A Banach space is both a vector space (over a normed field such as ) and a complete metric space, in a compatible way. Hence a complete normed vector space.
A source of simple Banach spaces comes from considering a Cartesian space (or where is the normed field) with the norm:
where (this doesn’t strictly make sense for , but taking the limit as and reading as the direct limit (as opposed to the inverse limit) we arrive at the formula ).
However, the theory of these spaces is not much more complicated than that of finite-dimensional vector spaces because they all have the same underlying topology. When we look at infinite-dimensional examples, however, things become trickier. Common examples are Lebesgue spaces, Hilbert spaces, and sequence spaces.
In the literature, one most often sees Banach spaces over the field of real numbers; Banach spaces over the field of complex numbers are not much different, since they are also over . But people do study them over p-adic numbers too. Unless otherwise stated, we assume below.
Let be a vector space over the field of real numbers. (One can generalise the choice of field somewhat.) A pseudonorm (or seminorm) on is a function
such that:
It follows from the above that ; in particular, . A norm is a pseudonorm that satisfies a converse to this: if .
A norm on is complete if, given any infinite sequence such that
there exists a (necessarily unique) sum such that
we write
(with the right-hand side undefined if no such sum exists).
Then a Banach space is simply a vector space equipped with a complete norm. As in the real line, we have in a Banach space that
with the left-hand side guaranteed to exist if the right-hand side exists as a finite real number (but the left-hand side may exist even if the right-hand side diverges, the usual distinction between absolute convergence and conditional convergence).
If we do not insist on the space being complete, we call it a normed (vector) space. If we have a topological vector space such that the topology comes from a norm, but we do not make an actual choice of such a norm, then we talk of a normable space.
The three axioms for a pseudonorm are very similar to the three axioms for a pseudometric.
Indeed, in any pseudonormed vector space, let the distance be
Then is a pseudometric, which is translation-invariant in that
always holds. Conversely, given any translation-invariant pseudometric on a vector space , let be
Then satisfies the axioms (1–3) for a pseudonorm, except that it may satisfy (2) only for . (In other words, it is only a G-pseudonorm.) It will actually be a pseudonorm iff the pseudometric satisfies a homogeneity rule:
Thus pseudonorms correspond precisely to homogeneous translation-invariant pseudometrics.
Similarly, norms correspond to homogenous translation-invariant metrics and complete norms correspond to complete homogeneous translation-invariant metrics. Indeed, (1) says that the sequence of partial sums is a Cauchy sequence, while (2) says that the sequence of partial sums converges to .
Thus a Banach space may equivalently be defined as a vector space equipped with a complete homogeneous translation-invariant metric. Actually, one usually sees a sort of hybrid approach: a Banach space is a normed vector space whose corresponding metric is complete.
If and are pseudonormed vector spaces, then the norm of a linear function may be defined in either of these equivalent ways:
(Some other forms are sometimes seen, but these may break down in degenerate cases.)
For finite-dimensional spaces, any linear map has a well-defined finite norm. In general, the following are equivalent:
In this case, we say that is bounded. If is not assumed to be linear, then the above conditions are no longer equivalent.
The bounded linear maps from to themselves form a pseudonormed vector space . This will be a Banach space if (and, except for degenerate cases of , only if) is a Banach space. In this way, the category of Banach spaces is a closed category with as the unit.
The clever reader will note that we have not yet defined as a category! (surprisingly in the nLab) There are many (nonequivalent) ways to do so.
In functional analysis, the usual notion of ‘isomorphism’ for Banach spaces is a bounded bijective linear map such that the inverse function (which is necessarily linear) is also bounded. In this case one can accept all bounded linear maps between Banach spaces as morphisms. Analysts sometimes refer to this as the “isomorphic category”.
Another natural notion of isomorphism is a surjective linear isometry. In this case, we take a morphism to be a short linear map, or linear contraction: a linear map such that . This category, which is what category theorists generally refer to as , is sometimes referred to as the “isometric category” by analysts. Note that this makes the ‘underlying set’ (in the sense of as a concrete category like any closed category) of a Banach space its (closed) unit ball
rather than the set of all vectors in (the underlying set of as a vector space).
Yemon Choi: This is really here to remind myself how to make query boxes. But while I’m at it, is it really OK to refer to the “unit ball functor” as “taking the underlying set”? I notice that on the discussion about internal homs at internal hom it is claimed that “Every closed category is a concrete category (represented by ), and the underlying set of the internal hom is the external hom” which seems to require “underlying set” to be interpreted in this looser sense.
Toby: Sure, but the point of putting ‘underlying set’ in scare quotes is precisely to point out that the category-theoretic underlying set is not what one would normally expect.
Mark Meckes: I’ve expanded this section in part to be consistent with analysts’ terminology. I’ve made some assumptions about category theorists’ conventions which might not be correct. (If I find time I might write about other categories of Banach spaces that analysts think about.)
