On the Number of Minimal 1-Steiner Trees

Similar documents
Arrangements And Duality

Solving Geometric Problems with the Rotating Calipers *

Mathematics Geometry Unit 1 (SAMPLE)

Binary Space Partitions

A Note on Maximum Independent Sets in Rectangle Intersection Graphs

Solutions to Homework 10

Approximation of an Open Polygonal Curve with a Minimum Number of Circular Arcs and Biarcs

COMBINATORIAL PROPERTIES OF THE HIGMAN-SIMS GRAPH. 1. Introduction

ON FIBER DIAMETERS OF CONTINUOUS MAPS

INCIDENCE-BETWEENNESS GEOMETRY

Computational Geometry. Lecture 1: Introduction and Convex Hulls

Algebra Geometry Glossary. 90 angle

Fast approximation of the maximum area convex. subset for star-shaped polygons

Angle - a figure formed by two rays or two line segments with a common endpoint called the vertex of the angle; angles are measured in degrees

Shortest Inspection-Path. Queries in Simple Polygons

Triangulation by Ear Clipping

Max-Min Representation of Piecewise Linear Functions

Chapter 3.1 Angles. Geometry. Objectives: Define what an angle is. Define the parts of an angle.

Shortcut sets for plane Euclidean networks (Extended abstract) 1

BALTIC OLYMPIAD IN INFORMATICS Stockholm, April 18-22, 2009 Page 1 of?? ENG rectangle. Rectangle

INDISTINGUISHABILITY OF ABSOLUTELY CONTINUOUS AND SINGULAR DISTRIBUTIONS

Strictly Localized Construction of Planar Bounded-Degree Spanners of Unit Disk Graphs

1. A student followed the given steps below to complete a construction. Which type of construction is best represented by the steps given above?

MCS 563 Spring 2014 Analytic Symbolic Computation Wednesday 9 April. Hilbert Polynomials

Actually Doing It! 6. Prove that the regular unit cube (say 1cm=unit) of sufficiently high dimension can fit inside it the whole city of New York.

The minimum number of distinct areas of triangles determined by a set of n points in the plane

Catalan Numbers. Thomas A. Dowling, Department of Mathematics, Ohio State Uni- versity.

High degree graphs contain large-star factors

GEOMETRY. Constructions OBJECTIVE #: G.CO.12

Disjoint Compatible Geometric Matchings

Small Maximal Independent Sets and Faster Exact Graph Coloring

1 if 1 x 0 1 if 0 x 1

The number of generalized balanced lines

How To Find The Ranking Of A Cormorant

arxiv: v1 [math.co] 7 Mar 2012

HOLES 5.1. INTRODUCTION

Binary Space Partitions of Orthogonal Subdivisions

3. Linear Programming and Polyhedral Combinatorics

Geometer s Sketchpad. Discovering the incenter of a triangle

Removing Even Crossings

1 Solution of Homework

COUNTING INDEPENDENT SETS IN SOME CLASSES OF (ALMOST) REGULAR GRAPHS

Geometry 1. Unit 3: Perpendicular and Parallel Lines

Geometry Chapter Point (pt) 1.1 Coplanar (1.1) 1.1 Space (1.1) 1.2 Line Segment (seg) 1.2 Measure of a Segment

Intersection of a Line and a Convex. Hull of Points Cloud

Duality of linear conic problems

A Short Proof that Compact 2-Manifolds Can Be Triangulated

Angles that are between parallel lines, but on opposite sides of a transversal.

SolidWorks Implementation Guides. Sketching Concepts

Closest Pair Problem

Diffuse Reflection Radius in a Simple Polygon

Seminar. Path planning using Voronoi diagrams and B-Splines. Stefano Martina

Volumetric Meshes for Real Time Medical Simulations

Outline 2.1 Graph Isomorphism 2.2 Automorphisms and Symmetry 2.3 Subgraphs, part 1

Elements of Plane Geometry by LK

GEOMETRY COMMON CORE STANDARDS

1. Prove that the empty set is a subset of every set.

DESARGUES THEOREM DONALD ROBERTSON

Line Segment Intersection

CS 598CSC: Combinatorial Optimization Lecture date: 2/4/2010

Optimal 3D Angular Resolution for Low-Degree Graphs

39 Symmetry of Plane Figures

x 2 f 2 e 1 e 4 e 3 e 2 q f 4 x 4 f 3 x 3

Special Situations in the Simplex Algorithm

Chapter 9. Systems of Linear Equations

Lecture 2: August 29. Linear Programming (part I)

Fiber Polytopes for the Projections between Cyclic Polytopes

Selected practice exam solutions (part 5, item 2) (MAT 360)

