5. Locally Compact Spaces

Total Page:16

File Type:pdf, Size:1020Kb

5. Locally Compact Spaces 5. Locally compact spaces Definition. A locally compact space is a Hausdorff topological space with the property (lc) Every point has a compact neighborhood. One key feature of locally compact spaces is contained in the following; Lemma 5.1. Let X be a locally compact space, let K be a compact set in X, and let D be an open subset, with K ⊂ D. Then there exists an open set E with: (i) E compact; (ii) K ⊂ E ⊂ E ⊂ D. Proof. Let us start with the following Particular case: Assume K is a singleton K = {x}. Start off by choosing a compact neighborhood N of x. Using the results from section 4, when equipped with the induced topology, the set N is normal. In particular, if we consider the closed sets A = {x} and B = N r D (which are also closed in the induced topology), it follows that there exist sets U, V ⊂ N, such that • U ⊃ {x}, V ⊃ B, U ∩ V = ∅; • U and V are open in the induced topology on N. The second property means that there exist open sets U0,V0 ⊂ X, such that U = N ∩ U0 and V = N ∩ V0. Let E = Int(U). By construction E is open, and E 3 x. Also, since E ⊂ U ⊂ N, it follows that (1) E ⊂ N = N. In particular this gives the compactness of E. Finally, since we obviously have E ∩ V0 ⊂ U ∩ V0 = N ∩ U0 ∩ V0 = U ∩ V = ∅, we get E ⊂ X r V0, so using the fact that X r V0 is closed, we also get the inclusion E ⊂ X r V0. Finally, combining this with (1) and with the inclusion N r D ⊂ V ⊂ V0, we will get E ⊂ N ∩ [X r V0] ⊂ N ∩ [X r (N r D)] = N ∩ D ⊂ D, and we are done. Having proven the particular case, we proceed now with the general case. For every x ∈ K we use the particular case to find an open set E(x), with E(x) compact, S and such that x ∈ E(x) ⊂ E(x) ⊂ D. Since we clearly have K ⊂ x∈K E(x), by compactness, there exist x1, . , xn ∈ K, such that K ⊂ E(x1)∪· · ·∪E(xn). Notice that if we take E = E(x1) ∪ · · · ∪ E(xn), then we clearly have K ⊂ E ⊂ E ⊂ E(x1) ∪ · · · ∪ E(xn) ⊂ D, and we are done. One of the most useful result in the analysis on locally compact spaces is the following. 25 26 CHAPTER I: TOPOLOGY PRELIMINARIES Theorem 5.1 (Urysohn’s Lemma for locally compact spaces). Let X be a locally compact space, and let K, F ⊂ X be two disjoint sets, with K compact, and F closed. Then there exists a continuous function f : X → [0, 1] such that f = 1 K and f F = 0. Proof. Apply Lemma 5.1 for the pair K ⊂ X rF and find an open set E, with E compact, such that K ⊂ E ⊂ E ⊂ X r F . Apply again Lemma 5.1 for the pair K ⊂ E and find another open set G with G compact, such that K ⊂ G ⊂ G ⊂ E. Let us work for the moment in the space E (equipped with the induced topol- ogy). This is a compact Hausdorff space, hence it is normal. In particular, using Urysohn Lemma (see section 1) there exists a continuous function g; E → [0, 1] such that g = 0 and g = 0. Let us now define the function f : X → [0, 1] by K ErG g(x) if x ∈ E f(x) = 0 if x ∈ X r E Notice that f = g , so f is continuous. If we take the open set A = X r G, E E E then it is also clear that f = 0. So now we have two open sets E and A, with A A∪E = X, and f and f both continuous. Then it is clear that f is continuous. A E The other two properties f K = 1 and f F = 0 are obvious. We now discuss an important notion which makes the linkage between locally compact spaces and compact spaces Definition. Let X be a locally compact space. By a compactification of X one means a pair (θ, T ) consisting of a compact Hausdorff space T , and of a continuous map θ : X → T , with the following properties (i) θ(X) is a dense open subset of T ; (ii) when we equip θ(X) with the induced topology, the map θ : X → θ(X) is a homeomorphism. Notice that, when X is already compact, any compactification (θ, T ) of X is nec- essarily made up of a compact space T , and a homeomorphism θ : X → T . Examples 5.1. A. Take [−∞, ∞] = R ∪ {−∞, ∞}, with the “usual” topology, in which a set D ⊂ [−∞, ∞] is open if D = D0 ∪ D1 ∪ D2, where D0 is open in R and D1,D2 ∈ ∅ ∪ (a, ∞]: a ∈ R ∪ [−∞, a): a ∈ R . Then [−∞, ∞] is a compactification of R B. (Alexandrov compactification) Suppose X is a locally compact space, which is not compact. We form a disjoint union with a singleton Xα = X t {∞}, and we equip the space Xα with the topology in which a subset D ⊂ Xα is declared to be open, if either D is an open subset of X, or there exists some compact subset K ⊂ X, such that D = (XrK)t{∞}. Define the inclsuion map ι : X,→ Xα. Then (ι, Xα) is a compactification of X, which is called the Alexandrov compactification. The fact that ι(X) is open in Xα, and ι : X → ι(X) is a homeomorphism, is clear. The density of ι(X) in Xα is also clear, since every open set D ⊂ Xα, with D 3 ∞, is of the form (X r K) t {∞}, for some compact set K ⊂ X, and then we have D ∩ ι(X) = ι(X r K), which is non-empty, because X is not compact. §5. Locally compact spaces 27 Remark that, if X is already compact, we can still define the topological space Xα = X t {∞}, but this time the singleton set {∞} will be also be open (equiv- alently ∞ is an isolated point in Xα). Although ι(X) will still be open in Xα, it will not be dense in Xα. Remark 5.1. Let K be some compact Hausdorff space, and let p ∈ K be some point with the property that the set X = K r {p} is non-compact. Of course, since X is open, it follows that X is a locally compact space, when equipped with the induced topology. If we denote by Xα the Alexandrov compactification of X, then the map Ψ : Xα → K, defined by x if x ∈ X Ψ(x) = p if x = ∞ is a homeomorphism. Indeed (see Corollary 4.1), since Xα is compact, and Ψ is obviously bijective, all we need to prove is the fact that Ψ is continuous. Start with some open set D ⊂ K, and let us show that Ψ−1(D) is open in Xα. If D 63 p, then D is open in X, so Ψ−1(D) = D is open in Xα. If D 3 p, then −1 Ψ (D) = [D r {p}] ∪ {∞} = [X r (K r D)] ∪ {∞}, Of course, since K is compact and D ⊂ K is open, it follows that K r D is closed, hence compact in K. Since K r D ⊂ X, it follows that K r D is also compact in X, so the above equality shows that f −1(D) is indeed open in Xα. One should regard the Alexandrov compactification as a minimal one. More precisely, one has the following. Proposition 5.1. Suppose X is a locally compact space which is non-compact. Let (θ, T ) be a compactification of X, and let Xα = X t {∞} be the Alexandrov compactification. Then there exists a unique continuous map Ψ: T → Xα, such that (Ψ◦θ)(x) = x, ∀ x ∈ X. Moreover, the map Ψ has the property that Ψ(y) = ∞, ∀ y ∈ T r θ(X). Proof. The uniqueness part is pretty obvious, since θ(X) is dense in T . For the existence, we use the map θ : X → T to identify X with an open dense subset of T , and we define Ψ : T → Xα by x if x ∈ X Ψ(x) = ∞ if x ∈ T r X so that all we have to prove is the fact that Ψ is continuous. Start with some open set D in Xα, and let us prove that Ψ−1 is open in T . There are two cases. Case I: D ⊂ X. This case is trivial, Φ−1(D) = D, and D is open in X, hence also open in T . Case II: D 6⊂ X. In this case D 3 ∞, so there exists some compact set K in X, such that D = Xα r K = (X r K) ∪ {∞}. We then have −1 −1 −1 Ψ (D) = Ψ (X r K) ∪ Ψ ({∞}) = (X r K) ∪ (T r X) = T r K. Since K is compact in X, it will be compact in T as well. In particular, K is closed −1 in T , hence the set Ψ (D) = T r K is indeed open in T . It turns out that there exists another compactification which is described below, which can be regarded as the largest. 28 CHAPTER I: TOPOLOGY PRELIMINARIES Theorem 5.2 (Stone-Cech)˘ . Let X be a locally compact space. Consider the set F = {f : X → [0, 1] : f continuous }, and consider the product space Y T = [0, 1], f∈F equipped with the product topology, and define the map θ : X → T by θ(x) = f(x) f∈F , ∀ x ∈ X.
