Inverse semigroup

Last updated

In group theory, an inverse semigroup (occasionally called an inversion semigroup [1] ) S is a semigroup in which every element x in S has a unique inversey in S in the sense that x = xyx and y = yxy, i.e. a regular semigroup in which every element has a unique inverse. Inverse semigroups appear in a range of contexts; for example, they can be employed in the study of partial symmetries. [2]

Contents

(The convention followed in this article will be that of writing a function on the right of its argument, e.g. x f rather than f(x), and composing functions from left to right—a convention often observed in semigroup theory.)

Origins

Inverse semigroups were introduced independently by Viktor Vladimirovich Wagner [3] in the Soviet Union in 1952, [4] and by Gordon Preston in the United Kingdom in 1954. [5] Both authors arrived at inverse semigroups via the study of partial bijections of a set: a partial transformation α of a set X is a function from A to B, where A and B are subsets of X. Let α and β be partial transformations of a set X; α and β can be composed (from left to right) on the largest domain upon which it "makes sense" to compose them:

where α1 denotes the preimage under α. Partial transformations had already been studied in the context of pseudogroups. [6] It was Wagner, however, who was the first to observe that the composition of partial transformations is a special case of the composition of binary relations. [7] He recognised also that the domain of composition of two partial transformations may be the empty set, so he introduced an empty transformation to take account of this. With the addition of this empty transformation, the composition of partial transformations of a set becomes an everywhere-defined associative binary operation. Under this composition, the collection of all partial one-one transformations of a set X forms an inverse semigroup, called the symmetric inverse semigroup (or monoid) on X, with inverse the functional inverse defined from image to domain (equivalently, the converse relation). [8] This is the "archetypal" inverse semigroup, in the same way that a symmetric group is the archetypal group. For example, just as every group can be embedded in a symmetric group, every inverse semigroup can be embedded in a symmetric inverse semigroup (see § Homomorphisms and representations of inverse semigroups below).

The basics

The inverse of an element x of an inverse semigroup S is usually written x1. Inverses in an inverse semigroup have many of the same properties as inverses in a group, for example, (ab)1 = b1a1. In an inverse monoid, xx1 and x1x are not necessarily equal to the identity, but they are both idempotent. [9] An inverse monoid S in which xx1 = 1 = x1x, for all x in S (a unipotent inverse monoid), is, of course, a group.

There are a number of equivalent characterisations of an inverse semigroup S: [10]

The idempotent in the -class of s is s1s, whilst the idempotent in the -class of s is ss1. There is therefore a simple characterisation of Green's relations in an inverse semigroup: [11]

Unless stated otherwise, E(S) will denote the semilattice of idempotents of an inverse semigroup S.

Examples of inverse semigroups

Multiplication table example. It is associative and every element has its own inverse according to aba = a, bab = b. It has no identity and is not commutative.

Inverse semigroup
abcde
aaaaaa
babcaa
caaabc
dadeaa
eaaade

The natural partial order

An inverse semigroup S possesses a natural partial order relation ≤ (sometimes denoted by ω), which is defined by the following: [12]

for some idempotent e in S. Equivalently,

for some (in general, different) idempotent f in S. In fact, e can be taken to be aa1 and f to be a1a. [13]

The natural partial order is compatible with both multiplication and inversion, that is, [14]

and

In a group, this partial order simply reduces to equality, since the identity is the only idempotent. In a symmetric inverse semigroup, the partial order reduces to restriction of mappings, i.e., αβ if, and only if, the domain of α is contained in the domain of β and = , for all x in the domain of α. [15]

The natural partial order on an inverse semigroup interacts with Green's relations as follows: if st and st, then s = t. Similarly, if st. [16]

On E(S), the natural partial order becomes:

so, since the idempotents form a semilattice under the product operation, products on E(S) give least upper bounds with respect to ≤.

