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

Perfect colorings of hypergraphs

A. A. Taranenko Sobolev Institute of Mathematics, Novosibirsk, Russia; [email protected]
(October 24, 2024)
Abstract

Perfect colorings (equitable partitions) of graphs are extensively studied, while the same concept for hypergraphs attracts much less attention. The aim of this paper is to develop basic notions and properties of perfect colorings for hypergraphs. Firstly, we introduce a multidimensional matrix equation for perfect colorings of hypergraphs and compare this definition with a standard approach based on the incidence graph. Next, we show that the eigenvalues of the parameter matrix of a perfect coloring are eigenvalues of the multidimensional adjacency matrix of a hypergraph. We consider coverings of hypergraphs as a special case of perfect colorings and prove a theorem on the existence of a common covering of two hypergraphs. As an example, we show that a kk-transversal in a hypergraph corresponds to a perfect coloring and calculate its parameters. At last, we find all perfect 22-colorings of the Fano’s plane hypergraph and compute some eigenvalues of this hypergraph.

Keywords: perfect coloring; eigenvalues of hypergraphs; coverings of hypergraphs; multidimensional adjacency matrix.

MSC2020: 05C15; 05C50; 05C65.

1 Introduction

A coloring of vertices of a graph is called perfect if a color of a vertex uniquely defines colors of adjacent vertices.

Perfect colorings of graphs have been studied for quite a long time and arose in the literature under different names. One of the most known of them is “equitable partition” that was introduced by Delsarte [dels.assch]. In book [CveDoobSac.gspectra] objects equivalent to perfect colorings are named as divisors of graphs. An algebraic definition of perfect colorings firstly appeared in book [godsil.algcomb]. At last, in [GodRoy.alggrath] one finds some information on graph coverings that can be considered as perfect colorings of a special type.

Graph perfect colorings have plenty of applications in coding theory and the theory of combinatorial designs. There are many results on enumeration or characterization of perfect colorings in certain families of graphs (see, e.g., survey papers [our.perfcolorHam, BorRifaZin.compregcodes]).

Meanwhile, the specifics of perfect colorings in hypergraphs are hardly studied. As it was noted in “Handbook of combinatorics” [GraGroLo.handbook, Ch. 31.2], a perfect coloring of a hypergraph is equivalent to a certain perfect coloring of its incidence graph. In [PotAv.hypercolor], it was shown that some combinatorial designs correspond to perfect colorings of hypergraphs. In particular, every transversal in a regular uniform hypergraph is a perfect 22-coloring.

Reduction of perfect colorings of hypergraphs to colorings of graphs implies immediately some of their properties. On the other hand, this approach uses redundant parameters of a coloring of hyperedges that can be deduced from a coloring of vertices.

The main aim of this paper is to propose and develop another perspective for perfect colorings of hypergraphs based on their multidimensional adjacency matrices.

The paper is structured as follows.

Section 2 is preliminary. It provides all definitions and auxiliary results on hypergraphs, their adjacency and incidence matrices, multidimensional matrices and their products, and eigenvalues of multidimensional matrices and hypergraphs. In this section we also briefly review the main results on perfect colorings in graphs that will be later generalized for hypergraphs.

Section 3 is the main part of the paper. Firstly, we consider in detail the definition of a hypergraph perfect coloring as a perfect coloring of its incidence graph. Although the concept of the hypergraph perfect coloring was introduced earlier, it is a debatable question how to define their parameters so that they still have the most nice properties of the graph case. Here we pay special attention to this question because it was omitted in previous works. Also we deduce some simple corollaries from the approach based on the incidence graph that were not given before. In particular, we get necessary and sufficient conditions for a pair of matrices to be parameters of a perfect coloring of some hypergraph (Theorem 3) and establish the analogue of the Weisfeiler–Leman–Vizing theorem on the refinement of perfect coloring (Theorem 4).

In the second part of Section 3, we propose a new definition of a hypergraph perfect coloring in terms of its multidimensional adjacency matrix. In Theorem 5 we show that for uniform hypergraphs this definition is equivalent to the previous one and introduce the multidimensional parameter matrix of a perfect coloring. We also prove that the multidimensional parameter matrix can be symmetrized (Theorem 6) and connect it with the 22-dimensional parameter matrices of the coloring (Corollary 1). The major asset of the multidimensional parameter matrix is that all its eigenvalues are also the eigenvalues of the adjacency matrix of a hypergraph (Theorem 8). We obtain it as a corollary from a more general statement for multidimensional matrices (Theorem 7). These results allow one to find several eigenvalues of a multidimensional matrix (or hypergraph) on the basis of eigenvalues of certain smaller matrices (hypergraphs).

In Section 4, we consider coverings of hypergraphs as a special case of perfect colorings. This allows us to easily get the following properties of coverings: characterization of hypergraphs that cover a given one (Theorem 9), preservation of parameters of a perfect coloring in coverings (Theorem 11), and inclusion of the spectrum of hypergraphs into the spectrum of their coverings (Theorem 10). The last result was recently obtained in [SongFanWang.hyperspeccover] by another method. At the end of the section we generalize Leighton’s theorem [leighton.comcover] on common coverings from graphs to hypergraphs (Theorem 12 and Corollary 4).

At last, in Section 5 we consider several examples of hypergraph perfect colorings. We find the parameter matrices of a hypergraph transversal as a perfect 22-coloring and calculate their eigenvalues (Theorem 13). Moreover, we show that multidimensional parameters may distinguish hypergraphs with no transversals, while the 22-dimensional approach does not. Finally, we compute the parameters of perfect 22-colorings of regular 33-uniform hypergraphs and find all perfect 22-colorings of the Fano’s plane hypergraph and some of its eigenvalues.

2 Definitions and preliminaries

A hypergraph 𝒢\mathcal{G} is a pair of sets X=X(𝒢)X=X(\mathcal{G}) and E=E(𝒢)E=E(\mathcal{G}), |X|=n|X|=n, |E|=m|E|=m. XX is called a set of vertices and EE is a set of hyperedges, each eEe\in E is some (nonempty) subset of XX. A vertex xXx\in X is said to be incident to a hyperedge ee if xex\in e. Vertices xx and yy are adjacent if there is eEe\in E incident to both of them. If EE is a multiset, then 𝒢\mathcal{G} is said to be a multihypergraph.

A hypergraph is connected if there is an interchanging sequence of incident vertices and hyperedges that connects a vertex xx with any other vertex yy. Such a sequence is called a Berge path from xx to yy.

A hypergraph 𝒢\mathcal{G} is said to be dd-uniform if every hyperedge of 𝒢\mathcal{G} consists of dd vertices. Simple graphs are exactly 22-uniform hypergraphs. A dd-uniform hypergraph is complete if its hyperedges are all dd-element subsets of the vertex set.

A degree of a vertex xx is the number of hyperedges containing xx. A hypergraph 𝒢\mathcal{G} is rr-regular if every vertex of 𝒢\mathcal{G} has degree rr.

A hypergraph 𝒢\mathcal{G} is said to be dd-partite if there is a partition of its vertex set into dd disjoint subsets (parts) such that each hyperedge of 𝒢\mathcal{G} contains no more than one vertex from each part.

A kk-transversal in a hypergraph 𝒢=(X,E)\mathcal{G}=(X,E) is a set of vertices YXY\subseteq X such that every hyperedge of 𝒢\mathcal{G} contains exactly kk vertices from YY. A kk-factor in 𝒢\mathcal{G} is a set of hyperedges UEU\subseteq E such that every vertex of 𝒢\mathcal{G} is incident to exactly kk hyperedges from UU. 11-factors are said to be perfect matchings and 11-transversals are known as transversals.

Transversals and factors are dual in the following sense. For every hypergraph 𝒢\mathcal{G} one can consider a dual hypergraph 𝒢\mathcal{G}^{*} obtained from 𝒢\mathcal{G} by interchanging sets of vertices and hyperedges and preserving the incidence relations between them. Then every kk-transversal in 𝒢\mathcal{G} is a kk-factor in 𝒢\mathcal{G}^{*} and vice versa.

Every hypergraph 𝒢=(X,E)\mathcal{G}=(X,E) corresponds to the incidence graph GG that is a bipartite graph with parts XX and EE: xXx\in X and eEe\in E are adjacent in GG if and only if they are incident in 𝒢\mathcal{G}. Incidence graph is also known as the Levi graph or the bipartite representation of a hypergraph.

2.1 Matrices and eigenvalues of hypergraphs

Let A={ai,j}A=\{a_{i,j}\}, i=1,,ni=1,\ldots,n, j=1,,mj=1,\ldots,m be a matrix of size n×mn\times m. If n=mn=m then AA is a matrix of order nn.

We will say that a matrix AA of size n×mn\times m is a block matrix with blocks Bi,jB_{i,j} of sizes ni×mjn_{i}\times m_{j}, i=1,,ki=1,\ldots,k, j=1,,lj=1,\ldots,l, i=1kni=n\sum\limits_{i=1}^{k}n_{i}=n, i=1lmj=m\sum\limits_{i=1}^{l}m_{j}=m, if (after, possibly, appropriate permutations of rows and columns) the matrix AA can be presented in a form

A=(B1,1B1,lBk,1Bk,l).A=\left(\begin{array}[]{ccc}B_{1,1}&\cdots&B_{1,l}\\ \vdots&\ddots&\vdots\\ B_{k,1}&\cdots&B_{k,l}\end{array}\right).

For shortness, we write A={Bi,j}A=\{B_{i,j}\}.

Let [n]={1,,n}[n]=\{1,\dots,n\}. A dd-dimensional matrix 𝔸\mathbb{A} of order nn is an array of entries (aα)(a_{\alpha}), aαa_{\alpha}\in\mathbb{R}, indexed by tuples α[n]d\alpha\in[n]^{d}, α=(α1,,αd)\alpha=(\alpha_{1},\ldots,\alpha_{d}). Vectors can be considered as 11-dimensional matrices, and usual matrices have 22 dimensions.

Let 𝔸\mathbb{A} be a dd-dimensional matrix of order nn. A hyperplane Γ\Gamma of direction ii in 𝔸\mathbb{A} is a (d1)(d-1)-dimensional submatrix obtained by fixing a value of index component αi\alpha_{i} and letting the other d1d-1 components vary.

A matrix 𝔸\mathbb{A} is said to be symmetric if for every permutation σSd\sigma\in S_{d} and for every index α[n]d\alpha\in[n]^{d} we have aα=aσ(α)a_{\alpha}=a_{\sigma(\alpha)}, where σ(α)=(ασ(1),,ασ(d))\sigma(\alpha)=(\alpha_{\sigma(1)},\ldots,\alpha_{\sigma(d)}). In other words, the matrix 𝔸\mathbb{A} does not change under permutations of directions of hyperplanes.

The dd-dimensional identity matrix 𝕀\mathbb{I} is the matrix with entries iα=1i_{\alpha}=1 if α1==αd\alpha_{1}=\cdots=\alpha_{d} and iα=0i_{\alpha}=0 otherwise.

Let 𝒢=(X,E)\mathcal{G}=(X,E) be hypergraph on nn vertices with mm hyperedges. The incidence matrix BB of the hypergraph 𝒢\mathcal{G} is a matrix of size n×mn\times m such that bx,e=1b_{x,e}=1 if a vertex xx is incident to a hyperedge ee and bx,e=0b_{x,e}=0 otherwise.

If 𝒢\mathcal{G} is a uniform hypergraph, then we can define its multidimensional adjacency matrix. The adjacency matrix 𝔸\mathbb{A} of a dd-uniform hypergraph 𝒢=(X,E)\mathcal{G}=(X,E) on nn vertices is a dd-dimensional matrix of order nn in which entries aαa_{\alpha} with index α=(x1,,xd)E\alpha=(x_{1},\ldots,x_{d})\in E are equal to (d1)!1(d-1)!^{-1} and all other entries of 𝔸\mathbb{A} are 0. Such scale of entries of the matrix 𝔸\mathbb{A} is taken for the sake of compactness of future expressions. By definition, the adjacency matrix 𝔸\mathbb{A} is symmetric.

If GG is a graph, then an eigenvalue of GG is an eigenvalue of its 22-dimensional adjacency matrix. For hypergraphs, there are several ways to define eigenvalues, the most popular approaches use the following:

  1. 1.

    maximization of some multilinear form generated by a hypergraph;

  2. 2.

    eigenvalues of the 22-dimensional signless Laplacian matrix BBTBB^{T};

  3. 3.

    eigenvalues of the multidimensional adjacency matrix 𝔸\mathbb{A}.

