Appendix on the Axiom of Choice

Total Page:16

File Type:pdf, Size:1020Kb

Appendix on the Axiom of Choice Appendix A Appendix on the Axiom of Choice We repeat the following definition from Chapter 1. Definition A.1. A partial ordering on a nonempty set E is a relation ≤ on E such that 1) ∀ a ∈ E, a ≤ a,(reflexive property) and 2) ∀ a, b, c ∈ E, a ≤ b and b ≤ c ⇒ a ≤ c.(transitive property) We sometimes write b ≥ a for a ≤ b. We call a partial ordering antisymmetric if 3) ∀ a, b ∈ E, a ≤ b and b ≤ a ⇒ a = b. For an antisymmetric ordering we write a < b for a ≤ b but a = b. A partial ordering ≤ is called a total ordering if 4) ∀ a, b ∈ E, either a ≤ b or b ≤ a. A total ordering ≤ is called a linear ordering if it is antisymmetric. Recall that for a set E with a partial ordering ≤,amaximal element of E is an element z such that for each y ∈ E, y need not be related to z,butify ≥ z, then we also have y ≤ z. For an antisymmetric ordering, this means that y = z. { α ∈ I } As noted in Chapter 1, the Axiom of Choice states that if Sα : is a nonempty collection of nonempty sets, then there is a function T : I → α∈I Sα such that for every α ∈ I , T(α) ∈ Sα . If the sets are disjoint, we may think of a parliament. Recall the example given by Bertrand Russell in terms of pairs of shoes and pairs of socks: Given a finite or even countably infinite set of pairs of shoes, one can always pick one shoe from each pair, e.g., the left shoe. Given a finite collection of pairs of socks, one can pick one sock from each pair, but what happens with an infinite collection of pairs of socks? The Axiom of Choice says there is a set consisting of exactly one sock from each pair. We start with a result which, as we shall show, is equivalent to the Axiom of Choice. It is a modification of results and proofs in [17]. We will call a nonempty family C of subsets of a set S a chain if for any A and B ∈ C either A ⊆ B or B ⊆ A. We will let ∪C denote the union of the sets in a chain C . © Springer International Publishing Switzerland 2016 221 P.A. Loeb, Real Analysis, DOI 10.1007/978-3-319-30744-2 222 A Appendix on the Axiom of Choice Theorem A.1 (Hausdorff Maximal Principle for Linear Orderings). Let ≤ be a partial ordering on a nonempty set E. Let F0 be a nonempty subset of E that is linearly ordered with respect to ≤. There is a maximal (with respect to ⊆ ) subset F ⊆ E such that F0 ⊆ F and ≤ is a linear ordering on F. Proof. We let F be the family of all F ⊆ E such that the restriction of the ordering ≤ to F is a linear ordering and F ⊇ F0.LetT (∅)=F0, and for each F ∈ F ,let T(F) be an element of F that strictly contains F if there is one, and otherwise let T(F)=F. Here we have used the Axiom of Choice to define T. We want to show that there is an F ∈ F with T(F)=F. By the definition of T,thisF is a maximal linear ordered subset of E containing F0. Now, the union of the sets in a chain C consisting of members of F is again in F . To see this, let x and y be members of this union. We have x ∈ A ∈ C , and y ∈ B ∈ C , and either A ⊆ B or B ⊆ A. Suppose the first. Then both x and y are in B, so since ≤ is a linear ordering on B, either x ≤ y or y ≤ x; if both then x = y.The same is true if B ⊆ A. Let W be the collection of all chains C consisting of members of F such that for each F ∈ C with F F0, ∪{A ∈ C : A F}∈C , and for each F ∈ C such that F = ∪{A ∈ C : A F}, F = T (∪{A ∈ C : A F}). This means that T uses the Axiom of Choice to supply the next step in the containment after the union. Note that the singleton {F0} is in W . Given C ∈ W , we say that a chain D is an initial chain of C if D ∈ W , each F ∈ D is a member of the chain C , and if A ∈ C and A ⊂ F, then A ∈ D. The union G of a nonempty family of initial chains of C is an initial chain of C . To see this, note that