Perfect Matchings: A Theory of Matching Covered Graphs (Algorithms and Computation in Mathematics, 31) [2024 ed.] 3031475038, 9783031475030

Beginning with its origins in the pioneering work of W.T. Tutte in 1947, this monograph systematically traces through so

119 56 8MB

English Pages 603 [584] Year 2024

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
Foreword
Preface
Contents
Terminology and Notation
Part I Basic Theory
Chapter 1 Perfect Matchings
Contents
1.1 Matchings and Perfect Matchings
Exercises
1.2 The Advent of Tutte
Exercises
1.3 Theorems of Tutte and Petersen
1.3.1 Matchable graphs
Complexity of Algorithms
1.3.2 Cubic matchable graphs
Exercises
1.4 Hall’s Theorem
Exercises
1.5 The Magic of Pfaffians
1.5.1 Pfaffians
1.5.2 The matrix of indeterminates
1.5.3 Tutte’s original proof of his theorem
Exercises
1.6 Notes
Chapter 2 Matching Covered Graphs
Contents
2.1 Matchable Edges in Matchable Graphs
Exercises
2.2 Matching Covered Graphs
2.2.1 Connectivity of matching covered graphs
2.2.2 Cubic matching covered graphs
2.2.3 Three special matching covered graphs
Exercises
2.3 Bipartite Matching Covered Graphs
2.3.1 Dulmage-Mendelsohn decomposition
Exercises
2.4 The Operation of Splicing
Exercises
2.5 Families of Matching Covered Graphs
Exercises
2.6 Notes
Chapter 3 Canonical Partitions
Contents
3.1 Barriers in Matchable Graphs
3.1.1 Properties of barriers of matchable graphs
3.1.2 Maximal barriers in matchable graphs
3.1.3 Maximal barriers in bipartite matching covered graphs
Exercises
3.2 Barriers and Cores
Exercises
3.3 Canonical Partitions
3.3.1 Bicritical graphs
3.3.2 Bisubdivisions
Exercises
3.4 Ear Decompositions of Bipartite Graphs
Exercises
3.5 Notes
Chapter 4 Tight Cuts
Contents
4.1 Separating Cuts
Exercises
4.2 Tight Cuts
4.2.1 Barrier cuts
Special barrier cuts
4.2.2 2-Separation cuts
4.2.3 Tight 3-edge cuts#
Tight cuts may be neither barrier nor 2-separation cuts
4.2.4 Separating cuts with bipartite shores
Exercises
4.3 Tight Cut Decompositions
4.3.1 Bricks and braces
4.3.2 Uncrossing cuts
Uncrossing separating cuts
Exercises
4.4 Uniqueness of Tight Cut Decompositions
4.4.1 The number of bricks
4.4.2 Near-bricks
Exercises
4.5 Subadditivity of the Number of Bricks
Exercises
4.6 Notes
Chapter 5 Characterizations of Bricks and Braces
Contents
5.1 Characterization of Bricks
5.2 Proof of the ELP Theorem##
5.2.1 Peripheral tight cuts
5.2.2 Dulmage-Mendelsohn barriers
5.2.3 A key lemma
5.2.4 The ELP Theorem
Exercises
5.3 Some Applications of the ELP Theorem
Tight cuts in cubic graphs
Tight cuts in bicritical graphs##
Exercises
5.4 Characterizations of Braces
Exercises
5.5 Tight Cuts in Graphs with Two Bricks
Barrier cuts and essentially 2-separation cuts
5.5.1 The laminar ELP Theorem
5.5.2 ELP tight cut decompositions
Exercises
5.6 Notes
Chapter 6 The Perfect Matching Polytope
Contents
6.1 Edmonds’ Theorem
6.1.1 Convex polytopes
6.1.2 The perfect matching polytope of a graph
Exercises
6.2 The Matching Space
6.2.1 Linear equality description
Exercises
6.3 The Dimension Formula
6.3.1 The merger operation
6.3.2 Matching lattice and the seminal work of Lovász
Exercises
6.4 Matching Spaces over Finite Fields
Exercises
6.5 Notes
Chapter 7 Solid Bricks
Contents
7.1 Solid Matching Covered Graphs
7.1.1 Odd-intercyclic bricks
7.1.2 Separating cut decompositions
Exercises
7.2 The Perfect Matching Polytope and Solid Matching Covered Graphs
7.2.1 Polytopes Poly (G) and Poly* (G )
7.2.2 Precedence relation on cuts
Exercises
7.3 A Characterization of Solid Bricks#
7.3.1 An infinite family of solid bricks that are not odd-intercyclic
Exercises
7.4 Notes
Chapter 8 Dependence Relation and Removable Classes
Contents
8.1 Deletions of Edges
8.1.1 Removable edges in bipartite graphs
8.1.2 Near-bipartite graphs
8.1.3 Removable sets and separating cuts
Exercises
8.2 The Dependence Relation
8.2.1 Dependence classes
Exercises
8.3 Removable Classes
8.3.1 Minimal classes
8.3.2 Sizes of removable classes
Exercises
8.4 Two Lemmas for Future Use##
8.4.1 The Zigzag Lemma
Exercises
8.5 Notes
Chapter 9 Dependence Classes in Bricks
Contents
9.1 Dependence Classes in Bricks
9.1.1 Monotonicity of functions b and b + p
9.1.2 Bricks with more than two removable doubletons
Exercises
9.2 Mutually Exclusive Classes
9.2.1 Removable edges in bricks
Exercises
9.3 Removable Edges in Near-Bipartite Bricks##
9.3.1 Nonremovable edges in near-bipartite bricks
Exercises
9.4 Notes
Chapter 10 Removable Classes in Solid Bricks
Contents
10.1 Separating Cuts and Dulmage-Mendelsohn Barriers
10.1.1 Nonremovable edges in bicritical graphs
10.1.2 Removable edges in solid bricks
10.1.3 The Lemma on wheels
Exercises
10.2 Some Implications of Being Solid
10.2.1 Removable edges in solid bricks and separating cuts
10.2.2 Propagation
10.2.3 b-Invariance (monotonicity revisited)
10.2.4 The exchange property
Exercises
10.3 Notes
Chapter 11 Ear Decompositions
Contents
11.1 Additions of Ears
11.1.1 Single and double ears
11.1.2 Conformal subgraphs
Exercises
11.2 Deletions of Ears
11.2.1 Removable ears
11.2.2 Series reductions of ears
11.2.3 The two ear theorem
Exercises
11.3 Ear Decompositions
11.3.1 Two equivalent definitions
11.3.2 An ear decomposition of the Petersen graph
Exercises
11.4 The Number of Ears in a Decomposition
11.4.1 Ear decompositions and perfect matchings
Exercises
11.5 Retract of a Graph
Exercises
11.6 Ear Decompositions of Critical Graphs#
Exercises
11.7 Notes
Chapter 12 Matching Minors
Contents
12.1 Matching Minors
12.1.1 Matching-minor-closed classes of graphs
12.1.2 Matching minors and tight cuts
Exercises
12.2 Conformal Minors
12.2.1 Unavoidable conformal minors
Exercises
12.3 Graphs With Specified Conformal Minors
12.3.1 4-free planar bricks
12.3.2 C6-free planar bricks
12.3.3 A special property of the Petersen graph
Exercises
12.4 Notes
Part II Brick and Brace Generation
Chapter 13 A Conjecture of Lov´asz Concerning Bricks
Contents
13.1 b-Invariant Edges in Bricks
13.1.1 b-Invariant ears in matching covered graphs
13.1.2 Our approach to Conjecture 13.2
Exercises
13.2 (b + p)-Invariant Edges in Bricks
13.2.1 The three case lemma
13.2.2 The Petersen graph
Exercises
13.3 Notes
Chapter 14 Robust Cuts in Bricks
Contents
14.1 Robust Cuts in Nonsolid Bricks#
14.1.1 Peripheral robust cuts
Exercises
14.2 Robust Cuts with a Shore of Minimum Size
Exercises
14.3 Quasi b-Invariant Edges
14.3.1 Barrier cut pairs and essentially 2-separation cut pairs
Exercises
14.4 Notes
Chapter 15 b-Invariant Edges in Bricks
Contents
15.1 Existence of b-Invariant Edges
15.2 Proof of Theorem 15.1#
15.2.1 Reduction to simple nonsolid bricks
15.2.2 Definitions of X, C, M0, G1 and G2
15.2.3 Reduction to the case in which G1 is an odd wheel
15.2.4 Reduction to the case in which G2 is a brick
15.2.5 Reduction to the case in which G2 is not a special brick
15.2.6 Choosing a b-invariant edge of e of G2
15.2.7 Reduction to the case in e which is a quasi -invariant edge of G
15.2.8 Reduction to case in which {C − e, D− e} is an essentially 2-separation cut pair of tight cuts of G − e
15.2.9 Working with the hypothesis that {C − e, D− e} is an essentially 2-separation cut pair of tight cuts of G − e
15.2.10 Reduction to the case in which G[X] and G[Y] are pentagons
15.2.11 The conclusion: G is the Petersen graph
Exercises
15.3 Notes
Chapter 16 The Matching Lattice and Optimal Ear Decompositions
Contents
16.1 The Matching Lattice and its Bases
16.1.1 Lattices, generating sets, and bases
16.1.2 Reduction to bricks and braces
16.1.3 The question of existence of bases
16.1.4 The matching lattice of the Petersen graph
Exercises
16.2 Optimal Ear Decompositions
16.2.1 The case of bipartite graphs and near-bricks
16.2.2 The difficulty in dealing with graphs which have more than one brick
16.2.3 A strengthening of Theorem 15.1
16.2.4 Exchange property of optimal removable classes
Exercises
16.3 Matching Integral Vectors
Exercises
16.4 Matching Spaces over Z2

16.4.1 Matching orthogonal vectors
Exercises
16.5 Notes
Chapter 17 Thin Edges in Bricks and Braces
Contents
17.1 Thin Edges and Thin Barriers in Bricks
17.1.1 Outline of the proof of the existence of thin edges in bricks
Exercises
17.2 Barriers and Tight Cuts in Near-Bricks#
17.2.1 Near-bricks inherited from bricks
Three Case Lemma–Revisited
Rank Computations
17.2.2 Removable edges in near-bricks inherited from bricks
Exercises
17.3 The Existence of Thin Edges in Bricks#
17.4 Building Bricks
17.4.1 The four expansion operations
Exercises
17.5 Thin Edges in Braces
17.5.1 Graphs obtained by the deletion of an edge from a brace
Tight cut decompositions of G − e
Removable edges in G − e
Maximal shores of tight cuts of G − e
Exercises
17.6 The Existence of Thin Edges in Braces#
The rank of an edge
Exercises
17.7 Building Braces
Exercises
17.8 Notes
Chapter 18 Strictly Thin Edges in Bricks and Braces
Contents
18.1 Strictly Thin Edges
18.1.1 Families of graphs without strictly thin edges
Exercises
18.2 Multiple Edges in Retracts
18.2.1 How do multiple edges in the retract of G − e arise?#
18.2.2 What is the status of those edges in G itself?
18.2.3 Two key lemmas
Exercises
18.3 Characterization of G ♯
18.3.1 Graphs G with index (G ) = 3
18.3.2 Graphs with index (G ) = 1
18.3.3 Graphs with index (G ) = 2
18.3.4 The characterization of G ♯
Exercises
18.4 Generating Simple Braces and Bricks
Exercises
18.5 Splitter Versions
18.6 Notes
Part III Pfaffian Orientations
Chapter 19 Pfaffian Orientations
Contents
19.1 The Pfaffian Orientation Problem
Exercises
19.2 Equivalent Formulations
19.2.1 Oddly oriented cycles and conformal subgraphs
19.2.2 An application: the even directed cycle problem
19.2.3 Non-Pfaffian graphs
Exercises
19.3 Pfaffian Orientations of Planar Graphs
19.3.1 Odd orientations of plane graphs
Exercises
19.4 Notes
Chapter 20 Similarity of Orientations and Characteristic Orientations
Contents
20.1 Similarity of Orientations
Determining whether a set of edges of a matching covered graph is a cut
20.1.1 Similarity and Pfaffian orientations
20.1.2 Similarity and removable ears
Exercises
20.2 Pfaffian Orientations and Tight Cuts
20.2.1 Pfaffian orientations and separating cuts
Exercises
20.3 The Number of Similarity Classes#
20.3.1 An upper bound for τ(G)
20.3.2 A lower bound for τ(G)
Exercises
20.4 Characteristic Orientations
20.4.1 Pfaffian extensions
20.4.2 Finding characteristic orientations
20.4.3 The equivalence of Problems 19.1 and 19.2
Exercises
20.5 Notes
Chapter 21 Excluded-Minor Characterizations of Pfaffian Graphs
Contents
21.1 S-Minors
Exercises
21.2 -Minimal non-Pfaffian Graphs
21.2.1 Properties of non-Pfaffian Graphs
21.2.2 Monochromatic chords of even cycles
Exercises
21.3 Pfaffian Bipartite graphs
Pfaffian solid bricks
21.4 Pfaffian Near-Bipartite Graphs#
21.4.1 Some immediate implications of the induction hypothesis
21.4.2 Removable edges of H
21.4.3 Degree of vertices
21.4.4 There is a unique ( M1, M2)-alternating cycle in G
21.4.5 Quadrilaterals in H
21.4.6 Triangles in G
21.4.7 The equality M1 ∩ M2 = R = {e, f}
21.4.8 The graphs Γ1 and Γ2
Exercises
21.5 Polynomial-time Algorithms
21.5.1 Efficient recognition of Pfaffian bipartite graphs
Description of the algorithm
21.5.2 Efficient recognition of Pfaffian near-bipartite graphs
Description of the algorithm
Exercises
21.6 An Infinite Family of Minimal Non-Pfaffian Graphs#
21.6.1 NT-minors
21.6.2 Non-Pfaffian NT-minimal graphs
21.6.3 The Norine and Thomas family NT
Exercises
21.7 Notes
Unsolved Problems
Some Classical Problems
Appendix A Solutions to Selected Exercises
Contents
A.1 Chapter 1
Exercise 1.2.2
Exercise 1.3.3
Exercise 1.3.5
Exercise 1.5.6
A.2 Chapter 2
Exercise 2.2.5
Exercise 2.2.6
Exercise 2.3.1
Exercise 2.3.2
Exercise 2.3.3
Exercise 2.4.1
Exercise 2.4.2
Exercise 2.4.6
Exercise 2.5.4
A.3 Chapter 3
Exercise 3.1.3
Exercise 3.1.4
Exercise 3.1.7
Exercise 3.2.2
Exercise 3.3.1
Exercise 3.3.5
Exercise 3.3.10
Exercise 3.4.4
A.4 Chapter 4
Exercise 4.2.2
Exercise 4.2.4
Exercise 4.3.8
Exercise 4.3.9
Exercise 4.3.11
Exercise 4.5.2
A.5 Chapter 5
Exercise 5.2.7
Exercise 5.2.8
Exercise 5.4.3
Exercise 5.4.4
Exercise 5.4.5
Exercise 5.4.6
A.6 Chapter 6
Exercise 6.1.2
Exercise 6.1.4
Exercise 6.1.7
Exercise 6.1.12
Exercise 6.2.1
Exercise 6.2.5
Exercise 6.2.6
Exercise 6.2.7
Exercise 6.3.3
A.7 Chapter 7
Exercise 7.2.3
Exercise 7.3.3
Exercise 7.3.5
A.8 Chapter 8
Exercise 8.1.7(iii)
Exercise 8.1.10
Exercise 8.3.8
Exercise 8.4.1
A.9 Chapter 9
Exercise 9.1.6
Exercise 9.2.3
Exercise 9.2.4
Exercise 9.2.5
A.10 Chapter 10
Exercise 10.1.6
Exercise 10.2.2
Exercise 10.2.3
Exercise 10.2.4
Exercise 10.2.5
Exercise 10.2.6
A.11 Chapter 11
Exercise 11.5.7
A.12 Chapter 12
Exercise 12.2.7
A.13 Chapter 13
Exercise 13.2.2
A.14 Chapter 14
Exercise 14.1.6
Exercise 14.3.1
A.15 Chapter 15
Exercise 15.2.6
A.16 Chapter 16
Exercise 16.2.6
Exercise 16.2.8
Algorithm A.2
Algorithm A.3
A.17 Chapter 17
Exercise 17.2.1
Exercise 17.2.2
Exercise 17.6.2
Exercise 17.6.4
Exercise 17.6.5
A.18 Chapter 18
Exercise 18.2.1
A.19 Chapter 19
Exercise 19.2.1
A.20 Chapter 20
Exercise 20.1.2
A.21 Chapter 21
Exercise 21.1.3
Exercise 21.5.1
Exercise 21.5.2
Exercise 21.5.3
Exercise 21.5.6
Exercise 21.6.4
References
List of Figures
Glossary
Index
Recommend Papers

Perfect Matchings: A Theory of Matching Covered Graphs (Algorithms and Computation in Mathematics, 31) [2024 ed.]
 3031475038, 9783031475030

  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

Algorithms and Computation in Mathematics  31

Cláudio L. Lucchesi U.S.R. Murty

Perfect Matchings

A Theory of Matching Covered Graphs

Algorithms and Computation in Mathematics Volume 31

Series Editors William J. Cook, Mathematics, University of Waterloo, Waterloo, ON, Canada David Eisenbud, Berkeley, CA, USA Bernhard Korte, Research Institute for Discrete Mathematics, University of Bonn, Bonn, Germany László Lovász, Számítógéptudományi Tanszék, Eötvös Loránd University, Budapest, Hungary Bernd Sturmfels, Berkeley, CA, USA Bianca Viray, Department of Mathematics, University of Washington, Seattle, WA, USA Avi Wigderson, School of Mathematics, Princeton University, Institute for Advanced Study, Princeton, NJ, USA Günter M. Ziegler , Institut für Mathematik, Freie Universität Berlin, Berlin, Germany

With this forward-thinking series Springer recognizes that the prevailing trend in mathematical research towards algorithmic and constructive processes is one of long-term importance. This series is intended to further the development of computational and algorithmic mathematics. In particular, Algorithms and Computation in Mathematics emphasizes the computational aspects of algebraic geometry, number theory, combinatorics, commutative, non-commutative and differential algebra, geometric and algebraic topology, group theory, optimization, dynamical systems and Lie theory. Proposals or manuscripts that center on content in non-computational aspects of one of these fields will also be regarded if the presentation gives consideration to the contents’ usefulness in algorithmic processes.

Cláudio L. Lucchesi • U. S. R. Murty

Perfect Matchings A Theory of Matching Covered Graphs

Cláudio L. Lucchesi University of Sao Paulo São Paulo, Brazil

U. S. R. Murty University of Waterloo Waterloo, ON, Canada

ISSN 1431-1550 Algorithms and Computation in Mathematics ISBN 978-3-031-47504-7 (eBook) ISBN 978-3-031-47503-0 https://doi.org/10.1007/978-3-031-47504-7 © The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now known or hereafter developed. The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore free for general use. The publisher, the authors, and the editors are safe to assume that the advice and information in this book are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the editors give a warranty, expressed or implied, with respect to the material contained herein or for any errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. This Springer imprint is published by the registered company Springer Nature Switzerland AG The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland Paper in this product is recyclable.

To the memory of our parents

Foreword

In recent decades the theory of matching covered graphs has blossomed. The concept was introduced by Tutte, who showed that every 2-connected cubic graph is matching covered, thereby strengthening a result of Petersen. The subject was studied in depth by Lov´asz, beginning in the 1980s. His contribution included the noteworthy theorem that any two tight cut decompositions of a given matching covered graph yield the same list of bricks and braces, up to multiple edges. The numbers of bricks and braces in this list are therefore important invariants of the original matching covered graph. Remarkably, it turns out that the number of bricks whose underlying simple graph is isomorphic to the Petersen graph is also significant. The authors’ interest in this subject was piqued by the early work of Lov´asz, and they began a close collaboration together that also included other authors. They developed an extensive theory of removable edges in matching covered graphs. These are edges whose removal preserves the matching covered property of the graph. The authors used this theory to unify the subject of matching covered graphs, simplify some proofs and make further significant advances. This volume is the result of that collaboration and presents a unified account of the most important results in this area. Its climax is a remarkable proof of a characterisation of Pfaffian near bipartite graphs. The authors present a subject of great beauty and depth. I now invite the reader to turn the page and begin to enjoy the treasures that await. Charles Little

vii

Preface

A graph is matchable if it has a perfect matching. A matching covered graph is a connected graph on at least two vertices in which each edge is covered by some perfect matching. The theory of matching covered graphs, though of relatively recent vintage, has an array of interesting results with elegant proofs, several surprising applications and challenging unsolved problems. Lovasz ´ and the matching lattice: It was the article Matching Structure and the

Matching Lattice by L´aszl´o Lov´asz (1987, [58]) that inspired us to learn the subject. The theorem which states that any two tight cut decompositions of a matching covered graph yield the same list of bricks and braces, up to multiple edges, is an example of an interesting and surprising assertion with an elegant proof. The crowning achievement of that paper was the characterization of the matching lattice. The matching lattice Lat(퐺) of a graph 퐺 is the set of all integer linear combinations of incidence vectors of perfect matchings of 퐺. A necessary condition for an integer vector x in the edge space of 퐺 to belong to Lat(퐺) is that it be regular, that is Í Í 푒∈휕(푣) x푒 , for any two vertices 푢 and 푣 of 퐺.1 But, in general, a 푒∈휕(푢) x푒 = regular integer vector in the edge space of a matching covered graph 퐺 need not belong to Lat(퐺). For example, the all 1’s vector 1 in the edge space of the Petersen graph is not in its matching lattice. Generalizing a theorem of Seymour (1979, [85]), Lov´asz proved the stunning statement that if 퐺 is any brick whose underlying simple graph is not the Petersen graph, then any regular integer vector in the edge space of 퐺 is in Lat(퐺). One cannot but be impressed by the sheer ‘audacity’ of a non-obvious assertion like this which says that ‘the Petersen graph is the only exception’! Tutte and the 1-factor theorem: The idea of a matching covered graph first ap-

peared in an article by William Thomas (Bill) Tutte who was one of the preeminent graph theorists and cryptographers of the twentieth century. In his high school days, Bill learnt basic graph theory from Rouse Ball’s book Mathematical Recreations and Essays and was familiar with Petersen’s Theorem, which states that every 2-connected cubic graph is matchable. In 1935 he went to Cambridge to study chemistry. There he made friends with Leonard Brooks, Cedric Smith and Arthur 1 For any subset 푆 of the vertex set of a graph 퐺, we denote the edge cut consisting of the set of edges of 퐺 with one end in 푆 and one end in 푆 by 휕(푆 ). ix

x

Preface

Stone, who were students of mathematics. The four of them worked on the problem of dissecting rectangles into squares no two of which are of the same size. They observed that there is a one-to-one correspondence between dissections of rectangles into squares and flows of currents in a certain associated electrical network. This observation made it possible for them to employ the determinantal formulas which are used for computing currents in an electrical network to find desired dissections of rectangles into squares. Their work was published in 1940 [4]. Bill Tutte switched to mathematics in 1940, but his studies were interrupted by call to wartime duty at Bletchley Park, the fabled organization of code breakers. We refer the reader to Younger (2012, [95]) for a biographical sketch of Bill Tutte and for an account of his pivotal contributions to code breaking when he was at Bletchley. After returning to Cambridge in 1945, Tutte resumed his research in graph theory. One of the problems he set himself to work on was to find a characterization of matchable graphs. He writes in his mathematical autobiography Graph Theory as I Have Known It (1998, [90]) that he was hoping to find a formula for the number of perfect matchings of a graph analogous to Kirchhoff’s determinantal formula for the number of spanning trees. He did not succeed in finding such a formula, but using properties of polynomials associated with matrices, called Pfaffians, he did succeed in finding a characterization of matchable graphs by showing that a graph 퐺 is matchable if and only if 표(퐺 − 푆) ≤ |푆|, for any subset 푆 of 푉, where 표(퐺 − 푆) denotes the number of odd components of 퐺 − 푆. (Tutte’s hope of using Pfaffians to count the number of perfect matchings in a graph was partially fulfilled by Kasteleyn (1963, [44]), who showed how they could be used to count the number of perfect matchings of a planar graph. This is described in Chapter 19.) Tutte’s fundamental result appeared in the paper The Factorization of Linear Graphs (1947, [89]). As a corollary, he established a strengthening of Petersen’s Theorem by showing that every 2-connected cubic graph is matching covered. Thus the notion of matching covered graphs goes back to this article by Tutte. Since Tutte referred to perfect matchings as 1-factors, some authors called such graphs 1-factor covered graphs. It appears that Lov´asz (1983, [57]) was the first to refer to them as matching covered graphs. Lov´asz has made many important contributions to combinatorics and computer science, but matching theory seems to have been his first love. His book Matching Theory with M. D. Plummer (1986, [59]) contains accounts of developments in the theory of matching covered graphs2 that took place up to the mid 1980’s. But that book was written before the appearance of his article Matching Structure and the Matching Lattice which established the uniqueness of tight cut decompositions of matching covered graphs, thereby providing an elegant theoretical foundation to this subject. 2 A matchable graph is elementary if the union of perfect matchings in it induces a connected subgraph. It should be noted that some results in [59] about matching covered graphs are presented in the more general context of elementary graphs.

Preface

xi

Tight Cut Decompositions and Bricks and Braces: Given any cut 퐶 := 휕 ( 푋) of a

graph 퐺 one can obtain two graphs 퐺/푋 and 퐺/ 푋 by contracting, respectively, the shores 푋 and 푋 of 퐶 to single vertices. These two graphs are called the 퐶-contractions of 퐺. A cut 퐶 of a matching covered graph 퐺 is said to be tight if |푀 ∩ 퐶| = 1 for every perfect matching 푀 of 퐺. It is easy to see that if 퐶 is a tight cut of a matching covered graph 퐺, then the two 퐶-contractions of 퐺 are also matching covered. A matching covered graph which is free of nontrivial tight cuts is a brick if it is nonbipartite and a brace if it is bipartite. If a graph 퐺 has a nontrivial tight cut 퐶 := 휕 ( 푋), then both 퐺/푋 and 퐺/푋 have fewer vertices than 퐺. Thus, by repeatedly contracting shores of nontrivial tight cuts, any matching covered graph may be reduced to a list of bricks and braces. Depending on the choice of tight cuts, this ‘tight cut decomposition procedure’ may yield different lists of bricks and braces. However, Lov´asz [58] showed that, up to multiple edges, any two applications of this procedure result in the same list of bricks and braces. This result allows us to define two invariants 푏(퐺) and 푝(퐺) associated with matching covered graphs, where 푏(퐺) is the total number of bricks in a tight cut decomposition of 퐺 and 푝(퐺) is the number of bricks in such a decomposition whose underlying simple graphs are isomorphic to the Petersen graph. Initial focus of our research: An edge 푒 of a matching covered graph 퐺 is removable if 퐺 − 푒 is also matching covered, and a removable edge 푒 is 푏-invariant if 푏(퐺 − 푒) = 푏(퐺). We started working systematically on the following conjecture: any brick 퐺, which is different from the complete graph 퐾4 , the triangular prism 퐶6 , and the Petersen graph P, has a 푏-invariant edge. (This conjecture was communicated to us by Lov´asz with the observation that its validity would simplify his proof of the theorem in [58] which established a characterization of the matching lattice.) As our work progressed, we realized that a proof of the above conjecture would also settle another (folklore!) conjecture which states that every brick whose underlying simple graph is not P has an ear decomposition which involves just one double ear addition. What led us to write this book?: We succeeded in confirming the above-stated

conjecture of Lov´asz in collaboration with Marcelo Carvalho, a former graduate student of Cl´audio Lucchesi. Our proofs appeared in the three papers [9], [10] and [11], which were published in 2002. (We were actually able to show that every brick other than 퐾4 , 퐶6 , P and a brick on eight vertices which we call the bicorn has at least two 푏-invariant edges and use this result to show that any matching covered graph 퐺 has an ear decomposition which involves just 푏(퐺) + 푝(퐺) double ear additions and that this result is best possible.) We continued to work on other problems in this area and have been able to make further contributions to the subject. Our efforts have led us to discover some interesting new facts and to develop some new methods of proof. Our object in writing this book has been to share with readers what we have learnt about this fascinating subject. We list below some of the many ideas which have enhanced our understanding of this subject.

xii

Preface

Barriers and matchable edges: A subset 퐵 of the vertex set of a matchable graph 퐺

is a barrier if |퐵| = 표(퐺 − 퐵). Theorem 2.1, which is a straightforward consequence of Tutte’s Theorem 1.3, shows that if 푢 and 푣 are any two vertices of a matchable graph 퐺, then 퐺 − 푢 − 푣 is also matchable if and only if there is no barrier of 퐺 which contains both 푢 and 푣. Graphs which do not have any nontrivial barriers are said to be bicritical graphs. As we shall see, all bricks are bicritical, but not all bicritical graphs are bricks. (The bicritical graph shown in Figure 3.5 is not a brick because it has nontrivial tight cuts.) An edge 푒 of a matchable graph 퐺 is matchable if there is some perfect matching of 퐺 which contains 푒. Theorem 2.1 stated above implies that 푒 is matchable if and only if there is no barrier of 퐺 which contains both ends of 푒. Thus a matching covered graph is a connected matchable graph in which each barrier is a stable set. A connected bipartite graph 퐺 := 퐺 [ 퐴, 퐵], with | 퐴| = |퐵| ≥ 2, is matching covered if and only if |푁 ( 푋)| ≥ | 푋 | + 1, for all 푋 ⊂ 퐴 such that 1 ≤ | 푋 | ≤ | 퐴| − 1. It is easy to prove this result, by adapting Hall’s well-known characterization of matchable bipartite graphs. Corollary 2.10, which follows from this characterization, states that in any matchable bipartite graph 퐺 := 퐺 [ 퐴, 퐵] there exists a subset 푆 of 퐴 such that the subgraph of 퐺 induced by 푆 ∪ 푁 (푆) is matching covered. This useful observation is the basis of what is known as the Dulmage-Mendelsohn decomposition of matchable bipartite graphs. (See Section 2.3.1.) Maximal barriers and the Canonical Partition Theorem: A barrier 퐵 of a matchable

graph 퐺 is maximal if there is no barrier of 퐺 which properly contains it. Theorem 3.2 states that if 퐵 is any maximal barrier of a matchable graph then each component of 퐺 − 퐵 is critical (factor-critical, hypomatchable); in particular, this means that no component of 퐺 − 퐵 is even. As noted earlier, all barriers in a matching covered graph are stable sets; in particular, all maximal barriers of such a graph are stable sets. One of the basic results in this theory, known as the canonical partition theorem (Lemma 5.2.1 in Lov´asz and Plummer [59]), states that the maximal barriers of a matching covered graph 퐺 constitute a partition of the vertex set of 퐺. In Chapter 3 we give a simple proof of this theorem as described below. We first consider the case in which 퐺 is a bipartite matching covered graph with bipartition {퐴, 퐵}. If 푆 is any maximal barrier of 퐺, then each component of 퐺 − 푆 has just one vertex because there are no critical bipartite graphs of order greater than one. Consequently 푆 := 푉 − 푆 is also a stable set, implying that {푆, 푆} is a bipartition of 퐺. However, being a matching covered graph, 퐺 is connected, and every connected bipartite graph has a unique bipartition. It follows that either 푆 = 퐴 or 푆 = 퐵. This also implies that if 푎 ∈ 퐴 and 푏 ∈ 퐵, then there is no barrier which includes both 푎 and 푏 and, furthermore, by Theorem 2.1, that 퐺 − 푎 − 푏 is matchable. To deal with the case in which 퐺 is not bipartite, we consider a maximal barrier 퐵 of 퐺 and associate with it the bipartite matching covered graph H(퐵) obtained from 퐺 by shrinking each component of 퐺 − 퐵 to a single vertex. It follows from the fact that each component of 퐺 − 퐵 is critical, and the above stated result concerning bipartite matching covered graphs, that if 푏 ∈ 퐵 and 푎 ∉ 퐵, then the graph 퐺 − 푎 − 푏 is matchable. Thus, if there is a maximal barrier 퐵′ ≠ 퐵 such that 푏 ∈ 퐵 ∩ 퐵′

Preface

xiii

and 푎 ∈ 퐵′ − 퐵, the graph 퐺 − 푎 − 푏 would be matchable. By Theorem 2.1 this is impossible because both 푎 and 푏 belong to the barrier 퐵′ . The core of a matchable graph with respect to a barrier: The bipartite H(퐵) as-

sociated with a maximal barrier 퐵 in a matching covered graph is a special case of a bipartite graph associated with any barrier 퐵 of a matchable graph. The core of a matchable graph 퐺 with respect to a barrier 퐵, denoted by H(퐵), is the bipartite graph obtained from 퐺 by deleting all edges with both ends in 퐵, all the vertices belonging to even components of 퐺 − 퐵, and then shrinking each odd component of 퐺 − 퐵 to a single vertex. If 퐺 is matching covered, then, clearly, H(퐵) is also matching covered. More generally, every matchable graph 퐺 has a barrier 퐵 such that H(퐵) is matching covered, with the additional property that each odd component of 퐺 − 퐵 is critical. We refer to such a barrier as a Dulmage-Mendelsohn barrier as it is possible to use Corollary 2.10 to show that any maximal barrier of a matchable graph includes such a barrier (see Lemma 5.4). This result plays a crucial role in our proof of an important result in this theory which is described below (and also in the proof of a useful result in Chapter 10). Edmonds-Lovasz-Pulleyblank ´ (ELP) Theorem: Recall that bricks and braces are,

respectively, nonbipartite and bipartite matching covered graphs which are free of nontrivial tight cuts. As mentioned earlier, Lov´asz [58] proved that any matching covered graph may be decomposed into bricks and braces by means of the tight cut decomposition procedure. But how does one recognize whether or not a given matching covered graph is free of nontrivial tight cuts? For example, how can we tell whether or not a given nonbipartite graph is a brick? ELP-cuts (barrier cuts and 2-separation cuts): To answer this question raised above, it would be convenient to know how tight cuts arise. There is no known simple answer to this question, but two scenarios which yield nontrivial tight cuts of a matching covered graph 퐺 were recognized early on: (i) if 퐵 is a barrier of 퐺, and 퐾 is an odd component of 퐺 − 퐵, then 휕 (퐾) is a tight cut; such a cut is called a barrier cut; (ii) if {푢, 푣} is a 2-vertex cut which is not a barrier, then all components of 퐺 − 푢 − 푣 are even, and if 퐻 is the union of any nonempty proper subset of the components of 퐺 − 푢 − 푣, then the cut 휕 ({푢} ∪ 푉 (퐻)), and similarly 휕 ({푣} ∪ 푉 (퐻)), are tight cuts of 퐺, such cuts are called 2-separation cuts. We refer to these two types of tight cuts as ELP-cuts. A fundamental result of matching theory, due to Edmonds, Lov´asz, and Pulleyblank (1982, [32]) states that if a matching covered graph has a nontrivial tight cut, then it also has a nontrivial ELP-cut. Their proof of this result was based on linear programming techniques. An easier and purely graph theoretical proof was given by Szigeti (2002, [87]). We give an alternative proof of the Edmonds-Lov´asz-Pulleyblank (ELP) Theorem in Chapter 5. Our proof is inspired by the one due to Szigeti mentioned above and relies on certain key properties of (i) Dulmage-Mendelsohn (DM) barriers of matchable graphs, and of (ii) peripheral tight cuts in matching covered graphs. (A nontrivial tight cut 휕 ( 푋) of a matching covered graph 퐺 is peripheral if there is no nontrivial tight cut 휕 ( 푋 ′) of 퐺 such that 푋 ′ ⊂ 푋. Clearly, any matching covered graph which has nontrivial tight cuts must have a peripheral tight cut.)

xiv

Preface

Lemma 5.5 (key lemma related to the existence of DM-barriers) states that if 퐺 is a matchable graph and 푋 is a nonempty proper subset of 푉 (퐺) such that both subgraphs 퐺 [푋] and 퐺 [ 푋] are connected and no edge of 휕 ( 푋) is matchable in 퐺, then there exists a DM-barrier 퐵 of 퐺 such that 퐵 and the vertex sets of all the components of 퐺 − 퐵 are entirely contained in one of the shores of 휕 ( 푋). Lemma 5.1 (key Lemma related to peripheral tight cuts) states that if 퐶 := 휕 ( 푋) is a peripheral tight cut of a matching covered graph 퐺 then there exists an edge 푢푣 of 퐺, with 푢 ∈ 푋 and 푣 ∈ 푋, such that both the subgraphs 퐺 [푋 − 푢] and 퐺 [푋 − 푣] are connected. To see how the two lemmas stated above may be used to prove the ELP-Theorem, let us consider (using the notation of Lemma 5.1) the special case in which 푣 is the only neighbour of 푢 in 푋 and 푢 is the only neighbour of 푣 in 푋. In this case, it is easy to see that 퐺 ′ = 퐺 − 푢 − 푣 is a matchable graph in which no edge of the cut 휕 ( 푋 − 푢) is matchable. By Lemma 5.5, there is a Dulmage-Mendelsohn barrier 퐵′ of 퐺 ′ such that 퐵′ as well as the vertex sets of all the components of 퐺 ′ − 퐵′ are contained either in 푋 − 푢 or 푋 − 푣. Assuming, without loss of generality that 퐵′ ⊂ 푋 − 푢, it is not difficult to show that 퐵′ + 푢 is a nontrivial barrier of 퐺. There are some other cases to be examined because 푣 may not be the only neighbour of 푢 in 푋 and 푢 may not be the only neighbour of 푣 in 푋. We analyse all possible cases and deduce that 퐺 either (i) has a 2-separation or (ii) a nontrivial barrier, say 퐵. In the first case 퐺 has a 2-separation cut. In the second case either 퐺 − 퐵 has a nontrivial component, say 퐾, implying that 휕 (퐾) is a barrier cut of 퐺; or all components of 퐺 − 퐵 are trivial, implying that 퐺 is a bipartite matching covered graph. By Theorem 4.7, whose proof is fairly simple, all tight cuts in a bipartite matching covered graph are barrier cuts. Thus, in all possible cases, 퐺 has an ELP-cut. As a consequence of the ELP Theorem, it follows that a matching covered graph 퐺 is a brick if and only if it is bicritical and 3-connected. Thus, to check that 퐺 is a brick, it suffices to verify that, for any two vertices 푢 and 푣 of 퐺 the graph 퐺 − 푢 − 푣 is connected and is matchable. The ELP Theorem does not provide a characterization of braces. But Theorem 5.17 provides a characterization of braces. According to this theorem, a bipartite graph 퐺 := 퐺 [ 퐴, 퐵] with | 퐴| = |퐵| ≥ 3 is a brace if and only if 퐺 − 푢 1 − 푢 2 − 푣 1 − 푣 2 is matchable for any two distinct vertices 푢 1 and 푢 2 in 퐴, and any two distinct vertices 푣 1 and 푣 2 in 퐵. A strengthening of the ELP Theorem: Using a refinement of the method we have used in our proof of the ELP Theorem, Chen, Feng, Lu, Lucchesi and Zhang (2021, [20]) have been able to establish a strengthening of that result; they showed that, given any tight cut 퐶 of a matching covered graph 퐺, there exists an ELP-cut 퐷 of 퐺 which is laminar to 퐶. Separating Cuts and Solid Bricks: A cut 퐶 of a matching covered graph 퐺 is a separating cut if both the 퐶-contractions of 퐺 are also matching covered. All tight cuts are separating cuts, but not every separating cut is tight. For example, the brick P (the Petersen graph), being a brick, has no nontrivial tight cuts but, for any pentagon 푄 of P, the cut 휕 (푄) is a separating cut.

Preface

xv

A matching covered graph 퐺 is solid if it has no separating cuts other than those which are tight. All bipartite graphs are solid; a brick is solid if it has no nontrivial separating cuts. We discovered this concept in connection with our work on the conjecture of Lov´asz concerning the existence of 푏-invariant edges in bricks; we were able to show that in a solid brick 퐺 any edge which is removable is also 푏-invariant. Our inductive proof of Lov´asz’s conjecture was based on this result. The perfect matching polytope of a graph 퐺 is the set of all convex linear combinations of incidence vectors of perfect matchings of 퐺. Edmonds (1965 ,[30]) showed that a vector x in the edge space R퐸 of 퐺 is in this polytope if and only if (i) x(푒) ≥ 0, for each 푒 ∈ 퐸 (non-negativity); (ii) x(휕 (푣)) = 1, for all 푣 ∈ 푉 (degree constraints); and (iii) x(휕 (푆)) ≥ 1, for each odd subset 푆 of 푉 (odd set constraints). We were pleasantly surprised to discover that the only bricks whose perfect matching polytopes can be described without using the odd set constraints are the solid bricks. It is not known if the problem of deciding whether or not a given brick is solid is in the complexity class NP. This is one of the challenging unsolved problems of this theory. The dependence relation: An edge 푒 of a matching covered graph 퐺 is said to depend on another edge 푓 if every perfect matching of 퐺 which contains 푒 also contains 푓 . Two edges 푒 and 푓 are mutually dependent if they depend on each other. Mutual dependence is an equivalence relation, and we refer to equivalence classes with respect to this relation as dependency classes. The partial order of dependence among the edges could be extended to partial order among dependency classes. This leads to the definition of a removable class, which is a generalization of the idea of a removable edge. Lemma 5.4.5 in Lov´asz and Plummer [59] states that any minimal removable class in a matching covered graph consists of either a single edge or a pair of edges. In Chapter 8 we give a new proof of that Lemma. (The notion of a removable edge extends naturally to a removable single ear, and that of a removable doubleton to a removable double ear. We have found it convenient to adopt a ‘top-down’ approach to the theory of ear decompositions of matching covered graphs. Lov´asz’s conjecture is clearly suggestive of this viewpoint!) Most of the above described material is confined to the first part of the book which deals with what we regard as basic theory. In the remaining two parts, there are many new results whose proofs involve intricate proof techniques. We hope that all readers, regardless of the degree of their familiarity with this subject, will find something of interest among these pages. Notation and Terminology: General mathematical notation which we use is mostly

standard except, when there is no scope for confusion, we write 푋 + 푌 for the union 푋 ∪ 푌 of two sets 푋 and 푌 and write 푋 − 푌 for their difference 푋 \ 푌 . Graph theoretical notation and terminology which we use, with minor differences, are mostly the same as the ones used in the book Graph Theory by Bondy and Murty (2008, [3]). Whenever in doubt, readers should either refer to the section Notation and Terminology which appears in the Front Matter or to the index at the end.

xvi

Preface

Broad outline: There are twenty-one chapters, divided into three parts. Each chapter begins with a mini-table of contents which lists the titles of all the sections and subsections in that chapter. Throughout the book, there are a number of exercises of varying difficulty. At the end of each chapter there is a Notes section which includes references to useful related work and to statements of unsolved problems. There is an appendix at the end of the book which includes solutions to some of the harder exercises. How to use the book: The theory of matching covered graphs is a fascinating

subject. It is our hope that those who take a course based on this book will come to appreciate the scope and beauty of this area and that some of them may even end up doing research in it. However, a warning is in order. If one tries to learn all the results and their proofs in a chapter before proceeding to the next, one would not get very far with the subject in just one term or two. For this reason, it is better for the instructor to judiciously select topics to be covered in class and to avoid presenting detailed proofs of all the theorems. A better approach would probably be to convey through examples the ideas underlying concepts introduced and statements of some of the harder theorems and their proofs. It would also be good to encourage student participation. This could be done by assigning selected material for classroom presentation. (For this purpose, we have highlighted certain sections of the book and have indicated them by the symbol ♯.) (There are some sections which are marked by ♯♯. We recommend that they be skipped on a first reading.) It might also be a good idea to encourage students to work together on selected exercises, which are marked with ⊲, a small triangle. Some exercises have solutions in the appendix, and are marked with ∗, an asterisk. Acknowledgements: First and foremost, we would like to express our gratitude to L´aszl´o Lov´asz whose article Matching Structure and the Matching Lattice, published in 1987, served as an inspiration for us to pursue research in this field. The first proof of Lov´asz’s conjecture concerning the existence of 푏-invariant edges in bricks was a part of Marcelo Carvalho’s PhD thesis, written under the supervision of Cl´audio Lucchesi, and finished in 1996. A strengthening of this result, which determined the minimum number of double ear additions required in any ear decomposition of a matching covered graph, appeared in 2002 in a series of papers co-authored by CLM.3 We are thankful to Marcelo for his many contributions to this subject and his warm friendship. Charles Little has done pioneering work in the theory of Pfaffian orientations of graphs. We have known him for a number of years and have learned much about this subject from him. When we finished a draft of this book and were looking for knowledgeable readers, he was the first person whose name occurred to us. We were happy that he undertook the arduous task of going through our work. He read the manuscript we sent to him very thoroughly, literally line-by-line, and sent us detailed comments. Apart from the more substantive observations about the mathematical content, he pointed out many misprints, instances of improper punctuation and grammatical errors. We owe a huge debt of gratitude to Charles. 3 CLM is an abbreviation of Carvalho, Lucchesi and Murty.

Preface

xvii

Arnaldo Mandel is another friend who agreed to read our manuscript. He has an extensive knowledge of many areas of mathematics and computer science. He is jovial and is fond of cryptic multilingual puns. Luckily, he communicated his comments to us in plain language. We thank him for his pertinent suggestions regarding terminology and style. Adrian Bondy has been a dear friend to us for many years. Clarity of his thinking and his artistic sensibilities are reflected in his lucid expository writing. We have learned a lot from him. Thank you dear Adrian. Alberto Miranda, who did his doctorate under the supervision of Cl´audio Lucchesi, and Nishad Kothari, who did his under the partial supervision of Rama Murty, have provided valuable input. We appreciate their help. Gabriel Morete de Azevedo, Sebasti˜ao F. Nachtergaele Gomes Navarro, Thiago Lima Oliveira and Ariana Maite Quispe Porras, who are four students in the University of S˜ao Paulo, have read the proofs of our book with utmost care and found some errors or misprints and, more importantly, suggested various changes to the text to make it more readable. We appreciate their help and wish them the very best in their academic careers. Cl´audio would like to express his gratitude to Lu´ısa Ang´elica (in memoriam), to Rosa Marilyn, and to Gabriela, Ricardo and Andrea, for their personal and emotional support. On his part, Rama has been the grateful beneficiary of affection and unwavering support from his friends Daksha and Kirti Shah and their entire family, and his siblings and their families. Finally, we wish to acknowledge the partial financial support of the Brazilian capes, and of cnpq, grant 309451/2021-3. Cl´audio Lucchesi and Rama Murty June 2023

Contents

Part I Basic Theory 1

Perfect Matchings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2

Matching Covered Graphs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

3

Canonical Partitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

4

Tight Cuts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

5

Characterizations of Bricks and Braces . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

6

The Perfect Matching Polytope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

7

Solid Bricks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139

8

Dependence Relation and Removable Classes . . . . . . . . . . . . . . . . . . . . . . 155

9

Dependence Classes in Bricks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183

3

10 Removable Classes in Solid Bricks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207 11 Ear Decompositions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229 12 Matching Minors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255 Part II Brick and Brace Generation 13 A Conjecture of Lov´asz Concerning Bricks . . . . . . . . . . . . . . . . . . . . . . . . 277 14 Robust Cuts in Bricks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 289 15 풃-Invariant Edges in Bricks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303 16 The Matching Lattice and Optimal Ear Decompositions . . . . . . . . . . . . 323 xix

xx

Contents

17 Thin Edges in Bricks and Braces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 347 18 Strictly Thin Edges in Bricks and Braces . . . . . . . . . . . . . . . . . . . . . . . . . . 393 Part III Pfaffian Orientations 19 Pfaffian Orientations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 417 20 Similarity of Orientations and Characteristic Orientations . . . . . . . . . . 437 21 Excluded-Minor Characterizations of Pfaffian Graphs . . . . . . . . . . . . . . 461 Unsolved Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 497 A

Solutions to Selected Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 499

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 555 List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 561 Glossary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 569 Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 573

Terminology and Notation

The reader is urged to peruse the glossary, which describes the notation used in the book.

1 Sets and Subsets Much of the notation we use in dealing with sets and their subsets is standard. However, in some cases, we have found it natural and convenient to depart from the general conventions. For example, if 푆 and 푇 are two sets, we denote the set of all elements in 푆 which are not also elements of 푇 by 푆 − 푇. This subset of 푆 is commonly represented by 푆 \ 푇.

If 퐴 is any subset of a set 푆, and 푥 is any element of 푆 that may or may not be an element of 퐴, we write 퐴 − 푥 instead of 퐴 \ {푥}, and 퐴 + 푥 instead of 퐴 ∪ {푥}.

2 Graphs and subgraphs All graphs considered in this book are loopless, but they may have multiple edges. For graph theoretical notation and terminology, we essentially follow Graph Theory (2008, [3]) by Bondy and Murty, or an earlier version Graph Theory with Applications (1976, [2]) of that book.

We shall assume that the reader is familiar with the basic notions of graph theory such as subgraphs, induced subgraphs, spanning subgraphs, paths, cycles, trees, connected graphs, cut vertices, cut edges and blocks. However, for the convenience of the reader, we list below some of the graph theoretical terms (and the symbols that pertain to them) that are used in this work. In some cases, our notation and terminology vary from the ones used in the standard texts of graph theory. These differences are also noted below. Neighbour set: For a vertex 푣 of a graph 퐺, the neighbour set of 푣, denoted by 푁퐺 (푣), is the set of all vertices of 퐺 that are adjacent to 푣. More generally, for a subset 푆 of the vertex set 푉 (퐺) of 퐺, the neighbour set 푁퐺 (푆) of 푆 is the set of all vertices of 퐺 in 푉 (퐺) − 푆 which are adjacent to at least one vertex in 푆.

xxi

xxii

Terminology and Notation

When graph 퐺 is evident from the context, we simply write 푁 (푣) for 푁퐺 (푣) and 푁 (푆) for 푁퐺 (푆). We use this convention of leaving out the suffix 퐺 from graph theoretical parameters whenever there is no scope for ambiguity. For example, we simply write 푉 for 푉 (퐺) and 퐸 for 퐸 (퐺). Order and size: The order of a graph is the number of its vertices, and its size is the number of its edges. When the graph under consideration is denoted by 퐺, we denote its order, which is |푉 |, by 푛, and its size, which is |퐸 |, by 푚. Deletions of vertices and edges: For a subset 푆 of the vertex set 푉 of a graph 퐺, the subgraph obtained by deleting all the vertices in 푆 is denoted by 퐺 − 푆. In the special case when 푆 has just one vertex 푣, we simplify the cumbersome notation 퐺 − {푣} to 퐺 − 푣. Similarly, for a subset 퐹 of the edge set 퐸 of 퐺, we denote the subgraph of 퐺 obtained by deleting all edges in 퐹 by 퐺 − 퐹. Again, if 퐹 happens to have a single edge 푒, we write 퐺 − 푒 instead of 퐺 − {푒}. Connectivity: A vertex cut of a graph 퐺 is a subset 푆 of 푉 such that 퐺 − 푆 is disconnected. A 푘-vertex-cut is a vertex cut of cardinality 푘. A graph that has a complete graph as a spanning subgraph has no vertex cuts, all other graphs do. If a graph 퐺 has at least one pair of nonadjacent vertices, the connectivity of 퐺, denoted by 휅(퐺), is the minimum value of 푘 for which 퐺 has a 푘-vertex-cut; otherwise we define 휅(퐺) to be 푛 − 1. (This definition is the same as the one given in [2], but for graphs with a complete graph as a spanning subgraph, it differs from the one in [3].) A graph is 푘-connected if 휅(퐺) ≥ 푘. By Menger’s theorem (9.1 in [3]), a graph 퐺 of order 푘 + 1 or more is 푘-connected if and only if there are at least 푘 internally-disjoint paths between any two nonadjacent vertices of 퐺. Definitions and notation: edge cuts and shores For a nonempty proper subset 푆 of the set of vertices of a graph 퐺, we follow [3] and denote the set of edges with precisely one end in 푆 by 휕 (푆), and refer to it as the edge cut, or simply the cut, associated with 푆. If 푣 is a vertex, then, we denote 휕 ({푣}), for brevity, by 휕 (푣). In the same spirit, if 퐻 is a subgraph of 퐺, we denote the cut 휕 (푉 (퐻)) simply by 휕 (퐻).

The shores of a cut 휕 (푆) of a connected graph 퐺 are the two sets 푆 and 푆 := 푉 −푆.

A cut 휕 (푆) of a graph is trivial if either |푆| or |푆| (or both) have exactly one vertex.

A cut 휕 (푆) in a graph 퐺 is a bond if both its shores induce connected subgraphs of 퐺. Edge Connectivity: A 푘-edge-cut of a graph 퐺 is an edge cut with 푘 edges. All nontrivial graphs do have edge cuts. If a graph 퐺 is nontrivial its edge connectivity, denoted by 휅 ′ (퐺), is the minimum value of 푘 for which 퐺 has a 푘-edge-cut; if 퐺 is trivial, then we define 휅 ′ (퐺) to be zero. A graph is 푘-edge-connected if its edge-connectivity is at least 푘. By the edge version of Menger’s theorem (9.7 in [3]), a nontrivial graph 퐺 is 푘-edge-connected if and only if there are at least 푘 edge-disjoint paths between any two distinct vertices of 퐺.

Terminology and Notation

xxiii

For any graph 퐺, the inequality 휅 ′ ≥ 휅 holds. There are graphs for which this inequality is strict. However, if 퐺 is a cubic graph, then 휅 ′ = 휅. Following Bondy and Murty [3], the minimum and maximum degrees of the vertices of 퐺 are denoted respectively 훿(퐺) and Δ(퐺).

3 Operations Contraction of a Set of Vertices: Let 푋 be a set of vertices of a graph 퐺. The graph obtained from 퐺 by contracting 푋 to a single new vertex 푥 (and removing all resulting loops) is denoted by 퐺/( 푋 → 푥). If the name of the contraction vertex 푥 is irrelevant we just write 퐺/푋.

Part I

Basic Theory

Chapter 1

Perfect Matchings

Contents 1.1 1.2 1.3

1.4 1.5

1.6

Matchings and Perfect Matchings . . . . . . . . . . . . . . . . . . . . . . . . . . . . The Advent of Tutte . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Theorems of Tutte and Petersen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3.1 Matchable graphs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3.2 Cubic matchable graphs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Hall’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The Magic of Pfaffians . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.5.1 Pfaffians . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.5.2 The matrix of indeterminates . . . . . . . . . . . . . . . . . . . . . . . . . 1.5.3 Tutte’s original proof of his theorem . . . . . . . . . . . . . . . . . . . Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

3 5 7 7 10 13 15 16 18 20 22

1.1 Matchings and Perfect Matchings The objects of principal interest in this book are perfect matchings in graphs. A matching in a graph 퐺 is a subset 푀 of 퐸 (퐺) such that no vertex of 퐺 is incident with more than one edge in 푀. A perfect matching in 퐺 is a matching such that every vertex of 퐺 is incident with precisely one edge in it. In other words, a subset 푀 of 퐸 (퐺) is a perfect matching in 퐺 if the subgraph 퐺 [푀] of 퐺 induced by 푀 is a spanning 1-regular subgraph of 퐺. For this reason, perfect matchings in a graph are also known as its 1-factors. The Petersen graph (Figure 1.1(a)), which plays a prominent role in our subject, has six perfect matchings. On the other hand, the Sylvester graph (Figure 1.1(b)) has no perfect matchings. Perfect matchings in graphs have been of interest ever since Tait showed that the four colour conjecture was equivalent to the statement that every 3-connected cubic planar graph is 3-edge-colourable or, equivalently, that such a graph has three pairwise edge-disjoint perfect matchings (see Bondy and Murty [3]). In fact, the four colour conjecture was later shown to be equivalent to the seemingly stronger © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7_1

3

4

1 Perfect Matchings

(푎)

(푏)

Fig. 1.1 (a) The Petersen graph; (b) The Sylvester graph

statement that every 2-connected cubic planar graph is 3-edge-colourable (see [3]). The Sylvester graph shows that a cubic planar graph need not have even a single perfect matching if there were no restriction on its connectivity. As we shall see (Theorem 1.4), Petersen [80] showed that any 2-connected cubic graph (planar or not) has at least one perfect matching. However, a nonplanar 2-connected cubic graph need not have two edge-disjoint perfect matchings; the Petersen graph is an example of such a graph (Exercise 1.1.2). This book is concerned mainly with graphs which have perfect matchings. For brevity, we shall refer to such graphs by the term defined in the following inset. Matchable graphs A graph is matchable if it has a perfect matching.

Before working on the exercises below, the readers should familiarize themselves with the notation we use for edge cuts in graphs. (If 푋 is a subset of the vertex set 푉 of a graph 퐺, then 휕 ( 푋) is the set of edges of 퐺 with one end in 푋 and one end in 푉 − 푋. For a subgraph 퐻 of 퐺, we simply write 휕 (퐻) for 휕 (푉 (퐻)); and for a vertex 푣, we write 휕 (푣) for 휕 ({푣}).) The following proposition plays a useful role in arguments involving perfect matchings and cuts in a graph. Proposition 1.1 For any perfect matching 푀 and for any subset 푋 of the vertex set 푉 of a graph 퐺 the following identity holds: |휕 ( 푋) ∩ 푀 | ≡ | 푋 |

(mod 2).

We leave the proof of the above proposition as Exercise 1.1.3(ii).

ƒ

1.2 The Advent of Tutte

5

Exercises 1.1.1 Show that the Sylvester graph is not matchable. 1.1.2 (i) Show that the Petersen graph is the unique cubic graph of girth five on ten vertices. (ii) For any pentagon 푄 and any perfect matching 푀 of the Petersen graph, show that |휕 (푄) ∩ 푀 | is either one or five. (iii) Show that every perfect matching of the Petersen graph is one of its 5-edge cuts. (iv) Deduce that every edge of the Petersen graph is in exactly two perfect matchings, and that any two perfect matchings have exactly one edge in common. (This also shows that the Petersen graph is nonhamiltonian.) (v) List all the perfect matchings of the Petersen graph. ⊲1.1.3 Let 퐺 be a graph. (i) For any subset 푋 of 푉, show that |휕 ( 푋)| ≡ 푟 (mod 2), where 푟 is the number of vertices in 푋 that have odd degree in 퐺. (ii) Deduce Proposition 1.1 from part (i).

1.2 The Advent of Tutte Frobenius (1917, [35]) and Hall (1935, [40]) established necessary and sufficient conditions for the existence of perfect matchings in bipartite graphs. But the more general problem of characterizing graphs which have a perfect matching remained unsolved until W. T. Tutte entered the field. Tutte started working in graph theory during his undergraduate days at Cambridge, but his studies were interrupted by the call to wartime duty at Bletchley Park, the fabled organization of codebreakers. During his tenure there (1941-45), he made pivotal contributions to cryptography (see Younger (2012, [95]). He returned to Cambridge after the war to resume his research in graph theory. Within a short period, he produced a number of brilliant results, one of them is a solution to the problem mentioned above. In his mathematical autobiography “Graph Theory As I Have Known It” (1998, [90]), he describes in a delightfully engaging manner how he was led to his solution of that problem. Obviously, in order for a graph 퐺 to have a perfect matching, it must have an even number of vertices. Thus, for any odd subset 푋 of 푉, the subgraph 퐺 [푋] induced by 푋 cannot have a perfect matching. Consequently, if 푀 is any perfect matching of 퐺, and 푋 is any odd subset of 푉, then |휕 ( 푋) ∩ 푀 | ≥ 1. (In fact |휕 ( 푋) ∩ 푀 | must be odd.) This simple observation is the basis of a necessary condition for the existence of a perfect matching in a graph 퐺. The notation described in the following inset is convenient for stating that condition.

6

1 Perfect Matchings

Notation related to odd components: O(퐺) and 표(퐺) We denote by O(퐺) the set of odd components of a graph 퐺 and by 표(퐺) the number |O(퐺)| of odd components of 퐺. Suppose that 퐺 has a perfect matching 푀 and let 푆 ⊆ 푉. Then, by the previous observation, for any odd component 퐾 of 퐺 − 푆, there must be at least one edge of 푀 in 휕 (퐾). As 퐾 is a component of 퐺 − 푆, every edge in 휕 (퐾) has one end in 퐾 and one end in 푆. A necessary condition for the existence of a perfect matching in 퐺 Since no two distinct edges of a perfect matching of 퐺 can have a common end vertex, the following condition 표(퐺 − 푆) ≤ |푆|,

for any subset 푆 of 푉

(1.1)

must be satisfied in order for 퐺 to have a perfect matching. Thus, if there exists a subset 푆 of 푉 such that 표(퐺 − 푆) > |푆|, then 퐺 cannot be matchable. See Example 1.2. Example 1.2 Figure 1.2 depicts a subset 푆 of the vertex set of a graph 퐺 (indicated by solid dots) which violates condition 1.1. Thus 퐺 is not matchable.

Fig. 1.2 |푆 | = 5, and 표 (퐺 − 푆 ) = 7

Tutte (1947, [89]) showed that the necessary condition (1.1) for the existence of a perfect matching in a graph 퐺 is also sufficient. Tutte’s proof is based on properties of invariants, known as Pfaffians, associated with skew-symmetric matrices. (They are described in Section 1.5.) As this proof involves many ideas that may be unfamiliar to readers who are more accustomed to combinatorial methods than algebraic ones, we have chosen to present, in the next section, a purely graph theoretic proof due to Lov´asz (1975, [56]) which essentially follows the same line of reasoning used by

7

1.3 Theorems of Tutte and Petersen

Tutte but which avoids the use of Pfaffians. Part-III of the book (Pfaffian Orientations) is partly dedicated to questions arising from Tutte’s seminal work.

Exercises ⊲1.2.1 Let 퐺 be a graph, and let 푋 be a subset of 푉. Show that: | 푋 | − 표(퐺 − 푋) ≡ |푉 |

(mod 2).

Deficiency of a matching The deficiency of a matching 푀 in a graph 퐺, denoted def(푀), is the number of vertices of 퐺 that are not covered by 푀. ∗ 1.2.2 Let 푀 be a matching of a graph 퐺. (i) Show that for any subset 푆 of 푉, def(푀) ≥ 표(퐺 − 푆) − |푆|.

(1.2)

(ii) Conclude that if equality holds in (1.2) for some set 푆 then 푀 is a maximum matching of 퐺 (that is, a matching of largest possible cardinality). 1.2.3 Find a maximum matching in the graph shown in Figure 1.2 and justify your answer.

1.3 Theorems of Tutte and Petersen We now proceed to describe Tutte’s characterization of graphs which have a perfect matching.

1.3.1 Matchable graphs

Tutte’s Theorem [89] 1.3 A graph 퐺 has a perfect matching if and only if 표(퐺 − 푆) ≤ |푆|, for every subset 푆 of 푉.

(1.3)

8

1 Perfect Matchings

Proof The proof we give here is essentially the same as it is in Tutte’s original paper, but we use the approach of Lov´asz (1975, [56]). See Bondy and Murty [2] or Lov´asz and Plummer (1986, [59]). We have already seen that if 퐺 has a perfect matching, then (1.3) holds. Conversely, suppose that 퐺 is a graph that has no perfect matchings. It is incumbent on us to establish the existence of a subset 푆 of 푉 for which 표(퐺 − 푆) > |푆|. We shall refer to such a set as a Tutte set. Clearly, we may assume without loss of generality that 퐺 is simple. If |푉 | is odd, then the empty set is a Tutte set. So, we may assume that |푉 | is even. Let 퐺 ∗ be a maximal simple graph, with 푉 (퐺 ∗ ) = 푉 (퐺) and 퐸 (퐺 ∗ ) ⊇ 퐸 (퐺), such that 퐺 ∗ also has no perfect matchings. (If 푢 and 푣 are two nonadjacent vertices of 퐺 such that 퐺 − 푢 − 푣 has no perfect matchings, then 퐺 + 푢푣, obtained from 퐺 by adding the edge 푢푣, also has no perfect matchings. A graph 퐺 ∗ satisfying the required property may be obtained by repeatedly adding edges in this manner.) We note that, for any subset 푆 of 푉, 표(퐺 − 푆) ≥ 표(퐺 ∗ − 푆). Thus, any Tutte set 푆 of 퐺 ∗ would also be a Tutte set of 퐺. Let 푆 denote the set of vertices of degree 푛 − 1 in 퐺 ∗ , where 푛 is the number of vertices of 퐺. (Thus, any vertex in 푆 is joined in 퐺 ∗ to all other vertices in 푉.) If 푆 = 푉, then 퐺 ∗ is a complete graph and hence would be matchable. Therefore 푆 is a proper subset of 푉 (it is possible for 푆 to be empty). Clearly the subgraph of 퐺 ∗ induced by 푆 is a complete graph. 1.3.1 Every component of 퐺 ∗ − 푆 is also complete. Proof Consider any component 퐽 of 퐺 ∗ − 푆. If it is not complete, it would have two nonadjacent vertices 푥 and 푦 with a common neighbour 푧 (see Exercise 1.3.1). Since 푧 is in 푉 (퐽), and not in 푆, the degree of 푧 is less than 푛 − 1. Let 푤 be a vertex that is not a neighbour of 푧. Clearly 푤 is not in 푆, but it may not be in 푉 (퐽) either. See Figure 1.3. 푥

푦

푤

푧

Fig. 1.3 The case in which 퐽 is not complete

Since 퐺 ∗ is a maximal simple graph that is not matchable, the graph 퐺 ∗ + 푢푣 is matchable, for any two nonadjacent vertices 푢 and 푣 of 퐺 ∗ . Let 푀1 and 푀2 denote perfect matchings in 퐺 ∗ +푥푦 and 퐺 ∗ +푤푧, respectively, and denote by 퐻 the subgraph of 퐺 ∗ + 푥푦 + 푤푧 induced by the symmetric difference 푀1 △ 푀2 . Since each vertex

9

1.3 Theorems of Tutte and Petersen

of 퐻 has degree two, 퐻 is a disjoint union of cycles. Furthermore, all these cycles are even, since edges of 푀1 alternate with edges of 푀2 around them. We distinguish two cases: Case 1 The edges 푥푦 and 푤푧 are in the same cycle 퐶 of 퐻 (Figure 1.4(a)).

푤

푥

푦

푧

(푎)

푥

푤

푦

푧

(푏)

Fig. 1.4 Perfect matching 푀1 is indicated by heavy lines, and 푀2 by dashed lines

By the symmetry of 푥 and 푦, we may assume that the vertices 푧, 푤, 푥 and 푦 occur in 퐶 in that cyclic order. The edges of 푀1 in the segment 푧푥 of 퐶, together with the edge 푥푧 and the edges of 푀2 not in the segment 푧푥 constitute a perfect matching of 퐺 ∗ . This is a contradiction to the definition of 퐺 ∗ . Case 2 The edges 푥푦 and 푤푧 are in distinct components of 퐻 (Figure 1.4(b)). Let 퐶 denote the cycle of 퐻 that contains the edge 푤푧. The edges of 푀1 in 퐶, together with the edges of 푀2 not in 퐶, constitute a perfect matching of 퐺 ∗ . This is a contradiction to the definition of 퐺 ∗ . ƒ We now use the fact that every component of 퐺 ∗ − 푆 is complete. For brevity, let us write 푡 := 표(퐺 ∗ − 푆). Our objective is to show that 푡 > |푆|. Assume to the contrary that 푡 ≤ |푆|. In this case, let 푇 be a subset of 푆 of cardinality 푡, and let 푇 ′ be a set of 푡 vertices consisting of distinct representatives of the vertex sets of the odd components of 퐺 ∗ − 푆. Since all vertices in 푆 have degree 푛 − 1, there exists a matching, say 푀, which matches the vertices in 푇 with vertices in 푇 ′ . As |푉 | is even, |푆| and 표(퐺 ∗ − 푆) have the same parity (see Exercise 1.2.1). Thus 푆 − 푇 is even, and the subgraph of 퐺 ∗ induced by this set is a complete graph. Furthermore, each component of 퐺 ∗ − (푆 ∪ 푇 ′ ) is a complete graph on an even number of vertices. Hence the matching 푀 can be extended to a perfect matching of 퐺 ∗ . This contradicts the assumption that 퐺 ∗ has no perfect matchings. It follows that 푡 = 표(퐺 ∗ − 푆) > |푆|. ƒ The problem of deciding whether or not a given graph 퐺 has a perfect matching is obviously in NP; if 퐺 has a perfect matching, that fact can be demonstrated by displaying a perfect matching in 퐺. The import of Tutte’s theorem is that the above problem is also in co-NP; if 퐺 does not have perfect matchings, that fact can

1 Perfect Matchings

10

be demonstrated by displaying a Tutte set in 퐺. In fact, the literature has several polynomial-time algorithms for determining maximum perfect matchings. The first such algorithm was given by Edmonds (1965, [31]). That article, and another article published by Cobham (1965, [24]), are regarded by computer scientists as fundamental contributions to the theory of Computational Complexity, by the introduction of the class P, the class of problems for which there are polynomial-time algorithms.

Complexity of Algorithms In a pioneering paper called “Paths, Trees and Flowers” (1965, [31]), Jack Edmonds persuasively argued the importance of polynomial-time algorithms and, as an example, presented an algorithm which finds a maximum matching in a graph with running time 푂 (푛2 푚). Hopcroft and Karp (1973, [43]) discovered √ an algorithm to ( 푂 time in graph determine a maximum matching of a bipartite 푛푚). Micali and √ Vazirani (1980, [70]) discovered an 푂 ( 푛푚)-time algorithm for finding a maximum matching of a graph.

1.3.2 Cubic matchable graphs Tutte had known Petersen’s Theorem on cubic graphs as a schoolboy. (A very interesting account of his encounter with a published proof of this theorem, in his own words, can be found in Chapter 6 of McCarthy and Stanton’s book (1979, [67]).) With the aid of his characterization of graphs which have a perfect matching, he was able to present a simple proof of that result. Petersen’s Theorem (1891, [80]) 1.4 Every 2-connected cubic graph is matchable. Proof Let 퐺 be a 2-connected cubic graph. As all vertices of 퐺 have degree three, |푉 | is even and 표(퐺) = 0. Thus, 표(퐺 − ∅) = 0. Let 푆 be any nonempty proper subset of 푉. For every odd component 퐾 of 퐺 − 푆, |휕 (퐾)| is odd (Exercise 1.1.3). On the other hand, by hypothesis, 퐺 is 2-connected. Hence |휕 (퐾)| ≥ 3, for each odd component 퐾 of 퐺 − 푆. We thus have: Õ 3|O(퐺 − 푆)| ≤ |휕 (퐾)| ≤ |휕 (푆)| ≤ 3|푆|, (1.4) 퐾 ∈ O (퐺−푆)

implying that 표(퐺 − 푆) ≤ |푆|, for all subsets 푆 of 푉. Hence, by Theorem 1.3, 퐺 has a perfect matching. ƒ

1.3 Theorems of Tutte and Petersen

11

Tutte not only deduced from his Theorem 1.3 that every 2-connected cubic graph is matchable, but also showed that, in fact, every such graph has the property that every edge is in some perfect matching. This stronger assertion follows from a more general result about matchable graphs which will be discussed in Chapter 2 (see Theorem 2.8). But a keen student should be able to work out its proof without much difficulty. The above-mentioned property of 2-connected cubic graphs suggests the definition of the following more general class of graphs whose name appears as the subtitle of this book: a nontrivial connected graph is matching covered if, given any edge of the graph, there is some perfect matching which contains that edge. Basic terminology related to this class and several interesting families of examples will be described in the next chapter.

Exercises ⊲1.3.1 Show that if a connected graph 퐺 is not a complete graph, then it would have two nonadjacent vertices 푥 and 푦 with a common neighbour 푧. Alternating paths and cycles, augmenting paths Let 푀 and 푁 be matchings in a graph 퐺. A path or a cycle of 퐺 is 푀-alternating if its edges are alternately in 푀 and in 퐸 − 푀, and is (푀, 푁)-alternating if its edges are alternately in 푀 − 푁 and in 푁 − 푀. A nontrivial 푀-alternating path in 퐺 is 푀-augmenting if neither of its ends is covered by 푀. ⊲1.3.2 (i) Let 푀 and 푁 be two matchings in a graph 퐺. Show that each component of the subgraph 퐺 [푀 △ 푁] of 퐺 induced by 푀 △ 푁 is either an (푀, 푁)-alternating path or an (푀, 푁)-alternating cycle. Moreover, if |푁 | > |푀 |, show that some component of 퐺 [푀 △ 푁] is an (푀, 푁)-alternating path whose first and last edges are in 푁. (ii) Deduce from part (i) that a matching 푀 in a graph 퐺 is maximum (that is, one of maximum size) if and only if 퐺 has no 푀-augmenting paths. (Berge, 1962) ∗ 1.3.3 Let 푀0 be a maximal matching of 퐺 (that is, one which is not a proper subset of any other matching of 퐺). Show that |푀 | ≤ 2|푀0 |, for every matching 푀 of 퐺. In particular, if 푀 ∗ denotes a maximum matching of 퐺, deduce that |푀 ∗ | ≤ 2|푀0 |. 1.3.4 Let 퐺 be a cubic graph that has a path which contains all cut edges of 퐺. Show that 퐺 has a perfect matching. (The Sylvester graph shows that this is the best possible statement one can make about cubic graphs which have a perfect matching.) ∗ 1.3.5 For each 푟 ≥ 3, construct an (푟 − 2)-edge-connected 푟-regular graph on an even number of vertices which is not matchable.

12

1 Perfect Matchings

푟-graphs For an integer 푟 ≥ 2, an 푟-graph is an (푟 − 1)-edge-connected 푟-regular graph on an even number of vertices. 1.3.6 Show that every 푟-graph is matchable. ⊲1.3.7 Show that the graph depicted in Figure 1.5 is a 4-edge connected 4-graph and that it is not 4-edge-colourable. (For an example of a simple graph with these properties, see Exercise 2.4.7.)

Fig. 1.5 A 4-graph which is not 4-edge-colourable

Critical graphs A graph 퐺 is factor-critical, or simply critical, if, for any 푣 ∈ 푉, the graph 퐺 − 푣 has a perfect matching. (For example, the odd cycle 퐶2푘+1 is critical for all 푘 ≥ 1.) Critical graphs are also known as hypomatchable graphs. Clearly, a critical graph has an odd number of vertices. 1.3.8 Let 퐺 be a critical graph on three or more vertices. Show that 퐺 is connected and contains an odd cycle. 1.3.9 Using Tutte’s Theorem 1.3 show that a graph 퐺 is critical if and only if: 표(퐺 − 푆) ≤ |푆| − 1 for every nonempty subset 푆 of 푉.

13

1.4 Hall’s Theorem

1.4 Hall’s Theorem ♯ We now present an attractive characterization of bipartite matchable graphs which is usually attributed to Hall [40]. We shall denote by 퐺 [ 퐴, 퐵] a bipartite graph 퐺 with bipartition ( 퐴, 퐵). Hall established necessary and sufficient conditions under which a bipartite graph 퐺 [ 퐴, 퐵] has a matching which matches each vertex in 퐴 with some vertex in 퐵 (see Exercise 1.4.1). When | 퐴| = |퐵|, such a matching is precisely a perfect matching of 퐺, and this special case is what is relevant to our purposes. Hall’s Theorem 1.5 A bipartite graph 퐺 := 퐺 [ 퐴, 퐵], with | 퐴| = |퐵|, has a perfect matching if and only if (1.5) |푁 ( 푋)| ≥ | 푋 |, for all 푋 ⊆ 퐴. Proof Suppose first that 퐺 has a perfect matching 푀. Then 푀 matches distinct vertices in 퐴 with distinct vertices of 퐵. This shows that (1.5) must hold for each 푋 ⊆ 퐴. Conversely, suppose that 퐺 has no perfect matching. In this case, we must establish the existence of a subset 푋 of 퐴 such that | 푋 | > |푁 ( 푋)|. Towards this end, let 퐺 ′ be the spanning supergraph of 퐺 obtained by replacing the stable set 퐵 in 퐺 by a complete graph on 퐵. (See Figure 1.6 for an illustration.) Clearly, as 퐺 has no perfect matching, the graph 퐺 ′ does not have one either. Therefore, by Tutte’s Theorem 1.3 and Exercise 1.2.1, applied to 퐺 ′ , there must exist a subset 푆 of 푉 (퐺 ′ ) such that 표(퐺 ′ − 푆) ≥ |푆| + 2.

(1.6)

푎1

푎2

푎3

푎4

푎5

푎1

푎2

푎3

푎4

푎5

푏1

푏2

푏3

푏4

푏5

푏1

푏2

푏3

푏4

푏5

퐴

퐵



퐺 [ 퐴, 퐵]

퐺 ′

Fig. 1.6 Graph 퐺 is obtained from the bipartite graph 퐺 [ 퐴, 퐵] by joining vertices in 퐵 to each other. The set 푆 := {푏1 , 푏2 } is a Tutte set of 퐺 ′ as 표 (퐺 ′ − 푆 ) has four odd components; 푎1 , 푎2 and 푎4 are the isolated vertices of 퐺 ′ − 푆. If we set 푋 := {푎1 , 푎2 , 푎4 }, then |푋 | = 3, whereas | 푁 (푋) | = 2.

1 Perfect Matchings

14

As 퐺 ′ [퐵] is complete, it follows that all vertices in 퐵 − 푆 will be in the same component of 퐺 ′ − 푆. Every other component of 퐺 ′ − 푆 is of order one and is the subgraph induced by an isolated vertex of 퐺 ′ − 푆 which belongs to 퐴. Let 푋 denote the subset of 퐴 consisting of the isolated vertices of 퐺 ′ − 푆. Then it follows from (1.6) that | 푋 | ≥ 표(퐺 ′ − 푆) − 1 ≥ |푆| + 1 ≥ |퐵 ∩ 푆| + 1. (1.7)

Now note that a vertex in 퐴 is isolated in 퐺 ′ − 푆 only if all its neighbours in 퐺 are in 퐵 ∩ 푆; hence, 퐵 ∩ 푆 ⊇ 푁 ( 푋). This observation combined with (1.7) implies that | 푋 | > |푁 ( 푋)|. ƒ

There are many simple and elegant proofs of Hall’s Theorem which do not rely on Tutte’s Theorem. A constructive proof, based on the notion of alternating paths and cycles, is sketched in Exercises 1.3.2 and 1.4.3. For more details, see Bondy and Murty [3].

Exercises A matching 푀 in a graph 퐺 is said to cover a vertex 푣 if 푣 is incident with some edge in 푀. The assertion that is generally known as Hall’s Theorem is the following: A bipartite graph 퐺 [ 퐴, 퐵] has a matching which covers all vertices in 퐴 if and only if | 푁 (푋) | ≥ |푋 |, for all 푋 ⊆ 퐴.

⊲1.4.1 Show that the above assertion is equivalent to Theorem 1.5. 1.4.2 (i) Show that, for any 푘 > 0, every 푘-regular bipartite graph has a perfect matching. (ii) Deduce that every 푘-regular bipartite graph is 푘-edge-colourable. ⊲1.4.3 Let 퐺 [ 퐴, 퐵] be a bipartite graph, and let 푀 be a maximum matching in 퐺. Suppose that 푢 is a vertex in 퐴 which is not covered by 푀, and that 푍 denotes the set of all vertices in 퐺 which are reachable from 푢 by 푀-alternating paths, and let 푋 := 퐴 ∩ 푍. Show that |푁 ( 푋)| < | 푋 |. 1.4.4 (Deduction of Tutte’s Theorem from Hall’s Theorem) Let 퐺 be a graph and suppose that 퐺 satisfies the Tutte condition 표(퐺 − 푆) ≤ |푆|, for each 푆 ⊆ 푉, and let us refer to a subset 퐵 of 푉 as a barrier of 퐺 if 표(퐺 − 퐵) = |퐵|. Clearly, the empty set is a barrier of 퐺. Let 퐵 denote a maximal barrier of 퐺. Assume, as inductive hypothesis, the sufficiency of Tutte’s condition for graphs having fewer vertices than 퐺.

15

1.5 The Magic of Pfaffians

(i) Prove that every component of 퐺−퐵 is odd. Hint: if 퐺−퐵 has an even component 퐾, add a vertex of 퐾 to 퐵. (ii) Let 퐻 be the bipartite graph obtained from 퐺 by removing all the edges having both ends in 퐵 and by contracting each component of 퐺 − 퐵 to a single vertex. Prove that 퐻 has a perfect matching. Hint: apply Hall’s Theorem to 퐻, by proving that for each set 푋 of contracted vertices of 퐻, the set of neighbours of 푋 (in 퐵) is at least as large as 푋. (iii) Prove that if 퐾 is a component of 퐺 − 퐵 that has a nonempty set 푋 of vertices such that 표(퐾 − 푋) ≥ | 푋 | + 1, then the set 퐵 ∪ 푋 is a barrier of 퐺. (iv) Prove that every component 퐾 is critical. Hint: use the inductive hypothesis and the result in Exercise 1.3.9. (v) Prove that 퐺 is matchable. Hint: use the fact that every component is critical to extend the perfect matching of 퐻 to a perfect matching of 퐺. (Younger, 1969) ⊲1.4.5 (Analogue of Exercise 1.2.2) We denote by 퐼 (퐺) the set of isolated vertices of a graph 퐺. Let 푀 be a matching of a bipartite graph 퐺. (i) Prove the following inequality involving the deficiency of 푀, for every set 푆 ⊂ 푉. def(푀) ≥ |퐼 (퐺 − 푆)| − |푆|.

(1.8)

(ii) Prove that if equality holds in (1.8) for some set 푆 then 푀 is a maximum matching of 퐺.

1.5 The Magic of Pfaffians ♯♯ An 푛 × 푛 matrix A := (푎 푖 푗 ) is skew-symmetric if 푎 푖 푗 = −푎 푗푖 , for 1 ≤ 푖, 푗 ≤ 푛. An example of a skew-symmetric matrix that is appropriate to the present context is the adjacency matrix of an orientation 퐷 of a simple graph 퐺. In fact, every skewsymmetric matrix A := (푎 푖 푗 ) of order 푛 may be viewed as the ‘adjacency matrix’ of an arc-weighted tournament 퐷 which is an orientation of an edge-weighted complete graph 퐺 on {1, 2, . . . , 푛}. The weight assigned an edge 푖 푗 in 퐺 is |푎 푖 푗 | and, in 퐷, edge 푖 푗 is oriented from 푖 to 푗 if 푎 푖 푗 ≥ 0, and from 푗 to 푖 if 푎 푖 푗 < 0. Note that each edge in 퐺 corresponds to two entries of the matrix; when 푎 푖 푗 ≠ 0, one of those entries is positive and the other is negative. The role of the orientation is to indicate which entry is positive (and, implicitly, which one is negative). See Figure 1.7 for an example. If all the entries 푎 푖 푗 of A are integers, one may view A as the incidence matrix of an orientation of a graph with multiple edges (see Exercise 1.5.2). This process can be reversed in an obvious manner; given any arc-weighted digraph, one may associate a skew-symmetric matrix with it.

16

1 Perfect Matchings 1

 0    −2   2  휋 

2 −2 − 휋  0 0 − 휋  0 0 −3  휋 3 0 

휋 2

2 휋

2

4

3

0

3

Fig. 1.7 A skew-symmetric matrix and the corresponding weighted digraph

1.5.1 Pfaffians It is easy to show that when 푛 is odd, the determinant det(A) of a skew-symmetric matrix A of order 푛 is zero (Exercise 1.5.1). When 푛 is even, say 푛 = 2푘, then det(A) is the square of a polynomial Pf (A) in the entries of A known as the Pfaffian of A. In this polynomial, there is one term for each perfect matching of the graph 퐺 associated with A. Let 푀 := {푒 1 , 푒 2 , . . . , 푒 푘 } be a perfect matching of 퐺, and suppose that in the orientation 퐷 of 퐺, the tail and the head of the edge 푒 푖 , for 1 ≤ 푖 ≤ 푘, are 푢 푖 and 푣 푖 , respectively. We may then associate with 푀 the permutation 휋(푀) as defined below:   1 2 3 4 . . . 2푘 − 1 2푘 휋(푀) := . 푢 1 푣 1 푢 2 푣 2 . . . 푢 푘 푣 푘 The sign of 푀, denoted by sign(푀), is +1 if 휋(푀) is an even permutation, and is −1 if 휋(푀) is an odd permutation. It can be seen that sign(푀) is independent of the order in which the edges of 푀 are enumerated. However, the sign does depend on the orientation 퐷 of 퐺. Thus, the same perfect matching of 퐺 may have different signs relative to different orientations of 퐺. Example 1.6 Consider the transitive tournament on {1, 2, 3, 4} (an orientation of 퐾4 in which 푖 precedes 푗 whenever 1 ≤ 푖 < 푗 ≤ 4). The three perfect matchings of this tournament, listed according to the convention mentioned above, are: 푀1 := {(1, 2), (3, 4)}, 푀2 := {(1, 3), (2, 4)}, 푀3 := {(1, 4), (2, 3)}. The permutations (1, 2, 3, 4) and (1, 4, 2, 3) are even permutations of (1, 2, 3, 4); and the permutation (1, 3, 2, 4) is odd. Thus, sign(푀1 ) = +1, sign(푀2 ) = −1, and sign(푀3 ) = +1. Example 1.7 Figure 1.8 shows the signs of the three perfect matchings in two orientations of 퐾4 .

1.5 The Magic of Pfaffians

17 1

4 2

(1, 2) (3, 4) (1, 3) (4, 2) (1, 4) (3, 2)

+ + −

(1, 2) (4, 3) (1, 3) (2, 4) (1, 4) (3, 2)

− − −

3 1

4 2

3

Fig. 1.8 The signs of the perfect matchings in two orientations of 퐾4

Pfaffian of a skew-symmetric matrix Let A be a skew-symmetric matrix of order 푛 = 2푘, and let 퐺 be the associated weighted graph, and its orientation 퐷 as defined in the first paragraph of this section. Let M denote the set of all perfect matchings of 퐺. For a perfect matching 푀 := {푢 1 푣 1 , 푢 2 푣 2 , ..., 푢 푘 푣 푘 } of 퐺, where, for each edge in 푀, its tail in 퐷 is listed first, let sign(푀) denote the sign of 푀. The Pfaffian of A, denoted by Pf (A), is the following polynomial in the entries of A: Õ Pf (A) := sign(푀) 푎 푢1 푣1 푎 푢2 푣2 . . . 푎 푢푘 푣푘 , (1.9)

where the sum is taken over all 푀 in M, the set of perfect matchings of 퐺.

Example 1.8 Consider the incidence matrix A = (푎 푖 푗 ) of the transitive tournament described in Example 1.6. Then, Pf (A) = 푎 12 푎 34 − 푎 13 푎 24 + 푎 14 푎 23 = 1 − 1 + 1 = 1. The following classical theorem, due to Cayley, expresses the magical relationship between the determinant and the Pfaffian of a skew-symmetric matrix.

1 Perfect Matchings

18

Cayley’s Theorem 1.9 For any skew-symmetric matrix A, its determinant det(A) is the square of ƒ its Pfaffian Pf (A). For a readable account of Pfaffians, in particular for a proof of Theorem 1.9, we refer the reader to Godsil (1993, [38]). See also Halton (1966, [41]). When Í A is the adjacency matrix of an orientation 퐷 of a graph 퐺, the Pfaffian of A is 푀 ∈ M sign(푀), where M is the set of all perfect matchings of 퐺. Thus, each nonzero term in the expansion of Pf (A) is either +1 or −1. If all the perfect matchings happen to have the same sign, then the absolute value of Pf (A) would be equal to the number of perfect matchings of 퐺, and that number could be computed by evaluating the determinant of 퐴 and taking its square root. However, there is no guarantee that all perfect matchings have the same sign. For some graphs, it is possible to choose an orientation which would make the signs of all perfect matchings the same. Such an orientation is called a Pfaffian orientation of the graph. The smallest graph that does not admit a Pfaffian orientation is 퐾3,3 . (The study of Pfaffian orientations is the subject of this chapter and also of Chapters 19, 20 and 21). Using Theorem 1.9, Tutte was hoping to find a determinantal formula for the number of perfect matchings. But he soon realized the difficulties involved in this approach, as explained in the previous paragraph. However, he found an ingenious way of circumventing this difficulty when the objective is just to determine whether or not the number of perfect matchings of 퐺 is positive.

1.5.2 The matrix of indeterminates Let A be the adjacency matrix of an orientation 퐷 of a simple graph 퐺. Replace all the nonzero entries above the diagonal by algebraically independent indeterminates, and each nonzero entry below the diagonal by the negative of the corresponding entry above the diagonal. We shall denote the resulting skew-symmetric matrix also by A (so as not to proliferate notation), and refer to it as the matrix of indeterminates associated with 퐺. Since, corresponding to each edge of 퐺, there is one nonzero entry above the diagonal of the incidence matrix, the total number of indeterminates above the diagonal of A is |퐸 |. The Pfaffian of A is then a polynomial in these |퐸 | indeterminates. Proposition 1.10 The Pfaffian Pf (A) (of the matrix A of indeterminates) vanishes if and only if 퐺 has no perfect matchings. Proof By the assumption that the |퐸 | upper diagonal indeterminate entries of A are algebraically independent, it follows that Pf (A), which is a polynomial in those |퐸 | indeterminates, would vanish only if each term in that polynomial is zero. This would happen precisely when 퐺 has no perfect matchings, regardless of the orientation chosen. (For concreteness, a student might wish to take the vertex set of

1.5 The Magic of Pfaffians

19

퐺 to be {1, 2, . . . , 푛}, and the orientation to be the transitive orientation where, for ƒ 1 ≤ 푖 < 푗 ≤ 푛, vertex 푖 precedes vertex 푗.) The above observation is the basis of Tutte’s characterization of graphs which have a perfect matching. For any pair 푥 and 푦 of vertices, we shall denote the minor of the matrix A obtained by deleting the two rows and two columns corresponding to 푥 and 푦 by A 푥 푦 , and the Pfaffian of A 푥 푦 by 푃 푥 푦 . (Thus A 푥 푦 is the matrix of indeterminates associated with the subgraph 퐺 − 푥 − 푦, and 푃 푥 푦 is its Pfaffian.) Similar notation applies to any even subset of vertices; however, we shall simply write 푃 for the Pfaffian of A itself. As an example, consider the 6 × 6 matrix A shown in equation (1.10). 0 © ­ −푎 12 ­ ­ −푎 A = ­ 13 ­ −푎 14 ­ ­ −푎 15 « −푎 16

Then

A13

푎 12 0 −푎 23 −푎 24 −푎 25 −푎 26

0 푎 24 푎 25 © ­ −푎 24 0 푎 45 =­ ­ −푎 25 −푎 45 0 « −푎 26 −푎 46 −푎 56

푎 13 푎 23 0 −푎 34 −푎 35 −푎 36

푎 14 푎 24 푎 34 0 −푎 45 −푎 46

푎 15 푎 25 푎 35 푎 45 0 −푎 56

푎 16 ª 푎 26 ® ® 푎 36 ® ®. 푎 46 ® ® 푎 56 ® 0 ¬

푎 26   ª 0 푎 56 푎 46 ® ® and A1234 = −푎 56 0 푎 56 ® 0 ¬

(1.10)

(1.11)

with 푃13 = 푎 24 푎 56 − 푎 25 푎 46 + 푎 26 푎 45 , and 푃1234 = 푎 56 .

An important ingredient in Tutte’s proof is an identity involving Pfaffians of minors of A of the above type. To get a feel for its statement, consider the skewsymmetric matrix of order six, shown in equation (1.10). Then the following identity in the fifteen variables 푎 푖 푗 may be verified by routine calculations (see Exercise 1.5.3). 푃12 푃34 − 푃13 푃24 + 푃14 푃23 = 푃 푃1234 .

(1.12)

If we set some of the 푎 푖 푗 ’s to be zero, this equation would clearly still hold. Thus equation (1.12) is valid for any graph on six vertices. Using a classical determinantal identity due to Jacobi, Tutte proved the following Pfaffian identity for any graph 퐺, and any four vertices 푥, 푦, 푧, 푤 of 퐺: ± 푃 푥 푦 푃 푧푤 ± 푃 푥푧 푃 푦푤 ± 푃 푥푤 푃 푦푧 = 푃 푃 푥 푦푧푤 .

(1.13)

Because of the ambiguity of the values of signs, some explanation is in order. What equation (1.13) means is that 푃 푃 푥 푦푧푤 is equal to one of the eight polynomials obtained from the expression on the left hand side by specific valuations of the three signs. (See Exercise 1.5.5 for a proof of this identity.) The ambiguities of signs arise because this Pfaffian identity is obtained from three determinantal identities by taking square roots, and the Pfaffian of a skew-symmetric matrix is one of the

20

1 Perfect Matchings

two square roots of its determinant. The precise values of the signs turned out to be irrelevant for the application Tutte had in his mind. The ambiguity of the values of the signs was resolved in 1966 by Halton (1966, [42]). The following lemma, which plays a crucial role in Tutte’s proof of Theorem 1.3, is a consequence of the above identity. (Recall that a graph is matchable if it has a perfect matching.) Lemma 1.11 Let 퐺 be a graph which is not matchable, and let 푥, 푦, 푧 and 푤 be any four vertices of 퐺. If neither 퐺 − 푥 − 푧 nor 퐺 − 푦 − 푧 is matchable, then at least one of 퐺 − 푥 − 푦 and 퐺 − 푧 − 푤 is also not matchable.

Proof Let A be the matrix of indeterminates associated with 퐺. Since 퐺 is not matchable, the Pfaffian 푃 of 퐺 is zero. Hence the right hand side of equation (1.13) is zero. Also, as 퐺 − 푥 − 푧 and 퐺 − 푦 − 푧 are not matchable, both 푃 푥푧 and 푃 푦푧 are zero. By equation (1.13) it now follows that at least one of 푃 푥 푦 and 푃 푧푤 must be zero. Hence, by Proposition 1.10, at least one of 퐺 − 푥 − 푦 and 퐺 − 푧 − 푤 is not matchable. ƒ

1.5.3 Tutte’s original proof of his theorem In Chapter 1, Tutte’s Theorem 1.3, in the proof of statement 1.3.1, we adopted the approach of Lov´asz (1975, [56]). In that proof, we consider a graph 퐺 ∗ , and the four vertices 푤, 푥, 푦 and 푧 of 퐺 ∗ . It is assumed that 퐺 ∗ is not matchable, and that 푥푧 and 푦푧 are edges of 퐺 ∗ , but neither 푥푦 nor 푤푧 are edges of 퐺 ∗ . It is then shown that at least one of the graphs 퐺 ∗ + 푥푦 and 퐺 ∗ + 푤푧 is not matchable. This result is equivalent to the statement of Lemma 1.11.

Exercises 1.5.1 Show that the determinant of an odd order skew-symmetric matrix is zero. (Hint: For any skew-symmetric matrix A, the matrix obtained from A by multiplying each of its columns by −1 is the transpose of A.)

1.5.2 Let 퐺 be a graph which is not necessarily simple, and let 퐷 be an orientation of 퐺 such that, for any adjacent pair {푢, 푣} of vertices, all edges joining 푢 and 푣 are oriented the same way, that is, either 푢 is the tail of all the corresponding arcs in 퐷, or 푢 is their head. In this case, in the adjacency matrix A = (푎 푢푣 ) of 퐷, we set 푎 푢푣 = 0, if 푢 and 푣 are not adjacent in 퐺, and 푎 푢푣 is the number of edges joining 푢 to 푣, or is the negative of that number, according to whether 푢 is the tail, or is the head, of the corresponding arcs in 퐷. The digraph shown in Figure 1.9 is such an orientation of a graph on four vertices. (It happens to be a Pfaffian orientation of its underlying graph.) Compute the determinant and Pfaffian of the adjacency matrix shown in Figure 1.9 and verify that the number of perfect matchings in the graph is the absolute value of the Pfaffian.

1.5 The Magic of Pfaffians

21 1

4

2

3

 0 2 −2 −1     −2 0 1 −1     2 −1 0 −3     1 1 3 0  

Fig. 1.9 A digraph and its adjacency matrix

1.5.3 Write down the expansions of 푃, 푃12 , 푃34 , 푃13 , 푃24 , 푃14 , 푃23 , and 푃1234 corresponding to the 6 × 6 matrix A shown in equation (1.10), and verify the identity equation (1.12). 1.5.4 (The cofactor matrix) Let A = (푎 푖 푗 ) be an 푛 × 푛 matrix.For 1 ≤ 푖, 푗 ≤ 푛, let A[푖, 푗] denote the matrix obtained from A by deleting the 푖 푡 ℎ row and 푗 푡 ℎ column. (This is not to be confused with A푖 푗 , which is the matrix obtained from A by deleting the 푖 푡 ℎ and 푗 푡 ℎ rows and the 푖 푡 ℎ and 푗 푡 ℎ columns.) The cofactor matrix C = (푐 푖 푗 ) is defined by 푐 푖 푗 := (−1) 푖+ 푗 det A[푖, 푗]. (Thus the adjugate Adj(A) of A is the transpose of the 푛 × 푛 matrix C := (푐 푖 푗 ).) Verify that if A is a skew-symmetric matrix of even order then C is a skew-symmetric matrix by showing that: (i) 푐 푖푖 = 0, for 1 ≤ 푖 ≤ 푛, and (ii) 푐 푖 푗 = −푐 푗푖 , for 1 ≤ 푖 < 푗 ≤ 푛. 1.5.5 (Jacobi’s Identity) Let A = (푎 푖 푗 ) be a skew-symmetric matrix of even order and let C be its cofactor matrix (defined in Exercise 1.5.4). Suppose that 푅 = {1, 2, . . . , 푟}, where 푟 is even, and let B denote the matrix obtained from the identity matrix I푛 by replacing its first 푟 columns by the first 푟 rows of C. (i) By considering the matrix AB, and using the fact that AC푡 = det(A)I푛 , show that det(A) det(C푅 ) = (det(A)) 푟 det(A푅 ), (1.14) where 푅 = {1, 2, . . . , 푛} − 푅. (For clarity of visualization, we have stated this identity taking 푅 to be {1, 2, . . . , 푟}, but it can be seen that it holds for any 푟-subset 푅 of {1, 2, . . . , 푛}.) (ii) Considering the special case in which A is the matrix of indeterminates associated with an orientation of the complete graph 퐾푛 , and 푅 is equal to {푥, 푦, 푧, 푤}, deduce that: (푐 푥 푦 푐 푧푤 − 푐 푥푧 푐 푦푤 + 푐 푥푤 푐 푦푧 ) 2 = (det(A)) 3 det(A 푥 푦푧푤 ).

22

1 Perfect Matchings

∗ 1.5.6 (i) By taking 푅 to be {푖, 푗 } in equation (1.14), deduce that 푐 푖 푗 = ±푃 푃푖 푗 . (ii) Now taking 푅 to be {푥, 푦, 푧, 푤}, and applying equation (1.14) deduce the identity stated in equation (1.13).

1.6 Notes We defer to the treatises by Lov´asz and Plummer [59], and Schrijver (2003, [84]) for an account of the history of matching theory. Efficient algorithms related to matchings can be found in the books by Schrijver [84], by Cormen, Leiserson, Rivest and Stein (2009, [25]) and also in the book by Bondy and Murty [3]. Tutte’s Theorem 1.3 was clearly a leap from Petersen’s Theorem 1.4 on cubic graphs and Hall’s Theorem 1.5 on bipartite graphs. As such it was received enthusiastically. As his original proof used Pfaffian identities, some authors felt that it was unnatural to use such algebraic machinery to prove a graph theoretical statement; they argued that graphs were simple ‘geometric’ objects, and it ought to be possible to prove statements concerning them by manipulations of diagrams that could be visualized. Alternative proofs of Theorem 1.3 which did not depend on properties of Pfaffians were discovered. But, as it turned out, Pfaffians continued, and continue, to play an important role in matching theory. We will present an account of some these developments in Part III. As noted in the introduction to this chapter, a perfect matching of a graph 퐺 is the edge set of a spanning 1-regular subgraph of 퐺, or, to use Tutte’s terminology, a 1-factor of 퐺. Having been able to characterize graphs which have 1-factors, Tutte was led to the more general problem of characterizing graphs which have spanning subgraphs with specified degree sequences, which he called 푓 -factors. An account of Tutte’s work, and of subsequent developments in this area, can be found in Lov´asz and Plummer [59].

Chapter 2

Matching Covered Graphs

Contents 2.1 2.2

2.3 2.4 2.5 2.6

Matchable Edges in Matchable Graphs . . . . . . . . . . . . . . . . . . . . . . . . Matching Covered Graphs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2.1 Connectivity of matching covered graphs . . . . . . . . . . . . . . 2.2.2 Cubic matching covered graphs . . . . . . . . . . . . . . . . . . . . . . . 2.2.3 Three special matching covered graphs . . . . . . . . . . . . . . . . Bipartite Matching Covered Graphs . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3.1 Dulmage-Mendelsohn decomposition . . . . . . . . . . . . . . . . . The Operation of Splicing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Families of Matching Covered Graphs . . . . . . . . . . . . . . . . . . . . . . . . . Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

23 26 26 27 27 29 30 31 34 39

2.1 Matchable Edges in Matchable Graphs This book is primarily concerned with matching covered graphs which were briefly introduced at the end of Section 1.3. We begin our exploration of their rich theory with a study of notions and statements applicable to the more general class of matchable graphs. By definition, a graph is matchable if it has a perfect matching. In any such graph 퐺, by Tutte’s Theorem 1.3, 표(퐺 − 푆) ≤ |푆|, for all 푆 ⊆ 푉. Barriers in matchable graphs A subset 퐵 of the vertex set of a matchable graph 퐺 is a barrier of 퐺 if 표(퐺 − 퐵) = |퐵|. The empty set and all singletons are barriers in any matchable graph. We shall refer to a barrier of cardinality at most one as a trivial barrier. The study of properties of barriers plays an important role in this theory.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7_2

23

2 Matching Covered Graphs

24

The graph 퐺 shown in Figure 2.1 has several nontrivial barriers. For example the set 퐵 := {푢 1 , 푢 2 , 푢 3 } is a barrier of cardinality three as 퐺 − 퐵 has three odd components (and one even component). The two sets {푢 1 , 푢 2 , 푢 3 , 푢 4 , 푢 5 } and {푣 7 , 푣 8 , 푣 9 } are also nontrivial barriers. 푣3 푣1

푣5

푣2

푣6

푣4

푣7

푣8

푣9

퐵 푢1

푢2

푢3

푢4

푢5

Fig. 2.1 Barriers in a matchable graph

Matchable edges An edge 푒 of a graph 퐺 is matchable in 퐺 if there is some perfect matching of 퐺 that contains it, and is unmatchable otherwise. Corollary 2.2 provides a characterization of matchable edges in a matchable graph. We derive that corollary from the following more general statement which will be of use in the next chapter. Theorem 2.1 Let 푢 and 푣 be any two vertices in a matchable graph 퐺. Then the graph 퐺 − 푢 − 푣 is matchable if and only if there is no barrier of 퐺 which contains both 푢 and 푣. Proof Firstly suppose that 퐺 − 푢 − 푣 is matchable. We wish to show that no subset of 푉 containing both 푢 and 푣 is a barrier of 퐺. Towards this end, let 퐵 be any subset of 푉 which includes both 푢 and 푣. By setting 푆 := 퐵 − 푢 − 푣, we have: 표(퐺 − 퐵) = 표((퐺 − 푢 − 푣) − 푆) ≤ |푆| (because 퐺 − 푢 − 푣 is matchable) = |퐵| − 2, implying that 표(퐺 − 퐵) < |퐵|, and hence that 퐵 is not a barrier of 퐺.

Conversely, suppose that 퐺 − 푢 − 푣 is not matchable. Then, by Theorem 1.3, and Exercise 1.2.1, there is a subset 푆 of 푉 (퐺 − 푢 − 푣) such that 표(퐺 − 푢 − 푣 − 푆) ≥ |푆| + 2.

(2.1)

Now, letting 퐵 := 푆 + 푢 + 푣, we have: 표(퐺 − 퐵) ≥ |퐵|.

(2.2)

2.1 Matchable Edges in Matchable Graphs

25

By hypothesis, 퐺 itself is matchable. Thus, by Theorem 1.3, strict inequality cannot hold in (2.2). Therefore 표(퐺 − 퐵) = |퐵|, implying that 퐵 is a barrier of 퐺 which contains both 푢 and 푣. ƒ If 푒 = 푢푣 is an edge of 퐺, then 푒 is matchable in 퐺 if and only if 퐺 − 푢 − 푣 is matchable. Thus, as a consequence of the above theorem, we have: Matchable edges and barriers Corollary 2.2 An edge of a matchable graph is matchable if and only if there is no barrier of the graph which contains both ends of that edge. ƒ The following proposition is a simple consequence of the above corollary. Proposition 2.3 Suppose that 퐵 is a barrier of a matchable graph 퐺 and that 퐾 is an even component of 퐺 − 퐵. Then any edge 푢푣 of 퐺 with 푢 ∈ 퐵 and 푣 ∈ 푉 (퐾) is unmatchable in 퐺. Proof Consider the set 퐵′ := 퐵 + 푣. Clearly, each odd component of 퐺 − 퐵 is also an odd component of 퐺 − 퐵′ . In addition, any odd component of 퐾 − 푣 is also an odd component of 퐺 − 퐵′ . There must be at least one odd component of 퐾 − 푣 because 퐾 is an even component of 퐺 − 퐵. Thus 표(퐺 − 퐵′ ) ≥ |퐵| + 1 = |퐵′ |. However, as 퐺 is a matchable graph, equality must hold implying that 퐵′ is a barrier of 퐺. Since both ends of the edge 푢푣 belong to this barrier, it follows from Corollary 2.2 that 푢푣 is unmatchable. ƒ Example 2.4 Consider the barrier 퐵 := {푢 1 , 푢 2 , 푢 3 } of the matchable graph 퐺 shown in Figure 2.1. The subgraph of order two induced by {푣 1 , 푣 2 } is an even component of 퐺 − 퐵. Adding any vertex of this even component of 퐺 − 퐵 to 퐵 yields another barrier of 퐺. The two edges 푢 1 푣 1 , 푢 2 푣 1 have both their ends in the barrier 퐵 + 푣 1 , and the two edges 푢 1 푣 2 and 푢 3 푣 2 have both their ends in the barrier 퐵 + 푣 2 . Thus the four edges 푢 1 푣 1 , 푢 2 푣 1 , 푢 1 푣 2 , and 푢 3 푣 2 are unmatchable in 퐺.

Exercises 2.1.1 Find all the matchable edges in the graph shown in Figure 2.1. ⊲2.1.2 Describe a polynomial-time algorithm to determine whether an edge 푣푤 of a graph 퐺 is matchable. Hint: determine whether the graph 퐻 := 퐺 − 푣 − 푤 is matchable; if a perfect matching 푀 of 퐺 is known, the complexity is reduced by considering the restriction of 푀 to 퐻. Special barriers A barrier 퐵 in a matchable graph 퐺 is a special barrier if 퐺 − 퐵 has precisely one nontrivial odd component. 2.1.3 Show that in the graph 퐺 obtained by deleting an edge from the Petersen graph every nontrivial barrier is a special barrier.

2 Matching Covered Graphs

26

2.2 Matching Covered Graphs Adopting the terminology introduced in Section 2.1 we may now define a matching covered graph as follows: Matching covered graphs A matching covered graph is a nontrivial connected graph in which every edge is matchable. The complete graph 퐾2 is the smallest matching covered graph. A plethora of examples of graphs and families of graphs which play important roles in this theory can be found at the end of this section and in Sections 2.4 and 2.5. Thus matching covered graphs are special types of matchable graphs. We make use of this fact implicitly in establishing statements concerning matching covered graphs. A subset of the edge set of a graph 퐺 is a perfect matching of 퐺 if and only it is a union of perfect matchings of the components of 퐺. Also, clearly, the set of perfect matchings of 퐺 is the same as the set of perfect matchings of the graph obtained from 퐺 by deleting all unmatchable edges. For this reason, the theory of matching covered graphs is the natural setting for studying problems such as counting or packing or covering problems involving perfect matchings. Corollary 2.2 implies the following distinguishing property of the class of matching covered graphs. (Recall that a subset 푆 of the vertex set of a graph 퐺 is stable if no edge of 퐺 has both its ends in 푆.) Theorem 2.5 A connected matchable graph is matching covered if and only if each of its barriers is a stable set. ƒ Hall’s Theorem 1.5 can be used to derive a characterization of matching covered graph that is specific to bipartite graphs, see Section 2.3.

2.2.1 Connectivity of matching covered graphs If 퐵 is a barrier of a matchable graph 퐺 then, by definition, the subgraph 퐺 − 퐵 has |퐵| odd components. The following proposition asserts that, if 퐺 is matching covered, then 퐺 − 퐵 cannot have any even components. Proposition 2.6 For each nonempty barrier 퐵 of a matching covered graph 퐺, each component of 퐺 − 퐵 is odd. Proof Assume that 퐺 − 퐵 has an even component, say 퐾. The graph 퐺, being matching covered, is connected. Thus, 퐺 has an edge 푢푣 with 푢 ∈ 퐵 and 푣 ∈ 푉 (퐾). By Proposition 2.3, edge 푢푣 is unmatchable in 퐺. This is absurd because 퐺 is matching covered. We conclude that every component of 퐺 − 퐵 is odd. ƒ

2.2 Matching Covered Graphs

27

The following assertion is a basic property of matching covered graphs: Corollary 2.7 Every matching covered graph on four or more vertices is 2-connected. Proof Let 푣 be any vertex of a matching covered graph 퐺. Then 퐺 − 푣 has odd components. As 퐺 is matchable, it has precisely one odd component. By Proposition 2.6, 퐺 − 푣 has no even component. Therefore 퐺 − 푣 has just one component, implying that 푣 is not a cut vertex of 퐺. It follows that if the order of 퐺 is at least four, then 퐺 ƒ is 2-connected.

2.2.2 Cubic matching covered graphs As noted at the end of Section 1.3, Tutte [89] deduced from his Theorem 1.3 that every 2-connected cubic graph is matching covered. Here we present a proof which is based on the characterization of matching covered graphs given by Theorem 2.5. Cubic matching covered graphs Theorem 2.8 Every 2-connected cubic graph is matching covered.

Proof Let 퐺 be a 2-connected cubic graph and let 푆 be any subset of 푉. Then, as in the proof of Theorem 1.4, the inequality 3 ≤ |휕 (퐾)| holds for each odd component 퐾 of 퐺 − 푆. We thus have: 3 표(퐺 − 푆) ≤ |휕 (푆)| = 3|푆| − 2|퐸 (퐺 [푆])| ≤ 3|푆|.

(2.3)

We deduce that 표(퐺 − 푆) ≤ |푆|. This conclusion holds for each subset 푆 of 푉. By Tutte’s Theorem 1.3, 퐺 is matchable. Moreover, from equation (2.3), we deduce that 표(퐺 − 푆) = |푆| only if 푆 is stable. This conclusion holds for each subset 푆 of 푉. Thus, every barrier of 퐺 is stable. The assertion now follows from Theorem 2.5. ƒ A similar argument shows that, for 푟 ≥ 2, any 푟-graph (as defined in Exercise 1.3.6) is matching covered (Exercise 2.2.3).

2.2.3 Three special matching covered graphs Even cycles and wheels (Exercise 2.2.4) are some examples of matching covered graphs which shall come across frequently. However, most of the graphs which play prominent roles in this theory are 3-connected cubic graphs. Figure 2.2 shows three such graphs (퐾4 , the complete graph on four vertices, 퐶6 , the triangular prism, and P, the Petersen graph) which play special roles in our theory. (These three graphs feature in the conjecture of Lov´asz mentioned in the preface.) The operation of

2 Matching Covered Graphs

28

(a) 퐾4

(b) 퐶6

(c) P

Fig. 2.2 The trio of important cubic matching covered graphs

splicing which can be used to ‘combine’ two matching covered graphs to obtain a larger matching covered graph is described in Section 2.4.

Exercises ⊲2.2.1 Let 푅 be the set of unmatchable edges of a graph 퐺. Show that every nontrivial component of 퐺 − 푅 is matching covered and show that 퐺 is matchable if and only if each component of 퐺 − 푅 is nontrivial. ⊲2.2.2 Describe a polynomial-time algorithm, which, given a graph 퐺, determines whether 퐺 is matching covered. Hint: check in linear-time whether 퐺 is connected, determine in polynomial-time a maximum matching of 퐺 and, if 퐺 is matchable then, for each edge 푣푤 ∈ 퐸, determine in polynomial-time whether 푣푤 is matchable (Exercise 2.1.2.). 2.2.3 Show that any 푟-graph (defined in Exercise 1.3.6), with 푟 ≥ 2, is matching covered. Wheels For an integer 푘 ≥ 3, the wheel 푊 푘 is the graph obtained from a cycle 푅 of length 푘 by adding a new vertex ℎ and joining it to all vertices of 푅. Thus 푊 푘 is a graph on 푘 + 1 vertices. (When a graph is denoted by a symbol with a subscript, that subscript usually indicates the order of the graph. However, here we are deviating from this general convention so as to keep our notation for wheels consistent with the one in standard use.) The cycle 푅 is the rim of 푊 푘 , the vertex ℎ is its hub, and the edges incident with ℎ are its spokes. Clearly, the degree of the hub of 푊 푘 is 푘. We note that 푊3 is isomorphic to the complete graph 퐾4 , and any one of its vertices may be regarded as its hub. For all 푘 ≥ 4, the wheel 푊 푘 has a unique hub. An odd wheel is a wheel with an odd number of spokes.

