Chains of Singular h-cyclic Matrices

Lemma 5.17.

If A𝖬n()A\in\mathsf{M}_{n}(\mathbb{C}) is hh-cyclic, then AA is singular if and only if at least one of the matrices BiB_{i} is singular.

Proof.

In view of (3), it follows that σ(Ah)=i=1hσ(Bi)\sigma(A^{h})=\bigcup_{i=1}^{h}\sigma(B_{i}). Thus, 0σ(Ah)0\in\sigma(A^{h}) if and only if one of the submatrices BiB_{i} of AhA^{h} is singular, and the latter holds if and only if 0σ(A)0\in\sigma(A). ∎

Lemma 5.18.

Let A𝖬n()A\in\mathsf{M}_{n}(\mathbb{C}) be hh-cyclic and of the form (2) with consecutive partition PP. For x|Vi|x\in\mathbb{C}^{|V_{i}|}, where ihi\in\langle h\rangle, let vnv\in\mathbb{C}^{n} be the conformably partitioned vector defined by

vk={x,k=i0,ki.\displaystyle v_{k}=\begin{cases}x,&k=i\\ 0,&k\neq i\end{cases}.

If pp\in\mathbb{N}, then

[Apv]k={Bipx,k=αp(i)0,kαp(i).\displaystyle[A^{p}v]_{k}=\begin{cases}B_{ip}x,&k=\alpha^{-p}(i)\\ 0,&k\neq\alpha^{-p}(i).\end{cases}
Proof.

Proceed by induction on pp. If p=1p=1 and i=1i=1, then αp(i)=α1(1)=h\alpha^{-p}(i)=\alpha^{-1}(1)=h. Thus, [Apv]k=[Av]k=0[A^{p}v]_{k}=[Av]_{k}=0, whenever 1k<h1\leq k<h, and [Apv]h=[Av]h=Ah1x[A^{p}v]_{h}=[Av]_{h}=A_{h1}x. Since

Bip=B11=j=hhAαj1(1),αj(1)=Aαh1(1),αh(1)=Ah1,B_{ip}=B_{11}=\prod_{j=h}^{h}A_{\alpha^{j-1}(1),\alpha^{j}(1)}=A_{\alpha^{h-1}(1),\alpha^{h}(1)}=A_{h1},

the result holds when i=p=1i=p=1.

If 1<ih1<i\leq h and p=1p=1, then αp(i)=α1(i)=i1\alpha^{-p}(i)=\alpha^{-1}(i)=i-1. Thus, [Apv]k=[Av]k=0[A^{p}v]_{k}=[Av]_{k}=0, whenever ki1k\neq i-1 and [Apv]i1=[Av]i1=Ai1,ix[A^{p}v]_{i-1}=[Av]_{i-1}=A_{i-1,i}x. Since

Bip=Bi1=j=hhAαj1(i),αj(i)=Aαh1(i),αh(i)=Ai1,i,B_{ip}=B_{i1}=\prod_{j=h}^{h}A_{\alpha^{j-1}(i),\alpha^{j}(i)}=A_{\alpha^{h-1}(i),\alpha^{h}(i)}=A_{i-1,i},

the result holds when 1<ih1<i\leq h and p=1p=1, which completes the base-case.

For the induction-step, assume that

[Apv]k={Bipx,k=αp(i)0,kαp(i),\displaystyle[A^{p}v]_{k}=\begin{cases}B_{ip}x,&k=\alpha^{-p}(i)\\ 0,&k\neq\alpha^{-p}(i),\end{cases}

for some positive integer pp.

If αp(i)=1\alpha^{-p}(i)=1, then α(p+1)(i)=h\alpha^{-(p+1)}(i)=h and

[Ap+1v]α(p+1)(i)\displaystyle[A^{p+1}v]_{\alpha^{-(p+1)}(i)} =[A(Apv)]h\displaystyle=[A(A^{p}v)]_{h}
=Ah1Bipx\displaystyle=A_{h1}B_{ip}x
=Ah1(j=h+1phAαj1(i),αj(i)x)\displaystyle=A_{h1}\left(\prod_{j=h+1-p}^{h}A_{\alpha^{j-1}(i),\alpha^{j}(i)}x\right)
=Aα(p+1)(i),αp(i)(j=h+1phAαj1(i),αj(i)x)\displaystyle=A_{\alpha^{-(p+1)}(i),\alpha^{-p}(i)}\left(\prod_{j=h+1-p}^{h}A_{\alpha^{j-1}(i),\alpha^{j}(i)}x\right)
=Aαh(p+1)(i),αhp(i)(j=h+1phAαj1(i),αj(i)x)\displaystyle=A_{\alpha^{h-(p+1)}(i),\alpha^{h-p}(i)}\left(\prod_{j=h+1-p}^{h}A_{\alpha^{j-1}(i),\alpha^{j}(i)}x\right) (αh=ε\alpha^{h}=\varepsilon)
=(j=hphAαj1(i),αj(i)x)\displaystyle=\left(\prod_{j=h-p}^{h}A_{\alpha^{j-1}(i),\alpha^{j}(i)}x\right)
=Bi,p+1x,\displaystyle=B_{i,p+1}x,

as desired.

