Fact-checked by Grok 2 weeks ago

Egorov's theorem

Egorov's theorem, also known as Egoroff's theorem, is a fundamental result in measure theory stating that if a sequence of measurable functions on a finite measure space converges pointwise almost everywhere to a limit function, then for every ε > 0, there exists a measurable subset of the space with measure less than ε outside of which the convergence is uniform. This theorem bridges the gap between pointwise almost everywhere convergence, which is weak, and uniform convergence, which is stronger, by guaranteeing the latter on a "large" subset where the exceptional set has arbitrarily small measure. The theorem is named after the Russian mathematician Dmitri Fyodorovich Egorov (1869–1931), who published a proof in 1911 in the Comptes Rendus de l'Académie des Sciences under the title "Sur les suites des fonctions mesurables." An independent proof had appeared earlier in 1910 by the Italian mathematician Carlo Severini, but the result became widely known through Egorov's work. Egorov, a professor at Moscow University and president of the Moscow Mathematical Society from 1922, made significant contributions to , , and the , though his devout faith led to under Soviet authorities, culminating in his arrest in 1930 and death in prison the following year. Egorov's theorem plays a key role in by facilitating the interchange of limits and integrals for sequences of functions, particularly in proofs involving the and . It applies specifically to finite measure spaces, highlighting the importance of finite measure in controlling behaviors, and has extensions to more general settings like σ-finite measures under additional conditions. The theorem underscores the subtle differences between various modes of in and remains a in advanced courses on measure and integration.

Introduction and Background

Overview of the Theorem

Egorov's theorem is a fundamental result in measure theory that addresses the convergence of sequences of measurable functions on finite measure spaces. Consider a measure space (X, \mathcal{A}, \mu) where \mu(X) < \infty, and a sequence of measurable functions \{f_n\} converging pointwise almost everywhere to a measurable function f, meaning \lim_{n \to \infty} f_n(x) = f(x) for all x \in X except on a set of \mu-measure zero. Pointwise almost everywhere convergence ensures the functions agree with the limit at nearly every point but does not guarantee uniformity, where the supremum of the differences \sup_{x \in X} |f_n(x) - f(x)| approaches zero as n \to \infty. Intuitively, the theorem asserts that on such finite measure spaces, the pointwise almost everywhere convergence can be strengthened to uniform convergence on a large of X, by excluding a set of arbitrarily small measure. Specifically, for any \epsilon > 0, there exists a measurable E \subseteq X with \mu(X \setminus E) < \epsilon such that \{f_n\} converges uniformly to f on E. This refinement allows for controlled approximation of the limit function while preserving most of the space's measure. The theorem holds significant importance in real analysis, particularly for analyzing sequences of functions in Lebesgue integration and L^p spaces, where it facilitates interchanging limits and integrals under weaker convergence conditions. By bridging pointwise almost everywhere and uniform convergence, it provides essential tools for proving results like the bounded convergence theorem.

Historical Development

Egorov's theorem, concerning the uniform convergence of sequences of measurable functions, traces its origins to the early 20th century amid the rapid advancement of measure theory following Henri Lebesgue's introduction of the integral in 1902. The first proof was given by Italian mathematician Carlo Severini in 1910, in a paper addressing sequences of orthogonal functions and their convergence properties on measurable sets. Severini's work demonstrated that pointwise convergence almost everywhere implies uniform convergence on a subset of large measure, though it remained relatively obscure outside Italian mathematical circles at the time. Independently, Russian mathematician Dmitri Egorov provided a proof in 1911, published in French, which gained wider recognition in the international community and led to the theorem being named after him. Egorov's contribution appeared in the context of his research on measurable functions and their limits, solidifying the result's place in . In 1924, Leonida Tonelli published a short note explicitly crediting Severini with the earlier proof, highlighting Severini's priority and prompting the alternative designation as the Severini–Egorov theorem in subsequent literature. The theorem's development continued through early generalizations that extended its scope within measure theory. In 1916, Nikolai Luzin relaxed certain conditions on the measure space, broadening applicability. Frigyes Riesz reformulated the result in the abstract measure space setting in 1922 and further refined it in 1928. Wacław Sierpiński offered another extension in 1928, focusing on properties of perfect sets. Later contributions included Pavel Korovkin's 1947 generalization to non-finite measures and Emmanuel Mokobodzki's 1970 version for more general convergence notions. These advancements reflected the evolving rigor of measure theory during a period when foundational questions about convergence and measurability were central to mathematical progress.