Recommended publications
  • 1.1.5 Hausdorff Spaces
    1. Preliminaries Proof. Let xn n N be a sequence of points in X convergent to a point x X { } 2 2 and let U (f(x)) in Y . It is clear from Definition 1.1.32 and Definition 1.1.5 1 2F that f − (U) (x). Since xn n N converges to x,thereexistsN N s.t. 1 2F { } 2 2 xn f − (U) for all n N.Thenf(xn) U for all n N. Hence, f(xn) n N 2 ≥ 2 ≥ { } 2 converges to f(x). 1.1.5 Hausdor↵spaces Definition 1.1.40. A topological space X is said to be Hausdor↵ (or sepa- rated) if any two distinct points of X have neighbourhoods without common points; or equivalently if: (T2) two distinct points always lie in disjoint open sets. In literature, the Hausdor↵space are often called T2-spaces and the axiom (T2) is said to be the separation axiom. Proposition 1.1.41. In a Hausdor↵space the intersection of all closed neigh- bourhoods of a point contains the point alone. Hence, the singletons are closed. Proof. Let us fix a point x X,whereX is a Hausdor↵space. Denote 2 by C the intersection of all closed neighbourhoods of x. Suppose that there exists y C with y = x. By definition of Hausdor↵space, there exist a 2 6 neighbourhood U(x) of x and a neighbourhood V (y) of y s.t. U(x) V (y)= . \ ; Therefore, y/U(x) because otherwise any neighbourhood of y (in particular 2 V (y)) should have non-empty intersection with U(x).
    [Show full text]
  • The Non-Hausdorff Number of a Topological Space
    THE NON-HAUSDORFF NUMBER OF A TOPOLOGICAL SPACE IVAN S. GOTCHEV Abstract. We call a non-empty subset A of a topological space X finitely non-Hausdorff if for every non-empty finite subset F of A and every family fUx : x 2 F g of open neighborhoods Ux of x 2 F , \fUx : x 2 F g 6= ; and we define the non-Hausdorff number nh(X) of X as follows: nh(X) := 1 + supfjAj : A is a finitely non-Hausdorff subset of Xg. Using this new cardinal function we show that the following three inequalities are true for every topological space X d(X) (i) jXj ≤ 22 · nh(X); 2d(X) (ii) w(X) ≤ 2(2 ·nh(X)); (iii) jXj ≤ d(X)χ(X) · nh(X) and (iv) jXj ≤ nh(X)χ(X)L(X) is true for every T1-topological space X, where d(X) is the density, w(X) is the weight, χ(X) is the character and L(X) is the Lindel¨ofdegree of the space X. The first three inequalities extend to the class of all topological spaces d(X) Pospiˇsil'sinequalities that for every Hausdorff space X, jXj ≤ 22 , 2d(X) w(X) ≤ 22 and jXj ≤ d(X)χ(X). The fourth inequality generalizes 0 to the class of all T1-spaces Arhangel ski˘ı’s inequality that for every Hausdorff space X, jXj ≤ 2χ(X)L(X). It is still an open question if 0 Arhangel ski˘ı’sinequality is true for all T1-spaces. It follows from (iv) that the answer of this question is in the afirmative for all T1-spaces with nh(X) not greater than the cardianilty of the continuum.
