Unit Interval Graphs: Properties and Algorithms

Total Page:16

File Type:pdf, Size:1020Kb

Unit Interval Graphs: Properties and Algorithms Unit Interval Graphs, Properties and Algorithms Sayyed Bashir Sadjad Hamid Zarrabi-Zadeh School of Computer Science University of Waterloo {bssadjad, hzarrabi}@uwaterloo.ca February 19, 2004 1 Introduction In this article we address some of the results on unit interval graphs. In short, a unit interval graph is an interval graph in which all intervals have the same length. These graphs have many applications in bioinformatics, databases, scheduling, measurement theory, etc. [KS96, BFMY83, PY79, Gol80]. We will study some of the problems that arises from such applications later in this article, after addressing properties of this class and available recognition algorithms. There are some other classes of graphs with different definitions which are in fact the same as the class of unit interval graphs. We introduce them in Section 2 and mention theorems that prove their equivalence. There are some efficient algorithms to check whether a graph is unit interval graph. Some of these algorithms are addressed in Section 3. A method for constructing efficient representations of unit interval graphs is given in Section 4. Finally, in Section 5 we investigate some problems on unit interval graphs that have faster algorithms than the general case. We also mention an interesting application of this class in bioinformatics in that section. 2 Definitions and Properties We start this section by giving three different definitions of three graph classes and then mention theorems showing that all these three classes are the same. When there are short proofs for the claims, we will give details of them, but for long proofs we only give references. Throughout this article we assume that the graphs are simple (with no loops or multiple edges) and connected with at least two vertices. To show an edge between vertices v and w we use (v, w) whether G is directed or undirected. So, when it is clear from the context that G is undirected, (v, w) and (w, v) are the same edges. Definition 2.1 An undirected graph G is an interval graph if there is a mapping f from vertices of G to a set I of real intervals such that (v, w) ∈ E(G) if and only if f(v) ∩ f(w) 6= φ. Such a set I is called an interval model for G. 1 v3 I3 I2 I1 I0 v0 v2 v1 Figure 1: A proper or unit interval graph cannot contain a K1,3. An interval graph is both chordal and co-comparability [Gol80]. On the other hand, if a graph is both chordal and co-comparability, then it is an interval graph. Therefore a nice characterization of the interval graphs is the following theorem: Theorem 2.2 (Golumbic [Gol80]) A graph G is an interval graph if and only if G is chordal and G¯ is a comparability graph. Definition 2.3 An interval graph G is a unit interval graph if there exists an interval model I for G in which all intervals are of equal size. In this case, I is called a unit interval model for G. Definition 2.4 An interval graph G is a proper interval graph if there exists an interval model I for G in which no interval properly contains another interval. In this case, I is called a proper interval model for G. It is not hard to show that the class of proper interval graphs is the same as the class of unit interval graphs. First of all, we show that a graph in any of these classes could not contain a K1,3 as an induced subgraph. Theorem 2.5 If G is a unit interval graph or a proper interval graph then G contains no K1,3 as an induced subgraph. Proof: The proof is given in Figure 1. In this figure, a K1,3 is shown on the left and an interval model for it given on the right (Ii is the interval of vi). As it is shown, I3 should be a proper subset of I0, so, K1,3 is not a proper interval graph. Moreover, it is clear that I0 and I3 cannot be of equal size. Therefore, K1,3 is not a unit interval graph. 2 The following theorem shows that this condition is also sufficient and implies that the class of unit interval graphs is the same as proper interval graphs. This theorem is from [BW99] but an extended version of it, which will be addressed later, is in [Gol80, BLS99, Rob69]. 2 Theorem 2.6 For a simple graph G the following three statements are equivalent: 1. G is a unit interval graph. 2. G is an interval graph with no K1,3 as induced subgraph. 