module 1Lab.intro where

Introduction🔗

This page aims to present a first introduction to cubical type theory, from the perspective of a mathematician who has heard about type theory but has no previous familiarity with it. Specifically, the kind of mathematician that we are appealing to is one who is familiar with some of the ideas in category theory and homotopy theory — however, the text also presents the concepts syntactically, in a way that can be read without any prior mathematical knowledge.

Note that whenever code appears, all of the identifiers are hyperlinked to their definitions, which are embedded in articles that describe the concepts involved. For instance, in the code block below, you can click the symbol to be taken to the page on path types.

_ : 2 + 2  4
_ = refl

For some operations, we use the built-in Agda definitions, which do not have explanations attached. This is the case for the + operation above. However, in any case, those definitions are either built-ins (in which case only the type matters), or defined as functions, in which case the implementation is visible. The most important built-ins for Cubical Agda, and those most likely to lead you to a built-in Agda page, have their own explanations, linked below:

  • The 1Lab.Path page explains the primitives I, i0, i1, _∧_, _∨_, ~, PathP, Partial, _[_↦_], and hcomp.

With that out of the way, the first thing to do is the aforementioned refresher on type theory. If you’re already familiar with type theory, feel free to skip to What’s next?. If you’re not familiar with some of the connections that HoTT has to other areas of mathematics, like higher category theory and topology, I recommend that you read the type theory section anyway!

Terminology On the term “homotopy type theory”.

The text below refers to the foundational system that it is presenting as “homotopy type theory”, while it would be technically more correct to refer to it as a homotopy type theory. HoTT is not a specific type theory, but a family of type theories that put emphasis on the homotopy-theoretic interpretation of type theory. In particular, a homotopy type theory is expected to validate univalence and have some class of higher inductive types.

Below, we describe cubical type theory, which is one particular homotopy type theory. In cubical type theory, we interpret types as the Kan complices in the category of De Morgan cubical sets, This interpretation gives rise to a model structure on the category of cubical sets, referred to as a “cubical-type model structure”.

Type theory🔗

\ Warning

The text below is subject to change! It’s still very much a document in flux. In particular, the type theory section doesn’t talk about induction, which is something that will be expanded on in the future.

Type theory is a foundational system for mathematics which, in contrast to a set-theoretic foundation like ZFC, formalises mathematical constructions rather than mathematical proofs. That is, instead of specifying which logical deductions are valid, and then giving a set of axioms which characterise the behaviour of mathematical objects, a type-theoretic foundation specifies directly which mathematical constructions are valid.

Formally speaking, it is impossible to construct objects in set theoretic foundations; Rather, one can, by applying the deduction rules of the logical system and appealing to the axioms of the set theory, prove that an object with the desired properties exists. As an example, if we have a way of referring to two objects and we can prove that there exists a set representing the ordered pair using the axiom of pairing. Writing quasi-formally, we could even say that this set is constructed using the expression but this is only a notational convenience.

In contrast, a type-theoretic foundational system, such as HoTT (what we use here), specifies directly how to construct mathematical objects, and proving a theorem becomes a special case of constructing an object. To specify what constructions are well-behaved, we sort the mathematical objects into boxes called types. Just like in set theory, the word “set” is a primitive notion, in type theory, the word “type” is a primitive notion. However, we can give an intuitive explanation: “A type is a thing that things can be”. For instance, things can be natural numbers, but things can not be “two” — hence, the natural numbers are a type, but two is not. Formally, a type is specified by giving:

  • A formation rule, which tells us which expressions denote valid types at all. For instance, the formation rule for product types tells us that denotes a well-formed type, but “” does not. The formation rules also have the duty of preventing us from running into size issues like Russell’s paradox, by constraining the use of universes.

  • A handful of introduction rules, which tell us how to piece previously-constructed objects together to form an object of a different type. Sticking with products as an example, if we have previously constructed objects called and we can construct an ordered pair

    Since the introduction rules are inseparable from their types, it is impossible to construct an object without knowing in which type it lives. If we used the rule which introduces ordered pairs, we know for sure that we built an object of the product type. To specify that an object lives in a certain type, we use the notation — which can be read as “ inhabits ”, or “ is a point of ”, or “ has type ”.

    The notation for stating inhabitation agrees with mathematical parlance in when both make sense. For example, a function from the reals to the reals is commonly introduced with the notation In type theory, this is made formal as a typing declaration: is a point of the type of functions hence we know that the rules of function types can be used with However, when introducing a more complex object, we’re forced to break into informality and conventional shorthand: “Let G be a group”. In HoTT, we can directly write 1, which tells us that the rules for the type of groups apply to

  • A handful of elimination rules, which say how inhabitants of that type can be manipulated. Keeping with the running example, the product type has elimination rules corresponding to projecting either the first or second component of a pair.

    Furthermore, each elimination rule comes with specified behaviour for when it “interacts” with the introduction rules, a process which we generally call reduction. Repeatedly applying these “interactions” tells us how a given construction can have computational behaviour — by performing a delicate dance of elimination and introduction, we can go from to

Now that we have an intuitive motivation for the notions of type theory, we can more precisely discuss the structure of formal type theory and how it differs from formal (material) set theory. The first (and perhaps most important) difference is that formal set theory is a “two layered” system, but type theory only has “one layer”. In (for example) ZFC, we have the deductive framework of first-order logic (the “first layer”), and, using the language of that logical system, we assert the axioms of set theory. This remains valid if we switch set theories: For instance, IZF would substitute first-order logic with first-order intuitionistic logic, and remove the assertion that the axiom of choice holds.