2.3 Bipartite Matching Covered Graphs

29

2.2.4 Consider now a wheel 푊 푘 , where 푘 is odd. (i) Show that every spoke of 푊 푘 lies in precisely one perfect matching of 푊 푘 . (ii) Show that 푊 푘 has precisely 푘 perfect matchings. (iii) Show that every edge in the rim of 푊 푘 lies in precisely (푘−1)/2 perfect matchings of 푊 푘 . Conclude that 푊 푘 is matching covered. (iv) Prove that, if 푘 ≥ 5, and 푒 is any edge of 푊 푘 , then 푊 푘 − 푒 is matching covered if and only if 푒 is a spoke of 푊 푘 . ∗ 2.2.5 Prove that if 퐶 := 휕 ( 푋) is a cut of a matching covered graph 퐺 and 퐺 [푋] has an odd cycle that is hamiltonian then the graph 퐺 1 , obtained from 퐺 by contracting 푋 to a single vertex 푥 and removing all resulting loops, is matching covered. ∗ 2.2.6 Let 퐺 [ 퐴, 퐵] be a bipartite graph and let 푀 be a perfect matching of 퐺. Let 퐻 be the graph obtained from 퐺 by the addition of a parallel edge to each edge of 푀. Let 퐷 be the directed graph obtained from 퐻 by directing every edge of 퐸 (퐻) − 푀 from 퐴 to 퐵 and by directing every edge of 푀 from 퐵 to 퐴. (i) Show that 퐺 is matching covered if and only if 퐷 is strongly connected. Hint: a digraph is strongly connected if and only if it is connected and each one of its edges is in a directed cycle. (ii) Design a linear-time algorithm to decide if a given bipartite graph is matching covered, assuming as input the bipartite graph 퐺 and a perfect matching 푀 of 퐺.

2.3 Bipartite Matching Covered Graphs

A Hall-type characterization of bipartite matching covered graphs Theorem 2.9 Let 퐺 := 퐺 [ 퐴, 퐵] be a bipartite graph of order four or more, such that | 퐴| = |퐵|. The graph 퐺 is matching covered if and only if |푁 (푆)| ≥ |푆| + 1, for all subsets 푆 of 퐴 such that 1 ≤ |푆| ≤ | 퐴| − 1. Proof Suppose that there exists a nonempty proper subset 푆 of 퐴 such that |푁 (푆)| ≤ |푆|. Then, every edge (if any) joining 푁 (푆) with 퐴 − 푆 is not matchable. If there were no such edge then 퐺 would not be connected, and consequently, not matching covered. In any case, 퐺 is not matching covered. Now, suppose that |푁 (푆)| ≥ |푆| + 1, for all nonempty proper subsets 푆 of 퐴. Let 푎푏 be an edge of 퐺, where 푎 ∈ 퐴 and 푏 ∈ 퐵. Then, in 퐺 − 푎 − 푏, we have |푁 ( 푋)| ≥ | 푋 |, for all 푋 ⊆ 퐴 − 푎. By Hall’s Theorem, 퐺 − 푎 − 푏 is matchable. It follows that 푎푏 is matchable in 퐺. This conclusion holds for each edge 푎푏 of 퐺. Moreover, 퐺 is matchable. Finally, for any connected component 퐾 of 퐺, as 퐺 is matchable we have that | 퐴 ∩ 푉 (퐾)| = |퐵 ∩ 푉 (퐾)|. As |푁 (푆)| ≥ |푆| + 1 for every nonempty proper subset 푆 of 퐴, it follows that 퐴 ∩ 푉 (퐾) = 퐴, hence 퐵 ∩ 푉 (퐾) = 퐵. Thus, 퐺 = 퐾, and 퐺 is connected. We conclude that 퐺 is matching covered. ƒ

2 Matching Covered Graphs

30

We leave the proof of the next assertion as Exercise 2.3.1. Corollary 2.10 A matchable bipartite graph 퐺 := 퐺 [ 퐴, 퐵] is not matching covered if and only if there exists a subset 푆 of 퐴 such that 1 ≤ |푆| ≤ | 퐴| −1 and |푁 (푆)| = |푆|. ƒ

2.3.1 Dulmage-Mendelsohn decomposition In analyzing the properties of a given matching covered graph we are sometimes led to considering the properties of a related matchable graph. For example, if 퐺 is a matching covered graph with at least two edges, and 푒 is an edge of 퐺, then the graph 퐺 − 푒 may not be matching covered, but it is matchable because any perfect matching of 퐺 that does not include the edge 푒 is a perfect matching of 퐺 − 푒. If 퐵 is a barrier of 퐺 − 푒 then either 퐵 is also a barrier of 퐺 or the two ends of the edge 푒 lie in different odd components of 퐺 − 푒 − 퐵. We conclude this section by noting a fundamental property of matchable bipartite graphs which we shall have occasion to use in future chapters. For this, we need the following auxiliary result, whose proof is left as Exercise 2.3.2. Proposition 2.11 In any matchable bipartite graph 퐺 [ 퐴, 퐵] there exists a nonempty subset 푆 of 퐴, possibly equal to 퐴, such that the subgraph of 퐺 induced by 푆 ∪ 푁 (푆) is matching covered. (Likewise, there is a nonempty subset 푇 of 퐵 such that the subgraph induced by 푁 (푇) ∪ 푇 is matching covered.) See Figure 2.3 for an illustration. ƒ 퐵 − 푁 (푆 )

푁 (푆 )

퐴 − 푆

푆

Fig. 2.3 The subgraph of 퐺 [ 퐴, 퐵] induced by 푆 ∪ 푁 (푆 ) is matching covered

The Dulmage-Mendelsohn decomposition A Dulmage-Mendelsohn decomposition of a matchable bipartite graph 퐺 := 퐺 [ 퐴, 퐵] consists of a partition ( 퐴1 , . . . , 퐴푟 ) (푟 ≥ 1) of 퐴 and a partition (퐵1 , . . . , 퐵푟 ) of 퐵 such that

2.4 The Operation of Splicing

31

(i) for 푖 = 1, 2, . . . , 푟, the subgraph 퐺 [ 퐴푖 ∪ 퐵푖 ] of 퐺 induced by 퐴푖 ∪ 퐵푖 is matching covered, and (ii) if 푎 푖 푏 푗 , with 푎 푖 ∈ 퐴푖 and 푏 푗 ∈ 퐵 푗 , is an edge of 퐺, then 푖 ≥ 푗. Proposition 2.11 is a special case of the following more general assertion. Theorem 2.12 (Dulmage and Mendelsohn (1958, [28])) Every matchable bipartite graph has a Dulmage-Mendelsohn decomposition. We leave the proof of Theorem 2.12 as Exercise 2.3.3.

Exercises ∗ 2.3.1 Prove Corollary 2.10. ∗ 2.3.2 Deduce Proposition 2.11. ∗ 2.3.3 Prove Theorem 2.12. Hint: use Proposition 2.11 and apply induction.

2.4 The Operation of Splicing Let 퐺 and 퐻 be two vertex-disjoint graphs and let 푢 and 푣 be vertices of 퐺 and 퐻, respectively, such that 푑퐺 (푢) = 푑 퐻 (푣). In addition, let 휃 be a given bijection between the set 휕퐺 (푢) of edges incident with 푢 in 퐺 and the set 휕퐻 (푣) of edges incident with 푣 in 퐻. We define (퐺 ⊙ 퐻)푢,푣, 휃 to be the graph obtained from the union of 퐺 − 푢 and 퐻 − 푣 by joining, for each edge 푒 in 휕퐺 (푢), the end of 푒 in 퐺 belonging to 푉 (퐺) − 푢 to the end of 휃 (푒) in 퐻 belonging to 푉 (퐻) − 푣; and refer to (퐺 ⊙ 퐻)푢,푣, 휃 as the graph obtained by splicing 퐺 at 푢, with 퐻 at 푣, with respect to the bijection 휃. Figure 2.4 illustrates this operation, with the bijection 휃 between 휕퐺 (푢) and 휕퐻 (푣), where 휃 (푒 푖 ) = 푓푖 , for 푖 = 1, 2, . . . , 5. It is not always necessary to use the cumbersome notation (퐺 ⊙ 퐻)푢,푣, 휃 for describing a splicing of two graphs 퐺 and 퐻; in most cases of interest either the parameters (푢, 푣, and 휃) are implicit, or the statements made about the graph resulting from splicing does not depend on the parameters. We leave verification of the following assertion as Exercise 2.4.1. When both 퐺 and 퐻 happen to be simple graphs, the bijection 휃 in the definition of splicing may be regarded just as a bijection between the neighbour set 푁퐺 (푢) of 푢 in 퐺 and the neighbour set 푁 퐻 (푣) of 푣 in 퐻, and (퐺 ⊙ 퐻)푢,푣, 휃 may be defined as the graph obtained from the union of 퐺 − 푢 and 퐻 − 푣 by joining each vertex 푥 in 푁퐺 (푢) to the vertex 휃 (푥) in 푁 퐻 (푣). Both the pentagonal prism and the Petersen graph may be obtained by splicing two copies of the 5-wheel at their hubs under suitable bijections between the vertex sets of their rims (see Exercise 2.4.2).

2 Matching Covered Graphs

32

푒1 푓1

푒2

푓2

푒3

푢

푓3

푣

푒4

푓4 푓5 푒5

(푎)

(푏)

(푐)

Fig. 2.4 (a) 퐺; (b) 퐻; (c) (퐺 ⊙ 퐻 )푢,푣, 휃

As the example mentioned above shows, in general the result of splicing two graphs 퐺 and 퐻 depends on the choices of 푢, 푣, 휃. However, if 퐺 is a vertextransitive cubic graph, then the result of splicing 퐺 with 퐻 = 퐾4 does not depend, up to isomorphism, on the choices of 푢, 푣, and 휃, and we denote it simply by 퐺 ⊙ 퐾4 . More generally, for any cubic graph 퐺, the result of splicing 퐺 and 퐾4 depends, up to isomorphism, only on the orbit of the automorphism group of 퐺 to which 푢 belongs (and the choices of 푣 and 휃 are immaterial); and we denote it simply by (퐺 ⊙ 퐾4 )푢 . For example, since both 퐾4 and the triangular prism 퐶6 are vertex-transitive, there is only one way of splicing 퐾4 with itself or with 퐶6 . The result of splicing 퐾4 with itself is clearly 퐶6 . The result of splicing 퐶6 with 퐾4 is shown in Figure 2.5(c). The automorphism group of this graph has three different orbits. The result of splicing a 퐾4 with it at this vertex is shown in Figure 2.5(d). We refer to the two graphs in Figure 2.5(c) and (d), respectively, as the bicorn and the tricorn as they resemble, in our imagination, to the two-cornered and three-cornered hats worn by pirates. We denote the bicorn by H8 and the tricorn by H10 . Splicing two matching covered graphs 2.13 Any graph obtained by splicing together two matching covered graphs is also matching covered. ƒ

Exercises ∗ 2.4.1 Give a proof of statement 2.13. ∗ 2.4.2 Consider two copies 퐺 and 퐻 of the 5-wheel 푊5 , where 푢 and 푣 denote, respectively, the hubs in the two copies. Describe bijections 휃 and 휙 between 푁퐺 (푢) and 푁 퐻 (푣) such that (퐺 ⊙ 퐻)푢,푣, 휃 is the pentagonal prism, and (퐺 ⊙ 퐻)푢,푣, 휙 is

33

2.4 The Operation of Splicing

(푎)

(푐)

(푏)

(푑)

Fig. 2.5 (a) 퐾4 ; (b) 퐶6 ; (c) H8 , the bicorn; (d) H10 , the tricorn

the Petersen graph. Show that all such splicings are 3-edge-colourable, with the exception of the Petersen graph. ⊲2.4.3 Show that, up to isomorphism, there are four graphs that are obtainable by the splicing of two 5-wheels at their hubs. ⊲2.4.4 Let 퐺 be a graph, let 푢 and 푤 be two vertices of 퐺, both having degree three. Show that if there is an automorphism 휃 of 퐺 that maps 푢 to 푤 then the graph obtained by splicing 퐺 and 퐾4 at 푢 is isomorphic to the one obtained by splicing 퐺 and 퐾4 at 푤. Conclude that, up to isomorphism, there is only one graph that can be obtained by splicing a vertex-transitive graph 퐺 to 퐾4 . ⊲2.4.5 Find all the graphs, up to isomorphism, that can be obtained by splicing the bicorn with 퐾4 . ∗ 2.4.6 Given any family G푛 of cubic graphs of order 푛, let G푛+2 denote the family of cubic graphs of order 푛 + 2 (up to isomorphism) that can be obtained by splicing members of G푛 with 퐾4 . Taking G4 to be {퐾4 }, determine all members of G6 , G8 , and G10 . ⊲2.4.7 The Meredith graph, shown in Figure 2.6, is obtained from the 4-graph in Figure 1.5 by splicing a 퐾4,4 at each of its vertices as indicated. Show that this graph is not 4-edge-colourable. (For 푟 ≥ 4, similar ideas can be used to construct 푟-connected 푟-graphs which are not 푟-edge-colourable.)

2 Matching Covered Graphs

34

Fig. 2.6 The Meredith graph

2.5 Families of Matching Covered Graphs ♯ In addition to wheels (see Exercise 2.2.4), there are several other families of matching covered graphs which play important roles in this theory. We introduce some of them in this section. These families play significant roles in the works of McCuaig [68] and Norine and Thomas [78], which are described in Chapter 18. Prisms The prism P2푛 , 푛 ≥ 3, is the graph obtained from two disjoint cycles 푢 1 푢 2 . . . 푢 푛 푢 1 and 푣 1 푣 2 . . . 푣 푛 푣 1 of length 푛 by the addition of the 푛 edges 푢 푖 푣 푖 , 푖 = 1, 2, . . . , 푛. Prisms are so named because they are the graphs associated with the familiar geometrical objects known as ‘prisms’. For this reason, we have special names for those of small orders. Thus P6 , which is the same as 퐶6 , is the triangular prism; P8 is the cube; and P10 , shown in Figure 2.7(a), is the pentagonal prism. The family of all prisms is denoted by P. ¨ Mobius ladders The M¨obius ladder M2푛 , 푛 ≥ 2, is the graph obtained from a cycle 푣 1 푣 2 . . . 푣 2푛 푣 1

35

2.5 Families of Matching Covered Graphs

of length 2푛 by the addition of the 푛 chords 푣 푖 푣 푖+푛 , 1 ≤ 푖 ≤ 푛, joining antipodal pairs of vertices of the cycle. All M¨obius ladders are embeddable in the projective plane. The two graphs M4 , and M6 are isomorphic to 퐾4 and 퐾3,3 , respectively. Figure 2.7(b) shows the M¨obius ladder M10 . The family of all M¨obius ladders is denoted ML. (We are using ML to denote the family of M¨obius ladders since M is reserved for the set of all perfect matchings of a graph.) 푢1

푣1 푣2

푣10 푣1 푢5

푢2

푣9

푣3

푣8

푣4

푣2

푣5 푣4

푣3

푣7 푢4

푢3

푣5 푣6 (푏)

(푎) Fig. 2.7 (a) prism P10 ; (b) M¨obius ladder M10

Prisms and M¨obius ladders may be defined alternatively as supergraphs of ladders, which are defined below. Ladders For 푛 ≥ 2, the ladder L2푛 is the graph obtained from two disjoint paths 푢 1 푢 2 . . . 푢 푛 and 푣 1 푣 2 . . . 푣 푛 of length 푛 − 1 by the addition of the 푛 rungs 푢 푖 푣 푖 , 푖 = 1, 2, . . . , 푛. Figure 2.8(a) shows the ladder L8 . Clearly, the prism P2푛 , 푛 ≥ 3, is just the planar graph obtained from L2푛 by adding the two edges 푢 1 푢 푛 and 푣 1 푣 푛 . Although it is less obvious, it can be shown that the M¨obius ladder M2푛 , 푛 ≥ 3 is isomorphic to the graph obtained from L2푛 by adding the two edges 푢 1 푣 푛 and 푣 1 푢 푛 (see Exercise 2.5.1). This explains the ‘ladder’ part in their names.

2 Matching Covered Graphs

36 푢1 푢1

푥

푣1 푣1

푢2

푣2

푢3

푢2

푣2

푢3

푣3

푣3 푣4

푢4

푣4

푦

푢4

(푎)

(푏)

Fig. 2.8 (a) Ladder L8 ; (b) Staircase S10

Staircases Consider the ladder L2푛−2 obtained from two disjoint paths 푢 1 푢 2 . . . 푢 푛−1 and 푣 1 푣 2 . . . 푣 푛−1 by adding, for 1 ≤ 푖 ≤ 푛 − 1, an edge joining 푢 푖 and 푣 푖 . For 푛 ≥ 3, the staircase S2푛 is the graph obtained from L2푛−2 by adding two new vertices 푥 and 푦, and then joining 푥 to 푢 1 and 푣 1 , 푦 to 푢 푛−1 and 푣 푛−1 , and 푥 and 푦 to each other. See Figure 2.8(b) for a drawing of S10 . The staircase S6 on six vertices is isomorphic to the triangular prism 퐶6 . The staircase S8 on eight vertices is the bicorn H8 . The family of all staircases is denoted ST . Biwheels The biwheel B2푛 , 푛 ≥ 4, is the bipartite graph obtained from a cycle 푣 1 푣 2 . . . 푣 2푛−2 푣 1 of length 2푛 − 2, called the rim of B2푛 , by the addition of two vertices, ℎ1 and ℎ2 , called the hubs of B2푛 , and by the addition of edges ℎ1 푣 1 , ℎ1 푣 3 , . . . , ℎ1 푣 2푛−3 and edges ℎ2 푣 2 , ℎ2 푣 4 , . . . , ℎ2 푣 2푛−2 . The family of biwheels is denoted B. Figure 2.9 shows two drawings of the biwheel B10 where the second drawing depicts it as a ‘wheel’ with two hubs. We note that the cube, apart from being the prism P8 , is also the biwheel B8 . Truncated biwheels The truncated biwheel T2푛 , 푛 ≥ 3, is the graph obtained from a path 푣 1 푣 2 . . . 푣 2푛−2 of length 2푛 − 3, by the addition of two vertices, ℎ1 and ℎ2 , and

37

2.5 Families of Matching Covered Graphs ℎ1 푣1

푣1

푣3

푣5

푣7

푣2

푣8

푣3 ℎ2

ℎ1

푣2

푣6

푣4

푣8

푣7

푣4

푣6

푣5

ℎ2 Fig. 2.9 Two drawings of the biwheel B10

by the addition of edges ℎ1푣 1 , ℎ1 푣 3 , . . . , ℎ1 푣 2푛−3 , edges ℎ2 푣 2 , ℎ2 푣 4 , . . . , ℎ2 푣 2푛−2 and edges ℎ1 푣 2푛−2 and ℎ2 푣 1 . Two drawings of T10 are shown in Figure 2.10. Note that for 푛 ≥ 4, the truncated biwheel T2푛 is isomorphic to the graph obtained from the biwheel B2푛 , by deleting the edge 푣 1 푣 2푛−2 and adding the two edges ℎ1 푣 2푛−2 and ℎ2 푣 1 . We remark that the graph 퐶6 is the truncated biwheel T6 , the triangular prism P6 , as well as the staircase S6 . We denote the family of truncated biwheels by T B. Norine and Thomas [79] refer to truncated biwheels as prismoids. ℎ1 ℎ1

푣1

푣3

푣5

푣7 푣2 푣1

푣2

푣4

푣6

푣6

푣4 푣3

푣5

푣8 ℎ2

ℎ2 Fig. 2.10 Two drawings of the truncated biwheel T10

푣7

푣8

2 Matching Covered Graphs

38

Exercises 2.5.1 Show that the two definitions of the M¨obius ladder M2푛 given above result in the same graph (up to isomorphism). 2.5.2 Show that: (i) the prism P2푛 is bipartite when 푛 is even and nonbipartite when 푛 is odd, (ii) the M¨obius ladder M2푛 is bipartite when 푛 is odd, and nonbipartite when 푛 is even. 2.5.3 (Ladders and Fibonacci Numbers) We denoted by Φ(퐺) the number of perfect matchings of a graph 퐺. Show that: Φ(L2푛 ) = 퐹 (푛 + 1), for 푛 ≥ 2.

(2.4)

(The Fibonacci numbers 퐹 (푛) are the numbers in the sequence defined by the recursion 퐹 (푛) = 퐹 (푛 −1) + 퐹 (푛 −2), for 푛 ≥ 3, with initial values 퐹 (1) = 퐹 (2) = 1. Thus, 퐹 (1) = 1, 퐹 (2) = 1, 퐹 (3) = 2, 퐹 (4) = 3, 퐹 (5) = 5, 퐹 (6) = 8, 퐹 (7) = 13, etc.) ∗ 2.5.4 Using (2.4), deduce the following: (i) for 푛 ≥ 3, Φ(P2푛 ) = (ii) for 푛 ≥ 2, Φ(M2푛 ) =



퐹 (푛 + 1) + 퐹 (푛 − 1) if 푛 is odd 퐹 (푛 + 1) + 퐹 (푛 − 1) + 2 if 푛 is even



퐹 (푛 + 1) + 퐹 (푛 − 1) + 2 if 푛 is odd 퐹 (푛 + 1) + 퐹 (푛 − 1) if 푛 is even

(iii) for 푛 ≥ 2, Φ(S2푛 ) = 퐹 (푛) + 2. 2.5.5 Contraction of a triangle in the staircase S2푛 results in the staircase S2(푛−1) . Using this argument, write a proof that Φ(S2푛 ) = 퐹 (푛) + 2, for 푛 ≥ 2, by induction on the order of S2푛 . 2.5.6 Show that Φ(B2푛 ) = (푛 − 1) 2 . (Hint: Show that any edge incident with hub ℎ1 , and any edge incident with hub ℎ2 are contained together in a unique perfect matching. 2.5.7 Contraction of a triangle in the truncated biwheel T2푛 results in the truncated biwheel T2(푛−1) plus an edge joining the two hubs. Use this argument and induction to deduce that 푛(푛 − 1) + 1, 푛 ≥ 3. Φ(T2푛 ) = 2

39

2.6 Notes

2.5.8 For each even integer 푛 ≥ 6, let 퐴푛 be the graph obtained from a matching {푢 1 푢 2 , 푢 3 푢 4 , . . . , 푢 푛−3 푢 푛−2 }, by adjoining two vertices, ℎ1 and ℎ2 , and then joining ℎ1 to each of 푢 1 , 푢 3 , . . . , 푢 푛−3 by a pair of multiple edges, and ℎ2 to each of 푢 2 , 푢 4 , . . . , 푢 푛−2 by a pair of multiple edges. Figure 2.11 shows a drawing of 퐴10 . 푢1

푢2

푢3

푢4

ℎ1

ℎ2 푢5

푢6

푢7

푢8

Fig. 2.11 The graph 퐴10

Show that: (i) the graph 퐴푛 is a bipartite matching covered graph for each even integer 푛 ≥ 6, and (ii) Φ( 퐴푛 ) = 2푛 − 4.

2.6 Notes In the treatise by Lov´asz and Plummer [59], matching covered graphs are referred to as 1-extendable graphs as any edge of such a graph can be ‘extended’ to a perfect matching of the graph. (More generally, a connected graph 퐺 is 푘-extendable if any set of 푘 edges which constitutes a matching can be extended to a perfect matching of 퐺.) The term ‘matching covered graphs’ has been widely accepted since the appearance of Lov´asz’s seminal paper (1987, [58]). A matchable graph is elementary if the set of matchable edges in it induces a connected subgraph. Every matching covered graph is elementary, but not every elementary graph is matching covered. However a connected bipartite graph is matching covered if and only if it is elementary. In Chapters 4 and 5 in Lov´asz and Plummer’s book [59] many results about matching covered graphs are derived as special cases of results about elementary graphs. In this work, we have chosen to confine ourselves to matching covered graphs. CLM (2013, [17]) showed that, for each even integer 푛 ≥ 14, any bipartite matching covered graph 퐺 푛 on 푛 vertices with 훿(퐺 푛 ) ≥ 3 has at least 2푛 − 4 perfect matchings and that the graph 퐴푛 described in Exercise 2.5.8 achieves this lower bound. We do not know the best possible lower bound to Φ(퐺 푛 ) if we require, as an additional

40

2 Matching Covered Graphs

hypothesis, that 퐺 푛 be simple. In particular, we do not know any infinite family {퐺 푛 } of simple bipartite matching covered graphs of minimum degree at least three such that Φ(퐺 푛 ) grows linearly. The number of perfect matchings of biwheels grows quadratically. For a special class of bipartite graphs known as braces, which will be introduced later, we have shown that the number of perfect matchings grows quadratically with the number of vertices. This result also appears in the same paper mentioned above. See the Notes Section of Chapter 6 for a related conjecture.

Chapter 3

Canonical Partitions

Contents 3.1

3.2 3.3

3.4 3.5

Barriers in Matchable Graphs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1.1 Properties of barriers of matchable graphs . . . . . . . . . . . . . . 3.1.2 Maximal barriers in matchable graphs . . . . . . . . . . . . . . . . . 3.1.3 Maximal barriers in bipartite matching covered graphs . . . Barriers and Cores . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Canonical Partitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3.1 Bicritical graphs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3.2 Bisubdivisions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Ear Decompositions of Bipartite Graphs . . . . . . . . . . . . . . . . . . . . . . . Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

41 42 42 42 44 46 46 47 50 55

3.1 Barriers in Matchable Graphs A barrier 퐵 in a matchable graph 퐺 is maximal if there is no other barrier of 퐺 that properly contains 퐵. The barrier 퐵 = {푢 1 , 푢 2 , 푢 3 } in the graph shown in Figure 2.1 is not maximal because there are barriers of the graph which properly contain 퐵. It can be verified that {푣 1 , 푢 1 , 푢 2 , 푢 3 , 푢 4 , 푢 5 } is a maximal barrier of that graph. The objective of this chapter is to establish one of the basic results of our subject, which states that the maximal barriers of a matching covered graph constitute a partition of its vertex set1. Our strategy is to first verify its validity for bipartite graphs, and then deduce the general result from this special instance. 1 According to Lov´asz and Plummer [59, Chapter 5], this result seems to have originated in the works of Kotzig (that appeared in a Slovenian journal in 1959 and 1960) and further developed by Lov´asz (in a paper that appeared in 1972).

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7_3

41

42

3 Canonical Partitions

3.1.1 Properties of barriers of matchable graphs We record below two properties of barriers of matchable graphs which are frequently used. We leave the proof of their validity as Exercise 3.1.1. Proposition 3.1 Let 퐵 be a barrier in a matchable graph 퐺, and let 푀 be any perfect matching of 퐺. Then: (i) if 퐾 is an odd component of 퐺 − 퐵, then 푀 ∩ 휕 (퐾) has precisely one edge; and if 푣 is the end of that edge in 푉 (퐾), then 푀 ∩ 퐸 (퐾) is a perfect matching of 퐾 − 푣, and (ii) if 퐿 is an even component of 퐺 − 퐵, then 푀 ∩ 퐸 (퐿) is a perfect matching of 퐿 ƒ and no edge in 휕 (퐿) is matchable in 퐺.

3.1.2 Maximal barriers in matchable graphs Theorem 3.2 establishes an interesting and useful property of maximal barriers. (Recall that a graph 퐺 is critical if 퐺 − 푣 is matchable for every vertex 푣 of 퐺. See Exercises 1.3.8 and 1.3.9.) Maximal barriers and critical subgraphs Theorem 3.2 Let 퐵 be a maximal barrier in a matchable graph 퐺. Then each component of 퐺 − 퐵 is a critical graph. (In particular, 퐺 − 퐵 has no even components.) Proof Let 퐽 be any component of 퐺 − 퐵. We consider separately the cases (i) 퐽 is even and (ii) 퐽 is odd but not critical. If 퐽 is even, then let 퐵′ := 퐵 + 푣, where 푣 is a vertex of 퐽; it follows that 표(퐺 − 퐵′ ) ≥ |퐵′ |. Alternatively, if 퐽 is odd but not critical, there exists a nonempty subset 푆 of 푉 (퐽) such that 표(퐽 − 푆) ≥ |푆| + 1 (by Exercises 1.2.1 and 1.3.9). Now consider the set 퐵′ = 퐵 ∪ 푆. It can be verified that 표(퐺 − 퐵′ ) ≥ |퐵′ |. In both alternatives, 표(퐺 − 퐵′ ) ≥ |퐵′ |. However, as 퐺 is matchable, we must in fact have 표(퐺 − 퐵′ ) = |퐵′ |, implying that 퐵′ is a barrier of 퐺, and contradicting the maximality of 퐵. The assertion follows. ƒ In fact, the converse of Theorem 3.2 is true (Exercise 3.1.4): Theorem 3.3 Let 퐵 denote a barrier of a matchable graph 퐺. If every component of 퐺 − 퐵 is critical then 퐵 is maximal. ƒ

3.1.3 Maximal barriers in bipartite matching covered graphs Let 푆 be any maximal barrier of a matching covered graph 퐺. Then each component of 퐺 − 푆 is a critical graph by Theorem 3.2. In the special case in which 푆 is a

3.1 Barriers in Matchable Graphs

43

maximal barrier in a bipartite graph 퐺 [ 퐴, 퐵], this means that all components of 퐺 − 푆 are trivial, because no nontrivial bipartite graph is critical. As a consequence we have the following simple result. Theorem 3.4 The only two maximal barriers in a bipartite matching covered graph 퐺 := 퐺 [ 퐴, 퐵] are its two parts 퐴 and 퐵. Proof Let 푆 be any maximal barrier of 퐺. Since 퐺 is matching covered, 푆 is a stable set. By the observation made above, each component of 퐺 − 푆 is trivial. Thus, 푆 := 푉 (퐺) − 푆 is another stable set, and (푆, 푆) is a bipartition of 퐺. However, the graph 퐺, being a nontrivial connected bipartite graph, has a unique bipartition (see Exercise 3.1.5), implying that 푆 is equal to either 퐴 or to 퐵. ƒ As any barrier is a subset of a maximal barrier, we have: Corollary 3.5 Every barrier in a bipartite matching covered graph 퐺 [ 퐴, 퐵] is a subset of one of the two sets 퐴 and 퐵. ƒ Another consequence of Theorem 3.4 is the following result which will prove to be very useful in the proof of Theorem 3.11. Corollary 3.6 Let 퐺 := 퐺 [ 퐴, 퐵] be a bipartite matching covered graph, and let 푎 and 푏 any two vertices of 퐺 with 푎 ∈ 퐴 and 푏 ∈ 퐵. Then 퐺 − 푎 − 푏 is matchable. Proof Suppose that 퐺 − 푎 − 푏 is not matchable. Then, by Theorem 2.1, there exists a barrier of 퐺 which contains both 푎 and 푏. By Corollary 3.5, this is impossible. ƒ As a straightforward consequence of Corollary 3.6, we have the following useful result: Corollary 3.7 Let 퐺 := 퐺 [ 퐴, 퐵] be a bipartite matching covered graph, and let 푎 and 푏 be any two vertices of 퐺 with 푎 ∈ 퐴 and 푏 ∈ 퐵. Then the graph 퐺 + 푎푏, obtained from 퐺 by adding an edge joining 푎 and 푏 is also a bipartite matching covered graph with bipartition [ 퐴, 퐵]. ƒ In fact, the converse of Corollary 3.6 also holds. In addition to the Hall-type characterization of bipartite matching covered graphs (Theorem 2.9), there is the following additional characterization: A second characterization of bipartite matching covered graphs Theorem 3.8 A bipartite graph 퐺 = 퐺 [ 퐴, 퐵] which has at least one edge is matching covered if and only if 퐺 − 푎 − 푏 is matchable, for every 푎 ∈ 퐴 and every 푏 ∈ 퐵. The proof of Theorem 3.8 is left as Exercise 3.1.7.

3 Canonical Partitions

44

Exercises 3.1.1 Prove Proposition 3.1. 3.1.2 Starting with the barrier 퐵 in the graph shown in Figure 2.1, and applying the idea of the proof of Theorem 3.2, find all the maximal barriers containing 퐵. ∗ 3.1.3 Draw the graph 퐾3,3 ⊙ 퐾3,3 and list all the barriers in it. ∗ 3.1.4 Prove Theorem 3.3. Hint: use the statement of Exercise 1.3.9 to prove that no proper superset of 퐵 is a barrier of 퐺. ⊲3.1.5 A graph 퐺 is said to have a unique bipartition if its vertex set 푉 has a unique partition into two stable sets. Show that a nonempty graph is a connected bipartite graph if and only if it has a unique bipartition. ⊲3.1.6 Prove Corollary 3.7. ∗ 3.1.7 Prove Theorem 3.8.

3.2 Barriers and Cores Although Theorem 3.4 is specific to bipartite graphs, it can be applied to prove that, in any matching covered graph, the maximal barriers of the graph partition its vertex set. We now proceed to introduce the device that makes this generalization possible. Core with respect to a barrier For a barrier 퐵 of a matchable graph 퐺, the bipartite graph H(퐵) is defined as the graph obtained by deleting all edges with both ends in 퐵, deleting all vertices in the even components of 퐺 − 퐵, and then shrinking the vertex set of each odd component of 퐺 − 퐵 to a single vertex. We shall refer to this bipartite graph H(퐵) as the core of 퐺 with respect to the barrier 퐵. See Example 3.9 for an illustration.

Example 3.9 Let 퐺 be the graph shown in Figure 3.1(a), and let 퐵 be the barrier {푢 1 , 푢 2 , 푢 3 }. Then 퐺 − 퐵 has one even component, and three odd components. The core H(퐵) with respect to the barrier 퐵 is shown in Figure 3.1(b), where 푎 1 is the vertex resulting from shrinking {푣 3 , 푣 4 , 푣 5 }, 푎 2 is the vertex resulting from shrinking {푣 6 }, and 푎 3 is the vertex resulting from shrinking {푢 4 , 푢 5 , 푣 7 , 푣 8 , 푣 9 }. It is an easy matter to verify that H(퐵) is matchable, and that it is matching covered when 퐺 is matching covered (see Exercise 3.2.1). This bipartite graph will play a crucial role in our approach to the subject. The proof of Theorem 3.11 provides a good illustration of its usefulness!

45

3.2 Barriers and Cores 푣5

푣3 푣1

푣2

푣6

푣7

푣8

푣9

푣4

퐵 푢1

푢2

푢3

푢4

푢5

(a) 퐺 푎1

푎2

푎3

푢1

푢2

푢3

퐵 (b) H( 퐵) Fig. 3.1 The core H( 퐵) of 퐺 with respect to the barrier 퐵

Exercises 3.2.1 Let 퐺 be a matchable graph and let 퐵 denote a barrier of 퐺. Show that the core H(퐵) with respect to 퐵 is matchable, and that if 퐺 is matching covered then H(퐵) is a bipartite matching covered graph. Hint: use Proposition 3.1. ∗ 3.2.2 (i) Show that if a bipartite nontrivial graph 퐻 := 퐻 [ 퐴, 퐵] has a unique perfect matching, then at least one vertex in 퐴, and at least one in 퐵, have degree one in 퐻. (Hint: Let 푀 be the only perfect matching in 퐻, and let 푃 be a maximal 푀-alternating path. Show that the ends of 푃 have degree one in 퐻 and show also that 푃 starts and ends with an edge in 푀.) (ii) Deduce that a 2-edge-connected matchable graph 퐺 has at least two perfect matchings. (Hint: Consider the core H(퐵) of 퐺 with respect to any maximal barrier 퐵 of 퐺.) (Kotzig) (iii) Show that if a bipartite graph 퐻 := 퐻 [ 퐴, 퐵] has a unique perfect matching 푀, then there is an enumeration 푎 1 , 푎 2 , . . . , 푎 푘 of the vertices of 퐴 and an enumeration 푏 1 , 푏 2 , . . . , 푏 푘 of the vertices of 퐵 such that (i) 푎 푖 푏 푖 ∈ 푀, for 푖 = 1, 2, . . . , 푘 and (ii) 푎 푖 푏 푗 ∉ 퐸 (퐻) for 1 ≤ 푖 < 푗 ≤ 푘.

3 Canonical Partitions

46

3.3 Canonical Partitions We are now ready to prove the result mentioned in the beginning of this chapter. Lemma 3.10 Let 퐵 be a maximal barrier of a matching covered graph 퐺, and let 푎 ∉ 퐵, and 푏 ∈ 퐵 be any two vertices of 퐺. Then 퐺 − 푎 − 푏 is matchable. Proof Since 푎 ∉ 퐵, it follows that it belongs to an odd component, say 퐾, of 퐺 − 퐵. Now consider the core H(퐵) of 퐺 relative to 퐵. Let 푎 퐾 denote the vertex in H(퐵) resulting from the shrinking of the odd component 퐾. It follows from Corollary 3.6 that H − 푎 퐾 − 푏 is matchable. Let 푁 be a perfect matching in H − 푎 퐾 − 푏. Using the fact that each component of 퐺 − 퐵 is critical, it is now easy to show that 푁 can be ƒ extended to a perfect matching of 퐺 − 푎 − 푏 (Exercise 3.3.1). The canonical partition theorem Theorem 3.11 The maximal barriers of a matching covered graph 퐺 partition its vertex set. Proof Since every singleton is a barrier, every vertex of 퐺 is contained in some maximal barrier of 퐺. So, to establish the assertion, it suffices to show that maximal barriers of 퐺 are pairwise disjoint. With this objective in mind, let 퐵 and 퐵′ be any two distinct maximal barriers of 퐺, and suppose that 푏 ∈ 퐵 ∩ 퐵′ . The fact that 퐵′ is maximal and is distinct from 퐵 implies that there must exist a vertex 푎 in 퐵′ − 퐵. By Lemma 3.10, it follows that the graph 퐺 − 푎 − 푏 is matchable. However, because 푎 and 푏 both belong to the barrier 퐵′ , this is impossible by Theorem 2.1. ƒ Canonical partition The partition of the vertex set 푉 of a matching covered graph 퐺 into maximal barriers is called its canonical partition. It follows from Corollary 3.5 that the canonical partitions of bipartite matching covered graphs have just two parts. Curiously, the canonical partition of any nonbipartite matching covered graph has at least four parts (see Exercise 3.3.5).

3.3.1 Bicritical graphs A nontrivial graph 퐺 is bicritical if 퐺 − 푢 − 푣 is matchable for any two vertices 푢 and 푣 of 퐺. Bicritical graphs are a special type of matching covered graphs (see Exercise 3.3.6). By Theorem 2.1 this is the case precisely when all maximal barriers of 퐺 are singletons. The three graphs shown in Figure 2.2 are bicritical; another example is shown in Figure 3.5. Three-connected bicritical graphs, which are known as bricks, will play an important role in our theory.

47

3.3 Canonical Partitions

3.3.2 Bisubdivisions A subdivision of an edge 푒 = 푢푣 of a graph 퐺 consists of replacing the edge 푒 by an 푢푣-path so that all the internal vertices (if any) of that path have degree two in the resulting graph. As the name suggests, a subdivision of 푒 consists of dividing 푒 by the insertion of a certain number (possibly zero) of vertices of degree two. If the number of vertices inserted is even, then we refer to the subdivision as a bisubdivision. Thus, in case of a bisubdivision, we replace 푒 by a path of odd length (with an even number of added vertices). For this reason, some authors refer to a bisubdivision of an edge as an odd subdivision (and some as an even subdivision). The following proposition is easy to verify (see Exercise 3.3.9). Proposition 3.12 Suppose that a graph 퐺˜ is obtained by bisubdividing an edge 푒 = 푢푣 of a matching covered graph 퐺 distinct from 퐾2 , effected by replacing the edge 푢푣 by the path 푃 := 푢푠1 푠2 . . . 푠2푘 푣, where 푘 > 0. Then the graph 퐺˜ is also matching covered. And if 퐵푢 and 퐵 푣 are the maximal barriers of 퐺 containing 푢 and 푣, respectively (see Figure 3.2), then: 푠1 퐵푢

푠2

푢

푠2푖−1

푠2푖

퐵푣

푣 푠2푘

푠2푘−1

Fig. 3.2 Bisubdivision of an edge 푒 = 푢푣

(i) all parts of the canonical partition of 퐺, except 퐵푢 and 퐵 푣 , are also parts in the ˜ and canonical partition of 퐺, (ii) 퐵˜ 푢 = 퐵푢 ∪ {푠2 , 푠4 , . . . , 푠2푘 } and 퐵˜ 푣 = 퐵 푣 ∪ {푠1 , 푠3 , . . . , 푠2푘−1 } are the maximal ˜ barriers containing 푢 and 푣, respectively, in the canonical partition of 퐺. ƒ A bisubdivision of a graph 퐺 is a graph obtained from 퐺 by bisubdividing each edge in a subset of 퐸. (We regard a graph as a bisubdivision of itself.) Figure 3.3 shows a spanning subgraph of the Petersen graph which is a bisubdivision of 퐾4 . (The edges of this subgraph are indicated by solid lines.)

48

3 Canonical Partitions

Fig. 3.3 A bisubdivision of 퐾4 contained in the Petersen graph

If 퐺˜ denotes a bisubdivision of a matching covered graph 퐺, then by the above proposition, the graph 퐺˜ is also matching covered. We refer to the vertices in 퐺˜ ˜ and those vertices of 퐺˜ which are also vertices of 퐺 as the branch vertices of 퐺, which are not vertices of 퐺 as the subdivision vertices. We conclude this subsection by noting that the graph obtained by subdividing an edge 푒 in a graph 퐺 may be constructed as the result of splicing 퐺 with a graph 퐻 whose underlying simple graph is an even cycle (see Exercise 3.3.11).

Exercises ∗ 3.3.1 In the proof of Lemma 3.10, supply the missing details in showing that 푁 can be extended to a perfect matching of 퐺 − 푎 − 푏. 3.3.2 Let 퐺 be a matching covered graph and let 푢 and 푣 be two vertices of 퐺 such that 퐺 + 푢푣 is matching covered. Prove that the canonical partition of 퐺 + 푢푣 is a refinement of the canonical partition of 퐺 in the sense that every maximal barrier of 퐺 + 푢푣 is a subset of a maximal barrier of 퐺. 3.3.3 Determine the canonical partition of the matching covered graph shown in Figure 3.4 (which is obtained by splicing 퐾3,3 with 퐾4 ). 3.3.4 Determine the canonical partitions of the following graphs: (i) the wheel 푊2푘+1 , 푘 ≥ 1, (ii) 푊2푘+1 − 푒, where 푘 ≥ 2, and 푒 is a spoke of 푊2푘+1 . ∗ 3.3.5 Show that the canonical partition of a nonbipartite matching covered graph has at least four parts. For each integer 푟 ≥ 4, give an example of a matching covered graph 퐺 whose canonical partition consists of precisely 푟 parts. (Carvalho) ⊲3.3.6 Prove that every bicritical graph is matching covered.

49

3.3 Canonical Partitions 1

2

4

3

6

5

7

8

Fig. 3.4 Figure for Exercise 3.3.3

Fig. 3.5 A bicritical graph of order six

3.3.7 Show that the graph shown in Figure 3.5 is bicritical. 3.3.8 Show that a graph obtained by splicing together two bicritical graphs is also bicritical. ⊲3.3.9 Prove Proposition 3.12. ∗ 3.3.10 Let 퐺 be a matching covered graph, and let {푢 1 , 푣 1 } and {푢 2 , 푣 2 } be any two pairs of nonadjacent vertices of 퐺. State and prove the necessary and sufficient conditions under which it is possible to add two new edges 푒 1 = 푢 1 푣 1 and 푒 2 = 푢 2 푣 2 to 퐺 such that neither 퐺 + 푒 1 nor 퐺 + 푒 2 is matching covered, but 퐺 + 푒 1 + 푒 2 is. 3.3.11 Let 푒 be an edge of a graph 퐺 incident with a vertex 푢 of degree 푘 > 2, and let 퐻 be a graph, disjoint from 퐺, which is obtained from a 2푟-cycle, 푟 ≥ 2, by replacing one of its edges 푣푣 ′ by 푘 − 1 multiple edges. (Thus, in 퐻, the vertices 푣 and 푣 ′ are adjacent and have degree 푘; the rest of the vertices have degree two.) Find a suitable bijection 휃 between 휕퐺 (푢) and 휕퐻 (푣) so that (퐺 ⊙ 퐻)푢,푣, 휃 is isomorphic to the graph obtained from 퐺 by subdividing 푒 by the insertion of 2푟 − 2 vertices.

3 Canonical Partitions

50

3.4 Ear Decompositions of Bipartite Graphs ♯ Given a matching covered graph 퐻, when is it possible to add an edge joining two vertices 푢 and 푣 of 퐻 by a new edge so that the resulting graph 퐻 +푢푣 is also matching covered? The answer to this question is provided by Lemma 3.10 and Theorem 3.11: the graph 퐻 + 푢푣 is matching covered if and only if 푢 and 푣 belong to different maximal barriers in the canonical partition of 퐻. Corollary 3.7 implies the following observation concerning bipartite matching covered graphs: Proposition 3.13 Let 퐻 [ 퐴, 퐵] be a spanning bipartite matching covered subgraph of a bipartite matching covered graph 퐺 [ 퐴, 퐵]. Then 퐺 can be obtained from 퐻 by adding the edges in 퐸 (퐺) − 퐸 (퐻), one at a time, in any order, so that all intermediate ƒ graphs in the sequence are also matching covered. The above result does not extend to nonbipartite graphs. Consider for example the spanning subgraph of the Petersen graph shown in Figure 3.3. There are three edges of the Petersen graph (indicated by dotted lines) which do not belong to this subgraph. None of those three edges can be added to the subgraph to obtain a matching covered graph. (We have to add at least two of those three edges simultaneously to obtain a matching covered graph.) Now let us suppose that 퐻 is a matching covered subgraph of a matching covered graph 퐺, with 푉 (퐻) ⊂ 푉 (퐺), and suppose that we wish to obtain 퐺 from 퐻 by means of elementary operations which preserve matching coveredness. Since the addition of one or more edges to 퐻 does not result in a graph of order larger than that of 퐻, it is also necessary to employ an operation that augments the number of vertices. An obvious candidate for such an operation is bisubdivision. Ear A (single) ear in a matching covered graph is a path of odd length2 all of whose internal vertices have degree two. The idea of adding an ear to a matching covered graph combines the ideas of edge additions and bisubdivisions. Ear path Suppose 퐻 is a subgraph of a matching covered graph 퐺. An odd path 푃 := 푢푠1 푠2 · · · 푠2푘 푣 in 퐺 connecting two vertices 푢 and 푣 of 퐻 is called an ear path of 퐻 in 퐺 if it is internally disjoint from 퐻. If 푃 is an ear path of 퐻 in 퐺, then it is an an ear in the graph 퐻 ∪ 푃. For this reason, we say that 퐻 + 푃 := 퐻 ∪ 푃 is obtained from 퐻 by adding the ear 푃 to 퐻. 2 On occasion, we shall use the term ‘even ear’ in a graph to mean a path of even length. Unless otherwise specifically stated, the term ‘ear’ should be taken to mean a path of odd length.

51

3.4 Ear Decompositions of Bipartite Graphs

Clearly, the above-defined operation is tantamount to first adding to 퐻 an edge joining 푢 and 푣, and then bisubdividing that edge by inserting new vertices 푠1 , 푠2 , ..., 푠2푘 . When 퐻 is matching covered, any graph obtained from 퐻 by adding an ear connecting two vertices in different maximal barriers of 퐻 is also matching covered. A natural question that arises is if it is possible to generate all matching covered graphs by starting with the smallest matching covered graph, namely 퐾2 , and sequentially adding ears (with the proviso that all intermediate graphs in the sequence are also matching covered). However, it follows from Proposition 3.12 that adding an ear to a bipartite matching covered graph results in another bipartite matching covered graph. Thus it is not possible to generate nonbipartite matching covered graphs in the above described manner. We shall however see that this is possible in case of bipartite graphs. Before formally defining what an ear decomposition of a bipartite graph is, let us consider the example shown in Figure 3.6. The paths 푃1 := 12,

푃2 := 1342,

푃4 := 3874,

푃5 := 58,

푃3 := 1562, and

푃6 := 67,

are paths of odd length in the cube and 퐺 푖 := 푃1 ∪ 푃2 ∪ · · · ∪ 푃푖 ,

for 1 ≤ 푖 ≤ 6.

The graph 퐺 1 is 퐾2 and, for 2 ≤ 푖 ≤ 6, the graph 퐺 푖 is obtained from 퐺 푖−1 by adding the ear 푃푖 . Ear decomposition of a bipartite graph A bottom-up ear decomposition of a bipartite matching covered graph 퐺 [ 퐴, 퐵] or, simply, an ear decomposition of 퐺 [ 퐴, 퐵]), is a sequence 퐺 1 ⊂ 퐺 2 ⊂ · · · ⊂ 퐺 푟 of matching covered subgraphs of 퐺, where (i) 퐺 1 = 퐾2 , 퐺 푟 = 퐺 and, (ii) for 2 ≤ 푖 ≤ 푟, 퐺 푖 = 퐺 푖−1 + 푃푖 , where 푃푖 is an ear of 퐺 푖 in 퐺. Note that if 푃 := 푢푠1 푠2 · · · 푠2푘 푣 is a path of odd length, then the set 퐸 푒푣푒푛 (푃) := {푠1 푠2 , 푠3 푠4 , . . . , 푠2푘−1 푠2푘 } of even numbered edges of 푃 is a perfect matching of 푃 − {푢, 푣} (when 푃 is a path of length one, then 퐸 푒푣푒푛 (푃) is empty). It follows that if a bipartite matching covered graph 퐺 has an ear decomposition as described above, where 푒 is the unique edge of 퐺 1 , then 푀푒 := {푒} ∪ 퐸 푒푣푒푛 (푃2 ) ∪ · · · ∪ 퐸 푒푣푒푛 (푃푟 )

52

3 Canonical Partitions 1

2

1

2

1

2

5

퐺1 1

3

퐺2

3 2

1

4

3

2

1

6

퐺3

4 2

5

6

5

6

5

6

8

7

8

7

8

7

퐺4

4

3

퐺5

4

3

퐺6

4

Fig. 3.6 An ear decomposition of the cube

is a perfect matching of 퐺 such that, for 1 ≤ 푖 ≤ 푟, the set 푀푒 ∩ 퐸 (퐺 푖 ) is a perfect matching of 퐺 푖 . This observation is the basis of the proof of the following fundamental theorem. Theorem: ear decomposition of a bipartite graph Theorem 3.14 Every bipartite matching covered graph 퐺 [ 퐴, 퐵] has an ear decomposition. Proof Let 푒 be any edge of 퐺, and let 푀푒 be any fixed perfect matching containing 푒. Take 퐺 1 to be the subgraph of order two whose only edge is 푒. Assume, inductively, that a partial ear decomposition 퐺 1 ⊂ 퐺 2 ⊂ · · · ⊂ 퐺 푖 of 퐺 has been constructed for some 푖 ≥ 1, with the property that the restriction of 푀푒 to 퐸 (퐺 푖 ) is a perfect matching of 퐺 푖 . If 퐺 푖 = 퐺, we have the required ear decomposition of 퐺 with 푖 = 푟. So, suppose that 퐺 푖 ⊂ 퐺. In this case, since 퐺 is connected, there must exist an edge 푓 belonging to 퐸 (퐺) − 퐸 (퐺 푖 ) which has at least one end in 푉 (퐺 푖 ). Now let 푀 푓 be any perfect matching of 퐺 containing 푓 . Such a perfect matching must exist because 퐺 is matching covered. Furthermore, 푀 푓 ≠ 푀푒 because 푀푒 ∩퐸 (퐺 푖 ) is a perfect matching of 퐺 푖 , and 푓 has an end in 푉 (퐺 푖 ) but is not in 퐸 (퐺 푖 ). Consider the (푀 푓 , 푀푒 )-alternating cycle starting with the edge 푓 , and let 푔 denote the edge with which this cycle re-enters 푉 (퐺 푖 ) for the first time. (If 푓 has both its ends in 푉 (퐺 푖 ), then 푔 = 푓 .) This edge 푔 must also be in the perfect matching 푀 푓 . Thus the segment of the (푀 푓 , 푀푒 )-alternating cycle starting with 푓 and ending with 푔 is a path of odd length which has its ends in 푉 (퐺 푖 ), but which is internally disjoint from 퐺 푖 .

53

3.4 Ear Decompositions of Bipartite Graphs

Now, by taking this path to be 푃푖+1 , we obtain the (matching covered) subgraph 퐺 푖+1 := 퐺 푖 + 푃푖+1 of 퐺 such that 푀푒 ∩ 퐸 (퐺 푖+1 ) is a perfect matching of 퐺 푖+1 . Thus we can extend the given partial decomposition of 퐺 by appending 퐺 푖+1 . This process can clearly be continued until we obtain an ear decomposition of 퐺. ƒ The above proof implicitly contains an algorithm for finding an ear decomposition of 퐺. The fact that, for 1 ≤ 푖 < 푟, 퐺 푖 can be extended to 퐺 푖+1 by adding an ear is guaranteed by the fact that 푀푒 is a perfect matching of 퐺 such that its restriction to 퐺 푖 is a perfect matching of 퐺 푖 . For this reason we refer to 푀푒 as the reference matching; it may be chosen arbitrarily but, once chosen, it guides the process of finding an ear decomposition. A bipartite graph need not have a unique ear decomposition. (For example, it is easy to find an ear decomposition of the cube in which the second graph 퐺 2 is a hamiltonian cycle in it.) However, it is easy to see that the number of subgraphs in an ear decomposition is independent of the choices of ears in obtaining the decomposition (Exercise 3.4.2). Theorem 3.15 In any ear decomposition (퐺 1 , 퐺 2 , . . . , 퐺 푟 ) of a bipartite matching covered graph 퐺, 푟 = 푚 − 푛 + 2,

where 푚 = |퐸 | and 푛 = |푉 |.

We conclude this chapter with a simple application of the above theorem. Corollary 3.16 Any bipartite matching covered graph 퐺 has at least 푚 − 푛 + 2

distinct perfect matchings.

Proof Let 퐾2 = 퐺 1 , 퐺 2 , . . . , 퐺 푟 = 퐺 be an ear decomposition of 퐺. We prove the assertion by induction on 푟. The assertion holds if 퐺 = 퐾2 . Assume that 퐺 ≠ 퐾2 . Let 푃푟 denote the ear added to 퐺 푟 −1 to obtain 퐺 푟 = 퐺. By induction, 퐺 푟 −1 has at least 푚푟 −1 − 푛푟 −1 + 2 distinct perfect matchings, where 푚푟 −1 = |퐸 (퐺 푟 −1 )| and 푛푟 −1 = |푉 (퐺 푟 −1 )|. Each of these perfect matchings can be extended to a perfect matching of 퐺 by adding the edges in the set of even numbered edges of 푃푟 . In addition, the set of odd numbered edges of 푃푟 are contained in at least one perfect matching of 퐺. (The restriction of such a perfect matching of 퐺 to 퐸 (퐺 푟 −1 ) cannot be a perfect matching of 퐺 푟 −1 .) Thus, altogether, 퐺 has at least 푚푟 −1 − 푛푟 −1 + 2 + 1 = 푚 − 푛 + 2 distinct perfect matchings.

ƒ

The proof of the above theorem illustrates how the ear decomposition procedure for bipartite graphs can be used as an inductive tool. There is an ear decomposition procedure which is applicable to all matching covered graphs where simultaneous addition of two disjoint ears is permitted. Establishing the existence of such a procedure is much harder and will be discussed in Chapter 11. In the next chapter we turn to a different type of decomposition which is of central importance in this theory.

3 Canonical Partitions

54

Exercises 3.4.1 Show that any bipartite matching covered graph 퐺 with minimum degree at least three has an edge 푒 such that 퐺 − 푒 is matching covered. Give an example to show that the above statement is invalid without the assumption that 퐺 is bipartite. 3.4.2 Give a proof of Theorem 3.15. Conformal subgraphs A matchable subgraph 퐻 of a matchable graph 퐺 is conformal if the graph 퐺 − 푉 (퐻) has a perfect matching (equivalently, if any perfect matching of 퐻 may be extended to a perfect matching of 퐺). ⊲3.4.3 (i) Prove that if 퐺 1 is a conformal subgraph of 퐺 2 and if 퐺 2 is a conformal subgraph of 퐺 3 then 퐺 1 is also a conformal subgraph of 퐺 3 . (ii) Find all the conformal cycles in the graph shown in Figure 3.7.

Fig. 3.7 Graph for Exercise 3.4.3.

(iii) Find all conformal cycles in the Petersen graph. ∗ 3.4.4 Let 퐻 be a matching covered subgraph of a bipartite matching covered graph 퐺. Show that any ear decomposition of 퐻 may be extended to an ear decomposition of 퐺 if and only if 퐻 is a conformal subgraph of 퐺. ⊲3.4.5 Let 퐻 be any conformal matching covered subgraph of a matching covered graph 퐺 (which may or may not be bipartite), and let 푔 be an edge in 퐸 (퐺) − 퐸 (퐻) which has an end in 푉 (퐻). Suppose that 푀 is a perfect matching of 퐺 such that 푀 ∩ 퐸 (퐻) is a perfect matching of 퐻, let 푁 be a perfect matching of 퐺 containing the edge 푔, and let 퐶 be an (푁, 푀)-alternating cycle containing the edge 푔. Show that 퐻 ∪ 퐶 is a conformal matching covered subgraph of 퐺. Conformal minors A matching covered graph 퐽 is a conformal minor of another matching covered graph 퐺 if some bisubdivision 퐻 of 퐽 is a conformal subgraph of 퐺. (Figure 3.3 shows that 퐾4 is a conformal minor of the Petersen graph.)

3.5 Notes

55

⊲3.4.6 Show that the triangular prism (the graph 퐶6 ) is not a conformal minor of the Petersen graph P.

3.5 Notes We shall see in Chapter 6 that the set of perfect matchings obtained as in the proof of Corollary 3.16 have a special significance, namely that their incidence vectors form a basis for the vector space generated by the set of incidence vectors of perfect matchings of 퐺 (see Exercise 6.3.3). Conformal subgraphs are known by various names in the literature. Lov´asz and Plummer [59] referred to them as ‘nice’ subgraphs, and Robertson, Seymour and Thomas [82] called them ‘central’ subgraphs. McCuaig [68] had the more descriptive name ‘well-fitted’ subgraphs. The term ‘conformal’ subgraphs that we are using (and have used in several of our papers) is in the same spirit as McCuaig’s terminology. The notion of a conformal minor of a matching covered graph (and related notions which will be introduced in due course) plays a role in this theory that is akin to that of the notion of a minor in the theory of graphs. For example, the structure of bipartite matching covered graphs which do not have 퐾3,3 as a conformal minor plays a pivotal role in the theory of Pfaffian orientations which will be discussed in Part III.

Chapter 4

Tight Cuts

Contents 4.1 4.2

4.3

4.4

4.5 4.6

Separating Cuts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Tight Cuts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.1 Barrier cuts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.2 2-Separation cuts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.3 Tight 3-edge cuts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2.4 Separating cuts with bipartite shores . . . . . . . . . . . . . . . . . . Tight Cut Decompositions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3.1 Bricks and braces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3.2 Uncrossing cuts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Uniqueness of Tight Cut Decompositions . . . . . . . . . . . . . . . . . . . . . . 4.4.1 The number of bricks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4.2 Near-bricks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Subadditivity of the Number of Bricks . . . . . . . . . . . . . . . . . . . . . . . . Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

57 59 60 61 61 63 66 66 68 74 76 76 77 80

4.1 Separating Cuts In Chapter 2 we introduced the operation of splicing which can be used to ‘combine’ two matching covered graphs to obtain another matching covered graph. Here we introduce the related notion of a separating cut which leads to a way of ‘decomposing’ a matching covered graph into two matching covered graphs. Tight cuts are a special type of separating cuts. They play a pivotal role in many aspects of this theory, including the study of questions concerning the existence of an edge in a matching covered graph whose deletion results in another matching covered graph. Before we proceed further, it is necessary to establish the notation and terminology we use in connection with cuts in graphs. Familiarity with these is crucial for understanding much of the rest of the book. © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7_4

57

4 Tight Cuts

58

Given a nonempty set 푋 of vertices of a graph 퐺, we denote by 퐺/( 푋 → 푥) the graph obtained from 퐺 by contracting 푋 to a single new vertex 푥 (and removing all resulting loops). If the name of the contraction vertex 푥 is not important, we simply write 퐺/푋. Cuts and cut-contractions Given any cut 퐶 := 휕 ( 푋) of a connected graph 퐺, where 푋 is a nonempty proper subset of 푉, we refer to the two graphs 퐺/( 푋 → 푥) and 퐺/( 푋 → 푥) as the 퐶-contractions of 퐺, where 푋 = 푉 − 푋. With the aid of the above notation, we are now in a position to introduce the important notion of a separating cut in a matching covered graph. Separating cut A cut 퐶 := 휕 ( 푋) in a matching covered graph 퐺 is a separating cut if both 퐶-contractions 퐺/푋 and 퐺/푋 are also matching covered. Figure 4.1 shows two examples of separating cuts which are indicated by dotted lines. Figure 4.1(a) depicts a separating cut in the Petersen graph, both the contractions with respect to this cut are 5-wheels. Referring to Figure 4.1(b), one of the contractions with respect to the indicated separating cut is 퐾3,3 , and the other is 퐾4 .

(푎)

(푏)

Fig. 4.1 Examples of separating cuts

Note that if 퐺 1 and 퐺 2 are the two 퐶-contractions of a matching covered graph 퐺 with respect to a separating cut 퐶, then the sets of edges incident with the contraction vertices in 퐺 1 and 퐺 2 are both equal to 퐶. Thus, 퐺 can be recovered from 퐺 1 and 퐺 2 by splicing them together in the obvious manner. The following assertion can be deduced from the definition of a separating cut and the fact that no matching covered graph can have a cut vertex (see Exercise 4.1.1):

4.2 Tight Cuts

59

Proposition 4.1 For any separating cut 퐶 of a matching covered graph 퐺, the subgraphs of 퐺 induced by the shores of 퐶 are both connected. In other words, 퐶 is ƒ a bond. The ensuing result provides a useful characterization of separating cuts. Characterization of separating cuts Theorem 4.2 A cut 퐶 = 휕 ( 푋) of a matching covered graph 퐺 is separating if and only if each edge of 퐺 lies in a perfect matching that contains precisely one edge in 퐶. Proof Let 퐺 1 := 퐺/( 푋 → 푥) and 퐺 2 := 퐺/( 푋 → 푥) be the two 퐶-contractions of 퐺. Suppose first that any edge 푒 of 퐺 lies in a perfect matching 푀푒 of 퐺 which meets 퐶 in exactly one edge. The graph 퐺, a matching covered graph, is connected, hence every contraction of 퐺 is connected. In particular, 퐺 1 is connected. Furthermore, for any edge 푒 of 퐺 1 , the fact that |푀푒 ∩ 퐶| = 1 implies that 푀푒 ∩ 퐸 (퐺 1 ) is a perfect matching of 퐺 1 containing 푒. We conclude that 퐺 1 is matching covered. Likewise, 퐺 2 is also matching covered. Hence 퐶 is a separating cut. Conversely, suppose that 퐶 is a separating cut. Then, by definition, both 퐺 1 and 퐺 2 are matching covered. Let 푒 be any edge of 퐺. Adjust notation so that 푒 is an edge of 퐺 1 . Then, since 퐺 1 is matching covered, there is a perfect matching of 퐺 1 , say 푀1 , that contains 푒. Let 푓 (possibly equal to 푒) be the edge of 푀1 incident with the contraction vertex 푥 of 퐺 1 . Then 푓 belongs to 퐶, and thus, is an edge of 퐺 2 incident with the contraction vertex 푥. Now, since 퐺 2 is matching covered, there is a perfect matching of 퐺 2 , say 푀2 , that contains 푓 . We conclude that 푀푒 := 푀1 ∪ 푀2 is a perfect matching of 퐺 that contains the edge 푒 and has just one edge, namely 푓 , in cut 퐶. ƒ

Exercises ⊲4.1.1 Prove Proposition 4.1. 4.1.2 Show that in a cubic matching covered graph every 3-edge cut is a separating cut.

4.2 Tight Cuts Tight cut A cut 퐶 of a matching covered graph 퐺 is tight if |푀 ∩ 퐶| = 1, for every perfect matching 푀 of 퐺.

4 Tight Cuts

60

Simplest examples of tight cuts are the trivial cuts. The cut shown in Figure 4.1(b) is a nontrivial tight cut. By Theorem 4.2, every tight cut of a matching covered graph is also a separating cut. But, in general, a separating cut need not be tight. (For example, the separating cut in the Petersen graph shown in Figure 4.1(a) is not tight!) We describe below two special types of tight cuts in matching covered graphs.

4.2.1 Barrier cuts Let 퐵 be any barrier of a matching covered graph 퐺. It follows from Proposition 3.1 that, for each (odd) component 퐾 of 퐺 − 퐵, the cut 휕 (퐾) is a tight cut in 퐺. Tight cuts which arise in this manner are known as barrier cuts. Every trivial cut is a barrier cut. The cut shown in Figure 4.1(b) is an example of a nontrivial barrier cut. Since bicritical graphs (defined in Section 3.3) do not have nontrivial barriers, the trivial cuts are the only barrier cuts in a bicritical graph.

Special barrier cuts Recall that a barrier 퐵 in a matching covered graph 퐺 is special if 퐺 − 퐵 has precisely one nontrivial odd component. A tight cut 퐶 of 퐺 is a special barrier cut if there is some special barrier 퐵 of 퐺 such that 퐶 = 휕 (퐾), where 퐾 is the unique nontrivial component of 퐺 − 퐵. The cut 퐶 shown in Figure 4.2 is the special barrier cut associated with the barrier {푢, 푣, 푤, 푥}. 푢

푣

푤

퐶

푥 Fig. 4.2 A special barrier cut

We leave the proof of the following simple proposition as Exercise 4.2.1. Proposition 4.3 A tight cut 퐶 in a matching covered graph of order four or more is a special barrier cut if and only if at least one of the two shores of 퐶 is bipartite. ƒ

61

4.2 Tight Cuts

4.2.2 2-Separation cuts A 2-separation of a matching covered graph 퐺 is a 2-vertex cut {푢, 푣} of 퐺 that is not a barrier. If {푢, 푣} is a 2-separation of 퐺, then each component of 퐺 − 푢 − 푣 is even. Suppose that 퐻1 is the union of a nonempty proper subset of the components of 퐺 − 푢 − 푣, and 퐻2 is the union of the remaining components of 퐺 − 푢 − 푣. Then 퐻1 and 퐻2 are two nonempty vertex-disjoint even order subgraphs whose union is 퐺 − 푢 − 푣. With any such expression of 퐺 −푢 −푣 as the union of 퐻1 and 퐻2 , we may associate the cuts 퐶 := 휕 (푉 (퐻1) + 푣) and 퐷 := 휕 (푉 (퐻2) + 푣), and it is easy to see that: 4.4 The two cuts 퐶 and 퐷 defined above are tight (Exercise 4.2.2).

ƒ

Tight cuts of a graph which arise in this manner are known as 2-separation cuts of 퐺. Figure 4.3 depicts a 2-separation {푢, 푣} in a matching covered graph 퐺, a partition {퐻1 , 퐻2 } of 퐺 − 푢 − 푣 and the pair {퐶, 퐷} of 2-separation cuts associated with the partition {퐻1 , 퐻2 }. 푢

퐶

퐷

퐻1 퐻2 푣 Fig. 4.3 Two 2-separation cuts in a matching covered graph

Clearly, 3-connected graphs cannot have any 2-separation cuts.

4.2.3 Tight 3-edge cuts ♯ A 3-edge cut of a graph is an edge cut with three edges. A 3-edge cut in a matching covered graph need not be tight. (For example, the graph 퐶6 has a 3-edge cut that is not tight.) However, those 3-edge cuts which happen to be tight satisfy the following interesting and useful property. Theorem 4.5 In a matching covered graph, every 3-edge cut that is tight is a barrier cut. Proof Let 퐶 := 휕 ( 푋) be a tight 3-edge cut of a matching covered graph 퐺. Let 퐶 := {푢 푖 푣 푖 : 1 ≤ 푖 ≤ 3, 푢 푖 ∈ 푋, 푣 푖 ∈ 푋}.

4 Tight Cuts

62

Case 1 The cut 퐶 is not a matching. Without loss of generality, we may assume that the vertices 푢 푖 , 푖 = 1, 2, 3 are not pairwise distinct. If the three vertices coincide then 퐶 is trivial, hence a barrier cut. If two of them are the same, adjust notation so that 푢 1 = 푢 2 ≠ 푢 3 . The shore 푋 of 퐶, being a shore of a tight cut, is odd. Thus {푢 1 , 푢 3 } is a barrier of 퐺 where 푋 − {푢 1 , 푢 3 } and 푋 are the sets of vertices of the two components of 퐺 − 푢 1 − 푢 3 . In both alternatives, the ends of the edges of 퐶 in 푋 constitute a barrier of 퐺 and 퐶 is a barrier cut associated with that barrier. Case 2 The cut 퐶 is a matching. See Figure 4.4.

푋

푢1

푣1

푢2

푣2 푋

푢3

푣3

Fig. 4.4 The case in which the 3-cut 퐶 is a matching

Let 퐺 1 := 퐺/( 푋 → 푥) and let 퐺 2 := 퐺/( 푋 → 푥). 4.5.1 At least one of the graphs 퐺 1 − 푢 1 − 푢 2 and 퐺 2 − 푣 1 − 푣 2 is not matchable. Proof Assume, to the contrary, that both graphs 퐺 1 − 푢 1 − 푢 2 and 퐺 2 − 푣 1 − 푣 2 are matchable. Let 푀1 and 푀2 be perfect matchings of 퐺 1 − 푢 1 − 푢 2 and 퐺 2 − 푣 1 − 푣 2 , respectively. Then, 푀 := 푀1 ∪ 푀2 ∪ {푢 1 푣 1 , 푢 2 푣 2 } is a perfect matching of 퐺 which contains all the three edges of 퐶, a contradiction to the hypothesis that 퐶 is tight in 퐺. ƒ Adjust notation so that 퐺 1 − 푢 1 − 푢 2 is not matchable. The graph 퐺 1 , being a 퐶-contraction of 퐺, is matching covered. Thus, 퐺 1 has a barrier that contains both 푢 1 and 푢 2 . Let 퐵 denote the maximal barrier of 퐺 1 that contains both 푢 1 and 푢 2 . As 퐺 1 is matching covered, the set 퐵 is independent, and thus the contraction vertex 푥 does not lie in 퐵. We deduce that 퐺 1 − 퐵 has a component, say 퐾, which contains the vertex 푥. 4.5.2 Vertex 푢 3 also lies in 퐵. Proof Assume the contrary. As 푢 3 푣 3 is an edge and 푥 belongs to 푉 (퐾), it follows that 푢 3 also belongs to 푉 (퐾). In fact, 푢 3 is the only vertex of 퐾 that is adjacent to 푥. Thus, 퐾 − 푢 3 is not matchable, hence 퐾 is not critical. This is a contradiction, as 퐵 is a maximal barrier of 퐺 1 . ƒ In sum, the vertex 푥 is isolated in the graph 퐺 1 − 퐵, therefore 퐶 is a barrier cut of 퐺, associated with 퐵, which is also a barrier of 퐺. ƒ

63

4.2 Tight Cuts

Tight cuts may be neither barrier nor 2-separation cuts A matching covered graph may have a tight cut which is neither a barrier cut, nor a 2-separation cut; Figure 4.5 depicts such a cut (see Exercise 4.2.6). However, as we shall now proceed to show, every separating cut in a bipartite matching covered graph is (or may be constructed as) a barrier cut. We shall in fact see that this statement is more generally true for a separating cut in any matching covered graph if that cut happens to have a shore that is bipartite.

푢

푡 푤 푣 푋 Fig. 4.5 A tight cut that is neither a barrier cut nor a 2-separation cut

4.2.4 Separating cuts with bipartite shores Majority and minority parts When the set 푋 of vertices of a connected bipartite graph 퐻 is odd, the two parts of the bipartition of 퐻 have distinct cardinalities; the larger part is called the majority part, the other the minority part; we denote the majority part of 푋 by 푋+ , and the minority part by 푋− . Lemma 4.6 Suppose that 퐺 is a matching covered graph and that 푣 is a vertex of 퐺. If the subgraph 퐺 − 푣 is bipartite, then 퐺 is also bipartite. Furthermore, the majority part of 퐺 − 푣 has just one more vertex than its minority part. Proof Suppose that 퐺 − 푣 is a bipartite graph with bipartition (푈, 푊). Assume without loss of generality that |푈| > |푊 |. Since 퐺 is connected, the vertex 푣 is joined to at least one vertex in 퐺 − 푣. If 푣 is joined in 퐺 to a vertex 푤 in 푊, then no matching containing the edge 푣푤 can cover all the vertices in 푈, and hence 푣푤 would not be matchable in 퐺. This is impossible because 퐺 is matching covered. Thus all

4 Tight Cuts

64

neighbours of 푣 in 퐺 are in 푈, implying that 퐺 is a bipartite graph with bipartition ƒ (푈, 푊 + 푣). As 퐺 is matching covered, it must be the case that |푈| = |푊 | + 1. Separating cuts with a bipartite shore Theorem 4.7 Let 퐶 := 휕 ( 푋) be a separating cut in a matching covered graph 퐺. If the subgraph 퐺 [푋] of 퐺 induced by 푋 is bipartite, and if the subgraph 퐺 [ 푋] of 퐺 induced by 푋 is nontrivial then the majority part 푋+ of 퐺 [푋] is a barrier of 퐺 and 퐶 is a special barrier cut of 퐺. Proof Consider the 퐶-contraction 퐺 ′ := 퐺/( 푋 → 푥). Then 퐺 ′ − 푥 = 퐺 [푋], which is bipartite by the hypothesis. Thus, by Lemma 4.6, 퐺 ′ is also bipartite, and the vertex 푥 is adjacent in 퐺 ′ only to the vertices in 푋+ . Furthermore, | 푋+ | = | 푋− | + 1. Thus, 퐺 − 푋+ has precisely | 푋+ | odd components, one of them being the nontrivial graph 퐺 [푋], and the remaining being the trivial components corresponding to the vertices in the minority part 푋− of 푋. ƒ It is easy to deduce from the above theorem the following useful characterization of separating cuts in bipartite matching covered graphs. We leave its proof as Exercise 4.2.8. Tight cuts in bipartite graphs Theorem 4.8 Let 퐶 := 휕 ( 푋) be a separating cut in a bipartite matching covered graph 퐺 [ 퐴, 퐵]. Then: (i) | 푋+ | = | 푋− | + 1, (and | 푋 + | = | 푋 − | + 1), (ii) the cut 퐶 is tight, (iii) 푋+ and 푋 + are subsets of different parts of the bipartition of 퐺, and (iv) every edge in the cut 퐶 has one end in 푋+ and one end in 푋 + . (In other words, all edges in 퐶 join vertices in the majority part of one shore of 퐶 to vertices in the majority part of the other shore. See Figure 4.6.)

푋

푋−

푋+

푋 +

푋

퐶

Fig. 4.6 A tight cut in a bipartite matching covered graph

푋 −

4.2 Tight Cuts

65

The fact that all tight cuts in bipartite matching covered graphs conform to the pattern described in the statement of Theorem 4.8 will be found to be very useful in many contexts. As immediate consequences of Theorem 4.7, we have: Corollary 4.9 Every separating cut in a bipartite matching covered graph is tight. ƒ Corollary 4.10 If a separating cut of a matching covered graph is not tight then both ƒ its shores induce graphs that have odd cycles. Corollary 4.11 Both shores of any nontrivial separating cut of a bicritical graph induce graphs that have odd cycles. Consequently, if there do not exist two vertexdisjoint odd cycles in a bicritical graph then the graph is free of nontrivial separating cuts. ƒ

Exercises 4.2.1 Prove Proposition 4.3 ∗ 4.2.2 Give a proof of statement 4.4. By Theorem 4.2, a cut 퐶 of a matching covered graph 퐺 is separating if and only if each edge of 퐺 is in a perfect matching which contains only one edge in 퐶. We now extend the notion of separating cut to a collection of cuts. Cohesive collection of cuts A collection C of cuts of a matching covered graph 퐺 is cohesive if every edge of 퐺 lies in a perfect matching that contains precisely one edge in each cut in C. Consequently, each cut of a cohesive collection of cuts is separating. ⊲4.2.3 Let 퐺 be a matching covered graph and let 퐶 be a separating cut of 퐺. (i) Prove that every collection of tight cuts of 퐺 is cohesive. (ii) Prove that the addition of 퐶 to a collection of tight cuts of 퐺 is cohesive. ∗ 4.2.4 Let 퐶 := 휕 ( 푋) be a separating cut of a matching covered graph 퐺, and let 퐷 be a separating cut of a 퐶-contraction 퐻 := 퐺/푋 of 퐺. (i) Prove that {퐶, 퐷} is cohesive. (ii) Prove that if 퐶 is tight in 퐺 and 퐷 is tight in 퐺/푋 then 퐷 is tight in 퐺. (iii) Give an example where 퐷 is tight in 퐺/푋 but not tight in 퐺. (iv) Give an example of a graph and a pair of separating cuts which is not cohesive. Hint: consider the Petersen graph.

66

4 Tight Cuts

⊲4.2.5 Let 퐺 be a cubic matching covered graph. Show that: (i) every 2-separation cut of 퐺 is also a barrier cut of 퐺, and that (ii) every barrier cut of 퐺 is a 3-edge-cut. ⊲4.2.6 Show that the cut shown in Figure 4.5 is tight, but it is neither a barrier cut, nor a 2-separation cut. 4.2.7 (i) Verify that the Petersen graph P is bicritical. (Hint: use the 4-transitivity of the Petersen graph and the fact that its diameter is equal to two, in order to reduce to two the number of pairs of vertices {푣, 푤} to check that P − 푣 − 푤 has a perfect matching.) (ii) List all the nontrivial separating cuts of the Petersen graph. (Hint: use the result of Corollary 4.10.) (iii) Show that the Petersen graph is free of nontrivial tight cuts. ⊲4.2.8 Give a proof of Theorem 4.8. ⊲4.2.9 Use Theorem 4.8 to show that all tight cuts in a bipartite cubic matching covered graph are 3-edge cuts.

4.3 Tight Cut Decompositions 4.3.1 Bricks and braces Some matching covered graphs do not have nontrivial tight cuts. The reader would be able to verify easily that 퐾2 , 퐶4 , 퐾4 , 퐶6 , 퐾3,3 and the Petersen graph are examples of such graphs. We refer to a matching covered graph which is free of nontrivial tight cuts as a brace if it is bipartite, and as a brick if it is nonbipartite. Among the graphs listed above, 퐾2 , 퐶4 , and 퐾3,3 are braces, and 퐾4 , 퐶6 and the Petersen graph are bricks. Simple ways of recognizing bricks and braces will be described in Chapter 5. (Bricks turn out to be 3-connected bicritical graphs, and braces turn out to be connected 2-extendable bipartite graphs.) A 3-connected cubic graph is essentially 4-edge-connected if it has no 3-edge cuts other than those which are trivial. It follows from Exercise 4.2.9 that every essentially 4-edge-connected bipartite cubic graph is a brace. It is easy to verify that prisms P2푛 , for even 푛 ≥ 4, and M¨obius ladders M2푛 , for odd 푛 ≥ 3, are essentially 4-edge-connected bipartite cubic graphs and, as such, they are braces. Every prism P2푛 , for odd 푛 ≥ 3, is bicritical and has a unique nontrivial separating cut and that cut is not tight (Exercise 4.3.1). M¨obius ladders M2푛 , for even 푛 ≥ 2, are also bicritical and have no nontrivial separating cuts at all, and hence are free of nontrivial tight cuts (Exercise 4.3.2). It follows that prisms P2푛 , for odd 푛 ≥ 3, and M¨obius ladders M2푛 , for even 푛 ≥ 2, are bricks.

67

4.3 Tight Cut Decompositions

Tight cut decomposition Given any matching covered graph 퐺, we may apply to it a procedure, called a tight cut decomposition of 퐺, which produces a list of bricks and braces. If 퐺 itself is a brick or a brace then the list consists of just 퐺. Otherwise, let 퐶 be any nontrivial tight cut of 퐺. Then, both 퐶-contractions of 퐺 are matching covered. One may recursively apply the tight cut decomposition procedure to each 퐶-contraction of 퐺, and then combine the resulting lists to produce a tight cut decomposition of 퐺 itself. As one might expect, the list of bricks and braces obtained by an application of this procedure might depend on the choice of tight cuts. Consider, for example, the graph 퐺 shown in Figure 4.7. The cut 휕 ( 푋) is a tight cut in 퐺 and the corresponding contractions of 퐺 are 퐺 1 := 퐺/( 푋 → 푥), and 퐺 2 := 퐺/( 푋 → 푥).

퐺1 푋

퐺

퐶

푥 푥

퐷

퐺2

Fig. 4.7 A graph 퐺 and its two 휕(푋)-contractions

Note that 퐺 1 is a brace and cannot be decomposed any further because it has no nontrivial tight cuts. However, 퐺 2 has two nontrivial tight cuts, namely 퐶 and 퐷, and 퐺 2 could be decomposed using either 퐶 or 퐷. Let us consider the cut 퐶 first, and denote the two 퐶-contractions of 퐺 2 by 퐺 3 and 퐺 4 . They are both bricks whose underlying simple graphs are 퐾4 ’s. Thus {퐺 1 , 퐺 3 , 퐺 4 } is a tight cut decomposition of 퐺. Considering the second option 퐷, let 퐺 5 and 퐺 6 denote the two 퐷-contractions of 퐺 2 . They are also bricks whose underlying simple graphs are 퐾4 ’s, and thus {퐺 1 , 퐺 5 , 퐺 6 } is another tight cut decomposition of 퐺. The reader will be able to check that although the underlying simple graphs of {퐺 1 , 퐺 3 , 퐺 4 } are the same, in some order, as the underlying simple graphs of {퐺 1 , 퐺 5 , 퐺 6 }, the two lists of graphs are not exactly the same (Exercise 4.3.3). The above example shows that different applications of the tight cut decomposition on a matching covered graph may yield different lists of bricks and braces. However, a remarkable result of Lov´asz (1987, [58]) states that any two applications of the tight cut decomposition procedure on a given graph 퐺 yield the same list of bricks

4 Tight Cuts

68

and braces, up to multiple edges. A proof of this fundamental result will be presented in the next section. It relies on the notion of uncrossing cuts which we define below.

4.3.2 Uncrossing cuts Let 퐺 be a matching covered graph. Let 퐶 := 휕 ( 푋) and 퐷 := 휕 (푌 ) be two odd cuts of 퐺. The four sets 푋 ∩ 푌 , 푋 ∩ 푌 , 푋 ∩ 푌 and 푋 ∩ 푌 are the quadrants defined by 퐶 and 퐷. The cuts 퐶 and 퐷 cross if each of these four quadrants is nonnull. (The cuts 퐶 and 퐷 shown in Figure 4.7 are two tight cuts in 퐺 2 that cross each other.) A collection C of cuts of 퐺 is laminar if no two of its members cross. We note that if two cuts 퐶 and 퐷 are laminar, then one of the shores of 퐶 is a subset of a shore of 퐷. Figure 4.8 shows two crossing cuts, 퐶 and 퐷. 푌

푋

푌

푋 ∩ 푌

푋 ∩ 푌 퐶

푋

푋 ∩ 푌

푋 ∩ 푌

퐷 Fig. 4.8 Crossing cuts 퐶 := 휕(푋), 퐷 := 휕(푌 ), 퐼 := 휕(푋 ∩ 푌 ), 푈 := 휕(푋 ∩ 푌 ) and the four quadrants

If 퐶 := 휕 ( 푋) and 퐷 := 휕 (푌 ) are crossing odd cuts, then one of the two quadrants 푋 ∩ 푌 and 푋 ∩ 푌 is odd and the other is even (and nonempty). We shall often find it convenient to adjust notation so that the quadrant 푋 ∩ 푌 is odd, which implies that 푋 ∩ 푌 and 푋 ∩ 푌 are even and 푋 ∩ 푌 is odd. In addition, if 퐶 and 퐷 are both tight, pictorially we display the four quadrants as in Figure 4.8. An unbroken line between two quadrants indicates the presence of at least one edge between them; a broken line between the quadrants 푋 ∩ 푌 and 푋 ∩ 푌 indicates the possible presence of edges between them; Figure 4.8 depicts the situation in which the four subgraphs 퐺 [푋], 퐺 [푋], 퐺 [푌 ], and 퐺 [푌 ] are connected, and all the four quadrants are nonempty. The following result provides a crucial step in the proof of Lov´asz’s fundamental result mentioned at the beginning of this section.

69

4.3 Tight Cut Decompositions

Uncrossing tight cuts Theorem 4.12 Let 퐶 := 휕 ( 푋) and 퐷 := 휕 (푌 ) be two tight cuts of a matching covered graph 퐺, where | 푋 ∩ 푌 | is odd. Then: (i) no edge of 퐺 joins a vertex in 푋 ∩ 푌 to a vertex in 푋 ∩ 푌 , and (ii) the cuts 퐼 := 휕 ( 푋 ∩ 푌 ) and 푈 := 휕 ( 푋 ∩ 푌 ) are both tight in 퐺. Proof Suppose first that 퐶 and 퐷 do not cross and, without loss of generality, that 푋 ∩ 푌 = ∅. In this case 푋 ⊂ 푌 . Thus 퐼 = 휕 ( 푋) = 퐶 and 푈 = 휕 (푌 ) = 퐷, and hence the assertion holds. So, we may assume that 퐶 and 퐷 do cross. Let 푀 be any perfect matching of 퐺. Denote by 퐹 the set of edges of 퐺 that have an end in 푋 ∩ 푌 and an end in 푋 ∩ 푌 . Clearly, |푀 ∩ 퐶| + |푀 ∩ 퐷| = |푀 ∩ 퐼 | + |푀 ∩ 푈| + 2|푀 ∩ 퐹|. (4.1) Let 푒 be any edge of 퐺, let 푀푒 be a perfect matching of 퐺 that contains the edge 푒. By hypothesis, 퐶 and 퐷 are tight, thus 푀푒 contains just one edge in each of 퐶 and 퐷. The cuts 퐼 and 푈 are both odd, hence 푀푒 contains at least one edge in each cut in {퐼, 푈}. From (4.1) we have that 2 = |푀푒 ∩ 퐶| + |푀푒 ∩ 퐷| = |푀푒 ∩ 퐼 | + |푀푒 ∩ 푈| + 2|푀푒 ∩ 퐹| ≥ 2 + 2|푀푒 ∩ 퐹|. (4.2) We conclude that for each edge 푒 of 퐺 and any perfect matching 푀푒 that contains edge 푒 we have (a) 푒 ∈ 푀푒 , |푀푒 ∩ 퐹| = 0, (b) |푀푒 ∩ 퐶| = 1, |푀푒 ∩ 퐷| = 1, |푀푒 ∩ 퐼 | = 1 and |푀푒 ∩ 푈| = 1. From (a) we deduce that 푒 ∉ 퐹. This conclusion holds for each edge 푒 of 퐺, hence 퐹 is empty. That is, part (i) of the assertion holds. From (b) we conclude that every perfect matching that contains edge 푒 contains just one edge in 퐼 and just one edge in 푈. This conclusion holds for each edge 푒 of 퐺, hence 퐼 and 푈 are both tight cuts. That is, part (ii) of the assertion holds. ƒ The following corollary will prove to be very useful in the proof of Theorem 4.17. Corollary 4.13 Let 퐺 be a matching covered graph, and let 퐶 := 휕 ( 푋) and 퐷 := 휕 (푌 ) be two tight cuts of 퐺 that cross, where | 푋 ∩푌 | is odd. Then the underlying simple graphs of the two graphs 퐺/푋/푋 ∩ 푌 and 퐺/푌 /푋 ∩ 푌 are isomorphic. Proof For the convenience of referring to the two graphs, let us write: 퐺 1 := 퐺/( 푋 → 푥)/( 푋 ∩ 푌 → 푠), and 퐺 2 := 퐺/(푌 → 푦)/( 푋 ∩ 푌 → 푡) See Figure 4.9. Observe that 푉 (퐺 1 ) = ( 푋 ∩ 푌 ) ∪ {푥, 푠}, and 푉 (퐺 2 ) = ( 푋 ∩ 푌 ) ∪ {푦, 푡}. As 퐷 is a tight cut, 퐺 [푌 ] is connected. This implies that 푥 and 푠 are adjacent in 퐺 1 . Similarly, since 퐶 is a tight cut, 퐺 [푋] is connected, implying that 푡 and 푦 are adjacent in 퐺 2 .

4 Tight Cuts

70 푡

푥

푦

푠 푋 ∩ 푌

푋 ∩ 푌

퐺1

퐺2

Fig. 4.9 Isomorphic contractions

Furthermore, as there are no edges between 푋 ∩ 푌 and 푋 ∩ 푌 by Theorem 4.12, it follows that the mapping 휃, where 휃 (푣) := 푣, for each 푣 ∈ 푋 ∩ 푌 ,

휃 (푥) := 푡

and

휃 (푠) := 푦

is an isomorphism between the underlying simple graphs of 퐺 1 and 퐺 2 . We leave the details as Exercise 4.3.4. ƒ Although the underlying simple graphs of 퐺 1 and 퐺 2 are isomorphic, the graphs 퐺 1 and 퐺 2 may not be isomorphic, as the number of edges joining 푥 and 푠 in 퐺 1 may not coincide with the number of edges joining 푡 and 푦 in 퐺 2 . In fact, the number of edges between 푥 and 푠 in 퐺 1 is the number of edges between 푋 and 푋 ∩ 푌 in 퐺, whereas the number of edges between 푡 and 푦 in 퐺 2 is the number of edges between 푋 ∩ 푌 and 푌 in 퐺. Clearly, there is no reason for these two numbers to be the same. Thus 퐺 1 and 퐺 2 may not be isomorphic, but they are isomorphic up to multiple edges!

Uncrossing separating cuts Since tight cuts are special types of separating cuts, it is natural to wonder if the assertion of Theorem 4.12 concerning tight cuts holds more generally for separating cuts. Alas, this is not the case; the cuts 퐼 := 휕 ( 푋 ∩ 푌 ) and 푈 := 휕 ( 푋 ∩ 푌 ) obtained by uncrossing two separating cuts 퐶 := 휕 ( 푋) and 퐷 := 휕 (푌 ) that cross need not be separating cuts (see Exercise 4.3.7). As stated in the next result, what is really required of separating cuts 퐶 and 퐷 for 퐼 and 푈 to be also separating is that each edge of the graph must be in some perfect matching that contains just one edge in 퐶 and just one edge in 퐷. In other words, it is required that the pair {퐶, 퐷} be cohesive. (Recall that a collection C of cuts of a matching covered graph 퐺 is cohesive if each edge of 퐺 is in a perfect matching that contains precisely one edge in each cut in C; see Exercise 4.2.4).

4.3 Tight Cut Decompositions

71

The proof of Theorem 4.14 requires quite minor modifications in the proof of Theorem 4.12 and is left as Exercise 4.3.8. Uncrossing cohesive pairs Theorem 4.14 Let 퐶 := 휕 ( 푋) and 퐷 := 휕 (푌 ) be two separating cuts of a matching covered graph 퐺, where | 푋 ∩ 푌 | is odd. If {퐶, 퐷} is cohesive then: (i) no edge of 퐺 joins a vertex in 푋 ∩ 푌 to a vertex in 푋 ∩ 푌 , (ii) for each perfect matching 푀 of 퐺, |푀 ∩ 퐶| + |푀 ∩ 퐷| = |푀 ∩ 퐼 | + |푀 ∩ 푈|, (iii) the collection {퐶, 퐷, 퐼, 푈} is cohesive, and (iv) the cuts 퐼 and 푈 are separating.

(4.3)

ƒ

If a cut 퐶 is separating and a cut 퐷 is tight then the pair {퐶, 퐷} is cohesive (Exercise 4.2.3). Thus, Corollary 4.15, below, is a consequence of Theorem 4.14. Corollary 4.15 Let 퐶 := 휕 ( 푋) be a separating cut of 퐺, and let 퐷 := 휕 (푌 ) be a tight cut of 퐺. Suppose that | 푋 ∩ 푌 | is odd. Then: (i) no edge of 퐺 joins a vertex in 푋 ∩ 푌 to a vertex in 푌 ∩ 푋, (ii) both 퐼 := 휕 ( 푋 ∩ 푌 ) and 푈 := 휕 ( 푋 ∩ 푌 ) are separating cuts of 퐺, and (iii) if 퐶 is tight then both 퐼 and 푈 are tight whereas if 퐶 is not tight, then at least one of 퐼 and 푈 is not tight. ƒ Note that Corollary 4.15 implies immediately Theorem 4.12. We leave the proof of Corollary 4.15 as Exercise 4.3.9. Corollary 4.16 Let 퐺 be a matching covered graph, and let 퐷 be a tight cut of 퐺. If 퐺 has a separating cut that is not tight, then one of the 퐷-contractions of 퐺 also has a separating cut that is not tight. ƒ The above corollary is used in the proof an important result (Theorem 7.1) in Chapter 7.

Exercises 4.3.1 Consider a prism 퐺 := P2푛 , where 푛 is odd. (i) Prove that 퐺 is bicritical. (ii) Prove that 퐺 has a unique pair of disjoint odd cycles, both of length 푛. (iii) Prove that 퐺 is a brick. (Hint: use the result of Corollary 4.11.)

4 Tight Cuts

72

4.3.2 Consider a M¨obius ladder 퐺 := M2푛 , where 푛 is even. (i) Prove that 퐺 does not have two disjoint odd cycles. (ii) Prove that 퐺 is a brick. (Hint: use the result of Corollary 4.11.) ⊲4.3.3 Show that the graph 퐺 in Figure 4.7 has two tight cut decompositions which are not the same if we take multiplicities of edges into consideration. ⊲4.3.4 Supply the missing details in the proof of Corollary 4.13. ⊲4.3.5 Let 퐺 be a matching covered graph of order 푛 ≥ 4, let L and C be, respectively, a maximal laminar collection of nontrivial tight cuts of 퐺 and the collection of braces and bricks of the corresponding tight cut decomposition of 퐺. (i) Prove that |L| ≤ 푛/2 − 2 and |C| ≤ 푛/2 − 1. Hint: use induction on 푛. (ii) Prove that for each even integer 푛 ≥ 4 there exists a bipartite matching covered graph of order 푛 such that both upper bounds are attained exactly. (iii) Similarly, prove that for each even integer 푛 ≥ 4 there exists a bicritical matching covered graph of order 푛 such that both upper bounds are attained exactly. 4.3.6 Consider the two cuts 퐶 and 퐷 of the graph 퐺 depicted in Figure 4.10. Show that 퐶 is a separating cut which is not tight, that 퐷 is a tight cut, and that 퐶 and 퐷 cross. 퐷

퐶

Fig. 4.10 Two cuts that cross (Exercise 4.3.6)

⊲4.3.7 Give an example to show that the cuts 퐼 and 푈 obtained by uncrossing two crossing separating cuts 퐶 and 퐷 of a matching covered graph need not be separating cuts. (Hint: Try the Petersen graph. See also Exercise 4.2.4.) ∗ 4.3.8 Make minor modifications to the proof of Theorem 4.12 in order to obtain a proof of Theorem 4.14. ∗ 4.3.9 Deduce Corollary 4.15 from Theorem 4.14.

73

4.3 Tight Cut Decompositions

Characteristic of a separating cut Let 퐶 be a separating cut of a matching covered graph 퐺. If 퐶 is tight, then |푀 ∩ 퐶| = 1, for every perfect matching 푀 of 퐺; otherwise, there is some perfect matching 푀 of 퐺 such that |푀 ∩ 퐶| ≥ 3. We define the characteristic 휆(퐶) of 퐶 to be ∞ if 퐶 is tight and, otherwise, to be the minimum of |푀 ∩ 퐶|, where the minimum is taken over the set of all those perfect matchings 푀 of 퐺 which have at least three edges in 퐶.

⊲4.3.10 (i) Show that the Petersen graph has precisely six nontrivial separating cuts (each consisting, coincidentally, of a perfect matching). (ii) Find the characteristics of all the nontrivial separating cuts in the Petersen graph. (iii) Find the characteristics of the four separating cuts 퐶1 , 퐶2 , 퐶3 and 퐶4 in the graph shown in Figure 4.11.

퐶1

퐶2

퐶4

퐶3

Fig. 4.11 Graph for Exercise 4.3.10

∗ 4.3.11 (Uncrossing Separating Cuts) Let 퐶, 퐷, 퐼 and 푈 be as in the statement of Corollary 4.15. Prove that 휆(퐶) = min{휆(퐼), 휆(푈)}. 4.3.12 Let 퐷 be a tight cut of a matching covered graph 퐺, and let 퐶 be a cut of a 퐷-contraction 퐻 of 퐺. (i) Prove that 퐶 is separating in 퐺 if and only if it is separating in 퐻. (ii) Prove that if 퐶 is separating in 퐺 then its characteristics in 퐺 and in 퐻 coincide.

4 Tight Cuts

74

Characteristic of a matching covered graph The characteristic of a matching covered graph 퐺, denoted by 휆(퐺), is the minimum of the characteristics of its separating cuts.

4.3.13 (i) Determine the characteristics of the Petersen graph and the graph in Figure 4.11. (ii) Using the assertion of Exercise 4.3.11, show that the characteristic of a matching covered graph 퐺 is the minimum of the characteristics of its bricks and braces. 4.3.14 Let 퐺 be a matching covered graph, let 퐶 := 휕 ( 푋) be a separating cut of 퐺 which is not tight. Let 퐻 be the underlying simple graph of 퐺/( 푋 → 푥), let 푑 denote the degree of 푥 in 퐻. Show that 휆(퐺) ≤ 푑. Deduce that if 푑 = 3 then 휆(퐺) = 3.

4.4 Uniqueness of Tight Cut Decompositions Suppose that 퐶 is a tight cut of a matching covered graph 퐺. Then any tight cut of a 퐶-contraction of 퐺 is a tight cut of 퐺 that does not cross 퐶. Thus, associated with any tight cut decomposition of 퐺 there is a maximal laminar collection of nontrivial tight cuts of 퐺, and vice versa. Using Theorem 4.12, and Corollary 4.13, Lov´asz [58] proved the following remarkable result on tight cut decompositions. The unique decomposition theorem (1987, [58]) 4.17 Any two applications of the tight cut decomposition procedure to a matching covered graph 퐺 produce the same list of bricks and braces, up to multiple edges. Proof We shall refer to two maximal laminar families C and D of nontrivial tight cuts of 퐺 as equivalent, and write C ≡ D, if they produce the same list of bricks and braces, up to multiple edges. We prove that any two maximal laminar collections of nontrivial tight cuts of 퐺 are equivalent, by induction on the number of vertices. For 푖 = 1, 2, let C푖 be two maximal laminar collections of nontrivial tight cuts of 퐺. We consider various cases. Case 1 Collections C1 and C2 contain a common cut 퐶. Let 퐺 1 and 퐺 2 denote the two 퐶-contractions of 퐺. For 푖, 푗 ∈ {1, 2}, let C푖 푗 denote the restriction of C푖 − 퐶 to 퐺 푗 . For 푖 = 1, 2, any tight cut of 퐺 푖 is also a tight cut of 퐺, it follows that C푖 푗 is a maximal laminar collection of nontrivial tight cuts of 퐺 푗 . By induction, for 푗 = 1, 2, C1 푗 and C2 푗 are equivalent. Thus, C1 and C2 are also equivalent. The assertion holds in this case.

4.4 Uniqueness of Tight Cut Decompositions

75

Case 2 There are cuts 퐶1 ∈ C1 and 퐶2 ∈ C2 that do not cross. By hypothesis, {퐶1 , 퐶2 } is laminar. Let C3 denote a maximal laminar collection of nontrivial tight cuts of 퐺 that includes {퐶1 , 퐶2 }. For 푖 = 1, 2, C푖 and C3 have cut 퐶푖 in common. By Case 1, C1 ≡ C3 ≡ C2 . Thus, C1 and C2 are equivalent. The assertion holds in this case. Case 3 There are cuts 퐶1 := 휕 ( 푋1 ) ∈ C1 and 퐶2 := 휕 ( 푋2 ) ∈ C2 such that | 푋1 ∩ 푋2 | is odd and nontrivial. If 퐶1 and 퐶2 do not cross then Case 2 is applicable. We may thus assume that 퐶1 and 퐶2 cross. Let 퐶3 := 휕 ( 푋1 ∩ 푋2 ). Then, 퐶3 is nontrivial. By Theorem 4.12, cut 퐶3 is tight in 퐺. Let C4 denote a maximal laminar collection of nontrivial tight cuts of 퐺 that contains 퐶3 . Cuts 퐶1 and 퐶3 do not cross. By Case 2, collections C1 and C4 are equivalent. Likewise, cuts 퐶2 and 퐶3 do not cross. By Case 2, collections C2 and C4 are equivalent. In sum, C1 ≡ C4 ≡ C2 . Thus, C1 and C2 are equivalent. The assertion holds in this case. Case 4 None of the previous cases is applicable. If 퐺 is a brick or a brace then the assertion holds trivially. We may thus assume that 퐺 has nontrivial tight cuts. Then, C1 and C2 are both nonempty. For 푖 = 1, 2, let 퐶푖 := 휕 ( 푋푖 ) denote a cut in C푖 . If 퐶1 and 퐶2 do not cross then Case 2 applies. We may thus assume that 퐶1 and 퐶2 cross. Adjust notation so that | 푋1 ∩ 푋2 | is odd, whereupon | 푋1 ∩ 푋2 | is also odd. If | 푋1 ∩ 푋2 | > 1 or if | 푋1 ∩ 푋2 | > 1 then Case 3 applies. We may thus assume that 푋1 ∩ 푋2 and 푋1 ∩ 푋2 are both singletons. Let 푢 denote the only vertex of 푋1 ∩ 푋2 , let 푣 denote the only vertex of 푋1 ∩ 푋2 . Assume that C푖 = {퐶푖 }, for 푖 = 1, 2. By Corollary 4.13, 퐺/푋1 and 퐺/푋2 are isomorphic, up to multiple edges. Likewise, 퐺/푋1 and 퐺/푋2 are also isomorphic, up to multiple edges. Thus, C1 ≡ C2 . The assertion holds in this case. We now prove that C푖 = {퐶푖 }, thereby completing the proof. Assume, to the contrary, that one of C1 and C2 contains two or more cuts. Adjust notation so that C1 − 퐶1 contains a cut 퐷 := 휕 (푌 ). By hypothesis, C1 is laminar, therefore 퐶1 and 퐷 do not cross. Adjust notation so that one of 푌 and 푋1 is a subset of the other. Adjust notation, by complementing the three sets 푋1 , 푋2 and 푌 if necessary, so that 푌 ⊂ 푋1 . Then, (4.4) 푌 ∩ 푋2 ⊆ 푋1 ∩ 푋2 = {푢} and {푣} = 푋1 ∩ 푋2 ⊆ 푌 ∩ 푋2 . Cuts 퐶2 and 퐷 cross, otherwise Case 2 applies. Thus, 푌 ∩ 푋2 is nonempty. From (4.4) we deduce that 푌 ∩ 푋2 = 푋1 ∩ 푋2 = {푢}. Then, |푌 ∩ 푋2 | is odd. If |푌 ∩ 푋2 | > 1 then Case 3 applies. We may thus assume that |푌 ∩ 푋2 | = 1. From (4.4), we deduce that 푌 ∩ 푋2 = 푋1 ∩ 푋2 , whence 푌 ∪ 푋2 = 푋1 ∪ 푋2 . In sum, 푌 ∩ 푋2 = 푋1 ∩ 푋2

and 푌 ∪ 푋2 = 푋1 ∪ 푋2 .

76

4 Tight Cuts

We conclude that 푌 and 푋1 coincide, a contradiction. As assumed, C푖 = {퐶푖 }, for ƒ 푖 = 1, 2.

4.4.1 The number of bricks One particular consequence of Lov´asz’s Theorem is that any two tight cut decompositions of a given matching covered graph 퐺 yield the same numbers of bricks and braces. We refer to these two invariants as the numbers of bricks and braces of 퐺, respectively. For example, the graph 퐺 shown in Figure 4.7 has two bricks and one brace. We denote the number of bricks of a matching covered graph 퐺 by 푏(퐺), or simply 푏 if 퐺 is understood. Let 퐻 be any given simple brick or brace, such as the Petersen graph P or 퐾3,3 . A specific implication of Theorem 4.17 is that the number of graphs in a tight cut decomposition of 퐺 whose underlying simple graphs are isomorphic to 퐻 does not depend on the choice of tight cuts used in the decomposition. In a later chapter, the number of ‘Petersen bricks’ will play an important role.

4.4.2 Near-bricks A near-brick is a matching covered graph with a single brick. Near-bricks feature prominently in many fundamental results which will be described in Part II. The following result is a simple consequence of Proposition 4.3. Proposition 4.18 Every tight cut of a near-brick is a special-barrier cut. Proof Let 퐶 := 휕 ( 푋) be a tight cut of a near-brick 퐺. By definition, 푏(퐺) = 1. It follows that precisely one of the 퐶-contractions of 퐺 is bipartite. Adjusting notation, if necessary, we may suppose that 퐺 1 := 퐺/( 푋 → 푥) is not bipartite and 퐺 2 := 퐺/( 푋 → 푥) is bipartite, with bipartition, say ( 퐴, 퐵), with 푥 belonging to 퐴. Then 퐵 is a special barrier of 퐺, 퐺 [푋] is the nontrivial component of 퐺 − 퐵 and 퐶 is a special barrier cut associated with 퐵. ƒ As bicritical graphs do not have nontrivial barriers, we have the following obvious corollary to the above assertion: Corollary 4.19 Every bicritical near-brick is a brick.

ƒ

If 퐵 is any barrier of a near-brick 퐺, then one of the components of 퐺 − 퐵 is nonbipartite. But 퐺 − 퐵 may also have nontrivial bipartite components (see, for example, the near-brick shown in Figure 4.2). Thus, not every barrier of a near-brick is special. However, every maximal barrier in a near-brick is special (Exercise 4.4.5).

77

4.5 Subadditivity of the Number of Bricks

Exercises 4.4.1 Construct a matching covered graph with two bricks and three braces such that (i) the underlying simple graphs of the bricks are 퐶6 and P, and (ii) the underlying simple graphs of the braces are 퐾3,3 and two copies of 퐾2,2 . ⊲4.4.2 Show that if 퐺 is bipartite, then 푏(퐺) = 0. (Hint: Use Theorem 4.7.) ⊲4.4.3 For any matching covered graph 퐺, let 푏 0 (퐺) denote the number of braces of 퐺. (i) Prove that if 퐺 has order 푛 ≥ 4 then 푏(퐺) + 푏 0 (퐺) ≤ 푛/2 − 1. (ii) Give an infinite family of bipartite matching covered graphs 퐺 of order 푛 such that 푏 0 (퐺) = 푛/2 − 1. (iii) Give an infinite family of matching covered graphs 퐺 of order 푛 such that 푏(퐺) = 푛/2 − 1. 4.4.4 Let 퐺 be a nonbipartite matching covered graph. Prove that 퐺 has a laminar collection C of 푏 − 1 nontrivial tight cuts such that the (possibly partial) tight cut decomposition produced by these 푏 − 1 cuts consists of 푏 near-bricks. 4.4.5 Prove that every maximal barrier in a near-brick is special.

4.5 Subadditivity of the Number of Bricks ♯ The number of bricks 푏(퐺) of a matching covered graph 퐺 is additive over the tight cuts of 퐺 in the sense that if 퐺 1 and 퐺 2 are the two cut-contractions with respect to a tight cut of 퐺, then 푏(퐺) = 푏(퐺 1 ) + 푏(퐺 2 ). This property, as we shall see, distinguishes tight cuts from separating cuts which are not tight; the function 푏(퐺) is subadditive over separating cuts of 퐺. This property will only be used in the second part of the book, in the proof of a conjecture of Lovasz ´ concerning bricks.

Subadditivity of function 푏(퐺) Theorem 4.20 Let 퐺 be a matching covered graph. Let 퐶 be a separating cut of 퐺, and let 퐺 1 and 퐺 2 be the two 퐶-contractions of 퐺. Then 푏(퐺) ≤ 푏(퐺 1 ) + 푏(퐺 2 ),

(4.5)

with equality if, and only if, 퐶 is a tight cut.

Proof If 퐶 is a tight cut of 퐺 then, by Theorem 4.17, it follows that (4.5) holds with equality. So, suppose that 퐶 is a separating cut that is not tight. Let 푋 be a shore of 퐶, and take 퐺 1 to be 퐺/푋 and 퐺 2 to be 퐺/푋. We shall prove the desired inequality by induction on |푉 (퐺)|.

4 Tight Cuts

78

Firstly observe that if either 퐺 1 or 퐺 2 is bipartite then 퐶 would be a barrier cut of 퐺, implying that 퐶 is tight in 퐺. We may therefore assume that 푏(퐺 1 ) ≥ 1 and 푏(퐺 2 ) ≥ 1. Now, if 퐺 is free of nontrivial tight cuts then the desired inequality is satisfied. Therefore we may assume that 퐺 has a nontrivial tight cut, say 퐷 := 휕 (푌 ). Let 퐻1 = 퐺/푌 and 퐻2 = 퐺/푌 denote the two 퐷-contractions of 퐺. The case in which 퐶 and 퐷 do not cross is easier and is left as Exercise 4.5.1. See Figure 4.12.

푋 ∩ 푌

푋

퐺11

퐺1

푋 ∩ 푌

퐺12

푌

푋 푌

퐻1

퐶

퐷

푌

퐻2

Graph 퐺

푋 ∩ 푌

푋

푋 ∩ 푌

퐺21

퐺2

퐺22

Fig. 4.12 Illustration for the proof of Theorem 4.20

Adjust notation, by interchanging 푌 with 푌 if necessary, so that | 푋 ∩ 푌 | is odd. Let 퐼 := 휕 ( 푋 ∩ 푌 ) and 푈 := 휕 ( 푋 ∩ 푌 ). The cut 퐶 is separating and the cut 퐷 is tight. By Corollary 4.15, (i) no edge of 퐺 joins a vertex in 푋 ∩ 푌 to a vertex in 푋 ∩ 푌 , (ii) each one of the cuts 퐼 and 푈 is separating in 퐺, and (iii) at least one of 퐼 and 푈 is not tight. Let 푀 be any perfect matching of 퐺. From (i) and recalling that 퐷 is tight we infer that

79

4.5 Subadditivity of the Number of Bricks

|푀 ∩ 퐼 | + |푀 ∩ 푈| = |푀 ∩ 퐶| + 1.

(4.6)

Every perfect matching of 퐺 1 can be extended to a perfect matching of 퐺 that meets 퐶 in just one edge. Thus, it follows from (4.6) that 퐼 is a tight cut of 퐺 1 . Similarly, 푈 is a tight cut of 퐺 2 . Let 퐺 11 := 퐺/( 푋 ∪ 푌 ), and observe that 퐺 11 is one of the 퐼-contractions of the graph 퐺 1 . Now let 퐺 12 denote the other 퐼-contraction of 퐺 1 . Similarly, let 퐺 22 := 퐺/( 푋 ∪ 푌 ), and observe that 퐺 22 is one of the 푈-contractions of 퐺 2 . Now let 퐺 21 denote the other 푈-contraction of 퐺 2 . Finally, note that 퐺 11 and 퐺 21 are the 퐼-contractions of 퐻1 ; and similarly, 퐺 12 and 퐺 22 are the 푈-contractions of 퐻2 . See Figure 4.12. We then have 푏(퐺) = 푏(퐻1 ) + 푏(퐻2 ), 푏(퐻1 ) ≤ 푏(퐺 11 ) + 푏(퐺 21 ),

푏(퐻2 ) ≤ 푏(퐺 12 ) + 푏(퐺 22 ), 푏(퐺 11 ) + 푏(퐺 12 ) = 푏(퐺 1 ), 푏(퐺 21 ) + 푏(퐺 22 ) = 푏(퐺 2 ).

(4.7) (4.8) (4.9) (4.10) (4.11)

The validity of (4.7) follows from the fact that 퐷 is tight in 퐺. Since the cut 퐼 is a separating cut in 퐺 and 퐷 is tight in 퐺, it follows that 퐼 is a separating cut in 퐻1 . Likewise, the cut 푈 is a separating cut in 퐻2 . The validity of the inequalities (4.8) and (4.9) follows from the induction hypothesis; moreover, since at least one of 퐼 and 푈 is not tight in 퐺 it follows that either 퐼 is not tight in 퐻1 or 푈 is not tight in 퐻2 . Therefore, at least one of (4.8) and (4.9) is strict. Finally, the facts that 퐼 is tight in 퐺 1 and that 푈 is tight in 퐺 2 , imply the two equations (4.10) and (4.11). By adding up inequalities (4.7)-(4.11), and simplifying, we obtain the asserted (strict) inequality. ƒ Robust Cuts A separating cut 퐶 of a matching covered graph 퐺 is robust if 퐶 is not tight and both 퐶-contractions of 퐺 are near-bricks. From the subadditivity of function 푏 we now derive the following consequence which indicates the importance of robust cuts. 4.21 If a matching covered graph has a robust cut then it is a near-brick. Proof Let 퐺 be a matching covered graph and assume that 퐺 has a robust cut 퐶. Let 퐺 1 and 퐺 2 denote the two 퐶-contractions of 퐺. By definition, 퐶 is not tight, and both 퐺 1 and 퐺 2 are near-bricks. By the subadditivity of function 푏, we have that 푏(퐺) < 푏(퐺 1 ) + 푏(퐺 2 ) = 2. Thus, 푏(퐺) ∈ {0, 1}. The cut 퐶 is separating but not tight. In a bipartite matching covered graph every separating cut is tight (Corollary 4.9). Thus, 퐺 is not bipartite. We conclude that 푏(퐺) = 1. As asserted, 퐺 is a near-brick. ƒ

80

4 Tight Cuts

Exercises ⊲4.5.1 Complete the proof of Theorem 4.20 in the case where 퐶 and the nontrivial tight cut 퐷 of 퐺 do not cross. ∗ 4.5.2 For any integer 푘 ≥ 2, show that there exists a brick 퐺, with a separating cut 퐶, such that 푏(퐺 ′ ) + 푏(퐺 ′′ ) = 푘, where 퐺 ′ and 퐺 ′′ are the two 퐶-contractions of 퐺.

