Hilbert C*-module

{{Short description|Mathematical objects that generalise the notion of Hilbert spaces}}

Hilbert C*-modules are mathematical objects that generalise the notion of Hilbert spaces

(which are themselves generalisations of Euclidean space),

in that they endow a linear space with an "inner product" that takes values in a

C*-algebra.

They were first introduced in the work of Irving Kaplansky in 1953,

which developed the theory for commutative,

unital algebras

(though Kaplansky observed that the assumption of a unit element was not "vital").{{cite journal| last = Kaplansky| first = I.| authorlink = Irving Kaplansky |title = Modules over operator algebras| journal = American Journal of Mathematics| volume = 75| issue = 4| pages = 839–853| year = 1953| doi = 10.2307/2372552| jstor = 2372552}}

In the 1970s the theory was extended to non-commutative C*-algebras independently by William Lindall Paschke{{cite journal| last = Paschke| first = W. L.| title = Inner product modules over B*-algebras| journal = Transactions of the American Mathematical Society| volume = 182| pages = 443–468| year = 1973| doi = 10.2307/1996542| jstor = 1996542}}

and Marc Rieffel,

the latter in a paper that used Hilbert C*-modules to construct a theory of induced representations of C*-algebras.{{cite journal| last = Rieffel| first = M. A.| title = Induced representations of C*-algebras| journal = Advances in Mathematics| volume = 13| pages = 176–257| year = 1974| doi = 10.1016/0001-8708(74)90068-1| doi-access=free | issue = 2}}

Hilbert C*-modules are crucial to Kasparov's formulation of KK-theory,{{cite journal| last = Kasparov| first = G. G.| title = Hilbert C*-modules: Theorems of Stinespring and Voiculescu| journal = Journal of Operator Theory| volume = 4| pages = 133–150| publisher = Theta Foundation| year = 1980}}

and provide the right framework to extend the notion

of Morita equivalence to C*-algebras.{{cite journal| last = Rieffel| first = M. A.| title = Morita equivalence for operator algebras| journal = Proceedings of Symposia in Pure Mathematics| volume = 38| pages = 176–257| publisher = American Mathematical Society| year = 1982}}

They can be viewed as the generalization

of vector bundles to noncommutative C*-algebras and as such play an important role in noncommutative geometry,

notably in C*-algebraic quantum group theory,{{cite journal| last = Baaj| first = S.|author2=Skandalis, G.| title = Unitaires multiplicatifs et dualité pour les produits croisés de C*-algèbres| journal = Annales Scientifiques de l'École Normale Supérieure| volume = 26| issue = 4| pages = 425–488| year = 1993| doi = 10.24033/asens.1677| doi-access = free}}{{cite journal| last = Woronowicz|

first = S. L.| authorlink = S. L. Woronowicz| title = Unbounded elements affiliated with C*-algebras and non-compact quantum groups| journal = Communications in Mathematical Physics| volume = 136| pages = 399–432| year = 1991| doi = 10.1007/BF02100032|bibcode =

1991CMaPh.136..399W| issue = 2 | s2cid = 118184597}}

and groupoid C*-algebras.

Definitions

= Inner-product C*-modules =

Let A be a C*-algebra (not assumed to be commutative or unital), its involution denoted by {}^*. An inner-product A-module (or pre-Hilbert A-module) is a complex linear space E equipped with a compatible right A-module structure, together with a map

: \langle \, \cdot \, , \, \cdot \,\rangle_A : E \times E \rightarrow A

that satisfies the following properties:

  • For all x, y, z in E, and \alpha, \beta in \mathbb{C}:

:: \langle x ,y \alpha + z \beta \rangle_A = \langle x, y \rangle_A \alpha + \langle x, z \rangle_A \beta

:(i.e. the inner product is \mathbb{C}-linear in its second argument).

  • For all x, y in E, and a in A:

:: \langle x, y a \rangle_A = \langle x, y \rangle_A a

  • For all x, y in E:

:: \langle x, y \rangle_A = \langle y, x \rangle_A^*,

:from which it follows that the inner product is conjugate linear in its first argument (i.e. it is a sesquilinear form).

