<<

Florida State University Libraries

Electronic Theses, Treatises and Dissertations The Graduate School

2018 The 1-Type of Algebraic K-Theory as a Multifunctor Yaineli Valdes

Follow this and additional works at the DigiNole: FSU's Digital Repository. For more information, please contact [email protected] FLORIDA STATE UNIVERSITY

COLLEGE OF ARTS AND SCIENCES

THE 1-TYPE OF ALGEBRAIC K-THEORY AS A MULTIFUNCTOR

By

YAINELI VALDES

A Dissertation submitted to the Department of in partial fulfillment of the requirements for the degree of Doctor of Philosophy

2018

Copyright © 2018 Yaineli Valdes. All Rights Reserved. Yaineli Valdes defended this dissertation on April 10, 2018. The members of the supervisory committee were:

Ettore Aldrovandi Professor Directing Dissertation

John Rawling University Representative

Amod Agashe Committee Member

Paolo Aluffi Committee Member

Kate Petersen Committee Member

Mark van Hoeij Committee Member

The Graduate School has verified and approved the above-named committee members, and certifies that the dissertation has been approved in accordance with university requirements.

ii ACKNOWLEDGMENTS

The completion of this research was only made possible through the assistance and support from so many throughout my educational career. I am very thankful to my committee members for their dedication as they not only served in my supervisory committee for my dissertation, but as well for my candidacy exam. I also had the immense pleasure of taking some of their courses as a graduate student and I owe much of my research achievements to their great teaching. I am especially thankful to my major professor Dr. Aldrovandi who has been my insight, professor, and guide throughout this research project, and has eloquently taught me so much of what I know today. This would definitely have not been possible if it were not for his support throughout these last several years. I am proud of the work Dr. Aldrovandi and I have completed together and am so grateful to him and everyone else that invested even a minute of their life to help me with this mission. I cannot leave out my family who have always been there for me unconditionally, especially my husband Alejandro who has supported and encouraged me to help me realize my dream. Finally, and most importantly, to my Lord and Savior, the One I owe everything to, for His unconditional love. I am beyond blessed and very hopeful for what the future holds in store!

iii TABLE OF CONTENTS

Abstract ...... vi

1 Introduction 1

2 Preliminaries 3

3 (Closed, Symmetric) Multicategories 6 3.1 Definitions ...... 6 3.2 Examples and Properties ...... 8

4 Waldhausen Categories 9 4.1 Definitions and Examples ...... 9 4.2 The closed structure on Wald ...... 11 4.3 The S∗ Construction ...... 12 4.4 K-theory for a Waldhausen ...... 13 4.5 K-theory Simplices from Total Simplicial ...... 15 4.6 Classical K-theory of a Ring ...... 18

5 Picard 19 5.1 Definitions and Examples ...... 19 5.2 The closed structure on Pic ...... 23

6 Stable Quadratic Modules 25 6.1 Definitions and Properties ...... 25 6.2 An Algebraic Model for the 1-type of K(W) as a Stable Quadratic Module . . . . . 26

7 Crossed Modules 29 7.1 Definitions and Examples ...... 29 7.2 Properties ...... 30

8 Determinant 32 8.1 Definitions and Examples ...... 32 8.2 Universal Determinant Functor for a Waldhausen Category ...... 34

9 P1K as a Multifunctor 36 9.1 Proof of Main Theorem ...... 40

iv Appendix A Computations for Proofs 44 A.1 ω defines a on morphisms ...... 44 A.2 ω is a well-defined symmetric, ...... 50

Bibliography ...... 52 Biographical Sketch ...... 54

v ABSTRACT

It is known that the category of Waldhausen categories is a closed symmetric multicategory and algebraic K-theory is a functor from the category of Waldhausen categories to the category of spectra. By assigning to any Waldhausen category the fundamental of the 1-type of its K-theory , we get a functor from the category of Waldhausen categories to the category of Picard groupoids, since stable 1-types are classified by Picard groupoids. We prove that this functor is a multifunctor to a corresponding multicategory of Picard groupoids.

vi CHAPTER 1

INTRODUCTION

Given any Waldhausen category, W, we can construct its K-theory spectrum, K(W), and for any of Waldhausen categories W → V, we get a morphism of spectra K(W) → K(V). It is well known that given an infinite loop space X its 1-type can be realized as a Picard groupoid,

P1(X) [5]. Therefore, realizing the 1-type of the K-theory of a Waldhausen category as a Picard groupoid gives us a functor from the category of Waldhausen categories to the category of Picard groupoids:

Wald −−−→P1K Pic

W 7→ P1(K(W))

We also know that although the category of Waldhausen categories does not have a symmetric monoidal structure, it does have a (closed) symmetric multicategorical structure. (Blumberg and Mandell gave a proof of this in [4], as well as Zakharevich in [2].) We call the multifunctors of Waldhausen categories multiexact . Similarly, we can show that Picard groupoids also form a (closed) symmetric multicategory when equipped with multimonoidal functors. Therefore, since P1K is a functor between categories underlying multicategories, a natural question to ask is if it extends to a multifunctor. Our main result answers this question in the affirmative, namely we prove the following theorem:

Theorem 9.1. Wald −−−→P1K Pic extends to a multifunctor.

To be clear, if a map between multicategories is a just a functor between their underlying categories, then we say it “extends to a multifunctor” if there is a multifunctor between the multi- categories such that when seen as a functor between the underlying categories, it agrees with the original one. This is defined in Definition 3.3. This theorem is based on two main ingredients: (1) the existence of a universal determinant functor

detW : wW → P1K(W)

1 that satisfies certain properties, described in detail in section 8.1, where wW is the subcategory of weak equivalences of W; and (2) the closed multicategorical structures on Wald and Pic. The former is known thanks to [2], [4] and we give some attention to the latter.

With these premises, our key ingredient is that the functor P1K “respects the closed structures” in the following sense:

P1K(Wald(W, V)) → Pic(P1K(W),P1K(V)) which allows us to prove the theorem. An application of this result is to use this multifunctor to study the algebraic structures on the crossed modules, which model Picard groupoids, coming from multiexact maps of Waldhausen categories. Aldrovandi shows there is an equivalence between bimonoidal functors and biexten- sions of crossed modules, which in fact extends to multimonoidal functors and multiextensions [3]. Therefore, an application of our results is to study the multiextensions corresponding to the multimonoidal functors of Picard groupoids under the multifunctor, PD·.

∗ ∗ ∗

This thesis is organized as follows: we start with reminding the reader of relevant algebraic structures such as Waldhausen categories, Picard groupoids, and stable quadratic modules. We cover universal determinant functors as they play a major role in proving the main result. We also recall the K-theory construction for a Waldhausen category by forming the S-construction. Finally, we end with stating and proving two propositions that allow us to prove the main theorem. Certain cumbersome but necessary computations are in the appendix.

2 CHAPTER 2

PRELIMINARIES

Definition 2.1. Let [n] denote the finite, nonempty, totally ordered set {0 < 1 < ··· < n} considered as a category. This means the objects are 0, 1, . . . , n and  ∅ j > k  Hom[n](j, k) = idj j = k j → k j < k

Definition 2.2. For any two categories C and D, we can form another category Fun(C, D) called the whose objects are functors F : C → D and whose morphism are natural transformations between these functors.

Definition 2.3. The arrow category of some category C is Ar C := Fun([1], C). This is truly the category of arrows in C because any object in Ar C is a functor from [1] → C and such a functor just chooses two objects in C and a morphism between them since there is only two objects in [1] and one non-identity morphism.

One particular arrow category used in this thesis is Ar[n].

Definition 2.4. The ordinal category ∆ is the category whose objects are the ordered sets [n] and whose morphisms are order preserving maps. Specifically, f :[m] → [n] is a morphism in ∆ if whenever i < j in [m], then f(i) ≤ f(j) in [n].

There are special maps in ∆ called the coface and codegeneracy maps. For n ≥ 0, the n + 1 coface maps are the injections di :[n − 1] → [n], which send j 7→ j for all j < i and j 7→ j + 1 for j ≥ i. For n ≥ 0, the n + 1 codegeneracy maps are the surjections si :[n + 1] → [n], which send j 7→ j for all j ≤ i and j 7→ j − 1 for j > i. These morphisms satisfy several relations and it can be verified that every morphism in ∆ is a composite of coface and codegeneracy maps. A nice reference on this is in [16].

3 Definition 2.5. A simplicial set is a contravariant functor from the ordinal category to the . More generally, for any category C, a simplicial object in C, X, is a contravariant functor from the ordinal category to C X : ∆op → C

So the data of a simplicial object in C comprises of an object in C for every natural number and face and degeneracy maps, denoted di and si respectively, which are the images of the coface and codegeneracy maps in ∆. The of [n] under this functor is an object Xn in C. If C = Set the category of sets, then Xn can be thought of as a set of simplices. The category of simplicial sets will be denoted sSet.

Definition 2.6. A simplex x ∈ Xn is called nondegenerate if x 6= si(y) for any y ∈ Xn−1 for all i.

n n+1 Definition 2.7. The topological n-simplex ∆ is a subspace of R whose underlying set is n n n+1 X ∆ := {(x0, . . . , xn) ∈ R | xi = 1 and ∀i, xi ≥ 0} i=0

n+1 whose topology is the subspace topology induced from the standard topology in R . This is homeomorphic to the n-dimensional ball. The i-th n-face is the inclusion

Di : ∆n → ∆n+1

(x0, . . . , xn) 7→ (x0, . . . , xi, 0, xi+1, . . . , xn) and the i-th degenerate n-simplex is the surjective map

Si : ∆n → ∆n−1

(x0, . . . , xn) 7→ (x0, . . . , xi + xi+1, . . . , xn)

Definition 2.8. The geometric realization of a simplicial set X is

∞ G n |X| := ∆ × Xn/ ∼ n=0 where ∼ is the equivalence relation generated by the relations (p, di(x)) ∼ (Di(p), x) and the n relations (p, si(y)) ∼ (Si(p), y) for all p ∈ ∆ , x ∈ Xn+1, and y ∈ Xn−1.

4 The topological space |X|, as Milnor showed in [17], is a CW-complex with a topological n- simplex for every nondegenerate simplex in Xn.

Definition 2.9. For a topological space X, the loop space of X at some chosen basepoint x ∈ X is the space of maps Map((S1, ∗), (X, x))

Definition 2.10. A pointed category is a category C with a zero object ∗. This means for every x ∈ Ob C, there are unique maps from and to the zero object: ∗ → x, x → ∗

Definition 2.11. A wide subcategory of a category C is a subcategory whose objects are all the objects of C.

5 CHAPTER 3

(CLOSED, SYMMETRIC) MULTICATEGORIES

Multicategories generalize categories in that they allow for morphisms with multiple objects in the source. If one thinks of morphisms in a category as functions, then morphisms in a multicategory can be thought of as multi- functions. A functor between multicategories, called a multifunctor, should preserve these multiple arity morphisms. We formally define these concepts now.

