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

Solutions to the matrix Yang-Baxter equation

Mukherjee, Himadri BITS Pilani K. K. Birla Goa Campus, Goa, India [email protected]  and  Ali M, Askar BITS Pilani K. K. Birla Goa Campus, Goa, India [email protected]
Abstract.

In this article, we give a few classes of solutions for the Yang-Baxter type matrix equation, AXA=XAXAXA=XAX. We provide all solutions for the cases when AA is equivalent to a Jordan block or has precisely two Jordan blocks. We also have given a few general properties of the solutions of the YB-equation.

Key words and phrases:
Yang-Baxter equation, Matrix equation
AMS Classification 2020. Primary: 15A24, 16T25

1. Introduction

The Yang-Baxter equation occurs naturally in the context of statistical mechanics. It is also a natural equation in the context of braid groups, where it is a simple identity of two different compositions of braid patterns. The equation in the matrix form has been studied by many authors recently. We will call the following equation the matrix YB-equation, where AMn(K)A\in M_{n}(K).

(1) AXA=XAXAXA=XAX

The Yang-Baxter equations have been studied across the fields of mathematics with numerous fascinating interconnections. From topology to physics to linear algebra and the theory of quantum groups, this equation has generated a feverish rush of quality research among mathematicians and physicists alike. Solving the equations in the setting of linear algebra poses an interesting problem that has a number of applications in other fields. An understanding of the solution in the linear algebra setting will definitely give important insight into the solution of YBE in general.

A number of interesting works have come up since the question posed by Drinfeld in [7], namely the works in the YBE in set-theoretic set up [1] exploring the equation and solutions in the category of SetsSets.

The complexity of having n2n^{2} non-linear equation in n2n^{2} variable to solve, makes it challenging to find complete solutions for arbitrary coefficient matrix. This compelled many researchers to find solutions for particular A, especially having algebraic properties like diagonalizability and commutativity with the solution matrix. Moreover, the non-linearity of the questions paves ways for geometric (or commutative algebraic) techniques such as manifold theory ( Grobner basis) to be applied to get further insight into the space of solutions. We will summarise a few results that deal with YBE. In [9],the authors discuss the solution to YBE when A is an idempotent matrix, by means of diagonalizing A for two distinct eigenvalues. Also, the authors explain the technique to find solutions when A=A2-A=A^{2}. The article [2] contains solutions when A is a diagonalizable matrix with spectrum contained in {1,α,0}\{1,\alpha,0\}, and extended these results for the case of idempotent and rank-one matrices.
The authors in [10], found all commuting solutions for A in Jordan canonical form, and corresponding to each Jordan block, authors described solutions in Toeplitz forms. In [3], authors found complete solutions when A is an elementary matrix of the form A=IuvTA=I-uv^{T}, where vTu0v^{T}u\neq 0, which makes A diagonalizable. In a similar passion, authors in [13] found solutions when A=IPQTA=I–PQ^{T}, where P and Q are two n×2n\times 2 complex matrices of full column rank such that det(QTP)=0det(Q^{T}P)=0. In [18], authors have found solutions for A having eigenvalues 0,λ,μ0,\lambda,\mu and A diagonalizable by Jordan decomposition of A.
Irrespective of its simple appearance, when the coefficient matrix is nilpotent poses additional challenges to solving it. In [6], authors have found all commuting solutions for A being a nilpotent matrix. Later, one of the same authors, along with others, gave an equivalent system for finding commuting solutions to the YBE, when A is nilpotent and has an index 3[16].
In [12] author finds complete solutions for YBE when A is rank 1 matrix using the spectral decomposition A=pqTA=pq^{T}, where p and q are non-zero complex vectors. In [15] the authors have studied the solution when A=PQTA=PQ^{T}, where P and Q are n×2n\times 2 vectors, with an assumption QPTQP^{T} non-singular. In an extension study [14] the same authors have studied the solutions for QPTQP^{T} being singular with the algebraic multiplicity of eigenvalue 0 being more than n2n-2. In [17] authors have studied non-commuting solutions when A satisfies A4=AA^{4}=A, by exploring various possibilities for its minimal polynomials.
In [5], authors have given a new horizon for the discussion of the solution to the YBE, by tracing the path-connected subsets of the solution. Also, the authors explain techniques to generate infinitely many solutions from existing solutions using generalized inverse.

In [8], author has given manifold structure for the solution set of matrix YBE when the coefficient matrix is having rank 1. In [4], authors have found all commuting solutions of YBE when the coefficient matrix is non-singular.

In our paper, the results are organized in the manner that there are global results, i.e the results without any assumption on the coefficient matrix or the type of the solutions and special results which come by with assumptions on the coefficient matrix and/or the type of solutions. Before every result, we have mentioned whether it is a global or a local result to facilitate the reader. There are a few minor overlaps in the results, especially with the newest articles, but each time we have presented a new way of looking at it or have extended the result. Which is why we have not removed those results from our paper but have mentioned whenever such overlaps occurred.

2. Main Results

For the coefficient matrix AA, having a single Jordan block with non-zero eigenvalue, we have the following theorem completely characterizing the solution of the Yang-Baxter equation, AXA=XAXAXA=XAX.

Theorem 2.1.

If A=λI+BA=\lambda I+B is a Jordan block, then for the YBE, AXA=XAXAXA=XAX, the following are true.

  1. (1)

    If λ0\lambda\neq 0 and det(X)0det(X)\neq 0 then XAX\simeq A.

  2. (2)

    If λ0\lambda\neq 0 and det(X)=0det(X)=0 then X=0X=0.

  3. (3)

    If λ=0\lambda=0, then the YBE do not have any invertible solution X.

For the case where AA has two Jordan blocks of the same size, and the determinant of AA is not zero, we have the following theorem completely characterizing the solutions of the YBE, AXA=XAXAXA=XAX. Let SolASol_{A} denote the set of all solutions of the YBE.

Theorem 2.2.

Let 𝒜\mathcal{A} be a matrix having only two eigenvalues λ1,λ20\lambda_{1},\lambda_{2}\neq 0, with the same algebraic multiplicity, and geometric multiplicity 1. Further, let us assume 0 is an eigenvalue of XX. Let UGLn(K)U\in GL_{n}(K) be defined as, the columns of UU are generalized eigenvectors of 𝒜\mathcal{A}, such that U𝒜U1U\mathcal{A}U^{-1} is Jordan-canonical form of 𝒜\mathcal{A}. Then one of the following is true.

  1. (1)

    X=0X=0

  2. (2)

    X=U1(0ZS1A0Y2)UX=U^{-1}\begin{pmatrix}0&ZS^{-1}A\\ 0&Y_{2}\end{pmatrix}U or U1(000Y2)UU^{-1}\begin{pmatrix}0&0\\ 0&Y_{2}\end{pmatrix}U, where Y2SolAY_{2}\in Sol_{A}, ZK[A]Z\in K[A], AA is the Jordan block corresponding to eigenvalue λ=λ1=λ2\lambda=\lambda_{1}=\lambda_{2}, and for some SGLn(K)S\in GL_{n}(K)

  3. (3)

    X=U1(Y10ZS1A0)UX=U^{-1}\begin{pmatrix}Y_{1}&0\\ ZS^{-1}A&0\end{pmatrix}U or U1(Y1000)UU^{-1}\begin{pmatrix}Y_{1}&0\\ 0&0\end{pmatrix}U, where Y1SolAY_{1}\in Sol_{A}, ZK[A]Z\in K[A], AA is the Jordan block corresponding to eigenvalue λ=λ1=λ2\lambda=\lambda_{1}=\lambda_{2}, and for some SGLn(K)S\in GL_{n}(K)

3. preliminaries

Let us recall the following notations, which will be used throughout the discussion.

  • Let KK be a field. For a coefficient matrix AMn(K)A\in M_{n}(K), the set of all solutions to the Yang-Baxter equation AXA=XAXAXA=XAX is denoted by SolA:={XMn(K)|AXA=XAX}Sol_{A}:=\{X\in M_{n}(K)|AXA=XAX\}.

  • For a matrix MMn(K)M\in M_{n}(K), σ(M)\sigma(M) denotes the spectrum of MM.

