Abstract

Let be a commutative ring and a finite locally inverse semigroup. It is proved that the semigroup algebra is isomorphic to the direct product of Munn algebras with , where is the number of -classes in , the number of -classes in , and a maximum subgroup of . As applications, we obtain the sufficient and necessary conditions for the semigroup algebra of a finite locally inverse semigroup to be semisimple.

1. Main Results

A regular semigroup is called a locally inverse semigroup if for all idempotent , the local submonoid is an inverse semigroup under the multiplication of . Inverse semigroups are locally inverse semigroups. Inverse semigroup algebras are a class of semigroup algebras which is widely investigated. One of fundamentally important results is that a finite inverse semigroup algebra is the direct product of full matrix algebras over group algebras of the maximum subgroups of this finite inverse semigroup. Consider that all local submonoids of a locally inverse semigroup are inverse semigroups, it is a very natural problem whether a finite locally inverse semigroup algebra has a similar representation to inverse semigroup algebras. This is the main topic of this note.

Let be an -algebra. Let and be positive integers, and let be a fixed matrix over . Let be the vector space of all matrices over . Define a product in by where is the usual matrix product of , , and . Then is an algebra over . Following [1], we call the Munn   matrix algebra over    with sandwich matrix  .

By a semisimple semigroup, we mean a semigroup each of whose principal factor is either a completely 0-simple semigroup or a completely simple semigroup. It is well known that a finite regular semigroup is semisimple. The Rees theorem tells us that any completely -simple semigroup (completely simple semigroup) is isomorphic to some Rees matrix semigroup (), and vice versa (for Rees matrix semigroups, refer to [1]). In what follows, by the phrase “Let be a finite regular semigroup,” we mean that is a finite regular semigroup in which the principal factor of determined by the -class is isomorphic to the Rees matrix semigroup or for any .

The following is the main result of this paper.

Theorem 1.1. Let be a finite locally inverse semigroup. Then the semigroup algebra is isomorphic to the direct product of with .

Based on Theorem 1.1 and [1, Lemma 5.17, page 162, and Lemma 5.18, page 163], the following corollary is straightforward.

Corollary 1.2. Let be a finite locally inverse semigroup. Then the semigroup algebra has an identity if and only if and is invertible in the full matrix algebra for all .

Reference [1, Lemma 5.18, page 163] told us that is isomorphic to the full matrix algebra if has an identity. Now, we have the following.

Corollary 1.3. Let be a finite locally inverse semigroup. If has an identity, then is isomorphic to the direct product of the full matrix algebras with .

The following corollary is a consequence of Corollary 1.3.

Corollary 1.4. Let be a finite locally inverse semigroup. Then the semigroup algebra is semisimple if and only if for all
(1);(2) is invertible in the full matrix algebra ;(3) is semisimple.

2. Proof of Theorem 1.1

For our purpose, we have the Möbius inversion theorem [2].

Lemma 2.1. Let () be a locally finite partially ordered set (i.e., intervals are finite) in which each principal ideal has a maximum and be an Abelian group. Suppose that is a function and define by Then where is a Möbius function.

Now assume that is a regular semigroup and . Define Then is a partial order on . Following [3], we call the natural partial order on . Equivalently, if and only if for every (for some) , there exists such that and . Moreover, Nambooripad [3, 4] proved that is a locally inverse semigroup if and only if the natural partial order is compatible with respect to the multiplication of .

Lemma 2.2. Let be a locally inverse semigroup and . Then for any , there exist and such that , and .

Proof. For any , we have and . Let be an inverse of . Clearly, . Note that . It is easy to check that , and . Hence and there exists such that and . Thus . On the other hand, since is a left congruence and since , we have ; while since , we have . These imply that . Dually, we have such that , and . Since , we know that and are the required elements and .

Define a multiplication on by where is the product of and in . By the arguments of [4, page 9], is a semigroup. We denote by the semigroup . For any , we denote . It is easy to check that is a subsemigroup of , which is isomorphic to the principal factor of determined by . We will denote the semigroup by . By the definition of , it is easy to see that in the semigroup ,

(i) for all ;(ii) for all such that . Thus is the direct sum of the contracted semigroup algebras with . Note that is isomorphic to some principal factor of . We observe that is a completely -simple semigroup since is a semisimple semigroup, and thus is isomorphic to some Rees matrix semigroup . By a result of [1], is isomorphic to . Consequently, to verify Theorem 1.1, we need only to prove that is isomorphic to .

For the convenience of description, we introduce the semigroup . Put . Define a multiplication on as follows: where we will identify with . It is easy to see that is isomorphic to . Hence the contracted semigroup algebra is isomorphic to the contracted semigroup algebra . For , we denote . It is easy to check that is a subsemigroup of isomorphic to the semigroup . So, for any , we have

For Theorem 1.1, it remains to prove the following lemma.

Lemma 2.3. .

Proof. We consider the mapping given on the basis by . Clearly, is well defined. Of course, and may be regarded as the mappings of the ordered set () into the additive group of . Now, by applying the Möbius inversion theorem to the mappings and , we have where is the Möbius function for (). Hence is surjective.
We will prove that is injective. For , we denote by the set and by the set of maximal elements in the set with respect to the partial order . In recurrence, we define , where . Let with . If , then by the definition of , , where and , and hence , thus and for any . This can imply the following.
Fact 2.4. If , then and by the definition of , .
By the definition of , the following facts are immediate.
Fact 2.5. if and only if and .
Fact 2.6. If and , then .
Note that and . We thus have a smallest integer such that . Clearly, . This means that is the smallest integer such that . Similarly, there exists the smallest integer such that and . Now, assume . By using Fact 2.4 repeatedly, But , we have and by the definition of , . Thus by the minimality of , and . Similarly, . Therefore . Since , by Fact 2.6, we have since and . Again by the hypothesis , and by Fact 2.4, ; and by (2.6), . By Fact 2.5, and imply ; moreover, by using Fact 2.5 repeatedly, and . We have now proved that is injective.
Finally, for any , by (2.4), we have and by Lemma 2.2, Moreover, by Lemma 2.2, we have Thus is a homomorphism of into . Consequently, is an isomorphism of onto .

Acknowledgment

The research is supported by the NSF of Jiangxi Province, the SF of Education Department of Jiangxi Province, and the SF of Jiangxi Normal University.