1 Introduction and preliminaries

The investigation and characterization of digraphs that are Cayley graphs of certain algebraic structures have a long history, documented for example in Maschke’s Theorem from 1896 about groups of genus zero, that is, groups G which possess systems A of generating elements, such that the Cayley graphs Cay(G,A) are planar. A modern presentation of this can be found in [16]. Stimulated by the rich results of research on the Cayley graphs of groups, the Cayley graphs of semigroups have also been considered by many authors, see, for example, [1, 58, 13, 14, 18]. Some of the earliest references on this subject are [2, 19, pp. 93–101]. The Cayley graphs of semigroups are closely related to the finite state automata and have many valuable applications (see the survey [11] and Sect. 2.4 of the book [4]). In addition, similar questions concerning divisibility graphs and power graphs of semigroups were considered in [9, 10], respectively.

If S is a semigroup, and T is a nonempty subset of S, the so-called connection set, then the Cayley graph Cay(S,T) of S relative to T is usually defined as the graph with vertex set S and edge set E(Cay(S,T)) consisting of those ordered pairs (x,y), where xt=y for some tT, see [1, 5, 8, 1214]. Note that in all of these articles the Cayley graph is defined by right translations which is quite usual. But at the same time, in some other articles, the Cayley graph is defined by left translations, see, for example, [6, 7, 11, 18].

However, the reflection of algebraic properties of S may depend strongly on the choice of left actions or right actions. Evidently, for a given semigroup S and a fixed subset T of S, the Cayley graphs Cay(S,T) defined by left actions and by right actions may not be the same. In order to unify these two cases, and to get more applications, we shall generalize the definition of Cayley graphs of semigroups in the present paper.

As we can see in [1, 58, 1114, 18], when studying Cayley graphs of semigroups, at least the following three basic aspects are of significance: (1) to describe the Cayley graph of a given semigroup; (2) given a Cayley graph of a semigroup, to determine the structure or study the properties of this semigroup; and (3) to find or construct a semigroup whose Cayley graph is a given graph. Our work on generalized Cayley graphs of semigroups will follow a similar line.

In Sect. 1 of the present paper, we introduce the notion of generalized Cayley graphs of semigroups. Some fundamental properties of generalized Cayley graphs are discussed in Sect. 2. At last, in Sect. 3, we focus on a special case, the universal Cayley graphs of semigroups so that some general results are given. Here we give some fundamental concepts and simple facts for preliminaries. As Example 1.4 below shows, the generalized Cayley graphs of a given semigroup may include more universal classes of graphs than the usual Cayley graphs.

Recall that if S is an ideal of a semigroup T, then we call T an ideal extension of S. For any semigroup T, let

$$T^1 = \left\{ \begin{array}{l@{\quad}l}T & \mbox{if}\ T\ \mbox{has the identity}\\T\cup \{1\}& \mbox{otherwise}.\end{array} \right.$$

Now we present the main definition of this paper, which generalizes the Cayley graph of a semigroup S relative to a subset of S to that relative to a relation on T 1 with T an ideal extension of S.

Definition 1.1

Let T be an ideal extension of a semigroup S and ρT 1×T 1. The Cayley graph Cay(S,ρ) of S relative to ρ is defined as the graph with vertex set S and edge set E(Cay(S,ρ)) consisting of those ordered pairs (a,b), where xay=b for some (x,y)∈ρ.

We also call the Cayley graphs defined in this way the generalized Cayley graphs, in order to distinguish them from the usual ones. Some examples are displayed below.

Example 1.2

If T is a nonempty subset of a semigroup S, then generalized Cayley graph Cay(S,T×{1}) is actually the usual Cayley graph Cay(S,T) of S relative to T defined by the left actions, while generalized Cayley graph Cay(S,{1}×T) is precisely the usual Cayley graph Cay(S,T) of S relative to T defined by the right actions.

Example 1.3

Assume that S is a semigroup. Let ω l =S 1×{1} (the left universal relation on S 1), ω r ={1}×S 1 (the right universal relation on S 1) and ω=S 1×S 1 (the universal relation on S 1). Then the generalized Cayley graphs Cay(S,ω l ), Cay(S,ω r ) and Cay(S,ω) are called left universal, right universal and universal Cayley graphs of S, respectively.

