100
Determinants in K -theory and operator algebras by Joseph Migler B.S., University of California, Santa Barbara, 2010 M.A., University of Colorado, Boulder, 2012 A thesis submitted to the Faculty of the Graduate School of the University of Colorado in partial fulfillment of the requirements for the degree of Doctor of Philosophy Department of Mathematics 2015

Determinants in K-theory and operator algebras

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Determinants in K-theory and operator algebras

Determinants in K-theory and operator algebras

by

Joseph Migler

B.S., University of California, Santa Barbara, 2010

M.A., University of Colorado, Boulder, 2012

A thesis submitted to the

Faculty of the Graduate School of the

University of Colorado in partial fulfillment

of the requirements for the degree of

Doctor of Philosophy

Department of Mathematics

2015

Page 2: Determinants in K-theory and operator algebras

This thesis entitled:Determinants in K-theory and operator algebras

written by Joseph Miglerhas been approved for the Department of Mathematics

Professor Alexander Gorokhovsky

Professor Carla Farsi

Professor Judith Packer

Professor Arlan Ramsay

Professor Bahram Rangipour

Date

The final copy of this thesis has been examined by the signatories, and we find that both thecontent and the form meet acceptable presentation standards of scholarly work in the above

mentioned discipline.

Page 3: Determinants in K-theory and operator algebras

iii

Migler, Joseph (Ph.D., Mathematics)

Determinants in K-theory and operator algebras

Thesis directed by Professor Alexander Gorokhovsky

A determinant in algebraic K-theory is associated to any two Fredholm operators which

commute modulo the trace class. This invariant is defined in terms of the Fredholm determinant,

which itself extends the usual notion of matrix determinant. On the other hand, one may consider

a homologically defined invariant known as joint torsion. This thesis answers in the affirmative a

conjecture of R. Carey and J. Pincus, namely that these two invariants agree.

The strategy is to analyze how joint torsion transforms under an action by certain groups

of pairs of invertible operators. This allows one to reduce the calculation to the finite dimensional

setting, where joint torsion is shown to be trivial. The equality implies that joint torsion has

continuity properties, satisfies the expected Steinberg relations, and depends only on the images

of the operators modulo trace class. Moreover, we show that the determinant invariant of two

commuting operators can be computed in terms of finite dimensional data.

The second main goal of this thesis is to investigate how joint torsion behaves under the

functional calculus. We study the extent to which the functional calculus commutes, modulo

operator ideals, with projections in a finitely summable Fredholm module. As an application, we

recover in particular some results of R. Carey and J. Pincus on determinants of Toeplitz operators

and Tate tame symbols. In addition, we obtain variational formulas and explicit integral formulas

for joint torsion.

Page 4: Determinants in K-theory and operator algebras

iv

Acknowledgements

I am indebted to Alexander Gorokhovsky, whom it is my honor to have as an advisor. His

endless patience and mathematical insight are a source of inspiration for me. I am grateful for his

guidance, for countless enlightening conversations, and for his generosity in sharing both his time

and knowledge. His help was never more than a hallway away.

I wish to thank Carla Farsi, Judith Packer, Arlan Ramsay, and Bahram Rangipour for serving

on my dissertation committee. I would also like to thank Carla Farsi for her encouragement and

guidance during several classes I taught as a graduate student. I wish to thank Judith Packer for

organizing many fascinating seminars which I had the pleasure of attending. I would like to thank

Arlan Ramsay for his careful reading of this dissertation. I wish to express my gratitude to Bahram

Rangipour for generously sharing his time with me during his stay in Boulder.

I am grateful for the help of many mathematicians, in particular Richard Carey, Guillermo

Cortinas, Raul Curto, Jorg Eschmeier, Karl Gustafson, Nigel Higson, Matthias Lesch, Jens Kaad,

Jerry Kaminker, Ryszard Nest, Joel Pincus, Raphael Ponge, Dan Voiculescu, and Mariusz Wodz-

icki. Their comments on the subject of this thesis and related topics have greatly enhanced my

understanding.

Finally I wish to thank my family – this degree is as much theirs as it is mine. I would like to

thank my parents for their support at every stage of my education. They taught me integrity and

most everything I know today. I am especially grateful for their encouragement during my years

of studying mathematics. I would like to thank my sisters and their families for the joy they have

shared with me and for their help throughout my life.

Page 5: Determinants in K-theory and operator algebras

Contents

Chapter

1 Introduction 1

2 Preliminaries 8

2.1 Fredholm theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2.1.1 Schatten classes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2.1.2 Fredholm determinant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

2.1.3 Fredholm index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

2.2 Algebraic K-theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

2.2.1 Algebraic K0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

2.2.2 Algebraic K1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

2.2.3 Algebraic K2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

2.2.4 The determinant invariant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

2.3 Toeplitz operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

2.3.1 Index and spectral theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

2.3.2 Commutators of Toeplitz operators . . . . . . . . . . . . . . . . . . . . . . . . 22

3 Joint torsion 24

3.1 Background: The work of Carey and Pincus . . . . . . . . . . . . . . . . . . . . . . . 24

3.2 Commuting operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

3.2.1 Algebraic torsion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

Page 6: Determinants in K-theory and operator algebras

vi

3.2.2 Joint torsion of commuting operators . . . . . . . . . . . . . . . . . . . . . . . 28

3.2.3 The case of two commuting operators . . . . . . . . . . . . . . . . . . . . . . 29

3.3 Almost commuting operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

3.3.1 Perturbation vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

3.3.2 Joint torsion of two almost commuting operators . . . . . . . . . . . . . . . . 35

4 Equality 36

4.1 The finite dimensional case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

4.1.1 Torsion of a double complex . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

4.1.2 Joint torsion in finite dimensions . . . . . . . . . . . . . . . . . . . . . . . . . 39

4.2 Factorization results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

4.2.1 Perturbation vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

4.2.2 Joint torsion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

4.3 Main results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

4.3.1 Group actions on commuting squares . . . . . . . . . . . . . . . . . . . . . . . 51

4.3.2 A proof of equality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

4.4 Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

4.4.1 Properties of the determinant invariant . . . . . . . . . . . . . . . . . . . . . 53

4.4.2 Properties of joint torsion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

4.4.3 Joint torsion and Hilbert-Schmidt operators . . . . . . . . . . . . . . . . . . . 59

5 Explicit formulas for joint torsion 62

5.1 Transformation rules for joint torsion . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

5.1.1 Commutators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

5.1.2 Perturbations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

5.1.3 Joint torsion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

5.1.4 Positivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

5.2 Fredholm modules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

Page 7: Determinants in K-theory and operator algebras

vii

5.3 Toeplitz operators and tame symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

5.3.1 H∞ symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

5.3.2 L∞ symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

5.3.3 An integral formula . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

Bibliography 85

Appendix

A The existence of perturbations 89

Page 8: Determinants in K-theory and operator algebras

Chapter 1

Introduction

Let A and B be two invertible operators on a complex separable Hilbert space H that com-

mute modulo the trace class L1(H), which consists of compact operators with summable singular

values. The multiplicative commutator ABA−1B−1 is an invertible determinant class operator,

that is it differs from the identity by a trace class operator, and therefore has a nonzero Fredholm

determinant. The assignment (A,B) 7→ det(ABA−1B−1) is bimultiplicative and skew-symmetric.

Moreover, det(ABA−1B−1) = det(ABA−1B−1) for any invertible trace class perturbations A and

B of A and B, respectively.

L. Brown observed in [8] that this is a special case of a more general phenomenon. Indeed,

any two bounded Fredholm operators A and B that commute modulo trace class have invertible

commuting images a and b, respectively, in the quotient B/L1. Here B = B(H) is the algebra

of bounded linear operators on H, and L1 = L1(H) is the ideal of trace class operators. As we

shall see in Section 2.2, these images determine a canonical element a, b, known as a Steinberg

symbol, in the second algebraic K-group K2(B/L1). Consequently, there is a determinant invariant

d(a, b) = det ∂a, b ∈ C×. Brown showed that when A and B are invertible, this invariant recovers

the classical determinant of a multiplicative commutator, that is, d(a, b) = det(ABA−1B−1). This

immediately yields the remarkable fact that the determinant of the multiplicative commutator

depends only on the Steinberg symbol in K-theory. See also the paper of J. W. Helton and

R. Howe [32] for work in this area around the same time.

As an example, consider two non-vanishing smooth functions f and g on the unit circle. Then

Page 9: Determinants in K-theory and operator algebras

2

one may form the Toeplitz operators Tf and Tg, which are compressions of multiplication operators

by f and g on L2(S1) to the Hardy space H2(S1) (Section 2.3). Since f and g are non-vanishing

and continuous, it turns out that Tf and Tg are Fredholm. Moreover, since f and g are smooth,

we shall see that Tf and Tg commute with each other modulo trace class. In [15] R. Carey and J.

Pincus obtain an integral formula for the determinant invariant of Tf and Tg:

d(Tf + L1, Tg + L1) = exp1

2πi

(∫S1

log f d(log g)− log g(p)

∫S1

d(log f)

). (1.1)

The integrals are taken counterclockwise starting at any point p ∈ S1. If h(eiθ) = |h(eiθ)|eiφ(θ) for

a continuous real-valued function φ, then we take log h(eiθ) = log |h| + iφ(θ). Any other choice of

log h will differ from this one by a multiple of 2πi and hence will leave the quantity in the formula

unaffected.

Now let A and B be commuting operators with images a and b in B/L1. In [13] Carey and

Pincus expressed the determinant invariant d(a − z1, b − z2) in terms of multiplicative Lefschetz

numbers of the form

det(B − z2)|ker(A−z1)

det(B − z2)|coker (A−z1)

det(A− z1)|coker (B−z2)

det(A− z1)|ker(B−z2)(1.2)

whenever a− z1 and a− z2 are invertible and (z1, z2) /∈ σT (A,B), the Taylor joint spectrum of A

and B. However, the determinant invariant is defined even if (z1, z2) ∈ σT (A,B). Thus, Carey and

Pincus introduce an invariant τ(A,B), known as joint torsion, for any two commuting Fredholm

operators A and B. This generalizes the multiplicative Lefschetz number in (1.2). By replacing A

by exp(A) and B by I, joint torsion recovers the Fredholm index of A, so the subject of this thesis

may be seen as a type of multiplicative index theory.

In order to define joint torsion, Carey and Pincus use the notion of algebraic torsion from

homological algebra. For any finite length exact sequence (V•, d•) of finite dimensional vector

spaces, the algebraic torsion τ(V•, d•) is a canonical generator of the determinant line

detV• =(

ΛdimVnVn

)∗⊗(

ΛdimVn−1Vn−1

)⊗ · · ·

Section 3.2.1 reviews the details of this construction. For example, the torsion of an automorphism

of a finite dimensional vector space is its determinant. As another example, J.-M. Bismut, H. Gillet,

Page 10: Determinants in K-theory and operator algebras

3

and C. Soule have shown in [6] that the Ray-Singer analytic torsion can be calculated as the norm

of τ(V•, d•).

J. Kaad has generalized the notion of joint torsion to n ≥ 2 commuting operators [37].

Moreover, he shows that joint torsion is multiplicative, satisfies cocycle identities, and is trivial

under appropriate Fredholm assumptions. He has also investigated the relationship between joint

torsion on the one hand, and determinant functors and K-theory of triangulated categories on

the other. In the case of n = 2 commuting operators, he has shown in [36] that the determinant

invariant coincides with the Connes-Karoubi multiplicative character [20]. Furthermore, Kaad has

constructed a product in relative K-theory and investigated the relative Chern character with

values in continuous cyclic homology [35]. He uses these results to calculate the multiplicative

character applied to Loday products of exponentials. See also Section 3.2 below for a discussion of

this commuting case.

Carey and Pincus extended their definition of joint torsion, in a different direction, to two

almost commuting Fredholm operators [16]. Thus, let A and B be Fredholm operators with trace

class commutator, and moreover, assume the existence of operators C and D such that AB = CD,

A − D ∈ L1, and B − C ∈ L1. They then proceed as before, defining τ(A,B,C,D) in terms of

short exact sequences of Koszul complexes. However, the result is no longer a scalar, but rather an

element of a certain determinant line. To obtain a scalar, Carey and Pincus associate a perturbation

vector σA,A′ to each pair of Fredholm operators A and A′ with A − A′ ∈ L1. See Section 3.3 for

details on the constructions. More recently, J. Kaad and R. Nest have generalized the notion of

perturbation vectors to finite rank perturbations of Fredholm complexes [39].

Perturbation vectors can be seen as a generalization to the singular setting of the classical

perturbation determinant det(A−1A′). Carey and Pincus have shown in [16] that perturbation

vectors form a non-vanishing section of a Quillen determinant line bundle [48]. Moreover, they

have applied joint torsion to Toeplitz operators, especially problems where standard techniques

only apply to symbols with zero winding number. They prove Szego-type limit theorems on the

asymptotic behavior of determinants of Toeplitz operators whose symbols have nonzero winding

Page 11: Determinants in K-theory and operator algebras

4

number [16]. See Section 3.1 for a discussion of this and other related work by Carey and Pincus.

In the case when f, g ∈ H∞(S1), the joint torsion τ(Tf , Tg) is the product of tame symbols [15],

and can be expressed in terms of Deligne cohomology [23]. In particular, the determinant invariant

d(Tf + L1, Tg + L1) is equal to the joint torsion τ(Tf , Tg) when f and g are smooth functions in

H∞(S1). More generally, Carey and Pincus state in [16, Section 8, p. 345]:

The existence of the identification map (in the existence theorem for the pertur-

bation vector) has uncovered a basic problem – which deserves to be stated as a

question or perhaps as a conjecture:

Let a, b be commuting units in B(H)/L1(H). Let A,B,C,D be elements in B(H)

so that AB = CD and let A,D denote lifts of a and B,C denote lifts of b. In what

generality is it true that det ∂a, b = τ(A,B,C,D)?

This is quickly seen to be true for invertible operators A, B, C, and D [16, Section 8]. Carey

and Pincus proved in [13] that this is also true for commuting Fredholm operators A and B with

acyclic Koszul complex K•(A,B). We will see later that this follows from (1.2). More generally,

they have shown in unpublished work that d(a, b) = τ(A,B,B,A) for any commuting Fredholm

operators A and B (see [15, Theorem 2]). One of the main results of this thesis is Theorem 4.3.3,

which answers the above question in full generality.

Let us take a moment to outline the strategy, which is carried out in Chapter 4. Since the

determinant invariant is trivial in finite dimensions, the first order of business is to show that this

true for joint torsion as well. This is done in Section 4.1, closely following the work of Kaad [37].

Our main tool is the double complex (4.4), which is composed of the homology spaces of modified

Koszul complexes. The resulting vertical and horizontal torsion vectors agree, up the sign of a

permutation which appears in the definition of joint torsion. By picking generators carefully, we

will see that these vectors comprise the joint torsion, which is consequently trivial.

Then in Section 4.2 we consider multiplicative properties of perturbation vectors and joint

torsion with respect to invertible operators. This will allow us to investigate the transformation

Page 12: Determinants in K-theory and operator algebras

5

of joint torsion under certain group actions in Section 4.3. More specifically, consider a quadruple

(A,B,C,D) where A, B, C, and D are Fredholm operators such that

A−D ∈ L1, B − C ∈ L1, and AB = CD.

Let U and V be invertible operators such that U − V ∈ L1 and the commutator [U,B] ∈ L1 as

well. The set of all such pairs (U, V ) forms a group which acts on quadruples by

(A,B,C,D) • (U, V ) = (U−1A,B,U−1CU, V −1D).

We will see that joint torsion transforms according to the rule

τ((A,B,C,D) • (U, V )) = d(U−1 + L1, B + L1) · τ(A,B,C,D).

A similar formula holds for pairs of invertibles that commute with the first operator A modulo L1.

These group actions will allow us to reduce the general case to operators A,B,C,D in the coset of

the identity modulo finite rank operators. The proof of Theorem 4.3.3 then follows quickly from

the finite dimensional case.

In Section 4.4 we record a number of consequences of this equality. On the one hand, joint

torsion is a finite dimensional object, at least for commuting operators. The determinant invariant

on the other hand is defined in terms of the infinite Fredholm determinant. Hence it is somewhat

surprising that the determinant invariant turns out to be expressible in terms of finite dimensional

data. Moreover, we will see that joint torsion enjoys continuity properties and satisfies the usual

Sternberg relations (e.g. it is multiplicative and skew-symmetric). Finally, we will use these results

to investigate the behavior of joint torsion with respect to Hilbert-Schmidt operators.

The second main goal of this thesis is to study the transformation of joint torsion under the

functional calculus. J. Kaad and R. Nest have investigated local indices of commuting tuples of

operators under the holomorphic functional calculus [38], and they obtain a global index theorem

originally due to J. Eschmeier and M. Putinar [29]. In Section 5.1, we establish a multiplicative

analogue of such transformation rules: the joint torsion τ(f(A), B) is given by∏λ∈σ(A) | f(λ)=0

τ(A− λ,B)ordλ(f) · τ(q(A), B).

Page 13: Determinants in K-theory and operator algebras

6

Here q(A) is an invertible operator, so the second factor may be viewed as a type of multiplicative

Lefschetz number. In addition, we investigate variational formulas for joint torsion (Corollaries

4.4.5, 4.4.9, and 5.1.15).

In [26], T. Ehrhardt generalizes the Helton-Howe-Pincus formula on determinants of expo-

nentials (Proposition 2.1.13) by showing that

eAeB − eA+B ∈ L1 (1.3)

whenever [A,B] ∈ L1, and moreover,

det(eAeBe−A−B

)= e

12

tr[A,B].

Now let P : L2(S1) → H2(S1) be the orthogonal projection onto the Hardy space (Section 2.3).

Let φ ∈ L∞(S1). With A = T(I−P )φ and B = TPφ, and under suitable regularity assumptions on

φ, one may use (1.3) to show that

Teφ − eTφ ∈ L1. (1.4)

We investigate the following question: to what extent does (1.4) hold with the exponential

replaced by other functions? In Section 5.2 we consider entire functions in the more general setting

of finitely summable Fredholm modules, before specializing to Toeplitz operators in Section 5.3. In

particular, Theorem 5.3.12 establishes (1.4) when

(1) f is holomorphic on a neighborhood of σ(Tφ), or

(2) φ is real-valued and f is C∞ on φ(S1).

Along the way, we investigate the functional calculus modulo ideals of compact operators. Under

suitable assumptions on f , we find

(1) f(A)− f(A′) ∈ Lp if A−A′ ∈ Lp (Proposition 5.1.6)

(2) [f(A), B] ∈ Lp if [A,B] ∈ Lp (see Proposition 5.1.2)

(3) Tf(φ) − f(Tφ) ∈ Lp in a 2p-summable Fredholm module (Proposition 5.2.7)

Page 14: Determinants in K-theory and operator algebras

7

Thus we obtain functional calculi on the Calkin-type algebra B/Lp of bounded operators B modulo

the Schatten ideal Lp of compact operators with p-summable singular values. Result (2) is due

to A. Connes [19]. See also J. W. Helton and R. Howe’s work [32] for the self-adjoint case. We

also obtain expressions for the trace and estimates on the Lp-norms of operators as above, and we

apply these results to obtain the integral formula (1.1) for the joint torsion τ(Tf , Tg) of Toeplitz

operators Tf and Tg.

In Section 5.3 we obtain explicit formulas for determinants of Toeplitz operators. First,

we recall the notion of a symbol in arithmetic [58], which is a bimultiplicative map c(·, ·) on the

multiplicative group of a field such that c(a, 1 − a) = 1. One example is Brown’s determinant

invariant described above. Another example is the Tate tame symbol on a field of meromorphic

functions. This symbol, denoted ca(f, g), is defined as a weighted ratio of two functions f and g

evaluated at a point a (Definition 5.3.3). It turns out that the tame symbol is closely related to

the determinant invariant. Indeed, Theorem 5.3.11 expresses the joint torsion of Toeplitz operators

in terms of their tame symbols. This extends a result due to R. Carey and J. Pincus [15]: if

f, g ∈ H∞(S1), then the joint torsion of Tf and Tg is the product of tame symbols

τ(Tf , Tg) =∏|a|<1

ca(f, g).

Let us conclude by noting that J. Kaad and R. Nest have recently investigated the local behavior of

joint torsion transition numbers associated to commuting tuples of operators [40]. They generalize

the above Carey-Pincus formula in a different direction and extend the notion of tame symbol to

the setting of transversal functions on a complex analytic curve.

Page 15: Determinants in K-theory and operator algebras

Chapter 2

Preliminaries

In this chapter we introduce the basic objects of study in this thesis and collect a number of

results that will prove useful later. First we define the Fredholm determinant, a generalization of

the matrix determinant to infinite dimensions. Next we review the Fredholm index, which may be

viewed as a measure of uniqueness and existence of solutions to vector equations. Then in Section

2.2 we discuss algebraic K-theory and use the Fredholm determinant to define one of main objects

of study in this thesis, the determinant invariant. Section 2.3 reviews Toeplitz operators, which

provide a rich source of examples, and indeed motivation, for our work.

2.1 Fredholm theory

2.1.1 Schatten classes

Let us begin by reviewing some classes of compact operators which will be important for this

thesis, namely the Schatten ideals Lp of p-summable operators. Our reference for this section is

[51].

