diff --git "a/39E4T4oBgHgl3EQfAwsS/content/tmp_files/2301.04845v1.pdf.txt" "b/39E4T4oBgHgl3EQfAwsS/content/tmp_files/2301.04845v1.pdf.txt" new file mode 100644--- /dev/null +++ "b/39E4T4oBgHgl3EQfAwsS/content/tmp_files/2301.04845v1.pdf.txt" @@ -0,0 +1,5345 @@ +A groupoid approach to regular ∗-semigroups +James East∗ and P.A. Azeef Muhammed +Centre for Research in Mathematics and Data Science, +Western Sydney University, Locked Bag 1797, Penrith NSW 2751, Australia. +J.East@WesternSydney.edu.au, A.ParayilAjmal@WesternSydney.edu.au +Abstract +In this paper we develop a new groupoid-based structure theory for the class of regular ∗- +semigroups. This class occupies something of a ‘sweet spot’ between the important classes of +inverse and regular semigroups, and contains many important and natural examples. Some +of the most significant families include the partition, Brauer and Temperley-Lieb monoids, +among other diagram monoids. +Our main result shows that the category of regular ∗-semigroups is isomorphic to the +category of so-called ‘chained projection groupoids’. Such a groupoid is in fact a pair (G, ε), +where: +• G is an ordered groupoid, whose object set P is a projection algebra (in the sense of +Imaoka and Jones), and +• ε : C → G is a special functor, where C is a certain natural ‘chain groupoid’ constructed +from P. +Roughly speaking: +the groupoid G remembers only the ‘easy’ products in a regular ∗- +semigroup S; the projection algebra P remembers only the ‘conjugation action’ of the projec- +tions of S; and the functor ε tells us how G and P ‘fit together’ in order to recover the entire +structure of S. In this way, our main result contains the first completely general structure +theorem for regular ∗-semigroups. +Among other applications, we use our structure theorem to obtain new, and perhaps +more transparent, constructions of fundamental regular ∗-semigroups. We also use the chain +groupoids to demonstrate the existence of free (idempotent-generated) regular ∗-semigroups +associated to arbitrary projection algebras. Specialising to inverse semigroups, we also obtain +a new proof of the celebrated Ehresmann–Schein–Nambooripad Theorem. +We consider several examples along the way, and pose a number of problems that we +believe are worthy of further attention. +Keywords: Regular ∗-semigroups; inverse semigroups; groupoids; projection algebras; parti- +tion monoids. +MSC: 20M10, 20M50, 18B40, 20M17, 20M20, 20M05. +Contents +1 +Introduction +3 +2 +Preliminaries +9 +2.1 +Semigroups +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +9 +2.2 +Categories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +10 +2.3 +Regular ∗-semigroups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +15 +2.4 +Case study: diagram monoids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +22 +∗Supported by ARC Future Fellowship FT190100632. +1 +arXiv:2301.04845v1 [math.RA] 12 Jan 2023 + +I +Structure +30 +3 +Projection algebras +30 +3.1 +Definitions and basic properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . +31 +3.2 +The path category +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +34 +3.3 +Linked pairs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +38 +3.4 +The chain groupoid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +40 +3.5 +The category of projection algebras . . . . . . . . . . . . . . . . . . . . . . . . . . +42 +4 +Projection groupoids +43 +4.1 +Projection groupoids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +43 +4.2 +Chained projection groupoids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +47 +4.3 +The regular ∗-semigroup associated to a chained projection groupoid . . . . . . . +51 +4.4 +Products of projections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +55 +5 +Regular ∗-semigroups +57 +5.1 +The chained projection groupoid associated to a regular ∗-semigroup . . . . . . . +58 +5.2 +Examples +. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +62 +6 +The category isomorphism +66 +II +Applications +71 +7 +Idempotent-generated regular ∗-semigroups +71 +7.1 +The chain semigroup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . +72 +7.2 +Presentation by generators and relations . . . . . . . . . . . . . . . . . . . . . . . +76 +7.3 +Chain semigroups as free objects +. . . . . . . . . . . . . . . . . . . . . . . . . . . +79 +8 +Fundamental regular ∗-semigroups +83 +8.1 +The maximum idempotent-separating congruence . . . . . . . . . . . . . . . . . . +83 +8.2 +Maximum fundamental regular ∗-semigroups . . . . . . . . . . . . . . . . . . . . . +85 +8.3 +Idempotent-generated fundamental regular ∗-semigroups . . . . . . . . . . . . . . +91 +9 +Inverse semigroups and inductive groupoids +92 +9.1 +The chained projection groupoid associated to an inverse semigroup +. . . . . . . +93 +9.2 +The Ehresmann–Schein–Nambooripad Theorem . . . . . . . . . . . . . . . . . . . +94 +9.3 +Fundamental inverse semigroups +. . . . . . . . . . . . . . . . . . . . . . . . . . . +97 +2 + +1 +Introduction +‘All of mathematics is the study of symmetry’ — H.S.M. Coxeter. +The broadest possible setting in which to place the current work is in the study of mathematical +symmetry. This should practically go without saying, given the above quote, attributed to Cox- +eter by Roberts in [125]. Nevertheless, the algebraic and categorical structures we are concerned +with had their origins in the concerted efforts of mathematicians and scientists to pin down the +exact meaning of ‘symmetry’, and to devise the best ways to encode and study it. +We could not hope to adequately recount the fascinating history here, but fortunately we +can instead refer readers to the introduction of Lawson’s monograph [87], and also to the work +of Hollings [66]. We pick up the story in the late nineteenth century, long after the discovery +of group theory, when mathematicians were facing a plethora of new non-Euclidean geometries. +Some of these had trivial automorphism groups, yet were intuitively highly symmetric. Two main +approaches to capture this non-global or partial symmetry eventually emerged in the middle of +the twentieth century: the inductive groupoids of Charles Ehresmann [45, 46], and the inverse +semigroups of Viktor Wagner [131, 132] and Gordon Preston [121, 122]. Far from being rival +approaches, the two theories were eventually united in one of the landmark results of inverse +semigroup theory: +Ehresmann–Schein–Nambooripad Theorem. The categories of inverse semigroups (with +semigroup homomorphisms) and inductive groupoids (with inductive functors) are isomorphic. +This theorem contains too much information to fully unpack just now, but more will be said in +Chapter 9. The ESN Theorem was first formulated by Lawson in [87, Theorem 4.1.8], who named +it after the three mathematicians who had contributed most to its development. We have already +encountered Ehresmann. The second named author, Boris Schein, was largely responsible for +establishing the explicit connection between inverse semigroups and inductive groupoids [128]. +The third, K.S. Subramonian Nambooripad [110], was credited for his categorical framework +(albeit in a different context), wherein the theorem is couched in terms of a relationship between +entire categories (including their morphisms) of semigroups and groupoids, rather than simply +showing how the algebraic and categorical structures are inter-definable. +The ESN Theorem has had many far-reaching consequences, both within the fields of inverse +semigroups and inductive groupoids (see [87]), and also beyond. Before we comment on the +latter, it is worth addressing the former, by confronting an elephant in the room: +• Inverse semigroups are relatively simple mathematical objects. In what sense does it ‘help’ +to view them through the lens of groupoids, and category theory? +The precise answer to this question would again take us too far afield, but some brief comments +should suffice for the time being. First, an inverse semigroup S can in principle be completely +understood in terms of its multiplication table. Each pair of elements a, b ∈ S gives rise to the +product ab ∈ S. This binary operation must be associative, i.e. satisfy the law (ab)c = a(bc). +Further, each element a ∈ S must have a unique (semigroup-theoretic) inverse in S, i.e. an +element a′ ∈ S satisfying a = aa′a and a′ = a′aa′. Uniqueness means that we can think of +a �→ a′ as an additional (unary) operation of S, and we generally write a′ = a−1. +Among +other things, the ESN Theorem says that one can associate a groupoid G = G(S) to such an +inverse semigroup S. We will postpone the full definition of G for now, but two key points bear +emphasising: +• the morphisms of G are precisely the elements of S, and +• the (partial) composition in G is a restriction of the (total) product in S. +3 + +Roughly speaking, the groupoid G ‘remembers’ only the ‘easy’ products of S, and the ESN +Theorem tells us that (combined with some order-theoretic data) these easy products are enough +to reconstruct the entire multiplication table of S. (The exact meaning of ‘easy’ will be explored +a little below, and in more detail in Sections 2.3 and 2.4.) In a sense, the latter point encapsulates +the deepest part of the theorem: Inductive groupoids are precisely the (abstract) categories G +whose composition can be extended in a natural way to construct an inverse semigroup S = S(G). +Moreover, the constructions S �→ G(S) and G �→ S(G) are mutually inverse functors between +the categories of inverse semigroups and inductive groupoids. This category isomorphism is not +merely a convenient way to say that inverse semigroups and inductive groupoids are ‘one and +the same’, however. The fact that the isomorphism sends morphisms to morphisms means that +a map φ : S → S′ between inverse semigroups is a semigroup homomorphism if and only if it is +a (so-called) inductive functor G(φ) = φ : G(S) → G(S′). Less effort is required to show that φ is +a functor, as opposed to a full semigroup homomorphism, as the former involves checking that +(a ◦ b)φ = (aφ) ◦ (bφ) +for composable pairs a, b ∈ G(S) ≡ S, +and we emphasise that the composable pairs in G are precisely the easy products in S. +The ESN Theorem has had a major impact outside of inverse semigroup theory, with one +particularly fruitful direction being the application of inverse semigroups to C∗-algebras. For +example, the articles [12,14,47,80,83,89,94,95,105,117] display a mixture of semigroup-theoretic +and groupoid-based approaches. For more extensive discussions, and many more references, see +Lawson’s survey [90] and Paterson’s monograph [116]. +Another way to measure the influence of the ESN Theorem is to consider its extensions and +generalisations. When faced with an important class of semigroups, one naturally looks for an +‘ESN-type link’ with an equally-natural class of categories, and vice versa. Thus, we have ESN- +type theorems for regular semigroups, restriction semigroups, Ehresmann semigroups, concordant +semigroups, and others; see for example [1, 18, 49, 50, 53–55, 65, 66, 86, 88, 91, 110, 133–135]. As +with other ‘dualities’ in mathematics [8, 15, 119, 120, 123, 124, 129], such correspondences allow +problems in one field to be translated into another, where they can hopefully be solved, and the +solution then reinterpreted in the original context. +Arguably the most significant of the papers just cited was Nambooripad’s 1979 memoir [110]. +As well as pioneering the categorical approach (which led to the ‘N’ in ‘ESN’), this paper majorly +extended the ESN Theorem to the class of regular semigroups. These are semigroups in which +every element has at least one inverse. Dropping the uniqueness restriction on inverses leads +to a far more general class of semigroups, which contains many additional natural examples, +including semigroups of mappings, linear transformations, and more. The increase in generality +of the semigroups led inevitably to an increase in the complexity of the categorical structures +modelling them. Although Nambooripad still represented a regular semigroup S by an ordered +groupoid G = G(S), the groupoid G was not enough to completely recover the semigroup; an +additional layer of structure was required. +The need for this extra data is due to the fact +that the object set of G is not a semilattice (as for inverse semigroups), but instead a (regular) +biordered set: a partial algebra with a pair of intertwined pre-orders satisfying a fairly complex +set of axioms. The increase in generality also led to the sacrifice of the category isomorphism. +Nambooripad’s main result, [110, Theorem 4.14], is that the category of regular semigroups is +equivalent to a certain category of groupoids (which Nambooripad also called inductive). +One way to understand this sacrifice is via the loss of symmetry when moving from inverse +to regular semigroups. As we discussed above, inverse semigroups were devised to model partial +symmetries that were unrecognisable by groups. +This is formalised in the Wagner–Preston +Theorem, which states that any inverse semigroup is (isomorphic to) a semigroup of partial +symmetries of some mathematical structure; see [87, Theorem 1.5.1]. But an inverse semigroup +is itself an extremely symmetrical mathematical structure in its own right, so much so that the +canonical proof of the Wagner–Preston Theorem involves showing that an inverse semigroup +4 + +is (isomorphic to) a semigroup of partial symmetries of itself ! +This ‘internal symmetry’ is +manifested in the properties of the inversion operation a �→ a−1, speficially the fact that this is +an involution, i.e. satisfies the familiar laws +(a−1)−1 = a +and +(ab)−1 = b−1a−1, +(1.1) +beyond the defining property that a and a−1 are mutual inverses: +a = aa−1a +and +a−1 = a−1aa−1. +(1.2) +Inverse semigroups actually satisfy an additional law: +aa−1 · bb−1 = bb−1 · aa−1, +(1.3) +and this is in fact equivalent to the commutativity of idempotents. This in turn means that the +set E = E(S) of idempotents of an inverse semigroup S forms a semilattice, with the (order- +theoretic) meet of two idempotents being their (semigroup) product: e∧f = ef. This then feeds +into the definition of the groupoid G = G(S). The object set of G is E, and each element a ∈ S +is thought of as a morphism from a ‘domain idempotent’ d(a) = aa−1 to a ‘range idempotent’ +r(a) = a−1a. Beyond the obvious law +d(a) · a = a = a · r(a), +(1.4) +which follows immediately from (1.2), the key point making everything work is that +r(a) = d(b) +⇒ +d(ab) = d(a) and r(ab) = r(b). +(1.5) +This allows us to define the composition +a ◦ b = ab +when r(a) = d(b). +(1.6) +These are the ‘easy’ products alluded to above. Going in the reverse direction, the semilattice +structure of the object set E of an inductive groupoid G is crucial in extending the (partial) +composition to a (total) product. Given two morphisms a, b ∈ G, the range e = r(a) and domain +f = d(b) have a meet (greatest lower bound) in E: g = e ∧ f. The ordered structure of G means +that we have (right and left) restrictions a′ = a⇂g and b′ = g⇃b, and since r(a′) = g = d(b′), these +are composable in G, so that we can define +a • b = a′ ◦ b′ = a⇂g ◦ g⇃b, +and in this way reconstruct the entire structure of S = S(G). The operation • is often called the +pseudo-product in the literature. +The major difficulty in extending the above ideas to arbitrary regular semigroups is in the +loss of uniqueness of inverses, and hence in the possible expansion of the available domain and +range idempotents. Indeed, if an element a of a regular semigroup S possesses distinct inverses +a′, a′′ ∈ S, then one is faced with a dilemma: whether to think of a as a morphism aa′ → a′a +or aa′′ → a′′a, and in general it is quite possible to have aa′ ̸= aa′′ and/or a′a ̸= a′′a. Thus, +Nambooripad was led to consider the pairs (a, a′) and (a, a′′) as separate morphisms, essentially +splitting each semigroup element into several morphisms, one for each inverse, and then consider- +ing equivalence-classes of such pairs. This also necessitated a far more elaborate pseudo-product. +Given semigroup elements a, b ∈ S, one would like to identify suitable domain and range idem- +potents, e = r(a) and f = d(b), and then to find composable restrictions a⇂g and g⇃b (with g +‘below’ e and f) in order to define a • b = a⇂g ◦ g⇃b, as above. This is not generally possible, +however, and one inevitably has to work with the biordered sets alluded to above. Specifically, +Nambooripad located g in the so-called sandwich set S(e, f). The restrictions a⇂g and g⇃b still +5 + +may or may not exist, but a⇂eg and gf⇃b do. These latter restrictions still may or may not be +composable, but it is possible to find a distinguished element x for which the composition +a⇂eg ◦ x ◦ gf⇃b +does exist, and can be taken as the definition of the pseudo-product a • b. This element x has +the form +x = ε(eg, g) ◦ ε(g, gf) = ε(eg, g, gf), +where ε is a special functor into G from a certain category of ‘E-chains’. In the end, Namboori- +pad’s ‘groupoids’ are in fact pairs (G, ε), where G is a special ordered groupoid with a biordered +object set, and where ε is a special functor from the E-chain groupoid into G. The ingredients G +and ε are linked via a certain coherence condition, and a property of so-called singular squares +in the biordered set. It would be impossible to adequately cover all the details and subtleties of +Nambooripad’s work, so instead we refer the reader to the survey [104]. In any case, the step up +in complexity from the inverse case should be apparent from our brief treatment. +Let us now take a small step back. Recall that inverse semigroups are the semigroups with +a unary operation a �→ a−1 satisfying the laws (1.1), (1.2) and (1.3). Associated to such an +inverse semigroup S, one can define a groupoid G = G(S) with morphism set S, and with +composition (1.6). As it happens, law (1.3) is not a necessary ingredient in the construction of +the groupoid G. In other words, if S is a semigroup satisfying laws (1.1) and (1.2), then the above +procedure works without modification to produce a groupoid G = G(S). If law (1.3) does not +hold, then the groupoid may or may not be inductive, but it still has many remarkable properties +shared by inductive groupoids. This observation is, in essence, the starting point for the current +work. +Semigroups satisfying the laws (1.1) and (1.2) are precisely the regular ∗-semigroups of Nor- +dahl and Scheiblich [114]. (These semigroups are not to be confused with the ∗-regular semigroups +of Drazin [26], and are known by some other names in the literature; for example, they are called +special ∗-semigroups in [113].) As the terminology suggests, the involution on such a semigroup +is generally denoted a �→ a∗ (reserving a−1 for inverse semigroups), and this must satisfy the +laws +(a∗)∗ = a = aa∗a +and +(ab)∗ = b∗a∗. +The class of regular ∗-semigroups has received a great deal of attention in recent years, partly +because of the prototypical examples of the so-called diagram semigroups. These semigroups +include families such as the Brauer, Temperley-Lieb, Kauffman and partition monoids, and have +their origins and applications in a wide range of mathematical and scientific disciplines, from +representation theory, low-dimensional topology, statistical mechanics, and many more; see for +example [2–4, 6, 7, 9, 10, 20, 21, 39–41, 43, 44, 51, 62, 74–79, 81, 82, 85, 92, 93, 97–103, 130, 136, 137]. +These monoids have provided a strong bridge between semigroup theory and these other branches +of mathematics; for a fuller discussion of this fruitful dialogue, and for many more references, +see the introduction to [39]. +As we will see, regular ∗-semigroups occupy something of a ‘sweet spot’ between the classes of +inverse and regular semigroups. On the one hand, their involution provides the ‘internal symme- +try’ possessed by inverse semigroups, allowing for the natural groupoid representations discussed +above. On the other hand, the non-commutativity of idempotents (i.e. the dropping of law (1.3)) +leads to a much richer idempotent structure, which in turn means that the groupoid representa- +tion is not faithful, and necessitates a more intricate approach when defining the pseudo-product. +In a way, this ‘balance’ can be summed up in an observation about the proliferation of inverses. +In an inverse semigroup, the inverse ‘picks itself’, in the sense that each element has exactly one. +In a regular semigroup there is some degree of ‘chaos’; while inverses are guaranteed to exist, +there could be an over-abundance (there even exist regular (∗-)semigroups in which every ele- +ment is an inverse of every other element). In regular ∗-semigroups, inverses are not necessarily +6 + +unique, but the involution ‘picks one for you’. The involution also leads to a canonical way to +‘pick’ domain and range idempotents; for a ∈ S, one can take +d(a) = aa∗ +and +r(a) = a∗a. +(1.7) +Such idempotents have the additional property that they are invariant under the involution, as +for example (aa∗)∗ = (a∗)∗a∗ = aa∗. Such an idempotent is called a projection, and the set of +all such projections is denoted +P = P(S) = {p ∈ S : p2 = p = p∗}. +This is of course contained in the set +E = E(S) = {e ∈ S : e2 = e} +of all idempotents, but we do not have equality in general; in fact, it is well known (and easy to +see) that the regular ∗-semigroup S is inverse if and only if E = P. Although P = P(S) is not a +subsemigroup of the regular ∗-semigroup S in general, it can be given an algebraic structure via +the ‘conjugation action’ of projections on each other. Each element p ∈ P determines a unary +operation +θp : P → P +given by +qθp = pqp +for q ∈ P. +These so-called projection algebras have been used by Imaoka (under a different name) and +Jones (in a different, but equivalent, form) in [72] and [73], and are the appropriate ‘∗-analogues’ +of semilattices in the theory of regular ∗-semigroups. (Semilattices are indeed a special case, +wherein we define eθf = e ∧ f.) In particular, the groupoid G = G(S) associated to a regular +∗-semigroup S has the projection algebra P = P(S) for its object set. Moreover, (1.4) and (1.5) +both hold (with respect to the domains and ranges given in (1.7)), allowing us to define the +composition in G exactly as in (1.6). Conversely, in attempting to define a pseudo-product • +from such a ‘projection groupoid’ G, one first identifies the range p = r(a) and domain q = d(b), +calculates two further projections p′ = qθp and q′ = pθq (which collectively play the role of a +meet p ∧ q, which may or may not exist), and then forms the restrictions a⇂p′ and q′⇃b. These +are still not composable in general, as we do not always have p′ = q′, but there always exists a +special morphism e : p′ → q′, allowing us to define +a • b = a⇂p′ ◦ e ◦ q′⇃b. +This morphism e is again in the image of a special functor ε : C → G, where here C = C (P) is a +certain ‘chain groupoid’ associated to the projection algebra P. This all results in what we call +a chained projection groupoid, which is a pair (G, ε), consisting of: +• an ordered groupoid G, whose object set P is an (abstract) projection algebra, with strong +links between the categorical and algebraic structures of G and P, and +• a functor ε : C = C (P) → G obeying a natural coherence condition. +Roughly speaking, our main result (Theorem 6.7) states that this construction is invertible, +and furnishes an isomorphism between the categories of regular ∗-semigroups and (abstract) +chained projection groupoids. Specialising to the inverse case, we obtain the ESN Theorem as +a corollary. We emphasise, however, that Theorem 6.7 is not a specialisation of Nambooripad’s +result on regular semigroups [110]. +In fact, such a specialisation would invariably lead to a +category equivalence (to a very different category of groupoids) rather than an isomorphism, as +inverses are generally not unique in regular ∗-semigroups. +Any further discussion would necessitate additional technicalities that are undesirable at this +stage. So instead, let us now give a brief summary of the structure of the article. After this +introductory chapter, we begin in Chapter 2 with preliminaries on (regular ∗-)semigroups and +ordered categories, as well as a fairly detailed discussion of the special case of partition monoids. +The paper then splits into two main parts: +7 + +• Part I is devoted to the proof of our main result, Theorem 6.7, which establishes the isomor- +phism between the categories of regular ∗-semigroups and chained projection groupoids. +• Part II then applies the machinery developed in the first part to (free) idempotent-generated +regular ∗-semigroups, fundamental regular ∗-semigroups and inverse semigroups. +Part I contains Chapters 3–6, which build towards our main structural result, stated in Theo- +rem 6.7. +• Chapter 3 begins with the most basic building block, the so-called projection algebras. In +particular, we show how to build a number of categorical structures on top of an abstract +projection algebra P, culminating in the construction of the chain groupoid C = C (P). +• In Chapter 4 we turn to the class of projection groupoids. These are ordered groupoids G, +whose object set P is a projection algebra, with strong links between the structures of G (as a +groupoid) and P (as an algebra). The key idea is that of a chained projection groupoid, which +is a pair (G, ε) consisting of a projection groupoid G, and a special functor ε : C → G called +an evaluation map, where here C = C (P) is the chain groupoid of the projection algebra P. +The main result of the chapter is Theorem 4.36, which shows how to construct a regular +∗-semigroup S = S(G, ε) from a chained projection groupoid (G, ε). +• Chapter 5 goes in the opposite direction, and shows how a regular ∗-semigroup S gives rise to +a chained projection groupoid (G, ε) = G(S); see Theorem 5.19. In a sense, this breaks down +the structure of S into simpler parts: the groupoid G remembers only the ‘easy products’ +in S; the projection algebra P remembers the projections of S and their conjugation action; +and the evaluation map ε tells us how these parts fit together. This decomposition can be +thought of as a structure theorem for arbitrary regular ∗-semigroups (see Remark 6.8); as far +as we are aware it is the first of its kind. +• We then bring these ideas together in Chapter 6 in order to complete the proof of Theorem 6.7. +This involves showing that the S and G constructions are in fact mutually inverse functors +between the categories of regular ∗-semigroups and chained projection groupoids. +Part II contains Chapters 7–9, which comprise a number of applications of the theory developed +in Part I. +• Chapter 7 concerns idempotent-generated regular ∗-semigroups. The main construction here is +the so-called chain semigroup associated to an arbitrary (abstract) projection algebra. These +semigroups enjoy several categorical ‘free-ness’ properties, as described in Theorems 7.11, 7.20 +and 7.23. Theorem 7.14 gives a presentation by generators and relations. +• We then consider fundamental regular ∗-semigroups in Chapter 8. These have been previously +studied by a number of authors [70,72,113,138], but we show how our general machinery leads +to a new way to build these semigroups; see Theorems 8.20 and 8.21. In Theorem 8.24 we +show that (up to isomorphism) there is a unique idempotent-generated fundamental regular +∗-semigroup with a given (abstract) projection algebra, and show how to construct it. +• Finally, in Chapter 9 we show how the entire theory simplifies in the case of inverse semi- +groups. As an important application, we give a new proof of the Ehresmann–Schein–Nambooripad +Theorem, stated above. +Throughout the paper, we consider a number of examples, which serve both to illustrate the +general theory, and also to highlight some of the subtleties that arise. For the same reasons, at +times we will compare our constructions and results with the regular and inverse cases. We also +pose a number of open problems, which we believe are worthy of further study. +8 + +2 +Preliminaries +In this chapter we gather the preliminary definitions and basic results we require in the rest of +the paper, and establish notation. We begin in Section 2.1 with the main ideas from semigroup +theory, and then in Section 2.2 we discuss (ordered) categories and groupoids. +Section 2.3 +contains some basic background on regular ∗-semigroups, as well as a foreshadowing of various +ideas that will occur throughout the paper. +This material is included to motivate the more +abstract results of later chapters. Finally, Section 2.4 concerns a particularly important concrete +class of regular ∗-semigroups, namely the partition monoids. +This is again included to give +more motivation for our later ideas and results, and to show how these can be understood +diagrammatically for these monoids. We also comment on a number of known results, and show +how they can be interpreted through the groupoid lens. +For further general background, we refer the reader to texts such as [16,68] for semigroups, [87] +for inverse semigroups, [5,96] for categories, and [13] for universal algebra, though the latter is +only used tangentially. +2.1 +Semigroups +A semigroup is a set with an associative binary operation, which will typically be denoted by +juxtaposition. A monoid is a semigroup with an identity element. As usual we write S1 for the +monoid completion of the semigroup S. So S1 = S if S happens to be a monoid; otherwise, +S1 = S ∪ {1}, where 1 is a symbol not belonging to S, acting as an adjoined identity element. +Green’s relations are five equivalence relations, defined on an arbitrary semigroup S as fol- +lows [59]. First, for a, b ∈ S we have +a R b ⇔ aS1 = bS1, +a L b ⇔ S1a = S1b, +a J b ⇔ S1aS1 = S1bS1. +Note that a R b precisely when either a = b or else a = bx and b = ay for some x, y ∈ S; similar +comments apply to L and J . The remaining two of Green’s relations are defined by +H = R ∩ L +and +D = R ∨ L , +where the latter is the join in the lattice of equivalences on S, i.e. the least equivalence contain- +ing R ∪ L . It is well known that D = R ◦ L = L ◦ R, and that D = J if S is finite. We +denote the R-class of a ∈ S by +Ra = {b ∈ S : a R b}, +and similarly for L -classes La, and so on. +An element a of a semigroup S is regular if a = axa for some x ∈ S. The element y = xax +then has the property that a = aya and y = yay; such an element y is called an inverse of x. +We say S is regular if every element of S is regular. The set of idempotents of S is denoted +E(S) = {e ∈ S : e = e2}. +Idempotents are of course regular. An inverse semigroup is a semigroup in which every element +has a unique inverse. It is well known that S is inverse if and only if it is regular and its idempo- +tents commute; in this case, E(S) is a semilattice, i.e. a semigroup of commuting idempotents. +In this paper we are concerned with a class of semigroups contained strictly between regular +and inverse semigroups, the so-called regular ∗-semigroups of Nordahl and Scheiblich [114]. We +postpone their definition until Section 2.3. +A congruence on a semigroup S is an equivalence relation σ that is compatible with the +product of S, meaning that +a σ b +⇒ +ax σ bx and xa σ xb +for all a, b, x ∈ S, +9 + +or equivalently that +a σ b and x σ y +⇒ +ax σ by +for all a, b, x, y ∈ S. +Given a congruence σ, the quotient semigroup S/σ consists of all σ-classes under the induced +product, [a][b] = [ab]. The kernel of a semigroup homomorphism φ : S → T is the congruence +ker(φ) = {(a, b) ∈ S × S : aφ = bφ}. +The fundamental homomorphism theorem for semigroups says that the quotient S/ ker(φ) is +isomorphic to im(φ), the image of S under φ. +2.2 +Categories +Unless stated otherwise, the categories we are concerned with are assumed to be small. We +typically identify a (small) category C with its set of morphisms. We identify the objects of C with +the identities, the set of which is denoted vC. We denote the domain and codomain (a.k.a. range) +of a ∈ C by d(a) and r(a), respectively. We compose morphisms left to right, so a◦b is defined if +and only if r(a) = d(b), in which case d(a◦b) = d(a) and r(a◦b) = r(b). (Sometimes functors will +be written to the left of their arguments, and so composed right to left; it should always be clear +which convention is being used.) For p, q ∈ vC, we write C(p, q) = {a ∈ C : d(a) = p, r(a) = q} +for the set of all morphisms p → q. +All the (small) categories we study will have an involution and an order, as made precise in +the next two standard definitions. +Definition 2.1. A ∗-category is a (small) category C with an involution, i.e. a map C → C : a �→ a∗ +satisfying the following, for all a, b ∈ C: +(I1) d(a∗) = r(a) and r(a∗) = d(a). +(I2) (a∗)∗ = a. +(I3) If r(a) = d(b), then (a ◦ b)∗ = b∗ ◦ a∗. +A groupoid is a ∗-category for which we additionally have +(I4) a ◦ a∗ = d(a) (and hence also a∗ ◦ a = r(a)) for all a ∈ C. +In a groupoid, we typically write a∗ = a−1 for a ∈ C. +It is easy to show that p∗ = p for all p ∈ vC, when C is a ∗-category. It is also clear that (I1) +is a consequence of (I4), so a groupoid is a category with a map a �→ a∗ satisfying (I2)–(I4). +Definition 2.2. An ordered ∗-category is a ∗-category C equipped with a partial order ≤ satis- +fying the following, for all a, b, c, d ∈ C and p ∈ vC: +(O1) If a ≤ b, then d(a) ≤ d(b) and r(a) ≤ r(b). +(O2) If a ≤ b, then a∗ ≤ b∗. +(O3) If a ≤ b and c ≤ d, and if r(a) = d(c) and r(b) = d(d), then a ◦ c ≤ b ◦ d. +(O4) For all p ≤ d(a), there exists a unique u ≤ a with d(u) = p. +It is easy to see that (O1)–(O4) imply the following, which is a dual of (O4): +10 + +(O4)∗ For all q ≤ r(a), there exists a unique v ≤ a with r(v) = q. +(Here we have q ≤ d(a∗), and if w is the unique element with w ≤ a∗ with d(w) = q, then we +take v = w∗.) +The elements u and v in (O4) and (O4)∗ are denoted u = p⇃a and v = a⇂q, respectively, and +called the left restriction of a to p and the right restriction of a to q. Some authors call a⇂q a +restriction, and p⇃a a co-restriction; we prefer the left/right terminology, however, as we feel that +it does not ‘prioritise’ one over the other. +An ordered groupoid is a groupoid with a partial order satisfying (O1)–(O4). In fact, when C +is a groupoid, (O2), (O3) and (I4) together imply (O1). +It is easy to see that the object set vC is an order ideal in any ordered ∗-category C, meaning +that the following holds: +• For all a ∈ C and p ∈ vC, a ≤ p ⇒ a ∈ vC. +Indeed, if a ≤ p, and if we write q = d(a) ∈ vC, then by (O1) we have q ≤ d(p) = p. But then +a, q ≤ p and d(a) = q = d(q), so by uniqueness in (O4) we have a = q ∈ vC. It also follows from +this that q⇃p = q for any q ≤ p. We typically use facts such as these without explicit reference. +In what follows, it is typically more convenient to construct an ordered ∗-category C by: +• defining an order ≤ on the object set vC, +• defining left restrictions p⇃a, for a ∈ C and p ≤ d(a), +• specifying that a ≤ b (for morphisms a, b ∈ C) if a is a restriction of b. +The next lemma axiomatises the conditions required to ensure that we do indeed obtain an +ordered ∗-category in this way. +Lemma 2.3. Suppose C is a ∗-category for which the following two conditions hold: +(i) There is a partial order ≤ on the object set vC. +(ii) For all a ∈ C, and for all p ≤ d(a), there exists a morphism p⇃a ∈ C, such that the following +hold, for all a, b ∈ C and p, q ∈ vC: +(O1)′ If p ≤ d(a), then d(p⇃a) = p and r(p⇃a) ≤ r(a). +(O2)′ If p ≤ d(a), and if q = r(p⇃a), then (p⇃a)∗ = q⇃a∗. +(O3)′ d(a)⇃a = a. +(O4)′ For all p ≤ q ≤ d(a), we have p⇃q⇃a = p⇃a. +(O5)′ If p ≤ d(a) and r(a) = d(b), and if q = r(p⇃a), then p⇃(a ◦ b) = p⇃a ◦ q⇃b. +Then C is an ordered ∗-category with order given by +a ≤ b +⇔ +a = p⇃b +for some p ≤ d(b). +(2.4) +Moreover, any ordered ∗-category has the above form. +Proof. Beginning with the final assertion, suppose C is an ordered ∗-category. Then vC is a +sub-poset of C, and hence (i) holds. For (ii), we take p⇃a to be the morphism u ≤ a from (O4), +and (O1)′–(O5)′ are all easily checked. For example, to verify (O2)′, suppose a ∈ C and p ≤ d(a), +and let q = r(p⇃a). +Then since p⇃a ≤ a, it follows from (O2) that (p⇃a)∗ ≤ a∗, and we +11 + +have d((p⇃a)∗) = r(p⇃a) = q. But q⇃a∗ is the unique element below a∗ with domain q, so in +fact (p⇃a)∗ = q⇃a∗. +Conversely, suppose conditions (i) and (ii) both hold. We first check that the relation ≤ +in (2.4) is a partial order. Reflexivity follows immediately from (O3)′, and transitivity from (O4)′. +For anti-symmetry, suppose a ≤ b and b ≤ a, so that a = p⇃b and b = q⇃a for some p ≤ d(b) and +q ≤ d(a). Then +a = p⇃b = p⇃q⇃a = p⇃a +⇒ +d(a) = d(p⇃a) = p ≤ d(b) +and similarly +d(b) = q ≤ d(a). +It follows that p = q = d(a) = d(b), and so a = p⇃b = d(b)⇃b = b. Now that we know ≤ is a +partial order, we verify conditions (O1)–(O4). +(O1) and (O2). Suppose a ≤ b, so that a = p⇃b for some p ≤ d(b). Then (O1)′ gives +d(a) = d(p⇃b) = p ≤ d(b) +and +r(a) = r(p⇃b) ≤ r(b). +We also have a∗ = (p⇃b)∗ = q⇃b∗ by (O2)′, where q = r(p⇃b), so that a∗ ≤ b∗. +(O3). Suppose a ≤ b and c ≤ d are such that r(a) = d(c) and r(b) = d(d). So a = p⇃b and +c = q⇃d for some p ≤ d(b) and q ≤ d(d). Since +q = d(q⇃d) = d(c) = r(a) = r(p⇃b), +it follows from (O5)′ that a ◦ c = p⇃b ◦ q⇃d = p⇃(b ◦ d), so that a ◦ c ≤ b ◦ d. +(O4). Given p ≤ d(a), we certainly have u ≤ a and d(u) = p, where u = p⇃a. For uniqueness, +suppose also that x ≤ a for some x ∈ C with d(x) = p. Since x ≤ a, we have x = q⇃a for some +q ≤ d(a). But then q = d(q⇃a) = d(x) = p, so in fact x = q⇃a = p⇃a = u. +Remark 2.5. The previous result referred only to left restrictions. Right restrictions can be +defined from the left, using the involution: +a⇂q = (q⇃a∗)∗ +for a ∈ C and q ≤ r(a). +As explained in Definition 2.2, a⇂q is the unique element below a with codomain q. In particular, +the following holds in any ordered ∗-category: +(O6)′ If a ∈ C and p ≤ d(a), and if q = r(p⇃a), then p⇃a = a⇂q. +Each of (O1)′–(O6)′ of course have duals. For example, the dual of (O5)′ says: +• If q ≤ r(b) and r(a) = d(b), and if p = d(b⇂q), then (a ◦ b)⇂q = a⇂p ◦ b⇂q. +It is also worth noting that for a, b ∈ C we have +a ≤ b ⇔ a = p⇃b +for some p ≤ d(b) +⇔ a = b⇂q +for some q ≤ r(b) +⇔ a = p⇃b = b⇂q +for some p ≤ d(b) and q ≤ r(b). +The p, q ∈ P here are of course p = d(a) and q = r(a). +12 + +Definition 2.6. A v-congruence on a category C is an equivalence relation ≈ on C satisfying +the following, for all a, b, u, v ∈ C: +(C1) a ≈ b ⇒ [d(a) = d(b) and r(a) = r(b)], +(C2) a ≈ b ⇒ u ◦ a ≈ u ◦ b, whenever the stated compositions are defined, +(C3) a ≈ b ⇒ a ◦ v ≈ b ◦ v, whenever the stated compositions are defined. +If C is a ∗-category, we say that ≈ is a ∗-congruence if it satisfies (C1)–(C3) and the following, +for all a, b ∈ C: +(C4) a ≈ b ⇒ a∗ ≈ b∗. +If C is an ordered ∗-category, we say that ≈ is an ordered ∗-congruence if it satisfies (C1)–(C4) +and the following, for all a, b ∈ C and p ∈ vC: +(C5) [a ≈ b and p ≤ d(a)] ⇒ +p⇃a ≈ p⇃b. +Given a v-congruence ≈ on a category C, the quotient category C/≈ consists of all ≈-classes, +under the induced composition. +We typically write [a] for the ≈-class of a ∈ C. +It follows +immediately from (C1) that p ≈ q ⇒ p = q for objects p, q ∈ vC, so we can identify the object +sets of C and C/≈, viz. p ≡ [p]. In this way, for a ∈ C we have d[a] = d(a) and r[a] = r(a) for all +a ∈ C, and +[a] ◦ [b] = [a ◦ b] +whenever r[a] = d[b]. +Lemma 2.7. +(i) If ≈ is a ∗-congruence on a ∗-category C, then C/≈ is a ∗-category, with +involution given by +[a]∗ = [a∗] +for all a ∈ C. +(2.8) +If also a ◦ a∗ ≈ d(a) for all a ∈ C, then C/≈ is a groupoid. +(ii) If ≈ is an ordered ∗-congruence on an ordered ∗-category C, then C/≈ is an ordered ∗- +category, with involution (2.8), and order given by +α ≤ β +⇔ +a ≤ b +for some a ∈ α and b ∈ β. +(2.9) +Proof. Part (i) is routine, and (ii) only a little more difficult, so we just give a sketch for the +latter. +For this, one begins by showing that α ≤ β is equivalent to the ostensibly stronger +condition: +• For all b ∈ β, there exists a ∈ α such that a ≤ b. +It is then easy to check that ≤ is indeed a partial order on C/≈. Conditions (O1)–(O4) for C/≈ +follow quickly from the corresponding conditions for C. +We will shortly return to congruences in Lemma 2.17 below, where we give simpler criteria +for checking that a v-congruence satisfies (C4) or (C5). The statement of the lemma will require +certain maps defined on any ordered ∗-category. These maps will also be used extensively in the +rest of the paper. To define them, fix some such ordered ∗-category C with object set P = vC. +For p ∈ P, we write +p↓ = {q ∈ P : q ≤ p} +for the down-set of p in the poset (P, ≤). Consider a morphism a ∈ C. A restriction p⇃a is defined +precisely when p ≤ d(a), and we write pϑa = r(p⇃a) for the codomain of p⇃a; by (O1)′, we have +pϑa ≤ r(a). In other words, we have a map +ϑa : d(a)↓ → r(a)↓ +given by +pϑa = r(p⇃a). +(2.10) +13 + +Note that for q ≤ r(a), we have +d(a⇂q) = d((q⇃a∗)∗) = r(q⇃a∗) = qϑa∗. +(2.11) +Lemma 2.12. If C is an ordered ∗-category, then for any a ∈ C, +ϑa : d(a)↓ → r(a)↓ +and +ϑa∗ : r(a)↓ → d(a)↓ +are mutually inverse bijections. +Proof. By symmetry, it is enough to show that ϑaϑa∗ = idd(a)↓. To do so, let p ≤ d(a), and +write q = pϑa = r(p⇃a). By (O6)′ we have p⇃a = a⇂q. Combining this with (O1)′ and (2.11), we +deduce that +p = d(p⇃a) = d(a⇂q) = qϑa∗ = pϑaϑa∗, +as required. +Note that property (O5)′ and its dual (cf. Remark 2.5) can be rephrased in terms of the ϑ +maps. Specifically, if a, b ∈ C (an ordered ∗-category) are such that r(a) = d(b), then (keep- +ing (2.11) and Lemma 2.12 in mind) we have +p⇃(a◦b) = p⇃a◦pϑa⇃b +and +(a◦b)⇂q = a⇂qϑ−1 +b ◦b⇂q +for all p ≤ d(a) and q ≤ r(b). (2.13) +This can be extended to restrictions of compositions with an arbitrary number of terms. In a +sense, the first part of the next result is a formalisation of the previous observation. +Lemma 2.14. Let C be an ordered ∗-category, and let a, b ∈ C and p ∈ vC. +(i) If r(a) = d(b), then ϑa◦b = ϑaϑb. +(ii) ϑp = idp↓. +(iii) If p ≤ d(a), then ϑp⇃a = ϑa|p↓. +Proof. (i). Write p = d(a), q = r(a) = d(b) and r = r(b). Note that +ϑa : p↓ → q↓, +ϑb : q↓ → r↓ +and +ϑa◦b : p↓ → r↓, +so that ϑa◦b and ϑaϑb are defined on the same domain, i.e. on p↓. Now let s ≤ p, and put +t = sϑa = r(s⇃a). Then by (O5)′ we have +sϑa◦b = r(s⇃(a ◦ b)) = r(s⇃a ◦ t⇃b) = r(t⇃b) = tϑb = sϑaϑb. +(ii). By part (i), we have ϑp = ϑp◦p = ϑpϑp, and the result follows since the only idempotent +bijection p↓ → p↓ is the identity. +(iii). Both maps have domain p↓, so it suffices to show that tϑp⇃a = tϑa for all t ≤ p. For this +we use (O4)′ to calculate +tϑp⇃a = r(t⇃p⇃a) = r(t⇃a) = tϑa. +We will also need the following simple property of the ϑ maps: +Lemma 2.15. If C is an ordered ∗-category, and if a ∈ C, then ϑa is order-preserving, in the +sense that +p ≤ q +⇒ +pϑa ≤ qϑa +for all p, q ∈ d(a)↓. +Proof. If p ≤ q ≤ d(a), then p⇃a = p⇃q⇃a ≤ q⇃a, and so r(p⇃a) ≤ r(q⇃a), i.e. pϑa ≤ qϑa. +14 + +In later chapters we will define v-congruences by specifying sets of generating pairs: +Definition 2.16. Suppose C is a category, and Ω ⊆ C × C a set of pairs satisfying +• (u, v) ∈ Ω ⇒ [d(u) = d(v) and r(u) = r(v)]. +The (v-)congruence generated by Ω, denoted Ω♯, is the least congruence on C containing Ω. +Specifically, we have (a, b) ∈ Ω♯ if and only if there is a sequence +a = a1 → · · · → ak = b +such that for each 1 ≤ i < k we have +ai = bi ◦ ui ◦ ci +and +ai+1 = bi ◦ vi ◦ ci +for some bi, ci ∈ C and (ui, vi) ∈ Ω ∪ Ω−1. +(Here Ω−1 = {(v, u) : (u, v) ∈ Ω}.) +The next result shows how conditions (C4) and (C5) regarding a congruence ≈ = Ω♯ can be +deduced from properties of its generating set Ω. In the case of (C5) we utilise the above maps ϑa. +Lemma 2.17. Suppose ≈ = Ω♯ is a v-congruence on an ordered ∗-category C. +(i) If Ω satisfies the condition +(a, b) ∈ Ω +⇒ +a∗ ≈ b∗, +(2.18) +then ≈ satisfies condition (C4). +(ii) If Ω satisfies the condition +(a, b) ∈ Ω +⇒ +ϑa = ϑb +and +p⇃a ≈ p⇃b for all p ≤ d(a), +(2.19) +then ≈ satisfies condition (C5). +Proof. We just prove the second part, as the first is easier. To show that (C5) holds, suppose +a ≈ b, and let the ai, bi, ci, ui, vi be as in Definition 2.16. Of course it suffices to show that +p⇃ai ≈ p⇃ai+1 for each i. For this we define +q = pϑbi +and +r = qϑui. +Since ϑui = ϑvi by (2.19), we also have r = qϑvi. We then use (O5)′ and (2.19) to calculate +p⇃ai = p⇃(bi ◦ ui ◦ ci) = p⇃bi ◦ q⇃ui ◦ r⇃ci ≈ p⇃bi ◦ q⇃vi ◦ r⇃ci = p⇃(bi ◦ vi ◦ ci) = p⇃ai+1. +2.3 +Regular ∗-semigroups +As we have already mentioned, our primary interest is in the class of regular ∗-semigroups, as +defined by Nordahl and Scheiblich in [114]. Here we revise their definition, and list some of their +basic properties. The main new results will be proved later in the paper, but we will give a few +‘appetisers’ here to motivate these later results. +Definition 2.20. A regular ∗-semigroup is a semigroup S with a unary operation ∗ : S → S : s �→ s∗ +satisfying +(a∗)∗ = a = aa∗a +and +(ab)∗ = b∗a∗ +for all a, b ∈ S. +15 + +Note that a regular ∗-semigroup S is considered to be an algebra of type (2, 1), i.e. ∗ is a basic +operation of S. There are semigroups with multiple unary operations, giving rise to pairwise +non-isomorphic regular ∗-semigroups [126]. +Any inverse semigroup is a regular ∗-semigroup, +with the involution being inversion, a∗ = a−1. In fact, it has long been known [127] that inverse +semigroups are precisely the regular ∗-semigroups satisfying the additional identity +aa∗bb∗ = bb∗aa∗ +for all a, b ∈ S. +(2.21) +As we will see, inverse semigroups represent a somewhat ‘degenerate’ case in the theory we +develop; see Chapter 9. +Arguably the most important families of non-inverse regular ∗-semigroups are the so-called +diagram monoids, including the Brauer monoids [10], Temperley-Lieb monoids [130], partition +monoids [76,98], Motzkin monoids [6] and several others. In Section 2.4 we will look extensively +at the partition monoids, but before this we give the following simple example, coming from a +standard semigroup-theoretic construction; we will return to it on a number of occasions. Another +standard generalisation of the construction is given in Example 5.24 (see also Example 5.21); +more examples of this kind can also be found in [118]. +Example 2.22. Let P be an arbitrary set, and let S = P × P = {(p, q) : p, q ∈ P} be the +cartesian product of two copies of P. We define binary and unary operations on S by: +(p, q)(r, s) = (p, s) +and +(p, q)∗ = (q, p). +It is then routine to check that S is a regular ∗-semigroup. This semigroup is known as the +square band over P. +(More generally, a rectangular band has the form P × Q, for possibly +different sets P and Q, with the same product as above; this is still a semigroup, but need not +have an involution.) +For a regular ∗-semigroup S, we write +E = E(S) = {e ∈ S : e2 = e} +and +P = P(S) = {p ∈ S : p2 = p = p∗}. +So E consists of all idempotents of S. The elements of P are called projections, and of course +we have P ⊆ E. +Projections play a very important role in virtually every study of regular +∗-semigroups in the literature, and this is also true in the current work. The following result +gathers some of the basic properties of idempotents and projections. Proofs can be found in many +places (e.g., [70]), but we give some simple arguments here to keep the paper self-contained. +Lemma 2.23. If S is a regular ∗-semigroup, with sets of idempotents E = E(S) and projections +P = P(S), then +(i) P = {aa∗ : a ∈ S} = {a∗a : a ∈ S}, +(ii) E = P 2 = {pq : p, q ∈ P}, and consequently ⟨E⟩ = ⟨P⟩, +(iii) a∗Pa ⊆ P for all a ∈ S. +Proof. (i). It is enough to show that P = {aa∗ : a ∈ S}. If p ∈ P, then p = pp = pp∗. +Conversely, if a ∈ S, then (aa∗)2 = aa∗aa∗ = aa∗, and (aa∗)∗ = (a∗)∗a∗ = aa∗. +(ii). If e ∈ E, then e = ee∗e = e(ee)∗e = ee∗ · e∗e, with ee∗, e∗e ∈ P, by part (i). Conversely, for +any p, q ∈ P, we have pq = pq(pq)∗pq = pqq∗p∗pq = pqqppq = pqpq = (pq)2. +(iii). If p ∈ P, then a∗pa = a∗p∗pa = (pa)∗pa ∈ P by part (i). +16 + +Comparing (2.21) with Lemma 2.23(i), it follows that inverse semigroups are precisely the +regular ∗-semigroups with commuting projections. In fact, it is easy to see that any idempotent of +an inverse semigroup is a projection, so another equivalent condition for a regular ∗-semigroup S +to be inverse is that E(S) = P(S). +If p, q ∈ P, then it follows from Lemma 2.23(iii) that pqp = p∗qp ∈ P. Thus, for all p ∈ P, +we have a well-defined map +θp : P → P +given by +qθp = pqp +for q ∈ P. +(2.24) +The next result summarises some of the key properties of these maps. +Lemma 2.25. If S is a regular ∗-semigroup, then for any p, q ∈ P we have +pθp = p, +θpθp = θp, +pθqθp = qθp, +θpθqθp = θqθp +and +θpθqθpθq = θpθq. +Proof. The first two follow from the fact that projections are idempotents. The third and fifth +follow from the fact that the product of two projections is an idempotent (cf. Lemma 2.23(ii)). +For the fourth, given any t ∈ P we have +tθqθp = tθpqp = pqp · t · pqp = tθpθqθp. +When S is inverse, commutativity of projections implies that qθp = pθq = pq = qp for +all p, q ∈ P. This then leads to a significant simplification in the theory developed below, as we +will explore in detail in Chapter 9. +Next we define a relation ≤ on P = P(S) by +p ≤ q +⇔ +p = pθq = qpq +for p, q ∈ P. +(2.26) +Note that p ≤ q precisely when q is a left and right identity for p. Using this it is easy to see +that ≤ is a partial order. It is also possible to deduce this fact from the identities listed in +Lemma 2.25, as we do in Lemma 3.7 below. In fact, since p = qpq ⇔ p = pq = qp, and since +p = pq ⇒ p = p∗ = (pq)∗ = q∗p∗ = qp, +(2.27) +and conversely by symmetry, it follows that +p ≤ q +⇔ +p = pq +⇔ +p = qp. +(2.28) +The above properties of projections, and their associated θ mappings, will be of fundamen- +tal importance in all that follows. +In particular, even though P = P(S) is generally not a +subsemigroup of the regular ∗-semigroup S, we can regard it as a unary algebra, whose basic +(unary) operations are the θp (p ∈ P). In Chapter 3, we take the properties from Lemma 2.25 +as the axioms for what we will call a projection algebra (see Definition 3.1). Ultimately, we will +see that (abstract) projection algebras are precisely the unary algebras of projections of regular +∗-semigroups. This has in fact been known for some time (see for example [72]), but our new +approach leads to another way to see this. When S is inverse, P is simply a subsemigroup of S, +and there is no real advantage in considering P as a unary algebra. In a sense, a projection +algebra is the ‘∗-analogue’ of the semilattice of idempotents of an inverse semigroup. +As a foreshadowing of things to come, we explain now how to use the projections of a regular +∗-semigroup S to construct a groupoid G = G(S). The groupoid G has the same underlying set +as S, and the (partially defined) composition in G is a restriction of the (totally defined) product +in S. Roughly speaking, G remembers only the ‘nice’ or ‘easy’ products; see Section 2.4 for some +examples justifying our use of these words. +17 + +As it happens, the construction of G = G(S) is exactly the same as in the inverse case; see +[87, Chapter 4]. However, we will see that the groupoid G(S) is not a total invariant of the +regular ∗-semigroup S, unlike the case for inverse semigroups; see Section 5.2 for more details. +The definition of G = G(S) begins with the following neat equational (rather than existential) +characterisation of Green’s relations on a regular ∗-semigroup. +Lemma 2.29. If S is a regular ∗-semigroup, and if a, b ∈ S, then +a R b ⇔ aa∗ = bb∗ +and +a L b ⇔ a∗a = b∗b. +Proof. Even though this has been proved in a number of places, we prove the statement con- +cerning R to keep the paper self contained. The statement for L is of course dual. +Suppose first that a R b, so that a = bx for some x ∈ S. Using (2.28), it then follows that +aa∗ = bxa∗ = bb∗bxa∗ = bb∗aa∗ +⇒ +aa∗ ≤ bb∗. +By symmetry, we also have bb∗ ≤ aa∗, so that aa∗ = bb∗. +Conversely, if aa∗ = bb∗, then a = aa∗a = b(b∗a), and similarly b = a(a∗b), so that a R b. +For a regular ∗-semigroup S, and for a ∈ S, we write +d(a) = aa∗ +and +r(a) = a∗a. +Both of these elements are projections (cf. Lemma 2.23(i)), and the identity a = aa∗a gives +a = d(a) · a = a · r(a). +Moreover, Lemma 2.29 says that +a R b ⇔ d(a) = d(b) +and +a L b ⇔ r(a) = r(b). +Since p = pp∗ = p∗p for any projection p, we have d(p) = r(p) = p for such p. It then follows +from Lemma 2.29 that the R- and L -classes of p are given by +Rp = {a ∈ S : aa∗ = pp∗} = {a ∈ S : d(a) = p} +and similarly +Lp = {a ∈ S : r(a) = p}. +(2.30) +Lemma 2.31. If S is a regular ∗-semigroup, and if a, b ∈ S are such that r(a) = d(b), then +d(ab) = d(a) +and +r(ab) = r(b). +Proof. We just prove the first claim, as the second is symmetrical. If r(a) = d(b), then a∗a = bb∗, +and then +d(ab) = ab(ab)∗ = abb∗a = aa∗aa∗ = aa∗ = d(a). +This allows us to make the following definition. +Definition 2.32. Given a regular ∗-semigroup S, we define the category G = G(S) as follows. +• The object set of G is vG = P = P(S). +• For a ∈ G we have d(a) = aa∗ and r(a) = a∗a. +• For a, b ∈ G with r(a) = d(b), we have a ◦ b = ab. +18 + +For projections p, q ∈ P, it follows from (2.30) that +G(p, q) = {a ∈ S : d(a) = p, r(a) = q} = Rp ∩ Lq. +Since D = R ◦ L , such a morphism set is non-empty precisely when p D q. It follows that the +connected components of G = G(S) are in one-one correspondence with the D-classes of S. +Since the underlying set of G = G(S) is S, the unary operation ∗ : S → S can be thought of +as a map ∗ : G → G. It is a routine matter to verify conditions (I1)–(I4) from Definition 2.1, so +we have the following: +Proposition 2.33. If S is a regular ∗-semigroup, then the category G = G(S) is a groupoid, with +inversion given by −1 = ∗. +Remark 2.34. We will see later that the groupoid G = G(S) associated to a regular ∗- +semigroup S has rather a lot more structure than is evident at this stage. For example, the +fact that the object set vG = P is a (so-called) projection algebra will allow us to construct an +order on G. This ordering can then be used to reduce the calculation of an arbitrary product ab +in S to an associated composition in G: +ab = a′ ◦ e ◦ b′, +(2.35) +where a′ ≤ a and b′ ≤ b, and where e ∈ G is a special morphism r(a′) → d(b′). Although we do +not need to understand the ordering on G yet, we can at least verify that a composition of the +form (2.35) does indeed exist. Specifically, let us write +p = r(a) = a∗a +and +q = d(b) = bb∗, +noting that a = ap and b = qb. We also define the additional projections +p′ = qθp = pqp +and +q′ = pθq = qpq. +Then since pq is an idempotent, by Lemma 2.23(ii), we have p′q′ = (pq)3 = pq, and so +ab = ap · qb = a · p′q′ · b = ap′ · p′q′ · q′b. +Then with a′ = ap′, e = p′q′(= pq) and b′ = q′b, one can check that +r(a′) = p′ = d(e) +and +r(e) = q′ = d(b′). +For example, +r(a′) = (ap′)∗ap′ = (apqp)∗apqp = pqpa∗apqp = pqp · p · pqp = pqpqp = pqp = p′. +Thus, continuing from above, it follows that indeed +ab = ap′ · p′q′ · q′b = a′ · e · b′ = a′ ◦ e ◦ b′. +In what follows, we will think of a′ = ap′ and b′ = q′b as ‘restrictions’ a′ = a⇂p′ and b′ = q′⇃b in +the groupoid G. +Remark 2.36. There is a rather subtle point regarding Remark 2.34 that is far from obvious at +this stage, but ought to be mentioned now. It may seem as if we are implying that the structure +of the regular ∗-semigroup S is completely determined by that of the groupoid G = G(S). That +is, if we are given a complete description of the groupoid G, including its composition, inversion, +ordering, and the (so-called) projection algebra structure of its object set P = vG, then it might +seem that we ought to be able to construct the entire multiplication table of S. However, this +is far from the truth. Indeed, in Section 5.2, we will give examples of non-isomorphic regular +19 + +∗-semigroups S1 and S2 with exactly the same groupoids G(S1) = G(S2). (We postpone the +definition of these semigroups, as we have already gone on a somewhat lengthy tangent, and +also since we will only be able to fully appreciate their properties once we have developed more +theory.) +If the reader is worried that this contradicts the previous remark, the source of the subtlety +is in the precise identity of the element e from (2.35) that allowed us to reduce a product ab in S +to a composition in G: +ab = a⇂p′ ◦ e ◦ q′⇃b. +(On the other hand, the projections p′, q′ can be found directly using the projection algebra +structure of P = vG.) We were able to ‘locate’ this element e in Remark 2.34 becase we began +with full knowledge of the semigroup S; we simply took e = pq = p′q′. However, if we begin +with an ordered groupoid G, it is not so obvious what this element e should be, even if we know +that G = G(S) for some (unknown) regular ∗-semigroup S. Certainly e should be a morphism +p′ → q′, but when p′ ̸= q′, it is not so easy to distinguish any such morphism. Getting around this +problem is one of the main sources of difficulty encountered in the current work. Our eventual +solution utilises what we will call the ‘chain groupoid’ C = C (P) associated to an (abstract) +projection algebra P, and a certain ‘evaluation map’ ε : C → G. We will find our element e in +the image of this map. +Remark 2.37. Before moving on, it is worth commenting on another related notion, namely +the so-called trace of a semigroup. This goes back to a classical result of Miller and Clifford. +Among other things, [106, Theorem 3] says that for elements a, b of a semigroup S, we have +ab ∈ Ra ∩ Lb +⇔ +La ∩ Rb contains an idempotent. +Such products ab ∈ Ra ∩ Lb are often called trace products in the literature. We can then define +a partial binary operation ⊙ on S by +a ⊙ b = +� +ab +if La ∩ Rb contains an idempotent +undefined +otherwise, +and the resulting partial algebra (S, ⊙) is often called the trace of S. Following Definition 2.32, +it is not hard to see that in a regular ∗-semigroup S, a composition a ◦ b exists precisely when +La ∩ Rb contains a projection, which must of course be a∗a = bb∗, i.e. r(a) = d(b). +Since +projections are idempotents, it follows that a ◦ b being defined forces a ⊙ b to be defined, and +then of course a ◦ b = a ⊙ b = ab, meaning that ◦ ⊆ ⊙, i.e. that ⊙ is an extension of ◦. We have +already noted that S is inverse if and only if P(S) = E(S), and it now quickly follows that this +is also equivalent to having ◦ = ⊙. In this inverse case, the trace (S, ⊙) is then precisely the +groupoid G(S) from Definition 2.32, but this is not true of arbitrary regular ∗-semigroups. +As we have already mentioned, projections have played a crucial role in virtually all studies +of regular ∗-semigroups. +Of particular significance in papers such as [22, 33, 39] are pairs of +projections p, q ∈ P = P(S) that are mutual inverses, i.e. pairs that satisfy p = pqp and q = qpq. +This is very much the case in the current work: so much so, in fact, that we define relations ≤F +and F on P by +p ≤F q +⇔ +p = pqp = qθp +and +p F q +⇔ +[p ≤F q and q ≤F p]. +(2.38) +We think of pairs of F-related projections as friends (hence the symbol F). It is clear that ≤F +and F are both reflexive, and that F is symmetric. But neither ≤F nor F is transitive in +general (and neither is friendship in ‘real life’); for an example, see Section 2.4. It is easy to +see that p ≤ q +⇒ +p ≤F q, where ≤ is the partial order in (2.26). It follows quickly from +Lemma 2.29 that +p F q +⇔ +p R pq L q +⇔ +p L qp R q. +(2.39) +20 + +It of course follows from this that +p F q +⇒ +p D q, +(2.40) +though the converse does not hold in general; again see Section 2.4. Combining (2.39) with +[68, Proposition 2.3.7], it follows that +p F q +⇔ +Lp ∩ Rq contains an idempotent +⇔ +Rp ∩ Lq contains an idempotent, +and of course these idempotents are +qp ∈ Lp ∩ Rq +and +pq ∈ Rp ∩ Lq. +This all means that a pair of F-related projections p, q ∈ P(S) generates a 2 × 2 rectangular +band in S: +p +pq +qp +q +One can define a graph Γ(S) with vertex set P = P(S), and with (undirected) edges {p, q} +whenever p ̸= q and p F q. Connectivity properties of (certain subgraphs of) these graphs +played a vital role in [39], in studies of idempotent-generated regular ∗-semigroups and their +ideals. We will say more about this in Section 2.4. +The (standard) proof of Lemma 2.23(ii) given above involved showing that any idempotent +e ∈ E = E(S) of a regular ∗-semigroup S satisfies e = ee∗ · e∗e, with ee∗, e∗e ∈ P = P(S). It +quickly follows that the idempotents e, e∗ ∈ E(S) also generate a 2 × 2 rectangular band in S: +e +ee∗ +e∗e +e∗ +This is in fact a characterisation of the biordered sets arising from regular ∗-semigroups, in a +sense made precise in [113, Corollary 2.7]. In any case, it quickly follows that ee∗ F e∗e for +any e ∈ E, so this shows that any idempotent is a product of two F-related projections, and it +follows that not only do we have E = P 2 (cf. Lemma 2.23(ii)), but in fact +E = {pq : (p, q) ∈ F}. +Actually, such factorisations for idempotents are unique. Indeed, if e = pq, where e ∈ E and +(p, q) ∈ F, then +ee∗ = pq(pq)∗ = pqq∗p∗ = pqqp = pqp = p +and similarly +e∗e = q. +(2.41) +Consequently, we have |E| = |F|. A number of studies have enumerated the idempotents in +various classes of regular ∗-semigroups [20,21,84]. The identity |E| = |F| played an implicit role +in [21]. +So idempotents are products of F-related projections. +As an application of the theory +developed in this paper, we will be able to prove the following result, which extends this to +arbitrary products of idempotents, though uniqueness is lost, in general (cf. (2.48)); see Proposi- +tion 4.39(iii) and Remark 4.40. The result might be known, but we are not aware of its existence +in the literature; it does bear some resemblance, however, to a classical result of FitzGerald [48]. +Moreover, we believe it could be useful in studies of idempotent-generated regular ∗-semigroups, +and might have even led to some simpler arguments in existing studies such as [22,38,39]. +21 + +Proposition 2.42. Let S be a regular ∗-semigroup, with E = E(S) and P = P(S). Then for +any a ∈ ⟨E⟩ = ⟨P⟩ we have +a = p1 · · · pk +for some p1, . . . , pk ∈ P with p1 F · · · F pk. +We will leave our preliminary discussion of regular ∗-semigroups here. We will return to them +again in Chapter 5. But for now, we will discuss an important special case. +2.4 +Case study: diagram monoids +The theory of regular ∗-semigroups has seen something of a resurgence in recent years, partly due +to the importance of so-called diagram monoids. Here we recall the definition of these monoids, +and use them to illustrate some of the ideas discussed in the previous section. We will also return +to them at various points during the rest of the paper. Here we focus exclusively on the partition +monoids [76,98], although similar comments could be made for other diagram monoids, such as +(partial) Brauer, Temperley-Lieb and Motzkin monoids [6,10,22,39,100,130]. +Let X be a set, and let X′ = {x′ : x ∈ X} be a disjoint copy of X. The partition monoid +over X, denoted PX, is defined as follows. The elements of PX are the set partitions of X ∪ X′, +and the operation will be defined shortly. A partition from PX will be identified with any graph +on vertex set X ∪ X′ whose connected components are the blocks of the partition. The elements +of X and X′ are called upper and lower vertices, respectively, and are generally displayed in two +parallel rows, as in the various figures below. +To define the product on PX, consider two partitions α, β ∈ PX. Let X′′ = {x′′ : x ∈ X} be +another disjoint copy of X, and define three new graphs: +• α∨, the graph on vertex set X ∪ X′′ obtained from α by changing each lower vertex x′ to x′′, +• β∧, the graph on vertex set X′′ ∪ X′ obtained from β by changing each upper vertex x to x′′, +• Π(α, β), the graph on vertex set X ∪ X′′ ∪ X′, whose edge set is the union of the edge sets +of α∨ and β∧. +The graph Π(α, β) is called the product graph of α and β; it is generally drawn with X′′ as the +middle row of vertices. The product αβ is defined to be the partition of X ∪X′ with the property +that u, v ∈ X ∪ X′ belong to the same block of αβ if and only if u, v are connected by a path +in Π(α, β). +If X = {1, . . . , n} for a non-negative integer n, we write Pn = PX. +Example 2.43. To illustrate the product above, consider the partitions α, β ∈ P6 defined by +α = +� +{1, 2, 3, 1′}, {4, 4′, 5′, 6′}, {5}, {6}, {2′, 3′} +� +, +β = +� +{1, 4′, 6′}, {2, 3}, {4, 5, 6, 1′, 2′, 3′}, {5′} +� +. +Figure 1 illustrates (graphs representing) α and β, their product +αβ = +� +{1, 2, 3, 4′, 6′}, {4, 1′, 2′, 3′}, {5}, {6}, {5′} +� +, +as well as the product graph Π(α, β). Here and elsewhere, vertices are assumed to increase from +left to right, 1, . . . , n, and similarly for (double) dashed vertices. +A reader familiar with partition monoids might protest that we have ‘cheated’ in Exam- +ple 2.43, and that our choice of α and β was too ‘easy’ to fully illustrate the complexities of +calculating products of partitions. This is indeed the case, as the bottom half of α ‘matches’ the +top half of β in a way that can hopefully be understood by examining Figure 1, but which will +be made precise shortly. Before this, we briefly consider a more ‘difficult’ product: +22 + +α = +β = += αβ +Figure 1. Left to right: partitions α, β ∈ P6, the product graph Π(α, β), and the product αβ ∈ P6. +For more information, see Example 2.43. +Example 2.44. Figure 2 gives another product, this time with α, β ∈ P20. One can immediately +see that there is no such ‘matching’ between the bottom of α and the top of β. Rather, to calculate +the product αβ, one needs to follow paths in the product graph Π(α, β), often alternating several +times between edges coming from α or β. For example, to see that {1′, 4′} is a block of αβ, one +needs to trace the following path (or its reverse): +1′ +β +−−→ 1′′ +α +−−→ 2′′ +β +−−→ 3′′ +α +−−→ 4′′ +β +−−→ 4′. +α = +Π(α, β) +β = +αβ = +Figure 2. +Top: partitions α, β ∈ P20, already connected to create the product graph Π(α, β). +Bottom: the product αβ ∈ P20. For more information, see Example 2.44. +We hope that Example 2.44 convinces the reader that products in PX can be ‘messy’. Of +course things get worse when we increase the number of vertices, even more so when X is +infinite [35]. Nevertheless, it is actually quite easy to see that PX is