2000 Mathematics Subject Classification. Primary: 18G, 19K, 46H, 46L80, 46M, 58B34.
Algebraic
-theory and locally convex algebras
Joachim Cuntz
Andreas Thom
Joachim Cuntz, Mathematisches Institut, Einsteinstr.62, 48149 Munster, Germany E-mail address : cuntz@math.uni-muenster.de URL: http://www.math.uni-muenster.de/u/cuntz/cuntz Andreas Thom, SFB 478, Hittorfstr.27, 48149 Munster, Germany E-mail address : thoman@math.uni-muenster.de URL: http://www.math.uni-muenster.de/u/thoman
Contents
1 Introduction
The category of locally convex algebras is very vast. Besides natural Fréchet-algebras arising from differential calculus on manifolds it also contains algebras defined purely algebraically (any complex algebra with a countable basis is a locally convex algebra in a natural way with the fine topology). The category is therefore very flexible and allows constructions mixing analytic and algebraic structures. It seems to cover nearly all examples of (usually noncommutative) algebras associated with differential geometric objects, such as algebras of differential or pseudodifferential operators, algebras of differential forms, deformation algebras etc., etc. In connection with index theory and noncommutative geometry it therefore seems highly desirable to have at our disposal tools such as a computable version of topological
-theory preferably in bivariant form as it had been developed for
-algebras by Kasparov and others and associated invariants, also in connection with the various classes of “infinitesimals” used by Connes [4] .
In[9] , the first named author had developed such a theory for the smaller category consisting of all locally convex algebras that can be topologized by a family of submultiplicative seminorms (i.e. the category of projective limits of Banach algebras).
This theory has all the desirable properties and covers already many of the important examples. However, there are still very natural examples of locally convex algebras that do not fit into this category. Particularly relevant examples are given by algebras of differential operators and in particular, in the simplest case, by the so called Weyl algebra.
In [7] a bivariant
-theory
on the category of all locally convex algebras was then indeed developed. This bivariant homology theory again has very good formal properties, in particular the usual properties of (differentiable) homotopy invariance, long exact sequences associated with extensions and stability under tensoring by the algebra
of rapidly decreasing matrices. Therefore the
-invariants can be computed for many interesting examples of locally convex algebras as modules over the graded coefficient ring
. However, the determination of
and more generally, a
-theoretic description of
remained open in [7] (it was shown though that
admits a non-trivial unital homomorphism into
).
In the present paper we show that the problem of determining the coefficient ring can be overcome simply by stabilizing by the Schatten ideal
rather than by
.
In fact for the
-stabilized theory
we show that
for
and thus, in particular, that
, with a generator
in degree
(see Corollary 6.2.3 ). Stabilization by
is a very natural operation in Connes' setting for noncommutative geometry. Similar results hold for stabilization with more general symmetrically normed multiplicative Banach ideals (see definition 2.2.1 ) including the dual
of the Macaev ideal.
The key to our result is a locally convex version of an old homotopy invariance theorem proved by Higson in [13] on the basis of previous work by Kasparov and one of us. One ingredient in our proof of this result is the development of a setup in which the technology of abstract Kasparov modules can be applied algebraically (and thus in particular to locally convex algebras). Technically, we prove smooth homotopy-invariance of
-stable and split-exact functors (see definition 3.1.1 ), i.e. in particular of negative algebraic
-theory, for “weakly
-stable” locally convex algebras,
being a harmonic Banach ideal (see definition 2.2.2 ). All algebras of the form
, where
is an arbitrary locally convex algebra, are weakly
-stable. The result implies that, for such an algebra
, the canonical evaluation maps from the algebra
of smooth
-valued functions on
induce the same map
for each non-positive
. In particular, since
, we conclude that
is
-regular. We mention that the original version of Higson's theorem has been used in the category of
-algebras in a similar spirit by Suslin-Wodzicki in their proof of the Karoubi conjecture [17] , see also [16] for a nice exposition of this result and of related ideas.
Our conclusions are not obvious, even in the case
, since we do not make any assumption concerning openness of the group of invertibles, stability under functional calculus etc.. Indeed, as it turns out, it is wrong for
, contradicting a claim in [14] concerning
. We include a computation of
for any harmonic Banach ideal.
For higher algebraic
-theory, smooth homotopy invariance fails in general for weakly
-stable algebras. This fact is, from our point of view, due to the lack of excision. Cortin͂as (see [5] ) has given a long exact sequence relating the algebraic and topological
-theory of such algebras.
Our results imply the existence of a homomorphism of graded rings
. It seems that a proof of the assertion that this homomorphism is an isomorphism could be derived from the results claimed in [18] . However, we were not able to verify the arguments in [18] .
There is independent partially published work by Wodzicki. There are hints to a deep relationship between the results announced in [21] , see also [11] pp.3, and implications of our results. It should be interesting to pursue this further.
The second author wants to thank G. Cortin͂as for fruitful discussions about the last section. We thank Chr. Valqui and G. Cortin͂as for helpful comments.
2
-stable bivariant
-theory
2.1 Definitions
By a locally convex algebra we mean an algebra over
equipped with a complete locally convex topology such that the multiplication
is (jointly) continuous.
This means that, for every continuous seminorm
on
, there is another continuous seminorm
such that
for all
. Equivalently, the multiplication map induces a continuous linear map
from the completed projective tensor product
. All homomorphisms between locally convex algebras will be assumed to be continuous.
Every Banach algebra or projective limit of Banach algebras obviously is a locally convex algebra. But so is every algebra over
with a countable basis if we equip it with the “fine” locally convex topology, see e.g. [10] . The fine topology on a complex vector space
is given by the family of all seminorms on
.
Let
be an interval in
. We denote by
the algebra of complex-valued
-functions
on
, all of whose derivatives vanish in
and in
(while
itself may take arbitrary values in
and
). Also the subalgebras
and
of
, which, by definition consist of functions
, that vanish in
, in
, or in
and
, respectively, will play an important role. The topology on these algebras is the usual Fréchet topology.
