Quick viewing(Text Mode)

Arxiv:2008.06852V2 [Math.RT] 7 Mar 2021 Rs Rjciemdls Ata Functions

Arxiv:2008.06852V2 [Math.RT] 7 Mar 2021 Rs Rjciemdls Ata Functions

arXiv:2008.06852v2 [math.RT] 7 Mar 2021 rs rjciemdls ata functions. Partial modules, Projective bras, Keywords: 16G10 20M30, Classification. Subject hemn eirusWoeCtgre r IadTheir and EI are Categories Whose Ehresmann ∗∗ ∗ ahmtc nt hmo olg fEgneig 74 As 77245 Engineering, of College Shamoon Unit, Mathematics eateto ahmtc,BrIa nvriy 20 Rama 52900 University, Ilan Bar Mathematics, of Department santrleape ecnie h the consider we example, rel Green’s natural generalized a using As indecompo but way the similar characteristic a good in described with fields over that show ftemxmlsbrusof subgroups maximal the of eiruswoectgre r Ii suoait.W pr We pseudovariety. algebra finite a of is collection the EI the of are that categories show whose We semigroups isomorphism. an EI- is an phism is Ehresmann corresponding whose semigroup tde eeaiaino nes eirus Let semigroups. inverse of generalization studied rjciemdlsadoti oml o hi descr matrix. their natural Cartan for a its formula give in a entries we obtain numbers, and complex modules of projective field the Over . esuysml n rjciemdlso eti ls fE of class certain a of modules projective and simple study We hemn eirus Ictgre,Rgtrsrcins restriction Right EI-categories, semigroups, Ehresmann ersnainTer xeddVersion Extended : Theory Representation k S S

oe n field any (over i h orsodn cüznegrmdl.Mroe,we Moreover, module. Schützenberger corresponding the via

[email protected] [email protected] tatMargolis Stuart tmrStein Itamar Abstract k 1 r omdb nuigtesml modules simple the inducing by formed are ) PT S eafiiergt(et etito Ehresmann restriction (left) right finite a be n falprilfntoso an on functions partial all of ∗∗ dd Israel hdod, a,Israel Gan, t ∗ tosisedo h tnadones. standard the of instead ations oevr efidcranzero certain find we Moreover, . aeoy hti,eeyendomor- every is, that category, al rjciemdlscnbe can modules projective sable v httesml modules simple the that ove rsansmgop,awell- a semigroups, hresmann pino t indecomposable its of iption ih etito Ehresmann restriction right mgop,Smgopalge- Semigroup emigroups, n -element 1 Introduction

A semigroup S is called inverse if every element a ∈ S has a unique inverse, that is, an element b ∈ S such that aba = a and bab = b. Inverse semigroups are one of the central objects of study in semigroup theory (see [20]) and in particular their representation theory is well-studied (see [40, Part IV]). There are several generalizations of inverse semigroups that keep some of their properties and structure. In this paper we discuss representations of a generalization called Ehresmann semigroups. Let E be a subsemilattice of a semigroup S (that is, a commutative subsemigroup of idempotents).

Define two equivalence relations LE and RE in the following way: Elements a,b ∈ S satisfy aLEb

(aRE b) if a and b have the same set of right (respectively, left) identities from E. We also define e e e HE = RE ∩ LE . Assume that every LE and RE class contains precisely one idempotent from E e denoted a∗ and a+ respectively. The semigroup S is called Ehresmann (or E-Ehresmann if the set e e e e e E is to be emphasized) if LE is a right congruence and RE is a left congruence. An Ehresmann semigroup is called right restriction if the ea = a(ea)∗ holds for every e ∈ E and a ∈ S. The e e main class of semigroups we consider in this paper are finite right restriction Ehresmann semigroups with the additional property that the HE -class of every e ∈ E is a . We call such semigroups right restriction EI-Ehresmann because the corresponding Ehresmann category is an EI-category e (that is, every is an isomorphism). This is a generalization of a class known in the literature as (finite) weakly ample semigroups (see [9, Section 4]). Natural examples of such semigroups are the monoid PT n of all partial functions on the set {1,...,n} and the monoid of full domain partitions on a finite set introduced in [6]. In Section 3 we prove that this class is a pseudovariety in signature (2, 1, 1). We also discuss cases when such semigroups are embeddable in

PT n as a bi-unary semigroup. In Section 4 we show that the simple modules of the semigroup algebra kS (over any field k) of a finite right restriction EI-Ehresmann semigroup are given by induced Schützenberger modules of the simple modules of the maximal subgroups of S. Unlike inverse semigroups, Ehresmann semigroups algebras need not be semisimple, even over fields of “good” characteristic. Therefore, an Ehresmann semigroup can have non-semisimple projective modules. Another goal of this paper is to describe the indecomposable projective modules of finite right restriction EI-Ehresmann semigroups. For this we need a construction that replaces Green’s L-classes by LE-classes as the basis of induction. Let S be a finite semigroup and let E ⊆ S be a subset of idempotents. Choose an e ∈ E and let e LE(e) be its LE-class. In Section 5 we characterize certain cases where Ge (the maximal subgroup with unit element e) acts on the right of LE(e). In particular we prove that Ge acts on the right e e of LE(e) for every Ehresmann semigroup. If we denote by kLE(e) the k- with basis e LE(e), this implies that kLE(e) V is a kS-module for every kGe-module V . e kG e Oe Ae very successful way to studye algebras of inverse semigroups or their generalizations is to relate

2 them to algebras of an associated category (or a “partial semigroup”). This is often done with an appropriate Mï¿œbius as in [11, 12, 17, 31, 33, 38, 39, 42]. In particular, the second author has proved in [34, 35] the following theorem. Let S be a finite right restriction Ehresmann semigroup. Then the semigroup algebra kS (over any field k) is isomorphic to the category algebra kC for the associated Ehresmann category C. This result has led to several applications in the representation theory of of partial functions [36, 37]. Now assume again that S is a finite right restriction EI-Ehresmann semigroup and also assume that the order of every subgroup of S is invertible in k. In Section 6 we use the isomorphism between the semigroup algebra and the corresponding category algebra mentioned above to prove that any module of the form kLE (e) V (where V is a kG Oe simple kGe-module) is an indecomposable projective module of kS. Moreover,e any indecomposable projective module is of this form. This gives a purely semigroup theoretic construction and it is very similar to other known constructions in the representation theory of semigroups [8, 26]. We also give a formula for the dimension of this module over an algebraically closed field. As an application, we consider the monoid PT n in the case k = C. It was already proved that the simple and indecomposable injective modules of C PT n can be described using induced and co-induced representations respectively (essentially this is a consequence of [24, Theorem 4.4]). In Section 7 we complete this picture by giving a description of the indecomposable projective modules of C PT n along with the natural epimorphism onto their simple images. The general formula for the dimension of the indecomposable projective modules boils down in this case to a combinatorial formula that sums up certain Kostka numbers. We also draw some conclusions regarding the Cartan matrix of

C PT n by showing that certain entries of it are zero. Finally, we give a counter-example to the above isomorphism between the algebras of an Ehresmann semigroup and the associated category in the case where S is not right (or left) restriction hence showing that this requirement cannot be omitted. We also give an example of an EI-Ehresmann semigroup that is neither left nor right restriction but its algebra is isomorphic to the algebra of the corresponding Ehresmann category.

2 Preliminaries

2.1 Semigroups

Let S be a semigroup and let S1 = S ∪{1} be the monoid formed by adjoining a formal unit element. We denote by H, L, R and J the usual Green’s equivalence relations:

a R b ⇐⇒ aS1 = bS1 a L b ⇐⇒ S1a = S1b a J b ⇐⇒ S1aS1 = S1bS1

3 and H = R ∩ L. In a finite semigroup, it is known that J = R ◦ L = L ◦ R, which is Green’s relation D. We denote by E(S) the set of all idempotents of S. It is well-known that if e ∈ E(S) then its H-class forms a group that we will denote by Ge. This is the maximal subgroup of S with unit element e. An element a ∈ S is called regular if there exists b ∈ S such that aba = a. Now assume that S is finite. A J -class J is called regular if it contains a regular element. It is known that in this case all its elements are regular and every R and every L class contains an idempotent.

Moreover, if J is a regular J -class then for every two idempotents e1,e2 ∈ J we have Ge1 ≃ Ge2 .

So we can denote this group by GJ - the unique maximal group associated with J. Let J be a regular J -class of S. The corresponding regular factor is a 0-simple finite semigroup and thus by 0 Rees Theorem is isomorphic to a regular Rees matrix semigroup M (Ge, A, B, P ). See [14, Chapter

3] and the Appendix of [28] for details. We denote by Bn the monoid of all on the set {1,...,n} with composition as operation and by PT n its submonoid of all partial functions.

The submonoid of all total functions is denoted T n and ISn will be our notation for the submonoid consisting of all injective partial functions. We remark that we compose binary relations (and in particular partial functions) from right to left in this paper. A standard textbook for elementary semigroup theory is [14]. For a text devoted to finite semigroups see [28].

2.2 Ehresmann semigroups and categories

Let S be a semigroup and let E ⊆ S be a subset of idempotents. We define two preorders ≤ e and LE ≤ e on S by RE a ≤ e b ⇐⇒ (∀e ∈ E be = b ⇒ ae = a) LE

a ≤ e b ⇐⇒ (∀e ∈ E eb = b ⇒ ea = a). RE

The corresponding equivalence relations are denoted LE and RE.

aLEb ⇐⇒ (∀e ∈ E bee= b ⇔ aee = a)

aReEb ⇐⇒ (∀e ∈ E eb = b ⇔ ea = a).