3. G is a proper interval graph. Proof: The proof is by going from (2) to (3) and then from (3) to (1) (we know that (1) to (2) holds). To show that (2)⇒(3), we choose an interval model I for G, in which every vertex v of G is represented by an interval Iv of I. Starting from left to right, for any proper inclusion we move the interval with start point at right, to right without violating any edge condition. Suppose that there are two intervals Ix = [a, b] and Iy = [c, d] such that a < c ≤ d ≤ b (that is, Ix properly contains Iy). Since there is no induced K1,3 in G, in at least one of intervals [a, c] and [d, b] there is no endpoint of an interval Iz that does not intersect Iy. So, we can change the intervals Ix and Iy, without violating any edge condition, so that there is no proper inclusion. A similar idea works for going from (3) to (1). Here for each interval, we do some expansion and shrinking to make it unit and continue with the interval on the right. 2 Now, we define a third class of graphs which comes from measurement theories (attributed in [BW99] and [PD03] to [SS58]) and has close relations to unit and proper interval graphs. We have already seen a close connection between comparability graphs and interval graphs. We will see a closer relation between unit interval graphs and semiorders, which are a sub-class of partial orders. The following definition of semiorders is from [Gol80]. Definition 2.7 A partial order S = (V, P ) with elements V and a binary order relation P (on V ) is semiorder, if there is a function f : V → R from V to real numbers s.t. (u, v) ∈ P if and only if f(u) ≥ f(v) + 1. In the above definition, there could be any δ > 0 instead of 1 in the condition f(u) ≥ f(v) + 1. In Theorem 2.8, it is shown that all these definitions are equivalent. Historically, this is the root of unit interval graphs and Theorem 2.8 was proved in [Rob69] many years before [BW99]. Theorem 2.8 (Roberts [Rob69]) Let G = (V, E) be an undirected graph, then the following statements are equivalent: 1. There exists a function f : V → R s.t. (u, w) ∈ E(G) ⇔ |f(u) − f(w)| < 1. 2. There is a semiorder on S(V, P ) on the vertices of V such that (u, w) ∈ P ⇔ (u, w) 6∈ E(G). 3. G¯ is a comparability graph and every transitive orientation of G¯ is a semiorder. 4. G is an interval graph with no induced K1,3. 5. G is a proper interval graph. 3 z w w A. B. x y x y z Figure 2: A representation of Theorem 2.9 6. G is a unit interval graph. In [Rob69] it is mentioned that the (4)⇒(2) part was first observed by Marley through a personal communication with Roberts. Also, from historical point of view, the following theorem of Scott and Suppes [SS58] on semiorders is interesting since it is exactly the condition of having no K1,3 or induced Ck (for k ≥ 4) in the undirected complement of the relation. This is from [Gol80] but is attributed to [SS58]. Theorem 2.9 (Scott and Suppes [SS58]) A partial order relation S defined on V is a semiorder if the following conditions hold: 1. S is irreflexive. 2. (x,y) ∈ S and (z, w) ∈ S imply (x, w) ∈ S or (z,y) ∈ S. 3. (x,y) ∈ S and (y, z) ∈ S imply (x, w) ∈ S or (w, z) ∈ S. To see why these conditions are the same as the previous condition (interval graph with no K1,3), it is easy to check that condition (2) is equivalent to the condition that there is no Ck (for k ≥ 4) in G¯, where (u, w) is an edge in G = (V, E) if and only if neither (u, w) nor (w, u) is in S. This is shown in part A of Figure 2. Notice that in this figure, if there is no other directed edges (relation) between x,y,w and z, then in G there will be a C4 with no chord. On the other hand, notice that any other relation (i.e. other than xSy and zSw) then because of transitivity one of the white arrows should be present in S as well. So this is the necessary and sufficient condition for G to be chordal. Since S is a partial order, G¯ is both chordal and co-comparability, and hence G is interval. On the other hand, condition (3) is equivalent to the condition that there is no induced K1,3 in G (Figure 2 part B). Notice that since S is a partial order it is transitive so if xSy and ySz then xSz and the precondition of (3) means that x,y,z forms a clique in G¯ and the postcondition is that any vertex w in V should be connected to at least one of them in G¯ which means G does not contain any induced K1,3.
