Hilbert C*-module

(Redirected from Hilbert module)

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").[1]

In the 1970s the theory was extended to non-commutative C*-algebras independently by William Lindall Paschke[2] and Marc Rieffel, the latter in a paper that used Hilbert C*-modules to construct a theory of induced representations of C*-algebras.[3]

Hilbert C*-modules are crucial to Kasparov's formulation of KK-theory,[4] and provide the right framework to extend the notion of Morita equivalence to C*-algebras.[5] 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,[6][7] and groupoid C*-algebras.


Definitions

edit

Inner-product C*-modules

edit

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

 

that satisfies the following properties:

  • For all  ,  ,   in  , and  ,   in  :
 
(i.e. the inner product is  -linear in its second argument).
  • For all  ,   in  , and   in  :
 
  • For all  ,   in  :
 
from which it follows that the inner product is conjugate linear in its first argument (i.e. it is a sesquilinear form).
  • For all   in  :
 
in the sense of being a positive element of A, and
 
(An element of a C*-algebra   is said to be positive if it is self-adjoint with non-negative spectrum.)[8][9]

Hilbert C*-modules

edit

An analogue to the Cauchy–Schwarz inequality holds for an inner-product  -module  :[10]

 

for  ,   in  .

On the pre-Hilbert module  , define a norm by

 

The norm-completion of  , still denoted by  , is said to be a Hilbert  -module or a Hilbert C*-module over the C*-algebra  . The Cauchy–Schwarz inequality implies the inner product is jointly continuous in norm and can therefore be extended to the completion.

The action of   on   is continuous: for all   in  

 

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

 

Whence it follows that   is dense in  , and   when   is unital.

Let

 

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

Examples

edit

Hilbert spaces

edit

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

Vector bundles

edit

If   is a locally compact Hausdorff space and   a vector bundle over   with projection   a Hermitian metric  , then the space of continuous sections of   is a Hilbert  -module. Given sections   of   and   the right action is defined by

 

and the inner product is given by

 

The converse holds as well: Every countably generated Hilbert C*-module over a commutative unital C*-algebra   is isomorphic to the space of sections vanishing at infinity of a continuous field of Hilbert spaces over  . [citation needed]

C*-algebras

edit

Any C*-algebra   is a Hilbert  -module with the action given by right multiplication in   and the inner product  . By the C*-identity, the Hilbert module norm coincides with C*-norm on  .

The (algebraic) direct sum of   copies of  

 

can be made into a Hilbert  -module by defining

 

If   is a projection in the C*-algebra  , then   is also a Hilbert  -module with the same inner product as the direct sum.

The standard Hilbert module

edit

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

 

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

The standard Hilbert module plays an important role in the proof of the Kasparov stabilization theorem which states that for any countably generated Hilbert  -module   there is an isometric isomorphism   [11]

Maps between Hilbert modules

edit

Let   and   be two Hilbert modules over the same C*-algebra  . These are then Banach spaces, so it is possible to speak of the Banach space of bounded linear maps  , normed by the operator norm.

The adjointable and compact adjointable operators are subspaces of this Banach space defined using the inner product structures on   and  .

In the special case where   is   these reduce to bounded and compact operators on Hilbert spaces respectively.

Adjointable maps

edit

A map (not necessarily linear)   is defined to be adjointable if there is another map  , known as the adjoint of  , such that for every   and  ,

 

Both   and   are then automatically linear and also  -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,   is unique (if it exists) and itself adjointable with adjoint  . If   is a second adjointable map,   is adjointable with adjoint  .

The adjointable operators   form a subspace   of  , which is complete in the operator norm.

In the case  , the space   of adjointable operators from   to itself is denoted  , and is a C*-algebra.[12]

Compact adjointable maps

edit

Given   and  , the map   is defined, analogously to the rank one operators of Hilbert spaces, to be

 

This is adjointable with adjoint  .

The compact adjointable operators   are defined to be the closed span of

 

in  .

As with the bounded operators,   is denoted  . This is a (closed, two-sided) ideal of  .[13]

C*-correspondences

edit

If   and   are C*-algebras, an   C*-correspondence is a Hilbert  -module equipped with a left action of   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,[14] and can be employed to put the structure of a bicategory on the collection of C*-algebras.[15]

Tensor products and the bicategory of correspondences

edit

If   is an   and   a   correspondence, the algebraic tensor product   of   and   as vector spaces inherits left and right  - and  -module structures respectively.

