Locally convex topological vector space

In functional analysis and related areas of mathematics, locally convex topological vector spaces (LCTVS) or locally convex spaces are examples of topological vector spaces (TVS) that generalize normed spaces. They can be defined as topological vector spaces whose topology is generated by translations of balanced, absorbent, convex sets. Alternatively they can be defined as a vector space with a family of seminorms, and a topology can be defined in terms of that family. Although in general such spaces are not necessarily normable, the existence of a convex local base for the zero vector is strong enough for the Hahn–Banach theorem to hold, yielding a sufficiently rich theory of continuous linear functionals.

Fréchet spaces are locally convex spaces that are completely metrizable (with a choice of complete metric). They are generalizations of Banach spaces, which are complete vector spaces with respect to a metric generated by a norm.

History edit

Metrizable topologies on vector spaces have been studied since their introduction in Maurice Fréchet's 1902 PhD thesis Sur quelques points du calcul fonctionnel (wherein the notion of a metric was first introduced). After the notion of a general topological space was defined by Felix Hausdorff in 1914,[1] although locally convex topologies were implicitly used by some mathematicians, up to 1934 only John von Neumann would seem to have explicitly defined the weak topology on Hilbert spaces and strong operator topology on operators on Hilbert spaces.[2][3] Finally, in 1935 von Neumann introduced the general definition of a locally convex space (called a convex space by him).[4][5]

A notable example of a result which had to wait for the development and dissemination of general locally convex spaces (amongst other notions and results, like nets, the product topology and Tychonoff's theorem) to be proven in its full generality, is the Banach–Alaoglu theorem which Stefan Banach first established in 1932 by an elementary diagonal argument for the case of separable normed spaces[6] (in which case the unit ball of the dual is metrizable).

Definition edit

Suppose   is a vector space over   a subfield of the complex numbers (normally   itself or  ). A locally convex space is defined either in terms of convex sets, or equivalently in terms of seminorms.

Definition via convex sets edit

A topological vector space (TVS) is called locally convex if it has a neighborhood basis (that is, a local base) at the origin consisting of balanced, convex sets.[7] The term locally convex topological vector space is sometimes shortened to locally convex space or LCTVS.

A subset   in   is called

  1. Convex if for all   and     In other words,   contains all line segments between points in  
  2. Circled if for all   and scalars   if   then   If   this means that   is equal to its reflection through the origin. For   it means for any     contains the circle through   centred on the origin, in the one-dimensional complex subspace generated by  
  3. Balanced if for all   and scalars   if   then   If   this means that if   then   contains the line segment between   and   For   it means for any     contains the disk with   on its boundary, centred on the origin, in the one-dimensional complex subspace generated by   Equivalently, a balanced set is a circled cone (in the TVS R^2, C=ball centered at the origin of radius sqrt{2}, x=(1,1) belongs, s=2, sx=(2,2) does not belong, C is not a cone but C is balanced, sx is in C, for all x belonging to C and scalar s:|s|<1) .
  4. A cone (when the underlying field is ordered) if for all   and    
  5. Absorbent or absorbing if for every   there exists   such that   for all   satisfying   The set   can be scaled out by any "large" value to absorb every point in the space.
    • In any TVS, every neighborhood of the origin is absorbent.[7]
  6. Absolutely convex or a disk if it is both balanced and convex. This is equivalent to it being closed under linear combinations whose coefficients absolutely sum to  ; such a set is absorbent if it spans all of  

In fact, every locally convex TVS has a neighborhood basis of the origin consisting of absolutely convex sets (that is, disks), where this neighborhood basis can further be chosen to also consist entirely of open sets or entirely of closed sets.[8] Every TVS has a neighborhood basis at the origin consisting of balanced sets, but only a locally convex TVS has a neighborhood basis at the origin consisting of sets that are both balanced and convex. It is possible for a TVS to have some neighborhoods of the origin that are convex and yet not be locally convex because it has no neighborhood basis at the origin consisting entirely of convex sets (that is, every neighborhood basis at the origin contains some non-convex set); for example, every non-locally convex TVS   has itself (that is,  ) has a convex neighborhood of the origin.

Because translation is continuous (by definition of topological vector space), all translations are homeomorphisms, so every base for the neighborhoods of the origin can be translated to a base for the neighborhoods of any given vector.