F0 is in G , and each member of G is a member of C . Moreover, if A and B are members of G , then as elements of the chain C , either A ⊆ B or B ⊆ A.Also, each A ∈ G is an element of an initial chain D ∈ W ,soifB ∈ C and B ⊂ A, then B ∈ D.ItfollowsthatG ∈ W . If C and D are chains in W , then either C is an initial chain of D or D is an initial chain of C . To see this, let G be the union of all chains in W that are both initial chains of C and initial chains of D. The singleton {F0} is such a chain. As before, G ∈ W , and G is the largest initial chain of both C and D. Suppose C = G . Let F = ∪G .IfF ∈/ G ,letG = G ∪{F}.IfF ∈ G ,letG = G ∪{T (F)}. Then G ∈ W , and G is strictly larger than G initial chain of C . Similarly, if D = G , then G is strictly larger than G initial chain of D. Since we cannot have both, either G = C or G = D, or both. Now let G be the union of all chains in W . Then G ∈ W . To see this, note that the singleton {F0} is in G .IfA and B are in G , then A ∈ C ∈ W , and B ∈ D ∈ W . If C is an initial chain of D, then both A and B are in D, so either A ⊆ B or B ⊆ A. Similarly, if D is an initial chain of C , then both A and B are in C , so either A ⊆ B or B ⊆ A. Therefore, G is a chain. If F ∈ G , then F ∈ C for some C ∈ W . Therefore, if F = F0, ∪{A ∈ C : A F}∈C ⊆ G .IfF = ∪{A ∈ C : A F}, then F = T (∪{A ∈ C : A F}). Since G ∈ W and G is the union of all chains in W , ∪G ∈ G . Since G ∪{T (∪G )} must equal G , ∪G is a maximal linearly ordered subset of E containing F0. A Appendix on the Axiom of Choice 223 Remark A.1. A similar proof, left to the reader, establishes the Hausdorff Maximal Principle for total orderings. Theorem A.2. The following principles are equivalent: 1) Axiom of Choice. 2) Hausdorff Maximal Principle for Linear Orderings. 3) Zorn’s Lemma: Let ≤ be a partial ordering on a nonempty set E.Ifevery linearly ordered subset of E has an upper bound in E, then there is a maximal element with respect to ≤ contained in E. 4) Well-Ordering Principle: Every set can be ordered with a well-ordering. Proof. 1) ⇒ 2): This is Theorem A.1. 2) ⇒ 3): By the assumption, we can find a maximum (under containment) lin- early ordered subset F of E.Letz be an upper bound of F. We will show that z is a maximal element of E. Consider any y such that z ≤ y. We must show that we also have y ≤ z; they do not have to be equal. Now since z is an upper bound of F,for each x ∈ F, x ≤ z ≤ y.SoF ∪{y} is totally ordered with respect to ≤. Since F is a maximal linearly ordered set in E, it must also be true that for some x ∈ F, y ≤ x, whence, y ≤ z. 3) ⇒ 4): Given a nonempty set E, consider subsets F of E with a well-orderings ≤F .LetF be the set of all such pairs (F,≤F ). (A given set may appear more than once but with different orderings. In this sense, the subscript on the ordering is a bit misleading.) We can put an antisymmetric, partial ordering on F by setting (F,≤F ) (G,≤G) when i) F ⊆ G, ii) ≤G=≤F on F × F, [i.e., if x, z ∈ F, then x ≤F z, if and only if x ≤G z], and iii) for every pair (x,y) with x ∈ F and y ∈ G \ F, x ≤G y. We wish to show that is, in fact, an antisymmetric, partial ordering on F . Clearly, for each element (F,≤F ) in F , (F,≤F ) (F,≤F ) [(iii) is vacuously satis- fied]. Moreover, (F,≤F ) (G,≤G) and (G,≤G) (H,≤H ) ⇒ (F,≤F ) (H,≤H ). To see this, we note that Properties i and ii are clear, so we only check (iii). If y ∈ H \ F and x ∈ F, then for the case y ∈ H \ G,wehavex ≤H y since x ∈ G.If y ∈ G\F, then we have x ≤G y,sox ≤H y by (ii). To see that is antisymmetric, we assume that we have both (F,≤F ) (G,≤G) and (G,≤G) (F,≤F ).By(i),F = G. By (ii), ≤G= ≤F . Fix a subset G = {(Fα ,≤F )} of F , linearly ordered with respect to .Weset G = ∪α Fα , and we define an ordering on G by setting x ≤G y if x ≤Fα y for some α.
