Fact-checked by Grok 2 weeks ago

Non-measurable set

In measure theory, a non-measurable set is a of a , such as the real line equipped with , that cannot be assigned a measure while satisfying the defining properties of the measure, including countable additivity, translation invariance, and the Carathéodory measurability criterion, which requires that for every set E, the \mu^*(E) = \mu^*(E \cap A) + \mu^*(E \cap A^c). The existence of such sets depends on the and was first established in by Giuseppe Vitali, who constructed an explicit example known as the within the interval [0,1]. The is formed by partitioning the real numbers into equivalence classes under the relation x \sim y if x - y \in \mathbb{Q}, and selecting exactly one representative from each class that intersects [0,1] using the . To show its non-measurability, consider the countable collection of translates E + q_i, where q_i are the rational numbers in [-1,1]; these sets are disjoint, their union contains [0,1], and is contained in [-1,2]. If the Vitali set E were measurable with positive measure m(E) > 0, the measure of the union would be infinite due to countable additivity, exceeding the measure 3 of [-1,2]; conversely, if m(E) = 0, the union would have measure 0, contradicting the measure 1 of [0,1]. This construction demonstrates that the Lebesgue \sigma-algebra does not include all subsets of \mathbb{R}, leaving a proper subclass of the power set measurable. Non-measurable sets have profound implications in and , illustrating the limitations of extending measures to all subsets while preserving invariance under isometries. In 1914, proved that no countably additive measure defined on all subsets of \mathbb{R}^n (for n \geq 1) can be invariant under isometries of the space. Such sets also underpin paradoxical decompositions, like the Banach-Tarski theorem (1924), which shows that a solid ball in three dimensions can be partitioned into finitely many non-measurable pieces that can be reassembled into two balls of the same size, relying on the . Despite their pathological nature, non-measurable sets are essential for understanding the foundations of measure theory and the trade-offs in axiomatic .

Foundations of Measure Theory

Measurable Sets and σ-Algebras

In measure theory, a σ-algebra on a set X is a nonempty collection \mathcal{F} of subsets of X such that X \in \mathcal{F}, \mathcal{F} is closed under taking complements (if A \in \mathcal{F}, then X \setminus A \in \mathcal{F}), and \mathcal{F} is closed under countable unions (if A_n \in \mathcal{F} for n \in \mathbb{N}, then \bigcup_{n=1}^\infty A_n \in \mathcal{F}). This structure ensures that the collection is also closed under countable intersections, as the De Morgan laws relate unions of complements to intersections. The pair (X, \mathcal{F}) forms a measurable space, where the elements of \mathcal{F} are called the measurable sets. Measurable sets are the subsets belonging to some on X, providing the domain over which measures can be consistently defined while preserving additivity and other properties. Typically, a \mathcal{F} is generated by a smaller , such as an \mathcal{A} on X, which is a nonempty collection containing \emptyset and X, closed under finite unions, finite intersections, and complements. The generated , denoted \sigma(\mathcal{A}), is the smallest containing \mathcal{A}, obtained by including all sets formed through countable unions, intersections, and complements of elements in \mathcal{A}. To construct measures on such σ-algebras, plays a central role: given a semi-ring \mathcal{S} of subsets of X (a nonempty collection closed under finite intersections, where the difference of any two sets is a finite disjoint union of sets in \mathcal{S}) and a σ-finite premeasure \mu_0: \mathcal{S} \to [0, \infty) (countably additive on \mathcal{S}), there exists a unique measure \mu on the σ-algebra \sigma(\mathcal{S}) generated by \mathcal{S} that extends \mu_0. This theorem enables the extension of elementary content functions to full measures while ensuring uniqueness under σ-finiteness. A prominent example is the Borel σ-algebra \mathcal{B}(\mathbb{R}) on the real line \mathbb{R}, defined as the σ-algebra generated by the collection of all open intervals (a, b) with a, b \in \mathbb{R}. Equivalently, it is the smallest σ-algebra containing all open sets in the standard topology on \mathbb{R}. The Lebesgue σ-algebra arises as the completion of this Borel σ-algebra with respect to the .