Fix a separable complex Hilbert space H. Let B denote the algebra of bounded operators

on H and let K be the ideal of compact operators. If K ∈ K, then by the spectral theorem for

compact self-adjoint operators, K∗K is diagonalizable with respect to a basis of eigenvectors. The

eigenvalues µk (listed according to algebraic multiplicity and known as the singular values of K)

are non-negative and have no accumulation points, except possibly at zero.

Page 16: Determinants in K-theory and operator algebras

9

Definition 2.1.1. Let 0 < p <∞. A compact operator K is in the Schatten p-class Lp if

∑µp/2 <∞.

Elements of L1 are known as trace class operators, and elements of L2 are known as Hilbert-

Schmidt operators. Notice that every Lp contains the finite rank operators.

Proposition 2.1.2.

(1) The Schatten classes Lp are two-sided ideals of B.

(2) If p ≥ 1, then ‖K‖p =(∑

µp/2)1/p

defines a norm on Lp.

(3) If A ∈ Lp and B ∈ B, then ‖AB‖p ≤ ‖A‖p‖B‖.

(4) (Lp, ‖ · ‖p) is a Banach space.

Example 2.1.3.

(1) Let ei∞i=0 be an orthonormal basis for H. Let S be the unilateral shift Sei = ei+1. Then

S∗e0 = 0

S∗ei = ei−1, i > 0

The self-commutator [S∗, S] is evidently the rank one projection onto the span of e0, and

hence is in every Schatten class. More generally, we will see below that commutators

of Toeplitz operators are in the trace class, under conditions on the smoothness of their

symbols.

(2) Let X be a locally compact Hausdorff space with a positive Borel measure. Assume that

L2(X) is separable. If k ∈ L2(X ×X), then the operator K defined on L2(X) by

(Kf)(x) =

∫Xk(x, y)f(y) dy

is a Hilbert-Schmidt operator with ‖K‖2 = ‖k‖L2.

Page 17: Determinants in K-theory and operator algebras

10

(3) A pseudodifferential operator T on a closed n-dimensional manifold M is trace class if its

order is less than −n. More generally, T ∈ Lp if ord T < −n/p.

Remark 2.1.4. In light of the example above, it may be of interest in the future to investigate the

subject of this thesis in the context of the ideals

Lp− =⋃q<p

Lp.

Definition 2.1.5. Let L ∈ L1 be a trace class operator on a Hilbert space, say with orthonormal

basis ei. The trace of L is defined to be

trL =∑〈Lei, ei〉.

Notice that in finite dimensions this recovers the usual matrix trace.

Proposition 2.1.6.

(1) If L ∈ L1, then trL is well-defined, that is, its defining sum converges and does not depend

on the choice of basis.

(2) The trace is a continuous linear functional on (L1, ‖ · ‖1).

(3) If A ∈ L1 and B ∈ B, then tr(AB) = tr(BA).

(4) If A,B ∈ L2, then tr(AB) = tr(BA).

Example 2.1.7. The trace of a pseudodifferential operator T ∈ L1 from Example 2.1.3(3) is given

by integrating its Schwarz kernel along the diagonal (see e.g. [43, Section 4.1]:

trT =

∫MKT (x, x) dx.

2.1.2 Fredholm determinant

The goal of this section is to describe the Fredholm determinant, which extends the usual

matrix determinant to the coset I + L1. Our reference for this section and the next is [62].

Page 18: Determinants in K-theory and operator algebras

11

In order to define the Fredholm determinant, we will need the notion of exterior power of a

vector space H. The k-th exterior power ΛkH is the vector space spanned by elements of the form

v1 ∧ · · · ∧ vk, vi ∈ H

subject to the relation

vσ(1) ∧ · · · ∧ vσ(k) = (−1)sign(σ)v1 ∧ · · · ∧ vk

for any permutation σ on the set 1, · · · , k. See [51, Section 1.5] for more details. By definition,

Λ0H = C. If H is finite dimensional, then ΛkH is evidently trivial for k > dimH. An inner product

〈·, ·〉 on H induces an inner product on ΛkH given by

〈v1 ∧ · · · ∧ vk, w1 ∧ · · · ∧ wk〉 = det(〈vi, wj〉).

If H is a Hilbert space, then ΛkH is as well. Moreover, any operator T ∈ B(H) induces an operator

ΛkT ∈ B(ΛkH) by

ΛkT (v1 ∧ · · · ∧ vk) = Tv1 ∧ · · · ∧ Tvk.

Definition 2.1.8. Let L ∈ L1(H). The Fredholm determinant of I + L is given by

det(I + L) =∞∑k=0

tr(

ΛkL).

One readily checks that ΛkL ∈ L1(ΛkH) and that the above series converges. In fact, we

have the following:

Proposition 2.1.9. The function z 7→ det(I + zL) is entire with

| det(I + zL)| ≤ e|z|‖L‖1 .

Corollary 2.1.10. The map L 7→ det(I + L) is continuous on L1.

In finite dimensions, the Fredholm determinant recovers the usual determinant of a matrix.

Indeed, let V be a finite dimensional vector space with basis ei. Let L be a linear transformation

Page 19: Determinants in K-theory and operator algebras

12

on V , and let T = U−1LU be its Jordan canonical form, with eigenvalues ai, repeated according

to algebraic multiplicities. Then

ΛkT (ei1 ∧ · · · ∧ eik) = ai1 · · · aik(ei1 ∧ · · · ∧ eik)

and consequently,

tr(

ΛkL)

= tr(

ΛkT)

=∑

i1<···<ik

ai1 · · · aik .

Hence the Fredholm determinant of I + L is

∑k

∑i1<···<ik

ai1 · · · aik .

On the other hand, one calculates the determinant as

det(I + L) =∏i

(1 + ai)

which agrees with the Fredholm determinant above.

Remark 2.1.11. Indeed, an equivalent and perhaps more familiar definition of the Fredholm de-

terminant is given by

det(I + L) =∏i

(1 + λi)

where λi are the eigenvalues of L, repeated according to multiplicity.

Let G denote the set of of invertible operators in the coset I + L1, with the operation of

composition. That G is closed under taking inverses is immediate from the identity

(I + L)−1 = I − L(I + L)−1.

Proposition 2.1.12.

(1) If L ∈ L1, then eL ∈ G and

det(eL)

= etrL.

(2) If A ∈ I + L1 and X is invertible, then X−1AX ∈ I + L1 and

det(X−1AX) = detA.

Page 20: Determinants in K-theory and operator algebras

13

(3) If A,B ∈ I + L1, then AB ∈ I + L1 and

det(AB) = detA · detB.

(4) Let A ∈ I + L1. Then detA 6= 0 if and only if A ∈ G.

Proposition 2.1.13 (Helton-Howe-Pincus). If A and B are operators such that [A,B] ∈ L1, then

det(eAeBe−Ae−B) = etr[A,B].

2.1.3 Fredholm index

On the coset I+L1 considered above, or more generally I+K, one has the famous Fredholm

alternative, one formulation of which may be stated as follows: if K is compact, then I + K is

either invertible or has nontrivial kernel and cokernel. This may be viewed as a manifestation of

the fact that I +K is a Fredholm operator with index zero, which we describe next.

Definition 2.1.14. An operator A on a Hilbert space is known as a Fredholm operator if its kernel

and cokernel are finite dimensional. In this case, the index of A is the difference of these two

dimensions:

indA = dim kerA− dim cokerA.

Proposition 2.1.15 (Atkinson’s criterion). The following are equivalent:

(1) A is Fredholm.

(2) There exist operators Q and R such that I −AQ and I −RA are compact.

(3) There exist operators Q and R such that I −AQ and I −RA have finite rank.

Property (2) above may be rephrased as saying that A is invertible in the Calkin algebra

B/K. Property (3) asserts that the same is true even modulo finite rank, and hence modulo any

operator ideal.

Proposition 2.1.16.

Page 21: Determinants in K-theory and operator algebras

14

(1) The set of Fredholm operators is open in B with respect to the operator norm.

(2) The map A 7→ indA on the set of Fredholm operators is continuous, and hence, locally

constant. In fact, the collections

Uk = A | indA = k

are precisely the connected components of the set of Fredholm operators in B.

(3) If A and B are Fredholm operators, then AB and A⊕B are Fredholm with

ind(AB) = ind(A⊕B) = indA+ indB.

(4) If A is Fredholm and K is compact, then A+K is Fredholm with

ind(A+K) = indA.

2.2 Algebraic K-theory

For any unital ringR and ideal J , there are algebraicK-groupsKi(R), Ki(R/J), andKi(R, J)

that fit into Quillen’s long exact sequence

· · · → Ki+1(R/J)∂−→ Ki(R, J)→ Ki(R)→ Ki(R/J)

∂−→ . . .

For the purposes of this thesis, we will mainly be interested in the groups K1(R, J) and K2(R/J),

as well as the boundary map between them. This section reviews the relevant definitions, closely

following [50].

2.2.1 Algebraic K0

For the sake of completeness, let us begin with a discussion of K0. Recall that for a ring R,

a projective R-module can be characterized as an R-module that embeds as a direct summand in a

free R-module. The set of isomorphism classes of finitely generated projective R-modules forms a

semigroup Proj(R) under the operation of direct sum. In fact, the assignment R→ Proj(R) defines

Page 22: Determinants in K-theory and operator algebras

15

a covariant functor from rings to abelian semigroups since a ring homomorphism ϕ : R → R′

induces a map

ϕ : Proj(R)→ Proj(R′), [M ] 7→ [R′ ⊗ϕM ].

Then one canonically obtains a group K0(R) via the Grothendieck construction. Indeed, for any

abelian semigroup (S,+), let G(S) be the abelian group generated by [x], x ∈ S, subject to the

relations

[x] + [y] = [z] if x+ y = z in S.

This construction S → G(S) defines a covariant functor since a homomorphism of semigroups

ϕ : S → S′ defines a group homomorphism

ϕ : G(S)→ G(S′), [x] 7→ [φ(x)].

Definition 2.2.1. K0(R) = G(Proj(R)).

Example 2.2.2. Finitely generated projective modules over a field F are simply finite dimensional

F -vector spaces. Up to isomorphism, these are classified by the dimension – a non-negative integer.

Hence K0(F ) = Z.

On the other hand, one may also consider the topological K-theory of a compact topological

space X. The group K0(X) is defined to be the Grothendieck group of the semigroup of finite rank

vector bundles over X. By the Serre-Swan Theorem [52], the category of finite rank vector bundles

over X is naturally equivalent to the category of finitely generated projective modules over C(X).

Hence we find

K0(C(X)) = K0(X).

In higher degrees, however, topological and algebraic K-theory differ in general. For example, one

may define topological K-theory on the category of C∗-algebras. The algebraic group K1, which

we describe next, turns out to be a quotient of the topological group Ktop1 .

Page 23: Determinants in K-theory and operator algebras

16

2.2.2 Algebraic K1

Let R be a unital ring and let GLn(R) be the group of invertible n × n matrices over R.

Define the group GL(R) to be the direct limit of the groups GLn(R) under the embedding

GLn(R) → GLn+1(R), a 7→

a 0

0 1

.

Let En(R) ⊂ GLn(R) be the subgroup of elementary matrices, which are generated by matrices

that differ from the identity by at most one off-diagonal entry. Then one has an embedding

En(R) → En+1(R) as above, and hence a direct limit E(R) ⊂ GL(R).

Proposition 2.2.3 (see Proposition 2.1.4 in [50]). The commutator subgroup [GL(R), GL(R)] =

E(R). In particular, E(R) is normal in GL(R).

Definition 2.2.4. For a unital ring R, define

K1(R) =GL(R)

E(R).

By the preceding proposition, K1(R) coincides with the abelianization GL(R)ab of GL(R).

The assignment R → K1(R) defines a covariant functor from the category of unital rings to

the category of abelian groups. Indeed, a morphism R → S induces a map GLn(R) → GLn(S),

and hence a map on the direct limits GL(R)→ GL(S) as well as their abelianizations GL(R)ab →

GL(S)ab.

Example 2.2.5 (see Section 2.2 of [50]). If R is a commutative ring, the determinant GLn(R)→

R× extends to a surjection GL(R)→ R× and hence to a surjection

det : K1(R)→ R×.

If R is a field, this is in fact an isomorphism.

Let GL(R, J) denote the kernel of the group homomorphism GL(R) → GL(R/J) induced

by the quotient map R → R/J . Denote by E(R, J) the subgroup of elementary matrices gen-

erated by matrices that differ from the identity by at most one off-diagonal element of J . Let

Page 24: Determinants in K-theory and operator algebras

17

[GL(R), E(R, J)] ⊆ GL(R, J) denote the subgroup generated by all elements of the form

ghg−1h−1, g ∈ GL(R), h ∈ E(R, J).

This subgroup is normal in GL(R, J), and we make the following definition:

Definition 2.2.6. K1(R, J) = GL(R, J)/[GL(R), E(R, J)].

2.2.3 Algebraic K2

The Steinberg group Stn(R) is the group with generators xij(a) for a ∈ R, i 6= j, 1 ≤ i, j ≤ n,

and relations

xij(a)xij(b) = xij(a+ b)

xij(a)xkl(b)xij(a)−1xkl(b)−1 = 1, j 6= k, i 6= l

xij(a)xjk(b)xij(a)−1xjk(b)−1 = xik(ab), i, j, k distinct

Since the generators of the group of elementary matrices En(R) satisfy these same relations, we

have a map Stn(R)→ En(R), and hence a natural map φ on the inductive limits St(R) and E(R).

Definition 2.2.7. K2(R) = ker(φ : St(R)→ E(R)).

In fact, St(R) is the universal central extension of E(R), and K2(R) is isomorphic to the

second homology group H2(E(R),Z) [50, Theorem 4.2.7 and Corollary 4.2.10].

Next we define specific elements in K2(R) known as Steinberg symbols, which turn out to be

quite useful. For example, K2 of a field is generated by its Steinberg symbols. For any a ∈ R× and

i 6= j, define elements wij(a), hij(a) ∈ St(R) by

wij(a) = xij(a)xji(−a−1)xij(a), hij(a) = wij(a)wij(−1).

Then we have

φ(w12(a)) =

0 a

−a−1 0

and φ(h12(a)) =

a 0

0 a−1

.

Page 25: Determinants in K-theory and operator algebras

18

Definition 2.2.8. For commuting units a, b ∈ R, the Steinberg symbol a, b ∈ K2(R) is

a, b = h12(a)h13(b)h12(a)−1h13(b)−1.

Proposition 2.2.9 (see Lemma 4.2.14 and Theorem 4.2.17 of [50]). For any units a, a′, and b in

R such that b commutes with a and a′, the following identities hold:

(1) a, b = b, a−1

(2) aa′, b = a, ba′, b

(3) a, 1− a = 1, whenever 1− a is invertible

(4) a,−a = a, 1 = 1

The boundary map ∂ : K2(R/J) −→ K1(R, J) is defined as follows. Any element u ∈ K2(R/J)

can be expressed in terms of generators xij(uk) ∈ St(R/J). We obtain an element r ∈ St(R) by

lifting each uk ∈ R/J to an element rk ∈ R. Then φ(r) ∈ GL(R, J), and we define ∂(u) to be the

image of φ(r) in K1(R, J). One then checks that this is independent of the choice of lifts.

2.2.4 The determinant invariant

In the case when R = B, the ring of bounded linear operators on a Hilbert space, and

J = L1, the ideal of trace class operators, the Fredholm determinant induces a surjective group

homomorphism

det : K1(B,L1)→ C×.

Indeed, let g ∈ GL(B) and let h ∈ E(B,L1). Then h is represented by a determinant class operator,

and by Proposition 2.1.12,

det(ghg−1h−1) = det(ghg−1) det(h−1).

The first factor is deth, again by Proposition 2.1.12, so det(ghg−1h−1) = 1 as desired. In fact,

K1(B,L1) = V ⊕ C×, where V is the additive group of a vector space of uncountable linear

dimension, and det is the projection onto the second factor [1].

Page 26: Determinants in K-theory and operator algebras

19

Definition 2.2.10. For any commuting units a, b ∈ B/L1, the determinant invariant d(a, b) is the

nonzero number

d(a, b) = det ∂a, b.

In particular, d(a, b) satisfies the relations in Proposition 2.2.9.

In calculating the determinant invariant, a ∈ B/L1 is lifted to an operator A ∈ B, which is

necessarily Fredholm. Moreover, a−1 is lifted to a parametrix Q of A modulo trace class. Here,

and in the sequel, a parametrix modulo an ideal is an inverse modulo that ideal. Thus I −AQ and

I − QA are trace class operators. When a and b have invertible lifts, Brown made the following

observation in [8]:

Proposition 2.2.11. Let a and b be commuting units in B/L1. If a and b have invertible lifts

A ∈ B and B ∈ B, respectively, then d(a, b) = detABA−1B−1. In particular, the determinant of

the multiplicative commutator depends only on the Steinberg symbol a, b.

Proof. Since a−1 and b−1 can be lifted to A−1 and B−1, we have

d(a, b) = det ∂ h12(σ(A))h13(σ(B))h12(σ(A))−1h13(σ(B))−1

= det

A 0 0

0 A−1 0

0 0 I

B 0 0

0 I 0

0 0 B−1

A 0 0

0 A−1 0

0 0 I

−1

B 0 0

0 I 0

0 0 B−1

−1

= detABA−1B−1

Example 2.2.12. Suppose A = expα and B = expβ for operators α and β with [α, β] ∈ L1. Then

A and B are invertible and [A,B] ∈ L1. By the Helton-Howe-Pincus formula (Proposition 2.1.13),

d(a, b) = exp tr [α, β].

Lemma 2.2.13. For any commuting units ai and bi in B/L1, i = 1, 2, we have

d(a1 ⊕ a2, b1 ⊕ b2) = d(a1, b1) d(a2, b2).

Page 27: Determinants in K-theory and operator algebras

20

Proof. In calculating the boundary map, ai is lifted to a Fredholm operator Ai, and a−1i is lifted to

any parametrix Qi of Ai modulo trace class. Then one may lift a1⊕a2 to A1⊕A2, and (a1⊕a2)−1

to Q1 ⊕ Q2. Similarly for bi, b1 ⊕ b2, and (b1 ⊕ b2)−1. The result then follows since determinants

are multiplicative over direct sums.

In fact, the determinant invariant can always be calculated in terms of a multiplicative

commutator. To see this, let a, b ∈ B/L1 be invertible commuting elements, and pick lifts A and B

in B of a and b, respectively. Let QA be an operator with index opposite that of A. For example, we

may take QA to be a parametrix for A, a unilateral shift, or the adjoint A∗. Pick QB similarly. Then

A⊕QA⊕I has index zero, so we may pick a finite rank operator FA such that A = A⊕QA⊕I+FA

is invertible. Form B = B ⊕ I ⊕ QB + FB similarly. Then [A, B] ∈ L1(H ⊕ H ⊕ H), and by the

preceding lemma, we immediately obtain the following:

Proposition 2.2.14. With A and B as above, one has

d(a, b) = det(ABA−1B−1).

Example 2.2.15. Let H be a Hilbert space with orthonormal basis ei∞i=0, and let S be the uni-

lateral shift Sei = ei+1. Let F2 and F3 be the rank one operators on H⊕H⊕H such that

F2(0, e0, 0) = (e0, 0, 0) and F3(0, 0, e0) = (e0, 0, 0).

Then S⊕S∗⊕I+F2 and S⊕I⊕S∗+F3 are invertible. By the preceding proposition, one calculates

d(S + L1, S + L1) = −1.

By replacing the first argument by S∗S, Proposition 2.2.9 (2) and (4) imply that d(S∗, S) = −1.

Similarly, d(S∗, S∗) = −1. By a similar technique, one may show more generally that for any

invertible a ∈ B/L1,

d(a, a) = (−1)ind a

where ind a is the index of any lift of a to B.

Page 28: Determinants in K-theory and operator algebras

21

2.3 Toeplitz operators

A special class of operators known as Toeplitz operators will provide a rich source of examples

in this thesis. For example, their index and spectral properties are particularly well-understood

(Section 2.3.1). Furthermore, as we shall see in Section 2.3.2, Toeplitz operators are prototypical

examples of operators with trace class commutators. Our reference in this section is [24, Chapter

7].

To begin, consider the Hilbert space L2(S1) of square-integrable functions on the unit circle

S1 with respect to the Lebesgue measure. The trigonometric monomials

zn, −∞ < n <∞

form a basis for L2(S1). The Hardy space H2(S1) is the closed span of the basis vectors

zn, n ≥ 0.

Let P : L2(S1)→ H2(S1) be the orthogonal projection. The Hardy space can be characterized in

a number of other ways, for example, as the closure in L2(S1) of the boundary values of functions

that are continuous on the closed unit disk and holomorphic on the interior.

Definition 2.3.1. Let f ∈ L∞(S1). The Toeplitz operator Tf : H2(S1)→ H2(S1) is the compres-

sion to H2(S1) of multiplication by f :

Tf (φ) = P (fφ).

One may check that Toeplitz operators satisfy the following properties:

(1) ‖Tf‖ = ‖f‖∞

(2) Tλf+g = λTf + Tg

(3) T ∗f = Tf

Example 2.3.2. The Toeplitz operator Tz acts as a unilateral shift on H2(S1) since

Tz(zn) = zn+1.