Definition via seminorms edit

A seminorm on   is a map   such that

  1.   is nonnegative or positive semidefinite:  ;
  2.   is positive homogeneous or positive scalable:   for every scalar   So, in particular,  ;
  3.   is subadditive. It satisfies the triangle inequality:  

If   satisfies positive definiteness, which states that if   then   then   is a norm. While in general seminorms need not be norms, there is an analogue of this criterion for families of seminorms, separatedness, defined below.

If   is a vector space and   is a family of seminorms on   then a subset   of   is called a base of seminorms for   if for all   there exists a   and a real   such that  [9]

Definition (second version): A locally convex space is defined to be a vector space   along with a family   of seminorms on  

Seminorm topology edit

Suppose that   is a vector space over   where   is either the real or complex numbers. A family of seminorms   on the vector space   induces a canonical vector space topology on  , called the initial topology induced by the seminorms. By definition, it is the coarsest topology on   for which all maps in   are continuous. It will not carry the structure of a topological vector space (TVS) though.

The vector space operations fail to be continuous in this topology, as all sets in the semi-norm topology will be symmetric, where as the lcs topology allows for more sets.

It is possible for a locally convex topology on a space   to be induced by a family of norms but for   to not be normable (that is, to have its topology be induced by a single norm).

Basis and subbases edit

Let   denote the open ball of radius   in  . The family of sets   as   ranges over a family of seminorms   and   ranges over the positive real numbers is a subbasis at the origin for the topology induced by  . These sets are convex, as follows from properties 2 and 3 of seminorms. Intersections of finitely many such sets are then also convex, and since the collection of all such finite intersections is a basis at the origin it follows that the topology is locally convex in the sense of the first definition given above.

Recall that the topology of a TVS is translation invariant, meaning that if   is any subset of   containing the origin then for any     is a neighborhood of the origin if and only if   is a neighborhood of  ; thus it suffices to define the topology at the origin. A base of neighborhoods of   for this topology is obtained in the following way: for every finite subset   of   and every   let

 
Bases of seminorms and saturated families edit

If   is a locally convex space and if   is a collection of continuous seminorms on  , then   is called a base of continuous seminorms if it is a base of seminorms for the collection of all continuous seminorms on  .[9] Explicitly, this means that for all continuous seminorms   on  , there exists a   and a real   such that  [9] If   is a base of continuous seminorms for a locally convex TVS   then the family of all sets of the form   as   varies over   and   varies over the positive real numbers, is a base of neighborhoods of the origin in   (not just a subbasis, so there is no need to take finite intersections of such sets).[9][proof 1]

A family   of seminorms on a vector space   is called saturated if for any   and   in   the seminorm defined by   belongs to  

If   is a saturated family of continuous seminorms that induces the topology on   then the collection of all sets of the form   as   ranges over   and   ranges over all positive real numbers, forms a neighborhood basis at the origin consisting of convex open sets;[9] This forms a basis at the origin rather than merely a subbasis so that in particular, there is no need to take finite intersections of such sets.[9]

Basis of norms edit

The following theorem implies that if   is a locally convex space then the topology of   can be a defined by a family of continuous norms on   (a norm is a seminorm   where   implies  ) if and only if there exists at least one continuous norm on  .[10] This is because the sum of a norm and a seminorm is a norm so if a locally convex space is defined by some family   of seminorms (each of which is necessarily continuous) then the family   of (also continuous) norms obtained by adding some given continuous norm   to each element, will necessarily be a family of norms that defines this same locally convex topology. If there exists a continuous norm on a topological vector space   then   is necessarily Hausdorff but the converse is not in general true (not even for locally convex spaces or Fréchet spaces).

Theorem[11] — Let   be a Fréchet space over the field   Then the following are equivalent:

  1.   does not admit a continuous norm (that is, any continuous seminorm on   can not be a norm).
  2.   contains a vector subspace that is TVS-isomorphic to  
  3.   contains a complemented vector subspace that is TVS-isomorphic to  
Nets edit

Suppose that the topology of a locally convex space   is induced by a family   of continuous seminorms on  . If   and if   is a net in  , then   in   if and only if for all    [12] Moreover, if   is Cauchy in  , then so is   for every  [12]

Equivalence of definitions edit

Although the definition in terms of a neighborhood base gives a better geometric picture, the definition in terms of seminorms is easier to work with in practice. The equivalence of the two definitions follows from a construction known as the Minkowski functional or Minkowski gauge. The key feature of seminorms which ensures the convexity of their  -balls is the triangle inequality.

