Chapter 5
Structure of Fundamental Regular Semigroups

Definition 5.1. We call a regular semigroup S a fundamental regular semigroup if and only if the greatest idempotent separating congruence on S is the identity congruence. Given a regular semigroup S, it can be easily seen that S|μ(s) is a fundamental regular semigroup.

Throughout this chapter, we denote a biordered set by E and a partially transitive set on E by T. We give below some additional notations needed in the sequel. For e,f E define

T (e,f) = {α Teα = e,fα = f}. (5.1)

Then

T (e,f)0 eδ(T )f.

If α T (e,f),β T (g,h), then by axiom (T3) we have

αβ T (e,fβ)  if fωg T (gα1,h) if gωf. (5.2)

Also for eωreα and fωlfα(α 1) define

e α = τr(e,eτr(e α)) α,α f = α τl(fτl(f 2),f). (5.3)

Then if α T ,e α T (e,(eτr(eα))α) and α f T ((fτl(fα))α1,f). Also for any α,β T and h S(fα,eβ), it follows from (5.2) and (5.3) that

α h h β T ((hτl(f α))α1,(hτr(e β))β). (5.4)

For eωeα(fωfα), e α is the domain restriction (α f is the range restriction) of α to ω(e)(ω(f)). Therefore when fαωeβ(eβωfα), (5.4) reduces to (5.2) if we choose h = fα(h = eβ).

In 1 define a releation σ as follows:

ασβ eαreβ,fαlfβ andeβ α = β fα. (5.5)

It is not difficult to see that σ is an equivalence relation on 1. We denote the σ-class of α 1 by ᾱ. If α T since 0 T it is clear from (5.5) that β T for all βσα or ᾱ T. Hence

T̄ = {ᾱα T}̄1 = 1σ (5.6)

and therefore if T ,TPT (E)

T TT̄ T̄. (5.7)

Following is the main structure theorem of this chapter.

Theorem 5.1. Let E be a biordered set and T PT (E). Define for α,β T,

ᾱ β̄ = α h h β¯ where h S(fα,eβ). (5.8)

Then T̄ is a fundamental regular semigroup under the operation defined by (5.8) and E(T̄) is isomorphic to E.

Conversely if S is a fundamental regular semigroup, then S is isomorphic to T̄s.

We prove this theorem through a sequence of lemmas.

Lemma 5.2. Let α 1,

  1. If f1,f2 ωl(fα),f1rf2 then τr(e 1,e 2) (α f 2) = (α f1) τr(f 1,f2)

    where ei = (fiτl(fα))α1,i = 1,2.

  2. If e1,e2 ωr(eα),e1le2 then τl(e 1,e2) (e2 α) = (e1 α) τl(f 1,f 2)

    where fi = (eiτr(eα))α,i = 1,2.

Proof. By hypothesis e1re and for all g ω(e1),

(g)τr(e 1,e 2) α = ((g)τr(e 2))α = (gα)τr(e 2α).

Since gα ω(e1α) = ω(f1τl(fα)),

(g)τr(e 1,e 2) α = (g)α τr(f 1τl(f 2),f2τl(f α)).

Hence the required result is equivalent to the following:

α τr(f,τl(f α),f2τl(fα)) τl(f 2τl(f α),f2) = α τl(f,τl(fα),f 1) τr(f 1,f2).

Since α is a partial ω-isomorphism, this is equivalent to the equality established in Lemma 3.7. Statement (2), which is the dual of (1), can be proved similarly.

Lemma 5.3. Let α,β T and h1,h2 S(fα,eβ). Then

α h1 h1 βσα h2 h2 β.

Proof. First suppose that h1rh2. Then h1τr(eβ) = h2τr(eβ) and therefore

h1 β = τr(h 1,h1τr(e β)) β = τr(h 1,h2τr(e β)) β = τr(h 1,h2) h2 β.

If hi = (hiτl(f2))α1,i = 1,2 then h1rh2, and by Lemma 5.2,

τr(h 1,h 2) (α h 2 h2 β) = α h1 τr(h 1,h2) h2 β = α h1 h1 β.

