Lady Margaret Hall \degreeMSc Mathematical Sciences \degreedateTrinity 2024
Synthetic Homotopy Theory
Abstract
The goal of this dissertation is to present results from synthetic homotopy theory based on homotopy type theory (HoTT). After an introduction to Martin-Löf’s dependent type theory and homotopy type theory, key results include a synthetic construction of the Hopf fibration, a proof of the Blakers–Massey theorem, and a derivation of the Freudenthal suspension theorem, with calculations of some homotopy groups of -spheres.
Keywords: homotopy type theory, homotopy theory, algebraic topology, type theory, constructive mathematics
Acknowledgements
I would like to thank my dissertation supervisor, Prof. Kobi Kremnitzer, for his invaluable guidance, inspirations and encouragement during this journey. His expertise and support have been instrumental in the completion of this work.
My gratitude also goes to my college supervisor, Dr. Rolf Suabedissen, for providing insightful advice on the structure and direction of this dissertation.
I am profoundly grateful to my parents for their unconditional and unwavering support in my academic pursuits and, indeed, all aspects of my life. Their efforts in their own field of research and teaching have been a constant source of motivation for me.
Finally, special thanks are due to all of my friends who helped me decide my academic path in pure mathematics or supported me in different ways during the writing of this dissertation. An incomplete list of their names follows: Qixuan Fang, Franz Miltz, Samuel Knutsen, Quanwen Chen, Sifei Li, Phemmatad Tansoontorn, Hao Chen and Haoyang Liu. This dissertation would not have been possible without each and every one of them.
Note
A slightly abridged version of this dissertation, originally written in Typst, was submitted in fulfillment of the requirements for the degree of MSc Mathematical Sciences at University of Oxford. That version was awarded the Departmental Dissertation Prize by the Mathematical Institute, University of Oxford. The current version has been extended and reformatted in LaTeX.
Chapter 1 Introduction
The goal of this dissertation is to present results in synthetic homotopy theory based on homotopy type theory (HoTT, also known as univalent foundations), including the Hopf fibration and Blakers–Massey theorem.
Compared to set theory, type theory is often considered as a better foundation for constructive mathematics and mechanised reasoning [Hou17, Section 1.2]. Martin-Löf’s intuitionistic type theory (ITT, also known as dependent type theory) was first proposed in the 1970s [ML75]. In ITT, any mathematical object is seen as inherently a term of some type. In particular, ITT follows the Curry–Howard correspondence, or the paradigm of propositions-as-types, where any proposition is regarded as a type and proving it is equivalent to constructing a term of that type. Since all terms of types can be seen as computer programmes and be executed, proof assistants, e.g. Coq, Agda and Lean, have been developed to ensure the correctness of proofs through type-checking [Bru16, Introduction].
HoTT is based on the intensional version of ITT, where a term of any identity type (i.e., the proposition that two terms of a type are equal) is not necessarily unique. This seemingly counterintuitive situation is addressed by interpreting types as topological spaces and identities as paths between points. The link between homotopy theory and type theory was first discovered in 2006 independently by Voevodsky and by Awodey and Warren [AW09]. In 2009, Voevodsky proposed the univalence axiom (UA), which formalises the idea that ‘isomorphic structures have the same properties’. In 2011, the concept of higher inductive types (HIT) emerged [Lum11], which allows the use of path constructors in addition to point constructors, so that many fundamental topological spaces can be elegantly constructed. In summary, the components of HoTT are
HoTT and -topoi form a syntax-semantics duality, and all results proven in HoTT hold in the -category of homotopy types [Car22]. This lays the foundation for our synthetic approach to homotopy theory, where all constructions are invariant under homotopy equivalences. Unlike in classical (or analytical) homotopy theory, notions such as spaces, points, continuous functions, and continuous paths are not based on set-theoretic constructions but are axiomatised as primitive ones, showing the formal abstraction provided by type theory [Hou17, Section 1.2].
Many results of synthetic homotopy theory, including the Hopf fibration, Freudenthal suspension theorem, and the van Kampen theorem, can be found in [Uni13, Section 8]. More recent efforts in this field include [Bru16], the Blakers–Massey theorem [HFLL16], the Cayley–Dickson construction [BR16], the construction of the real projective spaces [BR17], and the long exact sequence of homotopy -groups [BR21]. Additionally, synthetic (co)homology theory has been investigated by [Cav15], [BH18] and [Gra18].
This dissertation begins with a brief summary of dependent type theory in Chapter 2 and an overview of HoTT in Chapter 3. Chapter 4 mainly discusses homotopy -types and -connected types. Chapter 5 develops the fibre sequence, Hopf construction and Hopf fibration. Chapter 6 proves the Blakers–Massey theorem. Along the journey, we will calculate the homotopy groups of -spheres in Table 1.1.
Although no result is essentially new, calculations and discussions original to this dissertation are marked with a red square ( ). For instance, Chapter 4 contains discussions on path inductions, Corollary 4.4 and Corollary 4.6, and novel examples, Example 4.2 and Example 4.25. Most noticeably, Chapter 6 ‘informalises’ a mechanised Blakers–Massey theorem proof [HFLL16] by integrating details from a Freudenthal suspension theorem proof [Uni13, Section 8.6], before explicitly deriving the Freudenthal suspension theorem as a corollary. Additionally, visualisation diagrams are plotted for certain definitions and theorems to aid understanding.
Chapter 2 Dependent type theory
Dependent type theory is a formal system that organises all mathematical objects, structures and knowledge by expressing them as types or terms of types [Rij22, p. 1]. A type is defined by rules specifying its own formation and the construction of its terms. Types resemble sets; however, a major difference is that membership in a type is not a predicate but is part of the inherent nature of its term.
We write if is a term of type , where we also say that type is inhabited and is an inhabitant of . All types are considered terms of a special type, , and we denote if is a type111We suppress the underlying universe level (see [Uni13, Section 1.3]).. If and , we write to indicate that and are judgementally equal, meaning that and are only symbolically different. When defining a new term from , we use the notation . Judgemental equalities are to be contrasted with propositional equalities introduced in Section 2.3.
We shall briefly sketch the fundamentals of dependent type theory, based on [Rij22, Chapter I] and [Uni13, Chapter 1]. Importantly, these constructions all bear interpretations in logic and topology in Table 2.1.
Type theory | Logic | Topology |
---|---|---|
a proposition | a topological space | |
a proof of the proposition | a point in the space | |
if then | a continuous map | |
a predicate for each | a fibration over with fibre for each | |
for all , | a continuous section of fibration | |
there exists , 222This type carries more information than merely a ’true or false’ judgement of the existence, since a term of this type provides a concrete as well as a term , i.e., a proof for . | the total space of fibration | |
and | the product space | |
or 333A term of also carries the information ’which one of and is true’. | the disjoint union | |
True | the contractible space | |
False | the empty space | |
is a proof of the equality of and | is a continuous path from to | |
the reflexivity of equality at | the constant path at |
2.1 Function types
Let . A dependent type (or a family of types) over assigns a type to each term .