Moreover, we define HE = LE e∩ RE . We remark that L ⊆ LE, R ⊆ RE and H ⊆ HE. A subset E ⊆ S of idempotents is called a subsemilattice if it is a commutative subsemigroup. It is well e e e e e e known that any commutative semigroup of idempotents has the structure of a semilattice (i.e. a poset where every two elements have a meet) if one defines a ≤ b whenever ab = ba = a. A semigroup S with a subsemilattice E ⊆ S is called right E-Ehresmann if every LE-class contains a unique idempotent from E and LE is a right congruence. We denote the unique idempotent in ∗ ∗ e the LE -class of a by a . Note that a is the unique minimal element e ∈ E such that ae = a. It is e ∗ ∗ ∗ well known that LE is a right congruence if and only if the identity (ab) = (a b) holds for every e e 4 a,b ∈ S.

Dually, we can consider semigroups for which every RE class contains a unique idempotent. We + denote the unique idempotent in the RE class of a by a . Such semigroup is called left E-Ehresmann + e + + if RE is a left congruence, or equivalently if (ab) = (ab ) for every a,b ∈ S. A semigroup S e with a subsemilattice E ⊆ S is called E-Ehresmann if it is both left and right E-Ehresmann. The e semilattice E is also called the set of projections of S. If S is an E-Ehresmann semigroup we can ∗ + define two partial orders by a ≤l b (a ≤r b) if and only if a = ba (respectively, a = a b) which is also equivalent to a = be for some e ∈ E (respectively, a = eb for some e ∈ E). It is known that E-Ehresmann semigroups form a variety of (2, 1, 1)-algebras (where + and ∗ are the unary operations). See [10] for a list of the identities. We use them without reference in this paper. Therefore, we can drop the explicit mention of the set E and consider an Ehresmann semigroup S to be a bi-unary semigroup where the set of projections E is determined implicitly by the unary operations: E = {a∗ | a ∈ S} = {a+ | a ∈ S}

A right (left) Ehresmann semigroup S is called right (respectively, left) restriction if the “right ample” (respectively, “left ample”) identity ea = a(ea)∗ (respectively, ae = (ae)+a) holds for every a ∈ S and e ∈ E. For every Ehresmann semigroup S, we can associate a category C(S) = C in the following way. The object set of C(S) is the set E of projections and the of C(S) are in one-to-one correspondence with elements of S. For every a ∈ S, we associate a C(a) ∈ C1 with domain a+ and range a∗. If the range of C(a) is the domain of C(b) the composition C(a) · C(b) is defined to be C(ab). The convention is to compose morphisms in Ehresmann categories “from left to right”. However, for us it will be more convenient to use composition “from right to left”. In this case, we will associate to every a ∈ S a morphism C(a) ∈C1 with domain a∗ and range a+ and if the range of C(a) is the domain of C(b) the composition C(b) · C(a) is defined to be C(ba). For more facts and proofs on Ehresmann semigroups and Ehresmann categories the reader is referred to [9, 10].

2.3 Algebras and representations

Let A be a k-algebra where k is a field. Unless stated otherwise, we assume that algebras are unital and finite dimensional. Likewise, when we say that M is a module over A (or an A-module) we mean that M is a finite dimensional left module over A. We will denote the set of simple or irreducible modules of A by Irr A. Basic facts about the representation theory of finite dimensional algebras are found in [2], which we will use without further mention. We will be interested in semigroup algebras and category algebras. The semigroup algebra kS of a finite semigroup S is defined in the

5 following way. It is a vector space over k with basis the elements of S, that is, it consists of all formal linear combinations:

{k1s1 + ... + knsn | ki ∈ k, si ∈ S}.

The multiplication in kS is the linear extension of the semigroup multiplication. The category algebra kC of a finite category C is defined in the following way. It is a vector space over k with basis the morphisms of C, that is, it consists of all formal linear combinations:

1 {k1m1 + ... + knmn | ki ∈ k, mi ∈C }.

The multiplication in kC is the linear extension of the following:

m′m if m′ · m is defined m′ · m = 0 otherwise.   2.4 Representations of finite semigroups and groups

The main reference for the representation theory of finite monoids is [40]. Let S be a finite semi- group, let J be regular J -class and fix an idempotent e ∈ J. Let k be a field. The semigroup S acts by partial functions on L(e), the L-class of e, according to

sx sx ∈ L(e) s · x = undefined otherwise  for s ∈ S and x ∈ L(e). Moreover, the group Ge ≃ GJ acts on L(e) by right multiplication.

Therefore, k L(e) is a kS − kGJ bi-module (where k L(e) is the set of all formal linear combinations of elements of L(e)). Dually, k R(e) is a kGJ − kS bi-module. Let V be a kGJ -module, we define the induced module corresponding to J and V as follows:

IndGJ (V )= k L(e) V kG OJ It can be shown that this module does not depend on the specific choice of e ∈ J. This justifies Irr the notation IndGJ (V ). If V ∈ kGJ then the maximal semisimple image of IndGJ (V ) is a simple kS-module. Moreover, the Clifford-Munn-Ponizovskii theorem states that this induces a one-to-one correspondence between simple modules of kS and pairs (J, V ) where J is regular J class and

V ∈ Irr GJ . Therefore, semisimple images of induced modules of the above form give a complete list of all simple modules of kS up to isomorphism. More details on Clifford-Munn-Ponizovskii theory

6 can be found in [8] and [40, Chapter 5]. We assume familiarity with basic results on the representation theory of finite groups [4]. For the representation theory of the symmetric group and its connection to Young Tableaux, we refer to [16, 29]. We use these classical results without further mention. We say that a Young tableau is semi-standard if its columns are increasing and its rows are non- decreasing. The Kostka number Kλµ is the number of semistandard Young tableaux with shape λ and content µ (see [29, Section 2.11]).  For every composition λ = [λ1,...,λr] k we can associate the subgroup Sλ = Sλ1 ×···× Sλr of

Sk, called the Young subgroup corresponding to λ. Let µ = [µ1,...,µr]  k and denote by trµ the Sk trivial module of the algebra of the Young subgroup Sµ. The module IndSµ trµ is called the Young module corresponding to µ. For every partition λ ⊢ k the multiplicity of the Specht module Sλ in

Sk the decomposition of IndSµ trµ is precisely Kλµ - the Kostka number of λ, µ.

3 Some classes of Ehresmann semigroups

By an endomorphism of a category C we mean a morphism f whose domain and range are the same object. For an object c of C we call C(c,c), the endomorphism monoid at c.

Definition 3.1. A category is called an EI-category if every endomorphism is an isomorphism, that is, the endomorphism monoids are groups.

There is a vast literature about representations of EI-categories and their applications (for some examples see [21, 23, 41, 43]).

Definition 3.2. Let S be an Ehresmann semigroup. We call S an EI-Ehresmann semigroup if the associated Ehresmann category C(S) is an EI-category.

Note that a ∈ HE (e) if and only if C(a) is an endomorphism of e in the associated Ehresmann category C(S). Therefore S is EI-Ehresmann if and only if HE (e) is a group for every projection e e ∈ E. e

3.1 EI-Ehresmann semigroups

We start with some characterizations of EI-Ehresmann semigroups. Recall that the sandwich set of two idempotents f1,f2 ∈ S is defined by

S(f1,f2)= {h ∈ E(S) | f2hf1 = h, f1hf2 = f1f2}.

7 Let V (s) be the set of inverses of an element s ∈ S. It is well known that the sandwich set is non-empty if and only if f1f2 is a regular element of S. Furthermore,

S(f1,f2)= {h ∈ V (f1f2) ∩ E(S) | f2hf1 = h}.

In particular, if f1 and f2 commute, then S(f1,f2) is non-empty since in this case, f1f2 is an idempotent. See [14, Chapter 2.5] for details.

Lemma 3.3. Let S be a finite Ehresmann semigroup, The following are equivalent.

1. S is an EI-Ehresmann semigroup.

2. Every idempotent f ∈ E(S) satisfies f ∗ff + = f ∗f +.

3. The projection f ∗f + is an inverse of f for every idempotent f ∈ E(S).

4. Every idempotent f ∈ E(S) satisfies f ∈ S(f ∗,f +).

Proof. (1 =⇒ 2) Observe that f ∗ff + is an idempotent since

f ∗ff + · f ∗ff + = f ∗ff ∗f +ff + = f ∗fff + = f ∗ff +.

Moreover,

+ + + + f ∗ff + = f ∗ ff + = f ∗ ff + = f ∗ (ff)+ = f ∗f +      where the first and third equalities follow from the left congruence identity (ab)+ = (ab+)+. Likewise

∗ ∗ f ∗ff + = (f ∗f)∗ f + = (f ∗f)∗ f + = (ff)∗ f + = f ∗f +   where the first and third equalities follow from the right congruence identity (ab)∗ = ∗ ∗ ∗ + ∗ + ∗ + (a b) . Therefore f ff is HE-equivalent to f f . Since HE(f f ) has a unique idempotent we obtain that f ∗ff + = f ∗f +. e e + ∗ (2 =⇒ 1) Let e ∈ E and let f ∈ E(S) be an idempotent such that fHEe so f = f = e. The assumption f ∗ff + = f ∗f + now reduces to e efe = ee = e

and clearly efe = f +ff ∗ = f so f = e as required.

8 (2 ⇐⇒ 3 ⇐⇒ 4) Immediate from the properties of sandwich sets noted before this Lemma.