If h = (h1τr(eβ))β = (h2τr(eβ))β, by (5.4),

α hi hi β T (hi,h),i = 1,2,

and hence by (5.5) α h1 h1 βσα h2 h2 β.

If h1lh2, the required relation can be proved analogously. If h1 and h2 are any two members of S(fα,eβ) then by Lemma 2.7 there exists h2S(fα,eβ) such that h1rh2lh2 and hence the lemma.

Lemma 5.4. For i = 1,2, let αi T (ei,fi),βi T (gi,hi) and ki S(fi,gi). Suppose that α1σα2 and β1σβ2. Then

α1 k1 k1 β1σα2 k2 k2 β2.

Proof. The hypotheses imply that

e1re2,f1lf2 e1 α2 = α1 fα

and

g1rg2,h1lh2 g1 β2 = β1 h2.

Then by Lemma 2.6, S(f1,g1) = S(f2,g2). In view of Lemma 5.3, we need only establish that for any k S(f1,g1)

α1 k k β1σα2 k k β2.

The statement α1σα2 implies that

α21 = τl(f 2,f1) α1 τr(e 1,e2).

Hence if ei = (kτl(fi))αi1 and hi = (kτr(gi))βi

e2 = (kτl(f 2))τl(f 2,f1) α11 τr(e 1,e2) = ((kτl(f 1))α11)τr(e 1,e2) = e1τr(e 1,e2).

Dually h2 = h1τl(h1,h2) and thus we get e1re2 and h1lh2.

Since αi k k βi T (ei,hi) we need only establish the following assertion:

τl(e 1,e 2) α 2 k k β2 = α1 k k β1 τl(h 1,h 2).

For Lemma 3.1, taking c = (f1,f2),h = kτl(f1) we get

τl(kτl(f 1),kτl(f 1)) τl(f 1,f2) = τl(kτl(f 1),kτl(f 2)).

Similarly it can be seen that

τr(kτr(g1),kτr(g1)) τr(g1,g2) = τr(kτr(g1),kτr(g2)) τr(e1,e1) τr(e1,e2) = τr(e1,e2)

and

τl(h 1,h 1) τl(h 1,h2) = τl(h 1,h 2).

Hence if e ω(e1)

(e)τr(e 1,e 2) α 2 = (e)τr(e 1,e2) α2 = (e)α 1 τl(f 1,f2) = (e)α 1 τl(kτl(f 1),kτl(f 2)).

Therefore

τr(e 1,e 2) α 2 k = α1 τl(kτl(f 1),kτl(f 2)) τl(kτl(f 2),k) = α1 τl(kτl(f 1),k) = α1 k.

The dual statement

k β2 = k β1 τl(h 1,h 2),

can be proved similarly. Thus

α1 k k β1 τl(h 1,h 2) = τr(e 1,e 2) α 2 k k β2,

and hence the lemma.

Lemma 5.5. Let α,β T and for ki ωl(fα) ωr(eβ),i = 1,2 define

ei = (k iτl(f α))α1 andh i = (k iτr(e β))β.

If

(k2τl(f α),k2τr(e β))ωr × ωl(k 1τl(f α),k1τr(e β))

then

e2 (α k 1 k1 β)σα k2 k2 βσ(α k1 k1 β) h2.

Proof. By (5.4), we have for i = 1,2, α ki ki β T (ei,hi). Also e2ωre1 and h2ωlh1. We obtain the lemma if we establish the following equality:

τr(e 2,e 2τr(e 1)) α k 1 k1 β τl(h 2τl(h 1),h 2) = α k 2 k2 β. (5.9)

From the fact that α is a partial ω-isomorphism we have

τr(e 2,e 2τr(e 1) α = α αr(k 2τl(f α),k1τl(f α))

where αr(e1,e2) is defined by (3.26). Similarly

β τl(h 2τl(h 1),h 2) = αl(k 2τr(e β),k1τr(e β))1 β.

