Fact-checked by Grok 2 weeks ago

Lusin's theorem

Lusin's theorem is a fundamental result in real analysis that characterizes measurable functions by their approximation by continuous functions on sets of large measure. Specifically, for a measurable set E \subseteq \mathbb{R}^d of finite Lebesgue measure and a measurable function f: E \to \mathbb{C}, given any \varepsilon > 0, there exists a closed set F \subseteq E such that the measure of E \setminus F is less than \varepsilon and the restriction of f to F is continuous. Equivalently, f is measurable if and only if it satisfies this continuity condition on nearly all of its domain. Named after the Russian mathematician (1883–1950), the theorem was first proved by him in 1912 as an application of the Severini-Egorov theorem, with further developments appearing in his 1915 PhD thesis on integrals and trigonometric series. Lusin's theorem plays a central role in measure theory, underscoring one of Littlewood's three principles of , which highlight the intuitive properties of Lebesgue measurable functions: they are nearly continuous, nearly differentiable, and nearly convergent. The result extends to more general settings, such as \sigma-finite measures on locally compact Hausdorff spaces, where measurable functions can be approximated by continuous functions except on sets of small measure. In its proof, the theorem typically relies on approximating the measurable function by simple functions, applying Egorov's theorem for uniform convergence on a large set, and extending continuous restrictions appropriately. This approximation property is crucial for applications in , , and partial differential equations, as it bridges the gap between the abstract notion of measurability and the more concrete continuous functions. Variations and extensions of Lusin's theorem appear in modern contexts, such as C^k approximations for smooth functions or generalizations to non-Euclidean spaces.

Background

Prerequisites in Measure Theory

In measure theory on \mathbb{R}^n, the Borel \sigma-algebra \mathcal{B}(\mathbb{R}^n) forms a foundational structure, defined as the smallest \sigma-algebra containing all open subsets of \mathbb{R}^n. A \sigma-algebra on \mathbb{R}^n is a collection of subsets that includes the empty set and \mathbb{R}^n, and is closed under complements and countable unions. The Borel \sigma-algebra is generated by taking all countable unions, intersections, and complements starting from the open sets, resulting in the Borel sets, which include all closed sets, G_\delta sets (countable intersections of opens), and F_\sigma sets (countable unions of closeds). This structure ensures that topological properties of \mathbb{R}^n are preserved in a measurable framework, distinguishing Borel sets from more general Lebesgue measurable sets. To construct Lebesgue measure, one begins with the Lebesgue outer measure m^*, which assigns to any E \subset \mathbb{R}^n the infimum over all countable coverings of E by open rectangles (or cubes) of the sum of their volumes. Formally, for E \subset \mathbb{R}^n, m^*(E) = \inf \left\{ \sum_{k=1}^\infty |I_k| : E \subset \bigcup_{k=1}^\infty I_k, \, I_k \text{ open rectangles} \right\}, where |I_k| denotes the volume of I_k. This outer measure is subadditive and translation-invariant, providing a notion of "size" for all sets, though it is not always additive. The Carathéodory extension theorem then defines Lebesgue measurability and constructs the from the outer measure. A set A \subset \mathbb{R}^n is Lebesgue measurable if, for every E \subset \mathbb{R}^n, m^*(E) = m^*(E \cap A) + m^*(E \cap A^c), where A^c is the complement of A. The collection of all such measurable sets forms the Lebesgue \sigma-algebra \mathcal{L}(\mathbb{R}^n), which properly contains the Borel \sigma-algebra. The m is the restriction of m^* to \mathcal{L}(\mathbb{R}^n), which is countably additive, complete (subsets of measure-zero sets are measurable), and agrees with the elementary volume on rectangles. This extension uniquely determines Lebesgue measure on Borel sets via the uniqueness part of the theorem. A function f: \mathbb{R}^n \to \overline{\mathbb{R}} (where \overline{\mathbb{R}} is the extended reals) is Lebesgue measurable if the preimage f^{-1}(U) is Lebesgue measurable for every open U \subset \mathbb{R}, or equivalently, if \{x \in \mathbb{R}^n : f(x) > a\} is measurable for all a \in \mathbb{R}. Key properties include: the pointwise limit of a sequence of measurable functions is measurable; the composition of a with a continuous is measurable; and sums, products, and quotients (where defined) of measurable functions are measurable. Continuous functions on \mathbb{R}^n are measurable because preimages of opens under continuous maps are open, hence Borel and thus Lebesgue measurable. However, measurable functions differ fundamentally from continuous ones, as they may exhibit discontinuities on sets of positive measure, allowing for phenomena like the Dirichlet function (1 on rationals, 0 on irrationals), which is discontinuous everywhere but measurable. In \mathbb{R}^n with the , compact sets are precisely the closed and bounded subsets, by the Heine-Borel theorem. Compact sets have finite Lebesgue measure and play a crucial role in approximation arguments, as they allow for inner approximations of measurable sets. Specifically, for any Lebesgue measurable set A \subset \mathbb{R}^n with m(A) < \infty and any \varepsilon > 0, there exists a compact set K \subset A such that m(A \setminus K) < \varepsilon. This property stems from the regularity of Lebesgue measure, enabling the "sandwiching" of measurable sets between compact subsets and open supersets for precise measure calculations and convergence results. A related concept is that of a function being continuous on a set except for a set of measure zero: for a measurable set E \subset \mathbb{R}^n with m(E) < \infty, a function f: E \to \mathbb{R} is continuous except on a set of measure zero if the set of points in E where f fails to be continuous has zero. Such functions agree almost everywhere with continuous functions when restricted to E, preserving integrability properties while allowing isolated pathologies on negligible sets. Lusin's theorem establishes that every measurable function on a finite-measure set is continuous except on a set of measure zero, bridging measurability and continuity.