If E(S) is finite and forms a chain (i.e., E(S) is totally ordered by ≤), then S is a union of groups. [17] If E(S) is an infinite chain it is possible to obtain an analogous result under additional hypotheses on S and E(S). [18]

Homomorphisms and representations of inverse semigroups

A homomorphism (or morphism) of inverse semigroups is defined in exactly the same way as for any other semigroup: for inverse semigroups S and T, a function θ from S to T is a morphism if ()() = (st)θ, for all s,t in S. The definition of a morphism of inverse semigroups could be augmented by including the condition ()1 = s1θ, however, there is no need to do so, since this property follows from the above definition, via the following theorem:

Theorem. The homomorphic image of an inverse semigroup is an inverse semigroup; the inverse of an element is always mapped to the inverse of the image of that element. [19]

One of the earliest results proved about inverse semigroups was the Wagner–Preston Theorem, which is an analogue of Cayley's theorem for groups:

Wagner–Preston Theorem. If S is an inverse semigroup, then the function φ from S to , given by

dom () = Sa1 and x() = xa

is a faithful representation of S. [20]

Thus, any inverse semigroup can be embedded in a symmetric inverse semigroup, and with image closed under the inverse operation on partial bijections. Conversely, any subsemigroup of the symmetric inverse semigroup closed under the inverse operation is an inverse semigroup. Hence a semigroup S is isomorphic to a subsemigroup of the symmetric inverse semigroup closed under inverses if and only if S is an inverse semigroup.

Congruences on inverse semigroups

Congruences are defined on inverse semigroups in exactly the same way as for any other semigroup: a congruenceρ is an equivalence relation that is compatible with semigroup multiplication, i.e.,

[21]

Of particular interest is the relation , defined on an inverse semigroup S by

there exists a with [22]

It can be shown that σ is a congruence and, in fact, it is a group congruence, meaning that the factor semigroup S/σ is a group. In the set of all group congruences on a semigroup S, the minimal element (for the partial order defined by inclusion of sets) need not be the smallest element. In the specific case in which S is an inverse semigroup σ is the smallest congruence on S such that S/σ is a group, that is, if τ is any other congruence on S with S/τ a group, then σ is contained in τ. The congruence σ is called the minimum group congruence on S. [23] The minimum group congruence can be used to give a characterisation of E-unitary inverse semigroups (see below).

A congruence ρ on an inverse semigroup S is called idempotent pure if

[24]

E-unitary inverse semigroups

One class of inverse semigroups that has been studied extensively over the years is the class of E-unitary inverse semigroups: an inverse semigroup S (with semilattice E of idempotents) is E-unitary if, for all e in E and all s in S,

Equivalently,

[25]

One further characterisation of an E-unitary inverse semigroup S is the following: if e is in E and es, for some s in S, then s is in E. [26]

Theorem. Let S be an inverse semigroup with semilattice E of idempotents, and minimum group congruence σ. Then the following are equivalent: [27]

where is the compatibility relation on S, defined by

are idempotent.

McAlister's Covering Theorem. Every inverse semigroup S has a E-unitary cover; that is there exists an idempotent separating surjective homomorphism from some E-unitary semigroup T onto S. [28]

Central to the study of E-unitary inverse semigroups is the following construction. [29] Let be a partially ordered set, with ordering ≤, and let be a subset of with the properties that

Now let G be a group that acts on (on the left), such that

The triple is also assumed to have the following properties:

Such a triple is called a McAlister triple. A McAlister triple is used to define the following:

together with multiplication

.

