Fact-checked by Grok 2 weeks ago

Almost everywhere

In measure theory, a property holds almost everywhere (often abbreviated as a.e.) with respect to a measure \mu on a if the set of points where the property fails has \mu-measure zero. This concept allows mathematicians to disregard exceptional sets of negligible size when analyzing functions or properties over continuous domains. The notion is fundamental to , where two measurable functions that agree almost everywhere are considered equivalent, as they yield the same over any measurable set. This forms the basis for the space of integrable functions L^p, enabling rigorous treatment of limits and without concern for values on measure-zero sets. For instance, the Lebesgue of a non-negative is defined via functions that approximate it pointwise almost everywhere. In , "almost everywhere" corresponds to "," reflecting events or properties that occur with probability 1, excluding outcomes in a set of zero. This analogy underscores its applications in stochastic processes, where convergence ensures reliable limits for random variables. Notable theorems, such as Egorov's theorem, guarantee that can be made uniform on sets of positive measure by excluding small exceptional subsets. Examples abound in analysis: the fundamental theorem of calculus holds for absolutely continuous functions, with derivatives existing almost everywhere; similarly, monotone functions are differentiable almost everywhere with respect to Lebesgue measure. These properties highlight the concept's role in bridging classical with modern measure-theoretic rigor, facilitating proofs in and partial differential equations.

Introduction and Definition

Formal definition

In measure theory, the notion of a property holding "almost everywhere" is formulated in the context of a (X, \Sigma, \mu), where X is a set, \Sigma is a of subsets of X (a collection closed under complements and countable unions, containing \emptyset and X), and \mu: \Sigma \to [0, \infty] is a measure satisfying \mu(\emptyset) = 0 and countable additivity on disjoint measurable sets: \mu\left(\bigcup_{n=1}^\infty A_n\right) = \sum_{n=1}^\infty \mu(A_n) for pairwise disjoint A_n \in \Sigma. A property P holds \mu-almost everywhere on X, denoted P holds \mu-a.e., if the exceptional set where P fails has measure zero: \mu(\{x \in X \mid \neg P(x)\}) = 0. Equivalently, P(x) holds for all x \in X \setminus N, where N \in \Sigma and \mu(N) = 0. This framework allows functions or sets to be identified up to negligible differences. For instance, two measurable functions f, g: X \to \mathbb{R} satisfy f = g \mu-a.e. if \mu(\{x \in X \mid f(x) \neq g(x)\}) = 0. A simple case arises with indicator functions: the indicators \chi_A and \chi_B of measurable sets A, B \subseteq X are equal \mu-a.e. if \mu(A \triangle B) = 0, where A \triangle B = (A \setminus B) \cup (B \setminus A) is the symmetric difference.

Relation to measure zero sets

In measure theory, a null set, also known as a set of measure zero, is defined as a measurable set N in a measure space (X, \mathcal{M}, \mu) such that \mu(N) = 0. This means that N contributes negligibly to the total measure, as its measure is exactly zero, though N may be nonempty. Null sets can be constructed in various ways within standard measure spaces. The empty set \emptyset is the trivial example of a null set, since \mu(\emptyset) = 0 by the axioms of measure. In the context of Lebesgue measure on \mathbb{R}, any countable set—such as the rational numbers \mathbb{Q} or the set of integers \mathbb{Z}—is a null set, because it can be covered by countably many intervals with arbitrarily small total length. The concept of s is central to the notion of almost everywhere properties, where a property holds almost everywhere if it fails only on a . Consequently, the complement of a N in the space X has full measure, meaning \mu(X \setminus N) = \mu(X), so any that holds on X \setminus N is satisfied almost everywhere with respect to \mu. A key property ensuring the robustness of null sets is their closure under countable unions: if \{N_i\}_{i=1}^\infty is a countable collection of disjoint null sets, then \mu\left(\bigcup_{i=1}^\infty N_i\right) = \sum_{i=1}^\infty \mu(N_i) = 0 by countable additivity of the measure. More generally, even for non-disjoint null sets, subadditivity implies \mu\left(\bigcup_{i=1}^\infty N_i\right) \leq \sum_{i=1}^\infty \mu(N_i) = 0, confirming that the countable union remains a .