Historical Development

Nikolai Lusin first formulated the key ideas underlying his theorem during his studies in Paris from 1909 to 1912, where he was deeply influenced by 's foundational work on measurable functions and their relation to Riemann integrability in the early 1900s. Lusin's 1912 paper "Sur les propriétés des fonctions mesurables," published in the Comptes Rendus de l'Académie des Sciences, presented an initial version of the theorem as an application of the , establishing that measurable functions are nearly continuous. This work built directly on and subsequent results characterizing Riemann-integrable functions as bounded and continuous almost everywhere. Between 1912 and 1916, Lusin extended these ideas in his research on Denjoy integrals and what would later be termed Luzin sets, exploring the structure of measurable functions and sets of finite measure in connection with trigonometric series and integration theory. His 1915 doctoral thesis, The Integral and Trigonometric Series, submitted at Moscow University under Dmitri Egorov, culminated this period and was published in 1916 in Matematicheskii Sbornik (volume 30, pages 1–242), where he relaxed finiteness assumptions and refined the theorem's scope for functions on sets of finite measure. This publication appeared amid the emerging Russian school of analysis at Moscow University, where Lusin, mentored by Egorov, began fostering a seminar on function theory that attracted international talent. Lusin's efforts were shaped by the vibrant Russian analytical tradition, with key interactions including collaborations with Polish mathematician , who spent time in Moscow during World War I (1915–1918), including joining Lusin's seminar around 1916, and contributed an elementary proof of the theorem in 1922 using set-theoretic definitions of measurability. Lusin also engaged with through correspondence on descriptive set theory, influencing his approaches to Luzin sets—uncountable sets with countable intersections with every nowhere dense set—which emerged from this era's focus on effective sets without the axiom of choice. By the 1920s, Lusin's theorem evolved into more general forms within the burgeoning field of functional analysis, with Stefan Banach and others extending it to abstract measure spaces and linear operators, integrating it into frameworks for Banach spaces and Riesz representation theorems. These developments solidified the theorem's role in modern measure theory, as reflected in subsequent textbooks.