Then is an inverse semigroup under this multiplication, with (A, g)1 = (g1A, g1). One of the main results in the study of E-unitary inverse semigroups is McAlister's P-Theorem:

McAlister's P-Theorem. Let be a McAlister triple. Then is an E-unitary inverse semigroup. Conversely, every E-unitary inverse semigroup is isomorphic to one of this type. [30]

F-inverse semigroups

An inverse semigroup is said to be F-inverse if every element has a unique maximal element above it in the natural partial order, i.e. every σ-class has a maximal element. Every F-inverse semigroup is an E-unitary monoid. McAlister's covering theorem has been refined by M.V. Lawson to:

Theorem. Every inverse semigroup has an F-inverse cover. [31]

McAlister's P-theorem has been used to characterize F-inverse semigroups as well. A McAlister triple is an F-inverse semigroup if and only if is a principal ideal of and is a semilattice.

Free inverse semigroups

A construction similar to a free group is possible for inverse semigroups. A presentation of the free inverse semigroup on a set X may be obtained by considering the free semigroup with involution, where involution is the taking of the inverse, and then taking the quotient by the Vagner congruence

The word problem for free inverse semigroups is much more intricate than that of free groups. A celebrated result in this area due to W. D. Munn who showed that elements of the free inverse semigroup can be naturally regarded as trees, known as Munn trees. Multiplication in the free inverse semigroup has a correspondent on Munn trees, which essentially consists of overlapping common portions of the trees. (see Lawson 1998 for further details)

Any free inverse semigroup is F-inverse. [31]

Connections with category theory

The above composition of partial transformations of a set gives rise to a symmetric inverse semigroup. There is another way of composing partial transformations, which is more restrictive than that used above: two partial transformations α and β are composed if, and only if, the image of α is equal to the domain of β; otherwise, the composition αβ is undefined. Under this alternative composition, the collection of all partial one-one transformations of a set forms not an inverse semigroup but an inductive groupoid, in the sense of category theory. This close connection between inverse semigroups and inductive groupoids is embodied in the Ehresmann–Schein–Nambooripad Theorem, which states that an inductive groupoid can always be constructed from an inverse semigroup, and conversely. [32] More precisely, an inverse semigroup is precisely a groupoid in the category of posets that is an étale groupoid with respect to its (dual) Alexandrov topology and whose poset of objects is a meet-semilattice.

Generalisations of inverse semigroups

As noted above, an inverse semigroup S can be defined by the conditions (1) S is a regular semigroup, and (2) the idempotents in S commute; this has led to two distinct classes of generalisations of an inverse semigroup: semigroups in which (1) holds, but (2) does not, and vice versa.

Examples of regular generalisations of an inverse semigroup are: [33]

The class of generalised inverse semigroups is the intersection of the class of locally inverse semigroups and the class of orthodox semigroups. [34]

Amongst the non-regular generalisations of an inverse semigroup are: [35]

Inverse category

This notion of inverse also readily generalizes to categories. An inverse category is simply a category in which every morphism f : XY has a generalized inverse g : YX such that fgf = f and gfg = g. An inverse category is selfdual. The category of sets and partial bijections is the prime example. [36]

Inverse categories have found various applications in theoretical computer science. [37]

