This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Categorical representation of DRC-semigroups

Abstract

DRC-semigroups model associative systems with domain and range operations, and contain many important classes, such as inverse, restriction, Ehresmann, regular *-, and *-regular semigroups. In this paper we show that the category of DRC-semigroups is isomorphic to a category of certain biordered categories whose object sets are projection algebras in the sense of Jones. This extends the recent groupoid approach to regular *-semigroups of the first and third authors. We also establish the existence of free DRC-semigroups by constructing a left adjoint to the forgetful functor into the category of projection algebras.

Keywords: DRC-semigroup, biordered category, projection algebra.

MSC: 20M50, 20M10, 18B40, 20M05.

James East,111Centre for Research in Mathematics and Data Science, Western Sydney University, Locked Bag 1797, Penrith NSW 2751, Australia. Emails: [email protected], [email protected], [email protected]. Matthias Fresacher,1 P.A. Azeef Muhammed,1 Timothy Stokes222Department of Mathematics, University of Waikato, Hamilton 3216, New Zealand. Email: [email protected].

1 Introduction

DRC-semigroups were introduced in [33], and first studied systematically in [50] where they were used to model associative systems with domain and range operations.333DRC-semigroups were referred to in [33] as reduced UU-semiabundant semigroups satisfying the congruence conditions. The current terminology stems from usage in [50, 30]. Such a semigroup SS has additional unary operations DD and RR satisfying the laws

D(a)a\displaystyle D(a)a =a,\displaystyle=a, D(ab)\displaystyle D(ab) =D(aD(b)),\displaystyle=D(aD(b)), D(ab)\displaystyle D(ab) =D(a)D(ab)D(a),\displaystyle=D(a)D(ab)D(a), R(D(a))\displaystyle R(D(a)) =D(a),\displaystyle=D(a),
aR(a)\displaystyle aR(a) =a,\displaystyle=a, R(ab)\displaystyle R(ab) =R(R(a)b),\displaystyle=R(R(a)b), R(ab)\displaystyle R(ab) =R(b)R(ab)R(b),\displaystyle=R(b)R(ab)R(b), D(R(a))\displaystyle D(R(a)) =R(a).\displaystyle=R(a).

DRC-semigroups contain many important classes, such as inverse [34], restriction [19], Ehresmann [33], regular *- [43], and *-regular semigroups [12]. For example, any inverse semigroup is a DRC-semigroup under the operations D(a)=aa1D(a)=aa^{-1} and R(a)=a1aR(a)=a^{-1}a, or more generally any *-regular semigroup under D(a)=aaD(a)=aa^{\dagger} and R(a)=aaR(a)=a^{\dagger}a, in terms of the Moore–Penrose inverse. Prototypical examples of *-regular semigroups are the multiplicative monoids Mn()M_{n}(\mathbb{R}) and Mn()M_{n}(\mathbb{C}) of real or complex n×nn\times n matrices [46]. The above classes are shown in Figure 1; we refer to [50, 8] for more on these and other classes, as well as more examples and historical background.

One of the strong motivations for studying classes of semigroups such as those above stems from the fact that they can be understood categorically, going back to the celebrated Ehresmann–Nambooripad–Schein (ESN) Theorem, as formulated by Lawson in [34]. This result states that the category of inverse semigroups is isomorphic to the category of inductive groupoids, i.e. the ordered groupoids whose object sets are semilattices. The main significance of this result is that one can capture the entire semigroup by limited knowledge of the product (the groupoid only ‘remembers’ abab when R(a)=D(b)R(a)=D(b)) and the natural partial order (aba=D(a)ba=bR(a){a\leq b\ \Leftrightarrow\ a=D(a)b\ \Leftrightarrow\ a=bR(a)}). The ESN Theorem has far-reaching consequences and applications, for example to representation theory [47, 49, 48] and C*-algebras [36, 45, 31].

inverseregular *-restrictionDRC-restrictionEhresmann*-regularDRC
Figure 1: Certain (sub)classes of DRC-semigroups, and the containments among them.

One of the first generalisations of the ESN Theorem, to the important class of Ehresmann semigroups and categories, was due to Lawson himself [33]. These include many non-inverse semigroups, such as monoids of binary relations (which are not even regular) and partition monoids [14, 39]. See also [37] for a recent alternative approach, and [58, 35, 7, 17, 22, 16, 1, 56, 55, 41, 26, 21, 20, 25, 13, 57, 18, 51, 53, 52] for other generalisations. Of course the main challenge in obtaining an ‘ESN-type’ theorem for a class of semigroups is to identify and axiomatise the appropriate categorical structures used in the representation. A major part of this is to ascertain the precise structure to place on the object sets, which play the role of the semilattices of Lawson’s inductive groupoids. Typically these are (proper) subsets of the idempotents of the semigroup in question, but they need not be commutative or even closed under multiplication, and in particular they need not be semilattices.

Recent work of the first and third authors [13] represents regular *-semigroups by certain ordered groupoids whose object sets are the projection algebras of Imaoka [28]. These are special cases of Jones’ projection algebras, defined in [30] to obtain transformation representations of fundamental DRC-semigroups. These were subsequently used by Wang in [57] to obtain an ESN-type theorem for DRC-semigroups, although he did not use categories for his representation, but rather generalisations in which many more products/compositions are required to exist.

The main purpose of the current work is to extend the approaches of [13] and [33] to obtain a purely categorical representation for the class of DRC-semigroups. Our main result, Theorem 7.1, obtains an isomorphism between the category of DRC-semigroups and what we call chained projection categories. These are natural biordered categories whose object sets form projection algebras in the sense of Jones [30]. As an application, we establish the existence of free DRC-semigroups, thereby also extending the results of [15]. We note that special cases of these results, on so-called DRC-restriction semigroups, have been independently obtained by Die and Wang [8]. This is the class of DRC-semigroups for which our biordered categories have only a single order, and represents the limit to which the techniques of [13] generalise without significant modification.

The paper is organised as follows.

  • We begin in Section 2 with the preliminary definitions and results we need on biordered categories.

  • In Section 3 we recall the definitions and basic properties of DRC semigroups, and construct a functor 𝐃𝐑𝐂𝐁𝐂{\bf DRC}\to{\bf BC} (in Proposition 3.13), where these are the categories of DRC-semigroups and biordered categories.

  • We then turn to projection algebras in Section 4. After establishing several preliminary facts, we construct a forgetful functor 𝐏:𝐃𝐑𝐂𝐏𝐀{\bf P}:{\bf DRC}\to{\bf PA} (in Proposition 4.19) into the category of projection algebras. At the object level, this maps a DRC-semigroup SS to its underlying projection algebra 𝐏(S){\bf P}(S), as defined by Jones [30]. We also define the chain category 𝒞(P)\mathscr{C}(P) associated to a projection algebra PP (see Definition 4.28), which will play an important role in subsequent constructions.

  • In Section 5 we introduce the chained projection categories, and the category 𝐂𝐏𝐂{\bf CPC} formed by them. Such a category is in fact a triple (P,𝒞,ε)(P,\mathcal{C},\varepsilon), where 𝒞\mathcal{C} is a biordered category whose object set PP is a projection algebra (with close structural ties to 𝒞\mathcal{C}), and where ε:𝒞(P)𝒞\varepsilon:\mathscr{C}(P)\to\mathcal{C} is a certain functor from the chain category of PP. The main result here is Theorem 5.20, which constructs a functor 𝐂:𝐃𝐑𝐂𝐂𝐏𝐂{\bf C}:{\bf DRC}\to{\bf CPC}.

  • In Section 6 we then define a functor 𝐒:𝐂𝐏𝐂𝐃𝐑𝐂{\bf S}:{\bf CPC}\to{\bf DRC} in the opposite direction; see Theorem 6.7. The main work involves the construction of a DRC-semigroup 𝐒(P,𝒞,ε){\bf S}(P,\mathcal{C},\varepsilon) associated to a chained projection category (P,𝒞,ε)(P,\mathcal{C},\varepsilon); see Definition 6.1 and Theorem 6.5.

  • We then prove in Section 7 that the functors 𝐂{\bf C} and 𝐒{\bf S} are in fact mutually inverse isomorphisms between the categories 𝐃𝐑𝐂{\bf DRC} and 𝐂𝐏𝐂{\bf CPC}, thereby establishing our main result, Theorem 7.1.

  • In Section 8, we construct a left adjoint to the forgetful functor 𝐏:𝐃𝐑𝐂𝐏𝐀{\bf P}:{\bf DRC}\to{\bf PA} from Section 4, thereby establishing the existence of free projection-generated DRC-semigroups. These free semigroups are defined by presentations in terms of generators and defining relations. Among other things, this has the consequence that every (abstract) projection algebra is the algebra of projections of some DRC-semigroup, a fact that was first established by Jones in [30] by entirely different methods. Along the way, we also construct the unique (up to isomorphism) fundamental projection-generated DRC-semigroup with projection algebra PP.

  • Finally, in Sections 9 and 10 we examine the various simplifications that occur for the important special cases of regular *-semigroups [13], *-regular semigroups [12], Ehresmann semigroups [33] and DRC-restriction semigroups [8].

2 Preliminaries on biordered categories

Throughout the paper, we identify a small category 𝒞\mathcal{C} with its morphism set, and identify the objects of 𝒞\mathcal{C} with the identities, the set of which is denoted v𝒞v\mathcal{C}. Thus, we have domain and range maps 𝐝,𝐫:𝒞v𝒞{\bf d},{\bf r}:\mathcal{C}\to v\mathcal{C}, for which the following hold, for all a,b,c𝒞a,b,c\in\mathcal{C}:

  • aba\circ b exists if and only if 𝐫(a)=𝐝(b){\bf r}(a)={\bf d}(b), in which case 𝐝(ab)=𝐝(a){\bf d}(a\circ b)={\bf d}(a) and 𝐫(ab)=𝐫(b){\bf r}(a\circ b)={\bf r}(b). (So we are composing morphisms ‘left to right’.)

  • 𝐝(a)a=a=a𝐫(a){\bf d}(a)\circ a=a=a\circ{\bf r}(a).

  • If aba\circ b and bcb\circ c exist, then (ab)c=a(bc)(a\circ b)\circ c=a\circ(b\circ c).

  • 𝐝(p)=𝐫(p)=p{\bf d}(p)={\bf r}(p)=p for all pv𝒞p\in v\mathcal{C}.

A left-ordered category is a pair (𝒞,)(\mathcal{C},\leq), where 𝒞\mathcal{C} is a small category and \leq a partial order on 𝒞\mathcal{C} satisfying the following, for all a,b,c,d𝒞a,b,c,d\in\mathcal{C}:

  • If aba\leq b, then 𝐝(a)𝐝(b){\bf d}(a)\leq{\bf d}(b) and 𝐫(a)𝐫(b){\bf r}(a)\leq{\bf r}(b).

  • If aba\leq b and cdc\leq d, and if 𝐫(a)=𝐝(c){\bf r}(a)={\bf d}(c) and 𝐫(b)=𝐝(d){\bf r}(b)={\bf d}(d), then acbda\circ c\leq b\circ d.

  • For any pv𝒞p\in v\mathcal{C} with p𝐝(a)p\leq{\bf d}(a), there exists a unique uau\leq a with 𝐝(u)=p{\bf d}(u)=p. This element is denoted u=pau={}_{p}{\downharpoonleft}a, and called the left restriction of aa to pp.

(These are the Ω\Omega-structured categories with restrictions, in the terminology of [33].) Right-ordered categories are defined analogously, where the third item above is replaced by:

  • For any qv𝒞q\in v\mathcal{C} with q𝐫(a)q\leq{\bf r}(a), there exists a unique vav\leq a with 𝐝(v)=q{\bf d}(v)=q. This element is denoted v=aqv=a{\downharpoonright}_{q}, and called the right restriction of aa to qq.

A biordered category is a triple (𝒞,,r)(\mathcal{C},\leq_{\ell},\leq_{r}), where:

  • (𝒞,)(\mathcal{C},\leq_{\ell}) is a left-ordered category, and (𝒞,r)(\mathcal{C},\leq_{r}) a right-ordered category, and

  • \leq_{\ell} and r\leq_{r} restrict to the same order on v𝒞v\mathcal{C}, meaning that pqprqp\leq_{\ell}q\ \Leftrightarrow\ p\leq_{r}q for all p,qv𝒞p,q\in v\mathcal{C}.

(The Ehresmann categories considered in [33] are certainly biordered categories in the above sense.) Unless there is a chance of confusion, we typically speak of ‘a biordered category 𝒞\mathcal{C}’, and assume the orders are named \leq_{\ell} and r\leq_{r}. It is important to note that left and right restrictions in a biordered category are quite independent of each other. For example, if 𝐫(pa)=q{\bf r}({}_{p}{\downharpoonleft}a)=q, then we need not have pa=aq{}_{p}{\downharpoonleft}a=a{\downharpoonright}_{q}; in fact, aqa{\downharpoonright}_{q} need not even have domain pp. (Some concrete examples are discussed in Section 9.)

We write 𝐁𝐂{\bf BC} for the (large) category of biordered categories. Morphisms in 𝐁𝐂{\bf BC} are the biordered morphisms. Such a morphism (𝒞,,r)(𝒞,,r)(\mathcal{C},\leq_{\ell},\leq_{r})\to(\mathcal{C}^{\prime},\leq_{\ell}^{\prime},\leq_{r}^{\prime}) is a functor ϕ:𝒞𝒞\phi:\mathcal{C}\to\mathcal{C}^{\prime} preserving both orders, in the sense that

abaϕbϕandarbaϕrbϕfor all a,b𝒞.a\leq_{\ell}b\ \Rightarrow\ a\phi\leq_{\ell}^{\prime}b\phi\qquad\text{and}\qquad a\leq_{r}b\ \Rightarrow\ a\phi\leq_{r}^{\prime}b\phi\qquad\text{for all $a,b\in\mathcal{C}$.} (2.1)

Although a biordered category need not be particularly symmetric in its own right, there is a certain symmetry/duality in the category 𝐁𝐂{\bf BC} itself. Specifically, we have an isomorphism 𝐎:𝐁𝐂𝐁𝐂{\bf O}:{\bf BC}\to{\bf BC}, given at the object level by 𝐎(𝒞,l,r)=(𝒞op,r,l){\bf O}(\mathcal{C},\leq_{l},\leq_{r})=(\mathcal{C}^{\operatorname{op}},\leq_{r},\leq_{l}). Here 𝒞op\mathcal{C}^{\operatorname{op}} denotes the opposite category to 𝒞\mathcal{C}, in which domains and codomains are swapped, and composition is reversed. (At the morphism level we simply have 𝐎(ϕ)=ϕ{\bf O}(\phi)=\phi.)

Throughout the paper, we will typically construct biordered categories by first defining an ordering on objects, and then defining suitable restrictions, as formalised in the next result. We omit the proof, as it is directly analogous to that of [13, Lemma 2.3].

Lemma 2.2.

Suppose 𝒞\mathcal{C} is a small category for which the following two conditions hold:

  1. (i)

    There is a partial order \leq on the object set v𝒞v\mathcal{C}.

  2. (ii)

    For all a𝒞a\in\mathcal{C}, and for all p𝐝(a)p\leq{\bf d}(a) and q𝐫(a)q\leq{\bf r}(a), there exist morphisms pa{}_{p}{\downharpoonleft}a and aqa{\downharpoonright}_{q} in 𝒞\mathcal{C} such that the following hold, for all a,b𝒞a,b\in\mathcal{C} and p,q,r,sv𝒞p,q,r,s\in v\mathcal{C}:

    1. (O1)

      If p𝐝(a)p\leq{\bf d}(a), then 𝐝(pa)=p{\bf d}({}_{p}{\downharpoonleft}a)=p and 𝐫(pa)𝐫(a){\bf r}({}_{p}{\downharpoonleft}a)\leq{\bf r}(a).

    2. If q𝐫(a)q\leq{\bf r}(a), then 𝐫(aq)=q{\bf r}(a{\downharpoonright}_{q})=q and 𝐝(aq)𝐝(a){\bf d}(a{\downharpoonright}_{q})\leq{\bf d}(a).

    3. (O2)

      𝐝(a)a=a=a𝐫(a){}_{{\bf d}(a)}{\downharpoonleft}a=a=a{\downharpoonright}_{{\bf r}(a)}.

    4. (O3)

      If pq𝐝(a)p\leq q\leq{\bf d}(a), then pqa=pa{}_{p}{\downharpoonleft}{}_{q}{\downharpoonleft}a={}_{p}{\downharpoonleft}a.

    5. If rs𝐫(a)r\leq s\leq{\bf r}(a), then asr=ara{\downharpoonright}_{s}{\downharpoonright}_{r}=a{\downharpoonright}_{r}.

    6. (O4)

      If p𝐝(a)p\leq{\bf d}(a) and 𝐫(a)=𝐝(b){\bf r}(a)={\bf d}(b), and if q=𝐫(pa)q={\bf r}({}_{p}{\downharpoonleft}a), then p(ab)=paqb{}_{p}{\downharpoonleft}(a\circ b)={}_{p}{\downharpoonleft}a\circ{}_{q}{\downharpoonleft}b.

    7. If r𝐫(b)r\leq{\bf r}(b) and 𝐫(a)=𝐝(b){\bf r}(a)={\bf d}(b), and if s=𝐝(br)s={\bf d}(b{\downharpoonright}_{r}), then (ab)r=asbr(a\circ b){\downharpoonright}_{r}=a{\downharpoonright}_{s}\circ b{\downharpoonright}_{r}.

Then (𝒞,,r)(\mathcal{C},\leq_{\ell},\leq_{r}) is a biordered category with orders given by

aba=pbfor some p𝐝(b)a=𝐝(a)b,arba=bqfor some q𝐫(b)a=b𝐫(a).\begin{array}[]{ccccc}a\leq_{\ell}b&\qquad\Leftrightarrow&\qquad a={}_{p}{\downharpoonleft}b\ \ \text{for some $p\leq{\bf d}(b)$}&\qquad\Leftrightarrow&\qquad a={}_{{\bf d}(a)}{\downharpoonleft}b,\\[5.69054pt] a\leq_{r}b&\qquad\Leftrightarrow&\qquad a=b{\downharpoonright}_{q}\ \ \text{for some $q\leq{\bf r}(b)$}&\qquad\Leftrightarrow&\qquad a=b{\downharpoonright}_{{\bf r}(a)}.\end{array} (2.3)

Moreover, any biordered category has the above form. ∎

For biordered categories 𝒞\mathcal{C} and 𝒞\mathcal{C}^{\prime}, it is easy to check that a functor ϕ:𝒞𝒞\phi:\mathcal{C}\to\mathcal{C}^{\prime} is a biordered morphism, i.e. satisfies (2.1), if and only if it satisfies

(pa)ϕ=pϕaϕand(aq)ϕ=aϕqϕfor all a𝒞, and all p𝐝(a) and q𝐫(a).({}_{p}{\downharpoonleft}a)\phi={}_{p\phi}{\downharpoonleft}a\phi\qquad\text{and}\qquad(a{\downharpoonright}_{q})\phi=a\phi{\downharpoonright}_{q\phi}\qquad\text{for all $a\in\mathcal{C}$, and all $p\leq{\bf d}(a)$ and $q\leq{\bf r}(a)$.} (2.4)

Here, and henceforth, we denote by \leq the common restriction of \leq_{\ell} and r\leq_{r} to v𝒞v\mathcal{C}.

Consider a biordered category 𝒞\mathcal{C}, and let a𝒞a\in\mathcal{C}. We then have two maps

ϑa:𝐝(a)𝐫(a)anda:𝐫(a)𝐝(a),\vartheta_{a}:{\bf d}(a)^{\downarrow}\to{\bf r}(a)^{\downarrow}\qquad\text{and}\qquad\partial_{a}:{\bf r}(a)^{\downarrow}\to{\bf d}(a)^{\downarrow},

given by

pϑa=𝐫(pa)andqa=𝐝(aq)for p𝐝(a) and q𝐫(a).p\vartheta_{a}={\bf r}({}_{p}{\downharpoonleft}a)\qquad\text{and}\qquad q\partial_{a}={\bf d}(a{\downharpoonright}_{q})\qquad\text{for $p\leq{\bf d}(a)$ and $q\leq{\bf r}(a)$.} (2.5)

Here for a partially ordered set (P,)(P,\leq), we write t={sP:st}t^{\downarrow}=\{s\in P:s\leq t\} for the down-set of tPt\in P. It follows quickly from (ii)(O3) that

sϑpa=sϑaandtaq=tafor all a𝒞, and all sp𝐝(a) and tq𝐫(a).s\vartheta_{{}_{p}{\downharpoonleft}a}=s\vartheta_{a}\qquad\text{and}\qquad t\partial_{a{\downharpoonright}_{q}}=t\partial_{a}\qquad\text{for all $a\in\mathcal{C}$, and all $s\leq p\leq{\bf d}(a)$ and $t\leq q\leq{\bf r}(a)$.} (2.6)

A vv-congruence on a small category 𝒞\mathcal{C} is an equivalence relation \approx on 𝒞\mathcal{C} satisfying the following, for all a,b,u,v𝒞a,b,u,v\in\mathcal{C}:

  • ab[𝐝(a)=𝐝(b)a\approx b\ \Rightarrow\ [{\bf d}(a)={\bf d}(b) and 𝐫(a)=𝐫(b)]{\bf r}(a)={\bf r}(b)],

  • abuauba\approx b\ \Rightarrow\ u\circ a\approx u\circ b, whenever the stated compositions are defined,

  • abavbva\approx b\ \Rightarrow\ a\circ v\approx b\circ v, whenever the stated compositions are defined.

If 𝒞\mathcal{C} is a biordered category, we say that the vv-congruence \approx is a biordered congruence if it additionally satisfies the following, for all a,b𝒞a,b\in\mathcal{C}:

  • ab[papba\approx b\ \Rightarrow\ [{}_{p}{\downharpoonleft}a\approx{}_{p}{\downharpoonleft}b and aqbqa{\downharpoonright}_{q}\approx b{\downharpoonright}_{q}] for all p𝐝(a)p\leq{\bf d}(a) and q𝐫(a)q\leq{\bf r}(a).

In this case the quotient 𝒞/\mathcal{C}/{\approx} is a biordered category with orders given by

αβ\displaystyle\alpha\leq_{\ell}\beta abfor some aα and bβ,\displaystyle\qquad\Leftrightarrow\qquad a\leq_{\ell}b\ \ \text{for some $a\in\alpha$ and $b\in\beta$,}
γrδ\displaystyle\gamma\leq_{r}\delta crdfor some cγ and dδ.\displaystyle\qquad\Leftrightarrow\qquad c\leq_{r}d\ \ \text{for some $c\in\gamma$ and $d\in\delta$.}

3 DRC-semigroups

We now come to the main object of our study, the class of DRC-semigroups. We begin in Section 3.1 by recalling the definitions, and listing some basic properties; while many of these preliminary lemmas exist in the literature (see for example [30, 57, 50]), we give brief proofs to keep the paper self-contained. In Section 3.2 we show how to construct a biordered category from a DRC-semigroup, leading to a functor 𝐃𝐑𝐂𝐁𝐂{\bf DRC}\to{\bf BC}.

3.1 Definitions and basic properties

Definition 3.1.

A DRC-semigroup is an algebra (S,,D,R)(S,\cdot,D,R), where (S,)(S,\cdot) is a semigroup, and where DD and RR are unary operations SSS\to S satisfying the following laws, for all a,bSa,b\in S:

  1. (DRC1)

    D(a)a=aD(a)a=a,

  2. (DRC2)

    D(ab)=D(aD(b))D(ab)=D(aD(b)),

  3. (DRC3)

    D(ab)=D(a)D(ab)D(a)D(ab)=D(a)D(ab)D(a),

  4. (DRC4)

    R(D(a))=D(a)R(D(a))=D(a),

  5. aR(a)=aaR(a)=a,

  6. R(ab)=R(R(a)b)R(ab)=R(R(a)b),

  7. R(ab)=R(b)R(ab)R(b)R(ab)=R(b)R(ab)R(b),

  8. D(R(a))=R(a)D(R(a))=R(a).

Throughout the article, we will almost always identify an algebra with its underlying set, including for a DRC-semigroup S(S,,D,R)S\equiv(S,\cdot,D,R). Thus, we typically speak of ‘a DRC-semigroup SS’, write multiplication as juxtaposition, and assume the unary operations are denoted DD and RR.

We denote by 𝐃𝐑𝐂{\bf DRC} the category of all DRC-semigroups, with DRC-morphisms. Such a morphism (S,,D,R)(S,,D,R)(S,\cdot,D,R)\to(S^{\prime},\cdot^{\prime},D^{\prime},R^{\prime}) in 𝐃𝐑𝐂{\bf DRC} is a function ϕ:SS\phi:S\to S^{\prime} preserving all the operations, meaning that

(ab)ϕ=(aϕ)(bϕ),D(a)ϕ=D(aϕ)andR(a)ϕ=R(aϕ)for all a,bS.(a\cdot b)\phi=(a\phi)\cdot^{\prime}(b\phi),\qquad D(a)\phi=D^{\prime}(a\phi)\qquad\text{and}\qquad R(a)\phi=R^{\prime}(a\phi)\qquad\text{for all $a,b\in S$.}

As noted in the introduction, Lawson first introduced DRC-semigroups in [33], where he referred to them as reduced UU-semiabundant semigroups satisfying the congruence conditions. (The laws in axiom (DRC2) are typically called the congruence conditions as, for example, the identity D(ab)=D(aD(b))D(ab)=D(aD(b)) is equivalent, in the presence of the other laws, to the relation {(a,b)S×S:D(a)=D(b)}\{(a,b)\in S\times S:D(a)=D(b)\} being a left congruence on SS.) Lawson was mainly interested in Ehresmann semigroups, which additionally satisfy the law D(a)D(b)=D(b)D(a){D(a)D(b)=D(b)D(a)}, and hence also R(a)R(b)=R(b)R(a)R(a)R(b)=R(b)R(a), which makes (DRC3) redundant in the presence of the other laws.

As with 𝐁𝐂{\bf BC}, the category 𝐃𝐑𝐂{\bf DRC} posesses a natural symmetry/duality. Specifically, if (S,,D,R)(S,\cdot,D,R) is a DRC-semigroup, and if (S,)(S,\star) is the opposite semigroup of SS, where ab=baa\star b=b\cdot a for a,bSa,b\in S, then (S,,R,D)(S,\star,R,D) is a DRC-semigroup. Thus, any equality that holds in (S,,D,R)(S,\cdot,D,R) can be converted to a dual equality by reversing the order of all products, and interchanging DD and RR.

Remark 3.2.

Some authors use superscripts for the unary operations on a DRC-semigroup. For example, Jones [30] and Wang [57] write a+=D(a)a^{+}=D(a) and a=R(a)a^{*}=R(a). Here we prefer the functional notation, as it emphasises the connection with domains and ranges in the corresponding categories.

Note that (DRC4) says RD=DRD=D and DR=RDR=R. It follows from these that

D=RD=(DR)D=D(RD)=DD,and similarlyR=RR.D=RD=(DR)D=D(RD)=DD,\qquad\text{and similarly}\qquad R=RR.

That is, we have the following consequence of (DRC1)(DRC4):

  1. (DRC5)

    D(D(a))=D(a)D(D(a))=D(a),

  2. R(R(a))=R(a)R(R(a))=R(a).

In calculations to follow, we use =1=_{1} to indicate equality in a DRC-semigroup by one or more applications of (DRC1), and similarly for (DRC2)(DRC5).

It follows from (DRC4) that im(D)=im(R)\operatorname{im}(D)=\operatorname{im}(R). We denote this common image by 𝐏(S){\bf P}(S), so

𝐏(S)={D(a):aS}={R(a):aS}.{\bf P}(S)=\{D(a):a\in S\}=\{R(a):a\in S\}.

The elements of 𝐏(S){\bf P}(S) are called projections. We will typically use the next result without explicit reference.

Lemma 3.3 (cf. [50, Proposition 1.2]).

We have 𝐏(S)={pS:p2=p=D(p)=R(p)}{\bf P}(S)=\{p\in S:p^{2}=p=D(p)=R(p)\}.

Proof.

Certainly every element of the stated set belongs to im(D)=𝐏(S)\operatorname{im}(D)={\bf P}(S). Conversely, let p𝐏(S)p\in{\bf P}(S), so that p=D(a)p=D(a) for some aSa\in S. Then p=D(a)=5D(D(a))=D(p)p=D(a)=_{5}D(D(a))=D(p), and similarly p=R(p)p=R(p). From p=D(p)p=D(p), it also follows that p2=D(p)p=1pp^{2}=D(p)p=_{1}p. ∎

Since every projection is an idempotent, it follows that (DRC3) is equivalent (in the presence of the other axioms) to:

  1. (DRC3)

    D(ab)=D(a)D(ab)=D(ab)D(a)D(ab)=D(a)D(ab)=D(ab)D(a),     R(ab)=R(b)R(ab)=R(ab)R(b)R(ab)=R(b)R(ab)=R(ab)R(b).

Lemma 3.4.

For any aSa\in S and p𝐏(S)p\in{\bf P}(S) we have

D(pa)=pD(pa)=D(pa)pandR(ap)=pR(ap)=R(ap)p.D(pa)=pD(pa)=D(pa)p\qquad\text{and}\qquad R(ap)=pR(ap)=R(ap)p.
Proof.

The first follows from (DRC3) and p=D(p)p=D(p); the second is dual. ∎

Since projections are idempotents, the natural partial order on idempotents restricts to a partial order on 𝐏(S){\bf P}(S):

pqp=qpqp=pq=qpfor p,q𝐏(S).p\leq q\quad\Leftrightarrow\quad p=qpq\quad\Leftrightarrow\quad p=pq=qp\qquad\text{for $p,q\in{\bf P}(S)$.} (3.5)

The next result shows that this is equivalent to either of p=pqp=pq or p=qpp=qp. This is referred to as the reduced property in the literature [50].

Corollary 3.6.

For p,q𝐏(S)p,q\in{\bf P}(S), we have pqp=pqp=qpp\leq q\ \Leftrightarrow\ p=pq\ \Leftrightarrow\ p=qp.

Proof.

In light of (3.5), and by symmetry, it suffices to check that p=pqp=qpp=pq\ \Rightarrow\ p=qp. But if p=pqp=pq, then using Lemma 3.4 we have p=R(p)=R(pq)=qR(pq)=qR(p)=qpp=R(p)=R(pq)=qR(pq)=qR(p)=qp. ∎

Remark 3.7.

An equivalent definition of DRC-semigroups stems from this order-theoretic perspective. Thus, a DR-semigroup is defined in [50] to be a semigroup SS in which a (sub)set of idempotents PP is distinguished such that for each aSa\in S there exist smallest p,qPp,q\in P under the natural partial order for which a=pa=aqa=pa=aq, called p=D(a)p=D(a) and q=R(a)q=R(a). This defines unary operations DD and RR on such SS (both with image PP), and it follows that DRC-semigroups are precisely the DR-semigroups satisfying the congruence conditions (DRC2).

The set 𝐏(S){\bf P}(S) can be given the structure of a projection algebra, as defined by Jones in [30]. We will recall the formal definition in Section 4, but the next result will be used to verify the axioms. For p𝐏(S)p\in{\bf P}(S) we define the functions θp,δp:𝐏(S)𝐏(S)\theta_{p},\delta_{p}:{\bf P}(S)\to{\bf P}(S) by

qθp=R(qp)andqδp=D(pq)for q𝐏(S).q\theta_{p}=R(qp)\qquad\text{and}\qquad q\delta_{p}=D(pq)\qquad\text{for $q\in{\bf P}(S)$.}
Lemma 3.8 (cf. [30, Proposition 7.2]).

For any p,q,𝐏(S)p,q,\in{\bf P}(S), we have

  1. (i)

    pθp=pp\theta_{p}=p,   pθqθp=qθpp\theta_{q\theta_{p}}=q\theta_{p},   θqθqθp=θqθp\theta_{q}\theta_{q\theta_{p}}=\theta_{q}\theta_{p},   θpδp=θp\theta_{p}\delta_{p}=\theta_{p},   θpδqθp=θqθp\theta_{p\delta_{q}}\theta_{p}=\theta_{q}\theta_{p},

  2. (ii)

    pδp=pp\delta_{p}=p,   pδqδp=qδpp\delta_{q\delta_{p}}=q\delta_{p},   δqδqδp=δqδp\delta_{q}\delta_{q\delta_{p}}=\delta_{q}\delta_{p},   δpθp=δp\delta_{p}\theta_{p}=\delta_{p},   δpθqδp=δqδp\delta_{p\theta_{q}}\delta_{p}=\delta_{q}\delta_{p}.

Proof.

We just prove (i), as (ii) is dual. In what follows, we write ==_{*} to indicate an application of Lemma 3.4 (and continue to use =1=_{1}, =2=_{2}, etc.). For the first item we have pθp=R(pp)=R(p)=pp\theta_{p}=R(pp)=R(p)=p, and for the second,

pθqθp=R(pR(qp))=R(R(qp))=5R(qp)=qθp.p\theta_{q\theta_{p}}=R(pR(qp))=_{*}R(R(qp))=_{5}R(qp)=q\theta_{p}.

For the remaining three items, fix tPt\in P, and first note that tθqθp=R(R(tq)p)=2R(tqp)t\theta_{q}\theta_{p}=R(R(tq)p)=_{2}R(tqp). The third and fifth items follow by combining this with

tθqθqθp=R(R(tq)R(qp))=2R(tqR(qp))=R(tqpR(qp))=1R(tqp)t\theta_{q}\theta_{q\theta_{p}}=R(R(tq)R(qp))=_{2}R(tqR(qp))=_{*}R(tqpR(qp))=_{1}R(tqp)

and

tθpδqθp=R(R(tD(qp))p)=2R(tD(qp)p)=R(tD(qp)qp)=1R(tqp).t\theta_{p\delta_{q}}\theta_{p}=R(R(tD(qp))p)=_{2}R(tD(qp)p)=_{*}R(tD(qp)qp)=_{1}R(tqp).

For the fourth we have tθpδp=D(pR(tp))=D(R(tp))=4R(tp)=tθpt\theta_{p}\delta_{p}=D(pR(tp))=_{*}D(R(tp))=_{4}R(tp)=t\theta_{p}. ∎

3.2 From DRC-semigroups to biordered categories

We now wish to show that a DRC-semigroup SS naturally induces a biordered category 𝒞(S)\mathcal{C}(S) with morphism set SS, and object/identity set 𝐏(S){\bf P}(S). For this we need the following.

Lemma 3.9.

