Chapter 8 Ordered Sets

Total Page:16

File Type:pdf, Size:1020Kb

Chapter 8 Ordered Sets Chapter VIII Ordered Sets, Ordinals and Transfinite Methods 1. Introduction In this chapter, we will look at certain kinds of ordered sets. If a set \ is ordered in a reasonable way, then there is a natural way to define an “order topology” on \. Most interesting (for our purposes) will be ordered sets that satisfy a very strong ordering condition: that every nonempty subset contains a smallest element. Such sets are called well-ordered. The most familiar example of a well-ordered set is and it is the well-ordering property that lets us do mathematical induction in In this chapter we will see “longer” well ordered sets and these will give us a new proof method called “transfinite induction.” But we begin with something simpler. 2. Partially Ordered Sets Recall that a relation V\ on a set is a subset of \‚\ (see Definition I.5.2 ). If ÐBßCÑ−V, we write BVCÞ An “order” on a set \ is refers to a relation on \ that satisfies some additional conditions. Order relations are usually denoted by symbols such asŸ¡ß£ , , or . Definition 2.1 A relation V\ on is called: transitive ifÀ a +ß ,ß - − \ Ð+V, and ,V-Ñ Ê +V-Þ reflexive ifÀa+−\+V+ antisymmetric ifÀ a +ß , − \ Ð+V, and ,V+ Ñ Ê Ð+ œ ,Ñ symmetric ifÀ a +ß , − \ +V, Í ,V+ (that is, the set V is “symmetric” with respect to thediagonal ? œÖÐBßBÑÀB−\ש\‚\). Example 2.2 1) The relation “œ\ ” on a set is transitive, reflexive, symmetric, and antisymmetric. Viewed as a subset of \‚\, the relation “ œ ” is the diagonal set ? œÖÐBßBÑÀB−\×Þ 2) In ‘, the usual order relation is transitive and antisymmetric, but not reflexive or symmetric. 3) In ‘, the usual order Ÿ is transitive, reflexive and antisymmetric. It is not symmetric. 4) On any set of cardinal numbers V we have a relation Ÿ . It is transitive, reflexive and antisymmetric (by the Cantor-Schroeder-Bernstein Theorem I.10.2 ), but not symmetric Ðllœ"ÑÞunless V 319 Definition 2.3 A relation Ÿ\ on a set is called a partial order if Ÿ is transitive, reflexive and antisymmetric. The pair Ð\ß Ÿ Ñ is called a partially ordered set (or, for short, poset). A relation Ÿ\ on a set is called a linear order if Ÿ is a partial order and, in addition, any two elements in \ are comparable: a+ß,−\ either +Ÿ, or ,Ÿ+. In this case, the pair Ð\ߟ) is called a linearly ordered set. For short, a linearly ordered set is also called a chain. We write +, to mean that +Ÿ, and +Á,ÞFor any sort of order relation Ÿon \, we can invert the order notation and write , + Ð,+Ñto mean the same thing as +Ÿ, Ð+,ÑÞ In some books, a partial order is defined as a “strict” relation which is transitive and irreflexiveÐa+ − \ , a Î +ÑÞ In that case, we can define +Ÿ, to mean “ +, or +œ,” to get a partial order in the sense defined above. This variation in terminology creates no real mathematical problems: the difference is completely analogous to worrying about whether “Ÿ ” or “ ” should be called the“usual order” on ‘ . Example 2.4 1) Suppose E©\ÞFor any kind of order Ÿon \ß we can get an order ŸE on E©\ by restricting the order ŸEÞ to More formally, ŸœŸ∩E‚EE ( ). We always assume that a subset of an ordered set has this natural “inherited” ordering unless something else is explicitly stated. With that understanding, we usually omit the subscript and also write ŸŸE for the order E on . If Ð\ß Ÿ Ñ is a poset (or chain), and E © \, then ÐEß Ÿ Ñ is also a poset (or chain). For example, every subset of Ðߟё is a chain. 