In this paper we will utilize the last of them. For a brief survey on other definitions of hypergraph eigenvalues see, e.g., the introduction of [coop.specrandcomp]. To define eigenvalues of multidimensional matrices, we use the following operation.

Let 𝔸\mathbb{A} be a dd-dimensional matrix of order nn and 𝔹\mathbb{B} be a tt-dimensional matrix of the same order. Define the product 𝔸𝔹\mathbb{A}\circ\mathbb{B} to be the ((d1)(t1)+1)((d-1)(t-1)+1)-dimensional matrix \mathbb{C} of order nn with entries

ci,β2,,βd=i2=1nid=1nai,i2,,idbi2,β2bid,βd,c_{i,\beta^{2},\ldots,\beta^{d}}=\sum\limits_{i_{2}=1}^{n}\cdots\sum\limits_{i_{d}=1}^{n}a_{i,i_{2},\ldots,i_{d}}\cdot b_{i_{2},\beta^{2}}\cdots b_{i_{d},\beta^{d}},

where indices β2,,βd[n]t1\beta^{2},\ldots,\beta^{d}\in[n]^{t-1}, i[n]i\in[n].

The following properties of the product \circ can be derived directly from the definition or found in [shao.tensprod].

Proposition 1.

Let 𝔸\mathbb{A}, 𝔹\mathbb{B}, and \mathbb{C} be multidimensional matrices of appropriate sizes.

  1. 1.

    If dimensions of matrices 𝔸\mathbb{A} and 𝔹\mathbb{B} do not exceed 22, then \circ is a standard matrix (dot) product: 𝔸𝔹=𝔸𝔹\mathbb{A}\circ\mathbb{B}=\mathbb{A}\mathbb{B}.

  2. 2.

    The product of multidimensional matrices is associative: (𝔸𝔹)=𝔸(𝔹)(\mathbb{A}\circ\mathbb{B})\circ\mathbb{C}=\mathbb{A}\circ(\mathbb{B}\circ\mathbb{C}).

  3. 3.

    If λ\lambda\in\mathbb{R} and 𝔸\mathbb{A} is a dd-dimensional matrix, then 𝔸(λ𝔹)=λd1(𝔸𝔹\mathbb{A}\circ(\lambda\mathbb{B})=\lambda^{d-1}(\mathbb{A}\circ\mathbb{B}).

Let 𝔸\mathbb{A} be a dd-dimensional matrix of order nn. We will say that λ\lambda is an eigenvalue of 𝔸\mathbb{A} if there is a vector x=(x1,,xn)x=(x_{1},\ldots,x_{n}) such that Ax=λ𝕀xA\circ x=\lambda\mathbb{I}\circ x, where 𝕀\mathbb{I} is a dd-dimensional identity matrix of order nn and 𝕀x\mathbb{I}\circ x is a vector (x1d1,,xnd1)(x_{1}^{d-1},\ldots,x_{n}^{d-1}). The vector xx is called an eigenvector corresponding to the eigenvalue λ\lambda, a pair (λ,x)(\lambda,x) is said to be an eigenpair for 𝔸\mathbb{A}.

If 𝒢\mathcal{G} is a dd-uniform hypergraph with the dd-dimensional adjacency 𝔸\mathbb{A}, then the eigenvalues of the matrix 𝔸\mathbb{A} are said to be the eigenvalues of 𝒢\mathcal{G}, the eigenvectors of 𝔸\mathbb{A} are the eigenvectors of 𝒢\mathcal{G}.

The set V(λ)nV(\lambda)\subset\mathbb{C}^{n} of all eigenvectors for an eigenvalue λ\lambda is a complex algebraic variety, and the geometric multiplicity of the eigenvalue λ\lambda is the dimension of V(λ)V(\lambda).

The present definition of eigenvalues and eigenvectors for tensors was proposed in 2005 by Lim [lim.eigentensor] and was studied for symmetric matrices by Qi [qi.eigentensor]. They used the theory of determinants of multidimensional matrices and resultants of multilinear systems developed in book [GeKaZe.multdet] by Gelfand, Kapranov, and Zelevinsky. In particular, they proved that all eigenvalues of a multidimensional matrix 𝔸\mathbb{A} are roots of its characteristic polynomial φ𝔸\varphi_{\mathbb{A}}.

Theorem 1 ([qi.eigentensor]).

For a dd-dimensional matrix 𝔸\mathbb{A} of order nn, there is a characteristic polynomial φ𝔸\varphi_{\mathbb{A}} of degree n(d1)n1n(d-1)^{n-1} such that a number λ\lambda\in\mathbb{C} is an eigenvalue of 𝔸\mathbb{A} if and only if λ\lambda is a root of φ𝔸\varphi_{\mathbb{A}}.

The algebraic multiplicity of an eigenvalue λ\lambda is said be its multiplicity as a root of the characteristic polynomial. It is well known that for 22-dimensional matrices geometric and algebraic multiplicities coincide. Unfortunately, in the multidimensional case it is not true. For more information on relations between these multiplicities see [HuYe.multeigen].

The study of eigenvalues of multidimensional matrices and hypergraphs grows now extensively, so we mention here only several papers. In [CoopDut.hyperspec] Cooper and Dutle proved some properties of hypergraph spectra and calculate eigenvalues of certain classes of hypergraphs. In [shao.tensprod] it was introduced several products of hypergraphs and multidimensional matrices and then considered the eigenvalues of the result. Algebraic properties of eigenvalues and determinats of multidimensional matrices with respect to various matrix operations were studied in [ShaoShanZhang.detoftens]. At last, for some survey on the spectral theory of nonnegative matrices see [ChaQiZhang.survspecthyper].

2.2 Perfect colorings and coverings of graphs

Let us review some basic definitions and facts on perfect colorings of graphs that we aim to generalize for hypergraphs. Most of them can be found in [my.perfstrct].

A coloring of a graph G=(X,E)G=(X,E) in kk colors is a surjective function f:X{1,,k}f:X\rightarrow\{1,\ldots,k\}. To every kk-coloring ff, we can put into a correspondence a rectangular color matrix PP of size |X|×k|X|\times k with entries px,i=1p_{x,i}=1 if f(x)=if(x)=i and px,i=0p_{x,i}=0 otherwise. Note that each row of PP contains exactly one nonzero entry. A coloring of a graph parts its vertex set XX into disjoint color classes {C1,,Ck}\{C_{1},\ldots,C_{k}\}, where Ci={x:f(x)=i}C_{i}=\{x:f(x)=i\}. In what follows, we usually define a coloring with the help of the color matrix PP, but in some cases we also refer to the coloring function ff and the color classes {C1,,Ck}\{C_{1},\ldots,C_{k}\}.

If GG is a bipartite graph and ff is a coloring of GG, we will say that ff is a bipartite coloring if there are no color classes of ff that intersect both parts of GG.

A kk-coloring ff of a graph GG is called perfect if there exist integer si,js_{i,j}, i,j{1,,k}i,j\in\{1,\ldots,k\}, such that every vertex of color ii in ff is adjacent to exactly si,js_{i,j} vertices of color jj. The matrix S=(si,j)S=(s_{i,j}) of order kk is called the parameter matrix of the perfect coloring.

It is not hard to see that if AA is the adjacency matrix of a graph GG, then PP is a perfect coloring with the parameter matrix SS if and only if AP=PSAP=PS. Moreover, if yy is an eigenvector of the parameter matrix SS with an eigenvalue λ\lambda, then PyPy is the eigenvector of AA with the same eigenvalue λ\lambda.

Given a coloring PP, let us denote by NN the matrix PTPP^{T}P. It can be checked that NN is a diagonal matrix with entries ni,in_{i,i} equal to the number of vertices of color ii and that the matrix NS=PTPSNS=P^{T}PS is symmetric.

We will say that a coloring ff of a graph GG with color classes {C1,,Ck}\{C_{1},\ldots,C_{k}\} is a refinement of a coloring gg with color classes {D1,,Dl}\{D_{1},\ldots,D_{l}\} if each color class DiD_{i} is a union of some classes CjC_{j}. For a given graph GG, its colorings (and perfect colorings) compose a partially ordered set with respect to the refinement.

For every coloring of a graph, there is the unique coarsest refinement that will be a perfect coloring. The method of finding such refinements is known as Weisfeiler–Leman algorithm [WeiLeh.algo].

Theorem 2 ([WeiLeh.algo]).

Given a graph GG, there is a perfect coloring ff such that every other perfect coloring of GG is a refinement of ff. Moreover, for every coloring gg of GG there is a refinement hh of gg such that hh is a perfect coloring, and any other perfect coloring that refines gg is a refinement of hh.

A graph G=(X,E)G=(X,E) covers a graph H=(Y,U)H=(Y,U) if there exists a surjective function φ:XY\varphi:X\rightarrow Y such that for each xXx\in X the equality {φ(x)|(x,x)E}={y|(y,φ(x))U}\{\varphi(x^{\prime})|(x,x^{\prime})\in E\}=\{y|(y,\varphi(x))\in U\} holds. A covering φ\varphi of a graph HH by GG is a kk-covering if for every yYy\in Y there are exactly kk vertices xXx\in X such that φ(x)=y\varphi(x)=y. It is well known that every covering is a kk-covering for some kk\in\mathbb{N}.

It is not hard to see that a graph GG covers a graph HH if and only if there is a perfect coloring of GG with the parameter matrix equal to the adjacency matrix of HH. It implies, for example, the following simple properties of coverings:

  • If a graph GG covers a graph HH and a graph HH covers a graph FF, then GG covers FF.

  • If a graph GG covers a graph HH and λ\lambda is an eigenvalue of HH, then λ\lambda is an eigenvalue of GG.

In [leighton.comcover, AngGard.comcover], it was proved that if graphs H1H_{1} and H2H_{2} have perfect colorings with the same parameter matrices, then there exists a graph GG that covers both H1H_{1} and H2H_{2}.

For more information of graph coverings and their applications see [GodRoy.alggrath].

3 Perfect colorings of hypergraphs

There are two equivalent ways to define perfect colorings in uniform hypergraphs. The first of them uses a perfect coloring of the incidence graph, and the second one is based on an equation for the multidimensional adjacency matrix of a hypergraph. The first approach is more general and can be applied to non-uniform hypergraphs. Although it appeared in [GraGroLo.handbook] and [PotAv.hypercolor], the parameters of such colorings were not discussed before. The approach based on multidimensional matrices is completely new. In this section we also study the interrelations between these approaches.

3.1 Colorings of the incidence graph

Let 𝒢(X,E)\mathcal{G}(X,E) be a hypergraph. We will say that a surjective function f:X{1,,k}f:X\rightarrow\{1,\ldots,k\} is a coloring of 𝒢\mathcal{G} into kk colors (or kk-coloring). In other words, ff defines a partition of the set XX in kk color classes. Given a coloring ff and a hyperedge ee, let the color range f(e)f(e) be the multiset of colors of all incident vertices: f(e)={f(x)|xe}f(e)=\{f(x)|x\in e\}.

Let GG be the incidence graph of the hypergraph 𝒢\mathcal{G}. To each coloring ff of 𝒢\mathcal{G}, we associate an induced coloring gg of GG such that for every xXx\in X we have g(x)=f(x)g(x)=f(x) and for every eEe\in E we define g(e)=f(e)g(e)=f(e). Here the color ranges f(e)f(e) of hyperedges are considered as colors of the coloring gg. Note that every induced coloring of GG is bipartite, but not every bipartite coloring of GG is induced by some coloring of 𝒢\mathcal{G}.

A coloring ff of a hypergraph 𝒢\mathcal{G} is said to be perfect if each two vertices xx and yy of the same color have the same set of color ranges of incident hyperedges. Directly from the definitions, it follows that ff is a perfect coloring of 𝒢\mathcal{G} if and only if the induced bipartite coloring gg of GG is perfect.

We will say that a rectangular (0,1)(0,1)-matrix PP is a color matrix if every row of PP contains exactly one 11. To every kk-coloring ff of a hypergraph 𝒢(X,E)\mathcal{G}(X,E) on nn vertices, we associate the color matrix PP of size n×kn\times k such that px,i=1p_{x,i}=1 if f(x)=if(x)=i and px,i=0p_{x,i}=0 otherwise.

Assume that a hypergraph 𝒢(X,E)\mathcal{G}(X,E) has nn vertices, mm hyperedges, and the incidence matrix BB. Let ff be a perfect kk-coloring of 𝒢\mathcal{G} such that the number of different color ranges γ\gamma of hyperedges is equal to ll. Then the induced coloring gg of the incidence graph GG gives us the following matrix equation:

(0BBT0)(0PR0)=(0PR0)(0WV0)\left(\begin{array}[]{cc}0&B\\ B^{T}&0\end{array}\right)\left(\begin{array}[]{cc}0&P\\ R&0\end{array}\right)=\left(\begin{array}[]{cc}0&P\\ R&0\end{array}\right)\left(\begin{array}[]{cc}0&W\\ V&0\end{array}\right)

or, equivalently,

BR=PV;BTP=RW,BR=PV;\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ B^{T}P=RW, (1)

where

  • PP and RR are vertex and hyperedge color matrices of sizes n×kn\times k and m×lm\times l, respectively; the coloring RR is induced by the coloring PP.