Lebesgue Measure Properties

The \lambda is the standard measure on \mathbb{R}^n defined on the Lebesgue \sigma-algebra, which comprises all Lebesgue measurable subsets of \mathbb{R}^n. It extends the classical notion of volume for rectangles, where for a rectangle R = \prod_{i=1}^n [a_i, b_i], \lambda(R) = \prod_{i=1}^n (b_i - a_i). It is the unique complete translation-invariant measure on the Lebesgue \sigma-algebra that assigns measure 1 to the unit cube. Key properties of the Lebesgue measure include \sigma-additivity, which states that for a countable collection of pairwise disjoint measurable sets \{E_i\}_{i=1}^\infty, \lambda\left(\bigcup_{i=1}^\infty E_i\right) = \sum_{i=1}^\infty \lambda(E_i). Translation invariance holds such that for any measurable set E \subseteq \mathbb{R}^n and vector x \in \mathbb{R}^n, \lambda(E + x) = \lambda(E). Monotonicity ensures that if E \subseteq F with both sets measurable, then \lambda(E) \leq \lambda(F). Additionally, the Lebesgue measure is complete: any subset of a measurable set of measure zero is itself measurable with measure zero. The construction begins with the Lebesgue outer measure \lambda^*, defined for any set E \subseteq \mathbb{R}^n by \lambda^*(E) = \inf\left\{ \sum_{k=1}^\infty \lambda(R_k) : E \subseteq \bigcup_{k=1}^\infty R_k, \, R_k \text{ rectangles} \right\}, where the infimum is over all countable covers of E by rectangles and \lambda(R_k) is the volume of R_k. This outer measure approximates the "size" of E from above using volumes of covering rectangles. A set E \subseteq \mathbb{R}^n is Lebesgue measurable if and only if : for every test set A \subseteq \mathbb{R}^n, \lambda^*(A) = \lambda^*(A \cap E) + \lambda^*(A \cap E^c). Non-measurable sets violate this condition for some A, failing to split the outer measure additively and thus lying outside the domain of \lambda. The measurable sets form the Lebesgue \sigma-algebra, on which \lambda coincides with \lambda^* and inherits the listed properties.

Existence and Construction

Role of the Axiom of Choice

The (AC) asserts that, given any collection of nonempty sets, there exists a choice function that selects exactly one element from each set in the collection. This principle, formulated by in 1904 to prove the , enables the formalization of selections from infinite families of sets without explicit construction. AC plays a pivotal role in establishing the existence of non-Lebesgue measurable sets within Zermelo-Fraenkel with choice (ZFC). A key proof sketch relies on equivalence relations to partition the unit interval [0,1]. Consider the relation x \sim y if x - y \in \mathbb{Q}, ; this yields uncountably many equivalence classes, each countable and dense in [0,1]. By AC, a selector set V \subseteq [0,1] exists, containing precisely one representative from each class. The rational translates \{V + q : q \in \mathbb{Q} \cap [-1,1]\} form a countable disjoint cover of [0,1] (up to measure zero sets), and invariance under translations implies all translates have equal measure. If V were measurable with measure \mu(V) > 0, then \mu([0,1]) \geq \sum \mu(V + q) = \infty, a ; if \mu(V) = 0, the cover would have measure zero, contradicting \mu([0,1]) = 1. Thus, V is non-measurable. This argument demonstrates that AC implies the existence of non-Lebesgue measurable sets, highlighting its foundational necessity in measure theory. Historically, Zermelo's 1904 formulation of AC quickly influenced early measure theory, with Giuseppe Vitali employing a similar selection principle in 1905 to construct pathological sets, though explicit acknowledgment of AC grew prominent by the 1920s amid debates over paradoxes like Banach-Tarski. In the absence of AC—such as in Zermelo-Fraenkel set theory (ZF) alone—the existence of non-measurable sets is not provable; Robert Solovay constructed a model of ZF plus the axiom of dependent choices (DC) in 1970 where every set of reals is Lebesgue measurable, assuming the consistency of an inaccessible cardinal. This independence result underscores AC's indispensability for asserting non-measurability in standard set-theoretic foundations.