    [Show full text]
  • Topological Description of Riemannian Foliations with Dense Leaves
    TOPOLOGICAL DESCRIPTION OF RIEMANNIAN FOLIATIONS WITH DENSE LEAVES J. A. AL¶ VAREZ LOPEZ*¶ AND A. CANDELy Contents Introduction 1 1. Local groups and local actions 2 2. Equicontinuous pseudogroups 5 3. Riemannian pseudogroups 9 4. Equicontinuous pseudogroups and Hilbert's 5th problem 10 5. A description of transitive, compactly generated, strongly equicontinuous and quasi-e®ective pseudogroups 14 6. Quasi-analyticity of pseudogroups 15 References 16 Introduction Riemannian foliations occupy an important place in geometry. An excellent survey is A. Haefliger’s Bourbaki seminar [6], and the book of P. Molino [13] is the standard reference for riemannian foliations. In one of the appendices to this book, E. Ghys proposes the problem of developing a theory of equicontinuous foliated spaces parallel- ing that of riemannian foliations; he uses the suggestive term \qualitative riemannian foliations" for such foliated spaces. In our previous paper [1], we discussed the structure of equicontinuous foliated spaces and, more generally, of equicontinuous pseudogroups of local homeomorphisms of topological spaces. This concept was di±cult to develop because of the nature of pseudogroups and the failure of having an in¯nitesimal characterization of local isome- tries, as one does have in the riemannian case. These di±culties give rise to two versions of equicontinuity: a weaker version seems to be more natural, but a stronger version is more useful to generalize topological properties of riemannian foliations. Another relevant property for this purpose is quasi-e®ectiveness, which is a generalization to pseudogroups of e®ectiveness for group actions. In the case of locally connected fo- liated spaces, quasi-e®ectiveness is equivalent to the quasi-analyticity introduced by Date: July 19, 2002.
    [Show full text]
  • MTH 304: General Topology Semester 2, 2017-2018
    MTH 304: General Topology Semester 2, 2017-2018 Dr. Prahlad Vaidyanathan Contents I. Continuous Functions3 1. First Definitions................................3 2. Open Sets...................................4 3. Continuity by Open Sets...........................6 II. Topological Spaces8 1. Definition and Examples...........................8 2. Metric Spaces................................. 11 3. Basis for a topology.............................. 16 4. The Product Topology on X × Y ...................... 18 Q 5. The Product Topology on Xα ....................... 20 6. Closed Sets.................................. 22 7. Continuous Functions............................. 27 8. The Quotient Topology............................ 30 III.Properties of Topological Spaces 36 1. The Hausdorff property............................ 36 2. Connectedness................................. 37 3. Path Connectedness............................. 41 4. Local Connectedness............................. 44 5. Compactness................................. 46 6. Compact Subsets of Rn ............................ 50 7. Continuous Functions on Compact Sets................... 52 8. Compactness in Metric Spaces........................ 56 9. Local Compactness.............................. 59 IV.Separation Axioms 62 1. Regular Spaces................................ 62 2. Normal Spaces................................ 64 3. Tietze's extension Theorem......................... 67 4. Urysohn Metrization Theorem........................ 71 5. Imbedding of Manifolds..........................
