Topological Dimensions, Hausdorff Dimensions & Fractals

Total Page:16

File Type:pdf, Size:1020Kb

Topological Dimensions, Hausdorff Dimensions & Fractals Topological dimensions, Hausdor dimensions & fractals Yuval Kohavi, Hadar Davdovich May 2006 1 Abstract We present some basic facts about topological dimension, the motiva- tion, necessary denitions and their interrelations. Finally we discuss the Hausdor dimension and fractals. 1 Topological dimension 1.1 Motivation Topological dimensions denes the basic dierence between related topological sets such as In and Im when n 6= m. The lack of that denition is especially highlighted because of the easy explanation of the geometric dimension. When trying to dene a dimension for a topological space, you might run into several diculties - Unlike a vector space, you can not state that the dimension is the maximum of linear independent vector. Therefore in order to get an intuitive denition for a topological dimension we should look for dierent properties that has the eects of the geometrical dimension. 1.2 General A topological dimension has values in ¡1; 0; 1; 2; 3;::: , and is topological, i.e: if X and Y are homeomorphic then they have the same "dimension". It will also be nice, if Rn gets value n, for each n. There are 3 commonly used denitions: 1. Small inductive dimension (ind) 2. Large inductive dimension (Ind) 3. Lebesgue covering dimension (dim) All these denitions have the required properties, and we will see that they are the same for separable metrizable spaces (ind = Ind = dim). Remainder: metrizable space is a topological space that is homeomorphic to a metric space. 2 2 The inductive dimension 2.1 Intuition Lets look at a cube. What is the (geometrical) dimension of the cube? 3, and what is the (geometrical) dimension of its boundary [square]? 2. Now let's look at the square. What is the dimension of its boundary? well its boundaries are lines, so their dimension is 1. And what about the boundary of the lines? Well their boundary is composed of dots, there dimension is 0. We get a pattern - the dimension of a spaces, is 1 + the dimension of its boundary. As a boundary is well dened in topology, this notion can be easily applied to topological spaces, to create a topological dimension. 3 The small inductive dimension Denition 3.1. The small inductive dimension of X is notated ind(X), and is dened as follows: 1. We say that the dimension of a space X (ind(X)) is -1 i X = ; 2. ind(X) ·n if for every point x 2 X and for every open set U exists an open V, x 2 V such that V ⊆ U, and ind(@V ) · n ¡ 1. Where @V is the boundary of V . 3. ind(X) = n if (2) is true for n, but false for n ¡ 1. 4. ind(X) = 1if for every n, ind(X) · n is false. Remark 3.1. The ind dimension is indeed a topological dimension: X is home- omorphic to Y implies that ind(X) = ind(Y ). This can be shown easily (and inductively) as the denition relays only of open\closed sets. Remark 3.2. An equivalent condition to condition (2) is: ² The space has a base B, and every U 2 B has ind(@U) · n ¡ 1 [You can construct this base using the sets U from condition (2)]. Theorem 3.1. As we would expect, if Y ⊆ X then ind(Y ) · ind(X). Proof. By induction, it is true for ind(X) = ¡1. if ind(X) = n, for every point y 2 X there is an nbd V , with an open set U ⊆ V , such that ind(@V ) · n ¡ 1. Note that VY = Y \ V is a nbd in Y , and UY is open in Y . By induction, it is enough to show that @UY ⊆ @U, Because then ind(@UY ) · n ¡ 1, and therefore ind(Y ) · n = ind(X) (by denition). And indeed Uy ⊆ U; @UY = UY n UY ⊆ U n U = @U Example 3.1. Let's show that ind(R) is 1. for each x 2 R, lets select a nbd, V and a set U = (a; b) ⊆ V , ind(@U) = ind(fa; bg) = 0. This implies that ind(R) · 1. 