Vitali Set Construction

The Vitali set provides the first explicit example of a non-Lebesgue measurable subset of the real line, constructed by Giuseppe Vitali in his 1905 doctoral thesis. Vitali's construction relies on partitioning the interval [0,1] into equivalence classes based on rational translations and selecting representatives using the axiom of choice. To construct the Vitali set V \subset [0,1], first define an equivalence relation on \mathbb{R} by declaring x \sim y if x - y \in \mathbb{Q}. This partitions \mathbb{R} into uncountably many equivalence classes = \{x + q \mid q \in \mathbb{Q}\}. Restricting to [0,1], each class intersects [0,1] in a countable dense set, specifically \cap [0,1] = \{x + q \mod 1 \mid q \in \mathbb{Q} \cap [-1,1]\}, where the modulo 1 operation ensures elements remain in [0,1]. Using the axiom of choice, select exactly one representative from each distinct equivalence class intersecting [0,1] to form the set V. This V is uncountable, as there are uncountably many such classes (since \mathbb{R}/\mathbb{Q} has cardinality $2^{\aleph_0}). The non-measurability of V follows from considering its rational translates. For each q \in \mathbb{Q} \cap [0,1), the sets V + q = \{v + q \mod 1 \mid v \in V\} are pairwise disjoint, because if (v_1 + q_1) \mod 1 = (v_2 + q_2) \mod 1 with q_1 \neq q_2, then v_1 - v_2 = q_2 - q_1 \in \mathbb{Q}, implying v_1 = v_2 and q_1 = q_2 by the choice of distinct representatives. Moreover, the countable union \bigcup_{q \in \mathbb{Q} \cap [0,1)} (V + q) equals [0,1], as every x \in [0,1] belongs to some equivalence class with representative v \in V, so x = v + q \mod 1 for a unique q \in \mathbb{Q} \cap [0,1). Assume for contradiction that V is Lebesgue measurable with measure \lambda(V) = m. Then each translate V + q has the same measure m, since Lebesgue measure is translation-invariant. The union is a countable disjoint cover of [0,1], so \lambda([0,1]) = \sum_{q \in \mathbb{Q} \cap [0,1)} m = 1. If m = 0, the sum is 0, contradicting \lambda([0,1]) = 1; if m > 0, the sum is infinite (as there are countably infinitely many terms), again a contradiction. Thus, V cannot be Lebesgue measurable. This construction highlights the "pathological" nature of V, which evades intuitive geometric intuition while highlighting limitations in extending measure to all subsets of \mathbb{R}.

Key Examples and Paradoxes

Banach-Tarski Paradox

The Banach-Tarski paradox asserts that, given the , it is possible to decompose the closed unit ball in \mathbb{R}^3 into a finite number of disjoint pieces and reassemble them using only rigid motions (rotations and translations) to form two copies of the original unit ball. This counterintuitive result, first proved in 1924 by and , highlights the profound implications of non-measurable sets in , as the decomposition violates the intuitive notion of preservation. The pieces involved are necessarily non-measurable with respect to the , since any measurable decomposition would preserve total under isometries, making it impossible to double the ball without adding material. The proof hinges on the notion of equidecomposability, where two sets A and B are equidecomposable if A can be partitioned into finitely many subsets that can be rearranged via isometries to form B. Banach and Tarski's begins by considering the unit S^2 (excluding a of poles to avoid fixed points) and exploits the action of the special SO(3) on it, which contains a on two generators corresponding to independent rotations (e.g., by \arccos(1/3) around perpendicular axes). This admits a paradoxical into finitely many pieces that can be reassembled to form two copies of itself, a property derived from the non-amenability of the group. Extending this to the ball involves radial projections from the sphere to fill the interior, yielding the full decomposition into five pieces, which is the minimal number. This result builds on an earlier paradoxical decomposition discovered by in 1914, who showed that S^1 in \mathbb{R}^2 minus a can be partitioned into finitely many pieces equidecomposable via rotations to two copies of itself. Banach and Tarski extended Hausdorff's ideas from the one-dimensional to three dimensions, incorporating the to select representatives from equivalence classes under group actions, but without providing an explicit geometric description of the pieces themselves. The reliance on such abstract selectors underscores the paradox's dependence on set-theoretic , as the pieces defy intuitive and .