For an absorbing set   such that if   then   whenever   define the Minkowski functional of   to be

 

From this definition it follows that   is a seminorm if   is balanced and convex (it is also absorbent by assumption). Conversely, given a family of seminorms, the sets

 
form a base of convex absorbent balanced sets.

Ways of defining a locally convex topology edit

Theorem[7] — Suppose that   is a (real or complex) vector space and let   be a filter base of subsets of   such that:

  1. Every   is convex, balanced, and absorbing;
  2. For every   there exists some real   satisfying   such that  

Then   is a neighborhood base at 0 for a locally convex TVS topology on  

Theorem[7] — Suppose that   is a (real or complex) vector space and let   be a non-empty collection of convex, balanced, and absorbing subsets of   Then the set of all of all positive scalar multiples of finite intersections of sets in   forms a neighborhood base at the origin for a locally convex TVS topology on  


Example: auxiliary normed spaces

If   is convex and absorbing in   then the symmetric set   will be convex and balanced (also known as an absolutely convex set or a disk) in addition to being absorbing in   This guarantees that the Minkowski functional   of   will be a seminorm on   thereby making   into a seminormed space that carries its canonical pseudometrizable topology. The set of scalar multiples   as   ranges over   (or over any other set of non-zero scalars having   as a limit point) forms a neighborhood basis of absorbing disks at the origin for this locally convex topology. If   is a topological vector space and if this convex absorbing subset   is also a bounded subset of   then the absorbing disk   will also be bounded, in which case   will be a norm and   will form what is known as an auxiliary normed space. If this normed space is a Banach space then   is called a Banach disk.

Further definitions edit

  • A family of seminorms   is called total or separated or is said to separate points if whenever   holds for every   then   is necessarily   A locally convex space is Hausdorff if and only if it has a separated family of seminorms. Many authors take the Hausdorff criterion in the definition.
  • A pseudometric is a generalization of a metric which does not satisfy the condition that   only when   A locally convex space is pseudometrizable, meaning that its topology arises from a pseudometric, if and only if it has a countable family of seminorms. Indeed, a pseudometric inducing the same topology is then given by
     
    (where the   can be replaced by any positive summable sequence  ). This pseudometric is translation-invariant, but not homogeneous, meaning   and therefore does not define a (pseudo)norm. The pseudometric is an honest metric if and only if the family of seminorms is separated, since this is the case if and only if the space is Hausdorff. If furthermore the space is complete, the space is called a Fréchet space.
  • As with any topological vector space, a locally convex space is also a uniform space. Thus one may speak of uniform continuity, uniform convergence, and Cauchy sequences.
  • A Cauchy net in a locally convex space is a net   such that for every   and every seminorm   there exists some index   such that for all indices     In other words, the net must be Cauchy in all the seminorms simultaneously. The definition of completeness is given here in terms of nets instead of the more familiar sequences because unlike Fréchet spaces which are metrizable, general spaces may be defined by an uncountable family of pseudometrics. Sequences, which are countable by definition, cannot suffice to characterize convergence in such spaces. A locally convex space is complete if and only if every Cauchy net converges.
  • A family of seminorms becomes a preordered set under the relation   if and only if there exists an   such that for all     One says it is a directed family of seminorms if the family is a directed set with addition as the join, in other words if for every   and   there is a   such that   Every family of seminorms has an equivalent directed family, meaning one which defines the same topology. Indeed, given a family   let   be the set of finite subsets of   and then for every   define
     
    One may check that   is an equivalent directed family.
  • If the topology of the space is induced from a single seminorm, then the space is seminormable. Any locally convex space with a finite family of seminorms is seminormable. Moreover, if the space is Hausdorff (the family is separated), then the space is normable, with norm given by the sum of the seminorms. In terms of the open sets, a locally convex topological vector space is seminormable if and only if the origin has a bounded neighborhood.

Sufficient conditions edit

Hahn–Banach extension property edit

Let   be a TVS. Say that a vector subspace   of   has the extension property if any continuous linear functional on   can be extended to a continuous linear functional on  .[13] Say that   has the Hahn-Banach extension property (HBEP) if every vector subspace of   has the extension property.[13]

The Hahn-Banach theorem guarantees that every Hausdorff locally convex space has the HBEP. For complete metrizable TVSs there is a converse:

Theorem[13] (Kalton) — Every complete metrizable TVS with the Hahn-Banach extension property is locally convex.

