Banach space

From Wikipedia the free encyclopedia

In mathematics, more specifically in functional analysis, a Banach space (pronounced [ˈbanax]) is a complete normed vector space. Thus, a Banach space is a vector space with a metric that allows the computation of vector length and distance between vectors and is complete in the sense that a Cauchy sequence of vectors always converges to a well-defined limit that is within the space.

Banach spaces are named after the Polish mathematician Stefan Banach, who introduced this concept and studied it systematically in 1920–1922 along with Hans Hahn and Eduard Helly.[1] Maurice René Fréchet was the first to use the term "Banach space" and Banach in turn then coined the term "Fréchet space".[2] Banach spaces originally grew out of the study of function spaces by Hilbert, Fréchet, and Riesz earlier in the century. Banach spaces play a central role in functional analysis. In other areas of analysis, the spaces under study are often Banach spaces.

Definition[edit]

A Banach space is a complete normed space A normed space is a pair[note 1] consisting of a vector space over a scalar field (where is commonly or ) together with a distinguished[note 2] norm Like all norms, this norm induces a translation invariant[note 3] distance function, called the canonical or (norm) induced metric, defined for all vectors by[note 4]

This makes into a metric space A sequence is called Cauchy in or -Cauchy or -Cauchy if for every real there exists some index such that
whenever and are greater than The normed space is called a Banach space and the canonical metric is called a complete metric if is a complete metric space, which by definition means for every Cauchy sequence in there exists some such that
where because this sequence's convergence to can equivalently be expressed as:

The norm of a normed space is called a complete norm if is a Banach space.

L-semi-inner product

For any normed space there exists an L-semi-inner product on such that for all ; in general, there may be infinitely many L-semi-inner products that satisfy this condition. L-semi-inner products are a generalization of inner products, which are what fundamentally distinguish Hilbert spaces from all other Banach spaces. This shows that all normed spaces (and hence all Banach spaces) can be considered as being generalizations of (pre-)Hilbert spaces.

Characterization in terms of series

The vector space structure allows one to relate the behavior of Cauchy sequences to that of converging series of vectors. A normed space is a Banach space if and only if each absolutely convergent series in converges in [3]

Topology[edit]

The canonical metric of a normed space induces the usual metric topology on which is referred to as the canonical or norm induced topology. Every normed space is automatically assumed to carry this Hausdorff topology, unless indicated otherwise. With this topology, every Banach space is a Baire space, although there exist normed spaces that are Baire but not Banach.[4] The norm is always a continuous function with respect to the topology that it induces.

The open and closed balls of radius centered at a point are, respectively, the sets

Any such ball is a convex and bounded subset of but a compact ball/neighborhood exists if and only if is a finite-dimensional vector space. In particular, no infinite–dimensional normed space can be locally compact or have the Heine–Borel property. If is a vector and is a scalar then
Using shows that this norm-induced topology is translation invariant, which means that for any and the subset is open (respectively, closed) in if and only if this is true of its translation Consequently, the norm induced topology is completely determined by any neighbourhood basis at the origin. Some common neighborhood bases at the origin include:
where is a sequence in of positive real numbers that converges to in (such as or for instance). So for example, every open subset of can be written as a union
indexed by some subset where every may be picked from the aforementioned sequence (the open balls can be replaced with closed balls, although then the indexing set and radii may also need to be replaced). Additionally, can always be chosen to be countable if is a separable space, which by definition means that contains some countable dense subset.

Homeomorphism classes of separable Banach spaces

All finite–dimensional normed spaces are separable Banach spaces and any two Banach spaces of the same finite dimension are linearly homeomorphic. Every separable infinite–dimensional Hilbert space is linearly isometrically isomorphic to the separable Hilbert sequence space with its usual norm

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).[5][6] Thus all infinite–dimensional separable Fréchet spaces are homeomorphic to each other (or said differently, their topology is unique up to a homeomorphism). Since every Banach space is a Fréchet space, this is also true of all infinite–dimensional separable Banach spaces, including In fact, is even homeomorphic to its own unit sphere which stands in sharp contrast to finite–dimensional spaces (the Euclidean plane is not homeomorphic to the unit circle, for instance).