  • Let VV be a vector space, then we say AMn(K)A\in M_{n}(K) preserves VV, if A(V)VA(V)\subseteq V.

  • For a MMn(K)M\in M_{n}(K), minMmin_{M} represent the minimal polynomial of MM.

  • K[A]K[A] represent the polynomial ring on AMn(K)A\in M_{n}(K), over KK.

  • Throughout the discussion, BB represent the nilpotent block (01000010...00010000),\begin{pmatrix}0&1&0&\dots&0\\ 0&0&1&\dots&0\\ .\\ .\\ .\\ 0&0&0&\dots&1\\ 0&0&0&\dots&0\end{pmatrix}, with appropriate dimension.

Theorem 3.1 ([11]).

Given matrices An×nA\in\mathbb{C}^{n\times n} and Bm×mB\in\mathbb{C}^{m\times m}, the Sylvester equation AX+XB=CAX+XB=C has a unique solution Xn×mX\in\mathbb{C}^{n\times m} for any Cn×mC\in\mathbb{C}^{n\times m}, if and only if A and -B do not share any eigenvalue.

Lemma 3.2.

For any gGLn(K)g\in GL_{n}(K), gSolAg1=SolgAg1gSol_{A}g^{-1}=Sol_{gAg^{-1}}.

Proof.

Let B=gAg1B=gAg^{-1}, and xgSolAg1x\in gSol_{A}g^{-1}. Then x=gXg1x=gXg^{-1}, for some XSolAX\in Sol_{A}. Now for y=BxB=gAg1xgAg1y=BxB=gAg^{-1}xgAg^{-1}, that is g1yg=A(g1xg)A=AXA=XAX=g1xgAg1xg=g1xBxgg^{-1}yg=A(g^{-1}xg)A=AXA=XAX=g^{-1}xgAg^{-1}xg=g^{-1}xBxg that is, y=xBxy=xBx. which implies BxB=xBxBxB=xBx

As an immediate consequence of the above Lemma 3, we have the following.

Corollary 3.2.1.

Without loss of generality, we can assume A to be in the Jordan-canonical form.

Corollary 3.2.2.

Let XSolAX\in Sol_{A} , then for any gGA={gGLn(K)/gAg1=A}g\in G_{A}=\{g\in GL_{n}(K)/gAg^{-1}=A\}, gXg1gXg^{-1} is a solution to the YBE, AXA=XAXAXA=XAX.

Proof.

Let XSolAX\in Sol_{A}. For a gGAg\in G_{A}, we have gSolAg1=SolgAg1=SolAgSol_{A}g^{-1}\ =\ Sol_{gAg^{-1}}=Sol_{A}. ∎

Lemma 3.3.

Le XSolAX\in Sol_{A} for the YBE with coefficient matrix AA, and λ\lambda be an eigenvalue of AA with eigenvector vv, then either λ\lambda is an eigenvalue of XX or AXv=0AXv=0.

Proof.

We have, Av=λvAv=\lambda v, thus XAv=λXvXAv=\lambda Xv or AXAv=λAXvAXAv=\lambda AXv. By utilizing the YBE, we get XAXv=λAXvXAXv=\lambda AXv. Thus, if AXv0AXv\neq 0, then λ\lambda is an eigenvalue of XX. Further, by the symmetry of the YBE, we can see, if λ\lambda is an eigenvalue of XX with eigenvector vv, then either λ\lambda is an eigenvalue of AA or XAv=0XAv=0. ∎

Corollary 3.3.1.

Let coefficient matrix A be invertible and, XSolAX\in Sol_{A}, then σ(X)σ(A){0}\sigma(X)\subseteq\sigma(A)\cup\{0\}.

Proof.

From the above Lemma 3.3, we know that if λ\lambda is an eigenvalue of XX, then either it is an eigenvalue of AA or XAv=0XAv=0, for vv, an eigenvector of XX, with the eigenvalue λ\lambda. Now, since AA is invertible, we get Av0Av\neq 0. Thus, XAv=0XAv=0, implies AvAv is an eigenvector of XX with eigenvalue 0. ∎

Lemma 3.4.

If det(A)0det(A)\neq 0, then AA preserves the kernel of XX.

Proof.

Let vKer(X)v\in Ker(X), we have Xv=0AXv=0XAXv=0AXAv=0XAv=0Xv=0\iff AXv=0\implies XAXv=0\iff AXAv=0\iff XAv=0 AvKer(X)\Rightarrow Av\in Ker(X). This gives, A(Ker(X))Ker(X)A(Ker(X))\subseteq Ker(X). Since AA is invertible, we get dim(A(Ker(X)))=dim(Ker(X))dim(A(Ker(X)))=dim(Ker(X)), gives, A(Ker(X))=Ker(X)A(Ker(X))=Ker(X). ∎

Lemma 3.5.

Let the coefficient matrix AA be invertible with λσ(A)\lambda\in\sigma(A) and EλE_{\lambda}, the eigenspace of AA corresponding to λ\lambda, be of one dimension. If Ker(X)=EλKer(X)=E_{\lambda}, then λσ(X)\lambda\notin\sigma(X).

Proof.

If possible, let us assume λσ(X)\lambda\in\sigma(X). Let ww be an eigenvector for λ\lambda, thus we have Xw=λwXw=\lambda w. Now since λσ(A)\lambda\in\sigma(A), let us choose an eigenvector for it, say vv, we have Av=λvAv=\lambda v. Since we have Eλ=Ker(X)E_{\lambda}=Ker(X) we obtain Xv=0Xv=0. We also have XAXw=λXAwXAXw=\lambda XAw, or AXAw=λXAwAXAw=\lambda XAw, using the YBE. Which gives us XAwEλXAw\in E_{\lambda}, now since the eigenspace is one dimensional (generated by vv as a basis), we have XAw=μvXAw=\mu v for some μ\mu (could also be zero). This gives us X2AXw=0X^{2}AXw=0 or X(XAX)w=0X(XAX)w=0. Or XAX(Aw)=0XAX(Aw)=0, which gives us AXA2w=0AXA^{2}w=0, thus XA2w=0XA^{2}w=0 (as AA is invertible). So we have A2wker(X)A^{2}w\in ker(X). Now since ker(X)=Eλker(X)=E_{\lambda} we have A2w=lvA^{2}w=lv for some lKl\in K. So we have:

(2) A2w\displaystyle A^{2}w =(lv=l/λ)Av\displaystyle=(lv=l/\lambda)Av
Aw\displaystyle Aw =(l/λ)v\displaystyle=(l/\lambda)v
w\displaystyle w =(l/λ)A1v\displaystyle=(l/\lambda)A^{-1}v
w\displaystyle w =(l/λ2)v\displaystyle=(l/\lambda^{2})v

Thus wEλw\in E_{\lambda} but then Xw=0=λwXw=0=\lambda w giving us w=0w=0 a contradiction.

Corollary 3.5.1.

Let the coefficient matrix AA be invertible, and if λσ(A),dim(Eλ)=1\ \forall\ \lambda\in\sigma(A),\ dim(E_{\lambda})\ =1, and X(Eλ)0X(E_{\lambda})\neq 0, then σ(A)σ(X)\sigma(A)\subseteq\sigma(X).

Proof.

By the Lemma 3.3, if X(Eλ)0X(E_{\lambda})\neq 0, then λσ(X)\lambda\in\sigma(X). If this is true for all λσ(A)\lambda\in\sigma(A), then we have σ(A)σ(X)\sigma(A)\subseteq\sigma(X). ∎

Lemma 3.6.

Let A be an invertible coefficient matrix, λσ(A)\lambda\in\sigma(A) with geometric multiplicity 1 and λσ(X)\lambda\notin\sigma(X). If PλP_{\lambda} is the generalized eigenspace of AA corresponding to λ\lambda , then X(Pλ)=0X(P_{\lambda})=0.

Proof.