Hausdorff Paradox

The Hausdorff paradox provides an early and striking example of a paradoxical decomposition in the context of group actions on spheres, highlighting the challenges in defining consistent measures for all subsets. In 1914, showed that the 2-sphere S^2 \subset \mathbb{R}^3 admits a decomposition into non-measurable sets that defy intuitive notions of surface area under rotations. Specifically, there exists a countable subset D \subset S^2 such that S^2 \setminus D can be partitioned into three disjoint sets A, B, and C, where there are rotations \phi, \psi \in \mathrm{SO}(3) satisfying \phi(A) = C and \psi(B) = C. This implies that the union A \cup B is congruent via rotation to C, yielding the paradoxical equidecomposability where two disjoint pieces together match a single piece in size. Such a decomposition contradicts the existence of a finitely additive, rotation-invariant measure defined on all subsets of S^2 that agrees with the standard surface measure on measurable sets. If such a measure \mu existed, it would satisfy \mu(A) + \mu(B) = \mu(C), but also \mu(C) = \mu(A) and \mu(C) = \mu(B), implying $2\mu(A) = \mu(A), so \mu(A) = 0; repeating for B and extending to the whole sphere leads to \mu(S^2 \setminus D) = 0, which is impossible for a positive measure on a set of full surface area. Thus, the sets A, B, and C must be non-measurable with respect to the Lebesgue surface measure on S^2, underscoring the necessity of restricting measures to σ-algebras excluding such pathological sets. The construction of this decomposition leverages the group action of \mathrm{SO}(3), the group of rotations in three dimensions, on S^2. Hausdorff embedded a free group on two generators into \mathrm{SO}(3) by selecting specific rotations, such as a 180-degree rotation \phi around one axis and a 120-degree rotation \psi around another axis that do not commute and generate a free subgroup F_2. The set D consists of the countable collection of fixed points of nontrivial elements in this subgroup (the poles of the rotation axes). The action of F_2 on S^2 \setminus D is free, partitioning it into orbits; using the axiom of choice, one selects a transversal (system of representatives) for these orbits. A countably infinite paradoxical decomposition of F_2 itself—where the group is the disjoint union of sets of words beginning with powers of the generators and their inverses, such as F_2 = W(\phi) \cup W(\phi^{-1}) \cup W(\psi \phi^{-1}) with appropriate shifts—translates via the transversal to yield the desired partition of S^2 \setminus D into A, B, and C. Hausdorff extended a similar paradoxical construction to the 1-sphere S^1, the unit in \mathbb{R}^2. Here, S^1 minus a can be partitioned into two disjoint non-measurable pieces, each congruent via rotation in \mathrm{SO}(2) to the entire minus a single point. This variant again implies non-measurability, as it would violate additivity for any rotation-invariant length measure on all subsets. The Hausdorff paradox served as a crucial precursor to higher-dimensional analogs, directly inspiring Stefan Banach and Alfred Tarski to prove in 1924 that the three-dimensional ball admits a paradoxical decomposition into finitely many pieces, each congruent to the whole via rotations and translations.

Theoretical Implications

Measure Extension Problems

The Lebesgue measure is defined on the σ-algebra of Lebesgue measurable sets, which has cardinality $2^{2^{\aleph_0}}, the same as the power set of \mathbb{R}. Despite this, the Lebesgue σ-algebra does not contain all subsets of \mathbb{R}, as the existence of non-measurable sets, such as the Vitali set, serves as a fundamental obstacle to such an extension. The core challenge lies in extending the measure to the full power set while preserving key properties like σ-additivity and translation invariance. A central result in this area is Banach's measure problem, which asks whether there exists a countably additive, translation-invariant measure defined on all subsets of \mathbb{R} that extends the and satisfies \lambda([0,1]) = 1. The answer is negative in with the (ZFC): no such extension exists, as demonstrated by the contradiction arising from assigning measures to equivalence classes under rational translations, which would violate additivity. This impossibility highlights the tension between the desire for a complete and the preservation of geometric invariances inherent to the real line. Ulam's theorem provides a deeper insight, stating that there is no σ-additive extension of the to all subsets of \mathbb{R} that is translation-invariant and finite on compact sets. Specifically, if such an extension existed, the would need to be a real-valued measurable , which is inconsistent with standard axioms like ZFC without additional assumptions. This result underscores the foundational limitations of measure theory on the real line, linking extension problems directly to cardinal characteristics. Regarding consistency, while ZFC proves the existence of non-measurable sets, weaker systems allow for models where all subsets of \mathbb{R} are Lebesgue measurable. In particular, it is consistent with ZF plus the axiom of dependent choices (ZF + DC) that every set of reals is Lebesgue measurable, relative to the existence of an inaccessible cardinal; this is achieved in Solovay's model, constructed via a forcing extension that collapses the inaccessible to the continuum while preserving measurability for all sets. In such models, extensions trivially exist since the Lebesgue σ-algebra coincides with the power set, but translation invariance may not hold universally without the axiom of choice. Historical attempts to broaden measurability include Suslin's work on analytic sets, introduced in response to a flaw in Lebesgue's assumption that projections of Borel sets are Borel measurable. Suslin proved that analytic sets—continuous images of Borel sets—are always Lebesgue measurable, providing a significant enlargement of the measurable class beyond the Borel while maintaining desirable properties like the perfect set theorem. This development influenced later efforts to characterize measurable sets but did not resolve the full extension problem to arbitrary subsets.