This pattern in homeomorphism classes extends to generalizations of metrizable (locally Euclidean) topological manifolds known as metric Banach manifolds, which are metric spaces that are around every point, locally homeomorphic to some open subset of a given Banach space (metric Hilbert manifolds and metric Fréchet manifolds are defined similarly).[6] For example, every open subset of a Banach space is canonically a metric Banach manifold modeled on since the inclusion map is an open local homeomorphism. Using Hilbert space microbundles, David Henderson showed[7] in 1969 that every metric manifold modeled on a separable infinite–dimensional Banach (or Fréchet) space can be topologically embedded as an open subset of and, consequently, also admits a unique smooth structure making it into a Hilbert manifold.

Compact and convex subsets

There is a compact subset of whose convex hull is not closed and thus also not compact (see this footnote[note 5] for an example).[8] However, like in all Banach spaces, the closed convex hull of this (and every other) compact subset will be compact.[9] But if a normed space is not complete then it is in general not guaranteed that will be compact whenever is; an example[note 5] can even be found in a (non-complete) pre-Hilbert vector subspace of

As a topological vector space

This norm-induced topology also makes into what is known as a topological vector space (TVS), which by definition is a vector space endowed with a topology making the operations of addition and scalar multiplication continuous. It is emphasized that the TVS is only a vector space together with a certain type of topology; that is to say, when considered as a TVS, it is not associated with any particular norm or metric (both of which are "forgotten"). This Hausdorff TVS is even locally convex because the set of all open balls centered at the origin forms a neighbourhood basis at the origin consisting of convex balanced open sets. This TVS is also normable, which by definition refers to any TVS whose topology is induced by some (possibly unknown) norm. Normable TVSs are characterized by being Hausdorff and having a bounded convex neighborhood of the origin. All Banach spaces are barrelled spaces, which means that every barrel is neighborhood of the origin (all closed balls centered at the origin are barrels, for example) and guarantees that the Banach–Steinhaus theorem holds.

Comparison of complete metrizable vector topologies

The open mapping theorem implies that if and are topologies on that make both and into complete metrizable TVS (for example, Banach or Fréchet spaces) and if one topology is finer or coarser than the other then they must be equal (that is, if or then ).[10] So for example, if and are Banach spaces with topologies and and if one of these spaces has some open ball that is also an open subset of the other space (or equivalently, if one of or is continuous) then their topologies are identical and their norms are equivalent.

Completeness[edit]

Complete norms and equivalent norms

Two norms, and on a vector space are said to be equivalent if they induce the same topology;[11] this happens if and only if there exist positive real numbers such that for all If and are two equivalent norms on a vector space then is a Banach space if and only if is a Banach space. See this footnote for an example of a continuous norm on a Banach space that is not equivalent to that Banach space's given norm.[note 6][11] All norms on a finite-dimensional vector space are equivalent and every finite-dimensional normed space is a Banach space.[12]

Complete norms vs complete metrics

A metric on a vector space is induced by a norm on if and only if is translation invariant[note 3] and absolutely homogeneous, which means that for all scalars and all in which case the function defines a norm on and the canonical metric induced by is equal to

Suppose that is a normed space and that is the norm topology induced on Suppose that is any metric on such that the topology that induces on is equal to If is translation invariant[note 3] then is a Banach space if and only if is a complete metric space.[13] If is not translation invariant, then it may be possible for to be a Banach space but for to not be a complete metric space[14] (see this footnote[note 7] for an example). In contrast, a theorem of Klee,[15][16][note 8] which also applies to all metrizable topological vector spaces, implies that if there exists any[note 9] complete metric on that induces the norm topology on then is a Banach space.

A Fréchet space is a locally convex topological vector space whose topology is induced by some translation-invariant complete metric. Every Banach space is a Fréchet space but not conversely; indeed, there even exist Fréchet spaces on which no norm is a continuous function (such as the space of real sequences with the product topology). However, the topology of every Fréchet space is induced by some countable family of real-valued (necessarily continuous) maps called seminorms, which are generalizations of norms. It is even possible for a Fréchet space to have a topology that is induced by a countable family of norms (such norms would necessarily be continuous)[note 10][17] but to not be a Banach/normable space because its topology can not be defined by any single norm. An example of such a space is the Fréchet space whose definition can be found in the article on spaces of test functions and distributions.