  • VV and WW are matrices of sizes k×lk\times l and l×kl\times k, respectively. An entry vi,γv_{i,\gamma} of VV is equal to the number of hyperedges of color range γ\gamma in 𝒢\mathcal{G} incident to a vertex of color ii, entry wγ,iw_{\gamma,i} of WW is equal to the number of vertices of color ii in 𝒢\mathcal{G} contained in a hyperedge of color range γ\gamma.

We will say that the pair (V,W)(V,W) is the incidence parameters of the perfect coloring ff: the matrix VV is said to be the XE-parameter matrix (describing incidence of vertices xx to hyperedges ee), and WW is the EX-parameter matrix (describing incidence of hyperedges ee to vertices xx).

Example 1. Let 𝒢\mathcal{G} be a 33-uniform hypergraph with the vertex set X={x1,x2,x3,x4,x5,x6}X=\{x_{1},x_{2},x_{3},x_{4},x_{5},x_{6}\} and the hyperedge set E={e1,e2,e3,e4}E=\{e_{1},e_{2},e_{3},e_{4}\}, where e1={x1,x2,x3}e_{1}=\{x_{1},x_{2},x_{3}\}, e2={x1,x4,x5}e_{2}=\{x_{1},x_{4},x_{5}\}, e3={x2,x4,x6}e_{3}=\{x_{2},x_{4},x_{6}\}, e4={x3,x5,x6}e_{4}=\{x_{3},x_{5},x_{6}\}.

Let ff be a coloring of XX into two colors such that f(x1)=f(x2)=f(x3)=1f(x_{1})=f(x_{2})=f(x_{3})=1 and f(x4)=f(x5)=f(x6)=2f(x_{4})=f(x_{5})=f(x_{6})=2. Then the color ranges of hyperedges are f(e1)={1,1,1}f(e_{1})=\{1,1,1\} and f(e2)=f(e3)=f(e4)={1,2,2}f(e_{2})=f(e_{3})=f(e_{4})=\{1,2,2\}. Note that each vertex of color 11 is incident to one hyperedge of color range {1,1,1}\{1,1,1\} and one hyperedge of color range {1,2,2}\{1,2,2\}, and each vertex of color 22 is incident to two hyperedges of color range {1,2,2}\{1,2,2\}. Therefore, ff is a perfect coloring of 𝒢\mathcal{G}.

The incidence graph GG of 𝒢\mathcal{G} with the induced perfect coloring gg is given at Figure 1.

[Uncaptioned image]

Figure 1: The perfect coloring gg of the incidence graph GG for the hypergraph 𝒢\mathcal{G}

The incidence matrix of the hypergraph GG is

B=(110010101001011001010011).B=\left(\begin{array}[]{cccc}1&1&0&0\\ 1&0&1&0\\ 1&0&0&1\\ 0&1&1&0\\ 0&1&0&1\\ 0&0&1&1\\ \end{array}\right).

By the definition, the vertex color matrix PP and the hyperedge color matrix RR for the perfect coloring ff are

P=(101010010101);R=(10010101),P=\left(\begin{array}[]{cc}1&0\\ 1&0\\ 1&0\\ 0&1\\ 0&1\\ 0&1\end{array}\right);\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ R=\left(\begin{array}[]{cc}1&0\\ 0&1\\ 0&1\\ 0&1\end{array}\right),

and the incidence parameters (V,W)(V,W) are

V=(1102);W=(3012).V=\left(\begin{array}[]{cc}1&1\\ 0&2\end{array}\right);\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ W=\left(\begin{array}[]{cc}3&0\\ 1&2\end{array}\right).

It can be checked directly that BR=PVBR=PV and BTP=RWB^{T}P=RW.

Let us derive several simple observations from the given definition of a perfect coloring of hypergraphs.

First of all, if vertices xx and yy of a hypergraph 𝒢\mathcal{G} have different degrees, then f(x)f(y)f(x)\neq f(y) for every perfect coloring ff of 𝒢\mathcal{G} because the same is true for perfect colorings of graphs. The sum of entries in the ii-th row of the XE-parameter matrix VV is equal to the degree of vertices of color ii, and the sum of entries in the γ\gamma-th row of the EX-parameter matrix WW is equal to the size of a hyperedge of the color range γ\gamma.

Next, if PP is a coloring of vertices and RR is a coloring of hyperedges, then N=PTPN=P^{T}P and M=RTRM=R^{T}R are diagonal matrices. Entries ni,in_{i,i} of NN are equal to the numbers of vertices colored by color ii in PP, and entries mj,jm_{j,j} of MM are the numbers of hyperedges colored by color jj in RR.

We also prove the following results on the structure of the incidence matrix BB and a relation between the incidence parameters (V,W)(V,W) and matrices NN and MM.

Theorem 3.
  1. 1.

    Let 𝒢\mathcal{G} be a hypergraph with the incidence matrix BB, PP be a coloring of 𝒢\mathcal{G}, RR be an induced coloring of hyperedges, N=PTPN=P^{T}P and M=RTRM=R^{T}R be the diagonal matrices with entries nin_{i} and mjm_{j} at the diagonal. If a coloring PP is perfect for 𝒢\mathcal{G} with incidence parameters (V,W)(V,W), then NV=WTMNV=W^{T}M and BB is a block matrix {Ai,j}\{A_{i,j}\}, where Ai,jA_{i,j} are (0,1)(0,1)-matrices of sizes ni×mjn_{i}\times m_{j} with row sums vi,jv_{i,j} and column sums wj,iw_{j,i}.

  2. 2.

    Let VV and WW be integer nonnegative matrices and NN and MM be integer diagonal matrices with ni>0n_{i}>0 and mj>0m_{j}>0 such that NV=WTMNV=W^{T}M. Then there exist a hypergraph 𝒢\mathcal{G} and its perfect coloring PP having incidence parameters (V,W)(V,W).

Proof.

(1). Assume that PP is a perfect coloring of 𝒢\mathcal{G}. Recall that equations (1) for perfect colorings give

BR=PV;BTP=RW.BR=PV;\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ B^{T}P=RW.

Multiplying the first equation by PTP^{T} and the second equation by RTR^{T}, we have

PTBR=PTPV=NV;RTBTP=RTRW=MW.P^{T}BR=P^{T}PV=NV;\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ R^{T}B^{T}P=R^{T}RW=MW.

Note that the left-hand side of one equation is the transposition of the other, so NV=WTMNV=W^{T}M. On the other hand, these equations mean that matrices PP and RR define a partition of the incidence matrix BB into blocks Ai,jA_{i,j}, where Ai,jA_{i,j} has sizes ni×mjn_{i}\times m_{j}, row sums vi,jv_{i,j}, and column sums wj,iw_{j,i}.

(2). Equality NV=WTMNV=W^{T}M means that nivi,j=wj,imjn_{i}v_{i,j}=w_{j,i}m_{j} for all ii and jj. Choose a nonnegative integer tt so that for all ii and jj there exist (0,1)(0,1)-matrices Ai,jA_{i,j} of sizes tni×tmjtn_{i}\times tm_{j} such that each row of Ai,jA_{i,j} contains exactly vi,jv_{i,j} ones and each column contains exactly wj,iw_{j,i} ones.

Let a rectangular matrix BB is given by the block matrix {Ai,j}\{A_{i,j}\}. We treat BB as the incidence matrix of a hypergraph 𝒢\mathcal{G}. It can be verified directly that a kk-coloring of 𝒢\mathcal{G}

P=(10100101),P=\left(\begin{array}[]{ccc}1&\cdots&0\\ \vdots&\vdots&\vdots\\ 1&\cdots&0\\ \vdots&\vdots&\vdots\\ 0&\cdots&1\\ \vdots&\vdots&\vdots\\ 0&\cdots&1\\ \end{array}\right),

where the ii-th column contains exactly tnitn_{i} ones, is a perfect coloring with the incidence parameters (V,W)(V,W). ∎

At last, we define refinements of perfect colorings in hypergraphs similar to graphs. We will say that a coloring ff of a hypergraph 𝒢\mathcal{G} is a refinement of a coloring gg if the coloring of the incidence graph of 𝒢\mathcal{G} induced by ff is a refinement of the coloring induced by gg. The set of perfect colorings of a hypergraph with respect to the refinement forms a partially ordered set. Theorem 2 implies the following.

Theorem 4.

Given a hypergraph 𝒢=(X,E)\mathcal{G}=(X,E), there is a perfect coloring ff of 𝒢\mathcal{G} such that any other perfect coloring is a refinement of ff. Moreover, for every coloring gg of 𝒢\mathcal{G} there is a refinement hh of gg such that hh is a perfect coloring and any other perfect coloring that refines gg is a refinement of hh.

Proof.

To construct the coloring ff, it is sufficient to apply Theorem 2 to the trivial bipartite coloring of the incidence graph GG, in which the part XX is colored by one color, and the part EE is colored by another color.

The second statement of the theorem is obtained by the application of Theorem 2 to the coloring of the incidence graph induced by the coloring gg. ∎

We will say that the coloring ff from Theorem 4 is the minimal perfect coloring of the hypergraph 𝒢\mathcal{G}.

3.2 The multidimensional matrix equation for perfect colorings

Let 𝒢(X,E)\mathcal{G}(X,E) be a dd-uniform hypergraph on nn vertices, 𝔸\mathbb{A} be the dd-dimensional adjacency matrix of 𝒢\mathcal{G}, and PP be a coloring of 𝒢\mathcal{G} into kk colors (color matrix of vertices). We will say that PP is perfect if there exists a dd-dimensional matrix 𝕊\mathbb{S} of order kk such that

𝔸P=P𝕊,\mathbb{A}\circ P=P\circ\mathbb{S},

where the product \circ of multidimensional matrices is defined in Section 2. The matrix 𝕊\mathbb{S} is called the parameter matrix of the perfect coloring PP.

Let us show that this definition of perfect colorings in case of uniform hypergraphs is equivalent to the previous one. Moreover, we express entries of the parameter matrix 𝕊\mathbb{S} by the incidence parameters (V,W)(V,W).

Theorem 5.

Let 𝒢\mathcal{G} be dd-uniform hypergraph with the dd-dimensional adjacency matrix 𝔸\mathbb{A}. Then a coloring PP of 𝒢\mathcal{G} into kk colors is a perfect if and only if 𝔸P=P𝕊\mathbb{A}\circ P=P\circ\mathbb{S}. Moreover, the entries sγs_{\gamma}, γ=(γ1,γ2,,γd)\gamma=(\gamma_{1},\gamma_{2},\ldots,\gamma_{d}), of 𝕊\mathbb{S} are

sγ=vγ1,γ(d1d1,,dk)1,s_{\gamma}=v_{\gamma_{1},\gamma}\cdot{d-1\choose d_{1},\ldots,d_{k}}^{-1},

where vγ1,γv_{\gamma_{1},\gamma} is the number of hyperedges of color range γ\gamma incident to a vertex of color γ1\gamma_{1}, every color ll appears in the multiset {γ2,,γd}\{\gamma_{2},\ldots,\gamma_{d}\} exactly dld_{l} times, and (d1d1,,dk){d-1\choose d_{1},\ldots,d_{k}} is the multinomial coefficient.

Proof.

\Leftarrow: Suppose that ff is a kk-coloring of the hypergraph 𝒢\mathcal{G} with a color matrix PP such that

𝔸P=P𝕊.\mathbb{A}\circ P=P\circ\mathbb{S}.