Hence (5.9) is equivalent to

α αl(k 2,fα)1 αr(k 2,eβ) β = α T β

where

T = αr(k 2τl(f α),k1τl(f α)) αl(k 1,f2)1 αr(k 1,eβ) αl(k 2τr(e β),k1τr(e β)).

Since α and β are partial ω-isomorphisms, this is equivalent to

T = αl(k 2,fα)1 αr(k 2,eβ),

which is the same as the equality in Lemma 3.8.

Lemma 5.6. Let α,β,γ T ,h1 S(fα,eβ)h2 S(fβ,er),h1 = (h1τr(eβ))β and h2 = (h2τl(fβ))β1. If h S(fα,h2) and hS(h1,er) then

(α h1 h1 β) h h rσα h h (β h 2 h2 r).

Proof. By (5.4)

α h1 h1 β T (h1τl(f α))α1,h,

and

β h2 h0 r T (h2,(h 2τr(e r))r).

Hence the expressions on either side of σ in the required relation are defined. Further by Lemma 3.9, there exist h S(fα,h2) and hS(h1,er) such that

(hτr(e β))β = hτl(f β).

By Lemma 5.3 we may without loss of generality choose h and h so as to satisfy the above equality.

Since h ωl(fα) ωr(eβ) and h1 S(fα,eβ),h and h1 satisfy conditions of Lemma 5.5 and hence we get by the equality (hτr(eβ))β = hτl(fβ),

(α h1 h1 β) hτl(f β)σα h h β.

It is clear that

(α h1 h1 p) hτl(f β)σ(α h1 h1 β) h

and

(α h h β) hσα h h β.

Hence

(α h1 h1 β) h h γσ(α h h β) h h γ.

But

(α h h β) h = (α h τr(h 1hτr(e β)) β) τl(hτl(f β),h).

Since (hτr(eβ))β = hτl(fβ), the range of h β is ω(hτl(fβ)) and the domain of β h is ω(hτr(eβ)), and therefore,

(h β) h = h (β h).

Hence

(α h1 h1 β) h h γσα h (h β) h h γ

and the dual statement

α h h (β h2 h2 γ)σα h h (β h) h γ

can be obtained in the same way. Since (h β) h = h (β h), the lemma is proved.

Lemma 5.7. If T P(E), T̄ = T σ is a regular semigroup under the operation defined by Lemma 5.8.

Proof. If x,y T̄, for α x,β y and h S(fα,eβ), by Lemma 5.8

x y = α h h β¯.

By Lemma 5.4, the σ-class α h h β does not depend on the choice of α in x, β in y and h S(fα,eβ). Hence α h h β¯ is a σ-class of T depending only on the classes x and y, and therefore Lemma 5.8 defines a binary operation on T̄.

Let x,y,z T̄, α x, β y and γ z. Then if h1 S(fα,eβ), by (5.4)

α h1 h1 β T (h1τl(f α)α1,(h 1τr(e β))β).

Now for hS(h1,er), h1 = (h1τr(eβ))β

(x y) z = (α h1 h1 β) h h γ¯.

Similarly, if h2 S(fβ,eγ), h S(f2,h2) where h2 = (h2τ2(fβ))β1 then

(x (y z)) = α h h (β h2 h2 γ)¯.

Thus by Lemma 5.6,

(x y) z = x (y z).

If x T̄, α x then α1 ,(fα,eα) and by axiom (T2) of Definition 3.1, α1 T. Also if x = α1¯

x x x = α α1 α¯ = ᾱ = x,

x x x = α1 α α1¯ = α1¯ = x.

Hence T̄ is a regular semigroup.

Lemma 5.8. For every e E. ̄(e,e) E(T̄) and

eωrf ϵ̄(e,e)ωrϵ̄(f,f). (5.10)

Moreover, the mapping

αᾱ : T (e,f)T̄ (5.11)

is a bijection of T (e,f) onto the -class Hϵ̄(e,e),ϵ̄(f,f) of T̄. Consequently,

eδ(T )f ϵ̄(e,e)δ(T̄)ϵ̄(f,f). (5.12)

Proof. That ϵ̄(e,e) E(T̄) for all e E is clear from the definition of operation in T̄.

