Fréchet space

In functional analysis and related areas of mathematics, Fréchet spaces, named after Maurice Fréchet, are special topological vector spaces. They are generalizations of Banach spaces (normed vector spaces that are complete with respect to the metric induced by the norm). All Banach and Hilbert spaces are Fréchet spaces. Spaces of infinitely differentiable functions are typical examples of Fréchet spaces, many of which are typically not Banach spaces.

A Fréchet space X is defined to be a locally convex metrizable topological vector space (TVS) that is complete as a TVS,[1] meaning that every Cauchy sequence in X converges to some point in X (see footnote for more details).[note 1]

Important note: Not all authors require that a Fréchet space be locally convex (discussed below).

The topology of every Fréchet space is induced by some translation-invariant complete metric. Conversely, if the topology of a locally convex space X is induced by a translation-invariant complete metric then X is a Fréchet space.

Fréchet was the first to use the term "Banach space" and Banach in turn then coined the term "Fréchet space" to mean a complete metrizable topological vector space, without the local convexity requirement (such a space is today often called an "F-space").[1] The condition of locally convex was added later by Nicolas Bourbaki.[1] It's important to note that a sizable number of authors (e.g. Schaefer) use "F-space" to mean a (locally convex) Fréchet space while others do not require that a "Fréchet space" be locally convex. Moreover, some authors even use "F-space" and "Fréchet space" interchangeably. When reading mathematical literature, it is recommended that a reader always check whether the book's or article's definition of "F-space" and "Fréchet space" requires local convexity.[1]

Definitions

Fréchet spaces can be defined in two equivalent ways: the first employs a translation-invariant metric, the second a countable family of semi-norms.

Invariant metric definition

A topological vector space X is a Fréchet space if and only if it satisfies the following three properties:

  1. It is locally convex.[note 2]
  2. Its topology can be induced by a translation-invariant metric, i.e. a metric d: X × XR such that d(x, y) = d(x+a, y+a) for all a, x, yX. This means that a subset U of X is open if and only if for every uU there exists an ε > 0 such that { v : d(v, u) < ε } is a subset of U.
  3. Some (or equivalently, every) translation-invariant metric on X inducing the topology of X is complete.
    • Assuming that the other two conditions are satisfied, this condition is equivalent to X being a complete topological vector space, meaning that X is a complete uniform space when it is endowed with its canonical uniformity (this canonical uniformity is independent of any metric on X and is defined entirely in terms of vector subtraction and X's neighborhoods of the origin; moreover, the uniformity induced by any (topology-defining) translation invariant metric on X is identical to this canonical uniformity).

Note there is no natural notion of distance between two points of a Fréchet space: many different translation-invariant metrics may induce the same topology.

Countable family of semi-norms definition

The alternative and somewhat more practical definition is the following: a topological vector space X is a Fréchet space if and only if it satisfies the following three properties:

  1. It is a Hausdorff space,
  2. Its topology may be induced by a countable family of semi-norms k = 0, 1, 2, ... This means that a subset is open if and only if for every there exists and such that is a subset of
  3. it is complete with respect to the family of semi-norms.

A family of seminorms on yields a Hausdorff topology if and only if[2]

A sequence (xn) in X converges to x in the Fréchet space defined by a family of semi-norms if and only if it converges to x with respect to each of the given semi-norms.

As webbed Baire spaces

Theorem[3] (de Wilde 1978)  A topological vector space X is a Fréchet space if and only if it is both a webbed space and a Baire space.

Comparison to Banach spaces

In contrast to Banach spaces, the complete translation-invariant metric need not arise from a norm. The topology of a Fréchet space does, however, arise from both a total paranorm and an F-norm (the F stands for Fréchet).

Even though the topological structure of Fréchet spaces is more complicated than that of Banach spaces due to the potential lack of a norm, many important results in functional analysis, like the open mapping theorem, the closed graph theorem, and the Banach–Steinhaus theorem, still hold.

Constructing Fréchet spaces

Recall that a seminorm ǁ ⋅ ǁ is a function from a vector space X to the real numbers satisfying three properties. For all x, yX and all scalars c,

If ǁxǁ = 0 actually implies that x = 0, then ǁ ⋅ ǁ is in fact a norm. However, seminorms are useful in that they enable us to construct Fréchet spaces, as follows:

To construct a Fréchet space, one typically starts with a vector space X and defines a countable family of semi-norms ǁ ⋅ ǁk on X with the following two properties:

  • if xX and ǁxǁk = 0 for all k ≥ 0, then x = 0;
  • if (xn) is a sequence in X which is Cauchy with respect to each semi-norm ǁ ⋅ ǁk, then there exists xX such that (xn) converges to x with respect to each semi-norm ǁ ⋅ ǁk.