Page 29: Determinants in K-theory and operator algebras

22

2.3.1 Index and spectral theory

Let T be the C∗-algebra generated by the compact operators and Toeplitz operators Tf with

continuous symbol f . Equivalently, T is the C∗-algebra generated by the unilateral shift Tz. Thus

we have a C∗-algebra extension of C(S1) by the compact operators:

0→ K → T → C(S1)→ 0.

The following index theorem may be interpreted in terms of the boundary map of the resulting

six-term cyclic exact sequence in K-theory, although it long predates these developments.

Theorem 2.3.3 (F. Noether). A Toeplitz operator Tφ is Fredholm if and only if φ is continuous

and non-vanishing. In this case, the index of Tφ is opposite the winding number of φ:

indTφ = − 1

2πi

∫dφ

φ.

This formula can be verified, for example, by calculating the commutator of a Toeplitz opera-

tor with a shift operator, approximating continuous functions by polynomials, and using homotopy

invariance of the index. See [34] for more details.

We will need the following well-known results on the spectra of multiplication operators and

Toeplitz operators in Chapter 5.

Proposition 2.3.4.

(1) If φ ∈ L∞(S1), the spectrum of the multiplication operator φ ∈ B(L2) is the essential range

of φ.

(2) If φ ∈ C(S1), the essential spectrum of the Toeplitz operator Tφ ∈ B(H2) is the range

φ(S1). The spectrum of Tφ in addition consists of the connected components of C− φ(S1)

about which φ has nonzero winding number.

2.3.2 Commutators of Toeplitz operators

Lemma 2.3.5. If φ ∈ L∞(S1) is in the Sobolev space W12,2(S1) = H

12 (S1), then

(I − P )φP, Pφ(I − P ), [φ, P ] ∈ L2(L2(S1)).

Page 30: Determinants in K-theory and operator algebras

23

with

‖[φ, P ]‖2 ≤ ‖φ‖W

12 ,2(S1)

.

Proof. Write φ =∑cnz

n. A straightforward calculation shows that (I − P )φP ∈ L2, with

‖(I − P )φP‖22 =∑n>0

n|cn|2,

By taking adjoints, Pφ(I − P ) ∈ L2 as well, with

‖Pφ(I − P )‖22 = −∑n<0

n|cn|2.

Hence [φ, P ] = (I − P )φP − Pφ(I − P ) ∈ L2, and

‖[φ, P ]‖22 =∑n 6=0

|n||cn|2.

In this case, Toeplitz operators have trace class commutators, and the Berger-Shaw formula

calculates this trace:

Theorem 2.3.6. If f, g ∈ L∞(S1) ∩W12,2(S1), then [Tf , Tg] ∈ L1. If f, g ∈ C1(S1), then

tr[Tf , Tg] =1

2πi

∫f dg.

Proof. First notice that [Tf , Tg] = Pg(I − P )fP − Pf(I − P )gP . Both terms are trace class since

they are products of two operators which are Hilbert-Schmidt by the preceding lemma. The trace

formula then follows by writing f and g in the basis einθ.

Example 2.3.7. For example, if f and g are smooth functions on the unit circle, then the Toeplitz

operators Tef and Teg are Fredholm and commute modulo L1. Then Tef = expTf and Teg = expTg

modulo L1 by Proposition 5.3.12, and one may use the Helton-Howe-Pincus formula and the Berger-

Shaw formula to show that

d(Tef + L1, Teg + L1) = exp1

2πi

∫f dg.

Page 31: Determinants in K-theory and operator algebras

Chapter 3

Joint torsion

The goal of this chapter is to define an invariant known as joint torsion, and to do so in a

way that will be most convenient for our purposes. Moreover, we establish some basic results that

we will need in Chapter 4 to prove that joint torsion agrees with Brown’s determinant invariant.

Since joint torsion was introduced by Carey and Pincus, let us begin by discussing some of their

seminal work in this area.

3.1 Background: The work of Carey and Pincus

In a series of papers beginning with [11], Carey and Pincus carried out a study of almost

normal operators, that is, operators T with trace class self-commutator: [T, T ∗] ∈ L1. Notice that

we may write T = X + iY , for self-adjoint operators

X =T + T ∗

2and Y =

T − T ∗

2i

and [X,Y ] ∈ L1. Conversely, if X and Y are self-adjoint operators with [X,Y ] ∈ L1, then T =

X + iY is almost normal. For any polynomial p ∈ C[x, y], the operator p(X,Y ) is well-defined

modulo L1. (One may instead form p(T, T ∗) and carry out an equivalent analysis.) If q is another

polynomial, then the commutator [p(X,Y ), q(X,Y )] has a well-defined trace. Helton and Howe

showed in [32] that there exists a measure mT supported on σ(T ) such that

tr[p(X,Y ), q(X,Y )] =1

2πi

∫p, q dmT

Page 32: Determinants in K-theory and operator algebras

25

where p, q denotes the Poisson bracket

∂p

∂x

∂q

∂y− ∂p

∂y

∂q

∂x.

Moreover, mT is absolutely continuous with respect to Lebesgue measure, and its Radon-Nikodym

derivative is the Pincus principal function [12]. In the case of the unilateral shift, for example, the

Pincus principal function is the characteristic function of the unit disk.

Brown’s determinant invariant provides a multiplicative analogue of the trace of commuta-

tors considered above via the Helton-Howe-Pincus formula. Thus Carey and Pincus turned their

attention to the determinant invariant, first for commuting operators A and B with trivial joint

Koszul homology. As we alluded to in the introduction, they use intricate linear algebra arguments

and a careful choice of parametrices for A and B in [13] to calculate the determinant invariant

d(A+ L1, B + L1) in terms of multiplicative Lefschetz numbers

det(B|kerA)

det(B|cokerA)

det(A|cokerB)

det(A|kerA).

The above expression is not necessarily defined when A and B have nontrivial joint Koszul

homology, even though the determinant invariant is defined. Carey and Pincus were thus led to

introduce a homological invariant known as joint torsion, which generalizes the above expression

to the singular case. In [15] they express the joint torsion of Toeplitz operators Tf and Tg with

f, g ∈ H∞ as a product of tame symbols

∏|a|<1

ca(f, g).

Here the tame symbol ca(f, g) of any meromorphic functions f and g is defined to be the value at

a of the function

(−1)dega(f)·dega(g) fdega(g)

gdega(f).

In [16] Carey and Pincus extend joint torsion to operators A and B with trace class com-

mutator. They use this invariant to obtain generalizations of the classical Szego limit theorems

on the asymptotic behavior of determinants of Toeplitz matrices. Let us begin by recalling the

Page 33: Determinants in K-theory and operator algebras

26

groundbreaking work of Szego in [53]. Let Pn : H2(S1) → H2(S1) be the orthogonal projection

onto the first n basis vectors zk, 0 ≤ k < n. If φ ∈ L∞(S1) with Fourier coefficients φ(k), then

the truncated Toeplitz operator PnTφPn is expressed as a Toeplitz – that is, constant diagonal –

matrix

φ(0) φ(1) φ(2) · · · φ(n− 1)

φ(−1) φ(0) φ(1) · · · φ(n− 2)

φ(−2) φ(−1) φ(0) · · · φ(n− 3)

......

.... . .

...

φ(−n+ 1) φ(−n+ 2) φ(−n+ 3) · · · φ(0)

If φ > 0 and log φ ∈ L1(S1), then Szego showed that

limn→∞

PnTφPnPn−1TPn−1

= exp1

2πi

∫log φ.

Let G(φ) denote the above quantity. If in addition φ′ is Lipschitz continuous, then Szego showed

in [54] that

limn→∞

PnTφPnG(φ)n

= exp∞∑k=1

k log f(k) log f(−k)

in response to a question of Onsager and Yang on spontaneous magnetization in the Ising model

in statistical mechanics.

In [16] Carey and Pincus investigate the more general case when φ has nonzero winding

number. They found that quantities like Brown’s determinant invariant appeared quite naturally

in their analysis. For example,

PnTφPnd(Tφ + L1, Sn+1 + L1)

= det(TφS

∗n+1QφSn+1 + T ∗φS

n+1).

where S = Tz is the unilateral shift and Qφ is a parametrix for Tφ. To aid in these calculations,

Carey and Pincus extended their joint torsion to the case of Toeplitz operators, and more generally,

Fredholm operators that commute modulo the trace class. Joint torsion may be viewed heuristically

as a replacement for the logarithm in Szego’s limit theorem. This invariant is the subject of the

present chapter.

Page 34: Determinants in K-theory and operator algebras

27

3.2 Commuting operators

3.2.1 Algebraic torsion

To define their new invariant, Carey and Pincus use the notion of algebraic torsion, which

we describe in this section. Let (V•, d•) be an exact sequence of finite dimensional vector spaces,

0 −→ Vndn−→ Vn−1

dn−1−−−→ Vn−2 −→ · · · −→ V0 → 0.

Denote by detVk the top exterior power ΛdimVkVk, and define the determinant line

detV• = detV ∗n ⊗ detVn−1 ⊗ detV ∗n−2 ⊗ . . .

For each k, pick a nonzero element tk ∈ Λrank dkVk such that dktk 6= 0. By exactness, dktk ∧ tk−1 is

a nonzero element of detVk−1.

Definition 3.2.1. The torsion τ(V•, d•) of the complex (V•, d•) is the volume element

τ(V•, d•) = (tn)∗ ⊗ dntn ∧ tn−1 ⊗ (tn−2 ∧ dn−1tn−1)∗ ⊗ dn−2tn−2 ∧ tn−3 ⊗ . . .

This is nonzero and is independent of the choices of the tk [41]. Hence τ(V•, d•) defines a

canonical generator of detV•.

For a finite dimensional vector space V , we will make frequent use of the isomorphism detV ∗⊗

detV ∼= C given by

(v∗1 ∧ v∗2 ∧ . . . )⊗ (w1 ∧ w2 ∧ . . . ) 7→ det (v∗i (wj)) . (3.1)

For finite dimensional vector spaces V and W , we also identify the line detV ⊗detW with detW ⊗

detV according to the rule

s⊗ t 7→ (−1)dimV ·dimW t⊗ s. (3.2)

See [40] for a formulation of torsion in terms of the Picard category of graded lines, among other

insights.

Example 3.2.2. The torsion of an automorphism of a finite dimensional vector space is its deter-

minant.

Page 35: Determinants in K-theory and operator algebras

28

3.2.2 Joint torsion of commuting operators

For a collection of commuting operators, Carey and Pincus [15] and Kaad [37] have defined

invariants known as joint torsion. This section reviews their constructions. We begin with the

notion of a mapping cone from homological algebra. Thus, let (V•, dV• ) and (W•, d

W• ) be chain

complexes, i.e.

dVk−1dVk = dWk−1d

Wk = 0

and let f• : V• →W• be a morphism of chain complexes, i.e.

fk−1dVk = dWk fk.

Then the mapping cone (C(f)•, dC• ) is the chain complex with C(f)k = Vk−1 ⊕Wk and

dCk =

−dVk−1 0

fk−1 dWk

.

Now let A = (A1, . . . , An) be a collection of commuting operators on a vector space H. The

Koszul complex K•(A) is the chain complex with

Ki(A) = H⊗ ΛiCn

and differential di : Ki(A)→ Ki−1(A) given by

di =n∑k=1

Ak ⊗ ε∗k

where εk : ΛiCn → Λi+1Cn is the operation of exterior multiplication by the unit vector ek, and ε∗k

is its adjoint. The n-tuple A is said to be Fredholm if (K•(A), d•) has finite dimensional homology.

We also have maps ιk : ΛiCn → ΛiCn+1 induced by the inclusion

Cn → Cn+1, (a1, . . . , an) 7→ (a1, . . . , ak−1, 0, ak, . . . , an).

For any j ∈ 1, . . . , n, let j(A) = (A1, . . . , Aj , . . . , An). Since A = (A1, . . . , An) is commu-

tative, each Aj defines a morphism of chain complexes Aj := Aj ⊗ 1 : K•(j(A))→ K•(j(A)). The

Page 36: Determinants in K-theory and operator algebras

29

mapping cone of this morphism is isomorphic to the Koszul complex K•(A) via the isomorphismι∗jε∗jι∗j

: ΛkCn ∼−→ Λk−1Cn−1 ⊕ ΛkCn−1.

We thus obtain a triangle of chain complexes

K•(j(A))→ K•(j(A))→ K•(A)→ K•(j(A))[1]

and hence a long exact sequence in homology

Ej : 0→ Hn(A)→ Hn−1(j(A)) −→ Hn−1(j(A))→ Hn−1(A)→ . . .

The torsion vector τ(Ej) ∈ detH•(A)⊗detH•(j(A))∗⊗detH•(j(A)). Here, and in the sequel,

we use the notation

detH•(A) = detHn(A)∗ ⊗ detHn−1(A)⊗ . . .

for a sequence of homology spaces H•(A). We regard τ(Ej) ∈ detH•(A) using the isomorphism in

(3.1) applied to the line detH•(j(A)). Carrying out the same process for some i 6= j yields an exact

sequence Ei. If i(A) and j(A) are Fredholm, then Ei and Ej consist of finite dimensional vector

spaces, so we can form two torsion vectors τ(Ei), τ(Ej) ∈ detH•(A). Then τ(Ei) ⊗ τ(Ej)∗ can be

identified with a nonzero scalar, which up to a sign is the joint torsion defined by Carey and Pincus

for n = 2 [15] and by Kaad for n ≥ 2 [37].

3.2.3 The case of two commuting operators

Let us describe this construction more explicitly in the case n = 2. Thus, let A1 = A and

A2 = B be two commuting Fredholm operators. Upon choosing bases for the exterior algebras, we

have Koszul complexes

K•(A) : H A−→ H, K•(B) : H B−→ H

K•(A,B) : H

(−BA

)−−−−→ H2 (A B )−−−−→ H

Then H0(A) = cokerA, H1(A) = kerA, H2(A,B) = kerA∩kerB, H0(A,B) = H/(AH+BH), and

H1(A,B) =(y, z) |Ay +Bz = 0(−Bx,Ax) |x ∈ H

.

Page 37: Determinants in K-theory and operator algebras

30

The two exact sequences EA and EB are given by

EA : 0 −→ (kerA ∩ kerB)−ι−→ kerB

−A−−→ kerB−ι2∗−−−→ H1(A,B)→

π1∗−−→ cokerBA−→ cokerB

π−→ H/(AH+BH) −→ 0

EB : 0 −→ (kerA ∩ kerB)ι−→ kerA

−B−−→ kerA−ι1∗−−−→ H1(A,B)→

π2∗−−→ cokerAB−→ cokerA

π−→ H/(AH+BH) −→ 0

The map ιk∗ is induced by inclusion into the k-th coordinate, and πk∗ is induced by projection onto

the k-th coordinate. Here we have chosen signs to simplify formulas later.

Definition 3.2.3. The joint torsion τ(A,B) of A and B is the nonzero scalar

τ(A,B) = (−1)λ(A,B) τ(EA)⊗ τ(EB)∗.

Here λ(A,B) = sgn θ + γ, where θ is the permutation that maps the factors in (4.6) to (4.5)

and γ is given by (4.7). See Remark 4.1.3. Definition 3.3.4 deals with the more general situation,

i.e. when [A,B] is not necessarily zero, but we may write AB = CD. We note that our sign

conventions differ from those of [16]. In the above definition, τ(A,B) is identified with a scalar

according to (3.1) and (3.2).

Lemma 3.2.4. If Hi(A,B) = 0 for i = 0, 1, 2, then we have

τ(A,B) =detB|kerA

detB|cokerA

detA|cokerB

detA|kerB.

In particular, τ(A, I) = τ(I, A) = 1.

Proof. Since Hi(A,B) = 0, the exact sequence EA breaks up into the isomorphisms

−A|kerB : kerB → kerB and A|cokerB : cokerB → cokerB

and EB breaks up as

−B|kerA : kerA→ kerA and B|cokerA : cokerA→ cokerA.

Page 38: Determinants in K-theory and operator algebras

31

Joint torsion is the alternating product of the torsion vectors of these isomorphisms, which are

simply the determinants by Example 3.2.2. The sign (−1)λ(A,B) is calculated directly.

Example 3.2.5. Joint torsion generalizes the Fredholm index. Indeed, by Lemma 3.2.4,

τ(A, expB) =det expB|kerA

det expB|cokerA

The ratio on the right hand side is seen to be exp tr(B|kerA−B|cokerA), which is the exponential of

the Lefschetz number of B as an endomorphism of the chain complex K•(A). By taking B = I, we

obtain exp indA.

3.3 Almost commuting operators

One might wish to calculate joint torsion for operators A and B that do not necessarily

commute with one another. The difficulty is that there is no longer a well-defined Koszul complex

K•(A,B). Carey and Pincus circumvent this problem in the case n = 2 by introducing auxiliary

operators C and D that are perturbations of the original operators A and B. Thus, consider two

Fredholm operators A and B with [A,B] ∈ L1. Furthermore, assume the existence of operators C

and D such that AB = CD,A −D ∈ L1, and B − C ∈ L1. (See Appendix A for a discussion of

the existence of such perturbations.) Thus, one has the following commutative diagram:

H C−−−−→ H

D

x xAH B−−−−→ H

By analogy with the commuting case, consider the mapping cone K•(A,B,C,D) of the

vertical chain map (A,D). Explicitly,

K•(A,B,C,D) : H

(−BD

)−−−−→ H2 (A C )−−−−→ H

This yields a triangle of modified Koszul complexes, and hence a long exact sequence in homology:

EA,D : 0 −→ (kerB ∩ kerD)−ι−→ kerB

−D−−→ kerC−ι2∗−−−→ H1(A,B,C,D)→

π1∗−−→ cokerBA−→ cokerC

π−→ H/(AH+ CH) −→ 0

Page 39: Determinants in K-theory and operator algebras

32

The homology space H1(A,B,C,D) is given by

H1(A,B,C,D) =(y, z) |Ay + Cz = 0(−Bx,Dx) |x ∈ H

.

The map ι2∗ is induced by inclusion into the second coordinate, and π1∗ is induced by projection

onto the first coordinate. We have again picked signs to simplify formulas later. Likewise, there is

a mapping cone of the horizontal chain map (B,C), and hence a long exact sequence

EB,C : 0 −→ (kerB ∩ kerD)ι−→ kerD

−B−−→ kerA−ι1∗−−−→ H1(A,B,C,D)→

π2∗−−→ cokerDC−→ cokerA

π−→ H/(AH+ CH) −→ 0

However, τ(EA,D)⊗τ(EB,C)∗ ∈ detH•(A)⊗detH•(D)∗⊗detH•(B)⊗detH•(C)∗ is no longer

canonically identified with a scalar. To obtain a scalar, Carey and Pincus introduce canonical

generators of these determinant lines, known as perturbation vectors [16, Section 3]. We will find

it convenient to give a slightly different definition. Proposition 3.3.2 below verifies that these two

definitions agree.

3.3.1 Perturbation vectors

Let A and A′ be Fredholm operators such that A−A′ ∈ L1. First assume that A (and hence

also A′) has index zero. Let L and L′ be trace class operators such that

A+ L and A′ + L′ are invertible (3.3)

L(kerA) ∩ imA = L′(kerA′) ∩ imA′ = 0 (3.4)

Then L and L′ induce isomorphisms

πL = π L|kerA : kerA∼−→ cokerA

π′L′ = π′ L′|kerA′ : kerA′∼−→ cokerA′

where π and π′ are the quotient maps by imA and imA′, respectively. Let τ(πL) and τ(π′L′) be

the torsion vectors of the above isomorphisms. Let P : H → imA be the continuous projection

Page 40: Determinants in K-theory and operator algebras

33

along L(kerA). Then ker((I − P )L) ∩ kerA = 0, and hence A defines an isomorphism

ker((I − P )L)→ imA

since A is assumed to have index zero. Let A† be the inverse of this map, followed by the inclusion

into H. Define

D(L) = det(I +A†PL).

Now in general, if the index of A and A′ is possibly nonzero, let Q be a Fredholm operator

with index negative that of A. Then A⊕Q and A′⊕Q are Fredholm operators with index zero on

H⊕H. Choose L and L′ as above, now for A⊕Q and A′ ⊕Q, respectively. Since (A⊕Q+ L)−1

is a parametrix for (A′ ⊕ Q + L′) modulo L1(H ⊕ H), we obtain an invertible determinant class

operator

Σ = (A⊕Q+ L)−1(A′ ⊕Q+ L′).

Definition 3.3.1. The perturbation vector σA,A′ ∈ detH(A)⊗ detH(A′)∗ is defined by

σA,A′ = D(L)D(L′)−1 det Σ · τ(πL)⊗ τ(π′L′)∗.

Proposition 3.3.2. The above definition agrees with that of Carey and Pincus [16, Section 3,

Equation 41]. In particular, σA,A′ is independent of the choices of Q, L, and L′.

Proof. First we note that this definition is exactly Carey and Pincus’s perturbation vector σA⊕Q,A′⊕Q

for A ⊕ Q and A′ ⊕ Q, provided that we choose L so that its range is linearly independent from

that of A⊕Q, and similarly for L′.

Now suppose that L and L′ only satisfy (3.3) and (3.4). Then (I − P )L and (I − P ′)L′ are

as in the preceding paragraph, where P is the projection onto im(A ⊕ Q) along L(ker(A ⊕ Q)).