Given two complete locally convex spaces
and
, we denote by
their completed projective tensor product (see [20] , [10] ). We note that
is nuclear in the sense of Grothendieck [20] and that, for any complete locally convex space
, the space
is isomorphic to the space of
-functions on
with values in
, whose derivatives vanish in both endpoints, [20] , § 51.
Given a locally convex algebra
, we write
,
and
for the locally convex algebras
,
and
(their elements are
valued
-functions whose derivatives vanish at the endpoints). The algebra
is called the suspension of
and denoted by
.
We denote by
the algebra of
-matrices over
and abbreviate, as usual,
by
. Given continuous homomomorphisms
, we denote by
the diagonal sum of the
.
We also consider the algebra
It is a locally convex algebra with the fine topology (which is also the inductive limit topology in the representation as inductive limit of the algebras
). Note that, if
is a finitely generated algebra, every homomorphism from
to
factors through
for some
. Finally, given a locally convex algebra
, we denote by
its unitization (as a locally convex space
is the direct sum
).
2.2 Banach Ideals
Let
be an infinite dimensional separable Hilbert space. Denote by
the algebra of bounded operators on
. Given
, the Schatten ideal
is defined as
Equivalently, a compact operator
is in
if the sequence
of its singular values is in
.
is a Banach algebra with the norm
We denote by
the algebra of all compact operators on
.
Definition 2.2.1.
A symmetrically normed, multiplicative Banach ideal
assigns to each infinite-dimensional separable Hilbert space
a normed ideal
in the algebra
of bounded operators such that
-
the norm
on
is complete,
-
for all
and
,
-
there is a natural continuous homomorphism
, compatible with the homomorphism
, and
-
the assignment
is functorial under unitary transformations.
By abuse of notation, in the following we also denote by
the algebra
(which depends on
only up to unitary isomorphism).
We will abbreviate the term symmetrically normed, multiplicative Banach ideal to Banach ideal, since we are considering only such. The main examples we have in mind are of course the Schatten ideals
and the ideal
, but also the Macaev ideal and its dual, which we denote by
. The Banach ideal
is of major importance in Connes' non-commutative geometry. It consists of all compact operators with singular values
, satisfying
(see [4] , pp.439). Although our main result does not apply to
it still applies to
.
The only non-trivial Banach ideal which is closed in operator norm in
, is the ideal
. For our purposes, in particular for the proof of theorem 4.2.1 , we need to consider a Banach ideal as a closed ideal in some algebra related to
. This can be achieved using the following construction.
Given a Banach ideal
, we construct the algebra
. Topologically, this algebra is defined to be
with
-norm and the multiplication is such that as many summands as possible are put in
. Since
is symmetrically normed,
is clearly a Banach algebra. Note that
is a closed ideal in
and that there is a split extension (see definition 3.1.1 )
Definition 2.2.2.
A Banach ideal
is called harmonic if
contains a compact operator with singular values given by the harmonic series for one (and hence each) Hilbert space
. If the Hilbert
space has a standard basis
, we denote by
the operator which multiplies the
-th basis vector
by
.
is harmonic if and only if
.
Note that
for
and
are harmonic, whereas
is clearly not.
Let
be a harmonic Banach ideal. We have the following chain of inclusions of sub-algebras of
in which only the three right-most algebras are Banach ideals. Clearly, every smooth compact operator is trace class. The fact, that
is contained in any harmonic Banach ideal is due to the following chain of inequalities
This implies, given any
, that there is a bounded operator
such that
, for some partial isometry
, hence
.
2.3 Bivariant
-theory
The bivariant homology theory
associates with any pair
of locally convex algebras abelian groups
,
, see section 6.1 for a definition and [7] for complete proofs. We list some important properties of
.
Theorem 2.3.1.
(a) Every continuous homomorphism
determines an element
in
. Given two homomorphisms
and
, we have
.
(b) Every extension (see section 3.1 for a definition)
determines canonically an element
in
. The class of the cone extension
is the identity element in
.
If
| |
| |
is a morphism of extensions (a commutative diagram where the rows are extensions), then
.
(c)
satisfies Bott periodicity:
.
(d) Let
be any locally convex algebra. Every extension
induces exact sequences in
and
of the following form:
|
(1)
|
and
|
(2)
|
The vertical arrows in ( 1 ) and ( 2 ) are (up to a sign) given by right and left multiplication, respectively, by the class
. (e) For each locally convex algebra
, there is a multiplicative transformation
such that
, for any homomorphism
.
For the proof see [7] .
Definition 2.3.2.
Let
and
be locally convex algebras and
be a Banach ideal. We define
A priori, the definition depends on an additional variable, the Hilbert space. To see the independence, we only need to know that the action of the unitary group of the Hilbert space
on
is trivial after applying
. This is easy to see and a particular case of Lemma 3.2.3 .
Having the independence of the Hilbert space, we can use the natural map
and its obvious associativity property in combination with property 2.3.1 (e), to define an associative product
.
The properties listed in 2.3.1 extend to the bivariant homology theory
, since the projective tensor product is exact on extensions with continuous split.
The following theorem, which is of independent interest, will show that for the ideals
(
or
), the theories
are all naturally isomorphic.
Theorem 2.3.3.
Let
and
be locally convex algebras. We assume that there is a continuous homomorphism
and a continuous map
, such that
-
is the multiplication on
and
-
is the multiplication on
.
Under these conditions,
is invertible.
-
Proof.
Consider the locally convex vector space
. Let
be a smooth homeomorphism from
to itself with vanishing derivatives at the endpoints and satisfying
. Endowed with the multiplication
it is a locally convex algebra which we denote by
. There is a natural homomorphism
given by
Similarly, there is a natural map
given by the assignment
Consider the following diagram of extensions.
| |
| |
| |
By 2.3.1 , (b), the commutativity of the diagram implies that the class in
determined by the extension in the middle is left and right inverse to
. □
Remark 2.3.4.