3 So, it is enough to prove that ind(R) is not 0. But that is easy! if ind(R) = 0, that means that a set U exists, such that: ind(@U) = ¡1 , @U = ;, U is clopen! but if U is clopen, R is disconnected! Because R is connected, we get that ind(R) > 0. So nally, ind(R) = 1 ¦. The same proof can be done for the Sphere S1 and the interval I1. Corollary 3.2. A zero-dimensional space is disconnected. Corollary 3.3. if for every nbd V of x 2 X exists a clopen set U ⊆ V then ind(X) = 0 Example 3.2. What is the dimension of C the cantor set? Lets look at C in R, C has no interior points (fact). That means, that for each nbd U = (a; b) of x 2 C, we can nd two points c; d such that a · c · x · d · b, and c; d2 = C. Lets look at C as a subspace, V = (c; d) is open, and closed ) ind(C) = 0 Note that these results are not exactly what we expect. The cantor set has the same 'size' as R: 2@0 , so we should expect it to have the same dimension. But then again, the cantor set has no interval in it. Its dimension should be then, somewhere between 0 and 1. We shall later learn of a dimension denition that suits our expectations. It is worth to mention, that the cantor set is universal 0-dimensional space of separable metrizable spaces. meaning that every other space separable metric space , is homeomorphic to a sub-space of the cantor set. X; ind(X) = 0 P Remember that a number is in the Cantor set It has the form xi , i2N 3i ; xi 2 f0; 2g. for example: 2 0 2 0 . 3 + 9 + 27 + 81 + ¢ ¢ ¢ We can display this number simply as a fraction of radix 3 0:2020:::. This set of fractions has one-to-one and onto mapping to the set of binary fractions 0:1010:::. We simply replaced all the 2's with 1's. The latter set of fractions is more familiar to topology, it is the space:D@0 ;D = f0; 1g, D with the discrete topology. Proposition 3.4. The D@0 is homeomorphic to the Cantor set. Now it is enough to show that every zero dimensional separable metrizable space is homeomorphic to a sub-spaces of D@0 . Remark 3.3. Every zero dimensional spaces has a clopen base (by the alternative denition). If the space is also metrizable and separable, it has a countable clopen base (because then every base has a countable base contained in it). Theorem 3.5. The Cantor set is a universal space for all the zero-dimensional metrizable separable space. Proof. Let X be a separable metrizable space. It is enough to show that X is homeomorphic to a sub-set of D@0 . Let be its clopen countable base. We dene to be: ½ B = fBig f fi(x) = 1 x 2 B i . Dene f(x) = (f (x); f (x);:::). f is the homeomorphism 0 otherwise 1 2 that we wanted. ¦ 4 An interesting question we can ask about this dimension, is wether ind(Rn) = n. The answer is yes,we will start by showing that ind(Rn) · n. Proposition 3.6. The dimension of Rn;Sn;In · n. We can show this by induction. we have already shown that the dimension of R1;S1;I1is 1. For every point x 2 Rn;Sn;In, there is a nbd U with an open set V which is homeomorphic to Sn¡1, and by induction the proof is complete ¦ In order to show that ind(Rn) ¸ n, We shall have some denitions and theorems. Let's start by dening a partition between sets: Denition 3.2. A Partition L on between A and B exists, if there are open sets U; W , A ⊆ U; B ⊆ W , such that W \ U = ; and L = U c \ W c. Theorem 3.7. ind(X) · n ,For every point x and every closed set B there is a partition L, such that:ind(L) · n ¡ 1. Proof. ()) If x 2 X and B is closed, Then there is a set V , such that V \ B = ;. By denition of ind, there is a U ⊆ V , x 2 U, ind(@U) · n ¡ 1 now, let W = U and L = @U. (() Let x 2 X and V an nbd of x. The set B = V c is closed, and therefore there are U; W such that x 2 U, B ⊆ W . Note that by denition of B, U ⊆ V . Now W c is closed, therefore U ⊆ W c, and obviously @U ⊆ U ⊆ W c, and by denition @U = U n U_ = U n U ⊆ U c. Therefore @U ⊆ U c \ W c = L ind(U) · n ¡ 1 Theorem 3.8. If X is a metric space and Z is a zero dimensional separable subspace of X, then for all closed set A; B of XA \ B = ;, there is a partition between them L, such that L \ Z = ; Theorem 3.9. X is a separable metrizable space, then ind(X) · n , X is the union of two sub spaces Y; Z such that ind(Y ) · n ¡ 1; ind(Z) · 0. From this we can easily see, that for a separable metrizable space X ind(X) · n , X is the union of Z1;:::;Zn+1, ind(Zi) · 0.
