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

On monoid graphs

Kolja Knauer    Gil Puig i Surroca
Abstract

We investigate Cayley graphs of finite semigroups and monoids. First, we look at semigroup digraphs, i.e., directed Cayley graphs of semigroups, and give a Sabidussi-type characterization in the case of monoids. We then correct a proof of Zelinka from ’81 that characterizes semigroup digraphs with outdegree 11. Further, answering a question of Knauer and Knauer, we construct for every k2k\geq 2 connected kk-outregular non-semigroup digraphs. On the other hand, we show that every sink-free directed graph is a union of connected components of a monoid digraph.

Second, we consider monoid graphs, i.e., underlying simple undirected graphs of Cayley graphs of monoids. We show that forests and threshold graphs form part of this family. Conversely, we construct the – to our knowledge – first graphs, that are not monoid graphs. We present non-monoid graphs that are planar, have arboricity 22, and treewidth 33 on the one hand, and non-monoid graphs of arbitrarily high connectivity on the other hand.

Third, we study generated monoid trees, i.e., trees that are monoid graphs with respect to a generating set. We give necessary and sufficient conditions for a tree to be in this family, allowing us to find large classes of trees inside and outside the family.

1 Introduction

After being introduced in 1878 [9], Cayley graphs soon became a prominent tool in both graph and group theory. As of today Cayley graphs of groups play a prominent role in (books devoted to) algebraic graph theory, see e.g. [15, 38]. On one hand, they are useful for constructions. They are crucial in the proof of Frucht’s theorem [11] and the related representation problems [5], i.e., how to represent a group as the automorphism group of a graph. On the other hand, Cayley graphs endow groups with metric or topological properties. For instance, White defined the genus of a group as the minimum genus for any connected Cayley graph of the group [56]. This created intricate links to symmetry groups of surfaces, see [17, 57].

Cayley graphs of monoids and semigroups (usually considered as directed Cayley graphs) have been studied less, but still there is a considerable amount of work, see e.g. [38] for a book and the references therein and [26] for a survey on this subject. Characterizations of Cayley graphs of certain classes of semigroups have been subject to some research effort, see [30, 28, 32, 31, 43, 54, 34, 33, 19, 20, 41, 21, 35, 55]. Also Cayley graphs from certain restricted families have been studied, such as transitively acyclic, directed ones also known as Cayley posets [14] as well as generalized Petersen graphs [20, 19, 39, 13] and others [53, 44].

Along the lines of White’s genus of a group characterizations of semigroups admitting Cayley graphs with topological embeddability properties have been investigated extensively [51, 52, 59, 36, 37]. Note that for such topological questions edge directions, multiplicities and loops can usually be suppressed, which leads to the notion of underlying simple undirected Cayley graph. Also other questions of the type which semigroups admit a Cayley graph from a given class have been considered [25, 47, 27, 29]. This latter type of question is related to the representation problem for monoids. Indeed, a standard strategy to represent a monoid is to take its colored Cayley graph, see Figure 1, and mimic the colors with certain rigid gadgets on the edges. Then the monoid is isomorphic to the endomorphism monoid of the resulting graph, see [22, 23, 24]. Thus, if we know that a monoid has a Cayley graph in a given class, then (depending on the edge gadgets) it is the endomorphism monoid of a graph from that class. There are classic and recent questions and results on how sparse a class can be be while being rich enough to represent all groups [11, 16, 3] or monoids, see [46, Problem 19.2] or [6].

Refer to caption
Figure 1: Left: the multiplication table of a monoid MM with neutral element ee. Right: the colored Cayley graph Caycol(M,{a,b,c})\mathrm{Cay}_{\mathrm{col}}(M,\{a,b,c\}) with connection set {a,b,c}\{a,b,c\}.

In this paper we study the related question of how large the class of (finite) Cayley graphs is and if some structural properties can be shown similar to the case of Cayley graphs of groups. In particular, a starting point was the question whether sparsity can ensure Cayleyness. A similar study has been carried out with respect to Cayley posets in [14].

First, we study the class of directed Cayley graphs of semigroups and monoids called semigroup digraphs and monoid digraphs, respectively. In Section 2, we present an easy but to our knowledge unpublished analogue of Sabidussi’s Lemma [50] for monoid digraphs. This answers a question of [38, Chapter 11.3]. Section 3 further studies the class of semigroup and monoid digraphs. We review results of Zelinka [58], who characterized such digraphs with outdegree 11. In particular, we correct his proof. Answering a question of Knauer and Knauer [38, below Proposition 11.3.2], we construct for every k2k\geq 2 connected kk-outregular digraphs (of high (strong) connectivity) that are not semigroup digraphs. On the other hand, we show that every sink-free directed graph, i.e. with minimum outdegree at least 11, is a union of connected components of a monoid digraph. This can be seen as an analogue to the result that every graph is isomorphic to some induced subgraph of a Cayley graph of a group [4].

Second, we focus on the underlying simple undirected graphs of Cayley graphs of monoids, called monoid graphs. As mentioned earlier removing directions, orientations, and loops is natural when considering embeddability and sparsity questions of the resulting graph and has prompted questions about this class, see e.g. [13, 36, 37]. On the other hand, one difficulty that arises in their study is that they carry less categorical and algebraic structure than their directed non-simple counterparts, see e.g. [38, Chapter 11.1]. In Section 4 we show that forests and threshold graphs are monoid graphs. After that, we show to our knowledge the first graphs, that are not monoid graphs. These results were the starting point for this paper. In particular, we present families of planar non-monoid graphs of arboricity 22 and treewidth 33 and construct non-monoid graphs of arbitrarily high connectivity based on Ramanujan graphs.

Finally, we study monoid graphs where the connection set of the Cayley graph generates the monoid. In Section 5 we give a necessary and a sufficient condition for a tree to be in that class. This allows us to find large families of trees inside and outside of the family.

There are still many intriguing open questions, which we survey in Section 6.

2 Preliminaries, properties, and characterizations

For a semigroup SS and a subset CSC\subseteq S called connection set, we consider three different Cayley graphs. The directed graph with vertex set SS and arc set {(s,sc)|sS,cC}\{(s,sc)\ |\ s\in S,\,c\in C\}, denoted Cay(S,C)\mathrm{Cay}(S,C), is a digraph with possible loops and anti-parallel arcs, but without parallel arcs. We call the resulting digraphs semigroup and monoid digraphs, respectively. Note that the construction allowing parallel arcs has been studied from a categorical point of view, see e.g. [38, Chapter 11.1]. Suppressing multiplicities still yields a faithful covariant functor as in the case with parallel arcs.

The underlying simple undirected graph of Cay(S,C)\mathrm{Cay}(S,C), which we denote Cay¯(S,C)\underline{\mathrm{Cay}}(S,C), carries less information. It looses categorical and algebraic properties, but as justified above, is natural from a graph theoretical point of view. We call the resulting graphs semigroup and monoid graphs. In the last section we will further restrict this class by requiring that the connection set CC is a generating set, yielding the notion of generated monoid graphs.

The richest representation of the pair (S,C)(S,C) is the multi-digraph with colored arcs Caycol(S,C)\mathrm{Cay}_{\mathrm{col}}(S,C), where the colors correspond to the elements of CC generating the arcs. Here for each cc there is an arc (s,sc)(s,sc) of color cc, in particular, we allow multiple parallel arcs.

We denote by End(G)\mathrm{End}(G) and Aut(G)\mathrm{Aut}(G) the endomorphism monoid and automorphism group of GG, respectively. Here GG can be colored and directed, only directed, or undirected. In directed and undirected graphs endomorphisms just have to map arcs to arcs and edges to edges, respectively. If GG is colored and directed then an endomorphism has to map any arc to an arc of the same color. Automorphisms are just bijective endomorphisms. Clearly, the endomorphism monoid of a colored digraph is a submonoid of the endomorphism monoid of the digraph without colors. However, suppressing loops may remove some endomorphisms. This is, for a semigroup SS and CSC\subseteq S, we have End(Caycol(S,C))End(Cay(S,C))\mathrm{End}(\mathrm{Cay}_{\mathrm{col}}(S,C))\leq\mathrm{End}(\mathrm{Cay}(S,C)) but there might be φEnd(Cay¯(S,C))End(Cay(S,C))\varphi\in\mathrm{End}(\underline{\mathrm{Cay}}(S,C))\setminus\mathrm{End}(\mathrm{Cay}(S,C)).

The multiplication law of a semigroup yields information shared by all the (colored) Cayley digraphs it defines. See the following basic lemma.

Lemma 2.1.

If SS is a semigroup and CSC\subseteq S, then mapping every sSs\in S to left-multiplication φs\varphi_{s} with ss is a homomorphism from SS to End(Caycol(S,C))\mathrm{End}(\mathrm{Cay}_{\mathrm{col}}(S,C)).

Proof.

If s,tSs,t\in S, cCc\in C and (t,tc)(t,tc) is an arc of Caycol(S,C)\mathrm{Cay}_{\mathrm{col}}(S,C) having color cc, then (st,stc)(st,stc) is also an arc of Caycol(S,C)\mathrm{Cay}_{\mathrm{col}}(S,C) having color cc. Thus, left-multiplication φs\varphi_{s} with ss is an element of End(Caycol(S,C))\mathrm{End}(\mathrm{Cay}_{\mathrm{col}}(S,C)). Clearly, φst=φsφt\varphi_{st}=\varphi_{s}\circ\varphi_{t}, so this is a semigroup homomorphism. \square

A well-known strengthening of Lemma 2.1 is the following property of generated colored Cayley graphs of monoids, see e.g. [38, Theorem 7.3.7]:

Lemma 2.2.

Let MM be a monoid and be CMC\subseteq M such that M=CM=\langle C\rangle. Then left-multiplication yields an isomorphism from MM to Caycol(M,C)\mathrm{Cay}_{\mathrm{col}}(M,C).

Results that complement the above lemmas with a converse direction are sometimes called Sabidussi-type characterizations. They rely on certain properties of the automorphism group or endomorphism monoid. The first such result is due to Sabidussi [50, Lemma 4]:

Lemma 2.3.

A (colored, directed) graph GG is the (colored, directed) Cayley graph of a group if and only if Aut(G)\mathrm{Aut}(G) has a subgroup that acts regularly on GG.

Here we prove a Sabidussi-type characterization of monoid digraphs, that was asked for in [38, Chapter 11.3].

Lemma 2.4.

A directed graph G=(V,A)G=(V,A) is a monoid digraph if and only if there exists a vertex eVe\in V and a submonoid MEnd(G)M\leq\mathrm{End}(G) such that for each xVx\in V there is a unique φxM\varphi_{x}\in M with φx(e)=x\varphi_{x}(e)=x. Moreover, φx\varphi_{x} satisfies that for each (x,y)A(x,y)\in A there is (e,c)A(e,c)\in A such that φx(c)=y\varphi_{x}(c)=y.

Proof.

Suppose that there is a vertex ee and a submonoid MEnd(G)M\leq\mathrm{End}(G) satisfying this property. Let C={φM|(e,φ(e))A}C=\{\varphi\in M\ |\ (e,\varphi(e))\in A\}. We claim that there is an isomorphism GCay(M,C)G\cong\mathrm{Cay}(M,C) given by xφxx\mapsto\varphi_{x}. It is clear that this map is injective. It is also surjective: the preimage of φM\varphi\in M is φ(e)\varphi(e). Now, if (x,y)A(x,y)\in A, there is an out-neighbor cc of ee with φxφc(e)=φx(c)=y\varphi_{x}\circ\varphi_{c}(e)=\varphi_{x}(c)=y; thus φxφc=φy\varphi_{x}\circ\varphi_{c}=\varphi_{y}. Since φcC\varphi_{c}\in C, (φx,φy)(\varphi_{x},\varphi_{y}) is an arc of Cay(M,C)\mathrm{Cay}(M,C). Conversely, if (φx,φy)(\varphi_{x},\varphi_{y}) is an arc of Cay(M,C)\mathrm{Cay}(M,C) then φxφ=φy\varphi_{x}\circ\varphi=\varphi_{y} for some φC\varphi\in C. This implies that the arc (e,φ(e))A(e,\varphi(e))\in A is mapped by φx\varphi_{x} to (x,φxφ(e))=(x,y)(x,\varphi_{x}\circ\varphi(e))=(x,y), which must be also in AA, because φx\varphi_{x} is an endomorphism.