See also

Notes

  1. Weisstein, Eric W. (2002). CRC Concise Encyclopedia of Mathematics (2nd ed.). CRC Press. p. 1528. ISBN   978-1-4200-3522-3.
  2. Lawson 1998
  3. Since his father was German, Wagner preferred the German transliteration of his name (with a "W", rather than a "V") from Cyrillic – see Schein 1981.
  4. First a short announcement in Wagner 1952, then a much more comprehensive exposition in Wagner 1953.
  5. Preston 1954a,b,c.
  6. See, for example, Gołab 1939.
  7. Schein 2002 , p. 152
  8. Howie 1995 , p. 149
  9. Howie 1995 , Proposition 5.1.2(1)
  10. Howie 1995 , Theorem 5.1.1
  11. Howie 1995 , Proposition 5.1.2(1)
  12. Wagner 1952
  13. Howie 1995 , Proposition 5.2.1
  14. Howie 1995 , pp. 152–3
  15. Howie 1995 , p. 153
  16. Lawson 1998 , Proposition 3.2.3
  17. Clifford & Preston 1967 , Theorem 7.5
  18. Gonçalves, D; Sobottka, M; Starling, C (2017). "Inverse semigroup shifts over countable alphabets". Semigroup Forum. 96 (2): 203–240. arXiv: 1510.04117 . doi:10.1007/s00233-017-9858-5Corollary 4.9{{cite journal}}: CS1 maint: postscript (link)
  19. Clifford & Preston 1967 , Theorem 7.36
  20. Howie 1995 , Theorem 5.1.7 Originally, Wagner 1952 and, independently, Preston 1954c.
  21. Howie 1995 , p. 22
  22. Lawson 1998 , p. 62
  23. Lawson 1998 , Theorem 2.4.1
  24. Lawson 1998 , p. 65
  25. Howie 1995 , p. 192
  26. Lawson 1998 , Proposition 2.4.3
  27. Lawson 1998 , Theorem 2.4.6
  28. Grillet, P. A. (1995). Semigroups: An Introduction to the Structure Theory. CRC Press. p. 248. ISBN   978-0-8247-9662-4.
  29. Howie 1995 , pp. 193–4
  30. Howie 1995 , Theorem 5.9.2. Originally, McAlister 1974a,b.
  31. 1 2 Lawson 1998 , p. 230
  32. Lawson 1998 , 4.1.8
  33. Howie 1995 , Section 2.4 & Chapter 6
  34. Howie 1995 , p. 222
  35. Fountain 1979, Gould
  36. Grandis, Marco (2012). Homological Algebra: The Interplay of Homology with Distributive Lattices and Orthodox Semigroups. World Scientific. p. 55. ISBN   978-981-4407-06-9.
  37. Hines, Peter; Braunstein, Samuel L. (2010). "The Structure of Partial Isometries". In Gay and, Simon; Mackie, Ian (eds.). Semantic Techniques in Quantum Computation. Cambridge University Press. p. 369. ISBN   978-0-521-51374-6.

Related Research Articles

<span class="mw-page-title-main">Monoid</span> Algebraic structure with an associative operation and an identity element

In abstract algebra, a branch of mathematics, a monoid is a set equipped with an associative binary operation and an identity element. For example, the nonnegative integers with addition form a monoid, the identity element being 0.

<span class="mw-page-title-main">Semigroup</span> Algebraic structure consisting of a set with an associative binary operation

In mathematics, a semigroup is an algebraic structure consisting of a set together with an associative internal binary operation on it.

In mathematics, the concept of an inverse element generalises the concepts of opposite and reciprocal of numbers.

In mathematics, a join-semilattice is a partially ordered set that has a join for any nonempty finite subset. Dually, a meet-semilattice is a partially ordered set which has a meet for any nonempty finite subset. Every join-semilattice is a meet-semilattice in the inverse order and vice versa.

In mathematics, Green's relations are five equivalence relations that characterise the elements of a semigroup in terms of the principal ideals they generate. The relations are named for James Alexander Green, who introduced them in a paper of 1951. John Mackintosh Howie, a prominent semigroup theorist, described this work as "so all-pervading that, on encountering a new semigroup, almost the first question one asks is 'What are the Green relations like?'". The relations are useful for understanding the nature of divisibility in a semigroup; they are also valid for groups, but in this case tell us nothing useful, because groups always have divisibility.

In mathematics, the bicyclic semigroup is an algebraic object important for the structure theory of semigroups. Although it is in fact a monoid, it is usually referred to as simply a semigroup. It is perhaps most easily understood as the syntactic monoid describing the Dyck language of balanced pairs of parentheses. Thus, it finds common applications in combinatorics, such as describing binary trees and associative algebras.

In mathematics, a band is a semigroup in which every element is idempotent. Bands were first studied and named by A. H. Clifford (1954).

