Bochner integral

Last updated

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.

Contents

Definition

Let be a measure space, and be a Banach space. The Bochner integral of a function is defined in much the same way as the Lebesgue integral. First, define a simple function to be any finite sum of the form

where the are disjoint members of the -algebra the are distinct elements of and χE is the characteristic function of If is finite whenever then the simple function is integrable, and the integral is then defined by

exactly as it is for the ordinary Lebesgue integral.

A measurable function is Bochner integrable if there exists a sequence of integrable simple functions such that

where the integral on the left-hand side is an ordinary Lebesgue integral.

In this case, the Bochner integral is defined by

It can be shown that the sequence is a Cauchy sequence in the Banach space hence the limit on the right exists; furthermore, the limit is independent of the approximating sequence of simple functions These remarks show that the integral is well-defined (i.e independent of any choices). It can be shown that a function is Bochner integrable if and only if it lies in the Bochner space

Properties

Elementary properties

Many of the familiar properties of the Lebesgue integral continue to hold for the Bochner integral. Particularly useful is Bochner's criterion for integrability, which states that if is a measure space, then a Bochner-measurable function is Bochner integrable if and only if

Here, a function  is called Bochner measurable if it is equal -almost everywhere to a function taking values in a separable subspace of , and such that the inverse image of every open set  in  belongs to . Equivalently, is the limit -almost everywhere of a sequence of countably-valued simple functions.

Linear operators

If is a continuous linear operator between Banach spaces and , and is Bochner integrable, then it is relatively straightforward to show that is Bochner integrable and integration and the application of may be interchanged:

for all measurable subsets .

A non-trivially stronger form of this result, known as Hille's theorem, also holds for closed operators. [1] If is a closed linear operator between Banach spaces and and both and are Bochner integrable, then

for all measurable subsets .

Dominated convergence theorem

A version of the dominated convergence theorem also holds for the Bochner integral. Specifically, if is a sequence of measurable functions on a complete measure space tending almost everywhere to a limit function , and if

for almost every , and , then

as and

for all .

If is Bochner integrable, then the inequality

holds for all In particular, the set function

defines a countably-additive -valued vector measure on which is absolutely continuous with respect to .

Radon–Nikodym property

An important fact about the Bochner integral is that the Radon–Nikodym theorem fails to hold in general, and instead is a property (the Radon–Nikodym property) defining an important class of nice Banach spaces.

Specifically, if is a measure on then has the Radon–Nikodym property with respect to if, for every countably-additive vector measure on with values in which has bounded variation and is absolutely continuous with respect to there is a -integrable function such that

for every measurable set [2]

The Banach space has the Radon–Nikodym property if has the Radon–Nikodym property with respect to every finite measure. [2] Equivalent formulations include:

It is known that the space has the Radon–Nikodym property, but and the spaces for an open bounded subset of and for an infinite compact space, do not. [5] Spaces with Radon–Nikodym property include separable dual spaces (this is the Dunford–Pettis theorem)[ citation needed ] and reflexive spaces, which include, in particular, Hilbert spaces. [2]

See also

Related Research Articles

<span class="mw-page-title-main">Measure (mathematics)</span> Generalization of mass, length, area and volume

In mathematics, the concept of a measure is a generalization and formalization of geometrical measures and other common notions, such as magnitude, mass, and probability of events. These seemingly distinct concepts have many similarities and can often be treated together in a single mathematical context. Measures are foundational in probability theory, integration theory, and can be generalized to assume negative values, as with electrical charge. Far-reaching generalizations of measure are widely used in quantum physics and physics in general.

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.

In calculus, absolute continuity is a smoothness property of functions that is stronger than continuity and uniform continuity. The notion of absolute continuity allows one to obtain generalizations of the relationship between the two central operations of calculus—differentiation and integration. This relationship is commonly characterized in the framework of Riemann integration, but with absolute continuity it may be formulated in terms of Lebesgue integration. For real-valued functions on the real line, two interrelated notions appear: absolute continuity of functions and absolute continuity of measures. These two notions are generalized in different directions. The usual derivative of a function is related to the Radon–Nikodym derivative, or density, of a measure. We have the following chains of inclusions for functions over a compact subset of the real line:

In mathematics, the Radon–Nikodym theorem is a result in measure theory that expresses the relationship between two measures defined on the same measurable space. A measure is a set function that assigns a consistent magnitude to the measurable subsets of a measurable space. Examples of a measure include area and volume, where the subsets are sets of points; or the probability of an event, which is a subset of possible outcomes within a wider probability space.

