1 Introduction and preliminaries

Let S be an inverse semigroup with semilattice of idempotents E. Define an inverse subsemigroup N of S to be normal if it is full (i.e., EN), closed (i.e., =N, where ω:2S→2S is a closure operator given by ={sS:∃ aA [asA]} for all AS), and self-conjugate (i.e., s −1 NsN for every sS). It follows from [11] (see p. 181) that there exists an inclusion-preserving bijection between the set of normal subsemigroups of S and the set of group congruences on S. In fact, the relation ρ N ={(a,b)∈S×S:ab −1N} is a group congruence on S and \(\operatorname{ker}\rho_{N} = N\). These results were generalized in [9] and [16]. It is easy to see that (a,b)∈ρ N if and only if ax,bxN for some xS.

The main purpose of the next section is a description of group congruences on a semigroup S in the terms of some special subsemigroups of S. Our description is simpler than that of Dubreil (see 10.2 [1]) and a little more general than the description of Gomes [6] (nota bene our proof is simpler). We apply this description to determine group congruences (in particular, the least group congruence) on some special classes of semigroups; namely: E-inversive (E-)semigroups (in particular, idempotent-surjective (E-)semigroups), eventually regular semigroups.

We divide this paper into seven sections. In Sect. 2 we describe group congruences on an arbitrary semigroup S in the terms of normal subsemigroups of S (see below for the definition). In Sect. 3 we investigate group congruences on E-inversive semigroups. In particular, we show that the least group congruence on an E-inversive semigroup exists (in general, this is false: see Example 1.2). In Sect. 4 and 5 we study group congruences on E-inversive E-semigroups and E-unitary E-inversive semigroups, respectively. Further, in Sect. 6 we use the results of Sect. 2 for an easy description of all group congruences on eventually regular semigroups (in terms of full, closed and self-conjugate subsemigroups) and we give some remarks on group congruences on inverse semigroups. Finally, in Sect. 7, some remarks on the hypercore of a semigroup are given (see [8]).

Let S be a semigroup. Denote by Reg(S) the set of regular elements of S, that is, Reg(S)={aS:aaSa} and by V(a) the set of inverses of aS, i.e., the set {xS:a=axa,x=xax}. Note that if aS is regular, say a=axa for some xS, then xaxV(a). Also, S is called regular if V(a)≠∅ for every aS. Further, S is said to be eventually regular if every element a of S has a regular power. In such a case, by r(a) we shall mean the regular index of a, i.e., the least positive integer n for which a nReg(S).

Let S be a semigroup, aS. The set W(a)={xS:x=xax} is called the set of all weak inverses of a and so the elements of W(a) will be called weak inverse elements of a. A semigroup S is said to be E-inversive if for every aS there exists xS such that axE S , where E S (or briefly E) is the set of idempotents of S (more generally, if AS, then E A denotes the set of idempotents of A). It is easy to see that a semigroup S is E-inversive if and only if W(a) is nonempty for all aS. Hence if S is E-inversive, then for every aS there is xS such that ax,xaE S (see [19, 20]). Clearly, eventually regular semigroups are E-inversive. We remark that the class of eventually regular semigroups is very wide and contains the class of regular, group-bound (in particular, periodic, finite) semigroups. In [7] Hall observed that the set Reg(S) of a semigroup S with E S ≠∅ forms a regular subsemigroup of S if and only if the product of any two idempotents of S is regular. In that case, S is said to be an R-semigroup. Also, we say that S is an E-semigroup if E S is a subsemigroup of S. Evidently, every E-semigroup is an R-semigroup. Finally, [eventually] regular E-semigroups are called [eventually] orthodox.

A generalization of the concept of eventually regular will also prove convenient. Define a semigroup S to be idempotent-surjective if whenever ρ is a congruence on S and is an idempotent of S/ρ, then contains some idempotent of S. It is well known that eventually regular semigroups are idempotent-surjective [2]. Further, we have the following known result [10] (we include a simple proof for completeness).

Result 1.1

Every idempotent-surjective semigroup S is E-inversive.

Proof

Let aS. From the definition of a Rees congruence on S follows that the ideal SaS has at least one idempotent, that is, xay=eE S , where x,yS. Hence exaye=e. Thus yex=(yex)a(yex), so yexW(a), as required. □

A subset A of S is said to be (respectively) full; reflexive and dense if E S A; ∀a,bS [abAbaA] and ∀sS ∃ x,yS [sx,ysA]. Also, define the closure operator ω on S by ={sS:∃ aA [asA]} (AS). We shall say that AS is closed (in S) if =A. Finally, a subsemigroup N of a semigroup S is normal if it is full, dense, reflexive and closed (if N is normal, then we shall write NS). Moreover, if a subsemigroup of S is full, dense and reflexive, then it is called seminormal [6].

By the kernel \(\operatorname{ker}\rho\) of a congruence ρ on a semigroup S we shall mean the set {xS:(x,x 2)∈ρ}. Finally, denote by \(\mathcal{C}(S)\) the complete lattice of all congruences on a semigroup S.

Example 1.2

Consider the semigroup of positive integers (ℕ,+) (with respect to addition). It is well known that every group congruence on ℕ is of the following form: ρ n ={(k,l)∈ℕ×ℕ:n|(kl)} (n>0). Note that E =∅, so a semigroup without idempotents possesses group congruences but ℕ has not least group congruence. Also, \(\operatorname{ker}\rho_{n} = n\rho_{n} = \{n, 2n, 3n, \ldots\}\).

2 Group congruences—general case

Let S be a semigroup, \(\rho \in \mathcal{C}(S)\). We say that ρ is a group congruence if S/ρ is a group. Denote by \(\mathcal{GC}(S)\) the set of group congruences on S. Clearly, if \(\rho \in \mathcal{GC}(S)\), then \(\operatorname{ker}\rho\) is the identity of the group S/ρ. Finally, by \(\mathcal{N}(S)\) we shall mean the set of all normal subsemigroups of S.

The following two lemmas are almost evident and we omit their easy proofs.

Lemma 2.1

Let ρ be a group congruence on a semigroup S. Then \(\operatorname{ker}\rho \lhd S\).

Lemma 2.2

Let ρ 1,ρ 2 be group congruences on a semigroup S. Then ρ 1ρ 2 if and only if \(\operatorname{ker}\rho_{1} \subset \operatorname{ker}\rho_{2}\).

Let B be a nonempty subset of a semigroup S. Consider four relations on S:

Lemma 2.3

Let a subsemigroup B of a semigroup S be dense and reflexive. Then ρ 1,B =ρ 2,B =ρ 3,B =ρ 4,B .

Proof

Let (a,b)∈ρ 2,B . Then ax=yb for some x,yB. Also, asB for some sS, since B is dense and so saB, since B is reflexive. Hence asyB and so (sy)aB. It follows that (sy)b=s(yb)=s(ax)=(sa)xB. Thus (sy)a, (sy)bB. We have just shown that ρ 2,B ρ 3,B .

Conversely, if xa,xbB for some xS, then ax,bxB (since B is reflexive), so a(xb)=(ax)b, where ax,xbB. Hence (a,b)∈ρ 2,B . Thus ρ 2,B =ρ 3,B .

Dually, ρ 1,B =ρ 4,B . Since B is reflexive, then ρ 1,B =ρ 3,B . □

If B is a dense, reflexive subsemigroup of S, then we denote the above four relations by ρ B . We have the following theorem.