4.6 Notes Tight cut decompositions and the elegant Uniqueness Theorem 4.17 were first described in the landmark paper by Lov´asz (1987, [58]). But a related procedure known as a ‘brick decomposition procedure’ predates this paper. This procedure is essentially the tight cut decomposition procedure where one restricts oneself to cut-contractions with respect to barrier cuts and 2-separation cuts. The term ‘brace’ appeared for the first time in [58], but the notion of a ‘brick’ appeared in earlier papers (see [59]). In those works, a brick was defined as a 3-connected bicritical graph. It turns out that this is equivalent to the definition that a brick (as defined in this chapter) is a nonbipartite matching covered graph which is free of nontrivial tight cuts. (This is an important result due to Edmonds, Lov´asz and Pulleyblank (1982, [32]) which, together with a new proof, will be described in the next chapter.) In Chapter 5 of their book entitled Matching Theory (1982, [59]), Lov´asz and Plummer remark that the collection of bricks determined by the brick decomposition procedure is not unique. However, they present, in Chapter 7 of their book, a proof of the fact that any two applications of the brick decomposition procedure on a matching covered graph 퐺 yield the same number of bricks, implying that this number 푏(퐺) is an invariant of 퐺. This is established, indirectly, by relating it to the dimension of the perfect matching polytope of 퐺. (This connection will be described in Chapter 6.) The power and beauty of the Unique Decomposition Theorem 4.17 is that any two applications of the brick decomposition procedure (more generally, of the tight cut decomposition procedure) on a matching covered graph 퐺 yield the same collection of bricks up to multiple edges. This directly implies that 푏(퐺) is an invariant of 퐺.

Chapter 5

Characterizations of Bricks and Braces

Contents 5.1 5.2

5.3 5.4 5.5

5.6

Characterization of Bricks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81 Proof of the ELP Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83 5.2.1 Peripheral tight cuts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83 5.2.2 Dulmage-Mendelsohn barriers . . . . . . . . . . . . . . . . . . . . . . . 85 5.2.3 A key lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86 5.2.4 The ELP Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87 Some Applications of the ELP Theorem . . . . . . . . . . . . . . . . . . . . . . . 92 Characterizations of Braces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96 Tight Cuts in Graphs with Two Bricks . . . . . . . . . . . . . . . . . . . . . . . . . 99 5.5.1 The laminar ELP Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . 106 5.5.2 ELP tight cut decompositions . . . . . . . . . . . . . . . . . . . . . . . . 108 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108

5.1 Characterization of Bricks Recall that a matching covered graph which is free of nontrivial tight cuts is a brace if it is bipartite, and is a brick if it is nonbipartite. In Chapter 4 we presented a proof of Lov´asz’s Uniqueness Theorem (4.17) for tight cut decompositions, which is one of the cornerstones of this theory. According to that theorem, every matching covered graph has a unique decomposition into bricks and braces (up to multiple edges). The proof of this fundamental result simply relied on the fact that bricks and braces are free of nontrivial tight cuts. But this fact by itself is inadequate to suggest efficient procedures for recognizing these basic objects. In this chapter, we address these issues by establishing characterizations of bricks and braces which lead to polynomial-time recognition algorithms. We deal with bricks in this section and braces in Section 5.4.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7_5

81

5 Characterizations of Bricks and Braces

82

Let us begin by recalling the definitions of barrier cuts and 2-separation cuts. For any barrier 퐵 of a matching covered graph 퐺, and any odd component 퐾 of 퐺 − 퐵, the cut 휕 (퐾) is a tight cut of 퐺 (see Figure 4.1(b) for an example). Cuts which arise in this manner are barrier cuts of 퐺. (The set {휕 (퐾), 퐾 ∈ O(퐺 − 퐵)} is the set of barrier cuts associated with the barrier 퐵.) As bicritical graphs do not have nontrivial barriers, they do not have nontrivial barrier cuts. Recall that a 2-separation of a matching covered graph 퐺 is a 2-vertex cut of 퐺 which is not a barrier. Thus, if {푢, 푣} is a 2-separation of 퐺, the graph 퐺 − 푢 − 푣 has no odd components. Suppose that 퐻1 is the union of a nonempty proper subset of the components of 퐺 − 푢 − 푣, and 퐻2 is the union of the remaining components of 퐺 − 푢 − 푣. Then 퐻1 and 퐻2 are two vertex-disjoint even order subgraphs whose union is 퐺 − 푢 − 푣. The cuts 휕 (푉 (퐻1) + 푣) and 휕 (푉 (퐻2 ) + 푣) are two tight cuts of 퐺 which cross each other (see Figure 4.3 for an example). Cuts which arise in this manner are 2-separation cuts. Clearly, 3-connected graphs cannot have 2-separation cuts. Edmonds, Lov´asz and Pulleyblank (1982, [32]) proved that if a matching covered graph has a nontrivial tight cut then it has a nontrivial barrier cut or a (nontrivial) 2-separation cut. We refer to this assertion as the ELP Theorem and use the term ELP cut to mean either a barrier cut or a 2-separation cut. As we noted in Chapter 4, not all tight cuts in a matching covered graph need be ELP cuts. For example, the cut 퐶 in the matching covered graph shown in Figure 5.1 is not an ELP cut; however, the cut 퐷 is.

퐷 퐶 Fig. 5.1 The cut 퐶 is not an ELP cut, but 퐷 is a 2-separation cut

The proof of the above mentioned theorem by its three authors was based on linear programming duality. About twenty years later, Szigeti (2002, [87]) presented a simpler and purely graph theoretical proof. The proof we give here is inspired by the one given by Szigeti. It relies on the properties of special types of barriers in matchable graphs, which we call Dulmage-Mendelsohn barriers. But before we introduce them, we describe a property that is shared by all matching covered graphs with nontrivial tight cuts.

5.2 Proof of the ELP Theorem

83

5.2 Proof of the ELP Theorem ♯♯ 5.2.1 Peripheral tight cuts A nontrivial tight cut 퐶 of a matching covered graph 퐺 is peripheral if one of its shores, say 푋, has the property that there is no nontrivial proper subset 푋 ′ of 푋 such that 휕 ( 푋 ′) is a tight cut of 퐺. In other words, 퐶 is peripheral if at least one of the 퐶-contractions of 퐺 is free of nontrivial tight cuts. Clearly, any matching covered graph which has nontrivial tight cuts has a peripheral tight cut. We shall establish a property which all such tight cuts have. This will prove to be a key step in our proof of the ELP Theorem. A property of peripheral tight cuts Lemma 5.1 Let 퐶 := 휕 ( 푋) be a peripheral tight cut of 퐺. Then there exists an edge 푒 := 푢푣 ∈ 퐶, with 푢 ∈ 푋 and 푣 ∈ 푋, such that both the subgraphs 퐺 [푋 −푢] and 퐺 [푋 − 푣] are connected. Proof By hypothesis, 퐶 is peripheral. Adjust notation so that no cut 휕 ( 푋 ′) is tight, for any nontrivial proper subset 푋 ′ of 푋. Consider the two 퐶-contractions 퐺/푋 := 퐺/( 푋 → 푥) and 퐺/푋 := 퐺/( 푋 → 푥) of 퐺, and note that 퐺 [푋] = (퐺/푋) − 푥 and 퐺 [푋] = (퐺/푋) − 푥. As 퐶 is a tight cut of 퐺, both 퐺/푋 and 퐺/푋 are matching covered. Furthermore, by the minimality of 푋, the graph 퐺/푋 is free of nontrivial tight cuts. Let us first show that 퐶 contains an edge 푒 := 푢푣, where 푣 lies in 푋, such that 퐺 [푋 − 푣] is connected. The graph 퐺/푋, being matching covered, is 2-connected, and therefore the graph 퐺 [푋] = (퐺/푋) − 푥 is connected. If 퐺 [푋] happens to be 2-connected then 퐺 [푋 − 푣] would be connected, for every vertex 푣 of 푋. We may thus assume that 퐺 [푋] has two or more blocks. Then it has an end block 퐹 that contains precisely one cut vertex, say 푤, of 퐺 [푋]. Now, as 퐺/푋 is 2-connected, it follows that 퐹 contains a vertex 푣, different from 푤, which is incident with an edge 푒 = 푢푣 of 퐶 (otherwise, 푤 would be a cut vertex of 퐺). Since 푣 ≠ 푤, it is not a cut vertex of 퐺 [푋], and hence 퐺 [푋 − 푣] is connected (see Figure 5.2). To complete the proof, we now proceed to show that the graph 퐺 [푋 − 푢] is connected. For this, assume the contrary, that 퐺 [푋 − 푢] is disconnected. As the graph 퐺/푋 is matching covered, and 퐺 [푋 − 푢] = (퐺/푋) − 푥 − 푢, it follows that {푢, 푥} is either a barrier or a 2-separation of 퐺/푋. But the ends of 푒 in 퐺/푋 are 푢 and 푥. Moreover, as 퐺/푋 is matching covered, the edge 푒 is matchable in 퐺/푋. Consequently, {푢, 푥} is not a barrier of 퐺/푋. It follows that {푢, 푥} is a 2-separation of 퐺/푋. Thus, 퐺/푋 has nontrivial tight cuts, contradicting the minimality of 푋. This proves the assertion. ƒ

5 Characterizations of Bricks and Braces

84 푋

푋

퐹

퐺 [푋]

퐺 [푋]

푣 푢 푤

Fig. 5.2 An edge 푢푣 in 퐶 such that 퐺 [푋 − 푣 ] is connected

It should be noted that the property of peripheral tight cuts established above is not valid in general for all nontrivial tight cuts in matching covered graphs. For example, the cut 퐶 which is a nonperipheral tight cut in the matching covered graph shown in Figure 5.3 does not have the desired property (see Exercise 5.2.2). 퐶

푋

푋

Fig. 5.3 A nonperipheral cut in a matching covered graph

We shall use Lemma 5.1 to show that any matching covered graph 퐺 which has nontrivial tight cuts either has a nontrivial barrier (associated with a nontrivial barrier cut), or it has a 2-separation (in which case, it has a 2-separation cut). We are able to do this by considering matchable graphs which are closely related to the graph 퐺. For example, with the notation used in the statement of Lemma 5.1, the graph 퐺 − 푢 − 푣 is a matchable graph in which no edge in the cut 휕 ( 푋 − 푢) is matchable (Exercise 5.2.1). Under certain conditions, it turns out that this graph 퐺 − 푢 − 푣 has a barrier which can be extended to a barrier of 퐺 by adding to it either 푢 or 푣. This line of inquiry has led us to study the existence of a special type of barriers in matchable graphs which we now proceed to introduce.

5.2 Proof of the ELP Theorem

85

5.2.2 Dulmage-Mendelsohn barriers Recall that if 퐵 is a barrier of a matchable graph 퐺, the core of 퐺 with respect to B, denoted by H(퐵), is the bipartite graph obtained from 퐺 by deleting edges with both ends in 퐵, deleting the vertices in the even components of 퐺 − 퐵, and then shrinking each odd component of 퐺 − 퐵 to a single vertex (see Section 3.2). Dulmage-Mendelsohn barriers A nonempty barrier 퐵 of a matchable graph 퐺 is a Dulmage-Mendelsohn barrier if: • the core H(퐵) of 퐺 with respect to the barrier 퐵 is matching covered, and • each odd component of 퐺 − 퐵 is critical. For brevity, we shall refer to a Dulmage-Mendelsohn barrier as a DM-barrier. DM-barriers have a fundamental property that will be used in our proof of the ELP Theorem (Exercise 5.2.3). Lemma 5.2 Let 퐵 denote a DM-barrier of a matchable graph 퐺. Then every edge of H(퐵) is matchable in 퐺. ƒ An essential ingredient in our proof of the ELP Theorem is the assertion that every maximal barrier of a matchable graph contains a DM-barrier as a subset. Thus, let 퐺 be any matchable graph, and let 퐵∗ be a maximal barrier of 퐺. Then 퐵∗ is one of the parts of the bipartition of the core H(퐵∗ ) of 퐺 with respect to 퐵∗ . The other part of the bipartition of H(퐵∗ ), which we denote by 퐴∗ , consists of vertices resulting from shrinking the odd components of 퐺 − 퐵∗ . Proposition 2.11, applied to the bipartite graph H := H[ 퐴∗ , 퐵∗ ], yields the following result: 5.3 There exists a nonempty subset 푆 of 퐴∗ such that the subgraph of H induced by the set 푆 ∪ 푁 (푆), where 푁 (푆) is the neighbour set of 푆 in H, is matching covered. ƒ Maximal barriers and DM-barriers Lemma 5.4 Every maximal barrier 퐵∗ of a matchable graph 퐺 includes a Dulmage-Mendelsohn barrier 퐵. Moreover, every odd component of 퐺 − 퐵 is an odd component of 퐺 − 퐵∗ . Proof As described above, let H := H[ 퐴∗ , 퐵∗ ] denote the core of 퐺 with respect to the barrier 퐵∗ , where 퐴∗ is the set of vertices resulting from shrinking odd components of 퐺 − 퐵∗ . By Theorem 3.2 every component of 퐺 − 퐵∗ is odd and critical. Furthermore, for any perfect matching of 퐺, the set 푀 ∩ 퐸 (H) is a perfect matching of the core. Thus, H[ 퐴∗ , 퐵∗ ] is a matchable bipartite graph. Hence, by (5.3), there exists a nonempty subset 푆 of 퐴∗ such that the subgraph of H induced by 푆 ∪ 푁 (푆) is matching covered. It follows that 퐵 := 푁 (푆) is a nonempty barrier of

86

5 Characterizations of Bricks and Braces

H. Moreover, the core H(퐵) of 퐺 with respect to 퐵 is the same as the subgraph of H[ 퐴∗ , 퐵∗ ] induced by 푆 ∪ 퐵, implying that H(퐵) is matching covered. As the barrier 퐵∗ is maximal, each odd component of 퐺 − 퐵∗ is critical by Theorem 3.2. Thus, each odd component of 퐺 − 퐵, which is also an odd component of 퐺 − 퐵∗ , is critical. We ƒ conclude that 퐵 is a DM-barrier of 퐺.

5.2.3 A key lemma There is one more intermediate result that is needed before we are able to present a proof of the ELP Theorem. It asserts the existence of special types of DM-barriers in matchable graphs with certain very specific properties. DM-barriers included in shores of cuts Lemma 5.5 Let 퐺 be a matchable graph, and let 푋 be a nonempty proper subset of 푉 (퐺) such that: 1. both the subgraphs 퐺 [푋] and 퐺 [ 푋] are connected, and 2. no edge in the cut 휕 ( 푋) is matchable in 퐺. Then 퐺 has a DM-barrier 퐵 which is a subset of one of the shores of 휕 ( 푋) (that is, a subset of 푋 or of 푋). Furthermore, the vertex sets of all the odd components of 퐺 − 퐵 are also subsets of that same shore. Proof By hypothesis, 퐺 has perfect matchings but no edge of 휕 ( 푋) is matchable. It follows that the graphs 퐺 [푋] and 퐺 [푋] both have perfect matchings. Consider first the case in which 휕 ( 푋) is empty. Let 퐵★ be any maximal barrier of 퐺 [푋]. By Lemma 5.4, 퐺 [푋] has a DM-barrier 퐵 that is a subset of 퐵★. This barrier 퐵 of 퐺 [푋] is also a DM-barrier of 퐺. The assertion holds in this case. We may thus assume that 휕 ( 푋) is nonempty. Let 푒 := 푢푣 denote an edge of 휕 ( 푋), where 푢 ∈ 푋 and 푣 ∈ 푋. By hypothesis, 푒 is not matchable. By Corollary 2.2, 퐺 has a barrier that contains both 푢 and 푣. Let 퐵★ be a maximal barrier of 퐺 that contains both 푢 and 푣. By Lemma 5.4, some subset 퐵 of 퐵★ is a DM-barrier of 퐺. Let us first show that if 퐾 is any odd component of 퐺 − 퐵 then its vertex set 푉 (퐾) is a subset of one of the shores of the cut 휕 ( 푋). Suppose that this is not the case. Then, both 푉 (퐾) ∩ 푋 and 푉 (퐾) ∩ 푋 are nonempty. One of these sets has to be even and the other odd because 푉 (퐾) is an odd set. Without loss of generality, assume that |푉 (퐾) ∩ 푋 | is even. Clearly, 퐾 is also an odd component of 퐺 − 퐵∗ . Since 퐵∗ ∩ 푋 is not empty, 푉 (퐾) ∩ 푋 is a nonempty proper subset of 푋. As 퐺 [푋] is connected by the hypothesis, and 퐾 is a component of 퐺 − 퐵, it follows that there is some edge, say 푒 1 , which joins a vertex in 푉 (퐾) ∩ 푋 to a vertex in 퐵 ∩ 푋. That edge 푒 1 , being an edge of the graph H(퐵), is matchable in 퐺, by Lemma 5.2. Let 푀1 be a perfect matching of 퐺 containing 푒 1 . Since 퐾 is an odd component of 퐺 − 퐵, the edge 푒 1 is the only edge of 푀1 in 휕 (푉 (퐾)). However, since 푉 (퐾) ∩ 푋 is an even set, |푀1 ∩ 휕 ( 푋 ∩ 푉 (퐾))| is even. This implies that some edge with one

5.2 Proof of the ELP Theorem

87

end in 푉 (퐾) ∩ 푋 and one end in 푉 (퐾) ∩ 푋 is in 푀1 . This is impossible because, by hypothesis, no edge in 휕 ( 푋) is matchable. We conclude that 푉 (퐾) is a subset of one of the shores of 휕 ( 푋). Now observe that since 퐵 is a DM-barrier, graph H(퐵) is matching covered. Thus H(퐵) is connected. Moreover, by Lemma 5.2, each edge of H(퐵) is matchable in 퐺. As no edge of 휕 ( 푋) is matchable in 퐺, it follows that 퐵 and the vertex sets of all the odd components of 퐺 − 퐵 are all subsets of one and the same shore of 휕 ( 푋). ƒ

5.2.4 The ELP Theorem

´ Edmonds-Lovasz-Pulleyblank (1982, [32]) Theorem 5.6 If a matching covered graph has a nontrivial tight cut, then it has a nontrivial tight cut which is an ELP cut. Proof We begin the proof by making the following observation: 5.6.1 If 퐺 is not bicritical then it has a nontrivial barrier cut. Proof Suppose that 퐺 is not bicritical; let 퐵 denote a nontrivial barrier of 퐺. As 퐺 is matching covered, it follows that every component of 퐺 − 퐵 is odd. For each component 퐾 of 퐺 − 퐵, the cut 휕 (퐾) is a barrier cut. If some component 퐾 of 퐺 − 퐵 is nontrivial then 휕 (퐾) is also nontrivial, because 퐵 is nontrivial. Suppose then that every component of 퐺 − 퐵 is trivial. In that case, the graph 퐺 is bipartite, where 퐵 is one of the parts of the bipartition of 퐺. By Theorem 4.8, every tight cut of 퐺 is a barrier cut. By hypothesis, 퐺 has nontrivial tight cuts. ƒ We shall now analyse several possibilities, all but one of which leads to the conclusion that 퐺 is not bicritical, a case already considered. Finally, one last case will lead to the conclusion that 퐺 has a 2-separation. Let 퐶 := 휕 ( 푋) be a nontrivial tight cut of 퐺, and let 푒 := 푢푣, with 푢 ∈ 푋 and 푣 ∈ 푋, be an edge of 퐶 such that both the subgraphs graphs 퐺 [푋 − 푢] and 퐺 [푋 − 푣] are connected. (The existence of such a cut 퐶, and edge 푒, is guaranteed by Lemma 5.1.) As 퐶 is tight, both 퐶-contractions of 퐺 are matching covered, hence 2-connected. Thus, 퐺 [푋] and 퐺 [푋] are also connected. In sum, the four graphs 퐺 [푋], 퐺 [푋 − 푢], 퐺 [푋] and 퐺 [푋 − 푣] are connected. The rest of the proof consists of analysing various cases depending on the neighbourhoods of 푢 and 푣. Case 1 The vertex 푣 is the only neighbour of 푢 in 푋 and 푢 is the only neighbour of 푣 in 푋. Let 퐺 ′ := 퐺 −푢 −푣. As 퐺 is matching covered, it has perfect matchings that contain 푒, and the restriction of any such perfect matching of 퐺 to 퐸 (퐺 ′ ) is a perfect matching of 퐺 ′ . Therefore, 퐺 ′ is matchable. Moreover,

5 Characterizations of Bricks and Braces

88

퐶 − 푒 = 휕퐺 ′ ( 푋 − 푢) = 휕퐺 ′ ( 푋 − 푣) is a cut of 퐺 ′ . For every perfect matching 푀 ′ of 퐺 ′ , the set 푀 ′ + 푒 is a perfect matching of 퐺. As 퐶 is tight in 퐺, the edge 푒 has to be the only edge of 푀 ′ + 푒 in 퐶. It follows that no edge of 퐶 − 푒 is matchable in 퐺 ′ . By Lemma 5.5, the graph 퐺 ′ has a DM-barrier 퐵′ such that 퐵′ , as well as the vertex sets of all the odd components of 퐺 ′ − 퐵′ , are subsets of one of 푋 − 푢 and 푋 − 푣. Adjust notation so that 퐵′ ⊆ 푋 − 푢. (See Figure 5.4 for an illustration, and Exercise 5.2.4 for a concrete example.)

푋

퐵′

푋 푢 퐶

푒 푋

푣

퐶

푋

퐺

퐺 − 푢 − 푣

Fig. 5.4 휕(푢) ∩ 푋 = {푣 } and 휕(푣 ) ∩ 푋 = {푢}

Let 퐵 := 퐵′ + 푢. By hypothesis, 푢 is the only vertex of 푋 adjacent to 푣. Thus, 푣 is not adjacent to any vertex in any of the |퐵| − 1 odd components of 퐺 ′ − 퐵′ . It follows that all the |퐵| − 1 odd components of 퐺 ′ − 퐵′ are also odd components of 퐺 − 퐵. By parity, 퐵 is a nontrivial barrier of 퐺. We deduce that 퐺 is not bicritical. By statement 5.6.1, 퐺 has a nontrivial barrier cut. Case 2 Either 푢 has two or more neighbours in 푋, or 푣 has two or more neighbours in 푋. We may adjust notation so that 푢 has at least two neighbours in 푋, and let 푅 := 휕 (푢) − 퐶. As 퐺 [푋] is connected and nontrivial, the set 푅 is not empty. Now consider the graph 퐺 ′′ := 퐺 − 푅, together with the cut 퐷 := 휕 ( 푋 −푢) − 푅 = 휕 ( 푋 +푢) − 푅. (See Figure 5.5.) 5.6.2 The graphs 퐺 ′′ [푋 − 푢] and 퐺 ′′ [푋 + 푢] are both connected. Furthermore, the graph 퐺 ′′ is matchable and no edge in the cut 퐷 is matchable in 퐺 ′′ . Proof Note that the graph 퐺 ′′ [푋 − 푢] is the same as the graph 퐺 [푋 − 푢], which is connected by the choice of the cut 퐶 and the edge 푒. The graph 퐺 [푋] is connected and the vertex 푢 is adjacent to vertices of 푋 (푣 is one such vertex). Thus, the second graph 퐺 ′′ [푋 + 푢] is connected as well. Every perfect matching of 퐺 that contains edge 푒 is also a perfect matching of 퐺 ′′ . Therefore, the graph 퐺 ′′ is matchable. If possible, let 푀 be a perfect matching of 퐺 ′′ which has an edge in the cut 퐷. Since 퐷 is an even cut of 퐺 ′′ , this perfect

89

5.2 Proof of the ELP Theorem 푅

푅 퐷

푋 − 푢

푋 퐶

푢

푢

푒

푒

푣

푋

푣

(a) 퐺

푋 + 푢 (b) 퐺 ′′

Fig. 5.5 The graphs 퐺 and 퐺 ′′ ; dashed lines indicate removed edges

matching 푀 would have to meet 퐷 in at least two edges, and we would have |푀 ∩ 퐶| > |푀 ∩ 퐷| ≥ 2. As 푀 is also a perfect matching of 퐺, and 퐶 is a tight cut of 퐺, this is impossible. We conclude that no edge of 퐷 is matchable in 퐺 ′′ . ƒ By Lemma 5.5, it now follows that 퐺 ′′ has a DM-barrier 퐵′′ such that 퐵′′ and the vertex sets of all the odd components of 퐺 ′′ − 퐵′′ are subsets of one of 푋 − 푢 and 푋 + 푢. The rest of the analysis depends on where the vertex 푢 is in relation to the barrier 퐵′′ . Case 2.1 Vertex 푢 lies in an even component of 퐺 ′′ − 퐵′′ . In this case, 퐵 := 퐵′′ + 푢 is a nontrivial barrier of 퐺 ′′ . But 퐺 ′′ − 퐵 = 퐺 − 퐵, and therefore 퐵 is a nontrivial barrier of 퐺. We conclude that 퐺 is not bicritical. By statement 5.6.1, 퐺 has a nontrivial barrier cut. Suppose now that 푢 is either in 퐵′′ or in the vertex set of one of the odd components of 퐺 ′′ − 퐵′′ . Since 푢 is in 푋 + 푢, it follows that 퐵′′ and the vertex sets of all the odd components of 퐺 ′′ − 퐵′′ are subsets of 푋 + 푢. As 퐺 ′′ [푋 − 푢] is connected, we conclude that 푋 − 푢 is a subset of the vertex set of an even component, say 퐿, of the graph 퐺 ′′ − 퐵′′ . Case 2.2 Vertex 푢 lies in 퐵′′ . In this case, 퐺 ′′ − 퐵′′ = 퐺 − 퐵′′ . Consequently, 퐿 is an even component of 퐺 − 퐵′′ . Let 푤 be any vertex of 퐿. Then, 퐵′′ + 푤 is a nontrivial barrier of 퐺 and we conclude that 퐺 is not bicritical. By statement 5.6.1, 퐺 has a nontrivial barrier cut. Now there is just one possibility that remains to be analysed, namely the case in which 푢 lies in some odd component of 퐺 ′′ − 퐵′′ . Case 2.3 Vertex 푢 lies in some odd component, 퐾, of 퐺 ′′ − 퐵′′ . In this case, the even component 퐿 of 퐺 ′′ − 퐵′′ contains the graph 퐺 [푋 − 푢]. See Figure 5.6.

90

5 Characterizations of Bricks and Braces 퐵′′ 푤

푅 ... 퐾

Odd components of 퐺 ′′ − 퐵′′

퐿

푢 푋 − 푢 ⊆ 푉 ( 퐿)

Fig. 5.6 Vertex 푢 is in an odd component 퐾 of 퐺 ′′ − 퐵′′

Clearly, no edge of 퐺 ′′ joins vertices in two different components of 퐺 ′′ − 퐵′′ . But the edges in 푅 = 퐸 (퐺) − 퐸 (퐺 ′′ ) may join vertices belonging to different components of 퐺 ′′ − 퐵′′ . However, each edge in 푅 joins 푢 to a vertex in 푋 − 푢. Thus, each odd component of 퐺 ′′ − 퐵′′ other than 퐾 is also an odd component of 퐺 − 퐵′′ . We may thus conclude that 표(퐺 − 퐵′′ ) ≥ |퐵′′ | − 1, and hence, by parity, that 퐵′′ is in fact a barrier of 퐺 itself. If 퐵′′ is nontrivial, then 퐺 is not bicritical. Assume that 퐵′′ has just one vertex, say 푤. By the hypothesis of the case under consideration, the vertex 푢 has at least two neighbours in 푋. Thus 푢 has a neighbour in 푋 that is different from 푤, and that neighbour has to be a vertex in 퐾, implying that 퐾 is nontrivial. In this case, 퐺 − 푢 − 푤 is not connected; hence either {푢, 푤} is a nontrivial barrier of 퐺 or {푢, 푤} is a 2-separation of 퐺, implying that 휕 (푉 (퐿) + 푢) is a 2-separation cut of 퐺. If 퐺 is not bicritical then, by statement 5.6.1, 퐺 has a nontrivial barrier cut. We conclude that 퐺 has a nontrivial ELP cut. In all cases considered, we deduced that the existence of a nontrivial tight cut in 퐺 implies that 퐺 has a nontrivial ELP cut. The proof of the ELP Theorem is complete. ƒ The ELP Theorem implies the following characterization of bricks, as discovered by Edmonds, Lov´asz and Pulleyblank [32]. Characterization of bricks Theorem 5.7 A nonbipartite matching covered graph is a brick if and only if it is 3-connected and bicritical. ƒ The above result suggests an obvious polynomial-time algorithm for recognizing bricks. (Exercise 5.2.8). The ELP Theorem simply says that every matching covered graph that has a nontrivial tight cut also has an ELP cut. Theorem 4.8 provides an explicit characterization of tight cuts in bipartite matching covered graphs. But the nature of tight cuts

91

5.2 Proof of the ELP Theorem

in nonbipartite matching covered graphs remains a mystery. In Section 5.5, we shall establish a precise description of tight cuts in matching covered graphs with at most two bricks and present an extension of that result, concerning tight cuts in arbitrary matching covered graphs.

Exercises ⊲5.2.1 Let 퐺 be a matching covered graph, let 휕 ( 푋) be a tight cut of 퐺, and let 푢푣 be an edge of 퐺, with 푢 ∈ 푋 and 푣 ∈ 푋 (as in Lemma 5.1). Show that 퐺 − 푢 − 푣 is a matchable graph in which no edge in the cut 휕 ( 푋 − 푢) is matchable. ⊲5.2.2 Show that the cut 퐶 = 휕 ( 푋) in the matching covered graph 퐺 shown in Figure 5.3 is a nonperipheral tight cut and, furthermore, that for any edge 푢푣 of the cut with 푢 ∈ 푋 and 푣 ∈ 푋, either 퐺 [푋 − 푢] or 퐺 [푋 − 푣] is disconnected. ⊲5.2.3 Prove Lemma 5.2. (Hint: prove that every perfect matching of H(퐵) is extendable to a perfect matching of 퐺.) 5.2.4 Consider the graph 퐺 shown in Figure 5.7 with the indicated cut 퐶. Use the method of the proof of Theorem 5.6, Case 1 to find a barrier in 퐺.

푢 푒

푋 퐶

푣

Fig. 5.7 휕(푢) ∩ 푋 = {푣 } and 휕(푣 ) ∩ 푋 = {푢}

⊲5.2.5 Use Theorem 5.6 to show that the following five families of matching covered graphs defined in Section 2.5 are bricks: (i) odd wheels (푊2푘+1 , for 푘 ≥ 1), (ii) prisms of order 2푝 (mod 4) (P2푛 , for odd 푛 ≥ 3), (iii) M¨obius ladders of order 0 (mod 4) (M2푛 , for even 푛 ≥ 2), (iv) staircases (S2푛 , 푛 ≥ 3), and (v) truncated biwheels (T2푛 , 푛 ≥ 3).

92

5 Characterizations of Bricks and Braces

The bricks belonging to the above five families are known as Norine-Thomas bricks and will play an important role in Part II. 5.2.6 Let 퐺 be a brick and let 퐺 + 푒 be a graph obtained from 퐺 by adding an edge 푒 joining two vertices of 퐺. Show that 퐺 + 푒 is also a brick. ∗ 5.2.7 Let 퐺 be a brick and let 퐶 := 휕 ( 푋) be a 3-cut of 퐺. Prove that if the subgraph 퐺 [푋] of 퐺 induced by 푋 is bipartite then 푋 is trivial. ∗ 5.2.8 Based on Theorem 5.7, describe a polynomial-time algorithm, which, given a nonbipartite matching covered graph 퐺, either attests that 퐺 is a brick or finds a nontrivial tight cut of 퐺.

5.3 Some Applications of the ELP Theorem In this section we note two interesting consequences of the ELP Theorem.

Tight cuts in cubic graphs Theorem 5.8 In a cubic matching covered graph every tight cut is a 3-edge-cut. Proof The smallest cubic matching covered graph is the 휃-graph (two vertices joined by three parallel edges), and the statement clearly holds for this graph. We shall prove the validity of the statement in general by induction on the number of vertices. Thus let 퐺 be any cubic matching covered graph on four or more vertices. If 퐺 is either a brick or a brace, then the only tight cuts of 퐺 are the trivial cuts which are obviously 3-edge-cuts. If not, by the ELP Theorem, 퐺 has either a nontrivial barrier cut or a 2-separation cut. However, by Exercise 4.2.5(i), every 2-separation cut in 퐺 is also a barrier cut. Also, by the second part of the same exercise, every barrier cut in 퐺 is a 3-edge-cut. Let 퐷 := 휕 (푌 ) denote any nontrivial barrier cut of 퐺. Let 퐶 = 휕 ( 푋) be any nontrivial tight cut of 퐺. Our objective is to show that 퐶 is a 3-edge cut. Since 퐷 is a 3-edge-cut, both 퐷-contractions of the cubic graph 퐺 are also cubic. If 퐶 does not cross 퐷, then it is a tight cut in one of the 퐷-contractions of 퐺 and hence, by induction, is a 3-edge of that 퐷-contraction which clearly is also a 3-edgecut of 퐺 itself. So, we may assume that 퐶 and 퐷 cross. As usual, fix notation so that 푋 ∩푌 is odd. Then, by Theorem 4.12, the two cuts 퐼 := 휕 ( 푋 ∩푌 ) and 푈 := 휕 ( 푋 ∩푌 ) are tight, and there are no edges joining vertices in 푋 ∩ 푌 to vertices in 푋 ∩ 푌 , implying the following equality: |퐶| + |퐷| = |퐼 | + |푈|. The cut 퐷, being a barrier cut of the cubic graph 퐺, is a 3-edge-cut. The two cuts 퐼 and 푈, being tight cuts of the 퐷-contractions of 퐺, are (by induction) 3-edge-cuts. The above equation now implies the stated result. ƒ

5.3 Some Applications of the ELP Theorem

93

Corollary 5.9 In a cubic matching covered graph, every tight cut is a barrier cut. Proof Let 퐶 be any tight cut in a cubic matching covered graph 퐺. By the above theorem, 퐶 is a 3-edge cut, and by Theorem 4.5, 퐶 is a barrier cut. ƒ Essentially 4-edge-connected graphs We recall this definition, given in Section 4.3. A 3-edge-connected graph is essentially 4-edge-connected if all its 3-edge cuts are trivial. Thus, a 3-edgeconnected graph 퐺 is essentially 4-edge-connected if |퐶| ≥ 4, for any nontrivial cut 퐶 of 퐺. For future use, we note the following consequences of Theorem 5.8. Corollary 5.10 Every essentially 4-edge-connected cubic graph is either a brick or a brace. ƒ Corollary 5.11 Every nonbipartite essentially 4-edge-connected cubic graph is a brick. ƒ It should be noted that the above-stated condition is not a necessary condition for a cubic graph to be a brick. This is because nonbipartite cubic graphs, such as 퐶6 , may have 3-edge cuts which are separating but not tight. However, any 3-edge cut in a cubic graph is a separating cut (Exercise 4.1.2), and every separating cut in a bipartite graph is tight (Corollary 4.9). These two facts, together with Theorem 5.8 now imply the following characterization of cubic braces of order six or more. Corollary 5.12 A bipartite cubic graph of order six or more is a brace if and only if it is essentially 4-edge-connected. ƒ Tight cuts in bicritical graphs ♯♯ Theorem 5.13 Any tight cut decomposition of a bicritical matching covered graph consists entirely of bricks. Proof Let 퐺 be a bicritical graph. If 퐺 is itself a brick, then there is nothing more to prove. We shall establish the assertion in general by induction on |푉 |.

We first note that since 퐺 is bicritical, it follows that it cannot have any nontrivial barriers and, consequently, it cannot have nontrivial barrier cuts. Thus, by the ELP Theorem, 퐺 must have a 2-separation cut. Let 퐶 be such a cut and let 퐺 1 and 퐺 2 denote the two 퐶-contractions of 퐺. Using the fact that 퐺 is bicritical and 퐶 is a 2-separation cut, it is easy to show that both 퐺 1 and 퐺 2 are also bicritical (Exercise 5.3.1). By the induction hypothesis, it follows that any tight cut decompositions of 퐺 1 and of 퐺 2 consists entirely of bricks, implying that any tight cut decomposition of 퐺 starting with 퐶 consists just of bricks. Now Lov´asz’s Unique Decomposition Theorem 4.17, implies that every tight cut decomposition of 퐺 consists only of bricks. ƒ

94

5 Characterizations of Bricks and Braces

We note that the analogue of the above statement for separating cut decompositions does not hold, in the sense that a separating cut decomposition of a bicritical graph may not consist exclusively of bicritical graphs (see Exercise 5.3.2). However, interestingly, the converse of the above statement holds not only for tight cut decompositions, but more generally for separating cut decompositions, in the sense that any graph which has a separating cut decomposition consisting exclusively of bicritical graphs is itself bicritical. This is a consequence of the following proposition: Proposition 5.14 Let 퐶 be a separating cut of a matching covered graph 퐺. If both 퐶-contractions of 퐺 are bicritical, then 퐺 is also bicritical. ƒ The above assertion is equivalent to the statement of Exercise 3.3.8 which says that any matching covered graph obtained by splicing two bicritical graphs is also bicritical. The following is an interesting and useful result concerning graphs obtained by splicing two bricks. Proposition 5.15 Let 퐶 be a separating cut of a matching covered graph 퐺. If both 퐶-contractions of 퐺 are bricks, then 퐺 is also a brick unless the cut 퐶 is a 2-separation cut of 퐺. Proof Let 퐺 1 := 퐺/( 푋 → 푥) and 퐺 2 := 퐺/( 푋 → 푥) be the two 퐶-contractions of 퐺. By the hypothesis, both 퐺 1 and 퐺 2 are bricks which, by the ELP Theorem, are bicritical and also 3-connected. Using the fact that both 퐺 1 and 퐺 2 are bicritical and Proposition 5.14 we deduce that 퐺 is bicritical. If 퐺 is also 3-connected, then it would follow from the ELP Theorem that 퐺 is a brick. Being matching covered, 퐺 is 2-connected and hence has no cut vertices. Thus, if 퐺 is not 3-connected, it must have a 2-vertex cut. Since 퐺 1 is 3-connected, it follows that for any 2-subset 푆 of 푉 (퐺 1 ) − 푥 the graph 퐺 − 푆 is connected. Similarly, if 푆 is any 2-subset of 푉 (퐺 2 ) − 푥, the graph 퐺 − 푆 is connected. Thus, if 퐺 does have a 2-vertex cut 푆, then 푆 has one vertex, say 푢 in 푉 (퐺 1 ) − 푥, and one vertex, say 푣 in 푉 (퐺 2 ) − 푥. As 퐺 1 is 3-connected, the graph 퐺 [푋 − 푢] = 퐺 1 − 푥 − 푢 is connected. Likewise, 퐺 [푋 − 푣] = 퐺 2 − 푥 − 푣 is connected. As 퐺 − 푢 − 푣 is not connected, we conclude that every edge of 퐶 is incident with a vertex in {푢, 푣}. In this case, 퐺 − 푢 − 푣 has two even components, one with the vertex set 푉 (퐺 1 ) − 푥 − 푢 and one with the vertex set 푉 (퐺 2 ) − 푥 − 푣, and 퐶 is a 2-separation cut of 퐺 associated with the 2-separation {푢, 푣}. ƒ

Exercises 5.3.1 Let 퐶 be a 2-separation of a bicritical graph 퐺, and let 퐺 1 and 퐺 2 be the two 퐶-contractions of 퐺. Show that both 퐺 1 and 퐺 2 are also bicritical. 5.3.2 Consider the graph 퐺 in Figure 5.8. Show that 퐺 is bicritical and that 퐶 is a separating cut in 퐺 such that one of the 퐶-contractions of 퐺 is not bicritical. ⊲5.3.3 Prove that a matching covered graph of order four or more is bicritical if and only if its tight cut decomposition consists solely of bricks.

5.3 Some Applications of the ELP Theorem

95

퐶 Fig. 5.8 The graph 퐺 is bicritical, but one of its 퐶-contractions is not

5.3.4 Let 퐺 be a cubic matching covered graph of order 푛. (i) Show that if 퐺 is not a brick and 푛 ≥ 4 then 퐺 is not bicritical. (ii) Using the previous part and Theorem 5.8, prove that 푏(퐺) ≤ 푛/4, (and show that this bound is best possible by giving for each 푘 ≥ 1, an example of a cubic graph on 4푘 vertices with exactly 푘 bricks). Edge-extensions Let 퐻 be a cubic graph, and let 푢 1 푣 1 and 푢 2 푣 2 be two disjoint edges of 퐻. The cubic graph obtained from 퐻 by (i) subdividing the edge 푢 1 푣 1 by inserting a vertex 푤 1 between 푢 1 and 푣 1 , (ii) subdividing the edge 푢 2 푣 2 by inserting a vertex 푤 2 between 푢 2 and 푣 2 , and (iii) adding a new edge 푤 1 푤 2 joining 푤 1 and 푤 2 , is said to be obtained from 퐻 by an edge-extension based on {푢 1 푣 1 , 푢 2 푣 2 }. (The graph in Figure 5.9 is a graph obtained from the cube by a sequence of two edge-extensions.) ⊲5.3.5 (i) Let 퐻 be an essentially 4-edge-connected cubic graph. Show that any graph 퐺 obtained from 퐻 by an edge-extension is also an essentially 4-edge-connected cubic graph. (ii) Wormald (1979, [94]) showed that any essentially 4-edge-connected cubic graph of order four or more can be obtained from either 퐾4 or from the cube by a sequence of edge-extensions. Show how 퐾3,3 , M8 , and the Petersen graph P can be obtained from 퐾4 by means of edge-extensions. ⊲5.3.6 Let 퐺 be a matching covered graph, let 퐶 := 휕 ( 푋) be an odd cut of 퐺, and let 퐺 [푋] denote the subgraph of 퐺 induced by 푋. Prove the following statements: (i) if 퐺 is bicritical then 퐺/푋 is bicritical if and only if 퐺 [푋] is critical, (ii) if 퐺 is 3-connected then 퐺/푋 is 3-connected if and only if 퐺 [푋] is 2-connected, (iii) if 퐺 is a brick then 퐺/푋 is a brick if and only if 퐺 [푋] is critical and 2-connected.

5 Characterizations of Bricks and Braces

96

Fig. 5.9 A graph obtained from the cube by two edge-extensions

5.4 Characterizations of Braces We now turn to the problem of characterizing braces. Any bipartite matching covered graph of order at most four is a brace. This is not the case with graphs of order six or more. In the ensuing discussion, we shall restrict ourselves to graphs with at least six vertices. We present two characterizations of braces; the first is a characterization in terms of a ‘Hall-type’ property, and the second implies a polynomial-time recognition algorithm for braces (and also shows that braces are the same as 2-extendable bipartite graphs). Both these characterizations rely on the fact that all tight cuts in bipartite matching covered graphs are barrier cuts, and conform to the pattern described in Theorem 4.8. For the convenience of the reader, a profile of a tight cut 퐶 := 휕 ( 푋) in a bipartite matching covered graph is displayed in Figure 5.10. 푋−

푋 +

푋+

퐶

푁 (푋− ) = 푋+ and |푋+ | = |푋− | + 1 Fig. 5.10 Profile of a tight cut in a bipartite matching covered graph

The next result is similar to Theorem 2.9.

푋 −

5.4 Characterizations of Braces

97

A Hall-type characterization of braces Theorem 5.16 A bipartite graph 퐺 [ 퐴, 퐵], with | 퐴| = |퐵| ≥ 3, is a brace if and only if |푁 ( 푋)| ≥ | 푋 | + 2, for all 푋 ⊂ 퐴 such that 0 < | 푋 | < | 퐴| − 1.

(5.1)

Proof Let us first suppose that 퐺 is a brace. Then 퐺 is, in particular, a matching covered graph. By Theorem 2.9, |푁 (푆)| ≥ |푆| + 1 for each nonempty proper subset 푆 of 퐴. Suppose that for some subset 푆 of 퐴, with 0 < |푆| < | 퐴| − 1, the equality |푁 (푆)| = |푆| + 1 holds. Then it is easy to see that 휕 ( 푋), where 푋 := 푆 ∪ 푁 (푆), is a nontrivial tight cut of 퐺 (Exercise 5.4.1). We conclude that the property (5.1) must hold. Conversely, suppose that the property (5.1) holds. Then it is straightforward to verify (Exercise 5.4.1) that 퐺 is matching covered. If 퐺 were not a brace, it would have a nontrivial tight cut 휕 ( 푋). By interchanging 푋 and 푋, if necessary, we may assume that 푋− is a subset of 퐴. By Theorem 4.8, the set 푋− violates (5.1). We conclude that 퐺 is a brace. ƒ The next result is similar to Theorem 3.8. A second characterization of braces Theorem 5.17 A bipartite graph 퐺 [ 퐴, 퐵], with | 퐴| = |퐵| ≥ 3, is a brace if and only if 퐺 − 푢 1 − 푢 2 − 푣 1 − 푣 2 is matchable, (5.2) for any pair {푢 1 , 푢 2 } of distinct vertices in 퐴 and any pair {푣 1 , 푣 2 } of distinct vertices in 퐵. Proof Let us first suppose that 퐺 is a brace. Then, by Theorem 5.16, the property (5.1) must hold. Now let 푢 1 and 푢 2 be any two distinct vertices of 퐴, and let 푣 1 and 푣 2 be any two distinct vertices of 퐵. Then 퐻 := 퐺 − 푢 1 − 푢 2 − 푣 1 − 푣 2 is a bipartite graph with bipartition ( 퐴 − 푢 1 − 푢 2 , 퐵 − 푣 1 − 푣 2 ). For any subset 푆 of 퐴 − 푢 1 − 푢 2 , clearly, 푁 퐻 (푆) = 푁퐺 (푆) − 푣 1 − 푣 2 . Thus, it follows from inequality (5.1) that |푁 퐻 (푆)| ≥ |푆| for any subset 푆 of 퐴−푢 1 −푢 2 and, hence, by Hall’s Theorem 1.5, that 퐻 is matchable. Conversely, suppose that the property (5.2) holds. Let 푋 be a proper subset of 퐴 such that 0 < | 푋 | < | 퐴| − 1. Then | 퐴 − 푋 | ≥ 2; let 푢 1 and 푢 2 be two distinct vertices of 퐴 − 푋. Let 푣 1 and 푣 2 be two distinct vertices of 퐵 such that 퐼 := {푣 1 , 푣 2 } ∩ 푁퐺 ( 푋) is as large as possible. Let 퐻 := 퐺 − 푢 1 − 푢 2 − 푣 1 − 푣 2 . By hypothesis, 퐻 is matchable. By Hall’s Theorem, |푁퐺 ( 푋)| = |푁 퐻 ( 푋)| + |퐼 | ≥ | 푋 | + |퐼 |.

(5.3)

98

5 Characterizations of Bricks and Braces

We now show that |퐼 | = 2. By definition, 푋 is nonempty. From (5.3) we infer that |푁퐺 ( 푋)| ≥ |퐼 | + 1.

(5.4)

From inequality (5.4) we deduce that 푁퐺 ( 푋) is nonempty. By the maximality of 퐼, it follows that |퐼 | ≥ 1. Again, from inequality (5.4) we infer that |푁퐺 ( 푋)| ≥ 2. By the maximality of 퐼, we conclude that |퐼 | = 2. From (5.3) we infer that (5.1) holds for 푋. This conclusion holds for each proper ƒ subset 푋 of 퐴 such that 0 < | 푋 | < | 퐴| − 1. By Theorem 5.16, 퐺 is a brace. Braces are 2-extendable Recall that a graph 퐺 is said to be 2-extendable if, given any two disjoint edges 푒 and 푓 of 퐺, there exists a perfect matching of 퐺 which contains {푒, 푓 }. We leave the proof of the following theorem as an exercise. Theorem 5.18 A connected bipartite graph of order six or more is a brace if ƒ and only if it is 2-extendable. The above theorem suggests an obvious polynomial-time algorithm for determining whether or not a given bipartite graph is a brace. Let 퐺 [ 퐴, 퐵] be a bipartite graph on 푛 vertices and 푚 edges, and assume that a perfect matching 푀 has been determined. It can be checked in linear time 푂 (푚) if 퐺 is matching covered in the following way: let 퐷 be a directed graph obtained from 퐺 by directing every edge of 퐺 − 푀 from 퐴 to 퐵 and directing every edge in 푀 from 퐵 to 퐴. Then 퐺 is matching covered if and only if 퐷 is strongly connected (see Exercise 2.2.6). The strongly connected property is well known and an 푂 (푚)-time procedure is described, for instance, in Section 22.5 of [25]. Now, to check if a bipartite graph 퐺 is a brace, it suffices to remove the ends of every edge of 퐺 and decide if the resulting graph is matching covered. This procedure can be computed in polynomial-time. A more mindful reader may consider a faster algorithm (Exercise 5.4.4). The following three factors imply the existence of a polynomial-time algorithm to determine a tight cut decomposition of a matching covered graph (Exercise 5.4.5): (i) the polynomial-time algorithm for recognition of braces (Exercise 5.4.4), (ii) the polynomial-time algorithm for recognition of bricks (Exercise 5.2.8), and (iii) the fact that the number of bricks and braces in a tight cut decomposition of a matching covered graph of order 푛 is bounded above by 푛/2−1 (Exercise 4.3.5).

Exercises ⊲5.4.1 Supply the missing details in the proof of Theorem 5.16. 5.4.2 Let 퐻 [ 퐴, 퐵] be a brace on four or more vertices. Let 푎 1 , 푎 2 be any two vertices in 퐴, and 푏 1 , 푏 2 be any two vertices in 퐵. Show that the graph 퐺 obtained from 퐻 by adding the two edges 푎 1 푎 2 and 푏 1 푏 2 is a brick.

5.5 Tight Cuts in Graphs with Two Bricks

99

∗ 5.4.3 Prove Theorem 5.18. ∗ 5.4.4 Let 퐺 be a connected bipartite graph of order six or more and let 푀 be a perfect matching of 퐺. Show that 퐺 is a brace if and only if 퐺 − 푢 − 푣 is matching covered, for every edge 푢푣 of 푀. Based on this property, design a polynomial-time algorithm, which, given 퐺 and a perfect matching of 퐺, either attests that 퐺 is a brace, or determines a nontrivial tight cut of 퐺. Hint: Exercise 2.2.6. ∗ 5.4.5 Describe a polynomial-time algorithm to determine a tight cut decomposition of a given matching covered graph. ∗ 5.4.6 Let 퐺 [ 퐴, 퐵] be any brace of order at least six. Show that, for any two vertices 푢 and 푣 of 퐺, the graph 퐺 − 푢 − 푣 is connected (and thus, that 퐺 is 3-connected). Furthermore show that if 푆 is any subset of 푉 of cardinality three which is neither a subset of 퐴, nor a subset of 퐵, then 퐺 − 푆 is connected.

5.5 Tight Cuts in Graphs with Two Bricks ♯♯ Barrier cuts and essentially 2-separation cuts Every tight cut in a matching covered graph with at most one brick is an ELP cut. More precisely, if 퐶 is a tight cut in a matching covered graph 퐺 with 푏(퐺) ≤ 1 then one of the two shores of 퐶 is bipartite, and it follows from Theorem 4.7 that 퐶 is a special barrier cut. But, as we have seen, a matching covered graph with more than two bricks may have tight cuts which are not ELP cuts. In their attempts to confirm a conjecture of Lov´asz (Chapter 15), CLM found it necessary to have a precise description of all tight cuts in matching covered graphs with two bricks. In this special case, they found that every tight cut is either an ELP cut or is close to being one in the sense to be described below1. In Chapter 4, we defined a barrier 퐵 of a matching covered graph 퐺 to be special if 퐺 − 퐵 has precisely one nontrivial component. Thus, if 퐺 is nonbipartite, and 퐵 is a special barrier of 퐺, then 퐺 − 퐵 has one nonbipartite component, say 퐾, while the remaining |퐵| − 1 components are trivial. We also referred to 휕 (퐾) as the special barrier cut associated with 퐵. Essentially 2-separation cuts Let 퐺 be a nonbipartite matching covered graph. An essential 2-separation of 퐺 is a pair {푆, 푇 } of two disjoint odd subsets 푆 and 푇 of 푉 (퐺) such that: • both 퐺 [푆] and 퐺 [푇] are bipartite subgraphs of 퐺 such that their majority parts 푆+ and 푇+ are special barriers of 퐺, 1 We recommend to the first-time readers who wish to gain an overview of the basic results of the subject to skip this section and return to it at a later stage.

100

5 Characterizations of Bricks and Braces

• the cut 휕 (푆) is the special barrier cut associated with the barrier 푆+ and the cut 휕 (푇) is the special barrier cut associated with the barrier 푇+ , and • the pair {푠, 푡} is a 2-separation of the graph 퐺/(푆 → 푠)/(푇 → 푡). Clearly, if {푆, 푇 } is an essential 2-separation of 퐺, than any 2-separation cut of 퐺/(푆 → 푠)/(푇 → 푡) associated with the 2-separation {푠, 푡} corresponds naturally to a tight cut of 퐺 itself, and we refer to it as an essentially 2-separation cut of 퐺. See Figure 5.11 for an illustration. 퐷 3

4

2

5 1 퐶 6 10

7 9

8

Fig. 5.11 Essentially 2-separation cuts in a graph

In Figure 5.11, the pair {푆, 푇 } of subsets of 푉, where 푆 := {1, 2, 3} and 푇 = {6, 7, 8}, constitutes an essential 2-separation of 퐺 and the indicated tight cuts 퐶 and 퐷 of 퐺 are the two essentially 2-separation tight cuts of 퐺 associated with {푆, 푇 }. A second example of an essential 2-separation of 퐺 is the pair {푆 ′ , 푇 }, where 푆 ′ = {1} and 푇 = {6, 7, 8}. Note that one of the two essentially 2-separation cuts associated with {푆 ′ , 푇 } is the same cut 퐶 while the other cut, 퐷 ′ := 휕 ({1, 9, 10}), is different from 퐷. The above-given definition may be phrased more briefly as follows: a tight cut 퐶 := 휕 ( 푋) of a matching covered graph 퐺 is an essentially 2-separation cut if there is a companion tight cut 퐷 := 휕 (푌 ) which crosses 퐶 such that 푋 ∩ 푌 is odd and both 퐺 [푋 ∩ 푌 ] and 퐺 [푋 ∩ 푌 ] are bipartite subgraphs of 퐺. In the above example, 퐶 is an essentially 2-separation cut with either 퐷 or 퐷 ′ as its companion. Types of nontrivial tight cuts in graphs with two bricks Theorem 5.19 In a matching covered graph with just two bricks, every nontrivial tight cut is either a barrier cut or is an essentially 2-separation cut.

101

5.5 Tight Cuts in Graphs with Two Bricks

Proof Let 퐺 be a matching covered graph such that 푏(퐺) = 2, and let 퐶 := 휕 ( 푋) be a nontrivial tight cut of 퐺. All matching covered graphs on fewer than six vertices are either bricks or braces. Thus, a graph with two bricks has to have at least six vertices. It can be easily checked that if 푛 = 6 and 푏 = 2, then 퐶 is a 2-separation cut (Exercise 5.5.1). We prove, by induction on |푉 |, that 퐶 is either a barrier cut or is an essentially 2-separation cut. To begin with, we may assume that both 퐶-contractions of 퐺 are nonbipartite; otherwise, by Theorem 4.7, the cut 퐶 would be a barrier cut. Thus, since 푏(퐺) = 2, we may assume that both 퐶-contractions of 퐺 are near-bricks. Lemma 5.20 If the graph 퐺 is bicritical, then 퐶 is a 2-separation cut. Proof Suppose that 퐺 is bicritical. Theorem 5.13 and the fact that 푏(퐺) = 2 together imply that both 퐶-contractions of 퐺 are bricks. As 퐺 itself is not a brick, it follows from Proposition 5.15 that 퐶 is a 2-separation cut and the assertion holds. ƒ When 퐺 is not bicritical, it has nontrivial barriers. The proof of the theorem then turns out to be relatively simple if the graph 퐺 happens to have a barrier with a special relationship with the cut 퐶. 퐶-avoiding barrier We define a nontrivial barrier 퐵 of 퐺 to be 퐶-avoiding if 퐶 does not cross any of the barrier cuts associated with 퐵. Figure 5.12 shows two examples of such barriers. 퐶 퐵

퐵

(푎)

(푏)

퐶

Fig. 5.12 퐶-avoiding barriers

We note that if a nontrivial barrier 퐵 is 퐶-avoiding, then either 퐶 is a tight cut of the bipartite graph 퐻 := H(퐵) obtained by shrinking each odd component of 퐺 − 퐵 to a single vertex (as in the example shown in Figure 5.12(a)) or there is a nontrivial odd component 퐾 of 퐺 − 퐵 such that 퐶 is a tight cut of the graph 퐺/(푉 (퐾)) (as in the example shown in Figure 5.12(b)). See Exercise 5.5.2. Case: there is a 퐶-avoiding barrier 퐵 Lemma 5.21 If the graph 퐺 has a 퐶-avoiding barrier 퐵, then 퐶 is either a barrier cut or is an essentially 2-separation cut of 퐺.

102

5 Characterizations of Bricks and Braces

Proof We consider the two alternatives mentioned above. Firstly, suppose that 퐶 is a cut of the bipartite graph 퐻, and let 푋 be the shore of 퐶 in 퐻 such that 푋+ ⊂ 퐵. Then, 푋+ is a barrier of 퐺, and 퐶 is a barrier cut of 퐺 associated with 푋+ . If 퐶 is not a cut of the bipartite graph 퐻, the fact that 퐵 is 퐶-avoiding implies that there an odd component, say 퐾, of 퐺 − 퐵, such that 퐶 is a nontrivial tight cut of the graph 퐽 := 퐺/(푉 (퐾) → 푘). 5.21.1 If 퐶 is a barrier cut of 퐽, then it is also a barrier cut of 퐺. Proof Suppose that the cut 퐶 is a barrier cut associated with a barrier 퐵 퐽 of 퐽. If the contraction vertex 푘 lies in 퐵 퐽 then 퐵 ∪ (퐵 퐽 − 푘) is a barrier of 퐺 and the cut 퐶 is associated with that barrier. Alternatively, if the contraction vertex 푘 of 퐽 does not lie in 퐵 퐽 then 퐵 퐽 is a barrier of 퐺 and cut 퐶 is associated with 퐵 퐽 in 퐺. In both alternatives, 퐶 is a barrier cut of 퐺. ƒ In view of the above statement, we may assume that 퐶 is not a barrier cut of 퐽. As any tight cut of a graph with at most one brick is a barrier cut, it follows that 푏(퐽) ≥ 2. However, as 푏(퐺) = 2, it follows that 푏(퐽) = 2 and hence that 퐺/푉 (퐾) is bipartite. As 퐽 clearly has fewer vertices than 퐺, the induction hypothesis implies that 퐶 is either a barrier cut of 퐽 or is an essentially 2-separation cut of 퐽. In the former case, by statement 5.21.1, the cut 퐶 is a barrier cut of 퐺. The following statement deals with the latter case. 5.21.2 If 퐶 is an essentially 2-separation cut of 퐽 then it is also an essentially 2-separation cut of 퐺. Proof Let {푆, 푇 } be an essential 2-separation of 퐽 such that 퐶 is a 2-separation cut of 퐽 associated with it. If the contraction vertex 푘 does not lie in either 푆 or 푇, then {푆, 푇 } is an essential 2-separation of 퐺 itself, and 퐶 is an essentially 2-separation cut of 퐺. So we may assume that 푘 is in one of the sets 푆 and 푇. Adjust notation so that 푘 ∈ 푆. In this case, the pair {푆 ′ , 푇 }, where 푆 ′ := (푆 − 푘) ∪ 푉 (퐾), is an essential 2-separation of 퐺 and 퐶 is also an essentially 2-separation cut of 퐺 associated with {푆 ′ , 푇 }. ƒ This completes the proof of the assertion of the lemma.

ƒ

How does one go about finding a 퐶-avoiding nontrivial barrier of 퐺? The proof of the following lemma shows that given any maximal nontrivial barrier 퐵∗ of 퐺 one can either find a 퐶-avoiding barrier or deduce that 퐶 is a cut of the required type. Case: a maximal barrier 퐵∗ is not 퐶-avoiding Lemma 5.22 If the graph 퐺 has a maximal nontrivial barrier 퐵∗ which is not 퐶-avoiding, then: • either 퐺 has another nontrivial barrier 퐵 which is 퐶-avoiding, • or 퐶 is a 2-separation cut of 퐺.

103

5.5 Tight Cuts in Graphs with Two Bricks

Proof Since 퐵∗ is not 퐶-avoiding, there exists a barrier cut 퐷 := 휕 (푌 ) associated with 퐵∗ such that 퐶 and 퐷 cross, where 푌 is the vertex set of an odd component of 퐺 − 퐵∗ (and thus, the barrier 퐵∗ is a subset of 푌.) Adjust notation so that 푋 ∩ 푌 and 푋 ∩푌 are odd and, as usual, let 퐼 := 휕 ( 푋 ∩푌 ) and 푈 := 휕 ( 푋 ∩푌). By Theorem 4.12, both 퐼 and 푈 are tight cuts and there are no edges of 퐺 between 푋 ∩ 푌 and 푋 ∩ 푌 . We shall use the following notation to denote the various graphs obtained from 퐺 by tight cut contractions (see Figure 5.13): 퐺 1 := 퐺/( 푋 → 푥), 퐻1 := 퐺/(푌 → 푦),

퐺 11 := 퐺/( 푋 ∪ 푌 ),

퐺 12 := 퐺 1 /( 푋 ∩ 푌 ),

퐺 2 := 퐺/( 푋 → 푥), 퐻2 := 퐺/(푌 → 푦),

퐺 22 := 퐺/( 푋 ∪ 푌 ),

퐺 21 := 퐺 2 /( 푋 ∩ 푌 ).

(퐺 12 is the same as 퐻2 /( 푋 ∩ 푌 ) up to multiple edges.) (퐺 21 is the same as 퐻1 /( 푋 ∩ 푌 ) up to multiple edges.) If either of the two shores of 퐶 is bipartite, then the majority part of that shore would be a 퐶-avoiding barrier of 퐺 and there would be nothing more to prove. So, we may assume that both shores of 퐶 are nonbipartite. Since 푏(퐺) = 2, this implies that both 퐺 1 and 퐺 2 are near-bricks. Keeping this in mind, we now proceed to draw conclusions about the various graphs displayed in Figure 5.13. 5.22.1 The vertex 푦 does not lie in any nontrivial barrier of 퐻1 . Proof Let 퐵′ denote a barrier of 퐻1 that contains vertex 푦. Then the set 퐵′′ := (퐵′ − 푦) ∪ 퐵∗ is a barrier of 퐺. By the maximality of 퐵∗ , it follows that 퐵′ = {푦}. Indeed, the only barrier of 퐻1 that contains the vertex 푦 is trivial. ƒ 5.22.2 The graph 퐺 21 = 퐺/(푌 → 푦)/( 푋 ∩ 푌 → 푠) is not bipartite. Proof Assume the contrary. Let ( 퐴′ , 퐵′ ) denote the bipartition of 퐺 21 and adjust notation so that 푠 lies in 퐴′ . By Proposition 4.1, the subgraph 퐺 [푋] induced by 푋 is connected. This implies that 푦 and 푠 are adjacent in 퐺 21 implying that 푦 lies in 퐵′ . Then, 퐵′ is a nontrivial barrier of 퐻1 that contains vertex 푦, which contradicts statement 5.22.1. ƒ 5.22.3 The graph 퐺 22 is bipartite. Moreover, | 푋 ∩ 푌 | = 1. Proof The graph 퐺 2 , a 퐶-contraction of 퐺, is a near-brick. Moreover, the cut 푈 := 휕 ( 푋 ∩ 푌 ) is tight in 퐺 2 , and the 푈-contraction 퐺 21 of 퐺 2 is not bipartite. Therefore, 퐺 22 , the other 푈-contraction of 퐺 2 , is bipartite implying that the graph 퐺 [푋 ∩ 푌 ] is bipartite.

If | 푋 ∩ 푌 | ≥ 3, the majority part of 푋 ∩ 푌 would be a nontrivial special barrier of 퐺 which is 퐶-avoiding and there would be nothing more to prove. Therefore, we may assume that | 푋 ∩ 푌 | = 1. ƒ

104

5 Characterizations of Bricks and Braces

푋 ∩ 푌

푋 ∩ 푌

푋

퐼 푥 퐺1

퐺11

퐺12

푌

푋 푦

푦 푌

푌

퐶

퐻1

퐻2

퐷 퐺

푠 푥 푈

푋 ∩ 푌

푦

퐺21

푋

푋 ∩ 푌

퐺2

퐺22

Fig. 5.13 The graph 퐺 and its cut contractions

We shall let 푡 denote the unique vertex in 푋 ∩ 푌 . Now we consider two cases depending on whether or not the graph 퐺 11 is bipartite. Case 1 The graph 퐺 11 is bipartite. In this case, 퐺 [푋 ∩ 푌 ] is the bipartite shore of the cut 퐼 = 휕 ( 푋 ∩ 푌 ) in 퐺. If | 푋 ∩ 푌 | ≥ 3, then the majority part of 푋 ∩ 푌 would be a nontrivial special barrier of 퐺 which is 퐶-avoiding. On the other hand, if | 푋 ∩ 푌 | = 1, say 푋 ∩ 푌 = {푠}, then {푠, 푡} would be a 2-separation of 퐺, and 퐶 would be a 2-separation cut of 퐺. Both possibilities lead to a confirmation of the assertion of the lemma. Case 2 The graph 퐺 11 is not bipartite. In this case, neither of the two 퐼-contractions of 퐻1 is bipartite. As 퐻1 is a 퐷contraction of 퐺 which has two bricks, it follows that 퐺 11 and 퐺 21 are near-bricks and, furthermore, that 푏(퐻1 ) = 2 and that 퐻2 is bipartite. Since 푏(퐻1 ) = 2 and 퐼 is a tight cut of 퐻1 , by the induction hypothesis, 퐼 is either a barrier cut of 퐻1 , or is an essentially 2-separation cut of 퐻1 .

5.5 Tight Cuts in Graphs with Two Bricks

105

Case 2.1 퐼 is a barrier cut of 퐻1 . Let 퐵1 denote a barrier of 퐻1 such that the cut 퐼 = 휕 ( 푋 ∩ 푌 ) is associated with 퐵1 . Since 퐺 11 and 퐺 21 are the two 퐼-contractions of 퐻1 , it follows that either (i) 퐵1 is a subset of ( 푋 ∩ 푌 ) + 푦 and 푋 ∩ 푌 is the vertex set of an odd component of 퐻1 − 퐵1 , or (ii) 퐵1 is a subset of 푋 ∩ 푌 and 푉 (퐺 21 ) − 푠 is the vertex set of an odd component of 퐻1 − 퐵1 . If 퐵1 were a subset of ( 푋 ∩ 푌 ) + 푦 and 푋 ∩ 푌 the vertex set of an odd component of 퐻1 − 퐵1 , the contraction vertex 푦 would have to belong to 퐵1 because there are edges in 퐺 between 푋 ∩ 푌 and 푌. Likewise, some vertex of 푋 ∩ 푌 would also have to belong to 퐵1 , because there are edges that join vertices in 푋 ∩ 푌 to vertices in 푋 ∩ 푌 . But, in this case, 퐵1 would be a nontrivial barrier of 퐻1 that contains vertex 푦, in contradiction to statement 5.22.1. So, this situation cannot arise. Thus, we may suppose that 퐵1 is a subset of 푋 ∩ 푌 and 푉 (퐺 21 ) − 푠 is the vertex set of an odd component of 퐻1 − 퐵1 . In this case, 퐵1 is a barrier of 퐺 such that the shore 푋 of 퐶 is a subset of the vertex set of one of the odd components of 퐺 − 퐵1 , implying that 퐵1 is a 퐶-avoiding barrier of 퐺 and hence the assertion of the lemma. Case 2.2 퐼 is an essentially 2-separation cut of 퐻1 . Let {푋1 , 푋2 } denote an essential 2-separation of 퐻1 such that 퐼 is a 2-separation cut of 퐻1 /( 푋1 → 푥1 )/( 푋2 → 푥2 ) associated with the pair {푥1 , 푥2 }. For 푖 = 1, 2, let ( 퐴푖 , 퐵푖 ) denote the bipartition of 퐻1 [푋푖 ], where 퐵푖 is the majority part. One of 푋1 and 푋2 is a subset of 푋 ∩ 푌 , the other a subset of ( 푋 ∩ 푌 ) + 푦. Adjust notation so that 푋1 is a subset of 푋 ∩ 푌 . If 푋1 is not a singleton, the majority part 퐵1 of 푋1 , being a subset of 푋 ∩ 푌 , does not contain the contraction vertex 푦. Hence 퐵1 would be a 퐶-avoiding nontrivial barrier of 퐺 itself, and the first alternative in the assertion of the lemma follows. So, we may assume that | 푋1 | = 1. Let 푥1 be the vertex of 푋1 . 5.22.4 Every edge of 휕 ( 푋 ∩ 푌 ) is incident in 퐺 with a vertex in {푥1 , 푡}. Proof Let 푒 be an edge of 휕 ( 푋 ∩ 푌 ). The cuts 퐶 and 퐷 are tight in 퐺; therefore, no edge of 퐺 joins a vertex of 푋 ∩ 푌 to a vertex of 푋 ∩ 푌 . As 푡 is the only vertex of 푋 ∩ 푌 , we infer that 푒 is incident in 퐺 with a vertex in ( 푋 ∩ 푌 ) + 푡. Suppose that 푒 is not incident in 퐺 with vertex 푡. In that case, 푒 is incident with a vertex of 푋 ∩ 푌 . Thus, 푒 is in 퐼 and has both ends in 푋. Being an edge of 퐼, the edge 푒 is incident in 퐻1 with a vertex in 푋2 + 푥1 . The set 푋2 is disjoint with 푋. The edge 푒 has both ends in 푋. Thus, 푒 is incident with 푥1 . This conclusion holds for each edge 푒 of 휕 ( 푋 ∩ 푌 ) which is not incident with 푡. We deduce that every edge of 휕 ( 푋 ∩ 푌 ) is incident with a vertex in {푥1 , 푡}, as asserted. ƒ Graph 퐻2 is bipartite; therefore 퐺 [푌 ] is also bipartite. The barrier 퐵∗ is one of the parts in the bipartition of 퐺 [푌 ]. Let 퐴∗ denote the other part. The vertex 푡 is the only vertex of 푋 ∩ 푌, and it lies in 퐵∗ . See Figure 5.14. It follows that 퐴∗ ∪ (퐵∗ − 푡) = 푋 ∩ 푌. By statement (5.22.4), every edge of 휕 ( 퐴∗ ∪ (퐵∗ − 푡)) is incident with a vertex in {푥1 , 푡}. The vertices of 퐵∗ − 푡 are not adjacent to vertex 푡. Thus, every edge of 휕 (퐵∗ − 푡) is incident with 퐴∗ + 푥1 . It

106

5 Characterizations of Bricks and Braces 푌 퐵∗ − 푡

푥1

퐴∗

푋 퐶

푡

퐷 ∗



Fig. 5.14 The bipartition ( 퐴 , 퐵 ) of 퐺 [푌 ]

follows that the | 퐴∗ | vertices of 퐵∗ − 푡 are isolated in 퐺 − 퐴∗ − 푥1 . We conclude that 퐵 := 퐴∗ + 푥1 is a special barrier of 퐺, and the only nontrivial component of 퐺 − 퐵 contains all the vertices of 푋. In sum, 퐵 is a nontrivial 퐶-avoiding barrier of 퐺. ƒ It is now straightforward to deduce the assertion of the theorem from Lemmas 5.20, 5.21, and 5.22. ƒ We leave the proof of the following corollary as Exercise 5.5.3. Corollary 5.23 Let 퐺 be a matching covered graph such that 푏(퐺) = 2. Let 퐶 be a tight cut of 퐺. Then, one of the following alternatives holds: (i) one of the shores of 퐶 is bipartite, (ii) the cut 퐶 is a member of a pair of nontrivial barrier cuts, (iii) the cut 퐶 is a member of an essentially 2-separation cut pair.

ƒ

5.5.1 The laminar ELP Theorem As mentioned before, the need for a precise characterization of all tight cuts in matching covered graphs with two bricks arose in connection with the work of CLM on a conjecture of Lov´asz which will be described in Chapter 15. This naturally led them to the question of whether a similar description of tight cuts in arbitrary matching covered graphs was possible. For many years they were not able to find an answer to this question, and were led to conjecture that, if 퐶 is any tight cut in a matching covered graph 퐺, then there exists an ELP cut of 퐺 which is laminar to 퐶 (CLM (2018, [19])). This conjecture was confirmed in 2021 by Guantao Chen, Xing Feng, Fuliang Lu, Cl´audio L. Lucchesi and Lianzhu Zhang (2021, [20]). They were able to establish this assertion by showing that a statement which is in the same spirit as Theorem 5.19 is true for tight cuts in all matching covered graphs. We briefly sketch below the main ingredients in their proof.

107

5.5 Tight Cuts in Graphs with Two Bricks

퐶-sheltered and 퐶-avoiding sets Let 퐺 be a matching covered graph, let 퐶 be a tight cut of 퐺 and let 푆 be a set of vertices of 퐺 which is either a barrier or a 2-separation. The set 푆 is 퐶-sheltered if 푆 is a subset of a shore of 퐶, and is 퐶-avoiding if each ELP cut associated with 푆 is laminar with 퐶. If 푆 is 퐶-sheltered then, as each shore of 퐶 induces a connected subgraph of 퐺, one of the components of 퐺 − 푆 contains all the vertices of the shore of 퐶 disjoint with 푆. It follows that if 푆 is 퐶-sheltered then 푆 is 퐶-avoiding. We record this result for later reference. Proposition 5.24 Let 푆 be either a 2-separation or a barrier of a matching covered graph 퐺 and let 퐶 be a tight cut of 퐺. If 푆 is 퐶-sheltered then some cut associated with 푆 has a shore that is a superset of a shore of 퐶, say, 푋, and all the other cuts associated with 푆 have a shore that is a subset of 푋. Consequently, if 푆 is 퐶-sheltered then it is 퐶-avoiding. ƒ Theorem 5.25 (The Laminar ELP Theorem) Let 퐶 be a nontrivial tight cut of a matching covered graph 퐺. Then either 퐺 has a 퐶-sheltered nontrivial barrier or a 2-separation cut which is laminar with 퐶 (see Example 5.26). ƒ Example 5.26 Consider the graph depicted in Figure 5.15. The tight cut 퐶 is laminar with the cut 퐷, which is a 2-separation cut associated with the pair {푢 1 , 푢 2 }. The cut 퐶 is also laminar with the 2-separation cut 퐹, which is associated with the pair {1, 푏 3 }. The barriers {푏 1 , 푏 2 , 푏 3 } and {2, 3} are 퐶-sheltered, whereas the barrier {1, 2, 3} is not 퐶-avoiding. 푢2

1 퐶

푏1

푏2

2

푏3

3

퐹 퐷

Fig. 5.15 The graph of Example 5.26

푢1

108

5 Characterizations of Bricks and Braces

5.5.2 ELP tight cut decompositions Given any matching covered graph 퐺, we may apply to it a procedure, called an ELP tight cut decomposition of 퐺, which produces a list of bricks and braces. If 퐺 itself is a brick or a brace then the list consists of just 퐺. Otherwise, let 퐶 be any nontrivial ELP cut of 퐺. Then, both 퐶-contractions of 퐺 are matching covered. One may recursively apply the ELP tight cut decomposition procedure to each 퐶-contraction of 퐺, and then combine the resulting lists to produce an ELP tight cut decomposition of 퐺 itself. An interesting consequence of Theorem 5.25 is that any nontrivial tight cut 퐶 of a matching covered graph 퐺 participates in some ELP tight cut decomposition of 퐺. In this sense, every tight cut in a matching covered graph is ultimately an ELP cut. It is our hope that Theorem 5.25 which provides a characterization of all tight cuts could be used as a tool for establishing interesting results concerning the structure of matching covered graphs!

Exercises 5.5.1 Show that any nontrivial tight cut in a matching covered graph 퐺 with 푛 = 6 and 푏 = 2 is a 2-separation cut. 5.5.2 Show that if 퐵 is a 퐶-avoiding barrier, then either 퐶 is a tight cut of the core 퐻 of 퐺 with respect to 퐵, or that it is a nontrivial tight cut of 퐺/푉 (퐾) for some odd component 퐾 of 퐺 − 퐵. ⊲5.5.3 Prove Corollary 5.23.

5.6 Notes Recall that a matching covered graph 퐺 is 푘-extendable if any matching of size 푘 in 퐺 can be extended to a perfect matching of 퐺. Every matching covered graph is 1extendable and, by Theorem 5.18, a bipartite matching covered graph is 2-extendable if and only if it is a brace. The graph in Figure 5.16 is 2-extendable, but is not 3-extendable. For 푘 ≥ 3, the graph obtained from 퐾2푘 by deleting the edges of a perfect matching is known as a cocktail party graph as it can be viewed as representing the pattern of intermingling among 푛 couples at a cocktail party where each guest socializes with all people except with his or her own spouse. It is fairly easy to show that if a matching covered graph is 2-extendable, then it has to be either a brick or a brace. The complete graph 퐾4 is a 2-extendable brick. Prisms of order 푛, where 푛 ≡ 2 (mod 4), M¨obius ladders of order 푛 ≥ 8, where 푛 ≡ 0 (mod 4), and the Petersen graph are all examples of bricks which are not 2-extendable.

5.6 Notes

109

Fig. 5.16 The cocktail party graph of order eight

Figure 5.16 depicts the cocktail party graph of order eight. For 푘 ≥ 3, the cocktail party graph of order 2푘 is (푘 − 2)-extendable, but not (푘 − 1)-extendable. We refer the reader to the appendix in the second edition of Lov´asz and Plummer (1989,[60]) for further information on 푘-extendable graphs.

Chapter 6

The Perfect Matching Polytope

Contents 6.1

6.2 6.3

6.4 6.5

Edmonds’ Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111 6.1.1 Convex polytopes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112 6.1.2 The perfect matching polytope of a graph . . . . . . . . . . . . . . 114 The Matching Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121 6.2.1 Linear equality description . . . . . . . . . . . . . . . . . . . . . . . . . . 123 The Dimension Formula . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127 6.3.1 The merger operation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128 6.3.2 Matching lattice and the seminal work of Lov´asz . . . . . . . . 132 Matching Spaces over Finite Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . 135 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135

6.1 Edmonds’ Theorem A geometrical object of great interest in the theory of combinatorial optimization is the perfect matching polytope of a graph. It turns out that tight cut decompositions of a matching covered graph, which were described in Chapter 4, are very relevant to understanding the properties of its perfect matching polytope. The purpose of this chapter is to explore some of these connections. The edge space R퐸 of a graph 퐺 The edge space R퐸 of a graph 퐺 is the set of all real-valued functions on the edge set 퐸 of 퐺. It is convenient to view each member x of R퐸 as vector whose coordinates are indexed by the edges of 퐺, with x(푒) representing the value of x in the position corresponding to edge 푒. © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7_6

111

112

6 The Perfect Matching Polytope

Thus, any vector x in R퐸 may be specified by a drawing of 퐺 in which, for each edge 푒, the value of x(푒) is indicated by juxtaposition. Figure 6.1 displays two members of the edge space of 퐶6 (the missing entries are to be taken as zeros). The support of a vector x in R퐸 is the subset { 푓 ∈ 퐸 : x( 푓 ) ≠ 0}. For example, the support of the vector displayed in Figure 6.1(a) has six edges. Incidence vectors and a useful notation With each subset 퐹 of the edge set 퐸 of a graph 퐺 we associate a vector 휒 퐹 in R퐸 called its incidence vector, where 휒 퐹 (푒) = 1 if 푒 ∈ 퐹, and 휒 퐹 (푒) = 0 if 푒 ∉ 퐹. The vector displayed in Figure 6.1(b) is the incidence vector of the indicated perfect matching in 퐶6 . 퐸 Í Given any vector x in R , and a subset 퐹 of 퐸, we write 퐹x(퐹) to denote {x(푒) : 푒 ∈ 퐹}, which is the same as the dot product of 휒 and x. As an example, take x to be the vector in the edge space of 퐶6 shown in Figure 6.1(a) and 푀 to be the perfect matching indicated by dotted lines in Figure 6.1(b); then x(푀) = 1.

1 2

1 2 1 2

1 2 1 2

1

1

1

1 2

(푎)

(푏)

Fig. 6.1 Two vectors in the edge space of 퐶6 (zero entries not shown)

6.1.1 Convex polytopes The perfect matching polytope of a graph 퐺 is a convex polytope in its edge space. Pertinent material relating to these objects can be found in various books on combinatorial optimization such as the treatises by Lov´asz and Plummer [59] and Schrijver [84]. For the convenience of the readers, we provide here a brief account of the relevant definitions and the notation we use. (Readers who are familiar with the basic theory of convex polytopes may skip this subsection.)

6.1 Edmonds’ Theorem

113

Convex Linear Combinations: Recall that a linear combination of 푟 vectors x1 , x2 , . . . , x푟 in a Euclidean space R푛 over the field R of real numbers is an expression of the form 푟 Õ 훼푖 x푖 , where 훼푖 ∈ R, for 1 ≤ 푖 ≤ 푟. (6.1) Í푟

푖=1

A linear combination 푖=1 훼푖 x푖 of vectors x1 , x2 , . . . , x푟 is a convex linear combiÍ nation if (i) 0 ≤ 훼푖 ≤ 1, for 1 ≤ 푖 ≤ 푟, and (ii) 푟푖=1 훼푖 = 1. (To visualize these notions geometrically, let x1 and x2 be two distinct points in a Euclidean space R푛 . Then a point x in R푛 is a convex linear combination of x1 and x2 if and only if it is on the line segment joining x1 and x2 ; in particular the vector 21 x1 + 21 x2 represents the midpoint of that line segment.) Convex Sets and Convex hulls: A subset 푆 of R푛 is convex if it is closed under the operation of taking convex linear combinations. It is easy to show that the intersection of any two convex sets is a convex set.

The convex hull of x1 , x2 , . . . , x푟 is the set of all convex linear combinations of these vectors. Convex hulls of finite sets are examples of convex sets. Not all convex sets are convex hulls of finite sets. For example, a disc is a convex set but it is not the convex hull of a finite set of points. Extreme Points of Convex Sets: A point x in a convex set 푆 is an extreme point of 푆 if it cannot be expressed as a convex combination of two distinct points in 푆. For example, the line segment joining two distinct points x and y in R푛 has precisely two extreme points, namely x and y. Not all convex sets have extreme points. For example, an open disc (that is, a disc without its boundary) is such a convex set. Convex Polytopes: A subset 푃 of a Euclidean space R푛 is a convex polytope (or, simply, a polytope) if it is the convex hull of some finite subset of R푛 . It should be noted that a polytope may happen to be the convex hull of two different finite subsets of R푛 . For example, the convex hulls of the two sets {(1, 0, 0), (0, 1, 0), (0, 0, 1)} and {(1, 0, 0), (0, 1, 0), (0, 0, 1), ( 12 , 16 , 31 )} are the same, namely the triangle in R3 with the three unit vectors as its vertices. Clearly, every polytope is uniquely defined by the set of its extreme points; it is the convex hull of that set. Many important problems in combinatorial optimization may be expressed as a problem of optimizing a linear function over the set of points in a polytope. For example, the assignment problem [59] may be expressed as the problem of maximizing a linear function over the polytope which is the convex hull of the set of incidence vectors of perfect matchings of a bipartite graph. For applying linear programming techniques to solve such problems, it is necessary to be able to describe polytopes in terms of linear inequalities. All polytopes afford such descriptions as we now proceed to describe. Hyperplanes and half-spaces: A hyperplane in R푛 is the set of all solutions to a single linear equation 푛 Õ 푎 푖 푥푖 = 푏, 푖=1

6 The Perfect Matching Polytope

114

where not all the coefficients 푎 푖 are zero. (In R2 , hyperplanes are lines, in R3 , hyperplanes are planes, etc.) A half-space in R푛 is the set of all solutions to a single linear inequality 푛 Õ 푎 푖 푥푖 ≥ 푏. 푖=1

half-space’. (If the inequality is strict, then the set of its solutions is called an ‘open Í 푛 푎 푖 푥푖 = Here we restrict ourselves to ‘closed half-spaces’.) Note that a hyperplane 푖=1 Í푛 Í푛 푏 is the intersection of the two half-spaces 푖=1 푎 푖 푥푖 ≥ 푏 and 푖=1 푎 푖 푥푖 ≤ 푏.

It is easy to see that hyperplanes and half-spaces are convex sets. Hence the set of all solutions to a system of linear inequalities, being the intersection of a finite number of half-spaces, is a convex set. A classical result due to Minkowski and Weyl (see [84]) states that every convex polytope is the intersection of a finite number of half-spaces. Equivalently, every convex polytope is the set of solutions to a system of linear inequalities. For example, the convex polytope in R3 whose extreme points are the three unit vectors (1, 0, 0), (0, 1, 0) and (0, 0, 1) is the set of solutions to the following system: 푥1 , 푥2 , 푥3 ≥ 0

푥1 + 푥2 + 푥3 = 1

(6.2a) (6.2b)

This establishes the alluded connection between convex polytopes and systems of linear inequalities. Optimizing a linear function over a convex polytope: A fundamental theorem in the theory of convex polytopes states that a the optimal value of a linear function Í푛 푛 푖 푐 푗 푥 푗 over a convex polytope in R occurs at an extreme point of that polytope. Thus, if one could find a system of linear inequalities which describe that polytope, then one would be able to apply linear programming techniques to find the desired optimal solution. For example, suppose that with each edge 푒 of a graph 퐺 there is an associated cost 푐 푒 and our aim is to find a minimum cost perfect matching Í of 퐺. This amounts to finding the minimum value of the function 푒∈퐸 푐 푒 푥 푒 over the set of incidence vectors of perfect matchings of 퐺. Thus, if one could find a linear inequality description of the polytope whose extreme points are the incidence vectors of perfect matchings of 퐺, then it would be possible to find a minimum cost perfect matching of 퐺 by using linear programming techniques. Such a description is given in the following subsection.

6.1.2 The perfect matching polytope of a graph Recall that the set of all perfect matchings of a graph 퐺 is denoted by M (퐺), or simply by M when the graph 퐺 is understood from the context.

6.1 Edmonds’ Theorem

115

The perfect matching polytope The perfect matching polytope of 퐺, denoted by Poly(퐺), is the set of all convex linear combinations of incidence vectors of perfect matchings of 퐺. In other words: Õ Õ  Poly(퐺) = x ∈ R퐸 : x = 훼 푀 휒 푀 , 0 ≤ 훼 푀 ≤ 1, 훼 푀 = 1 . 푀 ∈ M

푀 ∈ M

Jack Edmonds (1965, [30]) established the following description of the perfect matching polytope of a graph in terms of linear inequalities. Edmonds’ characterization of Poly(퐺) Theorem 6.1 The perfect matching polytope Poly(퐺) of a graph 퐺 is the set of solutions to the following system of linear inequalities: x(푒) ≥ 0, for each 푒 ∈ 퐸,

x(휕 (푣)) = 1, for each 푣 ∈ 푉, x(휕 (푆)) ≥ 1, for each odd subset 푆 of 푉.

(6.3a) (6.3b) (6.3c)

We refer to the constraints in (6.3a) as non-negativity constraints, those in (6.3b) as the degree constraints, and those in (6.3c) as the odd set constraints. In case of bipartite graphs, (6.3a) and (6.3b) together imply (6.3c); see Exercise 6.1.1. However, in general, this is not the case. For example, the vector x in the edge space of 퐶6 which is displayed in Figure 6.1(a) satisfies both the non-negativity and degree constraints, but it fails to satisfy (6.3c) for the set 푆 consisting of the three vertices of a triangle in 퐶6 . For each edge 푒 of a graph 퐺, let M 푒 denote the set of perfect matchings of 퐺 that contain the edge 푒. The following observation is a straightforward consequence of the definition of the characteristic vector 휒 푀 : Í Í Proposition 6.2 If x := 푀 ∈ M 훼 푀 휒 푀 then x(푒) = 푀 ∈ M 푒 훼 푀 . ƒ Proof (of Theorem 6.1) Let P(퐺) denote the set of all solutions to the system of linear inequalities consisting of (6.3a), (6.3b) and (6.3c). Our task is to show that the two convex polytopes P(퐺) and Poly(퐺) are one and the same. It is easy to see that, for any perfect matching 푀 of 퐺, its incidence vector 휒 푀 is in P(퐺). This implies that Poly(퐺) is a subset of P(퐺) (Exercise 6.1.2). Our objective is to show that P(퐺) ⊆ Poly(퐺), and deduce that the two polytopes P(퐺) and Poly(퐺) are the same. This is clearly true if P(퐺) = ∅. On the other hand, if P(퐺) is not empty, it can be shown that 퐺 is matchable (Exercise 6.1.3). Since P(퐺) is a convex polytope, every vector in it is a convex linear combination of its extreme points. Thus, to show that P(퐺) ⊆ Poly(퐺), it suffices to show that every extreme point of P(퐺) is in Poly(퐺). Towards this end, let w be an extreme

116

6 The Perfect Matching Polytope

point of P(퐺). Our proof is divided into two main cases. In the first case, we use induction on the number of vertices to show that w is in Poly(퐺). In the second case, we show directly that w is the incidence vector of a perfect matching of 퐺. Case 1 There exists a nontrivial odd cut 퐶 := 휕 (푆) of 퐺 such that w(퐶) = 1. Let 퐺 1 := 퐺/( 푆 → 푠) and 퐺 2 := 퐺/(푆 → 푠) denote the two 퐶-contractions of 퐺. Also let M1 and M2 denote, respectively, the sets of perfect matchings of 퐺 1 and 퐺 2 and define, for 푖 = 1, 2, the vector w푖 as the restriction of w to the edge set 퐸 (퐺 푖 ) of 퐺 푖 . (To make sense of the notation used below, it might be helpful to the reader to work through the details using the graph shown in Figure 6.2 as an example. See Exercise 6.1.4.) Now, using the fact that w ∈ P(퐺) and that w(퐶) = 1, it is easy to check that w1 and w2 satisfy the non-negativity, degree, and odd-set constraints in graphs 퐺 1 and 퐺 2 , respectively (Exercise 6.1.5). Therefore, by the induction hypothesis, w1 and w2 are in Poly(퐺 1 ) and Poly(퐺 2 ), respectively. Thus, there exist coefficients 훽 푀1 , for 푀1 ∈ M1 and 휆 푀2 , for 푀2 ∈ M2 , such that Õ Í w1 = 훽 푀1 휒 푀1 , 훽 푀1 ≥ 0, 훽 푀1 = 1, 푀Õ 1 ∈ M1 Í w2 = 휆 푀2 휒 푀2 , 휆 푀2 ≥ 0, 휆 푀2 = 1. 푀2 ∈ M2

Whenever 푀1 and 푀2 are perfect matchings of 퐺 1 and 퐺 2 , respectively, such that 푀1 ∩ 퐶 = 푀2 ∩ 퐶, the union of 푀1 and 푀2 is a perfect matching of 퐺 which meets 퐶 in just one edge. Consider the set N of all perfect matchings of 퐺 which have exactly one edge in common with 퐶, and define the coefficient 훼 푀 for 휒 푀 , for each 푀 ∈ N, as follows: ( 훽 푀1 휆 푀2 , if w(푒) > 0 w(푒) 훼 푀 := (6.4) 0 otherwise where (i) 푀1 = 푀 ∩ 퐸 (퐺 1 ), (ii) 푀2 = 푀 ∩ 퐸 (퐺 2 ), and (iii) 푒 is the unique edge of 푀 ∩ 퐶. There may be perfect matchings of 퐺 with more than one edge in the cut 퐶. If 푀 is any such perfect matching of 퐺, we take 훼 푀 to be zero. Using (i) the fact that w1 and w2 belong, respectively, to Poly(퐺 1 ) and Poly(퐺 2 ), (ii) Í the definition of the coefficients 훼 푀 , and (iii) the validity of the equality 푒∈퐶 푤(푒) = 1, we now proceed to verify that: Õ Õ (a) 훼 푀 ≥ 0, (b) 훼 푀 = 1, and (c) w = 훼 푀 휒 푀 . 푀 ∈ N

푀 ∈ N

(a) It is easy to see 훼 푀 ≥ 0, for each 푀 ∈ N.

(b) We shall use the following notation in our verification of parts (b) and (c). For each 푒 ∈ 퐶 and for 푖 = 1, 2, we let M푖푒 denote the set of all perfect matchings of 퐺 푖 whose intersection with 퐶 is {푒}. Every perfect matching in N is of the form 푀1 ∪ 푀2 , where 푀1 belongs to M1푒 and 푀2 belongs to M2푒 . Thus, for any 푒 ∈ 퐶

6.1 Edmonds’ Theorem

117 1 2

퐶

1 6 1 3 1 3

퐺

1 2

1 6

1 6 1 2

1 3

1 3

1 3 1 3

퐺1

1 6

퐺2

1 2

1 3

1 6

1 3

1 2

1 6

1 6

1 2

1 2 1 3

1 3

1 3

1 3 1 3

Fig. 6.2 Proof of Edmonds’ Theorem, Case 1

with w(푒) ≠ 0, using the definitions of the 훼, 훽, and 휆-coefficients (6.4), it can be seen (Exercise 6.1.6) that: Õ

푀1 ∈ M1푒 , 푀2 ∈ M2푒

훼 푀1 ∪푀2 =

Õ

푀1 ∈ M1푒 , 푀2 ∈ M2푒

훽 푀1 휆 푀2 = w(푒). w(푒)

(6.5)

Í Í By the hypothesis of the case under consideration, 푒∈퐶 w(푒) = 1. Thus 푀 ∈ N 훼 푀 = 1. Í (c) It remains to be seen that w = 푀 ∈ N 훼 푀 휒 푀 . The proof of this relies on Proposition 6.2. Equation 6.5 and Proposition 6.2 show Í that the two vectors 푀 ∈ N 훼 푀 휒 푀 and w agree in the coordinates corresponding to edges in the cut 퐶. To complete the proof of (c), it is necessary to show that this agreement also holds for edges of 퐺 which are not in 퐶. We leave the details as Exercise 6.1.7. The three observations made above imply that w is a convex linear combination of the incidence vectors of perfect matchings of 퐺; in other words that w is in Poly(퐺). In fact, as w is an extreme point of the polytope P(퐺), it follows that w is itself the incidence vector of a perfect matching of 퐺.

118

6 The Perfect Matching Polytope

Case 2 w(퐶) > 1 for every nontrivial odd cut 퐶 := 휕 (푆) of 퐺. Let 퐺 [w] denote the subgraph of 퐺 induced by the support of w. The constraints 6.3a and 6.3b ensure that 퐺 [w] is a spanning subgraph of 퐺 with no isolated vertices. Our objective is to show that the edge set of 퐺 [w] is a perfect matching of 퐺. We start with the following simple observation. 6.2.1 The graph 퐺 [w] has no even cycles. Proof Assume to the contrary that 퐺 [w] has an even cycle 푄 whose edges in cyclic order are 푒 1 , 푒 2 , . . . , 푒 2푘 . Now define two vectors x and y in ‘close-proximity’ of w as follows:    w(푒) + 휖  x(푒) = w(푒) − 휖    w(푒)

if 푒 = 푒 1 , 푒 3 , . . . , 푒 2푘−1 if 푒 = 푒 2 , 푒 4 , . . . , 푒 2푘 if 푒 ∉ 퐸 (푄)

  w(푒) − 휖 if 푒 = 푒 1 , 푒 3 , . . . , 푒 2푘−1   w(푒) + 휖 if 푒 = 푒 2 , 푒 4 , . . . , 푒 2푘   w(푒) if 푒 ∉ 퐸 (푄).  Using the fact that w belongs to P(퐺) and that w(퐶) > 1 for each nontrivial odd cut 퐶 of 퐺, it is now easy to verify that there exists a positive 휖 such that both x and y are also in P(퐺). Also, w = 21 x + 12 y, for any value of 휖. This is impossible because w is an extreme point of P(퐺). ƒ and

y(푒) =

It is an elementary fact that if a graph is free of even cycles then each of its blocks is either 퐾1 , or 퐾2 , or is an odd cycle (Exercise 6.1.8). It follows that every block of 퐺 [w] of order three or more is an odd cycle. Suppose that an odd cycle 푄 is an end block of a component of 퐺 [w]. Then either 푄 by itself is that component, or it is attached to the rest of that component at a single cut vertex 푣. In the former case 휕 (푄) is empty and w(휕 (푄)) = 0 < 1, and in the latter case, 휕 (푄) is a proper subset of 휕 (푣) and w(휕 (푄)) < 휕 (푣) = 1 (see Figure 6.3). Thus, in either case, w(휕 (푄)) < 1 which violates (6.3c). It follows that no end block of 퐺 [w] is an odd cycle.

휕(푄)

푣 푄

Fig. 6.3 If an odd cycle 푄 is an end block of 퐺 [w], then w(휕(푄) ) < 1

6.1 Edmonds’ Theorem

119

On the other hand, it follows from 6.3a and 6.3b that the graph 퐺 [w] has no isolated vertices, and if it has a vertex of degree one, then the edge incident with that vertex is the only edge of a component of 퐺 [w]. We conclude that 퐺 [w] is a spanning 1-regular subgraph of 퐺 or, equivalently, that w is the incidence vector of a perfect matching of 퐺. ƒ Edmonds’ own proof of the above theorem was based on a polynomial-time algorithm for finding a maximum weight perfect matching in a weighted graph (1965, [30]). For a description of that algorithm, and other proofs of this theorem, we refer the reader to Schrijver’s treatise [84]. The following result concerning perfect matching polytopes of matching covered graphs may be proved by adapting the arguments used in Case 1 of the proof of Theorem 6.1. Theorem 6.3 Let 퐶 be a tight cut of a matching covered graph 퐺 and let 퐺 1 and 퐺 2 be the two 퐶-contractions of 퐺. A vector x in R퐸 is in Poly(퐺) if and only if the restrictions of x to 퐸 (퐺 1 ) and 퐸 (퐺 2 ) are in Poly(퐺 1 ) and Poly(퐺 2 ), respectively. ƒ Thus, when 퐺 is a matching covered graph, to check if a vector x is in the perfect matching polytope of 퐺 it suffices to check if the restrictions of x to the edge sets of the bricks and braces of 퐺 are in their respective perfect matching polytopes.

Exercises ⊲6.1.1 Let 퐺 be a bipartite graph, and let x be any vector in its edge space. Show that if x satisfies (6.3a) and (6.3b), then it also satisfies (6.3c). Hint: Use the fact that for any subset 푆 of 푉: x(휕 (푆)) ≥ x(휕 (푆+)) − x(휕 (푆 − )) = x(푆+ ) − x(푆 − ) = |푆+ | − |푆 − |. ∗ 6.1.2 In the proof of Theorem 6.1, explain why Poly(퐺) is a subset of P(퐺). ⊲6.1.3 Show that if the polytope P(퐺) defined in the proof of Theorem 6.1 is nonempty, then 퐺 is matchable. Hint: let x ∈ P(퐺) and show that for each subset 푆 of 푉, Õ 표(퐺 − 푆) ≤ x(휕 (퐾)) ≤ x(휕 (푆)) ≤ x(푆) = |푆|. 퐾 ∈ O (퐺−푆)

∗ 6.1.4 Consider the graph 퐺 and its two 퐶-contractions 퐺 1 and 퐺 2 shown in Figure 6.2. The weights attached to the edges of 퐺 are the components of a vector w which satisfies the conditions necessary for it to belong to Poly(퐺). (i) For 푖 = 1, 2, express the restriction w푖 of w to 퐸 (퐺 푖 ) as a convex combination of incidence vectors of perfect matchings of 퐺 푖 . (ii) Then, using the method of proof of Theorem 6.1 find a convex linear combination of incidence vectors of perfect matchings of 퐺 which is equal to the vector w.

6 The Perfect Matching Polytope

120

⊲6.1.5 In Case 1 of the proof of Theorem 6.1, show that, for 푖 = 1, 2, the vector w푖 is in P(퐺 푖 ). ⊲6.1.6 Verify the validity of Equation (6.5). Hint: Note that

Õ

푀1 ∈ M1푒 , 푀2 ∈ M2푒

Í

and use the fact that

푀1 ∈ M1푒

훽 푀1 휆 푀2 = w(푒)

훽 푀1 =

Í

Õ

푀1 ∈ M1푒

푀2 ∈ M2푒

훽 푀1 w(푒)

Õ

휆 푀2

푀2 ∈ M2푒

휆 푀2 = w(푒).

∗ 6.1.7 In Case 1 of the proof of Theorem 6.1, verify that w =

Í

푀 ∈ M

훼 푀 .

6.1.8 Show that if a graph 퐺 has no even cycles, then each block of 퐺 is either 퐾1 , 퐾2 , or an odd cycle. 6.1.9 Give a proof of Theorem 6.3. 6.1.10 Let 퐺 be an 푟-graph (as defined in Exercise 1.3.6). Show that the vector ( 푟1 , 푟1 , · · · , 푟1 ) is in Poly(퐺). The perfect matching cone The Perfect Matching Cone of a graph 퐺, denoted Cone(퐺), is the set of all nonnegative linear combinations of incidence vectors of perfect matchings of 퐺. In other words: Õ Cone(퐺) := {x ∈ R퐸 : x = 훼 푀 휒 푀 , 훼 푀 ≥ 0}. 푀 ∈ M

Theorem 6.4 (A characterization of Cone(퐺)) Let 퐺 be a graph. A vector x ∈ R퐸 is in Cone(퐺) iff it satisfies the three restraints below, for some 푟 ∈ R: x(푒) ≥ 0, for each 푒 ∈ 퐸, x(휕 (푣)) = 푟, for each 푣 ∈ 푉,

(6.6a) (6.6b)

x(휕 (푆)) ≥ 푟, for each odd subset 푆 of 푉.

(6.6c)

⊲6.1.11 Prove that Theorem 6.1 implies Theorem 6.4 and, conversely, that Theorem 6.4 implies Theorem 6.1. (Hint: let C(퐺) denote the set of all solutions to the system of linear inequalities consisting of (6.6a), (6.6b) and (6.6c), let x ∈ R+퐸 − 0, let 푟 := x(휕 (푣)) for some vertex 푣 and consider the equivalence of the following statements if 푟 > 0: x ∈ Cone(퐺),

y :=

x ∈ Poly(퐺), 푟

y ∈ P(퐺),

and x ∈ C(퐺).)

∗ 6.1.12 The objective of this exercise is to give a proof of Theorem 6.4, due to Seymour (1979, [85]). For this, adopt the definition of C(퐺) given in Exercise 6.1.11. Let x ∈ C(퐺), let 푟 := x(휕 (푣)) for some vertex 푣 and let 푆 be the support of x. Prove that x ∈ Cone(퐺) by induction on |푉 | + |푆|.

6.2 The Matching Space

121

(i) Prove that if 푆 is nonempty then the graph 퐺 [푆] of 퐺 induced by 푆 is a spanning matchable subgraph of 퐺. Thus, let 푀 be a perfect matching of 퐺 [푆] (and of 퐺). (ii) Suppose that 퐺 has a nontrivial odd cut 퐶 such that x(퐶) = 푟; apply the inductive hypothesis to both 퐶-contractions of 퐺, as in the first part of the proof of Theorem 6.1. (iii) Suppose that x(퐶) > 푟 > 0 for each nontrivial odd cut of 퐺. Let   x(퐶) − 푟 휖 := min {x(푒) : 푒 ∈ 푀 } ∪ { : |푀 ∩ 퐶| > 1, 퐶 odd} . |푀 ∩ 퐶| − 1

Let y := x − 휖 · 휒 푀 and prove that (i) y ∈ C(퐺) and (ii) either the support of y is a proper subset of 푆 or the previous case is applicable with y playing the role of x.

6.2 The Matching Space ♯

The perfect matching space The perfect matching space of a graph 퐺 is the linear subspace of R퐸 generated by the set { 휒 푀 } := { 휒 푀 : 푀 ∈ M} of incidence vectors of perfect matchings of 퐺. We denote this space by Lin(퐺) and, for brevity, refer to it simply as the matching space. The purpose of this section is to present a characterization of the matching space of a matching covered graph 퐺, and to determine a formula for its dimension (which, interestingly, involves the parameter 푏(퐺), the number of bricks of the graph 퐺.) Regular vectors We denote the set of all tight cuts in a matching covered graph 퐺 by T (퐺). A vector x in the edge space R퐸 of 퐺 is 푟-regular if x(퐶) = 푟, for each 퐶 ∈ T (퐺).

(6.7)

A regular vector is one which is 푟-regular for some real number 푟. The following result establishes a necessary condition for a vector to be in the matching space of a matching covered graph 퐺. Proposition 6.5 Every vector x in Lin(퐺) is regular.

122

6 The Perfect Matching Polytope

Proof Since x is in Lin(퐺), by definition, it may be expressed as a linear combination of incidence vectors of perfect matchings of 퐺. Thus, there exist coefficients Í 훼 푀 , 푀 ∈ M, such that x = 푀 ∈ M 훼 푀 휒 푀 . Now let 퐶 be any tight cut of 퐺. Then, using the fact that |푀 ∩ 퐶| = 1, for each 푀 ∈ M, we have: Õ Õ (6.8) x(퐶) = 훼 푀 . 훼 푀 휒 푀 (퐶) = 푀 ∈ M

It follows that x is 푟-regular, with 푟 =

Í

푀 ∈ M

푀 ∈ M

훼 푀 .

ƒ

Every vector in Poly(퐺) is a non-negative 1-regular vector, by Proposition 6.2; but not every 1-regular vector is in Poly(퐺). For example, the vectors x and y displayed in Figure 6.4 are 1-regular vectors in the edge space of 퐺 := 퐶6 . But, only one of them, namely y, is in Poly(퐺) (see Exercise 6.2.1). However, as the following Lemma shows, both x and y and, in fact, all 1-regular vectors are in Lin(퐺). 1 2

−1 1

1 −1

1 1

1 4

1 4

1

1 2

1 4 1 4

1

1 4 1 4

1 2

−1 x

y

Fig. 6.4 Two 1-regular vectors in the edge space of 퐶6

Lemma 6.6 Every 1-regular vector x in R퐸 belongs to Lin(퐺). Proof Every 1-regular vector in Poly(퐺) is clearly in Lin(퐺). Suppose that x is a 1-regular vector that is not in Poly(퐺). To show that x is in Lin(퐺), it suffices to show that it can be expressed as a linear combination of two vectors y and z in Poly(퐺). Í Let y = 푀 ∈ M 훼 푀Í 휒 푀 be a vector in Poly(퐺) in which each 훼 푀 is strictly positive. (For example, 푀 ∈ M |M | −1 휒 푀 is such a vector.) Then, y is 1-regular and strictly positive. Furthermore, for each odd cut 퐶 that is not tight, y(퐶) =

Õ

푀 ∈ M

|푀 ∩ 퐶| 훼 푀 >

Õ

훼 푀 = 1,

푀 ∈ M

because |푀 ∩ 퐶| ≥ 1 for all perfect matchings 푀, with strict inequality for at least one perfect matching. Now it can be verified that, for any ‘sufficiently’ small positive real number 휖, the vector z := (1 − 휖)y + 휖x

6.2 The Matching Space

123

is also a vector in Poly(퐺) (see Exercise 6.2.1). Since both y and z are in Poly(퐺), they are in Lin(퐺). We conclude that   1 1 x= 1− y+ z 휖 휖 is also in Lin(퐺).

ƒ Characterization of the matching space

Theorem 6.7 A vector x in the edge space of a matching covered graph 퐺 belongs to Lin(퐺) if and only if it is regular. Proof If x is in Lin(퐺) then, by Proposition 6.5, it is regular. To prove the converse, we must show that every regular vector x in R퐸 is in Lin(퐺). By Lemma 6.6, every 1-regular vector is in Lin(퐺). So, suppose that x is an 푟regular vector, where 푟 ≠ 1. Let y be any 1-regular vector, and consider the vector w defined as follows:  1 w := x−y (6.9) 푟 − 1 It is easy to check that this vector w is 1-regular (Exercise 6.2.2). Therefore, by Lemma 6.6, the vector w is in Lin(퐺); hence the vector x = (푟 − 1)w + y is also in Lin(퐺). ƒ We shall now proceed to derive a formula for the dimension of the matching space of a matching covered graph. It turns out that in the case of braces (and more generally, of all bipartite graphs), and in the case of bricks, this formula may be derived using elementary linear algebra. After dealing with these two special cases, we shall use the tight cut decomposition procedure to deduce the formula for the dimension of the matching space of an arbitrary matching covered graph by induction.

6.2.1 Linear equality description As Lin(퐺) is a subspace of R퐸 , it is the set of solutions to a system of linear homogeneous equations. We begin by introducing the notation necessary for describing such a system of equations. (We shall adopt the convention that all vectors in R퐸 are row vectors.) Let 퐺 be any matching covered graph. Recall that T := T (퐺) denotes the set of all tight cuts (including the trivial cuts) of 퐺. We shall denote the |T | × 푚 matrix whose rows are the incidence vectors of the tight cuts of 퐺 by T. Figure 6.5 shows an example of a graph 퐺 and the matrix T, where 퐶 = 휕 ({푤, 푥, 푦}), and 퐷 = 휕 ({푢, 푤, 푥}). We note that the rows of the 푛 × 푚 incidence matrix A of 퐺 are

6 The Perfect Matching Polytope

124

just the incidence vectors of the trivial tight cuts of 퐺. Thus A is an 푛 × 푚 submatrix of T, and if 퐺 is free of nontrivial tight cuts, then T = A. 푢

푣

푤

푥

푦

푧

푢푣 푢푤 푢푥 푣푧 푤푥 푤푦 푥 푦 푦 푧 휕(푢) : 1 1 1 0 0 0 0 0 휕(푣 ) : 1

0

0

1

0

0

0

0

휕(푤) : 0

1

0

0

1

1

0

0

휕( 푥 ) : 0

0

1

0

1

0

1

0

휕( 푦 ) : 0

0

0

0

0

1

1

1

휕(푧) : 0

0

0

1

0

0

0

1

퐶

: 0

1

1

0

0

0

0

1

퐷

: 1

0

0

0

0

1

1

0

Fig. 6.5 The matrix T(퐺 ) (the first six rows constitute A(퐺 ))

Suppose that 푟 is a fixed real number. By definition, a vector x in R퐸 is 푟-regular if and only x(퐶) = 푟, for each 퐶 ∈ T , which amounts to saying that, for each 퐶 ∈ T , the dot product of 휒퐶 and x is 푟. Thus, in matrix notation, the set of all 푟-regular vectors is the set of all solutions to the system Tx푡 = r,

(6.10)

of linear equations where (i) T is the |T | × 푚 matrix defined above, (ii) x is the 1 × 푚 vector of variables, one per each edge, and (iii) r is the |T | × 1 column vector each of whose entries is the real number 푟. (By our convention, all vectors in R퐸 are row vectors. Thus, for the sake of consistency of matrix multiplication, it is necessary to take the transpose of x.) Let Lin0 (퐺) denote the subset of Lin(퐺) consisting of all 0-regular vectors. It follows from the above observation that Lin0 (퐺) is the subspace of R퐸 consisting of all solutions to the system Tx푡 = 0 of linear homogeneous equations. Since the number of variables is 푚 (one per each edge), it follows that dim(Lin0 (퐺)) = 푚 − rank(T).

6.2 The Matching Space

125

A basis of Lin0 (퐺) together with any 1-regular vector clearly constitutes a basis of Lin(퐺) (Exercise 6.2.5). We therefore have: Theorem 6.8 The dimension of the matching space of 퐺 is: dim(Lin(퐺)) = 푚 + 1 − rank(T).

(6.11)

In view of the above theorem, determining a formula for the dimension of Lin(퐺) reduces to determining a formula for the rank of the matrix T. (This was the approach used by Naddef (1982, [76]) to determine the dimension of Lin(퐺).) Let us first consider the case in which 퐺 is a graph which is free of nontrivial tight cuts. In this case, the matrix T is identical to the incidence matrix A of 퐺. Using the fact that 퐺 is connected, it is not difficult to show (Exercise 6.2.7) the following:  푛 − 1 if 퐺 is bipartite, rank(A) = (6.12) 푛 otherwise. This, combined with Equation (6.11), immediately yields the following result: Dimension formulas for bricks and braces Theorem 6.9 The dimension of Lin(퐺) is 푚 −푛 +2 if 퐺 is a brace, and 푚 −푛 +1 ƒ if 퐺 is a brick.

For example the dimension of the matching space of 퐾3,3 is 9 − 6 + 2 = 5, and the dimension of the matching space of the pentagonal prism is 15 − 10 + 1 = 6.

