Left reductive regular semigroups
Abstract.
In this paper, we develop an ideal structure theory for the class of left reductive regular semigroups and apply it to several subclasses of popular interest. In these classes, we observe that the right ideal structure of the semigroup is ‘embedded’ inside the left ideal one, and so we can construct these semigroups starting with only one object (unlike in other more general cases). To this end, we introduce an upgraded version of the Nambooripad’s normal category [41] as our building block, which we call a connected category.
The main theorem of the paper describes a category equivalence between the category of left reductive regular semigroups and the category of connected categories. Then, we specialise our result to describe constructions of -unipotent semigroups, right regular bands, inverse semigroups and arbitrary regular monoids. Exploiting the left-right duality of semigroups, we also construct right reductive regular semigroups and use that to describe the more particular subclasses of -unipotent semigroups and left regular bands. Finally, we provide concrete (and rather simple) descriptions to the connected categories that arise from finite transformation semigroups, linear transformation semigroups (over a finite dimensional vector space) and symmetric inverse monoids.
Key words and phrases:
1. Introduction
The most important algebraic invariants of any given semigroup are its Green relations which describe the ideal structure of the semigroup. Introduced in a seminal paper [22] in 1951, Green relations are certain equivalence relations defined on a semigroup which partition the semigroup elements into an ‘egg-box’ diagram (for example, see Figures 3, 4 and 5). In this partitioning, the elements generating the same principal left ideals fall in the same column of the egg-box and those generating the same principal right ideal fall in the same row of the diagram. This captures a lot of information regarding the local and global structure of the semigroup. In fact, it is precisely this partitioning that makes semigroups manageable, in spite of them being rather general objects. Hence, it is no surprise that Howie would remark the Green relations are “so all-pervading that, on encountering a new semigroup, almost the first question one asks is ‘What are the Green relations like?’” [30]. Therefore, it is very natural that any structure theorem for semigroups may aim to begin with building blocks that abstract the principal left (and right) ideals. To make any headway in this direction, one needs to understand the inter-relationship between principal left and principal right ideal structures. In general, this is quite complicated and invariably involves two ‘ordered objects’ (each abstracting left and right ideal structures of the semigroup) interconnected in a non-trivial fashion. Unsurprisingly, this rather difficult question is still open in many general cases. Such a quest leads naturally towards the special class of regular semigroups.
In regular semigroups, each principal left (or right) ideal is generated by idempotent elements, giving some control over the structure of the semigroup. Indeed there is a very close relationship between the ideal structure and the idempotent structure, and in fact, we can obtain one from the other [38, 6, 7]. Recall that these semigroups were introduced by Green in [22], wherein he credited Rees for the suggestion to adopt von Neumann’s definition [51] from ring theory.
Definition 1.1.
A semigroup is said to be (von Neumann) regular semigroup if for every element in , there exists such that .
Historically, one of the first major leaps into regular semigroups was by Hall [25] who extended Munn’s [35] construction of fundamental inverse semigroups to fundamental regular semigroups generated by idempotents. To this end, Hall considered certain transformations on the partially ordered set (poset) of the principal left ideals and on that of the principal right ideals. Later, Grillet [23] gave an abstract characterisation of these posets as regular posets, and introduced the notion of cross-connection to describe the exact relationship between the left and the right posets of a regular semigroup. Simultaneously, Nambooripad [36, 37, 39] developed the idea of (regular) biordered set, as an abstract model of the set of idempotents of a (regular) semigroup, and using groupoids gave a general structure theorem for regular semigroups. This seminal work also described an equivalence between the category of regular semigroups and the category of certain groupoids, and in the process, it puts final touches to the celebrated ESN (Ehresmann-Schein-Nambooripad) Theorem [32]. Although extremely clever, Nambooripad’s description of the sets of idempotents as biordered sets is still complicated and pretty cumbersome to work with, especially for constructions. In 1978, Nambooripad [38] showed that regular biordered sets and cross-connected regular posets are equivalent. Elaborating on this fact, he developed his theory of cross-connections [41] by replacing regular posets with normal categories (which contain regular posets as subcategories). In this way, he proved that the category of regular semigroups is equivalent to the category of cross-connected normal categories.
A major problem with such a general approach is that the theory developed is too heavy to be applied to the vast majority of the objects which may have a rather simple structure! In fact, we believe this is one of the major reasons why Nambooripad’s cross-connection theory has not achieved the popularity and acclaim that such a deep work deserves. Addressing this wide gap in the literature is one of the main motivations behind this paper.
As the reader may observe, the entire discussion in this paper can be traced back to the Cayley’s theorem for groups. Just as any group may be realised as a subgroup of the symmetric group on the set , it is well-known that any semigroup can be looked at as a transformation semigroup on the set , [29]. This may be achieved by considering the regular representation of [13, Section 1.3], which is the homomorphism , , where , for every . Adjoining the element to the set (if is not a monoid) is sufficient to ensure injectivity of the representation, and so in this case is isomorphic .
So, a question arises: for which classes of semigroups do we have injectivity without adjoining ? This leads us to left reductive semigroups.
Definition 1.2.
A semigroup is said to be left reductive if the regular representation is injective.
In this paper, we discuss the ideal structure of the class of left reductive regular semigroups111All semigroups considered in this paper are regular. So, in the sequel, we often write just left reductive semigroups instead of left reductive regular semigroups. and apply our structural result to obtain constructions for several popular subclasses.
Left reductive semigroups include in particular, all regular monoids, the -unipotent semigroups, the inverse semigroups, the right regular bands, the full (linear) transformation semigroups, the singular transformation semigroups on a finite set, and the semigroup of singular linear transformations over a finite dimensional vector space. Each of these classes occurs naturally in various branches of mathematics, statistics or physics. We will discuss these in detail. Observe that left reductive semigroups exclude several ‘simple’ classes like completely simple semigroups [4], bands [44] and regular- semigroups [15].
In left reductive semigroups, the relationship between the left and the right ideals is relatively simpler and more transparent than in arbitrary regular semigroups. Roughly speaking, in the left reductive case, the left and the right ideals are very tightly interconnected, and the -structure totally restricts the -structure. More precisely, given a left reductive semigroup with the category of principal left ideals, we may show that the poset of principal right ideals of is isomorphic to an order ideal of the poset of right ideals of the semigroup that arises from the category (Proposition 3.17). As a result, we can avoid the use of several complicated notions needed in a general discussion of arbitrary regular semigroups (see [41]).
Just as an element of a group may be represented as a permutation of the set , in this paper we represent an element of a left reductive semigroup by a cone in the category (see Definition 2.3). In an arbitrary regular semigroup [41], Nambooripad used a pair of cross-connected cones to represent a typical element. Notice that in [41], Nambooripad also briefly outlined the construction of left reductive semigroups222Nambooripad had reversed the convention and called these semigroups as right reductive in [41]. We shall follow Clifford and Preston [13, Section 1.3] wherein these semigroups are named left reductive. This will coincide with the conventions in -unipotent semigroups later (see Section 4). inside his framework of cross-connections. However, his construction involved two normal categories and their duals, certain -valued functors, and the rather sophisticated definition of cross-connection relating all these categories (see [5, Section 5] for a concrete construction of a left reductive semigroup using cross-connections).
In contrast, taking advantage of the less complicated structure on hand, our construction uses just one normal category and bypasses most of the complicated tools of [41] including the cross-connections. We believe that our construction may drastically reduce the entry threshold of the ideal approach to the theory of the structure of semigroups.
The paper is divided into eight sections. After this introduction, in Section 2, we briefly recall the essential preliminary notions regarding semigroups and categories. We also discuss the initial layer of our construction including the notion of normal category from Nambooripad’s treatise [41]. In Section 3, we bifurcate ourselves from [41] and introduce connected categories. We describe the structure of left reductive semigroups using connected categories in Section 3.3 and prove a category equivalence in Section 3.4. We specialise the discussion to the category of -unipotent semigroups in Section 4, culminating in another category equivalence with the category of supported categories. In Section 5, we further particularise to describe a category adjunction between the category of supported categories and the category of right regular bands. As mentioned earlier, the left and the right Green relations induce a natural left-right duality in semigroups. We use this duality to show that the category of connected categories is also equivalent to the category of right reductive semigroups in Section 6. In Section 7, we discuss the arguably most important class of semigroups: the class of inverse semigroups. Here, we introduce self-supported categories which capture the isomorphism between the left and the right ideal structures in inverse semigroups. This leads to a category equivalence just as in [8] but this is much simpler and sheds more light on the symmetry of these semigroups. Finally, in Section 8, we discuss regular monoids and totally left reductive semigroups. One of the interesting results in this section is Theorem 8.4 which identifies the category of semigroups corresponding to the category of normal categories. We describe regular monoids using bounded above normal categories and provide concrete descriptions of connected categories in some of important semigroups like transformation semigroups, linear transformation semigroups and symmetric inverse monoids. These new descriptions subsume (and improve) the discussions in [5, 2] and also illustrate the precision of our construction.