If SS is a DRC-semigroup, and if a,bSa,b\in S are such that R(a)=D(b)R(a)=D(b), then D(ab)=D(a)D(ab)=D(a) and R(ab)=R(b)R(ab)=R(b).

Proof.

We have D(ab)=2D(aD(b))=D(aR(a))=1D(a)D(ab)=_{2}D(aD(b))=D(aR(a))=_{1}D(a); the other is dual. ∎

As a consequence, we can make the following definition.

Definition 3.10.

Given a DRC-semigroup SS, we define the (small) category 𝒞=𝒞(S)\mathcal{C}=\mathcal{C}(S) as follows:

  • The underlying (morphism) set of 𝒞\mathcal{C} is SS.

  • The object/identity set is v𝒞=P=𝐏(S)v\mathcal{C}=P={\bf P}(S).

  • For a𝒞a\in\mathcal{C} we have 𝐝(a)=D(a){\bf d}(a)=D(a) and 𝐫(a)=R(a){\bf r}(a)=R(a).

  • For a,b𝒞a,b\in\mathcal{C} with 𝐫(a)=𝐝(b){\bf r}(a)={\bf d}(b) we have ab=aba\circ b=ab.

For a𝒞a\in\mathcal{C}, and for p𝐝(a)p\leq{\bf d}(a) and q𝐫(a)q\leq{\bf r}(a) we define pa=pa{}_{p}{\downharpoonleft}a=pa and aq=aqa{\downharpoonright}_{q}=aq. (Here \leq is the partial order on PP in (3.5).)

Proposition 3.11.

For any DRC-semigroup SS, the category 𝒞=𝒞(S)\mathcal{C}=\mathcal{C}(S) is biordered, with \leq_{\ell} and r\leq_{r} as in (2.3).

Proof.

We just need to check conditions (ii)(O1)(ii)(O4) from Lemma 2.2. By symmetry, we just need to check the properties of left restrictions.

(ii)(O1). For p𝐝(a)=D(a)p\leq{\bf d}(a)=D(a) we have p=pD(a)p=pD(a), and then

𝐝(pa)=D(pa)=2D(pD(a))=D(p)=p.{\bf d}({}_{p}{\downharpoonleft}a)=D(pa)=_{2}D(pD(a))=D(p)=p.

We also have R(pa)=3R(a)R(pa)R(a)R(pa)=_{3}R(a)R(pa)R(a), which says that 𝐫(pa)=R(pa)R(a)=𝐫(a){\bf r}({}_{p}{\downharpoonleft}a)=R(pa)\leq R(a)={\bf r}(a).

(ii)(O2). We have 𝐝(a)a=D(a)a=1a{}_{{\bf d}(a)}{\downharpoonleft}a=D(a)a=_{1}a.

(ii)(O3). From pqp\leq q we have p=pqp=pq, and so pqa=p(qa)=pa=pa{}_{p}{\downharpoonleft}{}_{q}{\downharpoonleft}a=p(qa)=pa={}_{p}{\downharpoonleft}a.

(ii)(O4). Here we have q=𝐫(pa)=R(pa)q={\bf r}({}_{p}{\downharpoonleft}a)=R(pa), and also q=𝐝(qb)=𝐝(qb)q={\bf d}({}_{q}{\downharpoonleft}b)={\bf d}(qb), so

p(ab)=pab=1paR(pa)b=paqb=paqb=paqb.{}_{p}{\downharpoonleft}(a\circ b)=pab=_{1}paR(pa)b=paqb=pa\circ qb={}_{p}{\downharpoonleft}a\circ{}_{q}{\downharpoonleft}b.\qed
Remark 3.12.

Note that our orders \leq_{\ell} and r\leq_{r} were denoted r\leq_{r} and \leq_{\ell} (the other way round) by Lawson in the context of Ehresmann semigroups [33]. We have chosen the current terminology, as our \leq_{\ell} is contained in Green’s \leq_{\mathrel{\mathscr{L}}} preorder, which is defined by abaS1ba\leq_{\mathrel{\mathscr{L}}}b\ \Leftrightarrow\ a\in S^{1}b. An analogous statement applies to r\leq_{r} and \leq_{\mathrel{\mathscr{R}}}. These containments are strict in general. For example, if MM is a DRC-monoid with identity 11, then every element is \leq_{\mathrel{\mathscr{L}}} and \leq_{\mathrel{\mathscr{R}}} below 11, but only the projections are \leq_{\ell} or r\leq_{r} below 11.

The construction of 𝒞(S)\mathcal{C}(S) from SS is an object map from 𝐃𝐑𝐂{\bf DRC} to 𝐁𝐂{\bf BC}. For a DRC-morphism ϕ\phi, we define 𝒞(ϕ)=ϕ\mathcal{C}(\phi)=\phi. (Since the underlying set of 𝒞(S)\mathcal{C}(S) is just SS, this is well defined.)

Proposition 3.13.

𝒞\mathcal{C} is a functor 𝐃𝐑𝐂𝐁𝐂{\bf DRC}\to{\bf BC}.

Proof.

It remains to check that any DRC-morphism ϕ:SS\phi:S\to S^{\prime} is a biordered morphism 𝒞𝒞\mathcal{C}\to\mathcal{C}^{\prime}, where 𝒞=𝒞(S)\mathcal{C}=\mathcal{C}(S) and 𝒞=𝒞(S)\mathcal{C}^{\prime}=\mathcal{C}(S^{\prime}). This amounts to checking (cf. (2.4)) that

  1. (i)

    (ab)ϕ=(aϕ)(bϕ)(a\circ b)\phi=(a\phi)\circ^{\prime}(b\phi) for all a,b𝒞a,b\in\mathcal{C} with 𝐫(a)=𝐝(b){\bf r}(a)={\bf d}(b),

  2. (ii)

    (pa)ϕ=pϕaϕ({}_{p}{\downharpoonleft}a)\phi={}_{p\phi}{\downharpoonleft}a\phi for all a𝒞a\in\mathcal{C} and p𝐝(a)p\leq{\bf d}(a), and

  3. (iii)

    (aq)ϕ=aϕqϕ(a{\downharpoonright}_{q})\phi=a\phi{\downharpoonright}_{q\phi} for all a𝒞a\in\mathcal{C} and q𝐫(a)q\leq{\bf r}(a).

Here, we use dashes to distinguish the parameters and operations associated to 𝒞\mathcal{C} and 𝒞\mathcal{C}^{\prime}. We will, however, denote multiplication in both SS and SS^{\prime} by juxtaposition. Next we note that for any a𝒞a\in\mathcal{C},

𝐝(aϕ)=D(aϕ)=D(a)ϕ=𝐝(a)ϕ,and similarly𝐫(aϕ)=𝐫(a)ϕ.{\bf d}^{\prime}(a\phi)=D^{\prime}(a\phi)=D(a)\phi={\bf d}(a)\phi,\qquad\text{and similarly}\qquad{\bf r}^{\prime}(a\phi)={\bf r}(a)\phi.

(i). If 𝐫(a)=𝐝(b){\bf r}(a)={\bf d}(b), then 𝐫(aϕ)=𝐫(a)ϕ=𝐝(b)ϕ=𝐝(bϕ){\bf r}^{\prime}(a\phi)={\bf r}(a)\phi={\bf d}(b)\phi={\bf d}^{\prime}(b\phi), and so

(ab)ϕ=(ab)ϕ=(aϕ)(bϕ)=(aϕ)(bϕ).(a\circ b)\phi=(ab)\phi=(a\phi)(b\phi)=(a\phi)\circ^{\prime}(b\phi).

(ii). If pqp\leq q, where for simplicity we write q=𝐝(a)q={\bf d}(a), then p=pqp=pq, and so pϕ=(pϕ)(qϕ)p\phi=(p\phi)(q\phi). This gives pϕqϕ=𝐝(aϕ)p\phi\leq q\phi={\bf d}^{\prime}(a\phi), and it follows that (pa)ϕ=(pa)ϕ=(pϕ)(aϕ)=pϕaϕ({}_{p}{\downharpoonleft}a)\phi=(pa)\phi=(p\phi)(a\phi)={}_{p\phi}{\upharpoonleft}a\phi.

(iii). This is dual to (ii). ∎

4 Projection algebras

The set 𝐏(S){\bf P}(S) of projections of a DRC-semigroup SS need not be a subsemigroup of SS. Nevertheless, it can be given the structure of a projection algebra, as axiomatised by Jones [30]. These algebras will play a crucial role in all that follows, and here we gather the information we need. In Section 4.1 we recall Jones’ definition, albeit in an alternative format, and prove some preliminary results (some of which can again be found in the existing literature). In Section 4.2 we consider the category 𝐏𝐀{\bf PA} of projection algebras, and show how the construction of 𝐏(S){\bf P}(S) from a DRC-semigroup SS leads to a functor 𝐃𝐑𝐂𝐏𝐀{\bf DRC}\to{\bf PA}. In Section 4.3 we show how a projection algebra PP induces two biordered categories, the path category 𝒫(P)\mathscr{P}(P) and the chain category 𝒞(P)\mathscr{C}(P). These will be crucial ingredients in constructions of subsequent sections.

4.1 Definitions and basic properties

Definition 4.1.

A projection algebra is an algebra (P,θ,δ)(P,\theta,\delta), where θ={θp:pP}\theta=\{\theta_{p}:p\in P\} and δ={δp:pP}\delta=\{\delta_{p}:p\in P\} are families of unary operations satisfying the following laws, for all p,qPp,q\in P:

  1. (P1)

    pθp=pp\theta_{p}=p,

  2. (P2)

    pθqθp=qθpp\theta_{q\theta_{p}}=q\theta_{p},

  3. (P3)

    θqθqθp=θqθp\theta_{q}\theta_{q\theta_{p}}=\theta_{q}\theta_{p},

  4. (P4)

    θpδp=θp\theta_{p}\delta_{p}=\theta_{p},

  5. (P5)

    θpδqθp=θqθp\theta_{p\delta_{q}}\theta_{p}=\theta_{q}\theta_{p},

  6. pδp=pp\delta_{p}=p,

  7. pδqδp=qδpp\delta_{q\delta_{p}}=q\delta_{p},

  8. δqδqδp=δqδp\delta_{q}\delta_{q\delta_{p}}=\delta_{q}\delta_{p},

  9. δpθp=δp\delta_{p}\theta_{p}=\delta_{p},

  10. δpθqδp=δqδp\delta_{p\theta_{q}}\delta_{p}=\delta_{q}\delta_{p}.

The elements of a projection algebra are called projections. We typically speak of ‘a projection algebra PP’, and assume that its operations are denoted θ\theta and δ\delta.

Again there is a natural symmetry/duality, as whenever (P,θ,δ)(P,\theta,\delta) is a projection algebra, so too is (P,δ,θ)(P,\delta,\theta). As usual, this allows us to shorten proofs, by omitting the proof of any statement whose dual has already been proved. This will often be done without explicit comment.

Remark 4.2.

Jones’ original formulation of projection algebras in [30] utilised two binary operations, denoted ×\times and \star, which are related to our unary operations by:

p×q=qδpandpq=pθqfor p,qP.p\times q=q\delta_{p}\qquad\text{and}\qquad p\star q=p\theta_{q}\qquad\text{for $p,q\in P$.}

Jones’ axioms (with his labelling) are as follows:

  1. (LP1)

    e×e=ee\times e=e,

  2. (LP2)

    (e×f)×e=e×f(e\times f)\times e=e\times f,

  3. (LP3)

    g×(f×e)=(g×f)×(f×e)g\times(f\times e)=(g\times f)\times(f\times e),

  4. (LP4)

    (g×f)×e=(g×f)×(g×e)(g\times f)\times e=(g\times f)\times(g\times e),

  1. (RP1)

    ee=ee\star e=e,

  2. (RP2)

    e(fe)=fee\star(f\star e)=f\star e,

  3. (RP3)

    (ef)g=(ef)(fg)(e\star f)\star g=(e\star f)\star(f\star g),

  4. (RP4)

    e(fg)=(eg)(fg)e\star(f\star g)=(e\star g)\star(f\star g),

  1. (PA1)

    (f×e)f=f×e(f\times e)\star f=f\times e and f×(ef)=eff\times(e\star f)=e\star f,

  2. (PA2)

    (e(f×g))g=(ef)g(e\star(f\times g))\star g=(e\star f)\star g and g×((gf)×e)=g×(f×e)g\times((g\star f)\times e)=g\times(f\times e).

Our axioms (P1)(P3) correspond to Jones’ (LP1)(LP3) and (RP1)(RP3), and our (P4)(P5) are Jones’ (PA1)(PA2). For example, directly translating (RP3) to the unary setting yields eθfθg=eθfθfθge\theta_{f}\theta_{g}=e\theta_{f}\theta_{f\theta_{g}}, which is equivalent to the equality of maps, θfθg=θfθfθg\theta_{f}\theta_{g}=\theta_{f}\theta_{f\theta_{g}} (requiring one less parameter), the first part of our (P3).

Jones’ (LP4) and (RP4) are absent from our list, as they are consequences of the other axioms; see Lemma 4.5 and Remark 4.6. The reason Jones included those axioms is that (LP1)(LP4) are axioms for ‘left projection algebras’, which are used to model projections of ‘DC-semigroups’, and (LP4) cannot be deduced from (LP1)(LP3), as can be shown for example with Mace4 [40]. Similar comments apply to (RP1)(RP4).

One reason we have opted for the unary approach is due to associativity of function composition, meaning that bracketing becomes unnecessary for complex projection algebra terms. Another is to emphasise the connection with certain natural mappings ϑa\vartheta_{a}, Θa\Theta_{a}, a\partial_{a} and Δa\Delta_{a} on associated categories, which will be introduced below. For a more detailed comparison of the binary and unary approaches, we refer to [13, Remark 4.2].

Here then is the connection between DRC-semigroups and projection algebras. The following is well defined because of Lemma 3.8.

Definition 4.3 (cf. [30, Proposition 7.2]).

The projection algebra of a DRC-semigroup SS has

  • underlying set 𝐏(S)=im(D)=im(R)={pS:p2=p=D(p)=R(p)}{\bf P}(S)=\operatorname{im}(D)=\operatorname{im}(R)=\{p\in S:p^{2}=p=D(p)=R(p)\}, and

  • θ\theta and δ\delta operations defined by

qθp=R(qp)andqδp=D(pq)for p,q𝐏(S).q\theta_{p}=R(qp)\qquad\text{and}\qquad q\delta_{p}=D(pq)\qquad\text{for $p,q\in{\bf P}(S)$.}

For the rest of this section, we fix a projection algebra PP. In calculations to follow, we use =1=_{1} to indicate equality by one or more applications of (P1), and similarly for the other axioms.

Lemma 4.4.

For any p,qPp,q\in P we have

p=pθqp=qθpandp=pδqp=qδp.p=p\theta_{q}\ \Rightarrow\ p=q\theta_{p}\qquad\text{and}\qquad p=p\delta_{q}\ \Rightarrow\ p=q\delta_{p}.
Proof.

For the first implication (the second is dual), suppose p=pθqp=p\theta_{q}. Then

qθp=2pθqθp=pθqθqθp=3pθqθp=pθp=1p.q\theta_{p}=_{2}p\theta_{q\theta_{p}}=p\theta_{q}\theta_{q\theta_{p}}=_{3}p\theta_{q}\theta_{p}=p\theta_{p}=_{1}p.\qed

The next lemma gathers some fundamental consequences of (P1)(P5), which will be used so frequently as to warrant naming them (P6)(P10).

Lemma 4.5.

For any p,qPp,q\in P we have

  1. (P6)

    θpθp=θp\theta_{p}\theta_{p}=\theta_{p},

  2. (P7)

    θqθp=θpθqθp=θqθpθp\theta_{q\theta_{p}}=\theta_{p}\theta_{q\theta_{p}}=\theta_{q\theta_{p}}\theta_{p},

  3. (P8)

    qθqθp=qθpq\theta_{q\theta_{p}}=q\theta_{p},

  4. (P9)

    pδqθp=qθpp\delta_{q}\theta_{p}=q\theta_{p},

  5. (P10)

    δqθqθp=δqθp\delta_{q}\theta_{q\theta_{p}}=\delta_{q}\theta_{p},

  6. δpδp=δp\delta_{p}\delta_{p}=\delta_{p},

  7. δqδp=δpδqδp=δqδpδp\delta_{q\delta_{p}}=\delta_{p}\delta_{q\delta_{p}}=\delta_{q\delta_{p}}\delta_{p},

  8. qδqδp=qδpq\delta_{q\delta_{p}}=q\delta_{p},

  9. pθqδp=qδpp\theta_{q}\delta_{p}=q\delta_{p},

  10. θqδqδp=θqδp\theta_{q}\delta_{q\delta_{p}}=\theta_{q}\delta_{p}.

Proof.

(P6). We have θp=4θpδp=4θpδpθp=4θpθp\theta_{p}=_{4}\theta_{p}\delta_{p}=_{4}\theta_{p}\delta_{p}\theta_{p}=_{4}\theta_{p}\theta_{p}.

(P7). Using the now-established (P6), we have

θpθqθp=5θqθpδpθqθp=4θqθpθqθp=6θqθpandθqθpθp=3θqθpθqθpθp=6θqθpθqθp=6θqθp.\theta_{p}\theta_{q\theta_{p}}=_{5}\theta_{q\theta_{p}\delta_{p}}\theta_{q\theta_{p}}=_{4}\theta_{q\theta_{p}}\theta_{q\theta_{p}}=_{6}\theta_{q\theta_{p}}\qquad\text{and}\qquad\theta_{q\theta_{p}}\theta_{p}=_{3}\theta_{q\theta_{p}}\theta_{q\theta_{p}\theta_{p}}=_{6}\theta_{q\theta_{p}}\theta_{q\theta_{p}}=_{6}\theta_{q\theta_{p}}.

(P8). We have qθqθp=1qθqθqθp=3qθqθp=1qθpq\theta_{q\theta_{p}}=_{1}q\theta_{q}\theta_{q\theta_{p}}=_{3}q\theta_{q}\theta_{p}=_{1}q\theta_{p}.

(P9). Since pδq=4(pδq)θqp\delta_{q}=_{4}(p\delta_{q})\theta_{q}, it follows by Lemma 4.4 that pδq=qθpδqp\delta_{q}=q\theta_{p\delta_{q}}, and then

pδqθp=qθpδqθp=5qθqθp=1qθp.p\delta_{q}\theta_{p}=q\theta_{p\delta_{q}}\theta_{p}=_{5}q\theta_{q}\theta_{p}=_{1}q\theta_{p}.

(P10). We have δqθqθp=4δqθqθqθp=3δqθqθp=4δqθp\delta_{q}\theta_{q\theta_{p}}=_{4}\delta_{q}\theta_{q}\theta_{q\theta_{p}}=_{3}\delta_{q}\theta_{q}\theta_{p}=_{4}\delta_{q}\theta_{p}. ∎

Remark 4.6.

The θqθp=θpθqθp\theta_{q\theta_{p}}=\theta_{p}\theta_{q\theta_{p}} and δqδp=δpδqδp\delta_{q\delta_{p}}=\delta_{p}\delta_{q\delta_{p}} parts of (P7) are Jones’ axioms (RP4) and (LP4); cf. Remark 4.2.

Lemma 4.7.

For p,qPp,q\in P, the following are equivalent:

  1. (i)

    p=pθqp=p\theta_{q},

  2. (ii)

    p=rθqp=r\theta_{q} for some rPr\in P, i.e. pim(θq)p\in\operatorname{im}(\theta_{q}),

  3. (iii)

    p=pδqp=p\delta_{q},

  4. (iv)

    p=rδqp=r\delta_{q} for some rPr\in P, i.e. pim(δq)p\in\operatorname{im}(\delta_{q}),

  5. (v)

    p=pθq=pδq=qθp=qδpp=p\theta_{q}=p\delta_{q}=q\theta_{p}=q\delta_{p}.

Proof.

The following implications are clear:

(v)(i)(ii)and(v)(iii)(iv).\ref{leqP5}\ \Rightarrow\ \ref{leqP1}\ \Rightarrow\ \ref{leqP2}\qquad\text{and}\qquad\ref{leqP5}\ \Rightarrow\ \ref{leqP3}\ \Rightarrow\ \ref{leqP4}.

We will show that (ii)(i)(v)\ref{leqP2}\ \Rightarrow\ \ref{leqP1}\ \Rightarrow\ \ref{leqP5}, and then (iv)(iii)(v)\ref{leqP4}\ \Rightarrow\ \ref{leqP3}\ \Rightarrow\ \ref{leqP5} will follow by duality.

(ii)(i)\ref{leqP2}\ \Rightarrow\ \ref{leqP1}. If p=rθqp=r\theta_{q} for some rPr\in P, then p=rθq=6rθqθq=pθqp=r\theta_{q}=_{6}r\theta_{q}\theta_{q}=p\theta_{q}.

(i)(v)\ref{leqP1}\ \Rightarrow\ \ref{leqP5}. If p=pθqp=p\theta_{q}, then p=pθq=4pθqδq=pδqp=p\theta_{q}=_{4}p\theta_{q}\delta_{q}=p\delta_{q}. Lemma 4.4 then gives p=qθp=qδpp=q\theta_{p}=q\delta_{p}. ∎

Definition 4.8.

For p,qPp,q\in P we say that pqp\leq q if any (and hence all) of conditions (i)(v) of Lemma 4.7 hold. That is,

pqpim(θq)pim(δq)p=pθq=pδq=qθp=qδp.p\leq q\quad\Leftrightarrow\quad p\in\operatorname{im}(\theta_{q})\quad\Leftrightarrow\quad p\in\operatorname{im}(\delta_{q})\quad\Leftrightarrow\quad p=p\theta_{q}=p\delta_{q}=q\theta_{p}=q\delta_{p}. (4.9)
Proposition 4.10 (cf. [30, Lemma 5.2]).

\leq is a partial order.

Proof.

Reflexivity follows from (P1). For anti-symmetry, suppose pqp\leq q and qpq\leq p, so that p=pθqp=p\theta_{q} and q=qθpq=q\theta_{p}. Then p=pθq=2qθpθq=qθp=qp=p\theta_{q}=_{2}q\theta_{p\theta_{q}}=q\theta_{p}=q. For transitivity, suppose pqp\leq q and qrq\leq r, so that p=pθqp=p\theta_{q} and q=qθrq=q\theta_{r}. Then p=pθq=6pθqθq=pθqθqθr=3pθqθr=pθrp=p\theta_{q}=_{6}p\theta_{q}\theta_{q}=p\theta_{q}\theta_{q\theta_{r}}=_{3}p\theta_{q}\theta_{r}=p\theta_{r}, which gives prp\leq r. ∎

It follows from (4.9) that

im(θp)=im(δp)=p={qP:qp}for all pP.\operatorname{im}(\theta_{p})=\operatorname{im}(\delta_{p})=p^{\downarrow}=\{q\in P:q\leq p\}\qquad\text{for all $p\in P$.}

The next basic result will be used several times.

Lemma 4.11.

If pqp\leq q, then θp=θpθq=θqθp\theta_{p}=\theta_{p}\theta_{q}=\theta_{q}\theta_{p} and δp=δpδq=δqδp\delta_{p}=\delta_{p}\delta_{q}=\delta_{q}\delta_{p}.

Proof.

If pqp\leq q, then p=pθqp=p\theta_{q}, and so θp=θpθq=7θqθpθq=θqθp\theta_{p}=\theta_{p\theta_{q}}=_{7}\theta_{q}\theta_{p\theta_{q}}=\theta_{q}\theta_{p}. Essentially the same calculation gives θp=θpθq\theta_{p}=\theta_{p}\theta_{q}. ∎

In addition to the partial order \leq, a crucial role in all that follows will be played by the following relation.

Definition 4.12.

We define the relation \mathrel{\mathscr{F}} on a projection algebra PP by

pqp=qδpandq=pθqfor p,qP.p\mathrel{\mathscr{F}}q\qquad\Leftrightarrow\qquad p=q\delta_{p}\quad\text{and}\quad q=p\theta_{q}\qquad\text{for $p,q\in P$.}

Note that \mathrel{\mathscr{F}} is reflexive by (P1), but need not be symmetric or transitive.

In the next sequence of results we gather some basic facts about the relation \mathrel{\mathscr{F}}, and its interaction with the partial order \leq.

Lemma 4.13.

Let p,qPp,q\in P, and put p=qδpp^{\prime}=q\delta_{p} and q=pθqq^{\prime}=p\theta_{q}. Then:

  1. (i)

    ppp^{\prime}\leq p, qqq^{\prime}\leq q and pqp^{\prime}\mathrel{\mathscr{F}}q^{\prime},

  2. (ii)

    θpθq=θpθq\theta_{p}\theta_{q}=\theta_{p^{\prime}}\theta_{q^{\prime}} and δqδp=δqδp\delta_{q}\delta_{p}=\delta_{q^{\prime}}\delta_{p^{\prime}},

  3. (iii)

    if prqp\leq r\mathrel{\mathscr{F}}q for some rPr\in P, then p=pp^{\prime}=p,

  4. (iv)

    if psqp\mathrel{\mathscr{F}}s\geq q for some sPs\in P, then q=qq^{\prime}=q.

Proof.

(i). We obtain ppp^{\prime}\leq p and qqq^{\prime}\leq q by definition. We also have

qδp=pθqδqδp=10pθqδp=9qδp=p.q^{\prime}\delta_{p^{\prime}}=p\theta_{q}\delta_{q\delta_{p}}=_{10}p\theta_{q}\delta_{p}=_{9}q\delta_{p}=p^{\prime}.

We obtain pθq=qp^{\prime}\theta_{q^{\prime}}=q^{\prime} by symmetry, and hence pqp^{\prime}\mathrel{\mathscr{F}}q^{\prime}.

(ii). We have θpθq=3θpθpθq=5θpθqδpθpθq=9θqδpθpθq=θpθq\theta_{p}\theta_{q}=_{3}\theta_{p}\theta_{p\theta_{q}}=_{5}\theta_{p\theta_{q}\delta_{p}}\theta_{p\theta_{q}}=_{9}\theta_{q\delta_{p}}\theta_{p\theta_{q}}=\theta_{p^{\prime}}\theta_{q^{\prime}}.

(iii). The assumptions give p=rδpp=r\delta_{p} and r=qδrr=q\delta_{r} (among other things). We also have δp=δrδp\delta_{p}=\delta_{r}\delta_{p} by Lemma 4.11. But then p=qδp=qδrδp=rδp=pp^{\prime}=q\delta_{p}=q\delta_{r}\delta_{p}=r\delta_{p}=p. ∎

Corollary 4.14.

If pqp\mathrel{\mathscr{F}}q, then for any rpr\leq p and sqs\leq q we have rrθqr\mathrel{\mathscr{F}}r\theta_{q} and sδpss\delta_{p}\mathrel{\mathscr{F}}s.

Proof.

For the first (the second is dual), part (i) of Lemma 4.13 gives rqr^{\prime}\mathrel{\mathscr{F}}q^{\prime}, where r=qδrr^{\prime}=q\delta_{r} and q=rθqq^{\prime}=r\theta_{q}. Part (iii) of the same lemma gives r=rr^{\prime}=r. ∎

Lemma 4.15.

If p1,,pk,q1,,qlPp_{1},\ldots,p_{k},q_{1},\ldots,q_{l}\in P are such that p1pkp_{1}\mathrel{\mathscr{F}}\cdots\mathrel{\mathscr{F}}p_{k} and q1qkq_{1}\mathrel{\mathscr{F}}\cdots\mathrel{\mathscr{F}}q_{k}, then

θp1θpk=θq1θqlpk=qlandδpkδp1=δqlδq1p1=q1.\theta_{p_{1}}\cdots\theta_{p_{k}}=\theta_{q_{1}}\cdots\theta_{q_{l}}\ \Rightarrow\ p_{k}=q_{l}\qquad\text{and}\qquad\delta_{p_{k}}\cdots\delta_{p_{1}}=\delta_{q_{l}}\cdots\delta_{q_{1}}\ \Rightarrow\ p_{1}=q_{1}.
Proof.

For the first (the second is dual), we have

pk=p1θp1θpk=p1θq1θqlql,and similarlyqlpk.p_{k}=p_{1}\theta_{p_{1}}\cdots\theta_{p_{k}}=p_{1}\theta_{q_{1}}\cdots\theta_{q_{l}}\leq q_{l},\qquad\text{and similarly}\qquad q_{l}\leq p_{k}.\qed

The k=l=1k=l=1 case of the previous result has the following simple consequence.

Corollary 4.16.

For p,qPp,q\in P, we have θp=θqδp=δqp=q\theta_{p}=\theta_{q}\ \Leftrightarrow\ \delta_{p}=\delta_{q}\ \Leftrightarrow\ p=q. ∎

It will be convenient to record a result concerning products of projections in a DRC-semigroup.

Lemma 4.17.

If p1,,pk𝐏(S)p_{1},\ldots,p_{k}\in{\bf P}(S) for a DRC-semigroup SS, then

  1. (i)

    D(p1pk)=pkδpk1δp1D(p_{1}\cdots p_{k})=p_{k}\delta_{p_{k-1}}\cdots\delta_{p_{1}} and R(p1pk)=p1θp2θpkR(p_{1}\cdots p_{k})=p_{1}\theta_{p_{2}}\cdots\theta_{p_{k}},

  2. (ii)

    D(p1pk)=p1D(p_{1}\cdots p_{k})=p_{1} and R(p1pk)=pkR(p_{1}\cdots p_{k})=p_{k} if p1pkp_{1}\mathrel{\mathscr{F}}\cdots\mathrel{\mathscr{F}}p_{k}

Proof.

(i). The k=1k=1 case says D(p1)=p1D(p_{1})=p_{1}, which holds by (DRC5). For k2k\geq 2 we have

D(p1p2pk)\displaystyle D(p_{1}p_{2}\cdots p_{k}) =D(p1D(p2pk))\displaystyle=D(p_{1}D(p_{2}\cdots p_{k})) by (DRC2)
=D(p2pk)δp1\displaystyle=D(p_{2}\cdots p_{k})\delta_{p_{1}} by definition
=pkδpk1δp2δp1\displaystyle=p_{k}\delta_{p_{k-1}}\cdots\delta_{p_{2}}\delta_{p_{1}} by induction.

(ii). This follows from (i) and the definition of the \mathrel{\mathscr{F}} relation. ∎

4.2 The category of projection algebras

Definition 4.18.

We write 𝐏𝐀{\bf PA} for the category of projection algebras. A projection algebra morphism (P,θ,δ)(P,θ,δ){(P,\theta,\delta)\to(P^{\prime},\theta^{\prime},\delta^{\prime})} in 𝐏𝐀{\bf PA} is a map ϕ:PP\phi:P\to P^{\prime} such that

(qθp)ϕ=(qϕ)θpϕand(qδp)ϕ=(qϕ)δpϕfor all p,qP.(q\theta_{p})\phi=(q\phi)\theta^{\prime}_{p\phi}\qquad\text{and}\qquad(q\delta_{p})\phi=(q\phi)\delta^{\prime}_{p\phi}\qquad\text{for all $p,q\in P$.}

In Definition 4.3 (cf. Lemma 3.8) we saw that a DRC-semigroup SS gives rise to a projection algebra 𝐏(S){\bf P}(S). In this way, 𝐏{\bf P} can be thought of as an object map 𝐃𝐑𝐂𝐏𝐀{\bf DRC}\to{\bf PA}. Since a DRC-morphism ϕ:SS\phi:S\to S^{\prime} maps projections to projections (by the law D(a)ϕ=D(aϕ)D(a)\phi=D^{\prime}(a\phi)), it follows that we can define

𝐏(ϕ)=ϕ|𝐏(S):𝐏(S)𝐏(S){\bf P}(\phi)=\phi|_{{\bf P}(S)}:{\bf P}(S)\to{\bf P}(S^{\prime})

to be the (set-theoretic) restriction of ϕ\phi to 𝐏(S){\bf P}(S).

Proposition 4.19.

𝐏{\bf P} is a functor 𝐃𝐑𝐂𝐏𝐀{\bf DRC}\to{\bf PA}.

Proof.

It only remains to show that 𝐏(ϕ):𝐏(S)𝐏(S){\bf P}(\phi):{\bf P}(S)\to{\bf P}(S^{\prime}) is a projection algebra morphism for any DRC-morphism ϕ:SS\phi:S\to S^{\prime}. But for any p,q𝐏(S)p,q\in{\bf P}(S) we have

(qθp)ϕ=R(qp)ϕ=R((qp)ϕ)=R((qϕ)(pϕ))=(qϕ)θpϕ,and similarly(qδp)ϕ=(qϕ)δpϕ.(q\theta_{p})\phi=R(qp)\phi=R^{\prime}((qp)\phi)=R^{\prime}((q\phi)(p\phi))=(q\phi)\theta^{\prime}_{p\phi},\qquad\text{and similarly}\qquad(q\delta_{p})\phi=(q\phi)\delta^{\prime}_{p\phi}.\qed

4.3 The chain category of a projection algebra

Definition 4.20.

Let PP be a projection algebra. A (PP-)path is a non-empty tuple 𝔭=(p1,,pk)\mathfrak{p}=(p_{1},\ldots,p_{k}) where p1,,pkPp_{1},\ldots,p_{k}\in P and p1pkp_{1}\mathrel{\mathscr{F}}\cdots\mathrel{\mathscr{F}}p_{k}. We write 𝐝(𝔭)=p1{\bf d}(\mathfrak{p})=p_{1} and 𝐫(𝔭)=pk{\bf r}(\mathfrak{p})=p_{k}. The set 𝒫=𝒫(P)\mathscr{P}=\mathscr{P}(P) of all such paths is the path category of PP, under the composition defined by

(p1,,pk)(pk,,pl)=(p1,,pk1,pk,pk+1,,pl).(p_{1},\ldots,p_{k})\circ(p_{k},\ldots,p_{l})=(p_{1},\ldots,p_{k-1},p_{k},p_{k+1},\ldots,p_{l}).

For pPp\in P, we identify the path (p)(p) with pp itself, and in this way we have v𝒫=Pv\mathscr{P}=P.

