Derivations and relation modules for inverse semigroups
We define the derivation module for a homomorphism of inverse semigroups, generalizing a construction for groups due to Crowell. For a presentation map from a free inverse semigroup, we can then define its relation module as the kernel of a canonical map from the derivation module to the augmentati...
Збережено в:
Дата: | 2011 |
---|---|
Автор: | |
Формат: | Стаття |
Мова: | English |
Опубліковано: |
Інститут прикладної математики і механіки НАН України
2011
|
Назва видання: | Algebra and Discrete Mathematics |
Онлайн доступ: | http://dspace.nbuv.gov.ua/handle/123456789/154764 |
Теги: |
Додати тег
Немає тегів, Будьте першим, хто поставить тег для цього запису!
|
Назва журналу: | Digital Library of Periodicals of National Academy of Sciences of Ukraine |
Цитувати: | Derivations and relation modules for inverse semigroups / N.D. Gilbert// Algebra and Discrete Mathematics. — 2011. — Vol. 12, № 1. — С. 1–19. — Бібліогр.: 23 назв. — англ. |
Репозитарії
Digital Library of Periodicals of National Academy of Sciences of Ukraineid |
irk-123456789-154764 |
---|---|
record_format |
dspace |
spelling |
irk-123456789-1547642019-06-16T01:32:11Z Derivations and relation modules for inverse semigroups Gilbert, N.D. We define the derivation module for a homomorphism of inverse semigroups, generalizing a construction for groups due to Crowell. For a presentation map from a free inverse semigroup, we can then define its relation module as the kernel of a canonical map from the derivation module to the augmentation module. The constructions are analogues of the first steps in the Gruenberg resolution obtained from a group presentation. We give a new proof of the characterization of inverse monoids of cohomological dimension zero, and find a class of examples of inverse semigroups of cohomological dimension one. 2011 Article Derivations and relation modules for inverse semigroups / N.D. Gilbert// Algebra and Discrete Mathematics. — 2011. — Vol. 12, № 1. — С. 1–19. — Бібліогр.: 23 назв. — англ. 1726-3255 2000 Mathematics Subject Classification:20M18,20M50,18G20. http://dspace.nbuv.gov.ua/handle/123456789/154764 en Algebra and Discrete Mathematics Інститут прикладної математики і механіки НАН України |
institution |
Digital Library of Periodicals of National Academy of Sciences of Ukraine |
collection |
DSpace DC |
language |
English |
description |
We define the derivation module for a homomorphism of inverse semigroups, generalizing a construction for groups due to Crowell. For a presentation map from a free inverse semigroup, we can then define its relation module as the kernel of a canonical map from the derivation module to the augmentation module. The constructions are analogues of the first steps in the Gruenberg resolution obtained from a group presentation. We give a new proof of the characterization of inverse monoids of cohomological dimension zero, and find a class of examples of inverse semigroups of cohomological dimension one. |
format |
Article |
author |
Gilbert, N.D. |
spellingShingle |
Gilbert, N.D. Derivations and relation modules for inverse semigroups Algebra and Discrete Mathematics |
author_facet |
Gilbert, N.D. |
author_sort |
Gilbert, N.D. |
title |
Derivations and relation modules for inverse semigroups |
title_short |
Derivations and relation modules for inverse semigroups |
title_full |
Derivations and relation modules for inverse semigroups |
title_fullStr |
Derivations and relation modules for inverse semigroups |
title_full_unstemmed |
Derivations and relation modules for inverse semigroups |
title_sort |
derivations and relation modules for inverse semigroups |
publisher |
Інститут прикладної математики і механіки НАН України |
publishDate |
2011 |
url |
http://dspace.nbuv.gov.ua/handle/123456789/154764 |
citation_txt |
Derivations and relation modules for inverse semigroups / N.D. Gilbert// Algebra and Discrete Mathematics. — 2011. — Vol. 12, № 1. — С. 1–19. — Бібліогр.: 23 назв. — англ. |
series |
Algebra and Discrete Mathematics |
work_keys_str_mv |
AT gilbertnd derivationsandrelationmodulesforinversesemigroups |
first_indexed |
2025-07-14T06:52:23Z |
last_indexed |
2025-07-14T06:52:23Z |
_version_ |
1837604215893000192 |
fulltext |
Algebra and Discrete Mathematics RESEARCH ARTICLE
Volume 12 (2011). Number 1. pp. 1 – 19
c© Journal “Algebra and Discrete Mathematics”
Derivations and relation modules
for inverse semigroups
N. D. Gilbert
Communicated by V. I. Sushchansky
Abstract. We define the derivation module for a homo-
morphism of inverse semigroups, generalizing a construction for
groups due to Crowell. For a presentation map from a free inverse
semigroup, we can then define its relation module as the kernel of
a canonical map from the derivation module to the augmentation
module. The constructions are analogues of the first steps in the
Gruenberg resolution obtained from a group presentation. We give
a new proof of the characterization of inverse monoids of coho-
mological dimension zero, and find a class of examples of inverse
semigroups of cohomological dimension one.
1. Introduction
A cohomology theory for inverse semigroups was established in the funda-
mental work of Lausch [10] and Loganathan [14] but since then, there have
been limited applications of homological algebra to inverse semigroups.
However, important contributions have been made in closely related ar-
eas. These include an approach, due to Steinberg [18], to the study of
homomorphisms of inverse semigroups based on derived categories, and
developed in [18] in the wider context of morphisms of ordered groupoids;
the topology of 2–complex models of inverse semigroup presentations and
its application to amalgams, also due to Steinberg [19]; homotopy theory
in the category of ordered groupoids [12], leading to an alternative proof
I am very grateful for constructive email discussions with Benjamin Steinberg and
with Jonathon Funk about this work.
2000 Mathematics Subject Classification: 20M18,20M50,18G20.
Key words and phrases: inverse semigroup, cohomology, derivation, relation
module.
2 Derivations and relation modules
of Steinberg’s Factorization Theorem; and Funk’s [4] study of the topos
of representations of an inverse semigroup.
A key step in [14] was the reformulation of the module theory for an
inverse semigroup S, due to Lausch [10], as the module theory of a left
cancellative category L(S). An S–module is then a functor L(S) → Ab
from L(S) to the category of abelian groups. This enables the application
of the (co)homology theory of categories to be applied directly to the
study of inverse semigroups: for example, Loganathan [14, Theorem 4.5(i)]
shows that free inverse monoids have cohomological dimension one. Recent
work of Webb [22, 23] on the cohomology of categories has developed
ideas from group representation theory, such as relation modules and
the Schur multiplier, and formulated them for categories. Webb’s results,
when applied to Loganathan’s category L(S), give us an entrée to the
ideas of augmentation and relation modules for inverse semigroups: the
starting point in [23] is the Gruenberg resolution, whose construction we
now recall.