Moreover, one calculates

(A⊕Q+ (I − P )L)−1 (A⊕Q+ L) = I + (A⊕Q)†PL

the determinant of which is D(L), and similarly for D(L′). Hence we find that this definition agrees

with Carey and Pincus’s definition of σA⊕Q,A′⊕Q. In particular, it is independent of the choices of

L and L′ by [16, Theorem 11].

Page 41: Determinants in K-theory and operator algebras

34

Next we show that specific choices of L and L′ (and hence all choices) recover Carey and

Pincus’s perturbation vector σA,A′ . First assume indA ≥ 0. Write kerA = X0 ⊕ X1 with

dim cokerA = dimX0, and write imQ⊥ = X0 ⊕ X1 with dim kerQ = dim X0. Choose sub-

spaces X ′0, X′1 for A′ similarly. Pick isomorphisms L11 : X0 −→ imA⊥, L′11 : X ′0 −→ imA′⊥,

L22 : kerQ −→ X0, L21 : ker (A + L11) → im (Q + L22)⊥, and N : ker (A′ + L′11) → ker (A + L11).

Extend these operators by zero orthogonally to all of H. Define the operators

L =

L11 0

L21 L22

L′ =

L′11 0

L21N L22

Then A⊕Q+ L and A′ ⊕Q+ L′ are invertible. Moreover,

Σ = (A⊕Q+ L)−1(A′ ⊕Q+ L′)

=

(A+ L11)R L−121

0 (Q+ L22)L

A′ + L′11 0

L21N Q+ L22

so det Σ = det

((A+ L11)R(A′ + L′11) +N

). Here, (A + L11)R denotes the right inverse such that

(A+ L11)R(A+ L11) is the orthogonal projection onto X⊥1 and similarly for the left inverse. This

agrees with the corresponding term in Carey and Pincus’s definition. The factors associated with

Q in τ(πL) and τ(π′L′) cancel by the identity v∗(v) = 1. We are then left with the definition of

Carey and Pincus.

The case when indA ≤ 0 is similar. The last statement in the proposition follows from [16,

Theorem 11].

In particular, we have the following recipe for perturbation vectors in the case of index zero

operators:

Lemma 3.3.3. Let A and A′ be index zero Fredholm operators with A − A′ ∈ L1. Let π : H →

cokerA and π′ : H → cokerA′ be the quotient maps. Take L and L′ to be trace class operators

satisfying (3.3) and (3.4). Then we have

σA,A′ = det(A+ L)−1(A′ + L′) · τ(πL)⊗ τ(π′L′)∗.

Page 42: Determinants in K-theory and operator algebras

35

For example, let ZA be a subspace complementary to imA, and let F : kerA → ZA be

any isomorphism. Then we may take L to be the extension of F by zero to all of H along any

complement of kerA. We may define L′ similarly.

3.3.2 Joint torsion of two almost commuting operators

Definition 3.3.4. Suppose A,B,C, and D are Fredholm operators with AB = CD,A −D ∈ L1,

and B − C ∈ L1. The joint torsion τ(A,B,C,D) is the nonzero scalar

τ(A,B,C,D) = (−1)λ(A,B,C,D)τ(EA,D)⊗ σA,D ⊗ σB,C ⊗ τ(EB,C)∗.

Here λ(A,B,C,D) = sgn θ+ γ, where θ is the permutation that maps the factors in (4.6) to

(4.5) and γ is given by (4.7). See Remark 4.1.3. Once again, we note that our sign conventions

differ from those of [16] and that τ(A,B,C,D) has been identified with a scalar according to (3.1)

and (3.2).

If A and B commute, then τ(A,B,B,A) = τ(A,B) since σA,A and σB,B are trivial. By

Proposition 3.3.2 and the exact sequences EA,D and EB,C , we find that the above definition agrees

with that of Carey and Pincus [16, Section 5, Equation 51], up to the evident difference in our sign

conventions. By the multiplicativity of the determinant, we have the following:

Lemma 3.3.5. For any Fredholm operators Ai, Bi, Ci, Di with AiBi = CiDi, Ai − Di ∈ L1, and

Bi − Ci ∈ L1, i = 1, 2, we have

τ(A1 ⊕A2, B1 ⊕B2, C1 ⊕ C2, D1 ⊕D2) = τ(A1, B1, C1, D1) · τ(A2, B2, C2, D2).

Remark 3.3.6. For n ≥ 2 almost commuting operators, one can define joint torsion by introducing

auxiliary operators as in Definition 3.3.4 and using the constructions in [39]. It is not known in

what generality this can be done. In fact, not every pair of almost commuting Fredholm operators

A and B have trace class perturbations D and C, respectively, such that AB = CD. See Appendix

A for a more detailed discussion.

Page 43: Determinants in K-theory and operator algebras

Chapter 4

Equality

In this chapter we show that joint torsion is equal to the determinant invariant (Theorem

4.3.3). We begin by showing that joint torsion is trivial in finite dimensions. Next we establish

factorization results for joint torsion under actions by invertible operators in Section 4.2. Then

in Section 4.3, we will see that the proof of equality follows quickly from the preceding sections.

Finally, we obtain a number of consequences of Theorem 4.3.3, many of which will be used in the

following chapter. Most of the results of this chapter have appeared in the author’s work [45].

4.1 The finite dimensional case

In this section we show that joint torsion is trivial in a finite dimensional space, closely

following Kaad [37]. In fact, for commuting operators, this is a special case of [37, Theorem 4.3.3].

Later we will see that joint torsion is also trivial for operators in the coset I + L1. First consider

the exact sequence

0 −→ kerA −→ H A−→ H −→ cokerA −→ 0.

If H is finite dimensional, the torsion vector τ(A) of the above sequence is defined.

Lemma 4.1.1. If A and D are linear operators on a finite dimensional space H, then

σA,D = τ(A)⊗ τ(D)∗.

Proof. First we calculate the perturbation vector σA,D as

σA,D = det(A+ LA)−1(D + LD) · τ(πALA)⊗ τ(πDLD)∗.

Page 44: Determinants in K-theory and operator algebras

37

where LA, LD, πA, and πD are as in Lemma 3.3.3.

On the other hand, choose any nonzero t0 and t1 with t0 ∈ det kerA and, for example,

t1 ∈ det (kerA)⊥. Then we calculate

τ(A) = (t0)∗ ⊗ t0 ∧ t1 ⊗ (LAt0 ∧At1)∗ ⊗ (πALAt0).

We move the fourth factor past the middle two factors according to (3.2) (applied twice) and obtain

τ(πALA). For the middle factors, we calculate

t0 ∧ t1 ⊗ (LAt0 ∧At1)∗ = (−1)dimH ((A+ LA)(t0 ∧ t1))∗ ⊗ (t0 ∧ t1)

= (−1)dimH det(A+ LA)−1

The same calculation for τ(D) completes the proof.

4.1.1 Torsion of a double complex

Now consider a double complex X•• consisting of finite dimensional vector spaces Xij , 0 ≤

i ≤ m, 0 ≤ j ≤ n, with exact rows (Xi•, h) and exact columns (X•j , v), where h : Xi• → Xi,•+1

and v : X•j → X•+1,j . Assume that the horizontal and vertical maps h and v anti-commute, i.e.

vh+ hv = 0.

For each i, one may form the torsion vector τ(Xi•) of row i and combine all these to obtain the

horizontal torsion vector

τh = τ(X0•)⊗ τ(X1•)∗ ⊗ . . .

One may also form the vertical torsion vector

τv = τ(X•0)⊗ τ(X•1)∗ ⊗ . . .

Both of these vectors are generators of the same determinant line of X••, and moreover by the work

of F. F. Knudsen and D. Mumford [41], they agree up to a sign.

Page 45: Determinants in K-theory and operator algebras

38

One rather direct proof of this fact begins by noticing that each component Xi,j of the double

complex is isomorphic to

im(vh)⊕ ker v ∩ kerh

im(vh)⊕ ker v

ker v ∩ kerh⊕ kerh

ker v ∩ kerh⊕ ker(vh)

ker v + kerh⊕ Xi,j

ker(vh)(4.1)

Here, and below, the indices on v and h may be inferred, and hence are suppressed. We note of

course that some of these spaces may be trivial. The vertical map v induces isomorphisms

kerh

ker v ∩ kerh∼= im(vh)

ker(vh)

ker v + kerh∼=

ker v ∩ kerh

im(vh)

Xi,j

ker(vh)∼=

ker v

ker v ∩ kerh

Similarly, h induces isomorphisms

ker v

ker v ∩ kerh∼= im(vh)

ker(vh)

ker v + kerh∼=

ker v ∩ kerh

im(vh)

Xi,j

ker(vh)∼=

kerh

ker v ∩ kerh

For each Xi,j , one may pick representative subspaces for each of the six quotients in (4.1) above.

Thus:

X1i,j = im(vh) (4.2)

X2i,j is a complement of im(vh) in ker v ∩ kerh

X3i,j is a complement of ker v ∩ kerh in ker v

X4i,j is a complement of ker v ∩ kerh in kerh

X5i,j is a complement of ker v + kerh in ker(vh)

X6i,j is a complement of ker(vh) in Xi,j

These give an algebraic decomposition of Xi,j in which X1i,j , X

2i,j , X

3i,j span ker v and X4

i,j , X5i,j , X

6i,j

span a complement. Furthermore, X1i,j , X

2i,j , X

4i,j span kerh and X3

i,j , X5i,j , X

6i,j span a complement.

Page 46: Determinants in K-theory and operator algebras

39

Next we note that given a decomposition of the spaces Xi,j on the n-th diagonal, i.e. i+j = n,

we may find a compatible decomposition for the (n+1)-st diagonal. Indeed, let i+ j = n+1. Then

we may take

X1i,j = v(X4

i−1,j) = h(X3i,j−1) (4.3)

X2i,j = v(X5

i−1,j) = h(X5i,j−1)

X3i,j = v(X6

i−1,j)

X4i,j = h(X6

i,j−1)

In (4.3), X5i−1,j and X5

i,j−1 have already been chosen in this way so that v(X5i−1,j) = h(X5

i,j−1).

Moreover, having chosen X5i−1,j+1, we may take X5

i,j to be a lift of of v(X5i−1,j+1) along h. Finally,

we may take X6i,j to be any complement of ker(vh) in Xi,j .

Thus, one may inductively pick generators for all of the spaces Xki,j in order to calculate both

the vertical and horizontal torsion according to Definition 3.2.1. The same generators are used,

only in a different order. Hence these torsion vectors are the same, up to the sign given by repeated

application of (3.2). The proof of the proposition below follows the same strategy.

4.1.2 Joint torsion in finite dimensions

Now let A, B, C, and D be operators on a finite dimensional vector space H such that

AB = CD. Consider the Koszul complexes K•(A), K•(B), K•(C), K•(D), and K•(A,B,C,D).

At the level of homology, we obtain the following bicomplex of finite dimensional vector spaces with

exact rows and columns. Write Hi = Hi(A,B,C,D) for the homology spaces.

Page 47: Determinants in K-theory and operator algebras

40

0 0 0...y y y y−ι1∗

0 −−−−→ H2−ι−−−−→ kerB

−D−−−−→ kerC−ι2∗−−−−→ H1

π1∗−−−−→ · · ·yι yι yι yπ2∗

0 −−−−→ kerDι−−−−→ H D−−−−→ H π−−−−→ cokerD −−−−→ 0y−B yB yC yC

0 −−−−→ kerAι−−−−→ H −A−−−−→ H −π−−−−→ cokerA −−−−→ 0y−ι1∗ yπ yπ yπ

· · · −ι2∗−−−−→ H1π1∗−−−−→ cokerB

A−−−−→ cokerCπ−−−−→ H0 −−−−→ 0yπ2∗

y y y... 0 0 0

(4.4)

Here the upper right and lower left corners are identified. The above diagram therefore

consists of three exact rows and three exact columns. Two of the rows are the sequences from

Lemma 4.1.1 corresponding to D and A (up to a sign) and hence will be combined below to

form the perturbation vector σA,D. The other, seven term, row is the exact sequence EA,D in the

definition of joint torsion. Two of the columns correspond to B and C, and the other is the exact

sequence EB,C . Thus we obtain horizontal and vertical torsion vectors

τ(EA,D)∗ ⊗ τ(D)⊗ τ(A)∗

τ(EB,C)∗ ⊗ τ(B)⊗ τ(C)∗

The theorem below asserts that these agree, up to the signs in Definition 3.3.4 and Lemma 4.1.1.

We note that the above bicomplex arises from an odd homotopy exact bitriangle of Z2-graded

chain complexes that appears in [37, Equation 5.1]:

Page 48: Determinants in K-theory and operator algebras

41

H −−−−→ H[1] −−−−→ K•(B) −−−−→ Hy y y yH[1] −−−−→ H −−−−→ K•(C)[1] −−−−→ H[1]y y y yK•(D) −−−−→ K•(A)[1] −−−−→ K•(A,B,C,D) −−−−→ K•(D)y y y yH −−−−→ H[1] −−−−→ K•(B) −−−−→ H

Here the vertical maps are induced by A and D, and the horizontal maps are induced by B and C.

The spaceH is given the grading with trivial odd part, and the notation X[1] denotes the Z2-graded

chain complex X with the grading reversed and the differential negated. Thus, the horizontal and

vertical arrows are odd chain maps which anticommute with the differential, and the squares are

anticommutative.

Theorem 4.1.2. If H is finite dimensional and A,B,C, and D are operators on H such that

AB = CD, then τ(A,B,C,D) = 1.

Proof. For convenience, let us temporarily rename the spaces in (4.4):

0 0 0...y y y y

0 −−−−→ X1,1 −−−−→ X1,2 −−−−→ X1,3 −−−−→ Y −−−−→ · · ·y y y y0 −−−−→ X2,1 −−−−→ X2,2 −−−−→ X2,3 −−−−→ X2,4 −−−−→ 0y y y y0 −−−−→ X3,1 −−−−→ X3,2 −−−−→ X3,3 −−−−→ X3,4 −−−−→ 0y y y y· · · −−−−→ Y −−−−→ X4,2 −−−−→ X4,3 −−−−→ X4,4 −−−−→ 0y y y y

... 0 0 0

Page 49: Determinants in K-theory and operator algebras

42

As described in (4.2) and (4.3), we may pick generators

tkij ∈ detXki,j

for all of the spaces except the diagonal containing Y , that is, all spaces Xi,j with i+ j 6= 5. One

checks that we may also pick generators for subspaces

X1Y , X

3Y , X

4Y , X

6Y

X12,3, X

32,3, X

42,3, X

62,3

X13,2, X

33,2, X

43,2, X

63,2

In particular, we may pick all these generators to be compatible with one another in the sense of

(4.3). Moreover, there exist

x = t5Y ∈ detY 5, w = t523 ∈ detX52,3, z = t532 ∈ detX5

3,2

such that

v(z) = h(x), h(w) = v(x), h(z) = v(w)

for example by [37, Equations (4.5) and (4.6)]. Similarly we find that there exist compatible

generators

t2Y ∈ detY 2, t223 ∈ detX22,3, t232 ∈ detX2

3,2,

Having picked generators for the top exterior powers of all spaces in the bicomplex, we may

now use the columns to calculate τ(EB,C)⊗ σ∗B,C as:

(t611

)∗ ⊗ (vt611 ∧ t621

)⊗(t531 ∧ t631 ∧ vt612

)∗ ⊗ (vt531 ∧ vt631 ∧ ht613 ∧ t5Y)

(4.5)

⊗(ht632 ∧ vht613 ∧ vt5Y

)∗ ⊗ (vht632 ∧ ht633

)⊗(vht633

)∗⊗[ (ht611 ∧ t612

)∗ ⊗ (vht611 ∧ vt612 ∧ ht621 ∧ t522 ∧ t622

)⊗(ht631 ∧ t532 ∧ t632 ∧ vht621 ∧ vt522 ∧ vt622

)∗ ⊗ (vht631 ∧ vt532 ∧ vt632

) ]∗⊗(ht612 ∧ t513 ∧ t613

)∗ ⊗ (vht612 ∧ vt513 ∧ vt613 ∧ ht622 ∧ t523 ∧ t623

)⊗(ht632 ∧ t633 ∧ vht622 ∧ vt523 ∧ vt623

)∗ ⊗ (vht632 ∧ vt633

)

Page 50: Determinants in K-theory and operator algebras

43

Likewise, we use the rows to calculate τ(EA,D)⊗ σA,D as:

(t611

)∗ ⊗ (ht611 ∧ t612

)⊗(t513 ∧ t613 ∧ ht612

)∗ ⊗ (vt531 ∧ ht613 ∧ vt631 ∧ t5Y)

(4.6)

⊗(vt632 ∧ hvt631 ∧ vt532

)∗ ⊗ (hvt632 ∧ vt633

)⊗(hvt633

)∗⊗[ (vt611 ∧ t621

)∗ ⊗ (hvt611 ∧ ht621 ∧ vt612 ∧ t522 ∧ t622

)⊗(vt613 ∧ t523 ∧ t623 ∧ hvt612 ∧ vt513 ∧ ht622

)∗ ⊗ (hvt613 ∧ vt5Y ∧ ht623

) ]∗⊗(vt621 ∧ t531 ∧ t631

)∗ ⊗ (vt632 ∧ t633 ∧ hvt622 ∧ vt523 ∧ ht632

)∗⊗(hvt621 ∧ vt522 ∧ ht631 ∧ vt622 ∧ t532 ∧ t632

)⊗(hvt632 ∧ ht633

)Note that we have accounted for the sign in the third row of (4.4) by switching the order of

the factors corresponding to X3,2 and X3,3. The vector (4.6) consists of the same factors as (4.5),

only in a different order. Let θ denote this permutation. In more detail, the spaces Xij appear in

a different order, and moreover the subspaces X3ij and X4

ij are switched. In addition hv = −vh, so

we find that (4.6) and (4.5) further differ by a factor of (−1)γ , where

γ = dimX12,2 + dimX1

2,3 + dimX12,4 + dimX1

3,2

+ dimX13,3 + dimX1

3,4 + dimX14,2 + dimX1

4,3 + dimX14,4 (4.7)

Setting λ(A,B,C,D) = sgn θ + γ, we have

τ(EA,D)⊗ σA,D = (−1)λ(A,B,C,D)τ(EB,C)⊗ σ∗B,C

so that τ(A,B,C,D) = 1 as desired.

Remark 4.1.3. The subspaces X62,2, X4

2,3, X33,2, and X1

3,3 all have the same dimension. One finds

that this dimension appears once in both γ and the permutation θ. Hence this dimension does not

appear in the factor (−1)λ(A,B,C,D). Since this is the only dimension in the double complex (4.4)

that may be infinite in general, we find that (−1)λ(A,B,C,D) is indeed well-defined, regardless of

whether or not H is finite dimensional.

Page 51: Determinants in K-theory and operator algebras

44

Corollary 4.1.4. Suppose A,B,C, and D are operators on a Hilbert space H, each of which differs

from the identity by a finite rank operator. If AB = CD, then τ(A,B,C,D) = 1.

Proof. With respect to a decomposition H = H0 ⊕ V for some finite dimensional subspace V , the

operators A,B,C and D are of the form I ⊕ FA, I ⊕ FB, I ⊕ FC , and I ⊕ FD, respectively. In

particular, FAFB = FCFD. By Lemma 3.3.5, we have

τ(A,B,C,D) = τ(I, I, I, I) · τ(FA, FB, FC , FD)

The first factor is one, and the second factor is also one by the preceding proposition.

4.2 Factorization results

For an invertible operator U and finite dimensional subspace V of H, denote by U |V the

isomorphism U |V : V → U(V ). Let τ(U |V ) denote the torsion of this isomorphism. Also, for a

Fredholm operator T , let U |cokerT denote the isomorphism U |cokerT : cokerT → cokerUT given by

v + TH 7→ Uv + UTH. Let τ(U |cokerT ) denote the torsion of this isomorphism.

4.2.1 Perturbation vectors

Lemma 4.2.1. Let a and u be commuting units in B/L1, and let A,D ∈ B be lifts of a. If u has

an invertible lift U ∈ B, then

σA,U−1DU = d(a, u) · σA,D ⊗ τ(U−1|kerD

)⊗ τ

(U−1|cokerD

)∗.

Proof. First we note that A − U−1DU ∈ L1, so the perturbation vector σA,U−1DU is defined. Let

us begin by proving the lemma in the case when A, and hence also D, has index zero.

For T = A,D, choose a subspace ZT complementary to imT , and choose isomorphisms FT :

kerT → ZT . Define LT as in Lemma 3.3.3 and the subsequent paragraph. Let πT : H → cokerT

be the quotient map. The operator U−1FDU defines an isomorphism

U−1FDU : kerU−1DU = U−1(kerD)→ U−1(ZD) (4.8)

Page 52: Determinants in K-theory and operator algebras

45

and U−1(ZD) is a subspace complementary to im(U−1DU) = U−1(imD). The torsion of the

isomorphism induced by (4.8) is identified as

τ(πU−1DUU−1FDU) = τ

(U−1|kerD

)∗ ⊗ τ(πDFD)⊗ τ(U−1|cokerD

).