We now wish to give the path category 𝒫=𝒫(P)\mathscr{P}=\mathscr{P}(P) the structure of a biordered category. To this end, consider a path 𝔭=(p1,,pk)𝒫\mathfrak{p}=(p_{1},\ldots,p_{k})\in\mathscr{P}, and let q𝐝(𝔭)=p1q\leq{\bf d}(\mathfrak{p})=p_{1} and r𝐫(𝔭)=pkr\leq{\bf r}(\mathfrak{p})=p_{k}. We then define the restrictions

q𝔭\displaystyle{}_{q}{\downharpoonleft}\mathfrak{p} =(q1,,qk),\displaystyle=(q_{1},\ldots,q_{k}), where qi\displaystyle q_{i} =qθp1θpi=qθp2θpi\displaystyle=q\theta_{p_{1}}\cdots\theta_{p_{i}}=q\theta_{p_{2}}\cdots\theta_{p_{i}} for each 1ik1\leq i\leq k, (4.21)
and
𝔭r\displaystyle\mathfrak{p}{\downharpoonright}_{r} =(r1,,rk),\displaystyle=(r_{1},\ldots,r_{k}), where ri\displaystyle r_{i} =rδpkδpi=rδpk1δpi\displaystyle=r\delta_{p_{k}}\cdots\delta_{p_{i}}=r\delta_{p_{k-1}}\cdots\delta_{p_{i}} for each 1ik1\leq i\leq k. (4.22)

Note here that qθp1=qq\theta_{p_{1}}=q and rδpk=rr\delta_{p_{k}}=r, since qp1q\leq p_{1} and rpkr\leq p_{k}, so in particular we have q1=qq_{1}=q and rk=rr_{k}=r. We now prove a sequence of results that show these restrictions satisfy conditions (ii)(O1)(ii)(O4) from Lemma 2.2.

Lemma 4.23.

If 𝔭𝒫\mathfrak{p}\in\mathscr{P}, and if q𝐝(𝔭)q\leq{\bf d}(\mathfrak{p}), then q𝔭𝒫{}_{q}{\downharpoonleft}\mathfrak{p}\in\mathscr{P}, and we have 𝐝(q𝔭)=q{\bf d}({}_{q}{\downharpoonleft}\mathfrak{p})=q and 𝐫(q𝔭)𝐫(𝔭){\bf r}({}_{q}{\downharpoonleft}\mathfrak{p})\leq{\bf r}(\mathfrak{p}).

Proof.

Let 𝔭=(p1,,pk)\mathfrak{p}=(p_{1},\ldots,p_{k}), and write q𝔭=(q1,,qk){}_{q}{\downharpoonleft}\mathfrak{p}=(q_{1},\ldots,q_{k}) as in (4.21). Since 𝔭𝒫\mathfrak{p}\in\mathscr{P} we have

pipi+1,i.e.piθpi+1=pi+1 and pi+1δpi=pifor all 1i<k.p_{i}\mathrel{\mathscr{F}}p_{i+1},\qquad\text{i.e.}\qquad p_{i}\theta_{p_{i+1}}=p_{i+1}\text{ \ and \ }p_{i+1}\delta_{p_{i}}=p_{i}\qquad\text{for all $1\leq i<k$.}

To show that q𝔭𝒫{}_{q}{\downharpoonleft}\mathfrak{p}\in\mathscr{P}, we must show that qiqi+1q_{i}\mathrel{\mathscr{F}}q_{i+1} for all 1i<k1\leq i<k, i.e. that qiθqi+1=qi+1q_{i}\theta_{q_{i+1}}=q_{i+1} and qi+1δqi=qiq_{i+1}\delta_{q_{i}}=q_{i}. By definition, we note that qi+1=qiθpi+1q_{i+1}=q_{i}\theta_{p_{i+1}}. We then have

qiθqi+1=qiθqiθpi+1=8qiθpi+1=qi+1.q_{i}\theta_{q_{i+1}}=q_{i}\theta_{q_{i}\theta_{p_{i+1}}}=_{8}q_{i}\theta_{p_{i+1}}=q_{i+1}.

Since qipiq_{i}\leq p_{i}, we have piδqi=qip_{i}\delta_{q_{i}}=q_{i}, and also δqi=δpiδqi\delta_{q_{i}}=\delta_{p_{i}}\delta_{q_{i}} by Lemma 4.11, so

qi+1δqi=qiθpi+1δqi=9pi+1δqi=pi+1δpiδqi=piδqi=qi.q_{i+1}\delta_{q_{i}}=q_{i}\theta_{p_{i+1}}\delta_{q_{i}}=_{9}p_{i+1}\delta_{q_{i}}=p_{i+1}\delta_{p_{i}}\delta_{q_{i}}=p_{i}\delta_{q_{i}}=q_{i}.

By definition, we have 𝐝(q𝔭)=q1=q{\bf d}({}_{q}{\downharpoonleft}\mathfrak{p})=q_{1}=q, and 𝐫(q𝔭)=qk=qθp1θpkpk{\bf r}({}_{q}{\downharpoonleft}\mathfrak{p})=q_{k}=q\theta_{p_{1}}\cdots\theta_{p_{k}}\leq p_{k}. ∎

Lemma 4.24.

If 𝔭𝒫\mathfrak{p}\in\mathscr{P}, and if q=𝐝(𝔭)q={\bf d}(\mathfrak{p}), then q𝔭=𝔭{}_{q}{\downharpoonleft}\mathfrak{p}=\mathfrak{p}.

Proof.

Let 𝔭=(p1,,pk)\mathfrak{p}=(p_{1},\ldots,p_{k}), and write q𝔭=(q1,,qk){}_{q}{\downharpoonleft}\mathfrak{p}=(q_{1},\ldots,q_{k}) as in (4.21). We need to show that qi=piq_{i}=p_{i} for all 1ik1\leq i\leq k. This is true for i=1i=1, as q1=q=𝐝(𝔭)=p1q_{1}=q={\bf d}(\mathfrak{p})=p_{1}. If i2i\geq 2, then by induction, and since pi1pip_{i-1}\mathrel{\mathscr{F}}p_{i}, we have qi=qi1θpi=pi1θpi=piq_{i}=q_{i-1}\theta_{p_{i}}=p_{i-1}\theta_{p_{i}}=p_{i}. ∎

Lemma 4.25.

If 𝔭𝒫\mathfrak{p}\in\mathscr{P}, and if rq𝐝(𝔭)r\leq q\leq{\bf d}(\mathfrak{p}), then rq𝔭=r𝔭{}_{r}{\downharpoonleft}{}_{q}{\downharpoonleft}\mathfrak{p}={}_{r}{\downharpoonleft}\mathfrak{p}.

Proof.

Write

𝔭=(p1,,pk),q𝔭=(q1,,qk),r𝔭=(r1,,rk)andrq𝔭=(s1,,sk).\mathfrak{p}=(p_{1},\ldots,p_{k}),\qquad{}_{q}{\downharpoonleft}\mathfrak{p}=(q_{1},\ldots,q_{k}),\qquad{}_{r}{\downharpoonleft}\mathfrak{p}=(r_{1},\ldots,r_{k})\qquad\text{and}\qquad{}_{r}{\downharpoonleft}{}_{q}{\downharpoonleft}\mathfrak{p}=(s_{1},\ldots,s_{k}).

So

qi=qθp1θpi,ri=rθp1θpiandsi=rθq1θqifor all 1ik,q_{i}=q\theta_{p_{1}}\cdots\theta_{p_{i}},\qquad r_{i}=r\theta_{p_{1}}\cdots\theta_{p_{i}}\qquad\text{and}\qquad s_{i}=r\theta_{q_{1}}\cdots\theta_{q_{i}}\qquad\text{for all $1\leq i\leq k$,} (4.26)

and we must show that ri=sir_{i}=s_{i} for all ii. For i=1i=1 we have r1=r=s1r_{1}=r=s_{1}. For i2i\geq 2,

si\displaystyle s_{i} =si1θqi\displaystyle=s_{i-1}\theta_{q_{i}} by (4.26)
=si1θqi1θpi\displaystyle=s_{i-1}\theta_{q_{i-1}\theta_{p_{i}}} by (4.26)
=(si1θqi1)θqi1θpi\displaystyle=(s_{i-1}\theta_{q_{i-1}})\theta_{q_{i-1}\theta_{p_{i}}} by (4.26) and (P6)
=si1θqi1θpi\displaystyle=s_{i-1}\theta_{q_{i-1}}\theta_{p_{i}} by (P3)
=si1θpi\displaystyle=s_{i-1}\theta_{p_{i}} by (4.26) and (P6)
=ri1θpi\displaystyle=r_{i-1}\theta_{p_{i}} by induction
=ri\displaystyle=r_{i} by (4.26).\displaystyle\text{by \eqref{eq:pqrsi}.}\qed
Lemma 4.27.

If 𝔭,𝔮𝒫\mathfrak{p},\mathfrak{q}\in\mathscr{P} with 𝐫(𝔭)=𝐝(𝔮){\bf r}(\mathfrak{p})={\bf d}(\mathfrak{q}), and if r𝐝(𝔭)r\leq{\bf d}(\mathfrak{p}), then with s=𝐫(r𝔭)s={\bf r}({}_{r}{\downharpoonleft}\mathfrak{p}) we have

r(𝔭𝔮)=r𝔭s𝔮.{}_{r}{\downharpoonleft}(\mathfrak{p}\circ\mathfrak{q})={}_{r}{\downharpoonleft}\mathfrak{p}\circ{}_{s}{\downharpoonleft}\mathfrak{q}.
Proof.

Write 𝔭=(p1,,pk)\mathfrak{p}=(p_{1},\ldots,p_{k}) and 𝔮=(q1,,ql)\mathfrak{q}=(q_{1},\ldots,q_{l}), noting that pk=q1p_{k}=q_{1}. Then

r𝔭\displaystyle{}_{r}{\downharpoonleft}\mathfrak{p} =(r,rθp2,rθp2θp3,,rθp2θpk),sos=rθp2θpk.\displaystyle=(r,r\theta_{p_{2}},r\theta_{p_{2}}\theta_{p_{3}},\ldots,r\theta_{p_{2}}\cdots\theta_{p_{k}}),\qquad\text{so}\qquad s=r\theta_{p_{2}}\cdots\theta_{p_{k}}.
It follows that
r(𝔭𝔮)\displaystyle{}_{r}{\downharpoonleft}(\mathfrak{p}\circ\mathfrak{q}) =r(p1,,pk,q2,,ql)\displaystyle={}_{r}{\downharpoonleft}(p_{1},\ldots,p_{k},q_{2},\ldots,q_{l})
=(r,rθp2,rθp2θp3,,rθp2θpk=s,sθq2,sθq2θq3,,sθq2θql)\displaystyle=(r,r\theta_{p_{2}},r\theta_{p_{2}}\theta_{p_{3}},\ldots,\underbrace{r\theta_{p_{2}}\cdots\theta_{p_{k}}}_{=s},s\theta_{q_{2}},s\theta_{q_{2}}\theta_{q_{3}},\ldots,s\theta_{q_{2}}\cdots\theta_{q_{l}})
=(r,rθp2,rθp2θp3,,rθp2θpk)(s,sθq2,sθq2θq3,,sθq2θql)=r𝔭s𝔮.\displaystyle=(r,r\theta_{p_{2}},r\theta_{p_{2}}\theta_{p_{3}},\ldots,r\theta_{p_{2}}\cdots\theta_{p_{k}})\circ(s,s\theta_{q_{2}},s\theta_{q_{2}}\theta_{q_{3}},\ldots,s\theta_{q_{2}}\cdots\theta_{q_{l}})={}_{r}{\downharpoonleft}\mathfrak{p}\circ{}_{s}{\downharpoonleft}\mathfrak{q}.\qed

Lemmas 4.234.27 and their duals show that the restrictions given in (4.21) and (4.22) satisfy the conditions of Lemma 2.2. It follows that 𝒫=𝒫(P)\mathscr{P}=\mathscr{P}(P) is a biordered category, under the orders \leq_{\ell} and r\leq_{r} defined, for 𝔭,𝔮𝒫\mathfrak{p},\mathfrak{q}\in\mathscr{P}, by

𝔭𝔮𝔭=𝐝(𝔭)𝔮and𝔭r𝔮𝔭=𝔮𝐫(𝔭).\mathfrak{p}\leq_{\ell}\mathfrak{q}\ \Leftrightarrow\ \mathfrak{p}={}_{{\bf d}(\mathfrak{p})}{\downharpoonleft}\mathfrak{q}\qquad\text{and}\qquad\mathfrak{p}\leq_{r}\mathfrak{q}\ \Leftrightarrow\ \mathfrak{p}=\mathfrak{q}{\downharpoonright}_{{\bf r}(\mathfrak{p})}.
Definition 4.28.

Let PP be a projection algebra, and let 𝒫=𝒫(P)\mathscr{P}=\mathscr{P}(P) be the path category. We let \approx be the congruence on 𝒫\mathscr{P} generated by the relations (p,p)p(p,p)\approx p for all pPp\in P. Since

q(p,p)=(q,q)q=qp,and similarly(p,p)qpqfor qp,{}_{q}{\downharpoonleft}(p,p)=(q,q)\approx q={}_{q}{\downharpoonleft}p,\qquad\text{and similarly}\qquad(p,p){\downharpoonright}_{q}\approx p{\downharpoonright}_{q}\qquad\text{for $q\leq p$,}

it follows that \approx is a biordered congruence. Consequently, the quotient

𝒞=𝒞(P)=𝒫/\mathscr{C}=\mathscr{C}(P)=\mathscr{P}/{\approx}

is a biordered category under the induced \leq_{\ell} and r\leq_{r} orders. We call 𝒞=𝒞(P)\mathscr{C}=\mathscr{C}(P) the chain category of PP. An element of 𝒞\mathscr{C} is called a (PP-)chain, and is an \approx-class of a path. For a path 𝔭=(p1,,pk)𝒫\mathfrak{p}=(p_{1},\ldots,p_{k})\in\mathscr{P}, we write [𝔭]=[p1,,pk][\mathfrak{p}]=[p_{1},\ldots,p_{k}] for the corresponding chain, so that 𝒞={[𝔭]:𝔭𝒫}\mathscr{C}=\{[\mathfrak{p}]:\mathfrak{p}\in\mathscr{P}\}. We again identify a projection pPp\in P with the chain [p]𝒞[p]\in\mathscr{C}, so that v𝒞=Pv\mathscr{C}=P. Restrictions in 𝒞\mathscr{C} are given by

q[𝔭]=[q𝔭]and[𝔭]r=[𝔭r]for 𝔭𝒫, and q𝐝(𝔭) and r𝐫(𝔭).{}_{q}{\downharpoonleft}[\mathfrak{p}]=[{}_{q}{\downharpoonleft}\mathfrak{p}]\qquad\text{and}\qquad[\mathfrak{p}]{\downharpoonright}_{r}=[\mathfrak{p}{\downharpoonright}_{r}]\qquad\text{for $\mathfrak{p}\in\mathscr{P}$, and $q\leq{\bf d}(\mathfrak{p})$ and $r\leq{\bf r}(\mathfrak{p})$.}

5 Chained projection categories

In Proposition 3.11 we saw that a DRC-semigroup SS gives rise to a biordered category 𝒞(S)\mathcal{C}(S) whose object set is the projection algebra 𝐏(S){\bf P}(S). It turns out that the category 𝒞(S)\mathcal{C}(S) carries rather a lot more information, and has the structure of what we will call a chained projection category. In this section we introduce this class of categories, and the category 𝐂𝐏𝐂{\bf CPC} formed by them. We do so sequentially, by first defining weak projection categories (Section 5.1), projection categories (Section 5.2), weak chained projection categories (Section 5.3), and finally chained projection categories (Section 5.4). At each step of the way we will show that a DRC-semigroup gives rise to a structure of the relevant kind, ultimately leading to a functor 𝐂:𝐃𝐑𝐂𝐂𝐏𝐂{\bf C}:{\bf DRC}\to{\bf CPC}; see Theorem 5.20. We will eventually see in Section 7 that 𝐂{\bf C} is an isomorphism.

5.1 Weak projection categories

Definition 5.1.

A weak projection category is a pair (P,𝒞)(P,\mathcal{C}), consisting of a biordered category 𝒞=(𝒞,,r)\mathcal{C}=(\mathcal{C},\leq_{\ell},\leq_{r}) and a projection algebra P=v𝒞P=v\mathcal{C}, for which the restriction of both orders \leq_{\ell} and r\leq_{r} to PP is the order \leq from (4.9).

Let (P,𝒞)(P,\mathcal{C}) be a weak projection category, and let a𝒞a\in\mathcal{C}. As in (2.5), we have the two maps

ϑa:𝐝(a)𝐫(a)anda:𝐫(a)𝐝(a),\vartheta_{a}:{\bf d}(a)^{\downarrow}\to{\bf r}(a)^{\downarrow}\qquad\text{and}\qquad\partial_{a}:{\bf r}(a)^{\downarrow}\to{\bf d}(a)^{\downarrow},

given by pϑa=𝐫(pa)p\vartheta_{a}={\bf r}({}_{p}{\downharpoonleft}a) and qa=𝐝(aq)q\partial_{a}={\bf d}(a{\downharpoonright}_{q}), for p𝐝(a)p\leq{\bf d}(a) and q𝐫(a)q\leq{\bf r}(a). Since im(θ𝐝(a))=𝐝(a)\operatorname{im}(\theta_{{\bf d}(a)})={\bf d}(a)^{\downarrow} and im(δ𝐫(a))=𝐫(a)\operatorname{im}(\delta_{{\bf r}(a)})={\bf r}(a)^{\downarrow}, we can also define the maps

Θa=θ𝐝(a)ϑa:P𝐫(a)andΔa=δ𝐫(a)a:P𝐝(a).\Theta_{a}=\theta_{{\bf d}(a)}\vartheta_{a}:P\to{\bf r}(a)^{\downarrow}\qquad\text{and}\qquad\Delta_{a}=\delta_{{\bf r}(a)}\partial_{a}:P\to{\bf d}(a)^{\downarrow}. (5.2)
Lemma 5.3.

If (P,𝒞)(P,\mathcal{C}) is a weak projection category, and if a𝒞a\in\mathcal{C}, then

  1. (i)

    ϑaθ𝐫(a)=ϑa\vartheta_{a}\theta_{{\bf r}(a)}=\vartheta_{a} and aδ𝐝(a)=a\partial_{a}\delta_{{\bf d}(a)}=\partial_{a},

  2. (ii)

    Θaθ𝐫(a)=Θa\Theta_{a}\theta_{{\bf r}(a)}=\Theta_{a} and Δaδ𝐝(a)=Δa\Delta_{a}\delta_{{\bf d}(a)}=\Delta_{a},

  3. (iii)

    Θpa=θpΘa\Theta_{{}_{p}{\downharpoonleft}a}=\theta_{p}\Theta_{a} and Δaq=δqΔa\Delta_{a{\downharpoonright}_{q}}=\delta_{q}\Delta_{a} for any p𝐝(a)p\leq{\bf d}(a) and q𝐫(a)q\leq{\bf r}(a),

  4. (iv)

    ϑa=Θa|𝐝(a)\vartheta_{a}=\Theta_{a}|_{{\bf d}(a)^{\downarrow}} and a=Δa|𝐫(a)\partial_{a}=\Delta_{a}|_{{\bf r}(a)^{\downarrow}}.

Proof.

(i). For any tdom(ϑa)=𝐝(a)t\in\operatorname{dom}(\vartheta_{a})={\bf d}(a)^{\downarrow}, we have tϑa𝐫(a)t\vartheta_{a}\leq{\bf r}(a), and so tϑa=(tϑa)θ𝐫(a)t\vartheta_{a}=(t\vartheta_{a})\theta_{{\bf r}(a)}.

(ii). This follows from (i), as Θa=θ𝐝(a)ϑa\Theta_{a}=\theta_{{\bf d}(a)}\vartheta_{a} and Δa=δ𝐫(a)a\Delta_{a}=\delta_{{\bf r}(a)}\partial_{a}.

(iii). For any tPt\in P we have

tΘpa\displaystyle t\Theta_{{}_{p}{\downharpoonleft}a} =tθpϑpa\displaystyle=t\theta_{p}\vartheta_{{}_{p}{\downharpoonleft}a} by definition
=tθpϑa\displaystyle=t\theta_{p}\vartheta_{a} by (2.6)
=tθpθ𝐝(a)ϑa\displaystyle=t\theta_{p}\theta_{{\bf d}(a)}\vartheta_{a} by Lemma 4.11, as p𝐝(a)p\leq{\bf d}(a)
=tθpΘa\displaystyle=t\theta_{p}\Theta_{a} by definition.

(iv). The maps ϑa\vartheta_{a} and Θa|𝐝(a)\Theta_{a}|_{{\bf d}(a)^{\downarrow}} both have domain 𝐝(a){\bf d}(a)^{\downarrow}, and for p𝐝(a)p\in{\bf d}(a)^{\downarrow} we have p=pθ𝐝(a)p=p\theta_{{\bf d}(a)} by definition, and so pΘa=pθ𝐝(a)ϑa=pϑap\Theta_{a}=p\theta_{{\bf d}(a)}\vartheta_{a}=p\vartheta_{a}. ∎

Lemma 5.4.

If SS is a DRC-semigroup, then (𝐏(S),𝒞(S))({\bf P}(S),\mathcal{C}(S)) is a weak projection category.

Proof.

Write P=𝐏(S)P={\bf P}(S) and 𝒞=𝒞(S)\mathcal{C}=\mathcal{C}(S). By Lemma 3.8 and Proposition 3.11, PP is a projection algebra and 𝒞\mathcal{C} a biordered category, and by construction we have P=v𝒞P=v\mathcal{C}. It remains to check that:

pqpqprqfor all p,qP.p\leq_{\ell}q\ \ \Leftrightarrow\ \ p\leq q\ \ \Leftrightarrow\ \ p\leq_{r}q\qquad\text{for all $p,q\in P$.}

As ever, it is enough to establish the equivalence involving \leq_{\ell} and \leq. For this, first suppose pqp\leq_{\ell}q, so that p=rq=rqp={}_{r}{\downharpoonleft}q=rq for some rqr\leq q. It follows that p=pqp=pq, so that pqp\leq q (cf. (3.5)). Conversely, if pq(=𝐝(q))p\leq q\ (={\bf d}(q)), then the restriction pq{}_{p}{\downharpoonleft}q exists, and pq=pq=p{}_{p}{\downharpoonleft}q=pq=p; it follows that pqp\leq_{\ell}q. ∎

In what follows, we will need to understand the ϑ/Θ\vartheta/\Theta and /Δ\partial/\Delta maps associated to the weak projection category (P,𝒞)=(𝐏(S),𝒞(S))(P,\mathcal{C})=({\bf P}(S),\mathcal{C}(S)) arising from a DRC-semigroup SS. For this, consider a morphism a𝒞(=S)a\in\mathcal{C}\ (=S). For p𝐝(a)=D(a)p\leq{\bf d}(a)=D(a) we have

pϑa=𝐫(pa)=R(pa).p\vartheta_{a}={\bf r}({}_{p}{\downharpoonleft}a)=R(pa).

It follows that for arbitrary qPq\in P we have

qΘa=qθ𝐝(a)ϑa=R(qD(a))ϑa=R(R(qD(a))a)=2R(qD(a)a)=1R(qa).q\Theta_{a}=q\theta_{{\bf d}(a)}\vartheta_{a}=R(qD(a))\vartheta_{a}=R(R(qD(a))a)=_{2}R(qD(a)a)=_{1}R(qa).

Similar calculations apply to the \partial and Δ\Delta maps, and in summary we have:

pϑa\displaystyle p\vartheta_{a} =R(pa) for p𝐝(a),\displaystyle=R(pa)\text{ \ for $p\leq{\bf d}(a)$,} pΘa=R(pa) for pP,\displaystyle p\Theta_{a}=R(pa)\text{ \ for $p\in P$,}
pa\displaystyle p\partial_{a} =D(ap) for p𝐫(a),\displaystyle=D(ap)\text{ \ for $p\leq{\bf r}(a)$,} pΔa=D(ap) for pP.\displaystyle p\Delta_{a}=D(ap)\text{ \ for $p\in P$.} (5.5)

5.2 Projection categories

We saw in Lemma 5.3(iii) that the Θ\Theta and Δ\Delta maps in a weak projection category (P,𝒞)(P,\mathcal{C}) behave well with regard to left and right restrictions, respectively, in the sense that Θpa=θpΘa\Theta_{{}_{p}{\downharpoonleft}a}=\theta_{p}\Theta_{a} and Δaq=δqΔa\Delta_{a{\downharpoonright}_{q}}=\delta_{q}\Delta_{a} for any p𝐝(a)p\leq{\bf d}(a) and q𝐫(a)q\leq{\bf r}(a). We are particularly interested in the situation in which the analogous statements hold for the other restrictions.

Definition 5.6.

A projection category is a weak projection category (P,𝒞)(P,\mathcal{C}) satisfying:

  1. (C1)

    For every a𝒞a\in\mathcal{C}, and every p𝐝(a)p\leq{\bf d}(a) and q𝐫(a)q\leq{\bf r}(a), we have

    Θaq=ΘaθqandΔpa=Δaδp.\Theta_{a{\downharpoonright}_{q}}=\Theta_{a}\theta_{q}\qquad\text{and}\qquad\Delta_{{}_{p}{\downharpoonleft}a}=\Delta_{a}\delta_{p}.

We write 𝐏𝐂{\bf PC} for the category of projection categories. A morphism (P,𝒞)(P,𝒞)(P,\mathcal{C})\to(P^{\prime},\mathcal{C}^{\prime}) in 𝐏𝐂{\bf PC} is a biordered morphism ϕ:𝒞𝒞\phi:\mathcal{C}\to\mathcal{C}^{\prime} whose object map vϕ:PPv\phi:P\to P^{\prime} is a projection algebra morphism.

Proposition 5.7.

The assignment S(𝐏(S),𝒞(S))S\mapsto({\bf P}(S),\mathcal{C}(S)) is the object part of a functor 𝐃𝐑𝐂𝐏𝐂{\bf DRC}\to{\bf PC}.

Proof.

We first check that (P,𝒞)=(𝐏(S),𝒞(S))(P,\mathcal{C})=({\bf P}(S),\mathcal{C}(S)) is a projection category for any DRC-semigroup SS. By Lemma 5.4, it remains to check that (C1) holds. By symmetry, it suffices to show that Θaq=Θaθq\Theta_{a{\downharpoonright}_{q}}=\Theta_{a}\theta_{q} for all a𝒞a\in\mathcal{C}, and all q𝐫(a)q\leq{\bf r}(a). For this we fix tPt\in P, and use (5.5) to calculate

tΘaq=tΘaq=R(taq)=2R(R(ta)q)=R(ta)θq=tΘaθq.t\Theta_{a{\downharpoonright}_{q}}=t\Theta_{aq}=R(taq)=_{2}R(R(ta)q)=R(ta)\theta_{q}=t\Theta_{a}\theta_{q}.

Now we know that 𝐂{\bf C} maps objects of 𝐃𝐑𝐂{\bf DRC} to objects of 𝐏𝐂{\bf PC}. It remains to check that any DRC-morphism ϕ:SS\phi:S\to S^{\prime} is a projection category morphism from (P,𝒞)=(𝐏(S),𝒞(S))(P,\mathcal{C})=({\bf P}(S),\mathcal{C}(S)) to (P,𝒞)=(𝐏(S),𝒞(S))(P^{\prime},\mathcal{C}^{\prime})=({\bf P}(S^{\prime}),\mathcal{C}(S^{\prime})), i.e. that:

  • ϕ\phi is a biordered category morphism 𝒞𝒞\mathcal{C}\to\mathcal{C}^{\prime}, and

  • vϕv\phi is a projection algebra morphism PPP\to P^{\prime}.

But these follow from Propositions 3.13 and 4.19, respectively. ∎

5.3 Weak chained projection categories

We now bring in an extra layer of structure.

Definition 5.8.

Given a projection category (P,𝒞)(P,\mathcal{C}), an evaluation map is a biordered vv-functor ε:𝒞𝒞\varepsilon:\mathscr{C}\to\mathcal{C}, where 𝒞=𝒞(P)\mathscr{C}=\mathscr{C}(P) is the chain category of PP, and where vv-functor means that ε(p)=p\varepsilon(p)=p for all pP=v𝒞p\in P=v\mathcal{C}. (We write evaluation maps to the left of their arguments.) It quickly follows that 𝐝(ε(𝔠))=𝐝(𝔠){\bf d}(\varepsilon(\mathfrak{c}))={\bf d}(\mathfrak{c}) and 𝐫(ε(𝔠))=𝐫(𝔠){\bf r}(\varepsilon(\mathfrak{c}))={\bf r}(\mathfrak{c}) for all 𝔠𝒞\mathfrak{c}\in\mathscr{C}.

Within the image of ε\varepsilon, we will mainly be interested in elements of the form ε[p,q]\varepsilon[p,q], for (p,q)(p,q)\in{\mathrel{\mathscr{F}}}. The next lemma gathers some simple properties of these elements, all of which follow quickly from the definitions.

Lemma 5.9.

Let (P,𝒞)(P,\mathcal{C}) be a projection category, and ε:𝒞𝒞\varepsilon:\mathscr{C}\to\mathcal{C} an evaluation map.

  1. (i)

    For all pPp\in P we have ε[p,p]=p\varepsilon[p,p]=p.

  2. (ii)

    For all (p,q)(p,q)\in{\mathrel{\mathscr{F}}} we have 𝐝(ε[p,q])=p{\bf d}(\varepsilon[p,q])=p and 𝐫(ε[p,q])=q{\bf r}(\varepsilon[p,q])=q,

  3. (iii)

    For all (p,q)(p,q)\in{\mathrel{\mathscr{F}}}, and for all rpr\leq p and sqs\leq q, we have

    rε[p,q]=ε[r,rθq]andε[p,q]s=ε[sδp,s].{}_{r}{\downharpoonleft}\varepsilon[p,q]=\varepsilon[r,r\theta_{q}]\qquad\text{and}\qquad\varepsilon[p,q]{\downharpoonright}_{s}=\varepsilon[s\delta_{p},s].
Definition 5.10.

A weak chained projection category is a triple (P,𝒞,ε)(P,\mathcal{C},\varepsilon), where (P,𝒞)(P,\mathcal{C}) is a projection category, and ε:𝒞𝒞\varepsilon:\mathcal{C}\to\mathcal{C} is an evaluation map. We write 𝐖𝐂𝐏𝐂{\bf WCPC} for the category of weak chained projection categories. A morphism (P,𝒞,ε)(P,𝒞,ε)(P,\mathcal{C},\varepsilon)\to(P^{\prime},\mathcal{C}^{\prime},\varepsilon^{\prime}) in 𝐖𝐂𝐏𝐂{\bf WCPC} is called a chained projection functor, and is a projection category morphism ϕ:(P,𝒞)(P,𝒞)\phi:(P,\mathcal{C})\to(P^{\prime},\mathcal{C}^{\prime}) that respects the evaluation maps, in the sense that

(ε[p1,,pk])ϕ=ε[p1ϕ,,pkϕ]for p1,,pkP with p1pk.(\varepsilon[p_{1},\ldots,p_{k}])\phi=\varepsilon^{\prime}[p_{1}\phi,\ldots,p_{k}\phi]\qquad\text{for $p_{1},\ldots,p_{k}\in P$ with $p_{1}\mathrel{\mathscr{F}}\cdots\mathrel{\mathscr{F}}p_{k}$.}

Consider a DRC-semigroup SS, and write (P,𝒞)=(𝐏(S),𝒞(S))(P,\mathcal{C})=({\bf P}(S),\mathcal{C}(S)). Since projections are idempotents, there is a well-defined map

ε=ε(S):𝒞=𝒞(P)𝒞(=S)given byε[p1,,pk]=p1pk,\varepsilon=\varepsilon(S):\mathscr{C}=\mathscr{C}(P)\to\mathcal{C}\ (=S)\qquad\text{given by}\qquad\varepsilon[p_{1},\ldots,p_{k}]=p_{1}\cdots p_{k},

where the latter product is taken in SS.

Lemma 5.11.

ε=ε(S)\varepsilon=\varepsilon(S) is an evaluation map.

Proof.

We first claim that for 𝔠=[p1,,pk]𝒫\mathfrak{c}=[p_{1},\ldots,p_{k}]\in\mathscr{P} we have

𝐝(ε(𝔠))=p1=𝐝(𝔠)and𝐫(ε(𝔠))=pk=𝐫(𝔠),{\bf d}(\varepsilon(\mathfrak{c}))=p_{1}={\bf d}(\mathfrak{c})\qquad\text{and}\qquad{\bf r}(\varepsilon(\mathfrak{c}))=p_{k}={\bf r}(\mathfrak{c}),

i.e. that

D(p1pk)=p1andR(p1pk)=pkwhen p1pk.D(p_{1}\cdots p_{k})=p_{1}\qquad\text{and}\qquad R(p_{1}\cdots p_{k})=p_{k}\qquad\text{when $p_{1}\mathrel{\mathscr{F}}\cdots\mathrel{\mathscr{F}}p_{k}$.} (5.12)

By symmetry it suffices to prove the first. The k=1k=1 case being clear, we assume that k2k\geq 2. We then use (DRC2), induction and p1p2p_{1}\mathrel{\mathscr{F}}p_{2} to calculate

D(p1p2pk)=D(p1D(p2pk))=D(p1p2)=p2δp1=p1.D(p_{1}p_{2}\cdots p_{k})=D(p_{1}D(p_{2}\cdots p_{k}))=D(p_{1}p_{2})=p_{2}\delta_{p_{1}}=p_{1}.

Next we check that ε\varepsilon is a vv-functor. Certainly ε(p)=p\varepsilon(p)=p for all pP=v𝒞p\in P=v\mathcal{C}. Now consider composable chains 𝔠=[p1,,pk]\mathfrak{c}=[p_{1},\ldots,p_{k}] and 𝔡=[pk,,pl]\mathfrak{d}=[p_{k},\ldots,p_{l}] in 𝒞\mathscr{C}. We then have

ε(𝔠𝔡)=ε[p1,,pk,,pl]=p1pkpl=p1pkpkpl=ε(𝔠)ε(𝔡)=ε(𝔠)ε(𝔡).\varepsilon(\mathfrak{c}\circ\mathfrak{d})=\varepsilon[p_{1},\ldots,p_{k},\ldots,p_{l}]=p_{1}\cdots p_{k}\cdots p_{l}=p_{1}\cdots p_{k}\cdot p_{k}\cdots p_{l}=\varepsilon(\mathfrak{c})\cdot\varepsilon(\mathfrak{d})=\varepsilon(\mathfrak{c})\circ\varepsilon(\mathfrak{d}).

Note that the final product is a composition in 𝒞\mathcal{C} since 𝐫(p1pk)=pk=𝐝(pkpl){\bf r}(p_{1}\cdots p_{k})=p_{k}={\bf d}(p_{k}\cdots p_{l}) by (5.12).