Example 1.4

Let N be the semigroup of all natural numbers with the usual multiplication, and S the ideal of N consisting of all nonnegative even integers. Then on the one hand, for any subset T of S, the usual Cayley graph Cay(S,T) can not contain any edge (a,b) such that b is not divided by 4. But on the other hand, there are indeed such edges in the generalized Cayley graph Cay(S,N×N).

In what follows, we give some simple facts for preliminaries. Assume that S is a semigroup and aS. Then L(a)=S 1 a, R(a)=aS 1 and J(a)=S 1 aS 1 is the principal left ideal, principal right ideal and principal ideal generated by a, respectively. The next lemma is a direct consequence of Definition 1.1.

Lemma 1.5

Let T be an ideal extension of a regular semigroup S and ρT 1×T 1. Then the following statements hold:

  1. (1)

    if a,bS such that (a,b)∈E(Cay(S,ρ)), then J(a)⊇J(b);

  2. (2)

    there is a relation ρ 0S×S such that Cay(S,ρ) is a subgraph of Cay(S,ρ 0).

Throughout the paper, graphs are directed graphs without multiple edges, but possibly with loops, or digraphs in terms of [1, 14]. For a graph Γ, denote by V(Γ) and E(Γ) its vertex set and edge set, respectively. For any aV(Γ), Let

$$ \overrightarrow{a}=\{b\in V(\varGamma )|(a, b)\in E(\varGamma )\}, \qquad \overleftarrow{a}=\{b\in V(\varGamma )|(b, a)\in E(\varGamma )\}.$$
(1.1)

Let T be an ideal extension of a semigroup S and ρT 1×T 1. If a,bS such that (a,b)∈E(Cay(S,ρ)), then by Definition 1.1, there exists (x,y)∈ρ such that xay=b. Let

$$ \rho(a)=\{xay|(x, y)\in \rho\},$$
(1.2)

and set

$$ \rho(a)^1=\rho(a)\cup \{a\}.$$
(1.3)

We call ρ(a)1 the ρ-class of a. It is evident that (a,b)∈E(Cay(S,ρ)) if and only if bρ(a). So by (1.1), we obtain

$$ \rho (a)=\overrightarrow{a}.$$
(1.4)

The next lemma will be used repeatedly later.

Lemma 1.6

Let T be an ideal extension of a semigroup S, ρT 1×T 1 and a,bS. If ρ 1(a)⊇ρ 1(b) and ab, then (a,b)∈E(Cay(S,ρ)).

Proof

Assume that ρ 1(a)⊇ρ 1(b) and ab. By (1.3), bρ 1(b)⊆ρ 1(a)=ρ(a)∪{a}. Since ab, we have that bρ(a). By (1.2), there exists (x,y)∈ρ such that xay=b. Hence by Definition 1.1, (a,b)∈E(Cay(S,ρ)) as required. □

As usual, a directed graph without loops and multiple edges is called simple. Note that although generalized Cayley graphs defined in the present paper have no multiple edges, they may have loops, and so they don’t have to be simple. Sometimes, we shall equate two graphs each of which is isomorphic to the other one. That is, if Γ 1Γ 2, then we may identify Γ 1 with Γ 2.

We use standard terminology of graph theory following, for example, [17]. For semigroup theory, we refer the reader to [3, 15, 20].

2 Fundamental properties of generalized Cayley graphs of semigroups

The aim of this section is to discuss in general the fundamental properties of generalized Cayley graphs of semigroups by introducing some new concepts such as ‘I-compatible’ and ‘stable’. A necessary and sufficient condition is given in Theorem 2.2 for a relation on a semigroup to be I-compatible, and the semigroup whose generalized Cayley graph is a chain of complete graphs with loops is characterized in Theorem 2.10.

Let us begin by introducing the first new notion of this section.

Definition 2.1

Let T be a semigroup and ρT×T. If that (a,b),(c,d)∈ρ for any a,b,c,dT always implies (ca,bd)∈ρ, then we call ρ inversely compatible, or briefly, I-compatible.

For example, if T is a semigroup with A and B subsemigroups of T, then it is clear that ρ=A×B is I-compatible. Particularly, if S is a semigroup, then S×{1}, S 1×{1}, {1}×S, {1}×S 1, S×S and S 1×S 1 are all I-compatible.

Suppose that S and T are semigroups. A mapping ϕ:ST is called an anti-homomorphism if ϕ(ab)=ϕ(b)ϕ(a) for all a,bS. Furthermore, if ϕ is a bijection, then we call ϕ an anti-isomorphism, or say that S is anti-isomorphic to T.

