Skip to the Main Content

Note:These pages make extensive use of the latest XHTML and CSS Standards. They ought to look great in any standards-compliant modern browser. Unfortunately, they will probably look horrible in older browsers, like Netscape 4.x and IE 4.x. Moreover, many posts use MathML, which is, currently only supported in Mozilla. My best suggestion (and you will thank me when surfing an ever-increasing number of sites on the web which have been crafted to use the new standards) is to upgrade to the latest version of your browser. If that's not possible, consider moving to the Standards-compliant and open-source Mozilla browser.

September 5, 2023

A Notion of ‘Location’ for Processes in Monoidal Categories

Posted by Emily Riehl

guest post by Ariel Rosenfield and Ariadne Si Suo

One can view a monoidal category as a collection of resources (objects) and processes for transforming resources (morphisms). A central idempotent in a monoidal category (C,,I)(\mathbf{C}, \otimes, I) is a way to speak about the “location” of a process taking place in C\mathbf{C}, but without referring to points, space, or distance. Under mild conditions on C\mathbf{C}, its set ZI(C)ZI(\mathbf{C}) of central idempotents can be viewed as a sort of “topological space internal to C\mathbf{C},” which lets us use some tools and intuition from topology while working in C\mathbf{C}.

Our Adjoint School group spent a lot of time the past few months with this paper by Rui Barbosa and Chris Heunen, which outlines a particularly cool application of this idea, namely the representation of “nice enough” monoidal categories (those satisfying the aforementioned mild conditions) as the global sections of certain sheaves. In this post, we’ll take a close look at central idempotents and two related notions, namely the restriction and support of morphisms in C\mathbf{C}.

Preliminaries

People are always asking, “how can we use tools from topology in categories other than the category of spaces?” Since a topology on a space is usually defined in terms of sets of points, we might, toward answering this question, try to look for some pattern of relationships underlying a topology which we can describe without talking about points, which leads to the idea of a frame. Recall that the categories of frames, locales, and complete Heyting algebras all have the same objects, namely partially ordered sets having arbitrary joins \vee and finite meets \wedge, and satisfying

(1)x iIy i iI(xy i) x \wedge \bigvee_{i \in I} y_i \leq \bigvee_{i \in I} (x \wedge y_i)

for all xx and all (maybe infinite) families {y i} iI\{y_i\}_{i \in I}. This should look familiar - if XX is a topological space and we consider the topology 𝒯\mathcal{T} on XX as a partially ordered set where

(2)uvuv, u \leq v \; \iff \; u \subseteq v,

joins are set unions, and meets are set intersections, the first two conditions above are exactly the conditions for 𝒯\mathcal{T} to constitute a topology - the third is automatically satisfied in the case where xx and y iy_i are sets. Thus it makes sense to say that a frame is a type of lattice that captures the behavior of the open sets of a space. The category of locales is opposite to the category of frames, and so we might similarly view a locale as a type of lattice that captures the behavior of the closed sets of a space. Because our definition above makes no mention of elements of XX, we can find frames (or locales, if you prefer) in categories where we can’t easily speak of “elements of objects,” or speak of them at all.

Central idempotents in monoidal categories

If all that’s the case, and a frame is just the disembodied system of relationships between the open sets of a space, which we’re viewing now as dots in some category, what should serve as our new notion of open set? What should the nameless dots be?

Central idempotents are one answer! A good analogy to keep in mind for the rest of this discussion is “central idempotent \leftrightsquigarrow open set of a topological space,” and we’ll see an example soon that justifies this analogy more rigorously. They are defined as follows: A central idempotent in a small monoidal category (C,,I)(\mathbf{C}, \otimes, I) with left unitor λ\lambda and right unitor ρ\rho is a morphism u:UIu : U \to I satisfying

  • (Idempotency 1) λ U(uid U)=ρ U(id Uu):UUU\lambda_U \circ (u \otimes \text{id}_U) = \rho_U \circ (\text{id}_U \otimes u) : U \otimes U \to U;

  • (Idempotency 2) the above morphism is invertible;

  • (Centrality) the object U carries a half-braiding satisfying:

    a string diagram on half-braiding

Okay, so that’s cool, but how should you think about these conditions? If you squint a little, you can see that \otimes is kind of the “intersection” of two objects of C\mathbf{C} - this makes sense from the perspective on C\mathbf{C} as a collection of resources, where we think of ABA \otimes B as “all instances of (AA and BB).” What we’re asking for with the idempotency conditions is thus something like “UU=UU \cap U = U for an open UU.”

Still squinting a little, you can think of the centrality condition as asking for something like “UV=VUU \cap V = V \cap U for opens UU and VV.” We’ll skip the definition of a half-braiding - all of the example categories we’ll see are braided, and it’s a fact that braidings coincide with half-braidings when they’re present, so for now, we don’t need to worry about the centrality condition being satisfied.