Conversely, suppose that G=Cay(M,C)G=\mathrm{Cay}(M,C) for a monoid MM and CMC\subseteq M. For each vertex xx consider the endomorphism φx\varphi_{x} of GG defined by left-multiplication by xx. All these endomorphisms together form a submonoid MEnd(Caycol(M,C))End(G)M^{\prime}\leq\mathrm{End}(\mathrm{Cay}_{\mathrm{col}}(M,C))\leq\mathrm{End}(G), by Lemma 2.1. Let ee be the neutral element of MM. It is clear that for any vertex xx, the mapping φx\varphi_{x} is the only of them that maps ee to xx. Moreover, for any out-neighbor yy of xx there is an out-neighbor cCc\in C of ee with xc=yxc=y, so φx(c)=y\varphi_{x}(c)=y. \square

We end the section with a lemma which shows how the connectivity of a semigroup digraph yields information about its defining semigroup(s). A directed graph is (strongly) connected if for any two vertices x,yx,y there is a (directed) path from xx to yy. A set of arcs AAA^{\prime}\subseteq A of a connected directed graph G=(V,A)G=(V,A) is a directed edge cut if (V,AA)(V,A\setminus A^{\prime}) has exactly two connected components V1V_{1} and V2V_{2} and all arcs of AA^{\prime} are directed from V1V_{1} to V2V_{2}. It is an easy observation, that a connected directed graph G=(V,A)G=(V,A) is strongly connected if and only if it has no directed edge cut.

Lemma 2.5.

If G=Cay(S,C)G=\mathrm{Cay}(S,C) is strongly connected, then SS is left cancellative.

Proof.

For any sSs\in S and any XSX\subseteq S with XCXXC\subseteq X we have that sXCsXsXC\subseteq sX. In particular, sSsS has no outgoing arcs. Hence if sSSsS\subsetneq S, then there is a directed edge cut defined by the arcs from SsSS\setminus sS to sSsS. Thus, strong connectivity yields sS=SsS=S. By Lemma 2.1 and the fact that SS is finite, left-multiplication by ss is an automorphism of GG. This means that SS is left cancellative. \square

3 Semigroup digraphs and monoid digraphs

In this section we study semigroup and monoid digraphs. If for a directed graph GG there is a semigroup SS and a non-empty CSC\subseteq S with G=Cay(S,C)G=\mathrm{Cay}(S,C), then the outdegree of any vertex of GG is between 11 and |C||C|. In the 11-outregular case, i.e. when all vertices have outdegree 11, one can reduce CC to any of its members without changing GG:

Observation 3.1.

A semigroup digraph G=Cay(S,C)G=\mathrm{Cay}(S,C) is 11-outregular if and only if G=Cay(S,{a})G=\mathrm{Cay}(S,\{a\}) for some aCa\in C.

The 11-outregular semigroup digraphs were studied by Zelinka [58]. He gave a very concrete characterization of this class, but his construction of the corresponding semigroups is flawed. However, an alternative non-trivial construction can be given inspired by [38, Proposition 11.3.2]. This is the first objective of this section. We give a monoid version of the result (Theorem 3.2), and then obtain Zelinka’s theorem as a corollary (Corollary 3.3).

In a 11-outregular digraph GG each connected component 𝒞\mathcal{C} has exactly one cycle (possibly a loop) ZZ. All vertices in 𝒞\mathcal{C} have a unique shortest directed path to ZZ. For v𝒞v\in\mathcal{C} denote by (v)\ell(v) the length of this path and by (𝒞)\ell(\mathcal{C}) the maximum among all these lengths. Moreover, denote by z(𝒞)z(\mathcal{C}) the length of ZZ.

Theorem 3.2.

A 11-outregular digraph GG is a monoid digraph if and only if GG has a component 𝒞\mathcal{C} such that z(𝒟)z(\mathcal{D}) divides z(𝒞)z(\mathcal{C}) and (𝒟)(𝒞)\ell(\mathcal{D})\leq\ell(\mathcal{C}) for all components 𝒟\mathcal{D} of GG.

Proof.

By Observation 3.1, if G=(V,A)G=(V,A) is a monoid digraph, then G=Cay(M,{a})G=\mathrm{Cay}(M,\{a\}) for some monoid MM and aMa\in M. Let kk and hh be the pre-period and the period of aa, respectively. This is, e,a,,ak+h1e,a,...,a^{k+h-1} are pairwise distinct and ak+h=aka^{k+h}=a^{k}. Hence, the component 𝒞\mathcal{C} containing aa satisfies z(𝒞)=hz(\mathcal{C})=h and (𝒞)k\ell(\mathcal{C})\geq k. Now let 𝒟\mathcal{D} be any component of GG, x𝒟x\in\mathcal{D} at distance p=(𝒟)p=\ell(\mathcal{D}) from the cycle of 𝒟\mathcal{D}, and q=z(𝒟)q=z(\mathcal{D}). If qq does not divide hh, then kk and k+hk+h are not congruent modulo qq and thus xakxak+hxa^{k}\neq xa^{k+h}, a contradiction. Hence z(𝒟)z(\mathcal{D}) divides z(𝒞)z(\mathcal{C}). Now suppose pk+1p\geq k+1. Then xakxa^{k} is distinct from xalxa^{l} for each lkl\neq k, but this is also a contradiction. Hence pk(𝒞)p\leq k\leq\ell(\mathcal{C}).

Now suppose that the condition is fulfilled. For x,yVx,y\in V denote by d(x,y)d(x,y) the length of the shortest directed path from xx to yy, if it exists. Moreover, for x𝒟x\in\mathcal{D} denote z(x):=z(𝒟)z(x):=z(\mathcal{D}). In 𝒞\mathcal{C} choose a vertex at distance (𝒞)\ell(\mathcal{C}) from the cycle Z𝒞Z\subseteq\mathcal{C}. Conveniently call this vertex ee and call its out-neighbor aa. For any vertex xVx\in V and any kk\in\mathbb{N}, define x+kx+k as the end vertex of the directed walk of length kk starting at xx. It is clear that for any k,k,\ell\in\mathbb{N} the vertices x+k,x+x+k,x+\ell are equal if and only if d(x,x+k)=d(x,x+)d(x,x+k)=d(x,x+\ell).

Claim.

For every xVx\in V and every kk\in\mathbb{N} we have d(x,x+k)=d(x,x+d(e,e+k))d(x,x+k)=d(x,x+d(e,e+k)).

Proof.

Observe that

d(x,x+k)={kif k<(x)(x)+((k(x))modz(x))otherwise,d(x,x+k)=\begin{cases}k&\text{if }k<\ell(x)\\ \ell(x)+((k-\ell(x))\!\!\!\mod z(x))&\text{otherwise,}\end{cases}

where (k(x))modz(x)(k-\ell(x))\!\mod z(x) is the remainder of the Euclidean division of k(x)k-\ell(x) by z(x)z(x). Now we distinguish two cases. If k<(e)k<\ell(e) then d(e,e+k)=kd(e,e+k)=k. If otherwise k(e)k\geq\ell(e), write k=(e)+z(e)qk+rkk=\ell(e)+z(e)q_{k}+r_{k}, where qk,rkq_{k},r_{k} are the integers determined by 0rk<z(e)0\leq r_{k}<z(e). By hypothesis, k(e)(x)k\geq\ell(e)\geq\ell(x) and z(x)z(e)z(x)\mid z(e), so d(x,x+d(e,e+k))=d(x,x+(e)+rk)=d(x,x+(e)+z(e)qk+rk)=d(x,x+k)d(x,x+d(e,e+k))=d(x,x+\ell(e)+r_{k})=d(x,x+\ell(e)+z(e)q_{k}+r_{k})=d(x,x+k). \square

Now, let ω=e+((𝒞)+z(𝒞)1)\omega=e+(\ell(\mathcal{C})+z(\mathcal{C})-1). Observe that ωZ𝒞\omega\in Z\subseteq\mathcal{C}. Therefore ω\omega can be reached from any vertex vv of 𝒞\mathcal{C}, so we can set r(v):=d(e,ω)d(v,ω)r(v):=d(e,\omega)-d(v,\omega). Note that we have d(v,ω)=d(e+r(v),ω)d(v,\omega)=d(e+r(v),\omega). This implies that d(v+k,ω)=d(e+r(v)+k,ω)d(v+k,\omega)=d(e+r(v)+k,\omega) for any kk\in\mathbb{N}, so r(v+k)=r(e+r(v)+k)=d(e,e+r(v)+k)r(v+k)=r(e+r(v)+k)=d(e,e+r(v)+k) for any kk\in\mathbb{N}. We regard the vertices of GG as the elements of a monoid MM equipped with the following multiplication:

yV\displaystyle\forall y\in V\ \ ey=y;\displaystyle ey=y;
xV{e}yV\displaystyle\forall x\in V\!\setminus\!\{e\}\ \;\forall y\in V\ \ xy={x+r(y)if y𝒞yotherwise.\displaystyle xy=\begin{cases}x+r(y)&\text{if }y\in\mathcal{C}\\ y&\text{otherwise.}\end{cases}

By definition, ee is the right-neutral element, but also xe=x+r(e)=xxe=x+r(e)=x. Hence ee is the neutral element of this operation. Let us check that the operation is associative. Pick any x,y,zVx,y,z\in V. If y𝒞y\notin\mathcal{C}, then yz𝒞yz\notin\mathcal{C} and (xy)z=yz=x(yz)(xy)z=yz=x(yz). If z𝒞z\notin\mathcal{C} then (xy)z=z=x(yz)(xy)z=z=x(yz). So suppose y,z𝒞y,z\in\mathcal{C}. Then using the above Claim we can transform, (xy)z=(x+r(y))+r(z)=x+d(x,x+r(y)+r(z))=x+d(x,x+d(e,e+r(y)+r(z)))=x+d(e,e+r(y)+r(z))=x+r(y+r(z))=x+r(yz)=x(yz)(xy)z=(x+r(y))+r(z)=x+d(x,x+r(y)+r(z))=x+d(x,x+d(e,e+r(y)+r(z)))=x+d(e,e+r(y)+r(z))=x+r(y+r(z))=x+r(yz)=x(yz). Finally, note that (x,y)(x,y) is an arc of GG if and only if y=x+1=xay=x+1=xa. Therefore, G=Cay(M,{a})G=\mathrm{Cay}(M,\{a\}). \square

Corollary 3.3 (Zelinka’s Theorem).

A 11-outregular digraph GG is a semigroup digraph if and only GG has a component 𝒞\mathcal{C} such that z(𝒟)z(\mathcal{D}) divides z(𝒞)z(\mathcal{C}) and (𝒟)(𝒞)+1\ell(\mathcal{D})\leq\ell(\mathcal{C})+1 for all components 𝒟\mathcal{D} of GG.

Proof.