For characterizing the I-compatibility, suppose that T is a semigroup and ρT×T, and let

$$l_{\rho}=\{a|(a,b)\in\rho\},\qquad r_{\rho}=\{b|(a,b)\in\rho\},$$
(2.1)

If a,bl ρ , then exist c,dT such that (a,c),(b,d)∈ρ. If ρ is I-compatible, then (ba,cd)∈ρ which means that bal ρ and cdr ρ . Thus both l ρ and r ρ are subsemigroups of T.

Now ρl ρ ×r ρ . It is evident that the mapping ρl ρ , (a,b)↦a is an onto anti-homomorphism. Let \(l_{\rho}^{*}=l_{\rho}\). For any \(a,b\in l_{\rho}^{*}\), define ab=ba. Then \((l_{\rho}^{*}, \ast)\) is a semigroup anti-isomorphic to l ρ . We call \((l_{\rho}^{*}, \ast)\) the anti-semigroup of l ρ . The anti-semigroup of a subsemigroup of a semigroup T is called a anti-subsemigroup of T. Thus \((l_{\rho}^{*}, \ast)\) is a anti-subsemigroup of T. Now, \(\rho\longrightarrow l_{\rho}^{*}\), (a,b)↦a is an onto homomorphism. Of course, the mapping ρr ρ , (a,b)↦b is also an onto homomorphism. So ρ is a subdirect product of \(l_{\rho}^{*}\) and r ρ .

Conversely, assume that ρ is a subdirect product of \((l_{\rho}^{*}, \ast)\), an anti-subsemigroup of T, and l ρ , a subsemigroup of T. Then ρ is a subsemigroup of the direct product \(l_{\rho}^{*}\times r_{\rho}\). Thus for any (a,b),(c,d)∈ρ, we have that (ca,bd)=(ac,bd)=(a,b)(c,d)∈ρ. It follows that ρ is I-compatible.

Summarizing the above arguments, we obtain

Theorem 2.2

Assume that T is a semigroup and ρT×T. Then ρ is I-compatible if and only if there exist an anti-subsemigroup A and a subsemigroup B such that ρ is the subdirect product of A and B.

Let us turn our attention to the second new notion of this section.

Definition 2.3

Let Γ be a graph. If that (a,b),(b,c)∈E(Γ) for a,b,cV(Γ) always implies that (a,c)∈E(Γ), then Γ is said to be edge-transitive.

The relation of the above two notions is indicated in

Proposition 2.4

Assume that T is an ideal extension of a semigroup S. If ρT 1×T 1 is I-compatible, then Cay(S,ρ) is edge-transitive.

Proof

Assume that T is an ideal extension of a semigroup S and ρT 1×T 1 is I-compatible. If (a,b),(b,c)∈E(Cay(S,ρ)), then there exist (x,y),(x 0,y 0)∈ρ such that b=xay and c=x 0 by 0. Thus

$$c=x_0by_0=x_0(xay)y_0=(x_0x)a(yy_0),$$

where (x 0 x,yy 0)∈ρ by the assumption that ρ is I-compatible. It follows that (a,c)∈E(Cay(S,ρ)). Therefore, Cay(S,ρ) is edge-transitive and the proof is complete. □

Corollary 2.5

Assume that T is an ideal extension of a semigroup S and ρT 1×T 1 is I-compatible. If (a,b)∈E(Cay(S,ρ)), then ρ(a)⊇ρ(b) and ρ 1(a)⊇ρ 1(b).

Proof

Suppose that (a,b)∈E(Cay(S,ρ)) and that ρ is I-compatible. Together with (1.2) and Definition 1.1, we obtain that bρ(a). According to Proposition 2.4, Cay(S,ρ) is edge-transitive. Now, for any cρ(b), (b,c)∈E(Cay(S,ρ)). It follows that (a,c)∈E(Cay(S,ρ)), that is, cρ(a). Therefore, ρ(a)⊇ρ(b). By (1.3), ρ 1(b)=ρ(b)∪{b}. It follows that ρ(a)⊇ρ 1(b). Hence ρ 1(a)⊇ρ(a)⊇ρ 1(b) and the proof is complete. □

To obtain the converse implication of Corollary 2.5, we introduce the third new concept of this section as follows.

Definition 2.6

Assume that T is an ideal extension of a semigroup S and ρT 1×T 1. An element a of S is called stable under ρ if aρ(a). If all elements of S are stable under ρ, then S is called stable under ρ.