Core Properties

Equivalence relations

In measure theory, on a (X, \mathcal{A}, \mu), two measurable functions f, g: X \to \overline{\mathbb{R}} are said to be equal almost everywhere, denoted f \sim g, if the set \{x \in X \mid f(x) \neq g(x)\} has \mu-measure zero. This defines an on the set of measurable functions, as it is reflexive (the differing set is empty, hence null), symmetric (the differing sets coincide), and transitive. To verify transitivity, suppose f \sim g and g \sim h. Then the set where f(x) \neq h(x) is contained in the union of the sets where f(x) \neq g(x) and where g(x) \neq h(x), both of which are null sets. Since the countable union of null sets is null, this union has measure zero, so f \sim h. The equivalence classes under \sim partition the measurable functions, and the L^p spaces for $1 \leq p \leq \infty are defined as the quotient spaces of these classes, where elements are cosets = \{g \mid g \sim f\} with finite p-norm (understood almost everywhere). This quotient structure identifies functions that agree \mu-almost everywhere, ensuring well-defined operations and norms. The notion extends to measurable sets: two sets A, B \in \mathcal{A} satisfy A \sim B if \mu(A \Delta B) = 0, where \Delta denotes the (A \setminus B) \cup (B \setminus A). This relation is also an , with null symmetric difference implying A and B coincide almost everywhere.

Preservation under operations

In measure theory, properties that hold almost everywhere (a.e.) exhibit stability under various set operations, particularly in complete measure spaces where null sets are handled rigorously. Specifically, if sequences of sets A_n and B_n satisfy A_n \subseteq B_n a.e. for each n, meaning \mu(B_n \setminus A_n) = 0, then the countable \bigcup_n A_n \subseteq \bigcup_n B_n a.e., as the exceptional set \bigcup_n (B_n \setminus A_n) is a countable union of null sets and thus null. Similarly, for intersections, if A_n \supseteq B_n a.e., then \bigcap_n A_n \supseteq \bigcap_n B_n a.e., with the exceptional set again a countable union of null sets. These preservations follow from the countable of measures, ensuring that a.e. inclusions remain valid under countable unions and intersections. For operations on functions, almost everywhere convergence is preserved under pointwise addition and . If sequences of measurable functions f_n \to f a.e. and g_n \to g a.e., then f_n + g_n \to f + g a.e. and \alpha f_n \to \alpha f a.e. for any scalar \alpha. Moreover, if h is and f_n \to f a.e., then the compositions h \circ f_n \to h \circ f a.e., leveraging the continuity to control the exceptional sets where convergence fails. These properties ensure that a.e. limits behave algebraically like pointwise limits, modulo null sets, facilitating analysis in L^p spaces. A key inequality in integration theory is the for the Lebesgue : for a f on a (X, \mathcal{M}, \mu), \left| \int_X f \, d\mu \right| \leq \int_X |f| \, d\mu, with if f \geq 0 \mu-a.e., since then |f| = f almost everywhere and thus \int_X |f| \, d\mu = \int_X f \, d\mu = \left| \int_X f \, d\mu \right|. This underscores the non-negativity condition as essential for the to capture the full "size" without cancellation effects. Almost everywhere properties are also invariant under measure-preserving transformations. If T: X \to X is a measurable transformation that preserves the measure \mu, meaning \mu(T^{-1}(E)) = \mu(E) for all measurable E \subseteq X, then a property P holds \mu-a.e. on X if and only if P \circ T holds \mu-a.e. on X. This follows because the preimage under T of a null set is null, ensuring that exceptional sets are mapped to sets of measure zero. Such invariance is fundamental in ergodic theory, where it preserves integrals and convergence behaviors under dynamical systems.