Statement

Classical Version

Lusin's theorem, originally formulated by Nikolai Lusin in 1912, addresses the regularity properties of Lebesgue measurable functions on sets of finite measure in the real line. Let E \subseteq \mathbb{R} be a measurable set with finite Lebesgue measure and let f: E \to \mathbb{R} be a Lebesgue measurable function that is finite almost everywhere on E; that is, the set \{x \in E : |f(x)| = \infty\} has Lebesgue measure zero. For every \epsilon > 0, there exists a F \subseteq E such that m(E \setminus F) < \epsilon and the restriction f|_F is continuous, where m denotes Lebesgue measure. This result implies that any Lebesgue measurable function on a set of finite measure is continuous when restricted to closed subsets whose complements (within the domain) have arbitrarily small Lebesgue measure, thereby concentrating the potential discontinuities on sets of negligible size. Lusin's theorem is equivalent to the characterization that a function finite almost everywhere on \mathbb{R} is Lebesgue measurable if and only if it is approximately continuous almost everywhere.

General Form

Lusin's theorem in its general form addresses the approximation of measurable functions by continuous ones in abstract measure-theoretic settings, extending beyond the Euclidean case to spaces equipped with suitable measures and topologies. Specifically, let X be a locally compact Hausdorff space with a Radon measure \mu such that \mu(X) < \infty. For a \mu-measurable function f: X \to \mathbb{C} that is finite \mu-almost everywhere, and for every \epsilon > 0, there exists a compact K \subset X such that \mu(X \setminus K) < \epsilon and the restriction f|_K is continuous (with respect to the product topology on \mathbb{C}). A Radon measure \mu on the locally compact Hausdorff space X is defined as a Borel measure that is finite on every compact subset of X, outer regular on all Borel sets (meaning every Borel set can be approximated from above by open sets of arbitrarily small measure difference), and inner regular on all open sets (approximable from below by compact sets). This regularity ensures that the measure interacts well with the topology, enabling the compact approximations central to the theorem. The theorem extends to \sigma-finite regular measures on normal Hausdorff spaces, where the space can be covered by countably many sets of finite measure, allowing the result to hold by applying the finite-measure case to each component and combining via the \sigma-finiteness. In complete measure spaces, the theorem applies directly to measurable functions without additional modifications, as completeness ensures that null sets are handled appropriately in the approximation. For complex-valued functions, the continuity of f|_K is understood with respect to the standard Euclidean topology on \mathbb{C} \cong \mathbb{R}^2, which is second-countable and metrizable, facilitating the uniform approximation on compact sets. The theorem fails in settings lacking local compactness, such as certain infinite-dimensional topological vector spaces, where compact sets may not suffice to approximate the measure of the space, preventing the existence of such continuous restrictions on sets of large measure. The classical version on the real line with Lebesgue measure arises as a special case of this general form.

Proof Overview

Core Ideas

Lusin's theorem asserts that every measurable function on a space with finite measure can be approximated by a continuous function in the sense that they agree except on a set of arbitrarily small measure. At its core, the theorem relies on the foundational approximation of measurable functions by simple functions, which are finite linear combinations of characteristic functions of measurable sets. This stepwise refinement allows for building towards continuous approximations, particularly when restricted to compact subsets where topological properties can be leveraged to ensure continuity. A key conceptual tool in this process is , which guarantees that pointwise convergence of measurable functions to a limit can be made uniform on subsets of large measure within spaces of finite measure. This uniform convergence facilitates the control needed to extend approximations from simple functions to continuous ones, ensuring the error is confined to negligible sets. Complementing this, the inner regularity of the plays a crucial role by allowing measurable sets to be approximated from within by compact sets, on which measurability can more readily imply continuity-like behavior. Central to the theorem's intuition is the notion of "Lusin sets," which are compact subsets where the measurable function restricts to a continuous function, while the complement has controlled small measure. These sets embody the theorem's strategy of localizing continuity on "large" compact domains, exploiting the structure of the measure space. Ultimately, Lusin's theorem bridges the global, set-theoretic property of measurability—which concerns behavior across entire spaces—with the local, pointwise property of continuity, revealing that measurable functions are "nearly continuous" in a measure-theoretic sense and thus more amenable to analysis and computation.