Examples

Let’s see what central idempotents look like in some familiar categories:

Central idempotents in sheaf categories

Our motivating example is the category Sh(X)\text{Sh}(X) of sheaves on a space XX, whose central idempotents are exactly the open sets of XX. To see this, we first note that Sh(X)\text{Sh}(X) is a cartesian category whose monoidal unit is the sheaf 1\mathbf{1} which sends every object of the locale 𝒪(X) op\mathcal{O}(X)^\text{op} to the single-element set. It turns out that in any cartesian category, an object is the domain of a central idempotent if and only if it is a subobject of the monoidal unit. We know moreover that any locale is isomorphic to the locale of subsheaves of the terminal sheaf, so in particular, 𝒪(X) op\mathcal{O}(X)^\text{op} is isomorphic as a locale to Sub(1)\text{Sub}(\mathbf{1}) - but this is just the locale ZI(Sh(X))ZI(\text{Sh}(X)). For purposes of a future example, we’ll note here that open subsets uXu \subseteq X are represented as subsheaves of 1\mathbf{1} by sheaves F uF_u of the form

(3)F u(v):={{*} vu vu.(2.1) F_u(v) := \begin{cases} \{*\} & v \subseteq u \\ \varnothing & v \nsubseteq u \end{cases}. \qquad (2.1)

Central idempotents in module categories

For another example, say RR is a commutative ring with unity 1 R1_R, and consider the monoidal category ( RMod, R,R)(_R\text{Mod}, \otimes_R, R). To characterize central idempotents in RMod_R\text{Mod}, we want to characterize all morphisms a:ARa : A \to R from RR-modules AA satisfying the idempotency and centrality conditions in our definition above. Since RMod_R\text{Mod} is braided monoidal, centrality is automatically satisfied. To check the idempotency conditions, recall that the left- and right- unitors in RMod_R\text{Mod} are respectively

(4)1 Rx1 Rx,x1 Rx1 R,(2.2) 1_R \otimes x \mapsto 1_R \cdot x, \; x \otimes 1_R \mapsto x \cdot 1_R, \qquad (2.2)

so the given equality holds for aa exactly when

(5)a(x)y=a(y)x a(x) \cdot y = a(y) \cdot x

for every x,yAx,y \in A. Recall also that isomorphisms in the category of RR-modules are bijective module homomorphisms, so it’s enough to check that (2.2) is a bijection; i.e. that for every xAx \in A, there exist x i,y iAx_i, y_i \in A (say for i=1,...,ni = 1,...,n) such that

(6)x= i=1 na(x i)y i,(2.3)x = \sum_{i=1}^n a(x_i) \cdot y_i, \qquad (2.3)

and that whenever a(x)y=0a(x) \cdot y = 0, we have xy=0x \otimes y = 0.

It turns out that central idempotents in RMod_R\text{Mod} correspond to idempotent ideals of RR: First, note that (2.3) implies that for a central idempotent a:ARa : A \to R in RMod_R\text{Mod}, a(A)a(A) is always an idempotent ideal of RR (to see this, use the homomorphism property of aa!). Conversely, if IRI \subseteq R is an idempotent ideal, the inclusion IRI \hookrightarrow R is a central idempotent as defined above.

If you’re an algebraist, you might be wondering if there’s some connection between central idempotent elements in a ring and the categorical notion. When RR is assumed to be sufficiently nice, we can say that idempotent elements of RR correspond to central idempotents: For example, if RR is Noetherian, a consequence of Nakayama’s lemma from commutative ring theory is that any idempotent ideal IRI \subseteq R has the form I=eRI = e R for some idempotent element eRe \in R. As above, the inclusion a:eRRa : e R \hookrightarrow R of any such ideal into RR is a central idempotent in RMod_R\text{Mod}.

Restriction and support of morphisms

With the analogy “central idempotent \leftrightsquigarrow open set” in mind, we’d hope that whenever a morphism ff in our category is “defined” on a central idempotent uu, we can somehow restrict it to uu, in a similar sense to how a sheaf on a space XX yields a sheaf on any open set uXu \subseteq X, viewed as a subspace. In a general monoidal category, we say that a morphism f:ABf: A \to B restricts to a central idempotent s:SIs : S \to I when it factors through λ B(sid B)\lambda_B \circ (s \otimes \text{id}_B), as in the commuting diagram

support

We’d also hope to be able to find the “smallest” central idempotent to which a morphism f:ABf: A \to B restricts, and for that we need to talk about support! Recall that the support of a function ff on a manifold MM taking values in a vector space VV is the closed set

(7) MKS,K closed X, \bigcap_{M \supset K \supset S, \; K \text{ closed }} X,

where S={xM:f(x)0}S = \{x \in M : f(x) \neq 0\}. We often regard this object as the “smallest” closed subset of XX containing SS.