Theorem 2.4

Let B be a dense and reflexive subsemigroup of a semigroup S. Then the relation ρ B is a group congruence on S. Moreover, \(B\subseteq B\omega = \operatorname{ker}\rho_{B}\). If B is normal, then \(B = \operatorname{ker}\rho_{B}\).

Conversely, if ρ is a group congruence on S, then there exists a normal subsemigroup N of S such that ρ=ρ N (in fact, \(N = \operatorname{ker}\rho\)). Thus there exists an inclusion-preserving bijection between the set of all normal subsemigroups of S and the set of all group congruences on S.

Proof

Let aS. Since B is dense, then there exists xS such that xaB. Hence ρ B is reflexive. Obviously, ρ B is symmetric. Also, since B is a semigroup, then ρ B is transitive. Consequently, ρ B is an equivalence relation on S. Moreover, ρ B is a left congruence on S. Indeed, let (a,b)∈ρ B ,cS. Then ax,bxB and zcB for some x,zS, so zcax,zcbxB. It follows that (ca)(xz),(cb)(xz)∈B, since B is reflexive. Therefore (ca,cb)∈ρ B . By symmetry, ρ B is a right congruence on S. Finally, S/ρ B is a group. Indeed, let aS,bB and ax,xaB for some xS. Then baxB. Hence xa,x(ba)∈B, so (ba,a)∈ρ B . Since B is dense, then S/ρ B is a group, as required.

Since b(bb)=(bb)b for every bB, then \(B \subset \operatorname{ker}\rho_{B}\). Also, \(B\omega = \operatorname{ker}\rho_{B}\). Indeed, let \(s \in \operatorname{ker}\rho_{B}\). Then (s,b)∈ρ B for some bB. Hence b 1 s=bb 2 for some b 1,b 2B. Thus s, so \(\operatorname{ker}\rho_{B} \subset B\omega\). Conversely, let s. Then bsB for some bB. Since bbB, then (s,b)∈ρ B , so \(s \in \operatorname{ker}\rho_{B}\). Thus \(B\omega \subset \operatorname{ker}\rho_{B}\), as exactly required. Finally, if B is normal, then B=. Hence \(B = \operatorname{ker}\rho_{B}\).

Conversely, let ρ be a group congruence on S. By Lemma 2.1, \(\operatorname{ker}\rho \lhd S\). Put \(\operatorname{ker}\rho = N\). Then by Lemma 2.2, ρ=ρ N , since \(N = \operatorname{ker}\rho_{N} = \operatorname{ker}\rho\). It is now easy to see that the map \(\phi : \mathcal{N}(S) \to \mathcal{GC}(S)\), where =ρ N for every \(N \in \mathcal{N}(S)\), is an inclusion-preserving bijection between the set of all normal subsemigroups of S and the set of all group congruences on S (with the inverse \(\phi^{- 1} : \mathcal{GC}(S) \to \mathcal{N}(S)\), where \(\rho \phi^{- 1} = \operatorname{ker}\rho\) for all \(\rho \in \mathcal{GC}(S)\)). Note that ϕ −1 is an inclusion-preserving mapping, too. □

Since the first part of Theorem 2.4 is true for an arbitrary dense and reflexive subsemigroup of S, then we get the following corollary.

Corollary 2.5

Let B be a dense and reflexive subsemigroup of S. Then S.

Example 2.6

Let S={a,b,c,e,f} be the semigroup with the multiplication table given below:

figure a

It is easy to see that E is a dense and reflexive subsemigroup of S but E is not closed, since eaE and aE. Also, N={a,e,f} is normal. Indeed, the group congruence ρ E has two ρ E -classes: N and {b,c}, since ae,ee,bb,bcE and (e,b)∉ρ E . Note also that \(E \subset \operatorname{ker}\rho_{N} = N, E \not= N\) and ρ E =ρ N . It follows that there is no a one-to-one correspondence between the set of all seminormal subsemigroups of S and the set of all group congruences on S.

Remark 1

Obviously, every subgroup of a group is full and unitary but not every subgroup of a group is reflexive (for example: each two element subgroup of the group of all permutations of the six-element set X is not reflexive). It is well known that a subgroup H of a group G is normal if and only if the relation ρ H is a congruence on G. We have a corresponding result:

Let A be a closed subsemigroup of a semigroup S. Then A is normal if and only if \(\rho_{A} \in \mathcal{GC}(S)\).

Indeed, let \(\rho_{A} \in \mathcal{GC}(S)\). From A= and the second paragraph of the proof of Theorem 2.4 we obtain that \(A = \operatorname{ker}\rho_{A}\). Thus AS (Lemma 2.1). The converse of the result follows from Theorem 2.4.

The set of all group congruences on a semigroup S (in general) does not form a lattice. Indeed, let (ℝ,+) be the semigroup of real positive numbers with respect to addition. Put M=ℕ and N={x,2x,3x,…}, where x∈ℝ∖ℚ. Then M,NS but MN=∅.

We generalize now the results of Howie [12], LaTorre [16] and Hanumantha Rao and Lakshmi [9].

Theorem 2.7

Let B be a seminormal subsemigroup of a semigroup \(S, \rho \in \mathcal{C}(S)\). Then:

  1. (i)

    ρρ B =ρ B ρρ B ;

  2. (ii)

    \(\rho \lor \rho_{B} \in \mathcal{GC}(S)\);

  3. (iii)

    (x,y)∈ρρ B if and only if (ax,yb)∈ρ for some a,bB.

Proof

(i). Since ρ,ρ B ρρ B , ρ B ρρ B ρρ B . Also, ρ B ρρ B is a reflexive, symmetric and compatible relation on S. We show that ρ B ρρ B is transitive. Then ρρ B =ρ B ρρ B . Indeed, let (r,s),(s,t)∈ρ B ρρ B . Then (a) (w,s),(s,x)∈ρ B ; (b) (y,w),(x,z)∈ρ; (c) (r,y),(z,t)∈ρ B for some w,x,y,zS. From (a) we obtain (w,x)∈ρ B , so aw=xb for some a,bB. From (b) follows that (aw,ay),(xb,zb)∈ρ. Hence (ay,zb)∈ρ, since aw=xb. Finally, by (c), (r,ay),(zb,t)∈ρ B , since \(B \subset \operatorname{ker}\rho_{B}\), so (r,ay)∈ρ B ,(ay,zb)∈ρ,(zb,t)∈ρ B . Thus (r,t)∈ρ B ρρ B , as required.

(ii). This is evident.

(iii). Let (x,y)∈ρρ B . Then (x,r)∈ρ B ,(r,s)∈ρ and (s,y)∈ρ B for some r,sS. Hence ax=rb,cs=yd for some elements a,b,c,d of B. Therefore (ca)x=c(ax)=c(rb)=(crb)ρ(csb)=(cs)b=(yd)b=y(db), where ca,dbB. Conversely, let (ax,yb)∈ρ for some a,bB. Since (x,ax),(yb,y)∈ρ B , then (x,y)∈ρ B ρρ B =ρρ B (by (i)). □

Let A be a nonempty subset of a semigroup \(S, \rho \in \mathcal{C}(S)\). Put

$$A\rho = \bigl\{s \in S: \exists\, a \in A~[(s, a) \in \rho]\bigr\}. $$

Corollary 2.8

Let B be a seminormal subsemigroup of a semigroup \(S, \rho \in \mathcal{C}(S)\). Then \(\operatorname{ker}(\rho \lor \rho_{B}) = (B\rho)\omega\). In particular, ()ωS.