Let us consider entries of the matrix =𝔸P\mathbb{C}=\mathbb{A}\circ P. Using the definition of the multidimensional matrix product, for a given vertex xx and colors j1,,jd1j_{1},\ldots,j_{d-1} we have

cx,j1,,jd1=x1,,xd1=1nax,x1,,xd1px1,j1pxd1,jd1.c_{x,j_{1},\ldots,j_{d-1}}=\sum\limits_{x_{1},\ldots,x_{d-1}=1}^{n}a_{x,x_{1},\ldots,x_{d-1}}p_{x_{1},j_{1}}\cdots p_{x_{d-1},j_{d-1}}. (2)

Note that an entry ax,x1,,xd1a_{x,x_{1},\ldots,x_{d-1}} of 𝔸\mathbb{A} is equal to 1(d1)!\frac{1}{(d-1)!} if and only if x,x1,,xd1x,x_{1},\ldots,x_{d-1} are different vertices of 𝒢\mathcal{G} and constitute a hyperedge, and the entries pxi,ji=1p_{x_{i},j_{i}}=1 if and only if f(xi)=jif(x_{i})=j_{i}. Let tx,j1,,jd1t_{x,j_{1},\ldots,j_{d-1}} be the number of hyperedges (x,x1,,xd1)(x,x_{1},\ldots,x_{d-1}) such that {f(x1),,f(xd1)}\{f(x_{1}),\ldots,f(x_{d-1})\} and {j1,,jd1}\{j_{1},\ldots,j_{d-1}\} coincide as multisets. We also assume that each color ll appears in these multisets exactly dld_{l} times, l=1kdl=d1\sum\limits_{l=1}^{k}d_{l}=d-1. Then equation (2) is equivalent to cx,j1,,jd1=tx,j1,,jd1(d1d1,,dk)1c_{x,j_{1},\ldots,j_{d-1}}=t_{x,j_{1},\ldots,j_{d-1}}{d-1\choose d_{1},\ldots,d_{k}}^{-1}, because every ordering of vertices x1,,xd1x_{1},\ldots,x_{d-1} that does not change the sequence of colors f(x1),,f(xd1)f(x_{1}),\ldots,f(x_{d-1}) gives a summand 1(d1)!\frac{1}{(d-1)!} in counting of cx,j1,,jd1c_{x,j_{1},\ldots,j_{d-1}}.

On the other hand, using =P𝕊\mathbb{C}=P\circ\mathbb{S} and again the definition of \circ-operation, we have

cx,j1,,jd1=i=1kpx,isi,j1,,jd1.c_{x,j_{1},\ldots,j_{d-1}}=\sum\limits_{i=1}^{k}p_{x,i}s_{i,j_{1},\ldots,j_{d-1}}.

Since px,i=1p_{x,i}=1 if and only if f(x)=if(x)=i, we have that if a vertex xx has color ii, then cx,j1,,jd1=si,j1,,jd1c_{x,j_{1},\ldots,j_{d-1}}=s_{i,j_{1},\ldots,j_{d-1}} and, consequently, si,j1,,jd1=tx,j1,,jd1(d1d1,,dk)1s_{i,j_{1},\ldots,j_{d-1}}=t_{x,j_{1},\ldots,j_{d-1}}{d-1\choose d_{1},\ldots,d_{k}}^{-1}. Therefore, the color ranges {i,j1,,jd1}\{i,j_{1},\ldots,j_{d-1}\} of hyperedges incident to a vertex xx of color ii are uniquely defined by entries of 𝕊\mathbb{S}, and so the coloring PP is perfect. Replacing vi,(i,j1,,jd1)=tx,j1,,jd1v_{i,(i,j_{1},\ldots,j_{d-1})}=t_{x,j_{1},\ldots,j_{d-1}} for all vertices xx of color ii, we have the statement of the theorem.

\Rightarrow: Assume that a coloring PP is perfect. Repeating the calculation of entries of the matrix 𝔸P\mathbb{A}\circ P and using the fact that the values tx,j1,,jd1t_{x,j_{1},\ldots,j_{d-1}} are uniquely defined by the color of vertex xx, we see that there exists the required dd-dimensional matrix 𝕊\mathbb{S} satisfying 𝔸P=P𝕊\mathbb{A}\circ P=P\circ\mathbb{S}.

Note that a perfect coloring PP defines only the factor vγ1,γv_{\gamma_{1},\gamma} in the entries sγs_{\gamma} of the matrix 𝕊\mathbb{S}, while the factor (d1d1,,dk)1{d-1\choose d_{1},\ldots,d_{k}}^{-1} in sγs_{\gamma} depends only on the index γ\gamma and it is the same for all dd-dimensional parameter matrices 𝕊\mathbb{S}.

Corollary 1.

Let 𝒢\mathcal{G} be a dd-uniform hypergraph, PP be a perfect coloring of 𝒢\mathcal{G}, nln_{l} be the number of vertices of color ll, and mγm_{\gamma} be the number of hyperedges of a color range γ\gamma. Suppose that W=(wγ,l)W=(w_{\gamma,l}) is the EX-parameter matrix and 𝕊\mathbb{S} is the multidimensional parameter matrix of the perfect coloring PP. Then entries sγs_{\gamma} of 𝕊\mathbb{S} are

sγ=dmγnγ1(dwγ,1,,wγ,k)1.s_{\gamma}=d\cdot\frac{m_{\gamma}}{n_{\gamma_{1}}}\cdot{d\choose w_{\gamma,1},\ldots,w_{\gamma,k}}^{-1}.
Proof.

By Theorem 5, the entries sγs_{\gamma} of the parameter matrix 𝕊\mathbb{S} are equal to vγ1,γ(d1d1,,dk)1v_{\gamma_{1},\gamma}\cdot{d-1\choose d_{1},\ldots,d_{k}}^{-1}, where vγ1,γv_{\gamma_{1},\gamma} are entries of the XEXE-parameter matrix VV and each color ll appears in {γ2,,γd}\{\gamma_{2},\ldots,\gamma_{d}\} exactly dld_{l} times. By the definition of incidence parameters, γ={γ1,,γd}\gamma=\{\gamma_{1},\ldots,\gamma_{d}\} contains the ll-th color exactly wγ,lw_{\gamma,l} times, so dγ1=wγ,γ11d_{\gamma_{1}}=w_{\gamma,\gamma_{1}}-1 and dγi=wγ,γid_{\gamma_{i}}=w_{\gamma,\gamma_{i}} for all other γi\gamma_{i}.

By Theorem 3(1), nγ1vγ1,γ=wγ,γ1mγn_{\gamma_{1}}v_{\gamma_{1},\gamma}=w_{\gamma,\gamma_{1}}m_{\gamma}. Consequently,

sγ=vγ1,γd1!dk!(d1)!=vγ1,γwγ,γ1wγ,1!wγ,k!(d1)!=dmγnγ1(dwγ,1,,wγ,k)1.s_{\gamma}=v_{\gamma_{1},\gamma}\cdot\frac{d_{1}!\cdots d_{k}!}{(d-1)!}=\frac{v_{\gamma_{1},\gamma}}{w_{\gamma,\gamma_{1}}}\cdot\frac{w_{\gamma,1}!\cdots w_{\gamma,k}!}{(d-1)!}=d\cdot\frac{m_{\gamma}}{n_{\gamma_{1}}}\cdot{d\choose w_{\gamma,1},\ldots,w_{\gamma,k}}^{-1}.

Example 2. Let 𝒢\mathcal{G} be the 33-uniform hypergraph on 66 vertices from Example 1 and PP be its perfect coloring:

P=(101010010101)P=\left(\begin{array}[]{cc}1&0\\ 1&0\\ 1&0\\ 0&1\\ 0&1\\ 0&1\end{array}\right)

Recall that the incidence parameters (V,W)(V,W) for this coloring are

V=(1102);W=(3012).V=\left(\begin{array}[]{cc}1&1\\ 0&2\end{array}\right);\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ W=\left(\begin{array}[]{cc}3&0\\ 1&2\end{array}\right).

The adjacency matrix 𝔸\mathbb{A} of the hypergraph 𝒢\mathcal{G} is the following 33-dimensional matrix of order 66:

𝔸=12(000000001000010000001000000000100000010000100000000000000010000001000000000100000000000001000000000100000010\displaystyle\mathbb{A}=\frac{1}{2}\left(\begin{array}[]{cccccc|cccccc|cccccc|}0&0&0&0&0&0&0&0&1&0&0&0&0&1&0&0&0&0\\ 0&0&1&0&0&0&0&0&0&0&0&0&1&0&0&0&0&0\\ 0&1&0&0&0&0&1&0&0&0&0&0&0&0&0&0&0&0\\ 0&0&0&0&1&0&0&0&0&0&0&1&0&0&0&0&0&0\\ 0&0&0&1&0&0&0&0&0&0&0&0&0&0&0&0&0&1\\ 0&0&0&0&0&0&0&0&0&1&0&0&0&0&0&0&1&0\\ \end{array}\right.
000010000100000000000001000000000100000000000001000010000000100000010000100000000000001000010000001000000000).\displaystyle\left.\begin{array}[]{|cccccc|cccccc|cccccc}0&0&0&0&1&0&0&0&0&1&0&0&0&0&0&0&0&0\\ 0&0&0&0&0&1&0&0&0&0&0&0&0&0&0&1&0&0\\ 0&0&0&0&0&0&0&0&0&0&0&1&0&0&0&0&1&0\\ 0&0&0&0&0&0&1&0&0&0&0&0&0&1&0&0&0&0\\ 1&0&0&0&0&0&0&0&0&0&0&0&0&0&1&0&0&0\\ 0&1&0&0&0&0&0&0&1&0&0&0&0&0&0&0&0&0\\ \end{array}\right).

It can be checked directly that 𝔸P=P𝕊\mathbb{A}\circ P=P\circ\mathbb{S} for the 33-dimensional parameter matrix

𝕊=(10010110),\mathbb{S}=\left(\begin{array}[]{cc|cc}1&0&0&1\\ 0&1&1&0\end{array}\right),

where s1,1,1=v1,1/1=1s_{1,1,1}=v_{1,1}/1=1, s1,2,2=v1,2/1=1s_{1,2,2}=v_{1,2}/1=1, s2,2,1=s2,1,2=v2,2/2=1s_{2,2,1}=s_{2,1,2}=v_{2,2}/2=1. Note that s1,1,2=s1,2,1=s2,1,1=s2,2,2=0s_{1,1,2}=s_{1,2,1}=s_{2,1,1}=s_{2,2,2}=0 because there are no hyperedges of color ranges {1,1,2}\{1,1,2\} and {2,2,2}\{2,2,2\} in the perfect coloring PP.

Using the obtained results, let us prove several properties of the parameter matrix 𝕊\mathbb{S}.

Proposition 2.

Let PP be a perfect coloring in a dd-uniform hypergraph 𝒢\mathcal{G} with the parameter matrix 𝕊\mathbb{S}. Then the sum of entries of 𝕊\mathbb{S} along the ii-th hyperplane of the first direction is equal to the degree of vertices of color ii.

Proof.

By Theorem 5, the entries sγs_{\gamma} of the parameter matrix 𝕊\mathbb{S} are equal to vγ1,γ(d1d1,,dk)1v_{\gamma_{1},\gamma}\cdot{d-1\choose d_{1},\ldots,d_{k}}^{-1}, where vγ1,γv_{\gamma_{1},\gamma} is the number of hyperedges of color range γ\gamma incident to a vertex of color γ1\gamma_{1}. Note that every hyperedge with color range γ\gamma gives exactly (d1d1,,dk){d-1\choose d_{1},\ldots,d_{k}} nonzero entries sγs_{\gamma} in the γ1\gamma_{1}-th hyperplane of the first direction for the parameter matrix 𝕊\mathbb{S}. So the total contribution of such hyperedges to this sum is vγ1,γv_{\gamma_{1},\gamma}, and the total sum of entries along this hyperplane is the degree of a vertex of color γ1\gamma_{1}. ∎

Theorem 6.

Let 𝒢\mathcal{G} be a dd-uniform hypergraph. Assume that PP is a perfect coloring of 𝒢\mathcal{G} with the parameter matrix 𝕊\mathbb{S} and N=PTPN=P^{T}P. Then the matrix =N𝕊\mathbb{H}=N\circ\mathbb{S} is symmetric.

Proof.

Recall that N=PTPN=P^{T}P is a diagonal matrix whose ii-th diagonal entry is the number nin_{i} of vertices of color ii in the coloring PP.

By Corollary 1, an entry sγs_{\gamma} of the matrix 𝕊\mathbb{S} is equal to dmγnγ1(dwγ,1,,wγ,k)1d\cdot\frac{m_{\gamma}}{n_{\gamma_{1}}}\cdot{d\choose w_{\gamma,1},\ldots,w_{\gamma,k}}^{-1}, where wγ,lw_{\gamma,l} is the number of appearances of symbol ll in γ\gamma and mγm_{\gamma} is the number of hyperedges of 𝒢\mathcal{G} having the color range γ\gamma.

Using the definition of the product of multidimensional matrices, we see that =N𝕊\mathbb{H}=N\circ\mathbb{S} is a dd-dimensional matrix with entries

hβ=j=1knβ1,jsj,β2,,βd=dmβ(dwβ,1,,wβ,k)1.h_{\beta}=\sum\limits_{j=1}^{k}n_{\beta_{1},j}\cdot s_{j,\beta_{2},\ldots,\beta_{d}}=d\cdot m_{\beta}\cdot{d\choose w_{\beta,1},\ldots,w_{\beta,k}}^{-1}.