Core Theorem

Formal Statement

Egorov's theorem states that if (X, \mathcal{A}, \mu) is a measure space and E \subseteq X satisfies \mu(E) < \infty, and if \{f_n\} is a sequence of measurable functions from E to a separable metric space M that converges pointwise almost everywhere to a function f: E \to M, then f is measurable and for every \varepsilon > 0 there exists a measurable F \subseteq E such that \mu(E \setminus F) < \varepsilon and f_n converges uniformly to f on F. Pointwise almost everywhere convergence means that \mu(\{x \in E : \lim_{n \to \infty} f_n(x) \neq f(x)\}) = 0. The assumption that the codomain M is a separable metric space ensures that the limit function f is measurable with respect to the given \sigma-algebra on E.

Assumptions and Counterexamples

The finite measure assumption in is essential, as it ensures that pointwise almost everywhere convergence of a sequence of measurable functions implies nearly uniform convergence on the space. Without finite measure, the theorem fails, as the exceptional set required for uniform convergence cannot be controlled to have arbitrarily small measure. A standard counterexample illustrating this limitation occurs on the infinite measure space (\mathbb{R}, \mathcal{B}, m), where m denotes Lebesgue measure and \mathcal{B} the Borel \sigma-algebra. Consider the sequence of functions f_n = \chi_{[n, n+1]}, the characteristic function of the interval [n, n+1]. This sequence converges pointwise to 0 almost everywhere on \mathbb{R}, since for any fixed x \in \mathbb{R}, f_n(x) = 1 only when n \leq x < n+1, which holds for at most one n, and f_n(x) = 0 for all sufficiently large n. However, the convergence is not nearly uniform: to achieve uniform convergence to 0 on a set F \subseteq \mathbb{R} with m(\mathbb{R} \setminus F) < \varepsilon, F can omit only a set of measure less than \varepsilon. For \sup_{x \in F} f_n(x) \to 0, F must contain no points of [n, n+1] for all sufficiently large n, i.e., [n, n+1] \subseteq \mathbb{R} \setminus F for large n. But each [n, n+1] has measure 1, so excluding infinitely many requires m(\mathbb{R} \setminus F) = \infty, contradicting m(\mathbb{R} \setminus F) < \varepsilon. Thus, no such F exists for any \varepsilon > 0. The assumption that the codomain is a separable is also necessary to guarantee the measurability of the functions involved and to facilitate the proof's construction of measurable sets where uniform convergence holds. In non-separable s, pointwise limits of measurable functions may fail to be measurable, undermining the theorem's applicability. For instance, if the codomain lacks separability, the Borel \sigma-algebra may not align well with the measure space's structure, leading to potential breakdowns in the measurability of sets defined by convergence criteria like \{x : d(f_n(x), f(x)) < 1/k\}. Finally, Egorov's theorem requires only almost everywhere pointwise convergence rather than everywhere convergence, as null sets of non-convergence can be incorporated into the exceptional set of small measure without affecting the result. This relaxation is valid because the finite measure assumption allows the measure of such null sets to be absorbed into the controllable exceptional set, preserving the nearly uniform convergence on the complement.