Recommended publications
  • Lecture 10: April 20, 2005 Perfect Graphs
    Re-revised notes 4-22-2005 10pm CMSC 27400-1/37200-1 Combinatorics and Probability Spring 2005 Lecture 10: April 20, 2005 Instructor: L´aszl´oBabai Scribe: Raghav Kulkarni TA SCHEDULE: TA sessions are held in Ryerson-255, Monday, Tuesday and Thursday 5:30{6:30pm. INSTRUCTOR'S EMAIL: [email protected] TA's EMAIL: [email protected], [email protected] IMPORTANT: Take-home test Friday, April 29, due Monday, May 2, before class. Perfect Graphs k 1=k Shannon capacity of a graph G is: Θ(G) := limk (α(G )) : !1 Exercise 10.1 Show that α(G) χ(G): (G is the complement of G:) ≤ Exercise 10.2 Show that χ(G H) χ(G)χ(H): · ≤ Exercise 10.3 Show that Θ(G) χ(G): ≤ So, α(G) Θ(G) χ(G): ≤ ≤ Definition: G is perfect if for all induced sugraphs H of G, α(H) = χ(H); i. e., the chromatic number is equal to the clique number. Theorem 10.4 (Lov´asz) G is perfect iff G is perfect. (This was open under the name \weak perfect graph conjecture.") Corollary 10.5 If G is perfect then Θ(G) = α(G) = χ(G): Exercise 10.6 (a) Kn is perfect. (b) All bipartite graphs are perfect. Exercise 10.7 Prove: If G is bipartite then G is perfect. Do not use Lov´asz'sTheorem (Theorem 10.4). 1 Lecture 10: April 20, 2005 2 The smallest imperfect (not perfect) graph is C5 : α(C5) = 2; χ(C5) = 3: For k 2, C2k+1 imperfect.
    [Show full text]
  • I?'!'-Comparability Graphs
    Discrete Mathematics 74 (1989) 173-200 173 North-Holland I?‘!‘-COMPARABILITYGRAPHS C.T. HOANG Department of Computer Science, Rutgers University, New Brunswick, NJ 08903, U.S.A. and B.A. REED Discrete Mathematics Group, Bell Communications Research, Morristown, NJ 07950, U.S.A. In 1981, Chv&tal defined the class of perfectly orderable graphs. This class of perfect graphs contains the comparability graphs. In this paper, we introduce a new class of perfectly orderable graphs, the &comparability graphs. This class generalizes comparability graphs in a natural way. We also prove a decomposition theorem which leads to a structural characteriza- tion of &comparability graphs. Using this characterization, we develop a polynomial-time recognition algorithm and polynomial-time algorithms for the clique and colouring problems for &comparability graphs. 1. Introduction A natural way to color the vertices of a graph is to put them in some order v*cv~<” l < v, and then assign colors in the following manner: (1) Consider the vertxes sequentially (in the sequence given by the order) (2) Assign to each Vi the smallest color (positive integer) not used on any neighbor vi of Vi with j c i. We shall call this the greedy coloring algorithm. An order < of a graph G is perfect if for each subgraph H of G, the greedy algorithm applied to H, gives an optimal coloring of H. An obstruction in a ordered graph (G, <) is a set of four vertices {a, b, c, d} with edges ab, bc, cd (and no other edge) and a c b, d CC. Clearly, if < is a perfect order on G, then (G, C) contains no obstruction (because the chromatic number of an obstruction is twu, but the greedy coloring algorithm uses three colors).