Since for every permutation σSd\sigma\in S_{d} it holds hβ=hσ(β)h_{\beta}=h_{\sigma(\beta)}, we have that the matrix \mathbb{H} is symmetric.

It is well known that for graphs the spectrum of the parameter matrix of a perfect coloring is contained in the spectrum of the adjacency matrix. A similar fact (regarding geometric multiplicities) is true for hypergraphs. Moreover, this property holds not only for perfect colorings of hypergraphs, but for general matrices satisfying the same equation. To prove this statement, we need the following auxiliary result.

Proposition 3.

Let PP be a color matrix of size n×kn\times k and 𝕀\mathbb{I} be the dd-dimensional identity matrix. Then

𝕀P=P𝕀.\mathbb{I}\circ P=P\circ\mathbb{I}.
Proof.

The proof is based on the definitions of \circ-operation and matrices 𝕀\mathbb{I} and PP.

Let us show that (α1,α2,,αd)(\alpha_{1},\alpha_{2},\ldots,\alpha_{d})-th entries of matrices in left-hand and right-hand sides of the equation 𝕀P=P𝕀\mathbb{I}\circ P=P\circ\mathbb{I} are equal to 11 if α2==αd\alpha_{2}=\cdots=\alpha_{d} and pα1,α2=1p_{\alpha_{1},\alpha_{2}}=1 and they are 0 otherwise.

Indeed, an (α1,α2,,αd)(\alpha_{1},\alpha_{2},\ldots,\alpha_{d})-th entry of the matrix 𝕀P\mathbb{I}\circ P is

β2,,βd=1niα1,β2,,βdpβ2,α2pβd,αd.\sum\limits_{\beta_{2},\ldots,\beta_{d}=1}^{n}i_{\alpha_{1},\beta_{2},\ldots,\beta_{d}}\cdot p_{\beta_{2},\alpha_{2}}\cdots p_{\beta_{d},\alpha_{d}}.

Note that iα1,β2,,βd=1i_{\alpha_{1},\beta_{2},\ldots,\beta_{d}}=1 if and only if α1=β2==βd\alpha_{1}=\beta_{2}=\cdots=\beta_{d}. Otherwise, iα1,β2,,βd=0i_{\alpha_{1},\beta_{2},\ldots,\beta_{d}}=0. By the definition of a color matrix, there is a unique jj such that pα1,j=1p_{\alpha_{1},j}=1.

On the other hand, an (α1,α2,,αd)(\alpha_{1},\alpha_{2},\ldots,\alpha_{d})-th entry of the matrix P𝕀P\circ\mathbb{I} is

j=1kpα1,jij,α2,,αd.\sum\limits_{j=1}^{k}p_{\alpha_{1},j}\cdot i_{j,\alpha_{2},\ldots,\alpha_{d}}.

Again, there is a unique jj such that pα1,j=1p_{\alpha_{1},j}=1 and ij,α2,,αd=1i_{j,\alpha_{2},\ldots,\alpha_{d}}=1 if and only if j=α2==αdj=\alpha_{2}=\cdots=\alpha_{d}. ∎

Theorem 7.

Let 𝔸\mathbb{A} be a dd-dimensional matrix of order nn, PP be a color matrix of size n×kn\times k, and 𝔹\mathbb{B} be a dd-dimensional matrix of order kk such that 𝔸P=P𝔹\mathbb{A}\circ P=P\circ\mathbb{B}. If (λ,x)(\lambda,x) is an eigenpair for the matrix 𝔹\mathbb{B}, then (λ,Px)(\lambda,Px) is an eigenpair for 𝔸\mathbb{A}.

Proof.

By the definition, (λ,x)(\lambda,x) is an eigenpair for the matrix 𝔹\mathbb{B} if and only if 𝔹x=λ(𝕀x).\mathbb{B}\circ x=\lambda(\mathbb{I}\circ x). Using Proposition 1 and the equality P𝕀=𝕀PP\circ\mathbb{I}=\mathbb{I}\circ P from Proposition 3, we conclude that

𝔸(Px)=𝔸(Px)=(𝔸P)x=(P𝔹)x=P(𝔹x)=\displaystyle\mathbb{A}\circ(Px)=\mathbb{A}\circ(P\circ x)=(\mathbb{A}\circ P)\circ x=(P\circ\mathbb{B})\circ x=P\circ(\mathbb{B}\circ x)=
P(λ𝕀x)=λ(P𝕀)x=λ(𝕀P)x=λ𝕀(Px)=λ𝕀(Px).\displaystyle P\circ(\lambda\mathbb{I}\circ x)=\lambda(P\circ\mathbb{I})\circ x=\lambda(\mathbb{I}\circ P)\circ x=\lambda\mathbb{I}\circ(P\circ x)=\lambda\mathbb{I}\circ(Px).

This theorem easily implies that all eigenvalues of the parameter matrix of a perfect coloring are also eigenvalues of the adjacency matrix of a hypergraph.

Theorem 8.

Let 𝒢\mathcal{G} be a dd-uniform hypergraph with the adjacency matrix 𝔸\mathbb{A} and PP be a perfect coloring with the dd-dimensional parameter matrix 𝕊\mathbb{S}. If (λ,x)(\lambda,x) is an eigenpair for the matrix 𝕊\mathbb{S}, then (λ,Px)(\lambda,Px) is an eigenpair for the matrix 𝔸\mathbb{A}.

Proof.

It is sufficient to note that 𝔸P=P𝕊\mathbb{A}\circ P=P\circ\mathbb{S} and apply Theorem 7. ∎

Corollary 2.

Assume that PP is a perfect coloring of a dd-uniform hypergraph 𝒢\mathcal{G} into kk colors with the parameter matrix 𝕊\mathbb{S} and λ\lambda is an eigenvalue of 𝕊\mathbb{S}. Then there is an eigenvector yy of 𝒢\mathcal{G} for eigenvalue λ\lambda such that components of yy attain at most kk different values.

Proof.

Let xx be an eigenvector of 𝕊\mathbb{S} for the eigenvalue λ\lambda. Set y=Pxy=Px and repeat the proof of Theorem 7. ∎

4 Coverings of hypergraphs

We will say that a hypergraph 𝒢\mathcal{G} covers a hypergraph \mathcal{H} if there is a map (covering) φ:X(𝒢)X()\varphi:X(\mathcal{G})\rightarrow X(\mathcal{H}) such that for each hyperedge eE(𝒢)e\in E(\mathcal{G}) the set {φ(x)|xe}\{\varphi(x)|x\in e\} is a hyperedge of \mathcal{H} and φ\varphi saves a collection of incident hyperedges for each vertex.

Note that a covering preserves degrees of vertices and sizes of hyperedges, so every covering of a dd-uniform rr-regular hypergraph is also dd-uniform and rr-regular.

Every covering φ\varphi of a hypergraph \mathcal{H} by a hypergraph 𝒢\mathcal{G} may be considered as a coloring of 𝒢\mathcal{G} by vertices of \mathcal{H}. Moreover, φ\varphi colors the hyperedges of 𝒢\mathcal{G} by colors corresponding to hyperedges of \mathcal{H}. It gives us the following statement.

Proposition 4.

A map φ:X(𝒢)X()\varphi:X(\mathcal{G})\rightarrow X(\mathcal{H}) is a covering of a hypergraph \mathcal{H} by a hypergraph 𝒢\mathcal{G} if and only if φ\varphi is a perfect coloring of 𝒢\mathcal{G} with the parameter matrix 𝕊\mathbb{S} equal to the adjacency matrix 𝔸H\mathbb{A}_{H} of \mathcal{H}. The incidence parameters of this coloring are (C,CT)(C,C^{T}), where CC is the incidence matrix of \mathcal{H}.

An equivalent definition of coverings of hypergraphs was proposed in [LiHou.hypercover]. In particular, in [LiHou.hypercover, Theorem 7] it was stated that hypergraph coverings are equivalent to coverings of their incidence graphs. It is also straightforward from our notions and Proposition 4.

Proposition 5.

A coloring φ\varphi is a covering of a hypergraph \mathcal{H} by a hypergraph 𝒢\mathcal{G} if and only if the induced bipartite perfect coloring φ¯\overline{\varphi} of the incidence graph GG is a covering of the incidence graph HH.

We will say that a covering φ\varphi of a hypergraph \mathcal{H} by 𝒢\mathcal{G} is a kk-covering if for every yX()y\in X(\mathcal{H}) there are exactly kk vertices xX(𝒢)x\in X(\mathcal{G}) such that φ(x)=y\varphi(x)=y. Proposition 5 implies that every covering of a hypergraph is a kk-covering.

Proposition 6.

For every covering φ\varphi of a hypergraph =(Y,U)\mathcal{H}=(Y,U) by 𝒢=(X,E)\mathcal{G}=(X,E) there is kk\in\mathbb{N} such that φ\varphi is kk-covering. In particular, we have |X|=k|Y||X|=k|Y| and |E|=k|U||E|=k|U|.

At last, we note that kk-coverings of hypergraphs can be nicely described in terms of incidence matrices.

Theorem 9.

Let 𝒢\mathcal{G} and \mathcal{H} be hypergraphs with incidence matrices BB and CC, respectively. If there is a kk-covering of \mathcal{H} by 𝒢\mathcal{G}, then the matrix BB is a block matrix {Di,j}\{D_{i,j}\}, where Di,jD_{i,j} is a permutation matrix of order kk if ci,j=1c_{i,j}=1 and Di,jD_{i,j} is the zero matrix of order kk otherwise.

Proof.

By Proposition 4, we can consider a kk-covering φ\varphi as a perfect coloring of 𝒢\mathcal{G} with the incidence parameters (C,CT)(C,C^{T}). In particular, we have that the size of each color class in φ\varphi is kk. To prove the theorem, it only remains to apply Theorem 3(1) to the coloring φ\varphi. ∎

Example 3. Let 𝒢\mathcal{G} be the 33-uniform hypergraph with the vertex set X={x1,x2,,x8}X=\{x_{1},x_{2},\ldots,x_{8}\} and eight hyperedges e1={x3,x6,x7}e_{1}=\{x_{3},x_{6},x_{7}\}, e2={x4,x5,x8}e_{2}=\{x_{4},x_{5},x_{8}\}, e3={x1,x6,x7}e_{3}=\{x_{1},x_{6},x_{7}\}, e4={x2,x5,x8}e_{4}=\{x_{2},x_{5},x_{8}\}, e5={x1,x4,x8}e_{5}=\{x_{1},x_{4},x_{8}\}, e6={x2,x3,x7}e_{6}=\{x_{2},x_{3},x_{7}\}, e7={x2,x3,x5}e_{7}=\{x_{2},x_{3},x_{5}\}, e8={x1,x4,x6}e_{8}=\{x_{1},x_{4},x_{6}\}, and \mathcal{H} be the 33-uniform hypergraph with the vertex set X={x1,x2,x3,x4}X^{\prime}=\{x^{\prime}_{1},x^{\prime}_{2},x^{\prime}_{3},x^{\prime}_{4}\} and four hyperedges e1={x2,x3,x4}e^{\prime}_{1}=\{x^{\prime}_{2},x^{\prime}_{3},x^{\prime}_{4}\}, e2={x1,x3,x4}e^{\prime}_{2}=\{x^{\prime}_{1},x^{\prime}_{3},x^{\prime}_{4}\}, e3={x1,x2,x4}e^{\prime}_{3}=\{x^{\prime}_{1},x^{\prime}_{2},x^{\prime}_{4}\}, e4={x1,x2,x3}e^{\prime}_{4}=\{x^{\prime}_{1},x^{\prime}_{2},x^{\prime}_{3}\}.

Then the map φ:XX\varphi:X\rightarrow X^{\prime} such that φ(x1)=φ(x2)=x1\varphi(x_{1})=\varphi(x_{2})=x^{\prime}_{1}, φ(x3)=φ(x4)=x2\varphi(x_{3})=\varphi(x_{4})=x^{\prime}_{2}, φ(x5)=φ(x6)=x3\varphi(x_{5})=\varphi(x_{6})=x^{\prime}_{3}, φ(x7)=φ(x8)=x4\varphi(x_{7})=\varphi(x_{8})=x^{\prime}_{4} is a 22-covering of the hypergraph \mathcal{H} by the hypergraph 𝒢\mathcal{G}. In particular, φ(e1)=φ(e2)=e1\varphi(e_{1})=\varphi(e_{2})=e^{\prime}_{1}, φ(e3)=φ(e4)=e2\varphi(e_{3})=\varphi(e_{4})=e^{\prime}_{2}, φ(e5)=φ(e6)=e3\varphi(e_{5})=\varphi(e_{6})=e^{\prime}_{3}, and φ(e7)=φ(e8)=e4\varphi(e_{7})=\varphi(e_{8})=e^{\prime}_{4}.