Proof

To prove Egorov's theorem, consider a measure space (X, \mathcal{M}, \mu) with a measurable set E \subset X such that \mu(E) < \infty, and a sequence of measurable functions \{f_n\} converging pointwise almost everywhere to a function f on E. The limit f is measurable since the codomain M is a separable metric space and the f_n are measurable. Without loss of generality, assume the convergence holds pointwise everywhere on E by ignoring a null set. Fix \varepsilon > 0. For each positive integer k \geq 1 and each n \geq 1, define the set E_{n,k} = \{ x \in E : \sup_{m \geq n} d(f_m(x), f(x)) \geq 1/k \}. Each E_{n,k} is measurable since the functions are measurable. Now, for fixed k, consider the decreasing sequence of sets F_n = \bigcup_{m \geq n} E_{m,k}, so F_1 \supset F_2 \supset \cdots. The intersection \bigcap_{n=1}^\infty F_n = \limsup_{n \to \infty} E_{n,k} is the set of points x \in E for which \sup_{m \geq n} d(f_m(x), f(x)) \geq 1/k for infinitely many n. However, by the pointwise convergence of \{f_n\} to f on E, for almost every x \in E there exists N_x such that \sup_{m \geq N_x} d(f_m(x), f(x)) < 1/k, so x \notin F_{N_x + 1}. Thus, \mu(\bigcap_{n=1}^\infty F_n) = 0. Since \mu(E) < \infty, the continuity of measure from above implies \mu(F_n) \to 0 as n \to \infty. For each k \geq 1, choose an integer N_k \geq 1 such that \mu(F_{N_k}) = \mu\left( \bigcup_{m \geq N_k} E_{m,k} \right) < \varepsilon / 2^k. Define the exceptional set B = \bigcup_{k=1}^\infty \bigcup_{m \geq N_k} E_{m,k} = \bigcup_{k=1}^\infty F_{N_k}. By countable subadditivity of \mu, \mu(B) \leq \sum_{k=1}^\infty \mu(F_{N_k}) < \sum_{k=1}^\infty \frac{\varepsilon}{2^k} = \varepsilon. The set A = E \setminus B thus satisfies \mu(E \setminus A) < \varepsilon. It remains to verify that f_n \to f uniformly on A. Fix \delta > 0 and choose an integer K \geq 1 such that $1/K < \delta. Let M = \max\{N_1, N_2, \dots, N_K\}. For any x \in A and any n \geq M, since M \geq N_k for each k = 1, \dots, K, it follows that x \notin F_{N_k} for each such k, so x \notin E_{n,k} (as n \geq N_k). Thus, \sup_{m \geq n} d(f_m(x), f(x)) < 1/k for each k = 1, \dots, K. In particular, for k = K, d(f_n(x), f(x)) \leq \sup_{m \geq n} d(f_m(x), f(x)) < \frac{1}{K} < \delta. Since this holds for all x \in A, \sup_{x \in A} d(f_n(x), f(x)) < \delta for all n \geq M, establishing uniform convergence on A. To see that the convergence is uniform via the Cauchy criterion, note that for n, l \geq M with l \geq n, d(f_n(x), f_l(x)) \leq d(f_n(x), f(x)) + d(f_l(x), f(x)) < 2\delta for all x \in A, so the sequence is uniformly Cauchy on A, and hence uniformly convergent to f there, consistent with the pointwise limit.

Generalizations

Luzin's Statement

Nikolai Luzin provided a generalization of Egorov's theorem to σ-finite measure spaces, extending the result beyond sets of finite measure. In Luzin's version, let (E, \mathcal{A}, \mu) be a measure space where E is σ-finite, meaning E = \bigcup_{k=1}^\infty E_k with each E_k \in \mathcal{A} and \mu(E_k) < \infty. Suppose \{f_n\} is a sequence of measurable functions from E to a separable metric space that converges pointwise almost everywhere to a measurable function f: E \to the separable metric space. Then there exists a measurable set H \subseteq E with \mu(H) = 0 such that E \setminus H = \bigcup_{k=1}^\infty F_k, where each F_k \subseteq E is measurable and f_n converges uniformly to f on F_k. This generalization is obtained by decomposing the σ-finite set E into countably many subsets E_k of finite measure, applying the original Egorov's theorem on each E_k to obtain exceptional null sets H_k and subsets F_k \subseteq E_k \setminus H_k with uniform convergence, and then setting H = \bigcup_{k=1}^\infty H_k, which has measure zero as a countable union of null sets. The resulting F_k cover almost all of E outside the single null set H. The statement preserves the assumptions of measurability for the functions and the separable metric codomain, ensuring that uniform convergence is well-defined in the complete metric induced by the space.