Let 퐺 ( 퐴, 퐵) be a bipartite matching covered graph, and let 퐶 := 휕 ( 푋) be a tight cut of 퐺. Then, by Theorem 4.8, there exist partitions of 퐴 into ( 퐴1 , 퐴2 ) and 퐵 into (퐵1 , 퐵2 ) such that (i) | 퐴1 | = |퐵1 | + 1 and | 퐴2 | = |퐵2 | − 1, and (ii) every edge of the cut 퐶 has one end in 퐴1 and one end in 퐵2 . It follows that the incidence vector of 퐶 is the sum of the incidence vectors of the trivial cuts associated with vertices in 퐴1 minus the sum of the trivial cuts of the vertices in 퐵1 . Hence the incidence vector of any tight cut of 퐺 is in the row space of the incidence matrix of 퐺. Since 퐺 is bipartite, the rank of the incidence matrix of 퐺 is 푛 − 1. It now follows from Theorem 6.8 that the dimension of the matching space of 퐺 is 푚 − 푛 + 2. Similarly, when 퐺 is a near-brick (a graph with 푏(퐺) = 1), the incidence vector of any tight cut may be expressed as an affine linear combination of incidence vectors of trivial cuts, and the dimension of Lin(퐺) is 푚 − 푛 + 1 (Exercise 6.2.3). The next section is dedicated to finding a formula for the dimension of the matching space of a matching covered graph that takes into account the number of its bricks.

126

6 The Perfect Matching Polytope

Exercises ∗ 6.2.1

(i) Consider the two vectors y and x in the edge space of 퐺 := 퐶6 shown in Figure 6.4. Show that both y and x are 1-regular vectors, where y is in Poly(퐺) and x is not. Furthermore, find a value of 휖 such that z := (1 − 휖)y + 휖x is in Poly(퐺). (ii) In the proof of Lemma 6.6, show that z is in Poly(퐺) for all small enough positive values of 휖. 6.2.2 Show that the vector w defined in Equation (6.9) is a 1-regular vector.

⊲6.2.3 Let 퐺 be a matching covered graph with at most one brick. (i) Show that the incidence vector of any tight cut can be expressed as an affine linear combination of incidence vectors of trivial tight cuts. (ii) Deduce that a vector x in Z퐸 is regular if and only if x(휕 (푢)) = x(휕 (푣)), for any two vertices 푢 and 푣 of 퐺. ⊲6.2.4 Consider the matching covered graph 퐺 with two bricks which is shown in Figure 6.6.

퐶1

퐶2

Fig. 6.6 A matching covered graph 퐺 with two bricks

(i) Find a non-negative vector x in R퐸 such that x(휕 (푣)) = 1, for all 푣 ∈ 푉, with the property that neither x(퐶1 ) nor x(퐶2 ) is equal to 1. (Such a vector satisfies the non-negativity and degree constraints stipulated in the statement of Edmonds’ Theorem 6.1 without being a 1-regular vector.) (ii) For 푖 = 1, 2, show that the incidence vector of the cut 퐶푖 can be expressed as an affine linear combination of incidence vectors of trivial tight cuts and the cut 퐶3−푖 . ∗ 6.2.5 Show that a basis for Lin0 (퐺) together with any 1-regular vector y constitutes a basis for Lin(퐺). (Hint: Let x be any 푟-regular vector, where 푟 ≠ 0; then 푟1 x − y is a 0-regular vector.)

6.3 The Dimension Formula

127

Matching orthogonal vectors A vector x in the edge space of a matching covered graph 퐺 is matching orthogonal if it belongs to (Lin(퐺)) ⊥ ; equivalently, if x(푀) = 0, for each perfect matching 푀 of 퐺. ∗ 6.2.6 Let 퐺 be a matching covered graph and let T be the matrix whose rows are the incidence vectors of the tight cuts of 퐺. Let 퐷 be a tight cut of 퐺 and let B denote the matrix obtained from T by replacing each row 휒퐶 , 퐶 ≠ 퐷, by 휒퐶 − 휒 퐷 . (i) Prove that Bx = 0 if and only if x ∈ Lin(퐺) and deduce that the rows of B span Lin(퐺) ⊥ . Í (ii) Let p be any vector in R푉 such that 푣 ∈푉 푝(푣) = 0. Then show that the vector 퐸 x in R induced by p where, for each edge 푢푣 of 퐺, x(푢푣) = p(푢) + p(푣), is a matching orthogonal vector. (iii) Now suppose that 퐺 is a brick. Using the characterization of the matching space of a brick and part (i), show that every matching orthogonal vector in R퐸 is a function as described in part (ii). ∗ 6.2.7 Give a proof of the formula (6.12) for the rank of A. (Hint: First show that rank(A) ≥ 푛 − 1 by showing that the columns corresponding to the edges of a spanning tree of 퐺 are linearly independent. Then note that any linear combination Í of the rows of A is of the form 푣 ∈푉 푝(푣)휕 (푣). In such a linear combination, the component corresponding to an edge 푢푤 is 푝(푢) + 푝(푤). Hence the rows of A are linearly dependent if and only if there is an assignment of weights 푝(푣) to 푣 ∈ 푉 such that the ‘induced’ weights (푝(푢) + 푝(푤)) on the edges of 퐺 are all zero.) 6.2.8 Find two different bases for the row space of the matrix T(퐺), where 퐺 is the graph shown in Figure 6.5. 6.2.9 (i) Express the dimension of the matching space of a cubic brick of order 푛 as a function of 푛. (ii) Find bases for the matching spaces of the Petersen graph and the pentagonal prism.

6.3 The Dimension Formula ♯ If a matching covered 퐺 is neither a brace nor a brick, then it has a nontrivial tight cut 퐶. In this case, it is possible to obtain a basis for the matching space of 퐺 by combining in a certain manner bases for the matching spaces of the two 퐶-contractions of 퐺. This involves the operation defined below.

128

6 The Perfect Matching Polytope

6.3.1 The merger operation

Let 퐶 be a separating cut of a matching covered graph 퐺, and let 퐺 1 and 퐺 2 be the two 퐶-contractions of 퐺. For 푖 = 1, 2, let B푖 denote a set of perfect matchings of 퐺 푖 such that { 휒 푀 : 푀 ∈ B푖 } is a basis of Lin(퐺 푖 ). We shall now describe a procedure which may be used to obtain from B1 and B2 a set B := B1 ∨ B2 of |B푖 | + |B2 | − |퐶| perfect matchings of 퐺 whose incidence vectors are linearly independent and, when 퐶 is a tight cut, constitute a basis of Lin(퐺). (Although our interest here is only in tight cuts, with future applications in mind, we allow the possibility of 퐶 being a separating cut which may not be tight.) In describing the above-mentioned procedure and its properties, for brevity, we shall sometimes informally refer to a set of perfect matchings as being independent if their incidence vectors are independent. Similarly, we shall refer to a set of perfect matchings as a basis of the matching space of a graph to mean that the set of their incidence vectors is a basis of that space. For 푖 = 1, 2, each perfect matching of 퐺 푖 clearly contains precisely one edge incident with the contraction vertex in 퐺 푖 , and the edges of 퐺 푖 incident with this vertex are in one-to-one correspondence with the edges of 퐶. We obtain B1 ∨ B2 as follows: • For 푖 = 1, 2, group the members of B푖 according to the edge of 퐶 they contain. For each 푒 ∈ 퐶, denote the subset of perfect matchings in B푖 containing 푒 by B푖푒 . • For each 푒 ∈ 퐶, and 푖 = 1, 2, fix (select) a perfect matching 퐹푖푒 from B푖푒 . • For any perfect matching 푀1 of 퐺 1 in B1푒 , and the fixed perfect matching 퐹2푒 of 퐺 2 , the set 푀1 ∪ 퐹2푒 is a perfect matching of 퐺. Similarly, for the fixed perfect matching 퐹1푒 of 퐺 1 and for any perfect matching 푀2 of 퐺 2 in B2푒 , the set 퐹1푒 ∪ 푀2 is a perfect matching of 퐺. Define the set B 푒 of perfect matchings and the perfect matching 퐹 푒 as below: B 푒 := {푀1 ∪ 퐹2푒 : 푀1 ∈ B1푒 } ∪ {퐹1푒 ∪ 푀2 : 푀2 ∈ B2푒 }

and

퐹 푒 := 퐹1푒 ∪ 퐹2푒 .

• Now let B := ∪푒∈퐶 B 푒 . We denote this set of perfect matchings of 퐺 by B1 ∨ B2 and refer to it as the merger of B1 and B2 . It can be seen that |B 푒 | = |B1푒 | + |B2푒 | − 1, since 퐹1푒 ∪ 퐹2푒 is counted twice in the sum |B1푒 | + |B2푒 |. Consequently, |B1 ∨ B2 | = |B1 | + |B2 | − |퐶|.

(6.13)

Example 6.10 Consider the graph 퐺 shown in Figure 6.7(a) in which the cut 퐶 is a tight cut. The graph 퐺 1 , one of the two 퐶-contractions of 퐺, is the pentagonal prism; it has a basis B1 consisting of six perfect matchings of which two, 푀1 and 푀2 , contain the edge 푒. In Figure 6.7(b), the edges labeled 푖 are the edges in 푀푖 , 푖 = 1, 2. The graph 퐺 2 , the other 퐶-contraction of 퐺, is the graph obtained by splicing 퐾3,3 and 퐾4 ; it has a basis B2 consisting of five perfect matchings of which two, 푁1 and 푁2 , contain 푒. In Figure 6.7(c), the edges labeled 푖 are the edges in 푁푖 , 푖 = 1, 2.

6.3 The Dimension Formula

129

The result of the merger operation on edge 푒 results in three perfect matchings of 퐺, 푀1 ∪푁1 , 푀1 ∪푁2 and 푀2 ∪푁1 , labeled respectively 11, 12 and 21 in Figure 6.7(a); the perfect matchings 푀1 of 퐺 1 and the perfect matching 푁1 of 퐺 2 play, respectively, the roles of 퐹1푒 and 퐹2푒 .

푋

11, 21

21 11, 12

12

21 11, 12

11, 12, 21 푒

11, 12 21

11, 21 12

11, 12 12

21

11, 21 퐶 (a) 퐺 1

2 2

1 2 1

1, 2 푒

1

푥

푥

1, 2 푒

2 1

1

2

2 1

2 (b) 퐺1

(c) 퐺2

Fig. 6.7 Merger of bases of tight cut-contractions

Theorem 6.11 Let 퐺 1 and 퐺 2 be the two 퐶-contractions of a matching covered graph 퐺 with respect to a separating cut 퐶, and, for 푖 = 1, 2, let B푖 denote a set of perfect matchings of 퐺 푖 whose incidence vectors constitute a basis of Lin(퐺 푖 ). Then: (i) the set of incidence vectors of B := B1 ∨ B2 is an independent subset of Lin(퐺); (ii) furthermore, if 퐶 is a tight cut, then this set is a basis of Lin(퐺).

6 The Perfect Matching Polytope

130

Proof Let us first show that { 휒 푀 : 푀 ∈ B} is a linearly independent set. This requires us to show 훼 푀 = 0, for all 푀 ∈ B, is the only solution to the equation Õ 훼 푀 휒 푀 = 0, (6.14) 푀 ∈ B

where 0 is the vector of all zeros. We shall establish this by considering the restrictions of the above equation to 퐸 (퐺 1 ) and to 퐸 (퐺 2 ) and using the fact that B1 and B2 are ‘bases’ of Lin(퐺 1 ) and Lin(퐺 2 ), respectively.

Every perfect matching in B is either of the form 푀1푒 ∪ 퐹2푒 , for some 푀1푒 ∈ B1푒 , or of the form 퐹1푒 ∪ 푀2푒 for some 푀2푒 ∈ B2푒 . Thus, the restrictions of members of B to 퐸 (퐺 1 ) are perfect matchings in B1 . For each 푒 ∈ 퐶, a perfect matching 푀1푒 in B1푒 which is different from 퐹1푒 , appears as the restriction of just one perfect matching in B, namely 푀1푒 ∪ 퐹2푒 . However, for each 푒 ∈ 퐶, the perfect matching 퐹1푒 arises as the restriction of each of the perfect matchings in B of the form 퐹1푒 ∪ 푀2푒 where 푀2푒 ∈ B2푒 . (Similar statements apply to restrictions of members of B to 퐸 (퐺 2 ).) In view of the above observation, by restricting the equation (6.14) to 퐸 (퐺 1 ) we obtain the equation:  푒    Õ  Õ Õ 푒 (6.15) 훼퐹 푒 ∪ 푀 푒 휒 퐹1 = 0. 훼 푀 푒 ∪ 퐹 푒 휒 푀1 + 푒∈퐶

푀1푒 ∈ B1푒 −퐹1푒

1

2

푀2푒 ∈ B2푒

1

2

Since, by hypothesis, the set of incidence vectors of B1 is a linearly independent subset of Lin(퐺 1 ), it follows that 훼 푀 = 0 for all perfect matchings 푀 in B of the form 푀1푒 ∪ 퐹2푒 , where 푀1푒 ∈ B1푒 − 퐹1푒 , for any 푒 ∈ 퐶. Likewise, 훼 푀 = 0 for all perfect matchings 푀 in B of the form 퐹1푒 ∪ 푀2푒 , where 푀2푒 ∈ B2푒 − 퐹2푒 , for any 푒 ∈ 퐶. 푒 When we take these two facts into account, the coefficient of 휒 퐹1 in equation (6.15) reduces to 훼퐹1푒 ∪퐹2푒 . Now we can infer that 훼 푀 = 0, for all 푀 ∈ B, establishing that { 휒 푀 : 푀 ∈ B} is a linearly independent set. Now we turn to the proof of the second part of the statement which asserts that { 휒 푀 : 푀 ∈ B} is a basis of Lin(퐺) under the assumption that 퐶 is a tight cut. To prove this, it suffices to show that the incidence vector of any perfect matching of 퐺 is in the span of { 휒 푀 : 푀 ∈ B}. So, let 푁 be a perfect matching of 퐺. Since 퐶 is a tight cut, 푁 has precisely one edge in 퐶. Let 푁 ∩ 퐶 = {푒}, and let 푁1 and 푁2 denote the restrictions of 푁 to 퐸 (퐺 1 ) and 퐸 (퐺 2 ), respectively. Since, by hypothesis, { 휒 푀1 : 푀1 ∈ B1 } is a basis of Lin(퐺 1 ), it follows that there exist coefficients 훼 푀1 such that Õ (6.16) 훼 푀1 휒 푀1 . 휒 푁1 = 푀1 ∈ B1

The set 푁1 is a perfect matching of 퐺 1 . Thus, 휒 푁1 is 1-regular in 퐺 1 . By Proposition 6.5, Õ 훼 푀1 = 1. 푀1 ∈ B1

6.3 The Dimension Formula

131

Now, using the above 훼 푀1 ’s, define the vector x in R퐸 as follows: Õ 푒 훼 푀1 휒 푀1 ∪퐹2 . x :=

(6.17)

푀1 ∈ B1

Clearly, x is in the span of { 휒 푀 : 푀 ∈ B}. By (6.16), the restriction of x to 퐸 (퐺 1 ) 푒 is 휒 푁1 . And, by (6.17), the restriction of x to 퐸 (퐺 2 ) is 휒 퐹2 . By an analogous argument, it follows that there exists a vector Õ 푒 훽 푀2 휒 퐹1 ∪푀2 y := 푀2 ∈ B2

in the span of { 휒 푀 : 푀 ∈ B} such that the restriction of y to 퐸 (퐺 2 ) is 휒 푁2 , and the 푒 restriction of y to 퐸 (퐺 1 ) is 휒 퐹1 . To complete the proof, we note the easily verified identity (Exercise 6.3.1): 푒

푒

푒

푒

휒 푁 = 휒 푁1 ∪퐹2 + 휒 퐹1 ∪푁2 − 휒 퐹1 ∪퐹2 . Hence

푒

(6.18)

푒

휒 푁 = x + y − 휒 퐹1 ∪퐹2

and it follows that 휒 푁 is in the span of { 휒 푀 : 푀 ∈ B}. We conclude that ƒ { 휒 푀 : 푀 ∈ B} is a basis of Lin(퐺). When 퐶 is a separating cut that is not tight, the above arguments do not apply, and B is not a basis for Lin(퐺). However, the following statement may be easily deduced (Exercise 6.3.4). Theorem 6.12 Let 퐺 be a matching covered graph, and let 퐶 be a separating cut of 퐺 which is not tight, and let 퐺 1 , 퐺 2 and B be as in the statement of Theorem 6.11. Then B is a basis for the subspace Lin(퐺, 퐶) of Lin(퐺) generated by the set of incidence vectors of those perfect matchings of 퐺 that have just one edge in common ƒ with 퐶. The dimension formula The next result establishes the dimension of the matching space of a matching covered graph, in terms of its number of vertices, edges and bricks (Edmonds, Lov´asz and Pulleyblank [32], see also Naddef (1982, [76])). Theorem 6.13 The dimension of the matching space Lin(퐺) of a matching covered graph 퐺 with 푚 edges, 푛 vertices and 푏 bricks is 푚 − 푛 + 2 − 푏. Proof By induction on the number of edges of 퐺. We have already seen that if 퐺 is bipartite (that is, if 푏 = 0), dim(Lin(퐺)) = 푚 − 푛 + 2, and that if 퐺 is a near-brick (that is, if 푏 = 1), then dim(Lin(퐺)) = 푚 − 푛 + 1. Suppose that 푏 ≥ 2, and that 퐶 is a nontrivial tight cut of 퐺. Let 퐺 1 and 퐺 2 denote the two 퐶-contractions of 퐺 and,

6 The Perfect Matching Polytope

132

for 푖 = 1, 2, let 푚 푖 , 푛푖 , and 푏 푖 denote the numbers of edges, vertices and bricks of 퐺 푖 , respectively. By the induction hypothesis, dim(Lin(퐺 1 )) = 푚 1 − 푛1 + 2 − 푏 1 ,

and

dim(Lin(퐺 2 )) = 푚 2 − 푛2 + 2 − 푏 2 .

Furthermore, by Theorem 6.11 and equation (6.13), dim(Lin(퐺)) = dim(Lin(퐺 1 )) + dim(Lin(퐺 2 )) − |퐶|. We thus have dim(Lin(퐺)) = (푚 1 − 푛1 + 2 − 푏 1 ) + (푚 2 − 푛2 + 2 − 푏 2 ) − |퐶| = (푚 1 + 푚 2 − |퐶|) − (푛1 + 푛2 − 2) + 2 − (푏 1 + 푏 2 ) = 푚 − 푛 + 2 − 푏, where the last equality follows from Theorem 4.17.

ƒ

Since Lin(퐺) is the subspace of R퐸 spanned by { 휒 푀 : 푀 ∈ M}, the set of incidence vectors of perfect matchings of 퐺, there clearly exist bases of Lin(퐺) which are subsets of { 휒 푀 : 푀 ∈ M}. Thus, if dim(Lin(퐺)) = 푟, there exists an 푟 × 푚 matrix of rank 푟 whose rows are incidence vectors of perfect matchings of 퐺. We refer to any such matrix M as a basis matrix of Lin(퐺).

The problem of finding a basis for Lin(퐺) from among the incidence vectors of perfect matchings of 퐺 is related to the theory of ear decompositions. In case of bipartite graphs, this is straightforward. In fact, when 퐺 is bipartite, using the method of proof of Corollary 3.16, one can find a basis for Lin(퐺) such that any integer vector in Lin(퐺) may be expressed as an integer linear combination of vectors in that basis (see Exercise 6.3.3). However, dealing with nonbipartite graphs is much harder and will require several tools which are yet to be developed.

6.3.2 Matching lattice and the seminal work of Lov´asz

The perfect matching integer cone of a graph 퐺, (or simply the matching integer cone), denoted by Int Cone(퐺), is the set of all non-negative integer linear combinations of incidence vectors of perfect matchings of 퐺. It is easy to observe that, for any integer 푟 ≥ 1, an 푟-graph 퐺 is 푟-edge-colourable if and only if the vector 1 of all 1’s is in Int Cone(퐺). Thus, in particular, the 4-colour theorem is equivalent to the statement that if 퐺 is any planar 3-graph then 1 ∈ Int Cone(퐺). This equivalent formulation of the the 4-colour theorem does not make it any easier to prove it, but it allows us to consider relaxations of this problem, and more generally, of the problem of deciding which 푟-graphs are 푟-edge-colourable. We have already come across one such relaxation; Exercise 6.1.10 shows that if 퐺 is any 푟-graph then 1 ∈ Cone(퐺). In an article with many novel ideas, Seymour (1979,[85]) considered a more challenging relaxation which he referred to as ‘subtractive edge-colourings’. This led to the idea of the matching lattice described below.

6.3 The Dimension Formula

133

The perfect matching lattice The perfect matching lattice (or simply the matching lattice) of a matching covered graph 퐺, denoted by Lat(퐺), is the set of all integer linear combinations of the incidence vectors of perfect matchings of 퐺. Clearly, Lat(퐺) is a subset of Lin(퐺) ∩ Z퐸 , for any graph 퐺. Thus, every vector in Lat(퐺) is a regular integer vector. It follows from the observation made in the previous subsection that when 퐺 is bipartite, every regular integer vector in Z퐸 is in Lat(퐺). However, in general, not every regular integer vector in Z퐸 is in Lat(퐺). For example, if 퐺 is the Petersen graph, then 1 is a 3-regular integer vector but it cannot be expressed as an integer linear combination of incidence vectors of perfect matchings of 퐺 (see Exercise 6.3.5). One of the many interesting results proved by Seymour in the paper cited above implies that if 퐺 is any planar 3-graph then 1 ∈ Lat(퐺). In what may be regarded as the watershed paper of this entire subject, Lov´asz (1987, [58]) established a characterization of the matching lattice of a matching covered graph. In particular, he proved: ´ Lovasz’s Theorem Theorem 6.14 If 퐺 is any brick whose underlying simple graph is not the Petersen graph, then Lat(퐺) = Lin(퐺) ∩ Z퐸 . We shall present a proof of this remarkable result in Chapter 16. Many of the ideas introduced in Section 6.3.1 will play a crucial role there. For example, the proof of Theorem 6.11 can be adapted to show that if 퐶 is a tight cut of 퐺, and B1 and B2 are ‘bases’ of Lat(퐺 1 ) and Lat(퐺 2 ), respectively, then B is a ‘basis’ of Lat(퐺).

Exercises 6.3.1 Give a proof of the identity (6.18). ⊲6.3.2 (Merger of bases of separating cut-contractions) The graph 퐺 shown in Figure 6.8 is a brick, and the cut 퐶 is a separating cut in 퐺. One of the 퐶-contractions, 퐺 1 , is the pentagonal prism with one edge duplicated, and the other 퐶-contraction, 퐺 2 , is the brick 푆8 (shown in Figure 7.2). The dimensions of the matching spaces of 퐺 1 and 퐺 2 are 7 and 6. (i) Find a basis B1 of Lin(퐺 1 ) and a basis B2 of 퐺 2 . (Such bases can be found by ‘extending’ the bases of the pentagonal prism and 퐾3,3 ⊙ 퐾4 ). (ii) Determine the merger B := B1 ∨ B2 . (iii) Show that B does not constitute a basis of Lin(퐺). (iv) Extend B to a basis of Lin(퐺). (This would require adding to B one perfect matching which has more than one edge in the cut 퐶.)

134

6 The Perfect Matching Polytope 푋

퐶 Fig. 6.8 Merger across a separating cut

∗ 6.3.3 Let 퐺 be a bipartite matching covered graph. Using the method of the proof of Corollary 3.16 show that there exist perfect matchings 푀1 , 푀2 , . . . , 푀푟 in 퐺, where 푟 = 푚 − 푛 + 2, such that every integer vector in Lin(퐺) may be expressed as an integer linear combination of { 휒 푀1 , 휒 푀2 , . . . , 휒 푀푟 }. ⊲6.3.4 Give a proof of Theorem 6.12. 6.3.5 Show that the vector 1 cannot be expressed as an integer linear combination of incidence vectors of perfect matchings of the Petersen graph. (Hint: Observe that the set M of the six perfect Í matchings of the Petersen graph is a basis for its matching space, and that 1 = 푀 ∈ M 12 휒 푀 .) 6.3.6 Using Theorem 6.13 and Exercise 5.3.4, show that the number of perfect matchings in a cubic matching covered graph 퐺 is at least 푛/4 + 2.

⊲6.3.7 (A basis for the matching space of a complete graph) Let 퐺 be the complete graph on the vertex set {ℎ, 푣 1 , 푣 2 , . . . , 푣 2푟 −1 }, 푟 ≥ 2. (i) Show that 퐺 can be expressed as the union of 푟 − 1 spanning subgraphs 퐺 1 , 퐺 2 , . . . , 퐺 푟 −1 where, for 1 ≤ 푖 ≤ 푟 − 1, the graph 퐺 푖 is a wheel with the vertex ℎ as its hub and a cycle of length 2푟 −1 on the vertex set {푣 1 , 푣 2 , . . . , 푣 2푟 −1 } as its rim. (Thus, for 1 ≤ 푖 < 푗 ≤ 푟 − 1, 퐸 (퐺 푖 ) ∩ 퐸 (퐺 푗 ) consists of the 2푟 − 1 edges incident with ℎ, but the rims of 퐺 푖 and 퐺 푗 are edge-disjoint (2푟 − 1)-cycles on {푣 1 , 푣 2 , . . . , 푣 2푟 −1 }.) (ii) For 1 ≤ 푖 ≤ 푟 − 1, let B푖 denote the set of all perfect matchings of the wheel 퐺 푖 , and let B := B1 ∪ B2 ∪ · · · ∪ B푟 −1 . Show that the set of the incidence vectors of the perfect matchings in B constitutes a basis for the matching space of 퐺. ⊲6.3.8 Consider the brick in Figure 6.9. Show that the number of perfect matchings in this brick is precisely the same as the dimension of its matching space.

6.5 Notes

135

Fig. 6.9 The brick of Exercise 6.3.8.

6.4 Matching Spaces over Finite Fields ♯ The matching space Lin(퐺) associated with a matching covered graph 퐺, as defined in the previous section, is the vector space over the field R of real numbers generated by the set of incidence vectors of perfect matchings of 퐺. Clearly, one may define matching spaces over any field F, where one regards the incidence vector 휒 푀 of a perfect matching 푀 as a (0, 1)-vector with its entries in F. We denote by Lin(퐺, F) the linear space over F generated by the set of incidence vectors of perfect matchings of 퐺, and refer to it as the matching space of 퐺 over F. It is a routine exercise to show that Theorem 6.11 extends to matching spaces over arbitrary fields (Exercise 6.4.1). However, the dimension formula (Theorem 6.13) does not extend to fields of characteristic 2. Once again, it is the Petersen graph P that is the exception. As we have seen, over the reals, the dimension of Lin(P) is six. However, the dimension of Lin(P, Z2 ), where Z2 is the field of integers modulo 2, is only five (Exercise 6.4.2). If 퐺 is any brick other than a Petersen brick, we shall see in a Chapter 16 that the dimension of Lin(퐺, Z2 ) is 푚 − 푛 + 1. The techniques for proving this result turn out to be the same as the one we use to prove Theorem 6.14!

Exercises ⊲6.4.1 Prove that Theorem 6.11 extends to matching spaces over arbitrary fields. 6.4.2 Show that the dimension of Lin(P, Z2 ) is equal to five. (Hint: Use the fact that each edge of the Petersen graph is in precisely two of its perfect matchings, and that any two of its perfect matchings have exactly one edge in common.)

6.5 Notes A good reference for useful proof techniques and interesting unsolved problems related to the perfect matching polytope and edge colourings of graphs is the paper by Seymour (1979, [85]). In that paper Seymour asked if it is possible to use the four colour theorem to show that every planar 4-graph is 4-edge-colourable. This leads

136

6 The Perfect Matching Polytope

to the more general question: For each 푟 ≥ 4, does the four colour theorem imply that every planar 푟-graph is 푟-edge-colourable? Guenin proved that this is the case for 푟 = 4, 5 (2003, [39]). Dvoˇra´ k et al. proved the case 푟 = 6 (2016, [29]). The cases 푟 = 7, 8 were proved by Chudnovsky et al. (2016, [22, 23]). In that same paper [85], Seymour proposed several conjectures. We state two of them now. 6.15 (Conjecture 3.3 in [85]) Any 푟-graph is (푟 + 1)-edge-colourable. 6.16 (Conjecture 3.5 in [85]) If 푟 ≥ 4 then every 푟-graph has a perfect matching whose deletion gives an (푟 − 1)-graph. To the best of our knowledge, Conjecture 6.15 is open. Conjecture 6.16 implies Conjecture 6.15, because every 3-graph has a perfect matching and every graph 퐺 having Δ(퐺) = 2 has a 3-edge colouring. However, Conjecture 6.16 would imply that if 푟 ≥ 4 then every 푟-graph has two disjoint perfect matchings. The latter is false: for each 푟 ≥ 4 there exist 푟-graphs which do not have two disjoint perfect matchings, as proved by Rizzi (1999, [81]). The number of perfect matchings of a graph 퐺 is denoted by Φ(퐺). Suppose that 퐺 is a 2-connected cubic graph. The fact that every such graph 퐺 is matching covered (Theorem 2.8) implies that Φ(퐺) ≥ 3. Since, clearly, Φ(퐺) ≥ dim(Lin(퐺)), the dimension formula (Theorem 6.13) implies a lower bound for Φ(퐺) that is a linear function of 푛 (see Exercise 6.3.6). CLM (2005, [13]) defined a matching covered graph 퐺 to be extremal if Φ(퐺) = dim(Lin(퐺)). Figure 6.9 depicts an example of an extremal brick. It turned out that the only extremal bipartite cubic graph is the 휃-graph on three edges. All other extremal cubic graphs are bricks, and they were able to show that there are precisely ten extremal cubic bricks. One of them is the Petersen graph. Of the remaining nine, the complete graph 퐾4 , the triangular prism 퐶6 , the bicorn H8 , and the tricorn H10 are the ones of order four, six, eight and ten. The largest extremal cubic graph is the cubic brick of order sixteen shown in Figure 6.10.

Fig. 6.10 Extremal cubic brick of order 16

6.5 Notes

137

Some years later, confirming a conjecture of Lov´asz and Plummer, Esperet, Kardoˇs, King, Kr´al’ and Norine (2011, [33]) showed that any 2-connected cubic graph on 푛 vertices has at least 2푛/3656 perfect matchings. The work on the number of perfect matchings in bipartite matching covered graphs [17], mentioned in the Notes section of Chapter 2, led CLM to the following conjecture: The number of perfect matchings in a brace Conjecture 6.17 There exists a positive integer 푁 such that Φ(퐺) ≥

(푛 − 2) 2 , 4

for any brace 퐺 of order 푛 ≥ 푁. The biwheel on 푛 vertices has (푛 − 2) 2 /4 perfect matchings. Using a Theorem of McCuaig [68], which will be described in Part II, CLM were able to show in [17] that any brace on 푛 vertices has at least (푛 − 2) 2 /8 perfect matchings.

An odd wheel of order 푛 has 푛 − 1 perfect matchings. It may be true that there exists a positive integer 푁 such that Φ(퐺) ≥ 푛 − 1, for any brick 퐺 on 푛 ≥ 푁 vertices!

Chapter 7

Solid Bricks

Contents 7.1

7.2

7.3 7.4

Solid Matching Covered Graphs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139 7.1.1 Odd-intercyclic bricks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140 7.1.2 Separating cut decompositions . . . . . . . . . . . . . . . . . . . . . . . 142 The Perfect Matching Polytope and Solid Matching Covered Graphs144 7.2.1 Polytopes Poly푡 (퐺) and Poly∗ (퐺) . . . . . . . . . . . . . . . . . . . . 144 7.2.2 Precedence relation on cuts . . . . . . . . . . . . . . . . . . . . . . . . . . 145 A Characterization of Solid Bricks . . . . . . . . . . . . . . . . . . . . . . . . . . . 147 7.3.1 An infinite family of solid bricks that are not odd-intercyclic149 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153

7.1 Solid Matching Covered Graphs The fact that bipartite graphs are free of odd cycles often makes the task of analyzing their properties easier, and makes it possible to state assertions concerning them more simply, and prove them more easily, than the corresponding assertions that are valid for all graphs. For example, the description of the perfect matching polytope of a bipartite graph is simpler, and its proof easier, than the one that applies to all graphs (see Exercise 6.1.1). Solid matching covered graphs defined below have many properties akin to those of bipartite matching covered graphs. Solid matching covered graphs A matching covered graph is solid if every separating cut in it is a tight cut. By Corollary 4.9, every bipartite matching covered graph is solid. But not all solid matching covered graphs are bipartite. (The smallest solid matching covered graph that is not bipartite is 퐾4 .) © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7_7

139

7 Solid Bricks

140

Solid bricks By definition, bricks are free of nontrivial tight cuts. Thus a brick is solid if it is free of nontrivial separating cuts. In other words, a brick is solid if it cannot be obtained by splicing together two smaller matching covered graphs. (In this sense, solid bricks are ‘unbreakable’!) The brick 퐶6 is not solid because it is equal to 퐾4 ⊙ 퐾4 . According to Corollary 4.16, if 퐷 is a tight cut of a matching covered graph 퐺, and 퐺 has a separating cut 퐶 that is not tight, then one of the 퐷-contractions of 퐺 also has a separating cut which is not tight. It follows that 퐺 is not solid if and only if some member in a tight cut decomposition of 퐺 is not solid. The assertion stated below follows on noting that braces, being bipartite, are solid (see Exercise 7.1.1). Theorem 7.1 A matching covered graph is solid if and only if each of its bricks is solid. ƒ Thus, the problem of recognizing whether or not a given matching covered graph is solid reduces to the problem of recognizing whether or not a given brick is solid.

7.1.1 Odd-intercyclic bricks By Corollary 4.10, if a separating cut in a matching covered graph is not tight, then both shores of that cut induce subgraphs which are not bipartite. This observation leads to the following definition. Odd-intercyclic graphs A graph is said to be odd-intercyclic if any two odd cycles in it have at least one vertex in common. It follows from Corollary 4.10 that every odd-intercyclic brick is solid. Odd wheels, which are easily seen to be odd-intercyclic bricks, will play a special role in future chapters. Example 7.2 (Odd-intercyclic projective planar graphs) Let 퐻 be a 2-connected planar bipartite graph and let 푄 := 푣 1 푣 2 . . . 푣 2푘 푣 1 be a facial cycle of 퐻. Obtain 퐺 from 퐻 by joining, for 1 ≤ 푖 ≤ 푘, the vertices 푣 푖 and 푣 푖+푘 by a new edge which we shall refer to as a chord of the cycle 푄. For 1 ≤ 푖 < 푗 ≤ 푘, the two chords 푣 푖 푣 푖+푘 and 푣 푗 푣 푗+푘 cross in the sense that 푄 traverses the ends of the these chords in the cyclic order 푣 푖 푣 푗 푣 푖+푘 푣 푗+푘 . (See Figure 7.1 for an illustration.) Each graph 퐺 obtained in this way has an embedding in the projective plane so that all faces are even. (When 퐻 is just an even cycle of length 2푘, the resulting graph is the M¨obius ladder M2푘 .) We leave it to the reader as an exercise to show that all graphs that can be obtained by this construction are odd-intercyclic (Exercise 7.1.5).

141

7.1 Solid Matching Covered Graphs

푣8

푣1

푣7

푣2 푣3

푣6

푣5

푣4

퐻

퐺

Fig. 7.1 A projective planar odd-intercyclic graph

Subject to some connectivity conditions, every odd-intercyclic graph turns out to be a projective planar graph of the type described in Example 7.2. This had been known for a long time to many prominent mathematicians, but no proof was ever published. Eventually, a paper by Kawarabayashi and Ozeki [45], which appeared in 2013, gave an elegant proof of this result. They showed that if 퐺 is any internally 4-connected graph, then 퐺 has no two vertex-disjoint odd cycles if and only if 퐺 satisfies one of the following: (i) 퐺 −푣 is bipartite for some 푣 ∈ 푉, (ii) 퐺 −{푒 1 , 푒 2 , 푒 3 } is bipartite for some three edges which constitute the edges of a triangle of 퐺, (iii) |푉 | ≤ 5, or (iv) 퐺 can be embedded in the projective plane so that each face boundary has even length. The above-mentioned result leads to a polynomial-time algorithm for recognizing solid bricks which are odd-intercyclic bricks. But, alas, not all solid bricks are oddintercyclic! The brick depicted in Figure 7.2 is a member of an infinite family of non-odd-intercyclic solid bricks which will be described in the next section.

Fig. 7.2 Σ8 , a solid brick of order eight

7 Solid Bricks

142

The problem of recognizing whether a given brick is solid is clearly in the complexity class co-NP; if a brick happens to be nonsolid, to certify that it is nonsolid it suffices to display a nontrivial separating cut in it. However, no polynomial-time algorithm for recognizing solid bricks is yet known. It is not even known if the related decision problem is in NP! Unsolved problem: recognizing solid bricks Problem 7.3 Is the problem of recognizing whether a given brick is solid in the complexity class NP? Is it in P?

7.1.2 Separating cut decompositions Analogous to a tight cut decomposition of a matching covered graph, one may define a separating cut decomposition by considering contractions with respect to nontrivial separating cuts, instead of the more restrictive tight cuts. Using such a decomposition, any matching covered graph may be reduced to a list of graphs which are free of nontrivial separating cuts. Since any separating cut in a bipartite graph is also tight, such a list would consist of braces and solid bricks. However, no statement similar to Theorem 4.17 holds for separating cuts: distinct separating cut decompositions may even yield distinct numbers of graphs. Consider, for example, the brick shown in Figure 7.3.

퐶3

퐶4

퐶1

퐶2

Fig. 7.3 A brick which does not admit a unique separating cut decomposition

One separating cut decomposition of this brick, using the three cuts 퐶1 , 퐶2 , and 퐶3 yields a collection of four 퐾4 ’s, up to multiple edges, whereas another separating cut decomposition that uses only the cut 퐶4 produces two 푊5 ’s, up to multiple edges.

7.1 Solid Matching Covered Graphs

143

Do separating cut decompositions of a matching covered graph shed any light on the structure of that graph? There is no known evidence to suggest that they do. But, as we shall see in future chapters, properties of solid bricks do have many uses. It is conceivable that a parameter associated with separating cut decompositions, such as the minimum possible number of solid bricks in a decomposition of a graph, might prove to be relevant in some context!

Exercises ⊲7.1.1 Give a proof of Theorem 7.1. 7.1.2 Show that the brick shown in Figure 7.2 is not odd-intercyclic, but is solid. ⊲7.1.3 Prove that the odd wheel 푊2푘+1 is an odd-intercyclic brick for all 푘 ≥ 1. ⊲7.1.4 Prove that the M¨obius ladder M2푛 (defined in Section 2.5) is an odd-intercyclic brick, for all even 푛 ≥ 2. ⊲7.1.5 Let 퐺 be a graph obtained by the procedure described in Example 7.2. Show that: (i) every odd cycle of 퐺 must include an odd number of chords of 푄, and (ii) if 퐶 and 퐷 are two odd cycles in 퐺, then there exist chords of 푄 belonging to 퐶 and 퐷, respectively, which cross. Now, using the planarity of 퐻, deduce that 퐺 is odd-intercyclic. 7.1.6 Determine an infinite family of odd-intercyclic cubic graphs which are not M¨obius ladders. (Hint: Note that in the construction described in Example 7.2, it is not necessary to add all the 푘 chords of the cycle 푄 to obtain an odd-intercyclic graph. Consider a suitable 2-connected planar bipartite graph 퐻 in which there is a facial cycle 푄 with four vertices of degree two and add two crossing chords of 푄 to obtain a cubic graph. A M¨obius ladder is obtained by adding two edges to a ladder!) 7.1.7 Consider the graph Σ8 ⊙ Σ8 obtained by a splicing two copies of the brick Σ8 (Figure 7.2). Show that it is a brick and show that it has two distinct separating cut decompositions.

Fig. 7.4 Σ8 ⊙ Σ8

7 Solid Bricks

144

7.2 The Perfect Matching Polytope and Solid Matching Covered Graphs By Edmonds’ Theorem 6.1 there are three types of constraints that every vector x in the perfect matching polytope Poly(퐺) of a matching covered graph 퐺 must satisfy. We consider two other polytopes associated with 퐺 which are obtained by relaxing the odd set constraints.

7.2.1 Polytopes Poly풕 (푮) and Poly∗ (푮) The first polytope, which we denote by Poly푡 (퐺), is obtained by restricting the odd set constraints to equalities corresponding to the tight cuts of 퐺, and the second one, which we denote by Poly∗ (퐺), is obtained by dropping all the odd set constraints. Thus Poly푡 (퐺) is the set of solutions to the following system of linear inequalities: x(푒) ≥ 0, for each 푒 ∈ 퐸

x(휕 (푆)) = 1, for each tight cut 휕 (푆) of 퐺,

(7.1a) (7.1b)

and Poly∗ (퐺), which is known as the fractional perfect matching polytope, is the set of solutions to: x(푒) ≥ 0, for each 푒 ∈ 퐸

x(휕 (푣)) = 1, for each vertex 푣 of 퐺.

(7.2a) (7.2b)

For want of a better name, we shall refer to Poly푡 (퐺) as the tight-cut perfect matching polytope. The following assertion is an easy consequence of the above definitions: Proposition 7.4 For any matching covered graph 퐺 Poly(퐺) ⊆ Poly푡 (퐺) ⊆ Poly∗ (퐺).

(7.3)

It is easy to see that for any bipartite graph 퐺, both inclusions in (7.3) hold as equalities. There are examples of nonbipartite matching covered graphs which show that either one or both inclusions in (7.3) may hold strictly (Exercise 7.2.6). We are thus led to the natural questions of characterizing matching covered graphs for which these inclusions hold as equalities. Interestingly, the notion of solid matching covered graphs plays a crucial role in the answers to these questions. We shall see, for example, that Poly(퐺) = Poly푡 (퐺) if and only if 퐺 is solid, and that Poly(퐺) = Poly∗ (퐺) if and only if 퐺 is a solid matching covered graph with at most one brick. We shall first deal with bricks and then return to the general case later on. Thus let 퐺 be a brick. As bricks do not have nontrivial tight cuts it follows that the equality Poly푡 (퐺) = Poly∗ (퐺) holds. Thus, every nonnegative vector in R퐸 which satisfies the degree constraints is in Poly푡 (퐺); it follows that Poly(퐺) = Poly∗ (퐺) if and only if every such vector is also in Poly(퐺). This happens to be the case if and only

7.2 The Perfect Matching Polytope and Solid Matching Covered Graphs

145

if 퐺 is a solid brick. The proof of this surprising fact requires the notion of a certain partial order on the set of odd cuts of a matching covered graph. (This notion will also feature as a crucial tool in the proofs of several important theorems in future chapters.)

7.2.2 Precedence relation on cuts

Precedence relation A cut 퐶 of a matching covered graph 퐺 precedes a cut 퐷 (written 퐶  퐷) if |푀 ∩ 퐶| ≤ |푀 ∩ 퐷|,

for all 푀 ∈ M.

If strict inequality holds for at least one perfect matching 푀, then we say that 퐶 strictly precedes 퐷 and write 퐶 ≺ 퐷. On the other hand, if |푀 ∩ 퐶| = |푀 ∩ 퐷| for all 푀 ∈ M, then 퐶 and 퐷 are matching equivalent.

For example, in the brick 퐺 shown in Figure 7.5, the cut 퐶 strictly precedes the cut 퐷; whereas, in the graph 퐺 − 푒, the two cuts 퐶 and 퐷 are matching equivalent (Exercise 7.2.1). We say that a cut 퐶 is minimal with respect to the relation  if there is no cut that strictly precedes 퐶. 퐷

푒

퐶 Fig. 7.5 퐶 precedes 퐷

Theorem 7.5 The perfect matching polytope Poly(퐺) of a brick 퐺 consists of all non-negative 1-regular vectors if and only if 퐺 is solid. (In other words, the equality Poly(퐺) = Poly∗ (퐺) holds for a brick 퐺 if and only 퐺 is solid.) Proof Firstly suppose that 퐺 is not solid. We wish to show that there is some nonnegative 1-regular vector in R퐸 that does not belong to Poly(퐺). Since 퐺 is nonsolid, we first note that it has a nontrivial separating cut, say 퐶. Let 푀0 be a perfect matching of 퐺 such that |푀0 ∩퐶| > 1. (Such a perfect matching must exist; otherwise 퐶 would

7 Solid Bricks

146

be a tight cut in a brick which is absurd.) Also, since 퐶 is separating, for every edge 푒 of 퐺, there is a perfect matching 푀푒 of 퐺 that contains edge 푒 and just one edge in 퐶, by Theorem 4.2. Now let ! ! Õ 1 푀푒 푀0 x := 휒 − 휒 (7.4) |퐸 | − 1 푒∈퐸 It can be verified that x is non-negative 1-regular with x(퐶) < 1 (Exercise 7.2.2). Conversely, suppose that 퐺 is solid. We wish to prove that every non-negative 1-regular vector in R퐸 belongs to Poly(퐺). Assume to the contrary that there is a nonnegative 1-regular vector x that does not belong to Poly(퐺). Then, by Theorem 6.1 there must exist at least one odd cut 퐶 such that x(퐶) is strictly less than one. Let C denote the set of all odd cuts 퐶 for which x(퐶) < 1 and let 퐷 := 휕 (푌 ) be a cut in C that is minimal with respect to the precedence relation  defined above. We shall show that 퐷 is a separating cut and thereby obtain a contradiction to the hypothesis that 퐺 is solid. Let 퐺 1 := 퐺/(푌 → 푦) and 퐺 2 := 퐺/(푌 → 푦) denote the two 퐷-contractions of 퐺. We wish to show that both 퐺 1 and 퐺 2 are matching covered. Let us first consider 퐺 1 . Clearly it is connected. If it is not matching covered, then it is either not matchable, or is matchable but has an unmatchable edge. In the former case, by Tutte’s Theorem 1.3, there exists a subset 푆 of 푉 (퐺 1 ) such that 표(퐺 1 − 푆) > |푆| (Figure 7.6(a)). In the latter case, by Theorem 2.1, there exists a subset 푆 of 푉 (퐺 1 ) such that 표(퐺 1 − 푆) = |푆|, but there is an edge 푒 of 퐺 1 with both its ends in 푆 (Figure 7.6(b)). As 퐺 is matching covered, on both alternatives the contraction vertex 푦 is in 푆 (see Figure 7.6).

푆

푒

푆

푦 퐷

휕(퐾 )

푦 퐷

휕(퐾 ) (a)

(b)

Fig. 7.6 Finding a separating cut using precedence

In either case, a simple counting argument shows that there must be an odd component 퐾 of 퐺 1 − 푆 for which x(휕 (퐾)) < 1 (Exercise 7.2.3(i)). One may now verify that 휕 (퐾) ≺ 퐷 (Exercise 7.2.3(ii)), contradicting the choice of 퐷. Therefore 퐺 1 is matching covered. Similarly, 퐺 2 is also matching covered; hence 퐷 is a

7.3 A Characterization of Solid Bricks

147

separating cut. As x(퐷) < 1, the cut 퐷 is nontrivial. In sum, 퐷 is a nontrivial separating cut of a brick. Hence 퐷 is not tight, in contradiction to the hypothesis that 퐺 is solid. ƒ The following statements related to the perfect matching polytope of a matching covered graph are easy consequences of the above theorem. We leave their proofs as Exercise 7.2.5. Corollary 7.6 For every matching covered graph 퐺, the following properties hold: (i) Poly(퐺) = Poly푡 (퐺) if and only if 퐺 is solid, (ii) Poly(퐺) = Poly∗ (퐺) if and only if 퐺 is a solid matching covered graph with at ƒ most one brick.

Exercises ⊲7.2.1 Verify that in the brick shown in Figure 7.5 the cut 퐶 strictly precedes the cut 퐷. ⊲7.2.2 Verify that x(퐶) < 1 for the vector x defined in Equation (7.4). ∗ 7.2.3 Consider the graph 퐺 1 := 퐺/푌 defined in the proof of Theorem 7.5, and a subset 푆 of 푉 (퐺 1 ) with the properties specified there. Show that: (i) the graph 퐺 1 − 푆 has an odd component 퐾 such that x(휕 (퐾)) < 1, and (ii) the cut 휕 (퐾) strictly precedes 퐷. 7.2.4 Describe the perfect matching polytope of the Petersen graph using as few odd set constraints as possible. Justify your answer. 7.2.5 Give a proof of Corollary 7.6. 7.2.6 Give examples of matching covered graphs to show that each of the following instances may occur: (i) Poly(퐺) ⊂ Poly푡 (퐺) = Poly∗ (퐺), (ii) Poly(퐺) = Poly푡 (퐺) ⊂ Poly∗ (퐺), (iii) Poly(퐺) ⊂ Poly푡 (퐺) ⊂ Poly∗ (퐺).

7.3 A Characterization of Solid Bricks ♯ The characterization of solid bricks presented in this section does not seem to lead in any obvious way to a polynomial-time algorithm for recognizing solid bricks. However, it provides a convenient means for showing that certain bricks are solid.

7 Solid Bricks

148

It is straightforward to verify that the incidence vector of any perfect matching of any matching covered graph 퐺 is an extreme point of the polytope Poly∗ (퐺) (Exercise 7.3.1). If all the extreme points of Poly∗ (퐺) were incidence vectors of perfect matchings of 퐺, then Poly(퐺) and Poly∗ (퐺) would be the same polytope. Thus if 퐺 is a brick that is not solid, the polytope Poly∗ (퐺) must have an extreme point which is not the incidence vector of a perfect matching. The following result concerning extreme points of Poly∗ (퐺) may be proved by means of arguments analogous to the ones used in our proof of Theorem 6.1 (Exercise 7.3.2). Extreme points of Poly∗ (퐺) Theorem 7.7 Every extreme point of Poly∗ (퐺) is a (0, 21 , 1)-vector. Furthermore, a vector w is an extreme point of Poly∗ (퐺) if and only if the subgraph 퐺 [w] of 퐺 induced by the support of w is a spanning subgraph of 퐺 in which each component is either an odd cycle or has just one edge, and    0 w(푒) = 21  1 

if 푒 is not an edge of 퐺 [w] if 푒 is an edge of an odd cycle in 퐺 [w] if 푒 is the only edge of a component of 퐺 [w].

ƒ

For example, the polytope Poly∗ (퐶6 ) has five extreme points; four of them are the incidence vectors of the four perfect matchings of 퐶6 , and the fifth is the (0, 12 )-valued vector displayed in Figure 6.1(a). With the help of the above theorem we can now present the result mentioned at the beginning of this section. It was communicated to us in 2003 by four friends from the University of S˜ao Paulo: de Pina, Fernandes, Reed and Wakabayashi (2003, [27]). It first appeared in a paper by CLM (2004, [12]), which presented a purely graph theoretic proof. That proof relies on results concerning robust cuts in nonsolid bricks that will be described in a later chapter. Here we shall derive it as a consequence of Theorem 7.7. It may also be deduced from more general results by Balas (1981, [1]).

A characterization of solid bricks Theorem 7.8 A brick 퐺 is nonsolid if and only if it has two disjoint odd cycles 퐶1 and 퐶2 such that 퐺 − (푉 (퐶1 ) ∪ 푉 (퐶2 )) has a perfect matching. Proof By Theorem 7.5, 퐺 is nonsolid if and only if Poly(퐺) is a proper subset of Poly∗ (퐺). As observed earlier this is the case if and only if Poly∗ (퐺) has an extreme point w which is not the incidence vector of a perfect matching of 퐺. It follows from Theorem 7.7 that each component of the subgraph 퐺 w of 퐺 induced by the support of w is either an odd cycle or has just two vertices. Let 퐶1 , 퐶2 , . . . , 퐶 푘 denote the components of 퐺 w which are odd cycles. Since w is not the incidence vector of a perfect matching of 퐺, and since |푉 | is even, it follows that 푘 is greater than zero and

7.3 A Characterization of Solid Bricks

149

that it is even. Since all components of 퐺 w other than 퐶1 , 퐶2 , . . . , 퐶 푘 have just one edge each, it follows that 퐺 − (푉 (퐶1 ) ∪푉 (퐶2 ) ∪ · · · ∪푉 (퐶 푘 )) has a perfect matching. If 푘 = 2, there is nothing more to prove. So, suppose that 푘 ≥ 4. Let 푢 be any vertex of 퐶 푘 , let 푀 푘 be a maximum matching of 퐶 푘 − 푢, and let 퐿 denote a perfect matching of 퐺−(푉 (퐶1 ) ∪푉 (퐶2 ) ∪· · ·∪푉 (퐶 푘 )). Then 푀 = 푀 푘 ∪퐿 is a matching of 퐺 which leaves all vertices of 퐶1 , 퐶2 , . . . , 퐶 푘−1 and just one vertex of 퐶 푘 , namely 푢, exposed. Since 퐺 is matchable, there exists a 푀-augmenting path, say 푃, starting at 푢. Let 푣 denote the other end of 푃 and adjust notation, if necessary, so that 푣 is a vertex of 퐶 푘−1 . Denote by 푀 푘−1 the perfect matching of 퐶 푘−1 − 푣, and let 푀 ′ := 푀 푘−1 ∪ (푀 △ 푃). This is a perfect matching of 퐺 − (푉 (퐶1 ) ∪ 푉 (퐶2 ) ∪ · · · ∪ 푉 (퐶 푘−2 )). See Figure 7.7.

푢

퐶1

퐶3

퐶2

퐶4

푣

Fig. 7.7 Proof of Theorem 7.8. The edges of 푀 are indicated by thick lines. The edges of 퐸 ( 푃 ) − 푀 and those of the perfect matching 푀3 of 퐶3 − 푣, indicated by dotted lines, constitute the perfect matching 푀 ′ of 퐺 − (푉 (퐶1 ) ∪ 푉 (퐶2 ) ).

By applying this procedure (푘 − 2)/2 times we can reduce the list of 푘 odd cycles to just two odd cycles 퐶1 and 퐶2 such that 퐺 − (푉 (퐶1 ) ∪ 푉 (퐶2 )) has a perfect matching. ƒ

7.3.1 An infinite family of solid bricks that are not odd-intercyclic The graph shown in Figure 7.2 is the smallest solid brick that is not odd-intercyclic. A description of an infinite family of such bricks, which includes the one in Figure 7.2 can be found in the article by Lucchesi, Carvalho, Kothari and Murty (2018, [64]). At the time of writing that paper, the authors were unaware of any examples of cubic solid bricks with disjoint odd cycles. In a recent manuscript, Chen, Feng, Lu and Zhang described an infinite family of non-odd-intercyclic cubic solid bricks (2019, [21]). The graph shown in Figure 5.9 is the first member of that infinite family. Using the idea of edge-extensions, defined in Exercise 5.3.5, we are now able to present a simpler description of another infinite family of non-odd-intercyclic cubic solid bricks. Each graph in our family is obtained from a bipartite prism by means of two edge-extensions.

7 Solid Bricks

150

Let 푘 ≥ 3 be any integer and let P4푘 be the prism obtained from two disjoint even cycles 푢 1 푢 2 · · · 푢 2푘 푢 1 and 푣 1 푣 2 · · · 푣 2푘 푣 1 by the addition of the 2푘 edges 푢 푖 푣 푖 , 푖 = 1, 2, . . . , 2푘. Now obtain 퐺 4푘 from 푃4푘 by deleting the two edges 푢 3 푣 3 and 푢 2푘 푣 2푘 , and then adding an edge joining 푢 3 to 푣 2푘 and an edge joining 푣 3 to 푢 2푘 . Figure 7.8 depicts the graph 퐺 20 obtained from the prism P20 . (The graph shown in Figure 5.9 is isomorphic to 퐺 12 .) 푢1

푢10

푢2

푣1

푢3

푣2 푣3

푣10 푣9

푢9

푢4

푣4

푣8

푣5 푣7

푢8

푣6

푢7

푢5

푢6

Fig. 7.8 퐺20 : a cubic solid brick with disjoint odd cycles

Lemma 7.9 For each 푘 ≥ 3, the graph 퐺 4푘 obtained from the prism P4푘 as described above is a brick with disjoint odd cycles. Proof Let us first note that 퐺 4푘 is cubic and that it is not bipartite. (The 5-cycle 푢 1 푣 1 푣 2푘 푢 3 푢 2 푢 1 and the (2푘 − 1)-cycle 푢 2푘 푣 3 푣 4 · · · 푣 2푘−1 푢 2푘−1 푢 2푘 are two disjoint odd cycles in 퐺 4푘 .) Since 퐺 4푘 is cubic and nonbipartite, to conclude that it is a brick it suffices to show, by Corollary 5.11, that 퐺 4푘 is essentially 4-edge-connected. This is indeed the case. Consider the prism 퐻 obtained from the two (2푘 − 2)-cycles 푢 1 푢 2 푢 4 푢 5 · · · 푢 2푘−1 푢 1 and 푣 1 푣 2 푣 4 푣 5 · · · 푣 2푘−1 푣 1 by adding the 2푘 −2 edges 푢 푖 푣 푖 , for 푖 = 1, 2, 4, 5, . . . , 2푘 −1. Clearly, 퐻 is isomorphic to the prism P4푘−4 . (It is important to note that neither 푢 3 nor 푣 3 are vertices of 퐻.) It can be seen that the graph 퐺 4푘 is the graph obtained from 퐻 by two edge-extensions; one based on the pair {푢 2 푢 4 , 푣 2푘−1 푣 1 }, and the other based on pair {푣 2 푣 4 , 푢 2푘−1 푢 1 }. As 푘 ≥ 3, the prism P4푘−4 is a (cubic) brace; hence 퐻 is a cubic brace. By Corollary 5.12, 퐻 is essentially 4-edge connected. As 퐺 4푘 is obtained from 퐻 by two edge extensions, it is essentially 4-edge connected (Exercise 5.3.5(i)). Indeed, 퐺 4푘 is a brick. ƒ Theorem 7.10 For each 푘 ≥ 3, the graph 퐺 := 퐺 4푘 obtained from the prism P4푘 is a cubic solid brick with disjoint odd cycles.

7.3 A Characterization of Solid Bricks

151

Proof It follows from Lemma 7.9 that 퐺 is a brick with disjoint odd cycles. To show that 퐺 is solid we use Theorem 7.8. Thus let 퐶1 and 퐶2 be any two disjoint odd cycles in 퐺. Note that the graph 퐺 − {푢 3 푣 2푘 , 푣 3 푢 2푘 } is bipartite. Since 퐶1 and 퐶2 are disjoint odd cycles, it follows that one of these two cycles must pass through 푢 3 푣 2푘 and the other through 푣 3 푢 2푘 . The set of the vertices of the 4-cycle 푄 := 푢 1 푢 2 푣 2 푣 1 푢 1 is a 4-vertex cut of 퐺. Furthermore, it is easy to see that the graph 퐺 − 푉 (푄) is isomorphic to a bisubdivision of the M¨obius ladder M4푘−8 (Exercise 7.3.4). Since M4푘−8 is odd-intercyclic, not both 퐶1 and 퐶2 , which are disjoint odd cycles, can be confined to the graph 퐺 − 푉 (푄). Without loss of generality, we may assume that the cycle 퐶1 passes through the edge 푢 3 푣 2푘 and that it has edges in common with the 4-cycle 푄. Since 퐶1 , being disjoint from 퐶2 , cannot pass through either 푣 3 or 푢 2푘 , it follows that 퐶1 is either 푢 3 푣 2푘 푣 1 푢 1 푢 2 푢 3 or 푢 3 푣 2푘 푣 1 푣 2 푢 2 푢 3 . Thus, in the graph 퐺 − 푉 (퐶1 ∪ 퐶2 ), one of the two vertices 푢 1 and 푣 2 is an isolated vertex, implying that 퐺 − 푉 (퐶1 ∪ 퐶2 ) is not matchable. Hence, by Theorem 7.8, 퐺 is solid. ƒ

Exercises ⊲7.3.1 For any perfect matching 푀 of a graph 퐺, show that 휒 푀 is an extreme point of Poly∗ (퐺). ⊲7.3.2 Let 퐺 be any graph, let w be any extreme point of its fractional perfect matching polytope Poly∗ (퐺), and let 퐺 [w] be the subgraph of 퐺 induced by the support of w. Give a proof of Theorem 7.7 along the following lines. (i) The graph 퐺 [w] is free of even cycles (see proof of statement 6.2.1). (ii) Each block of 퐺 [w] is either an odd cycle or is 퐾2 . (iii) If a component of 퐺 [w] has a vertex 푣 of degree one, then the edge incident with 푣 is the only edge of that component (implying that if a component of 퐺 [w] has more than one edge, then all end blocks of that component are odd cycles). (iv) Every component of 퐺 [w] is either 퐾2 or an odd cycle. (Suppose that, to the contrary, there is a component of 퐺 [w] with more than one end block. Such a component will contain two odd cycles, say 푄 1 and 푄 2 , and a path 푃 which connects a vertex 푢 of 푄 1 with a vertex 푣 of 푄 2 , but which is internally-disjoint from both 푄 1 and 푄 4 . Now obtain a contradiction to the assumption that w is an extreme point of Poly∗ (퐺) by showing that there exist two vectors x and y in Poly∗ (퐺) such w = 21 x + 21 y. It may be convenient to consider two cases depending on the parity of the length of 푃.) (v) Now show that w is (0, 12 , 1)-valued and is as described in the statement of Theorem 7.7. ∗ 7.3.3 Describe all the extreme points of the fractional perfect matching polytope Poly∗ (P) of the Petersen graph. ⊲7.3.4 In the proof of Theorem 7.10 it was claimed that the graph 퐺 − 푉 (푄), where 푄 = 푢 1 푢 2 푣 2 푣 1 푢 1 , is a bisubdivision of the M¨obius ladder M4푘−4 . Justify this claim.

7 Solid Bricks

152

∗ 7.3.5 Show that the brick 퐺 depicted in Figure 7.9 does not have two disjoint odd cycles 퐶1 and 퐶2 such that 퐺 − (푉 (퐶1 ) ∪푉 (퐶2 )) has a perfect matching, and deduce that it is solid. (This is one of an infinite family of solid bricks, discovered by Nishad Kothari, that are not odd-intercyclic which will be described in a later chapter.) 푓

0

1

ℎ1 푣1

6

2 푒

ℎ2

푣2 5

ℎ3

3 4

Fig. 7.9 The solid brick 퐻12

7.3.6 All the members of the family {퐺 4푘 : 푘 ≥ 3} described in Section 7.3.1 are near-bipartite, that is, have a pair of edges whose deletion results in a bipartite matching covered graph. The graph shown in Figure 7.10 was found by Nishad Kothari by exhaustive computer search. It is not near-bipartite. Show that it is solid.

Fig. 7.10 A cubic solid brick of order 14

7.4 Notes

153

7.4 Notes CLM discovered solid matching covered graphs in connection with their work on a conjecture of Lov´asz, and have since found them to have many interesting and useful properties (see (2002, [9]), (2005, [13]) and (2012, [16])).

The most puzzling unsolved problem concerning solid matching covered graphs is of course Problem 7.3. But there are several other intriguing questions related to them. Let 푓 (2푛) denote the maximum number of edges a simple solid matching covered graph on 2푛 vertices can have. It is easy to see that 푓 (4) = 6, and not too difficult to show that 푓 (6) = 11. A simple bipartite graph of order 2푛 has at most 푛2 edges, which is the number of edges of 퐾푛.푛 . It seems highly plausible that the following conjecture is true: Conjecture 7.11 There exists a positive integer 푁 such that 푓 (2푛) = 푛2 , for all 푛 ≥ 푁. We do not know any solid bricks which are ‘dense’. In particular, we do not know any solid bricks whose connectivity is greater than three. Problem 7.12 Do there exist solid bricks which are 4-connected? Apart from the fact that there are many intriguing problems concerning solid matching covered graphs, their properties will prove to be very useful in inductive proofs of several important results of the subject. In particular, solid matching covered graphs play a fundamental role in Chapter 15, where it is proved Lov´asz’s conjecture, on the existence of 푏-invariant edges in every brick distinct from 퐾4 , 퐶6 and the Petersen graph (an edge 푒 of a matching covered graph 퐺 is 푏-invariant if 퐺 − 푒 is matching covered and 푏(퐺 − 푒) = 푏(퐺)).

Chapter 8

Dependence Relation and Removable Classes

Contents 8.1

8.2 8.3

8.4 8.5

Deletions of Edges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155 8.1.1 Removable edges in bipartite graphs . . . . . . . . . . . . . . . . . . . 157 8.1.2 Near-bipartite graphs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160 8.1.3 Removable sets and separating cuts . . . . . . . . . . . . . . . . . . . 161 The Dependence Relation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164 8.2.1 Dependence classes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167 Removable Classes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168 8.3.1 Minimal classes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168 8.3.2 Sizes of removable classes . . . . . . . . . . . . . . . . . . . . . . . . . . . 170 Two Lemmas for Future Use . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176 8.4.1 The Zigzag Lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181

8.1 Deletions of Edges Deletions and contractions of edges are two common inductive tools in graph theory. For example, the famous theorem of Kuratowski states that a graph is nonplanar if and only if it can be reduced to either 퐾5 or to 퐾3,3 by means of deletions and contractions of edges. There are several theorems of similar flavour in the theory of matching covered graphs. But here the deletion and contraction operations used are, of necessity, more restrictive as the deletion of an arbitrary edge from a matching covered graph need not result in a matching covered graph, and the contraction of a single edge does not even preserve the parity of the number of vertices. Recall that the contraction of the shore 푋 of a separating cut 휕 ( 푋) of a matching covered graph 퐺 amounts to contracting each edge of 퐺 with both its ends in 푋 and then deleting all the resulting loops. The graph 퐺/푋 thus obtained is also matching covered. We have effectively used tight-cut contractions as inductive tools for establishing properties of matching covered graphs such as the formula for the © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7_8

155

156

8 Dependence Relation and Removable Classes

dimension of Lin(퐺) (Theorem 6.13). Separating cut-contractions and tight-cut contractions of a very special type, known as ‘bicontractions’, will play prominent roles in future chapters. In this chapter and the next two, our focus will be on deletions of edges. Removable edges An edge 푒 of a matching covered graph 퐺 is removable if the graph 퐺 − 푒 is also matching covered. For example, every edge of 퐾3,3 is removable, but no edge of 퐾4 or of 퐶6 is. In the context of matching covered graphs, by the deletion of an edge, we always mean the deletion of a removable edge. More generally, a nonempty (proper) subset 푅 of edges is removable if 퐺 − 푅 is matching covered. Bipartite matching covered graphs have the following special property: Theorem 8.1 In a bipartite matching covered graph, every minimal removable set of edges is a singleton. Proof Let 퐺 := 퐺 [ 퐴, 퐵] be a bipartite matching covered graph with bipartition ( 퐴, 퐵). Let 푅 be a removable set of edges of 퐺. Suppose that |푅| ≥ 2, and let 푒 := 푎푏, with 푎 ∈ 퐴 and 푏 ∈ 퐵, be an edge in the set 푅. By hypothesis 퐺 ′ := 퐺 − 푅 is a bipartite matching covered graph, and 푎 and 푏 are two vertices which belong to different parts of the bipartition of 퐺 ′ . It follows from Corollary 3.7 (with 퐺 ′ playing the role of 퐺) that 퐺 ′ + 푒 is matching covered. This implies that 푅 is not minimal because 퐺 ′ + 푒 = (퐺 − 푅) + 푒 = 퐺 − (푅 − 푒). We conclude that every minimal removable set of edges of 퐺 is a singleton. ƒ A nonbipartite graph 퐺 may contain removable pairs of edges of the following type: Removable doubletons A pair {푒, 푓 } of edges of a matching covered graph 퐺 is a removable doubleton if 퐺 − 푒 − 푓 is matching covered, but neither 퐺 − 푒 nor 퐺 − 푓 is. For example, each of 퐾4 and 퐶6 have three removable doubletons, and the bicorn H8 has two. In the above definition, the fact that neither 퐺 − 푒 nor 퐺 − 푓 is matching covered implies that every perfect matching of 퐺 which contains one of the two edges 푒 and 푓 also contains the other. In the next section we shall introduce an equivalence relation on the edge set of a matching covered graph that is suggested by the above observation, and a partial order on the set of equivalence classes with respect to this relation. Using the properties of these equivalence classes, we shall establish the striking fact (Corollary 8.18) that if a matching covered graph 퐺 has a removable set 푅 of edges, with the proviso that no proper subset of 푅 is removable, then 푅 is either a removable edge or is a removable

157

8.1 Deletions of Edges

doubleton of 퐺. In other words, every minimal removable set of a matching covered graph contains just one or two edges. It should be noted that, in general, a matching covered graph 퐺 need not have any removable sets of edges at all. For example, if each edge of 퐺 is incident with a vertex of degree two and |퐸 | > 2, then the deletion of any nonempty subset of 퐸 (퐺) results in a graph which is not matching covered. However, it follows from the theory of ear decompositions to be described in Chapter 11 that every matching covered graph of minimum degree at least three has either a removable edge or a removable doubleton (Corollary 11.9). (In case of bipartite graphs, this is just the statement of Exercise 3.4.1.)

8.1.1 Removable edges in bipartite graphs Let 퐺 := 퐺 [ 퐴, 퐵] be a bipartite matching covered graph. Suppose that there exist partitions ( 퐴0 , 퐴1 ) of 퐴 and (퐵0 , 퐵1 ) of 퐵 into nonempty subsets such that | 퐴0 | = |퐵0 |, | 퐴1 | = |퐵1 |, and 푎푏, with 푎 ∈ 퐴0 and 푏 ∈ 퐵1 , is the only edge of 퐺 with one end in 퐴0 and one end in 퐵1 . (See Figure 8.1 for an illustration.) 퐵0

퐵1

푣

푏

푎 퐴0

푢 퐴1

Fig. 8.1 A nonremovable edge in a bipartite graph

Clearly, 퐺 has four or more vertices and hence is 2-connected, by Corollary 2.7. Thus, 푎푏 cannot be the only edge in the cut 휕 ( 퐴0 ∪ 퐵0 ). So, there must be at least one edge, say 푢푣, which has one end in 퐴1 and one end in 퐵0 . It is easy to see that every perfect matching of 퐺 which contains the edge 푢푣 also contains 푎푏. (In the terminology to be introduced in the next section, the edge 푢푣 ‘depends’ on the edge 푎푏.) Thus, the deletion of 푎푏 would result in a graph in which 푢푣 would be unmatchable, implying that 푎푏 is not removable. The following lemma establishes necessary and sufficient conditions under which an edge in a bipartite matching covered graph is not removable.

158

8 Dependence Relation and Removable Classes

Nonremovable edges in bipartite matching covered graphs Lemma 8.2 Let 퐺 := 퐺 [ 퐴, 퐵] be a bipartite matching covered graph with |퐸 | ≥ 2. An edge 푎푏 of 퐺, with 푎 ∈ 퐴 and 푏 ∈ 퐵, is not removable in 퐺 if and only if there exist nonempty proper subsets 퐴1 and 퐵1 of 퐴 and 퐵, respectively, such that: (i) the subgraph 퐺 [ 퐴1 ∪ 퐵1 ] induced by 퐴1 ∪ 퐵1 is matching covered, and (ii) 푎 ∈ 퐴 − 퐴1 and 푏 ∈ 퐵1 , and 푎푏 is the only edge joining a vertex in 퐴 − 퐴1 to a vertex in 퐵1 . We shall refer to the cut 휕 ( 퐴1 ∪ 퐵1 ) as a certificate for the nonremovability of the edge 푎푏. Proof Suppose first that there exist nonempty proper subsets 퐴1 and 퐵1 of 퐴 and 퐵, respectively, satisfying the required properties. Then, as explained above, there must exist an edge, say 푢푣, with 푢 ∈ 퐴1 and 푣 ∈ 퐵 − 퐵1 , and any perfect matching of 퐺 which contains edge 푢푣 must also contain the edge 푎푏, implying that 푎푏 is not removable in 퐺. Suppose now that 푎푏 is not removable. Then, by definition, the subgraph 퐻 := 퐺 − 푎푏 is not matching covered. But it is matchable because the hypothesis that |퐸 | ≥ 2 implies that the vertex 푎 has degree at least two, and any perfect matching of 퐺 which does not contain the edge 푎푏 is also a perfect matching of 퐻. Therefore, by Theorem 2.12, there exists a nonempty proper subset 푇 of 퐵 containing 푏 such that the subgraph of 퐻 induced by 푁 퐻 (푇) ∪ 푇 is matching covered. Since 퐺 itself is matching covered, it must be the case that 푎푏 has one end in 퐴 − 푁 퐻 (푇) and one end in 푇, for, otherwise, |푁퐺 (푇)| = |푇 | which is impossible because 퐺 is matching covered (Theorem 2.9). Thus, if we let 퐴1 := 푁 퐻 (푇) and 퐵1 := 푇, we have the two subsets 퐴1 and 퐵1 of 퐴 and 퐵, respectively, satisfying both the two required conditions. ƒ As a consequence of the above results, we have the following special property of braces. Removable edges in braces Theorem 8.3 Every edge in a brace of order at least six is removable. Proof Let 퐺 [ 퐴, 퐵] be a brace of order at least six. If 퐺 has a nonremovable edge 푎푏 then it follows from Lemma 8.2 that there exist subsets 퐴1 and 퐵1 of 퐴 and 퐵, respectively, satisfying the two conditions specified in its statement. If we set 퐴0 := 퐴−퐴1 and 퐵0 := 퐵−퐵1 , it is easy to see that both 휕 ( 퐴0 ∪(퐵0 +푏)) and 휕 (( 퐴1 +푎)∪퐵1 ) are tight cuts of 퐺 and that at least one of them is nontrivial (Exercise 8.1.6). But this contradicts the hypothesis that 퐺 is a brace. ƒ In an unpublished paper, Lov´asz and Vempala [61] proved the following useful generalization of Lemma 8.2 which provides a way of simultaneously certifying the nonremovability of each of the edges in a specified subset of 휕 (푎).

159

8.1 Deletions of Edges

Nonremovable edges in bipartite matching covered graphs Theorem 8.4 Let 퐺 [ 퐴, 퐵] be a matching covered bipartite graph and let 푎 be a vertex in 퐴 with 푏 1 , 푏 2 , . . . , 푏 푑 , 푑 ≥ 3, as its neighbours. Suppose that the edges 푎푏 1 , 푎푏 2 , . . . , 푎푏푟 , 푟 ≤ 푑, incident with the vertex 푎, are not removable in 퐺. Then there exist partitions ( 퐴0 , 퐴1 , . . . , 퐴푟 ) of 퐴 and (퐵0 , 퐵1 , . . . , 퐵푟 ) of 퐵 such that, for 1 ≤ 푖 ≤ 푟, (i) the subgraph 퐺 [ 퐴푖 ∪ 퐵푖 ] induced by 퐴푖 ∪ 퐵푖 is matching covered, and (ii) 푎 ∈ 퐴0 , 푏 푖 ∈ 퐵푖 and 푎푏 푖 is the only edge that joins a vertex of 퐴 − 퐴푖 to a vertex in 퐵푖 . Consequently, | 퐴푖 | = |퐵푖 |, for 푖 = 0, 1, 2, . . . , 푟 and 푏푟+1 , 푏푟+2 , . . . , 푏 푑 ∈ 퐵0 . (Figure 8.2 illustrates the case in which 푟 = 2 and 푑 = 3.)

퐵1

퐵0

푏1

퐵2 푏3

푏2

푎 퐴1

퐴0

퐴2

Fig. 8.2 Illustration for the case 푟 = 2 and 푑 = 3 (Theorem 8.4)

Proof The edge 푎푏 푖 is not removable in 퐺, for 1 ≤ 푖 ≤ 푟. Therefore, by Lemma 8.2, there are nonempty sets 퐴푖 ⊂ 퐴 and 퐵푖 ⊂ 퐵 satisfying the two properties described in the statement of the theorem, with 푏 푖 , 퐴푖 and 퐵푖 playing respectively the roles of 푏, 퐴1 and 퐵1 . So, for any index 푖, where 1 ≤ 푖 ≤ 푟, the subgraph 퐺 [ 퐴푖 ∪ 퐵푖 ] is matching covered, 푎 ∈ 퐴 − 퐴푖 , 푏 푖 ∈ 퐵푖 and 푎푏 푖 is the only edge that joins a vertex of 퐴 − 퐴푖 to a vertex in 퐵푖 . Let 퐴0 := 퐴 −

푟 Ø 푖=1

퐴푖

and

퐵0 := 퐵 −

푟 Ø

퐵푖 .

푖=1

The vertex 푎 does not belong to any of the sets 퐴1 , 퐴2 , . . . , 퐴푟 . Thus, 푎 ∈ 퐴0 . Of the neighbours of the vertex 푎, for 1 ≤ 푖 ≤ 푟, the vertex 푏 푖 is the only vertex of 퐵푖 adjacent to 푎; hence it belongs only to 퐵푖 (and not to 퐵 푗 for 푗 ≠ 푖). To establish the desired conclusion, it thus suffices to show that the 퐵푖 , as well as the 퐴푖 , are disjoint. Consider, say, 퐵1 and 퐵2 , and let 푆 := 퐵1 ∩ 퐵2 . The vertex 푏 1 is in 퐵1 − 퐵2 ; hence 푁 (푆) ⊆ 퐴1 . Likewise, 푁 (푆) ⊆ 퐴2 . Thus, 푁 (푆) ⊆ 퐴1 ∩ 퐴2 .

8 Dependence Relation and Removable Classes

160

As 푑 ≥ 3, the graph 퐺 has a perfect matching, 푀, that contains the edge 푎푏 3 . As 푎푏 1 is the only edge that joins a vertex in 퐴 − 퐴1 to a vertex in 퐵1 , it follows that the vertices of 퐵1 are matched by 푀 to vertices of 퐴1 . Likewise, the vertices of 퐵2 are matched by 푀 with vertices of 퐴2 . The vertices of 퐵1 ∪ 퐵2 are matched by 푀 with vertices of 퐴1 ∪ 퐴2 . Thus, |퐵1 ∪ 퐵2 | ≤ | 퐴1 ∪ 퐴2 |. Likewise, every vertex of 푆 is matched by 푀 with a vertex of 퐴1 ∩ 퐴2 . Thus, |푆| ≤ | 퐴1 ∩ 퐴2 |. As 퐺 [ 퐴1 ∪ 퐵1 ] is matching covered, we deduce that |퐵1 | = | 퐴1 |. Likewise, |퐵2 | = | 퐴2 |. In sum, |퐵1 | |퐵2 | |퐵1 ∪ 퐵2 | |푆|

= = ≤ ≤

| 퐴1 |, | 퐴2 |, | 퐴1 ∪ 퐴2 | and | 퐴1 ∩ 퐴2 |.

It follows that |퐵1 | + |퐵2 | = |퐵1 ∪ 퐵2 | + |푆| ≤ | 퐴1 ∪ 퐴2 | + | 퐴1 ∩ 퐴2 | = | 퐴1 | + | 퐴2 | = |퐵1 | + |퐵2 |; hence equality holds throughout. In particular, |푆| = | 퐴1 ∩ 퐴2 |. We have seen that 푁 (푆) ⊆ 퐴1 ∩ 퐴2 ; hence |푁 (푆)| = |푆|. As 퐺 is matching covered, it follows that 푆 is either 퐵 or the empty set (Theorem 2.9). As 푏 1 is not in 푆, we conclude that 푆 is empty; hence so too is 퐴1 ∩ 퐴2 . ƒ We deduce that 퐵1 and 퐵2 are disjoint, as are 퐴1 and 퐴2 .

8.1.2 Near-bipartite graphs The deletion of a removable edge from a nonbipartite matching covered graph never results in a bipartite matching covered graph (Exercise 8.1.3), but the deletion of a removable doubleton may. Near-bipartite graphs A matching covered graph 퐺 is near-bipartite if it has a removable doubleton 푅 such that the subgraph 퐺 − 푅 obtained by the deletion of 푅 is a bipartite (matching covered) graph. Many prominent classes of bricks, such as prisms of order 2 (mod 4) and M¨obius ladders of order 0 (mod 4), are near-bipartite. Proposition 8.5 Let 퐺 be a near-bipartite matching covered graph and let 푅 be a removable doubleton of 퐺 such that 퐺 − 푅 is bipartite. Let ( 퐴, 퐵) denote the bipartition of 퐺 − 푅. Then, one of the edges of 푅 has both ends in 퐴 and the other edge of 푅 has both ends in 퐵. Proof Let 푒 and 푓 denote the two edges of 푅. Assume that one of the edges in 푅, say 푒, has one end in 퐴 and one end in 퐵. As 퐺 − 푒 − 푓 is a matching covered graph with bipartition ( 퐴, 퐵), and 푒 is an edge that has one end in 퐴 and one end in 퐵, it follows

8.1 Deletions of Edges

161

from Corollary 3.7 that (퐺 − 푒 − 푓 ) + 푒 = 퐺 − 푓 is matching covered, implying that 푓 is removable in 퐺. This is contrary to the definition of a removable doubleton. We deduce that 푒 has both ends in the same part of the bipartition of 퐺 − 푒 − 푓 , say 퐴. Likewise, edge 푓 has both its ends in the same part of 퐺 − 푒 − 푓 . It cannot have both ends also in 퐴 because, in that case, both 푒 and 푓 would be unmatchable in 퐺. ƒ We deduce that 푓 has both ends in 퐵. By an ingenious use of elementary linear algebra, Lov´asz (1987, [58]) showed that every brick with a removable doubleton is near-bipartite; see Exercise 8.1.7. We shall present an alternative proof of this result in Section 9.1.1 in a more general context. But, it should be noted that not every matching covered graph with a removable doubleton is near-bipartite. For example, the two edges 26 and 59 constitute a removable doubleton in the graph shown in Figure 8.4, but that graph is not nearbipartite. We conclude this section with a result which will be found to be useful in Section 9.3. Its proof makes use of Theorem 8.4. Lemma 8.6 Let 퐺 be a near-bipartite brick, where 푅 is a removable doubleton of 퐺 such that 퐻 := 퐺 − 푅 is bipartite. Let 푣 be a vertex of 퐺 that has degree three or more in 퐻. Then, at most two edges incident with 푣 are not removable in 퐻. Proof Let 푒 and 푓 denote the two edges of 푅, and let ( 퐴, 퐵) denote the bipartition of 퐻. Adjust notation so that 푒 has both ends in 퐴 and 푓 has both ends in 퐵. Let 푎 denote a vertex in 퐴 having degree 푑 ≥ 3 in 퐻. Let 푎푏 1 , 푎푏 2 , . . . , 푎푏푟 , 푟 ≤ 푑, denote 푟 edges of 퐻 incident with 푎 that are not removable in 퐻. Let ( 퐴0 , 퐴1 , . . . , 퐴푟 ) and (퐵0 , 퐵1 , . . . , 퐵푟 ) be partitions of 퐴 and 퐵, respectively, as in the statement of Theorem 8.4. For 푖 = 1, 2, . . . , 푟, the set 퐴푖 ∪ {푎} is a nontrivial barrier of 퐻, where the | 퐴푖 | vertices of 퐵푖 are isolated in 퐺 − 퐴푖 − 푎 (See Figure 8.2). By hypothesis, 퐺 is bicritical; therefore edge 푓 must have an end in the set 퐵푖 . (Otherwise, 퐴푖 ∪ {푎} would be a nontrivial barrier of 퐺 itself which is absurd because 퐺 is bicritical.) This conclusion holds for each 푖 such that 1 ≤ 푖 ≤ 푟. Since 푓 has just two ends, it ƒ follows that 푟 ≤ 2. Making use of their Theorem 8.4, and the results in Exercises 8.1.8 and 3.2.2, Lov´asz and Vempala [61] proved that every near-bipartite brick other than 퐾4 , 퐶6 and the bicorn H8 has at least two removable edges. In fact, every simple near-bipartite brick other than the three bricks mentioned above has at least two nonadjacent removable edges. We shall present a proof of this interesting result in the next chapter.

8.1.3 Removable sets and separating cuts The following useful assertion, showing a connection between removable sets of edges and separating cuts, is easy to verify.

162

8 Dependence Relation and Removable Classes

Proposition 8.7 Let 퐺 be a matching covered graph, let 퐶 be a separating cut of 퐺, let 퐺 1 and 퐺 2 be the two 퐶-contractions of 퐺 and let 푅 be a set of edges of 퐺. If 퐺 푖 − 푅 is matching covered, for 푖 = 1, 2, then 퐺 − 푅 is also matching covered and ƒ 퐶 − 푅 is a separating cut of 퐺 − 푅. The converse of the above assertion does not, in general, hold. For instance, consider the cut 퐶 of the bicorn H8 shown in Figure 8.3. The edge 푒 is removable in H8 , but it is not removable in 퐶6 , which is one of its 퐶-contractions. 퐶 푒

Fig. 8.3 Edge 푒 is removable in H8 but not in one of its 퐶-contraction

If 퐶 is a tight cut of 퐺, and 푅 is a removable set of edges of 퐺, then 퐶 − 푅 is a tight cut of 퐺 − 푅, because every perfect matching of 퐺 − 푅 is also a perfect matching of 퐺. This observation implies that the converse of Proposition 8.7 does hold when 퐶 is a tight cut. Thus: Proposition 8.8 Let 퐺 be a matching covered graph, let 퐶 be a tight cut of 퐺, let 퐺 1 and 퐺 2 be the two 퐶-contractions of 퐺 and let 푅 be a set of edges of 퐺. The graph 퐺 − 푅 is matching covered if and only if, for 푖 = 1, 2, the graph 퐺 푖 − 푅 is matching covered. ƒ Let 퐶 be a separating cut of a matching covered graph 퐺. We shall adopt the convention that if an edge 푒 does not belong to the edge set of a 퐶-contraction 퐻 of 퐺, then 푒 is removable in 퐻. Corollary 8.9 Let 퐺 be a matching covered graph, and let 퐶 be a separating cut of 퐺. If an edge 푒 is removable in both 퐶-contractions of 퐺 then 퐶 − 푒 is a separating cut of 퐺 − 푒, and 푒 is removable in 퐺. ƒ

Exercises 8.1.1 Find all the removable edges and doubletons in the graph shown in Figure 8.4. 8.1.2 Find all the removable edges and doubletons in each of the families of matching covered graphs described in Section 2.5.

8.1 Deletions of Edges

163 1 2

9

3

8

푥

4

7 5

6

Fig. 8.4 Graph obtained by deleting an edge from the Petersen graph

⊲8.1.3 Let 퐺 be a matching covered graph and let 푒 be a removable edge of 퐺. Show that 퐺 is bipartite if and only if 퐺 − 푒 is bipartite. (Compare with Lemma 4.6 in Chapter 4.) ⊲8.1.4 Prove Propositions 8.7 and 8.8. 8.1.5 Let 퐺 [ 퐴, 퐵] be a bipartite matching covered graph on four or more vertices, and let 푥 be a vertex of 퐺. Show that either there is an edge of 퐺, not incident with 푥, which is removable in 퐺; or there is a subset 푌 of 푉 − 푥, |푌 | ≥ 3, such that 휕 (푌 ) is a tight cut of 퐺. ⊲8.1.6 Supply the missing details in the proof of Theorem 8.3. ∗ 8.1.7 (Graphs with a removable doubleton, Lov´asz [58]) Let 푒 and 푓 be two edges of a matching covered graph 퐺 such that 퐺 − 푒 − 푓 is matching covered, but neither 퐺 − 푒 nor 퐺 − 푓 is. Show that: (i) every perfect matching of 퐺 which contains one of 푒 and 푓 also contains the other, (ii) deduce that if x = 휒 푒 − 휒 푓 , where 휒 푒 and 휒 푓 are the incidence vectors of the singleton sets {푒} and { 푓 }, respectively, then x ◦ 휒 푀 = 0, for any perfect matching 푀 of 퐺, or, in other words, x is matching orthogonal (this is just a translation of the statement in the first item), (iii) now, using the result in Exercise 6.2.6, show that if 퐺 is a brick, then 퐺 − 푒 − 푓 is bipartite. An alternative proof of (iii) in a more general context will be presented in Section 9.1. ⊲8.1.8 Let 퐻 [ 퐴, 퐵] be a bipartite matching covered graph, and let 푣 be a vertex of degree three or more in 퐻. Suppose that 퐻 has a 4-cycle 푄 that contains vertex 푣. Using Theorem 8.4, show that some edge of 휕 (푣) ∩ 퐸 (푄) is removable in 퐻. ⊲8.1.9 (The Exchange Property) We say that a matching covered graph 퐺 has the exchange property with respect to the deletion of edges if 푒 is any removable edge of 퐺, and if 푓 is any removable edge of 퐺 − 푒, then the edge 푓 is removable in 퐺 itself (and 푒 is removable in 퐺 − 푓 ).

8 Dependence Relation and Removable Classes

164

(i) Show that all bipartite matching covered graphs have the exchange property. (We shall show in Chapter 10 that all solid matching covered graphs have the exchange property.) (ii) Find two suitable edges 푒 and 푓 of the tricorn H10 (Figure 2.5(d)) to show that the tricorn does not have the exchange property. ∗ 8.1.10 Let 퐺 be a brace of order at least six, and let 푒 := 푎푏 be any edge of 퐺 such that 퐺 − 푒 is not a brace. (i) Let 퐶 := 휕 ( 푋) and 퐷 := 휕 (푌 ) be nontrivial cuts of 퐺 such that the cuts 퐶 −푒 and 퐷 − 푒 are tight in 퐺 − 푒 and 푎 ∈ 푋 ∩푌 . Let 퐼 := 휕 ( 푋 ∩푌 ) and let 푈 := 휕 ( 푋 ∩푌 ). Show that 퐼 − 푒 and 푈 − 푒 are both nontrivial tight cuts in 퐺 − 푒. (ii) Show that there exists a nested family 푋1 ⊂ 푋2 ⊂ · · · ⊂ 푋 푘 of subsets of 푉, each containing 푎, but not 푏, such that  (퐺 − 푒)/푋1 is a brace,    (퐺 − 푒)/푋푖 /푋푖+1 is a brace for 2 ≤ 푖 < 푘,   (퐺 − 푒)/푋 푘 is a brace. 

(8.1)

(iii) In the above decomposition, show that 푋1 and 푋 푘 are uniquely determined and deduce that if 퐺 − 푒 has at most three braces, then 퐺 − 푒 has a unique tight cut decomposition, that is, 퐺 − 푒 has a unique maximal laminar family of nontrivial tight cuts. (CLM (2015, [18]))

8.2 The Dependence Relation In this section we introduce an equivalence relation on the edge set of a matching covered graph 퐺 which was first defined and used in a paper by Carvalho and Lucchesi (1996, [7]) and later studied more extensively in a paper by CLM (1999, [8]). It turns out that every minimal nonempty removable set of edges of 퐺 is an equivalence class with respect to this relation. Let 퐺 be a matching covered graph, and let 푒 and 푓 be any two edges of 퐺. Then we say that 푒 depends on 푓 , or 푒 implies 푓 , if every perfect matching that contains 푒 also contains 푓 . Equivalently, 푒 depends on 푓 if and only if 푒 is not matchable in 퐺 − 푓 . (In the graph depicted in Figure 8.1, the edge 푢푣 depends on the edge 푎푏. In fact, every edge between 퐴1 and 퐵0 depends on the edge 푎푏.) We write 푒 ⇒ 푓 to indicate that 푒 depends on 푓 . Clearly, ⇒ is a reflexive and transitive relation on the edge set 퐸 (퐺). Two edges 푒 and 푓 are mutually dependent if 푒 ⇒ 푓 and 푓 ⇒ 푒. In this case we write 푒 ⇔ 푓 . Clearly mutual dependence is an equivalence relation on 퐸 (퐺). We refer to the equivalence classes with respect to this relation as dependence classes . It is convenient to visualize the relation ⇒ by means of the digraph it defines on the set of edges of 퐺, called the dependence digraph, and denoted by 퐷 (퐺). Figure 8.5 depicts the dependence digraph 퐷 (H8 ) of the bicorn H8 , which is a brick of order eight. For every edge 푒 of H8 distinct from the edge 27 there is an edge distinct

8.2 The Dependence Relation

165

from 푒 that depends on 푒. In other words, if 푒 is any edge other than 27, then some edge of H8 − 푒 is unmatchable in that subgraph. But H8 − 27 has no unmatchable edges. Thus 27 is the only removable edge of H8 . 1

2

6

3

8 7

5

4 (a) H8

67

23

12

78

45 38

16

15

48

34 (b) 퐷 (H8 )

56

27

Fig. 8.5 The bicorn H8 and the associated dependence digraph 퐷 (H8 )

As already noted, an edge 푒 depends on another edge 푓 if and only if 푒 is not matchable in 퐺 − 푓 . Thus, by Corollary 2.2: Proposition 8.10 An edge 푒 of a matching covered graph 퐺 depends on another edge 푓 of 퐺 if and only if 퐺 − 푓 has a barrier 퐵 that contains both ends of 푒. ƒ Proposition 8.11 Let 푒 and 푓 be distinct edges of a matching covered graph 퐺 such that 퐺 − 푓 has a barrier 퐵 that contains both ends of 푒. Then, the edge 푓 has its ends in distinct odd components of 퐺 − 푓 − 퐵. Proof The graph 퐺 is matching covered, hence 푒 is matchable in 퐺. As 푒 has both ends in 퐵, then 퐵 is not a barrier of 퐺, hence 푓 has its ends in distinct components of 퐺 − 푓 − 퐵. If 푓 has an end, 푣, in an even component of 퐺 − 푓 − 퐵 then 퐵 + 푣 is a barrier of 퐺 that contains both ends of 푒. This is not possible, because 푒 is matchable in 퐺. We deduce that 푓 has its ends in distinct odd components of 퐺 − 푓 − 퐵. ƒ The next assertion is rather technical, but will play a fundamental role in several proofs in the book.

8 Dependence Relation and Removable Classes

166

Lemma 8.12 Let 푓 be an edge of a matching covered graph 퐺, let 퐵 denote a barrier of 퐺 − 푓 such that 푓 has its ends in distinct odd components of 퐺 − 푓 − 퐵. Let 퐸 퐵 denote the set of edges of 퐺 that have both ends in 퐵 and let 퐶 be the set of edges that join a vertex of 퐵 to a vertex of some even component of 퐺 − 푓 − 퐵, as illustrated in Figure 8.6. For each perfect matching 푀 of 퐺, 2|푀 ∩ 퐸 퐵 | + |푀 ∩ 퐶| ≤ 2|푀 ∩ { 푓 }|,

(8.2)

with equality if and only if |푀 ∩ 휕 (퐾)| = 1, for each odd component 퐾 of 퐺 − 푓 − 퐵.

푒1

푒2

퐶

퐵

y

푓 E

O

Fig. 8.6 Illustration for Lemma 8.12, where 퐸퐵 = {푒1 , 푒2 }, and O and E denote respectively the collection of odd and even components of 퐺 − 푓 − 퐵

Proof (of Lemma 8.12) Let O denote the collection of odd components of the graph 퐺 − 푓 − 퐵 and let 퐹 denote the set of edges that join a vertex of 퐵 to a vertex of some odd component of 퐺 − 푓 − 퐵. Then, Õ |퐵| ≤ |푀 ∩ 휕 (퐾)|, (8.3) Õ

퐾 ∈ O

퐾 ∈ O

|푀 ∩ 휕 (퐾)| = 2|푀 ∩ { 푓 }| + |푀 ∩ 퐹|

|푀 ∩ 퐹| + |푀 ∩ 퐶| = |푀 ∩ 휕 (퐵)|

|푀 ∩ 휕 (퐵)| + 2|푀 ∩ 퐸 퐵 | = |퐵|.

Note that equality holds in (8.3) if and only if |푀 ∩ 휕 (퐾)| = 1 for each 퐾 ∈ O. Addition of the inequalities and simplification yields the asserted inequality (8.2). Moreover, equality holds in (8.2) if and only if |푀 ∩ 휕 (퐾)| = 1 for each 퐾 ∈ O. ƒ

8.2 The Dependence Relation

167

8.2.1 Dependence classes The following theorem describes properties of a matching covered graph which are implied by the existence of two mutually dependent edges. Theorem 8.13 Let 푒 and 푓 be two distinct mutually dependent edges of a matching covered graph 퐺, and let 퐵 be a barrier of 퐺 − 푓 that contains both ends of 푒. Then, the following properties hold: (i) edge 푓 has its ends in distinct odd components of 퐺 − 푓 − 퐵, (ii) for each component 퐾 of 퐺 − 푓 − 퐵, the cut 휕 (퐾) is tight in 퐺, (iii) edge 푒 is the only edge of 퐺 that has both ends in 퐵, and (iv) every component of 퐺 − 푓 − 퐵 is odd. Proof Let O denote the collection of odd components of 퐺 − 푓 − 퐵. The validity of part (i) follows immediately from Proposition 8.11. To prove the other parts, let 푀 be a perfect matching of 퐺. By hypothesis, the edges 푒 and 푓 are mutually dependent, hence |푀 ∩ {푒}| = |푀 ∩ { 푓 }|. Let 퐸 퐵 and 퐶 be as in the statement of Lemma 8.12. Then, 2|푀 ∩ {푒}| + 2|푀 ∩ (퐸 퐵 − 푒)| + |푀 ∩ 퐶| ≤ 2|푀 ∩ { 푓 }| = 2|푀 ∩ {푒}|,

(8.4)

with equality if and only if |푀 ∩ 휕 (퐾)| = 1, for each component 퐾 ∈ O. Equality holds throughout in (8.4). Thus, 2|푀 ∩ (퐸 퐵 − 푒)| + |푀 ∩ 퐶| = 0 and |푀 ∩ 휕 (퐾)| = 1, for each 퐾 ∈ O. These equalities hold for each perfect matching 푀 of 퐺. Thus, 휕 (퐾) is tight, for each 퐾 ∈ O. This proves part (ii). As 퐺 is matching covered, every edge of 퐺 is matchable. We deduce that 퐸 퐵 − 푒 and 퐶 are both empty. By hypothesis, 푒 ∈ 퐸 퐵 , thus 퐸 퐵 = {푒}. This proves part (iii). As 퐺 is matching covered, it is connected. As 퐶 is empty, we conclude that 퐺 − 푓 − 퐵 ƒ does not have any even components. This proves part (iv).

Exercises 8.2.1 Draw the dependence digraphs of 퐾4 , 퐶6 , 퐾3,3 , and the graph shown in Figure 8.4 which is obtained by deleting an edge from the Petersen graph. Even and odd 2-edge cuts ⊲8.2.2 Let 푒 and 푓 be two edges of a matching covered graph 퐺 such that {푒, 푓 } is a cut of 퐺. Show that 푒 and 푓 depend on each other if and only if {푒, 푓 } is an even cut of 퐺. (Hint: Use the parity condition stated in Exercise 1.2.1.)

8 Dependence Relation and Removable Classes

168

⊲8.2.3 Let 퐺 1 and 퐺 2 denote the two 퐶-contractions with respect to a tight cut 퐶 of a matching covered graph 퐺. Show that: (i) for 푖 = 1, 2, the dependence relation in 퐺 푖 is the restriction to 퐸 (퐺 푖 ) of the dependence relation in 퐺, and (ii) if 푄 is any equivalence class in 퐺, then |푄 ∩ 퐶| ≤ 1. 8.2.4 Let 푢, 푣, and 푤 be the vertices of a triangle in a matching covered graph 퐺, and suppose that the degree of 푢 in 퐺 is three and that 푢푣, 푢푤, and 푢푢 ′ , where 푢 ′ ≠ 푣, 푤, are the three edges incident with 푢 (see Figure 8.7). (i) Show that the edge 푣푤 depends on the edge 푢푢 ′ . (ii) Deduce that 푢푢 ′ is not removable in 퐺. 푣 푢

푢′

푤 Fig. 8.7 Edge 푣푤 depends on the edge 푢푢′

8.2.5 Consider the wheel 푊2푘+1 , where 푘 ≥ 2. (i) Show that every spoke of 푊2푘+1 is removable. (ii) Using Exercise 8.2.4, show that no edge in the rim of 푊2푘+1 is removable.

8.3 Removable Classes 8.3.1 Minimal classes By identifying the vertices in the equivalence classes in the dependence digraph, we b (퐺). The reduced dependence digraph of obtain the reduced dependence digraph 퐷 the bicorn H8 shown in Figure 8.5 is the digraph shown in Figure 8.8. b (퐺) is acyclic. The sources in 퐷 b (퐺) are called minimal dependence Clearly 퐷 b (퐺) classes, or simply minimal classes1. For an edge 푒 of 퐺, any source 푄 of 퐷 that contains an edge 푓 of 퐺 that depends on 푒 is said to be a minimal dependence b (H8 ) shown in Figure 8.8, the minimal class class induced by 푒. For example, in 퐷 induced by the edges 67, 34 and 45 are, respectively, {15, 48}, {34, 56} and {27}. 1 Minimality here is with respect to the partial order ⇒ and it should not be confused with minimality with respect to set containment.

8.3 Removable Classes

169

{67}

{23}

{12}

{78}

{45} {38}

{16}

{15, 48}

{27}

{34, 56}

b (H8 ) corresponding to the graph H8 in Figure 8.5 Fig. 8.8 The digraph 퐷

Proposition 8.14 Every edge of a matching covered graph induces at least one minimal dependence class. ƒ In general, an edge may induce several minimal classes. For example, the graph shown in Figure 8.9 illustrates this phenomenon. 0

푒1

1

2

3

4

푒2 푒3 5

6

7

8

9

Fig. 8.9 The set 푄 := {푒1 , 푒2 , 푒3 } is a dependence class and 35 and 36 are removable edges. Any edge in 푄 induces both {35} and {36}.

Removable classes An equivalence class 푅 of the dependence relation of a matching covered graph 퐺 is a removable class if the graph 퐺 − 푅 is matching covered. Clearly, every removable class is a minimal class. By the definition of the dependence relation, a minimal class 푄 of a matching covered graph 퐺 is a minimal nonempty subset of 퐸 such that every edge in 퐸 − 푄 is matchable in 퐺 − 푄. Thus, if 퐺 − 푄 happens to be connected, then it is also matching covered; in this case, 푄 is a removable class. For example, if 퐺 has at least two edges and |푄| = 1, then 푄 is a removable class of 퐺 because any matching covered graph 퐺 with at least two edges is 2-connected. The bicorn H8 shown in Figure 8.5 has three minimal classes, and each of them is also a removable class. Although, in general, every removable class is minimal, a minimal class need not always be a removable class. As an example, consider the cycle 퐶2푘 , where 푘 ≥ 2. It has precisely two minimal classes, namely its perfect matchings, but neither of them is a removable class. As another example, consider

170

8 Dependence Relation and Removable Classes

the graph shown in Figure 8.9. In this graph {푒 1 , 푒 2 , 푒 3 } is a minimal class, but it is not removable! When a removable class 푅 of a matching covered graph 퐺 consists of just one edge, that edge is a removable edge of 퐺, and when 푅 consists of two edges, then those two edges form a removable doubleton of 퐺. The bicorn H8 in Figure 8.5 has one removable edge and two removable doubletons.

8.3.2 Sizes of removable classes All removable classes in a bipartite matching covered graph are singletons (Theorem 8.1). But, as we have seen, nonbipartite matching covered graphs may have removable classes which have more than one edge. However, as we shall demonstrate in this section, all removable classes in any matching covered graph are either removable edges or removable doubletons. This result (Corollary 8.18) plays an important role in the theory of ear decompositions of matching covered graphs (Chapter 11). The above-mentioned fundamental fact concerning sizes of removable classes may be easily deduced from Lemma 5.4.5 in Lov´asz and Plummer [59]. But the statement of that lemma in [59] and the proof they give involve concepts and techniques that are not included in this book. Here we shall deduce that result by first establishing, entirely within the framework of this chapter, a property of all minimal classes of size three or more. Our proof of this more general result (Theorem 8.17) is inductive and depends crucially on Theorem 8.13 and the following lemma. Lemma 8.15 Let 퐺 1 and 퐺 2 denote the two 퐶-contractions with respect to a tight cut 퐶 of a matching covered graph 퐺, and let 푄 be a minimal class in 퐺. Then: (i) for 푖 = 1, 2, the set 푄 ∩ 퐸 (퐺 푖 ) is either empty or a minimal class in 퐺 푖 , (ii) |푄 ∩ 퐶| ≤ 1, and (iii) if 푄 and 퐶 are disjoint then 푄 is entirely contained in one of 퐸 (퐺 1 ) − 퐶 and 퐸 (퐺 2 ) − 퐶. Proof The first two parts of the assertion are covered by Exercise 8.2.3. Turning to the last part, suppose that 푄 and 퐶 are disjoint. If possible, let 푒 1 and 푒 2 be two edges of 푄 such that 푒 1 ∈ 퐸 (퐺 1 ) − 퐶 and 푒 2 ∈ 퐸 (퐺 2 ) − 퐶. Let 푀1 be a perfect matching of 퐺 containing 푒 1 , and let 푓 be the edge of 푀1 in 퐶. As 푄 and 퐶 are disjoint, 푓 is not in 푄. Thus, as 푄 is a minimal class, edge 푓 is matchable in 퐺 − 푄. If 푀 푓 is any perfect matching of 퐺 in 퐺 − 푄 that contains the edge 푓 , the matching (푀1 ∩ 퐸 (퐺 1 )) ∪ (푀 푓 ∩ 퐸 (퐺 2 )) would be a perfect matching of 퐺 that contains 푒 1 but not 푒 2 . This would imply that 푒 1 and 푒 2 are not equivalent, which is absurd because 푒 1 and 푒 2 are both edges of the minimal class 푄. ƒ We have noted that minimal classes of matching covered graphs can be arbitrarily large. This is the case even when the graph under consideration has no vertices of degree two. For example, in the graph on 16 vertices shown in Figure 8.10, the set of the eight edges indicated by dotted lines is a minimal class.

8.3 Removable Classes

171

Fig. 8.10 A minimal class with eight edges (indicated by dotted lines)

Any two edges in a matching covered graph which constitute an even 2-cut are equivalent; see Exercise 8.2.2. (The cut shown in Figure 8.10 is an even 2-cut.) The following result indicates a very important property of minimal classes that are doubletons. Proposition 8.16 Let 푄 be a minimal class of a matching covered graph 퐺. If |푄| = 2 then either 푄 is removable in 퐺 or 푄 is an even 2-cut of 퐺. Proof Suppose that |푄| = 2. As 푄 is minimal, every edge of the graph 퐺 − 푄 is matchable in 퐺 −푄. If 퐺 −푄 is connected then the graph 퐺 −푄 is matching covered; hence 푄 is a removable class. Alternatively, suppose that 퐺 − 푄 is not connected. Then, 푄 includes some nonempty cut, 퐶, of 퐺. As |푄| = 2, the (matching covered) graph 퐺 has two or more edges; hence it is 2-connected. Thus, |퐶| ≥ 2. We deduce that 퐶 = 푄. That is, 푄 is a 2-cut of 퐺. It is now easy to conclude that 푄 is an even cut of 퐺. See Exercise 8.2.2. ƒ Let 푄 be a minimal class of a matching covered graph 퐺 with at least two edges. We have seen that if |푄| = 1 then 푄 is a removable class of 퐺, and that if |푄| = 2 then either 푄 is a removable class or it is an even 2-cut of 퐺. We now turn our attention to the case in which |푄| ≥ 3. Minimal classes and even 2-cuts Theorem 8.17 Any minimal class 푄 of a matching covered graph 퐺, with |푄| ≥ 3, includes an even 2-cut of 퐺. Proof Suppose that |푄| ≥ 3. We shall prove, by induction on |푉 |, that 푄 includes an even 2-cut. Towards this end, let 푒 1 and 푒 2 be two edges which belong to 푄. As this means that 푒 1 and 푒 2 are mutually dependent, Proposition 8.10 implies that there exists a barrier 퐵 of 퐺 − 푒 2 which contains both ends of 푒 1 . By Theorem 8.13, the only edge of 퐺 with both ends in 퐵 is 푒 1 , for each component 퐾 of 퐺 − 푒 2 − 퐵, the

8 Dependence Relation and Removable Classes

172

cut 휕 (퐾) is tight in 퐺, and the edge 푒 2 has its ends in two different components of 퐺 − 푒 2 − 퐵. We first of all consider the case in which all components of 퐺 − 푒 2 − 퐵 are trivial. It turns out that it is possible to deal with this case without recourse to induction. Case 1 Every component of 퐺 − 푒 2 − 퐵 is trivial. In this case, the graph 퐻 := 퐺 − 푒 1 − 푒 2 is bipartite with bipartition ( 퐴, 퐵), where 퐴 is the set of isolated vertices of 퐺 − 푒 2 − 퐵. Since 푄 is a dependence class, any perfect matching which contains one of the edges of 푄 contains all edges of 푄. Thus, since every perfect matching of 퐻 is also a perfect matching of 퐺, it follows that no edge of 푄 − 푒 1 − 푒 2 is matchable in 퐻. As |푄| ≥ 3, we conclude that 퐻 is not matching covered. On the other hand, any perfect matching of 퐺 that does not include the edges in 푄 is a perfect matching of 퐻. So, 퐻 is matchable. Moreover, since 푄 is a minimal class, all edges in 퐸 − 푄 are matchable in 퐺 − 푄 and hence also in the graph 퐻. As 퐻 is matchable but not matching covered, it follows (from Theorem 2.9) that there exists a nonempty proper subset 푆 of 퐵 such that |푁 (푆)| = |푆|, where 푁 (푆) is the neighbour set of 푆 in 퐻. Take 푋 to be the set 푆 ∪ 푁 (푆) and consider the cut 휕퐻 ( 푋). All the edges of 퐻 in this cut have one end in 푁 (푆) and one end in 퐵 − 푆. As |푁 (푆)| = |푆|, no such edge is matchable in 퐻. By the above observation, it follows that all edges in the cut 휕퐻 ( 푋) are in 푄. Now let 푀 denote a perfect matching of 퐺 which includes all the edges of 푄. A simple counting argument implies that |(푀 −푒 1 −푒 2 ) ∩휕퐺 ( 푋)| is equal to the number of ends of 푒 1 in 푆 minus the number of ends of 푒 2 in 푁 (푆). Since, as noted above, all edges in 휕퐺 ( 푋) are in 푄, it follows that |푄 ∩ 휕퐺 ( 푋)| ≤ 2. Furthermore |푄 ∩ 휕퐺 ( 푋)| has to be even because 휕퐺 ( 푋) is an even cut, and it cannot be zero because if that were the case graph 퐺 would be disconnected. It is now easy to conclude that 푄 ∩ 휕퐺 ( 푋) is an even 2-cut of 퐺. The four possible scenarios which give rise to even 2-cuts are illustrated in Figure 8.11. (We leave the details as Exercise 8.3.5.) Case 2 The graph 퐺 − 푒 2 − 퐵 has a nontrivial component, 퐾. By Theorem 8.13, 퐶 := 휕 (퐾) is a tight cut of 퐺. Now let 퐺 1 := 퐺/(푉 (퐾) → 푘),

and 퐺 2 := 퐺/(푉 (퐾) → 푘)

be the two 퐶-contractions of 퐺, and, for 푖 = 1, 2, let 푄 푖 := 푄 ∩ 퐸 (퐺 푖 ) denote the restrictions of the minimal class 푄 of 퐺 to the edge set of 퐺 푖 . Case 2.1 Either 푄 1 or 푄 2 has three or more edges. Let 푖 ∈ {1, 2} be such that |푄 푖 | ≥ 3. By Lemma 8.15(i) the set 푄 푖 is a minimal class of 퐺 푖 . Since |푉 (퐺 푖 )| < |푉 | and |푄 푖 | ≥ 3, it follows by induction that 푄 푖 includes an even 2-cut of 퐺 푖 . This validates the assertion of the Theorem, because an even 2-cut of 퐺 푖 is also an even 2-cut of 퐺 itself.

