Injective tensor product

In mathematics, the injective tensor product of two topological vector spaces was introduced by Alexander Grothendieck and was used by him to define nuclear spaces.

Preliminaries and notation

Throughout let X,Y, and Z be topological vector spaces and be a linear map.

  • is a topological homomorphism or homomorphism, if it is linear, continuous, and is an open map, where , the image of L, has the subspace topology induced by Y.
    • If S is a subspace of X then both the quotient map and the canonical injection are homomorphisms. In particular, any linear map can be canonically decomposed as follows: where defines a bijection.
  • The set of continuous linear maps (resp. continuous bilinear maps ) will be denoted by L(X, Z) (resp. B(X, Y; Z)) where if Z is the scalar field then we may instead write L(X) (resp. B(X, Y)).
  • The set of separately continuous bilinear maps (i.e. continuous in each variable when the other variable is fixed) will be denoted by where if is the scalar field then we may instead write .
  • We will denote the continuous dual space of X by X* or and the algebraic dual space (which is the vector space of all linear functionals on X, whether continuous or not) by .
    • To increase the clarity of the exposition, we use the common convention of writing elements of with a prime following the symbol (e.g. denotes an element of and not, say, a derivative and the variables x and need not be related in any way).

Notation for topologies

  • σ(X, X′) denotes the coarsest topology on X making every map in X′ continuous and or denotes X endowed with this topology.
  • σ(X′, X) denotes weak-* topology on X* and or denotes X′ endowed with this topology.
    • Note that every induces a map defined by . σ(X′, X) is the coarsest topology on X′ making all such maps continuous.
  • b(X, X′) denotes the topology of bounded convergence on X and or denotes X endowed with this topology.
  • b(X′, X) denotes the topology of bounded convergence on X′ or the strong dual topology on X′ and or denotes X′ endowed with this topology.
    • As usual, if X* is considered as a topological vector space but it has not been made clear what topology it is endowed with, then the topology will be assumed to be b(X′, X).
  • 𝜏(X, X′) denotes the Mackey topology on X or the topology of uniform convergence on the convex balanced weakly compact subsets of X′ and or denotes X endowed with this topology. is the finest locally convex TVS topology on X whose continuous dual space is equal to .
  • 𝜏(X′, X) denotes the Mackey topology on X′ or the topology of uniform convergence on the convex balanced weakly compact subsets of X and or denotes X endowed with this topology.
    • Note that .[1]
  • ε(X, X′) denotes the topology of uniform convergence on equicontinuous subsets of X′ and or denotes X endowed with this topology.
    • If H is a set of linear mappings then H is equicontinuous if and only if it is equicontinuous at the origin; that is, if and only if for every neighborhood V of 0 in Y, there exists a neighborhood U of 0 in X such that for every .
  • A set H of linear maps from X to Y is called equicontinuous if for every neighborhood V of 0 in Y, there exists a neighborhood U of 0 in X such that for all .[2]

Definition

Throughout let X and Y be topological vector spaces with continuous dual spaces and . Note that almost all results described are independent of whether these vector spaces are over or but to simplify the exposition we will assume that they are over the field .

Continuous bilinear maps as a tensor product

Note that the question of whether or not one vector space is a tensor product of two other vector spaces is a purely algebraic one (i.e. the answer does not depend on the topologies of X or Y). Nevertheless, the vector space of continuous bilinear functionals is always a tensor product of X and Y, as we now describe.

For every we now define bilinear form, denoted by the symbol xy, from into the underlying field (i.e. ) by

.

This induces a canonical map defined by sending to the bilinear form . The span of the range of this map is . The following theorem may be used to verify that together with the above map ⊗ is a tensor product of X and Y.

Theorem Let X, Y, and Z be vector spaces and let be a bilinear map. Then the following are equivalent:[3]

  1. (Z, T) is a tensor product of X and Y;
  2. (a) the image of T spans all of Z, and (b) X and Y are T-linearly disjoint (this means that for all positive integers n and all elements and such that , (i) if all are linearly independent then all are 0, and (ii) if all are linearly independent then all are 0).
  • Equivalently, X and Y are T-linearly disjoint if and only if for all linearly independent sequences in X and all linearly independent sequences in Y, the vectors are linearly independent.