Toby: Looks good to me!
From a category-theorist's perspective, the isomorphic category is really the full image of the inclusion functor from to (the category of topological vector spaces), which may be denoted . If you're working in , then you only care about the topological linear structure of your space (although you do also care that it can be derived from some metric); if you're working in , then you care about all of the structure on the space.
Since the forgetful functor , taking each Banach space to its closed unit ball, is faithful, it's natural to consider a Banach space as this ball with extra structure. We can always do this by abstract nonsense: define a Banach space to be a set equipped with a bijection to the closed unit ball of some complete normed vector space.
However, we can also do this in a more explicit way, constructing a Banach space as a structured set. The key is that, while a vector space supports any finite linear combinations (and a complete TVS supports some infinite ones), the unit ball of a Banach space supports linear combinations (finite or infinite) whose coefficients' absolute values sum to at most . (As a special case, this includes convex-linear combinations, where all coefficients are positive and sum to , so the unit ball is a convex space.)
So we define a Banach space to be set (the underlying set, or closed unit ball to be unambiguous, whose elements are called short vectors to distinguish them from the more general elements of the underlying vector space of ) equipped with, for each multiset of scalars whose absolute values sum to at most , an -indexed operation on , such that these operations are compatible as linear combinations, together with a function from to the unit interval satisfying homogeneity, definiteness, normalizability (defined below), and the infinitary triangle inequality that makes into a complete metric space.
To be more explicit:
Given a set (thought of as an index set), a short -sequence of scalars is a scalar-valued function on (whose values are written for in ) such that (with, in constructive mathematics, the classically redundant assumption that the left-hand side of this inequality exists). In other words, is in the sequence space .
Similarly, an -sequence of short vectors is a -valued function on (whose values are written for in ), with no requirements.
For each set , each short -sequence of scalars, and each -sequence of short vectors, we require a short vector . (Special cases include the empty sum , , and , but not .)
For each short vector , we require and (although the latter is redundant with a homogenous definite norm).
For each set , each -indexed family of sets (so is a set for each in ), each short -sequence of scalars, each -indexed family of short -sequences of scalars (so is a scalar for each in and in , and for each in ), and each -indexed family of -sequences of short vectors (so is an element of for each in and in ), we require (which makes sense since ). (This is what it means for the linear-combination operations to be compatible as linear combinations.)
For each short vector , we require a norm , a scalar ranging from to .
For each short scalar (so ) and each short vector , we require . (This is homogeneity.)
For each set , each short -sequence of scalars, and each -sequence of short vectors, we require . (This is the infinitary triangle inequality.)
For each short vector , if , then we require (this is definiteness).
For each short vector , if , then we require a short vector such that (this is what I'm calling normalizability). (By homogeneity, .)
For each short vector and short vector , although we cannot form , we can define the distance to be . (Then for each scalar , , using homogeneity.)
If is a Cauchy sequence (or net in constructive mathematics), we require that the sequence converges. (We automatically get from what has come before, but not every Cauchy sequence takes this form.)
We can then define a vector to be a formal scalar multiple of a short vector. Specifically, a vector is an equivalence class of pairs , where is a scalar and is a short vector, where is equivalent to iff for some nonzero scalar such that and , which without loss of generality can be . Similarly, we can define , where now or any other nonzero scalar such that . (More generally, we can define finitary linear combinations of vectors, although not always infinitary ones, even if the coefficients are a short sequence; that’s an advantage of the unit ball.) And we can define even more easily as , and . This makes the set of vectors into a Banach space in the usual sense, and its closed unit ball is equivalent to , consisting of all vectors of the form .
Many examples of Banach spaces are parametrised by an exponent . (Sometimes one can also try , but these generally don't give Banach spaces.)
The Cartesian space is a Banach space with
(We can allow by taking a limit; the result is that .) Every finite-dimensional Banach space is isomorphic to this for some and ; in fact, once you fix , the value of is irrelevant up to isomorphism.
The sequence space is the set of infinite sequences of real numbers such that
exists as a finite real number. (The only question is whether the sum converges. Again is a limit, with the result that .) Then is a Banach space with that norm. These are all versions of , but they are no longer isomorphic for different values of . (See isomorphism classes of Banach spaces.)
More generally, let be any set and let be the set of functions from to such that
exists as a finite real number. (Again, .) Then is a Banach space. (This example includes the previous examples, for a countable set.)
On any measure space , the Lebesgue space is the set of measurable almost-everywhere-defined real-valued functions on such that
exists as a finite real number. (Again, the only question is whether the integral converges. And again is a limit, with the result that is the essential supremum of .) As such, is a complete pseudonormed vector space; but we identify functions that are equal almost everywhere to make it into a Banach space. (This example includes the previous examples, for a set with counting measure.)