A dependent function is a function whose output types may vary depending on its input. The dependent function type, also known as -type, is specified by:
-
(1)
The formation rule: Given a type and a dependent type , the dependent function type can be formed.
-
(2)
The introduction rule: To introduce a dependent function of type , a term is needed for each . This dependent function is then denoted by .
-
(3)
The elimination rule: Given a dependent function and a term , a term is obtained by evaluating at .
-
(4)
The computation rule: Combining the introduction and elimination rules, .
In particular, when is a constant type family assigning the same type to every , we obtain the (non-dependent) function type .
2.2 Inductive types
A wide range of constructions are based on inductive types. An inductive type is specified by the following:
-
(1)
The constructors (or construction rules) of state the ways to construct a term of .
-
(2)
The induction principle states the necessary data to construct a dependent function for a given dependent type . Particularly, the recursion principle is the induction principle when is a constant type.
-
(3)
The computation rule states how the dependent function produced by the induction principle acts on the constructors.
2.2.1 -types
Given a type and a dependent type , we can form the -type, written as whose terms are called dependent pairs.
-
(1)
The construction rule states that given and , we can construct the dependent pair .
-
(2)
The induction principle states that given
-
•
a dependent type and
-
•
a function 444Arrows are right-associative.,
there is a dependent function .
-
•
-
(3)
The computation rule states that the above satisfies .
The function above is the curried form of and is the uncurried form of , and they can be used interchangeably. Unlike the convention in [Uni13], we use the notation that reflects the actual curriedness of a function, i.e., is not written as .
Using the induction principle, we define the first projection function by and the second projection function by .
The case when is a constant type gives the cartesian product . In this case, we write and for the projection maps onto and , respectively.
2.2.2 Other inductive types
Some other inductive types are specified as follows based on [Esc22, Section 2].
-
•
The empty type has no construction rule. The induction principle states that given , there is a dependent function . In particular, the recursion principle states that for any , we can define a function , which can be logically interpreted as the principle of explosion, or ex falso quodlibet. There is also no computation rule for . For any type , is defined as the function type .
-
•
The unit type has a single constructor . The induction principle for states that given a dependent type and a term , there is a dependent function . The computation rule555Henceforth, we may concisely state the computation rule as a part of the induction principle. states that such constructed satisfies .
-
•
Let . The disjoint sum of and has construction rules and . The induction principle states that given
-
–
a dependent type ,
-
–
two dependent functions and
-
–
,
there is a dependent function such that and .
-
–
-
•
The boolean type is defined as , and its two constructors are written as and .
-
•
The type of natural numbers has construction rules and . The induction principle states that given
-
–
a dependent type ,
-
–
a term , and
-
–
a dependent function ,
there is a dependent function such that and .
-
–
-
•
The type of integers has three constructors, representing zero, the positive and negative numbers, and its induction principle and computation rules are similar to [Bru16, Section 1.3].
2.3 Identity types
Given a type and two terms and , we can form the identity type , whose terms are called (propositional) equalities or paths. The constructor for a family of identity types indexed by is given by a dependent function
However, Axiom K, the assumption that for any , does not hold in general.
For a fixed , the induction principle (called the path induction) for a family of identity types indexed by states that given
-
•
a dependent type and
-
•
a term ,
there is a dependent function
such that . The path induction implies that to construct a dependent function for a path from a fixed endpoint to a free , it suffices to assume that is and is . Importantly, one endpoint should be free, since otherwise we would have Axiom K. Intuitively, the path induction reflects the fact that the space of all paths starting from a fixed point is contractible. We shall revisit this idea in Proposition 4.3.
The reader is advised to review Table 2.1 at this point.
Chapter 3 Homotopy type theory
The implications of the identity type are profound, especially when it interacts with other types. In Section 3.1, the identity type equips each type with an -groupoid structure. In Section 3.2, we explore how equalities can be passed along by non-dependent and dependent functions. The notions of ‘equivalence’ between functions and between types are subtle ones treated in Section 3.3, and their relationship with paths in function types or in is handled in Section 3.4 by function extensionality and univalence axiom. In Section 3.5, we introduce higher inductive types that allow path constructors. Together, these form the basic components of HoTT.
3.1 The -groupoid structure of types
In classical mathematics, equalities are reflexive, symmetric and transitive. While reflexivity is expressed as the constructor for identity types, symmetry and transitivity appear in the form of path inversion and concatenation.
Lemma 3.1 ([Uni13, Lemma 2.1.1]).
For every type and , there is a function , denoted , such that , where is called the inverse of .
Remark 3.2.
By the paradigm of propositions-as-types, this lemma is equivalent to the type
The proof shows how we construct a term of this type.
Proof.
Fix and . By path induction, it suffices to assume is and is . Then it suffices to give a term of , which is fulfilled by . We thus have . ∎
Lemma 3.3 ([Uni13, Lemma 2.1.2]).
For every type and , there is a function , denoted , such that , where is called the concatenation of and .
Proof.
By path induction twice.∎
Lemma 3.4 ([Uni13, Lemma 2.1.4]).
Path concatenation is associative, i.e., for any , , , and , there is a -dimensional path .
Proof.
By path induction, we assume are all . Then by Lemma 3.3, so we give . ∎
We similarly can show that path concatenation satisfies inverse and identity laws [Uni13, Lemma 2.1.4]. The above are all examples of coherence operations on paths or higher-dimensional paths. We omit further discussions and direct to [Uni13, Section 2.1] or [Bru16, Section 1.4] for details. Henceforth, we shall omit proofs which only require path inductions.
We then observe that the identity types can be constructed recursively: for any type , we can form , , and , etc. This observation, combined with the next definition, pointed types, gives rise to the notion of loop spaces.
Definition 3.5.
The type of pointed types is defined as . Alternatively, a pointed type is a dependent pair where and , and is called the basepoint of this pointed type. We often leave the basepoint implicit.
Example 3.6.
Let , then define a pointed type . This is called adjoining a disjoint basepoint.
Definition 3.7.
The loop space of a pointed type is defined as the pointed type
Then for , define the -th loop space of by and (basepoints implicit).
In summary, coherence operations and loop spaces demonstrate that the identity type equips each pointed type with a (weak) -groupoid structure (see [Bru16, Appendix A] for a formal definition).
3.2 Dependent paths
In classical mathematics, for any function between sets and , implies . The analogue in HoTT is expressed as a path induced by the path , demonstrating the proof relevance feature of type theory.