Convergence and Sequences

Almost everywhere convergence

In measure theory, a sequence of measurable functions \{f_n\} on a measure space (X, \mathcal{M}, \mu) is said to converge \mu-almost everywhere (or pointwise almost everywhere) to a function f if the exceptional set E = \{x \in X : \lim_{n \to \infty} f_n(x) \neq f(x)\} satisfies \mu(E) = 0. This notion strengthens pointwise convergence by allowing failures only on a set of measure zero, which is negligible in the context of integration and limits under the measure \mu. Equivalently, almost everywhere convergence holds if, for every \varepsilon > 0, the set \{x \in X : \limsup_{n \to \infty} |f_n(x) - f(x)| > \varepsilon\} has \mu-measure zero. This characterization emphasizes that the sequence approaches f arbitrarily closely outside sets of arbitrarily small measure, capturing the "typical" behavior in the space. A key result characterizing almost everywhere convergence on spaces of finite measure is Egoroff's theorem, which states that if \mu(X) < \infty and \{f_n\} converges \mu-almost everywhere to f, then for every \delta > 0, there exists a measurable set E \subset X with \mu(E) < \delta such that \{f_n\} converges uniformly to f on X \setminus E. The theorem, originally proved by Dmitrii Egorov in 1911, bridges pointwise almost everywhere convergence and uniform convergence by isolating the non-convergent behavior on a controllable exceptional set. The explicit construction of the exceptional set E in Egoroff's theorem proceeds as follows: Given \delta > 0, for each integer k \geq 1, let \varepsilon_k = 1/k and \delta_k = \delta / 2^k. For each m \in \mathbb{N}, define E_m^{(k)} = \left\{ x \in X : \sup_{n \geq m} |f_n(x) - f(x)| \geq \varepsilon_k \right\}. Since \{f_n\} converges almost everywhere to f, \mu(E_m^{(k)}) \to 0 as m \to \infty. Choose m_k \in \mathbb{N} such that \mu(E_{m_k}^{(k)}) < \delta_k. Set E^{(k)} = E_{m_k}^{(k)}, and let E = \bigcup_{k=1}^\infty E^{(k)}. Then \mu(E) \leq \sum_{k=1}^\infty \delta_k = \delta. On X \setminus E, for every k \geq 1, \sup_{n \geq m_k} |f_n(x) - f(x)| < 1/k. To verify uniform convergence, fix \eta > 0 and choose K \in \mathbb{N} such that $1/K < \eta. Let M = \max\{ m_1, \dots, m_K \}. For all n \geq M and all x \in X \setminus E, |f_n(x) - f(x)| \leq \sup_{j \geq m_K} |f_j(x) - f(x)| < 1/K < \eta, since n \geq M \geq m_K and x \notin E^{(K)}. Thus, \{f_n\} converges uniformly to f on X \setminus E. This construction exploits the finite measure of X to bound the measures of the tail sets via countable subadditivity.

Relations to other convergence types

Almost everywhere (a.e.) convergence, also known as pointwise convergence almost everywhere, does not imply convergence in L^1, even on finite measure spaces. A standard counterexample is the sequence of functions f_n = n \chi_{[1/n, 2/n]} on [0,1] with Lebesgue measure, where each f_n converges pointwise to 0 almost everywhere, but \|f_n\|_1 = 1 for all n, so the sequence does not converge to 0 in L^1. In the context of probability spaces, a.e. convergence corresponds to almost sure (a.s.) convergence of random variables. A.s. convergence implies convergence in probability, but the converse does not hold. For the latter direction, consider the "typewriter sequence" on [0,1] with Lebesgue measure, defined by f_n = \chi_{[(n - 2^k)/2^k, (n - 2^k + 1)/2^k]} where k is chosen such that $2^k \leq n < 2^{k+1}; this sequence converges to 0 in probability (and even in L^1), but not pointwise almost everywhere, as every point in [0,1] is visited infinitely often by the indicators. The Borel–Cantelli lemmas provide key tools for establishing a.s. convergence in probability theory, particularly for sequences of events. The first lemma states that if \sum P(A_n) < \infty, then P(\limsup A_n) = 0, implying the events occur only finitely often almost surely; the second lemma, under independence, shows that if \sum P(A_n) = \infty, then P(\limsup A_n) = 1. These results are instrumental in proving a.s. convergence for sums or maxima of independent random variables. On probability spaces, a.s. convergence combined with uniform integrability of \{|X_n|\} implies L^1 convergence to the same limit. This is a consequence of Vitali's convergence theorem, which characterizes L^1 convergence via a.e. convergence, uniform integrability, and domination by an integrable function on sets of finite measure.