The following table lists the various categories of left reductive regular semigroups considered in the paper and their corresponding categories of connected categories:
Semigroups | Categories | |
---|---|---|
- left reductive regular semigroups | - connected categories | |
- -unipotent semigroups | - supported categories | |
- right regular bands | — | |
- right reductive regular semigroups | - connected categories | |
- -unipotent semigroups | - supported categories | |
- left regular bands | — | |
Inverse semigroups | Self-supported categories | |
Totally left reductive semigroups | Normal categories | |
Regular monoids | Bounded above normal categories | |
- full transformation monoid | (or just ) - full powerset category | |
- singular transformation semigroup | - powerset category | |
- linear transformation monoid | - full subspace category | |
- singular linear transformation semigroup | - subspace category | |
- symmetric inverse monoid | - partial bijection subsets category |
2. Preliminaries
We assume familiarity with some basic ideas from category theory and semigroup theory. For undefined notions, we refer to [33, 28] for category theory and [13, 29, 24] for semigroups and biordered sets. In the sequel, all mappings, morphisms and functors shall be written in the order of their composition, i.e., from left to right.
2.1. Semigroups and categories
As mentioned earlier, all the semigroups we discuss in this paper are regular. Given a regular semigroup , we can define two quasi-orders and on as follows. For :
Then the Green relations and are equivalence relations defined on the semigroup as follows:
Given an element , we shall denote the and classes of using and , respectively. The natural partial order on a regular semigroup can be given by [40]. In the sequel, we shall denote the restrictions of the above mentioned relations to the set of idempotents of , also by the same symbols. Given an idempotent , we shall denote the downset of by .
In this paper, we shall deal with two types of categories. The first type are locally small categories whose objects are algebraic structures and morphisms are structure preserving mappings (for example, the category of left reductive semigroups). These categories will be dealt with in a standard way and we will be concerned about adjunctions and equivalences between such categories. The second type are small categories which are treated as algebraic structures by themselves (for example, a normal category). So, when comparing such small categories, we shall employ stronger notions like normal category isomorphisms. It is worth mentioning that we shall also consider locally small categories whose objects are small categories (for example, the category of connected categories).
Given any category , the class of objects of is denoted by , and the set of morphisms by itself. Hence, given two objects , the set of all morphisms from to is denoted by .
2.2. Regular semigroups and normal categories
Now, we proceed to give a quick introduction regarding the notion of normal categories and how these categories characterise the principal left ideals of a regular semigroup.
Recall that a morphism in a category is called a monomorphism if it is right cancellable and an epimorphism if it is left cancellable. A morphism in a category is said to be an isomorphism if there exists a morphism in such that and .
A preorder is a category such that there is at most one morphism between any two given objects (possibly equal). A strict preorder is a preorder in which the only isomorphisms are the identity morphisms. In a strict preorder, the relation on the class defined by:
is a partial order. Hence, a small strict preorder category is equivalent to a partially ordered set (poset) .
Definition 2.1.
Let be a small category and be a subcategory of . Then the pair (often denoted by just ) is said to be a category with subobjects if:
-
(1)
is a strict preorder with .
-
(2)
Every is a monomorphism in .
-
(3)
If and if for some , then .
In a category with subobjects, the morphisms in are called inclusions. If is an inclusion, we have , and we denote this inclusion by . An inclusion splits if there exists such that , and a morphism satisfying such an equality is called a retraction.
Definition 2.2.
Let be a category with subobjects. A morphism in is said to have a normal factorisation if , where is a retraction, is an isomorphism and is an inclusion, respectively in .
Given a normal factorisation the morphism does not depend on the factorisation, it is known as the epimorphic component of the morphism and is denoted in the sequel by . Similarly, the morphism is known as the inclusion of and denoted by . The codomain of is called the image of and shall be denoted as . Likewise, the codomain of the retraction is called the coimage of and denoted by . We collect the following results as a lemma which will be quite useful in the sequel for manipulating expressions involving morphisms.
Lemma 2.1 ([41, Corollary II.4, Proposition II.5 and II.7]).
Let be a category with normal factorisation property where inclusions split.
-
(1)
Every morphism has a unique epimorphic component i.e., is independent of the chosen normal factorisation of .
-
(2)
If is an epimorphism, then the epimorphic component .
-
(3)
If and are composable morphisms such that the inclusion of is , then
-
(4)
The inclusion of an epimorphism is the identity morphism, and so every normal factorisation of is of the form , where is a retraction and is an isomorphism.
-
(5)
Dually, the retraction of a monomorphism is the identity morphism, and so every normal factorisation of is of the form , where is an isomorphism and is an inclusion. In particular, the epimorphic component of an inclusion is the identity morphism.
Remark 2.1.
Given a morphism in a category with the normal factorisation property where inclusions split, by Lemma 2.1(1), the morphism has a unique factorisation of the form where is an epimorphism and is an inclusion. Such a factorisation is called as a canonical factorisation of the morphism .
Definition 2.3.
Let be a category with subobjects and . A mapping from to defined by for each , is said to be a cone333cones were called as normal cones in [41]. with vertex if:
-
(1)
whenever , ;
-
(2)
there exists at least one such that is an isomorphism.
Given a cone , we denote by the vertex of and the morphism is called the component of the cone at the object . The figure 1 illustrates a typical cone with vertex in a category .
Definition 2.4.
[41, Section III.1.3] A category is said to be a normal category if:
-
(NC 1)
is a category with subobjects;
-
(NC 2)
every inclusion in splits;
-
(NC 3)
every morphism in admits a normal factorisation;
-
(NC 4)
for each there exists a cone with vertex such that .
Naturally, for two normal categories to be isomorphic, we need an inclusion preserving functor which is fully-faithful such that the object map is a bijective order isomorphism.
Let be a normal category and let be a cone in , if is an epimorphism with , then as in [41, Lemma I.1], we can easily see that the map
(1) |
is a cone such that the vertex . Hence given cones and ,
(2) |
where is the epimorphic component of the morphism , defines a binary composition on the set of all cones in . This binary composition on the set of cones is illustrated in Figure 2 wherein the components of the composed cone are drawn in red colour. So, for instance, the component of the cone at the object is the morphism . Observe that the vertices but the inclusion need not always be identity morphism. In the sequel, we may often denote the binary composition of cones by juxtaposition.
Lemma 2.2 ([41, Theorem I.2]).
Let be a normal category. Then the set of all cones forms a regular semigroup under the binary composition defined in (2). A cone in is an idempotent if and only if .
The next two lemmas follow from the discussion in [41, Section III.2].
Lemma 2.3.
Let be cones in the regular semigroup . Then the quasi-orders in are characterised as follows.
-
(1)
, and so .
-
(2)
Hence, we have if and only if is an isomorphism such that
Lemma 2.4.
Let be idempotent cones in the semigroup and let be the natural partial order on the set of idempotents of . Then if and only if is a retraction such that
Now, we briefly describe how normal categories come from regular semigroups. Given a regular semigroup , we can define the category of the principal left ideals, called the left category, by
and the set of all morphisms from the object to the object is the set
where , for each .
Given any morphisms and , they are equal if and only if , and (or ) ; and they are composable if (i.e., if ) in which case
Observe that has a particular subcategory defined by and there exists in a morphism from to if and only if , this morphism being exactly . The morphisms of correspond to the inclusions of principal ideals. By definition, is therefore a strict preorder and is a category with subobjects. Given an inclusion , it has a right inverse in ; every inclusion in the category splits.
Let be an arbitrary morphism in , then as shown in [41, Corollary III.14], we can see that there exists and for ,
where is a retraction, is an isomorphism and is an inclusion. This is a normal factorisation of the morphism . Observe that the image of the morphism is uniquely determined and it is the principal left ideal , but there is a choice for the coimage . This shows that in an arbitrary regular semigroup, although the image of a morphism is unique, the coimage need not be unique.
Further, if is an arbitrary element of a regular semigroup , then for each , the mapping defined by
(3) |
is a cone with vertex , usually referred to as a principal cone in the category . Observe that, for an idempotent , we have a principal cone with vertex such that . Summarising the above discussion, it can be easily verified that is a normal category [41, Theorem III.16].
Conversely, given an abstractly defined normal category , we obtain a regular semigroup of cones in . Then the left category of the semigroup is isomorphic to [41, Theorem III.19]. We shall give an independent proof of this fact, later as a consequence of our results (see Proposition 3.14 and Proposition 8.1).
It is worth mentioning here that although for a normal category , we do not have isomorphic to for an arbitrary regular semigroup . This relationship in general is more subtle as described in Theorem 2.6 below. So, every normal category comes from a regular semigroup although not every regular semigroup can be constructed from a normal category.
Theorem 2.5 ([41, Corollary III.20]).
A small category is normal if and only if is isomorphic to a category , for some regular semigroup .
Now, we shift our focus to the subclass of regular semigroups which can indeed be constructed using just one normal category. Recall that a regular semigroup is said to be left reductive if the regular representation is injective. In the category , the cones are direct abstractions of the regular representation of a semigroup. In fact, it can be shown that the semigroup of all principal cones in is isomorphic to the image of the regular representation of the semigroup . Roughly speaking, the left ‘part’ of the regular semigroup is captured by the normal category .
Theorem 2.6 ([41, Theorem III.16]).
Let be a regular semigroup. There is a homomorphism given by . Also, is isomorphic to a subsemigroup of (via the map ) if and only if is left reductive.
Dually, we define the normal category of principal right ideals of a regular semigroup by:
(4) |
where a morphism from to is the mapping for each .
3. Left reductive regular semigroups
We proceed to give a construction for a left reductive regular semigroup as a subsemigroup of the semigroup of cones of a normal category . This is where we bifurcate ourselves from Nambooripad’s construction.