In contrast, type theory only has one layer: The layer of types. What in formal set theory would be a deduction in FOL, in type theory becomes the construction of an inhabitant of a certain type, where propositions are identified with certain types as in the table below. Note that up to a first approximation, we can read each type former as denoting a specific set, or as denoting a specific space.

Types Logic Sets Spaces
proposition set space
proof element point
predicate family of sets fibration
conditional proof family of elements section
Cartesian product product space
disjoint union coproduct space
function set function space
n-ary product space of sections
n-ary disjoint union total space
equality diagonal path space

This segues nicely into another difference between type theory and set theory, which concerns the setup of their deductive systems. A deductive system is a specification for how to derive judgements using rules. We can think of the deductive system as a game of sorts, where the judgements are to be thought of as board positions, and the rules specify the valid moves.

In the deductive system of first-order logic, there is only one kind of judgement, stating that a proposition has a proof. If is a proposition, then the judgement “ has a proof” is written “”. Note that the judgement is a part of the metatheory, and the proposition is a part of the theory itself. That is, is an external notion, and is an internal notion.

In type theory, the basic judgement is inhabitation, written If is a type that denotes a proposition, then this is analogous to the judgement of first-order logic, and we refer to as a witness of or simply as a proof of Even better, if is a purely logical proposition such that holds without any appeal to the principle of excluded middle, then there is a term for which holds in type theory.

When is being thought of as a collection of things, we might imagine that is analogous to However, this is not the case: While is a proposition (an internal statement), is a judgement, an external statement. That is, we can not use with the propositional connectives mentioned in the table above — we can not make statements of the form “if it is not the case that then ”. This is the second difference we will mention between set theory and type theory: The proposition is an internal relation between previously-constructed terms and but the external judgement is indivisible, and it does not make sense to talk about without also bringing up that it’s in

The third and final difference is the treatment of equality, which is arguably also the most important part of homotopy type theory. In set theory, given sets and we have a proposition which is given meaning by the axiom of extensionality: whenever and behave the same with respect to Correspondingly, in type theory, we have an equality type, written (where giving us the internal notion of equality. But we also have an external notion of equality, which encodes the aforementioned interactions between introduction and elimination rules. This is written and it’s meant to indicate that and are the same by definition — hence we call it definitional equality.

Since definitional equality is a judgement, it’s also not subject to the internal propositional connectives — we can not prove, disprove, assume or negate it when working inside type theory, for it does not make sense to say “if and are equal by definition, then […]”. Correspondingly, in the Agda formalisation, definitional equality is invisible, since it’s represented by the computation rules of the type theory.

In the rest of this section, we will recall the presentation of the most ubiquitous types. The idea is that if you’ve encountered type theory before, then skimming the explanation here should help snap your mind back into focus, while if this is your first time interacting with it, then the content here should be enough for you to understand the rest of the 1lab development.

Functions🔗

If we have two previously-constructed types and we can form the type of functions from to , written Often, functions will also be referred to as maps. A function is, intuitively, a rule prescribing how to obtain an inhabitant of given an inhabitant of In type theory, this is not only an intuition, but rather a definition. This is in contrast with set theory, where functions are defined to be relations satisfying a certain property.

Given a function we can apply it to an inhabitant resulting in an inhabitant of the codomain called the value of at . For clarity, we sometimes write for the value of at but in Agda, the parentheses are not necessary.

The most general form for constructing an inhabitant of is called lambda abstraction, and it looks like though shorthands exist. In this construction, the subexpression denotes the body of a function definition, and is the parameter. You can, and should, think of the expression as being the same as the notation often used for specifying a map. Indeed, Agda supports definition of functions using a more traditional, “rule-like” notation, as shorthand for using abstraction. For instance, the following definitions of inhabitants f and g of are definitionally the same:

In general, we refer to a way of constructing an inhabitant of a type as an introduction rule, and to a way of consuming an inhabitant of a type as an elimination rule. Concisely, the introduction rule for is abstraction, and the elimination rule is function application. A general principle of type theory is that whenever an elimination rule “meets” an introduction rule, we must describe how they interact.

In the case of meeting function application, this interaction is called 2, and it tells us how to compute a function application. This is the usual rule for applying functions defined by rules: The value of is but with replacing whenever appears in Using the notation that you’d find in a typical3 type theory paper, we can concisely state this as:

In addition, function types enjoy a definitional uniqueness rule, which says “any function is a expression”. Symbolically, this means that if we have the terms and are equal, definitionally. This is called and when applied right-to-left as a reduction rule,

Functions of multiple arguments are defined by iterating the unary function type. Since functions are types themselves, we can define functions whose codomain is another function — where we abbreviate by By the uniqueness rule for function types, we have that any inhabitant of this type is going to be of the form which we abbreviate by In Agda, we can write this as λ x y → ..., or by mentioning multiple variables in the left-hand side of a rule, as shown below.

  f : Nat  Nat  Nat
  f x y = x + 3 * y

Universes🔗

Instead of jumping right into the syntactic definition (and motivation) for universes, I’m going to take a longer route, through topos theory and eventually higher topos theory, which gives a meaning explanation for the idea of universes by generalising from a commonly-understood idea: the correspondence between predicates and subsets. Initially, we work entirely in the category of sets, assuming excluded middle for simplicity. Then we pass to an arbitrary elementary topos, and finally to an arbitrary higher topos. If you don’t want to read about categories, click here

First, let’s talk about subsets. I swear I’m going somewhere with this, so bear with me. We know that subsets correspond to predicates, but what even is a subset?