Let P = 〈X : R〉 be a group presentation of a group G. Gruenberg
[6] (and see also [7, chapter 3]) gave a functorial construction from P , of
a free resolution of the trivial G–module Z. This Gruenberg resolution
has a number of interesting properties and we outline the construction,
referring to [6, 7] for details.
Let F = F (X) be the free group on X, let N be the normal closure of
R in F (X), and let θ : F → G be the natural map with kernel N . Then θ
induces a map θ : ZF → ZG of integral group rings, and we let r be its
kernel. The augmentation ideal f of F is the kernel of the augmentation
map ZF → Z induced by mapping w 7→ 1 for all w ∈ F : clearly r and f
are two-sided ideals.
Theorem (Gruenberg, [6]). The complex of ZG–modules
. . .→ r2/r3 → fr/fr2 → r/r2 → f/fr → ZG→ Z → 0
is a G–free resolution of Z, and this construction gives a functor from the
category of free presentations of G to the category of G–free resolutions
of Z.
Gruenberg [8] went on to show that his resolution gave rise to explicit
formulae for the homology groups of G in terms of r and f, generalising
the Hopf formula for H2(G,Z).
The kernel of the map f/fr → ZG is r/fr and Gruenberg [6] shows
that this is isomorphic to the relation module Nab. (Indeed, more is true,
as shown in [6]: the kernel of frn−1/frn → rn−1/rn is isomorphic to the
tensor product Nab⊗· · ·⊗Nab of n copies of the relation module, with the
diagonal G–action.) Defining the relation module in this way permits the
introduction of the concept in other algebraic settings where the operation
N. D. Gilbert 3
of abelianisation has no obvious counterpart.
For a small category C, the integral category algebra is the free abelian
group ZC having the morphisms of C as a basis. Multiplication is defined
by the Z–linear extension of the composition of morphisms in C, with
undefined compositions being set equal to zero. An inverse semigroup S
therefore determines the category algebra ZL(S). However, Loganathan
defines ZS as a L(S)–module determined by the free abelian groups on
Green’s L–classes in S (see section 2 below). Beginning with Loganathan’s
definition leads to definitions of the augmentation ideal and relation
modules differing from those that result from following the constructions
of [23] for the category algebra ZL(S).
We shall begin with Loganathan’s ZS, and blend some of the ideas of
[14] with Webb’s approach to the cohomology of categories, so that we are
able to define the relation module of an inverse semigroup presentation.
The key ingredient is the construction of an L(S)–module D that corre-
sponds to the module f/fr in Gruenberg’s resolution. Our construction
generalises Crowell’s derived module of a group homomorphism [3], and
is based on the derived module of a morphism of groupoids as defined
by Brown and Higgins in [2]. After some preliminaries on Loganathan’s
category L(S) and L(S)–modules in section 2, we define the derivation
module Dφ of an inverse semigroup homomorphism φ in section 3 and
establish its adjoint relationship to the semidirect products of S with
L(S)–modules. For subsequent applications, the most important result of
section 3 is that if S is an inverse monoid presented by 〈X : R〉, and if
θ : FIM(X) → S is the presentation map from the free inverse monoid on
X to S, then the derivation module Dθ is a projective L(S)–module.
There is a canonical L(S)–map from Dθ to the augmentation module
of S, and its kernel is defined to be the relation module of the presentation
〈X : R〉. We show that the set of relations R gives rise to a natural
generating set for the relation module, and show that the relation module
may also be interpreted as the first homology group of the Schützenberger
graph of S with respect to the generating set X. This naturally leads to
consideration of the class of inverse monoids whose Schützenberger graph
is a forest: we call such an inverse monoid arboreal. The analogue of the
Gruenberg resolution in this case shows that an arboreal inverse monoid
has cohomological dimension one.
2. Modules and augmentation
Let S be an inverse semigroup, with semilattice of idempotents E(S).
Recall that the natural partial order on S may be defined as follows: for
s, t ∈ S we have s 6 t if only if s = et for some e ∈ E(S). We have the
following useful alternative characterisations of the natural partial order
(see [11, Lemma 1.4.6] for example):
4 Derivations and relation modules
Lemma 2.1. Let S be an inverse semigroup. Then the following are
equivalent for s, t ∈ S:
• s 6 t,
• s = tf for some f ∈ E(S),
• s = ss−1t,
• s = ts−1s.
Loganathan’s category L(S) is now defined as follows. Its set of objects
is E(S), and the set of arrows is {(e, s) : e ∈ E(S), s ∈ S, e > ss−1}. The
identity arrow at e ∈ E(S) is (e, e), and the arrow (e, s) has domain
d(e, s) = e and range r(e, s) = s−1s. If (e, s), (f, t) ∈ L(S) and s−1s = f
then their composite is (e, s)(f, t) = (e, st). It is now straightforward to
check that L(S) is a left-cancellative category.
An L(S)–module is then a functor A : L(S) → Ab to the category
Ab of abelian groups. Modules for the category algebra ZL(S), as defined
in [22], will be L(S)–modules in our sense, but the converse will be true
if and only if E(S) is finite. An L(S)–module A determines an abelian
group Ae for each e ∈ E(S) and a homomorphism α(e,s) : Ae → As−1s
for each arrow (e, s), such that α(e,e) is the identity map on Ae, and such
that α(e,s)α(f,t) = α(e,st) whenever s−1s = f . If A is an L(S)–module and
a ∈ Ae, we shall write a⊳ (e, s) for aα(e,s) ∈ As−1s.
We may regard the poset E(S) as a category in the usual way, with a
unique morphism (e, f) : e→ f whenever e 6 f in E(S). Then E(S)op is
a subcategory of L(S), and the restriction of an L(S)–module to E(S)op
determines a presheaf of abelian groups A over E(S). Hence an L(S)–
module A may be considered as a presheaf of abelian groups over E(S)
with an S–action. Moreover, for any L(S)–module A, the disjoint union
A⊔ = ⊔e∈E(S)Ae is a commutative inverse semigroup, with the product
a ∗ b of a ∈ Ae and b ∈ Af defined by a ∗ b = (a⊳ (e, ef)) + (b⊳ (f, ef)).
The natural partial order on A⊔ then coincides with that induced by the
action of E(S)op: for a ∈ Ae and b ∈ Af ,
a 6 b⇐⇒ e 6 f and a = b⊳ (f, e).