Now let eωrf. Then eτr(f) S(e,f) and ϵ̄(f,f) ϵ̄(e,e) = ϵ(eτr(f),eτr(f)) ϵ(eτr(f),e)¯ = ϵ̄(eτr(f),e). But clearly ϵ(eτr(f),e)σϵ(e,e) and hence

ϵ̄(f,f) ϵ̄(e,e) = ϵ̄(e,e)

which implies that ϵ̄(e,e)ωrϵ̄(f,f) in E(T̄), by (1.10)

On the other hand if ϵ̄(e,e)ωrϵ̄(f,f) and if h S(f,e) then

ϵ̄(e,e) = ϵ̄(f,f) ϵ̄(e,e) = ϵ(hτl(f)),h ϵ(h,hτr(e))¯.

Hence

hτl(f)relhτr(e).

Since hτr(e)ωe,elhτr(e) implies e = hτr(e) or hre. Hence h = hτl(f).

erhωf.

Let α T (e,f). Then α1 T (f,e) and we have

ᾱ α1¯ = ϵ̄(e,e),α1¯ ᾱ = ϵ̄(f,f)

ϵ̄(e,e) ᾱ = ᾱ = ᾱ ϵ̄(f,f).

Hence ᾱ Hϵ̄(e,e),ϵ̄(f,f) and so the mapping αᾱ maps T (e,f) into Hϵ̄(e,e),ϵ̄(f,f). Let x HĒ(e,e),ϵ̄(f,f). Then if α x as observed earlier,

x = ᾱ Hϵ̄(eα,eα),ϵ̄(fα,fα).

Hence by (5.10) and (5.10)*, ereα, flfα. Then if α = τr(e,e2) α τl(fα,f),ασα and αT (e,f). Thus the mapping defined by (5.11) is onto Hϵ̄(e,e),ϵ̄(f,f). From the definition of σ it is clear that

α,αT (e,f),ασαα = α;

which shows that the mapping defined by (5.11) is one-to-one.

Further

eδ(T )f T (e,f) Hϵ̄(e,e),ϵ̄(f,f) ϵ̄(e,e)δ(T̄)ϵ̄(f,f),

which is statement (5.12).

Lemma 5.9. Let ψ denote the mapping of E into E(T ) defined by

eψ = ϵ̄(e,e) (5.13)

for every e E. Then ψ is an isomorphism of E onto the biordered set of idempotents E(T̄) of T̄. Moreover,

α ψ = ψ a(ᾱ,α1¯) (5.14)

for every α T.

Proof. We have already noted that ψ maps E into E(T̄) and that ψ and ψ1 preserve both ωr and ωl (by (5.10) and (5.10)*).

Let ᾱ E(T̄) where α T (e,f). Then if h S(f,e) we have

ᾱ ᾱ = α h h α¯ = ᾱ.

Hence

er(hτl(f))α1andfl(hτr(e))α.

Since α is a partial ω-isomorphism we have

hτl(f)reαandhτr(e)lfα1.

Hence

hτl(f) = fandhτr(e) = e.

Therefore

flhre.

By Lemma 5.8, we see that ϵ¯(h,h) and α¯ are -equivalent idempotents and consequently

α¯ = ϵ¯(h,h).

This proves that ψ is onto. Being one-to-one, it is a bijection.

Let eωrf. Then e S(e,f) and by (4.1),

eψτr(fψ) = eψ fψ = ϵ(e,eτr(f)) ϵ(f,f)¯ = ϵ¯(e,eτr(f)).

But

ϵ(e,eτr(f))σϵ(eτr(f),eτr(f))

and therefore

eψτr(fψ) = ϵ¯(eτr(f),eτr(f)) = (eτr(f))ψ.

Let e,f E and h S(e,f). Then h S(e,h) S(h,f) and hence

eψ hψ fψ = ϵ(e,e) ϵ(hτl(e),ϵ(h,hτr(f))) ϵ(f,f)¯ = ϵ(e,e) ϵ¯(f,f).

Then by (4.4) hψ S(eψ,fψ).