Topology

Henceforth, all topological vector spaces considered will be assumed to be locally convex. If Z is any locally convex topological vector space, then for any equicontinuous subsets and , and any neighborhood in Z, define

It can be shown that every set is bounded, which is necessary and sufficient for the collection of all such to form a locally convex TVS topology on called the ε-topology. When (resp. , , etc.) is endowed with the ε-topology, it will be denoted by (resp. , , etc.). (Note that ).

In particular, when Z is the underlying scalar field then since , we will denote by , which is called the injective tensor product of X and Y. The space is not necessarily complete so we denote its completion by . The space is complete if and only if both X and Y are complete, in which case the completion of is a subvector space (denoted by ) of . If X and Y are normed then so is . And is a Banach space if and only if both X and Y are Banach spaces.[4]

Equicontinuous sets

One reason for converging on equicontinuous subsets (of all possibilities) is the following important fact:

A set of continuous linear functionals H on a TVS X (not necessarily Hausdorff or locally convex) is equicontinuous if and only if it is contained in the polar of some neighborhood U of 0 in X (i.e. ).

Since a TVS's topology is completely determined by the open neighborhoods of the origin, when combined with the bipolar theorem, this means that via the operation of taking the polar of a set, the collection of equicontinuous subsets of "encodes" all information about X's topology: distinct TVS topologies on X produce distinct collections of equicontinuous subsets, and given any such collection one may recover the TVS's original topology by taking the polars of sets in the collection. Thus uniform convergence on the collection of equicontinuous subsets is essentially convergence on the very topologies of the TVSs and allows one to more directly relate this topology with the topology of the original spaces. Furthermore, the topology of a locally convex Hausdorff space X is identical to the topology of uniform convergence on the equicontinuous subsets of .[5]

For this reason, we now list some properties of equicontinuous sets that are relevant for dealing with the injective tensor product. Throughout X and Y are arbitrary TVSs and H is a collection of linear maps from X into Y.

  • If is equicontinuous then the subspace topologies that H inherits from the following topologies on are identical:[6]
  1. the topology of precompact convergence
  2. the topology of compact convergence
  3. the topology of pointwise convergence
  4. the topology of pointwise convergence on a given dense subset of X
  • An equicontinuous set is bounded in the topology of bounded convergence (i.e. bounded in ).[6] So in particular, H will also bounded in every TVS topology that is finer coarser than the topology of bounded convergence.
  • If X is a barrelled space and Y Y is locally convex then for any subset , the following are equivalent:
  1. H is equicontinuous;
  2. H is bounded in the topology of pointwise convergence (i.e. bounded in );
  3. H is bounded in the topology of bounded convergence (i.e. bounded in ).

In particular, to show that a set H is equicontinuous it suffices to show that it's bounded in the topology of pointwise converge.[7]

  • If X is a Baire space then any subset that is bounded in is necessarily equicontinuous.[7]
  • If X is separable, Y is metrizable, and D is a dense subset of X, then the topology of pointwise convergence on D makes metrizable so that in particular, the subspace topology that any equicontinuous subset inherits from is metrizable.[6]

We now restrict our attention to properties of equicontinuous subsets of the continuous dual space (where Y is now the underlying scalar field of X).

  • The weak closure of an equicontinuous set of linear functionals on X is a compact subspace of .[6]
  • If X is separable then every weakly closed equicontinuous subset of is a metrizable compact space when it is given the weak topology (i.e. the subspace topology inherited from ).[6]
  • If X is a normable space then a subset is equicontinuous if and only if it is strongly bounded (i.e. bounded in ).[6]
  • If X is a barrelled space then for any subset , the following are equivalent:[7]
  1. H is equicontinuous;
  2. H is relatively compact in the weak dual topology.
  3. H is weakly bounded;
  4. H is strongly bounded;

We mention some additional important basic properties relevant to the injective tensor product:

  • Suppose that is a bilinear map where is a Fréchet space, is metrizable, and Y is locally convex. If B is separately continuous then it is continuous.[8]

Canonical identification of separately continuous bilinear maps with linear maps

One may show that the set equality holds (i.e. if is a linear map, then is continuous if and only if is continuous, where here Y has its original topology).[9]

There is also a canonical vector space isomorphism .[9] To define it, for every separately continuous bilinear form defined on and every , let be defined by . Since is canonically vector space-isomorphic to Y (via the canonical map value at y), we will identify as an element of Y, which we'll denote by . This defines a map given by and so the canonical isomorphism is of course defined by .

