Locally integrable function

Last updated

In mathematics, a locally integrable function (sometimes also called locally summable function) [1] is a function which is integrable (so its integral is finite) on every compact subset of its domain of definition. The importance of such functions lies in the fact that their function space is similar to Lp spaces, but its members are not required to satisfy any growth restriction on their behavior at the boundary of their domain (at infinity if the domain is unbounded): in other words, locally integrable functions can grow arbitrarily fast at the domain boundary, but are still manageable in a way similar to ordinary integrable functions.

Contents

Definition

Standard definition

Definition 1. [2] Let Ω be an open set in the Euclidean space and f : Ω → be a Lebesgue measurable function. If f on Ω is such that

i.e. its Lebesgue integral is finite on all compact subsets K of Ω, [3] then f is called locally integrable. The set of all such functions is denoted by L1,loc(Ω):

where denotes the restriction of f to the set K.

The classical definition of a locally integrable function involves only measure theoretic and topological [4] concepts and can be carried over abstract to complex-valued functions on a topological measure space (X, Σ, μ): [5] however, since the most common application of such functions is to distribution theory on Euclidean spaces, [2] all the definitions in this and the following sections deal explicitly only with this important case.

An alternative definition

Definition 2. [6] Let Ω be an open set in the Euclidean space . Then a function f : Ω → such that

for each test function φC 
c
 
(Ω)
is called locally integrable, and the set of such functions is denoted by L1,loc(Ω). Here C 
c
 
(Ω)
denotes the set of all infinitely differentiable functions φ : Ω → with compact support contained in Ω.

This definition has its roots in the approach to measure and integration theory based on the concept of continuous linear functional on a topological vector space, developed by the Nicolas Bourbaki school: [7] it is also the one adopted by Strichartz (2003) and by Maz'ya & Shaposhnikova (2009 , p. 34). [8] This "distribution theoretic" definition is equivalent to the standard one, as the following lemma proves:

Lemma 1. A given function f : Ω → is locally integrable according to Definition 1 if and only if it is locally integrable according to Definition 2 , i.e.

Proof of Lemma 1

If part: Let φC 
c
 
(Ω)
be a test function. It is bounded by its supremum norm ||φ||, measurable, and has a compact support, let's call it K. Hence

by Definition 1 .

Only if part: Let K be a compact subset of the open set Ω. We will first construct a test function φKC 
c
 
(Ω)
which majorises the indicator function χK of K. The usual set distance [9] between K and the boundary ∂Ω is strictly greater than zero, i.e.

hence it is possible to choose a real number δ such that Δ > 2δ > 0 (if ∂Ω is the empty set, take Δ = ∞). Let Kδ and K2δ denote the closed δ-neighborhood and 2δ-neighborhood of K, respectively. They are likewise compact and satisfy

Now use convolution to define the function φK : Ω → by

where φδ is a mollifier constructed by using the standard positive symmetric one. Obviously φK is non-negative in the sense that φK ≥ 0, infinitely differentiable, and its support is contained in K2δ, in particular it is a test function. Since φK(x) = 1 for all xK, we have that χKφK.

Let f be a locally integrable function according to Definition 2 . Then

Since this holds for every compact subset K of Ω, the function f is locally integrable according to Definition 1 . □

Generalization: locally p-integrable functions

Definition 3. [10] Let Ω be an open set in the Euclidean space and f : Ω → be a Lebesgue measurable function. If, for a given p with 1 ≤ p ≤ +∞, f satisfies

i.e., it belongs to Lp(K) for all compact subsets K of Ω, then f is called locallyp-integrable or also p-locally integrable. [10] The set of all such functions is denoted by Lp,loc(Ω):

An alternative definition, completely analogous to the one given for locally integrable functions, can also be given for locally p-integrable functions: it can also be and proven equivalent to the one in this section. [11] Despite their apparent higher generality, locally p-integrable functions form a subset of locally integrable functions for every p such that 1 < p ≤ +∞. [12]

Notation

Apart from the different glyphs which may be used for the uppercase "L", [13] there are few variants for the notation of the set of locally integrable functions

Properties

Lp,loc is a complete metric space for all p ≥ 1

Theorem 1. [14] Lp,loc is a complete metrizable space: its topology can be generated by the following metric:

where {ωk}k≥1 is a family of non empty open sets such that

In references ( Gilbarg & Trudinger 2001 , p. 147), ( Maz'ya & Poborchi 1997 , p. 5), ( Maz'ja 1985 , p. 6) and ( Maz'ya 2011 , p. 2), this theorem is stated but not proved on a formal basis: [15] a complete proof of a more general result, which includes it, is found in ( Meise & Vogt 1997 , p. 40).