Then the topology induced by these seminorms (as explained above) turns X into a Fréchet space; the first property ensures that it is Hausdorff, and the second property ensures that it is complete. A translation-invariant complete metric inducing the same topology on X can then be defined by

The function uu/(1 + u) maps [0, ∞) monotonically to [0, 1), and so the above definition ensures that d(x, y) is "small" if and only if there exists K "large" such that ǁx - yǁk is "small" for k = 0, , K.

Examples

From pure functional analysis

  • Every Banach space is a Fréchet space, as the norm induces a translation-invariant metric and the space is complete with respect to this metric.
  • The space ω of all real valued sequences becomes a Fréchet space if we define the k-th semi-norm of a sequence to be the absolute value of the k-th element of the sequence. Convergence in this Fréchet space is equivalent to element-wise convergence.

From smooth manifolds

  • The vector space of all infinitely differentiable functions becomes a Fréchet space with the seminorms
    for every non-negative integer k. Here, ƒ(k) denotes the k-th derivative of ƒ, and ƒ(0) = ƒ. In this Fréchet space, a sequence of functions converges towards the element if and only if for every non-negative integer the sequence converges uniformly.
  • The vector space of all infinitely differentiable functions becomes a Fréchet space with the seminorms
    for all integers Then, a sequence of functions converges if and only if for every the sequences converge compactly.
  • The vector space of all -times continuously differentiable functions becomes a Fréchet space with the seminorms
    for all integers n ≥ 0 and k = 0, ..., m.
  • If M is a compact C-manifold and B is a Banach space, then the set C(M, B) of all infinitely-often differentiable functions ƒ : MB can be turned into a Fréchet space by using as seminorms the suprema of the norms of all partial derivatives. If M is a (not necessarily compact) C-manifold which admits a countable sequence Kn of compact subsets, so that every compact subset of M is contained in at least one Kn, then the spaces Cm(M, B) and C(M, B) are also Fréchet space in a natural manner. As a special case, every smooth finite-dimensional complete manifold M can be made into such a nested union of compact subsets: equip it with a Riemannian metric g which induces a metric d(x, y), choose xM, and let

    Let X be a compact C-manifold and V a vector bundle over X. Let C(X, V) denote the space of smooth sections of V over X. Choose Riemannian metrics and connections, which are guaranteed to exist, on the bundles TX and V. If s is a section, denote its jth covariant derivative by Djs. Then

    (where |⋅| is the norm induced by the Riemannian metric) is a family of seminorms making C(M, V) into a Fréchet space.

From holomorphicity

  • Let H be the space of entire (everywhere holomorphic) functions on the complex plane. Then the family of seminorms
    makes H into a Fréchet space.
  • Let 'H be the space of entire (everywhere holomorphic) functions of exponential type τ. Then the family of seminorms
    makes H into a Fréchet space.

Not all vector spaces with complete translation-invariant metrics are Fréchet spaces. An example is the space Lp([0, 1]) with p < 1. Although this space fails to be locally convex, it is an F-space.

Properties and further notions

If a Fréchet space admits a continuous norm, we can take all the seminorms to be norms by adding the continuous norm to each of them. A Banach space, C([a,b]), C(X, V) with X compact, and H all admit norms, while ω and C(ℝ) do not.

A closed subspace of a Fréchet space is a Fréchet space. A quotient of a Fréchet space by a closed subspace is a Fréchet space. The direct sum of a finite number of Fréchet spaces is a Fréchet space.

Several important tools of functional analysis which are based on the Baire category theorem remain true in Fréchet spaces; examples are the closed graph theorem and the open mapping theorem.

All Fréchet spaces are stereotype. In the theory of stereotype spaces Fréchet spaces are dual objects to Brauner spaces.

Every bounded linear operator from a Fréchet space into another topological vector space (TVS) is continuous.[4]

There exists a Fréchet space X having a bounded subset B and also a dense vector subspace M such that B is not contained in the closure (in X) of any bounded subset of M.[5]

If X is a non-normable Fréchet space on which there exists a continuous norm, then X contains a closed vector subspace that has no topological complement.[6]

All metrizable Montel spaces are separable.[7] A separable Fréchet space is a Montel space if and only if each weak-* convergent sequence in its continuous dual converges is strongly convergent.[7]

The following theorem implies that if X is a locally convex space then the topology of X can be a defined by a family of continuous norms on X (a norm is an injective seminorm) if and only if there exists at least one continuous norm on X.[8]

Theorem[8]  Let X be a Fréchet space over the field 𝕂. Then the following are equivalent:

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

Anderson–Kadec theorem

Anderson–Kadec theorem  Every infinite-dimensional, separable real Fréchet space is homeomorphic to , the Cartesian product of countably many copies of the real line .

Note that the homeomorphism described in the Anderson–Kadec theorem is not necessarily linear.

Eidelheit theorem  A Fréchet space is either isomorphic to a Banach space, or has a quotient space isomorphic to .