When is given the topology of uniform convergence on equicontinous subsets of X′, the canonical map becomes a TVS-isomorphism .[9] In particular, can be canonically TVS-embedded into ; furthermore the image in of under the canonical map J consists exactly of the space of continuous linear maps whose image is finite dimensional.[4]

One always has . If X is normed then is in fact a topological vector subspace of . And if in addition Y is Banach then so is (even if X is not complete).[4]

Properties

  • The canonical map is continuous.[10]
  • If X and Y are normed then is normable.
  • is Hausdorff if and only if both X and Y are Hausdorff.[10]
  • Suppose that and are two linear maps between locally convex spaces. If both u and v are continuous then so is their tensor product .[11]
    • If u and v are both TVS-embeddings then so is .[12]
    • If (resp. ) is a linear subspace of (resp. ) then is canonically isomorphic to a linear subspace of and is canonically isomorphic to a linear subspace of .[13]
    • There are examples of u and v such that both u and v are surjective homomorphisms but is not a homomorphism.[14]
    • If all four spaces are normed then .[15]
  • The ε-topology is always finer than the π-topology and coarser than the inductive topology (which is the finest locally convex TVS topology making separately continuous).
  • If X and Y are normed then for all , .[15]

Relation to projective tensor product and nuclear spaces

The strongest locally convex topology on making the canonical map (defined by sending to the bilinear form ) continuous is called the projective topology or the π-topology. When is endowed with this topology then it will be denoted by and called the projective tensor product of X and Y.

The following definition was used by Grothendieck to define nuclear spaces.[16]

Definition 0: Let X be a locally convex topological vector space. Then X is nuclear if for any locally convex space Y, the canonical vector space embedding is an embedding of TVSs whose image is dense in the codomain.

Canonical identifications of bilinear and linear maps

In this section we describe canonical identifications between spaces of bilinear and linear maps. These identifications will be used to define important subspaces and topologies (particularly those that relate to nuclear operators and nuclear spaces).

Dual spaces of the injective tensor product and its completion

Suppose that denotes the TVS-embedding of into its completion and let be its transpose, which is a vector space-isomorphism. This identifies the continuous dual space of as being identical to the continuous dual space of .

The identity map is continuous (by definition of the π-topology) so there exists a unique continuous linear extension . If X and Y are Hilbert spaces then is injective and the dual of is canonically isometrically isomorphic to the vector space of nuclear operators from X into Y (with the trace norm).

Injective tensor product of Hilbert spaces

There is a canonical map that sends to the linear map defined by , where it may be shown that the definition of does not depend on the particular choice of representation of z. The map is continuous and when is complete, it has a continuous extension .

When X and Y are Hilbert spaces then is a TVS-embedding and isometry (when the spaces are given their usual norms) whose range is the space of all compact linear operators from X into Y (which is a closed vector subspace of . Hence is identical to space of compact operators from into Y (note the prime on X). The space of compact linear operators between any two Banach spaces (which includes Hilbert spaces) X and Y is a closed subset of .[17]

Furthermore, the canonical map is injective when X and Y are Hilbert spaces. [17]

Integral forms and operators

Integral bilinear forms

Let denote the identity map and denote its transpose, which is a continuous injection. Recall that is canonically identified with , the space of continuous bilinear maps on . In this way, the continuous dual space of can be canonically identified as a subvector space of , denoted by . The elements of are called integral (bilinear) forms on . The following theorem justifies the word integral.

Theorem[18][19] The dual J(X, Y) of consists of exactly those continuous bilinear forms v on that can be represented in the form of a map

where S and T are some closed, equicontinuous subsets of and , respectively, and is a positive Radon measure on the compact set with total mass . Furthermore, if A is an equicontinuous subset of J(X, Y) then the elements can be represented with fixed and running through a norm bounded subset of the space of Radon measures on .