Additionally, one has

det(A+ LA)−1(U−1DU + U−1LDU) = det(A+ LA)−1(D + LD)

· det(D + LD)−1U−1(D + LD)U

= det(A+ LA)−1(D + LD) · d(a, u)

Combining the two preceding equations, we calculate

σA,U−1DU = det(A+ LA)−1(U−1DU + U−1LDU) · τ(πAFA)⊗ τ(πU−1FDU)∗

= d(a, u) · det(A+ LA)−1(D + LD) · τ(πAFA)⊗ τ(πDFD)∗

⊗ τ(U−1|kerD

)⊗ τ

(U−1|cokerD

)∗= d(a, u) · σA,D ⊗ τ

(U−1|kerD

)⊗ τ

(U−1|cokerD

)∗In the second equality, we have used (3.2) several times.

Now if indA is not necessarily zero, let Q be any Fredholm operator with indQ = −indA.

Let q be the image of in B/L1 of Q. Let A = A⊕Q, D = D ⊕Q, and U = U ⊕ I. Then A and D

have index zero, A− D ∈ L1(H2), and [D, U ] ∈ L1(H2). Hence we calculate

σA,U−1DU = d(a, u) · σA,D ⊗ τ(U−1|ker D

)⊗ τ

(U−1|coker D

)∗= d(a, u) · d(q, 1) · σA,D ⊗ σS,S ⊗ τ

(U−1|kerD

)⊗ τ

(U−1|cokerD

)∗= d(a, u) · σA,D ⊗ τ

(U−1|kerD

)⊗ τ

(U−1|cokerD

)∗On the other hand,

σA,U−1DU = σA,U−1DU

so we have

σA,U−1DU = d(a, u) · σA,D ⊗ τ(U−1|kerD

)⊗ τ

(U−1|cokerD

)∗.

Page 53: Determinants in K-theory and operator algebras

46

Lemma 4.2.2. Let a and u be units in B/L1 and let A,D ∈ B be lifts of a. If u has an invertible

lift U ∈ B, then

(1) σAU,DU = τ(U−1|kerA

)∗ ⊗ σA,D ⊗ τ (U−1|kerD

)(2) σUA,UD = τ (U |cokerA)⊗ σA,D ⊗ τ (U |cokerD)∗

Proof. First we note that AU −DU,UA−UD ∈ L1. Let us begin with the case when A has index

zero. With FA, FD, LA, LD as before, we have

det(AU + LAU)−1(DU + LDU) = det(A+ LA)−1(D + LD).

Furthermore, kerAU = U−1(kerA) and kerDU = U−1(kerD), and we calculate

τ(πAFAU) = τ(πAFA)⊗ τ(U−1|kerA)∗

τ(πDFDU) = τ(πDFD)⊗ τ(U−1|kerD)∗

Therefore σAU,DU = τ(U−1|kerA

)∗ ⊗ σA,D ⊗ τ (U−1|kerD

).

In general, if indA is not necessarily zero, let Q, A, D, U be as in Lemma 4.2.1. Then

σAU ,DU = τ(U−1|kerA

)∗ ⊗ σA,D ⊗ τ (U−1|kerD

).

On the other hand,

σAU ,DU = σAU,DU .

Therefore σAU,DU = τ(U−1|kerA

)∗⊗σA,D⊗ τ (U−1|kerD

). The second part is proved similarly.

Lemma 4.2.3. Let a be a unit in B/L1, and let A,D ∈ B be lifts of a. If U ∈ B is an invertible

lift of 1 ∈ B/L1, then

(1) σA,DU = σA,D ⊗ τ(U−1|kerD

)∗ · detU

(2) σA,UD = σA,D ⊗ τ (U |cokerD) · detU

Proof. First we note that U is an invertible determinant class operator by assumption. Moreover,

A−DU,A− UD ∈ L1, so the perturbation vectors are defined. The proof then proceeds as in the

previous lemma.

Page 54: Determinants in K-theory and operator algebras

47

4.2.2 Joint torsion

In this section, we use the previous three lemmas to calculate the joint torsion of quadruples

(A,B,C,D) in terms of quadruples modified by an invertible operator. This will allow us to reduce

the calculation of joint torsion to a determinant invariant and a finite dimensional calculation,

which has already been dealt with in Section 4.1. In the following section, we will interpret the

proposition below as calculating the transformation of the joint torsion of quadruples (A,B,C,D)

under an action by a group of invertibles.

Proposition 4.2.4. Let a and b be commuting units in B/L1. Let A,D ∈ B be lifts of a, and let

B,C ∈ B be lifts of b such that AB = CD. Suppose u ∈ B/L1 has an invertible lift U ∈ B.

(1) If a and u commute, then

τ(A,BU,CU,U−1DU) = d(a, u) · τ(A,B,C,D).

(2) If b and u commute, then

τ(UA,B,UCU−1, UD) = d(u, b) · τ(A,B,C,D).

Proof. First we note that A(BU) = (CU)(U−1DU) and A − U−1DU ∈ L1, and BU − CU ∈ L1.

Hence the joint torsion in (1) is defined, and likewise for (2). Let us first prove (1). By Lemmas

4.2.1 and 4.2.2, we have the factorization

σA,U−1DU ⊗ σBU,CU = d(a, u) · σA,D ⊗ σB,C ⊗ τ(U−1|kerB

)∗ ⊗ τ (U−1|kerC

)⊗

⊗ τ(U−1|kerD

)⊗ τ

(U−1|cokerD

)∗(4.9)

One uses (3.2) to check that the torsion vectors can indeed be moved past each other with impunity.

To calculate τ(EA,D), we choose generators t0, . . . , t5 appropriately:

• t0 ∈ det(kerB ∩ kerD)

• t0 ∧ t1 ∈ det kerB

Page 55: Determinants in K-theory and operator algebras

48

• Dt1 ∧ t2 ∈ det kerC

• ι2∗t2 ∧ t3 ∈ detH1(A,B,C,D)

• π1∗t3 ∧ t4 ∈ det cokerB

• At4 ∧ t5 ∈ det cokerC

• πt5 ∈ det (H/(AH+ CH))

Then τ(EA,D) = t∗0 ⊗ (t0 ∧ t1) ⊗ · · · ⊗ (At4 ∧ t5) ⊗ (πt5)∗. Here, H1 = H1(A,B,C,D) is the first

Koszul homology space, which we recall is expressed as:

H1 =(y, z) |Ay = −Cz(−Bx,Dx) |x ∈ H

.

The map ι2∗ is induced by inclusion into the second coordinate, and π1∗ is induced by projection

onto the first coordinate. Let v ∈ kerC be such that v /∈ D(kerC). Then ι2∗v = [(0, v)] 6= 0 in H1,

so we can take t2 to be the product of sufficiently many such vectors, say ∧ivi. On the other hand,

let w /∈ imB be such that Aw = Cu ∈ imC for some u. Then [(w,−u)] 6= 0 in H1 and π1∗[(w,−u)],

so we can take t3 to be the product of sufficiently many such vectors, say ∧j [(wj ,−uj)]. Of course,

τ(EA,D) is independent of these specific choices.

Now we would like to calculate the torsion vectors τ(EA,U−1DU ) and τ(EBU,CU ) in terms

of τ(EA,D) and τ(EB,C). The only potential difficulty is in the H1 position, so let us compare

H ′1 = H1(A,BU,CU,U−1DU) with the discussion of the previous paragraph. First,

H ′1 =(y, U−1z) |Ay = −Cz(−Bx,U−1Dx) |x ∈ H

.

The invertible operator I ⊕ U−1 on H ⊕ H induces an isomorphism from H1 onto H ′1, which we

denote by (I ⊕ U−1)|H1 . Moreover, we have

Page 56: Determinants in K-theory and operator algebras

49

U−1(Dt1 ∧ t2) = U−1Dt1 ∧ U−1t2

∈ det kerCU

(I ⊕ U−1)|H1(ι2∗t2 ∧ t3) =(∧i[(0, U−1vi)]

)∧(∧j [(wj ,−U−1uj)]

)= ι2∗U

−1t2 ∧ (I ⊕ U−1)|H1t3

∈ detH ′1

π1∗(I ⊕ U−1)|H1t3 ∧ t4 = π1∗t3 ∧ t4

∈ det cokerBU

Hence, we calculate

τ(EA,U−1DU ) = (U−1t0)∗ ⊗ ((−U−1t0) ∧ U−1t1)⊗ (U−1t2 ∧ −U−1Dt1)∗⊗

⊗ ((I ⊕ U−1)|H1(−ι2∗)t2 ∧ t3)⊗ (t4 ∧ π1∗t3)∗ ⊗ (At4 ∧ t5)⊗ (πt5)∗

= τ(EA,D)⊗ τ((I ⊕ U−1)|H1)⊗ τ(U−1|kerB

)⊗ τ

(U−1|kerC

)∗⊗ (4.10)

⊗ τ(U−1|kerB∩kerD

)∗Similarly, we find

τ(EBU,CU ) = τ(EB,C)⊗ τ((I ⊕ U−1)|H1)⊗ τ(U−1|kerD

)⊗ τ

(U−1|cokerD

)∗⊗ (4.11)

⊗ τ(U−1|kerB∩kerD

)∗Notice that λ = λ(A,B,C,D) is the same as λ(A,BU,CU,U−1DU) in Definition 3.3.4 since

all the spaces involved have the same dimension. Combining (4.9), (4.10), and (4.11), we find

τ(A,BU,CU,U−1DU) = (−1)λτ(EA,U−1DU )⊗ τ(EBU,CU )∗ ⊗ σA,U−1DU ⊗ σBU,CU

= (−1)λd(a, u) · τ(EA,D)⊗ τ(EB,C)∗ ⊗ σA,D ⊗ σB,C

= d(a, u) · τ(A,B,C,D)

This completes the proof of part 1. The second part follows by a similar analysis.

Page 57: Determinants in K-theory and operator algebras

50

Proposition 4.2.5. Let A,B,C, and D be Fredholm operators with AB = CD, A−D ∈ L1, and

B − C ∈ L1. For any invertible determinant class operator U , we have

τ(A,B,CU,U−1D) = τ(A,B,C,D).

Proof. First we note that I − U−1 = (U − I)U−1 ∈ L1, so A− U−1D ∈ L1 and B −CU ∈ L1. We

calculate perturbation vectors as in (4.9), this time using using Lemma 4.2.3:

σA,U−1D ⊗ σB,CU = σA,D ⊗ σB,C ⊗ τ(U−1|cokerD)∗ ⊗ τ(U |ker C)∗. (4.12)

Next, pick generators tk as in Proposition 4.2.4. As before, we find that

H1(A,B,CU,U−1D) =(y, U−1z) |Ay = −Cz(−Bx,U−1Dx) |x ∈ H

.

The invertible operator I ⊕ U−1 on H ⊕ H induces an isomorphism from H1(A,B,C,D) onto

H1(A,B,CU,U−1D), which we denote by (I ⊕ U−1)|H1 . Moreover, we have

U−1(Dt1 ∧ t2) = U−1Dt1 ∧ U−1t2

∈ det kerCU

(I ⊕ U−1)|H1(ι2∗t2 ∧ t3) = ι2∗U−1t2 ∧ (I ⊕ U−1)|H1t3

∈ detH1(A,B,CU,U−1D)

π1∗(I ⊕ U−1)|H1t3 ∧ t4 = π1∗t3 ∧ t4

∈ det cokerB

Hence we calculate

τ(EA,U−1D) = τ(EA,D)⊗ τ((I ⊕ U−1)|H1)⊗ τ(U−1|kerC

)∗ ⊗ τ (U−1|kerB∩kerD

)∗(4.13)

τ(EB,CU ) = τ(EB,C)⊗ τ((I ⊕ U−1)|H1)⊗ τ(U−1|cokerD

)∗ ⊗ τ (U−1|kerB∩kerD

)∗(4.14)

Combining equations (4.12), (4.13), and (4.14), we find that τ(A,B,CU,U−1D) = τ(A,B,C,D),

as desired.

Page 58: Determinants in K-theory and operator algebras

51

4.3 Main results

4.3.1 Group actions on commuting squares

As usual, let π : B → B/L1 denote the projection. Denote by S the set of all quadruples of

Fredholm operators (A,B,C,D) on a fixed Hilbert space such that AB = CD, A −D ∈ L1, and

B − C ∈ L1.

Definition 4.3.1. For a, b ∈ B/L1, define the following sets:

S1,a = (A,B,C,D) ∈ S |π(A) = a

S2,b = (A,B,C,D) ∈ S |π(B) = b

and the following group:

Ga = (U, V ) ∈ B(H)× × B(H)× |π(U) = π(V ), [π(U), a] = 0

Then Ga acts on S1,a on the right:

(A,B,C,D) •1,a (U, V ) = (A,BU,CV, V −1DU)

and Gb acts on S2,b on the right:

(A,B,C,D) •2,b (U, V ) = (U−1A,B,U−1CV, V −1D)

Now we can rephrase our factorization results from the previous section in terms of these

group actions:

Theorem 4.3.2.

(1) If (U, V ) ∈ Ga and (A,B,C,D) ∈ S1,a, then

τ ((A,B,C,D) •1,a (U, V )) = d(a, π(U)) · τ(A,B,C,D).

(2) If (U, V ) ∈ Gb and (A,B,C,D) ∈ S2,b, then

τ ((A,B,C,D) •2,b (U, V )) = d(π(U)−1, b) · τ(A,B,C,D).

Page 59: Determinants in K-theory and operator algebras

52

4.3.2 A proof of equality

We are now in a position to prove the main result of this chapter, namely that joint torsion

is equal to the determinant invariant:

Theorem 4.3.3. Let a and b be commuting units in B/L1. Let A,D ∈ B be lifts of a, and let

B,C ∈ B be lifts of b such that AB = CD. Then d(a, b) = τ(A,B,C,D).

Proof. First suppose A and B have index zero. Let a = π(A), b = π(B). Let U be an invertible

parametrix for B modulo finite rank operators. For example, we may take U = (B + F )−1 for any

appropriate finite rank operator F . Similarly, let V be an invertible parametrix for C modulo finite

rank operators. Let U ′ and V ′ be invertible finite rank perturbations of A and D, respectively.

Then (U, V ) ∈ Ga and (U ′, V ′) ∈ Gb, and we find

(A,B,C,D) •1,a (U, V ) •2,b (U ′, V ′)

= (U ′−1A,BU,U ′−1CV V ′, V ′−1V −1DU)

Notice that all four of the above operators are finite rank perturbations of the identity. Hence

τ((A,B,C,D) •1,a (U, V ) •2,b (U ′, V ′)

)= 1

by Theorem 4.1.2. On the other hand, the above theorem gives

τ((A,B,C,D) •1,a (U, V ) •2,b (U ′, V ′)

)= d(π(U ′)−1, π(BU)) · d(a, π(U)) · τ(A,B,C,D)

Notice that d(π(U ′)−1, π(BU)) = 1 and d(a, π(U)) = d(a, b)−1 since π(U) = b−1. By comparing

the two preceding equations, we find

τ(A,B,C,D) = d(a, b).

In general, if indA = n and indB = m, let Q be a Fredholm operator with index −n and let

R be a Fredholm operator with index −m. Then the operators A = A ⊕ Q ⊕ I, B = B ⊕ I ⊕ R,

Page 60: Determinants in K-theory and operator algebras

53

C = C ⊕ I ⊕R, D = D ⊕Q⊕ I are index zero operators on H⊕H⊕H such that AB = CD. Let

a and b be the images of A and B, respectively, modulo trace class. By the above, we find that

τ(A, B, C, D) = d(a, b).

Moreover,

τ(A,B,C,D) = τ(A, B, C, D)

and

d(a, b) = d(a, b).

The result follows by combining the above three equations.

4.4 Applications

In the first two subsections, we apply Theorem 4.3.3 to derive some properties of the deter-

minant invariant and joint torsion. This provides new proofs of known results, namely Theorem

4.4.1 and Lemma 4.4.6 in the commutative case, as well as other results that appear to be new.

4.4.1 Properties of the determinant invariant

By Theorem 4.3.3, the determinant invariant enjoys the same properties as joint torsion. Al-

though the determinant invariant is defined in terms of infinite dimensional Fredholm determinants,

Theorem 4.3.3 guarantees that for commuting operators, it can actually be calculated in terms of

finite dimensional data.

Next, we note that in the case when the Koszul complex K•(A,B) is acyclic, we have the

following consequence of Theorem 4.3.3 and Lemma 3.2.4, which was first obtained in [13].

Theorem 4.4.1. If A and B are commuting Fredholm operators and the Koszul complex K•(A,B)

is acyclic, we have

d(a, b) =detB|kerA

detB|cokerA

detA|cokerB

detA|kerB.

Page 61: Determinants in K-theory and operator algebras

54

There are analogues of Theorem 4.4.1 more generally in the case of almost commuting oper-

ators. Let a and b be commuting units in B/L1. Let A,D ∈ B be lifts of a, and let B,C ∈ B be

lifts of b such that AB = CD. For simplicity, assume that the Koszul complex K•(A,B,C,D) is

acyclic. Pick Fredholm operators Q and R such that indQ = −indA and indR = −indB. Then

A = A⊕Q⊕ I, B = B⊕ I ⊕R, C = C ⊕ I ⊕R, and D = D⊕Q⊕ I all have index zero. Moreover,

the new Koszul complex K•(A, B, C, D) is acyclic, and d(A,B) = d(A, B) = τ(A, B). Then we

find

τ(EA,D) = τ(B : ker D → ker A)⊗ τ(C : coker D → coker A)

τ(EB,C) = τ(D : ker B → ker C)⊗ τ(A : coker B → coker C)

For T = A, B, C, D, pick trace class operators LT and let πT be the quotient map as in

Lemma 3.3.3. Thus we have isomorphisms πTLT : kerT → cokerT , and we calculate

σA,D = det(A+ LA)−1(D + LD)τ(πALA)⊗ τ(πDLD)∗

σB,C = det(B + LB)−1(C + LC)τ(πBLC)⊗ τ(πBLC)∗

Corollary 4.4.2. In the situation above, d(a, b) is given by

(−1)λ(A,B,C,D) det(A+ LA)−1(D + LD)(B + LB)−1(C + LC)

·det(πDLD)−1(C|coker D)−1(πALA)(B|ker D)

det(πBLB)−1(A|coker B)−1(πCLC)(D|ker B)

4.4.2 Properties of joint torsion

Now let us record a number of properties of joint torsion that follow from Theorem 4.3.3.

First, we note that for any two Fredholm operators A and B with commuting images a and b,

respectively, in B/L1, we are justified in writing τ(a, b), regardless of the existence of perturbations

C,D required in the definition of joint torsion.

Next we study the continuity of joint torsion. Consider the space of almost commuting pairs

M = (A,B) |A and B are Fredholm and [A,B] ∈ L1

Page 62: Determinants in K-theory and operator algebras

55

endowed with the metric

d((A1, B1), (A2, B2)) = ‖A1 −A2‖+ ‖B1 −B2‖+ ‖[A1, B1]− [A2, B2]‖1.

One may check that (M,d) is complete. In light of Theorem 4.3.3, joint torsion may be defined on

M by

τ(A,B) = τ(a, b)

where a and b as usual denote the images of A and B in B/L1. The following corollary also follows

from the recent work [39].

Corollary 4.4.3. Joint torsion is a continuous map from (M,d) into C.

Proof. The strategy is to prove this result first on the subset of pairs of invertible operators, then

for index zero operators, and finally in general. Thus, suppose A0 and B0 are invertible. Then

τ(A0, B0) = detA0B0A−10 B−1

0 = det(I + [A0, B0]A−1

0 B−10

).

Suppose (A,B)→ (A0, B0) in M . Since A→ A0 and B → B0 in norm, we may assume that all A

and B are invertible. Hence

[A,B]A−1B−1 → [A0, B0]A−10 B−1

0

in L1. Consequently, τ(A,B)→ τ(A0, B0) since the map

L1 → C, X 7→ det(I +X)

in continuous.

Now suppose that A0 and B0 merely have index zero. Let FA and FB be finite rank operators

such that A0 + FA and B0 + FB are invertible. Suppose (A,B) → (A0, B0) in M . Then we may

assume that all A+FA and B+FB are invertible. Moreover (A+FA, B+FB)→ (A0 +FA, B0 +FB)

in M , so by the above argument,

τ(A+ FA, B + FB)→ τ(A0 + FA, B0 + FB).

Page 63: Determinants in K-theory and operator algebras

56

Hence τ(A,B)→ τ(A0, B0).

Finally suppose indA0 = n and indB0 = m are possibly nonzero. Let Q be a Fredholm

operator with index −n, and let R be a Fredholm operator with index −m. Suppose (A,B) →

(A0, B0) in M . Then we may assume that all A have index n and all B have index m. Define the

following index zero operators on H⊕H⊕H:

A0 = A0 ⊕Q⊕ I and B0 = B0 ⊕ I ⊕R

and A and B similarly. Then (A, B) → (A0, B0) in M . By Lemmas 3.2.4 and 3.3.5, τ(A, B) =

τ(A,B) and τ(A0, B0) = τ(A0, B0). Then the above argument shows that τ(A,B)→ τ(A0, B0).

Recall, however, that joint torsion is defined in terms of the finite dimensional homology

spaces H•(A), H•(B), H•(C), H•(D), and H•(A,B,C,D). In particular, the dimensions of these

spaces are by no means continuous. Thus, the continuity of joint torsion may be seen as analogous

to the continuity of the Fredholm index.