Step-by-Step Outline

The proof of the classical version of Lusin's theorem, stated for Lebesgue measurable functions on \mathbb{R}^d, relies on the inner regularity of Lebesgue measure, which ensures that every measurable set E of finite measure contains a compact subset K \subseteq E such that \mu(E \setminus K) < \epsilon for any \epsilon > 0. Step 1: Bounded functions and approximation by simple functions. Assume first that f: E \to \mathbb{R} is bounded and measurable, where E \subset \mathbb{R}^d is measurable with \mu(E) < \infty. By the standard approximation theorem for measurable functions, there exists a sequence of simple functions s_n = \sum_{k=1}^{m_n} a_{n,k} \chi_{E_{n,k}} such that s_n \to f pointwise almost everywhere on E and \|s_n\|_\infty \leq \|f\|_\infty. It suffices to prove the theorem for each simple function s_n, as the result for bounded f follows by applying Egorov's theorem to ensure uniform convergence on a large compact set where the continuous approximations align closely with f. Step 2: Simple functions and compact subsets via open covers. Consider a simple function s = \sum_{k=1}^m c_k \chi_{A_k}, where the A_k are disjoint measurable sets partitioning the support of s and each c_k \in \mathbb{R}. For each k, the inner regularity of yields a compact set K_k \subseteq A_k with \mu(A_k \setminus K_k) < \epsilon / (m \|s\|_\infty). To construct a continuous extension, cover the space with open sets U_k containing K_k such that the U_k are disjoint and s is constant on each U_k \cap \mathrm{supp}(s); compactness of each K_k allows selection of such a large compact subset inside one dominant open set where s is constant, enabling the definition of a continuous function g that agrees with s on \bigcup K_k by or direct construction on the compact set. Step 3: Unbounded functions via truncation and \sigma-additivity. For an unbounded measurable f: E \to \mathbb{R}, given \epsilon > 0, choose n large enough so that \mu(\{x \in E : |f(x)| > n\}) < \epsilon/2. Let E_n = \{x \in E : |f(x)| \leq n\}, on which f|_{E_n} is bounded and measurable. Apply the bounded case to f|_{E_n} to obtain a compact K \subseteq E_n with \mu(E_n \setminus K) < \epsilon/2, so \mu(E \setminus K) < \epsilon, and f|_K is continuous. Step 4: Extension to the general form. In the general setting of a locally compact X equipped with a \mu, the theorem extends by leveraging the separation properties of Hausdorff spaces. For a measurable f: X \to \mathbb{C} vanishing outside a set of finite measure, provides continuous functions with compact support that separate points and match f on compact subsets of almost full measure, reducing to the classical case via local compactness and regularity.

Examples and Applications

Basic Illustrations

A classic illustration of Lusin's theorem involves the characteristic function of the rational numbers, \chi_{\mathbb{Q}}, defined on the interval [0,1] by \chi_{\mathbb{Q}}(x) = 1 if x is rational and $0 otherwise. This function is Lebesgue measurable because the set of rationals has measure zero, yet it is discontinuous at every point in [0,1], as every neighborhood contains both rationals and irrationals. Lusin's theorem guarantees that for any \varepsilon > 0, there exists a compact set K \subset [0,1] with m([0,1] \setminus K) < \varepsilon such that \chi_{\mathbb{Q}}|_K is continuous. To construct such a K, enumerate the rationals in [0,1] as \{q_n\}_{n=1}^\infty and remove open intervals I_n centered at q_n with total length less than \varepsilon, ensuring the complement K = [0,1] \setminus \bigcup_n I_n is compact and contains no rationals, so \chi_{\mathbb{Q}} \equiv 0 on K, which is constant and thus continuous. For instance, with \varepsilon = 0.1, m(K) > 0.9. Another straightforward example is a on [0,1], such as f(x) = 0 for x \in [0, 0.5) and f(x) = 1 for x \in [0.5, 1], which is discontinuous only at the jump point x=0.5. This function is measurable as a , and Lusin's theorem applies directly: for \varepsilon > 0, remove a small open interval around $0.5 of length less than \varepsilon, yielding a compact K consisting of two closed subintervals where f is constant (hence continuous) on each component. The restriction f|_K is then continuous on the disconnected set K, as continuity is checked relative to the . Thomae's function, t(x) = 0 if x is irrational and t(x) = 1/q if x = p/q in lowest terms with q > 0, provides a computational example on [0,1]. It is continuous at irrationals but discontinuous at , a of measure zero. By Lusin's theorem, for \varepsilon > 0, a compact K \subset [0,1] exists with m([0,1] \setminus K) < \varepsilon such that t|_K is continuous. An explicit construction uses a fat Cantor set, a compact nowhere dense perfect set of positive measure containing only irrationals, obtained by iteratively removing open intervals around with total length \varepsilon; on this K, t \equiv 0, so it is continuous. For example, the Smith-Volterra-Cantor set has measure $1/2 and can be adjusted to achieve m(K) > 0.9. These examples visualize how Lusin's theorem "patches" discontinuities by restricting to a compact set of large measure, effectively ignoring a small exceptional set where irregularities occur, thereby rendering the continuous in a measure-theoretic sense. This patching underscores the theorem's role in bridging measurable and continuous s without altering the 's values substantially.