Examples and Applications

Classical examples

A fundamental example in Lebesgue measure theory is the set of rational numbers \mathbb{Q} in \mathbb{R}. This set is countable, and any countable set has Lebesgue measure zero. Consequently, the complement—the irrational numbers—has full measure in any interval, meaning properties holding on the irrationals hold almost everywhere on \mathbb{R}. Another illustrative example is the Dirichlet function on [0,1], defined as f(x) = 1 if x is rational and f(x) = 0 if x is irrational. This function equals the zero function almost everywhere because the rationals have Lebesgue measure zero. However, it is discontinuous at every point in [0,1], as every neighborhood contains both rational and irrational points where the function takes different values. Despite this, the Lebesgue integral \int_0^1 f(x) \, dx = 0, highlighting how almost everywhere equality enables integrability even for functions with unbounded variation in the Riemann sense. For sequences, consider f_n(x) = n^2 x (1 - x)^n on [0,1] with Lebesgue measure. For each fixed x \in [0,1), f_n(x) \to 0 as n \to \infty because the exponential decay of (1 - x)^n dominates the polynomial growth of n^2 x; at x=1, f_n(1) = 0 for all n. Thus, the sequence converges pointwise (and hence almost everywhere) to the zero function. However, the convergence is not uniform, as \sup_{x \in [0,1]} |f_n(x) - 0| \to \infty, with the maximum occurring near x \approx 1/n. Moreover, \int_0^1 f_n(x) \, dx = n^2 / ((n+1)(n+2)) \to 1 \neq 0, so the sequence does not converge to zero in L^1([0,1]), illustrating that almost everywhere convergence does not preserve integrals without additional conditions.

Applications in integration theory

In integration theory, the concept of almost everywhere properties plays a pivotal role in establishing foundational results for Lebesgue integrals on \mathbb{R}^n. The Lebesgue differentiation theorem asserts that for a locally integrable function f, the average value over balls shrinking to a point x recovers f(x) almost everywhere; specifically, at Lebesgue points, which form a set of full measure, f(x) = \lim_{r \to 0} \frac{1}{m(B(x,r))} \int_{B(x,r)} f \, dm, where m is the Lebesgue measure and B(x,r) is the ball of radius r centered at x. This theorem underpins the ability to interchange limits and integrals in many contexts, ensuring that pointwise behavior aligns with integral averages except on measure-zero sets. Lusin's theorem further illustrates the utility of almost everywhere continuity for measurable functions. It states that for any measurable function f on a finite measure set E and any \epsilon > 0, there exists a compact K \subset E with m(E \setminus K) < \epsilon such that the restriction of f to K is continuous. This "almost continuity" allows measurable functions to be approximated by continuous ones on sets of nearly full measure, facilitating the extension of Riemann integration techniques to the Lebesgue framework. The density of continuous functions in L^p spaces, for $1 \leq p < \infty, relies on almost everywhere to ensure robust approximations. Continuous functions with compact are dense in L^p(\mathbb{R}^n) in the L^p , and moreover, for any f \in L^p, there exists a of such continuous functions converging pointwise almost everywhere to f. This pointwise almost everywhere approximation strengthens the norm , enabling the use of smoother functions in limit processes while preserving essential . A key application arises in the analysis of monotone functions, where almost everywhere differentiability yields a variant of the fundamental theorem of calculus. Any monotone function f: [a,b] \to \mathbb{R} is differentiable almost everywhere, and its derivative f' satisfies \int_a^b f' \, dm = f(b) - f(a), with f' integrable over [a,b]. This result connects pointwise derivatives to global integral behavior, excluding discontinuities on sets of measure zero, and forms the basis for more general differentiation theorems in integration theory.

