Schauder basis

Last updated

In mathematics, a Schauder basis or countable basis is similar to the usual (Hamel) basis of a vector space; the difference is that Hamel bases use linear combinations that are finite sums, while for Schauder bases they may be infinite sums. This makes Schauder bases more suitable for the analysis of infinite-dimensional topological vector spaces including Banach spaces.

Contents

Schauder bases were described by Juliusz Schauder in 1927, [1] [2] although such bases were discussed earlier. For example, the Haar basis was given in 1909, and Georg Faber discussed in 1910 a basis for continuous functions on an interval, sometimes called a Faber–Schauder system. [3]

Definitions

Let V denote a topological vector space over the field  F. A Schauder basis is a sequence {bn} of elements of V such that for every element vV there exists a unique sequence {αn} of scalars in F so that

The convergence of the infinite sum is implicitly that of the ambient topology, i.e.,

but can be reduced to only weak convergence in a normed vector space (such as a Banach space). [4] Unlike a Hamel basis, the elements of the basis must be ordered, since the series may not converge unconditionally.

Note that some authors define Schauder bases to be countable (as above), while others use the term to include uncountable bases. In either case, the sums themselves always are countable. An uncountable Schauder basis is a linearly ordered set rather than a sequence, and each sum inherits the order of its terms from this linear ordering. They can and do arise in practice. As an example, a separable Hilbert space can only have a countable Schauder basis, but a non-separable Hilbert space may have an uncountable one.

Though the definition above technically does not require a normed space, a norm is necessary to say almost anything useful about Schauder bases. The results below assume the existence of a norm.

A Schauder basis {bn}n ≥ 0 is said to be normalized when all the basis vectors have norm 1 in the Banach space V.

A sequence {xn}n ≥ 0 in V is a basic sequence if it is a Schauder basis of its closed linear span.

Two Schauder bases, {bn} in V and {cn} in W, are said to be equivalent if there exist two constants c > 0 and C such that for every natural number N ≥ 0 and all sequences {αn} of scalars,

A family of vectors in V is total if its linear span (the set of finite linear combinations) is dense in V. If V is a Hilbert space, an orthogonal basis is a total subset B of V such that elements in B are nonzero and pairwise orthogonal. Further, when each element in B has norm 1, then B is an orthonormal basis of V.

Properties

Let {bn} be a Schauder basis of a Banach space V over F = R or C. It is a subtle consequence of the open mapping theorem that the linear mappings {Pn} defined by

are uniformly bounded by some constant C. [5] When C = 1, the basis is called a monotone basis. The maps {Pn} are the basis projections .

Let {b*n} denote the coordinate functionals, where b*n assigns to every vector v in V the coordinate αn of v in the above expansion. Each b*n is a bounded linear functional on V. Indeed, for every vector v in V,

These functionals {b*n} are called biorthogonal functionals associated to the basis {bn}. When the basis {bn} is normalized, the coordinate functionals {b*n} have norm ≤ 2C in the continuous dual V of V.

A Banach space with a Schauder basis is necessarily separable, but the converse is false. Since every vector v in a Banach space V with a Schauder basis is the limit of Pn(v), with Pn of finite rank and uniformly bounded, such a space V satisfies the bounded approximation property.

A theorem attributed to Mazur [6] asserts that every infinite-dimensional Banach space V contains a basic sequence, i.e., there is an infinite-dimensional subspace of V that has a Schauder basis. The basis problem is the question asked by Banach, whether every separable Banach space has a Schauder basis. This was negatively answered by Per Enflo who constructed a separable Banach space failing the approximation property, thus a space without a Schauder basis. [7]

Examples

The standard unit vector bases of c0, and of p for 1 p < ∞, are monotone Schauder bases. In this unit vector basis {bn}, the vector bn in V = c0 or in V = ℓp is the scalar sequence [bn, j]j where all coordinates bn, j are 0, except the nth coordinate:

where δn, j is the Kronecker delta. The space ℓ is not separable, and therefore has no Schauder basis.

Every orthonormal basis in a separable Hilbert space is a Schauder basis. Every countable orthonormal basis is equivalent to the standard unit vector basis in ℓ2.