3.1 Definitions

Definition 3.1. A symmetric multicategory, M consists of a of objects, Ob M, and for any k ≥ 0 and A1,...,Ak,B ∈ Ob M, a set of k-morphisms, HomM(A1,...,Ak; B).

There is a distinguished identity morphism 1A ∈ HomM(A; A) for all A ∈ Ob M. There is also a right action of the symmetric Σk on the set of k-morphisms:

∗ σ : HomM(A1,...,Ak; B) → HomM(Aσ(1),...,Aσ(k); B)

Finally, there is a composition law

k Y ◦ : HomM(B1,...,Bk; C) × HomM(Ai1,...,Aili ; Bi) → HomM(A11,...,Aklk ; C) i=1 subject to the following axioms: associativity

j ∀ θi ∈ HomM(Aij1,...,Aijmij ; Aij), θi ∈ HomM(Ai1,...,Aiki ; Bi)

and

∀θ ∈ HomM(B1,...,Bn; C) we have

1 k1 1 kn 1 k1 1 kn θ ◦ (θ1 ◦ (θ1, . . . , θ1 ), . . . , θn ◦ (θn, . . . , θn )) = (θ ◦ (θ1, . . . , θn)) ◦ (θ1, . . . , θ1 , . . . , θn, . . . , θn ) identity for every θ ∈ HomM(A1,...,Ak; B), we have

θ ◦ (1A1 ,..., 1Ak ) = θ = 1B ◦ θ

6 compatibility ∀σ ∈ Σk, ∀τi ∈ Σli for 1 ≤ i ≤ k and

∀θ ∈ HomM(A1,...,Ak; B), ∀θi ∈ HomM(Ai1,...,Aili ; Bi)

we have ∗ ∗ σ (θ ◦ (θ1, . . . , θk)) = σ (θ) ◦ (θσ(1), . . . , θσ(k))

and ∗ ∗ ∗ θ ◦ (τ1 (θ1), . . . τk (θk)) = (τ1 ⊕ · · · ⊕ τk) (θ ◦ (θ1, . . . , θk))

Definition 3.2. A multifunctor F between symmetric multicategories M, N is a map sending each object in M to an object in N and ∀k each k-morphism f : A1 ×· · ·×Ak → B to a k-morphism

F (f): F (A1) × · · · × F (Ak) → F (B) that preserves the units, the σk action on the collection of morphisms, and the composition of morphisms.

There is a from the category of (symmetric) multicategories to the category of categories by simply forgetting the set of k−morphisms for k > 1. The image of any multicategory M under this functor is called the underlying category of M.

Definition 3.3. Let M, N be symmetric multicategories and F : M → N be a functor between their underlying categories. Then we say F extends to a multifunctor if there is a multifunctor F 0 : M → N so that its image under the forgetful functor from the category of multicategories to the category of categories is F .

Definition 3.4. A symmetric multicategory M is closed if ∀A1,...,Ak,B ∈ M, there is an internal hom object, M(A1,...,Ak; B), and an evaluation map

evA1,...,Ak;B ∈ HomM(A1,...,Ak, M(A1,...,Ak; B); B) such that the following holds:

i) ∀A1,...,Ak,C1,...,Cl,B ∈ M there is a bijection

=∼ HomM(C1,...,Cl; M(A1,...,Ak; B)) −→ HomM(A1,...,Ak,C1,...,Cl; B)

which depends on these objects and is defined by sending

f ∈ HomM(C1,...,Cl; M(A1,...,Ak; B))

to the composite

evA1,...,Ak;B ◦(idA1 ,..., idAk , f)

7 ii) The following diagram commutes ∀(σ, τ) ∈ Σk × Σl: =∼ HomM(C1,...,Cl; M(A1,...,Ak; B)) HomM(A1,...,Ak,C1,...,Cl; B)

f7→σ◦(f◦τ) g7→g·(σ×τ) =∼ HomM(Cτ(1),...,Cτ(l); M(Aσ(1),...,Aσ(k); B)) HomM(Aσ(1),...,Aσ(k),Cτ(1),...,Cτ(l); B)

Definition 3.5. If M, N are closed multicategories and F : M → N is a functor between their underlying categories then we say F respects the closed structure if the following morphism ϕA,B exists for any objects A, B of M:

ϕA,B : F (M(A, B)) → N (F (A),F (B))

3.2 Examples and Properties

Example 3.6. We can view any (symmetric) (definition 5.5), (M, ⊗), as a (sym- metric) multicategory by defining the k-morphisms to be HomM(A1,...,Ak; B) := HomM(A1 ⊗

· · · ⊗ Ak,B).

If M is non-strict, then we have to be careful about how we choose A1 ⊗ · · · ⊗ Ak (see, for instance, [10]). One way to deal with this ambiguity is to choose a certain bracketing; for example, for n = 5 choose (((A1⊗A2)⊗A3)⊗A4)⊗A5. In this case, the multicategory is called a representable multicategory. We can similarly view any category with finite products as a multicategory.

Example 3.7. An is a multicategory with one object.

Remark 3.8. Not every multicategory is a representable multicategory. In fact, one of the main categories we will deal with in this paper, the category of Waldhausen categories, is an example of a non-representable multicategory (proof given in [2]). However, there is a nice relationship between multicategories and monoidal categories by . For any (symmetric) multicategory M, even non-representable ones, we can construct a (symmetric) monoidal category whose objects (respectively, morphisms) are lists of objects (respectively, morphisms) in M where the monoidal structure is given by concatenation. This functor from the category of multicategories to the category of monoidal categories is left adjoint to the representable functor that takes any monoidal category to its representable multicategory.

8 CHAPTER 4

WALDHAUSEN CATEGORIES

Waldhausen defined what we now call Waldhausen categories. These generalize exact categories and are the most general input for K-theory which agrees with the classical K-theory notions. We will discuss this in more detail later, but we begin with formally defining Waldhausen categories.

4.1 Definitions and Examples

Definition 4.1. A Waldhausen category, W, is a pointed category with two wide subcategories, ∼ 0 wW (whose morphisms A → A are called weak equivalences) and coW (whose morphisms A  B are called cofibrations) subject to the following axioms:

i) All isomorphisms are cofibrations and weak equivalences

ii) the unique map from the zero object to any object A is a cofibration: ∗  A ∈ coW iii) Push-outs over cofibrations exist, meaning every such diagram can be completed with the bottom morphism being a cofibration:

A B

C B F C A iv) For any such diagram: C A B ∼ ∼ ∼ C0 A0 B0

the unique map from B F C ∼ B0 F C0 is a weak equivalence. A A0

Remark 4.2. 1. Every Waldhausen category has coproducts by pushing out along the unique morphism from the zero object to any object, which is required to be a cofibration by axiom

9 iii). More specifically, we have A F B = A F B by: ∗ ∗ B

A A F B ∗

2. Waldhausen categories have cofibration for every cofibration A  B: A B

∗ ∗ F B A Letting B/A = ∗ F B, then A A  B  B/A is a cofibration .

Definition 4.3. An exact functor of Waldhausen categories, F : W → V, is a functor that preserves the zero object, weak equivalences, cofibrations, and push-out diagrams.

Definition 4.4. The southern arrow condition of a commutative diagram of cofibrations A B

C D is that the unique map from B F C to D is also a cofibration: A A B

B F C A ∃!

C D

Definition 4.5. A biexact functorf of Waldhausen categories, F : W1 ×W2 → V is a functor which 0 0 is exact in each factor and for any pair of cofibrations, f1 : A  A and f2 : B  B in W1 and

W2 respectively, we can form the cube, [(f1, f2)]F : F (f B) F (A, B) 1 F (A0,B)

0 F (A f2) F (A f2) F (f B0) F (A, B0) 1 F (A0,B0)

10 This cube should satisfy the southern arrow condition.

Definition 4.6. A multiexact functor generalizes a biexact functor; namely, it should be exact in each factor and for each cofibration in each factor, there is a n-cube that should satisfy a generalized version of the southern arrow condition above. Full details on this are in [2].

The (multi)category of Waldhausen categories will be denoted Wald.

Example 4.7. The category of pointed finite sets is a Waldhausen category by letting the weak equivalences be the bijections and the cofibrations be the inclusions. We require pointed sets so that there is a zero object, namely the singletons.

Example 4.8. A typical example of a Waldhausen category is any abelian category, or more gen- erally, any Quillen exact category where one views the admissible monomorphisms as cofibrations and the isomorphisms as weak equivalences. The category of finitely generated projective modules over a ring is a Quillen exact category and therefore Waldhausen. (Notice, this is not abelian since kernels may not exist, for example, if the ring is non-Noetherian.)

Example 4.9. The typical non-exact example of a Waldhausen category is the example motivating Waldhausen in [5]: the category R(X) of retractive spaces over X, namely the category of spaces having X as a retract. The cofibrations here are the typical cofibration of spaces and the weak equivalences are weak homotopy equivalences (or classes of maps that induce isomorphisms in any homology theory). Technically, we have to take the objects in R(X) to be spaces Y that are finite CW complexes relative to X. This way, the inclusion from the zero object, X, is a cofibration as it should be. More generally, one can view a full subcategory of cofibrant objects in a pointed model category as a Waldhausen category.

4.2 The closed structure on Wald

The closed structure on Wald is given by defining the internal hom

Wald(W1,..., Wk; V) to be the Waldhausen category whose objects are k-exact functors F : W1×· · ·×Wk → V and whose morphisms are natural transformations between k-exact functors. The weak equivalences are the

11 natural weak equivalences between k-exact functors (which means they induce weak equivalences for all objects) and the cofibrations are the natural transformations α : F → G such that for all cofibrations fi ∈ coWi (1 ≤ i ≤ k), we can construct a k +1 cube which should satisfy a generalized version of the southern arrow condition. Details for the general case can be found in [2]; the case for k = 1 is needed for the proof of Proposition 9.3 so we define it now. The internal Hom object Wald(W; V) as a Waldhausen category has as its objects exact functors ϕ : W → V and as its morphisms natural transformations between such exact functors. The weak equivalences are defined to be the natural weak equivalences and the cofibrations are the natural 0 transformations, α : ϕ0 =⇒ ϕ1 : W −→ V, such that for every cofibration f : X  X in W, all the maps along with the southern arrow of the following diagram are cofibrations:

ϕ0(f) 0 ϕ0(X) ϕ0(X )

αX αX0 ϕ1(f) 0 ϕ1(X) ϕ1(X )

(Zakharevich proves that this is well-defined in [2].)

4.3 The S∗ Construction

Waldhausen’s S∗-construction, which generalizes Quillen’s Q-construction for exact categories, was defined in [5]. This is used to define K-theory for a Waldhausen category.