Luzin's Proof

To prove Luzin's generalization of in a σ-finite measure space (X, \mathcal{A}, \mu), where f_n \to f pointwise almost everywhere, decompose X = \bigcup_{k=1}^\infty E_k with each E_k \in \mathcal{A} and \mu(E_k) < \infty. Let N \in \mathcal{A} be a measurable set with \mu(N) = 0 such that f_n(x) \to f(x) for all x \in X \setminus N. For each fixed k, consider the finite-measure set E_k \setminus N. Apply the core iteratively in a nested fashion to construct a set of full measure in E_k \setminus N on which the convergence is uniform. Specifically, start with H_0 = E_k \setminus N. For j = 1, 2, \dots, apply the core theorem to H_{j-1} with tolerance \varepsilon_j > 0 to obtain H_j \subseteq H_{j-1} such that \mu(H_{j-1} \setminus H_j) < \varepsilon_j and f_n \to f uniformly on H_j. Choose \varepsilon_j = \delta / 2^j for arbitrary \delta > 0. Then, \mu((E_k \setminus N) \setminus \bigcap_{j=1}^\infty H_j) \le \sum_{j=1}^\infty \varepsilon_j = \delta. Since \delta > 0 is arbitrary, it follows that \mu((E_k \setminus N) \setminus \bigcap_{j=1}^\infty H_j) = 0. Let G_k = \bigcap_{j=1}^\infty H_j \subseteq E_k \setminus N, so \mu(E_k \setminus G_k) = 0 and f_n \to f uniformly on G_k. The sets \{G_k\}_{k=1}^\infty are measurable, X = N \cup \bigcup_{k=1}^\infty G_k \cup \bigcup_{k=1}^\infty (E_k \setminus G_k), and the exceptional set H = N \cup \bigcup_{k=1}^\infty (E_k \setminus G_k) satisfies \mu(H) = 0. Moreover, f_n \to f uniformly on each G_k.

Korovkin's Statement

Korovkin's statement provides a of Egorov's that applies to abstract classes of measurable sets possessing specific measure-theoretic properties, thereby extending the control over exceptional sets beyond the standard sigma-algebra framework. Let E be a measurable set in a (X, \Sigma, \mu) with \mu(E) < \infty, and let \mathfrak{A} be a class of measurable subsets of E satisfying the following conditions: (i) E \in \mathfrak{A}; (ii) if A \in \mathfrak{A} and B \subseteq A, then either B \in \mathfrak{A} or \mu(A \setminus B) < \varepsilon for every \varepsilon > 0; (iii) \mathfrak{A} is closed under countable unions of sets each of finite measure. Suppose \{f_n\} is a sequence of s from E to a separable that converges to a f: E \to the separable . Then, for every \varepsilon > 0, there exists A' \in \mathfrak{A} such that $0 \leq \mu(E \setminus A') < \varepsilon and \{f_n\} converges uniformly to f on A'. This formulation generalizes the original theorem by incorporating classes \mathfrak{A} that behave like ideals or filters with respect to sets of small measure, allowing the exceptional sets to be managed through the structure of \mathfrak{A} rather than arbitrary measurable sets of small measure. Such abstraction enables applications in settings where the notion of "negligible" sets is defined via the properties of \mathfrak{A}, providing flexibility in controlling the measure of exceptions while preserving the essence of nearly uniform convergence. The statement retains key assumptions from the classical version, including the finite measure of E, the separability of the metric space for the function values (to ensure uniform convergence is well-defined), and the measurability of the functions involved.

Korovkin's Proof

Korovkin's proof begins by adapting the standard construction of exceptional sets from the core proof of Egorov's theorem to ensure compatibility with the class \mathfrak{A}. Given a sequence of measurable functions f_n converging \mu-almost everywhere to f on a set E \in \mathfrak{A} with \mu(E) < \infty, define the exceptional sets B_k = \limsup_{n \to \infty} \{ x \in E \mid |f_n(x) - f(x)| > 1/k \} for each positive integer k. Almost everywhere convergence implies that \mu(B_k) \to 0 as k \to \infty. To control the measure while staying within \mathfrak{A}, property (ii) of the class \mathfrak{A}—which guarantees that sets of arbitrarily small measure can be complemented by elements of \mathfrak{A}—ensures that for any \varepsilon > 0 and sufficiently large k, \mu(B_k) < \varepsilon and B_k^c \in \mathfrak{A}. This allows the complement B_k^c, where the convergence is "nearly uniform" in the sense of avoiding large deviations infinitely often, to be selected from the class. Property (iii), involving closure under relevant operations, further supports including such complements in \mathfrak{A}. For uniformity on a large subset, consider a finite intersection \bigcap_{j=1}^m B_j^c for fixed m. Since \mathfrak{A} is closed under finite intersections, this set A_m = \bigcap_{j=1}^m B_j^c \in \mathfrak{A}. On A_m, the sequence f_n converges uniformly to f with |f_n(x) - f(x)| \leq 1/m for all sufficiently large n, independent of x \in A_m. Moreover, \mu(E \setminus A_m) = \mu\left( \bigcup_{j=1}^m B_j \right) \leq \sum_{j=1}^m \mu(B_j) < \varepsilon for large enough m, by the subadditivity of the measure. Thus, A_m captures almost all of E while achieving the desired uniform bound. To integrate the almost everywhere convergence, note that the set of non-convergence points has measure zero and is contained in \bigcup_{k=1}^\infty B_k. Leveraging the countable union property of \mathfrak{A}, there exists a set N \in \mathfrak{A} with \mu(N) = 0 containing this null set, allowing exclusion within the class without affecting the measure. The uniform convergence then holds on A_m \setminus N, which remains in \mathfrak{A} up to negligible adjustment, completing the argument for arbitrary \varepsilon > 0.