Finally, we need to check that ε\varepsilon is biordered, i.e. that

ε(q𝔠)=qε(𝔠)andε(𝔠r)=ε(𝔠)rfor all 𝔠𝒞, and for all q𝐝(𝔠) and r𝐫(𝔠).\varepsilon({}_{q}{\downharpoonleft}\mathfrak{c})={}_{q}{\downharpoonleft}\varepsilon(\mathfrak{c})\quad\text{and}\quad\varepsilon(\mathfrak{c}{\downharpoonright}_{r})=\varepsilon(\mathfrak{c}){\downharpoonright}_{r}\qquad\text{for all $\mathfrak{c}\in\mathscr{C}$, and for all $q\leq{\bf d}(\mathfrak{c})$ and $r\leq{\bf r}(\mathfrak{c})$.}

For the first (the second is dual), write 𝔠=[p1,,pk]\mathfrak{c}=[p_{1},\ldots,p_{k}], so that q𝔠=[q1,,qk]{}_{q}{\downharpoonleft}\mathfrak{c}=[q_{1},\ldots,q_{k}], where each qi=qθp2θpiq_{i}=q\theta_{p_{2}}\cdots\theta_{p_{i}}. Noting that ε(q𝔠)=q1qk\varepsilon({}_{q}{\downharpoonleft}\mathfrak{c})=q_{1}\cdots q_{k} and qε(𝔠)=qp1pk{}_{q}{\downharpoonleft}\varepsilon(\mathfrak{c})=q\cdot p_{1}\cdots p_{k}, we need to show that

qp1pk=q1qk.q\cdot p_{1}\cdots p_{k}=q_{1}\cdots q_{k}.

When k=1k=1 we have qp1=q=q1q\cdot p_{1}=q=q_{1} (as qp1q\leq p_{1}), so now suppose k2k\geq 2. We then have qp1pk1pk=q1qk1pkq\cdot p_{1}\cdots p_{k-1}p_{k}=q_{1}\cdots q_{k-1}\cdot p_{k} by induction, so it remains to show that qk1pk=qk1qkq_{k-1}p_{k}=q_{k-1}q_{k}. For this we use Lemma 3.4 to calculate

qk1pk=1qk1pkR(qk1pk)=qk1R(qk1pk)=qk1qk1θpk=qk1qk.q_{k-1}p_{k}=_{1}q_{k-1}p_{k}R(q_{k-1}p_{k})=q_{k-1}R(q_{k-1}p_{k})=q_{k-1}\cdot q_{k-1}\theta_{p_{k}}=q_{k-1}q_{k}.\qed
Proposition 5.13.

The assignment S(𝐏(S),𝒞(S),ε(S))S\mapsto({\bf P}(S),\mathcal{C}(S),\varepsilon(S)) is the object part of a functor 𝐃𝐑𝐂𝐖𝐂𝐏𝐂{\bf DRC}\to{\bf WCPC}.

Proof.

It follows from Proposition 5.7 and Lemma 5.11 that (𝐏(S),𝒞(S),ε(S))({\bf P}(S),\mathcal{C}(S),\varepsilon(S)) is a weak chained projection category for any DRC-semigroup SS. It remains to check that any DRC-morphism ϕ:SS\phi:S\to S^{\prime} is a chained projection functor (P,𝒞,ε)(P,𝒞,ε)(P,\mathcal{C},\varepsilon)\to(P^{\prime},\mathcal{C}^{\prime},\varepsilon^{\prime}). We already know from Proposition 5.7 that ϕ\phi is a projection category morphism (P,𝒞)(P,𝒞)(P,\mathcal{C})\to(P^{\prime},\mathcal{C}^{\prime}), so it remains to check that ϕ\phi preserves the evaluation maps. But for p1,,pkPp_{1},\ldots,p_{k}\in P with p1pkp_{1}\mathrel{\mathscr{F}}\cdots\mathrel{\mathscr{F}}p_{k} we have

(ε[p1,,pk])ϕ=(p1pk)ϕ=(p1ϕ)(pkϕ)=ε[p1ϕ,,pkϕ].(\varepsilon[p_{1},\ldots,p_{k}])\phi=(p_{1}\cdots p_{k})\phi=(p_{1}\phi)\cdots(p_{k}\phi)=\varepsilon^{\prime}[p_{1}\phi,\ldots,p_{k}\phi].\qed

5.4 Chained projection categories

We have almost achieved the main objective of this section, defining chained projection categories. These will be the weak chained projection categories (P,𝒞,ε)(P,\mathcal{C},\varepsilon) satisfying a natural coherence condition stating that certain diagrams in 𝒞\mathcal{C} commute. A diagrammatic representation of this condition can be seen in Figure 2. To see that the relevant morphisms are well defined requires a preliminary definition and lemma.

In what follows, for a small category 𝒞\mathcal{C}, and objects p,qv𝒞p,q\in v\mathcal{C}, it will be convenient to write

𝒞(p,q)={a𝒞:𝐝(a)=p,𝐫(a)=q}\mathcal{C}(p,q)=\{a\in\mathcal{C}:{\bf d}(a)=p,\ {\bf r}(a)=q\}

for the set of all morphisms pqp\to q.

Definition 5.14.

Let (P,𝒞)(P,\mathcal{C}) be a projection category, let b𝒞(q,r)b\in\mathcal{C}(q,r), and let p,sPp,s\in P. We define the projections

e(p,b,s)\displaystyle e(p,b,s) =sΔbδp,\displaystyle=s\Delta_{b}\delta_{p}, e1(p,b,s)\displaystyle e_{1}(p,b,s) =sΔbδpθq,\displaystyle=s\Delta_{b}\delta_{p\theta_{q}}, f1(p,b,s)\displaystyle f_{1}(p,b,s) =sδpΘb,\displaystyle=s\delta_{p\Theta_{b}}, f(p,b,s)\displaystyle f(p,b,s) =pΘbθs,\displaystyle=p\Theta_{b}\theta_{s},
e2(p,b,s)\displaystyle e_{2}(p,b,s) =pθsΔb,\displaystyle=p\theta_{s\Delta_{b}}, f2(p,b,s)\displaystyle f_{2}(p,b,s) =pΘbθsδr.\displaystyle=p\Theta_{b}\theta_{s\delta_{r}}. (5.15)

When no confusion arises, we will abbreviate these to e=e(p,b,s)e=e(p,b,s), e1=e1(p,b,s)e_{1}=e_{1}(p,b,s), and so on.

Lemma 5.16.

Let (P,𝒞)(P,\mathcal{C}) be a projection category, and let b𝒞(q,r)b\in\mathcal{C}(q,r) and p,sPp,s\in P. Then the projections in (5.15) satisfy the following:

  1. (i)

    epe\leq p and e1,e2qe_{1},e_{2}\leq q,

  2. (ii)

    ee1,e2e\mathrel{\mathscr{F}}e_{1},e_{2},

  3. (iii)

    (pθqb)f1({}_{p\theta_{q}}{\downharpoonleft}b){\downharpoonright}_{f_{1}} exists, and has domain e1e_{1},

  4. (iv)

    fsf\leq s and f1,f2rf_{1},f_{2}\leq r,

  5. (v)

    f1,f2ff_{1},f_{2}\mathrel{\mathscr{F}}f,

  6. (vi)

    e2(bsδr){}_{e_{2}}{\downharpoonleft}(b{\downharpoonright}_{s\delta_{r}}) exists, and has range f2f_{2}.

Proof.

(i). We have

e=sΔbδpp,e1=sΔbδpθqpθqqande2=pθsΔbsΔb𝐝(b)=q.e=s\Delta_{b}\delta_{p}\leq p,\qquad e_{1}=s\Delta_{b}\delta_{p\theta_{q}}\leq p\theta_{q}\leq q\qquad\text{and}\qquad e_{2}=p\theta_{s\Delta_{b}}\leq s\Delta_{b}\leq{\bf d}(b)=q.

(ii). For ee2e\mathrel{\mathscr{F}}e_{2}, we have

eθe2=sΔbδpθpθsΔb=10sΔbδpθsΔb=9pθsΔb=e2,e\theta_{e_{2}}=s\Delta_{b}\delta_{p}\theta_{p\theta_{s\Delta_{b}}}=_{10}s\Delta_{b}\delta_{p}\theta_{s\Delta_{b}}=_{9}p\theta_{s\Delta_{b}}=e_{2},

and

e2δe=pθsΔbδsΔbδp=10pθsΔbδp=9sΔbδp=e.e_{2}\delta_{e}=p\theta_{s\Delta_{b}}\delta_{s\Delta_{b}\delta_{p}}=_{10}p\theta_{s\Delta_{b}}\delta_{p}=_{9}s\Delta_{b}\delta_{p}=e.

For ee1e\mathrel{\mathscr{F}}e_{1}, let p=qδpp^{\prime}=q\delta_{p} and q=pθqq^{\prime}=p\theta_{q}, so that pqp^{\prime}\mathrel{\mathscr{F}}q^{\prime} by Lemma 4.13(i). Since e1=sΔbδqqe_{1}=s\Delta_{b}\delta_{q^{\prime}}\leq q^{\prime}, it then follows from Corollary 4.14 that e1δpe1e_{1}\delta_{p^{\prime}}\mathrel{\mathscr{F}}e_{1}, so we can complete the proof of this part by showing that e1δp=ee_{1}\delta_{p^{\prime}}=e. For this we use Lemmas 4.13(ii) and 5.3(ii) to calculate

e1δp=sΔbδqδp=sΔbδqδp=sΔbδp=e.e_{1}\delta_{p^{\prime}}=s\Delta_{b}\delta_{q^{\prime}}\delta_{p^{\prime}}=s\Delta_{b}\delta_{q}\delta_{p}=s\Delta_{b}\delta_{p}=e.

(iii). As in the previous part we write q=pθqq^{\prime}=p\theta_{q}, noting that qq=𝐝(b)q^{\prime}\leq q={\bf d}(b). This means that qb{}_{q^{\prime}}{\downharpoonleft}b exists, and we denote this by b=qbb^{\prime}={}_{q^{\prime}}{\downharpoonleft}b. Note that

𝐫(b)=qϑb=pθqϑb=pΘb.{\bf r}(b^{\prime})=q^{\prime}\vartheta_{b}=p\theta_{q}\vartheta_{b}=p\Theta_{b}.

Since f1=sδpΘbpΘb=𝐫(b)f_{1}=s\delta_{p\Theta_{b}}\leq p\Theta_{b}={\bf r}(b^{\prime}), it follows that bf1b^{\prime}{\downharpoonright}_{f_{1}} exists, and it remains to show that this has domain e1e_{1}. For this we have

𝐝(bf1)=f1b=sδpΘbb\displaystyle{\bf d}(b^{\prime}{\downharpoonright}_{f_{1}})=f_{1}\partial_{b^{\prime}}=s\delta_{p\Theta_{b}}\partial_{b^{\prime}} =sΔb\displaystyle=s\Delta_{b^{\prime}} as pΘb=𝐫(b)p\Theta_{b}={\bf r}(b^{\prime})
=sΔbδq=e1\displaystyle=s\Delta_{b}\delta_{q^{\prime}}=e_{1} by (C1), as b=qbb^{\prime}={}_{q^{\prime}}{\downharpoonleft}b.

(iv)(vi). These are dual to (i)(ii). ∎

Definition 5.17.

Let (P,𝒞)(P,\mathcal{C}) be a projection category, and let b𝒞(q,r)b\in\mathcal{C}(q,r) and p,sPp,s\in P. Then with the projections in (5.15), it follows from Lemma 5.16 that 𝒞\mathcal{C} contains the following well-defined morphisms:

λ(p,b,s)=ε[e,e1](pθqb)f1ε[f1,f]andρ(p,b,s)=ε[e,e2]e2(bsδr)ε[f2,f].\lambda(p,b,s)=\varepsilon[e,e_{1}]\circ({}_{p\theta_{q}}{\downharpoonleft}b){\downharpoonright}_{f_{1}}\circ\varepsilon[f_{1},f]\qquad\text{and}\qquad\rho(p,b,s)=\varepsilon[e,e_{2}]\circ{}_{e_{2}}{\downharpoonleft}(b{\downharpoonright}_{s\delta_{r}})\circ\varepsilon[f_{2},f]. (5.18)

These morphisms are shown in Figure 2.

Here then is the main definition of this section:

Definition 5.19.

A chained projection category is a weak chained projection category (P,𝒞,ε)(P,\mathcal{C},\varepsilon) satisfying the following coherence condition:

  1. (C2)

    For every b𝒞b\in\mathcal{C}, and for every p,sPp,s\in P, we have λ(p,b,s)=ρ(p,b,s)\lambda(p,b,s)=\rho(p,b,s), where these morphisms are as in (5.18).

We denote by 𝐂𝐏𝐂{\bf CPC} the full subcategory of 𝐖𝐂𝐏𝐂{\bf WCPC} consisting of all chained projection categories, and all chained projection functors between them.

eee1e_{1}e2e_{2}fff1f_{1}f2f_{2}ppqqrrssbb(pθqb)f1({}_{p\theta_{q}}{\downharpoonleft}b){\downharpoonright}_{f_{1}}e2(bsδr){}_{e_{2}}{\downharpoonleft}(b{\downharpoonright}_{s\delta_{r}})ε[e,e1]\varepsilon[e,e_{1}]ε[e,e2]\varepsilon[e,e_{2}]ε[f1,f]\varepsilon[f_{1},f]ε[f2,f]\varepsilon[f_{2},f]
Figure 2: The projections and morphisms associated to b𝒞(q,r)b\in\mathcal{C}(q,r) and p,sPp,s\in P; see Definitions 5.145.17 and 5.19, and Lemma 5.16. Dashed lines indicate \leq relationships. Axiom (C2) says that the hexagon at the bottom of the diagram commutes.

For a DRC-semigroup SS we write 𝐂(S)=(𝐏(S),𝒞(S),ε(S)){\bf C}(S)=({\bf P}(S),\mathcal{C}(S),\varepsilon(S)) for the weak chained projection category from Proposition 5.13.

Theorem 5.20.

𝐂{\bf C} is a functor 𝐃𝐑𝐂𝐂𝐏𝐂{\bf DRC}\to{\bf CPC}.

Proof.

During the proof, we use (5.5) freely.

Given Proposition 5.13, we just need to check that (P,𝒞,ε)=𝐂(S)(P,\mathcal{C},\varepsilon)={\bf C}(S) satisfies (C2) for any DRC-semigroup SS. So fix b𝒞(q,r)b\in\mathcal{C}(q,r) and p,sPp,s\in P, let e,e1,e2,f,f1,f2Pe,e_{1},e_{2},f,f_{1},f_{2}\in P be as in (5.15), and write

b=(pθqb)f1andb′′=e2(bsδr).b^{\prime}=(_{p\theta_{q}}{\downharpoonleft}b){\downharpoonright}_{f_{1}}\qquad\text{and}\qquad b^{\prime\prime}={}_{e_{2}}{\downharpoonleft}(b{\downharpoonright}_{s\delta_{r}}).

By Lemma 5.16(iii) we have D(b)=e1D(b^{\prime})=e_{1}, and of course R(b)=f1R(b^{\prime})=f_{1}, so it follows from (DRC1) that b=e1bf1b^{\prime}=e_{1}b^{\prime}f_{1}, and so

λ(p,b,s)=ee1bf1f=ebf,and similarlyρ(p,b,s)=eb′′f.\lambda(p,b,s)=ee_{1}\cdot b^{\prime}\cdot f_{1}f=eb^{\prime}f,\qquad\text{and similarly}\qquad\rho(p,b,s)=eb^{\prime\prime}f.

It therefore remains to show that ebf=eb′′feb^{\prime}f=eb^{\prime\prime}f, and we claim that

ebf=ebfandeb′′f=ebf.eb^{\prime}f=ebf\qquad\text{and}\qquad eb^{\prime\prime}f=ebf. (5.21)

It suffices by symmetry to prove the first. For this we first note that b=(pθqb)f1=R(pq)bf1b^{\prime}=({}_{p\theta_{q}}{\downharpoonleft}b){\downharpoonright}_{f_{1}}=R(pq)bf_{1}, and we have

eb\displaystyle eb^{\prime} =epR(pq)bf1\displaystyle=ep\cdot R(pq)bf_{1} since epe\leq p by Lemma 5.16(i); cf. (3.5)
=epqR(pq)bf1\displaystyle=ep\cdot qR(pq)bf_{1} by Lemma 3.4
=epqbf1\displaystyle=epqbf_{1} by (DRC1)
=epbf1\displaystyle=epbf_{1} by (DRC1), since q=D(b)q=D(b). (5.22)
Next we note that
f1f\displaystyle f_{1}f =f1sf\displaystyle=f_{1}sf since fsf\leq s by Lemma 5.16(iv); cf. (3.5)
=D(R(pb)s)sf\displaystyle=D(R(pb)s)\cdot sf by the definition of f1f_{1}
=D(R(pb)s)R(pb)sf\displaystyle=D(R(pb)s)R(pb)\cdot sf by Lemma 3.4
=R(pb)sf\displaystyle=R(pb)sf by (DRC1)
=R(pb)f\displaystyle=R(pb)f using fsf\leq s again. (5.23)
Putting everything together, it follows that
ebf\displaystyle eb^{\prime}f =epbf1f\displaystyle=epbf_{1}f by (5.22)
=epbR(pb)f\displaystyle=epbR(pb)f by (5.23)
=epbf\displaystyle=epbf by (DRC1)
=ebf\displaystyle=ebf as epe\leq p.

This completes the proof of (5.21), and hence of the theorem. ∎

6 From chained projection categories to DRC-semigroups

In Section 5 we defined the category 𝐂𝐏𝐂{\bf CPC} of chained projection categories, and constructed a functor 𝐂:𝐃𝐑𝐂𝐂𝐏𝐂{\bf C}:{\bf DRC}\to{\bf CPC}. We will see in Section 7 that 𝐂{\bf C} is an isomorphism. The proof involves constructing an inverse functor 𝐒:𝐂𝐏𝐂𝐃𝐑𝐂{\bf S}:{\bf CPC}\to{\bf DRC} in the opposite direction, and that is the purpose of this section. The main part of the work is undertaken in Section 6.1, where we show how to construct a DRC semigroup from a chained projection category; see Definition 6.1 and Theorem 6.5. We then construct the functor 𝐒{\bf S} in Section 6.2; see Theorem 6.7.

6.1 The DRC-semigroup associated to a chained projection category

Definition 6.1.

Given a chained projection category (P,𝒞,ε)(P,\mathcal{C},\varepsilon), we define 𝐒(P,𝒞,ε)=(S,,D,R){\bf S}(P,\mathcal{C},\varepsilon)=(S,\bullet,D,R) to be the (2,1,1)(2,1,1)-algebra with:

  • underlying set S=𝒞S=\mathcal{C},

  • unary operations DD and RR given by D(a)=𝐝(a)D(a)={\bf d}(a) and R(a)=𝐫(a)R(a)={\bf r}(a) for a𝒞a\in\mathcal{C}, and

  • binary operation \bullet defined for a,b𝒞a,b\in\mathcal{C}, with p=𝐫(a)p={\bf r}(a) and q=𝐝(b)q={\bf d}(b), by

    ab=apε[p,q]qb,wherep=qδpandq=pθq.a\bullet b=a{\downharpoonright}_{p^{\prime}}\circ\varepsilon[p^{\prime},q^{\prime}]\circ{}_{q^{\prime}}{\downharpoonleft}b,\qquad\text{where}\qquad p^{\prime}=q\delta_{p}\quad\text{and}\quad q^{\prime}=p\theta_{q}.

    These elements and compositions are well defined because of Lemma 4.13(i). See Figure 3 for a diagrammatic representation of the product aba\bullet b.

ppqqpp^{\prime}qq^{\prime}aabbapa{\downharpoonright}_{p^{\prime}}qb{}_{q^{\prime}}{\downharpoonleft}bε[p,q]\varepsilon[p^{\prime},q^{\prime}]aba\bullet b
Figure 3: Formation of the product aba\bullet b; see Definition 6.1. Dashed lines indicate \leq relationships.

For the remainder of this section we fix a chained projection category (P,𝒞,ε)(P,\mathcal{C},\varepsilon). In Theorem 6.5 we will show that 𝐒(P,𝒞,ε){\bf S}(P,\mathcal{C},\varepsilon) is indeed a DRC-semigroup. Our first task in this direction is to show that \bullet is associative, the proof of which requires the following lemma.

Lemma 6.2.

For any a,b𝒞a,b\in\mathcal{C} we have

𝐝(ab)=𝐝(b)Δaand𝐫(ab)=𝐫(a)Θb.{\bf d}(a\bullet b)={\bf d}(b)\Delta_{a}\qquad\text{and}\qquad{\bf r}(a\bullet b)={\bf r}(a)\Theta_{b}.
Proof.

For the first (the second is dual), we keep the notation of Definition 6.1, and we have 𝐝(ab)=𝐝(ap)=pa=qδpa=qΔa=𝐝(b)Δa{\bf d}(a\bullet b)={\bf d}(a{\downharpoonright}_{p^{\prime}})=p^{\prime}\partial_{a}=q\delta_{p}\partial_{a}=q\Delta_{a}={\bf d}(b)\Delta_{a}. ∎

Lemma 6.3.

For any a,b,c𝒞a,b,c\in\mathcal{C} we have (ab)c=a(bc)(a\bullet b)\bullet c=a\bullet(b\bullet c).

Proof.

During the proof, we write

p=𝐫(a),q=𝐝(b),r=𝐫(b)ands=𝐝(c).p={\bf r}(a),\qquad q={\bf d}(b),\qquad r={\bf r}(b)\qquad\text{and}\qquad s={\bf d}(c).

We also denote the projections from Definition 5.14 by e=e(p,b,s)e=e(p,b,s), e1=e1(p,b,s)e_{1}=e_{1}(p,b,s), and so on. Given (C2), we can prove the lemma by showing that

(ab)c=aeλfcanda(bc)=aeρfc,(a\bullet b)\bullet c=a{\downharpoonright}_{e}\circ\lambda\circ{}_{f}{\downharpoonleft}c\qquad\text{and}\qquad a\bullet(b\bullet c)=a{\downharpoonright}_{e}\circ\rho\circ{}_{f}{\downharpoonleft}c,

where λ=λ(p,b,s)\lambda=\lambda(p,b,s) and ρ=ρ(p,b,s)\rho=\rho(p,b,s) are as in Definition 5.17. We just do the first, as the second follows by duality.

Writing p=qδpp^{\prime}=q\delta_{p} and q=pθqq^{\prime}=p\theta_{q}, we first note that

ab=agb,wherea=ap,g=ε[p,q]andb=qb,a\bullet b=a^{\prime}\circ g\circ b^{\prime},\qquad\text{where}\qquad a^{\prime}=a{\downharpoonright}_{p^{\prime}},\quad g=\varepsilon[p^{\prime},q^{\prime}]\quad\text{and}\quad b^{\prime}={}_{q^{\prime}}{\downharpoonleft}b,

and we also let t=𝐫(ab)=pΘbt={\bf r}(a\bullet b)=p\Theta_{b} (cf. Lemma 6.2). Then

(ab)c=(ab)tε[t,s]sc,wheret=sδtands=tθs.(a\bullet b)\bullet c=(a\bullet b){\downharpoonright}_{t^{\prime}}\circ\varepsilon[t^{\prime},s^{\prime}]\circ{}_{s^{\prime}}{\downharpoonleft}c,\qquad\text{where}\qquad t^{\prime}=s\delta_{t}\ \ \text{and}\ \ s^{\prime}=t\theta_{s}.

We also use (ii)(O4) to calculate

(ab)t=(agb)t=augvbt,(a\bullet b){\downharpoonright}_{t^{\prime}}=(a^{\prime}\circ g\circ b^{\prime}){\downharpoonright}_{t^{\prime}}=a^{\prime}{\downharpoonright}_{u}\circ g{\downharpoonright}_{v}\circ b^{\prime}{\downharpoonright}_{t^{\prime}},

where v=𝐝(bt)=tbv={\bf d}(b^{\prime}{\downharpoonright}_{t^{\prime}})=t^{\prime}\partial_{b^{\prime}} and u=𝐝(gv)=vgu={\bf d}(g{\downharpoonright}_{v})=v\partial_{g}. Next we note that

t=sδt=sδpΘb=f1ands=tθs=pΘbθs=f.t^{\prime}=s\delta_{t}=s\delta_{p\Theta_{b}}=f_{1}\qquad\text{and}\qquad s^{\prime}=t\theta_{s}=p\Theta_{b}\theta_{s}=f.

Since t=𝐫(ab)=𝐫(agb)=𝐫(b)t={\bf r}(a\bullet b)={\bf r}(a^{\prime}\circ g\circ b^{\prime})={\bf r}(b^{\prime}), we also have

v=tb=sδtb=sΔb=sΔqb=sΔbδq=e1,v=t^{\prime}\partial_{b^{\prime}}=s\delta_{t}\partial_{b^{\prime}}=s\Delta_{b^{\prime}}=s\Delta_{{}_{q^{\prime}}{\downharpoonleft}b}=s\Delta_{b}\delta_{q^{\prime}}=e_{1},

using (C1) in the second-last step. Combining this with Lemma 5.9(iii) it follows that

gv=ε[p,q]v=ε[vδp,v].g{\downharpoonright}_{v}=\varepsilon[p^{\prime},q^{\prime}]{\downharpoonright}_{v}=\varepsilon[v\delta_{p^{\prime}},v].

We then have

u=𝐝(gv)=vδp\displaystyle u={\bf d}(g{\downharpoonright}_{v})=v\delta_{p^{\prime}} =sΔbδqδp\displaystyle=s\Delta_{b}\delta_{q^{\prime}}\delta_{p^{\prime}} as shown above
=sΔbδqδp\displaystyle=s\Delta_{b}\delta_{q}\delta_{p} by Lemma 4.13(ii)
=sΔbδp=e\displaystyle=s\Delta_{b}\delta_{p}=e by Lemma 5.3(ii), noting that q=𝐝(b)q={\bf d}(b).

Putting everything together, we have

(ab)c=(ab)tε[t,s]sc\displaystyle(a\bullet b)\bullet c=(a\bullet b){\downharpoonright}_{t^{\prime}}\circ\varepsilon[t^{\prime},s^{\prime}]\circ{}_{s^{\prime}}{\downharpoonleft}c =augvbtε[t,s]sc\displaystyle=a^{\prime}{\downharpoonright}_{u}\circ g{\downharpoonright}_{v}\circ b^{\prime}{\downharpoonright}_{t^{\prime}}\circ\varepsilon[t^{\prime},s^{\prime}]\circ{}_{s^{\prime}}{\downharpoonleft}c
=apeε[e,e1](qb)f1ε[f1,f]fc\displaystyle=a{\downharpoonright}_{p^{\prime}}{\downharpoonright}_{e}\circ\varepsilon[e,e_{1}]\circ({}_{q^{\prime}}{\downharpoonleft}b){\downharpoonright}_{f_{1}}\circ\varepsilon[f_{1},f]\circ{}_{f}{\downharpoonleft}c =aeλfc,\displaystyle=a{\downharpoonright}_{e}\circ\lambda\circ{}_{f}{\downharpoonleft}c,

as required. ∎

Lemma 6.4.
  1. (i)

    If a,b𝒞a,b\in\mathcal{C} are such that 𝐫(a)=𝐝(b){\bf r}(a)={\bf d}(b), then ab=aba\bullet b=a\circ b.

  2. (ii)

    If p,qPp,q\in P, then pq=ε[p,q]p\bullet q=\varepsilon[p^{\prime},q^{\prime}], where p=qδpp^{\prime}=q\delta_{p} and q=pθqq^{\prime}=p\theta_{q}.

  3. (iii)

    If a𝒞a\in\mathcal{C}, and if p𝐝(a)p\leq{\bf d}(a) and q𝐫(a)q\leq{\bf r}(a), then pa=pap\bullet a={}_{p}{\downharpoonleft}a and aq=aqa\bullet q=a{\downharpoonright}_{q}.

Proof.

(i). Keeping the notation of Definition 6.1, the assumption here is that p=qp=q. In this case we also have p=p=q=qp^{\prime}=p=q=q^{\prime} by (P1), and then

ab=apε[p,p]pb=apb=ab.a\bullet b=a{\downharpoonright}_{p}\circ\varepsilon[p,p]\circ{}_{p}{\downharpoonleft}b=a\circ p\circ b=a\circ b.

(ii). Noting that p=𝐫(p)p={\bf r}(p) and q=𝐝(q)q={\bf d}(q), we have

pq=ppε[p,q]qq=pε[p,q]q=ε[p,q].p\bullet q=p{\downharpoonright}_{p^{\prime}}\circ\varepsilon[p^{\prime},q^{\prime}]\circ{}_{q^{\prime}}{\downharpoonleft}q=p^{\prime}\circ\varepsilon[p^{\prime},q^{\prime}]\circ q^{\prime}=\varepsilon[p^{\prime},q^{\prime}].

(iii). For the first (the second is dual), we write s=𝐝(a)s={\bf d}(a), and we have

pa=ppε[p,s]sa,wherep=sδpands=pθs.p\bullet a=p{\downharpoonright}_{p^{\prime}}\circ\varepsilon[p^{\prime},s^{\prime}]\circ{}_{s^{\prime}}{\downharpoonleft}a,\qquad\text{where}\qquad p^{\prime}=s\delta_{p}\quad\text{and}\quad s^{\prime}=p\theta_{s}.

But since psp\leq s, we have p=s=pp^{\prime}=s^{\prime}=p, so in fact

pa=ppε[p,p]pa=pppa=pa.p\bullet a=p{\downharpoonright}_{p}\circ\varepsilon[p,p]\circ{}_{p}{\downharpoonleft}a=p\circ p\circ{}_{p}{\downharpoonleft}a={}_{p}{\downharpoonleft}a.\qed
Theorem 6.5.

If (P,𝒞,ε)(P,\mathcal{C},\varepsilon) is a chained projection category, then 𝐒(P,𝒞,ε){\bf S}(P,\mathcal{C},\varepsilon) is a DRC-semigroup.

Proof.

Given Lemma 6.3, it remains to verify (DRC1)(DRC4). By symmetry, only the first part of each needs to be treated.

(DRC1). We have D(a)a=𝐝(a)a=𝐝(a)a=aD(a)\bullet a={\bf d}(a)\bullet a={\bf d}(a)\circ a=a, where we used Lemma 6.4(i), and the fact that 𝐫(𝐝(a))=𝐝(a){\bf r}({\bf d}(a))={\bf d}(a).

(DRC2). Writing p=𝐝(b)p={\bf d}(b), we use Lemma 6.2 to calculate

D(aD(b))=𝐝(ap)=𝐝(p)Δa=pΔa=𝐝(b)Δa=𝐝(ab)=D(ab).D(a\bullet D(b))={\bf d}(a\bullet p)={\bf d}(p)\Delta_{a}=p\Delta_{a}={\bf d}(b)\Delta_{a}={\bf d}(a\bullet b)=D(a\bullet b).

(DRC3). Write p=𝐝(a)p={\bf d}(a) and q=𝐝(ab)q={\bf d}(a\bullet b). We must show that pq=q=qpp\bullet q=q=q\bullet p. By Lemma 6.4(ii), we have

pq=ε[p,q],wherep=qδpandq=pθq.p\bullet q=\varepsilon[p^{\prime},q^{\prime}],\qquad\text{where}\qquad p^{\prime}=q\delta_{p}\quad\text{and}\quad q^{\prime}=p\theta_{q}.

By Lemma 6.2, and since im(Δa)p\operatorname{im}(\Delta_{a})\subseteq p^{\downarrow}, we have q=𝐝(ab)=𝐝(b)Δapq={\bf d}(a\bullet b)={\bf d}(b)\Delta_{a}\leq p, so in fact p=q=qp^{\prime}=q^{\prime}=q. Thus,

pq=ε[p,q]=ε[q,q]=q.p\bullet q=\varepsilon[p^{\prime},q^{\prime}]=\varepsilon[q,q]=q.

Essentially the same argument gives qp=qq\bullet p=q.

(DRC4). This follows from 𝐫(𝐝(a))=𝐝(a){\bf r}({\bf d}(a))={\bf d}(a). ∎

We will also need the following information concerning the projections of 𝐒(P,𝒞,ε){\bf S}(P,\mathcal{C},\varepsilon).

Proposition 6.6.

Let (P,𝒞,ε)(P,\mathcal{C},\varepsilon) be a chained projection category, and let S=𝐒(P,𝒞,ε)S={\bf S}(P,\mathcal{C},\varepsilon). Then

  1. (i)

    𝐏(S)=P{\bf P}(S)=P,

  2. (ii)

    D(pq)=qδpD(p\bullet q)=q\delta_{p} and R(pq)=pθqR(p\bullet q)=p\theta_{q} for all p,qPp,q\in P.

Proof.

(i). We have 𝐏(S)={D(a):aS}={𝐝(a):a𝒞}=v𝒞=P{\bf P}(S)=\{D(a):a\in S\}=\{{\bf d}(a):a\in\mathcal{C}\}=v\mathcal{C}=P.

(ii). This follows from combining Lemma 6.4(ii) with Lemma 5.9(ii). ∎

6.2 A functor

We can think of the construction of the DRC-semigroup 𝐒(P,𝒞,ε){\bf S}(P,\mathcal{C},\varepsilon) from the chained projection category (P,𝒞,ε)(P,\mathcal{C},\varepsilon) as an object map 𝐂𝐏𝐂𝐃𝐑𝐂{\bf CPC}\to{\bf DRC}. At the level of morphisms, any chained projection functor ϕ:(P,𝒞,ε)(P,𝒞,ε)\phi:(P,\mathcal{C},\varepsilon)\to(P^{\prime},\mathcal{C}^{\prime},\varepsilon^{\prime}) is a map 𝒞𝒞\mathcal{C}\to\mathcal{C}^{\prime}. Since the underlying sets of the semigroups S=𝐒(P,𝒞,ε)S={\bf S}(P,\mathcal{C},\varepsilon) and S=𝐒(P,𝒞,ε)S^{\prime}={\bf S}(P^{\prime},\mathcal{C}^{\prime},\varepsilon^{\prime}) are 𝒞\mathcal{C} and 𝒞\mathcal{C}^{\prime}, respectively, we can think of ϕ\phi as a map ϕ=𝐒(ϕ):SS\phi={\bf S}(\phi):S\to S^{\prime}. We now show that 𝐒:𝐂𝐏𝐂𝐃𝐑𝐂{\bf S}:{\bf CPC}\to{\bf DRC}, interpreted in this way, is a functor.

Theorem 6.7.

𝐒{\bf S} is a functor 𝐂𝐏𝐂𝐃𝐑𝐂{\bf CPC}\to{\bf DRC}.

Proof.