Lemma 3.8 ([Uni13, Lemma 2.2.1]).
Let , , and . Then there is a path , such that .
However, for a dependent function , there is not necessarily a path from to , which generally belong to different types and , even when there is . We have to ‘transport’ from to ‘along’ first, shown by the following two lemmas, visualised in Figure 3.1.
Lemma 3.9 ([Uni13, Lemma 2.3.1]).
Let , , and . Then there is a function , such that .
Lemma 3.10 ([Uni13, Lemma 2.3.4]).
Let , , , and . Then there is a dependent path such that , where we define the type
for any and .
The next lemma relates the dependent and non-dependent cases, visualised in Figure 3.2.
Lemma 3.11 ([Uni13, Lemma 2.3.5, Lemma 2.3.8]).
Let , , and . Let be a constant family of types given by . Then for any , there is a path
Let , then .
The next result shows the functoriality of .
Proposition 3.12 ([Uni13, Lemma 2.3.9]).
Let , and . Then
It is helpful to characterise when the dependent type is of particular forms, as in the following lemmas, visualised in Figure 3.3, Figure 3.4 and Figure 3.5.
Lemma 3.13.
Let , , , and . Let . Then
Lemma 3.14 ([Uni13, (2.9.4)]).
Let and . Let . Let , . Let . Then
Lemma 3.15 ([Uni13, Lemma 2.11.2]).
Let , and . Then
3.3 Homotopies and equivalences
The next problem is how to construct equalities between functions and between types. In this section, we shall present two constructions, function homotopy and type equivalence, based on the identity types.
Definition 3.16.
Let , and . A (function) homotopy from to is a dependent function of type
i.e., a family of pointwise equalities.
Definition 3.17.
Let , . A quasi-inverse of is a triple , where , and .
Proposition 3.18 ([Uni13, Section 2.4]).
Let and . Define type
We say ‘ is an equivalence’ if is inhabited. Then
-
(1)
is an equivalence if and only if has a quasi-inverse;
-
(2)
Given any , there is a path .
In the language of Section 4.2, (2) states that is a mere proposition. In practice, to prove that a function is an equivalence, by (1), we only have to find a quasi-equivalence of . An example follows.
Lemma 3.19 ([Uni13, Example 2.4.9]).
Let , , and . Then is an equivalence.
Proof.
By Proposition 3.12, forms a quasi-inverse with .∎
Definition 3.20.
Let . Then we define the type of equivalences between and as
We now characterise the identity type in certain types using the language of equivalence:
Proposition 3.21 ([Bru16, Proposition 1.6.8]).
Let , and two terms , then there is an equivalence of types
The function from right to left is written as , i.e., given and , there is a path .
3.4 Univalence axiom and function extensionality
Given , it is natural to wonder what is the relationship between type equality and type equivalence . It turns out that we can derive the latter from the former:
Proposition 3.22 ([Uni13, Lemma 2.10.1]).
Let . Then there is a function .
Proof.
Let . The identity function can be regarded as a dependent type, so by Lemma 3.9, we have a function , which is an equivalence by Lemma 3.19. ∎
Desirable as it is, we however cannot go the other way around with what we currently have. Our wish is instead formulated as an axiom:
Axiom 3.23 (Univalence).
Let . Then is an equivalence.
Remark 3.24.
The axiom means that we assume the type is inhabited without constructing its term.
In particular, this implies that there is a function
which forms a quasi-inverse of . By abuse of notation, we write if is an equivalence. In practice, to find a path between two types and , it suffices to give a function and show that is an equivalence.
A similar situation arises with function equality and function homotopy.
Proposition 3.25.
Let , and . Then there is a function .
Proof.
Let . For any , the evaluation function is a non-dependent one, so we can apply Lemma 3.8 and get ∎
The next theorem can be proved with univalence [Uni13, Section 4.9].
Theorem 3.26 (Function extensionality).
The function is an equivalence.
This implies that there is a function
which forms a quasi-inverse of . In practice, to give an equality between two functions and , it suffices to give a family of pointwise equalities.
3.5 Higher inductive types
In classical dependent type theory, any constructor for an inductive type should produce a term of . This is not the case for higher inductive types (HIT), where path constructors are allowed which produce paths, or ‘customised equalities’, between points of . We now introduce two HITs, the circle and pushouts.
3.5.1 Circle
The circle is the HIT generated by a (point) constructor and a path constructor . The wording here is ‘generated’, in the sense that by coherence operations, from the constructor we automatically obtain paths such as and . Note that is non-trivial, in the sense that 111Note that this means the type is inhabited.[Uni13, Lemma 6.4.1].
The induction principle states that given
-
•
a dependent type ,
-
•
a term and
-
•
a dependent path ,
there is a dependent function such that and 222Here the computation rule for the path is a propositional equality instead of a judgemental one; see [Uni13, Section 6.2]..
Remark 3.27 ([Uni13, Remark 6.2.4]).
When informally applying circle induction on given , we may use expressions ‘when is ’ and ‘when varies along ’ to introduce the constructions of and as above.
The recursion principle states that given
-
•
,
-
•
and
-
•
,
there is a function such that and [Uni13, Lemma 6.2.5].
3.5.2 Pushouts
Let , and . Then we get a diagram , which we call a span. The pushout of this span is the HIT generated by two point constructors and and a path constructor
as visualised by Figure 3.6.
The induction principle states that given
-
•
a dependent type ,
-
•
two dependent functions and
-
•
, and
-
•
a family of dependent paths
there is a dependent function such that , and .
By specialising the span, we can obtain many interesting constructions from the pushout.
-
(1)
Let . The suspension of is the pushout of
We denote , , and for any , . can be seen as two points (’north and south poles’) connected by a family of paths (’meridians’) indexed by the terms of .
-
(2)
Let . The join of and is the pushout of
-
(3)
Let and be two pointed types. The wedge sum of and is the pushout of
With suspension, we define the -sphere by induction on : and . Note that [Uni13, Lemma 6.5.1], where is in the sense of Section 3.5.1.
For any type , we define its pointed suspension as This turns into pointed types for . For , we view as . We write the basepoint of as .
Chapter 4 Homotopy -types and -connected types
Section 4.1 discusses contractible types, ‘trivial types’ that are equivalent to . Section 4.2 generalises contractible types to (homotopy) -types: types with trivial ‘homotopy information’ in any dimension greater than (as in Theorem 4.8). Section 4.3 discusses how to truncate any given type into an -type. Then our tools at hand will allow us to define and calculate the homotopy groups of the circle in Section 4.4. Finally, Section 4.5 focuses on the ‘dual’ of -types, -connected types, which have trivial ‘homotopy information’ in any dimension less than or equal to .