  • For all x in E:

:: \langle x, x \rangle_A \geq 0

:in the sense of being a positive element of A, and

:: \langle x, x \rangle_A = 0 \iff x = 0.

:(An element of a C*-algebra A is said to be positive if it is self-adjoint with non-negative spectrum.){{cite book| last = Arveson| first = William| authorlink = William Arveson|title = An Invitation to C*-Algebras| publisher = Springer-Verlag| year = 1976| page = 35}}In the case when A is non-unital, the spectrum of an element is calculated in the C*-algebra generated by adjoining a unit to A.

= Hilbert C*-modules =

An analogue to the Cauchy–Schwarz inequality holds for an inner-product A-module E:This result in fact holds for semi-inner-product A-modules, which may have non-zero elements A such that \langle x , x \rangle_A

= 0, as the proof does not rely on the nondegeneracy property.

:\langle x, y \rangle_A \langle y, x \rangle_A \leq \Vert \langle y, y \rangle_A \Vert \langle x, x \rangle_A

for x, y in E.

On the pre-Hilbert module E, define a norm by

:\Vert x \Vert = \Vert \langle x, x \rangle_A \Vert^\frac{1}{2}.

The norm-completion of E, still denoted by E, is said to be a Hilbert A-module or a Hilbert C*-module over the C*-algebra A.

The Cauchy–Schwarz inequality implies the inner product is jointly continuous in norm and can therefore be extended to the completion.

The action of A on E is continuous: for all x in E

:a_{\lambda} \rightarrow a \Rightarrow xa_{\lambda} \rightarrow xa.

Similarly, if (e_\lambda) is an approximate unit for A (a net of self-adjoint elements of A for which a e_\lambda and e_\lambda a tend to a for each a in A), then for x in E

: xe_\lambda \rightarrow x.

Whence it follows that EA is dense in E, and x 1_A = x when A is unital.

Let

: \langle E, E \rangle_A = \operatorname{span} \{ \langle x, y \rangle_A \mid x, y \in E \},

then the closure of \langle E, E \rangle_A is a two-sided ideal in A. Two-sided ideals are C*-subalgebras and therefore possess approximate units. One can verify that E \langle E, E \rangle_A is dense in E. In the case when \langle E , E \rangle_A is dense in A, E is said to be full. This does not generally hold.

Examples

= Hilbert spaces =

Since the complex numbers \mathbb{C} are a C*-algebra with an involution given by complex conjugation, a complex Hilbert space \mathcal{H} is a Hilbert \mathbb{C} -module under scalar multipliation by complex numbers and its inner product.

=Vector bundles=

If X is a locally compact Hausdorff space and E a vector bundle over X with projection \pi \colon E \to X a Hermitian metric g , then the space of continuous sections of E is a Hilbert C(X) -module. Given sections \sigma, \rho of E and f \in C(X) the right action is defined by

:: \sigma f (x) = \sigma(x) f(\pi(x)),

and the inner product is given by

:: \langle \sigma,\rho\rangle_{C(X)} (x):=g(\sigma(x),\rho(x)).

The converse holds as well: Every countably generated Hilbert C*-module over a commutative unital C*-algebra A = C(X) is isomorphic to the space of sections vanishing at infinity of a continuous field of Hilbert spaces over X . {{cn|date=May 2023}}

= C*-algebras =

Any C*-algebra A is a Hilbert A -module with the action given by right multiplication in A and the inner product \langle a , b \rangle = a^*b . By the C*-identity, the Hilbert module norm coincides with C*-norm on A .

The (algebraic) direct sum of n copies of A

: A^n = \bigoplus_{i=1}^n A

can be made into a Hilbert A -module by defining

:\langle (a_i), (b_i) \rangle_A = \sum_{i=1}^n a_i^* b_i.

If p is a projection in the C*-algebra M_n(A), then pA^n is also a Hilbert A-module with the same inner product as the direct sum.

= The standard Hilbert module =

One may also consider the following subspace of elements in the countable direct product of A

: \ell_2(A)= \mathcal{H}_A = \Big\{ (a_i) | \sum_{i=1}^{\infty} a_i^{*}a_i\text{ converges in }A \Big\}.