Recommended publications
  • A Guide to Topology
    i i “topguide” — 2010/12/8 — 17:36 — page i — #1 i i A Guide to Topology i i i i i i “topguide” — 2011/2/15 — 16:42 — page ii — #2 i i c 2009 by The Mathematical Association of America (Incorporated) Library of Congress Catalog Card Number 2009929077 Print Edition ISBN 978-0-88385-346-7 Electronic Edition ISBN 978-0-88385-917-9 Printed in the United States of America Current Printing (last digit): 10987654321 i i i i i i “topguide” — 2010/12/8 — 17:36 — page iii — #3 i i The Dolciani Mathematical Expositions NUMBER FORTY MAA Guides # 4 A Guide to Topology Steven G. Krantz Washington University, St. Louis ® Published and Distributed by The Mathematical Association of America i i i i i i “topguide” — 2010/12/8 — 17:36 — page iv — #4 i i DOLCIANI MATHEMATICAL EXPOSITIONS Committee on Books Paul Zorn, Chair Dolciani Mathematical Expositions Editorial Board Underwood Dudley, Editor Jeremy S. Case Rosalie A. Dance Tevian Dray Patricia B. Humphrey Virginia E. Knight Mark A. Peterson Jonathan Rogness Thomas Q. Sibley Joe Alyn Stickles i i i i i i “topguide” — 2010/12/8 — 17:36 — page v — #5 i i The DOLCIANI MATHEMATICAL EXPOSITIONS series of the Mathematical Association of America was established through a generous gift to the Association from Mary P. Dolciani, Professor of Mathematics at Hunter College of the City Uni- versity of New York. In making the gift, Professor Dolciani, herself an exceptionally talented and successfulexpositor of mathematics, had the purpose of furthering the ideal of excellence in mathematical exposition.
    [Show full text]
  • On Dimension and Weight of a Local Contact Algebra
    Filomat 32:15 (2018), 5481–5500 Published by Faculty of Sciences and Mathematics, https://doi.org/10.2298/FIL1815481D University of Nis,ˇ Serbia Available at: http://www.pmf.ni.ac.rs/filomat On Dimension and Weight of a Local Contact Algebra G. Dimova, E. Ivanova-Dimovaa, I. D ¨untschb aFaculty of Math. and Informatics, Sofia University, 5 J. Bourchier Blvd., 1164 Sofia, Bulgaria bCollege of Maths. and Informatics, Fujian Normal University, Fuzhou, China. Permanent address: Dept. of Computer Science, Brock University, St. Catharines, Canada Abstract. As proved in [16], there exists a duality Λt between the category HLC of locally compact Hausdorff spaces and continuous maps, and the category DHLC of complete local contact algebras and appropriate morphisms between them. In this paper, we introduce the notions of weight wa and of dimension dima of a local contact algebra, and we prove that if X is a locally compact Hausdorff space then t t w(X) = wa(Λ (X)), and if, in addition, X is normal, then dim(X) = dima(Λ (X)). 1. Introduction According to Stone’s famous duality theorem [43], the Boolean algebra CO(X) of all clopen (= closed and open) subsets of a zero-dimensional compact Hausdorff space X carries the whole information about the space X, i.e. the space X can be reconstructed from CO(X), up to homeomorphism. It is natural to ask whether the Boolean algebra RC(X) of all regular closed subsets of a compact Hausdorff space X carries the full information about the space X (see Example 2.5 below for RC(X)).