3.1. Connected categories
First, recall that given any regular semigroup , the set forms a poset under the usual set inclusion as follows:
(5) |
In fact, the poset has been characterised by Grillet as a regular poset444Since the definition involves several new notions and as we do not explicitly use any of the properties of regular posets, we omit the formal definition. in [23]. Now given a normal category , since is a regular semigroup (Lemma 2.2), the poset is a regular poset. We are now ready to give the most important definition of this paper.
Definition 3.1.
Let be a normal category and let be an order ideal of the poset . Then is said to be connected by if for every , there is some such that contains some idempotent cone with vertex . We denote such a category by and say that the regular poset connects the normal category .
Given a normal category , we define
Observe that each idempotent cone in the set may be uniquely represented as such that the vertex and where and . In this case, we shall say that the object is connected by . Hence, we have:
(6) |
Remark 3.1.
The definition of the set involves picking certain -classes from the semigroup , using the order ideal . So, we could have equivalently defined connected categories by letting be an order ideal isomorphic to an order ideal of the poset . Admittedly, this would complicate the discussion substantially and so we avoid it at this stage. However, we shall indeed use this identification in Example 3.1 below and later in Section 8, where we discuss concrete cases, as such an identification will lead to simpler descriptions of the connecting posets .
Remark 3.2.
By Definition 3.1, every is connected by at least one and conversely each connects at least one . As we shall see later, this is a reflection of the fact that every and class of a regular semigroup contains at least one idempotent. Notice that, in general, one object may be connected by multiple , and also different objects in may be connected to the same .
Before proceeding further, we shall recall some ideas from [5, 42, 43] and use them to illustrate a concrete example of a connected category. We also rectify an error in these papers by assuming the underlying set to be finite rather than arbitrary. The following Lemmas 3.2 and 3.3 could have been obtained as a Corollary of [5, Theorem 3.1], we include their proofs here as they can act as a roadmap for the reader to follow the more abstract construction. This example will be revisited, later in the Section 8.3.
Example 3.1.
(Full power set category, ) Let . Then the set of all subsets of forms a small category with mappings as morphisms, i.e.,
Also, given , the identity map is the identity morphism at the object in . Observe that is a small, full subcategory of the large (in fact, locally small) category .
Lemma 3.2.
is a normal category.
Proof.
To begin with, we can realise as a category with subobjects as follows. Let and for subsets , we define
where is the set inclusion map from to . Then is a strict preorder category, or equivalently is a partially ordered set. It is routine to verify that the pair satisfies Definition 2.1 and hence forms a category with subobjects. Observe that given an inclusion map in , we can always find a retraction map in the category such that . Observe that the retraction need not be unique, in general. But any mapping has a uniquely defined image, and so the image of the morphism in the category is the image of the mapping.
Further, given a mapping in from to , let be the image of the mapping so that is an inclusion map. Now, the map determines a partition of given by:
Let be a cross-section of the partition , and given an arbitrary , let be the equivalence class of in the set containing . Define as the surjection given by and then we have . Also, will be a bijection from to . Hence we have as illustrated in the diagram below where is a retraction, is an isomorphism and is an inclusion in the category .
Hence any morphism in has a normal factorisation. Finally, for any object , and any mapping such that , for each subset , letting , we have that is a cone in with vertex such that . Moreover, using Lemma 2.2, we can see that is an idempotent cone. Hence is a normal category. ∎
Observe that in the last part of the proof above, a cone in the category is determined by a mapping from to itself. We shall see below that in fact, this relationship is much stronger. Recall that the semigroup of all mappings from a finite set to itself, under mapping composition is known as the full transformation monoid .
Lemma 3.3.
The semigroup of cones in the category is isomorphic to the full transformation monoid .
Proof.
First, observe that the normal category has a largest object, namely . So given a cone in the category with vertex , we may define
where the mapping is an element of the semigroup , and so is well-defined. We proceed to prove that is an isomorphism. For a cone , by Definition 2.3(2) there is some such that is a bijection. However, since , the component is always a surjection and so by Lemma 2.1(2), we have . Hence the expression is in fact the unique canonical factorisation of the mapping . Also notice that by Definition 2.3(1), for each , we have .
Now, to verify that is a homomorphism, let be cones in the category with vertices and , respectively. If we denote the vertex of the cone by , we see that . Then using equations (1) and (2), and Lemma 2.1 (3), we have
Also by the definition of , since
we see that is a homomorphism.
To show that is injective, let . Then as this is the unique canonical factorisation, we have . Now since has a largest object , every cone is uniquely determined by its component , whence .
Finally, to verify that is a surjection, given an arbitrary mapping in the monoid , for each , the map is a cone with vertex such that . We conclude that is a semigroup isomorphism. ∎
To realise the category as a connected category, we need the characterisation of the Green -relation on the regular semigroup . By the above lemma, the poset is order isomorphic to . So, we recall the following well known results regarding the monoid .
Lemma 3.4 ([13, Section 2.2]).
Let be arbitrary mappings in .
-
(1)
For principal left ideals, if and only if . Hence if and only if .
-
(2)
For principal right ideals, if and only if . Hence if and only if .
We denote the poset of all partitions of the set by . Given an idempotent cone in , define a map by . Using Lemmas 3.3 and 3.4(2), we can routinely verify that is an order isomorphism. This leads to the following characterisation of the poset .
Lemma 3.5.
Let be cones in the semigroup . Then if and only if . Hence the regular poset is order isomorphic to the poset of all partitions of the set .
By the above lemma, we may identify the -classes of the semigroup by the partitions . Summarising, given a finite set , the set of subsets of forms a normal category such that the poset of -classes of the semigroup is isomorphic to the set of partitions of . This leads us to our first example of a connected category.
Proposition 3.6.
Given a finite set with powerset and partitions , the category is connected by , and so is a connected category.
Proof.
Given any subset , let be such that . Then is an idempotent in . Then as in the last part of the proof of Lemma 3.2, for each subset , define . Now, is an idempotent cone in such that and . Hence, the subset is connected by and so, the normal category is connected by the poset . ∎
Remark 3.3.
In the above example of a connected category , we have . As discussed in Remark 3.1, the relaxation that the ideal is an isomorphic copy of (rather than ) leads to a concrete characterisation of as the poset . Strictly speaking, with the terminology of Definition 3.1, Proposition 3.6 says that the category is connected by the poset such that is isomorphic to the poset .
3.2. The connection semigroup
Having digressed a bit, we now return back to the abstract construction of a left reductive semigroup from connected category . We shall see that the required semigroup is in fact , which is realised as the subsemigroup of the semigroup of cones in the category . The following lemma is crucial for the sequel.
Lemma 3.7.
Let be a connected category. Then every cone in the set can be expressed as , for some idempotent cone and an isomorphism . Conversely, every cone in which can be expressed in this form belongs to .
Proof.
First, observe that given a cone in , we have and let . Now, since connects the category , by Remark 3.2, there is some such that connects . Let the associated idempotent cone be . So we have in the semigroup . Then by Lemma 2.3(2), we get such that is an isomorphism.
Conversely, if , then by Lemma 2.3(2), we obtain and so . Thus . ∎
Remark 3.4.
Given a cone with vertex , the decomposition of as above is not unique, in general. In fact, given idempotents and in such that , we can see that for isomorphisms and . Figure 3 illustrates this situation, wherein we consider the ‘egg-box’ diagram of a typical -class of the regular semigroup . Observe that by Lemma 2.3, the -classes of are determined by the vertices of the cones.
Proposition 3.8.
Let be a connected category. Then is a regular semigroup.
Proof.
First we need to show that is a closed subset of . Let and be two cones in the set . Then by Lemma 3.7, there are an idempotent cone and an isomorphism such that . Let . Using equation (2), and Lemma 2.1(2) and (3), we see that
As is an isomorphism and is an epimorphism, their composition is an epimorphism. Since we are in a normal category, by Lemma 2.1(4), an epimorphism has a normal factorisation such that the inclusion component is the identity, thus where is retraction and is an isomorphism. Hence . Since a retraction is, in particular an epimorphism, by equation (2), we see that is a cone in with vertex . Let and in the semigroup . Now observe that
Thus is an idempotent, and with a retraction. Therefore, by Lemma 2.4, we have in the semigroup . In particular, . Since is an order ideal and , we get . So and as shown in Figure 4, we have . Next, using Lemma 3.7, we obtain and so is a subsemigroup of .
Finally, to see that is regular, let be a cone with vertex . By definition of a connected category, there is an idempotent cone in the semigroup with vertex . Using Lemma 3.7, we can write for some , and an isomorphism from to . Then let so that and . Since , by Lemma 3.7 the cone . Also, observe that and , and are both isomorphisms. Then
Similarly, , and so is an inverse of (see Figure 5). Hence is a regular subsemigroup of .
∎
The following variant of the Lemma 3.7 uses the technique applied in the proof of Proposition 3.8 and will be useful in the sequel.
Lemma 3.9.
Given a connected category , any cone in the semigroup has a representation of the form for an idempotent cone and an epimorphism . Conversely, any cone of this form belongs to .
Proof.
