Blog Archive

Friday, May 19, 2023

05-19-2023-0245 - uniform norm, triangle inequality, discontinuous linear map, weak topology , coherent topology, dual system, dual pair, Transpose of a linear map with respect to pairings, Metrizability and separability, polarity, embedding, equicontinuous, subset, image, adjoint, canonical duality, Space of finite sequences, Reductive dual pair, Role of the axiom of choice, impact, Discontinuous_linear_map, constructivism, A linear map from a finite-dimensional space is always continuous, Classical Invariant Theory, general linear group, centralizer, subgroup, representation theory, isometry group, symplectic vector space, Bounded subsets , strongest polar, bounded convergence, weak convergence, compact subsets, balanced, injective, absolute polar, disk, ring, complex conjugate vector space, coordinates, bilinear, annihilator, orthogonal complement, analogously, degeneracy, convex, initial topology, weakly continuous, etc. (draft)

In topology and related areas of mathematics, the set of all possible topologies on a given set forms a partially ordered set. This order relation can be used for comparison of the topologies

https://en.wikipedia.org/wiki/Comparison_of_topologies

From Wikipedia, the free encyclopedia

In mathematics, weak topology is an alternative term for certain initial topologies, often on topological vector spaces or spaces of linear operators, for instance on a Hilbert space. The term is most commonly used for the initial topology of a topological vector space (such as a normed vector space) with respect to its continuous dual. The remainder of this article will deal with this case, which is one of the concepts of functional analysis.

One may call subsets of a topological vector space weakly closed (respectively, weakly compact, etc.) if they are closed (respectively, compact, etc.) with respect to the weak topology. Likewise, functions are sometimes called weakly continuous (respectively, weakly differentiable, weakly analytic, etc.) if they are continuous (respectively, differentiable, analytic, etc.) with respect to the weak topology. 

https://en.wikipedia.org/wiki/Weak_topology

From Wikipedia, the free encyclopedia

In mathematics, a dual system, dual pair, or duality over a field is a triple consisting of two vector spaces and over and a non-degenerate bilinear map . Duality theory, the study of dual systems, is part of functional analysis.

According to Helmut H. Schaefer, "the study of a locally convex space in terms of its dual is the central part of the modern theory of topological vector spaces, for it provides the deepest and most beautiful results of the subject."[1]

Definition, notation, and conventions

Pairings

A pairing or pair over a field is a triple which may also be denoted by consisting of two vector spaces and over (which this article assumes is either the real numbers or the complex numbers ) and a bilinear map which is called the bilinear map associated with the pairing[2] or simply the pairing's map/bilinear form.

For every define

and for every define
Every is a linear functional on and every is a linear functional on Let
where each of these sets forms a vector space of linear functionals.

It is common practice to write instead of in which case the pair is often denoted by rather than
However, this article will reserve use of for the canonical evaluation map (defined below) so as to avoid confusion for readers not familiar with this subject.

Dual pairings

A pairing is called a dual system, a dual pair,[3] or a duality over if the bilinear form is non-degenerate, which means that it satisfies the following two separation axioms:

  1. separates/distinguishes points of : if is such that then ; or equivalently, for all non-zero the map is not identically (i.e. there exists a such that );
  2. separates/distinguishes points of : if is such that then ; or equivalently, for all non-zero the map is not identically (i.e. there exists an such that ).

In this case say that is non-degenerate, say that places and in duality (or in separated duality), and is called the duality pairing of the [2][3]

Total subsets

A subset of is called total if for every

implies A total subset of is defined analogously (see footnote).[note 1]

Orthogonality

The vectors and are called orthogonal, written if Two subsets and are orthogonal, written if ; that is, if for all and The definition of a subset being orthogonal to a vector is defined analogously.

The orthogonal complement or annihilator of a subset is

Polar sets

Throughout, will be a pairing over The absolute polar or polar of a subset of is the set:[4]

Dually, the absolute polar or polar of a subset of is denoted by and defined by

In this case, the absolute polar of a subset of is also called the absolute prepolar or prepolar of and may be denoted by

The polar is necessarily a convex set containing where if is balanced then so is and if is a vector subspace of then so too is a vector subspace of [5]

If then the bipolar of denoted by is the set Similarly, if then the bipolar of is

If is a vector subspace of then and this is also equal to the real polar of

Dual definitions and results

Given a pairing define a new pairing where for all [2]

There is a repeating theme in duality theory, which is that any definition for a pairing has a corresponding dual definition for the pairing

Convention and Definition: Given any definition for a pairing one obtains a dual definition by applying it to the pairing This conventions also apply to theorems.
Convention: Adhering to common practice, unless clarity is needed, whenever a definition (or result) for a pairing is given then this article will omit mention of the corresponding dual definition (or result) but nevertheless use it.

For instance, if " distinguishes points of " (resp, " is a total subset of ") is defined as above, then this convention immediately produces the dual definition of " distinguishes points of " (resp, " is a total subset of ").