Lp is a subspace of L1,loc for all p ≥ 1

Theorem 2. Every function f belonging to Lp(Ω), 1 ≤ p ≤ +∞, where Ω is an open subset of , is locally integrable.

Proof. The case p = 1 is trivial, therefore in the sequel of the proof it is assumed that 1 < p ≤ +∞. Consider the characteristic function χK of a compact subset K of Ω: then, for p ≤ +∞,

where

Then for any f belonging to Lp(Ω), by Hölder's inequality, the product K is integrable i.e. belongs to L1(Ω) and

therefore

Note that since the following inequality is true

the theorem is true also for functions f belonging only to the space of locally p-integrable functions, therefore the theorem implies also the following result.

Corollary 1. Every function in , , is locally integrable, i. e. belongs to .

Note: If is an open subset of that is also bounded, then one has the standard inclusion which makes sense given the above inclusion . But the first of these statements is not true if is not bounded; then it is still true that for any , but not that . To see this, one typically considers the function , which is in but not in for any finite .

L1,loc is the space of densities of absolutely continuous measures

Theorem 3. A function f is the density of an absolutely continuous measure if and only if .

The proof of this result is sketched by ( Schwartz 1998 , p. 18). Rephrasing its statement, this theorem asserts that every locally integrable function defines an absolutely continuous measure and conversely that every absolutely continuous measures defines a locally integrable function: this is also, in the abstract measure theory framework, the form of the important Radon–Nikodym theorem given by Stanisław Saks in his treatise. [16]

Examples

is not locally integrable in x = 0: it is indeed locally integrable near this point since its integral over every compact set not including it is finite. Formally speaking, : [19] however, this function can be extended to a distribution on the whole as a Cauchy principal value. [20]
does not define any distribution on . [21]
where k1 and k2 are complex constants, is a general solution of the following elementary non-Fuchsian differential equation of first order
Again it does not define any distribution on the whole , if k1 or k2 are not zero: the only distributional global solution of such equation is therefore the zero distribution, and this shows how, in this branch of the theory of differential equations, the methods of the theory of distributions cannot be expected to have the same success achieved in other branches of the same theory, notably in the theory of linear differential equations with constant coefficients. [22]

Applications

Locally integrable functions play a prominent role in distribution theory and they occur in the definition of various classes of functions and function spaces, like functions of bounded variation. Moreover, they appear in the Radon–Nikodym theorem by characterizing the absolutely continuous part of every measure.

See also

Notes

  1. According to Gel'fand & Shilov (1964 , p. 3).
  2. 1 2 See for example ( Schwartz 1998 , p. 18) and ( Vladimirov 2002 , p. 3).
  3. Another slight variant of this definition, chosen by Vladimirov (2002 , p. 1), is to require only that K ⋐ Ω (or, using the notation of Gilbarg & Trudinger (2001 , p. 9), K ⊂⊂ Ω), meaning that Kis strictly included inΩ i.e. it is a set having compact closure strictly included in the given ambient set.
  4. The notion of compactness must obviously be defined on the given abstract measure space.
  5. This is the approach developed for example by Cafiero (1959 , pp. 285–342) and by Saks (1937 , chapter I), without dealing explicitly with the locally integrable case.
  6. See for example ( Strichartz 2003 , pp. 12–13).
  7. This approach was praised by Schwartz (1998 , pp. 16–17) who remarked also its usefulness, however using Definition 1 to define locally integrable functions.
  8. Be noted that Maz'ya and Shaposhnikova define explicitly only the "localized" version of the Sobolev space Wk,p(Ω), nevertheless explicitly asserting that the same method is used to define localized versions of all other Banach spaces used in the cited book: in particular, Lp,loc(Ω) is introduced on page 44.
  9. Not to be confused with the Hausdorff distance.
  10. 1 2 See for example ( Vladimirov 2002 , p. 3) and ( Maz'ya & Poborchi 1997 , p. 4).
  11. As remarked in the previous section, this is the approach adopted by Maz'ya & Shaposhnikova (2009), without developing the elementary details.
  12. Precisely, they form a vector subspace of L1,loc(Ω): see Corollary 1 to Theorem 2 .
  13. See for example ( Vladimirov 2002 , p. 3), where a calligraphic is used.
  14. See ( Gilbarg & Trudinger 2001 , p. 147), ( Maz'ya & Poborchi 1997 , p. 5) for a statement of this results, and also the brief notes in ( Maz'ja 1985 , p. 6) and ( Maz'ya 2011 , p. 2).
  15. Gilbarg & Trudinger (2001 , p. 147) and Maz'ya & Poborchi (1997 , p. 5) only sketch very briefly the method of proof, while in ( Maz'ja 1985 , p. 6) and ( Maz'ya 2011 , p. 2) it is assumed as a known result, from which the subsequent development starts.
  16. According to Saks (1937 , p. 36), "If E is a set of finite measure, or, more generally the sum of a sequence of sets of finite measure (μ), then, in order that an additive function of a set (𝔛) on E be absolutely continuous on E, it is necessary and sufficient that this function of a set be the indefinite integral of some integrable function of a point of E". Assuming (μ) to be the Lebesgue measure, the two statements can be seen to be equivalent.
  17. See for example ( Hörmander 1990 , p. 37).
  18. See ( Strichartz 2003 , p. 12).
  19. See ( Schwartz 1998 , p. 19).
  20. See ( Vladimirov 2002 , pp. 19–21).
  21. See ( Vladimirov 2002 , p. 21).
  22. For a brief discussion of this example, see ( Schwartz 1998 , pp. 131–132).