4.1 Contractible types
Definition 4.1.
For , we define the predicate
which reads ‘ is contractible’, and we say is the centre of contraction.
Clearly, is a contractible type. In fact, is contractible if and only if [Uni13, Lemma 3.11.3]. Informally, contractibility means having trivial ‘homotopy information’ in all dimensions (observe that for all ). We first give a non-example.
Example 4.2.
is not contractible. Otherwise, taking as the centre of contraction, we have to prove . By circle induction, when is , we need . When ‘varies along’ , we need . By Lemma 3.15, , so by coherence operations, we need , but this would give us absurdity by the non-triviality of [Uni13, Lemma 6.4.1].
As mentioned earlier, the space of all paths from a fixed point is contractible:
Proposition 4.3.
Let and . Then the type is contractible.
Proof.
The centre of contraction is given by . For any , we have and by Lemma 3.15. Then we have by Proposition 3.21. ∎
This lends us some further insights into path induction, as in the two following corollaries.
Corollary 4.4 (Path induction, strengthened).
Let , , and . Then for any , and , there is a function .
Proof.
Let be the uncurried form of . Then by the induction principle of , it suffices to take any and to obtain . The desired is the curried form of . ∎
Remark 4.5.
In the usual path induction, and should be and respectively.
We may also use to express functions defined using path induction:
Corollary 4.6.
Proof.
By path induction, assume is and is , then both sides are .∎
4.2 Homotopy -types
Definition 4.7.
We recursively define the predicate
for any , which reads ‘ is an -type’. We also define
and in particular, we write .
A -type is by definition a contractible type; we also define a mere proposition as a -type and a set as a -type. They are characterised as follows:
-
•
is contractible, if and only if is a pointed mere proposition, if and only if [Uni13, Lemma 3.11.3].
- •
-
•
is a set, if and only if for any and we have , if and only if satisfies Axiom K, i.e., is contractible for any [Uni13, Theorem 7.2.1]. This aligns with the ‘equality-irrelevant’ sets in classical mathematics and allows us to talk about set-theoretic notions such as subsets. The types , and are sets [Bru16, Proposition 1.10.7].
As a generalisation of Axiom K, we can characterise -types in terms of loop spaces:
Theorem 4.8 ([Uni13, Theorem 7.2.9]).
For any , a type is an -type if and only if is contractible for all .
Noticeably, -types have the cumulativity property:
Proposition 4.9 ([Uni13, Theorem 7.1.7]).
Let . An -type is also an -type.
4.3 -truncations
Any type can be turned into an -type , called the -truncation of , by a map . The construction of as an HIT is elaborated in Appendix A.
The induction principle (called the truncation induction) for states that given
-
•
a dependent type and
-
•
a dependent function ,
there is a dependent function , such that for all [Uni13, Theorem 7.3.2]. The recursion principle states that given
-
•
an -type and
-
•
a function ,
there is a function such that .
Particularly, the -truncation turns any type into a mere proposition. This allows us to get rid of the ‘superfluous’ information carried by the -type and disjoint union, as footnoted in Table 2.1, and perform logic in the classical sense. For and , we shall use the expression ‘there merely exists such that ’ to represent the type (see [Uni13, Sections 3.6, 3.7, 3.10]).
The next result shows that truncation operation is a functor:
Proposition 4.10.
Given and , there is a map such that .
Proof.
Since is an -type, define by by the recursion principle. ∎
Unsurprisingly, -truncating an -type gives an equivalence:
Proposition 4.11 ([Uni13, Corollary 7.3.7]).
Let . is an -type if and only if the map is an equivalence.
Proof.
Suppose is an -type. Then we are allowed to define by . By definition, . Then we would like to define a dependent function . Notice that is an -type (since it is by definition an -type), so by its induction principle it suffices to give a dependent function , which can be given by since . Therefore, forms a quasi-inverse for and is an equivalence.
Conversely, if is an equivalence, then since is an -type, so is . ∎
The following two results deal with -truncations of identity types and loop spaces:
Proposition 4.12 ([Uni13, Theorem 7.3.12]).
Let , and . Then
Proposition 4.13 ([Uni13, Corollary 7.3.14]).
Let , , and be a pointed type. Then
Finally, -truncations have the cumulativity property:
Lemma 4.14 ([Uni13, Lemma 7.3.15]).
For and ,
4.4 Interlude: homotopy groups of the circle
In this interlude, we first take groups as an example to demonstrate how to construct algebraic structures. Then we define homotopy groups and calculate them for .
Definition 4.15.
A group is a set with functions (called the multiplication function), (called the inversion function) and a term (called the neutral element), such that the equalities hold for all :
-
•
;
-
•
;
-
•
;
-
•
;
-
•
.
The group is abelian if further hold for all .
Remark 4.16.
Since a group is by definition a set, any of these equalities has at most one inhabitant. Hence, the predicate ‘ is a group’ is a mere proposition. This allows the groups we define to behave identically as in classical mathematics.
Definition 4.17.
Let and be two groups. Then a function is called a group homomorphism if for all . A group homomorphism is a group isomorphism if it is an equivalence.
Example 4.18.
is trivially a group, written as .
Importantly, set-truncation enables us to turn a loop space into a group:
Definition 4.19.
Let be a pointed type and . Define the -th homotopy group of as .
When , the set is a group with path concatenation, path inversion and reflexivity. When , by the Eckmann-Hilton argument [Uni13, Theorem 2.1.6], is an abelian group. We may also write , although it is not a group in general. Finally, is a functor, as truncation and loop space are both functors by Lemma 5.6 and Proposition 4.10.
We know that is not contractible by Example 4.2. For the rest of this section, we briefly sketch the calculation of for all . We proceed by first showing with the encode-decode method based on [Uni13, Section 8.1.4]. This technique will be used again in Chapter 6.
The initial idea would be to give two functions and show that they are quasi-inverses. The function is given by , where denotes the -time concatenation of if or if and if . However, we cannot easily define a function , as path induction is unavailable when the two endpoints are both fixed. We thus generalise the claim to the following.
Lemma 4.20.
There exists a dependent type , such that
and .
Proof sketch.
To finish the definition of , as is an equivalence, is a path . We then set . represents the universal cover of the circle: the fibre over is and going around once corresponds to going up by one in .
From this definition, we can prove that
(4.1) |
for any (by [Uni13, Lemma 8.1.2] and Proposition 3.12).
It then remains to define four functions of the types:
-
(1)
,
-
(2)
,
-
(3)
, and
-
(4)
,
where (3) and (4) state that for each , and are quasi-inverses.
To proceed, (1) is given by
(2) uses circle induction, where is given by the function defined earlier; (3) uses path induction and (4) circle induction, and they both apply Equality 4.1. We omit the details. ∎
Now setting to in Lemma 4.20 yields .