Definition 4.10. Let W be a Waldhausen category. The S∗-construction on W, denoted Sn(W), is the category whose objects are functors:

Fn : Ar[n] −→ W

(i, j) 7→ Ai,j ∈ Ob(W) such that

i) Fn(j, j) = 0

ii) Aij  Aik  Ajk is a cofibration sequence ∀i ≤ j ≤ k

12 The idea is for any A ∈ Sn(W), A := A01  A02  ...  A0n is a sequence of cofibrations with choice of sub-quotients Aij = A0j/A0i.

An arrow in Sn(W) is a diagram of the form

A01 A02 ... A0n

α1 α2 αn (4.1)

0 0 0 A01 A02 ... A0n

We let Sn(W) be a category with cofibrations and weak equivalences as follows. The diagram (4.1) above is:

1. a cofibration if ∀i ∈ {1, . . . , n − 1}, each αi is a cofibration in W and each unique map

0 G 0 A0i A0(i+1)  A0(i+1) A0i is a cofibration in W;

2. a weak equivalence if ∀i ∈ {1, . . . , n}, each αi is a weak-equivalence in W.

It is easily verified that Sn(W), with the above notions of cofibrations and weak equivalences, becomes a Waldhausen category.

It is also clear by the definition of Sn(W) that

op S∗(W) : ∆ −→ Wald

[n] 7→ Sn(W) is a simplicial Waldhausen category, as shown by Waldhausen in [5]. The face map di : Sn(W) →

Sn−1(W) omits A0i for i = 1, . . . , n and quotients everything by A01 for i = 0. The degeneracy id map si : Sn(W) → Sn+1(W) extends A0i to A0i −→ A0i for i = 1, . . . , n and includes ∗  A01 at beginning for i = 0.

4.4 K-theory for a Waldhausen Category

Before we can define K-theory for a Waldhausen category, we need to recall some definitions.

Definition 4.11. For any category C, we can construct its nerve which is a simplicial set

op Ner∗ C : ∆ → Set

13 where Nern C is a set of n-composable morphisms between n + 1 objects in C where the face maps th di ignore the i object Xi and replaces the sequence

Xi−1 → Xi → Xi+1 with the composition

Xi−1 → Xi+1

= and the degeneracy maps si replaces each object Xi with the identity morphism Xi −→ Xi.

Let ssSet be the category of bisimplicial sets, i.e. the category of contravariant functors ∆op → sSet. There is a Diag : ssSet → sSet by precomposing with (id, id) : ∆op → ∆op × ∆op

Therefore, the set of n-simplices of the diagonal of a bisimplicial set X∗,∗ is Xn,n. Waldhausen gave a definition of K-theory as an infinite loop space as well as a connective spectrum (spectra whose negative homotopy groups vanish). We will give both definitions here.

Definition 4.12. For any Waldhausen category W, its K-theory as an infinite loop space is defined as:

K(W) = Ω| Diag(Ner∗ wS∗W)| where | · | : sSet → Top is the geometric realization of simplicial sets and Ω is the loop space functor. We can define the K-groups as the homotopy groups of this space:

Ki(W) = πi(K(W)) = πi+1(| Diag(Ner∗ wS∗W)|)

Definition 4.13. For any Waldhausen category W, its K-theory as a connective spectrum of spaces is: ∞ ∞ K(W) = Ω | Diag(Ner∗ wS∗ W)|

14 n where S∗ is the multi-simplicial Waldhausen category obtained by iterating the S-construction to the simplicial Waldhausen category S∗. We could also define the spectrum as a connective spectrum of simplicial sets where:

∞ K(W) = Diag(wS∗ W)

n So the n-th simplicial set is Diag(wS∗ W)

This agrees with the classical definitions of K-groups as well as the higher K-theory defined for symmetric monoidal categories and Quillen exact categories. However, if one is interested in the lower K-groups K0 and K1, then computing them via this construction may be tedious and complicated. This is where Picard groupoids and determinant functors play a role since the 1-type of this K-theory space will have the same homotopy groups (hence K-groups) in dimensions 0 and 1. These 1-types are known to be realized by Picard groupoids and there are many algebraic models which allow us to construct these 1-types directly from the Waldhausen category itself.

These algebraic models are a lot nicer and easier to construct than the S∗-construction so it gives a better route for computing K0 and K1.

4.5 K-theory Simplices from Total Simplicial Set

There is another way to go from bisimplicial sets to simplicial sets, namely the functor

Tot : ssSet → sSet originally defined by Artin and Mazur in [14]. It is known that there is a natural weak equivalence between the diagonal simplicial set and the total simplicial set of any bisimplicial set. Although the diagonal functor is used to define the K-theory of a Waldhausen category, any weakly equivalent model will work just as well. In particular, the simplices and relations that Muro and Tonks [1] use to construct their model for the 1-type of the K-theory of a Waldhausen category arise from the total simplicial set, rather than the diagonal simplicial set. We remind the reader of the construction along with the face and degeneracy maps which is given in [14].

n , Y 0 i+1 (Tot X)n = Xi,n−i dvxi = dh xi+1 i=1

15 n Q where (x0, x1, . . . , xn) ∈ Xi,n−i. The face and degeneracy maps (for j = 0, . . . , n) are: i=0

j D : (Tot X)n → (Tot X)n−1

j j−1 1 j j (x0, . . . , xn) 7→ (dvx0, dv x1, . . . , dvxj−1, dhxj+1, . . . , dhxn) j S : (Tot X)n → (Tot X)n+1

j j−1 0 j j (x0, . . . , xn) 7→ (svx0, sv x1, . . . , svxj, shxj, . . . , shxn)

Take the bisimplicial set Ner∗ wS∗W, given by the nerve of the simplicial Waldhausen category given by the S-construction for a Waldhausen category, W. (Draw the bisimplicial set so S∗ changes horizontally.) The set of 0-simplices is the trivial set and the set of 1-simplices is the set of objects in W. The 2-simplices are ∼ 0 0 0 (∗,A → A ,A  B ) which can be drawn as A ∼ A0 B0 0 0 0 whose face maps are D0 = B /A , D1 = B , and D2 = A.

The elements of the third simplicial set Tot(Ner∗ w S∗W)3 are

A0 B0 ! ∼ 0 ∼ ∗,A → A → A”, ∼ ∼ ,A”  B”  C” A” B”

Instead, we can draw this 3-simplex as:

A ∼ A0 B0 ∼ ∼ A” B” C”

16 And its faces are:

0 0 B /A A 0 2 D : ∼ D : ∼ B”/A” C”/A” A” C” B0 A 1 3 D : ∼ D : ∼ B” C” A0 B0

Similarly, to go one step further, we can draw a 4-simplex and define its face maps as follows:

A ∼ A0 B0 ∼ ∼ A” B” C” ∼ ∼ ∼ A”0 B”0 C”0 E”0

The faces of this 4-simplex are:

0 0 B /A A ∼ ∼ 0 2 D : B”/A” C”/A” D : A” C”

∼ ∼ ∼ ∼ B”0/A”0 C”0/A”0 E”0/A”0 A”0 C”0 E”0 B0 A ∼ ∼ 1 3 D : B” C” D : A0 B0 ∼ ∼ ∼ ∼ B”0 C”0 E”0 A”0 B”0 E”0 A ∼ 4 D : A0 B0 ∼ ∼ A” B” C”

It is easy to see how this generalizes for higher simplices.

17 4.6 Classical K-theory of a Ring

Classical K-theory more or less began with Grothendieck’s K-group for a ring or topological space; this is what we now call K0. Other lower K-groups were then defined, such as Milnor’s

Whitehead group K1, until Quillen defined higher K-theory as a connective spectrum of spaces (or simplicial sets) whose homotopy groups computed the classical lower K-groups and more; this construction was done for Quillen exact categories. Waldhausen generalized this construction for

Waldhausen categories which is what we have described above. Now we can compute K0 of a ring using Grothendieck’s definition or Waldhausen’s generalization.

Using Grothendieck’s method, we construct the abelian group K0R for some ring R by taking fg ProjR to be the set of isomorphism classes of finitely generated projective modules; this is actually an abelian monoid where the monoidal structure is given by:

[P ] + [Q] = [P ⊕ Q], 0 = [0R]

(In fact, this is a semi-ring since tensoring two finitely generated projective modules over R is still a finitely generated projective module.) Now we can define K0R:

fg K0R = Gr(ProjR ) where Gr is the group completion of a module. Recall that this group satisfies an obvious and is constructed by forming the free group on the monoid as a set and identifying the new monoidal structure from the free group construction with the original monoidal structure of the monoid.

Alternatively, we can use Waldhausen’s K-theory construction to define K0R. Abusing notation fg a bit, take ProjR now to be the category whose objects are finitely generated projective modules over R. This is clearly an exact category, therefore it is Waldhausen. Now, we can take K0R fg to be π0K(ProjR ), which is equivalently the fundamental group of the geometric realization of fg the diagonal of the bisimplicial set formed by the nerve of the wS∗ ProjR . These two groups are isomorphic, although it may not be obvious right now. We will use the 1-type to get a better sense of what K0R is in terms of this K-theory construction. We will continue this discussion in example 8.5 in the section of Universal Determinant Functors after we have carefully constructed a model for the 1-type.

18 CHAPTER 5

PICARD GROUPOIDS

We will now define the structures that make-up a Picard groupoid. Recall that a groupoid is a category for which every morphism is an isomorphism. The Picard structure means it has a braided, symmetric, group-like structure. In particular, this means Picard categories have a monoidal structure, so we start with formally defining monoidal categories. A nice reference for these notions is [11].

5.1 Definitions and Examples

Definition 5.1. A monoidal category is a category M equipped with the following data: a) an object 1 ∈ Ob M, called the “unit object” b) a functor: ⊗ : M × M → M called the “ product”

=∼ c) a natural isomorphism: αx,y,z :(x ⊗ y) ⊗ z −→ x ⊗ (y ⊗ z) called the “associator”

=∼ d) a natural isomorphism: lx : x ⊗ 1 −→ x, called the “left unitor”

=∼ e) a natural isomorphism: rx : 1 ⊗ x −→ x, called the “right unitor” such that the following diagrams commute: pentagon identity

(w ⊗ x) ⊗ (y ⊗ z) αw⊗x,y,z αw,x,y⊗z

((w ⊗ x) ⊗ y) ⊗ z w ⊗ (x ⊗ (y ⊗ z))

αw,x,y⊗idz idw ⊗αx,y,z α (w ⊗ (x ⊗ y)) ⊗ z w,x⊗y,z w ⊗ ((x ⊗ y) ⊗ z) triangle identity α (x ⊗ 1) ⊗ y x,1,y x ⊗ (1 ⊗ y)

lx⊗idy idx ⊗ry x ⊗ y

19 Definition 5.2. A group-like groupoid G is a monoidal groupoid such that for every object x of G, the functors x ⊗ − : G → G

− ⊗ x : G → G are equivalences of categories. In particular, this means for every object x of G, we can find an “inverse” object x0, so there are isomorphisms ∼ x ⊗ x0 −→= 1

∼ x0 ⊗ x −→= 1

Remark 5.3. In any group-like category P we get the following bijection

−1 ∼ HomP (x ⊗ z ⊗ y, x) = HomP (y, z) for all objects x, y, z in P. This means that if we want to define a monoidal functor F : P → G on −1 HomP (x ⊗ z ⊗ y, x), we only need to define it on HomP (y, z). This will be useful later.