It remains only to check that a morphism ϕ:(P,𝒞,ε)(P,𝒞,ε)\phi:(P,\mathcal{C},\varepsilon)\to(P^{\prime},\mathcal{C}^{\prime},\varepsilon^{\prime}) in 𝐂𝐏𝐂{\bf CPC} is also a morphism SSS\to S^{\prime} in 𝐃𝐑𝐂{\bf DRC}, where S=𝐒(P,𝒞,ε)S={\bf S}(P,\mathcal{C},\varepsilon) and S=𝐒(P,𝒞,ε)S^{\prime}={\bf S}(P^{\prime},\mathcal{C}^{\prime},\varepsilon^{\prime}). For simplicity, we will write a¯=aϕ\overline{a}=a\phi for aS(=𝒞)a\in S\ (=\mathcal{C}). We also use dashes to distinguish between the various parameters and operations on 𝒞\mathcal{C} and 𝒞\mathcal{C}^{\prime}, and on SS and SS^{\prime}.

If aSa\in S, then writing p=D(a)=𝐝(a)p=D(a)={\bf d}(a), we have

D(a¯)=𝐝(a¯)=𝐝(pa¯)=𝐝(p¯a¯)=𝐝(p¯)=p¯=D(a)¯,and similarlyR(a¯)=R(a)¯.D^{\prime}(\overline{a})={\bf d}^{\prime}(\overline{a})={\bf d}^{\prime}(\overline{p\circ a})={\bf d}^{\prime}(\overline{p}\circ^{\prime}\overline{a})={\bf d}^{\prime}(\overline{p})=\overline{p}=\overline{D(a)},\qquad\text{and similarly}\qquad R^{\prime}(\overline{a})=\overline{R(a)}.

It therefore remains to show that ab¯=a¯b¯\overline{a\bullet b}=\overline{a}\bullet^{\prime}\overline{b} for all a,bSa,b\in S. So fix some such a,ba,b, and let

p=𝐫(a),q=𝐝(b),p=qδpandq=pθq.p={\bf r}(a),\qquad q={\bf d}(b),\qquad p^{\prime}=q\delta_{p}\qquad\text{and}\qquad q^{\prime}=p\theta_{q}.

As shown above, we have 𝐫(a¯)=𝐫(a)¯=p¯{\bf r}^{\prime}(\overline{a})=\overline{{\bf r}(a)}=\overline{p}, and similarly 𝐝(b¯)=q¯{\bf d}^{\prime}(\overline{b})=\overline{q}. It follows that

ab=apε[p,q]qbanda¯b¯=a¯p¯ε[p¯,q¯]q¯b¯,a\bullet b=a{\downharpoonright}_{p^{\prime}}\circ\varepsilon[p^{\prime},q^{\prime}]\circ{}_{q^{\prime}}{\downharpoonleft}b\qquad\text{and}\qquad\overline{a}\bullet^{\prime}\overline{b}=\overline{a}{\downharpoonright}_{\overline{p}^{\prime}}\circ^{\prime}\varepsilon^{\prime}[\overline{p}^{\prime},\overline{q}^{\prime}]\circ^{\prime}{}_{\overline{q}^{\prime}}{\downharpoonleft}\overline{b},

where p¯=q¯δp¯\overline{p}^{\prime}=\overline{q}\delta^{\prime}_{\overline{p}} and q¯=p¯θq¯\overline{q}^{\prime}=\overline{p}\theta^{\prime}_{\overline{q}}. Since vϕ:PPv\phi:P\to P^{\prime} is a projection algebra morphism, we have

p¯=qδp¯=q¯δp¯=p¯,and similarlyq¯=q¯.\overline{p^{\prime}}=\overline{q\delta_{p}}=\overline{q}\delta^{\prime}_{\overline{p}}=\overline{p}^{\prime},\qquad\text{and similarly}\qquad\overline{q^{\prime}}=\overline{q}^{\prime}.

Since ϕ\phi respects evaluation maps, we have

ε[p,q]¯=ε[p¯,q¯]=ε[p¯,q¯].\overline{\varepsilon[p^{\prime},q^{\prime}]}=\varepsilon^{\prime}[\overline{p^{\prime}},\overline{q^{\prime}}]=\varepsilon^{\prime}[\overline{p}^{\prime},\overline{q}^{\prime}].

Since ϕ\phi is a biordered functor, we have

ap¯=a¯p¯=a¯p¯,and similarlyqb¯=q¯b¯.\overline{a{\downharpoonright}_{p^{\prime}}}=\overline{a}{\downharpoonright}_{\overline{p^{\prime}}}=\overline{a}{\downharpoonright}_{\overline{p}^{\prime}},\qquad\text{and similarly}\qquad\overline{{}_{q^{\prime}}{\downharpoonleft}b}={}_{\overline{q}^{\prime}}{\downharpoonleft}\overline{b}.

Putting everything together, we have

ab¯=apε[p,q]qb¯=ap¯ε[p,q]¯qb¯=a¯p¯ε[p¯,q¯]q¯b¯=a¯b¯.\overline{a\bullet b}=\overline{a{\downharpoonright}_{p^{\prime}}\circ\varepsilon[p^{\prime},q^{\prime}]\circ{}_{q^{\prime}}{\downharpoonleft}b}=\overline{a{\downharpoonright}_{p^{\prime}}}\circ^{\prime}\overline{\varepsilon[p^{\prime},q^{\prime}]}\circ^{\prime}\overline{{}_{q^{\prime}}{\downharpoonleft}b}=\overline{a}{\downharpoonright}_{\overline{p}^{\prime}}\circ^{\prime}\varepsilon^{\prime}[\overline{p}^{\prime},\overline{q}^{\prime}]\circ^{\prime}{}_{\overline{q}^{\prime}}{\downharpoonleft}\overline{b}=\overline{a}\bullet^{\prime}\overline{b}.\qed

7 The category isomorphism

In this section we will show that the functors

𝐂:𝐃𝐑𝐂𝐂𝐏𝐂and𝐒:𝐂𝐏𝐂𝐃𝐑𝐂{\bf C}:{\bf DRC}\to{\bf CPC}\qquad\text{and}\qquad{\bf S}:{\bf CPC}\to{\bf DRC}

from Theorems 5.20 and 6.7 are mutually inverse isomorphisms, thereby proving the main result of the paper:

Theorem 7.1.

The category 𝐃𝐑𝐂{\bf DRC} of DRC-semigroups (with DRC-morphisms) is isomorphic to the category 𝐂𝐏𝐂{\bf CPC} of chained projection categories (with chained projection functors).

Proof.

Since the functors 𝐂{\bf C} and 𝐒{\bf S} both act identically on morphisms, we just need to show that

𝐂(𝐒(P,𝒞,ε))=(P,𝒞,ε)and𝐒(𝐂(S))=S{\bf C}({\bf S}(P,\mathcal{C},\varepsilon))=(P,\mathcal{C},\varepsilon)\qquad\text{and}\qquad{\bf S}({\bf C}(S))=S

for any chained projection category (P,𝒞,ε)(P,\mathcal{C},\varepsilon), and any DRC-semigroup SS. We do this in Propositions 7.2 and 7.3. ∎

Proposition 7.2.

For any chained projection category (P,𝒞,ε)(P,\mathcal{C},\varepsilon) we have 𝐂(𝐒(P,𝒞,ε))=(P,𝒞,ε){\bf C}({\bf S}(P,\mathcal{C},\varepsilon))=(P,\mathcal{C},\varepsilon).

Proof.

Throughout the proof we will write

S=𝐒(P,𝒞,ε)and(P,𝒞,ε)=𝐂(S)=(𝐏(S),𝒞(S),ε(S)).S={\bf S}(P,\mathcal{C},\varepsilon)\qquad\text{and}\qquad(P^{\prime},\mathcal{C}^{\prime},\varepsilon^{\prime})={\bf C}(S)=({\bf P}(S),\mathcal{C}(S),\varepsilon(S)).

To prove the result we need to show that:

  1. (i)

    P=PP=P^{\prime} as projection algebras,

  2. (ii)

    𝒞=𝒞\mathcal{C}=\mathcal{C}^{\prime} as biordered categories, and

  3. (iii)

    ε=ε\varepsilon=\varepsilon^{\prime} as maps.

As usual we use dashes in what follows to denote the various parameters and operations on 𝒞\mathcal{C}^{\prime}.

(i). This follows immediately from Proposition 6.6.

(ii). By construction, the underlying set of 𝒞\mathcal{C}^{\prime} is the same as that of SS and 𝒞\mathcal{C}. We also have v𝒞=𝐏(S)=P=v𝒞v\mathcal{C}^{\prime}={\bf P}(S)=P=v\mathcal{C}, and

𝐝(a)=D(a)=𝐝(a),and similarly𝐫(a)=𝐫(a)for all a𝒞(=𝒞).{\bf d}^{\prime}(a)=D(a)={\bf d}(a),\qquad\text{and similarly}\qquad{\bf r}^{\prime}(a)={\bf r}(a)\qquad\text{for all $a\in\mathcal{C}^{\prime}\ (=\mathcal{C})$.}

It follows that for any a,b𝒞a,b\in\mathcal{C}^{\prime} with 𝐫(a)=𝐝(b){\bf r}^{\prime}(a)={\bf d}^{\prime}(b), we have 𝐫(a)=𝐝(b){\bf r}(a)={\bf d}(b), and Lemma 6.4(i) then gives

ab=ab=ab.a\circ^{\prime}b=a\bullet b=a\circ b.

It remains to check that the orders on 𝒞\mathcal{C} and 𝒞\mathcal{C}^{\prime} coincide, and for this it suffices to show that

pa=paandaq=aqfor all a𝒞(=𝒞), and all p𝐝(a) and q𝐫(a).{}_{p}{\upharpoonleft}a={}_{p}{\downharpoonleft}a\qquad\text{and}\qquad a{\upharpoonright}_{q}=a{\downharpoonright}_{q}\qquad\text{for all $a\in\mathcal{C}^{\prime}\ (=\mathcal{C})$, and all $p\leq{\bf d}(a)$ and $q\leq{\bf r}(a)$.}

Here we write {\upharpoonleft} and {\upharpoonright} for restrictions in 𝒞\mathcal{C}^{\prime}, and {\downharpoonleft} and {\downharpoonright} for restrictions in 𝒞\mathcal{C}. For this we use Lemma 6.4(iii) to calculate

pa=pa=pa,and similarlyaq=aq.{}_{p}{\upharpoonleft}a=p\bullet a={}_{p}{\downharpoonleft}a,\qquad\text{and similarly}\qquad a{\upharpoonright}_{q}=a{\downharpoonright}_{q}.

(iii). We must show that ε(𝔠)=ε(𝔠)\varepsilon^{\prime}(\mathfrak{c})=\varepsilon(\mathfrak{c}) for any chain 𝔠𝒞(P)\mathfrak{c}\in\mathcal{C}(P). So fix some such 𝔠=[p1,,pk]{\mathfrak{c}=[p_{1},\ldots,p_{k}]}. First observe that for any 1i<k1\leq i<k, Lemma 6.4(ii) gives

pipi+1=ε[pi+1δpi,piθpi+1]=ε[pi,pi+1],as pipi+1.p_{i}\bullet p_{i+1}=\varepsilon[p_{i+1}\delta_{p_{i}},p_{i}\theta_{p_{i+1}}]=\varepsilon[p_{i},p_{i+1}],\qquad\text{as $p_{i}\mathrel{\mathscr{F}}p_{i+1}$.}

We then have

ε(𝔠)\displaystyle\varepsilon^{\prime}(\mathfrak{c}) =p1pk\displaystyle=p_{1}\bullet\cdots\bullet p_{k} by definition
=(p1p2)(p2p3)(pk1pk)\displaystyle=(p_{1}\bullet p_{2})\bullet(p_{2}\bullet p_{3})\bullet\cdots\bullet(p_{k-1}\bullet p_{k}) as each pip_{i} is an idempotent of SS
=ε[p1,p2]ε[p2,p3]ε[pk1,pk]\displaystyle=\varepsilon[p_{1},p_{2}]\bullet\varepsilon[p_{2},p_{3}]\bullet\cdots\bullet\varepsilon[p_{k-1},p_{k}] as just observed
=ε[p1,p2]ε[p2,p3]ε[pk1,pk]\displaystyle=\varepsilon[p_{1},p_{2}]\circ\varepsilon[p_{2},p_{3}]\circ\cdots\circ\varepsilon[p_{k-1},p_{k}] by Lemma 6.4(i)
=ε[p1,p2,p3,,pk]=ε(𝔠)\displaystyle=\varepsilon[p_{1},p_{2},p_{3},\ldots,p_{k}]=\varepsilon(\mathfrak{c}) as ε\varepsilon is a functor.

This completes the proof. ∎

Proposition 7.3.

For any DRC-semigroup SS we have 𝐒(𝐂(S))=S{\bf S}({\bf C}(S))=S.

Proof.

This time we write

(P,𝒞,ε)=𝐂(S)=(𝐏(S),𝒞(S),ε(S))andS=𝐒(P,𝒞,ε).(P,\mathcal{C},\varepsilon)={\bf C}(S)=({\bf P}(S),\mathcal{C}(S),\varepsilon(S))\qquad\text{and}\qquad S^{\prime}={\bf S}(P,\mathcal{C},\varepsilon).

Again the underlying sets of SS and SS^{\prime} are the same, and so are the unary operations. It therefore remains to show that

ab=abfor all a,bS(=S).a\bullet b=ab\qquad\text{for all $a,b\in S^{\prime}\ (=S)$.}

For this, write p=𝐫(a)=R(a)p={\bf r}(a)=R(a) and q=𝐝(b)=D(b)q={\bf d}(b)=D(b). Then with p=qδp=D(pq)p^{\prime}=q\delta_{p}=D(pq) and q=pθq=R(pq)q^{\prime}=p\theta_{q}=R(pq), we have

ab=apε[p,q]qb=(ap)(pq)(qb)=apqb.a\bullet b=a{\downharpoonright}_{p^{\prime}}\circ\varepsilon[p^{\prime},q^{\prime}]\circ{}_{q^{\prime}}{\downharpoonleft}b=(ap^{\prime})\cdot(p^{\prime}q^{\prime})\cdot(q^{\prime}b)=ap^{\prime}q^{\prime}b.

We then use Lemma 3.4 and (DRC1) to calculate

pq=D(pq)R(pq)=D(pq)pqR(pq)=pq.p^{\prime}q^{\prime}=D(pq)\cdot R(pq)=D(pq)p\cdot qR(pq)=pq. (7.4)

Thus, continuing from above, we have

ab=apqb=apqb=ab,a\bullet b=ap^{\prime}q^{\prime}b=ap\cdot qb=ab,

by (DRC1), as p=R(a)p=R(a) and q=D(b)q=D(b). ∎

8 Free and fundamental projection-generated DRC-semigroups

Theorem 7.1 shows that the categories 𝐃𝐑𝐂{\bf DRC} and 𝐂𝐏𝐂{\bf CPC} are isomorphic. It also follows from the definition of the functors 𝐂:𝐃𝐑𝐂𝐂𝐏𝐂{\bf C}:{\bf DRC}\to{\bf CPC} and 𝐒:𝐂𝐏𝐂𝐃𝐑𝐂{\bf S}:{\bf CPC}\to{\bf DRC} that the algebras of projections of DRC-semigroups are precisely the same as the (structured) object sets of chained projection categories. But the theorem does not answer the following:

Question 8.1.

Given an abstract projection algebra PP, as in Definition 4.1, does there exist a DRC-semigroup SS with 𝐏(S)=P{\bf P}(S)=P? Or equivalently, does there exist a chained projection category (P,𝒞,ε)(P,\mathcal{C},\varepsilon)?

The semigroup version of this question was answered in the affirmative by Jones in his study of fundamental DRC-semigroups [30], i.e. the DRC-semigroups with no non-trivial projection-separating congruences. In [30, Section 9], Jones constructed a DRC-semigroup, denoted C(P)C(P), from a projection algebra PP, and showed it to be the maximum fundamental DRC-semigroup with projection algebra PP, meaning that:

  • C(P)C(P) is fundamental, and has projection algebra P=𝐏(C(P))P={\bf P}(C(P)), and

  • any fundamental DRC-semigroup with projection algebra PP embeds canonically into C(P)C(P).

In a sense, one of the purposes of the current section is to provide another way to answer Question 8.1, in that we construct two different DRC-semigroups, denoted FPF_{P} and MPM_{P}, both with projection algebra PP. However, we have a deeper purpose than this (as did Jones), and our main objective here is to demonstrate the existence of free (projection-generated) DRC-semigroups. Specifically, we show that the assignment PFPP\mapsto F_{P} is the object part of a functor 𝐅:𝐏𝐀𝐃𝐑𝐂{\bf F}:{\bf PA}\to{\bf DRC}, and that this is in fact a left adjoint to the forgetful functor 𝐏:𝐃𝐑𝐂𝐏𝐀{\bf P}:{\bf DRC}\to{\bf PA} from Proposition 4.19, which maps a DRC-semigroup SS to its underlying projection algebra 𝐏(S){\bf P}(S). It follows that FPF_{P} is the free (projection-generated) DRC-semigroup with projection algebra PP.

The semigroup FPF_{P} is defined abstractly in terms of a presentation by generators and relations. By contrast, the semigroup MPM_{P} is a concrete semigroup consisting of pairs of self-maps of PP, built from the underlying θp\theta_{p} and δp\delta_{p} operations of PP. The main initial purpose of MPM_{P} is to provide a concrete homomorphic image of FPF_{P}, which will be useful in certain proofs concerning the latter. But we will also see that MPM_{P} is very special in its own right, in that it is the unique projection-generated fundamental DRC-semigroup with projection algebra PP.

The definitions of FPF_{P} and MPM_{P} are given in Section 8.1, and their key properties are established in Theorem 8.11. Categorical freeness of FPF_{P} is established in Section 8.2; see Theorem 8.17. The unique fundamentality properties of MPM_{P} are established in Section 8.3; see Theorem 8.25.

Before we begin, we briefly revise our notation for semigroup presentations. Let XX be an alphabet, and write X+X^{+} for the free semigroup over XX, which is the semigroup of all non-empty words over XX, under concatenation. Also let RX+×X+R\subseteq X^{+}\times X^{+} be a set of pairs of words over XX, and write RR^{\sharp} for the congruence on X+X^{+} generated by RR. We then write

X:R=X+/R,\langle X:R\rangle=X^{+}/R^{\sharp},

and call this the semigroup defined by the presentation X:R\langle X:R\rangle. Elements of XX are called letters or generators, and elements of RR are called relations. A relation (u,v)R(u,v)\in R is often written as an equality, u=vu=v.

8.1 Definition of 𝑭𝑷F_{P} and 𝑴𝑷M_{P}

Fix a projection algebra PP for the duration of this section. Also fix an alphabet

XP={xp:pP}X_{P}=\{x_{p}:p\in P\}

in one-one correspondence with PP, and let RPR_{P} be the set of relations, quantified over all p,qPp,q\in P:

  1. (R1)

    xp2=xpx_{p}^{2}=x_{p},

  2. (R2)

    xpxq=xpxpθqx_{p}x_{q}=x_{p}x_{p\theta_{q}},

  3. (R3)

    xpxq=xqδpxqx_{p}x_{q}=x_{q\delta_{p}}x_{q}.

Write =RP{\sim}=R_{P}^{\sharp} for the congruence on XP+X_{P}^{+} generated by the relations (R1)(R3), and denote by w¯\overline{w} the \sim-class of a word wXP+w\in X_{P}^{+}. We then take FPF_{P} to be the semigroup defined by the presentation

FP=XP:RP=XP+/={w¯:wXP+}.F_{P}=\langle X_{P}:R_{P}\rangle=X_{P}^{+}/{\sim}=\{\overline{w}:w\in X_{P}^{+}\}.

We will soon see that FPF_{P} is a DRC-semigroup with projection algebra (isomorphic to) PP.

We also define a second semigroup associated to PP. First, we denote by 𝒯P\mathcal{T}_{P} the full transformation semigroup over PP, i.e. the semigroup of all maps PPP\to P under composition. As before, we also write 𝒯Pop\mathcal{T}_{P}^{\operatorname{op}} for the opposite semigroup. Our second semigroup will be a subsemigroup of the direct product 𝒯P×𝒯Pop\mathcal{T}_{P}\times\mathcal{T}_{P}^{\operatorname{op}}, in which the operation is given by (α,α)(β,β)=(αβ,βα)(\alpha,\alpha^{\prime})(\beta,\beta^{\prime})=(\alpha\beta,\beta^{\prime}\alpha^{\prime}). For each pPp\in P, we define the pair

p^=(θp,δp)𝒯P×𝒯Pop,\widehat{p}=(\theta_{p},\delta_{p})\in\mathcal{T}_{P}\times\mathcal{T}_{P}^{\operatorname{op}},

and the semigroup

MP=p^:pP𝒯P×𝒯Pop.M_{P}=\langle\widehat{p}:p\in P\rangle\leq\mathcal{T}_{P}\times\mathcal{T}_{P}^{\operatorname{op}}.

A typical element of MPM_{P} has the form

p^1p^k=(θp1θpk,δpkδp1)for p1,,pkP.\widehat{p}_{1}\cdots\widehat{p}_{k}=(\theta_{p_{1}}\cdots\theta_{p_{k}},\delta_{p_{k}}\cdots\delta_{p_{1}})\qquad\text{for $p_{1},\ldots,p_{k}\in P$.}

Define the (surjective) semigroup homomorphism

ψ:XP+MP:xpp^=(θp,δp).\psi:X_{P}^{+}\to M_{P}:x_{p}\mapsto\widehat{p}=(\theta_{p},\delta_{p}).
Lemma 8.2.

We have RPker(ψ)R_{P}\subseteq\ker(\psi).

Proof.

We need to check that each relation from RPR_{P} is preserved by ψ\psi, in the sense that uψ=vψu\psi=v\psi for each such relation (u,v)RP(u,v)\in R_{P}. To do so, let p,qPp,q\in P. For (R1) we have

xp2ψ=p^p^=(θpθp,δpδp)=(θp,δp)=p^=xpψ,x_{p}^{2}\psi=\widehat{p}\cdot\widehat{p}=(\theta_{p}\theta_{p},\delta_{p}\delta_{p})=(\theta_{p},\delta_{p})=\widehat{p}=x_{p}\psi,

by (P6). For (R2) we use (P3) and (P5) to calculate:

(xpxq)ψ=p^q^=(θpθq,δqδp)=(θpθpθq,δpθqδp)=p^pθq^=(xpxpθq)ψ.(x_{p}x_{q})\psi=\widehat{p}\cdot\widehat{q}=(\theta_{p}\theta_{q},\delta_{q}\delta_{p})=(\theta_{p}\theta_{p\theta_{q}},\delta_{p\theta_{q}}\delta_{p})=\widehat{p}\cdot\widehat{p\theta_{q}}=(x_{p}x_{p\theta_{q}})\psi.

We deal with (R3) in similar fashion. ∎

It follows that ψ\psi induces a well-defined (surjective) semigroup homomorphism

Ψ:FPMPgiven byw¯Ψ=wψfor wXP+.\Psi:F_{P}\to M_{P}\qquad\text{given by}\qquad\overline{w}\Psi=w\psi\qquad\text{for $w\in X_{P}^{+}$.}

Note in particular that x¯pΨ=p^=(θp,δp)\overline{x}_{p}\Psi=\widehat{p}=(\theta_{p},\delta_{p}) for pPp\in P.

In what follows, we write 1\sim_{1} to denote \sim-equivalence of words from XP+X_{P}^{+} by one or more applications of relations from (R1), with similar meanings for 2\sim_{2} and 3\sim_{3}.

Lemma 8.3.

If p,qPp,q\in P, then xpxqxpxqx_{p}x_{q}\sim x_{p^{\prime}}x_{q^{\prime}}, where p=qδpp^{\prime}=q\delta_{p} and q=pθqq^{\prime}=p\theta_{q}.

Proof.

We have xpxq2xpxpθq3x(pθq)δpxpθq=xqδpxpθqx_{p}x_{q}\sim_{2}x_{p}x_{p\theta_{q}}\sim_{3}x_{(p\theta_{q})\delta_{p}}x_{p\theta_{q}}=x_{q\delta_{p}}x_{p\theta_{q}}, using (P9) in the last step. ∎

For a path 𝔭=(p1,,pk)\mathfrak{p}=(p_{1},\ldots,p_{k}) from the path category 𝒫=𝒫(P)\mathscr{P}=\mathscr{P}(P), we define the word

w𝔭=xp1xpkXP+.w_{\mathfrak{p}}=x_{p_{1}}\cdots x_{p_{k}}\in X_{P}^{+}.
Lemma 8.4.

Any word over XPX_{P} is \sim-equivalent to w𝔭w_{\mathfrak{p}} for some 𝔭𝒫\mathfrak{p}\in\mathscr{P}.

Proof.

Consider an arbitrary word w=xp1xpkXP+w=x_{p_{1}}\cdots x_{p_{k}}\in X_{P}^{+}. We will show by induction on kk that

wxp1xpkfor some pipi with p1pk.w\sim x_{p_{1}^{\prime}}\cdots x_{p_{k}^{\prime}}\qquad\text{for some $p_{i}^{\prime}\leq p_{i}$ with $p_{1}^{\prime}\mathrel{\mathscr{F}}\cdots\mathrel{\mathscr{F}}p_{k}^{\prime}$.}

The k=1k=1 case is trivial, so we now assume that k2k\geq 2. First, Lemma 8.3 gives

xpk1xpkxpk1′′xpk′′,wherepk1′′=pkδpk1andpk′′=pk1θpk,x_{p_{k-1}}x_{p_{k}}\sim x_{p_{k-1}^{\prime\prime}}x_{p_{k}^{\prime\prime}},\qquad\text{where}\qquad p_{k-1}^{\prime\prime}=p_{k}\delta_{p_{k-1}}\quad\text{and}\quad p_{k}^{\prime\prime}=p_{k-1}\theta_{p_{k}},

and by Lemma 4.13(i) we have pk1′′pk1p_{k-1}^{\prime\prime}\leq p_{k-1}, pk′′pkp_{k}^{\prime\prime}\leq p_{k} and pk1′′pk′′p_{k-1}^{\prime\prime}\mathrel{\mathscr{F}}p_{k}^{\prime\prime}. By induction, we have

xp1xpk2xpk1′′xp1xpk2xpk1,x_{p_{1}}\cdots x_{p_{k-2}}x_{p_{k-1}^{\prime\prime}}\sim x_{p_{1}^{\prime}}\cdots x_{p_{k-2}^{\prime}}x_{p_{k-1}^{\prime}},

for some pipip_{i}^{\prime}\leq p_{i} (i=1,,k2i=1,\ldots,k-2) and pk1pk1′′p_{k-1}^{\prime}\leq p_{k-1}^{\prime\prime}, with p1pk1p_{1}^{\prime}\mathrel{\mathscr{F}}\cdots\mathrel{\mathscr{F}}p_{k-1}^{\prime}. We also have

xpk1xpk′′2xpk1xpk,wherepk=pk1θpk′′.x_{p_{k-1}^{\prime}}x_{p_{k}^{\prime\prime}}\sim_{2}x_{p_{k-1}^{\prime}}x_{p_{k}^{\prime}},\qquad\text{where}\qquad p_{k}^{\prime}=p_{k-1}^{\prime}\theta_{p_{k}^{\prime\prime}}.

Since pk1pk1′′pk′′p_{k-1}^{\prime}\leq p_{k-1}^{\prime\prime}\mathrel{\mathscr{F}}p_{k}^{\prime\prime}, it follows from Corollary 4.14 that pk1pkp_{k-1}^{\prime}\mathrel{\mathscr{F}}p_{k}^{\prime}. Putting everything together, we have

w=xp1xpkxp1xpk2xpk1′′xpk′′xp1xpk2xpk1xpk′′xp1xpk2xpk1xpk,w=x_{p_{1}}\cdots x_{p_{k}}\sim x_{p_{1}}\cdots x_{p_{k-2}}x_{p_{k-1}^{\prime\prime}}x_{p_{k}^{\prime\prime}}\sim x_{p_{1}^{\prime}}\cdots x_{p_{k-2}^{\prime}}x_{p_{k-1}^{\prime}}x_{p_{k}^{\prime\prime}}\sim x_{p_{1}^{\prime}}\cdots x_{p_{k-2}^{\prime}}x_{p_{k-1}^{\prime}}x_{p_{k}^{\prime}},

with all conditions met. (Note that pipi′′pip_{i}^{\prime}\leq p_{i}^{\prime\prime}\leq p_{i} for i=k1,ki=k-1,k.) ∎

For a path 𝔭=(p1,,pk)𝒫\mathfrak{p}=(p_{1},\ldots,p_{k})\in\mathscr{P}, we define

𝔭^=w𝔭ψ=p^1p^k=(θp1θpk,δpkδp1)MP.\widehat{\mathfrak{p}}=w_{\mathfrak{p}}\psi=\widehat{p}_{1}\cdots\widehat{p}_{k}=(\theta_{p_{1}}\cdots\theta_{p_{k}},\delta_{p_{k}}\cdots\delta_{p_{1}})\in M_{P}.

It follows from Lemma 8.4, and surjectivity of Ψ\Psi, that

FP={w¯𝔭:𝔭𝒫}andMP={𝔭^:𝔭𝒫}.F_{P}=\{\overline{w}_{\mathfrak{p}}:\mathfrak{p}\in\mathscr{P}\}\qquad\text{and}\qquad M_{P}=\{\widehat{\mathfrak{p}}:\mathfrak{p}\in\mathscr{P}\}.
Lemma 8.5.

If 𝔭,𝔮𝒫\mathfrak{p},\mathfrak{q}\in\mathscr{P}, then

  1. (i)

    𝔭^=𝔮^𝐝(𝔭)=𝐝(𝔮)\widehat{\mathfrak{p}}=\widehat{\mathfrak{q}}\ \Rightarrow\ {\bf d}(\mathfrak{p})={\bf d}(\mathfrak{q}) and 𝐫(𝔭)=𝐫(𝔮){\bf r}(\mathfrak{p})={\bf r}(\mathfrak{q}),

  2. (ii)

    w¯𝔭=w¯𝔮𝐝(𝔭)=𝐝(𝔮)\overline{w}_{\mathfrak{p}}=\overline{w}_{\mathfrak{q}}\ \Rightarrow\ {\bf d}(\mathfrak{p})={\bf d}(\mathfrak{q}) and 𝐫(𝔭)=𝐫(𝔮){\bf r}(\mathfrak{p})={\bf r}(\mathfrak{q}).

Proof.

Since w¯𝔭=w¯𝔮w𝔭w𝔮𝔭^=w𝔭ψ=w𝔮ψ=𝔮^\overline{w}_{\mathfrak{p}}=\overline{w}_{\mathfrak{q}}\ \Rightarrow\ w_{\mathfrak{p}}\sim w_{\mathfrak{q}}\ \Rightarrow\ \widehat{\mathfrak{p}}=w_{\mathfrak{p}}\psi=w_{\mathfrak{q}}\psi=\widehat{\mathfrak{q}}, it suffices to prove (i). For this, write 𝔭=(p1,,pk){\mathfrak{p}=(p_{1},\ldots,p_{k})} and 𝔮=(q1,,ql)\mathfrak{q}=(q_{1},\ldots,q_{l}), so that

𝔭^=(θp1θpk,δpkδp1)and𝔮^=(θq1θql,δqlδq1).\widehat{\mathfrak{p}}=(\theta_{p_{1}}\cdots\theta_{p_{k}},\delta_{p_{k}}\cdots\delta_{p_{1}})\qquad\text{and}\qquad\widehat{\mathfrak{q}}=(\theta_{q_{1}}\cdots\theta_{q_{l}},\delta_{q_{l}}\cdots\delta_{q_{1}}).

It then follows from Lemma 4.15 that 𝔭^=𝔮^\widehat{\mathfrak{p}}=\widehat{\mathfrak{q}} implies p1=q1p_{1}=q_{1} and pk=qlp_{k}=q_{l}. ∎

It follows that we have well-defined maps D,R:FPFPD,R:F_{P}\to F_{P} and D,R:MPMPD,R:M_{P}\to M_{P} given by

D(w¯𝔭)\displaystyle D(\overline{w}_{\mathfrak{p}}) =x¯𝐝(𝔭)=x¯p1,\displaystyle=\overline{x}_{{\bf d}(\mathfrak{p})}=\overline{x}_{p_{1}}, D(𝔭^)\displaystyle D(\widehat{\mathfrak{p}}) =𝐝(𝔭)^=p^1=(θp1,δp1),\displaystyle=\widehat{{\bf d}(\mathfrak{p})}=\widehat{p}_{1}=(\theta_{p_{1}},\delta_{p_{1}}),
R(w¯𝔭)\displaystyle R(\overline{w}_{\mathfrak{p}}) =x¯𝐫(𝔭)=x¯pk,\displaystyle=\overline{x}_{{\bf r}(\mathfrak{p})}=\overline{x}_{p_{k}}, R(𝔭^)\displaystyle R(\widehat{\mathfrak{p}}) =𝐫(𝔭)^=p^k=(θpk,δpk),\displaystyle=\widehat{{\bf r}(\mathfrak{p})}=\widehat{p}_{k}=(\theta_{p_{k}},\delta_{p_{k}}), for 𝔭=(p1,,pk)𝒫\mathfrak{p}=(p_{1},\ldots,p_{k})\in\mathscr{P}. (8.6)

There should be no confusion in using DD and RR to denote these operations on both semigroups FPF_{P} and MPM_{P}, as they take on different kinds of arguments in the two semigroups. Note also that

D(w¯𝔭)Ψ=(x¯𝐝(𝔭))Ψ=𝐝(𝔭)^=D(𝔭^)=D(w¯𝔭Ψ),and similarlyR(w¯𝔭)Ψ=R(w¯𝔭Ψ).D(\overline{w}_{\mathfrak{p}})\Psi=(\overline{x}_{{\bf d}(\mathfrak{p})})\Psi=\widehat{{\bf d}(\mathfrak{p})}=D(\widehat{\mathfrak{p}})=D(\overline{w}_{\mathfrak{p}}\Psi),\qquad\text{and similarly}\qquad R(\overline{w}_{\mathfrak{p}})\Psi=R(\overline{w}_{\mathfrak{p}}\Psi). (8.7)
Lemma 8.8.

If 𝔭𝒫\mathfrak{p}\in\mathscr{P}, and if s𝐝(𝔭)s\leq{\bf d}(\mathfrak{p}) and t𝐫(𝔭)t\leq{\bf r}(\mathfrak{p}), then

xsw𝔭ws𝔭andw𝔭xtw𝔭t.x_{s}w_{\mathfrak{p}}\sim w_{{}_{s}{\downharpoonleft}\mathfrak{p}}\qquad\text{and}\qquad w_{\mathfrak{p}}x_{t}\sim w_{\mathfrak{p}{\downharpoonright}_{t}}.
Proof.