This following notation is almost ubiquitous and it allows us to avoid having to assign a symbol to

Convention and Notation: If a definition and its notation for a pairing depends on the order of and (e.g. the definition of the Mackey topology on ) then by switching the order of and then it is meant that definition applied to (e.g. actually denotes the topology ).

For instance, once the weak topology on is defined, which is denoted by then this definition will automatically be applied to the pairing so as to obtain the definition of the weak topology on where this topology will be denoted by rather than

Identification of with

Although it is technically incorrect and an abuse of notation, this article will also adhere to the following nearly ubiquitous convention of treating a pairing interchangeably with and also of denoting by

Examples

Restriction of a pairing

Suppose that is a pairing, is a vector subspace of and is a vector subspace of Then the restriction of to is the pairing If is a duality then it's possible for a restrictions to fail to be a duality (e.g. if and ).

This article will use the common practice of denoting the restriction by

Canonical duality on a vector space

Suppose that is a vector space and let denote the algebraic dual space of (that is, the space of all linear functionals on ). There is a canonical duality where which is called the evaluation map or the natural or canonical bilinear functional on Note in particular that for any is just another way of denoting ; i.e.

If is a vector subspace of then the restriction of to is called the canonical pairing where if this pairing is a duality then it is instead called the canonical duality. Clearly, always distinguishes points of so the canonical pairing is a dual system if and only if separates points of The following notation is now nearly ubiquitous in duality theory.

The evaluation map will be denoted by (rather than by ) and will be written rather than

Assumption: As is common practice, if is a vector space and is a vector space of linear functionals on then unless stated otherwise, it will be assumed that they are associated with the canonical pairing

If is a vector subspace of then distinguishes points of (or equivalently, is a duality) if and only if distinguishes points of or equivalently if is total (that is, for all implies ).[2]

Canonical duality on a topological vector space

Suppose is a topological vector space (TVS) with continuous dual space Then the restriction of the canonical duality to × defines a pairing for which separates points of If separates points of (which is true if, for instance, is a Hausdorff locally convex space) then this pairing forms a duality.[3]

Assumption: As is commonly done, whenever is a TVS then, unless indicated otherwise, it will be assumed without comment that it's associated with the canonical pairing

Polars and duals of TVSs

The following result shows that the continuous linear functionals on a TVS are exactly those linear functionals that are bounded on a neighborhood of the origin.

Theorem[2] — Let be a TVS with algebraic dual and let be a basis of neighborhoods of at the origin. Under the canonical duality the continuous dual space of is the union of all as ranges over (where the polars are taken in ).

Inner product spaces and complex conjugate spaces

A pre-Hilbert space is a dual pairing if and only if is vector space over or has dimension Here it is assumed that the sesquilinear form is conjugate homogeneous in its second coordinate and homogeneous in its first coordinate.

  • If is a real Hilbert space then forms a dual system.
  • If is a complex Hilbert space then forms a dual system if and only if If is non-trivial then does not even form pairing since the inner product is sesquilinear rather than bilinear.[2]

Suppose that is a complex pre-Hilbert space with scalar multiplication denoted as usual by juxtaposition or by a dot Define the map

where the right hand side uses the scalar multiplication of Let denote the complex conjugate vector space of where denotes the additive group of (so vector addition in is identical to vector addition in ) but with scalar multiplication in being the map (instead of the scalar multiplication that is endowed with).

The map defined by is linear in both coordinates[note 2] and so forms a dual pairing.

Other examples

  • Suppose and for all let
    Then is a pairing such that distinguishes points of but does not distinguish points of Furthermore,
  • Let (where is such that ), and Then is a dual system.
  • Let and be vector spaces over the same field Then the bilinear form places and in duality.[3]
  • A sequence space and its beta dual with the bilinear map defined as for forms a dual system.

Weak topology

Suppose that is a pairing of vector spaces over If then the weak topology on induced by (and ) is the weakest TVS topology on denoted by or simply making all maps continuous as ranges over [2] If is not clear from context then it should be assumed to be all of in which case it is called the weak topology on (induced by ). The notation or (if no confusion could arise) simply is used to denote endowed with the weak topology Importantly, the weak topology depends entirely on the function the usual topology on and 's vector space structure but not on the algebraic structures of

Similarly, if then the dual definition of the weak topology on induced by (and ), which is denoted by or simply (see footnote for details).[note 3]

Definition and Notation: If "" is attached to a topological definition (e.g. -converges, -bounded, etc.) then it means that definition when the first space (i.e. ) carries the topology. Mention of or even and may be omitted if no confusion will arise. So for instance, if a sequence in "-converges" or "weakly converges" then this means that it converges in whereas if it were a sequence in then this would mean that it converges in ).