    [Show full text]
  • Proper Interval Graphs and the Guard Problem 1
    DISCRETE N~THEMATICS ELSEVIER Discrete Mathematics 170 (1997) 223-230 Note Proper interval graphs and the guard problem1 Chiuyuan Chen*, Chin-Chen Chang, Gerard J. Chang Department of Applied Mathematics, National Chiao Tung University, Hsinchu 30050, Taiwan Received 26 September 1995; revised 27 August 1996 Abstract This paper is a study of the hamiltonicity of proper interval graphs with applications to the guard problem in spiral polygons. We prove that proper interval graphs with ~> 2 vertices have hamiltonian paths, those with ~>3 vertices have hamiltonian cycles, and those with />4 vertices are hamiltonian-connected if and only if they are, respectively, 1-, 2-, or 3-connected. We also study the guard problem in spiral polygons by connecting the class of nontrivial connected proper interval graphs with the class of stick-intersection graphs of spiral polygons. Keywords." Proper interval graph; Hamiltonian path (cycle); Hamiltonian-connected; Guard; Visibility; Spiral polygon I. Introduction The main purpose of this paper is to study the hamiltonicity of proper interval graphs with applications to the guard problem in spiral polygons. Our terminology and graph notation are standard, see [2], except as indicated. The intersection graph of a family ~ of nonempty sets is derived by representing each set in ~ with a vertex and connecting two vertices with an edge if and only if their corresponding sets intersect. An interval graph is the intersection graph G of a family J of intervals on a real line. J is usually called the interval model for G. A proper interval graph is an interval graph with an interval model ~¢ such that no interval in J properly contains another.
    [Show full text]
  • Cantor Groups
    Integre Technical Publishing Co., Inc. College Mathematics Journal 42:1 November 11, 2010 10:58 a.m. classroom.tex page 60 Cantor Groups Ben Mathes ([email protected]), Colby College, Waterville ME 04901; Chris Dow ([email protected]), University of Chicago, Chicago IL 60637; and Leo Livshits ([email protected]), Colby College, Waterville ME 04901 The Cantor set C is simultaneously small and large. Its cardinality is the same as the cardinality of all real numbers, and every element of C is a limit of other elements of C (it is perfect), but C is nowhere dense and it is a nullset. Being nowhere dense is a kind of topological smallness, meaning that there is no interval in which the set is dense, or equivalently, that its complement is a dense open subset of [0, 1]. Being a nullset is a measure theoretic concept of smallness, meaning that, for every >0, it is possible to cover the set with a sequence of intervals In whose lengths add to a sum less than . It is possible to tweak the construction of C that results in a set H that remains per- fect (and consequently has the cardinality of the continuum), is still nowhere dense, but the measure can be made to attain any number α in the interval [0, 1). The con- struction is a special case of Hewitt and Stromberg’s Cantor-like set [1,p.70],which goes like this: let E0 denote [0, 1],letE1 denote what is left after deleting an open interval from the center of E0.IfEn−1 has been constructed, obtain En by deleting open intervals of the same length from the center of each of the disjoint closed inter- vals that comprise En−1.
    [Show full text]
  • 1. Definition. Let I Be the Unit Interval. If I, G Are Maps on I Then 1 and G Are Conjugate Iff There Is a Homeomorphism H of I Such That 1 = H -I Gh
    TRANSACTIONS OF THE AMERICAN MATHEMATICAL SOCIETY Volume 319. Number I. May 1990 A COMPLETE CLASSIFICATION OF THE PIECEWISE MONOTONE FUNCTIONS ON THE INTERVAL STEWART BALDWIN ABSTRACT. We define two functions f and g on the unit interval [0, I] to be strongly conjugate iff there is an order-preserving homeomorphism h of [0, I] such that g = h-1 fh (a minor variation of the more common term "conjugate", in which h need not be order-preserving). We provide a complete set of invariants for each continuous (strictly) piecewise monotone function such that two such functions have the same invariants if and only if they are strongly conjugate, thus providing a complete classification of all such strong conjugacy classes. In addition, we provide a criterion which decides whether or not a potential invariant is actually realized by some 'piecewise monotone continuous function. INTRODUCTION 1. Definition. Let I be the unit interval. If I, g are maps on I then 1 and g are conjugate iff there is a homeomorphism h of I such that 1 = h -I gh. A more convenient term for our purposes will be: I, g are strongly conjugate (written 1 ~ g) iff there is an order preserving homeomorphism h as above. Given a map 1 on I, define the map j by I*(x) = 1 - 1(1 - x) (essentially a 1800 rotation of the graph). The main result of this paper will be to give a complete set of invariants for the strong conjugacy classes of the piecewise monotone continuous functions on the unit interval I = [0, I].