Proof

Let \(x \in \operatorname{ker}(\rho \lor \rho_{B})\). Then there exists bB such that (x,b)∈ρρ B , since \(B \subset \operatorname{ker}(\rho \lor \rho_{B})\). Hence (ax,bc)∈ρ for some a,cB (by Theorem 1.6(iii)). Thus ax. It follows that x∈()ω. Conversely, if x∈()ω, then ax for some a, so (ax,b),(a,c)∈ρ for some b,cB. It follows that (cx,b)∈ρ. Hence ((cc)x,cb)∈ρ. Thus (x,c)∈ρρ B . Consequently, \(x \in\operatorname{ker}(\rho \lor \rho_{B})\). □

Also, by Theorem 1.6(i) and Proposition 2.3(ii) in [15] we obtain the following (see Corollary 3.2 [15]) corollary.

Corollary 2.9

Every group congruence on a semigroup S is dually right modular element of \(\mathcal{C}(S)\).

Corollary 2.10

Let B be a seminormal subsemigroup of a semigroup \(S, \rho \in \mathcal{C}(S)\). Then ρρ B =S×S if and only if ()ω=S.

Let B be a seminormal subsemigroup of a semigroup \(S, \rho_{1}, \rho_{2} \in \mathcal{C}(S)\). Suppose that (x,y)∈(ρ 1ρ B )∩(ρ 2ρ B ). Then (ax)ρ 2(yb), where a,bB. Moreover, ax(ρ 1ρ B )x,x(ρ 1ρ B )y,y(ρ 1ρ B )yb, so ax(ρ 1ρ B )yb. Thus (cax,ybd)∈ρ 1, where c,dB. It follows that (caxd,cybd)∈ρ 1. Moreover, (caxd,cybd)∈ρ 2. Hence (xd,cy)∈(ρ 1ρ 2)∨ρ B . Thus (x,y)∈(ρ 1ρ 2)∨ρ B , since (ρ 1ρ 2)∨ρ B is a group congruence on S and \(c, d \in B \subset \operatorname{ker}((\rho_{1} \cap \rho_{2}) \lor \rho_{B})\). We have just shown that (ρ 1ρ B )∩(ρ 2ρ B )⊂(ρ 1ρ 2)∨ρ B . The converse inclusion is evident. Thus we may conclude that (ρ 1ρ B )∩(ρ 2ρ B )=(ρ 1ρ 2)∨ρ B .

We have the following theorem (see Theorem III.5.6 [21] and Theorem 4 [23]).

Theorem 2.11

Let B be a seminormal subsemigroup of a semigroup S. Then the mapping \(\phi : \mathcal{C}(S) \to \mathcal{GC}(S)\), where

$$\rho \phi = \rho \lor \rho_B $$

for every \(\rho \in \mathcal{C}(S)\), is a (lattice) homomorphism of \(\mathcal{C}(S)\) onto the (modular) lattice [ρ B ,S×S] of all group congruences on S containing ρ B .

Proof

We have just proved that (ρ 1ρ 2)ϕ=ρ 1 ϕρ 2 ϕ for all \(\rho_{1}, \rho_{2} \in \mathcal{C}(S)\). Clearly, (ρ 1ρ 2)ϕ=ρ 1 ϕρ 2 ϕ for all \(\rho_{1}, \rho_{2} \in \mathcal{C}(S)\) and evidently ϕ is onto [ρ B ,S×S]. □

We have the following corollary (see Theorem 4.5 [15]).

Corollary 2.12

Let B be a seminormal subsemigroup of a semigroup S. Then ρ B distributes over meet.

Let S be a semigroup, NS. Put

$$\mathcal{P}(S; N) = \bigl\{A \subseteq S: A^2 \subseteq A, N \subseteq A, A\omega = A\bigr\}. $$

Also, denote S/ρ N by S/N. In particular, \(\mathcal{P}(S/N; \{N\})\) is the set of all subgroups of the group S/N. Remark that if \(A \in \mathcal{P}(S; N)\), then A is full and dense.

The proofs of the following two propositions are standard and so we omit the proofs.

Proposition 2.13

Let S be a semigroup, NS. Then there exists an inclusion-preserving bijection ϕ between the set \(\mathcal{P}(S; N)\) and the set \(\mathcal{P}(S/N; \{N\})\). Moreover, \(M \in \mathcal{P}(S; N)\) and MS if and only if S/N.

Proposition 2.14

Let ϕ be an epimorphism of a semigroup S onto a group (G,⋅,1). Then:

  1. (i)

    Ker(ϕ)=ϕϕ −1 is a group congruence on S;

  2. (ii)

    N={1}ϕ −1S;

  3. (iii)

    Ker(ϕ)=ρ N .

Conversely, if NS, then N is the kernel of the canonical homomorphism of S onto S/N.

Example 2.15

We now describe all normal subsemigroups of the bicyclic semigroup S=ℕ0×ℕ0, where (k,l)(m,n)=(kl+max{l,m},nm+max{l,m}). It is known that every (non-identical) homomorphic image of the bicyclic semigroup is a cyclic group. Also, it is almost evident that E S ={(0,0),(1,1),(2,2),…}◁S and (k,l)ρ E (m,n) if and only if k+n=l+m, so S/ρ E ≅(ℤ,+). It follows that (iℤ)ϕ −1={(m,n)∈S:(m) i =(n) i } for every i∈ℕ. The conclusion is that every cyclic group is a homomorphic image of the bicyclic semigroup.

We have also the following well known proposition (from group theory).

Proposition 2.16

Let S be a semigroup; M,NS and MN. Then:

  1. (i)

    MN;

  2. (ii)

    N/MS/M;

  3. (iii)

    (S/M)/(N/M)≅S/N.

Every full and closed subsemigroup A of an E-inversive semigroup S is itself E-inversive. Indeed, let aA. Then axE S =E A for some xS, so x=A. Consequently, there is xA such that axE A .

Finally, by way of contrast, we prove in the present section the following proposition which is valid for the class of all E-inversive semigroups.

Proposition 2.17

Let S be an E-inversive semigroup, NS. Suppose also that a subsemigroup M of S is full and closed. Then:

  1. (i)

    MNM;

  2. (ii)

    N◁(MN)ω;

  3. (iii)

    M/(MN)≅(MN)ω/N.

Proof

(i). It is clear that E S MN, so MN is a full subsemigroup of M. Let a,bM be such that abMN. Then baM and baN (since N is reflexive in S). Hence baMN. Hence MN is reflexive in M. Further, if x∈(MN)ω, then yxMN for some yMN, so xMN (because N and M are closed). Since MN is full and closed, then it is E-inversive, so it is dense in M. Thus MNM.

(ii). We show that (MN)ω is a subsemigroup of S. Let a,b∈(MN)ω. Then m 1 n 1 a=m 2 n 2 for some m 1,m 2M,n 1,n 2N. Since S is E-inversive, then \(W(m_{1}) \not= \emptyset\). Hence mm 1,m 1 mE S M for some mS. Thus mM (since M is closed), (mm 1)n 1 a=(mm 2)n 2. Therefore (n 1 a,mm 2)∈ρ N , since mm 1E S N, so (a,m 3)∈ρ N (m 3M). Similarly, (b,m 4)∈ρ N for some m 4M. It follows that (ab,m 5)∈ρ N , where m 5M. Hence n 3 ab=m 5 n 4 for some n 3,n 4N. Thus (m 5 n 3)ab=(m 5 m 5)n 4. Consequently, ab∈(MN)ω. Furthermore, N⊂(MN)ω. Indeed, let nN. Then n 1 n=en 2 for some eE S ,n 1,n 2N. Hence we have (en 1)n=en 2MN, so n∈(MN)ω. Consequently, N◁(MN)ω (since NS).