Lemma 4.4.4. If [A,B] ∈ L1, then

τ(eA, eB) = etr[A,B].

Proof. This follows from Theorem 4.3.3 and Example 2.2.12.

It is convenient to state the following variational formula using the logarithmic derivative.

Thus ddz log u should be interpreted as u−1 d

dzu.

Corollary 4.4.5. Suppose A(z) is a differentiable family of operators such that [A(z), B] ∈ L1 for

every z. If, in addition, [A(z), B] is differentiable in L1, then

d

dzlog τ(eA(z), eB) = log τ(e

ddzA(z), eB).

Lemma 4.4.6. Whenever the following joint torsion numbers are defined, we have:

(1) τ(A,B1B2) = τ(A,B1) · τ(A,B2)

Page 64: Determinants in K-theory and operator algebras

57

(2) τ(A, I) = 1

(3) τ(A,B)−1 = τ(B,A)

(4) τ(A, I −A) = 1

(5) τ(A,−A) = 1

(6) τ(A,B) = τ(A∗, B∗)−1

(7) τ(A,B−1) = τ(A,B)−1

(8) τ(A,A) = (−1)indA

Proof. Properties (1)-(6) follow from the corresponding properties of the determinant invariant.

See for instance Lemma 4.2.14 and Theorem 4.2.17 of [50]. Property (7) follows from (1) and (2).

To verify (8), notice that the two torsion factors in the definition of joint torsion are the same.

Thus we are left with (−1)ν(A,A), where ν(A,A) is the sign in the definition of joint torsion. The

result follows since ν(A,A) = indA.

Lemma 4.4.7. Whenever the following joint torsion numbers are defined, we have:

(1) τ(A,A∗) ∈ R.

(2) If A and B are self-adjoint, then |τ(A,B)| = 1.

(3) If B is an idempotent, i.e. B2 = B, then τ(A,B) = 1.

(4) If A is self-adjoint and B is a partial isometry, then τ(A,B) ∈ R.

(5) If A and B are partial isometries, then |τ(A,B)| = 1.

Proof.

(1) By properties (6) and (3) of Lemma 4.4.6,

τ(A,A∗) = τ(A∗, A)−1 = τ(A,A∗).

Page 65: Determinants in K-theory and operator algebras

58

(2) Since A and B are self-adjoint, Lemma 4.4.6(6) implies that

τ(A,B) = τ(A,B)−1.

(3) By Lemma 4.4.6(1), τ(A,B) = τ(A,B)2, and the result follows since joint torsion is nonzero.

(4) First let T be any Fredholm operator which commutes with B modulo L1. Since B is a

Fredholm partial isometry, T also commutes with B∗ modulo L1. Since B∗B is a projection,

(3) implies that

τ(T ∗, B∗) · τ(T ∗, B) = τ(T ∗, B∗B) = 1

so by Lemma 4.4.6(6),

τ(T ∗, B) = τ(T,B). (4.15)

The result follows by setting T = A since A∗ = A.

(5) Applying (4.15) to both A and B yields

τ(A,B) = τ(A∗, B∗)

and the result follows by Lemma 4.4.6(6).

In (4), if A is in fact positive, we will use the behavior of joint torsion under the functional

calculus to show that τ(A,B) > 0 (Proposition 5.1.11).

Lemma 4.4.8. If A and B are commuting Fredholm operators, then for any µ 6= 0,

τ(A,µB) = µindA−dimH0+dimH2τ(A,B)

where H0 = H0(A,B) and H2 = H2(A,B) as usual denote the joint Koszul homology spaces.

Proof. The two long exact sequences EA and EµB in the definition of τ(A,µB) are the same as

those for τ(A,B), except for a factor of µ, given by the exponent on µ above.

Corollary 4.4.9. If A and B are commuting Fredholm operators, then

d

dµlog τ(A,µB) =

indA− dimH0 + dimH2

µ.

Page 66: Determinants in K-theory and operator algebras

59

4.4.3 Joint torsion and Hilbert-Schmidt operators

This section represents a first attempt at understanding torsion invariants in the context of

operator ideals other than the trace class. This work was motivated by Voiculescu’s conjectures on

the existence of lifts of almost normal operators [61]. See Proposition 4.4.10 below.

Smooth non-vanishing functions on the circle may be factored in terms of exponentials and

the function z 7→ z. As we will see, the joint torsion of Toeplitz operators may be calculated as

follows:

(1) τ(Tef , Teg) = exp 12πi

∫D df ∧ dg

(2) τ(Tef , Tz) = exp 12πi

∫S1 f d log z

(3) τ(Tz, Tz) = −1

In (1), f and g are extended to the unit disk D appropriately. Hence (1) recovers the Pincus

principal function [11, 12]. Terms like (3) carry the index data, a fact which seems to hold more

generally. Terms like (2) appear to be new “cross-terms.”

Suppose T is such that [T, T ∗] ∈ L1. As described in Section 3.1, the operator p(T, T ∗) is

well-defined for any p ∈ C[z, z], modulo L1. The exponential of p(T, T ∗) is similarly well-defined,

modulo L1, via the usual power series.

Proposition 4.4.10. Suppose [T, T ∗] ∈ L1, and suppose there exist a unitary U , a normal operator

N , and L ∈ L2 such that T = U∗NU + L. If p, q ∈ C[x, y], then

τ(exp p(T, T ∗), exp q(T, T ∗)) = 1.

The proof of this proposition proceeds by a series of lemmas. If L ∈ L2, then L2 ∈ L1, and

consequently I + L− eL ∈ L1. Hence,

τ(I + L1, I + L2) = τ(eL1 , eL2) = exp tr [L1, L2] = 1

whenever L1, L2 ∈ L2. The third equality follows from Proposition 2.1.6. More generally, one has

the following:

Page 67: Determinants in K-theory and operator algebras

60

Lemma 4.4.11. Suppose A is a Fredholm operator and L1, L2 ∈ L2 such that [A,L2] ∈ L1. Then

τ(A+ L1, I + L2) = τ(A, I + L2).

Proof. If A is invertible, then

τ(A+ L1, I + L2) = τ(A, I + L2) · τ(I +A−1L1, I + L2)

and the result follows by the observation above. In general, let Q be any Fredholm operator with

index opposite that of A, and let F be a finite rank operator such that A ⊕ Q + F is invertible.

The result follows from the invertible case since

τ(A+ L1, I + L2) = τ((A+ L1)⊕Q+ F, (I + L2)⊕ I)

and

τ(A, I + L2) = τ(A⊕Q+ F, (I + L2)⊕ I).

Lemma 4.4.12. Suppose N is a normal Fredholm operator and L ∈ L2 such that [N,L] ∈ L1.

Then

τ(N, I + L) = 1.

Proof. Since N is normal, by [60] we may write N = D+K for some K ∈ L2 and a diagonalizable

operator D. By changing basis, we find that

τ(N, I + L) = τ(D′ +K ′, I + L′)

where L′,K ′ ∈ L2 and D′ is diagonal. By the preceding lemma,

τ(D′ +K ′, I + L′) = τ(D′, I + L′).

By adding a finite rank digaonal operator, we may assume that D′ is invertible and diagonal. Write

D′ = exp d for a diagonal operator d. Then [d, L′] ∈ L1 and

τ(D′, I + L′) = τ(exp d, expL′) = exp tr [d, L′].

Moreover, tr [d, L′] = 0 since d is normal and L′ ∈ L2, and the result follows.

Page 68: Determinants in K-theory and operator algebras

61

More generally, one has the following:

Lemma 4.4.13. Suppose N1 and N2 are normal Fredholm operators and L1, L2 ∈ L2 are such that

[N1, N2], [N1, L2], and [L1, N2] ∈ L1. Then

τ(N1 + L1, N2 + L2) = τ(N1, N2).

Proof. By adding the orthogonal projections onto the kernels of N1 and N2, necessarily finite

dimensional, we may assume N1 and N2 are invertible normal operators. Hence

τ(N1 + L1, N2 + L2) = τ(N1, N2) · τ(I +N−11 L1, N2)

· τ(N1, I +N−12 L2) · τ(I +N−1

1 L1, I +N−12 L2)

The second, third, and fourth factors are all 1 by the preceding lemmas.

Proof of Proposition 4.4.10. We must show that tr[p(T, T ∗), q(T, T ∗)] = 0 since

τ(exp p(T, T ∗), exp q(T, T ∗)) = exp tr[p(T, T ∗), q(T, T ∗)].

First we note that p(T, T ∗) = N1 + L1 and q(T, T ∗) = N2 + L2, modulo L1, for some L1, L2 ∈ L2

and normal operators N1 and N2. If [N1, N2], [N1, L2], [L1, N2] ∈ L1, then the result follows from

the previous lemma. More generally, we find that both UN1U∗ and UN2U

∗ can be expressed as

sums and products of N and N∗. Moreover, [N1, L2] + [L1, N2] ∈ L1. Since N is the sum of a

diagonalizable operator and a Hilbert-Schmidt operator, one has tr([N1, L2]+ [L1, N2]) = 0 and the

result follows.

Page 69: Determinants in K-theory and operator algebras

Chapter 5

Explicit formulas for joint torsion

In this chapter we investigate the transformation of joint torsion under the functional calculus.

Combined with the results of the previous chapter, this will allow us to obtain explicit formulas for

the joint torsion of Toeplitz operators. We also investigate the relationship between joint torsion

and Tate tame symbols. The results in this chapter have appeared in the author’s work [46].

5.1 Transformation rules for joint torsion

5.1.1 Commutators

Let A and B be bounded operators on a fixed Hilbert space. Connes shows in [19] that if

[A,B] ∈ Lp, and either

(1) f is holomorphic on a neighborhood of σ(A), or

(2) A is self-adjoint and f is C∞ on σ(A),

then [f(A), B] ∈ Lp. In this section, we calculate the trace of such a commutator.

Lemma 5.1.1. If [A,B] ∈ L1 and f is an entire function, then

tr[f(A), B] = tr(f ′(A)[A,B]

).

Proof. Write f(z) =∑ckz

k. Then [f(A), B] =∑ck[A

k, B]. Using the identity

[Ak, B] =k∑l=1

Al−1[A,B]Ak−l (5.1)

Page 70: Determinants in K-theory and operator algebras

63

we find that

tr[Ak, B] = tr(k Ak−1[A,B]).

Hence

tr[f(A), B] = tr∑

kckAk−1[A,B]

= tr(f ′(A)[A,B]

)Let f be holomorphic on a neighborhood of the spectrum σ(A) of an operator A. By an

admissible contour Γ for defining f(A), we mean a collection of Jordan curves in the neighborhood

that enclose σ(A) on the left. Thus

f(A) =1

2πi

∫Γ(λ−A)−1f(λ) dλ. (5.2)

Proposition 5.1.2. Suppose [A,B] ∈ L1. If either

(1) f is holomorphic on a neighborhood of σ(A), or

(2) A is self-adjoint and f is C∞ on σ(A)

then

tr[f(A), B] = tr(f ′(A)[A,B]

).

Proof. Recall that in both cases [f(A), B] ∈ L1 by [19], and we adapt arguments therein.

(1) Let Γ be an admissible contour for defining f(A). Since

[(λ−A)−1, B] = (λ−A)−1[A,B](λ−A)−1

we find that λ 7→ [(λ − A)−1, B] is a continuous map from the resolvent set of A into L1,

and

[f(A), B] =1

2πi

∫Γ(λ−A)−1[A,B](λ−A)−1f(λ) dλ.

Moreover,

tr((λ−A)−1[A,B](λ−A)−1

)= tr

((λ−A)−2[A,B]

).

Page 71: Determinants in K-theory and operator algebras

64

Hence

tr[f(A), B] = tr

(1

2πi

∫Γ(λ−A)−2f(λ) dλ [A,B]

)= tr

(f ′(A)[A,B]

)(2) We may assume that f has compact support, so that f = g, the Fourier transform of a

Schwartz class function g. Hence

[f(A), B] =1√2π

∫[e−itA, B]g(t) dt.

By the preceding lemma,

tr[e−itA, B] = tr(−ite−itA[A,B]

)and again by continuity,

tr[f(A), B] = tr1√2π

∫−ite−itAg(t) dt[A,B]

= tr(f ′(A)[A,B]

)Corollary 5.1.3. With the same hypotheses as above,

τ(ef(A), eB) = τ(eA, ef′(A)B).

Proof. Since A and f ′(A) commute, we have f ′(A)[A,B] = [A, f ′(A)B]. The result then follows by

Lemma 4.4.4.

5.1.2 Perturbations

Analogues of Lemma 5.1.1 and Proposition 5.1.2 hold for suitable functions applied to Lp-

perturbations. We will need the following estimate for the exponential function:

Proposition 5.1.4. If A and A′ are self-adjoint with A−A′ ∈ Lp, then eitA − eitA′ ∈ Lp with

‖eitA − eitA′‖p ≤ C (|t|+ 1)

Page 72: Determinants in K-theory and operator algebras

65

where

C = max0≤t≤1

‖eitA − eitA′‖p.

Proof. Using the identity

rn − sn =n∑k=1

sk−1(r − s)rn−k (5.3)

we find that

‖eitnA − eitnA′‖p ≤ n‖eitA − eitA′‖p.

The result then follows by scaling.

Proposition 5.1.5. Let K ∈ Lp. If either

(1) f is holomorphic on a neighborhood of σ(K), or

(2) K is self-adjoint and f is C∞ on σ(K),

then f(K)− f(0)I ∈ Lp.

Note that in (2), σ(K) consists of 0 and real eigenvalues possibly accumulating to 0 by the

spectral theorem for compact self-adjoint operators.

Proof.

(1) Let Γ be an admissible contour for defining f(K). Then

f(K)− f(0)I =

∫Γ

[(λ−K)−1 − (λI)−1

]f(λ) dλ

= K

∫Γ(λ2 − λK)−1f(λ) dλ

The latter integral converges in norm, and the result follows.

(2) We may assume that f has compact support, so that f = g for a Schwartz class function

g. Then

f(K)− f(0)I =

∫ (e−itK − I

)g(t) dt.

The integral converges in the Lp-norm by the preceding proposition with A = K and

A′ = 0.

Page 73: Determinants in K-theory and operator algebras

66

Proposition 5.1.6. Let A−A′ ∈ Lp. If either

(1) f is holomorphic on a neighborhood of σ(A)∪σ(A′) and there is a contour that defines both

f(A) and f(A′), or

(2) A and A′ are self-adjoint and f is C∞ on σ(A) ∪ σ(A′),

then f(A)− f(A′) ∈ Lp.

Proof. The proof proceeds as in the previous proposition. For part (1), one uses the identity

(λ−A)−1 − (λ−A′)−1 = (λ−A)−1(A−A′)(λ−A′)−1.

5.1.3 Joint torsion

For a given Fredholm operator A, we begin with a simple characterization of holomorphic

functions f that preserve the Fredholmness of A. We will use the following factorization of holo-

morphic functions:

Definition 5.1.7. Let f be a holomorphic function on a neighborhood of a compact set K. Then

the collection of zeros λ ∈ K | f(λ) = 0 is finite. Define the polynomial

pK(z) =∏

λ∈K | f(λ)=0

(z − λ)ordλ(f)

where ordλ(f) is the order of the zero of f at λ. Then

f = pKqK

for a holomorphic function qK with no zeros in K.

The index formula (5.4) below is a special case of [29, Theorem 10.3.13]. See also [38, Theorem

1.1].

Proposition 5.1.8. Let A be a Fredholm operator and let f be holomorphic on a neighborhood of

σ(A). Then f(A) is Fredholm if and only if f−1(0) is disjoint from the essential spectrum σe(A).

In this case,

ind f(A) =∑

λ∈σ(A) | f(λ)=0

ordλ(f) · ind(A− λ). (5.4)

Page 74: Determinants in K-theory and operator algebras

67

Proof. Let p = pσ(A) and q = qσ(A) from the definition above. Then f(A) = p(A)q(A) and q

is invertible on a neighborhood of σ(A), so q(A) is invertible. The first assertion then follows

by factoring p, and the index formula follows by the additive property of the index: ind(ST ) =

indS + indT .

More generally one has the following sufficient condition for the Borel functional calculus to

preserve Fredholmness:

Proposition 5.1.9. Let A be a normal Fredholm operator, and let f ∈ L∞(σ(A)). If the sets

f−1(0) and f−1(±∞) are finite and disjoint from σe(A), then f(A) is Fredholm.

Proof. The strategy is to excise the sets f−1(0) and f−1(±∞) and use the resulting function to

construct a parametrix for f(A). Suppose f(λ) = 0, +∞, or −∞. Let Un be a nested sequence of

open subsets of σ(A) such that ∩Un = λ. Then χUn converges to χλ pointwise, so Pn = χUn(A)

converges to P = χλ(A) strongly. Now P is either 0 or the projection onto the λ-eigenspace of A,

which is finite dimensional since A−λ is Fredholm. Since Pn is a descending sequence of projections

that converge to a finite rank projection, there is an N for which Pn is finite dimensional for all

n > N .

Let Uλ = Un and χλ = χUλ for some n > N . By taking n large enough, we may assume

that the open sets Uλ are pairwise disjoint, where λ ranges over all the singularities and zeros of

f . Then

g = (1−∑λ

χλ)f +∑λ

χλ

is invertible in L∞(σ(A)), and g(A)−f(A) is a finite rank operator. Hence, g(A)−1 is a parametrix

for f(A) modulo finite rank operators, so f(A) is Fredholm.

Next we obtain a multiplicative analogue of (5.4):

Proposition 5.1.10. Suppose A and B are Fredholm operators with [A,B] ∈ L1. If f is holomor-

phic on a neighborhood of σ(A) and f(A) is Fredholm, then

τ(f(A), B) =∏

λ∈σ(A) | f(λ)=0

τ(A− λ,B)ordλ(f) · τ(q(A), B)

Page 75: Determinants in K-theory and operator algebras

68

with q = qσ(A) as in Definition 5.1.7, so that in particular q(A) is invertible.

Proof. First we note that [f(A), B] ∈ L1 by Proposition 5.1.2(1). Writing f = pq, we have

[p(A), B] ∈ L1, so [q(A), B] ∈ L1 as well. By multiplicativity,

τ(f(A), B) = τ(p(A), B) · τ(q(A), B).

Since p(A) is a product of factors A − λ, we find that τ(p(A), B) further factors as the product

above.

5.1.4 Positivity

In this section we investigate general conditions under which joint torsion is positive. This

is used to clarify the relationship between joint torsion and the polar decomposition, and also to

obtain variational formulas.

Proposition 5.1.11. Suppose A and B are Fredholm operators and [A,B] ∈ L1. If A is positive

and B is a partial isometry, then τ(A,B) > 0.

Proof. Let F = PkerA be the orthogonal projection onto kerA = imA⊥. Then A + F is positive-

definite. By Proposition 5.1.2(2), B commutes with T = (A+ F )1/2 modulo L1. Hence

τ(A+ F,B) = τ(T,B)2.

By Lemma 4.4.7, τ(T,B) ∈ R since T is self-adjoint. Hence

τ(A,B) = τ(A+ F,B) > 0.

Proposition 5.1.12. Suppose A and B are Fredholm operators with [A,B] ∈ L1 and [A,B∗] ∈ L1.

Then with respect to the polar decompositions

A = PAVA, B = PBVB

we have

|τ(A,B)| = τ(PA, VB) · τ(VA, PB)

Page 76: Determinants in K-theory and operator algebras

69

and consequently,

τ(A,B)

|τ(A,B)|= τ(PA, PB) · τ(VA, VB).

Proof. First notice that PA and VA are Fredholm since A is. Similarly, PB and VB are Fredholm.

We must show that the four joint torsion numbers above are well-defined, that is, the appropriate

commutators lie in L1. Our strategy is to show first that [PA, B], [A,PB], [PA, PB] ∈ L1, then

[VA, PB], [PA, VB] ∈ L1, and finally [VA, VB] ∈ L1.

If FA = PkerA, then A∗A+FA is invertible and commutes with B modulo L1. By Proposition

5.1.2(2), [(A∗A + FA)1/2, B] ∈ L1, so [PA, B] ∈ L1 as well, with PA = (A∗A)1/2. By reversing the

roles of A and B, we find that [A,PB] ∈ L1. Moreover, by replacing B by PB, we find that

[PA, PB] ∈ L1.

Next, we calculate modulo L1:

[VA, PB] ≡ [(PA + FA)−1PAVA, PB]

≡ [(PA + FA)−1A,PB]

≡ (PA + FA)−1[A,PB] + [(PA + FA)−1, PB]A

The first term is in L1 since [A,PB] ∈ L1, and the second term is in L1 since [PA, PB] ∈ L1 as well.

Similarly, [PA, VB] ∈ L1 by reversing the roles of A and B.

Again we calculate modulo L1:

[VA, VB] ≡ [(PA + FA)−1A, (PB + FB)−1B]

≡ (PA + FA)−1[A, (PB + FB)−1]B + (PA + FA)−1(PB + FB)−1[A,B]

+ [(PA + FA)−1, (PB + FB)−1]BA+ (PB + FB)−1[(PA + FA)−1, B]A

As before, all four of the above terms are evidently in L1.

By the multiplicative property of joint torsion,

τ(A,B) = τ(PA, VB) · τ(VA, PB) · τ(PA, PB) · τ(VA, VB).

The first two factors are positive by preceding proposition. The third factor has magnitude one by

Lemma 4.4.7(2), as does the last factor by Lemma 4.4.7(5).