Notice that being able to call Supp(f)\text{Supp}(f) a closed subset of our topological space MM requires that the intersection of an arbitrary family of closed sets is itself closed - a property indeed possessed by the locale of closed sets of a topological space! We can realize this notion in the generality of a monoidal category C\mathbf{C} for which ZI(C)ZI(\mathbf{C}) is a frame using central idempotents: For a morphism ff in C\mathbf{C}, define

(8)Supp(f)={sZI(C):f restricts to s}. \text{Supp}(f) = \bigwedge \{s \in ZI(\mathbf{C}) : f\, \text{ restricts to }\, s\}.

In this setting, the correct analogue of “a closed set containing S={xM:f(x)0}S = \{x \in M : f(x) \neq 0\}” is a certain morphism called a support datum. We’ll skip the precise definition, but note that for our new definition of support to capture the property of being the “smallest” closed set, we’d want Supp(f)\text{Supp}(f) to be universal amongst all support data, and it turns out that it is - any support datum factors through Supp(f)\text{Supp}(f) uniquely.

Returning to our example of Sh(X)\text{Sh}(X), and considering an open set uXu \subseteq X, observe that a morphism f:ABf : A \Rightarrow B of sheaves satisfies the diagram (3.1) exactly when there is a natural transformation

(9)h:AF u×B, h: A \Rightarrow F_u \times B,

where F uF_u is the sheaf defined in (2.1), which makes (3.1) commute. Such an hh exists exactly when

(10)A(v)= for every vu: A(v) = \varnothing \; \text{ for every } \; v \nsubseteq u: \qquad

If this condition holds, define the vv-component of hh by

(11)h v={ A(v)= x(*,f v(x)) A(v). h_v = \begin{cases} \varnothing \to \varnothing & A(v) = \varnothing \\ x \mapsto (*,f_v(x)) & A(v) \neq \varnothing \end{cases}.

No such hh exists if A(v)A(v) \neq \varnothing for some vuv \nsubseteq u, since then F u(v)=F_u(v) = \varnothing, and there can be no map of sets A(v)F u(v)×B(v)=A(v) \to F_u(v) \times B(v) = \varnothing. We see that the support of ff is just

(12)Supp(f)={uX:A(u)= for every v¬u}. \text{Supp}(f) = \bigwedge \{u \subseteq X : A(u) = \varnothing \,\text{ for every }\, v \not \subset u\}.

Lattice-theoretic properties of ZI(C)ZI(\mathbf{C}); sheaf representation theorems

If central idempotents are supposed to be like open sets, do they satisfy the same system of relationships satisfied by the open sets of a topological space? In other words, do they constitute the elements of a frame? The answer here is “only sometimes.”

We write ZI(C)ZI(\mathbf{C}) for the set of central idempotents of C\mathbf{C}, recalling that we take C\mathbf{C} to be small. We identify central idempotents u:UIu: U \to I and v:VIv: V \to I when u=vmu = v \circ m for an isomorphism m:UVm: U \to V that respects the half-braiding.

ZI(C)ZI(\mathbf{C}) can be given the structure of a lattice. For central idempotents u:UIu: U \to I and v:VIv: V \to I in a monoidal category C\mathbf{C}, we say uvu \leq v if u=vmu = v \circ m for a necessarily unique m:UVm: U \to V that respects the half-braiding. This condition is equivalent to either of the following: - Uv:UVUIU \otimes v: U \otimes V \to U \otimes I is invertible. - vU:VUIUv \otimes U: V \otimes U \to I \otimes U is invertible.

Under mild conditions on C\mathbf{C}, ZI(C)ZI(\mathbf{C}) has nice lattice-theoretic properties. We’ll now define these conditions on C\mathbf{C}, starting from the weakest one.

A monoidal category is stiff if

commutative diagram and string diagram on stiffness

is a pullback for all objects AA and central idempotents uu and vv. Roughly, this says that pullbacks and monoidal products are two ways to take “intersections” of central idempotents, and that they coincide. Stiffness is a very weak condition - almost every monoidal category is stiff (except probably artificially constructed counterexamples)!

A monoidal category has finite universal joins / \vee-joins of central idempotents if - it has an initial object 00 satisfying A00A \otimes 0 \simeq 0 for all objects AA - ZI(C)ZI(\mathbf{C}) has binary joins such that

commutative diagram and string diagram on finite universal joins

is both a pullback and pushout for all objects AA and central idempotents uu and vv.