Recommended publications
  • A Proof of Cantor's Theorem
    Cantor’s Theorem Joe Roussos 1 Preliminary ideas Two sets have the same number of elements (are equinumerous, or have the same cardinality) iff there is a bijection between the two sets. Mappings: A mapping, or function, is a rule that associates elements of one set with elements of another set. We write this f : X ! Y , f is called the function/mapping, the set X is called the domain, and Y is called the codomain. We specify what the rule is by writing f(x) = y or f : x 7! y. e.g. X = f1; 2; 3g;Y = f2; 4; 6g, the map f(x) = 2x associates each element x 2 X with the element in Y that is double it. A bijection is a mapping that is injective and surjective.1 • Injective (one-to-one): A function is injective if it takes each element of the do- main onto at most one element of the codomain. It never maps more than one element in the domain onto the same element in the codomain. Formally, if f is a function between set X and set Y , then f is injective iff 8a; b 2 X; f(a) = f(b) ! a = b • Surjective (onto): A function is surjective if it maps something onto every element of the codomain. It can map more than one thing onto the same element in the codomain, but it needs to hit everything in the codomain. Formally, if f is a function between set X and set Y , then f is surjective iff 8y 2 Y; 9x 2 X; f(x) = y Figure 1: Injective map.
    [Show full text]
  • A Convenient Category for Higher-Order Probability Theory
    A Convenient Category for Higher-Order Probability Theory Chris Heunen Ohad Kammar Sam Staton Hongseok Yang University of Edinburgh, UK University of Oxford, UK University of Oxford, UK University of Oxford, UK Abstract—Higher-order probabilistic programming languages 1 (defquery Bayesian-linear-regression allow programmers to write sophisticated models in machine 2 let let sample normal learning and statistics in a succinct and structured way, but step ( [f ( [s ( ( 0.0 3.0)) 3 sample normal outside the standard measure-theoretic formalization of proba- b ( ( 0.0 3.0))] 4 fn + * bility theory. Programs may use both higher-order functions and ( [x] ( ( s x) b)))] continuous distributions, or even define a probability distribution 5 (observe (normal (f 1.0) 0.5) 2.5) on functions. But standard probability theory does not handle 6 (observe (normal (f 2.0) 0.5) 3.8) higher-order functions well: the category of measurable spaces 7 (observe (normal (f 3.0) 0.5) 4.5) is not cartesian closed. 8 (observe (normal (f 4.0) 0.5) 6.2) Here we introduce quasi-Borel spaces. We show that these 9 (observe (normal (f 5.0) 0.5) 8.0) spaces: form a new formalization of probability theory replacing 10 (predict :f f))) measurable spaces; form a cartesian closed category and so support higher-order functions; form a well-pointed category and so support good proof principles for equational reasoning; and support continuous probability distributions. We demonstrate the use of quasi-Borel spaces for higher-order functions and proba- bility by: showing that a well-known construction of probability theory involving random functions gains a cleaner expression; and generalizing de Finetti’s theorem, that is a crucial theorem in probability theory, to quasi-Borel spaces.