For the first (the second is dual), write 𝔭=(p1,,pk)\mathfrak{p}=(p_{1},\ldots,p_{k}), and fix sp1s\leq p_{1}. We then have

s𝔭=(s1,,sk),wheresi=sθp1θpifor each i,{}_{s}{\downharpoonleft}\mathfrak{p}=(s_{1},\ldots,s_{k}),\qquad\text{where}\qquad s_{i}=s\theta_{p_{1}}\cdots\theta_{p_{i}}\qquad\text{for each $i$,}

and we must show that

xsxp1xpkxs1xsk.x_{s}x_{p_{1}}\cdots x_{p_{k}}\sim x_{s_{1}}\cdots x_{s_{k}}.

If k=1k=1, then xsxp12xsxsθp1=xsxs1xs=xs1x_{s}x_{p_{1}}\sim_{2}x_{s}x_{s\theta_{p_{1}}}=x_{s}x_{s}\sim_{1}x_{s}=x_{s_{1}} (using sp1s\leq p_{1} in the second step). For k2k\geq 2, we apply induction to calculate

xsxp1xpk1xpkxs1xsk1xpk2xs1xsk1xsk1θpk,x_{s}x_{p_{1}}\cdots x_{p_{k-1}}x_{p_{k}}\sim x_{s_{1}}\cdots x_{s_{k-1}}x_{p_{k}}\sim_{2}x_{s_{1}}\cdots x_{s_{k-1}}x_{s_{k-1}\theta_{p_{k}}},

and we are done, since sk1θpk=sks_{k-1}\theta_{p_{k}}=s_{k}. ∎

For paths 𝔭=(p1,,pk)\mathfrak{p}=(p_{1},\ldots,p_{k}) and 𝔮=(q1,,ql)\mathfrak{q}=(q_{1},\ldots,q_{l}) with pkq1p_{k}\mathrel{\mathscr{F}}q_{1}, the concatenation of 𝔭\mathfrak{p} with 𝔮\mathfrak{q} is a path, which we denote by

𝔭𝔮=(p1,,pk,q1,,ql).\mathfrak{p}\oplus\mathfrak{q}=(p_{1},\ldots,p_{k},q_{1},\ldots,q_{l}).
Lemma 8.9.

For paths 𝔭=(p1,,pk)\mathfrak{p}=(p_{1},\ldots,p_{k}) and 𝔮=(q1,,ql)\mathfrak{q}=(q_{1},\ldots,q_{l}), we have

w𝔭w𝔮w𝔭st𝔮,wheres=q1δpk and t=pkθq1.w_{\mathfrak{p}}w_{\mathfrak{q}}\sim w_{\mathfrak{p}{\downharpoonright}_{s}\oplus{}_{t}{\downharpoonleft}\mathfrak{q}},\qquad\text{where}\qquad\text{$s=q_{1}\delta_{p_{k}}$ and $t=p_{k}\theta_{q_{1}}$.}
Proof.

Using (R1), and Lemmas 8.3 and 8.8, we have

w𝔭w𝔮w𝔭xpkxq1w𝔮w𝔭xsxtw𝔮w𝔭swt𝔮=w𝔭st𝔮.w_{\mathfrak{p}}w_{\mathfrak{q}}\sim w_{\mathfrak{p}}\cdot x_{p_{k}}x_{q_{1}}\cdot w_{\mathfrak{q}}\sim w_{\mathfrak{p}}\cdot x_{s}x_{t}\cdot w_{\mathfrak{q}}\sim w_{\mathfrak{p}{\downharpoonright}_{s}}\cdot w_{{}_{t}{\downharpoonleft}\mathfrak{q}}=w_{\mathfrak{p}{\downharpoonright}_{s}\oplus{}_{t}{\downharpoonleft}\mathfrak{q}}.\qed
Proposition 8.10.

FPF_{P} is a DRC-semigroup.

Proof.

We need to check the axioms (DRC1)(DRC4), and as usual we only have to prove one part of each. To do so, fix paths 𝔭=(p1,,pk)\mathfrak{p}=(p_{1},\ldots,p_{k}) and 𝔮=(q1,,ql)\mathfrak{q}=(q_{1},\ldots,q_{l}).

(DRC1). We have D(w¯𝔭)w¯𝔭=x¯p1x¯p1x¯pk=1x¯p1x¯pk=w¯𝔭D(\overline{w}_{\mathfrak{p}})\overline{w}_{\mathfrak{p}}=\overline{x}_{p_{1}}\cdot\overline{x}_{p_{1}}\cdots\overline{x}_{p_{k}}=_{1}\overline{x}_{p_{1}}\cdots\overline{x}_{p_{k}}=\overline{w}_{\mathfrak{p}}.

(DRC2). Let s=q1δpks=q_{1}\delta_{p_{k}} and t=pkθq1t=p_{k}\theta_{q_{1}}, and write 𝔭s=(s1,,sk)\mathfrak{p}{\downharpoonright}_{s}=(s_{1},\ldots,s_{k}) and t𝔮=(t1,,tl){}_{t}{\downharpoonleft}\mathfrak{q}=(t_{1},\ldots,t_{l}). Lemma 8.9 then gives

w¯𝔭w¯𝔮=w¯𝔭st𝔮andw¯𝔭D(w¯𝔮)=w¯𝔭x¯q1=w¯𝔭w¯(q1)=w¯𝔭st(q1).\overline{w}_{\mathfrak{p}}\overline{w}_{\mathfrak{q}}=\overline{w}_{\mathfrak{p}{\downharpoonright}_{s}\oplus{}_{t}{\downharpoonleft}\mathfrak{q}}\qquad\text{and}\qquad\overline{w}_{\mathfrak{p}}D(\overline{w}_{\mathfrak{q}})=\overline{w}_{\mathfrak{p}}\cdot\overline{x}_{q_{1}}=\overline{w}_{\mathfrak{p}}\cdot\overline{w}_{(q_{1})}=\overline{w}_{\mathfrak{p}{\downharpoonright}_{s}\oplus{}_{t}{\downharpoonleft}(q_{1})}.

Since 𝔭st𝔮=(s1,,sk,t1,,tl)\mathfrak{p}{\downharpoonright}_{s}\oplus{}_{t}{\downharpoonleft}\mathfrak{q}=(s_{1},\ldots,s_{k},t_{1},\ldots,t_{l}) and 𝔭st(q1)=(s1,,sk,t)\mathfrak{p}{\downharpoonright}_{s}\oplus{}_{t}{\downharpoonleft}(q_{1})=(s_{1},\ldots,s_{k},t), it then follows that

D(w¯𝔭w¯𝔮)=x¯s1=D(w¯𝔭D(w¯𝔮)).D(\overline{w}_{\mathfrak{p}}\overline{w}_{\mathfrak{q}})=\overline{x}_{s_{1}}=D(\overline{w}_{\mathfrak{p}}D(\overline{w}_{\mathfrak{q}})).

(DRC3). Keeping the notation of the previous part, and keeping in mind s1p1s_{1}\leq p_{1}, we have

D(w¯𝔭w¯𝔮)D(w¯𝔭)=x¯s1x¯p1=2x¯s1x¯s1θp1=x¯s1x¯s1=1x¯s1=D(w¯𝔭w¯𝔮).D(\overline{w}_{\mathfrak{p}}\overline{w}_{\mathfrak{q}})D(\overline{w}_{\mathfrak{p}})=\overline{x}_{s_{1}}\overline{x}_{p_{1}}=_{2}\overline{x}_{s_{1}}\overline{x}_{s_{1}\theta_{p_{1}}}=\overline{x}_{s_{1}}\overline{x}_{s_{1}}=_{1}\overline{x}_{s_{1}}=D(\overline{w}_{\mathfrak{p}}\overline{w}_{\mathfrak{q}}).

An analogous calculation gives D(w¯𝔭)D(w¯𝔭w¯𝔮)=D(w¯𝔭w¯𝔮)D(\overline{w}_{\mathfrak{p}})D(\overline{w}_{\mathfrak{p}}\overline{w}_{\mathfrak{q}})=D(\overline{w}_{\mathfrak{p}}\overline{w}_{\mathfrak{q}}).

(DRC4). We have R(D(w¯𝔭))=R(x¯p1)=x¯p1=D(w¯𝔭)R(D(\overline{w}_{\mathfrak{p}}))=R(\overline{x}_{p_{1}})=\overline{x}_{p_{1}}=D(\overline{w}_{\mathfrak{p}}). ∎

Theorem 8.11.

If PP is a projection algebra, then FPF_{P} and MPM_{P} are projection-generated DRC-semigroups, with

𝐏(FP)𝐏(MP)P.{\bf P}(F_{P})\cong{\bf P}(M_{P})\cong P.
Proof.

We saw in Proposition 8.10 that FPF_{P} is a DRC-semigroup. It follows that

𝐏(FP)=im(D)={x¯p:pP},{\bf P}(F_{P})=\operatorname{im}(D)=\{\overline{x}_{p}:p\in P\},

so we have a surjective map

ξ:P𝐏(FP):px¯p.\xi:P\to{\bf P}(F_{P}):p\mapsto\overline{x}_{p}.

Combining the definitions with Corollary 4.16 shows that ξ\xi is also injective, as for p,qPp,q\in P we have

pξ=qξx¯p=x¯qp^=x¯pΨ=x¯qΨ=q^(θp,δp)=(θq,δq)p=q.p\xi=q\xi\ \Rightarrow\ \overline{x}_{p}=\overline{x}_{q}\ \Rightarrow\ \widehat{p}=\overline{x}_{p}\Psi=\overline{x}_{q}\Psi=\widehat{q}\ \Rightarrow\ (\theta_{p},\delta_{p})=(\theta_{q},\delta_{q})\ \Rightarrow\ p=q. (8.12)

So ξ:P𝐏(FP)\xi:P\to{\bf P}(F_{P}) is a bijection, and we now wish to show that it is a projection algebra morphism, meaning that

(pθq)ξ=(pξ)θqξand(pδq)ξ=(pξ)δqξfor all p,qP,(p\theta_{q})\xi=(p\xi)\theta^{\prime}_{q\xi}\qquad\text{and}\qquad(p\delta_{q})\xi=(p\xi)\delta^{\prime}_{q\xi}\qquad\text{for all $p,q\in P$,}

where here we write θ\theta^{\prime} and δ\delta^{\prime} for the projection algebra operations in 𝐏(FP){\bf P}(F_{P}). For the first (the second is dual), we write p=qδpp^{\prime}=q\delta_{p} and q=pθqq^{\prime}=p\theta_{q}, and use Lemma 8.3 to calculate

(pξ)θqξ=(x¯p)θx¯q=R(x¯px¯q)=R(x¯px¯q)=R(w¯(p,q))=x¯q=qξ=(pθq)ξ.(p\xi)\theta^{\prime}_{q\xi}=(\overline{x}_{p})\theta^{\prime}_{\overline{x}_{q}}=R(\overline{x}_{p}\overline{x}_{q})=R(\overline{x}_{p^{\prime}}\overline{x}_{q^{\prime}})=R(\overline{w}_{(p^{\prime},q^{\prime})})=\overline{x}_{q^{\prime}}=q^{\prime}\xi=(p\theta_{q})\xi.

By definition, FP=XP:RPF_{P}=\langle X_{P}:R_{P}\rangle is generated by the projections x¯p\overline{x}_{p} (pPp\in P).

This completes the proof of the assertions regarding FPF_{P}. Those for MPM_{P} follow quickly, in light of the fact that the surjective homomorphism Ψ:FPMP\Psi:F_{P}\to M_{P} preserves the DD and RR operations on the two semigroups (cf. (8.7)), and maps 𝐏(FP){\bf P}(F_{P}) bijectively (as x¯pΨ=x¯qΨp=q\overline{x}_{p}\Psi=\overline{x}_{q}\Psi\ \Rightarrow\ p=q, as in (8.12)). ∎

Remark 8.13.

By identifying each projection pPp\in P with the \sim-class x¯pFP\overline{x}_{p}\in F_{P}, we can identify PP itself with the projection algebra 𝐏(FP){\bf P}(F_{P}). It will be convenient to do so in the next section.

8.2 Free projection-generated DRC-semigroups

Recall that we have a functor 𝐏:𝐃𝐑𝐂𝐏𝐀{\bf P}:{\bf DRC}\to{\bf PA}, which at the object level maps a DRC-semigroup SS to its projection algebra

𝐏(S)=im(D)=im(R)={pS:p2=p=D(p)=R(p)},{\bf P}(S)=\operatorname{im}(D)=\operatorname{im}(R)=\{p\in S:p^{2}=p=D(p)=R(p)\},

with operations given by

qθp=R(qp)andqδp=D(pq)for p,q𝐏(S).q\theta_{p}=R(qp)\qquad\text{and}\qquad q\delta_{p}=D(pq)\qquad\text{for $p,q\in{\bf P}(S)$.}

We can think of 𝐏{\bf P} as a forgetful functor, as when we construct 𝐏(S){\bf P}(S) from SS, we remember only a subset of the elements of SS, and retain only partial information about their products.

It turns out that 𝐏{\bf P} has a left adjoint, 𝐅:𝐏𝐀𝐃𝐑𝐂{\bf F}:{\bf PA}\to{\bf DRC}, which involves the semigroups constructed in Section 8.1. At the object level, we define

𝐅(P)=FPfor a projection algebra P.{\bf F}(P)=F_{P}\qquad\text{for a projection algebra $P$.}

To see how 𝐅{\bf F} acts on a projection algebra morphism ϕ:PP\phi:P\to P^{\prime}, we begin by defining a semigroup morphism

φ:XP+FP=XP+/RPbyxpφ=x¯pϕfor pP.\varphi:X_{P}^{+}\to F_{P^{\prime}}=X_{P^{\prime}}^{+}/R_{P^{\prime}}^{\sharp}\qquad\text{by}\qquad x_{p}\varphi=\overline{x}_{p\phi}\qquad\text{for $p\in P$.}

It is easy to see that RPker(φ)R_{P}\subseteq\ker(\varphi). For example, if p,qPp,q\in P then

(xpxq)φ=x¯pϕx¯qϕ\displaystyle(x_{p}x_{q})\varphi=\overline{x}_{p\phi}\overline{x}_{q\phi} =x¯pϕx¯(pϕ)θqϕ\displaystyle=\overline{x}_{p\phi}\overline{x}_{(p\phi)\theta^{\prime}_{q\phi}} as RPR_{P^{\prime}} contains the relation xpϕxqϕ=xpϕx(pϕ)θqϕx_{p\phi}x_{q\phi}=x_{p\phi}x_{(p\phi)\theta^{\prime}_{q\phi}}
=x¯pϕx¯(pθq)ϕ\displaystyle=\overline{x}_{p\phi}\overline{x}_{(p\theta_{q})\phi} as ϕ\phi is a projection algebra morphism
=(xpxpθq)φ,\displaystyle=(x_{p}x_{p\theta_{q}})\varphi,

which shows that φ\varphi preserves (R2). It follows that φ\varphi induces a well-defined semigroup homomorphism

Φ:FP=XP+/RPFPgiven byw¯Φ=wφfor wXP+.\Phi:F_{P}=X_{P}^{+}/R_{P}^{\sharp}\to F_{P^{\prime}}\qquad\text{given by}\qquad\overline{w}\Phi=w\varphi\qquad\text{for $w\in X_{P}^{+}$.} (8.14)

We then define 𝐅(ϕ)=Φ{\bf F}(\phi)=\Phi. It is essentially trivial to check that Φ\Phi preserves the DD and RR operations on FPF_{P} and FPF_{P^{\prime}}, in light of the definitions of these operations in (8.6), meaning that 𝐅(ϕ)=Φ{\bf F}(\phi)=\Phi is indeed a well-defined DRC-morphism 𝐅(P)𝐅(P){\bf F}(P)\to{\bf F}(P^{\prime}).

Proposition 8.15.

𝐅{\bf F} is a functor 𝐏𝐀𝐃𝐑𝐂{\bf PA}\to{\bf DRC}, and we have 𝐏𝐅=id𝐏𝐀{\bf P}{\bf F}=\operatorname{id}_{{\bf PA}}.

Proof.

The first claim follows quickly from the above discussion. For the second, we must show that:

  • 𝐏(𝐅(P))=P{\bf P}({\bf F}(P))=P for all projection algebras PP, and

  • 𝐏(𝐅(ϕ))=ϕ{\bf P}({\bf F}(\phi))=\phi for all projection algebra morphisms ϕ\phi.

This is routine, in light of the identification of PP with 𝐏(FP){\bf P}(F_{P}), as in Remark 8.13; cf. (8.14). ∎

Proposition 8.16.

For every projection algebra PP, every DRC-semigroup SS, and every projection algebra morphism ϕ:P𝐏(S)\phi:P\to{\bf P}(S), there exists a unique DRC-morphism Φ:𝐅(P)S\Phi:{\bf F}(P)\to S such that ϕ=𝐏(Φ)\phi={\bf P}(\Phi), i.e. such that the following diagram of maps commutes (with vertical arrows being inclusions):

P{P}𝐏(S){{\bf P}(S)}𝐅(P){{\bf F}(P)}S.{S.}ϕ\scriptstyle{\ \ \phi}Φ\scriptstyle{\Phi\ \ }
Proof.

Fix a projection algebra morphism ϕ:P𝐏(S)\phi:P\to{\bf P}(S). For convenience we write

p¯=pϕ𝐏(S)for pP.\underline{p}=p\phi\in{\bf P}(S)\qquad\text{for $p\in P$.}

We write DD and RR for the unary operations on FP=𝐅(P)F_{P}={\bf F}(P), and θ\theta and δ\delta for the projection algebra operations in P=𝐏(FP)P={\bf P}(F_{P}), and use dashes to distinguish the corresponding operations in SS and 𝐏(S){\bf P}(S).

We begin by defining

φ:XP+Sbyxpφ=p¯for pP.\varphi:X_{P}^{+}\to S\qquad\text{by}\qquad x_{p}\varphi=\underline{p}\qquad\text{for $p\in P$.}

To see that RPker(φ)R_{P}\subseteq\ker(\varphi), note for example that if p,qPp,q\in P then

(xpxq)φ=p¯q¯\displaystyle(x_{p}x_{q})\varphi=\underline{p}\;\;\!\!\underline{q} =p¯q¯R(p¯q¯)\displaystyle=\underline{p}\;\;\!\!\underline{q}\cdot R^{\prime}(\underline{p}\;\;\!\!\underline{q}) by (DRC1)
=p¯R(p¯q¯)\displaystyle=\underline{p}\cdot R^{\prime}(\underline{p}\;\;\!\!\underline{q}) by Lemma 3.4
=p¯p¯θq¯\displaystyle=\underline{p}\cdot\underline{p}\theta^{\prime}_{\underline{q}} by definition
=p¯pθq¯\displaystyle=\underline{p}\cdot\underline{p\theta_{q}} as ϕ\phi is a projection algebra morphism
=(xpxpθq)φ,\displaystyle=(x_{p}x_{p\theta_{q}})\varphi,

showing that φ\varphi preserves (R2). As usual, it follows that φ\varphi induces a well-defined semigroup morphism

Φ:FPSgiven byw¯Φ=wφfor wXP+.\Phi:F_{P}\to S\qquad\text{given by}\qquad\overline{w}\Phi=w\varphi\qquad\text{for $w\in X_{P}^{+}$.}

To see that Φ\Phi preserves DD (RR is dual), consider a typical element w¯𝔭\overline{w}_{\mathfrak{p}} of FPF_{P} (cf. Lemma 8.4), where 𝔭=(p1,,pk)𝒫\mathfrak{p}=(p_{1},\ldots,p_{k})\in\mathscr{P}. We then have

D(w¯𝔭)Φ=x¯p1Φ=p¯1andD(w¯𝔭Φ)=D(p¯1p¯k),D(\overline{w}_{\mathfrak{p}})\Phi=\overline{x}_{p_{1}}\Phi=\underline{p}_{1}\qquad\text{and}\qquad D^{\prime}(\overline{w}_{\mathfrak{p}}\Phi)=D^{\prime}(\underline{p}_{1}\cdots\underline{p}_{k}),

so we must show that D(p¯1p¯k)=p¯1D^{\prime}(\underline{p}_{1}\cdots\underline{p}_{k})=\underline{p}_{1}. But this follows from Lemma 4.17(ii), together with the fact that p1pkp_{1}\mathrel{\mathscr{F}}\cdots\mathrel{\mathscr{F}}p_{k} implies p¯1p¯k\underline{p}_{1}\mathrel{\mathscr{F}}\cdots\mathrel{\mathscr{F}}\underline{p}_{k} (as ϕ\phi is a projection algebra morphism).

So Φ\Phi is indeed a DRC-morphism FPSF_{P}\to S. Since Φ\Phi maps px¯pp\equiv\overline{x}_{p} to p¯=pϕ\underline{p}=p\phi, it is clear that 𝐏(Φ)=Φ|P=ϕ{{\bf P}(\Phi)=\Phi|_{P}=\phi}. This establishes the existence of Φ\Phi. Uniqueness follows from the fact that FP=P{F_{P}=\langle P\rangle}. ∎

Propositions 8.15 and 8.16 verify the assumptions of [15, Lemma 5.6], which immediately gives the following result. For the definitions of the (standard) categorical terms in the statement see for example [2, 38], and also [15, Section 5].

Theorem 8.17.

The functor 𝐏𝐀𝐃𝐑𝐂:PFP{\bf PA}\to{\bf DRC}:P\mapsto F_{P} is a left adjoint to the forgetful functor 𝐃𝐑𝐂𝐏𝐀:S𝐏(S){\bf DRC}\to{\bf PA}:S\mapsto{\bf P}(S), and 𝐏𝐀{\bf PA} is coreflective in 𝐃𝐑𝐂{\bf DRC}. ∎

Remark 8.18.

One might wonder if the assignment PMPP\mapsto M_{P} is also the object part of a functor 𝐏𝐀𝐃𝐑𝐂{{\bf PA}\to{\bf DRC}}, but this turns out not to be the case, for exactly the same reason discussed in [13, Remark 4.29].

Remark 8.19.

The study of free idempotent-generated semigroups over biordered sets is a topic of considerable interest in its own right [23, 24, 11, 9, 6, 60, 10, 3], and the corresponding theory for regular *-semigroups has been recently initiated in [15]. We hope that the free projection-generated DRC-semigroups introduced here will likewise inspire further studies.

8.3 Fundamental projection-generated DRC-semigroups

Following [30], a (DRC-)congruence on a DRC-semigroup SS is an equivalence relation σ\sigma on SS respecting all of the operations, in the sense that

[a𝜎b and a𝜎b]aa𝜎bbanda𝜎b[D(a)𝜎D(b) and R(a)𝜎R(b)].[a\mathrel{\sigma}b\text{ and }a^{\prime}\mathrel{\sigma}b^{\prime}]\ \Rightarrow\ aa^{\prime}\mathrel{\sigma}bb^{\prime}\qquad\text{and}\qquad a\mathrel{\sigma}b\ \Rightarrow\ [D(a)\mathrel{\sigma}D(b)\text{ and }R(a)\mathrel{\sigma}R(b)].

The congruence σ\sigma is projection-separating if p𝜎qp=qp\mathrel{\sigma}q\ \Rightarrow\ p=q for all projections p,q𝐏(S)p,q\in{\bf P}(S). Among other things, [30, Proposition 2.3] shows that SS has a maximum projection-separating congruence, denoted μS\mu_{S}; that is, μS\mu_{S} is projection-separating, and any other projection-separating congruence is contained in μS\mu_{S}. We say SS is (DRC-)fundamental if μS\mu_{S} is the trivial congruence ιS={(a,a):aS}\iota_{S}=\{(a,a):a\in S\}. In general, S/μSS/\mu_{S} is fundamental, the maximum fundamental image of SS.

Remark 8.20.

Here we follow Jones’ convention in dropping the ‘DRC-’ prefix, and speak simply of ‘fundamental DRC-semigroups’. But the reader should be aware that a ‘fundamental’ (ordinary) semigroup is usually defined as a semigroup with no non-trivial congruence contained in Green’s \mathrel{\mathscr{H}}-relation. For regular semigroups, this is equivalent to having no non-trivial idempotent-separating congruence [32].

Proposition 8.21.

If SS is a DRC-semigroup, then

μS\displaystyle\mu_{S} ={(a,b)S×S:Θa=Θb and Δa=Δb}\displaystyle=\{(a,b)\in S\times S:\Theta_{a}=\Theta_{b}\text{ and }\Delta_{a}=\Delta_{b}\}
={(a,b)S×S:ϑa=ϑb and a=b}\displaystyle=\{(a,b)\in S\times S:\vartheta_{a}=\vartheta_{b}\text{ and }\partial_{a}=\partial_{b}\}
Proof.

Writing P=𝐏(S)P={\bf P}(S), Proposition 2.3 of [30] says that for a,bSa,b\in S we have (a,b)μS(a,b)\in\mu_{S} if and only if

D(ap)=D(bp)andR(pa)=R(pb)for all pP1,D(ap)=D(bp)\quad\text{and}\quad R(pa)=R(pb)\qquad\text{for all $p\in P^{1}$,}

where P1P^{1} is PP with an adjoined identity element. Keeping (5.5) in mind, this is equivalent to all of the following holding:

Θa=Θb,Δa=Δb,D(a)=D(b)andR(a)=R(b).\Theta_{a}=\Theta_{b},\qquad\Delta_{a}=\Delta_{b},\qquad D(a)=D(b)\qquad\text{and}\qquad R(a)=R(b).

Remembering also that 𝐝=D{\bf d}=D and 𝐫=R{\bf r}=R, we can therefore prove the first equality in the current proposition by showing that

[Θa=Θb and Δa=Δb][𝐝(a)=𝐝(b) and 𝐫(a)=𝐫(b)].[\Theta_{a}=\Theta_{b}\text{ and }\Delta_{a}=\Delta_{b}]\qquad\Rightarrow\qquad[{\bf d}(a)={\bf d}(b)\text{ and }{\bf r}(a)={\bf r}(b)]. (8.22)

To do so, suppose Θa=Θb\Theta_{a}=\Theta_{b} and Δa=Δb\Delta_{a}=\Delta_{b}. Writing q=𝐫(a)q={\bf r}(a), we have

𝐝(a)=𝐝(aq)=qa=qδqa=qΔa=qΔb𝐝(b),and similarly𝐝(b)𝐝(a).{\bf d}(a)={\bf d}(a{\downharpoonright}_{q})=q\partial_{a}=q\delta_{q}\partial_{a}=q\Delta_{a}=q\Delta_{b}\leq{\bf d}(b),\qquad\text{and similarly}\qquad{\bf d}(b)\leq{\bf d}(a).

The proof that 𝐫(a)=𝐫(b){\bf r}(a)={\bf r}(b) is dual.

For the second equality in the proposition, we will show that

[Θa=Θb and Δa=Δb][ϑa=ϑb and a=b]for a,bS.[\Theta_{a}=\Theta_{b}\text{ and }\Delta_{a}=\Delta_{b}]\qquad\Leftrightarrow\qquad[\vartheta_{a}=\vartheta_{b}\text{ and }\partial_{a}=\partial_{b}]\qquad\text{for $a,b\in S$.} (8.23)

In light of the identities Θa=θ𝐝(a)ϑa\Theta_{a}=\theta_{{\bf d}(a)}\vartheta_{a} and ϑa=Θa|𝐝(a)\vartheta_{a}=\Theta_{a}|_{{\bf d}(a)^{\downarrow}} (the second of which is Lemma 5.3(iv)), and the analogous identities for Δa\Delta_{a} and a\partial_{a}, we can prove (8.23) by showing that

[Θa=Θb and Δa=Δb]\displaystyle[\Theta_{a}=\Theta_{b}\text{ and }\Delta_{a}=\Delta_{b}] [𝐝(a)=𝐝(b) and 𝐫(a)=𝐫(b)],\displaystyle\qquad\Rightarrow\qquad[{\bf d}(a)={\bf d}(b)\text{ and }{\bf r}(a)={\bf r}(b)],
and[ϑa=ϑb and a=b]\displaystyle\text{and}\qquad[\vartheta_{a}=\vartheta_{b}\text{ and }\partial_{a}=\partial_{b}]\hskip 6.544pt [𝐝(a)=𝐝(b) and 𝐫(a)=𝐫(b)].\displaystyle\qquad\Rightarrow\qquad[{\bf d}(a)={\bf d}(b)\text{ and }{\bf r}(a)={\bf r}(b)].

The first implication is (8.22), which was proved above. The second follows from the fact that dom(ϑa)=𝐝(a)\operatorname{dom}(\vartheta_{a})={\bf d}(a)^{\downarrow} and dom(a)=𝐫(a)\operatorname{dom}(\partial_{a})={\bf r}(a)^{\downarrow}. ∎

Fundamental DRC-semigroups were classified in [30] by means of a transformation representation. We will not attempt to reprove this here, but instead we prove a result concerning projection-generated fundamental DRC-semigroups (see Theorem 8.25 below), which was not given in [30]. For its proof, we will make use of the following result, which shows that the maps Θa\Theta_{a} and Δa\Delta_{a} have very natural forms when aa is a product of projections.

Lemma 8.24.

If SS is a DRC-semigroup, and if p1,,pk𝐏(S)p_{1},\ldots,p_{k}\in{\bf P}(S), then

Θp1pk=θp1θpkandΔp1pk=δpkδp1.\Theta_{p_{1}\cdots p_{k}}=\theta_{p_{1}}\cdots\theta_{p_{k}}\qquad\text{and}\qquad\Delta_{p_{1}\cdots p_{k}}=\delta_{p_{k}}\cdots\delta_{p_{1}}.
Proof.

To prove the first statement (the second is dual), let t𝐏(S)t\in{\bf P}(S). Then using (5.5) and Lemma 4.17(i), we have tΘp1pk=R(tp1pk)=tθp1θpkt\Theta_{p_{1}\cdots p_{k}}=R(tp_{1}\cdots p_{k})=t\theta_{p_{1}}\cdots\theta_{p_{k}}. ∎

We are now ready to prove our final result of this section:

Theorem 8.25.

For any projection algebra PP there is a unique (up to isomorphism) projection-generated fundamental DRC-semigroup with projection algebra PP, namely MPFP/μFPM_{P}\cong F_{P}/\mu_{F_{P}}.

Proof.

By Theorem 8.11, we can identify the projection algebras 𝐏(FP){\bf P}(F_{P}) and 𝐏(MP){\bf P}(M_{P}) with PP itself, and in this way the semigroups FPF_{P} and MPM_{P} are both generated by PP. To avoid confusion, we will denote the products in these semigroups by \bullet and \star, respectively, so that typical elements have the form

p1pkxp1xpk¯FPandp1pkp^1p^k=(θp1θpk,δpkδp1)MP,p_{1}\bullet\cdots\bullet p_{k}\equiv\overline{x_{p_{1}}\cdots x_{p_{k}}}\in F_{P}\quad\text{and}\quad p_{1}\star\cdots\star p_{k}\equiv\widehat{p}_{1}\cdots\widehat{p}_{k}=(\theta_{p_{1}}\cdots\theta_{p_{k}},\delta_{p_{k}}\cdots\delta_{p_{1}})\in M_{P},

for some p1,,pkPp_{1},\ldots,p_{k}\in P. In what follows, we use Proposition 8.21 freely.

We have already noted that MPM_{P} is a projection-generated DRC-semigroup with projection algebra PP. To see that it is fundamental, we need to show that μMP\mu_{M_{P}} is trivial. To do so, suppose (a,b)μMP(a,b)\in\mu_{M_{P}}; we must show that a=ba=b. As above, we have

a=p1pk=(θp1θpk,δpkδp1)andb=q1ql=(θq1θql,δqlδq1),a=p_{1}\star\cdots\star p_{k}=(\theta_{p_{1}}\cdots\theta_{p_{k}},\delta_{p_{k}}\cdots\delta_{p_{1}})\qquad\text{and}\qquad b=q_{1}\star\cdots\star q_{l}=(\theta_{q_{1}}\cdots\theta_{q_{l}},\delta_{q_{l}}\cdots\delta_{q_{1}}),

for some p1,,pk,q1,,qlPp_{1},\ldots,p_{k},q_{1},\ldots,q_{l}\in P. Since (a,b)μMP(a,b)\in\mu_{M_{P}}, we have Θa=Θb\Theta_{a}=\Theta_{b} and Δa=Δb\Delta_{a}=\Delta_{b}. It then follows from Lemma 8.24 that

θp1θpk=Θa=Θb=θq1θql,and similarlyδpkδp1=δqlδq1.\theta_{p_{1}}\cdots\theta_{p_{k}}=\Theta_{a}=\Theta_{b}=\theta_{q_{1}}\cdots\theta_{q_{l}},\qquad\text{and similarly}\qquad\delta_{p_{k}}\cdots\delta_{p_{1}}=\delta_{q_{l}}\cdots\delta_{q_{1}}.

But then a=(θp1θpk,δpkδp1)=(θq1θql,δqlδq1)=ba=(\theta_{p_{1}}\cdots\theta_{p_{k}},\delta_{p_{k}}\cdots\delta_{p_{1}})=(\theta_{q_{1}}\cdots\theta_{q_{l}},\delta_{q_{l}}\cdots\delta_{q_{1}})=b.

Now let SS be an arbitrary projection-generated fundamental DRC-semigroup with projection algebra 𝐏(S)=P{\bf P}(S)=P; we denote the product in SS simply by juxtaposition. The proof will be complete if we can show that SFP/μFPS\cong F_{P}/\mu_{F_{P}}. Applying Proposition 8.16 to the identity morphism ϕ:PP=𝐏(S)\phi:P\to P={\bf P}(S), we see that there is a DRC-morphism

Φ:FPS:pp.\Phi:F_{P}\to S:p\mapsto p.