It can also be endowed with the  -valued sesquilinear form defined on pure tensors by

 

This is positive semidefinite, and the Hausdorff completion of   in the resulting seminorm is denoted  . The left- and right-actions of   and   extend to make this an   correspondence.[16]

The collection of C*-algebras can then be endowed with the structure of a bicategory, with C*-algebras as objects,   correspondences as arrows  , and isomorphisms of correspondences (bijective module maps that preserve inner products) as 2-arrows.[17]

Toeplitz algebra of a correspondence

edit

Given a C*-algebra  , and an   correspondence  , its Toeplitz algebra   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.[18]

In particular, graph algebras , crossed products by   , and the Cuntz algebras are all quotients of specific Toeplitz algebras.

Toeplitz representations

edit

A Toeplitz representation[19] of   in a C*-algebra   is a pair   of a linear map   and a homomorphism   such that

  •   is "isometric":
  for all  ,
  •   resembles a bimodule map:
  and   for   and  .

Toeplitz algebra

edit

The Toeplitz algebra   is the universal Toeplitz representation. That is, there is a Toeplitz representation   of   in   such that if   is any Toeplitz representation of   (in an arbitrary algebra  ) there is a unique *-homomorphism   such that   and  .[20]

Examples

edit

If   is taken to be the algebra of complex numbers, and   the vector space  , endowed with the natural  -bimodule structure, the corresponding Toeplitz algebra is the universal algebra generated by   isometries with mutually orthogonal range projections.[21]

In particular,   is the universal algebra generated by a single isometry, which is the classical Toeplitz algebra.

See also

edit

Notes

edit
  1. ^ Kaplansky, I. (1953). "Modules over operator algebras". American Journal of Mathematics. 75 (4): 839–853. doi:10.2307/2372552. JSTOR 2372552.
  2. ^ Paschke, W. L. (1973). "Inner product modules over B*-algebras". Transactions of the American Mathematical Society. 182: 443–468. doi:10.2307/1996542. JSTOR 1996542.
  3. ^ Rieffel, M. A. (1974). "Induced representations of C*-algebras". Advances in Mathematics. 13 (2): 176–257. doi:10.1016/0001-8708(74)90068-1.
  4. ^ Kasparov, G. G. (1980). "Hilbert C*-modules: Theorems of Stinespring and Voiculescu". Journal of Operator Theory. 4. Theta Foundation: 133–150.
  5. ^ Rieffel, M. A. (1982). "Morita equivalence for operator algebras". Proceedings of Symposia in Pure Mathematics. 38. American Mathematical Society: 176–257.
  6. ^ Baaj, S.; Skandalis, G. (1993). "Unitaires multiplicatifs et dualité pour les produits croisés de C*-algèbres". Annales Scientifiques de l'École Normale Supérieure. 26 (4): 425–488. doi:10.24033/asens.1677.
  7. ^ Woronowicz, S. L. (1991). "Unbounded elements affiliated with C*-algebras and non-compact quantum groups". Communications in Mathematical Physics. 136 (2): 399–432. Bibcode:1991CMaPh.136..399W. doi:10.1007/BF02100032. S2CID 118184597.
  8. ^ Arveson, William (1976). An Invitation to C*-Algebras. Springer-Verlag. p. 35.
  9. ^ In the case when   is non-unital, the spectrum of an element is calculated in the C*-algebra generated by adjoining a unit to  .
  10. ^ This result in fact holds for semi-inner-product  -modules, which may have non-zero elements   such that  , as the proof does not rely on the nondegeneracy property.
  11. ^ Kasparov, G. G. (1980). "Hilbert C*-modules: Theorems of Stinespring and Voiculescu". Journal of Operator Theory. 4. ThetaFoundation: 133–150.
  12. ^ Wegge-Olsen 1993, pp. 240-241.
  13. ^ Wegge-Olsen 1993, pp. 242-243.
  14. ^ Brown, Ozawa 2008, section 4.6.
  15. ^ Buss, Meyer, Zhu, 2013, section 2.2.
  16. ^ Brown, Ozawa 2008, pp. 138-139.
  17. ^ Buss, Meyer, Zhu 2013, section 2.2.
  18. ^ Brown, Ozawa, 2008, section 4.6.
  19. ^ Fowler, Raeburn, 1999, section 1.
  20. ^ Fowler, Raeburn, 1999, Proposition 1.3.
  21. ^ Brown, Ozawa, 2008, Example 4.6.10.

References

edit
edit