Page 77: Determinants in K-theory and operator algebras

70

Proposition 5.1.13. Suppose A and B are Fredholm operators with [A,B] ∈ L1. If A is positive,

and B is a partial isometry, then for all t ≥ 0,

τ(At, B) = τ(A,B)t.

If A is positive-definite, then the formula holds for all t ∈ R.

Proof. First we note that τ(A,B) > 0 by Proposition 5.1.11, [At, B] ∈ L1 by Proposition 5.1.2(2),

and At is Fredholm with parametrix (A + PkerA)−t. The formula holds for positive integers by

repeated application of Lemma 4.4.6(1), and for t = 0 by Lemma 4.4.6(2). The formula also holds

for all positive rational numbers: if p and q are any positive integers, then

τ(Ap/q, B)q = τ(A,B)p.

If F = PkerA, then A+ F is positive and invertible, and τ(At, B) = τ((A+ F )t, B). We will

show that the map t 7→ ((A + F )t, B), t > 0, is a continuous map into the space M in Corollary

4.4.3. Since t 7→ (A+ F )t is continuous in norm, it suffices to show that

limt→0‖[(A+ F )t, B]‖1 = 0.

Since [log(A+ F ), B] ∈ L1 by Proposition 5.1.2(2), this follows from the estimate

‖[(A+ F )t, B]‖1 ≤ t et‖ log(A+F )‖‖[log(A+ F ), B]‖1.

Joint torsion is continuous on M by Corollary 4.4.3, so the map t 7→ τ(At, B) is continuous.

Thus the result extends from rational t to all t ≥ 0. Finally, if A is positive definite, then A−t is

also positive definite for any t > 0. By the above result for positive t, we find

τ(A−t, B)t = τ(A,B).

A similar result holds when A and B are positive. In this case, |τ(A,B)| = 1 by Lemma

4.4.7. Suppose A and B are positive Fredholm operators with [A,B] ∈ L1. If FA = PkerA and

FB = PkerB, then

φ(A,B) = −i tr [log(A+ FA), log(B + FB)] ∈ R

Page 78: Determinants in K-theory and operator algebras

71

is well-defined by Proposition 5.1.2. Since τ(A,B) = τ(A+ FA, B + FB), Lemma 4.4.4 gives

τ(A,B) = eiφ(A,B)

and φ(A,B) enjoys the additive versions of the properties in Lemma 4.4.6. Moreover, we have:

Proposition 5.1.14. If A and B are positive Fredholm operators with [A,B] ∈ L1, then for all

t > 0,

τ(At, B) = eitφ(A,B) = τ(A,B)t.

If A is positive-definite, then the formula holds for all t ∈ R.

Proof. This follows by noticing that τ(At, B) = τ((A + FA)t, B + FB), then using the fact that

A+ FA and B + FB have logarithms.

Corollary 5.1.15. Suppose A and B are Fredholm operators with [A,B] ∈ L1. If A is positive and

B is either positive or a partial isometry, then

d

dtlog τ(At, B) = log τ(A,B).

5.2 Fredholm modules

Let (A,H, F ) be a 2p-summable Fredholm module, i.e. [φ, F ] ∈ L2p for any φ ∈ A. Let

P = 12(F + I) be the projection onto the +1-eigenspace of F , so in particular, [φ, P ] ∈ L2p for any

φ ∈ A.

Definition 5.2.1. For φ ∈ A, form the abstract Toeplitz operator Tφ = PφP ∈ B(PH).

The main goal of this section is to show that f(Tφ) − Tf(φ) ∈ Lp for suitable functions f .

First let us record a corresponding result for the continuous functional calculus modulo compact

operators:

Proposition 5.2.2. Let T ∈ B be normal and let π : B → B/K be the quotient map onto the Calkin

algebra. If f ∈ C(σ(T )), then π(f(T )) = f(π(T )).

Page 79: Determinants in K-theory and operator algebras

72

Proof. The polynomial functional calculus commutes with the quotient map, so the result follows

from the Stone-Weierstrass Theorem by approximating f by polynomials.

Next we show that the entire functional calculus commutes with the symbol map modulo L2p.

This complements the results of [26] and Proposition 5.2.7 below. Assume that A is closed under

the entire functional calculus. Otherwise, we may replace A by the algebra generated by f(a), for

all a ∈ A and entire functions f . The resulting algebra still has the property that [φ, P ] ∈ L2p.

Indeed, we have already seen that if [a, P ] ∈ L2p and f is holomorphic on a neighborhood of σ(a),

then [f(a), P ] ∈ L2p.

Lemma 5.2.3. For any φ ∈ A and integer k > 1, T kφ − Tφk ∈ L2p, with

‖T kφ − Tφk‖2p ≤k(k − 1)

2‖φ‖k−1‖[φ, P ]‖2p.

Proof. Each term in the identity

(Pφ)kP − PφkP =k−1∑l=1

(Pφ)k−l[φl, P ]P

contains a commutator, so (Pφ)kP − PφkP ∈ L2p. Using the identity (5.1), we estimate

‖[φl, P ]‖2p ≤ l‖φ‖l−1‖[φ, P ]‖2p.

Hence

‖(Pφ)kP − PφkP‖2p ≤k−1∑l=1

l‖φ‖k−1‖[φ, P ]‖2p

and the result follows.

Definition 5.2.4. For an entire function f(z) =∑ckz

k, let f(z) =∑|ck|zk.

Proposition 5.2.5. For any φ ∈ A and any entire function f , Tf(φ) − f(Tφ) ∈ L2p with

‖Tf(φ) − f(Tφ)‖2p ≤‖[φ, P ]‖2p

2‖φ‖f ′′(‖φ‖)

Page 80: Determinants in K-theory and operator algebras

73

Proof. Write f(z) =∑ckz

k. The first two terms in the expansion

Tf(φ) − f(Tφ) =∞∑k=0

ck

(PφkP − (Pφ)kP

)vanish, and by Lemma 5.2.3 we estimate

‖Tf(φ) − f(Tφ)‖2p ≤∞∑k=2

|ck|k(k − 1)

2‖φ‖k−1‖[φ, P ]‖2p

≤ ‖[φ, P ]‖2p2‖φ‖

∞∑k=2

k(k − 1)|ck|‖φ‖k−2

≤ ‖[φ, P ]‖2p2‖φ‖

f ′′(‖φ‖)

In fact, the entire functional calculus commutes with the symbol map modulo Lp. First we

prove the following analogue of Lemma 5.2.3:

Lemma 5.2.6. For any φ ∈ A and integer k > 1, T kφ − Tφk ∈ Lp, with

‖T kφ − Tφk‖p ≤k(k − 1)

2‖φ‖k−2‖[φ, P ]‖22p.

Proof. First one verifies that

(Pφ)kP − PφkP =k−1∑l=1

P [P, φl][P, φ](Pφ)k−l−1P

using the identity P [P,ψ][P, χ]P = Pψ(P − I)χP . Each term of the sum contains a product of

commutators, so it is in Lp. Moreover, by (5.1),

‖[P, φl]‖2p ≤ l‖φ‖l−1‖[φ, P ]‖2p.

Hence

‖(Pφ)kP − PφkP‖p ≤k−1∑l=1

l‖φ‖k−2‖[φ, P ]‖22p

and the result follows.

Proposition 5.2.7. For any φ ∈ A and any entire function f , Tf(φ) − f(Tφ) ∈ Lp with

‖Tf(φ) − f(Tφ)‖p ≤1

2‖[φ, P ]‖22pf ′′(‖φ‖).

Page 81: Determinants in K-theory and operator algebras

74

Proof. Write f(z) =∑ckz

k. The first two terms in the expansion

Tf(φ) − f(Tφ) =∞∑k=0

ck

(PφkP − (Pφ)kP

)vanish, and by Lemma 5.2.6 we estimate

‖Tf(φ) − f(Tφ)‖p ≤∞∑k=2

|ck|k(k − 1)

2‖[φ, P ]‖22p‖φ‖k−2

=1

2‖[φ, P ]‖22p

∞∑k=2

k(k − 1)|ck|‖φ‖k−2

=1

2‖[φ, P ]‖22pf ′′(‖φ‖)

We will need a sharper estimate for the exponential function:

Proposition 5.2.8. If φ is self-adjoint, then eTitφ − Teitφ ∈ Lp for any t ∈ R, and

‖eTitφ − Teitφ‖p ≤ (|t|+ 1)2(c1 + c2)

where

c1 = max0≤t≤1

‖eTitφ − Teitφ‖p and c2 = max0≤t≤1

‖[P, eitφ]‖22p

Proof. Setting r = eTitφ and s = Teitφ in identity (5.3), we obtain

‖eTintφ − (Teitφ)n‖p ≤ ‖eTitφ − Teitφ‖pn∑k=1

‖Teitφ‖k−1‖eTitφ‖n−k.

Since ‖eTitφ‖ = 1 and ‖Teitφ‖ ≤ 1, we find

‖eTintφ − (Teitφ)n‖p ≤ n‖eTitφ − Teitφ‖p. (5.5)

By Lemma 5.2.6,

‖(Tf )n − Tfn‖p ≤n(n− 1)

2‖[P, f ]‖22p‖f‖n−2.

Setting f = eitφ, we have ‖f‖ = 1, so

‖(Teitφ)n − Teintφ‖p ≤n(n− 1)

2‖[P, eitφ]‖22p. (5.6)

Combining (5.5) and (5.6), we find

‖eTintφ − Teintφ‖p ≤ n2(‖eTitφ − Teitφ‖p + ‖[P, eitφ]‖22p)

and the result follows by scaling.

Page 82: Determinants in K-theory and operator algebras

75

We are now able to obtain an analogue of Proposition 5.1.6 for summable Fredholm modules.

Below we regard φ as an operator on H and Tφ as an operator on PH, and we view σ(φ), σ(Tφ),

f(φ), and f(Tφ) accordingly.

Theorem 5.2.9. If either

(1) f is holomorphic on a neighborhood of σ(φ) ∪ σ(Tφ) and there is a contour Γ that defines

both f(φ) and f(Tφ), or

(2) φ is self-adjoint and f is C∞ on σ(φ) ∪ σ(Tφ),

then f(Tφ)− Tf(φ) ∈ Lp.

Proof.

(1) Notice that (λ− PφP )−1 − P (λ− φ)−1P can be written as

P [(λ− φ)−1, P ][φ, P ]P (λ− PφP )−1.

Since [P, φ] ∈ L2p, the assignment

λ 7→ (λ− PφP )−1 − P (λ− φ)−1P

is a continuous map into Lp. Hence

f(Tφ)− Tf(φ) =1

2πi

∫Γ

((λ− PφP )−1 − P (λ− φ)−1P

)f(λ) dλ

converges in Lp.

(2) As in Proposition 5.1.5, we may assume that f has compact support, so that f = g for a

Schwartz class function g. Then

f(Tφ)− Tf(φ) =1√2π

∫ (eT−itφ − Te−itφ

)g(t) dt

and the result follows by Proposition 5.2.8.

Page 83: Determinants in K-theory and operator algebras

76

5.3 Toeplitz operators and tame symbols

In this section we apply our techniques to Toeplitz operators and obtain formulas for joint

torsion in terms of Tate tame symbols. Let P : L2(S1) → H2(S1) be the orthogonal projection

onto the Hardy space H2(S1). Any function φ ∈ L∞(S1) defines a bounded operator on L2(S1) by

multiplication by φ. Then one has the Toeplitz operator Tφ = PφP .

5.3.1 H∞ symbols

Proposition 5.3.1. Suppose φ ∈ C(S1) ∩H∞(S1) is invertible in H∞(S1).

(1) If |λ| > 1, then τ(Tφ, Tz − λ) = 1.

(2) If |λ| < 1, then τ(Tφ, Tz−λ) = φ(λ), with φ extended holomorphically to the interior of the

unit disk.

Proof. First notice that Tφ is invertible with inverse T1/φ. If |λ| > 1, then z − λ is invertible in

H∞(S1) as well. The operators Tφ and Tz − λ commute, so in this case τ(Tφ, Tz − λ) = 1.

Now suppose |λ| < 1. By Lemma 4.4.6(6), it is enough to show that τ(Tz − λ, Tφ) = φ(λ).

In this case, coker(Tz − λ) = 0 and

ker(Tz − λ) = span

(1

1− λz=

∞∑k=0

(λz)k

).

The operator Tφ acts as multiplication by φ(λ) on the one dimensional subspace ker(Tz − λ). In

particular,

detTφ|ker(Tz−λ) = φ(λ).

This is the joint torsion by Lemma 3.2.4 since Tφ is invertible and commutes with Tz − λ.

Proposition 5.3.2. Let λ, µ ∈ C.

(1) If |λ1| > 1 and |λ2| > 1, then τ(Tz − λ1, Tz − λ2) = 1.

(2) If |λ1| < 1 and |λ2| > 1, then τ(Tz − λ1, Tz − λ2) = (λ1 − λ2)−1.

Page 84: Determinants in K-theory and operator algebras

77

(3) If |λ1| > 1 and |λ2| < 1, then τ(Tz − λ1, Tz − λ2) = λ2 − λ1.

(4) If |λ1| < 1 and |λ2| < 1, then τ(Tz − λ1, Tz − λ2) = −1.

Proof. In case (1), both Tz − λ1 and Tz − λ2 are invertible in H∞(S1) and commute with each

other, so τ(Tz − λ1, Tz − λ2) = 1.

Cases (2) and (3) follow from the preceding proposition.

For (4), we use the multiplicative property of joint torsion:

τ(Tz − λ1, Tz − λ2) = τ(Tz, Tz) · τ(Tz, Tz(z−λ2))−1

· τ(Tz(z−λ1), Tz)−1 · τ(Tz(z−λ1), Tz(z−λ2))

The first term is −1, and the last term is 1 since Tz(z−λ1) and Tz(z−λ2) are invertible and commute

with each other. The middle two terms are both 1 by the above proposition. Hence, τ(Tz−λ1, Tz−

λ2) = −1.

Let us recall the notion of tame symbol mentioned in Section 3.1. If f and g are meromorphic

at λ ∈ C, then the quotient

fordλ(g)

gordλ(f)

is regular at λ. Here, ordλ denotes the order of the zero or pole at λ.

Definition 5.3.3. The tame symbol cλ(f, g) of f and g at λ is defined as

cλ(f, g) = (−1)ordλ(f)·ordλ(g) fordλ(g)

gordλ(f)(λ).

Definition 5.3.4. If a ∈ C is nonzero, the Blaschke factor Ba is

Ba(z) =|a|a

a− z1− az

.

Let B0(z) = z and B∞(z) = z. A product of Blaschke factors is known as a Blaschke product.

Notice that for z ∈ S1, we have

Ba(z) = Ba(z) = B1/a(z). (5.7)

The preceding propositions may be rephrased in terms of tame symbols, and in fact we have:

Page 85: Determinants in K-theory and operator algebras

78

Proposition 5.3.5. Suppose f and g are products of

(1) invertible functions in C(S1) ∩H∞(S1),

(2) polynomials, and

(3) Blaschke factors Ba with |a| < 1.

If f and g are non-vanishing on S1, then

τ(Tf , Tg) =∏|λ|<1

cλ(f, g)

Proof. A straightforward calculation with f(z) = z − λ1 and g(z) = z − λ2 verifies that

∏|λi|<1

cλi(z − λ1, z − λ2)

agrees with (1)-(4) in Proposition 5.3.2. The result then holds for polynomials since both joint

torsion and the tame symbol are multiplicative. By Proposition 5.3.1, we find that the result

holds for factors of type (1) and (2). If |a| < 1, then Ba is the product of a polynomial and

(1− az)−1 ∈ H∞(S1). Hence factors of type (3) are products of types (1) and (2).

We will need the following Beurling-Szego factorization into inner and outer functions. See

for instance [18].

Theorem 5.3.6. If f ∈ H∞(S1) is continuous and non-vanishing on S1, then there exists an outer

function φ that is invertible in H∞(S1) such that

f = φ ·∏

Ba

where the above product is taken over finitely many zeros a with |a| < 1.

The following result was first obtained in [15, Proposition 1]. See also [47], and see [40] for a

generalization to the multivariable setting.

Page 86: Determinants in K-theory and operator algebras

79

Theorem 5.3.7. If f, g ∈ H∞(S1) are continuous and non-vanishing on S1, then τ(Tf , Tg) is the

product of tame symbols:

τ(Tf , Tg) =∏|a|<1

ca(f, g).

Proof. As in the preceding theorem, write

f = φf ·∏

Ba, g = φg ·∏

Bb.

By Proposition 5.3.5, the joint torsion numbers

τ(φf , φg), τ(φf , Bb), τ(Ba, φg), τ(Ba, Bb)

agree with the corresponding tame symbols. The result then follows since both joint torsion and

the tame symbol are bimultiplicative.

5.3.2 L∞ symbols

In this section we extend the above result to the noncommutative setting.

Proposition 5.3.8. Suppose φ ∈ C(S1) ∩H∞(S1) is invertible in H∞(S1).

(1) If |λ| > 1, then τ(Tφ, Tz − λ) = φ(1/λ)φ(0) .

(2) If |λ| < 1, then τ(Tφ, Tz − λ) = 1φ(0) .

Proof. If λ = 0, then

τ(Tφ, Tz) · τ(Tφ, Tz) = τ(Tφ, I) = 1.

By Proposition 5.3.1, the second factor is φ(0), so the result follows in this case.

If λ 6= 0, we may write

− 1

λ(z − λ)z = z − 1

λ

so that

τ(Tφ, T−1/λ) · τ(Tφ, Tz−λ) · τ(Tφ, Tz) = τ(Tφ, Tz−1/λ)

Page 87: Determinants in K-theory and operator algebras

80

The first factor is 1 since φ is invertible and the third factor is φ(0). Hence

τ(Tφ, Tz−λ) =τ(Tφ, Tz−1/λ)

φ(0)

and result follows by the Proposition 5.3.1.

Proposition 5.3.9. Let λ, µ ∈ C.

(1) If |λ1| > 1 and |λ2| > 1, then τ(Tz − λ1, Tz − λ2) = 1− (λ1λ2)−1.

(2) If |λ1| < 1 and |λ2| > 1, then τ(Tz − λ1, Tz − λ2) = −λ−12 .

(3) If |λ1| > 1 and |λ2| < 1, then τ(Tz − λ1, Tz − λ2) = −λ−11 .

(4) If |λ1| < 1 and |λ2| < 1, then τ(Tz − λ1, Tz − λ2) = (λ1λ2 − 1)−1.

Proof. This result follows from Proposition 5.3.2, as the preceding proposition follows from Propo-

sition 5.3.1.

Notice that z − λ extends meromorphically to the interior of the unit disk as

1

z− λ

with a simple pole at 0 and a simple zero at 1λ . A straightforward verification shows that the

previous two propositions express the joint torsion as a product of tame symbols. Since a Blaschke

factor is the ratio of two linear factors, we have the following noncommutative generalization of

Proposition 5.3.5:

Proposition 5.3.10. Suppose f and g are products of

(1) invertible functions in C(S1) ∩H∞(S1),

(2) trigonometric polynomials in z and z, and

(3) Blaschke factors Ba with a ∈ C ∪ ∞.

Page 88: Determinants in K-theory and operator algebras

81

If f and g are non-vanishing on S1, then

τ(Tf , Tg) =∏|λ|<1

cλ(f, g).

Here f and g have been extended meromorphically to the interior of the unit disk.

Now suppose f ∈ L∞(S1) such that Tf is Fredholm. Then f is continuous and non-vanishing

on S1, say with winding number n. The function z−nf(z) has winding number zero, so there is a

continuous function f such that

ef(z) = z−nf(z).

Let f+ = P f , f− = (I − P )f , where P : L2(S1) → H2(S1) is the orthogonal projection as usual.

Then

f(z) = z−nef−ef+ . (5.8)

Thus we may write f = f−f+ with f+, f− ∈ H∞ continuous and non-vanishing. By Theorem 5.3.6,

we may write

f+ = f1 ·∏

Ba, f− = f2 ·∏

Bb

where f1, f2 ∈ H∞ are invertible in H∞, the zeros a satisfy |a| < 1, and the zeros b satisfy |b| > 1.

Letting f0 be the product of Blaschke factors above, we have the factorization

f = f0f1f2.

If f is smooth, then so is z−nf , and we can take f to be smooth as well. Consequently f+

and f− are smooth, for example because the projection P can be expressed in terms of the Hilbert

transform, which preserves regularity. Hence the related functions f±, fi, i = 0, 1, 2, are smooth as

well. Define gi, i = 0, 1, 2, similarly. As in Proposition 5.1.10, we see that joint torsion factors as a

discrete part (tame symbols) and a continuous part (a determinant):

Theorem 5.3.11. If f, g ∈ C∞(S1) are non-vanishing on S1, then

τ(Tf , Tg) =∏|a|<1

ca(f0f1, g0g1) · ca(g0, f2)

ca(f0, g2)·τ(Tf1 , Tg2)

τ(Tg1 , Tf2).

Page 89: Determinants in K-theory and operator algebras

82

Here fi, gi are as above, and

τ(Tf1 , Tg2) = exp

(1

2πi

∫log f1 d(log g2)

).

for continuous choices of logarithms of f1 and g2, and similarly for τ(Tg1 , Tf2).