Since every isomorphism is an epimorphism, the first part of the lemma is obvious from Lemma 3.7. Conversely, let be a cone in of the form , where is an idempotent cone with vertex and is an epimorphism. Using Lemma 2.1(4), let be the normal factorisation of the epimorphism . Then . Then, as argued above in the proof of Proposition 3.8 (see Figure 4), we see that is an idempotent cone such that . So, we have , where is an idempotent cone and is an isomorphism. Hence, by Lemma 3.7, the cone belongs to the semigroup . ∎
Proposition 3.10.
Let be a connected category. The semigroup is left reductive.
Proof.
Recall that, to prove that is left reductive, given such that
(7) |
we need to show . Also, observe that by Definition 2.3 and Lemma 2.1(1), to prove that two cones and are equal in a category , it suffices to show that:
-
(1)
the vertices of the cones are same, i.e., , and,
-
(2)
the epimorphic components of the respective morphisms are same, i.e.,
To begin with, observe that since the semigroup is regular, there exists an idempotent cone such that , and so . Then using the assumption (7) and letting , we have . So, , i.e., . Hence, using Lemma 2.3(1), we see that the vertices of the cones, satisfy . Similarly, using an idempotent such that , we can show that . Thus .
Next, given an arbitrary , by definition there is an idempotent cone in the semigroup with vertex such that . Letting in the assumption of (7), we get , and so using equation (2) we get . Now, comparing the component of these cones at the object , we obtain for the morphisms, . However, since , we have , for each object . Thereby we conclude that the cones and coincide, and so the regular semigroup is left reductive. ∎
Summarising the above discussion, given a connected category , we constructed a left reductive regular semigroup . We shall refer to this semigroup as the connection semigroup of the category . Now, to take the discussion forward, we need to explore the left and right ideal structure of . Since is a regular subsemigroup of the semigroup , the Green relations in get inherited from (see [29, Proposition 2.4.2]). So using Lemma 2.3, we have the following:
Lemma 3.11.
Let be cones in the connection semigroup . Then
-
(1)
, and ;
-
(2)
.
Since the Green relations in and in coincide ( see Remark 3.1), we readily conclude that the poset is order isomorphic to . For later use, we denote this order isomorphism by the map .
By Lemma 3.11(1), it is clear that the poset is order isomorphic to the poset . To completely describe the left ideal structure of the semigroup , we need a deeper understanding. To this end, we employ normal categories and dive one additional layer deeper. Recall that since is a regular semigroup, the principal left ideals of form a normal category .
To simplify the notation, fix for the remaining of the section. We can define a functor as follows: given and a morphism in such that , let
(8) |
where is the inclusion morphism in . Observe that given a morphism in the category from to , since , we have . So, using Lemma 3.11(2), we see that is an epimorphism in such that . Also, by Remark 2.1, the expression is the unique canonical factorisation of the corresponding morphism belonging to .
Lemma 3.12.
is a well-defined functor from the normal category to .
Proof.
Suppose that , then by Lemma 3.11(1) we have , and is well-defined on objects. To verify that is well-defined on morphisms, suppose that in the left category . Then, from Section 2.2, this equality of morphisms implies , and . By Lemma 3.11(1), we have and . Further, since is an epimorphism, by Lemma 2.1(2), we get Hence, using equations (1) and (2), we see that
Now, since every morphism in the category has a unique canonical factorisation, we see that the morphism and so is well-defined.
To see that is a functor from to , first observe that given the identity morphism in the category , we have . Hence and the identities are preserved by .
Next, let and be morphisms in such that . Then is a morphism such that . Also, and . Thus since , and so by Definition of cone 2.3(1), we have . Moreover, as , using Remark 2.1, we can write . Therefore, the morphisms and are composable and
Finally, from and , we obtain
Thus, the assignment preserves the composition also, whence is a functor. ∎
Lemma 3.13.
The functor is a normal category isomorphism.
Proof.
By Lemma 3.11(1), the map is clearly a bijection. Given an inclusion in the category , we can easily see that is an inclusion in the category . Hence is inclusion preserving.
To see that is faithful, suppose that , i.e., in the category . Then and and so by Lemma 3.11(1), we have and . On another hand, using the canonical factorisation property of morphisms in , we get . Then applying (2), we obtain
Hence, in the category , and so is faithful.
To show that is full, given a morphism , let be its canonical factorisation. Since is a connected category, there exist idempotent cones with vertices and , respectively. Let . By Lemma 3.9, the cone is in and by Lemma 3.11, we have . So is a morphism in the normal category such that
We conclude that is a normal category isomorphism, as required. ∎
We now summarise the above discussion.
Proposition 3.14.
Given a connected category , the left ideal category is isomorphic to and the regular poset is order isomorphic to .
Remark 3.5.
Since is a regular semigroup, the principal right ideals of form a normal category as defined in equation (4). The poset of objects of the right normal category is, in fact, order isomorphic to the poset .
From the discussion above, it follows the next characterisation of the quasi-orders on the set of idempotents of .
Lemma 3.15.
Let and be idempotents in the semigroup , then:
-
(1)
-
(2)
.
In this last lemma, observe the abuse of notation as is the partial order on , however it coincides with the partial order on .
3.3. Structure of left reductive regular semigroups
In the previous section, we saw a left reductive semigroup constructed using a connected category. Now, we proceed to show how a left reductive semigroup gives rise to a connected category. To this end, recall from Theorem 2.6 that given a left reductive semigroup , we have via the map . In fact, more is true.
Lemma 3.16.
If is a left reductive semigroup, for an element , the map is a bijection onto , wherein denotes the -class of the cone in the semigroup .
Proof.
We have already seen that the map is injective, when is a left reductive semigroup. To see that the map has image , let . Thus and for some . Then since is an injective homomorphism, there exists such that in , and in the semigroup . Now, by Lemma 2.3(2), there is an isomorphism in , say such that . Then since is an isomorphism in , we have in the semigroup (see Figure 6 and [41, Proposition III.13(c)]). So, we obtain and , whence maps onto . ∎
From the discussion above and using still the notation , we see that : , where maps to , is such that, for any element , the sets and are in bijection. Hence in the sequel, for ease of notation we shall denote the -class in containing the cone by just . Observe that there might be -classes in which are not of the form for some . For instance, in the Figure 6, the bottom -class of the semigroup does not have a preimage under . As the reader may have already realised (also see Remark 3.1), the definition of a connection semigroup involves excluding from the semigroup the -classes that are not of the form . So, let
(9) |
Then, and for each , there is such that contains an idempotent cone with vertex , namely . Observe that this idempotent cone may be denoted by , i.e., the object is connected by . So, we can get a connected category from a left reductive semigroup. Hence we have proved the following proposition:
Proposition 3.17.
Let be a left reductive semigroup. The normal category is connected by the regular poset , that is, is a connected category.
Given the connected category , by Propositions 3.8 and 3.10, we know that the connection semigroup is a left reductive semigroup. Moreover, by equation (6), the idempotents of are given by:
Proposition 3.18.
Given a left reductive semigroup , the connection semigroup
is isomorphic to .
Proof.
First, recall that there is an injective homomorphism given by . Now given any , for some and , we have . Here is an idempotent cone in and is an isomorphism in . Now, using Lemma 3.7, we see that is an injective homomorphism which is also surjective (see Figure 6). Hence the connection semigroup is isomorphic to . ∎
3.4. Category equivalence
We have seen that given a connected category , we get a left reductive semigroup such that the category is isomorphic to and the order ideal is isomorphic to the regular poset (Proposition 3.14). Conversely, given a left reductive semigroup , we have obtained a connected category such that its connection semigroup is isomorphic to (Proposition 3.18). Next, we proceed to extend this correspondence to a category equivalence.
First observe that left reductive semigroups form a full subcategory, say of the category of regular semigroups, with semigroup homomorphisms as morphisms.
Definition 3.2.
Given connected categories and , we define a CC-morphism as an ordered pair such that is an inclusion preserving functor and is an order preserving map satisfying:
(10) |
for every .
Remark 3.6.
Given a CC-morphism from to , by definition the functor maps an idempotent cone in the category to the idempotent cone in the category . This makes the relation between the categories and via the functor rather strong. Roughly speaking, the semigroup homomorphism associated with the morphism will be an ‘extension’ of this mapping .
Remark 3.7.
Also, given an arbitrary pair from to such that is a normal category isomorphism (i.e., a category isomorphism that preserves inclusions) from to and is an order isomorphism from to , we can easily construct examples such that the semigroups and are not isomorphic. Hence the condition (10) is crucial in the definition of a CC-morphism, to obtain isomorphic semigroups.
It is routine to verify that the class of all connected categories with CC-morphisms form a category. This category will be denoted by in the sequel.
Recall from Proposition 3.17 that given an object in , the connected category is an object in the category . Now we proceed to make this correspondence functorial.
Lemma 3.19.
Given a semigroup homomorphism in the category , define a functor and a map as follows: for idempotents and ,
Then is a CC-morphism from to .
Proof.
It is a routine matter to verify that is an inclusion preserving functor from to and is an order preserving map from to (see equation (9)). To verify (10), let the object be connected to in , we are taking to be an idempotent in such that is an idempotent in . Let . As is a homomorphism, is an idempotent in . Hence in the category , we have connected to , where denotes the -class of the cone in the semigroup . Further, for every , observe that . Since is a homomorphism we have so that
Also, in . Hence
Therefore satisfies equation (10) and so is a CC-morphism. ∎
Further, it is routine to verify that this assignment preserves identities and compositions. Hence, we have the following proposition.
Proposition 3.20.