    [Show full text]
  • Jointly Measurable and Progressively Measurable Stochastic Processes
    Jointly measurable and progressively measurable stochastic processes Jordan Bell [email protected] Department of Mathematics, University of Toronto June 18, 2015 1 Jointly measurable stochastic processes d Let E = R with Borel E , let I = R≥0, which is a topological space with the subspace topology inherited from R, and let (Ω; F ;P ) be a probability space. For a stochastic process (Xt)t2I with state space E, we say that X is jointly measurable if the map (t; !) 7! Xt(!) is measurable BI ⊗ F ! E . For ! 2 Ω, the path t 7! Xt(!) is called left-continuous if for each t 2 I, Xs(!) ! Xt(!); s " t: We prove that if the paths of a stochastic process are left-continuous then the stochastic process is jointly measurable.1 Theorem 1. If X is a stochastic process with state space E and all the paths of X are left-continuous, then X is jointly measurable. Proof. For n ≥ 1, t 2 I, and ! 2 Ω, let 1 n X Xt (!) = 1[k2−n;(k+1)2−n)(t)Xk2−n (!): k=0 n Each X is measurable BI ⊗ F ! E : for B 2 E , 1 n [ −n −n f(t; !) 2 I × Ω: Xt (!) 2 Bg = [k2 ; (k + 1)2 ) × fXk2−n 2 Bg: k=0 −n −n Let t 2 I. For each n there is a unique kn for which t 2 [kn2 ; (kn +1)2 ), and n −n −n thus Xt (!) = Xkn2 (!). Furthermore, kn2 " t, and because s 7! Xs(!) is n −n left-continuous, Xkn2 (!) ! Xt(!).
    [Show full text]
  • An Update on the Four-Color Theorem Robin Thomas
    thomas.qxp 6/11/98 4:10 PM Page 848 An Update on the Four-Color Theorem Robin Thomas very planar map of connected countries the five-color theorem (Theorem 2 below) and can be colored using four colors in such discovered what became known as Kempe chains, a way that countries with a common and Tait found an equivalent formulation of the boundary segment (not just a point) re- Four-Color Theorem in terms of edge 3-coloring, ceive different colors. It is amazing that stated here as Theorem 3. Esuch a simply stated result resisted proof for one The next major contribution came in 1913 from and a quarter centuries, and even today it is not G. D. Birkhoff, whose work allowed Franklin to yet fully understood. In this article I concentrate prove in 1922 that the four-color conjecture is on recent developments: equivalent formulations, true for maps with at most twenty-five regions. The a new proof, and progress on some generalizations. same method was used by other mathematicians to make progress on the four-color problem. Im- Brief History portant here is the work by Heesch, who developed The Four-Color Problem dates back to 1852 when the two main ingredients needed for the ultimate Francis Guthrie, while trying to color the map of proof—“reducibility” and “discharging”. While the the counties of England, noticed that four colors concept of reducibility was studied by other re- sufficed. He asked his brother Frederick if it was searchers as well, the idea of discharging, crucial true that any map can be colored using four col- for the unavoidability part of the proof, is due to ors in such a way that adjacent regions (i.e., those Heesch, and he also conjectured that a suitable de- sharing a common boundary segment, not just a velopment of this method would solve the Four- point) receive different colors.
    [Show full text]
  • Shape Analysis, Lebesgue Integration and Absolute Continuity Connections
    NISTIR 8217 Shape Analysis, Lebesgue Integration and Absolute Continuity Connections Javier Bernal This publication is available free of charge from: https://doi.org/10.6028/NIST.IR.8217 NISTIR 8217 Shape Analysis, Lebesgue Integration and Absolute Continuity Connections Javier Bernal Applied and Computational Mathematics Division Information Technology Laboratory This publication is available free of charge from: https://doi.org/10.6028/NIST.IR.8217 July 2018 INCLUDES UPDATES AS OF 07-18-2018; SEE APPENDIX U.S. Department of Commerce Wilbur L. Ross, Jr., Secretary National Institute of Standards and Technology Walter Copan, NIST Director and Undersecretary of Commerce for Standards and Technology ______________________________________________________________________________________________________ This Shape Analysis, Lebesgue Integration and publication Absolute Continuity Connections Javier Bernal is National Institute of Standards and Technology, available Gaithersburg, MD 20899, USA free of Abstract charge As shape analysis of the form presented in Srivastava and Klassen’s textbook “Functional and Shape Data Analysis” is intricately related to Lebesgue integration and absolute continuity, it is advantageous from: to have a good grasp of the latter two notions. Accordingly, in these notes we review basic concepts and results about Lebesgue integration https://doi.org/10.6028/NIST.IR.8217 and absolute continuity. In particular, we review fundamental results connecting them to each other and to the kind of shape analysis, or more generally, functional data analysis presented in the aforeme- tioned textbook, in the process shedding light on important aspects of all three notions. Many well-known results, especially most results about Lebesgue integration and some results about absolute conti- nuity, are presented without proofs.