Any Hilbert space is Banach space; this includes all of the above examples for .
The category of Banach spaces is small complete, small cocomplete, and symmetric monoidal closed with respect to its standard internal hom (described at internal hom). Some details follow.
The category of Banach spaces admits small products. Given a small family of Banach spaces , its product in is the subspace of the vector-space product
consisting of -tuples which are uniformly bounded (i.e., there exists such that ), taking the least such upper bound as the norm of . This norm is called the -norm; in particular, the product of an -indexed family of copies of or is what is normally denoted as .
The category of Banach spaces admits equalizers. Indeed, the equalizer of a pair of maps in is the kernel of under the norm inherited from (the kernel is closed since is continuous, and is therefore complete). In fact every equalizer is even a section by the Hahn-Banach theorem. Every extremal monomorphism is even already an equalizer (and a section): Let be an extremal monomorphism, the embedding of into the codomain of and with restricted codomain. Since is an epimorphism, , and extremal, is an isomorphism, thus is an embedding.
The category of Banach spaces admits small coproducts. Given a small family of Banach spaces , its coproduct in is the completion of the vector space coproduct
with respect to the norm given by
where is finite and denotes the norm of an element in . This norm is called the -norm; in particular, the coproduct of an -indexed family of copies of or is what is normally denoted as .
The category of Banach spaces admits coequalizers. Though one may expect the coequalizer of a pair of maps to be the cokernel of under the quotient norm (in which the norm of a coset is the minimum norm attained by elements of ; here is the image ), these spaces are not Banach spaces in general, as the image of a map need not be closed (indeed, the inclusion has dense image in ), and so the quotient space may not be complete. However, the quotient by the closure of suffices.
To describe the tensor product of two Banach spaces (making symmetric monoidal closed with respect to its usual internal hom), let be the free vector space generated by the set , with norm on a typical element defined by
Let denote its completion with respect to this norm. Then take the cokernel of by the closure of the subspace spanned by the obvious bilinear relations. This quotient is .
In the literature on Banach spaces, tensor product above is usually called the projective tensor product of Banach spaces; see other tensor product of Banach spaces. The product and coproduct are considered direct sums; see other direct sums of Banach spaces.
To be described:
This paragraph describes some aspects of integration theory in Banach spaces that are relevant to understand the literature about AQFT. In the given context, elements of a Banach space are sometimes called vectors, a function or measure taking values in are therefore called vector functions and vector measures. Functions and measures taking values in the field that the Banach space is defined upon as a vector space are called scalar functions and scalar measures.
We will consider two types of integrals:
integrals of vector functions with respect to a scalar measure, specifically the Bochner integral,
integrals of scalar functions with respect to a vector measure, specifically the spectral integral of a normal operator on a Hilbert space.
The Bochner integral is a direct generalization of the Lesbegue integral to functions that take values in a Banach space. Whenever you encounter an integral of a function taking values in a Banach space in the AQFT literature, it is safe to assume that it is meant to be a Bochner integral. Two points already explained by Wikipedia are of interest:
reference: Joseph Diestel: “Sequences and Series in Banach Spaces” (ZMATH entry), chapter IV.
The integral with respect to the spectral measure of a bounded normal operator on a Hilbert space is an example of a Banach space integral with respect to a vector measure. In this paragraph we present a well known, but somewhat less often cited result, that is of use in some proofs in some approaches to AQFT, it is the version of the dominated convergence theorem for the given setting.
Let A be a bounded normal operator on a Hilbert space and E be it’s spectral measure (the “resolution of identity” in the terms of Dunford and Schwartz). Let be the spectrum of A. For a bounded complex Borel function f we then have
If the uniformly bounded sequence of complex Borel functions converges at each point of to the function , then in the strong operator topology.
See Dunford, Schwartz II, chapter X, corollary 8.
Every inductive limit of Banach spaces is a bornological vector space. (Alpay-Salomon 13, prop. 2.3)
Conversely, every bornological vector space is an inductive limit of normed spaces, and of Banach spaces if it is quasi-complete (Schaefer-Wolff 99)
Named after Stefan Banach.
Walter Rudin, Functional analysis
Dunford, Nelson; Schwartz, Jacob T.: “Linear operators. Part I: General theory.” (ZMATH entry), “Linear operators. Part II: Spectral theory, self adjoint operators in Hilbert space.” (ZMATH entry)
Z. Semadeni, Banach spaces of continuous functions, vol. I, Polish scientific publishers. Warszawa 1971
Daniel Alpay, Guy Salomon, On algebras which are inductive limits of Banach spaces (arXiv:1302.3372)
H. H. Schaefer with M. P. Wolff, Topological vector spaces, Springer 1999
Jiří Rosický, Are Banach spaces monadic?, (arXiv:2011.07543)
Last revised on March 11, 2024 at 08:20:21. See the history of this page for a list of all contributions to it.