The assignment
constitutes a functor, say from the category of left reductive semigroups to the category of connected categories.
To build a functor in the opposite direction, we have seen that given a connected category in , by Propositions 3.8 and 3.10, the semigroup . Now, given a CC-morphism in , we need to construct a semigroup homomorphism. To this end, recall that given a connected category , for each , there is an associated idempotent cone in such that . By Lemma 3.7, every cone in may be written as , for an idempotent cone and an isomorphism in . Let be a morphism in from to . Define by:
(11) |
Lemma 3.21.
is a well-defined map from semigroup to .
Proof.
First, observe that by Lemma 3.7, any cone in admits a representation, not necessarily unique, of the form . Then will be an idempotent cone in . Note that given an inclusion preserving functor preserves normal factorisations (see [41, Proof of Lemma V.4] for a routine verification). So will be an isomorphism and using Lemma 3.7, we see that will be a cone in . Now let be a cone in with vertex and . As in Remark 3.4, suppose where and are idempotent cones, and and are isomorphisms. Let , , , , , and . We need to show that , whence .
Lemma 3.22.
is a semigroup homomorphism.
Proof.
Using Lemma 3.7, let and be two arbitrary cones in such that and are isomorphisms. We need to show that .
To this end, given a morphism in the category , we shall denote its image in under by dashed versions without further comment, i.e., we shall denote by . Then since , we have
Now, using Lemma 2.1(3), we see that
So if we let and , we obtain
Hence by definition of , we get
Then as is a functor, using equation (10), we reach
Also, as preserves normal factorisations, using Lemma 2.1(3) and equation (10), we conclude that
Therefore putting everything together and using equation (2),
as required. ∎
After having constructed a semigroup homomorphism from a CC-morphism, we may now routinely verify the following assertion.
Proposition 3.23.
The assignment
constitutes a functor, say from the category of connected categories to the category of left reductive semigroups.
Now, we have all the ingredients to prove the category equivalence of and .
Lemma 3.24.
The identity functor is naturally isomorphic to the functor .
Proof.
We need to illustrate a natural transformation between the functors and such that each of its components is a semigroup isomorphism. By Proposition 3.18 we know that the map is an isomorphism from to . So, for each object , we denote this map by . Then since
we see that the map is a semigroup isomorphism from to . Further, given a semigroup homomorphism , we can routinely verify that the following diagram commutes:
Hence the assignment constitutes a natural isomorphism between the functors and . ∎
Lemma 3.25.
The identity functor is naturally isomorphic to the functor .
Proof.
From Proposition 3.14, given a connected category , the left ideal category is a normal category isomorphic to the category and the regular poset is order isomorphic to , via the functor and the map respectively. Further, we can routinely (but admittedly a bit cumbersome in terms of notation) verify that satisfies condition (10), and both and are bijections; hence is a CC-isomorphism. So, for each object , if we denote this morphism by , then since
we see that is a CC-isomorphism from to . Further, given a CC-morphism , we can routinely verify that the following diagram in the category commutes:
As a meticulous reader may see, the condition (10) is again crucial in this verification. Hence the assignment constitutes a natural isomorphism between the functors and . ∎
Theorem 3.26.
The category of left reductive semigroups is equivalent to the category of connected categories.
4. -unipotent semigroups
The remaining of the paper is essentially dedicated to applications of the results in Section 3 to various classes of left reductive semigroups. In this section, we specialise the construction in the previous section to give an abstract construction of -unipotent semigroups using supported normal categories. -unipotent semigroups were introduced and studied initially by Venkatesan [47, 48, 49, 50] under the name of right inverse semigroups. Over the years, various facets of this class of semigroups (and of its dual, the class of -unipotent semigroups) have been studied by many people including the second author [11, 31, 27, 17, 16, 46, 9, 10, 26, 19, 21, 20]. We begin by recalling some basic properties of -unipotent semigroups.
As usual by an -unipotent semigroup we mean a regular semigroup in which each -class contains a unique idempotent.
Proposition 4.1 ([48, Theorem 1][16, Corollary 1.2, 1.3]).
Let be a regular semigroup. The following are equivalent:
-
(1)
is an -unipotent semigroup;
-
(2)
, for all ;
-
(3)
, for all , i.e., is a right regular band;
-
(4)
, for all and ;
-
(5)
each -class contains a unique idempotent , for all and ;
-
(6)
, for all and .
-
(7)
, for all , and .
By [48, Theorem 4(1)], it is known that an -unipotent semigroup is left reductive, i.e., the regular representation is injective. Since this is a cornerstone of this section, we record this formally amongst some other useful results regarding -unipotent semigroups and provide a more transparent proof for the left-reductivity. In the process, we also characterise the natural partial order on an -unipotent semigroup.
Recall that a regular semigroup whose idempotents form a band is said to be orthodox, and this is the case of any -unipotent semigroup. In any orthodox semigroup , for all , and , the elements and are idempotents (cf. [29, Proposition 6.2.2]).
On another hand, in any regular semigroup , the natural partial order [40] is given by, for all ,
When is -unipotent, a possible form taken by is, for all ,
Proposition 4.2.
Let be an -unipotent semigroup. For every pair of distinct elements in , there exists an idempotent in such that . In particular, the semigroup is left reductive.
Proof.
We prove by contradiction. Suppose for every idempotent . Since is an idempotent, we have ; similarly we get . This implies and respectively, whence and we get a contradiction. This concludes the first part of the lemma.
Given in , we have an idempotent in such that . Hence, the map is injective, whence is left reductive. ∎
Next, given an -unipotent semigroup , we shift our attention to the normal category . The following lemma is a simple consequence of Proposition 4.1(2).
Lemma 4.3.
In the category , given morphisms and we have if and only if , and . Also, if is an inclusion in , the corresponding unique retraction is .
Recall also that in any regular semigroup , the -classes of form a regular poset under usual set inclusion. If is, in addition -unipotent, then by Proposition 4.1(3), we have , and so the regular poset becomes a semilattice where the meet operation is given by set intersection:
The condition (3) in the Proposition 4.1 may make one wonder if a regular semigroup such that its poset is a semilattice is always an -unipotent semigroup (also see equation (5)). This need not be the case as the following simple example shows. Consider a three element semigroup given by the following -class picture (on the left):
This semigroup is clearly not -unipotent (as and are distinct -related idempotents) although the regular poset forms a three element semilattice (given on the right side). Observe that .
Notice that this semilattice which appears in the context of the left regular bands has been referred to as the support semilattice of the semigroup [34, 12]. We shall be discussing the case of right and left regular bands in Sections 5 and 6, respectively. With this terminology in mind, we proceed to give the construction of an -unipotent semigroup by introducing supported categories, as specialisations of connected categories.
Definition 4.1.
Let be a normal category and let be a sub-semilattice of the poset . Then is said to be supported by if each is connected by a unique .
Remark 4.1.
By the definition above, there is a well-defined mapping . This is a surjection and will be referred to as the support map in the sequel. Observe that the support map need not be injective, in general, but we shall later see that is always order preserving (see Proposition 4.5). Also note that in contrast to the support map of [34], our map is not a homomorphism as there is no semigroup structure in the set .
Remark 4.2.
The assumption that is a semilattice is not necessary in Definition 4.1. One can show that the regular poset will necessarily form a semilattice, once we establish the following proposition.
Proposition 4.4.
Let be a supported category. Then the connection semigroup is -unipotent.
Proof.
Given a supported category , it is connected and so by Proposition 3.8, we know that is a regular semigroup. Now, using Lemmas 3.11 and 3.15, given a cone with vertex , we can see that any idempotent in the -class of is of the form , for some connecting . However, is a supported category, there is a unique with this property. Hence each -class of contains a unique idempotent and by Proposition 4.1(2), the semigroup is -unipotent. ∎
We now point out to the following specialisations of Lemmas 3.11 and 3.15, in the case of supported categories.
Proposition 4.5.
Given a supported category , let and be idempotents in the semigroup , then:
-
(1)
, and there is a bijection between the sets and ;
-
(2)
.
In particular, the support map defined by is an order preserving surjection.
Proof.
(1) follows from the fact that there is a unique idempotent in the semigroup with a given vertex. To prove (2), first observe that is a semilattice and in , we have if and only if . Also, by Proposition 3.14, we know that is order isomorphic to . So,
To show the last part of this proposition, first observe that the map is well-defined by (1) and by definition, it is a surjection. Now if , then by Lemma 3.15(1), we have . Using biorder properties [39, Definition 1.1(B21)], we have in the -unipotent semigroup . As each -class in contains a unique idempotent, we have , i.e., . So by Lemma 3.15(2), we obtain and therefore is order preserving. ∎
We have seen above how a supported category gives rise to an -unipotent semigroup. The proposition below shows the converse, i.e. every -unipotent semigroup determines a supported category.
Proposition 4.6.
Let be an -unipotent semigroup . Then is a supported category. The support map is given by .
Proof.
Given an -unipotent semigroup , by Proposition 4.2, it is left reductive. As in Section 3.3, we can show that the category is normal and the semigroup is regular. Now, we define the order ideal as:
We know that is order isomorphic to via , and in addition it is a a meet semilattice when is -unipotent. Hence is a meet semilattice with
Also, since each -class in contains a unique idempotent, each object in is connected by a unique . Further, in an -unipotent semigroup, as if and only if , we have:
Hence the support map is an order preserving surjection from to . ∎
Specialising Proposition 3.18, we get:
Proposition 4.7.