    [Show full text]
  • Professor Smith Math 295 Lecture Notes
    Professor Smith Math 295 Lecture Notes by John Holler Fall 2010 1 October 29: Compactness, Open Covers, and Subcovers. 1.1 Reexamining Wednesday's proof There seemed to be a bit of confusion about our proof of the generalized Intermediate Value Theorem which is: Theorem 1: If f : X ! Y is continuous, and S ⊆ X is a connected subset (con- nected under the subspace topology), then f(S) is connected. We'll approach this problem by first proving a special case of it: Theorem 1: If f : X ! Y is continuous and surjective and X is connected, then Y is connected. Proof of 2: Suppose not. Then Y is not connected. This by definition means 9 non-empty open sets A; B ⊂ Y such that A S B = Y and A T B = ;. Since f is continuous, both f −1(A) and f −1(B) are open. Since f is surjective, both f −1(A) and f −1(B) are non-empty because A and B are non-empty. But the surjectivity of f also means f −1(A) S f −1(B) = X, since A S B = Y . Additionally we have f −1(A) T f −1(B) = ;. For if not, then we would have that 9x 2 f −1(A) T f −1(B) which implies f(x) 2 A T B = ; which is a contradiction. But then these two non-empty open sets f −1(A) and f −1(B) disconnect X. QED Now we want to show that Theorem 2 implies Theorem 1. This will require the help from a few lemmas, whose proofs are on Homework Set 7: Lemma 1: If f : X ! Y is continuous, and S ⊆ X is any subset of X then the restriction fjS : S ! Y; x 7! f(x) is also continuous.
    [Show full text]
  • Hausdorff Dimension of Singular Vectors
    HAUSDORFF DIMENSION OF SINGULAR VECTORS YITWAH CHEUNG AND NICOLAS CHEVALLIER Abstract. We prove that the set of singular vectors in Rd; d ≥ 2; has Hausdorff dimension d2 d d+1 and that the Hausdorff dimension of the set of "-Dirichlet improvable vectors in R d2 d is roughly d+1 plus a power of " between 2 and d. As a corollary, the set of divergent t t −dt trajectories of the flow by diag(e ; : : : ; e ; e ) acting on SLd+1 R= SLd+1 Z has Hausdorff d codimension d+1 . These results extend the work of the first author in [6]. 1. Introduction Singular vectors were introduced by Khintchine in the twenties (see [14]). Recall that d x = (x1; :::; xd) 2 R is singular if for every " > 0, there exists T0 such that for all T > T0 the system of inequalities " (1.1) max jqxi − pij < and 0 < q < T 1≤i≤d T 1=d admits an integer solution (p; q) 2 Zd × Z. In dimension one, only the rational numbers are singular. The existence of singular vectors that do not lie in a rational subspace was proved by Khintchine for all dimensions ≥ 2. Singular vectors exhibit phenomena that cannot occur in dimension one. For instance, when x 2 Rd is singular, the sequence 0; x; 2x; :::; nx; ::: fills the torus Td = Rd=Zd in such a way that there exists a point y whose distance in the torus to the set f0; x; :::; nxg, times n1=d, goes to infinity when n goes to infinity (see [4], chapter V).
    [Show full text]
  • Base-Paracompact Spaces
    View metadata, citation and similar papers at core.ac.uk brought to you by CORE provided by Elsevier - Publisher Connector Topology and its Applications 128 (2003) 145–156 www.elsevier.com/locate/topol Base-paracompact spaces John E. Porter Department of Mathematics and Statistics, Faculty Hall Room 6C, Murray State University, Murray, KY 42071-3341, USA Received 14 June 2001; received in revised form 21 March 2002 Abstract A topological space is said to be totally paracompact if every open base of it has a locally finite subcover. It turns out this property is very restrictive. In fact, the irrationals are not totally paracompact. Here we give a generalization, base-paracompact, to the notion of totally paracompactness, and study which paracompact spaces satisfy this generalization. 2002 Elsevier Science B.V. All rights reserved. Keywords: Paracompact; Base-paracompact; Totally paracompact 1. Introduction A topological space X is paracompact if every open cover of X has a locally finite open refinement. A topological space is said to be totally paracompact if every open base of it has a locally finite subcover. Corson, McNinn, Michael, and Nagata announced [3] that the irrationals have a basis which does not have a locally finite subcover, and hence are not totally paracompact. Ford [6] was the first to give the definition of totally paracompact spaces. He used this property to show the equivalence of the small and large inductive dimension in metric spaces. Ford also showed paracompact, locally compact spaces are totally paracompact. Lelek [9] studied totally paracompact and totally metacompact separable metric spaces. He showed that these two properties are equivalent in separable metric spaces.