    [Show full text]
  • General Topology
    General Topology Tom Leinster 2014{15 Contents A Topological spaces2 A1 Review of metric spaces.......................2 A2 The definition of topological space.................8 A3 Metrics versus topologies....................... 13 A4 Continuous maps........................... 17 A5 When are two spaces homeomorphic?................ 22 A6 Topological properties........................ 26 A7 Bases................................. 28 A8 Closure and interior......................... 31 A9 Subspaces (new spaces from old, 1)................. 35 A10 Products (new spaces from old, 2)................. 39 A11 Quotients (new spaces from old, 3)................. 43 A12 Review of ChapterA......................... 48 B Compactness 51 B1 The definition of compactness.................... 51 B2 Closed bounded intervals are compact............... 55 B3 Compactness and subspaces..................... 56 B4 Compactness and products..................... 58 B5 The compact subsets of Rn ..................... 59 B6 Compactness and quotients (and images)............. 61 B7 Compact metric spaces........................ 64 C Connectedness 68 C1 The definition of connectedness................... 68 C2 Connected subsets of the real line.................. 72 C3 Path-connectedness.......................... 76 C4 Connected-components and path-components........... 80 1 Chapter A Topological spaces A1 Review of metric spaces For the lecture of Thursday, 18 September 2014 Almost everything in this section should have been covered in Honours Analysis, with the possible exception of some of the examples. For that reason, this lecture is longer than usual. Definition A1.1 Let X be a set. A metric on X is a function d: X × X ! [0; 1) with the following three properties: • d(x; y) = 0 () x = y, for x; y 2 X; • d(x; y) + d(y; z) ≥ d(x; z) for all x; y; z 2 X (triangle inequality); • d(x; y) = d(y; x) for all x; y 2 X (symmetry).
    [Show full text]
  • Dense Embeddings of Sigma-Compact, Nowhere Locallycompact Metric Spaces
    proceedings of the american mathematical society Volume 95. Number 1, September 1985 DENSE EMBEDDINGS OF SIGMA-COMPACT, NOWHERE LOCALLYCOMPACT METRIC SPACES PHILIP L. BOWERS Abstract. It is proved that a connected complete separable ANR Z that satisfies the discrete n-cells property admits dense embeddings of every «-dimensional o-compact, nowhere locally compact metric space X(n e N U {0, oo}). More gener- ally, the collection of dense embeddings forms a dense Gs-subset of the collection of dense maps of X into Z. In particular, the collection of dense embeddings of an arbitrary o-compact, nowhere locally compact metric space into Hubert space forms such a dense Cs-subset. This generalizes and extends a result of Curtis [Cu,]. 0. Introduction. In [Cu,], D. W. Curtis constructs dense embeddings of a-compact, nowhere locally compact metric spaces into the separable Hilbert space l2. In this paper, we extend and generalize this result in two distinct ways: first, we show that any dense map of a a-compact, nowhere locally compact space into Hilbert space is strongly approximable by dense embeddings, and second, we extend this result to finite-dimensional analogs of Hilbert space. Specifically, we prove Theorem. Let Z be a complete separable ANR that satisfies the discrete n-cells property for some w e ÍV U {0, oo} and let X be a o-compact, nowhere locally compact metric space of dimension at most n. Then for every map f: X -» Z such that f(X) is dense in Z and for every open cover <%of Z, there exists an embedding g: X —>Z such that g(X) is dense in Z and g is °U-close to f.
    [Show full text]
  • DEFINITIONS and THEOREMS in GENERAL TOPOLOGY 1. Basic
    DEFINITIONS AND THEOREMS IN GENERAL TOPOLOGY 1. Basic definitions. A topology on a set X is defined by a family O of subsets of X, the open sets of the topology, satisfying the axioms: (i) ; and X are in O; (ii) the intersection of finitely many sets in O is in O; (iii) arbitrary unions of sets in O are in O. Alternatively, a topology may be defined by the neighborhoods U(p) of an arbitrary point p 2 X, where p 2 U(p) and, in addition: (i) If U1;U2 are neighborhoods of p, there exists U3 neighborhood of p, such that U3 ⊂ U1 \ U2; (ii) If U is a neighborhood of p and q 2 U, there exists a neighborhood V of q so that V ⊂ U. A topology is Hausdorff if any distinct points p 6= q admit disjoint neigh- borhoods. This is almost always assumed. A set C ⊂ X is closed if its complement is open. The closure A¯ of a set A ⊂ X is the intersection of all closed sets containing X. A subset A ⊂ X is dense in X if A¯ = X. A point x 2 X is a cluster point of a subset A ⊂ X if any neighborhood of x contains a point of A distinct from x. If A0 denotes the set of cluster points, then A¯ = A [ A0: A map f : X ! Y of topological spaces is continuous at p 2 X if for any open neighborhood V ⊂ Y of f(p), there exists an open neighborhood U ⊂ X of p so that f(U) ⊂ V .