The topology is locally convex since it is determined by the family of seminorms defined by as ranges over [2] If and is a net in then -converges to if converges to in [2] A net -converges to if and only if for all converges to If is a sequence of orthonormal vectors in Hilbert space, then converges weakly to 0 but does not norm-converge to 0 (or any other vector).[2]

If is a pairing and is a proper vector subspace of such that is a dual pair, then is strictly coarser than [2]

Bounded subsets

A subset of is -bounded if and only if

where

Hausdorffness

If is a pairing then the following are equivalent:

  1. distinguishes points of ;
  2. The map defines an injection from into the algebraic dual space of ;[2]
  3. is Hausdorff.[2]

Weak representation theorem

The following theorem is of fundamental importance to duality theory because it completely characterizes the continuous dual space of

Weak representation theorem[2] — Let be a pairing over the field Then the continuous dual space of is

Furthermore,

  1. If is a continuous linear functional on then there exists some such that ; if such a exists then it is unique if and only if distinguishes points of
    • Note that whether or not distinguishes points of is not dependent on the particular choice of
  2. The continuous dual space of may be identified with the quotient space where
    • This is true regardless of whether or not distinguishes points of or distinguishes points of

Consequently, the continuous dual space of is

With respect to the canonical pairing, if is a TVS whose continuous dual space separates points on (i.e. such that is Hausdorff, which implies that is also necessarily Hausdorff) then the continuous dual space of is equal to the set of all "evaluation at a point " maps as ranges over (i.e. the map that send to ). This is commonly written as

This very important fact is why results for polar topologies on continuous dual spaces, such as the strong dual topology on for example, can also often be applied to the original TVS ; for instance, being identified with means that the topology on can instead be thought of as a topology on Moreover, if is endowed with a topology that is finer than then the continuous dual space of will necessarily contain as a subset. So for instance, when is endowed with the strong dual topology (and so is denoted by ) then
which (among other things) allows for to be endowed with the subspace topology induced on it by, say, the strong dual topology (this topology is also called the strong bidual topology and it appears in the theory of reflexive spaces: the Hausdorff locally convex TVS is said to be semi-reflexive if and it will be called reflexive if in addition the strong bidual topology on is equal to 's original/starting topology).

Orthogonals, quotients, and subspaces

If is a pairing then for any subset of :

  • and this set is -closed;[2]
  • ;[2]
    • Thus if is a -closed vector subspace of then
  • If is a family of -closed vector subspaces of then
    [2]
  • If is a family of subsets of then [2]

If is a normed space then under the canonical duality, is norm closed in and is norm closed in [2]

Subspaces

Suppose that is a vector subspace of and let denote the restriction of to The weak topology on is identical to the subspace topology that inherits from

Also, is a paired space (where means ) where is defined by

The topology is equal to the subspace topology that inherits from [6] Furthermore, if is a dual system then so is [6]

Quotients

Suppose that is a vector subspace of Then is a paired space where is defined by

The topology is identical to the usual quotient topology induced by on [6]

Polars and the weak topology

If is a locally convex space and if is a subset of the continuous dual space then is -bounded if and only if for some barrel in [2]

The following results are important for defining polar topologies.

If is a pairing and then:[2]

  1. The polar of is a closed subset of
  2. The polars of the following sets are identical: (a) ; (b) the convex hull of ; (c) the balanced hull of ; (d) the -closure of ; (e) the -closure of the convex balanced hull of
  3. The bipolar theorem: The bipolar of denoted by is equal to the -closure of the convex balanced hull of
    • The bipolar theorem in particular "is an indispensable tool in working with dualities."[5]
  4. is -bounded if and only if is absorbing in
  5. If in addition distinguishes points of then is -bounded if and only if it is -totally bounded.

If is a pairing and is a locally convex topology on that is consistent with duality, then a subset of is a barrel in if and only if is the polar of some -bounded subset of [7]

Transposes

Transpose of a linear map with respect to pairings

Let and be pairings over and let be a linear map.

For all let be the map defined by It is said that 's transpose or adjoint is well-defined if the following conditions are satisfied:

  1. distinguishes points of (or equivalently, the map from into the algebraic dual is injective), and
  2. where and .

In this case, for any there exists (by condition 2) a unique (by condition 1) such that ), where this element of will be denoted by This defines a linear map

called the transpose or adjoint of with respect to and (this should not to be confused with the Hermitian adjoint). It is easy to see that the two conditions mentioned above (i.e. for "the transpose is well-defined") are also necessary for to be well-defined. For every the defining condition for is

that is,
     for all

By the conventions mentioned at the beginning of this article, this also defines the transpose of linear maps of the form [note 4] [note 5] [note 6] [note 7] etc. (see footnote for details).

Properties of the transpose

Throughout, and be pairings over and will be a linear map whose transpose is well-defined.

  • is injective (i.e. ) if and only if the range of is dense in [2]
  • If in addition to being well-defined, the transpose of is also well-defined then
  • Suppose is a pairing over and is a linear map whose transpose is well-defined. Then the transpose of which is is well-defined and
  • If is a vector space isomorphism then is bijective, the transpose of which is is well-defined, and [2]
  • Let and let denotes the absolute polar of then:[2]
    1. ;
    2. if for some then ;
    3. if is such that then ;
    4. if and are weakly closed disks then if and only if ;
These results hold when the real polar is used in place of the absolute polar.