It is clear that a is stable under ρ if and only if ρ(a)1=ρ(a). For example, it is easy to show that if S is a regular semigroup or a monoid, then S is stable under ω l , ω r and ω, respectively. Combining Lemma 1.6 and Corollary 2.5, one may easily deduce

Proposition 2.7

Assume that T is an ideal extension of a semigroup S. If ρT 1×T 1 is I-compatible and S is stable under ρ, then for any a,bS the following conditions are equivalent:

  1. (1)

    (a,b)∈E(Cay(S,ρ));

  2. (2)

    ρ(a)⊇ρ(b);

  3. (3)

    ρ 1(a)⊇ρ 1(b).

In what follows, we shall characterize semigroups whose generalized Cayley graphs are some special graphs.

Let (Y,≥) be a partially ordered set. For any aY, set

$$ a\downarrow=\{b\in Y|a\geq b\},\qquad a\uparrow=\{b\in Y|b\geq a\}.$$
(2.2)

Recall that for any a,bY, a,b are called comparable if ab or ba. We call (Y,≥) a totally ordered set if a,b are comparable for all a,bY. Also we call the total order ≥ a linear order. The relation graph of a linear order is called a linear graph. It is easily seen that a graph Γ is a linear graph if and only if the following conditions are satisfied: (1) Γ does not contain multiple edges; (2) for any aV(Γ), (a,a)∈E(Γ); (3) for any a,bV(Γ), (a,b)∈E(Γ) or (b,a)∈E(Γ); (4) if abV(Γ) and (a,b)∈E(Γ), then \((b,a)\ \bar{\in}\ E(\varGamma )\); (5) Γ is edge-transitive. The next lemma characterizes semigroups whose Cayley graphs are linear graphs.

Lemma 2.8

Assume that T is an ideal extension of a semigroup S and that ρT 1×T 1 is I-compatible such that S is stable under ρ. Then the following conditions are equivalent:

  1. (1)

    Cay(S,ρ) is a linear graph;

  2. (2)

    There exists a linear orderon S such that for any aS, ρ(a)=a↓;

  3. (3)

    There exists a linear orderon S such that for any aS, ρ 1(a)=a↓.

Proof

Assume that T is an ideal extension of a semigroup S and ρT 1×T 1 is I-compatible such that S is stable under ρ. Then by Proposition 2.4 and (1.4) Cay(S,ρ) is edge-transitive and for any aS, \(\rho ^{1}(a)=\rho(a)=\overrightarrow{a}\).

(1) ⇒ (2). Suppose that Cay(S,ρ) is a linear graph. Then define an order ≥ on S by that ab if and only ρ(a)⊇ρ(b). It is clear that ≥ is reflexive and transitive. Assume that ab and ba. Then ρ(a)⊇ρ(b) and ρ(b)⊇ρ(a). Thus ρ(a)=ρ(b). By Proposition 2.7, (b,a)∈E(Cay(S,ρ)) and (a,b)∈E(Cay(S,ρ)). Since Cay(S,ρ) is a linear graph, we have a=b. Therefore ≥ is anti-symmetric and furthermore it is a partial order. For any a,bS, (a,b)∈E(Cay(S,ρ)) or (b,a)∈E(Cay(S,ρ)). It follows from Proposition 2.7 that ρ(a)⊇ρ(b) or ρ(b)⊇ρ(a). That is, ab or ba, which means that ≥ is a total order. By (2.2) and Proposition 2.7, for any aS, \(a\downarrow=\{b\in S|a\geq b\}=\{b\in S|\rho (a)\supseteq\rho (b)\}=\{b\in S|(a,b)\in E(\mathrm{Cay}(S,\rho))\}=\overrightarrow{a}=\rho(a)\), as required.

(2) ⇒ (1). Assume that ≥ is a linear order on S such that for any aS, ρ(a)=a↓. Then for any aS, aa. It follows that aa↓=ρ(a) which means that (a,a)∈E(Cay(S,ρ)). For any a,bS, ab or ba. It follows that ba↓=ρ(a) or ab↓=ρ(b). That is, (a,b)∈E(Cay(S,ρ)) or (b,a)∈E(Cay(S,ρ)). Thus by Proposition 2.7, ρ(a)=ρ(b) which deduce that a↓=b↓. Hence ab and ba. Since ≥ is anti-symmetric, we have a=b. Therefore, Cay(S,ρ) is a linear graph.

