Fact-checked by Grok 2 weeks ago

Monotone convergence theorem

The monotone convergence theorem is a cornerstone result in measure theory, asserting that if a sequence of non-negative measurable functions \{f_n\} on a (X, \mathcal{A}, \mu) increases to a f, then the of the equals the of the integrals: \int_X f \, d\mu = \lim_{n \to \infty} \int_X f_n \, d\mu. This theorem, first proved by in his 1902 dissertation and later generalized by Beppo Levi in 1906, extends the basic property of Riemann integrals to the more general Lebesgue framework, enabling the rigorous handling of under integration for monotone sequences. It applies specifically to non-decreasing sequences of functions taking values in [0, \infty], requiring measurability and , and serves as a prerequisite for more advanced results like the . The theorem's proof typically relies on and the monotonicity of the Lebesgue , highlighting its role in establishing continuity of the operator with respect to monotone . In applications, it justifies interchanging and integrals in , partial differential equations, and , where constructing integrals via of simple functions is common.

Sequences and Series in Real Numbers

Monotone Sequence Convergence

A sequence of real numbers is one that is either non-decreasing or non-increasing. Specifically, a \{a_n\}_{n=1}^\infty is increasing if a_n \leq a_{n+1} for all n \in \mathbb{N}, and strictly increasing if the inequality is strict; it is decreasing if a_n \geq a_{n+1} for all n \in \mathbb{N}, and strictly decreasing if strict. The convergence theorem for states that every bounded of real numbers converges. More precisely, if \{a_n\} is increasing and bounded above, then it converges to its least upper bound, or supremum: \lim_{n \to \infty} a_n = \sup \{a_n : n \in \mathbb{N}\}. Similarly, if \{a_n\} is decreasing and bounded below, it converges to its greatest lower bound, or infimum. Conversely, a converges it is bounded. To prove this, consider the case of a monotone increasing sequence \{a_n\} that is bounded above. Let S = \{a_n : n \in \mathbb{N}\}, and let L = \sup S, which exists by the completeness axiom of the real numbers (the least upper bound property). For any \epsilon > 0, L - \epsilon is not an upper bound for S, so there exists some N \in \mathbb{N} such that a_N > L - \epsilon. Since the sequence is increasing, a_n \geq a_N > L - \epsilon for all n \geq N. Also, a_n \leq L for all n, so L - \epsilon < a_n \leq L for all n \geq N, proving that \lim_{n \to \infty} a_n = L. The decreasing case follows analogously by considering -a_n. The "only if" direction holds because every convergent sequence is bounded. This proof relies on the Dedekind completeness of the reals, ensuring every nonempty subset bounded above has a least upper bound. This foundational result in real analysis relies on the completeness of the real numbers, developed in the 19th century.

Monotone Series Convergence

A monotone series is defined as an infinite series \sum_{n=1}^\infty a_n where each term a_n \geq 0 for all n \in \mathbb{N}, ensuring that the sequence of partial sums S_n = \sum_{k=1}^n a_k is non-decreasing, or increasing if all a_n > 0. The fundamental theorem for such series states that \sum_{n=1}^\infty a_n converges the partial sums \{S_n\} are bounded above by some finite M > 0. If bounded, the series converges to \lim_{n \to \infty} S_n = \sup \{ S_n : n \in \mathbb{N} \}; otherwise, the partial sums diverge to \infty, and the series diverges. The proof follows directly from the monotone convergence theorem for sequences applied to the partial sums. Since \{S_n\} is monotone increasing (as S_{n+1} = S_n + a_{n+1} \geq S_n) and bounded above the series converges, the limit exists and is finite precisely when bounded. This result applies exclusively to series with non-negative terms and differs from , which requires the series of absolute values \sum |a_n| to converge for general real or complex terms, guaranteeing unconditional convergence but not vice versa in the non-negative case.