We can interpret the last result in the language of universal algebra, namely (pseudo)varieties and quasivarieties. We refer to [1, 3] for the basics of varieties, quasivarieties and pseudovarieties and other concepts from Universal Algebra. If x is an element of a free profinite semigroup, then xω is the unique idempotent in the closed subsemigroup generated by x. In particular, if S is a finite semigroup, then xω is the unique idempotent in the subsemigroup generated by x. The equivalence 1 ⇐⇒ 2 of Lemma 3.3 immediately implies the following corollary in the language of pseudovarieties.

Corollary 3.4. The class of all finite EI-Ehresmann semigroups is a pseudovariety of finite bi- unary semigroups defined by the identities of Ehresmann semigroups (see [10]) and the profinite identity (xω)∗ xω (xω)+ = (xω)∗ (xω)+ .

Proof. Let S is a finite Ehresmann semigroup. If S is EI-Ehresmann then the required profinite identity holds by Lemma 3.3 because xω is an idempotent for every x ∈ S . In the other direction, take f ∈ E(S). Then f ω = f so the assumption yields

f ∗ff + = f ∗f + which implies that S is EI-Ehresmann by Lemma 3.3.

Let V be a pseudovariety of finite groups and let V be the pseudovariety of finite semigroups whose subgroups belong to V . Note that if S is a finite EI-Ehresmann semigroups then every HE(e)- class is a group and in particular, HE(e) = H(e) for every e ∈ E. Therefore the pseudovariety of e all finite Ehresmann semigroups whose HE-classes (or equivalently, the endomorphism groups in e the category C(S)) belongs to V is precisely the intersection between V and the pseudovariety of e Corollary 3.4.

Example 3.5. Consider the pseudovariety of all Ehresmann semigroups such that the correspond- ing Ehresmann category C(S) is locally trivial (i.e, the only are the identity func- tions). Some natural monoids which belong to this pseudovariety are discussed in [37]. In this case the HE-classes are trivial so the subgroups of S are trivial. Such semigroups are called aperiodic (or H-trivial). This pseudovariety can be described by the Ehresmann identities, and the profinite e identities (xω)∗ xω (xω)+ = (xω)∗ (xω)+ , xωx = xω where the second one is for the aperiodicity of S.

9 3.2 Right restriction EI-Ehresmann semigroups

Now we turn to the main class of semigroups that we want to discuss in this paper: Semigroups that are both EI-Ehresmann and right restriction. We start with the following lemma.

Lemma 3.6. Let S be a right restriction Ehresmann semigroup. Then f + ≤ f ∗ for every idempo- tent f ∈ E(S).

Proof. 1Let f ∈ E(S) be an idempotent. According to the right congruence identity

f ∗f + = (f ∗f +)+ = (f ∗f)+ and according to the right ample identity

(f ∗f)+ = (f(f ∗f)∗)+ .

Now, by the left congruence identity

(f(f ∗f)∗)+ . = (f(ff)∗)+ = (ff ∗)+ = f + so f ∗f + = f + and thus f + ≤ f ∗ as required.

Lemma 3.7. Let S be a finite right restriction Ehresmann semigroup. The following are equivalent.

1. S is an EI-Ehresmann semigroup.

2. Every idempotent f ∈ E(S) satisfies ff + = f +.

3. Every regular element a ∈ S satisfies a R a+.

Proof. (1 =⇒ 2) If S is EI-Ehresmann then f ∗ff + = f ∗f + by Lemma 3.3. Lemma 3.6 implies that f ∗f = f ∗f +f = f +f = f

so we establish ff + = f + as required.

(2 =⇒ 3) If a is regular then a R f for some idempotent f ∈ E(S) and clearly f + = a+. Now ff + = f + implies f R f + hence a R a+ as required.

(3 =⇒ 1) Every idempotent f ∈ E(S) is regular, so fRf + which says that fx = f + for some x ∈ S. Now ff + = ffx = fx = f +

1We thank Professor Victoria Gould for this proof.

10 so f ∗ff + = f ∗f + hence S is EI-Ehresmann by Lemma 3.3.

Remark 3.8. A sandwich set S(f1,f2), if not empty, is a rectangular band ([14, Proposition 2.5.3]), + + i.e., a subsemigroup satisfying ghg = g for every g,h ∈ S(f1,f2). The property ff = f implies that S(f ∗,f +) is a right zero semigroup (a rectangular band with one R-class). Indeed if h ∈ S(f ∗,f +) then fh = ff +hf ∗ = f +hf ∗ = h and hence f = fhf = hf so every element of S(f ∗,f +) is R-equivalent to f.

Example 3.9. The following example, taken from [34, Example 5.12], shows that not every EI- Ehresmann semigroup is right or left restriction, so right restriction EI-Ehresmann semigroups form a proper subclass of EI-Ehresmann semigroups.

Recall that T n is the monoid of all total functions on an n-element set. Denote by id the identity function and by k the constant functions that sends every element to k. Define S to be the op subsemigroup of T 2 × T 2 containing the six elements

(1, 1), (2, 1), (1, 2), (2, 2), (1, id), (id, 1)

It can be checked that S is an E-Ehresmann semigroup with

E = {(1, 1), (1, id), (id, 1)} as its set of projections. The corresponding Ehresmann category C is given by the following drawing (recall that composition is “right to left”):

(1, id) (id, 1) (2, 2)

(1, 2) (2, 1)

(1, 1)

Clearly, C is an EI-category (in fact, it is even locally trivial) so S is an EI-Ehresmann semigroup.

11 However, it is easy to see that (2, 2)+ = (1, id) and (2, 2)∗ = (id, 1) but (2, 2) is not L-related to (1, id) and not R-related to (id, 1) so S is neither left nor right restriction. Note also that the sandwich set S((id, 1), (1, id)) = {(1, 1), (2, 1), (1, 2), (2, 2)} is a rectangular band but not a right\left zero semigroup.

As in the case of general EI-Ehresmann semigroups, we have an immediate corollary of Lemma 3.7.

Corollary 3.10. The class of all finite right restriction EI-Ehresmann semigroups is a pseudovari- ety of finite bi-unary semigroups defined by the identities of Ehresmann semigroups, the right ample identity b∗a = a(b∗a)∗ and the profinite identity

xω (xω)+ = (xω)+ .

The most natural example of a right restriction EI-Ehresmann semigroup is the monoid PT n of all partial functions on the set {1,...,n}. In Section 7 we will consider at length its representation theory. Let A ⊆ {1,...,n} and denote by idA the partial identity function on A. It is clear that

E = {idA | A ⊆ {1,...,n}} is a subsemilattice of PT n and it is well known that PT n is a right restriction E-Ehresmann semigroup. Let t : A → B be a . We denote by dom(t) ⊆ A + ∗ and im(t) ⊆ B the domain and image of t and it is clear that t =1im(t) and t =1dom(t). It is also clear that an idempotent f ∈ E(PT n) must satisfy f|im(f) =1im(f) so the only idempotent f ∈ PT n with dom(f)= im(f) is f =1dom(f). Therefore, PT n is indeed an EI-Ehresmann semigroup. Since finite right restriction EI-Ehresmann semigroups form a pseudovariety, it is clear that any (2, 1, 1)- subalgebra of PT n is also a right restriction EI-Ehresmann semigroup. We now show that at least for regular semigroups the converse also holds.

Proposition 3.11. Let S be a regular finite right restriction EI-Ehresmann semigroup. There is an embedding of bi-unary semigroups Φ: S → PT n for n = |S|.

Proof. The “Cayley theorem” for right restriction semigroups (see [9, Theorem 6.2]) says that there ∗ ∗ is a semigroup monomorphism Φ: S → PT n such that Φ(a )=Φ(a) for every a ∈ S. It remains to show that Φ(a+)=Φ(a)+ as well. First note that

∗ ∗ Φ(a+)=Φ a+ =Φ a+     + + so Φ(a ) is a partial identity of PT n. Lemma 3.7 and the fact that S is regular imply that a R a + + for every a ∈ S. Therefore Φ(a) R Φ(a ) in PT n. Since Φ(a) is the only partial identity which is + + R equivalent to Φ(a) in PT n we must conclude that Φ(a )=Φ(a) as required.

12 On the other hand, not every right restriction EI-Ehresmann semigroup is embeddable in PT n as a bi-unary semigroup. It is known from [30, 15] that a right restriction Ehresmann semigroup embeds + + in PT n as a bi-unary semigroup if and only if it satisfies the quasiidentity xz = yz =⇒ xz = yz . We now give an example of an Ehresmann right restriction EI-semigroup that does not satisfy this quasiidentity.

2 Example 3.12. Recall that Bn is the monoid of all binary relations on an n-element set. Denote by idA the partial identity relation on the set A. Choosing the partial identities as projections, it + ∗ is well known that Bn is Ehresmann with α =1im(a) and a =1dom(a) for every α ∈Bn. Let S be the subsemigroup of B3 consisting of the following 5 elements:

e = id{1,2}, f = id{3}, a = {(3, 1), (3, 2)}, g = {(1, 2), (2, 1)} and the empty relation 0. It easy to check that S is a bi-unary subsemigroup of B3 and hence E-Ehresmann for E = {e, f, 0}. It is a routine to verify that S is also right restriction (in fact, it is also left restriction) and it follows from Lemma 3.7 that it is EI-Ehresmann since every idempotent is a projection. On the other hand, we have ea = a = ga but ea+ = ee = e 6= g = ge = ga+ so the quasi-identity xz = yz =⇒ xz+ = yz+ does not hold in S.