8.3 Removable Classes 푒1

173 푆

푒1

푆

퐵

퐴 푁 (푆 ) 푆

푒2 푒1

푁 (푆 ) 푆

푒2 푒1

퐵

퐴 푁 (푆 )

푒2

푁 (푆 )

푒2

Fig. 8.11 The four scenarios which give rise to even 2-cuts (Case 1)

Case 2.2 |푄 1 | ≤ 2 and |푄 2 | ≤ 2. In this case, the following assertion holds. 8.17.1 The edge 푒 2 belongs to the cut 퐶, |푄| = 3, and |푄 1 | = |푄 2 | = 2. Proof Since |푄| ≥ 3, it follows that both 푄 1 and 푄 2 are strict subsets of 푄. Lemma 8.15 now implies that |푄 ∩ 퐶| = 1. Let 푓 be the edge of 푄 in 퐶. If 푓 ≠ 푒 2 , the three edges 푒 1 , 푒 2 and 푓 belong to 푄 1 , contradicting the hypothesis that |푄 1 | ≤ 2. Hence 푒 2 must be the edge of 푄 in 퐶, and it is the only edge in 푄 1 ∩ 푄 2 . Since 푄 1 ∪ 푄 2 = 푄, it follows that |푄 1 | + |푄 2 | = |푄| + 1 ≥ 4. Since both 푄 1 and 푄 2 have fewer than three edges, this implies that |푄| = 3 and |푄 1 | = |푄 2 | = 2. ƒ The minimal class 푄 then has just one edge other than 푒 1 and 푒 2 . Let 푒 3 be that edge. This edge has both its ends in 푉 (퐾), the set 푄 1 is equal to {푒 1 , 푒 2 } and the set 푄 2 is equal to {푒 2 , 푒 3 }. Let 푥 be the end of 푒 2 in 푉 (퐾), let 푦 be the end of 푒 2 in 푉 (퐾). See Figure 8.12. As 푄 2 is a minimal dependence class in 퐺 2 there exists a barrier 퐵′ in 퐺 2 − 푒 3 such that (i) the ends of 푒 2 in 퐺 2 , namely 푥 and the contraction vertex 푘, belong to 퐵′ , and (ii) edge 푒 3 joins vertices in different components of 퐺 2 − 푒 3 − 퐵′ . Now let 퐵′′ := (퐵′ − 푘) ∪ 퐵. It can be verified (Exercise 8.3.6) that (i) 퐵′′ is a barrier of 퐺 − 푒 3 which contains both ends of 푒 1 and (ii) the vertex 푥 is the only end of 푒 2 in 퐵′′ . The components of 퐺 − 푒 3 − 퐵′′ are all those of 퐺 2 − 푒 3 − 퐵′ and those of 퐺 − 푒 2 − 퐵 other than 퐾. See Figure 8.13. If all components of 퐺 − 푒 3 − 퐵′′ are trivial, then Case 1 applies, with 푒 3 and 퐵′′ playing respectively the roles of 푒 2 and 퐵. Alternatively, if the graph 퐺 − 푒 3 − 퐵′′ has a nontrivial component, 퐿, then 퐺/푉 (퐿) contains all the three edges of 푄 and Case 2.1 applies, with 푒 3 , 퐵′′ and 퐿 playing respectively the roles of 푒 2 , 퐵 and 퐾. In all alternatives considered, we conclude that 푄 includes an even 2-cut. ƒ

8 Dependence Relation and Removable Classes

174 푒1

퐵

퐶 푒2

퐾 푥

푦

푒3

Fig. 8.12 The case in which 푄1 = {푒1 , 푒2 } and 푄2 = {푒2 , 푒3 } 푉 (퐾 ) 퐵′ − 푘 퐵′′ 퐺2 − 푘

푉 (퐾 )

퐶

퐵

푒1 푥

퐺1 − 푘

푒2 푦

푒3 Fig. 8.13 Barrier 퐵′′ of 퐺 − 푒3 containing the two ends of 푒1

Size of removable classes Corollary 8.18 Every removable class of a matching covered graph is either a singleton or a doubleton. Proof Let 푅 be a removable class of a matching covered graph 퐺. As 푅 is a removable class, it is necessarily a minimal class with respect to the dependence relation. The graph 퐺 − 푅 is matching covered, hence connected. Thus, 푅 does not include any nonempty cut of 퐺. In particular, 푅 does not include any even 2-cut. By Theorem 8.17, it follows that |푅| ≤ 2. ƒ From now on, for the above reason, we shall employ the term removable class to mean either a removable edge or a removable doubleton. A matching covered graph 퐺 need not have any removable classes. For example, as noted in the first section of this chapter, if |퐸 | ≥ 3 and every edge of 퐺 is incident with a vertex of degree two, then 퐺 cannot have removable classes. However, as also mentioned in the first section, we shall see in Chapter 11 that if the minimum degree of a matching covered graph 퐺 is at least three, then 퐺 does have at least one removable class.

8.3 Removable Classes

175

Exercises ⊲8.3.1 Draw the reduced dependence digraph of the matching covered graph shown in Figure 8.4, and determine all the minimal classes of that graph. ⊲8.3.2 Show that in a wheel of order 10 every edge in the rim induces four minimal classes, each of which consists of a single spoke. ⊲8.3.3 The objective of this exercise is to prove that every near-bipartite matching covered graph has exactly one brick. (In other words, every near-bipartite graph is a near-brick.) For this, let 퐺 be a near-bipartite graph, let 푅 := {푒, 푓 } be a removable doubleton of 퐺 such that 퐺 − 푅 is bipartite (and matching covered). (i) Prove that 퐺 is not bipartite. Hint: otherwise the set 퐸 △ (퐸 − 푅) = 푅 is a cut of 퐺. (ii) Suppose that 퐺 is not a brick. Let 퐶 := 휕 ( 푋) be a nontrivial tight cut of 퐺, note that 퐶 − 푅 is tight in 퐺 − 푅 and consider all the possible scenarios for the edges 푒 and 푓 in a way similar to that in Case 1 of the proof of Theorem 8.17. Conclude that one 퐶-contraction of 퐺 is bipartite and the other is near-bipartite. ⊲8.3.4 Show that, in the graph shown in Figure 8.14: (i) the set 푄 := {푒 1 , 푒 2 , 푒 3 } of edges is a dependence class and (ii) both { 푓1 } and { 푓2 } are minimal classes induced by 푒 푖 , for 푖 = 1, 2, 3. Hint: Observe that {푒 1 , 푓1 , 푓2 } is an even cut, and use the fact that every perfect matching meets an even cut in an even number of edges.

푒1 푢

푒2 푒3 푓1 푓2 푣 Fig. 8.14 The set {푒1 , 푒2 , 푒3 } is a dependence class which is not minimal. For any 푖 ∈ {1, 2, 3}, both { 푓1 } and { 푓2 } are minimal classes induced by the edge 푒푖 .

⊲8.3.5 In Case 1 of the proof of Theorem 8.17, supply all the details that are required to show that 휕퐺 ( 푋) is an even 2-cut which is a subset of 푄.

8 Dependence Relation and Removable Classes

176

⊲8.3.6 In the proof of Theorem 8.17, verify the fact that 퐵′′ = (퐵′ − 푘) ∪ 퐵 is a barrier of 퐺 − 푒 3 which contains the two ends of 푒 1 and that 푒 2 has precisely one end in 퐵′′ . Hint: Use the fact that 퐵 is a barrier of 퐺 − 푒 2 containing both ends of 푒 1 and that 퐵′ is a barrier of 퐺 2 containing the ends 푥 and 푘 of the edge 푒 2 in 퐺 2 .

⊲8.3.7 Let 퐵 denote a barrier of a matching covered graph and let 퐻 be the bipartite graph obtained from 퐺 by shrinking each component of 퐺 − 퐵 to a single vertex. Let 푅 be a removable class of 퐺. Prove that |푅 ∩ 퐸 (퐻)| ≤ 1 and also that 퐺 − 퐵 has precisely one component 퐾 such that 푅 ⊂ 퐸 (퐺 1 ), where 퐺 1 := 퐺/푉 (퐾). ∗ 8.3.8 Let 퐺 be a matching covered graph such that 훿(퐺) ≥ 3. Let C be a (possibly empty) maximal laminar collection of even 2-cuts of 퐺. Prove that 퐺 has at least Δ(퐺) + |C| removable classes. 8.3.9 Consider the brick 퐺 depicted in Figure 8.15. 퐶 푒

푓 Fig. 8.15 The doubleton {푒, 푓 } is a nonremovable dependence class

(i) Show that {푒, 푓 } is a dependence class in 퐺 and that any edge of the cut 퐶 depends on {푒, 푓 }. (ii) Deduce that {푒, 푓 } is not minimal (with respect to the partial order ⇒) and hence not removable. (iii) Show that every edge of 퐺 distinct from 푒 and 푓 is removable.

8.4 Two Lemmas for Future Use ♯♯ When 퐶 is the barrier cut associated with a special barrier of a matching covered graph, one of the 퐶-contractions of that graph is bipartite. Using Lemma 8.2 and Theorem 8.4 we now derive a lemma concerning removable edges in bipartite graphs which arise in this way in a very particular context. It will play a crucial role in the proofs of several important results in Part II, but is not needed for an understanding of the basic theory.

8.4 Two Lemmas for Future Use

177

Lemma 8.19 Let 퐺 be a bicritical graph. Let 푒 be a removable edge of 퐺 and let 퐶 := 휕 ( 푋) be a cut of 퐺. Suppose that the graph 퐻 := (퐺 − 푒)/( 푋 → 푥) is bipartite and matching covered. The following properties hold: (i) every edge of 퐶 − 푒 is removable in 퐻, and (ii) for every vertex 푣 of the minority part 푋− of 푋, if 푣 has degree three or more in 퐺 − 푒 then at most one edge of 휕 (푣) − 푒 is not removable in 퐺 − 푒. Proof Let ( 퐴, 퐵) denote the bipartition of 퐻, where 퐵 := 푋+ and 퐴 := 푋− + 푥. 8.19.1 Let ( 퐴′ , 퐴′′ ) be a partition of 퐴 and let (퐵′ , 퐵′′ ) be a partition of 퐵, such that | 퐴′ | = |퐵′ | and 퐻 has just one edge, say 푓 , that joins a vertex in 퐴′ to a vertex in 퐵′′ . Then, the contraction vertex 푥 is in 퐴′′ . (See Figure 8.16(a).) 퐵′

퐵′′

푓

푣 ′

퐴′′

퐴

(a) 퐵′

퐵′′

푓

푤 퐴′

− 푣

푋 퐴′′ (b)

Fig. 8.16 Illustration for Lemma 8.19

Proof Suppose that 푥 ∈ 퐴′ . Let 푣 denote the end of 푓 in 퐴. If 푓 ∈ 퐶, let 푤 denote the end of 푓 in 푋 (see Figure 8.16(b)). If 푓 is not in 퐶, then 푣 is an end of 푓 in 퐺, let 푤 := 푣. In either case, the set 퐴′′ + 푤 is a nontrivial barrier of 퐺. This conclusion is a contradiction to the hypothesis that 퐺 is bicritical. ƒ

178

8 Dependence Relation and Removable Classes

Let 푓 be any edge of 퐶 − 푒, and assume that 푓 is not removable in 퐻. By Lemma 8.2, there exists a partition ( 퐴′ , 퐴′′ ) of 퐴 and a partition (퐵′ , 퐵′′ ) of 퐵 such that | 퐴′ | = |퐵′ | and 푓 is the only edge of 퐻 that joins a vertex in 퐴′ to a vertex in 퐵′′ . In that case, as 푓 ∈ 퐶, the end of 푓 in 퐴 is 푥; hence 푥 is not in 퐴′′ , a contradiction to statement (8.19.1). We conclude that 푓 is removable in 퐻. This conclusion holds for each edge 푓 in 퐶 − 푒. Let 푣 be a vertex of 퐴 − 푥 that has degree three or more in 퐺 − 푒. Then, 푣 has degree three or more in 퐻. Suppose that 휕 (푣) − 푒 contains two edges, say 푣푏 1 and 푣푏 2 , that are not removable in 퐻. By Theorem 8.4, with 푣 playing the role of 푎, there exist a partition ( 퐴0 , 퐴1 , 퐴2 ) of 퐴 and a partition (퐵0 , 퐵1 , 퐵2 ) of 퐵 such that, for 1 ≤ 푖 ≤ 2, | 퐴푖 | = |퐵푖 | and 푣 ∈ 퐴0 , 푏 푖 ∈ 퐵푖 and 푣푏 푖 is the only edge that joins a vertex of 퐴 − 퐴푖 to a vertex in 퐵푖 . The sets 퐴1 and 퐴2 are disjoint; therefore one of them does not contain the contraction vertex 푥. Adjust notation so that 푥 ∉ 퐴2 . We then have a contradiction to statement (8.19.1), with 퐴0 ∪ 퐴1 , 퐴2 , 퐵0 ∪ 퐵1 and 퐵2 playing respectively the roles of 퐴′ , 퐴′′ , 퐵′ and 퐵′′ . We conclude that at most one edge of 휕 (푣) − 푒 is not removable in 퐻. The vertex 푣 and the contraction vertex 푥 are in the same part of the bipartition of 퐻. Thus, no edge of 휕 (푣) − 푒 is in 퐶 − 푒. Moreover, as 퐻 is matching covered, the cut 퐶 − 푒, a barrier cut of 퐺 − 푒, is tight in 퐺 − 푒. It follows that every edge of 휕 (푣) − 푒 which is removable in 퐻 is also removable in 퐺 − 푒. We deduce that at most one edge of 휕 (푣) − 푒 is not removable in 퐺 − 푒. This conclusion holds for each vertex 푣 in 퐴 − 푥 that has degree three or more in 퐺 − 푒. ƒ

8.4.1 The Zigzag Lemma We shall now make use of Theorem 8.4 and Lemma 8.2 to derive another lemma concerning removable edges in bipartite graphs. This result will have a crucial role in our proof of an important result in Part III, but is not needed for an understanding of the basic theory. Lemma 8.20 (The Zigzag Lemma) Let 퐺 := 퐺 [ 퐴, 퐵] be a bipartite matching covered graph and let 푃 := (푣 1 , 푣 2 , 푣 3 , 푣 4 ), 푣 1 ∈ 퐴, be a path in 퐺 such that no edge of 푃 is removable in 퐺 and the vertices 푣 2 and 푣 3 are adjacent to three or more vertices. See Figure 8.17. Then, 퐴 has a partition ( 퐴1 , 퐴2 , 퐴3 , 퐴4 ) and 퐵 has a partition (퐵1 , 퐵2 , 퐵3 , 퐵4 ) such that (i) 푣 1 ∈ 퐴1 , 푣 2 ∈ 퐵2 , 푣 3 ∈ 퐴3 , 푣 4 ∈ 퐵4 , (ii) for 푖 = 1, 2, 3, 4, | 퐴푖 | = |퐵푖 |, (iii) 푣 1 푣 2 is the only edge that joins a vertex in 퐴1 to a vertex of 퐵 − 퐵1 , (iv) 푣 3 푣 4 is the only edge that joins a vertex of 퐴 − 퐴4 to a vertex of 퐵4 , and (v) 푣 2 푣 3 is the only edge that joins a vertex of 퐵1 ∪ 퐵2 to a vertex of 퐴3 ∪ 퐴4 . Proof The edges 푣 2 푣 3 and 푣 3 푣 4 are not removable in 퐺 and the vertex 푣 3 is adjacent to three or more vertices in 퐺. From Theorem 8.4, with 푣 3 , 푣 2 and 푣 4 playing

8.4 Two Lemmas for Future Use 퐵1

179 퐵2

퐵3

퐵4

푣2

푣4

푣1

푣3

퐴1

퐴2

퐴3

퐴4

Fig. 8.17 The path 푃 and the partitions of 퐴 and 퐵

respectively the role of 푎, 푏 1 and 푏 2 , we infer that 퐴 has a partition ( 퐴12 , 퐴3 , 퐴4 ) and 퐵 has a partition (퐵12 , 퐵3 , 퐵4 ) such that (i) (ii) (iii) (iv)

푣 2 ∈ 퐵12 , 푣 3 ∈ 퐴3 and 푣 4 ∈ 퐵4 , | 퐴12 | = |퐵12 |, | 퐴3 | = |퐵3 | and | 퐴4 | = |퐵4 |, 푣 3 푣 4 is the only edge that joins a vertex in 퐴 − 퐴4 to a vertex in 퐵4 , and 푣 2 푣 3 is the only edge that joins a vertex in 퐵12 to a vertex in 퐴 − 퐴12 .

See Figure 8.18. 푋 퐵12 − 푣2

퐵3 푣2

퐵4 푣4

푒 푣1

푣3

퐴12

퐴3

퐴4

퐶 Fig. 8.18 The partitions ( 퐴12 , 퐴3 , 퐴4 ) and ( 퐵12 , 퐵3 , 퐵4 )

Let 푒 := 푣 1 푣 2 in 퐺. Note that 푣 1 ∈ 퐴12 and 푣 2 ∈ 퐵12 . Let 퐶 := 휕 ( 푋), where 푋 := 퐴3 ∪ 퐴4 ∪ 퐵3 ∪ 퐵4 ∪ {푣 2 }. The cut 퐶 is tight in 퐺. Let 퐻 := 퐺/( 푋 → 푥). Then, ( 퐴12 , (퐵12 − 푣 2 ) + 푥) is the bipartition of 퐻. See Figure 8.19(a). By hypothesis, the vertex 푣 2 is adjacent to three or more vertices in 퐺. Thus, 푣 2 is joined to one or more vertices in 퐴12 − 푣 1 ; hence 푒 is a multiple edge in 퐻 ′ := 퐺/( 푋 → 푥). As 푒 is not removable in 퐺, we conclude that 푒 is not removable in 퐻. From Lemma 8.2, with 퐴12 , (퐵12 − 푣 2 ) + 푥, 푣 1 and 푥 playing respectively the roles of 퐴, 퐵, 푎 and 푏, we infer that there is a partition ( 퐴1 , 퐴2′ ) of 퐴12 and a partition

8 Dependence Relation and Removable Classes

180

퐵1

퐵12 − 푣2

푥

푥

퐵2′ − 푥

푒 푒 푣1

푣1

퐴12

퐴′2

퐴1

(푎)

(푏)

Fig. 8.19 The graph 퐻 and the partitions of 퐴퐻 and 퐵퐻

(퐵1 , 퐵′2 ) of (퐵12 − 푣 2 + 푥) such that 푒 is the only edge that joins a vertex of 퐴1 to a vertex in 퐵′2 , | 퐴1 | = |퐵1 | and | 퐴2′ | = |퐵′2 |. See Figure 8.19(b). Let us now expand vertex 푥. Let 푓 be an edge of 퐶 − 푒. As can be seen in Figure 8.19(b), 푓 is incident with a vertex in 퐴2′ and as can be seen in Figure 8.18, 푓 is in 휕 (퐵3 + 푣 2 ). See Figure 8.20. 퐵2′ − 푥

퐵1

푣2

퐵3

퐵4 푣4

푒

푣1 퐴1

푣3 퐴′2

퐴3

퐴4

Fig. 8.20 The expansion of vertex 푥

The assertion holds, with 퐴2 := 퐴2′ and 퐵2 := (퐵′2 − 푥) + 푣 2 .

ƒ

Recall the concept of certificate for nonremovability of an edge in a bipartite graph, defined in the statement of Lemma 8.2. In a paper submitted for publication, Lucchesi [63, 2023] proved the following generalization of that Lemma. Theorem 8.21 Let 퐺 [ 퐴, 퐵] be a bipartite matching covered graph of order four or more and let 푇 be a nontrivial tree which is a subgraph of 퐺. If (i) no edge of 푇 is removable in 퐺 and (ii) each internal vertex of 푇 has degree three or more in 퐺, then there is a laminar collection C of cuts, each of which is a certificate of an edge of 푇. ƒ

181

8.5 Notes

Example 8.22 In Figure 8.21, the tree 푇 consists of the four edges 푒 1 := 푣 1 푣 2 , 푒 2 := 푣 2 푣 3 , 푒 3 := 푣 3 푣 4 , and 푒 4 := 푣 3 푣 5 . For each 푒 푖 , 푖 = 1, 2, 3, 4, the corresponding certificate cut 퐶푖 is defined as follows: 퐶1 := 휕 (푆1),

퐶2 := 휕 (푆1 ∪ 푆2 ),

퐶3 := 휕 (푆4)

푣2

푣4

푣5

푣3

푣1 푆1

and 퐶4 := 휕 (푆5).

푆2

푆3

푆4

푆5

Fig. 8.21 An illustration for Theorem 8.21.

Theorem 8.21 may be proved by induction on the number of edges of the tree (Exercise 8.4.2). Interestingly enough, Theorem 8.4 and Lemma 8.20 my be easily derived from Theorem 8.21 (Exercises 8.4.3 and 8.4.4).

Exercises ∗ 8.4.1 Let 퐺 be brace of order at least six and let 푒 be an edge of 퐺. Show that: (i) for any vertex 푣 that has degree at least three in 퐺 − 푒, at most one edge of 퐺 − 푒 incident with 푣 is not removable in 퐺 − 푒, and (ii) if 퐶 := 휕 ( 푋) − 푒 is a nontrivial tight cut of 퐺 − 푒 then each edge of 퐶 is removable in 퐺 − 푒. (CLM [18]) ⊲8.4.2 Prove Theorem 8.21 by induction on the number of edges of 퐺. ⊲8.4.3 Deduce Theorem 8.4 from Theorem 8.21. ⊲8.4.4 Deduce Lemma 8.20 from Theorem 8.21.

8.5 Notes The paper written by CLM (1999, [8]) is a self-contained introduction to the theory of removable classes in matching covered graphs. That paper can be thought of as a useful supplement to the material in this chapter and in Chapters 9 and 11.

8 Dependence Relation and Removable Classes

182

The next chapter is devoted to the study of removable classes in bricks. We shall see that in a brick every dependence class, minimal or not, has at most two edges. In 2014, CLM found the graph in Figure 8.9 to show that 3-connected matching covered graphs which are not bricks could have dependence classes of size more than two. But both of us forgot about it. Recently, Lu, Kothari, Feng and Zhang (2020, [62]) showed that, given any 푘 ≥ 2 and ℓ ≥ 1, there exists a 푘-connected matching covered graph which has a dependence class of size ℓ. The graph depicted in Figure 8.22 is a 3-connected matching covered graph in which {푒 1 , 푒 2 , 푒 3 } is a dependence class of size three. 푒1

푒2 푒3

Fig. 8.22 A 3-connected matching covered graph with a dependence class of size three

The above graph is obtained from the one shown in Figure 8.14 by splicing bipartite graphs at vertices 푢 and 푣. Interestingly, this trick preserves the dependence class {푒 1 , 푒 2 , 푒 3 } while boosting the connectivity up by one! Lu, Kothari, Feng and Zhang [62] use similar ideas to construct, for any given 푘 ≥ 2 and ℓ ≥ 1, a 푘connected matching covered graph which has a dependence class of specified size ℓ.

Chapter 9

Dependence Classes in Bricks

Contents 9.1

9.2 9.3 9.4

Dependence Classes in Bricks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183 9.1.1 Monotonicity of functions 푏 and 푏 + 푝 . . . . . . . . . . . . . . . . . 186 9.1.2 Bricks with more than two removable doubletons . . . . . . . . 190 Mutually Exclusive Classes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192 9.2.1 Removable edges in bricks . . . . . . . . . . . . . . . . . . . . . . . . . . . 197 Removable Edges in Near-Bipartite Bricks . . . . . . . . . . . . . . . . . . . . . 199 9.3.1 Nonremovable edges in near-bipartite bricks . . . . . . . . . . . . 204 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205

9.1 Dependence Classes in Bricks We continue our study of dependence classes in matching covered graphs. But here our attention is mainly devoted to bricks. We have seen that, in general, dependence classes in matching covered graphs can be arbitrarily large (see Figure 8.10), although no removable class can have more than two edges (Corollary 8.18). In case of bricks, as we shall see, all dependence classes, regardless of whether they are removable or not, are either singletons or doubletons. This is a consequence of the following theorem, which deals with mutually dependent edges in bricks, and is a generalization of a result proved by Lov´asz [58]. Mutually dependent edges in bricks Theorem 9.1 (CLM(1999, [8])) Let 퐺 be a brick and let 푒 and 푓 be two mutually dependent edges of 퐺. Then 퐺 − 푒 − 푓 is a bipartite graph, with parts of equal cardinality, such that both ends of 푒 lie in one part of its bipartition and both ends of 푓 lie in the other. © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7_9

183

184

9 Dependence Classes in Bricks

Proof As 푒 and 푓 are mutually dependent by the hypothesis, it follows from Proposition 8.10 that there exists a barrier, say 퐵, of 퐺 − 푓 which contains both ends of the edge 푒. Furthermore, by Theorem 8.13, • edge 푓 has its ends in different components of 퐺 − 푓 − 퐵, • edge 푒 is the only edge of 퐺 having both ends in 퐵, • for each component 퐾 of 퐺 − 푓 − 퐵, the cut 휕 (퐾) is tight in 퐺. But the graph 퐺, being a brick, is free of nontrivial tight cuts. We conclude that each component of 퐺 − 푓 − 퐵 is in fact trivial. Let 퐴 denote the set 푉 − 퐵. Then, | 퐴| = |퐵| and 퐺 − 푒 − 푓 is a bipartite graph with bipartition ( 퐴, 퐵). Hence the assertion. ƒ Sizes of dependence classes in bricks Corollary 9.2 Any dependence class of a brick 퐺 has at most two edges. Proof Assume to the contrary that 퐺 has a dependence class, 푄, such that |푄| ≥ 3, and let 푒, 푓 and 푔 denote three distinct edges of 푄. Since 푒 ⇔ 푓 , it follows from Theorem 9.1 that 퐸 − 푒 − 푓 is a cut of 퐺. Likewise, 퐸 − 푒 − 푔 is also a cut of 퐺. As the symmetric difference of any two cuts of 퐺 is a cut of 퐺, it follows that (퐸 − 푒 − 푓 ) △ (퐸 − 푒 − 푔) = { 푓 , 푔} is a cut of 퐺. This is absurd because 퐺, being a brick, is 3-edge-connected. ƒ The brick shown in Figure 8.15 (Exercise 8.3.9) shows that, in general, a dependence class in a brick need not be minimal and hence may not be removable. However, using the fact all bricks are 3-connected, it is easy to deduce the following corollary, whose proof is left as Exercise 9.1.2. Corollary 9.3 Every minimal dependence class in a brick is removable.

ƒ

The following is another simple consequence of Theorem 9.1. Corollary 9.4 The union 푄 1 ∪ 푄 2 of any two dependence classes 푄 1 and 푄 2 of size two of a brick 퐺 is a 4-cut of 퐺. Proof Clearly, the two doubletons 푄 1 and 푄 2 are disjoint; hence |푄 1 ∪ 푄 2 | = 4. As in the proof of Corollary 9.2, 퐸 −푄 1 and 퐸 −푄 2 are cuts; hence so is their symmetric difference 푄 1 ∪ 푄 2 . We conclude that 푄 1 ∪ 푄 2 is a 4-cut of 퐺. ƒ For example, the prism P2푛 , where 푛 ≥ 3 is odd, and the M¨obius ladder M2푛 , where 푛 ≥ 4 is even, are bricks. Each of these bricks has 푛 dependence classes of size two, and the union of any two of these classes is a 4-cut. The brick 퐾4 is a M¨obius ladder; it has three dependence classes of size two and the union of any two of them is a 4-cut of 퐾4 . Theorem 9.10 describes the structure of bricks that have three or more removable doubletons.

9.1 Dependence Classes in Bricks

185

In all the examples mentioned above, each dependence class of size two is also a removable class. But, as the example in Figure 8.15 shows, a dependence class 푅 of size two in a brick 퐺 need not be removable in that brick. However, if 푅 is removable, that is, if 퐺 − 푅 is matching covered, then we obtain the following result, due to Lov´asz [58], as a consequence of Theorem 9.1. (See Exercise 8.1.7 for Lov´asz’s original proof.) Removable doubletons in bricks Theorem 9.5 If 푅 is a removable doubleton of a brick 퐺, then the graph 퐺 − 푅 ƒ is a bipartite matching covered graph. Thus bricks with removable doubletons are the same as near-bipartite bricks, implying that if 푅 is a removable doubleton in a brick 퐺, then 푏(퐺 − 푅) = 0. A more general result to be proved later (Theorem 9.9) shows that if 푅 is a removable doubleton in any (nonbipartite) matching covered graph 퐺, then 푏(퐺 − 푅) = 푏(퐺) − 1. Recall that a minimal class 푅 of a matching covered graph 퐺 is said to be induced by an edge 푒 of 퐺 if 푅 contains an edge 푓 such that 푓 depends on 푒, that is, if every perfect matching of 퐺 containing 푓 also contains 푒. Clearly minimal classes which are induced by two adjacent edges have to be distinct. This observation leads to the following assertion. A lower bound on the number of removable classes in bricks Theorem 9.6 Every brick 퐺 has at least Δ(퐺) removable classes, where Δ(퐺) is the maximum of the degrees of vertices in 퐺. Proof Let 푒 1 , 푒 2 , . . . , 푒 Δ be the edges of 퐺 incident with a vertex, say 푣, of degree Δ, and, for 1 ≤ 푖 ≤ Δ, let 푅푖 denote a minimal class of 퐺 induced by 푒 푖 . Since any two edges incident with 푣 are adjacent, it follows that 푅1 , 푅2 , . . . , 푅Δ are distinct minimal classes of 퐺. Moreover, since 퐺, being a brick, is 3-connected, it follows from Corollary 9.2 that each 푅푖 is either a singleton or a doubleton, and hence that 퐺 − 푅푖 is connected. We conclude that 푅1 , 푅2 , . . . , 푅Δ are distinct removable classes. ƒ If a brick has no removable doubletons, then all its removable classes are removable edges. Thus: Corollary 9.7 Any brick 퐺 that is not near-bipartite has at least Δ(퐺) removable edges. ƒ It follows from the above corollary that a cubic brick which is not near-bipartite has at least three removable edges. The tricorn H10 (Figure 2.5(d)) is an example of such a cubic brick with precisely three removable edges (Exercise 9.1.1). We do not know if there exist arbitrarily large cubic bricks which are free of removable

186

9 Dependence Classes in Bricks

doubletons but have just three removable edges. See Conjecture 9.19 in the Notes Section at the end of this chapter. This result clearly does not, in general, hold for near-bipartite bricks. (For example 퐾4 and 퐶6 are cubic bricks with no removable edges, and the bicorn H8 is a cubic brick with just one removable edge. See Figure 2.5.) In the next section we shall study the structure of near-bipartite bricks which will help us in deducing lower bounds on the number of removable edges in near-bipartite bricks. We shall see in particular that any simple near-bipartite brick distinct from the three mentioned above has at least two disjoint removable edges.

9.1.1 Monotonicity of functions 풃 and 풃 + 풑 The invariant 푏(퐺) of a matching covered graph 퐺, which denotes the number of bricks of 퐺, was introduced in Chapter 4, and its significance in determining the dimension of the matching space was described in Chapter 6. There is another invariant of 퐺 which will prove to be of special interest in later chapters: a Petersen brick is any graph whose underlying simple graph is the Petersen graph. Theorem 4.17 implies that the number of Petersen bricks in a tight cut decomposition of 퐺 is independent of the tight-cut-contractions employed in the decomposition. We denote this number by 푝(퐺). We may consider counting the number of bricks, Petersen bricks counted twice. This is the invariant (푏 + 푝) (퐺) := 푏(퐺) + 푝(퐺). As we shall see in later chapters, it plays a special role in various aspects of this theory, especially in the theory of ear decompositions. (See the Notes Section at the end of this chapter for an example of an assertion about matching covered graphs which involves the parameter 푏 + 푝.) When the graph 퐺 under consideration is clear from the context, we shall simply write 푏 and 푝 for 푏(퐺) and 푝(퐺). The purpose of this subsection is to study the effect of deletions of removable classes from a matching covered graph on the two parameters 푏 and 푏 + 푝. Monotonicity with respect to deletions of removable edges Theorem 9.8 (CLM (2002, [9])) Let 퐺 be a matching covered graph and let 푒 be a removable edge of 퐺. Then, 푏(퐺 − 푒) ≥ 푏(퐺) and (푏 + 푝) (퐺 − 푒) ≥ (푏 + 푝) (퐺).

(9.1)

Proof By induction on |푉 |. Consider first the case in which 퐺 has a nontrivial tight cut, say 퐶. Let 퐺 ′ and 퐺 ′′ be the two 퐶-contractions of 퐺. By Proposition 8.8, both 퐺 ′ − 푒 and 퐺 ′′ − 푒 are matching covered. Moreover, every perfect matching of 퐺 − 푒 is a perfect matching of 퐺, whence 퐶 − 푒 is tight in 퐺 − 푒. By induction, using the uniqueness of the tight cut decomposition, we then have 푏(퐺 − 푒) = 푏(퐺 ′ − 푒) + 푏(퐺 ′′ − 푒) ≥ 푏(퐺 ′ ) + 푏(퐺 ′′ ) = 푏(퐺).

9.1 Dependence Classes in Bricks

187

Likewise, (푏 + 푝) (퐺 − 푒) = (푏 + 푝) (퐺 ′ − 푒) + (푏 + 푝) (퐺 ′′ − 푒)

≥ (푏 + 푝) (퐺 ′ ) + (푏 + 푝) (퐺 ′′ ) = (푏 + 푝) (퐺).

The assertion holds in this case. We may thus assume that 퐺 is free of nontrivial tight cuts. That is, 퐺 is either a brick or a brace. If 퐺 is a brace then clearly 퐺 − 푒 is bipartite, whence 푏(퐺 − 푒) = (푏 + 푝) (퐺 − 푒) = 푏(퐺) = (푏 + 푝) (퐺) = 0 and the assertion holds. We may thus assume that 퐺 is a brick. If the underlying simple graph of 퐺 is the Petersen graph then 푝(퐺) = 1. In that case, for every edge 푒 of 퐺, if 푒 is a multiple edge then 푏(퐺 − 푒) = 푝(퐺 − 푒) = 1, whereas if 푒 is not a multiple edge then 푏(퐺 − 푒) = 2. In both alternatives, the assertion holds. Finally, we may assume that 퐺 is a brick and 푝(퐺) = 0. It now suffices to prove that 퐺 − 푒 is not bipartite. This follows from Exercise 8.1.3. ƒ In all cases considered, the asserted inequalities hold. The inequality established in the above theorem need not be tight; by deleting an edge from a brick it is possible to obtain a matching covered graph with an arbitrarily large number of bricks (see Exercise 9.1.3). In contrast, as is shown by the following assertion, the changes in the values of the functions 푏 and 푏 + 푝 effected by the deletion of a removable doubleton are predictable. We note that, by definition, a removable doubleton is a minimal removable class. By Theorem 8.1, if a matching covered graph has a removable doubleton then it is nonbipartite. Monotonicity with respect to deletions of removable doubletons Theorem 9.9 (CLM (2002, [11])) Let 퐺 be a (nonbipartite) matching covered graph, and let 푅 be a removable doubleton of 퐺. Then, 푏(퐺 − 푅) = 푏(퐺) − 1 and (푏 + 푝) (퐺 − 푅) = (푏 + 푝) (퐺) − 1.

(9.2)

Proof By induction on |푉 |. Case 1 Graph 퐺 is a brick. As 퐺 is a brick, 푏(퐺) = 1. Furthermore, since 퐺 has a removable doubleton, it cannot be a Petersen brick because such bricks do not have removable doubletons, and thus (푏 + 푝) (퐺) = 1. Since 푅 is a removable doubleton of the brick 퐺, by Theorem 9.5, the graph 퐺 − 푅 is a bipartite matching covered graph. Hence 푏(퐺 − 푅) = 0 = 푏(퐺) − 1, and (푏 + 푝) (퐺 − 푅) = 0 = (푏 + 푝) (퐺) − 1.

9 Dependence Classes in Bricks

188

Case 2 Graph 퐺 has a nontrivial tight cut 퐶 such that |퐶 ∩ 푅| = 0. Let 퐺 1 and 퐺 2 denote the two 퐶-contractions of 퐺. As 푅 and 퐶 are disjoint, it follows, from Lemma 8.15, that 푅 is is entirely contained in one of 퐸 (퐺 1 ) − 퐶 and 퐸 (퐺 2 ) − 퐶, and also that if 푅 ⊂ 퐸 (퐺 푖 ) − 퐶, then 푅 is a removable doubleton of 퐺 푖 . Assume without loss of generality that 푅 ⊂ 퐸 (퐺 2 ) − 퐶. Clearly every perfect matching of 퐺 − 푅 is a perfect matching of 퐺, and therefore 퐶 is a tight cut of 퐺 − 푅. By induction and Theorem 4.20, we have: 푏(퐺 − 푅) = 푏(퐺 1 ) + 푏(퐺 2 − 푅) = 푏(퐺 1 ) + 푏(퐺 2 ) − 1 = 푏(퐺) − 1. Likewise, (푏 + 푝) (퐺 − 푅) = (푏 + 푝) (퐺 1 ) + (푏 + 푝) (퐺 2 − 푅)

= (푏 + 푝) (퐺 1 ) + (푏 + 푝) (퐺 2 ) − 1 = (푏 + 푝) (퐺) − 1.

Hence the required assertion holds in this case. Case 3 There is no nontrivial tight cut 퐶 of 퐺 such that |퐶 ∩ 푅| = 0. As 푅 is a removable doubleton, every perfect matching which contains one of the edges of 푅 also contains the other. Hence no tight cut of 퐺 can contain both the edges of 푅. Thus, if 퐶 is a tight cut such that |퐶 ∩ 푅| ≠ 0, then |퐶 ∩ 푅| = 1. Let 푅 := {푒, 푓 } and suppose that 푒 is the edge of 푅 in 퐶. Let 퐺 1 and 퐺 2 be the two 퐶-contractions of 퐺, and assume without loss of generality that 퐸 (퐺 1 ) ∩ 푅 = {푒} and 퐸 (퐺 2 ) ∩ 푅 = 푅. Since 푅 is a removable class in 퐺 it follows that 푒 is a removable edge of 퐺 1 and 푅 is a removable doubleton of 퐺 2 . At this stage one might think that the desired identities can be obtained by induction as in the case above. But this is not quite so straightforward. Although one can assert, using induction, that 푏(퐺 2 − 푅) = 푏(퐺 2 ) − 1, there is no obvious way to assert that 푏(퐺 1 − 푒) = 푏(퐺 1 ). For this reason, rather than considering arbitrary tight cuts of 퐺, we restrict our attention to barrier cuts and 2-separation cuts. (By the ELP Theorem 5.6, any matching covered graph that has a nontrivial tight cut either has a nontrivial barrier cut or a 2-separation cut.) Case 3.1 Graph 퐺 has a nontrivial barrier 퐵. Let 퐺 1 denote the bipartite matching covered graph obtained by shrinking the vertex set of each odd component of 퐺 − 퐵 to a single vertex. The graph 퐺 − 퐵 must have one component 퐾 such that 푅 ⊂ 퐸 (퐺 2 ), where 퐺 2 := 퐺/푉 (퐾) (see Exercise 8.3.7). By the hypothesis of the case, the cut 휕 (퐾) contains an edge of 푅, say 푒. Thus, 푓 has both ends in 푉 (퐾). For any component 퐿 of 퐺 − 퐵 distinct from 퐾, the (tight) cuts 휕 (퐾) and 휕 (퐿) are disjoint; therefore 휕 (퐿) and 푅 are disjoint. By the hypothesis of the case, 퐿 is trivial. It follows that 퐾 is the unique nontrivial odd component of 퐺 − 퐵. Consequently, the bipartite graph 퐺 1 is equal to 퐺/푉 (퐾).

9.1 Dependence Classes in Bricks

189

퐵 푒

푓 퐾 Fig. 9.1 퐺 − 퐵 has one nontrivial component 퐾 and 퐶퐾 ∩ 푅 = {푒}

Since 푅 is removable in 퐺, it follows that 퐸 (퐺 1 ) ∩ 푅 = {푒} is a removable singleton of 퐺 1 , and 퐸 (퐺 2 ) ∩ 푅 = 푅 is a removable doubleton of 퐺 2 . Since 퐺 1 is bipartite, 푏(퐺 1 − 푒) = 푏(퐺 1 ) = 0. By induction, 푏(퐺 2 − 푅) = 푏(퐺 2 ) − 1 and (푏 + 푝) (퐺 2 − 푒) = (푏 + 푝) (퐺 2 ) − 1. The asserted equalities may now be deduced as before. Case 3.2 Graph 퐺 is bicritical but it has a 2-separation {푢, 푣}. Let 퐻1 and 퐻2 be two vertex disjoint subgraphs of 퐺 such that 퐺 − 푢 − 푣 is the union of 퐻1 and 퐻2 . Then, as described in Chapter 4, the two cuts 퐶 := 휕 (푉 (퐻1) + 푣) and 퐷 := 휕 (푉 (퐻2) + 푣) are the two 2-separation cuts associated with {푢, 푣}. We shall first show that both the vertices 푢 and 푣 must be adjacent to at least two distinct vertices in each of the even sets 푉 (퐻1 ) and 푉 (퐻2 ). Suppose that this is not the case and assume that 푢 is adjacent to just one vertex 푤 in 푉 (퐻1 ). In this case the graph 퐺 − 푣 − 푤 is disconnected and has an odd component whose vertex set is contained in 푉 (퐻1 ) − 푤. It follows that {푣, 푤} is a nontrivial barrier of 퐺. This is not possible because, by hypothesis of the subcase, 퐺 is bicritical. Thus each of 푢 and 푣 must be adjacent to at least two distinct vertices in each of the sets 푉 (퐻1 ) and 푉 (퐻2 ). A consequence of this observation is that each edge of 퐶 is a multiple edge in one of the 퐶-contractions, and an analogous statement holds for 퐷. Furthermore, since Case 2 does not apply, |퐶 ∩ 푅| = |퐷 ∩ 푅| = 1. (It is easy to see that any edge joining 푢 and 푣 is removable in 퐺 and hence cannot be in the doubleton 푅 (Exercise 9.1.4). So, 퐶 ∩ 푅 ≠ 퐷 ∩ 푅.) Adjust notation so that 퐶 ∩ 푅 = {푒}. In one of the two 퐶-contractions of 퐺, say 퐺 1 , the edge 푒 is a multiple edge, and in the other 퐶-contraction of 퐺, say 퐺 2 , the set 푅 is a removable doubleton. In the illustration depicted in Figure 9.2 edge 푒 of the doubleton 푅 is in 퐶 := 휕 ( 푋), where 푋 := 푉 (퐻1 ) + 푣, and the second edge 푓 of 푅 has both its ends in 푋 := 푉 (퐻2 ) + 푢. As the edge 푒 is multiple in 퐺 1 , it is removable in 퐺 1 . Moreover, the underlying simple graphs of 퐺 1 and of 퐺 1 − 푒 are the same. Therefore, 푏(퐺 1 − 푒) = 푏(퐺 1 ) and (푏 + 푝) (퐺 1 − 푒) = (푏 + 푝) (퐺 1 ). On the other hand, as 푅 is a subset of 퐸 (퐺 2 ), it is a removable doubleton of 퐺 2 . Therefore, by induction, 푏(퐺 2 − 푅) = 푏(퐺 2 ) − 1 and (푏 + 푝) (퐺 2 − 푅) = (푏 + 푝) (퐺 2 ) − 1. The desired equalities in (9.2) now follow, as in the previous case. ƒ

190

9 Dependence Classes in Bricks 퐶

퐷 푢 푓

퐻1

퐻2

푒 푣 푋

푋

Fig. 9.2 Edge 푒 is a multiple edge in one of the 퐶-contractions of 퐺, and 푅 is a removable doubleton in the other 퐶-contraction of 퐺

9.1.2 Bricks with more than two removable doubletons The next assertion describes the structure of bricks that have three or more removable doubletons. Its proof is left as Exercise 9.1.6. Theorem 9.10 (Carvalho, Kothari, Lucchesi, Murty (2015, [26])) Let 퐺 be a brick distinct from 퐾4 which has precisely 푘 ≥ 3 removable doubletons, and let 푅 be the union of the 푘 doubletons. Then, 퐺 − 푅 has precisely 푘 components, each of which is bipartite and matchable. Moreover, the 푘 components may be enumerated as 퐻1 , 퐻2 , . . . , 퐻 푘 and the 푘 removable doubletons may be enumerated as 푅1 , 푅2 , . . . , 푅 푘 such that 퐻푖 is joined to 퐻푖+1 by the two edges of 푅푖 , as illustrated in Figure 9.3 (indices are taken modulo 푘). ƒ

퐶1

퐶2

퐶3 푒4

퐻1

퐻2

퐻3

퐻4

푓4 Fig. 9.3 The general structure of 퐺, where 푘 = 4, 푅4 = {푒4 , 푓4 } and 퐶푖 := 푅푖 ∪ 푅4 , for 푖 = 1, 2, 3

A brick need not have any removable doubletons. But it may have as many as 푛/2 removable doubletons. Moreover, if a simple brick has 푛/2 removable doubletons, then it is either a prism or a M¨obius ladder, as stated below.

9.1 Dependence Classes in Bricks

191

Corollary 9.11 Every brick of order 푛 ≥ 6 has at most 푛/2 removable doubletons. Moreover, a simple brick of order 푛 has 푛/2 removable edges if and only if it is a prism or a M¨obius ladder. ƒ

Exercises 9.1.1 Show that the tricorn has no removable doubletons and that it has precisely three removable edges. ⊲9.1.2 Prove Corollary 9.3. ⊲9.1.3 For each positive integer 푘, find a brick 퐺 with a removable edge 푒 such that 푏(퐺 − 푒) = 푘. ⊲9.1.4 Let {푢, 푣} be a separation of a matching covered graph 퐺 as in Case 3.2 of the proof of Theorem 9.9. Show that any edge joining 푢 and 푣 is removable in 퐺 and hence cannot be in a removable doubleton of 퐺. ⊲9.1.5 For each positive integer 푘, find a matching covered graph 퐺 with a removable doubleton 푅 such that 푏(퐺) = 푘 and 푏(퐺 − 푅) = 푘 − 1. ∗ 9.1.6 The objective of this exercise is to obtain a proof of Theorem 9.10. Let 퐺 be a brick, let R be the collection of the 푘 ≥ 3 removable doubletons of 퐺, let 푅 be the union of the 푘 doubletons, let 푅 푘 be one of the doubletons, and let C := {푅 푘 ∪ 푆 : 푆 ∈ R − 푅 푘 }. Note that each one of the 푘 − 1 sets in C is a 4-cut of 퐺, by Corollary 9.4. (i) Prove that if C is not laminar then 퐺 is 퐾4 . Hint: suppose that C contains two cuts, 퐶 := 휕 ( 푋) and 퐷 := 휕 (푌 ), that cross. (a) Consider the four quadrants 푋 ∩ 푌 , 푋 ∩ 푌, 푋 ∩ 푌 and 푋 ∩ 푌 and the corresponding collection F of four cuts each of which has one of the quadrants as a shore. (b) Each edge of 퐶 ∪ 퐷 is in precisely two cuts in F and each edge in 푅 푘 is in 퐶 ∩ 퐷. (c) Deduce that the sum of the number of edges of each one of the four cuts in F is equal to 12 and conclude that each cut in F is a 3-cut. (d) Apply the result of Exercise 5.2.7 and conclude that each quadrant is a singleton. (ii) Let 푒 푘 := 푣 푘 푤 푘 be an edge of 푅 푘 . For each cut 퐶 in C, let 푋 be the shore of 퐶 that contains the vertex 푣 푘 . Let X be the collection of the 푘 − 1 shores thereby obtained. Prove that if 퐺 is not 퐾4 then there is an enumeration 푋1 , 푋2 , . . . , 푋 푘−1 of the shores in X such that 푋1 ⊂ 푋2 ⊂ · · · ⊂ 푋 푘−1 . Hint:use the laminarity of C. (iii) Suppose that 퐺 is not 퐾4 . Complete the proof of the Theorem. Hint: see the general structure of 퐺 in Figure 9.3.

192

9 Dependence Classes in Bricks

9.2 Mutually Exclusive Classes ♯♯ By definition, if 푄 is an equivalence class of the dependence relation on the edge set of a matching covered graph 퐺, every perfect matching of 퐺 which contains one edge of 푄 contains all edges of 푄. Two distinct equivalence classes 푄 1 and 푄 2 are mutually exclusive if no perfect matching of 퐺 contains all the edges of the set 푄 1 ∪ 푄 2 . In other words, 푄 1 and 푄 2 are mutually exclusive if any perfect matching of 퐺 which contains all edges of one of 푄 1 and 푄 2 contains none of the edges of the other. Two examples should serve to clarify the above definition. In the Petersen graph (Figure 9.4(a)), all equivalence classes are singletons, as each edge is removable. There is no perfect matching containing both 푒 and 푓 , but there is a perfect matching containing both 푒 and 푔 (Exercise 9.2.1). Therefore, {푒} and { 푓 } are mutually exclusive but {푒} and {푔} are not. The pentagonal prism (Figure 9.4(b)) has five removable edges; no two of these are mutually exclusive because there is a perfect matching containing all of them. But it also has five pairs of equivalence classes corresponding to the five removable doubletons. It is easy to see that the two equivalence classes {푒 1 , 푓1 } and {푒 2 , 푓2 } are mutually exclusive whereas {푒 1 , 푓1 } and {푒 3 , 푓3 } are not.

푒

푓

푓3

푓1 푒1

푔

푓2

푒3

푒2 (푎)

(푏)

Fig. 9.4 Mutually exclusive equivalence classes–examples

If some edge of 푄 1 is adjacent to some edge of 푄 2 then, obviously, 푄 1 and 푄 2 are mutually exclusive. As the example of the Petersen graph shows, this condition of adjacency is not necessary for two singleton equivalence classes to be mutually exclusive. However, if two doubleton equivalence classes 푄 1 and 푄 2 of a brick 퐺 are mutually exclusive then the edges of 푄 1 and 푄 2 lie in close proximity to each other in a special way. To make this precise, we need the definition given below. Aitch configuration Let 푄 1 := {푒 1 , 푓1 } and 푄 2 := {푒 2 , 푓2 } be two pairs of edges of a graph 퐺 such that 푄 1 ∩ 푄 2 = ∅. We say that 푄 1 and 푄 2 induce an aitch configuration in 퐺

9.2 Mutually Exclusive Classes

193

if 퐺 has a subgraph 퐻, where 푉 (퐻) is the set of ends of edges in 푄 1 ∪ 푄 2 , 퐸 (퐻) = 푄 1 ∪푄 2 ∪ {ℎ}, and ℎ is an edge not in 푄 1 ∪푄 2 , satisfying the following properties: (i) the edges 푒 1 and 푒 2 share precisely one common end, say 푢 12 , (ii) the edges 푓1 and 푓2 share precisely one common end, say 푣 12 , (iii) the vertices 푢 12 and 푣 12 both have degree three and are joined by edge ℎ. Figure 9.5 illustrates an aitch configuration. For specific examples, consider the pentagonal prism which is shown in Figure 9.4(b), and, for 푖 = 1, 2, 3, let 푄 푖 := {푒 푖 , 푓푖 }. Then 푄 1 and 푄 2 induce an aitch configuration, and so do 푄 2 and 푄 3 .

푒2

푓2 ℎ

푢12

푣12

푒1

푓1

Fig. 9.5 The doubletons {푒1 , 푓1 } and {푒2 , 푓2 } induce an aitch configuration

The brick 퐾4 induces an extreme form of an aitch graph, in which 푄 1 and 푄 2 are two distinct perfect matchings of 퐾4 . See Figure 9.6. ℎ 푢12

푣12 푒2

푓2

푒1

푓1

Fig. 9.6 An extreme form of aitch graph induced by 퐾4

Lemma 9.12 Let 푄 1 and 푄 2 be two (not necessarily removable) doubletons that are equivalence classes of the dependence relation of a simple brick 퐺. If 푄 1 and 푄 2 are mutually exclusive then they induce an aitch configuration in 퐺. Proof Let 푄 푖 := {푒 푖 := 푢 푖 푤 푖 , 푓푖 := 푣 푖 푥푖 }, 푖 = 1, 2, be two mutually exclusive doubletons of 퐺. By Theorem 9.1, 퐸 (퐺) − 푄 푖 is a cut of 퐺; it is the set of edges of a

194

9 Dependence Classes in Bricks

bipartite graph 퐻푖 ( 퐴푖 , 퐵푖 ), with | 퐴푖 | = |퐵푖 |. By Corollary 9.4, 푄 1 ∪ 푄 2 = 휕 ( 퐴1 Δ퐴2 ) is a 4-cut of 퐺. Adjust notation so that 푒 1 has both ends in 퐴1 , implying that 푓1 has both ends in 퐵1 . The ends of 푒 1 are 푢 1 and 푤 1 . One of them is in 퐴2 , the other in 퐵2 . Adjust notation so that 푢 1 ∈ 퐴2 , implying that 푤 1 ∈ 퐵2 . Likewise, note that the ends of 푓1 are 푣 1 and 푥1 , and adjust notation so that 푥1 ∈ 퐴2 and 푣 1 ∈ 퐵2 . Adopt similar notational adjustments for the ends of the two edges 푒 2 and 푓2 in 푄 2 (Figure 9.7). Any edge of 퐺 which does not belong to the cut 푄 1 ∪ 푄 2 belongs to both 휕 ( 퐴1 ) and 휕 ( 퐴2 ). Therefore, any such edge either joins a vertex in 퐴1 ∩ 퐴2 to a vertex in 퐵1 ∩ 퐵2 , or a vertex in 퐴1 ∩ 퐵2 to a vertex in 퐵1 ∩ 퐴2 , as indicated by dotted lines in Figure 9.7. Note that, up to this point, we have not used the hypothesis that 푄 1 and 푄 2 are mutually exclusive. In what follows, we shall use this hypothesis to show that 푄 1 ∪ 푄 2 is not a matching. (In other words, not all the eight ends of the four edges in 푄 1 ∪ 푄 2 are distinct as Figure 9.7 might suggest.) 푣2

푣1

퐵1 ∩ 퐵2

푓2

퐴1 ∩ 퐴2

푢2

푢1

푥1

푓1

푤2 퐵1 ∩ 퐴2

푒2

푒1

푤1

푥2

퐴1 ∩ 퐵2

Fig. 9.7 The cut 푄1 ∪ 푄2

9.12.1 The following equalities hold: | 퐴1 ∩ 퐴2 | = |퐵1 ∩ 퐵2 | and | 퐴1 ∩ 퐵2 | = |퐵1 ∩ 퐴2 |.

(9.3)

Moreover, the set 푄 1 ∪ 푄 2 is not a matching. Proof Let 푀1 be a perfect matching of 퐺 that contains 푒 1 . The doubleton 푄 1 is an equivalence class and therefore 푀1 contains 푓1 as well. The two doubletons 푄 1 and 푄 2 are mutually exclusive. Thus 푀1 and 푄 2 are disjoint. A simple counting argument then establishes the asserted equalities in (9.3). Now let 푋 and 푌 denote the two subsets of 푉 (퐺) as defined below:   푋 := ( 퐴1 ∩ 퐴2 ) ∪ (퐵1 ∩ 퐵2 ) − {푢 1 , 푢 2 , 푣 1 , 푣 2 },   푌 := ( 퐴1 ∩ 퐵2 ) ∪ (퐵1 ∩ 퐴2 ) − {푤 1 , 푤 2 , 푥1 , 푥2 }.

Assume, contrary to the second part of the assertion, that 푄 1 ∪ 푄 2 is a matching. Graph 퐺, being a brick, is bicritical. Therefore 퐺 − 푢 1 − 푢 2 has a perfect matching, say 퐹. As | 퐴1 ∩ 퐴2 | = |퐵1 ∩ 퐵2 |, the matching 퐹 must contain both 푓1 and 푓2 . Hence 퐹 includes a matching, say 푀, that saturates precisely the vertices of 푋. Likewise, 퐺 − 푥1 − 푤 2 has a perfect matching that contains both 푒 1 and 푓2 , whence it includes

9.2 Mutually Exclusive Classes

195

a matching, say 푁, that saturates precisely the vertices of 푌 . It now follows that 푀 ∪ 푁 ∪ 푄 1 ∪ 푄 2 is a perfect matching of 퐺 that includes 푄 1 ∪ 푄 2 . This is absurd because, by hypothesis, 푄 1 and 푄 2 are mutually exclusive. Hence 푄 1 ∪ 푄 2 cannot be a matching. ƒ 9.12.2 Suppose that the edge 푒 1 of 푄 1 is adjacent to the edge 푒 2 of 푄 2 . Then: • the two edges 푓1 of 푄 1 and 푓2 of 푄 2 are also adjacent, and • the common end 푢 1 = 푢 2 of 푒 1 and 푒 2 is adjacent to the common end 푣 1 = 푣 2 of 푓1 and 푓2 . Moreover, as illustrated in Figure 9.8, | 퐴1 ∩ 퐴2 | = 1 = |퐵1 ∩ 퐵2 |, 푁 (푢 1 ) = {푣 1 , 푤 1 , 푤 2 } and 푁 (푣 1 ) = {푢 1 , 푥1 , 푥2 }.

푣1 = 푣2

푓2

푢1 = 푢2

푥1

푓1

푤2 퐵1 ∩ 퐴2

푒2

푒1

푤1

푥2

퐴1 ∩ 퐵2

Fig. 9.8 Edges 푓1 and 푓2 adjacent

Proof Assume to the contrary that 푣 1 ≠ 푣 2 . In this case, since 푢 1 = 푢 2 by the hypothesis, and since | 퐴1 ∩ 퐴2 | = |퐵1 ∩ 퐵2 | by (9.12.1), it follows that 퐺 − 푣 1 − 푣 2 cannot have a perfect matching. This cannot be the case because 퐺 is bicritical. Therefore 푣 1 = 푣 2 must be the common end of 푓1 and 푓2 . If 퐴1 ∩ 퐴2 had any vertex other than 푢 1 , the pair {푢 1 , 푣 1 } would be a 2-vertex-cut of 퐺. This cannot be the case because 퐺, being a brick, is 3-connected. We conclude that 퐴1 ∩ 퐴2 = {푢 1 } and 퐵1 ∩ 퐵2 = {푣 1 }. The last part of the assertion, that 푢 1 and 푣 1 must be adjacent follows from the fact that in a brick every vertex has at least three distinct neighbours. ƒ Each of the doubletons 푄 1 and 푄 2 is a matching. Thus, as 푄 1 ∪ 푄 2 is not a matching, some edge of 푄 1 = {푒 1 , 푓1 } must be adjacent to some edge of 푄 2 = {푒 2 , 푓2 }. By renaming the edges of 푄 1 and 푄 2 , if necessary, we may assume that 푒 1 and 푒 2 are adjacent. Then, by the above result (9.12.2), the edges 푒 1 and 푒 2 share a common end 푢 12 := 푢 1 = 푢 2 , the edges 푓1 and 푓2 share a common end 푣 12 := 푣 1 = 푣 2 , and the vertices 푢 12 and 푣 12 both have degree three and are joined by an edge, say ℎ, which is not in 푄 1 ∪ 푄 2 . Indeed, 푄 1 and 푄 2 induce an aitch configuration. ƒ As an application of Lemma 9.12, we now derive the following consequence. Bricks with three mutually exclusive doubletons Theorem 9.13 If a simple brick 퐺 has three pairwise mutually exclusive doubletons then 퐺 is one of 퐾4 and 퐶6 .

196

9 Dependence Classes in Bricks

Proof Let 푄 푖 := {푒 푖 , 푓푖 }, 푖 = 1, 2, 3 be three mutually exclusive doubletons of a simple brick 퐺. By Lemma 9.12 the pair 푄 1 , 푄 2 induces an aitch configuration. Adjust notation so that the edges 푒 1 and 푒 2 share a common end labelled 푢 12 , and the edges 푓1 and 푓2 share a common end labelled 푣 12 . The vertices 푢 12 and 푣 12 have degree three and are adjacent in 퐺. Let ℎ denote the edge that joins 푢 12 and 푣 12 . See Figure 9.5. Case 1 The edge ℎ is in 푄 3 . Adjust notation so that ℎ = 푒 3 . Then, the vertex 푢 12 is the common end of the edges 푒 1 and 푒 3 . By Lemma 9.12, with 푄 3 playing the role of 푄 2 , the edge 푒 2 joins the common end 푢 12 of 푒 1 and 푒 3 to the common end of 푓1 and 푓3 , say 푣 13 . See Figure 9.9(a). Also, the vertex 푣 12 is the common end of 푓1 and 푒 3 . Thus, 푓2 joins 푣 12 to the common end of 푒 1 and 푓3 , say, 푢 13 . Moreover, the vertices 푢 13 and 푣 13 are adjacent and have degree three. We conclude that 퐺 is 퐾4 . See Figure 9.9(b).

푓2

푒3

푒3

푢12

푣12

푢12

푒2

푣12 푒2

푒1

푓1 푣13 푓3 (a)

푒1

푓2

푓1

푢13

푣13 푓3 (b)

Fig. 9.9 The case in which ℎ ∈ 푄3

Case 2 The edge ℎ is not in 푄 3 . By Lemma 9.12, with 푄 3 playing the role of 푄 2 , one of the edges of 푄 3 is adjacent to 푒 1 , the other is adjacent to 푓1 . Adjust notation so that 푒 3 is adjacent to 푒 1 and 푓3 is adjacent to 푓1 . As ℎ ∉ 푄 3 , the edges 푒 1 and 푒 3 share a common end, say 푢 13 , distinct from 푢 12 . Likewise, 푓1 and 푓3 share a common end, say 푣 13 , distinct from 푣 12 . Moreover, the vertices 푢 13 and 푣 13 are adjacent and have degree three. See Figure 9.10(a). Again, by Lemma 9.12, with 푄 3 playing the role of 푄 1 , we deduce that one of 푒 3 and 푓3 shares an end, say 푢 23 , with 푒 2 and the other shares an end, say 푣 23 , with 푓2 . Moreover, 푢 23 and 푣 23 are adjacent and have degree three. It follows that the six vertices 푢 12 , 푣 12 , 푢 13 , 푣 13 , 푢 23 , and 푣 23 are all the vertices of 퐺 and that 퐺 is a cubic

9.2 Mutually Exclusive Classes

197 푢23

푒2

푓2

푒2

푓2





푢12

푣12

푒3

푢12 푓3

푒1

푓1 푢13

푣23

푣13 (a)

푣12

푒3

푓3 푒1

푓1 푢13

푣13 (b)

Fig. 9.10 The case in which ℎ ∉ 푄3

graph. Also, as 퐺 is a brick, it is not bipartite. Therefore 푢 23 is not an end of 푓2 and 푣 23 is not an end of 푒 2 ; otherwise 퐺 would be the complete bipartite graph 퐾3,3 . We conclude that 푢 23 is an end of 푒 2 and 푣 23 is an end of 푓2 , and hence that 퐺 is 퐶6 . See Figure 9.10(b). ƒ Prisms of order 2푛, for each odd 푛 ≥ 3, M¨obius ladders of order 2푛, for each even 푛 ≥ 4, and staircases of order 2푛, for 푛 ≥ 4 are examples of bricks which have two, but not three, pairwise mutually exclusive removable doubletons. But one can construct other examples (Exercise 9.2.3).

9.2.1 Removable edges in bricks Theorem 9.13 proved above has the following interesting implication: A lower bound on the number of removable edges in bricks Theorem 9.14 Every simple brick 퐺 distinct from 퐾4 and 퐶6 has at least Δ(퐺) − 2 removable edges, where Δ(퐺) is the maximum of the degrees of vertices of 퐺. Proof Let 퐺 be a simple brick distinct from 퐾4 and 퐶6 . Let 푣 be a vertex of 퐺 of degree Δ(퐺). For each edge 푒 of 휕 (푣), let 푅(푒) denote a minimal class induced by 푒. Class 푅(푒) has at most two edges. Moreover, as 퐺 is a brick, it is 3-edge-connected. Thus, 푅(푒) is a removable class. It is easy to see that, for any two distinct edges 푒 and 푓 in 휕 (푣), the classes 푅(푒) and 푅( 푓 ) are mutually exclusive (Exercise 9.2.2). Thus, 퐺 has Δ(퐺) mutually exclusive removable classes. By hypothesis, 퐺 is neither 퐾4 nor 퐶6 , implying that at most two of the removable classes induced by edges in 휕 (푣)

198

9 Dependence Classes in Bricks

are doubletons. Thus, at least Δ(퐺) − 2 removable classes among those induced by ƒ the Δ edges in 휕 (푣) are singletons. We shall pursue in the next section a more detailed study of removable edges in near-bipartite bricks.

Exercises ⊲9.2.1 Referring to the Figure 9.4(a) of the Petersen graph, find a perfect matching containing 푒 and 푔, and show that there is no perfect matching containing both 푒 and 푓 . 9.2.2 For any vertex 푣 of a matching covered graph 퐺, show that any two minimal classes of 퐺 induced by different edges 푒 and 푓 in 휕 (푣) are mutually exclusive. ∗ 9.2.3 Find an example of a brick which is not a prism or a M¨obius ladder or a staircase which has two, but not three mutually exclusive doubletons. Staircases Theorem 9.15 Let 퐺 be a brick that has two removable, mutually exclusive doubletons 푅1 := {푒 1 , 푓1 } and 푅2 := {푒 2 , 푓2 }. For 푖 = 1, 2, let 푀푖 be a perfect matching of 퐺 that includes the edges of 푅푖 . If 푀1 and 푀2 are unique and disjoint then either 퐺 is 퐾4 or 퐺 is a staircase. ∗ 9.2.4 Prove Theorem 9.15, following the steps listed below. (i) Prove that 푅1 and 푅2 induce an aitch configuration of 퐺; adjust notation so that (i) 푒 1 and 푒 2 share a common end, 푢 12 , (ii) 푓1 and 푓2 share a common end, 푣 12 , and (iii) the vertices 푢 12 and 푣 12 both have degree three in 퐺 and are joined by an edge ℎ not in 푅1 ∪ 푅2 . (ii) Let 퐶 be an (푀1 , 푀2 )-alternating cycle in 퐺 that contains the edge 푒 1 . Prove that 퐶 is a hamiltonian cycle of 퐺, where ℎ is a chord of 퐶. Hint: use the fact that 푀1 and 푀2 are unique and disjoint. (iii) Let 푃 and 푄 denote the two subpaths of 퐶, both from 푢 12 to 푣 12 , denoted as follows: 푃 := 푢 12 푤 1 푤 2 · · · 푤 푠 푣 12 , 푠 ≥ 1

and 푄 := 푢 12 푥1 푥2 · · · 푥푡 푣 12 , 푡 ≥ 1.

Prove that every chord of 퐶 distinct from ℎ joins an internal vertex of 푃 to an internal vertex of 푄. Hint: use the uniqueness of 푀푖 and the fact that 퐺 − 푅푖 is bipartite, for 푖 = 1, 2. (iv) Prove that for 푖 = 1, 2, . . . , min{푠, 푡}, 푤 푖 and 푥푖 both have degree three and are adjacent. Hint: use induction on 푖. (v) Conclude that 푠 = 푡 and also that 퐺 is either 퐾4 or is a staircase.

9.3 Removable Edges in Near-Bipartite Bricks

199

Bricks with at most one removable edge Theorem 9.16 Let 퐺 be a brick that has at most one removable edge. Then, 퐺 is one of the following three bricks: 퐾4 , 퐶6 or H8 . ∗ 9.2.5 Using Theorem 9.15, prove Theorem 9.16, following the steps listed below. (i) Prove that 퐺 has a pair of mutually exclusive removable doubletons, 푅1 and 푅2 . Hint: for any vertex 푣, consider the minimal dependence classes induced by the edges of 휕 (푣). (ii) For 푖 = 1, 2, let 푀푖 be a perfect matching of the brick 퐺 that includes 푅푖 and let 퐻푖 be the (bipartite) graph 퐺 − 푅푖 . Prove that every removable edge of 퐻푖 which is not in 푀푖 is removable in 퐺. (iii) Prove that every edge of 푀푖 − 푅푖 is removable in 퐻푖 . Hint: suppose the contrary, let 푒 := 푣푤 be an edge in 푀푖 − 푅푖 which is not removable in 퐻푖 and apply Lemma 8.6 to vertices 푣 and 푤, to obtain two edges of 퐻푖 which are not in 푀푖 but are removable in 퐻푖 and then deduce a contradiction, using part (ii). (iv) Prove that 푀1 and 푀2 are unique. Hint: suppose that 푀푖 is not unique, let 퐶 be an 푀푖 -alternating cycle that does not contain any edge in 푅푖 , and use parts (iii) and (ii) to prove that every edge of 퐸 (퐶) − 푀푖 is removable in 퐺. (v) Prove that 푀1 and 푀2 are disjoint. Hint: suppose that 푀1 ∩푀2 contains an edge, 푒, and use part (iii) to prove that 푒 is a removable edge of 퐻1 and conclude that 퐺 has a perfect matching that contains both edges of 푅2 but does not contain 푒, contradicting the uniqueness of 푀2 . (vi) Apply Theorem 9.15 to prove that 퐺 is either 퐾4 or a staircase. (vii) Conclude that 퐺 has order eight or less and is thus 퐾4 , 퐶6 or H8 . 9.2.6 Prove the converse of Theorem 9.15.

9.3 Removable Edges in Near-Bipartite Bricks ♯♯ The bricks 퐾4 and 퐶6 have no removable edges. The bicorn H8 has only one removable edge. In fact, they are the only three bricks that have fewer than two removable edges (Exercise 9.2.5). The following theorem in fact shows that every simple nearbipartite brick distinct from these three bricks has two nonadjacent removable edges. The proof we present here is also from the unpublished paper [61] by Lov´asz and Vempala. Removable edges in near-bipartite bricks Theorem 9.17 Every simple near-bipartite brick distinct from 퐾4 , 퐶6 and the bicorn H8 has two nonadjacent removable edges.

200

9 Dependence Classes in Bricks

Proof Let 퐺 be a simple near-bipartite brick. We shall prove that either 퐺 has two nonadjacent removable edges or 퐺 is either 퐾4 , 퐶6 or the bicorn. By hypothesis, 퐺 is near-bipartite. Let 푅 := {푒, 푓 } denote a removable doubleton of 퐺 such that 퐻 := 퐺 − 푅 is bipartite. The next assertion is immediate. 9.17.1 Let 푀 be a perfect matching of 퐺 that contains both edges of 푅. Every edge of 퐸 − 푀 which is removable in 퐻 is also removable in 퐺. ƒ As 푅 is a dependence class, every perfect matching of 퐺 that contains one edge of 푅 contains both edges of 푅. Let 푀 be a perfect matching of 퐺 that contains both edges of 푅. We now consider several cases. In each case we assume that the previous cases do not apply. Case 1 The matching 푀 is not unique. The analysis of this case is quite similar to that used in the proof required in Exercise 9.2.5(iii). Assume that 퐺 has a perfect matching 푀 ′ , distinct from 푀, that contains the edges 푒 and 푓 . Let 퐶 be an (푀, 푀 ′ )-alternating cycle. Consider first the case in which every edge in 퐶 ∩ 푀 is removable in 퐻. As 퐺 is simple, the cycle 퐶 has length four or more; hence 퐶 ∩ 푀 contains two or more edges. The edges of 퐶 ∩ 푀 are not in 푀 ′ ; hence they are removable in 퐺, by statement 9.17.1. The assertion holds in this case, as the edges of 퐶 ∩ 푀 are not adjacent. We may thus assume that some edge of 퐶 ∩ 푀, say ℎ := 푣푤, is not removable in 퐻. Clearly, the edges 푒 and 푓 are not adjacent to edge ℎ. Thus, 푣 has degree three or more in the bipartite graph 퐻. By Lemma 8.6, 휕 (푣) contains an edge, say 푒 ′ , that is removable in 퐻. The edge 푒 ′ is not the edge ℎ, because ℎ is not removable in 퐻. Thus, 푒 ′ ∉ 푀. By statement 9.17.1, 푒 ′ is removable in 퐺. Likewise, 휕 (푤) − ℎ contains an edge, 푒 ′′ , which is removable in 퐺. As 퐺 is simple and since ℎ is neither 푒 ′ nor 푒 ′′ , we conclude that 푒 ′ and 푒 ′′ are two removable edges of 퐺 that are not adjacent. The assertion holds in this case. We may thus assume that the matching 푀 is unique.

(9.4)

Let ( 퐴, 퐵) denote the bipartition of 퐻. One of 푒 and 푓 has both ends in 퐴, the other has both ends in 퐵. Adjust notation so that 푒 has both ends in 퐴 and 푓 has both ends in 퐵. Let 푒 := 푢 1 푢 2 and 푓 := 푣 1 푣 2 . Consider the subgraph 퐻 ′ := 퐺 − {푢 1 , 푢 2 , 푣 1 , 푣 2 } obtained by deleting the ends of the two edges 푒 and 푓 . Now let 퐴′ := 퐴 − {푢 1 , 푢 2 }, let 퐵′ := 퐵 − {푣 1 , 푣 2 }, and let 푀 ′ := 푀 − {푒, 푓 }. The graph 퐻 ′ is bipartite; the pair ( 퐴′ , 퐵′ ) is a bipartition of 퐻 ′ . The set 푀 ′ is a perfect matching of 퐻 ′ . From equation 9.4, 푀 is the only perfect matching of 퐺 that contains the edges 푒 and 푓 . Thus, 푀 ′ is the only perfect matching of 퐻 ′ . This implies that there is an enumeration 푎 1 , 푎 2 , . . . , 푎 푘 of the vertices of 퐴′ and an enumeration 푏 1 , 푏 2 , . . . .푏 푘 of the vertices of 퐵′ such that (i) 푎 푖 푏 푖 ∈ 푀, for 푖 = 1, 2, . . . , 푘 and (ii) 푎 푖 푏 푗 ∉ 퐸 (퐻) for 1 ≤ 푖 < 푗 ≤ 푘 (Exercise 3.2.2). If 푘 = 0 then 퐺 has order four; hence 퐺 = 퐾4 . We may thus assume that 푘 ≥ 1. 9.17.2 푁 (푎 1 ) = {푣 1 , 푣 2 , 푏 1 } and 푁 (푏 푘 ) = {푢 1 , 푢 2 , 푎 푘 }. See Figure 9.11.

9.3 Removable Edges in Near-Bipartite Bricks 푣1

푓

푣2

201 푏1

푏 푘

...

푢1

푒

푢2

푎1

푎 푘

Fig. 9.11 The adjacencies of 푎1 and 푏 푘 .

Proof In 퐻 ′ , the vertex 푎 1 has degree one. Thus, in 퐺, 푎 1 is adjacent to both 푣 1 and 푣 2 ; hence 푁 (푎 1 ) = {푣 1 , 푣 2 , 푏 1 }. Likewise, 푁 (푏 푘 ) = {푢 1 , 푢 2 , 푎 푘 }. ƒ Case 2 푘 = 1. If 푘 = 1 then 퐺 has 퐶6 as a spanning subgraph. In that case, either 퐺 = 퐶6 or it has two nonadjacent removable edges. (Exercise 9.3.1.) We may thus assume that 푘 ≥ 2.

(9.5)

In 퐻 ′ , the vertex 푎 2 is adjacent to 푏 2 and possibly to 푏 1 , but it is not adjacent to any vertex in 퐵′ − {푏 1 , 푏 2 }. Thus, 푁 (푎 2 ) ⊆ {푏 1 , 푏 2 , 푣 1 , 푣 2 }. As 푎 2 has degree three or more, it is adjacent to a vertex in {푣 1 , 푣 2 }. Adjust notation so that 푎 2 and 푣 2 are adjacent. The vertices 푎 2 and 푏 2 are also adjacent. Thus, {푏 2 , 푣 2 } ⊂ 푁 (푎 2 ). It follows that 푁 (푎 2 ) − {푏 2 , 푣 2 } is a nonempty subset of {푏 1 , 푣 1 }. See Figure 9.12. 푣1

푓

푣2

푏1

푏2

푢1

푒

푢2

푎1

푎2

Fig. 9.12 The adjacencies of 푎1 and 푎2 . At least one of the two dashed lines represents an edge of 퐺.

Likewise the vertex 푏 푘−1 is adjacent to 푎 푘−1 and possibly to 푎 푘 , but it is not adjacent to any vertex in 퐴′ − {푎 푘−1 , 푎 푘 }. Thus, 푁 (푏 푘−1 ) ⊆ {푎 푘 , 푎 푘−1 , 푢 1 , 푢 2 }. As 푏 푘−1 has degree three or more, it is adjacent to a vertex in {푢 1 , 푢 2 }. Adjust notation so that 푏 푘−1 and 푢 2 are adjacent. The vertices 푏 푘−1 and 푎 푘−1 are also adjacent. Thus, {푎 푘−1 , 푢 2 } ⊂ 푁 (푏 푘−1 ). It follows that 푁 (푏 푘−1 ) − {푎 푘−1 , 푢 2 } is a nonempty subset of {푎 푘 , 푢 1 }. See Figure 9.13. Exercise 8.1.8 and statement 9.17.1 immediately imply the next assertion (Exercise 9.3.2).

9 Dependence Classes in Bricks

202 푣1

푓

푣2

푏 푘−1

푏 푘

푢1

푒

푢2

푎 푘−1

푎 푘

Fig. 9.13 The adjacencies of 푏 푘−1 and 푏 푘 . At least one of the two dashed lines represents an edge of 퐺.

9.17.3 Let 퐶 be a quadrilateral of 퐻 and let 푣 be a vertex of 퐶. If 푣 has degree three or more in 퐻 and the sets 휕 (푣) ∩ 퐸 (퐶) and 푀 are disjoint then at least one edge of 휕 (푣) ∩ 퐸 (퐶) is removable in 퐺. ƒ 9.17.4 If 푎 2 and 푣 1 are adjacent then one of the edges 푎 1 푣 1 and 푎 1 푣 2 is removable in 퐺. Proof Suppose that 푎 2 and 푣 1 are adjacent. Then 퐻 has the quadrilateral 퐶 := 푎 2 푣 1 푎 1 푣 2 푎 2 . Neither 푎 1 푣 1 nor 푎 1 푣 2 is in 푀; thus one of them is removable in 퐺, by statement 9.17.3. ƒ 9.17.5 If 푎 2 and 푏 1 are adjacent then one of the edges 푎 2 푏 1 and 푎 2 푣 2 is removable in 퐺. Proof Suppose that 푎 2 and 푏 1 are adjacent. Then 퐻 has the quadrilateral 퐶 := 푎 2 푏 1 푎 1 푣 2 푎 2 . Neither 푎 2 푏 1 nor 푎 2 푣 2 is in 푀, thus one of them is removable in 퐺, by statement 9.17.3. ƒ A reasoning similar to that used in the proofs of statements 9.17.4 and 9.17.5 may be used to prove the next two assertions (Exercise 9.3.3). 9.17.6 If 푏 푘−1 and 푢 1 are adjacent then one of the edges 푏 푘 푢 1 and 푏 푘 푢 2 is removable in 퐺. ƒ 9.17.7 If 푏 푘−1 and 푎 푘 are adjacent then one of the edges 푏 푘−1 푎 푘 and 푏 푘−1 푢 2 is removable in 퐺. ƒ Case 3 Either 푎 2 is adjacent to 푣 1 or 푏 푘−1 is adjacent to 푢 1 . Suppose that 푎 2 is adjacent to 푣 1 . From statement 9.17.4 we infer that one of the edges 푎 1 푣 1 and 푎 1 푣 2 , say, 푒 ′ , is removable in 퐺. We have seen that 푏 푘−1 is adjacent to a vertex in {푎 푘 , 푢 1 }. Thus, statements 9.17.6 and 9.17.7 imply that at least one of the edges 푏 푘 푢 1 , 푏 푘 푢 2 , 푏 푘−1 푎 푘 and 푏 푘−1 푢 2 , say, 푒 ′′ , is removable in 퐺. The edges 푒 ′ and 푒 ′′ are distinct and nonadjacent. Likewise, if 푏 푘−1 and 푢 1 are adjacent then one of the edges 푏 푘 푢 1 and 푏 푘 푢 2 , say, 푓 ′ , is removable in 퐺 and one of the edges 푎 1 푣 1 , 푎 1 푣 2 , 푎 2 푏 1 and 푎 2 푣 2 , say, 푓 ′′ , is removable in 퐺. The edges 푓 ′ and 푓 ′′ are distinct

9.3 Removable Edges in Near-Bipartite Bricks

203

and nonadjacent. In both alternatives, we concluded that 퐺 has two nonadjacent removable edges. The assertion holds in this case. We may thus assume that 푁 (푎 2 ) = {푏 1 , 푏 2 , 푣 2 }

and

푁 (푏 푘−1 ) = {푎 푘 , 푎 푘−1 , 푢 2 }.

(9.6)

Case 4 푘 ≥ 3.

From statement 9.17.5 we infer that one of the edges 푎 2 푏 1 and 푎 2 푣 2 , say, 푒 ′ , is removable in 퐺. Likewise, statement 9.17.7 implies that one of the edges 푏 푘−1 푎 푘 and 푏 푘−1 푢 2 , say, 푒 ′′ , is removable in 퐺. As 푘 ≥ 3, the edges 푒 ′ and 푒 ′′ are distinct and nonadjacent. We conclude that 퐺 has two nonadjacent removable edges. The assertion holds in this case. Case 5 푘 = 2. In this case, in view of statement 9.17.2 and equation (9.6), we infer that either 푢 1 푣 1 ∈ 퐸 or {푢 1 푣 2 , 푢 2 푣 1 } ⊂ 퐸. See Figure 9.14. 푣1

푓

푣2

푏1

푏2

푢1

푒

푢2

푎1

푎2

Fig. 9.14 The adjacencies of 푢1 , 푢2 , 푣1 and 푣2 . Either 푢1 푣1 ∈ 퐸 or {푢1 푣2 , 푢2 푣1 } ⊂ 퐸. Vertices 푢2 and 푣2 may also be adjacent.

Suppose that 푢 1 푣 1 ∉ 퐸. Then, 푢 1 푣 2 , 푢 2 푣 1 ∈ 퐸. In that case, 퐻 has two vertex-disjoint quadrilaterals, 퐶1 := 푎 1 푣 1 푢 2 푏 1 푎 1 and 퐶2 = 푎 2 푣 2 푢 1 푏 2 푎 2 . By statement 9.17.3, one of the edges 푢 2 푏 1 and 푢 2 푣 1 is removable in 퐺. Likewise, one of the edges 푣 2 푎 2 and 푣 2 푢 1 is removable in 퐺. (In fact, 푢 2 푏 1 and 푣 2 푎 2 are both removable in 퐺.) Alternatively, if 푢 1 푣 1 ∈ 퐸 then H8 is a spanning subgraph of 퐺 and the edge 푎 2 푏 1 is removable in H8 and in 퐺. Thus, either 퐺 is the bicorn or any edge of 퐺 − 퐸 (H8 ) is an edge that joins a vertex in {푢 1 , 푢 2 } to a vertex in {푣 1 , 푣 2 } and is thus removable in 퐺 and not adjacent to 푎 2 푏 1 . In all cases considered, we concluded that either 퐺 ∈ {퐾4 , 퐶6 , H8 } or 퐺 has two nonadjacent removable edges. ƒ We note that the conclusion of Theorem 9.17 does not hold without the hypothesis that the brick 퐺 is near-bipartite. There exist infinite families of bricks in which all removable edges are incident with a single vertex. One such family consists of the odd wheels 푊푛 , 푛 ≥ 5. A particular example is shown in Figure 9.15 (see Exercise 9.3.5). In Chapter 11 we shall present a construction of another infinite family of bricks which do not have nonadjacent removable edges.

204

9 Dependence Classes in Bricks 푣3 푣1 푣4 푣5

푣0

푣6 푣2 푣7 Fig. 9.15 A brick in which all removable edges are incident with one vertex.

9.3.1 Nonremovable edges in near-bipartite bricks We now consider near-bipartite bricks 퐺, with a removable doubleton 푅, in which an edge is removable in 퐺 − 푅 but is not removable in 퐺. Theorem 9.18 (CLM [8]) Let 퐺 be a brick and let 푅 := {푒, 푓 } be a removable doubleton of 퐺. Let ( 퐴, 퐵) denote the bipartition of 퐺 − 푅, where 푒 has both ends in 퐴 and 푓 has both ends in 퐵. Let ℎ be a removable edge of 퐺 − 푅. Edge ℎ is not removable in 퐺 if and only if 퐴 and 퐵 have partitions ( 퐴1 , 퐴2 ) and (퐵1 , 퐵2 ), as illustrated in Figure 9.16, such that: (i) | 퐴1 | = |퐵1 | + 1 (and |퐵2 | = | 퐴2 | + 1), (ii) edge 푒 has both ends in 퐴1 , (iii) edge 푓 has both ends in 퐵2 , and (iv) edge ℎ is the only edge of 퐺 that joins a vertex of 퐵1 to a vertex of 퐴2 . 푓 퐵1

퐵2

ℎ 퐴1

퐴2

푒 Fig. 9.16 Edge ℎ is removable in 퐺 − 푒 − 푓 , but not in 퐺.

Proof The existence of the asserted configuration easily implies that ℎ is not removable in 퐺. Conversely, suppose that ℎ is not removable in 퐺. Then, 푒 and 푓 are the

9.4 Notes

205

only edges of 퐺 that depend on ℎ (in addition to ℎ itself). Let 퐵2 denote a maximal barrier of 퐺 − ℎ that contains both ends of edge 푓 . The graph 퐺 − 푓 − ℎ is not bipartite. To see this, suppose the contrary. Then, both pairs {푒, 푓 } and { 푓 , ℎ} would be complements of cuts of 퐺; hence {푒, ℎ} = (퐸 − {푒, 푓 }) △ (퐸 − { 푓 , ℎ}) would be a cut of 퐺, a contradiction, as 퐺, a brick, is 3-connected. Thus, 퐽 := 퐺 − ℎ − 퐵2 has nontrivial components. Let 퐾 be a nontrivial component of 퐽. The maximality of 퐵2 implies that 퐾 is critical; thus, 퐾 is not bipartite. As 퐺 − 푅 is bipartite, it follows that 퐾 − 푒 is bipartite. We deduce that 푒 ∈ 퐸 (퐾). This conclusion implies that 퐾 is the only nontrivial component of 퐽. Let 퐴2 denote the set of isolated vertices of 퐽. Then, |퐵2 | = | 퐴2 | + 1. Let 퐴1 := 퐴 − 퐴2 and let 퐵1 := 퐵 − 퐵2 . Then, | 퐴1 | = |퐵1 | + 1. The edge 푒 has both ends in 퐾 and also both ends in 퐴. Thus, 푒 has both ends in 퐴1 . The edge ℎ has one end in 퐴 and one end in 퐵. But ℎ has at least one end in 퐴2 and no end in 퐵2 . We conclude that ℎ joins a vertex in 퐵1 to a vertex in 퐴2 . Finally, each edge of 휕퐺 (퐾) − ℎ is incident with a vertex in 퐵2 and has one end in 퐴 − 퐴2 ; ƒ hence each edge of 휕퐺 (퐾) − ℎ joins a vertex of 퐴1 to a vertex of 퐵2 .

Exercises ⊲9.3.1 In the proof of Theorem 9.17, in the case where 푀 is unique and 푘 = 1, prove that (i) 퐺 has 퐶6 as a spanning subgraph and (ii) if 퐶6 is a proper subgraph of 퐺 then 퐺 has two nonadjacent removable edges. 9.3.2 Prove statement 9.17.3. 9.3.3 Prove statements 9.17.6 and 9.17.7. ⊲9.3.4 Let 퐺 be a matching covered graph distinct from 퐾4 , 퐶6 and H8 , and such that 훿(퐺) ≥ 3. Prove the following properties: (i) if 퐺 is a near-brick then it has at least two removable edges, and (ii) if 퐺 is a simple near-bipartite graph then 퐺 has at least two nonadjacent removable edges. ⊲9.3.5 The graph shown in Figure 9.15 is a brick. Show that all the removable edges in it are incident with the vertex 푣 0 .

9.4 Notes By Theorem 9.13, every brick 퐺 which is distinct from 퐾4 and 퐶6 has a removable edge, that is, an edge 푒 such that 퐺 − 푒 is matching covered. This result by itself is not sufficiently strong to be useful as an inductive tool for establishing properties of bricks because the deletion of an edge from a brick may result in a matching covered

206

9 Dependence Classes in Bricks

graph with more than one brick. A removable edge 푒 in a brick 퐺 is 푏-invariant if 푏(퐺 − 푒) = 1. (Recall that a matching covered graph with one brick is called a near-brick. Most properties of bricks that are of interest to us in this theory are also shared by near-bricks.) Motivated by his work on the matching lattice, Lov´asz conjectured in 1987 that every brick distinct from 퐾4 , 퐶6 and the Petersen graph has a 푏-invariant edge. We shall present a proof of this result and several of its implications in a future chapter. One of the interesting implications is the fact that the dimension of the matching space Lin(퐺) of a matching covered graph 퐺 over the field GF(2) is 푚 − 푛 + 2 − (푏(퐺) + 푝(퐺)). (This formula is an example of an assertion about matching covered graphs in which the parameter 푏(퐺) + 푝(퐺) plays an essential role.) The lower bounds for the number of removable edges in a brick established in Corollary 9.7 and Theorem 9.14 are both obtained by just considering the minimal classes induced by the Δ edges incident with one vertex of maximum degree. We find it difficult to believe that such a ‘local argument’ could provide a tight bound for the number of removable edges in bricks of large orders! Conjecture 9.19 There exist a positive real number 푐 and an integer 푁 such that any brick 퐺 of order 푛 ≥ 푁 has at least 푐푛 removable edges. It would be of interest to characterize bricks in which all removable edges are incident with a single vertex.

Chapter 10

Removable Classes in Solid Bricks

Contents 10.1

10.2

10.3

Separating Cuts and Dulmage-Mendelsohn Barriers . . . . . . . . . . . . . 207 10.1.1 Nonremovable edges in bicritical graphs . . . . . . . . . . . . . . . 208 10.1.2 Removable edges in solid bricks . . . . . . . . . . . . . . . . . . . . . . 210 10.1.3 The Lemma on wheels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213 Some Implications of Being Solid . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217 10.2.1 Removable edges in solid bricks and separating cuts . . . . . 218 10.2.2 Propagation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220 10.2.3 푏-Invariance (monotonicity revisited) . . . . . . . . . . . . . . . . . . 220 10.2.4 The exchange property . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228

10.1 Separating Cuts and Dulmage-Mendelsohn Barriers Recall that a matching covered graph is solid if every separating cut in it is a tight cut and, in particular, that a brick is solid if it is free of nontrivial separating cuts. If a brick 퐺 is not solid, then it is easy to see that it has a nontrivial separating cut 퐶 such that at least one of the two 퐶-contractions of 퐺 is a solid brick (Exercise 10.1.4). Because of this, properties of bricks which are solid may sometimes be used to analyse properties of those which are not. It is this idea that first led CLM (2002, [9]) to the study of solid bricks. There was one particular result concerning removable edges in solid bricks (called the Lemma on Odd Wheels (2002, [10])) which served as an essential tool in our resolution of a conjecture of Lov´asz in the above cited papers. According to that result, if 퐺 is a simple solid brick in which all removable edges are incident with a vertex 푣 0 , then 퐺 is a wheel with 푣 0 as its hub. In this section, we shall derive a stronger version of that result (called here the Lemma on Wheels) from a general assertion related to nonremovable edges in bicritical graphs. The proof of this assertion makes crucial use of Dulmage-Mendelsohn barriers which were introduced in Chapter 5. © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7_10

207

208

10 Removable Classes in Solid Bricks

10.1.1 Nonremovable edges in bicritical graphs If an edge 푓 of a nontrivial matching covered graph 퐺 is not removable, by definition, there must be unmatchable edges in 퐺 − 푓 . Thus, 퐺 − 푓 must necessarily contain a barrier which includes both ends of some edge. All the known criteria for deciding whether or not a given edge of a matching covered graph is removable are based on the above observation. Let 퐵 denote a barrier of a matchable graph 퐺. In Section 3.2 we defined the core of 퐺 associated with 퐵 to be the bipartite graph H(퐵) obtained by deleting all edges with both ends in 퐵, deleting all vertices in the even components of 퐺 − 퐵, and then shrinking the vertex set of each odd component of 퐺 − 퐵 to a single vertex. In Section 5.1 we defined a barrier 퐵 to be a Dulmage-Mendelsohn barrier of 퐺 if • the core H(퐵) of 퐺 with respect to 퐵 is matching covered and • each odd component of 퐺 − 퐵 is critical. Nonremovable edges and Dulmage-Mendelsohn barriers Theorem 10.1 (CLM (2012, [16])) Let 퐺 be a bicritical graph of order at least four, and let 푓 be a nonremovable edge of 퐺. Then, 퐺 − 푓 contains a DulmageMendelsohn barrier 퐵 that satisfies the following properties: (i) edge 푓 has its ends in distinct odd components of 퐺 − 푓 − 퐵, and (ii) for each odd component 퐾 of 퐺 − 푓 − 퐵, the cut 퐶퐾 := 휕퐺 (퐾) is separating in 퐺. Proof Since 퐺 has order four or more and as 푓 is not removable, the graph 퐺 − 푓 has nonmatchable edges. Let 푒 be an edge of 퐺 − 푓 that depends on 푓 in 퐺. By Proposition 8.10, 퐺 − 푓 has barriers that contain both ends of 푒. Let 퐵★ be a maximal barrier of 퐺 − 푓 that contains both ends of 푒. By Lemma 5.4, 퐵★ includes a Dulmage-Mendelsohn barrier, say 퐵. Let H := H(퐵) denote the core of 퐺 associated with 퐵. By Proposition 8.11, 푓 has its ends in distinct odd components of 퐺 − 푓 − 퐵★. If 푓 has an end 푣 in an odd component of 퐺 − 푓 − 퐵★ which is not also an odd component of 퐺 − 푓 − 퐵, then 퐵 + 푣 would be a nontrivial barrier of 퐺, in contradiction to the hypothesis that 퐺 is bicritical. Thus, 푓 has its two ends in distinct odd components of 퐺 − 푓 − 퐵. This proves part (i). See Figure 10.1. The proof of part (ii) is less straightforward. We begin by establishing the notation which is convenient for explaining the steps involved in its proof: • 퐸 퐵 denotes the set of edges that have both ends in 퐵, • 퐶 denotes the set of edges that join a vertex in 퐵 to a vertex in some even component of 퐺 − 푓 − 퐵, • O denotes the set of odd components of 퐺 − 푓 − 퐵,

209

10.1 Separating Cuts and Dulmage-Mendelsohn Barriers 퐵★

퐵 replacements

− 퐵

퐶퐾 퐾 퐶 푓 Fig. 10.1 A DM barrier 퐵 in 퐺 − 푓

• for each 퐾 ∈ O, – 퐶퐾 denotes the cut 휕퐺 (퐾) of 퐺, – 퐺 퐾 denotes the 퐶퐾 -contraction 퐺/(푉 (퐾) → 푣 퐾 ) of 퐺, – 퐺 ′퐾 denotes the 퐶퐾 -contraction 퐺/(푉 (퐾) → 푣 퐾 ) of 퐺, and • L denotes the graph obtained from 퐺 by contracting the vertex set of each component in O to a single vertex. 10.1.1 The set 퐸 퐵 ∪ 퐶 is nonempty. Proof If 퐵 = 퐵★ then 푒, an edge having both ends in 퐵★, has both ends in 퐵. In that case, 푒 ∈ 퐸 퐵 . Alternatively, if 퐵 ⊂ 퐵★ then, as 퐺 is connected, the cut 퐶 is nonempty. ƒ In order to show that 퐶퐾 is a separating cut of 퐺 for each 퐾 ∈ O, we show that both 퐶퐾 -contractions of 퐺 are matching covered. The following lemma shows that 퐺 퐾 is bicritical, which clearly implies that it is matching covered. Lemma 10.2 For any component 퐾 in O, the graph 퐺 퐾 is bicritical. Proof Let 퐾 be any component in O. As 퐵 is a DM-barrier, by definition of DMbarrier we infer that 퐾 is critical. The vertex set of 퐺 퐾 is 푉 (퐾) + 푣 퐾 . So, for any vertex 푣 ∈ 푉 (퐾), the graph 퐺 퐾 − {푣, 푣 퐾 } has a perfect matching. Hence 푣 퐾 is not contained in any nontrivial barrier of 퐺 퐾 . On the other hand, any subset of 푉 (퐾) that is a nontrivial barrier of 퐺 퐾 is also a nontrivial barrier of 퐺. This is impossible because, by hypothesis, 퐺 is bicritical. We conclude that 퐺 퐾 is bicritical. ƒ For any particular 퐾 ∈ O, the 퐶퐾 -contraction 퐺 ′퐾 of 퐺 is obtained by shrinking 푉 (퐾) to a single vertex. The following lemma shows that L, which is obtained from 퐺 by shrinking each 푉 (퐾), for 퐾 ∈ O, to a single vertex, is matching covered. We shall see then that this, together with the above lemma, implies that 퐺 ′퐾 is matching covered. Lemma 10.3 The graph L is matching covered. Proof The graph L is obtained from 퐺 by the contraction of some sets of vertices. Thus, L is connected. Let us prove that every edge of L is matchable. Towards this end, let 푔 be an edge of L − 푓 , and let 푀푔 be a perfect matching of 퐺 that contains it.

10 Removable Classes in Solid Bricks

210

Case 1 푀푔 contains an edge in 퐸 퐵 ∪ 퐶. By Lemma 8.12, 2|푀푔 ∩ 퐸 퐵 | + |푀푔 ∩ 퐶| ≤ 2|푀푔 ∩ { 푓 }|,

(10.1)

with equality if and only if |푀푔 ∩ 휕퐺 (퐾)| = 1 for each component 퐾 in O. If 푀푔 contains an edge in 퐶, in turn an even cut, then in fact it contains two edges in 퐶. It follows that equality holds in (10.1); hence 푀푔 ∩ 퐸 (L) is a perfect matching of L that contains edges 푓 and 푔. Case 2 푀푔 and 퐸 퐵 ∪ 퐶 are disjoint. Let 퐽 be the graph L − 푉 (H). As 푀푔 and 퐶 are disjoint, the set 푀 퐽 := 푀푔 ∩ 퐸 (퐽) is a perfect matching of 퐽. Suppose that 푔 is an edge of 퐽. The core H is matching covered. Let 푀 be any perfect matching of H. The set 푀 ∪ 푀 퐽 is a perfect matching of L that contains the edge 푔. Alternatively, suppose that 푔 is not an edge of 퐽. As 푀푔 and 퐸 퐵 ∪ 퐶 are disjoint, we infer that 푔 is an edge of H. As H is matching covered, let 푁 be a perfect matching of H that contains the edge 푔. Then, 푁 ∪ 푀 퐽 is a perfect matching of L that contains the edge 푔. In all cases considered, every edge of L − 푓 is matchable in L. Moreover, as noted before, any perfect matching of L that contains an edge in the nonempty set 퐸 퐵 ∪ 퐶 also contains the edge 푓 . Every edge of the connected graph L is matchable, hence L is matching covered. ƒ Using Lemma 10.3, we now proceed to prove that the 퐶퐾 -contraction 퐺 ′퐾 of 퐺 is matching covered for any component 퐾 ∈ O. For every 퐾 ∈ O, the graph 퐺 퐾 is matching covered, by Lemma 10.2. Thus, splicing L with 퐺 퐽 , for each 퐽 ∈ O − 퐾, yields the matching covered graph 퐺 ′퐾 = 퐺/(푉 (퐾) → 푣 퐾 ). The other 퐶퐾 -contraction 퐺 퐾 = 퐺/(푉 (퐾) → 푣 퐾 ) is, as noted earlier, also a matching covered graph. We conclude that 퐶퐾 is a separating cut. As this is true for any 퐾 ∈ O, it establishes the validity of part (ii) of the assertion of the theorem. ƒ We note that part (ii) of the above theorem is not valid in general for components of 퐺 − 푓 − 퐵∗ . In Figure 10.2, we give an example of a bicritical graph 퐺, a nonremovable edge 푓 of 퐺, a maximal barrier 퐵★ of 퐺 − 푓 , a component 퐾 of 퐺 − 푓 − 퐵★ and the unique perfect matching of 퐺 that contains edge 푔. One of the 퐶퐾 -contractions of 퐺 is 퐾4 , up to multiple edges, but the other 퐶퐾 -contraction 퐺/(푉 (퐾) → 푘) of 퐺 is not matching covered, and hence 퐶퐾 is not a separating cut of 퐺 (Exercise 10.1.1).

10.1.2 Removable edges in solid bricks We shall now use Theorem 10.1 to derive useful results concerning removable edges in solid bricks. We begin with a simple consequence of that theorem, with edge 푒 playing the role of 푓 . (We leave its proof as Exercise 10.1.5.)

10.1 Separating Cuts and Dulmage-Mendelsohn Barriers

211 푒

퐵★ 푔

푓

퐾

Fig. 10.2 A nonremovable edge in a bicritical graph

Corollary 10.4 Let 퐺 be a solid brick, and let 푒 be a nonremovable edge of 퐺. Then, 퐺 − 푒 has a Dulmage-Mendelsohn barrier 퐵 such that every odd component of 퐺 − 푒 − 퐵 is trivial and edge 푒 has its ends in distinct odd components of 퐺 − 푒 − 퐵. ƒ Theorem 10.5 Let 퐺 be a solid brick and let 푣 be a vertex of 퐺. Suppose that 푣푣 1 and 푣푣 2 are distinct nonremovable edges of 퐺. Then, 푣 has exactly three neighbours, 푣 1 , 푣 2 and 푣 3 . Moreover, for 푖 = 1, 2, there exists an equipartition (퐵푖 , 퐼푖 ) of 푉 (퐺) such that (i) 푣푣 푖 is the only edge of 퐺 that has both ends in 퐼푖 , (ii) every edge that has both ends in 퐵푖 is incident with 푣 3 , (iii) the subgraph 퐻푖 of 퐺, obtained by the removal of 푒 푖 and each edge having both ends in 퐵푖 , is matching covered and bipartite, with bipartition {퐵푖 , 퐼푖 }, and (iv) 퐵1 = 퐼2 − 푣 + 푣 3 and 퐵2 = 퐼1 − 푣 + 푣 3 . (See Figure 10.3 for an illustration.) Proof Let 푑 be the degree of 푣. Enumerate the 푑 edges of 휕 (푣) as 푒 1 , 푒 2 , . . . , 푒 푑 , where 푒 푖 joins 푣 to 푣 푖 , for 푖 = 1, 2, . . . , 푑. By hypothesis, neither 푒 1 nor 푒 2 is removable in 퐺. By Corollary 10.4, for 푖 = 1, 2, graph 퐺 − 푒 푖 has a DulmageMendelsohn barrier 퐵푖 such that (i) each odd component of 퐺 − 푒 푖 − 퐵푖 is trivial and (ii) edge 푒 푖 has its ends in distinct odd components of 퐺 − 푒 푖 − 퐵푖 . The following result will be used in this proof. (Exercise 10.1.3). Lemma 10.6 Let 퐺 be a bicritical graph, let 푒 1 and 푒 2 be two adjacent edges of 퐺 and let 퐵푖 be a barrier of 퐺 − 푒 푖 , for 푖 = 1, 2. Then, |퐵1 ∩ 퐵2 | ≤ 1. ƒ Let us now prove that 푣 1 , 푣 2 and 푣 3 are the only three neighbours of 푣. Let us also prove that 퐵1 ∩ 퐵2 = {푣 3 }. As 푒 푖 has both ends in odd components of 퐺 − 푒 푖 − 퐵푖 , and since each such odd component is trivial, it follows that 푣 is the vertex of a trivial component of 퐺 − 푒 푖 − 퐵푖 . Clearly, 푒 푖 is the only edge of 퐺 that has its ends in distinct components of 퐺 − 푒 푖 − 퐵푖 . It follows that {푣 3 , 푣 4 , . . . , 푣 푑 } is a subset of 퐵1 ∩ 퐵2 . By Lemma 10.6, 퐵1 ∩ 퐵2 is a singleton. We deduce that 푣 has only three neighbours and 퐵1 ∩ 퐵2 = {푣 3 }.

10 Removable Classes in Solid Bricks

212

푣1

푣6

푣11

푣

푣5

푣3

푣2 푣10

푣7

푣4

푣8

푣9

(a) 퐺 푣

푣1

푣5

푣7

푣9

푣11 퐼1

푣3

푣2

푣4

푣6

푣8

푣10

푣9

푣11

퐵1

(b) 퐵1 and 퐼1

푣3

푣1

푣5

푣7

퐵2

푣

푣2

푣4

푣6

푣8

푣10

퐼2

(c) 퐵2 and 퐼2

Fig. 10.3 The solid brick 퐺 and the pairs 퐵1 , 퐼1 and 퐵2 , 퐼2

For 푖 = 1, 2, if 푥 is a vertex in an even component of 퐺 − 푒 푖 − 퐵푖 , then 퐵푖 + 푥 is a barrier of 퐺 − 푒 푖 . Using this fact, we now proceed to show that 퐺 − 푒 푖 − 퐵푖 has no even components. Suppose that 퐺 − 푒 1 − 퐵1 has even components and that 푥푦 is an edge of such a component. If 푥 is in an even component of 퐺 − 푒 2 − 퐵2 , then 퐵1 + 푥 and 퐵2 + 푥 are barriers of 퐺 − 푒 1 and 퐺 − 푒 2 , respectively, and (퐵1 + 푥) ∩ (퐵2 + 푥) = {푣 3 , 푥}. Similarly, if 푥 is in 퐵2 , 퐵1 + 푥 and 퐵2 are barriers of 퐺 − 푒 1 and 퐺 − 푒 2 , respectively, and (퐵1 + 푥) ∩ 퐵2 = {푣 3 , 푥}. In either case, we have a contradiction to Lemma 10.6.

10.1 Separating Cuts and Dulmage-Mendelsohn Barriers

213

It follows that 푥 is neither in an even component of 퐺 − 푒 2 − 퐵2 , nor in 퐵2 ; hence 푥 is in an odd component of 퐺 − 푒 2 − 퐵2 . The odd components of 퐺 − 푒 2 − 퐵2 are trivial; hence 푥 is the only vertex of an odd component of 퐺 − 푒 2 − 퐵2 . Likewise, 푦 is also the only vertex of an odd component of 퐺 − 푒 2 − 퐵2 . As the vertices 푥 and 푦 are adjacent, we derive a contradiction. We conclude that 퐺 − 푒 1 − 퐵1 and, similarly, 퐺 − 푒 2 − 퐵2 , have no even components. For 푖 = 1, 2, now let 퐼푖 := 푉 (퐺) − 퐵푖 . Then, 퐼푖 is the set of isolated vertices of 퐺 − 푒 푖 − 퐵푖 and edge 푒 푖 is the only edge of 퐺 that has both ends in 퐼푖 . Moreover, 퐻푖 is the core of 퐺 associated with 퐵푖 , in turn a Dulmage-Mendelsohn barrier; therefore 퐻푖 is matching covered. As 퐵1 ∩ 퐵2 = {푣 3 }, and since 푣 lies in 퐼1 ∩ 퐼2 , it follows that 퐵1 = 퐼2 − 푣 + 푣 3 and 퐵2 = 퐼1 − 푣 + 푣 3 . As 푒 1 is the only edge of 퐺 having both ends in 퐼1 , it follows that every edge having both ends in 퐵2 is incident with 푣 3 . Likewise, every edge having both ends in 퐵1 is incident with 푣 3 . ƒ We leave the proofs of the following two corollaries, proved in [16], as Exercise 10.1.6. Corollary 10.7 If 퐺 is a solid brick with six vertices or more then every vertex of 퐺 ƒ is incident with at most two nonremovable edges of 퐺. Corollary 10.8 If 퐺 is a solid brick of maximum degree three or four, then, for every vertex 푣 of 퐺 at most one edge incident with 푣 does not lie in a removable class of 퐺. ƒ

10.1.3 The Lemma on wheels For a fixed vertex 푣 0 of a graph 퐺, a subset 푀 of the edges of 퐺 is a 푣 0 -matching if |푀 ∩ 휕 (푣)| = 1 for each vertex 푣 distinct from 푣 0 , and if |푀 ∩ 휕 (푣 0 )| > 1. We say that 퐺 is a 푣 0 -wheel if 퐺 is a wheel having 푣 0 as a hub. The following is easily verified using a counting argument: Proposition 10.9 If a matching covered graph 퐺 has a 푣 0 -matching for some vertex 푣 0 then 퐺 is not bipartite. ƒ Lemma on wheels Lemma 10.10 Let 푣 0 be a vertex of a simple solid brick 퐺, and let 푀0 be a 푣 0 -matching of 퐺. Then either 퐺 is a 푣 0 -wheel or 퐺 has two removable edges 푒 1 , 푒 2 ∉ 푀0 ∪ 휕 (푣 0 ). Proof Let 퐻 := 퐺 − 푣 0 . For each vertex 푣 of 퐻, let 푑 퐻 (푣) denote the degree of 푣 in 퐻. As 퐺 is a brick, every vertex of 퐺 has degree three or more. Thus, for every vertex 푣 of 퐻, 푑 퐻 (푣) ≥ 2, with equality only if 푣 is adjacent to 푣 0 .

214

10 Removable Classes in Solid Bricks

Case 1 The graph 퐻 is 2-regular. Each component of 퐻 is then a cycle. As 퐺 is matching covered, it is 2-connected; hence 퐻 is connected. It follows that 퐻 is a cycle. Every vertex of 퐻 is adjacent to 푣 0 ; hence 퐺 is a 푣 0 -wheel. 10.10.1 Let 푣 be a vertex of 퐻 such that 푑 퐻 (푣) ≥ 3. Then, at most one edge of 휕퐻 (푣) − 푀0 is not removable in 퐺. Proof Let 푑 be the degree of 푣 in 퐺 and let 푒 1 , 푒 2 , . . . , 푒 푑 be the edges of 휕퐺 (푣), where each 푒 푖 joins 푣 to a vertex 푣 푖 . Assume, to the contrary, that 휕퐻 (푣) − 푀0 contains two edges, say 푒 1 and 푒 2 , that are not removable in 퐺. As 퐺 is simple, from Theorem 10.5 we infer that 푑 = 3 and 퐺 has an equipartition (퐵1 , 퐼1 ) such that 푒 1 is the only edge of 퐺 having both ends in 퐼1 and every edge of 퐺 that has both ends in 퐵1 is incident with 푣 3 . Moreover, the graph 퐻1 , obtained from 퐺 by removing 푒 1 and each edge having both ends in 퐵1 , is matching covered. The edges 푒 1 and 푒 2 are not in 푀0 ; hence 푒 3 is the edge of 푀0 incident with 푣. The edge 푒 3 is an edge of 퐻; hence 푣 3 ≠ 푣 0 . It follows that 푣푣 3 is the only edge of 푀0 incident with 푣 3 . Thus, no edge of 푀0 has both ends in 퐵1 . We conclude that 푀0 is a 푣 0 -matching of the bipartite matching covered graph 퐻1 , in contradiction to ƒ Proposition 10.9. Case 2 The graph 퐻 has a vertex 푣 such that 푑 퐻 (푣) ≥ 4. The set 휕퐻 (푣) − 푀0 contains at least three edges, of which at least two are removable in 퐺, by statement 10.10.1. The assertion holds in this case. We have seen that 푑 퐻 (푣) ≥ 2, for each vertex 푣 of 퐻. We may assume that the previous cases do not apply; therefore 푑 퐻 (푣) ∈ {2, 3} for each vertex 푣 of 퐻. Moreover, at least one vertex 푣 has degree three in 퐻. Case 3 The graph 퐻 has a vertex 푣 such that 푑 퐻 (푣) = 3. We shall now prove that 퐺 has two removable edges which are not in 푀0 ∪ 휕 (푣 0). We have seen that 푑 퐻 (푢) ∈ {2, 3}, for each vertex 푢 in 퐻. The number of vertices of odd degree in 퐻 is even. Thus, 퐻 has a vertex 푣 ′ distinct from 푣 such that 푑 퐻 (푣 ′ ) = 3. At least two edges of 휕퐻 (푣) are not in 푀0 . By statement 10.10.1, 휕퐻 (푣) − 푀0 contains an edge, 푒, which is removable in 퐺. Likewise, 휕 (푣 ′ ) − 푀0 contains an edge, 푒 ′ , which is removable in 퐺. If 푒 and 푒 ′ are distinct then the assertion holds. We may thus assume that 푒 = 푒 ′ . If 퐻 has a third vertex, 푣 ′′ , such that 푑 퐻 (푣 ′′ ) = 3 then again 휕퐻 (푣 ′′ ) − 푀0 has an edge, 푒 ′′ , which is removable in 퐺, and distinct from 푒. In that case, 퐺 has at least two removable edges not in 푀0 ∪ 휕 (푣 0). The assertion holds in every case considered, except the case in which 푣 and 푣 ′ are the only two vertices of 퐻 that have degree three in 퐻; all the other vertices of 퐻 have degree two. Moreover, 푣 and 푣 ′ are joined by an edge, 푒 = 푒 ′ , which is removable in 퐺.

10.1 Separating Cuts and Dulmage-Mendelsohn Barriers

215

As 퐺 − 푒 is matching covered, it is 2-connected; hence 퐻 − 푒 is connected. Every vertex of 퐻 − 푒 has degree two. We deduce that 퐻 − 푒 is a cycle, say 푄 := 푣 1 푣 2 · · · 푣 푠 푣 1 ,

푠 ≥ 3,

푠 is odd.

Adjust notation so that 푣 = 푣 1 , let 푡 be such that 푣 ′ = 푣 푡 . We may consider the reverse of 푄, if necessary, so that 푡 is odd. Thus, 푠 and 푡 are both odd. As 퐺 is simple, the vertices 푣 and 푣 ′ are not adjacent in 퐻 − 푒. Thus, 3 ≤ 푡 ≤ 푠 − 2, hence 푠 ≥ 5. Figure 10.4 illustrates the case in which 푡 = 5 and 푠 = 9. 푒 푋 푣 ′

푣 퐶

퐺

퐺/푋

퐺/푋

Fig. 10.4 The case 푡 = 5 and 푠 = 9. Dashed lines indicate the possibility of an edge.

Let 푋 := {푣 1 , 푣 2 , . . . , 푣 푡 } and let 퐶 := 휕 ( 푋). As 3 ≤ 푡 ≤ 푠 − 2, the cut 퐶 is nontrivial. We now show that 퐶 is separating, therefore obtaining a contradiction to the hypothesis that 퐺 is solid. The underlying simple graph of the 퐶-contraction 퐺/푋 is clearly a wheel of order 푠 − 푡 + 2 ≥ 4. The underlying simple graph of the 퐶-contraction 퐺/푋 is also a wheel of order 푡 + 1 ≥ 4. We deduce that 퐺 is not solid, a contradiction. ƒ Figure 10.5 illustrates the necessity for 퐺 to be simple.