(2) ⇔ (3). This is trivial because ρ 1(a)=ρ(a) for every aS by the assumption that S is stable under ρ, and so the lemma is proved. □

A complete graph with loops are defined as a graph Γ with the edge set E(Γ)={(x,y)|x,yV(Γ)}. Assume that T is an ideal extension of a semigroup S and ρT 1×T 1. If ρ 1(a)=ρ 1(b) for all a,bS, then S is called ρ-simple. It is easy to check that S is ρ-simple if and only if for every aS, ρ 1(a)=S. For example, if T is an ideal extension of a simple (rep. left simple, right simple) semigroup S, then S is ω- (resp., ω l -, ω r -) simple. The next lemma uses the concept ‘ρ-simple’ to characterize semigroups whose Cayley graphs are complete graphs with loops.

Lemma 2.9

Assume that T is an ideal extension of a semigroup S and ρT 1×T 1 is an I-compatible relation such that S is stable under ρ. Then the following conditions are equivalent:

  1. (1)

    Cay(S,ρ) is a complete graph with loops;

  2. (2)

    S is ρ-simple.

Proof

Assume that T is an ideal extension of a semigroup S and ρT 1×T 1 is I-compatible such that S is stable under ρ. Then the conditions of Proposition 2.7 are satisfied.

(1) ⇒ (2). Suppose that Cay(S,ρ) is a complete graph with loops. Then for any a,bS, (a,b)∈E(Cay(S,ρ)) and (b,a)∈E(Cay(S,ρ)). According to Proposition 2.7, ρ 1(a)=ρ 1(b) which means that S is ρ-simple.

(2) ⇒ (1). Assume that S is ρ-simple. Then for any a,bS, ρ 1(a)=ρ 1(b). Again, by Proposition 2.7, (a,b)∈E(Cay(S,ρ)) and (b,a)∈E(Cay(S,ρ)). It is shown that Cay(S,ρ) is a complete graph with loops. □

Together with Proposition 2.7, Lemmas 2.8 and 2.9, we can easily prove the main theorem of this section:

Theorem 2.10

Assume that T is an ideal extension of a semigroup S and ρT 1×T 1 is an I-compatible relation such that S is stable under ρ. Then the following conditions are equivalent:

  1. (1)

    Cay(S,ρ) is a chain of complete graph with loops;

  2. (2)

    S is a chain of ρ-simple semigroups.

3 Universal Cayley graphs of semigroups

In this section, we concentrate on the special case ρ=ω, that is, we shall investigate the universal Cayley graph Cay(S,ω) of a semigroup S, where ω is the universal relation on S 1. The main results of this section are Theorems 3.9 and 3.13. The former establishes the connection between a completely regular semigroup and a semilattice of complete graphs with loops, and the latter generalizes Theorem 3.9 and characterizes the nonidempotent-trivial -partial order of complete graphs with loops by means of a new construction similar with Clifford semigroups, so called semigroup-partially-ordered sets of semigroups which we define in Definition 3.11.

Observing that ω is I-compatible and S is stable under ω, and that for any aS, ω 1(a)=ω(a)=J(a), the principal ideal of a, according to Proposition 2.7, we get the following result.

Lemma 3.1

For any semigroup S, the following conditions are equivalent:

  1. (1)

    (a,b)∈E(Cay(S,ω));

  2. (2)

    J(a)⊇J(b).

As a consequence of this lemma or Lemma 2.9, we have

Lemma 3.2

Assume that S is a semigroup. Then Cay(S,ω) is a complete graph with loops if and only if S is simple.

Let us turn our attention to universal Cayley graphs of semilattices.

The relation graph of the inverse relation ≥ of the natural order ≤ on a semilattice is called a semilattice graph. It is easily seen that a graph Γ is a semilattice graph if and only if the following conditions are satisfied: (1) for any aV(Γ), (a,a)∈E(Γ); (2) for any a,bV(Γ), \(\overrightarrow{a}\cap\overrightarrow{b}\neq\emptyset\) and it has an initial element, that is to say, there exists \(c\in \overrightarrow{a}\cap\overrightarrow{b}\) such that (c,d)∈E(Γ) for all \(d\in \overrightarrow{a}\cap\overrightarrow{b}\).

Lemma 3.3

If S is a semilattice, then Cay(S,ω) is a semilattice graph. Actually, Cay(S,ω) is exactly the relation graph of inverse relationof the natural orderon S.