Since λσ(X)\lambda\notin\sigma(X), we know that XX annihilates the eigenspace of λ\lambda so EλKer(X)E_{\lambda}\subset Ker(X). Let {v1,,vr}\{v_{1},\dots,v_{r}\} be the canonical basis for the generalised eigen space PλP_{\lambda}, such that Av1=λv1Av_{1}=\lambda v_{1}, and Avi=vi1+λviAv_{i}=v_{i-1}+\lambda v_{i}, for i=2,,ri=2,\dots,r. Then we have, Xv1=0Xv_{1}=0. Now, XAv2=Xv1+λXv2XAv_{2}=Xv_{1}+\lambda Xv_{2}, or AXAv2=λAXv2AXAv_{2}=\lambda AXv_{2}. Then by YBE, XAXv2=λAXv2XAXv_{2}=\lambda AXv_{2}. This gives, if AXv20AXv_{2}\neq 0, then λσ(X)\lambda\in\sigma(X), which is not true. Hence Xv2=0Xv_{2}=0. Continuing this process, by induction we get Xvi=0Xv_{i}=0, for i=1,2,,ri=1,2,\dots,r. Which implies XPλ=0XP_{\lambda}=0. ∎

Corollary 3.6.1.

Let AA be an invertible coefficient matrix for the YBE, and for a λσ(A)\lambda\in\sigma(A), the Jordan block corresponding to λ\lambda, JλJ_{\lambda} has size strictly greater than 1 and has geometric multiplicity 1, then for any XSolAX\in Sol_{A}, such that λσ(X)\lambda\notin\sigma(X), Ker(X)EλKer(X)\neq E_{\lambda}.

Proof.

If, for a λσ(A)\lambda\in\sigma(A), λσ(X)\lambda\notin\sigma(X), with dim(Eλ)=1dim(E_{\lambda})=1, then, by Lemma 3.6, we have XX annihilate any cyclic subspace of generalised eigenspace of AA, corresponding to λ\lambda. Let PλP_{\lambda} be the maximum cyclic subspace of such generalised eigenspace. Then, PλKer(X)P_{\lambda}\subseteq Ker(X). Since, dim(Pλ)dim(P_{\lambda}) is same as Jordan block corresponding to λ\lambda, which is strictly greater than one, and EλE_{\lambda} has dimension one, we have Ker(X)EλKer(X)\neq E_{\lambda}.

Lemma 3.7.

Let A be an invertible coefficient matrix, and XSolAX\in Sol_{A} is also invertible, then AXA\simeq X. In particular, they have the same Jordan form.

Proof.

The result follows from the fact,

AXA=XAXX1AX=AXA1AXA=XAX\iff X^{-1}AX=AXA^{-1}

Lemma 3.8.

If the coefficient matrix AA is a Jordan block, λI+B\lambda I+B, such that λ0\lambda\neq 0, and if a solution of the YB equation XX is not the zero matrix, then σ(X)={λ}\sigma(X)=\{\lambda\}.

Proof.

We have σ(X)σ(A){0}\sigma(X)\subseteq\sigma(A)\cup\{0\}, and σ(A)={λ}\sigma(A)=\{\lambda\}. Also, we have A(Ker(X))Ker(X)A(Ker(X))\subset Ker(X), by Corollary 3.3.1. So, Ker(X)Ker(X) is an invariant subspace of AA. i.e., Ker(X)=Eλ or KnKer(X)=E_{\lambda}\mbox{ or }K^{n}, as λ\lambda is the only eigenvalue of AA, Pλ=KnP_{\lambda}=K^{n}. Now, if Ker(X)=KnKer(X)=K^{n}, then X=0X=0. If Ker(X)=EλKer(X)=E_{\lambda}, then by Lemma 3.5, we have λσ(X)\lambda\notin\sigma(X); which gives σ(X){0}\sigma(X)\subset\{0\}. i.e., X=0X=0. As a result, if X0X\neq 0, then σ(X)={λ}\sigma(X)=\{\lambda\}. ∎

Lemma 3.9.

AXA=XAXAXA=XAX if and only if AXAn=XnAXAXA^{n}=X^{n}AX for all nn\in\mathbb{N}. Similarly, AXA=XAXAXA=XAX if and only if AnXA=XAXn,nA^{n}XA=XAX^{n},\ \forall\ n\in\mathbb{N}.

Proof.
AXA=XAXAXA=XAX
AXA2=XAXA=X2AX.\Rightarrow AXA^{2}=XAXA=X^{2}AX.

Then by induction, we have, n\forall\ n\in\mathbb{N}, AXAn=XnAXAXA^{n}=X^{n}AX. Conversely, if AXAn=XnAXAXA^{n}=X^{n}AX, is true for any nn\in\mathbb{N}, then in particular, it is true for n=1n=1. i.e., AXA=XAXAXA=XAX. Similar way, we can prove, AXA=XAXAnXA=XAXn,nAXA=XAX\ \iff\ A^{n}XA=XAX^{n},\ \forall\ n\in\mathbb{N}

Lemma 3.10.

If XAX=AXAXAX=AXA, and XX and A commute, then (Iet(AX))XA=0(I-e^{t(A-X)})XA=0, for any tt\in\mathbb{R}.

Proof.

If XAX=AXAXAX=AXA, then as a consequence of above lemma, we have XAetX=etAXAXAe^{tX}=e^{tA}XA for any tt\in\mathbb{R}.
Which gives

XAetXX\displaystyle XAe^{tX}X =etAXAX\displaystyle=e^{tA}XAX
XAXetX\displaystyle XAXe^{tX} =etAXAX\displaystyle=e^{tA}XAX
AXAetX\displaystyle AXAe^{tX} =etAAXA\displaystyle=e^{tA}AXA
=AetAXA\displaystyle=Ae^{tA}XA

This gives XA=et(AX)XAXA=e^{t(A-X)}XA, implies (Iet(AX))XA=0(I-e^{t(A-X)})XA=0. ∎

Lemma 3.11.

Let XSolAX\in Sol_{A}, be a commuting solution, such that AXA-X is invertible, then AX=0AX=0.

Proof.

If XX is a commuting solution in SolASol_{A}, then we have, A2X=AX2A^{2}X=AX^{2}. By taking Y=AXY=AX, we have, AYYX=0AY-YX=0. By Sylvester’s Theorem 3.1, Y=0Y=0 if and only if, σ(A)σ(X)=\sigma(A)\cap\sigma(X)=\varnothing, i.e., AXA-X is invertible. ∎

Theorem 3.12.

Let ϕA(x)\phi_{A}(x) be the characteristic polynomial of A (respectively ϕX(x)\phi_{X}(x) be the characteristic polynomial of X ). Then the following is true,

  • i

    XAϕA(X)=0XA\phi_{A}(X)=0

  • ii

    ϕA(X)AX=0\phi_{A}(X)AX=0

Proof.

Let

ϕA(x)=a0+a1x++anxn.\phi_{A}(x)=a_{0}+a_{1}x+\dots+a_{n}x^{n}.

Then we have,

XAϕA(X)\displaystyle XA\phi_{A}(X) =a0XA+a1XAX+a2XAX2++anXAXn\displaystyle=a_{0}XA+a_{1}XAX+a_{2}XAX^{2}+\dots+a_{n}XAX^{n}
=a0XA+a1AXA+a2A2XA++anAnXA\displaystyle=a_{0}XA+a_{1}AXA+a_{2}A^{2}XA+\dots+a_{n}A^{n}XA
=(a0I+a1A+a2A2++anAn)XA\displaystyle=(a_{0}I+a_{1}A+a_{2}A^{2}+\dots+a_{n}A^{n})XA
=ϕA(A)XA=0\displaystyle=\phi_{A}(A)XA=0

Similarly, we have ϕA(X)AX=0\phi_{A}(X)AX=0. ∎

Corollary 3.12.1.

If An=0A^{n}=0, then XAXm=XmAX=0XAX^{m}=X^{m}AX=0, for every mnm\geq n.

Lemma 3.13.

Let A,XA,X be two square matrices of the same dimension. ϕA(X)\phi_{A}(X) is invertible if and only if σ(A)σ(X)=\sigma(A)\cap\sigma(X)=\varnothing.

Proof.

Let ϕA(X)=(Xλ1I)(XλnI)\phi_{A}(X)=(X-\lambda_{1}I)\dots(X-\lambda_{n}I). Then ϕA(X)\phi_{A}(X) is invertible if and only if for any ii, λiσ(X)\lambda_{i}\notin\sigma(X). ∎

Theorem 3.14.