Hence ψ is a bimorphism and similarly ψ1 is also a bimorphism. Thus ψ is an isomorphism.

Let α T and e ω(eα). Then

(eψ)a(α¯,α1¯) = α1¯ eψ α¯ = α1 ϵ(e,e) α¯.

Now

α1 ϵ(e,e) α = ϵ(eα,eα)

and therefore

(eψ)a(α¯,α1¯) = (eα)ψ.

This proves (5.14).

Remark 5.10: We have already noted that given any regular semigroup S we can associate with S a partially transitive set Ts on E(S) (of Theorem 4.12). For any T PT (E),T¯ is a regular semigroup and hence we can define a partially transitive set TT¯ on E(T¯) by (4.13). It will be of interest to find out the relation between T and TT¯. In fact the mapping

α a(α¯,α1¯) (5.15)

is a bijection of T onto TT¯ which reduces to identity mapping when E and E(T¯) are identified by ψ.

That (5.15) defines a mapping of T into TT¯, is clear. Now suppose that x T¯ and x i(x). Let eψ = x x,fψ = x x so that x Heψ,fψ and x Hfψ,eψ. By Lemma 5.8, there exists α T (e,f) such that α¯ = x. Then α1¯ is an inverse of x in Hfψ,eψ and so α1¯ = x. Thus

a(α¯,α1¯) = a(x,x)

and this proves that the mapping defined by (5.15) is onto. Now if for some x T¯ and x i(x), we have

a(α¯1,α11) = a(x,x)

where α1 T (e1,f1), then α¯,x,α11¯,x and therefore there exists αT (e1,f1) such that

α¯ = x,α11¯ = x.

However, for all e ω(e1), by (5.14)

(eα1)ψ = (eψ)a(α1¯,α11¯) = (eψ)a(α¯,α11¯) = (eα)ψ.

Thus eα1 = eα and this being true for all e ω(e1) we conclude that α = α1. Hence the mapping defined by (5.15) is also one-to-one.

Lemma 5.11. T ¯ is a fundamental regular semigroup.

Proof. For x,y T¯, by Theorem 4.13,

xμ(T¯)y a(x,x) = a(y,y)

for some x i(x) and y i(y). T ¯ is fundamental if we establish that

x,y T¯,a(x,x) = a(y,y) x = y. (5.16)

Suppose that x,y T¯,x i(x),y i(y) and a(x,x) = a(y,y). Choose α,β T so that α¯ = x,α1¯ = x,β¯ = y and β1¯ = y.

Then

a(α¯,α1¯) = a(β¯,β1¯).

Hence by Remark 5.10 α = β and consequently

x = α¯ = β¯ = y.

Thus we have established the first part of Theorem 5.1 (i.e., for all T PT (E),T¯ under the operation defined by (5.8), is a fundamental regular semigroup and E is isomorphic to E(T¯)). We obtain the converse part of Theorem 5.1 from the following theorem.

Theorem 5.12. Let S be a regular semigroup and TS denote the partially transitive set on E = E(S) defined by (4.13). Then for x S and x i(x)

xϕS = a(x,x)¯ (5.17)

defines a homomorphism of S onto TS such that

ϕS ϕS1 = μ(S). (5.18)

Proof. It follows from the definition of σ and Lemma 4.11 (c), that for all x S and x,x i(x),

a(x,x)σa(x,x)

and thus ϕs defines a mapping of S into TS. For x,y S, h S(x,y),x i(x) and y i(y) we have by Theorem (4.8) yhx i(xy). Therefore by Lemma 4.11 (b) and (d),

a(xy,yhx) = a(xh,hx) a(hy,yh) = a(x,x) τl(hτl(f),h) τr(h,hτr(e)) a(yy) = a(x,x) h h a(y,y)

where e = xx and f = yy. By (5.8)

a(x,x) h h a(y,y)¯ = a(x,x)¯ a(y,y)¯

and so

(x y)ϕs = a(xy,yhx)¯ = a(x,x)¯ a(y,y)¯ = xϕs yϕs.