ISOMETRIES OF R n KEITH CONRAD

(Basic definitions and properties; Separation theorems; Characterizations) 1.1 Definition, examples, inner description, algebraic properties

MA 323 Geometric Modelling Course Notes: Day 02 Model Construction Problem

ME 111: Engineering Drawing

NP-Hardness Results Related to PPAD

Class One: Degree Sequences

Geometry and Measurement

1 The Line vs Point Test

4. How many integers between 2004 and 4002 are perfect squares?

Section 9-1. Basic Terms: Tangents, Arcs and Chords Homework Pages : 1-18

Approximated Distributed Minimum Vertex Cover Algorithms for Bounded Degree Graphs

No: Bilkent University. Monotonic Extension. Farhad Husseinov. Discussion Papers. Department of Economics

2. If C is the midpoint of AB and B is the midpoint of AE, can you say that the measure of AC is 1/4 the measure of AE?

28 Closest-Point Problems

MATH STUDENT BOOK. 8th Grade Unit 6

SHARP BOUNDS FOR THE SUM OF THE SQUARES OF THE DEGREES OF A GRAPH

GEOMETRIC MENSURATION

Situation: Proving Quadrilaterals in the Coordinate Plane

Handout #Ch7 San Skulrattanakulchai Gustavus Adolphus College Dec 6, Chapter 7: Digraphs

Geometry Review Flash Cards

Euclidean Minimum Spanning Trees Based on Well Separated Pair Decompositions Chaojun Li. Advised by: Dave Mount. May 22, 2014

arxiv: v5 [math.co] 26 Dec 2014

Smooth functions statistics

The Essentials of CAGD

Angles and Quadrants. Angle Relationships and Degree Measurement. Chapter 7: Trigonometry

Geometry: Unit 1 Vocabulary TERM DEFINITION GEOMETRIC FIGURE. Cannot be defined by using other figures.

POLYNOMIAL RINGS AND UNIQUE FACTORIZATION DOMAINS

The positive minimum degree game on sparse graphs

CSC2420 Fall 2012: Algorithm Design, Analysis and Theory

Largest Fixed-Aspect, Axis-Aligned Rectangle

Approximation Algorithms

The Graphical Method: An Example

Transcription:

On the Number of Minimal 1-Steiner Trees Boris Aronov Marshall Bern David Eppstein Abstract We count the number of nonisomorphic geometric minimum spanning trees formed by adding a single point to an n-point set in d-dimensional space, by relating it to a family of convex decompositions of space. The O(n d log 2d2 d n) bound that we obtain significantly improves previously known bounds and is tight to within a polylogarithmic factor. The minimum Steiner tree of a point set is the tree of minimum total length, spanning the input points and possibly having some additional vertices. Georgakopoulos and Papadimitriou [1] define the minimum 1-Steiner tree as the minimum spanning tree whose set of vertices consists of the input points and exactly one extra point. They describe an algorithm for constructing this tree in the plane by enumerating all combinatorially distinct minimal 1-Steiner trees, that is, trees formed by fixing the position of the extra vertex and considering the minimum spanning tree of the resulting set of points. They give an O(n 2 ) bound on the number of such trees and construct a matching lower bound. Independently, Monma and Suri [3] proved the same bounds in the plane and gave bounds of O(n 2d ) and Ω(n d ) in any dimension d. We significantly improve Monma and Suri s upper bounds. We show that the maximum possible number of combinatorially distinct minimal 1-Steiner trees on n points in d-space is O(n d log 2d2 d n) for all dimensions d. Our bounds require as do the bounds in previous papers either an assumption of general position or a tie-breaking rule, so that the minimum spanning tree is unique; otherwise even in the plane there can be exponentially many distinct topologies. We begin by stating a theorem of Kalai [2]: Theorem 1. In a family of n convex sets in d-space with the property that no c +1of the sets share a point, the number of k-tuples of sets that do have a point in common is bounded by d ( c d )( n+d c ) i=0 k i i, whenever 0 k c. Here ( c d k i) =0ifc d<k i. So in particular, putting k = c we have: Corollary 1. Given a family of convex decompositions of d-space with a total of n cells, the number of cells in the decomposition formed by overlaying them is less than ( n d). A slightly weaker bound can be proved by an elementary argument that we now sketch. Assume that the convex decompositions are in general position with respect to each other; Computer Science Dept., Polytechnic University, Six MetroTech Center, Brooklyn, NY 11201. Xerox Palo Alto Research Center, 3333 Coyote Hill Rd., Palo Alto, CA 94304. Dept. of Information and Computer Science, University of California, Irvine, CA 92717; research performed in part while visiting Xerox PARC. 1