    [Show full text]
  • Fibonacci, Kronecker and Hilbert NKS 2007
    Fibonacci, Kronecker and Hilbert NKS 2007 Klaus Sutner Carnegie Mellon University www.cs.cmu.edu/∼sutner NKS’07 1 Overview • Fibonacci, Kronecker and Hilbert ??? • Logic and Decidability • Additive Cellular Automata • A Knuth Question • Some Questions NKS’07 2 Hilbert NKS’07 3 Entscheidungsproblem The Entscheidungsproblem is solved when one knows a procedure by which one can decide in a finite number of operations whether a given logical expression is generally valid or is satisfiable. The solution of the Entscheidungsproblem is of fundamental importance for the theory of all fields, the theorems of which are at all capable of logical development from finitely many axioms. D. Hilbert, W. Ackermann Grundzuge¨ der theoretischen Logik, 1928 NKS’07 4 Model Checking The Entscheidungsproblem for the 21. Century. Shift to computer science, even commercial applications. Fix some suitable logic L and collection of structures A. Find efficient algorithms to determine A |= ϕ for any structure A ∈ A and sentence ϕ in L. Variants: fix ϕ, fix A. NKS’07 5 CA as Structures Discrete dynamical systems, minimalist description: Aρ = hC, i where C ⊆ ΣZ is the space of configurations of the system and is the “next configuration” relation induced by the local map ρ. Use standard first order logic (either relational or functional) to describe properties of the system. NKS’07 6 Some Formulae ∀ x ∃ y (y x) ∀ x, y, z (x z ∧ y z ⇒ x = y) ∀ x ∃ y, z (y x ∧ z x ∧ ∀ u (u x ⇒ u = y ∨ u = z)) There is no computability requirement for configurations, in x y both x and y may be complicated.
    [Show full text]
  • The Open Handbook of Formal Epistemology
    THEOPENHANDBOOKOFFORMALEPISTEMOLOGY Richard Pettigrew &Jonathan Weisberg,Eds. THEOPENHANDBOOKOFFORMAL EPISTEMOLOGY Richard Pettigrew &Jonathan Weisberg,Eds. Published open access by PhilPapers, 2019 All entries copyright © their respective authors and licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License. LISTOFCONTRIBUTORS R. A. Briggs Stanford University Michael Caie University of Toronto Kenny Easwaran Texas A&M University Konstantin Genin University of Toronto Franz Huber University of Toronto Jason Konek University of Bristol Hanti Lin University of California, Davis Anna Mahtani London School of Economics Johanna Thoma London School of Economics Michael G. Titelbaum University of Wisconsin, Madison Sylvia Wenmackers Katholieke Universiteit Leuven iii For our teachers Overall, and ultimately, mathematical methods are necessary for philosophical progress. — Hannes Leitgeb There is no mathematical substitute for philosophy. — Saul Kripke PREFACE In formal epistemology, we use mathematical methods to explore the questions of epistemology and rational choice. What can we know? What should we believe and how strongly? How should we act based on our beliefs and values? We begin by modelling phenomena like knowledge, belief, and desire using mathematical machinery, just as a biologist might model the fluc- tuations of a pair of competing populations, or a physicist might model the turbulence of a fluid passing through a small aperture. Then, we ex- plore, discover, and justify the laws governing those phenomena, using the precision that mathematical machinery affords. For example, we might represent a person by the strengths of their beliefs, and we might measure these using real numbers, which we call credences. Having done this, we might ask what the norms are that govern that person when we represent them in that way.