If a vector space   has uncountable dimension and if we endow it with the finest vector topology then this is a TVS with the HBEP that is neither locally convex or metrizable.[13]

Properties edit


Throughout,   is a family of continuous seminorms that generate the topology of  

Topological closure

If   and   then   if and only if for every   and every finite collection   there exists some   such that  [14] The closure of   in   is equal to  [15]

Topology of Hausdorff locally convex spaces

Every Hausdorff locally convex space is homeomorphic to a vector subspace of a product of Banach spaces.[16] The Anderson–Kadec theorem states that every infinite–dimensional separable Fréchet space is homeomorphic to the product space   of countably many copies of   (this homeomorphism need not be a linear map).[17]

Properties of convex subsets edit

Algebraic properties of convex subsets

A subset   is convex if and only if   for all  [18] or equivalently, if and only if   for all positive real  [19] where because   always holds, the equals sign   can be replaced with   If   is a convex set that contains the origin then   is star shaped at the origin and for all non-negative real    

The Minkowski sum of two convex sets is convex; furthermore, the scalar multiple of a convex set is again convex.[20]

Topological properties of convex subsets

  • Suppose that   is a TVS (not necessarily locally convex or Hausdorff) over the real or complex numbers. Then the open convex subsets of   are exactly those that are of the form   for some   and some positive continuous sublinear functional   on  [21]
  • The interior and closure of a convex subset of a TVS is again convex.[20]
  • If   is a convex set with non-empty interior, then the closure of   is equal to the closure of the interior of  ; furthermore, the interior of   is equal to the interior of the closure of  [20][22]
    • So if the interior of a convex set   is non-empty then   is a closed (respectively, open) set if and only if it is a regular closed (respectively, regular open) set.
  • If   is convex and   then[23]   Explicitly, this means that if   is a convex subset of a TVS   (not necessarily Hausdorff or locally convex),   belongs to the closure of   and   belongs to the interior of   then the open line segment joining   and   belongs to the interior of   that is,  [22][24][proof 2]
  • If   is a closed vector subspace of a (not necessarily Hausdorff) locally convex space    is a convex neighborhood of the origin in   and if   is a vector not in   then there exists a convex neighborhood   of the origin in   such that   and  [20]
  • The closure of a convex subset of a locally convex Hausdorff space   is the same for all locally convex Hausdorff TVS topologies on   that are compatible with duality between   and its continuous dual space.[25]
  • In a locally convex space, the convex hull and the disked hull of a totally bounded set is totally bounded.[7]
  • In a complete locally convex space, the convex hull and the disked hull of a compact set are both compact.[7]
    • More generally, if   is a compact subset of a locally convex space, then the convex hull   (respectively, the disked hull  ) is compact if and only if it is complete.[7]
  • In a locally convex space, convex hulls of bounded sets are bounded. This is not true for TVSs in general.[26]
  • In a Fréchet space, the closed convex hull of a compact set is compact.[27]
  • In a locally convex space, any linear combination of totally bounded sets is totally bounded.[26]

Properties of convex hulls edit

For any subset   of a TVS   the convex hull (respectively, closed convex hull, balanced hull, convex balanced hull) of   denoted by   (respectively,      ), is the smallest convex (respectively, closed convex, balanced, convex balanced) subset of   containing  