3.3 EI-restriction semigroups

If a semigroup S is both EI-Ehresmann and restriction (i.e., left and right restriction), the situation reduces to a familiar one. A restriction semigroup S with the property that E = E(S) is called weakly ample (see [9, Section 4]).

Lemma 3.13. Let S be a finite restriction semigroup. Then S is EI-Ehresmann if and only if S is weakly ample.

Proof. Clearly E = E(S) implies that e is the unique idempotent of HE(e) for every e ∈ E hence

HE(e) is a group. In the other direction Lemma 3.6 and its dual implies that for every idempotent e f ∈ E(S) we have f + ≤ f ∗ and f ∗ ≤ f + hence f + = f ∗. This says that every idempotent f is e HE-equivalent to a projection e ∈ E and being EI-Ehremsann implies f = e. e 3.4 The infinite case

In this paper we focus on finite semigroups. However all the results in Section 3 can be adapted to infinite semigroups as well. A monoid is called unipotent if the identity is its only idempotent and a category is called EU if all its endomorphisms monoids are unipotent. We say that an Ehresmann

2We thank Professor Michael Kinyon for this example, obtained using Prover9[27].

13 semigroup S is EU-Ehresmann if its associated Ehresmann category is an EU-category (i.e., HE(e) is a unipotent monoid for every e ∈ E). Clearly a finite monoid is unipotent if and only if it is a e group and a finite category is EU if and only if it is EI. The arguments in the previous subsections prove the following statements (where now the semigroups might be infinite):

Proposition 3.14. The class of all EU-Ehresmann semigroups is a quasivariety of bi-unary semi- groups defined by the identities of Ehresmann semigroups and the quasiidentity

f 2 = f → f ∗ff + = f ∗f +.

.

Proposition 3.15. The class of all right restriction EU-Ehresmann semigroups is a quasivariety of bi-unary semigroups defined by the identities of Ehresmann semigroups, the right ample identity b∗a = a(b∗a)∗ and the quasiidentity f 2 = f → ff + = f +.

We denote by PT X the monoid of all partial functions on the (possible infinite) set X.

Proposition 3.16. Let S be a regular right restriction EU-Ehresmann semigroup. There is an embedding of bi-unary semigroups Φ: S → PT S .

Lemma 3.17. Let S be a restriction semigroup. Then S is EU-Ehresmann if and only if S is weakly ample.

4 Simple modules of finite right restriction EI-Ehresmann semigroups

From now on the focus will be on the case of finite right restriction EI-Ehresmann semigroups. In this section we would like to describe the simple modules of algebras of such semigroups and in certain cases also the indecomposable injective ones. We remark that certain observations on the semisimple image of such semigroup algebras and their ordinary quiver can be found in [34, Proposition 5.17] and [25, Section 6.3].

Lemma 4.1. Let S be a finite right restriction EI-Ehresmann semigroup and let e ∈ E. Then e is the only idempotent in L(e).

Proof. Assume f ∈ L(e) is an idempotent so ef = e and fe = f. Lemma 3.7 implies that ff + = f + and clearly f +f = f. Now,

14 e = ef = ef +f = f +ef = f +e f + = ff + = fef + = ff +e = f +e so e = f +. Therefore f L e and f R f + = e so f H e which implies f = e.

Proposition 4.2. Let S be a finite right restriction EI-semigroup. Let J be a regular J -class of

S. Then the sandwich matrix of J is left invertible over kGJ for every field k.

Proof. Let EJ = E ∩ J = {e1,...,ek} be the set of projections which belongs to J. An R-class cannot have two projections and a regular R-class has at least one by Lemma 3.7 so J has precisely k

R-classes. Moreover, e1,...,ek are also in different L-classes. Fix e = e1 and choose representatives

ρ1,...,ρk for the H-classes of L(e) and λ1,...,λl for the H-classes of R(e) (note that l ≥ k). Since projections of EJ are in different L and R-classes, we can order the representatives such that ei ∈ R(ρi) ∩ L(λi). Lemma 4.1 implies that there is no idempotent in R(ρi) ∩ L(λj ) for i 6= j and

1 ≤ i, j ≤ k since ej is the unique idempotent in L(λj ) (note that there can be idempotents in

R(ρi) ∩ L(λj ) for j > k). Therefore, the sandwich matrix is a l × k matrix whose upper k × k block is a diagonal matrix and all the entries on the main diagonal are non-zero. Therefore, the sandwich matrix is left invertible over kGJ .

If the sandwich matrix of a regular J -class J is left invertible then the module induced from a left

Schützenberger module of J and V ∈ Irr kGJ is simple, see [40, Corollary 4.22] and [40, Lemma 5.20]. This fact and the Clifford-Munn-Ponizovskii theorem yields the following corollary.

Theorem 4.3. Let S be a finite right restriction EI-Ehresmann semigroup and let k be a field. Let

J be a regular J -class and let V ∈ Irr kGJ . Then

IndGJ (V )= k L(e) V kG OJ (for some idempotent e ∈ J) is a simple module of kS. In fact,

Irr {IndGJ (V ) | J is a regular J -class, V ∈ kGJ } is a list of all the simple modules of kS up to isomorphism with no isomorphic copies.

Remark 4.4. Assume in addition that S is regular and that the orders of all subgroups of S are invertible in k. Then the left invertibility of the sandwich matrices implies that the indecomposable injective modules of S are given by the co-induced modules for regular J classes and V ∈ Irr GJ

15 (see [24, Theorem 4.4]). The co-induced module is define by

CoindGJ (V ) = HomkGJ (k R(e), V )

(for some idempotent e ∈ J). In this case CoindGJ (V ) is the injective envelope of IndGJ (V ). In addition, it follows that in this case the global dimension of the algebra of S is bounded by one less than the longest chain of J -classes of S. The above results on indecomposable injective modules are true even if we replace the requirement of regularity by being left Fountain, see [26, Theorem 4.8].

5 Construction of LE - modules

Our next goal is to obtain a descriptione of indecomposable projective modules of finite right re- striction EI-Ehresmann semigroups. The description involves an induction using the LE relation instead of L. In this intermediate section we consider several cases where such a construction yields e a well defined kS-module structure. Let S be a fixed semigroup and let E ⊆ S be a subset of idempotents.

Lemma 5.1. Let a ∈ S and denote by LE(a) the LE -class of a. Then S acts by partial functions on LE(a) according to e e sx sx ∈ LE(a) e s · x = undefined otherwise  e for s ∈ S and x ∈ LE (a). 

Proof. The set L =e {x ∈ S | x ≤ e a} is a left ideal of S. For if x ∈ L and s ∈ S then 1 LE 1

ae = a =⇒ xe = x =⇒ sxe = sx for every e ∈ E, so sx ∈ L as well. The set L = {x ∈ S | x  e a} is also a left ideal of S. 1 2 LE Indeed, assume x ∈ L2 but sxLE a for some s ∈ S. Then

exe = x =⇒ sxe = sx =⇒ ae = a

for every e ∈ E, so xLE a, a contradiction. Therefore, S acts by left multiplication on L1 and L2.

It is now clear that the required action is the Rees quotient L1/L2. e

Now, choose a projection e ∈ E. Recall that we denote the group H-class of e by Ge. In general, right multiplication does not induce a right action of the group Ge on LE(e). For instance consider

e 16 the following example.

Example 5.2. Let S = PT 2 and recall that we compose from right to left. Choose

E = {id = id{1,2}, id{1}, id∅} and e = id is the identity function. We have id{2} LE id but we can take the transposition (12) ∈ Ge = S2 and acting on the right we obtain e 1 2 id{2}(12) = 2 ∅ ! which is not in the LE -class of id.

We want to considere cases where the group Ge acts on the right of LE(e) by right multiplication. Definition 5.3. A semigroup S is called right Fountain if every L class contains an idempotent, E(Se) where E(S) is the set of all idempotents of S. e Proposition 5.4 ([26, Corollary 3.4]). Let S be a finite right Fountain semigroup and let e ∈ E(S) be an idempotent. Then Ge acts on the right of LE(S)(e).

Another case is where LE is a right congruence. e Lemma 5.5. Let S be a semigroup such that L is a right congruence. Then G acts on the right e E e of LE(e) by right multiplication for every e ∈ E. e

Proof.e Let x ∈ LE(e) and g ∈ Ge. Then xLE e so the right congruence property implies xgLE eg.

Now g H e implies gLEe and eg = g so we have also xgLEe as required. e e e Example 5.6. Anothere trivial case is where S is an eH-trivial semigroup hence Ge = {e} and clearly acts trivially on LE(e).

Recall that kLE(e) is thee k-vector space with basis the elements of LE(e). The following is an immediate corollary of Lemma 5.1. e e Corollary 5.7. If Ge acts on the right of LE(e) as in the above cases, then kLE(e) is a kS − kGe bimodule. Therefore, if V is a kGe-module, the tensor product e e

kLE(e) V kG Oe e is a kS-module.

Remark 5.8. As mentioned above, if LE is replaced by the standard Green’s L class then the above module is the induced Schützenberger module of J and V (where e belongs to the J -class J) . e

17 6 Projective modules of right restriction EI-Ehresmann semi- groups

In this section we study projective modules of finite Ehresmann and right restriction semigroups. The main result will be in the case of right restriction EI-Ehresmann semigroups. We start with some lemmas that will be of later use.

Lemma 6.1. Let S be an Ehresmann and right restriction semigroup and let s,m ∈ S. Then,

+ ∗ s · m ∈ LE(m) ⇐⇒ m ≤ s

(where ≤ is the natural partial order on thee subsemilattice E).