This roughly says that the functor AA \otimes - preserves joins. If C\mathbf{C} has finite universal joins of central idempotents, then ZI(C)ZI(\mathbf{C}) is a distributive lattice with least element 00.

A monoidal category has universal joins / \bigvee-joins of central idempotents when ZI(C)ZI(\mathbf{C}) has all joins such that the cocone

string diagram

is a colimit of the diagram with morphisms

string diagram

if u iu ju_i \leq u_j for any set {u i}\{ u_i \} of central idempotents satisfying {u i}={u iu j}\{ u_i \} = \{ u_i \wedge u_j\}. This roughly says that the functor AA \otimes - preserves suprema. If C\mathbf{C} has universal joins of central idempotents, then ZI(C)ZI(\mathbf{C}) is a frame.

It turns out that any stiff monoidal category can be freely completed with universal (finite) joins of central idempotents.

We just need a few more definitions before stating two of our aformentioned representation theorems. We say a partially ordered set is \vee-local if umu \vee m implies u=1u = 1 or v=1v = 1. A monoidal category C\mathbf{C} is \vee-local if ZI(C)ZI(\mathbf{C}) is \vee-local.

We say a partially ordered set is \bigvee-local if u i=1\bigvee u_i = 1 implies that there exists an ii with u i=1u_i = 1. A monoidal category C\mathbf{C} is \bigvee-local if ZI(C)ZI(\mathbf{C}) is \bigvee-local.

As promised, here are the representation theorems:

  • Theorem. Any small monoidal category with universal \vee-joins of central idempotents is monoidally equivalent to the category of global sections of a sheaf of \vee-local monoidal categories.

  • Theorem. Any small monoidal category C\mathbf{C} in which ZI(C)ZI(\mathbf{C}) forms a spatial frame is monoidally equivalent to the category of global sections of a sheaf of \bigvee-local monoidal categories.

Future work

In a way, a morphism ff restricts to a central idempotent ss says that ff “only lives over”ss. We can think of the support of a morphism as telling us precisely where the morphism lives.

We can extend this intuition to say that an object AA only lives over a central idempotent ss if id A\text{id}_A restricts to ss. This is in fact equivalent to sAAs \otimes A \cong A. We can define the support of an object AA as the support of id A\text{id}_A, which tells us the precise location of AA.

This gives us a notion of space in a monoidal category. If we think of morphisms in a category as processes, and compositon of morphisms as processes happening one after another, then the monoidal product captures the idea of “processes happening simultaneously.” This is the natural notion of time in a monoidal category.

It seems that monoidal categories provide the perfect setting for concurrency. A classic example is the producer-consumer problem. The producer has a process of making cookies, and the consumer has a process of eating cookies. The consumer can be eating the last batch while the producer is making new ones.

Monoidal categories are perfect for modelling concurrent processes because of their natural notion of simultaneity. With the help of central idempotents and supports, we can now also talk about the space in which the processes take place. What if the producer is making cookies at the bakery, and the consumer is attending the ACT (i.e. processes have different supports)?

This motivated our research problem: What structure can we add to a monoidal category so that concurrent processes supported on disjoint central idempotents can exchange resources?

Posted at September 5, 2023 2:14 PM UTC

TrackBack URL for this Entry:   https://golem.ph.utexas.edu/cgi-bin/MT-3.0/dxy-tb.fcgi/3494

2 Comments & 0 Trackbacks

Re: A Notion of ‘Location’ for Processes in Monoidal Categories

For example, if RR is Noetherian, a consequence of Nakayama’s lemma from commutative ring theory is that any idempotent ideal IRI \subseteq R has the form I=eRI = e R for some idempotent element eRe \in R.

Hmm! By coincidence I just needed this and blogged about it here last week. It’s Lemma 17: I used it to show that if AA is an finite-dimensional algebra over a field where the short exact sequence

0IAAmA0 0 \to I \hookrightarrow A \otimes A \stackrel{m}{\to} A \to 0

splits as a sequence of A,AA,A-bimodules, then the ideal IAAI \subseteq A \otimes A has the form eRe R for some idempotent ee. This would also work when AA is a Noetherian commutative ring. (I guess you’re assuming RR is commutative.)

I’m too scared to try to connect why I cared about this with why you care about it, but they could be related.

Posted by: John Baez on September 6, 2023 10:40 AM | Permalink | Reply to this

Re: A Notion of ‘Location’ for Processes in Monoidal Categories

Oh cool! I’ll have to go back and read the post. Yes, RR is assumed to be commutative - this is just because otherwise the category of RR-modules isn’t necessarily monoidal.

Posted by: Ari on September 6, 2023 11:20 PM | Permalink | Reply to this

Post a New Comment