  • The convex hull of compact subset of a Hilbert space is not necessarily closed and so also not necessarily compact. For example, let   be the separable Hilbert space   of square-summable sequences with the usual norm   and let   be the standard orthonormal basis (that is   at the  -coordinate). The closed set   is compact but its convex hull   is not a closed set because   belongs to the closure of   in   but   (since every sequence   is a finite convex combination of elements of   and so is necessarily   in all but finitely many coordinates, which is not true of  ).[28] However, like in all complete Hausdorff locally convex spaces, the closed convex hull   of this compact subset is compact. The vector subspace   is a pre-Hilbert space when endowed with the substructure that the Hilbert space   induces on it but   is not complete and   (since  ). The closed convex hull of   in   (here, "closed" means with respect to   and not to   as before) is equal to   which is not compact (because it is not a complete subset). This shows that in a Hausdorff locally convex space that is not complete, the closed convex hull of compact subset might fail to be compact (although it will be precompact/totally bounded).
  • In a Hausdorff locally convex space   the closed convex hull   of compact subset   is not necessarily compact although it is a precompact (also called "totally bounded") subset, which means that its closure, when taken in a completion   of   will be compact (here   so that   if and only if   is complete); that is to say,   will be compact. So for example, the closed convex hull   of a compact subset of   of a pre-Hilbert space   is always a precompact subset of   and so the closure of   in any Hilbert space   containing   (such as the Hausdorff completion of   for instance) will be compact (this is the case in the previous example above).
  • In a quasi-complete locally convex TVS, the closure of the convex hull of a compact subset is again compact.
  • In a Hausdorff locally convex TVS, the convex hull of a precompact set is again precompact.[29] Consequently, in a complete Hausdorff locally convex space, the closed convex hull of a compact subset is again compact.[30]
  • In any TVS, the convex hull of a finite union of compact convex sets is compact (and convex).[7]
    • This implies that in any Hausdorff TVS, the convex hull of a finite union of compact convex sets is closed (in addition to being compact[31] and convex); in particular, the convex hull of such a union is equal to the closed convex hull of that union.
    • In general, the closed convex hull of a compact set is not necessarily compact. However, every compact subset of   (where  ) does have a compact convex hull.[31]
    • In any non-Hausdorff TVS, there exist subsets that are compact (and thus complete) but not closed.
  • The bipolar theorem states that the bipolar (that is, the polar of the polar) of a subset of a locally convex Hausdorff TVS is equal to the closed convex balanced hull of that set.[32]
  • The balanced hull of a convex set is not necessarily convex.
  • If   and   are convex subsets of a topological vector space   and if   then there exist     and a real number   satisfying   such that  [20]
  • If   is a vector subspace of a TVS     a convex subset of   and   a convex subset of   such that   then  [20]
  • Recall that the smallest balanced subset of   containing a set   is called the balanced hull of   and is denoted by   For any subset   of   the convex balanced hull of   denoted by   is the smallest subset of   containing   that is convex and balanced.[33] The convex balanced hull of   is equal to the convex hull of the balanced hull of   (i.e.  ), but the convex balanced hull of   is not necessarily equal to the balanced hull of the convex hull of   (that is,   is not necessarily equal to  ).[33]
  • If   are subsets of a TVS   and if   is a scalar then    [34]   and   Moreover, if   is compact then  [35] However, the convex hull of a closed set need not be closed;[34] for example, the set   is closed in   but its convex hull is the open set  
  • If   are subsets of a TVS   whose closed convex hulls are compact, then  [35]
  • If   is a convex set in a complex vector space   and there exists some   such that   then   for all real   such that   In particular,   for all scalars   such that  
  • Carathéodory's theorem: If   is any subset of   (where  ) then for every   there exist a finite subset   containing at most   points whose convex hull contains   (that is,   and  ).[36]

Examples and nonexamples edit

Finest and coarsest locally convex topology edit

Coarsest vector topology edit

Any vector space   endowed with the trivial topology (also called the indiscrete topology) is a locally convex TVS (and of course, it is the coarsest such topology). This topology is Hausdorff if and only   The indiscrete topology makes any vector space into a complete pseudometrizable locally convex TVS.

In contrast, the discrete topology forms a vector topology on   if and only   This follows from the fact that every topological vector space is a connected space.

Finest locally convex topology edit

If   is a real or complex vector space and if   is the set of all seminorms on   then the locally convex TVS topology, denoted by   that   induces on   is called the finest locally convex topology on  [37] This topology may also be described as the TVS-topology on   having as a neighborhood base at the origin the set of all absorbing disks in  [37] Any locally convex TVS-topology on   is necessarily a subset of     is Hausdorff.[15] Every linear map from   into another locally convex TVS is necessarily continuous.[15] In particular, every linear functional on   is continuous and every vector subspace of   is closed in  ;[15] therefore, if   is infinite dimensional then   is not pseudometrizable (and thus not metrizable).[37] Moreover,   is the only Hausdorff locally convex topology on   with the property that any linear map from it into any Hausdorff locally convex space is continuous.[38] The space   is a bornological space.[39]

Examples of locally convex spaces edit

Every normed space is a Hausdorff locally convex space, and much of the theory of locally convex spaces generalizes parts of the theory of normed spaces. The family of seminorms can be taken to be the single norm. Every Banach space is a complete Hausdorff locally convex space, in particular, the   spaces with   are locally convex.