Examples of Monotone Sums

The harmonic series \sum_{n=1}^\infty \frac{1}{n} provides a classic example of a . Its partial sums H_n = \sum_{k=1}^n \frac{1}{k} form a monotonically increasing because each is positive, yet H_n is unbounded above, as H_n > \ln n + \gamma where \gamma \approx 0.577 is the Euler-Mascheroni constant, implying divergence to infinity. The for the base of the natural logarithm, e = \sum_{k=0}^\infty \frac{1}{k!}, exemplifies monotone convergence of a series with positive terms. The partial sums S_n = \sum_{k=0}^n \frac{1}{k!} are monotonically increasing and bounded above by 3, since the tail satisfies e - S_n = \sum_{k=n+1}^\infty \frac{1}{k!} < \frac{1}{n!} \sum_{j=0}^\infty \frac{1}{(n+1)^j} = \frac{1}{n! (n)} for n \geq 1, implying S_n < e < S_n + \frac{1}{n!}. Thus, by the monotone convergence theorem for series, S_n converges to e \approx 2.71828.

Measure-Theoretic Generalization

Beppo Levi's Lemma for Integrals

Beppo Levi's lemma, named after the Italian mathematician Beppo Levi (1875–1961), emerged in the early as a cornerstone of the developing theory of , building directly on Henri Lebesgue's foundational work from 1902. Levi's contributions, particularly in his 1906 publications, addressed key aspects of integration for non-negative functions and series, providing rigorous justifications that complemented and extended Lebesgue's ideas. This lemma represents the measure-theoretic generalization of the monotone convergence theorem for real sequences, adapting the discrete concept to integrals over abstract measure spaces. In the context of , consider a (X, \mathcal{M}, \mu). A is a finite s = \sum_{k=1}^m c_k \chi_{E_k}, where c_k \geq 0, the E_k are disjoint measurable sets in \mathcal{M}, and \chi_{E_k} is the of E_k. The of such a is \int s \, d\mu = \sum_{k=1}^m c_k \mu(E_k). For a non-negative measurable function f: X \to [0, \infty], the Lebesgue is defined as \int_X f \, d\mu = \sup \left\{ \int_X s \, d\mu : s \text{ simple}, \, 0 \leq s \leq f \right\}, which may equal \infty. This definition extends the from to all non-negative measurable functions via approximation from below. Beppo Levi's lemma asserts that if \{f_n\}_{n=1}^\infty is a of non-negative s on X such that f_n \uparrow f (meaning $0 \leq f_1 \leq f_2 \leq \cdots and \lim_{n \to \infty} f_n(x) = f(x) for \mu- x \in X, where f is the equal to this ), then \int_X f \, d\mu = \lim_{n \to \infty} \int_X f_n \, d\mu. The integrals take values in [0, \infty], and the exists (possibly infinite) because the \{\int_X f_n \, d\mu\}_{n=1}^\infty is non-decreasing by the monotonicity of the Lebesgue integral for non-negative functions. This equality follows from the definition of the , as the partial approximations align under the : \int_X f \, d\mu = \sup \left\{ \int_X s \, d\mu : s \text{ simple}, \, 0 \leq s \leq f \right\} = \lim_{n \to \infty} \int_X f_n \, d\mu, ensuring the interchange of and under monotonicity.

Monotonicity of the Lebesgue Integral

The monotonicity property of the Lebesgue integral states that if $0 \leq g \leq f are non-negative measurable functions on a (X, \mathcal{A}, \mu), then \int_X g \, d\mu \leq \int_X f \, d\mu. This property is fundamental to the theory of integration and serves as a prerequisite for results like Beppo Levi's lemma. To prove this, first consider the case where f and g are simple functions. Express f = \sum_{i=1}^n a_i \chi_{E_i}, where a_i \geq 0, the E_i are disjoint measurable sets partitioning the support of f, and \chi_{E_i} is the of E_i. The Lebesgue integral is then \int_X f \, d\mu = \sum_{i=1}^n a_i \mu(E_i). Since $0 \leq g \leq f, on each E_i, g \leq a_i, so \int_{E_i} g \, d\mu \leq a_i \mu(E_i). Summing over i gives \int_X g \, d\mu \leq \sum_{i=1}^n a_i \mu(E_i) = \int_X f \, d\mu. For general non-negative measurable functions f and g with g \leq f, the result follows directly from the : any simple s with $0 \leq s \leq g also satisfies s \leq f, so \int_X g \, d\mu = \sup\left\{ \int_X s \, d\mu : s \text{ simple}, 0 \leq s \leq g \right\} \leq \sup\left\{ \int_X t \, d\mu : t \text{ simple}, 0 \leq t \leq f \right\} = \int_X f \, d\mu. A direct consequence is that if \{f_n\} is a of non-negative measurable functions with f_n \uparrow f pointwise , then \int_X f_n \, d\mu \uparrow \int_X f \, d\mu. This follows by repeated application of the monotonicity property to the differences f_{n+1} - f_n \geq 0. Unlike the , which may fail for functions with many discontinuities due to reliance on interval partitions, the Lebesgue integral's monotonicity holds robustly through measure-theoretic approximation, accommodating arbitrary non-negative measurable functions.