푣0

Fig. 10.5 The brick 퐺 must be simple – heavy lines indicate the edges of 푀0

10 Removable Classes in Solid Bricks

216

Corollary 10.11 Let 퐺 be a solid brick and let 푣 0 be a vertex of 퐺. Either 퐺 is a wheel having 푣 0 as a hub, up to multiple spokes, or 퐺 has two removable edges not incident with 푣 0 . Proof Suppose that 퐺 is simple. Let 푣 0 푤 and 푣 0 푥 be two edges incident with 푣 0 . The graph 퐺, a brick, is bicritical. Thus, 퐺 − 푤 − 푥 has a perfect matching, 푀. The set 푀0 := 푀 ∪ {푣 0 푤, 푣 0 푥} is a 푣 0 -matching of 퐺. By Lemma 10.10, either 퐺 is a 푣 0 -wheel or 퐺 has two removable edges not incident with 푣 0 . The assertion holds if 퐺 is simple. We may thus assume that 퐺 has multiple edges. If 퐺 has a pair of parallel edges not incident with 푣 0 then each one of these edges is removable in 퐺 and not incident with 푣 0 . In that case, the assertion holds. We may thus assume that every pair of parallel edges of 퐺 is incident with 푣 0 . Let 퐻 be the underlying simple graph of 퐺. Clearly, 퐻 is a simple solid brick. We have seen that the assertion holds for simple solid bricks. Thus, either 퐻 is a 푣 0 -wheel or 퐻 has two removable edges not incident with 푣 0 . If 퐻 is a 푣 0 -wheel then 퐺 is a wheel having 푣 0 as a hub, up to multiple spokes. Alternatively, 퐻 has two removable edges not incident with 푣 0 . These two edges are removable in 퐺. The assertion holds in both alternatives. ƒ

Exercises ⊲10.1.1 Show that the cut 퐶퐾 := 휕 (퐾) is not a separating cut in the bicritical graph shown in Figure 10.2. ⊲10.1.2 Verify that the graph 퐺 in Figure 10.3 is solid. Hint: see Example 7.2 and Exercise 7.1.6 to prove that 퐺 is odd-intercyclic. ⊲10.1.3 Prove Lemma 10.6. Hint: for any two vertices 푤 and 푥 of 퐵푖 , the graph 퐺 − 푤 − 푥 has a perfect matching that contains the edge 푒 푖 . ⊲10.1.4 Show that every nonsolid brick 퐺 has a nontrivial separating cut 퐶 such that at least one of the two 퐶-contractions of 퐺 is a solid brick. ⊲10.1.5 Give a proof of Corollary 10.4. ∗ 10.1.6 Give proofs of Corollaries 10.7 and 10.8. Wheel-like bricks A graph 퐺 is a wheel-like-brick if 퐺 is a brick and has a vertex ℎ, called its hub, such that every removable class of 퐺 has an edge in 휕 (ℎ). By Corollary 10.11 every solid wheel-like brick having ℎ as a hub is an odd ℎ-wheel. Example 10.12 shows a wheel-like brick which is not solid. We do not know any characterization of wheel-like bricks.

10.2 Some Implications of Being Solid

217

Example 10.12 Figure 10.6 depicts an 푥-wheel 퐺 1 , an 푥-wheel 퐺 2 and the graph 퐺, which is a splicing of 퐺 1 at 푥 with 퐺 2 at vertex 푥. The graph 퐺 is a wheel-like brick having ℎ as its hub. Thus, the converse of Corollary 10.11 does not hold in general.

3

1

4 ℎ

푥

푥

5 6

2

7 (a) 퐺1

(b) 퐺2 1

3 4



5 6

2

7 (c) 퐺

Fig. 10.6 The wheel-like brick 퐺.

⊲10.1.7 Prove that the graph 퐺 depicted in Figure 10.6 is a wheel-like brick having ℎ as its hub.

10.2 Some Implications of Being Solid All bipartite matching covered graphs are solid, but not all solid matching covered graphs are bipartite. However, solid matching covered graphs have many properties which are akin to those of bipartite graphs. For example, perfect matching polytopes of solid bricks, as those of bipartite graphs, afford linear inequality descriptions which do not involve odd set constraints (see Theorem 7.5). In this section we shall establish three other properties of solid bricks which exemplify this phenomenon. The proofs of the first two of these three theorems will follow from the assertion in the next section, whose proof, in turn, makes use of three results:

218

10 Removable Classes in Solid Bricks

• Separating cuts with a bipartite shore. Theorem 4.7 asserts that if 퐶 := 휕 ( 푋) is a separating cut of a matching covered graph 퐺 and the subgraph 퐺 [푋] of 퐺 induced by 푋 is bipartite then 퐶 is a barrier cut of 퐺; hence 퐶 is tight. • The precedence relation defined in Section 7.2. Recall that a cut 퐶 of a matching covered graph 퐺 is said to precede another cut 퐷 of 퐺 if |푀 ∩ 퐶| ≤ |푀 ∩ 퐷|, for each perfect matching 푀 of 퐺. If, in addition, |푁 ∩ 퐶| < |푁 ∩ 퐷|, for some perfect matching 푁 of 퐺, then 퐶 strictly precedes 퐷. We write 퐶  퐷 to indicate that 퐶 precedes 퐷, and 퐶 ≺ 퐷 to indicate that 퐶 strictly precedes 퐷. • Theorem 4.2. According to that theorem, a cut 퐶 of a matching covered graph 퐺 is a separating cut of 퐺 if and only if, for any edge 푒 of 퐺, there exists a perfect matching 푀푒 of 퐺 such that 푒 ∈ 푀푒 and |푀푒 ∩ 퐶| = 1.

10.2.1 Removable edges in solid bricks and separating cuts Lemma 10.13 Let 퐺 be a solid brick, let 푒 be a removable edge of 퐺 and let 퐷 be a cut of 퐺. If 퐷 − 푒 is a separating cut in 퐺 − 푒 then (precisely) one of the (퐷 − 푒)-contractions of 퐺 − 푒 is bipartite. Proof Let us first prove that at least one (퐷 − 푒)-contraction of 퐺 − 푒 is nonbipartite. According to Theorem 9.8, if 푒 is any removable edge of a matching covered graph 퐺, then 푏(퐺 − 푒) ≥ 푏(퐺). In our case, as 퐺 is a brick, we must have that 푏(퐺 − 푒) ≥ 1. By Theorem 4.20, at least one of the (퐷 − 푒)-contractions of 퐺 − 푒 is nonbipartite. Having established that at least one of the (퐷 − 푒)-contractions of 퐺 − 푒 is not bipartite, assume, to the contrary, that both (퐷 − 푒)-contractions of 퐺 − 푒 are nonbipartite. Let C denote the collection of those odd cuts 퐶 of 퐺 that have the following properties: (i) both (퐶 − 푒)-contractions of 퐺 − 푒 are nonbipartite, and (ii) 퐶  퐷. 10.13.1 For every cut 퐶 in C, the cut 퐶 − 푒 is separating in 퐺 − 푒. Proof By hypothesis, 퐷 − 푒 is separating in 퐺 − 푒. Let 푓 be any edge of 퐺 − 푒. By Theorem 4.2, 퐺 − 푒 has a perfect matching 푀 푓 such that 푓 ∈ 푀 푓 and |푀 푓 ∩ 퐷| = 1. As 퐶  퐷, and since 퐶 is odd, we deduce that |푀 푓 ∩ 퐶| = 1. This conclusion holds ƒ for each edge 푓 of 퐺 − 푒. By Theorem 4.2, 퐶 − 푒 is separating in 퐺 − 푒. Let 퐶 := 휕 ( 푋) be a minimal cut in C, relative to the precedence relation. Let 퐺 1 := 퐺/( 푋 → 푥) and 퐺 2 := 퐺/( 푋 → 푥) be the two 퐶-contractions of 퐺. As both (퐶 − 푒)-contractions of 퐺 − 푒 are nonbipartite, the cut 퐶 of 퐺 is nontrivial. As 퐺 is a solid brick by hypothesis, the cut 퐶 cannot be a separating cut of 퐺. Thus the edge 푒 must be unmatchable in one of the two 퐶-contractions of 퐺. (All other edges are matchable in both 퐶-contractions of 퐺 because 퐶 − 푒 is separating in 퐺 − 푒.) Assume without loss of generality that the edge 푒 is unmatchable in 퐺 1 . By Corollary 2.2, there must exist a maximal barrier of 퐺 1 , say 퐵, which contains both ends of 푒.

10.2 Some Implications of Being Solid

219

Since 퐺 is a brick, 퐵 cannot be a barrier of 퐺 itself. This means that the contraction vertex 푥 (which may or may not be an end of the edge 푒 in 퐺 1 ) is in 퐵. Let O denote the set of all odd components of 퐺 1 − 퐵 and, for 퐾 ∈ O, let 퐶퐾 denote the cut 휕 (퐾). See Figure 10.7. 퐶

퐶퐾

푥

퐾

푒

퐵

퐶퐽

퐽

Fig. 10.7 Illustration for Lemma 10.13: 퐵 is a maximal barrier of 퐺1 which includes both ends of the edge 푒.

It can be verified by simple counting arguments that, for any perfect matching 푀 of 퐺, the following identity holds: Õ |푀 ∩ 퐶퐾 | = |푀 ∩ 퐶| + (|O| − 1) − 2|푀 ∩ {푒}|. (10.2) 퐾 ∈ O

The above identity has the following implications: 10.13.2 For each 퐾 ∈ O, the cut 퐶퐾 strictly precedes 퐶. Proof Let 퐾 be a component in O. From (10.2) we infer that |푀 ∩ 퐶퐾 | ≤ |푀 ∩ 퐶|, for each perfect matching 푀 of 퐺. Moreover, equality holds only if 푒 ∉ 푀. Thus, for any perfect matching 푀푒 of 퐺 that contains the edge 푒, |푀푒 ∩ 퐶퐾 | < |푀푒 ∩ 퐶|. We conclude that 퐶퐾 ≺ 퐶. This conclusion holds for each 퐾 ∈ O. ƒ 10.13.3 There exists a nontrivial component 퐽 ∈ O such that both (퐶 퐽 − 푒)contractions of 퐺 − 푒 are nonbipartite. Proof Let us first observe that, by hypothesis, 퐺 1 − 푒 is not bipartite; hence 퐺 1 − 퐵 has a nontrivial component, 퐽. The maximality of 퐵 implies that 퐽 is critical, hence nonbipartite. Thus, the 퐶 퐽 -contraction (퐺 − 푒)/푉 (퐽) is nonbipartite. By hypothesis, the (퐶 − 푒)-contraction 퐺 2 − 푒 of 퐺 − 푒 is nonbipartite. By Lemma 4.6, the shore 푋 of 퐶 induces a nonbipartite graph in 퐺 − 푒. Thus, the 퐶 퐽 -contraction (퐺 − 푒)/푉 (퐽) is nonbipartite. ƒ By statement 10.13.2, 퐶 퐽 ≺ 퐶  퐷, and by statement 10.13.3, both (퐶 퐽 − 푒)contractions of 퐺 − 푒 are nonbipartite. This is in contradiction to the criteria used for choosing 퐶. Thus, (precisely) one of the (퐷 −푒)-contractions of 퐺 −푒 is nonbipartite. ƒ

10 Removable Classes in Solid Bricks

220

10.2.2 Propagation If 퐺 is any bipartite matching covered graph and 푒 is a removable edge of 퐺 then, clearly, 퐺 − 푒 is also a bipartite matching covered graph. Analogously, a solid matching covered graph bequeaths the property of being solid to any graph obtained from it by the deletion of a removable class. We first consider the case which involves the deletion of removable edges of a brick. Property of being solid is hereditary: deletions of removable edges of a solid brick Theorem 10.14 (CLM (2005, [13])) Every graph obtained from a solid brick by the deletion of a removable edge is also solid. Proof Let 푒 be a removable edge of a solid brick 퐺. Let 퐶 be a cut of 퐺 such that 퐶 − 푒 is a separating cut of 퐺 − 푒. By Lemma 10.13, one of the (퐶 − 푒)-contractions of 퐺 − 푒 is bipartite. By Theorem 4.7, 퐶 − 푒 is tight in 퐺 − 푒. This conclusion holds for each cut 퐶 of 퐺 such that 퐶 − 푒 is separating in 퐺 − 푒. We deduce that 퐺 − 푒 is solid. ƒ The following result can be proved fairly easily by induction on the number of edges of 퐺. We leave it to the reader as Exercise 10.2.2. Property of being solid is hereditary: deletions of removable classes Theorem 10.15 If 퐺 is a solid matching covered graph and 푅 is a removable class of 퐺, then 퐺 − 푅 is solid. ƒ Theorem 10.15 may also be deduced from the result stated in Exercise 10.2.6.

10.2.3 풃-Invariance (monotonicity revisited) According to Theorem 9.8, if 푒 is any removable edge of a matching covered graph 퐺, then 푏(퐺 − 푒) ≥ 푏(퐺). Recall that if equality holds, that is, if 푏(퐺 − 푒) = 푏(퐺), then we refer to 푒 as a 푏-invariant edge. (See the Notes Section of Chapter 9 for an account of the origin of this notion.) Every brick distinct from 퐾4 and 퐶6 has removable edges, but not every removable edge in a such a brick need be 푏-invariant. In fact, given any positive integer 푘, one can find an example of a brick 퐺 and a removable edge 푒 of 퐺 such that 푏(퐺 − 푒) = 푘 (see Exercise 9.1.3). However, a striking property of solid matching covered graphs is that every removable edge in such a graph is 푏-invariant. This can be deduced from the following assertion concerning solid bricks.

10.2 Some Implications of Being Solid

221

푏-invariance of removable edges in solid bricks Theorem 10.16 (CLM (2002, [9])) In a solid brick, every removable edge is 푏-invariant. Proof Let 퐺 be a solid brick, and let 푒 be a removable edge of 퐺. Assume that 푏(퐺 − 푒) ≥ 2. Since 푒 is a removable edge of 퐺, the graph 퐺 − 푒 is matching covered, and since 푏(퐺 − 푒) ≥ 2, the graph 퐺 − 푒 must have a nontrivial tight cut such that both cut contractions with respect to it are nonbipartite matching covered graphs. This is a contradiction to Lemma 10.13. ƒ The following result can be proved fairly easily by induction on the number of edges of 퐺. We leave it to the reader as Exercise 10.2.3. Theorem 10.17 Let 퐺 be a solid matching covered graph. For each removable edge 푒 of 퐺, 푏(퐺 − 푒) = 푏(퐺) and (푏 + 푝) (퐺 − 푒) = (푏 + 푝) (퐺). ƒ Corollary 10.18 Let 퐺 be a solid matching covered graph and let 푅 be a nonempty set of edges of 퐺 such that 퐺 − 푅 is matching covered. Then, 푏(퐺 − 푅) ≥ 푏(퐺) + 1 − |푅|, with equality if and only if the set 푅 is a removable class of 퐺. Proof by induction on |푅|. If 푅 is a removable class of 퐺 then equality holds, by Theorems 10.17 and 9.9. Suppose then that 푅 is not a removable class of 퐺. As 퐺 − 푅 is matching covered, the set 푅 includes a removable class, 푆, of 퐺. Let 퐻 := 퐺 − 푆 and let 푇 := 푅 − 푆. The graph 퐻 − 푇 is equal to the graph 퐺 − 푅, in turn a matching covered graph. Thus, the set 푇 is a nonempty subset of 퐸 (퐻) such that 퐻 − 푇 is matching covered. Moreover, by Theorem 10.15, the graph 퐻 is solid. By induction, 푏(퐺 − 푅) = 푏(퐻 − 푇) ≥ 푏(퐻) + 1 − |푇 | = 푏(퐺) + 2 − |푆| − |푇 | = 푏(퐺) + 2 − |푅|. The asserted inequality is strict if 푅 is not a removable class.

ƒ

10.2.4 The exchange property Let 퐺 be a matching covered graph, let 푅 be a removable class of 퐺 and let 푆 be a removable class of 퐺 − 푅. The pair (푅, 푆) has the exchange property if 퐺 has a removable class 푇 and 퐺 − 푇 has a removable class 푊 such that 푇 ⊆ 푆 and 퐺 − 푇 − 푊 = 퐺 − 푅 − 푆.

10 Removable Classes in Solid Bricks

222

A matching covered graph 퐺 is said to have the exchange property with respect to deletions of removable classes if, for any given removable class 푅 of 퐺, and any removable class 푆 of 퐺 − 푅, the pair (푅, 푆) has the exchange property. A particular case of the exchange property holds for bipartite graphs (Exercise 8.1.9): if 퐺 is a bipartite matching covered graph, if an edge 푒 is removable in 퐺 and an edge 푓 is removable in 퐺 − 푒 then 푓 is removable in 퐺 itself (and 푒 is removable in 퐺 − 푓 ). We shall prove (Theorem 10.23) that, more generally, all solid matching covered graphs have the exchange property. Example 10.19 Figure 10.8(a) depicts the graph 퐺 := Σ8 , and 푅 := {푒} is a removable class of 퐺; the graph 퐺 − 푒 is shown in Figure 10.8(b). The pair 푆 := { 푓1 , 푓2 } is a removable doubleton of 퐺 − 푒. Figure 10.8(c) depicts the graph 퐺 − 푒 − 푆. The set 푇 := { 푓1 } is a removable class of 퐺. The graph 퐺 − 푓1 is shown in Figure 10.8(d). The set 푊 := {푒, 푓2 }, which is equal to 푅 ∪ (푆 − 푇), is a removable doubleton of 퐺 − 푓1 . Indeed, (푅, 푆) has the exchange property. 푒

푓1

푓1

푓2

푓2

(a) 퐺 푒

(b) 퐺 − 푒

푓2

(c) 퐺 − 푒 − 푆

(d) 퐺 − 푓1

Fig. 10.8 Illustration for Example 10.19

Every removable class of a matching covered graph has at most two edges, by Corollary 8.18. This fundamental property implies the next assertion. Proposition 10.20 Let 푅 be a removable class of a matching covered graph 퐺 and let 푆 be a removable class of 퐺 − 푅. Suppose that (푅, 푆) has the exchange property, let 푇 ⊆ 푆 be a removable class of 퐺 and let 푊 := 푅 ∪ (푆 − 푇) be a removable class of 퐺 − 푇. The number of doubletons in {푅, 푆} is equal to the number of doubletons in {푇, 푊 }. Proof Each one of 푅 and 푆 has either one or two edges. Thus, the number of doubletons in {푅, 푆} is equal to |푅| + |푆| − 2. Likewise, the number of doubletons

10.2 Some Implications of Being Solid

223

in {푇, 푊 } is equal to |푊 | + |푇 | − 2. Moreover, |푅| + |푆| = |푊 | + |푇 |. The assertion ƒ holds. Let us now make a comment about requiring that 푊 be a removable class, in the above definition of exchange property. One might ask that if 푇 is a removable class of 퐺 then 퐺 −푇 −푊 = 퐺 − 푅 − 푆, and 퐺 − 푅 − 푆 is matching covered. The problem is that 푊 may not be a removable class of 퐺 −푇. This is illustrated in the next example. Example 10.21 Consider the graph 퐺 depicted in Figure 10.9. The edge 푒 is removable in 퐺 and the doubleton { 푓1 , 푓2 } is removable in 퐺 − 푒. The edge 푓1 is removable in 퐺 and the edge 푓2 is removable in 퐺 − 푓1 and, finally, the edge 푒 is removable in 퐺 − 푓1 − 푓2 . Thus, ({푒}, { 푓1 , 푓2 }) does not have the exchange property, because 퐺 − 푓1 − {푒, 푓2 } is matching covered, but {푒, 푓2 } is not a removable doubleton of 퐺 − 푓1 . 푒

푓1

푓2

Fig. 10.9 Illustration for Example 10.21

The graph in Example 10.21 is not solid. In fact, it is not possible to exhibit a solid graph which does not have the exchange property, as the next assertion shows. Proposition 10.22 Let 퐺 be a solid matching covered graph, let 푅 be a removable class of 퐺 and let 푆 be a removable class of 퐺 − 푅. If 푆 includes a removable class of 퐺 then (푅, 푆) has the exchange property. Proof Assume that 푆 includes a removable class, 푇, of 퐺, let 퐻 := 퐺 − 푇 and let 푊 := 푅 ∪ (푆 −푇). The graph 퐻 −푊 is equal to the graph 퐺 − 푅 − 푆, in turn a matching covered graph. Thus, 푊 is a nonempty subset of 퐸 (퐻) such that 퐻 − 푊 is matching covered. By Theorem 10.15, the graphs 퐺 − 푅 and 퐻 are solid. By Corollary 10.18, 푏(퐻 − 푊) = 푏(퐺 − 푅 − 푆) = 푏(퐺 − 푅) + 1 − |푆| = 푏(퐺) + 2 − |푅| − |푆| = [푏(퐻) − 1 + |푇 |] + 2 − |푅| − |푆| = 푏(퐻) + 1 − |푊 |. Thus, 푊 is a removable class of 퐻. We deduce that (푅, 푆) has the exchange property.ƒ Exchange property of solid matching covered graphs Theorem 10.23 Every solid matching covered graph has the exchange property with respect to deletions of removable classes.

10 Removable Classes in Solid Bricks

224

Proof Let 퐺 be a solid matching covered graph, let 푅 be a removable class of 퐺 and let 푆 be a removable class of 퐺 − 푅. We shall prove that (푅, 푆) has the exchange property. In view of Proposition 10.22, it suffices to prove that 푆 includes a removable class of 퐺. We shall prove this by induction on |푉 |. The analysis of the basis of the induction hypothesis is done if 퐺 is free of nontrivial tight cuts. Case 1 The set 푆 is a singleton, { 푓 }. 10.24 No edge of 푅 depends in 퐺 on edge 푓 . Proof Assume, to the contrary, that 푅 contains an edge, 푒, that depends on 푓 in 퐺. The graph 퐺 − 푅 − 푆, which is matching covered, is a spanning subgraph of 퐺 − 푓 ; hence 퐺 − 푓 is matchable. By Proposition 8.10, 퐺 − 푓 has barriers that contain both ends of 푒. Let 퐵 denote a maximal barrier of 퐺 − 푓 containing both ends of 푒. Let 퐸 퐵 denote the set of edges that contain both ends in 퐵. By Proposition 8.11, the edge 푓 joins vertices in two different odd components of 퐺 − 푓 − 퐵. By Theorem 3.2, the maximality of 퐵 implies that each component of 퐺 − 푓 − 퐵 is odd. Let O denote the set of all (odd) components of 퐺 − 푓 − 퐵, and for each 퐾 ∈ O let 퐶퐾 := 휕 (퐾). See Figure 10.10. 푒 퐵

퐶퐾 퐾 푓 Fig. 10.10 The edge 푒 of 푅 depends on edge 푓 of 푆.

10.24.1 For each 퐾 ∈ O, the cut 퐶퐾 is separating in 퐺. Proof By hypothesis, the graph 퐺 − 푅 − 푓 is matching covered. Hence every edge of 퐺 − 푅 − 푓 is matchable in 퐺 − 푓 . Thus, 퐵 is stable in 퐺 − 푅 − 푓 . The edge 푓 has both ends in distinct components in O. Thus, 퐸 퐵 ⊆ 푅. Let 푀 be a perfect matching of 퐺. The set [∪퐾 ∈ O (푀 ∩ 퐶퐾 )] − (푀 ∩ { 푓 }) is equal to 푀 ∩ 휕 (퐵). Moreover, |퐵| = |O|. Thus, Õ |푀 ∩ 퐶퐾 | − 2|푀 ∩ { 푓 }| = |O| − 2|푀 ∩ 퐸 퐵 |. (10.3) 퐾 ∈ O

To prove the assertion, let ℎ be an edge of 퐺; let us prove that 퐺 has a perfect matching that contains edge ℎ and just one edge in 퐶퐾 , for each 퐾 ∈ O. For this, it suffices to prove that ℎ is in a perfect matching 푀 of 퐺 such that

10.2 Some Implications of Being Solid

Õ

퐾 ∈ O

225

|푀 ∩ 퐶퐾 | = |O|.

(10.4)

Consider first the case in which ℎ is an edge of 퐺 − 푅 − 푓 . The graph 퐺 − 푅 − 푓 is matching covered; let 푀 be a perfect matching of 퐺 − 푅 − 푓 that contains the edge ℎ. As 퐸 퐵 ⊆ 푅, we conclude that 푀 and 퐸 퐵 are disjoint. From equation (10.3) we deduce the validity of equality (10.4). Suppose then that ℎ ∈ 푅 + 푓 . Let 푀푒 be a perfect matching of 퐺 that contains the edge 푒. As 푒 depends on 푓 , 푓 ∈ 푀푒 . Moreover, 푒 ∈ 퐸 퐵 . From equation (10.3) we infer that 퐸 퐵 = {푒} and also the validity of equality (10.4). As 푒 ∈ 푅, and since 푅 is a removable class of 퐺 that contains edge 푒, it follows that 푅 ⊆ 푀푒 . In sum, 푅 + 푓 ⊆ 푀푒 . Indeed, every edge of 퐺 is in a perfect matching 푀 that satisfies equality (10.4).ƒ As 퐺 − 푅 is matching covered, the edge 푓 is in a perfect matching, 푀 푓 , of 퐺 − 푅. As 퐸 퐵 = {푒}, no edge of 푀 푓 has both ends in 퐸 퐵 . From equation (10.3) we conclude that 푀 푓 has precisely three edges in 퐶 퐿 , for some 퐿 ∈ O. Thus, 퐶 퐿 is separating but ƒ not tight in 퐺. This is not possible, because 퐺 is solid. The graph 퐺 − 푅 − 푓 is matching covered. Thus, each edge of 퐺 − 푅 − 푓 is matchable in 퐺 − 푓 . Let 푒 be an edge of 푅. We have seen that 푒 does not depend on 푓 ; hence 퐺 has a perfect matching, 푁 푒 , that contains the edge 푒 but does not contain the edge 푓 . As 푅 is an equivalence class of the dependence relation of the edges in 퐺, 푅 ⊂ 푁 푒 . Thus, every edge of 푅 is matchable in 퐺 − 푓 . In sum, every edge of 퐺 − 푓 is matchable in 퐺 − 푓 . We conclude 푓 is removable in 퐺. Indeed, the only edge 푓 of 푆 is removable in 퐺. We may thus assume that 푆 is a doubleton. Case 2 The graph 퐺 is free of nontrivial tight cuts. From Corollary 10.18 we deduce that 푏(퐺) = 푏(퐺 − 푅) − 1 + |푅|. By Theorem 9.9, 푏(퐺 − 푅 − 푆) = 푏(퐺 − 푅) − 1. Moreover, as 퐺 is free of nontrivial tight cuts, 푏(퐺) ≤ 1. Thus, 1 ≥ 푏(퐺) = 푏(퐺 − 푅) − 1 + |푅| = 푏(퐺 − 푅 − 푆) + |푅| ≥ |푅| ≥ 1. Equality holds throughout, hence 퐺 is a brick, 푅 is a singleton and 퐺 − 푅 − 푆 is bipartite. Let 푒 be the edge of 푅, let 푓1 and 푓2 be the two edges of 푆 and let ( 퐴, 퐵) denote the bipartition of 퐺 − 푅 − 푆. The bipartite graph 퐺 − 푅 − 푆 is matching covered. Addition to 퐺 − 푅 − 푆 of an edge having one end in 퐴 and the other in 퐵 yields a matching covered graph, by Corollary 3.7. As 푆 is a removable doubleton of 퐺 − 푅, no edge of 푆 has one end in 퐴, the other in 퐵. For if, say, 푓1 has an end in 퐴 and the other in 퐵 then the edge 푓2 is removable in 퐺 − 푅, a contradiction. As 퐺 − 푅 is matching covered, one of 푓1 and 푓2 has both ends in 퐴 and the other has its ends in 퐵. If 푒 has one end in 퐴 and the other in 퐵, the graph 퐺 − 푆 is matching covered. We may thus assume that 푒 has both ends in the same part of ( 퐴, 퐵). Adjust notation so that 푓1 and 푒 have both ends in 퐴, whereas 푓2 has both ends in 퐵. A simple counting argument

226

10 Removable Classes in Solid Bricks

then shows that every perfect matching of 퐺 that contains the edge 푒 contains also the edge 푓2 , but not the edge 푓1 . As 퐺 − 푒 − 푆 is matching covered, we deduce that every edge of 퐺 − 푓1 is matchable in 퐺 − 푓1 . Thus, 퐺 − 푓1 is matching covered. In both alternatives, 푆 includes a removable class of 퐺. Case 3 The graph 퐺 has a nontrivial tight cut, 퐶 := 휕 ( 푋). Let 퐺 1 := 퐺/( 푋 → 푥) and let 퐺 2 := 퐺/( 푋 → 푥) be the two 퐶-contractions of 퐺. As 퐺 is solid, each brick of 퐺 is solid, by Theorem 7.1. Each brick of 퐺 1 is a brick of 퐺, hence solid. By Theorem 7.1, 퐺 1 is solid. Likewise, 퐺 2 is solid. For 푖 = 1, 2, let 푅푖 := 푅 ∩ 퐸 (퐺 푖 ) and let 푆푖 := 푆 ∩ 퐸 (퐺 푖 ). The cut 퐶 − 푅 is tight in 퐺 − 푅, because every perfect matching of 퐺 − 푅 is a perfect matching of 퐺 and 퐶 − 푅 is odd in 퐺 − 푅. The graph 퐺 푖 − 푅푖 is a (퐶 − 푅)-contraction of 퐺 − 푅; hence 퐺 푖 − 푅푖 is matching covered. By Lemma 8.15, (푖) 푅푖 is either empty or a removable class of 퐺 푖 , (ii) 푅 has at most one edge in 퐶 and (iii) if 푅 and 퐶 are disjoint then 푅 is entirely contained in one of 퐸 (퐺 1 ) − 퐶 and 퐸 (퐺 2 ) − 퐶. Likewise, the cut 퐶 − 푅 − 푆 is tight in 퐺 − 푅 − 푆; hence 퐺 푖 − 푅푖 − 푆푖 , a (퐶 − 푅 − 푆)-contraction of 퐺 − 푅 − 푆, is matching covered. By Lemma 8.15, (푖) 푆푖 is either empty or is a removable class of 퐺 푖 − 푅푖 , (ii) 푆 has at most one edge in 퐶 and (iii) if 푆 and 퐶 are disjoint then 푆 is entirely contained in one of 퐸 (퐺 1 ) − 퐶 and 퐸 (퐺 2 ) − 퐶. The set 푆 has at most two edges; hence it does not have edges in both 퐸 (퐺 1 ) − 퐶 and 퐸 (퐺 2 ) − 퐶. Adjust notation so that 푆 and 퐸 (퐺 2 ) − 퐶 are disjoint. If 푅1 is empty then 푆 is a removable class of 퐺 1 . Alternatively, if 푅1 is not empty then 푅1 is a removable class of 퐺 1 and 푆 is a removable class of 퐺 1 − 푅1 . As 퐺 1 is solid, we infer, by induction, that 푆 includes a removable class of 퐺 1 . In both alternatives, 푆 includes a removable class, 푇, of 퐺 1 . If 푇 and 퐶 are disjoint then 퐺 −푇 is the splicing of 퐺 1 −푇 and 퐺 2 , both matching covered graphs; hence 퐺−푇 is matching covered and therefore 푆 includes a removable class of 퐺. Let us thus assume that 푇 and 퐶 are not disjoint. The doubleton 푆 contains at most one edge in 퐶 and 푇 ⊆ 푆. Then, 푇 ∩퐶 is a singleton, { 푓 }. Note that 푆2 = { 푓 }. If 푅2 is empty then 푓 is a removable edge of 퐺 2 . Alternatively, if 푅2 is not empty then 푅2 is a removable class of 퐺 2 and 푓 is a removable edge of 퐺 2 − 푅2 . As 퐺 2 is solid, we infer, by induction, that 푓 is removable in 퐺 2 . In both alternatives, 푓 is a removable edge of 퐺 2 . Thus, the graph 퐺 − 푇 is matching covered, because it is the splicing of 퐺 1 − 푇 and 퐺 2 − 푓 , both matching covered graphs. The assertion holds by induction, if 퐺 has nontrivial tight cuts. ƒ There are essentially two reasons that justify the fact that a nonsolid graph 퐺 may not have the exchange property. One reason is given in Example 10.21. A second reason is the fact that given a removable class 푅 of 퐺, the graph 퐺 − 푅 may have a removable class which does not include any removable class of 퐺. This is illustrated by the tricorn H10 (Exercise 10.2.7). There are, however, simple nonsolid graphs that have the exchange property; see Exercise 10.2.8.

10.2 Some Implications of Being Solid

227

Exercises 10.2.1 (i) Show that every near-bipartite graph is a near-brick. (ii) Deduce from Theorem 9.17 that every near-bipartite brick distinct from 퐾4 , 퐶6 and H8 has at least two nonadjacent 푏-invariant removable edges. ∗ 10.2.2 Give a proof of Theorem 10.15. In fact, prove that if 퐺 is a solid matching covered graph and 푅 is a set of edges of 퐺 such that 퐺 − 푅 is matching covered, then 퐺 − 푅 is solid. ∗ 10.2.3 Give a proof of Theorem 10.17. ∗ 10.2.4 Let 푅 be a removable class in a matching covered graph 퐺, and suppose that 퐷 is a cut in 퐺 such that 퐷 − 푅 is a separating cut in 퐺 − 푅. Now let C denote the family of all cuts 퐶 in 퐺 satisfying the following properties: (i) 퐶  퐷, (ii) 퐶 − 푅 is a separating cut in 퐺 − 푅, (iii) 휆퐺−푅 (퐶 − 푅) ≤ 휆퐺−푅 (퐷 − 푅), (iv) subject to (i), (ii) and (iii), 퐶 is minimal. Show that every member of C is a separating cut of 퐺. ∗ 10.2.5 Let 퐺 be a solid matching covered graph, let 푅 be a removable class of 퐺 and let 푆 be a removable class of 퐺 − 푅 such that 퐺 − 푆 is matching covered. The objective of this exercise is to prove that 푆 is a removable class of 퐺. This is certainly the case if 푆 is a singleton. So we may suppose that 푆 is a doubleton. We then prove that 푆 is removable in 퐺 by proving that 푏(퐺 − 푆) ≤ 푏(퐺) − 1 and then applying Theorems 9.8 and 9.9. In fact, these two theorems are used several times in the proof. (i) Prove that in the particular case in which 푅 is a singleton, {푒}, then 푏(퐺 − 푆) ≤ 푏(퐺 − 푒 − 푆) = 푏(퐺 − 푒) − 1 = 푏(퐺) − 1. Hint: use the fact that 퐺 is solid and apply Theorem 10.17. (ii) Prove that in the particular case in which 푅 is a doubleton, 푏(퐺 − 푆) ≤ 푏(퐺 − 푅 − 푆) + 1 = 푏(퐺 − 푅) = 푏(퐺) − 1. Monotonicity of the function 휆 [9] ∗ 10.2.6 Let 푅 be a removable class in a matching covered graph 퐺. (i) Using the result in Exercise 10.2.4, show that 휆(퐺) ≤ 휆(퐺 − 푅). (ii) Deduce that if 퐺 is solid, then so is 퐺 − 푅.

228

10 Removable Classes in Solid Bricks

⊲10.2.7 Define a 푘-corn, 푘 ≥ 3, 푘 odd, to be the graph obtained from the odd wheel 푊 푘 by splicing each vertex in the rim with a 퐾4 . The 3-corn is the tricorn. Prove that the collection 푅 푘 of removable classes of the 푘-corn consists of 푘 removable edges and no removable doubletons. Prove that, for 푘 ≥ 5, removal of any edge in 푅 푘 yields a graph with 푘 removable edges. Verify that in the case of the 3-corn, the removal of any edge of 푅3 produces a graph with 5 removable edges. Conclude that the 푘-corn does not have the exchange property. ⊲10.2.8 Prove that every prism of order 푛 ≡ 2 (mod 4), 푛 ≥ 10, has the exchange property. ⊲10.2.9 Let 퐺 be a near-bipartite brick, let 푅 be a removable doubleton of 퐺 and let ℎ be a removable edge of 퐺 − 푅. Suppose that ℎ is not removable in 퐺. By Theorem 10.23, 퐺 is not solid. Apply Theorem 9.18 to determine a nontrivial separating cut of 퐺.

10.3 Notes Using Theorem 10.23, it can be shown that any solid matching covered graph 퐺, with 훿 ≥ 3, has at least 푚 − 푛 + 1 removable classes. Proving such a result by induction on the number of edges poses difficulties because a graph obtained from 퐺 by the deletion of removable classes may have vertices of degree two and may not have any removable classes at all. Thus, the theory of ear decompositions, which is the subject matter of the next chapter, turns out to be the appropriate context for proving such results by induction. (Ear decompositions of bipartite graphs were introduced in Section 3.4.) This theory incorporates all the ideas related to removable classes, but is unfettered by constraints on degrees!

Chapter 11

Ear Decompositions

Contents 11.1

11.2

11.3

11.4 11.5 11.6 11.7

Additions of Ears . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229 11.1.1 Single and double ears . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229 11.1.2 Conformal subgraphs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231 Deletions of Ears . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233 11.2.1 Removable ears . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233 11.2.2 Series reductions of ears . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235 11.2.3 The two ear theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235 Ear Decompositions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237 11.3.1 Two equivalent definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . 237 11.3.2 An ear decomposition of the Petersen graph . . . . . . . . . . . . 238 The Number of Ears in a Decomposition . . . . . . . . . . . . . . . . . . . . . . 240 11.4.1 Ear decompositions and perfect matchings . . . . . . . . . . . . . 240 Retract of a Graph . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245 Ear Decompositions of Critical Graphs . . . . . . . . . . . . . . . . . . . . . . . . 249 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253

11.1 Additions of Ears 11.1.1 Single and double ears We now return to the topic of ear decompositions of matching covered graphs which was initiated in Chapter 3. Recall that a single ear in a graph 퐺 is a path of odd length all of whose internal vertices have degree two in 퐺. Suppose that 퐻 is a subgraph of 퐺, and that 푃 is an odd path in 퐺, which is edge-disjoint from 퐻, and is such that its ends are in 푉 (퐻) but its internal vertices, if any, are not. Then 푃 is an ear path of 퐻 and, noting that 푃 © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7_11

229

11 Ear Decompositions

230

is a single ear in 퐻 ∪ 푃, we say that the graph 퐻 + 푃 := 퐻 ∪ 푃 is obtained by adding the ear 푃 to 퐻. One convenient way of remembering ‘ears’ is by thinking of them as bisubdivided ‘edges’. Thus adding an ear to a graph amounts to first adding an edge and then bisubdividing that edge. We showed in Chapter 3 that if 퐻 is a matching covered subgraph of a bipartite matching covered graph 퐺, then any bipartite graph obtained by adding a single ear 푃 to 퐻 is also a matching covered subgraph of 퐺. This statement does not extend to nonbipartite graphs. Consider, for example, the case in which 퐺 is the complete graph 퐾4 , and 퐻 is a 4-cycle in 퐺. There are two ear paths of 퐻 in 퐺 (both being of length one), but adding any one of them singly to 퐻 results in a graph which is not matching covered; it is necessary to add both of them simultaneously to obtain 퐺 from 퐻. Double ears A double ear in a graph 퐺 is a set of two vertex-disjoint single ears in 퐺. When 퐻 is a matching covered subgraph of 퐺, and 푃 and 푄 are two vertex-disjoint ear paths of 퐻 in 퐺, the graph 퐻 + {푃 ∪ 푄} := 퐻 ∪ (푃 ∪ 푄) is said to be obtained from 퐻 by adding the double ear {푃, 푄} to 퐻. The principal purpose of this chapter is to show that any matching covered graph may be obtained from 퐾2 by a sequence of (single or double) ear additions such that all intermediate graphs are also matching covered. Before formally stating and proving this result, let us consider the example shown in Figure 11.1. The paths 푃1 := 12, 푃2 := 157862, 푃3 := 1342, 푃4 := 38, 푃4′ := 47, and 푃5 := 56 are paths of odd length in the M¨obius ladder M8 and the graphs 퐺 1 := 푃1 , 퐺 2 := 퐺 1 + 푃2 , 퐺 3 := 퐺 2 + 푃3 , 퐺 4 := 퐺 3 + {푃4 ∪ 푃4′ }, and 퐺 5 := 퐺 4 + 푃5 are matching covered. The graph 퐺 1 is 퐾2 and, for 2 ≤ 푖 ≤ 5, 푖 ≠ 4, the graph 퐺 푖 is obtained from 퐺 푖−1 by adding the single ear 푃푖 , whereas 퐺 4 is obtained from 퐺 3 by adding the double ear {푃4 , 푃4′ }. The graph 퐺 5 is the M¨obius ladder M8 . As a preliminary to proving the above-mentioned principal result, we shall first establish the following weaker statement: Theorem 11.1 Given any matching covered graph 퐺, there exists a sequence (퐾2 = 퐺 1 ⊂ 퐺 2 ⊂ · · · ⊂ 퐺 푟 = 퐺)

(11.1)

of matching covered subgraphs of 퐺 such that, for 2 ≤ 푖 ≤ 푟, the graph 퐺 푖 is obtained from 퐺 푖−1 by simultaneously adding a set of vertex-disjoint ears. The proof of this result that we shall give is based on the notion of conformal subgraphs introduced in Exercise 3.4.3.

11.1 Additions of Ears 1

231 2

1

2

3

2

5

6

5

6

8

7

8

7

퐺1

퐺2

1

1

2

퐺3

3 1

2

5

6

5

6

8

7

8

7

퐺4

4

4

3

퐺5

4

Fig. 11.1 A bottom-up ear decomposition of the M¨obius ladder M8

11.1.2 Conformal subgraphs Conformal subgraphs Recall that a matchable subgraph 퐻 of a matchable graph 퐺 is a conformal subgraph of 퐺 if the graph 퐺 − 푉 (퐻) has a perfect matching. In other words, 퐻 is a conformal subgraph of 퐺 if any perfect matching of 퐻 may be extended to a perfect matching of 퐺. Suppose that 퐻 is a matchable subgraph of a graph 퐺, let 푃 := 푢푠1 푠2 . . . 푠2푘 푣 be an ear path of 퐻 in 퐺, and let 퐸 even (푃) := {푠1 푠2 , 푠3 푠4 , . . . , 푠2푘−1 푠2푘 } be the set of even numbered edges of 푃. Likewise, 퐸 odd (푃) is the set of odd numbered edges of 푃. We extend these notations to a pair 푅 of vertex disjoint ear paths of 퐻, by defining 퐸 even (푅) and 퐸 odd (푅) to be respectively the set of even numbered and odd numbered edges of each path in 푅. As we noted in Chapter 3, for any perfect matching 푀 퐻 of 퐻, the set 푀 퐻 ∪퐸 even (푃) is a perfect matching of 퐻 + 푃. An easy consequence of this observation is the following: Proposition 11.2 If (퐺 1 , 퐺 2 , . . . , 퐺 푟 ) is any sequence of subgraphs of a matching covered graph 퐺 as described in the statement of Theorem 11.1, then 퐺 푖 is a conformal subgraph of 퐺 푗 for 1 ≤ 푖 ≤ 푗 ≤ 푟. ƒ

11 Ear Decompositions

232

Let 퐻 be a conformal matching covered proper subgraph of a matching covered graph 퐺. We shall see that by adding a suitable set of vertex-disjoint single ears simultaneously to 퐻, we can obtain a conformal matching covered subgraph of 퐺 which properly contains 퐻. Towards this end, let 푀 be any perfect matching of 퐺 whose restriction to 퐸 (퐻) is a perfect matching of 퐻 (such a perfect matching must exist because 퐻 is a conformal subgraph of 퐺). Now let 푓 be any edge in 퐸 (퐺) − 퐸 (퐻) which has at least one end in 푉 (퐻) (such an edge must exist because 퐺 is connected). Let 푀 푓 be any perfect matching of 퐺 containing the edge 푓 , and let 퐶 denote the (푀 푓 , 푀)-alternating cycle which includes the edge 푓 (see Figure 11.2, where edges of 푀 are indicated by dotted lines, and those of 푀 푓 ∩ 퐸 (퐶) are indicated by solid lines). It is not difficult to see that 퐻 ∪ 퐶 is a conformal matching covered subgraph of 퐺 (see Exercise 11.1.1). But the subgraph of 퐺 induced by the edge set 퐸 (퐶) − 퐸 (퐻) consists of vertex-disjoint ear paths of 퐻, and 퐻 ∪ 퐶 is obtained by simultaneously adding those paths as ears to 퐻. (In the Example shown in Figure 11.2, these are the ears 푃1 , 푃2 , and 푃3 .)

푉 (퐻 )

퐶

푓 푃1

푃2

푃3

푉 (퐺 ) − 푉 (퐻 ) Fig. 11.2 Simultaneous addition of vertex-disjoint single ears

Starting with any given conformal subgraph 퐹 of 퐺, one may recursively apply the procedure described above to obtain a sequence of conformal matching covered subgraphs where each member of the sequence, subsequent to 퐹, is obtained from the previous one by simultaneous addition of a set of ears. We thus have the following assertion: (We leave its proof as Exercise 11.1.2.)

11.2 Deletions of Ears

233

Theorem 11.3 Given any matching covered graph 퐺 and a conformal matching covered subgraph 퐹 of 퐺, there exists a sequence (퐹 = 퐺 1 ⊂ 퐺 2 ⊂ · · · ⊂ 퐺 푟 = 퐺)

(11.2)

of conformal matching covered subgraphs of 퐺 such that, for 2 ≤ 푖 ≤ 푟, the graph 퐺 푖 is obtained from 퐺 푖−1 by simultaneously adding a set of vertex-disjoint ears. ƒ Observe that Theorem 11.1 is a just special case of the above theorem in which the first graph 퐹 is 퐾2 . Our aim is to show that a decomposition of the type described in Theorem 11.3 can be achieved by never having to add more than two single ears simultaneously. For establishing this stronger result, we shall find it convenient to view an ear decomposition as a way of reducing a given matching covered graph 퐺 to 퐾2 by ‘deleting’ ears. This approach makes it possible for us to use the theory of removable classes described in the previous chapter.

Exercises ⊲11.1.1 Let 퐻 be a conformal matching covered subgraph of a matching covered graph 퐺, and suppose that 푀 is a perfect matching of 퐺 whose restriction to 퐸 (퐻) is a perfect matching of 퐻. Let 푓 be an edge in 퐸 (퐺) − 퐸 (퐻) which has at least one end in 푉 (퐻), and let 푀 푓 be a perfect matching of 퐺 containing the edge 푓 . Show that: (i) the graph 퐻 ∪ 퐶, where 퐶 is the (푀 푓 , 푀)-alternating cycle containing the edge 푓 , is a a conformal matching covered subgraph of 퐺, (ii) the subgraph of 퐺 induced by 퐸 (퐶) − 퐸 (퐻) consists of vertex disjoint ear paths of 퐻, and (iii) the graph 퐻 ∪ 퐶 is the same as the graph obtained from 퐻 by adding those ear paths to it. ⊲11.1.2 Give a proof of Theorem 11.3.

11.2 Deletions of Ears 11.2.1 Removable ears Deletion of an ear When 푃 is a single ear in a graph 퐺, we denote by 퐺 − 푃 the graph obtained from 퐺 by deleting all the edges and internal vertices of 푃.

11 Ear Decompositions

234

The following fact, which will be found to be very useful, is a consequence of Theorem 11.3. Corollary 11.4 Let 퐺 be a matching covered graph, and let 퐹 be a proper conformal matching covered subgraph of 퐺. Then 퐺 has a set {푃1, 푃2 , . . . , 푃 푘 } of vertex-disjoint single ears such that the graph 퐺 − 푃1 − 푃2 − · · · − 푃 푘 is matching covered and contains 퐹 as a conformal subgraph. Proof By Theorem 11.3 there exists a sequence (퐹 = 퐺 1 ⊂ 퐺 2 ⊂ · · · ⊂ 퐺 푟 −1 ⊂ 퐺 푟 = 퐺) of matching covered subgraphs of 퐺 such that, for 2 ≤ 푖 ≤ 푟, the graph 퐺 푖 is obtained from 퐺 푖−1 by adding a set of vertex-disjoint single ears. In particular, 퐺 푟 = 퐺 is obtained from 퐺 푟 −1 by adding a set of vertex-disjoint single ears, say 푃1 , 푃2 , . . . , 푃 푘 . By deleting these 푘 ears from 퐺 we obtain 퐺 푟 −1 , which is matching covered. By ƒ Proposition 11.2, 퐹 is a conformal subgraph of 퐺 푟 −1 . The assertion follows. Removable ears A single ear 푃 in a matching covered graph 퐺 is removable if 퐺 − 푃 is matching covered. In the particular case in which a removable single ear 푃 is a path of length one, the edge of 푃 is removable in 퐺. A double ear 푅 := {푃, 푄} of a matching covered graph 퐺 is removable if 퐺 − 푅 is matching covered, but neither 퐺 − 푃 nor 퐺 − 푄 is matching covered. In the particular case in which both the single ears 푃 and 푄 of a removable double ear 푅 = {푃, 푄} are paths of length one, the two edges of 푃 and 푄 together constitute a removable doubleton of 퐺. For example, in the graph 퐺 depicted in Figure 11.3, the ear indicated by dotted lines constitutes a removable single ear, and the two paths indicated by solid lines together constitute the two ears of a removable double ear. Henceforth, we shall use the term ‘ear’ to designate either a single or double ear. Accordingly, a removable ear of 퐺 is a removable single ear or a removable double ear of 퐺. For a removable ear 푅 in a matching covered graph 퐺, we say that the graph 퐺 − 푅 is obtained by deleting 푅.

Fig. 11.3 Examples of removable ears

11.2 Deletions of Ears

235

Many statements that apply to a path which is a single ear also apply to each of the two paths that are the constituent single ears of a double ear. For this reason, we shall find it convenient to refer to a path 푃 as a constituent path of an ear 푅 if either 푅 is a single ear and 푅 = 푃, or 푅 is a double ear and 푃 is one of the two paths that constitute 푅. For future reference, we note two simple properties of removable ears in a matching covered graph. Proposition 11.5 Let 퐺 be a matching covered graph. Suppose that 푃 is a constituent path of a removable ear 푅. Then, either 퐺 is an even cycle (and 푅 = 푃 is a removable single ear), or both ends of 푃 have degree at least three in 퐺 (Exercise 11.2.1). ƒ Proposition 11.6 Suppose that 푢 and 푣 are two vertices of degree at least three in a matching covered graph 퐺, and that 푄 is an 푢푣-path in 퐺 all of whose internal vertices have degree two. Then, for any removable ear 푅 of 퐺, either 푄 is one of the constituent paths of 푅, or 퐸 (푄) ∩ 퐸 (푅) = ∅ (Exercise 11.2.1). ƒ

11.2.2 Series reductions of ears In Chapter 3 we noted that the operation of adding a single ear 푢푠1 푠2 . . . 푠2푘 푣 to a graph 퐻 may be viewed as the composite operation of adding an edge joining 푢 and 푣 first, and then bisubdividing that edge by inserting the vertices 푠1 , 푠2 , . . . , 푠2푘 . In much the same way, the operation of deleting a removable single ear 푃 := 푢푠1 푠2 . . . 푠2푘 푣 from a matching covered graph 퐺 may be construed as the composite operation involving the 휕 ( 푋)-contraction 퐻 := 퐺/( 푋 → 푥), where 푋 := {푠1 , 푠2 , . . . , 푠2푘 , 푣}, followed by the deletion of the edge 푢푥. Note that 휕 ( 푋) is a tight cut of 퐺; hence 퐻 is matching covered. Moreover, 퐻 − 푢푥 is isomorphic to 퐺 − 푃; hence 푢푥 is a removable edge of 퐻. The above idea may also be described as consisting of a series reduction of the path 푃 to a path of length one, and then deleting the edge 푒(푃) of that path. (Similarly, the deletion of a removable double ear 푅 = {푃, 푄} from 퐺 may be effected by first reducing 푃 and 푄 to single edges by series reductions, and then deleting the removable doubleton to which 푅 is reduced.)

11.2.3 The two ear theorem We present a proof of the following basic result due to Lov´asz and Plummer [59]. The two ear theorem Theorem 11.7 Given a matching covered graph 퐺 and a proper conformal matching covered subgraph 퐹 of 퐺, there exists a removable (single or double) ear 푅 of 퐺 such that 퐹 is a conformal subgraph of 퐺 − 푅.

11 Ear Decompositions

236

Proof This assertion is clearly valid if 퐺 is an even cycle. So, suppose that it is not an even cycle. By Corollary 11.4, 퐺 has a set {푃1 , 푃2 , . . . , 푃 푘 } of vertex-disjoint single ears such that 퐺 − 푃1 − 푃2 − · · · − 푃 푘 is matching covered and contains 퐹 as a conformal subgraph. Let 퐻 be the graph obtained from 퐺 by the application of series reductions to each ear 푃푖 , thereby obtaining the edges 푒 푖 , 푖 = 1, 2, . . . , 푘. Since the graph 퐺 − 푃1 − 푃2 − · · · − 푃 푘 is matching covered, it follows that the graph 퐻 − 푒 1 − 푒 2 − · · · − 푒 푘 is matching covered. The set {푒 1 , 푒 2 , · · · , 푒 푘 } must include a removable class, which, by Corollary 8.18, is a removable edge or a removable doubleton. If 푒 푖 is a removable edge of 퐻, then 푅 := 푃푖 is a removable single ear in 퐺, because 퐺 − 푃푖 is obtained from 퐻 − 푒 푖 by the bisubdivisions of the ears 푃 푗 , 푗 ≠ 푖. Likewise, if {푒 푖 , 푒 푗 } is a removable doubleton of 퐻, then 푅 := {푃푖 , 푃 푗 } is a removable double ear in 퐺. In both cases, clearly, 퐹 is a conformal subgraph of ƒ 퐺 − 푅. The assertion holds. By taking 퐹 to be the subgraph spanned by a single edge of 퐺, and applying the above theorem, we have: Existence of removable ears Theorem 11.8 Every matching covered graph 퐺 ≠ 퐾2 has a removable ear. ƒ Corollary 11.9 Every matching covered graph 퐺 whose minimum degree is at ƒ least three has a removable class.

Exercises 11.2.1 Give proofs of Propositions 11.5 and 11.6. ⊲11.2.2 (Monotonicity of functions 푏 and 푏 + 푝) Apply Theorems 9.8 and 9.9 to deduce the following inequalities: Monotonicity with respect to deletions of removable ears (i) for any removable single ear 푅 of 퐺: 푏(퐺 − 푅) ≥ 푏(퐺) and (푏 + 푝) (퐺 − 푅) ≥ (푏 + 푝) (퐺);

(11.3)

(ii) for any removable double ear 푅 of 퐺: 푏(퐺 − 푅) = 푏(퐺) − 1 and (푏 + 푝) (퐺 − 푅) = (푏 + 푝) (퐺) − 1.

(11.4)

⊲11.2.3 Let 퐻 be a conformal, nonbipartite, matching covered subgraph of a matching covered graph 퐺, and let 푅 be a removable single ear of 퐻. Show that 퐻 − 푅 is also a conformal nonbipartite matching covered graph of 퐺.

237

11.3 Ear Decompositions

⊲11.2.4 Let 푅 be a removable single ear of a solid matching covered graph 퐺. Using Theorem 10.17, show that 푏(퐺 − 푅) = 푏(퐺) and (푏 + 푝) (퐺 − 푅) = (푏 + 푝) (퐺). ⊲11.2.5 Let 퐺 be any matching covered graph and let 푅 be either a removable double ear of 퐺, or a removable single ear such that 푏(퐺 − 푅) = 푏(퐺). Show that the dimension of Lin(퐺) is just one more than the dimension of Lin(퐺 − 푅).

11.3 Ear Decompositions 11.3.1 Two equivalent definitions Theorem 11.8 immediately implies, by induction, the following assertion: Theorem 11.10 Given any matching covered graph 퐺 there exists a sequence G := (퐺 1 = 퐺 ⊃ 퐺 2 · · · ⊃ 퐺 푟 = 퐾2 )

(11.5)

of subgraphs of 퐺 such that, for 1 ≤ 푖 ≤ 푟 − 1, 퐺 푖+1 = 퐺 푖 − 푅푖 , where 푅푖 is a removable ear of 퐺 푖 .

(11.6)

We define a top-down ear decomposition of a matching covered graph 퐺 as a sequence (11.5) of subgraphs of 퐺 which satisfies the property (11.6). Reversing the order of the sequence we obtain the more traditional definition of an ear decomposition of a matching covered graph 퐺 as a sequence G := (퐺 1 = 퐾2 ⊂ 퐺 2 · · · ⊂ 퐺 푟 = 퐺)

(11.7)

of matching covered subgraphs of 퐺 such that, for 2 ≤ 푖 ≤ 푟, 퐺 푖 = 퐺 푖−1 + 푅푖 , where 푅푖 is an ear of 퐺 푖 .

(11.8)

We define a bottom-up ear decomposition (or, simply, ear decomposition) of a matching covered graph 퐺 as a sequence (11.7) of subgraphs of 퐺 which satisfies the property (11.8). Note that the stipulation that in a top-down ear decomposition each 푅푖 is a removable ear ensures that all subgraphs in the sequence are matching covered. Conversely, in a bottom-up ear decomposition, the stipulation that all the graphs in the sequence are matching covered ensures that each ear 푅푖 is removable. So, the two types of decompositions differ only on a viewpoint. Suppose that 퐹 is a matching covered subgraph of 퐺 such that every ear decomposition of 퐹 may be extended to an ear decomposition of 퐺. In this case, there exists a sequence G := (퐹 = 퐺 1 ⊂ 퐺 2 · · · ⊂ 퐺 푟 = 퐺) (11.9)

238

11 Ear Decompositions

of subgraphs of 퐺 satisfying the property (11.6). We refer to such a sequence as a partial ear decomposition of 퐺 starting with 퐹. The following result is an immediate consequence of Theorem 11.7 and Proposition 11.2: Theorem 11.11 A matching covered graph 퐺 has an ear decomposition starting with a given graph 퐹 if and only if 퐹 is a conformal matching covered subgraph of ƒ 퐺. An immediate corollary of the above theorem is that an ear decomposition of a matching covered graph 퐺 may start with any edge 푒 of 퐺. More precisely, for any edge 푒 of 퐺, the subgraph induced by 푒 could be the first graph 퐺 1 in an ear decomposition of 퐺. The second graph 퐺 2 in any ear decomposition of 퐺 is an even cycle. The above theorem implies that any conformal cycle in 퐺 could be the second graph 퐺 2 in an ear decomposition of 퐺. If 퐺 has at least two edges, and 푀 is any perfect matching of 퐺, it is easy to see that any given edge of 퐺 is contained in some 푀-alternating cycle, and hence in a conformal cycle of 퐺 (see Exercise 11.3.3). Little (1974, [53]) has strengthened this statement to show that any two distinct edges of 퐺 are contained together in a conformal cycle of 퐺 (see Exercise 11.3.5). Thus, given any two edges 푒 and 푓 of 퐺, there exists an ear decomposition (퐺 1 , 퐺 2 , . . . , 퐺 푟 ) of 퐺 such that (i) 퐸 (퐺 1 ) = {푒}, and (ii) 퐺 2 is a conformal cycle containing both 푒 and 푓 .

11.3.2 An ear decomposition of the Petersen graph We have already seen an ear decomposition of the M¨obius ladder M8 (Figure 11.1). An ear decomposition of the Petersen graph is shown in Figure 11.4. For 2 ≤ 푖 ≤ 5, the ear added to 퐺 푖−1 to obtain 퐺 푖 (or, equivalently, the ear deleted from 퐺 푖 to obtain 퐺 푖−1 ) is indicated by broken lines. The graph 퐺 3 , which is a bisubdivision of 퐾4 , is obtained by adding a double ear to 퐺 2 , a conformal 8-cycle of the Petersen graph. The graph 퐺 4 is obtained from 퐺 3 by the addition of a double ear. All other ear additions involved are single ear additions. One of the many distinguishing properties of the Petersen graph is that, up to isomorphism, it has a unique ear decomposition (Exercise 11.3.2).

Exercises ⊲11.3.1 Show that every nonbipartite matching covered graph 퐺 contains a bisubdivision of a near-bipartite graph as a conformal subgraph. 11.3.2 Show that, up to isomorphism, the Petersen graph has a unique ear decomposition.

239

11.3 Ear Decompositions

퐺1

퐺2

퐺4

퐺3

퐺5

Fig. 11.4 An ear decomposition of the Petersen graph

11.3.3 Let 퐺 be a matching covered graph on at least two edges, and let 푀 be any perfect matching of 퐺. Given any edge 푒 of 퐺, show that there is some 푀-alternating cycle that contains 푒. 11.3.4 Let 퐻 be a proper conformal matching covered subgraph of a matching covered graph 퐺, and let 푀 be a perfect matching of 퐺 whose restriction to 퐸 (퐻) is a perfect matching of 퐻. Show that 퐺 has a removable (single or double) ear 푅 such that (i) 퐻 ⊆ 퐺 − 푅, and (ii) if 푓 is any edge of 푀 in 퐸 (푅), then both ends of 푓 have degree two in 퐺. 11.3.5 Let 푒 and 푓 be two edges of a matching covered graph 퐺, and let 푀 푓 be a perfect matching of 퐺 containing the edge 푓 . (i) Show that there exists an 푀 푓 -alternating cycle 퐶푒 containing the edge 푒. (Hint: See Exercise 11.3.3.) (ii) If 푓 ∉ 퐸 (퐶푒 ), show that there is a removable ear 푅 of 퐺 such that either 푓 does not belong to 퐸 (푅), or 푓 belongs to 퐸 (푅) but both ends of 푓 have degree two in 퐺. (Hint: Apply the result in Exercise 11.3.4 by taking 퐶푒 to be 퐻, and the perfect matching 푀 푓 to be 푀.) (iii) Now use induction on the number of edges to show that 푒 and 푓 are contained together in a conformal cycle of 퐺. (Little (1974, [53]))

11 Ear Decompositions

240

11.3.6 Prove that every matching covered graph distinct from 퐾2 and 퐶2푛 , (푛 ≥ 2) has two edge-disjoint removable ears. (We shall come across a much stronger statement (Theorem 16.12) in Part II. A proof of that statement appeared in [11].) Hint: Use induction on the number of edges. First deal with the simple cases in which 퐺 either has multiple edges or has vertices of degree two. Then show that the statement is valid if 퐺 is either a brick or a brace. What remains then is dealing with the case in which 퐺 has nontrivial tight cuts. By Theorem 5.6, any matching covered graph that has nontrivial tight cuts, either has a nontrivial barrier cut or 2-separation cut. Consider these two cases to complete the proof.

11.4 The Number of Ears in a Decomposition In general, a matching covered graph may have many ear decompositions. This is the case even when the graph under consideration is bipartite. However, we noted in Chapter 3, the number 푟 of graphs in an ear decomposition of a bipartite graph 퐺 is always 푚 − 푛 + 2 and hence is an invariant of 퐺. This is not the case with nonbipartite graphs, as the number of graphs in an ear decomposition G of 퐺 depends on the number of double ears used in the decomposition. The following relationship may be established by a simple counting argument (Exercise 11.4.1). Proposition 11.12 Let 푑 (G) denote the number of double ear additions used in an ear decomposition G, and let 푟 (G) be the number of graphs in G. Then: 푟 (G) = 푚 − 푛 + 2 − 푑 (G).

(11.10)

Every ear decomposition of the Petersen graph uses exactly two double ears, but there are graphs which admit ear decompositions with different numbers of double ears. See Exercise 11.4.5.

11.4.1 Ear decompositions and perfect matchings A list (푀1 , 푀2 , . . . , 푀푟 ) of perfect matchings of a matching covered graph 퐺 is said to be in stepwise descending order if there are edges 푒 1 , 푒 2 , ..., 푒푟 such that, for 1 ≤ 푖 ≤ 푟, the edge 푒 푖 is in 푀푖 but not in any 푀 푗 with 푗 < 푖. The set { 휒 푀1 , 휒 푀2 , . . . , 휒 푀푟 } of incidence vectors of such a list of perfect matchings is a linearly independent subset of the matching space Lin(퐺). (To see this, consider the 푟 × 푚 matrix whose 푖 푡 ℎ -row is 휒 푀푖 . The 푟 × 푟 submatrix of this matrix corresponding to the edges 푒 1 , 푒 2 , . . . , 푒푟 is a lower diagonal matrix with ones on its main diagonal.) If { 휒 푀1 , 휒 푀2 , . . . , 휒 푀푟 } also happens to generate Lin(퐺), we shall refer to it as a stepwise canonical basis of Lin(퐺). We saw in Chapter 3 how an ear decomposition of a bipartite matching covered graph can be used to generate a set of perfect matchings whose incidence vectors

11.4 The Number of Ears in a Decomposition

241

are linearly independent (see Corollary 3.16 and Exercise 6.3.3). More generally, an ear decomposition of any matching covered graph 퐺 may be used to generate a list of perfect matchings which is in stepwise descending order, thereby yielding a linearly independent subset of Lin(퐺). This assertion can be proved by induction on the number of subgraphs in the ear decomposition and relies on the following Proposition 11.13. Recall that if 푅 is a path of odd length, then 퐸 even (푅) and 퐸 odd (푅) denote the sets of even and odd numbered edges of 푅, respectively; if 푅 is a union of two disjoint odd paths 푃1 and 푃2 , then 퐸 even (푅) = 퐸 even (푃1 ) ∪ 퐸 even (푃2 ), and 퐸 odd (푅) = 퐸 odd (푃1 ) ∪ 퐸 odd (푃2 ). Proposition 11.13 Let 퐺 be a matching covered graph which is obtained from one of its matching covered subgraphs 퐻 by the addition of a (single or double) ear 푅. Then, given any list (푁1 , 푁2 , . . . , 푁 푘 ) of perfect matchings of 퐻 in stepwise descending order, there exists a list (푀1 , 푀2 , . . . , 푀 푘 , 푀 푘+1 ) of perfect matchings of 퐺 in stepwise descending order such that 푁푖 = 푀푖 ∩ 퐸 (퐻), for 1 ≤ 푖 ≤ 푘. Proof Let 퐸 even (푅) and 퐸 odd (푅) be defined as above. Clearly, any perfect matching of 퐺 which contains one edge of 퐸 even (푅) contains all edges of 퐸 even (푅). Likewise, any perfect matching of 퐺 which contains one edge of 퐸 odd (푅) contains all edges 퐸 odd (푅). Furthermore, 푀푖 := 푁푖 ∪ 퐸 even (푅) is a perfect matching of 퐺 for 1 ≤ 푖 ≤ 푘. Let 푀 푘+1 be a perfect matching of 퐺 which contains an edge of 퐸 odd (푅). Then 푀 푘+1 includes all the edges in 퐸 odd (푅). It is easy to see that (푀1 , 푀2 , . . . , 푀 푘 , 푀 푘+1 ) is a list of perfect matchings of 퐺 ƒ which is in stepwise descending order (Exercise 11.4.2). The following theorem can be easily derived by applying the above proposition recursively (see Exercise 11.4.3). Theorem 11.14 Given a matching covered graph 퐺, and an ear decomposition G := (퐺 1 , 퐺 2 , . . . , 퐺 푟 ) of 퐺, there exists an associated list (푀1 , 푀2 , . . . , 푀푟 ) of perfect matchings of 퐺 which is in stepwise descending order such that, for 1 ≤ 푖 ≤ 푟, (푀1 ∩ 퐸 (퐺 푖 ), 푀2 ∩ 퐸 (퐺 푖 ), . . . , 푀푖 ∩ 퐸 (퐺 푖 )) is a list of perfect matchings of 퐺 푖 which is in stepwise descending order. ƒ The stepwise descending property of the 푟 perfect matchings (푀1 , 푀2 , . . . , 푀푟 ) of 퐺, as described in the above theorem, implies that their incidence vectors { 휒 푀1 , 휒 푀2 , . . . , 휒 푀푟 } are linearly independent over any field F. This observation, together with Equation 11.10, implies the following corollary. Corollary 11.15 If a matching covered graph 퐺 has an ear decomposition G consisting of 푟 subgraphs, then the dimension of the matching space over any field F is at least 푟. Furthermore, since, 푟 (G) = 푚 − 푛 + 2 − 푑 (G) by Equation 11.10, where 푑 (G) is the number of double ear additions used obtaining G, it follows that 푑 (G) ≥ 푚 − 푛 + 2 − dim(Lin(퐺, F)),

(11.11)

where Lin(퐺, F) is the matching space of 퐺 over the field F and dim(Lin(퐺, F)) is ƒ its dimension.

11 Ear Decompositions

242

Example 11.16 Figure 11.1 shows an ear decomposition of the M¨obius ladder M8 . Using this decomposition, one can obtain the following five perfect matchings of the brick M8 : 푀1 := {12, 57, 68, 34}      푀2 := {15, 78, 26, 34}    푀3 := {13, 24, 57, 68} (11.12)   푀4 := {38, 47, 15, 26}    푀5 := {56, 78, 12, 34}   If we let 푒 1 := 12, 푒 2 := 15, 푒 3 := 13, 푒 4 := 38, and 푒 5 := 56, it can be verified that the edge 푒 푖 appears in 푀푖 but not in 푀 푗 for 푗 < 푖. Thus (푀1 , 푀2 , 푀3 , 푀4 , 푀5 ) is in stepwise descending order. The incidence vectors of these five perfect matchings are the rows of the matrix shown in Table 11.1. 12 15 13 38 56 57 68 34 78 26 24 47 휒 푀1 : 1 0 0 0 0 1 1 1 0 0 0 0 휒 푀2 : 0 1 0 0 0 0 0 1 1 1 0 0 휒 푀3 : 0 0 1 0 0 1 1 0 0 0 1 0 휒 푀4 : 0 1 0 1 0 0 0 0 0 1 0 1 휒 푀5 : 1 0 0 0 1 0 0 1 1 0 0 0 The first five columns of the above matrix form a lower triangular matrix with 1’s on the main diagonal. It follows that the rows of this matrix are linearly independent. Since the dimension of Lin(M8 ) is five, this matrix is thus a basis matrix of Lin(M8 ). Table 11.1 The incidence vectors of the five perfect matchings

Example 11.17 Figure 11.4 shows an ear decomposition P := (퐺 1 , 퐺 2 , . . . , 퐺 5 ) of the Petersen graph P. This decomposition may be used to obtain five perfect matchings of P whose incidence vectors are linearly independent over any field. Thus the dimension of the matching space of P over any field F is at least five. This lower bound falls short of the actual dimension over the reals by one, but over the field Z2 , it is tight. To see this, note that the Petersen graph has, altogether, six perfect matchings. As each edge is in precisely two of those six matchings, their incidence vectors are linearly dependent over Z2 . By Theorem 11.14, given any ear decomposition G of a matching covered graph 퐺 with 푟 := 푟 (G) subgraphs, one can find a list of 푟 perfect matchings which is in stepwise descending order. The set of incidence vectors in that list is always linearly independent. If one’s aim is to find a basis for Lin(퐺) by this means, one must clearly employ an ear decomposition of 퐺 with the largest possible number 푟 of subgraphs. Since 푟 = 푚 − 푛 + 2 − 푑, where 푑 := 푑 (G) is the number of double ear additions used in G, it would thus be of advantage to find an ear decomposition G which involves as few double ear additions as possible.

11.4 The Number of Ears in a Decomposition

243

All ear decompositions of bipartite graphs utilize only single ear additions. Any ear decomposition of a nonbipartite matching covered graph, in particular that of a brick, must use at least one double ear addition; and any ear decomposition of a Petersen brick requires two double ear additions. The following theorem establishes a lower bound on the number of double ear additions in an ear decomposition of a matching covered graph. Theorem 11.18 Let 퐺 be a matching covered graph. For any ear decomposition G := (퐾2 = 퐺 1 ⊂ 퐺 2 · · · ⊂ 퐺 푟 = 퐺) of 퐺, 푑 (G) ≥ (푏 + 푝) (퐺) Proof By induction on the number of edges. If 푟 = 1, then 퐺 = 퐾2 , and the statement is clearly true. So, suppose that 푟 ≥ 2, and let 푅 be the ear deleted from 퐺 = 퐺 푟 to obtain 퐺 푟 −1 . Clearly, G ′ := (퐾2 = 퐺 1 ⊂ 퐺 2 · · · ⊂ 퐺 푟 −1 ), the subsequence of G consisting of its first 푟 − 1 terms, is an ear decomposition of 퐺 푟 −1 . By the induction hypothesis, 푑 (G ′ ) ≥ (푏 + 푝) (퐺 푟 −1 ) = (푏 + 푝) (퐺 − 푅)

Note that 푑 (G) = 푑 (G ′ ) if 푅 is a single ear, and 푑 (G) = 푑 (G ′ ) + 1 if 푅 is a double ear. The assertion now follows from (11.3) and (11.4) in Exercise 11.2.2. ƒ We shall prove in a later chapter that the above bound is tight in the sense that, given any matching covered graph 퐺, there exists an optimal ear decomposition of 퐺, namely an ear decomposition of which 퐺 involves precisely (푏 + 푝) (퐺) double ear additions. (In particular, we shall show that any brick which is not a Petersen brick has an ear decomposition which uses just one double ear addition.) Returning to the example of the M¨obius ladder M8 , we note that, the ear decomposition we found contains just one double ear addition, and the number 푟 of graphs is equal to 푚 − 푛 + 2 − 푑 = 12 − 8 + 2 − 1 = 5. This number coincides with the the dimension of the matching space of M8 . Lo and behold! we have a stepwise canonical basis for Lin(M8 ). Unfortunately, things don’t work so well for the Petersen graph. Every ear decomposition of this brick uses two double ear additions, and hence the above method yields only 푟 = 푚 − 푛 + 2 − 푑 = 15 − 10 + 2 − 2 = 5 perfect matchings whose incidence vectors are linearly independent. This is one short of the required number because the dimension of the matching space of the Petersen graph is six. But all is not lost because the Petersen graph has just six perfect matchings, and their incidence vectors are linearly independent!

Exercises 11.4.1 Give a proof of Proposition 11.12. ⊲11.4.2 Show that the list (푀1 , 푀2 , . . . , 푀 푘 , 푀 푘+1 ) of perfect matchings defined in the proof of Proposition 11.13 is in stepwise descending order.

11 Ear Decompositions

244

⊲11.4.3 Give a proof of Theorem 11.14. 11.4.4 (i) Show that the basis of Lin(퐾6 ) defined in Exercise 6.3.7 is not a stepwise canonical basis; (ii) Find a stepwise canonical basis of Lin(퐾6 ). ⊲11.4.5 Consider the graph 퐺 by adding an edge to the Petersen graph (see Figure 11.5(a)). 1

1

9

2

8

9

3

푥

7

4

6

2

8

3

푥

7

5

4

6

(푎)

5 (푏)

Fig. 11.5 (a) 퐺 := P + 47; (b) 퐻, a bisubdivision of 퐶6 in 퐺

(i) Using the fact that the Petersen graph is a conformal subgraph of 퐺, find an ear decomposition of 퐺 which involves two double ear additions. (ii) Find an ear decomposition of the graph 퐻 shown in Figure 11.5(b) which uses just one double ear. (iii) Complete the ear decomposition of 퐻 to an ear decomposition of 퐺 by adding the three ‘missing edges’ as single ears. (iv) Now use the ear decomposition you have found to obtain a basis for the matching space of 퐺. 11.4.6 Consider the solid brick Σ8 shown in Figure 7.2. (i) Find an ear decomposition of Σ8 using just one double ear deletion; and (ii) using it, find a stepwise canonical basis for Lin(Σ8 ), and justify your answer.

⊲11.4.7 (Ear Decompositions of Solid Matching Covered Graphs) Let 퐺 be a solid matching covered graph. Using the identities in Exercises 11.2.2 and 11.2.4 prove that if G is any ear-decomposition of 퐺, then: 푑 (G) = (푏 + 푝) (퐺) = 푏(퐺). Deduce that any ear decomposition of a solid matching covered graph is an optimal ear decomposition.

245

11.5 Retract of a Graph

11.5 Retract of a Graph Given any matching covered graph 퐺, there exists a related matching covered graph b which can be obtained from 퐺 by means of special types of tight cut contractions 퐺 b called the retract of 퐺, is known as bicontractions. The significance of this graph 퐺, that, when 퐺 is not an even cycle, there exists a one-to-one correspondence between b the set of all removable ears of 퐺 and the set of all removable classes of 퐺. We shall first define bicontractions, explain their relevance, and then describe how the retract of a matching covered graph can be obtained using bicontractions. Bicontraction of a vertex of degree two Let 푣 0 be a vertex of degree two in a matching covered graph 퐺 of order four or more, and let 푣 1 and 푣 2 be its two neighbours. Then 휕 ( 푋), where 푋 := {푣 0 , 푣 1 , 푣 2 }, is a tight cut of 퐺, and the matching covered graph 퐺/푋 is said to be obtained from 퐺 by the bicontraction of 푣 0 . See Figure 11.6 for an illustration. 푣0

푣1

푋

푣2

(푎)

푥

(푏)

Fig. 11.6 Bicontraction of vertex 푣0 : (a) 퐺, (b) 퐺/(푋 → 푥 )

Motivated by the above considerations, we now define the notion of the retract of a matching covered graph. The retract of a matching covered graph b is a matching covered The retract of a matching covered graph 퐺, denoted by 퐺, graph obtained from 퐺 by the following recursive procedure: If 퐺 has order b := 퐺. Otherwise, 퐺 b := 퐺 c푣 , two, or if 퐺 has no vertices of degree two, then 퐺 where 퐺 푣 is the graph obtained by bicontracting a vertex 푣 of degree two in 퐺. Clearly, if 퐺 has more than one vertex of degree two, this recursive procedure

11 Ear Decompositions

246

is not uniquely determined. However, it can be shown that, up to isomorphism, b does not depend on the order in which the bicontractions are the retract 퐺 performed. (An inductive proof of this result first appeared in CLM (2005, [13]). Theorem 11.19 and Corollary 11.20 provide a constructive proof.) Since each b bicontraction preserves the property of being matching covered, we note that 퐺 is matching covered. See Figure 11.7 for an illustration. 푢

푣

퐺

푤

b of a matching covered graph 퐺 Fig. 11.7 The retract 퐺

푥

b 퐺

Let us extend the notion of retract to paths of a matching covered graph. Let 푃 be a path of a matching covered graph 퐺 joining vertices of degree three or more whose internal vertices, if any, have degree two in 퐺. The path 푃 has an image in b of 퐺. That image is obtained by bicontracting the internal vertices of the retract 퐺 b and is called the retract of 푃 in 퐺. 푃, is denoted 푃 Let 퐺 be a matching covered graph of order at least four which is not an even cycle. We shall refer to vertices of degree at least three as the branch vertices of 퐺, and those of degree two, if any, as the subdivision vertices of 퐺. Every edge of 퐺 clearly lies on a path whose ends are branch vertices but whose internal vertices, if any, are not. We shall refer to such paths as lobes in 퐺, and note that all internal vertices of a lobe are subdivision vertices. A lobe 푃 is an even lobe or odd lobe according to the parity of its length. It should be noted that odd lobes are the same as ears whose ends are branch vertices, and also that any two lobes of 퐺 are internally disjoint. The matching covered graph 퐺 shown in Figure 11.7 has one even lobe, namely the path 푢푣푤 connecting the two branch vertices 푢 and 푤. All the edges other than 푢푣 and 푣푤 belong to odd lobes (or ears) in 퐺. The following theorem elucidates the interaction between even and odd lobes in a matching covered graph 퐺. Theorem 11.19 Let 퐺 be a matching covered graph of order at least four which is not an even cycle. Let 퐺 even denote the subgraph of 퐺 induced by the set of edges that belong to even lobes of 퐺. Then the set of branch vertices of 퐺 in any component of 퐺 even is a barrier of 퐺. Furthermore,

11.5 Retract of a Graph

247

(i) the graph 퐺 even is acyclic, and (ii) no odd lobe connects two branch vertices of 퐺 that belong to the same component of 퐺 even . Proof Let 퐻 be a component of 퐺 even . By definition, 퐻 is the union of a subset of the set of even lobes of 퐺. Let 퐵(퐻) denote the set of vertices of 퐻 which are branch vertices of 퐺. Since 퐻 is connected, it follows that there are at least |퐵(퐻)| − 1 even lobes that connect pairs of vertices in 퐵(퐻). Since each even lobe has an odd number of internal vertices, it follows that 퐺 − 퐵(퐻) has at least |퐵(퐻)| − 1 odd components. Using Exercise 1.2.1 we now conclude that 퐵(퐻) is a barrier of 퐺. Suppose that a connected component 퐻 of 퐺 even has a cycle. Then, there are at least |퐵(퐻)| even lobes that connect pairs of branch vertices of 퐺 in 퐻, and hence 퐺 − 퐵(퐻) has at least |퐵(퐻)| odd components. It must in fact have precisely |퐵(퐻)| odd components because 퐺 is matching covered, implying that 퐻 is a cycle. Since each vertex in 퐵(퐻) is a branch vertex of 퐺, and since 퐺 is not a cycle itself, it follows that the graph 퐺 − 퐵(퐻) would have components other than the |퐵(퐻)| odd components of 퐻 − 퐵(퐻). This is impossible because 퐺 is matching covered and 퐵(퐻) is a barrier of 퐺. We conclude that each component of 퐺 even is a tree, and hence that 퐺 even is acyclic. This proves statement (i). Let 퐻 be a component of 퐺 even . As already noted, the set 퐵(퐻) of branch vertices of 퐺 in 퐻 is a barrier of 퐺. Suppose that there is an an odd lobe 푃 which connects two vertices 푢 and 푣 belonging to 퐵(퐻). If the length of 푃 is one, then 푢푣 would be an edge which joins two vertices belonging to the barrier 퐵(퐻) of 퐺. And, if 푃 has length three or more, then 푃 − {푢, 푣} would be an even component of 퐺 − 퐵(퐻). Both these are impossible because 퐵(퐻) is a barrier of 퐺. This proves statement (ii). ƒ Readers may use the graph shown in Figure 11.8, on page 248, to identify its even lobes, its odd lobes and its removable ears (Exercise 11.5.1). Series reduction of an odd lobe (or ear) 푃 connecting two branch vertices 푠 and b whose single edge connects 푡 has the effect of replacing the path 푃 by its retract 푃, the two ends 푠 and 푡 of 푃. (This does not affect the edge sets of the other lobes, odd or even.) And, the series reduction of an even lobe 푃 has the effect of deleting the internal vertices of 푃 and then identifying its ends 푠 and 푡 to form a single vertex, b of 푃. (This does not lead to the identification which is the only vertex of the retract 푃 of the ends of any lobe (odd or even) other than 푃, and thus, the edge sets of all lobes other than 푃 remain intact. This follows from Theorem 11.19.) We leave the proof of the following corollary to the above theorem as Exercise 11.5.3. Corollary 11.20 Let 퐺 be a matching covered graph on at least four vertices which is not an even cycle. Then, for each component 퐻 of 퐺 even , the cut 휕 (퐻) is a tight b is, up to isomorphism, the graph obtained from 퐺 by first cut of 퐺. The retract 퐺 shrinking each component of 퐺 even to a single vertex, and then replacing each odd b of 푃 in 퐺. lobe 푃 by a single edge, which is the only edge of the retract 푃 ƒ

248

11 Ear Decompositions

Clearly the graph 퐾2 has no removable ears. In the even cycle 퐶2푛 there are 2푛 removable single ears, and the retract of 퐶2푛 is 퐶2 . Suppose now that 퐺 is a matching covered graph of order at least four which is not an even cycle. And, suppose that 푅 is a removable ear (single or double) of 퐺. It follows from Proposition 11.6 that each constituent single ear of 푅 is an odd lobe whose ends are branch vertices. For a b the graph obtained from 퐺 by replacing single or double ear 푅 of 퐺, denote by 퐺 ( 푅) b The following conclusion can be drawn each constituent path 푃 of 푅 by its retract 푃. from the preceding analysis: Theorem 11.21 Let 퐺 be a matching covered graph which has at least four vertices and is not an even cycle. Then

b is a removable edge (i) an ear 푃 in 퐺 is a removable single ear of 퐺 if and only if 푃 b of 퐺, and b b 푄} (ii) a double ear {푃, 푄} in 퐺 is a removable double ear of 퐺 if and only if { 푃, b is a removable doubleton of 퐺. ƒ

b shown in Figure 11.7 is the Note that the underlying simple graph of the graph 퐺 b b has bicorn H8 ; 퐺 is obtained by duplicating the only removable edge of H8 . Thus 퐺 two removable edges, and two removable doubletons. Correspondingly, the graph 퐺 has two removable single ears, and two removable double ears.

Exercises 11.5.1 Identify (i) even lobes, (ii) odd lobes, and (iii) removable ears in the graph shown in Figure 11.8.

Fig. 11.8 Graph for Exercise 11.5.1

11.6 Ear Decompositions of Critical Graphs

249

⊲11.5.2 Show that the retract of a near-bipartite graph is also near-bipartite. ⊲11.5.3 Give a proof of Corollary 11.20. ⊲11.5.4 Let 퐻 be any cubic matching covered graph. Show that any matching covered graph 퐺 whose retract is 퐻 is a bisubdivision of 퐻. ⊲11.5.5 Let 퐺 be a matching covered graph such that the maximum degree Δ of its vertices is three or more. (i) Prove that 퐺 has at least Δ edge-disjoint removable ears. (ii) Let 푀 be a perfect matching of 퐺. Prove that 퐺 has at least Δ − 1 edge-disjoint removable ears 푅 such that 푀 − 푅 is a perfect matching of 퐺 − 푅. Hint: apply Theorem 11.21 and Exercise 8.3.8. b be the retract of a matching covered graph 퐺. Show that ⊲11.5.6 Let 퐺 b b and (푏 + 푝) (퐺) = (푏 + 푝) ( 퐺) 푏(퐺) = 푏( 퐺),

∗ 11.5.7 (Number of removable ears of solid graphs) Let 퐺 be a solid matching covered graph. Prove each of the two statements below. (i) If 퐺 is neither 퐾2 nor an even cycle then 퐺 has at least 푚 − 푛 + 2 − 푏 edge-disjoint removable ears. In particular, if 퐺 is bipartite and is neither 퐾2 nor an even cycle then 퐺 has at least 푚 − 푛 + 2 edge-disjoint removable single ears. (ii) If 퐺 is a solid near-brick which is not a bisubdivision of 퐾4 then 퐺 has at least 푚 − 푛 − 1 edge-disjoint removable single ears.

11.6 Ear Decompositions of Critical Graphs ♯ Recall that a graph 퐺 is critical if, for any vertex 푣 of 퐺, the subgraph 퐺 − 푣 has a perfect matching. It turns out that critical graphs afford decompositions which are somewhat analogous to ear decompositions of matching covered graphs. In this section we present a brief discussion of these decompositions. It is easy to see that all critical graphs are connected, and that every nontrivial critical graph is nonbipartite. A critical graph may have cut vertices; however, all blocks of a critical graph are also critical (Exercise 11.6.1). Proposition 11.22 Suppose that 퐺 is a critical graph, and 퐻 is a proper critical subgraph of 퐺. (i) For any (odd) ear path 푃 of 퐻 in 퐺, the graph 퐻 + 푃 := 퐻 ∪ 푃 is also critical. (ii) For any odd cycle 퐶 in 퐺 which has just one vertex in common with 퐻, the graph 퐻 + 퐶 := 퐻 ∪ 퐶 is also critical. ƒ

11 Ear Decompositions

250

We leave the proof of above assertion as Exercise 11.6.2. In the latter case of the statement, we refer to 퐶 as a closed ear and, in the former case, for contrast, we refer to 푃 as an open ear. With the exception that, in dealing with decompositions of critical graphs, we permit closed ears, the terminology we use here is analogous to the one introduced in the context of ear decompositions of matching covered graphs. Thus, a critical subgraph 퐻 of a critical graph 퐺 is conformal if the graph 퐺 − 푉 (퐻) has a perfect matching. A near-perfect matching of a graph 퐺 is a matching 푀 of 퐺 which covers all but one of the vertices of 퐺. So, a subgraph 퐻 of 퐺 is conformal, in the abovedefined sense, if there is a near-perfect matching 푀 of 퐺 such that 푀 ∩ 퐸 (퐻) is a near-perfect matching of 퐻, and 푀 − 퐸 (퐻) is a perfect matching of 퐺 − 푉 (퐻). An ear decomposition of a critical graph 퐺 is a sequence G := (퐾1 = 퐺 1 , 퐺 2 , . . . , 퐺 푟 = 퐺)

(11.13)

of conformal critical subgraphs of 퐺 such that for 1 ≤ 푖 < 푟, the graph 퐺 푖+1 is obtained from 퐺 푖 by adding to it either an open ear or a closed ear. (Note that if 퐺 is nontrivial, the second graph 퐺 2 is an odd cycle and is necessarily obtained from 퐺 1 by the addition of a closed ear.) For an example, see Figure 11.9.

퐺1

퐺2

퐺3

퐺 = 퐺4

Fig. 11.9 An ear decomposition of a critical graph

In the above example, the first two ear additions are those of closed ears. Clearly, it is possible to find an ear decomposition of this graph 퐺 which involves just one closed ear addition (Exercise 11.6.3). Although a critical graph may have different ear decompositions, the number of graphs in the sequence is simply a function of the order and size of the graph. Proposition 11.23 Let G := (퐾1 = 퐺 1 , 퐺 2 , . . . , 퐺 푟 = 퐺) be an ear decomposition of a critical graph 퐺. Then, 푖 = |퐸 (퐺 푖 )| − |푉 (퐺 푖 )| + 2, for 1 ≤ 푖 ≤ 푟. In particular, the number 푟 of graphs in G is equal to |퐸 |−|푉 |+2 (see Exercise 11.6.4). ƒ

11.6 Ear Decompositions of Critical Graphs

251

Theorem 11.24 Given any conformal critical subgraph 퐻 of a critical graph 퐺, there exists an ear decomposition of 퐺 starting with 퐻. That is, there exists a sequence (퐻 = 퐺 푖 , 퐺 푖+1 , . . . , 퐺 푟 = 퐺), where 푖 = |퐸 (퐻)| − |푉 (퐻)| + 2, such that, for 푖 ≤ 푗 < 푟, the graph 퐺 푗+1 is obtained from 퐺 푗 by adding either an open ear or a closed ear. Proof If 퐻 is a spanning subgraph of 퐺, the edges in 퐸 (퐺) −퐸 (퐻) can be added, one by one, as open ears. So, we may assume that 푉 (퐻) ⊂ 푉 (퐺). Since 퐺 is connected, there must be a vertex 푣 of 퐻 adjacent to a vertex 푢 in 푉 (퐺) − 푉 (퐻). Let 푀푣 be a near-perfect matching of 퐺 such that 푀푣 ∩ 퐸 (퐻) is a near-perfect matching of 퐻 which covers all vertices of 퐻 except 푣. Such a perfect matching must exist because 퐻 is a conformal critical subgraph of 퐺. In addition, since 퐺 is critical, there is a perfect matching 푀푢 of 퐺 −푢. Clearly, there must exist an (푀푣 , 푀푢 )-alternating path 푃푢푣 in 퐺 connecting 푢 and 푣. Let 푤 be the first vertex of 푃푢푣 in 푉 (퐻), and let 푃푢푤 be the 푢푤-segment of 푃푢푣 . If 푤 = 푣, then 푣푢푃푢푣 is a cycle of odd length which can be added to 퐻 as a closed ear to obtain a conformal critical subgraph of 퐺 which has more edges than 퐻. On the other hand, if 푤 ≠ 푣, then 푣푢푃푢푤 is an ear path of 퐻 and it can be added to 퐻 as an open ear to the same effect. This procedure can be clearly carried out recursively to obtain the desired type of ear decomposition of 퐺. ƒ For any vertex 푣 of a critical graph 퐺, the subgraph induced by {푣} is clearly a conformal subgraph of 퐺. Thus, the above theorem implies the existence of an ear decomposition of 퐺 starting with 퐾1 . Lov´asz and Plummer (see [59, page 197]) established the following strengthening of the above theorem for 2-connected critical graphs. Theorem 11.25 Every 2-connected critical graph 퐺 admits an ear decomposition in which all ear additions, subsequent to the first one, are those of open ears. Proof By induction on |퐸 | − |푉 |. If |퐸 | − |푉 | = 0, then 퐺 is an odd cycle, and the assertion clearly holds in this case. So, suppose that |퐸 | − |푉 | ≥ 1. Since 퐺 is 2-connected, in any ear decomposition of 퐺, the last ear added is necessarily an open ear. Thus, if there is a decomposition of 퐺 in which the penultimate graph is 2connected, then the induction hypothesis may be applied to that graph and the validity of the assertion may be deduced. With this in view, choose an ear decomposition G := (퐺 1 , 퐺 2 , . . . , 퐺 푟 ) of 퐺 such that the block of the largest size of 퐺 푟 −1 has the maximum possible number of edges. If 퐺 푟 −1 is 2-connected, there is nothing more to prove. Suppose that 퐺 푟 −1 has more than one block, and let 푃푢푣 be the ear added to 퐺 푟 −1 to obtain 퐺 = 퐺 푟 . Since 퐺 is 2-connected, it follows that 퐺 푟 −1 has precisely two end blocks and that 푃푢푣 connects a vertex 푢 in one end block to a vertex 푣 in the other end block. Now let 푀푢 be a perfect matching of 퐺 푟 −1 − 푢, let 푀푣 be a perfect matching of 퐺 푟 −1 − 푣, and let 푄 푣푢 be the (푀푢 , 푀푣 )-alternating 푣푢-path in 퐺 푟 −1 . Since 푄 푣푢 connects vertices belonging to the two end blocks of 퐺 푟 −1 , it must traverse through each block of 퐺 푟 −1 exactly once. For each block 퐵 the first edge of 푄 푣푢 ∩ 퐸 (퐵) is an edge of 푀푣 and the last is an edge of 푀푢 . Let 퐿 denote a block of 퐺 푟 −1 with the largest possible number of edges, and let 푠 and 푡 be the first and last vertices of 푄 푣푢 in 푉 (퐿). See Figure 11.10 for an illustration in which 퐺 푟 −1 has three blocks with 퐿 being the

11 Ear Decompositions

252

middle block (the edges indicated by solid lines belong to 푀푢 , and those indicated by dashed lines belong to 푀푣 ). 푃푢푣

푣

푄푣푠

푢

푄푡푢 푠

푄푠푡

푡

퐿

Fig. 11.10 The graph 퐺푟 −1 , largest block 퐿 of 퐺푟 −1 , and the last ear 푃푢푣

Now let 푅푡 푠 denote the path which is composed of the 푡푣-segment 푄 푡푢 of 푄 푣푢 , the path 푃푢푣 , and the 푣푠-segment 푄 푣푠 of 푄 푣푢 . It can be verified that 퐿 ∪ 푅푡 푠 is a 2connected conformal subgraph of 퐺 (Exercise 11.6.5). If G ′ is an ear decomposition of 퐺 starting with 퐿 ∪ 푅푡 푠 , the largest block of the penultimate member of G ′ has more edges than 퐿. This contradicts the choice of G. ƒ Although the above proof is not constructive, it is possible to recast it as a constructive procedure for finding an ear decomposition of a 2-connected critical graph of the required type. We leave it as an exercise to the readers to describe such a procedure.

Exercises 11.6.1 Show that all blocks of a critical graph are critical. 11.6.2 Give a proof of Proposition 11.22. 11.6.3 Find an ear decomposition of the critical graph 퐺 in Figure 11.9 which involves a single closed ear addition. 11.6.4 Give a proof of Proposition 11.23. 11.6.5 Supply the missing details in the proof of Theorem 11.25.

11.7 Notes

253

11.7 Notes The idea of ear decompositions, though not the terminology, originated in the work of Whitney (1932). He showed that any 2-connected graph may be obtained from a cycle by adding a sequence of ears (see Chapter 5 of Bondy and Murty [3]). The theory of ear decompositions of matching covered graphs seems to have had its origins in the works of Kotzig published mostly in Slovenian journals in the nineteenfifties, and of Hetyei published in 1964 in a Hungarian journal. (References to the works of these two authors can be found in the treatises by Lov´asz and Plummer [59] and Schrijver [84]). Many significant advances were made by Lov´asz and Plummer and a cohesive account was presented for the first time in their book. But some of their proofs were rather involved. Szigeti [86] gave a simpler proof of the Two Ear Theorem 11.7. Our own approach to this theory has been in the general framework of the dependence relation and removable classes described in Chapters 8 and 9. This approach has helped us in finding simpler proofs of several basic results in the theory, and also led us to generalizations of some. For example, CLM [8] proved that every matching covered graph 퐺 which is different from 퐾2 and 퐶2푛 has at least Δ(퐺) removable ears, which implies that 퐺 has Δ(퐺)! ear decompositions. In the next chapter we shall present a very simple proof of a result of Lov´asz which states that every nonbipartite matching covered graph has an ear decomposition in which either the third graph is a bisubdivision of 퐾4 or the fourth graph is a bisubdivision of 퐶6 . One of the main results in Part II is that every brick which is not a Petersen brick has an ear decomposition which involves just one double ear addition.

Chapter 12

Matching Minors

Contents 12.1

12.2 12.3

12.4

Matching Minors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255 12.1.1 Matching-minor-closed classes of graphs . . . . . . . . . . . . . . . 258 12.1.2 Matching minors and tight cuts . . . . . . . . . . . . . . . . . . . . . . . 258 Conformal Minors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263 12.2.1 Unavoidable conformal minors . . . . . . . . . . . . . . . . . . . . . . . 265 Graphs With Specified Conformal Minors . . . . . . . . . . . . . . . . . . . . . 267 12.3.1 퐾4 -free planar bricks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269 12.3.2 퐶6 -free planar bricks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271 12.3.3 A special property of the Petersen graph . . . . . . . . . . . . . . . 272 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 273

12.1 Matching Minors In this final chapter of Part I, we bring together ideas of bicontractions of vertices of degree two and deletions of removable classes described in the preceding chapters to introduce two closely related types of minors that are specific to the theory of matching covered graphs. They play a role in this subject that is similar to that of the familiar type of minors in graph theory. We consider first the more general of these two types of minors which was introduced by Norine and Thomas (2007, [78]). Matching minors A matching covered graph 퐽 is a matching minor of another matching covered graph 퐺 if there is some conformal subgraph 퐹 of 퐺 such that 퐽 is obtainable from 퐹 by means of bicontractions.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7_12

255

12 Matching Minors

256

As an example, consider the tricorn H10 shown in Figure 12.1(a). The wheel 푊5 can be obtained from the conformal subgraph of H10 depicted in Figure 12.1(b) by means of bicontractions. Likewise, 퐾4 can be obtained from the conformal subgraph depicted in Figure 12.1(c) by means of the same operations. Thus 푊5 and 퐾4 are matching minors of H10 .

(푎)

(푏)

(푐)

Fig. 12.1 The wheels 푊5 and 푊3 = 퐾4 are matching minors of H10

The following observation is an obvious consequence of the definition of a matching minor: Proposition 12.1 Any matching minor of a conformal subgraph of a matching covered graph 퐺 is also a matching minor of 퐺. ƒ Theorem 11.11 implies that a matching covered subgraph 퐹 of a matching covered graph 퐺 is a conformal subgraph of 퐺 if and only if it can be obtained from 퐺 by a sequence of deletions of removable ears. But the deletion of a removable ear 푅 from 퐺 (or, from any conformal subgraph of 퐺) can be effected by first applying suitable bicontractions to reduce 푅 to a removable class in the resulting graph and then deleting it from that graph. Thus any conformal subgraph of 퐺 can be obtained from it by means of a sequence of bicontractions of vertices of degree two and deletions of removable classes. It follows that if 퐽 is a matching minor of 퐺, then it can be obtained from 퐺 by a sequence of bicontractions and deletions of removable classes. We shall see that, conversely, any graph 퐽 which can be obtained from 퐺 by a sequence of such operations is a matching minor of 퐺. We need the operation of bisplitting a vertex defined below for establishing this converse. Bisplitting a vertex Let 푥 be a vertex of degree at least two in a matching covered graph 퐺. The operation of bisplitting the vertex 푥 consists of first replacing 푥 by two nonadjacent vertices 푣 1 and 푣 2 , distributing the edges in 휕 (푥) between 푣 1 and 푣 2 so that each of these two vertices is incident with at least one edge in 휕 (푥), then adding a new vertex 푣 0 and joining it to both 푣 1 and 푣 2 . See Figure 12.2.

12.1 Matching Minors

257 푣0 푥 ⇒

푣1

푣2

Fig. 12.2 Bisplitting a vertex 푥

It is easy to check that the resulting graph 퐺 ′ is matching covered in which the vertex 푣 0 has degree two and that 퐺 can be recovered from 퐺 ′ by bicontracting 푣 0 . (In this context, we refer to the new vertex 푣 0 as the inner vertex and the two new vertices 푣 1 and 푣 2 as the outer vertices.) Theorem 12.2 A matching covered graph 퐽 is a matching minor of another matching covered graph 퐺 if and only if there exists a sequence 퐺 = 퐺 1 ⊃ 퐺 2 ⊃ · · · ⊃ 퐺 푟 = 퐽

(12.1)

of matching covered graphs such that, for 2 ≤ 푖 ≤ 푟, the graph 퐺 푖 is obtained from the graph 퐺 푖−1 by either deleting a removable class or by bicontracting a vertex of degree two. Proof We have already seen that any matching minor of 퐺 can be obtained from it by means of a sequence of bicontractions and deletions of removable classes. This proves the ‘only if’ part of the assertion. To prove the converse, suppose that there exists a sequence of matching covered graphs as in (12.1) satisfying the stated properties. We shall prove by induction on the number 푟 of graphs in (12.1) that 퐽 is a matching minor of 퐺. This is obvious if 퐽 = 퐺. So, we may assume that 퐽 ≠ 퐺. By hypothesis, (퐺 2 ⊃ · · · ⊃ 퐺 푟 = 퐽) is a sequence of matching covered subgraphs such that, for 3 ≤ 푖 ≤ 푟, the graph 퐺 푖 is obtained from the graph 퐺 푖−1 by either deleting a removable class from 퐺 푖−1 , or by bicontracting a vertex of degree two in 퐺 푖−1 . By the induction hypothesis, it follows that 퐽 is a matching minor of 퐺 2 . Thus there exists a conformal subgraph 퐹 of 퐺 2 such that 퐽 is obtainable from 퐹 by means of bicontractions alone. Case 1 Graph 퐺 2 is obtained from 퐺 1 = 퐺 by the deletion of a removable class 푅. Clearly, any conformal subgraph of 퐺 2 = 퐺 − 푅 is also a conformal subgraph of 퐺. Thus, 퐹 is a conformal subgraph of 퐺. We conclude that 퐽 is a matching minor of 퐺. Case 2 Graph 퐺 2 is obtained from 퐺 1 = 퐺 by bicontracting a vertex, say 푣 0 , of degree two. If the contraction vertex 푥 resulting from the bicontraction of 푣 0 is not a vertex of 퐹, then 퐹 is a subgraph of 퐺. It can be shown easily that it is also a conformal subgraph of 퐺. On the other hand, if 푥 is a vertex of 퐹, then 퐹 is not a conformal subgraph

258

12 Matching Minors

of 퐺. However, by a suitable bisplitting of 푥, one can obtain a graph 퐹 ′ which is a conformal subgraph of 퐺. Furthermore, 퐽 can be obtained from 퐹 ′ by a sequence of bicontractions, showing that 퐽 is a matching minor of 퐺. We leave the details as Exercise 12.1.1(i). ƒ The following result is a consequence of Theorem 12.2. We leave its proof as Exercise 12.1.1(ii). Corollary 12.3 A matching covered graph 퐽 is a matching minor of a matching covered graph 퐺 ≠ 퐽 if and only if (i) either there is a removable class 푅 of 퐺 such that 퐽 is a matching minor of 퐻 := 퐺 − 푅; (ii) or there is a vertex 푣 of degree two in 퐺 such that 퐽 is a matching minor of the ƒ graph 퐻 obtained from 퐺 by the bicontraction of 푣.

12.1.1 Matching-minor-closed classes of graphs A class G of matching covered graphs is matching-minor-closed if all matching minors of any member of G are also members of G. An obvious example of such a class is the one consisting of all planar matching covered graphs. A less obvious example is the class of all solid matching covered graphs. (It follows from Theorem 7.1 that a graph obtained from the bicontraction of a vertex of degree two in a solid matching covered graph is solid. And, it follows from Theorems 10.14 and 10.15 that the deletion of a removable class from a solid matching covered graph results in a solid matching covered graph. Thus, any matching minor of a solid matching covered graph is also solid.) The fact that the class of solid matching covered graphs is matching-minor-closed implies that no nonsolid matching covered graph can be a matching minor of a solid matching covered graph. This suggests the possibility of a forbidden matching minor characterization of solid matching covered graphs. One such forbidden matching minor is 퐶6 , which happens to be the smallest nonsolid matching covered graph. However, not all ‘퐶6 -free’ matching covered graphs are solid. For example, 퐶6 is not a matching minor of P, the Petersen graph, but P is not solid. An account of what we know about the relationship between matching covered graphs which are solid and those which are ‘퐶6 -free’ is described in Section 12.3. We shall come across several important examples of matching-minor-closed classes of matching covered graphs and their forbidden matching minor characterizations in Part III. In such problems the graph or graphs which we wish to forbid tend to be either simple bricks or simple braces. The results described in the following subsection play a useful role in finding solutions to such problems.

12.1.2 Matching minors and tight cuts We leave the proof of the following preliminary proposition as Exercise 12.1.2.

12.1 Matching Minors

259

Proposition 12.4 Let 퐺 be a matching covered graph. A simple graph 퐽 is a matching minor of 퐺 if and only if 퐽 is a matching minor of the underlying simple graph of 퐺. ƒ We first consider the case in which the matching minor of interest is a simple brick 퐽. Our objective is to show that 퐽 is a matching minor of a matching covered graph 퐺 if and only if it is a matching minor of one of the bricks of 퐺. We present the proof of this assertion through a sequence of lemmas. Before we do that, it is in order to clarify the terminology we use concerning bricks of a matching covered graph 퐺. By a brick of 퐺 we shall mean a brick resulting from any tight cut decomposition of 퐺 and, by a simple brick of 퐺 we shall mean the underlying simple graph of a brick of 퐺. By Theorem 4.17, the list of simple bricks of 퐺 does not depend on the choice of the tight cut decomposition employed. Lemma 12.5 Let 퐶 := 휕 ( 푋) be a nontrivial tight cut of a matching covered graph 퐺. If a simple brick 퐽 is a matching minor of a 퐶-contraction of 퐺 then it is also a matching minor of 퐺. Proof Assume that 퐽 is a matching minor of a 퐶-contraction of 퐺. We prove, by induction on |퐸 |, that 퐽 is a matching minor of 퐺. Let 퐺 1 := 퐺/( 푋 → 푥) and 퐺 2 := 퐺/( 푋 → 푥) denote the two 퐶-contractions of 퐺. Adjust notation so that 퐽 is a matching minor of 퐺 1 . Let 푌 be a maximal superset of 푋 such that 퐷 := 휕 (푌 ) is a nontrivial tight cut of 퐺. Then, 퐷 is a tight cut of 퐺 2 . Moreover, if 퐷 is a trivial cut of 퐺 2 then 퐷 = 퐶. Let 퐻1 := 퐺/(푌 → 푦) and 퐻2 := 퐺/(푌 → 푦) be the two 퐷-contractions of 퐺. 12.5.1 The brick 퐽 is a matching minor of 퐻1 . Proof If 퐶 = 퐷 then 퐻1 = 퐺 1 ; in that case, 퐽 is a matching minor of 퐻1 . Suppose thus that 퐶 and 퐷 are distinct. The cut 퐶 is a nontrivial tight cut of 퐻1 . The brick 퐽 is a matching minor of 퐺 1 , in turn a 퐶-contraction of 퐻1 . By induction, with 퐻1 playing the role of 퐺, we conclude that the brick 퐽 is a matching minor of 퐻1 . ƒ Case 1 The graph 퐻2 has a removable class 푅2 disjoint with 퐷. The set 푅2 is then a removable class of 퐺, the cut 퐷 is tight in 퐺 − 푅2 and 퐻1 is intact as a 퐷-contraction of 퐺 − 푅2 . By induction, 퐽 is a matching minor of 퐺 − 푅2 . By Corollary 12.3, 퐽 is a matching minor of 퐺. Case 2 The graph 퐻2 has a vertex 푣 2 of degree two such that 휕 (푣 2) and 퐷 are disjoint. Let 퐺 ′ be the graph obtained from 퐺 by bicontracting vertex 푣 2 . The cut 퐷 is tight in 퐺 ′ and 퐻1 is intact as a 퐷-contraction of 퐺 ′ . By induction, 퐽 is a matching minor of 퐺 ′ . By Corollary 12.3, 퐽 is a matching minor of 퐺. 12.5.2 If 퐻2 is a brace then either Case 1 or Case 2 applies.

12 Matching Minors

260

Proof If 퐻2 has a removable edge which is not incident with the contraction vertex 푦 then Case 1 applies. Suppose then that 퐻2 does not have a removable edge which is not incident with the contraction vertex 푦. By Theorem 8.3, 퐻2 has only four vertices. Moreover, 퐻2 has a vertex 푣 of degree two in the same part of the bipartition of 퐻2 ƒ that contains vertex 푦. Case 2 then applies. By the maximality of 푌 , the graph 퐻2 does not have nontrivial tight cuts. We may thus assume that 퐻2 is a brick. 12.5.3 If the graph 퐻1 is also a brick then it has multiple edges. Proof Suppose that 퐻1 is also a brick. Then, 퐺 is the splicing of two bricks. By hypothesis, the cut 퐷 is tight and nontrivial; therefore 퐺 is not a brick. By Proposition 5.15, there exists a 2-separation 푆 := {푢, 푣} of 퐺, with 푢 ∈ 푌 and 푣 ∈ 푌 , such that every edge of the cut 퐷 is incident with a vertex in 푆. (See Figure 12.3.) 푢 푌

푌

푤

퐷

푤 ′

푣

Fig. 12.3 The case in which 퐺 is not a brick but both 퐻1 and 퐻2 are

If 푢 is adjacent to just one vertex 푤 ∈ 푌 − 푣, then {푣, 푤} would be a barrier of the brick 퐻2 , which is absurd. Therefore, 푢 is adjacent to at least two vertices in 푌 − 푣. It follows that 퐻1 = 퐺/푌 has multiple edges incident with the contraction vertex 푦. ƒ By hypothesis, 퐽 is a simple brick; hence 퐽 is not isomorphic to 퐻1 . As 퐽 is a matching minor of 퐻1 , it follows that a sequence such as (12.1), starting with 퐻1 and ending with 퐽, has two or more graphs. We now consider two cases, depending on whether the second graph of the sequence is obtained by the deletion of a removable class or by the bicontraction of a vertex of degree two. Case 3 The graph 퐻1 has a removable class 푅1 such that 퐽 is a matching minor of 퐻1 − 푅1 . Suppose that 푅1 and 퐷 are disjoint. The set 푅1 is then a removable class of 퐺, the cut 퐷 is tight in 퐺 − 푅1 and 퐻1 − 푅1 is a 퐷-contraction of 퐺 − 푅1 . By induction, 퐽 is a matching minor of 퐺 − 푅1 . By Corollary 12.3, 퐽 is a matching minor of 퐺.

12.1 Matching Minors

261

Suppose then that 푅1 and 퐷 are not disjoint. In that case, 푅1 has one edge, say 푒, in 퐷. By Proposition 8.14, edge 푒 induces a minimal class, say, 푅2 , of the dependence relation on the edges of 퐻2 . Thus, only edges of 푅2 depend on edges of 푅2 and every edge of 푅2 depends on edge 푒. As 퐻2 , a brick, is 3-connected, 푅2 is a removable class of 퐻2 , by Corollary 9.2. If 푅2 does not contain the edge 푒, then 푅2 would be a subset of 퐸 (퐻2 ) − 퐷, 퐻2 − 푅2 would be matching covered and Case 1 would apply. Therefore, we may assume that 푅2 contains the edge 푒. In this case, 퐷 − 푒 is tight in 퐺 − 푅1 − 푅2 , and 퐻1 − 푅1 is a (퐷 − 푒)-contraction of 퐺 − 푅1 − 푅2 . By induction, 퐽 is a matching minor of 퐺 − 푅1 − 푅2 . But 퐺 − 푅1 − 푅2 is a conformal subgraph of 퐺 with 퐽 as a matching minor, implying that 퐽 is a matching minor of 퐺 itself (see Proposition 12.1). Case 4 퐻1 has a vertex 푣 of degree two such that 퐽 is a matching minor of the graph 퐻1′ obtained by bicontracting 푣 in 퐻1 . Suppose that 휕 (푣) and 퐷 are disjoint. Let 퐺 ′ be the graph obtained from 퐺 by bicontracting vertex 푣. The cut 퐷 is tight in 퐺 ′ and 퐻1′ is a 퐷-contraction of 퐺 ′ . By induction, 퐽 is a matching minor of 퐺 ′ . By Corollary 12.3, 퐽 is a matching minor of 퐺. Suppose thus that 휕 (푣) and 퐷 are not disjoint. Since 퐻2 is a brick, the cut 퐷 has at least three edges and, hence 푣 ≠ 푦. Therefore, precisely one of the two neighbours of 푣, say 푤, is in 푌. See Figure 12.4. 푍

푌 푤 푣

푢

Fig. 12.4 The case in which 퐻1 has a vertex of degree two

Let 푢 be the other neighbour of 푣, and let 푍 := 푌 − {푢, 푣}. The cut 휕 (푍) is tight in 퐺 as well as in 퐺 ′ , the graph obtained from 퐺 by bicontracting 푣. Furthermore, the graph 퐺 ′ /(푍 → 푧) is isomorphic to 퐻1′ . By induction, 퐽 is a matching minor of 퐺 ′ . Hence 퐽 is a matching minor of 퐺, by Corollary 12.3. ƒ Corollary 12.6 Every simple brick of a matching covered graph 퐺 is a matching minor of 퐺. ƒ Theorem 12.7 Let 퐺 be a matching covered graph. A simple brick 퐽 is a matching minor of 퐺 if and only if 퐽 is a matching minor of some simple brick of 퐺.