If 1<αp(i)h1<\alpha^{-p}(i)\leq h, then α(p+1)(i)=i1\alpha^{-(p+1)}(i)=i-1 and

[Ap+1v]α(p+1)(i)\displaystyle[A^{p+1}v]_{\alpha^{-(p+1)}(i)} =[A(Apv)]i1\displaystyle=[A(A^{p}v)]_{i-1}
=Ai1,iBipx\displaystyle=A_{i-1,i}B_{ip}x
=Ai1,i(j=h+1phAαj1(i),αj(i)x)\displaystyle=A_{i-1,i}\left(\prod_{j=h+1-p}^{h}A_{\alpha^{j-1}(i),\alpha^{j}(i)}x\right)
=Aα(p+1)(i),αp(i)(j=h+1phAαj1(i),αj(i)x)\displaystyle=A_{\alpha^{-(p+1)}(i),\alpha^{-p}(i)}\left(\prod_{j=h+1-p}^{h}A_{\alpha^{j-1}(i),\alpha^{j}(i)}x\right)
=Aαh(p+1)(i),αhp(i)(j=h+1phAαj1(i),αj(i)x)\displaystyle=A_{\alpha^{h-(p+1)}(i),\alpha^{h-p}(i)}\left(\prod_{j=h+1-p}^{h}A_{\alpha^{j-1}(i),\alpha^{j}(i)}x\right) (αh=ε\alpha^{h}=\varepsilon)
=(j=hphAαj1(i),αj(i)x)\displaystyle=\left(\prod_{j=h-p}^{h}A_{\alpha^{j-1}(i),\alpha^{j}(i)}x\right)
=Bi,p+1x,\displaystyle=B_{i,p+1}x,

which completes the induction-step. The entire result now follows by the principle of mathematical induction. ∎

Theorem 5.19.

Let A𝖬n()A\in\mathsf{M}_{n}(\mathbb{C}) be hh-cyclic with consecutive partition PP. Suppose that BiB_{i} is singular. If xNull(Bi),x0x\in\operatorname{Null}(B_{i}),\ x\neq 0, then there is a positive integer php\leq h such that Bipx=0B_{ip}x=0 and Biqx0B_{iq}x\neq 0 for all 1q<p1\leq q<p. Furthermore, any Jordan canonical form of AA has a p×pp\times p singular Jordan block.

Proof.

Let vnv\in\mathbb{C}^{n} be the conformably partitioned vector defined by

vk={x,k=i0,ki.\displaystyle v_{k}=\begin{cases}x,&k=i\\ 0,&k\neq i\end{cases}.

By Lemma 5.18 and the assumption that xNull(Bi)x\in\operatorname{Null}(B_{i}), we have

[Ahv]k={Bihx,k=αh(i)0,kαh(i)={Bix,k=i0,ki,=0.\displaystyle[A^{h}v]_{k}=\begin{cases}B_{ih}x,&k=\alpha^{-h}(i)\\ 0,&k\neq\alpha^{-h}(i)\end{cases}=\begin{cases}B_{i}x,&k=i\\ 0,&k\neq i,\end{cases}=0.

Thus, the set

{k(A0In)kv=Akv=0}\begin{Bmatrix}k\in\mathbb{N}\mid(A-0I_{n})^{k}v=A^{k}v=0\end{Bmatrix}\neq\emptyset

and contains a minimal element—say pp—such that php\leq h and Aqv0A^{q}v\neq 0 whenever 1q<p1\leq q<p. Hence, in view of Lemma 5.18, Bipx=0B_{ip}x=0 and Biqx0B_{iq}x\neq 0.

If wk:=Apkvw_{k}:=A^{p-k}v, then {w1,,wp}\{w_{1},\dots,w_{p}\} is a Jordan chain corresponding to the eigenvalue zero. Hence, any Jordan canonical form of AA has a p×pp\times p singular Jordan block. ∎

Remark 5.20.

For ihi\in\langle h\rangle, let i={b1(i),,bdi(i)}\mathcal{B}_{i}=\left\{b_{1}^{(i)},\ldots,b_{d_{i}}^{(i)}\right\} be a basis for Null(Bi)\operatorname{Null}(B_{i}). In Theorem 5.19, for xNull(Bi)x\in\operatorname{Null}(B_{i}), if pxp_{x} denotes the length of the Jordan chain corresponding to xx, then

px=max{p1(i),,pdi(i)},p_{x}=\max\left\{p_{1}^{(i)},\ldots,p_{d_{i}}^{(i)}\right\},

where pk(i)p_{k}^{(i)} is the length of the Jordan chain corresponding to bk(i)b_{k}^{(i)}, k{1,,di}k\in\{1,\ldots,d_{i}\}. Thus, it suffices to only consider elements of i\mathcal{B}_{i} to determine the Jordan chains of AA corresponding to the eigenvalue zero.

Example 5.21.