  • If we regard sets as material collections of objects, equipped with a membership relation then is a subset of if every element of is also an element of However, this notion does not carry over directly to structural set theory (where sets are essentially opaque objects of a category satisfying some properties), or indeed to type theory (where there is no membership relation) — but we can salvage it.

    Instead of thinking of a subset directly as a subcollection, we define a subset to be an injective function Note that the emphasis is on the function and not on the index set since we can have many distinct injective functions with the same domains and codomains (exercise: give two distinct subsets of with the same domain4).

    To see that this definition is a generalisation of the material definition in terms of we note that any injective function can be restricted to a bijective function by swapping the original codomain for the image of the function. Since the image is a subset of the codomain, we can interpret any injective as expressing a particular way that can be regarded as a subset of

  • Alternatively, we can think of a subset as being a predicate on the larger set which holds of a given element when is in the subset Rather than directly thinking of a predicate as a formula with a free variable, we think of it as a characteristic function from which we can recover the subset as

    This definition works as-is in structural set theory (where the existence of solution sets for equations — called “equalisers” — is assumed), but in a more general setting (one without excluded middle) like an arbitrary elementary topos, or indeed type theory, we would replace the 2-point set by an object of propositions, normally denoted with

These two definitions are both appealing in their own right — the “subsets are injective functions” can be generalised to arbitrary categories, by replacing the notion of “injective function” with “monomorphism”, while “subsets are predicates” is a very clean notion conceptually. Fortunately, it is well-known that these notions coincide; For the purposes of eventually generalising to universes, I’d like to re-explain this correspondence, in a very categorical way.

Consider some sets and an arbitrary function between them. We define the fibre of over to be the set that is, the set of points of the domain that maps to — in the context of set theory, this would be referred to as the preimage of but we’re about to generalise beyond sets, so the more “topological” terminology will be useful. We can often characterise properties of by looking at all of its fibres. What will be useful here is the observation that the fibres of an injective function have at most one element.

More precisely, let be injective, and fix a point so we can consider the fibre over y. If we have two inhabitants then we know (by the definition of fibre) that hence by injectivity of we have — if has any points at all, then it has exactly one. It turns out that having at most one point in every fibre precisely characterises the injective functions: Suppose Then we have two inhabitants of and but since we assumed that has at most one inhabitant for any we have so is injective.

We call a set (type) with at most one element (at most one point) a subsingleton, since it is a subset of the singleton set (singleton type), or a proposition, since a choice of point doesn’t give you any more information than merely telling you “ is inhabited” would. Since there is a 1-to-1 correspondence between subsingletons and propositions “ is inhabited”, the assignment of fibres defines a predicate where we have implicitly translated between a subsingleton and its corresponding proposition.5

Conversely, if we have such an assignment we can make a map into with subsingleton fibres by carefully choosing the domain. We take the domain to be the disjoint union of the family that is, the set The elements of this set can be described, moving closer to type-theoretic language, as pairs where and is a witness that the subset predicate holds of

I claim: the map which takes has subsingleton fibres. This is because an element of is described as a pair and since it’s in the fibre over we have Since (by assumption) takes values in subsingletons, concluding that is enough to establish that any two and must be equal.

The point of the entire preceding discussion is that the notions of “injections into ” and “maps ” are equivalent, and, more importantly, this is enough to characterise the object 6.

Let me break up this paragraph and re-emphasise, since it’ll be the key idea later: A type of propositions is precisely a classifier for maps with propositional fibres.

Since the subobjects of are defined to be injections and maps represent these subobjects (in a coherent way), we call a subobject classifier, following in the tradition of “classifying objects”. A subobject classifier is great to have, because it means you can talk about logical propositions — and, in the presence of functions, predicates — as if they were points of a type.

More on general classifying objects.

A classifying object for some data is an object — normally written — such that maps correspond to data “over” in the sense that the data admits some map into Above, we were discussing the subobject classifier that one can find in an elementary topos; Maps (the object of propositions) correspond to maps into with “propositional data”.

To find another example in category theory, we pass to higher category theory, i.e.  theory for high values of such as As a 1-dimensional structure, we can consider the data of a category lying over another category to be a functor which has certain lifting properties - some sort of categorical fibration.

One would like to find a correspondence between these “categories over categories” and some data that we already know how to conceptualise; It turns out that Grothendieck fibrations over — i.e., the functors satisfying said “certain lifting properties” — correspond to weak 2-functors into

Conversely, this equivalence lets us think of a family of categories indexed by — a 2-dimensional concept — corresponds precisely to 1-dimensional data (a special class of functors between ordinary categories). Furthermore, the equivalence is induced by a construction on (weak 2-)functors — the Grothendieck construction, written — that exactly mirrors the disjoint union that we used above!

To summarise:

  • A classifying object is an object where maps corresponds to data over (“pointing towards”)

  • The kind of classifying objects that we care about here are ubiquitous in all levels of theory, but especially in theory, where some behave like universes of a given truncation level.

  • Since the objects in a act like the “most complex” classifier that a can have is for internal the of internal and this all works better in the case where we’re actually talking about

However, we often want to talk about sets and families of sets as if they were points of a type, instead of just subsets. We could pass to a higher category that has a discrete object classifier — this would be a “type of sets” —, but then the same infelicity would be repeated one dimension higher: we’d have a classifier for “sets”, but our other objects would look like “categories”, and we wouldn’t be able to classify those!