Since SS is projection-generated, Φ\Phi is surjective, and so SFP/ker(Φ)S\cong F_{P}/\ker(\Phi) by the fundamental homomorphism theorem. It therefore remains to show that ker(Φ)=μFP\ker(\Phi)=\mu_{F_{P}}. Since Φ\Phi maps P=𝐏(FP)P={\bf P}(F_{P}) identically, ker(Φ)\ker(\Phi) is projection-separating, and so ker(Φ)μFP\ker(\Phi)\subseteq\mu_{F_{P}}. Conversely, let (a,b)μFP(a,b)\in\mu_{F_{P}}, so that Θa=Θb\Theta_{a}=\Theta_{b} and Δa=Δb\Delta_{a}=\Delta_{b}. Since FPF_{P} is projection-generated, we have

a=p1pkandb=q1qlfor some p1,,pk,q1,,qlP.a=p_{1}\bullet\cdots\bullet p_{k}\qquad\text{and}\qquad b=q_{1}\bullet\cdots\bullet q_{l}\qquad\text{for some $p_{1},\ldots,p_{k},q_{1},\ldots,q_{l}\in P$.}

But then applying Lemma 8.24 in both FPF_{P} and SS, we have

Θa=Θp1pk=θp1θpk=Θp1pk=ΘaΦ,and similarlyΘb=ΘbΦ.\Theta_{a}=\Theta_{p_{1}\bullet\cdots\bullet p_{k}}=\theta_{p_{1}}\cdots\theta_{p_{k}}=\Theta_{p_{1}\cdots p_{k}}=\Theta_{a\Phi},\qquad\text{and similarly}\qquad\Theta_{b}=\Theta_{b\Phi}.

It follows that ΘaΦ=Θa=Θb=ΘbΦ\Theta_{a\Phi}=\Theta_{a}=\Theta_{b}=\Theta_{b\Phi}, and simlarly ΔaΦ=ΔbΦ\Delta_{a\Phi}=\Delta_{b\Phi}, so that (aΦ,bΦ)μS(a\Phi,b\Phi)\in\mu_{S}. Since SS is fundamental, it follows that aΦ=bΦa\Phi=b\Phi, i.e. that (a,b)ker(Φ)(a,b)\in\ker(\Phi). ∎

9 Regular *- and *-regular semigroups

As mentioned in Section 1, the current paper extends to DRC-semigroups the groupoid-based approach to regular *-semigroups from [13]. In the case of regular *-semigroups (whose definition will be recalled below), many of the current categorical constructions simplify, sometimes drammatically. In this section we outline many of these simplifications, and indicate why the more elaborate approach of the current paper is necessary to encompass the full generality of DRC-semigroups. In doing so, we will contrast the situation with Drazin’s broader class of *-regular semigroups [12], which induce natural DRC-structures. These are already complex enough to require the more complicated setup, and a ready source of (counter)examples will be provided by the real matrix monoids Mn()M_{n}(\mathbb{R}), which are known to be *-regular [46]. (Due to quirks of terminology, *-regular and regular *-semigroups are distinct classes, though the former contains the latter.) We will also observe in Section 10 that many of the simplifications hold in the DRC-restriction semigroups considered by Die and Wang in [8].

Before we begin, we recall some basic semigroup theoretical background; for more information see for example [27, 4]. Let SS be a semigroup, and S1S^{1} its monoid completion. So S1=SS^{1}=S if SS is a monoid, or else S1=S{1}S^{1}=S\cup\{1\}, where 11 is a symbol not belonging to SS, acting as an adjoined identity. Green’s \mathrel{\mathscr{L}}\mathrel{\mathscr{R}} and 𝒥\mathrel{\mathscr{J}} equivalences are defined, for a,bSa,b\in S, by

abS1a=S1b,abaS1=bS1anda𝒥bS1aS1=S1bS1.a\mathrel{\mathscr{L}}b\ \Leftrightarrow\ S^{1}a=S^{1}b,\qquad a\mathrel{\mathscr{R}}b\ \Leftrightarrow\ aS^{1}=bS^{1}\qquad\text{and}\qquad a\mathrel{\mathscr{J}}b\ \Leftrightarrow\ S^{1}aS^{1}=S^{1}bS^{1}.

From these are defined the further equivalences ={\mathrel{\mathscr{H}}}={\mathrel{\mathscr{L}}}\cap{\mathrel{\mathscr{R}}} and 𝒟={\mathrel{\mathscr{D}}}={\mathrel{\mathscr{L}}}\vee{\mathrel{\mathscr{R}}}, where the latter is the join in the lattice of all equivalences on SS. We have 𝒟=={\mathrel{\mathscr{D}}}={\mathrel{\mathscr{L}}}\circ{\mathrel{\mathscr{R}}}={\mathrel{\mathscr{R}}}\circ{\mathrel{\mathscr{L}}}, and if SS is finite then 𝒟=𝒥{\mathrel{\mathscr{D}}}={\mathrel{\mathscr{J}}}.

An element aa of a semigroup SS is regular if a=axaa=axa for some xSx\in S. This is equivalent to having a=ayaa=aya and y=yayy=yay for some ySy\in S (if a=axaa=axa, then take y=xaxy=xax), and such an element yy is called a (semigroup) inverse of aa. Note then that aaya\mathrel{\mathscr{R}}ay and ayaa\mathrel{\mathscr{L}}ya, with ayay and yaya idempotents. In fact, an element of a semigroup is regular if and only if it is 𝒟\mathrel{\mathscr{D}}-related to an idempotent.

9.1 Matrix monoids

Fix an integer n1n\geq 1, and let Mn()M_{n}(\mathbb{R}) be the monoid of n×nn\times n real matrices, under ordinary matrix multiplication. (We consider real matrices merely for convenience, though what we say can be adapted to the complex field.) Denoting the row and column spaces of aMn()a\in M_{n}(\mathbb{R}) by Row(a)\operatorname{Row}(a) and Col(a)\operatorname{Col}(a), and writing rank(a)=dim(Row(a))=dim(Col(a))\operatorname{rank}(a)=\dim(\operatorname{Row}(a))=\dim(\operatorname{Col}(a)) for the rank of aa, Green’s relations on Mn()M_{n}(\mathbb{R}) are given by

ab\displaystyle a\mathrel{\mathscr{L}}b Row(a)=Row(b),\displaystyle\ \Leftrightarrow\ \operatorname{Row}(a)=\operatorname{Row}(b),
ab\displaystyle a\mathrel{\mathscr{R}}b Col(a)=Col(b)anda𝒥ba𝒟brank(a)=rank(b).\displaystyle\ \Leftrightarrow\ \operatorname{Col}(a)=\operatorname{Col}(b)\qquad\text{and}\qquad a\mathrel{\mathscr{J}}b\ \Leftrightarrow\ a\mathrel{\mathscr{D}}b\ \Leftrightarrow\ \operatorname{rank}(a)=\operatorname{rank}(b). (9.1)

(See for example [44, Lemma 2.1].) Since idempotents of arbitrary rank 0rn0\leq r\leq n clearly exist, Mn()M_{n}(\mathbb{R}) is regular.

In fact, any matrix aMn()a\in M_{n}(\mathbb{R}) can be assigned a special inverse [46]. Using T{}^{\operatorname{T}} to denote transpose, there exists a unique matrix xMn()x\in M_{n}(\mathbb{R}) satisfying

a=axa,x=xax,(ax)T=axand(xa)T=xa.a=axa,\qquad x=xax,\qquad(ax)^{\operatorname{T}}=ax\qquad\text{and}\qquad(xa)^{\operatorname{T}}=xa. (9.2)

This xx is known as the Moore–Penrose inverse, and is denoted x=ax=a^{\dagger}. There is no simple formula for aa^{\dagger}; its existence is established in [46] by using the linear dependence of the sequence (aTa)k(a^{\operatorname{T}}a)^{k}, k=1,2,3,k=1,2,3,\ldots. Nevertheless, many computational packages exist for working with the Moore–Penrose inverse, e.g. Matlab [54]. For example, with a=[111100000]M3()a=\left[\begin{smallmatrix}1&1&1\\ 1&0&0\\ 0&0&0\end{smallmatrix}\right]\in M_{3}(\mathbb{R}), Matlab tells us that a=[0101/21/201/21/20]a^{\dagger}=\left[\begin{smallmatrix}0&1&0\\ \nicefrac{{1}}{{2}}&-\nicefrac{{1}}{{2}}&0\\ \nicefrac{{1}}{{2}}&-\nicefrac{{1}}{{2}}&0\end{smallmatrix}\right]. One can readily check the identities (9.2), and we note that aa=[100010000]aa^{\dagger}=\left[\begin{smallmatrix}1&0&0\\ 0&1&0\\ 0&0&0\end{smallmatrix}\right] and aa=[10001/21/201/21/2]a^{\dagger}a=\left[\begin{smallmatrix}1&0&0\\ 0&\nicefrac{{1}}{{2}}&\nicefrac{{1}}{{2}}\\ 0&\nicefrac{{1}}{{2}}&\nicefrac{{1}}{{2}}\end{smallmatrix}\right] are symmetric idempotents, with aaaaaaa^{\dagger}\mathrel{\mathscr{R}}a\mathrel{\mathscr{L}}a^{\dagger}a; cf. (9.1).

9.2 Regular *- and *-regular semigroups as DRC-semigroups

The defining properties of the Moore–Penrose inverse in Mn()M_{n}(\mathbb{R}) led to the introduction in [12] of the class of *-regular semigroups, to which we now turn.

First, a *-semigroup is an algebra (S,,)(S,\cdot,{}^{*}), where (S,)(S,\cdot) is a semigroup, and where is an involution, i.e. a unary operation SS:aaS\to S:a\mapsto a^{*} satisfying the laws

a=aand(ab)=bafor all a,bS.a^{**}=a\qquad\text{and}\qquad(ab)^{*}=b^{*}a^{*}\qquad\text{for all $a,b\in S$.}

For example, Mn()M_{n}(\mathbb{R}) becomes a *-semigroup with involution given by transpose: a=aTa^{*}=a^{\operatorname{T}}. The category of *-semigroups with *-morphisms (i.e. the semigroup morphisms preserving ) contains two important (full) subcategories, whose definitions we will shortly recall:

  • 𝐑𝐒𝐒{\bf RSS}, the category of regular *-semigroups [43], and

  • 𝐒𝐑𝐒{\bf SRS}, the category of *-regular semigroups [12].

Beginning with the latter, a regular *-semigroup is a *-semigroup (S,,)(S,\cdot,{}^{*}) additionally satisfying

a=aaafor all aS.a=aa^{*}a\qquad\text{for all $a\in S$.}

It follows from the axioms that also aaa=aa^{*}aa^{*}=a^{*} in a regular *-semigroup, meaning that aa^{*} is a (semigroup) inverse of aa.

A *-regular semigroup is a *-semigroup (S,,)(S,\cdot,{}^{*}) for which every element aSa\in S has an inverse that is \mathrel{\mathscr{H}}-related to aa^{*}. It follows from basic semigroup-theoretic facts that this inverse is unique; it is denoted by aa^{\dagger}, and called the Moore–Penrose inverse of aa [46]. The terminology stems from the fact (see [42, Corollary 1.2]) that aa^{\dagger} can also be characterised as the unique element xx satisfying

a=axa,x=xax,(ax)=axand(xa)=xa.a=axa,\qquad x=xax,\qquad(ax)^{*}=ax\qquad\text{and}\qquad(xa)^{*}=xa.

In this way, a *-regular semigroup can be equivalently defined as an algebra (S,,,)(S,\cdot,{}^{*},{}^{\dagger}), where (S,)(S,\cdot) is a semigroup, and and are unary operations satisfying:

a=a,(ab)=ba,aaa=a,aaa=a,(aa)=aaand(aa)=aa.a^{**}=a,\quad(ab)^{*}=b^{*}a^{*},\quad aa^{\dagger}a=a,\quad a^{\dagger}aa^{\dagger}=a^{\dagger},\quad(aa^{\dagger})^{*}=aa^{\dagger}\quad\text{and}\quad(a^{\dagger}a)^{*}=a^{\dagger}a. (9.3)

Various other laws can be deduced as consequences (see for example [42, Theorem 1.3]), including:

a=aanda=a.a^{\dagger\dagger}=a\qquad\text{and}\qquad a^{*\dagger}=a^{\dagger*}. (9.4)

Regular *-semigroups are the *-regular semigroups in which a=a{a^{\dagger}=a^{*}} for all aa. The matrix monoid Mn()M_{n}(\mathbb{R}) is *-regular (cf. (9.2)), but it is not a regular *-semigroup, as aTa^{\operatorname{T}} need not be an inverse of aa, even when n=1n=1.

It is well known (and easy to see) that any *-morphism ϕ:ST\phi:S\to T between *-regular semigroups preserves Moore–Penrose inverses, meaning that (a)ϕ=(aϕ)(a^{\dagger})\phi=(a\phi)^{\dagger} for all aSa\in S. Less obvious is the following:

Theorem 9.5 (Jones [30, Proposition 4.2]).

Any *-regular semigroup S(S,,)S\equiv(S,\cdot,{}^{*}) gives rise to a DRC-semigroup 𝐈(S)=(S,,D,R){\bf I}(S)=(S,\cdot,D,R), under the unary operations

D(a)=aaandR(a)=aafor aS.D(a)=aa^{\dagger}\qquad\text{and}\qquad R(a)=a^{\dagger}a\qquad\text{for $a\in S$.}
Remark 9.5.

The proof of Theorem 9.5 is completely routine in the special case that SS is a regular *-semigroup. We also note that the elements D(a)=aaD(a)=aa^{\dagger} and R(a)=aaR(a)=a^{\dagger}a appearing in the theorem are the unique projections of SS that are \mathrel{\mathscr{R}}- and \mathrel{\mathscr{L}}-related to aa, respectively. This was observed, for example, in [5].

We can think of the assignment S𝐈(S)S\mapsto{\bf I}(S) as an object map from 𝐒𝐑𝐒{\bf SRS} to 𝐃𝐑𝐂{\bf DRC}, and this can be easily extended to a functor 𝐈:𝐒𝐑𝐒𝐃𝐑𝐂{\bf I}:{\bf SRS}\to{\bf DRC}. Given a *-morphism ϕ:ST\phi:S\to T in 𝐒𝐑𝐒{\bf SRS}, we simply define 𝐈(ϕ)=ϕ{\bf I}(\phi)=\phi. Since ϕ\phi preserves , it also preserves DD and RR.

The functor 𝐈:𝐒𝐑𝐒𝐃𝐑𝐂{\bf I}:{\bf SRS}\to{\bf DRC} is not full (surjective on morphisms), however. Indeed, consider an abelian group GG with identity 11, in which at least one element is not its own inverse (e.g. the cyclic group of order 33). Then S1=(G,,)1S_{1}=(G,\cdot,{}^{-1}) and S2=(G,,)S_{2}=(G,\cdot,{}^{\star}) are both *-regular semigroups, where a=aa^{\star}=a for all aGa\in G. Moreover, we have 𝐈(S1)=𝐈(S2){\bf I}(S_{1})={\bf I}(S_{2}), since D(a)=R(a)=1D(a)=R(a)=1 in both semigroups for all aGa\in G. It follows that the identity map ι:GG\iota:G\to G is a DRC-morphism 𝐈(S1)𝐈(S2){\bf I}(S_{1})\to{\bf I}(S_{2}), but it is not a *-morphism, as (a1)ι(aι)(a^{-1})\iota\not=(a\iota)^{\star} for any non-self-inverse aGa\in G.

On the other hand, the restriction of 𝐈{\bf I} to the subcategory 𝐑𝐒𝐒{\bf RSS} of regular *-semigroups is full, as follows from the next result (recall that coincides with in a regular *-semigroup).

Proposition 9.6.

If ϕ:𝐈(S)𝐈(T)\phi:{\bf I}(S)\to{\bf I}(T) is a DRC-morphism, for *-regular semigroups SS and TT, then (a)ϕ=(aϕ)(a^{\dagger})\phi=(a\phi)^{\dagger} for all aSa\in S.

Proof.

Since ϕ\phi is a DRC-morphism, we have

aϕaϕ=(aa)ϕ=D(a)ϕ=D(aϕ)=aϕ(aϕ),and similarlyaϕaϕ=(aϕ)aϕ.a\phi\cdot a^{\dagger}\phi=(aa^{\dagger})\phi=D(a)\phi=D(a\phi)=a\phi\cdot(a\phi)^{\dagger},\qquad\text{and similarly}\qquad a^{\dagger}\phi\cdot a\phi=(a\phi)^{\dagger}\cdot a\phi.

Combining these gives

aϕ=(aaa)ϕ=aϕaϕaϕ=aϕaϕ(aϕ)=(aϕ)aϕ(aϕ)=(aϕ).a^{\dagger}\phi=(a^{\dagger}\cdot a\cdot a^{\dagger})\phi=a^{\dagger}\phi\cdot a\phi\cdot a^{\dagger}\phi=a^{\dagger}\phi\cdot a\phi\cdot(a\phi)^{\dagger}=(a\phi)^{\dagger}\cdot a\phi\cdot(a\phi)^{\dagger}=(a\phi)^{\dagger}.\qed

It follows that 𝐑𝐒𝐒{\bf RSS} is (isomorphic to) a full subcategory of 𝐃𝐑𝐂{\bf DRC}, and is hence isomorphic to its image under the isomorphism 𝐂:𝐃𝐑𝐂𝐂𝐏𝐂{\bf C}:{\bf DRC}\to{\bf CPC} from Theorem 5.20. One could then go on to show that this image is isomorphic to the category 𝐂𝐏𝐆{\bf CPG} of chained projection groupoids, as defined in [13], thereby proving the main result of that paper. We will not give the full details here, as a similar deduction was considered in [13] itself, where the isomorphism 𝐑𝐒𝐒𝐂𝐏𝐆{\bf RSS}\cong{\bf CPG} was used to reprove the Ehresmann–Nambooripad–Schein Theorem on inverse semigroups and inductive groupoids.

The situation for the more general *-regular semigroups, however, is more complicated. Since the functor 𝐈:𝐒𝐑𝐒𝐃𝐑𝐂{\bf I}:{\bf SRS}\to{\bf DRC} is not an isomorphism onto a full subcategory, we are left with what seems to be a very interesting and important open problem:

Problem 9.7.

Obtain an ‘ESN-type’ theorem for the category of *-regular semigroups.

9.3 Projection algebras of regular *- and *-regular semigroups

For the rest of this section we fix a *-regular semigroup S(S,,,)S\equiv(S,\cdot,{}^{*},{}^{\dagger}), and denote the corresponding DRC-semigroup from Theorem 9.5 by S=𝐈(S)=(S,,D,R)S^{\prime}={\bf I}(S)=(S,\cdot,D,R). Using (9.3) and (9.4) we see that

R(a)=(a)a=(a)a=(aa)=aa=D(a),and similarlyD(a)=R(a),R(a^{*})=(a^{*})^{\dagger}a^{*}=(a^{\dagger})^{*}a^{*}=(aa^{\dagger})^{*}=aa^{\dagger}=D(a),\qquad\text{and similarly}\qquad D(a^{*})=R(a), (9.8)

meaning that SS^{\prime} is an involutory DRC-semigroup in the terminology of Jones [30, Section 4]. Among other things, this leads to a simplification in the structure of the projection algebra P=𝐏(S)P={\bf P}(S^{\prime}). First, we note that

P={pS:p2=p=p}={pS:p2=p=p}.P=\{p\in S:p^{2}=p=p^{*}\}=\{p\in S:p^{2}=p=p^{\dagger}\}.

Indeed, the first equality was shown in [30, Proposition 4.2]. For the second, suppose first that p2=p=pp^{2}=p=p^{*}. It follows that pp is an inverse of pp that is \mathrel{\mathscr{H}}-related (indeed, equal) to pp^{*}, so that p=p{p=p^{\dagger}}. Conversely, if p2=p=pp^{2}=p=p^{\dagger}, then p=pp=pp=(pp)=pp=pp=pp^{\dagger}=(pp^{\dagger})^{*}=p^{*}. Using (9.8), it then follows that for any p,qPp,q\in P we have

qθp=R(qp)=R(qp)=R((pq))=D(pq)=qδp.q\theta_{p}=R(qp)=R(q^{*}p^{*})=R((pq)^{*})=D(pq)=q\delta_{p}.

In other words, we have θp=δp\theta_{p}=\delta_{p} for all pp, meaning that the two families of operations of PP reduces to one, and PP is symmetric in the terminology of Jones [30, Section 11].

In fact, not only do the θp\theta_{p} and δp\delta_{p} operations coincide in a *-regular semigroup, but they can be defined equationally in terms of the operation. To describe this, and for later use, we need the following basic result:

Lemma 9.9.

If SS is a *-regular semigroup, then for any aSa\in S and p𝐏(S)p\in{\bf P}(S) we have

(pa)=(pa)pand(ap)=p(ap).(pa)^{\dagger}=(pa)^{\dagger}p\qquad\text{and}\qquad(ap)^{\dagger}=p(ap)^{\dagger}.
Proof.

Since (pa)(pa)=ap(pa)^{\dagger}\mathrel{\mathscr{H}}(pa)^{*}=a^{*}p, it follows that pp is a right identity for (pa)(pa)^{\dagger}, which gives the first identity. The second is dual. ∎

Proposition 9.10.

If SS is a *-regular semigroup, then for any p,q𝐏(S)p,q\in{\bf P}(S) we have

qθp=qδp=p(pq)=(qp)p.q\theta_{p}=q\delta_{p}=p(pq)^{\dagger}=(qp)^{\dagger}p.
Proof.

Using Lemma 9.9 we see that qθp=R(qp)=(qp)qp=(qp)pq\theta_{p}=R(qp)=(qp)^{\dagger}qp=(qp)^{\dagger}p. A symmetrical calculation gives qδp=p(pq){q\delta_{p}=p(pq)^{\dagger}}, and we have already seen that qθp=qδpq\theta_{p}=q\delta_{p}. ∎

Remark 9.11.

In the case that SS is a regular *-semigroup, we obtain the simpler expression

qθp=p(pq)=p(pq)=pqp.q\theta_{p}=p(pq)^{\dagger}=p(pq)^{*}=pqp. (9.12)

This does not hold in *-regular semigroups in general, however. For example, with the projections p=[1/21/21/21/2]p=\left[\begin{smallmatrix}\nicefrac{{1}}{{2}}&\nicefrac{{1}}{{2}}\\ \nicefrac{{1}}{{2}}&\nicefrac{{1}}{{2}}\end{smallmatrix}\right] and q=[1000]q=\big{[}\begin{smallmatrix}1&0\\ 0&0\end{smallmatrix}\big{]}, both from M2()M_{2}(\mathbb{R}), the matrix pqp=[1/41/41/41/4]pqp=\left[\begin{smallmatrix}\nicefrac{{1}}{{4}}&\nicefrac{{1}}{{4}}\\ \nicefrac{{1}}{{4}}&\nicefrac{{1}}{{4}}\end{smallmatrix}\right] is not an idempotent, and hence not a projection.

The projection algebras of regular *-semigroups were axiomatised by Imaoka [28], as the algebras P(P,θ)P\equiv(P,\theta), where θ={θp:pP}\theta=\{\theta_{p}:p\in P\} is a set of unary operations satisfying

  1. (P1)

    pθp=pp\theta_{p}=p,

  2. (P2)

    θpθp=θp\theta_{p}\theta_{p}=\theta_{p},

  3. (P3)

    pθqθp=qθpp\theta_{q}\theta_{p}=q\theta_{p},

  4. (P4)

    θpθqθp=θqθp\theta_{p}\theta_{q}\theta_{p}=\theta_{q\theta_{p}},

  5. (P5)

    θpθqθpθq=θpθq\theta_{p}\theta_{q}\theta_{p}\theta_{q}=\theta_{p}\theta_{q}.

It is easy to see that these axioms hold in a regular *-semigroup (using (9.12)), and that they imply axioms (P1)(P5), keeping θp=δp\theta_{p}=\delta_{p} in mind. Note that (P1) and (P2) are exactly (P1) and (P6), respectively, while (P3) is the θ=δ\theta=\delta version of (P9). In particular, axioms (P1)(P3) hold in the projection algebra of a *-regular semigroup. Axioms (P4) and (P5) do not, however. For example, in M3()M_{3}(\mathbb{R}) we have

tθpθqθptθqθpandtθpθqθpθqtθpθqt\theta_{p}\theta_{q}\theta_{p}\not=t\theta_{q\theta_{p}}\qquad\text{and}\qquad t\theta_{p}\theta_{q}\theta_{p}\theta_{q}\not=t\theta_{p}\theta_{q}

for

t=[1/21/201/21/20001],p=[1/201/20101/201/2]andq=[10001/21/201/21/2].t=\left[\begin{smallmatrix}\nicefrac{{1}}{{2}}&\nicefrac{{1}}{{2}}&0\\ \nicefrac{{1}}{{2}}&\nicefrac{{1}}{{2}}&0\\ 0&0&1\end{smallmatrix}\right],\qquad p=\left[\begin{smallmatrix}\nicefrac{{1}}{{2}}&0&\nicefrac{{1}}{{2}}\\ 0&1&0\\ \nicefrac{{1}}{{2}}&0&\nicefrac{{1}}{{2}}\end{smallmatrix}\right]\qquad\text{and}\qquad q=\left[\begin{smallmatrix}1&0&0\\ 0&\nicefrac{{1}}{{2}}&\nicefrac{{1}}{{2}}\\ 0&\nicefrac{{1}}{{2}}&\nicefrac{{1}}{{2}}\end{smallmatrix}\right].

(This example was found, and can be checked, with Matlab [54], as with all the examples to come.)

As far as we are aware, there is no known axiomatisation for the projection algebras of *-regular semigroups. Obtaining such an axiomatisation is a necessary first step in tackling Problem 9.7. We also believe it is of interest in its own right, as among other things it could also be useful in developing a theory of fundamental *-regular semigroups; cf. [28, 30, 29, 59].

Problem 9.13.

Give an (abstract) axiomatisation for the class of projection algebras of *-regular semigroups.

9.4 Chained projection categories of regular *- and *-regular semigroups

We continue to fix the *-regular semigroup S(S,,,)S\equiv(S,\cdot,{}^{*},{}^{\dagger}), and its associated DRC-semigroup S=𝐈(S)=(S,,D,R)S^{\prime}={\bf I}(S)=(S,\cdot,D,R) from Theorem 9.5. Also let (P,𝒞,ε)=𝐂(S)(P,\mathcal{C},\varepsilon)={\bf C}(S^{\prime}) be the chained projection category associated to SS^{\prime}. In particular, P=𝐏(S)P={\bf P}(S^{\prime}) is the projection algebra of SS^{\prime}. Note that for any a𝒞(=S)a\in\mathcal{C}\ (=S) we have

𝐝(a)=D(a)=aa,and similarly𝐫(a)=aa.{\bf d}(a)=D(a)=aa^{\dagger},\qquad\text{and similarly}\qquad{\bf r}(a)=a^{\dagger}a.
Proposition 9.14.

If SS is a *-regular semigroup, then 𝒞=𝒞(S)\mathcal{C}=\mathcal{C}(S^{\prime}) is a groupoid, in which a1=aa^{-1}=a^{\dagger} for all a𝒞a\in\mathcal{C}.

Proof.

Fix aSa\in S, and first note that (9.4) gives

𝐝(a)=aa=aa=𝐫(a),and similarly𝐫(a)=𝐝(a).{\bf d}(a^{\dagger})=a^{\dagger}a^{\dagger\dagger}=a^{\dagger}a={\bf r}(a),\qquad\text{and similarly}\qquad{\bf r}(a^{\dagger})={\bf d}(a).

It follows that the compositions aaa\circ a^{\dagger} and aaa^{\dagger}\circ a exist in 𝒞\mathcal{C}. Specifically, we have

𝐝(a)=aa=aa,and similarly𝐫(a)=aa,{\bf d}(a)=aa^{\dagger}=a\circ a^{\dagger},\qquad\text{and similarly}\qquad{\bf r}(a)=a^{\dagger}\circ a,

so that aa^{\dagger} is indeed the inverse of aa in 𝒞\mathcal{C}. ∎

Using (9.3) and (9.4), it is also worth noting that for aSa\in S we have

𝐝(a)=aa=aa=(aa)=aa=𝐫(a),and similarly𝐫(a)=𝐝(a).{\bf d}(a^{*})=a^{*}a^{*\dagger}=a^{*}a^{\dagger*}=(a^{\dagger}a)^{*}=a^{\dagger}a={\bf r}(a),\qquad\text{and similarly}\qquad{\bf r}(a^{*})={\bf d}(a).

In what follows, we will typically use the laws 𝐝(a)=𝐝(a)=𝐫(a){\bf d}(a^{\dagger})={\bf d}(a^{*})={\bf r}(a) and 𝐫(a)=𝐫(a)=𝐝(a){\bf r}(a^{\dagger})={\bf r}(a^{*})={\bf d}(a) without explicit reference.

A further simplification occurs in the special case that SS is a regular *-semigroup (where we recall that and coincide).

Proposition 9.15.

If SS is a regular *-semigroup, then the orders \leq_{\ell} and r\leq_{r} on 𝒞\mathcal{C} coincide.

Proof.

By symmetry it suffices to show that abarba\leq_{\ell}b\ \Rightarrow\ a\leq_{r}b for a,b𝒞a,b\in\mathcal{C}, so suppose aba\leq_{\ell}b. This means that a=pb=pba={}_{p}{\downharpoonleft}b=pb, where p=𝐝(a)𝐝(b)=bbp={\bf d}(a)\leq{\bf d}(b)=bb^{*}. It follows from the latter that p=bbpp=bb^{*}p. Write q=𝐫(a)=aaq={\bf r}(a)=a^{*}a, and note that the biordered category axioms give q𝐫(b)q\leq{\bf r}(b). It then follows that

bq=bq=baa=b(pb)a=bbpa=pa=a,b{\downharpoonright}_{q}=bq=ba^{*}a=b(pb)^{*}a=bb^{*}pa=pa=a,

so that arba\leq_{r}b. ∎

Remark 9.16.

The orders \leq_{\ell} and r\leq_{r} are generally distinct for *-regular semigroups. For example, consider the following matrices from M2()M_{2}(\mathbb{R}):

a=[1100],b=[1101],p=[1000],q=[1/21/21/21/2]ande=[1001].a=\big{[}\begin{smallmatrix}1&1\\ 0&0\end{smallmatrix}\big{]},\qquad b=\big{[}\begin{smallmatrix}1&1\\ 0&1\end{smallmatrix}\big{]},\qquad p=\big{[}\begin{smallmatrix}1&0\\ 0&0\end{smallmatrix}\big{]},\qquad q=\left[\begin{smallmatrix}\nicefrac{{1}}{{2}}&\nicefrac{{1}}{{2}}\\ \nicefrac{{1}}{{2}}&\nicefrac{{1}}{{2}}\end{smallmatrix}\right]\qquad\text{and}\qquad e=\big{[}\begin{smallmatrix}1&0\\ 0&1\end{smallmatrix}\big{]}.

Then pp, qq and ee are projections with p,qep,q\leq e, and we have D(a)=pD(a)=p, R(a)=qR(a)=q and D(b)=R(b)=e{D(b)=R(b)=e}. We also have D(a)b=pb=a{}_{D(a)}{\downharpoonleft}b=pb=a, so that aba\leq_{\ell}b. But arba\not\leq_{r}b, since bR(a)=bq=[111/21/2]ab{\downharpoonright}_{R(a)}=bq=\left[\begin{smallmatrix}1&1\\ \nicefrac{{1}}{{2}}&\nicefrac{{1}}{{2}}\end{smallmatrix}\right]\not=a.

Another special property enjoyed by the category 𝒞\mathcal{C} has to do with the ϑ/Θ\vartheta/\Theta and /Δ\partial/\Delta maps from (2.5) and (5.2). We summarise this in Proposition 9.18 below, but we first note that for a𝒞a\in\mathcal{C} we have

pϑa\displaystyle p\vartheta_{a} =(pa)a for p𝐝(a),\displaystyle=(pa)^{\dagger}a\text{ \ for $p\leq{\bf d}(a)$,} pΘa=(pa)a for pP,\displaystyle p\Theta_{a}=(pa)^{\dagger}a\text{ \ for $p\in P$,}
pa\displaystyle p\partial_{a} =a(ap) for p𝐫(a),\displaystyle=a(ap)^{\dagger}\text{ \ for $p\leq{\bf r}(a)$,} pΔa=a(ap) for pP.\displaystyle p\Delta_{a}=a(ap)^{\dagger}\text{ \ for $p\in P$.} (9.17)

Indeed, if p𝐝(a)p\leq{\bf d}(a), then using (5.5) and Lemma 9.9 we have

pϑa=R(pa)=(pa)pa=(pa)a.p\vartheta_{a}=R(pa)=(pa)^{\dagger}pa=(pa)^{\dagger}a.

The other formulae are checked analogously. If SS is a regular *-semigroup, then these reduce further; for example, pΘa=apap\Theta_{a}=a^{*}pa for a𝒞a\in\mathcal{C} and pPp\in P. Such simplifications do not hold in general for *-regular semigroups; for example, for the projections p,qM2()p,q\in M_{2}(\mathbb{R}) from Remark 9.11, we have qΘp=qθppqp=pqpq\Theta_{p}=q\theta_{p}\not=pqp=p^{*}qp.

Proposition 9.18.

If SS is a *-regular semigroup, then for any a𝒞a\in\mathcal{C}:

  1. (i)

    ϑa\vartheta_{a} and a\partial_{a} are bijections, and ϑa=ϑa1\vartheta_{a^{\dagger}}=\vartheta_{a}^{-1} and a=a1\partial_{a^{\dagger}}=\partial_{a}^{-1},

  2. (ii)

    ϑa=a\vartheta_{a^{*}}=\partial_{a} and a=ϑa\partial_{a^{*}}=\vartheta_{a},

  3. (iii)

    Θa=Δa\Theta_{a^{*}}=\Delta_{a} and Δa=Θa\Delta_{a^{*}}=\Theta_{a}.

Proof.

For each item, it suffices by symmetry to prove the first assertion.

(i). It suffices by symmetry to show that ϑaϑa=id𝐝(a)\vartheta_{a}\vartheta_{a^{\dagger}}=\operatorname{id}_{{\bf d}(a)^{\downarrow}}, i.e. that pϑaϑa=pp\vartheta_{a}\vartheta_{a^{\dagger}}=p for all p𝐝(a)p\leq{\bf d}(a). To do so, fix some such pp, and write q=pϑa=𝐫(pa)q=p\vartheta_{a}={\bf r}({}_{p}{\downharpoonleft}a). We then have

p=p𝐝(a)=p(aa)=paqa,p={}_{p}{\downharpoonleft}{\bf d}(a)={}_{p}{\downharpoonleft}(a\circ a^{\dagger})={}_{p}{\downharpoonleft}a\circ{}_{q}{\downharpoonleft}a^{\dagger},

from which it follows that p=𝐫(p)=𝐫(paqa)=𝐫(qa)=qϑa=pϑaϑap={\bf r}(p)={\bf r}({}_{p}{\downharpoonleft}a\circ{}_{q}{\downharpoonleft}a^{\dagger})={\bf r}({}_{q}{\downharpoonleft}a^{\dagger})=q\vartheta_{a^{\dagger}}=p\vartheta_{a}\vartheta_{a^{\dagger}}.