a regular ∗-semigroup. The +involution +∗ : PX → PX : α �→ α∗ +can be defined diagrammatically as a reflection in a horizontal axis; see Figure 3. Formally, α∗ +is obtained from α by swapping dashed and undashed vertices, x ↔ x′. It is not hard to see that +(α∗)∗ = α = αα∗α +and +(αβ)∗ = β∗α∗ +for all α, β ∈ PX. +α = += α∗ +Figure 3. A partition α ∈ P6 (left) and its image α∗ under the involution of P6. +23 + +At this point it is convenient to recall some further notation and terminology. First, we say +a non-empty subset ∅ ̸= A ⊆ X ∪ X′ is: +• a transversal if A ∩ X ̸= ∅ and A ∩ X′ ̸= ∅ (i.e. if A contains both upper and lower vertices), +• an upper non-transversal if A ⊆ X (i.e. if A contains only upper vertices), or +• a lower non-transversal if A ⊆ X′ (i.e. if A contains only lower vertices). +We can then describe a partition α ∈ PX using a convenient two-line notation from [38]. Specif- +ically, we write +α = +�Ai Cj +Bi Dk +� +i∈I, j∈J, k∈K +to indicate that α has transversals Ai ∪ B′ +i (i ∈ I), upper non-transversals Cj (j ∈ J) and +lower non-transversals Dk (k ∈ K). We often abbreviate this to α = +�Ai Cj +Bi Dk +� +, with the indexing +sets I, J and K being implied, rather than explicitly listed. When X is finite we list the blocks +of partitions, viz. α = +�A1 . . . Aq C1 . . . Cs +B1 . . . Bq D1 . . . Dt +� +. Thus, for example, the partition β ∈ P6 from Figure 1 +can be expressed as +β = +� 1 +4, 5, 6 2, 3 +4, 6 1, 2, 3 +5 +� +. +With this notation, the identity of PX is the partition idX = +� +x +x +� +x∈X, and the involution is given +by +α = +�Ai Cj +Bi Dk +� +�→ α∗ = +�Bi Dk +Ai Cj +� +. +Moreover, one can easily see that with α = +�Ai Cj +Bi Dk +� +, and with the notation of Section 2.3, we +have +d(α) = αα∗ = +�Ai Cj +Ai Cj +� +and +r(α) = α∗α = +�Bi Dk +Bi Dk +� +. +Thus, consulting Lemma 2.29, it follows that partitions α, β ∈ PX are R-related (or L -related) +precisely when the ‘top halves’ (or ‘bottom halves’) of α, β ‘match’ in a sense that we again hope +is clear. Nevertheless, this ‘matching’ can be formalised by defining some further notation. +For α ∈ PX, we define the (co)domain and (co)kernel of α by: +dom(α) = {x ∈ X : x belongs to a transversal of α}, +codom(α) = {x ∈ X : x′ belongs to a transversal of α}, +ker(α) = {(x, y) ∈ X × X : x and y belong to the same block of α}, +coker(α) = {(x, y) ∈ X × X : x′ and y′ belong to the same block of α}. +The rank of α, denoted rank(α), is the number of transversals of α. Thus, dom(α) and codom(α) +are subsets of X, ker(α) and coker(α) are equivalences on X, and rank(α) is a cardinal between 0 +and |X|. For example, with α ∈ P6 from Figure 1, we have +dom(α) = {1, 2, 3, 4}, +ker(α) = (1, 2, 3 | 4 | 5 | 6), +codom(α) = {1, 4, 5, 6}, +coker(α) = (1 | 2, 3 | 4, 5, 6), +rank(α) = 2, +where we have indicated equivalences by listing their classes. The various parts of the following +result are contained in [42,51,137], though some of those papers use different terminology. +24 + +Lemma 2.45. For α, β ∈ PX, we have +(i) α R β ⇔ [dom(α) = dom(β) and ker(α) = ker(β)], +(ii) α L β ⇔ [codom(α) = codom(β) and coker(α) = coker(β)], +(iii) α D β ⇔ α J β ⇔ rank(α) = rank(β). +The D = J -classes of PX are the sets +Dµ = Dµ(PX) = {α ∈ PX : rank(α) = µ} +for cardinals 0 ≤ µ ≤ |X|. +Group H -classes contained in Dµ are isomorphic to the symmetric group Sµ. +Let us now return to the example products considered earlier. +Example 2.46. Consider again the partitions α, β ∈ P6 from Example 2.43 and Figure 1. The +projections +r(α) = α∗α +and +d(β) = ββ∗ +are calculated in Figure 4. Since we have r(α) = d(β), we see that α and β are composable in +the groupoid G(P6) from Definition 2.32, so that in fact αβ = α◦β. This explains the ‘matching’ +phenomenon discussed above, and gives the reason for the ‘ease’ of forming the product αβ. +α∗ = +α = +r(α) = +β = +β∗ = +d(β) = +Figure 4. The projections r(α) = α∗α and d(β) = ββ∗, where α, β ∈ P6 are as in Figure 1. For +more information, see Examples 2.43 and 2.46. +Example 2.47. On the other hand, the partitions α, β ∈ P20 from Example 2.44 and Figure 2 +are not composable in the groupoid G = G(P20). Indeed, one can check that the projections +r(α) = α∗α +and +d(β) = ββ∗ +are as shown in Figure 5, and we clearly do not have r(α) = d(β). On the other hand, it will +follow from results of later chapters that there exist certain partitions α′ ≤ α and β′ ≤ β (the ≤ +relation will be defined later) that can ‘almost’ be composed in G. By this we mean that while +we still do not have r(α′) = d(β′), we do have r(α′) F d(β′). (The F relation was defined +in (2.38).) It will follow from the general theory that +α = α′ ◦ ε ◦ β′, +where ε ∈ G is a special element with d(ε) = r(α′) and r(ε) = d(β′). The partitions α′, ε, β′ ∈ P20 +are shown in Figure 6. The reader might like to check that ε is an idempotent of P20, i.e. that ε2 = ε +(but we note that the composition ε ◦ ε does not exist in G). +25 + +r(α) = +d(β) = +Figure 5. The projections r(α) = α∗α and d(β) = ββ∗, where α, β ∈ P20 are as in Figure 2. For +more information, see Examples 2.44 and 2.47. +ε = +α′ = +Π(α′, ε, β′) +β′ = +α′ ◦ ε ◦ β′ = +Figure 6. Top: partitions α′, ε, β′ ∈ P20, with the edges of ε shown in red. Bottom: the composition +α′ ◦ ε ◦ β′ in the groupoid G(P20). Note that α′ ◦ ε ◦ β′ = αβ, where α, β ∈ P20 are as in Figure 2. +For more information, see Examples 2.44 and 2.47. +We conclude this section with a brief discussion of the relations ≤, ≤F and F on the set +P = P(PX) of projections of a partition monoid PX. These relations were defined in (2.26) +and (2.38). First, one can check that the projections of the partition monoid P2 are the six +partitions shown in Figure 7. +Figure 8 shows the partial order ≤ on P(P2), as defined in (2.26); as usual for Hasse dia- +grams, only covering relationships are shown, and the rest can be deduced from transitivity (and +reflexivity). Figure 9 shows the relations ≤F and F on P(P2), as defined in (2.38), and we +remind the reader that these are not transitive. For example, we have +F +F +, +even though the first and third of the above projections are not F-related, or even ≤F-related. +In both Figures 8 and 9, the elements of P = P(P2) have been arranged so that each row +consists of D-related elements; cf. Lemma 2.45. So from top to bottom, the rows correspond to +the projections from the D-classes D2, D1 and D0, where +Di = Di(P2) = {α ∈ P2 : rank(α) = i} +for i = 0, 1, 2. +26 + +Figure 7. The projections of the partition monoid P2. +Figure 8. Hasse diagram of the poset P(P2), with respect to the order ≤ given in (2.26). An arrow +p → q means p ≤ q. +For i = 0, 1, 2, we write Pi = P ∩ Di for the set of projections from Di. The right-hand graph in +Figure 9 is the graph Γ(P2), as defined at the end of Section 2.3. We denote this by Γ = Γ(P2); +so the vertex set of Γ is P = P(P2), and Γ has an undirected edge {p, q} precisely when p ̸= q +and p F q. As in (2.40), F-related projections are D-related. Thus, +Γ = Γ0 ⊔ Γ1 ⊔ Γ2 +decomposes as the disjoint union of the induced subgraphs Γi on the vertex sets Pi, for i = 0, 1, 2. +It is clear that each Γi is connected, but this is not necessarily the case for an arbitrary regular +∗-semigroup S. +In general, for a regular ∗-semigroup S, we still have the graph Γ(S), and we still have the +decomposition +Γ(S) = +� +D∈S/D +Γ(D), +where Γ(D) is the induced subgraph of Γ(S) on vertex set P(S) ∩ D, for each D-class D of S. +However, the subgraphs Γ(D) need not be connected in general. For example, if S is inverse, then +the F-relation is trivial; it follows that each Γ(D) is discrete (has empty edge set), and hence is +disconnected if D contains more than one idempotent. On the other hand, it follows from results +of [39] that the graphs Γ(D) are connected for any D-class D of a finite partition monoid Pn, +and this turns out to be equivalent to certain facts about minimal idempotent-generation of the +proper ideals of Pn. We have seen this connectivity property in the case n = 2, above. Figure 10 +shows the graph Γ(P3), produced using the Semigroups package for GAP [52, 107]. From left +to right, the connected components of the graph Γ(P3) in Figure 10 are the induced subgraphs +Γ(D3), Γ(D2), Γ(D1) and Γ(D0). +27 + +Figure 9. Left: the relation ≤F on P(P2); an arrow p → q means p ≤F q. Right: the relation F +on P(P2); an edge p − q means p F q. These relations are given in (2.38). In both diagrams, loops +are omitted. +Figure 10. The graph Γ(P3), produced by GAP. +As a special case, the induced subgraph Γ(Dn−1) of the graph Γ(Pn) corresponding to the +D-class Dn−1 = Dn−1(Pn) has a very interesting structure. First, one can easily check that the +projections of Pn contained in Dn−1 have the form +· · · +· · · +· · · +· · · +1 +i +n +πi = +· · · +· · · +· · · +· · · +· · · +· · · +1 +j +k +n +and +πjk = +for each 1 ≤ i ≤ n and 1 ≤ j < k ≤ n. It is also easy to check that the only non-identical +F-relations among these projections are +πi F πij F πj. +The graph Γ(Dn−1) is shown in Figure 11 in the case n = 5. In the figure, vertices representing +projections πi or πjk are labelled simply with the subscripts i or jk. +28 + +1 +2 +3 +4 +5 +12 +13 +14 +15 +23 +24 +25 +34 +35 +45 +Figure 11. The graph Γ(D4), where D4 = D4(P5). +The idempotent-generated subsemigroup E(PX) = ⟨E(PX)⟩ = ⟨P(PX)⟩ of a partition monoid PX +was described for finite and infinite X in [34] and [38], respectively. One of the main results of [34] +is that finite E(Pn) is (minimally) generated as a monoid by the set +Ω = {πi, πjk : 1 ≤ i ≤ n, 1 ≤ j < k ≤ n} +of all projections of rank n − 1, and that in fact +E(Pn) = {idn} ∪ Sing(Pn), +where here Sing(Pn) = Pn \ Sn is the singular ideal of Pn, consisting of all non-invertible ele- +ments. The second main result was a presentation for Sing(Pn) in terms of the above generating +set Ω. For example, one of the defining relations was +πiπijπjπjkπkπkiπi = πiπikπkπkjπjπjiπi +for distinct 1 ≤ i, j, k ≤ n, +(2.48) +where here we use symmetrical notation πst = πts. Examining Figure 11, one can see that the +two words in this equation represent two different triangular paths in the graph Γ(Dn−1): +i ⇝ j ⇝ k ⇝ i +and +i ⇝ k ⇝ j ⇝ i. +(In representing the above paths we have omitted the intermediate vertices, so i ⇝ j is shorthand +for i → ij → j, and so on.) Since paths in this graph correspond to lists of sequentially F-related +projections, the above words represent factorisations of the form described in Proposition 2.42, +and the corresponding tuples (πi, πij, πj, . . .) and (πi, πik, πk, . . .) are examples of what we will +later call P-paths; see Section 3.2. +The two factorisations represent the following partition, +pictured in the case i < j < k: +· · · +· · · +· · · +· · · +· · · +· · · +· · · +· · · +1 +i +j +k +n +Much more could be said, but we will conclude our preliminary discussion of partition monoids +here. We will return to them at various stages throughout the text. +29 + +Part I +Structure +This first part of the paper is devoted to our main structural result, Theorem 6.7. This theorem +states that the category of regular ∗-semigroups is isomorphic to the category of so-called chained +projection groupoids. Such groupoids play the same role in the theory of regular ∗-semigroups +that inductive groupoids play in inverse semigroup theory. Roughly speaking, we do not assume +that the object sets of our groupoids are semilattices (as per inductive groupoids), but rather +projection algebras. These are unary algebras that abstractly model the projections of regular +∗-semigroups, together with their ‘conjugation actions’ as in (2.24). In fact, we will see that a +groupoid simply having a projection algebra for its object set is not enough. Rather, our chained +projection groupoids are pairs (G, ε), where G is a groupoid with vG = P a projection algebra, +and where ε is a certain special functor that encapsulates the strong relationship between the +groupoid structure of G and the algebra structure of P. The domain of the functor ε is another +groupoid, C = C (P), which we call the chain groupoid of P. In a sense, C models the behaviour +of P at the most ‘free’ level, somehow recording precisely the information that holds in every +regular ∗-semigroup with projection algebra P. The full extent of this ‘free-ness’ will be explored +in more detail in the second part of the paper; see Chapter 7. +This part of the paper contains four chapters: +• Chapter 3 introduces projection algebras, and their chain groupoids. +• Chapter 4 defines chained projection groupoids, and shows how to associate a regular ∗- +semigroup to each such groupoid; see Theorem 4.36. +• Chapter 5 goes in the reverse direction, and shows how a regular ∗-semigroup gives rise to a +chained projection groupoid; see Theorem 5.19. +• Chapter 6 contains the main result, Theorem 6.7, which shows that the constructions of +Chapters 4 and 5 are in fact mutually inverse isomorphisms between the categories of regular +∗-semigroups and chained projection groupoids. +The introduction of each chapter contains a detailed summary of its structure, and of the results +it contains. +3 +Projection algebras +One of the key ideas in this paper is that of a projection algebra. Such an algebra consists of a +set P, along with a family θp (p ∈ P) of unary operations, one for each element of P. These +operations are required to satisfy certain axioms, as listed in Definition 3.1 below, and are meant +to abstractly model the unary algebras of projections of regular ∗-semigroups. This is in fact +not a new concept. Indeed, projection algebras have appeared in a number of settings, under a +variety of names, including the P-groupoids of Imaoka [72], the P-sets of Yamada [138], (certain +special) P-sets of Nambooripad and Pastijn [113], and the (left and right) projection algebras +of Jones [73]. (We also mention Yamada’s p-systems [139]; these are defined in a very different +way, but for a similar purpose. Strictly speaking, Yamada’s P-sets are different to Imaoka’s +P-groupoids, but are ultimately equivalent.) We prefer Jones’ terminology ‘projection algebras’, +since these are indeed algebras, not just sets, though we note that Jones considered binary +algebras rather than Imaoka’s unary algebras, which we use here; see Remark 3.2. We also only +use the term ‘groupoid’ in its usual categorical sense. Imaoka used it in one of its less-common +30 + +meanings, stemming from the fact that a projection algebra also has a partially defined binary +operation; curiously this partial operation plays no role in the current work. +Although the concept of a projection algebra is not new, here we build several new categorical +structures on top of such algebras. The most general situation is covered in Chapter 4 below, +where we have an ordered groupoid G whose object set vG = P is a projection algebra. The +current chapter lays the foundation for this, by showing how to construct various categories +directly from a projection algebra. The main such construction is the chain groupoid C = C (P); +this will be used in the definition of the groupoids G in Chapter 4, but we will uncover its deeper +categorical significance in Chapter 7. +We begin in Section 3.1 by giving/recalling the definition of projection algebras (see Def- +inition 3.1), and establishing some of their important properties. In Sections 3.2 and 3.4, re- +spectively, we introduce the path category P = P(P) and the chain groupoid C = C (P) of a +projection algebra P (see Definitions 3.17 and 3.39). The groupoid C is defined as a quotient +of the category P by a certain congruence ≈, whose definition requires the somewhat-technical +notion of linked pairs, which are the subject of Section 3.3. +3.1 +Definitions and basic properties +Here then is the key definition: +Definition 3.1. A projection algebra is a set P, together with a collection of unary opera- +tions θp (p ∈ P) satisfying the following axioms, for all p, q ∈ P: +(P1) pθp = p, +(P2) θpθp = θp, +(P3) pθqθp = qθp, +(P4) θpθqθp = θqθp, +(P5) θpθqθpθq = θpθq. +The elements of a projection algebra are called projections. +Strictly speaking, one should refer to ‘a projection algebra (P, θ)’, but we will almost always +use the symbol θ to denote the unary operations of such an algebra, and so typically refer to +‘a projection algebra P’. On the small number of occasions we need to refer simultaneously to +more than one projection algebra, we will be careful to distinguish the operations. +Remark 3.2. It is worth noting at the outset that the collection of unary operations θp (p ∈ P) +could be replaced by a single binary operation ⋄, defined by q ⋄p = qθp. For example, Jones took +this binary approach in [73], in his study of the more general P-restriction semigroups, though +his preferred symbol was ⋆, which we reserve for other uses throughout the paper. We prefer the +unary approach of Imaoka [72], however, for three main reasons: +• First, we feel that the meaning of (some of) axioms (P1)–(P5) is clearer in the unary context. +For example, the binary form of (P4) is +r ⋄ (q ⋄ p) = ((r ⋄ p) ⋄ q) ⋄ p +for all p, q, r ∈ P. +As we will discuss in more detail in Remark 4.9, axiom (P4) can be thought of as a rule for +‘iterating’ the unary operations, and we feel that this intuition is somewhat lost in the binary +form of (P4) above, although this is of course entirely subjective. (We also note that Jones’ +axioms are different to (P1)–(P5), though they are equivalent, as he explains in [73, Section 7].) +31 + +• The non-associativity of ⋄ means that brackets are necessary when working with ⋄-terms. On +the other hand, associativity of function composition allows us to dispense with bracketing +when using the θ maps, and this is a significant advantage when dealing with lengthy terms. +• In later chapters, we will consider groupoids G whose object sets are projection algebras, vG = P. +One of the key tools when studying such groupoids are certain maps +Θa = θd(a)ϑa : P → P +that are built from the unary operations of P and the maps ϑa : d(a)↓ → r(a)↓ defined +in (2.10). These Θ maps could certainly be defined in terms of ⋄ and the ϑ maps, viz. +pΘa = (p ⋄ d(a))ϑa, +but their definition as a composition Θa = θd(a)ϑa is more direct, and leads to more suc- +cinct statements and proofs, as for example with Proposition 4.6. The direct definition as a +composition also helps to clarify the very formulation of our so-called (chained) projection +groupoids; see Definitions 4.8 and 4.22. +As we have already observed, the axioms (P1)–(P5) are abstractions of the properties of +projections of regular ∗-semigroups; cf. Lemma 2.25. +It will transpire (see especially Chap- +ters 7 and 8) that abstract projection algebras are precisely the projection algebras of regular +∗-semigroups, as is already known [72]. For the eager reader, Example 3.3 below provides a +construction (without proof at this stage) of a regular ∗-semigroup with projection algebra P. +In any case, the reader can keep in mind that (abstract) projection algebras are supposed to +‘model’ the behaviour of algebras of projections of regular ∗-semigroups; thus, one can interpret +every result in this chapter in the context of regular ∗-semigroups, and can gain an intuition for +their meaning in that context. +Example 3.3. Consider a projection algebra P. Since an operation θp is a map P → P, we +can think of it as an element of the full transformation monoid TP . (This is the monoid of all +maps P → P, under composition.) The maps θp generate the subsemigroup +SP = ⟨θp : p ∈ P⟩ = {θp1 · · · θpk : k ≥ 1, p1, . . . , pk ∈ P} ≤ TP . +So SP is an idempotent-generated semigroup (cf. (P2)), but it is generally not a regular ∗- +semigroup. However, we can alter it to create a regular ∗-semigroup. Indeed, first we write T op +P +for the opposite semigroup to TP . So T op +P +has the same underlying set as TP , and its product ⋆ +is given by α ⋆ β = βα. We then consider the subsemigroup of the direct product TP × T op +P +generated by all pairs (θp, θp): +FP = ⟨(θp, θp) : p ∈ P⟩ = {(θp1 · · · θpk, θpk · · · θp1) : k ≥ 1, p1, . . . , pk ∈ P} ≤ TP × T op +P . +It turns out that FP is a regular ∗-semigroup with projection algebra (isomorphic to) P. In +fact, in Chapter 8 we will see that (up to isomorphism) FP is the unique idempotent-generated +fundamental regular ∗-semigroup with projection algebra P; see Theorem 8.24 and Remark 8.25. +Note that FP is a subdirect product of SP and Sop +P . +For the rest of this section we fix a projection algebra P. In what follows, for x, y ∈ P, we +write x =1 y to indicate that x = y by an application of (P1), and similarly for =2, and so on. +Lemma 3.4. For any p1, . . . , pk, q, r ∈ P we have +θqθp1···θpk = θpk · · · θp1θqθp1 · · · θpk. +Proof. This follows by iterating (P4). +32 + +A number of relations on P will play a crucial role in all that follows. The first of these is +defined by +p ≤ q +⇔ +p = pθq. +(3.5) +By (P2), it quickly follows that +p ≤ q +⇔ +p = rθq +for some r ∈ P. +(3.6) +Lemma 3.7. ≤ is a partial order on P. +Proof. Reflexivity follows from (P1). For anti-symmetry, suppose p ≤ q and q ≤ p, so that +p = pθq and q = qθp. Then +p =1 pθp = pθqθp =3 qθp = q. +For transitivity, suppose p ≤ q and q ≤ r, so that p = pθq and q = qθr. Then +p = pθq = pθqθr =4 pθrθqθr. +By (3.6), this implies p ≤ r. +By (3.6), the image of θp : P → P is precisely +im(θp) = p↓ = {q ∈ P : q ≤ p}, +(3.8) +the down-set of p in the poset (P, ≤). +Lemma 3.9. If p ≤ q, then θp = θpθq = θqθp. +Proof. Since p = pθq, we have θp = θpθq =4 θqθpθq. The claim then quickly follows from (P2). +An equally important role will be played by two further relations, ≤F and F, defined as +follows. For p, q ∈ P, we say that +p ≤F q +⇔ +p = qθp. +(3.10) +By (P1), ≤F is reflexive, but it need not be transitive. We also define +F = ≤F ∩ ≥F, +(3.11) +which is the largest symmetric (and reflexive) relation contained in ≤F. So +p F q +⇔ +p = qθp and q = pθq. +Lemma 3.12. For any p, q ∈ P, +(i) pθq ≤F p, +(ii) p ≤ q ⇒ p ≤F q, +(iii) p ≤F q ⇒ θp = θpθqθp. +(iv) p F q ⇒ θp = θpθqθp and θq = θqθpθq. +Proof. (i). We have pθpθq =4 pθqθpθq =3 qθpθq =3 pθq. +(ii). If p ≤ q, then p = pθq, so qθp = qθpθq =4 qθqθpθq =1 qθpθq =3 pθq = p. +(iii) and (iv). These follow immediately from (P4). +33 + +Remark 3.13. The partial order ≤ from (3.5) was used by Imaoka in [72], where it was defined +by +p ≤ q +⇔ +p = pθq = qθp. +Note that part (ii) of the lemma just proved means that the ‘= qθp’ part of Imaoka’s definition +is superfluous. Jones also used the order ≤ in [73], defined exactly as in (3.5), albeit in binary +form, p ≤ q ⇔ p = p ⋄ q (cf. Remark 3.2). We are not aware of any previous use of the ≤F +relation in the literature, however, despite its central importance in the current work. +The following simple result will be crucial in later chapters. Roughly speaking, it says that +any pair of projections p, q ∈ P sit above an F-related pair p′, q′ ∈ P. +Lemma 3.14. Let p, q ∈ P be arbitrary, and let p′ = qθp and q′ = pθq. Then +p′ ≤ p, +q′ ≤ q +and +p′ F q′. +Proof. We obtain p′ ≤ p and q′ ≤ q directly from (3.6). We also have +p′θq′ = qθpθpθq =4 qθpθqθpθq =5 qθpθq =3 pθq = q′ +and similarly +q′θp′ = p′. +Although ≤F is not transitive, ≤ and ≤F have some ‘transitivity-like’ properties in combi- +nation: +Lemma 3.15. If p, q, r ∈ P are such that p ≤ q ≤F r or p ≤F q ≤ r, then +(i) p ≤F r, +(ii) p F pθr. +Proof. Throughout the proof we assume that p ≤ q ≤F r. The arguments for the case of +p ≤F q ≤ r are similar, and are omitted. +Since p ≤ q and q ≤F r, we have p = pθq and q = rθq. Since also p ≤F q by Lemma 3.12(ii), +we also have p = qθp. +(i). We must show that p = rθp. Since p ≤ q, Lemma 3.9 gives θp = θqθp. Combining all of the +above gives rθp = rθqθp = qθp = p. +(ii). We have pθr ≤F p by Lemma 3.12(i), so we just need to show that p ≤F pθr, i.e. that +p = (pθr)θp. This follows from the axioms and the above-mentioned consequences of p ≤ q ≤F r: +pθrθp =3 rθp = rθpθq =4 rθqθpθq = qθpθq =3 pθq = p. +It remains an open problem to determine whether a projection algebra P (including its θ +operations) can be somehow characterised by the purely order-theoretic properties of the relations +≤ and ≤F. The next result is included in case it is of use in such a characterisation. +Lemma 3.16. If p, q ∈ P are such that p ≤F q, then we have p F p′ ≤ q for some p′ ∈ P. +Proof. It is a routine matter to verify that p′ = pθq has the stated properties. +3.2 +The path category +For the duration of this section we fix a projection algebra P, as in Definition 3.1, including the +operations θp (p ∈ P), and the relations ≤, ≤F and F. +Roughly speaking, we wish to provide an abstract setting in which we can think about ‘prod- +ucts’ of projections p1 · · · pk, even though P itself does not have a binary operation. Guided by +the (as-yet unproved) Proposition 2.42, we are (for now) solely interested in such ‘products’ in +34 + +the case that p1 F · · · F pk. The path category P (see Definition 3.17) is the first attempt +at doing so; it represents such a ‘product’ as a tuple (p1, . . . , pk), which is considered as a mor- +phism p1 → pk. In Section 3.4, we will define the chain groupoid C (see Definition 3.39) as a +quotient C = P/≈ by a certain congruence ≈. The intuition here is that the chain groupoid +records more information about such products than the factors alone, but only such informa- +tion that is present in any regular ∗-semigroup with projection algebra P. For example, (p, p) +and (p) should always represent the same ‘product’, since projections are idempotents; so too +should (p, q, p) and (p) when p F q (cf. (2.38)). These pairs of paths (and one other family of +pairs) are taken as generators for the congruence ≈. But before we get ahead of ourselves, here +is the main definition for the current section: +Definition 3.17. A P-path in a projection algebra P (cf. Definition 3.1) is a tuple +p = (p1, p2, . . . , pk) ∈ P k +for some k ≥ 1, such that p1 F p2 F · · · F pk. +(Since F is not transitive, this does not imply that the pi are all F-related.) We say that p is +a P-path from p1 to pk, and write d(p) = p1 and r(p) = pk. We identify each p ∈ P with the +path (p) of length 1. +The path category of P is the ∗-category P = P(P) of all P-paths, with object set vP = P, +under the following operations: +• For p = (p1, . . . , pk) and q = (q1, . . . , ql) with pk = q1, we define +p ◦ q = (p1, . . . , pk−1, pk = q1, q2, . . . , ql). +• For p = (p1, . . . , pk) ∈ P, we define +prev = (pk, . . . , p1), +the reverse of p. (It is convenient to write prev instead of p∗, and it is clear that conditions +(I1)–(I3) from Definition 2.1 all hold.) +So a morphism set P(p, q) consists of all P-paths from p to q. The identities are the paths +of the form p ≡ (p). Although P is a ∗-category, it is not a groupoid, as when p has length +k ≥ 2, p ◦ prev has length 2k − 1 > k. However, and as noted above, a very important role will +be played by a certain groupoid quotient of P; see Definition 3.39. +Remark 3.18. It is also worth noting that P is the free ∗-category over the relation F in the +following sense. We define Γ = ΓP to be the graph with vertex set P, and an edge xpq : p → q for +each (p, q) ∈ F with p ̸= q. As in [96, p. 49], the free category C = C(Γ) has object set vC = P, +and its morphisms are the paths in Γ, including the empty path at each vertex p ∈ P (which are +identified with p, as usual). Non-empty paths can be thought of as words xp1p2xp2p3 · · · xpk−1pk +(note the matching subscripts between successive letters), and composition of paths is given by +concatenation when the endpoints match. The involution of C is given by reversal of paths; this +is well-defined since F is symmetric. It is then clear that +xp1p2xp2p3 · · · xpk−1pk → (p1, p2, . . . , pk) +defines a ∗-isomorphism C → P. +We now wish to show that P is an ordered ∗-category, using Lemma 2.3. To apply the lemma, +we need a partial order on the object set vP = P, and a collection of (left) restrictions q⇃p. +We already have the order ≤ on P given in (3.5). To define the restrictions, consider a P-path +p = (p1, . . . , pk), and suppose q ∈ P is such that q ≤ d(p) = p1. We define a tuple +q⇃p = (q1, . . . , qk) +where +q1 = q +and +qi = qi−1θpi for 2 ≤ i ≤ k. +(3.19) +35 + +Note that qi = qθp2 · · · θpi for all 1 ≤ i ≤ k. In fact, since q ≤ p1 we have q = qθp1, so that +qi = qθp2 · · · θpi = qθp1 · · · θpi +for all 1 ≤ i ≤ k. +(3.20) +It follows immediately that qi ≤ pi for all i. +Lemma 3.21. If p ∈ P, and if q ≤ d(p), then q⇃p ∈ P. +Moreover, d(q⇃p) = q and +r(q⇃p) ≤ r(p). +Proof. Write p = (p1, . . . , pk) and q⇃p = (q1, . . . , qk), as above. Since p ∈ P, we have pi F pi+1 +for all 1 ≤ i < k, and we must show that qi F qi+1 for all such i, i.e. that +qiθqi+1 = qi+1 +and +qi+1θqi = qi. +For the first, we have +qiθqi+1 = qiθqiθpi+1 =4 qiθpi+1θqiθpi+1 =3 pi+1θqiθpi+1 =3 qiθpi+1 = qi+1. +For the second, it follows from (3.20) that qi =1 qiθpi. +Combining this with pi = pi+1θpi +(as pi F pi+1), we have +qi+1θqi = qiθpi+1θqi =3 pi+1θqi = pi+1θqiθpi =4 pi+1θpiθqiθpi = piθqiθpi =3 qiθpi = qi. +This all shows that q⇃p ∈ P. +By definition, we have d(q⇃p) = q1 = q, and r(q⇃p) = qk ≤ pk = r(p), where the latter follows +from the fact, noted above, that qi ≤ pi for all i. +Lemma 3.22. If p ∈ P and q ≤ d(p), and if r = r(q⇃p), then (q⇃p)rev = r⇃prev. +Proof. Write p = (p1, . . . , pk) and q⇃p = (q1, . . . , qk), so that r = qk. We also of course have +prev = (pk, . . . , p1) and (q⇃p)rev = (qk, . . . , q1). To make the subscripts match up, it is convenient +to write r⇃prev = (rk, . . . , r1). So +qi = qθp1 · · · θpi +and +ri = rθpk · · · θpi +for all 1 ≤ i ≤ k. +(3.23) +We show by descending induction that qi = ri for all i. The i = k case is trivial, as rk = r = qk. +For 1 ≤ i < k, we have +ri = ri+1θpi +by (3.23) += qi+1θpi +by induction += qθp1 · · · θpiθpi+1θpi +by (3.23) += qθp1 · · · θpi +by Lemma 3.12(iv) += qi +by (3.23). +Lemma 3.24. If p ∈ P and q = d(p), then q⇃p = p. +Proof. Write p = (p1, . . . , pk), so that q = p1. +Then q⇃p = (q1, . . . , qk), where each qi = +qθp1 · · · θpi, and we must show that pi = qi for all i. This is clear for i = 1. For i ≥ 2, we have +qi = qi−1θpi = pi−1θpi = pi, +by definition, induction, and the fact that pi F pi−1. +Lemma 3.25. If p ∈ P and r ≤ q ≤ d(p), then r⇃q⇃p = r⇃p. +36 + +Proof. Write +p = (p1, . . . , pk), +q⇃p = (q1, . . . , qk), +r⇃p = (r1, . . . , rk) +and +r⇃q⇃p = (s1, . . . , sk). +So +qi = qθp1 · · · θpi, +ri = rθp1 · · · θpi +and +si = rθq1 · · · θqi +for all 1 ≤ i ≤ k, +(3.26) +and we must show that ri = si for all i. For i = 1 we have r1 = r = s1. For i ≥ 2, +si = si−1θqi +by (3.26) += (si−1θqi−1)θqi−1θpi +by (3.26), noting that si−1 ≤ qi−1 += si−1θqi−1θpiθqi−1θpi +by (P4) += si−1θqi−1θpi +by (P5) += si−1θpi +by (3.26) += ri−1θpi +by induction += ri +by (3.26). +Lemma 3.27. If p, q ∈ P with r(p) = d(q), and if r ≤ d(p) and s = r(r⇃p), then +r⇃(p ◦ q) = r⇃p ◦ s⇃q. +Proof. Write p = (p1, . . . , pk) and q = (q1, . . . , ql), noting that pk = q1. Then +r⇃p = (r, rθp2, rθp2θp3, . . . , rθp2 · · · θpk), +so +s = rθp2 · · · θpk. +It follows that +r⇃(p ◦ q) = r⇃(p1, . . . , pk, q2, . . . , ql) += (r, rθp2, rθp2θp3, . . . , rθp2 · · · θpk = s, sθq2, sθq2θq3, . . . , sθq2 · · · θql) += (r, rθp2, rθp2θp3, . . . , rθp2 · · · θpk) ◦ (s, sθq2, sθq2θq3, . . . , sθq2 · · · θql) = r⇃p ◦ s⇃q. +Proposition 3.28. For any projection algebra P, the path category P = P(P) is an ordered +∗-category, with ordering given by +p ≤ q +⇔ +p = r⇃q +for some r ≤ d(q). +Proof. This follows from an application of Lemma 2.3. +The ordering on vP = P is given +in (3.5). Properties (O1)′–(O5)′ were established in Lemmas 3.21–3.27. +As usual (cf. Remark 2.5), we can use the involution to define a right-handed restriction: +p⇂q = (q⇃prev)rev +for p ∈ P and q ≤ r(p). +(3.29) +Specifically, if p = (p1, . . . , pk) and q ≤ r(p) = pk, then +p⇂q = (q1, . . . , qk) +where +qi = qθpk · · · θpi +for all 1 ≤ i ≤ k. +(3.30) +37 + +3.3 +Linked pairs +In the next section we will define the chain groupoid C = C (P) of a projection algebra P as a +quotient of the path category P = P(P) by a certain congruence ≈. For the definition of ≈ we +require the concept of linked pairs: +Definition 3.31. Consider a projection p ∈ P. A pair of projections (e, f) ∈ P 2 is said to be +p-linked if +f = eθpθf +and +e = fθpθe. +(3.32) +Associated to such a p-linked pair (e, f) we define the tuples +λ(e, p, f) = (e, eθp, f) +and +ρ(e, p, f) = (e, fθp, f). +The next two results gather some important basic properties of p-linked pairs. +Lemma 3.33. If (e, f) is p-linked, then +(i) (f, e) is also p-linked, and we have λ(e, p, f)rev = ρ(f, p, e) and ρ(e, p, f)rev = λ(f, p, e), +(ii) e, f ≤F p, +(iii) λ(e, p, f) and ρ(e, p, f) both belong to P(e, f). +Proof. (i). This follows directly by inspecting Definition 3.31. +(ii). By the symmetry afforded by part (i), it suffices to show that e ≤F p. For this we use (3.32), +Lemma 3.4 and (P3) to calculate +pθe = pθfθpθe = pθeθpθfθpθe = eθpθfθpθe = fθpθe = e. +(iii). +By symmetry, it suffices to show that λ(e, p, f) ∈ P, i.e. that e F eθp F f. +From +e ≤ e ≤F p it follows from Lemma 3.15(ii) that e F eθp. By (3.32) and Lemma 3.12(i), we have +f = (eθp)θf ≤F eθp, so we are left to show that eθp ≤F f, and for this we use (P4) and (3.32): +fθeθp = fθpθeθp = eθp. +Remark 3.34. Consider a projection p ∈ P, and a p-linked pair (e, f). By Lemma 3.33(ii) +we have e, f ≤F p, and of course we also have eθp, fθp ≤ p. These relationships are all shown +in Figure 12. In the diagram, each arrow s → t stands for the P-path (s, t) ∈ P. Thus, the +upper and lower paths from e to f in the bottom part of the diagram correspond to λ(e, p, f) +and ρ(e, p, f), respectively. +The next result has an obvious dual, but we will not state it. +Lemma 3.35. If (e, f) is p-linked, and if e′ ≤ e, then (e′, f′) is p-linked, where f′ = e′θpθf, and +we have +e′⇃λ(e, p, f) = λ(e′, p, f′) +and +e′⇃ρ(e, p, f) = ρ(e′, p, f′). +Proof. To show that (e′, f′) is p-linked, we must show that +f′ = e′θpθf′ +and +e′ = f′θpθe′. +For the first we use the definition of f′ and Lemma 3.4 several times to calculate +e′θpθf′ = e′θpθe′θpθf = e′θpθfθpθe′θpθf = fθpθe′θpθf = e′θpθf = f′. +38 + +e +p +eθp +fθp +f +Figure 12. A projection p ∈ P, and a p-linked pair (e, f), as in Definition 3.31. Dotted and dashed +lines indicate ≤F and ≤ relationships, respectively. See Remark 3.34 for more details. +For the second, we begin with the projection algebra axioms, and calculate +f′θpθe′ = e′θpθfθpθe′ =4 e′θfθpθe′ =3 fθpθe′ = fθpθeθe′ +by Lemma 3.9, as e′ ≤ e += eθe′ +by (3.32) += e′ +as e′ ≤F e by Lemma 3.12(ii). +It remains to show that e′⇃λ = λ′ and e′⇃ρ = ρ′, where for convenience we write +λ = λ(e, p, f), +ρ = ρ(e, p, f), +λ′ = λ(e′, p, f′) +and +ρ′ = ρ(e′, p, f′). +Using (3.19) and Definition 3.31, we have +e′⇃λ = (e′, e′θeθp, e′θeθpθf), +λ′ = (e′, e′θp, f′), +e′⇃ρ = (e′, e′θfθp, e′θfθpθf), +ρ′ = (e′, f′θp, f′), +so it remains to check that +e′θeθp = e′θp, +e′θeθpθf = f′, +e′θfθp = f′θp +and +e′θfθpθf = f′. +These are all easily dealt with: +• e′θeθp =4 e′θpθeθp = e′θeθpθeθp =5 e′θeθp = e′θp, using e′ ≤ e, +• (e′θeθp)θf = e′θpθf = f′, using the previous calculation, +• e′θfθp =4 e′θpθfθp = f′θp, and +• e′θfθpθf =4 e′θpθfθpθf =5 e′θpθf = f′. +39 + +3.4 +The chain groupoid +We are now almost ready to define the chain groupoid C = C (P) associated to a projection +algebra P. This groupoid is defined below as a certain quotient C = P/≈ of the path category +P = P(P) from Definition 3.17. The congruence ≈ is defined by specifying a generating set: +Definition 3.36. Given a projection algebra P (cf. Definition 3.1), let Ω = Ω(P) be the set of +all pairs (s, t) ∈ P × P of the following three forms: +(Ω1) s = (p, p) and t = (p) ≡ p, for some p ∈ P, +(Ω2) s = (p, q, p) and t = (p) ≡ p, for some (p, q) ∈ F, +(Ω3) s = λ(e, p, f) and t = ρ(e, p, f), for some p ∈ P, and some p-linked pair (e, f). +We define ≈ = Ω♯ to be the congruence on P generated by Ω. +Since d(s) = d(t) and r(s) = r(t) for all (s, t) ∈ Ω, it follows that ≈ is a v-congruence. +Lemma 3.37. The set Ω satisfies (2.18). Consequently, the congruence ≈ satisfies (C4). +Proof. By Lemma 2.17, it suffices to prove the first claim. +To do so, we must show that +(s, t) ∈ Ω +⇒ +srev ≈ trev for all (s, t) ∈ Ω. This is immediate when (s, t) has the form (Ω1) +or (Ω2). For (Ω3), it follows from Lemma 3.33(i). +Lemma 3.38. The set Ω satisfies (2.19). Consequently, the congruence ≈ satisfies (C5). +Proof. By Lemma 2.17, it suffices to prove the first claim. To do so, let (s, t) ∈ Ω. We must +show that +r⇃s ≈ r⇃t +and +r(r⇃s) = r(r⇃t) +for all r ≤ d(s). +(Note that the equality ϑs = ϑt, which is part of (2.19), is equivalent to r(r⇃s) = r(r⇃t) for all +r ≤ d(s).) We do this separately for each of the three forms the pair (s, t) can take. This is very +easy for (Ω1), so we just treat the other two cases. +(Ω2). Let r ≤ p. To calculate r⇃s = r⇃(p, q, p), we first note that rθqθp = rθpθqθp = rθp = r, +where we used r ≤ p in the first and third steps, and Lemma 3.12(iv) in the second. We then +have +r⇃s = (r, rθq, rθqθp) = (r, rθq, r) +and of course +r⇃t = (r). +This shows that in fact (r⇃s, r⇃t) ∈ Ω, so certainly r⇃s ≈ r⇃t. We also have r(r⇃s) = r = r(r⇃t). +(Ω3). Let e′ ≤ d(s) = e, and let f′ = e′θpθf. By Lemma 3.35 we have +e′⇃s = λ(e′, p, f′) +and +e′⇃t = ρ(e′, p, f′), +so again (e′⇃s, e′⇃t) ∈ Ω, and r(e′⇃s) = f′ = r(e′⇃t). +It follows quickly from iterating (Ω2) that +p ◦ prev ≈ d(p) +for all p ∈ P. +Combining this with Lemmas 2.7, 3.37 and 3.38, it follows that the quotient P/≈ is an ordered +groupoid. +Definition 3.39. The chain groupoid of a projection algebra P (cf. Definition 3.1) is the quotient +C = C (P) = P/≈, +where P = P(P) is the path category of P (cf. Definition 3.17), and where ≈ is the congruence +given in Definition 3.36. +40 + +• The elements of C , which are ≈-classes of P-paths, are called P-chains. For p ∈ P, we write +[p] ∈ C for the ≈-class of p. If p = (p1, . . . , pk), then we write [p] = [p1, . . . , pk]. We then have +d[p] = d(p) = p1 and r[p] = r(p) = pk. +• For p = (p1, . . . , pk) and q = (q1, . . . , ql) with pk = q1, we have +[p] ◦ [q] = [p ◦ q] = [p1, . . . , pk−1, pk = q1, q2, . . . , ql]. +• For p = (p1, . . . , pk) ∈ P, we have +[p]−1 = [prev] = [pk, . . . , p1]. +• The order in C is given by +c ≤ d +⇔ +p ≤ q +for some p ∈ c and q ∈ d. +Before we move on, it is worth commenting on the definition of the congruence ≈. The chain +groupoid C = C (P) will be used in Chapter 4 to provide an environment in which to ‘interpret’ +products of projections in certain abstract groupoids G with object set P. Since distinct pro- +jections/objects have distinct co/domains, they cannot be composed in such a groupoid. But a +P-chain [p1, . . . , pk] ∈ C is meant to be thought of as a ‘product’ p1 · · · pk, and the ‘interpreta- +tion’ mentioned above is a functor C → G. Thus, ≈ is meant to equate such ‘products’ when +they ‘should be equal’ in any regular ∗-semigroup with projection algebra P. +In this way, it is clear why [p, p] and [p] ‘should’ be equal for any p ∈ P; cf. (Ω1). Similar +comments apply to [p, q, p] and [p] when p F q; cf. (Ω2) and (2.38). Pairs of the form (Ω3) are +not quite as obvious. For the purposes of Chapter 4, one could replace ≈ with the congruence ∼ +generated by pairs only of the form (Ω1) and (Ω2), and hence replace C = P/≈ with the +quotient P/∼. However, as we will see in Chapter 5 (see especially the proof of Lemma 5.15), +pairs of the form (Ω3) do indeed become equal when interpreted as products of projections in +any regular ∗-semigroup. At the very least, this means that there is ‘no harm’ in including such +pairs in the definition of ≈. +However, there is a more compelling reason for including such pairs. Specifically, we will see +in Chapter 7 that C = P/≈ gives rise to a ‘free regular ∗-semigroup’ with projection algebra P. +For this to work, we need C to be a (so-called) chained projection groupoid, and for this we +require pairs of the form (Ω3) to be equivalent; see especially the proof of Proposition 7.4. +This of course leads to the question of whether the equivalence of pairs of type (Ω3) is implied +by those of type (Ω1) and (Ω2), i.e. whether the congruences ≈ and ∼ are in fact equal. But +it is actually fairly easy to see that they are not. Indeed, it is easy to check that the rewriting +system generated by the rules +(p, p) → (p) +and +(p, q, p) → (p) +for p, q ∈ P with p F q +is complete and Notherian, in the sense of [69]. It follows that any P-path is ∼-equivalent to a +unique ‘reduced’ path, i.e. one in which there are no sub-paths of the form (p, p) or (p, q, p). This +reduced path can be obtained by repeatedly but arbitrarily replacing any sub-path of the form +(p, p) or (p, q, p) by (p). In particular, two P-paths are ∼-equivalent if and only if they have the +same reduced word. Moreover, it is not hard to show that there are pairs of the form (Ω3) that +are reduced but not equal, and hence not ∼-equivalent. +For a concrete example, consider the projection algebra P = P(P3) of the partition monoid P3, +and define the following projections from P: +p = +, +e = +and +f = +. +41 + +One can easily check that (e, f) is p-linked, and that +eθp = pep = +and +fθp = pfp = +. +In particular, the paths λ(e, p, f) = (e, eθp, f) and ρ(e, p, f) = (e, fθp, f) are reduced. Since +eθp ̸= fθp, these reduced paths are not equal, and hence not ∼-equivalent, even though they are +≈-equivalent by definition; cf. (Ω3). +Although we will not elaborate on this further, it is worth noting that linked pairs in pro- +jection algebras are somewhat akin to singular squares in regular biordered sets; cf. [110, p. 20]. +Equating pairs of the form (Ω3) is somewhat akin to requiring singular squares to commute; +cf. [110, p40]. +3.5 +The category of projection algebras +Let PA be the (large) category of projection algebras. A morphism in PA is a projection algebra +morphism φ : P → P ′, by which we mean a map satisfying +(pθq)φ = (pφ)θ′ +qφ +for all p, q ∈ P. +Here we use θ and θ′ to denote the unary operations on P and P ′, respectively. One can also +think of a projection algebra morphism as a morphism of binary algebras, in the sense discussed +in Remark 3.2. +Indeed, using ⋄ and ⋄′ to denote the binary operations of P and P ′ (as in +Remark 3.2), φ : P → P ′ is a projection algebra morphism if and only if +(p ⋄ q)φ = (pφ) ⋄′ (qφ) +for all p, q ∈ P. +In the next result we also write ≤ and ≤′ for the partial orders on P and P ′, as in (3.5), and +similarly for the ≤F and F relations from (3.10) and (3.11). +Lemma 3.40. If φ : P → P ′ is a projection algebra morphism, then for any p, q ∈ P, +p ≤ q ⇒ pφ ≤′ qφ, +p ≤F q ⇒ pφ ≤′ +F qφ +and +p F q ⇒ pφ F ′ qφ. +Proof. The proofs are all essentially the same, so we just prove the first. For this we have +p ≤ q +⇒ +p = pθq +⇒ +pφ = (pθq)φ = (pφ)θ′ +qφ +⇒ +pφ ≤′ qφ. +The next result shows that any projection algebra morphism naturally induces a functor +between the corresponding chain groupoids. +Proposition 3.41. If φ : P → P ′ is a projection algebra morphism, then there is a well-defined +ordered groupoid functor +Φ : C (P) → C (P ′) +given by +[p1, . . . , pk]Φ = [p1φ, . . . , pkφ]. +Proof. During the proof we write C = C (P) and C ′ = C (P ′), and similarly for P and P′. By +Lemma 3.40, we have a well-defined functor +ϕ : P → C ′ +given by +(p1, . . . , pk)ϕ = [p1φ, . . . , pkφ]. +We first show that ≈ ⊆ ker(ϕ). To do so, it suffices to show that sϕ = tϕ for any pair (s, t) ∈ Ω. +This is clear if (s, t) has type (Ω1). We now consider the other two types. +42 + +(Ω2). Next suppose s = (p, q, p) and t = (p) for some (p, q) ∈ F. By Lemma 3.40, we have +pφ F qφ (in P ′), and so sϕ = [pφ, qφ, pφ] = [pφ] = tϕ. +(Ω3). +Finally, suppose s = λ(e, p, f) = (e, eθp, f) and t = ρ(e, p, f) = (e, fθp, f) for some +p ∈ P, and some p-linked pair (e, f). +It is then easy to see that (eφ, fφ) is pφ-linked, and +sϕ = [λ(eφ, pφ, fφ)] and tϕ = [ρ(eφ, pφ, fφ)], so again we have sϕ = tϕ. +Now that we know ≈ ⊆ ker ϕ, we have an induced functor +Φ : C = P/≈ → C ′ +given by +[p]Φ = pϕ, +and this is of course the map in the statement of the proposition. It is easy to check that φ being +a morphism implies (q⇃c)Φ = qφ⇃(cΦ) for all c ∈ C and q ≤ d(c), so that Φ is ordered. +4 +Projection groupoids +Recall from Section 2.3 (see Definition 2.32 and Proposition 2.33) that to any regular ∗-semigroup S +we can associate a groupoid G = G(S). The object set of this groupoid is vG = P = P(S), the +projection algebra of S, and the (partial) composition in G is a restriction of the (total) product +in S. Nevertheless, as explained in Remark 2.34, the groupoid G contains enough information +to completely recover the entire product, in the sense that an arbitrary product ab in S can be +written as a composition in G, viz. ab = a′ ◦ e ◦ b′ for suitable a′, e, b′ ∈ G(= S). This is not +without subtleties, however, as discussed in Remark 2.36. +Roughly speaking, the purpose of the current chapter is to provide an abstract framework +for the above ideas. In Section 4.1 we introduce projection groupoids (see Definition 4.8) as +certain ordered groupoids G whose object set vG = P is an (abstract) projection algebra, as +in Definition 3.1. Section 4.2 introduces the concept of an evaluation map, which is a special +functor ε : C → G, where C = C (P) is the chain groupoid of P, as in Definition 3.39. We then +introduce the key idea of a chained projection groupoid (G, ε); here G is a projection groupoid, +and ε : C → G is an evaluation map obeying a certain coherence condition; see Definition 4.22. +In Section 4.3 we show that an arbitrary chained projection groupoid (G, ε) gives rise to a +regular ∗-semigroup S = S(G, ε) on the same underlying set as G, with product • extending the +composition ◦, with involution given by groupoid inversion, and with projection algebra P(S) +equal to the object set P = vG; see Theorem 4.36. We conclude, in Section 4.4, with some +basic properties concerning • products of projections in S, including a characterisation of the +idempotent-generated subsemigroup E(S) = ⟨E(S)⟩ in Proposition 4.39. +In coming chapters we will take this idea much further. +Specifically, in Chapter 5 we +show, conversely, that any regular ∗-semigroup S gives rise to a chained projection groupoid +G(S) = (G, ε). In Chapter 6 we show that S and G are in fact mutually inverse functors be- +tween the categories of regular ∗-semigroups and chained projection groupoids (with suitable +morphisms), so that these categories are isomorphic; see Theorem 6.7. +4.1 +Projection groupoids +Let G be an ordered groupoid, and suppose the object set P = vG is a projection algebra (cf. Def- +inition 3.1), with the ordering on G inherited from that of P (cf. (3.5)) in the sense described in +Lemma 2.3. That is, for each a ∈ G and each p ≤ d(a) we have a left restriction p⇃a ∈ G, with re- +spect to which conditions (O1)′–(O5)′ hold. We also have the right restrictions a⇂q = (q⇃a−1)−1, +defined for q ≤ r(a), and these satisfy the duals of (O1)′–(O5)′. We typically use these conditions +without explicit reference, and also (O6)′ and its dual, which follow from the others. +At this point, the only ‘link’ between the structures of the groupoid G and the projection +algebra P = vG is via the order ≤ on P. This order is defined in terms of the θ operations +43 + +in (2.26). Conversely, however, we cannot recover the θ operations from the order ≤; indeed, we +will see in Example 5.21 that different projection algebras (with the same underlying set) can +give rise to exactly the same ordering. Thus, we will be particularly interested in groupoids G +(as above) with a stronger link to their object algebra P = vG. Such a link manifests itself in +a number of equivalent properties listed in Proposition 4.6 below. But before we get to these +properties, we first consider the more general situation in which the only assumed link between G +and P = vG is via the order ≤ on P, as above. +Consider a morphism a ∈ G. As in (2.10), we have the map +ϑa : d(a)↓ → r(a)↓ +given by +pϑa = r(p⇃a). +Since d(a) ∈ P, we also have the unary operation θd(a), and by (3.8) its image is d(a)↓. It follows +that we can compose θd(a) with ϑa, and we denote this composition by +Θa = θd(a)ϑa : P → r(a)↓. +(4.1) +As a special case, note that by Lemma 2.14(ii), we have +Θp = θd(p)ϑp = θp idp↓ = θp +for any projection p ∈ P. +(4.2) +We begin by recording some basic properties of the Θ maps. +Lemma 4.3. For any a ∈ G we have: +(i) ϑaθr(a) = ϑa, +(ii) Θaθr(a) = Θa = θd(a)Θa, +(iii) Θaϑa−1 = θd(a), +(iv) Θa = θd(a)ϑb for any a ≤ b, +(v) Θp⇃a = θpΘa for any p ≤ d(a). +Proof. (i). Since im(ϑa) = r(a)↓, this follows immediately from the fact that each operation θp +fixes each element of im(θp) = p↓. +(ii). Since Θa = θd(a)ϑa, we obtain Θaθr(a) = Θa from part (i), and θd(a)Θa = Θa from (P2). +(iii). Since im(θd(a)) = d(a)↓, it follows from Lemma 2.12 that +Θaϑa−1 = θd(a)ϑaϑa−1 = θd(a) idd(a)↓ = θd(a). +(iv). Since a ≤ b, we have a = p⇃b, where p = d(a). Now let s ��� P be arbitrary; we must show +that sΘa = sθpϑb. For this, we write t = sθp, and use (O4)′ to calculate +sΘa = sθpϑa = tϑa = r(t⇃a) = r(t⇃p⇃b) = r(t⇃b) = tϑb = sθpϑb. +(v). We have +Θp⇃a = θd(p⇃a)ϑa = θpϑa +by part (iv), as p⇃a ≤ a += θpθd(a)ϑa +by Lemma 3.9, as p ≤ d(a) += θpΘa. +Another important property is that when a and b are composable in G, the map Θa◦b is the +composition of Θa and Θb. +44 + +Lemma 4.4. If a, b ∈ G are such that r(a) = d(b), then Θa◦b = ΘaΘb. +Proof. Write p = d(a) and q = r(a) = d(b). Then by Lemmas 2.14(i) and 4.3(i) we have +ΘaΘb = θpϑaθqϑb = θpϑaθr(a)ϑb = θpϑaϑb = θpϑa◦b = Θa◦b. +Part (v) of Lemma 4.3 above shows how the Θ maps interact with left restrictions, but the +lemma did not include a corresponding statement concerning right restrictions. +Of course if +q ≤ r(a), then a⇂q = p⇃a where p = qϑa−1 (cf. (2.11)), and then +Θa⇂q = Θp⇃a = θpΘa = θqϑa−1Θa. +(4.5) +However, the groupoids we will be concerned with satisfy the neater identity Θa⇂q = Θaθq. This +is in fact equivalent to a number of additional properties linking the groupoid structure of G to +the projection algebra structure of P = vG: +Proposition 4.6. If G is an ordered groupoid, whose object set P = vG is a projection algebra, +as above, then the following are equivalent: +(G1a) θpϑa = Θa−1θpΘa for all a ∈ G and p ≤ d(a), +(G1b) θpΘa = Θa−1θpΘa for all a ∈ G and p ∈ P, +(G1c) Θa⇂q = Θaθq for all a ∈ G and q ≤ r(a), +(G1d) ϑa is a projection algebra morphism (and hence isomorphism) d(a)↓ → r(a)↓ for all a ∈ G. +Proof. For the duration of the proof we fix a ∈ G, and we write s = d(a) and t = r(a). On a +number of occasions we will use the fact that +Θa−1Θa = Θa−1◦a = Θr(a) = Θt = θt. +(4.7) +In the above calculation we used Lemma 4.4 and (4.2). In what follows, we show that (G1a) +implies each of (G1b)–(G1d), and conversely that any of (G1b)–(G1d) implies (G1a). +(G1a) ⇒ (G1b)–(G1d). Suppose first that (G1a) holds. To verify (G1b), let p ∈ P be arbitrary. +Then +θpΘa = θ(pθs)ϑa +by definition += Θa−1θpθsΘa +by (G1a), as pθs ≤ s = d(a) += Θa−1θsθpθsΘa +by (P4) += Θa−1θr(a−1)θpθd(a)Θa += Θa−1θpΘa +by Lemma 4.3(ii). +For (G1c), if q ≤ r(a) then +Θa⇂q = θqϑa−1Θa +by (4.5) += ΘaθqΘa−1Θa +by (G1a) += Θaθqθr(a) +by (4.7) += Θaθq +by Lemma 3.9, as q ≤ r(a). +For (G1d), since ϑa is a bijection s↓ → t↓ by Lemma 2.12, it is enough to show that ϑa is a +projection algebra morphism, i.e. that +(pθq)ϑa = (pϑa)θqϑa +for all p, q ≤ s = d(a). +45 + +But for any such p, q, we have +(pϑa)θqϑa = pϑaΘa−1θqΘa +by (G1a) += pϑaθtϑa−1θqθsϑa +by definition of the Θ maps += pϑaϑa−1θqϑa +as pϑa ≤ r(a) = t and q ≤ s (cf. Lemma 3.9) += pθqϑa +by Lemma 2.12. +(G1b) ⇒ (G1a). This is immediate, since for any p ≤ d(a) = s we have pϑa = (pθs)ϑa = pΘa . +(G1c) ⇒ (G1a). If (G1c) holds, then for any p ≤ d(a) = r(a−1) we have +Θa−1θpΘa = Θa−1⇂pΘa +by (G1c) += θpϑaΘa−1Θa +by (4.5) += θpϑaθr(a) +by (4.7) += θpϑa +by Lemma 3.9, as pϑa ≤ r(a). +(G1d) ⇒ (G1a). Suppose (G1d) holds, and let p ≤ s. We must show that +eθpϑa = eΘa−1θpΘa +for any e ∈ P, +so fix some such e. +To make the following calculation easier to read, let f = eθt. +Since +f ≤ t = d(a−1), we can also define g = fϑa−1. +By Lemma 2.12 we have f = gϑa, and we +note also that g ≤ r(a−1) = s. We then have +eθpϑa = eθtθpϑa = fθpϑa +by Lemma 3.9, as pϑa ≤ r(a) = t += (gϑa)θpϑa +as observed above += (gθp)ϑa +by (G1d), as g, p ≤ s = d(a) += eθtϑa−1θpϑa +as g = fϑa−1 and f = eθt += eθtϑa−1θpθsϑa +by Lemma 3.9, as p ≤ s += eΘa−1θpΘa +by definition of the Θ maps. +Here then is the main definition of this section: +Definition 4.8. A projection groupoid is an ordered groupoid G, whose object set P = vG is a +projection algebra (cf. Definition 3.1), with the ordering on G inherited from that of P (cf. (3.5)) +in the sense described in Lemma 2.3, and for which: +(G1) any (and hence all) of the conditions (G1a)–(G1d) hold. +Remark 4.9. The reader should notice a resemblance between (G1a) and (P4). In fact, this is +more than just superficial. Recall that (abstract) projection algebras are supposed to ‘model’ pro- +jections of regular ∗-semigroups, and that the θp operations model ‘conjugation’, viz. qθp ≡ pqp. +In this way, axiom (P4) can be thought of as a recipe for ‘iterating’ conjugation by projections. +By the same token, restrictions p⇃a in G (with p ≤ d(a)) are supposed to model ‘restric- +tions’ pa in a regular ∗-semigroup S, as in Remark 2.34. ‘Pretending’ that we are working in a +regular ∗-semigroup, we can then think of +pϑa = r(p⇃a) ≡ r(pa) ≡ (pa)∗pa = a∗p∗pa = a∗pa +46 + +as a kind of ‘conjugate of p by a’. In this sense, the ϑa maps can be thought of as analogues of +conjugation in an arbitrary (abstract) projection groupoid. For arbitrary t ∈ P (not necessarily +below d(a)), we can then also think of +tΘa = tθd(a)ϑa ≡ a∗ · d(a) · t · d(a) · a ≡ a∗ · aa∗ · t · aa∗ · a = a∗ta, +so that Θa also acts as a kind of ‘conjugation’ by morphisms. Condition (G1a) can then be +thought of as a way to iterate this kind of conjugation, as can condition (G1b). +We close the section with a simple technical result that will help us shorten some later +arguments. +Lemma 4.10. If G is a projection groupoid, and if a ∈ G and p, q ∈ P, then +pΘa−1θqΘaθp = qΘaθp. +Proof. By (G1b) and (P3) we have pΘa−1θqΘaθp = pθqΘaθp = qΘaθp. +4.2 +Chained projection groupoids +At this point it is worth pausing to consider again the broad goal of this chapter, and to draw +attention to one of the major hurdles we currently face. Recall that (broadly speaking) we are +aiming to give an abstract description of the groupoids that correspond to regular ∗-semigroups. +Starting with a regular ∗-semigroup S, in Definition 2.32 we constructed a groupoid G = G(S) +with underlying set S, and whose composition was a restriction of the product in S. As we +explained in Remark 2.34, this (partial) composition is enough to reconstruct the entire product +of S, in the sense that for any a, b ∈ S we have ab = a′ ◦ e ◦ b′ for suitable a′, e, b′ ∈ G(= S). +As noted in that remark, the element e is in fact the idempotent e = p′q′, where p′, q′ ∈ P are +certain special projections below r(a) = a∗a and d(b) = bb∗. Being an idempotent of course +means that e2 = e in S. However, since d(e) = p′ and r(e) = q′ (cf. (2.41)), and since p′ and q′ +are not necessarily equal (though always F-related), the composition e◦e is generally not defined +in G. Returning to the abstract setting of projection groupoids, in which we are operating in +this chapter, it is not at all obvious at the outset which morphism p′ → q′ ought to play the role +of this idempotent e. The precise mechanism for solving this problem (which was explored at +greater length in Remark 2.36) is encoded in what we will call an evaluation map, and in the +properties we will require of them below. +Definition 4.11. Let G be a projection groupoid (cf. Definition 4.8), and let C = C (P) be +the chain groupoid of P = vG (cf. Definition 3.39). An evaluation map is an ordered v-functor +ε : C → G, meaning that the following hold: +(E1) ε(p) = p for all p ∈ P (where as usual we identify p ≡ [p] ∈ C ), +(E2) ε(c ◦ d) = ε(c) ◦ ε(d) if r(c) = d(d), +(E3) c ≤ d ⇒ ε(c) ≤ ε(d). +We will soon define a chained projection groupoid to be a projection groupoid with an eval- +uation map possessing a certain coherence property. +But first we note a number of simple +consequences of (E1)–(E3). +Remark 4.12. As with group homomorphisms, it is easy to see that +(E4) ε(c−1) = ε(c)−1 for all c ∈ C . +47 + +It also follows that: +(E5) d(ε(c)) = d(c) and r(ε(c)) = r(c) for all c ∈ C . +For example, d(ε(c)) = ε(c) ◦ ε(c)−1 = ε(c) ◦ ε(c−1) = ε(c ◦ c−1) = ε(d(c)). +It is also easy to see that (E3) is equivalent (in the presence of the other axioms) to: +(E6) ε(q⇃c) = q⇃ε(c) if q ≤ d(c). +For example, if (E6) holds, and if c ≤ d, then c = p⇃d for some p ≤ d(d); it follows from this that +ε(c) = ε(p⇃d) = p⇃ε(d) ≤ ε(d). +Consider a projection groupoid G with P = vG, and let ε : C → G be an evaluation map. +For p, q ∈ P with p F q, we have the P-chain [p, q] ∈ C , and the elements ε[p, q] ∈ G will play +a crucial role in all that follows. For one thing, these elements generate the image of C under ε +(as the [p, q] generate C ). Specifically, if c = [p1, p2, . . . , pk] ∈ C , then +ε(c) = ε[p1, p2] ◦ ε[p2, p3] ◦ · · · ◦ ε[pk−1, pk]. +(4.13) +The next lemma gathers some important basic properties of the ε[p, q]; in what follows, we will +typically use these without explicit reference. We write f|A for the (set-theoretic) restriction of +a function f to a subset A of its domain. +Lemma 4.14. Let G be a projection groupoid, and ε : C → G an evaluation map. +(i) For any p ∈ P we have ε[p, p] = p. +(ii) If p, q ∈ P are such that p F q, then +d(ε[p, q]) = p, +ε[p, q]−1 = ε[q, p], +ϑε[p,q] = θq|p↓, +r(ε[p, q]) = q, +Θε[p,q] = θpθq. +(iii) If p, q, r, s ∈ P are such that p F q, r ≤ p and s ≤ q, then +r⇃ε[p, q] = ε[r, rθq] +and +ε[p, q]⇂s = ε[sθp, s]. +Proof. (i). This follows quickly from (E1) and [p, p] = [p], keeping in mind the identification of +p ∈ P with the chain [p] ∈ C . +(iii). This follows from (E6), together with (3.19) and (3.30). +(ii). The claims concerning domains, ranges and inversion follow from (E4) and (E5), and the +fact that [p, q]−1 = [q, p] in C . +Since the maps ϑε[p,q] and θq|p↓ both have domain p↓, we can show the maps are equal by +showing that +rϑε[p,q] = rθq +for all r ≤ p. +But for such r, we use the definition of the ϑ maps, and parts of the current lemma that have +already been proved, to calculate +rϑε[p,q] = r(r⇃ε[p, q]) = r(ε[r, rθq]) = rθq. +Finally, again using already-proved parts of the current lemma, we have +Θε[p,q] = θpϑε[p,q] = θp ◦ θq|p↓ = θpθq, +where in the last step we used the fact that p↓ = im(θp). +48 + +Before we move on, it will also be convenient to record the following result. +Lemma 4.15. Let G be a projection groupoid, and ε : C → G an evaluation map. If p, q ∈ P +are such that p F q, and if a ≤ ε[p, q], then +a = ε[r, s] +where +r = d(a) +and +s = r(a). +Proof. Noting that a ≤ ε[p, q], we use Lemma 4.14 to calculate +a = r⇃ε[p, q] = ε[r, rθq] +and +rθq = r(ε[r, rθq]) = r(a) = s. +To state the coherence property alluded to above, we require the concept of linked pairs, +which we now define. We have reused this term from Section 3.3 (where p-linked pairs were used +to define the congruence ≈ on the path category P), because there is a strong tie between the +two concepts, as we will see later. +Definition 4.16. Let G be a projection groupoid, and consider a morphism b ∈ G. A pair of +projections (e, f) ∈ P 2 is said to be b-linked if +f = eΘbθf +and +e = fΘb−1θe. +(4.17) +We will soon associate two morphisms λ(e, b, f) and ρ(e, b, f) to the b-linked pair (e, f). But +first we prove the next result, which will ensure these morphisms are well defined. +Lemma 4.18. Let G be a projection groupoid, and suppose (e, f) is b-linked, where b ∈ G(q, r). +Define further projections +e1 = eθq, +e2 = fΘb−1, +f1 = eΘb +and +f2 = fθr. +(4.19) +Then +(i) e ≤F q and f ≤F r, +(ii) ei ≤ q and fi ≤ r for i = 1, 2, +(iii) e F ei and f F fi for i = 1, 2, +(iv) ei⇃b = b⇂fi for i = 1, 2. +Proof. It is clear from Definition 4.16 that (e, f) is b-linked if and only if (f, e) is b−1-linked. If +we replace b ↔ b−1 and (e, f) ↔ (f, e), then the projections defined in (4.19) are interchanged +accordingly, viz. e1 ↔ f2 and e2 ↔ f1. Because of this symmetry, for parts (i)–(iii), it suffices to +prove the claims concerning e, e1, e2. (Alternatively, the arguments below can be easily adapted +to prove the claims regarding f, f1, f2.) +(ii). This follows from im(Θb−1) = im(θq) = q↓. +(iii). From e ≤ e ≤F q, it follows from Lemma 3.15(ii) that e F eθq = e1. By Lemma 3.12(i), +we have e = e2θe ≤F e2. It remains to show that e2 ≤F e, and for this we use (G1b), (4.17) +and (4.19): +eθe2 = eθfΘb−1 = eΘbθfΘb−1 = fΘb−1 = e2. +(iv). We must show that fi = eiϑb (i = 1, 2). Keeping q = d(b) and r = d(b−1) in mind, we have +e1ϑb = eθqϑb = eΘb = f1 +and +e2ϑb = fΘb−1ϑb = fθr = f2, +where we used Lemma 4.3(iii) in the second calculation. +(i). Combining parts (ii) and (iii), we have e ≤F e1 ≤ q, and Lemma 3.15(i) then gives e ≤F q. +49 + +Definition 4.20. Let G be a projection groupoid (cf. Definition 4.8), and ε : C → G an evaluation +map (cf. Definition 4.11). Let (e, f) be a b-linked pair, where b ∈ G(q, r), and let e1, e2, f1, f2 ∈ P +be as in (4.19). By Lemma 4.18, we have two well-defined morphisms +λ(e, b, f) = ε[e, e1] ◦ e1⇃b ◦ ε[f1, f] +and +ρ(e, b, f) = ε[e, e2] ◦ e2⇃b ◦ ε[f2, f] += ε[e, e1] ◦ b⇂f1 ◦ ε[f1, f] += ε[e, e2] ◦ b⇂f2 ◦ ε[f2, f]. +(4.21) +These morphisms are shown in Figure 13. +e +e1 +e2 +f +f1 +f2 +q +r +b +e1⇃b = b⇂f1 +e2⇃b = b⇂f2 +ε[e, e1] +ε[e, e2] +ε[f1, f] +ε[f2, f] +Figure 13. The projections and morphisms associated to a b-linked pair (e, f), for b ∈ G(q, r); see +Definitions 4.16, 4.20 and 4.22, and Lemma 4.18. Dotted and dashed lines indicate ≤F and ≤ rela- +tionships, respectively. Axiom (G2) says that the hexagon at the bottom of the diagram commutes. +The above-mentioned coherence property states that the two terms in (4.21) must be equal: +Definition 4.22. A chained projection groupoid is a pair (G, ε), where G is a projection groupoid +(cf. Definition 4.8), and ε : C → G is an evaluation map (cf. Definitions 3.39 and 4.11) satisfying +the following condition: +(G2) For every b ∈ G, and for every b-linked pair (e, f), we have λ(e, b, f) = ρ(e, b, f), where +these morphisms are as in (4.21). +Remark 4.23. One might wonder if the concept of linked pairs could be avoided when defining +chained projection groupoids. Specifically, one might wonder if (G2) is equivalent to: +(G2)′ For any morphism b ∈ G(q, r), and for any projections e, e1, e2, f, f1, f2 ∈ P satisfying +conditions (i)–(iv) from Lemma 4.18, we have ε[e, e1]◦e1⇃b◦ε[f1, f] = ε[e, e2]◦e2⇃b◦ε[f2, f]. +While (G2)′ of course implies (G2), the converse does not hold. Most significantly, the stronger +condition (G2)′ does not (generally) hold in the all-important case that G = G(S) is the groupoid +associated to a regular ∗-semigroup S, as in Definition 2.32. We will say more about this in +Remark 5.20. +50 + +4.3 +The regular ∗-semigroup associated to a chained projection groupoid +For the duration of this section, we fix a chained projection groupoid (G, ε), as in Definition 4.22, +and we continue to write P = vG, C = C (P), and so on. Our aim here is to construct a regular +∗-semigroup S = S(G, ε), built from G and ε in a natural way. The underlying set of S will +simply be G, and the involution of S will simply be inversion in G: +a∗ = a−1 +for a ∈ G. +The definition of the product in S, which we will denote by •, is more involved. Its inspiration +is drawn from the properties of regular ∗-semigroups discussed in Remark 2.34. +Definition 4.24. Let (G, ε) be a chained projection groupoid (cf. Definition 4.22), and consider +an arbitrary pair of morphisms a, b ∈ G. Let p = r(a) and q = d(b), and define the projections +p′ = qθp +and +q′ = pθq. +By Lemma 3.14, we have p′ ≤ p, q′ ≤ q and p′ F q′. In particular, the morphisms a⇂p′, ε[p′, q′] +and q′⇃b exist, and we define +a • b = a⇂p′ ◦ ε[p′, q′] ◦ q′⇃b. +(4.25) +This is illustrated in Figure 14. We define S(G, ε) to be the (2, 1)-algebra with underlying set G, +and with binary operation • and unary operation ∗ = −1. +p +q +p′ +q′ +a +b +a⇂p′ +q′⇃b +ε[p′, q′] +a • b +Figure 14. Construction of the product a • b, as in Definition 4.24. +To show that S = S(G, ε) is a regular ∗-semigroup, with respect to the binary operation • +and unary operation ∗ = −1, we must establish the identities: +(a • b) • c = a • (b • c), +(a∗)∗ = a = (a • a∗) • a +and +(a • b)∗ = b∗ • a∗. +Associativity of • is quite difficult to establish, and is finally achieved in Lemma 4.34 below. The +identities involving ∗ are comparatively easy, however, and we verify these shortly in Lemma 4.27. +For its proof, we need the following lemma, which shows that the (total) product • extends the +(partial) composition ◦, in the sense that these operations agree whenever the latter is defined. +Before we begin, it is worth noting that the operation • from Definition 4.24 makes sense in +the more general case of G being a projection groupoid with an evaluation map ε. All of the +coming lemmas apart from the final Lemma 4.34 hold in this more general situation as well; the +coherence property (G2) is needed only at the very last stage of the proof of associativity. Still, +for simplicity, we do continue to assume that (G, ε) is a chained projection groupoid. +51 + +Lemma 4.26. If a, b ∈ G are such that r(a) = d(b), then a • b = a ◦ b. +Proof. In the notation of Definition 4.24, we have p = q, so also p′ = q′ = p. But then +a • b = a⇂p ◦ ε[p, p] ◦ p⇃b = a ◦ p ◦ b = a ◦ b. +Lemma 4.27. For any a, b ∈ G, we have +(a∗)∗ = a = (a • a∗) • a +and +(a • b)∗ = b∗ • a∗. +Proof. Of course (a∗)∗ = (a−1)−1 = a. It follows from Lemma 4.26 that a•a∗ = a◦a−1 = d(a), +and so (a • a∗) • a = d(a) • a = d(a) ◦ a = a. +Next, let p, q, p′, q′ be as in Definition 4.24. Since q = r(b∗) and p = d(a∗), we have +a • b = a⇂p′ ◦ ε[p′, q′] ◦ q′⇃b +and +b∗ • a∗ = b∗⇂q′ ◦ ε[q′, p′] ◦ p′⇃a∗, +and so +(a • b)∗ = (q′⇃b)∗ ◦ ε[p′, q′]∗ ◦ (a⇂p′)∗ = b∗⇂q′ ◦ ε[q′, p′] ◦ p′⇃a∗ = b∗ • a∗. +The symmetry/duality afforded by the identity (a • b)∗ = b∗ • a∗ will allow for some simplifi- +cation in some of the proofs to follow. This is the case with the proof of the next result, which +identifies the domain and range of a product a • b. +Lemma 4.28. For any a, b ∈ G we have +d(a • b) = d(b)Θa−1 +and +r(a • b) = r(a)Θb. +Proof. Let p, q, p′, q′ be as in Definition 4.24, so that a • b is as in (4.25). Then +r(a • b) = r(q′⇃b) = q′ϑb = pθqϑb = r(a)θd(b)ϑb = r(a)Θb. +The identity involving domains may be proved in similar fashion (using (2.11)). Alternatively, +it follows from the range identity and duality: +d(a • b) = r((a • b)∗) = r(b∗ • a∗) = r(b∗)Θa∗ = d(b)Θa−1. +Our next result establishes an important identity. +Lemma 4.29. For any a, b ∈ G we have Θa•b = ΘaΘb. +Proof. Let p, q, p′, q′ be as in Definition 4.24. Then by the projection algebra axioms we have +θp′θq′ = θqθpθpθq =4 θpθqθpθqθpθq =5 θpθq. +We then calculate +Θa•b = Θa⇂p′ ◦ Θε[p′,q′] ◦ Θq′⇃b +by (4.25) and Lemma 4.4 += Θaθp′ ◦ θp′θq′ ◦ θq′Θb +by (G1c) and Lemmas 4.3(v) and 4.14(ii) += Θa ◦ θp′θq′ ◦ Θb +by (P2) += Θa ◦ θpθq ◦ Θb +by the above observation += Θaθr(a) ◦ θd(b)Θb += ΘaΘb +by Lemma 4.3(ii). +52 + +We are almost ready to show that • is associative, i.e. that (a • b) • c and a • (b • c) are equal, +for all a, b, c ∈ G. Before we do so, we note that Lemmas 4.28 and 4.29 can be used to show that +these two terms have equal domains, and equal ranges. For example, +r((a • b) • c) = r(a • b)Θc = r(a)ΘbΘc = r(a)Θb•c = r(a • (b • c)). +(4.30) +A similar calculation gives +d((a • b) • c) = d(a �� (b • c)) = d(c)Θb−1Θa−1. +(4.31) +As a stepping stone to associativity, the next result gives expansions of (a•b)•c and a•(b•c), +expressing each term as a composition of five morphisms. +Lemma 4.32. If a, b, c ∈ G, and if p = r(a), q = d(b), r = r(b) and s = d(c), then +(i) (a • b) • c = a⇂e ◦ ε[e, e1] ◦ e1⇃b ◦ ε[f1, f] ◦ f⇃c, where +e = sΘb−1θp, +e1 = eθq, +f1 = eΘb +and +f = pΘbθs, +(ii) a • (b • c) = a⇂e ◦ ε[e, e2] ◦ b⇂f2 ◦ ε[f2, f] ◦ f⇃c, where +e = sΘb−1θp, +e2 = fΘb−1, +f2 = fθr +and +f = pΘbθs. +The above factorisations are illustrated in Figure 15. +Proof. (i). Put p′ = qθp and q′ = pθq, so that +a • b = a′ ◦ ε′ ◦ b′ +where +a′ = a⇂p′, +ε′ = ε[p′, q′] +and +b′ = q′⇃b. +Keeping Lemma 4.28 in mind, we now define +t = r(a • b) = pΘb. +Then +(a • b) • c = (a • b)⇂t′ ◦ ε[t′, s′] ◦ s′⇃c +where +t′ = sθt and s′ = tθs. +We now focus on the term (a • b)⇂t′. By (2.13) we have +(a • b)⇂t′ = (a′ ◦ ε′ ◦ b′)⇂t′ = a′⇂u ◦ ε′⇂v ◦ b′⇂t′ +where +v = t′ϑ−1 +b′ +and +u = t′ϑ−1 +b′ ϑ−1 +ε′ . +Now, +• a′⇂u = a⇂p′⇂u = a⇂u, +• ε′⇂v ≤ ε′ = ε[p′, q′], d(ε′⇂v) = r(a′⇂u) = u and r(ε′⇂v) = v, so ε′⇂v = ε[u, v] by Lemma 4.15, +and +• b′⇂t′ ≤ b′ ≤ b and d(b′⇂t′) = r(ε′⇂v) = v, so b′⇂t′ = v⇃b. (For later use, since b′⇂t′ and b⇂t′ are +both below b and have codomain t′, it also follows that b⇂t′ = b′⇂t′ = v⇃b.) +Putting everything together, it follows that +(a • b) • c = (a • b)⇂t′ ◦ ε[t′, s′] ◦ s′⇃c += a′⇂u ◦ ε′⇂v ◦ b′⇂t′ ◦ ε[t′, s′] ◦ s′⇃c += a⇂u ◦ ε[u, v] ◦ v⇃b ◦ ε[t′, s′] ◦ s′⇃c. +53 + +It therefore remains to check that +(a) u = e, +(b) v = e1, +(c) t′ = f1, +(d) s′ = f, +where e, e1, f1, f are as defined in the statement of the lemma. +(a). First note that d(a⇂u) = d((a • b) • c) = sΘb−1Θa−1 by (4.31), so a⇂u = sΘb−1Θa−1⇃a. But +then +u = r(a⇂u) = r(sΘb−1Θa−1⇃a) = sΘb−1Θa−1ϑa = sΘb−1θp = e, +where we used Lemma 4.3(iii) in the second-last step. +(c). Here we use (G1b) to calculate t′ = sθt = sθpΘb = sΘb−1θpΘb = eΘb = f1. +(b). We noted above that v⇃b = b⇂t′. Combining this with (2.11), the just-proved item (c), and +Lemma 4.3(iii), it follows that +v = d(v⇃b) = d(b⇂t′) = t′ϑb−1 = f1ϑb−1 = (eΘb)ϑb−1 = eθq = e1. +(d). Finally, s′ = tθs = pΘbθs = f. +As noted above, this completes the proof of (i). +(ii). This can be proved in similar fashion to part (i), but in fact we can also deduce it from (i). +Indeed, we first use Lemma 4.27 to expand +(a • (b • c))−1 = (b • c)−1 • a−1 = (c−1 • b−1) • a−1. +Note that s = r(c−1), r = d(b−1), q = r(b−1) and p = d(a−1). It follows from part (i) that +(c−1 • b−1) • a−1 = c−1⇂e′ ◦ ε[e′, e′ +1] ◦ e′ +1⇃b−1 ◦ ε[f′ +1, f′] ◦ f′⇃a−1, +where +e′ = pΘbθs, +e′ +1 = e′θr, +f′ +1 = e′Θb−1 +and +f′ = sΘb−1θp. +But then +a • (b • c) = ((c−1 • b−1) • a−1)−1 += (c−1⇂e′ ◦ ε[e′, e′ +1] ◦ e′ +1⇃b−1 ◦ ε[f′ +1, f′] ◦ f′⇃a−1)−1 += (f′⇃a−1)−1 ◦ ε[f′ +1, f′]−1 ◦ (e′ +1⇃b−1)−1 ◦ ε[e′, e′ +1]−1 ◦ (c−1⇂e′)−1 += a⇂f′ ◦ ε[f′, f′ +1] ◦ b⇂e′ +1 ◦ ε[e′ +1, e′] ◦ e′⇃c. +(4.33) +We note immediately that f′ = e and e′ = f (where e, f are as in the statement of the lemma). +We also have +e′ +1 = e′θr = fθr = f2 +and +f′ +1 = e′Θb−1 = fΘb−1 = e2. +Thus, continuing from (4.33), we have +a • (b • c) = a⇂e ◦ ε[e, e2] ◦ b⇂f2 ◦ ε[f2, f] ◦ f⇃c, +as required. +54 + +e +e1 +e2 +f +f1 +f2 +p +q +r +s +a +b +c +e1⇃b = b⇂f1 +e2⇃b = b⇂f2 +ε[e, e1] +ε[e, e2] +ε[f1, f] +ε[f2, f] +a⇂e +f⇃c +Figure 15. Top: three morphisms a, b, c ∈ G. Bottom: the upper and lower paths represent (a•b)•c +and a • (b • c), respectively; cf. Lemma 4.32. Dashed lines indicate ≤ relationships. +Lemma 4.34. For any a, b, c ∈ G, we have (a • b) • c = a • (b • c). +Proof. By Lemma 4.32, it suffices to show (in the notation of that lemma) that +ε[e, e1] ◦ e1⇃b ◦ ε[f1, f] = ε[e, e2] ◦ b⇂f2 ◦ ε[f2, f]. +(4.35) +Comparing the definitions of the ei, fi from Lemma 4.32 (in terms of e, f, q, r, b) with (4.19), an +application of (G2) will give us (4.35) if we can show that (e, f) is b-linked (cf. Definitions 4.16 +and 4.20). In other words, the proof of the lemma will be complete if we can show that +f = eΘbθf +and +e = fΘb−1θe. +For the first, we have +eΘbθf = sΘb−1θpΘbθpΘbθs +by definition += sΘb−1θpΘbθsθpΘbθs +by (P4) += pΘbθsΘb−1θpΘbθs +by (G1b) and Lemma 4.10 += sΘb−1θpΘbθs +by Lemma 4.10 += pΘbθs = f +by Lemma 4.10. +The proof that e = fΘb��1θe is virtually identical. +Lemmas 4.27 and 4.34 give the following: +Theorem 4.36. If (G, ε) is a chained projection groupoid, then S(G, ε) is a regular ∗-semigroup. +4.4 +Products of projections +We conclude this chapter with some simple results concerning • products involving projections. +These will be of use later, but we also apply them here to describe the idempotent-generated +subsemigroup of the regular ∗-semigroup S = S(G, ε). In the following statements we continue +to fix the chained projection groupoid (G, ε), and write P = vG. +55 + +Lemma 4.37. If a ∈ G(p, q), and if t ∈ P, then +(i) a • t = a⇂q′ ◦ ε[q′, t′], where q′ = tθq and t′ = qθt, +(ii) t • a = ε[t′, p′] ◦ p′⇃a, where p′ = tθp and t′ = pθt, +(iii) a • t = a ◦ ε[q, t] if q F t, +(iv) t • a = ε[t, p] ◦ a if p F t, +(v) a • t = a⇂t if t ≤ q, +(vi) t • a = t⇃a if t ≤ p. +Proof. By symmetry, it suffices to prove (i), (iii) and (v). +(i). We have a • t = a⇂q′ ◦ ε[q′, t′] ◦ t′⇃t = a⇂q′ ◦ ε[q′, t′] ◦ t′ = a⇂q′ ◦ ε[q′, t′]. +(iii). This follows from part (i), as q′ = q and t′ = t if q F t. +(v). If t ≤ q, then we also have t ≤F q by Lemma 3.12(ii), so that t = tθq = q′ and t = qθt = t′. +It then follows from part (i) that +a • t = a⇂q′ ◦ ε[q′, t′] = a⇂t ◦ ε[t, t] = a⇂t ◦ t = a⇂t. +Lemma 4.38. If p, q ∈ P, then +(i) p • q = ε[p′, q′], where p′ = qθp and q′ = pθq, +(ii) p • q = ε[p, q] if p F q. +(iii) p • q • p = qθp. +Proof. (i). By Lemma 4.37(i) we have p • q = p⇂p′ ◦ ε[p′, q′] = p′ ◦ ε[p′, q′] = ε[p′, q′]. +(ii). This follows from part (i), as p = p′ and q = q′ when p F q. +(iii). Here we have +p • q • p = ε[p′, q′] • p +by part (i), where p′ = qθp and q′ = pθq += ε[p′, q′]⇂q′′ ◦ ε[q′′, p′′] +by Lemma 4.37(i), where q′′ = pθq′ and p′′ = q′θp. +Using the projection algebra axioms, it is easy to show that in fact p′′ = p′ and q′′ = q′. Thus, +continuing from above, we have +p • q • p = ε[p′, q′]⇂q′′ ◦ ε[q′′, p′′] = ε[p′, q′]⇂q′ ◦ ε[q′, p′] = ε[p′, q′] ◦ ε[q′, p′] = p′ = qθp. +Recall that the sets of idempotents and projections of a regular ∗-semigroup S are denoted +E(S) = {e ∈ S : e2 = e} +and +P(S) = {p ∈ S : p2 = p = p∗}. +Although these sets are not equal in general, Lemma 2.23(ii) says that they generate the same +subsemigroup of S, i.e. ⟨E(S)⟩ = ⟨P(S)⟩. We write E(S) for this idempotent-generated subsemi- +group. The results proved above allow us to give the following description of E(S) in the case +that S = S(G, ε) arises from a chained projection groupoid (G, ε). +56 + +Proposition 4.39. Let (G, ε) be a chained projection groupoid, and let P = vG, C = C (P) and +S = S(G, ε). Then +(i) P(S) = P, +(ii) E(S) = {ε[p, q] : (p, q) ∈ F}, +(iii) E(S) = im(ε) = {ε(c) : c ∈ C }, and consequently S is idempotent-generated if and only +if ε is surjective. +Proof. (i). By Lemmas 2.23(i) and 4.26 we have +P(S) = {a • a∗ : a ∈ S} = {a ◦ a−1 : a ∈ G} = {d(a) : a ∈ G} = vG = P. +(ii). By Lemma 2.23(ii), and part (i), we have E(S) = {p • q : p, q ∈ P}. The claim then follows +from Lemma 4.38. +(iii). If c = [p1, . . . , pk] ∈ C , then +ε(c) = ε[p1, p2] ◦ ε[p2, p3] ◦ · · · ◦ ε[pk−1, pk] +by (4.13) += ε[p1, p2] • ε[p2, p3] • · · · • ε[pk−1, pk] +by Lemma 4.26 += (p1 • p2) • (p2 • p3) • · · · • (pk−1 • pk) +by Lemma 4.38(ii) += p1 • p2 • · · · • pk ∈ E(S). +Conversely, fix some a ∈ E(S), so that a = p1 • · · · • pk for some p1, . . . , pk ∈ P. We show +that a ∈ im(ε) by induction on k. The k = 1 case being trivial, we assume that k ≥ 2, and let +b = p1 • · · · • pk−1. By induction we have b = ε(c) for some c ∈ C , and we write r = r(b) = r(c). +Then with r′ = pkθr and p′ +k = rθpk, it follows from Lemma 4.37(i) and properties of ε from +Definition 4.11 that +a = b • pk = b⇂r′ ◦ ε[r′, p′ +k] = ε(c)⇂r′ ◦ ε[r′, p′ +k] = ε(c⇂r′) ◦ ε[r′, p′ +k] = ε(c⇂r′ ◦ [r′, p′ +k]) ∈ im(ε). +Remark 4.40. We will see in Chapter 6 (see Theorem 6.7) that every regular ∗-semigroup has +the form S(G, ε) for some chained projection groupoid (G, ε). +Combining this with Proposi- +tion 4.39(iii), we obtain a proof of Proposition 2.42, stated in Chapter 2. +5 +Regular ∗-semigroups +In the previous chapter we showed that a chained projection groupoid (G, ε) gives rise to a +regular ∗-semigroup S = S(G, ε); see Theorem 4.36. In the current chapter we go in the opposite +direction. Starting from a regular ∗-semigroup S, we show how to construct a chained projection +groupoid G(S) = (G, ε) from S; see Definitions 5.4 and 5.16, and Theorem 5.19. In the next +chapter we will show that the S and G constructions are inverse processes, and in fact define +isomorphisms between the categories of regular ∗-semigroups and chained projection groupoids. +The construction of G(S) is given in Section 5.1. We then pause in Section 5.2 to consider +a number of examples. These are intended to aid with understanding of the theory developed +thus far, and also to help uncover some subtleties that arise. +57 + +5.1 +The chained projection groupoid associated to a regular ∗-semigroup +For the duration of this chapter, we fix a regular ∗-semigroup S, as in Definition 2.20. As ever, +we write +P = P(S) = {p ∈ S : p2 = p = p∗} +for the set of projections of S. As in Section 2.3, every projection p ∈ P gives rise to a map +θp : P → P +defined by +qθp = pqp. +(5.1) +Comparing Lemma 2.25 with Definition 3.1, we immediately deduce the following: +Proposition 5.2. If S is a regular ∗-semigroup, then P = P(S) is a projection algebra. +Consequently, we have the partial order ≤ on P given in (3.5), and the relations ≤F and F +on P given in (3.10) and (3.11). In particular (keeping (2.27) in mind), we note that +p ≤ q +⇔ +p = pθq = qpq +⇔ +p = pq = qp +⇔ +p = pq +⇔ +p = qp. +(5.3) +We also have the path category P = P(P) and the chain groupoid C = C (P), as in Defini- +tions 3.17 and 3.39. +Recall that we wish to construct a chained projection groupoid G(S) = (G, ε) from S. The +definition of the groupoid G = G(S) has already been given in Section 2.3; see Definition 2.32 and +Proposition 2.33. Since we wish to show that G is in fact an ordered groupoid, it is convenient +to give the following expanded definition. +Definition 5.4. Given a regular ∗-semigroup S, we define the ordered groupoid G = G(S) as +follows. +• The object set of G is vG = P = P(S). +• For a ∈ G we have d(a) = aa∗ and r(a) = a∗a. +• For a, b ∈ G with r(a) = d(b), we have a ◦ b = ab. +• For a ∈ G we have a−1 = a∗. +The ordering on P = vG is given in (5.3), and left restrictions in G are defined by +p⇃a = pa +for p ≤ d(a). +As usual, right restrictions are given by a⇂q = (q⇃a−1)−1, for q ≤ r(a), and in fact we have +a⇂q = (qa∗)∗ = aq +for such q. For a, b ∈ G, we have +a ≤ b +⇔ +a = p⇃b +for some p ≤ d(b) +⇔ +a = b⇂q +for some q ≤ r(b) +⇔ +a = p⇃b = b⇂q +for some p ≤ d(b) and q ≤ r(b), +and then of course p = d(a) and q = r(a). +Lemma 5.5. If S is a regular ∗-semigroup, then G = G(S) is an ordered groupoid. +58 + +Proof. By Proposition 2.33, we just need to verify conditions (O1)′–(O5)′ from Lemma 2.3, with +respect to the ordering (5.3), and the restrictions given in Definition 5.4. +(O1)′. Consider a morphism a ∈ G, and let p ≤ d(a) = aa∗. It follows from p ≤ aa∗ that +p = paa∗, and so +d(p⇃a) = d(pa) = pa(pa)∗ = paa∗p∗ = pp = p. +We also have +r(p⇃a) = r(pa) = (pa)∗pa = a∗p∗pa = a∗pa. +(5.6) +It then follows that +r(p⇃a) = a∗pa = a∗a · a∗pa · a∗a = r(a) · r(p⇃a) · r(a). +By (5.3), the previous conclusion says precisely that r(p⇃a) ≤ r(a). +(O2)′. By (5.6) we have q = a∗pa, and again p = paa∗ follows from p ≤ d(a). It follows that +(p⇃a)∗ = (pa)∗ = a∗p∗ = a∗p = a∗paa∗ = qa∗ = q⇃a∗. +(O3)′. Here d(a)⇃a = d(a)a = aa∗a = a. +(O4)′. It follows from p ≤ q that p = pq, and so p⇃q⇃a = p(qa) = (pq)a = pa = p⇃a. +(O5)′. Here we again have q = a∗pa. Since p ≤ d(a) we have p = aa∗p, so pa = aa∗pa = aq. It +follows that +p⇃(a ◦ b) = p(ab) = pp(ab) = paqb = p⇃a ◦ q⇃b. +Remark 5.7. In particular, the relation ≤ on the regular ∗-semigroup S(= G) given in Defini- +tion 5.4 is a partial order. This order was in fact used by Imaoka in [71], but in a different form. +Imaoka’s order, which for clarity we will denote by ≤′, was defined by +a ≤′ b +⇔ +a ∈ Pb ∩ bP +where P = P(S). +(5.8) +In light of the rules p⇃b = pb and b⇂q = bq, it is clear that a ≤ b ⇒ a ≤′ b. For the converse, +suppose a ≤′ b, so that +a = pb = bq +for some p, q ∈ P. +In the following calculations, we make extensive use of (5.3). We first note that +a = bq +⇒ +a = bb∗a +⇒ +aa∗ = bb∗aa∗ +⇒ +aa∗ ≤ bb∗. +We also have +a = pb +⇒ +a = pa +⇒ +aa∗ = paa∗ +⇒ +aa∗ ≤ p +⇒ +aa∗ = aa∗p. +We then calculate +a = aa∗a = aa∗(pb) = (aa∗p)b = aa∗b. +Writing r = aa∗ ∈ P, we have already seen that r = aa∗ ≤ bb∗ = d(b), so that in fact +a = aa∗b = rb = r⇃b, and hence a ≤ b. +Remark 5.9. Recall that any regular semigroup S has a so-called natural partial order. This +order, which for clarity we will denote by ⪯, has many different formulations; see for example +[63,64,111], and especially [108] for a discussion of the variations, and an extension to arbitrary +semigroups. +The most straightforward definition of the natural partial order on the regular +semigroup S is: +a ⪯ b +⇔ +a ∈ Eb ∩ bE +where E = E(S). +59 + +It is natural to ask how the orders ≤ and ⪯ are related when S is a regular ∗-semigroup, where ≤ +is the order from Definition 5.4. Since P ⊆ E, and since ≤ is equal to ≤′ (cf. (5.8)), the order ≤ +is of course contained in ⪯, meaning that a ≤ b ⇒ a ⪯ b. But the converse does not hold, +in general. For example, if S is a regular ∗-monoid with identity 1, and if e ∈ E \ P is any +non-projection idempotent, then e ⪯ 1 but e ̸≤ 1. +Next we wish to show that G is a projection groupoid, and to do this we need to understand +the ϑ and Θ maps on G. For this, let a ∈ G. For p ≤ d(a) = aa∗, it follows from (5.6) that +pϑa = r(p⇃a) = a∗pa. +(5.10) +It follows that for arbitrary p ∈ P, +pΘa = pθd(a)ϑa = a∗(d(a) · p · d(a))a = a∗(aa∗ · p · aa∗)a = a∗pa. +(5.11) +In other words, ϑa and Θa both act by ‘conjugation’ on projections (cf. Remark 4.9), but we +note that these maps have different domains: dom(ϑa) = d(a)↓ and dom(Θa) = P. +Lemma 5.12. If S is a regular ∗-semigroup, then G = G(S) is a projection groupoid. +Proof. By Lemma 5.5, it remains to check (G1). Specifically, we show that (G1a) holds, i.e. that +θpϑa = Θa∗θpΘa +for any a ∈ G and p ≤ d(a). +But for any such a and p, and for arbitrary t ∈ P, we use (5.10) and (5.11) to calculate +tθpϑa = tθa∗pa = a∗pa · t · a∗pa = tΘa∗θpΘa. +Next we wish to construct an evaluation map ε = ε(S) : C → G, where C = C (P) is the +chain groupoid of P = P(S). To do so, it is first convenient to define a map +π = π(S) : P → G(= S) +by +π(p1, . . . , pk) = p1 · · · pk, +(5.13) +where P = P(P) is the path category of P. So π(p) is simply the product (in S) of the entries +of p (taken in order). We begin with the following fact: +Lemma 5.14. For any p ∈ P, we have d(π(p)) = d(p) and r(π(p)) = r(p). +Proof. Write p = (p1, . . . , pk). A routine calculation shows that +d(π(p)) = π(p) · π(p)∗ = p1 · · · pk−1 · pk · pk−1 · · · p1. +Since p1 F · · · F pk, this reduces to p1 = d(p). We obtain r(π(p)) = pk = r(p) by symmetry. +The next result concerns the congruence ≈ = Ω♯ on P from Definition 3.36. +Lemma 5.15. The map π = π(S) : P → G is a v-functor, and ≈ ⊆ ker(π). +Proof. For the first statement, it suffices by Lemma 5.14 to show that π is a functor. For this, +consider composable paths p = (p1, . . . , pk) and q = (pk, . . . , pl). Then since pk is an idempotent, +and since r(π(p)) = pk = d(π(q)) by Lemma 5.14, we have +π(p ◦ q) = π(p1, . . . , pl) = p1 · · · pkpk+1 · · · pl += p1 · · · pk · pkpk+1 · · · pl = π(p)π(q) = π(p) ◦ π(q). +To show that ≈ ⊆ ker(π), it suffices to show that Ω ⊆ ker(π), i.e. that π(s) = π(t) for all +(s, t) ∈ Ω. This is clear if (s, t) has the form (Ω1) or (Ω2); for the latter, recall that p = qθp = pqp +60 + +when p F q. So suppose now that (s, t) has the form (Ω3). Consulting Definitions 3.36 and 3.31, +this means that +s = (e, eθp, f) +and +t = (e, fθp, f) +for some p ∈ P, and some p-linked pair (e, f). +Since ep, pf ∈ E(S) by Lemma 2.23(ii), it follows that +π(s) = e · pep · f = epf = e · pfp · f = π(t). +Definition 5.16. Since C = P/≈, it follows from Lemma 5.15 that we have a well-defined +v-functor +ε = ε(S) : C → G +given by +ε[p] = π(p) +for p ∈ P. +That is, ε[p1, . . . , pk] = p1 · · · pk whenever p1 F · · · F pk. +Lemma 5.17. The functor ε = ε(S) : C → G is an evaluation map. +Proof. Conditions (E1) and (E2) hold because ε is a v-functor. As explained in Remark 4.12, +we can prove (E6) in place of (E3), and we note that (E6) says +ε(q⇃[p]) = q⇃ε[p] +for all p ∈ P and q ≤ d[p] = d(p). +Since ε(q⇃[p]) = ε[q⇃p] = π(q⇃p) and q⇃ε[p] = q⇃π(p), this is equivalent to +π(q⇃p) = q⇃π(p) +for all p ∈ P and q ≤ d(p). +(5.18) +We prove this by induction on k, the length of the path p = (p1, . . . , pk). When k = 1, we have +π(q⇃p) = π(q⇃(p1)) = π(q) = q = qp1 = qπ(p) = q⇃π(p), +where we used the fact that q ≤ d(p) = p1 +⇒ +q = qp1. +Now suppose k ≥ 2, and let +q⇃p = (q1, . . . , qk) be as in (3.19). Let p′ = (p1, . . . , pk−1) ∈ P, noting that q⇃p′ = (q1, . . . , qk−1). +Then +q⇃π(p) = q · p1 · · · pk = qp1 · · · pk · (qp1 · · · pk)∗ · qp1 · · · pk += qp1 · · · pk · pk · · · p1q · qp1 · · · pk += q · (p1 · · · pk−1) · (pk · · · p1 · q · p1 · · · pk) += q⇃π(p′) · qθp1 · · · θpk += π(q⇃p′) · qk +by induction and (3.20) += q1 · · · qk−1 · qk = π(q⇃p). +This completes the proof of (5.18), and hence of the lemma. +Note in particular that ε[p, q] = pq for p, q ∈ P with p F q. We are now ready to prove the +main result of this chapter: +Theorem 5.19. If S is a regular ∗-semigroup, then G(S) = (G, ε) is a chained projection +groupoid. +Proof. By Lemmas 5.12 and 5.17, it remains to verify (G2). To do so, let (e, f) be a b-linked +pair, where b ∈ G(= S). Write +q = d(b) = bb∗ +and +r = r(b) = b∗b, +and let the ei, fi be as in (4.19). Keeping (5.11) in mind, we have +e1 = qeq, +e2 = bfb∗, +f1 = b∗eb +and +f2 = rfr. +61 + +We then calculate ee1 = eqeq = eq and bf1 = bb∗eb = qeb. Since e ≤F q by Lemma 4.18(i), we +also have e = qθe = eqe. Putting this all together, we have +λ(e, b, f) = ε[e, e1] ◦ e1⇃b ◦ ε[f1, f] = ee1 · e1b · f1f = ee1 · bf1 · f = eq · qeb · f = eqe · bf = ebf. +A similar calculation gives ρ(e, b, f) = ebf, and the proof is complete. +Remark 5.20. In the above proof, verification of (G2) boils down to checking that +ee1bf1f = ee2bf2f, +where b ∈ G(q, r), (e, f) is b-linked, and the ei, fi are as in (4.19). We are now in a position to see +that condition (G2) cannot be replaced by the stronger (G2)′ discussed in Remark 4.23. Indeed, +consider the partition monoid P4, and the elements defined by +b = +, +e = e1 = f1 = +, +e2 = f2 = +, +f = +. +These elements are all projections (so in particular q = r = b), and they satisfy conditions (i)–(iv) +of Lemma 4.18. However, we have +ee1bf1f = +̸= += ee2bf2f. +5.2 +Examples +At this point it is worth pausing to consider a number of examples in a fair amount of detail. +These are included to illustrate the general theory developed above, but also to demonstrate +some subtleties. +Our first example shows that it is possible for two non-isomorphic regular ∗-semigroups S1 +and S2 to have exactly the same ordered groupoids G(S1) = G(S2). +Example 5.21. Let P be an arbitrary set, and to avoid trivialities we assume that |P| ≥ 2. We +will define two semigroups S1 and S2, both with underlying set (P × P) ∪ {0}. To distinguish +the products, we will denote them by ⋆1 and ⋆2. The element 0 will act as a zero element in +both semigroups, and the rest of the products are defined by +(p, q) ⋆1 (r, s) = (p, s) +and +(p, q) ⋆2 (r, s) = +� +(p, s) +if q = r +0 +otherwise. +So these are in fact special cases of some basic semigroup constructions: +• S1 is a P × P square band with a zero attached, and +• S2 is a P × P combinatorial Brandt semigroup. +Both S1 and S2 are regular ∗-semigroups with respect to the involution +(p, q)∗ = (q, p) +and +0∗ = 0. +In fact, S2 is inverse (as a Brandt semigroup), while S1 is not (as the idempotents (p, p) and +(q, q) do not commute, for distinct p, q ∈ P). In particular, S1 and S2 are not isomorphic; this +also follows from the fact that S1 has no non-trivial zero divisors, while every element of S2 is a +zero divisor. +62 + +The regular ∗-semigroups S1 and S2 have the same projection sets, and the projections are +precisely 0 and the pairs (p, p), for p ∈ P. So, identifying (p, p) ≡ p, we have +P(S1) = P(S2) ≡ P ∪ {0}. +On the other hand, the idempotents are the products of two projections (cf. Lemma 2.23(ii)). +Of course any product involving 0 is 0, while for p, q ∈ P we have +p ⋆1 q = (p, q) +and +p ⋆2 q = +� +p +if p = q +0 +otherwise. +This shows that +E(S1) = S1 +and +E(S2) = P(S2) = P ∪ {0}. +(Of course these facts are also easy to see directly.) This difference is also reflected at the level +of projection algebras. In what follows, we use θi to denote the projection algebra operations +on Si (i = 1, 2), as in (5.1). Of course θ1 +0 = θ2 +0 is the constant map with image {0}, while for +p ∈ P we have +xθ1 +p = +� +p +if x ∈ P +0 +if x = 0 +and +xθ2 +p = +� +p +if x = p +0 +if x = 0 or x ∈ P \ {p}. +(5.22) +Thus, the projection algebras P1 = P(S1) and P2 = P(S2) are different, despite having the same +underlying sets. It is also easy to see that the F relations on P1 and P2 are very different. +Indeed, denoting these by F1 and F2, we have +F1 = ∇P ∪ {(0, 0)} +and +F2 = ∆P∪{0}. +On the other hand, the partial orders on P1 and P2 (see (5.3)) are the same. These orders have +the following Hasse diagram, shown in the case P = {p1, . . . , p5}: +p1 +p2 +p3 +p4 +p5 +0 +(5.23) +It turns out that the groupoids G1 = G(S1) and G2 = G(S2) are exactly the same as well. Indeed, +to see this, we note that the object sets are equal: vG1 = vG2 = P ∪ {0}. Next we note that +Green’s R and L relations are the same on S1 and S2. Indeed, in both semigroups we have +R0 = L0 = {0}, and +(p, q) R (r, s) +⇔ +p = r +and +(p, q) L (r, s) +⇔ +q = s. +It then follows that the non-empty morphism sets in Gi (i = 1, 2) are +Gi(0, 0) = {0} +and +Gi(p, q) = Rp ∩ Lq = {(p, q)} +for p, q ∈ P. +The operations in the groupoids Gi (i = 1, 2) are given by +(p, q) ◦ (q, r) = (p, r) +and +(p, q)−1 = (p, q)∗ = (q, p). +Moreover, the order in both groupoids are the same. Indeed, keeping the order on P1 = P2 +in mind (cf. (5.23)), the morphism 0 is below every other morphism, and there are no other +non-trivial relations. +63 + +We have just seen that it is possible for two non-isomorphic regular ∗-semigroups S1 ̸∼= S2 +to have exactly the same ordered groupoids, G(S1) = G(S2). +It follows from this that the +groupoid alone is not enough information to distinguish regular ∗-semigroups; in other words, +the groupoid is not a total invariant of a regular ∗-semigroup. +However, the semigroups S1 +and S2 in Example 5.21 had different projection algebra structures. +Thus, even though the +groupoids G1 and G2 are identical (in terms of their compositions, inversions and orderings), we +can still distinguish them by means of the underlying projection algebra structure of their object +sets, vG1 ̸∼= vG2. +This may seem like a strange point, as the projection algebra structure is somewhat ‘invisible’ +in the structure of the groupoid G. In a sense, the projection algebra is merely used as a means +to define the order on vG, which then feeds in to the order on G itself. However, G is not only +an ordered groupoid, but in fact a projection groupoid (cf. Definition 4.8), and this means that +there is a stronger link between the structures of G and vG than merely the order-theoretic one +just described. Indeed, the defining properties (G1a)–(G1d) from Definition 4.8 involve the Θa +maps from (4.1), which are defined in terms of the order-theoretic properties of G (specifically, +the ϑa maps) and the projection algebra structure of vG (specifically, the θp maps). And it is +easy to see that the groupoids G1 = G(S1) and G2 = G(S2) have different Θ maps. Indeed, again +distinguishing these by using the symbols Θ1 and Θ2, we even have Θ1 +p = θ1 +p ̸= θ2 +p = Θ2 +p for +p ∈ P; cf. (4.2) and (5.22). +In any case, one may wonder if the pair (G, P) consisting of the groupoid G = G(S) and the +projection algebra P = P(S), including its structure as determined by the θ operations, might +be a total invariant for a regular ∗-semigroup. The next example shows that this is still not the +case, and illustrates the need for the evaluation map in the chained projection groupoid (G, ε). +Example 5.24. Fix a group G, and an arbitrary set P. Also let M = (mpq)p,q∈P be a P × P +‘sandwich matrix’ over G with the property that +mqp = m−1 +pq +for all p, q ∈ P. +(5.25) +Note that (5.25) forces mpp = 1 for all p ∈ P, where here 1 denotes the identity of the group G. +The Rees matrix semigroup SM = M (G, P, M) has underlying set P × G × P, and product +(p, g, q)(r, h, s) = (p, gmqrh, s). +One can easily check (using (5.25)) that SM becomes a regular ∗-semigroup under the involution +given by +(p, g, q)∗ = (q, g−1, p). +The projections of SM are the elements (p, 1, p) for each p ∈ P, and we identify p ≡ (p, 1, p). +With this identification, one can easily check that pqp = p for all p, q ∈ P, so that each θp map +(as in (5.1)) is the constant map with image {p}. It follows that the projection algebra structure +of P(SM) is independent of the sandwich matrix M, and depends only on the set P. Note also +that the law p = pqp = qθp gives p ≤F q for all p, q ∈ P, which means that F = ∇P is the +universal relation on P. On the other hand, the order ≤ on P is trivial, as p ≤ q ⇔ p = pθq = q. +As ever, the idempotents of SM are the products of two projections: +pq ≡ (p, 1, p)(q, 1, q) = (p, mpq, q) +for p, q ∈ P. +In particular, the idempotents of SM depend on the sandwich matrix M. +We now consider the groupoid G = G(SM). (The reason for not including an M-subscript +on G will become clear shortly.) As ever, the object set of G is vG = P. Green’s R and L +relations on SM are given by +(p, g, q) R (r, h, s) +⇔ +p = r +and +(p, g, q) L (r, h, s) +⇔ +q = s. +64 + +In particular, the morphism sets of G are given by +G(p, q) = Rp ∩ Lq = {p} × G × {q} = {(p, g, q) : g ∈ G} +for each p, q ∈ P. +The operations in the groupoid G are given by +(p, g, q) ◦ (q, h, r) = (p, gh, r) +and +(p, g, q)−1 = (p, g, q)∗ = (q, g−1, p). +Note that the above rule for composition in G holds because of the identity mqq = 1 (cf. (5.25)). +In particular, these operations are independent of the sandwich matrix M. Since the ordering +on P is trivial, so too is the ordering on G. +Thus, given distinct matrices M1 and M2 satisfying (5.25), the semigroups S1 = SM1 and +S2 = SM2 will give rise to identical projection algebras P(S1) = P(S2), including the θ operations, +and identical groupoids G(S1) = G(S2). +However, the semigroups S1 and S2 are of course +different, as they will have different products (since M1 ̸= M2). Moreover, S1 and S2 need not +even be isomorphic. Indeed, if every entry of M1 is 1 (the identity of G), then S1 is isomorphic +to the direct product of G with the square band P ×P, and in particular the idempotents E(S1) +form a subsemigroup of S1. But this need not be the case in general; indeed, if |G| ≥ 2 and +|P| ≥ 3, and if we choose the entries of M2 = (mpq) in such a way that mpq = mqr = 1 ̸= mpr +for distinct p, q, r ∈ P, then (p, 1, q) and (q, 1, r) are both idempotents of S2, but their product +(p, 1, q)(q, 1, r) = (p, 1, r) ̸= (p, mpr, r) +is not. Thus, S1 ̸∼= S2 in this case. +It follows from this that the pair (G, P) consisting of the groupoid G = G(S) and projection +algebra P = P(S) is not a total invariant of a regular ∗-semigroup. It turns out that the missing +ingredient is the evaluation map ε = ε(S) from Definition 5.16. Indeed, we will see that the +pair (G, ε), i.e. the chained projection groupoid associated to S, is a total invariant. +Going back to the Rees matrix semigroup SM, we now consider the evaluation map +εM = ε(SM) : C → G, +where here C = C (P) is the chain groupoid of P. As in (4.13), and keeping F = ∇P in mind, εM +is determined entirely by the elements εM[p, q], for p, q ∈ P. Following Definition 5.16, we have +εM[p, q] = pq ≡ (p, mpq, q). +In particular, the evaluation map εM depends on the sandwich matrix M. +In a sense, the previous conclusion is reversible. Indeed, if we start from the pair (G, P), +consisting of the groupoid G, and the projection algebra P (with each θp being the constant map +with image {p}), we have immense freedom in defining an evaluation map ε : C → G. Indeed, we +can do so by defining the elements ε[p, q], for p, q ∈ P, to be arbitrary morphisms from G(p, q), as +long as we ensure that ε[q, p] = ε[p, q]−1 for all p, q. Since ε[p, q] must be of the form (p, mpq, q) +for some mpq ∈ G, and since then ε[p, q]−1 = (q, m−1 +pq , p), this condition on inverses is of course +equivalent to (5.25). Thus, roughly speaking, evaluation maps C → G and sandwich matrices +satisfying (5.25) are one and the same thing. +On the other hand, the idempotent structure of SM, as determined by the biordered set E(SM), +is independent of M. The biordered set of a semigroup S consists of the set E(S) of idempotents, +together with a partial binary operation (e, f) �→ ef (a restriction of the product in S), defined +precisely when ef and/or fe is equal to one of e or f. These ‘basic products’ can be conveniently +depicted using Easdown’s arrow notation from [31]; we write +e +f +or +e +f +65 + +to indicate that ef = e (e is a left zero for f) or fe = e (e is a right zero for f), respectively. In +the case that S = SM, we have seen that E = E(S) consists of the elements epq = (p, mpq, q), +for p, q ∈ P. One can easily check that +epq +ers +⇔ +p = r, +epq +ers +⇔ +q = s. +The entire biordered structure of E(S) is then as follows, in the case P = {1, 2, 3, 4}, and +remembering that the +and +relations are transitive: +e11 +e21 +e31 +e41 +e12 +e22 +e32 +e42 +e13 +e23 +e33 +e43 +e14 +e24 +e34 +e44 +6 +The category isomorphism +We are now ready to bring all of the above ideas together. The previous two chapters gave +constructions for: +• building a regular ∗-semigroup S(G, ε) from a chained projection groupoid (G, ε), and +• building a chained projection groupoid G(S) from a regular ∗-semigroup S. +As we have already discussed, one of the goals of the current chapter is to show that the S and G +constructions are inverse processes, meaning that +G(S(G, ε)) = (G, ε) +and +S(G(S)) = S +for any chained projection groupoid (G, ε), and any regular ∗-semigroup S. In fact, we go further +than this, and show in Theorem 6.7 below that S and G are mutually inverse isomorphisms +between the categories of regular ∗-semigroups and chained projection groupoids. +Let RSS be the (large) category of all regular ∗-semigroups. Morphisms in RSS are the +∗-semigroup homomorphisms, i.e. the maps +φ : S1 → S2 +such that +(ab)φ = (aφ)(bφ) +and +(a∗)φ = (aφ)∗ +for all a, b ∈ S1. +We also define CPG to be the (large) category of all chained projection groupoids. Mor- +phisms in CPG are what we call chained projection functors: +66 + +Definition 6.1. Consider two chained projection groupoids (G1, ε1) and (G2, ε2). For i = 1, 2, +we write Pi = vGi, Ci = C (Pi), and so on. A chained projection functor from (G1, ε1) to (G2, ε2) +is an ordered functor φ : G1 → G2 satisfying the following two conditions: +(F1) The object map vφ : P1 → P2 (i.e. the restriction of φ to P1 = vG1) is a projection algebra +morphism, meaning that +(pθq)φ = (pφ)θqφ +for all p, q ∈ P1. +(Here we use θ to denote the projection algebra operations on both P1 and P2.) +(F2) The functor φ respects the evaluation maps, meaning that the following diagram commutes: +C1 +C2 +G1 +G2, +Φ +ε1 +ε2 +φ +where here Φ : C1 → C2 is the induced functor given by [p1, . . . , pk]Φ = [p1φ, . . . , pkφ], as +in Proposition 3.41. Commutativity of the diagram means that +(ε1(c))φ = ε2(cΦ) +for all c ∈ C1. +The main result of this chapter, Theorem 6.7 below, states that the categories RSS and +CPG are isomorphic. We build towards the theorem with a series of lemmas. +Lemma 6.2. If S1 and S2 are regular ∗-semigroups, then any ∗-semigroup homomorphism +S1 → S2 is a chained projection functor G(S1) → G(S2). +Proof. Fix a ∗-semigroup homomorphism φ : S1 → S2, and write G(Si) = (Gi, εi), Pi = P(Si) = vGi, +Ci = C (Pi), and so on. Since the underlying sets of G1 and G2 are S1 and S2, we can think of φ +as a map G1 → G2. Since φ is a ∗-semigroup homomorphism, it follows that for a ∈ G1 we have +d(aφ) = (aφ)(aφ)∗ = (aφ)(a∗φ) = (aa∗)φ = d(a)φ +and similarly +r(aφ) = r(a)φ. +But then if a, b ∈ G1 are such that r(a) = d(b), then r(aφ) = d(bφ), so we have +(a ◦ b)φ = (ab)φ = (aφ)(bφ) = (aφ) ◦ (bφ). +Thus, φ is a functor. +Before we check that φ is order-preserving, we note that (F1) holds, as for any p, q ∈ P1, +(pθq)φ = (qpq)φ = (qφ)(pφ)(qφ) = (pφ)θqφ. +To see that φ preserves the order, suppose a, b ∈ G1 are such that a ≤ b. Then a = p⇃b = pb +for some p ≤ d(b). Note that +p ≤ d(b) +⇒ +p = pθd(b) +⇒ +pφ = (pφ)θd(b)φ +⇒ +pφ ≤ d(b)φ = d(bφ). +It then follows that +aφ = (pb)φ = (pφ)(bφ) = pφ⇃(bφ) ≤ bφ. +It remains to check that (F2) holds. For this, let c = [p1, . . . , pk] ∈ C1. Then +(ε1(c))φ = (p1 · · · pk)φ = (p1φ) · · · (pkφ) = ε2[p1φ, . . . , pkφ] = ε2(cΦ). +67 + +Lemma 6.3. If (G1, ε1) and (G2, ε2) are chained projection groupoids, then any chained projection +functor (G1, ε1) → (G2, ε2) is a ∗-semigroup homomorphism S(G1, ε1) → S(G2, ε2). +Proof. Fix a chained projection functor φ : G1 → G2. Also write S1 = S(G1, ε1) and S2 = S(G2, ε2). +Again we can think of φ as a map S1 → S2. Since φ is a groupoid functor, we have +(a∗)φ = (a−1)φ = (aφ)−1 = (aφ)∗ +for all a ∈ S1. +To make the rest of this proof notationally concise, we will write a = aφ for a ∈ S1. It remains +to show that φ is a semigroup homomorphism, i.e. that a • b = a • b for all a, b ∈ S1. (Here we +write • for the product on both S1 and S2.) For this, write as usual +p = r(a), +q = d(b), +p′ = qθp +and +q′ = pθq, +so that +a • b = a⇂p′ ◦ ε1[p′, q′] ◦ q′⇃b. +Since φ is a functor, we have +r(a) = r(a ◦ p) = r(a ◦ p) = r(p) = p +and similarly +d(b) = q. +Thus, +a • b = a⇂p′ ◦ ε2[p′, q′] ◦ q′⇃b +where +p′ = qθp +and +q′ = pθq. +By (F1) we have +p′ = qθp = qθp = p′ +and similarly +q′ = q′. +By (F2) we have +ε1[p′, q′] = (ε1[p′, q′])φ = ε2([p′, q′]Φ) = ε2[p′, q′] = ε2[p′, q′]. +Since φ is an ordered functor, we have +a⇂p′ = a⇂p′ = a⇂p′ +and similarly +q′⇃b = q′⇃b. +Putting everything together, we have +a • b = a⇂p′ ◦ ε1[p′, q′] ◦ q′⇃b = a⇂p′ ◦ ε1[p′, q′] ◦ q′⇃b = a⇂p′ ◦ ε2[p′, q′] ◦ q′⇃b = a • b. +Lemma 6.4. If (G, ε) is a chained projection groupoid, then G(S(G, ε)) = (G, ε). +Proof. Write +S = S(G, ε) +and +(G′, ε′) = G(S). +We must show that G′ = G and ε′ = ε. By the equality G′ = G, we mean that these have +the same underlying sets, and all the same operations and relations (composition, inversion, +order and projection algebra operations). The underlying sets of S and G′ are simply G, and +the unary operations of S and G′ are simply inversion in G. We also have vG = P(S) = vG′ +(cf. Proposition 4.39(i)), and we denote this set of projections by P. +Recall that the product • in S is given by Definition 4.24. The groupoid G′ is constructed +from S, as outlined in Definition 5.4. To avoid ambiguity, we will denote the composition on G′ +by ⊚ in order to distinguish it from the original composition ◦ in G. For the same reason, we +also denote the domain and range maps in G′ by d′ and r′. So for a, b ∈ G′(= G), we have +d′(a) = a • a−1, +r′(a) = a−1 • a +and +a ⊚ b = a • b when r′(a) = d′(b). +By Lemma 4.26, and since r(a) = d(a−1), it follows that +d′(a) = a • a−1 = a ◦ a−1 = d(a) +and similarly +r′(a) = r(a). +(6.5) +68 + +It also follows that +r′(a) = d′(b) +⇔ +r(a) = d(b) +and then +a ⊚ b = a • b = a ◦ b. +Thus, the compositions in the categories G and G′ coincide. +Next we consider the projection algebra operations. We denote those in G as usual by θp +(p ∈ P), and those in G′ by θ′ +p (p ∈ P), remembering that vG′ = vG = P. We must show that +qθp = qθ′ +p for all p, q ∈ P, and this in fact follows from Lemma 4.38(iii): +qθ′ +p = p • q • p = qθp. +Since θp = θ′ +p for all p ∈ P = vG = vG′, it follows that p ≤ q in G precisely when p ≤ q in G′. +We now move on to the orders on G and G′. We use the symbols ⇃ and ⇂ to denote restrictions +in G as usual, and ↿ and ↾ for restrictions in G′ (note the placement of the arrow heads). Let +a ∈ G and write q = d(a). By (6.5), p⇃a is defined in G precisely when p↿a is defined in G′, +i.e. when p ≤ q. And for such p, it follows from Definition 5.4 and Lemma 4.37(vi) that +p↿a = p • a = p⇃a. +But then for a, b ∈ G, and writing ≤ and ≤′ for the orders in G and G′, we have +a ≤ b +⇔ +a = p⇃b for some p ≤ d(b) +⇔ +a = p↿b for some p ≤ d′(b) +⇔ +a ≤′ b. +All of the above shows that the ordered groupoids G and G′ coincide. It remains to show that +ε = ε′. To do so, fix some c = [p1, . . . , pk] ∈ C . By Definition 5.16 we have +ε′(c) = p1 • · · · • pk, +and we show that ε(c) = ε′(c) by induction on k. The k = 1 case being clear, suppose k ≥ 2, +and write d = [p1, . . . , pk−1] and a = ε(d). Note that c = d ◦ [pk−1, pk], and by induction we have +a = ε(d) = ε′(d) = p1 • · · · • pk−1. +It follows that ε′(c) = a•pk. Since r(a) = r(d) = pk−1 (as ε′ is a v-functor), and since pk−1 F pk +(as c ∈ C ), it follows from Lemma 4.37(iii) that +ε′(c) = a • pk = a ◦ ε[pk−1, pk] = ε(d) ◦ ε[pk−1, pk] = ε(d ◦ [pk−1, pk]) = ε(c), +as required. +Lemma 6.6. If S is a regular ∗-semigroup, then S(G(S)) = S. +Proof. Write +(G, ε) = G(S) +and +S′ = S(G, ε). +Again, by construction the underlying sets of S and S′ coincide, and so do the involutions. It +therefore remains to show that the products in S and S′ coincide as well. We denote the product +in S by juxtaposition, as usual. The product in S′ is •, which is constructed from (G, ε) as in +Definition 4.24, while (G, ε) is in turn constructed from S as in Definitions 5.4 and 5.16. +So let a, b ∈ S; we must show that a • b = ab. Write p = r(a) and q = d(b), so that +a • b = a⇂p′ ◦ ε[p′, q′] ◦ q′⇃b +where +p′ = qθp +and +q′ = pθq. +Following the definitions in Chapter 5, and using Lemma 2.23, we have +p = a∗a, +p′ = pqp, +a⇂p′ = ap′ = apqp, +q = bb∗, +q′ = qpq, +q′⇃b = q′b = qpqb, +ε[p′, q′] = p′q′ = pqpqpq = pq. +Combining all of the above, and keeping Definition 5.4 and Lemma 2.23 in mind, it follows that +a • b = apqp ◦ pq ◦ qpqb = apqpqpqb = apqb = aa∗abb∗b = ab, +as required. +69 + +We now have all we need to prove the main result. +Theorem 6.7. The category RSS of regular ∗-semigroups (with ∗-semigroup homomorphisms) +is isomorphic to the category CPG of chained projection groupoids (with chained projection +functors). +Proof. By Theorems 4.36 and 5.19 we have maps +S : CPG → RSS : (G, ε) �→ S(G, ε) +and +G : RSS → CPG : S �→ G(S). +By Lemmas 6.2 and 6.3, S and G are functors, and by Lemmas 6.4 and 6.6 they are mutually +inverse. +Remark 6.8. We can use Theorem 6.7 to deduce a structure theorem for regular ∗-semigroups, in +the sense that it shows how the entire structure of such a semigroup can be entirely, and uniquely, +determined by ‘simpler’ objects. Specifically, given a regular ∗-semigroup S, one constructs: +• the projection groupoid G = G(S), as in Definition 5.4, including its object set vG = P(S), +which is the projection algebra of S, and +• the evaluation map ε = ε(S), as in Definition 5.16. +Lemma 6.6 tells us that S is completely determined by these two simpler objects, in the sense +that S = S(G, ε) can be constructed from G and ε in the manner described in Definition 4.24. +Examples 5.21 and 5.24 demonstrate the subtlety of the decomposition, showing how seemingly- +small changes to G and/or ε can result in very different semigroups. +Theorem 6.7 shows that the categories of regular ∗-semigroups and chained projection groupoids +are in a sense the same. It also follows from the definition of the functors S and G that the +algebras of projections of regular ∗-semigroups are precisely the object sets of chained projection +groupoids. However, we are still left with the following two (equivalent) questions, which we will +explore in some depth in the second part of the paper: +Question 6.9. Given an (abstract) projection algebra P, as in Definition 3.1: +• does there exist a regular ∗-semigroup S with P(S) = P? +• does there exist a chained projection groupoid (G, ε) with vG = P? +Question 6.9 (or at least the first version of it) is already known to have an affirmative +answer. +For example, Imaoka [70, 72] and Yamada [138] have both shown how to construct +a maximum fundamental regular ∗-semigroup with a given projection algebra P (definitions +are given below), though Yamada worked with a somewhat different formulation of projection +algebras. This fundamental semigroup was also studied by Jones in [73, Section 4], but he was +not concerned with maximality. As an application of our results, we will give another (ultimately +equivalent) construction of this maximum fundamental semigroup in Chapter 8. We also answer +Question 6.9 by an alternative route in Chapter 7, by constructing a free (projection-generated) +regular ∗-semigroup with projection algebra P. As the name suggests, these semigroups have +many remarkable (categorical) ‘free-ness’ properties, which will be elaborated upon further in +Chapter 7, and can be constructed in a variety of ways, including by a presentation by generators +and relations (see Theorem 7.14). +70 + +Part II +Applications +The first part of this paper was devoted to a proof of our main structural result, Theorem 6.7, +which established the isomorphism of the categories of regular ∗-semigroups and chained projec- +tion groupoids. In this second part, we explore a number of consequences of this theorem: +• Chapter 7 uses chain groupoids to construct the free regular ∗-semigroup over an arbitrary +projection algebra; see Theorems 7.11, 7.14, 7.20 and 7.23. +• Chapter 8 uses projection groupoids to give a new, transparent, construction of fundamental +regular ∗-semigroups; see Theorems 8.20, 8.21 and Theorem 8.24. +• Chapter 9 applies our theory to the special case of inverse semigroups, leading to a new proof +of the celebrated Ehresmann–Schein–Nambooripad Theorem, stated in Theorem 9.1. +Again, the introduction of each chapter contains a detailed outline of its structure, and a summary +of the main results. +7 +Idempotent-generated regular ∗-semigroups +As discussed at the end of Chapter 6, one of our motivations at this point is to show how to +construct, from an arbitrary projection algebra P (cf. Definition 3.1), a regular ∗-semigroup S +with projection algebra P(S) = P or, equivalently, a chained projection groupoid (G, ε) with +object set vG = P. As it happens, we have already constructed such a groupoid, namely the chain +groupoid C = C (P) from Definition 3.39, with evaluation map simply the identity idC , although +we still have to verify quite a few non-trivial details. The chained projection groupoid (C , idC ) +leads as usual to the regular ∗-semigroup S(C , idC ), which we will call the chain semigroup of P, +and denote by CP . As we will see, this semigroup possesses many important properties, beyond +simply being a regular ∗-semigroup with projection algebra P(CP ) = P. +In Section 7.1 we prove Theorem 7.11, which shows that CP is universal among idempotent- +generated regular ∗-semigroups in the sense that: +• CP is an idempotent-generated regular ∗-semigroup with projection algebra P, and +• any idempotent-generated regular ∗-semigroup with projection algebra P is a homomorphic +image of CP . +(Recall from Lemma 2.23(ii) that a regular ∗-semigroup is idempotent-generated if and only if it +is projection-generated.) As explained in Remark 7.12, this shows that CP is a free/initial object +in the category IGRSSP of idempotent-generated regular ∗-semigroups with (fixed) projection +algebra P. In Section 7.2 we give a presentation for CP in terms of generators and relations; +see Theorem 7.14. +We then return in Section 7.3 to our discussion of the ‘free-ness’ of CP . +Theorem 7.20 demonstrates another universal property of CP among all regular ∗-semigroups: +• For any projection algebra P and regular ∗-semigroup S, any projection algebra morphism +φ : P → P(S) extends uniquely to a ∗-semigroup morphism CP → S. +This is then used to prove Theorem 7.23, which says that CP is the free regular ∗-semigroup +with projection algebra P, in the sense that the semigroups CP are precisely the objects in the +image of a left adjoint of a suitable forgetful functor. (The relevant definitions are given below.) +We then conclude the chapter with some open questions. +71 + +7.1 +The chain semigroup +For the duration of this section we fix a projection algebra P, as in Definition 3.1, and all of +the data that comes with it, i.e. the unary operations θp, the relations ≤, ≤F and F, and so +on. We also write P = P(P) and C = C (P) for the path category and chain groupoid of P; +cf. Definitions 3.17 and 3.39. +Our first goal is to show that C is a projection groupoid, for which it remains to ver- +ify (G1), specifically (G1a). To do so, we need to understand the maps ϑc and Θc, for a P-chain +c = [p1, . . . , pk] ∈ C . +Keeping d(c) = p1 and r(c) = pk in mind, and using (3.20), the map +ϑc : p↓ +1 → p↓ +k is given by +qϑc = r(q⇃c) = qθp2 · · · θpk = qθp1 · · · θpk +for q ≤ p1. +(7.1) +It follows from this that qΘc = qθd(c)ϑc = qθp1θp2 · · · θpk for arbitrary q ∈ P. This all shows that +Θc = θp1 · · · θpk +for c = [p1, . . . , pk] ∈ C . +(7.2) +Lemma 7.3. For any projection algebra P, the chain groupoid C = C (P) is a projection +groupoid. +Proof. To verify (G1a), consider a P-chain c = [p1, . . . , pk] ∈ C , and let q ≤ d(c) = p1. We +must show that +θqϑc = Θc−1θqΘc. +For this we use (7.1), Lemma 3.4 and (7.2) to calculate +θqϑc = θqθp1···θpk = θpk · · · θp1θqθp1 · · · θpk = Θc−1θqΘc, +where in the last step we recall that c−1 = [pk, . . . , p1]. +Next, it is clear that the identity map ε = idC is an evaluation map (cf. Definition 4.11). +Proposition 7.4. For any projection algebra P, (C , idC ) is a chained projection groupoid, where +C = C (P) is the chain groupoid of P. +Proof. It remains to verify (G2). To do so, fix a c-linked pair (e, f), where c = [p] ∈ C for +p = (p1, . . . , pk) ∈ P. Following Definitions 4.16, 4.20 and 4.22 this means that +f = eΘcθf +and +e = fΘc−1θe, +(7.5) +and we must show that λ(e, c, f) = ρ(e, c, f). Keeping ε = idC in mind, these morphisms are +defined by +λ(e, c, f) = [e, e1] ◦ e1⇃c ◦ [f1, f] +and +ρ(e, c, f) = [e, e2] ◦ c⇂f2 ◦ [f2, f], +(7.6) +in terms of the projections +e1 = eθp1, +e2 = fΘc−1, +f1 = eΘc +and +f2 = fθpk. +For convenience, we will write +e1⇃c = [u1, . . . , uk] +and +c⇂f2 = [v1, . . . , vk]. +Using (3.19) and (3.30), and remembering that e1 = eθp1 and f2 = fθpk, we have +ui = e1θp2 · · · θpi = eθp1 · · · θpi +and +vi = f2θpk−1 · · · θpi = fθpk · · · θpi, +72 + +for each 1 ≤ i ≤ k. Keeping in mind that the compositions in (7.6) exist, it follows from all this +that +λ(e, c, f) = [e, u1, . . . , uk, f] +and +ρ(e, c, f) = [e, v1, . . . , vk, f], +and we must show that these chains are equal. It will also be convenient to additionally write +u0 = e and vk+1 = f. To assist with understanding the coming arguments, these projections are +shown in Figure 16 (in the case k = 4). Our task is essentially to show that the large rectangle +at the bottom of the diagram commutes, modulo ≈. +Next we claim that (ui−1, vi+1) is pi-linked for each 1 ≤ i ≤ k, as in Definition 3.31. To prove +this, we must show that +vi+1 = ui−1θpiθvi+1 +and +ui−1 = vi+1θpiθui−1. +First we note that (7.2) and (7.5) give +f = eθp1 · · · θpkθf +and +e = fθpk · · · θp1θe. +Combining this with Lemma 3.4, we obtain +ui−1θpiθvi+1 = eθp1 · · · θpi−1 · θpi · θfθpk···θpi+1 += eθp1 · · · θpi−1 · θpi · θpi+1 · · · θpkθfθpk · · · θpi+1 = fθpk · · · θpi+1 = vi+1. +The proof that ui−1 = vi+1θpiθui−1 is essentially identical. Now that we have proved the claim, +these linked pairs lead to the paths +λ(ui−1, pi, vi+1) = (ui−1, ui−1θpi, vi+1) +and +ρ(ui−1, pi, vi+1) = (ui−1, vi+1θpi, vi+1) += (ui−1, ui, vi+1) += (ui−1, vi, vi+1), +as in Definition 3.31. It then follows by the form of Ω in Definition 3.36 (see (Ω3)) that +[ui−1, vi, vi+1] = [ui−1, ui, vi+1] +for each 1 ≤ i ≤ k. +(7.7) +In other words, each of the small rectangles at the bottom of Figure 16 commute. It follows, +therefore, that the large rectangle commutes. Formally, we repeatedly use (7.7) in the indicated +places, as follows: +ρ(e, c, f) = [e, v1, . . . , vk, f] = [u0, v1, v2, v3, v4, . . . , vk, vk+1] += [u0, u1, v2, v3, v4, . . . , vk, vk+1] += [u0, u1, u2, v3, v4, . . . , vk, vk+1] +... += [u0, u1, u2, u3, u4, . . . , uk, vk+1] = [e, u1, . . . , uk, f] = λ(e, c, f), +and the proof is complete. +As a result of Proposition 7.4 and Theorem 4.36, we have a regular ∗-semigroup S(C , idC ), +which we will denote by CP . To give an explicit description of the • product in CP , consider +an arbitrary pair of P-chains c = [p1, . . . , pk] and d = [q1, . . . , ql]. Also write p = r(c) = pk and +q = d(d) = q1. Following Definition 4.24, and remembering that the evaluation map is idC , we +have +c • d = c⇂p′ ◦ [p′, q′] ◦ q′⇃d +where +p′ = qθp +and +q′ = pθq. +As in (3.19) and (3.30), we have c⇂p′ = [p′ +1, . . . , p′ +k] and q′⇃d = [q′ +1, . . . , q′ +l], where +p′ +i = p′θpk · · · θpi +and +q′ +j = q′θq1 · · · θqj +for 1 ≤ i ≤ k and 1 ≤ j ≤ l, +73 + +e = u0 +v5 = f +p1 +u1 +v1 +p2 +u2 +v2 +p3 +u3 +v3 +p4 +u4 +v4 +Figure 16. The projections e, f, pi, ui, vi from the proof of Proposition 7.4, shown here in the case +k = 4. Dashed lines indicate ≤ relationships. Each arrow s → t represents the P-path (s, t) ∈ P, so +the upper and lower paths e → f represent (e, u1, . . . , uk, f) and (e, v1, . . . , vk, f), respectively. +and where p′ = p′ +k and q′ = q′ +1. Then +c • d = c⇂p′ ◦ [p′, q′] ◦ q′⇃d = [p′ +1, . . . , p′ +k] ◦ [p′ +k, q′ +1] ◦ [q′ +1, . . . , q′ +l] = [p′ +1, . . . , p′ +k, q′ +1, . . . , q′ +l] +is simply the concatenation of c⇂p′ and q′⇃d. We denote the concatenation of a, b ∈ C by a ⊕ b, +but we note that this only belongs to C if r(a) F d(b). As special cases we have +c • d = c ⊕ d if r(c) F d(d) +and +c • d = c ◦ d if r(c) = d(d). +Definition 7.8. Given a projection algebra P (cf. Definition 3.1), we define the regular ∗- +semigroup +CP = S(C , idC ), +where C = C (P) is the chain groupoid of P (cf. Definition 3.39). Explicitly: +• The elements of CP are the P-chains, [p1, . . . , pk], as in Definition 3.39. +• The product • in CP is given, for c, d ∈ CP with r(c) = p and d(d) = q, by +c • d = c⇂p′ ⊕ q′⇃d +where +p′ = qθp +and +q′ = pθq, +and where ⊕ denotes concatenation, as above. +• The involution in CP is given by reversal of P-chains, [p1, . . . , pk]∗ = [pk, . . . , p1]. +• The projections of CP are the trivial chains [p] ≡ p, for p ∈ P, and consequently P(CP ) ≡ P. +• The idempotents of CP are the chains [p, q], for (p, q) ∈ F. +We call CP the chain semigroup of P. +For the coming results of this chapter, and for later use, we require the following definition: +Definition 7.9. Let S and S′ be regular ∗-semigroups with the same projection algebras, +P(S) = P(S′) = P. +A ∗-semigroup homomorphism φ : S → S′ is said to be canonical if +φ|P = idP , i.e. if pφ = p for all p ∈ P. +74 + +Proposition 7.10. For any chained projection groupoid (G, ε), the evaluation map ε is a chained +projection functor (C , idC ) → (G, ε), and (consequently) a canonical ∗-semigroup homomorphism +CP → S(G, ε). +Proof. By Lemma 6.3 (and since the v-functor ε : C → G has object map vε = idP ), it suffices +to prove the first claim. Since ε is an ordered v-functor C → G (cf. Definition 4.11), we just need +to check that conditions (F1) and (F2) both hold. The first follows immediately from vε = idP . +For (F2), vε = idP means that the induced map Φ : C → C at the top of the diagram in +Definition 6.1 is the identity map. Thus, this diagram is +C +C +C +G, +idC +idC +ε +ε +which obviously commutes. +Theorem 7.11. If P is a projection algebra, then +(i) CP is an idempotent-generated regular ∗-semigroup with projection algebra P, +(ii) any idempotent-generated regular ∗-semigroup with projection algebra P is a canonical im- +age of CP . +Proof. (i). +By Theorem 4.36, CP = S(C , idC ) is a regular ∗-semigroup. +It is idempotent- +generated by Proposition 4.39(iii), since the evaluation map idC : C → C is obviously surjective. +We have already observed that P(CP ) ≡ P. +(ii). Let S be an idempotent-generated regular ∗-semigroup with projection algebra P. By Propo- +sition 7.10 (and Theorem 6.7), ε is a canonical ∗-semigroup homomorphism CP → S(G(S)) = S, +and by Proposition 4.39(iii), it is surjective. +Remark 7.12. We now briefly comment on a categorical interpretation of Theorem 7.11. For a +fixed projection algebra P, we denote by IGRSSP the (typically large) category of idempotent- +generated regular ∗-semigroups with projection algebra P. The morphisms in IGRSSP are the +canonical ∗-homomorphisms from Definition 7.9. Since every object in IGRSSP is generated +by P (cf. Lemma 2.23(ii)), and since every morphism in IGRSSP maps P identically, there can +be at most one such morphism S → S′, for any pair of objects S, S′ of IGRSSP , which must +then be surjective. +Theorem 7.11 essentially states that CP is a free/initial object in IGRSSP . We will say +more about the free-ness of CP (in the full category RSS) in Section 7.3; see in particular +Theorems 7.20 and 7.23. +Before we move on, we record the following simple consequence of our previous results. +Proposition 7.13. Any projection algebra morphism φ : P → P ′ extends to a (unique) well- +defined ∗-semigroup homomorphism +Φ : CP → CP ′ +given by +[p1, . . . , pk]Φ = [p1φ, . . . , pkφ]. +Proof. By Proposition 3.41, the mapping Φ is an ordered groupoid functor C → C ′, where as +usual we write C = C (P) and C ′ = C (P ′). By Lemma 6.3, it suffices to show that Φ is in +fact a chained projection functor (C , idC ) → (C ′, idC ′), as it will then also be a ∗-semigroup +75 + +homomorphism CP = S(C , idC ) → S(C ′, idC ′) = CP ′. Consulting Definition 6.1, we are left to +show that Φ satisfies conditions (F1) and (F2). +(F1). By definition we have vΦ = φ, and this is a projection algebra morphism by assumption. +(F2). Keeping in mind that the evaluation maps are the identities, the diagram in Definition 6.1 +becomes +C +C ′ +C +C ′, +Φ +idC +idC′ +Φ +and this obviously commutes. +7.2 +Presentation by generators and relations +Our next result is a presentation by generators and relations for the chain semigroup CP . Roughly +speaking, we take an abstract copy of P as a generating set, and impose the bare minimum of +relations to ensure that the quotient of the free semigroup on P is a regular ∗-semigroup with +projection algebra P. This idea is akin to the constructions of Pastijn [115] and Easdown [31] of +the free (regular) idempotent-generated semigroup over an arbitrary (regular) biordered set E. +They both take a semigroup defined by a presentation with generating set (an abstract copy +of) E, and relations coming from the ‘basic products’, and the ‘sandwich sets’ in the regular +case. +Before we state the result, we briefly establish some notation. For a set X, we denote by X+ +the free semigroup over X, which consists of all non-empty words over X, under concatenation. +For a set R ⊆ X+ × X+ of pairs of words, we write R♯ for the congruence on X+ generated +by R. We say a semigroup S has presentation ⟨X : R⟩ if S ∼= X+/R♯, i.e. if there is a surjective +semigroup homomorphism X+ → S with kernel R♯. At times we identify ⟨X : R⟩ with the semi- +group X+/R♯ itself. The elements of X and R are called generators and relations, respectively. +A relation (u, v) ∈ R is typically displayed as an equation: u = v. +Theorem 7.14. For any projection algebra P, the chain semigroup CP has presentation +CP ∼= ⟨XP : RP ⟩, +where XP = {xp : p ∈ P} is an alphabet in one-one correspondence with P, and where RP is the +set of relations +x2 +p = xp +for all p ∈ P, +(R1) +(xpxq)2 = xpxq +for all p, q ∈ P, +(R2) +xpxqxp = xqθp +for all p, q ∈ P. +(R3) +To prove the theorem, we require some technical lemmas. But first, it is worth observing that +relations (R1)–(R3) closely resemble projection algebra axioms (P2), (P4) and (P5), i.e. those +that are stated purely in terms of the θ maps. +For the rest of this section, we fix the projection algebra P, and also the alphabet XP and +relations RP from Theorem 7.14. We also write ∼ = R♯ +P for the congruence on X+ +P generated +by relations (R1)–(R3). For words u, v ∈ X+ +P , we write u ∼1 v to indicate that u and v differ by +one or more applications of (R1), and similarly for ∼2 and ∼3. +76 + +Lemma 7.15. If p1, . . . , pk, q ∈ P are such that p1 F · · · F pk and q ≤ pk, then +xp1 · · · xpk−1xq ∼ xq1 · · · xqk−1xq +for some q1, . . . , qk−1 ∈ P with q1 F · · · F qk−1 F q. +Proof. The proof is by induction on k. When k = 1 the result is vacuously true, so we as- +sume k ≥ 2. Then with qk−1 = qθpk−1, we have +xpk−1xq ∼2 xpk−1xqxpk−1xq ∼3 xqθpk−1xq = xqk−1xq. +By induction, noting that qk−1 ≤ pk−1, we have +xp1 · · · xpk−2xqk−1 ∼ xq1 · · · xqk−2xqk−1 +for some q1, . . . , qk−2 ∈ P with q1 F · · · F qk−2 F qk−1. It follows that +xp1 · · · xpk−2xpk−1xq ∼ xp1 · · · xpk−2xqk−1xq ∼ xq1 · · · xqk−2xqk−1xq, +and it remains only to show that qk−1 F q. +But from q ≤ pk F pk−1 it follows from +Lemma 3.15(ii) that q F qθpk−1 = qk−1. +Lemma 7.16. For any w ∈ X+ +P of length k, we have +w ∼ xp1 · · · xpk +for some p1, . . . , pk ∈ P with p1 F · · · F pk. +Proof. We prove the lemma by induction on k. The k = 1 case being trivial, we assume k ≥ 2, +and we write w = xq1 · · · xqk. By induction we have +xq1 · · · xqk−1 ∼ xr1 · · · xrk−1 +for some r1, . . . , rk−1 ∈ P with r1 F · · · F rk−1. Next we define +pk−1 = qkθrk−1 +and +pk = rk−1θqk, +noting that pk−1 F pk, by Lemma 3.14. We then calculate +xrk−1xqk ∼2 xrk−1xqkxrk−1xqkxrk−1xqk ∼3 xqkθrk−1xrk−1θqk = xpk−1xpk. +Since pk−1 ≤ rk−1, it follows from Lemma 7.15 that +xr1 · · · xrk−2xpk−1 ∼ xp1 · · · xpk−2xpk−1 +for some p1, . . . , pk−2 ∈ P with p1 F · · · F pk−2 F pk−1. Putting everything together, we have +w = xq1 · · · xqk−2xqk−1xqk ∼ xr1 · · · xrk−2xrk−1xqk ∼ xr1 · · · xrk−2xpk−1xpk ∼ xp1 · · · xpk−2xpk−1xpk, +and the proof is complete. +Given a P-path p = (p1, . . . , pk) ∈ P = P(P), we define the word +wp = xp1 · · · xpk ∈ X+ +P . +So Lemma 7.16 says that every word over XP is ∼-equivalent to some wp. Using (R1), it is easy +to see that +wpwq ∼ wp◦q +for any p, q ∈ P with r(p) = d(q). +(7.17) +(In fact, we have wpwq = wp⊕q when r(p) F d(q), where ⊕ again denotes the concatenation +operation.) The next result refers to the congruence ≈ on P from Definition 3.36. +77 + +Lemma 7.18. For any p, q ∈ P, we have p ≈ q ⇒ wp ∼ wq. +Proof. It suffices to assume that p and q differ by a single application of (Ω1)–(Ω3), i.e. that +p = p′ ◦ s ◦ p′′ +and +q = p′ ◦ t ◦ p′′ +for some p′, p′′ ∈ P and (s, t) ∈ Ω ∪ Ω−1. +Since wp ∼ wp′wswp′′ and wq ∼ wp′wtwp′′, by (7.17), it is in fact enough to prove that +ws ∼ wt +for all (s, t) ∈ Ω. +We consider the three forms the pair (s, t) ∈ Ω can take. +(Ω1). This follows immediately from (R1). +(Ω2). If s = (p, q, p) and t = (p) for some (p, q) ∈ F, then ws = xpxqxp ∼3 xqθp = xp = wt. +(Ω3). Finally, suppose s = λ(e, p, f) = (e, eθp, f) and t = ρ(e, p, f) = (e, fθp, f) for some p ∈ P, +and some p-linked pair (e, f). Then +ws = xexeθpxf ∼3 xexpxexpxf ∼2 xexpxf ∼2 xexpxfxpxf ∼3 xexfθpxf = wt. +We can now tie together the loose ends. +Proof of Theorem 7.14. Define the homomorphism +Ψ : X+ +P → CP +by +xpΨ = p ≡ [p] +for p ∈ P. +To see that Ψ is surjective, let c ∈ CP , so that c = [p] for some p = (p1, . . . , pk) ∈ P. Then as +in the proof of Proposition 4.39(iii), and remembering that CP = S(C , idC ), we have +c = idC (c) = p1 • · · · • pk = (xp1 · · · xpk)Ψ = wpΨ. +(7.19) +Next, we note that Ψ preserves the relations RP , meaning that +(u, v) ∈ RP +⇒ +uΨ = vΨ (in CP ). +Indeed, this follows immediately from Lemma 2.23(ii) when (u, v) has type (R1) or (R2), and +from Lemma 4.38(iii) for type (R3). It follows from this that R♯ +P ⊆ ker(Ψ). +It remains to show that ker(Ψ) ⊆ R♯ +P . To do so, fix some (u, v) ∈ ker(Ψ), so that u, v ∈ X+ +P +and uΨ = vΨ; we must show that u ∼ v. By Lemma 7.16, we have u ∼ wp and v ∼ wq for +some p, q ∈ P. Using (7.19), and remembering that ∼ ⊆ ker(Ψ), we have +[p] = wpΨ = uΨ = vΨ = wqΨ = [q], +meaning that p ≈ q. But then wp ∼ wq by Lemma 7.18, so u ∼ wp ∼ wq ∼ v, as required. +In what follows, we will typically denote the R♯ +P -class of a word w ∈ X+ +P by [w]. In this way, +we have +⟨XP : RP ⟩ = X+ +P /R♯ +P = {[w] : w ∈ X+ +P }. +The product in ⟨XP : RP ⟩ is given by [u][v] = [uv] for u, v ∈ X+ +P , and the involution by +[xp1 · · · xpk]∗ = [xpk · · · xp1] for p1, . . . , pk ∈ P. The projections of ⟨XP : RP ⟩ are the R♯ +P -classes +[xp] ≡ p (p ∈ P), so the projection algebra of ⟨XP : RP ⟩ is (isomorphic to) P. +78 + +7.3 +Chain semigroups as free objects +In Remark 7.12 we discussed a certain ‘free-ness’ property of the chain semigroup CP in the +category of idempotent-generated regular ∗-semigroups with (fixed) projection algebra P. In +category theory, ‘free’ objects are more formally defined as the objects in the image of a ‘left +adjoint to a forgetful functor’. We will soon recall the meaning of these terms, and show that +our chain semigroups fit this definition, with respect to an appropriate forgetful functor. +We begin by proving the following result. +It will be used to verify the above-mentioned +categorical condition, but we hope that the reader can already sense a ‘flavour’ of free-ness from +the statement, and the universal property it describes. +Theorem 7.20. If P is a projection algebra and S is a regular ∗-semigroup, then for any projec- +tion algebra morphism φ : P → P(S), there is a unique ∗-semigroup homomorphism Φ : CP → S +such that the following diagram commutes (where both ‘ι’s denote inclusion maps): +P +P(S) +CP +S. +φ +ι +ι +Φ +Proof. This could be proved by constructing a suitable chained projection functor (C , idC ) → G(S), +where C = C (P) is the chain groupoid, and then applying Lemma 6.3. Alternatively, we can +use Theorem 7.14, and prove the result with ⟨XP : RP ⟩ = X+ +P /R♯ +P in place of CP . Taking this +second route, we begin by defining a semigroup homomorphism +ϕ : X+ +P → S +by +xpϕ = pφ +for p ∈ P. +Next we check that RP ⊆ ker(ϕ), meaning that uϕ = vϕ for all (u, v) ∈ RP . Indeed, this is +essentially trivial (modulo Lemma 2.23(ii)) when (u, v) has type (R1) or (R2). For type (R3), +we must show that (xpxqxp)ϕ = xqθpϕ for any p, q ∈ P. But for any such p, q, and using θ′ to +denote the unary operations in P(S), as in (5.1), we have +(xpxqxp)ϕ = (pφ)(qφ)(pφ) = (qφ)θ′ +pφ = (qθp)φ = xqθpϕ. +(In the penultimate step we used the fact that φ is a projection algebra morphism.) +It follows from RP ⊆ ker(ϕ) that R♯ +P ⊆ ker(ϕ). We therefore have a well-defined semigroup +homomorphism +Φ : ⟨XP : RP ⟩ → S +given by +[w]Φ = wϕ +for w ∈ X+ +P . +This is a ∗-homomorphism, as for any word w = xp1 · · · xpk ∈ X+ +P , +[w]∗Φ = [xpk · · · xp1]Φ = (pkφ) · · · (p1φ) = (pkφ)∗ · · · (p1φ)∗ = ((p1φ) · · · (pkφ))∗ = ([w]Φ)∗. +Commutativity of the diagram amounts to the fact that Φ|P = φ (where as above we identify +p ≡ [xp] for p ∈ P). This also implies uniqueness, as CP is generated by P ≡ {[xp] : p ∈ P}. +Consider again the two (large) categories: +• RSS, of regular ∗-semigroups with ∗-semigroup homomorphisms, and +• PA, of projection algebras with projection algebra morphisms. +79 + +The process of taking the projection algebra P(S) of a regular ∗-semigroup S can be thought of +as a ‘forgetful functor’; we forget the non-projections of S, and remember only the ‘conjugation’ +actions of projections on each other. Guided by standard letter usage (see for example [96, p. 79]), +we denote this functor by +U : RSS → PA. +So the action of U on an object S ∈ vRSS and morphism φ ∈ RSS(S, S′) is given by: +• U(S) = P(S) is the projection algebra of S, with unary operations as in (5.1), and +• U(φ) : U(S) → U(S′) is the (set-theoretic) restriction of φ to U(S), which is easily seen to be +a projection algebra morphism. +In the other direction, the process of constructing the chain semigroup CP of a projection alge- +bra P is a functor +F : PA → RSS. +The action of F on an object P ∈ vPA and morphism φ ∈ PA(P, P ′) is given by: +• F(P) = CP is the chain semigroup of P, as in Definition 3.39, and +• F(φ) : F(P) → F(P ′) is the induced ∗-semigroup homomorphism Φ : CP → CP ′ from +Proposition 7.13. +So we have the two functors +U : RSS → PA +and +F : PA → RSS. +We have already observed that U(F(P)) = P(CP ) = P for any projection algebra P. It is +also clear that U(F(φ)) = φ for any morphism φ in PA (this is precisely what it means for Φ to +extend φ in Proposition 7.13). In other words, UF = idPA is the identity functor PA → PA. The +other composition FU : RSS → RSS is not the identity. However, we will show in Theorem 7.23 +that