2) For Dß A −‚ Ð œthe set of complex numbers Ñ , define D £ A iff lDl Ÿ lAlß where Ÿ is the usual order in ‘‚. Ðᜥ is not a poset. (Why? ) 3) Let Ð\ßg Ñ be a topological space. For 0ß 1 − GÐ\Ñ, define 0Ÿ1 iff aB−\, 0ÐBÑŸ1ÐBÑ. As a set, Ÿ œ ÖÐ0ß 1Ñ − GÐ\Ñ ‚ GÐ\Ñ À aB −‘ ß 0ÐBÑ Ÿ 1ÐBÑ×. Notice that, in contrast to part 2), we are allowing ourselves an ambiguity in the notation here because we are using “ŸŸGÐ\Ñ ” with two different meanings: we are defining an order “ ” in , but the comparison “0ÐBÑ Ÿ 1ÐBÑ ” refers to the usual order a different ordered set, ‘. Of course, we could be more careful and write 0£1for the new order on GÐ\Ñ, but usually we won't be that fussy when the context makes clear which meaning of “Ÿ ” we have in mind. ÐGÐ\Ñß Ÿ Ñ is a poset but usually not a chain: for example, if \ œ‘ and 0ß 1 are given by 0ÐBÑ œ B and 1ÐBÑ œ B#, then 0 ŸÎ1 and 1ŸÎ 0. When is ÐGÐ\Ñß Ÿ Ñ a chain? ÐThe answer is not “ iff l\l Ÿ "Þ” Ñ 320 4) The following two diagrams represent posets, each with 5 elements. Line segments upward from B to C indicate that BŸC. In Figure (i), for example, .Ÿ-and -Ÿ+(so .Ÿ+); in Figure (i), +, and are not comparable. Figure (ii) shows a chain: +Ÿ,Ÿ-Ÿ.Ÿ/Þ 5) Suppose VV is a collection of sets. We can define ŸEŸFE©F on by iff . Then ÐߟÑVV is a poset. In this case, we say that has been ordered by inclusion. In particular, for any set \, we can order cc Ð\Ñby inclusion. What conditions on \ will guarantee Ð Ð\Ñß Ÿ ) is a chain ? 6) Suppose VV is a collection of sets. We can define ŸEŸFEªF on by iff . Then ÐߟÑVV is a poset. In this case, we say that has been ordered by reverse inclusion. In particular, for any set \Ð\Ñ, we can order c by reverse inclusion. What conditions on \ will guarantee that ÐÐ\Ñߟc ) is a chain ? For a given collection V, Examples 5) and 6) are quite similar: one is a “mirror image” of the other. The identity map 3ÀVV Ä is an “order-reversing isomorphism” between the posets. For our purposes, the “reverse inclusion” ordering on a collection of sets will turn out to be more useful. For a point B\ in a topological space , we can order the neighborhood system aB by reverse inclusion ŸÐߟÑ\. The “order structure” of the poset aB reflects some topological properties of and indicates just how complicated the “neighborhood structure” at B is. For example, i) If B is isolated, then ÖB×−RBB and RŸÖB× for every R−R . So the poset ÐߟÑaB has a largest element. ÐIs the converse true? Ñ ii) If \ is first countable and UB"#5 œ ÖR ß R ß ÞÞÞß R ß ÞÞÞ× is a countable neighborhood base at BR−5, then for every aB there is a such that RŸR5B. Thus the poset Ða ßŸÑ contains a countable subset ÖR"# ß R ß ÞÞÞß R 5 ß ÞÞÞ× whose members become “arbitrarily large” in the poset. iii) The poset ÐߟÑaB is usually not a chain. But it does have interesting 321 order property: for any RßR−"#aa B ßbR− $ B such that RŸR " $ and RŸR # $ (let RœR∩R$"#ÑÞ It will turn out (in Chapter 9) that this