Moreover, A⊔ also admits a semigroup action of S by endomorphisms,
defined for a ∈ Ae and s ∈ S by a · s = a⊳ (e, es).
A homomorphism of L(S)–modules is just a natural transformation
of functors. We call such a natural transformation an L(S)–map. Hence
an L(S)–map µ : A → B is determined by a family of abelian group
homomorphisms µe : Ae → Be indexed by E(S), such that the following
N. D. Gilbert 5
square commutes:
Ae
⊳(e,s)
��
µe
// Be
⊳(e,s)
��
As−1s µ
s−1s
// Bs−1s
We denote the category of L(S)–modules and L(S)–maps by ModS . The
kernel of an L(S)–map µ is the L(S)–module kerµ defined by (kerµ)e =
ker(µe : Ae → Be). Clearly kerµ is then an L(S)–submodule of A.
A fixed abelian group A determines a trivial L(S)–module A, with
A e = A for all e ∈ E(S), and α(e,s) = id for all (e, s). In particular, the
group of integers determines the trivial module Z (denoted by ∆Z in [14]).
Given an L(S)–module A we can define the semidirect product S ⋉A,
which will again be an inverse semigroup. The construction is a special
case of both Billhardt’s λ–semidirect product of inverse semigroups [1]
and of Steinberg’s semidirect product of ordered groupoids, and so as an
instance of [18, proposition 3.3] we have the following result.
Proposition 2.2. Let S be an inverse semigroup and A an L(S)–module.
Define
S ⋉A = {(s, a) : s ∈ S, a ∈ As−1s}.
Then S ⋉A is an inverse semigroup, with multiplication
(s, a)(t, b) = (st, (a⊳ (s−1s, s−1st)) + (b⊳ (t−1t, t−1s−1st)))
= (st, (a · t) ∗ b).
Clearly the projection (s, a) 7→ s is a homomorphism of inverse semi-
groups, and the construction A → S⋉A is a functor from the category of
L(S)–modules to the slice category (IS ↓ S) of inverse semigroups over
S. In section 3 we shall construct a left adjoint, modifying a construc-
tion originally due to Crowell [3] and then generalised in [2]. This left
adjoint will then give us the term corresponding to f/fr in the Gruenberg
resolution.
Now Loganathan [14] constructs the L(S)–module ZS as follows. For
each idempotent e ∈ E(S), let Le be the L–class of e and let ZSe be the
free abelian group with basis Le. Now if a ∈ Le and (e, s) ∈ L(S) we
define a⊳ (e, s) = as. Since e = a−1a > ss−1, it follows that (as)−1(as) =
s−1a−1as = s−1s, so that as ∈ Ls−1s and the basis transformation a 7→ as
induces a homomorphism ZSe → ZSs−1s. Loganathan notes [14, Remark
4.2] that if S is an inverse monoid then ZS is a free L(S)–module. In this
setting, a basis for a free L(S)–module is an E(S)–set, that is a family of
disjoint sets Xe indexed by e ∈ E(S).
The augmentation map εS : ZS → Z is the L(S)–map defined on the
basis Le of ZSe by s 7→ 1 ∈ Ze. It is clear that this is an L(S)–map, and
6 Derivations and relation modules
its kernel is the augmentation module s of S.
Lemma 2.3. For each e ∈ E(S), the abelian group se is freely generated
by the elements s− e with e 6= s ∈ Le.
Proof. Let x =
∑
i∈I nisi ∈ se, so that each si ∈ Le. Since xεS = 0, we
have
∑
i∈I ni = 0 and hence
∑
i∈I nie = 0. Then
x =
∑
i∈I
nisi −
∑
i∈I
nie =
∑
i∈I
ni(si − e).
Hence the elements s− e with s ∈ Le generate se, and since ZSe is freely
generated by Le, it is clear that the elements s− e with e 6= s ∈ Le are a
basis.
We remark that essentially the same definitions of ZG and its aug-
mentation module were given for a groupoid G by Brown and Higgins
[2]. In the next section we shall adapt another definition for groupoids
from [2] in order to define the relation module of an inverse semigroup
presentation.
3. The derivation module of a homomorphism
The second step in Gruenberg’s resolution is the module f/fr, which is
G–free on the cosets of the elements 1 − x , (x ∈ X). It is therefore
isomorphic to the module f ⊗F ZG, which is Crowell’s derived module
of the homomorphism θ : F → G: indeed, Crowell gives the short exact
sequence
0 → Nab → f⊗F ZG→ g → 0
in [3]. We shall adapt a generalisation of Crowell’s derived module due
to Brown and Higgins [2] and apply it to a homomorphism of inverse
semigroups. From a homomorphism φ : T → S of inverse semigroups, we
shall construct an L(S)–module Dφ, called the derivation module of φ,
with a universal property detailed in Proposition 3.5, and with a canonical
L(S)–map to the augmentation module of S.
Let A be an L(S)–module. We first define the notion of a (L(T ), φ)–
derivation from L(T ) to A, following the definition of a derivation to
a module for a category algebra given in [22]. The properties of the
composition of an (L(T ), φ)–derivation with the map T → L(T ) that
carries t 7→ (tt−1, t) will then lead us to the formulation of the notion of a
φ–derivation. Given a functor φ : L(T ) → L(S), we define an (L(T ), φ)–
derivation to A to be a function ζ : L(T ) → A⊔, defined on the arrows of
L(T ), such that
• if (p, a) is an arrow of L(T ) then (p, a)ζ ∈ A(a−1a)φ,
N. D. Gilbert 7
• if (p, a), (a−1a, b) are composable arrows in L(T ), so that a−1a >
bb−1, then
(p, ab)ζ = ((p, a)(a−1a, b))ζ = (p, a)ζ ⊳ (a−1a, b)φ+ (a−1a, b)ζ.
Now let φ be a homomorphism of inverse semigroups. Of course, φ
induces a functor L(T )
φ
→ L(S) mapping (p, a) 7→ (pφ, aφ). A φ–derivation
η : T → A is a function T → A⊔ such that
• if a ∈ T then aη ∈ A(a−1a)φ,
• if a, b ∈ T and a−1a > bb−1, then
(ab)η = aη ⊳ ((a−1a)φ, bφ) + bη. (3.1)
Example 3.1. If φ : T → S is a homomorphism of inverse semigroups
then the function η : T → s, t 7→ tφ − (t−1t)φ is a φ–derivation. If
t−1t > uu−1 then
(tu)η = (tu)φ− (u−1u)φ
= (tφ)(uφ) − uφ+ uφ− (u−1u)φ
= (tφ− (t−1t)φ) ⊳ ((t−1t)φ, uφ) + uφ− (u−1u)φ
= (tη) ⊳ ((t−1t)φ, uφ) + uη.