The Haar system is an example of a basis for Lp([0, 1]), when 1 p < ∞. [2] When 1 <p <, another example is the trigonometric system defined below. The Banach space C([0, 1]) of continuous functions on the interval [0, 1], with the supremum norm, admits a Schauder basis. The FaberSchauder system is the most commonly used Schauder basis for C([0, 1]). [3] [8]

Several bases for classical spaces were discovered before Banach's book appeared (Banach (1932)), but some other cases remained open for a long time. For example, the question of whether the disk algebra A(D) has a Schauder basis remained open for more than forty years, until Bočkarev showed in 1974 that a basis constructed from the Franklin system exists in A(D). [9] One can also prove that the periodic Franklin system [10] is a basis for a Banach space Ar isomorphic to A(D). [11] This space Ar consists of all complex continuous functions on the unit circle T whose conjugate function is also continuous. The Franklin system is another Schauder basis for C([0, 1]), [12] and it is a Schauder basis in Lp([0, 1]) when 1 ≤ p<. [13] Systems derived from the Franklin system give bases in the space C1([0, 1]2) of differentiable functions on the unit square. [14] The existence of a Schauder basis in C1([0, 1]2) was a question from Banach's book. [15]

Relation to Fourier series

Let {xn} be, in the real case, the sequence of functions

or, in the complex case,

The sequence {xn} is called the trigonometric system. It is a Schauder basis for the space Lp([0, 2π]) for any p such that 1 <p<. For p = 2, this is the content of the Riesz–Fischer theorem, and for p  2, it is a consequence of the boundedness on the space Lp([0, 2π]) of the Hilbert transform on the circle. It follows from this boundedness that the projections PN defined by

are uniformly bounded on Lp([0, 2π]) when 1 <p<. This family of maps {PN} is equicontinuous and tends to the identity on the dense subset consisting of trigonometric polynomials. It follows that PNf tends to f in Lp-norm for every fLp([0, 2π]). In other words, {xn} is a Schauder basis of Lp([0, 2π]). [16]

However, the set {xn} is not a Schauder basis for L1([0, 2π]). This means that there are functions in L1 whose Fourier series does not converge in the L1 norm, or equivalently, that the projections PN are not uniformly bounded in L1-norm. Also, the set {xn} is not a Schauder basis for C([0, 2π]).

Bases for spaces of operators

The space K(ℓ2) of compact operators on the Hilbert space ℓ2 has a Schauder basis. For every x, y in ℓ2, let xy denote the rank one operator v ∈ ℓ2<v, x>y. If {en}n ≥ 1 is the standard orthonormal basis of ℓ2, a basis for K(ℓ2) is given by the sequence [17]

For every n, the sequence consisting of the n2 first vectors in this basis is a suitable ordering of the family {ejek}, for 1 ≤ j, kn.

The preceding result can be generalized: a Banach space X with a basis has the approximation property, so the space K(X) of compact operators on X is isometrically isomorphic [18] to the injective tensor product

If X is a Banach space with a Schauder basis {en}n ≥ 1 such that the biorthogonal functionals are a basis of the dual, that is to say, a Banach space with a shrinking basis, then the space K(X) admits a basis formed by the rank one operators e*jek : ve*j(v) ek, with the same ordering as before. [17] This applies in particular to every reflexive Banach space X with a Schauder basis

On the other hand, the space B(ℓ2) has no basis, since it is non-separable. Moreover, B(ℓ2) does not have the approximation property. [19]

Unconditionality

A Schauder basis {bn} is unconditional if whenever the series converges, it converges unconditionally. For a Schauder basis {bn}, this is equivalent to the existence of a constant C such that

for all natural numbers n, all scalar coefficients {αk} and all signs εk = ±1. Unconditionality is an important property since it allows one to forget about the order of summation. A Schauder basis is symmetric if it is unconditional and uniformly equivalent to all its permutations: there exists a constant C such that for every natural number n, every permutation π of the set {0, 1, ..., n}, all scalar coefficients {αk} and all signs k},

The standard bases of the sequence spaces c0 and ℓp for 1 p < ∞, as well as every orthonormal basis in a Hilbert space, are unconditional. These bases are also symmetric.