The argument in the proof of 2.3.3 shows at the same time that
induces an isomorphism
for any half-exact (see 4.1.1 ) functor
on the category of locally convex algebras which is diffotopy invariant.
Corollary 2.3.5.
For
or
, the natural map
defines an isomorphism for all
and
.
-
Proof.
Note that the multiplication
factors continuously through
by Hölders inequality. By 2.3.3 the inclusion map
induces an isomorphism in
whenever
and, by iteration, whenever
. □
3
-stable and split-exact functors
3.1 Split exactness
We consider functors from the category of locally convex algebras with continuous homomorphisms, which we denote by
, to the category of abelian groups, which we denote by
. There is a natural forgetful functor from the category of locally convex algebras to the category of locally convex vector spaces. A sequence
of locally convex algebras is called an extension, if it is a split-extension in the category of locally convex vector spaces, i.e. the middle term is a topological direct sum of kernel and co-kernel. An extension is called split-extension if there is a continuous splitting
which at the same time is a homomorphism.
Definition 3.1.1.
Let
be a functor.
-
(a)
The functor
is called split-exact, if any split extension of locally convex algebras is mapped to a split extension of abelian groups, i.e. if, for every extension
of locally convex algebras with a homomorphism splitting
, the induced sequence
is exact (and then automatically also split).
-
(b)
The functor
is called
-stable, if for any locally convex algebra
, the natural inclusion map
which embeds
in the upper left corner induces an isomorphism under
.
The most important example of a split-exact and
-stable functor is of course the algebraic
-theory functor
, but there are other examples which are of independent interest. Cortin͂as has heavily used the diffotopy invariance result (Theorem 4.2.1 ), which we will derive for all weakly
-stable, split-exact and
-stable functors, for the functors
,
and
in order to obtain results about the structure of higher algebraic
-theory of weakly
-stable algebras, see [5] and the definitions in section 4 .
Split exact functors satisfy the following familiar property, the proof of which is trivial.
Proposition 3.1.2.
Let
be a split exact functor and let
be homomorphisms with
for all
. We have that the linear map
is a homomorphism and that
.
We will show that every split exact and
-stable functor admits a pairing with abstract Kasparov modules. Kasparov modules are a convenient way of encoding the extended functoriality of
-stable and split-exact functors. The underlying construction is the one of quasihomomorphisms, which were introduced in [8] . Next, we give a brief introduction to the basics of quasihomomorphisms.
3.2 Quasi-homomorphisms
Let
and
be two homomorphisms
between locally convex algebras.
Assume that
is a closed subalgebra of
such that
and
,
for all
. We call such a pair
a quasihomomorphism from
to
relative to
and denote it by
.
We will show that
induces a homomorphism
in the following way. Define
by
and denote by
the subalgebra of
generated by all elements
and by
. We obtain an extension with two splitting homomorphisms
and
:
where the map
by definition maps
to
and
to
. The map
is defined to be
(this uses split-exactness). Note that
is independent of
in the sense that we can enlarge
without changing
as long as
maintains the properties above.
Proposition 3.2.1.
The assignment
has the following properties:
-
(a)
-
(b)
If the linear map
is a homomorphism and satisfies
for all
, then
.
-
Proof.
(a) This is obvious from the definition. (b) This follows from proposition 3.1.2 and the fact that
. □
Definition 3.2.2.
We say that a functor
is
-
invariant under inner automorphisms, if
for any invertible element
in a unital algebra
,
-
invariant under idealizing automorphisms, if the following holds:
Whenever
is isomorphic to a subalgebra of a unital algebra
and
is an invertible element in
such that
, then
for the automorphism
.
Lemma 3.2.3.
-
(a)
Every split-exact functor that is invariant under inner automorphisms is also invariant under idealizing automorphisms.
-
(b)
Every
-stable functor is invariant under inner automorphisms.
-
(c)
Every split-exact and
-stable functor is invariant under idealizing automorphisms.
-
(d)
Assume that
is invariant under inner automorphisms and let
be a quasihomomorphism relative to
. If there is an invertible element
in
such that
and such that moreover
,
for all
, then
.
-
Proof.
(a) Let
be a locally convex algebra and
be an idealizing element of
. Let
be the sub-algebra of
which is generated by
and
(we view
as a locally convex algebra with the fine topology). There is a natural split extension
The inner automorphism determined by
induces the identity on
and thus on
.
(b) Let
induce an inner automorphism of the algebra
. Consider the inner automorphism of
induced by
. Since, by an easy argument, the inclusions of
into the upper left corner and the inclusion into the lower right corner of
determine the same isomorphism under
, this implies that
is invariant under idealizing automorphisms.
(c) Combine (b) and (a).
(d) Note that
defines an idealizing automorphism of the algebra
that appears in the split-extension defining the map
. By (c), we know that
. Hence, we get
. □
3.3 Kasparov modules
Definition 3.3.1.
Let
and
be locally convex algebras. Assume that
is unital and contains
as a closed ideal. An abstract Kasparov
-module relative to
is a triple
where
-
is a continuous homomorphism from
into
.
-
is an invertible element and
is an idempotent element in
such that the following commutators are in
for all
:
Remark 3.3.2.
(a) The condition that
is a closed ideal in
could be weakened to demanding that the inclusion
and the homomorphisms
and,
which are given by mapping
to
and multiplying in
are continuous for the projective topology on the tensor product.
Furthermore, just as for quasihomomorphisms, the morphisms that we construct from a Kasparov module will be independent of the algebra
in the sense that
can be enlarged as long as the conditions on the commutators and the ideal are maintained.
(b) In Definition 3.3.1 one could replace the condition that
by the condition that
and
are in
for all
.
In the sequel, given an idempotent
in a unital algebra, we will denote by
the idempotent
. Given an abstract Kasparov module
, we define invertible matrices
and
in
as follows
and
We obtain in this way a quasihomomorphism
.
Definition 3.3.3.
Let
be a functor that is split exact and
-stable and let
be an abstract
-Kasparov module. We denote by
the map
associated with the quasihomomorphism
and the natural inclusion
.