Definition 5.4. A braided monoidal category is a monoidal category equipped additionally with a natural isomorphism =∼ βx,y : x ⊗ y −→ y ⊗ x called the “braiding” such that the following diagrams commute (called the hexagon identities):

α B (x ⊗ y) ⊗ z x,y,z x ⊗ (y ⊗ z) x,y⊗z (y ⊗ z) ⊗ x

βx,y⊗idz αy,z,x α id ⊗B (y ⊗ x) ⊗ z y,x,z y ⊗ (x ⊗ z) y x,zy ⊗ (z ⊗ x) and

α−1 B x ⊗ (y ⊗ z) x,y,z (x ⊗ y) ⊗ z x⊗y,z z ⊗ (x ⊗ y)

−1 idx ⊗βy,z αz,x,y α−1 B ⊗id x ⊗ (z ⊗ y) x,z,y (x ⊗ z) ⊗ y x,z y (z ⊗ x) ⊗ y

20 Definition 5.5. A symmetric monoidal category is a braided monoidal category such that the braiding must satisfy the condition

βy,x ◦ βx,y = idx⊗y for all objects x, y in the category. A consequence of this condition is that the two hexagon identities now become one.

Definition 5.6. A Picard groupoid P is a symmetric, group-like groupoid.

Remark 5.7. 1. A nice way to think of Picard groupoids is that they categorify abelian groups.

2. One should note that the symmetric (or more generally the braided) structure and group-like structure in a Picard groupoid are independent of one another.

We denote the set of isomorphism classes of objects in P by π0P; the Picard structure gives this set an abelian group structure. The automorphisms of the identity object always form an abelian group denoted π1P. (See [11] for details.)

Definition 5.8. A monoidal functor between Picard groupoids is a functor F : P → Q such that the following exist:

=∼ a) an isomorphism in Q:  : 1Q −→ F (1P )

=∼ b) a natural isomorphism: µx,y : F (x) ⊗P F (y) −→ F (x ⊗Q y) making the following diagrams commute:

associativity

αF (x),F (y),F (z) (F (x) ⊗Q F (y)) ⊗Q F (z) F (x) ⊗Q (F (y) ⊗Q F (z))

µx,y⊗idF (z) idF (x) ⊗µy,z

F (x ⊗P y) ⊗Q F (z) F (x) ⊗Q (F (y ⊗P z))

µx⊗y,z µx,y⊗z

F (αx,y,z) F ((x ⊗P y) ⊗P z) F (x ⊗P (y ⊗P z))

unitality ⊗idF (x) 1Q ⊗Q F (x) F (1P ) ⊗Q F (x)

rF (x) µ1,x

F (x) F (1P ⊗P x) F (rx)

21 idF (x)⊗ F (x) ⊗Q 1Q F (x) ⊗Q F (1P )

lF (x) µx,1

F (x) F (x ⊗P 1P ) F (lx) symmetry BF (x),F (y) F (x) ⊗Q F (y) F (y) ⊗Q F (x)

µx,y µy,x

F (Bx,y) F (x ⊗P y) F (y ⊗P x)

We will denote the (multi)category of Picard groupoids by Pic. This is also a 2-category if we let the 2-arrows be the natural transformations between monoidal functors. We form the Ho(Pic) by identifying all naturally isomorphic monoidal functors and discarding the natural transformations.

Definition 5.9. A bimonoidal functor of Picard groupoids is a bifunctor F : P1 × P2 → Q such that: a) it is monoidal in each factor, meaning the following isomorphisms exist:

µ ,id id ,µ 0 x,x0 y 0 0 x y,y0 0 F (x, y) ⊗Q F (x , y) −−−−−→ F (x ⊗P1 x , y),F (x, y) ⊗Q F (x, y ) −−−−−→ F (x, y ⊗P2 y ) b) it has a compatibility condition between the monoidal structure in each factor, meaning the following diagram commutes

0 µ 0 ,idy ⊗µ 0 ,y (F (x, y) ⊗ F (x0y)) ⊗ (F (x, y0) ⊗ F (x0y0)) x,x x,x F (x ⊗ x0, y) ⊗ F (x ⊗ x0, y0)

idx⊗x0 ,µy,y0 0 0 BF (x0,y),F (x,y0) F (x ⊗ x , y ⊗ y )

µx,x0 ,idy⊗y0 idx,µ 0 ⊗id 0 ,µ 0 (F (x, y) ⊗ F (x, y0)) ⊗ (F (x0, y) ⊗ F (x0y0)) y,y x y⊗y F (x, y ⊗ y0) ⊗ F (x0, y ⊗ y0)

=∼ c) The two ways to form the isomorphism F (1P1 , 1P2 ) −→ 1Q by using the monoidal structure in each factor must coincide, meaning the following diagram must commute

F (1P1 −)

F (1P1 , 1P2 ) 1Q

F (− 1P2 )

22 For general n, an n-monoidal or multimonoidal functor of Picard groupoids is a multifunctor

F : P1 × · · · × Pn → Q that generalizes the notion of bimonoidal functors, meaning fixing all but two factors produces a bimonoidal functor.

Example 5.10. If X is a scheme then L(X) is the groupoid of all line bundles on X with the usual monoidal structure, ⊗.

Example 5.11. Take PicR to be the category of invertible R-modules over some commutative ring R. This means the objects are rank 1 R-modules M such that there is an R-module N with ∼ M ⊗R N = R. (These are always projective.) This is a Picard groupoid under the

⊗R. Notice:

∼ π0(PicR) = Pic(SpecR) ∼ × π1(PicR) = R

5.2 The closed structure on Pic

The closed structure on Pic is given by defining the internal hom Pic(P1,..., Pk; Q) to be the

Picard groupoid whose objects are k-monoidal functors F : P1 ×· · ·×Pk → Q and whose morphisms are natural isomorphism between k-monoidal functors. The “unit object” is 1Q where

1Q : P1 × · · · × Pk → Q

(x1, . . . , xk) 7→ 1Q

and the tensor product of F,G ∈ Ob Pic(P1,..., Pk; Q) is

F ⊗ G : P1 × · · · × Pk → Q

(x1, . . . , xk) 7→ F (x1, . . . , xk) ⊗ Q(x1, . . . , xk)

This makes Pic(P1,..., Pk; Q) a Picard groupoid because for any F : P1 × · · · × Pk → Q, there is −1 an inverse F : P1 × · · · × Pk → Q such that:

−1 =∼ F (x1, . . . , xk) ⊗Q F (x1, . . . , xk) −→ 1Q

Stable can be done in the category of spectra, Spec. Stable 1-types are the objects of the full subcategory of Spec that have trivial homotopy groups in all levels except 0 and

23 1 1, Spec0. Picard groupoids are known to classify stable 1-types under the and fundamental groupoid functors. Specifically, we have an equivalence of categories:

1 Ho(Spec0)  Ho(Pic)

24 CHAPTER 6

STABLE QUADRATIC MODULES

We are particularly interested in the 1-type of the K-theory of a Waldhausen category, which has many algebraic models. Muro and Tonks [1] construct a model for this 1-type by ways of stable quadratic modules which we define now.

6.1 Definitions and Properties

Definition 6.1. A stable quadratic module, C∗, is a diagram of group homomorphisms

ab ab h·,·i ∂ C0 ⊗ C0 −−→ C1 −→ C0 such that ∀ai, bi ∈ Ci for i = 0, 1,

• ∂ha0, b0i = [b0, a0] = −b0 − a0 + b0 + a0

• h∂a1, ∂b1i = [b1, a1]

• ha0, b0i + hb0, a0i = 0

The homotopy groups of C∗ are π1C∗ = ker ∂ and π0C∗ = coker ∂.

Remark 6.2. 1. The third condition is the stability condition. If we instead only require:

h∂a1, b0i + hb0, ∂a1i = 0

then C∗ would be a reduced quadratic module.

2. For any stable quadratic module C∗, we get a group action

C0 × C1 → C1

c0 (c0, c1) 7→ c1 := c1 + hc0, ∂c1i

Definition 6.3. A morphism between stable quadratic modules, f : C∗ → D∗ is given by group homomorphisms fi : Ci → Di, i = 0, 1, such that ∀ci, di ∈ Ci

• ∂D∗ f1(c1) = f0∂C∗ (c1)

25 • hf0(c0), f0(d0)iD∗ = f1hc0, d0iC∗

Weak equivalences of stable quadratic modules are morphisms that induce isomorphisms on πi for i = 0, 1.

We will denote the category of stable quadratic modules as squad and its homotopy category, Ho squad, formally inverts the weak equivalences. The following are some simple examples, see [7]:

Example 6.4. Take any group homomorphism between abelian groups: f : A → B. Then

h·,·i B ⊗ B −−→ A −→∂ B where w = 0A and ∂ = f is a stable quadratic module. Notice the action of B on A here is trivial and π0 = coker f and π1 = ker f.

Example 6.5. Take C0 = Z and C1 = Z/2Z. Let Z be generated by γ, then Z/2Z is generated by hγ, γi. The map ∂ : Z/2Z → Z is the zero map. Notice the action here is again trivial and π0 = Z and π1 = Z/2Z. This is an algebraic model for the 1-type of the sphere spectrum.

Example 6.6. A generalization of the last example is

h·,·i × 0 Z ⊗ Z −−→ F −→ Z

× with h1, 1i = −1. The action again is trivial and π0 = Z and π1 = F . This is an algebraic model for the K-theory of a field (or more generally, a commutative local ring).