Again by Observation 3.1 we can suppose that G=Cay(S,{a})G=\mathrm{Cay}(S,\{a\}) for a semigroup SS and aSa\in S. Let 𝒞\mathcal{C} be the connected component of aa and SS^{\prime} the monoid obtained from SS by adjoining a neutral element ee. We can establish a bijective correspondence 𝒟𝒟\mathcal{D}\mapsto\mathcal{D}^{\prime} between connected components of GG and Cay(S,{a})\mathrm{Cay}(S^{\prime},\{a\}) as follows: 𝒟=𝒟\mathcal{D}^{\prime}=\mathcal{D} if 𝒟𝒞\mathcal{D}\neq\mathcal{C}, and 𝒞\mathcal{C}^{\prime} is obtained from 𝒞\mathcal{C} by adding the vertex ee and the arc (e,a)(e,a). Observe that z(𝒞)=z(𝒞)z(\mathcal{C}^{\prime})=z(\mathcal{C}) and (𝒞)(𝒞)+1\ell(\mathcal{C}^{\prime})\leq\ell(\mathcal{C})+1. By Theorem 3.2, z(𝒟)z(𝒞)z(\mathcal{D})\mid z(\mathcal{C}) and (𝒟)(𝒞)+1\ell(\mathcal{D})\leq\ell(\mathcal{C})+1 for any connected component 𝒟\mathcal{D} of GG.

Now suppose that G=(V,A)G=(V,A) fulfills the condition. Pick a vertex vv in 𝒞\mathcal{C} with (v)=(𝒞)\ell(v)=\ell(\mathcal{C}), and from GG construct a new 11-outregular graph GG^{\prime} by adding a new vertex uu and the arc (u,v)(u,v). By Theorem 3.2, GG^{\prime} is isomorphic to Cay(M,{a})\mathrm{Cay}(M,\{a\}) for some monoid MM and some aMa\in M. It follows from its proof that uu can be assumed to be the neutral element ee and vv to be aa. Additionally, since ee has no in-neighbors, it can be checked that M{e}M\setminus\{e\} is closed under the constructed product. Therefore, it is a semigroup, and GG is isomorphic to Cay(M{e},{a})\mathrm{Cay}(M\!\setminus\!\{e\},\{a\}). \square

The following has been claimed before [38, Proposition 11.3.2]. However, also in this proof there is a slight problem in the monoid construction. Here we get it as a direct consequence of Theorem 3.2:

Corollary 3.4.

Any connected 11-outregular digraph is a monoid digraph.

It was asked in [38, below Proposition 11.3.2] if Corollary 3.4 could be extended to kk-outregular semigroup digraphs. Here we give a negative answer in a strong sense. A subset of vertices VV^{\prime} of a (strongly) connected (directed) graph GG is called (directed) vertex cut if GVG\setminus V^{\prime} is not (strongly) connected. If κ\kappa is an integer such that all (directed) vertex cuts of GG are of order at least κ\kappa and κ+1n\kappa+1\leq n, where nn is the order of GG, then GG is (strongly) κ\kappa-connected.

Theorem 3.5.

For every k>1k>1 there exist infinitely many strongly connected kk-outregular non-semigroup digraphs. Moreover, for every κ>0\kappa>0 with kκ>1\left\lfloor\frac{k}{\kappa}\right\rfloor>1 there are such digraphs that are strongly κ\kappa-connected, and whose underlying undirected graph is (k+κ)(k+\kappa)-connected. For kκ>2\left\lfloor\frac{k}{\kappa}\right\rfloor>2 such digraphs exist that cannot even be turned into semigroup digraphs by adding loops.

Proof.

Given two positive integers k,k,\ell, let Gk,G_{k,\ell} be the directed graph with vertex set {(0,j)2| 0jk21}{(i,j)2| 1i1, 0jk1}\{(0,j)\in\mathbb{Z}^{2}\ |\ 0\leq j\leq k^{2}-1\}\cup\{(i,j)\in\mathbb{Z}^{2}\ |\ 1\leq i\leq\ell-1,\,0\leq j\leq k-1\} which has an arc from (i1,j1)(i_{1},j_{1}) to (i2,j2)(i_{2},j_{2}) if and only if i1+1=i2i_{1}+1=i_{2} or alternatively i1=1i_{1}=\ell-1, i2=0i_{2}=0 and j1=j2kj_{1}=\left\lfloor\frac{j_{2}}{k}\right\rfloor. Given a positive integer κ\kappa, let κ\sim_{\kappa} be an equivalence relation on {(0,j)2| 0jk21}\{(0,j)\in\mathbb{Z}^{2}\ |\ 0\leq j\leq k^{2}-1\} defined as (0,j1)κ(0,j2)(0,j_{1})\sim_{\kappa}(0,j_{2}) if and only if the following conditions are satisfied:

  • (i)

    j1j2modkj_{1}\equiv j_{2}\mod k;

  • (ii)

    j1kκ=j2kκ\left\lfloor\frac{j_{1}}{k\kappa}\right\rfloor=\left\lfloor\frac{j_{2}}{k\kappa}\right\rfloor, or κk\kappa\nmid k and j1kκ+1=j2kκ=k21kκ\left\lfloor\frac{j_{1}}{k\kappa}\right\rfloor+1=\left\lfloor\frac{j_{2}}{k\kappa}\right\rfloor=\left\lfloor\frac{k^{2}-1}{k\kappa}\right\rfloor (or the symmetric condition).

We call Gk,,κG_{k,\ell,\kappa} the directed graph obtained from Gk,G_{k,\ell} by merging all vertices of each κ\sim_{\kappa}-class. See Figure 2 for an illustration.

Refer to caption
Figure 2: Left: the directed graph G4,5G4,5,1G_{4,5}\cong G_{4,5,1}. Right: the merged directed graph G4,5,2G_{4,5,2}.

Note that for 2\ell\geq 2 the digraph Gk,,κG_{k,\ell,\kappa} is kk-outregular and strongly connected. Denote by ViV_{i} the set of vertices of Gk,,κG_{k,\ell,\kappa} with first coordinate congruent to ii modulo \ell if i\ell\nmid i, and the set of κ\sim_{\kappa}-classes if i\ell\mid i. In particular, |V0|=kκk|V_{0}|=\left\lfloor\frac{k}{\kappa}\right\rfloor k and |V1|=|V1|=k|V_{1}|=...|V_{\ell-1}|=k. Assume that kκ>1\left\lfloor\frac{k}{\kappa}\right\rfloor>1 and 4\ell\geq 4.

It is clear that Gk,,κG_{k,\ell,\kappa} is not strongly (κ+1)(\kappa+1)-connected: for instance, all in-neighbors of any vertex of V0V_{0} form a directed vertex cut. Now let XX be a minimal directed vertex cut. Suppose that there is a vertex xXVix\in X\cap V_{i} with i1modi\not\equiv-1\mod\ell. Then since |X|κ<k|X|\leq\kappa<k there is in ViXV_{i}\setminus X a vertex with the same in-neighbors and out-neighbors as xx, so X{x}X\setminus\{x\} is also a directed vertex cut, a contradiction. Therefore XV1X\subseteq V_{-1}. Hence there must be a vertex in V0V_{0} with all in-neighbors in XX, or XX would not be a directed vertex cut. This implies that |X|=κ|X|=\kappa.

Moreover the underlying undirected graph of Gk,,κG_{k,\ell,\kappa} is (k+κ)(k+\kappa)-connected. First note that V1YV_{1}\cup Y, where YY is the set of in-neighbors of a vertex of V0V_{0}, is a vertex cut. Now let XX be a minimal vertex cut; it is clear that ViXV_{i}\subseteq X for some ii. This ii is unique (modulo \ell); more precisely, it is the only index satisfying |XVi|k|X\cap V_{i}|\geq k. Suppose that there is a vertex xXVjx\in X\cap V_{j} with j1,imodj\not\equiv-1,i\mod\ell. Then there is a vertex in VjXV_{j}\setminus X with the same neighbors as xx, so X{x}X\setminus\{x\} is also a vertex cut, a contradiction. Therefore XV1ViX\subseteq V_{-1}\cup V_{i}. Since 3\ell\geq 3 it is clear that V1,V0ViV_{-1},V_{0}\neq V_{i}. Moreover, since 4\ell\geq 4 there is a vertex yV0Xy\in V_{0}\setminus X with all its in-neighbors in XX. This implies that k+κ|Vi|+|Y||X|k+κk+\kappa\leq|V_{i}|+|Y|\leq|X|\leq k+\kappa, where YY is the set of in-neighbors of yy.

We will now show that:

  • (a)

    under the assumed hypotheses on k,,κk,\ell,\kappa, Gk,,κG_{k,\ell,\kappa} is not a semigroup digraph;

  • (b)

    if additionally kκ>2+1k\left\lfloor\frac{k}{\kappa}\right\rfloor>2+\frac{1}{k}, then there is no semigroup SS and CSC\subseteq S for which Gk,,κ=Cay(S,C)G_{k,\ell,\kappa}=\mathrm{Cay}^{\mathrlap{\circlearrowleft}\,\setminus}(S,C), where Cay(S,C)\mathrm{Cay}^{\mathrlap{\circlearrowleft}\,\setminus}(S,C) denotes the graph obtained from Cay(S,C)\mathrm{Cay}(S,C) by removing all loops.

In both cases a hypothetical representation with a semigroup SS and a connection set CC leads to a contradiction. Indeed, let xSx\in S and XSX\subseteq S. By Lemma 2.5 SS is left cancellative, so |xX|=|X||xX|=|X|. Now suppose that xV2x\notin V_{-2} and let yV2y\in V_{-2}. We have k=|xC2|=|C2|=|yC2|=kκkk=|xC^{2}|=|C^{2}|=|yC^{2}|=\left\lfloor\frac{k}{\kappa}\right\rfloor k if there is a semigroup representation in case (a) and 2k+1|xC2|=|C2|=|yC2|kκk2k+1\geq|xC^{2}|=|C^{2}|=|yC^{2}|\geq\left\lfloor\frac{k}{\kappa}\right\rfloor k if there is a semigroup representation in case (b), the desired contradictions. \square

On the other hand we also know a smallest outregular non-semigroup digraph.

Proposition 3.6.

There is an outregular non-semigroup digraph on three vertices and this is smallest possible.

Proof.

Consider the directed graph GG from Figure 3 with vertex set {x,y,z}\{x,y,z\} and arc set

{(x,x),(x,y),(y,x),(y,z),(z,y),(z,z)}.\{(x,x),(x,y),(y,x),(y,z),(z,y),(z,z)\}.

Suppose that G=Cay(S,C)G=\mathrm{Cay}(S,C) for a semigroup SS and some CSC\subseteq S. It is clear that there is no uSu\in S with ux=yux=y or uz=yuz=y. On the other hand, c,cCxc=zc=y\exists c,c^{\prime}\in C\ \,xc=zc^{\prime}=y. Therefore, c=c=yCc=c^{\prime}=y\in C. Now suppose that y2=xy^{2}=x. Then yx=xy=yyx=xy=y, a contradiction. Similarly, y2zy^{2}\neq z. Hence y2=yy^{2}=y, but this is also a contradiction, because yy is loopless.

Refer to caption
Figure 3: A smallest outregular non-semigroup digraph.

Moreover, there is no kk-outregular non-semigroup digraph with less than three vertices. Since all 0-outregular digraphs and all 11-outregular connected digraphs are monoid digraphs (by Theorem 3.2), we only have to worry about 11-outregular disconnected digraphs and 22-outregular digraphs. But with less than three vertices there is only one of each kind: the digraph formed by two vertices with a loop each, and the complete digraph with loops of two vertices; and they are respectively isomorphic to Cay(2,{0})\mathrm{Cay}(\mathbb{Z}_{2},\{0\}) and Cay(2,{0,1})\mathrm{Cay}(\mathbb{Z}_{2},\{0,1\}), where 2\mathbb{Z}_{2} is the additive group on the integers modulo 22. \square

Remark 3.7.

By examining all cases, one can show that there are in total exactly 33 outregular non-semigroup digraphs on 33 vertices. We omit this discussion here for the sake of brevity.

It is known that every graph is isomorphic to some induced subgraph of a Cayley graph of a group [4]. Here we show that any sink-free directed graph can be obtained from a monoid digraph by removing a connected component. Note that it is necessary to require the digraph to be sink-free. Similar constructions have been used before (see e.g. [14, Theorem 2.2], also see the standard concept of transition monoid in automata theory [49, Chapter IV.3]). Given a graph GG (directed or undirected) and a vertex vv of GG, we denote by GvG_{v} the connected component of vv.