Corollary 4.21.
and for .
Proof.
Since is a set, . Thus, . Further, since is a set, is also a set, which implies that is trivial for any , and therefore so is . ∎
4.5 -connected types
Definition 4.22.
Let and . The fibre of over is defined as the type
Definition 4.23.
A type is -connected if is contractible, and a function is -connected if for all , is contractible. In particular, a function is surjective, connected, or simply-connected if it is -, -, or -connected, respectively.
Remark 4.24.
An -connected function can be thought of as ‘being surjective up to dimension ’. In particular, preserves ‘homotopy information’ up to dimension . This is formalised in Lemma 4.32 and Corollary 5.14. We also remark that different conventions in defining the -connectedness of functions exist.
Easily seen from the definitions are:
-
(1)
A type is -connected if and only if the unique function is -connected;
-
(2)
Every type is -connected and so is every function;
-
(3)
Every pointed type is -connected.
Example 4.25 ( Inspired by [Hou17, Figure 3.4]).
Consider an HIT generated by a non-empty graph : each vertex gives a term and each edge connecting vertices and gives a path , where the order of and is arbitrary thanks to path inversion.
If has only one connected component, then is -connected. Indeed, as is pointed, it suffices to show that is contractible for any . The goal is a mere proposition, so we assume is and is for . By Proposition 4.12, , but is pointed by assumption and is thus contractible.
If further is a tree, then is -connected. Indeed, using the same reasoning twice, it suffices to show that is contractible for any and , which holds since any two vertices on a tree has a unique path between them up to homotopy.
Example 4.26.
is pointed and is -connected. Example 4.25 indicates that is -connected. In fact, for all , the sphere is -connected [Uni13, Corollary 8.2.2].
We observe that -connected types and functions have the (downward) cumulativity property:
Proposition 4.27 ([Bru16, Proposition 2.3.7]).
Let . Any -connected type (resp. function) is also an -connected type (resp. function).
Proof.
Let be -connected. Then by Lemma 4.14, and by Proposition 4.11 since is contractible and a fortiori an -type. So is contractible. The case for -connected functions is similar. ∎
We can also say something about the lower homotopy groups of an -connected type:
Lemma 4.28 ([Uni13, Lemma 8.3.2]).
If is a pointed -connected type, then for all .
Proof.
Corollary 4.29.
for .
Proof.
By Example 4.26 and Lemma 4.28. ∎
We have ‘induction principles’ for -connected types and functions:
Proposition 4.30 (Induction principle for -connected types).
Suppose and is an -connected type. Given
-
•
a dependent type ,
-
•
and
-
•
,
there is a dependent function such that .
Proof.
Since is an -type, by the recursion principle of , there is a dependent type such that . Since is contractible, there is a function such that . Therefore, we obtain . ∎
Proposition 4.31 (‘Induction principle’ for -connected maps, [Uni13, Lemma 7.5.7]).
Suppose , , and . Then is -connected, if and only if for any , the map
is an equivalence.
Intuitively, Proposition 4.30 says that an -connected type ‘looks like’ a contractible type to any family of -types, and Proposition 4.31 says that in the presence of an -connected map , a family of -types over sees as being ‘generated’ by [HFLL16, Section 2].
Finally, we mention that an -connected function induces an equivalence on -truncations:
Lemma 4.32 ([Uni13, Lemma 7.5.14]).
If is -connected, then it induces an equivalence .
Chapter 5 Hopf fibration
The centre of this chapter is the following definition:
Definition 5.1.
Let , be pointed types. Define the type of pointed maps from to as
We may say that is a pointed map witnessed by path .
For any pointed map , the homotopy groups of , and are related by a long exact sequence, obtained from a fibre sequence, shown in Section 5.1. In particular, given a ‘pointed fibration’ (in the sense of Corollary 5.3), , , correspond respectively to the fibre, total space and base space of the fibration. In Section 5.2, we work to build the Hopf construction, a fibration based on H-spaces. In Section 5.3, we demonstrate an H-space structure on and derive the Hopf fibration, whose long exact sequence fascinatingly connects the homotopy groups of , and .
5.1 Fibre sequences
We begin with a simple yet powerful observation.
Lemma 5.2.
Let and be pointed types and let be witnessed by . Let and . Then we have pointed types and pointed maps:
Proof.
is pointed by , and since , is a pointed map witnessed by . ∎
A special case is the following, a fibration over with fibre and total space .
Corollary 5.3.
Let be a pointed type and be a dependent type such that is a pointed type. Then we have pointed types and pointed maps:
called the fibration sequence, where and .
Proof.
has basepoint and is thus a pointed map. Then we can apply Lemma 5.2, but by [Uni13, Lemma 4.8.1] (which uses Proposition 4.3). The map is as above under this equivalence. ∎
The rest of this section is based on the more general Lemma 5.2, although later in practice we only use the case Corollary 5.3. Since is again a pointed map, we apply Lemma 5.2 recursively to yield the following construction:
Definition 5.4.
Let and be pointed types. The fibre sequence of a pointed map witnessed by is the infinite sequence of pointed types and pointed maps
We denote as the basepoint of and as the witnessing path of . The sequence is then defined recursively by
-
•
, , , , , ;
-
•
If ,
We denote .
This fibre sequence can be (magically) turned into an equivalent form where the loop spaces of , and show up in a periodic fashion. The following two ingredients are needed before we move on.
Definition 5.5.
If is a type equivalence, then is a pointed equivalence and in this case and are pointed equivalent.
Lemma 5.6.
The looping of defined as is also a pointed map.
Proof.
. ∎
We now work to reveal the first step of the alleged equivalence.
Lemma 5.7 ([Uni13, Lemma 8.4.4]).
Let and be pointed types and let . Then in the fibre sequence of . Further, under the equivalence, is identified with
Proof.
See Figure 5.1. Using the definition of fibres twice, a term of is of the form , where , and . We first define a function . Take any . Then . Thus since , we can produce . Conversely, we define by
Now take any , then
where the last equality is by Proposition 3.21 and Lemma 3.13. Conversely, for any ,
Thus forms a quasi-inverse of . Also and are both pointed maps, so and are pointed equivalent. We therefore identify with , which can be simplified as , easily a pointed map. ∎
By applying Lemma 5.7 recursively to the fibre sequence, we have:
Corollary 5.8 ([Uni13, Lemma 8.4.4]).
The fibre sequence of can be equivalently written as
where and the minus sign represents composition with path inversion.
We will then show that this fibre sequence, when set-truncated, becomes an exact sequence, defined below.
Definition 5.9.
Let be sets and let , then the image of is defined as If are pointed sets and let , then the kernel of is defined as .