6.2 An Algebraic Model for the 1-type of K(W) as a Stable Quadratic Module

We briefly recall Muro and Tonks constructed a stable quadratic module, D∗W, from a Wald- hausen category, W, by letting the group D0W be the group of nilpotency class 2 of the free group generated by objects of W, and D1W be the group generated by the weak equivalences and cofi- brations of W subject to several relations outlined below. (D1W will also end up being a group of nilpotency class 2.) There is a functor (details in section 7.2)

P : squad → Pic

26 which induces an equivalence of categories

P : Ho squad → Ho Pic between the homotopy category of stable quadratic modules and the homotopy category of Picard groupoids. In particular, ∼ πiC∗ = πi(PC∗), i = 0, 1

Therefore, as Picard groupoids classify stable 1-types, we can alternatively use stable quadratic modules as algebraic models for stable 1-types. We denote as PD·W the Picard groupoid corre- sponding to the stable quadratic module D·W and, by composition, we get a functor: PD· : Wald → Pic

Definition 6.7. Given a Waldhausen category, define a stable quadratic module D∗W as follows:

D0W is generated by [A] ∀A ∈ W with [∗] = 0

∼ 0 ∼ 0 D1W is generated by [A −→ A ] and [A  B  B/A], for all A −→ A weak equivalences and A  B cofibrations in W subject to the following relations:

a) ∂[A −→∼ A0] = −[A0] + [A]

b) ∂[A  B  B/A] = −[B] + [B/A] + [A] id c)[ A −−→A A] = 0

idA d)[ A  A  ∗] = [∗  A  A] = 0 e) For all composable weak equivalences A −→∼ B −→∼ C we have

[A −→∼ C] = [B −→∼ C] + [A −→∼ B]

f) For all commutative diagrams:

A B B/A

∼ ∼ ∼ A0 B0 B0/A0

we have

[A −→∼ A0] + [B/A −→∼ B0/A0] + h[A], −[B0/A0] + [B/A]i 0 0 0 0 ∼ 0 = −[A  B  B /A ] + [B −→ B ] + [A  B  B/A]

27 g) For all commutative diagrams:

C/B

B/A C/A

A B C

we have the relations

[B  C  C/B] + [A  B  B/A] =[A  C  C/A] + [B/A  C/A  C/B]+ h[A], −[C/A] + [C/B] + [B/A]i

i2 F p1 i1 F p2 h) h[A], [B]i = −[B  A B  A] + [A  A B  B]

We have a functor: D· : Wald → squad since for any exact functor F : W → V, we get a morphism of stable quadratic modules

D·F : D·W → D·V defined on generators by:

(D0F )[A] = [F (A)]

∼ 0 ∼ 0 (D1F )[A −→ A ] = [F (A) −→ F (A )]

(D1F )[A  B  B/A] = [F (A)  F (B)  F (B/A)]

Notice π0D·W is the group generated by [A] for all objects A in W and whose relations are generated by [A] = [A0] for all weak equivalences A −→∼ A0 and by [B/A] + [A] = [B] for all cofibrations A  B  B/A. π1D·W is the group generated by generators of automorphisms: [A → A].

28 CHAPTER 7

CROSSED MODULES

7.1 Definitions and Examples

∂ Definition 7.1. A crossed module is a group homomorphism C1 −→ C0 along with a right action

a0 C1 × C0 → C1 (where (a1, a0) 7→ a1 ) such that

a0 • ∂(a1 ) = −a0 + ∂(a1) + a0 ∀ai ∈ Ci, i = 0, 1

∂(b1) • a1 = −b1 + a1 + b1 ∀a1, b1 ∈ C1

Again, we can define πiC∗ (i = 0, 1) for crossed modules as we did for stable quadratic modules.

Definition 7.2. A morphism between crossed modules, f : C∗ → D∗, is given by group homomor-

c0 f0(c0) phisms fi : Ci → Di, i = 0, 1, such that ∀ci ∈ Ci, f1(c1 ) = f1(c1) . We will denote the category of crossed modules as cross.

Example 7.3.

i) For any normal subgroup N of a group G, the inclusion

i : N,→ G

is a crossed module where the action is given by conjugation.

ii) For any central extension of groups

1 → A → E → G → 1

the (surjective) map E → G is a crossed module with the action of G on E. iii) For a pointed pair of spaces (A, X, x), where A is a subspace of X with x ∈ A, the homotopy boundary from the second relative homotopy group to the fundamental group of A is a crossed module:

d : π2(X, A, x) → π1(A, x)

29 7.2 Properties

Lemma 7.4. squad is a full reflective subcategory of cross. (Lemma 4.15 in [1])

Any crossed module, C∗, gives rise to a Picard groupoid, PC∗, by letting the objects of PC∗ be the group C0 and letting the morphisms be the semidirect product C0 n C1. This means for any pair (c0, d1) ∈ C0 n C1, we take this to be a morphism in HomPC∗ (c0 + ∂d1, c0). The isomorphism between c0 ⊗ d0 and d0 ⊗ c0 is given by the morphism (d0 ⊗ c0, hd0, c0i). A morphism of crossed modules corresponds to a functor of Picard groupoids in a functorial way and the homotopy groups of C∗ will agree with the homotopy groups of PC∗. Altogether, this gives us a functor

P : cross → Pic which reduces to a functor on squad.

Remark 7.5. Any crossed module C∗ has a classifying space BC∗ whose homotopy groups in dimension 1 is π0C∗, in dimension 2 is π1C∗, and is 0 for all dimensions above 2. This is used to prove that crossed modules classify homotopy 2-types.

Example 7.6. Now that we have defined a stable quadratic module (hence a crossed module) D·W for any Waldhausen category W, we can define the corresponding Picard groupoid. PD·W is the Picard groupoid whose objects is a group generated by [X] for every X ∈ W and whose morphisms are generated by

([X], [f : Y →∼ Y 0]) : [X] − [Y 0] + [Y ] → [X] and

([X], [g : Y  Z]) : [X] − [Z] + [Z/Y ] + [Y ] → [X]

∼ 0 for every weak equivalence Y → Y and every cofibration Y  Z. In the proof of Proposition 9.3, we need to define a monoidal functor on PD·W for a particular Waldhausen category W. By the remark 5.3, defining the functor on the generator

([X], [f : Y →∼ Y 0]) : [X] − [Y 0] + [Y ] → [X] is equivalent to defining it on the generator

([Y 0], [f : Y →∼ Y 0]) : [Y ] → [Y 0]

30 and defining it on the generator

([X], [g : Y  Z]) : [X] − [Z] + [Z/Y ] + [Y ] → [X] is equivalent to defining it on the generator

([Z], [g : Y  Z]) : [Z/Y ] + [Y ] → [Z]

This will significantly simplify the definition of the functor which is essential for the main theorem.

31 CHAPTER 8

DETERMINANT FUNCTOR

Knudsen and Mumford categorified the determinant of invertible matrices as a determinant functor in [15]. Deligne in [9] axiomatized these properties to define a determinant functor on any exact category. He also explicitly constructed a Picard groupoid, which is the target of a universal deter- minant functor on an exact category. Muro and Tonks generalized this definition and construction tp Waldhausen categories in [7]. We now recall determinant functors on Waldhausen categories.

8.1 Definitions and Examples

Definition 8.1. Let W be a Waldhausen category and P a Picard groupoid. A determinant is a functor det : w W → P equipped, for every cofibration sequence ∆ : X  Y  Y/X in W, with an isomorphism

det(∆) : det(Y/X) ⊗ det(X) → Y which is natural with respect to commutative diagrams in W of the following form:

X Y Y/X

∼ ∼ ∼ X0 Y 0 Y 0/X0

and further subject to the following axioms: associativity For any commutative diagram consisting of four cofibration sequences

Z/Y

Y/X Z/X

X Y Z

32 we get the following commutative diagram in P

det(Z)

det(Z/Y ) ⊗ det(Y ) det(Z/X) ⊗ det(X)

det(Z/Y ) ⊗ (det(Y/X) ⊗ det(X)) (det(Z/Y ) ⊗ det(Y/X)) ⊗ det(X) commutativity For any two objects X,Y ∈ W, we have the following cofibration sequences G G ∆1 : X  X Y  Y, ∆2 : Y  X Y  X

and the following diagram in P commutes:

det(X F Y )

det(Y ) ⊗ det(X) det(X) ⊗ det(Y )

Definition 8.2. A universal determinant functor is a determinant functor, det : W → V(W), such that any determinant functor det0 : W → P must factor through det in an essentially unique way. Specifically, there exists a monoidal functor f :V(W) → P and a natural transformation α : f ◦ det ⇒ det0 such that the following diagram commutes:

det W V(W)

α f det0 P

The uniqueness of f and α are detailed in [7].

fd Example 8.3. Fix some field F . Take VectF to be the category of finite dimensional vector spaces over F . This is an exact category which we can view as an exact category. The universal determinant functor on this category is given by

fd det : VectF → linesZ V 7→ (∧dim V V, dim V )

33 where linesZ is the category of Z-graded 1-dimensional vector spaces over F . This means an object is given by a pair, (L, n) for some 1-dimensional L and some integer n. This is a

Picard groupoid by (L, n) ⊗ (K, m) = (L ⊗F K, n + m). The determinant is defined on weak equivalences, which in this case means isomorphisms. For =∼ any isomorphisms f : V −→ W of finite dimensional vector spaces where {v1, . . . , vn} is a basis for n n V and {w1, . . . , wn} is a basis for W , we get a morphism of the Z-graded lines (∧ V, n) → (∧ W, n) by defining the morphism

∧nV → ∧nW

v1 ∧ · · · ∧ vn 7→ det Af (w1 ∧ · · · ∧ wn) where Af is the matrix corresponding to the isomorphism f. The determinant also tells us that for any short exact sequence of finite dimensional vector spaces:

0 → U → V → W → 0 we get an isomorphism in linesZ:

(∧nU, n) ⊗ (∧mW, m) → (∧n+mV, n + m)

8.2 Universal Determinant Functor for a Waldhausen Category

Theorem 8.4 (Muro and Tonks [1]). The following is a universal determinant functor: det : W → PD·W X 7→ [X]

X −→∼ X0 7→ ([X0], [X → X0])

∆ 7→ ([Y ], [∆])

for any cofibration sequence ∆ : X  Y  Y/X

In fact, any model for the 1-type of the K-theory of W, P1K(W), is part of a universal de- terminant functor. This is very useful because now we can compute K0,K1 algebraically from W itself.

K0(W) = π0PD·W = π0D·W K1(W) = π1PD·W = π1D·W

34 fg Example 8.5. Fix a (commutative) ring R. Let ProjR be the category of finitely generated pro- jective modules over R. This is an exact category, hence a Waldhausen category. Every exact category of finitely generated projective modules splits and the weak equivalences are the isomor- phisms. For the rest of this example, let P be a finitely generated projective module over R. fg fg The group D0 ProjR is generated by [P ] for every P where [0R] = 0 and the group D1 ProjR is generated by the isomorphisms and split exact sequences. Therefore, defining K (Projfg) as π D Projfg, we get the abelian group generated by classes 0 R 0 · R ∼ of [P ] where [P ] = [P 0] for all isomorphisms P −→= P 0 and [P ⊕ Q] = [P ] + [Q] for every short exact sequence isomorphic to 0 → P → P ⊕ Q → Q → 0. Recalling the Grothendieck group of a ring from 4.6, we see this is precisely K0(R). Similarly, defining K (Projfg) as π D Projfg we get the abelian group that is generated by 1 R 1 · R =∼ classes of isomorphisms [P −→ P ]. K1(R), defined by Milnor, is the abelianization of the stable F general linear group over R, GL(R) = GLn(R) where n

GL1(R) ,→ GL2(R) ,→ · · · ,→ GLn(R) ,→ GLn+1(R) ,→ · · ·

A 0 and every A ∈ GL (R) is identified with in GL (R). n 0 1 n+1 Whitehead’s lemma states that the commutator subgroup of GL(R) is exactly the subgroup generated by elementary matrices where an elementary matrix differs from the identity matrix by some non-diagonal. Every matrix in K1(R) can be represented by a diagonal matrix which ∼ n = n fg represents an isomorphism R −→ R . This is precisely K1(ProjR ). Weibel explains this in more detail in [18].