Endowed with the obvious inner product (analogous to that of A^n ), the resulting Hilbert A -module is called the standard Hilbert module over A .

The fact that there is a unique separable Hilbert space

has a generalization to Hilbert modules in the form of the

Kasparov stabilization theorem, which states

that if E is a countably generated Hilbert A-module, there is an isometric isomorphism E \oplus \ell^2(A) \cong \ell^2(A). {{cite journal| last = Kasparov| first = G. G.| title = Hilbert C*-modules: Theorems of Stinespring and Voiculescu| journal = Journal of Operator Theory| volume = 4| pages = 133–150| publisher = ThetaFoundation| year = 1980}}

Maps between Hilbert modules

Let E and F be two Hilbert modules over the same

C*-algebra A. These are then Banach spaces, so it is possible to

speak of the Banach space of bounded linear maps \mathcal{L}(E,F),

normed by the operator norm.

The adjointable and compact adjointable operators are subspaces of this Banach space

defined using the inner product structures on E and F.

In the special case where A is \mathbb{C} these reduce to

bounded and compact operators on Hilbert spaces respectively.

= Adjointable maps =

A map (not necessarily linear)

T \colon E \to F is defined to be adjointable if there

is another map T^* \colon F \to E, known as the adjoint

of T, such that for every

e \in E and f \in F,

: \langle f, Te \rangle = \langle T^* f, e \rangle.

Both T and T^* are then automatically linear

and also A-module maps. The

closed graph theorem can be used to show that they are also bounded.

Analogously to the adjoint of operators on Hilbert spaces, T^*

is unique (if it exists) and itself adjointable with adjoint T. If S \colon F \to G

is a second adjointable map, ST is adjointable with adjoint

S^* T^*.

The adjointable operators E \to F form a subspace \mathbb{B}(E,F)

of \mathcal{L}(E,F), which is complete in the operator norm.

In the case F = E, the space \mathbb{B}(E,E) of

adjointable operators from E to itself is denoted \mathbb{B}(E), and is a

C*-algebra.Wegge-Olsen 1993, pp. 240-241.

= Compact adjointable maps =

Given e \in E and f \in F, the map

| f \rangle \langle e | \colon E \to F is defined, analogously to the

rank one operators of Hilbert spaces, to be

:g \mapsto f \langle e, g \rangle.

This is adjointable with adjoint | e \rangle \langle f |.

The compact adjointable operators \mathbb{K}(E,F) are defined to be the closed span

of

:\{ | f \rangle \langle e | \mid e \in E, \; f \in F \}

in \mathbb{B}(E,F).

As with the bounded operators, \mathbb{K}(E,E) is denoted

\mathbb{K}(E). This is a

(closed, two-sided) ideal of

\mathbb{B}(E).Wegge-Olsen 1993, pp. 242-243.

C*-correspondences

If A and B are C*-algebras, an (A,B) C*-correspondence

is a Hilbert B-module equipped with a left action of A by

adjointable maps that is faithful. (NB: Some authors require the left action to be

non-degenerate instead.) These objects are used in the formulation of Morita equivalence

for C*-algebras, see applications in the construction of Toeplitz and Cuntz-Pimsner algebras,Brown, Ozawa 2008, section 4.6.

and can be employed to put the structure of a bicategory on the collection of C*-algebras.Buss, Meyer, Zhu, 2013, section 2.2.

= Tensor products and the bicategory of correspondences =

If E is an (A,B) and F a (B,C) correspondence,

the algebraic tensor product E \odot F of E and F

as vector spaces inherits left and right A- and C-module

structures respectively.

It can also be endowed with the C-valued sesquilinear form defined on

pure tensors by

: \langle e \odot f, e' \odot f' \rangle_C := \langle f, \langle e, e' \rangle_B f \rangle_C.

This is positive semidefinite, and the Hausdorff completion of E \odot F

in the resulting seminorm is denoted E \otimes_B F. The left- and right-actions of

A and C extend to make this an (A,C) correspondence.Brown, Ozawa 2008, pp. 138-139.