Complete norms vs complete topological vector spaces

There is another notion of completeness besides metric completeness and that is the notion of a complete topological vector space (TVS) or TVS-completeness, which uses the theory of uniform spaces. Specifically, the notion of TVS-completeness uses a unique translation-invariant uniformity, called the canonical uniformity, that depends only on vector subtraction and the topology that the vector space is endowed with, and so in particular, this notion of TVS completeness is independent of whatever norm induced the topology (and even applies to TVSs that are not even metrizable). Every Banach space is a complete TVS. Moreover, a normed space is a Banach space (that is, its norm-induced metric is complete) if and only if it is complete as a topological vector space. If is a metrizable topological vector space (such as any norm induced topology, for example), then is a complete TVS if and only if it is a sequentially complete TVS, meaning that it is enough to check that every Cauchy sequence in converges in to some point of (that is, there is no need to consider the more general notion of arbitrary Cauchy nets).

If is a topological vector space whose topology is induced by some (possibly unknown) norm (such spaces are called normable), then is a complete topological vector space if and only if may be assigned a norm that induces on the topology and also makes into a Banach space. A Hausdorff locally convex topological vector space is normable if and only if its strong dual space is normable,[18] in which case is a Banach space ( denotes the strong dual space of whose topology is a generalization of the dual norm-induced topology on the continuous dual space ; see this footnote[note 11] for more details). If is a metrizable locally convex TVS, then is normable if and only if is a Fréchet–Urysohn space.[19] This shows that in the category of locally convex TVSs, Banach spaces are exactly those complete spaces that are both metrizable and have metrizable strong dual spaces.

Completions[edit]

Every normed space can be isometrically embedded onto a dense vector subspace of some Banach space, where this Banach space is called a completion of the normed space. This Hausdorff completion is unique up to isometric isomorphism.

More precisely, for every normed space there exist a Banach space and a mapping such that is an isometric mapping and is dense in If is another Banach space such that there is an isometric isomorphism from onto a dense subset of then is isometrically isomorphic to This Banach space is the Hausdorff completion of the normed space The underlying metric space for is the same as the metric completion of with the vector space operations extended from to The completion of is sometimes denoted by

General theory[edit]

Linear operators, isomorphisms[edit]

If and are normed spaces over the same ground field the set of all continuous -linear maps is denoted by In infinite-dimensional spaces, not all linear maps are continuous. A linear mapping from a normed space to another normed space is continuous if and only if it is bounded on the closed unit ball of Thus, the vector space can be given the operator norm

For a Banach space, the space is a Banach space with respect to this norm. In categorical contexts, it is sometimes convenient to restrict the function space between two Banach spaces to only the short maps; in that case the space reappears as a natural bifunctor.[20]

If is a Banach space, the space forms a unital Banach algebra; the multiplication operation is given by the composition of linear maps.

If and are normed spaces, they are isomorphic normed spaces if there exists a linear bijection such that and its inverse are continuous. If one of the two spaces or is complete (or reflexive, separable, etc.) then so is the other space. Two normed spaces and are isometrically isomorphic if in addition, is an isometry, that is, for every in The Banach–Mazur distance between two isomorphic but not isometric spaces and gives a measure of how much the two spaces and differ.

Continuous and bounded linear functions and seminorms[edit]

Every continuous linear operator is a bounded linear operator and if dealing only with normed spaces then the converse is also true. That is, a linear operator between two normed spaces is bounded if and only if it is a continuous function. So in particular, because the scalar field (which is or ) is a normed space, a linear functional on a normed space is a bounded linear functional if and only if it is a continuous linear functional. This allows for continuity-related results (like those below) to be applied to Banach spaces. Although boundedness is the same as continuity for linear maps between normed spaces, the term "bounded" is more commonly used when dealing primarily with Banach spaces.

If is a subadditive function (such as a norm, a sublinear function, or real linear functional), then[21] is continuous at the origin if and only if is uniformly continuous on all of ; and if in addition then is continuous if and only if its absolute value is continuous, which happens if and only if is an open subset of [21][note 12] And very importantly for applying the Hahn–Banach theorem, a linear functional is continuous if and only if this is true of its real part and moreover, and the real part completely determines which is why the Hahn–Banach theorem is often stated only for real linear functionals. Also, a linear functional on is continuous if and only if the seminorm is continuous, which happens if and only if there exists a continuous seminorm such that ; this last statement involving the linear functional and seminorm is encountered in many versions of the Hahn–Banach theorem.

Basic notions[edit]

The Cartesian product of two normed spaces is not canonically equipped with a norm. However, several equivalent norms are commonly used,[22] such as

which correspond (respectively) to the coproduct and product in the category of Banach spaces and short maps (discussed above).[20] For finite (co)products, these norms give rise to isomorphic normed spaces, and the product (or the direct sum ) is complete if and only if the two factors are complete.