Take a moment to convince yourself that the interpretation of a discrete object classifier as a type of sets makes sense.

Note that we defined “proposition” to mean “at most one element”. We can generalise this to a notion which applies to categories by defining a discrete category to be one that has propositional hom-sets, i.e. one in which the only morphisms are identities. Note that a discrete category is automatically a discrete groupoid, since the identity map is an isomorphism.

A discrete object classifier (say would then be a classifier for functors which have discrete groupoids as fibres — the fibres have no categorical information, being nothing more than a set. Given any set we’d have a name for it as the interpretation of the unique map as a map The name was chosen because in the the discrete object classifier is exactly the category of sets.

Thus we might hope to find, in addition to a subobject classifier, we could have a general object classifier, an object such that maps correspond to arbitrary maps Let’s specify this notion better using the language of higher category theory:

A subobject classifier is a representing object for the functor which takes an object to its poset of subobjects. Since is a subcategory of restricted to the objects which are (-1)-truncated, an object classifier would be a representing object for the functor which sends an object to the category of arbitrary families”, the slice category over If capturing the directed structure isn’t possible, then, at the very least, it should represent — the core of so that the isomorphisms between families are accurately represented.

However, there is an issue here: there is, a priori, an entire category but only a set Objects in a category can have non-trivial isomorphisms, but there’s no such notion between elements in a set! Let’s suppose that we had such an equivalence, in the context setting aside the size issues for now: we have an object equipped with an isomorphism There’s an element classifying the type of booleans, which should have a non-trivial identification corresponding to the involution on However, this is not possible, since a set (by definition) only has trivial identifications.

One might attempt to remedy the issue by looking for an object classifier in a bicategory, where would be a genuine category. However, this didn’t actually fix anything, since we only shifted the issue one level up: we now have an entire bicategory which can’t be classified by the mere category of maps

The issue does eventually resolve itself, if we’re very persistent: if we have an where can be a genuine then it’s actually perfectly possible for there to be an object with an isomorphism — well, at least as long as we restrict ourselves to We’ve finally found a setting where object classifiers can exist:

Again, the importance of object classifiers is that they let us talk about arbitrary objects as if they were points of an object Specifically, any object has a name, a map which represents the unique map To make this connection clearer, let’s pass back to a syntactic presentation of type theory, to see what universes look like “from the inside”.

Universes, internally🔗

Inside type theory, object classifiers present themselves as types which contain types, which we call universes. Every type is contained in some universe, but it is not the case that there is a universe containing all types; In fact, if we did have some magical universe we could reproduce Russell’s paradox, as is done here.

The solution to this is the usual stratification of objects into “sizes”, with the collection of all “small” objects being “large”, and the collection of all “large” objects being “huge”, etc. On the semantic side, a has a sequence of object classifiers for maps with fibres, where is a regular cardinal; Syntactically, this is reflected much more simply as having a tower of universes with So, in Agda, we have:

  _ : Type
  _ = Nat

  _ : Type₁
  _ = Type

In Agda, we do not have automatically that a “small” object can be considered a “large” object. When this is the case, we say that universes are cumulative. Instead, we have an explicit Lifting construction that lets us treat types as being in higher universes:

  _ : Type₂
  _ = Lift (lsuc (lsuc lzero)) Type

To avoid having Lift everywhere, universes are indexed by a (small) type of countable ordinals, called Level, and constructions can be parametrised by the level(s) of types they deal with. The collection of all types is which is itself a type in the next universe up, etc. Since levels are themselves inhabitants of a type, we do in fact have a definable function To avoid further Russellian paradoxes, functions out of Level must be very big, so they live in the “first infinite universe”, There is another hierarchy of infinite universes, with inhabiting To avoid having even more towers of universes, you can’t quantify over the indices in To summarize, we have:

To represent a collection of types varying over an value of we use a function type into a universe, called a type family for convenience: A type family over is a function for some choice of level An example of a type family are the finite types, regarded as a family — where is a type with elements. Type families are also used to model type constructors, which are familiar to programmers as being the generic types. And finally, if we have a type we can consider the constant family at defined to be the function

To cap off the correspondence between universes and object classifiers, thus of type families and maps, we have a theorem stating that the space of maps into is equivalent to the space of type families indexed by As a reminder, you can click on the name Fibration-equiv to be taken to the page where it is defined and explained.

  _ :  {B : Type}  (Σ[ E  Type ] (E  B))  (B  Type)
  _ = Fibration-equiv

Universes and size issues🔗

Perhaps one of the most famous paradoxes in the history of formal logic, Russell’s paradox reminds us that, for most reasonable kinds of “collection” — be they sets, types, or categories — considering the collection of all collections that do not contain themselves is a road that leads to madness. The standard way of getting around these issues, at least in set-theoretic foundations as applied to category theory, is to refer to such problematic collections as “classes”, and to only axiomatise the sets which are not too big.

In the branches of mathematics that Alexander Grothendieck influenced, a more common approach is to work instead with Grothendieck universes: A Grothendieck universe is a set that forms a transitive model of ZF, closed under indexed products. The classification of models of ZF implies that any Grothendieck universe is equivalent to a segment of the von Neumann universe where is a strongly inaccessible cardinal — essentially, an upper bound on the size of the sets in

In traditional mathematical writing, it is common to entirely ignore size issues, save perhaps for an offhand mention of Grothendieck universes (or strongly inaccessible cardinals) within the first few sections. Working in a proof assistant forces us to be honest about the size of our constructions: Correspondingly, we try to be precise in our prose as well. As mentioned above, Agda universes are stratified into a pair of hierarchies and where we’re using to stand for a variable of Level type. The Level is a built-in type which contains a constant representing is closed under taking successors, and is closed under taking binary maxima.