Integral linear operators

Given a linear map , one can define a canonical bilinear form , called the associated bilinear form on , by . A continuous map is called integral if its associated bilinear form is an integral bilinear form.[20] An integral map is of the form, for every and :

for suitable weakly closed and equicontinuous aubsets and of and , respectively, and some positive Radon measure of total mass .

Canonical map into L(X; Y)

There is a canonical map that sends to the linear map defined by , where it may be shown that the definition of does not depend on the particular choice of representation of z.

Examples

Space of summable families

Throughout this section we fix some arbitrary (possibly uncountable) set A, a TVS X, and we let be the directed set of all finite subsets of A directed by inclusion .

Let be a family of elements in a TVS X and for every finite subset H of A, let . We call summable in X if the limit of the net converges in X to some element (any such element is called its sum). The set of all such summable families is a vector subspace of denoted by .

We now define a topology on S in a very natural way. This topology turns out to be the injective topology taken from and transferred to S via a canonical vector space isomorphism (the obvious one). This is a common occurrence when studying the injective and projective tensor products of function/sequence spaces and TVSs: the "natural way" in which one would define (from scratch) a topology on such a tensor product is frequently equivalent to the injective or projective tensor product topology.

Let denote a base of convex balanced neighborhoods of 0 in X and for each , let denote its Minkowski functional. For any such U and any , let

where defines a seminorm on S. The family of seminorms generates a topology making S into a locally convex space. The vector space S endowed with this topology will be denoted by .[21] The special case where X is the scalar field will be denoted by .

There is a canonical embedding of vector spaces defined by linearizing the bilinear map defined by .[21]

Theorem:[21] The canonical embedding (of vector spaces) becomes an embedding of topological vector spaces when is given the injective topology and furthermore, its range is dense in its codomain. If is a completion of X then the continuous extension of this embedding is an isomorphism of TVSs. So in particular, if X is complete then is canonically isomorphic to .

Space of continuously differentiable vector-valued functions

Throughout, let be an open subset of , where is an integer and let be a locally convex topological vector space (TVS).

Definition[22] Suppose and is a function such that with a limit point of . Then we say that f is differentiable at if there exist n vectors in Y, called the partial derivatives of f, such that
in Y
where .

One may naturally extend the notion of continuously differentiable function to Y-valued functions defined on . For any , let denote the vector space of all Y-valued maps defined on and let denote the vector subspace of consisting of all maps in that have compact support.

One may then define topologies on and in the same manner as the topologies on and are defined for the space of distributions and test functions (see the article: Differentiable vector-valued functions from Euclidean space). All of this work in extending the definition of differentiability and various topologies turns out to be exactly equivalent to simply taking the completed injective tensor product:

Theorem[23] If Y is a complete Hausdorff locally convex space, then is canonically isomorphic to the injective tensor product .

Spaces of continuous maps from a compact space

If Y is a normed space and if K is a compact set, then the -norm on is equal to .[23] If H and K are two compact spaces, then , where this canonical map is an isomorphism of Banach spaces.[23]

Spaces of sequences converging to 0

If Y is a normed space, then let denote the space of all sequences in Y that converge to the origin and give this space the norm . Let denote . Then for any Banach space Y, is canonically isometrically isomorphic to .[23]

Schwartz space of functions

We will now generalize the Schwartz space to functions valued in a TVS. Let be the space of all such that for all pairs of polynomials P and Q in n variables, is a bounded subset of Y. To generalize the topology of the Schwartz space to , we give the topology of uniform convergence over of the functions , as P and Q vary over all possible pairs of polynomials in n variables.[23]

Theorem:[23] If Y is a complete locally convex space, then is canonically isomorphic to .