The proof of the condition (iii) is standard. □

3 Group congruences on an E-inversive semigroup

Note that if a semigroup S is E-inversive, then every full subsemigroup of S is dense (since E S is dense), so a subsemigroup A of S is normal if and only if A is full, reflexive and closed. It follows that S has a least normal subsemigroup U. Thus the least group congruence on an arbitrary E-inversive semigroup exists. Denote it by σ or σ S . Then σ=ρ U and \(\operatorname{ker}\sigma = U\) (Theorem 2.4).

Firstly, we have the following proposition.

Proposition 3.1

Let S be an E-inversive semigroup. Then \(\mathcal{GC}(S) = [\sigma, S \times S]\). Thus \(\mathcal{GC}(S)\) is a complete sublattice of \(\mathcal{C}(S)\).

Also, ρ M ρ N =ρ M ρ N =ρ N ρ M for all M,NS. Hence the lattice

$$\bigl(\mathcal{GC}(S), \subseteq, \cap, \circ\bigr) $$

is modular.

Proof

The first part of the above proposition is clear. We show its second part. Let a(ρ M ρ N )b. Then (a,c)∈ρ M ,(c,b)∈ρ N , where cS. Take any xW(c). Then xc,cxE S ,(cxa)ρ N (bxa),(bxa)ρ M (bxc), so (a,bxa)∈ρ N ,(bxa,b)∈ρ M . Hence (a,b)∈ρ N ρ M . Therefore ρ M ρ N ρ N ρ M . We may equally well show the opposite inclusion. Consequently, ρ M ρ N =ρ M ρ N =ρ N ρ M . In the light of Proposition I.8.5 [11], the lattice \((\mathcal{GC}(S), \subseteq, \cap, \circ)\) is modular. □

Let M,N be normal subsemigroups of a semigroup S. From Proposition 3.1 and Corollary 2.8 we obtain that \(\operatorname{ker}(\rho_{M} \rho_{N}) = \operatorname{ker}(\rho_{N} \rho_{M}) = (M\rho_{N})\omega = (N\rho_{M})\omega\). In fact, if S is E-inversive, then \(\operatorname{ker}(\rho_{M} \rho_{N}) = \operatorname{ker}(\rho_{N} \rho_{M}) = M\rho_{N} = N\rho_{M}\). Indeed, let \(x \in \operatorname{ker}(\rho_{M} \rho_{N})\). Then (x,e)∈ρ M ρ N for some eE S . Hence (x,n)∈ρ M , (n,e)∈ρ N , where nS (in fact, \(n \in \operatorname{ker}\rho_{N} = N\)). Thus x M . Conversely, if x M , then (x,n)∈ρ M for some nN. Hence (x,n)∈ρ M ,(n,e)∈ρ N , where eE S . Thus (x,e)∈ρ M ρ N , that is, \(x \in \operatorname{ker}(\rho_{M} \rho_{N})\), so \(\operatorname{ker}(\rho_{M} \rho_{N}) = N\rho_{M}\). Similarly, \(\operatorname{ker}(\rho_{N} \rho_{M}) = M\rho_{N}\). This implies the required equalities. Also, \(\operatorname{ker}(\rho_{M} \rho_{N}) = (MN)\omega\). Indeed, let x N . Then n 1 x=mn 2 for some n 1,n 2N,mM. Hence (mn 1)xMN. Thus x∈(MN)ω. We have proved that \(\operatorname{ker}(\rho_{M} \rho_{N}) \subset (MN)\omega\). Conversely, let x∈(MN)ω. Then m 1 n 1 x=m 2 n 2 for some m 1,m 2M,n 1,n 2N. Since S is E-inversive, then mm 1=eE S M for some mS. It follows that mM (since M is closed), so en 1 x=mm 2 n 2. Hence (x,mm 2)∈ρ N . Thus \(x \in M\rho_{N} = \operatorname{ker}(\rho_{M} \rho_{N})\), so \((MN)\omega \subset \operatorname{ker}(\rho_{M} \rho_{N})\), as exactly required.

In fact, we have just shown that in an arbitrary E-inversive semigroup S, ρ (MN)ω =ρ M ρ N =ρ N ρ M =ρ (NM)ω for all M,NS. Moreover, notice that \(\operatorname{ker}(\rho_{M} \cap \rho_{N}) = \operatorname{ker}\rho_{M} \cap \operatorname{ker}\rho_{N} = M \cap N\) (M,NS), so ρ M ρ N =ρ MN for M,NS. Consequently, the lattice \((\mathcal{N}(S), \subseteq, \cap, \lor)\), where MN=(MN)ω for all M,NS, is isomorphic to the lattice \((\mathcal{GC}(S), \subseteq, \cap, \circ)\) (by the inclusion-preserving bijection ϕ, see the proof of Theorem 2.4). Note also that the lattice \((\mathcal{N}(S), \subseteq, \cap, \lor)\) is complete (since it has the greatest element S and the intersection of any nonempty family of normal subsemigroups of S is a normal subsemigroup of S).

For terminology and elementary facts about lattices the reader is referred to the book [21] (Sect. I.2). The following result will be useful (see Exercise I.2.15(iii) in [21]).

Lemma 3.2

Every lattice isomorphism of complete lattices is a complete lattice isomorphism.

From the above consideration we obtain the following theorem.

Theorem 3.3

Let S be an E-inversive semigroup. Then there exists a (lattice) isomorphism ϕ between the lattice \((\mathcal{N}(S), \subseteq, \cap, \lor)\), where MN=(MN)ω for all M,NS, and the lattice \((\mathcal{GC}(S), \subseteq, \cap, \circ)\). In fact, ϕ is defined by =ρ N for every \(N \in \mathcal{N}(S)\). Moreover, ϕ is a complete lattice isomorphism.

Finally, we have the following proposition.

Proposition 3.4

Let S be an E-inversive semigroup, NS. Then (a,b)∈ρ N if and only if ab N for some (all) b W(b).

Proof

(⟹). Let na=bm, where n,mN, and b W(b). Then nab =bmb . Since b bmN and N is reflexive, then nab N. Hence ab =N.

(⟸). Let ab =nN for some b W(b). Then a(b b)=nb, so (a,b)∈ρ N (by Lemma 2.3). □

4 Group congruences on an E-semigroup

First, we “generalize” some results from orthodox semigroups to E-semigroups (see Theorem VI.1.1 [11]).

Proposition 4.1

Let S be a semigroup. The following conditions are equivalent:

  1. (i)

    S is an E-semigroup;

  2. (ii)

    a,bS [W(b)W(a)⊆W(ab)].

Moreover, the condition (i) implies the following condition:

  1. (iii)

    eE S  [W(e)⊆E S ].

If in addition S is an R-semigroup, then the conditions (i)(iii) are equivalent.

Proof

The proof is closely similar to the proof of Theorem VI.1.1 [11]. □

Corollary 4.2

Let S be an E-semigroup. Then:

  1. (i)

    eE S  [W(e),V(e)⊆E S ];

  2. (ii)

    aS,a W(a),eE S  [aea ,a eaE S ];

  3. (iii)

    aS,a W(a),e,fE S  [ea ,a e,ea fW(a)].