Advanced Uses in Analysis

Lusin's theorem plays a pivotal role in the proof of the by facilitating the approximation of integrable by ones, thereby reducing the general case to scenarios where is more straightforward. Specifically, for a locally integrable f on \mathbb{R}^n, the theorem allows the construction of a g that agrees with f outside a set of arbitrarily small measure, enabling the extension of Tietze's theorem to show that the averages over shrinking balls converge to f(x) . This approximation step is essential because satisfy the property directly via the or density arguments, and the error on the exceptional set is controlled by the maximal . In the context of the Riemann-Lebesgue lemma, Lusin's theorem supports proofs by enabling the dense approximation of L^1 functions by continuous compactly supported functions, for which the Fourier transform vanishes at infinity through integration by parts. For an f \in L^1(\mathbb{R}), the lemma states that \hat{f}(\xi) \to 0 as |\xi| \to \infty; Lusin's theorem ensures that f can be modified on a small measure set to become continuous, preserving the L^1 norm and allowing the integral against e^{-2\pi i \xi x} to be bounded and decay appropriately. This approach highlights the theorem's utility in bridging measurable and smooth approximations for oscillatory integrals. Lusin's theorem is integral to the Riesz representation theorem, particularly in representing positive linear functionals on C_c(X) for locally compact Hausdorff spaces X as Radon measures. In the proof, it is combined with extension principles to approximate measurable sets and functions defining the functional, ensuring that the induced measure agrees with the functional on continuous functions while controlling discrepancies on negligible sets. For a functional \Lambda on C_c(X), Lusin's theorem aids in constructing Lusin sets—compact subsets where measurable functions behave continuously—thus verifying the measure's regularity and uniqueness in the representation. This connection underscores the theorem's role in unifying functional analysis with measure theory. In Sobolev spaces, Lusin's theorem facilitates embeddings by providing Lusin-type approximations that embed weakly differentiable functions into continuous ones on compact subsets, essential for regularity results in W^{k,p}(\Omega) where kp > n. For instance, functions in Sobolev spaces can be approximated in the Lusin sense by or smooth functions outside sets of small measure, preserving norms and enabling compact embeddings into C(\overline{\Omega}) via the Rellich-Kondrachov theorem. This approximation is crucial for establishing higher regularity and boundedness in applications like PDE solutions, where measurable weak derivatives are "smoothed" without altering the space's properties. Counterexamples in lower regularity cases, such as W^{1,p} with p \leq n, illustrate boundaries where such embeddings fail, but Lusin's theorem delineates the for . Despite its versatility, Lusin's theorem has limitations in non-regular measures, where it may fail, leading to counterexamples in pathological settings. For non-\sigma-finite or non-inner regular measures, such as certain capacities or non-additive set functions on infinite-dimensional spaces, measurable functions cannot always be approximated by continuous ones on nearly full measure sets; for example, in Gaussian measures on Hilbert spaces, the theorem requires additional smoothness assumptions to hold. These failures highlight the necessity of regularity conditions, like those in measures, and have spurred extensions to "Lusin measures"—a class of \sigma-finite Riesz measures where the approximation property persists—emphasizing the theorem's dependence on the underlying measure's topological regularity.