In abstract algebra, a semiheap is an algebraic structure consisting of a non-empty set H with a ternary operation denoted that satisfies a modified associativity property:

In algebra, a presentation of a monoid is a description of a monoid in terms of a set Σ of generators and a set of relations on the free monoid Σ generated by Σ. The monoid is then presented as the quotient of the free monoid by these relations. This is an analogue of a group presentation in group theory.

In mathematics, a regular semigroup is a semigroup S in which every element is regular, i.e., for each element a in S there exists an element x in S such that axa = a. Regular semigroups are one of the most-studied classes of semigroups, and their structure is particularly amenable to study via Green's relations.

In mathematics, the congruence lattice problem asks whether every algebraic distributive lattice is isomorphic to the congruence lattice of some other lattice. The problem was posed by Robert P. Dilworth, and for many years it was one of the most famous and long-standing open problems in lattice theory; it had a deep impact on the development of lattice theory itself. The conjecture that every distributive lattice is a congruence lattice is true for all distributive lattices with at most 1 compact elements, but F. Wehrung provided a counterexample for distributive lattices with ℵ2 compact elements using a construction based on Kuratowski's free set theorem.

A biordered set is a mathematical object that occurs in the description of the structure of the set of idempotents in a semigroup.

In mathematics, particularly in abstract algebra, a semigroup with involution or a *-semigroup is a semigroup equipped with an involutive anti-automorphism, which—roughly speaking—brings it closer to a group because this involution, considered as unary operator, exhibits certain fundamental properties of the operation of taking the inverse in a group: uniqueness, double application "cancelling itself out", and the same interaction law with the binary operation as in the case of the group inverse. It is thus not a surprise that any group is a semigroup with involution. However, there are significant natural examples of semigroups with involution that are not groups.

In abstract algebra, a cover is one instance of some mathematical structure mapping onto another instance, such as a group (trivially) covering a subgroup. This should not be confused with the concept of a cover in topology.

In mathematics, Brandt semigroups are completely 0-simple inverse semigroups. In other words, they are semigroups without proper ideals and which are also inverse semigroups. They are built in the same way as completely 0-simple semigroups:

In mathematics, the Munn semigroup is the inverse semigroup of isomorphisms between principal ideals of a semilattice. Munn semigroups are named for the Scottish mathematician Walter Douglas Munn (1929–2008).

In mathematics, the oscillator representation is a projective unitary representation of the symplectic group, first investigated by Irving Segal, David Shale, and André Weil. A natural extension of the representation leads to a semigroup of contraction operators, introduced as the oscillator semigroup by Roger Howe in 1988. The semigroup had previously been studied by other mathematicians and physicists, most notably Felix Berezin in the 1960s. The simplest example in one dimension is given by SU(1,1). It acts as Möbius transformations on the extended complex plane, leaving the unit circle invariant. In that case the oscillator representation is a unitary representation of a double cover of SU(1,1) and the oscillator semigroup corresponds to a representation by contraction operators of the semigroup in SL(2,C) corresponding to Möbius transformations that take the unit disk into itself.

In mathematics, an invariant convex cone is a closed convex cone in a Lie algebra of a connected Lie group that is invariant under inner automorphisms. The study of such cones was initiated by Ernest Vinberg and Bertram Kostant.

In mathematics, the four-spiral semigroup is a special semigroup generated by four idempotent elements. This special semigroup was first studied by Karl Byleen in a doctoral dissertation submitted to the University of Nebraska in 1977. It has several interesting properties: it is one of the most important examples of bi-simple but not completely-simple semigroups; it is also an important example of a fundamental regular semigroup; it is an indispensable building block of bisimple, idempotent-generated regular semigroups. A certain semigroup, called double four-spiral semigroup, generated by five idempotent elements has also been studied along with the four-spiral semigroup.

In abstract algebra, an E-dense semigroup is a semigroup in which every element a has at least one weak inversex, meaning that xax = x. The notion of weak inverse is weaker than the notion of inverse used in a regular semigroup.

References

Further reading