    [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]
  • 23. Dimension Dimension Is Intuitively Obvious but Surprisingly Hard to Define Rigorously and to Work With
    58 RICHARD BORCHERDS 23. Dimension Dimension is intuitively obvious but surprisingly hard to define rigorously and to work with. There are several different concepts of dimension • It was at first assumed that the dimension was the number or parameters something depended on. This fell apart when Cantor showed that there is a bijective map from R ! R2. The Peano curve is a continuous surjective map from R ! R2. • The Lebesgue covering dimension: a space has Lebesgue covering dimension at most n if every open cover has a refinement such that each point is in at most n + 1 sets. This does not work well for the spectrums of rings. Example: dimension 2 (DIAGRAM) no point in more than 3 sets. Not trivial to prove that n-dim space has dimension n. No good for commutative algebra as A1 has infinite Lebesgue covering dimension, as any finite number of non-empty open sets intersect. • The "classical" definition. Definition 23.1. (Brouwer, Menger, Urysohn) A topological space has dimension ≤ n (n ≥ −1) if all points have arbitrarily small neighborhoods with boundary of dimension < n. The empty set is the only space of dimension −1. This definition is mostly used for separable metric spaces. Rather amazingly it also works for the spectra of Noetherian rings, which are about as far as one can get from separable metric spaces. • Definition 23.2. The Krull dimension of a topological space is the supre- mum of the numbers n for which there is a chain Z0 ⊂ Z1 ⊂ ::: ⊂ Zn of n + 1 irreducible subsets. DIAGRAM pt ⊂ curve ⊂ A2 For Noetherian topological spaces the Krull dimension is the same as the Menger definition, but for non-Noetherian spaces it behaves badly.
    [Show full text]
  • Chapter 7 Separation Properties
    Chapter VII Separation Axioms 1. Introduction “Separation” refers here to whether or not objects like points or disjoint closed sets can be enclosed in disjoint open sets; “separation properties” have nothing to do with the idea of “separated sets” that appeared in our discussion of connectedness in Chapter 5 in spite of the similarity of terminology.. We have already met some simple separation properties of spaces: the XßX!"and X # (Hausdorff) properties. In this chapter, we look at these and others in more depth. As “more separation” is added to spaces, they generally become nicer and nicer especially when “separation” is combined with other properties. For example, we will see that “enough separation” and “a nice base” guarantees that a space is metrizable. “Separation axioms” translates the German term Trennungsaxiome used in the older literature. Therefore the standard separation axioms were historically named XXXX!"#$, , , , and X %, each stronger than its predecessors in the list. Once these were common terminology, another separation axiom was discovered to be useful and “interpolated” into the list: XÞ"" It turns out that the X spaces (also called $$## Tychonoff spaces) are an extremely well-behaved class of spaces with some very nice properties. 2. The Basics Definition 2.1 A topological space \ is called a 1) X! space if, whenever BÁC−\, there either exists an open set Y with B−Y, CÂY or there exists an open set ZC−ZBÂZwith , 2) X" space if, whenever BÁC−\, there exists an open set Ywith B−YßCÂZ and there exists an open set ZBÂYßC−Zwith 3) XBÁC−\Y# space (or, Hausdorff space) if, whenever , there exist disjoint open sets and Z\ in such that B−YC−Z and .