Proof. First note that ∗ ∗ s · m ∈ LE (m) ⇐⇒ (sm) = m .

If (sm)∗ = m∗ then by the right ample identitye

s∗m = m(s∗m)∗ and by the right congruence identity

m(s∗m)∗ = m(sm)∗ = mm∗ = m.

Since m+ is the minimal projection which is a left identity of m we obtain m+ ≤ s∗. In the other direction, assume m+ ≤ s∗ which implies s∗m = m. By the right ample and right congruence identities we have m = s∗m = m(s∗m)∗ = m(sm)∗ which implies that m∗ ≤ (sm)∗. However, smm∗ = sm so (sm)∗ ≤ m∗ hence (sm)∗ = m∗ as required.

Definition 6.2. Let ≤ be a poset on a set S and let s ∈ S. The principal down ideal generated by s is the set s ↓= {x ∈ S | x ≤ s}.

Now consider the poset ≤l on an Ehresmann semigroup S. Recall that on the set of projections E, the poset ≤l coincide with the natural semilattice order denoted by ≤. Therefore

s ↓= {x ∈ S | x ≤l s}

18 and s∗ ↓= {e ∈ E | e ≤ s∗} for every s ∈ S. The following lemma is well-known and we prove it for the sake of completeness.

Lemma 6.3. Let S be an Ehresmann semigroup and let s ∈ S. The function F : s∗ ↓→ s ↓ defined by f(e)= se is an isomorphism of posets with inverse G : s ↓→ s∗ ↓ defined by G(x)= x∗.

Proof. First note that e1 ≤ e2 implies se1 ≤l se2 since se2e1 = se1 so F is a of ∗ posets . Now, x1 ≤l x2 says that x1 = x2x1. By the right congruence identity

∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ x2x1 = (x2x1) = (x2x1) = x1 so ∗ ∗ G(x1)= x1 ≤ x2 = G(x2) hence G is also a homomorphism of posets. Finally, we show that F and G are inverses of each other: G(F (e)) = G(se) = (se)∗ = s∗e = e since e ≤ s∗ and F (G(x)) = F (x∗)= sx∗ = x since x ≤l s.

Lawson [19] proved that the category of all Ehresmann semigroups is isomorphic to the category of all Ehresmann categories (with an appropriate choice of morphisms). In fact, the function C described in Section 2.2 is the object part of this isomorphism. Another link between certain Ehresmann semigroups and their corresponding categories appear when we consider their algebras as seen in the following theorem.

Theorem 6.4 ([35, Theorem 1.5]). Let k be any commutative unital ring, let S be a finite Ehres- mann and right restriction semigroup and let C be the corresponding Ehresmann category. There is an isomorphism of k-algebras kS ≃ kC. Explicit isomorphisms ϕ : kS → kC, ψ : kC → kS are defined (on basis elements) by ϕ(s)= C(t) t s X≤l

19 ψ(C(x)) = µ(y, x)y y x X≤l where µ is the Mï¿œbius function of the poset ≤l (see [32, Chapter 3]). Remark 6.5. This theorem is a generalization of several results due to Solomon [31], Steinberg [38], Guo and Chen [12] and the second author [33]. It was further generalized by Wang [42] to the class of P -Ehresmann and right\left P -restriction semigroups. P -Ehresmann and P -restriction semigroups were introduced by Jones in [18]. We want to use the isomorphism of Theorem 6.4 in order to study the projective modules of kS where S is an Ehresmann and right restriction semigroup. Let k be a field, and let e ∈ E be a projection. Recall that we consider the composition in the Ehresmann category C “from right to left” so a morphism C(a) has domain a∗ and range a+. The idempotent e is an object of C with identity morphism C(e). Consider the set C· C(e) of all the morphisms whose domain is e. The category C acts on the left of C· C(e) by concatenation of morphisms. More precisely, recall that an action of C is a functor F from C to the and functions. Here, for every object c of C we set F(c)= C(e,c) to be the set of all morphisms with domain e and range c. For every morphism C(m) ∈ C(c1,c2) the morphism F(C(m)) : C(e,c1) → C(e,c2) is defined by left composition by C(m). This action yields a kC-module denoted by kC· C(e) whose elements are linear combinations of morphisms in C· C(e). Again, we can describe this module as a functor from C to the category of k-vector spaces and linear transformations, but we prefer the description via a category algebra. It is clear that kC· C(e) is a projective module since C(e) is an idempotent of kC. Since ϕ of Theorem 6.4 is an isomorphism, we can view kC· C(e) also as a kS-module whose action ⋆ is defined by

s ⋆ C(m)= ϕ(s) · C(m)= C(t) · C(m)   t s X≤l   where s,m ∈ S and · is the action of kC on kC· C(e). However, this description is not natural from the semigroup point of view, so we would like to describe this module using the structure of the semigroup itself. Let LE(e) be the LE -class of e. Observe that mLEe if and only if the domain of

C(m) is e. Therefore, the sets LE(e) and C· C(e) are in one-to-one correspondence. Furthermore e e e we claim the following. e Proposition 6.6. There is an isomorphism of kS-modules

Φ: kLE (e) → kC· C(e) defined (on basis elements) by e Φ(m)= C(m)

20 for every m ∈ S.

Proof. It is clear that Φ is an isomorphism of k-vector spaces. For every s ∈ S it is left to prove that Φ(s · m)= s⋆ Φ(m) where s⋆ Φ(m)= ϕ(s)Φ(m)= C(r) · C(m). r s X≤l First we prove that s · m =0 ⇐⇒ C(r) · C(m)=0. According to Lemma 6.1, s · m =0 implies

r≤ls + ∗ X ∗ + m 6≤ s so by Lemma 6.3 there is no element t ≤l s such that t = m hence

C(r) · C(m)=0. r s X≤l The converse implication follows in the same manner. Now, assume s · m 6= 0 so m+ ≤ s∗. By ∗ + Lemma 6.3, there exists a unique t ≤l s with t = m . In other words, there is a unique t ≤l s such that C(t) · C(m) 6=0. Therefore,

s⋆ Φ(m)= C(r) · C(m)= C(t) · C(m)= C(tm). r s X≤l Note that tm = st∗m = sm+m = sm hence

Φ(sm)= C(sm)= C(tm)= s⋆ Φ(m) as required.

Let Ge be the group H-class of e and let g ∈ Ge. According to Lemma 5.5 ,Ge acts on the right of

LE(e) since LE is a right congruence. It is also clear that the domain and range of C(g) is e, so G acts on the right of C· C(e) as well. Moreover, for every m ∈ LE (e) we have e e C(mg)= C(m)C(g) e so Φ is not only a kS -modules isomorphism but a kS − kGe bimodule isomorphism. Therefore, we obtain the following corollary.

Corollary 6.7. Let S be a finite Ehresmann and right restriction semigroup and let e ∈ E. Denote by Ge the group H-class of e and let V be any Ge-module. Then there is an isomorphism of kS

21 modules

kLE (e) V ≃ kC· C(e) V. kG kG Oe Oe e In particular, if p ∈ kGe is a primitive idempotent of kGe, and C(p) is the corresponding linear combination of morphisms, there is an isomorphism

kCC(p)= kCC(e)C(p) ≃ kC· C(e) kGeC(p) ≃ kLE (e) kGep. kG kG Oe Oe e Clearly, kCC(p) is a projective module since C(p) is an idempotent. So this gives a semigroup theoretic description of certain projective modules of S. However, there is no reason for the in- decomposable projective modules to be of this form. The situation is much better if we consider (finite) right restriction EI-Ehresmann semigroups. A proof of the following lemma can be found in [23, Lemma 9.31] or [43, Corollary 4.5].

Lemma 6.8. Let C be a finite EI-category and given an object e let Ge = C(e,e) to be its automor- e e phism group and Pe = {p1,...,pme } to be a complete set of primitive orthogonal idempotents for kGe. The set

Pe e [ (where the union is taken over all objects of C) is a complete set of primitive orthogonal idempotents for kC.

From now on we assume that the order of every (maximal) subgroup of S is invertible in k. This implies that kGe is semisimple for every e ∈ E by Maschke’s theorem.

Theorem 6.9. Let S be a finite right restriction EI-Ehresmann semigroup. Let e ∈ E and Ge =

C(e,e) be as above and let V ∈ Irr kGe. Then

kLE(e) V kG Oe e is an indecomposable projective module of kS. Moreover, every indecomposable projective module of kS is isomorphic to a module of this form for an appropriate choice of a projection e ∈ E and a simple Ge-module V .

Proof. This is immediate from Corollary 6.7 and the fact that if kGe is semisimple then V ≃ kGep where p is a primitive idempotent of kGe which is also a primitive idempotent of kC by Lemma 6.8.

Remark 6.10. A similar result for indecomposable projective modules of a certain type of right Fountain monoids is obtained in [26, Theorem 4.7].

22 Isomorphic copies of indecomposable projective modules The set

{kLE(e) V | e ∈ E, V ∈ Irr Ge} kG Oe e usually contains isomorphic modules. We want to obtain a list of all indecomposable projective modules up to isomorphism but without isomorphic copies. From the categorical point of view we can say that

kLE (e1) V1 ≃ kLE (e2) V2 kG kG Oe1 Oe2 e e if and only if e1 is isomorphic to e2 (as objects in a category) and V1 ≃ V2 as Ge1 -modules (note that Ge1 ≃ Ge2 ) - see [43, Corollary 4.2 and Proposition 4.3]. From the semigroup point of view we can say even more. We start with two lemmas that will make the situation more transparent.