Proof of Beppo Levi's Lemma

Beppo Levi's lemma states that if \{f_n\}_{n=1}^\infty is an increasing sequence of nonnegative measurable functions on a (X, \mathcal{A}, \mu), then f = \lim_{n \to \infty} f_n (defined ) is measurable and \lim_{n \to \infty} \int_X f_n \, d\mu = \int_X f \, d\mu, where the integrals may be infinite. The proof proceeds in two main steps, beginning with the case where the f_n are simple functions. In this case, each f_n = \sum_{i=1}^{m_n} a_{n,i} \chi_{E_{n,i}} with a_{n,i} \geq 0 and E_{n,i} \in \mathcal{A}. Since the sequence is increasing, f(x) = \sup_n f_n(x) for \mu-almost every x \in X, and the \int_X f_n \, d\mu = \sum_{i=1}^{m_n} a_{n,i} \mu(E_{n,i}) increases with n by monotonicity of the measure. The limit f is a nonnegative , and since each f_n is simple with f_n \leq f, \lim \int f_n \leq \int f; the reverse holds by the sup definition, as simple functions below f are eventually approximated by the f_n. For the general case, approximate each nonnegative measurable f_n by an increasing sequence of simple functions s_{n,k} \uparrow f_n as k \to \infty, such that \int_X s_{n,k} \, d\mu \uparrow \int_X f_n \, d\mu by the definition of the Lebesgue integral for nonnegative functions. The double-indexed sequence \{s_{n,k}\} then satisfies s_{n,k} \uparrow f as n,k \to \infty almost everywhere. By monotonicity of the integral, \lim_{n \to \infty} \int_X f_n \, d\mu = \lim_{n \to \infty} \lim_{k \to \infty} \int_X s_{n,k} \, d\mu = \lim_{k \to \infty} \lim_{n \to \infty} \int_X s_{n,k} \, d\mu = \lim_{k \to \infty} \int_X \lim_{n \to \infty} s_{n,k} \, d\mu = \int_X f \, d\mu, where the iterated limits commute due to the joint monotonicity in both indices, as the double sequence increases to f. Monotonicity of the Lebesgue integral ensures that \int_X f_n \, d\mu \leq \int_X f \, d\mu for each n, so \lim_{n \to \infty} \int_X f_n \, d\mu \leq \int_X f \, d\mu. The reverse inequality follows from the approximation: for any \phi with $0 \leq \phi \leq f, define A_n = \{x : f_n(x) \geq t \phi(x)\} for $0 < t < 1; then A_n \uparrow X almost everywhere, and \int_X f_n \, d\mu \geq t \int_{A_n} \phi \, d\mu. Taking n \to \infty yields \lim_{n \to \infty} \int_X f_n \, d\mu \geq t \int_X \phi \, d\mu by continuity of the measure from below. Letting t \to 1^- gives \lim_{n \to \infty} \int_X f_n \, d\mu \geq \int_X \phi \, d\mu. Supremum over all such \phi yields \lim_{n \to \infty} \int_X f_n \, d\mu \geq \int_X f \, d\mu, establishing equality by squeezing. If \lim_{n \to \infty} \int_X f_n \, d\mu = \infty, then \int_X f \, d\mu = \infty, as the reverse always holds. This completes the proof, with the monotonicity of the serving as the key tool throughout.