262

12 Matching Minors

Proof If 퐽 is a matching minor of some simple brick of 퐺 then 퐽 is a matching minor of 퐺, by Corollary 12.6. We prove the converse by induction on |퐸 |. Assume that 퐽 is a matching minor of 퐺. If 퐺 is a brick, then 퐽 is a matching minor of the underlying simple graph of 퐺, which is the simple brick of 퐺. We may thus assume that 퐺 is not a brick. In particular, this implies that 퐺 and 퐽 are distinct. By Corollary 12.3, 퐽 is a matching minor of a graph 퐺 ′ , obtained from 퐺 either by bicontracting a vertex of degree two or by the removal of a removable class of 퐺. Consider first the case in which 퐺 ′ is the result of the bicontraction of a vertex of degree two of 퐺. By induction, 퐽 is a matching minor of a simple brick of 퐺 ′ . But, every simple brick of 퐺 ′ is a simple brick of 퐺. Thus, 퐽 is a matching minor of a simple brick of 퐺. We may thus assume that 퐺 ′ = 퐺 − 푅, where 푅 is a removable class of 퐺. As 퐺 is not a brick, it follows that 퐺 has a nontrivial tight cut 퐶 := 휕 ( 푋). Let 퐺 1 and 퐺 2 denote the two 퐶-contractions of 퐺. By induction, 퐽 is a matching minor of some simple brick of 퐺 − 푅, say 퐻. The cut 퐶 ′ := 퐶 − 푅 is nontrivial and tight in 퐺 − 푅. Thus, 퐻 is a simple brick of some 퐶 ′ -contraction of 퐺 − 푅. The two 퐶 ′ -contractions of 퐺 − 푅 are 퐺 1 − 푅 and 퐺 2 − 푅. Adjust notation so that 퐻 is a simple brick of 퐺 1 − 푅. By Corollary 12.6, 퐽 is a matching minor of 퐺 1 − 푅. Thus, 퐽 is a matching minor of 퐺 1 . By induction, 퐽 is a matching minor of some simple brick 퐿 of 퐺 1 . The graph 퐿 is a simple brick of 퐺. Thus, 퐽 is a matching minor of a simple brick ƒ of 퐺. Corollary 12.8 Let 퐺 be a matching covered graph and let 푣 be a vertex of degree two of 퐺. A simple brick 퐽 is a matching minor of 퐺 if and only if 퐽 is a matching ƒ minor of the graph obtained from 퐺 by the bicontraction of 푣. Curiously, the analogue of Corollary 12.6 for braces is not valid: a simple brace 퐽 of a matching covered graph 퐺 need not be a matching minor of 퐺. For example, 퐾3,3 is a simple brace of 퐾3,3 ⊙ 퐾4 , but it is not a matching minor of 퐾3,3 ⊙ 퐾4 (Exercise 12.1.6). However, if 퐺 is a bipartite matching covered graph, then every simple brace of 퐽 of 퐺 is a matching minor of 퐺 (Exercise 12.1.8).

Exercises 12.1.1 Provide the missing details in the proofs of: (i) Theorem 12.2; (ii) Corollary 12.3. 12.1.2 Give a proof of Proposition 12.4. ⊲12.1.3 Give examples of simple bricks 퐽 and separating cuts 퐶 of a brick 퐺 such that 퐽 is a matching minor of some 퐶-contraction of 퐺 but it is not a matching minor of 퐺. 12.1.4 Prove Corollary 12.8.

12.2 Conformal Minors

263

⊲12.1.5 Show that the Petersen graph is a matching minor of the Blanuˇsa snark depicted in Figure 12.5.

Fig. 12.5 The Blanuˇsa snark

12.1.6 Show that 퐾3,3 is not a matching minor of 퐾3,3 ⊙ 퐾4 . 12.1.7 Let 퐺 be a matching covered graph and let 퐽 be a simple brick that is a matching minor of 퐺. Prove that if 퐺 and 퐽 are not isomorphic then one of the following properties holds: (i) either 퐺 has vertices of degree two and, for every vertex 푣 of degree two, 퐽 is a matching minor of the graph obtained from 퐺 by the bicontraction of 푣, or (ii) 퐺 has minimum degree three or more and 퐺 has a removable class 푅 such that 퐽 is a matching minor of 퐺 − 푅. ⊲12.1.8 Show that a simple brace 퐽 is a matching minor of a bipartite matching covered graph 퐺 if and only if it is a matching minor of a simple brace of 퐺.

12.2 Conformal Minors We now turn to a special class of matching minors which we refer to as conformal minors that have been in vogue, though not under this name, since they were introduced by Little (1975, [54]). They play a role in this theory that is akin to that of topological minors in the theory of embeddings of graphs in surfaces. We recall the definition of conformal minor, originally given in Chapter 3.

12 Matching Minors

264

Conformal minors A matching covered graph 퐽 is a conformal minor of another matching covered graph 퐺 if some bisubdivision of 퐽 is a conformal subgraph of 퐺. As Figure 12.6(a) shows, 퐾4 is a conformal minor of P, the Petersen graph. But 퐶6 is not a conformal minor of P (Exercise 12.2.3). However, if P + 푒 is any graph obtained by adding an edge 푒 to P joining two nonadjacent vertices, then 퐶6 is a conformal minor of P + 푒 as illustrated in 12.6(b), and also of 퐾4 ⊙ P, as illustrated in 12.6(c). 푒

(푎)

(푏)

(푐)

Fig. 12.6 Conformal minors: (a) 퐾4 of P; (b) 퐶6 of P + 푒; (c) 퐶6 of 퐾4 ⊙ P

A restricted bicontraction is the bicontraction of a vertex of degree two that has at least one neighbour which is also of degree two. It follows from 11.11 that if 퐻 is a conformal matching covered subgraph of a matching covered graph 퐺, then 퐻 can be obtained from 퐺 by a sequence of deletions of removable ears (single or double). But the deletion of an ear amounts to first reducing that ear to one of length one by means of restricted bicontractions, and then deleting the only edge of that ear. Likewise, if 퐻 is a bisubdivision of a graph 퐽 then 퐽 can in fact be obtained from 퐻 by means of restricted bicontractions. Thus any conformal minor 퐽 of a matching covered graph 퐺 can be obtained from it by means of a sequence of deletions of removable classes and restricted bicontractions. Conversely, it is not difficult to see that any graph 퐽 obtainable from 퐺 by a sequence of such operations is a conformal minor of 퐺 (Exercise 12.2.1). The following proposition, which is analogous to Theorem 12.2, records these observations. Proposition 12.9 A matching covered graph 퐽 is a conformal minor of another matching covered graph 퐺 if and only if there exists a sequence 퐺 = 퐺 1 ⊃ 퐺 2 ⊃ · · · ⊃ 퐺 푟 = 퐽

(12.2)

of matching covered graphs such that, for 2 ≤ 푖 ≤ 푟, the graph 퐺 푖 is obtained from the graph 퐺 푖−1 by either deleting a removable class from 퐺 푖−1 , or by a restricted bicontraction of a vertex of degree two in 퐺 푖−1 . ƒ

12.2 Conformal Minors

265

As an immediate consequence of the above proposition, we have: Corollary 12.10 Every conformal minor of a matching covered graph 퐺 is also ƒ matching minor of 퐺. The converse is not true in general, due to the fact that unrestricted bicontractions are permissible in obtaining a matching minor of 퐺. (For example, the wheel 푊5 is a matching minor of the tricorn H10 as we have already observed (see Figure 12.1). But it is not a conformal minor of H10 for the simple reason that H10 has no vertices of degree greater than three and hence cannot have a bisubdivision of 푊5 as a conformal subgraph.) However, it is easily seen that if a cubic matching covered graph 퐽 is obtained from a matching covered graph 퐻 by means of bicontractions, then 퐻 must be a bisubdivision of 퐽. Consequently, we have the following: Corollary 12.11 A cubic matching covered graph 퐽 is a matching minor of a matchƒ ing covered graph 퐺 if and only if 퐽 is a conformal minor of 퐺.

12.2.1 Unavoidable conformal minors The following fundamental theorem due to Lov´asz provides the motivation and also serves as a prototype for many of our investigations. Unavoidable conformal minors: 퐾4 , 퐶6 Theorem 12.12 (Lov´asz (1983, [57])) Every nonbipartite matching covered graph has a bisubdivision of either 퐾4 or of 퐶6 as a conformal subgraph. ƒ In other words, every nonbipartite matching covered graph has at least one of 퐾4 and 퐶6 as a conformal minor. In view of Corollary 12.11, the following result implies Theorem 12.12: Theorem 12.13 Every nonbipartite matching covered has either 퐾4 or 퐶6 as a matching minor. Proof By induction on |퐸 |. Consider first the case in which 퐺 has a removable edge, 푒. By the monotonicity of function 푏, the graph 퐺 − 푒 is nonbipartite. By induction, 퐺 − 푒 has either 퐾4 or 퐶6 as a matching minor. In that case, the assertion holds. Consider next the case in which 퐺 has a nontrivial tight cut 퐶 := 휕 ( 푋). At least one 퐶-contraction of 퐺, say, 퐺 1 , is nonbipartite. By induction, 퐺 1 has either 퐾4 or 퐶6 as a matching minor. By Lemma 12.5, 퐺 also has one of 퐾4 or 퐶6 as a matching minor. Suppose then that 퐺 is free of nontrivial tight cuts and also that it has no removable edges. As 퐺 is nonbipartite, we deduce that 퐺 is a brick. As it has no removable edges, it is simple. Moreover, by Theorem 9.14, 퐺 itself is either 퐾4 or 퐶6 . ƒ

266

12 Matching Minors

The original proof of the above theorem given by Lov´asz was somewhat involved. Carvalho and Lucchesi (1996, [7]) found a simpler proof based on the result of Little stated in Exercise 11.3.5. CLM (1999, [8]) deduced this result from Theorem 9.18 by showing that every near-bipartite brick 퐺 different from 퐾4 and 퐶6 has an edge 푒 such that 퐺 − 푒 is near-bipartite. An ear decomposition (퐾2 = 퐺 1 ⊂ 퐺 2 ⊂ · · · ⊂ 퐺 푟 = 퐺) of a nonbipartite matching covered graph 퐺 is primary if either the third graph 퐺 3 is a bisubdivision of 퐾4 or the fourth graph 퐺 4 is a bisubdivision of 퐶6 . Lov´asz [57] uses the term canonical to indicate that an ear decomposition is primary. Theorems 12.12 and 11.11 imply the following fundamental result concerning ear decompositions. Theorem 12.14 (Primary Ear Decomposition Theorem) Every nonbipartite matching covered graph has a primary ear decomposition. ƒ For example, the ear decomposition of the Petersen graph P described in Figure 11.4 is a primary ear decomposition because the third graph is a bisubdivision of 퐾4 ; and the ear decomposition of P + 47 described in Exercise 11.4.5(iii) is a primary ear decomposition because the fourth graph is a bisubdivision of 퐶6 . ƒ

Exercises 12.2.1 Show that a matching covered graph 퐽 is a conformal minor of another matching covered graph 퐺 if and only if 퐽 can be obtained from 퐺 by means of deletions of removable classes and restricted bicontractions. 12.2.2 Consider the graph 퐺 shown in Figure 12.7. Find an ear decomposition of 퐺 which is primary, and one which is not primary.

Fig. 12.7 Primary Ear Decomposition (Exercise 12.2.2)

12.2.3 Show that the Petersen graph does not have a conformal subgraph that is a bisubdivision of 퐶6 . Hint: note that P is a transitive brick, that 퐶6 is cubic, and use Corollaries 12.11 and 12.8. 12.2.4 Show that neither the bicorn H8 nor the tricorn H10 has 퐶6 as a conformal minor.

12.3 Graphs With Specified Conformal Minors

267

12.2.5 Consider the family S2푛 , of staircases defined in Section 2.5. Show that S2푛 , (푛 ≥ 3), has 퐶6 as a conformal minor if and only if 푛 is odd. 12.2.6 Show that 퐶6 is a conformal minor of the graph shown in Figure 12.8 which is obtained by adding an edge to the staircase S12 .

푒

Fig. 12.8 퐶6 is a conformal minor of S12 + 푒

∗ 12.2.7 The objective of this exercise is to prove, by induction, that every brace of order six or more has either 퐾3,3 or the cube as a conformal minor. (i) Prove that for every simple brace 퐺 of order ten or more and for every (removable) edge 푒 of 퐺, the underlying simple graph 퐽 of the retract 퐻 of 퐺 − 푒 has minimum degree three or more. Hint: consider the ways in which 퐻 may have multiple edges and prove that if 퐽 has minimum degree two then 퐺 has a nontrivial barrier containing fewer than five vertices. (ii) Prove that the only simple brace of order six is 퐾3,3 and also prove that every brace of order eight has the cube as a spanning subgraph. (iii) Prove the assertion by induction. If the brace has order ten or more, delete an edge 푒 from 퐺, use the first item to prove that 퐺 − 푒 has a brace of order six or more and then apply Corollary 12.11 and the result stated in Exercise 12.1.8.

12.3 Graphs With Specified Conformal Minors

퐽-Based and 퐽-free graphs Let 퐽 be any specific matching covered graph. We say that a matching covered graph 퐺 is 퐽-based if there exists a conformal subgraph of 퐺 which is isomorphic to a bisubdivision of 퐽; and is 퐽-free if there is no such subgraph of 퐺. In other words, 퐺 is 퐽-based if 퐽 is a conformal minor of 퐺; otherwise, it is 퐽-free.

12 Matching Minors

268

Using the above terminology, Theorem 12.12 may be restated as follows: Theorem 12.15 Every nonbipartite matching covered graph is either 퐾4 -based or is 퐶6 -based (or both). ƒ For example, the Petersen graph P is 퐾4 -based, but is 퐶6 -free; the pentagonal prism P10 is 퐶6 -based, but 퐾4 -free; and the graph P + 47 shown in Figure 11.5 is both 퐾4 -based and 퐶6 -based. The problems of deciding which matching covered graphs are 퐾4 -based and, similarly, which are 퐶6 -based are far from being solved. The following assertion, which is relevant to the solutions of these problems, first appeared in a paper by Kothari and Murty (2015, [50]). But it follows directly from Theorem 12.7 and the result in Exercise 11.5.4. Theorem 12.16 ([50]) Suppose that 퐽 is a cubic brick and that 퐶 is a tight cut of a matching covered graph 퐺. Then 퐺 is 퐽-free if and only if each 퐶-contraction of 퐺 is 퐽-free. (In particular, 퐺 is 퐽-free if and only if each brick of 퐺 is 퐽-free.) ƒ The restriction that 퐽 be a cubic brick, hence a simple brick, is crucial for the validity of the above statement. (As already noted, it is not valid even for cubic braces.) Two unsolved problems Problem 12.17 Characterize 퐾4 -free bricks. Problem 12.18 Characterize 퐶6 -free bricks. Since both 퐾4 and 퐶6 are cubic bricks, it follows from Theorem 12.16 that solutions to the above problems would provide characterizations of 퐾4 -free and 퐶6 -free matching covered graphs. Using a brick generation theorem of Norine and Thomas, which will be described in Part-II, Kothari and Murty [50] were able to resolve them in the special case of planar bricks. In the remainder of this section we describe their results with some explanation but without proofs. We do this to give the reader an inkling of the topics to be studied in Parts II and III. Adopting the terminology of Bondy and Murty [3], we shall refer to a planar embedding of a planar graph as a plane graph. A well-known theorem of Whitney (1933) asserts that every simple 3-connected planar graph has a unique embedding in the plane (see Theorem 10.28 in [3]). The import of this assertion is that if 퐺 ′ and 퐺 ′′ are two plane graphs obtained by embeddings of a simple 3-connected plane graph 퐺, then a cycle of 퐺 is a facial cycle in 퐺 ′ if and only if it is also a facial cycle in 퐺 ′′ . Clearly this statement is also true for subdivisions of simple 3-connected planar graphs.

12.3 Graphs With Specified Conformal Minors

269

12.3.1 푲4 -free planar bricks Suppose that 퐺 is a plane 퐾4 -based graph. Then there exists a sequence (퐺 1 , 퐺 2 , . . . , 퐺 푟 ) of plane graphs such that 퐺 1 is a bisubdivision of 퐾4 and, for 1 ≤ 푖 ≤ 푟 − 1, the plane graph 퐺 푖+1 is obtained by adding an ear to the plane graph 퐺 푖 . Clearly, as 퐺 is a plane graph, any ear path of 퐺 푖 is confined to a face of 퐺 푖 . Thus adding a single ear 푃푖 to 퐺 푖 amounts to dividing a face of 퐺 푖 into two faces by drawing ear path 푃푖 joining two vertices on the boundary of that face. It is easy to verify that 퐺 1 has 4 odd faces, and the number of odd faces of 퐺 푖 is nondecreasing along the sequence. Consequently, any plane matching covered graph that is 퐾4 -based must have at least four odd faces (Exercise 12.3.1(ii)). But not every plane matching covered graph with four (or more) odd faces is 퐾4 -based. For example, the graph depicted in Figure 12.9 has four odd faces but it is 퐾4 -free (Exercise 12.3.2).

Fig. 12.9 A planar matching covered graph that is 퐾4 -free

However, any plane brick with four or more odd faces is 퐾4 -based. This follows from the following characterization of 퐾4 -free plane bricks. A characterization of planar 퐾4 -free bricks Theorem 12.19 (Kothari and Murty [50]) A simple planar brick is 퐾4 -free if and only if its (unique) planar embedding has precisely two odd faces. A cubic nonplanar 퐾4 -free graph of order 14, found by Nishad Kothari, is shown in Figure 12.10. It happens to be the smallest nonplanar 퐾4 -free brick. It is easy to see that the trellis shown in Figure 12.10 is a nonplanar brick (Exercise 12.3.3(i)). Apart from the obvious ‘left-right’ and ‘top-bottom’ reflectional symmetries, it also has the ‘insider-out’ symmetry which maps the quadrilateral 푄 1 := 푢 1 푢 2 푣 2 푣 1 푢 1 onto 푄 2 := 푠1 푠2 푡2 푡1 푠1 , and vice-versa. We shall briefly sketch here a proof, by contradiction, to the effect that the trellis is 퐾4 -free. Let 퐺 denote the trellis and assume that it is 퐾4 -based. This assumption implies that brick 퐺 has a subgraph 퐻 of 퐺 such that (i) 퐻 is a bisubdivision of

12 Matching Minors

270 푢2

푣2 푠2

푡2

푢3

푣3

푢4

푣4

푢5

푣5 푠1

푡1

푢1

푣1

Fig. 12.10 Trellis, the smallest nonplanar 퐾4 -free brick

퐾4 , and (ii) the graph 퐺 − 푉 (퐻) has a perfect matching, say 푀. Let 퐻 ∗ denote the spanning subgraph of 퐺 induced by the edge set 퐸 (퐻) ∪ 푀. In 퐻 ∗ , the four branch vertices of 퐻 have degree three, all the subdivision vertices of 퐻 have degree two, and all the vertices which are incident with edges in the matching 푀 have degree one. It is easy to establish the following statements (assuming that 퐺 is 퐾4 -based) (Exercise 12.3.3(ii)). Proposition 12.20 Any two cycles of the graph 퐻 ∗ have at least one vertex in common. ƒ Proposition 12.21 If 푒 is any removable edge of 퐺 that is not in 퐸 (퐻 ∗ ), then the matching covered graph 퐺 − 푒 is also 퐾4 -based. ƒ Corollary 12.22 Any removable edge 푒 of 퐺 such that 퐺 − 푒 is 퐾4 -free must be an edge of 퐻 ∗ . ƒ Consider for example the two edges 푢 1 푢 2 and 푢 2 푣 2 . The bricks of 퐺 − 푒 1 and 퐺 − 푒 2 happen to be just their retracts: 퐺 1 := 퐺œ − 푢 1 푢 2 = 퐺 − 푢 1 푢 2 /({푢 5 , 푢 1 , 푣 1 } → 푥)/({푢 3 , 푢 2 , 푣 2 } → 푦) and 퐺 2 := 퐺œ − 푢 2 푣 2 = 퐺 − 푢 2 푣 2 /({푢 3 , 푢 2 , 푢 1 } → 푥)/({푣 3 , 푣 2 , 푣 1 } → 푦).

These bricks are shown in Figure 12.11

Note that the bricks of both 퐺 − 푢 1 푢 2 and 퐺 − 푢 2 푣 2 are planar bricks with two odd faces. Hence, by Theorem 12.19, they are 퐾4 -free. Clearly, the edge 푣 1 푣 2 is similar to 푢 1 푢 2 , and the edge 푢 1 푣 1 to 푢 2 푣 2 . Hence 퐺 − 푣 1 푣 2 and 퐺 − 푢 1 푣 1 are also 퐾4 -free. It now follows from Corollary 12.22 that the four edges 푢 1 푢 2 , 푢 2 푣 2 , 푣 2 푣 1 and 푣 1 푢 1 all belong to 퐸 (퐻 ∗ ), and hence that the 4-cycle 푄 1 := 푢 1 푢 2 푣 2 푣 1 푢 1 is a cycle in 퐻 ∗ .

The ‘inside-out’ symmetry of the trellis mentioned earlier implies that the 4-cycle 푄 2 := 푠1 푠2 푡2 푡1 푠1 is also a cycle in 퐻 ∗ . The two cycles 푄 1 and 푄 2 are vertex-disjoint cycles in 퐻 ∗ . This contradicts Proposition 12.20. We conclude that the trellis is 퐾4 -free. A description of an infinite family of 퐾4 -free nonplanar bricks, which includes the trellis, is given in Exercise 12.3.4. Other constructions of 퐾4 -free bricks and fascinating conjectures relating to them were given by Kothari (2019, [46]).

12.3 Graphs With Specified Conformal Minors

271

푣4

푣5

푢4

푥

푠1 푡1

푢5

푢4

푠2

(a) 퐺1 := 퐺œ − 푢1 푢2

푥

푣3

푦

푠2

푡2

푠1

푡1

(b) 퐺2 := 퐺œ − 푢2 푣2

푡2

푦

푣5

푣4

Fig. 12.11 The bricks of 퐺 − 푢1 푢2 and 퐺 − 푢2 푣2

12.3.2 푪6 -free planar bricks Now let us turn to the problem of characterizing 퐶6 -free planar bricks. Odd wheels, being solid bricks, are 퐶6 -free. It can be verified that H10 , the tricorn, is 퐶6 -free (Exercise 12.2.4), and also that any staircase S2푛 , where 푛 is an even integer which is at least four, is 퐶6 -free (Exercise 12.2.5). Kothari and Murty showed that these are the only simple 퐶6 -free planar bricks: The list of all simple 퐶6 -free planar bricks Theorem 12.23 (Kothari and Murty [50]) The only simple planar 퐶6 -free bricks are the odd wheels, the tricorn and staircases of order 4푘. ƒ Since every solid brick is 퐶6 -free and since the tricorn and staircases are not solid, we have the following characterization of simple planar solid bricks, originally proved by CLM (2006, [15]).

272

12 Matching Minors

Corollary 12.24 (Kothari and Murty [50]) The only simple planar solid bricks are ƒ the odd wheels.

12.3.3 A special property of the Petersen graph The problem of characterizing nonplanar 퐶6 -free matching covered graphs and the seemingly unrelated problem of characterizing nonplanar solid bricks are now known to be equivalent. This is a consequence of the following theorem due to Lucchesi, Carvalho, Kothari and Murty [64]: Theorem 12.25 The Petersen graph is the only simple nonplanar 퐶6 -free brick that is not solid. ƒ We conclude this part with the assertion above, which establishes a property of the Petersen graph that is not shared by any other simple brick. This graph will continue to play a prominent role in the chapters to come.

Exercises 12.3.1 Let 퐺 be a 퐾4 -based plane matching covered graph. Let (퐺 1 , 퐺 2 , . . . , 퐺 푟 ) be an ear decomposition of 퐺 which starts with a bisubdivision of 퐾4 . (i) Verify that the number of odd faces of 퐺 푖 is at least as large as the number of faces of 퐺 푖−1 , for 푖 = 2, . . . , 푟. (ii) Deduce that any plane matching covered graph that is 퐾4 -based must have at least four odd faces. 12.3.2 The planar matching covered graph shown in Figure 12.9 has four odd faces. Using Theorem 12.16, show that this graph is 퐾4 -free. 12.3.3 (i) Show that the trellis is nonplanar. (ii) Give proofs of Propositions 12.20, 12.21, and Corollary 12.22. ⊲12.3.4 Let 푘 ≥ 2 be any positive integer, and consider the prism P4푘+2 obtained from the two odd cycles 푢 1 푢 2 · · · 푢 2푘+1 푢 1 and 푣 1 푣 2 · · · 푣 2푘+1 푣 1 by adding an edge joining 푢 푖 to 푣 푖 , for 1 ≤ 푖 ≤ 2푘 + 1. Now replace the edge 푢 2푘+1 푣 2푘+1 by the path 푢 2푘+1 푠1 푡1 푣 2푘+1 , and the edge 푢 3 푣 3 by the path 푢 3 푠2 푡2 푣 3 . Add the edges 푠1 푠2 and 푡1 푡2 . We shall denote the resulting graph by Tr4푘+6 . Note that Tr14 is the trellis shown in Figure 12.10. (i) Draw the graph Tr18 . (ii) Show that Tr4푘+6 is a nonplanar 퐾4 -free brick for each 푘 ≥ 2.

12.4 Notes

273

⊲12.3.5 The Heawood graph, shown in Figure 12.12, is the unique cubic graph of girth six with the fewest number of vertices. Show that it is 퐾3,3 -free. (Hint: At least three edges of 퐾3,3 have to be bisubdivided to obtain a bipartite matching covered graph that is free of 4-cycles.)

Fig. 12.12 The Heawood graph (a nonplanar 퐾3,3 -free brace)

12.4 Notes The first author to recognize the importance of the idea of conformal minors of matching covered graphs was Little (1972, [51]). Following the lead of Kasteleyn (1963, [44]), he considered graphs which admit a certain special type of orientation called a Pfaffian orientation. He showed that a bipartite matching covered graph admits a Pfaffian orientation if and only if it is 퐾3,3 -free. (The Heawood graph shown in Figure 12.12 is an example of such a graph.) Later on, Robertson, Seymour and Thomas (1999, [82]) and, independently, McCuaig (2001, [68]), established a structural characterization of 퐾3,3 -free bipartite matching covered graphs which led to a polynomial-time algorithm for recognizing such graphs. These results will be described in greater detail in Part III. Norine and Thomas (2007, [78]) introduced the notion of matching minor. In that paper they established an important theorem concerning matching minors known as the ‘splitter theorem’. This theorem was one of the tools used in the proof of Theorem 12.25. Brick and brace generation procedures are a source for the inductive tools that are needed for analyzing properties of matching covered graphs with specified excluded matching minors. These procedures are the subject matter of Part II. Some of the future advances in this theory may well come from a deeper understanding of the structure of certain matching-minor-closed classes of matching covered graphs!

Part II

Brick and Brace Generation

Chapter 13

A Conjecture of Lov´asz Concerning Bricks

Contents 13.1

13.2

13.3

푏-Invariant Edges in Bricks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 277 13.1.1 푏-Invariant ears in matching covered graphs . . . . . . . . . . . . 278 13.1.2 Our approach to Conjecture 13.2 . . . . . . . . . . . . . . . . . . . . . . 279 (푏 + 푝)-Invariant Edges in Bricks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 280 13.2.1 The three case lemma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281 13.2.2 The Petersen graph . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 284 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287

13.1 풃-Invariant Edges in Bricks The properties of matching covered graphs with nontrivial tight cuts may often be deduced from those of their bricks and braces. But to deal with bricks and braces themselves, which are free of such cuts, one requires subtler inductive tools. In case of bricks, for example, optimal ear decompositions, that is, ear decompositions with as few double ear additions as possible, turn out to be useful for analyzing their properties. The seminal work of Lov´asz on the matching lattice (mentioned in Section 6.3.2), and a conjecture posed by him in that context, served as the motivation for much of the work of CLM (2002, [9], [10] and [11]). This work led to the development of the theory in several different directions. Independently, motivated by questions related to Pfaffian orientations of graphs, Little (1975, [54]), McCuaig (2001, [68]), Robertson, Seymour and Thomas (1999, [82]), and Norine and Thomas (2007,[78]) established deep results on bricks and braces. The objective of this part of the book is to describe some of these developments. Recall that the matching lattice of a matching covered graph 퐺, denoted by Lat(퐺), consists of all integer linear combinations of incidence vectors of perfect matchings of 퐺. Clearly, Lat(퐺) is a subset of Lin(퐺) ∩ Z퐸 , where Lin(퐺) is © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7_13

277

278

13 A Conjecture of Lov´asz Concerning Bricks

the matching space of 퐺. Thus, every x in Lat(퐺) is a regular integer vector (see Theorem 6.7). But, as we have noted, the converse does not always hold; the vector 1 of all 1’s is in the matching space of the Petersen graph P, but it does not belong to the matching lattice of P. We shall refer to a brick whose underlying simple graph is P as a Petersen brick. The essential part of Lov´asz’s characterization of the matching lattice is the following theorem which we mentioned in Chapter 6: Theorem 13.1 (Lov´asz [58]) For any matching covered graph 퐺, Lat(퐺) = Lin(퐺) ∩ Z퐸

(13.1)

if and only if no brick of 퐺 is a Petersen brick.

ƒ

The proof of the above theorem given by Lov´asz [58] was somewhat involved. His search for a simpler proof led him to surmise the existence of special types of removable edges in bricks. Recall that a removable edge 푒 in a matching covered graph 퐺 is 푏-invariant if the equality 푏(퐺 − 푒) = 푏(퐺) holds. For brevity, we shall refer to a 푏-invariant removable edge simply as a 푏-invariant edge. ´ (1987, [58]) Lovasz Conjecture 13.2 Every brick different from 퐾4 , 퐶 6 , and P has a 푏-invariant edge. The above conjecture has several relevant and interesting ramifications involving special types of removable ears in matching covered graphs. One such type is described below.

13.1.1 풃-Invariant ears in matching covered graphs Suppose that 푅 is a removable ear in a matching covered graph 퐺. Recall that if 푅 is a double ear, then 푏(퐺 − 푅) = 푏(퐺) − 1, whereas if 푅 is a single ear, then 푏(퐺 − 푅) ≥ 푏(퐺) (see Exercise 11.2.2). The latter inequality suggests the following definition: a removable single ear 푅 of a matching covered graph 퐺 is 푏-invariant if 푏(퐺 − 푅) = 푏(퐺).

A simple calculation shows that if 푅 is either a removable double ear, or is a 푏-invariant single ear of a matching covered graph 퐺, then the dimension of the matching space of 퐻 := 퐺 − 푅 is just one less than that of 퐺 (see Exercise 11.2.5), and a basis for Lin(퐺) can be obtained from a basis for Lin(퐻) by using the procedure described in Proposition 11.13. Clearly, every removable ear of a bipartite matching covered graph is 푏-invariant. But, in general, an arbitrary removable single ear in a matching covered graph need not be 푏-invariant. The following statement may be regarded as a generalization of Lov´asz’s Conjecture:

13.1 푏-Invariant Edges in Bricks

279

Removable-ear-version of Conjecture 13.2 Conjecture 13.3 Any nontrivial matching covered graph 퐺, none of whose bricks is the Petersen graph, either has a removable double ear or has a 푏invariant single ear. Apart from the above, we shall come across many other statements which are closely related to Lov´asz’s Conjecture. One of the interesting facts that we shall learn is that this conjecture is equivalent to the statement that every brick that is not a Petersen brick admits an ear decomposition with just one double ear.

13.1.2 Our approach to Conjecture 13.2 How does one approach a conjecture such as this? Tight cut decompositions are not applicable because this is a statement about bricks. Thus it was necessary for us to look for new techniques and inductive tools for dealing with bricks. Any brick different from 퐾4 and 퐶6 has a removable edge. An examination of conditions under which an arbitrary removable edge in a brick is not 푏-invariant led us to the serendipitous discovery of solid bricks and the striking fact that every removable edge in such a brick is 푏-invariant (see Theorem 10.16). This left us with the problem of having to deal with bricks which have nontrivial separating cuts. A natural idea then was to take the cut-contractions with respect to a nontrivial separating cut and apply induction. Unfortunately there is a hitch; neither of the two cut-contractions with respect to an arbitrary separating cut of a brick need be a brick, or even a near-brick whose properties are closely akin to those of bricks. But, fortunately, we were able to show that every nonsolid brick 퐺 has a robust cut; that is nontrivial separating cut 퐶 such that both the 퐶-contractions of 퐺 are near-bricks. This was a major step. Existence of robust cuts in nonsolid bricks is established in the next chapter. The uniqueness of the role of the Petersen graph is perhaps the most mysterious and fascinating aspect of Lov´asz’s characterization of the matching lattice. The notion of the characteristic of a separating cut introduced in Exercise 4.3.10 was helpful in resolving this mystery; we were able to show that every nonsolid brick different from a Petersen brick has a robust separating cut whose characteristic is three. Further techniques and ideas were necessary to prove Conjecture 13.2. (The first proof of this conjecture appeared in the Ph.D. thesis (1996, [6]) of Marcelo de Carvalho). Article [10] has a proof of a strengthened form of this conjecture, and several consequences of this result were described in [11]. One of them, not unexpectedly, is a proof of Theorem 13.1. All these will be discussed in due course.

13 A Conjecture of Lov´asz Concerning Bricks

280

We shall present a proof of Conjecture 13.2 in Chapter 15 that does not rely on the notion of characteristic. This became possible after we discovered Lemma 10.10 which is a much stronger property of odd wheels used in earlier proofs of Lov´asz’s conjecture.

Exercises ⊲13.1.1 Verify that no edge of the Petersen graph P is 푏-invariant. ⊲13.1.2 List all the removable edges in the two bricks 퐺 and 퐻 shown in Figure 13.1, and the 푏-invariant edges among them.

퐺

퐻

Fig. 13.1 푏-Invariant edges, Exercise 13.1.2

13.2 (풃 + 풑)-Invariant Edges in Bricks Recall that a Petersen brick is one whose underlying simple graph is the Petersen graph; and also that for a matching covered graph 퐺, the symbol 푝(퐺) denotes the number of its Petersen bricks. A removable edge 푒 of a brick 퐺 is (푏 + 푝)-invariant if (푏 + 푝) (퐺 − 푒) = (푏 + 푝) (퐺). In a Petersen brick, every edge is (푏 + 푝)-invariant (Exercise 13.2.1). If 퐺 is not a Petersen brick, then 푏(퐺) = 1 and 푝(퐺) = 0. Thus, in such a brick 퐺, a removable edge 푒 is (푏 + 푝)-invariant if it is 푏-invariant and the brick of 퐺 − 푒 is not a Petersen brick. As an example, consider the brick 퐺 := P + 푒 shown in Figure 13.2, with 푏(퐺) = 1 and 푝(퐺) = 0. The edge 푒 is 푏-invariant because 푏(퐺 − 푒) = 1; however, it is not (푏 + 푝)-invariant because (푏 + 푝) (퐺 − 푒) = 2; but the edges

13.2 (푏 + 푝)-Invariant Edges in Bricks

281

푓1 and 푓2 , indicated by dashed lines, are (푏 + 푝)-invariant because, for 푖 = 1, 2, (푏 + 푝) (퐺 − 푓푖 ) = 1 (Exercise 13.2.2(ii)). (We also note that, although P has no separating cuts of characteristic three, the graph P + 푒 does; the cut 휕 ( 푋) has characteristic three, Exercise 13.2.2(i).) 푒 푢1

푣1 푢5

푓2

푓1

푢2

푣2

푣5 푣4

푣3 푋

푢4

푢3

Fig. 13.2 The edges 푓1 and 푓2 are (푏 + 푝)-invariant in P + 푒

In any brick which is not a Petersen brick, every (푏 + 푝)-invariant edge is also 푏-invariant. But, as we have noted above, a 푏-invariant edge in a brick need not be (푏 + 푝)-invariant. However, as we shall see later in this section, if a brick has a 푏-invariant edge then it also has a (푏 + 푝)-invariant edge. We first present a lemma which is crucial for establishing this assertion.

13.2.1 The three case lemma Let 퐺 be a brick, and let 푒 be a 푏-invariant edge of 퐺. Then 퐺 − 푒 is a near-brick; and therefore, one of the two contractions with respect to any tight cut of 퐺 − 푒 is bipartite. It turns out that the brick of 퐺 −푒 can be obtained from 퐺 −푒 by contracting the bipartite shores of at most two nontrivial tight cuts. Let 퐶 := 휕 ( 푋) be a cut of 퐺 such that 퐶 − 푒 is a nontrivial tight cut of 퐺 − 푒, and suppose that 퐽 := (퐺 − 푒)/( 푋 → 푥) is bipartite. We refer to the part to which 푥 belongs as the inner part of 퐽, and the other part of 퐽 as its outer part. The following property is easily proved: Lemma 13.4 Let 퐺 be a brick and let 푒 be a 푏-invariant edge of 퐺. Suppose that 퐶 := 휕 ( 푋) is a cut of 퐺 such that 퐶 − 푒 is a nontrivial tight cut in 퐺 − 푒, and that 퐽 := (퐺 − 푒)/( 푋 → 푥) is bipartite. Then, no end of 푒 lies in the outer part of 퐽 and at least one end of 푒 lies in the inner part of 퐽.

282

13 A Conjecture of Lov´asz Concerning Bricks

Proof By hypothesis, 퐺 is a brick, 푒 is 푏-invariant and 퐶 − 푒 is a nontrivial tight cut of 퐺 − 푒. Thus, one of the (퐶 − 푒)-contractions of 퐺 − 푒 is bipartite. Let 퐵 and 퐼 denote the outer and inner parts of 퐽, respectively. If edge 푒 has an end in 퐵 or if 푒 has no end in 퐼, then 퐵 would be a nontrivial barrier of 퐺 which is absurd because 퐺 is a brick. Thus, no end of 푒 lies in 퐵 and at least one end of 푒 lies in 퐼. ƒ The three case lemma Lemma 13.5 Let 퐺 be a brick. Suppose that 푒 is a 푏-invariant edge of 퐺 such that 퐺 − 푒 is not a brick, and let 퐻 be the brick of 퐺 − 푒, obtained by a tight cut decomposition of 퐺 − 푒. Then, one of the following three alternatives holds (see Figure 13.3): (i) either 퐺 has a cut 퐶1 := 휕 ( 푋1 ) such that 퐶1 − 푒 is a nontrivial tight cut of 퐺 − 푒 and 퐽1 := (퐺 − 푒)/( 푋1 → 푥1 ) is bipartite, 퐻 = (퐺 − 푒)/( 푋1 → 푥1 ) and the edge 푒 has one end in the inner part of 퐽1 , and the other end in 푋1 = 푉 (퐻) − 푥1 , (ii) or 퐺 has two cuts cuts 퐶1 := 휕 ( 푋1) and 퐶2 := 휕 ( 푋2) such that 푋1 and 푋2 are disjoint, 퐶1 − 푒 and 퐶2 − 푒 are nontrivial tight cuts of 퐺 − 푒, and for 푖 = 1, 2, 퐽푖 := (퐺 − 푒)/( 푋푖 → 푥푖 ) is bipartite, 퐻 = (퐺 − 푒)/( 푋1 → 푥1 )/( 푋2 → 푥2 ) and the edge 푒 has one end in the inner part of 퐽1 , and the other end in the inner part of 퐽2 , (iii) or 퐺 has a cut 퐶1 := 휕 ( 푋1) such that 퐶1 − 푒 is a nontrivial tight cut of 퐺 − 푒 and 퐽1 := (퐺 − 푒)/( 푋1 → 푥1 ) is bipartite, 퐻 = (퐺 − 푒)/( 푋1 → 푥1 ) and the edge 푒 has both ends in the inner part of 퐽1 .

Proof Assume that 퐺 − 푒 is not a brick. Let 퐶1 := 휕 ( 푋1) be a nontrivial cut of 퐺 such that 퐶1 − 푒 is tight and one of the contraction vertices of 퐻 is 푥1 , obtained by contracting 푋1 to 푥1 . As 푒 is 푏-invariant, 퐽1 is bipartite. Moreover, either 푒 has both ends in the inner part of 퐽1 , or 푒 has one end in the inner part of 퐽1 , the other end in 푋1 . In the former case, we have the last of the three asserted cases. Assume thus that 푒 has one end in the inner part of 퐽1 , the other end in 푋1 . If 푥1 is the only contraction vertex of 퐻 then the first of the three cases holds. We may thus assume that 퐻 has more than one contraction vertex. Let 푥2 be another contraction vertex of 퐻, distinct from 푥1 . Let 퐽2 := (퐺 − 푒)/( 푋2 → 푥2 ). Then, 퐽2 is bipartite. Moreover, edge 푒 has one end in the inner part of 퐽2 . This conclusion holds for each contraction vertex 푥2 of 퐻 distinct from 푥1 . We deduce that 퐻 has precisely two contraction vertices. Moreover, the second of the three cases holds. ƒ We now introduce the notion of the index of a 푏-invariant edge 푒 in a brick 퐺 which depends on the pertinent case of Lemma 13.5. The relevance of this notion, which may not be evident at this stage, will hopefully become clear in Chapter 17 where special types of 푏-invariant edges called thin edges are discussed.

13.2 (푏 + 푝)-Invariant Edges in Bricks

283 푒

퐼1

푋1 퐵1 퐻 − 푥1

푋1 푒

퐼1

퐼2 푋1 ∪ 푋2

퐵1

퐵2 푋1

퐻 − 푥1 − 푥2

푋2

푒

퐼1

푋1 퐵1 푋1

퐻 − 푥1

Fig. 13.3 The three cases of Lemma 13.5

Let 퐺 be a brick and let 푒 be a 푏-invariant edge of 퐺. If 퐺 − 푒 is also a brick then we say that 푒 has index zero. If 퐺 − 푒 is not a brick, then we say that 푒 has index one, two or three, depending on which of the three cases stated in Lemma 13.5 holds. If 퐻 has one contraction vertex and one of the ends of 푒 lies in 푉 (퐻) then 푒 has index one. If 퐻 has two contraction vertices then 푒 has index two. Finally, if the last of the three cases holds then 푒 has index three. Whereas the index is at the moment just a pointer into the cases of the Three Case Lemma, this particular numbering will turn out to be mathematically useful in later chapters.

13 A Conjecture of Lov´asz Concerning Bricks

284

13.2.2 The Petersen graph With the aid of the Three Case Lemma, we now proceed to prove the main theorem of this section. In its proof, we shall implicitly make use of the fact that the diameter of P is two and that, given any two paths 푠0 푠1 푠2 and 푡0 푡1 푡2 of P, there is an automorphism of P which maps 푠푖 to 푡푖 , for 0 ≤ 푖 ≤ 2. The interested reader will note that the proof of the following theorem uses many basic results such as the subadditivity of the function 푏 (Theorem 4.20), the monotonicity of the function 푏 (Theorem 9.8) and the notion of the characteristic. Some of the ideas used in this proof will recur in the next two chapters. Theorem 13.6 Let 퐺 be a brick and let 푒 be a 푏-invariant edge of 퐺. If the edge 푒 is not (푏 + 푝)-invariant, then: (i) the characteristic of 퐺 is three, and (ii) some other edge of 퐺 is (푏 + 푝)-invariant (in fact 퐺 has at least two nonadjacent (푏 + 푝)-invariant edges). Proof Since 푒 is 푏-invariant, 푏(퐺 − 푒) = 1. Let 퐻 denote the brick of 퐺 − 푒 obtained by a tight cut decomposition. Suppose that 푒 is not (푏 + 푝)-invariant, implying that 퐺 itself is not a Petersen brick, but 퐻 is. By the Three Case Lemma, the index of 푒 is either zero, one, two, or three. Let us first suppose that the index of 푒 is zero. In this case 퐻 = 퐺 − 푒. If the edge 푒 were a multiple edge of 퐺, then it would itself be (푏 + 푝)-invariant. Thus, 푒 joins vertices which are not adjacent in 퐻. The underlying simple graph of 퐻 is P, the Petersen graph. Considering the symmetries of the Petersen graph, we may thus assume that, up to multiple edges, 퐺 is isomorphic to the graph P + 푒 depicted in Figure 13.2. It has (푏 + 푝)-invariant edges different from 푒 and is of characteristic three (see Exercise 13.2.2(i) and (ii)). In dealing with the remaining cases, we shall adopt the notation used in the statement of the Three Case Lemma. Also, in our illustrations, we shall label the vertices of the Petersen brick 퐻 as in Figure 13.2, with the understanding that when the index is either one or three, one vertex of 퐻 is to be identified with the contraction vertex 푥1 , and that when the index is two, then two vertices of 퐻 would serve as contraction vertices 푥1 and 푥2 . Index of 푒 is three: We shall consider first the case in which the index of 푒 is three as this happens to be the simplest to deal with and yet brings out the essential ideas involved in all the cases. By the third item in the Three Case Lemma, the graph 퐺 − 푒 has a nontrivial tight cut 휕 ( 푋1) such that 퐽1 := (퐺 − 푒)/( 푋1 → 푥1 ) is bipartite, 퐻 = (퐺 − 푒)/( 푋1 → 푥1 ), and the edge 푒 has both ends in 퐼1 . See Figure 13.4. Note that in the figure 퐼1 is shown to have just two vertices, but 퐼1 may be larger. Also, note that a neighbour of the contraction vertex 푥1 = 푢 1 in 퐻, may be joined to more than one vertex in 퐵1 ,

13.2 (푏 + 푝)-Invariant Edges in Bricks

285

although our figures indicate just one edge between each of 푢 5 , 푣 1 and 푢 2 and 퐵1 . However, since 휕 ( 푋1 ) is not tight in 퐺, there is some perfect matching which meets this cut in at least three edges. This implies that there is a matching which matches the three vertices 푢 5 , 푣 1 and 푢 2 with distinct vertices in 퐵1 . We shall establish the first part of the assertion by showing that 퐶 := 휕 ( 푋), where 푋 := {푣 1 , 푣 2 , 푣 3 , 푣 4 , 푣 5 }, is a separating cut of 퐺 whose characteristic is three. The two 퐶-contractions of 퐺 − 푒 are near-bricks with 5-wheels as their bricks. Therefore, 퐶 is a separating cut in 퐺 − 푒. To show that 퐶 is a separating cut in 퐺, we must show that there is a perfect matching 푀푒 of 퐺 containing 푒 that meets 퐶 in precisely one edge. The existence of such a perfect matching is obvious and is shown in Figure 13.4(a). Now to show that 퐶 has characteristic three, one must find a perfect matching which meets 퐶 in three edges. Such a perfect matching 푀푒′ is shown in Figure 13.4(b). Finally, let us turn to showing that 푓1 is a (푏 + 푝)-invariant edge of 퐺. The graph (퐺 − 푒)/ 푋 is a 5-wheel and 푓1 is a spoke of that wheel. Therefore (퐺 − 푒 − 푓1 )/푋 is matching covered (see Exercise 2.2.4), and in fact a near-brick. Also, the brick of (퐺 − 푒)/푋 is a 5-wheel and 푓1 is a spoke of that wheel, and is not in the tight cut 휕 (퐼1 ∪ 퐵1 ). Thus (퐺 − 푒 − 푓1 )/푋 is a near-brick. The cut 퐶 − 푓1 is a robust cut of 퐺 − 푒 − 푓1 because, being of characteristic three, it is not tight, and both the (퐶 − 푓1 )-contractions of 퐺 − 푒 − 푓1 are near-bricks. By the subadditivity of the function 푏 (Theorem 4.20), 푏(퐺 − 푒 − 푓1 ) < 푏((퐺 − 푒 − 푓1 )/푋) + 푏((퐺 − 푒 − 푓1 )/푋) Since, as observed above both (퐺 − 푒 − 푓1 )/푋 and (퐺 − 푒 − 푓1 )/푋 are near-bricks, it follows that 퐺 − 푒 − 푓1 is a near-brick. The edge 푒 is matchable in 퐺 − 푓1 because the perfect matching 푀푒 , which contains 푒, does not contain 푓1 . By the monotonicity of the function 푏 (Theorem 9.8) it now follows that 퐺 − 푓1 is a near-brick. The characteristic of 퐺 − 푓1 is three because the cut 퐶 − 푓1 has characteristic three. Consequently, the only brick of 퐺 − 푓1 also has characteristic three, implying that it is not a Petersen brick. We conclude that 푓1 is a (푏 + 푝)-invariant edge of 퐺. Similarly, the edge 푓2 is also a (푏 + 푝)-invariant edge of 퐺. Analyzing the cases involving other indices is very similar except that, in some cases, asserting the existence of perfect matchings 푀푒 and 푀푒′ is not so straightforward as in the case discussed above. For example, if the index of the edge 푒 is one, the edge 푒 joins a vertex in 퐼1 to a vertex in 푉 (퐻) − 푥1 , but that vertex may or not be adjacent to the contraction vertex 푥1 of 퐻. In the former case, finding the appropriate perfect matchings 푀푒 and 푀푒′ is quite easy, but in the latter case, it is not straightforward. To illustrate the sort of arguments needed we shall discuss this case. Index of 푒 is one and 푒 has an end in 푉 (퐻) − 푥1 that is not adjacent to 푥1 . As we did in the previous case, let us identify the contraction vertex 푥1 with the vertex 푢 1 of 퐻 and take 푤 1 belonging to 퐼1 and 푢 3 as the ends of the edge 푒. To find the perfect matching 푀푒 such that (i) 푒 ∈ 푀푒 , (ii) 푓1 ∉ 푀푒 , and |푀푒 ∩ 퐶| = 1, consider first the graph 퐺 − 푣 1 − 푤 1 . Since 퐺 is a brick, 퐺 − 푣 1 − 푤 1 has a perfect matching, say 퐹. It matches 푢 2 and 푢 5 with distinct vertices, say 푤 2 and 푤 5 in 퐵1 , the vertices

13 A Conjecture of Lov´asz Concerning Bricks

286 퐼1

푒

푒

퐼1

퐵1

퐵1

푣1 푢5

푣1

푓1

푢2

푓2 푣5

푢5

푣2

푣4

푣5

푣3

푋

푢4

푓2

푓1 푣2

푣4

푢3

푢2

푣3

푋

푢4

푢3

(푎)

(푏)

Fig. 13.4 Index of 푒 is three; (a) | 푀푒 ∩ 퐶 | = 1; (b) | 푀푒′ ∩ 퐶 | = 3

in 퐼1 − 푤 1 with vertices in 퐵1 − 푤 2 − 푤 5 , and includes a perfect matching, say 퐹 ′ of the 6-cycle 푣 2 푣 5 푣 3 푢 3 푢 4 푣 4 푣 2 . Now define 푀푒 as follows: 푀푒 = (퐹 − 퐹 ′ ) ∪ {푒, 푣 1 푣 3 , 푣 2 푣 5 , 푢 4 푣 4 } It can be checked that 푀푒 satisfies the three properties stated above. An analogous argument starting with a perfect matching of 퐺 − 푢 2 − 푤 1 yields the matching 푀푒′ which meets the cut 휕 ( 푋) in three edges. Deducing that 퐺 has characteristic three and that 푓1 and 푓2 are (푏 + 푝)-invariant is very similar to the previous case. 푤1

퐼1 퐵1

푢5

퐵1

푒

푣1

푓1

푢2 푣5

푢5

푒

푣1

푓1

푣2

푣4

푣3 푋

푣3 푋 푓2

푓2 푢4

(푎)

푢2

푣5

푣2

푣4

푤1

퐼1

푢3

푢4

Fig. 13.5 Index of 푒 is one; (a) | 푀푒 ∩ 퐶 | = 1; (b) | 푀푒′ ∩ 퐶 | = 3

(푏)

푢3

13.3 Notes

We refer to paper [9] for full details concerning the cases not discussed here.

287

ƒ

Exercises 13.2.1 Show that in a Petersen brick every edge is (푏 + 푝)-invariant. ∗ 13.2.2 Let 퐺 := P + 푒 be the brick depicted in Figure 13.2. Show that: (i) the cut 퐶 := 휕 ( 푋) is a separating cut of characteristic three in 퐺; and (ii) (푏 + 푝) (퐺 − 푒) = 2, whereas (푏 + 푝) (퐺 − 푓1 ) = (푏 + 푝) (퐺 − 푓2 ) = 1. Hint: Find the retracts of 퐺 − 푒, 퐺 − 푓1 and 퐺 − 푓2 and use the fact that a matching covered graph and its retract have the same number of bricks, and the same number of Petersen bricks; see Exercise 11.5.6. (iii) Find all the (푏 + 푝)-invariant edges in the brick P + 푒.

13.3 Notes Laci Lov´asz mentioned Conjecture 13.2 to the second author in a private communication around 1988. In a talk given at a workshop in Ard`eche, France, in 1994, Santosh Vempala stated that he and Laci had established the validity of this conjecture and sketched some ideas. At this stage, the work on Marcelo Carvalho’s thesis, under the supervision of the first author, had also progressed towards a solution. The second author joined the team some time later, and the three authors, CLM, produced a preliminary report with the title “A Proof of the Lov´asz-Vempala Theorem” in May 1997. We sent Laci a copy of that report and inquired if his work with Santosh on this conjecture had been written up. With his characteristic kindness and generosity, he wrote to us in a letter dated 11 June, 1998: “we have not finished the writing up, and have no claim on the result”. (With that letter he enclosed two short notes. One of them is [61], whose contents have been described in Section 9.3.) It appears that both Laci and Santosh became interested in problems in other areas and abandoned this project. This subject is poorer for it!

Chapter 14

Robust Cuts in Bricks

Contents 14.1 14.2 14.3 14.4

Robust Cuts in Nonsolid Bricks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 289 14.1.1 Peripheral robust cuts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 293 Robust Cuts with a Shore of Minimum Size . . . . . . . . . . . . . . . . . . . . 295 Quasi 푏-Invariant Edges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 297 14.3.1 Barrier cut pairs and essentially 2-separation cut pairs . . . . 299 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 301

14.1 Robust Cuts in Nonsolid Bricks ♯

Robust cuts Recall that a separating cut 퐶 in a matching covered graph 퐺 is robust if both the 퐶-contractions of 퐺 are near-bricks. The purpose of this section is to establish the existence of robust cuts in nonsolid bricks. In the brick 퐺 shown in Figure 14.1, both 퐶 and 퐷 are nontrivial separating cuts. It can be seen that both 퐶-contractions of 퐺 are bricks, but one of the two 퐷-contractions of 퐺 has two bricks. Therefore 퐶 is a robust cut of 퐺, but 퐷 is not. Let 퐺 be a nonsolid brick. By definition, 퐺 has nontrivial separating cuts. Consider any nontrivial separating cut 퐷 of 퐺, and let 푀0 be a perfect matching in 퐺 such that |푀0 ∩ 퐷| > 1. Now let C be the collection of cuts 퐶 of 퐺 such that (i) |푀0 ∩ 퐶| > 1 and (ii) 퐶  퐷, where  is the precedence relation on cuts defined in Section 7.2.2. We leave as Exercise 14.1.1 the proof that every cut in C is separating.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7_14

289

14 Robust Cuts in Bricks

290 1

2

3

4

5

퐶 6 퐷 7

8

Fig. 14.1 Cut 퐶 is robust but 퐷 is not

Let C∗ be the subcollection of C consisting of those cuts that are minimal with respect to the precedence relation . We refer to a member of C∗ as a cut which is 푀0 -induced by 퐷. (See Exercise 14.1.4 for an example.) We shall show that every such cut is a robust cut of 퐺. We start with two preliminary lemmas. Lemma 14.1 Let 퐶 := 휕 ( 푋) be a cut in C∗ . Then every 2-vertex cut is a barrier of the 퐶-contraction 퐺 1 := 퐺/( 푋 → 푥) of 퐺. (Equivalently, no tight cut of the graph 퐺 1 is a 2-separation cut.) Proof No 2-vertex cut of 퐺 1 can be a vertex cut of 퐺, as 퐺, a brick, is 3-connected; it follows that every 2-vertex cut of 퐺 1 contains 푥. Suppose now that one such cut, {푥, 푣}, is not a barrier of 퐺 1 . Let 퐷 1 and 퐷 2 denote two crossing 2-separation tight cuts of 퐺 1 associated with the 2-separation {푥, 푣}. See Figure 14.2. 퐶

푥

퐷1

퐷2 푣

Fig. 14.2 2-separation cuts 퐷1 and 퐷2 in 퐺1

14.1 Robust Cuts in Nonsolid Bricks

291

Recall that M denotes the collection of perfect matchings of 퐺. For any perfect matching 푀 ∈ M, we clearly have: |푀 ∩ 퐷 1 | + |푀 ∩ 퐷 2 | = |푀 ∩ 휕 ( 푥)| + |푀 ∩ 휕 (푣)| = |푀 ∩ 퐶| + 1.

(14.1)

Since |푀0 ∩ 퐶| is greater than one, at least one of |푀0 ∩ 퐷 1 | and |푀0 ∩ 퐷 2 | is greater than one. Adjust notation so that |푀0 ∩ 퐷 1 | > 1.

It follows immediately from equation (14.1) that |푀 ∩ 퐷 1 | ≤ |푀 ∩ 퐶| for all 푀 ∈ M, implying that 퐷 1  퐶. Since 퐺 is a brick, 퐷 2 is not a tight cut of 퐺. Therefore, there must exist a perfect matching 푀2 of 퐺 such that |푀2 ∩ 퐷 2 | > 1. Hence, by equation (14.1), |푀2 ∩ 퐷 1 | < |푀2 ∩ 퐶|, which implies that 퐷 1 strictly precedes 퐶. This contradicts the minimality of 퐶. ƒ Lemma 14.2 Let 퐶 := 휕 ( 푋) be a cut in C∗ . Suppose that the 퐶-contraction 퐺 1 := 퐺/( 푋 → 푥) of 퐺 is not a brick. Then, 퐺 1 is not bicritical and, for any nontrivial barrier 퐵1 in 퐺 1 : (i) the contraction vertex 푥 of 퐺 1 is in 퐵1 , and (ii) 퐺 1 − 퐵1 has precisely one nontrivial odd component, 퐻1 , and the cuts 퐶 and 퐶1 := 휕 (퐻1 ) are matching-equivalent. Proof The graph 퐺 1 is matching covered because 퐶, a cut in C, is a separating cut of 퐺 and 퐺 1 is a 퐶-contraction of 퐺. Suppose that 퐺 1 is not a brick. By Lemma 14.1, 퐺 1 is not bicritical. Let 퐵1 denote a nontrivial barrier in 퐺 1 . (i) The contraction vertex 푥 has to be in 퐵1 ; because 퐺, a brick, is bicritical. (ii) As 퐺 1 is matching covered, 퐵1 , a barrier of 퐺 1 , cannot contain both ends of any edge of 퐺 1 . Therefore, if 퐻1 , 퐻2 , . . . , 퐻푡 , 푡 = |퐵1 |, denote the odd components of 퐺 1 − 퐵1 , then 푡 Õ |푀 ∩ 휕 (퐻푖 )| = |푀 ∩ 퐶| + (푡 − 1), (14.2) 푖=1

for any perfect matching 푀 of 퐺 (see Exercise 14.1.2). The above equation implies that 휕 (퐻푖 )  퐶, for 1 ≤ 푖 ≤ 푡, and also that |푀0 ∩ 휕 (퐻푖 )| > 1 for at least one value of the index 푖. Adjust notation so that |푀0 ∩ 휕 (퐻1 )| > 1, and let 휕 (퐻1 ) be the cut 퐶1 . The fact that |푀0 ∩ 퐶1 | > 1 implies that |푉 (퐻1 )| > 1, and also that 퐶1 ∈ C. (All cuts 휕 (퐻 푗 ), 푗 ≠ 1, turn out to be trivial.) See Figure 14.3. If possible, suppose that 퐻 푗 is nontrivial for some 푗 ≠ 1. Since 퐺 is a brick, the cut 휕 (퐻 푗 ) cannot be tight in 퐺. Thus, there exists a perfect matching 푀 푗 of 퐺 such that |푀 푗 ∩ 휕 (퐻 푗 )| > 1. It now follows from equation (14.2) that |푀 푗 ∩ 퐶1 | < |푀 푗 ∩ 퐶|. This implies that 퐶1 strictly precedes the cut 퐶, which contradicts the minimality of 퐶. We conclude that 퐻1 is the unique nontrivial component of 퐺 1 − 퐵. From equation (14.2) we infer that 퐶 and 퐶1 are matching equivalent. ƒ

14 Robust Cuts in Bricks

292 퐶

푥

| 퐵1 | = 푡

퐶1 푋1 Fig. 14.3 퐺1 := 퐺/(푋 → 푥 ) (Lemma 14.2)

Theorem 14.3 Let 퐺 be a nonsolid brick. Let C be the collection of nontrivial separating cuts of 퐺. Let C∗ be the subcollection of C which are minimal with respect to the precedence relation  on the cuts of 퐺. Let 퐶 := 휕 ( 푋) be a cut in C∗ and let 퐺 1 := 퐺/( 푋 → 푥). Then, there exists a subset 푋 ′ of 푋 (possibly equal to 푋) such that: (i) the cuts 퐶 and 퐶 ′ := 휕 ( 푋 ′) are matching-equivalent cuts of 퐺, (ii) the graph 퐺 ′ := 퐺/( 푋 ′ → 푥 ′ ) is a brick, and (iii) the 퐶 ′ -contraction 퐻 ′ := 퐺 1 /( 푋 ′ → 푥 ′ ) of 퐺 1 is bipartite and matching covered, and its two contraction vertices, 푥 and 푥 ′ , are in distinct parts of its bipartition. Consequently, 퐺 1 is a near-brick. Proof If 퐺 1 is a brick then the assertion holds immediately, with 푋 ′ := 푋. We may thus assume that 퐺 1 is not a brick. To prove part (i) we now apply Lemma 14.2. The graph 퐺 1 is not bicritical. Let 퐵1 denote a maximal nontrivial barrier of 퐺 1 . The contraction vertex 푥 of 퐺 1 is in 퐵1 . The graph 퐺 1 − 퐵1 has precisely one nontrivial component, say, 퐾 ′ . Let 푋 ′ := 푉 (퐾 ′ ). The cut 퐶 ′ := 휕 ( 푋 ′ ) is matching-equivalent to 퐶. This proves part (i). Let us now prove part (ii), which asserts that 퐺 ′ is a brick. For this, assume the contrary. As 퐶 and 퐶 ′ are matching-equivalent, the cut 퐶 ′ is also in C∗ . By Lemma 14.2, 퐺 ′ has a nontrivial barrier, 퐵′1 , which contains the contraction vertex 푥 ′ of 퐺 ′ . For any vertex 푣 of 퐵′1 − 푥 ′ , the graph 퐺 ′ − 푥 ′ − 푣 does not have a perfect matching. This implies that the graph 퐾 ′ − 푣 does not have a perfect matching. Thus, 퐾 ′ is not critical. By Theorem 3.2, 퐵1 is not maximal, a contradiction. This proves part (ii). The validity of part (iii) follows immediately from (i) 푥 ∈ 퐵1 , (ii) 퐾 ′ is the unique nontrivial component of 퐺 1 − 퐵1 and (iii) 푋 ′ = 푉 (퐾 ′ ). Finally, as the cuts 퐶 and 퐶 ′ are matching-equivalent in 퐺, it follows that 퐶 ′ is tight in 퐺 1 . One of the 퐶 ′ -contractions of 퐺 1 is the graph 퐺 ′ , a brick. The other 퐶 ′ -contraction of 퐺 1 is the bipartite matching covered graph 퐻 ′ . We conclude that 퐺 1 is a near-brick. ƒ

14.1 Robust Cuts in Nonsolid Bricks

293

Note that the theorem implies, in particular, the following: Corollary 14.4 Let 퐺 be a nonsolid brick. Let C be the collection of nontrivial separating cuts of 퐺. Let C∗ be the subcollection of C which are minimal with respect to the precedence relation  on the cuts of 퐺. Every cut in C∗ is robust. Consequently, every nonsolid brick has a robust cut. ƒ Underlying the above proof, there is a polynomial-time algorithm for finding a robust cut in a nonsolid brick 퐺 which accepts any nontrivial separating cut 퐷 of 퐺 as input and returns a robust cut of 퐺, induced by 퐷, as output (Exercise 14.1.3). However, there is a catch, we do not know if the problem of deciding whether or not a given brick has a nontrivial separating cut is in the complexity class P.

14.1.1 Peripheral robust cuts If 퐶 is a robust cut of a brick 퐺, then the two 퐶-contractions of 퐺 are near-bricks, but, in general, need not be bricks. In proving properties of bricks, we cannot directly apply the induction hypothesis to the 퐶-contractions if they were not bricks. Fortunately, one can infer from Theorem 14.3 a certain structural result that applies to all nonsolid bricks. Let 퐶 := 휕 ( 푋) be a robust cut of a brick 퐺. Let 푋 ′ , 퐶 ′ , 퐺 ′ and 퐻 ′ be as in the statement of Theorem 14.3. If we apply Theorem 14.3, with 푋 playing the role of 푋 and 퐺 2 := 퐺/( 푋 → 푥) playing the role of 퐺 1 , we deduce that (i) there exists a subset 푋 ′′ of 푋 such that 퐶 ′′ := 휕 ( 푋 ′′ ) is matching-equivalent to 퐶, (ii) the graph 퐺 ′′ := 퐺/( 푋 ′′ → 푥 ′′ ) is a brick and (iii) the 퐶 ′′ -contraction 퐻 ′′ := 퐺 2 /( 푋 ′′ → 푥 ′′ ) of 퐺 2 is bipartite and matching covered, and its two contraction vertices, 푥 and 푥 ′′ , are in distinct parts of its bipartition. Thus, the graph 퐻 := 퐺/( 푋 ′ → 푥 ′ )/( 푋 ′′ → 푥 ′′ ), which is the splicing of 퐻 ′ and 퐻 ′′ , is bipartite and matching covered, and its two contraction vertices, 푥 ′ and 푥 ′′ , are in distinct parts of the bipartition of 퐻. Moreover, the brick 퐺 is the result of the splicing of the bipartite matching covered graph 퐻 and the two bricks 퐺 ′ and 퐺 ′′ . The following assertion summarizes these observations. Corollary 14.5 Let 퐺 be a brick, and let 퐶 := 휕 ( 푋) be a robust cut of 퐺. Then, there exist a subset 푋 ′ of 푋 and a subset 푋 ′′ of 푋 such that: (i) the three cuts 퐶, 퐶 ′ := 휕 ( 푋 ′) and 퐶 ′′ := 휕 ( 푋 ′′ ) are matching-equivalent robust cuts of 퐺, (ii) the graphs 퐺/푋 ′ and 퐺/푋 ′′ are both bricks, and (iii) the graph 퐺/푋 ′ /푋 ′′ is bipartite and matching covered. ƒ We refer to 휕 ( 푋 ′ ) and 휕 ( 푋 ′′ ) as peripheral robust cuts, and to the bricks 퐺/푋 ′ and 퐺/푋 ′′ as the corresponding peripheral bricks. See Figure 14.4 for an illustration.

14 Robust Cuts in Bricks

294 푋 ′

푋

푋

푋 ′′

Fig. 14.4 Peripheral robust cuts (Corollary 14.5)

Exercises ⊲14.1.1 Prove that every cut in C is separating. Hint: apply Theorem 4.2. 14.1.2 Verify the validity of equation (14.2). (Compare this equation with equation (10.2).) ⊲14.1.3 Describe a polynomial-time algorithm which accepts a brick 퐺 and a nontrivial separating cut 퐷 and returns a robust cut in 퐺. Hint: there is no need to have the resulting robust cut to be 푀0 -induced, but it is possible to determine 푀0 , because there exists a pair {푢 1 푣 1 , 푢 2 푣 2 } of nonadjacent edges of 퐶, such that the graph 퐺 − 푢 1 − 푣 1 − 푢 2 − 푣 2 is matchable. 14.1.4 Consider the brick 퐺 shown in Figure 14.1, with the cuts 퐶 and 퐷 indicated by broken lines, and the perfect matching 푀0 , indicated by thick lines, such that |푀0 ∩ 퐷| = 3. Show that the cut 퐶 is 푀0 -induced by 퐷, and verify that 퐶 is a robust cut. ⊲14.1.5 Give an example to show that the peripheral bricks with respect to a robust cut of a nonsolid brick need not be solid bricks. A subset 푋 of the vertex set 푉 of a matching covered graph 퐺 is critical if 푋 is nontrivial and the subgraph 퐺 [푋] of 퐺 induced by 푋 is critical. ∗ 14.1.6

(i) Show that a brick 퐺 is nonsolid if and only if there are two disjoint critical subsets 푋 and 푌 of 푉 such that the graph 퐺 − ( 푋 ∪ 푌 ) is matchable. CLM (2004, [13]). (ii) Deduce the Reed-Wakabayashi Theorem 7.8 from the above statement.

⊲14.1.7 Let 퐶 := 휕 ( 푋) be a robust cut of a brick 퐺 such that 푋 is minimal. Show that the near-brick 퐺 1 := 퐺/( 푋 → 푥) is in fact a brick. 14.1.8 Let 퐶 = 휕 ( 푋) be a robust cut of a brick 퐺 such that 퐺 1 := 퐺/( 푋 → 푥) is a Petersen brick, and let 푄 be the vertex set of a pentagon of 퐺 1 which does not contain the contraction vertex 푥. Show that 휕 (푄) is a robust cut of 퐺 whose characteristic is three.

14.2 Robust Cuts with a Shore of Minimum Size

295

14.2 Robust Cuts with a Shore of Minimum Size By definition, if 퐶 is a robust cut of a nonsolid brick 퐺, then both 퐶-contractions of 퐺 are near-bricks. By Corollary 14.4, every nonsolid brick has robust cuts. As indicated in the statement of Exercise 14.1.7, it is always possible to find a robust cut 퐶 such that one of the 퐶-contractions of 퐶 is a brick. However, there is no guarantee that such a 퐶-contraction be a solid brick, as indicated in the statement of Exercise 14.1.5. The objective of this subsection is to prove the stronger statement that every nonsolid brick 퐺 has a robust cut 퐶 such that one of the two 퐶-contractions is a solid brick. Robust cuts having minimum shores Lemma 14.6 (CLM (2005, [13])) Let 퐺 be a nonsolid brick and let 푋 be a subset of 푉 of minimum possible cardinality such that 퐶 := 휕 ( 푋) is a robust cut of 퐺. Then the 퐶-contraction 퐺 1 := 퐺/( 푋 → 푥) is a solid brick. Proof It follows from the minimality of | 푋 | that 퐺 1 is a brick (see Exercise 14.1.7).

Suppose that 퐺 1 is a nonsolid brick, and let 퐷 0 be a nontrivial separating cut of 퐺 1 . By Exercise 4.2.4 the pair {퐷 0 , 퐶} is cohesive in 퐺.

If 퐷 0 is a robust cut of 퐺, we have an immediate contradiction as to the minimality of | 푋 |. If this is not the case, we shall show that there is a robust cut of 퐺 with a shore with fewer than | 푋 | vertices. Towards this end, let 푀 ′ be a perfect matching of 퐺 1 such that |푀 ′ ∩ 퐷 0 | ≥ 3. (Such a matching exists because 퐷 0 is a nontrivial separating cut of brick 퐺 1 .) Extend 푀 ′ to a perfect matching 푀 of 퐺. (Such a matching exists because 퐶 is separating in 퐺.) Let 퐷 be a cut 푀-induced by 퐷 0 in 퐺. By Corollary 14.4, cut 퐷 is robust in 퐺. Note that 푀 has only one edge in 퐶, yet it has three or more edges in 퐷. Consequently, 퐶 and 퐷 are not matching-equivalent. We now analyse three cases, depending on where 퐷 is located in relation to 퐶. Suppose first that cut 퐷 is a cut of 퐺 1 . Cuts 퐶 and 퐷 are not matching-equivalent; thus, they are distinct. Then, 퐷 is a robust cut of 퐺 and a nontrivial cut of 퐺 1 , in contradiction to the minimality of | 푋 |. Now suppose that 퐷 is a cut of 퐺 2 := 퐺/( 푋 → 푥). The perfect matching 푀 of 퐺 has three or more edges in 퐷 and only one edge in 퐶. Thus, 푀 ∩ 퐸 (퐺 2 ) is a perfect matching of 퐺 2 that may be extended to a perfect matching of 퐺 that contains just one edge in 퐷 0 . This is a contradiction, as 퐷  퐷 0 .

Finally, suppose that 퐶 and 퐷 cross, and let 푌 be the shore of 퐷 such that | 푋 ∩ 푌 | is odd. Let 퐼 := 휕 ( 푋 ∩ 푌 ) and let 푈 := 휕 ( 푋 ∩ 푌 ). The pair {퐶, 퐷} is cohesive. By Theorem 4.14, (i) no edge of 퐺 joins a vertex in 푋 ∩ 푌 to a vertex in 푋 ∩ 푌 , (ii) for each perfect matching 푀 of 퐺, |푀 ∩ 퐶| + |푀 ∩ 퐷| = |푀 ∩ 퐼 | + |푀 ∩ 푈|,

(14.3)

296

14 Robust Cuts in Bricks

(iii) the collection {퐶, 퐷, 퐼, 푈} is cohesive, and (iv) the cuts 퐼 and 푈 are separating. 14.6.1 Cut 푈 is trivial in 퐺. In fact, the shore 푋 ∩ 푌 is a singleton. Proof Let us first show that 푈 is tight in 퐺 2 . So, suppose that the cut 푈 is not tight in 퐺 2 . Then, a perfect matching of 퐺 2 containing three or more edges in 푈 may be extended to a perfect matching 푀 ′′ of 퐺 containing just one edge in cut 퐶 and just one edge in cut 퐷 0 . From (14.3) we deduce that 푀 ′′ contains three or more edges in 퐷, contradicting the fact that 퐷 precedes 퐷 0 . We deduce that 푈 is tight in 퐺 2 . The cut 퐶 is robust in 퐺 and 퐺 2 is a 퐶-contraction of 퐺. Thus, the graph 퐺 2 is a near-brick. As 푈 is tight in 퐺 2 , it follows that some 푈-contraction of 퐺 2 is bipartite. Thus, 푈 has a shore in 퐺 2 , say, 푍, such that the subgraph of 퐺 2 induced by 푍 is bipartite. Moreover, the majority part 푍+ of 푍 is a barrier of 퐺 2 , and 푈 is the only nontrivial barrier cut associated with 푍+ . We shall now prove that 푍 = 푋 ∩ 푌 and then deduce that 푍 is a singleton. Suppose first that the contraction vertex 푥 of 퐺 2 is in 푍+ . In that case, the cuts 푈 and 퐶 are matching-equivalent. From (14.3), we infer that cuts 퐷 and 퐼 are also matching-equivalent. By Corollary 14.4, the cut 퐼 is robust in 퐺. The shore 푋 ∩ 푌 of 퐼 is a proper subset of 푋. Thus, we have a contradiction to the minimality of | 푋 |. We may thus assume that 푥 is not in 푍+ . In that case, 푍+ is a barrier of 퐺. As 퐺, a brick, is bicritical, it follows that 푍+ is a singleton. The cut 푈 is a barrier cut associated with 푍+ . As 푍+ is a singleton, 푍+ is a shore of 푈 in 퐺 2 . The shore (푌 − 푋) + 푥 of 푈 in 퐺 2 is not a singleton, because 푌 − 푋 is nonempty. Thus, 푍+ must be the shore 푋 ∩ 푌 of 푈. Indeed, 푋 ∩ 푌 is a singleton and 푈 is trivial. ƒ No edge of 퐺 joins a vertex of 푋 ∩ 푌 to a vertex of 푋 ∩ 푌 . Thus, 푋 ∩ 푌 is not a singleton, else 퐶 and 퐷 would both be nontrivial tight cuts of the brick 퐺, which is absurd. Hence, |푌 | = | 푋 ∩ 푌 | + | 푋 ∩ 푌 | < | 푋 ∩ 푌 | + | 푋 ∩ 푌 | = | 푋 |, implying that |푌 | < | 푋 |. This contradicts the minimality of | 푋 |. We conclude that 퐺 1 is a solid brick. ƒ

Exercises 14.2.1 Consider the cut 퐶 in the brick 퐺 shown in Figure 14.5. Show that 퐶 is a robust cut of 퐺 and that neither 퐶-contraction of 퐺 is solid. (This example shows that minimality of | 푋 | in the proof of Lemma 14.6 cannot be replaced by minimality of 푋 as a set.)

14.3 Quasi 푏-Invariant Edges

297

퐼

퐶

퐷 Fig. 14.5 Minimality of |푋 | in Lemma 14.6 cannot be replaced by minimality of 푋 as a set.

14.3 Quasi 풃-Invariant Edges

Quasi 푏-invariant edges A removable edge 푒 in a nonsolid brick 퐺 is quasi 푏-invariant if 푏(퐺 − 푒) = 2. By Corollary 14.4, every nonsolid brick has robust cuts. As has already been mentioned, the strategy we use to prove the existence of 푏-invariant edges in a nonsolid brick 퐺, distinct from 퐾4 , 퐶6 and P, is to consider a robust cut 퐶 of 퐺 and use inductive arguments on the two 퐶-contractions 퐺 1 and 퐺 2 of 퐺. If 푒 is an edge that is 푏-invariant in both 퐺 1 and 퐺 2 , then both 퐺 1 − 푒 and 퐺 2 − 푒 are near-bricks. Thus, 푒 is a removable edge of 퐺 and 퐶 − 푒 is a separating cut of 퐺 − 푒. By the subadditivity of function 푏 (Theorem 4.20), 푏(퐺 − 푒) ≤ 푏(퐺 1 − 푒) + 푏(퐺 2 − 푒) = 2,

(14.4)

where equality holds if and only if the cut 퐶 − 푒 is tight in 퐺 − 푒. By the monotonicity of function 푏 (Theorem 9.8), 푏(퐺) ≤ 푏(퐺 − 푒). Consequently, edge 푒 is 푏-invariant in 퐺 if and only if 퐶 − 푒 is not a tight cut of 퐺 − 푒. The following lemma, which plays a crucial role in several steps of our proof of Conjecture 13.2, records this crucial observation. Lemma 14.7 Let 퐺 be a brick, let 퐶 be a robust cut of 퐺 and let 푒 be an edge that is 푏-invariant in both 퐶-contractions of 퐺. Then edge 푒 is 푏-invariant in 퐺 if and only if 퐶 − 푒 is not tight in 퐺 − 푒. Equivalently, 푒 is 푏-invariant in 퐺 if and only if there is a perfect matching 푀0 of 퐺 such that (i) 푒 ∉ 푀0 , and (ii) |푀0 ∩ 퐶| > 1. ƒ The following lemma and its corollary state easily proved sufficient conditions under which a cut 퐷 of 퐺 with the property that 퐷 − 푒 is tight in 퐺 − 푒 is a robust cut of 퐺 (Exercise 14.3.1).

14 Robust Cuts in Bricks

298

Lemma 14.8 Let 퐺 be a brick, let 푒 be a quasi 푏-invariant edge of 퐺, and let 퐶 be a nontrivial cut of 퐺 such that 퐶 − 푒 is a tight cut of 퐺 − 푒. Then 퐶 is a robust cut ƒ of 퐺 if and only if it is a separating cut of 퐺. By Theorem 4.2, a cut 퐷 of a matching covered graph 퐺 is a separating cut if and only if, given any edge of 퐺, there is a perfect matching of 퐺 containing that edge which meets 퐷 in precisely one edge. The following corollary is an obvious consequence of the hypothesis that 퐷 − 푒 is tight in 퐺 − 푒 and Theorem 4.2. Corollary 14.9 Let 푒 be a removable edge of a brick 퐺 and let 퐶 be a nontrivial cut of 퐺 such that 퐶 − 푒 is tight in 퐺 − 푒. If 퐺 has a perfect matching which contains the edge 푒 and meets 퐶 in precisely one edge then 퐶 is robust in 퐺 . ƒ Recall that the characteristic of a separating cut 퐶 in a matching covered graph 퐺, denoted by 휆(퐶), is ∞ if 퐶 is tight, and is the smallest integer 푘 > 1 such that there is a perfect matching of 퐺 which meets 퐶 in precisely 푘 edges if 퐶 is not tight. Lemma 14.10 Let 퐺 be a brick, let 퐶 and 퐷 be nontrivial odd cuts of 퐺, and let 푒 be a quasi 푏-invariant edge of 퐺 such that the equality |푀 ∩ 퐶| + |푀 ∩ 퐷| = 2 + 2|푀 ∩ {푒}|

(14.5)

holds for each perfect matching 푀 of 퐺. Then, the cuts 퐶 and 퐷 are robust in 퐺 and have characteristic three. Proof Since both 퐶 and 퐷 are odd cuts of 퐺, any perfect matching of 퐺 has at least one edge in each of these cuts. For each perfect matching 푀 of 퐺 − 푒, the equality (14.5) implies that |푀 ∩ (퐶 − 푒)| = 1 = |푀 ∩ (퐷 − 푒)|; hence 퐶 − 푒 and 퐷 − 푒 are both tight in 퐺 − 푒. Let us now prove that 퐶 and 퐷 are both separating cuts in 퐺 and have characteristic three. By Theorem 4.2, a cut of 퐺 is a separating cut of 퐺 if and only if, given any edge 푓 of 퐺, there is a perfect matching 푀 푓 of 퐺 containing 푓 and meeting that cut in exactly one edge. Since 퐶 − 푒 and 퐷 − 푒 are tight cuts of 퐺 − 푒, for any edge 푓 ≠ 푒, this condition is clearly satisfied. Since 퐺 is a brick, neither 퐶 nor 퐷 is a tight cut of 퐺. Thus, there is a perfect matching 푀퐶 containing the edge 푒 which meets 퐶 in at at least three edges. By equation (14.5), it follows that |푀퐶 ∩ 퐶| = 3 and |푀퐶 ∩ 퐷| = 1, and hence that 퐷 is a separating cut of 퐺. Likewise, there is a perfect matching 푀퐷 containing the edge 푒 which meets 퐷 in at at least three edges. By equation (14.5), it follows that |푀퐷 ∩ 퐷| = 3 and |푀퐷 ∩ 퐶| = 1, and hence that 퐶 is a separating cut of 퐺. In sum, both cuts 퐶 and 퐷 are separating in 퐺. Furthermore, as |푀퐶 ∩ 퐶| = 3 = |푀퐷 ∩ 퐷|, we deduce that 휆(퐶) = 3 = 휆(퐷). The cut 퐶 − 푒 is tight in 퐺 − 푒 and the perfect matching 푀퐷 of 퐺 contains the edge 푒 and has just one edge in 퐶. By Corollary 14.9, the cut 퐶 is robust in 퐺. Likewise, the cut 퐷 is also robust in 퐺. ƒ

14.3 Quasi 푏-Invariant Edges

299

14.3.1 Barrier cut pairs and essentially 2-separation cut pairs For the convenience of the readers we recall here the definitions of these two types of tight cuts. Barrier cut pairs: Let 퐺 be a nonsolid brick, let 푒 be a quasi 푏-invariant edge of 퐺,

and let 퐵 be a barrier of 퐺 − 푒 such that 퐺 − 푒 − 퐵 has precisely two nontrivial odd components, 퐾 and 퐿. Then, the pair (휕 (퐾) − 푒, 휕 (퐿) − 푒) is a barrier cut pair of 퐺 − 푒. It should be noted that, since 퐺 itself is a brick, the edge 푒 must have its ends in two different odd components of 퐺 − 푒 − 퐵. See Figure 14.6. (In this figure the edge 푒 has no end in either 푉 (퐾) or 푉 (퐿). But these possibilities are not precluded.) 퐵

퐷

퐶 퐾

퐿

푒

Fig. 14.6 The case in which (퐶 − 푒, 퐷 − 푒) is a barrier cut pair of 퐺 − 푒

Essentially 2-separation cut pairs: Let 푒 be a quasi 푏-invariant edge of a nonsolid

brick 퐺 and let {푆, 푇 } be an essential 2-separation of 퐺 − 푒. Thus, (i) the two subgraphs (퐺 − 푒) [푆] and (퐺 − 푒) [푇] of 퐺 − 푒 induced by 푆 and 푇, respectively, are bipartite subgraphs of 퐺 − 푒 such that their their majority parts 푆+ and 푇+ are special barriers of 퐺 − 푒, (ii) the cut 휕 (푆) is the special barrier cut associated with the barrier 푆+ and the cut 휕 (푇) is the special barrier cut associated with the barrier 푇+ , and (iii) the pair {푠, 푡} is a 2-separation of the graph (퐺 − 푒)/(푆 → 푠)/(푇 → 푡). See Figure 14.7. Let 푒 be a quasi 푏-invariant edge of a nonsolid brick 퐺 and let 퐶 and 퐷 be two cuts of 퐺 such that 퐶 − 푒 and 퐷 − 푒 are either a barrier cut pair or an essentially 2-separation cut pair. We now establish fundamental properties of the cuts 퐶 and 퐷. Lemma 14.11 Let 푒 be a quasi 푏-invariant edge of a brick 퐺, let 퐵 be a barrier of 퐺 − 푒 such that 퐺 − 푒 − 퐵 has precisely two nontrivial (odd) components, 퐾 and 퐿. Then, the cuts 퐶 := 휕퐺 (퐾) and 퐷 := 휕퐺 (퐿) are both robust and have characteristic three. Proof Let 푀 be a perfect matching of 퐺. A simple counting argument shows that |푀 ∩ 퐶| + |푀 ∩ 퐷| + (|퐵| − 2) = 2|푀 ∩ {푒}| + |퐵|;

14 Robust Cuts in Bricks

300 퐷 − 푒 푆−

푋 ∩ 푌

푆+ 퐶 − 푒 푋 ∩ 푌

푇+

푇− Fig. 14.7 Essentially 2-separation cuts 퐶 − 푒 and 퐷 − 푒 of 퐺 − 푒: 퐶 = 휕(푋), 퐷 = 휕(푌 ), 퐼 = 휕(푋 ∩ 푌 ), and 푈 = 휕(푋 ∩ 푌 ).

hence equation (14.5) holds. By Lemma 14.10, the cuts 퐶 and 퐷 are robust and have characteristic three. ƒ Lemma 14.12 Let 푒 be a quasi 푏-invariant edge of a brick 퐺 and let (퐶 − 푒, 퐷 − 푒) be a pair of tight cuts of 퐺 associated with an essentially 2-separation {푆, 푇 } of 퐺 − 푒. If either 푆 or 푇 has just one vertex, then both 퐶 and 퐷 are robust cuts of characteristic three. Proof Let 푋 be a shore of 퐶 and 푌 be a shore of 퐷 such that 푋 ∩ 푌 = 푆. Assume that one of 푆 and 푇 has just one vertex. Adjust notation so that 푇 is a singleton. If 푆 is also a singleton then 푒 joins a vertex in 푋 ∩ 푌 to a vertex in 푋 ∩ 푌 . In that case, equality (14.5) holds, for each perfect matching 푀 of 퐺. Alternatively, suppose that 푆 is not a singleton; let 푀 be any perfect matching of 퐺. The edge 푒 must have an end in the minority part 푆 − of 푆; hence |푀 ∩ 휕 (푆)| = 1 + 2|푀 ∩ {푒}|. Moreover, no edge of 퐺 joins a vertex in 푋 ∩ 푌 to a vertex in 푋 ∩ 푌 . Thus, |푀 ∩ 퐶| + |푀 ∩ 퐷| = |푀 ∩ 휕 (푆)| + |푀 ∩ 휕 (푇)| = (1 + 2|푀 ∩ {푒}|) + 1; hence equality (14.5) holds. In both alternatives, equality (14.5) holds for each perfect matching 푀 of 퐺. By Lemma 14.10, the cuts 퐶 and 퐷 are robust and have characteristic three. ƒ Unlike the case with barrier cut pairs of tight cuts of 퐺 − 푒, the fact that 퐶 − 푒 and 퐷 − 푒 are an essentially 2-separation cut pair does not ensure that 퐶 and 퐷 are both separating cuts of 퐺; see Exercise 14.3.2. The fact below plays an important role in our proof of Lov´asz’s conjecture which is presented in the next chapter.

14.4 Notes

301

A consequence of Corollary 5.23 Suppose that 푒 is a quasi-푏-invariant edge in a nonsolid brick 퐺, and let 퐶 − 푒 be a nontrivial tight cut of 퐺 − 푒 such that both (퐶 − 푒)-contractions of 퐺 − 푒 are near-bricks. Then it follows from Corollary 5.23, that 퐶 − 푒 is either a member of a barrier cut pair or is a member of an essentially 2-separation cut pair.

Exercises ∗ 14.3.1 Give a proof of Lemma 14.8. 14.3.2 Consider the brick 퐺, the quasi 푏-invariant edge 푒 in 퐺, and the two cuts 퐶 and 퐷 shown in Figure 14.8. Show that (i) (퐶 − 푒, 퐷 − 푒) is an essentially 2separation pair of tight cuts of 퐺 − 푒, (ii) 퐷 is not a separating cut of 퐺 although 퐶 is. 퐷 2

3

4

2

퐷 − 푒 3 4

푒 5

1

6

10

9

8 퐺

7

5

1 퐶

6

10

9

8 퐺 − 푒

퐶 − 푒

7

Fig. 14.8 The pair (퐶 − 푒), 퐷 − 푒) is an essentially 2-separation cut pair, but 퐷 is not separating in 퐺.

14.4 Notes The first published proof of the validity of Lov´asz’s conjecture (Theorem 15.1) appeared in Carvalho’s Ph. D. Thesis (1996, [6]), written under the supervision of Lucchesi and cosupervision of Murty. Later, it was published in a paper, by CLM (2002, [10]). That proof relied on a result concerning characteristics of bricks. It was shown (Theorem 4.1 in [10]) that if 퐺 is any nonsolid brick that is not a Petersen brick, then 휆(퐺) = 3. In other words, every such brick has a separating cut 퐶, and a perfect matching 푀0 , such that |푀0 ∩ 퐶| = 3. The inductive proof of that

302

14 Robust Cuts in Bricks

result made crucial use of the fact that if 푅 is any removable class of a matching covered graph 퐺, then 휆(퐺) ≤ 휆(퐺 − 푅) (Exercise 10.2.6). Thus, to prove the required result by induction, it sufficed to show that any nonsolid brick 퐺 distinct from the Petersen graph either has a removable doubleton or has a (푏 + 푝)-invariant edge 푒 such that 퐺 − 푒 is nonsolid. An unpublished report due to Campos and Lucchesi (2000, [5]) established the stronger result that if 퐺 is any nonsolid brick that is not a Petersen brick, and 퐶 is any nontrivial separating cut of 퐺, then there exists a perfect matching 푀0 of 퐺 such that |푀0 ∩ 퐶| = 3.

Chapter 15

풃-Invariant Edges in Bricks

Contents 15.1 15.2

15.3

Existence of 푏-Invariant Edges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303 Proof of Theorem 15.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 304 15.2.1 Reduction to simple nonsolid bricks . . . . . . . . . . . . . . . . . . . 304 15.2.2 Definitions of 푋, 퐶, 푀0 , 퐺 1 and 퐺 2 . . . . . . . . . . . . . . . . . . . 305 15.2.3 Reduction to the case in which 퐺 1 is an odd wheel . . . . . . . 306 15.2.4 Reduction to the case in which 퐺 2 is a brick . . . . . . . . . . . . 307 15.2.5 Reduction to the case in which 퐺 2 is not a special brick . . 309 15.2.6 Choosing a 푏-invariant edge 푒 of 퐺 2 . . . . . . . . . . . . . . . . . . 310 15.2.7 Reduction to the case in which 푒 is a quasi 푏-invariant edge of 퐺 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 311 15.2.8 Reduction to case in which {퐶 − 푒, 퐷 − 푒} is an essentially 2-separation cut pair of tight cuts of 퐺 − 푒 . . . . 311 15.2.9 Working with the hypothesis that (퐶 − 푒, 퐷 − 푒) is an essentially 2-separation cut pair of tight cuts of 퐺 − 푒 . . . . 313 15.2.10 Reduction to the case in which 퐺 [푋] and 퐺 [푌 ] are pentagons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 315 15.2.11 The conclusion: 퐺 is the Petersen graph . . . . . . . . . . . . . . . 317 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 321

15.1 Existence of 풃-Invariant Edges Using the results on robust cuts in nonsolid bricks established in the previous chapter, we now proceed to present a proof of Lov´asz’s Conjecture 13.2:

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7_15

303

15 푏-Invariant Edges in Bricks

304

The main theorem Theorem 15.1 Every brick distinct from 퐾4 , 퐶6 and the Petersen graph has a 푏-invariant edge. For convenience we shall refer to 퐾4 , 퐶6 , and P as special bricks. The evolution of the approach to the Conjecture 13.2, as explained in Chapter 13, started with the observation that if a removable edge in a brick is not 푏-invariant, then that brick must have a nontrivial separating cut or, equivalently, that any removable edge in a solid brick is 푏-invariant. This led to the notion of robust cuts and to the consideration of the use of contractions with respect to a robust cut as an inductive tool. All this has been described in Chapter 13. The proof we present here relies on many of the ideas developed in Carvalho’s Ph. D. Thesis (1996, [6]), and in subsequently published, in streamlined and refined form, in papers by CLM (2002, [9, 10]). It also relies on results published some years later in CLM (2005, [13]), and in Lucchesi, Carvalho, Kothari and Murty (2018,[64]). However, the very elegant climactic ending is new.

15.2 Proof of Theorem 15.1 ♯ The idea of the proof is to assume that (i) 퐺 is a brick distinct from 퐾4 and 퐶6 and (ii) 퐺 has no 푏-invariant edge and deduce that 퐺 is the Petersen graph. Induction hypothesis We shall prove Theorem 15.1 by assuming, inductively, that if 퐺 ′ is any brick that is not one of the three special bricks, with |푉 (퐺 ′ )| < |푉 (퐺)|, then 퐺 ′ has a 푏-invariant edge.

15.2.1 Reduction to simple nonsolid bricks Every solid brick 퐺 distinct from 퐾4 has removable edges by Theorem 9.14, and by Theorem 10.16, every removable edge of such a brick 퐺 is 푏-invariant. Also, if 퐺 is a brick with multiple edges then clearly any member of a pair of parallel edges is 푏-invariant in 퐺. We may thus suppose that 퐺 is a simple nonsolid brick distinct from 퐶6 .

305

15.2 Proof of Theorem 15.1

The idea of reductions Our proof of Theorem 15.1 involves series of steps which we refer to as reductions. In each of these steps, we show that unless 퐺 has a certain property that the Petersen graph has, there is some edge of 퐺 that is 푏-invariant in 퐺. The property that 퐺 is a simple nonsolid brick distinct from 퐶6 is an example of reduction. Let 퐺 be a simple nonsolid brick distinct from 퐶6 . By Corollary 14.4, 퐺 has robust cuts. We describe below how we choose a robust cut and also the notation that we use throughout the rest of the proof.

15.2.2 Definitions of 푿 , 푪 , 푴0 , 푮 1 and 푮 2 • Let 푋 be a set of vertices of 퐺 having minimum cardinality such that 퐶 := 휕 ( 푋) is robust. • Let 푀0 be a perfect matching of 퐺 such that |푀0 | = 휆(퐶). (Thus, among the perfect matchings of 퐺 that meet 퐶 in more than one edge, 푀0 has the fewest number of edges in common with 퐶.) • Let 퐺 1 := 퐺/( 푋 → 푥) and 퐺 2 := 퐺/( 푋 → 푥) denote the two 퐶-contractions of 퐺. The following assertion records an immediate consequence of the minimality of | 푋 | and Lemma 14.6: 15.1.1 The graph 퐺 1 is a solid brick.

ƒ

We shall prove that if 퐺 has no 푏-invariant edges and is distinct from 퐶6 then |푀0 ∩ 퐶| = 5 and 퐺 is the Petersen graph. The following proposition is crucial to the inductive proof of this statement. Proposition 15.2 Let 푒 be an edge of 퐺 such that both 퐺 1 − 푒 and 퐺 2 − 푒 are near-bricks. If 푒 ∉ 푀0 then 푒 is a 푏-invariant edge of 퐺. Proof By hypothesis, 푒 is not in 푀0 . Thus, 푀0 is a perfect matching of 퐺 − 푒. As 푀0 has three or more edges in 퐶, it follows that 퐶 − 푒 is not a tight cut of 퐺 − 푒. By hypothesis, both (퐶−푒)-contractions of 퐺−푒 are near-bricks. Hence, by subadditivity (Theorem 4.20), we deduce that 퐺 − 푒 is also a near-brick. ƒ

15 푏-Invariant Edges in Bricks

306

15.2.3 Reduction to the case in which 푮 1 is an odd wheel By statement 15.1.1, the 퐶-contraction 퐺 1 is a solid brick. In establishing the following assertion, we use the results concerning removable edges in solid bricks which are not odd wheels (up to multiple edges), described in Chapter 10, specifically Lemma 10.10 on Wheels. 15.2.1 Suppose that 퐺 1 is a solid brick that is not an odd wheel, up to multiple edges. Then some edge 푒 of 퐺 1 is 푏-invariant in 퐺. Proof The graph 퐺 is simple. Thus, any multiple edges of 퐺 1 must be in 퐶. We may then obtain an underlying simple graph, 퐺 ′1 , of 퐺 1 , by removing multiple edges of 퐺 1 in 퐶 − 푀0 . By Lemma 10.10 applied to 퐺 ′1 , the graph 퐺 ′1 has a removable edge, say 푒, which is not in 푀0 ∪ 퐶. The edge 푒 is removable in 퐺 1 , and since 퐺 1 is a solid brick, it follows from Theorem 10.16 that 푒 is 푏-invariant in 퐺 1 . The graph 퐺 2 − 푒 is also a near-brick, because 푒 ∉ 퐸 (퐺 2 ). Since 푒 ∉ 푀0 , it follows from Proposition 15.2 that 푒 is 푏-invariant in 퐺. ƒ We may thus assume that 퐺 1 is an odd wheel, up to multiple edges. Since 퐺 itself is simple, any multiple edges that 퐺 1 has must be incident with its hub 푥. If | 푋 | ≥ 5, then every edge of 퐺 1 incident with 푥 is 푏-invariant in 퐺 1 . On the other hand, if | 푋 | = 3, an edge incident with 푥 is 푏-invariant in 퐺 1 if and only if it is a multiple edge. The following statement analyses the case in which an edge 푒 incident with 푥 is 푏-invariant in 퐺 2 but not 푏-invariant in 퐺 1 .

Lemma 15.3 Let 푒 be an edge of 퐺 2 . If 푒 is 푏-invariant in 퐺 2 but is not 푏-invariant in 퐺 1 then | 푋 | = 3, 푒 ∈ 푀0 ∩ 퐶 and 퐺 1 has an edge that is 푏-invariant in 퐺. Proof Suppose that 푒 is 푏-invariant in 퐺 2 but not 푏-invariant in 퐺 1 . Every edge of 퐺 2 that is not in 퐶 is 푏-invariant in 퐺 1 . Thus, 푒 is in 퐶. The edge 푒 is in 퐶 and is not 푏-invariant in 퐺 1 . As 퐺 1 is a wheel, this implies that the underlying simple graph of 퐺 1 is 퐾4 and also that 푒 is not a multiple edge of 퐺 1 . It follows that | 푋 | = 3. Let 푢, 푣 and 푤 denote the three vertices of 푋 and adjust notation so that 푒 is incident with vertex 푢. See Figure 15.1. 푣

푓

푤 퐶

푢 푒 Fig. 15.1 Edge 푒 is in 퐶 and not 푏-invariant in 퐺1 . Solid lines indicate the edges of 푀0 ∩ 퐶

15.2 Proof of Theorem 15.1

307

The graph 퐺 1 has precisely one edge, 푓 , that joins the vertices 푣 and 푤. Let 푅 := {푒, 푓 }. One of the (퐶 − 푅)-contractions of 퐺 − 푅 is 퐺 1 − 푅, which is 퐶4 , up to multiple edges. The other (퐶 − 푅)-contraction of 퐺 − 푅 is 퐺 2 − 푒, a near-brick. Thus, the graph 퐺 − 푅 is a near-brick. Moreover, the cut 퐶 − 푒 is tight in 퐺 − 푅. As 푀0 has more than one edge in 퐶, and since 푒 is not a multiple edge in 퐶, it follows that |푀0 ∩ 퐶| = 3 and 푒 ∈ 푀0 . Moreover, 푓 ∉ 푀0 . Thus, 푀0 is a perfect matching of 퐺 − 푓 that contains edge 푒. As 퐺 − 푅 is matching covered, it follows that 퐺 − 푓 is matching covered. Thus, 푓 is removable in 퐺. By the monotonicity of ƒ function 푏, 1 ≤ 푏(퐺 − 푓 ) ≤ 푏(퐺 − 푅) = 1; hence 푓 is 푏-invariant in 퐺. Corollary 15.4 Let 푒 be an edge of 퐺 2 . If 푒 is 푏-invariant in 퐺 2 and is not in 푀0 then 푒 is 푏-invariant in 퐺 1 and in 퐺. Proof Assume that 푒 is 푏-invariant in 퐺 2 and is not in 푀0 . As 푒 is not in 푀0 , then 푒 is 푏-invariant in 퐺 1 , by Lemma 15.3. By Proposition 15.2, 푒 is 푏-invariant in 퐺.ƒ The following results rely on the previous two statements. Conditions under which the existence of a 푏-invariant edge in 퐺 2 implies the existence of a 푏-invariant edge in 퐺 Lemma 15.5 Let 푒 be a 푏-invariant edge of 퐺 2 . The following properties hold: (i) if 푒 is not 푏-invariant in 퐺 1 then 푒 is not a multiple edge of 퐺 1 , | 푋 | = 3 and 퐺 1 has an edge that is 푏-invariant in 퐺; and (ii) if 푒 is not in 푀0 then 푒 is 푏-invariant in 퐺 1 and in 퐺. ƒ The graph 퐺 2 , being a 퐶-contraction of 퐺 with respect to a robust cut, is a nearbrick. We would like to show that, unless 퐺 has a 푏-invariant edge, 퐺 2 is also an odd wheel. We are able to achieve this only by gradually unravelling the structure of 퐺.

15.2.4 Reduction to the case in which 푮 2 is a brick 15.5.1 If 퐺 2 is not a brick then some edge of 퐺 is 푏-invariant in 퐺. Proof Suppose that 퐺 2 is not a brick. Then, it has nontrivial tight cuts. Let 푌 be a minimal proper superset of 푋 such that 퐷 := 휕 (푌 ) is a nontrivial tight cut of 퐺 2 . The two shores of the cut 퐷 in 퐺 2 are 푌 ′ := (푌 − 푋) +푥 and 푌 , and the two 퐷-contractions of 퐺 2 are 퐻 := 퐺 2 /(푌 → 푦) and 퐻 ′ := 퐺 2 /(푌 ′ → 푦 ′ ). Since 퐺 2 is a near-brick and 퐷 is a tight cut of 퐺 2 , one of 퐻 and 퐻 ′ is bipartite. But 퐻 ′ cannot be bipartite because, if it were, one of its parts would be free of contraction vertices, and that part would be a nontrivial barrier of 퐺. This is impossible because 퐺 is a brick. Therefore 푌 ′ is the bipartite shore of 퐷, and 퐻 is bipartite with bipartition ( 퐴, 퐵), where 퐴 is the majority part 푌+′ of 푌 ′ , and 퐵 is the minority part 푌−′ of 푌 ′ together with 푦. Thus

15 푏-Invariant Edges in Bricks

308 퐴

퐶

퐵

replacements 푋

푥 푣

푣

퐷 푦 푌

(푎) 퐺

(푏) 퐻

Fig. 15.2 (a) Cuts 퐶 and 퐷 in 퐺; (b) 퐻 = 퐺2 /푌 . (If 푣 ≠ 푦 is any vertex in 퐵, then some edge incident with 푣 is 푏-invariant in both 퐺2 and 퐺.)

the contraction vertex 푥 belongs to 퐴 and the contraction vertex 푦 belongs to 퐵; all other vertices of 퐻 are vertices of 퐺. See Figure 15.2. The minimality of 푌 implies that the bipartite graph 퐻 = 퐺 2 /(푌 → 푦) is a brace. This can be proved quite easily by showing that if 퐻 is not a brace, then there is some set 푍 , with 푋 ⊂ 푍 ⊂ 푌 such that 휕 (푍) is a nontrivial tight cut of 퐺 2 (Exercise 15.2.1). Let 푣 ≠ 푦 be a vertex of 퐺 that belongs to the same part of the bipartition of 퐻 to which 푦 belongs. Suppose that, in 퐺, two or more edges join 푣 to vertices of 푋. Let 푒 and 푓 be two such edges incident with 푣. Since 푒 and 푓 are adjacent, not both can belong to the perfect matching 푀0 . Adjust notation so that 푒 ∉ 푀0 . In 퐺 2 , the edges 푒 and 푓 are multiple edges. Thus, 푒 is 푏-invariant in 퐺 2 . Moreover, edge 푒 does not belong to 푀0 . By Lemma 15.5, edge 푒 is 푏-invariant in 퐺. So, suppose that, in 퐺, vertex 푣 is adjacent to at most one vertex of 푋. As 퐺 is a brick, the vertex 푣 is adjacent to three or more vertices in 퐺. Thus, in 퐻, vertex 푣 is adjacent to two or more vertices of 퐴 − 푥. It follows that 퐻 has six or more vertices. As has already been noted, 퐻 is a brace. Thus, every edge incident with 푣 is removable in 퐻. This clearly implies that there is an edge 푒 which is incident with 푣 that does not belong to 퐶 ∪ 퐷 ∪ 푀0 . It is now easy to see that 푒 is 푏-invariant in both 퐺 2 and 퐺. (Exercise 15.2.2). ƒ We may thus assume that 퐺 2 is a brick.

15.2 Proof of Theorem 15.1

309

15.2.5 Reduction to the case in which 푮 2 is not a special brick 15.5.2 Suppose that brick 퐺 2 is a special brick. Then 퐺 2 is not 퐾4 ; it must be either 퐶6 or the Petersen graph. In either case, 퐺 has a 푏-invariant edge. Proof All the three special bricks are cubic. Thus, if 퐺 2 is a special brick, then |퐶| = 3 implying that 퐺 1 = 퐾4 . If 퐺 2 is 퐾4 then 퐺 is 퐶6 , a contradiction. If 퐺 2 is 퐶6 then 퐺 is the bicorn (Figure 15.3) and the only removable edge of 퐺 is 푏-invariant in 퐺. 퐶 푒

푋

푋

Fig. 15.3 Edge 푒 is the only 푏-invariant edge of the bicorn

Suppose that 퐺 2 is the Petersen graph (Figure 15.4).

푋

퐶

Fig. 15.4 Graph 퐺 in the case 퐺2 = P. Solid lines indicate 푏-invariant edges of 퐺

It is easy to prove that 퐺 has nine 푏-invariant edges (Exercise 15.2.3).

ƒ

15 푏-Invariant Edges in Bricks

310

15.2.6 Choosing a 풃-invariant edge 풆 of 푮 2 The reductions we have described so far allow us to assume that 퐺 1 is an odd wheel (up to multiple edges in 퐶) and that 퐺 2 is a brick, but not one of the three special bricks. By induction, 퐺 2 has a 푏-invariant edge. We select a 푏-invariant edge 푒 of 퐺 2 according to the following criterion: Rule for choosing 푏-invariant edge 푒 of 퐺 2 If 퐺 2 has 푏-invariant edges in 퐶, then choose 푒 to be an edge in 퐶. At this stage, we bring in a second perfect matching of 퐺 into play. Perfect matching 푀1 Let 푀1 be a perfect matching of 퐺 that contains edge 푒 and just one edge in 퐶. (Since 퐶 is a separating cut of 퐺, such a perfect matching must exist by Theorem 4.2.) We have assumed that 퐺 does not contain 푏-invariant edges. In particular, 푒 is not 푏-invariant in 퐺. Lemma 15.5 immediately implies the following consequence: Corollary 15.6 푒 ∈ 푀0 ∩ 푀1 .

ƒ

As the chosen 푏-invariant edge 푒 of 퐺 2 is not 푏-invariant in 퐺, it does not necessarily mean that 퐺 has no 푏-invariant edges. Thus it helps to have some flexibility in selecting a 푏-invariant edge of 퐺 2 . The following result is relevant in this context. Lemma 15.7 Any edge of 퐺 2 − 푒 which does not belong to 푀1 and is 푏-invariant in 퐺 2 − 푒 is also 푏-invariant in 퐺 2 . Proof Let 푓 ∉ 푀1 be an edge of 퐺 2 that is 푏-invariant in 퐺 2 −푒. Edge 푒 is 푏-invariant in 퐺 2 ; thus, 퐺 2 − 푒 is a near-brick. Edge 푓 is 푏-invariant in 퐺 2 − 푒; consequently, 퐺 2 − 푒 − 푓 is a near-brick. As matching 푀1 does not contain edge 푓 , 푀1 is a perfect matching of 퐺 2 − 푓 that contains edge 푒. We deduce that graph 퐺 2 − 푓 is matching covered. Moreover, by the monotonicity of function 푏, we have that 1 = 푏(퐺 2 ) ≤ 푏(퐺 2 − 푓 ) ≤ 푏(퐺 2 − 푒 − 푓 ) = 1. Hence, 푓 is 푏-invariant in 퐺 2 . This conclusion holds for each edge 푓 of 퐺 2 that is 푏-invariant in 퐺 2 − 푒 and is not in 푀1 . ƒ

15.2 Proof of Theorem 15.1

311

15.2.7 Reduction to the case in which 풆 is a quasi 풃-invariant edge of 푮 Suppose that the 푏-invariant edge 푒 of 퐺 2 is not 푏-invariant in the wheel 퐺 1 . By Lemma 15.5, the graph 퐺 1 contains an edge that is 푏-invariant in 퐺. We may thus assume that edge 푒 is 푏 -invariant in 퐺 1 .

The edge 푒, by our choice, is 푏-invariant in 퐺 2 . Thus, 퐺 1 − 푒 and 퐺 2 − 푒 are both near-bricks. If the cut 퐶 − 푒 is not tight in 퐺 − 푒 then, by the subadditivity of function 푏, the edge 푒 is 푏-invariant in 퐺. Hence we may assume that 퐶 − 푒 is tight in 퐺 − 푒. In this case, 푒 is not 푏-invariant in 퐺. By Lemma 15.5, edge 푒 lies in 푀0 . Moreover, 푏(퐺 − 푒) = 푏(퐺 1 − 푒) + 푏(퐺 2 − 푒) = 2. This implies that 푒 is a quasi 푏-invariant edge of 퐺. The above analysis shows that if 퐺 has no 푏-invariant edges, then the edge 푒, chosen to be a 푏-invariant edge of 퐺 2 , has to be a quasi 푏-invariant edge of 퐺. So, we may suppose that 푏(퐺 − 푒) = 2. In this case, as explained in Section 14.3, there exists a cut 퐷 := 휕 (푌 ) of 퐺 such that {퐶 − 푒, 퐷 − 푒} is either a barrier cut pair or an essentially 2-separation cut pair of 퐺 − 푒. The reduction described below eliminates the first possibility. In this reduction, and the rest of the proof of the Main Theorem, Lemma 8.19 plays a crucial role. The readers should review that lemma carefully before proceeding further.

15.2.8 Reduction to case in which {푪 − 풆, 푫 − 풆} is an essentially 2-separation cut pair of tight cuts of 푮 − 풆 Suppose that {퐶 − 푒, 퐷 − 푒} is a barrier cut pair of tight cuts of 퐺 − 푒 associated with a barrier 퐵 of 퐺 − 푒. By Lemma 14.11, the cut 퐷 is robust, and both cuts 퐶 and 퐷 have characteristic three in 퐺. The cut 퐷 is a nontrivial cut of 퐺 in one of the 퐶-contractions of 퐺. As the brick 퐺 1 is solid, it must be the case that 퐷 is a separating cut of 퐺 2 . Adjust notation so that 푌 is the shore of 퐷 that is a subset of 푋. (See Figure 15.5.) Let 퐻 denote the bipartite graph (퐺 − 푒)/( 푋 → 푥)/(푌 → 푦). Clearly, 퐻 = (퐺 2 − 푒)/(푌 → 푦). Proposition 15.8 |퐶 − 푒| ≥ 3. Proof If |퐶| ≥ 4 then the inequality holds. Assume thus that |퐶| = 3. In that case, | 푋 | = 3, 퐺 1 = 퐾4 and no edge of 퐶 is removable in 퐺 1 . The cut 퐶 − 푒 is tight in 퐺 − 푒; thus 푒 ∉ 퐶. We deduce that 퐶 − 푒 = 퐶; hence the inequality holds with equality. ƒ

15 푏-Invariant Edges in Bricks

312 퐵

퐷

퐶 푋

푌

푒

Fig. 15.5 The case in which (퐶 − 푒, 퐷 − 푒) is a barrier cut pair of 퐺 − 푒. The graph 퐻 = (퐺 − 푒)/(푋 → 푥 )/(푌 → 푦 ) = (퐺2 − 푒)/(푌 → 푦 ) is bipartite.