Lemma 3.3.4.
Let
be split exact and
-stable. We consider abstract
-Kasparov modules.
-
(a)
whenever all terms are defined.
-
(b)
If
, then
.
-
(c)
If
commutes with all
,
and if
for all
, then
.
-
(d)
If
for all
, then
.
-
(e)
If
, then
.
-
Proof.
(a) We have
. This implies
| |
| |
| |
| |
| |
The last equality follows, since
is an idealizing automorphism of
and we have shown the invariance under such in Lemma 3.2.3 (c).
(b) One easily checks that
. The assertion then follows from (a) combined with (d).
(c) A computation using the definition of a Kasparov module and
-stability shows that
.
This observation implies that
(d) The element
satisfies the hypotheses of lemma 3.2.3 (d) with respect to
.
(e) We have the following chain of congruences mod the ideal
:
(
denotes congruence mod
).
The explicit form of the matrix
is as follows:
and similarly for
. Therefore the matrices
and
are in
. These two matrices produce an algebraic equivalence in
between the idempotents
and
. Using the standard technique (e.g. [2] prop. 4.3.1) there exists an invertible element
in
such that
and
Moreover,
can be chosen such that
. This last observation, together with
-stability, implies the third equality in the following computation.
| |
| |
| |
| |
| |
The last equality is due to the observation that
is an idealizing automorphism of
. This proves the assertion
using (a) and (d). □
3.4 Diffotopy invariance via Kasparov modules
In this section we construct explicit Kasparov modules which induce the evaluation maps, stabilized by the smooth compact operators. They are used in the proof of the diffotopy invariance theorem (see Theorem 4.2.1 ).
We let
act in the usual way on
and denote by
the Hardy projection from
onto the subspace
.
We denote by
the subalgebra of
generated algebraically by
together with
and all smooth compact operators. There is a natural isomorphism of
with the Schwartz space of rapidly decreasing
-sequences
defined by the action on the basis vector
. A vector is called smooth if it belongs to the image of this embedding.
We list some elementary properties concerning the action of the smooth algebras that we consider on
.
Lemma 3.4.1.
-
(a)
There is an isomorphism of topological vector spaces
extending the map
. In particular, any finite-rank operator built out of smooth vectors is a smooth compact operator.
-
(b)
The action of
on the Hilbert space
is induced by
where
denotes the canonical bilinear form on
. In particular the image
of any
is contained in
. Any eigenvector for an eigenvalue
of an element
is smooth.
-
(c)
The algebra
(resp. its unitization
) is a subalgebra of the
-algebra
of compact operators (resp. of its unitization
) which is closed under functional calculus by holomorphic functions for arbitrary elements (in particular the spectrum of an element in
is the same as in
), and under functional calculus by
-functions for self-adjoint and normal elements.
-
(d)
The algebra
preserves the space of smooth vectors.
-
Proof.
The only non-trivial assertion is (c). Let
be the unbounded operator in
defined by
on the domain
consisting of all (finite) linear combinations of vectors
in the standard orthonormal basis for
. It is trivial to check that
can be identified with the set of operators
in
for which
extends from
to a bounded operator on
for all
.
Define norms
on
by
where
denotes the operator norm. Then, for
we have:
Thus, the family
defines a differential seminorm in the sense of [3] . The topology on
is described by the family of sub-multiplicative norms
defined by
The results in [3] then immediately imply that
is a subalgebra of the algebra
of compact operators on
which is closed under functional calculus by holomorphic functions for arbitrary elements (see [3] , 3.12) and by functional calculus by
-functions for self-adjoint or normal elements (see [3] , 6.4). □
Linearly,
splits into
via the linear isomorphism
We topologize
by norms
, where
is of the type described in Lemma 3.4.1 , (c) and
is the usual sub-multiplicative norm on
defined by
An easy computation shows that these norms are sub-multiplicative. This shows that
is a locally convex algebra and contains the algebra
of smooth compact operators as a closed ideal.
We identify
with the algebra of smooth periodic functions on the interval
and divide this interval into four succeeding intervals
of length
.
We define a homomorphism
in the following way:
| |
| |
where we identify the interval
with each of the intervals
and where we put
, for
.
Let now
,
be a smooth family of unitary elements in
such that each
has winding number 1 and such that
outside a closed interval
contained in the interior of
and
outside an interval
contained in the interior of
(there is an obvious explicit family with these properties).
Since
and
for each
and
,
defines an abstract Kasparov module for each
.
Lemma 3.4.2.
Let
be an
-stable and split exact functor. Then
where
are the evaluation maps and
is the natural inclusion.
-
Proof.
Let
be a function in
such that
on
,
outside
and
and put
We consider
as an idempotent in
. It commutes exactly with
and we have
and
We have, using the last identity and Lemma 3.3.4 (e), the following equalities:
| |
| |
| |
| |
An analogous argument shows that
.
Now, note that
is connected to the standard unitary
by a continuous path
in
with
and
. Since
commutes with all elements in
modulo compact operators we see that
describes a continuous path of operators, which are Fredholm on
. The operator
is clearly of index
, then so is
. Applying functional calculus by the holomorphic function, which is equal to
in a neighborhood of 1 and equal to 0 in a neighborhood of 0, to
, there is a positive element
in the algebra generated by
and
, which differs from
only by an element in
such that
is a partial isometry still of index
(in fact with the same kernel and cokernel as
). Choose a partial isometry
in
with support
and range contained in
. Since the range and support of
consist of smooth vectors,
is in
. Thus
is an isometry of index -1 on
which differs from
only by an element in
.
In the same way one gets a coisometry
of index 1 on
and equal to
modulo
. Adding a partial isometry of rank 1 in
, that connects
and
, to
one obtains a unitary
in
that differs from
only by an element in
such that
.
Using Lemma 3.3.4 (e) and the explicit form of the quasihomomorphism defined by the abstract Kasparov module
, we have
However, by the choice of
, we have
, so that
by 3.2.1 (b). The argument for
is exactly symmetric. □
4 Diffotopy invariance theorem
4.1 Half-exact functors and diffotopy invariance
We continue with some definitions.