The incidence matrices BB and CC of the hypergraphs 𝒢\mathcal{G} and \mathcal{H}, respectively, are

B=(0010100100010110100001100100100101010010101000011010010001011000);C=(0111101111011110).B=\left(\begin{array}[]{cccccccc}0&0&1&0&1&0&0&1\\ 0&0&0&1&0&1&1&0\\ 1&0&0&0&0&1&1&0\\ 0&1&0&0&1&0&0&1\\ 0&1&0&1&0&0&1&0\\ 1&0&1&0&0&0&0&1\\ 1&0&1&0&0&1&0&0\\ 0&1&0&1&1&0&0&0\\ \end{array}\right);\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ C=\left(\begin{array}[]{cccc}0&1&1&1\\ 1&0&1&1\\ 1&1&0&1\\ 1&1&1&0\\ \end{array}\right).

It can checked that they satisfy the condition of Theorem 9.

The correspondence between coverings and colorings allows us to establish several properties of coverings of hypergraphs. First of all, let us show that if a uniform hypergraph 𝒢\mathcal{G} covers a hypergraph \mathcal{H}, then eigenvalues of \mathcal{H} are also eigenvalues of 𝒢\mathcal{G}. This result was recently obtained in [SongFanWang.hyperspeccover, Theorem 4.1].

Theorem 10.

Assume that a dd-uniform hypergraph 𝒢\mathcal{G} with the adjacency matrix 𝔸G\mathbb{A}_{G} covers a hypergraph \mathcal{H} with the adjacency matrix 𝔸H\mathbb{A}_{H}. Then every eigenvalue of 𝔸H\mathbb{A}_{H} is an eigenvalue of 𝔸G\mathbb{A}_{G} (regarding geometric multiplicities).

Proof.

By the definition of coverings, there is a perfect coloring PP of 𝒢\mathcal{G} with the parameter matrix 𝔸H\mathbb{A}_{H}. By Theorem 8, every eigenvector of the matrix 𝔸H\mathbb{A}_{H} for an eigenvalue λ\lambda gives an eigenvector of the matrix 𝔸G\mathbb{A}_{G} for the same eigenvalue. ∎

Next, let us prove that a covering preserves the set of perfect colorings of hypergraphs.

Theorem 11.

Assume that a dd-uniform hypergraph 𝒢\mathcal{G} with the adjacency matrix 𝔸G\mathbb{A}_{G} covers a hypergraph \mathcal{H} with the adjacency matrix 𝔸H\mathbb{A}_{H}. Then for every perfect coloring of \mathcal{H} with the parameter matrix 𝕊\mathbb{S} there is a perfect coloring of 𝒢\mathcal{G} with the same parameter matrix 𝕊\mathbb{S}.

Proof.

Let PP be a perfect coloring of \mathcal{H} with the parameter matrix 𝕊\mathbb{S}:

𝔸HP=P𝕊.\mathbb{A}_{H}\circ P=P\circ\mathbb{S}.

Since 𝒢\mathcal{G} covers \mathcal{H}, there is a color matrix RR such that

𝔸GR=R𝔸H.\mathbb{A}_{G}\circ R=R\circ\mathbb{A}_{H}.

Then, using properties of the product of multidimensional matrices from Proposition 1, we have

𝔸G(RP)=(𝔸GR)P=(R𝔸H)P=R(𝔸HP)=R(P𝕊)=RP𝕊.\mathbb{A}_{G}\circ(RP)=(\mathbb{A}_{G}\circ R)\circ P=(R\circ\mathbb{A}_{H})\circ P=R\circ(\mathbb{A}_{H}\circ P)=R\circ(P\circ\mathbb{S})=RP\circ\mathbb{S}.

Consequently, RPRP is a perfect coloring of 𝒢\mathcal{G} with the parameter matrix 𝕊\mathbb{S}. ∎

At last, we establish a result on the existence of common coverings for hypergraphs with several corollaries. They can be obtained by an accurate combination of the theorem on the existence of common coverings for graphs (see [leighton.comcover]) and the correspondence between coverings of hypergraphs and their incidence graphs (Proposition 5). Meanwhile, here we provide a direct proof of this result, reformulating the ideas from [leighton.comcover] in terms of incidence matrices. For future applications, we state this result for multihypergraphs.

Theorem 12.

If connected (multi)hypergraphs \mathcal{H} and \mathcal{H}^{\prime} have perfect colorings with the same incidence parameters (V,W)(V,W), then there is a hypergraph 𝒢\mathcal{G} that covers both \mathcal{H} and \mathcal{H}^{\prime}.

Proof.

Assume that CC and CC^{\prime} are the incidence matrices of \mathcal{H} and \mathcal{H}^{\prime} and φ\varphi and φ\varphi^{\prime} are their perfect colorings with incidence parameters (V,W)(V,W), respectively. Let nin_{i} and mjm_{j} (nin_{i}^{\prime} and mjm_{j}^{\prime}), i=1,,ki=1,\ldots,k, j=1,,lj=1,\ldots,l be the number of vertices of color ii and the number of hyperedges of color jj in the induced bipartite perfect coloring of \mathcal{H} (of \mathcal{H}^{\prime}). By Theorem 3(1), there are numbers πi,j\pi_{i,j} and πi,j\pi^{\prime}_{i,j} such that

πi,j=nivi,j=wj,imj;πi,j=nivi,j=wj,imj.\pi_{i,j}=n_{i}v_{i,j}=w_{j,i}m_{j};\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \pi^{\prime}_{i,j}=n^{\prime}_{i}v_{i,j}=w_{j,i}m^{\prime}_{j}.

Moreover, by Theorem 3(1), the incidence matrix CC is the block matrix {Ai,j}\{A_{i,j}\} such that Ai,jA_{i,j} is a (0,1)(0,1)-matrix of size ni×mjn_{i}\times m_{j} with row sums vi,jv_{i,j} and column sums wj,iw_{j,i}. In particular, the block Ai,jA_{i,j} contains πi,j\pi_{i,j} ones. Let us index all nonzero entries ax,ea_{x,e} of Ai,jA_{i,j} by pairs (μ,ν)(\mu,\nu), μ{0,,wj,i1}\mu\in\{0,\ldots,w_{j,i}-1\}, ν{0,,vi,j1}\nu\in\{0,\ldots,v_{i,j}-1\}, so that μ(x)\mu(x) is the position of ax,ea_{x,e} in the ee-th column of Ai,jA_{i,j} with respect to other nonzero entries (from up to down), and ν(e)\nu(e) is the position of this entry in the xx-th row of Ai,jA_{i,j} (from left to right).

Similarly, the incidence matrix CC^{\prime} is a block matrix {Ai,j}\{A^{\prime}_{i,j}\} with blocks Ai,jA^{\prime}_{i,j} of sizes ni×mjn^{\prime}_{i}\times m^{\prime}_{j}, row sums vi,jv_{i,j}, and column sums wj,iw_{j,i}. We also label all nonzero entries of {Ai,j}\{A^{\prime}_{i,j}\} by pairs (μ,ν)(\mu^{\prime},\nu^{\prime}) in the same way.

Let Π\Pi denote the least common multiple of the numbers πi,j\pi_{i,j}: Π=lcmi,j(πi,j),\Pi=\mathrm{lcm}_{i,j}(\pi_{i,j}), and let ti=Π/nit_{i}=\Pi/n_{i} and sj=Π/mjs_{j}=\Pi/m_{j}. Note that all tit_{i} and sjs_{j} are integer.

We construct the incidence matrix BB of a covering (multi)hypergraph 𝒢\mathcal{G} as a block matrix {Di,j}\{D_{i,j}\}, i=1,,ki=1,\ldots,k, j=1,,lj=1,\ldots,l. The size of each block Di,jD_{i,j} will be tinini×sjmjmjt_{i}n_{i}n^{\prime}_{i}\times s_{j}m_{j}m^{\prime}_{j}.

Note that the sizes of blocks Di,jD_{i,j} remain the same, if we use tit^{\prime}_{i} and sjs^{\prime}_{j} instead tit_{i} and sjs_{j}. It is shown in the following claim.

Claim: For all i=1,,ki=1,\ldots,k, j=1,,lj=1,\ldots,l, we have that ti=tit_{i}=t^{\prime}_{i} and sj=sjs_{j}=s^{\prime}_{j}, where ti=Π/nit_{i}=\Pi/n_{i}, ti=Π/nit^{\prime}_{i}=\Pi^{\prime}/n^{\prime}_{i}, sj=Π/mjs_{j}=\Pi/m_{j}, sj=Π/mjs^{\prime}_{j}=\Pi/m^{\prime}_{j}, Π=lcm(πi,j)\Pi=\mathrm{lcm}(\pi_{i,j}), Π=lcm(πi,j)\Pi^{\prime}=\mathrm{lcm}(\pi^{\prime}_{i,j}).

Proof of claim.

Thanks to symmetry, it is sufficient to prove the equalities for all tit_{i} and tit^{\prime}_{i}. Without loss of generality, we show that t1=t1t_{1}=t^{\prime}_{1}, i.e., lcm(πi,j)n1=lcm(πi,j)n1\frac{\mathrm{lcm}(\pi_{i,j})}{n_{1}}=\frac{\mathrm{lcm}(\pi^{\prime}_{i,j})}{n^{\prime}_{1}}. Since \mathcal{H} is a connected (multi)hypergraph, there are Berge paths from a vertex of color 11 to any vertex of color c1c\neq 1. Using relations nivi,j=wj,imjn_{i}v_{i,j}=w_{j,i}m_{j} from Theorem 3(1) for colors of vertices ii and colors of hyperedges jj along this path, we conclude that there are rational numbers r1,cr_{1,c} depending only VV and WW such that nc=r1,cn1n_{c}=r_{1,c}n_{1} for every c1c\geq 1. Using these equalities, we conclude that there is a rational number qq depending only on VV and WW such that Π=lcm(πi,j)=qn1\Pi=\mathrm{lcm}(\pi_{i,j})=qn_{1}.

Acting similarly, in case of the \mathcal{H}^{\prime} we get the same qq such that Π=lcm(πi,j)=qn1\Pi^{\prime}=\mathrm{lcm}(\pi^{\prime}_{i,j})=qn^{\prime}_{1}. Consequently, t1=t1t_{1}=t^{\prime}_{1}. ∎

Let us describe the structure of blocks Di,jD_{i,j}. Given ii and jj, each Di,jD_{i,j} is a block matrix {Fα,β}\{F_{\alpha,\beta}\} with blocks of sizes ti×sjt_{i}\times s_{j}, each index α\alpha has a form (x,x)(x,x^{\prime}), xX()x\in X(\mathcal{H}), xX()x^{\prime}\in X(\mathcal{H}^{\prime}), and an index β\beta has a form (e,e)(e,e^{\prime}), eE()e\in E(\mathcal{H}), eE()e^{\prime}\in E(\mathcal{H}^{\prime}). Each block F=Fα,βF=F_{\alpha,\beta}, where α=(x,x)\alpha=(x,x^{\prime}), β=(e,e)\beta=(e,e^{\prime}), is a (0,1)(0,1)-matrix defined as follows.

  • Assume that for entries of Ai,jA_{i,j} and Ai,jA^{\prime}_{i,j} it holds ax,e=1a_{x,e}=1 and ax,e=1a^{\prime}_{x^{\prime},e^{\prime}}=1. If σν(e)ν(e)modvi,j\sigma\equiv\nu(e)-\nu^{\prime}(e^{\prime})\mod v_{i,j} and δμ(x)μ(x)modwj,i\delta\equiv\mu(x)-\mu^{\prime}(x^{\prime})\mod w_{j,i}, then the σ\sigma-th row and the δ\delta-th column of the block Fα,βF_{\alpha,\beta} contain exactly one 11, otherwise they are filled by 0.

  • If we have ax,e=0a_{x,e}=0 or ax,e=0a^{\prime}_{x^{\prime},e^{\prime}}=0, then the block Fα,βF_{\alpha,\beta} is the zero matrix.

Note that the nonzero blocks Fα,βF_{\alpha,\beta} contain exactly ti/vi,j=Πnivi,jt_{i}/v_{i,j}=\frac{\Pi}{n_{i}v_{i,j}} nonzero rows that coincides with the number sj/wj,i=Πmjwj,is_{j}/w_{j,i}=\frac{\Pi}{m_{j}w_{j,i}} of nonzero columns. Since both of these numbers are integers, the blocks Fα,βF_{\alpha,\beta} are well defined.