poset ÐߟÑaB is the inspiration for defining the notion of “convergent nets,” a kind of convergence that is more powerful than “convergent sequences.” Unlike sequences, convergent nets will be able to determine closures (and therefore the topology) in any topological space. Definition 2.5 Suppose Ð\ß Ÿ Ñ is a poset and let E © \Þ An element B − \ is called i) the largest (or last) element in \CŸB if for all C−\ ii) the smallest (or first) element in \BŸC if for all C−\ iii) a maximal element in \ if ÐC B ÊCœBÑ for all C−\ iv) a minimal element in \ if ÐCŸB ÊCœBÑ for all C−\. It is clear that a largest element in Ð\ß Ÿ Ñ, if it exists, is unique. (If DD"# and were both largest, then DŸD"# and DŸD #""# so DœD.) In Figure (i): both +ß , are maximal elements and .ß / are minimal elements. This poset has no largest or smallest element. Suppose a poset Ð\ß Ÿ Ñ has a unique maximal element D. Must D also be the largest element in Ð\ß Ÿ Ñ ? In Figure (ii): +/ is the smallest (and also a minimal element); is the largest (and also a maximal) element. v) Suppose E©\ and B−\ÞWe say that B−\ is an upper bound for E if +ŸB for all +−EàB is called aleast upper bound (sup) for E if B is the smallest upper bound for EÞ The set E\Emight have many upper bounds, one upper bound, or no upper bounds in . If has more than one upper bound, E\\ might or might not have a least upper bound in . But if has a least upper boundB , then the least upper bound is unique (why? ). vi) An element B−\ is called a lowerbound for E if BŸ+ for all +−EàB is called a greatest lower bound (inf) for EB if is the largest lower bound for EÞ The set E might have many lower bounds, one lower bound, or no lower bounds in \E. If has more than one lower bounds, E\\ might or might not have a greatest lower bound in , but if has a greatest lower bound B, then the greatest lower bound is unique (why? ). vii) If BC−\ and if bD−\Î with BDC, then B is called an immediate predecessor of CC and is called an immediate successor of B.
Recommended publications
  • Chapter 4. Homomorphisms and Isomorphisms of Groups
    Chapter 4. Homomorphisms and Isomorphisms of Groups 4.1 Note: We recall the following terminology. Let X and Y be sets. When we say that f is a function or a map from X to Y , written f : X ! Y , we mean that for every x 2 X there exists a unique corresponding element y = f(x) 2 Y . The set X is called the domain of f and the range or image of f is the set Image(f) = f(X) = f(x) x 2 X . For a set A ⊆ X, the image of A under f is the set f(A) = f(a) a 2 A and for a set −1 B ⊆ Y , the inverse image of B under f is the set f (B) = x 2 X f(x) 2 B . For a function f : X ! Y , we say f is one-to-one (written 1 : 1) or injective when for every y 2 Y there exists at most one x 2 X such that y = f(x), we say f is onto or surjective when for every y 2 Y there exists at least one x 2 X such that y = f(x), and we say f is invertible or bijective when f is 1:1 and onto, that is for every y 2 Y there exists a unique x 2 X such that y = f(x). When f is invertible, the inverse of f is the function f −1 : Y ! X defined by f −1(y) = x () y = f(x). For f : X ! Y and g : Y ! Z, the composite g ◦ f : X ! Z is given by (g ◦ f)(x) = g(f(x)).