If and are normed spaces under their canonical dualities and if is a continuous linear map, then [2]

Weak continuity

A linear map is weakly continuous (with respect to and ) if is continuous.

The following result shows that the existence of the transpose map is intimately tied to the weak topology.

Proposition — Assume that distinguishes points of and is a linear map. Then the following are equivalent:

  1. is weakly continuous (that is, is continuous);
  2. ;
  3. the transpose of is well-defined.

If is weakly continuous then

  • is weakly continuous, meaning that is continuous;
  • the transpose of is well-defined if and only if distinguishes points of in which case

Weak topology and the canonical duality

Suppose that is a vector space and that is its the algebraic dual. Then every -bounded subset of is contained in a finite dimensional vector subspace and every vector subspace of is -closed.[2]

Weak completeness

If is a complete topological vector space say that is -complete or (if no ambiguity can arise) weakly-complete. There exist Banach spaces that are not weakly-complete (despite being complete in their norm topology).[2]

If is a vector space then under the canonical duality, is complete.[2] Conversely, if is a Hausdorff locally convex TVS with continuous dual space then is complete if and only if ; that is, if and only if the map defined by sending to the evaluation map at (i.e. ) is a bijection.[2]

In particular, with respect to the canonical duality, if is a vector subspace of such that separates points of then is complete if and only if Said differently, there does not exist a proper vector subspace of such that is Hausdorff and is complete in the weak-* topology (i.e. the topology of pointwise convergence). Consequently, when the continuous dual space of a Hausdorff locally convex TVS is endowed with the weak-* topology, then is complete if and only if (that is, if and only if every linear functional on is continuous).

Identification of Y with a subspace of the algebraic dual

If distinguishes points of and if denotes the range of the injection then is a vector subspace of the algebraic dual space of and the pairing becomes canonically identified with the canonical pairing (where is the natural evaluation map). In particular, in this situation it will be assumed without loss of generality that is a vector subspace of 's algebraic dual and is the evaluation map.

Convention: Often, whenever is injective (especially when forms a dual pair) then it is common practice to assume without loss of generality that is a vector subspace of the algebraic dual space of that is the natural evaluation map, and also denote by

In a completely analogous manner, if distinguishes points of then it is possible for to be identified as a vector subspace of 's algebraic dual space.[3]

Algebraic adjoint

In the special case where the dualities are the canonical dualities and the transpose of a linear map is always well-defined. This transpose is called the algebraic adjoint of and it will be denoted by ; that is, In this case, for all [2][8] where the defining condition for is:

or equivalently,

Examples

If for some integer is a basis for with dual basis is a linear operator, and the matrix representation of with respect to is then the transpose of is the matrix representation with respect to of

Weak continuity and openness

Suppose that and are canonical pairings (so and ) that are dual systems and let be a linear map. Then is weakly continuous if and only if it satisfies any of the following equivalent conditions:[2]

  1. is continuous;
  2. the transpose of F, with respect to and is well-defined.

If is weakly continuous then will be continuous and furthermore, [8]

A map between topological spaces is relatively open if is an open mapping, where is the range of [2]

Suppose that and are dual systems and is a weakly continuous linear map. Then the following are equivalent:[2]

  1. is relatively open;
  2. The range of is -closed in ;

Furthermore,

  • is injective (resp. bijective) if and only if is surjective (resp. bijective);
  • is surjective if and only if is relatively open and injective.
Transpose of a map between TVSs

The transpose of map between two TVSs is defined if and only if is weakly continuous.

If is a linear map between two Hausdorff locally convex topological vector spaces then:[2]

  • If is continuous then it is weakly continuous and is both Mackey continuous and strongly continuous.
  • If is weakly continuous then it is both Mackey continuous and strongly continuous (defined below).
  • If is weakly continuous then it is continuous if and only if maps equicontinuous subsets of to equicontinuous subsets of
  • If and are normed spaces then is continuous if and only if it is weakly continuous, in which case
  • If is continuous then is relatively open if and only if is weakly relatively open (i.e. is relatively open) and every equicontinuous subsets of is the image of some equicontinuous subsets of
  • If is continuous injection then is a TVS-embedding (or equivalently, a topological embedding) if and only if every equicontinuous subsets of is the image of some equicontinuous subsets of

Metrizability and separability

Let be a locally convex space with continuous dual space and let [2]

  1. If is equicontinuous or -compact, and if is such that is dense in then the subspace topology that inherits from is identical to the subspace topology that inherits from
  2. If is separable and is equicontinuous then when endowed with the subspace topology induced by is metrizable.
  3. If is separable and metrizable, then is separable.
  4. If is a normed space then is separable if and only if the closed unit call the continuous dual space of is metrizable when given the subspace topology induced by
  5. If is a normed space whose continuous dual space is separable (when given the usual norm topology), then is separable.