Advanced Proof Techniques

Intermediate Lemmas for Convergence

In measure theory, several intermediate lemmas establish key properties of measures that facilitate proofs of theorems, particularly by linking set measures to via functions. These lemmas exploit the structure of increasing sequences of sets and the monotonicity of the underlying measure. A fundamental result is the continuity of measures from below, which states that if \{E_n\}_{n=1}^\infty is an increasing of measurable sets with E = \bigcup_{n=1}^\infty E_n, then \mu(E) = \lim_{n \to \infty} \mu(E_n) for a measure \mu on a . To prove this, consider the characteristic functions \chi_{E_n}, which form an increasing converging pointwise to \chi_E. Since the of a characteristic function equals the measure of the corresponding set, applying Beppo Levi's lemma to the nonnegative measurable functions \chi_{E_n} \uparrow \chi_E yields \int \chi_E \, d\mu = \lim_{n \to \infty} \int \chi_{E_n} \, d\mu, or equivalently, \mu(E) = \lim_{n \to \infty} \mu(E_n). This lemma can also be established using properties of outer measures, but the approach highlights its connection to functional . Building on this, another extends finite additivity to countable unions of disjoint measurable sets. Specifically, if \{E_n\}_{n=1}^\infty is a sequence of pairwise disjoint measurable sets with E = \bigcup_{n=1}^\infty E_n, then the measure satisfies countable additivity: \mu(E) = \sum_{n=1}^\infty \mu(E_n). The proof proceeds by defining partial unions F_k = \bigcup_{n=1}^k E_n, which form an increasing sequence converging to E and satisfy \mu(F_k) = \sum_{n=1}^k \mu(E_n) by finite additivity; continuity from below then implies \mu(E) = \lim_{k \to \infty} \mu(F_k) = \sum_{n=1}^\infty \mu(E_n). These lemmas serve as bridges between finite additivity—assumed in the definition of a —and the \sigma-additivity required for a full measure, enabling rigorous extensions to infinite processes in arguments.

Proof Using

Fatou's lemma provides a fundamental for interchanging limits and integrals of non-negative functions. Specifically, if \{f_n\} is a sequence of non-negative measurable functions on a (X, \mathcal{M}, \mu), then \int_X \liminf_{n \to \infty} f_n \, d\mu \leq \liminf_{n \to \infty} \int_X f_n \, d\mu. This result, originally established by in his 1906 work on trigonometric series, forms the basis for many theorems in measure theory. To prove the monotone convergence theorem using Fatou's lemma, assume \{f_n\} is a sequence of non-negative measurable functions such that $0 \leq f_1 \leq f_2 \leq \cdots and f_n \uparrow f pointwise, where f is also non-negative and measurable. Since the sequence is monotonically increasing, \liminf_{n \to \infty} f_n = \lim_{n \to \infty} f_n = f. Applying Fatou's lemma yields \int_X f \, d\mu \leq \liminf_{n \to \infty} \int_X f_n \, d\mu. The monotonicity of the functions implies that the sequence of integrals \{\int_X f_n \, d\mu\} is non-decreasing, so its limit exists (finite or infinite), and \liminf_{n \to \infty} \int_X f_n \, d\mu = \lim_{n \to \infty} \int_X f_n \, d\mu. Additionally, the inequality f_n \leq f for all n, combined with the monotonicity of the Lebesgue integral for non-negative functions, gives \int_X f_n \, d\mu \leq \int_X f \, d\mu for each n. Taking the as n \to \infty produces \lim_{n \to \infty} \int_X f_n \, d\mu \leq \int_X f \, d\mu. Combining this with the from results in \int_X f \, d\mu \leq \lim_{n \to \infty} \int_X f_n \, d\mu \leq \int_X f \, d\mu, which implies equality: \int_X f \, d\mu = \lim_{n \to \infty} \int_X f_n \, d\mu. This establishes the monotone convergence theorem, providing an alternative perspective to the direct construction via simple functions in Beppo Levi's original approach. The utility of this proof lies in its simplicity and generality; by leveraging , it readily extends to settings where monotonicity is relaxed, such as in the , by incorporating an integrable dominating function to control the limsup.