The collection of C*-algebras can then be endowed with

the structure of a bicategory, with C*-algebras as

objects, (A,B) correspondences as

arrows B \to A, and isomorphisms of correspondences (bijective module maps that preserve

inner products) as 2-arrows.Buss, Meyer, Zhu 2013, section 2.2.

= Toeplitz algebra of a correspondence =

Given a C*-algebra A, and an (A,A) correspondence E,

its Toeplitz algebra \mathcal{T}(E) is defined as the universal algebra

for Toeplitz representations (defined below).

The classical Toeplitz algebra can be recovered

as a special case, and the Cuntz-Pimsner algebras

are defined as particular quotients of Toeplitz algebras.Brown, Ozawa, 2008, section 4.6.

In particular, graph algebras , crossed products by \mathbb{Z} , and the

Cuntz algebras are all quotients of specific Toeplitz algebras.

== Toeplitz representations ==

A Toeplitz representationFowler, Raeburn, 1999, section 1. of E in a C*-algebra D

is a pair (S,\phi)

of a linear map S \colon E \to D and a homomorphism

\phi \colon A \to D such that

  • S is "isometric":

:S(e)^* S(f) = \phi(\langle e, f \rangle) for all e,f \in E,

  • S resembles a bimodule map:

:S(a e) = \phi(a) S(e) and S(ea) = S(e) \phi(a) for e \in E and a \in A.

== Toeplitz algebra ==

The Toeplitz algebra \mathcal{T}(E) is the universal Toeplitz representation.

That is, there is a Toeplitz representation (T, \iota) of E

in \mathcal{T}(E) such that if (S,\phi) is any Toeplitz representation

of E (in an arbitrary algebra D) there is a unique *-homomorphism

\Phi \colon \mathcal{T}(E) \to D such that S = \Phi \circ T

and \phi = \Phi \circ \iota.Fowler, Raeburn, 1999, Proposition 1.3.

== Examples ==

If A is taken to be the algebra of complex numbers, and E

the vector space \mathbb{C}^n, endowed with the natural

(\mathbb{C},\mathbb{C})-bimodule structure, the corresponding Toeplitz algebra

is the universal algebra generated by n isometries with mutually orthogonal

range projections.Brown, Ozawa, 2008, Example 4.6.10.

In particular, \mathcal{T}(\mathbb{C}) is the universal algebra generated by

a single isometry, which is the classical Toeplitz algebra.

See also

Notes

{{Reflist}}

References

  • {{cite book |last=Lance |first=E. Christopher|title=Hilbert C*-modules: A toolkit for operator algebraists |series=London Mathematical Society Lecture Note Series|year=1995 |publisher=Cambridge University Press |location=Cambridge, England}}
  • {{cite book| last = Wegge-Olsen | first = N. E.| title = K-Theory and C*-Algebras | publisher = Oxford University Press | year = 1993}}
  • {{cite book

| last1 = Brown

| first1 = Nathanial P.

| last2 = Ozawa

| first2 = Narutaka

| date = 2008

| title = C*-Algebras and Finite-Dimensional Approximations

| url = https://bookstore.ams.org/gsm-88

| publisher = American Mathematical Society

}}

  • {{cite journal

| last1 = Buss

| first1 = Alcides

| last2 = Meyer

| first2 = Ralf

| last3 = Zhu

| first3 = Chenchang

| date = 2013

| title = A higher category approach to twisted actions on c* -algebras

| url =

| journal = Proceedings of the Edinburgh Mathematical Society

| volume = 56

| issue = 2

| pages = 387–426

| doi = 10.1017/S0013091512000259

| arxiv = 0908.0455

}}

  • {{cite journal

| last1 = Fowler

| first1 = Neal J.

| last2 = Raeburn

| first2 = Iain

| date = 1999

| title = The Toeplitz algebra of a Hilbert bimodule

| url = https://www.jstor.org/stable/24900141

| journal = Indiana University Mathematics Journal

| volume = 48

| issue = 1

| pages = 155–181

| doi = 10.1512/iumj.1999.48.1639

| jstor = 24900141

| arxiv = math/9806093

}}