Proof

(i). This follows from Proposition 4.1.

(ii). This follows from the proof of Proposition VI.1.4 [11].

(iii). Let aS,a W(a),e,fE S . Since eW(e) and fW(f), then ea W(e)W(a)⊆W(ae). Hence ea =ea aeea =(ea )a(ea ). Therefore ea W(a). Similarly, a eW(a). Finally, ea fW(e)W(a)W(f)⊆W(fae) and so ea f=ea ffaeea f=(ea f)a(ea f). Hence ea fW(a). □

Proposition 4.3

Let S be an E-inversive E-semigroup. Then

$$\rho_{1, E} = \rho_{2, E} = \rho_{3, E} = \rho_{4, E}. $$

Proof

Let (a,b)∈ρ 2,E and a W(a). Then ae=fb for some e,fE. Moreover, a fW(a) (Corollary 4.2(iii)), so (a f)a,a(a f)∈E. Further, a fb=a aeE. We have just shown that xa,ax,xbE for some xS. Thus ρ 2,E ρ 4,E .

On the other hand, if xa,xbE for some xS, say xa=e,xb=f, then (efx)a(efx)=ef(xa)efx=efx, so efxW(a). Also, fxbfx=f(xb)fx=fx, i.e., fxW(b). Hence efxW(b) (Corollary 4.2(iii)). Thus \(W(a) \cap W(b) \not= \emptyset\). It follows that ay,by,ya,ybE for some yS. Dually, if ax,bxE for some xS, then ay,by,ya,ybE for some yS. Thus ρ 4,E =ρ 1,E . In fact, we get \(\rho_{4, E} = \rho_{1, E} = \{(a, b) \in S \times S: W(a) \cap W(b) \not= \emptyset \}\). Finally, if xW(a)∩W(b), then a(xb)=(ax)b and xb,axE. Thus ρ 2,E =ρ 4,E =ρ 1,E . We may equally well show that ρ 3,E =ρ 4,E =ρ 1,E . Consequently, ρ 1,E =ρ 2,E =ρ 3,E =ρ 4,E . □

Lemma 4.4

Let S be an E-inversive E-semigroup. Then:

  1. (i)

    aS ∃ e,fE S  [ea,afReg(S)];

  2. (ii)

    \(\forall a \in S~\exists\, r \in \mathit{Reg}(S)~[W(a) \cap W(r) \not= \emptyset]\).

Proof

Let aS,xW(a). Then (ax)a,a(xa)∈Reg(S), where ax,xaE S , so (i) holds. Also, r=axaReg(S) and xrx=x. Thus xW(a)∩W(r). □

Denote the above four relations from Proposition 4.3 by ρ E . Recall that from the proof of Proposition 4.3 follows that \(\rho_{E} = \{(a, b) \in S \times S: W(a) \cap W(b) \not= \emptyset \}\).

Theorem 4.5

In any E-inversive E-semigroup, σ=ρ E . Moreover, \(\operatorname{ker}\sigma = E_{S}\omega\). Thus E S ωS.

Proof

It is clear that ρ E is an equivalence relation on S. Let (a,b)∈ρ E ,cS. Then xW(a)∩W(b). Take any yW(c). In the light of Proposition 4.1,

$$xy \in W(a)W(c) \cap W(b)W(c) \subseteq W(ca) \cap W(cb). $$

Hence (ca,cb)∈ρ E . Thus ρ E is a left congruence on S. We may equally well show that ρ E is a right congruence on S. Also, if e,fE S , then ee,efE S . Consequently, (e,f)∈ρ E for all e,fE S . Lemma 4.4(ii) says that every ρ E -class of S contains a regular element. This implies that S/ρ E is a group.

Furthermore,

$$ x \in \operatorname{ker}\rho_E\ \Leftrightarrow\ \exists\, e \in E_S\ [(x, e) \in \rho_E]\ \Leftrightarrow\ \exists\, e, f, g \in E_S\ [fx = eg]\ \Leftrightarrow\ x \in E_S\omega, $$

so \(\operatorname{ker}\sigma = E_{S}\omega\). Thus E S ωS (Theorem 2.4). Finally, ρ E ρ N for ever NS. Indeed, E S N. Hence E S ω=N. Thus \(\rho_{E} = \rho_{E_{S}\omega} \subseteq \rho_{N}\) (Theorem 2.4). Consequently, σ=ρ E . □

Corollary 4.6

The least group congruence σ on an E-inversive E-semigroup is given by

$$\sigma = \bigl\{(a, b) \in S \times S: \exists\, e \in E_S~[eae = ebe]\bigr\}. $$

Remark 2

Note that the condition “∃ eE S  [eae=ebe]” from the above corollary is equivalent to the apparently weaker condition “∃ sS [sas=sbs]”.

From Result 1.1 and Theorem 4.5 we obtain the following theorem.

Theorem 4.7

In any idempotent-surjective E-semigroup, σ=ρ E .

Let S be a semigroup. A congruence ρ on S is called idempotent pure if E S for every eE S . Note that if S is idempotent-surjective, then ρ is idempotent pure if and only if \(\operatorname{ker}\rho = E_{S}\). Let \(\mathcal{E}\) be an equivalence relation on S induced by the partition: {E S ,SE S }. Then \(\mathcal{E}^{\flat}\) (defined in [13], see p. 27) is the greatest idempotent pure congruence on S. Put \(\tau = \mathcal{E}^{\flat}\). Then (see [13], p. 28)

$$\tau = \bigl\{(a, b)\in S \times S: \forall x, y\in S^1~[xay\in E_S \iff xby\in E_S]\bigr\}. $$

Finally, if S is E-inversive, then τσ. Indeed, let (a,b)∈τ and b W(b). Then bb E S ,(ab ,bb )∈τ. Hence \(ab^{*} \in E_{S} \subseteq \operatorname{ker}\sigma\). In the light of Proposition 3.4, (a,b)∈σ, as exactly required. In the following corollary we give an alternative proof of this fact.

Corollary 4.8

If ρ is a congruence on an idempotent-surjective E-semigroup S, then \(\operatorname{ker}(\rho \lor \sigma) = (\operatorname{ker}\rho)\omega\). In particular, τσ.

Proof

By Corollary 2.8, \(\operatorname{ker}(\rho \lor \sigma) = (E_{S}\rho)\omega = (\operatorname{ker}\rho)\omega\). In particular,

$$\operatorname{ker}(\tau \lor \sigma) = E_S\omega \subseteq \operatorname{ker}\sigma. $$

Hence τσ=σ. Thus τσ. □

Let ρ be a congruence on a semigroup S. By the trace \(\operatorname{tr}\rho\) of ρ we shall mean the restriction of ρ to E S . Also, we say that ρ is idempotent-separating if \(\operatorname{tr}\rho = 1_{E_{S}}\). Edwards in [3] shows that if S is an eventually regular semigroup, then the relation \(\theta = \{(\rho_{1}, \rho_{2}) \in \mathcal{C}(S) \times \mathcal{C}(S): \operatorname{tr}\rho_{1} = \operatorname{tr}\rho_{2}\}\) is a complete congruence on \(\mathcal{C}(S)\) and proves that every θ-class ρθ is a complete sublattice of \(\mathcal{C}(S)\) with the maximum element

$$\mu (\rho) = \bigl\{(a, b) \in S \times S: (a\rho, b\rho) \in \mu_{S/\rho}\bigr\} $$

and the minimum element 1(ρ). Edwards generalizes some of these results for the class of all idempotent-surjective semigroups [4]. In fact, if S is an arbitrary idempotent-surjective semigroup, then every θ-class ρθ is the interval [1(ρ),μ(ρ)], where μ is the maximum idempotent-separating congruence on S (see [4] for more details).