    [Show full text]
  • Classes of Graphs with Restricted Interval Models Andrzej Proskurowski, Jan Arne Telle
    Classes of graphs with restricted interval models Andrzej Proskurowski, Jan Arne Telle To cite this version: Andrzej Proskurowski, Jan Arne Telle. Classes of graphs with restricted interval models. Discrete Mathematics and Theoretical Computer Science, DMTCS, 1999, 3 (4), pp.167-176. hal-00958935 HAL Id: hal-00958935 https://hal.inria.fr/hal-00958935 Submitted on 13 Mar 2014 HAL is a multi-disciplinary open access L’archive ouverte pluridisciplinaire HAL, est archive for the deposit and dissemination of sci- destinée au dépôt et à la diffusion de documents entific research documents, whether they are pub- scientifiques de niveau recherche, publiés ou non, lished or not. The documents may come from émanant des établissements d’enseignement et de teaching and research institutions in France or recherche français ou étrangers, des laboratoires abroad, or from public or private research centers. publics ou privés. Discrete Mathematics and Theoretical Computer Science 3, 1999, 167–176 Classes of graphs with restricted interval models ✁ Andrzej Proskurowski and Jan Arne Telle ✂ University of Oregon, Eugene, Oregon, USA ✄ University of Bergen, Bergen, Norway received June 30, 1997, revised Feb 18, 1999, accepted Apr 20, 1999. We introduce ☎ -proper interval graphs as interval graphs with interval models in which no interval is properly con- tained in more than ☎ other intervals, and also provide a forbidden induced subgraph characterization of this class of ✆✞✝✠✟ graphs. We initiate a graph-theoretic study of subgraphs of ☎ -proper interval graphs with maximum clique size and give an equivalent characterization of these graphs by restricted path-decomposition. By allowing the parameter ✆ ✆ ☎ to vary from 0 to , we obtain a nested hierarchy of graph families, from graphs of bandwidth at most to graphs of pathwidth at most ✆ .
    [Show full text]
  • On Computing Longest Paths in Small Graph Classes
    On Computing Longest Paths in Small Graph Classes Ryuhei Uehara∗ Yushi Uno† July 28, 2005 Abstract The longest path problem is to find a longest path in a given graph. While the graph classes in which the Hamiltonian path problem can be solved efficiently are widely investigated, few graph classes are known to be solved efficiently for the longest path problem. For a tree, a simple linear time algorithm for the longest path problem is known. We first generalize the algorithm, and show that the longest path problem can be solved efficiently for weighted trees, block graphs, and cacti. We next show that the longest path problem can be solved efficiently on some graph classes that have natural interval representations. Keywords: efficient algorithms, graph classes, longest path problem. 1 Introduction The Hamiltonian path problem is one of the most well known NP-hard problem, and there are numerous applications of the problems [17]. For such an intractable problem, there are two major approaches; approximation algorithms [20, 2, 35] and algorithms with parameterized complexity analyses [15]. In both approaches, we have to change the decision problem to the optimization problem. Therefore the longest path problem is one of the basic problems from the viewpoint of combinatorial optimization. From the practical point of view, it is also very natural approach to try to find a longest path in a given graph, even if it does not have a Hamiltonian path. However, finding a longest path seems to be more difficult than determining whether the given graph has a Hamiltonian path or not.