Proposition 3.8.

Let G=(V,A)G=(V,A) be a digraph, \mathcal{F} a set of mappings VVV\rightarrow V satisfying

  1. i)

    fvV(v,f(v))A\forall f\in\mathcal{F}\ \ \forall v\in V\ \ (v,f(v))\in A,

  2. ii)

    (v,w)Aff(v)=w\forall(v,w)\in A\ \ \exists f\in\mathcal{F}\ \ f(v)=w,

and \langle\mathcal{F}\rangle the monoid generated by \mathcal{F} with respect to composition. The set M=VM=V\sqcup\mathcal{\langle}\mathcal{F}\rangle admits a monoid structure such that GCay(M,)Cay(M,)eG\cong\mathrm{Cay}(M,\mathcal{F})\setminus\mathrm{Cay}(M,\mathcal{F})_{e}.

Proof.

Consider the following multiplication \cdot on MM:

x,yV\displaystyle\forall x,y\in V\ \, xy=y,\displaystyle x\cdot y=y,
xVf\displaystyle\forall x\in V\ \,\forall f\in\langle\mathcal{F}\rangle\ \, xf=f(x),\displaystyle x\cdot f=f(x),
xVf\displaystyle\forall x\in V\ \,\forall f\in\langle\mathcal{F}\rangle\ \, fx=x,\displaystyle f\cdot x=x,
f,g\displaystyle\forall f,g\in\langle\mathcal{F}\rangle\ \, fg=gf.\displaystyle f\cdot g=g\circ f.

Note that id\textrm{id}\in\langle\mathcal{F}\rangle serves as the neutral element of \cdot. To see that \cdot is associative, pick any α,β,γM\alpha,\beta,\gamma\in M. If γV\gamma\in V, then (αβ)γ=γ=αγ=α(βγ)(\alpha\cdot\beta)\cdot\gamma=\gamma=\alpha\cdot\gamma=\alpha\cdot(\beta\cdot\gamma). Thus, let γ\gamma\in\langle\mathcal{F}\rangle. In the case βV\beta\in V we have (αβ)γ=γ(β)=αγ(β)=α(βγ)(\alpha\cdot\beta)\cdot\gamma=\gamma(\beta)=\alpha\cdot\gamma(\beta)=\alpha\cdot(\beta\cdot\gamma); in the case αV\alpha\in V, β\beta\in\langle\mathcal{F}\rangle we have (αβ)γ=γ(β(α))=α(γβ)=α(βγ)(\alpha\cdot\beta)\cdot\gamma=\gamma(\beta(\alpha))=\alpha\cdot(\gamma\circ\beta)=\alpha\cdot(\beta\cdot\gamma); and the case α,β\alpha,\beta\in\mathcal{\langle}\mathcal{F}\rangle is trivial. Let H=Cay(M,)H=\mathrm{Cay}(M,\mathcal{F}). It is clear that the subgraph H[]H[\langle\mathcal{F}\rangle] induced by \langle\mathcal{F}\rangle is a connected component of HH, and it coincides with HidH_{\textrm{id}}. The subgraph H[V]H[V] induced by VV is formed by the rest of the connected components, and by the hypotheses on \mathcal{F}, it is isomorphic to GG. \square

Proposition 3.8 has the last result of this section as a corollary.

Corollary 3.9.

Let GG be a directed graph in which each vertex has at most kk out-neighbors, and at least 11. Then there is a monoid MM and CMC\subseteq M with |C|=k|C|=k such that GCay(M,C)Cay(M,C)eG\cong\mathrm{Cay}(M,C)\setminus\mathrm{Cay}(M,C)_{e}.

Proof.

One can greedily cover the arcs of GG by kk spanning 11-outregular subdigraphs: to construct the iith such digraph GiG_{i} simply take any vertex xx that has not yet outdegree 11 in GiG_{i}, add any not yet covered arc (x,y)(x,y) to GiG_{i} or an already covered one (x,y)(x,y) if all are covered, and iterate.

Now each GiG_{i} yields a function fif_{i} that just maps every vertex xx to its out-neighbor yy in GiG_{i}. The resulting set \mathcal{F} has order kk and Proposition 3.8 yields the result. \square

4 Monoid graphs

Until here we have examined directed graphs, but now we forget about orientations and loops: we concentrate on the underlying simple undirected graphs. Also, we shift our attention from semigroups in general to just monoids. The aim of this section is to give several examples of both monoid and non-monoid graphs.

Before starting it is worth noting that in this particular setting some assumptions about the connection set can be made. Let MM be a monoid and CMC\subseteq M. First, the graph Cay¯(M,C)\underline{\mathrm{Cay}}(M,C) does not depend on the pertinence of ee to CC, so it can be assumed that eCe\notin C. Second, since now the direction of the arcs is not important, CC can be assumed to be closed under taking existing left-inverses. This is:

Lemma 4.1.

Let MM be a monoid, CMC\subseteq M and G=Cay¯(M,C)G=\underline{\mathrm{Cay}}(M,C). Then G=Cay¯(M,N(e))G=\underline{\mathrm{Cay}}(M,N(e)), where N(e)N(e) is the set of neighbors of the neutral element.

Proof.

Since CN(e)C\subseteq N(e) all we have to show is that every edge {v,w}\{v,w\} of Cay¯(M,N(e))\underline{\mathrm{Cay}}(M,N(e)) is an edge of GG. Suppose that w=vxw=vx for some xN(e)Cx\in N(e)\setminus C. Then xc=exc=e for some cCc\in C. This implies that {v,w}\{v,w\} is also an edge of GG, because wc=vxc=vwc=vxc=v. \square

We begin with some positive results. A graph G=(V,E)G=(V,E) is called a threshold graph if there exist non-negative reals tt and wvw_{v}, vVv\in V, such that for every UVU\subseteq V we have that vUwvt\sum_{v\in U}w_{v}\leq t if and only if UU is an independent set. Equivalently, a threshold graph is a graph that can be obtained from the one-vertex graph by repeatedly adding an isolated vertex or a vertex which is connected to all the others. More equivalent definitions can be found in [42, Theorem 1.2.4]. As a consequence of Proposition 4.2, these graphs are monoid graphs.

Proposition 4.2.

Let G=(V,E)G=(V,E) be a monoid graph and let xVx\notin V. Then both G=(V{x},E)G^{\prime}=(V\cup\{x\},E) and G′′=(V{x},EvV{v,x})G^{\prime\prime}=(V\cup\{x\},E\cup\bigcup_{v\in V}\{v,x\}) are monoid graphs.

Proof.

Let MM be a monoid and CMC\subseteq M with G=Cay¯(M,C)G=\underline{\mathrm{Cay}}(M,C). Define a monoid structure on M=V{x}M^{\prime}=V\cup\{x\} using the multiplication from MM whenever possible, and defining xv=vx=x2=xxv=vx=x^{2}=x for every vVv\in V. This new multiplication preserves the neutral element from MM, and it is also associative: let u,v,wMu,v,w\in M^{\prime} and suppose than at least one of them is equal to xx; then, (uv)w=x=u(vw)(uv)w=x=u(vw). It is straightforward to check that G=Cay¯(M,C)G^{\prime}=\underline{\mathrm{Cay}}(M^{\prime},C) and G′′=Cay¯(M,C{x})G^{\prime\prime}=\underline{\mathrm{Cay}}(M^{\prime},C\cup\{x\}). \square

Corollary 4.3.

Threshold graphs are monoid graphs.

Another class of monoid graphs is the class of forests. This is a corollary to the monoid version of Zelinka’s theorem (Theorem 3.2).

Corollary 4.4.

For every forest FF there is a monoid MM and aMa\in M such that FF is isomorphic to Cay¯(M,{a})\underline{\mathrm{Cay}}(M,\{a\}).

Proof.

Let FF be any forest. Pick in each component 𝒞\mathcal{C} an arbitrary vertex vv. Direct all edges of 𝒞\mathcal{C} towards vv and add a loop at vv. Clearly, the resulting digraph satisfies the hypothesis of Theorem 3.2: every component 𝒞\mathcal{C} has z(𝒞)=1z(\mathcal{C})=1, and some of them maximize (𝒞)\ell(\mathcal{C}). \square

Corollary 4.4 leads to the idea of covering a graph with several forests in order to find a monoid representation. More precisely, a pseudoforest is a graph having at most one cycle per connected component. The pseudoarboricity p(G)p(G) (resp. arboricity a(G)a(G)) of a graph GG is the minimum number of pseudoforests (resp. forests) needed to cover all the edges of GG. These invariants can be computed in polynomial time [12], and results from [45, 10, 48] yield the following formulas:

a(G)\displaystyle a(G) =maxSV(G)|E(G[S])||S|1\displaystyle=\underset{S\subseteq V(G)}{\max}\left\lceil\frac{|E(G[S])|}{|S|-1}\right\rceil
p(G)\displaystyle p(G) =maxSV(G)|E(G[S])||S|,\displaystyle=\underset{S\subseteq V(G)}{\max}\left\lceil\frac{|E(G[S])|}{|S|}\right\rceil\!,

where E(G[S])E(G[S]) is the set of edges of the subgraph of GG induced by SS. Not too surprisingly, in the undirected setting the pseudoarboricity has a similar role to that of kk-outregularity in Section 3. The pseudoarboricity of a semigroup graph is bounded by the size of the connection set in any semigroup representation. More precisely:

Observation 4.5.

If G=Cay¯(S,C)G=\underline{\mathrm{Cay}}(S,C) is a semigroup graph, then p(G)|C|p(G)\leq|C|.

Corollary 3.9 together with the fact that p(G)p(G) is the minimum, over all orientations GG^{\prime} of GG, of the maximum outdegree of GG^{\prime}, gives:

Corollary 4.6.

Let GG be a graph with pseudoarboricity kk. Then there is a monoid MM and CMC\subseteq M with |C|=k|C|=k such that GCay¯(M,C)Cay¯(M,C)eG\cong\underline{\mathrm{Cay}}(M,C)\setminus\underline{\mathrm{Cay}}(M,C)_{e}.

Corollary 4.4 can be restated by saying that graphs of arboricity 11 are monoid graphs. One could ask what happens with higher arboricities. In Proposition 4.7 we show an infinite family of planar non-monoid graphs of arboricity 22 and treewidth 33.

Proposition 4.7.

The disjoint union K4CK_{4}\cup C_{\ell} is non-monoid, for all >1\ell>1 with 2,32,3\nmid\ell.

Proof.

Assume for a contradiction that Cay¯(M,C)=K4C\underline{\mathrm{Cay}}(M,C)=K_{4}\cup C_{\ell} for some monoid MM and CMC\subseteq M. Denote by K4\overrightarrow{K_{4}}, C\overrightarrow{C_{\ell}} the corresponding components of Cay(M,C)\mathrm{Cay}(M,C), that are directed graphs with possible loops and anti-parallel arcs. Since \ell is odd, there are two consecutive arcs in C\overrightarrow{C_{\ell}}. Hence, there is a vertex uu and c,cCc,c^{\prime}\in C such that the vertices u,uc,uccu,uc,ucc^{\prime} are all different. If the neutral element ee is in K4K_{4}, then e,cce,cc^{\prime} are either neighbors or the same vertex. But then u,uccu,ucc^{\prime} should also be neighbors or the same vertex, a contradiction.