It is easily seen that the class of idempotent-surjective semigroups is closed under homomorphic images [10]. Using the obvious terminology we show next that every homomorphism of idempotent-surjective E-semigroups can be factored into a homomorphism preserving the maximal group homomorphic images and an idempotent-separating homomorphism. Firstly, we have need the following lemma.

Lemma 4.9

Let ρ be a congruence on an idempotent-surjective E-semigroup S, a,bS. Then (,)∈σ in S/ρ implies (a,b)∈σ if and only if ρσ.

Proof

The proof is closely similar to the proof of Lemma III.5.9 [21]. □

Let S be an idempotent-surjective E-semigroup, \(\rho \in \mathcal{C}(S)\). Clearly, (a,b)∈σ implies (,)∈σ. In the light of Lemma 4.9, if ρσ, then (a,b)∈σ if and only if (,)∈σ. Hence S/σ≅(S/ρ)/σ, that is, S and S/ρ have isomorphic maximal group homomorphic images. In that case, we may say that ρ preserves the maximal group homomorphic images. Since for any congruence ρ on S we have 1(ρ)⊆ρ, then we obtain the following factorization:

$$S \to S/1(\rho) \to S/\rho \cong \bigl(S/1(\rho)\bigr)\big/\bigl(\rho/1(\rho)\bigr). $$

The following proposition generalizes Proposition III.5.10 [21].

Proposition 4.10

Every homomorphism of idempotent-surjective E-semigroups can be factored into a homomorphism preserving the maximal group homomorphic images and an idempotent-separating homomorphism.

Proof

Let ρ be any congruence on an idempotent-surjective E-semigroup S. Since ρS×S, then 1(ρ)⊆1(S×S). Clearly, σ∈[1(S×S),S×S] and so 1(ρ)⊆σ. It follows that the canonical epimorphism of S onto S/1(ρ) preserves the maximal group homomorphic images. Finally, an epimorphism ϕ:S/1(ρ)→S/ρ (defined by the obvious way) is idempotent-separating, since \(\operatorname{tr}\rho = \operatorname{tr}(1(\rho))\). The thesis of the proposition is a consequence of the above factorization. □

5 Group congruences on an E-unitary semigroup

A nonempty subset A of a semigroup S is called left [right] unitary if asA [saA] implies sA for every aA,sS. Also, we say that A is unitary if it is both left and right unitary. Finally, a semigroup S with \(E_{S} \not= \emptyset\) is said to be E-unitary if E S is unitary.

Proposition 5.1

Let S be a semigroup with \(E_{S} \not= \emptyset\). The following conditions are equivalent:

  1. (i)

    S is E-unitary;

  2. (ii)

    E S is left unitary;

  3. (iii)

    E S is right unitary.

Also, if S is an E-unitary E-inversive semigroup, then S is an E-semigroup.

Proof

\(\mathrm{(i)} \Longrightarrow \mathrm{(ii)}\). This is trivial.

\(\mathrm{(ii)} \Longrightarrow \mathrm{(iii)}\). Let sS,eE S . If se=fE S , then fsef=f and so we get (efs)(efs)=efs, that is, efsE S . Hence fsE S . Thus sE S .

\(\mathrm{(iii)} \Longrightarrow \mathrm{(i)}\). We may equally well show like above that E S is left unitary. Thus the condition (i) holds.

Finally, let S be an E-unitary E-inversive semigroup. If e,fE S ,xW(ef), then xefE S . Hence xef,xE S . Thus efE S . □

Corollary 5.2

Let S be an E-inversive semigroup. Then the following conditions are equivalent:

  1. (i)

    S is E-unitary;

  2. (ii)

    \(\operatorname{ker}\sigma = E_{S}\);

  3. (iii)

    τ=σ.

In particular, if S is an E-unitary E-inversive semigroup, then E S S.

Proof

\(\mathrm{(i)} \Longrightarrow \mathrm{(ii)}\). In the light of Proposition 5.1 and Theorem 4.5, \(\operatorname{ker}\sigma = E_{S}\omega\). Also, S is left unitary, that is, E S is closed. Thus \(\operatorname{ker}\sigma = E_{S}\).

\(\mathrm{(ii)} \Longrightarrow \mathrm{(iii)}\). We have mentioned above that τσ. On the other hand, the main assumption implies that σ is idempotent pure. Hence στ. Thus τ=σ.

\(\mathrm{(iii)} \Longrightarrow \mathrm{(i)}\). Let aS,e,fE S . If ea=f, then \(a \in \operatorname{ker}\sigma = \operatorname{ker}\tau = E_{S}\), that is, E S is left unitary. In the light of Proposition 5.1, S is E-unitary. □

Remark 3

Notice that if a semigroup is not E-inversive, then Corollary 5.2 is false. Indeed, let F X 1 be the free monoid on the set X. Then F X 1 is E-unitary but τ is induced by the partition {F X ,{1}}. Thus τ is not a group congruence.

From Proposition 3.4 and Corollary 5.2 we obtain the following proposition.

Proposition 5.3

Let S be an E-unitary E-inversive semigroup. Then (a,b)∈σ if and only if ab E S for some (all) b W(b).

Corollary 5.4

Let A be an E-inversive subsemigroup of an E-unitary E-inversive semigroup S. Then σ A =σ S ∩(A×A).

Proof

Clearly, σ A σ S ∩(A×A). The converse follows from Proposition 5.3. □

In [14] Howie and Lallement showed that \(\sigma \cap \mathcal{H} = 1_{S}\), when S is an E-unitary regular semigroup. We prove a corresponding result.

Theorem 5.5

Let S be an E-unitary E-inversive semigroup. Then \(\sigma \cap \mathcal{H} = 1_{S}\). Moreover, if in addition E S forms a semilattice, then \(\sigma \cap \mathcal{L} = \sigma \cap \mathcal{R} = 1_{S}\).

Proof

Let S be an E-unitary E-inversive semigroup. Suppose also that E S forms a semilattice. Then E S is normal (Corollary 5.2), so if \((a, b) \in \sigma \cap \mathcal{L}\), then ax=e, bx=fE S for some xS (see Proposition 5.3) and sa=b,tb=a for some s,tS. Hence se=sax=bx=fE S ,tf=tbx=ax=eE S . Thus s,tE S (since E S is unitary), so since idempotents commute and ta=tb,

$$a = tb = t(sa) = (ts)a = (st)a = s(ta) = s(tb) = sa = b. $$

We may equally well show that \(\sigma \cap \mathcal{R} = 1_{S}\).

If S is E-unitary, then E S is normal, too. Let \((a, b) \in \sigma \cap \mathcal{H}\). By the above proof and its dual we conclude that a=eb=bf and b=ga=ah for some e,f,g,hE S . In the light of Proposition 2 in [18], a=b. □

Remark 4

The assumption that S is an E-inversive semigroup is important. Indeed, let S=(ℝ0,+) be the semigroup of nonnegative real numbers with respect to addition. Then S is an E-unitary commutative semigroup. Put M=ℕ0 and N={0,x,2x,3x,…} (where x∈ℝ∖ℚ). Then M,NS but MN={0} is not normal, so S has no least group congruence.