(ii). Let p𝐫(a)p\leq{\bf r}(a). Keeping in mind that pap\partial_{a} is a projection, we use (9.17) and (9.4) to calculate

pa=(pa)=(a(ap))=(ap)a=(ap)a=(pa)a=pϑa.p\partial_{a}=(p\partial_{a})^{*}=(a(ap)^{\dagger})^{*}=(ap)^{\dagger*}a^{*}=(ap)^{*\dagger}a^{*}=(pa^{*})^{\dagger}a^{*}=p\vartheta_{a^{*}}.

(iii). Using previously-established facts, we have Θa=θ𝐝(a)ϑa=δ𝐫(a)a=Δa\Theta_{a^{*}}=\theta_{{\bf d}(a^{*})}\vartheta_{a^{*}}=\delta_{{\bf r}(a)}\partial_{a}=\Delta_{a}. ∎

Remark 9.19.

When SS is a regular *-semigroup, Proposition 9.18(i) says that ϑa\vartheta_{a^{*}} is the inverse of ϑa\vartheta_{a}. This was proved in [13, Lemma 2.12] using the identity pa=aq{}_{p}{\downharpoonleft}a=a{\downharpoonright}_{q}, where p𝐝(a)p\leq{\bf d}(a) and q=𝐫(pa)q={\bf r}({}_{p}{\downharpoonleft}a). This does not hold in a general *-regular semigroup. For example, with b,p,qM2()b,p,q\in M_{2}(\mathbb{R}) as in Remark 9.16, we have p𝐝(b)p\leq{\bf d}(b) and q=𝐫(pb)q={\bf r}({}_{p}{\downharpoonleft}b), but pbbq{}_{p}{\downharpoonleft}b\not=b{\downharpoonright}_{q}.

Remark 9.20.

Another point of difference concerning the ϑa\vartheta_{a} and a\partial_{a} maps is that these are always projection algebra morphisms when SS is a regular *-semigroup, as shown in the proof [13, Proposition 6.11]. This property was called (G1d) in [13], and was shown to be equivalent to certain other natural conditions in the context of the projection groupoids of [13], including our (C1). However, the maps ϑa\vartheta_{a} and a\partial_{a} need not be projection algebra morphims when SS is *-regular.444This is also the case in general for Ehresmann semigroups; see for example [20, Section 4.2]. For example, we have (pθq)ϑa(pϑa)θqϑa(p\theta_{q})\vartheta_{a}\not=(p\vartheta_{a})\theta_{q\vartheta_{a}} for a=[1101]a=\big{[}\begin{smallmatrix}1&1\\ 0&1\end{smallmatrix}\big{]}, p=[1000]p=\big{[}\begin{smallmatrix}1&0\\ 0&0\end{smallmatrix}\big{]} and q=[0001]q=\big{[}\begin{smallmatrix}0&0\\ 0&1\end{smallmatrix}\big{]}, all from M2()M_{2}(\mathbb{R}). It is worth noting that one of the other equivalent conditions in (G1) is:

θpΘa=ΘaθpΘa(=ΔaθpΘa)for all pP and a𝒞.\theta_{p\Theta_{a}}=\Theta_{a^{*}}\theta_{p}\Theta_{a}\ (=\Delta_{a}\theta_{p}\Theta_{a})\qquad\text{for all $p\in P$ and $a\in\mathcal{C}$.} (9.21)

This also need not hold for *-regular semigroups. For example, for p=[1001]p=\big{[}\begin{smallmatrix}1&0\\ 0&1\end{smallmatrix}\big{]}, t=[1000]t=\big{[}\begin{smallmatrix}1&0\\ 0&0\end{smallmatrix}\big{]} and a=[1101]a=\big{[}\begin{smallmatrix}1&1\\ 0&1\end{smallmatrix}\big{]} from M2()M_{2}(\mathbb{R}) we have tθpΘa=[1000]t\theta_{p\Theta_{a}}=\big{[}\begin{smallmatrix}1&0\\ 0&0\end{smallmatrix}\big{]} and tΔaθpΘa=[1/21/21/21/2]t\Delta_{a}\theta_{p}\Theta_{a}=\left[\begin{smallmatrix}\nicefrac{{1}}{{2}}&\nicefrac{{1}}{{2}}\\ \nicefrac{{1}}{{2}}&\nicefrac{{1}}{{2}}\end{smallmatrix}\right].

The final issue we wish to address concerns axiom (C2), and is somewhat more technical. Recall that this is quantified over all morphisms b𝒞b\in\mathcal{C} and all projections p,sPp,s\in P, and asserts the equality of the morphisms

λ(p,b,s)=ε[e,e1](pθqb)f1ε[f1,f]andρ(p,b,s)=ε[e,e2]e2(bsδr)ε[f2,f]\lambda(p,b,s)=\varepsilon[e,e_{1}]\circ({}_{p\theta_{q}}{\downharpoonleft}b){\downharpoonright}_{f_{1}}\circ\varepsilon[f_{1},f]\qquad\text{and}\qquad\rho(p,b,s)=\varepsilon[e,e_{2}]\circ{}_{e_{2}}{\downharpoonleft}(b{\downharpoonright}_{s\delta_{r}})\circ\varepsilon[f_{2},f] (9.22)

in 𝒞\mathcal{C}. These morphisms were defined in terms of the projections q=𝐝(b)q={\bf d}(b), r=𝐫(b)r={\bf r}(b), and e,ei,f,fie,e_{i},f,f_{i} given by

e(p,b,s)\displaystyle e(p,b,s) =sΔbδp,\displaystyle=s\Delta_{b}\delta_{p}, e1(p,b,s)\displaystyle e_{1}(p,b,s) =sΔbδpθq,\displaystyle=s\Delta_{b}\delta_{p\theta_{q}}, f1(p,b,s)\displaystyle f_{1}(p,b,s) =sδpΘb,\displaystyle=s\delta_{p\Theta_{b}}, f(p,b,s)\displaystyle f(p,b,s) =pΘbθs,\displaystyle=p\Theta_{b}\theta_{s},
e2(p,b,s)\displaystyle e_{2}(p,b,s) =pθsΔb,\displaystyle=p\theta_{s\Delta_{b}}, f2(p,b,s)\displaystyle f_{2}(p,b,s) =pΘbθsδr.\displaystyle=p\Theta_{b}\theta_{s\delta_{r}}. (9.23)

All of the above projections and morphisms are shown in Figure 2. The idea behind axiom (C2) is that if we have three morphisms a,b,c𝒞a,b,c\in\mathcal{C}, with

p=𝐫(a),q=𝐝(b),r=𝐫(b)ands=𝐝(c),p={\bf r}(a),\qquad q={\bf d}(b),\qquad r={\bf r}(b)\qquad\text{and}\qquad s={\bf d}(c),

then

(ab)c=aeλ(p,b,s)fcanda(bc)=aeρ(p,b,s)fc.(a\bullet b)\bullet c=a{\downharpoonright}_{e}\circ\lambda(p,b,s)\circ{}_{f}{\downharpoonleft}c\qquad\text{and}\qquad a\bullet(b\bullet c)=a{\downharpoonright}_{e}\circ\rho(p,b,s)\circ{}_{f}{\downharpoonleft}c.

(See the proof of Lemma 6.3.)

In the case of regular *-semigroups, the corresponding axiom (G2) from [13] was quantified over all morphisms bb, and all bb-linked pairs (e,f)(e,f). Such a pair satisfies

eΘbθf=fandfΔbδe=e.e\Theta_{b}\theta_{f}=f\qquad\text{and}\qquad f\Delta_{b}\delta_{e}=e. (9.24)

(In the notation of [13] the latter equation was fΘbθe=ef\Theta_{b^{*}}\theta_{e}=e, which is equivalent to the above since δe=θe\delta_{e}=\theta_{e} and Δb=Θb\Delta_{b}=\Theta_{b^{*}}.) The idea here is that if p,sPp,s\in P are arbitrary, then e=e(p,b,s)e=e(p,b,s) and f=f(p,b,s)f=f(p,b,s) from (9.23) are bb-linked, and we have

e(e,b,f)=e,ei(e,b,f)=ei(p,b,s),fi(e,b,f)=fi(p,b,s)andf(e,b,f)=f,e(e,b,f)=e,\qquad e_{i}(e,b,f)=e_{i}(p,b,s),\qquad f_{i}(e,b,f)=f_{i}(p,b,s)\qquad\text{and}\qquad f(e,b,f)=f, (9.25)

which in turn leads to λ(p,b,s)=λ(e,b,f)\lambda(p,b,s)=\lambda(e,b,f) and ρ(p,b,s)=ρ(e,b,f)\rho(p,b,s)=\rho(e,b,f). For example, using (9.21), (9.23), (9.24) and θt=δt\theta_{t}=\delta_{t}, we have

f1(e,b,f)=fδeΘb=fθeΘb=fΔbθeΘb\displaystyle f_{1}(e,b,f)=f\delta_{e\Theta_{b}}=f\theta_{e\Theta_{b}}=f\Delta_{b}\theta_{e}\Theta_{b} =eΘb\displaystyle=e\Theta_{b}
=(sΔbδp)Θb=sΔbθpΘb=sθpΘb=sδpΘb=f1(p,b,s).\displaystyle=(s\Delta_{b}\delta_{p})\Theta_{b}=s\Delta_{b}\theta_{p}\Theta_{b}=s\theta_{p\Theta_{b}}=s\delta_{p\Theta_{b}}=f_{1}(p,b,s).

In the general case of a *-regular semigroup, the equalities involving eie_{i} and fif_{i} in (9.25) need not hold. For example, consider the following matrices from M3()M_{3}(\mathbb{R}):

b=[110110001],p=[10001/21/201/21/2]ands=[000000001].b=\left[\begin{smallmatrix}1&1&0\\ 1&1&0\\ 0&0&1\end{smallmatrix}\right],\qquad p=\left[\begin{smallmatrix}1&0&0&\\ 0&\nicefrac{{1}}{{2}}&\nicefrac{{1}}{{2}}\\ 0&\nicefrac{{1}}{{2}}&\nicefrac{{1}}{{2}}\end{smallmatrix}\right]\qquad\text{and}\qquad s=\left[\begin{smallmatrix}0&0&0\\ 0&0&0\\ 0&0&1\end{smallmatrix}\right]. (9.26)

Then pp and ss are projections, and we have

e=e(p,b,s)=[00001/21/201/21/2]andf=f(p,b,s)=s=[000000001].e=e(p,b,s)=\left[\begin{smallmatrix}0&0&0&\\ 0&\nicefrac{{1}}{{2}}&\nicefrac{{1}}{{2}}\\ 0&\nicefrac{{1}}{{2}}&\nicefrac{{1}}{{2}}\end{smallmatrix}\right]\qquad\text{and}\qquad f=f(p,b,s)=s=\left[\begin{smallmatrix}0&0&0\\ 0&0&0\\ 0&0&1\end{smallmatrix}\right]. (9.27)

Moreover, we have ei(p,b,s)=fi(p,b,s)=se_{i}(p,b,s)=f_{i}(p,b,s)=s for i=1,2i=1,2. However, although e2(e,b,f)=e2(p,b,s){e_{2}(e,b,f)=e_{2}(p,b,s)} and f2(e,b,f)=f2(p,b,s)f_{2}(e,b,f)=f_{2}(p,b,s), we have

e1(e,b,f)=[1/61/61/31/61/61/31/31/32/3]e1(p,b,s)andf1(e,b,f)=[1/31/31/31/31/31/31/31/31/3]f1(p,b,s).e_{1}(e,b,f)=\left[\begin{smallmatrix}\nicefrac{{1}}{{6}}&\nicefrac{{1}}{{6}}&\nicefrac{{1}}{{3}}\\ \nicefrac{{1}}{{6}}&\nicefrac{{1}}{{6}}&\nicefrac{{1}}{{3}}\\ \nicefrac{{1}}{{3}}&\nicefrac{{1}}{{3}}&\nicefrac{{2}}{{3}}\end{smallmatrix}\right]\not=e_{1}(p,b,s)\qquad\text{and}\qquad f_{1}(e,b,f)=\left[\begin{smallmatrix}\nicefrac{{1}}{{3}}&\nicefrac{{1}}{{3}}&\nicefrac{{1}}{{3}}\\ \nicefrac{{1}}{{3}}&\nicefrac{{1}}{{3}}&\nicefrac{{1}}{{3}}\\ \nicefrac{{1}}{{3}}&\nicefrac{{1}}{{3}}&\nicefrac{{1}}{{3}}\end{smallmatrix}\right]\not=f_{1}(p,b,s). (9.28)

Not only can axiom (C2) be quantified over a smaller set of projections when SS is a regular *-semigroup, but the projections eie_{i} and fif_{i} in (9.23) take on much simpler forms. Specifically, when (e,f)(e,f) is bb-linked (and SS is a regular *-semigroup), we have

e1(e,b,f)=eθq,e2(e,b,f)=fΔb,f1(e,b,f)=eΘbandf2(e,b,f)=fδr,e_{1}(e,b,f)=e\theta_{q},\qquad e_{2}(e,b,f)=f\Delta_{b},\qquad f_{1}(e,b,f)=e\Theta_{b}\qquad\text{and}\qquad f_{2}(e,b,f)=f\delta_{r}, (9.29)

where again q=𝐝(b)q={\bf d}(b) and r=𝐫(b)r={\bf r}(b). For example, using θt=δt\theta_{t}=\delta_{t}, (P4), Lemma 5.3(ii) and (9.24), we have

e1(e,b,f)=fΔbδeθq=fΔbθeθq=fΔbθqθeθq=fΔbδ𝐝(b)θeθq=fΔbθeθq=eθq.e_{1}(e,b,f)=f\Delta_{b}\delta_{e\theta_{q}}=f\Delta_{b}\theta_{e\theta_{q}}=f\Delta_{b}\theta_{q}\theta_{e}\theta_{q}=f\Delta_{b}\delta_{{\bf d}(b)}\theta_{e}\theta_{q}=f\Delta_{b}\theta_{e}\theta_{q}=e\theta_{q}.

The simplifications in (9.29) do not hold in general DRC-semigroups. Indeed, small counterexamples can be readily constructed using Mace4 [40], although we do not currently know of any *-regular counterexamples.

As well as the simplification in the projections (9.23) and (9.29), the ‘double restrictions’ in (9.22) are just single restrictions for regular *-semigroups, as equality of the \leq_{\ell} and r\leq_{r} orders leads to the identities (pa)q=aq({}_{p}{\downharpoonleft}a){\downharpoonright}_{q}=a{\downharpoonright}_{q} and r(as)=ra{}_{r}{\downharpoonleft}(a{\downharpoonright}_{s})={}_{r}{\downharpoonleft}a (for appropriate p,q,r,sp,q,r,s). As usual, these do not hold in general *-regular semigroups.

10 Other special cases

In this final section we consider another two important subclasses of DRC-semigroups, specifically the Ehresmann and DRC-restriction semigroups considered by Lawson [33] and Die and Wang [8]. Again, although we could in principle deduce the main results of [33, 8] from ours, we will not give all of the details. Rather, we just outline the ways in which our constructions simplify in these cases.

10.1 Ehresmann semigroups

We noted in Section 3.1 that the Ehresmann semigroups of Lawson [33] are precisely the DRC-semigroups SS satisfying the additional law

D(a)D(b)=D(b)D(a)andR(a)R(b)=R(b)R(a)for all a,bS.D(a)D(b)=D(b)D(a)\qquad\text{and}\qquad R(a)R(b)=R(b)R(a)\qquad\text{for all $a,b\in S$.}

Given (DRC4), either of the above identities implies the other, and of course they say that projections of SS commute. This is a very strong property, and to unpack it we first prove the following:

Proposition 10.1.

If p,q𝐏(S)p,q\in{\bf P}(S) for a DRC-semigroup SS, then the following are equivalent:

  1. (i)

    pq=qppq=qp,

  2. (ii)

    pθq=qθpp\theta_{q}=q\theta_{p} and pδq=qδpp\delta_{q}=q\delta_{p},

  3. (iii)

    pθq=qθp=pδq=qδpp\theta_{q}=q\theta_{p}=p\delta_{q}=q\delta_{p},

  4. (iv)

    pθq=qθp=pδq=qδp=pq=qpp\theta_{q}=q\theta_{p}=p\delta_{q}=q\delta_{p}=pq=qp.

Proof.

(i)\ \Rightarrow\ (iv). Suppose first that pq=qppq=qp. By Lemma 3.4 we have D(pq)=D(pq)pD(pq)=D(pq)p, and since D(pq)=D(qp){D(pq)=D(qp)}, the same lemma gives D(pq)=D(pq)qD(pq)=D(pq)q. Together with (DRC1) these imply D(pq)=D(pq)pq=pq{D(pq)=D(pq)pq=pq}, which says qδp=pqq\delta_{p}=pq. The remaining equalities are analogous.

(iv)\ \Rightarrow\ (iii)\ \Rightarrow\ (ii). These implications are clear.

(ii)\ \Rightarrow\ (i). Now suppose pθq=qθpp\theta_{q}=q\theta_{p} and pδq=qδpp\delta_{q}=q\delta_{p}, which means that R(pq)=R(qp)R(pq)=R(qp) and D(pq)=D(qp)D(pq)=D(qp). As in (7.4), we have pq=D(pq)R(pq)pq=D(pq)R(pq), so also

qp=D(qp)R(qp)=D(pq)R(pq)=pq.qp=D(qp)R(qp)=D(pq)R(pq)=pq.\qed
Corollary 10.2.

If p,q𝐏(S)p,q\in{\bf P}(S) for a DRC-semigroup SS, and if pq=qppq=qp, then

  1. (i)

    the meet pqp\wedge q exists in the poset (P,)(P,\leq), and pq=pq=qpp\wedge q=pq=qp,

  2. (ii)

    pqp=qp\mathrel{\mathscr{F}}q\ \Leftrightarrow\ p=q.

Proof.

(i). By Proposition 10.1 we have pq=pθqPpq=p\theta_{q}\in P, and moreover pq=pθqqpq=p\theta_{q}\leq q and similarly pqppq\leq p. So pqpq is a lower bound for pp and qq. If we had another lower bound rp,qr\leq p,q, then r=prp=qrqr=prp=qrq (cf. (3.5)); combined with pq=qppq=qp this leads to r=(pq)r(pq)r=(pq)r(pq), i.e. rpqr\leq pq.

(ii). If pqp\mathrel{\mathscr{F}}q, then p=qδpp=q\delta_{p} and q=pθqq=p\theta_{q}, and since qδp=pθqq\delta_{p}=p\theta_{q} (by Proposition 10.1) it follows that p=qp=q. ∎

Corollary 10.2 has some important simplifying consequences for an Ehresmann semigroup SS, since all projections commute. First, P=𝐏(S)P={\bf P}(S) is a meet semilattice, and hence a subsemigroup of SS. Second, =ιP{\mathrel{\mathscr{F}}}=\iota_{P} is the equality relation. Thus, the chain category 𝒞(P)\mathscr{C}(P) is simply a copy of PP, in which the only compositions are p=ppp=p\circ p (pPp\in P), and in which =r={\leq_{\ell}}={\leq_{r}}={\leq} is the order from (4.9). It follows that the chained projection category (P,𝒞,ε)=𝐂(S)(P,\mathcal{C},\varepsilon)={\bf C}(S) of the Ehresmann semigroup SS carries no more information than the projection category (P,𝒞)(P,\mathcal{C}) itself. Moreover, since =ιP{\mathrel{\mathscr{F}}}=\iota_{P} is trivial, axiom (C2) simplifies as well. Indeed, we first note that since ee1,e2e\mathrel{\mathscr{F}}e_{1},e_{2} and f1,f2ff_{1},f_{2}\mathrel{\mathscr{F}}f by Lemma 5.16, it follows that e=e1=e2e=e_{1}=e_{2} and f=f1=f2f=f_{1}=f_{2}. Thus, (C2) asserts the equality of the double restrictions:

(pθqb)f=e(bsδr)for all b𝒞(q,r) and p,sP,  where e=sΔbδp and f=pΘbθs.({}_{p\theta_{q}}{\downharpoonleft}b){\downharpoonright}_{f}={}_{e}{\downharpoonleft}(b{\downharpoonright}_{s\delta_{r}})\qquad\text{for all $b\in\mathcal{C}(q,r)$ and $p,s\in P$, \quad where $e=s\Delta_{b}\delta_{p}$ and $f=p\Theta_{b}\theta_{s}$.} (10.3)

This can be used to prove the following, which is one of the Ehresmann category axioms from [33].

Proposition 10.4.

In an Ehresmann semigroup SS, we have r=r{\leq_{\ell}}\circ{\leq_{r}}={\leq_{r}}\circ{\leq_{\ell}}.

Proof.

By symmetry, it suffices to prove the inclusion rr{\leq_{\ell}}\circ{\leq_{r}}\subseteq{\leq_{r}}\circ{\leq_{\ell}}. To do so, suppose a,b𝒞=𝒞(S)a,b\in\mathcal{C}=\mathcal{C}(S) are such that aurba\leq_{\ell}u\leq_{r}b for some u𝒞u\in\mathcal{C}. This means that

a=p(bs)for some s𝐫(b) and p𝐝(bs)=sb.a={}_{p}{\downharpoonleft}(b{\downharpoonright}_{s})\qquad\text{for some $s\leq{\bf r}(b)$ and $p\leq{\bf d}(b{\downharpoonright}_{s})=s\partial_{b}$.}

(So here u=bsu=b{\downharpoonright}_{s}.) We can complete the proof by showing that

a=(pb)t,wheret=𝐫(a).a=({}_{p}{\downharpoonleft}b){\downharpoonright}_{t},\qquad\text{where}\qquad t={\bf r}(a). (10.5)

To do so, we now write q=𝐝(b)q={\bf d}(b) and r=𝐫(b)r={\bf r}(b), and we work towards interpreting (10.3) in the current context. First note that

s𝐫(b)=randp𝐝(bs)𝐝(b)=q,s\leq{\bf r}(b)=r\qquad\text{and}\qquad p\leq{\bf d}(b{\downharpoonright}_{s})\leq{\bf d}(b)=q,

which imply that pθq=pp\theta_{q}=p and sδr=ss\delta_{r}=s. We also have

e=sΔbδp=sδrbδp=sbδp=p,e=s\Delta_{b}\delta_{p}=s\delta_{r}\partial_{b}\delta_{p}=s\partial_{b}\delta_{p}=p,

because psbp\leq s\partial_{b}; cf. (4.9). Thus, (10.3) gives

(pb)f=p(bs),({}_{p}{\downharpoonleft}b){\downharpoonright}_{f}={}_{p}{\downharpoonleft}(b{\downharpoonright}_{s}),

where f=pΘbθsf=p\Theta_{b}\theta_{s}. Since a=p(bs)=(pb)fa={}_{p}{\downharpoonleft}(b{\downharpoonright}_{s})=({}_{p}{\downharpoonleft}b){\downharpoonright}_{f}, it follows that f=𝐫(a)f={\bf r}(a), completing the proof of (10.5), and hence of the proposition. ∎

Lawson in [33] denotes by e\leq_{e} the composition r=r{\leq_{\ell}}\circ{\leq_{r}}={\leq_{r}}\circ{\leq_{\ell}}, and we note that the above proof shows that if aeba\leq_{e}b, then we have

a=(pb)t=p(bt),wherep=𝐝(a)andt=𝐫(a).a=({}_{p}{\downharpoonleft}b){\downharpoonright}_{t}={}_{p}{\downharpoonleft}(b{\downharpoonright}_{t}),\qquad\text{where}\qquad p={\bf d}(a)\quad\text{and}\quad t={\bf r}(a).

This is [33, Lemma 4.6].

Finally, we note that =ιP{\mathrel{\mathscr{F}}}=\iota_{P} has the consequence that the \bullet product on the category 𝒞\mathcal{C} from Definition 6.1 takes on the simpler form

ab=aeeb,wheree=𝐫(a)𝐝(b).a\bullet b=a{\downharpoonright}_{e}\circ{}_{e}{\downharpoonleft}b,\qquad\text{where}\qquad e={\bf r}(a)\wedge{\bf d}(b).

10.2 DRC-restriction semigroups

We conclude the paper by briefly commenting on the DRC-restriction semigroups considered by Die and Wang in [8]. These are the DRC-semigroups S(S,,D,R)S\equiv(S,\cdot,D,R) satisfying the additional laws

D(ab)a=aD(R(a)b)andaR(ba)=R(bD(a))afor all a,bS.D(ab)a=aD(R(a)b)\qquad\text{and}\qquad aR(ba)=R(bD(a))a\qquad\text{for all $a,b\in S$}.

It was observed in [8, Section 1] that these laws need only be quantified over aSa\in S and b𝐏(S)b\in{\bf P}(S), given (DRC2). The DRC-restriction laws might seem unwieldy, but it turns out that they are equivalent to the orders \leq_{\ell} and r\leq_{r} coinciding, as shown in [8, Lemma 5.8].

The full subcategory 𝐃𝐑𝐂𝐑{\bf DRCR} of 𝐃𝐑𝐂{\bf DRC} consisting of all DRC-restriction semigroups (and DRC-morphisms) is isomorphic to its image under the isomorphism 𝐂:𝐃𝐑𝐂𝐂𝐏𝐂{\bf C}:{\bf DRC}\to{\bf CPC} from Theorem 6.7 (cf. Theorem 7.1). One could then use this to deduce the results from [8]. Again we omit the details, but we do note that the constructions and results of [8] involve some (but not all) of the simplifications that we discussed in Section 9 for regular *-semigroups. As we have already mentioned, the \leq_{\ell} and r\leq_{r} orders coincide, so one works with ordered rather than biordered categories. However, the θp\theta_{p} and δp\delta_{p} operations need not be equal, i.e. 𝐏(S){\bf P}(S) need not be symmetric. Similarly, the ϑ/Θ\vartheta/\Theta and /Δ\partial/\Delta maps need not be related to each other; for example, ϑa\vartheta_{a} (aSa\in S) need not be equal to any b\partial_{b} map (for bSb\in S). On the other hand, ϑa\vartheta_{a} is always a projection algebra morphism 𝐝(a)𝐫(a){\bf d}(a)^{\downarrow}\to{\bf r}(a)^{\downarrow}, with a similar statement for a\partial_{a}; see [8, Lemmas 4.7 and 7.7]. This leads to the various equivalent conditions in the analogue of (C1) in [8], including (9.21); cf. Remark 9.20. Likewise, the simplifications in (C2) discussed at the end of Section 9.4 all hold in DRC-restriction semigroups; specifically, the axiom can be quantified over bb-linked pairs (cf. (9.24)), and the projections ei,fie_{i},f_{i} take on the simpler forms in (9.29).

References

  • [1] S. Armstrong. Structure of concordant semigroups. J. Algebra, 118(1):205–260, 1988.
  • [2] S. Awodey. Category theory, volume 52 of Oxford Logic Guides. Oxford University Press, Oxford, second edition, 2010.
  • [3] 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.
  • [4] 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.
  • [5] S. Crvenković and I. Dolinka. Congruences on \ast-regular semigroups. Period. Math. Hungar., 45(1-2):1–13, 2002.
  • [6] 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.
  • [7] D. DeWolf and D. Pronk. The Ehresmann-Schein-Nambooripad theorem for inverse categories. Theory Appl. Categ., 33:Paper No. 27, 813–831, 2018.
  • [8] Y. Die and S. Wang. Chain projection ordered categories and DRC-restriction semigroups. Preprint, 2024, arXiv:2407.11434.
  • [9] I. Dolinka. Free idempotent generated semigroups: the word problem and structure via gain graphs. Israel J. Math., 245(1):347–387, 2021.
  • [10] 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.
  • [11] 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.
  • [12] M. P. Drazin. Regular semigroups with involution. In Symposium on Regular Semigroups, pages 29–46. Northern Illinois University, DeKalb, 1979.
  • [13] J. East and P. A. Azeef Muhammed. A groupoid approach to regular \ast-semigroups. Adv. Math., 437:Paper No. 109447, 104 pp., 2024.
  • [14] J. East and R. D. Gray. Ehresmann theory and partition monoids. J. Algebra, 579:318–352, 2021.
  • [15] J. East, R. D. Gray, P. A. Azeef Muhammed, and N. Ruškuc. Projection algebras and free projection- and idempotent-generated regular *-semigroups. Preprint, 2024, arXiv:2406.09109.
  • [16] D. G. FitzGerald. Factorizable inverse monoids. Semigroup Forum, 80(3):484–509, 2010.
  • [17] D. G. FitzGerald. Groupoids on a skew lattice of objects. Art Discrete Appl. Math., 2(2):Paper No. 2.03, 14, 2019.
  • [18] D. G. FitzGerald and M. K. Kinyon. Trace- and pseudo-products: restriction-like semigroups with a band of projections. Semigroup Forum, 103(3):848–866, 2021.
  • [19] V. Gould. Notes on restriction semigroups and related structures.
    https://www-users.york.ac.uk/~varg1/restriction.pdf, 2010.
  • [20] V. Gould. Restriction and Ehresmann semigroups. In Proceedings of the International Conference on Algebra 2010, pages 265–288. World Sci. Publ., Hackensack, NJ, 2012.
  • [21] V. Gould and C. Hollings. Restriction semigroups and inductive constellations. Comm. Algebra, 38(1):261–287, 2010.
  • [22] V. Gould and Y. Wang. Beyond orthodox semigroups. J. Algebra, 368:209–230, 2012.
  • [23] R. Gray and N. Ruškuc. On maximal subgroups of free idempotent generated semigroups. Israel J. Math., 189:147–176, 2012.
  • [24] R. Gray and N. Ruškuc. Maximal subgroups of free idempotent-generated semigroups over the full transformation monoid. Proc. Lond. Math. Soc. (3), 104(5):997–1018, 2012.
  • [25] C. Hollings. Extending the Ehresmann-Schein-Nambooripad theorem. Semigroup Forum, 80(3):453–476, 2010.
  • [26] C. Hollings. The Ehresmann-Schein-Nambooripad theorem and its successors. Eur. J. Pure Appl. Math., 5(4):414–450, 2012.
  • [27] 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.
  • [28] 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.
  • [29] P. R. Jones. A common framework for restriction semigroups and regular *-semigroups. J. Pure Appl. Algebra, 216(3):618–632, 2012.
  • [30] P. R. Jones. Generalized Munn representations of DRC-semigroups. Southeast Asian Bull. Math., 45(5):591–619, 2021.
  • [31] M. Khoshkam and G. Skandalis. Regular representation of groupoid CC^{*}-algebras and applications to inverse semigroups. J. Reine Angew. Math., 546:47–72, 2002.
  • [32] G. Lallement. Demi-groupes réguliers. Ann. Mat. Pura Appl. (4), 77:47–129, 1967.
  • [33] M. V. Lawson. Semigroups and ordered categories. I. The reduced case. J. Algebra, 141(2):422–462, 1991.
  • [34] M. V. Lawson. Inverse semigroups. World Scientific Publishing Co., Inc., River Edge, NJ, 1998. The theory of partial symmetries.
  • [35] M. V. Lawson. Ordered groupoids and left cancellative categories. Semigroup Forum, 68(3):458–476, 2004.
  • [36] M. V. Lawson. Recent developments in inverse semigroup theory. Semigroup Forum, 100(1):103–118, 2020.
  • [37] M. V. Lawson. On Ehresmann semigroups. Semigroup Forum, 103(3):953–965, 2021.
  • [38] S. Mac Lane. Categories for the working mathematician, volume 5 of Graduate Texts in Mathematics. Springer-Verlag, New York, second edition, 1998.
  • [39] S. Margolis and I. Stein. Ehresmann semigroups whose categories are EI and their representation theory. J. Algebra, 585:176–206, 2021.
  • [40] W. McCune. Prover9 and Mace4. http://www.cs.unm.edu/~mccune/prover9/.
  • [41] K. S. S. Nambooripad. Structure of regular semigroups. I. Mem. Amer. Math. Soc., 22(224):vii+119, 1979.
  • [42] 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.
  • [43] T. E. Nordahl and H. E. Scheiblich. Regular \ast-semigroups. Semigroup Forum, 16(3):369–377, 1978.
  • [44] J. Okniński. Semigroups of matrices, volume 6 of Series in Algebra. World Scientific Publishing Co., Inc., River Edge, NJ, 1998.
  • [45] A. L. T. Paterson. Groupoids, inverse semigroups, and their operator algebras, volume 170 of Progress in Mathematics. Birkhäuser Boston, Inc., Boston, MA, 1999.
  • [46] R. Penrose. A generalized inverse for matrices. Proc. Cambridge Philos. Soc., 51:406–413, 1955.
  • [47] B. Steinberg. Möbius functions and semigroup representation theory. J. Combin. Theory Ser. A, 113(5):866–881, 2006.
  • [48] B. Steinberg. Möbius functions and semigroup representation theory. II. Character formulas and multiplicities. Adv. Math., 217(4):1521–1557, 2008.
  • [49] B. Steinberg. A groupoid approach to discrete inverse semigroup algebras. Adv. Math., 223(2):689–727, 2010.
  • [50] T. Stokes. Domain and range operations in semigroups and rings. Comm. Algebra, 43(9):3979–4007, 2015.
  • [51] T. Stokes. D-semigroups and constellations. Semigroup Forum, 94(2):442–462, 2017.
  • [52] T. Stokes. Left restriction monoids from left EE-completions. J. Algebra, 608:143–185, 2022.
  • [53] T. Stokes. Ordered Ehresmann semigroups and categories. Comm. Algebra, 50(11):4805–4821, 2022.
  • [54] The MathWorks Inc., Natick, Massachusetts, United States. MATLAB. https://www.mathworks.com.
  • [55] S. Wang. An Ehresmann-Schein-Nambooripad-type theorem for a class of PP-restriction semigroups. Bull. Malays. Math. Sci. Soc., 42(2):535–568, 2019.
  • [56] S. Wang. An Ehresmann-Schein-Nambooripad theorem for locally Ehresmann PP-Ehresmann semigroups. Period. Math. Hungar., 80(1):108–137, 2020.
  • [57] S. Wang. An Ehresmann-Schein-Nambooripad type theorem for DRC-semigroups. Semigroup Forum, 104(3):731–757, 2022.
  • [58] Y. Wang. Beyond regular semigroups. Semigroup Forum, 92(2):414–448, 2016.
  • [59] M. Yamada. On the structure of fundamental regular \ast-semigroups. Studia Sci. Math. Hungar., 16(3-4):281–288, 1981.
  • [60] D. Yang, I. Dolinka, and V. Gould. Free idempotent generated semigroups and endomorphism monoids of free GG-acts. J. Algebra, 429:133–176, 2015.