Let AXA=XAXAXA=XAX, if σ(A)σ(X)=\sigma(A)\cap\sigma(X)=\varnothing, then either

  • i

    A=0A=0 and XX is invertible
    or

  • ii

    AA is invertible and X=0X=0.

Proof.

As σ(A)σ(X)=\sigma(A)\cap\sigma(X)=\varnothing, only atmost one of σ(A),σ(X)\sigma(A),\sigma(X) can contain 0. If none of them contains 0, then both A,XA,X are invertible and hence XAX\simeq A, which implies σ(A)=σ(X)\sigma(A)=\sigma(X), which is not possible. Therefore, one must contain 0. If 0σ(A)0\in\sigma(A), then 0σ(X)0\notin\sigma(X) and X is invertible. Also, since σ(A)σ(X)=\sigma(A)\cap\sigma(X)=\varnothing, we have minA,minXmin_{A},min_{X} are co-prime. As a result, ϕA(X)\phi_{A}(X) is invertible. Then,

XAϕA(X)=ϕA(X)AX=0XA=0.XA\phi_{A}(X)=\phi_{A}(X)AX=0\\ \Longrightarrow XA=0.

Since X is invertible, A must be 0. Similarly, if 0σ(X)0\in\sigma(X), then X=0X=0 and A is invertible. ∎

Theorem 3.15.

If A=λI+BA=\lambda I+B is a Jordan block, then for the YBE, AXA=XAXAXA=XAX, the following are true.

  1. (1)

    If λ0\lambda\neq 0 and det(X)0det(X)\neq 0 then XAX\simeq A.

  2. (2)

    If λ0\lambda\neq 0 and det(X)=0det(X)=0 then X=0X=0.

  3. (3)

    If λ=0\lambda=0, then the YBE do not have any invertible solution X.

Proof.

The first case follows from Lemma 3.7. Now, by Lemma 3.8, if AA is invertible and X0X\neq 0, the σ(X)={λ}\sigma(X)=\{\lambda\}, i.e XX is invertible. Thus, if XX is singular, XX must be 0 matrix. Let, A=BA=B, i.e., λ=0\lambda=0, and assume if possible, XX is invertible. Then, σ(A)σ(X)=\sigma(A)\cap\sigma(X)=\varnothing. Thus, by Theorem 3.14, AA must be 0, which is not true. Hence, XX must be singular in this case. ∎

Corollary 3.15.1.

Let the coefficient matrix A=(A1000A20...00Am)A=\begin{pmatrix}A_{1}&0&\dots&0\\ 0&A_{2}&\dots&0\\ .\\ .\\ .\\ 0&0&\dots&A_{m}\end{pmatrix}, where AiA_{i}’s are square matrices, then X=(X1000X20...00Xm)SolAX=\begin{pmatrix}X_{1}&0&\dots&0\\ 0&X_{2}&\dots&0\\ .\\ .\\ .\\ 0&0&\dots&X_{m}\end{pmatrix}\in Sol_{A}, where XiSolAiX_{i}\in Sol_{A_{i}}.

3.1. Some special solutions

Example 3.15.1.

When A=λI+BA=\lambda I+B, for some λ0\lambda\neq 0, is a 2×22\times 2 Jordan block, non-trivial solution XX has one of the following form:
(λ10λ),(λ+λaaλ2λλa)or(λλaaλ2λ+λa)\begin{pmatrix}\lambda&1\\ 0&\lambda\end{pmatrix},\begin{pmatrix}\lambda+\lambda\sqrt{a}&a\\ -\lambda^{2}&\lambda-\lambda\sqrt{a}\end{pmatrix}or\begin{pmatrix}\lambda-\lambda\sqrt{a}&a\\ -\lambda^{2}&\lambda+\lambda\sqrt{a}\end{pmatrix} where a K\in K

Proof.

Let X=(x1x2x3x4)X=\begin{pmatrix}x_{1}&x_{2}\\ x_{3}&x_{4}\end{pmatrix}, then for A=λI2+BA=\lambda I_{2}+B

(3) AXA\displaystyle AXA =(λ10λ)(x1x2x3x4)(λ10λ)\displaystyle=\begin{pmatrix}\lambda&1\\ 0&\lambda\end{pmatrix}\begin{pmatrix}x_{1}&x_{2}\\ x_{3}&x_{4}\end{pmatrix}\begin{pmatrix}\lambda&1\\ 0&\lambda\end{pmatrix}
=(λ2x1+λx3λx1+x3+λ2x2+λx4λ2x3λx3+λ2x4)\displaystyle=\begin{pmatrix}\lambda^{2}x_{1}+\lambda x_{3}&\lambda x_{1}+x_{3}+\lambda^{2}x_{2}+\lambda x_{4}\\ \lambda^{2}x_{3}&\lambda x_{3}+\lambda^{2}x_{4}\end{pmatrix}
XAX\displaystyle XAX =(x1x2x3x4)(λ10λ)(x1x2x3x4)\displaystyle=\begin{pmatrix}x_{1}&x_{2}\\ x_{3}&x_{4}\end{pmatrix}\begin{pmatrix}\lambda&1\\ 0&\lambda\end{pmatrix}\begin{pmatrix}x_{1}&x_{2}\\ x_{3}&x_{4}\end{pmatrix}
=(λx12+x1x3+λx3x2λx1x2+x1x4+λx4x2λx1x3+x32+λx3x4λx2x3+x4x3+λx42)\displaystyle=\begin{pmatrix}\lambda x_{1}^{2}+x_{1}x_{3}+\lambda x_{3}x_{2}&\lambda x_{1}x_{2}+x_{1}x_{4}+\lambda x_{4}x_{2}\\ \lambda x_{1}x_{3}+x_{3}^{2}+\lambda x_{3}x_{4}&\lambda x_{2}x_{3}+x_{4}x_{3}+\lambda x_{4}^{2}\end{pmatrix}

Upon equating and solving, we have, x3=0x_{3}=0 or λ2-\lambda^{2}. Now if x3=0x_{3}=0, as a consequence from above equations, we get x1=x4=0x_{1}=x_{4}=0 or λ\lambda. But, x1,x4=0x_{1},x_{4}=0 is not possible, since by Theorem 3.15, when λ0\lambda\neq 0, non-trivial solution XX is similar to A, hence trace of AA and XX must be equal, which is 2λ2\lambda. As a result, we have x1,x4=λx_{1},x_{4}=\lambda, also which leads to x2=1.x_{2}=1.
Now, if x3=λ2x_{3}=-\lambda^{2}, then we have x1,x4=λ±λx2x_{1},x_{4}=\lambda\pm\lambda\sqrt{x_{2}}. But again, since AA and XX are similar, their trace must be equal. As a result, x1x_{1} should be conjugate of x4x_{4}, and x2x_{2} has a free choice.

Note that, when the dimension of the coefficient matrix increases, there will be more free variable as aa in this case, which hardly follows any pattern w.r.t the dimension. ∎

Example 3.15.2.

Let A be the Jordan block of size 2, with eigenvalue λ=0\lambda=0, we have the following.

(4) AXA\displaystyle AXA =(0100)(x1x2x3x4)(0100)=(0x300)\displaystyle=\begin{pmatrix}0&1\\ 0&0\end{pmatrix}\begin{pmatrix}x_{1}&x_{2}\\ x_{3}&x_{4}\end{pmatrix}\begin{pmatrix}0&1\\ 0&0\end{pmatrix}=\begin{pmatrix}0&x_{3}\\ 0&0\end{pmatrix}
XAX\displaystyle XAX =(x1x2x3x4)(0100)(x1x2x3x4)=(x1x3x1x4x32x4x3)\displaystyle=\begin{pmatrix}x_{1}&x_{2}\\ x_{3}&x_{4}\end{pmatrix}\begin{pmatrix}0&1\\ 0&0\end{pmatrix}\begin{pmatrix}x_{1}&x_{2}\\ x_{3}&x_{4}\end{pmatrix}=\begin{pmatrix}x_{1}x_{3}&x_{1}x_{4}\\ x_{3}^{2}&x_{4}x_{3}\end{pmatrix}

This gives x3=0,x1x4=0x_{3}=0,x_{1}x_{4}=0 and x2x_{2} has free choice over KK.
Hence, SolA={(aα0b)|a,b,αK,ab=0}Sol_{A}=\{\begin{pmatrix}a&\alpha\\ 0&b\end{pmatrix}|\ a,b,\alpha\in K,\ ab=0\}.