Applications in Probability Theory

In , a is defined as a countably additive on a of events that assigns values between 0 and 1, with the entire having measure 1; consequently, non-measurable sets are excluded from consideration, as they cannot be assigned a probability in this framework. This axiomatic approach, introduced by Kolmogorov in 1933, ensures that aligns with measure theory, providing tools for and limit theorems while deliberately avoiding the ambiguities introduced by non-measurable sets. To achieve consistent probability assignments for all subsets of the , including non-measurable ones, alternatives to countable additivity are employed, such as finitely additive extensions of the . Using the , Banach constructed such extensions that are translation-invariant and defined on the power set of \mathbb{R}^n for n=1,2, normalizing them to yield probability measures on all subsets while preserving values on measurable sets. These finitely additive measures, often termed Banach measures, allow probabilistic interpretations beyond σ-algebras but sacrifice countable additivity, which is essential for many results. In subjective probability, de Finetti further advocated finitely additive measures as coherent betting systems, extending to all events without reliance on σ-algebras. The requirement of measurability extends to random variables, which are defined as measurable functions from the probability space to \mathbb{R}; non-measurable functions lack well-defined Lebesgue integrals, rendering expectations undefined and preventing standard probabilistic analysis. For instance, if a function X: \Omega \to \mathbb{R} is not measurable with respect to the σ-algebra, the integral \int X \, dP cannot be computed via the usual approximation by simple functions, leading to inconsistencies in computing means or variances. In the study of processes, assumptions of measurability are imposed to circumvent arising from the , such as non-measurable sample paths that defy intuitive probabilistic behavior. Standard constructions require processes to be measurable with respect to the product on the path space or progressively measurable with respect to a , ensuring that integrals and limits remain well-defined. This approach maintains the rigor of Kolmogorov's framework, excluding axiom-of-choice-dependent anomalies while enabling applications like .