    [Show full text]
  • Complexity and Exact Algorithms for Vertex Multicut in Interval and Bounded Treewidth Graphs 1
    Complexity and Exact Algorithms for Vertex Multicut in Interval and Bounded Treewidth Graphs 1 Jiong Guo a,∗ Falk H¨uffner a Erhan Kenar b Rolf Niedermeier a Johannes Uhlmann a aInstitut f¨ur Informatik, Friedrich-Schiller-Universit¨at Jena, Ernst-Abbe-Platz 2, D-07743 Jena, Germany bWilhelm-Schickard-Institut f¨ur Informatik, Universit¨at T¨ubingen, Sand 13, D-72076 T¨ubingen, Germany Abstract Multicut is a fundamental network communication and connectivity problem. It is defined as: given an undirected graph and a collection of pairs of terminal vertices, find a minimum set of edges or vertices whose removal disconnects each pair. We mainly focus on the case of removing vertices, where we distinguish between allowing or disallowing the removal of terminal vertices. Complementing and refining previous results from the literature, we provide several NP-completeness and (fixed- parameter) tractability results for restricted classes of graphs such as trees, interval graphs, and graphs of bounded treewidth. Key words: Combinatorial optimization, Complexity theory, NP-completeness, Dynamic programming, Graph theory, Parameterized complexity 1 Introduction Motivation and previous results. Multicut in graphs is a fundamental network design problem. It models questions concerning the reliability and ∗ Corresponding author. Tel.: +49 3641 9 46325; fax: +49 3641 9 46002 E-mail address: [email protected]. 1 An extended abstract of this work appears in the proceedings of the 32nd Interna- tional Conference on Current Trends in Theory and Practice of Computer Science (SOFSEM 2006) [19]. Now, in particular the proof of Theorem 5 has been much simplified. Preprint submitted to Elsevier Science 2 February 2007 robustness of computer and communication networks.
    [Show full text]
  • Semisimplicial Unital Groups
    International Journal of Theoretical Physics, Vol. 43, Nos. 7/8, August 2004 (C 2004) Semisimplicial Unital Groups David J. Foulis1 and Richard J. Greechie2 If there is an order-preserving group isomorphism from a directed abelian group G with a finitely generated positive cone G+ onto a simplicial group, then G is called a semisimplicial group. By factoring out the torsion subgroup of a unital group having a finite unit interval, one obtains a semisimplicial unital group. We exhibit a represen- tation for the base-normed space associated with a semisimplicial unital group G as the Banach dual space of a finite dimensional order-unit space that contains G as an additive subgroup. In terms of this representation, we formulate necessary and sufficient conditions for G to be archimedean. KEY WORDS: partially ordered abelian group; directed group; unital group; sim- plicial group; semisimplicial unital group; effect algebra; state; base-normed space, order-unit space. 1. INTRODUCTION If H is a Hilbert space, then the set G of all bounded self-adjoint operators on H is organized in the usual way into a partially ordered real linear space (in fact, an order-unit Banach space), and the operators belonging to the interval E :={A ∈ G | 0 ≤ A ≤ 1} are called the effect operators on H. The set E, organized into a partial algebra under the restriction ⊕ of addition on G to E, is called the algebra of effect operators on H. In the contemporary theory of quantum measurement (Busch et al., 1991), an observable is represented by a positive-operator-valued (POV) measure defined on a σ-field of sets and taking on values in the effect algebra E.
    [Show full text]
  • Chapter 1: Interval Analysis, Fuzzy Set Theory and Possibility Theory In
    Draft 1.0 November 4, 2010 CHAPTER 1 of the book Lodwick, Weldon A. (2011). Flexible and Un- certainty Optimization (to be published) Interval Analysis, Fuzzy Set Theory and Possibility Theory in Optimization Weldon A. Lodwick University of Colorado Denver Department of Mathematical and Statistical Sciences, Campus Box 170 P.O. Box 173364 Denver, CO 80217-3364, U.S.A. Email: [email protected] Abstract: This study explores the interconnections between interval analysis, fuzzy interval analysis, and interval, fuzzy, and possibility optimization. The key ideas considered are: (1) Fuzzy and possibility optimization are impor- tant methods in applied optimization problems, (2) Fuzzy sets and possibility distributions are distinct which in the context of optimization lead to ‡exible optimization on one hand and optimization under uncertainty on the other, (3) There are two ways to view possibility analysis leading to two distinct types of possibility optimization under uncertainty, single and dual distribu- tion optimization, and (4) The semantic (meaning) of constraint set in the presence of fuzziness and possibility needs to be known a-priori in order to determine what the constraint set is and how to compute its elements. The thesis of this exposition is that optimization problems that intend to model real problems are very (most) often satis…cing and epistemic and that the math- ematical language of fuzzy sets and possibility theory are well-suited theories for stating and solving a wide classes of optimization problems. Keywords: Fuzzy optimization, possibility optimization, interval analysis, constraint fuzzy interval arithmetic 1 Introduction This exposition explores interval analysis, fuzzy interval analysis, fuzzy set theory, and possibility theory as they relate to optimization which is both a synthesis of the author’s previous work bringing together, transforming, 1 and updating many ideas found in [49],[50], [51], [52], [58], [63], [64], [118], [116], and [117] as well as some new material dealing with the computation of constraint sets.