    [Show full text]
  • Basic Properties of Filter Convergence Spaces
    Basic Properties of Filter Convergence Spaces Barbel¨ M. R. Stadlery, Peter F. Stadlery;z;∗ yInstitut fur¨ Theoretische Chemie, Universit¨at Wien, W¨ahringerstraße 17, A-1090 Wien, Austria zThe Santa Fe Institute, 1399 Hyde Park Road, Santa Fe, NM 87501, USA ∗Address for corresponce Abstract. This technical report summarized facts from the basic theory of filter convergence spaces and gives detailed proofs for them. Many of the results collected here are well known for various types of spaces. We have made no attempt to find the original proofs. 1. Introduction Mathematical notions such as convergence, continuity, and separation are, at textbook level, usually associated with topological spaces. It is possible, however, to introduce them in a much more abstract way, based on axioms for convergence instead of neighborhood. This approach was explored in seminal work by Choquet [4], Hausdorff [12], Katˇetov [14], Kent [16], and others. Here we give a brief introduction to this line of reasoning. While the material is well known to specialists it does not seem to be easily accessible to non-topologists. In some cases we include proofs of elementary facts for two reasons: (i) The most basic facts are quoted without proofs in research papers, and (ii) the proofs may serve as examples to see the rather abstract formalism at work. 2. Sets and Filters Let X be a set, P(X) its power set, and H ⊆ P(X). The we define H∗ = fA ⊆ Xj(X n A) 2= Hg (1) H# = fA ⊆ Xj8Q 2 H : A \ Q =6 ;g The set systems H∗ and H# are called the conjugate and the grill of H, respectively.
    [Show full text]
  • Nieuw Archief Voor Wiskunde
    Nieuw Archief voor Wiskunde Boekbespreking Kevin Broughan Equivalents of the Riemann Hypothesis Volume 1: Arithmetic Equivalents Cambridge University Press, 2017 xx + 325 p., prijs £ 99.99 ISBN 9781107197046 Kevin Broughan Equivalents of the Riemann Hypothesis Volume 2: Analytic Equivalents Cambridge University Press, 2017 xix + 491 p., prijs £ 120.00 ISBN 9781107197121 Reviewed by Pieter Moree These two volumes give a survey of conjectures equivalent to the ber theorem says that r()x asymptotically behaves as xx/log . That Riemann Hypothesis (RH). The first volume deals largely with state- is a much weaker statement and is equivalent with there being no ments of an arithmetic nature, while the second part considers zeta zeros on the line v = 1. That there are no zeros with v > 1 is more analytic equivalents. a consequence of the prime product identity for g()s . The Riemann zeta function, is defined by It would go too far here to discuss all chapters and I will limit 3 myself to some chapters that are either close to my mathematical 1 (1) g()s = / s , expertise or those discussing some of the most famous RH equiv- n = 1 n alences. Most of the criteria have their own chapter devoted to with si=+v t a complex number having real part v > 1 . It is easily them, Chapter 10 has various criteria that are discussed more brief- seen to converge for such s. By analytic continuation the Riemann ly. A nice example is Redheffer’s criterion. It states that RH holds zeta function can be uniquely defined for all s ! 1.
    [Show full text]
  • The General Linear Group
    18.704 Gabe Cunningham 2/18/05 [email protected] The General Linear Group Definition: Let F be a field. Then the general linear group GLn(F ) is the group of invert- ible n × n matrices with entries in F under matrix multiplication. It is easy to see that GLn(F ) is, in fact, a group: matrix multiplication is associative; the identity element is In, the n × n matrix with 1’s along the main diagonal and 0’s everywhere else; and the matrices are invertible by choice. It’s not immediately clear whether GLn(F ) has infinitely many elements when F does. However, such is the case. Let a ∈ F , a 6= 0. −1 Then a · In is an invertible n × n matrix with inverse a · In. In fact, the set of all such × matrices forms a subgroup of GLn(F ) that is isomorphic to F = F \{0}. It is clear that if F is a finite field, then GLn(F ) has only finitely many elements. An interesting question to ask is how many elements it has. Before addressing that question fully, let’s look at some examples. ∼ × Example 1: Let n = 1. Then GLn(Fq) = Fq , which has q − 1 elements. a b Example 2: Let n = 2; let M = ( c d ). Then for M to be invertible, it is necessary and sufficient that ad 6= bc. If a, b, c, and d are all nonzero, then we can fix a, b, and c arbitrarily, and d can be anything but a−1bc. This gives us (q − 1)3(q − 2) matrices.