Polar topologies and topologies compatible with pairing

Starting with only the weak topology, the use of polar sets produces a range of locally convex topologies. Such topologies are called polar topologies. The weak topology is the weakest topology of this range.

Throughout, will be a pairing over and will be a non-empty collection of -bounded subsets of

Polar topologies

Given a collection of subsets of , the polar topology on determined by (and ) or the -topology on is the unique topological vector space (TVS) topology on for which

forms a subbasis of neighborhoods at the origin.[2] When is endowed with this -topology then it is denoted by Y. Every polar topology is necessarily locally convex.[2] When is a directed set with respect to subset inclusion (i.e. if for all there exists some such that ) then this neighborhood subbasis at 0 actually forms a neighborhood basis at 0.[2]

The following table lists some of the more important polar topologies.

Notation: If denotes a polar topology on then endowed with this topology will be denoted by or simply (e.g. for we'd have so that and all denote endowed with ).

("topology of uniform convergence on ...")
Notation Name ("topology of...") Alternative name
finite subsets of
(or -closed disked hulls of finite subsets of )

pointwise/simple convergence weak/weak* topology
-compact disks
Mackey topology
-compact convex subsets compact convex convergence
-compact subsets
(or balanced -compact subsets)
compact convergence
-bounded subsets
bounded convergence strong topology
Strongest polar topology

Definitions involving polar topologies

Continuity

A linear map is Mackey continuous (with respect to and ) if is continuous.[2]

A linear map is strongly continuous (with respect to and ) if is continuous.[2]

Bounded subsets

A subset of is weakly bounded (resp. Mackey bounded, strongly bounded) if it is bounded in (resp. bounded in bounded in ).

Topologies compatible with a pair

If is a pairing over and is a vector topology on then is a topology of the pairing and that it is compatible (or consistent) with the pairing if it is locally convex and if the continuous dual space of [note 8] If distinguishes points of then by identifying as a vector subspace of 's algebraic dual, the defining condition becomes: [2] Some authors (e.g. [Trèves 2006] and [Schaefer 1999]) require that a topology of a pair also be Hausdorff,[3][9] which it would have to be if distinguishes the points of (which these authors assume).

The weak topology is compatible with the pairing (as was shown in the Weak representation theorem) and it is in fact the weakest such topology. There is a strongest topology compatible with this pairing and that is the Mackey topology. If is a normed space that is not reflexive then the usual norm topology on its continuous dual space is not compatible with the duality [2]

Mackey–Arens theorem

The following is one of the most important theorems in duality theory.

Mackey–Arens theorem I[2] — Let will be a pairing such that distinguishes the points of and let be a locally convex topology on (not necessarily Hausdorff). Then is compatible with the pairing if and only if is a polar topology determined by some collection of -compact disks that cover[note 9]

It follows that the Mackey topology which recall is the polar topology generated by all -compact disks in is the strongest locally convex topology on that is compatible with the pairing A locally convex space whose given topology is identical to the Mackey topology is called a Mackey space. The following consequence of the above Mackey-Arens theorem is also called the Mackey-Arens theorem.

Mackey–Arens theorem II[2] — Let will be a pairing such that distinguishes the points of and let be a locally convex topology on Then is compatible with the pairing if and only if

Mackey's theorem, barrels, and closed convex sets

If is a TVS (over or ) then a half-space is a set of the form for some real and some continuous real linear functional on

Theorem — If is a locally convex space (over or ) and if is a non-empty closed and convex subset of then is equal to the intersection of all closed half spaces containing it.[10]

The above theorem implies that the closed and convex subsets of a locally convex space depend entirely on the continuous dual space. Consequently, the closed and convex subsets are the same in any topology compatible with duality; that is, if and are any locally convex topologies on with the same continuous dual spaces, then a convex subset of is closed in the topology if and only if it is closed in the topology. This implies that the -closure of any convex subset of is equal to its -closure and that for any -closed disk in [2] In particular, if is a subset of then is a barrel in if and only if it is a barrel in [2]

The following theorem shows that barrels (i.e. closed absorbing disks) are exactly the polars of weakly bounded subsets.

Theorem[2] — Let will be a pairing such that distinguishes the points of and let be a topology of the pair. Then a subset of is a barrel in if and only if it is equal to the polar of some -bounded subset of

If is a topological vector space then:[2][11]

  1. A closed absorbing and balanced subset of absorbs each convex compact subset of (i.e. there exists a real such that contains that set).
  2. If is Hausdorff and locally convex then every barrel in absorbs every convex bounded complete subset of

All of this leads to Mackey's theorem, which is one of the central theorems in the theory of dual systems. In short, it states the bounded subsets are the same for any two Hausdorff locally convex topologies that are compatible with the same duality.