For each e ∈ E(S) we define
Xe = {(a, s) : a ∈ T, s ∈ Le, (a
−1a)φ > ss−1}.
We note that if (a, s) ∈ Xe then (aφ)s ∈ Le. Now the group Dφ,e is defined
as the (additive) abelian group generated by Xe subject to all relations of
the form
(ab, s) − (b, s) = (a, (bφ)s) (3.2)
where a, b ∈ T with a−1a > bb−1 and s ∈ S. We denote the image of (a, s)
in Dφ,e by 〈a, s〉. For subsequent use, we note the following consequences
of the relations in Dφ,e.
Lemma 3.2. (a) If (a, s) ∈ Xe then 〈aa−1, s〉 = 0 in Dφ,e.
(b) If a, b ∈ T with b > a, and s ∈ S with (a−1a)φ > ss−1, then
〈a, s〉 = 〈b, s〉 in Dφ,e.
Proof. For (a) we have
0 = 〈a, s〉 − 〈a, s〉 = 〈aa−1a, s〉 − 〈a, s〉
= 〈a, (a−1a)φs〉 + 〈a−1a, s〉 − 〈a, s〉
8 Derivations and relation modules
= 〈a, s〉 + 〈a−1a, s〉 − 〈a, s〉
= 〈a−1a, s〉.
For (b), we note that a = ba−1a with b−1b > a−1a, and therefore
〈a, s〉 = 〈ba−1a, s〉 = 〈b, (a−1a)φs〉 + 〈a−1a, s〉 = 〈b, s〉.
Lemma 3.3. Let s ∈ S with s−1s = e and suppose that 〈a, s〉 ∈ Dφ,e. If
b ∈ T and bφ = s then 〈ab, e〉 ∈ Dφ,e and
〈ab, e〉 = 〈a, s〉 + 〈b, e〉 .
Proof. Set f = (a−1a)φ: then f > ss−1 and so
((ab)φ)−1(ab)φ = s−1fs = s−1s = e.
Therefore 〈ab, e〉 ∈ Dφ. Since (a−1ab)(a−1ab)−1 = a−1abb−1 6 a−1a we
have
〈ab, e〉 = 〈a(a−1ab), e〉 = 〈a, (a−1ab)φ〉 + 〈a−1ab, e〉 (by (3.2))
= 〈a, fs〉 + 〈a−1ab, e〉
= 〈a, s〉 + 〈b, e〉
since fs = s and, by Lemma 3.2(b), 〈a−1ab, e〉 = 〈b, e〉.
Lemma 3.4. Suppose that T is generated as an inverse semigroup by a
subset X. Then Dφ,e is generated, as an abelian group, by the subset
{〈x, s〉 : x ∈ X, (x−1x)φ > ss−1, s−1s = e}.
Proof. Suppose that t ∈ T with t = ua for some u ∈ T and a ∈ X ∪X−1.
Then t = u(u−1ua) and u−1u > u−1uaa−1 = (u−1ua)(u−1ua)−1. The
relations (3.2) then imply that, if (t, s) ∈ Xe , then
〈t, s〉 = 〈u, ((u−1ua)φ)s〉 + 〈u−1ua, s〉.
By part (b) of Lemma 3.2, we have 〈u−1ua, s〉 = 〈a, s〉 and so
〈t, s〉 = 〈u, ((u−1ua)φ)s〉 + 〈a, s〉.
It now follows, by induction on the minimum length of a product of
elements of X ∪ X−1 representing t ∈ T , that Dφ,e is generated as an
abelian group by the subset {〈x, s〉 : x ∈ X ∪X−1}. But, by part (a) of
Lemma 3.2 and the relations (3.1), for any (a, s) ∈ Xe,
0 = 〈aa−1, s〉 = 〈a, (a−1φ)s〉 + 〈a−1, s〉
N. D. Gilbert 9
and so 〈x−1, s〉 = −〈x, (x−1φ)s〉.
We may now define the derivation module of φ, denoted by Dφ. For
e ∈ E(S) we have (Dφ)e = Dφ,e , and the action of (e, t) on a generator
〈a, s〉 of Dφ,e is given by 〈a, s〉 ⊳ (e, t) = 〈a, st〉. Since s−1s = e > tt−1,
(st)−1(st) = t−1s−1st = t−1t, and so 〈a, st〉 ∈ Dφ,t−1t . It is easy to check
that Dφ is indeed an L(S)–module.
Proposition 3.5. There exists a canonical φ–derivation δ : T → Dφ such
that, given any φ–derivation η : T → A to an L(S)–module A, there is a
unique L(S)–map ξ : Dφ → A such that η = δξ.
Proof. We define the map δ : T → Dφ by t 7→ 〈t, (t−1t)φ〉. Then tδ ∈
Dφ,(t−1t)φ and δ clearly satisfies the defining property for a φ–derivation
given in (3.1).
Let A be an L(S)–module. Suppose that an L(S)–map ξ : Dφ → A
satisfies η = δξ, and consider 〈a, s〉 ∈ D⊔
φ . Then (a−1a)φ > ss−1 and
〈a, s〉ξ = (〈a, (a−1a)φ〉⊳ ((a−1a)φ, s))ξ
= (〈a, (a−1a)φ〉)ξ ⊳ ((a−1a)φ, s)
= (aδ)ξ ⊳ ((a−1a)φ, s)
= aη ⊳ ((a−1a)φ, s) ,
so ξ is uniquely determined by η.
Given η : T → A, we may define a map ξ : Dφ → A by setting
〈a, s〉ξ = aη ⊳ ((a−1a)φ, s). Firstly, this will be well-defined on D⊔
φ since,
if a, b ∈ T with a−1a > bb−1, then
(〈ab, s〉 − 〈b, s〉)ξ = ((ab)η ⊳ ((b−1b)φ, s)) − (bη ⊳ ((b−1b)φ, s))
= ((ab)η − bη) ⊳ ((b−1b)φ, s)
= (aη ⊳ ((bb−1)φ, bφ)) ⊳ ((b−1b)φ, s)
= aη ⊳ ((bb−1)φ, (bφ)s)
= 〈a, (bφ)s〉ξ ,
and is an L(S)–map since
(〈a, s〉⊳ (e, x))ξ = 〈a, sx〉ξ = aη ⊳ ((a−1a)φ, sx)
= aη ⊳ ((a−1a)φ, s) ⊳ (e, x)
= 〈a, s〉ξ ⊳ (e, x).
Corollary 3.6. Given any inverse semigroup homomorphism φ : T → S,
there is an L(S)–map ∂1 : Dφ → ZS whose image is the augmentation
module s.