If is a closed linear subspace of a normed space there is a natural norm on the quotient space

The quotient is a Banach space when is complete.[23] The quotient map from onto sending to its class is linear, onto and has norm except when in which case the quotient is the null space.

The closed linear subspace of is said to be a complemented subspace of if is the range of a surjective bounded linear projection In this case, the space is isomorphic to the direct sum of and the kernel of the projection

Suppose that and are Banach spaces and that There exists a canonical factorization of as[23]

where the first map is the quotient map, and the second map sends every class in the quotient to the image in This is well defined because all elements in the same class have the same image. The mapping is a linear bijection from onto the range whose inverse need not be bounded.

Classical spaces[edit]

Basic examples[24] of Banach spaces include: the Lp spaces and their special cases, the sequence spaces that consist of scalar sequences indexed by natural numbers ; among them, the space of absolutely summable sequences and the space of square summable sequences; the space of sequences tending to zero and the space of bounded sequences; the space of continuous scalar functions on a compact Hausdorff space equipped with the max norm,

According to the Banach–Mazur theorem, every Banach space is isometrically isomorphic to a subspace of some [25] For every separable Banach space there is a closed subspace of such that [26]

Any Hilbert space serves as an example of a Banach space. A Hilbert space on is complete for a norm of the form

where
is the inner product, linear in its first argument that satisfies the following:

For example, the space is a Hilbert space.

The Hardy spaces, the Sobolev spaces are examples of Banach spaces that are related to spaces and have additional structure. They are important in different branches of analysis, Harmonic analysis and Partial differential equations among others.

Banach algebras[edit]

A Banach algebra is a Banach space over or together with a structure of algebra over , such that the product map is continuous. An equivalent norm on can be found so that for all

Examples[edit]

  • The Banach space with the pointwise product, is a Banach algebra.
  • The disk algebra consists of functions holomorphic in the open unit disk and continuous on its closure: Equipped with the max norm on the disk algebra is a closed subalgebra of
  • The Wiener algebra is the algebra of functions on the unit circle with absolutely convergent Fourier series. Via the map associating a function on to the sequence of its Fourier coefficients, this algebra is isomorphic to the Banach algebra where the product is the convolution of sequences.
  • For every Banach space the space of bounded linear operators on with the composition of maps as product, is a Banach algebra.
  • A C*-algebra is a complex Banach algebra with an antilinear involution such that The space of bounded linear operators on a Hilbert space is a fundamental example of C*-algebra. The Gelfand–Naimark theorem states that every C*-algebra is isometrically isomorphic to a C*-subalgebra of some The space of complex continuous functions on a compact Hausdorff space is an example of commutative C*-algebra, where the involution associates to every function its complex conjugate

Dual space[edit]

If is a normed space and the underlying field (either the real or the complex numbers), the continuous dual space is the space of continuous linear maps from into or continuous linear functionals. The notation for the continuous dual is in this article.[27] Since is a Banach space (using the absolute value as norm), the dual is a Banach space, for every normed space The Dixmier–Ng theorem characterizes the dual spaces of Banach spaces.

The main tool for proving the existence of continuous linear functionals is the Hahn–Banach theorem.

Hahn–Banach theorem — Let be a vector space over the field Let further

  • be a linear subspace,
  • be a sublinear function and
  • be a linear functional so that for all

Then, there exists a linear functional so that

In particular, every continuous linear functional on a subspace of a normed space can be continuously extended to the whole space, without increasing the norm of the functional.[28] An important special case is the following: for every vector in a normed space there exists a continuous linear functional on such that

When is not equal to the vector, the functional must have norm one, and is called a norming functional for

The Hahn–Banach separation theorem states that two disjoint non-empty convex sets in a real Banach space, one of them open, can be separated by a closed affine hyperplane. The open convex set lies strictly on one side of the hyperplane, the second convex set lies on the other side but may touch the hyperplane.[29]

A subset in a Banach space is total if the linear span of is dense in The subset is total in if and only if the only continuous linear functional that vanishes on is the functional: this equivalence follows from the Hahn–Banach theorem.

If is the direct sum of two closed linear subspaces and then the dual of is isomorphic to the direct sum of the duals of and [30] If is a closed linear subspace in one can associate the orthogonal of in the dual,

The orthogonal is a closed linear subspace of the dual. The dual of is isometrically isomorphic to The dual of is isometrically isomorphic to [31]

The dual of a separable Banach space need not be separable, but:

Theorem[32] — Let be a normed space. If is separable, then is separable.