References

  1. [1]
    [PDF] Lusin's Theorem
    Theorem 8 (Lusin's Theorem). Given a measurable set E ⊆ Rd and given f : E → C, the following statements are equivalent. (a) f is measurable.Missing: analysis | Show results with:analysis
  2. [2]
    [PDF] biography of nn luzin (1883-1950)
    Aug 13, 2012 · At approximately the same time, he proved what is now called Lusin's theorem in real analysis. His Ph.D. thesis entitled Integral and ...
  3. [3]
    [PDF] Extensions of the Lusin's Theorem, the Severini-Egorov's Theorem ...
    Oct 14, 2016 · The LT was first proved by Nikolai Lusin [3] in 1912 as an application of the SET. In 1916, Lusin [4] succeeded in slightly relaxing the ...
  4. [4]
    [PDF] Lecture Notes in Real Analysis - University of Texas at Austin
    Dec 8, 2014 · We now start with the proof of Lusin's theorem. Exercise 27. If f : E → R is simple and > 0 then there exists a closed set F ⊂ E such that m ...
  5. [5]
    [PDF] Proofs-3-3.pdf - Real Analysis
    Real Analysis. October 25, 2020. 8 / 10. Page 24. Lusin's Theorem. Lusin's Theorem. Lusin's Theorem. Let f be a real-valued measurable function on E. Then for ...Missing: statement | Show results with:statement
  6. [6]
    [PDF] A Ck LUSIN APPROXIMATION THEOREM FOR REAL ... - cvgmt
    Jun 3, 2022 · A Ck LUSIN APPROXIMATION THEOREM FOR REAL-VALUED. FUNCTIONS ... A large part of geometric analysis and geometric measure theory in Euclidean.
  7. [7]
    [PDF] Introduction to Real Analysis Chapter 10 - Christopher Heil
    Jan 25, 2020 · The Borel σ-algebra BRd on. Rd is the smallest σ-algebra that contains all of the open subsets of Rd.
  8. [8]
    [PDF] Sigma Algebras and Borel Sets.
    A σ-algebra contains ∅, its complement, and countable unions. Borel sets are the smallest σ-algebra containing open sets, and are the smallest σ-algebra ...
  9. [9]
    [PDF] Chapter 1. Open Sets, Closed Sets, and Borel Sets
    Aug 16, 2022 · Borel sets are the smallest σ-algebra containing all open sets of real numbers. An Fσ set is a countable union of closed sets, and a Gδ set is ...
  10. [10]
    [PDF] Chapter 2: Lebesgue Measure - UC Davis Math
    This theorem states that a set is Lebesgue measurable if and only if it can be approximated from the outside by an open set in such a way that the difference.
  11. [11]
    [PDF] Lebesgue Measure on Rn - UC Davis Math
    As we show below, L(Rn) is the completion of B(Rn) with respect to Lebesgue measure, meaning that we get all Lebesgue measurable sets by adjoining all subsets.
  12. [12]
    [PDF] The Caratheodory Construction of Measures
    Show that the measures provided by Theorem 5.2, from outer measures, are all complete. In particular, Lebesgue measure is complete. 9. If µ is a measure on (X, ...
  13. [13]
    [PDF] Lecture 3 - Measurable functions and their properties.
    If f : Rd → R is continuous then f is measurable. If f measurable and finite-valued and. Φ continuous, then Φ ◦ f is measurable1. 3. If {fn}n is measurable ...
  14. [14]
    [PDF] math 676 properties of measurable functions
    These properties follow from the definitions and the fact that the Lebesgue- measurable sets form a σ-algebra containing the open sets. Property 1. The ...<|control11|><|separator|>
  15. [15]
    [PDF] 3.4 Functions Equal Almost Everywhere - Christopher Heil
    Second, if we are given a function that is defined on all of X except for a subset Z that has measure zero then we can assign any values we like to f(x) for x ...
  16. [16]
    Nikolai Luzin (1883 - 1950) - Biography - MacTutor