Lemma 6.11 ([34, Corollary 5.5]). Let S be a finite Ehresmann semigroup and let e,f ∈ E be two projections. Then e and f are isomorphic (as objects in the corresponding Ehresmann category) if and only if e J f in S.

Lemma 6.12. Let S be a finite EI-Ehresmann semigroup. Then in every regular J -class of S there is a projection e ∈ E.

Proof. Let J be a regular J -class and choose an idempotent f ∈ J. Lemma 3.3 says that f ∗f + is an inverse of f. Therefore f J f ∗f + so e = f ∗f + is a projection in J.

Remark 6.13. The converse of Lemma 6.12 is not true. Recall that B2 is the monoid of all binary relations on the set {1, 2}, i.e., subsets of {1, 2}2. We have already mentioned (Example 3.12) that + this is an Ehresmann semigroup with respect to the semilattice of partial identities - α = 1im(a) ∗ and α = 1dom(a) for every α ∈ B2. Consider the submonoid M ⊆ B2 of the 12 relations whose sizes are 0, 1, 2 or 4 (as subsets of {1, 2}2). It is a routine matter to verify that this is indeed a submonoid of B2. Since E ⊆ M it is clear that M is also an Ehresmann semigroup. It is also routine to check that M has three J -classes: One of the two invertible elements, one with the zero element and another with the other 9 elements. Therefore, every J -class is regular and contains a projection. On the other hand, choose the idempotent f = {(1, 1), (2, 2), (1, 2), (2, 1)} ∈ M. Since dom(f) = im(f) = {1, 2} it is clear that the morphism C(f) is a non-invertible endomorphism of the object id ∈ E in the corresponding Ehresmann category. Therefore, the category is not an EI-category and M is not EI-Ehresmann.

According to Lemma 6.11 and Lemma 6.12 choosing one object from every isomorphism class of objects is precisely like choosing one projection from every regular J -class. So we can fix a set of representative projections I = {e1,...,ek}, one from every regular J -class whose corresponding

23 maximal groups are GJ1 ,...,GJk . Now,

kL (e ) V | i =1,...,k V ∈ Irr kG  E i Ji  kG  OJi  e is a list of all the indecomposable projective modules (up to isomorphism) without isomorphic copies. This is an explicit correspondence between indecomposable projective modules of S and pairs (J, V ) where J is a regular J -class and V ∈ Irr kGJ . From this point of view, the correspondence between a simple module and its projective cover is clear. It is easy to see that for every representative projection e ∈ I there is an kS-module epimorphism

ψ : kLE(e) → k L(e) defined on basis elements by e m m ∈ L(e) ψ(m)= 0 m∈ / L(e).  This induces an kS-module epimorphism 

ψ ⊗ id : kLE(e) V → k L(e) V kG kG OJ OJ e for every V ∈ Irr kGJ (where J is the regular J -class of e). This is an explicit description of the epimorphism from an indecomposable projective module and its simple image.

Dimension of indecomposable projective modules Next we want to describe the dimension of the projective modules. Recall that we assume that the order of every group being discussed is invertible in k hence its algebra is semisimple and here we also assume that k is algebraically closed. Recall that if V is a kG-module then the dual D(V ) = Homk(V, k) is a right kG-module. We will use a well-known fact whose proof we will briefly sketch.

Proposition 6.14. Let G be a finite group, let M be a right kG-module and let V ∈ Irr kG. The dimension of M V (as a k-vector space) is the multiplicity of D(V ) in the decomposition of M kG into simple (right)OkG-modules.

Proof. Let V1,...,Vr be the simple kG-modules up to isomorphism. Fix a set {p1,...,pr} of primitive orthogonal idempotents such that Vi ≃ kGpi. It is easy to check that

pikG kGpj = pikGpj kG O

24 as k-vector spaces. Moreover, since k is algebraically closed, the Wedderburn-Artin theorem implies that k i = j pikGpj ≃ 0 otherwise  hence,  1 i = j dim pikGpj = 0 otherwise  (this is actually Schur’s Lemma). Now, consider the decomposition of M into a sum of simple right kG-modules r M ≃ nlplkG l M=1 and let V ≃ kGpi for some i. It is clear that

r dim M kGpi = dim nlplkGpi = ni. CG l O M=1

So the dimension is indeed the multiplicity of D(V ) ≃ pikG in the decomposition of M as required.

As an immediate corollary of Theorem 6.9 and Proposition 6.14, we obtain the following result.

Corollary 6.15. Let S be a finite right restriction EI-Ehresmann semigroup. Let J be a regular

J -class and choose a projection e ∈ J (such exists by Lemma 6.12). Let V ∈ Irr kGJ where GJ is the maximal subgroup associated with J. Assume also that k is algebraically closed and the order of every subgroup of S is invertible in k. Then the dimension of kL˜E(e) V as a k-vector space CG OJ is the multiplicity of D(V ) in the decomposition of kLE (e) as a right GJ -module.

e 7 The monoid of partial functions

Recall that PT n denotes the monoid of all partial functions on the set {1,...,n}. In this section we apply the results of Section 6 in this specific case. For this section we fix k = C, the field of complex numbers. Let A ⊆ {1,...,n} and denote by idA the partial identity function on A. As already mentioned in Section 3.2, PT n is a right restriction EI-Ehresmann semigroup with

E = {idA | A ⊆{1,...,n}}

25 as a subsemilattice of projections. Recall that we are composing functions from right to left. We denote the corresponding Ehresmann category by En, and it can be described in the following way. The objects of En are subsets of {1,...,n} and for every A, B ⊆ {1,...,n} the hom-set

En(A, B) contains all the (total) onto functions from A to B. The endomorphism monoid of an object A is the symmetric group SA so the category En is indeed an EI-category. Let f : A → B be a partial function. We denote by dom(f) ⊆ A and im(f) ⊆ B the domain and image of f.

Recall that the (set theoretic) kernel of f is the on dom f defined by a1 ∼ a2 if f(a1)= f(a2). We denote it by ker f. Consider two partial functions f1,f2 ∈ PT n. It is easy to see that f1LEf2 if and only if dom(f1)= dom(f2). It is well known that f1 and f2 are J -equivalent if

| im(f1)| = | im(f2)| and L-equivalent if dom(f1)= dom(f2) and ker f1 = ker f2. All the J -classes of e PT n are regular and we choose a representative projection idk = id{1,...,k} ∈ E for every J -class.

The group corresponding to the J -class of idk is the symmetric group Sk. Therefore, the maximal subgroups of PT k are Sk for 0 ≤ k ≤ n (note that S0 ≃ S1). It is well known that simple modules of

CSk are indexed by partitions of k (or Young diagrams). Therefore, the simple modules of C PT n can be indexed by partitions λ ⊢ k for 0 ≤ k ≤ n. We denote the simple module of Sk corresponding λ to a partition λ ⊢ k by S . According to Theorem 4.3 all simple modules of C PT n are of the form λ C L(idk) S for 0 ≤ k ≤ n and λ ⊢ k (this is a known fact, see [7, Section 11.3], and this is also

CSk a corollaryO of [24, Theorem 4.4]). Theorem 6.9 yields the following result.

λ Theorem 7.1. The modules CLE(idk) S for 0 ≤ k ≤ n and λ ⊢ k form a list of all the CS Ok indecomposable projective modulese of C PT n up to isomorphism. (Note that LE(idk) consists of all the partial functions f ∈ PT n with dom(f)= {1,...,k}). e

Dimension of indecomposable projective modules Our goal now is to obtain a formula for λ the dimension (as a C-vector space) of the projective module CLE(idk) ⊗ S where λ ⊢ k. By λ Corollary 6.15 we need to consider the multiplicity of D(S ) in the right CSk module CLE(idk). e This is a permutation module, induced by the (right) action of Sk on LE (idk). We start by e examining this action. Let f : {1,...,k}→{b1,...,br} be an onto function. We associate with f e −1 a composition µf = [µ1,...,µr]  k where µi is the number of elements in the preimage f (bi). For instance, if f : {1, 2, 3, 4, 5, 6}→{2, 3, 6} is defined by

123456 f = 262636 ! then µf = [2, 1, 3]  6. It is clear that two functions f1,f2 ∈ LE (idk) are in the same orbit (under im im the right action of Sk) if and only if (f1) = (f2) and µf1 = µf2 . Therefore, we can index the e orbits by pairs (B,µ) where B ⊆ {1,...,n} and µ  k is a composition with |µ| = |B|. Denote

26 such an orbit by O(B,µ). It is also easy to check that if f ∈ O(B,µ) with µ = [µ1,...,µr] then the stabilizer of f is isomorphic to Sµ = Sµ1 × ... × Sµr (the Young subgroup corresponding to µ). Now, as a C-vector space we have a decomposition

λ λ CLE(idk) S ≃ CO(B,µ) S CS CS Ok M  Ok e where the sum is over all pairs (B,µ) where B ⊆ {1,...,n} and µ  k is a composition with |µ| = |B|. This is clearly isomorphic to

λ CO(B,µ) S . CS ! M Ok

Since Sk acts transitively on O(B,µ) and since Sµ is the stabilizer of this action it is known that O ≃ IndSk tr where tr is the trivial module of S . This is the Young module corresponding C (B,µ) Sµ µ µ C µ λ Sk λ to µ. The multiplicity of S in the decomposition of IndSµ trµ (or D(S ) if we consider the right module) is the Kostka number Kλµ. Therefore,

λ dim CO(B,µ) S = Kλµ CS Ok and λ dim CO(B,µ) S = Kλµ CS ! M Ok X where again the sum is over all pairs (B,µ) where B ⊆{1,...,n} and µ  k is a composition with

|µ| = |B|. Since Kλµ does not depend on B this equals

k n K . l λµ l=1 µk   X |Xµ|=l

So we have obtained the following result.