10 Derivations and relation modules
Proof. The map ∂1 is induced by the φ–derivation t 7→ tφ − (t−1t)φ of
Example 3.1, and so 〈a, s〉∂1 = (aφ− (a−1a)φ) ⊳ s = (aφ)s− s. Lemma
2.3 then implies that its image is s.
Suppose that we have inverse semigroup homomorphisms φ : T → S
and ψ : U → S and that λ : T → U is a homomorphism of inverse
semigroups such that φ = λψ. Then λ is a morphism in the slice category
(IS ↓ S) and induces a mapping λ∗ : Dφ → Dψ given by 〈a, s〉 7→ 〈aλ, s〉.
In this way the construction of the derivation module gives a functor
(IS ↓ S) → ModS .
Proposition 3.7. The derivation module functor (T
φ
→ S) 7→ Dφ is left
adjoint to the semidirect product functor A 7→ S ⋉A.
Proof. The first part of the proof is the standard verification, adapted
to our setting, of the correspondence between derivations and maps to
semidirect products. The details are as follows. Let φ : T → S be an
inverse semigroup homomorphism and let α : T → S ⋉A be a morphism
in the slice category (IS ↓ S), so that for all t ∈ T we have tα = (tφ, tζ)
for some function ζ : T → A⊔. Since (tφ, tζ) ∈ S ⋉ A, it follows that
tζ ∈ A(t−1t)φ.
Now assume that a, b ∈ T with a−1a > bb−1. Then
(aα)(bα) = (aφ, aζ)(bφ, bζ)
= (aφbφ, aζ ⊳ ((a−1a)φ, (a−1ab)φ)
+ bζ ⊳ ((b−1b)φ, (b−1a−1ab)φ))
and, since a−1ab = b,
= (aφbφ, aζ ⊳ ((a−1a)φ, bφ) + bζ ⊳ ((b−1b)φ, (b−1b)φ))
= (aφbφ, aζ ⊳ ((a−1a)φ, bφ) + bζ).
Now (aα)(bα) = (ab)α = ((ab)φ, (ab)ζ), and so if a−1a > bb−1, we have
(ab)ζ = aζ ⊳ ((a−1a)φ, bφ) + bζ
and ζ is a φ-derivation.
By Proposition 3.5, ζ induces a unique L(S)–map α̂ : Dφ → A defined
by
α̂ : 〈a, s〉 7→ aζ ⊳ ((a−1a)φ, s) . (3.3)
Now an L(S)–map γ : Dφ → A induces an inverse semigroup homo-
morphism γ† : T → S ⋉A in the slice category (IS ↓ S) given by
γ† : t 7→ (tφ, 〈t, (t−1t)φ〉γ). (3.4)
N. D. Gilbert 11
We check that γ† is indeed an inverse semigroup homomorphism as follows.
For all a, b ∈ T we have
(ab)γ† = ((ab)φ, 〈ab, (b−1a−1ab)φ〉γ).
Now we compute (aγ†)(bγ†):
(aγ†)(bγ†) = (aφ, 〈a, (a−1a)φ〉γ)(bφ, 〈b, (b−1b)φ〉γ)
= (aφbφ, 〈a, (a−1a)φ〉γ ⊳ ((a−1a)φ, (a−1ab)φ)
+ 〈b, (b−1b)φ〉γ ⊳ ((b−1b)φ, (b−1a−1ab)φ))
= (aφbφ, 〈a, (a−1ab)φ〉γ + 〈b, (b−1a−1ab)φ〉γ)
= (aφbφ, (〈a, (a−1ab)φ〉 + 〈b, (b−1a−1ab)φ〉)γ).
Now it follows from part (b) of Lemma 3.2 that
〈b, (b−1a−1ab)φ〉 = 〈a−1ab, (b−1a−1ab)φ〉
and then by Lemma 3.3 that
〈a, (a−1ab)φ〉 + 〈a−1ab, (b−1a−1ab)φ〉 = 〈ab, (b−1a−1ab)φ〉.
Hence
(aγ†)(bγ†) = (aφbφ, 〈ab, (b−1a−1ab)φ〉γ)
and so γ† is an inverse semigroup homomorphism.
The correspondences α 7→ α̂ (defined in (3.3)) and γ 7→ γ† (defined
in (3.4)) are mutually inverse bijections, giving a natural isomorphism
between the bifunctors ModS(D−,−) and (IS ↓ S)(−, S ⋉−) confirming
the adjunction given in the Proposition.
Theorem 3.8. Let S be an inverse monoid generated by a set X, let
FIM(X) be the free inverse monoid on X, and let θ : FIM(X) → S be the
presentation map.
(a) Dθ,e is generated, as an abelian group, by the subset
{〈x, s〉 : x ∈ X, (x−1x)θ > ss−1, s−1s = e}
and hence Dθ is generated as an L(S)–module by the set
{〈x, (x−1x)θ〉 : x ∈ X}.
(b) Dθ is a projective L(S)–module.
Proof. (a) This follows from Lemma 3.4.
12 Derivations and relation modules
(b) Let µ : A → B be an epimorphism of L(S)–modules. For any
e ∈ E(S), the evaluation functor evale : ModS → Ab given by A 7→ Ae is
exact, and hence µ : Ae → Beµ is surjective.
Suppose that β : Dθ → B is an L(S)–map. By Proposition 3.7, β
induces a inverse monoid homomorphism β† : FIM(X) → S ⋉ B, and
there is a surjection µ∗ : S ⋉ A → S ⋉ B of inverse monoids, given by
(s, a) 7→ (s, aµ). The freeness of FIM(X) then implies that we can lift β†
to α† : FIM(X) → S ⋉A, with ᆵ∗ = β†. For w ∈ FIM(X), it follows
that wα† = (wθ,wη), where η : FIM(X) → A is a θ–derivation such
that wηµ = 〈w, (w−1w)θ〉β. By Proposition 3.5 η induces an L(S)–map
α : Dθ → A defined by
α = (̂α†) : 〈w, s〉 7→ wη ⊳ ((w−1w)θ, s),
and
〈w, s〉αµ = (wη ⊳ ((w−1w)θ, s))µ
= wηµ⊳ ((w−1w)θ, s)
= 〈w, (w−1w)θ〉β ⊳ ((w−1w)θ, s)
= (〈w, (w−1w)θ〉⊳ ((w−1w)θ, s))β
= 〈w, s〉β.
This shows that Dθ is projective.