Consider the 3-cyclic matrix

A=[0A12000A23A3100]A=\begin{bmatrix}0&A_{12}&0\\ 0&0&A_{23}\\ A_{31}&0&0\end{bmatrix}

where A12A_{12} is the all-ones matrix of order four, A31=I4A_{31}=I_{4}, and

A23=[1000100010000001].A_{23}=\begin{bmatrix}1&0&0&0\\ 1&0&0&0\\ 1&0&0&0\\ 0&0&0&1\\ \end{bmatrix}.

The jordan command in MATLAB reveals that a Jordan canonical form of AA is

[01000000000000100000000000000000000000001000000000000000000000000010000000000000000000000000000000000000000000000000022/300000000000022/3ω00000000000022/3ω2].\left[\begin{array}[]{*{3}{c}|*{2}{c}|*{2}{c}|*{1}{c}|*{1}{c}|*{3}{c}}0&1&0&0&0&0&0&0&0&0&0&0\\ 0&0&1&0&0&0&0&0&0&0&0&0\\ 0&0&0&0&0&0&0&0&0&0&0&0\\ \hline\cr 0&0&0&0&1&0&0&0&0&0&0&0\\ 0&0&0&0&0&0&0&0&0&0&0&0\\ \hline\cr 0&0&0&0&0&0&1&0&0&0&0&0\\ 0&0&0&0&0&0&0&0&0&0&0&0\\ \hline\cr 0&0&0&0&0&0&0&0&0&0&0&0\\ \hline\cr 0&0&0&0&0&0&0&0&0&0&0&0\\ \hline\cr 0&0&0&0&0&0&0&0&0&2^{2/3}&0&0\\ \hline\cr 0&0&0&0&0&0&0&0&0&0&2^{2/3}\omega&0\\ \hline\cr 0&0&0&0&0&0&0&0&0&0&0&2^{2/3}\omega^{2}\end{array}\right].

A straightforward calculation reveals that

B1=B3=[3001300130013001]B_{1}=B_{3}=\begin{bmatrix}3&0&0&1\\ 3&0&0&1\\ 3&0&0&1\\ 3&0&0&1\end{bmatrix}

and

B2=[1111111111111111].B_{2}=\begin{bmatrix}1&1&1&1\\ 1&1&1&1\\ 1&1&1&1\\ 1&1&1&1\end{bmatrix}.

If

x=[1/3001],x=\begin{bmatrix}-1/3&0&0&1\end{bmatrix}^{\top},
y=[0100],y=\begin{bmatrix}0&1&0&0\end{bmatrix}^{\top},

and

z=[0010],z=\begin{bmatrix}0&0&1&0\end{bmatrix}^{\top},

then 1={x,y,z}\mathcal{B}_{1}=\{x,y,z\} is a basis for Null(B1)\operatorname{Null}(B_{1}). It can then be shown that Avx0,A2vx0Av_{x}\neq 0,A^{2}v_{x}\neq 0, and A3vx=0A^{3}v_{x}=0 while A2vy=0A^{2}v_{y}=0 and A2vz=0A^{2}v_{z}=0 with Avy0Av_{y}\neq 0 and Avz0Av_{z}\neq 0. As a consequence of Theorem 5.19, the matrix B1B_{1} contributes singular Jordan blocks of size three and two to the Jordan canonical form of AA.

On the other hand, the matrices B1B_{1} and B3B_{3} share the same null space but the vectors vx,vyv_{x},v_{y} and vzv_{z} corresponding to B3B_{3} differ from those of B1B_{1} in the positions of their non-zero entries and we find that B3B_{3} contributes Jordan chains of length one and two. The basis of Null(B2)\operatorname{Null}(B_{2}) produces three chains of length one.

Thus, there is potential for redundancy in producing Jordan chains via Theorem 5.19. However, calculating the Weyr characteristic of AA with respect to zero (see, e.g., Horn and Johnson [3, Lemma 3.1.18]) in conjunction with Theorem 5.19 yields an accurate procedure.

6 Acknowledgements

The second author thanks Daniel Szyld for his talk at the 2019 Meeting of the International Linear Algebra Society which provided the impetus for this work and Michael Tsatsomeros for pointing out that Lemma 4.1 of McDonald and Paparella [4] generalized the result for bipartite graphs.

References

  • [1] R. A. Brualdi and H. J. Ryser. Combinatorial matrix theory, volume 39 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge, 1991.
  • [2] P. J. Davis. Circulant matrices. A Wiley-Interscience Publication. John Wiley & Sons, New York-Chichester-Brisbane, 1979.
  • [3] R. A. Horn and C. R. Johnson. Matrix analysis. Cambridge University Press, Cambridge, second edition, 2013.
  • [4] J. J. McDonald and P. Paparella. Jordan chains of hh-cyclic matrices. Linear Algebra Appl., 498:145–159, 2016.
  • [5] L. Mirsky. An inequality for characteristic roots and singular values of complex matrices. Monatsh. Math., 70:357–359, 1966.
  • [6] V. Nikiforov. Hypergraphs and hypermatrices with symmetric spectrum. Linear Algebra Appl., 519:1–18, 2017.