Advanced Formulations

Ultrafilter-based definition

In abstract settings, particularly in topology and set theory, the notion of "almost everywhere" can be formulated using ultrafilters without invoking a measure structure. A property P on a set X holds almost everywhere with respect to the Fréchet filter \mathcal{F} of cofinite subsets if the set \{x \in X \mid P(x)\} belongs to every ultrafilter \mathcal{U} on X that contains \mathcal{F}. Since every non-principal (or free) ultrafilter on an infinite set extends \mathcal{F}, this is equivalent to the set where P holds being cofinite. This ultrafilter-based perspective generalizes the classical discrete case, where "almost everywhere" means "all but finitely many." For sequences (x_n)_{n \in \mathbb{N}} in a topological space, convergence almost everywhere to a limit L—meaning pointwise convergence except on a finite set—occurs precisely when the sequence converges to L along every free ultrafilter on \mathbb{N}. Here, convergence along a free ultrafilter \mathcal{U} means that for every neighborhood V of L, the set \{n \in \mathbb{N} \mid x_n \in V\} belongs to \mathcal{U}. Free ultrafilters on \mathbb{N} are constructed via the axiom of choice by extending \mathcal{F}, providing a non-constructive way to capture "generic" behavior excluding finite exceptions. In standard measure spaces, this ultrafilter formulation aligns with measure-theoretic almost everywhere for principal ultrafilters, which concentrate on single points and correspond to Dirac measures; thus, properties holding along principal ultrafilters reduce to evaluation, mirroring how measure-zero sets (like singletons in atomless spaces) are negligible. However, the ultrafilter approach extends beyond measures, defining in arbitrary topological spaces—including non-Hausdorff ones—via ultrafilter limits, where a net or converges if every finer ultrafilter converges to the same point. This measure-free generalization proves useful in settings like and general convergence spaces, where traditional measure-theoretic tools do not apply.

Extensions to non-standard measures

In incomplete measure spaces, where the sigma-algebra does not necessarily contain all subsets of null sets, the concept of almost everywhere is extended by completing the measure space. The completion \overline{\mathcal{A}} of the sigma-algebra \mathcal{A} consists of all sets of the form A \triangle N, where A \in \mathcal{A} and N is a subset of a measurable , and the extended measure \overline{\mu} assigns to such sets the measure \mu(A). A property holds almost everywhere with respect to \overline{\mu} if the exceptional set has \overline{\mu}-measure zero, thereby incorporating subsets of that were previously non-measurable. This completion ensures that in the original space generate the ideal of negligible sets in the completed space, preserving the intuitive notion of "everywhere except on a set of measure zero." An alternative approach to defining almost everywhere properties arises in the context of outer measures, which precede the construction of a full measure space via the Carathéodory extension theorem. For an outer measure \mu^* on a set X, a property holds almost everywhere if the outer measure of the exceptional set—where the property fails—is zero, i.e., \mu^*(E) = 0. This definition applies even before restricting to the measurable sets and is particularly useful in settings where the sigma-algebra is generated from the outer measure, ensuring that negligible sets align with those of zero outer measure. Such an formulation allows for the treatment of properties on non-measurable sets while maintaining consistency with the inner measure space. In non-\sigma-finite measure spaces, where the space cannot be covered by countably many sets of finite measure, almost everywhere properties are often defined relative to local supports or restricted to \sigma-finite subspaces to avoid pathologies. For instance, in the context of on locally compact groups, which may not be \sigma-finite globally (e.g., on non-\sigma-compact groups), the almost everywhere notion is applied locally on compact subsets or relative to the modular function, ensuring invariance under group translations. This localized approach maintains the utility of almost everywhere and while respecting the infinite total measure. A key result extending almost everywhere to product spaces is provided by Fubini's theorem for product measures. If (X, \mathcal{A}, \mu) and (Y, \mathcal{B}, \nu) are \sigma-finite measure spaces with product measure \mu \times \nu on X \times Y, then for a f: X \times Y \to \mathbb{R}, the iterated integrals satisfy \int_X \left( \int_Y |f(x,y)| \, \nu(dy) \right) \mu(dx) = \int_Y \left( \int_X |f(x,y)| \, \mu(dx) \right) \nu(dy), and Fubini's theorem implies that for \mu-almost every x \in X, the section y \mapsto f(x,y) is \nu-integrable (i.e., \int_Y |f(x,y)| \, \nu(dy) < \infty), and similarly for \nu-almost every y \in Y; the iterated integrals then equal the \int_{X \times Y} |f| \, (\mu \times \nu). This ensures that properties holding almost everywhere in the product space propagate to almost all sections, facilitating computations in multivariable settings.