    [Show full text]
  • Basic Topologytaken From
    Notes by Tamal K. Dey, OSU 1 Basic Topology taken from [1] 1 Metric space topology We introduce basic notions from point set topology. These notions are prerequisites for more sophisticated topological ideas—manifolds, homeomorphism, and isotopy—introduced later to study algorithms for topological data analysis. To a layman, the word topology evokes visions of “rubber-sheet topology”: the idea that if you bend and stretch a sheet of rubber, it changes shape but always preserves the underlying structure of how it is connected to itself. Homeomorphisms offer a rigorous way to state that an operation preserves the topology of a domain, and isotopy offers a rigorous way to state that the domain can be deformed into a shape without ever colliding with itself. Topology begins with a set T of points—perhaps the points comprising the d-dimensional Euclidean space Rd, or perhaps the points on the surface of a volume such as a coffee mug. We suppose that there is a metric d(p, q) that specifies the scalar distance between every pair of points p, q ∈ T. In the Euclidean space Rd we choose the Euclidean distance. On the surface of the coffee mug, we could choose the Euclidean distance too; alternatively, we could choose the geodesic distance, namely the length of the shortest path from p to q on the mug’s surface. d Let us briefly review the Euclidean metric. We write points in R as p = (p1, p2,..., pd), d where each pi is a real-valued coordinate. The Euclidean inner product of two points p, q ∈ R is d Rd 1/2 d 2 1/2 hp, qi = Pi=1 piqi.
    [Show full text]
  • The Modal Logic of Potential Infinity, with an Application to Free Choice
    The Modal Logic of Potential Infinity, With an Application to Free Choice Sequences Dissertation Presented in Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy in the Graduate School of The Ohio State University By Ethan Brauer, B.A. ∼6 6 Graduate Program in Philosophy The Ohio State University 2020 Dissertation Committee: Professor Stewart Shapiro, Co-adviser Professor Neil Tennant, Co-adviser Professor Chris Miller Professor Chris Pincock c Ethan Brauer, 2020 Abstract This dissertation is a study of potential infinity in mathematics and its contrast with actual infinity. Roughly, an actual infinity is a completed infinite totality. By contrast, a collection is potentially infinite when it is possible to expand it beyond any finite limit, despite not being a completed, actual infinite totality. The concept of potential infinity thus involves a notion of possibility. On this basis, recent progress has been made in giving an account of potential infinity using the resources of modal logic. Part I of this dissertation studies what the right modal logic is for reasoning about potential infinity. I begin Part I by rehearsing an argument|which is due to Linnebo and which I partially endorse|that the right modal logic is S4.2. Under this assumption, Linnebo has shown that a natural translation of non-modal first-order logic into modal first- order logic is sound and faithful. I argue that for the philosophical purposes at stake, the modal logic in question should be free and extend Linnebo's result to this setting. I then identify a limitation to the argument for S4.2 being the right modal logic for potential infinity.
    [Show full text]
  • TOPOLOGY and ITS APPLICATIONS the Number of Complements in The
    TOPOLOGY AND ITS APPLICATIONS ELSEVIER Topology and its Applications 55 (1994) 101-125 The number of complements in the lattice of topologies on a fixed set Stephen Watson Department of Mathematics, York Uniuersity, 4700 Keele Street, North York, Ont., Canada M3J IP3 (Received 3 May 1989) (Revised 14 November 1989 and 2 June 1992) Abstract In 1936, Birkhoff ordered the family of all topologies on a set by inclusion and obtained a lattice with 1 and 0. The study of this lattice ought to be a basic pursuit both in combinatorial set theory and in general topology. In this paper, we study the nature of complementation in this lattice. We say that topologies 7 and (T are complementary if and only if 7 A c = 0 and 7 V (T = 1. For simplicity, we call any topology other than the discrete and the indiscrete a proper topology. Hartmanis showed in 1958 that any proper topology on a finite set of size at least 3 has at least two complements. Gaifman showed in 1961 that any proper topology on a countable set has at least two complements. In 1965, Steiner showed that any topology has a complement. The question of the number of distinct complements a topology on a set must possess was first raised by Berri in 1964 who asked if every proper topology on an infinite set must have at least two complements. In 1969, Schnare showed that any proper topology on a set of infinite cardinality K has at least K distinct complements and at most 2” many distinct complements.