Related Research Articles

In mathematics, the Lp spaces are function spaces defined using a natural generalization of the p-norm for finite-dimensional vector spaces. They are sometimes called Lebesgue spaces, named after Henri Lebesgue, although according to the Bourbaki group they were first introduced by Frigyes Riesz.

<span class="mw-page-title-main">Fourier transform</span> Mathematical transform that expresses a function of time as a function of frequency

In physics, engineering and mathematics, the Fourier transform (FT) is an integral transform that takes as input a function and outputs another function that describes the extent to which various frequencies are present in the original function. The output of the transform is a complex-valued function of frequency. The term Fourier transform refers to both this complex-valued function and the mathematical operation. When a distinction needs to be made the Fourier transform is sometimes called the frequency domain representation of the original function. The Fourier transform is analogous to decomposing the sound of a musical chord into the intensities of its constituent pitches.

In vector calculus, the divergence theorem, also known as Gauss's theorem or Ostrogradsky's theorem, is a theorem relating the flux of a vector field through a closed surface to the divergence of the field in the volume enclosed.

In mathematical analysis, a function of bounded variation, also known as BV function, is a real-valued function whose total variation is bounded (finite): the graph of a function having this property is well behaved in a precise sense. For a continuous function of a single variable, being of bounded variation means that the distance along the direction of the y-axis, neglecting the contribution of motion along x-axis, traveled by a point moving along the graph has a finite value. For a continuous function of several variables, the meaning of the definition is the same, except for the fact that the continuous path to be considered cannot be the whole graph of the given function, but can be every intersection of the graph itself with a hyperplane parallel to a fixed x-axis and to the y-axis.

In mathematics and signal processing, the Hilbert transform is a specific singular integral that takes a function, u(t) of a real variable and produces another function of a real variable H(u)(t). The Hilbert transform is given by the Cauchy principal value of the convolution with the function (see § Definition). The Hilbert transform has a particularly simple representation in the frequency domain: It imparts a phase shift of ±90° (π/2 radians) to every frequency component of a function, the sign of the shift depending on the sign of the frequency (see § Relationship with the Fourier transform). The Hilbert transform is important in signal processing, where it is a component of the analytic representation of a real-valued signal u(t). The Hilbert transform was first introduced by David Hilbert in this setting, to solve a special case of the Riemann–Hilbert problem for analytic functions.

In mathematics, a Sobolev space is a vector space of functions equipped with a norm that is a combination of Lp-norms of the function together with its derivatives up to a given order. The derivatives are understood in a suitable weak sense to make the space complete, i.e. a Banach space. Intuitively, a Sobolev space is a space of functions possessing sufficiently many derivatives for some application domain, such as partial differential equations, and equipped with a norm that measures both the size and regularity of a function.

In mathematics, the total variation identifies several slightly different concepts, related to the (local or global) structure of the codomain of a function or a measure. For a real-valued continuous function f, defined on an interval [a, b] ⊂ R, its total variation on the interval of definition is a measure of the one-dimensional arclength of the curve with parametric equation xf(x), for x ∈ [a, b]. Functions whose total variation is finite are called functions of bounded variation.

