HOME

TheInfoList



OR:

In the mathematical field of
real analysis In mathematics, the branch of real analysis studies the behavior of real numbers, sequences and series of real numbers, and real functions. Some particular properties of real-valued sequences and functions that real analysis studies include conve ...
, the monotone convergence theorem is any of a number of related theorems proving the
convergence Convergence may refer to: Arts and media Literature *''Convergence'' (book series), edited by Ruth Nanda Anshen * "Convergence" (comics), two separate story lines published by DC Comics: **A four-part crossover storyline that united the four Wei ...
of
monotonic sequence In mathematics, a monotonic function (or monotone function) is a function between ordered sets that preserves or reverses the given order. This concept first arose in calculus, and was later generalized to the more abstract setting of orde ...
s (sequences that are non-decreasing or non-increasing) that are also bounded. Informally, the theorems state that if a sequence is increasing and bounded above by a
supremum In mathematics, the infimum (abbreviated inf; plural infima) of a subset S of a partially ordered set P is a greatest element in P that is less than or equal to each element of S, if such an element exists. Consequently, the term ''greatest lo ...
, then the sequence will converge to the supremum; in the same way, if a sequence is decreasing and is bounded below by an
infimum In mathematics, the infimum (abbreviated inf; plural infima) of a subset S of a partially ordered set P is a greatest element in P that is less than or equal to each element of S, if such an element exists. Consequently, the term ''greatest lo ...
, it will converge to the infimum.


Convergence of a monotone sequence of real numbers


Lemma 1

If a sequence of real numbers is increasing and bounded above, then its
supremum In mathematics, the infimum (abbreviated inf; plural infima) of a subset S of a partially ordered set P is a greatest element in P that is less than or equal to each element of S, if such an element exists. Consequently, the term ''greatest lo ...
is the limit.


Proof

Let (a_n)_ be such a sequence, and let \ be the set of terms of (a_n)_ . By assumption, \ is non-empty and bounded above. By the
least-upper-bound property In mathematics, the least-upper-bound property (sometimes called completeness or supremum property or l.u.b. property) is a fundamental property of the real numbers. More generally, a partially ordered set has the least-upper-bound property if eve ...
of real numbers, c = \sup_n \ exists and is finite. Now, for every \varepsilon > 0, there exists N such that a_N > c - \varepsilon , since otherwise c - \varepsilon is an upper bound of \, which contradicts the definition of c. Then since (a_n)_ is increasing, and c is its upper bound, for every n > N, we have , c - a_n, \leq , c - a_N, < \varepsilon . Hence, by definition, the limit of (a_n)_ is \sup_n \.


Lemma 2

If a sequence of real numbers is decreasing and bounded below, then its
infimum In mathematics, the infimum (abbreviated inf; plural infima) of a subset S of a partially ordered set P is a greatest element in P that is less than or equal to each element of S, if such an element exists. Consequently, the term ''greatest lo ...
is the limit.


Proof

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


Theorem

If (a_n)_ is a monotone
sequence In mathematics, a sequence is an enumerated collection of objects in which repetitions are allowed and order matters. Like a set, it contains members (also called ''elements'', or ''terms''). The number of elements (possibly infinite) is calle ...
of
real number In mathematics, a real number is a number that can be used to measure a ''continuous'' one-dimensional quantity such as a distance, duration or temperature. Here, ''continuous'' means that values can have arbitrarily small variations. Every r ...
s (i.e., if ''a''''n'' ≤ ''a''''n''+1 for every ''n'' ≥ 1 or ''a''''n'' ≥ ''a''''n''+1 for every ''n'' ≥ 1), then this sequence has a finite limit
if and only if In logic and related fields such as mathematics and philosophy, "if and only if" (shortened as "iff") is a biconditional logical connective between statements, where either both statements are true or both are false. The connective is bicond ...
the sequence is bounded.


Proof

* "If"-direction: The proof follows directly from the lemmas. * "Only If"-direction: By
(ε, δ)-definition of limit Although the function (sin ''x'')/''x'' is not defined at zero, as ''x'' becomes closer and closer to zero, (sin ''x'')/''x'' becomes arbitrarily close to 1. In other words, the limit of (sin ''x'')/''x'', as ''x'' approaches z ...
, every sequence (a_n)_ with a finite limit L is necessarily bounded.