    [Show full text]
  • Order Types and Structure of Orders
    ORDER TYPES AND STRUCTURE OF ORDERS BY ANDRE GLEYZALp) 1. Introduction. This paper is concerned with operations on order types or order properties a and the construction of order types related to a. The reference throughout is to simply or linearly ordered sets, and we shall speak of a as either property or type. Let a and ß be any two order types. An order A will be said to be of type aß if it is the sum of /3-orders (orders of type ß) over an a-order; i.e., if A permits of decomposition into nonoverlapping seg- ments each of order type ß, the segments themselves forming an order of type a. We have thus associated with every pair of order types a and ß the product order type aß. The definition of product for order types automatically associates with every order type a the order types aa = a2, aa2 = a3, ■ ■ ■ . We may further- more define, for all ordinals X, a Xth power of a, a\ and finally a limit order type a1. This order type has certain interesting properties. It has closure with respect to the product operation, for the sum of ar-orders over an a7-order is an a'-order, i.e., a'al = aI. For this reason we call a1 iterative. In general, we term an order type ß having the property that ßß = ß iterative, a1 has the following postulational identification: 1. a7 is a supertype of a; that is to say, all a-orders are a7-orders. 2. a1 is iterative.
    [Show full text]
  • MA/CSSE 473 Day 15 Return Exam
    MA/CSSE 473 Day 15 Return Exam Student questions Towers of Hanoi Subsets Ordered Permutations MA/CSSE 473 Day 13 • Student Questions on exam or anything else •Towers of Hanoi • Subset generation –Gray code • Permutations and order 1 Towers of Hanoi •Move all disks from peg A to peg B •One at a time • Never place larger disk on top of a smaller disk •Demo •Code • Recurrence and solution Towers of Hanoi code Recurrence for number of moves, and its solution? 2 Permutations and order number permutation number permutation •Given a permutation 0 0123 12 2013 of 0, 1, …, n‐1, can 1 0132 13 2031 2 0213 14 2103 we directly find the 3 0231 15 2130 next permutation in 4 0312 16 2301 the lexicographic 5 0321 17 2310 sequence? 6 1023 18 3012 7 1032 19 3021 •Given a permutation 8 1203 20 3102 of 0..n‐1, can we 9 1230 21 3120 determine its 10 1302 22 3201 11 1320 23 3210 permutation sequence number? •Given n and i, can we directly generate the ith permutation of 0, …, n‐1? Subset generation • Goal: generate all subsets of {0, 1, 2, …, N‐1} • Bottom‐up (decrease‐by‐one) approach •First generate Sn‐1, the collection of all subsets of {0, …, N‐2} •Then Sn = Sn‐1 { Sn‐1 {n‐1} : sSn‐1} 3 Subset generation • Numeric approach: Each subset of {0, …, N‐1} corresponds to an bit string of length N where the ith bit is 1 iff i is in the subset. •So each subset can be represented by N bits.