Definition 4.1.1.
Let
be a functor from the category of locally convex algebras to the category of abelian groups. Let
be a Banach ideal. We say that
-
is diffotopy invariant, if the maps
induced by the different evaluation maps for
are all the same (it is easy to see that this is the case if and only if the map induced by evaluation at
is an isomorphism).
-
is half-exact, if, for every extension
of locally convex algebras, the induced short sequence
is exact.
-
is
-stable, if the map
induced by the natural inclusion
, defined by any minimal idempotent in
, is an isomorphism for each locally convex algebra
(note that any Banach ideal contains minimal idempotents).
-
is weakly
-stable, if there is a natural map
, such that the composition
with the map induced by the inclusion map
defined by a minimal idempotent in
is the identity for each locally convex algebra
.
Remark 4.1.2.
We are using weak
-stability only for Banach ideals
, but note that weak stability is hereditary for certain subalgebras of
.
In particular, if
contains all smooth compact operators (for instance if
is harmonic, see end of section 2.2 ), then any weakly
-stable functor is also weakly
-stable.
Definition 4.1.3.
Two homomorphisms
between locally convex algebras are called diffotopic if there is a homomorphism
such that
Remark 4.1.4.
In [
9]
and elsewhere it was stated that the condition defining diffotopy in 4.1.3 was equivalent to the fact that there exists a family of homomorphisms
such that, for each
in
, the function
is in
and such that
,
.
The equivalence between the two conditions holds only if
and
are Fréchet, i.e. metrizable. A counterexample in the general case, due to Frerick and Shkarin, was communicated to us by L.Frerick [
12]
. Since only the condition given in the definition above was used in all the proofs, this does not affect any of the results in [
9]
or [
10]
.
It is easy to check that diffotopy is an equivalence relation. If
and
are diffotopic and
is a diffotopy invariant functor, then obviously
.
The next lemma illustrates the important consequences of the combination of the properties of diffotopy invariance and half-exactness. Note that the algebraic
-theory functor
is half-exact, in fact, it satisfies excision (see [1] ). Using the formally similar property of split-exactness, we are going to conclude that the
-stabilized version of the algebraic
-theory functor is diffotopy invariant. However, split-exactness does not imply half-exactness in general, nor does the reverse implication hold. For some of the notation, the reader is referred to section 6.1 .
Lemma 4.1.5.
Let
be a functor from the category of locally convex algebras to the category of abelian groups which is diffotopy invariant and half-exact. Then
-
(a)
has long exact sequences, i.e. for each extension
of locally convex algebras there is a long exact sequence (infinite to the left) of the form
-
(b)
There is a natural isomorphism
-
Proof.
(a) is well known and follows from a standard argument using mapping cones.
(b) Apply the long exact sequence from (a) to the extension
and use the fact that
by diffotopy invariance of
and the fact that
is smoothly contractible. □
4.2 Main theorem
In this section we state and prove the diffotopy invariance theorem.
Theorem 4.2.1.
Every functor from the category of locally convex algebras to the category of abelian groups which is split exact,
-stable and weakly
-stable for some harmonic Banach ideal, is diffotopy invariant.
-
Proof.
Given a weakly
-stable,
-stable and split-exact functor
, these properties are inherited by the functor
which assigns
, for any given locally convex algebra
. This implies that w.l.o.g. it is enough to show that the two evaluation maps from
to
are equal.
Using the result of Lemma 3.4.2 and the injectivity of
, which followed from weak
-stability and remark 4.1.2 , it suffices to show that we have an equality of
-Kasparov modules
Using the weak
-stability again we conclude that it also suffices to show that the following
-Kasparov modules are equal:
Note that
for some self-adjoint element
. Denote by
the unitary in
given by
where
is as in 3.4 .
We have to show that
. The element
is the product of elements
and
Here, the term with
appears precisely
times. The elements
lie in
since they are exponentials of elements
with
.
This is the only place, where we use the assumption that
is harmonic in an essential way. Note that the elements
commute with
and with
modulo
and hence define abstract Kasparov
-modules relative to
.
Note further that
with
and
Clearly,
for some permutation matrix
.
We see that
so that abstract Kasparov
-modules relative to
are defined, since
is a closed ideal in
. (For a definition of
see section 2.2 .) By lemma 3.3.4 (a) wee see that
.
A similar reasoning applies to
. This finishes the proof. □
5
of a stable algebra.
5.1 Stabilized functors
In this section we consider split-exact and
-stable functors
, defined on the category of locally convex algebras.
Definition 5.1.1.
Let
be a locally convex algebra and let
be a
-stable and split-exact functor. Let
be a Banach ideal. The algebra
is called weakly
-stable with respect to the functor
if the functor
is weakly
-stable (see definition 4.1.1 ).
Proposition 5.1.2 shows that
is weakly
-stable, so that there is always one obvious weakly
-stable algebra. Moreover, if
is weakly
-stable and
is any locally convex algebra, the
is weakly
-stable.
Let now
be a harmonic Banach ideal and
be a weakly
-stable algebra with respect to the functor
. In this section we show, as a corollary of theorem 4.2.1 , that the associated
-stabilized functor
satisfies diffotopy invariance. See the remark after definition 3.1.1 for important examples of split-exact and
-stable functors to which such a result could be applied.
Proposition 5.1.2.
The functor
which assigns
is weakly
-stable in the sense of definition 4.1.1 , i.e. the algebra
is weakly
-stable with respect to any
-stable and split exact functor.
-
Proof.
The natural map
induces a natural map
for every locally convex algebra
. We want to show that
for the natural map
induced by the inclusion
, i.e.
.
There is an isometry
in
such that
for
. Choose a second isometry
in
such that
.