Given an -unipotent semigroup , the connection semigroup is isomorphic to .
Further, the discussion in Section 3.4 carries over verbatim to the -unipotent case. It is clear that -unipotent semigroups form a full subcategory of , say and supported categories form a full subcategory of , say . We shall refer to the morphisms in the subcategory as SC-morphisms, in the sequel.
Theorem 4.8.
The category of -unipotent semigroups is equivalent to the category of supported categories.
5. Right regular bands
Now we further specialise the construction in Section 4 to describe right regular bands and, in this case, we obtain an adjunction. We show that right regular bands form a full coreflective subcategory of the category . To guide the readers to this end, we recall the following definitions which shall also be needed later in this section.
Definition 5.1.
Let and be two arbitrary categories. An adjunction is a triple , where and are functors, and is a natural transformation such that the following condition holds:
-
•
For every pair of objects and , and for every morphism in the category , there exists a unique morphism in the category such that the following diagram commutes:
In this case, and are called left and right adjoints , respectively, and is the unit of adjunction.
Definition 5.2.
A coreflective subcategory is a full subcategory of a category whose inclusion functor has a right adjoint. We shall say that a category is coreflective in if is equivalent to a coreflective subcategory of .
We refer the reader to [18, Proposition 1.3] for several equivalent characterisations of the (dual of) second definition above. Now, it is routine to verify the following lemma.
Lemma 5.1.
The category of right regular bands is a coreflective subcategory of the category of -unipotent semigroups.
Here the coreflector functor maps an -unipotent semigroup to its right regular band of idempotents. Combining Lemma 5.1 and Theorem 4.8, one can construct the adjunction between the categories and , but we shall briefly exposit this adjunction in a direct manner. Putting together Propositions 4.4 and 4.1(4), we are led to the next lemma.
Lemma 5.2.
Let be a supported category. Then the set of all idempotent cones in the connection semigroup
forms a right regular band.
Notice that by Proposition 4.5(1), the band is in bijection with the set . Further, given an SC-morphism from to , as shown in Lemmas 3.21 and 3.22 we can prove that given by:
is a semigroup homomorphism of right regular bands. Hence we obtain a functor as follows:
Conversely starting from a right regular band , since it is also -unipotent, we can easily see that (as shown in Section 4) the category constitutes a supported category. This correspondence is given by the functor . The functor is precisely the restriction of the functor defined in the Section 3.4 to the category . Further, we will show that the functor is a left adjoint to the functor .
Theorem 5.3.
There is an adjunction from the category of right regular bands to the category of supported categories. In particular, the category is coreflective in the category .
Proof.
To begin with, observe that given a right regular band , it is left reductive and by Proposition 3.18, the connection semigroup is isomorphic to via the map . Moreover, . Hence the assignment is a natural isomorphism from the functor to the functor .
Now, given an object , by Proposition 4.6, we can see that is a supported category. Let so that . Given a semigroup homomorphism in , by Lemma 3.19 and Proposition 3.14, we see that is the unique SC-morphism from to , where . Next, we may routinely verify that the following diagram commutes:
Hence constitutes an adjunction from the category to . The last part of the theorem follows directly from the fact that the left adjoint is fully-faithful (see [18, Proposition 1.3]). To conclude one can check that is equivalent to the category , and the latter is a coreflective subcategory of . ∎
6. Right reductive regular semigroups
In this section, we describe some straightforward applications of our results from the previous sections to their dual classes of semigroups. We begin with the class of right reductive semigroups and also consider their subclasses of -unipotent semigroups and of left regular bands.
In Section 2 (see equation (4)), we introduced the normal category of principal right ideals of a regular semigroup . Observe that given an arbitrary element , the principal cone is the cone in with vertex given by, for each
Recall from [13, Section 1.3] that the anti-regular representation of is the anti-homomorphism given by , and is said to be right reductive if is injective. We may then show the following dual statement of Theorem 2.6.
Theorem 6.1.
Let be a regular semigroup. There is an anti-homomorphism given by and the semigroup is isomorphic to a subsemigroup of if and only if is right reductive.
From Propositions 3.8 and 3.10, given a connected category , the connection semigroup is left reductive. Thus, the opposite semigroup where, for any cones ,
(12) |
is right reductive and regular. In the sequel, we shall refer to this semigroup as the dual connection semigroup.
Now by [41, Remark III.6], we may see that for the opposite semigroup of a regular semigroup , we get
(13) |
Moreover, the following relationships between its regular posets hold:
(14) |
Therefore, and using Proposition 3.14, we get isomorphic to . Hence we may state the next proposition.
Proposition 6.2.
Let be a connected category. The semigroup is a right reductive semigroup. The right category is a normal category isomorphic to the category and the regular poset is isomorphic to .
At this point recall that by Theorem 6.1, given a right reductive semigroup , we have and so, as in Section 3.3, we may isolate the -classes in of the form . Observe that is the -class in containing the principal cone , for . now, define
Since , we have and can verify the result below.
Proposition 6.3.
Let be a right reductive semigroup. The normal category is connected by the regular poset , and so is a connected category. Moreover, the dual connection semigroup is isomorphic to the semigroup .
The next step is to extend this to a category equivalence. We denote the category of right reductive semigroups by . Given a semigroup homomorphism in , as in Lemma 3.19, we obtain a CC-morphism from to . Notice that remains a covariant functor. Hence, we obtain a functor as follows:
Conversely, given a connected category , the semigroup is a right reductive, and as in Proposition 3.23, we find a functor . Imitating the proofs of Lemmas 3.24 and 3.25, we conclude the required equivalence.
Theorem 6.4.
The category of right reductive semigroups is equivalent to the category of connected categories.
Naturally, specialising our discussion on right reductive semigroups to -unipotent semigroups and left regular bands, we may emulating Sections 4 and 5.
Theorem 6.5.
The category of -unipotent semigroups is equivalent to the category of supported categories. The category of left regular bands is coreflective in the category of supported categories.
7. Inverse semigroups
In this section, we look at a class of regular semigroups, which are both left and right reductive, namely inverse semigroups. Inverse semigroups arguably form the most important class of regular semigroups, mainly due to their ability to capture partial symmetry [32]. Inverse semigroups are in fact -unipotent semigroups, which are left-right symmetrical.
In the joint work [8], the first author described a category equivalence between inverse semigroups and inversive categories. That construction used Nambooripad’s normal categories and admittedly, the description did not reflect the symmetrical nature of inverse semigroups. In contrast, when we employ supported categories, the symmetry of the semigroups gets manifested by the categories ‘supporting’ themselves; so we dub these self-supported categories.
Before continuing we recall some characterisations of inverse semigroups:
Proposition 7.1 ([24, Theorem II.2.6]).
Let be a regular semigroup. The following are equivalent:
-
(1)
is an inverse semigroup;
-
(2)
every element in has a unique inverse element;
-
(3)
is a semilattice;
-
(4)
there is a unique idempotent in each -class and each -class of .
As mentioned inverse semigroups form one of the most ‘symmetrical’ classes of semigroups. This symmetry is a reflection of the uniqueness of the inverse, which in turn, defines a natural involution on the semigroup given by . This leads to the proposition below.
Proposition 7.2.
Let be an inverse semigroup. Then the left category is normal category isomorphic to the right category . In particular, the semilattice is order isomorphic to .
Proof.
Define a functor by, for any inverse semigroups ,
(15) |
Since inverse semigroups have a unique idempotent in every -class and in every - class, it is easy to see that the map is a well-defined bijection. Now, by Proposition 7.1(3), the quasi-orders and on the idempotents of an inverse semigroups coincide with the natural partial order , and so
Hence is an order isomorphism between the semilattices and . Also, observe that is order isomorphic to the set of idempotents of .
Given such that is a morphism in from to , we can see that so that is a morphism in from to . Then using Lemma 4.3 and Proposition 7.1(2), we verify that the map is well defined. Now given two composable morphism and in the category such that , we know that and
On the other hand,
Hence is a covariant functor from to . Applying Lemma 4.3, one sees that is inclusion preserving and fully-faithful. Therefore, is a normal category isomorphism. ∎
From Proposition 4.6, the support map of an -unipotent semigroup is given by . In the case of inverse semigroups, the next corollary reflects the left-right symmetry of these semigroups.
Corollary 7.3.
Let be an inverse semigroup. The support map is an order isomorphism.
Proof.
The map in the previous proposition may be interpreted as an order isomorphism from to given by . As discussed in Section 3.3, we know that the map is an order isomorphism from the poset to . Hence is order isomorphic to via the map . ∎
Remark 7.1.
When is an inverse semigroup, we see that is order isomorphic to . In other words, the normal category is supported by a partially ordered set which is isomorphic to . Or by abuse of terminology, we can just say that the normal category is supported by , i.e., is self-supported.
Definition 7.1.
A supported category is said to be self-supported if the support map is an order isomorphism.
Remark 7.2.
In the above definition, we do not need to explicitly specify the supporting semilattice as we know that is the semilattice .
Proposition 7.4.
Given a self-supported category with , the connection semigroup is inverse.
Proof.
Since a self-supported category is supported, by Proposition 4.4 the connection semigroup is -unipotent. Now let and be -related idempotents of . Then the objects and are both connected by . As the support map is injective, we get . Hence and each -class in contains a unique idempotent. By Proposition 7.1(2), the semigroup is inverse. ∎
Proposition 7.5.