    [Show full text]
  • Handout from Today's Lecture
    MA532 Lecture Timothy Kohl Boston University April 23, 2020 Timothy Kohl (Boston University) MA532 Lecture April 23, 2020 1 / 26 Cardinal Arithmetic Recall that one may define addition and multiplication of ordinals α = ot(A, A) β = ot(B, B ) α + β and α · β by constructing order relations on A ∪ B and B × A. For cardinal numbers the foundations are somewhat similar, but also somewhat simpler since one need not refer to orderings. Definition For sets A, B where |A| = α and |B| = β then α + β = |(A × {0}) ∪ (B × {1})|. Timothy Kohl (Boston University) MA532 Lecture April 23, 2020 2 / 26 The curious part of the definition is the two sets A × {0} and B × {1} which can be viewed as subsets of the direct product (A ∪ B) × {0, 1} which basically allows us to add |A| and |B|, in particular since, in the usual formula for the size of the union of two sets |A ∪ B| = |A| + |B| − |A ∩ B| which in this case is bypassed since, by construction, (A × {0}) ∩ (B × {1})= ∅ regardless of the nature of A ∩ B. Timothy Kohl (Boston University) MA532 Lecture April 23, 2020 3 / 26 Definition For sets A, B where |A| = α and |B| = β then α · β = |A × B|. One immediate consequence of these definitions is the following. Proposition If m, n are finite ordinals, then as cardinals one has |m| + |n| = |m + n|, (where the addition on the right is ordinal addition in ω) meaning that ordinal addition and cardinal addition agree. Proof. The simplest proof of this is to define a bijection f : (m × {0}) ∪ (n × {1}) → m + n by f (hr, 0i)= r for r ∈ m and f (hs, 1i)= m + s for s ∈ n.
    [Show full text]
  • Math 501. Homework 2 September 15, 2009 ¶ 1. Prove Or
    Math 501. Homework 2 September 15, 2009 ¶ 1. Prove or provide a counterexample: (a) If A ⊂ B, then A ⊂ B. (b) A ∪ B = A ∪ B (c) A ∩ B = A ∩ B S S (d) i∈I Ai = i∈I Ai T T (e) i∈I Ai = i∈I Ai Solution. (a) True. (b) True. Let x be in A ∪ B. Then x is in A or in B, or in both. If x is in A, then B(x, r) ∩ A , ∅ for all r > 0, and so B(x, r) ∩ (A ∪ B) , ∅ for all r > 0; that is, x is in A ∪ B. For the reverse containment, note that A ∪ B is a closed set that contains A ∪ B, there fore, it must contain the closure of A ∪ B. (c) False. Take A = (0, 1), B = (1, 2) in R. (d) False. Take An = [1/n, 1 − 1/n], n = 2, 3, ··· , in R. (e) False. ¶ 2. Let Y be a subset of a metric space X. Prove that for any subset S of Y, the closure of S in Y coincides with S ∩ Y, where S is the closure of S in X. ¶ 3. (a) If x1, x2, x3, ··· and y1, y2, y3, ··· both converge to x, then the sequence x1, y1, x2, y2, x3, y3, ··· also converges to x. (b) If x1, x2, x3, ··· converges to x and σ : N → N is a bijection, then xσ(1), xσ(2), xσ(3), ··· also converges to x. (c) If {xn} is a sequence that does not converge to y, then there is an open ball B(y, r) and a subsequence of {xn} outside B(y, r).
    [Show full text]
  • Lecture 13: Basis for a Topology
    Lecture 13: Basis for a Topology 1 Basis for a Topology Lemma 1.1. Let (X; T) be a topological space. Suppose that C is a collection of open sets of X such that for each open set U of X and each x in U, there is an element C 2 C such that x 2 C ⊂ U. Then C is the basis for the topology of X. Proof. In order to show that C is a basis, need to show that C satisfies the two properties of basis. To show the first property, let x be an element of the open set X. Now, since X is open, then, by hypothesis there exists an element C of C such that x 2 C ⊂ X. Thus C satisfies the first property of basis. To show the second property of basis, let x 2 X and C1;C2 be open sets in C such that x 2 C1 and x 2 C2. This implies that C1 \ C2 is also an open set in C and x 2 C1 \ C2. Then, by hypothesis, there exists an open set C3 2 C such that x 2 C3 ⊂ C1 \ C2. Thus, C satisfies the second property of basis too and hence, is indeed a basis for the topology on X. On many occasions it is much easier to show results about a topological space by arguing in terms of its basis. For example, to determine whether one topology is finer than the other, it is easier to compare the two topologies in terms of their bases.
    [Show full text]