We refer to structured type (e.g. a category, or a group) as when its underlying type inhabits the universe; For the specific case of categories, we also use the terminology locally to refer to the universe in which the family lands. We use subscript Greek letters in prose to index our structures by their universe level. For example, is the ( locally category of sets.

Small (hah!) note on small categories.

For those familiar with the notions of internal and enriched categories, we might rephrase the classical definitions of and locally categories as follows:

  • A category is if it is an internal category in

  • A category is locally if it is enriched over the Cartesian monoidal category

Because every type is contained in some universe, we note that every category that appears in our development is locally for some universe level In particular, we have no “type of categories”, but instead a type of “ locally categories”.

\ Warning

Note that our use of the term “” is nonstandard. In set-theoretic foundations, where the only objects are sets anyway, this means “category internal to ”, as mentioned in the infobox above. In the 1Lab, the objects we manipulate are higher groupoids rather than sets, and so very few categories will be internal to a category of sets. Instead, when we describe a category is we mean that the type of objects of is an inhabitant of the universe and that the family is valued in Our shorthand for the traditional notion is a “strict ” category.

Our policy is to keep the use of universes “as precise as possible”: Definitions are polymorphic over as many levels as they can possibly be, and they are always placed in the smallest universe in which they fit. This is, however, an iterative process: For example, our definition of “sheaf topos” was originally parametrised over 5 levels, but after formalising more desirable properties of topoi, we have found that the definition only really mentions two levels.

Interlude: Basics of paths🔗

Since the treatment of paths is the most important aspect of homotopy type theory, rest assured that we’ll talk about it in more detail later. However, before discussing the dependent sum of a type family, we must discuss the fundamentals of paths, so that the categorical/homotopical connections can be made clear. Before we even get started, though, there is something that needs to be made very clear:

Paths are not equality!

The only conceptualisation of paths that is accurate in all cases is that types are and hence path spaces are the Hom- given by the structure of each type. However, this understanding is very bulky, and doesn’t actually help understanding paths all that much — so we present a bunch of other analogies, with the caveat that they only hold for some types, or are only one of the many possible interpretations of paths in those types.

The ideas presented below are grouped in blue <details> elements, but not because they aren’t important; You should read all of them.

Idea #1: Some types behave like sets, providing an embedding of “classical” equality behaviour into homotopy type theory. While these types may not be as “exciting” as others, they aren’t any less important!

Contradicting the big bold warning up there, in a very restricted class of types, paths do behave like the equality relation! We call these types sets, and their defining property is that they “only have identity paths” — hence they could also be called “discrete types”, by analogy with “discrete category”, but this terminology is not employed because “discrete” already has a distinct type-theoretic meaning.7

Types in this class include the natural numbers, the integers, the rationals and the reals; Also included is every finite type, and the power set of any type at all. The type of functions where is a set, is again a set. We can generalise this and say that “most” type formers preserve set-ness.

Note that the meaning of “is a set”/“is not a set” in HoTT is distinct from how that sentence is used in classical mathematics, where a collection is a set if it’s definable in the language of whatever foundational system you’re using. For example, in ZFC, the collection of all sets is not a set — because if it were, we could reproduce Russell’s paradox.

In HoTT, being a set has to do with spatial structure: a set has only trivial paths. Examples of non-sets would then be given by types that have interesting paths, like the circle, which has a non-trivial loop. Coincidentally, the type of all sets is not a set, since its spatial structure is given by the isomorphisms of sets — and sets can have non-trivial automorphism groups.

Idea #2: Some types are best understood as representing spaces, with intrinsic homotopic information, which we can investigate using type-theoretic methods. This interpretation is the core idea of “synthetic homotopy theory”, a whole field of maths!

We conceptualise some types (like the circle mentioned above) as being representatives for spaces, though this does not mean that they have any intrinsically-detectable topology in the sense of “open sets”-topological spaces. Instead, we think of these types as being an incarnation of spaces à la homotopy theory, with structure similar to what you’d find there: Homology groups, cohomology groups, but most easily seen are the homotopy groups. In these types, type-theoretic paths quack like topological paths.

Indeed, any type in homotopy type theory has an associated sequence of homotopy groups, and these behave precisely like the homotopy groups of classical algebraic topology — with the difference that the way we compute them in HoTT has a distinctly more “typey” flavour.

This interpretation of types as spaces leads us to look at the statement “there is a type with ” and rightly call “the circle”, even though a topologist might object that our circle is not any subset of Similarly, we might take the integer and apply the isomorphism above, to get a path in which “walks clockwise around the circle twice”.

Idea #3: HoTT provides a setting for doing category theory where all constructions are intrinsically isomorphism-invariant. Under this light, types provide the underlying groupoids of categories.

When doing category theory in HoTT, we have some types that naturally arise as the type of objects in a category, such as the type of groups relative to a universe. In category theory done set-theoretically, the collection of objects in has two different notions of sameness: The set-theoretic equality with behaviour given by the axiom of extensionality and divorced from the categorical structure, and the class of isomorphisms of which naturally has to do with the categorical structure.