The converse of Theorem 4.15 is not valid (in general). Indeed, let S=〈x〉, where x=(2 3 4 5 6 7 5) is a mapping of \(\mathcal{T}(\{1, 2, \ldots, 7\})\). Then S=M(4,3) is the monogenic semigroup with index 4 and period 3, say S={x,x 2,…,x 6}. Also, the cyclic subgroup K x of S with the unit e is equal {x 4,x 5,x 6=e}. Since x 3 e=x 7 x 2=x 4 x 2=e, then S is not E-unitary. On the other hand, σ is induced by the partition: {{x,x 4},{x 2,x 5},{x 3,e}} and \(\mathcal{H}\) by the partition: {K x ,{x},{x 2},{x 3}}. Thus \(\sigma \cap \mathcal{H} = 1_{S}\).

From Theorem 5.5 and Corollary 5.2 we have the following corollary.

Corollary 5.6

Let S be an E-unitary E-inversive semigroup. Then

$$\sigma \cap \mathcal{H} = \tau \cap \mathcal{H} = 1_S. $$

Moreover, if in addition E S forms a semilattice, then

$$\sigma \cap \mathcal{L} = \tau \cap \mathcal{L} = \sigma \cap \mathcal{R} = \tau \cap \mathcal{R} = 1_S. $$

Recall that a congruence ρ on a semigroup S is E-unitary if S/ρ is E-unitary. In [5] the author described the least E-unitary congruence κ on an idempotent-surjective semigroup. Also, for every congruence ρ on an idempotent-surjective semigroup S there exists the least E-unitary congruence κ ρ on S containing ρ [5].

Let S be an idempotent-surjective semigroup, NS. Define the relation \(\hat{\rho_{N}}\) on \(\mathcal{C}(S)\) by the following rule: \((\rho_{1}, \rho_{2}) \in \hat{\rho_{N}} \Leftrightarrow \rho_{1} \lor \rho_{N} = \rho_{2} \lor \rho_{N}\) (\(\rho_{1}, \rho_{2} \in \mathcal{C}(S)\)). Then \(\hat{\rho_{N}}\) is a congruence on \(\mathcal{C}(S)\), since \(\phi \phi^{- 1} = \hat{\rho_{N}}\) (see Theorem 2.11).

Also, we prove the following proposition.

Proposition 5.7

Let S be an idempotent-surjective semigroup, \(N \lhd S, \rho \in \mathcal{C}(S)\). Then the elements ρ,κ ρ ,ρρ N are \(\hat{\rho_{N}}\)-equivalent and ρκ ρ ρρ N . Moreover, the element ρρ N is the largest in the \(\hat{\rho_{N}}\)-class \(\rho \hat{\rho_{N}}\).

Proof

Since κ ρ is the least E-unitary congruence containing ρ and clearly ρρ N is E-unitary, then ρκ ρ ρρ N . Hence ρρ N κ ρ ρ N ρρ N . Therefore ρρ N =κ ρ ρ N . Thus \((\rho, \kappa_{\rho}) \in \hat{\rho_{N}}\). Evidently, \((\rho, \rho \lor \rho_{N}) \in \hat{\rho_{N}}\). This implies the first part of the proposition. The second part is clear. □

Remark 5

Recall from [22] that in the class of inverse semigroups not every \(\hat{\sigma}\)-class has a least element.

Finally, it is easy to see that the least E-unitary congruence κ on an arbitrary E-inversive semigroup exists, too. We show that \(\mathcal{H}\cap\sigma\subseteq\kappa\) in any E-inversive semigroup. Firstly, we have need the following useful proposition.

Proposition 5.8

Let B be the least seminormal subsemigroup of an E-inversive semigroup S. If ϕ is an epimorphism of S onto an E-unitary semigroup T, then E T .

Proof

Put A=(E T )ϕ −1. Clearly, A is a full subsemigroup of S, so A is dense. Further, if xyA, then E T ∋(xy)ϕ===(yx)ϕ (since E T is reflexive), so yxA. Hence BA. Thus ⊆((E T )ϕ −1)ϕE T . □

We may now prove the following equivalent theorem to Theorem 5.5.

Theorem 5.9

In any E-inversive semigroup S, \(\mathcal{H}\cap\sigma\subseteq\kappa\). If in addition E S forms a semilattice, then \(\mathcal{L}\cap\sigma\subseteq\kappa\) and \(\mathcal{R}\cap\sigma\subseteq\kappa\).

Proof

Indeed, σ=ρ B , where B is the least seminormal subsemigroup of S. Let \((a, b)\in\mathcal{H}\cap\sigma\). Then clearly \((a\kappa, b\kappa)\in\mathcal{H}^{S/\kappa}\). Also, ax=yb for some a,bB. In the light of Proposition 5.8, ()()=()(), where ,E S/κ . Hence \((a\kappa, b\kappa)\in\mathcal{H}^{S/\kappa}\cap\sigma_{S/\kappa}= 1_{S/\kappa}\) (Theorem 5.5). Thus \(\mathcal{H}\cap\sigma\subseteq\kappa\), as required. □

6 Group congruences on an eventually regular semigroup

Group congruences on eventually regular semigroups were described in [9] by Hanumantha Rao and Lakshmi. In the paper [9] the following definition was introduced: a subset A of S is called self-conjugate if x r(x)−1(x r(x)) AxA and xAx r(x)−1(x r(x))A for all xS,(x r(x))V(x r(x)). We say that A is self-conjugate if the former condition holds.

Lemma 6.1

Let N be a subsemigroup of an eventually regular semigroup S. Then N is normal if and only if N is full, self-conjugate and closed.

Proof

Let N be normal, xS,(x r(x))V(x r(x)). Then N is full and closed. Also, x r(x)(x r(x)) NENN, so x r(x)−1(x r(x)) NxN, since N is reflexive.

Let N be full, self-conjugate and closed, xyN,(x r(x))V(x r(x)). Then x r(x)−1(x r(x))(xy)xx r(x)−1(x r(x)) NxN, i.e., (x r(x)−1(x r(x)) x)(yx)∈N, where x r(x)−1(x r(x)) xE S N. Hence yx=N, so N is reflexive. Thus NS. □

Lemma 6.2

Let S be an eventually regular semigroup, NS. Then

$$\rho_N = \bigl\{(a, b) \in S \times S: \exists\, \bigl({b^{r(b)}}\bigr)^{*} \in V\bigl(b^{r(b)}\bigr)~\bigl[ab^{r(b) -1}\bigl({b^{r(b)}}\bigr)^{*} \in N\bigr]\bigr\}. $$

Proof

Let (a,b)∈ρ N and (b r(b))V(b r(b)). Then na=bm for some n,mN. Hence nab r(b)−1(b r(b))=bmb r(b)−1(b r(b)). Also, since b r(b)−1(b r(b)) bE S , then mb r(b)−1(b r(b)) bNE S N, so nab r(b)−1(b r(b))=bmb r(b)−1(b r(b))N, since N is reflexive. Consequently, ab r(b)−1(b r(b))=N.

Conversely, let a,bS,(b r(b))V(b r(b)) and ab r(b)−1(b r(b))=nN. Then a(b r(b)−1(b r(b)) b)=nb, where b r(b)−1(b r(b)) bE S N. Hence (a,b)∈ρ N . □

We have the following corollary (see Theorem 1 [9]).

Corollary 6.3

Let S be an eventually regular semigroup, NS. Then

$$\rho_N = \bigl\{(a, b) \in S \times S: \exists\, \bigl({b^{r(b)}}\bigr)^{*} \in V\bigl(b^{r(b)}\bigr)~\bigl[ab^{r(b) -1}\bigl({b^{r(b)}}\bigr)^{*} \in N\bigr]\bigr\} $$

is a group congruence on S.