Therefore ee must be in CC_{\ell}. Hence by Lemma 4.1 we can suppose that C={c,c}C=\{c,c^{\prime}\} with ccc\neq c^{\prime}. The periods of c,cc,c^{\prime} can only be 11, 22 or \ell, so by Theorem 3.2 both Cay¯(M,{c})\underline{\mathrm{Cay}}(M,\{c\}) and Cay¯(M,{c})\underline{\mathrm{Cay}}(M,\{c^{\prime}\}) are pseudoforests without cycles of length 33 or 44. Then each element of CC corresponds to at most 33 non-loop arcs of K4\overrightarrow{K_{4}}, and it must correspond to exactly 33, because K4K_{4} has 66 edges; moreover, none of the edges of K4K_{4} correspond to both c,cc,c^{\prime}. Therefore, if we group the edges of K4K_{4} depending on whether they come from cc or from cc^{\prime}, we obtain two edge-disjoint copies of the path P4P_{4}. We now distinguish two cases.

If there is an edge in CC_{\ell} corresponding to both c,cc,c^{\prime}, any endomorphism mapping ee to a vertex of K4K_{4} implies the existence of a loop in K4\overrightarrow{K_{4}}. Lemma 2.4 guarantees the existence of such an endomorphism.

If otherwise every edge in CC_{\ell} corresponds to either cc or cc^{\prime}, since \ell is odd then there is a loop in C\overrightarrow{C_{\ell}}. Thus, by Lemma 2.4 there is a loop in K4\overrightarrow{K_{4}}. We can suppose that this loop corresponds to cc. Let xx be the vertex of the loop. Wherever xx lies along the cc-copy of P4P_{4}, it implies there are vertices vwv\neq w, different from xx, such that vc=w,wc=xvc=w,wc=x. This implies that e,c,c2e,c,c^{2} are all different. We know that cc{e,c,c2}cc^{\prime}\in\{e,c,c^{2}\}, so xc=(xc)c=x(cc)=xxc^{\prime}=(xc)c^{\prime}=x(cc^{\prime})=x, and there cannot be more loops in K4\overrightarrow{K_{4}}. Now, wc=vcc{v,w,x}wc^{\prime}=vcc^{\prime}\in\{v,w,x\}, but that is a contradiction. \square

Refer to caption
Figure 4: The graph K4C5K_{4}\cup C_{5} is the smallest non-monoid graph we know of.
Remark 4.8.

We have checked by computer, that all graphs on up to 66 vertices are monoid graphs. The instance of Proposition 4.7 on 99 vertices depicted in Figure 4 is the smallest non-monoid graph we know of. Is it the smallest one?

The construction and the argument in Proposition 4.7 are based on the fact that the graph is disconnected. We dedicate the remainder of the section to the construction of kk-connected non-monoid graphs for arbitrarily large kk (Theorem 4.12).

We start with some more technical definitions and lemmas. Given an undirected graph G=(V,E)G=(V,E), consider the set ={f:VE|vV:vf(v)}\mathcal{F}=\{f:V\rightarrow E\ |\ \forall v\in V:v\in f(v)\} and define β(G,k)=maxf1,,fkα((V,Ei=1kfi(V)))\beta(G,k)=\underset{f_{1},...,f_{k}\in\mathcal{F}}{\max}\;\alpha(\,(V,E\setminus\bigcup_{i=1}^{k}f_{i}(V))\,), where α\alpha denotes the independence number.

Lemma 4.9.

Let G=(V,E)G=(V,E) be a triangle-free nn-vertex graph with minimum degree δ\delta and maximum degree Δ2\Delta\geq 2. If for k0k\geq 0, >2Δ+2k+1\ell>2\Delta+2k+1 there is a monoid MM and CMC\subseteq M such that Cay¯(M,C)\underline{\mathrm{Cay}}(M,C) consists of GG and KK_{\ell} joined by kk edges, then β(G,k)nΔ1(δ2k1)\beta(G,k)\geq\frac{n}{\Delta-1}\left(\frac{\delta}{2}-k-1\right).

Proof.

By Lemma 4.1, we can suppose that the elements of CC are exactly the neighbors of ee. Hence, ee has to be a vertex of KK_{\ell}: otherwise |C|Δ+k|C|\leq\Delta+k, but this would imply that there are at most (Δ+k)+k<(1)2+k\ell(\Delta+k)+k<\frac{\ell(\ell-1)}{2}+k edges incident to vertices of KK_{\ell}.

Let EE^{\prime} be the set of edges joining KK_{\ell} and GG, and C={cCdCcdC{e}}C^{\prime}=\{c\in C\mid\exists d\in C\ cd\notin C\cup\{e\}\}. For each cCc\in C^{\prime} let εc\varepsilon_{c} be the edge {e,c}E\{e,c\}\in E^{\prime} if cVc\in V, or any edge of EE^{\prime} adjacent to cc if cVc\notin V. The map CEC^{\prime}\rightarrow E^{\prime} defined by cεcc\mapsto\varepsilon_{c} is injective, so |C||E|=k|C^{\prime}|\leq|E^{\prime}|=k.

For each uVu\in V define

indegC(u)=|{vV{u}cCCvc=u}|,\textrm{indeg}^{C^{\prime}}(u)=\left|\,\{v\in V\setminus\{u\}\mid\exists c\in C\setminus C^{\prime}\ vc=u\}\,\right|,
outdegC(u)=|{vV{u}cCCuc=v}|.\textrm{outdeg}^{C^{\prime}}(u)=\left|\,\{v\in V\setminus\{u\}\mid\exists c\in C\setminus C^{\prime}\ uc=v\}\,\right|.

Consider the set O={uV|outdegC(u)=0}O=\{u\in V\,|\,\textrm{outdeg}^{C^{\prime}}(u)=0\}. Let vVOv\in V\setminus O and wV{v}w\in V\setminus\{v\} with vc=wvc=w for some cCCc\in C\setminus C^{\prime}. Suppose that there are two different u1,u2V{v}u_{1},u_{2}\in V\setminus\{v\} and c1,c2CCc_{1},c_{2}\in C\setminus C^{\prime} with u1c1=u2c2=vu_{1}c_{1}=u_{2}c_{2}=v. We can suppose that u1wu_{1}\neq w. Observe that c1cC{e}c_{1}c\in C\cup\{e\}, because otherwise c1Cc_{1}\in C^{\prime}. In fact, since u1,v,wu_{1},v,w are pairwise different, c1cCc_{1}c\in C, but then these vertices form a triangle, a contradiction. Therefore indegC(v)1\textrm{indeg}^{C^{\prime}}(v)\leq 1. Now we can write

nδ2nk|E|n|C|uVindegC(u)=uOindegC(u)+uVOindegC(u)\frac{n\delta}{2}-nk\,\leq\,|E|-n|C^{\prime}|\,\leq\,\sum_{u\in V}\textrm{indeg}^{C^{\prime}}(u)\,=\,\sum_{u\in O}\textrm{indeg}^{C^{\prime}}(u)+\sum_{u\in V\setminus O}\textrm{indeg}^{C^{\prime}}(u)\,\leq
Δ|O|+|VO|=(Δ1)|O|+n.\leq\,\Delta|O|+|V\setminus O|\,=\,(\Delta-1)|O|+n.

Note that OO is an independent set of the graph (V,EcCfc(V))(V,E\setminus\bigcup_{c\in C^{\prime}}f_{c}(V)), where fc:VEf_{c}:V\rightarrow E is defined by fc(u)={u,uc}f_{c}(u)=\{u,uc\}. This yields the desired result. \square

For λ>0\lambda>0 a graph GG is an (n,d,λ)(n,d,\lambda)-graph if GG has nn vertices, is dd-regular, and all the Eigenvalues of its adjacency matrix are in [λ,λ]{d,d}[-\lambda,\lambda]\cup\{d,-d\}. The expander mixing lemma (see, e.g., [1, 2][18, Section 5] or [8, Lemma 2.1]) asserts that if S,TS,T are two sets of vertices in such a graph then

|e(S,T)d|S||T|n|λ|S||T|(1|S|n)(1|T|n),\left|e(S,T)-d\frac{|S||T|}{n}\right|\leq\lambda\sqrt{|S||T|\left(1-\frac{|S|}{n}\right)\left(1-\frac{|T|}{n}\right)},

where e(S,T)e(S,T) is the number of (ordered) edges uvuv with uS,vTu\in S,v\in T.

A standard application of the expander mixing lemma is bounding from above the independence number α(G)\alpha(G). Here we show that β(G,k)\beta(G,k) can be bounded in a similar manner.

Lemma 4.10.

If G=(V,E)G=(V,E) is an (n,d,λ)(n,d,\lambda)-graph and d1d\geq 1, then β(G,k)nd(λ+2k)\beta(G,k)\leq\frac{n}{d}(\lambda+2k) for all k0k\geq 0.

Proof.

Let f1,,fk={f:VE|vVvf(v)}f_{1},...,f_{k}\in\mathcal{F}=\{f:V\rightarrow E\ |\ \forall v\in V\ v\in f(v)\} and let SS be an independent set of (V,Ei=1kfi(V))(V,E\setminus\bigcup_{i=1}^{k}f_{i}(V)). The number of edges of the subgraph of GG induced by SS is at most k|S|k|S|. The expander mixing lemma yields

|S|nd(λ+2k),|S|\leq\frac{n}{d}(\lambda+2k),

and this implies the result. \square

Our strategy will be to confront the bounds of Lemma 4.9 and Lemma 4.10 by using (n,d,λ)(n,d,\lambda)-graphs in the construction of the examples. For this, we will also need some control over the connectivity of (n,d,λ)(n,d,\lambda)-graphs.

Lemma 4.11.

If G=(V,E)G=(V,E) is an (n,d,λ)(n,d,\lambda)-graph, d1d\geq 1 and 0k<(dλ)2d+10\leq k<\frac{(d-\lambda)^{2}}{d}+1, then GG is kk-connected.

Proof.

Suppose that GG is not kk-connected. Since k<(dλ)2d+1<d+1nk<\frac{(d-\lambda)^{2}}{d}+1<d+1\leq n (it can be always assumed that λd\lambda\leq d), we know that there exist non-empty S,T,UVS,T,U\subseteq V with SΓTΓU=VS\,\mathaccent 0{\cdot}\cup\,T\,\mathaccent 0{\cdot}\cup\,U=V, |U|=k1|U|=k-1, e(S,T)=0e(S,T)=0 and |S||T||S|\leq|T|.

Applying the expander-mixing lemma to S,TΓUS,T\,\mathaccent 0{\cdot}\cup\,U we get that

(dλ)|S||TΓU|ne(S,TΓU).\frac{(d-\lambda)|S||T\,\mathaccent 0{\cdot}\cup\,U|}{n}\leq e(S,T\,\mathaccent 0{\cdot}\cup\,U).

On the other hand, e(S,TΓU)=e(S,U)|S|(k1)e(S,T\,\mathaccent 0{\cdot}\cup\,U)=e(S,U)\leq|S|(k-1), so

(dλ)(n|S|)nk1.\frac{(d-\lambda)(n-|S|)}{n}\leq k-1.

Finally, we must have that |S|nλd|S|\leq\frac{n\lambda}{d}; otherwise, applying the expander mixing lemma to S,TS,T would lead to a contradiction. This yields

(dλ)2d+1k.\frac{(d-\lambda)^{2}}{d}+1\leq k.

\square

We are ready to prove the main theorem of this section.

Theorem 4.12.

For every non-negative integer kk there exists a kk-connected graph that is not a monoid graph.

Proof.

The family of (triangle-free) non-bipartite Ramanujan graphs of unbounded degree given by Lubotzky, Phillips and Sarnak [40] consists of (n,d,λ)(n,d,\lambda)-graphs with λ=2d1\lambda=2\sqrt{d-1}. Let GG be one of such graphs, also satisfying d>(6k+6+2)2=6k+46k+6+10d>(\sqrt{6k+6}+2)^{2}=6k+4\sqrt{6k+6}+10, and let >2d+2k+1\ell>2d+2k+1. Choose kk different vertices u1,,uku_{1},...,u_{k} of GG and kk different vertices v1,,vkv_{1},...,v_{k} of KK_{\ell}. Consider the graph HH consisting in GG, KK_{\ell} and the kk additional edges {u1,v1},,{uk,vk}\{u_{1},v_{1}\},...,\{u_{k},v_{k}\}.