Mackey's theorem[11][2] — Suppose that is a Hausdorff locally convex space with continuous dual space and consider the canonical duality If is any topology on that is compatible with the duality on then the bounded subsets of are the same as the bounded subsets of

Examples

Space of finite sequences

Let denote the space of all sequences of scalars such that for all sufficiently large Let and define a bilinear map by

Then [2] Moreover, a subset is -bounded (resp. -bounded) if and only if there exists a sequence of positive real numbers such that for all and all indices (resp. and ).[2] It follows that there are weakly bounded (that is, -bounded) subsets of that are not strongly bounded (that is, not -bounded).

See also

Notes


  • A subset of is total if for all
    implies

  • That is linear in its first coordinate is obvious. Suppose is a scalar. Then which shows that is linear in its second coordinate.

  • The weak topology on is the weakest TVS topology on making all maps continuous, as ranges over The dual notation of or simply may also be used to denote endowed with the weak topology If is not clear from context then it should be assumed to be all of in which case it is simply called the weak topology on (induced by ).

  • If is a linear map then 's transpose, is well-defined if and only if distinguishes points of and In this case, for each the defining condition for is:

  • If is a linear map then 's transpose, is well-defined if and only if distinguishes points of and In this case, for each the defining condition for is:

  • If is a linear map then 's transpose, is well-defined if and only if distinguishes points of and In this case, for each the defining condition for is:

  • If is a linear map then 's transpose, is well-defined if and only if distinguishes points of and In this case, for each the defining condition for is:

  • Of course, there is an analogous definition for topologies on to be "compatible it a pairing" but this article will only deal with topologies on

    1. Recall that a collection of subsets of a set is said to cover if every point of is contained in some set belonging to the collection.

    References


  • Schaefer & Wolff 1999, p. 122.

  • Narici & Beckenstein 2011, pp. 225–273.

  • Schaefer & Wolff 1999, pp. 122–128.

  • Trèves 2006, p. 195.

  • Schaefer & Wolff 1999, pp. 123–128.

  • Narici & Beckenstein 2011, pp. 260–264.

  • Narici & Beckenstein 2011, pp. 251–253.

  • Schaefer & Wolff 1999, pp. 128–130.

  • Trèves 2006, pp. 368–377.

  • Narici & Beckenstein 2011, p. 200.

    1. Trèves 2006, pp. 371–372.

    Bibliography

    External links

     

     https://en.wikipedia.org/wiki/Dual_system#Weak_continuity

    In the mathematical field of representation theory, a reductive dual pair is a pair of subgroups (G, G′) of the isometry group Sp(W) of a symplectic vector space W, such that G is the centralizer of G′ in Sp(W) and vice versa, and these groups act reductively on W. Somewhat more loosely, one speaks of a dual pair whenever two groups are the mutual centralizers in a larger group, which is frequently a general linear group. The concept was introduced by Roger Howe in Howe (1979). Its strong ties with Classical Invariant Theory are discussed in Howe (1989a)

    https://en.wikipedia.org/wiki/Reductive_dual_pair

    In Weyl's wonderful and terrible1 book The Classical Groups [W] one may discern two main themes: first, the study of the polynomial invariants for an arbitrary number of (contravariant or covariant) variables for a standard classical group action; second, the isotypic decomposition of the full tensor algebra for such an action.

    1Most people who know the book feel the material in it is wonderful. Many also feel the presentation is terrible. (The author is not among these latter.)

    Howe (1989, p.539)

    The Classical Groups: Their Invariants and Representations is a mathematics book by Hermann Weyl (1939), which describes classical invariant theory in terms of representation theory. It is largely responsible for the revival of interest in invariant theory, which had been almost killed off by David Hilbert's solution of its main problems in the 1890s.

    Weyl (1939a) gave an informal talk about the topic of his book. There was a second edition in 1946. 

    https://en.wikipedia.org/wiki/The_Classical_Groups

    Representation theory is a branch of mathematics that studies abstract algebraic structures by representing their elements as linear transformations of vector spaces, and studies modules over these abstract algebraic structures.[1][2] In essence, a representation makes an abstract algebraic object more concrete by describing its elements by matrices and their algebraic operations (for example, matrix addition, matrix multiplication). The theory of matrices and linear operators is well-understood, so representations of more abstract objects in terms of familiar linear algebra objects helps glean properties and sometimes simplify calculations on more abstract theories.  

    https://en.wikipedia.org/wiki/Representation_theory

    In mathematics, specifically linear algebra, a degenerate bilinear form f (x, y ) on a vector space V is a bilinear form such that the map from V to V (the dual space of V ) given by v ↦ (xf (x, v )) is not an isomorphism. An equivalent definition when V is finite-dimensional is that it has a non-trivial kernel: there exist some non-zero x in V such that

    for all

     https://en.wikipedia.org/wiki/Degenerate_bilinear_form

    https://en.wikipedia.org/wiki/Strong_dual_space

    https://en.wikipedia.org/wiki/Normed_vector_space

    https://en.wikipedia.org/wiki/Coherent_topology

    https://en.wikipedia.org/wiki/Initial_topology

    https://en.wikipedia.org/wiki/Dual_space#Continuous_dual_space

    https://en.wikipedia.org/wiki/Dual_space#Continuous_dual_space

     

     https://en.wikipedia.org/wiki/Discontinuous_linear_map

     

    In mathematics, linear maps form an important class of "simple" functions which preserve the algebraic structure of linear spaces and are often used as approximations to more general functions (see linear approximation). If the spaces involved are also topological spaces (that is, topological vector spaces), then it makes sense to ask whether all linear maps are continuous. It turns out that for maps defined on infinite-dimensional topological vector spaces (e.g., infinite-dimensional normed spaces), the answer is generally no: there exist discontinuous linear maps. If the domain of definition is complete, it is trickier; such maps can be proven to exist, but the proof relies on the axiom of choice and does not provide an explicit example.