Proof

If for a,bS, ab, then ba which implies that ab=b. Thus bJ(a) and J(b)⊆J(a). By Lemma 3.1, (a,b)∈E(Cay(S,ω)).

Conversely, if (a,b)∈E(Cay(S,ω)), then by Lemma 3.1 again, J(b)⊆J(a). Thus we have bJ(a) which implies that b=xay=xya for some x,yS 1. Thus ba, or equivalently, ab.

Now we have proved that ab if and only if (a,b)∈E(Cay(S,ω)). So, Cay(S,ω) is exactly the relation graph of ≥ on S, and Cay(S,ω) is a semilattice graph. □

Remark 3.4

The converse of Lemma 3.3 is not true. For example, we consider the null semigroup with three elements 0,a,b. Evidently, J(0)={0}, J(a)={0,a} and J(b)={0,b}. So the Cay(S,ω) is a semilattice graph, but S is not a semilattice.

Lemma 3.5

For a semilattice graph Γ, there is a semilattice S such thatCay(S,ω)=Γ.

Proof

Since Γ is a semilattice graph, then there is a semilattice (S,≤) such that Γ is the relation graph of the inverse relation ≥ of the relation ≤. For any a,bS, define multiplication by ab=ab, the greatest lower bound of a and b in (S,≤). Then S becomes a semigroup which is a semilattice. By Lemma 3.3, Cay(S,ω) is exactly the relation graph of ≥ on S, where ≥ is the inverse relation of the natural order ≤ on S. That is, Cay(S,ω)=Γ, as required. □

To describe further the properties of universal Cayley graphs of semigroups, we require some new concepts which we introduce as follows.

A partially ordered set (Y,≥) is called a strong partially ordered set if for any α,βX, there is γY such that αγ and βγ. If (Y,≥) is a partially ordered set, and Y is a semigroup such that for any α,βY, α,βαβ, then we call the semigroup Y a partially ordered semigroup, and we call the partially ordered set Y a semigroup-partially-ordered set. If B is a partially ordered semigroup which is also a band (i.e., an idempotent semigroup), then we call B a band-partially-ordered set or a partially ordered band. If (Y,≥) is a partially ordered set, Y is a semigroup, T is an ideal extension of Y and ρT 1×T 1 such that for any αY, ρ 1(α)=α↓ where by α↓ we denote the set {β|αβ}, then we call the semigroup Y a ρ-semigroup, and we call the partially ordered set Y a ρ-partially-ordered set. In particular, if ρ=S 1×S 1 (rep. S 1×{1}, {1}×S 1), then a ρ-partially-ordered set or ρ-semigroup is called a -partially-ordered set or -semigroup (rep. -partially-ordered set or -semigroup, -partially-ordered set or -semigroup).

It is clear that a semilattice with its natural order is a -partially-ordered set, a -partially-ordered set is a semigroup-partially-ordered set, and a semigroup-partially-ordered set is a strong partially ordered set. The relation graph of a strong partially ordered set (rep. semigroup-partially-ordered set, -partially-ordered set, etc.) is called a strong partially ordered graph (rep. semigroup-partially-ordered graph, -partially-ordered graph, etc.). From definitions, it is easy to deduce

Lemma 3.6

If Y is a -semigroup, then Cay(Y,ω) is precisely the relation graph of the partially ordered set Y. Conversely, if Y is a partially-ordered graph as the universal Cayley graph of a semigroup, then Y is a -partially-ordered graph.

Now we are ready to analyze the characteristics of the universal Cayley graph of a semigroup.

According to Lemma 3.1, we know that (a,b)∈E(Cay(S,ω)) if and only if J(a)⊇J(b), and that (a,b),(b,a)∈E(Cay(S,ω)) if and only if J(a)=J(b). Denoted by J a the -class of a element a is S, as in [3]. Thus, (a,b)∈E(Cay(S,ω)) if and only if J b J a (see (2.1.3) of [3]), and (a,b),(b,a)∈E(Cay(S,ω)) if and only if . Since is an equivalence relation on S, S is decomposed into a union of all distinguished -classes each of which is in correspondence with a complete subgraph with loops of Cay(S,ω). Suppose that J b <J a and that Γ α ,Γ β are the complete subgraphs with loops of Cay(S,ω), which correspond with J a and J b respectively. Then for any a′∈Γ α and b′∈Γ β , . From this and Lemma 3.1, it follows that (a′,b′)∈E(Cay(S,ω)) but \((b',a')\bar{\in} E(\mathrm{Cay}(S,\omega))\).