otherwise, a small perturbation only increases the number of cells. Assume also that no edges are perpendicular to a vertical axis. First we count the number of cells in the overlay that are bounded below. The lowest point of a cell is the lowest point in the intersection of some collection of i cells, i d, each from a different convex decomposition. Adding an infinitely low hyperplane, that is, lying below all finite cells, allows us to count all other cells in the same way. Altogether, we can conclude that there are at most d i=0 ( n+1 i ) cells in the overlay. Let S(P ) be the number of distinct labeled minimum spanning tree topologies that can arise from the addition of one point to a set P of n points in d-space. Let S d (n) = max P =n S(P ). We now use the corollary above to bound S d (n). Let T be the minimum spanning tree of point set P, P = n. When new point x is added to P, the new minimum spanning tree T = T (x) has certain additional edges adjacent to x. The tree might also lack certain edges from T ; missing edges can be identified by adding new edges to T one by one, and removing the longest edge on each cycle thus created. The removed edges are completely determined by the set of the added edges. Below we construct a decomposition of space so that in each cell C of the decomposition the added edges come from a known set X(C) of size O(1). Once such a set is fixed, we build the 1-Steiner tree for a generic point x C from T by repeatedly adding an edge e X(C) to the current tree and deleting the longest edge in the cycle thus formed. Since there are O(1) edges in the candidate list, it is easily verified that only O(1) comparisons whose outcome depends on the exact position of x in C are needed and O(1) distinct outcomes are possible. In particular, cell C in our decomposition gives rise to only O(1) combinatorial types for T. Our overall strategy is as follows: (1) Construct a decomposition of space, and associate with each cell in the decomposition a small set of input points. The associated set of points is such that, for any point x in the cell, the set contains that point s neighbors N(x) in tree T. (2) Refine this partition so that for points x within a cell a constant-size superset of N(x) is fixed. (3) Use the above observation to deduce that each cell in the refined partition corresponds to O(1) possible combinatorial types of T. Our construction depends on the next lemma, which gives a necessary condition for a point to belong to neighbor set N(x). Define a frame to be a triangulation of the unit (d 1)-sphere. In the plane, a frame is a partition of the unit circle into circular arcs, and in three dimensions, it is a partition of a sphere into triangles with arcs of great circles for sides. Given a point x and a frame F d, we divide space into cones with common apex x and simplicial cross sections. The cones correspond to the simplices ( triangles ) of the frame: point y is in the cone corresponding to simplex s exactly when the unit vector parallel to xy is contained in s. Each cone is formed by intersecting d halfspaces that meet at x. Lemma 1 (Yao [5]). For any dimension d there exists a frame F d, with κ d simplices, having the following property. Let (x, y) be an edge in the minimum spanning tree of a point set. Point y is contained in some F d -cone with apex x; let K be that cone. Then y must be the closest input point to x in K. We now describe a method for constructing the initial subdivision of space, in which each cell corresponds to a small number of minimum spanning tree topologies. We start by building a family of convex decompositions and then apply Corollary 1. If x is newly added to the 2

Figure 1. (a) Cone-Voronoi diagram for the simplex (arc) [0,π/4]; (b) Boxes for the same simplex. point set, there can be O(1) new edges in T incident to x, namely, the edges (x, y) where y is the closest point to x in each of the O(1) cones determined by frame F d. We construct κ d cone-voronoi diagrams. Point x is in cell C y of diagram i if y is the nearest input point to x among the points in the cone with apex x determined by the ith simplex. One such diagram is depicted in Figure 1(a). Superimposing the κ d diagrams gives an arrangement in which, for each cell, the κ d possible edges specified by Yao s lemma are all determined. In the plane, this construction can be used to prove that O(n 2 ) topologies are possible. In higher dimensions, each cone-voronoi diagram has O(n 2 ) facets, giving an O(n 2d ) bound [3]. We cannot improve this directly because the cells in each diagram are not convex. Instead, we construct a more complicated structure to which Corollary 1 can be applied directly. Fix a single simplex of frame F d. A cone corresponding to the simplex is defined by d hyperplanes. We pass d hyperplanes through each of the n input points, one parallel to each of the defining hyperplanes of the simplex. This divides space into (n +1) d regions, in the form of parallelotopes (or boxes). Figure 1(b) depicts this subdivision in the plane. For any point x added as the new point in a 1-Steiner tree, the set of input points in the cone with apex x is determined only by the box containing x, and not by the exact location of x. Number each parallel set of hyperplanes from 1 to n, in order of the hyperplanes intersections with a line passing through them, with numbers increasing in the direction in which a cone corresponding to the simplex is unbounded (left-to-right, bottom-to-top in Figure 1(b)). Associate numbers 0 and n + 1 with hyperplanes at infinity on either side of the n hyperplanes. Each box is now determined by 2d indices, in d pairs corresponding to the d hyperplane directions. If each pair consists of two adjacent hyperplanes, then the box is a smallest or atomic box; nonadjacent pairs give larger boxes that are unions of atomic boxes. We say a box is regular if each pair is of the form (i2 j, (i + 1)2 j ) for some i and j. Regular boxes can be grouped by size, e.g., 1 1, 2 1, 1 2, etc. If we assume, without loss of generality, that n+1 is a power of 2, each generic point is contained in the same number of boxes, and is contained in c d (n) =O(log d n) regular boxes, one of each size. For each possible additional point x, we can find a nonoverlapping set of O(log d n) regular boxes, each of which is contained in the cone with apex x and and which together contain all the input points in the cone. (This construction generalizes the familiar one that realizes a ray as a union of O(log n) basic segments in a segment tree [4].) We say box b dominates box b if each lower index for b is at least as large as the corresponding higher index for b. For example, a box defined by pairs (4, 6) and (8, 16) dominates a box with pairs (0, 4) and (4, 5). Let the domination cone of box b be the union of all boxes dominated by b. 3