More generally, every Fréchet space is locally convex. A Fréchet space can be defined as a complete locally convex space with a separated countable family of seminorms.

The space   of real valued sequences with the family of seminorms given by

 
is locally convex. The countable family of seminorms is complete and separable, so this is a Fréchet space, which is not normable. This is also the limit topology of the spaces   embedded in   in the natural way, by completing finite sequences with infinitely many  

Given any vector space   and a collection   of linear functionals on it,   can be made into a locally convex topological vector space by giving it the weakest topology making all linear functionals in   continuous. This is known as the weak topology or the initial topology determined by   The collection   may be the algebraic dual of   or any other collection. The family of seminorms in this case is given by   for all   in  

Spaces of differentiable functions give other non-normable examples. Consider the space of smooth functions   such that   where   and   are multiindices. The family of seminorms defined by   is separated, and countable, and the space is complete, so this metrizable space is a Fréchet space. It is known as the Schwartz space, or the space of functions of rapid decrease, and its dual space is the space of tempered distributions.

An important function space in functional analysis is the space   of smooth functions with compact support in   A more detailed construction is needed for the topology of this space because the space   is not complete in the uniform norm. The topology on   is defined as follows: for any fixed compact set   the space   of functions   with   is a Fréchet space with countable family of seminorms   (these are actually norms, and the completion of the space   with the   norm is a Banach space  ). Given any collection   of compact sets, directed by inclusion and such that their union equal   the   form a direct system, and   is defined to be the limit of this system. Such a limit of Fréchet spaces is known as an LF space. More concretely,   is the union of all the   with the strongest locally convex topology which makes each inclusion map   continuous. This space is locally convex and complete. However, it is not metrizable, and so it is not a Fréchet space. The dual space of   is the space of distributions on  

More abstractly, given a topological space   the space   of continuous (not necessarily bounded) functions on   can be given the topology of uniform convergence on compact sets. This topology is defined by semi-norms   (as   varies over the directed set of all compact subsets of  ). When   is locally compact (for example, an open set in  ) the Stone–Weierstrass theorem applies—in the case of real-valued functions, any subalgebra of   that separates points and contains the constant functions (for example, the subalgebra of polynomials) is dense.

Examples of spaces lacking local convexity edit

Many topological vector spaces are locally convex. Examples of spaces that lack local convexity include the following:

  • The spaces   for   are equipped with the F-norm
     
    They are not locally convex, since the only convex neighborhood of zero is the whole space. More generally the spaces   with an atomless, finite measure   and   are not locally convex.
  • The space of measurable functions on the unit interval   (where we identify two functions that are equal almost everywhere) has a vector-space topology defined by the translation-invariant metric (which induces the convergence in measure of measurable functions; for random variables, convergence in measure is convergence in probability):
     
    This space is often denoted  

Both examples have the property that any continuous linear map to the real numbers is   In particular, their dual space is trivial, that is, it contains only the zero functional.

  • The sequence space     is not locally convex.

Continuous mappings edit

Theorem[40] — Let   be a linear operator between TVSs where   is locally convex (note that   need not be locally convex). Then   is continuous if and only if for every continuous seminorm   on  , there exists a continuous seminorm   on   such that  

Because locally convex spaces are topological spaces as well as vector spaces, the natural functions to consider between two locally convex spaces are continuous linear maps. Using the seminorms, a necessary and sufficient criterion for the continuity of a linear map can be given that closely resembles the more familiar boundedness condition found for Banach spaces.

Given locally convex spaces   and   with families of seminorms   and   respectively, a linear map   is continuous if and only if for every   there exist   and   such that for all  

 

In other words, each seminorm of the range of   is bounded above by some finite sum of seminorms in the domain. If the family   is a directed family, and it can always be chosen to be directed as explained above, then the formula becomes even simpler and more familiar:

 

The class of all locally convex topological vector spaces forms a category with continuous linear maps as morphisms.