Applications in Measure Theory

Egorov's theorem plays a pivotal role in the proof of , which asserts that any f that is finite on a locally compact can be approximated by s on sets of nearly full measure. To establish this, one constructs a of functions converging pointwise to f. Applying Egorov's theorem on compact subsets of finite measure ensures uniform convergence on a large measurable , excluding a set of arbitrarily small measure; this uniform limit can then be extended to a via the , yielding the desired approximation. In the context of the , Egorov's theorem provides essential control over to achieve almost everywhere differentiability of . Specifically, for an integrable function f \in L^1(\mathbb{R}^d), the theorem is applied to the sequence of averages of f over shrinking balls centered at points in a set of finite measure, converting almost everywhere to on a excluding a negligible exceptional set. This uniform control facilitates the verification that the limit equals f(x) almost everywhere, underpinning the theorem's conclusion that the derivative of the indefinite recovers f almost everywhere. Egorov's theorem also strengthens the connection between pointwise and L^1 convergence in the . Under the assumption of domination by an integrable g, the theorem allows selection of subsets where a pointwise convergent converges uniformly, bounding the difference on the exceptional set of small measure by the ; the uniform convergence then implies L^1 convergence on the complement, completing the proof for finite measure spaces. For \sigma-finite measure spaces, such as \mathbb{R}^n with , Luzin's version of Egorov's theorem extends its applicability by allowing the space to be covered (up to a set of measure zero) by countably many sets of finite measure on each of which the original applies, yielding on each such set. This enables the use of Egorov's in differentiation and results on infinite domains, such as establishing local uniform convergence properties essential for like the in \mathbb{R}^n. The provides a of norm convergence in L^p spaces ($1 \leq p < \infty) for sequences of measurable functions on a , requiring both convergence in measure and of the family \{|f_n|^p\}. This extends the pointwise-to-almost-uniform bridge of Egorov's theorem by ensuring L^p-norm convergence through the uniform integrability condition, which bounds the L^1 norms on sets of small measure and addresses cases where the ambient space may have infinite measure, unlike Egorov's finite-measure assumption. The dominated convergence theorem establishes that pointwise almost everywhere convergence of a sequence \{f_n\} to f, together with domination by an integrable function g (i.e., |f_n| \leq g a.e. for all n), implies convergence of the integrals \int f_n \to \int f. In contrast to Egorov's theorem, which facilitates uniform convergence on large-measure subsets without a global dominator, the dominated convergence theorem directly yields L^1 convergence under the stronger domination hypothesis; Egorov's result is frequently invoked in proofs of the dominated theorem to manage the lack of uniformity in the pointwise limit. In , almost sure convergence of random variables implies convergence in , and Egorov's theorem strengthens this by yielding almost uniform convergence on sets of close to 1, which is instrumental in establishing strong laws of where uniform control over sample paths is needed beyond mere probabilistic limits. It also plays a key role in the martingale convergence theorem, upgrading of martingales to almost uniform on sets of high , aiding applications of Doob's maximal inequalities. Modern variants include ideal versions of Egorov's theorem, where convergence is considered relative to ideals on the natural numbers (e.g., for statistical or convergence), ensuring almost uniform convergence holds within the ideal structure for sequences satisfying pointwise ideal limits. Additionally, in the Weyl-Hörmander calculus for pseudodifferential operators on , a 2024 generalization establishes Egorov-type propagation of singularities for evolution propagators under phase-space metric conditions, extending classical relations to quantum-mechanical settings.