It only remains to show that the constructed (0,1)(0,1)-matrix B={Di,j}B=\{D_{i,j}\} (with Di,j={Fα,β}D_{i,j}=\{F_{\alpha,\beta}\}) is the incidence matrix of a (multi)hypergraph 𝒢\mathcal{G} covering both \mathcal{H} and \mathcal{H}^{\prime}. Without loss of generality, let us prove that 𝒢\mathcal{G} covers \mathcal{H}.

For every (x,e)(x,e)-entry of the incidence matrix CC of \mathcal{H}, consider a submatrix (block) Jx,eJ_{x,e} of BB composed of blocks Fα,βF_{\alpha,\beta} such that α=(x,x)\alpha=(x,x^{\prime}), β=(e,e)\beta=(e,e^{\prime}), (x,e)(x^{\prime},e^{\prime}) runs over all entries of CC^{\prime}. It can be checked that Jx,eJ_{x,e} is a square matrix of order Π\Pi^{\prime}.

Let xx has color ii and ee has color jj in the perfect coloring φ\varphi of \mathcal{H}. Then Jx,eJ_{x,e} is a submatrix contained in the block Di,jD_{i,j} of BB.

The definition of blocks Fα,βF_{\alpha,\beta} implies that if (x,e)(x,e)-entry of CC is 0, then the corresponding block Jx,eJ_{x,e} is the zero matrix.

Suppose that (x,e)(x,e)-entry of CC is equal to 11. Then every row of the submatrix Jx,eJ_{x,e} contains exactly one 11, because, (by the construction of Fα,βF_{\alpha,\beta}) there is a unique (x,e)(x^{\prime},e^{\prime})-entry of CC^{\prime} that gives the block Fα,βF_{\alpha,\beta} having the unity entry exactly at this row. For similar reasons, every column of Jx,eJ_{x,e} contains exactly one 11. Consequently, Jx,eJ_{x,e} is a permutation matrix.

By Theorem 9, it means that the constructed (multi)hypergraph 𝒢\mathcal{G} covers \mathcal{H}. If 𝒢\mathcal{G} has multiple hyperedges (i.e. BB contains identical columns), then one can cover 𝒢\mathcal{G} by some hypergraph 𝒢\mathcal{G}^{\prime} that also covers \mathcal{H} and \mathcal{H}^{\prime}. ∎

Remark 1.

In definition of matrices Fα,βF_{\alpha,\beta} instead relations σν(e)ν(e)modvi,j\sigma\equiv\nu(e)-\nu^{\prime}(e^{\prime})\mod v_{i,j} and δμ(x)μ(x)modwj,i\delta\equiv\mu(x)-\mu^{\prime}(x^{\prime})\mod w_{j,i} we can use any other latin squares of orders vi,jv_{i,j} and wj,iw_{j,i}, whose entries we treat as triples (ν(e),ν(e),σ)(\nu(e),\nu^{\prime}(e^{\prime}),\sigma) and (μ(x),μ(x),δ)(\mu(x),\mu^{\prime}(x^{\prime}),\delta), respectively.

Let us obtain several corollaries of Theorem 12. Firstly, we reformulate it in terms of multidimensional matrices.

Corollary 3.

Let 𝔸\mathbb{A} and 𝔸\mathbb{A}^{\prime} be dd-dimensional adjacency matrices of dd-uniform hypergraphs \mathcal{H} and \mathcal{H}^{\prime}. If there are color matrices PP and PP^{\prime} and a dd-dimensional matrix 𝕊\mathbb{S} such that 𝔸P=P𝕊\mathbb{A}\circ P=P\circ\mathbb{S} and 𝔸P=P𝕊\mathbb{A}^{\prime}\circ P^{\prime}=P^{\prime}\circ\mathbb{S}, then there are color matrices RR and RR^{\prime} and a dd-dimensional matrix 𝕃\mathbb{L} such that 𝕃R=R𝔸\mathbb{L}\circ R=R\circ\mathbb{A} and 𝕃R=R𝔸\mathbb{L}\circ R^{\prime}=R^{\prime}\circ\mathbb{A}^{\prime}.

Proof.

The statement directly follows from the existence of a common covering (Theorem 12), an interpretation of a covering as a perfect coloring (Proposition 4), and the definition of perfect colorings by the means of multidimensional adjacency matrices (Theorem 5). ∎

Corollary 4.

Let \mathcal{H} and \mathcal{H}^{\prime} be dd-uniform connected hypergraphs. There are perfect colorings of \mathcal{H} and \mathcal{H}^{\prime} with the same parameter matrix 𝕊\mathbb{S} if and only if there exists a hypergraph 𝒢\mathcal{G} that covers both \mathcal{H} and \mathcal{H}^{\prime}.

Proof.

Necessity is proved in Corollary 3 and in Theorem 12.

Let us prove sufficiency. Suppose that a hypergraph 𝒢\mathcal{G} covers hypergraphs \mathcal{H} and \mathcal{H}^{\prime} with adjacency matrices 𝔸\mathbb{A} and 𝔸\mathbb{A}^{\prime}, respectively. Assume that there are no perfect colorings of \mathcal{H} and \mathcal{H}^{\prime} with the same parameter matrix 𝕊\mathbb{S}. In particular, the minimal perfect colorings PP and PP^{\prime} of these hypergraphs (that exist by Theorem 4) have different parameter matrices 𝕊\mathbb{S} and 𝕊\mathbb{S}^{\prime}:

𝔸P=P𝕊;𝔸P=P𝕊.\mathbb{A}\circ P=P\circ\mathbb{S};\leavevmode\nobreak\ \leavevmode\nobreak\ \mathbb{A}^{\prime}\circ P^{\prime}=P^{\prime}\circ\mathbb{S}^{\prime}.

By Theorem 11, the hypergraph 𝒢\mathcal{G} has perfect colorings with the parameter matrices 𝕊\mathbb{S} and 𝕊\mathbb{S}^{\prime}. Using again Theorem 4, find the minimal perfect coloring of the hypergraph 𝒢\mathcal{G}. It has the parameter matrix 𝕋\mathbb{T} such that 𝕋\mathbb{T} is different from at least one of the matrices 𝕊\mathbb{S} and 𝕊\mathbb{S}^{\prime}. Without loss of generality, assume that 𝕋\mathbb{T} is not equal to 𝕊\mathbb{S}. Then the perfect coloring of 𝒢\mathcal{G} with the parameter matrix 𝕊\mathbb{S} is a refinement of the minimal perfect coloring of 𝒢\mathcal{G}. Equivalently, there exists a color matrix RR such that 𝕊R=R𝕋\mathbb{S}\circ R=R\circ\mathbb{T}. Using Proposition 1, we get

𝔸(PR)=𝔸(PR)=(𝔸P)R=(P𝕊)R=\displaystyle\mathbb{A}\circ(PR)=\mathbb{A}\circ(P\circ R)=(\mathbb{A}\circ P)\circ R=(P\circ\mathbb{S})\circ R=
P(𝕊R)=P(R𝕋)=(PR)𝕋.\displaystyle P\circ(\mathbb{S}\circ R)=P\circ(R\circ\mathbb{T})=(PR)\circ\mathbb{T}.

Therefore, PRPR is a perfect coloring of \mathcal{H} in which some color classes are a union of color classes of PP: a contradiction with the minimality of the perfect coloring PP. ∎

We also prove that every regular uniform hypergraph can be covered by a multipartite hypergraph that can be partitioned into perfect matchings.

Corollary 5.

For every dd-uniform rr-regular hypergraph \mathcal{H} there is a hypergraph 𝒢\mathcal{G} such that 𝒢\mathcal{G} covers \mathcal{H} and 𝒢\mathcal{G} is dd-partite hypergraph that can be partitioned into rr perfect matchings.

Proof.

Consider an auxiliary multihypergraph \mathcal{F} with the incidence matrix BB of size d×rd\times r whose all entries equal to 11. It consists of dd vertices x1,,xdx_{1},\ldots,x_{d} and a hyperedge e=(x1,,xd)e=(x_{1},\ldots,x_{d}) of multiplicity rr. By the definition, each hyperedge of \mathcal{F} is a perfect matching and \mathcal{F} is a dd-partite hypergraph (each vertex is a part).

Colorings of all vertices into one color are perfect colorings of \mathcal{H} and \mathcal{F} that also induce monochrome colorings of hyperedges of these hypergraphs. The incidence parameter matrices WW and VV of these colorings are the same, have size 1×11\times 1 and, by the definition, W=(d)W=(d) and V=(r)V=(r). By Theorem 12, there is a hypergraph 𝒢\mathcal{G} covering both \mathcal{H} and \mathcal{F}. Since the covering preserves the incidence relations between vertices and hyperedges, we see that every vertex of \mathcal{F} corresponds to a part of 𝒢\mathcal{G} and every hyperedge gives a perfect matching in 𝒢\mathcal{G}. So 𝒢\mathcal{G} is a dd-partite hypergraph that can be partitioned into rr perfect matchings. ∎

5 Examples of perfect colorings of hypergraphs

5.1 Transversals and perfect matchings

A kk-transversal in uniform regular hypergraphs gives us the simplest nontrivial example of perfect colorings: the vertices of a hypergraph are colored into two colors (transversal and non-transversal), while all hyperedges have the same color. So if 𝒢\mathcal{G} is a dd-uniform rr-regular hypergraph, then a kk-transversal is a perfect coloring with the incidence parameters

V=(rr);W=(kdk).V=\left(\begin{array}[]{c}r\\ r\end{array}\right);\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ W=\left(\begin{array}[]{cc}k&d-k\end{array}\right).

By Theorem 5, the parameter matrix 𝕊\mathbb{S} of a kk-transversal is the dd-dimensional matrix of order 22 with entries

sγ={r(d1k1)1 if γ1=1 and #{γi:γi=1}=k;r(d1k)1 if γ1=2 and #{γi:γi=1}=k;0 otherwise.s_{\gamma}=\left\{\begin{array}[]{l}r{d-1\choose k-1}^{-1}\mbox{ if }\gamma_{1}=1\mbox{ and }\#\{\gamma_{i}:\gamma_{i}=1\}=k;\\ r{d-1\choose k}^{-1}\mbox{ if }\gamma_{1}=2\mbox{ and }\#\{\gamma_{i}:\gamma_{i}=1\}=k;\\ 0\mbox{ otherwise}.\end{array}\right.

For example, the parameter matrices of kk-transversals in dd-uniform rr-regular hypergraphs for some small dd and kk are the following:

d=2,k=1:𝕊=(0rr0);d=2,\leavevmode\nobreak\ k=1:\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \mathbb{S}=\left(\begin{array}[]{cc}0&r\\ r&0\end{array}\right);
d=3,k=1:𝕊=(0r/2r0r/2000);d=3,\leavevmode\nobreak\ k=1:\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \mathbb{S}=\left(\begin{array}[]{cc|cc}0&\nicefrac{{r}}{{2}}&r&0\\ \nicefrac{{r}}{{2}}&0&0&0\end{array}\right);
d=4,k=1:𝕊=(0r/3r0r/3000r/30000000);d=4,\leavevmode\nobreak\ k=1:\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \mathbb{S}=\left(\begin{array}[]{cc|cc}0&\nicefrac{{r}}{{3}}&r&0\\ \nicefrac{{r}}{{3}}&0&0&0\\ \hline\cr\nicefrac{{r}}{{3}}&0&0&0\\ 0&0&0&0\end{array}\right);
d=4,k=2:𝕊=(000r/30r/3r/300r/3r/30r/3000).d=4,\leavevmode\nobreak\ k=2:\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \mathbb{S}=\left(\begin{array}[]{cc|cc}0&0&0&\nicefrac{{r}}{{3}}\\ 0&\nicefrac{{r}}{{3}}&\nicefrac{{r}}{{3}}&0\\ \hline\cr 0&\nicefrac{{r}}{{3}}&\nicefrac{{r}}{{3}}&0\\ \nicefrac{{r}}{{3}}&0&0&0\end{array}\right).

Let us find eigenvalues of transversals as perfect colorings of hypergraphs.

Theorem 13.

Suppose that 𝕊\mathbb{S} is the parameter matrix of the perfect coloring corresponding to a kk-transversal in a dd-uniform rr-regular hypergraph 𝒢\mathcal{G}. Then the eigenvalues of 𝕊\mathbb{S} are λ0=0\lambda_{0}=0 and λj=rξjk\lambda_{j}=r\xi^{jk}, where ξ\xi is a dd-th primitive root of unity. In particular, the matrix 𝕊\mathbb{S} has dgcd(k,d)\frac{d}{\gcd(k,d)} different nonzero eigenvalues.