References

  1. [1]
    [PDF] Measure Theory John K. Hunter - UC Davis Math
    ... measure zero is said to hold almost everywhere, or a.e. for short. If we want to emphasize the measure, we say µ-a.e. In general, a subset of a set of ...
  2. [2]
    [PDF] MEASURE THEORY AND APPLICATIONS - ScholarWorks
    Since any two functions that are equal almost everywhere on a measurable set E, have the same integral on E, the distinctions between the functions are ...
  3. [3]
    [PDF] 4 Sequences of measurable functions - UCSB Math
    The Egorov's theorem is important in applications. It asserts that the. µ -a.e. convergence is uniform provided we throw away a set of arbitrary small measure.<|control11|><|separator|>
  4. [4]
    [PDF] applications of lebesgue measure to the cantor set and non ...
    “almost everywhere.” This is a measure-theoretic term which means that a property holds almost everywhere if it holds everywhere on a domain ex- cept on a ...
  5. [5]
    [PDF] Measure Theory John K. Hunter - UC Davis Math
    Measures are important not only because of their intrinsic geometrical and probabilistic significance, but because they allow us to define integrals.
  6. [6]
    [PDF] Measure Theory Princeton University MAT425 Lecture Notes
    Jan 10, 2025 · Definition 5.30 (Almost-everywhere). If for two measurable functions f,g : X → C we have. µ ({ x ∈ X | f (x) ̸= g (x) })=0 we say that f ...
  7. [7]
    [PDF] An Introduction to Measure Theory - Terry Tao
    every x, because the countable union of null sets is still a null set. Because of these properties, one can (as a rule of thumb) treat the almost universal ...
  8. [8]
    [PDF] An introduction to measure theory Terence Tao
    Definition 1.3.5 (Almost everywhere and support). A property P(x) of a point x ∈ Rd is said to hold (Lebesgue) almost everywhere in Rd, or for (Lebesgue) ...
  9. [9]
    [PDF] Real Analysis lecture notes for MA 645/646
    Unless stated otherwise “almost everywhere” means “almost everywhere with ... Here “almost everywhere” is with respect to Lebesgue measure. Sketch of ...
  10. [10]
    [PDF] Real Analysis
    In practice, we are thinking f as the equivalence class of all functions which are equal to f µ-almost everywhere in A. Thus Lp(A) actually consists of ...
  11. [11]
    [PDF] notes on measure theory and the lebesgue integral - People
    Let E∆F denote the symmetric difference of E and F: E∆F := (E \ F) ... (d) (Almost everywhere equivalence) If f(x) = g(x) for µ-almost every x ∈ X ...
  12. [12]
    [PDF] 1 Measurable Sets - TTU Math
    The symmetric difference m(E1∆E2) = 0 and any subset of a set of measure zero is measurable and has measure zero (Exercise 11.2#3). Therefore, E1 − E2 and E2 − ...
  13. [13]
    [PDF] Real Analysis
    Folland, Gerald B. Real analysis : modern techniques and their applications I Gerald B. Folland. - 2nd ed. p. em. - (Pure and applied mathematics).
  14. [14]
    On Iterated Limits of Measurable Mappings | Cambridge Core
    Nov 20, 2018 · Egoroff, D. T., Sur les suites de fonctions mesurables, C.R. Acad. Sci. Paris, 152(1911), 244-6.Google Scholar. 2. 2. Kelley, J., General ...<|control11|><|separator|>