References

  1. [1]
    [PDF] 3.9 Egoroff's Theorem - Christopher Heil
    Egoroff's Theorem essentially states that this example is typical, as long as we are dealing with a finite measure space. So it is important in the ex- ample ...
  2. [2]
    [PDF] LECTURE 13 Egoroff 's theorem (pointwise convergence is nearly ...
    Now we show that the theorems are valid if we replace convergence by convergence in measure. Theorem 0.4. Let fn : X. C be a sequence of measurable functions. →.
  3. [3]
    Dimitri Fedorovich Egorov (1869 - 1931) - Biography - MacTutor
    He graduated from the Sixth Moscow Gymnasium in 1887 having been awarded the gold medal. Then he entered Moscow University to study mathematics and physics.
  4. [4]
    [PDF] An introduction to measure theory Terence Tao
    A in Egorov's theorem cannot be taken to have zero measure, at least if one uses the bounded-set definition of local uniform convergence from Definition 1.3 ...
  5. [5]
    [PDF] Lecture Notes in Real Analysis - University of Texas at Austin
    Dec 8, 2014 · Theorem 7.2 (Third principle: Egorov's Theorem). Suppose {fn} is a sequence of measur- able functions defined on a set E ⊂ R of finite measure.
  6. [6]
    [PDF] MIRA.pdf - Measure, Integration & Real Analysis
    ... Egorov's Theorem 63. Approximation by Simple Functions 65. Luzin's Theorem 66. Lebesgue Measurable Functions 69. Exercises 2E 71. 3 Integration 73. 3A ...
  7. [7]
    [PDF] Real Analysis Nicola Arcozzi - https: //site.unibo.
    The Severini-Egorov Theorem. Almost everywhere, point- wise convergence is a weak notion. In 1910 Carlo Severini, and independently in 1911 Dmitri Egorov ...
  8. [8]
    The bounded convergence theorem. Amer. Math. Monthly 127 (2020 ...
    Dec 1924. Leonida Tonelli · View · An eigenvector proof of Fatou's lemma for continuous functions. Article. Jun 1995. Stephen Simons. In a first course on ...
  9. [9]
    None
    Summary of each segment:
  10. [10]
    [PDF] An Introduction to Measure Theory - Terry Tao
    Most of the material here is self-contained, assuming only an undergraduate knowledge in real analysis (and in particular, on the. Heine-Borel theorem, which we ...
  11. [11]
    Measure, Topology, and Differentiation (Chapter 7) - Real Analysis ...
    Egoroff, Dmitri (1911). Sur les suites de fonctions mesurables. C. R. Acad. Sci. Paris 152: 244–246. Fefferman, Charles (1971). On the convergence of ...
  12. [12]
    [PDF] Measure Theory John K. Hunter - UC Davis Math
    An alternative, and perhaps more illuminating, proof of the dominated conver- gence theorem may be obtained from Egoroff's theorem and the absolute continuity.
  13. [13]
    [PDF] Egorov's Theorem, a detailed proof.
    Theorem: Let (X, M,µ) be a measure space with µ(X) < ∞. Let {fn} be a sequence of measurable functions on X and let f be a measurable function on X. Assume ...Missing: Dmitri 1911
  14. [14]
    Egorov theorem - Encyclopedia of Mathematics
    Oct 18, 2012 · A theorem on the relation between the concepts of almost-everywhere convergence and uniform convergence of a sequence of functions. In ...Missing: Dmitri | Show results with:Dmitri
  15. [15]
  16. [16]
    real analysis - On Luzin's generalization of Egoroff's Theorem: the ...
    On Luzin's generalization of Egoroff's Theorem: the case of domain of infinite measure ; A)=+∞. Notice that, if ; A)=+∞, then it can not be true that ; Ak)<+∞ ...
  17. [17]
    Strange Functions in Real Analysis | Taylor & Francis Group
    Oct 11, 2017 · Kharazishvili, A. (2017). Strange Functions in Real Analysis (3rd ed.). Chapman and Hall/CRC. https://doi.org/10.1201/9781315154473. COPY ...
  18. [18]
  19. [19]
    [PDF] Lusin's theorem
    For each inte- ger N, Egorov's theorem implies the existence of a measurable set AN ⊆ (N,N +1) with λ(AN ) < 2−|N|ε so that Sn → f uniformly on (N,N +1)\ ...
  20. [20]
    [PDF] MATH 6337: Homework 6 Solutions 5.1. If f is a simple measurable ...
    Use Egorov's theorem to prove the bounded convergence theorem. Solution. Suppose {fk} is a sequence of measurable functions such that fk → f a.e. in E ...
  21. [21]
    [2412.04320] Egorov's theorem in the Weyl--Hörmander calculus
    We prove a general version of Egorov's theorem for evolution propagators in the Euclidean space, in the Weyl--Hörmander framework of metrics on the phase space.