    A linear map from a finite-dimensional space is always continuous

    Let X and Y be two normed spaces and a linear map from X to Y. If X is finite-dimensional, choose a basis in X which may be taken to be unit vectors. Then,

    and so by the triangle inequality,
    Letting
    and using the fact that
    for some C>0 which follows from the fact that any two norms on a finite-dimensional space are equivalent, one finds
    Thus, is a bounded linear operator and so is continuous. In fact, to see this, simply note that f is linear, and therefore for some universal constant K. Thus for any we can choose so that ( and are the normed balls around and ), which gives continuity.

    If X is infinite-dimensional, this proof will fail as there is no guarantee that the supremum M exists. If Y is the zero space {0}, the only map between X and Y is the zero map which is trivially continuous. In all other cases, when X is infinite-dimensional and Y is not the zero space, one can find a discontinuous map from X to Y.

    A concrete example

    Examples of discontinuous linear maps are easy to construct in spaces that are not complete; on any Cauchy sequence of linearly independent vectors which does not have a limit, there is a linear operator such that the quantities grow without bound. In a sense, the linear operators are not continuous because the space has "holes".

    For example, consider the space X of real-valued smooth functions on the interval [0, 1] with the uniform norm, that is,

    The derivative-at-a-point map, given by
    defined on X and with real values, is linear, but not continuous. Indeed, consider the sequence
    for This sequence converges uniformly to the constantly zero function, but

    as instead of which would hold for a continuous map. Note that T is real-valued, and so is actually a linear functional on X (an element of the algebraic dual space X*). The linear map XX which assigns to each function its derivative is similarly discontinuous. Note that although the derivative operator is not continuous, it is closed.

    The fact that the domain is not complete here is important. Discontinuous operators on complete spaces require a little more work.

    A nonconstructive example

    An algebraic basis for the real numbers as a vector space over the rationals is known as a Hamel basis (note that some authors use this term in a broader sense to mean an algebraic basis of any vector space). Note that any two noncommensurable numbers, say 1 and , are linearly independent. One may find a Hamel basis containing them, and define a map so that f acts as the identity on the rest of the Hamel basis, and extend to all of by linearity. Let {rn}n be any sequence of rationals which converges to . Then limn f(rn) = π, but By construction, f is linear over (not over ), but not continuous. Note that f is also not measurable; an additive real function is linear if and only if it is measurable, so for every such function there is a Vitali set. The construction of f relies on the axiom of choice.

    This example can be extended into a general theorem about the existence of discontinuous linear maps on any infinite-dimensional normed space (as long as the codomain is not trivial).

    General existence theorem

    Discontinuous linear maps can be proven to exist more generally, even if the space is complete. Let X and Y be normed spaces over the field K where or Assume that X is infinite-dimensional and Y is not the zero space. We will find a discontinuous linear map f from X to K, which will imply the existence of a discontinuous linear map g from X to Y given by the formula where is an arbitrary nonzero vector in Y.

    If X is infinite-dimensional, to show the existence of a linear functional which is not continuous then amounts to constructing f which is not bounded. For that, consider a sequence (en)n () of linearly independent vectors in X. Define

    for each Complete this sequence of linearly independent vectors to a vector space basis of X, and define T at the other vectors in the basis to be zero. T so defined will extend uniquely to a linear map on X, and since it is clearly not bounded, it is not continuous.

    Notice that by using the fact that any set of linearly independent vectors can be completed to a basis, we implicitly used the axiom of choice, which was not needed for the concrete example in the previous section but one.

    Role of the axiom of choice

    As noted above, the axiom of choice (AC) is used in the general existence theorem of discontinuous linear maps. In fact, there are no constructive examples of discontinuous linear maps with complete domain (for example, Banach spaces). In analysis as it is usually practiced by working mathematicians, the axiom of choice is always employed (it is an axiom of ZFC set theory); thus, to the analyst, all infinite-dimensional topological vector spaces admit discontinuous linear maps.

    On the other hand, in 1970 Robert M. Solovay exhibited a model of set theory in which every set of reals is measurable.[1] This implies that there are no discontinuous linear real functions. Clearly AC does not hold in the model.