Relaxing Monotonicity Assumptions

One key generalization of the monotone convergence theorem relaxes the strict monotonicity requirement by imposing a domination condition, leading to the . Specifically, if a of non-negative measurable functions f_n converges to f and there exists an integrable function g such that $0 \leq f_n \leq g for all n, then \lim_{n \to \infty} \int f_n \, d\mu = \int f \, d\mu. A further relaxation allows for sequences that are monotone almost everywhere. In this case, if f_n \uparrow f almost everywhere (i.e., f_{n+1}(x) \geq f_n(x) for all x outside a set of measure zero, and f_n \geq 0), the monotone convergence theorem still holds: \lim_{n \to \infty} \int f_n \, d\mu = \int f \, d\mu. This variant is particularly useful in L^1 spaces, where measure-zero exceptions do not affect integrability. Moreover, under the additional assumption that \sup_n \int f_n \, d\mu < \infty, the sequence converges in the L^1 norm, meaning \lim_{n \to \infty} \int |f_n - f| \, d\mu = 0. These relaxations can be established using Fatou's lemma, which provides a lower semicontinuity property for integrals without requiring full monotonicity. However, without a domination condition, even pointwise convergence of non-negative functions need not preserve integrals. A classic counterexample involves "moving bump" functions on \mathbb{R}, such as f_n(x) = \chi_{[n, n+1]}(x), which converge pointwise to 0 but satisfy \int f_n \, dx = 1 for all n, so the integrals do not converge to \int 0 \, dx = 0. Similar constructions, like f_n(x) = n \chi_{[1/n, 2/n]}(x) on [0,1], illustrate failures where no integrable dominator exists.