    [Show full text]
  • INTRODUCTION to ALGEBRAIC TOPOLOGY 1 Category And
    INTRODUCTION TO ALGEBRAIC TOPOLOGY (UPDATED June 2, 2020) SI LI AND YU QIU CONTENTS 1 Category and Functor 2 Fundamental Groupoid 3 Covering and fibration 4 Classification of covering 5 Limit and colimit 6 Seifert-van Kampen Theorem 7 A Convenient category of spaces 8 Group object and Loop space 9 Fiber homotopy and homotopy fiber 10 Exact Puppe sequence 11 Cofibration 12 CW complex 13 Whitehead Theorem and CW Approximation 14 Eilenberg-MacLane Space 15 Singular Homology 16 Exact homology sequence 17 Barycentric Subdivision and Excision 18 Cellular homology 19 Cohomology and Universal Coefficient Theorem 20 Hurewicz Theorem 21 Spectral sequence 22 Eilenberg-Zilber Theorem and Kunneth¨ formula 23 Cup and Cap product 24 Poincare´ duality 25 Lefschetz Fixed Point Theorem 1 1 CATEGORY AND FUNCTOR 1 CATEGORY AND FUNCTOR Category In category theory, we will encounter many presentations in terms of diagrams. Roughly speaking, a diagram is a collection of ‘objects’ denoted by A, B, C, X, Y, ··· , and ‘arrows‘ between them denoted by f , g, ··· , as in the examples f f1 A / B X / Y g g1 f2 h g2 C Z / W We will always have an operation ◦ to compose arrows. The diagram is called commutative if all the composite paths between two objects ultimately compose to give the same arrow. For the above examples, they are commutative if h = g ◦ f f2 ◦ f1 = g2 ◦ g1. Definition 1.1. A category C consists of 1◦. A class of objects: Obj(C) (a category is called small if its objects form a set). We will write both A 2 Obj(C) and A 2 C for an object A in C.
    [Show full text]
  • Completeness Properties of the Generalized Compact-Open Topology on Partial Functions with Closed Domains
    Topology and its Applications 110 (2001) 303–321 Completeness properties of the generalized compact-open topology on partial functions with closed domains L’ubica Holá a, László Zsilinszky b;∗ a Academy of Sciences, Institute of Mathematics, Štefánikova 49, 81473 Bratislava, Slovakia b Department of Mathematics and Computer Science, University of North Carolina at Pembroke, Pembroke, NC 28372, USA Received 4 June 1999; received in revised form 20 August 1999 Abstract The primary goal of the paper is to investigate the Baire property and weak α-favorability for the generalized compact-open topology τC on the space P of continuous partial functions f : A ! Y with a closed domain A ⊂ X. Various sufficient and necessary conditions are given. It is shown, e.g., that .P,τC/ is weakly α-favorable (and hence a Baire space), if X is a locally compact paracompact space and Y is a regular space having a completely metrizable dense subspace. As corollaries we get sufficient conditions for Baireness and weak α-favorability of the graph topology of Brandi and Ceppitelli introduced for applications in differential equations, as well as of the Fell hyperspace topology. The relationship between τC, the compact-open and Fell topologies, respectively is studied; moreover, a topological game is introduced and studied in order to facilitate the exposition of the above results. 2001 Elsevier Science B.V. All rights reserved. Keywords: Generalized compact-open topology; Fell topology; Compact-open topology; Topological games; Banach–Mazur game; Weakly α-favorable spaces; Baire spaces; Graph topology; π-base; Restriction mapping AMS classification: Primary 54C35, Secondary 54B20; 90D44; 54E52 0.