  15. [15]
    245A, Notes 4: Modes of convergence - Terry Tao - WordPress.com
    Oct 2, 2010 · ... almost uniform convergence, pointwise almost everywhere convergence, or convergence in measure can imply {L^1} convergence. The escape to ...<|separator|>
  16. [16]
    [PDF] THE BOREL-CANTELLI LEMMA - Mathematics and Statistics
    P(En) = ∞. =⇒. P. (. E(S). ) = 1. that is,. P[En occurs infinitely often ]=1. Note: This result is useful for assessing almost sure convergence.
  17. [17]
    [PDF] Section 4.6. Uniform Integrability: The Vitali Convergence Theorem
    Nov 24, 2020 · In this section, we introduce a new condition on a set of functions (uniform integrability) which produces another convergence theorem that is ...Missing: almost | Show results with:almost
  18. [18]
    [PDF] Lebesgue Outer Measure and Lebesgue Measure.
    Example. If C is a countable set then m∗(C) = 0. In particular, the outer measure of the rational numbers is zero.
  19. [19]
    [PDF] 2 Lebesgue integration
    Integral of any measurable function over a set of measure 0 vanishes. Hence ... Thus the Dirichlet function is Lebesgue integrable but not Riemann.
  20. [20]
    CounterExamples: From Elementary Calculus to the Beginnings of ...
    ... THEORY Franz Holter-Koch GROUP INVERSES OF M-MATRICES AND THEIR ... modified Riemann function: √ n , x ∈ Q, x = m/n , n = 2k 1/ √ ˜ (x) ...<|separator|>
  21. [21]
    Prove $f_n(x) = n^2 x (1-x)^n$ does not converges uniformly
    Jun 21, 2018 · So the sequence of functions (fn) does not converge uniformly (i.e. in the supremum = infinity norm) to the pointwise limit, the only chance of ...Convergence of sequence of functions, $f_n(x) = n^2 x(1-nx) \dotsHow do I show $f_n(x)=n^2 x^n(1-x)$ pointwise converges to $0$ on ...More results from math.stackexchange.comMissing: almost everywhere
  22. [22]
    Leçons sur l'intégration et la recherche des fonctions primitives ...
    Mar 28, 2006 · Leçons sur l'intégration et la recherche des fonctions primitives, professées au Collège de France. by: Lebesgue, Henri Léon, 1875-1941.
  23. [23]
    [PDF] Intégrale, Longueur, aire - Internet Archive
    Intégrale, Longueur, Aire. 2e THÈSE. — Propositions ...
  24. [24]
  25. [25]
    245A, Notes 6: Outer measures, pre-measures, and product measures
    Oct 30, 2010 · In this set of notes, we will give the Carathéodory lemma, which constructs a countably additive measure from any abstract outer measure.
  26. [26]
    254A, Notes 3: Haar measure and the Peter-Weyl theorem - Terry Tao
    Sep 27, 2011 · Haar measures will help us build useful representations and useful metrics on locally compact groups {G}.
  27. [27]
    [PDF] Existence and uniqueness of Haar measure - UChicago Math
    Aug 31, 2010 · Introduction. The purpose of this paper is to prove existence and uniqueness of Haar measure on locally compact groups.<|separator|>
  28. [28]
    [PDF] Chapter 6 - Product Measures and Fubini's Theorem
    3) V is integrable for almost all y € Y, and yau is an. (almost everywhere defined) v-integrable function on Y. If any of these three conditions holds then ƒ ...