In probability theory and related fields, Malliavin calculus is a set of mathematical techniques and ideas that extend the mathematical field of calculus of variations from deterministic functions to stochastic processes. In particular, it allows the computation of derivatives of random variables. Malliavin calculus is also called the stochastic calculus of variations. P. Malliavin first initiated the calculus on infinite dimensional space. Then, the significant contributors such as S. Kusuoka, D. Stroock, J-M. Bismut, Shinzo Watanabe, I. Shigekawa, and so on finally completed the foundations.

In mathematics, the stationary phase approximation is a basic principle of asymptotic analysis, applying to functions given by integration against a rapidly-varying complex exponential.

In mathematics, more particularly in functional analysis, differential topology, and geometric measure theory, a k-current in the sense of Georges de Rham is a functional on the space of compactly supported differential k-forms, on a smooth manifold M. Currents formally behave like Schwartz distributions on a space of differential forms, but in a geometric setting, they can represent integration over a submanifold, generalizing the Dirac delta function, or more generally even directional derivatives of delta functions (multipoles) spread out along subsets of M.

In complex analysis, functional analysis and operator theory, a Bergman space, named after Stefan Bergman, is a function space of holomorphic functions in a domain D of the complex plane that are sufficiently well-behaved at the boundary that they are absolutely integrable. Specifically, for 0 < p < ∞, the Bergman space Ap(D) is the space of all holomorphic functions in D for which the p-norm is finite:

In mathematics, the Bochner integral, named for Salomon Bochner, extends the definition of Lebesgue integral to functions that take values in a Banach space, as the limit of integrals of simple functions.

In mathematics, Weyl's lemma, named after Hermann Weyl, states that every weak solution of Laplace's equation is a smooth solution. This contrasts with the wave equation, for example, which has weak solutions that are not smooth solutions. Weyl's lemma is a special case of elliptic or hypoelliptic regularity.

In mathematics, a Caccioppoli set is a set whose boundary is measurable and has a finite measure. A synonym is set of (locally) finite perimeter. Basically, a set is a Caccioppoli set if its characteristic function is a function of bounded variation.

In mathematics, uniform integrability is an important concept in real analysis, functional analysis and measure theory, and plays a vital role in the theory of martingales.

In mathematics, the trace operator extends the notion of the restriction of a function to the boundary of its domain to "generalized" functions in a Sobolev space. This is particularly important for the study of partial differential equations with prescribed boundary conditions, where weak solutions may not be regular enough to satisfy the boundary conditions in the classical sense of functions.

In mathematics, the Pettis integral or Gelfand–Pettis integral, named after Israel M. Gelfand and Billy James Pettis, extends the definition of the Lebesgue integral to vector-valued functions on a measure space, by exploiting duality. The integral was introduced by Gelfand for the case when the measure space is an interval with Lebesgue measure. The integral is also called the weak integral in contrast to the Bochner integral, which is the strong integral.

In functional analysis, the Fréchet–Kolmogorov theorem gives a necessary and sufficient condition for a set of functions to be relatively compact in an Lp space. It can be thought of as an Lp version of the Arzelà–Ascoli theorem, from which it can be deduced. The theorem is named after Maurice René Fréchet and Andrey Kolmogorov.

In geometry, a valuation is a finitely additive function from a collection of subsets of a set to an abelian semigroup. For example, Lebesgue measure is a valuation on finite unions of convex bodies of Other examples of valuations on finite unions of convex bodies of are surface area, mean width, and Euler characteristic.

In the mathematical discipline of functional analysis, a differentiable vector-valued function from Euclidean space is a differentiable function valued in a topological vector space (TVS) whose domains is a subset of some finite-dimensional Euclidean space. It is possible to generalize the notion of derivative to functions whose domain and codomain are subsets of arbitrary topological vector spaces (TVSs) in multiple ways. But when the domain of a TVS-valued function is a subset of a finite-dimensional Euclidean space then many of these notions become logically equivalent resulting in a much more limited number of generalizations of the derivative and additionally, differentiability is also more well-behaved compared to the general case. This article presents the theory of -times continuously differentiable functions on an open subset of Euclidean space , which is an important special case of differentiation between arbitrary TVSs. This importance stems partially from the fact that every finite-dimensional vector subspace of a Hausdorff topological vector space is TVS isomorphic to Euclidean space so that, for example, this special case can be applied to any function whose domain is an arbitrary Hausdorff TVS by restricting it to finite-dimensional vector subspaces.

References

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