    [Show full text]
  • LEBESGUE MEASURE and L2 SPACE. Contents 1. Measure Spaces 1 2. Lebesgue Integration 2 3. L2 Space 4 Acknowledgments 9 References
    LEBESGUE MEASURE AND L2 SPACE. ANNIE WANG Abstract. This paper begins with an introduction to measure spaces and the Lebesgue theory of measure and integration. Several important theorems regarding the Lebesgue integral are then developed. Finally, we prove the completeness of the L2(µ) space and show that it is a metric space, and a Hilbert space. Contents 1. Measure Spaces 1 2. Lebesgue Integration 2 3. L2 Space 4 Acknowledgments 9 References 9 1. Measure Spaces Definition 1.1. Suppose X is a set. Then X is said to be a measure space if there exists a σ-ring M (that is, M is a nonempty family of subsets of X closed under countable unions and under complements)of subsets of X and a non-negative countably additive set function µ (called a measure) defined on M . If X 2 M, then X is said to be a measurable space. For example, let X = Rp, M the collection of Lebesgue-measurable subsets of Rp, and µ the Lebesgue measure. Another measure space can be found by taking X to be the set of all positive integers, M the collection of all subsets of X, and µ(E) the number of elements of E. We will be interested only in a special case of the measure, the Lebesgue measure. The Lebesgue measure allows us to extend the notions of length and volume to more complicated sets. Definition 1.2. Let Rp be a p-dimensional Euclidean space . We denote an interval p of R by the set of points x = (x1; :::; xp) such that (1.3) ai ≤ xi ≤ bi (i = 1; : : : ; p) Definition 1.4.
    [Show full text]
  • 1 Measurable Functions
    36-752 Advanced Probability Overview Spring 2018 2. Measurable Functions, Random Variables, and Integration Instructor: Alessandro Rinaldo Associated reading: Sec 1.5 of Ash and Dol´eans-Dade; Sec 1.3 and 1.4 of Durrett. 1 Measurable Functions 1.1 Measurable functions Measurable functions are functions that we can integrate with respect to measures in much the same way that continuous functions can be integrated \dx". Recall that the Riemann integral of a continuous function f over a bounded interval is defined as a limit of sums of lengths of subintervals times values of f on the subintervals. The measure of a set generalizes the length while elements of the σ-field generalize the intervals. Recall that a real-valued function is continuous if and only if the inverse image of every open set is open. This generalizes to the inverse image of every measurable set being measurable. Definition 1 (Measurable Functions). Let (Ω; F) and (S; A) be measurable spaces. Let f :Ω ! S be a function that satisfies f −1(A) 2 F for each A 2 A. Then we say that f is F=A-measurable. If the σ-field’s are to be understood from context, we simply say that f is measurable. Example 2. Let F = 2Ω. Then every function from Ω to a set S is measurable no matter what A is. Example 3. Let A = f?;Sg. Then every function from a set Ω to S is measurable, no matter what F is. Proving that a function is measurable is facilitated by noticing that inverse image commutes with union, complement, and intersection.