A big part of doing category theory in set-theoretic foundations is learning in which situations we should talk about the set-theoretic equality of objects, and when we should talk about isomorphism — what you end up discovering is that constructions that talk about equality of objects should be done very sparingly, so much so that they’re colloquially called “evil”.

Conversely, the possibility of talking about equality of objects at all means that we have to spend effort verifying that constructions appropriately respect isomorphisms, though of course this verification can be (and often ends up being) handwaved.

In HoTT, we can do better: We add a condition to the definition of category that says the paths of the type of objects must correspond to the categorical isomorphisms — which we call “univalence”. While this condition might initially seem hard to fulfill, it’s surprisingly common to come by univalent categories, since most constructions on categories preserve univalence! For instance:

  • If and are univalent categories, then so is the product and the disjoint union The initial and terminal categories are trivially univalent.

  • If is a univalent category, then so is the slice category and the coslice More general comma categories are also univalent.

  • If is a univalent category, then for any so is the functor category This implies that a very large class of categories is univalent, e.g. the models of any Lawvere theory in a univalent category.

    Further, since is a univalent category, the Yoneda lemma implies that every can be embedded into a univalent category - the functor category If we restrict to the subcategory of representable presheaves, we have that every admits a full, faithful and essentially surjective functor into a univalent category, called the Rezk completion of If was already univalent, this functor is an equivalence.

In addition to the benefit of univalent categories that will be presented outside the fold, they’re also interesting in their own right because they are better constructively behaved than categories with a set of objects: For instance, in the last bullet point above we appealed to the statement “every fully faithful, essentially surjective functor is an equivalence”. For univalent categories, this is a theorem without any form of choice axiom!

One benefit that HoTT brings to the table is that it very naturally leads to mixing and matching the interpretations above. For example, the underlying groupoid of a univalent category can just as well be interpreted as a space, mixing interpretations #2 and #3. The spatial interpretation of categories really shines when we pass to higher categories, where the notion of “unique” behaves very badly.

When thinking even for small values of one must first weaken “unique” to “unique up to unique isomorphism”, and then to something like “unique up to isomorphism, which is unique up to unique invertible 2-cell”. Reasoning spatially, we see that “unique up to isomorphism” really means that there is a contractible space of choices for that object — a notion which is well-behaved at any level of higher categorical complexity.

Every type has an associated family of path types — to form a path type, we must have two inhabitants of handy, called the endpoints. The type of paths between and in is written where the subscript is omitted when it can be inferred from context. The name path is meant to evoke the topological idea of path, after which the paths in cubical type theory were modelled. In topology, a path is defined to be a continuous function from the real unit interval so that we can interpret a path as being a point “continuously varying over time”.

It would be incredibly difficult (and oversized!) to formalise the entire real line and its topology to talk about paths, which are meant to be a fundamental concept in type theory, so we don’t bother! Instead, we look at which properties of the unit interval give paths their nice behaviour, and only add those properties to our type theory. The topological interval is contractible, and it has two endpoint inclusions, with domain the one-point space. Those are the properties that characterise it (for defining paths), so those are the properties we keep!

The interval type, has two introduction rules — but they are so simple that we refer to them as “closed inhabitants” instead. These are written and and they denote the left- and right- hand side of the unit interval. A classic result in the interpretation of constructive mathematics says that any function defined inside type theory is automatically continuous, and so, we define the type to be the type of functions restricted to those where and definitionally.

The introduction rule for paths says that if with and (those are definitional equalities), then we can treat it as a path Conversely, if we have a term then we can apply it to an argument to get a term of type The type of paths satisfies the same reduction rule ( as function types, but with an additional rule that says is definitionally equal to (resp. and even if has not reduced to a

By considering the constant function at a point we obtain a term of type called the trivial path, and referred to (for historical reasons) as refl, since for the types that are sets, it expresses reflexivity of equality. In the interpretation of types as refl is the identity morphism. The rest of the groupoid structure is here too! Any path has an inverse and for any and there is a composite path There are paths between paths which express the groupoid identities (e.g.  and those paths satisfy their own identities (up to a path between paths between paths), and so on.

This little introduction to paths was written so that we can make some of the ideas in the following section, about dependent sums, precise. It’s not meant to be a comprehensive treatment of the path type; For that, see the section Paths, in detail later on.

Sums🔗

Recall that in the construction of a map into from a predicate we interpreted as a family of sets with at most one element, and then took the disjoint union of that family, written which admits a projection onto That was actually a sneaky introduction of the dependent sum type former, Indeed, the short discussion there also made some mention of the introduction rule, but let’s reiterate with more clarity here.

If we have some index type and a family of types then we can form the dependent sum type When the index is clear from context, we let ourselves omit the variable, and write Since is a type, it lives in a universe; Since it admits a projection onto and a projection onto a fibre of it must be a universe bigger than both and but it need not be any bigger: The dependent sum of lives in the universe indexed by

The introduction rule for says that if you have an and a then the pair inhabits the type Thus the inhabitants of the dependent sum are ordered pairs, but where the second coordinate has a type which depends on the value of the first coordinate. If we take to be a constant type family, we recover a type of ordered pairs, which we call the product of and which is notated by

For the elimination rules, well, pairs would be rather useless if we couldn’t somehow get out what we put in, right? The first elimination rule says that if we have an inhabitant then we can project out the first component, which is written to get an inhabitant of The computation rule here says that if you project from a pair, you get back what you put in:

Projecting out the second component is trickier, so we start by considering the special case where we’re projecting from a literal pair. We know that in the terms have types and so in this case, the second projection is When we’re not projecting from a literal pair, though, we’re in a bit of a pickle. How do we state the type of the second projection, when we don’t know what the first component is? Well - we generalise. We know that for a pair, so that we may replace the aforementioned typing for with the wordy Now we don’t use any subterms of the pair in the typing judgement, so we can generalise:

A very common application of the dependent sum type is describing types equipped with some structure. Let’s see an example, now turning to Agda so that we may see how the dependent sum is notated in the formalisation. We define a pointed magma structure on a type to be a point of that type, and a binary function on that type. If we let the type vary, then we get a type family of pointed magma structures:

  Pointed-magma-on : Type  Type
  Pointed-magma-on A = A × (A  A  A)

Taking the dependent sum of this family gives us the type of pointed magmas, where inhabitants can be described as triples where is the underlying type, is the distinguished point, and is the binary operation. Note that, to be entirely precise, the pair would have to be written but since by far the most common of nested pairs is with the nesting on the right, we allow ourselves to omit the inner parentheses in this case.

  Pointed-magma : Type₁
  Pointed-magma = Σ Type Pointed-magma-on

Note that, since we fixed the family to and recalling that are the first two universes, the type of all pointed magmas in lives in the next universe, If the type of all pointed magmas were in then we could again reproduce Russell’s paradox.

Just like functions, pairs enjoy a uniqueness principle, which (since coming up with names is very hard), also goes by The rule says that every inhabitant of a is definitionally equal to the pairing of its projections: This essentially guarantees us that the inhabitants of a dependent sum type can’t have any “extra stuff” hiding from our projection maps; They are exactly the pairs, no more, and no less.

The dependent sum is a construction that goes by many names. In addition to the ones mentioned below (which are the names people actually use), you could call it the Grothendieck construction, denote it by and have a hearty chuckle — if it weren’t for the fact that the Grothendieck construction is generally taken for functors between categories rather than just groupoids.

  • , named after its shape;

  • Disjoint union, since if we take to be the 2-point type, then the dependent sum of is exactly the disjoint union

  • Thinking homotopically, you can consider a type family to be a fibration, since as we shall see later it satisfies the path lifting property; In that sense, gives the total space, and the actual fibration is the first projection map

This last correspondence will be particularly important later, and it inspires much of the terminology we use in HoTT. Since universes are object classifiers, we know that there is an equivalence between type families and maps into Dependent sums and paths let us be more precise about this equivalence: We can turn any type family into a map and looking at the fibre of first over a point recovers We thus blur this distinction a bit and refer to for some as a fibre of and we say that something happens fibrewise if it happens for each

One last thing to mention about is that they are the realization of general subsets. This is a special case of the idea of stuff-with-structure, where the structure is merely a witness for the truth of some predicate of the first value. For example, recall that we defined the fibre of over as In type theory, this is rendered with a as below:

  fibre : (A  B)  B  Type
  fibre f y = Σ[ x  A ] (f x  y)

Dependent products🔗

Imagine that we have two types, and For simplicity, they live in the same universe, which can be any universe of our choice. We know that we can consider to be a constant type family taking indices in Furthermore, we know that the total space of this family is the product type and that it comes equipped with a projection map

Given any function we define its graph to be the function which takes each point of the input space to an ordered pair where the first coordinate is the input, and the second coordinate is the output of at that input. The graph of a function is special because of its interaction with the projection map In particular, we can see that is the identity function! This property turns out to precisely characterise the functions which are the graphs of functions

By rewriting the equality we noticed above in terms of function composition, we obtain A function, like which composes on the right with some other function to give the identity function, is called a section of This lets us characterise the functions as the sections of In this sense, we see that a function is precisely a choice of values where is in the fibre over — in this case, the fibres over all points are

Given a map that we know is a section of we can recover a function Specifically, given a point we know that the second projection has type — so we can simply define to be Since we know that is a section of we conclude that

This idea generalises cleanly from the case where is just some type, to the case where is a type family indexed by We define a section of to be a function where Note that this terminology conflates the type family with the projection map keeping with the trend of identifying type families with fibrations. Generalising the argument from the previous paragraph, we show that a section gives a rule for assigning a point in to each point in We again take the second projection of but since now is a type family, this lives in the type — but since we know that is a section of we can correct this to be

We would like to package up such a rule into a function but this is not a well-formed type, since is a type family. Moreover, this would “forget” that is sent to the fibre Thus, we introduce a new type former, the dependent product type, written or more directly The inhabitants of this type are precisely the sections of More importantly, the introduction, elimination, computation and uniqueness rules of this type are exactly the same as for the function type.

The dependent product type takes on another very important role when universes are involved: Supporting polymorphism. In programming, we say that a function is polymorphic when it works uniformly over many types; Dependent products where the index type is a universe reflect this exactly. For example, using polymorphism, we can write out the types of the first and second projection maps:

  project-first : (A : Type) (B : A  Type)  Σ A B  A
  project-first A B x = x .fst

  project-second : (A : Type) (B : A  Type)
                  (p : Σ A B)  B (project-first A B p)
  project-second A B x = x .snd

In the logical interpretation, the dependent product type implements the universal quantifier For example (keep in mind that in general, we should not read path types as equality), we could read the type as the proposition “for all pairs of points and of is equal to ”. However, an inhabitant of this type has a much more interesting topological interpretation!

Worked example: Interpreting dependent sums as fibrations and dependent products as sections lets us derive topological interpretations for types. We show how to do this for the type and show that a space admitting an inhabitant of this type has the property of “being contractible if it is inhabited”.