    [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]
  • Lecture Notes on Foliation Theory
    INDIAN INSTITUTE OF TECHNOLOGY BOMBAY Department of Mathematics Seminar Lectures on Foliation Theory 1 : FALL 2008 Lecture 1 Basic requirements for this Seminar Series: Familiarity with the notion of differential manifold, submersion, vector bundles. 1 Some Examples Let us begin with some examples: m d m−d (1) Write R = R × R . As we know this is one of the several cartesian product m decomposition of R . Via the second projection, this can also be thought of as a ‘trivial m−d vector bundle’ of rank d over R . This also gives the trivial example of a codim. d- n d foliation of R , as a decomposition into d-dimensional leaves R × {y} as y varies over m−d R . (2) A little more generally, we may consider any two manifolds M, N and a submersion f : M → N. Here M can be written as a disjoint union of fibres of f each one is a submanifold of dimension equal to dim M − dim N = d. We say f is a submersion of M of codimension d. The manifold structure for the fibres comes from an atlas for M via the surjective form of implicit function theorem since dfp : TpM → Tf(p)N is surjective at every point of M. We would like to consider this description also as a codim d foliation. However, this is also too simple minded one and hence we would call them simple foliations. If the fibres of the submersion are connected as well, then we call it strictly simple. (3) Kronecker Foliation of a Torus Let us now consider something non trivial.
    [Show full text]
  • On Expanding Locally Finite Collections
    Can. J. Math., Vol. XXIII, No. 1, 1971, pp. 58-68 ON EXPANDING LOCALLY FINITE COLLECTIONS LAWRENCE L. KRAJEWSKI Introduction. A space X is in-expandable, where m is an infinite cardinal, if for every locally finite collection {Ha\ a £ A} of subsets of X with \A\ ^ m (car­ dinality of A S ni ) there exists a locally finite collection of open subsets {Ga \ a £ A} such that Ha C Ga for every a 6 A. X is expandable if it is m-expandable for every cardinal m. The notion of expandability is closely related to that of collection wise normality introduced by Bing [1], X is collectionwise normal if for every discrete collection of subsets {Ha\a € A} there is a discrete collec­ tion of open subsets {Ga\a £ A] such that Ha C Ga for every a 6 A. Expand­ able spaces share many of the properties possessed by collectionwise normal spaces. For example, an expandable developable space is metrizable and an expandable metacompact space is paracompact. In § 2 we study the relationship of expandability with various covering properties and obtain some characterizations of paracompactness involving expandability. It is shown that Xo-expandability is equivalent to countable paracompactness. In § 3, countably perfect maps are studied in relation to expandability and various product theorems are obtained. Section 4 deals with subspaces and various sum theorems. Examples comprise § 5. Definitions of terms not defined here can be found in [1; 5; 16]. 1. Expandability and collectionwise normality. An expandable space need not be regular (Example 5.6), and a completely regular expandable space need not be normal (W in Example 5.3).
    [Show full text]
  • Dimension and Dimensional Reduction in Quantum Gravity
    May 2017 Dimension and Dimensional Reduction in Quantum Gravity S. Carlip∗ Department of Physics University of California Davis, CA 95616 USA Abstract , A number of very different approaches to quantum gravity contain a common thread, a hint that spacetime at very short distances becomes effectively two dimensional. I review this evidence, starting with a discussion of the physical meaning of “dimension” and concluding with some speculative ideas of what dimensional reduction might mean for physics. arXiv:1705.05417v2 [gr-qc] 29 May 2017 ∗email: [email protected] 1 Why Dimensional Reduction? What is the dimension of spacetime? For most of physics, the answer is straightforward and uncontroversial: we know from everyday experience that we live in a universe with three dimensions of space and one of time. For a condensed matter physicist, say, or an astronomer, this is simply a given. There are a few exceptions—surface states in condensed matter that act two-dimensional, string theory in ten dimensions—but for the most part dimension is simply a fixed, and known, external parameter. Over the past few years, though, hints have emerged from quantum gravity suggesting that the dimension of spacetime is dynamical and scale-dependent, and shrinks to d 2 at very small ∼ distances or high energies. The purpose of this review is to summarize this evidence and to discuss some possible implications for physics. 1.1 Dimensional reduction and quantum gravity As early as 1916, Einstein pointed out that it would probably be necessary to combine the newly formulated general theory of relativity with the emerging ideas of quantum mechanics [1].
    [Show full text]