The trigonometric system is not an unconditional basis in Lp, except for p = 2.

The Haar system is an unconditional basis in Lp for any 1 < p < ∞. The space L1([0, 1]) has no unconditional basis. [20]

A natural question is whether every infinite-dimensional Banach space has an infinite-dimensional subspace with an unconditional basis. This was solved negatively by Timothy Gowers and Bernard Maurey in 1992. [21]

Schauder bases and duality

A basis {en}n≥0 of a Banach space X is boundedly complete if for every sequence {an}n≥0 of scalars such that the partial sums

are bounded in X, the sequence {Vn} converges in X. The unit vector basis for ℓp, 1 ≤ p < ∞, is boundedly complete. However, the unit vector basis is not boundedly complete in c0. Indeed, if an = 1 for every n, then

for every n, but the sequence {Vn} is not convergent in c0, since ||Vn+1Vn|| = 1 for every n.

A space X with a boundedly complete basis {en}n≥0 is isomorphic to a dual space, namely, the space X is isomorphic to the dual of the closed linear span in the dual X of the biorthogonal functionals associated to the basis {en}. [22]

A basis {en}n≥0 of X is shrinking if for every bounded linear functional f on X, the sequence of non-negative numbers

tends to 0 when n → ∞, where Fn is the linear span of the basis vectors em for mn. The unit vector basis for ℓp, 1 < p < ∞, or for c0, is shrinking. It is not shrinking in ℓ1: if f is the bounded linear functional on ℓ1 given by

then φnf(en) = 1 for every n.

A basis [en]n ≥ 0 of X is shrinking if and only if the biorthogonal functionals [e*n]n ≥ 0 form a basis of the dual X. [23]

Robert C. James characterized reflexivity in Banach spaces with basis: the space X with a Schauder basis is reflexive if and only if the basis is both shrinking and boundedly complete. [24] James also proved that a space with an unconditional basis is non-reflexive if and only if it contains a subspace isomorphic to c0 or ℓ1. [25]

A Hamel basis is a subset B of a vector space V such that every element v ∈ V can uniquely be written as

with αbF, with the extra condition that the set

is finite. This property makes the Hamel basis unwieldy for infinite-dimensional Banach spaces; as a Hamel basis for an infinite-dimensional Banach space has to be uncountable. (Every finite-dimensional subspace of an infinite-dimensional Banach space X has empty interior, and is nowhere dense in X. It then follows from the Baire category theorem that a countable union of bases of these finite-dimensional subspaces cannot serve as a basis. [26] )