First, we’ll shuffle the type so that we can phrase it in terms of a single dependent product, and a single type family. We define a type to be a subsingleton if there is a term

An of this type can be seen as a section of the type family defined by the rule

The total space of this family is the space of all paths in which will be written as to emphasise its nature as a path space. Since it is the total space of a type family, it comes equipped with a fibration Given that a term in can be seen to be a pair where we see that this fibration projects both endpoints of a path — so we will write it since it is the pairing of the maps which evaluate a path at the left and right endpoints of the interval.

As a very quick aside, there is a map making the diagram below commute. This diagram expresses that the diagonal can be factored as a weak equivalence followed by a fibration through which is the defining property of a path space object.

A section of this fibration — that is, a dependent function like — is then a continuous function with the property that Now assume that our space is pointed, say with a point and that we have a section We can then define the homotopy mapping between the identity function on and the constant function at exhbiting the space as contractible.

If you assume the law of excluded middle, every space is either pointed or empty. In this case, we can describe a space equipped with a map like as being “either empty or contractible”, but we prefer the terminology “contractible if inhabited” or subsingleton instead.

What next?🔗

While the text above was meant to serve as a linearly-structured introduction to the field of homotopy type theory, the rest of the 1Lab is not organised like this. It’s meant to be an explorable presentation of HoTT, where concepts can be accessed in any order, and everything is hyperlinked together. However, we can highlight the following modules as being the “spine” of the development, since everything depends on them, and they’re roughly linearly ordered.

Paths, in detail🔗

open import 1Lab.Path

The module 1Lab.Path develops the theory of path types, filling in the details that were missing in Interlude - Basics of paths above. In particular, the Path module talks about the structure of the interval type the definition of dependent paths, squares, and the symmetry involution on paths. It then introduce the transport operation, turns a path between types into an equivalence of types.

It also exposits about the structure of types as Kan complices; In particular, the Path module defines and talks about partial elements, which generalise the notion of “horn of a simplex” to any shape of partial cube; Then, it defines extensibility, which exhibits a partial element as being a subshape of some specific cube, and the composition and filling operations, which express that extensibility is preserved along paths. Finally, it introduces the Cartesian coercion operation, which generalises transport.

To reiterate, paths are the most important idea in cubical type theory, so you should definitely read through this module!

Equivalences🔗

The idea of subsingleton type, mentioned in passing in the discussion about universes and expanded upon in the section on dependent products, generalises to give the notion of h-level, which measures how much interesting homotopical information a type has. A h-level that comes up very often is the level of sets, since these are the types where equality behaves like a logical proposition. Due to their importance, we provide a module defining equivalent characterisations of sets.

open import 1Lab.HLevel

Contractible types can be used to give a definition of equivalence that is well-behaved in the context of higher category theory, but it is not the only coherent notion. We also have half-adjoint equivalences and bi-invertible maps. Finally, we have a module that explains why these definitions are needed in place of the simpler notion of “isomorphism”.

open import 1Lab.Equiv
open import 1Lab.Equiv.Biinv
open import 1Lab.Equiv.HalfAdjoint
open import 1Lab.Counterexamples.IsIso

Univalence🔗

open import 1Lab.Univalence

The module 1Lab.Univalence defines and explains the Glue type former, which serves as the Kan structure of the universe. It then concretises this idea by describing how Glue can be used to turn any equivalence into a path between types, which leads to a proof of the univalence principle. After, we prove that the type of equivalences satisfies the same induction principle as the type of paths, and that universes are object classifiers.

open import 1Lab.Univalence.SIP

While univalence guarantees that equivalence of types is equivalent to paths in types, it does not say anything about types-with-structure, like groups or posets. Fortunately, univalence implies the structure identity principle, which characterises the paths in spaces of “structured types” as being homomorphic equivalence.


  1. As usual, we do not have a type of all groups. Instead, we have a family of types of groups that depends on how “big” their underlying type is allowed to be: what universe it lives in.↩︎

  2. Consequently, a term of the form is referred to as a redex (plural redices), short for expression, but since we’re not really interested in the theory of programming languages here, this terminology will not come up.↩︎

  3. A survey paper of POPL proceedings by Guy L. Steele (Steele 2017) identified twenty-eight different notations for substitution, so the word “typical” is.. questionable, at best.↩︎

  4. Solution: Recall what it means for a map to be injective - that If the domain of a function is a set with a single point, then the consequent is automatically satisfied! Since the maps are in 1-to-1 correspondence with the elements of we can consider any two numbers to be distinct subsets with the same domain.↩︎

  5. In some foundational systems — including HoTT, the one we are discussing — “subsingleton” is literally taken as the definition of “proposition”, so this translation isn’t merely omitted, it’s not necessary at all.↩︎

  6. Let me make this more precise. In any (univalent) category with finite limits we have a functor which takes to the full subcategory of on the injections. We say that is a subobject classifier if there is a natural isomorphism i.e., if is a representing object for the subobject functor. Since the Yoneda lemma implies that representing objects are unique, this characterises ↩︎

  7. Discreteness, type-theoretically, has to do with whether we can decide equality in that type by an algorithm. For example, the natural numbers are discrete, because we can compare the numerals they represent in time bounded by the size of the number. On the other hand, sequences of natural numbers are not discrete, because an algorithm can not run forever. Any discrete type is a set, and assuming excluded middle, any set is discrete.↩︎


References