Define a relation on V(Cay(S,ω)) by setting ab if and only if

$$(a, b)\in E(\mathrm{Cay}(S,\omega))\quad \mbox{and}\quad (b,a)\in E(\mathrm{Cay}(S,\omega)).$$

It is clear that ab if and only if . Thus ∼ is an equivalence on V(Cay(S,ω)), and for every aS, the ∼-class [a] of a is exactly J a , the -class of a. Now we define the quotient graph Cay(S,ω)/∼ of Cay(S,ω) as the graph with vertex set V(Cay(S,ω))/∼ and the edge set

$$\{([a],[b])|(a,b)\in E(\mathrm{Cay}(S,\omega))\}.$$

Immediately, we have that , and that E(Cay(S,ω)/∼)={(J a ,J b )|J b J a )}. It is trivial that for any a,bS, J(ab)⊆J(a) and J(ab)⊆J(b). Thus ([a],[ab])∈E(Cay(S,ω)/∼) and ([b],[ab])∈E(Cay(S,ω)/∼). It follows that Cay(S,ω)/∼ is a strong partially ordered graph.

If the quotient graph Γ/∼ of a graph Γ respective to ∼ is a strong partially ordered graph (resp. semigroup-partially-ordered graph, -partially-ordered graph, semilattice graph, etc.), then we call Γ a strong partial order (resp. semigroup-partial order, -partial-order, semilattice, etc.) of complete graphs with loops. A strong partial order Γ of complete graphs Γ α with loops is called nonidempotent-trivial if Γ α is trivial (i.e., |V(Γ α )|=1) for every nonidempotent α of Y. In terms of these terminologies and in light of the above arguments, we obtain the following lemma, which characterizes the fundamental property of the universal Cayley graph of a semigroup.

Lemma 3.7

For any semigroup S, Cay(S,ω) is a strong partial order of complete graphs with loops.

For further discussion, we need the following lemma.

Lemma 3.8

For any completely regular semigroup S, Cay(S,ω) is a semilattice of complete graphs with loops.

Proof

Assume that S is a completely regular semigroup S. By Lemma 3.7, Cay(S,ω) is a strong partial order of complete graphs with loops. In light of [3], S is a semilattice Y of completely simple semigroups S α , that is S=⋃ αY S α . Each S α is actually a -class of S. The quotient graph Cay(S,ω)/∼ is exactly the relation graph of the inverse relation ≥ of the semilattice (Y,≤), thus it is a semilattice graph. Therefore, Cay(S,ω) is a semilattice of complete graphs with loops. □

Based on Lemma 3.8, the next theorem characterizes completely the universal Cayley graphs of completely regular semigroups:

Theorem 3.9

The following statements are valid:

  1. (1)

    if S is a completely regular semigroup, then Cay(S,ω) is a semilattice of complete graphs with loops;

  2. (2)

    if Γ is a semilattice of complete graphs with loops, then there is a (commutative) completely regular semigroup such that Cay(S,ω)=Γ.

Proof

First, (1) follows from Lemma 3.8 directly.

Second, to prove (2), suppose that Γ is a semilattice Y of complete graphs Γ α with loops. It is obvious that for any αY, there is an (Abelian) group G α such that |G α |=|V(Γ α )|. The identity element of G α is denoted by 1 α . Obviously, G α is a simple semigroup. So from Lemma 3.2, it follows that Cay(G α ,ω)=Γ α .

For any αY, define a mapping as follows: ϕ α,α :G α G α ,aa; and for any α>βY, define a mapping as follows: ϕ α,β :G α G β ,a↦1 β . It is verified that for αβγ, ϕ α,β ϕ β,γ =ϕ α,γ . Consequently, ({G α } αY ,{ϕ α,β } αβ ) is a strong semilattice of all groups G α ’s.

Let \(S=\dot{\bigcup}_{\alpha \in Y} G_{\alpha}\), the disjoint union of G α . Define a composition on S as follows: for aG α ,bG β , ab=(a)ϕ α,αβ (b)ϕ β,αβ . Then S is a (commutative) Clifford semigroup which is of course completely regular. For any aG α , J(a)=G α . For any aG α ,bG β , J(a) and J(b) are comparable if and only if α and β are comparable. Furthermore, J(a)⊇J(b) if and only if αβ. Thus by Lemma 3.1, Cay(S,ω)=Γ, as required. □

To generalize the result of Theorem 3.9, we need

Lemma 3.10

Let Γ be the relation graph of a partially ordered set (Y,≥). Then there is a composition on Y such that Y becomes a semigroup and Cay(Y,ω)=Γ if and only if (Y,≥) is a -partially-ordered set.