Differentiation of functions

If X and Y are Fréchet spaces, then the space consisting of all continuous linear maps from X to Y is not a Fréchet space in any natural manner. This is a major difference between the theory of Banach spaces and that of Fréchet spaces and necessitates a different definition for continuous differentiability of functions defined on Fréchet spaces, the Gateaux derivative:

Suppose X and Y are Fréchet spaces, U is an open subset of X, P: UY is a function, xU and hX. The map P is differentiable at x in the direction h if the limit

exists. The map P is said to be continuously differentiable in U if the map

is continuous. Since the product of Fréchet spaces is again a Fréchet space, we can then try to differentiate D(P) and define the higher derivatives of P in this fashion.

The derivative operator P : C([0,1]) → C([0,1]) defined by P(ƒ) = ƒ′ is itself infinitely differentiable. The first derivative is given by

for any two elements ƒ and h in C([0,1]). This is a major advantage of the Fréchet space C([0,1]) over the Banach space Ck([0,1]) for finite k.

If P : UY is a continuously differentiable function, then the differential equation

need not have any solutions, and even if does, the solutions need not be unique. This is in stark contrast to the situation in Banach spaces.

The inverse function theorem is not true in Fréchet spaces; a partial substitute is the Nash–Moser theorem.

Fréchet manifolds and Lie groups

One may define Fréchet manifolds as spaces that "locally look like" Fréchet spaces (just like ordinary manifolds are defined as spaces that locally look like Euclidean space n), and one can then extend the concept of Lie group to these manifolds. This is useful because for a given (ordinary) compact C manifold M, the set of all C diffeomorphisms ƒ: MM forms a generalized Lie group in this sense, and this Lie group captures the symmetries of M. Some of the relations between Lie algebras and Lie groups remain valid in this setting.

Another important example of a Fréchet Lie group is the loop group of a compact Lie group G, the smooth (C) mappings γ : S1G, multiplied pointwise by 1 γ2)(t) = γ1(t) γ2(t).[9][10]

Generalizations

If we drop the requirement for the space to be locally convex, we obtain F-spaces: vector spaces with complete translation-invariant metrics.

LF-spaces are countable inductive limits of Fréchet spaces.

See also

Notes

  1. Here "Cauchy" means Cauchy with respect to the canonical uniformity that every TVS possess. That is, a sequence x = (xm)
    m=1
    in a TVS X is Cauchy if and only if for all neighborhoods U of 0 in X, xm - xnU whenever m and n are sufficiently large. Note that this definition of a Cauchy sequence does not depend on any particular metric and doesn't even require that X be metrizable.
  2. Some authors do not include local convexity as part of the definition of a Fréchet space.

Citations

References

  • "Fréchet space", Encyclopedia of Mathematics, EMS Press, 2001 [1994]
  • Berberian, Sterling K. (1974). Lectures in Functional Analysis and Operator Theory. Graduate Texts in Mathematics. 15. New York: Springer. ISBN 978-0-387-90081-0. OCLC 878109401.
  • Bourbaki, Nicolas (1987) [1981]. Sur certains espaces vectoriels topologiques [Topological Vector Spaces: Chapters 1–5]. Annales de l'Institut Fourier. Éléments de mathématique. 2. Translated by Eggleston, H.G.; Madan, S. Berlin New York: Springer-Verlag. ISBN 978-3-540-42338-6. OCLC 17499190.
  • Conway, John (1990). A course in functional analysis. Graduate Texts in Mathematics. 96 (2nd ed.). New York: Springer-Verlag. ISBN 978-0-387-97245-9. OCLC 21195908.
  • 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.
  • Khaleelulla, S. M. (1982). Counterexamples in Topological Vector Spaces. Lecture Notes in Mathematics. 936. Berlin, Heidelberg, New York: Springer-Verlag. ISBN 978-3-540-11565-6. OCLC 8588370.
  • Köthe, Gottfried (1969). Topological Vector Spaces I. Grundlehren der mathematischen Wissenschaften. 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.
  • Pressley, Andrew; Segal, Graeme (1986). Loop groups. Oxford Mathematical Monographs. Oxford Science Publications. New York: Oxford University Press. ISBN 0-19-853535-X. MR 0900587.
  • Robertson, Alex P.; Robertson, Wendy J. (1980). Topological Vector Spaces. Cambridge Tracts in Mathematics. 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. 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. 8 (Second ed.). New York, NY: Springer New York Imprint Springer. ISBN 978-1-4612-7155-0. OCLC 840278135.
  • Sergeev, Armen (2010). Kähler Geometry of Loop Spaces. Mathematical Society of Japan Memoirs. 23. World Scientific Publishing. doi:10.1142/e023. ISBN 978-4-931469-60-0.
  • Adasch, Norbert; Ernst, Bruno; Keim, Dieter (1978). Topological Vector Spaces: The Theory Without Convexity Conditions. Lecture Notes in Mathematics. 639. Berlin New York: Springer-Verlag. ISBN 978-3-540-08662-8. OCLC 297140003.
  • 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.
This article is issued from Wikipedia. The text is licensed under Creative Commons - Attribution - Sharealike. Additional terms may apply for the media files.