3.1.1. Some General Solutions when λ=0\lambda=0

Example 3.15.3.

Let A be the 3×33\times 3 Jordan block with eigenvalue 0, and the variable matrix be X=(abcdefghi)X=\begin{pmatrix}a&b&c\\ d&e&f\\ g&h&i\end{pmatrix}. Upon equating AXA=XAXAXA=XAX, we have the following polynomials, whose zeros give SolASol_{A}.

(5) f1=ad+bg\displaystyle f_{1}=ad+bg f2=ae+bhd\displaystyle f_{2}=ae+bh-d f3=af+bie\displaystyle f_{3}=af+bi-e
f4=d2+eg\displaystyle f_{4}=d^{2}+eg f5=de+ehg\displaystyle f_{5}=de+eh-g f6=df+eih\displaystyle f_{6}=df+ei-h
f7=gd+hg\displaystyle f_{7}=gd+hg f8=ge+h2\displaystyle f_{8}=ge+h^{2} f9=gf+hi\displaystyle f_{9}=gf+hi

Let us find a Gro¨\ddot{o}bner basis for the ideal f1,f2,,f9\langle f_{1},f_{2},\dots,f_{9}\rangle in the ring [a,b,c,d,e,f,g,h,i]\mathbb{C}[a,b,c,d,e,f,g,h,i], with respect to the lexicographic order for a>b>>ia>b>\dots>i. On simplifying from the Gro¨\ddot{o}bner basis, we get g,h,d,e=0g,h,d,e=0 and af+bi=0af+bi=0, and other variables have free choice. Hence,

SolA={(abc00f00i)|af+bi=0}.Sol_{A}=\{\begin{pmatrix}a&b&c\\ 0&0&f\\ 0&0&i\end{pmatrix}|\ af+bi=0\}.
Example 3.15.4.

For the coefficient matrix A being Jordan block of size n with eigenvalue 0, one can generalise the solutions in Example 3.15.3, to the following solutions class. But note that, for n>3n>3, this does not give the whole set of solutions.

X=(0a1a2.an2α000i=1n3aibi+1b10000b2...............0000bn200000)X=\begin{pmatrix}0&a_{1}&a_{2}&....&a_{n-2}&\alpha\\ 0&0&0&\dots&\sum_{i=1}^{n-3}a_{i}b_{i+1}&b_{1}\\ 0&0&0&\dots&0&b_{2}\\ .&.&.&\dots&.&.\\ .&.&.&\dots&.&.\\ .&.&.&\dots&.&.\\ 0&0&0&\dots&0&b_{n-2}\\ 0&0&0&\dots&0&0\end{pmatrix}

We can directly verify the above matrix belongs SolASol_{A}.

Lemma 3.16.

Let AMn(K)A\in M_{n}(K) be in its Jordan-canonical form, then for a matrix, MMn(K)M\in M_{n}(K), AM=MAAM=MA implies MK[A]M\in K[A].

Proof.

Theorem 3.17.

Let A be the Jordan block of size n+1n+1, n3n\geq 3, with eigenvalue 0. Any commuting solution to the YBE, with the coefficient matrix AA, will have the form,

X=(0100αβ00100α000100..................000001000000) or (00αβ000α0000............00000000) where α,βKX=\begin{pmatrix}0&1&0&0&\dots&\alpha&\beta\\ 0&0&1&0&\dots&0&\alpha\\ 0&0&0&1&\dots&0&0\\ .&.&.&.&\dots&.&.\\ .&.&.&.&\dots&.&.\\ .&.&.&.&\dots&.&.\\ 0&0&0&0&\dots&0&1\\ 0&0&0&0&\dots&0&0\end{pmatrix}\text{ or }\begin{pmatrix}0&0&\dots&\alpha&\beta\\ 0&0&\dots&0&\alpha\\ 0&0&\dots&0&0\\ .&.&\dots&.&.\\ .&.&\dots&.&.\\ .&.&\dots&.&.\\ 0&0&\dots&0&0\\ 0&0&\dots&0&0\end{pmatrix}\text{ where }\alpha,\beta\in K
Proof.

Here we have A=BA=B, as λ=0\lambda=0, hence if XX is a commuting solution to the YB equation, then XX must be a polynomial in BB. Let, X=a0I+a1B++an1Bn1+anBnX=a_{0}I+a_{1}B+...+a_{n-1}B^{n-1}+a_{n}B^{n} (note Bn+1=0B^{n+1}=0 ). Then,

(6) BXB\displaystyle BXB =B(a0I+a1B++an1Bn1+anBn)B\displaystyle=B(a_{0}I+a_{1}B+\dots+a_{n-1}B^{n-1}+a_{n}B^{n})B
=a0B2+a1B3++an2Bn\displaystyle=a_{0}B^{2}+a_{1}B^{3}+\dots+a_{n-2}B^{n}
=i=2nai2Bi\displaystyle=\sum_{i=2}^{n}a_{i-2}B^{i}
XBX\displaystyle XBX =(a0I+a1B++anBn)B(a0I+a1B++anBn)\displaystyle=(a_{0}I+a_{1}B+\dots+a_{n}B^{n})B(a_{0}I+a_{1}B+\dots+a_{n}B^{n})
=(an1a0+an2a1++a0an1)Bn++(a1a0+a0a1)B2+a02B\displaystyle=(a_{n-1}a_{0}+a_{n-2}a_{1}+\dots+a_{0}a_{n-1})B^{n}+\dots+(a_{1}a_{0}+a_{0}a_{1})B^{2}+a_{0}^{2}B
=i=1nj+k=i1ajakBi\displaystyle=\sum_{i=1}^{n}\sum_{j+k=i-1}a_{j}a_{k}B^{i}

Upon equating, we get

(7) a0\displaystyle a_{0} =0, 2a0a1=a0\displaystyle=0,\ 2a_{0}a_{1}=a_{0}
a1\displaystyle a_{1} =2a2a0+a12,(a1=0 or 1)\displaystyle=2a_{2}a_{0}+a_{1}^{2},(\implies a_{1}=0\text{ or }1)
a2\displaystyle a_{2} =a3a0+a2a1+a1a2+a0a3,(a2=0)\displaystyle=a_{3}a_{0}+a_{2}a_{1}+a_{1}a_{2}+a_{0}a_{3},(\implies a_{2}=0)
a3\displaystyle a_{3} =a4a0+a3a1+a22+a1a3+a0a4,(a3=0)\displaystyle=a_{4}a_{0}+a_{3}a_{1}+a_{2}^{2}+a_{1}a_{3}+a_{0}a_{4},(\implies a_{3}=0)

Continuing this way, we obtain, in general,

an2=an1a0+an2a1++a1an2+a0an1,(an2=0)a_{n-2}=a_{n-1}a_{0}+a_{n-2}a_{1}+...+a_{1}a_{n-2}+a_{0}a_{n-1},(\implies a_{n-2}=0)

and an1,ana_{n-1},a_{n} has free choice over KK. Hence XX has a general form as stated above when a1=1a_{1}=1, and a1=0a_{1}=0 respectively. ∎

4. Two Jordan blocks

Let the coefficient matrix 𝒜\mathcal{A} has two Jordan blocks in its Jordan-canonical form, say (A100A2)\begin{pmatrix}A_{1}&0\\ 0&A_{2}\end{pmatrix} where Ai=λiIni+BiA_{i}=\lambda_{i}I_{n_{i}}+B_{i}. Rewriting the YB equation in block form with X=(X1X2X3X4)X=\begin{pmatrix}X_{1}&X_{2}\\ X_{3}&X_{4}\end{pmatrix}, and expanding, gives us the set of four equations, namely:

(8) X1A1X1+X2A2X3=A1X1A1\displaystyle X_{1}A_{1}X_{1}+X_{2}A_{2}X_{3}=A_{1}X_{1}A_{1}
X1A1X2+X2A2X4=A1X2A2\displaystyle X_{1}A_{1}X_{2}+X_{2}A_{2}X_{4}=A_{1}X_{2}A_{2}
X3A1X1+X4A2X3=A2X3A1\displaystyle X_{3}A_{1}X_{1}+X_{4}A_{2}X_{3}=A_{2}X_{3}A_{1}
X3A1X2+X4A2X4=A2X4A2\displaystyle X_{3}A_{1}X_{2}+X_{4}A_{2}X_{4}=A_{2}X_{4}A_{2}

Also, note that when the coefficient matrix is invertible, any invertible solution is similar to the coefficient matrix, as we stated in Lemma 3.7.