Now suppose that xμ(s)y. Then by the definition of μ(s) for some x i(x), y i(y)

a(x,x) = a(y,y)

and thus

xϕs = yϕs.

On the other hand if xϕs = yϕs, then for some x i(x),y i(y)

a(x,x)σa(y,y).

Thus by (5.5)

xx = ere1 = yy xx = flf1 = yy.

In other words xy and for y = i(y,f,e),

a(x,x),a(y,y) Ts(e,f)

and

a(x,x)σa(y,y).

We now conclude that

a(x,x) = a(y,y);

and thus xμ(s)y.

Proof of the theorem is now complete.

For any biordered set E, since PT (E), we have the following corollary:

Corollary 5.13. Every biordered set is isomorphic to the biordered set of idempotents of some regular semigroup.

Theorem 5.14. If

(E) = {T̄|T PT (E)} (5.19)

then (E) is a complete lattice under inclusion relation and the mapping

T T̄ (5.20)

is a lattice isomorphism of PT (E) onto (E).

Proof. The mapping defined by (5.20) is an order isomorphism of PT (E) onto (E) by (5.7) and hence the result follows from this and Theorem 3.5.

Remark 5.15: A different construction for the greatest and the least members of the lattice (E) is given by Hall [5].

Remark 5.16: If S is a regular semigroup, the operation of S induces a biordered structure on E(S) (cf. Theorem 4.5). By 5.13, the structure of any biordered set E may be registered as having been induced by the operation of a regular semigroup S that contains E isomorphically as its idempotent set. It is clear that the operation of S induces a partial operation of E and the partial groupoid so formed determines the structure of E (as a biordered set). But there are in general more than one such partial groupoids on E that determine the same biorder structure on E.

For example let G be any group, aij(i,j = 1,2) be distinct elements of G and e be its identity. Suppose that

P = a11a12 a21 a22 ,P = ee e e .

Then

S1 = M(G,2,2;P),S1 = M(G,2,2;P)

(where 2 denotes the index set {1, 2}) are two completely simple semigroups such that E(S1) and E(S1) are isomorphic as biordered sets but are different as partial groupoids. In fact, as partial groupoids E(S1) is a band, but E(S1) may not.

It is natural to define a biordered set E as a band if ̄0 is a band semigroup. Also E is a band in S if the partial groupoid E is a band semigroup. Clearly if E is a band in S it is a band (as a biordered set). But as the previous example shows, the converse of this is not true.

The structure of band semigroup was determined by McLean [8], and of those semigroups whose biordered set of idempotents is a band in that semigroup was determined by Yamada [13] (Yamada [13] calls such semigroups as strictly regular semigroups). The following theorem characterizes bands in the class of all biordered sets and strictly regular semigroups in terms of generalised Brandt groupoids (cf [14]).

Theorem 5.17. Let E be a biordered set. Then E is a band if and only if the following conditions are satisfied.

e,fϵE,er fl el fr (i)

where er(el) denotes the r (l)-class of E containing e.

ϵ,ϵ0(e,f) ϵ = ϵ. (ii)

If S is a regular semigroup, E(S) is a band in S if and only if the trace (cf. p. 92 of [2]) of each D-class of S is isomorphic to a generalized Brandt semigroup.

Proof. Suppose that the biordered set E satisfies the hypothesis and let eδ0f. Then by Theorem 3.3, there exists a δ0-chain c = (e0,e1,,en) such that e0 = e,fn = f. We may assume that for i = 1,2,,n 1, the elementary factors (ei1,ei) and (ei,ei+1) of are not similar. Let e0re1. Then by hypothesis

e0re1le2re3.

Hence there exists e1 e0l e2r such that

e0le1re3

and again by hypothesis we conclude that there exists e1 such that

e0re1le3.

Continuing this process, we obtain by induction that there exist g,g E such that

e = e0rglen = f

and

elgrf.

Now suppose that ϵ(c) 0(e,f). Then by the definition of σ, and (ii),

ϵ(c)σϵ(g,g),ϵ(c1)σϵ(g,g).

Thus

ϵ(c)¯ = ϵ(g,g)¯,ϵ(c1)¯ = ϵ(g),g¯.