Finally, we give some remarks concerning group congruences on inverse semigroups. Firstly, consider the following result (see Exercise 7(ii) [11], p. 181).

Statement 6.4

An inverse subsemigroup N of an inverse semigroup S is normal if and only if (Nx)ω=(xN)ω for every xS.

This result is false. Indeed, let S be a Clifford semigroup. Put \(N = \mathcal{Z}(S)\), where \(\mathcal{Z}(S) = \{s \in S: \forall a \in S~[sa = as]\}\). Clearly, N is a full subsemigroup of S. Also, N is self-conjugate. If the result is valid, then N is normal (since Nx=xN for every xS). Hence ρ N =S×S=ρ S , when S=S 0. It follows that every Clifford semigroup is commutative, a contradiction. Consequently, we conclude that the above result is false. Moreover, the assumptions of the result and the conditions: “N is full” and “N is self-conjugate” do not imply that (Nx)ω=(xN)ω for every xS.

It is clear that every subgroup of a group is full and closed. We prove now a correct version of the above statement.

Proposition 6.5

A full and closed inverse subsemigroup N of an inverse semigroup S is normal if and only if (Nx)ω=(xN)ω for every xS.

Proof

It is easy to see that if N is normal, then (Nx)ω=(xN)ω for every xS.

Conversely, let (Nx)ω=(xN)ω for every xS. It is easy to check that two relations ρ 1={(a,b)∈S×S:ab −1N} and ρ 2={(a,b)∈S×S:a −1 bN} are equivalences on S and that 1=(Nx)ω, 2=(xN)ω for every xS. Also, ρ 1 is right compatible and ρ 2 is left compatible. Indeed, we show first that the equality (A())ω=(AB)ω holds for all A,BS. Recall from [11] that

$$H\omega = \bigl\{s \in S: \exists\, h \in H~[h \leq s]\bigr\}\quad(H \subseteq S), $$

where ≤ is the so-called natural partial order on (an inverse semigroup) S (i.e., ab⇔∃ eE S  [a=eb]). Notice that ≤ is compatible. Let x∈(A())ω. Then ayx for some aA,y (that is, by for some bB). Hence abayx. Thus x∈(AB)ω. We have just proved that (A())ω⊂(AB)ω. The opposite inclusion is clear. Let now (a,b)∈ρ 2,cS. Then (aN)ω=(bN)ω and so (c(aN)ω)ω=(c(bN)ω)ω. Therefore (caN)ω=(cbN)ω. Thus ρ 2 is a left congruence on S. We may equally well show that ρ 1 is a right congruence on S. Since (Nx)ω=(xN)ω and 1=(Nx)ω, 2=(xN)ω for every xS, then ρ 1=ρ 2 is a congruence on S. Put for simplicity ρ=ρ 1=ρ 2. Finally, if eE S , then E S N==(eN)ω. Hence ρ is a group congruence on S and \(\operatorname{ker}\rho = N\). Thus NS, as required. □

Corollary 6.6

A Clifford semigroup S is commutative if and only if \(\mathcal{Z}(S)\) is closed in S (i.e., if and only if for every sS there exists \(z \in \mathcal{Z}(S)\) such that zs).

Lemma 6.7

Let S be a finite inverse semigroup with semilattice of idempotents E. Then =S if and only if S has zero.

Proof

It is clear that if S has zero, then =S. Conversely, let =S. Since E is finite, then E has the least idempotent with respect to the natural partial order, say 0. Let sS=. Then e=fs and e=sg for some e,f,gE (see Proposition V.2.2 in [11]). Hence 0=0s=s0. Thus S=S 0, as required. □

By an analogy to groups we may introduce the concept of a σ-simple inverse semigroup in the class of finite inverse semigroups without 0. From Lemma 6.7 follows that every finite inverse semigroup S without zero has at least one non-universal group congruence, so S has exactly one non-universal group congruence if and only if S/ is a simple group. Hence we may say that a finite inverse semigroup S without zero is σ-simple if S/ is a simple group. This definition is equivalent to the following definition: S is σ-simple if S has exactly two normal subsemigroups, namely: and S.

Example 6.8

Let (E,≤) be a chain with the least element 0. Put S=E∪{a}, where aE and aaa=a. Assume also that aa=0. Hence a=aaa=0a=a0. It is easy to see that if a binary operation on S is associative, then ea=ae=a for every eE S . For example, ea=e(0a)=(e0)a=0a=a. Conversely, it is straightforward to verify that such defined binary operation is associative. Thus S is a semigroup. Since a=a −1, then S is an inverse semigroup. Finally, E=, so S/E={E,{a}}.

7 The hypercore of a semigroup

In [8] Hall and Munn studied the hypercore of a semigroup. In this section we give some remarks on the hypercore of E-inversive E-semigroups and inverse semigroups.

Let S be a semigroup with \(E_{S} \not= \emptyset\). Denote by ℘ S the set of all subsemigroups A of S such that A has no cancellative congruences except the universal congruence. Note that {e}∈℘ S for every eE S . Define the hypercore \(\operatorname{hyp}(S)\) of S, as follows: \(\operatorname{hyp}(S) = \langle \bigcup \{A: A \in \wp_{S} \} \rangle\) [8]. Furthermore, by the core \(\operatorname{core}(S)\) of an E-inversive semigroup S we shall mean \(\operatorname{ker}\sigma\).

In [8] the authors showed the following two results.

Result 7.1

Let S be an E-inversive semigroup. Then:

  1. (i)

    \(\operatorname{hyp}(S) \in \wp_{S}\);

  2. (ii)

    \(\operatorname{hyp}(S)\) is full and unitary;

  3. (iii)

    \(\forall \rho \in \mathcal{GC}(S)~[\operatorname{hyp}(S) \subseteq \operatorname{ker}\rho]\).

Result 7.2

In any E-inversive semigroup S, \(\operatorname{hyp}(S)\) is the greatest E-inversive subsemigroup of S with no non-universal group congruence.

Let U be the least full unitary subsemigroup of an E-inversive semigroup S. Clearly, \(U \subseteq \operatorname{hyp}(S) \subseteq \operatorname{core}(S)\).

Finally, we have the following proposition.

Proposition 7.3

Let S be an E-inversive E-semigroup such that \(1_{S} \notin \mathcal{GC}(S)\). Then \(U = \operatorname{hyp}(S) = \operatorname{core}(S) = E_{S}\omega\). In particular, E S ω has no non-universal group congruence.

If in addition S is an inverse semigroup and E S ω is finite, then E S ω is an inverse semigroup with zero. In particular, every finite inverse semigroup S (which is not a group) contains exactly one normal inverse subsemigroup with zero.

Proof

Let S be an E-inversive E-semigroup. Then \(\operatorname{core}(S) = E_{S}\omega\) (Theorem 4.5). Since E S U and U is closed, then E S ωU, so \(U = \operatorname{hyp}(S) = \operatorname{core}(S) = E_{S}\omega\). In the light of Result 7.2, E S ω has no non-universal group congruence.

If S is an inverse semigroup, then obviously \(U = \operatorname{hyp}(S) =\operatorname{core}(S) = E_{S}\omega\) has no non-universal group congruence. Finally, if E S ω is finite, then E S ω has zero (Lemma 6.7). The rest of the proposition is now immediate. □