Proposition 3.7 is based on the adjunction result for groupoid actions
given in [2], but may also be seen as a special case of a result of Nico
[16] for the semidirect product of categories. This work is elaborated by
Steinberg and Tilson in [20], and by Steinberg in [18] for the category of
ordered groupoids. The results of [18] are of particular relevance since the
category of inverse semigroups is isomorphic to the subcategory of inductive
groupoids in the category of ordered groupoids. Steinberg constructs a left
adjoint Der(φ) to the semidirect product, where the latter is regarded as
a functor from the category of ordered groupoid actions to the category of
ordered functors: from a pair (G,H) of ordered groupoids withH acting on
G, the semidirect product H ⋉G is an ordered groupoid with a projection
map H ⋉G→ H. We have restricted attention to L(S)–modules, and so
have replaced the category of ordered groupoid actions by ModS , and the
category of ordered functors by the slice category (IS ↓ S). Steinberg’s
adjunction then tells us that morphisms T → S⋉A in (IS ↓ S) correspond
bijectively to a certain subclass of morphisms Der(φ) → A in the category
of ordered groupoid actions. Following through the details of Steinberg’s
construction, it is readily seen that the subclass of morphisms are those
that arise from φ–derivations T → A.
N. D. Gilbert 13
4. Presentations and relation modules
Let S be an inverse semigroup given by a presentation P = Inv[X : R],
and let θ : FIS(X) → S be the associated presentation map. Corollary 3.1
gives an L(S)–map ∂1 : Dθ → s: its kernel M is an L(S)–module called
the relation module of P . Our first task is to find a convenient generating
set for the relation module: we approach this in stages in our next result.
Theorem 4.1. (a) The relation module Me at e ∈ E(S) is generated,
as an abelian group, by all elements of the form 〈u, s〉− 〈v, s〉, where
uθ = vθ and s−1s = e.
(b) A smaller generating set for Me, as an abelian group, is given by
the set of all elements of the form 〈pr1q, s〉 − 〈pr2q, s〉, where p, q ∈
FIS(X), (r1, r2) ∈ R and s−1s = e.
(c) The relation module M is generated as an L(S)–module by the set of
all elements of the form 〈r1, e〉−〈r2, e〉 where (r1, r2) ∈ R, e ∈ E(S),
and (r−1
1 r2)θ = e.
Proof. (a) Let Pe be the subgroup of Me generated by the set of elements
specified in part (a) of the theorem. Now suppose that α =
∑
i∈I εi〈ui, si〉
is an element of the relation module Me at e (where εi = ±1), so that
α∂1 =
∑
i∈I
εi((uiθ)si − si) = 0. (4.1)
Since ZS is (additively) free on the elements of the L–class Le, the terms
in the sum in (4.1) must cancel in pairs. We fix such a pairing of cancelling
terms, and for each si occurring in the given expression for α we choose
wsi ∈ FIS(X) with wsiθ = si.
Suppose that in our pairing of cancelling terms, (uiθ)si is paired with
si. Then (uiθ)si = si and, by Lemma 3.2, 〈ui, si〉 = 〈uiwsiw
−1
si
, si〉 and
〈uiwsiw
−1
si
, si〉 = 0. It follows that
〈ui, si〉 = 〈uiwsiw
−1
si
, si〉 − 〈uiwsiw
−1
si
, si〉 ∈ Pe.
Now suppose that (uiθ)si is paired with (ujθ)sj : that is (uiθ)si =
(ujθ)sj and −εi = εj . Then by Lemma 3.3,
〈uiwsi , e〉 = 〈ui, si〉 + 〈wsi , e〉 and 〈ujwsj , e〉 = 〈uj , sj〉 + 〈wsj , e〉,
with 〈uiwsi , e〉 − 〈ujwsj , e〉 ∈ Pe. Therefore, we may write
εi〈ui, si〉+ εj〈uj , sj〉 = εi(〈uiwsi , e〉 − 〈ujwsj , e〉)− εi〈wsi , e〉 − εj〈wsj , e〉,
14 Derivations and relation modules
and modulo Pe, we see that εi〈ui, si〉+ εj〈uj , sj〉 is equal to −εi〈wsi , e〉 −
εj〈wsj , e〉.
Finally, if (ui)θsi is paired with sj , then 〈uiwsi , e〉 − 〈wsj , e〉 ∈ Pe and,
again by Lemma 3.3,
〈ui, si〉 = (〈uiwsi , e〉 − 〈wsj , e〉) + 〈wsj , e〉 − 〈wsi , e〉.
These considerations show that modulo Pe, the element α ∈ Me is
equal to a sum β =
∑
kmk〈wsk , e〉 with mk ∈ Z and the sk distinct. But
then β∂1 =
∑
kmk(sk− e) = 0 is a sum of Z–independent elements of ZS,
and so each mk = 0. Therefore β = 0, and we conclude that α is in Pe.
(b) Let Qe be the subgroup of Me generated by the set of elements
specified in part (b) of the theorem. Let 〈u, s〉 − 〈v, s〉, as in part (a), be a
generator of Pe. Now S is presented by Inv[X : R] and so, since uθ = vθ,
there exists a finite sequence of elements w0, w1, . . . , wm ∈ FIM(X) with
w0 = u, wm = v and such that for each i, 0 6 i < m, we may write
wi = piriqi and wi+1 = pir
′
iqi with pi, qi ∈ FIM(X) and with either (ri, r
′
i)
or (r′i, ri) in R. We show by induction on m that the element 〈u, s〉−〈v, s〉
of Dθ is in the subgroup Qe. If m = 1 then u = p0r0q0 and v = p0r
′
0q0
and so
〈u, s〉 − 〈v, s〉 = 〈p0r0q0, s〉 − 〈p0r
′
0q0, s〉 ∈ Qe.
If m > 1 we have u = p0r0q0 and w1 = p0r
′
0q0
〈u, s〉 − 〈v, s〉 = 〈p0r0q0, s〉 − 〈p0r
′
0q0, s〉 + 〈w1, s〉 − 〈v, s〉.
Now we have w1θ = vθ and the elements w1, v are linked by the sequence
w1, . . . , wm ∈ FIM(X), and by induction 〈w1, s〉 − 〈v, s〉 ∈ Qe. It follows
that 〈u, s〉 − 〈v, s〉 ∈ Qe.
(c) We show that a generator 〈pr1q, s〉− 〈pr2q, s〉 is an L(S)–translate
of the element 〈r1, f〉 − 〈r2, f〉, where f = (r−1
1 r2)θ. Now by Lemma 3.3,
〈pr1q, s〉 = 〈p(p−1p)r1q, s〉 = 〈p, (p−1pr1q)θs〉 + 〈p−1pr1q, s〉
and similarly
〈pr2q, s〉 = 〈p, (p−1pr2q)θs〉 + 〈p−1pr2q, s〉 .