Observe that the inequality (d2d1)2>(d2d)2(d-2\sqrt{d-1})^{2}>(d-2\sqrt{d})^{2} holds for d>8d>8. Hence we have (dλ)2d+1>(d2d)2d+1=(d2)2+1>6k+7>k\frac{(d-\lambda)^{2}}{d}+1>\frac{(d-2\sqrt{d})^{2}}{d}+1=(\sqrt{d}-2)^{2}+1>6k+7>k. By Lemma 4.11, GG is kk-connected, so HH is kk-connected.

Now suppose that HH is a monoid graph. Lemma 4.9 and Lemma 4.10 give

nd(d2k1)β(G,k)\displaystyle\frac{n}{d}\left(\frac{d}{2}-k-1\right)\leq\beta(G,k) nd(λ+2k)\displaystyle\leq\frac{n}{d}(\lambda+2k)
d2λ6k2\displaystyle d-2\lambda-6k-2 0\displaystyle\leq 0
d4d6k2\displaystyle d-4\sqrt{d}-6k-2 0,\displaystyle\leq 0,

but the set of solutions of the inequality x24x6k20x^{2}-4x-6k-2\leq 0 is [26k+6, 2+6k+6]\left[2-\sqrt{6k+6},\,2+\sqrt{6k+6}\,\right], so the above inequalities do not hold. This contradiction implies that HH is not a monoid graph. \square

5 Generated monoid trees

In the present section we investigate the more restricted class of monoid graphs, where the connection set generates the monoid. This restriction has received some attention with respect to semigroups, see [38, Chapter 11.3]. More precisely, we study the following problem: given a tree TT, determine whether there exist a monoid MM and CC with M=CM=\langle C\rangle such that T=Cay¯(M,C)T=\underline{\mathrm{Cay}}(M,C).

We know already, that trees are monoid graphs by Corollary 4.4. So the present section is an investigation on how much more restrictive the condition M=CM=\langle C\rangle is. In particular, we provide a sufficient condition (Theorem 5.1) and a necessary condition (Theorem 5.2) for a tree to be such a generated monoid graph, both stated in similar terms. However, a characterization remains open.

We start by introducing some notation borrowed from the context of distance-regular graphs [7, Section 20]. Let G=(V,E)G=(V,E) be a connected graph and ee a fixed vertex of GG. For every vertex xx we define the number of successors of xx with respect to ee as be(x)=|{yV|{x,y}E,d(e,y)=d(e,x)+1}|b_{e}(x)=\left|\{y\in V\ |\ \{x,y\}\in E,\ d(e,y)=d(e,x)+1\}\right|. To avoid unnecessary complications, throughout the section we will assume that in an undirected Cayley graph Cay¯(M,C)\underline{\mathrm{Cay}}(M,C) the neutral element is not in CC.

Theorem 5.1.

Let T=(V,E)T=(V,E) be a tree and eVe\in V. If for every x,yVx,y\in V with d(e,x)>d(e,y)d(e,x)>d(e,y) we have be(x)be(y)b_{e}(x)\leq b_{e}(y), then TT is a generated monoid graph.

Proof.

Let C={c1,,cn}C=\{c_{1},...,c_{n}\} the set of neighbors of ee. Let Σ={c¯1,,c¯n}\Sigma=\{\overline{c}_{1},...,\overline{c}_{n}\} be a set of nn different symbols. We construct a multi-digraph with Σ\Sigma-colored arcs T=(V,A)T^{\prime}=(V,A) which has TT as its underlying simple undirected graph. Given a vertex xVx\in V, call x1,,xbe(x)x_{1},...,x_{b_{e}(x)} its successors respect to ee; the successor xbe(x)x_{b_{e}(x)} will be also denoted by xix_{i} for i=be(x)+1,,ni=b_{e}(x)+1,...,n. If xx has no successor, then use the notation x=x1==xnx=x_{1}=...=x_{n}. The arcs of TT^{\prime} of color c¯i\overline{c}_{i} are precisely (x,xi)(x,x_{i}) for each xVx\in V. TT^{\prime} depends on the choice made when labeling the successors, but in what follows that will not matter.

Now, for each vertex xx and each finite word wΣw\in\Sigma^{*} from the alphabet Σ\Sigma, define xwx*w to be the ending vertex of the directed walk starting at xx and defined by ww. Observe that for any wΣw\in\Sigma^{*} d(e,ew)length(w)d(e,e*w)\leq\textrm{length}(w), which is an equality if ewe*w is not a leaf of TT. Let \sim be the equivalence relation in Σ\Sigma^{*} defined by wwew=eww\sim w^{\prime}\Leftrightarrow e*w=e*w^{\prime}. Construct a third graph T′′T^{\prime\prime}, also directed and with multiple Σ\Sigma-colored arcs, but this time with vertex set Σ/\Sigma^{*}/\!\sim: there is an arc of color c¯\overline{c} from [w1][w_{1}] to [w2][w_{2}] if and only if [w1c¯]=[w2][w_{1}\overline{c}]=[w_{2}]. Note that the map from Σ/\Sigma^{*}/\!\sim to VV sending [w][w] to ewe*w is an isomorphism between T′′T^{\prime\prime} and TT^{\prime}.

Endow Σ/\Sigma^{*}/\!\sim with the following multiplication: [w1][w2]=[w1w2][w_{1}][w_{2}]=[w_{1}w_{2}] for each w1,w2Σw_{1},w_{2}\in\Sigma^{*}. The following will be useful.

Claim.

For any w,wΣw,w^{\prime}\in\Sigma^{*} we have be(e(ww))be(ew)b_{e}(e*(ww^{\prime}))\leq b_{e}(e*w^{\prime}).

Proof.

In the case that d(e,e(ww))>d(e,ew)d(e,e*(ww^{\prime}))>d(e,e*w^{\prime}), this is implied by the assumed hypothesis. And in the opposite case, d(e,e(ww))d(e,ew)length(w)length(ww)d(e,e*(ww^{\prime}))\leq d(e,e*w^{\prime})\leq\textrm{length}(w^{\prime})\leq\textrm{length}(ww^{\prime}), so either there is a chain of equalities and ww is the empty word (in which case we are done), or e(ww)e*(ww^{\prime}) is a leaf, and then be(e(ww))=0b_{e}(e*(ww^{\prime}))=0. \square

We have to check that the definition is independent of the representatives. Let w1,w2,w1,w2Σw_{1},w_{2},w^{\prime}_{1},w^{\prime}_{2}\in\Sigma^{*} with w1w1,w2w2w_{1}\sim w^{\prime}_{1},w_{2}\sim w^{\prime}_{2}. Since e(w1w2)=(ew1)w2=(ew1)w2=e(w1w2)e*(w_{1}w_{2})=(e*w_{1})*w_{2}=(e*w^{\prime}_{1})*w_{2}=e*(w^{\prime}_{1}w_{2}), we only have to bother if w2w2w_{2}\neq w^{\prime}_{2}. We distinguish two cases:

If length(w2)=length(w2)\textrm{length}(w_{2})=\textrm{length}(w^{\prime}_{2}), write w2=d¯1d¯rw_{2}=\overline{d}_{1}...\overline{d}_{r} and w2=d¯1d¯rw^{\prime}_{2}=\overline{d}^{\prime}_{1}...\overline{d}^{\prime}_{r} with d¯i,d¯iΣ\overline{d}_{i},\overline{d}^{\prime}_{i}\in\Sigma for 1ir1\leq i\leq r. We have that, for each ii, either d¯i=d¯i\overline{d}_{i}=\overline{d}^{\prime}_{i} or be(e(d¯1d¯i1))j,jb_{e}(e*(\overline{d}_{1}...\overline{d}_{i-1}))\leq j,j^{\prime}, where d¯i=c¯j\overline{d}_{i}=\overline{c}_{j} and d¯i=c¯j\overline{d}^{\prime}_{i}=\overline{c}_{j^{\prime}}. We now use the Claim to conclude that be(e(w1d¯1d¯i1))be(e(d¯1d¯i1))b_{e}(e*(w_{1}\overline{d}_{1}...\overline{d}_{i-1}))\leq b_{e}(e*(\overline{d}_{1}...\overline{d}_{i-1})) for each ii. This implies that w1d¯1d¯rw1d¯1d¯rw_{1}\overline{d}_{1}...\overline{d}_{r}\sim w_{1}\overline{d}^{\prime}_{1}...\overline{d}^{\prime}_{r}.

If length(w2)<length(w2)\textrm{length}(w_{2})<\textrm{length}(w^{\prime}_{2}), write w2=w2′′w2′′′w^{\prime}_{2}=w^{\prime\prime}_{2}w^{\prime\prime\prime}_{2} with w2′′,w2′′′Σw^{\prime\prime}_{2},w^{\prime\prime\prime}_{2}\in\Sigma^{*} and length(w2′′)=length(w2)\textrm{length}(w^{\prime\prime}_{2})=\textrm{length}(w_{2}). We know that ew2=ew2e*w_{2}=e*w^{\prime}_{2} is a leaf, so e(w2w2′′′)=ew2e*(w_{2}w^{\prime\prime\prime}_{2})=e*w^{\prime}_{2}. Here length(w2w2′′′)=length(w2)\textrm{length}(w_{2}w^{\prime\prime\prime}_{2})=\textrm{length}(w^{\prime}_{2}), so ew2=ew2′′e*w_{2}=e*w^{\prime\prime}_{2}, which can be obtained following the argumentation of the first case. Again by the first case, e(w1w2)=e(w1w2′′)e*(w_{1}w_{2})=e*(w_{1}w^{\prime\prime}_{2}), and by the Claim, this vertex is a leaf of TT. Hence, e(w1w2)=(e(w1w2′′))w2′′′=e(w1w2)e*(w_{1}w_{2})=(e*(w_{1}w^{\prime\prime}_{2}))*w^{\prime\prime\prime}_{2}=e*(w_{1}w^{\prime}_{2}), and we are done.

The defined multiplication is associative; showing it is an easy task: ([w1][w2])[w3]=[(w1w2)w3]=[w1(w2w3)]=[w1]([w2][w3])([w_{1}][w_{2}])[w_{3}]=[(w_{1}w_{2})w_{3}]=[w_{1}(w_{2}w_{3})]=[w_{1}]([w_{2}][w_{3}]). Moreover, it has a neutral element, the \sim-class of the empty word. Therefore Σ/\Sigma^{*}/\!\sim with this operation is a monoid, and clearly T′′Caycol(Σ/,Σ/)T^{\prime\prime}\cong\mathrm{Cay}_{\mathrm{col}}(\Sigma^{*}/\!\sim,\Sigma/\!\sim). \square

Let T=(V,E)T=(V,E) be a tree and eVe\in V. For each cc neighbor of ee we define the branch of cc respect to ee as e(c)={xVd(c,x)<d(e,x)}\mathcal{B}_{e}(c)=\{x\in V\mid d(c,x)<d(e,x)\}. If xe(c)x\in\mathcal{B}_{e}(c), then we also say that e(c)\mathcal{B}_{e}(c) is the branch of xx respect to ee. The branches we will consider in generated monoid trees will be only with respect to (possible) neutral elements.

Given a monoid MM and xMx\in M, let φx:MM\varphi_{x}:M\rightarrow M be the map corresponding to left-multiplication by xx.

Proposition 5.2.