Denote by
the algebra generated algebraically by
and
in
and by
the algebra generated by
together with
and
inside
. We have a split extension
It is easy to see that there is a unitary
in
and hence in
such that
for
in the subalgebra
of
. Thus, by
stability of
it is clear that
on
. On the other hand, from the split extension above, we see that
so that the restriction of
induces the identity on
also. □
Corollary 5.1.3.
Let
be an
-stable and split exact functor and let
be a weakly
-stable algebra with respect to the functor
, for a fixed harmonic Banach ideal
. Under these circumstances, the functor
is diffotopy invariant.
-
Proof.
This follows now from 4.2.1 . □
5.2 Algebraic
-theory
We continue by applying Corollary 5.1.3 to the algebraic
-theory functor
and identify algebraic
-theory of stable algebras with a suitable group of homotopy classes of maps.
Consider the algebra
and denote by
the two generators
. The argument in [8] , 3.1, shows that
and that
for each
-stable and diffotopy invariant functor
. With the classical description of
in [1] , the generator of
is given by the difference of equivalence classes of the idempotent elements
and
in
:
and,
Note that
and that therefore
.
Proposition 5.2.1.
For each locally convex algebra
, one has natural isomorphisms
-
Proof.
First of all, note that the righthand side is an abelian semigroup by block sum. The second isomorphism follows from the fact that
is finitely generated and from properties of the fine topology on
.
We construct a map
It sends a diffotopy class
to the difference of equivalence classes of projections
in
. The difference lies in the direct summand
.
The map is well-defined by the preceding corollary.
This map is surjective since all generators of
and their negatives are hit and since the map is clearly a map of abelian semigroups. In order to prove injectivity we use Lemma
in [9] which we reprove for sake of completeness.
Lemma 5.2.2.
Let
be the restriction of the homomorphism
, which sends
to
and
to
. Then
is diffotopic to the inclusion map
.
-
Proof.
Let
be the restriction of the homomorphism
which is defined by
| |
| |
where
For each
the difference
lies in the ideal
. Therefore
defines a diffotopy, which connects
with
. □
To prove injectivity we now use Lemma 5.2.2 . Assume that
are homomorphisms, such that
in
(where
denotes the induced map
).
This means that the there exists a projector
and an invertible element
such that
. This element
can even be chosen to be connected to
by a differentiable family
, such that
for all
.
Consider the homomorphisms
defined as restrictions of the maps from
that map
to
and
, respectively, and
to
. Note that
.
According to Lemma 5.2.2 ,
is diffotopic to
and similarly
is diffotopic to
. On the other hand, the family
,
, of homomorphisms
, obtained as restrictions of the maps from
, which map
to
and
to
, defines a diffotopy connecting
to
. □
6 Determination of
6.1 Bivariant
-theory revisited
We now finally have to use the explicit definition of
. To this end we recall some constructions and notation from [10] .
Let
be a complete locally convex space. Consider the algebraic tensor algebra
with the usual product given by concatenation of tensors. There is a canonical linear map
mapping
into the first direct summand. We equip
with the locally convex topology given by the family of all seminorms of the form
, where
is any homomorphism from
into a locally convex algebra
such that
is continuous on
, and
is a continuous seminorm on
. We further denote by
the completion of
with respect to this locally convex structure.
For any locally convex algebra
we have the natural extension
|
(3)
|
Here
maps a tensor
to
and
is defined as Ker
. This extension is (uni)versal in the sense that, given any extension
of a locally convex algebra
, admitting a continuous linear splitting, and any continuous homomorphism
, there is a morphism of extensions
|
(4)
|
The map
is obtained by choosing a continuous linear splitting
in the given extension and mapping
to
, where
. Then
is the restriction of
.
Choosing
in place of the second extension
, we see that
is a functor, i.e. any homomorphism
induces a homomorphism
.
Using the universal property of
, one can associate a classifying map with any linearly split extension of locally convex algebras of the form
We consider such an extension as a complex, denoting the arrows (boundary maps) by
and we say that it is linearly split if there is a continuous linear map
of degree -1 such that
. Every such splitting
induces a commutative diagram of the form
| |
| |
The leftmost vertical arrow in this diagram is the classifying map for this
-step extension. It depends on
only up to diffotopy.
Recall then from [10] the following definitions.
Definition 6.1.1.
Let
and
be locally convex algebras. For any continuous homomorphism
, we denote by
the equivalence class of
with respect to diffotopy and we set
Given
we set
where the inductive limit is with respect to the natural maps
mapping the diffotopy class of
to the diffotopy class of
, where
is defined by the commutative diagram
| |
| |
where
.
Thus, by definition of
, we have
In the sequel we will also use stabilization by the finite matrix algebras
.
Proposition 6.1.2.
The natural maps, induced by the inclusion
,
define an isomorphism
of abelian groups:
-
Proof.
Let
be an isometry in
and
the inclusion map (into the upper left corner). A standard argument shows that
is diffotopic to
.
Denote by
the natural tensor product of operators and by
the natural inclusion. We claim that the inverse
to the map
above is induced by
The identity
follows from the fact that the composition
is of the form
as above.
To show that
, consider the inclusions
into the left, resp. right, factor using the standard rank 1 projector onto the first basis vector in
. By [10] , 2.2.1,
and
are diffotopic and in fact both diffotopic to the natural isomorphism
.
We want to show that the composition of the following maps
is diffotopic to the identity. This composition can also be factored as follows
Replacing in this composition
by the diffotopic map
we again obtain a map of
of the form
which is diffotopic to id (using the fact that
is diffotopic to the natural isomorphism
). □
6.2 Main theorem
Our main result in this section is the following computation. Its proof requires some preparation and is given in section 6.3 .
Theorem 6.2.1.
For every locally convex algebra
and for every harmonic Banach ideal
one has
.
Remark 6.2.2.
In particular this shows that
does not depend on
for
(see also [
14]
, 4.1).
Corollary 6.2.3.
Let
be a harmonic Banach ideal. The coefficient ring
is isomorphic to
.
-
Proof.