Now (p−1pr1q)θ = (p−1pr2q)θ, and so
〈pr1q, s〉 − 〈pr2q, s〉 = 〈p−1pr1q, s〉 − 〈p−1pr2q, s〉 = 〈r1q, s〉 − 〈r2q, s〉
by part (b) of Lemma 3.2. We see that 〈pr1q, s〉−〈pr2q, s〉 does not depend
on p. In the same manner,
〈r1q, s〉 = 〈r1(r
−1
1 r1q), s〉 = 〈r1, (r
−1
1 r1q)θs〉 + 〈r−1
1 r1q, s〉
N. D. Gilbert 15
= 〈r1, (r
−1
1 r1q)θs〉 + 〈q, s〉
since 〈r−1
1 r1q, s〉 = 〈q, s〉 by part (b) of Lemma 3.2. It follows that
〈r1q, s〉 − 〈r2q, s〉 = 〈r1, (r
−1
1 r1q)θs〉 − 〈r2, (r
−1
2 r2q)θs〉 = 〈r1, s
′〉 − 〈r2, s
′〉
where s′ = (r−1
1 r1q)θs = (r−1
2 r2q)θs. Then finally we have
〈r1, s
′〉 − 〈r2, s
′〉 = (〈r1, e〉 − 〈r2, e〉) ⊳ (e, s′).
Corollary 4.2. Let S be an inverse monoid presented by Inv[X : R] with
presentation map θ : FIM(X) → S. Then there is an exact sequence of
L(S)–modules
F(R) → Dθ → ZS → Z → 0
where F(R) is the free L(S)–module on the E(S)–set X = {Xe : e ∈
E(S)}, with Xe = {(r1, r2) ∈ R : (r−1
1 r2)θ = e}.
4.1. The Schützenberger representation
Suppose that S is an inverse monoid presented by Inv[X : R] with pre-
sentation map θ : FIM(X) → S. The (left) Schützenberger graph of S
with respect to X, denoted by SchL(S,X), has vertex set S and, for each
(x, s) ∈ X × S with (x−1x)θ > ss−1, has a directed edge labelled by (x, s)
with initial vertex s and terminal vertex (xθ)s. Equivalently, there is a
directed edge from s to xs labelled by (x, s) whenever sL(xθ)s in S. The
connected components of SchL(S,X) are the full subgraphs spanned by
the L–classes of S, and the connected component containing e ∈ E(S) is
denoted by SchL(S,X, e).
The usual convention for Schützenberger graphs [19, 21] is to have a
directed edge (s, x) from s to sx whenever sRsx. To get a natural right
L(S)–module, we have to use the left Schützenberger graphs.
Suppose that s−1s = e and that (e, t) ∈ L(S). Then by defining
s⊳ (e, t) = st and (a, s) ⊳ (e, t) = (a, st)
we get a graph map SchL(S,X, e) → SchL(S,X, t−1t) and in this way
we obtain a functor from L(S) to the category of directed graphs. The
cellular chain complex CL(S,X) of SchL(S,X) is then a complex of L(S)–
modules, with CL
0 (S,X) = ZS and with the group CL
1 (S,X)e free abelian
on the set
{(x, s) : x ∈ X, s ∈ S, (x−1x)θ > ss−1, s−1s = e}.
The boundary map CL
1 (S,X)e → CL
0 (S,X)e maps (x, s) 7→ (xθ)s− s.
Theorem 4.3. Suppose that S is an inverse monoid presented by Inv[X :
16 Derivations and relation modules
R] with presentation map θ : FIM(X) → S. Then the derivation module
Dθ is isomorphic to CL
1 (S,X) and the relation module M is isomorphic
to the first homology group of the Schützenberger graph SchL(S,X).
Proof. By Corollary 3.6 we have a commutative square of L(S)–maps
CL
1 (S,X)
κ
��
��
// CL
0 (S,X)
∼=
��
Dθ
∂1
// ZS
in which the left-hand map κ is the L(S)–map induced (as a homo-
morphism of abelian groups) by (x, s) 7→ 〈x, s〉, and κ restricts to a
surjection H1(SchL(S,X)) → M. Now the function X → S ⋉ CL
1 (S,X)
that maps x 7→ (xθ, (x, (x−1x)θ) induces an inverse monoid homomor-
phism FIM(X) → S ⋉ CL
1 (S,X) and by Proposition 3.7 there is then an
L(S)–map Dθ → CL
1 (S,X) mapping 〈x, (x−1x)θ〉 7→ (x, (x−1x)θ) giving
an inverse to κ.
4.2. Cohomological dimension
The classification of all small categories of cohomological dimension zero
was completed by Laudal [9], following a conjecture of Oberst [17]. Lau-
dal’s result was then used by Leech [13] to classify the inverse monoids
of cohomological dimension zero: the classification is the expected gener-
alisation of the result that a group has cohomological dimension zero if
and only if it is trivial. We present another proof here that makes direct
use of the fact that, if S has cohomological dimension zero, then Z is a
projective L(S)–module.
Theorem 4.4 (Leech [13]). An inverse monoid S has cohomological
dimension zero if and only if it is a semilattice.
Proof. If S is a semilattice then ZS = Z and so Z is a projective L(S)–
module, and hence cd(S) = 0.
Conversely, if cd(S) = 0 then Z has a projective resolution P0 → Z
and hence Z is a projective L(S)–module. In particular, the augmentation
map ε : ZS → Z splits as an L(S)–map, and so there is an L(S)–map
σ : Z → ZS such that σε = id. Let 1e denote the copy of 1 ∈ Z in Ze
(that is, in the copy of Z indexed by e ∈ E(S)) and let 1̂ = 11. We set
1̂σ = w ∈ ZS1. Then wε = 1̂ and w determines σ since
1eσ = (1̂ ⊳ (1, e))σ = w ⊳ (1, e) = we ∈ ZLSe .
N. D. Gilbert 17
Now for any s ∈ S, we have
ws = w ⊳ (1, s) = (1̂σ) ⊳ (1, s) = (1̂ ⊳ (1, s))σ = 1s−1sσ = ws−1s (4.2)
in the free abelian group ZSs−1s.
We may write w =
∑
t−1
i ti=1mtiti (with mti ∈ Z and at most finitely
many non-zero). Suppose that s ∈ S with ss−1 = 1. Then for each ti with
mti 6= 0, there exists a tj with mtj 6= 0 such that tis = tjs
−1s. But then
tit
−1
i = (tis)(tis)
−1 = (tjs
−1s)(tjs
−1s)−1 = tjs
−1st−1
j 6 tjt
−1
j .