Definition 5.10.
An exact sequence of pointed sets is a sequence of pointed sets and pointed maps:
such that for every . The sequence is an exact sequence of groups if further each is a group and each is a group homomorphism.
Lemma 5.11 ([Uni13, Theorem 8.4.6] 111 A few typos in [Uni13, Theorem 8.4.6] are fixed in this proof.).
Proof.
. Take any , then as witnessed by . Thus sends everything to , and so does by the functoriality of .
. For any and , we claim that there merely exists such that . As the goal is a mere proposition and a fortiori a -type, by truncation induction, it suffices to assume is for some . Then since , we have , which by Proposition 4.12, corresponds to some . Again by truncation induction, it suffices to assume is for some . Thus we take , and then . ∎
Theorem 5.12 ([Uni13, Theorem 8.4.6]).
Let be pointed types and let . Let . Then we have an exact sequence of pointed sets
This is further an exact sequence of groups except the terms, and an exact sequence of abelian groups except the and terms.
Proof.
By recursively applying Lemma 5.11 to the fibre sequence of , the above is an exact sequence of pointed sets. The rest of the proof, including some technicalities involving path inversion, is omitted. ∎
The following lemma is commonly known in classical algebraic topology.
Lemma 5.13 ([Uni13, Lemma 8.4.7]).
Let be an exact sequence of abelian groups. Then (1) is injective if is ; (2) is surjective if is ; (3) is an isomorphism if and are both .
From the next result, which characterises an -connected function in terms of homotopy groups, we formalise the idea that an -connected function is a generalisation of a surjective function.
Corollary 5.14.
Let be pointed types and let be -connected.
-
(1)
If , then is an isomorphism;
-
(2)
If , then is surjective.
Proof.
We omit the case when and suppose . We have an exact sequence
Since is -connected, by Proposition 4.27, is also -connected when . Thus is contractible, and (by Proposition 4.13) is also contractible for . Applying Lemma 5.13 yields the result. ∎
Remark 5.15.
Set to and Corollary 5.14 would imply Lemma 4.28.
5.2 Hopf construction
To trigger all the machinery in the previous section, we only need a fibration sequence as in Corollary 5.3. We now build the Hopf construction, a fibration sequence derived from a -connected H-space.
Definition 5.16.
A type is an H-space if it is equipped with a base point , a binary operation and equalities and for every .
Lemma 5.17.
Let be a -connected H-space. Then and are equivalences for every .
Proof.
Let be the dependent proposition which sends to the predicate that (resp. ) is an equivalence. By Proposition 4.30, it suffices to prove , which holds by definition. ∎
The flattening lemma, used in the next proof, is a result concerning the total space of a fibration over an HIT constructed via univalence [Uni13, Section 6.12].
Theorem 5.18 ([Uni13, Lemma 8.5.7]).
Let be a -connected -space. Then there is a dependent type (fibration) with fibre and total space , called the Hopf construction.
Proof.
We define a dependent type by , , and since each is an equivalence . By the flattening lemma (see [Uni13, Lemma 8.5.3] or [Bru16, Proposition 1.9.2]), is equivalent to the pushout of the span We define a function by , and is an equivalence because it has quasi-inverse . Therefore, we have an equivalence of spans
since both squares commute and each vertical line is an equivalence. Thus, the pushout of the two spans are also equivalent, but the pushout of the lower span is by definition . ∎
5.3 Hopf fibration
We finally present the Hopf construction induced by the H-space structure on .
Lemma 5.19 ([Uni13, Lemma 8.5.8]).
There is an -space structure on .
Proof.
Take . Now we define by circle induction on the first argument. We define ; then we need a path such that . By function extensionality, this is equivalent to a path for each . By induction on again, we define . It then remains to give a dependent path , which is equivalent to by Lemma 3.13.
Now since . We show that by circle induction on . When is , , so we take ; when ‘varies along’ , we need a dependent path
which is equivalent to by Lemma 3.13. By function extensionality,
Finally, we change to its uncurried form. ∎
The next result follows from the -lemma [Bru16, Section 1.8].
Lemma 5.20 ([Bru16, Proposition 1.8.8]).
For every , .
Theorem 5.21.
There is a fibration, called the Hopf fibration, over with fibre and total space .
Proof.
is -connected by Example 4.26. By Lemma 5.19 and Theorem 5.18, there is a fibration over with fibre and total space by Lemma 5.20. ∎
We put all the pieces together by writing the long exact sequence induced by the Hopf fibration:
Corollary 5.22 ([Uni13, Corollary 8.5.2]).
and for .
Proof.
By Corollary 5.3, Theorem 5.12 and Theorem 5.21, we have an exact sequence of groups
By Corollary 4.21 and Corollary 4.29, this is reduced to
Applying Lemma 5.13 yields the result. ∎
Chapter 6 Blakers–Massey theorem
In this chapter, we will ‘informalise’ the mechanised Agda proof of the Blakers–Massey theorem in [HFLL16] (or [Hou17, Sections 3.4, 4.3]) by combining [Uni13, Section 8.6]. For this, a generalised version of pushouts is first presented in Section 6.1. Then the theorem is stated, generalised and relaxed in Section 6.2. The proof employs a more sophisticated encode-decode method, where we define by the wedge connectivity lemma in Section 6.3 and then show the contractibility of in Section 6.4. Some consequences, including the Freudenthal suspension theorem and stability for spheres, are derived in Section 6.5.
Throughout the chapter, we shall omit proofs on coherence operations too technical for presentation, and write a square for an unspecified coherence path.
6.1 Generalised pushouts
Lemma 6.1 ([Uni13, Lemma 4.8.2]).
For any function , .
Proof.
By Proposition 4.3.∎
This result shows that we can interchange function types and families of types, i.e., any function is equivalent to the projection .
Recall that Section 3.5.2 defines a pushout for the span . By rewriting the functions as families of types and expanding the definition of fibres, this span is equivalent to
If we define a dependent type , then this span can be generalised as
Definition 6.2.
Let and . The (generalised) pushout is the HIT generated by two point constructors and and a path constructor
as visualised by Figure 6.1.
By previous discussions, when , this definition is equivalent to the one in Section 3.5.2. Henceforth, we only use the generalised pushout.
The induction principle states that given
-
•
a dependent type ,
-
•
two dependent functions and
-
•
and
-
•
a family of dependent paths
there is a dependent function such that , , and .
Henceforth, we may omit the first two parameters of for conciseness.
6.2 Theorem statement
Theorem 6.3 (Blakers–Massey theorem).
Let and . Let . Suppose the type is -connected for each , and the type is -connected for each . Then for each and each , the map
is -connected.