Lemma 4.1.

If the coefficient matrix 𝒜\mathcal{A} has two Jordan blocks, AiA_{i} with λi0\lambda_{i}\neq 0, for i=1,2i\ =1,2, with the cyclic subspaces of AiA_{i} being PiP_{i} respectively. Then, for XSol𝒜X\in Sol_{\mathcal{A}}, Ker(X)=P1Ker(X)=P_{1} or P2P_{2} or P1P2P_{1}\bigoplus P_{2}.

Proof.

By Lemma 3.4, if XSol𝒜X\in Sol_{\mathcal{A}}, Ker(X)Ker(X) is an invariant subspace of 𝒜\mathcal{A}. Hence, Ker(X)Ker(X) should be one of P1,P2,P1P2,Eλ1,Eλ2P_{1},P_{2},P_{1}\bigoplus P_{2},E_{\lambda_{1}},E_{\lambda_{2}}. Now, in light of the Corollary 3.6.1, Eλ1,Eλ2E_{\lambda_{1}},E_{\lambda_{2}} can not be the kernel, which leads to the result.

Lemma 4.2.

Let A1A_{1} and A2A_{2} be two Jordan blocks with eigenvalues λ1\lambda_{1} and λ2\lambda_{2}, respectively, and X2SolA2X_{2}\in Sol_{A_{2}}. Then, A1X1A2=X1A2X2A_{1}X_{1}A_{2}=X_{1}A_{2}X_{2} if and only if either σ(A1)σ(X2)ϕ\sigma(A_{1})\cap\sigma(X_{2})\neq\phi or X1A2=0X_{1}A_{2}=0. Further, if A1,A2A_{1},A_{2} are Jordan blocks with non-zero eigenvalues and of the same dimension, then either A1=A2A_{1}=A_{2} or X1=0X_{1}=0. Also, when A1=A2=AA_{1}=A_{2}=A, X1X_{1} will have the form ZS1A1ZS^{-1}A^{-1}, where ZK[A]Z\in K[A], and for some SGLn(K)S\in GL_{n}(K).

Proof.

Let Y=X1A2Y=X_{1}A_{2}, then the equation A1X1A2=X1A2X2A_{1}X_{1}A_{2}=X_{1}A_{2}X_{2} becomes, A1Y=YX2A_{1}Y=YX_{2}. Then by Sylvester’s theorem 3.1, Y has a non-zero solution if and only if σ(A1)σ(X2)\sigma(A_{1})\cap\sigma(X_{2})\neq\varnothing. Therefore, if X1A20X_{1}A_{2}\neq 0, then σ(A1)σ(X2)\sigma(A_{1})\cap\sigma(X_{2})\neq\varnothing, or if X1A2=0X_{1}A_{2}=0, then σ(A1)σ(X2)=\sigma(A_{1})\cap\sigma(X_{2})=\varnothing.
Let X1A20X_{1}A_{2}\neq 0, and A1,A2A_{1},A_{2} are Jordan blocks of the same dimension, with non-zero eigenvalues. As X2SolA2X_{2}\in Sol_{A_{2}}, if X2X_{2} is non-trivial, then X2A2X_{2}\simeq A_{2}, i.e, σ(X2)={λ2}=σ(A2)\sigma(X_{2})=\{\lambda_{2}\}=\sigma(A_{2}). Also, we have σ(A1)σ(X2)\sigma(A_{1})\cap\sigma(X_{2})\neq\varnothing, i.e. σ(X2)={λ1}=σ(A1)\sigma(X_{2})=\{\lambda_{1}\}=\sigma(A_{1}), leads to λ1=λ2\lambda_{1}=\lambda_{2}. Since A1A_{1} and A2A_{2} are Jordan blocks and of the same dimension, we have, A1=A2A_{1}=A_{2}.
If X1A2=0X_{1}A_{2}=0, as A2A_{2} is invertible, we have X1=0X_{1}=0.
Now, when A1=A2=AA_{1}=A_{2}=A, the equation A1X1A2=X1A2X2A_{1}X_{1}A_{2}=X_{1}A_{2}X_{2} implies AX1A=X1AX2AX_{1}A=X_{1}AX_{2}. By taking Y=X1AY=X_{1}A, and since X2AX_{2}\simeq A, we have X2=SAS1X_{2}=SAS^{-1} for some S, and AYYSAS1=0AY-YSAS^{-1}=0 AYSYSA=0\Longrightarrow AYS-YSA=0. Calling YSYS as ZZ, we have AZ=ZAAZ=ZA. Since A is Jordan block, ZK[A]Z\in K[A].

Theorem 4.3.

Let 𝒜\mathcal{A} be a matrix having only two eigenvalues λ1,λ20\lambda_{1},\lambda_{2}\neq 0, with the same algebraic multiplicity, and geometric multiplicity 1. Further, let us assume 0 is an eigenvalue of XX. Let UGLn(K)U\in GL_{n}(K) be defined as, the columns of UU are generalized eigenvectors of 𝒜\mathcal{A}, such that U𝒜U1U\mathcal{A}U^{-1} is Jordan-canonical form of 𝒜\mathcal{A}. Then one of the following is true.

  1. (1)

    X=0X=0

  2. (2)

    X=U1(0ZS1A0Y2)UX=U^{-1}\begin{pmatrix}0&ZS^{-1}A\\ 0&Y_{2}\end{pmatrix}U or U1(000Y2)UU^{-1}\begin{pmatrix}0&0\\ 0&Y_{2}\end{pmatrix}U, where Y2SolAY_{2}\in Sol_{A}, ZK[A]Z\in K[A], AA is the Jordan block corresponding to eigenvalue λ=λ1=λ2\lambda=\lambda_{1}=\lambda_{2}, and for some SGLn(K)S\in GL_{n}(K)

  3. (3)

    X=U1(Y10ZS1A0)UX=U^{-1}\begin{pmatrix}Y_{1}&0\\ ZS^{-1}A&0\end{pmatrix}U or U1(Y1000)UU^{-1}\begin{pmatrix}Y_{1}&0\\ 0&0\end{pmatrix}U, where Y1SolAY_{1}\in Sol_{A}, ZK[A]Z\in K[A], AA is the Jordan block corresponding to eigenvalue λ=λ1=λ2\lambda=\lambda_{1}=\lambda_{2}, and for some SGLn(K)S\in GL_{n}(K)

Proof.

Let us discuss the three cases for the kernel of XX. If the kernel of XX is P1P2P_{1}\bigoplus P_{2} then clearly X=0X=0. If the kernel is P1P_{1}, let us write the block form of XX in the basis of P1P2P_{1}\bigoplus P_{2}. We get the following block form X(0Y10Y2)X\simeq\begin{pmatrix}0&Y_{1}\\ 0&Y_{2}\end{pmatrix} for some matrices Y1,Y2Y_{1},Y_{2}. i.e., For some UGLn(K)U\in GL_{n}(K), we have,

Y=UXU1=U(X1X2X3X4)U1=(0Y10Y2),Y=UXU^{-1}=U\begin{pmatrix}X_{1}&X_{2}\\ X_{3}&X_{4}\end{pmatrix}U^{-1}=\begin{pmatrix}0&Y_{1}\\ 0&Y_{2}\end{pmatrix},

where the column vectors of UU are from the basis of P1P2P_{1}\bigoplus P_{2}, such that U𝒜U1=𝒜U\mathcal{A}U^{-1}=\mathcal{A^{\prime}} is the Jordan-canonical form of 𝒜\mathcal{A}. Then, rewriting the YBE as 𝒜Y𝒜=Y𝒜Y\mathcal{A^{\prime}}Y\mathcal{A^{\prime}}=Y\mathcal{A^{\prime}}Y, where 𝒜=U𝒜U1\mathcal{A^{\prime}}=U\mathcal{A}U^{-1}, gives the original solution XX, as X=U1YUX=U^{-1}YU.