Since the collection of such ti is finite, there must exist tk occurring in w
with tkt
−1
k = tks
−1st−1
k , or equivalently with tk = tks
−1s. Then
1 = t−1
k tk = s−1st−1
k tk = s−1s .
We see that ss−1 = 1 = s−1s, and so Green’s L–class at 1 ∈ E(S) coincides
with the H–class H1, and is a group. In particular, ZS1 is the integral
group ring ZH1. Now for all h ∈ H1 we have, by (4.2),
∑
t−1
i ti=1
mtitih =
∑
t−1
i ti=1
mtiti .
Taking h = t−1
i tj in the group H1, we see that mti = mtj , and since∑
imti = 1, it follows that |H1| = 1. In particular w = 1. From (4.2) we
see, for any s ∈ S, that s = s−1s and so S is a semilattice.
We note the contrast between this result and the theory of monoids
of cohomological dimension zero. Laudal [9, Lemma A] shows that if
a monoid M has (left) monoid cohomological dimension equal to zero
then M has a right zero. See [5] for further results on the cohomological
dimension of monoids.
Arboreal inverse monoids
An arboreal inverse monoid S is an inverse monoid given by a presentation
of the form
S = Inv[X : ei = fi , (i ∈ I)]
where I is some indexing set and ei, fi are idempotents in FIM(X). There-
fore ei, fi are Dyck words in (X∪X−1)∗, that is words whose freely reduced
form is equal to 1. Free inverse monoids and free groups are arboreal inverse
monoids, as is the bicyclic monoid B, presented by Inv[x : xx−1 = 1].
Arboreal inverse monoids are the main concern of [15], wherein it
is shown that arboreal inverse monoids with generating set X are pre-
18 Derivations and relation modules
cisely the E–unitary quotients of FIM(X) with maximum group image
F (X), and are also characterised as the inverse monoids each of whose
Schützenberger graphs is a tree. This latter characterization has motivated
the use of the adjective ‘arboreal’. A principal result of [15] is that finitely
presented arboreal inverse monoids have decidable word problem. The
following result may be readily deduced from the work of Steinberg [19]
and also follows from Funk’s topos-theoretic approach to inverse semigroup
representations [4].
Theorem 4.5. An arboreal inverse monoid S has cohomological dimension
equal to 1.
Proof. Let S be presented by S = Inv[X : ei = fi , (i ∈ I)], and let
θ : FIM(X) → S be the presentation map. By part (c) of Theorem 4.1,
the relation module Me at e ∈ E(S) is generated by L(S)–translates of
elements of the form 〈ei, eiθ〉− 〈fi, fiθ〉 (where eiθ = fiθ ∈ S). By Lemma
3.2, all such elements are equal to 0. It follows that Me is trivial, and
hence that ∂1 : Dθ → s is injective. By [14, Remark 4.2], for a monoid
S the L(S)–module ZS is free, and by Corollary 3.8 we know that Dθ is
projective. Therefore the short exact sequence
0 → Dθ
∂1→ ZS
εS→ Z → 0
is a L(S)–projective resolution of Z.
References
[1] B. Billhardt, On a wreath product embedding and idempotent pure congruences
on inverse semigroups. Semigroup Forum 45 (1992) 45-54.
[2] R. Brown and P.J. Higgins, Crossed complexes and chain complexes with operators.
Math. Proc. Camb. Phil. Soc. 107 (1990) 33-57.
[3] R.H. Crowell, The derived module of a homomorphism. Adv. in Math. 6 (1971)
210-238.
[4] J. Funk, Semigroups and toposes. Semigroup Forum 75 (2007) 481-520.
[5] V.S. Guba and S.J Pride, On the left and right cohomological dimension of monoids.
Bull. London Math. Soc 30 (1998) 391-396.
[6] K.W. Gruenberg, Resolutions by relations. J. London Math. Soc. 35 (1960) 481-
494.
[7] K.W. Gruenberg, Cohomological Topics in Group Theory. Lect. Notes in Math.
143. Springer-Verlag (1970)
[8] K.W. Gruenberg, The universal coefficient theorem in the cohomology of groups.
J. London Math. Soc. 43 (1968) 239-241.
[9] O.A. Laudal, Note on the projective limit of small categories. Proc. Amer. Math.
Soc. 33 (1972) 307-309.
[10] H. Lausch, Cohomology of inverse semigroups. J. Algebra 35 (1975) 273-303.
[11] M.V. Lawson, Inverse Semigroups. World Scientific (1998).
N. D. Gilbert 19
[12] M.V Lawson, J. Matthews and T. Porter, The homotopy theory of inverse semi-
groups. Internat. J. Algebra Comput. 12 (2002) 755-790.
[13] J. Leech, The D–category of a semigroup. Semigroup Forum 11 (1975) 283-296.
[14] M.Loganathan, Cohomology of inverse semigroups. J. Algebra (1981).
[15] S.W. Margolis and J.E. Meakin, Inverse monoids, trees, and context-free languages.
Trans. Amer. Math. Soc. 335 (1993) 259-276.
[16] W.R. Nico, Wreath products and extensions. Houston J. Math. 9 (1983) 71-99.
[17] U. Oberst, Exactness of inverse limits. Bull. Amer. Math. Soc. 74 (1968) 1129-1132.
[18] B. Steinberg, Factorization theorems for morphisms of ordered groupoids and
inverse semigroups. Proc. Edin. Math. Soc. 44 (2001) 549-569.
[19] B. Steinberg, A topological approach to inverse and regular semigroups. Pacific J.
Math. 208 (2003) 367-396.
[20] B. Steinberg and B. Tilson, Categories as algebra, II. Internat. J. Algebra Comput.
6 (2003) 627-703.
[21] J.B Stephen, Presentations of inverse monoids. J. Pure Appl. Algebra 63 (1990)
81-112.
[22] P.J. Webb, An introduction to the representations and cohomology of categories.
In Group representation theory M. Geck et al. (Eds), EPFL Press Lausanne (2007)
149–173, .
[23] P.J. Webb, Resolutions, relation modules and Schur multipliers for categories.
J.Algebra 326 (2011) 245-276..
Contact information
N. D. Gilbert School of Mathematical and Computer Sciences
& the Maxwell Institute for Mathematical Sci-
ences, Heriot-Watt University, Edinburgh EH14
4AS, U.K.
E-Mail: N.D.Gilbert@hw.ac.uk
URL: www.ma.hw.ac.uk/∼ndg
Received by the editors: 23.07.2010
and in final form 10.10.2011.
|