Let be an inverse semigroup. Then the category is self-supported and such that is order isomorphic to .
Finally, Theorem 4.8 allows us to obtain the following equivalence theorem for inverse semigroups.
Theorem 7.6.
The category of inverse semigroups is equivalent to the category of self-supported categories.
8. Totally left reductive semigroups and regular monoids
In this section we aim to study regular monoids which also happen to be both left and right reductive. To this end, we identify another special class of left reductive semigroups, which we call totally left reductive. In the process, we prove a couple of interesting results regarding the semigroup of all cones from an arbitrary normal category (Proposition 8.1 and Corollary 8.3). We introduce morphisms between arbitrary normal categories and thereby define the category of normal categories. The discussion in this section tells us that if a regular semigroup is totally left reductive (in particular, if it is a monoid), the cross-connection analysis is expendable and connected categories suffice to describe such semigroups completely.
8.1. Totally left reductive semigroups
Recall that, given a regular semigroup , we have the homomorphism , , and by Theorem 2.6, this is injective when is left reductive. The question of the surjectivity of leads to the following definition.
Definition 8.1.
A regular semigroup is said to be totally left reductive if is isomorphic to the semigroup of cones in the category .
It is not difficult to see that regular monoids are totally left reductive semigroups (see Proposition 8.5), but this class contains several non-monoids, too. It has been showed in [5, Theorem 3.1] and in [2, Theorem 2], respectively, that singular transformation semigroups and singular linear transformation semigroups are totally left reductive. The semigroups of order preserving mappings on finite chains [45] and the Clifford inverse semigroups [3] also fall into this class. All the above mentioned papers were written within the framework of Nambooripad’s cross-connection theory. Our next aim is to employ connected categories to look at these classes and as the reader may see, our analysis will take us to easier and stronger characterisations of each of these classes.
From Section 3, and by definition, a connected category connects a normal category with an order ideal . In particular, taking , we obtain . Observe that here we are treating an arbitrary normal category as a connected category. For the remaining part of the paper, we shall treat normal categories in this manner without further comment. By Proposition 3.10, we know that the semigroup of all cones in a normal category , is indeed left reductive. Then applying Proposition 3.14, we come to the next result, which we believe to be of more general interest.
Proposition 8.1.
The semigroup of all cones in a normal category is left reductive. Moreover, the left category is normal and isomorphic to .
The latter half of above proposition was already known and one can find an alternate proof in [41, Section III.3.3] where the functor is defined from to .
The following lemma is of a routine verification.
Lemma 8.2.
Let and be isomorphic normal categories. Then the regular semigroup is isomorphic to the semigroup .
Further, combining Proposition 8.1 and Lemma 8.2, we conclude that the semigroup is isomorphic to the semigroup of cones in the left category . Hence, we get the next statement.
Corollary 8.3.
Let be a normal category. The semigroup of all cones in is a totally left reductive semigroup.
Summarising the above discussion: A normal category can be realised as a connected category such that the connection semigroup is totally left reductive, and the left category is isomorphic to . Conversely, given a totally left reductive semigroup , the category is normal, and the semigroup is isomorphic to . Thus, the regular poset is isomorphic to , whence the normal category may be realised as the connected category and the connection semigroup is isomorphic to .
Now, normal categories (with CC-morphisms) form a full subcategory of the category , and totally left reductive semigroups (with regular semigroup homomorphisms) form a full subcategory of the category . Therefore, specialising Theorem 3.26, we obtain:
Theorem 8.4.
The category of normal categories is equivalent to the category of totally left reductive semigroups.
8.2. Regular monoids
Now we look at the most important class of totally left reductive semigroups, namely that of regular monoids. Although the construction we present in this special case, does not seem very insightful regarding the ideal structure of the monoids (as a monoid itself forms a two sided ideal, and consequently, the left and the right actions on this ideal determines the entire actions), our analysis does provide a category equivalence of regular monoids with small categories which we believe it may prove to be quite useful.
To begin with, let be a regular monoid with identity . The left ideal category of has a largest object, namely . Hence we have the corollary to Theorem 2.6 below.
Proposition 8.5 ([41, Corollary III.17]).
Let be a regular monoid. The normal category has a largest object, and is isomorphic to the semigroup of all cones in .
The above proposition leads us to identify certain special normal categories.
Definition 8.2.
A normal category is said to be bounded above if there exists an object such that , for every .
Lemma 8.6.
Let be a normal category which is bounded above. Then the semigroup of all cones in is a regular monoid.
Proof.
By Lemma 2.2, we know that is a regular semigroup. Since is bounded above, it has a largest object, say . Let be an idempotent cone in with vertex . Now, for , we have , and given an arbitrary cone , by Definition 2.3(1) we get . In particular, . Observe that will always be an epimorphism. Then using equation (2), we see that
Also for an arbitrary ,
Hence , for any . Thus the semigroup is a regular monoid with identity . ∎
Since regular monoids form a full subcategory of the category of totally left reductive semigroups, and bounded above categories form a full subcategory of normal categories, specialising Theorem 8.4, we get:
Theorem 8.7.
The category of regular monoids is equivalent to the category of bounded above normal categories (with CC-morphisms).
8.3. Transformation semigroups
The full transformation monoid is regular and it contains the symmetric group as its subgroup of units. From Example 3.1, the full powerset category of all subsets of is connected and normal with largest object , hence is bounded above. We denote simply by .
In the light of Theorem 8.7, the next proposition leads to a full description of the regular monoids , in terms of categories.
Proposition 8.8.
The normal category is isomorphic to the full powerset category , as bounded above categories.
Proof.
From Lemma 3.4(1) the left ideals in the monoid are determined by images. Hence, given and , define a functor as:
It may be routinely verified that is a normal category isomorphism from to .
Notice that the largest object in is and so, we can define a map as . Since is isomorphic to and using Lemma 3.4(2), we may verify that is an order isomorphism. Further, we may prove that the pair also satisfies the condition (10), and hence the category is isomorphic to as bounded above categories. ∎
The above proposition tells that the full transformation monoid is equivalent to the bounded above category . Hence, any question regarding the monoid may be translated to an equivalent question regarding the connected category . In particular, we can obtain the exact same descriptions of the biordered set and the sandwich sets of in terms of subsets and partitions (see [1, Section 5.1]) using the connected category description rather than cross-connections. Observe that the cross-connection description of previously known, [5], involved the rather cumbersome category of partitions[43], but our approach bypasses this by just using the poset .
Having settled the full monoid case , we move onto one of its most important regular subsemigroups, namely that of singular transformations . See [14] for a good overview regarding this semigroup. The ideal structure of was studied in detail inside the cross-connection theory in [5, 42, 43]. Naturally, our next objective is to use our theory of connected categories to realise the semigroup as a normal category.
It is easy to see that the set of proper subsets of the set forms a small category with mappings as morphisms. We shall refer to as the powerset category. Observe that is a full subcategory of the full powerset category (see Example 3.1).
At this point, we call upon some known results.
Lemma 8.9 ([5, Theorem 3.1][42, Theorem 3]).
Let be the singular transformation semigroup. The category is normal and isomorphic to the left category . The semigroup of all cones in the category is isomorphic to , and so is a totally left reductive.
Recall also that, since is a regular subsemigroup of , the Green relations in get inherited from . Using Lemma 3.4 and the fact that is totally left reductive, we see that the poset of right ideals of the semigroup may be characterised using non-identity partitions of .
Next, observe that the non-identity partitions of the set form an order ideal, say , of the poset . Also, given an element , we have a principal cone in the normal category . This discussion allows us to verify that the map gives an order isomorphism from the poset of -classes in the semigroup to the poset . In fact, emulating the proof of Proposition 8.8, we get a description of the semigroup of singular transformations as a normal category.
Proposition 8.10.
The normal category of the singular transformation semigroup is isomorphic to the powerset category .
8.4. Linear transformation semigroups
Continuing our list of applications, we move onto a brief discussion on the linear transformation semigroups, which are quite analogous to the transformation semigroups. Given a finite dimensional vector space over a field , the linear transformations on form a regular monoid . The group of invertible linear transformations on forms the subgroup of units of . In [2], the cross-connections of , its singular part and the variants of , were discussed. We refer the readers to [2] for a detailed discussion on the normal categories involved. It may be worth mentioning that the right ideal structure of was described using the annihilator category in [2]. However, employing our approach of connected categories, we can describe the right structure of with just the null spaces of .
Given a finite dimensional vector space , the subspaces of form a small category with linear transformations as morphisms. This category has a largest object and the Green relations and in are determined by subspaces and null spaces, respectively [13, Section 2.2]. Let denote the poset of null spaces of under reverse inclusion. Imitating the proofs in the case of transformation semigroups, we obtain the next results.
Proposition 8.11.
Let be a finite dimensional vector space over a field . The category is normal and bounded above. In particular, the poset of -classes in the semigroup is isomophic to the poset , and so is a connected category. The normal category is isomorphic to , as bounded above categories.
Next, let be the singular linear transformation semigroup and let be the category of proper subspaces of . Clearly, is a full subcategory of . Using [2, Theorem 2] and the discussion on the semigroup of singular transformations, we can also deal with the singular linear case.
Proposition 8.12.
Let be the singular linear transformation semigroup. The category is normal and isomorphic to , as normal categories. Moreover, the semigroup is isomorphic to the semigroup .