Let 𝒜=(A100A2)\mathcal{A^{\prime}}=\begin{pmatrix}A_{1}&0\\ 0&A_{2}\end{pmatrix} , where Ai=λiI+BiA_{i}=\lambda_{i}I+B_{i}, for i=1,2i=1,2, is the Jordan block corresponding to eigenvalue λi\lambda_{i}. Let us write the YB equation,

Y2A2Y2=A2Y2A2Y_{2}A_{2}Y_{2}=A_{2}Y_{2}A_{2}
A1Y1A2=A2Y2A2A_{1}Y_{1}A_{2}=A_{2}Y_{2}A_{2}

The first equation is just an instance of YBE for the single Jordan block case, which by Theorem 3.15 and the fact that det(A2)0det(A_{2})\neq 0, we get X2A2X_{2}\simeq A_{2} or 0. By Lemma 4.2, we have, either X1=0X_{1}=0 or X1=ZS1A1X_{1}=ZS^{-1}A^{-1}, where ZK[A]Z\in K[A]. This gives us the solution stated in the second case.

The proof of the third case is exactly similar to the second case. ∎

Example 4.3.1.

The following example will give insight into the significance of the Theorem 4.3, which otherwise would have been more complicated to solve. Let 𝒜=(A00A)\mathcal{A}=\begin{pmatrix}A&0\\ 0&A\end{pmatrix}, where A=(λ10λ)A=\begin{pmatrix}\lambda&1\\ 0&\lambda\end{pmatrix}, for some λ0\lambda\neq 0. Then by 4.3, one of the forms for the solution XSol𝒜X\in Sol_{\mathcal{A}} is (X10X20)\begin{pmatrix}X_{1}&0\\ X_{2}&0\end{pmatrix}, where X1SolAX_{1}\in Sol_{A}, and let X2X_{2} be the variable matrix (bcde)\begin{pmatrix}b&c\\ d&e\end{pmatrix}. Also, for X1X_{1}, let’s choose one of the forms given in the special solutions Example 3.15.1, say X1=(λ+λaaλ2λλa)X_{1}=\begin{pmatrix}\lambda+\lambda\sqrt{a}&a\\ -\lambda^{2}&\lambda-\lambda\sqrt{a}\end{pmatrix} for some aKa\in K. Then on simplifying the equation 𝒜X𝒜=X𝒜X\mathcal{A}X\mathcal{A}=X\mathcal{A}X, it reduces to the following equations:

(9) (1) λ2(abcλ)=bλ2+dλ\displaystyle\text{(1) }\lambda^{2}(\sqrt{a}b-c\lambda)=b\lambda^{2}+d\lambda (2) λ2(adcλ)=dλ2\displaystyle\text{(2) }\lambda^{2}(\sqrt{a}d-c\lambda)=d\lambda^{2}
(3) adλaλ(d+eλ)=0\displaystyle\text{(3) }ad\lambda-\sqrt{a}\lambda(d+e\lambda)=0 (4) d+eλ=abλaλ(b+cλ)\displaystyle\text{(4) }d+e\lambda=ab\lambda-\sqrt{a}\lambda(b+c\lambda)

Then for different cases of λ,a\lambda,a, we have the following solutions for X2X_{2}, which along with the given X1X_{1}, completes the solution XX.

  • i

    When a=0a=0, X2=(bebλeλe)X_{2}=\begin{pmatrix}b&\frac{e-b}{\lambda}\\ -e\lambda&e\end{pmatrix} where b,eKb,e\in K.

  • ii

    When a0,a1,λ=1a\neq 0,a\neq 1,\lambda=1, X2=(ca1+e(a1)2cea1e)X_{2}=\begin{pmatrix}\frac{c}{\sqrt{a}-1}+\frac{e}{(\sqrt{a}-1)^{2}}&c\\ \frac{e}{\sqrt{a}-1}&e\end{pmatrix} where c,eKc,e\in K.

  • iii

    When a0,a1,λ1a\neq 0,a\neq 1,\lambda\neq 1, X2=(e(a1)20ea1e)X_{2}=\begin{pmatrix}\frac{e}{(\sqrt{a}-1)^{2}}&0\\ \frac{e}{\sqrt{a}-1}&e\end{pmatrix} where eKe\in K.

  • iv

    When a=1,λ1a=1,\lambda\neq 1, X2=(b000)X_{2}=\begin{pmatrix}b&0\\ 0&0\end{pmatrix}

  • v

    When a=1,λ=1a=1,\lambda=1, X2=(bcc0)X_{2}=\begin{pmatrix}b&c\\ -c&0\end{pmatrix} where b,cKb,c\in K.

5. Pencils of solutions

Lemma 5.1.

Let XSolAX\in Sol_{A}, then for a matrix MM with MA=0=AMMA=0=AM, for any αK\alpha\in K, X+αMSolAX+\alpha M\in Sol_{A}.

Proof.

we have we have A(X+αM)A=AXA+αAMA=AXAA(X+\alpha M)A=AXA+\alpha AMA=AXA

(X+αM)A(X+αM)=XAX+αXAM+αMAX+αMAM=XAX(X+\alpha M)A(X+\alpha M)=XAX+\alpha XAM+\alpha MAX+\alpha MAM=XAX

Note that in this case, MM itself forms a solution. ∎

Lemma 5.2.

Let XSolAX\in Sol_{A}, for any α,βK\alpha,\beta\in K, αX+βMSolA\alpha X+\beta M\in Sol_{A}, if MA=0=AMMA=0=AM and AXA=0AXA=0.

Proof.

If XSolAX\in Sol_{A}, then for α0,1\alpha\neq 0,1, αX\alpha X is a solution if AXA=0AXA=0 [5]. Then the rest follows directly from the above Lemma 5.1. ∎

When A is non-singular, MA=0=AMMA=0=AM demands M to be 0. This discussion is interesting when A is singular. Let A be a singular matrix in its Jordan form. Say A=(J100J2)A=\begin{pmatrix}J_{1}&0\\ 0&J_{2}\end{pmatrix} where J1=(010001000)J_{1}=\begin{pmatrix}0&1&0\\ 0&0&1\\ 0&0&0\end{pmatrix}, J2=(λ10λ)J_{2}=\begin{pmatrix}\lambda&1\\ 0&\lambda\end{pmatrix} for some λ0K\lambda\neq 0\in K. Then we can see that MM has the form (00α0000000000000000000000)\begin{pmatrix}0&0&\alpha&0&0\\ 0&0&0&0&0\\ 0&0&0&0&0\\ 0&0&0&0&0\\ 0&0&0&0&0\end{pmatrix}, where αK\alpha\in K. When J2J_{2} is also Jordan block of 0, that is J2=(0100)J_{2}=\begin{pmatrix}0&1\\ 0&0\end{pmatrix}, M will have the following form. M=(00a10a2000000000000a30a400000)M=\begin{pmatrix}0&0&a_{1}&0&a_{2}\\ 0&0&0&0&0\\ 0&0&0&0&0\\ 0&0&a_{3}&0&a_{4}\\ 0&0&0&0&0\end{pmatrix}, where aiKa_{i}\in K. We can observe that when there is a Jordan block of non-zero eigenvalue, the corresponding rows and columns in M will be annihilated. Also, corresponding blocks in M are solutions to respective YBEs JiXJi=XJiXJ_{i}XJ_{i}=XJ_{i}X.

Theorem 5.3.

Let X0,X1SolAX_{0},X_{1}\in Sol_{A}, for any λK\lambda\neq\in K, X0+λX1SolAX_{0}+\lambda X_{1}\in Sol_{A} if and only if AX1A=0=X1AX1AX_{1}A=0=X_{1}AX_{1} and X0AX1+X1AX0=0X_{0}AX_{1}+X_{1}AX_{0}=0.

Proof.