Let be fixed. We also fix . By the definition of connectedness, the claim is equivalent to
Similar to Section 4.4, in order to apply path induction on , we need to generalise the right-hand side of to an arbitrary term of . Therefore, we claim the following lemma, which would imply Theorem 6.3 by setting to :
Lemma 6.4.
There exists a dependent type
such that
-
•
, and
-
•
for any and any , is contractible.
Additionally, we need the following relaxation:
Lemma 6.5.
To prove Lemma 6.4, we may assume that there exist some fixed and .
Proof.
By assumption, is -connected, so is contractible and thus inhabited. The proposition that is contractible is a mere proposition, and thus is a fortiori an -type as . Thus, by truncation induction, we may assume the inhabitant of is of the form for some and . ∎
In summary, our goal is to prove Lemma 6.4 while assuming some fixed and . This is broken into the definition of and its contractibility in the following two sections.
6.3 Definition of
We begin with the wedge connectivity lemma. For two pointed types and , there is a canonical map given by and . By Proposition 4.31 and the induction principle of wedge sums, the next lemma is an equivalent way to state that if is -connected and is -connected, then is -connected.
Lemma 6.6 (Wedge connectivity lemma, [Uni13, Lemma 8.6.2]).
Suppose and are - and -connected pointed types, respectively, with 111This lemma works for or although [Uni13, Lemma 3.1.3] requires [Hou17].. Then given
-
•
,
-
•
,
-
•
and
-
•
,
there is with homotopies and such that .
In shorter terms, to define the dependent function , it suffices to consider respectively the cases when coincides with by giving and when coincides with by giving , and then ensure the consistency when both cases happen simultaneously by giving .
The following is an application of the wedge connectivity lemma, used in defining .
Corollary 6.7.
Let and . Let , and . Define type and type . Assume that is -connected and is -connected. Define by
where and . Then there is a dependent function .
Proof.
is pointed by and is pointed by . By Lemma 6.6, to obtain , it suffices to consider the following cases:
(1) When coincides with , for each , we need a term of the type
which is inhabited by . (Here, is the coherence path .) (2) When coincides with , for each , we need to a term of the type
which is inhabited by .
(3) Choices from (1) and (2) are consistent by coherence operations. ∎
We are now ready to define .
Proposition 6.8.
Proof.
See Figure 6.2. When is , for , as mentioned earlier,
When is , for ,
When ‘varies along’ , we need to find a path
By function extensionality, it suffices to find for every , a path
Starting from the left-hand side, we have a path
where the first equality is by Lemma 3.14 and the second by Lemma 3.11 and Lemma 3.15. Therefore, by univalence, it remains to find a function
(6.1) |
and show that is an equivalence. By the definition of and and truncation induction, it suffices to find for each with , a term of . By coherence operations, we then eliminate , and it suffices to find for each , a term of
We now invoke Corollary 6.7 (by using which the order of the variables is effectively rearranged), and the desired term is given by . We can similarly define the function on the other direction and show that the two functions are quasi-inverses, but the details are omitted. ∎
6.4 Contractibility of
We now show that for any and any , is contractible, by first giving its centre of contraction.
Proposition 6.9.
Proof.
By path induction, it suffices to assume and . Then
which is inhabited by . Then is explicitly given by
by Corollary 4.6, where by Lemma 3.15 and
is the uncurried form of . ∎
Before we proceed, the next lemma allows us to extract some valuable information from the hard-earned definition of for calculating a certain function.
Lemma 6.10 ([Uni13, Lemma 8.6.10]).
Let , , , , , and . Denote . Also denote the uncurried form of as . Then the function
where is an obvious path, is equal to the equivalence obtained by applying to the composite of paths
where the first equality is by definition and by Lemma 3.11, the second is by Lemma 3.14, and the third is by .
We apply Lemma 6.10 with , , , , , , and . Then by Lemma 3.15. Then Lemma 6.10 says that to calculate the function
the major work lies in recovering our definition of , which is a part of the definition of . Noticing that the type of matches the type of (Function 6.1), to calculate , we can utilise our definition of , with replaced with . The definition of relies on Corollary 6.7 and we know explicitly the output under the two special cases in the proof of Corollary 6.7. In particular, suppose that coincides with and further coincides with . By case (2) in the proof of Corollary 6.7 and the full statement of Lemma 6.6, gives the term (up to a path), and thus so does and .
Finally, we show that each is contractible, written as:
Proposition 6.11.
Remark 6.12.
It is tempting to assume by path induction that is and is . The desired path is in and thus is an -type (and a fortiori an -type), so by truncation induction, we assume that is , where and . By the definition of , we need a path Now we can only attempt to prove , which is unattainable, as only indicates .
Proof.
By Corollary 4.4, it suffices to assume that is and is . We need a re-generalisation for the ‘s and claim the following:
By truncation induction, it suffices to assume is for some and . Noticing that since is a free point in the type , we use path induction on to assume that is and is . It then remains to show that, for any and any , we have
By the definition of , the left-hand side is
This is precisely the case we discussed after Lemma 6.10 (since now , and coincide with , , ), thus it is indeed equal to . ∎
We have hence completed the proof of Lemma 6.4 and thus Theorem 6.3.
6.5 Consequences
The contractibility of in fact implies more than Theorem 6.3, which only uses the case . Our definition for the other case, , immediately implies the following:
Proposition 6.13.
Proof.
For any , matches our definition for , which is contractible by 6.11. ∎
Curiously, it is this dual case, instead of Theorem 6.3, that directly gives rise to the Freudenthal suspension theorem as below.
Theorem 6.14 (Freudenthal suspension theorem).
Let be an -connected type with and . Then the map , given by , is -connected.
Proof.
We rewrite the suspension using the generalised pushout. The suspension is the pushout of the diagram , so , and are two trivial maps. Then in this case, is equivalent to and under this equivalence the map is identified with . The connectivity assumptions are easily seen to be satisfied, so by Proposition 6.13, the map is -connected. Hence, for any , the fibre is -connected. Notice that by the definition of fibres and coherence operations,
so is also -connected. ∎
Corollary 6.15 (Freudenthal equivalence, [Uni13, Corollary 8.6.14]).
Suppose that type is -connected and pointed with . Then .
Proof.
By Theorem 6.14 and Lemma 4.32.∎
We can now complete Table 1.1 as promised.
Corollary 6.16 (Stability for spheres).
If , then .
Proof.
By Example 4.26, is -connected. By Corollary 6.15 and by the definition of -spheres, . Since , by Lemma 4.14, . We then calculate
where we use Proposition 4.13 twice to exchange truncation and . ∎
Corollary 6.17.
for .
Proof.
by Corollary 4.21 and by Corollary 5.22. When , , so by induction on and Corollary 6.16, the result follows. ∎
Corollary 6.18.