    [Show full text]
  • FINITE TOPOLOGICAL SPACES 1. Introduction: Finite Spaces And
    FINITE TOPOLOGICAL SPACES NOTES FOR REU BY J.P. MAY 1. Introduction: finite spaces and partial orders Standard saying: One picture is worth a thousand words. In mathematics: One good definition is worth a thousand calculations. But, to quote a slogan from a T-shirt worn by one of my students: Calculation is the way to the truth. The intuitive notion of a set in which there is a prescribed description of nearness of points is obvious. Formulating the “right” general abstract notion of what a “topology” on a set should be is not. Distance functions lead to metric spaces, which is how we usually think of spaces. Hausdorff came up with a much more abstract and general notion that is now universally accepted. Definition 1.1. A topology on a set X consists of a set U of subsets of X, called the “open sets of X in the topology U ”, with the following properties. (i) ∅ and X are in U . (ii) A finite intersection of sets in U is in U . (iii) An arbitrary union of sets in U is in U . A complement of an open set is called a closed set. The closed sets include ∅ and X and are closed under finite unions and arbitrary intersections. It is very often interesting to see what happens when one takes a standard definition and tweaks it a bit. The following tweaking of the notion of a topology is due to Alexandroff [1], except that he used a different name for the notion. Definition 1.2. A topological space is an A-space if the set U is closed under arbitrary intersections.
    [Show full text]
  • 18 | Compactification
    18 | Compactification We have seen that compact Hausdorff spaces have several interesting properties that make this class of spaces especially important in topology. If we are working with a space X which is not compact we can ask if X can be embedded into some compact Hausdorff space Y . If such embedding exists we can identify X with a subspace of Y , and some arguments that work for compact Hausdorff spaces will still apply to X. This approach leads to the notion of a compactification of a space. Our goal in this chapter is to determine which spaces have compactifications. We will also show that compactifications of a given space X can be ordered, and we will look for the largest and smallest compactifications of X. 18.1 Proposition. Let X be a topological space. If there exists an embedding j : X → Y such that Y is a compact Hausdorff space then there exists an embedding j1 : X → Z such that Z is compact Hausdorff and j1(X) = Z. Proof. Assume that we have an embedding j : X → Y where Y is a compact Hausdorff space. Let j(X) be the closure of j(X) in Y . The space j(X) is compact (by Proposition 14.13) and Hausdorff, so we can take Z = j(X) and define j1 : X → Z by j1(x) = j(x) for all x ∈ X. 18.2 Definition. A space Z is a compactification of X if Z is compact Hausdorff and there exists an embedding j : X → Z such that j(X) = Z. 18.3 Corollary.
    [Show full text]
  • Is Being Hausdorff a Homotopy Invariant?
    Topology Class Is being Hausdor a homotopy invariant? Yash Chandra Published on: Oct 08, 2020 License: Creative Commons Attribution 4.0 International License (CC-BY 4.0) Topology Class Is being Hausdor a homotopy invariant? Hi Everyone, In class, the concepts of homotopy and homotopy equivalence were defined. In that discussion, homotopy invariants – properties of a topological space that remain invariant under an homotopy equivalence – were introduced. Prof. Terilla left this as an open-ended question whether being Hausdorff is a homotopy invariant or not. We will try to address that question here. I’d like to thank Prof. Terilla for his valuable insights into this. I’d also like to acknowledge fellow Graduate Center students Ben, James, Moshe, and Ryan for informal discussions. The answer to the question is No, being Hausdorff is not a homotopy invariant! We will give a particular example to show that, and then try to give a generic recipe for getting more examples. Before doing anything technical, I think it is reasonable to guess that being Hausdorff is not a homotopy invariant, since homotopy is basically a topological deformation and one would intuitively expect that a non-Hausdorff space can be deformed into a Hausdorff one and vice-versa. Notice that we know that the one-point space is vacuously Hausdorff, so maybe we can find a homotopy equivalence between a non-Hausdorff space and the one-point space. We will try to exploit this idea here. A particular example: Consider the set X = {a, b} with the topology τ = {∅, {a}, X}. (X, τ ) is also known as the Sierpinski Space.
    [Show full text]