Monotone convergence theorem

Last updated

In the mathematical field of real analysis, the monotone convergence theorem is any of a number of related theorems proving the convergence of monotonic sequences (sequences that are decreasing or increasing) that are also bounded. Informally, the theorems state that if a sequence is increasing and bounded above by a supremum, then the sequence will converge to the supremum; in the same way, if a sequence is decreasing and is bounded below by an infimum, it will converge to the infimum.

Contents

Convergence of a monotone sequence of real numbers

Lemma 1

If a sequence of real numbers is increasing and bounded above, then its supremum is the limit.

Proof

Let be such a sequence, and let be the set of terms of . By assumption, is non-empty and bounded above. By the least-upper-bound property of real numbers, exists and is finite. Now, for every , there exists such that , since otherwise is an upper bound of , which contradicts the definition of . Then since is increasing, and is its upper bound, for every , we have . Hence, by definition, the limit of is

Lemma 2

If a sequence of real numbers is decreasing and bounded below, then its infimum is the limit.

Proof

The proof is similar to the proof for the case when the sequence is increasing and bounded above.

Theorem

If is a monotone sequence of real numbers (i.e., if an  an+1 for every n ≥ 1 or an  an+1 for every n ≥ 1), then this sequence has a finite limit if and only if the sequence is bounded. [1]

Proof

Convergence of a monotone series

Theorem

If for all natural numbers j and k, aj,k is a non-negative real number and aj,k aj+1,k, then [2] :168

The theorem states that if you have an infinite matrix of non-negative real numbers such that

  1. the columns are weakly increasing and bounded, and
  2. for each row, the series whose terms are given by this row has a convergent sum,

then the limit of the sums of the rows is equal to the sum of the series whose term k is given by the limit of column k (which is also its supremum). The series has a convergent sum if and only if the (weakly increasing) sequence of row sums is bounded and therefore convergent.

As an example, consider the infinite series of rows

where n approaches infinity (the limit of this series is e). Here the matrix entry in row n and column k is

the columns (fixed k) are indeed weakly increasing with n and bounded (by 1/k!), while the rows only have finitely many nonzero terms, so condition 2 is satisfied; the theorem now says that you can compute the limit of the row sums by taking the sum of the column limits, namely .

Beppo Levi's lemma

The following result is due to Beppo Levi, who proved a slight generalization in 1906 of an earlier result by Henri Lebesgue. [3] In what follows, denotes the -algebra of Borel sets on . By definition, contains the set and all Borel subsets of

Theorem

Let be a measure space, and . Consider a pointwise non-decreasing sequence of -measurable non-negative functions , i.e., for every and every ,

Set the pointwise limit of the sequence to be . That is, for every ,

Then is -measurable and

Remark 1. The integrals may be finite or infinite.

Remark 2. The theorem remains true if its assumptions hold -almost everywhere. In other words, it is enough that there is a null set such that the sequence non-decreases for every To see why this is true, we start with an observation that allowing the sequence to pointwise non-decrease almost everywhere causes its pointwise limit to be undefined on some null set . On that null set, may then be defined arbitrarily, e.g. as zero, or in any other way that preserves measurability. To see why this will not affect the outcome of the theorem, note that since we have, for every

and

provided that is -measurable. [4] :section 21.38 (These equalities follow directly from the definition of the Lebesgue integral for a non-negative function).

Remark 3. Under the assumptions of the theorem,

(Note that the second chain of equalities follows from Remark 5).

Remark 4. The proof below does not use any properties of the Lebesgue integral except those established here. The theorem, thus, can be used to prove other basic properties, such as linearity, pertaining to Lebesgue integration.

Remark 5 (monotonicity of the Lebesgue integral). In the proof below, we apply the monotonic property of the Lebesgue integral to non-negative functions only. Specifically (see Remark 4), let the functions be -measurable.

Proof. Denote by the set of simple -measurable functions such that everywhere on

1. Since we have

By definition of the Lebesgue integral and the properties of supremum,

2. Let be the indicator function of the set It can be deduced from the definition of the Lebesgue integral that

if we notice that, for every outside of Combined with the previous property, the inequality implies

Proof

This proof does not rely on Fatou's lemma; however, we do explain how that lemma might be used. Those not interested in this independency of the proof may skip the intermediate results below.

Intermediate results

Lebesgue integral as measure

Lemma 1. Let be a measurable space. Consider a simple -measurable non-negative function . For a subset , define

Then is a measure on .