Proof.

Knowing the entries of the matrix 𝕊\mathbb{S} from Theorem 5, we find that the equation 𝕊x=λ𝕀x\mathbb{S}\circ x=\lambda\mathbb{I}\circ x on eigenvalues λ\lambda is equivalent to the following system:

{λx1d1+rx1k1x2dk=0;rx1kx2dk1λx2d1=0.\left\{\begin{array}[]{c}-\lambda x_{1}^{d-1}+rx_{1}^{k-1}x_{2}^{d-k}=0;\\ rx_{1}^{k}x_{2}^{d-k-1}-\lambda x_{2}^{d-1}=0.\end{array}\right.

From here it is easy to see that λ0=0\lambda_{0}=0 is an eigenvalue with the corresponding eigenvectors (x1,0)(x_{1},0) and (0,x2)(0,x_{2}).

Assume that x=(x1,x2)x=(x_{1},x_{2}) is an eigenvector with both nonzero components: x1=tx2x_{1}=tx_{2}, t0t\neq 0. Then the above system of equations can be rewritten as

{λtd1+rtk1=0;rtkλ=0.\left\{\begin{array}[]{l}-\lambda t^{d-1}+rt^{k-1}=0;\\ rt^{k}-\lambda=0.\end{array}\right.

Consequently, λ=rtk\lambda=rt^{k} and td=1t^{d}=1, so tt is a dd-th root of unity. So all nonzero eigenvalues of the matrix 𝕊\mathbb{S} are λ=rξjk\lambda=r\xi^{jk}, j=1,,dj=1,\ldots,d, where ξ\xi is a dd-th primitive root of unity. ∎

We believe that all nonzero eigenvalues of 𝕊\mathbb{S} have the same algebraic multiplicity. Moreover, we propose the following conjecture on the characteristic polynomial of 𝕊\mathbb{S}.

Conjecture 1.

Let 𝕊\mathbb{S} be the parameter matrix of a kk-transversal in a dd-uniform rr-regular hypergraph 𝒢\mathcal{G}. Then the characteristic polynomial of 𝕊\mathbb{S} is

φ(λ)=λd2i=jd(λξjkr),\varphi(\lambda)=\lambda^{d-2}\prod\limits_{i=j}^{d}(\lambda-\xi^{jk}r),

where ξ\xi is a dd-th primitive root of unity.

Combining Theorem 13 with Theorem 8, we obtain that every dd-uniform rr-regular hypergraph with a kk-transversal should have values rξjkr\xi^{jk} in its spectrum. It gives a spectral condition on existence of kk-transversals that cannot be reduced to the spectrum of the induced coloring of the incidence graph.

Indeed, consider complete 33-uniform hypergraphs on n4n\geq 4 vertices. It is obvious that they do not contain transversals.

Using computational results for [CoopDut.hyperspec] (available at [dutle.compres]), one can check that complete 33-uniform hypergraphs on nn vertices, 4n84\leq n\leq 8, do not have eigenvalues re±2πi3re^{\pm\frac{2\pi i}{3}}. It indicates that they do not contain a transversal.

On the other hand, the parameter matrix of the induced perfect coloring of the incidence graph of a dd-uniform rr-regular hypergraph is

(0WV0)=(01d1r00r00).\left(\begin{array}[]{cc}0&W\\ V&0\end{array}\right)=\left(\begin{array}[]{ccc}0&1&d-1\\ r&0&0\\ r&0&0\end{array}\right).

The eigenvalues of this matrix are λ1=0\lambda_{1}=0 and λ2,3=±dr\lambda_{2,3}=\pm\sqrt{dr}, where the latter values are contained in the eigenspectrum of every bipartite biregular graph with degrees of parts rr and dd. Eigenvalue 0 belongs to the spectrum of many of bipartite biregular graphs, for example, the incidence graph of the complete 33-uniform hypergraph on 55 vertices.

5.2 Perfect 22-colorings of 33-uniform hypergraphs

Let us completely describe the smallest case of hypergraph perfect colorings, i.e., perfect colorings PP of rr-regular 33-uniform hypergraphs in 22 colors.

Assume that n1n_{1} and n2n_{2} are the numbers of vertices of first and second colors and let χ=n1/n21\chi=n_{1}/n_{2}\leq 1. In a general case, the incidence parameters (V,W)(V,W) of the coloring PP are

V=(v1,1v1,2v1,300v2,2v2,3v2,4);W=(30211203),V=\left(\begin{array}[]{cccc}v_{1,1}&v_{1,2}&v_{1,3}&0\\ 0&v_{2,2}&v_{2,3}&v_{2,4}\end{array}\right);\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ W=\left(\begin{array}[]{cc}3&0\\ 2&1\\ 1&2\\ 0&3\end{array}\right),

where v1,1+v1,2+v1,3=v2,2+v2,3+v2,4=rv_{1,1}+v_{1,2}+v_{1,3}=v_{2,2}+v_{2,3}+v_{2,4}=r. If some color range of hyperedges is absent in the coloring PP, then the matrices WW and VV lack the corresponding row and column.

From the equality NV=WTMNV=W^{T}M (see Theorem 3(1)), we deduce that

n1v1,2=2n2v2,2;n2v2,3=2n1v1,3,n_{1}v_{1,2}=2n_{2}v_{2,2};\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ n_{2}v_{2,3}=2n_{1}v_{1,3},

and, consequently,

v2,2v1,2=χ/2;v2,3v1,3=2χ.\frac{v_{2,2}}{v_{1,2}}=\chi/2;\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \frac{v_{2,3}}{v_{1,3}}=2\chi. (3)

The parameter matrix of a perfect coloring PP is a 33-dimensional matrix of order 22:

𝕊=(s1,1,1s1,1,2s2,1,1s2,1,2s1,2,1s1,2,2s2,2,1s2,2,2).\mathbb{S}=\left(\begin{array}[]{cc|cc}s_{1,1,1}&s_{1,1,2}&s_{2,1,1}&s_{2,1,2}\\ s_{1,2,1}&s_{1,2,2}&s_{2,2,1}&s_{2,2,2}\end{array}\right).

With the help of Theorem 5, we find entries of the parameter matrix 𝕊\mathbb{S}:

s1,1,1=v1,1;s1,1,2=s1,2,1=v1,2/2;s1,2,2=v1,3;\displaystyle s_{1,1,1}=v_{1,1};\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ s_{1,1,2}=s_{1,2,1}=v_{1,2}/2;\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ s_{1,2,2}=v_{1,3};
s2,1,1=v2,2;s2,1,2=s2,2,1=v2,3/2;s2,2,2=v2,4.\displaystyle s_{2,1,1}=v_{2,2};\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ s_{2,1,2}=s_{2,2,1}=v_{2,3}/2;\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ s_{2,2,2}=v_{2,4}.

Using additionally equalities (3), we reduce the number of essential parameters of the matrix 𝕊\mathbb{S} and denote them by a,b,ca,b,c and dd:

𝕊=(abχbχcbcχcd).\mathbb{S}=\left(\begin{array}[]{cc|cc}a&b&\chi b&\chi c\\ b&c&\chi c&d\end{array}\right).

To calculate the characteristic polynomial φ(λ)\varphi(\lambda) of the matrix 𝕊\mathbb{S}, we use a formula from [MorShak.resultform] for resultants R2,2R_{2,2}:

φ(λ)=λ42λ3(a+d)+λ2(d2+4ad+a26χbc)+\displaystyle\varphi(\lambda)=\lambda^{4}-2\lambda^{3}(a+d)+\lambda^{2}(d^{2}+4ad+a^{2}-6\chi bc)+
λ((a+d)(6χbc2ad)4χ2c3χb3)+a2d23χ2b2c26χabcd+4χ2ac3+χb3d.\displaystyle\lambda((a+d)(6\chi bc-2ad)-4\chi^{2}c^{3}-\chi b^{3})+a^{2}d^{2}-3\chi^{2}b^{2}c^{2}-6\chi abcd+4\chi^{2}ac^{3}+\chi b^{3}d.

5.3 Fano’s plane hypergraph

Here we illustrate that perfect colorings can be used for finding eigenvalues of some hypergraphs and multidimensional matrices.

Consider a hypergraph \mathcal{F} corresponding to the Fano’s plane: \mathcal{F} is a 33-uniform hypergraph on 77 vertices with the following 77 hyperedges:

(x1,x2,x3),(x1,x4,x5),(x1,x6,x7),(x2,x4,x6),(x2,x5,x7),(x3,x4,x7),(x3,x5,x6).(x_{1},x_{2},x_{3}),\leavevmode\nobreak\ (x_{1},x_{4},x_{5}),\leavevmode\nobreak\ (x_{1},x_{6},x_{7}),\leavevmode\nobreak\ (x_{2},x_{4},x_{6}),\leavevmode\nobreak\ (x_{2},x_{5},x_{7}),\leavevmode\nobreak\ (x_{3},x_{4},x_{7}),\leavevmode\nobreak\ (x_{3},x_{5},x_{6}).

Due to the symmetry of the hypergraph \mathcal{F}, it is not hard to list all its 22-colorings and find that only two of them are perfect.

1. Color one vertex of \mathcal{F} into the first color, and all other six vertices into the second color (Figure 2):

[Uncaptioned image]

Figure 2: The first perfect 2-coloring of the Fano’s plane hypergraph.

The incidence parameters are

V=(3012);W=(1203).V=\left(\begin{array}[]{cc}3&0\\ 1&2\end{array}\right);\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ W=\left(\begin{array}[]{cc}1&2\\ 0&3\end{array}\right).

Using the help of Theorem 5, we find that the multidimensional parameter matrix of this perfect coloring is

𝕊1=(0001/2031/22).\mathbb{S}_{1}=\left(\begin{array}[]{cc|cc}0&0&0&\nicefrac{{1}}{{2}}\\ 0&3&\nicefrac{{1}}{{2}}&2\end{array}\right).

To calculate the characteristic polynomial φ1(λ)\varphi_{1}(\lambda) of the matrix 𝕊1\mathbb{S}_{1}, we use a formula from [MorShak.resultform] or the result for 22-colorings of 33-uniform hypergraphs from the previous section:

φ1(λ)=λ(λ3)(λ2λ+1).\varphi_{1}(\lambda)=\lambda(\lambda-3)(\lambda^{2}-\lambda+1).

So the parameter matrix 𝕊1\mathbb{S}_{1} has eigenvalues 0, 33, and 12(1±i3)\frac{1}{2}(1\pm i\sqrt{3}).

2. Another perfect coloring of the hypergraph \mathcal{F} is presented at Figure 3.

[Uncaptioned image]

Figure 3: The second perfect 2-coloring of the Fano’s plane hypergraph.

The incidence parameters of this coloring are

V=(1203);W=(3012).V=\left(\begin{array}[]{cc}1&2\\ 0&3\end{array}\right);\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ W=\left(\begin{array}[]{cc}3&0\\ 1&2\end{array}\right).

Theorem 5 gives the following multidimensional parameter matrix:

𝕊2=(1003/2023/20).\mathbb{S}_{2}=\left(\begin{array}[]{cc|cc}1&0&0&\nicefrac{{3}}{{2}}\\ 0&2&\nicefrac{{3}}{{2}}&0\end{array}\right).

Using the result for 22-colorings of 33-uniform hypergraphs from previous section, we find the characteristic polynomial for the matrix 𝕊2\mathbb{S}_{2}:

φ2(λ)=(λ3)(λ1)(λ2+2λ+6).\varphi_{2}(\lambda)=(\lambda-3)(\lambda-1)(\lambda^{2}+2\lambda+6).

So the parameter matrix 𝕊2\mathbb{S}_{2} has eigenvalues 11, 33, and 1±i5-1\pm i\sqrt{5}.

Thus, the hypergraph \mathcal{F} of the Fano’s plane has at least 77 different eigenvalues:

λ0=0,λ1=1,λ2=3,λ3,4=12(1±i3),λ5,6=1±i5.\lambda_{0}=0,\leavevmode\nobreak\ \lambda_{1}=1,\leavevmode\nobreak\ \lambda_{2}=3,\leavevmode\nobreak\ \lambda_{3,4}=\frac{1}{2}(1\pm i\sqrt{3}),\lambda_{5,6}=-1\pm i\sqrt{5}.

Acknowledgements

The author is grateful to the anonymous reviewer for valuable comments.

This work was funded by the Russian Science Foundation under grant No 22-21-00202, https://rscf.ru/project/22-21-00202/.

References

  • [] \bibselectbiblio