    Nikolai Nikolaevich Luzin ... He returned to Moscow in 1914 and he completed his thesis The integral and trigonometric series which he submitted in 1915.Missing: Denjoy 1912-1916
  17. [17]
    Nicolai Nicolaevich Luzin and the Moscow school of the theory of ...
    This paper describes Luzin's early mathematical education at Moscow University and the three year period he spent abroad (mainly in Paris)
  18. [18]
    (PDF) Development of the Theory of the Functions of Real Variables ...
    Egoroff Theorem! The very simple proof by Lusin is similar to the preceding ... 26 In (Banach 1924) there is a simple demonstration of this fundamental theorem.
  19. [19]
    [PDF] MIRA.pdf - Measure, Integration & Real Analysis
    He has been a member of the Council of the American Mathematical Society and ... Theorem 63. Approximation by Simple Functions 65. Luzin's Theorem 66.
  20. [20]
    [PDF] A Comprehensive Overview of the Lebesgue Differentiation ... - arXiv
    Jun 30, 2024 · Lusin's theorem states that, given a measurable function f over A (a measurable bounded set) and ε > 0, there exists a compact K ⊆ A such that µ ...<|separator|>
  21. [21]
    [PDF] Shape Analysis, Lebesgue Integration and Absolute Continuity ...
    Jul 7, 2018 · Proposition 2.24 (Lusin's Theorem): Let f be a real-valued (not ex- tended) measurable function on a measurable set E. Then given ǫ > 0, there.
  22. [22]
    [PDF] 1. Chapter 7: Radon Measures - Stony Brook University
    X = locally compact Hausdorff space (LCH space) . Cc(X) = continuous ... Lusin's Theorem: Suppose that µ is a Radon measure on X and f : X → C is a ...
  23. [23]
    [PDF] October 9, 2018 1. Measures on Locally compact Hausdorff spaces ...
    Even though Lusin's theorem requires the measure to be com- plete, a little thought shows that for Theorem 1.2.2 we do not require (M,µ) to be complete. The ...
  24. [24]
    Measure Theory
    Then we adapted the idea to give an actual proof of the theorem. Lusin's theorem turned out to be a relatively straightforward consequence of Egorov's theorem, ...Missing: core | Show results with:core
  25. [25]
    [PDF] lecture notes on measure theory fall 2020 gautam iyer
    Proof of Lusin's theorem. Previous two lemmas. □. Proof of Lemma 6.31 ... The main idea behind the proof is as follows: Let µ(A) = λ(φ(A)). Lemma 11.21 ...
  26. [26]
    [PDF] Measure and Integral
    and C is called a fat Cantor set. Note that C is a compact nowhere ... compare to Lusin's theorem 2.53. Moreover, the function is unbounded in any.
  27. [27]
    [PDF] Lusin's theorem
    There are several ways to state this as a precise theorem, all (or mostly) bearing Lusin's name. We shall state and prove two of these. But first, some lemmas.<|separator|>
  28. [28]
    [PDF] functional analysis lecture notes: egoroff and lusin's theorems
    Now, since Radon measures are inner and outer regular, we can find a compact set K ⊆ A and an open set U ⊇ E such that µ(A\K) < ε and µ(U\E) < ε. Note that f is ...
  29. [29]
    [PDF] MAT205a, Fall 2019 Part V: Lecture 16, Following Folland, ch 7.1
    Riesz representation theorem for positive functionals. First we will show ... We want to combine the Lusin theorem with the following extension theorem.
  30. [30]
    [PDF] Lecture 11 - Functional analysis
    Now, our goal is to complete the proof of the Riesz–Markov–Kakutani theorem (The- orem 3.23) in the complex case. It will be another application of the Radon– ...
  31. [31]
    A note on Lusin-type approximation of Sobolev functions on ...
    Mar 15, 2021 · We establish new approximation results in the sense of Lusin for Sobolev functions f with on infinite-dimensional spaces equipped with Gaussian measures.