Hence by Lemma 5.9, ϵ(c)¯,ϵ(c1)¯ E(̄0). It follows that ̄0 is a band semigroup, and hence E is a band.

Conversely let E be a band. Then ̄0 is a band semigroup. Let e,f E,er fl. Then there exists g E such that erglf and therefore x = τr(e,g) τl(g,f)¯ ̄0. Clearly

Ẽ0(e,f) = {τr(e,g) τl(g,f)}.

Further,

i(x;f,e) = τl(f,g) τr(g,e)¯

and this is an idempotent. Hence by Lemma 5.9, there exists g E such that

τl(f,g) τr(g,e)σ (y,y),

which implies that frgle

Let S be a regular semigroup such that E(S) is a band in S and D be a D-class of S. Fix a maximal subgroup He,e of S contained in D. We shall show that it is possible to choose elements

ref He,f,rfe Hf,e,f E(D),

such that ref = ref if and only if flfrfe = rfe if and only if frf and ref i(rfe)

For this let {eα|α E(D)|δ0} be a subset of E(D) such that for any f E(D),e δ0(f). Choose re,eα He,eα and let reα,e be its inverse in Heα,e. For each f δ(eα) define

ref = reeαf.

Clearly ref He,f and ref = reeα if and only if fleα. Also if flf

refeα = reeα = refeα

(since feα is an idempotent -equivalent to eα, refeα = reeα). Since the mapping x xeα of Lf onto Leα is one-to-one, we have ref = ref. Thus we have chosen ref He,f for all f E(D). In a similar way we can choose rfe Hf,e for all f E(D) also by defining rfe = freαe if f δ0(eα). Then we can show that rfe = rfe if and only if frf.

Now let f E(D). Then

rgerefrfe = freαereeαfreαe = feαfreαe.

Now feαrf and hence feαf = f. Therefore

rferefrfe = rfe.

Similarly

refrferef = ref.

Hence

ref i(rfe).

It is clear that if we define the matrix

Pe (P λie|λ E(D)l,i E(D)r)

by

Pλie = rfe ref if rfe ref He,e 0  otherwise

where λ is the l-class of E(D) containing f and i is the r-class of E(D) containing f, then every nonzero entry of Pe is the identity of He,e. Using the fact that fδ0f if and only if there exist f,fϵE(D) such that

frflf,flf1rf.

We can prove that the matrix constructed above is the (E(D)l) × (E(D)r) semi-identity matrix of type {(αl) × (αr)} (cf. Yamada [14]).

Now by Theorem 3.4 of [2], the trace of D is isomorphic to

M(H e,e,E(D)r,E(D)l;Pe)

and this is a generalized Brandt semigroup by lemma 5 of [14].

Suppose now that trace of each D-class of S is a generalized Brandt semigroup. Let e,f E(S). Then for any h S(e,f),eh,h,hf E(S) and ehlhrhf. Hence the product (eh)(hf) = ef is not zero in the trace of Dh. Since the set of idempotents of the trace of Dh is a band, it follows that ef is a nonzero idempotent of the trace of Dh and hence ef E(S). Hence E(S) is a band. This completes the proof of Theorem 5.17.

If E(S) is a semilattice (i.e., ωr = ωl) it can be easily seen that it is a semilattice in S (i.e., a commutative subsemigroup of S). Hence in this case the statement that E(S) is a semilattice is equivalent to the statement that E(S) is a semilattice in S.

Remark 5.18:

When E is a semilattice, it is clear that σ defined in 1 by (5.5) reduces to the identity relation. Moreover, the operation defined by (5.8) is then equivalent to the usual composition of relations. Therefore 1 is an inverse semigroup whose semilattice is isomorphic to E. Also, any partially transitive set T on E is a subsemigroup of 1 since axiom (T3) in this case is equivalent to the fact that α,β T implies α β T. Hence every fundamental inverse semigroup whose semilattice is isomorphic to E is isomorphic to a subsemigroup of 1.

The analogous results for bisimple inverse semigroup are due to W. D. Munn (cf. [11]).