See also

Notes

  1. see Schauder (1927).
  2. 1 2 Schauder, Juliusz (1928). "Eine Eigenschaft des Haarschen Orthogonalsystems". Mathematische Zeitschrift . 28: 317–320. doi:10.1007/bf01181164.
  3. 1 2 Faber, Georg (1910), "Über die Orthogonalfunktionen des Herrn Haar", Deutsche Math.-Ver (in German) 19: 104112. ISSN   0012-0456; http://www-gdz.sub.uni-goettingen.de/cgi-bin/digbib.cgi?PPN37721857X  ; http://resolver.sub.uni-goettingen.de/purl?GDZPPN002122553
  4. Karlin, S. (December 1948). "Bases in Banach spaces". Duke Mathematical Journal . 15 (4): 971–985. doi:10.1215/S0012-7094-48-01587-7. ISSN   0012-7094.
  5. see Theorem 4.10 in Fabian et al. (2011).
  6. for an early published proof, see p. 157, C.3 in Bessaga, C. and Pełczyński, A. (1958), "On bases and unconditional convergence of series in Banach spaces", Studia Math. 17: 151164. In the first lines of this article, Bessaga and Pełczyński write that Mazur's result appears without proof in Banach's book to be precise, on p. 238 but they do not provide a reference containing a proof.
  7. Enflo, Per (July 1973). "A counterexample to the approximation problem in Banach spaces". Acta Mathematica . 130 (1): 309–317. doi: 10.1007/BF02392270 .
  8. see pp. 4849 in Schauder (1927). Schauder defines there a general model for this system, of which the FaberSchauder system used today is a special case.
  9. see Bočkarev, S. V. (1974), "Existence of a basis in the space of functions analytic in the disc, and some properties of Franklin's system", (in Russian) Mat. Sb. (N.S.) 95(137): 318, 159. Translated in Math. USSR-Sb. 24 (1974), 1–16. The question is in Banach's book, Banach (1932) p. 238, §3.
  10. See p. 161, III.D.20 in Wojtaszczyk (1991).
  11. See p. 192, III.E.17 in Wojtaszczyk (1991).
  12. Franklin, Philip (1928). "A set of continuous orthogonal functions". Math. Ann. 100: 522–529. doi:10.1007/bf01448860.
  13. see p. 164, III.D.26 in Wojtaszczyk (1991).
  14. see Ciesielski, Z (1969). "A construction of basis in C1(I2)". Studia Math. 33: 243–247. and Schonefeld, Steven (1969). "Schauder bases in spaces of differentiable functions". Bull. Amer. Math. Soc. 75 (3): 586–590. doi: 10.1090/s0002-9904-1969-12249-4 .
  15. see p. 238, §3 in Banach (1932).
  16. see p. 40, II.B.11 in Wojtaszczyk (1991).
  17. 1 2 see Proposition 4.25, p. 88 in Ryan (2002).
  18. see Corollary 4.13, p. 80 in Ryan (2002).
  19. see Szankowski, Andrzej (1981). "B(H) does not have the approximation property". Acta Math. 147: 89–108. doi: 10.1007/bf02392870 .
  20. see p. 24 in Lindenstrauss & Tzafriri (1977).
  21. Gowers, W. Timothy; Maurey, Bernard (6 May 1992). "The unconditional basic sequence problem". arXiv: math/9205204 .
  22. see p. 9 in Lindenstrauss & Tzafriri (1977).
  23. see p. 8 in Lindenstrauss & Tzafriri (1977).
  24. See James (1950) and Lindenstrauss & Tzafriri (1977 , p. 9).
  25. See James (1950) and Lindenstrauss & Tzafriri (1977 , p. 23).
  26. Carothers, N. L. (2005), A short course on Banach space theory, Cambridge University Press ISBN   0-521-60372-2

This article incorporates material from Countable basis on PlanetMath, which is licensed under the Creative Commons Attribution/Share-Alike License.

Related Research Articles

In mathematics, more specifically in functional analysis, a Banach space 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.

In mathematics, any vector space has a corresponding dual vector space consisting of all linear forms on together with the vector space structure of pointwise addition and scalar multiplication by constants.

<span class="mw-page-title-main">Basis (linear algebra)</span> Set of vectors used to define coordinates

In mathematics, a set B of vectors in a vector space V is called a basis if every element of V may be written in a unique way as a finite linear combination of elements of B. The coefficients of this linear combination are referred to as components or coordinates of the vector with respect to B. The elements of a basis are called basis vectors.

In mathematics, an operator is generally a mapping or function that acts on elements of a space to produce elements of another space. There is no general definition of an operator, but the term is often used in place of function when the domain is a set of functions or other structured objects. Also, the domain of an operator is often difficult to characterize explicitly, and may be extended so as to act on related objects. See Operator (physics) for other examples.

In mathematics, the tensor product of two vector spaces V and W is a vector space to which is associated a bilinear map that maps a pair to an element of denoted

<span class="mw-page-title-main">Vector space</span> Algebraic structure in linear algebra

In mathematics and physics, a vector space is a set whose elements, often called vectors, may be added together and multiplied ("scaled") by numbers called scalars. Scalars are often real numbers, but can be complex numbers or, more generally, elements of any field. The operations of vector addition and scalar multiplication must satisfy certain requirements, called vector axioms. Real vector space and complex vector space are kinds of vector spaces based on different kinds of scalars: real coordinate space or complex coordinate space.

<span class="mw-page-title-main">Haar wavelet</span> First known wavelet basis

In mathematics, the Haar wavelet is a sequence of rescaled "square-shaped" functions which together form a wavelet family or basis. Wavelet analysis is similar to Fourier analysis in that it allows a target function over an interval to be represented in terms of an orthonormal basis. The Haar sequence is now recognised as the first known wavelet basis and is extensively used as a teaching example.