35 CHAPTER 9

P1K AS A MULTIFUNCTOR

The foregoing and Muro and Tonks’ results imply there exists a functor

PD· : Wald −→ Pic

Both Wald and Pic are closed, symmetric multicategories. We have the following theorem:

Theorem 9.1. PD· extends to a multifunctor.

This theorem follows from two propositions: Proposition 9.2 below and Proposition 9.3 on page 37.

Proposition 9.2. Given closed, multicategories, M, N , and a functor F : M → N , if there is a morphism F (M(A, B)) → N (F (A),F (B))

∀A, B ∈ Ob M, then we get a morphism

HomM(A1,...,Ak; B) → HomN (F (A1),...,F (Ak); F (B))

∀A1,...,Ak,B ∈ M and ∀k ∈ N.

Proof. (We will prove this inductively on k.) First, we want to define the morphism:

F : HomM(A, B; C) → HomN (F A, F B; FC)

Since M and N are closed, multicategories, we have the following bijections:

=∼ ϕA,B;C : HomM(B; M(A, C)) −→ HomM(A, B; C)

=∼ ϕF A,F B;FC : HomN (FB; N (F A, F C)) −→ HomN (F A, F B; FC)

Since HomM(B; M(A, C)) is a set of 1-morphisms, we can apply F and get the following morphism:

HomM(B; M(A, C)) → HomN (FB; F M(A, C))

36 So, all together, we have the following diagram:

F HomM(B, M(A; C)) HomN (FB,F (M(A; C)))

∼ = HomN (FB, N (FA; FC))

=∼

HomM(A, B; C) HomN (F A, F B; FC)

In order to get the bottom dotted map necessary for F to preserve the set of 2-morphisms, we need to a morphism

ω : HomN (FB,F (M(A; C))) → HomN (FB, N (FA; FC))

By Yoneda’s lemma, this is equivalent to having a morphism

ω : F (M(A; C)) → N (FA; FC)

This is precisely what Proposition 9.2 requires. Now we can extend the same argument to define a map

F : HomM(A1,...,Ak; B) → HomN (FA1,...,FAk; FB) by studying the diagram

F HomM(A2,...,Ak+1; M(A1; B)) HomN (FA2,...,Ak+1; F (M(A1; B)))

∼ = HomN (FA2,...,Ak+1; N (FA1; FB))

=∼

HomM(A1,...,Ak+1; B) HomN (FA1,...,FAk+1; FB)

Again, by Yoneda’s lemma, the missing gap is determined by a morphism

ω : F (M(A1,B)) → N (FA1,FB)

Proposition 9.3. There exists a functor of Picard groupoids:

PD·(Wald(W, V)) → Pic(PD·W; PD·(V))

37 Proof. We will start with explicitly describing each of the Picard groupoids in the functor. First, let’s describe the internal Hom object Wald(W; V) as a Waldhausen category. Its objects are exact functors ϕ : W → V and its morphisms are natural transformations between exact functors. The weak equivalences are defined to be the natural weak equivalences and the cofibrations are the 0 natural transformations, α : ϕ0 =⇒ ϕ1 : W −→ V, such that for every cofibration f : X  X in W, all the maps along with the southern arrow of the following diagram are cofibrations:

ϕ0(f) 0 ϕ0(X) ϕ0(X )

αX αX0 ϕ1(f) 0 ϕ1(X) ϕ1(X )

(Zakharevich proves that this is well-defined in [2].) We can construct the stable quadratic module, D·(Wald(W; V)) where D0(Wald(W; V)) is the group generated by exact functors and D1(Wald(W; V)) is generated by the weak equivalence and cofibration natural transformations defined above, subject to some relations. Finally, PD·(Wald(W; V)) is the Picard groupoid corre- sponding to this stable quadratic module. In particular, the objects are D0(Wald(W; V)) and the morphisms are given by the semi-direct product:

D0(Wald(W; V)) n D1(Wald(W; V)) = {([ϕ], [α]) : [ϕ] + ∂[α] −→ [ϕ]}

The group structure for the Hom sets in PD·(Wald(W; V)) is given by:

[ϕ1] ([ϕ0], [α0]) + ([ϕ1], [α1]) = ([ϕ0] + [ϕ1], [α0] + [α1]) and

1[ϕ] = ([ϕ], 0) We need to describe Pic(PD·W; PD·V) as an internal Hom object in Pic. Its objects are monoidal functors F : PD·W −→ PD·V and its morphisms are natural transformations between such monoidal functors. The monoidal structure on the set of objects is given by computing the monoidal operation pointwise in PD·V. Specifically, for every object X in W, we have F + G : PD·W −→ PD·V [X] 7→ F ([X]) + G([X])

38 We are now ready to define the functor:

ω : PD·(Wald(W; V)) −→ Pic(PD·W; PD·V) For every generator [ϕ : W −→ V] in PD·(Wald(W; V)), ω([ϕ : W −→ V]) = PD·ϕ. Therefore: ω([ϕ]) : PD·W −→ PD·V [X] 7→ [ϕ(X)]

[X] + [Y ] 7→ [ϕ(X)] + [ϕ(Y )] f ϕ(f) ([X], [Y −→ Y 0]) 7→ ([ϕ(X)], [ϕ(Y ) −−−→ ϕ(Y 0)])

0 The morphisms in PD·(Wald(W; V)) are generated by ([ϕ], [α : τ τ ]) : [ϕ] + ∂[α] → [ϕ] and ([ϕ], [β : η δ]) : [ϕ] + ∂[β] → [ϕ] for α natural weak equivalence and β a natural transformation that is a cofibration in the internal Hom Wald(W; V). By example 7.6, we can restrict to the 0 0 0 generators ([τ ], [α : τ τ ]) : [τ] → [τ ] and ([δ], [β : η δ]) : [δ/η] + [η] → [δ]. Define ω([τ 0], [α]) to be the following morphism in Pic(PD·W,PD·V):

0 PD·α : PD·τ PD·τ This is the natural transformation defined as follows for all generators [X] ∈ PD·W

0 0 PD·α[X] = ([τ (X)], [α(X)]) ∈ HomPD·V ([τ(X)], [τ (X)]) Define ω([δ], [β]) to be the following morphism in Pic(PD·W,PD·V):

PD·β : PD·δ/η + PD·η PD·δ This is the natural transformation defined as follows for all generators [X] ∈ PD·W

PD·β[X] = ([δ(X)], [β(X)]) ∈ HomPD·V ([δ/η(X)] + [η(X)], [δ(X)]) Normally we would need to check that PD·α and PD·β are natural transformation between Picard groupoids. However, it has been shown that the functors

D· : Wald → squad and P : squad → Pic

39 are 2-functors. Our definition of ω is really the composition of these two functors on the exact functors of Waldhausen categories and natural transformations between these functors. This means these natural transformations are well defined. Although these computations do not need to be shown, they are in the Appendix for the curious reader. We do need to check that ω is a well-defined symmetric, monoidal functor between Picard Groupoids. This is a matter of checking simple but tedious computations, which can be found the Appendix.

9.1 Proof of Main Theorem

Proof. From Propositions 9.2 and 9.3, we have the following maps for all k:

PD· PD· : HomWald(W1,..., Wk; V) −−−→ HomPic(PD·(W1),...,PD·(Wk); PD·(V))

Notice that the group of objects in PD·W is generated by detW (X) = [X]W for all objects X of W. So we can define morphisms in Pic on the generators alone. We will use the diagram from Proposition 9.2 and ω defined in Proposition 9.3 to determine how PD· maps k-morphisms in Wald to k-morphisms in Pic. This diagram involves the isomorphisms that exist from the closed structures on Wald and Pic:

=∼ η : HomWald(W1, W2,..., Wk+1; V) −→ HomWald(W2,..., Wk+1; Wald(W1, V))

∼= ϕ: HomPic(PD·(W2),...,P D·(Wk+1);Pic(PD·(W1),P D·(V)))−→HomPic(PD·(W1),...,P D·(Wk+1);PD·(V))

We will describe these isomorphisms point-wise.

For every f ∈ HomWald(W1, W2,..., Wk+1; V) and every Xi ∈ Wi,

η(f)(X2,...,Xk+1) ∈ Wald(W1, V) corresponds to the functor that maps X1 to f(X1,X2,...,Xk) ∈ V.

For every f ∈ HomPic(PD·(W2),...,PD·(Wk+1); Pic·(PD·(W1),PD·(V)))

ϕ(f) = ev(id , f) PD·W1

maps ([X1]W1 ,..., [Xk+1]Wk+1 ) to f([X2]W1 ,..., [Xk+1]Wk+1 )([X1]W1 )

40 The diagram also involves the functor PD· which is defined on 1-morphisms. The map

PD· : HomWald(W2; Wald(W1; V)) → HomPic(PD·(W2); PD·(Wald(W1, V))) maps f to

PD·(f): PD·(W2) → PD·(Wald(W1, V))

[X]W2 7→ [f(X): W1 → V]Wald(W1,V)

PD·, as defined in Proposition 9.2 on the set of 2-morphisms HomWald(W1, W2; V), is given by the following diagram:

PD· HomWald(W2; Wald(W1, V)) HomPic(PD·(W2); PD·(Wald(W1, V))) ω◦−

η HomPic(PD·(W2); Pic(PD·(W1),PD·(V))) ev(id −) PD·(W1) PD· HomWald(W1, W2; V) HomPic(PD·(W1),PD·(W2); PD·(V))

Therefore, for f ∈ HomWald(W1, W2; V),

PD f(det (X), det (Y )) = ev(id , ω ◦ PD ηf)(det (X), det (Y )) · W1 W2 PD·(W1) · W1 W2 = ev(det (X), ω ◦ PD ηf(det (Y ))) W1 · W2

= ev(detW1 (X), ω detWald(W1,V)(ηf(Y ))) = ev(det (X),PD ηf(Y )) W1 · = PD ηf(Y )(det (X)) · W1

= detV (ηf(Y )(X))

= detV (f(X,Y )) i.e. (PD f)([X] , [Y ] ) = [f(X,Y )] · W1 W2 V In fact, this extends and we get:

PD f(det (X ),..., det (X )) = det (f(X ,...,X )) · W1 1 Wk+1 k+1 V 1 k+1 which is the same as (PD·f)([X1],..., [Xk+1]) = [f(X1,...,Xk+1)])

41 We know PD· is a functor, but to show it is a multifunctor we have to check the composition and symmetric conditions on the set of k-morphisms. Namely, the following two diagrams should commute: Qk HomWald(V1,..., Vk; U) × i=1 HomWald(Wi1,..., Wili ; Vi)

PD· × Q PD·

Hom (PD V ,...,PD V ; PD U) × Qk Hom (PD W ,...,PD W ; PD V ) ◦ Pic · 1 · k · i=1 Pic · i1 · ili · i

PD Hom (PD W ,...,PD W ; PD U) · Hom (W ,..., W ; U) Pic · 11 · klk · Wald 11 klk

σ∗ HomWald(W1,..., Wk; V) HomWald(Wσ(1),..., Wσ(k); V) PD· PD· σ∗ HomPic(PD·W1,...,PD·Wk; PD·V) HomPic(PD·Wσ(1),...,PD·Wσ(k); PD·V) To show PD· respects composition for k = 2, there are two cases to check. Assume f ∈ 0 HomWald(W1, W2; V), gi ∈ HomWald(Wi, Wi) for i = 1, 2, f ∈ HomWald(W2, V), and g ∈

HomWald(W1, W2; V). We need the following identities:

PD·(f ◦ (g1, g2)) = PD·f ◦ (PD·g1,PD·g2) PD·(f ◦ g) = PD·f ◦ PD·g which we have by defining point-wise on objects [X] ∈ PD W and [Y ] ∈ PD W : W1 · 1 W2 · 2 PD (f ◦ (g , g ))([X] , [Y ] ) = [(f(g , g )(X,Y ))] · 1 2 W1 W2 1 2 V

= [f(g1(X), g2(Y ))]V

= PD f([g (X)] , [g (Y )] ) · 1 W1 2 W2 = PD f(PD g ([X] ),PD g ([Y ] )) · · 1 W1 · 2 W2 = PD f ◦ (PD g ,PD g )([X] , [Y ] ) · · 1 · 2 W1 W2

PD (f ◦ g)([X] , [Y ] ) = [f ◦ g(X,Y )] · W1 W2 V = PD f([g(X,Y )] ) · W2 = PD f(PD g([X] , [Y ] )) · · W1 W2 = PD f ◦ PD g([X] , [Y ] ) · · W1 W2

42 The proof for any k is very similar. To prove the symmetry condition for k = 2, we should note there is only one trivial element in

Σ2, which is the one that flips the two objects. So we need to show:

∗ σWald HomWald(W1, W2; V) HomWald(W2, W1; V) PD· PD· ∗ σPic HomPic(PD·W1,PD·W2; PD·V) HomPic(PD·W2,PD·W1; PD·V)

PD (σ∗ f)(det (Y ), det (X)) = det (σ∗ f(Y,X)) · Wald W2 W1 V Wald

= detV (f(X,Y ))

= PD f(det (X), det (Y )) · W1 W2 = σ∗ (PD f)(det (Y ), det (X)) Pic · W2 W1

Once again, this extends similarly for any k.

43 APPENDIX A

COMPUTATIONS FOR PROOFS

A.1 ω defines a natural transformation on morphisms

The next few computations are to check the natural transformations defined under ω respect the usual commutative diagram for any morphism in PD·W. Once again, we can restrict to generators ([Y 0], [f]) : [Y ] → [Y 0] and ([Z], [g]) : [Z/Y ] + [Y ] → [Z]. 0 0 0 ∼ 0 First, let’s start with ω([τ ], [α]) : PD·τ PD·τ on the morphism ([Y ], [f : Y → Y ]). We need the following diagram to commute: ([τ(Y 0)], [τ(f)]) [τ(Y )] [τ(Y 0)]

0 0 0 ([τ (Y )], [αY ]) ([τ (Y )], [αY 0 ]) ([τ 0(Y 0)], [τ 0(f)]) [τ 0(Y )] [τ 0(Y 0)]

So the composition both ways is:

0 0 ([τ (Y )], [αY 0 ] + [τ(f)]) and 0 0 0 ([τ (Y )], [τ (f)] + [αY ])

To prove these are the same morphism, we have to show the second factors are the same. (Since their image under ∂ is the same, any action on their difference is trivial.) Since α is a natural transformation, we have the following commutative diagram consisting of all weak equivalences for f : Y →∼ Y 0 in W:

τ(f) τ(Y ) τ(Y 0)

αY αY 0 τ 0(f) τ 0(Y ) τ 0(Y 0)

So, by the relations in D1V, 0 [αY 0 ] + [τ(f)] = [τ (f)] + [αY ]

44 which is what we needed to prove the composition both ways agree. 0 0 Now, let’s work with ω([τ ], [α]) : PD·τ PD·τ on the morphism ([Z], [g : Y  Z]). We need the following diagram to commute: ([τ(Z)], [τ(g)]) [τ(Z/Y )] + [τ(Y )] [τ(Z)]

0 0 0 ([τ (Z/Y )], [αZ/Y ]) + ([τ Y ], [αY ]) ([τ (Z)], [αZ ])

[τ 0(Z/Y )] + [τ 0(Y )] [τ 0(Z)] ([τ 0(Z)], [τ 0(g)])

So the composition both ways is: 0 ([τ (Z)], [αZ ] + [τ(g)]) and 0 0 [τ 0(Y )] ([τ (Z)], [τ (g)] + [αZ/Y ] + [αY ])

Therefore, we need

0 [τ(Y )] [αZ ] + [τ(g)] = [τ (g)] + [αZ/Y ] + [αY ] (A.1)

Since α is a natural transformation, we have the following commutative diagram consisting of weak equivalences for the vertical morphisms and cofibration sequences horizontally for g : Y  Z in W: τ(f) τ(Y ) τ(Z) τ(Z/Y )

αY αZ αZ/Y τ 0(f) τ 0(Y ) τ 0(Z) τ 0(Z/Y )

So, by the relations in D1V, we get

0 [τ(Y )] [τ (g)] + [αY ] + [αZ/Y ] = [αZ ] + [τ(g)]

45 Therefore, equation (A.1) becomes:

0 [τ(Y )] 0 [τ(Y )] [τ (g)] + [αY ] + [αZ/Y ] = [τ (g)] + [αZ/Y ] + [αY ]

0 =⇒ [αY ] + [αZ/Y ] + h[τ(Y )], ∂[αZ/Y ]i = [αZ/Y ] + [αY ] + h[τ (Y )], ∂[αY ]i

=⇒ [αY ] + [αZ/Y ] = [αZ/Y ] + [αY ] + h∂[αZ/Y ], ∂[αY ]i

= [αZ/Y ] + [αY ] + [[αY ], [αZ/Y ]]

= [αY ] + [αZ/Y ]

Now let’s move on to checking that ω([δ], [β]) : PD·δ/η + PD·η PD·δ. We will start on the morphism ([Y 0], [f : Y →∼ Y 0]) is a natural transformation. We need the following diagram to commute: ([δ/η(Y 0)], [δ/η(f)]) + ([η(Y 0)], [η(f)]) [δ/η(Y )] + [η(Y )] [δ/η(Y 0)] + [η(Y 0)]

0 ([δ(Y )], [βY ]) ([δ(Y )], [βY 0 ])

[δ(Y )] [δ(Y 0)] ([δ(Y 0)], [δ(f)])

So the composition both ways is:

0 [η(Y 0)] ([δ(Y )], [βY 0 ] + [δ/η(f)] + [η(f)]) and 0 ([δ(Y )], [δ(f)] + [βY ]

We need

[η(Y 0)] [βY 0 ] + [δ/η(f)] + [η(f)] = [δ(f)] + [βY ] (A.2)

Since β is a natural transformation, we have the following commutative diagram consisting of weak equivalences for the vertical morphisms and cofibration sequences horizontally for f : Y →∼ Y 0 in W: β η(Y ) Y δ(Y ) δ/η(Y )

η(f) δ(f) δ/η(f)

β 0 η(Y 0) Y δ(Y 0) δ/η(Y 0)

46 So, by the relations in D1V, we get

[η(Y )] [δ(f)] + [βY ] = [βY 0 ] + [η(f)] + [δ/η(f)]

Therefore, equation (A.2) becomes:

[η(Y 0)] [η(Y )] [βY 0 ] + [δ/η(f)] + [η(f)] = [βY 0 ] + [η(f)] + [δ/η(f)]

=⇒ [δ/η(f)] + [η(f)] + h[η(Y 0)], ∂[δ/η(f)]i = [η(f)] + [δ/η(f)] + h[η(Y )], ∂[δ/η(f)]i

=⇒ [δ/η(f)] + [η(f)] = [η(f)] + [δ/η(f)] + h∂[η(f)], ∂[δ/η(f)]i

= [η(f)] + [δ/η(f)] + [[δ/η(f)], [η(f)]]

= [δ/η(f)] + [η(f)]

Finally, we have to check ω([δ], [β]) : PD·δ/η +PD·η PD·δ on the morphism ([Z], [g : Y  Z]). We need the following diagram to commute: [δ/η(Z/Y )] + [η(Z/Y )] + [δ/η(Y )] + [η(Y )]([δ/η(Z)], [δ/η(g)]) + ([η(Z)], [η(g)] + h[δ/η(Y )], [η(Z/Y )]i) [δ/η(Z)] + [η(Z)]

([δ(Z/Y )], [βZ/Y ]) + ([δ(Y )], [βY ]) ([δ(Z)], [βZ ])

[δ(Z/Y )] + [δ(Y )] [δ(Z)] ([δ(Z)], [δ(g)])

Therefore, the composition both ways is:

[η(Z)] ([δ(Z)], [βZ ] + [δ/η(g)] + [η(g)] + h[δ/η(Y )], [η(Z/Y )]i) and [δ(Y )] ([δ(Z)], [δ(g)] + [βZ/Y ] + [βY ])

[η(Z)] [δ(Y )] =⇒ [βZ ] + [δ/η(g)] + [η(g)] + h[δ/η(Y )], [η(Z/Y )]i = [δ(g)] + [βZ/Y ] + [βY ] (A.3)

Since β is a natural transformation and is a cofibration, that means the southern arrow condition is satisfied: β η(Y ) Y δ(Y )

i η(Z) F δ(Y ) η(g) δ(g) η(Y ) j k

β η(Z) Z δ(Z)

47 And we get the following four diagrams:

δ/η(Z/Y )

δ/η(Y ) δ/η(Z) (A.4)

η(Z) η(Z) F δ(Y ) δ(Z) η(Y )

δ/η(Z/Y )

η(Z/Y ) δ(Z/Y ) (A.5)

δ(Y ) η(Z) F δ(Y ) δ(Z) η(Y )

η(Z/Y )