Linear functionals edit

Theorem[40] — If   is a TVS (not necessarily locally convex) and if   is a linear functional on  , then   is continuous if and only if there exists a continuous seminorm   on   such that  

If   is a real or complex vector space,   is a linear functional on  , and   is a seminorm on  , then   if and only if  [41] If   is a non-0 linear functional on a real vector space   and if   is a seminorm on  , then   if and only if  [15]

Multilinear maps edit

Let   be an integer,   be TVSs (not necessarily locally convex), let   be a locally convex TVS whose topology is determined by a family   of continuous seminorms, and let   be a multilinear operator that is linear in each of its   coordinates. The following are equivalent:

  1.   is continuous.
  2. For every   there exist continuous seminorms   on   respectively, such that   for all  [15]
  3. For every   there exists some neighborhood of the origin in   on which   is bounded.[15]

See also edit

  • Convex metric space – metric space with the property any segment joining two points in that space has other points in it besides the endpoints
  • Krein–Milman theorem – On when a space equals the closed convex hull of its extreme points
  • Linear form – Linear map from a vector space to its field of scalars
  • Locally convex vector lattice
  • Minkowski functional – Function made from a set
  • Seminorm – nonnegative-real-valued function on a real or complex vector space that satisfies the triangle inequality and is absolutely homogenous
  • Sublinear functional – Type of function in linear algebra
  • Topological group – Group that is a topological space with continuous group action
  • Topological vector space – Vector space with a notion of nearness
  • Vector space – Algebraic structure in linear algebra

Notes edit

  1. ^ Hausdorff, F. Grundzüge der Mengenlehre (1914)
  2. ^ von Neumann, J. Collected works. Vol II. pp. 94–104
  3. ^ Dieudonne, J. History of Functional Analysis Chapter VIII. Section 1.
  4. ^ von Neumann, J. Collected works. Vol II. pp. 508–527
  5. ^ Dieudonne, J. History of Functional Analysis Chapter VIII. Section 2.
  6. ^ Banach, S. Theory of linear operations p. 75. Ch. VIII. Sec. 3. Theorem 4., translated from Theorie des operations lineaires (1932)
  7. ^ a b c d e f g h Narici & Beckenstein 2011, pp. 67–113.
  8. ^ Narici & Beckenstein 2011, p. 83.
  9. ^ a b c d e f Narici & Beckenstein 2011, p. 122.
  10. ^ Jarchow 1981, p. 130.
  11. ^ Jarchow 1981, pp. 129–130.
  12. ^ a b Narici & Beckenstein 2011, p. 126.
  13. ^ a b c d Narici & Beckenstein 2011, pp. 225–273.
  14. ^ Narici & Beckenstein 2011, p. 149.
  15. ^ a b c d e f g Narici & Beckenstein 2011, pp. 149–153.
  16. ^ Narici & Beckenstein 2011, pp. 115–154.
  17. ^ Bessaga & Pełczyński 1975, p. 189
  18. ^ Rudin 1991, p. 6.
  19. ^ Rudin 1991, p. 38.
  20. ^ a b c d e f Trèves 2006, p. 126.
  21. ^ Narici & Beckenstein 2011, pp. 177–220.
  22. ^ a b Schaefer & Wolff 1999, p. 38.
  23. ^ Jarchow 1981, pp. 101–104.
  24. ^ Conway 1990, p. 102.
  25. ^ Trèves 2006, p. 370.
  26. ^ a b Narici & Beckenstein 2011, pp. 155–176.
  27. ^ Rudin 1991, p. 7.
  28. ^ Aliprantis & Border 2006, p. 185.
  29. ^ Trèves 2006, p. 67.
  30. ^ Trèves 2006, p. 145.
  31. ^ a b Rudin 1991, pp. 72–73.
  32. ^ Trèves 2006, p. 362.
  33. ^ a b Trèves 2006, p. 68.
  34. ^ a b Narici & Beckenstein 2011, p. 108.
  35. ^ a b Dunford 1988, p. 415.
  36. ^ Rudin 1991, pp. 73–74.
  37. ^ a b c Narici & Beckenstein 2011, pp. 125–126.
  38. ^ Narici & Beckenstein 2011, p. 476.
  39. ^ Narici & Beckenstein 2011, p. 446.
  40. ^ a b Narici & Beckenstein 2011, pp. 126–128.
  41. ^ Narici & Beckenstein 2011, pp. 126-–128.
  1. ^ Let   be the open unit ball associated with the seminorm   and note that if   is real then   and so   Thus a basic open neighborhood of the origin induced by   is a finite intersection of the form   where   and   are all positive reals. Let   which is a continuous seminorm and moreover,   Pick   and   such that   where this inequality holds if and only if   Thus   as desired.
  2. ^ Fix   so it remains to show that   belongs to   By replacing   with   if necessary, we may assume without loss of generality that   and so it remains to show that   is a neighborhood of the origin. Let   so that   Since scalar multiplication by   is a linear homeomorphism     Since   and   it follows that   where because   is open, there exists some   which satisfies   Define   by   which is a homeomorphism because   The set   is thus an open subset of   that moreover contains   If   then   since   is convex,   and   which proves that   Thus   is an open subset of   that contains the origin and is contained in   Q.E.D.