Proof

Monotonicity follows from Remark 5. Here, we will only prove countable additivity, leaving the rest up to the reader. Let , where all the sets are pairwise disjoint. Due to simplicity,

for some finite non-negative constants and pairwise disjoint sets such that . By definition of the Lebesgue integral,

Since all the sets are pairwise disjoint, the countable additivity of gives us

Since all the summands are non-negative, the sum of the series, whether this sum is finite or infinite, cannot change if summation order does. For that reason,

as required.

"Continuity from below"

The following property is a direct consequence of the definition of measure.

Lemma 2. Let be a measure, and , where

is a non-decreasing chain with all its sets -measurable. Then

Proof of theorem

Step 1. We begin by showing that is –measurable. [4] :section 21.3

Note. If we were using Fatou's lemma, the measurability would follow easily from Remark 3(a).

To do this without using Fatou's lemma, it is sufficient to show that the inverse image of an interval under is an element of the sigma-algebra on , because (closed) intervals generate the Borel sigma algebra on the reals. Since is a closed interval, and, for every , ,

Thus,

Being the inverse image of a Borel set under a -measurable function , each set in the countable intersection is an element of . Since -algebras are, by definition, closed under countable intersections, this shows that is -measurable, and the integral is well-defined (and possibly infinite).

Step 2. We will first show that

The definition of and monotonicity of imply that , for every and every . By monotonicity (or, more precisely, its narrower version established in Remark 5; see also Remark 4) of the Lebesgue integral,

and

Note that the limit on the right exists (finite or infinite) because, due to monotonicity (see Remark 5 and Remark 4), the sequence is non-decreasing.

End of Step 2.

We now prove the reverse inequality. We seek to show that

.

Proof using Fatou's lemma. Per Remark 3, the inequality we want to prove is equivalent to

But the latter follows immediately from Fatou's lemma, and the proof is complete.

Independent proof. To prove the inequality without using Fatou's lemma, we need some extra machinery. Denote by the set of simple -measurable functions such that on .

Step 3. Given a simple function and a real number , define

Then , , and .

Step 3a. To prove the first claim, let , for some finite collection of pairwise disjoint measurable sets such that , some (finite) non-negative constants , and denoting the indicator function of the set .

For every holds if and only if Given that the sets are pairwise disjoint,

Since the pre-image of the Borel set under the measurable function is measurable, and -algebras, by definition, are closed under finite intersection and unions, the first claim follows.

Step 3b. To prove the second claim, note that, for each and every ,

Step 3c. To prove the third claim, we show that .

Indeed, if, to the contrary, , then an element

exists such that , for every . Taking the limit as , we get

But by initial assumption, . This is a contradiction.

Step 4. For every simple -measurable non-negative function ,

To prove this, define . By Lemma 1, is a measure on . By "continuity from below" (Lemma 2),

as required.

Step 5. We now prove that, for every ,

Indeed, using the definition of , the non-negativity of , and the monotonicity of the Lebesgue integral (see Remark 5 and Remark 4), we have

for every . In accordance with Step 4, as , the inequality becomes

Taking the limit as yields

as required.

Step 6. We are now able to prove the reverse inequality, i.e.

Indeed, by non-negativity, and For the calculation below, the non-negativity of is essential. Applying the definition of the Lebesgue integral and the inequality established in Step 5, we have

The proof is complete.

Relaxing the monotonicity assumption

Under similar hypotheses to Beppo Levi's theorem, it is possible to relax the hypothesis of monotonicity. [5] As before, let be a measure space and . Again, will be a sequence of -measurable non-negative functions . However, we do not assume they are pointwise non-decreasing. Instead, we assume that converges for almost every , we define to be the pointwise limit of , and we assume additionally that pointwise almost everywhere for all . Then is -measurable, and exists, and

As before, measurability follows from the fact that almost everywhere. The interchange of limits and integrals is then an easy consequence of Fatou's lemma. One has

by Fatou's lemma, and then, by standard properties of limits and monotonicity,

Therefore , and both are equal to . It follows that exists and equals .