In abstract algebra, the direct sum is a construction which combines several modules into a new, larger module. The direct sum of modules is the smallest module which contains the given modules as submodules with no "unnecessary" constraints, making it an example of a coproduct. Contrast with the direct product, which is the dual notion.

In mathematics, specifically functional analysis, a trace-class operator is a linear operator for which a trace may be defined, such that the trace is a finite number independent of the choice of basis used to compute the trace. This trace of trace-class operators generalizes the trace of matrices studied in linear algebra. All trace-class operators are compact operators.

<span class="mw-page-title-main">Exterior algebra</span> Algebra of a vector space

In mathematics, the exterior algebra of a vector space is a graded associative algebra,

In mathematics, the rank, Prüfer rank, or torsion-free rank of an abelian group A is the cardinality of a maximal linearly independent subset. The rank of A determines the size of the largest free abelian group contained in A. If A is torsion-free then it embeds into a vector space over the rational numbers of dimension rank A. For finitely generated abelian groups, rank is a strong invariant and every such group is determined up to isomorphism by its rank and torsion subgroup. Torsion-free abelian groups of rank 1 have been completely classified. However, the theory of abelian groups of higher rank is more involved.

In mathematics, the modern component-free approach to the theory of a tensor views a tensor as an abstract object, expressing some definite type of multilinear concept. Their properties can be derived from their definitions, as linear maps or more generally; and the rules for manipulations of tensors arise as an extension of linear algebra to multilinear algebra.

In mathematics, more specifically in harmonic analysis, Walsh functions form a complete orthogonal set of functions that can be used to represent any discrete function—just like trigonometric functions can be used to represent any continuous function in Fourier analysis. They can thus be viewed as a discrete, digital counterpart of the continuous, analog system of trigonometric functions on the unit interval. But unlike the sine and cosine functions, which are continuous, Walsh functions are piecewise constant. They take the values −1 and +1 only, on sub-intervals defined by dyadic fractions.

<span class="mw-page-title-main">Approximation property</span> Mathematical concept

In mathematics, specifically functional analysis, a Banach space is said to have the approximation property (AP), if every compact operator is a limit of finite-rank operators. The converse is always true.

In functional analysis and related areas of mathematics, a sequence space is a vector space whose elements are infinite sequences of real or complex numbers. Equivalently, it is a function space whose elements are functions from the natural numbers to the field K of real or complex numbers. The set of all such functions is naturally identified with the set of all possible infinite sequences with elements in K, and can be turned into a vector space under the operations of pointwise addition of functions and pointwise scalar multiplication. All sequence spaces are linear subspaces of this space. Sequence spaces are typically equipped with a norm, or at least the structure of a topological vector space.

In mathematics, nuclear spaces are topological vector spaces that can be viewed as a generalization of finite dimensional Euclidean spaces and share many of their desirable properties. Nuclear spaces are however quite different from Hilbert spaces, another generalization of finite dimensional Euclidean spaces. They were introduced by Alexander Grothendieck.

In mathematics, there are usually many different ways to construct a topological tensor product of two topological vector spaces. For Hilbert spaces or nuclear spaces there is a simple well-behaved theory of tensor products, but for general Banach spaces or locally convex topological vector spaces the theory is notoriously subtle.

<span class="mw-page-title-main">Per Enflo</span> Swedish mathematician and concert pianist

Per H. Enflo is a Swedish mathematician working primarily in functional analysis, a field in which he solved problems that had been considered fundamental. Three of these problems had been open for more than forty years:

In functional analysis, a Markushevich basis is a biorthogonal system that is both complete and total.

In mathematics, the injective tensor product of two topological vector spaces (TVSs) was introduced by Alexander Grothendieck and was used by him to define nuclear spaces. An injective tensor product is in general not necessarily complete, so its completion is called the completed injective tensor products. Injective tensor products have applications outside of nuclear spaces. In particular, as described below, up to TVS-isomorphism, many TVSs that are defined for real or complex valued functions, for instance, the Schwartz space or the space of continuously differentiable functions, can be immediately extended to functions valued in a Hausdorff locally convex TVS without any need to extend definitions from real/complex-valued functions to -valued functions.

References

.

Further reading