See also

References

  1. Treves 2006, pp. 432-434.
  2. Treves 2006, pp. 338-345.
  3. Treves 2006, pp. 403-404.
  4. Treves 2006, pp. 432-433.
  5. Treves 2006, pp. 368-370.
  6. Treves 2006, pp. 338-343.
  7. Treves 2006, pp. 347-350.
  8. Treves 2006, pp. 351-354.
  9. Treves 2006, pp. 428-430.
  10. Treves 2006, p. 434.
  11. Treves 2006, p. 439.
  12. Treves 2006, p. 440.
  13. Treves 2006, p. 441.
  14. Treves 2006, p. 442.
  15. Treves 2006, p. 444.
  16. Schaefer 1999, p. 170.
  17. Treves 2006, p. 494.
  18. Schaefer 1999, p. 168.
  19. Treves 2006, pp. 500-502.
  20. Treves 2006, pp. 502-505.
  21. Schaefer 1999, pp. 179-184.
  22. Treves 2006, pp. 412-419.
  23. Treves 2006, pp. 446-451.
  • Diestel, Joe (2008). The metric theory of tensor products : Grothendieck's résumé revisited. Providence, R.I: American Mathematical Society. ISBN 0-8218-4440-7. OCLC 185095773.CS1 maint: ref=harv (link)
  • Dubinsky, Ed (1979). The structure of nuclear Fréchet spaces. Berlin New York: Springer-Verlag. ISBN 3-540-09504-7. OCLC 5126156.CS1 maint: ref=harv (link)
  • Grothendieck, Grothendieck (1966). Produits tensoriels topologiques et espaces nucléaires (in French). Providence: American Mathematical Society. ISBN 0-8218-1216-5. OCLC 1315788.CS1 maint: ref=harv (link)
  • Husain, Taqdir (1978). Barrelledness in topological and ordered vector spaces. Berlin New York: Springer-Verlag. ISBN 3-540-09096-7. OCLC 4493665.CS1 maint: ref=harv (link)
  • Khaleelulla, S. M. (July 1, 1982). Written at Berlin Heidelberg. Counterexamples in Topological Vector Spaces. Lecture Notes in Mathematics. 936. Berlin New York: Springer-Verlag. ISBN 978-3-540-11565-6. OCLC 8588370.CS1 maint: ref=harv (link) CS1 maint: date and year (link)
  • Narici, Lawrence; Beckenstein, Edward (2011). Topological Vector Spaces. Pure and applied mathematics (Second ed.). Boca Raton, FL: CRC Press. ISBN 978-1584888666. OCLC 144216834.
  • Nlend, H (1977). Bornologies and functional analysis : introductory course on the theory of duality topology-bornology and its use in functional analysis. Amsterdam New York New York: North-Holland Pub. Co. Sole distributors for the U.S.A. and Canada, Elsevier-North Holland. ISBN 0-7204-0712-5. OCLC 2798822.CS1 maint: ref=harv (link)
  • Nlend, H (1981). Nuclear and conuclear spaces : introductory courses on nuclear and conuclear spaces in the light of the duality. Amsterdam New York New York, N.Y: North-Holland Pub. Co. Sole distributors for the U.S.A. and Canada, Elsevier North-Holland. ISBN 0-444-86207-2. OCLC 7553061.CS1 maint: ref=harv (link)
  • Pietsch, Albrecht (1972). Nuclear locally convex spaces. Berlin,New York: Springer-Verlag. ISBN 0-387-05644-0. OCLC 539541.CS1 maint: ref=harv (link)
  • Robertson, A. P. (1973). Topological vector spaces. Cambridge England: University Press. ISBN 0-521-29882-2. OCLC 589250.CS1 maint: ref=harv (link)
  • Ryan, Raymond (2002). Introduction to tensor products of Banach spaces. London New York: Springer. ISBN 1-85233-437-1. OCLC 48092184.CS1 maint: ref=harv (link)
  • Schaefer, Helmut H.; Wolff, Manfred P. (1999). Topological Vector Spaces. GTM. 8 (Second ed.). New York, NY: Springer New York Imprint Springer. ISBN 978-1-4612-7155-0. OCLC 840278135.CS1 maint: ref=harv (link)
  • Trèves, François (August 6, 2006) [1967]. Topological Vector Spaces, Distributions and Kernels. Mineola, N.Y.: Dover Publications. ISBN 978-0-486-45352-1. OCLC 853623322.CS1 maint: ref=harv (link) CS1 maint: date and year (link)
  • Wong (1979). Schwartz spaces, nuclear spaces, and tensor products. Berlin New York: Springer-Verlag. ISBN 3-540-09513-6. OCLC 5126158.CS1 maint: ref=harv (link)
This article is issued from Wikipedia. The text is licensed under Creative Commons - Attribution - Sharealike. Additional terms may apply for the media files.