Proof Necessity.

If Y is a semigroup such that Cay(Y,ω)=Γ, then for any aY, \(J(a)=\omega ^{1}(a)=\omega (a)=\overrightarrow{a}=a\downarrow\). Thus (Y,≥) is a -partially-ordered set.

Sufficiency. Assume that (Y,≥) is -partially-ordered set. Then Y is a semigroup and for any aY, \(J(a)=a\downarrow =\overrightarrow{a}\). Thus for any a,bY, (a,b)∈E(Cay(Y,ω)) if and only if J(a)⊇J(b), if and only if a↓⊇b↓, if and only if ab. Then Cay(Y,ω) is just the relation graph of ≥, that is, Cay(Y,ω)=Γ. □

Also, we need the following construction, which is similar to that of Clifford semigroups.

Definition 3.11

Let (Y,≥) be a semigroup-partially-ordered set, {S α } αY a class of semigroups. If for any αβY, there is a homomorphism ϕ α,β :S α S β satisfying the following conditions:

  1. (1)

    \(\phi _{\alpha,\alpha}=i_{S_{\alpha}}\), the identity mapping;

  2. (2)

    ϕ α,β ϕ β,γ =ϕ α,γ (with mapping notations on the right of the variables).

then ({G α } αY ,{ϕ α,β } αβ ) is called a semigroup-partially-ordered set Y of semigroups S α ’s.

It is routine to verify the next lemma.

Lemma 3.12

With the same notations as in Definition 3.11, let \(S=\dot{\bigcup}_{\alpha \in Y} S_{\alpha}\), the disjoint union of G α , and define a composition on S as follows: for aS α ,bS β , ab=(a)ϕ α,αβ (b)ϕ β,αβ . Then (S,∗) is a semigroup.

Evidently, the composition ∗ in Definition 3.12 restricted to S α coincides with the original one of S α if and only if α is an idempotent of the semigroup Y. If Y and S α ’s are all commutative semigroups, then so is S.

Now we are ready to prove the following theorem, which generalizes Theorem 3.9, describes the basic properties of the universal Cayley graph of an arbitrary semigroup and characterizes the -partial order of complete graphs with loops.

Theorem 3.13

The following statements hold:

  1. (1)

    For any semigroup S, Cay(S,ω) is a strong partial order of complete graphs with loops, and for each aS and all bJ(a), (a,b)∈E(Cay(S,ω));

  2. (2)

    Let Γ be a nonidempotent-trivial -partial order Y of complete graphs with loops. Then there is a semigroup S such that Cay(S,ω)=Γ. Moreover, if Y is a commutative semigroup, then S can be chosen to be commutative.

Proof

From Lemmas 3.7 and 3.1, (1) follows. To prove (2), let Γ be a -partial order (Y,≥) of complete graphs Γ α ’s with loops. For any αY, it cleat that there is a simple monoid (in fact, an Abelian group) S α such that |S α |=|V(Γ α )|, and by Lemma 3.2, Cay(S α ,ω)=Γ α . In light of Lemma 3.12, we obtain a semigroup \((S=\dot{\bigcup}_{\alpha \in Y} S_{\alpha},\ast)\). If Y is a commutative semigroup, then S can be chosen to be commutative since all S α ’s can be chosen to be commutative.

Denote by X the set of all idempotents of Y. For any aS, there exists a unique αY such that aS α . If αX, then the composition ∗ restricted in S α coincides with the original one of S α . Since S α is a simple monoid, S α aS α =S α by Corollary 3.1.2 of [3]. If αYX, then Γ α is trivial by the assumption. Thus S α is trivial, in effect, S α ={1 α }={a}. So whether α is in X or not, we have S α J(a). Since Y is a -partially-ordered set, J(α)=α↓={β|αβ}. It follows that \(J(a)=\bigcup _{\alpha \geq\beta}S_{\beta}=\{b|(a,b)\in E(\varGamma )\}=\overrightarrow{a}\). Thus we have Cay(S,ω)=Γ as required. □