Proof. By the multiplicative property of joint torsion, we find that

τ(Tf , Tg) = τ(Tf0f1 , Tg0g1) · τ(Tf0f1 , Tg2) · τ(Tf2 , Tg0g1) · τ(Tf2 , Tg2).

The first factor is the product of tame symbols by Theorem 5.3.7. The fourth factor is 1 since Tf2

and Tg2 are invertible commuting operators. Next we calculate the second factor; the third factor

is dealt with similarly. Again using multiplicativity, the second factor is

τ(Tf0 , Tg2) · τ(Tf1 , Tg2).

For the first factor above, notice that f0 is still a Blaschke product and g2 ∈ H∞ is invertible

in H∞. Hence τ(Tf0 , Tg2) = τ(Tf0, Tg2)−1 by Lemma 4.4.6(6), and the latter is ca(f0, g2)−1 by

Proposition 5.3.10. In the second factor both Tf1 and Tg2 are invertible. Hence their joint torsion

is the multiplicative commutator

det(Tf1Tg2Tf−1

1Tg−1

2

)which is calculated as the claimed integral by Lemma 4.4.4 and Theorem 2.3.6.

5.3.3 An integral formula

Now we apply and refine the results of Section 5.2 in the case of Toeplitz operators. Let

L2 = L2(S1) and H2 = H2(S1). As a consequence of Proposition 2.3.4, we obtain the following:

Theorem 5.3.12. Let φ ∈ L∞(S1) ∩W12,2(S1). If either

(1) f is holomorphic on a neighborhood of φ(S1), or

(2) φ is real-valued and f is C∞ on φ(S1),

then f(Tφ)− Tfφ ∈ L1(H2).

Page 90: Determinants in K-theory and operator algebras

83

Proof. Let Γ be an admissible contour for defining f(Tφ) ∈ B(H2), as in (5.2). By the above

discussion Γ can also be used to define f(φ) ∈ B(L2), that is,

f(φ) =1

2πi

∫Γ(λ− φ)−1f(λ) dλ

The result then follows by Theorem 5.2.9 with p = 1.

We conclude with an illustration of the above results by deriving an integral formula for the

joint torsion of Toeplitz operators [15, Theorem 7]. See also [30].

Theorem 5.3.13. If f, g ∈ C∞(S1) are non-vanishing functions, then

τ(Tf , Tg) = exp1

2πi

(∫S1

log f d(log g)− log g(p)

∫S1

d(log f)

).

The integrals are taken counterclockwise starting at any point p = eiα ∈ S1. If h(eiθ) =

|h(eiθ)|eiφ(θ) for a continuous function φ : [α, α+ 2π]→ R, then we take log h(eiθ) = log |h|+ iφ(θ).

Any other choice of log h will differ by a multiple of 2πi and hence will leave the quantity in the

theorem unaffected.

Proof. Let n and m be the winding numbers of f and g, respectively. Define f , f+, and f− as in

(5.8), and similarly for g. By Theorem 5.3.12, Tef− eTf ∈ L1, so

τ(Tf , Tg) = τ(Tz, Tz)mn · τ(Tz, Teg)

n · τ(Tef, Tz)

m · τ(eTf , eTg).

The first factor is (−1)mn by Proposition 5.3.2. By applying Proposition 5.3.1 with λ = 0 and both

φ = eg+ and φ = eg− , we find that the second term is e−ng+(0). Similarly, the third term is emf+(0).

By Lemma 4.4.4 and Theorem 2.3.6, the fourth term is

exp

(1

2πi

∫f dg

).

Hence

τ(Tf , Tg) = exp

(πimn+mf+(0)− ng+(0) +

1

2πi

∫f dg

). (5.9)

Page 91: Determinants in K-theory and operator algebras

84

Now we calculate the last term in the exponential:∫f dg =

∫log(e−inθf) d(log(e−imθg)

=

∫−inθ dg +

∫log f d(log(e−imθg))

Integration by parts gives ∫−inθ dg = −inθg|α+2π

α +

∫ing dθ.

The first term is −2πing(p) since g has winding number zero. By writing g in terms of the

orthonormal basis elements eikθ, we see that the second term is 2πing+(0). Next we calculate∫log f d(log(e−imθg)) =

∫f · −imdθ +

∫inθ · −imdθ +

∫log f d(log g).

As before the first term is −2πimf+(0), and the second term is 2mnπ2 + 2πmnα. Combining this

with (5.9) gives

τ(Tf , Tg) = exp

(−ng(p)− imnα+

1

2πi

∫log f d(log g)

).

The result follows since the first term is

−n(−imα+ log g(p)) = imnα− 1

2πilog g(p)

∫d(log f).

Page 92: Determinants in K-theory and operator algebras

Bibliography

[1] J. Anderson and L. N. Vaserstein. Commutators in ideals of trace class operators. IndianaUniv. Math. J., 354(2):345–372, 1986.

[2] William B. Arveson. p-summable commutators in dimension d. J. Operator Theory, 54(1):101–117, 2005.

[3] M. F. Atiyah and R. Bott. A Lefschetz fixed point formula for elliptic complexes. Ann. ofMath., 86:374–407, 1967.

[4] A. A. Beılinson. Higher regulators and values of L functions. J. Soviet Math., 30:2036–2070,1985.

[5] Jean-Michel Bismut and Daniel S. Freed. The analysis of elliptic families. I. Metrics andconnections on determinant bundles. Comm. Math. Phys., 106(1):159–176, 1986.

[6] Jean-Michel Bismut, Henri Gillet, and Christophe Soule. Analytic torsion and holomorphicdeterminant bundles. i. bott-chern forms and analytic torsion. Comm. Math. Phys. 115,115(1):49–78, 1988.

[7] L. G. Brown. The determinant invariant for operators with compact self-commutators. In Proc.Conf. on Operator Theory, volume 345 of Lecture Notes in Math., pages 210–228. Springer-Verlag, Berlin, Heidelberg, and New York, 1973.

[8] L. G. Brown. Operator algebras and algebraic k-theory. Bull. Amer. Math. Soc., 81:1119–1121,1975.

[9] J. Bruning and M. Lesch. Hilbert complexes. J. Funct. Anal., 108(1):88–132, 1992.

[10] Ulrich Bunke. A regulator for smooth manifolds and an index theorem. arXiv:1407. 1379,pages 1–53, 2014.

[11] Richard Carey and Joel Pincus. An exponential function for determining functions. IndianaUniv. Math. J., 23:1155–1165, 1974.

[12] Richard Carey and Joel Pincus. Commutators, symbols and determining functions. J.Functional Analysis, 19:50–80, 1975.

[13] Richard Carey and Joel Pincus. Reciprocity for Fredholm operators. Integral Equations andOperator Theory, 9:469–501, 1986.

Page 93: Determinants in K-theory and operator algebras

86

[14] Richard Carey and Joel Pincus. Joint torsion. Preprint, pages 1–174, 1995-1996.

[15] Richard Carey and Joel Pincus. Joint torsion of Toeplitz operators with H∞ symbols. IntegralEquations and Operator Theory, 33:273–304, 1999.

[16] Richard Carey and Joel Pincus. Perturbation Vectors. Integral Equations and OperatorTheory, 35:271–365, 1999.

[17] L. A. Coburn. Toeplitz operators on odd spheres. In Proceedings of the Conference OperatorTheory (Dalhousie Univ. Halifax, N.S., 1973), volume 345 of Lecture Notes in Math., pages7–12. Springer, Berlin, 1973.

[18] Peter Colwell. Blaschke products. Bounded analytic functions. University of Michigan Press,Ann Arbor, MI, 1985.

[19] Alain Connes. Non-commutative differential geometry. IHES Publ. Math., 62:257–360, 1985.

[20] Alain Connes and Max Karoubi. Caractere multiplicatif d’un module de Fredholm. K-Theory,2(3):431–463, 1988.

[21] Alain Connes and Henri Moscovici. Cyclic cohomology, the Novikov conjecture and hyperbolicgroups. Topology, 29(3):345–388, 1990.

[22] Raul E. Curto. Fredholm and invertible n-tuples of operators. The deformation problem.Trans. Amer. Math. Soc., 266:129–159, 1981.

[23] P. Deligne. Le symbol modere. Inst. Hautes Etudes Sci. Publ. Math., 73:147–181, 1991.

[24] Ronald G. Douglas. Banach algebra techniques in operator theory. Springer-Verlag, 2nd ed.edition, 1998.

[25] Ronald G. Douglas. A new kind of index theorem. In Analysis, geometry and topology ofelliptic operators, pages 369–382. World Sci. Publ., Hackensack, NJ, 2006.

[26] T. Ehrhardt. A generalization of Pincus’ formula and Toeplitz operator determinants. Arch.Math. (Basel), 80(3):302–309, 2003.

[27] Jorg Eschmeier. Quasicomplexes and Lefschetz numbers. Acta. Math. Sci. (Szeged), 79(3-4):611–621, 2013.

[28] Jorg Eschmeier and Mihai Putinar. Spectral theory and sheaf theory. iii. J. Reine Agnew.Math., 354:150–163, 1984.

[29] Jorg Eschmeier and Mihai Putinar. Spectral decompositions and analytic sheaves, volume 10of London Mathematical Society Monographs. New Series. The Clarendon Press, OxfordUniversity Press, 1996.

[30] H. Esnault and E. Viehweg. Deligne-Beılinson cohomology. In Beılinson’s conjectures on specialvalues of L-functions, volume 4 of Perspect. Math., pages 43–91. Academic Press, Boston, MA,1988.

[31] Bjorn Gustafsson and Vladimir G. Tkachev. The resultant on compact Riemann surfaces.Comm. Math. Phys., 286(1):313–358, 2009.

Page 94: Determinants in K-theory and operator algebras

87

[32] J. W. Helton and R. Howe. Integral operators: commutators, traces, index and homology.In Proc. Conf. on Operator Theory, volume 345 of Lecture Notes in Math., pages 141–209.Springer-Verlag, 1973.

[33] J. William Helton and Roger E. Howe. Traces of commutators of integral operators. ActaMath., 135(3-4):271–305, 1975.

[34] Nigel Higson and John Row. Analytic K-Homology. Oxford University Press, 2000.

[35] Jens Kaad. A calculation of the multiplicative character. J. Noncommut. Geom., 5(3):351–385,2011.

[36] Jens Kaad. Comparison of secondary invariants of algebraic k-theory. J. K-Theory, 8(1):169–182, 2011.

[37] Jens Kaad. Joint torsion of several commuting operators. Adv. Math., 229:442–486, 2012.

[38] Jens Kaad and Ryszard Nest. A transformation rule for the index of commuting operators.arXiv:1208.1862, pages 1–28, 2012.

[39] Jens Kaad and Ryszard Nest. Canonical holomorphic sections of determinant line bundles.arXiv:1403.7937, pages 1–45, 2014.

[40] Jens Kaad and Ryszard Nest. Tate tame symbol and the joint torsion of commuting operators.arXiv:1408.3851, pages 1–26, 2014.

[41] F. F. Knudsen and D. Mumford. The projectivity of the moduli of stable curves. i. Math.Scand., 39(1):19–55, 1976.

[42] Maxim Kontsevich and Simeon Vishik. Geometry of determinants of elliptic operators. InFunctional analysis on the eve of the 21st century (New Brunswick, NJ, 1993), volume 131 ofProgr. Math., pages 173–197. Birkhauser Boston, Boston, MA, 1995.

[43] Matthias Lesch. Pseudodifferential operators and regularized traces. In Motives, QuantumField Theory, and Pseudodifferential Operators, volume 12 of Clay Mathematics Proceedings,pages 37–72. Amer. Math. Soc., 2010.

[44] Jean-Lois Loday. Cyclic homology. Springer-Verlag, 1992.

[45] Joseph Migler. Functional calculus and joint torsion of pairs of almost commuting operators.arXiv:1409.6289, 2014.

[46] Joseph Migler. Joint torsion equals the determinant invariant. To appear in the Journal ofOperator Theory, arXiv:1403.4882, 2015.

[47] Efton Park. The tame symbol and determinants of Toeplitz operators. arXiv:0905.4510, pages1–10, 2009.

[48] Daniel Quillen. Determinants of Cauchy-Riemann operators over a Riemann surface.Funktsional. Anal. i Prilozhen., 19(1):37–41, 1985.

[49] D. B. Ray and I. M. Singer. Analytic torsion for complex manifolds. Ann. of Math., 98:154–177,1973.

Page 95: Determinants in K-theory and operator algebras

88

[50] Jonathan Rosenberg. Algebraic K-theory and its applications. Springer-Verlag, 1994.

[51] Barry Simon. Trace ideals and their applications, 2nd ed. American Mathematical Society,2005.

[52] R. G. Swan. Vector bundles and projective modules. Trans. Amer. Math. Soc., 105:264–277,1962.

[53] G. Szego. Ein Grenzwertsatz uber die Toeplitzschen Determinanten einer reelen positivenfunction. Math. Ann., 76:490–503, 1915.

[54] G. Szego. On certain Hermitian forms associated with the Fourier series of a positive function.Festschrift Marcel Riesz, pages 228–238, 1952.

[55] N. Tarkhanov. The Euler characteristic of a Fredholm quasicomplex. Funktsional. Anal. iPrilozhen., 41(4):87–93, 2007.

[56] N. Tarkhanov and D. Wallenta. The Lefschetz number of sequences of trace class curvature.Mathematical Sciences, 6(44), 2012.

[57] N. N. Tarkhanov. Euler characteristic of Fredholm quasicomplexes. Functional Analysis andIts Applications, 41(4):318–322, 2007.

[58] John Tate. Symbols in arithmetic. In Actes du Congres International des Mathematiciens(Nice, 1970), Tome 1, pages 201–211. Gauthier-Villars, Paris, 1971.

[59] Joseph L. Taylor. A joint spectrum for several commuting operators. J. Functional Analysis,6:172–191, 1970.

[60] Dan Voiculescu. Some results on norm-ideal perturbations of Hilbert space operators. J.Operator Theory, 2:3–37, 1979.

[61] Dan Voiculescu. Almost normal operators mod Hilbert-Schmidt and the K-theory of thealgebras EΛ(Ω). arXiv:1112.4930, 2011.

[62] Antony Wassermann. Analysis of operators. www.dpmms.cam.ac.uk/%7Eajw, 2006.

[63] Mariusz Wodzicki. Algebraic K-theory and functional analysis. In First European Congress ofMathematics, Vol. II (Paris, 1992), volume 120 of Progr. Math., pages 485–496. Birkhauser,Basel, 1994.

Page 96: Determinants in K-theory and operator algebras

Appendix A

The existence of perturbations

Let R be a ring and let J be a two-sided ideal. Suppose a1, . . . , an ∈ R/J are commuting

elements. We begin by describing some obstructions to lifting the ai to commuting Ai ∈ R. The

strategy is to replace R and R/J by simple models – either a free ring or one with the minimal

number of relations – and look at homomorphisms out of them obtained by lifting commuting

elements. Later we specialize the case when R = B, the algebra of bounded operators on a Hilbert

space, and J = L1, the ideal of trace class operators.

Let q : R → R/J be the quotient map. First suppose R is a unital ring. Let ϕ :

Z[x1, . . . , xn] → R/J be the unital homomorphism that maps xi to ai. Let F be a covariant

functor on rings. If there exist commuting lifts Ai ∈ R of the ai, then we can define a homorphism

ϕ : Z[x1, . . . , xn]→ R as above, and the following diagram commutes:

F(R)

Fq

F(Z[x1, . . . , xn])

Fϕ66

Fϕ// F(R/J)

If F(Z[x1, . . . , xn]) 6= 0, pick any nonzero x ∈ F(Z[x1, . . . , xn]). Then Fϕ(x) ∈ imFq, thus

providing an obstruction to lifting the ai to commuting Ai.

Likewise, if G is a contravariant functor on rings, then the following diagram commutes:

G(R)Gϕ

ww

G(Z[x1, . . . , xn]) G(R/J)

Gq

OO

Gϕoo

Page 97: Determinants in K-theory and operator algebras

90

In general, if R is unital or otherwise, let

Xn = 〈x1, . . . , xn |xixj = xjxi〉

be the free commutative ring on n generators. Then one may replace Z[x1, . . . , xn] above everywhere

with Xn.

Next consider the “AB = CD” problem raised in Section 3.3. Thus let a1, a2 ∈ R/J be two

commuting elements and consider the ring

Y2 = 〈x1, x2, y1, y2 |x1x2 = y2y1〉.

Let ψ : Y2 → R/J be the homomorphism defined by xi, yi 7→ ai. This map factors through X2. If

there exist lifts Ai, Bi ∈ R of ai modulo J such that A1A2 = B2B1, then let ψ : Y2 → R be the

homomorphism defined by xi 7→ Ai, yi 7→ Bi. One may then proceed as above.

If one of the ai is invertible, it turns out that there is no obstruction to finding lifts Ai, Bi ∈ R

of ai modulo J such that A1A2 = B2B1. In fact, we have the following slightly stronger result,

which was obtained in [27] in the case when R = B and J = Lp for some p.

Proposition A.0.14. Let R be a unital ring and let J be a two-sided ideal in R. Let a1, a2 ∈ R/J

be commuting elements.

(1) If a1 is invertible, then for any lifts A1, B1 ∈ R of a1, there exist lifts A2, B2 of a2 such

that A1A2 = B2B1.

(2) If a2 is invertible, then for any lifts A2, B2 ∈ R of a2, there exist lifts A1, B1 of a1 such

that A1A2 = B2B1.

Proof. For the proof of (1), let C2 be a lift of a2 and let Q1 be any lift of a−11 . Then we may take

A2 = C2Q1B1, B2 = A1C2Q1

since

A1A2 = A1C2Q1B1

= B2B1

Page 98: Determinants in K-theory and operator algebras

91

The second claim is verified similarly.

In light of the assumptions in [16], there is also the case when A1, A2 are specified and

required to appear in the answer. Of course, Theorem 4.3.3 and the above proposition show that

these questions are no longer necessary for the study of joint torsion of two almost commuting

Fredholm operators.

So far the treatment has been purely algebraic. Now in the case when R = B and J is any

operator ideal, we have the following consequence of the Open Mapping Theorem:

Proposition A.0.15. Let a1, a2 ∈ B/J be commuting invertible elements. Suppose Ai ∈ B are lifts

of ai such that the operator −A2

A1

: H → H2

has closed range. Then there exist lifts Bi of ai such that

A1B2 = A2B1.

Let a1, a2 ∈ B/J be commuting invertible elements. We conclude with some partial positive

results and negative results to the following question: Given lifts Ai ∈ B of ai, when do there exist

lifts Bi such that A1A2 = B2B1?

Proposition A.0.16. Let R be a unital ring and let J be a two-sided ideal. Let a1, a2 ∈ R/J be

commuting elements with lifts A1, A2 ∈ R. If either

(1) a1 has a left invertible lift, or

(2) a2 has a right invertible lift,

then there exist lifts Bi ∈ R of ai such that A1A2 = B2B1.

Proof. In the case of (1), let F1 ∈ J be such that A1 + F1 has a left inverse (A1 + F1)L. Then

A1A2 =(A1A2(A1 + F1)L

)(A1 + F1)

Page 99: Determinants in K-theory and operator algebras

92

so we may take B1 = A1 + F1 and B2 = A1A2(A1 + F1)L. The proof the second claim follows

similarly.

Corollary A.0.17. Let J ∈ B be a two-sided ideal. Let a1, a2 ∈ B/J be commuting elements with

lifts A1, A2 ∈ B. If either

(1) A1 is Fredholm with indA1 ≤ 0, or

(2) A2 is Fredholm with indA2 ≥ 0,

then there exist lifts Bi ∈ B of ai such that A1A2 = B2B1.

Proof. Every two-sided ideal J contains the finite rank operators. Moreover, a Fredholm operator

with non-positive (resp. non-negative) index can be perturbed by a finite rank operator to an

injective (resp. surjective) operator. The Open Mapping Theorem then guarantees the existence of

a left (resp. right) inverse.

Example A.0.18. The above conditions are sufficient but not necessary. Indeed, let H = l2(Z+),

let J be any two-sided ideal, and let S be the unilateral shift with index −1. Let A1 = SS∗S∗,

A2 = SSS∗, B1 = S∗, and B2 = S. Then indA1 > 0, indA2 < 0, and

A1A2 = SS∗S∗SSS∗

= SS∗

= B2B1

Example A.0.19. This example shows that the conclusion of the above corollary is not true in

general. Let A1 be a bounded surjective operator with nontrivial kernel, for example a unilateral

shift, and let A2 be a bounded right inverse. Suppose there exist perturbations Bi of Ai modulo any

ideal such that

A1A2 = B2B1.

Page 100: Determinants in K-theory and operator algebras

93

Since A1A2 = I, this implies that B1 is injective, and hence has non-positive index. This contradicts

the fact that A1, and hence also B1, have positive index. Thus the conclusion of the corollary fails

to hold in this case.

We do not yet have a complete characterization of pairs (A1, A2) that satisfy the conclusion

of the above corollary, but at least we have the following:

Corollary A.0.20. Let J ∈ B be a two-sided ideal. Let a1, a2 ∈ B/J be commuting invertible

elements with lifts A1, A2 ∈ B. Then there exist lifts Bi ∈ B of ai such that either

A1A2 = B2B1 or A2A1 = B1B2.