Convergence of a monotone series


Theorem

If for all natural numbers ''j'' and ''k'', ''a''''j'',''k'' is a non-negative real number and ''a''''j'',''k'' ≤ ''a''''j''+1,''k'', then :\lim_ \sum_k a_ = \sum_k \lim_ a_. The theorem states that if you have an infinite matrix of non-negative real numbers such that #the columns are weakly increasing and bounded, and #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 In mathematics, the infimum (abbreviated inf; plural infima) of a subset S of a partially ordered set P is a greatest element in P that is less than or equal to each element of S, if such an element exists. Consequently, the term ''greatest lo ...
). 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 :: \left( 1+ \frac1 n\right)^n = \sum_^n \binom nk \frac 1 = \sum_^n \frac1 \times \frac nn \times \fracn\times\cdots\times\fracn, where ''n'' approaches infinity (the limit of this series is e). Here the matrix entry in row ''n'' and column ''k'' is :\binom nk \frac 1 =\frac1\times\frac nn\times\fracn\times\cdots\times\fracn; 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 (1+1/n)^n by taking the sum of the column limits, namely \frac1.


Beppo Levi's lemma

The following result is due to
Beppo Levi Beppo Levi (14 May 1875 – 28 August 1961) was an Italian mathematician. He published high-level academic articles and books, not only on mathematics, but also on physics, history, philosophy, and pedagogy. Levi was a member of the Bologna Ac ...
, who proved a slight generalization in 1906 of an earlier result by
Henri Lebesgue Henri Léon Lebesgue (; June 28, 1875 – July 26, 1941) was a French mathematician known for his theory of integration, which was a generalization of the 17th-century concept of integration—summing the area between an axis and the curve of ...
. In what follows, \operatorname_ denotes the \sigma-algebra of Borel sets on ,+\infty/math>. By definition, \operatorname_ contains the set \ and all Borel subsets of \R_.


Theorem

Let (\Omega,\Sigma,\mu) be a
measure space A measure space is a basic object of measure theory, a branch of mathematics that studies generalized notions of volumes. It contains an underlying set, the subsets of this set that are feasible for measuring (the -algebra) and the method that ...
, and X\in\Sigma. Consider a pointwise non-decreasing sequence \^\infty_ of (\Sigma,\operatorname_)-
measurable In mathematics, the concept of a measure is a generalization and formalization of geometrical measures ( length, area, volume) and other common notions, such as mass and probability of events. These seemingly distinct concepts have many si ...
non-negative functions f_k:X\to ,+\infty/math>, i.e., for every and every , : 0 \leq f_k(x) \leq f_(x)\leq\infty. Set the pointwise limit of the sequence \ to be f. That is, for every x\in X, : f(x):= \lim_ f_k(x). Then f is (\Sigma,\operatorname_)-measurable and :\lim_ \int_X f_k \,d\mu = \int_X f \,d\mu. Remark 1. The integrals may be finite or infinite. Remark 2. The theorem remains true if its assumptions hold \mu-almost everywhere. In other words, it is enough that there is a
null set In mathematical analysis, a null set N \subset \mathbb is a measurable set that has measure zero. This can be characterized as a set that can be covered by a countable union of intervals of arbitrarily small total length. The notion of null ...
N 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 f to be undefined on some null set N. On that null set, f 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 k, : \int_X f_k \,d\mu = \int_ f_k \,d\mu and \int_X f \,d\mu = \int_ f \,d\mu, provided that f is (\Sigma,\operatorname_)-measurable.See for instance (These equalities follow directly from the definition of Lebesgue integral for a non-negative function). Remark 3. Under 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 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 Lebesgue integral). In the proof below, we apply the monotonic property of Lebesgue integral to non-negative functions only. Specifically (see Remark 4), let the functions f,g : X \to ,+\infty/math> be (\Sigma,\operatorname_)-measurable. *If f \leq g everywhere on X, then :\int_X f\,d\mu \leq \int_X g\,d\mu. *If X_1,X_2 \in \Sigma and X_1 \subseteq X_2, then :\int_ f\,d\mu \leq \int_ f\,d\mu. Proof. Denote \operatorname(h) the set of simple (\Sigma, \operatorname_)-measurable functions s:X\to [0,\infty) such that 0\leq s\leq h everywhere on X. 1. Since f \leq g, we have : \operatorname(f) \subseteq \operatorname(g). By definition of Lebesgue integral and the properties of supremum, :\int_X f\,d\mu = \sup_\int_X s\,d\mu \leq \sup_\int_X s\,d\mu = \int_X g\,d\mu. 2. Let _ be the indicator function of the set X_1. It can be deduced from the definition of Lebesgue integral that : \int_ f\cdot _ \,d\mu = \int_ f \,d\mu if we notice that, for every s \in (f\cdot _), s=0 outside of X_1. Combined with the previous property, the inequality f\cdot _ \leq f implies : \int_ f \,d\mu = \int_ f\cdot _ \,d\mu \leq \int_ f \,d\mu.