    [Show full text]
  • WHY a DEDEKIND CUT DOES NOT PRODUCE IRRATIONAL NUMBERS And, Open Intervals Are Not Sets
    WHY A DEDEKIND CUT DOES NOT PRODUCE IRRATIONAL NUMBERS And, Open Intervals are not Sets Pravin K. Johri The theory of mathematics claims that the set of real numbers is uncountable while the set of rational numbers is countable. Almost all real numbers are supposed to be irrational but there are few examples of irrational numbers relative to the rational numbers. The reality does not match the theory. Real numbers satisfy the field axioms but the simple arithmetic in these axioms can only result in rational numbers. The Dedekind cut is one of the ways mathematics rationalizes the existence of irrational numbers. Excerpts from the Wikipedia page “Dedekind cut” A Dedekind cut is а method of construction of the real numbers. It is a partition of the rational numbers into two non-empty sets A and B, such that all elements of A are less than all elements of B, and A contains no greatest element. If B has a smallest element among the rationals, the cut corresponds to that rational. Otherwise, that cut defines a unique irrational number which, loosely speaking, fills the "gap" between A and B. The countable partitions of the rational numbers cannot result in uncountable irrational numbers. Moreover, a known irrational number, or any real number for that matter, defines a Dedekind cut but it is not possible to go in the other direction and create a Dedekind cut which then produces an unknown irrational number. 1 Irrational Numbers There is an endless sequence of finite natural numbers 1, 2, 3 … based on the Peano axiom that if n is a natural number then so is n+1.
    [Show full text]
  • Topological Properties
    CHAPTER 4 Topological properties 1. Connectedness • Definitions and examples • Basic properties • Connected components • Connected versus path connected, again 2. Compactness • Definition and first examples • Topological properties of compact spaces • Compactness of products, and compactness in Rn • Compactness and continuous functions • Embeddings of compact manifolds • Sequential compactness • More about the metric case 3. Local compactness and the one-point compactification • Local compactness • The one-point compactification 4. More exercises 63 64 4. TOPOLOGICAL PROPERTIES 1. Connectedness 1.1. Definitions and examples. Definition 4.1. We say that a topological space (X, T ) is connected if X cannot be written as the union of two disjoint non-empty opens U, V ⊂ X. We say that a topological space (X, T ) is path connected if for any x, y ∈ X, there exists a path γ connecting x and y, i.e. a continuous map γ : [0, 1] → X such that γ(0) = x, γ(1) = y. Given (X, T ), we say that a subset A ⊂ X is connected (or path connected) if A, together with the induced topology, is connected (path connected). As we shall soon see, path connectedness implies connectedness. This is good news since, unlike connectedness, path connectedness can be checked more directly (see the examples below). Example 4.2. (1) X = {0, 1} with the discrete topology is not connected. Indeed, U = {0}, V = {1} are disjoint non-empty opens (in X) whose union is X. (2) Similarly, X = [0, 1) ∪ [2, 3] is not connected (take U = [0, 1), V = [2, 3]). More generally, if X ⊂ R is connected, then X must be an interval.
    [Show full text]