In probability theory, the conditional expectation, conditional expected value, or conditional mean of a random variable is its expected value – the value it would take “on average” over an arbitrarily large number of occurrences – given that a certain set of "conditions" is known to occur. If the random variable can take on only a finite number of values, the “conditions” are that the variable can only take on a subset of those values. More formally, in the case when the random variable is defined over a discrete probability space, the "conditions" are a partition of this probability space.

In mathematics, the ba space of an algebra of sets is the Banach space consisting of all bounded and finitely additive signed measures on . The norm is defined as the variation, that is

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 mathematics, signed measure is a generalization of the concept of (positive) measure by allowing the set function to take negative values.

In mathematics, specifically measure theory, a complex measure generalizes the concept of measure by letting it have complex values. In other words, one allows for sets whose size is a complex number.

In mathematics, a positive (or signed) measure μ defined on a σ-algebra Σ of subsets of a set X is called a finite measure if μ(X) is a finite real number (rather than ∞), and a set A in Σ is of finite measure if μ(A) < ∞. The measure μ is called σ-finite if X is a countable union of measurable sets each with finite measure. A set in a measure space is said to have σ-finite measure if it is a countable union of measurable sets with finite measure. A measure being σ-finite is a weaker condition than being finite, i.e. all finite measures are σ-finite but there are (many) σ-finite measures that are not finite.

In mathematics, Bochner spaces are a generalization of the concept of spaces to functions whose values lie in a Banach space which is not necessarily the space or of real or complex numbers.

In measure theory, a pushforward measure is obtained by transferring a measure from one measurable space to another using a measurable function.

In probability theory, a random measure is a measure-valued random element. Random measures are for example used in the theory of random processes, where they form many important point processes such as Poisson point processes and Cox processes.

In mathematics—specifically, in functional analysis—a weakly measurable function taking values in a Banach space is a function whose composition with any element of the dual space is a measurable function in the usual (strong) sense. For separable spaces, the notions of weak and strong measurability agree.

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 probability theory and statistics, the law of the unconscious statistician, or LOTUS, is a theorem which expresses the expected value of a function g(X) of a random variable X in terms of g and the probability distribution of X.

In mathematics, lifting theory was first introduced by John von Neumann in a pioneering paper from 1931, in which he answered a question raised by Alfréd Haar. The theory was further developed by Dorothy Maharam (1958) and by Alexandra Ionescu Tulcea and Cassius Ionescu Tulcea (1961). Lifting theory was motivated to a large extent by its striking applications. Its development up to 1969 was described in a monograph of the Ionescu Tulceas. Lifting theory continued to develop since then, yielding new results and applications.

<span class="mw-page-title-main">Lebesgue integration</span> Method of integration

In mathematics, the integral of a non-negative function of a single variable can be regarded, in the simplest case, as the area between the graph of that function and the x-axis. The Lebesgue integral, named after French mathematician Henri Lebesgue, extends the integral to a larger class of functions. It also extends the domains on which these functions can be defined.

In mathematics – specifically, in functional analysis – a Bochner-measurable function taking values in a Banach space is a function that equals almost everywhere the limit of a sequence of measurable countably-valued functions, i.e.,

A Markov chain on a measurable state space is a discrete-time-homogeneous Markov chain with a measurable space as state space.

References

    1. Diestel, Joseph; Uhl, Jr., John Jerry (1977). Vector Measures. Mathematical Surveys. American Mathematical Society. doi:10.1090/surv/015. (See Theorem II.2.6)
    2. 1 2 3 Bárcenas, Diómedes (2003). "The Radon–Nikodym Theorem for Reflexive Banach Spaces" (PDF). Divulgaciones Matemáticas. 11 (1): 55–59 [pp. 55–56].
    3. 1 2 Bourgin 1983 , pp. 31, 33. Thm. 2.3.6-7, conditions (1,4,10).
    4. Bourgin 1983 , p. 16. "Early workers in this field were concerned with the Banach space property that each X-valued function of bounded variation on [0,1] be differentiable almost surely. It turns out that this property (known as the Gelfand-Fréchet property) is also equivalent to the RNP [Radon-Nikodym Property]."
    5. Bourgin 1983 , p. 14.