.
Proof.
by Corollary 5.22 and Corollary 6.17. ∎
Chapter 7 Conclusion and Further Discussions
Result | Location(s) |
---|---|
Homotopy groups of -spheres in Table 1.1 | Corollaries 4.21, 4.29, 5.22, 6.17, 6.18 |
Fibre sequence of a pointed map | Corollary 5.8 |
Long exact sequence of a pointed map | Theorem 5.12 |
Hopf construction | Theorem 5.18 |
Hopf fibration | Theorem 5.21 |
Blakers–Massey theorem | Theorem 6.3 |
Freudenthal suspension theorem | Theorem 6.14 |
Stability for spheres | Corollary 6.16 |
The main results in this dissertation are listed in Table 7.1. We briefly evaluate our synthetic approach to homotopy theory:
-
•
Formal abstraction allows us to focus on the homotopy properties without considering the underlying set-theoretic structures. For example, unlike in classical topology, the universal cover of (see Lemma 4.20) does not involve real numbers (a helix projecting onto the circle [Hat02, Theorem 1.7]) but only a fibration with fibre , a primitive notion expressed by a dependent type.
-
•
All notions are homotopy-invariant. For example, our formation of the Blakers–Massey theorem Theorem 6.3 replaces set intersection and union in classical text [Hat02, Theorem 4.23] with pushouts, leading to a more general statement.
-
•
Proof relevance has enabled us to effectively reuse pieces of previous results, such as the application of Lemma 6.10 in Section 6.4.
-
•
The rule-based nature of type theory has rendered many proofs routine applications of induction principles and existing lemmas. Whilst this allows for mechanised reasoning, occasionally proofs lack geometric intuition and appear overly technical. Nevertheless, there are methods, such as [Uni13, Section 8.1.5] for ) and [HFLL16, Section 4] for the Blakers–Massey theorem, that offer easier homotopy-theoretic interpretations.
These further topics are related to our work:
- •
-
•
The Blakers–Massey theorem is used in proving that there is a natural number such that [Bru16, Proposition 3.4.4].
- •
We also state some speculations based on current work:
-
•
Chapter 6 has left the proofs on coherence operations unattended. As remarked by [Bru16, p. 23], in general, these proofs can be done via recursive application of path inductions. Here we speculate that there may be automated methods or some ‘meta-theorem’ that reduce the need to spell out these proofs.
-
•
Our diagrams, combined with the topological interpretations of HoTT, suggest the potential for a systematic approach to HoTT-based graphical reasoning. For example, proof assistants may be equipped with a graphical interface, where the application of rules is visualised through shape transformations. This could potentially offer more intuition and simplify mechanised reasoning.
Appendix A The ‘hub and spoke’ construction of -truncations
Definition A.1.
Let and . Define , the -truncation of , as the higher inductive type generated by two point constructors and and a path constructor
The rationale behind and (‘hub and spoke’) in this definition will be clear in the following proof, visualised by Figure A.1.
Lemma A.2 ([Uni13, Lemma 7.3.1]).
For any and , is an -type.
Proof.
Take any and let be pointed by . By Theorem 4.8, it suffices to prove that is contractible. By a corollary of the ‘suspension-loop adjunction’ [Uni13, Lemma 6.5.4], it then suffices to prove that the type is contractible. The constant function is a pointed map of type witnessed by , and we claim it to be the centre of contraction, i.e., for any pointed function witnessed by , we have . By Proposition 3.21, it suffices to find a path such that . Using function extensionality, to find , it suffices to find a path for any . Now using the constructors we have a point along with two paths and . Thus the desired path is given by
Finally, by Lemma 3.13, is equivalent to , which holds by the definition of and coherence operations. ∎
References
- [ABFJ20] Mathieu Anel, Georg Biedermann, Eric Finster, and André Joyal. A generalized Blakers–Massey theorem. Journal of Topology, 13(4):1521–1553, 2020.
- [AW09] Steve Awodey and Michael A. Warren. Homotopy theoretic models of identity types. Mathematical Proceedings of the Cambridge Philosophical Society, 146:45–55, 2009.
- [BH18] Ulrik Buchholtz and Kuen-Bang (Favonia) Hou. Cellular cohomology in Homotopy Type Theory. In Proceedings of the 33rd Annual ACM/IEEE Symposium on Logic in Computer Science, pages 521–529. ACM, July 2018.
- [BR16] Ulrik Buchholtz and Egbert Rijke. The Cayley–Dickson construction in Homotopy Type Theory, 2016.
- [BR17] Ulrik Buchholtz and Egbert Rijke. The real projective spaces in Homotopy Type Theory. In 2017 32nd Annual ACM/IEEE Symposium on Logic in Computer Science (LICS), pages 1–8, June 2017.
- [BR21] Ulrik Buchholtz and Egbert Rijke. The long exact sequence of homotopy -groups, 2021.
- [Bru16] Guillaume Brunerie. On the Homotopy Groups of Spheres in Homotopy Type Theory. PhD thesis, Université de Nice, June 2016.
- [Car22] David Martínez Carpena. Semantics of Homotopy Type Theory, November 2022.
- [Cav15] Evan Cavallo. Synthetic cohomology in Homotopy Type Theory, December 2015.
- [Esc22] Martín Hötzel Escardó. Introduction to univalent foundations of mathematics with Agda, 2022.
- [Gra18] Robert Graham. Synthetic homology in Homotopy Type Theory, December 2018.
- [Hat02] Allen Hatcher. Algebraic Topology. Cambridge University Press, 2002.
- [HFLL16] Kuen-Bang (Favonia) Hou, Eric Finster, Dan Licata, and Peter LeFanu Lumsdaine. A mechanization of the Blakers–Massey connectivity theorem in Homotopy Type Theory, 2016.
- [Hou17] Kuen-Bang (Favonia) Hou. Higher-Dimensional Types in the Mechanization of Homotopy Theory. PhD thesis, Carnegie Mellon University, February 2017.
- [Lum11] Peter LeFanu Lumsdaine. Higher inductive types: a tour of the menagerie, April 2011. Blog post.
- [ML75] Per Martin-Löf. An intuitionistic theory of types: Predicative part. In H. E. Rose and J. C. Shepherdson, editors, Logic Colloquium ’73, volume 80, pages 73–118. Elsevier, 1975.
- [Rez15] Charles Rezk. Proof of the Blakers–Massey theorem, January 2015.
- [Rij22] Egbert Rijke. Introduction to Homotopy Type Theory, 2022.
- [Uni13] The Univalent Foundations Program. Homotopy Type Theory: Univalent Foundations of Mathematics. https://homotopytypetheory.org/book, Institute for Advanced Study, 2013.