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 (\Omega,\Sigma,\mu) be a measurable space. Consider a simple (\Sigma,\operatorname_)-measurable non-negative function s:\Omega\to. For a subset S\subseteq\Omega, define :\nu(S)=\int_Ss\,d\mu. Then \nu is a measure on \Omega.


Proof

Monotonicity follows from Remark 5. Here, we will only prove countable additivity, leaving the rest up to the reader. Let S=\bigcup^\infty_S_i, where all the sets S_i are pairwise disjoint. Due to simplicity, :s=\sum^n_c_i\cdot _, for some finite non-negative constants c_i\in_ and pairwise disjoint sets A_i\in\Sigma such that \bigcup^n_A_i=\Omega. By definition of Lebesgue integral, : \begin \nu(S) &=\sum^n_c_i\cdot\mu(S\cap A_i)\\ &=\sum^n_c_i\cdot\mu\left(\left(\bigcup^\infty_ S_j\right)\cap A_i\right)\\ &=\sum^n_c_i\cdot\mu\left(\bigcup^\infty_(S_j\cap A_i)\right) \end Since all the sets S_j\cap A_i are pairwise disjoint, the countable additivity of \mu gives us : \sum^n_ c_i\cdot\mu \left(\bigcup^\infty_(S_j\cap A_i)\right)=\sum^n_c_i\cdot\sum^\infty_ \mu(S_j\cap A_i). 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, : \begin \sum^n_c_i\cdot\sum^\infty_\mu(S_j\cap A_i)&=\sum^\infty_\sum^n_c_i\cdot \mu(S_j\cap A_i)\\ &=\sum^\infty_\int_ s\,d\mu\\ &=\sum^\infty_\nu(S_j), \end as required.


="Continuity from below"

= The following property is a direct consequence of the definition of measure. Lemma 2. Let \mu be a measure, and S = \bigcup^\infty_S_i, where : S_1\subseteq\cdots\subseteq S_i\subseteq S_\subseteq\cdots\subseteq S is a non-decreasing chain with all its sets \mu-measurable. Then :\mu(S)=\lim_i\mu(S_i).