    [Show full text]
  • Notes 2 : Measure-Theoretic Foundations II
    Notes 2 : Measure-theoretic foundations II Math 733-734: Theory of Probability Lecturer: Sebastien Roch References: [Wil91, Chapters 4-6, 8], [Dur10, Sections 1.4-1.7, 2.1]. 1 Independence 1.1 Definition of independence Let (Ω; F; P) be a probability space. DEF 2.1 (Independence) Sub-σ-algebras G1; G2;::: of F are independent for all Gi 2 Gi, i ≥ 1, and distinct i1; : : : ; in we have n Y P[Gi1 \···\ Gin ] = P[Gij ]: j=1 Specializing to events and random variables: DEF 2.2 (Independent RVs) RVs X1;X2;::: are independent if the σ-algebras σ(X1); σ(X2);::: are independent. DEF 2.3 (Independent Events) Events E1;E2;::: are independent if the σ-algebras c Ei = f;;Ei;Ei ; Ωg; i ≥ 1; are independent. The more familiar definitions are the following: THM 2.4 (Independent RVs: Familiar definition) RVs X, Y are independent if and only if for all x; y 2 R P[X ≤ x; Y ≤ y] = P[X ≤ x]P[Y ≤ y]: THM 2.5 (Independent events: Familiar definition) Events E1, E2 are indepen- dent if and only if P[E1 \ E2] = P[E1]P[E2]: 1 Lecture 2: Measure-theoretic foundations II 2 The proofs of these characterizations follows immediately from the following lemma. LEM 2.6 (Independence and π-systems) Suppose that G and H are sub-σ-algebras and that I and J are π-systems such that σ(I) = G; σ(J ) = H: Then G and H are independent if and only if I and J are, i.e., P[I \ J] = P[I]P[J]; 8I 2 I;J 2 J : Proof: Suppose I and J are independent.
    [Show full text]
  • The Axiom of Choice and Its Implications
    THE AXIOM OF CHOICE AND ITS IMPLICATIONS KEVIN BARNUM Abstract. In this paper we will look at the Axiom of Choice and some of the various implications it has. These implications include a number of equivalent statements, and also some less accepted ideas. The proofs discussed will give us an idea of why the Axiom of Choice is so powerful, but also so controversial. Contents 1. Introduction 1 2. The Axiom of Choice and Its Equivalents 1 2.1. The Axiom of Choice and its Well-known Equivalents 1 2.2. Some Other Less Well-known Equivalents of the Axiom of Choice 3 3. Applications of the Axiom of Choice 5 3.1. Equivalence Between The Axiom of Choice and the Claim that Every Vector Space has a Basis 5 3.2. Some More Applications of the Axiom of Choice 6 4. Controversial Results 10 Acknowledgments 11 References 11 1. Introduction The Axiom of Choice states that for any family of nonempty disjoint sets, there exists a set that consists of exactly one element from each element of the family. It seems strange at first that such an innocuous sounding idea can be so powerful and controversial, but it certainly is both. To understand why, we will start by looking at some statements that are equivalent to the axiom of choice. Many of these equivalences are very useful, and we devote much time to one, namely, that every vector space has a basis. We go on from there to see a few more applications of the Axiom of Choice and its equivalents, and finish by looking at some of the reasons why the Axiom of Choice is so controversial.
    [Show full text]
  • “Strange” Limits
    c Gabriel Nagy “Strange” Limits Notes from the Functional Analysis Course (Fall 07 - Spring 08) Prerequisites: The reader is assumed to be familiar with the Hahn-Banach Theorem and/or the theory of (ultra)filter convergence. References to some of my notes will be added later. Notations. Given a non-empty set S, we denote by `∞(S) the vector space of all bounded R functions x : S → . On occasion an element x ∈ `∞(S) will also be written as an “S-tuple” R R x = (xs)s∈S. If S is infinite, we denote by Pfin(S) the collection of all finite subsets of S. Assuming S is infinite, for an element x = (x ) ∈ `∞(S), we can define the quantities s s∈S R lim sup xs = inf sup xs , F ∈P (S) S fin s∈SrF lim inf xs = sup inf xs . S s∈S F F ∈Pfin(S) r Definition. Suppose S is infinite. A limit operation on S is a linear map Λ : `∞(S) → , R R satisfying the inequalities: lim inf x ≤ Λ(x) ≤ lim sup x , ∀ x = (x ) ∈ `∞(S). (1) s s s s∈S R S S Comment. Of course, the quantities lim supS xs and lim infS xs can be defined for ar- bitrary “S-tuples,” in which case these limits might equal ±∞. In fact, this allows one to use S-tuples which might take infinite values. In other words, one might consider maps x : S → [−∞, ∞]. We will elaborate on his point of view later in this note. With the above terminology, our problem is to construct limit operations.
    [Show full text]