8.5. Symmetric inverse monoids
We conclude the paper with a quick discussion on arguably the most important inverse semigroup: the monoid of all the partial bijections on an arbitrary set , denoted . We shall realise as a self-supported category which is also bounded above.
To begin with, recall that the Green relations and in the semigroup are determined by the images and domains of the partial mappings, respectively [29, Exercise 5.11.2]. Let be the category of all subsets of with partial bijections as morphisms. Clearly, is normal and the set is the largest object in , wence is bounded above. Now, define a functor as follows: for idempotents and ,
We may check that is a normal category isomorphism. Hence the semigroup of cones in is isomorphic to the semigroup . Moreover, since the -classes of the semigroup are determined by the domains, the poset of -classes of the semigroup is order isomorphic to . Hence the category is self-supported too. At last, we can verify that the isomorphism may be extended to an isomorphism of self-supported categories. We collect the above discussion in the final proposition, which describes how the symmetric inverse monoid may be realised as a category, just as in the ESN Theorem.
Proposition 8.13.
Let be a set. The category is normal and bounded above. The semigroup of cones in is isomorphic to the semigroup . The normal category and the category are isomorphic as self-supported categories.
References
- [1] P. A. Azeef Muhammed. Cross-connections and variants of the full transformation semigroup. Acta Sci. Math. (Szeged), 84(3-4):377–399, 2018.
- [2] P. A. Azeef Muhammed. Cross-connections of linear transformation semigroups. Semigroup Forum, 97(3):457–470, 2018.
- [3] P. A. Azeef Muhammed and C. S. Preenu. Cross-connections in Clifford semigroups. In Semigroups, algebras and operator theory, volume 436 of Springer Proc. Math. Stat., pages 3–9. Springer, Singapore, [2023] ©2023.
- [4] P. A. Azeef Muhammed and A. R. Rajan. Cross-connections of completely simple semigroups. Asian-European J. Math., 09(03):1650053, 2016.
- [5] P. A. Azeef Muhammed and A. R. Rajan. Cross-connections of the singular transformation semigroup. J. Algebra Appl., 17(3):1850047, 22, 2018.
- [6] P. A. Azeef Muhammed and M. V. Volkov. Inductive groupoids and cross-connections of regular semigroups. Acta Math. Hungar., 157(1):80–120, 2019.
- [7] P. A. Azeef Muhammed and M. V. Volkov. A tale of two categories: inductive groupoids and cross-connections. J. Pure Appl. Algebra, 226(7):Paper No. 106940, 40, 2022.
- [8] P. A. Azeef Muhammed, M. V. Volkov, and K. Auinger. Cross-connection structure of locally inverse semigroups. Internat. J. Algebra Comput., 33(1):123–159, 2023.
- [9] B. Billhardt, Paula Marques-Smith, and Paula Mendes Martins. Extensions of left regular bands by -unipotent semigroups. Period. Math. Hungar., 83(2):260–271, 2021.
- [10] M. J. J. Branco and Gracinda M. S. Gomes. The generalized Szendrei expansion of an -unipotent semigroup. Acta Math. Hungar., 123(1-2):11–26, 2009.
- [11] M. J. J. Branco, Gracinda M. S. Gomes, and Victoria Gould. Extensions and covers for semigroups whose idempotents form a left regular band. Semigroup Forum, 81(1):51–70, 2010.
- [12] K. S. Brown. Semigroups, rings, and Markov chains. J. Theoret. Probab., 13(3):871–938, 2000.
- [13] A. H. Clifford and G. B. Preston. The Algebraic Theory of Semigroups, Volume 1. Number 7 in Mathematical Surveys. American Mathematical Society, Providence, Rhode Island, 1961.
- [14] J. East. A presentation for the singular part of the full transformation semigroup. Semigroup Forum, 81(2):357–379, 2010.
- [15] J. East and P. A. Azeef Muhammed. A groupoid approach to regular -semigroups. Adv. Math., 437:Paper No. 109447, 104, 2024.
- [16] Constance C. Edwards. The minimum group congruence on an -unipotent semigroup. Semigroup Forum, 18(1):9–14, 1979.
- [17] Constance C. Edwards. The structure of -unipotent semigroups. Semigroup Forum, 18(3):189–199, 1979.
- [18] P. Gabriel and M. Zisman. Calculus of fractions and homotopy theory, volume Band 35 of Ergebnisse der Mathematik und ihrer Grenzgebiete [Results in Mathematics and Related Areas]. Springer-Verlag New York, Inc., New York, 1967.
- [19] Gracinda M. S. Gomes. -unipotent congruences on regular semigroups. Semigroup Forum, 31(3):265–280, 1985.
- [20] Gracinda M. S. Gomes. The lattice of -unipotent congruences on a regular semigroup. Semigroup Forum, 33(2):159–173, 1986.
- [21] Gracinda M. S. Gomes. Some results on congruences on -unipotent semigroups. Semigroup Forum, 33(2):175–185, 1986.
- [22] J. A. Green. On the structure of semigroups. Ann. of Math. (2), 54:163–172, 1951.
- [23] P. A. Grillet. Structure of regular semigroups: 1. A representation, 2. Cross-connections, 3. The reduced case. Semigroup Forum, 8:177–183, 254–259, 260–265, 1974.
- [24] P. A. Grillet. Semigroups: An Introduction to the Structure Theory. CRC Pure and Applied Mathematics. Taylor & Francis, 1995.
- [25] T. E. Hall. On regular semigroups. J. Algebra, 24(1):1–24, 1973.
- [26] A. Hayes. The structure of 0-bisimple -unipotent semigroups. Semigroup Forum, 65(3):405–427, 2002.
- [27] A. Hayes. The structure of -unipotent semigroups with zero. Comm. Algebra, 32(11):4465–4484, 2004.
- [28] P. J. Higgins. Notes on categories and groupoids. Van Nostrand Reinhold, 1971.
- [29] J. M. Howie. Fundamentals of semigroup theory, volume 12 of London Mathematical Society Monographs. New Series. The Clarendon Press, Oxford University Press, New York, 1995. Oxford Science Publications.
- [30] J. M. Howie. Semigroups, past, present and future. In Proceedings of the International Conference on Algebra and its Applications (ICAA 2002) (Bangkok), pages 6–20. Chulalongkorn Univ., Bangkok, 2002.
- [31] Y. Ji. -unipotent semigroup algebras. Comm. Algebra, 46(2):740–755, 2018.
- [32] M. V. Lawson. Inverse Semigroups: The Theory of Partial Symmetries. World Scientific Pub. Co. Inc., 1998.
- [33] S. MacLane. Categories for the Working Mathematician, volume 5 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1971.
- [34] S. Margolis, F. V. Saliola, and B. Steinberg. Cell complexes, poset topology and the representation theory of algebras arising in algebraic combinatorics and discrete geometry. Mem. Amer. Math. Soc., 274(1345):xi+135, 2021.
- [35] W. D. Munn. Fundamental inverse semigroups. Q. J. Math., 21(2):157–170, 1970.
- [36] K. S. S. Nambooripad. Structure of regular semigroups. I. Fundamental regular semigroups. Semigroup Forum, 9:354–363, 1975.
- [37] K. S. S. Nambooripad. Structure of regular semigroups. II. The general case. Semigroup Forum, 9:364–371, 1975.
- [38] K. S. S. Nambooripad. Relations between cross-connections and biordered sets. Semigroup Forum, 16:67–82, 1978.
- [39] K. S. S. Nambooripad. Structure of regular semigroups. I. Mem. Amer. Math. Soc., 22(224):vii+119, 1979.
- [40] K. S. S. Nambooripad. The natural partial order on a regular semigroup. Proc. Edinburgh Math. Soc. (2), 23(3):249–260, 1980.
- [41] K. S. S. Nambooripad. Theory of Cross-connections. Publication No. 28. Centre for Mathematical Sciences, Thiruvananthapuram, 1994.
- [42] A. R. Rajan and P. A. Azeef Muhammed. Normal categories from the transformation semigroup. Bull. Allahabad Math. Soc., 30(1):61–74, 2015.
- [43] A. R. Rajan and P. A. Azeef Muhammed. Normal category of partitions of a set. Southeast Asian Bull. Math., 41(5):735–746, 2017.
- [44] A. R. Rajan, C. S. Preenu, and K. S. Zeenath. Cross connections for normal bands and partial order on morphisms. Semigroup Forum, 104(1):125–147, 2022.
- [45] K. K. Sneha and P. G. Romeo. Normal category arising from the semigroup . Palest. J. Math., 13(2):285–292, 2024.
- [46] Mária B. Szendrei. -unitary -unipotent semigroups. Semigroup Forum, 32(1):87–96, 1985.
- [47] P. S. Venkatesan. Bisimple left inverse semigroups. Semigroup Forum, 4:34–45, 1972.
- [48] P. S. Venkatesan. Right (left) inverse semigroups. J. Algebra, 31:209–217, 1974.
- [49] P. S. Venkatesan. On right unipotent semigroups. Pacific J. Math., 63(2):555–561, 1976.
- [50] P. S. Venkatesan. On right unipotent semigroups. II. Glasgow Math. J., 19(1):63–68, 1978.
- [51] J. von Neumann. On regular rings. Proc. Natl. Acad. Sci. USA, 12(22):707–713, 1936.