References

  1. [1]
    [PDF] Real Analysis MAA 6616 Lecture 15 Monotone Convergence ...
    If {fn} is ↗ and converges to f we write fn ↗ f. Theorem ( Monotone Convergence Theorem 2). Let {fn}n be a sequence of measurable, nonnegative functions on a ...
  2. [2]
    [PDF] 2. Convergence theorems - KSU Math
    Convergence theorems. 305. Theorem 2.2 (General Lebesgue Monotone Convergence Theorem). Let. (X, Л,µ) be a measure space, and let f,fn : X → [0, ∞], n ≥ 1 ...
  3. [3]
    [PDF] The Monotone Convergence Theorem - UMD MATH
    Idea: We know that if a sequence converges then it must be bounded. We also know the reverse is not true. However in the case of monotone sequences it is.Missing: real | Show results with:real
  4. [4]
    [PDF] 3.4 Monotone Sequences
    Mar 3, 2018 · Using the Monotone Convergence Theorem show that the sequence hsni converges, and then find its limit. Solution.
  5. [5]
    Monotonic Sequences
    The monotone convergence theorem provides a powerful one-two punch that is sufficient to prove that a sequence converges by proving two (probably) ...
  6. [6]
    [PDF] Math 341 Lecture #8 §2.4: The Monotone Convergence Theorem ...
    Theorem 2.4.2 (Monotone Convergence Theorem). If a sequence is monotone and bounded, then it converges. Proof. Suppose (an) is monotone and bounded.
  7. [7]
    2.3 monotone sequences and the Bolzano-Weierstraß Theorem
    If X is a monotone sequence, then X converges if and only if X is bounded. 🔗. 🔗. Proof. Observe that one direction is Theorem 2.1.14 ...
  8. [8]
    The Early Development of Set Theory
    Apr 10, 2007 · The form of completeness axiom that Weierstrass preferred was Bolzano's principle that a sequence of nested closed intervals in \(\mathbf{R}\) ( ...
  9. [9]
    [PDF] Series - UC Davis Math
    We say that a series converges if its sequence of partial sums converges, and in that case we define the sum of the series to be the limit of its partial sums.
  10. [10]
    [PDF] The Harmonic Series Diverges Again and Again
    In this article, the authors survey some of these divergence proofs. Throughout, Hn is used to denote the nth partial sum of the harmonic series. That is ...
  11. [11]
    [PDF] The exponential function for matrices - UCR Math Department
    Matrix exponentials provide a concise way of describing the solutions to systems of homoge- neous linear differential equations that parallels the use of ...Missing: monotone spectral
  12. [12]
    [PDF] 4 The exponential map - UC Berkeley math
    Example 52 If A is nilpotent, then exp(tA) is a copy of the real line, and its elements consist of unipotent matrices. In this case the exponential series is ...
  13. [13]
    [PDF] Spectral Mapping Theorem for the Matrix Exponential
    In the last segment, we saw that this makes sense, that is, that the series converges for all numbers t and all matrices A.Missing: monotone sums radius
  14. [14]
    [PDF] 1 Math 311:02 Fall 2008 The Exponential Series
    By the Monotone Convergence Theorem, to show that it converges, it will be enough to find an upper bound for the sequence of partial sums.Missing: Taylor | Show results with:Taylor
  15. [15]
    Beppo Levi (1875 - 1961) - Biography - MacTutor
    He had also studied the theory of integration, partial differential equations and the Dirichlet Principle, producing the famous "Beppo Levi theorem" and spaces ...
  16. [16]
    (PDF) Development of the Theory of the Functions of Real Variables ...
    In this paper a technical survey of the arising function and measure theory is given with a particular regard to the contribution of the Italian mathematicians ...
  17. [17]
    [PDF] 18.102 S2021 Lecture 11. The Lebesgue Integral of a Nonnegative ...
    Mar 30, 2021 · We saw last time already that for a nonnegative measurable function f , we can always find a sequence of simple functions that increase ...
  18. [18]
    [PDF] Lecture 3 The Lebesgue Integral
    Sep 28, 2013 · Apply the monotone convergence theorem to the partial sums gn = f1 + ··· + fn, and use linearity of integration. Once we have established a ...
  19. [19]
    [PDF] Real Analysis MAA 6616 Lecture 14 The Lebesgue Integral of ...
    Proof. Let Eλ = {f > λ}. consider two cases depending on wether the measure of Eλ is finite or not.<|control11|><|separator|>
  20. [20]
    [PDF] B Lebesgue Measure and Integral
    If the functions fn are nonnegative but are not monotone increasing, then we may not be able to interchange a limit and an integral. However, the following ...
  21. [21]
    [PDF] Chapter 4 Integration
    If f is a non-negative simple function RX fdµ = Iµ(f). Theorem 4.5. (Beppo-Levi Theorem) Let (X, J,µ) be a measure space and 1fjlj2N be.
  22. [22]
    [PDF] Integration - UC Davis Mathematics
    To show the linearity, we will first derive one of the fundamental convergence theorem for the Lebesgue integral, the monotone convergence theorem.Missing: Beppo Levi's<|control11|><|separator|>
  23. [23]
    [PDF] Lecture note on Analysis II 1 Why study measure theory?
    By Lebesgue's monotone convergence theorem, one has. Z. E. ¯g dµ = lim n→∞. Z ... Proof of Lebesgue differentiation theorem (L1 case): By the density theorem,.Missing: history | Show results with:history
  24. [24]
    [PDF] Lecture Notes on Real Analysis
    Theorem 4.9 (Beppo Levi: fk ↑ f ⇒ R fk ↑ R f). Suppose fk : E → ¯R+, fk ... Theorem 5.2 (Continuity of signed measure). If Ek is increasing and E ...<|control11|><|separator|>
  25. [25]
    [PDF] 2.3 Basic Properties of Measures - Christopher Heil
    Prove the following statements. (a) µ is a measure if and only if it satisfies continuity from below. (b) If µ(X) < ...
  26. [26]
    [PDF] Countable additivity of the integral under ... - DSpace@MIT
    This is a countably additive measure, because of countable additivity ... (Monotone Convergence Theorem) We are in the measure space (X, F, µ). ... Proof.
  27. [27]
    [PDF] Measure Theory John K. Hunter - UC Davis Math
    As stated in the following definition, an outer measure is a monotone, countably ... their Lebesgue measure in a way that preserves countable additivity (or even ...
  28. [28]
    [PDF] Lecture 22: Fatou's Lemma. Monotone Convergence Theorem.
    (Monotone Convergence Theorem). If {f} == is a x=1 sequence of ... =>. V is a measure in X. Proof: Check. that: (i) v($)=0 ✓. (ii) {Ex} disjoint ...
  29. [29]
    [PDF] Proofs-4-3.pdf - Real Analysis
    Monotone Convergence Theorem. Let {fn} be an increasing sequence of ... Beppo Levi's Lemma. Let {fn} be an increasing sequence of nonnegative.Missing: history | Show results with:history
  30. [30]
    [PDF] Fatou's Lemma and the Monotone Convergence Theorem §8
    ϕ dµ. This gives countable additivity. D. The Monotone Convergence Theorem. If {fn} is a sequence of nonnegative measurable functions defined on a measurable ...
  31. [31]
    [PDF] Fatou's Lemma, Dominated Convergence
    Monotone Convergence Theorem. Suppose that 0 ≤ f1 ≤ f2 ≤··· is a monotonically increasing sequence of non-negative measurable functions on Rn, and let.
  32. [32]
    [PDF] An introduction to measure theory Terence Tao
    the dominated convergence theorem (Theorem 1.4.49). Both branches of the ... ing any of the moving bump examples by −1). While this lemma was stated ...