References edit

  • Aliprantis, Charalambos D.; Border, Kim C. (2006). Infinite Dimensional Analysis: A Hitchhiker's Guide (Third ed.). Berlin: Springer Science & Business Media. ISBN 978-3-540-29587-7. OCLC 262692874.
  • Berberian, Sterling K. (1974). Lectures in Functional Analysis and Operator Theory. Graduate Texts in Mathematics. Vol. 15. New York: Springer. ISBN 978-0-387-90081-0. OCLC 878109401.
  • Bessaga, C.; Pełczyński, A. (1975), Selected Topics in Infinite-Dimensional Topology, Monografie Matematyczne, Warszawa: Panstwowe wyd. naukowe.
  • Bourbaki, Nicolas (1987) [1981]. Topological Vector Spaces: Chapters 1–5. Éléments de mathématique. Translated by Eggleston, H.G.; Madan, S. Berlin New York: Springer-Verlag. ISBN 3-540-13627-4. OCLC 17499190.
  • Conway, John (1990). A course in functional analysis. Graduate Texts in Mathematics. Vol. 96 (2nd ed.). New York: Springer-Verlag. ISBN 978-0-387-97245-9. OCLC 21195908.
  • Dunford, Nelson (1988). Linear operators (in Romanian). New York: Interscience Publishers. ISBN 0-471-60848-3. OCLC 18412261.
  • Edwards, Robert E. (1995). Functional Analysis: Theory and Applications. New York: Dover Publications. ISBN 978-0-486-68143-6. OCLC 30593138.
  • Grothendieck, Alexander (1973). Topological Vector Spaces. Translated by Chaljub, Orlando. New York: Gordon and Breach Science Publishers. ISBN 978-0-677-30020-7. OCLC 886098.
  • Jarchow, Hans (1981). Locally convex spaces. Stuttgart: B.G. Teubner. ISBN 978-3-519-02224-4. OCLC 8210342.
  • Köthe, Gottfried (1983) [1969]. Topological Vector Spaces I. Grundlehren der mathematischen Wissenschaften. Vol. 159. Translated by Garling, D.J.H. New York: Springer Science & Business Media. ISBN 978-3-642-64988-2. MR 0248498. OCLC 840293704.
  • Narici, Lawrence; Beckenstein, Edward (2011). Topological Vector Spaces. Pure and applied mathematics (Second ed.). Boca Raton, FL: CRC Press. ISBN 978-1584888666. OCLC 144216834.
  • Robertson, Alex P.; Robertson, Wendy J. (1980). Topological Vector Spaces. Cambridge Tracts in Mathematics. Vol. 53. Cambridge England: Cambridge University Press. ISBN 978-0-521-29882-7. OCLC 589250.
  • Rudin, Walter (1991). Functional Analysis. International Series in Pure and Applied Mathematics. Vol. 8 (Second ed.). New York, NY: McGraw-Hill Science/Engineering/Math. ISBN 978-0-07-054236-5. OCLC 21163277.
  • Schaefer, Helmut H.; Wolff, Manfred P. (1999). Topological Vector Spaces. GTM. Vol. 8 (Second ed.). New York, NY: Springer New York Imprint Springer. ISBN 978-1-4612-7155-0. OCLC 840278135.
  • Swartz, Charles (1992). An introduction to Functional Analysis. New York: M. Dekker. ISBN 978-0-8247-8643-4. OCLC 24909067.
  • Trèves, François (2006) [1967]. Topological Vector Spaces, Distributions and Kernels. Mineola, N.Y.: Dover Publications. ISBN 978-0-486-45352-1. OCLC 853623322.
  • Wilansky, Albert (2013). Modern Methods in Topological Vector Spaces. Mineola, New York: Dover Publications, Inc. ISBN 978-0-486-49353-4. OCLC 849801114.