F is a left adjoint to U. +To make the previous statement precise, we need to recall some definitions. There are many +equivalent definitions of adjunctions, but for our purposes the most convenient is that of [5, +Definition 9.1]. +Definition 7.21. Consider two (possibly large) categories C and D, and a pair of functors +F, G : C → D. A natural transformation η : F → G is a family η = (ηC)C∈vC, where each +ηC : F(C) → G(C) is a morphism in D, and such that the following condition holds: +• For every pair of objects C, C′ ∈ vC, and for every morphism φ : C → C′ in C, the following +diagram commutes: +F(C) +G(C) +F(C′) +G(C′). +ηC +F(φ) +G(φ) +ηC′ +80 + +Definition 7.22. Consider two (possibly large) categories C and D. An adjunction C → D is a +triple (F, U, η), where F : C → D and U : D → C are functors, and η is a natural transformation +idC → UF, such that the following condition holds: +• For every pair of objects C ∈ vC and D ∈ vD, and for every morphism φ : C → U(D), there +exists a unique morphism φ : F(C) → D such that the following diagram commutes: +C +U(F(C)) +U(D). +φ +ηC +U(φ) +In this set-up, F and U are called the left and right adjoints, respectively, and η is the unit of +the adjunction. The C-free objects in D are the objects in the image of F, i.e. those of the form +F(C) for C ∈ vC. +Theorem 7.23. The functor +F : PA → RSS : P �→ CP +is a left adjoint to the forgetful functor +U : RSS → PA : S �→ P(S). +Consequently, the U-free objects in the category RSS are precisely the chain semigroups. +Proof. Consulting Definition 7.22, we need a natural transformation η : idPA → UF. Since we +have already observed that UF = idPA, we can take η = (idP )P∈vPA. For P, P ′ ∈ vPA and +φ : P → P ′, the diagram in Definition 7.21 becomes +P +P +P ′ +P ′, +idP +φ +φ +idP ′ +and this obviously commutes. +To verify that (F, U, η) is an adjunction PA → RSS, we need to show that: +• For every projection algebra P ∈ vPA and regular ∗-semigroup S ∈ vRSS, and for every +morphism φ : P → P(S), there exists a unique ∗-semigroup homomorphism φ : CP → S such +that the following diagram commutes: +P +P +P(S). +φ +idP +φ|P +For such P, S and φ, we take φ = Φ : CP → S, as in Theorem 7.20, and the required properties +(uniqueness and commutativity) follow from the theorem. +81 + +It follows from Theorem 7.23 that we may rightly speak of CP as ‘the free regular ∗-semigroup +with projection algebra P’. +There is a vast literature on free idempotent-generated semigroups FIG(E) over biordered +sets E; see for example [11, 17, 19, 23–25, 27–32, 56–58, 110, 112, 140]. A long-standing folklore +conjecture was that the maximal subgroups of any FIG(E) are all free groups. The first coun- +terexample was constructed in [11], and then the conjecture was proven to be maximally false +in [57], where it was shown that every group appears as the maximal subgroup of some free +idempotent-generated semigroup; this result has been reproved a number of times [25, 56, 140]. +Since then, a number of important studies have computed maximal subgroups arising from +biordered sets of natural families of semigroups, such as transformation monoids [58] and linear +monoids [23]. We believe it would be very interesting to study the free (idempotent/projection- +generated) regular ∗-semigroups CP along similar lines. Natural questions include the following: +Problem 7.24. +(i) Are maximal subgroups of CP always free? +(ii) Can any group appear as the maximal subgroup of some CP ? +(iii) What are the maximal subgroups of CP , when P is the projection algebra of some natural +family of regular ∗-semigroups (e.g. partition, Brauer or Temperley-Lieb monoids)? +(iv) Given a regular ∗-semigroup S, with projection algebra P = P(S) and biordered set +E = E(S), how are the free semigroups CP and FIG(E) related? Since E is a regular +biordered set, the same question can be asked of the free regular idempotent-generated +semigroup FRIG(E); cf. [115]. +Some of these questions will be explored in [37]. We will say a little about the last question in +Example 7.25 below. One could also study structural properties of CP , or consider decision prob- +lems, as in [17,19,24]. The above papers typically study free idempotent-generated semigroups +via Easdown’s presentation [31]. To study CP , one has the option of using its presentation from +Theorem 7.14 or its direct definition as the chain semigroup of P. +Example 7.25. Let P be an arbitrary set, and consider again the square band over P, i.e. the +regular ∗-semigroup S = P × P, with operations +(p, q)(r, s) = (p, s) +and +(p, q)∗ = (q, p). +The projections of S are of the form (p, p), for p ∈ P. +Identifying (p, p) ≡ p, we see that +P(S) ≡ P. For each p ∈ P, the operation θp is the constant map with image {p}, and F = ∇P +is the universal relation. +In the special case that P = {p, q} has size 2, the chain semigroup +CP = {[p], [q], [p, q], [q, p]} +has size 4, and of course CP ∼= S. This can also be seen by writing down the presentation from +Theorem 7.14, which simplifies (writing xp ≡ p and xq ≡ q) to +⟨p, q : p2 = p, q2 = q, pqp = p, qpq = q⟩. +It is clear that {p, q, pq, qp} is a set of normal forms, i.e. a set of representatives of equivalence- +classes of words. +On the other hand, the biordered set E = E(S) of S is simply E = S, and it is easy to +see that FIG(E) is infinite; indeed, simple computations in GAP [52, 107] show that maximal +subgroups of FIG(E) are infinite cyclic. This all shows that the answer to the fourth question +in Problem 7.24 will not simply be that CP and FIG(E) are always isomorphic. +82 + +8 +Fundamental regular ∗-semigroups +In this chapter we turn our attention to fundamental regular ∗-semigroups. Many of the ideas +discussed here exist in the literature in a different form [70, 72, 73, 138], but we feel that our +groupoid approach leads to some clearer constructions and proofs. +We begin in Section 8.1 by obtaining a number of formulations for the maximum projection- +separating congruence µS on a regular ∗-semigroup S; see Proposition 8.3 and Corollary 8.4. +In Section 8.2 we construct the maximum fundamental regular ∗-semigroup MP with a given +projection algebra P; see Definition 8.19. Theorems 8.20 and 8.21 establish various universal +properties of the semigroup MP . Specifically, we use the ϑ maps on chained projection groupoids +to show that for any regular ∗-semigroup S with projection algebra P(S) = P, there is a canonical +∗-semigroup homomorphism φS : S → MP , and that im(φS) ∼= S/µS is the fundamental image +of S. In particular, if S happens to be fundamental itself, then φS is an embedding of S in MP . +At times we compare our construction of MP with others from the literature. In Section 8.3 we +consider idempotent-generated fundamental regular ∗-semigroups. In particular, Theorem 8.24 +shows that (up to isomorphism) there is exactly one such semigroup with a given projection +algebra, and gives a number of ways to construct it. +8.1 +The maximum idempotent-separating congruence +Recall that a congruence σ on a semigroup S is idempotent-separating if +e σ f +⇒ +e = f +for all e, f ∈ E(S). +A ∗-congruence on a regular ∗-semigroup S is a congruence σ that is also compatible with ∗: +a σ b +⇒ +a∗ σ b∗ +for all a, b ∈ S. +We say that such a σ is projection-separating if +p σ q +⇒ +p = q +for all p, q ∈ P(S). +Any such projection-separating ∗-congruence σ is idempotent-separating, as for e, f ∈ E(S), +e σ f ⇒ e∗ σ f∗ ⇒ ee∗ σ ff∗ ⇒ ee∗ = ff∗ ⇒ e R f +and similarly +e L f. +It follows that e σ f ⇒ e H f ⇒ e = f, since an H -class can contain at most one idempotent. +Definition 8.1. A regular ∗-semigroup S is fundamental if it has no non-trivial idempotent- +separating (equivalently, projection-separating) ∗-congruences, i.e. if the only idempotent-separating +∗-congruence is the trivial relation ∆S = {(a, a) : a ∈ S}. +Given a regular ∗-semigroup S, it was shown in [70, Theorem 4] and [138, Theorem 2.4] that +the relation +µS = {(a, b) ∈ S × S : a∗pa = b∗pb and apa∗ = bpb∗ for all p ∈ P} +(8.2) +is the maximum projection-separating ∗-congruence on S, and also the maximum idempotent- +separating (ordinary semigroup) congruence on S. It follows that a regular ∗-semigroup is fun- +damental as a ∗-semigroup (as in Definition 8.1) if and only if it is fundamental as a semigroup +(i.e. has no non-trivial idempotent-separating (semigroup) congruences). It also follows that S/µS +is the unique fundamental quotient of S with the same projection algebra (up to isomorphism); +we call S/µS the fundamental image of S. +To keep the paper self-contained, it is worth briefly sketching why (8.2) holds, i.e. why the +stated relation is the maximum projection-separating congruence: +83 + +• First, it is easy to check that µS is a ∗-congruence. +• It is projection-separating because for p, q ∈ P, +(p, q) ∈ µS ⇒ p = p∗pp = q∗pq = qpq ⇒ p ≤ q +and similarly +q ≤ p, +which shows that (p, q) ∈ µS ⇒ p = q. +• It is the maximum such, because if σ is an arbitrary projection-separating ∗-congruence, then +for any a, b ∈ S and p ∈ P, +a σ b ⇒ a∗ σ b∗ ⇒ a∗pa σ b∗pb ⇒ a∗pa = b∗pb +and similarly +apa∗ = bpb∗, +which shows that (a, b) ∈ σ ⇒ (a, b) ∈ µS, i.e. that σ ⊆ µS. +It turns out that the congruence µS has a useful alternative description, in terms of our ϑ and Θ +maps: +Proposition 8.3. If S is a regular ∗-semigroup, then the maximum idempotent-separating con- +gruence of S is the relation +µS = {(a, b) ∈ S × S : ϑa = ϑb} = {(a, b) ∈ S × S : Θa = Θb and Θa∗ = Θb∗} += {(a, b) ∈ S × S : d(a) = d(b) and Θa = Θb}. +Proof. It follows immediately from (5.11) that Θa = Θb +⇔ +a∗pa = b∗pb for all p ∈ P. +Combining this with the definition of µS in (8.2), we have +µS = {(a, b) ∈ S × S : Θa = Θb and Θa∗ = Θb∗}. +It therefore remains to show that the following are equivalent, for all a, b ∈ S: +(i) ϑa = ϑb +(ii) Θa = Θb and Θa∗ = Θb∗, +(iii) d(a) = d(b) and Θa = Θb. +Equivalence of (i) and (iii) follows quickly from the fact that ϑa = Θa|d(a)↓ and Θa = θd(a)ϑa. +(i) ⇒ (ii). Suppose ϑa = ϑb. Considering the domains of these mappings (see (2.10)), it follows +that d(a) = d(b), and then +Θa = θd(a)ϑa = θd(b)ϑb = Θb. +By Lemma 2.12, we also have ϑa∗ = ϑ−1 +a += ϑ−1 +b += ϑb∗. But then the previous argument also +gives Θa∗ = Θb∗. +(ii) ⇒ (iii). +Now suppose Θa = Θb and Θa∗ = Θb∗. +Considering the range of Θa∗ = Θb∗ +(see (4.1)), it follows that d(a) = d(b). +This also leads to a somewhat simpler equational characterisation of µS, as compared to the +original in (8.2). For another equivalent formulation, see for example [114, Corollary 4.6]. +Corollary 8.4. If S is a regular ∗-semigroup, then the maximum idempotent-separating congru- +ence of S is the relation +µS = {(a, b) ∈ S × S : aa∗ = bb∗, a∗pa = b∗pb for all p ≤ aa∗} += {(a, b) ∈ S × S : a∗a = b∗b, apa∗ = bpb∗ for all p ≤ a∗a}. +84 + +Proof. By Proposition 8.3 we have (a, b) ∈ µS ⇔ ϑa = ϑb. This is of course equivalent to the +following two conditions: +(i) dom(ϑa) = dom(ϑb), and +(ii) pϑa = pϑb for all p ∈ dom(ϑa). +Since dom(ϑa) = (aa∗)↓ and dom(ϑb) = (bb∗)↓, (i) is equivalent to aa∗ = bb∗. Combining this +with (5.10), it follows that (ii) is equivalent to a∗pa = b∗pb for all p ≤ aa∗. +This gives the first claimed expression for µS. The second follows from the first because µS +is a ∗-congruence, or from the fact that ϑa = ϑb ⇔ ϑa∗ = ϑb∗ (cf. Lemma 2.12). +8.2 +Maximum fundamental regular ∗-semigroups +Coming from the other direction, it has also long been known that for any projection algebra P, +there is a unique ‘maximum’ fundamental regular ∗-semigroup with projection algebra P (up to +isomorphism). This maximum semigroup has been constructed in a number of ways; the papers +[70,72,73,113,138] start from projection algebras (or similar structures), while [113] also contains +an alternative approach using certain special biordered sets. As an application of our groupoid +approach, in this section we provide an alternative construction. Although we of course arrive at +the same semigroup (up to isomorphism), we believe our approach is a little more transparent. +See Definition 8.19 and Theorems 8.20 and 8.21, and also Remark 8.22. +For the rest of this section we fix a projection algebra P, as in Definition 3.1, and all of +the data that comes with it, i.e. the unary operations θp, the relations ≤, ≤F and F, and so +on. We also write P = P(P) and C = C (P) for the path category and chain groupoid of P; +cf. Definitions 3.17 and 3.39. +Recall that for p ∈ P, we have the down-set +p↓ = {q ∈ P : q ≤ p}. +If s, t ∈ p↓, then by Lemma 3.9 we have sθt = sθtθp ≤ p. This shows that p↓ is closed under each +operation θt (t ∈ p↓), and is hence a projection algebra in its own right. (This is not to say it is +a subalgebra of P, as it might not be closed under some θq for q ∈ P \ p↓. On the other hand, +any down-set p↓ is a ⋄-subalgebra of P; cf. Remark 3.2.) +Definition 8.5. For p, q ∈ P, a P-isomorphism p → q is a projection algebra isomorphism p↓ → q↓, +i.e. a bijection α : p↓ → q↓ satisfying +(sθt)α = (sα)θtα +for all s, t ∈ p↓. +(8.6) +We write M(p, q) for the set of all such P-isomorphisms p → q, and we set +M = M(P) = +� +p,q∈P +M(p, q). +It is easy to see that M is a groupoid, under ordinary function composition and inversion, and +with objects/identities vM = {idp↓ : p ∈ P}. As usual we identify vM ≡ P, viz. p ≡ idp↓. We +call M the Munn groupoid of P, in honour of the early work of Douglas Munn on fundamental +inverse semigroups [109]. +Munn’s semigroups were constructed using order-isomorphisms of +principal ideals/down-sets of semilattices, and this approach has been highly influential in studies +of inverse semigroups and their various generalisations. +We begin with the following simple lemma. +85 + +Lemma 8.7. Every α ∈ M is order-preserving, in the sense that +s ≤ t +⇒ +sα ≤ tα +for all s, t ≤ d(α). +Proof. We have s ≤ t ⇒ s = sθt ⇒ sα = (sθt)α = (sα)θtα ⇒ sα ≤ tα. +For α ∈ M, and for p ≤ d(α), we have p↓ ⊆ d(α)↓ = dom(α), so we can define the restriction +p⇃α = α|p↓. +Here as usual we write f|A for the restriction of a function f to a subset A of its domain. +Lemma 8.8. If α ∈ M, and if p ≤ d(α), then p⇃α ∈ M(p, pα). Consequently, p↓α = (pα)↓. +Proof. Of course it suffices to prove the first claim. By definition we have p⇃α = α|p↓. For any +t ∈ p↓, we have t ≤ p, and so tα ≤ pα by Lemma 8.7, and this says that tα ∈ (pα)↓. Thus, p⇃α +maps p↓ into (pα)↓. Since α is injective and satisfies (8.6), so too does p⇃α. Finally, p⇃α is +surjective onto (pα)↓, since for any t ≤ pα we have t = (tα−1)α, with tα−1 ≤ (pα)α−1 = p since +α−1 ∈ M is order-preserving by Lemma 8.7. +Lemma 8.9. For any projection algebra P, M = M(P) is an ordered groupoid. +Proof. Again the proof is by an application of Lemma 2.3. We have the usual order ≤ on +vM = P (cf. (3.5)), and we have already defined the restrictions p⇃α = α|p↓. It is then routine +to check that properties (O1)′–(O5)′ all hold. +As usual, the right-handed restrictions are defined by +α⇂q = (⇃qα−1)−1 = (α−1|q↓)−1 +for α ∈ M and q ≤ r(α). +Next we wish to verify that M = M(P) is a projection groupoid. To do so, we need to understand +the ϑ and Θ maps. It follows from Lemma 8.8 that for α ∈ M, the map +ϑα : d(α)↓ → r(α)↓ +is given by +pϑα = r(p⇃α) = pα +for all p ≤ d(α). +In other words, we have ϑα = α (!) for all α ∈ M, so also Θα = θd(α)ϑα = θd(α)α. To summarise: +ϑα = α +and +Θα = θd(α)α +for all �� ∈ M. +(8.10) +Lemma 8.11. For any projection algebra P, M = M(P) is a projection groupoid. +Proof. By Lemma 8.9 we just need to verify (G1). +For this, it is essentially trivial to see +that (G1d) holds, as ϑα = α is a projection algebra morphism by definition, for all α ∈ M. +Consider a pair p, q ∈ P with p F q. Since the operation θq maps into q↓, so too does the +restriction +γpq = θq|p↓. +Lemma 8.12. For any p, q ∈ P with p F q, we have γpq ∈ M(p, q), and γ−1 +pq = γqp. +Proof. This can be proved directly, but it also follows by combining results of earlier chapters. +Recall from Proposition 7.4 that (C , idC ) is a chained projection groupoid, where C = C (P) is +the chain groupoid of P. Applying Lemma 4.14(ii) to this groupoid, and remembering that the +evaluation map is ε = idC , we have +ϑ[p,q] = ϑidC [p,q] = θq|p↓ = γpq. +It follows from property (G1d) that γpq = ϑ[p,q] is a projection algebra isomorphism p↓ → q↓, +i.e. that γpq ∈ M(p, q). Combining the above with Lemma 2.12, it follows that +γqp = ϑ[q,p] = ϑ[p,q]−1 = ϑ−1 +[p,q] = γ−1 +pq . +86 + +To show that M is a chained projection groupoid, we need to define an evaluation map +ε = ε(P) : C → M. As usual, it is convenient to first define a functor +π = π(P) : P → M : (p1, . . . , pk) �→ γp1p2 · · · γpk−1pk. +By convention, when k = 1, we interpret this last expression simply as idp↓ +1 ≡ p1. This means +that π(p) = idp↓ ≡ p for all p ∈ P, i.e. that π is a v-functor. The next result concerns the +congruence ≈ = Ω♯ on P from Definition 3.36. +Lemma 8.13. We have ≈ ⊆ ker(π). +Proof. We need to check that π(s) = π(t) for each pair (s, t) ∈ Ω. As usual this is clear when +the pair has one of the forms (Ω1) or (Ω2). So suppose instead that +s = λ(e, p, f) = (e, eθp, f) +and +t = ρ(e, p, f) = (e, fθp, f) +for some projection p ∈ P, and some p-linked pair (e, f), as in Definition 3.31. Then +π(s) = γe,eθpγeθp,f +and +π(t) = γe,fθpγfθp,f. +Both π(s) and π(t) map e↓ → f↓, and we must show that these maps are equal, i.e. that +qπ(s) = qπ(t) for any q ≤ e. By definition of the γuv, and keeping q = qθe in mind (as q ≤ e), +we calculate +qπ(s) = qθeθpθf =4 (qθe)θpθeθpθf =5 qθeθpθf = qθpθf +We also have qπ(t) = qθfθpθf =4 qθpθfθpθf =5 qθpθf, so the proof is complete. +Definition 8.14. Since C = P/≈, it follows from Lemma 8.13 that we have a well-defined +functor +ε = ε(P) : C → M +given by +ε[p] = π(p) +for p ∈ P. +That is, ε[p1, . . . , pk] = γp1p2 · · · γpk−1pk whenever p1 F · · · F pk (and ε[p] = p ≡ idp↓ for p ∈ P). +The proof of the next result makes use of the basic fact that for bijections f : A → B and +g : B → C, and for X ⊆ A, we have +(f ◦ g)|X = f|X ◦ g|Y +where Y = Xf. +(8.15) +(This is essentially (O5)′ in the category of bijections.) +Lemma 8.16. The functor ε = ε(P) : C → M is an evaluation map. +Proof. As with Lemma 5.17, the proof boils down to showing that +π(q⇃p) = q⇃π(p) +for all p ∈ P and q ≤ d(p). +We prove this by induction on k, the length of the path p = (p1, . . . , pk). When k = 1, both sides +evaluate to q ≡ idq↓. We now assume that k ≥ 2, and we write q⇃p = (q1, . . . , qk) as in (3.19). +Also write p′ = (p1, . . . , pk−1), noting that q⇃p′ = (q1, . . . , qk−1). By induction we have +π(q⇃p) = π(q1, . . . , qk) = γq1q2 · · · γqk−2qk−1γqk−1qk = π(q⇃p′) ◦ γqk−1qk = q⇃π(p′) ◦ γqk−1qk. +On the other hand, using (8.15), and writing Y = q↓γp1p2 · · · γpk−2pk−1, we have +q⇃π(p) = (γp1p2 · · · γpk−1pk)|q↓ = (γp1p2 · · · γpk−2pk−1)|q↓ ◦ γpk−1pk|Y = q⇃π(p′) ◦ γpk−1pk|Y . +87 + +Examining the last two conclusions, it remains to show that γpk−1pk|Y = γqk−1qk. By Lemma 8.8 +and (3.19), we have +Y = q↓(γp1p2 · · · γpk−2pk−1) = (qγp1p2 · · · γpk−2pk−1)↓ = (qθp2 · · · θpk−1)↓ = q↓ +k−1, +and so +γpk−1pk|Y = (θpk|p↓ +k−1)|q↓ +k−1 = θpk|q↓ +k−1. +Since γqk−1qk = θqk|q↓ +k−1, it remains to show that tθpk = tθqk for all t ∈ q↓ +k−1. For this we use +t = tθqk−1 (as t ≤ qk−1), qk = qk−1θpk (by (3.19)), and the projection algebra axioms to calculate +tθqk = (tθqk−1)θqk−1θpk =4 tθqk−1θpkθqk−1θpk =5 tθqk−1θpk = tθpk. +Note in particular that ε[p, q] = γpq for p, q ∈ P with p F q. +Proposition 8.17. If P is a projection algebra, then (M, ε) is a chained projection groupoid. +Proof. By Lemmas 8.11 and 8.16, it remains to verify (G2). To do so, let (e, f) be a β-linked +pair, where β ∈ M(q, r). Let the ei, fi be as in (4.19). We must show that λ = ρ, where +λ = λ(e, β, f) = γee1 ◦ e1⇃β ◦ γf1f +and +ρ = ρ(e, β, f) = γee2 ◦ e2⇃β ◦ γf2f. +Keeping γuv = θv|u↓ in mind, this is equivalent to showing that +tθe1βθf = tθe2βθf +for all t ≤ e. +(8.18) +Starting with the left-hand side, we use t = tθe (as t ≤ e), e1 = eθq (by (4.19)), and the projection +algebra axioms to calculate +tθe1βθf = tθeθeθqβθf =4 tθeθqθeθqβθf =5 tθeθqβθf = tθqβθf. +On the other hand we have +tθe2βθf = tθfΘβ−1βθf +by (4.19) += tΘβθfΘβ−1βθf +by (G1b) += tθqβθfθrβ−1βθf +by (8.10) += tθqβθfθrθf += tθqβθf +by Lemma 3.12(iii), as f ≤F r (cf. Lemma 4.18). +Examining the previous two conclusions, we have completed the proof of (8.18), and hence of +the proposition. +Now that we know (M, ε) is a chained projection groupoid for any projection algebra P, we +can apply the functor S to obtain a regular ∗-semigroup (cf. Theorems 4.36 and 6.7): +Definition 8.19. For a projection algebra P, we define the regular ∗-semigroup +MP = S(M, ε), +where M = M(P) and ε = ε(P) are as in Definitions 8.5 and 8.14. Explicitly: +• The elements of MP are the P-isomorphisms p↓ → q↓ (p, q ∈ P), as in Definition 8.5. +• The product • in MP is given, for α, β ∈ MP with r(α) = p and d(β) = q, by +α • β = α⇂p′ ◦ γp′q′ ◦ q′⇃β = (αθq′β)|(p′α−1)↓ +where +p′ = qθp +and +q′ = pθq. +88 + +• The involution in MP is given by ordinary inversion of bijections, α∗ = α−1. +• The projections of MP are the identity maps idp↓ ≡ p, for p ∈ P, and consequently +P(MP ) ≡ P. +• The idempotents of MP are the maps p • q = γpq, for (p, q) ∈ F. +We call MP the Munn semigroup of P. +An ostensibly weaker version of the next result was proved in [70, Theorem 5], albeit with a +different formulation of MP ; cf. Remark 8.25. The statement, and the following one, refers to +the canonical homomorphisms of Definition 7.9. +Theorem 8.20. If S is a regular ∗-semigroup with projection algebra P(S) = P, then +(i) there is a canonical ∗-semigroup homomorphism +φS : S → MP +given by +aφS = ϑa +for all a ∈ S, +(ii) ker(φS) = µS is the maximum idempotent-separating congruence of S, and consequently S +is fundamental if and only if φS is injective, +(iii) im(φS) ∼= S/µS is the fundamental image of S. +Proof. Throughout the proof we write M = M(P) and ε = ε(P), as in Definitions 8.5 and 8.14. +We also write G(S) = (G, ε′). +(i). By (G1d), we have ϑa ∈ M(p, q) for any a ∈ G(p, q). It follows that we have a well-defined +mapping +φ : G → M : a �→ ϑa. +Our first goal is to show that φ is a chained projection functor (G, ε′) → (M, ε); cf. Definition 6.1. +It follows quickly from Lemma 2.14 that φ is an ordered v-functor. Thus, (F1) holds trivially. It +also follows from vφ = idP that the induced map Φ : C → C at the top of the diagram in (F2) +is also an identity map. Thus, the diagram becomes: +C +C +G +M, +idC +ε′ +ε +φ +and we must show that this commutes, i.e. that +(ε′(c))φ = ε(c) +for all c ∈ C . +This is clear if c = [p] ≡ p for some p ∈ P, so suppose instead that c = [p1, . . . , pk] with k ≥ 2. +Then using the definitions, and the fact that each of the maps in the diagram is a functor, we +calculate +(ε′(c))φ = (ε′[p1, p2] ◦ · · · ◦ ε′[pk−1, pk])φ += (ε′[p1, p2])φ ◦ · · · ◦ (ε′[pk−1, pk])φ += ϑε′[p1,p2] ◦ · · · ◦ ϑε′[pk−1,pk] += θp2|p↓ +1 ◦ · · · ◦ θpk|p↓ +k−1 +by Lemma 4.14(ii) += γp1p2 ◦ · · · ◦ γpk−1pk += ε[p1, p2] ◦ · · · ◦ ε[pk−1, pk] = ε(c), +89 + +as required. +We now know that φ is a chained projection functor G(S) = (G, ε′) → (M, ε). By Lemma 6.3 +(and Theorem 6.7), φ is also a ∗-semigroup homomorphism S = S(G(S)) → S(M, ε) = MP , and +of course φ = φS is the map in the statement of the theorem. It is canonical because φ is a +v-functor. +(ii). This follows immediately from Proposition 8.3 and the definition of φS. +(iii). This follows from the fundamental homomorphism theorem. +We can now quickly deduce the next result, which is essentially [72, Theorem 3.2]. It says that +not only is MP fundamental, but it is in fact the maximum fundamental regular ∗-semigroup +with projection algebra P. +Theorem 8.21. If P is a projection algebra, then +(i) MP is a fundamental regular ∗-semigroup with projection algebra P, +(ii) any fundamental regular ∗-semigroup with projection algebra P embeds canonically in MP . +Proof. (i). By (8.10), the homomorphism φMP : MP → MP : α �→ ϑα from Theorem 8.20(i) is +simply the identity map, and is therefore of course injective. Consequently, MP is fundamental +by Theorem 8.20(ii). We have already observed that P(MP ) is (isomorphic to) P. +(ii). This follows immediately from Theorem 8.20(ii). +Remark 8.22. The above construction of MP using P-isomorphisms is closest in spirit to the +work of Yamada [138] and Jones [73], although Yamada’s basic set-up was quite different; in +place of projection algebras, he used ‘P-sets in fundamental regular warps’. +An alternative construction of MP was given by Imaoka [70], and we briefly comment here +on how to interpret this in our set-up. Let TP denote the full transformation semigroup over P, +i.e. the semigroup of all (totally-defined) functions P → P, under composition. For any regular +∗-semigroup S with P(S) = P, Lemma 4.29 guarantees the existence of an (ordinary semigroup) +homomorphism +S → TP : a �→ Θa. +(Recall that the • operation in S(G(S)) = S is simply the original product in S.) Since the +involution of S is an antihomomorphism, the map a �→ Θa∗ is an antihomomorphism S → TP , +but an (ordinary) homomorphism S → T op +P . The latter denotes the opposite semigroup to TP ; +the product ⋆ in T op +P +is given by α ⋆ β = βα. Thus, we have a homomorphism into the direct +product +ξS : S → TP × T op +P +given by +aξS = (Θa, Θa∗) +for a ∈ S. +By Proposition 8.3, ker(ξS) = µS is the maximum idempotent-separating congruence of S, and +im(ξS) ∼= S/µS is the fundamental image of S. In particular, when S = MP , the image of ξMP +is isomorphic to MP . Keeping (8.10) in mind, this copy of MP , which we will denote by MP , +is precisely the subsemigroup +MP = im(ξMP ) = {α : α ∈ MP } ≤ TP × T op +P , +where we write α = (θpα, θqα−1) for α ∈ M(p, q). The product in MP is simply composition +(and reverse composition) of transformations. While TP and T op +P +are not regular ∗-semigroups +for |P| ≥ 2 (as their D-classes are not square), of course MP is, and its involution is given +by α∗ = α−1. Returning to the general case, the image of ξS (where S is an arbitrary regular +∗-semigroup with P(S) = P) is contained in MP , and ξS is an embedding S → MP if and only +if S is fundamental. +90 + +Another, very different, way to construct MP can be found in [113, Section 3], wherein the +projection algebra P is characterised as a specialised poset with a connection to a certain dual +poset P ◦ (in the sense of Grillet [60,61]), and MP is realised as a subdirect product, similar to +Imaoka’s construction. +We believe the following problem is of considerable interest: +Problem 8.23. Given a projection algebra P, describe the maximum fundamental regular ∗- +semigroup MP . +In particular, one could attempt to do this in the case that P = P(S) arises from some +natural regular ∗-semigroup S, or for some family of such semigroups. For example, the case +that S is a finite diagram monoid (including partition monoids) is the subject of an ongoing work +by the current authors [36]. +8.3 +Idempotent-generated fundamental regular ∗-semigroups +Finally, we combine the ideas of this chapter and the previous one. The next result concerns +the chain semigroup CP , as in Definition 7.8, and the proof utilises the map φCP : CP → MP +from Theorem 8.20. The fundamental image of CP is the quotient CP /µ, where µ = µCP is the +maximum idempotent-separating congruence. This quotient is of course idempotent-generated, +as CP is. For the following statement, recall that E(S) denotes the idempotent-generated sub- +semigroup of the semigroup S. To the best of our knowledge, no general results exist in the +literature concerning idempotent-generated fundamental regular semigroups, though they were +used tangentially in the proof of [110, Theorem 7.2] in the case of so-called solid biordered sets. +Theorem 8.24. Let P be a projection algebra, and let µ = µCP be the maximum idempotent- +separating congruence on the chain semigroup CP . +(i) Up to isomorphism, there is exactly one idempotent-generated fundamental regular ∗-semigroup +with projection algebra P, namely CP /µ. +(ii) For any fundamental regular ∗-semigroup S with projection algebra P = P(S), we have +E(S) ∼= CP /µ. +Proof. We prove both parts together. Suppose S is an arbitrary fundamental regular ∗-semigroup +with projection algebra P = P(S). By Proposition 7.10 and Theorem 8.20, we have the two +∗-semigroup homomorphisms +CP +S +MP . +ε +φS +Here, ε is the evaluation map C → G(S), and φS is injective by Theorem 8.20(ii). Since all +three of the above semigroups have projection algebra P, and since both of the above maps is +canonical, it follows that the composition +εφS : CP → MP +is also canonical. +But CP is projection-generated, so there can be at most one canonical ∗- +semigroup homomorphism CP → MP . It follows that in fact εφS = φCP . But then +ker(ε) = ker(εφS) +as φS is injective += ker(φCP ) +as εφS = φCP += µCP = µ +by Theorem 8.20(ii). +91 + +Combining this with Proposition 4.39(iii) and the fundamental homomorphism theorem, it fol- +lows that +E(S) = im(ε) ∼= CP / ker(ε) = CP /µ. +This of course gives (ii). Part (i) also follows, since if S happens to be idempotent-generated +(and fundamental, with P(S) = P), then S = E(S) ∼= CP /µ. +Remark 8.25. Theorem 8.24 shows that (up to isomorphism) there is a unique idempotent- +generated fundamental regular ∗-semigroup with a given projection algebra P. The theorem also +gives a number of ways to get hold of such a semigroup, which for now we will denote by FP . +First, we can construct FP as the quotient CP /µ, where µ = µCP . The elements of this quotient +are µ-classes of P-chains. For two P-chains c = [p1, . . . , pk] and d = [q1, . . . , ql], we combine +Proposition 8.3 with (7.1) and (7.2) to obtain +(c, d) ∈ µ ⇔ θp1 · · · θpk = θq1 · · · θql +and +θpk · · · θp1 = θql · · · θq1 +⇔ θp1 · · · θpk = θq1 · · · θql +and +p1 = q1 +⇔ θp2 · · · θpk = θq2 · · · θql +and +p1 = q1. +Alternatively, we could take FP to be the idempotent-generated subsemigroup E(MP ) of MP . +Elements of E(MP ) are of the form +idp↓ +1 • · · · • idp↓ +k +for p1, . . . , pk ∈ P with p1 F · · · F pk. +As in Definition 8.19, for such pi we have +idp↓ +1 • idp↓ +2 • · · · • idp↓ +k = γp1p2γp2p3 · · · γpk−1pk = (θp2 · · · θpk)|p↓ +1, +so E(MP ) consists of all such maps. +As another possibility, we could take FP = E(MP ), where MP ≤ TP ×T op +P +is the isomorphic +copy of MP discussed in Remark 8.22. +Using the over-line notation from that remark, and +using (4.2), the projections of MP are of the form +p = (Θp, Θp∗) = (θp, θp) +for p ∈ P. +Keeping in mind that the operation in MP is (ordinary and reverse) composition, it follows that +a general element of E(MP ) has the form +(θp1 · · · θpk, θpk · · · θp1) +for p1, . . . , pk ∈ P with p1 F · · · F pk. +9 +Inverse semigroups and inductive groupoids +We have noted on a number of occasions that any inverse semigroup is a regular ∗-semigroup +(with a∗ = a−1). In this final chapter, we look at how the theory developed above simplifies +in the case of inverse semigroups. In particular, we will see that our results allow us to deduce +the celebrated Ehresmann–Schein–Nambooripad (ESN) Theorem, stated below as Theorem 9.1. +This theorem was first explicitly formulated by Lawson in [87, Theorem 4.1.8], who named it as +such in order to honour the contributions of the three stated mathematicians to the development +of the result. We discussed this in some detail in Chapter 1, but see [87, Chapter 4] and [66,67] +for more. +In what follows, we write IS for the category of inverse semigroups. Morphisms in IS are sim- +ply the semigroup homomorphisms. (Any semigroup homomorphism between inverse semigroups +automatically respects the involutions.) We also write IG for the category of inductive groupoids, +i.e. the ordered groupoids whose object set is a semilattice (under the order inherited from the +containing groupoid). Morphisms in IG are the inductive functors, i.e. the ordered groupoid +functors whose object maps are semilattice morphisms. The ESN Theorem is as follows: +92 + +Theorem 9.1. The category IS of inverse semigroups (with semigroup homomorphisms) is iso- +morphic to the category IG of inductive groupoids (with inductive functors). +We will give a proof of this theorem in Section 9.2, relying on our above results on regular +∗-semigroups. But before we do, in Section 9.1 we show how we are somewhat-inevitably led to +the ESN Theorem by considering the simplifications that arise when we specialise our general +theory to inverse semigroups. +9.1 +The chained projection groupoid associated to an inverse semigroup +In Theorem 6.7 we showed that the functors +G : RSS → CPG +and +S : CPG → RSS +are mutually inverse isomorphisms between +• the category RSS of regular ∗-semigroups, with ∗-semigroup homomorphisms, and +• the category CPG of chained projection groupoids, with chained projection functors. +It follows that for any subcategory C of RSS, the functor G restricts to an isomorphism from C +onto its image G(C) in CPG: +RSS +C +CPG +G(C) +G +S +In particular, the category IS of inverse semigroups is isomorphic to its image G(IS). To under- +stand this image, fix an inverse semigroup S, and let (G, ε) = G(S) be the chained projection +groupoid associated to S, as in Definitions 5.4 and 5.16. As usual we write +E = E(S) = {e ∈ S : e2 = e} +and +P = P(S) = {p ∈ S : p2 = p = p−1} +for the semilattice of idempotents, and the projection algebra of S, respectively. Since each idem- +potent is self-inverse, it follows that in fact P = E. As there will be a number of orders in play, we +will (temporarily) write ≤′ for the natural partial order on E, defined by e ≤′ f ⇔ e = ef(= fe). +Note then that the (order-theoretic) meet of two idempotents is their product (in S), e∧f = ef, +and so e ≤′ f ⇔ e = e ∧ f. +As in (5.1), and remembering that idempotents commute, the θ maps on E(= P) are given +by: +eθf = fef = ef = e ∧ f = fθe +for all e, f ∈ E. +(9.2) +It follows quickly from this that the ≤ and ≤F relations, defined on E(= P) in (3.5) and (3.10), +coincide with ≤′ defined above. In particular, ≤F is a partial order, and so F = ≤F ∩≥F = ∆E +is the trivial relation on E. It follows that the only E-paths are of the form (e, e, . . . , e), and each +such path is ≈-equivalent to (e) ≡ e. This all shows that the chain groupoid C = C (E) is trivial, +meaning that it consists entirely of its identity morphisms. That is, we have C = vC = E, and +e, f ∈ E are composable in C precisely when e = f. It follows that the evaluation map ε : C → G +(being a v-functor) is just the inclusion ι : E �→ G(= S). This means that all the information +contained in the pair (G, ε) = (G, ι) is in fact contained in the groupoid G = G(S) alone. +Roughly speaking, the next result shows that the above situation is reversible. +93 + +Proposition 9.3. Let (G, ε) be a chained projection groupoid, and let P = vG and C = C (P). +Then (G, ε) = G(S) for some inverse semigroup S if and only if C = P is trivial (in which case ε +is the inclusion ε = ι : P �→ G). +Proof. We have already proved the forwards implication. Conversely, assume C = P is trivial, +and (hence) ε is the inclusion ε = ι : P �→ G. Let S = S(G, ι), so that S is a regular ∗-semigroup +with projection algebra P(S) = P, and (G, ι) = G(S); cf. Theorem 6.7. We must show that S is +inverse, and we can do this by showing that idempotents commute. As usual, we write E = E(S) +for the set of idempotents of S. By Proposition 4.39(ii), E is contained in im(ε) = im(ι), and +by assumption the latter is equal to P. Since P ⊆ E for any regular ∗-semigroup, it follows +that E = P. But also E = P 2 by Lemma 2.23(ii), so it follows that P = P 2, i.e. that P is +a subsemigroup of S. Now let e, f ∈ E = P. Since e, f and ef are all projections, we have +ef = (ef)∗ = f∗e∗ = fe, as required. +Remark 9.4. Given an inverse semigroup S, the groupoid G = G(S) is inductive, as the object +set vG = E = E(S) is a semilattice. At the outset, one might expect there to also be a version +of Proposition 9.3 stating that a chained projection groupoid (G, ε) corresponds to an inverse +semigroup if and only if G is inductive. However, it follows from Example 5.21 that this is not the +case. Indeed, there we considered two regular ∗-semigroups S1 and S2, with S2 inverse and S1 +non-inverse, and with the same groupoids G(S1) = G(S2). Since S2 is inverse, this groupoid +is inductive. It follows that the non-inverse regular ∗-semigroup S1 gives rise to an inductiive +groupoid G(S1). +The next result is essentially a reformulation of Proposition 9.3, but it seems to be worth +recording. +Proposition 9.5. Let S be a regular ∗-semigroup, with projection algebra P = P(S). Then S is +inverse if and only if F = ∆P . +Proof. Clearly F = ∆P is equivalent to C = P, and the result then follows from Proposition 9.3 +(and Theorem 6.7). +Now that we have some understanding of the image G(IS) of IS in CPG (cf. Proposition 9.3), +we are a step closer to proving Theorem 9.1. +One might now go on to show that G(IS) is +isomorphic to IG, the category of inductive groupoids. However, there are still some subtle +points remaining to be dealt with, some so subtle as to be almost invisible at this point. Rather +than continue on this path, we take a more direct/canonical route in the next section, though +we will make use of Proposition 9.3 on a number of occasions. +9.2 +The Ehresmann–Schein–Nambooripad Theorem +In Proposition 9.3 we characterised the chained projection groupoids (G, ε) corresponding to +inverse semigroups; essentially the proposition shows that these are precisely those in which the +evaluation map ε carries no information whatsoever. Consequently, if we wish to prove the ESN +Theorem (stated above as Theorem 9.1) we may as well dispense with evaluation maps altogether +(when dealing with inverse semigroups), and consider the groupoid construction S �→ G(S) as +a functor IS → IG, and prove that this is a category isomorphism. +This is of course the +standard/canonical approach to the ESN Theorem, but the details required for the proof will be +taken from our general theory developed in earlier chapters. Following this canonical approach, +we proceed by defining two functors +G : IS → IG +and +S : IG → IS, +and showing that they are mutually inverse isomorphisms. +94 + +First, for an inverse semigroup S, we take G(S) = G = G(S) to be the ordered groupoid +constructed in Definition 5.4. As in Section 9.1, we have E = P (where as usual E = E(S) and +P = P(S)). Moreover, following Definition 5.4, for e, f ∈ E we have +e ≤ f in G +⇔ +e = ef in S +⇔ +e ≤ f in E. +In particular, E = vG is a semilattice (under the order inherited from G), and so G(S) = G is +indeed an inductive groupoid. Next, it is easy to see that any morphism φ : S → S′ in IS is an +inductive functor G(S) → G(S′). Indeed, φ is an ordered functor by Lemma 6.2, and the object +map vφ is a semilattice morphism since for any e, f ∈ vG(S) = E(S) we have +(e ∧ f)φ = (ef)φ = (eφ)(fφ) = eφ ∧ fφ. +So we can simply take G(φ) = φ for any morphism φ from IS. +To define the functor S, consider an inductive groupoid G, with object semilattice E = vG. +The semigroup S = S(G) will have the same underlying set as G. To define the product on S, +which we will denote by ⋆, consider two morphisms a, b ∈ G(= S), and write e = r(a) and +f = d(b). Since E is a semilattice, the meet g = e ∧ f exists, and since g ≤ e, f we can define +a ⋆ b = a⇂g ◦ g⇃b. At this point it is not at all clear that S = S(G) = (G, ⋆) is a semigroup, let +alone inverse, but we will deal with this shortly. For any morphism φ : G → G′ in IG, we again +define S(φ) = φ. +Lemma 9.6. If G is an inductive groupoid, with object semilattice vG = E, then +(i) E is a projection algebra with respect to the maps +θe : E → E +given by +fθe = e ∧ f +for e, f ∈ E, +(ii) G is a projection groupoid. +Proof. (i). It is a routine matter to check that axioms (P1)–(P5) hold. +(ii). It remains to show that condition (G1) from Definition 4.8 holds, and for this we verify (G1d). +To do so, fix some a ∈ G, and write p = d(a) and q = r(a). We must show that ϑa : p↓ → q↓ is +a projection algebra morphism, i.e. that +(eθf)ϑa = (eϑa)θfϑa +for all e, f ≤ p. +By the definition of the θ maps, this amounts to showing that +(e ∧ f)ϑa = eϑa ∧ fϑa +for all e, f ≤ p, +i.e. that ϑa is a semilattice morphism. But this follows quickly from basic order-theoretic facts: +• p↓ and q↓ are both meet semilattices, as order ideals of the semilattice E, and +• ϑa : p↓ → q↓ and its inverse ϑ−1 +a += ϑa−1 : q↓ → p↓ are both order-preserving, by Lemma 2.15 +(cf. Lemma 2.12). +It follows that ϑa preserves meets. +95 + +Lemma 9.7. If G is an inductive groupoid, with object semilattice vG = E, then +(i) C (E) = E is trivial (where the projection algebra structure on E is as in Lemma 9.6), +(ii) (G, ι) is a chained projection groupoid, where ι : E �→ G is the inclusion, +(iii) S(G) = S(G, ι) is an inverse semigroup. +Proof. (i). For any e, f ∈ E we have fθe = e ∧ f = f ∧ e = eθf. We then obtain F = ∆E, and +hence C (E) = E, in exactly the same way as in the discussion following (9.2). +(ii). By Lemma 9.6, it remains to check that (G2) holds, and this is essentially trivial. Indeed, +consider some b-linked pair (e, f), where b ∈ G, and let e1, e2, f1, f2 ∈ E be as in (4.19). Since +F = ∆E, it follows from Lemma 4.18(iii) that in fact e1 = e2 = e and f1 = f2 = f. But then +λ(e, b, f) = ι[e, e] ◦ e⇃b ◦ ι[f, f] = e ◦ e⇃b ◦ f = e⇃b +and similarly +ρ(e, b, f) = e⇃b. +(iii). +The underlying sets of S(G) and S(G, ι) are both G. +To check that the operations ⋆ +and • coincide, let a, b ∈ G. Write e = r(a) and f = d(b), and also set g = e ∧ f. Following +Definition 4.24 (applied to the chained projection groupoid (G, ι)), we have +a • b = a⇂e′ ◦ ι[e′, f′] ◦ f′⇃b +where +e′ = fθe +and +f′ = eθf. +But e′ = fθe = e ∧ f = g, and similarly f′ = g, so in fact +a • b = a⇂g ◦ ι[g, g] ◦ g⇃b = a⇂g ◦ g ◦ g⇃b = a⇂g ◦ g⇃b = a ⋆ b. +It remains to check that the regular ∗-semigroup S = S(G, ι) is inverse, and this follows from +Proposition 9.3 and part (i) of the current lemma, as (G, ι) = G(S). +We call (G, ι), as in Lemma 9.7(ii), the trivial chained projective groupoid associated to the +inductive groupoid G. +Lemma 9.8. If G1 and G2 are inductive groupoids, then any inductive functor G1 → G2 is a +semigroup homomorphism S(G1) → S(G2). +Proof. Let φ : G1 → G2 be an inductive functor. By Lemmas 6.3 and 9.7(iii), it suffices to show +that φ is a chained projection functor (G1, ι1) → (G2, ι2), where these are the trivial chained +projective groupoids associated to G1 and G2. +Following Definition 6.1, it remains to verify +conditions (F1) and (F2). +(F1). This follows from the fact that vφ is a semilattice morphism (as φ is inductive), and keeping +the rule eθf = e ∧ f in mind. +(F2). Writing Ei = vGi (i = 1, 2), each Ci = Ei is trivial, so the induced map Φ at the top of +the diagram in Definition 6.1 is just the object map Φ = vφ. The diagram then becomes: +E1 +E2 +G1 +G2, +vφ +ι1 +ι2 +φ +and this obviously commutes. +96 + +It follows immediately from Lemmas 9.7(iii) and 9.8 that S is indeed a functor IG → IS. We +can now tie together the loose ends: +Proof of Theorem 9.1. It remains to show that S and G are mutually inverse, i.e. that +(i) S(G(S)) = S for any inverse semigroup S, and +(ii) G(S(G)) = G for any inductive groupoid G. +Beginning with (i), fix an inverse semigroup S, and let G = G(S) = G(S). By Proposition 9.3, +G(S) = (G, ι) is the trivial chained projection groupoid associated to G. Combining this with +Lemmas 9.7(iii) and 6.6, it follows that +S(G(S)) = S(G) = S(G, ι) = S(G(S)) = S. +For (ii), fix an inductive groupoid G, and write S = S(G). By Lemma 9.7(iii) we have S = S(G, ι), +where (G, ι) is the trivial chained projection groupoid associated to G. Combining this with +Lemma 6.4 (and Definitions 5.4 and 5.16), it follows that +(G, ι) = G(S(G, ι)) = G(S) = (G(S), ε(S)), +and so G = G(S) = G(S) = G(S(G)). +We have now seen that the Ehresmann–Schein–Nambooripad Theorem (Theorem 9.1) follows +from our Theorem 6.7. Alternatively, one could also obtain a direct (and quite transparent) proof +of the ESN Theorem by specialising our proof of Theorem 6.7 to the inverse case. We will not +give the full outline of this, but will comment briefly on the most involved step, which is to show +that the product • defined on the groupoid G is associative. In the inverse/inductive case, this +product is defined by +a • b = a⇂e ◦ e⇃b +for a, b ∈ G, where e = r(a) ∧ d(b). +It follows that for a, b, c ∈ G, we have +(a • b) • c = a⇂f ◦ b⇂g ◦ c⇂h +and +a • (b • c) = a⇂f′ ◦ b⇂g′ ◦ c⇂h′ +for some f, g, h, f′, g′, h′ ∈ E = vG. We can then show that (a • b) • c and a • (b • c) are equal +by showing that they have equal domains and equal ranges. This follows from (4.30) and (4.31), +which were proved quite early in Section 4.3. +9.3 +Fundamental inverse semigroups +So far we have examined the way that the general theory developed in Chapters 3–6 simplifies +in the case of inverse semigroups. Specifically, we were able to deduce the Ehresmann–Schein– +Nambooripad Theorem (stated in Theorem 9.1) from our results. It is then of course natural to +follow the same program for the results of Chapters 7 and 8. +The results of Chapter 7 become trivial/vacuous in the context of inverse semigroups, as +idempotent-generated inverse semigroups are simply inverse semigroups consisting entirely of +idempotents, i.e. semilattices. It is therefore no surprise that the free (projection-generated) +regular ∗-semigroup over a semilattice is simply the semilattice itself. There is another obvious +way to to think about this. Given a semilattice E, regarded as a projection algebra with θ +maps as in (9.2), we have already seen that the chain groupoid C = C (E) is trivial, in the +sense that C = vC = E, with e, f ∈ E being composable in C precisely when e = f. Recall +from Definition 3.39 that the chain semigroup of E is the semigroup CE = S(C , idC ). Since idC +97 + +can also be seen as the inclusion idC = ι : E = C �→ C , it follows from Lemma 9.7(iii) +that CE = S(C , idC ) = S(C , ι) = S(C ). This semigroup has underlying set C = E, and the +product • = ⋆ is given by +e • f = e ⋆ f = e⇂e∧f ◦ e∧f⇃f = (e ∧ f) ◦ (e ∧ f) = e ∧ f. +In other words, CE is just E itself. +On the other hand, the results of Chapter 8 do have interesting specialisations to inverse +semigroups, as we now briefly discuss. Fix a semilattice E, which we again consider as a pro- +jection algebra, with θ maps given in (9.2). As in the proof of Lemma 9.6(ii), the projection +algebra morphisms e↓ → f↓ (for e, f ∈ E) are precisely the semilattice morphisms e↓ → f↓. +Since the order and meet operations are inter-definable in E (viz. s ≤ t ⇔ s = s ∧ t), it follows +that these morphisms are also precisely the order isomorphisms e↓ → f↓. It follows that the +(inductive) groupoid M = M(E) from Definition 8.5 consists precisely of all such order isomor- +phisms e↓ → f↓ (e, f ∈ E). The resulting semigroup ME = S(M) has underlying set M, and +product • = ⋆ given as follows. For α, β ∈ ME with r(α) = e and d(β) = f, we have +α • β = α ⋆ β = α⇂e∧f ◦ e∧f⇃β. +This last expression is the composition in the groupoid M of the bijections α⇂e∧f and e∧f⇃β; the +(set-theoretic) range of the former is equal to the (set-theoretic) domain of the latter, and this +is precisely the order ideal e↓ ∩ f↓ = (e ∧ f)↓. However, we can also think of α and β as partial +bijections of E, i.e. as elements of the symmetric inverse monoid IE, and the above product +α • β = α ⋆ β is simply the product of α and β in IE, i.e. their composition as binary relations: +α • β = α ⋆ β = {(s, t) ∈ E × E : (s, g) ∈ α and (g, t) ∈ β for some g ∈ E}. +In this way, we can think of ME as an inverse subsemigroup of the symmetric inverse monoid IE. +This is in fact the original approach taken by Munn in his seminal paper [109], where the +semigroup ME (considered as a subsemigroup of IE) was denoted TE. +References +[1] S. Armstrong. Structure of concordant semigroups. J. Algebra, 118(1):205–260, 1988. +[2] K. Auinger. Krohn-Rhodes complexity of Brauer type semigroups. Port. Math., 69(4):341–360, 2012. +[3] K. Auinger. Pseudovarieties generated by Brauer type monoids. Forum Math., 26(1):1–24, 2014. +[4] K. Auinger, Y. Chen, X. Hu, Y. Luo, and M. V. Volkov. The finite basis problem for Kauffman monoids. +Algebra Universalis, 74(3-4):333–350, 2015. +[5] S. Awodey. Category theory, volume 52 of Oxford Logic Guides. Oxford University Press, Oxford, second +edition, 2010. +[6] G. Benkart and T. Halverson. Motzkin algebras. European J. Combin., 36:473–502, 2014. +[7] G. Benkart and T. Halverson. Partition algebras Pk(n) with 2k > n and the fundamental theorems of +invariant theory for the symmetric group Sn. J. Lond. Math. Soc. (2), 99(1):194–224, 2019. +[8] G. Birkhoff. Rings of sets. Duke Math. J., 3(3):443–454, 1937. +[9] M. Borisavljević, K. Došen, and Z. Petrić. Kauffman monoids. J. Knot Theory Ramifications, 11(2):127–143, +2002. +[10] R. Brauer. On algebras which are connected with the semisimple continuous groups. Ann. of Math. (2), +38(4):857–872, 1937. +[11] M. Brittenham, S. W. Margolis, and J. Meakin. Subgroups of the free idempotent generated semigroups +need not be free. J. Algebra, 321(10):3026–3042, 2009. +[12] N. Brownlowe, N. S. Larsen, and N. Stammeier. On C∗-algebras associated to right LCM semigroups. +Trans. Amer. Math. Soc., 369(1):31–68, 2017. +[13] S. Burris and H. P. Sankappanavar. A course in universal algebra, volume 78 of Graduate Texts in Mathe- +matics. Springer-Verlag, New York-Berlin, 1981. +98 + +[14] A. Buss, R. Exel, and R. Meyer. Reduced C∗-algebras of Fell bundles over inverse semigroups. Israel J. +Math., 220(1):225–274, 2017. +[15] D. M. Clark and B. A. Davey. Natural dualities for the working algebraist, volume 57 of Cambridge Studies +in Advanced Mathematics. Cambridge University Press, Cambridge, 1998. +[16] A. H. Clifford and G. B. Preston. The algebraic theory of semigroups. Vol. I. Mathematical Surveys, No. +7. American Mathematical Society, Providence, R.I., 1961. +[17] Y. Dandan, I. Dolinka, and V. Gould. A group-theoretical interpretation of the word problem for free +idempotent generated semigroups. Adv. Math., 345:998–1041, 2019. +[18] D. DeWolf and D. Pronk. The Ehresmann-Schein-Nambooripad theorem for inverse categories. Theory +Appl. Categ., 33:Paper No. 27, 813–831, 2018. +[19] I. Dolinka. Free idempotent generated semigroups: the word problem and structure via gain graphs. Israel +J. Math., 245(1):347–387, 2021. +[20] I. Dolinka, J. East, A. Evangelou, D. FitzGerald, N. Ham, J. Hyde, and N. Loughlin. Enumeration of +idempotents in diagram semigroups and algebras. J. Combin. Theory Ser. A, 131:119–152, 2015. +[21] I. Dolinka, J. East, A. Evangelou, D. FitzGerald, N. Ham, J. Hyde, N. Loughlin, and J. D. Mitchell. +Enumeration of idempotents in planar diagram monoids. J. Algebra, 522:351–385, 2019. +[22] I. Dolinka, J. East, and R. D. Gray. Motzkin monoids and partial Brauer monoids. J. Algebra, 471:251–298, +2017. +[23] I. Dolinka and R. D. Gray. Maximal subgroups of free idempotent generated semigroups over the full linear +monoid. Trans. Amer. Math. Soc., 366(1):419–455, 2014. +[24] I. Dolinka, R. D. Gray, and N. Ruškuc. On regularity and the word problem for free idempotent generated +semigroups. Proc. Lond. Math. Soc. (3), 114(3):401–432, 2017. +[25] I. Dolinka and N. Ruškuc. Every group is a maximal subgroup of the free idempotent generated semigroup +over a band. Internat. J. Algebra Comput., 23(3):573–581, 2013. +[26] M. P. Drazin. Regular semigroups with involution. In Symposium on Regular Semigroups, pages 29–46. +Northern Illinois University, DeKalb, 1979. +[27] D. Easdown. Biordered sets are biordered subsets of idempotents of semigroups. J. Austral. Math. Soc. +Ser. A, 37(2):258–268, 1984. +[28] D. Easdown. Biordered sets of bands. Semigroup Forum, 29(1-2):241–246, 1984. +[29] D. Easdown. Biordered sets of eventually regular semigroups. Proc. London Math. Soc. (3), 49(3):483–503, +1984. +[30] D. Easdown. +A new proof that regular biordered sets come from regular semigroups. +Proc. Roy. Soc. +Edinburgh Sect. A, 96(1-2):109–116, 1984. +[31] D. Easdown. Biordered sets come from semigroups. J. Algebra, 96(2):581–591, 1985. +[32] D. Easdown. Biordered sets of rings. In Monash Conference on Semigroup Theory (Melbourne, 1990), pages +43–49. World Sci. Publ., River Edge, NJ, 1991. +[33] J. East. Generators and relations for partition monoids and algebras. J. Algebra, 339:1–26, 2011. +[34] J. East. On the singular part of the partition monoid. Internat. J. Algebra Comput., 21(1-2):147–178, 2011. +[35] J. East. Infinite partition monoids. Internat. J. Algebra Comput., 24(4):429–460, 2014. +[36] J. East and P. A. Azeef Muhammed. Fundamental diagram monoids. In preparation. +[37] J. East, P. A. Azeef Muhammed, and N. Ruškuc. Free idempotent-generated regular ∗-semigroups. In +preparation. +[38] J. East and D. G. FitzGerald. The semigroup generated by the idempotents of a partition monoid. J. +Algebra, 372:108–133, 2012. +[39] J. East and R. D. Gray. Diagram monoids and Graham–Houghton graphs: Idempotents and generating +sets of ideals. J. Combin. Theory Ser. A, 146:63–128, 2017. +[40] J. East, J. D. Mitchell, N. Ruškuc, and M. Torpey. Congruence lattices of finite diagram monoids. Adv. +Math., 333:931–1003, 2018. +[41] J. East and N. Ruškuc. Classification of congruences of twisted partition monoids. Adv. Math., 395:Paper +No. 108097, 65 pp., 2022. +[42] J. East and N. Ruškuc. Congruences on infinite partition and partial Brauer monoids. Mosc. Math. J., +22(2):295–372, 2022. +[43] J. East and N. Ruškuc. Properties of congruences of twisted partition monoids and their lattices. J. Lond. +Math. Soc. (2), 106(1):311–357, 2022. +99 + +[44] J. East and N. Ruškuc. Congruence lattices of ideals in categories and (partial) semigroups. Mem. Amer. +Math. Soc., to appear, arXiv:2001.01909 . +[45] C. Ehresmann. Gattungen von lokalen Strukturen. Jber. Deutsch. Math.-Verein., 60(Abt. 1):49–77, 1957. +[46] C. Ehresmann. Catégories inductives et pseudo-groupes. Ann. Inst. Fourier (Grenoble), 10:307–332, 1960. +[47] C. Farthing, P. S. Muhly, and T. Yeend. Higher-rank graph C∗-algebras: an inverse semigroup and groupoid +approach. Semigroup Forum, 71(2):159–187, 2005. +[48] D. G. Fitz-Gerald. On inverses of products of idempotents in regular semigroups. J. Austral. Math. Soc., +13:335–337, 1972. +[49] D. G. FitzGerald. Factorizable inverse monoids. Semigroup Forum, 80(3):484–509, 2010. +[50] D. G. FitzGerald. Groupoids on a skew lattice of objects. Art Discrete Appl. Math., 2(2):Paper No. 2.03, +14, 2019. +[51] D. G. FitzGerald and K. W. Lau. On the partition monoid and some related semigroups. Bull. Aust. Math. +Soc., 83(2):273–288, 2011. +[52] The GAP Group. GAP – Groups, Algorithms, and Programming, Version 4.10.1, 2019. +[53] V. Gould. +Restriction and Ehresmann semigroups. +In Proceedings of the International Conference on +Algebra 2010, pages 265–288. World Sci. Publ., Hackensack, NJ, 2012. +[54] V. Gould and C. Hollings. Restriction semigroups and inductive constellations. Comm. Algebra, 38(1):261– +287, 2010. +[55] V. Gould and Y. Wang. Beyond orthodox semigroups. J. Algebra, 368:209–230, 2012. +[56] V. Gould and D. Yang. +Every group is a maximal subgroup of a naturally occurring free idempotent +generated semigroup. Semigroup Forum, 89(1):125–134, 2014. +[57] R. Gray and N. Ruskuc. On maximal subgroups of free idempotent generated semigroups. Israel J. Math., +189:147–176, 2012. +[58] R. Gray and N. Ruškuc. Maximal subgroups of free idempotent-generated semigroups over the full trans- +formation monoid. Proc. Lond. Math. Soc. (3), 104(5):997–1018, 2012. +[59] J. A. Green. On the structure of semigroups. Ann. of Math. (2), 54:163–172, 1951. +[60] P. A. Grillet. The structure of regular semigroups. I. A representation. Semigroup Forum, 8(2):177–183, +1974. +[61] P. A. Grillet. The structure of regular semigroups. II. Cross-connections. Semigroup Forum, 8(3):254–259, +1974. +[62] T. Halverson and A. Ram. Partition algebras. European J. Combin., 26(6):869–921, 2005. +[63] R. E. Hartwig. How to partially order regular elements. Math. Japon., 25(1):1–13, 1980. +[64] J. B. Hickey. Semigroups under a sandwich operation. Proc. Edinburgh Math. Soc. (2), 26(3):371–382, 1983. +[65] C. Hollings. Extending the Ehresmann-Schein-Nambooripad theorem. Semigroup Forum, 80(3):453–476, +2010. +[66] C. Hollings. The Ehresmann-Schein-Nambooripad theorem and its successors. Eur. J. Pure Appl. Math., +5(4):414–450, 2012. +[67] C. D. Hollings and M. V. Lawson. Wagner’s theory of generalised heaps. Springer, Cham, 2017. +[68] J. M. Howie. Fundamentals of semigroup theory, volume 12 of London Mathematical Society Monographs. +New Series. The Clarendon Press, Oxford University Press, New York, 1995. Oxford Science Publications. +[69] G. Huet. Confluent reductions: abstract properties and applications to term rewriting systems. J. Assoc. +Comput. Mach., 27(4):797–821, 1980. +[70] T. Imaoka. On fundamental regular ∗-semigroups. Mem. Fac. Sci. Shimane Univ., 14:19–23, 1980. +[71] T. Imaoka. Prehomomorphisms on regular ∗-semigroups. Mem. Fac. Sci. Shimane Univ., 15:23–27, 1981. +[72] T. Imaoka. Some remarks on fundamental regular ∗-semigroups. In Recent developments in the algebraic, +analytical, and topological theory of semigroups (Oberwolfach, 1981), volume 998 of Lecture Notes in Math., +pages 270–280. Springer, Berlin, 1983. +[73] P. R. Jones. A common framework for restriction semigroups and regular ∗-semigroups. J. Pure Appl. +Algebra, 216(3):618–632, 2012. +[74] V. F. R. Jones. Index for subfactors. Invent. Math., 72(1):1–25, 1983. +[75] V. F. R. Jones. Hecke algebra representations of braid groups and link polynomials. Ann. of Math. (2), +126(2):335–388, 1987. +100 + +[76] V. F. R. Jones. The Potts model and the symmetric group. In Subfactors (Kyuzeso, 1993), pages 259–267. +World Sci. Publ., River Edge, NJ, 1994. +[77] V. F. R. Jones. A quotient of the affine Hecke algebra in the Brauer algebra. Enseign. Math. (2), 40(3- +4):313–344, 1994. +[78] L. H. Kauffman. State models and the Jones polynomial. Topology, 26(3):395–407, 1987. +[79] L. H. Kauffman. An invariant of regular isotopy. Trans. Amer. Math. Soc., 318(2):417–471, 1990. +[80] M. Khoshkam and G. Skandalis. Regular representation of groupoid C∗-algebras and applications to inverse +semigroups. J. Reine Angew. Math., 546:47–72, 2002. +[81] N. V. Kitov and M. V. Volkov. +Identities of the Kauffman monoid K4 and of the Jones monoid J4. +In Fields of logic and computation. III, volume 12180 of Lecture Notes in Comput. Sci., pages 156–178. +Springer, Cham, [2020. +[82] G. Kudryavtseva and V. Mazorchuk. +On presentations of Brauer-type monoids. +Cent. Eur. J. Math., +4(3):413–434 (electronic), 2006. +[83] M. Laca, I. Raeburn, J. Ramagge, and M. F. Whittaker. Equilibrium states on operator algebras associated +to self-similar actions of groupoids on graphs. Adv. Math., 331:268–325, 2018. +[84] D. Larsson. +Combinatorics on Brauer-type semigroups. +U.U.D.M. Project Report 2006:3, Available at +http://www.diva-portal.org/smash/get/diva2:305049/FULLTEXT01.pdf. +[85] K. W. Lau and D. G. FitzGerald. Ideal structure of the Kauffman and related monoids. Comm. Algebra, +34(7):2617–2629, 2006. +[86] M. V. Lawson. Semigroups and ordered categories. I. The reduced case. J. Algebra, 141(2):422–462, 1991. +[87] M. V. Lawson. Inverse semigroups. World Scientific Publishing Co., Inc., River Edge, NJ, 1998. The theory +of partial symmetries. +[88] M. V. Lawson. Ordered groupoids and left cancellative categories. Semigroup Forum, 68(3):458–476, 2004. +[89] M. V. Lawson. Non-commutative Stone duality: inverse semigroups, topological groupoids and C∗-algebras. +Internat. J. Algebra Comput., 22(6):1250058, 47, 2012. +[90] M. V. Lawson. Recent developments in inverse semigroup theory. Semigroup Forum, 100(1):103–118, 2020. +[91] M. V. Lawson. On Ehresmann semigroups. Semigroup Forum, 103(3):953–965, 2021. +[92] G. Lehrer and R. Zhang. The second fundamental theorem of invariant theory for the orthogonal group. +Ann. of Math. (2), 176(3):2031–2054, 2012. +[93] G. Lehrer and R. B. Zhang. +The Brauer category and invariant theory. +J. Eur. Math. Soc. (JEMS), +17(9):2311–2351, 2015. +[94] X. Li. Semigroup C∗-algebras and amenability of semigroups. J. Funct. Anal., 262(10):4302–4340, 2012. +[95] X. Li. Partial transformation groupoids attached to graphs and semigroups. Int. Math. Res. Not. IMRN, +(17):5233–5259, 2017. +[96] S. Mac Lane. +Categories for the working mathematician, volume 5 of Graduate Texts in Mathematics. +Springer-Verlag, New York, second edition, 1998. +[97] V. Maltcev and V. Mazorchuk. Presentation of the singular part of the Brauer monoid. Math. Bohem., +132(3):297–323, 2007. +[98] P. Martin. Temperley-Lieb algebras for nonplanar statistical mechanics—the partition algebra construction. +J. Knot Theory Ramifications, 3(1):51–82, 1994. +[99] P. Martin. The structure of the partition algebras. J. Algebra, 183(2):319–358, 1996. +[100] P. Martin and V. Mazorchuk. +On the representation theory of partial Brauer algebras. +Q. J. Math., +65(1):225–247, 2014. +[101] P. P. Martin. The decomposition matrices of the Brauer algebra over the complex field. Trans. Amer. Math. +Soc., 367(3):1797–1825, 2015. +[102] P. P. Martin and G. Rollet. The Potts model representation and a Robinson-Schensted correspondence for +the partition algebra. Compositio Math., 112(2):237–254, 1998. +[103] V. Mazorchuk. Endomorphisms of Bn, PBn, and Cn. Comm. Algebra, 30(7):3489–3513, 2002. +[104] J. Meakin, P. A. Azeef Muhammed, and A. R. Rajan. The mathematical work of K. S. S. Nambooripad. +In Semigroups, categories, and partial algebras, volume 345 of Springer Proc. Math. Stat., pages 107–140. +Springer, Singapore, 2021. +[105] D. Milan and B. Steinberg. On inverse semigroup C∗-algebras and crossed products. Groups Geom. Dyn., +8(2):485–512, 2014. +101 + +[106] D. D. Miller and A. H. Clifford. Regular D-classes in semigroups. Trans. Amer. Math. Soc., 82:270–280, +1956. +[107] J. D. Mitchell et al. Semigroups - GAP package, Version 3.3.1, 2020. +[108] H. Mitsch. A natural partial order for semigroups. Proc. Amer. Math. Soc., 97(3):384–388, 1986. +[109] W. D. Munn. Fundamental inverse semigroups. Quart. J. Math. Oxford Ser. (2), 21:157–170, 1970. +[110] K. S. S. Nambooripad. Structure of regular semigroups. I. Mem. Amer. Math. Soc., 22(224):vii+119, 1979. +[111] K. S. S. Nambooripad. The natural partial order on a regular semigroup. Proc. Edinburgh Math. Soc. (2), +23(3):249–260, 1980. +[112] K. S. S. Nambooripad and F. Pastijn. Subgroups of free idempotent generated regular semigroups. Semi- +group Forum, 21(1):1–7, 1980. +[113] K. S. S. Nambooripad and F. J. C. M. Pastijn. Regular involution semigroups. In Semigroups (Szeged, +1981), volume 39 of Colloq. Math. Soc. János Bolyai, pages 199–249. North-Holland, Amsterdam, 1985. +[114] T. E. Nordahl and H. E. Scheiblich. Regular ∗-semigroups. Semigroup Forum, 16(3):369–377, 1978. +[115] F. Pastijn. The biorder on the partial groupoid of idempotents of a semigroup. J. Algebra, 65(1):147–187, +1980. +[116] A. L. T. Paterson. Groupoids, inverse semigroups, and their operator algebras, volume 170 of Progress in +Mathematics. Birkhäuser Boston, Inc., Boston, MA, 1999. +[117] A. L. T. Paterson. Graph inverse semigroups, groupoids and their C∗-algebras. J. Operator Theory, 48(3, +suppl.):645–662, 2002. +[118] M. Petrich. Certain varieties of completely regular ∗-semigroups. Boll. Un. Mat. Ital. B (6), 4(2):343–370, +1985. +[119] L. Pontrjagin. The general topological theorem of duality for closed sets. Ann. of Math. (2), 35(4):904–914, +1934. +[120] L. Pontrjagin. The theory of topological commutative groups. Ann. of Math. (2), 35(2):361–388, 1934. +[121] G. B. Preston. Inverse semi-groups. J. London Math. Soc., 29:396–403, 1954. +[122] G. B. Preston. Representations of inverse semi-groups. J. London Math. Soc., 29:411–419, 1954. +[123] H. A. Priestley. Representation of distributive lattices by means of ordered stone spaces. Bull. London +Math. Soc., 2:186–190, 1970. +[124] H. A. Priestley. Ordered topological spaces and the representation of distributive lattices. Proc. London +Math. Soc. (3), 24:507–530, 1972. +[125] S. Roberts. King of infinite space. Walker and Company, New York, 2006. Donald Coxeter, the man who +saved geometry, With a foreword by Douglas R. Hofstadter. +[126] H. E. Scheiblich. Projective and injective bands with involution. J. Algebra, 109(2):281–291, 1987. +[127] B. M. Schein. On the theory of generalized groups. Dokl. Akad. Nauk SSSR, 153:296–299, 1963. +[128] B. M. Schein. On the theory of generalized groups and generalized heaps. In Theory of Semigroups and +Appl. I (Russian), pages 286–324. Izdat. Saratov. Univ., Saratov, 1965. English translation in Twelve papers +in logic and algebra, Amer. Math. Soc. Translations Ser 2 113, AMS, 1979, pp. 89–122. +[129] M. H. Stone. The theory of representations for Boolean algebras. Trans. Amer. Math. Soc., 40(1):37–111, +1936. +[130] H. N. V. Temperley and E. H. Lieb. Relations between the “percolation” and “colouring” problem and other +graph-theoretical problems associated with regular planar lattices: some exact results for the “percolation” +problem. Proc. Roy. Soc. London Ser. A, 322(1549):251–280, 1971. +[131] V. V. Wagner. Generalized groups. Doklady Akad. Nauk SSSR (N.S.), 84:1119–1122, 1952. +[132] V. V. Wagner. The theory of generalized heaps and generalized groups. Mat. Sbornik N.S., 32(74):545–632, +1953. +[133] S. Wang. An Ehresmann-Schein-Nambooripad-type theorem for a class of P-restriction semigroups. Bull. +Malays. Math. Sci. Soc., 42(2):535–568, 2019. +[134] S. Wang. An Ehresmann-Schein-Nambooripad theorem for locally Ehresmann P-Ehresmann semigroups. +Period. Math. Hungar., 80(1):108–137, 2020. +[135] Y. Wang. Beyond regular semigroups. Semigroup Forum, 92(2):414–448, 2016. +[136] H. Wenzl. On the structure of Brauer’s centralizer algebras. Ann. of Math. (2), 128(1):173–193, 1988. +[137] S. Wilcox. Cellularity of diagram algebras as twisted semigroup algebras. J. Algebra, 309(1):10–31, 2007. +[138] M. Yamada. +On the structure of fundamental regular ∗-semigroups. +Studia Sci. Math. Hungar., 16(3- +4):281–288, 1981. +[139] M. Yamada. p-systems in regular semigroups. Semigroup Forum, 24(2-3):173–187, 1982. +[140] D. Yang, I. Dolinka, and V. Gould. Free idempotent generated semigroups and endomorphism monoids of +free G-acts. J. Algebra, 429:133–176, 2015. +102 +