()(\Rightarrow) We have X0AX0=AX0A,X1AX1=AX1AX_{0}AX_{0}=AX_{0}A,X_{1}AX_{1}=AX_{1}A. Which implies A(X0+λX1)A=AX0A+λAX1A=AX0AA(X_{0}+\lambda X_{1})A=AX_{0}A+\lambda AX_{1}A=AX_{0}A, as AX1A=0AX_{1}A=0. Also, we have (X0+λX1)A(X0+λX1)=X0AX0(X_{0}+\lambda X_{1})A(X_{0}+\lambda X_{1})=X_{0}AX_{0}.

()(\Leftarrow) For any λK\lambda\neq\in K, if X0+λX1X_{0}+\lambda X_{1} is a solution, then,

(10) A(X0+λX1)A\displaystyle A(X_{0}+\lambda X_{1})A =AX0A+λAX1A\displaystyle=AX_{0}A+\lambda AX_{1}A
=(X0+λX1)A(X0+λX1)\displaystyle=(X_{0}+\lambda X_{1})A(X_{0}+\lambda X_{1})
=X0AX0+λX0AX1+λX1AX0+λ2X1AX1.\displaystyle=X_{0}AX_{0}+\lambda X_{0}AX_{1}+\lambda X_{1}AX_{0}+\lambda^{2}X_{1}AX_{1}.

Then λAX1A=λX0AX1+λX1AX0+λ2X1AX1\lambda AX_{1}A=\lambda X_{0}AX_{1}+\lambda X_{1}AX_{0}+\lambda^{2}X_{1}AX_{1}. Since it’s true for every λ\lambda, it’s true for λ=1\lambda=1 also, which gives X0AX1+X1AX0=0X_{0}AX_{1}+X_{1}AX_{0}=0. This implies λAX1A=0=λ2X1AX1\lambda AX_{1}A=0=\lambda^{2}X_{1}AX_{1} which is true for any λ\lambda, gives AX1A=0=X1AX1AX_{1}A=0=X_{1}AX_{1}. ∎

Example 5.3.1.

When A is the 3×33\times 3 Jordan block, with eigenvalue zero, in Example 3.15.3 of special solutions, we have seen that X must have the form (abc00f00i)\begin{pmatrix}a&b&c\\ 0&0&f\\ 0&0&i\end{pmatrix}, such that af+bi=0af+bi=0. Then, by taking Xj=(ajbjcj00fj00ij)SolAX_{j}=\begin{pmatrix}a_{j}&b_{j}&c_{j}\\ 0&0&f_{j}\\ 0&0&i_{j}\end{pmatrix}\in Sol_{A}, for j=0,1j=0,1, we can easily verify the condition for X0+λX1SolAX_{0}+\lambda X_{1}\in Sol_{A}. Note that AXjA=0AX_{j}A=0. Now,
X1AX0+X0AX1=(00a1f0+a0f1+b1i0+b0i1000000)=0X_{1}AX_{0}+X_{0}AX_{1}=\begin{pmatrix}0&0&a_{1}f_{0}+a_{0}f_{1}+b_{1}i_{0}+b_{0}i_{1}\\ 0&0&0\\ 0&0&0\end{pmatrix}=0 , if λ(a1f0+a0f1+b1i0+b0i1)=0\lambda(a_{1}f_{0}+a_{0}f_{1}+b_{1}i_{0}+b_{0}i_{1})=0. This is agreeing with the general form of solution when

(a0+λa1b0+λb1c0+λc100f0+λf100i0+λi1)SolA.\begin{pmatrix}a_{0}+\lambda a_{1}&b_{0}+\lambda b_{1}&c_{0}+\lambda c_{1}\\ 0&0&f_{0}+\lambda f_{1}\\ 0&0&i_{0}+\lambda i_{1}\end{pmatrix}\in Sol_{A}.

6. Conclusion

In this paper, solutions to the Yang-Baxter equation has studied for coefficient matrix A being single Jordan blocks and A having two Jordan block in the canonical form. A complete characterization of solutions is still an open question. Although the solutions are harder to find, it inspire to approach through many parallel ways including via Gro¨\ddot{o}bner basis. In future we would like to look into those kinds of solutions more, also these theories ignite light to finding solutions for similar interesting matrix equations such as (i) ΦA(X)X=0(i)\text{ }\Phi_{A}(X)X=0, (ii) ΦA(X)X=ΦX(A)A(ii)\text{ }\Phi_{A}(X)X=\Phi_{X}(A)A.

7. Bibliography

References

  • [1] M. Castelli, G. Pinto, and W. Rump. On the indecomposable involutive set-theoretic solutions of the Yang-Baxter equation of prime-power size. Comm. Algebra, 48(5):1941–1955, 2020.
  • [2] D. Chen, Z. Chen, and X. Yong. Explicit solutions of the Yang-Baxter-like matrix equation for a diagonalizable matrix with spectrum contained in {1,α,0}\{1,\alpha,0\}. Appl. Math. Comput., 348:523–530, 2019.
  • [3] J. Ding and H. Tian. Solving the Yang-Baxter-like matrix equation for a class of elementary matrices. Comput. Math. Appl., 72(6):1541–1548, 2016.
  • [4] J. Ding, C. Zhang, and N. Rhee. Commuting solutions of the yang–baxter matrix equation. Applied Mathematics Letters, 44:1–4, 2015.
  • [5] N. v. Dinčić and B. D. Djordjević. On the intrinsic structure of the solution set to the Yang-Baxter-like matrix equation. Rev. R. Acad. Cienc. Exactas Fís. Nat. Ser. A Mat. RACSAM, 116(2):Paper No. 73, 24, 2022.
  • [6] Q. Dong, J. Ding, and Q. Huang. Commuting solutions of a quadratic matrix equation for nilpotent matrices. Algebra Colloquium, 25(1):31–44, 2018.
  • [7] V. G. Drinfeld. On some unsolved problems in quantum group theory. In Quantum groups (Leningrad, 1990), volume 1510 of Lecture Notes in Math., pages 1–8. Springer, Berlin, 1992.
  • [8] L. Lu. Manifold expressions of all solutions of the yang–baxter-like matrix equation for rank-one matrices. Applied Mathematics Letters, 132:108175, 2022.
  • [9] S. I. A. Mansour, J. Ding, and Q. Huang. Explicit solutions of the Yang-Baxter-like matrix equation for an idempotent matrix. Appl. Math. Lett., 63:71–76, 2017.
  • [10] D. Shen, M. Wei, and Z. Jia. On commuting solutions of the Yang-Baxter-like matrix equation. J. Math. Anal. Appl., 462(1):665–696, 2018.
  • [11] J. Sylvester. Sur l’equations en matrices px=xqpx=xq. C. R. Acad. Sci. Paris., 99(2):67–71, 1884.
  • [12] H. Tian. All solutions of the Yang-Baxter-like matrix equation for rank-one matrices. Appl. Math. Lett., 51:55–59, 2016.
  • [13] H.-H. Yin, X. Wang, X.-B. Tang, and L. Chen. On the commuting solutions to the Yang-Baxter-like matrix equation for identity matrix minus special rank-two matrices. Filomat, 32(13):4591–4609, 2018.
  • [14] D. Zhou, G. Chen, and J. Ding. On the Yang-Baxter-like matrix equation for rank-two matrices. Open Math., 15(1):340–353, 2017.
  • [15] D. Zhou, G. Chen, and J. Ding. Solving the yang-baxter-like matrix equation for rank two matrices. J. Comput. Appl. Math, 313(1):142–151, 2017.
  • [16] D. Zhou and J. Ding. Solving the Yang-Baxter-like matrix equation for nilpotent matrices of index three. Int. J. Comput. Math., 95(2):303–315, 2018.
  • [17] D.-M. Zhou and H.-Q. Vu. Some non-commuting solutions of the yang-baxter-like matrix equation. Open Mathematics, 18(1):948–969, 2020.
  • [18] D.-M. Zhou, X.-X. Ye, Q.-W. Wang, J.-W. Ding, and W.-Y. Hu. Explicit solutions of the Yang-Baxter-like matrix equation for a singular diagonalizable matrix with three distinct eigenvalues. Filomat, 35(12):3971–3982, 2021.