When is separable, the above criterion for totality can be used for proving the existence of a countable total subset in

Weak topologies[edit]

The weak topology on a Banach space is the coarsest topology on for which all elements in the continuous dual space are continuous. The norm topology is therefore finer than the weak topology. It follows from the Hahn–Banach separation theorem that the weak topology is Hausdorff, and that a norm-closed convex subset of a Banach space is also weakly closed.[33] A norm-continuous linear map between two Banach spaces and is also weakly continuous, that is, continuous from the weak topology of to that of [34]

If is infinite-dimensional, there exist linear maps which are not continuous. The space of all linear maps from to the underlying field (this space is called the algebraic dual space, to distinguish it from also induces a topology on which is finer than the weak topology, and much less used in functional analysis.

On a dual space there is a topology weaker than the weak topology of called weak* topology. It is the coarsest topology on for which all evaluation maps where ranges over are continuous. Its importance comes from the Banach–Alaoglu theorem.

Banach–Alaoglu theorem — Let be a normed vector space. Then the closed unit ball of the dual space is compact in the weak* topology.

The Banach–Alaoglu theorem can be proved using Tychonoff's theorem about infinite products of compact Hausdorff spaces. When is separable, the unit ball of the dual is a metrizable compact in the weak* topology.[35]

Examples of dual spaces[edit]

The dual of is isometrically isomorphic to : for every bounded linear functional on there is a unique element such that

The dual of is isometrically isomorphic to . The dual of Lebesgue space is isometrically isomorphic to when and

For every vector in a Hilbert space the mapping

defines a continuous linear functional on The Riesz representation theorem states that every continuous linear functional on is of the form for a uniquely defined vector in The mapping is an antilinear isometric bijection from onto its dual When the scalars are real, this map is an isometric isomorphism.

When is a compact Hausdorff topological space, the dual of is the space of Radon measures in the sense of Bourbaki.[36] The subset of consisting of non-negative measures of mass 1 (probability measures) is a convex w*-closed subset of the unit ball of The extreme points of are the Dirac measures on The set of Dirac measures on equipped with the w*-topology, is homeomorphic to

Banach–Stone Theorem — If and are compact Hausdorff spaces and if and are isometrically isomorphic, then the topological spaces and are homeomorphic.[37][38]

The result has been extended by Amir[39] and Cambern[40] to the case when the multiplicative Banach–Mazur distance between and is The theorem is no longer true when the distance is [41]

In the commutative Banach algebra the maximal ideals are precisely kernels of Dirac measures on

More generally, by the Gelfand–Mazur theorem, the maximal ideals of a unital commutative Banach algebra can be identified with its characters—not merely as sets but as topological spaces: the former with the hull-kernel topology and the latter with the w*-topology. In this identification, the maximal ideal space can be viewed as a w*-compact subset of the unit ball in the dual

Theorem — If is a compact Hausdorff space, then the maximal ideal space of the Banach algebra is homeomorphic to [37]

Not every unital commutative Banach algebra is of the form for some compact Hausdorff space However, this statement holds if one places in the smaller category of commutative C*-algebras. Gelfand's representation theorem for commutative C*-algebras states that every commutative unital C*-algebra is isometrically isomorphic to a space.[42] The Hausdorff compact space here is again the maximal ideal space, also called the spectrum of in the C*-algebra context.

Bidual[edit]

If is a normed space, the (continuous) dual of the dual is called bidual, or second dual of For every normed space there is a natural map,

This defines as a continuous linear functional on that is, an element of The map is a linear map from to As a consequence of the existence of a norming functional for every this map is isometric, thus injective.

For example, the dual of is identified with and the dual of is identified with the space of bounded scalar sequences. Under these identifications, is the inclusion map from to It is indeed isometric, but not onto.

If is surjective, then the normed space is called reflexive (see below). Being the dual of a normed space, the bidual is complete, therefore, every reflexive normed space is a Banach space.

Using the isometric embedding it is customary to consider a normed space as a subset of its bidual. When is a Banach space, it is viewed as a closed linear subspace of If is not reflexive, the unit ball of is a proper subset of the unit ball of The Goldstine theorem states that the unit ball of a normed space is weakly*-dense in the unit ball of the bidual. In other words, for every in the bidual, there exists a net in so that