By properties of
we are reduced to a computation of
and
. By Theorem 6.2.1 these groups are isomorphic to the algebraic
-groups
and
. Both algebras appearing are smooth sub-algebras of
-algebras whose
-theory is well-known, i.e.
and
This finishes the proof of the corollary. □
Corollary 6.2.4.
Let
be the Weyl algebra, i.e the unital algebra with two generators
and
satisfying the relation
(with the fine topology). Then for every harmonic Banach ideal
we have
and
.
-
Proof.
This follows from the result in [10] , 12.4 in combination with 6.2.3 above. □
Similarly, the
-stable
-theory of many other algebras can now be computed as an abelian group rather than just as a module over the coefficient ring.
In order to organize the notation in the following computations, we introduce a category
. The objects are locally convex algebras and the morphisms between two locally convex algebras are given by
Given two diffotopy classes of continuous homomorphisms
and
, their composition in
is defined to be the diffotopy class of
It is clear that composition is associative at the level of diffotopy classes. Moreover, note that endo-functors like
and
descend to endo-functors on
, since, for example, there is a natural map
.
In particular, using the new notation, we have
by proposition 6.1.2 and
by proposition 5.2.1 .
Proposition 6.2.5.
For all locally convex algebras
and
one has
This identity was noted in [10] , remark 8.4. In order to give an explicit proof, we have to introduce some notation. Throughout, we are working in the category
.
Denote by
the classifying map of the cone extension of
. Denote by
the classifying map of the Toeplitz extension tensored with
. We define
. (The map
has the important interpretation as the classifying map for the
-step extension given by the Yoneda product of the Toeplitz extension and the cone extension.) Furthermore, note that
by Korollar 3.1.1 in [9] , or by Lemma 4.6 in [10] , for all algebras
. For
, we also have the identity
We define inductively
by setting
and
by setting
. We also define
by setting
. Note that, in the definition of
, all other choices lead to the same definition of
, since
. Note that
, as the following induction argument shows.
| |
| |
| |
| |
| |
With this notation, the abelian group
is defined as the direct limit of a system of abelian groups
via a stabilization map
The
-th stabilization map is given by
which, again, follows from an easy induction argument.
The right hand side is defined as the direct limit of abelian semi-groups
via the stabilization map
-
Proof.
First of all, we define a map
by
This assignment induces a map of directed systems since it is compatible with the stabilization maps. Indeed,
is mapped to
and the following computation shows that this is the desired result.
| |
| |
| |
| |
| |
This implies that there is a well-defined map
Consider a class
which is represented by a homomorphism
. If the composition
is homotopic to zero for some
, then also
as element in
Note that, for any
and
, the functor
maps
isomorphically onto
. Furthermore,
and
are invertible for all
. This together implies that
as a class in
and hence that
is injective.
We still have to show that
is surjective. Consider a class
which is represented by a homomorphism
. The element
gives rise to a class in
. The natural map
which is induced by the assignment
is an isomorphism. I.e. there is some
and
such that
We claim that
does the job. Indeed, it is mapped to
Stabilizing yields
| |
| |
This shows the surjectivity of
. It is another easy check to show, that the assignment is compatible with the several composition products. □
6.3 Proof of the main theorem
We now proceed by proving Theorem 6.2.1 . By proposition 5.2.1 it suffices to shows that there is a natural isomorphism
The existence of a natural map will become apparent in the sequel of the proof.
-
Proof.
There is a natural map
given as the classifying map of a
-step extension which comes as the Yoneda-product of the Toeplitz extension and a canonical extension
where
Denote by
the restriction of
. It follows from the definition of
that
.
We define natural maps
.
Since
is a half-exact and diffotopy invariant functor by Theorem 4.2.1 , we conclude by Lemma 4.1.5 that there are natural isomorphisms
. However, the righthand side of the last equation is isomorphic to
and contains a canonical generator. Denote the canonical generators of
by
. The following identities are immediate, once we have the alternative description of
, given in proposition 5.2.1 .
| |
It is clear from the identities above that the maps
and
, which are induced from
and
by precomposition, are well defined. Furthermore,
is equal to the identity. We now show that
is also equal to the identity.
| |
| |
| |
| |
□
7 Computation of
for harmonic Banach ideals
In this section we want to give a computation of the algebraic
-theory group
for a harmonic Banach ideal
(see definition 2.2.2 ). The result contradicts an old result in [14] prop. 4.1. The error in the proof of proposition 4.1 in [14] was brought to our attention by Valqui and Cortin͂as. This concrete computation fits nicely with the far more general structure theorem about higher algebraic
-theory of locally convex algebras stabilized by harmonic Banach ideals, which was obtained by Cortin͂as in [5] using our diffotopy invariance theorem for weakly
-stable,
-stable, split-exact functors.
Let
be a locally convex algebra. In this section
denotes the unitization by
rather than by
. A priori, this difference matters and one has to be careful not to use the complex unitization.
The abelian group
is defined as
. Denote by
the (algebraic) square of
, i.e. the image of the algebraic tensor product
under the multiplication map. We consider the map
. Since
is abelian, we have a naturally defined determinant map
We get induced maps
(we use here the identification of the additive group
with the multiplicative group
).
We want to show that
is an isomorphism. Note, that it is obviously surjective, since
for any
.
Theorem 7.1.
Let
be a harmonic Banach ideal. The natural determinant map yields an isomorphism
The preceding result implies that our diffotopy result does not extend to higher algebraic
-theory. The first topological
-theory of the Schatten ideals is well-known to be zero. The proof of the preceding theorem is given after stating and proving a lemma.
Note that, every class
is represented by an element
which maps to the identity under the canonical evaluation onto
. Using the next lemma, we are able to show injectivity of the determinant map.
Lemma 7.2.
-
(a)
Every invertible element in
which maps to
is a product of exponentials of elements in
.
An invertible element in
which is connected by a norm continuous path of elements, which are invertible in
, to
is a product of exponentials of elements in
.
-
(b)
Let
. The element
is invertible in
and
in
.
-
(c)
Let
. We have that
in
.