F δ/η(Y ) δ/η(Y ) η(Z/Y ) (A.6)

η(Y ) δ(Y ) η(Z) F δ(Y ) η(Y )

δ/η(Y )

F η(Z/Y ) δ/η(Y ) η(Z/Y ) (A.7)

η(Y ) η(Z) η(Z) F δ(Y ) η(Y )

So, by the relations in D1V, we get the following four relations:

[η(Z)] [k] + [j] = [βZ ] + [δ/η(g)] (A.8)

[δ(Y )] [k] + [i] = [δ(g)] + [βZ/Y ] (A.9)

G [η(Y )] [i] + [βY ] = [i ◦ βY ] + [δ/η(Y )  δ/η(Y ) η(Z/Y )] (A.10)

48 G [η(Y )] [j] + [η(g)] = [j ◦ η(g)] + [η(Z/Y )  δ/η(Y ) η(Z/Y )] (A.11)

Using (A.6) and (A.7) as well as the fact that i ◦ βY = j ◦ η(g), we get:

G [η(Y )] G [η(Y )] [i] + [βY ] = [j] + [η(g)] − [η(Z/Y )  δ/η(Y ) η(Z/Y )] + [δ/η(Y )  δ/η(Y ) η(Z/Y )] = [j] + [η(g)] + h[δ/η(Y )], [η(Z/Y )]i[η(Y )]

Any action on the image of h·, ·i is trivial. Using (A.4) the right hand side further simplifies to

[η(Z)] −[k] + [βZ ] + [δ/η(g)] + [η(g)] + h[δ/η(Y )], [η(Z/Y )]i

Using (A.5), the left hand side simplifies to:

[δ(Y )] −[k] + [δ(g)] + [βZ/Y ] + [βY ]

Therefore, we get

[δ(Y )] [η(Z)] [δ(g)] + [βZ/Y ] + [βY ] = [βZ ] + [δ/η(g)] + [η(g)] + h[δ/η(Y )], [η(Z/Y )]i which is exactly what (A.3) needed. Technically, we must also check these are symmetric monoidal natural transformations. This means that for ω([τ 0], [α]), we need PD·τ([X]) + PD·τ([Y ]) PD·τ 0([X]) + PD·τ 0([Y ])

PD·τ([X] + [Y ]) PD·τ 0([X] + [Y ]) and 0

PD·τ(0) PD·τ 0(0) However, these commutativity diagrams are trivially satisfied since PD·τ([X] + [Y ]) := PD·τ([X]) + PD·τ([Y ]) and PD·τ(0) = 0. Similarly for ω([δ], [β]).

49 A.2 ω is a well-defined symmetric, monoidal functor

Let [ϕ], [ϕ0], [ϕ1], [ϕ2] ∈ PD·(Wald(W; V)). I) Functor Requirements:

a) Composition Preserved:

ω(([ϕ0], [α0]) ◦ ([ϕ0] + ∂[α0], [α1]))[X] = ω(([ϕ0], [α0] + [α1]))[X]

= ([ϕ0(X)], [α0(X)] + [α1(X)])

= ([ϕ0(X)], [α0(X)]) ◦ ([ϕ0(X)] + ∂[α0(X)], [α1(X)])

= ω(([ϕ0], [α0]))[X] ◦ ω(([ϕ0] + ∂[α0], [α1]))[X]

b) Identity Morphism Preserved: ω(([ϕ], 0))[X] = ([ϕ(X)], 0)

II) Monoidal Requirements:

a) Identity Object Preserved: ω(∗) = ω([0W ]) = [ϕ(0W )] = [0V ] = ∗ b) Monoidal Structure on Objects Preserved: True by definition. c) Monoidal Structure on Morphisms Preserved:

[ϕ1] ω(([ϕ0], [α0]) + ([ϕ1], [α1]))[X] = ω(([ϕ0] + [ϕ1], [α0] + [α1]))[X]

[ϕ1(X)] = ([ϕ0(X)] + [ϕ1(X)], [α0(X)] + [α1(X)])

= ([ϕ0(X)], [α0(X)]) + ([ϕ1(X)], [α1(X)])

= ω(([ϕ0], [α0]))[X] + ω(([ϕ1], [α1]))[X]

d) Associativity Respected: We need the following diagram to commute:

(ω([ϕ0]) + ω([ϕ1])) + ω([ϕ2]) ω([ϕ0]) + (ω([ϕ1]) + ω(ϕ2))

ω([ϕ0] + [ϕ1]) + ω([ϕ2]) ω([ϕ0]) + ω([ϕ1] + [ϕ2])

ω(([ϕ0] + [ϕ1]) + [ϕ2]) ω([ϕ0] + ([ϕ1] + [ϕ2])) which by definition really means:

(PD·ϕ0 + PD·ϕ1) + PD·ϕ2 PD·ϕ0 + (PD·ϕ1 + PD·ϕ2) id id

(PD·ϕ0 + PD·ϕ1) + PD·ϕ2 PD·ϕ0 + (PD·ϕ1 + PD·ϕ2) id id

(PD·ϕ0 + PD·ϕ1) + PD·ϕ2 PD·ϕ0 + (PD·ϕ1 + PD·ϕ2)

50 which is true since all vertical natural transformations are identities. e) Unitality Respected: We need the following diagrams to commute:

∗ + ω([ϕ]) id ω(∗) + ω([ϕ])

id id ω([ϕ]) ω(∗ + [ϕ]) id

ω([ϕ]) + ∗ id ω([ϕ]) + ω(∗)

id id ω([ϕ]) ω([ϕ] + ∗) id which is true since all natural transformations are identities.

III) Symmetry Requirements: We need the following diagram to commute:

ω([ϕ0]) + ω([ϕ1]) ω([ϕ1]) + ω([ϕ0])

ω([ϕ0] + [ϕ1]) ω([ϕ1] + [ϕ0])

which by definition really means:

[ϕ0(X)] + [ϕ1(X)] [ϕ1(X)] + [ϕ0(X)]

(PD·ϕ0 + PD·ϕ1)([X]) (PD·ϕ1 + PD·ϕ0)([X]) This means does

ω([ϕ1] + [ϕ0], h[ϕ1], [ϕ0]i)[X] = ([ϕ1(X)] + [ϕ0(X)], h[ϕ1(X)], [ϕ0(X)]i)

Well, by definition,

ω([ϕ1] + [ϕ0], h[ϕ1], [ϕ0]i)[X] = (([ϕ1] + [ϕ0]([X])), h[ϕ1], [ϕ0]i([X]))

= ([ϕ1(X)] + [ϕ0(X)], h[ϕ1(X)], [ϕ0(X)]i)

Together, this shows that ω is a well-defined symmetric, monoidal functor between PD·(Wald(W; V)) and Pic(PD·W; PD·V).

51 BIBLIOGRAPHY

[1] Fernando Muro and Andrew Tonks. The 1-Type of a Waldhausen K-Theory Spectrum. Ad- vances in Mathematics 216 (2007), no. 1, 178-211. arXiv:math/0603544

[2] Inna Zakharevich. The category of Waldhausen categories as a closed multicategory. arXiv:1410.4834

[3] Ettore Aldrovandi. Biextensions, bimonoidal functors, multilinear functor calculus, and cate- gorical rings. Theory and Applications of Categories 32 (2017), 889–969. arXiv:1501.04664

[4] Andrew J. Blumberg, Michael A. Mandell. Derived Koszul Duality and Involutions in the Algebraic K-Theory of Spaces. J. Topology 4 (2011), no. 2, 327-342. arXiv:0912.1670

[5] Friedhelm Waldhausen. Algebraic K-theory of spaces. In Algebraic and geometric topology (New Brunswick, N.J., 1983), volume 1126 of Lecture Notes in Math., pages 318-419. Springer, Berlin, 1985.

[6] Ettore Aldrovandi. Butterflies I: morphisms of 2-group stacks. Advances in Mathematics 221 (2009), 687-773. arXiv:0808.3627

[7] Fernando Muro, Andrew Tonks, Malte Witte. On determinant functors and K-theory. arXiv:1006.5399

[8] A. D. Elmendorf and M. A. Mandell. Rings, modules, and algebras in infinite loop space theory. Adv. Math., 205(1):163-228, 2006. arXiv:math/0403403

[9] P. Deligne. Le d´eterminantde la cohomologie. Contemp. Math., vol. 67, Amer. Math. Soc., Providence, RI, 1987, pp. 93-177.

[10] Tom Leinster. Higher , Higher Categories. Higher Operads, Higher Categories, by Tom Leinster, pp. 448. ISBN 0521532159. Cambridge, UK: Cambridge University Press, August 2004., 448 arXiv:math/0305049

[11] Andre Joyal, Ross Street. Braided Monoidal Categories. Macquarie Math Reports (1986).

[12] Niles Johnson, Angelica M. Osorno. Modeling Stable One-Types. Theory and Applications of Categories, Vol. 26, 2012, No. 20, pp 520-537. arXiv:1201.2686

[13] Bjorn Ian Dundas, Thomas G. Goodwillie, Randy McCarthy. Local Structure of Algebraic K-theory. ISBN: 9781447143925 (print)

[14] M. Artin, B. Mazur. On the Van Kampen theorem. Topology 5 (1966) 179189

52 [15] F.F. Knudsen, D. Mumford. The projectivity of the moduli space of stable curves I. Prelimi- naries on det and Div. Math. Scand. 39 (1976), no. 1, 1955.

[16] Paul Goerss, Rick Jardine. Simplicial homotopy theory. Progress in Mathematics, Birkhuser (1996)

[17] John Milnor, The geometric realization of a semi-simplicial complex, Ann. of Math. (2) 65 (1957), 357362.

[18] Charles A. Weibel. The K-book: An introduction to algebraic K-theory. Graduate Studies in Mathematics 145).

53 BIOGRAPHICAL SKETCH

I was born in Miami, FL by Cuban immigrants. My father was a math professor in Cuba who taught me at the age of three how to read and write, as well as some basic mathematics such as geometry and arithmetic. This is when I fell in love with the subject. I am the first in my family to attend college (in the United States), let alone graduate with a doctorate degree. I graduated from my local high school as Valedictorian and was a recipient of the Gates Millennium Scholarship which covered my cost of attendance for my entire undergraduate and graduate career. I attended Florida State University as an undergraduate Pure Mathematics major and continued at FSU as a graduate student, where I worked as a Teaching Assistant. I have had the pleasure of doing research for the last several years with my PhD advisor, Dr. Ettore Aldrovandi. Apart from several talks at FSU’s algebra seminars, I was invited to speak at two conferences: UF/FSU Topology conference in Gainesville, FL in Spring of 2017 and the 2017 Lloyd Roeling UL Lafayette Mathematics Conference in Fall of 2017. I enjoyed these conferences so much as I was able to meet so many peers from around the country and learn about their research interests. These experiences will stay with me forever.

54