    Solovay's result shows that it is not necessary to assume that all infinite-dimensional vector spaces admit discontinuous linear maps, and there are schools of analysis which adopt a more constructivist viewpoint. For example, H. G. Garnir, in searching for so-called "dream spaces" (topological vector spaces on which every linear map into a normed space is continuous), was led to adopt ZF + DC + BP (dependent choice is a weakened form and the Baire property is a negation of strong AC) as his axioms to prove the Garnir–Wright closed graph theorem which states, among other things, that any linear map from an F-space to a TVS is continuous. Going to the extreme of constructivism, there is Ceitin's theorem, which states that every function is continuous (this is to be understood in the terminology of constructivism, according to which only representable functions are considered to be functions).[2] Such stances are held by only a small minority of working mathematicians.

    The upshot is that the existence of discontinuous linear maps depends on AC; it is consistent with set theory without AC that there are no discontinuous linear maps on complete spaces. In particular, no concrete construction such as the derivative can succeed in defining a discontinuous linear map everywhere on a complete space.

    Closed operators

    Many naturally occurring linear discontinuous operators are closed, a class of operators which share some of the features of continuous operators. It makes sense to ask which linear operators on a given space are closed. The closed graph theorem asserts that an everywhere-defined closed operator on a complete domain is continuous, so to obtain a discontinuous closed operator, one must permit operators which are not defined everywhere.

    To be more concrete, let be a map from to with domain written We don't lose much if we replace X by the closure of That is, in studying operators that are not everywhere-defined, one may restrict one's attention to densely defined operators without loss of generality.

    If the graph of is closed in we call T closed. Otherwise, consider its closure in If is itself the graph of some operator is called closable, and is called the closure of

    So the natural question to ask about linear operators that are not everywhere-defined is whether they are closable. The answer is, "not necessarily"; indeed, every infinite-dimensional normed space admits linear operators that are not closable. As in the case of discontinuous operators considered above, the proof requires the axiom of choice and so is in general nonconstructive, though again, if X is not complete, there are constructible examples.

    In fact, there is even an example of a linear operator whose graph has closure all of Such an operator is not closable. Let X be the space of polynomial functions from [0,1] to and Y the space of polynomial functions from [2,3] to . They are subspaces of C([0,1]) and C([2,3]) respectively, and so normed spaces. Define an operator T which takes the polynomial function xp(x) on [0,1] to the same function on [2,3]. As a consequence of the Stone–Weierstrass theorem, the graph of this operator is dense in so this provides a sort of maximally discontinuous linear map (confer nowhere continuous function). Note that X is not complete here, as must be the case when there is such a constructible map.

    Impact for dual spaces

    The dual space of a topological vector space is the collection of continuous linear maps from the space into the underlying field. Thus the failure of some linear maps to be continuous for infinite-dimensional normed spaces implies that for these spaces, one needs to distinguish the algebraic dual space from the continuous dual space which is then a proper subset. It illustrates the fact that an extra dose of caution is needed in doing analysis on infinite-dimensional spaces as compared to finite-dimensional ones.

    Beyond normed spaces

    The argument for the existence of discontinuous linear maps on normed spaces can be generalized to all metrizable topological vector spaces, especially to all Fréchet spaces, but there exist infinite-dimensional locally convex topological vector spaces such that every functional is continuous.[3] On the other hand, the Hahn–Banach theorem, which applies to all locally convex spaces, guarantees the existence of many continuous linear functionals, and so a large dual space. In fact, to every convex set, the Minkowski gauge associates a continuous linear functional. The upshot is that spaces with fewer convex sets have fewer functionals, and in the worst-case scenario, a space may have no functionals at all other than the zero functional. This is the case for the spaces with from which it follows that these spaces are nonconvex. Note that here is indicated the Lebesgue measure on the real line. There are other spaces with which do have nontrivial dual spaces.

    Another such example is the space of real-valued measurable functions on the unit interval with quasinorm given by

    This non-locally convex space has a trivial dual space.

    One can consider even more general spaces. For example, the existence of a homomorphism between complete separable metric groups can also be shown nonconstructively.

    See also

    References


  • Solovay, Robert M. (1970), "A model of set-theory in which every set of reals is Lebesgue measurable", Annals of Mathematics, Second Series, 92: 1–56, doi:10.2307/1970696, MR 0265151.

  • Schechter, Eric (1996), Handbook of Analysis and Its Foundations, Academic Press, p. 136, ISBN 9780080532998.

    1. For example, the weak topology w.r.t. the space of all (algebraically) linear functionals.
    • Constantin Costara, Dumitru Popa, Exercises in Functional Analysis, Springer, 2003. ISBN 1-4020-1560-7.
    • Schechter, Eric, Handbook of Analysis and its Foundations, Academic Press, 1997. ISBN 0-12-622760-8.

     https://en.wikipedia.org/wiki/Discontinuous_linear_map

    https://en.wikipedia.org/wiki/Triangle_inequality

    https://en.wikipedia.org/wiki/Uniform_norm

     

    No comments:

    Post a Comment