Proposition 7.2. The dimension of the indecomposable projective module λ CLE(idk) ⊗ CS is given by the formula

k e n K . l λµ l=1 µk   X |Xµ|=l

Cartan matrix of C PT n

27 Definition 7.3. Let A be a C-algebra. Let {S(1),...,S(r)} be the irreducible representations of A up to isomorphism and let {P (1),...,P (r)} be the indecomposable projective modules of A ordered such that P (i) is the projective cover of S(i). The Cartan matrix of A is the r × r matrix whose (i, j) entry is the number of times that S(i) appears as a Jordan-Hï¿œlder factor in P (j).

We want to show that certain entries in the Cartan matrix of C PT n equal 0. The proof is similar to the proof of Proposition 7.2. The irreducible representations of C PT n (and hence the rows and columns of the Cartan matrix) are indexed by Young diagrams with k boxes for 0 ≤ k ≤ n. For every Young diagram α ⊢ k we denote by S(α) and P (α) the associated irreducible and indecomposable projective module respectively. We think of the Cartan matrix as an (n + 1) × (n + 1) block matrix where the (i, j) block contains pairs (α, β) of Young diagrams such that α ⊢ (i − 1) and β ⊢ (j − 1). Denote by E(k, r) the set of all onto functions f : {1,...,k}→{1,...,r}. It is clear that CE(k, r) has a natural structure of CSr − CSk bi-module. The following result regarding the Cartan matrix was obtained in [36, Corollary 4.2] (this is a specific case of a known formula for the Cartan matrix of any EI-category algebra).

Proposition 7.4. Let α ⊢ r and β ⊢ k be two Young diagrams. The number of times that S(α) appears as a Jordan-Hï¿œlder factor of P (β) is the number of times that Sα appears as an irreducible β constituent in the CSr module CE(k, r) S . CS Ok Remark 7.5. It is proved in [36] that the Cartan matrix of C PT n is block upper unitriangular. Moreover, a combinatorial description for the first and second block superdiagonals is given there.

Proposition 7.6. Let β ⊢ k be a Young diagram such that r< |β| then CE(k, r) Sβ =0. CS Ok Proof. According to Proposition 6.14,

dim CE(k, r) Sβ CS Ok β is the multiplicity of D(S ) in the right CSk-module CE(k, r). This is a permutation module, induced from the right action of Sk on E(k, r). It is clear that two functions f1,f2 ∈ E(k, r) are  −1 in the same orbit if and only if µf1 = µf2 (again, µf = [µ1,...,µr] k where µi = |f (i)|). Therefore, we can index the orbits by compositions, µ  k with |µ| = r. Denote such an orbit by

28 Oµ. As C-Vector spaces

β β CE(k, r) S ≃  Oµ S

CSk µk CSk O |Mµ|=r  O     β ≃ Oµ S .

µk CSk ! |Mµ|=r O

It is clear that the stabilizer of f ∈ Oµ is Sµ ≃ Sµ1 ×···×Sµr . As we have already seen, this implies that

β Sk β dim Oµ S = dim IndSµ trµ S = Kβµ. CS CS Ok Ok The assumption |µ| = r < |β| implies that Kβµ = 0 since there can be no semistandard Young tableau with shape β and content µ for |µ| < |β|.

Proposition 7.6 immediately implies the following corollary:

Proposition 7.7. Let α ⊢ r and β ⊢ k two young diagrams such that r < |β|. Then the (α, β) entry in the Cartan matrix of C PT n is 0.

Remark 7.8. This result is a generalization of some known facts about the projective modules of k C PT n. Consider the case where β = [1 ] for 1 ≤ k ≤ n. Proposition 7.7 (with the fact that the Cartan matrix is unitriangular) says that the only irreducible constituent in P (β) is S(β).

Therefore S(β) ≃ P (β) which is a known fact about PT n (see [36, Lemma 6.7]). In the case where β = [2, 1k−2] for some 2 ≤ k ≤ n, Proposition 7.7 is precisely [36, Lemma 6.4] which is a key step in the proof that the global dimension of C PT n is n − 1.

Appendix: Ehresmann semigroups and categories - two coun- terexamples

The isomorphism between the algebras of a right restriction Ehresmann semigroup and the corre- sponding category (Theorem 6.4) was presented in [34] without the requirement of S being right restriction. However, Shoufeng Wang has found out that the proof implicitly assumes it. In fact, the functions ψ and ϕ of Theorem 6.4 are k-algebra if and only if S is right re- striction [42, Lemma 4.3]. This led to a correction [35], but it was not clear whether being right restriction is necessary for the conclusion of Theorem 6.4 to be true. In this appendix we show that this requirement can not be omitted. A counterexample is given by the partition monoid which

29 arises in the study of partition algebras (see [13]). On the other hand, we show that the algebra of the semigroup in Example 3.9 is isomorphic to the corresponding category algebra despite being neither left nor right restriction.

The partition monoid Pn for n ∈ N can be described in the following way. Define n = {1,...,n} ′ ′ ′ ′ ′ and n = {1 ,...,n }. An element α ∈Pn is a partition of the set n ∪ n . For every x, y ∈ n ∪ n we write x ∼α y when x and y are in the same partition class of α. We refer the reader to [13] for the description of the operation of Pn which is best pictured with certain diagrams.

For any partition α ∈Pn we follow [5] and define two equivalence relations on n by

ker(α)= {(i, j) ∈ n × n | i ∼α j}

′ ′ coker(α)= {(i, j) ∈ n × n | i ∼α j }.

Note that there are two types of equivalence classes in ker(α). A kernel class K is in the domain of ′ ′ ′ α if there exists j ∈ n such that j ∼α i for some (and hence all) i ∈ K. Likewise a cokernel class ′ ′ ′ ′ K is in the of α if there exists i ∈ n such that i ∼α j for some (all) j ∈ K . Note also that the number of domain classes in ker(α) equals the number of codomain classes in coker(α).

For every equivalence relation σ on n we define a partition eσ ∈ Pn in the following way. Two ′ ′ elements x, y ∈ n ∪ n will be in the same eσ-class if (i, j) ∈ σ where x = i or x = i and y = j or ′ y = j . It is clear that eσ is an idempotent with ker(eσ) = coker(eσ)= σ. Moreover, eσ1 eσ2 = eσ3 where σ3 is the minimal equivalence relation such that σ1, σ2 ⊆ σ3 (i.e., it is the join in the lattice of equivalence relations on n). Therefore, the set

E = {eσ | σ is an equivalence relation on n}

is a subsemilattice of Pn (note that eσ1 ≤ eσ2 if σ2 ⊆ σ1). For every α ∈Pn it is clear that eker(α)

(ecoker(α)) is the minimal element e ∈ E such that eα = α (respectively, αe = α) hence we can write

+ ∗ α = eker(α), α = ecoker(α).

It is also possible to check that the identities

(αβ)+ = (αβ+)+, (αβ)∗ = (α∗β)∗

are satisfied for every α, β ∈ Pn. Therefore, Pn is an Ehresmann semigroup and there exists a corresponding Ehresmann category Cn. This fact was first noted by East and Gray [6]. In this specific case it will be more convenient to compose morphisms in a category “from left to right” because it fits better the standard description of the partition monoid. The objects of Cn are in

30 one-to-one correspondence with equivalence relations (or partitions) on n and for every α ∈ Pn there is a corresponding morphism C(α) with domain ker(α) and range coker(α). Note that Cn is not an EI-category. According to [13], the algebra C Pn is not semisimple (for n> 1). In order to complete our counterexample it is enough to show that CCn is semisimple, hence k Pn 6≃ kCn for 1 n k = C. Denote by n the identity relation on . Note that e1n is the identity of Pn which will also be denoted by 1n for the sake of simplicity.

Proposition 7.9. Let k be a field.

1. The following holds

kCn1nkCn = kCn.

2. There is an isomorphism of k-algebras

1nkCn1n ≃ k ISn

(where ISn is the symmetric inverse monoid).

Proof. 1. Let α ∈Pn. Denote by K1,...,Kr the kernel classes of the domain and by Kr+1,...,Ks ′ ′ the other kernel classes. Likewise, denote by K1,...,Kr the cokernel classes of the codomain ′ ′ ′ and by Kr+1,...,Ks′ the other cokernel classes (s 6= s in general). We order the classes such ′ that i ∼α j for i ∈ Kl and j ∈ Kl (for l ≤ r). Define a partition α1 ∈Pn by

′ ′ ′ ′ α1 = {K1 ∪{1 },...,Kr ∪{r },Kr+1,...Ks, {(r + 1) },..., {n }}

and another partition α2 ∈Pn by

′ ′ ′ ′ α2 = {{1} ∪ K1,..., {r} ∪ Kr, {r +1},..., {n},Kr+1,...,Ks′ }.

First note that α = α1α2 = α11nα2 and moreover,

ker(α1) = ker(α)

coker(α1)= 1n = ker(α2)

coker(α2) = coker(α).

This implies that

C(α)= C(α1)C(1n)C(α2)

since the composition of morphisms on the right hand side is defined. Therefore, kCn1nkCn =

kCn as required.