Let T=(V,E)T=(V,E) be a generated monoid tree, i.e., T=Cay¯(M,C)T=\underline{\mathrm{Cay}}(M,C) for some monoid MM and some CC with M=CM=\langle C\rangle. Then for each xV{e}x\in V\setminus\{e\} there is a yVy\in V with d(e,x)=d(e,y)+1d(e,x)=d(e,y)+1 and be(x)be(y)b_{e}(x)\leq b_{e}(y). Moreover, if Aut(T){φxxM}={id}\mathrm{Aut}(T)\cap\{\varphi_{x}\mid x\in M\}=\{\mathrm{id}\}, then for each vertex xx and each cCc\in C there is a vertex yce(c)y_{c}\in\mathcal{B}_{e}(c) with d(e,yc)d(e,x)+1d(e,y_{c})\leq d(e,x)+1 and be(yc)be(x)+εb_{e}(y_{c})\leq b_{e}(x)+\varepsilon, where ε={0 if xC{e},1 otherwise.\varepsilon=\begin{cases}0\text{ if }x\in C\cup\{e\},\\ 1\text{ otherwise.}\end{cases}

Proof.

We start with a useful fact.

Claim.

For each cCc\in C and each xMx\in M we have d(e,cx)d(e,x)+1d(e,cx)\leq d(e,x)+1.

Proof.

Let e=x0,x1,,xr=xe=x_{0},x_{1},...,x_{r}=x be the successive vertices of a shortest path between ee and xx, where r=d(e,x)r=d(e,x). Then the vertices e,c,cx1,,cxr=cxe,c,cx_{1},...,cx_{r}=cx, maybe after adding some loops, define a walk between ee and cxcx. Therefore d(e,cx)d(e,x)+1d(e,cx)\leq d(e,x)+1. \square

Given any xM{e}x\in M\setminus\{e\}, let cCc\in C with xe(c)x\in\mathcal{B}_{e}(c). Write x=c1crx=c_{1}...c_{r} with ciCc_{i}\in C, r=d(e,x)r=d(e,x) and c1=cc_{1}=c. Clearly y=c2cry=c_{2}...c_{r} satisfies d(e,y)+1rd(e,y)+1\leq r. By the Claim this is, in fact, an equality. For an element vMv\in M let o(v)=|vC{v}|\textrm{o}(v)=|vC\setminus\{v\}|. Observe that o(v)={be(v) if cCd(e,vc)=d(e,v)+1be(v)+1 otherwise\textrm{o}(v)=\begin{cases}b_{e}(v)&\text{ if }\forall c^{\prime}\in C\ d(e,vc^{\prime})=d(e,v)+1\\ b_{e}(v)+1&\text{ otherwise}\end{cases} and o(cv)o(v)\textrm{o}(cv)\leq\textrm{o}(v). Hence to show that be(x)be(y)b_{e}(x)\leq b_{e}(y) it will be enough to see that if o(y)=be(y)+1\textrm{o}(y)=b_{e}(y)+1 then o(x)=be(x)+1\textrm{o}(x)=b_{e}(x)+1. Now, o(y)=be(y)+1\textrm{o}(y)=b_{e}(y)+1 means that there is some zMz\in M and some c′′Cc^{\prime\prime}\in C with yc′′=zyc^{\prime\prime}=z and d(e,z)=d(e,y)1d(e,z)=d(e,y)-1. Again by the Claim, d(e,cz)d(e,z)+1=d(e,y)=d(e,x)1d(e,cz)\leq d(e,z)+1=d(e,y)=d(e,x)-1. Since xc′′=czxc^{\prime\prime}=cz, we have that o(x)=be(x)+1\textrm{o}(x)=b_{e}(x)+1.

For the second part, let yc=cxy_{c}=cx. Suppose that yce(c)y_{c}\notin\mathcal{B}_{e}(c). Take a walk between ee and xx, and let e=x0,x1,,xr=xe=x_{0},x_{1},...,x_{r}=x be its successive vertices. The vertices c=cx0,cx1,,cxr=cxc=cx_{0},cx_{1},...,cx_{r}=cx, maybe after adding some loops, define a walk between cc and cxcx. Therefore cxi=ecx_{i}=e for some ii. Since MM is finite, φc\varphi_{c} is an automorphism, contradicting that Aut(T){φxxM}={id}\mathrm{Aut}(T)\cap\{\varphi_{x}\mid x\in M\}=\{\mathrm{id}\}. Therefore, yce(c)y_{c}\in\mathcal{B}_{e}(c). The fact that be(yc)be(x)+1b_{e}(y_{c})\leq b_{e}(x)+1 is a consequence of o(cx)o(x)\textrm{o}(cx)\leq\textrm{o}(x). Furthermore, if xCx\in C then o(x)=be(x)\textrm{o}(x)=b_{e}(x), because otherwise φxAut(T)\varphi_{x}\in\mathrm{Aut}(T), so be(yc)be(x)b_{e}(y_{c})\leq b_{e}(x) in this case. If x=ex=e the same conclusion holds, because o(x)=be(x)\textrm{o}(x)=b_{e}(x) directly. Finally, what is left follows from the Claim. \square

The following lemma shows that the condition for the second part of Proposition 5.2 is not as restrictive as it may seem.

Lemma 5.3.

Let T=(V,E)T=(V,E) be a tree satisfying T=Cay¯(M,C)T=\underline{\mathrm{Cay}}(M,C) for some monoid MM and some CC with M=CM=\langle C\rangle. If Aut(T){φxxM}{id}\mathrm{Aut}(T)\cap\{\varphi_{x}\mid x\in M\}\neq\{\mathrm{id}\} then there is some cCc\in C such that c2=ec^{2}=e and Aut(T){φxxM}={id,φc}\mathrm{Aut}(T)\cap\{\varphi_{x}\mid x\in M\}=\{\mathrm{id,\varphi_{c}}\}.

Proof.

Let xM{e}x\in M\setminus\{e\} such that φxAut(T)\varphi_{x}\in\mathrm{Aut}(T). Write x=c1crx=c_{1}...c_{r} with ciCc_{i}\in C. Since MM is finite, φc1,,φcrAut(T)\varphi_{c_{1}},...,\varphi_{c_{r}}\in\mathrm{Aut}(T). This implies that c12==cr2=ec_{1}^{2}=...=c_{r}^{2}=e, or else TT would have a cycle. Now let c,cCc,c^{\prime}\in C with φc,φcAut(T)\varphi_{c},\varphi_{c^{\prime}}\in\mathrm{Aut}(T). If ccecc^{\prime}\neq e then there is a cycle in TT. Indeed, let kk be the order of cccc^{\prime}. It will be enough to show that the vertices e,c,cc,ccc,,(cc)k1,(cc)k1ce,c,cc^{\prime},cc^{\prime}c,...,(cc^{\prime})^{k-1},(cc^{\prime})^{k-1}c are all different. It is clear that e,cc,,(cc)k1e,cc^{\prime},...,(cc^{\prime})^{k-1} are pairwise different, and since right-multiplication by cc has to be bijective, so are c,(cc)c,,(cc)k1cc,(cc^{\prime})c,...,(cc^{\prime})^{k-1}c. Moreover there is no overlap between these two sets of vertices: the distance from ee is even in the first case and odd in the second (the presence of consecutive vertices in the succession e,c,cc,ccc,,(cc)k1,(cc)k1ce,c,cc^{\prime},cc^{\prime}c,...,(cc^{\prime})^{k-1},(cc^{\prime})^{k-1}c would imply that c=ec=e or c=ec^{\prime}=e after the cancellation of the left terms). Therefore, cc=ecc^{\prime}=e, which means that c=cc=c^{\prime}. \square

A drawback of Proposition 5.2 is that it can be difficult to use without knowing the (possible) location(s) of the neutral element. Luckily, the number of candidates can be sometimes restricted.

Lemma 5.4.

Let T=(V,E)T=(V,E) be a generated monoid tree, i.e., T=Cay¯(M,C)T=\underline{\mathrm{Cay}}(M,C) for some monoid MM and some CC with M=CM=\langle C\rangle and let Δ\Delta be the maximum degree of TT. We have Δ1deg(e)Δ\Delta-1\leq\deg(e)\leq\Delta.

Proof.

Let ueu\neq e be a vertex of TT. Suppose that in Cay(M,C)\mathrm{Cay}(M,C) there are two arcs (x1,u),(x2,u)(x_{1},u),(x_{2},u) incident to uu, and that the arcs (u,x1),(u,x2)(u,x_{1}),(u,x_{2}) do not exist. We know that in TT any vertex vv is reachable from ee by a unique shortest path P(v)P(v). The edge {x1,u}\{x_{1},u\} is the last one of either P(x1)P(x_{1}) or P(u)P(u). Since the corresponding paths in Cay(M,C)\mathrm{Cay}(M,C) are directed, from ee to the target vertex, {x1,u}\{x_{1},u\} is the last edge of P(u)P(u). By the same argument, so is {x2,u}\{x_{2},u\}, and this shows that in Cay(M,C)\mathrm{Cay}(M,C) vertex uu has at most one in-neighbor which is not an out-neighbor. Therefore deg(u)|C|+1\deg(u)\leq|C|+1, and the fact that deg(e)|C|\deg(e)\geq|C| yields the conclusion. \square

We apply the above results to present two families of trees, one being generated monoid graphs and the other not being generated monoid graphs. A perfect kk-ary tree of height hh, Tk,hT_{k,h} is a rooted tree in which the root has kk neighbors, every non-leaf vertex has also kk successors, i.e., neighbors that are further from the root than itself, and all the leaves are at distance exactly hh from the root. See Figure 5 for an illustration. A direct consequence of Proposition 5.1 is:

Observation 5.5.

Every perfect kk-ary tree Tk,hT_{k,h} is a generated monoid graph.

Refer to caption
Figure 5: The perfect 33-ary tree of height 33, and (with the grey edge added) the tree T3,3+T^{+}_{3,3}.

Denote by Tk,h+T^{+}_{k,h} the tree obtained from Tk,hT_{k,h} by adding an extra leaf at distance hh to the root. The tree T3,3+T^{+}_{3,3} is depicted in Figure 5.

Proposition 5.6.

For k3k\geq 3 and h2h\geq 2, the tree Tk,h+T^{+}_{k,h} is not a generated monoid graph.

Proof.

Suppose that Tk,h+=Cay¯(M,C)T^{+}_{k,h}=\underline{\mathrm{Cay}}(M,C) for some monoid MM and some CC with M=CM=\langle C\rangle. By Lemma 5.3, there is no non-trivial element of MM corresponding to an automorphism of Tk,h+T^{+}_{k,h}; such an automorphism would map the neutral element to a vertex which is at another depth (distance from the root). So Proposition 5.2 can be fully used. Let \ell be the depth of the neutral element. By Lemma 5.4, 1h11\leq\ell\leq h-1, because h2h\geq 2. We know that there is a leaf xx with d(e,x)=hd(e,x)=h-\ell (and be(x)=0b_{e}(x)=0), but none of the vertices yy in the branch of the root with d(e,y)h+1d(e,y)\leq h-\ell+1 is a leaf, and in fact they all satisfy be(y)k12b_{e}(y)\geq k-1\geq 2, contradicting the second part of Proposition 5.2. \square

In the last construction, due to the high symmetry, considering different possible placements of the neutral element was easy. If we add two leaves instead of just one to the same vertex, then by Lemma 5.4 in the resulting graph only one placement of the neutral element has to be analyzed.

Remark 5.7.

The above results are enough to determine, up to order 77, if a tree is a generated monoid graph or not. It turns out that all of them are generated monoid graphs, except one (see the left of Figure 6). On the other hand on 88 vertices we can find a first tree for which our results do not determine whether it is a generated monoid tree or not (see the right of Figure 6).

Refer to caption
Figure 6: Left: the smallest non-generated monoid tree. Right: a tree of order 88 which is not handled by above results. Still, they yield that the only candidate to neutral element is the squared vertex.

6 Concluding remarks

We have studied the classes of monoid and semigroup graphs and digraphs. One of our first results is a Sabidussi-type characterization of monoid digraphs (Lemma 2.4). Since from every semigroup SS one can obtain a monoid S+S^{+} by adjoining a neutral element, this can in a certain way be used to obtain a characterization of semigroup digraphs: a digraph is a semigroup digraph if and only if a source can be added, such that this new vertex plays the role of ee in Lemma 2.4. However, we wonder if there is an intrinsic characterization:

Question 6.1.