References

  1. [1]
    [PDF] Measure Theory John K. Hunter - UC Davis Math
    The following is the standard example of a non-Lebesgue measurable set, due to Vitali (1905). Example 2.17. Define an equivalence relation ∼ on R by x ∼ y ...
  2. [2]
    [PDF] An introduction to measure theory Terence Tao
    Measure theory a non-measurable set is to exhibit a bounded set which is not almost elementary. Intuitively, we want to build a set which has oscillatory.
  3. [3]
    Sigma-Algebra -- from Wolfram MathWorld
    Let X be a set. Then a sigma-algebra F is a nonempty collection of subsets of X such that the following hold: 1. X is in F. 2. If A is in F, then so is the ...
  4. [4]
    Measurable Set -- from Wolfram MathWorld
    If F is a sigma-algebra and A is a subset of X, then A is called measurable if A is a member of F. X need not have, a priori, a topological structure.
  5. [5]
    Algebra of sets - Encyclopedia of Mathematics
    Aug 16, 2013 · A collection A of subsets of some set X which contains the empty set and is closed under the set-theoretic operations of finite union, finite intersection and ...Algebra of sets · σ-Algebra · Relations to measure theory
  6. [6]
    Carathéodory's extension theorem - PlanetMath.org
    Mar 22, 2013 · In measure theory, Carathéodory's extension theorem is an important result used in the construction of measures , such as the Lebesgue ...
  7. [7]
    Borel Sigma-Algebra -- from Wolfram MathWorld
    A sigma-algebra which is related to the topology of a set. The Borel sigma -algebra is defined to be the sigma-algebra generated by the open sets.
  8. [8]
    [PDF] An Introduction to Measure Theory - Terry Tao
    Theorem 1.7.3 (Carathéodory extension theorem). Let µ∗ : 2X →. [0, +∞] be an outer measure on a set X, let B be the collection of all subsets of X that ...
  9. [9]
    [PDF] Chapter 2: Lebesgue Measure - UC Davis Math
    we define Lebesgue measure on E by restricting Lebesgue measure on Rn to E, as in Definition 1.10, and denote the corresponding σ-algebra of Lebesgue measurable.
  10. [10]
    [PDF] 1.4 Properties of Lebesgue Measure - Christopher Heil
    Now we will derive some of the important properties of Lebesgue measure. We begin by improving on what we know about monotonicity, which tells us.
  11. [11]
    245A, Notes 1: Lebesgue measure | What's new - Terence Tao
    Sep 9, 2010 · Lebesgue outer measure (also known as Lebesgue exterior measure) is a special case of a more general concept known as an outer measure.
  12. [12]
    [PDF] A Gentle Introduction to the Axiom of Choice - arXiv
    1 Introduction. Described by David Hilbert as the axiom “most attacked up to the present in the mathemati- cal literature” [Hil26], the axiom of choice (AC).
  13. [13]
    A Model of Set-Theory in Which Every Set of Reals is Lebesgue ...
    This paper presents a model of set theory where every set of reals is Lebesgue measurable, by Robert M. Solovay.
  14. [14]
    [PDF] 24.118_S19 Paradox and Infinity, Lecture Note 14: The Vitali Sets
    Giving up on Uniformity means changing the subject: the whole point of our enterprise is to find a way of extending the notion of Lebesgue. Measure without ...
  15. [15]
    Sur la décomposition des ensembles de points en parties ... - EUDML
    Sur la décomposition des ensembles de points en parties respectivement congruentes. Stefan Banach; Alfred Tarski · Fundamenta Mathematicae (1924).
  16. [16]
    Grundzüge der Mengenlehre : Hausdorff, Felix, 1868-1942
    Dec 2, 2008 · Grundzüge der Mengenlehre. by: Hausdorff, Felix, 1868-1942 ... PDF download · download 1 file · SCRIBE SCANDATA ZIP download · download 1 ...
  17. [17]
    Dimension und äußeres Maß | Mathematische Annalen
    Dimension und äußeres Maß ... Article PDF. Download to read the full article text. Use our pre-submission checklist.
  18. [18]
    Lebesgue Measurable But Not Borel - Math3ma
    Aug 9, 2015 · Still, in this case, there are Lebesgue-measurable sets which are not Borel-measurable, simply by cardinality reasoning: the Borel sigma-algebra ...
  19. [19]
    Show that Lebesgue σ-algebra has the same cardinality as P(R ...
    Nov 30, 2017 · Finally, the interval [0,1] is the same cardinality as R; it has the same cardinality as (0,1), and the latter has explicit bijections with R ...The cardinality of Lebesgue sets - Math Stack ExchangeDifference between Lebesgue Sigma Algebra and Borel Sigma ...More results from math.stackexchange.com
  20. [20]
    EXTENDING LEBESGUE MEASURE BY INFINITELY MANY SETS
    Measure extension axioms and small large cardinals. The proofs of this section use standard techniques or are modifications of proofs of well-known theorems so ...
  21. [21]
    Measure and set theory. - Mathematics Stack Exchange
    Apr 24, 2016 · ... and this was exactly why Ulam proved the theorem: to show there was no (countably additive) extension of Lebesgue measure to all sets.CH, countable additivity, and Ulam's Theorem - Math Stack ExchangeThe Banach-Tarski paradox and the notion of measureMore results from math.stackexchange.com
  22. [22]
    [PDF] The Measure Problem - Purdue Math
    Ulam [5] obtains the same conclusion for the classical continuum when the measure has no nonzero value other than one. The same conclusion is now obtained for ...Missing: line | Show results with:line
  23. [23]
    [PDF] A Model of Set-Theory in Which Every Set of Reals is Lebesgue ...
    Sep 10, 2003 · We show that the existence of a non-Lebesgue measurable set cannot be proved in Zermelo-Frankel set theory (ZF) if use of the axiom of choice is.
  24. [24]
    [PDF] Historical Remarks on Suslin's Problem - Boston University
    Suslin himself was a wunderkind who after finding a mistake in a paper of Lebesgue formulated the analytic sets and estab- lished [1917] fundamental results ...
  25. [25]
    [PDF] Probability and Measure - University of Colorado Boulder
    Measure and integral are used together in Chapters 4 and 5 for the study of random sums, the Poisson process, convergence of measures, characteristic functions, ...