We construct the Euclidean Voronoi diagram for the points in each regular box and truncate the diagram to lie within the box s domination cone. We turn the truncated Voronoi diagram into a convex decomposition of space by cutting the complement of the domination cone into O(d) convex pieces (for example, by extending one by one the hyperplanes bounding the domination cone). Each input point is included in (on the boundary or in the interior of) 2 d c d (n) =O(log d n) regular boxes, so there are O(n log d n) cells in all diagrams. We repeat this construction for each one of the κ d = O(1) simplices of F d and thus obtain a family of convex decompositions of space with a total of O(n log d n) cells. We now classify each (generic) point of space according to the set of cells containing it (that is, its cell in the overlay). The number of such classes by Corollary 1 is (O(n log d n)) d = O(n d log d2 n). For each cell C of the overlay, we can now fix a set of O(log d n) candidate input points with the following property: for any point x C and any simplex of F d, one of the candidates is the closest input point to x in the cone with apex x corresponding to that simplex. To form the set of candidates for cell C, we consider each simplex of F d in turn, and find the nonoverlapping set of O(log d n) regular boxes that exactly cover the input points in a cone with apex in C. (Since C is contained in one atomic box, the choice of apex does not matter.) We then add to the candidate set exactly those input points (at most one per box) whose Voronoi cells contain C. Thus we have constructed the initial decomposition of space, referred to in step (1) above. Actually, for any given simplex of F d, not all of the O(log d n) candidates are necessary. If one regular box lies in the domination cone of another regular box, then the candidate in the second box is further from x than the candidate in the first box. (A sphere centered at x through the furthest corner of the first box does not intersect the second box.) This observation reduces the size of the set of candidates for a given simplex to O(log d 1 n), because each nonempty maximally dominated box must have some lower index that is at least as small as the corresponding index for any other nonempty maximally dominated box. Constructing the Euclidean Voronoi diagram of the O(log d 1 n) candidate points partitions cell C according to the identity of the closest point. Overlaying the κ d diagrams corresponding to different simplices produces a partition of C in each cell of which the closest point in any cone is fixed; this is the refinement of step (2). The number of resulting sub-cells is (O(log d 1 n)) d = O(log d2 d ) by another application of Corollary 1, giving a total of O(n d log 2d2 d n) cells overall. For each of the resulting cells C, the neighbors of x in T come from a fixed set of cardinality κ d = O(1). By previous discussion, C must give rise to O(1) distinct combinatorial types of the minimal 1-Steiner tree. Thus we have shown: Theorem 2. S d (n) is O(n d log 2d2 d n). We mention again that this bound, apart from the polylogarithmic factor, is tight. Clearly, there is some room for improvement. In fact, it is conceivable that S d (n) isθ(n d ). In conclusion, we thank Janos Pach for providing a crucial pointer. 4

References [1] G. Georgakopoulos and C. Papadimitriou. The 1-Steiner tree problem. J. Algorithms 8 (1987) 122 130. [2] G. Kalai. Intersection patterns of convex sets. Israel J. Math. 48 (1984) 161 174. [3] C. Monma and S. Suri. Transitions in geometric spanning trees. 7th ACM Symp. Computational Geometry (1991) 239 249. [4] F.P. Preparata and M.I. Shamos. Computational Geometry: An Introduction. Springer- Verlag, 1985. [5] A.C. Yao. On constructing minimum spanning trees in k-dimensional space and related problems. SIAM J. Comput. 11 (1982) 721 736. 5