Is there a Sabidussi-type characterization of semigroup digraphs?

The central piece of this paper is about monoid graphs. Our knowledge about this class remains superficial and its structure seems hard to capture. Note in particular, that the class of monoid graphs is not closed under disjoint union by Proposition 4.7, which would have been a generalization of Proposition 4.2. The positive examples we know are forests and threshold graphs (Corollaries 4.3 and 4.4). Forests coincide with the class of graphs of treewidth 11. On the other hand the non-monoid graphs from Proposition 4.7 have treewidth 33. How about graphs of treewidth 22? E.g., outerplanar graphs or 22-trees: graphs that can be obtained from K3K_{3} by successively adding a new vertex with exactly two neighbors who form an edge. Figure 1 depicts an outerplanar 22-tree that is a monoid graph.

Question 6.2.

Are 22-trees monoid graphs? Are outerplanar graphs monoid graphs? Are there series-parallel non-monoid graphs?

Another direction are semigroup graphs; here the fundamental question is open. We formulate it in the more provocative way, even if we believe the answer to be negative as in the case of posets, see [14].

Question 6.3.

Is every graph a semigroup graph?

Finally, we have studied trees that are generated monoid graphs. We have found necessary and sufficient conditions for a tree to form part of this class (Theorem 5.1 and Proposition 5.2). These conditions are in terms of simple parameters related to the distance to the root of a tree TT. It would be nice to find a characterization in these terms. More generally both conditions are computationally easy to verify. We wonder:

Question 6.4.

Can it be tested in polynomial time if a tree is a generated monoid tree?

Acknowledgements

This research was initiated in the BGSMath course Fragments of algebraic graph theory 2021. Thanks to participants, lecturers, and organizers. K.K was partially supported by the Spanish Ministerio de Economía, Industria y Competitividad through grant RYC-2017-22701 and by MICINN through grant PID2019-104844GB-I00.

References

  • [1] N. Alon and F. R. K. Chung, Explicit construction of linear sized tolerant networks., Discrete Math., 72 (1988), pp. 15–19.
  • [2] N. Alon and J. H. Spencer, The probabilistic method. 4th edition., Hoboken, NJ: John Wiley & Sons, 4th edition ed., 2016.
  • [3] L. Babai, Automorphism groups of graphs and edge-contraction, Discrete Math., 8 (1974), pp. 13–20.
  • [4] L. Babai, Embedding graphs in Cayley graphs, in Probl. Comb. Théorie des Graphes (Proc. Conf. Paris-Orsay 1976), J-C. Bermond et al., ed., 1978, pp. 13–15.
  • [5] L. Babai, On the abstract group of automorphisms. In Combinatorics, vol. 52 of London Math. Soc. Lecture Notes, Cambridge University Press, 1981, pp. 1–40.
  • [6] L. Babai and A. Pultr, Endomorphism monoids and topological subgraphs of graphs, J. Comb. Theory, Ser. B, 28 (1980), pp. 278–283.
  • [7] N. Biggs, Algebraic Graph Theory, Cambridge Tracts in Mathematics, Cambridge University Press, 2nd ed., 1993.
  • [8] A. Bishnoi, S. Mattheus, and J. Schillewaert, Minimal multiple blocking sets, Electron. J. Combin., 25 (2018), pp. Paper No. 4.66, 14.
  • [9] A. Cayley, Desiderata and suggestions: No. 2. The Theory of groups: graphical representation, American Journal of Mathematics, 1 (1878), pp. 174–176.
  • [10] A. Frank and A. Gyárfás, How to orient the edges of a graph?, in Combinatorics (Proc. Fifth Hungarian Colloq., Keszthely, 1976), Vol. I, Colloq. Math. Soc. János Bolyai, vol. 18, North-Holland, 1978, pp. 353–364.
  • [11] R. Frucht, Herstellung von Graphen mit vorgegebener abstrakter Gruppe, Compositio Mathematica, 6 (1939), pp. 239–250.
  • [12] H. N. Gabow and H. H. Westermann, Forests, frames, and games: Algorithms for matroid sums and applications, Algorithmica, 7 (1992), pp. 407–421.
  • [13] I. García-Marco and K. Knauer, Beyond symmetry in generalized Petersen graphs. in preparation.
  • [14] I. García-Marco, K. Knauer, and G. Mercui-Voyant, Cayley posets, Mediterr. J. Math., 17 (2020).
  • [15] C. Godsil and G. Royle, Algebraic graph theory., vol. 207, New York, NY: Springer, 2001.
  • [16] M. Grohe, P. Schweitzer, and D. Wiebking, Automorphism groups of graphs of bounded Hadwiger number, 2020.
  • [17] J. Gross and T. Tucker, Topological Graph Theory, Wiley Series in Discrete Mathematics and Optimization, Wiley-Interscience, 1987.
  • [18] W. H. Haemers, Interlacing eigenvalues and graphs, Linear Algebra Appl., 226/228 (1995), pp. 593–616.
  • [19] Y. Hao, X. Gao, and Y. Luo, On Cayley graphs of symmetric inverse semigroups., Ars Comb., 100 (2011), pp. 307–319.
  • [20] Y. Hao, X. Gao, and Y. Luo, On the Cayley graphs of Brandt semigroups., Commun. Algebra, 39 (2011), pp. 2874–2883.
  • [21] Y. Hao and Y. Luo, On the Cayley graphs of left (right) groups., Southeast Asian Bull. Math., 34 (2010), pp. 685–691.
  • [22] Z. Hedrlin and J. Lambek, How comprehensive is the category of semigroups?, J. Algebra, 11 (1969), pp. 195–212.
  • [23] Z. Hedrlin and A. Pultr, Relations (graphs) with given finitely generated semigroups, Monatsh. Math., 68 (1964), pp. 213–217.
  • [24] Z. Hedrlin and A. Pultr, Symmetric relations (undirected graphs) with given semigroups, Monatsh. Math., 69 (1965), pp. 318–322.
  • [25] G. Jian and S. Fan, On undirected Cayley graphs of some completely simple semigroups., Southeast Asian Bull. Math., 33 (2009), pp. 741–749.
  • [26] A. Kelarev, J. Ryan, and J. Yearwood, Cayley graphs as classifiers for data mining: The influence of asymmetries, Discrete Mathematics, 309 (2009), pp. 5360–5369.
  • [27] A. V. Kelarev, On undirected Cayley graphs, Australas. J. Combin., 25 (2002), pp. 73–78.
  • [28] A. V. Kelarev, On Cayley graphs of inverse semigroups, Semigroup Forum, 72 (2006), pp. 411–418.
  • [29] A. V. Kelarev and C. E. Praeger, On transitive Cayley graphs of groups and semigroups, European J. Combin., 24 (2003), pp. 59–72.
  • [30] B. Khosravi, On Cayley graphs of left groups., Houston J. Math., 35 (2009), pp. 745–755.
  • [31] B. Khosravi, Some properties of Cayley graphs of cancellative semigroups., Proc. Rom. Acad., Ser. A, Math. Phys. Tech. Sci. Inf. Sci., 17 (2016), pp. 3–10.
  • [32] B. Khosravi, On the Cayley graphs of completely simple semigroups., Bull. Malays. Math. Sci. Soc. (2), 41 (2018), pp. 741–749.
  • [33] B. Khosravi and B. Khosravi, A characterization of Cayley graphs of Brandt semigroups., Bull. Malays. Math. Sci. Soc. (2), 35 (2012), pp. 399–410.
  • [34] B. Khosravi and B. Khosravi, On Cayley graphs of semilattices of semigroups., Semigroup Forum, 86 (2013), pp. 114–132.
  • [35] B. Khosravi and M. Mahmoudi, On Cayley graphs of rectangular groups., Discrete Math., 310 (2010), pp. 804–811.
  • [36] K. Knauer and U. Knauer, Toroidal embeddings of right groups, Thai J. Math., 8 (2010), pp. 483–490.
  • [37] K. Knauer and U. Knauer, On planar right groups., Semigroup Forum, 92 (2016), pp. 142–157.
  • [38] U. Knauer and K. Knauer, Algebraic graph theory. Morphisms, monoids and matrices. 2nd revised and extended edition., Berlin: De Gruyter, 2nd revised and extended edition ed., 2019.
  • [39] M. Lovrečič Saražin, A note on the generalized Petersen graphs that are also Cayley graphs, J. Comb. Theory, Ser. B, 69 (1997), pp. 226–229.
  • [40] A. Lubotzky, R. Phillips, and P. Sarnak, Ramanujan graphs, Combinatorica, 8 (1988), pp. 261–277.
  • [41] Y. Luo, Y. Hao, and G. T. Clarke, On the Cayley graphs of completely simple semigroups., Semigroup Forum, 82 (2011), pp. 288–295.
  • [42] N. V. R. Mahadev and U. N. Peled, Threshold Graphs and Related Topics, Annals of Discrete Mathematics 56, North Holland, 1 ed., 1995.
  • [43] J. Meksawang and S. Panma, Cayley digraphs of Brandt semigroups relative to Green’s equivalence classes., Southeast Asian Bull. Math., 39 (2015), pp. 815–827.
  • [44] J. Meksawang, S. Panma, and U. Knauer, Characterization of finite simple semigroup digraphs., Algebra Discrete Math., 12 (2011), pp. 53–68.
  • [45] C. S. J. A. Nash-Williams, Decomposition of finite graphs into forests, Journal of the London Mathematical Society, s1-39 (1964), pp. 12–12.
  • [46] J. Nešetřil and P. Ossona de Mendez, Sparsity. Graphs, structures, and algorithms, vol. 28, Berlin: Springer, 2012.
  • [47] S. Panma, U. Knauer, and S. Arworn, On transitive Cayley graphs of strong semilattices of right (left) groups., Discrete Math., 309 (2009), pp. 5393–5403.
  • [48] J.-C. Picard and M. Queyranne, A network flow solution of some non-linear 0-1 programming problems and applications to graph theory, Networks, 12 (2006), pp. 141–159.
  • [49] J.-É. Pin, Mathematical foundations of automata theory, 2012.
  • [50] G. Sabidussi, On a class of fixed-point-free graphs, Proceedings of the American Mathematical Society, 9 (1958), pp. 800–804.
  • [51] D. V. Solomatin, Direct products of cyclic semigroups admitting a planar Cayley graph., Sibirskie Ehlektronnye Matematicheskie Izvestiya [Siberian Electronic Mathematical Reports], 3 (2006), pp. 238–252.
  • [52] D. V. Solomatin, Semigroups with outerplanar Cayley graphs, Sibirskie Èlektronnye Matematicheskie Izvestiya [Siberian Electronic Mathematical Reports], 8 (2011), pp. 191–212.
  • [53] T. Suksumran and S. Panma, On connected Cayley graphs of semigroups., Thai J. Math., 13 (2015), pp. 641–652.
  • [54] S. Wang and Y. Li, On Cayley graphs of completely 0-simple semigroups., Cent. Eur. J. Math., 11 (2013), pp. 924–930.
  • [55] W. Wang and H. Hou, Cayley graphs of strong semilattices of left groups., J. Wuhan Univ., Nat. Sci. Ed., 55 (2009), pp. 633–636.
  • [56] A. White, On the genus of a group, Transactions of the American Mathematical Society, 1973 (1972), pp. 203–214.
  • [57] A. T. White, Graphs, Groups and Surfaces, North-Holland, completely rev. and enlarged ed., 1984.
  • [58] B. Zelinka, Graphs of semigroups, Časopis pro pěstování matematiky, 106 (1981), pp. 407–408.
  • [59] X. Zhang, Clifford semigroups with genus zero, in Semigroups, Acts and Categories with Applications to Graphs, Proceedings, Tartu 2007, vol. 3 of Mathematics Studies, Estonian Mathematical Society, Tartu, 2008, pp. 151–160.