Proof of theorem

Step 1. We begin by showing that f is (\Sigma, \operatorname_) –measurable. 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 ,t/math> under f is an element of the sigma-algebra \Sigma on X, because (closed) intervals generate the
Borel sigma algebra In mathematics, a Borel set is any set in a topological space that can be formed from open sets (or, equivalently, from closed sets) through the operations of countable union, countable intersection, and relative complement. Borel sets are named ...
on the reals. Since ,t/math> is a closed interval, and, for every k, 0\le f_k(x) \le f(x), :0\leq f(x)\leq t\quad \Leftrightarrow\quad \Bigl forall k\quad 0\leq f_k(x)\leq t\Bigr Thus, :\ = \bigcap_k \. Being the inverse image of a
Borel set In mathematics, a Borel set is any set in a topological space that can be formed from open sets (or, equivalently, from closed sets) through the operations of countable union, countable intersection, and relative complement. Borel sets are named ...
under a (\Sigma,\operatorname_)-measurable function f_k, each set in the countable intersection is an element of \Sigma. Since \sigma-algebras are, by definition, closed under countable intersections, this shows that f is (\Sigma,\operatorname_)-measurable, and the integral \textstyle \int_X f \,d\mu is well defined (and possibly infinite). Step 2. We will first show that \textstyle\int_X f \,d\mu \geq \lim_k \int_X f_k \,d\mu. The definition of f and monotonicity of \ imply that f(x)\geq f_k(x), for every k and every x\in X. By monotonicity (or, more precisely, its narrower version established in Remark 5; see also Remark 4) of Lebesgue integral, :\int_X f\,d\mu\geq\int_X f_k\,d\mu, and :\int_X f\,d\mu\geq\lim_k\int_X f_k\,d\mu. 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 : \int_X f \,d\mu \leq \lim_k \int_X f_k \,d\mu . Proof using Fatou's lemma. Per Remark 3, the inequality we want to prove is equivalent to :\int_X \liminf_k f_k(x) \,d\mu \leq \liminf_k \int_X f_k \,d\mu. 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 \operatorname(f) the set of simple (\Sigma,\operatorname_)-measurable functions s:X\to [0,\infty) such that 0\leq s\leq f on X. Step 3. Given a simple function s\in\operatorname(f) and a real number t\in (0,1), define :B^_k=\\subseteq X. Then B^_k\in\Sigma, B^_k\subseteq B^_, and \textstyle X=\bigcup_k B^_k. Step 3a. To prove the first claim, let \textstyle s=\sum^m_c_i\cdot_, for some finite collection of pairwise disjoint measurable sets A_i\in\Sigma such that \textstyle X=\cup^m_A_i, some (finite) non-negative constants c_i\in _, and _ denoting the indicator function of the set A_i. For every x\in A_i, t\cdot s(x)\leq f_k(x) holds if and only if f_k(x) \in [t\cdot c_i, +\infty]. Given that the sets A_i are pairwise disjoint, :B^_k=\bigcup^m_\Bigl(f^_k\Bigl([t\cdot c_i,+\infty]\Bigr)\cap A_i\Bigr). Since the pre-image f^_k\Bigl([t\cdot c_i,+\infty]\Bigr) of the Borel set \cdot c_i,+\infty/math> under the measurable function f_k is measurable, and \sigma-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 k and every x\in X, f_k(x)\leq f_(x). Step 3c. To prove the third claim, we show that \textstyle X\subseteq\bigcup_k B^_k. Indeed, if, to the contrary, \textstyle X\not\subseteq\bigcup_k B^_k, then an element :\textstyle x_0\in X\setminus\bigcup_k B^_k=\bigcap_k(X\setminus B^_k) exists such that f_k(x_0), for every k. Taking the limit as k\to\infty, we get :f(x_0)\leq t\cdot s(x_0) But by initial assumption, s\leq f. This is a contradiction. Step 4. For every simple (\Sigma,\operatorname_)-measurable non-negative function s_2, :\lim_n\int_s_2\,d\mu=\int_Xs_2\,d\mu. To prove this, define \textstyle\nu(S)=\int_S s_2\,d\mu. By Lemma 1, \nu(S) is a measure on \Omega. By "continuity from below" (Lemma 2), :\lim_n\int_s_2\,d\mu=\lim_n\nu(B^_n)=\nu(X)=\int_Xs_2\,d\mu, as required. Step 5. We now prove that, for every s\in\operatorname(f), :\int_X s\,d\mu\leq\lim_k\int_X f_k\,d\mu. Indeed, using the definition of B^_k, the non-negativity of f_k, and the monotonicity of Lebesgue integral (see Remark 5 and Remark 4), we have :\int_t\cdot s\,d\mu\leq\int_ f_k\,d\mu\leq\int_X f_k\,d\mu, for every k\geq 1. In accordance with Step 4, as k\to\infty, the inequality becomes :t\int_X s\,d\mu\leq\lim_k\int_X f_k\,d\mu. Taking the limit as t\uparrow 1 yields :\int_X s\,d\mu\leq\lim_k\int_X f_k\,d\mu, as required. Step 6. We are now able to prove the reverse inequality, i.e. : \int_X f \,d\mu \leq \lim_k \int_X f_k \,d\mu. Indeed, by non-negativity, f_+ = f and f_- = 0. For the calculation below, the non-negativity of f is essential. Applying the definition of Lebesgue integral and the inequality established in Step 5, we have : \int_X f \,d\mu=\sup_\int_X s\,d\mu\leq\lim_k\int_X f_k\,d\mu. The proof is complete.


See also

*
Infinite series In mathematics, a series is, roughly speaking, a description of the operation of adding infinitely many quantities, one after the other, to a given starting quantity. The study of series is a major part of calculus and its generalization, mat ...
*
Dominated convergence theorem In measure theory, Lebesgue's dominated convergence theorem provides sufficient conditions under which almost everywhere convergence of a sequence of functions implies convergence in the ''L''1 norm. Its power and utility are two of the primary t ...


Notes

{{Measure theory Articles containing proofs Theorems in calculus Sequences and series Theorems in real analysis Theorems in measure theory it:Passaggio al limite sotto segno di integrale#Integrale di Lebesgue