-
(d)
Let
. If
, then
in
.
-
Proof.
Since
is isomorphic to
and
we can restrict our reasoning, for a proof of (a),(b) and (c), to the case
.
(a) If
is close to one, then the logarithmic series converges to an element in
. An easy calculation shows that the logarithm lies in
if
(it can be written as a product of
by an element in
). The assertion follows by standard arguments using compactness and the fact that the group of invertible elements in
which map to
is connected.
(b) The proof of this lemma follows the idea in [14] , 4.1, using a scheme going back to [15] . As in the proof of the homotopy invariance theorem we consider certain invertible elements which we want to represent by commutators. Consider
as acting on
and choose an isometry between the second copy of
in this direct sum and
. Under this identification, we define
and
(again, with
-term appearing
times) can be considered as elements in
, since
is harmonic (compare to the proof of Theorem 4.2.1 ). Clearly
. In order to show that
, it suffices to show
in
. We want to show that
and
are products of commutators. This implies that
is also a product of commutators and hence finishes the proof.
We first concentrate on
. As before we construct matrices
and
which satisfy
. There is an explicit isomorphism
(preserving the ideal
) such that the matrix
is mapped to a matrix of the form
. The element
is in
(since the exponential series was already convergent in
) and therefore decomposes into
with
. To see this assume that
with
and
in
. Then there are
and
in
such that
. Identify then
with
using
for a suitable isometry
.
By the proof of Vaserstein's lemma (or by direct computation), we have that
where
denotes a multiplicative commutator. Thus,
is a product of two multiplicative commutators in
. A similar reasoning applies to
and matrices occurring in a similar decomposition of
. Thus
in
.
(c) Since the leading terms in the series expansion of
vanish, this element can be written as a sum of four terms where each term is a product of
respectively, by an element in
. Therefore
as well as its inverse lie in
. By (a)
is a product of exponentials of elements in
. Using (b), this implies the claim.
(d) A matrix with trace in
is a finite sum of
-
off-diagonal matrices with one entry in
,
-
matrices of the form
with
and
-
a matrix
with
.
The classes in
of their exponentials are zero, whence the claim by iterated application of (c). □
We now proceed with the proof of the theorem.
-
Proof.
Let
be an invertible element in
which maps to the identity in
and with determinant zero. By (a) of the preceding lemma it is of the form
and
. The matrix
has the same class in
and the trace of its logarithm is just a lift of its determinant
to
and hence in
. By (d) of the preceding lemma the class in
is zero. This shows injectivity of the determinant map. Surjectivity was obvious, hence the assertion. □
Remark 7.0.1.
It is of course always true that
maps surjectively onto
,
being a Banach ideal.
References
-
H.Bass, Algebraic
-theory, W.A. Benjamin, New York (1968).
-
B.Blackadar,
-theory for Operator Algebras, Springer-Verlag, Heidelberg / Berlin / New York / Tokyo (1986).
-
B.Blackadar and J. Cuntz, Differential Banach algebra norms and smooth subalgebras of
-algebras, J. Operator Theory 26 (1991), 255-282.
-
A.Connes, Non-commutative Geometry, Academic Press (1994).
-
G.Cortin͂as, Algebraic
-theory of locally convex algebras, in preparation
-
J.Cuntz, A class of
-algebras generated by isometries, Commun. Math. Phys. 57 (1977), 173-185.
-
J.Cuntz, K-theory and
-algebras, in ”Algebraic K-theory, Number theory, Geometry and Analysis”, Springer Lecture Notes in Math. 1046, 55-79.
-
J.Cuntz, A new look at KK-theory, K-theory 1 (1987), 31-52.
-
J.Cuntz, Bivariante
-theorie für lokalkonvexe Algebren und der bivariante Chern-Connes-Charakter, Docum. Math. J. DMV 2 (1997), 139-182, http://www.mathematik.uni-bielefeld.de/documenta
-
J.Cuntz, Bivariant
-theory and the Weyl algebra,
-theory (to appear).
-
K.Dykema, T.Figiel, G.Weiss and M.Wodzicki, Commutator structure of operator ideals, Adv. in Mathematics 185 (2004), 1-79.
-
L.Frerick and S.Shkarin, Completeness of certain function spaces, preprint.
-
N.Higson, Algebraic
-theory of stable
-algebras. Adv. in Math. 67 (1988), no. 1, 140 pp.
-
M.Karoubi, Homologie de groupes discrets associés à des algèbres d'opérateurs, J. Operator Theory 15 (1986), no. 1, 109–161.
-
C.Pearcy and D.Topping, On commutators in ideals of compact operators. Michigan Math. J. 18 (1971) 247–252
-
J.Rosenberg, Comparison between algebraic and topological
-theory for Banach algebras and
-algebras, preprint.
-
A.Suslin and M.Wodzicki, Excision in algebraic
-theory, Ann. of Math. (2) 136 (1992), no. 1, 51–122.
-
J.Tapia,
-théorie algébrique négative et
-théorie topologique de l'algèbre de Fréchet des opérateurs régularisants, Ann. Sci. Ecole Norm. Sup. (4) 30 (1997), no. 2, 241–277.
-
A.Thom, Connective
-theory and bivariant homology, 2003, preprint
-
F.Treves, Topological vector spaces, distributions and kernels, Academic Press, New York, London, 1967.
-
M.Wodzicki, Algebraic
-theory and functional analysis, in: Proceedings of the First European Congress of Mathematics, Paris, Vol. II (1992), 485-496, Progress in Mathematics vol. 120, Birkhäuser Verlag, Basel (1994).
Joachim Cuntz, Mathematisches Institut, Einsteinstr.62, 48149 Munster, Germany E-mail address : cuntz@math.uni-muenster.de URL: http://www.math.uni-muenster.de/u/cuntz/cuntz Andreas Thom, SFB 478, Hittorfstr.27, 48149 Munster, Germany E-mail address : thoman@math.uni-muenster.de URL: http://www.math.uni-muenster.de/u/thoman