See also

Notes

  1. A generalisation of this theorem was given by Bibby, John (1974). "Axiomatisations of the average and a further generalisation of monotonic sequences". Glasgow Mathematical Journal . 15 (1): 63–65. doi: 10.1017/S0017089500002135 .
  2. See for instance Yeh, J. (2006). Real Analysis: Theory of Measure and Integration. Hackensack, NJ: World Scientific. ISBN   981-256-653-8.
  3. Schappacher, Norbert; Schoof, René (1996), "Beppo Levi and the arithmetic of elliptic curves" (PDF), The Mathematical Intelligencer , 18 (1): 60, doi:10.1007/bf03024818, MR   1381581, S2CID   125072148, Zbl   0849.01036
  4. 1 2 See for instance Schechter, Erik (1997). Handbook of Analysis and Its Foundations. San Diego: Academic Press. ISBN   0-12-622760-8.
  5. coudy (https://mathoverflow.net/users/6129/coudy), Do you know important theorems that remain unknown?, URL (version: 2018-06-05): https://mathoverflow.net/q/296540

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 mathematical analysis and in probability theory, a σ-algebra on a set X is a nonempty collection Σ of subsets of X closed under complement, countable unions, and countable intersections. The ordered pair is called a measurable space.

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 probability theory, Chebyshev's inequality provides an upper bound on the probability of deviation of a random variable from its mean. More specifically, the probability that a random variable deviates from its mean by more than is at most , where is any positive constant.

<span class="mw-page-title-main">Law of large numbers</span> Averages of repeated trials converge to the [[expected value]]

In probability theory, the law of large numbers (LLN) is a mathematical theorem that states that the average of the results obtained from a large number of independent and identical random samples converges to the true value, if it exists. More formally, the LLN states that given a sample of independent and identically distributed values, the sample mean converges to the true mean.

In mathematical analysis, Hölder's inequality, named after Otto Hölder, is a fundamental inequality between integrals and an indispensable tool for the study of Lp spaces.

In mathematics, specifically measure theory, the counting measure is an intuitive way to put a measure on any set – the "size" of a subset is taken to be the number of elements in the subset if the subset has finitely many elements, and infinity if the subset is infinite.

In mathematics, Fatou's lemma establishes an inequality relating the Lebesgue integral of the limit inferior of a sequence of functions to the limit inferior of integrals of these functions. The lemma is named after Pierre Fatou.

<span class="mw-page-title-main">Jensen's inequality</span> Theorem of convex functions

In mathematics, Jensen's inequality, named after the Danish mathematician Johan Jensen, relates the value of a convex function of an integral to the integral of the convex function. It was proved by Jensen in 1906, building on an earlier proof of the same inequality for doubly-differentiable functions by Otto Hölder in 1889. Given its generality, the inequality appears in many forms depending on the context, some of which are presented below. In its simplest form the inequality states that the convex transformation of a mean is less than or equal to the mean applied after convex transformation; it is a simple corollary that the opposite is true of concave transformations.

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 mathematics, a Dirichlet series is any series of the form

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.

<span class="mw-page-title-main">Mixing (mathematics)</span> Mathematical description of mixing substances

In mathematics, mixing is an abstract concept originating from physics: the attempt to describe the irreversible thermodynamic process of mixing in the everyday world: e.g. mixing paint, mixing drinks, industrial mixing.

In mathematics, more precisely in measure theory, an atom is a measurable set which has positive measure and contains no set of smaller positive measure. A measure which has no atoms is called non-atomic or atomless.

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, more specifically measure theory, there are various notions of the convergence of measures. For an intuitive general sense of what is meant by convergence of measures, consider a sequence of measures μn on a space, sharing a common collection of measurable sets. Such a sequence might represent an attempt to construct 'better and better' approximations to a desired measure μ that is difficult to obtain directly. The meaning of 'better and better' is subject to all the usual caveats for taking limits; for any error tolerance ε > 0 we require there be N sufficiently large for nN to ensure the 'difference' between μn and μ is smaller than ε. Various notions of convergence specify precisely what the word 'difference' should mean in that description; these notions are not equivalent to one another, and vary in strength.

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, especially measure theory, a set function is a function whose domain is a family of subsets of some given set and that (usually) takes its values in the extended real number line which consists of the real numbers and

In mathematical analysis, Lorentz spaces, introduced by George G. Lorentz in the 1950s, are generalisations of the more familiar spaces.

In analytic number theory, a Dirichlet series, or Dirichlet generating function (DGF), of a sequence is a common way of understanding and summing arithmetic functions in a meaningful way. A little known, or at least often forgotten about, way of expressing formulas for arithmetic functions and their summatory functions is to perform an integral transform that inverts the operation of forming the DGF of a sequence. This inversion is analogous to performing an inverse Z-transform to the generating function of a sequence to express formulas for the series coefficients of a given ordinary generating function.