31 2. It is clear that C(α) ∈ 1nCn1n if and only if ker(α) = coker(α) = 1n which implies that

α ∈ ISn (according to the natural embedding of ISn in Pn, see [5, Section 3]). The claim follows immediately.

Corollary 7.10. kCn is a semisimple algebra if and only if n! is invertible in k.

Proof. It is well known (see [22, Theorem 2.8.7]) that for any algebra A and an idempotent e ∈ A, if A = AeA then the algebras A and eAe are Morita equivalent. In our case, we take A = CCn and e = 1n so Proposition 7.9 implies that kCn is Morita equivalent to k ISn. It is well known that k ISn is a semisimple algebra if and only if n! is invertible in k ([40, Corollary 9.7]) hence the same holds for kCn as well.

Corollary 7.10 for k = C settles the counterexample which is the main goal of this section. Finally, we would like to give an example of an EI-Ehresmann semigroup that is not left nor right restriction, but still has an algebra isomorphic to the algebra of its category. Consider Example 3.9 given above. The category C is a poset. That is, for any two objects a,b there is at most one morphism in the union of hom-sets C(a,b) ∪C(b,a). Therefore, its algebra is the incidence algebra of this poset. In this case the poset is the linear poset (id, 1) < (1, 1) < (1, id) so in this case the incidence algebra (over a field k) is the algebra UT3(k) of all upper triangular matrices over k.

Now, consider the map Φ: S → UT3(k) defined by

1 0 0 0 0 0 1 1 ( , id) →  0 1 0  , (id, ) →  0 1 0  0 0 0 0 0 1         0 0 0 0 0 0 1 1 2 1 ( , ) →  0 1 0  , ( , ) →  0 1 1  0 0 0 0 0 0         0 1 0 0 1 1 1 2 2 2 ( , ) →  0 1 0  , ( , ) →  0 1 1  . 0 0 0 0 0 0         It is not difficult to check that Φ is a one-to-one semigroup homomorphism and that its image is a basis of UT3(k). Therefore Φ extends to an isomorphism Φ: kS → UT3(k). This implies that kS ≃ kC in this case. We leave as an open problem to determine necessary and sufficient conditions for an EI-Ehresmann finite semigroup to have an algebra isomorphic to that of its associated category algebra. It may

32 be true that every finite EI-Ehresmann semigroup has an algebra isomorphic to the algebra of its category. Acknowledgments: The authors thank the referee for his\her helpful comments. We thank Professor Michael Kinyon and Professor Victoria Gould for a number of useful discussions and for the specific results we noted in the body of the paper. We thank Professor Timothy Stokes for pointing our attention to references [15] and [30] and for discussions related to their content and relation to this paper.

References

[1] Jorge Almeida. Finite semigroups and universal algebra, volume 3 of Series in Algebra. World Scientific Publishing Co. Inc., River Edge, NJ, 1994. Translated from the 1992 Portuguese original and revised by the author.

[2] Ibrahim Assem, Daniel Simson, and Andrzej Skowroński. Elements of the representation the- ory of associative algebras. Vol. 1, volume 65 of London Mathematical Society Student Texts. Cambridge University Press, Cambridge, 2006. Techniques of representation theory.

[3] Stanley Burris and H. P. Sankappanavar. A course in universal algebra, volume 78 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1981.

[4] Charles W Curtis and Irving Reiner. Representation theory of finite groups and associative algebras, volume 356. American Mathematical Soc., 1966.

[5] James East. On the singular part of the partition monoid. Internat. J. Algebra Comput., 21(1-2):147–178, 2011.

[6] James East and Robert D Gray. Ehresmann theory and partition monoids. arXiv preprint arXiv:2011.00663, 2020.

[7] Olexandr Ganyushkin and Volodymyr Mazorchuk. Classical finite transformation semigroups, volume 9 of Algebra and Applications. Springer-Verlag London Ltd., London, 2009. An intro- duction.

[8] Olexandr Ganyushkin, Volodymyr Mazorchuk, and Benjamin Steinberg. On the irreducible representations of a finite semigroup. Proc. Amer. Math. Soc., 137(11):3585–3592, 2009.

[9] Victoria Gould. Notes on restriction semigroups and related structures; formerly (weakly) left E-ample semigroups, 2010.

33 [10] Victoria Gould. Restriction and Ehresmann semigroups. In Proceedings of the International Conference on Algebra 2010, pages 265–288. World Sci. Publ., Hackensack, NJ, 2012.

[11] Xiaojiang Guo. A note on locally inverse semigroup algebras. Int. J. Math. Math. Sci., 2008. Art. ID 576061, 5.

[12] Xiaojiang Guo and Lin Chen. Semigroup algebras of finite ample semigroups. Proc. Roy. Soc. Edinburgh Sect. A, 142(2):371–389, 2012.

[13] Tom Halverson and Arun Ram. Partition algebras. European J. Combin., 26(6):869–921, 2005.

[14] John 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.

[15] Marcel Jackson and Tim Stokes. Partial maps with domain and range: extending Schein’s representation. Comm. Algebra, 37(8):2845–2870, 2009.

[16] Gordon James and Adalbert Kerber. The representation theory of the symmetric group, vol- ume 16 of Encyclopedia of Mathematics and its Applications. Addison-Wesley Publishing Co., Reading, Mass., 1981. With a foreword by P. M. Cohn, With an introduction by Gilbert de B. Robinson.

[17] Yingdan Ji and Yanfeng Luo. Locally adequate semigroup algebras. Open Math., 14:29–48, 2016.

[18] Peter R. Jones. A common framework for restriction semigroups and regular ∗-semigroups. J. Pure Appl. Algebra, 216(3):618–632, 2012.

[19] M. V. Lawson. Semigroups and ordered categories. I. The reduced case. J. Algebra, 141(2):422– 462, 1991.

[20] Mark V. Lawson. Inverse semigroups. World Scientific Publishing Co., Inc., River Edge, NJ, 1998. The theory of partial symmetries.

[21] Liping Li. A characterization of finite EI categories with hereditary category algebras. J. Algebra, 345:213–241, 2011.

[22] Markus Linckelmann. The block theory of finite group algebras. Vol. I, volume 91 of London Mathematical Society Student Texts. Cambridge University Press, Cambridge, 2018.

[23] Wolfgang Lück. Transformation groups and algebraic K-theory, volume 1408 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 1989. Mathematica Gottingensis.

34 [24] Stuart Margolis and Benjamin Steinberg. The quiver of an algebra associated to the Mantaci- Reutenauer descent algebra and the homology of regular semigroups. Algebr. Represent. The- ory, 14(1):131–159, 2011.

[25] Stuart Margolis and Benjamin Steinberg. Quivers of monoids with basic algebras. Compos. Math., 148(5):1516–1560, 2012.

[26] Stuart Margolis and Benjamin Steinberg. Projective indecomposable modules and quivers for monoid algebras. Comm. Algebra, 46(12):5116–5135, 2018.

[27] W. McCune. Prover9 and mace4. http://www.cs.unm.edu/~mccune/prover9/, 2005–2010.

[28] John Rhodes and Benjamin Steinberg. The q-theory of finite semigroups. Springer Monographs in Mathematics. Springer, New York, 2009.

[29] Bruce E. Sagan. The symmetric group, volume 203 of Graduate Texts in Mathematics. Springer- Verlag, New York, second edition, 2001. Representations, combinatorial algorithms, and sym- metric functions.

[30] Boris M Shain. Restrictively multiplicative algebras of transformations. Izvestiya Vysshikh Uchebnykh Zavedenii. Matematika, (4):91–102, 1970.

[31] Louis Solomon. The Burnside algebra of a finite group. J. Combinatorial Theory, 2:603–615, 1967.

[32] Richard P. Stanley. Enumerative combinatorics. Vol. 1, volume 49 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 1997. With a foreword by Gian-Carlo Rota, Corrected reprint of the 1986 original.

[33] Itamar Stein. The representation theory of the monoid of all partial functions on a set and related monoids as EI-category algebras. J. Algebra, 450:549–569, 2016.

[34] Itamar Stein. Algebras of Ehresmann semigroups and categories. Semigroup Forum, 95(3):509– 526, 2017.

[35] Itamar Stein. Erratum to: Algebras of Ehresmann semigroups and categories. Semigroup Forum, 96(3):603–607, 2018.

[36] Itamar Stein. The global dimension of the algebra of the monoid of all partial functions on an n-set as the algebra of the EI-category of epimorphisms between subsets. J. Pure Appl. Algebra, 223(8):3515–3536, 2019.

[37] Itamar Stein. Representation theory of order-related monoids of partial functions as locally trivial category algebras. Algebr. Represent. Theory, 23(4):1543–1567, 2020.

35 [38] Benjamin Steinberg. Möbius functions and semigroup representation theory. J. Combin. The- ory Ser. A, 113(5):866–881, 2006.

[39] Benjamin Steinberg. Möbius functions and semigroup representation theory. II. Character formulas and multiplicities. Adv. Math., 217(4):1521–1557, 2008.

[40] Benjamin Steinberg. Representation theory of finite monoids. Universitext. Springer, Cham, 2016.

[41] Tammo tom Dieck. Transformation groups, volume 8 of De Gruyter Studies in Mathematics. Walter de Gruyter & Co., Berlin, 1987.

[42] Shoufeng Wang. On algebras of P -Ehresmann semigroups and their associate partial semi- groups. Semigroup Forum, 95(3):569–588, 2017.

[43] Peter Webb. An introduction to the representations and cohomology of categories. In Group representation theory, pages 149–173. EPFL Press, Lausanne, 2007.

36