Lemma 15.9 At most one edge of 퐶 − 푀1 is not 푏-invariant in 퐺 2 . Proof By Proposition 15.8, |퐶 − 푒| ≥ 3. We may thus suppose that the degree of 푥 in 퐻, which is equal to |퐶 − 푒|, is at least three. By Lemma 8.19 (with the brick 퐺 2 and the cut 퐷, respectively, playing the roles of 퐺 and 퐶 in the statement of that lemma), at most one edge of 퐶 − 푒 is not removable in 퐻. By definition, 푀1 is a perfect matching containing 푒 and meeting 퐶 in exactly one edge. If 푒 ∈ 퐶 then 퐶 − 푀1 = 퐶 − 푒. Alternatively, if 푒 ∉ 퐶 then at most one edge of 퐶 is not removable in 퐻. In both alternatives, at most one edge of 퐶 − 푀1 is not removable in 퐻. Let 푓 be any edge of 퐶 − 푀1 that is removable in 퐻. Then, as 퐶 − 푒 and 퐷 − 푒 are disjoint, it follows that 푓 does not belong to cut 퐷, and therefore that 푓 is 푏-invariant in 퐺 2 − 푒. Moreover, as 푓 does not belong to 푀1 , it follows from Lemma 15.7 that edge 푓 is 푏-invariant in 퐺 2 . This conclusion holds for each edge of 퐶 − 푀1 that is removable in 퐻. ƒ We first observe that, by the criterion used for choosing 푒, the above conclusion implies that edge 푒 belongs to 퐶. Thus, 푒 is the edge of 퐶 ∩ 푀1 ; hence, again by the above conclusion, at most one edge of 퐶 − 푒 is not 푏-invariant in 퐺 2 . Moreover, as 푒 ∈ 푀0 , we have that 푒 is the edge of 푀0 ∩ 푀1 ∩ 퐶; hence 퐶 − 푀0 − 푀1 = 퐶 − 푀0 . Suppose that |퐶| ≥ 5. In that case, 퐶 − 푀0 consists of two or more edges, neither of which is 푒. Thus, 퐶 − 푀0 contains an edge, 푒 ′ , which is 푏-invariant in 퐺 2 . By Lemma 15.5, 푒 ′ is 푏-invariant in 퐺 1 and in 퐺. Consider now the case in which |퐶| ≤ 4. Hence | 푋 | = 3 and 퐺 1 is 퐾4 , up to multiple edges in 퐶. The edge 푒 is in 퐶 and the cut 퐶 − 푒 is tight in 퐺 − 푒; hence 푒 is a multiple edge in 퐺 1 . It follows that |퐶| = 4. The doubleton 푀0 ∩ 퐶 − 푒 contains an edge, 푒 ′′ , which is 푏-invariant in 퐺 2 . Moreover, 푒 ′′ is not removable in 퐺 1 . By Lemma 15.5, 퐺 1 contains an edge which is 푏-invariant in 퐺. We may thus assume that 푒 is a quasi 푏-invariant edge of 퐺, and also assume that the cut 퐶 − 푒 is tight in 퐺 − 푒 but is not a member of a barrier cut pair of 퐺 − 푒. Therefore, the brick 퐺 has a cut 퐷 such that (퐶 − 푒, 퐷 − 푒) is an essentially 2-separation cut pair of 퐺 − 푒.

15.2 Proof of Theorem 15.1

313

15.2.9 Working with the hypothesis that (푪 − 풆, 푫 − 풆) is an essentially 2-separation cut pair of tight cuts of 푮 − 풆 Let {푆, 푇 } be an essential 2-separation of 퐺 − 푒 such that 퐶 − 푒 and 퐷 − 푒 are the two tight cuts of 퐺 − 푒 associated with {푆, 푇 }. Adjust notation so that 푆 = 푋 ∩ 푌 and 푇 = 푋 ∩ 푌 and, as usual, let 퐼 and 푈 denote, respectively, 휕 (푆) and 휕 (푇). Subgraphs 퐺 [푋 ∩ 푌 ] and 퐺 [푋 ∩ 푌] As 퐺 1 := 퐺/( 푋 → 푥) is an odd wheel, it follows that: (i) the subgraph induced by 푋 ∩ 푌 is a path of even length and (ii) the subgraph induced by 푋 ∩ 푌 is one of odd length. Furthermore, (iii) the cardinality of 푋 ∩ 푌 is equal to either one or three because in 퐺 − 푒 the vertices in 푆 − have no neighbours outside 푆+ . At this stage, all that we can say is that | 푋 ∩ 푌 | is a positive even integer. But assuming that 퐺 has no 푏-invariant edges, we will show that | 푋 ∩ 푌 | and | 푋 ∩ 푌 | are, respectively 3 and 2 and thereby deducing that 퐺 [푋] is a pentagon. (This requires several intermediate steps.) The subset 퐹 of the cut 퐶 We denote by 퐹 the set of edges in the cut 퐶 which have both their ends in 푌 . In other words, 퐹 = 휕 ( 푋 ∩ 푌 ) ∩ 푈. See Figures 15.6(a) and 15.6(b) where the edges in 퐹 are indicated by dotted lines. 푌

푥

푋 퐶 − 푒

퐶 − 푒 푋

퐷 − 푒 (a) 퐺 − 푒 Fig. 15.6 The graphs 퐺 − 푒 and 퐺2 − 푒

푌 (b) 퐺2 − 푒

15 푏-Invariant Edges in Bricks

314

Lemma 15.10 |퐹 − 푒| ≥ | 푋 ∩ 푌 | − 1, with equality only if | 푋 ∩푌 | ≥ 4. Consequently, |퐹 − 푒| ≥ 2. Proof Up to multiple edges in 퐶, the graph 퐺 1 is an odd wheel having the contraction vertex 푥 as a hub. Thus, every vertex of 푋 ∩ 푌 is incident with an edge of 퐶. Since 퐶 − 푒 and 퐷 − 푒 are two tight cuts of 퐺 − 푒 that cross, by Theorem 4.12 it follows that there is no edge of 퐺 − 푒 with one end in 푋 ∩ 푌 and one end in 푋 ∩ 푌 . If | 푋 ∩ 푌 | ≥ 4 then |퐹 − 푒| ≥ | 푋 ∩ 푌 | − 1. If no end of 푒 is in 푋 ∩ 푌 then |퐹 − 푒| ≥ | 푋 ∩ 푌 |. We may thus assume that | 푋 ∩ 푌 | = 2 and 푒 is incident with a vertex in 푋 ∩ 푌 . In that case, as (i) 푒 ∈ 퐶, (ii) the cut 퐼 = 휕 (푆) is tight in 퐺 − 푒 and (iii) 퐺 [푆] − 푒 is bipartite, it follows that 푆 is a singleton. Thus, | 푋 | = 3. The graph 퐺 1 is then 퐾4 , up to multiple edges in 퐶. The edge 푒 is 푏-invariant in 퐺 1 ; hence it is a multiple edge in 퐺 1 . We deduce that |퐹 − 푒| ≥ 2 = | 푋 ∩ 푌 |. ƒ The two (푈 − 푒)-contractions of 퐺 2 − 푒 Now we consider the two (푈 − 푒)-contractions 퐻 ′ := (퐺 2 − 푒)/( 푋 ∩ 푌 ) and 퐻 ′′ := (퐺 2 − 푒)/(( 푋 ∩ 푌 ) + 푥) of the graph 퐺 2 − 푒. See Figure 15.7. 푥

퐶 − 푒 푢

푋

푋 ∩ 푌 푌 (a) 퐻 ′ = (퐺2 − 푒)/(푋 ∩ 푌 )

(b) 퐻 ′′ = (퐺2 − 푒)/( (푋 ∩ 푌 ) + 푥 )

Fig. 15.7 The two (푈 − 푒)-contractions of 퐺2 − 푒

The following statement is an obvious observation (Exercise 15.2.4). Lemma 15.11 The bipartite graph 퐻 ′′ = (퐺 2 − 푒)/(( 푋 ∩ 푌 ) + 푥) is isomorphic to the (푈 − 푒)-contraction (퐺 − 푒)/( 푋 ∪ 푌 ) of 퐺 − 푒. ƒ Recall that 푀0 is a perfect matching of 퐺 such that |푀0 ∩ 퐶| = 휆(퐶) ≥ 3 and that 푀1 is a perfect matching of 퐺 such that 푒 ∈ 푀1 and |푀1 ∩ 퐶| = 1. The following result is analogous to Lemma 15.9.

15.2 Proof of Theorem 15.1

315

Lemma 15.12 Every edge of 퐹 − 푀1 is 푏-invariant in 퐺 2 . Proof Let 푓 be any edge of 퐹 that does not belong to 푀1 . As the edge 푒 does belong to 푀1 , it follows that 푓 belongs to 퐹 − 푒. By Lemma 8.19 (with brick 퐺 2 playing the role of 퐺 and 푈 the role of 퐶), every edge of 푈 − 푒 is removable in the (bipartite) (푈 − 푒)-contraction 퐻 ′′ = (퐺 2 − 푒)/(( 푋 ∩ 푌 ) + 푥) of 퐺 2 − 푒 (Figure 15.7(b)). In the other (푈 − 푒)-contraction 퐻 ′ := (퐺 2 − 푒)/( 푋 ∩ 푌 ) of 퐺 2 − 푒 (Figure 15.7(a)), from Lemma 15.10 it follows that all edges in 퐹 − 푒 are multiple edges and hence are removable. As 푓 ∈ 퐹 − 푒, it follows that the edge 푓 is removable in 퐻 ′ as well. In sum, 푓 is removable in both (푈 − 푒)-contractions of 퐺 2 − 푒. Thus, 푓 is removable in 퐺 2 − 푒. The set 푋 ∩ 푌 is the bipartite shore of the tight cut 푈 − 푒 of 퐺 2 − 푒. The graph 퐺 2 − 푒 is a near-brick and 퐻 ′ is the (푈 − 푒)-contraction (퐺 2 − 푒)/( 푋 ∩ 푌 ) of 퐺 2 − 푒. Thus, 퐻 ′ is a near-brick. The edge 푓 is a multiple edge of 퐻 ′ ; hence 퐻 ′ − 푓 is a near-brick. That is, the (푈 − 푒 − 푓 )-contraction 퐻 ′ − 푓 of 퐺 2 − 푒 − 푓 is a near-brick. The other (푈 − 푒 − 푓 )-contraction of 퐺 2 − 푒 − 푓 is the bipartite matching covered graph 퐻 ′′ − 푓 . We conclude that 퐺 2 − 푒 − 푓 is a near-brick. That is, 푓 is 푏-invariant in 퐺 2 − 푒. By hypothesis, 푓 does not belong to 푀1 . Therefore, by Lemma 15.7, edge 푓 is 푏-invariant in 퐺 2 . This conclusion holds for each edge 푓 in 퐹 − 푀1 . ƒ Corollary 15.13 푒 ∈ 퐶. Proof Suppose that 푒 ∉ 퐶. By Lemma 15.10, |퐹| ≥ 2; hence 퐹 − 푀1 is nonempty. By Lemma 15.12 we have that 퐹 contains an edge, 푓 , that is 푏-invariant in 퐺 2 . The criterion adopted for choosing the edge 푒 would have given preference to 푓 ; a contradiction. ƒ

Reduction to the case in which 푭 ⊆ 푴0 Lemma 15.14 Any edge in 퐹 − 푀0 is 푏-invariant in 퐺. Proof Let 푓 be an edge in 퐹 − 푀0 . The edge 푒 is in 푀0 and is the only edge of 푀1 ∩ 퐶. Thus, 푓 ∈ 퐹 − 푀1 . By Lemma 15.12, 푓 is 푏-invariant in 퐺 2 . As 푓 ∉ 푀0 we conclude that 푓 is 푏-invariant in 퐺 1 and also 푏-invariant in 퐺, by Lemma 15.5. This conclusion holds for each edge 푓 ∈ 퐹 − 푀0 . ƒ We may thus assume that every edge of 퐹 is an edge in the perfect matching 푀0 . This clearly implies that 퐹 is a matching!

15.2.10 Reduction to the case in which 푮[푿] and 푮[풀] are pentagons In this final reductive step we shall show, with the proviso that 퐺 has no 푏-invariant edges, that both the subgraphs 퐺 [푋] and 퐺 [푌 ] of 퐺, induced by 푋 and 푌, respectively, are pentagons. For this, it is convenient to establish the following somewhat technical result.

15 푏-Invariant Edges in Bricks

316

Lemma 15.15 The cut 푈 is nontrivial, | 푋 ∩ 푌 | = 2 and the edge 푒 has one end in the minority part of the shore 푇 of 푈 and one end in 푋 ∩ 푌 . Proof Clearly, 퐹 ⊆ 푈. By Lemma 15.14, we have that 퐹 ⊆ 푀0 . Hence 퐹 − 푒 ⊆ 푀0 ∩ 푈 − 푒.

(15.1)

By Lemma 15.10, |퐹 − 푒| ≥ 2. It follows that 2 ≤ |퐹 − 푒| ≤ |푀0 ∩ 푈 − 푒|.

(15.2)

Hence 푈 is nontrivial and the shore 푇 of 푈 has three or more vertices. The subgraph 퐺 [푇] of 퐺 induced by the shore 푇 of 푈 is bipartite; hence its majority part 푇+ is a nontrivial barrier of 퐺 − 푒. As 퐺 is a brick, the edge 푒 has an end in 푇− . As 푒 ∈ 퐶, the edge 푒 has its other end in 푋; hence the perfect matching 푀0 , which contains 푒, must match each vertex in 푇− , other than the end of 푒, with a vertex in 푇+ . This implies that 푀0 has precisely three edges in the cut 푈; one of these edges is 푒. Thus, equality holds throughout in equation (15.2); hence |퐹 − 푒| = |(푀0 ∩ 푈) − 푒| = 2. Moreover, from equation (15.1), we deduce that 퐹 − 푒 = (푀0 ∩ 푈) − 푒. As |퐹 − 푒| = 2, it follows that | 푋 ∩ 푌 | = 2, by Lemma 15.10. The two edges of 퐹 − 푒 are in 푀0 and are incident with the two vertices of 푋 ∩ 푌 ; hence 푒 ∉ 퐹. In sum, 푒 is incident with a vertex in 푇− and is in 퐶 − 퐹. Thus, 푒 has an end in 푋 ∩ 푌 . ƒ Lemma 15.16 The graph 퐺 [푋] is a pentagon. Proof We already know that 퐺 1 is an odd wheel and hence that 퐺 [푋] is an odd polygon. Thus, to show that 퐺 [푋] is a pentagon, it suffices to show that | 푋 ∩ 푌 | = 3 and | 푋 ∩ 푌 | = 2. By Lemma 15.15, we already know that | 푋 ∩ 푌 | = 2. Let us now prove that | 푋 ∩ 푌 | = 3. Suppose that | 푋 ∩ 푌 | = 1. Then, | 푋 | = | 푋 ∩ 푌 | + | 푋 ∩ 푌 | = 3 and both the edges of the matching 퐹 are 푏-invariant in 퐺 2 and are not multiple edges. In this case, two edges of 퐺 1 are 푏-invariant in 퐺 by Lemma 15.5. Therefore | 푋 ∩ 푌 | ≥ 3 and each minority vertex of the bipartite graph 퐺 [푋 ∩ 푌 ] has to be incident with 푒. As we have already noted, this is possible only if there is just one minority vertex; equivalently, only if | 푋 ∩ 푌 | = 3. We conclude that | 푋 | = 5 and hence that 퐺 [푋] is a pentagon. ƒ Lemma 15.17 The cut 퐷 is robust and for every perfect matching 푀 of 퐺, |푀 ∩ 퐶|, |푀 ∩ 퐷| ∈ {1, 5}. Proof The two edges of 퐹 and the edge 푒 are edges of 푀0 ∩ 퐶. As 푀0 is a perfect matching, it follows that the two vertices of 푆 + are also incident with edges in 푀0 ∩퐶. Thus, |푀0 ∩ 퐶| = 5. By definition of 푀0 , the characteristic of 퐶 is |푀0 ∩ 퐶|. Thus, 퐶 has characteristic five. As | 푋 | = 5, |푀 ∩ 퐶| ∈ {1, 5}, for every perfect matching 푀 of 퐺. Neither 푆 nor 푇 is trivial. Thus, for each perfect matching 푀 of 퐺, |푀 ∩ 퐼 | = 1 + 2|푀 ∩ {푒}| and |푀 ∩ 푈| = 1 + 2|푀 ∩ {푒}|.

15.2 Proof of Theorem 15.1

317

Moreover, no edge of 퐺 joins a vertex in 푋 ∩ 푌 to a vertex in 푋 ∩ 푌 . Thus, |푀 ∩ 퐶| + |푀 ∩ 퐷| = |푀 ∩ 퐼 | + |푀 ∩ 푈| = 2 + 4|푀 ∩ {푒}|. As |푀 ∩ 퐶| ∈ {1, 5}, it follows that |푀 ∩ 퐷| ∈ {1, 5}. This conclusion holds for every perfect matching of 퐺. In particular, 푀0 contains the edge 푒 and has just one edge in 퐷. Moreover, 퐷 − 푒 is tight in 퐺 − 푒. By Corollary 14.9, 퐷 is robust. ƒ Lemma 15.18 The graph 퐺 [푌 ] is also a pentagon. Proof We have already seen that the cuts 퐼 and 푈 are nontrivial and 푒 has one end in the minority part of 푆 and the other end in the minority part of 푇. Thus, only one minority vertex of 푇 is incident with 푒. If possible, suppose that the shore 푇 of 푈 has five or more vertices. In this case, the minority part of 푇 has two or more vertices; hence the minority part of 푇 contains a vertex, 푣, not incident with edge 푒. Vertex 푣 has degree at least three in 퐺 2 − 푒. By Lemma 8.19, 휕 (푣) contains two or more edges that are removable in the bipartite graph (퐺 − 푒)/( 푋 ∪ 푌 ). At least one of these edges, say, 푓 , does not lie in the perfect matching 푀1 . Edge 푓 has both ends in 푇; hence 퐺 2 − 푒 − 푓 is a near-brick. As 푓 ∉ 푀1 , it follows from Lemma 15.7 that 퐺 2 − 푓 is a near-brick. As 푓 has both ends in 푋, it is 푏-invariant in 퐺 1 . In sum, both 퐶-contractions of 퐺 − 푓 are near-bricks. If 퐶 is not tight in 퐺 − 푓 then 푓 is 푏-invariant in 퐺, a contradiction. We may thus assume that 퐶 is tight in 퐺 − 푓 ; hence 푓 is quasi 푏-invariant in 퐺 and 퐶 is tight in 퐺 − 푓 . Then, 퐺 has a cut 퐷 ′ such that (퐶, 퐷 ′ − 푓 ) is either a barrier cut pair of 퐺 − 푓 , or (퐶, 퐷 ′ − 푓 ) is an essentially 2-separation cut pair of 퐺 − 푓 . If (퐶, 퐷 − 푓 ) is a barrier cut pair then 퐶 has characteristic three, by Lemma 14.11; this is a contradiction. So, suppose that (퐶, 퐷 − 푓 ) is associated with an essentially 2-separation pair {푆 ′ , 푇 ′ }. Adjust notation so that 푆 ′ ⊂ 푋 and 푇 ′ ⊂ 푋. The edge 푓 has both ends in 푋; hence 푆 ′ is a singleton. By Lemma 14.12, 퐶 has characteristic three, a contradiction. We deduce that 푇 consists of precisely three vertices. The edge 푒 is not in 퐹; hence 퐺 [푌 ] is a pentagon. ƒ

15.2.11 The conclusion: 푮 is the Petersen graph Making use of the various reductions described in the previous sections, we are now in a position to wrap up the proof of Lov´asz’s Conjecture by showing that the only brick that has no 푏-invariant edges and is different from 퐾4 and 퐶6 is the Petersen graph. By Lemmas 15.16 and 15.18, the subgraphs 퐺 [푋] and 퐺 [푌 ] of 퐺 induced by 푋 and 푌 are pentagons. Let 푣 0 푣 1 푣 2 푣 3 푣 4 푣 0 denote the pentagon 퐺 [푋], where 푣 0 is one of the ends of 푒, the two vertices 푣 1 and 푣 4 are the two majority vertices of 푋 ∩푌 , and 푣 2 and 푣 3 are the two vertices in 푋 ∩ 푌 . Also, let 푣 ′0 푣 ′1 푣 3 푣 2 푣 ′4 푣 ′0 denote the pentagon 퐺 [푌 ], where 푣 ′0 is the minority vertex of 푋 ∩ 푌 and 푣 ′1 and 푣 ′4 are its two majority vertices. See Figure 15.8.

15 푏-Invariant Edges in Bricks

318

The two edges 푣 3 푣 ′1 and 푣 2 푣 ′4 are the edges in 퐹. The perfect matching 푀0 contains the edge 푒 = 푣 0 푣 ′0 , the two edges of 퐹, and it matches 푣 1 and 푣 4 with two distinct vertices in 푋 ∩ 푌 . See Figure 15.8(a). Likewise, the perfect matching 푀1 contains the edges 푒, 푣 1 푣 2 , 푣 3 푣 4 and it matches 푣 ′1 and 푣 ′4 with two distinct vertices in 푋 ∩ 푌 . See Figure 15.8(b). 푌

푌

푣1 푋

푣1

푣2

푣0

푋 푣3

푣4

푣2

푣0 푣3

푣4 퐶

푣1′

푒

푣4′

퐶 푣1′

푒

푣0′ 퐷 (a) 푀0

푣4′ 푣0′

퐷 (b) 푀1

Fig. 15.8 The perfect matchings 푀0 and 푀1 .

At this stage we can infer that 퐺 has at least ten vertices, and that 푛 − 8 of them are in the quadrant 푋 ∩ 푌 . We shall now proceed to show that this quadrant has precisely two vertices and that both 퐺 [푋] and 퐺 [푌 ] are also pentagons by using the fact that the two edges 푣 3 푣 ′1 and 푣 2 푣 ′4 in 퐹 are 푏-invariant in both 퐺 1 and 퐺 2 but not 푏-invariant in 퐺 (Exercise 15.2.5). Making use of this fact to deduce that 퐺 is indeed the Petersen graph involves a subtle use of symmetry as explained below. Final step: cyclic rotations Let 푣 ′2 and 푣 ′3 denote two vertices of 푋 ∩ 푌 that are respectively adjacent to vertex ′′ 푣 1 and vertex 푣 4 . Likewise, let 푣 ′′ 2 and 푣 3 denote two vertices of 푋 ∩ 푌 that are ′ respectively adjacent to vertex 푣 1 and vertex 푣 ′4 . See Figure 15.9(a). We first observe that the four vertices 푣 0 , 푣 ′0 , 푣 2 and 푣 3 have degree three. As noted above, edge 푒 1 := 푣 3 푣 ′1 is also 푏-invariant in 퐺 2 and in 퐺 1 , but not in 퐺. Hence 퐺 has a robust cut 퐷 1 := 휕 (푌1 ) such that the pair {퐶 − 푒 1 , 퐷 1 − 푒 1 } is associated with an essential 2-separation {푆1 , 푇1 } of 퐺 − 푒 1 . Moreover, neither 푆1 nor 푇1 is trivial, and the graph 퐺 [푌1 ] is a pentagon. The minority vertex of 푋 ∩ 푌1 is the end 푣 3 of 푒 1 in 푋. Thus, the two majority vertices of 푋 ∩푌1 are the vertices 푣 2 and 푣 4 . Consequently, the two vertices of 푋 ∩푌1 are 푣 0 and 푣 1 . This corresponds to the rotation of the pentagon 퐺 [푋] in which every vertex 푣 푖 is now mapped to the vertex 푣 푖+3 , where the indices are taken modulo 5. See Figure 15.9(b).

15.2 Proof of Theorem 15.1

319 푌1

푌 푣1 푋

푣4

푣2 푋

푣0

푣1

푣3 푣2′′

퐶 푣1′

푣4′

푣3′ 푒

푣3 푣2

푣4

푣2′

푣0

푣3′′

푣3′′

푣3′ 푒1

퐶 푣2′

푣0′

푣4′ 푣1′

푣0′

퐷 (a)

퐷1 (b)

Fig. 15.9 Solid dots represent vertices known to have degree three. (a) The vertices 푣2′ , 푣3′ , 푣2′′ and 푣3′′ . (b) the vertices 푣2′ and 푣2′′ coincide.

We have seen that the vertices 푣 0 , 푣 ′0 , 푣 2 and 푣 3 have degree three. There is a correspondence between the edges of the pentagons 퐺 [푌 ] and 퐺 [푌1 ]. The vertices 푣 ′1 and 푣 1 correspond, with edge 푒 1 playing the role of 푒, respectively to the vertices 푣 ′0 and 푣 3 in the case of edge 푒. Thus, the vertices 푣 1 and 푣 ′1 also have degree three in 퐺. Moreover, the vertex 푣 ′2 , which is adjacent to the vertex 푣 1 , coincides with ′ ′ the vertex 푣 ′′ 2 , which is adjacent to 푣 1 . Thus, the vertices 푣 1 and 푣 1 are adjacent to precisely one and the same vertex 푣 ′2 of 푋 ∩ 푌 . A similar reasoning may be applied with 푒 2 := 푣 2 푣 ′4 playing the role of 푒, thereby concluding that both vertices 푣 4 and 푣 ′4 have degree three and have a common neighbour 푣 ′3 in 푋 ∩ 푌 . In sum, all the eight vertices of the graph 퐺 [푋 ∪ 푌] have degree three. Moreover, the ends of the edges of 휕 ( 푋 ∪ 푌) in 푋 ∩ 푌 are in the set 푅 := {푣 ′2 , 푣 ′3 }. See Figure 15.10(a). As 퐺 is a brick, it is 3-connected; thus, 푋 ∩푌 = 푅. As | 푋 ∩푌 | is even, the vertices 푣 ′2 and 푣 ′3 are distinct. Moreover, they must have degree three or more; hence they have degree three and are adjacent. We conclude that 퐺 is the Petersen graph. See Figure 15.10(b). This completes the proof of the validity of Lov´asz’s Conjecture.

ƒ

A removable class 푅 of a matching covered graph 퐺 is optimal if either 푅 is a removable doubleton or 푅 is a (푏 + 푝)-invariant singleton. The following corollary is an easy consequence of Theorems 15.1 and 13.6. Optimal removable classes Corollary 15.19 Every brick has an optimal removable class.

ƒ

15 푏-Invariant Edges in Bricks

320 푌

푌

푣1 푋

푣1

푣2

푣0

푋

푣0

푣3

푣4

푣2

푣3

푣4 퐶

푣3′

푣2′

푣1′

퐶

푣4′

푒

푣3′

푣2′

푣1′

푣4′

푒 푣0′ 퐷 (a)

푣0′ 퐷 (b)

Fig. 15.10 (a) The vertices 푣2′ and 푣3′ , where solid dots represent vertices known to have degree three; (b) the Petersen graph.

Exercises 15.2.1 Show that the bipartite graph 퐻 defined in the proof of statement 15.5.1 is a brace. 15.2.2 In the proof of statement 15.5.1 explain why there is an edge 푒 incident with 푣 that is 푏-invariant in 퐺. 15.2.3 Show that the nine edges indicated by solid lines in Figure 15.4 are the 푏-invariant edges in 퐺. 15.2.4 Prove Lemma 15.11. 15.2.5 Show that the two edges 푣 3 푣 ′1 and 푣 2 푣 ′4 are 푏-invariant in both 퐺 1 and 퐺 2 (Hint: Use Lemma 15.12). Lemma 15.20 (Inheritance of removable classes) Let 퐺 be a brick, let 푒 be a 푏-invariant edge of 퐺 and let 퐻 be a brick obtained by a tight cut decomposition of 퐺 − 푒. Every removable class of 퐻 is a removable class of 퐺 − 푒. Moreover, if 퐺 is regular then every removable doubleton of 퐻 is a removable doubleton of 퐺. ∗ 15.2.6 Prove Lemma 15.20. An edge of a matching covered graph 퐺 is solitary if it lies in precisely one perfect matching of 퐺. ⊲15.2.7 Prove that a brick 퐺 ∉ {퐾4 , 퐶6 , P} is extremal if and only if (i) every 푏-invariant edge of 퐺 is solitary and (ii) for any 푏-invariant edge of 퐺, the graph 퐺 − 푒 is extremal.

15.3 Notes

321

15.3 Notes As we have already mentioned, the first proof of Lov´asz’s conjecture appeared in Carvalho’s Ph.D. thesis [6]. As we shall see in the next chapter, using Theorem 15.1, it is fairly easy to show that every near-brick with 푝 = 0 has an optimal ear decomposition (as defined in Section 11.4). To show that every matching covered graph has an optimal ear decomposition we needed to show that every brick 퐺 different from 퐾4 , 퐶6 , the bicorn and the Petersen graph has two 푏-invariant edges. The proof of this stronger result appeared in [10]. We have chosen not to include that proof here. In Chapters 17 and 18 we shall present ramifications of Theorem 15.1 which lead to generation procedures for bricks. We shall also describe there analogous generation procedures for braces.

Chapter 16

The Matching Lattice and Optimal Ear Decompositions

Contents 16.1

16.2

16.3 16.4 16.5

The Matching Lattice and its Bases . . . . . . . . . . . . . . . . . . . . . . . . . . . 323 16.1.1 Lattices, generating sets, and bases . . . . . . . . . . . . . . . . . . . . 324 16.1.2 Reduction to bricks and braces . . . . . . . . . . . . . . . . . . . . . . . 324 16.1.3 The question of existence of bases . . . . . . . . . . . . . . . . . . . . 325 16.1.4 The matching lattice of the Petersen graph . . . . . . . . . . . . . . 327 Optimal Ear Decompositions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 332 16.2.1 The case of bipartite graphs and near-bricks . . . . . . . . . . . . 333 16.2.2 The difficulty in dealing with graphs which have more than one brick . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 334 16.2.3 A strengthening of Theorem 15.1 . . . . . . . . . . . . . . . . . . . . . 335 16.2.4 Exchange property of optimal removable classes . . . . . . . . 335 Matching Integral Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 337 Matching Spaces over Z2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 341 16.4.1 Matching orthogonal vectors . . . . . . . . . . . . . . . . . . . . . . . . . 342 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 345

16.1 The Matching Lattice and its Bases The objective of this chapter is to present a characterization of the matching lattice of a matching covered graph. Our approach to this characterization, envisaged by Lov´asz as explained in Section 13.1, makes essential use of Theorem 15.1. This approach also enables us to answer two related questions; one concerning bases of the matching lattices, and the other concerning optimal ear decompositions of matching covered graphs. In the last section we give a description of the ‘dual’ approach adopted by Lov´asz [58].

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7_16

323

324

16 The Matching Lattice and Optimal Ear Decompositions

16.1.1 Lattices, generating sets, and bases For the convenience of the reader, we start by recalling some basic definitions and results from the theory of lattices in Euclidean spaces. These are not necessary for understanding the material of this chapter, but we are including them here to give the readers an appreciation of the special nature of the matching lattice of a matching covered graph. A subset L of R푚 is a lattice if there exists a finite set S of rational vectors in R푚 , called a generating set of L, such that L consists of all integer linear combinations of vectors in S. (The matching lattice Lat(퐺) of a matching covered graph 퐺 is the lattice generated by { 휒 푀 : 푀 ∈ M}.)

The theory of lattices is a part of geometric theory of numbers, and there is an extensive literature which deals with it. In recent years, combinatorial optimizers have been interested in lattices because of their connection with integer programming. The book by Schrijver [83] has a good introduction to aspects of lattices that are relevant for the material in this section. Clearly, the same lattice L in R푚 may have more than one generating set. But it is not obvious that every lattice must have generating sets that are linearly independent. The following fundamental theorem (Corollary 4.1b in [83]) asserts this to be true: Theorem 16.1 Every lattice has linearly independent generating sets, called its bases. ƒ In the above respect, lattices are analogous to subspaces. However, there is one essential difference. Any generating set of a subspace contains a subset that is a basis of that subspace, but a generating set of a lattice may not contain a basis of that lattice. For example, {2, 3} generates the lattice Z in R1 , but it does not contain a subset that is a basis of Z.

16.1.2 Reduction to bricks and braces Now we return to the specific question of characterizing matching lattices of matching covered graphs. It would be helpful to the reader to review Section 6.2 and 6.3 before reading this section. Recall that a vector x in the edge space R퐸 of a matching covered graph 퐺 is 푟-regular if x(퐶) = 푟 for any tight cut 퐶 of 퐺. We showed in Chapter 6 that a vector x belongs to the matching space Lin(퐺) if and only if it is regular.

As the matching lattice Lat(퐺) of a matching covered graph 퐺 is the set of integer linear combinations of incidence vectors of perfect matchings of 퐺, it clearly is a subset of the matching space Lin(퐺) which consists of all (real) linear combinations of the same set of vectors. Thus, Lat(퐺) ⊆ Lin(퐺) ∩ Z퐸 .

(16.1)

16.1 The Matching Lattice and its Bases

325

Furthermore, as every vector in Lin(퐺) is regular, the above inequality implies that every vector in Lat(퐺) is a regular integer vector. However, as has been noted before, the converse is not always true; there are graphs, such as the Petersen graph, whose matching lattices do not include all the regular integer vectors. Given any regular integer vector x in Z퐸 , how do we decide whether or not it belongs to Lat(퐺)? The following observation is the first step in trying to find an answer to this question; we leave its proof, which is a straightforward adaptation of the proof of Theorem 6.11, as Exercise 16.1.2. Lemma 16.2 Let 퐶 be a tight cut of a matching covered graph 퐺, and let 퐺 1 and 퐺 2 be the two 퐶-contractions of 퐺. A vector x ∈ Z퐸 belongs to Lat(퐺) if and only if the restriction of x to 퐸 (퐺 푖 ) belongs to Lat(퐺 푖 ), for 푖 = 1, 2. ƒ Consequently, given any tight cut decomposition G of 퐺, the problem of deciding whether or not a given vector x is in the matching lattice of 퐺 reduces to deciding whether or not the restrictions of x to the edge sets of the bricks and braces in G are in their respective matching lattices. Our solution to the problem of characterizing Lat(퐺) goes hand-in-hand with an answer to a related question concerning the existence of bases of matching lattices.

16.1.3 The question of existence of bases Although Lat(퐺) is a proper subset of Lin(퐺), the dimensions of these two sets are clearly the same, which, as we saw in Section 6.3, is equal to 푚 − 푛 + 2 − 푏. Obviously, Lin(퐺) has bases which are subsets of { 휒 푀 : 푀 ∈ M}. However, a priori, it is not clear that Lat(퐺) has a basis consisting solely of incidence vectors of perfect matchings. As a counterpart to Conjecture 13.2, we were led to the premise that Lat(퐺) does indeed have such bases. (As observed in the previous subsection, in general, a generating set of a lattice need not contain a basis of the lattice! See Exercise 16.1.1 for an instructive example.) As shown in Section 6.3, a basis for the matching space of a matching covered graph 퐺 can be obtained from bases of the matching spaces of its bricks and braces by using the merger operation. Finding a basis of the matching space of a brace is easy enough. In fact, more generally, using the fact that every bipartite matching covered graph distinct from 퐾2 has a removable ear, one may easily deduce the following result (see Exercise 6.3.3): Lemma 16.3 In any bipartite matching covered graph 퐺 one can find a list B of 푚 − 푛 + 2 perfect matchings in stepwise canonical order such that any regular integer vector in Z퐸 may be expressed as an integer linear combination of vectors ƒ in { 휒 푀 : 푀 ∈ B}. We shall now prove Theorem 13.1 which says that if 퐺 is any matching covered graph none of whose bricks is a Petersen brick, then the equality Lat(퐺) = Lin(퐺) ∩ Z퐸 holds.

326

16 The Matching Lattice and Optimal Ear Decompositions

Matching lattice of graphs with 푝 = 0 Theorem 16.4 Let 퐺 be a matching covered graph none of whose bricks is a Petersen brick. Then, 퐺 has a basis for the matching lattice consisting of 푚 − 푛 + 2 − 푏 perfect matchings. Moreover, a vector x in the edge space Z퐸 of 퐺 belongs to Lat(퐺) if and only if it is regular. Proof Clearly, every x in Lat(퐺) is regular. Our task is thus to show that every regular integer vector x belongs to Lat(퐺), and also to establish a basis for Lat(퐺) with the asserted properties. We do this by induction on |퐸 |. Case 1 Graph 퐺 has a nontrivial tight cut 퐶. Let 퐺 1 and 퐺 2 denote the two 퐶-contractions of 퐺. As 퐺 is free of Petersen bricks so are both 퐺 1 and 퐺 2 . Furthermore, for 푖 = 1, 2, the trivial tight cut of 퐺 푖 associated with the contraction vertex is the same as 퐶, whereas all the other tight cuts of 퐺 푖 are also tight cuts of 퐺. It follows that the restriction of x to 퐸 (퐺 푖 ) is regular. By the induction hypothesis, it follows that the restrictions of x to 퐸 (퐺 1 ) and 퐸 (퐺 2 ) are in Lat(퐺 1 ) and Lat(퐺 2 ), respectively. Using Lemma 16.2 we now deduce that x is in Lat(퐺). Turning to the question of the existence of a required type of basis of Lat(퐺), we first note that for 푖 = 1, 2, by induction, there exists a basis B푖 of Lat(퐺 푖 ) consisting of 푚 푖 − 푛푖 + 2 − 푏 푖 perfect matchings of 퐺 푖 . It is now a straightforward matter to verify that B1 ∨ B2 consists of 푚 − 푛 + 2 − 푏 perfect matchings of 퐺 and that it is a basis of Lat(퐺). See Exercise 16.1.4. Case 2 Graph 퐺 is a brace. In this case, the assertion follows from Lemma 16.3. Case 3 Graph 퐺 is a brick which is not a Petersen brick. If 퐺 is either 퐾4 or 퐶6 , it is easy to see that every regular vector in Z퐸 is in Lat(퐺) and that { 휒 푀 : 푀 ∈ M} is a basis of Lat(퐺). We may therefore assume that 퐺 is not one of the three special bricks 퐾4 , 퐶6 , and P. By Theorem 15.1 퐺 has a 푏-invariant edge, and hence, by Theorem 13.6, has a (푏 + 푝)-invariant edge, say 푒. Thus, the graph 퐺 ′ = 퐺 − 푒 is a near-brick whose brick is not a Petersen brick. Let 푀푒 be any perfect matching of 퐺 containing the edge 푒, and let x′ denote the restriction of x − x(푒) 휒 푀푒 to 퐸 − 푒. 16.4.1 x′ is a regular integer vector in the edge space of 퐺 ′ .

Proof Certainly x′ is an integer vector in the edge space of 퐺 ′ . We must thus prove that x′ is regular. For this, let 푟 be the common value of x(휕 (푣)) for all 푣 ∈ 푉 and let 푟 ′ := 푟 − x(푒). Clearly, for each 푣 ∈ 푉, x′ (휕 (푣)) = 푟 − x(푒) = 푟 ′ . Let 퐶 be any nontrivial tight cut of 퐺 ′ . The graph 퐺 ′ is a near-brick; hence one 퐶contraction of 퐺 ′ , say, 퐻, is bipartite. Let ( 퐴, 퐵) denote the bipartition of 퐻, let 푦 denote the contraction vertex of 퐻, and adjust notation so that 푦 ∈ 퐴. Then, x′ (퐶) = |퐵|푟 ′ − (| 퐴| − 1)푟 ′ = 푟 ′ . We conclude that x′ (퐶) = 푟 ′ , for each tight cut of 퐺 ′ ; hence x′ is regular. ƒ

16.1 The Matching Lattice and its Bases

327

By induction, x′ is in Lat(퐺 ′ ) and Lat(퐺 ′ ) has a basis B ′ consisting of incidence vectors of perfect matchings of 퐺 ′ . It is now easy to see that x is in Lat(퐺) and that B ′ ∪ { 휒 푀푒 } is a basis for Lat(퐺). ƒ Lov´asz’s own proof of Theorem 13.1, given in [58], is via a characterization of the ‘dual’ of the matching lattice. He established a clear correspondence between the two points of view in that paper. It relies on some basic facts from the theory of lattices. We shall give a brief description of these ideas in Section 16.3. To complete our characterization of the matching lattices of all matching covered graphs, we need to present a description of matching lattices of Petersen bricks. Before we do that, we record here an attractive consequence of Theorem 16.4. Theorem 16.5 The edge set of any cubic brick 퐺 different from P may be written as the symmetric difference of a collection of perfect matchings of 퐺. Proof Consider the incidence vector 1 of the edge set 퐸 of 퐺. This vector is clearly a 3-regular vector. By Theorem 16.4 it follows that 1 is in Lat(퐺). Thus, there exist integers 훼 푀 : 푀 ∈ M such that Õ 1= 훼 푀 휒 푀 . 푀 ∈ M

Consider any such linear combination of 1, and let Modd and Meven denote, respectively, the subsets of M such that 훼 푀 is odd for each 푀 ∈ Modd and 훼 푀 is even for each 푀 ∈ Meven . We thus have: Õ Õ 훼 푀 휒 푀 , 훼 푀 휒 푀 + 1= 푀 ∈ Modd

implying that: 1≡

Õ

푀 ∈ Modd

푀 ∈ Meven

훼 푀 휒 푀

(mod 2).

The above congruence implies that 퐸 is the symmetric difference of the perfect ƒ matchings in the set 푀 ∈ Modd . It is not difficult to see that the edge set of P cannot be written as the symmetric difference of a set of perfect matchings of P (Exercise 16.1.10).

16.1.4 The matching lattice of the Petersen graph Not every regular integer vector x in the edge space of the Petersen graph P is in its matching lattice Lat(P). However, a simple condition on x, in addition to its regularity, ensures that it is in Lat(P). We start by noting that the Petersen graph has twelve pentagons. For any pentagon 푄 of P, we shall refer to 푁 := 휕 (푉 (푄)) as a central cut of P. There are six central cuts, and each of these cuts is also a perfect matching of P. Indeed, a set of edges

16 The Matching Lattice and Optimal Ear Decompositions

328

of P is a central cut if and only if it is a perfect matching. Moreover, (i) every edge is in precisely two of the six perfect matchings of P, and (ii) the intersection of any two distinct perfect matchings consists of precisely one edge. Suppose that x is an 푟-regular integer vector that belongs to Lat(P). Then, by definition, corresponding to each perfect matching 푀 there exists an integer 훼 푀 such that Õ 훼 푀 휒 푀 x= Í By equation (6.8), 푟 = 훼 푀 .

Given any fixed perfect matching 푁 of P, every edge of 푁 is in precisely one perfect matching of P that is different from 푁. It follows that Õ  훼 푀 + 5훼 푁 = 푟 + 4훼 푁 ≡ 푟 (mod 4). x(푁) = 푀 ∈ M− 푁

Not all regular integer vectors in the edge space of P satisfy this condition. For example, the 3-regular vector of all 1’s does not, because, on any central cut it takes the value 5 and 5 . 3 (mod 4). But the 7-regular vector x displayed in Figure 16.1(a) does (x(푁) = −1 ≡ 7 (mod 4)). (One of the curious consequences of the proof of Lemma 16.6 is that if x(푁) ≡ 푟 (mod 4) for any one particular perfect matching 푁, then x(푀) ≡ 푟 (mod 4) for any perfect matching 푀 of P. See Exercise 16.1.5.)

5

4

−1

5

3

−1 4

3

4

1 0

5

3

4

푁

푁 0

3 (a) x

2 2

3 4

1

4 5

2

3

3

1

3

2 3 (b) y

Fig. 16.1 (a) A 7-regular vector x in Z퐸 ; (b) the 9-regular vector y = x + 2휒 푁

Lemma 16.6 An 푟-regular integer vector x in the edge space of P, the Petersen graph, is in the matching lattice Lat(P) if and only if x(푁) ≡ 푟 (mod 4), for each perfect matching 푁 of P. Furthermore, the set of the incidence vectors of the six perfect matchings of P is a basis for Lat(P).

16.1 The Matching Lattice and its Bases

329

Proof The only if part follows from the observations made above. We now proceed to prove the converse. Let x be an 푟-regular vector on Z퐸 , and let 푁 be any fixed perfect matching (central cut) of P. Assume that x(푁) ≡ 푟 (mod 4). This assumption implies that 푑 := (x(푁) − 푟)/4 is an integer. Now consider the vector y := x − 푑 · 휒 푁 . Clearly y is an (푟 − 푑)-regular integer vector. Moreover, y(푁) = x(푁) − 5푑 = (4푑 + 푟) − 5푑 = 푟 − 푑. Let P1 and P2 denote the two 푁-contractions of P. The above property signifies that, not only is y an (푟 − 푑)-regular vector but that, for 푖 = 1, 2, the restriction y푖 of y to 퐸 (P푖 ) is also (푟 − 푑)-regular in P푖 . Both P1 and P2 are 5-wheels. By Theorem 16.4, y푖 ∈ Lat(P푖 ). Furthermore, for 푖 = 1, 2, each perfect matching of P푖 is the restriction of a perfect matching of P that has just one edge in common with 푁. It follows that y is an integer linear combination of the incidence vectors of the five perfect matchings of P that have just one edge in common with 푁. Since x = y + 푑휒 푁 , it follows that x is in Lat(P), and that { 휒 푀 : 푀 ∈ M} is a basis of Lat(P). ƒ In the example shown in Figure 16.1(a), the vector x is a 7-regular vector and x(푁) = −1. Thus, in this case 푑 = (−1 − 7)/4 = −2. The vector y = x + 2휒 푁 , which is displayed in Figure 16.1(b), is a 9-regular vector, with the additional property that y(푁) is also equal to 9. A Petersen brick 퐺 (a graph whose underlying simple graph is the Petersen graph P) also has twelve central cuts that correspond to central cuts of P. If 퐺 has multiple edges, then not all central cuts are perfect matchings of 퐺. However, each central cut 푄 is a union of perfect matchings of 퐺 that meet it in five edges, and all other perfect matchings of 퐺 have precisely one edge in common with the cut 푄. If 퐺 is simple then it is the Petersen graph. If it is not simple, it has multiple edges and each multiple edge is clearly 푏-invariant. The following theorem may thus be proved by induction on the number of edges using the statement of Lemma 16.6 as the basis for induction (Exercise 16.1.6). The matching lattice of a Petersen brick Theorem 16.7 Let 퐺 be a Petersen brick. Then an 푟-regular integer vector x in the edge space of 퐺 is in Lat(퐺) if and only if x(푄) ≡ 푟 (mod 4), for each central cut 푄 of 퐺. Furthermore, Lat(퐺) has a basis consisting of incidence vectors of 푚 − 푛 + 1 perfect matchings of 퐺. ƒ

16 The Matching Lattice and Optimal Ear Decompositions

330

As a corollary of Theorems 16.4 and 16.7 we have the following result which answers the question concerning the bases of matching lattices: Bases of Lat(퐺) Corollary 16.8 Every matching covered graph 퐺 has a basis for the matching lattice consisting of precisely 푚 − 푛 + 2 − 푏 perfect matchings. ƒ It should be noted that Lex Schrijver has a proof of Theorem 16.4 in Chapter 38 of his treatise [84]. This proof is different from the one due to Lov´asz and relies on the study of properties of special barriers in a matching covered graph.

Exercises 16.1.1 Consider the pentagonal prism 퐺 := P10 with its edges labelled as in Figure 16.2.

푎4

푒1

푎5

푒5

푒2 푏4 푏3

푎3

푏5 푋

푏1 푎1

푏2

푒4

푒3 푎2

Fig. 16.2 The pentagonal prism 퐺 := P10 .

Let B denote the set of the incidence vectors of the five perfect matchings listed below: 푀1 푀2 푀3 푀4 푀5

:= := := := :=

{푒 1 , 푎 1 , 푎 3 , 푏 1 , 푏 3 } {푒 2 , 푎 2 , 푎 4 , 푏 2 , 푏 4 } {푒 3 , 푎 3 , 푎 5 , 푏 3 , 푏 5 } {푒 4 , 푎 1 , 푎 4 , 푏 1 , 푏 4 } {푒 5 , 푎 2 , 푎 5 , 푏 2 , 푏 5 }

16.1 The Matching Lattice and its Bases

331

(i) Show that B is a basis of the sub-lattice of Lat(퐺) generated by the incidence vectors of perfect matchings that meet the robust cut 휕 ( 푋) in precisely one edge. (ii) Let 푀6 denote the perfect matching {푒 1 , 푒 2 , 푒 3 , 푒 4 , 푒 5 }. Show that B ∪ { 휒 푀6 } is a basis of Lin(퐺). (iii) Show that the vector 1 cannot be expressed as an integer linear combination of the the six vectors in B ∪ { 휒 푀6 }. (iv) Finally, conclude that B ∪ { 휒 푀6 } is a basis of Lin(퐺) but that it is not a basis of Lat(퐺). (v) Find a basis of Lat(퐺) that includes B. 16.1.2 Give a proof of Lemma 16.2. ⊲16.1.3 Figure 16.3 depicts a brick 퐺, and an integral vector x, where the unlabeled edges have weight one. Show that x lies in Lat(퐺), following the proof of Theorem 16.4. 0

0 0

0

0

0

Fig. 16.3 The brick for Exercise 16.1.3

⊲16.1.4 Adapt the proof of Theorem 6.11 to show that if 퐶 is a tight cut of 퐺, and B1 and B2 are bases of Lat(퐺 1 ) and Lat(퐺 2 ), respectively, then B is a basis of Lat(퐺). ⊲16.1.5 Let x be an integral vector in the matching space of P. Prove that if x(푁) ≡ 푟 (mod 4) for any one particular perfect matching 푁 of P, then x(푀) ≡ 푟 (mod 4) for any perfect matching 푀 of P. Hint: this statement is a consequence of the proof of Lemma 16.6. 16.1.6 Give a proof of Theorem 16.7. ⊲16.1.7 Prove Corollary 16.8. ⊲16.1.8 Give a polynomial-time algorithm which, given a matching covered graph 퐺 and a vector x in Z퐸 , determines whether x lies in Lat(퐺).

332

16 The Matching Lattice and Optimal Ear Decompositions

16.1.9 Consider the cubic brick 퐺 = P ⊙ 퐾4 , depicted in Figure 16.3.

(i) Find an expression of 휒 퐸 = 1 as an integer linear combination of the incidence vectors of perfect matchings of 퐺. (ii) Using the above linear combination, find a set of perfect matchings of 퐺 whose symmetric difference is the edge set 퐸.

16.1.10 Show that 퐸 (P) cannot be expressed as the symmetric difference of a collection of perfect matchings of P. 16.1.11 Let 퐺 be a cubic matching covered graph. Show that the vector 2 := (2, 2, . . . , 2) of all 2’s in the edge space of 퐺 belongs to Lat(퐺).

16.2 Optimal Ear Decompositions According to Theorem 11.18, the number of double ears in any ear decomposition of a matching covered graph 퐺 is at least 푏(퐺) + 푝(퐺). This bound is always attainable, as shown by CLM (2002, [11]). In other words, any matching covered graph 퐺 has an ear decomposition with precisely 푏 + 푝 double ears. We refer to such an ear decomposition as an optimal ear decomposition of 퐺. In seeking to find optimal ear decompositions of a matching covered graph 퐺, it is convenient to think in terms of top-down decompositions in which we start with 퐺 and sequentially delete removable ears to obtain 퐾2 . If 푅 is any removable double ear of 퐺, then (푏 + 푝) (퐺 − 푅) = (푏 + 푝) (퐺) − 1. Thus, deleting 푅 reduces the task of finding an optimal ear decomposition of 퐺 to one of finding such a decomposition of 퐺 − 푅. On the other hand, if 푅 is an arbitrary removable single ear, then (푏 + 푝) (퐺 − 푅) may be even larger than (푏 + 푝) (퐺) itself, and deleting 푅 would be a setback to the task of finding an optimal ear decomposition of 퐺. However, if 푅 were a (푏 + 푝)-invariant single ear, then (푏 + 푝) (퐺 − 푅) = (푏 + 푝) (퐺). In this case, deletion of 푅 reduces the task of finding an optimal ear decomposition of 퐺 to finding one of the graph 퐺 − 푅 with fewer edges than 퐺. This leads us to the notion of an optimal ear as defined in the following inset: Optimal removable ears and their existence A removable ear 푅 of a matching covered graph 퐺 is optimal if it is either a double ear or is a (푏 + 푝)-invariant single ear. (An optimal removable class of 퐺 is either a removable doubleton or is a (푏 + 푝)-invariant edge.). Our proof of the following theorem, which appeared in CLM [11], is along the lines of the hint given for establishing the weaker statement in Exercise 11.3.6. Theorem 16.9 Every matching covered graph 퐺 distinct from 퐾2 has an optimal removable ear. ƒ From Theorem 16.9 it is easy to prove the following result, by induction on the number of edges of 퐺 (Exercise 16.2.1).

16.2 Optimal Ear Decompositions

333

Corollary 16.10 Every matching covered graph 퐺 has an optimal ear decomposiƒ tion, which uses precisely (푏 + 푝)(G) double ears. As we shall see below, the validity of Theorem 16.9 for near-bricks is a consequence of Corollary 15.19.

16.2.1 The case of bipartite graphs and near-bricks Let us prove Theorem 16.9 for matching covered graphs 퐺 that are either bipartite or b of 퐺 is also a matching near-bricks. That is, 푏(퐺) ≤ 1. First note that the retract 퐺 b = (푏 + 푝) (퐺). Observe that, b = 푏(퐺) ≤ 1 and (푏 + 푝) ( 퐺) covered graph with 푏( 퐺) b by Theorem 11.21, there is a one-to-correspondence between removable ears of 퐺 b then the b is an optimal removable class of 퐺, and those of 퐺. Also note that if 푅 corresponding ear 푅 of 퐺 is an optimal removable ear of 퐺 (Exercise 16.2.2). In view of this, it is sufficient to prove the existence of optimal removable classes in bipartite graphs and near-bricks which are free of vertices of degree two. Furthermore, if 퐺 has multiple edges, then it clearly has a (푏 + 푝)-invariant edge. Thus, we may suppose that 퐺 is simple and that 훿(퐺) ≥ 3. (Figure 16.4 shows a near-brick and its retract.)

푋

푋

푢

푣

(푎)

(푏)

Fig. 16.4 (a) A near-brick; (b) its retract

If 퐺 has no nontrivial tight cuts, then 퐺 is a brace or a brick. If 퐺 is a brace, then as 퐺 is simple and 훿(퐺) ≥ 3, it follows that 퐺 has order six or more and every edge of 퐺 is removable, hence (푏 + 푝)-invariant. If 퐺 is a brick then the validity of Theorem 16.9 follows from Corollary 15.19. So, suppose that 퐺 has nontrivial tight cuts. Then, since 푏(퐺) ≤ 1, it follows that at least one of the shores of any tight cut is bipartite. Let 휕 ( 푋) be a nontrivial tight cut of 퐺 such that: (i) 퐺 1 := 퐺/( 푋 → 푥) is bipartite and, (ii) subject to (i), the shore 푋 is minimal. The minimality of 푋 implies that 퐺 1 is a brace, and the fact that 퐺 is a simple graph which is free of vertices of degree two implies that 퐺 1 has order at least six. The majority part 푋+ of 푋 is a special barrier of 퐺 and 퐺 [푋] is the only nontrivial component of 퐺 − 푋+ . Any edge

16 The Matching Lattice and Optimal Ear Decompositions

334

푒 incident with a vertex in 푋− is a removable edge in the brace 퐺 1 and is not an edge in the cut 휕 ( 푋). Consequently, 푒 is a removable edge of 퐺. Also, clearly, 휕 ( 푋) is a tight cut of 퐺 − 푒, implying that (푏 + 푝) (퐺 − 푒) = (푏 + 푝) (퐺) and hence that 푒 is in fact a (푏 + 푝)-invariant edge of 퐺. (In the near-brick shown in Figure 16.4(b), any edge incident with either 푢 or with 푣 is a (푏 + 푝)-invariant edge of that near-brick.) The validity of Theorem 16.9 for bipartite graphs and near-bricks implies the existence of optimal ear decompositions for matching covered graphs that are bipartite or near-bricks. In fact, if 푅 is an optimal removable ear of 퐺 and 푏(퐺) ≤ 1 then 푏(퐺 − 푅) ≤ 1. Hence, an inductive argument based on the number of edges of 퐺 establishes the validity of Corollary 16.10 for bipartite matching covered graphs and for near-bricks.

16.2.2 The difficulty in dealing with graphs which have more than one brick Suppose that 퐺 is a matching covered graph with two or more bricks. (See Figure 16.5 for an illustration.) Then, by the result established above, every brick of 퐺 has an optimal ear decomposition. It would be natural to somehow combine the optimal ear decompositions of the bricks of 퐺 to obtain an optimal ear decomposition of 퐺. 푢

푣

푒

푋

푓

푌

Fig. 16.5 A graph with two bricks 퐺/푋 and 퐺/푌 , and a brace

Unfortunately, this is not simple because a (푏 + 푝)-invariant single ear of a brick of 퐺 need not be a (푏 + 푝)-invariant single ear of 퐺. In fact, it may not even be removable in all the bricks and braces to whose edge sets it belongs. Consider for example a matching covered graph shown in Figure 16.5 with the indicated tight cut decomposition. The edge 푒 which belongs to 퐺/푋 as also to the unique brace of order four is not removable in 퐺 as 푣 is a cut vertex of 퐺 − 푒. Similarly, the edge 푓 which is shared by 퐺/푌 and the brace is not removable in 퐺. Thus, if 푒 were the only (푏 + 푝)-invariant edge of 퐺/푋, and 푓 were the only (푏 + 푝)-invariant edge of 퐺/푌 , we would be left with no obvious way of finding a (푏 + 푝)-invariant edge in 퐺!

16.2 Optimal Ear Decompositions

335

16.2.3 A strengthening of Theorem 15.1 To circumvent the difficulty pointed out above, it was necessary to prove a result which is stronger than Theorem 15.1. The following result was established by CLM (2002, [10]): Theorem 16.11 Every brick different from 퐾4 , 퐶6 , P, and the bicorn (Figure 15.3) has at least two (푏 + 푝)-invariant edges. ƒ The proof of this theorem is rather involved and we have chosen to not include that proof here with the hope that keen students might be led to pursue the subject further. The graphs 퐾4 and 퐶6 have three removable doubletons. In the Petersen graph, each edge is removable and (푏 + 푝)-invariant. The bicorn has two removable doubletons, in addition to a (푏 + 푝)-invariant edge. These observations, together with Theorem 16.11, can be used to show the following strengthening of Theorem 16.9 (Exercise 16.2.7). Theorem 16.12 Every matching covered graph distinct from 퐶2푛 (푛 ≥ 2) and 퐾2 has two optimal removable ears. ƒ

16.2.4 Exchange property of optimal removable classes Optimal removable ears in matching covered graphs satisfy an exchange property which will play a crucial role in the proofs of some important results in future chapters. The following lemma is a statement of that property. Lemma 16.13 Let 퐺 be a matching covered graph, let 푅 be an optimal removable class of 퐺 and let 푆 be an optimal removable class of 퐺 − 푅. If 퐺 has a perfect matching that contains the edges of 푅 but is disjoint with 푆 then 퐺 − 푆 is matching covered, 푆 is an optimal removable class of 퐺 and 푅 is an optimal removable class of 퐺 − 푆. Proof There are four cases to be considered depending whether 푅 is a singleton or a doubleton, and whether 푆 is a singleton or doubleton. Note that, in all these four cases, (i) (퐺 − 푅) − 푆 = (퐺 − 푆) − 푅 is matching covered because, by hypothesis, 푆 is removable in 퐺 − 푅, and (ii) there is a perfect matching of 퐺 − 푆 containing all edges of 푅. Therefore, in all cases, the graph 퐺 − 푆 is a matching covered graph. Thus, to show that 푆 is an optimal removable class of 퐺, it suffices to show that (푏 + 푝) (퐺 − 푆) = (푏 + 푝) (퐺) if 푆 is a singleton, and (푏 + 푝) (퐺 − 푆) = (푏 + 푝) (퐺) − 1 if 푆 is a doubleton. Similarly, to show that 푅 is an optimal removable removable class of 퐺 − 푆, it suffices to that (푏 + 푝) (퐺 − 푆 − 푅) = (푏 + 푝) (퐺 − 푆) if 푅 is a singleton, and (푏 + 푝) (퐺 − 푆 − 푅) = (푏 + 푝) (퐺 − 푆) − 1 if 푅 is a doubleton. Both these objectives can be achieved by using the monotonicity of the function (푏 + 푝) established in Section 9.1.1.

336

16 The Matching Lattice and Optimal Ear Decompositions

We illustrate this idea by considering the special case in which 푅 = {푒} and 푆 = { 푓 } are singletons. Since {푒} is an optimal class in 퐺, we have that (푏 + 푝) (퐺 − 푒) = (푏 + 푝) (퐺), and since { 푓 } is an optimal removable class in 퐺 − 푒, we have (푏 + 푝) ((퐺 − 푒) − 푓 ) = (푏 + 푝) (퐺 − 푒). As already noted, the fact that there is a perfect matching containing 푒 but not 푓 implies that 퐺 − 푓 is matching covered. Thus, to show that { 푓 } is an optimal removable class in 퐺, it is necessary to show that (푏 + 푝) (퐺) = (푏 + 푝) (퐺 − 푓 ). Similarly, to show that {푒} is an optimal removable class of 퐺 − 푓 , it is necessary to show that (푏 + 푝) (퐺 − 푓 − 푒) = (푏 + 푝) (퐺 − 푓 ). By the monotonicity of the function 푏 + 푝, we have: (푏 + 푝) (퐺) ≤ (푏 + 푝) (퐺 − 푓 ) ≤ (푏 + 푝) (퐺 − 푒 − 푓 ) = (푏 + 푝) (퐺 − 푒) = (푏 + 푝) (퐺). The two desired equalities follow. We leave the proofs of the remaining three cases as Exercise 16.2.6.

ƒ

Exercises 16.2.1 Prove that Theorem 16.9 implies Corollary 16.10. 16.2.2 Let 퐺 ≠ 퐾2 be a matching covered graph. Show that any optimal removable b of 퐺 corresponds to an optimal removable ear 푅 of 퐺. b of the retract 퐺 class 푅

16.2.3 Find an ear decomposition with two double ears of the near-brick shown in Figure 16.6.

Fig. 16.6 A near-brick with 푏 + 푝 = 2

16.2.4 Find an ear decomposition of P ⊕ 퐾4 with just one double ear. 16.2.5 Show that any near-brick 퐺 with 푝 = 1 has an ear decomposition with two double ears. ∗ 16.2.6 Provide proofs of the remaining three cases of Lemma 16.13.

337

16.3 Matching Integral Vectors

⊲16.2.7 Prove that Theorem 16.11 implies Theorem 16.12. ∗ 16.2.8 Give a polynomial-time algorithm which, given a matching covered graph 퐺, produces an optimal ear decomposition of 퐺 and the corresponding sequence of optimal removable ears.

16.3 Matching Integral Vectors ♯ The objective of this section is to provide a brief sketch of Lov´asz’s original approach (in [58]) to the problem of characterizing matching lattices which did not use the idea of 푏-invariant edges. For this reason, in the ensuing discussion, we shall not make use of Corollary 16.8, whose proof depends on the existence of 푏-invariant edges in nonspecial bricks. To do so would be like putting a horse before the cart. But Theorem 16.1 which asserts that every lattice has a basis is a crucial ingredient of the proofs of the main results in this section. Basis matrix of Lat(퐺) Let 퐺 be any matching covered graph. By Theorem 6.13 the dimension of Lin(퐺) is 푑 := 푚 − 푛 + 2 − 푏. Thus, as a consequence of Theorem 16.1, there exists a 푑 × 푚 matrix M whose rows constitute a basis of Lat(퐺). We shall refer to such a matrix as a basis matrix of Lat(퐺). For future reference we record below an observation that follows from the definition of a basis of a lattice. Proposition 16.14 Let M be a 푑 × 푚 basis matrix of Lat(퐺). A vector x in Z퐸 is in Lat(퐺) if and only if there is an integer vector a = (푎 1 , 푎 2 , . . . , 푎 푑 ) such that aM = x. ƒ Matching integral vectors A vector y in the edge space of a matching covered graph 퐺 is matching integral if xy푡 is an integer for each vector x in Lat(퐺). Equivalently, y is matching integral if y(푀) ∈ Z, for each 푀 ∈ M. We shall denote the set of all matching integral vectors of 퐺 by Lat∗ (퐺). (In the theory of ‘lattices’, Lat∗ (퐺) is referred to as the dual of Lat(퐺). See Schrijver [84].) Recall that a vector y0 in the edge space of a matching covered graph 퐺 is matching orthogonal (Exercise 6.2.6) if y0 (푀) = 0, for each perfect matching 푀 of 퐺. Clearly each matching orthogonal vector is matching integral. As specific examples of matching integral vectors, consider the three vectors y0 , z, and y0 + z in the edge space of 퐶6 shown in Figure 16.7. (The coefficients corresponding to edges with no attached labels should be taken as zeroes). The first

16 The Matching Lattice and Optimal Ear Decompositions

338

vector y0 is a matching orthogonal vector. The second vector z, being an integer vector, is clearly matching integral. The third vector, which is the sum of the first two, is thus also a matching integral vector.

1

1

2

− 12

1 1

− 21

− 12 − 12

− 12

1 2

2

− 12

3 2

− 12

3 − 12

− 12

− 12

1

1 2

y0

z

y0 + z

Fig. 16.7 Examples of matching integral vectors

The following proposition is just a restatement of what it means to say that a vector y in R퐸 is a matching integral vector. Proposition 16.15 Let M be a basis matrix of Lat(퐺). A vector y in R퐸 is in Lat★ (퐺) if and only if My푡 is an integer vector. ƒ For any matching covered graph 퐺, the set of matching orthogonal vectors is, by definition, equal to Lin⊥ := (Lin(퐺)) ⊥ , the orthogonal complement of the matching space Lin := Lin(퐺). As noted above, the sum of a matching orthogonal vector and an integer vector is matching integral. Thus: Lin⊥ + Z퐸 ⊆ Lat★.

(16.2)

There are graphs for which the above inequality holds strictly. For example, the 5cycle vector in the edge space of the Petersen graph P which is shown in Figure 16.8 is a matching integral vector, but it cannot be written as the sum of a matching orthogonal vector of P and an integer vector (Exercise 16.3.3). However, Lov´asz [58] showed that the inequality (16.2) holds as an equation if and only if Lat = Lin ∩ Z퐸 . His proof of this assertion relies on three results from linear algebra. The first is an elementary fact concerning solution sets to systems of linear equations. Lemma 16.16 Let 푆 be the set of all solutions to a system Myt = b of linear equations, let 푆0 be the set of all solutions to the corresponding homogeneous system Myt = 0, and suppose that s is any vector in 푆. Then 푆 = 푆 0 + s. ƒ The next two statements follow from a classical result concerning the existence of integer solutions to systems of linear equations with integer coefficients (Corollary 4.1a in [83]). We state them here in notation that is convenient for us:

339

16.3 Matching Integral Vectors

1 2

1 2

1 2

1 2

1 2

Fig. 16.8 A 5-cycle vector w in the edge space of P.

Lemma 16.17 Let M be a 푑 × 푚 integer matrix and let x be a 1 × 푚 integer vector. Then there exists a 1 × 푑 integer vector a satisfying the system aM = x of linear equations if and only if there is no 1 × 푚 vector y such that My푡 is an integer vector, but xy푡 is not an integer. ƒ Thus, to show that there is no integer vector a satisfying the equation aM = x, it suffices to display a vector y such My푡 is an integer vector, but xy푡 is not an integer. As an example, consider the following system of linear equations:     1 1 푎 1 푎 2 = 21 . −1 1

This system has no integral solution because:      1 1 1 12 = , 0 −1 1 12 is an integer vector, whereas (2, 1) ( 21 , 12 ) 푡 =

3 2

is not an integer.

The next lemma is a variant of the one above. Lemma 16.18 Let M be a 푑 × 푚 integer matrix and let b be a 푑 × 1 integer vector. Then there exists a 1 × 푚 integer vector y satisfying the system My푡 = b of linear equations if and only if there is no 1 × 푑 vector a such that aM is an integer vector, but ab is not an integer. ƒ We are now in a position to state and prove the main assertion of this section. In our proof M represents a basis matrix of Lat(퐺). Theorem 16.19 For any matching covered graph 퐺: Lat = Lin ∩ Z퐸 if and only if Lat∗ = Lin⊥ + Z퐸 .

340

16 The Matching Lattice and Optimal Ear Decompositions

Proof We have already noted that the inequality Lin⊥ + Z퐸 ⊆ Lat★ holds regardless of whether or not Lat = Lin ∩ Z퐸 . Let us first suppose that there is an integer vector x in Lin that is not in Lat. By Proposition 16.14, this is equivalent to saying that there is no integer vector a satisfying the system aM = x of linear equations. By Lemma 16.17, it follows that there is a vector y such that My푡 is an integer vector and xy푡 is not an integer. The fact that My푡 is an integer vector means that y is a matching integral vector. Since x is an integer vector in Lin, the fact that xy푡 is not an integer clearly means that y is not in Lin⊥ + Z퐸 (Exercise 16.3.2). Thus, if Lat ⊂ Lin∩Z퐸 , then the strict inequality Lin⊥ +Z퐸 ⊂ Lat★ holds. (Thus, a matching integral vector y such that xy푡 is not an integer serves as a certificate for showing that x is not in Lat(퐺). For example, the 5-cycle vector w shown in Figure 16.8 serves as a certificate to show that 1 is not in the matching lattice of P.) Turning to the converse, suppose that Lat = Lin ∩ Z퐸 . What we need to do is to show that this supposition implies that the equality Lin⊥ + Z퐸 = Lat★ holds. Towards this end, let z be any matching integral vector. Then, by definition, Mz푡 is an integer vector. Let b := Mz푡 and consider the equation: My푡 = b. Using Lemma 16.18, we shall first show that the above equation has an integer solution. Thus, suppose that a is any vector such that aM is an integer vector. As aM is a linear combination of the rows of the basis matrix M of Lin, it belongs to Lin. Thus, being an integer vector it belongs to Lin ∩ Z퐸 which, by our assumption, is equal to Lat(퐺). Therefore, aM = cM, for some integer vector c. Since the rows of M are linearly independent, it follows that a = c is an integer vector. Consequently, ab is an integer. Therefore My푡 = b has an integer solution. Suppose that y = s is an integral solution to My푡 = b. Noting that Lin⊥ is the solution set of the homogeneous system My푡 = 0, we conclude, by Lemma 16.16, that the set of all solutions to the equation My푡 = b is a subset of Lin⊥ + s. This means that if Lat = Lin ∩ Z퐸 then any matching integral vector may be expressed ƒ as the sum of a matching orthogonal vector and an integer vector! The main theorem established by Lov´asz in [58] was the following: Theorem 16.20 If 퐺 is any brick which is not a Petersen brick then Lat★ (퐺) = (Lin(퐺)) ⊥ + Z퐸 . (See Exercise 16.3.1 for an illustration.) Using the above result and Theorem 16.19, Lov´asz deduced that Lat(퐺) = Lin(퐺) ∩ Z퐸 for any such brick. He also showed how this theorem could be used to determine the dimension of the matching space of a matching covered graph over any finite field and, in particular, to derive an attractive statement concerning matching spaces over Z2 . These results are described in the next section.

16.4 Matching Spaces over Z2

341

Exercises 16.3.1 Consider the vector w in the edge space of the pentagonal prism shown in Figure 16.9. Show that w is a matching integral vector and express it as the sum of a matching orthogonal vector and an integer vector.

3/2

3/2

1 −1/2

−1/2

1

1 −1/2

−1/2 −1/2

3/2 1

3/2 1

3/2 Fig. 16.9 A vector in the edge space of the pentagonal prism.

16.3.2 Let x ∈ Lin ∩ Z퐸 and let y be a matching integral vector. Prove that if xy푡 is not an integer then y ∉ Lin⊥ + Z퐸 . 16.3.3 Consider the vector in the edge space of P which is indicated by the edge labels in Figure 16.8 (only nonzero labels are shown). Show that this vector cannot be written as the sum of a matching orthogonal vector and an integer vector.

16.4 Matching Spaces over Z2 ♯ Recall (from Section 6.4) that, given any field F, the matching space of a matching covered graph 퐺, denoted by Lin(퐺, F), is the linear space over F generated by the set of incidence vectors of perfect matchings of 퐺. There is a striking difference between matching spaces over fields of characteristic 2 and those over fields of odd characteristic. For example, if F is any field of characteristic 2, then dim(Lin(P, F)) = 5; on the other hand, if F is any field of odd characteristic, then dim(Lin(P, F)) = 6. In fact, by using the existence of optimal removable classes in bricks (Corollary 15.19), it may be shown that, for any matching covered graph 퐺, dim(Lin(퐺, F)) = 푚 − 푛 + 2 − 푏, if F has odd characteristic, whereas dim(Lin(퐺, Z2 )) = 푚 − 푛 + 2 − (푏 + 푝). See Exercises 16.4.1 and 16.4.2.

342

16 The Matching Lattice and Optimal Ear Decompositions

In this section, for the sake of simplicity and clarity, we shall restrict ourselves to matching spaces over Z2 . We shall describe how Lov´asz was able to use Theorem 16.20 to obtain an elegant characterization of the orthogonal complement of Lin(퐺, Z2 ).

16.4.1 Matching orthogonal vectors Let 퐺 be a matching covered graph, and let F be a field. A vector y in F퐸 is matching Í orthogonal over F if, for any perfect matching 푀 of 퐺, the sum y(푀) = 푒∈ 푀 y(푒) is zero in that field. In other words, matching orthogonal vectors over F are the vectors in the orthogonal complement Lin⊥ (퐺, F) of the matching space Lin(퐺, F). According to Theorem 16.20, any matching integral vector in the edge space R퐸 of a brick 퐺 is the sum of a matching orthogonal vector and an integer vector, provided that 퐺 is not a Petersen brick. Using this result which relates to vector spaces over R, Lov´asz presented in [58] a simple characterization of matching orthogonal vectors over Z2 . We describe below this ingenious application of Theorem 16.20. We begin with the observation that a subset 푄 of the edge set of 퐺 is the support of a vector in Z2퐸 which is matching orthogonal over Z2 if and only if |푄 ∩ 푀 | is even for every perfect matching 푀 of 퐺. (By the support of a vector y ∈ Z2퐸 , denoted by ||y||, we mean the subset of the edge set 퐸 corresponding to nonzero coordinates of vector y.) Example 16.21 Supports of matching orthogonal vectors over Z2 . By Proposition 1.1, for any subset 푆 of the vertex set of a matching covered graph 퐺, and for any perfect matching 푀 of 퐺, the parity of |휕 (푆) ∩ 푀 | coincides with the parity of |푆|. It follows that even cuts are supports of matching orthogonal vectors over Z2 . In addition, if 푛 ≡ 0 (mod 4), all perfect matchings of 퐺 have even cardinality, implying that, in this case, complements of even cuts are also subsets of 퐸 of the required type. On the other hand, if 푛 ≡ 2 (mod 4), all perfect matchings of 퐺 have odd cardinality, implying that, in this case, it is the complement of odd cuts that fit the bill. In the Petersen graph, the edge set of any pentagon is the support of a matching orthogonal vector over Z2 . This is neither a cut nor the complement of a cut! Proposition 16.22 Let y be a vector in Z2퐸 , and let w be the (0, 1)-vector in R퐸 such that ||w|| = ||y||. Vector y is a matching orthogonal vector over Z2 if and only if 21 w is a matching integral vector over R. Proof By definition, y is matching orthogonal over Z2 if y(푀) is zero in Z2 , for any perfect matching 푀 of 퐺. Since 1 + 1 = 0 in Z2 , this amounts to saying that y is matching orthogonal over Z2 if and only if ||y|| ∩ 푀 is an even subset of 퐸. This is clearly equivalent to the statement that w(푀) is an even integer, for any perfect matching 푀 of 퐺. The assertion follows. ƒ

343

16.4 Matching Spaces over Z2

Theorem 16.23 Let 퐺 be a brick which is not a Petersen brick, and let 푄 be a subset of 퐸 such that |푄 ∩ 푀 | is even for all 푀 ∈ M. Then either 푄 is an even cut, or it is the complement of an even cut when 푛 ≡ 0 (mod 4) and the complement of an odd cut when 푛 ≡ 2 (mod 4). Proof Let w denote the characteristic function 휒 푄 of the set 푄. The first step in the proof is to deduce a property of w regarded as a vector over R. 16.23.1 Vector w, regarded as a vector over R, is the sum of a matching orthogonal vector and an integer vector all of whose coordinates are even integers. Proof The hypothesis implies that 21 w, regarded as a vector over R, is a matching integral vector. It now follows from Theorem 16.20 that 12 w is the sum of a matching orthogonal vector and an integer vector. As any scalar multiple of a matching orthogonal vector is also a matching orthogonal vector, the assertion follows. ƒ Let w = x + z,

(16.3)

where x is a matching orthogonal vector and 푧 is a vector all of whose coordinates are even integers. At this stage we invoke a result concerning matching orthogonal vectors stated in Exercise 6.2.6. According to that result, any matching orthogonal vector in the edge space of a brick is vertex induced, which means that there exists a real-valued function 푝 on the vertex set of 퐺 such that x(푢푣) = 푝(푢) + 푝(푣), for any edge 푢푣 of 퐺. (We shall refer to the values of 푝 as vertex labels.) As each coordinate of z is an integer, and since each coordinate of w is 0 or 1, we deduce that each coordinate of x is also an integer. Thus, 푝(푢) + 푝(푣) is an integer, for each edge 푢푣 of 퐺. Moreover, as each coordinate of z is even, we conclude that 16.23.2 There exists a real-valued function 푝 on the vertex set of 퐺 such that, for any edge 푢푣 of 퐺, we have: w(푢푣) = 푝(푢) + 푝(푣)

(mod 2)

(16.4)

Furthermore, since w(푢푣) = 0 or 1, it follows that w(푢푣) = 1 if and only if 푝(푢)+푝(푣) is an odd integer. Equivalently, an edge 푢푣 belongs to the set 푄 if and only if 푝(푢) + 푝(푣) is an odd integer. Equation 16.4 does not imply that vertex labels themselves are integer-valued. However, we claim that all vertex labels are either integral, or all of them are halfintegral. (Each real number 푟 may be written as the sum of an integer 푧 and a non-negative real number 푡 less than 1. We shall refer to 푧 and 푡, respectively, as the integer and and fractional parts of 푟. For example −2.25 = −3 + .75 and, therefore, the integer part of −2.25 is −3 and its fractional part is .75. A number is half-integral if its fractional part is a half.)

344

16 The Matching Lattice and Optimal Ear Decompositions

16.23.3 If 푝 is a vertex labelling of the brick 퐺 which satisfies equation (16.4), then either all labels are integers or all labels are half-integral. Proof If all labels are integral then the assertion holds. Suppose then that not all labels are integral. Let 푢 and 푣 be any two adjacent vertices in 퐺. As noted before, 푝(푢) + 푝(푣) must be an integer. Suppose that 푝(푢) is not integral; let 푡 be its fractional part. Then, 0 < 푡 < 1. Moreover, 푝(푣) is not integral either and its fractional part is 1 − 푡. It follows that if 푡 ≠ 12 then 퐺 has to be bipartite: each edge joins a vertex having 푡 as the fractional part of its label to a vertex having 1 − 푡 as the fractional part of its label. As 퐺, a brick, is not bipartite, we deduce that 푡 = 12 . Thus, all the labels are half-integral. ƒ To obtain the desired conclusion, we now consider two cases depending on whether or not all vertex labels are integers. Case 1 All vertex labels are integers. Let 푆 denote the set of all vertices whose labels are odd integers. Then the labels of all vertices in 푆 are even integers. We then have that 푝(푢) + 푝(푣) is an odd integer if and only if 푢 and 푣 belong to different shores of the cut 휕 (푆). By Statement 16.23.2, this implies that 푄 = 휕 (푆). And, by the hypothesis of the theorem, every perfect matching of 퐺 meets 푄 in an even number of edges, implying that 푄 is an even cut of 퐺. Case 2 All vertex labels are half-integral. Let 푆 denote the set of all vertices whose integer parts are odd integers. In this case, for any two adjacent vertices 푢 and 푣 of 퐺, the sum 푝(푢) + 푝(푣) is an odd integer if and only if 푢 and 푣 are in the same shore of 휕 (푆). Thus, by Statement 16.23.2, it follows that 푄 is the complement of the cut 휕 (푆). Let 푀 be any perfect matching of 퐺; then, clearly, |푀 ∩ 휕 (푆)| = |푀 | − |푀 ∩ 푄|. By hypothesis, |푀 ∩ 푄| is even. If 푛 ≡ 0 (mod 4), then, being a perfect matching of 퐺, 푀 has an even number of edges. Thus, in this case, |푀 ∩ 휕 (푆)| is even, implying that 휕 (푆) is an even cut and that 푄 is the complement of 휕 (푆). A similar argument shows that, when 푛 ≡ 2 (mod 4), the cut 휕 (푆) is odd and 푄 is its complement. ƒ We leave the proof of the following corollary as Exercise 16.4.3. Corollary 16.24 If 퐺 is a brick which is not a Petersen brick, then: (i) when 푛 ≡ 0 (mod 4), Lin⊥ (퐺, Z2 ) is generated by 휒 퐸 = 1 and the incidence vectors of all even cuts; and (ii) when 푛 ≡ 0 (mod 2), Lin⊥ (퐺, Z2 ) is generated by the incidence vector of the complement of a trivial cut and the incidence vectors of all even cuts. ƒ The above corollary may now be used to derive a formula for the dimension of Lin⊥ (퐺, Z2 ). For this purpose, we make use of some known facts concerning the row space (over Z2 ) of the incidence matrix A of a graph 퐺. This space is known as the bond space of 퐺; the support of any vector in it is a cut of 퐺. If 휕 ( 푋) and 휕 (푌 ) are two even cuts, then 휕 ( 푋) △ 휕 (푌 ) is also an even cut. This means that the (mod 2) sum of

16.5 Notes

345

the incidence vectors of any two even cuts is also the incidence vector of some even cut of 퐺. Thus the incidence vectors of even cuts of 퐺 is a subspace of the bond space of 퐺. Furthermore, any basis of this subspace, together with the incidence vector any one odd cut generates the bond space (see Exercise 16.4.4). The dimension of the bond space of a connected graph 퐺 is known to be 푛 − 1 (Theorem 20.7 in [3]), and thus the dimension of its subspace generated by the incidence vectors of even cuts is 푛 − 2. This observation, together with Corollary 16.24 implies the following: Corollary 16.25 If 퐺 is any brick which is not a Petersen brick, then the dimension of Lin⊥ (퐺, Z2 ) is 푛 − 1 and the dimension of Lin(퐺, Z2 ) is 푚 − 푛 + 1. Theorem 16.5 shows that if 퐺 is any cubic brick different from P, then 1 is in the matching space Lin(퐺, Z2 ). We shall not attempt here to present a characterization of matching spaces over Z2 of arbitrary matching covered graphs. This is somewhat more involved than the above mentioned characterization results concerning matching orthogonal vectors. We leave it as a special topic for interested readers to pursue.

Exercises 16.4.1 Adapt the proof of Theorem 16.4 to prove that, for every matching covered graph 퐺, dim(Lin(퐺, Z2 )) = 푚 − 푛 + 2 − (푏 + 푝). ⊲16.4.2 Let F be a field of finite, odd characteristic. Prove that (i) the dimension of Lat(퐺, F) is equal to 푚 − 푛 + 2 − 푏 and (ii) a vector x ∈ F퐸 is in Lat(퐺, F) if and only if it is regular. Hint: in the case 푝 = 0, adapt the proof of Theorem 16.4 and in the case of Petersen bricks, adapt the proof of Lemma 16.6, as 4 has an inverse in F. 16.4.3 Prove Corollary 16.24. 16.4.4 Prove that the incidence vectors of even cuts of a graph 퐺 is a subspace of the bond space of 퐺. Furthermore, any basis of this subspace, together with the incidence vector any one odd cut generates the bond space of 퐺.

16.5 Notes Chapter 38 of Volume A of Schrijver’s three-volume treatise [84] contains a proof of Lov´asz’s characterization of the matching lattice of a brick (Theorem 6.14) without reference to the dual of the matching lattice. It is based on a detailed study of properties of special barriers in matching covered graphs. The inductive nature of our proof, which relies on the existence of 푏-invariant edges in nonspecial bricks (Theorem 15.1), has the advantage that it can be iteratively used to obtain a basis for the matching lattice of any matching covered graph.

346

16 The Matching Lattice and Optimal Ear Decompositions

Recall the perfect matching integer cone Int Cone(퐺) of a matching covered graph 퐺 is the set of all non-negative integer linear combinations of incidence vectors of perfect matchings of 퐺 (see Section 6.3.2). One of the longstanding unresolved conjectures in graph theory is the following: Fulkerson’s conjecture For any 2-connected cubic graph 퐺, the vector 2 = (2, 2, . . . , 2) of all 2’s in the edge space of 퐺 belongs to its perfect matching integer cone Int Cone(퐺). (It is easy to show that 2 belongs Lat(퐺) (see Exercise 16.1.11).) The above conjecture amounts to saying that if 퐺 is any 2-connected cubic graph then there exist six perfect matchings of 퐺 such that each edge of 퐺 belongs to two of these six matchings. The paper by Ma et al (2022, [65] has many references to results related to Fulkerson’s Conjecture. For any matching covered graph 퐺, let us define the Berge number, denoted by 푐(퐺), to be the minimum number of perfect matchings of 퐺 whose union is 퐸 (퐺). An obvious lower bound for 푐(퐺) is the maximum degree Δ(퐺). Clearly, 푐(P) = 5. It is not difficult to prove that 푐(퐺) = Δ(퐺) for bipartite graphs, but we do not know whether the same equality holds for solid graphs. Clearly, if 퐺 is an 푟-regular graph, then 푐(퐺) = Δ(퐺) if and only if 퐺 is 푟-edge colourable. Another interesting question: does the equality 푐(퐺) = Δ(퐺) hold, in general, for all planar matching covered graphs? Berge’s conjecture For every 2-connected cubic graph 퐺, 푐(퐺) ≤ 5. Fulkerson’s Conjecture clearly implies Berge’s Conjecture. The converse is also true, as proved by Mazzuoccolo (2011, [66]).

Chapter 17

Thin Edges in Bricks and Braces

Contents 17.1 17.2

17.3 17.4 17.5 17.6 17.7 17.8

Thin Edges and Thin Barriers in Bricks . . . . . . . . . . . . . . . . . . . . . . . 347 17.1.1 Outline of the proof of the existence of thin edges in bricks 349 Barriers and Tight Cuts in Near-Bricks . . . . . . . . . . . . . . . . . . . . . . . . 352 17.2.1 Near-bricks inherited from bricks . . . . . . . . . . . . . . . . . . . . . 354 17.2.2 Removable edges in near-bricks inherited from bricks . . . . 357 The Existence of Thin Edges in Bricks . . . . . . . . . . . . . . . . . . . . . . . . 361 Building Bricks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 368 17.4.1 The four expansion operations . . . . . . . . . . . . . . . . . . . . . . . . 369 Thin Edges in Braces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 377 17.5.1 Graphs obtained by the deletion of an edge from a brace . . 378 The Existence of Thin Edges in Braces . . . . . . . . . . . . . . . . . . . . . . . . 382 Building Braces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 390 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 392

17.1 Thin Edges and Thin Barriers in Bricks

Thin edges in bricks An edge 푒 of a brick 퐺 is thin if the retract of 퐺 − 푒 is a brick. Since 푏(퐺 − 푒) = 푏( 퐺š − 푒), it follows that every thin edge of a brick 퐺 is a 푏-invariant edge of 퐺. But, in general, not every 푏-invariant edge of a brick is thin (see Exercise 17.1.3). Figure 17.1 shows examples of thin edges, indicated by solid lines.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7_17

347

17 Thin Edges in Bricks and Braces

348

푒

푒 푣2

푣1

푢1 푒 푢2

(푎)

푣2

푣1 푣1

푣3

푣2

(푏)

(푐)

Fig. 17.1 Examples of thin edges in three different bricks

The notion of a thin edge in a brick was introduced by CLM (2006, [15]). However, the definition of a thin edge 푒 in a brick 퐺 given there is in terms of ‘sparsity’ of barriers of 퐺 − 푒; and it was that definition that led to the choice of the adjective ‘thin’. Let 퐺 be a brick, and let 푒 be a 푏-invariant edge of 퐺. Every barrier of 퐺 − 푒 is special (Exercise 17.1.1). Thus, for any barrier 퐵 of 퐺 − 푒, the graph 퐺 − 푒 − 퐵 has just one nontrivial component and it follows that |퐼 (퐵)| = |퐵| − 1, where 퐼 := 퐼 (퐵) denotes the set of isolated vertices of 퐺 − 푒 − 퐵. Thin barriers A barrier 퐵 of 퐺 − 푒 is thin if all vertices in 퐼 have degree two in 퐺 − 푒. In the brick 퐺 shown in Figure 17.1(a), the graph 퐺 − 푒 has a unique nontrivial barrier, namely {푣 1 , 푣 2 }; in the brick 퐺 shown in Figure 17.1(b), the graph 퐺 − 푒 has two nontrivial barriers, namely 퐵1 := {푢 1 , 푢 2 } and 퐵2 := {푣 1 , 푣 2 }; and in the brick 퐺 shown in Figure 17.1(c), the graph 퐺 − 푒 has a unique nontrivial maximal barrier, namely {푣 1 , 푣 2 , 푣 3 }. All these barriers are thin. The following proposition is straightforward to establish (Exercise 17.1.4). Proposition 17.1 Let 푒 be a 푏-invariant edge of a brick 퐺, and let 퐵 be a thin barrier of 퐺 − 푒. Then |퐵| ≤ 3. Furthermore: (i) if |퐵| = 2 then, in 퐺, one end of the edge 푒 is the unique vertex in 퐼 and the other end of 푒 is not in 퐵 ∪ 퐼, and (ii) if |퐵| = 3 then, in 퐺, the edge 푒 joins the two vertices in 퐼. ƒ By the Three Case Lemma 13.5, when 푒 is a 푏-invariant edge of a brick 퐺, the brick of 퐺 − 푒 is obtained by at most two tight cut-contractions, each of which is a barrier cut associated with a special barrier. Based on this result and the above proposition, it is now easy to prove the following theorem. We leave its proof as Exercise 17.1.5. Theorem 17.2 A 푏-invariant edge 푒 of a brick 퐺 is thin if and only if every barrier of 퐺 − 푒 is a thin barrier. ƒ

17.1 Thin Edges and Thin Barriers in Bricks

349

By Theorem 15.1, every brick which is not one of the three special bricks has a 푏-invariant edge. One of the principal results of this chapter is the following strengthening of that theorem: Thin edge theorem for bricks 17.3 Every brick distinct from 퐾4 , 퐶6 and the Petersen graph has a 푏-invariant edge that is thin. As a corollary of the above theorem, we have the interesting property of bricks stated below: Corollary 17.4 Every brick 퐺 of minimum degree greater than three has an edge 푒 such that 퐺 − 푒 is a brick. ƒ It follows from Theorem 17.3 that, given any brick 퐺, there exists a sequence (퐺 1 , 퐺 2 , . . . , 퐺 푘 ) of bricks such that (i) 퐺 1 = 퐺 and 퐺 푘 ∈ {퐾4 , 퐶6 , P}, and (ii) for 1 ≤ 푖 < 푘, the brick 퐺 푖+1 is obtained from 퐺 푖 by deleting an edge and then bicontracting the resulting vertices of degree two. Conversely, as we shall see in Section 17.4, given any brick 퐻, a larger brick can be obtained from 퐻 by means of one of four elementary operations involving bisplitting vertices (as defined in Chapter 12), and adding edges. This yields a simple procedure for generating bricks. In fact, some of these operations may also be used for generating braces.

17.1.1 Outline of the proof of the existence of thin edges in bricks We now recall the definition of index of a 푏-invariant edge of a brick 퐺, given in Chapter 13. Let 퐺 be a brick and let 푒 be a 푏-invariant edge of 퐺. If 퐺 − 푒 is also a brick then we say that 푒 has index zero. If 퐺 − 푒 is not a brick, then we say that 푒 has index one, two or three, depending on which of the three cases stated in Lemma 13.5 holds. Let 퐻 be the brick of 퐺 − 푒. If 퐻 has one contraction vertex and one of the ends of 푒 lies in 푉 (퐻) then 푒 has index one. If 퐻 has two contraction vertices then 푒 has index two. Finally, if 퐻 has one contraction vertex and neither end of 푒 lies in 푉 (퐻) then 푒 has index three. If 푒 is an edge of index zero, then 퐺 − 푒 is a brick and hence 푒 is thin. If a 푏-invariant edge has index greater than zero and is not thin, we shall need to show that there is some other edge that is thin. Let 푒 be a 푏-invariant edge of a brick 퐺 that is not thin. Then 퐺 − 푒 must have a barrier 퐵 that is not thin, and thus 퐼 must have a vertex 푣 of degree three or more in 퐺 − 푒. The set 휕 (푣) − 푒 can be shown to have at least one edge that is 푏-invariant in

350

17 Thin Edges in Bricks and Braces

both 퐺 and 퐺 − 푒. For any such edge 푒 ′ , the brick of 퐺 − 푒 ′ would have at least as many vertices as the brick of 퐺 − 푒. Moreover, if equality holds then 푒 has index two and in that case 휕 (푣) − 푒 would have a second edge, say 푒 ′′ , that is also 푏-invariant in both 퐺 and 퐺 − 푒. Finally, it turns out that equality of the numbers of vertices of the bricks of 퐺 − 푒 ′ and 퐺 − 푒 ′′ is impossible, which implies that the brick of 퐺 − 푒 ′′ has more vertices than the brick of 퐺 − 푒. This leads us to define the pre-rank of a 푏-invariant edge 푒, denoted by 푟 0 (푒), as the order of the brick of 퐺 − 푒. (For example, if 푒 is of index zero, then 푟 0 (푒) = |푉 |.) The idea behind this notion is that the larger the pre-rank of a 푏-invariant edge is, the closer it is to being thin. Let 푒 be a 푏-invariant edge that is not thin, let 퐵 be a maximal barrier of 퐺 − 푒 that is not thin and let 푠 be a vertex in 퐼 whose degree in 퐺 − 푒 is at least three. Then, some edge that is incident with 푠 is also 푏-invariant in 퐺. As explained above, for any such 푒 ′ , we would have 푟 0 (푒 ′ ) ≥ 푟 0 (푒). If perchance, the pre-ranks of 푒 and 푒 ′ happen to be equal, then the index of 푒 would have to be two. In this case, it turns out some other edge 푒 ′′ incident with 푠 is 푏-invariant and its pre-rank is larger than that of 푒. (Two 푏-invariant edges 푒 and 푓 of 퐺, of indices two and one, respectively, may have the same pre-rank, and it may be the case that 푒 is thin while 푓 is not. See Exercise 17.1.7.) Motivated by the above observation, we define the rank 푟 (푒) of a 푏-invariant edge 푒 as follows:  푟 (푒) + 1, if index(푒) = 2 푟 (푒) := 0 푟 0 (푒), otherwise We prove Theorem 17.3 by showing that a 푏-invariant edge of a brick that has the maximum possible rank is thin! As noted above, if a brick has a 푏-invariant edge that is not thin, then there is some other 푏-invariant edge of the brick whose rank is higher. In Section 17.2.2 we shall establish the tools necessary for finding such alternatives, and using these tools we shall present a proof of the existence of thin edges in bricks in Section 17.3.

Exercises ⊲17.1.1 Let 퐺 be a brick, let 푒 be a 푏-invariant edge of 퐺 and let 퐵 denote a barrier of 퐺 − 푒. The graph 퐺 − 푒 − 퐵 has precisely one nonbipartite component. Show that 퐵 is special. Hint: it suffices to show that all the bipartite components of 퐺 − 푒 − 퐵 are trivial. Assume, by way of contradiction, that 퐺 − 푒 − 퐵 has a nontrivial bipartite component 퐺 [푋]. See Figure 17.2. Derive a contradiction by showing that one of 푋+ and 퐵 ∪ 푋− is a nontrivial barrier of 퐺. 17.1.2 In each of the three bricks shown in Figure 17.1, show that the edge 푒 is a thin edge. 17.1.3 Consider the brick 퐺 shown in Figure 17.3. Verify the following statements concerning the edges of 퐺:

17.1 Thin Edges and Thin Barriers in Bricks

351

퐵

푋+ 푋− Fig. 17.2 Illustration for Exercise 17.1.1 푥

푦

푢

푣

푎

푠

푏

푤

푐

푡

Fig. 17.3 The brick 퐺 in Exercise 17.1.3

(i) 푎푏 is not removable; (ii) 푥푦 is 푏-invariant but not thin; (iii) 푏푣 and 푠푡 are thin. (Altogether, there are ten thin edges in this brick.) 17.1.4 Give a proof of Proposition 17.1. 17.1.5 Give a proof of Theorem 17.2. 17.1.6 Let 푒 be a thin edge of a brick 퐺 of order 푛. Verify the following statements: (i) If 푒 has index zero, then 푟 (푒) = 푟 0 (푒) = 푛; (ii) if 푒 has index one, then 푟 (푒) = 푟 0 (푒) = 푛 − 2; (iii) if 푒 has index two, then 푟 (푒) = 푛 − 3 whereas 푟 0 (푒) = 푛 − 4; and (iv) if 푒 has index three, then 푟 (푒) = 푟 0 (푒) = 푛 − 4. 17.1.7 Consider the brick 퐺 depicted in Figure 17.4. Show that 푒 = 48 and 푓 = 13 have the same pre-rank but that 푒 (which is of index 2) is thin whereas the edge 푓 (which is of index 1) is not.

17 Thin Edges in Bricks and Braces

352 4

푥

3

5 푒

1

푓 6 푦

2

7

9

푧

8 Fig. 17.4 Two edges of the same pre-rank, one thin and the other not

17.1.8 Consider the four bricks shown in Figure 17.5. In each case show that edge 푓 indicated by a solid line is thin.

› Hint: (a) Observe that 퐺 − 푓 is obtained by splicing two bricks (one of order four › and one of order six) and use Proposition 5.15; (b) observe that 퐺 − 푓 is isomorphic › to the brick in Figure 17.5(a); (c) observe that 퐺 − 푓 is isomorphic to the graph obtained by the addition of an edge to the brick in Figure 17.5(a); (d) observe that › 퐺 − 푓 is isomorphic to the brick shown in Figure 17.5(b).

17.2 Barriers and Tight Cuts in Near-Bricks ♯ In this section we shall establish the relationship between barriers and tight cuts in near-bricks. It follows from Proposition 4.18 that for any tight cut 퐶 in a near-brick 퐺, precisely one of the shores of 퐶 is bipartite and, furthermore, 퐶 is a barrier cut associated with a special barrier of 퐺. It also follows that every bicritical near-brick is a brick. We have used these facts in many proofs in previous chapters, and shall use them here without explicit reference to it. Let 퐺 be a near-brick, let 퐶 be a tight cut of 퐺, and let 푋 be the bipartite shore of 퐶. We shall refer to the majority part 퐵 of 퐺 [푋] as the external part of 푋, and its minority part 퐼 as the internal part. See Figure 17.6 (the vertices of the barrier associated with the cut 퐶 are indicated by solid dots.) Given any tight cut of a near-brick 퐺, as noted at the beginning of this section, there is a unique special barrier that corresponds to it. Conversely, there is a unique nontrivial tight cut that corresponds to any special barrier of 퐺. We shall use the following notation for nontrivial tight cuts associated with special barriers of nearbricks.

17.2 Barriers and Tight Cuts in Near-Bricks

353

푓 푓

(푎)

(푏)

푓

푓

(푐)

(푑 )

Fig. 17.5 Thin edges in four bricks related to the Petersen graph

푋

푋

퐵

퐼

퐵 퐶 Fig. 17.6 Cuts and their shores in near-bricks

354

17 Thin Edges in Bricks and Braces

Notation: 퐵, 퐼, 푋, 퐶 17.5 Let 퐶 be a tight cut of a near-brick 퐺. Then, we denote the bipartite shore of 퐶 by 푋퐺 (퐶), and its majority and minority parts by 퐵퐺 (퐶) and 퐼퐺 (퐶), respectively. Correspondingly, let 퐵 be a special barrier of a graph 퐺. Then we denote the set of isolated vertices of 퐺 − 퐵 by 퐼퐺 (퐵); the set 퐵 ∪ 퐼퐺 (퐵) by 푋퐺 (퐵); and the cut 휕퐺 ( 푋퐺 (퐵)) by 퐶퐺 (퐵). (In both the above cases, we adopt obvious notational simplifications when the parameters involved are understood from context. For example, if the near-brick 퐺 and tight cut 퐶 are understood, then we simply write 푋, 퐵 and 퐼, instead of 푋퐺 (퐶), 퐵퐺 (퐶) and 퐼퐺 (퐶), respectively. Also, we adopt the usual notational conventions when some of the parameters under consideration are denoted by letters with either subscripts or superscripts. For example, if the tight cut of 퐺 under consideration is 퐶1 , we denote 푋 (퐶1 ), 퐵(퐶1 ) and 퐼 (퐶1 ) simply by 푋1 , 퐵1 , and 퐼1 , respectively. Similarly, if 퐵′ is a special barrier of 퐺, we denote 퐼 (퐵′ ), 푋 (퐵′ ) and 퐶 (퐵′ ) simply by 퐼 ′ , 푋 ′ and 퐶 ′ , respectively.) It should be noted that, in general, not every barrier in a near-brick is special. However, as has been observed in Section 4.4, the following assertion holds (Exercise 4.4.5): Proposition 17.6 Every maximal barrier in a near-brick is special.

ƒ

17.2.1 Near-bricks inherited from bricks A near-brick obtained from the deletion of a 푏-invariant edge from a brick is said to be inherited from that brick. Many properties that do not hold in general for all near-bricks hold for near-bricks inherited from bricks. The purpose of this subsection is to describe some of those properties. Lemma 17.7 Let 퐺 be a brick and let 푒 be a 푏-invariant edge of 퐺. Then all barriers of 퐺 − 푒 are special (Exercise 17.1.1). ƒ Lemma 17.8 Let 퐺 be a brick, let 푒 be a 푏-invariant edge of 퐺 and let 퐵 be a (special) maximal barrier of 퐺 − 푒. Then, graph 퐺/( 푋 → 푥) is a brick. Proof According to the notational conventions we have adopted, the cut 퐶 := 휕 ( 푋) is a tight cut of 퐺 − 푒, the graph 퐺 [푋] induced by 푋 is the bipartite shore of 퐶, and 퐵 is its majority part. Let 퐺 ′ := 퐺/( 푋 → 푥). As every tight cut in a near-brick is a barrier cut, to establish that 퐺 ′ is a brick it suffices to show that (i) 퐺 ′ is a near-brick, and that (ii) 퐺 ′ has no nontrivial barriers.

17.2 Barriers and Tight Cuts in Near-Bricks

355

Let us first show that 퐺 ′ is a near-brick. As the graph 퐺 ′ − 푒 is the nonbipartite (퐶 − 푒)-contraction of the near-brick 퐺 − 푒, it follows that 퐺 ′ − 푒 is a near-brick. If 푒 ∉ 퐶 then 퐺 ′ = 퐺 ′ − 푒, and 퐺 ′ is a near-brick. Suppose then that 푒 ∈ 퐶. As 퐵 is maximal, it follows, by Theorem 3.2, that the component 퐺 [ 푋] of 퐺 − 푒 − 퐵 is critical. Thus, {푥} is a maximal barrier of 퐺 ′ − 푒. In 퐺 ′ , edge 푒 is incident with 푥. Thus, the graph 퐺 ′ is matching covered. As 퐺 ′ − 푒 is a near-brick, it follows, by the monotonicity of function 푏 (Theorem 9.8), that 퐺 ′ is also a near-brick. Now let us proceed to show that 퐺 ′ has no nontrivial barriers. Thus, let 퐵′ be any barrier of 퐺 ′ . Suppose first that 푥 ∈ 퐵′ . In this case, as 퐺 ′ − 푥 is critical, it follows that 퐵′ = {푥}. Alternatively, suppose that 푥 ∉ 퐵′ . In this case, 퐵′ is a barrier of 퐺, where the set 푋 is part of the set of vertices of some component of 퐺 − 퐵 and its contraction produces the corresponding component of 퐺 ′ − 퐵. As 퐺 is a brick, we deduce that 퐵′ is a singleton. In both alternatives, we deduced that 퐵′ is a singleton. As this conclusion holds for any barrier 퐵′ of 퐺 ′ , it follows that 퐺 ′ is bicritical. Hence the assertion holds. ƒ

Three Case Lemma–Revisited In Section 13.2.1 we observed that if 푒 is a 푏-invariant edge of a brick 퐺, then 푒 has to have at least one end in the minority part of the bipartite shore of any nontrivial tight cut of 퐺 − 푒. As a consequence, it follows that the brick of 퐺 − 푒 can be obtained from 퐺 − 푒 by at most two tight cut contractions. If 퐺 − 푒 is itself not a brick, this observation gives rise to the three cases to which the structure of 퐺 −푒 must conform; that is, to the Three Case Lemma 13.5. In much the same way, if 퐵 is any special barrier of 퐺 − 푒, then 푒 has to have at least one end in the set 퐼 of isolated vertices of 퐺 − 푒 − 퐵. As a consequence, it follows that 퐺 − 푒 has at most two maximal nontrivial special barriers. This observation also leads to the same conclusions concerning the structure of 퐺 − 푒. Since our emphasis here is on special barriers of 퐺 − 푒, we restate below that lemma using the notation introduced above. It will prove to be crucial for computing ranks of 푏-invariant edges in bricks. Suppose that 푒 is a 푏-invariant edge of a brick 퐺 and that 퐺 − 푒 is itself not a brick. In this case, 퐺 − 푒 must have nontrivial maximal special barriers. If there is only one such barrier, say 퐵, then the brick of 퐺 − 푒 is (퐺 − 푒)/( 푋 → 푥) which is obtained from 퐺 − 푒 by contracting the bipartite shore of the tight cut of 퐺 − 푒 associated with 퐵. However, if 퐵1 and 퐵2 are two nontrivial maximal special barriers of 퐺 − 푒, the tight cuts 퐶1 and 퐶2 associated with 퐵1 and 퐵2 , respectively, may cross each other. In this case, it would not be right to conclude that the brick of 퐺 − 푒 is (퐺 − 푒)/( 푋1 → 푥1 )/( 푋2 → 푥2 ). Clearly, being distinct maximal barriers of 퐺 − 푒, by Theorem 3.11, 퐵1 and 퐵2 are disjoint, but 푋1 := 퐵1 ∪ 퐼1 and 푋2 := 퐵2 ∪ 퐼2 need not be disjoint. This situation occurs when 퐵2 ∩ 퐼1 is nonempty (which implies that 퐵1 ∩ 퐼2 is also nonempty, see Exercise 17.2.1).

For example, consider the brick 퐺 shown in Figure 17.7. The edge 푒 is 푏-invariant in 퐺, and 퐺 − 푒 has two maximal nontrivial barriers which are indicated by large black and white discs, respectively. The bipartite shores of the tight cuts associated with these barriers are 푋1 and 푋2 , and the cuts 퐶1 and 퐶2 cross each other.

17 Thin Edges in Bricks and Braces

356 푋2 푎1

푎2

푎3

푎4

푋1 푏1

푏2

푎5

푎6

푏6

푏3

푒

푏4

푏5 푐

푑

Fig. 17.7 The case in which the tight cuts associated with two maximal barriers 퐵1 and 퐵2 of 퐺 − 푒 cross (퐵1 = {푏1 , 푏2 , 푏3 , 푏4 , 푏5 } with 퐼1 = {푎1 , 푎2 , 푎3 , 푎4 } and 퐵2 = {푎1 , 푎2 , 푎5 , 푎6 } with 퐼2 = {푏1 , 푏2 , 푏6 })

We now present the version of the Three Case Lemma, stated in terms of maximal nontrivial barriers of 퐺 − 푒, that we shall use in this chapter. Lemma 17.9 (The Three Case Lemma) Let 퐺 be a brick, and let 푒 be a 푏-invariant edge of 퐺. Then, one the following alternatives holds: (i) index(e) = 0: graph 퐺 − 푒 is a brick, or (ii) index(e) = 1: graph 퐺 − 푒 has precisely one maximal nontrivial barrier, 퐵, the edge 푒 has only one end in 퐼 and the graph (퐺 − 푒)/( 푋 → 푥) is a brick, or (iii) index(e) = 2: graph 퐺 − 푒 has precisely two maximal nontrivial barriers, 퐵1 and 퐵2 , and • edge 푒 has one end in 퐼1 − 푋2 and the other in 퐼2 − 푋1 , • the graph (퐺 − 푒)/( 푋1 → 푥1 ) is a near-brick, 퐵2 − 푋1 is its unique nontrivial maximal barrier and 푋2 − 푋1 is the bipartite shore of a nontrivial tight cut, • the graph (퐺 − 푒)/( 푋1 → 푥1 )/(( 푋2 − 푋1 ) → 푥2 ) is a brick (see Figure 17.7 — up to multiple edges, by Corollary 4.13, this is the same as the brick (퐺 − 푒)/(( 푋1 − 푋2 ) → 푥1 )/( 푋2 → 푥2 )), or (iv) index(e) = 3: graph 퐺 − 푒 has precisely one maximal nontrivial barrier, 퐵, the edge 푒 has both ends in 퐼 and the graph (퐺 − 푒)/( 푋 → 푥) is a brick. Rank Computations If an edge 푒 is of index zero, then 퐺 − 푒 is a brick. In this case, 푟 (푒) = 푟 0 (푒) = |푉 |. We describe below the procedure for finding the rank of 푒 when the index of 푒 is positive.

17.2 Barriers and Tight Cuts in Near-Bricks

357

Suppose that a 푏-invariant edge 푒 of a brick 퐺 has index one or three and suppose that 퐵 is the only nontrivial maximal barrier of 퐺 − 푒. Then the edge 푒 might have both its ends in 퐼 (index(푒) = 3) or one end in 퐼 and one end in 푋 (index(푒) = 1). In either case, the graph obtained from 퐺 − 푒 by contracting 푋 to a single vertex is the brick of 퐺 − 푒. Thus, in this case: 푟 (푒) = 푟 0 (푒) = |푉 | − | 푋 | + 1 = | 푋 | + 1.

(17.1)

Now suppose that a 푏-invariant edge 푒 of a brick 퐺 has index two and suppose that 퐵1 and 퐵2 are the two nontrivial maximal barriers of 퐺 − 푒. In this case, 퐵′2 := 퐵2 − 푋1 is the only maximal nontrivial barrier of 퐺 ′ := (퐺 − 푒)/( 푋1 → 푥1 ) and 퐺 ′ /(( 푋2 − 푋1 ) → 푥2 ) is the brick of 퐺 − 푒. Thus, the pre-rank 푟 0 (푒) of 푒 is equal to (|푉 | − | 푋1 |) + 1 − (| 푋2 − 푋1 |) + 1 = | 푋1 | − | 푋2 − 푋1 | + 2, implying the following formula for its rank: 푟 (푒) = 푟 0 (푒) + 1 = | 푋1 | − | 푋2 − 푋1 | + 3 ≤ | 푋1 |.

(17.2)

For example, if 푒 is the indicated 푏-invariant edge in the brick shown in Figure 17.7, 푟 (푒) = | 푋1 | − | 푋2 − 푋1 | + 3 = 5 − 3 + 3 = 5. The next proposition is a simple consequence of (17.1) and (17.2).

Proposition 17.10 Suppose that 퐵 is a maximal barrier of 퐺 − 푒 and that 휕 ( 푋) is the tight cut of 퐺 − 푒 associated with 퐵. Then 푟 (푒) ≤ | 푋 | + 1, with equality only if the index of e is not equal to two. ƒ Note that 푟 0 (푒) is always even as it is the order of a brick. Thus, 푟 (푒) is odd only when 푒 is of index 2.

17.2.2 Removable edges in near-bricks inherited from bricks Suppose that 퐺 is a brick and 푒 is a 푏-invariant edge of 퐺 that is not thin. Our objective then is to show that there is some other 푏-invariant edge of 퐺 that is thin. The first step is to show that there exist 푏-invariant edges (thin or not) other than 푒. This section is devoted to the development of tools that are necessary for finding such edges. If 퐵 is a nontrivial maximal barrier of 퐺 − 푒, it turns out that one can always find a 푏-invariant edge of 퐺 incident with a vertex in 퐼 that is of degree greater than two in 퐺 − 푒. The main tool used in establishing this result is Lemma 8.19 proved in Chapter 8. We shall also use the following easily proved proposition. Proposition 17.11 Let 퐺 be a matching covered graph, let 푒 be a removable edge of 퐺, and let 푒 ′ be an edge of 퐺 − 푒 that is removable in 퐺 − 푒. Edge 푒 ′ is removable in 퐺 if and only if 퐺 has a perfect matching that contains edge 푒 but does not contain edge 푒 ′ . ƒ

358

17 Thin Edges in Bricks and Braces

Alternative choices for 푏-invariant edges Theorem 17.12 Let 퐺 be a brick, let 푒 be a 푏-invariant edge of 퐺 and let 퐵 be a maximal nontrivial (special) barrier of 퐺 − 푒. Let 푣 be a vertex of 퐼 that has degree strictly greater than two in 퐺 − 푒. Then, the following properties hold: (i) At most one edge of 휕 (푣) − 푒 is not 푏-invariant in 퐺 − 푒. (ii) At most one edge of 휕 (푣)−푒 is 푏-invariant in 퐺−푒 without being 푏-invariant in 퐺. (iii) If 푒 is of index two, then every edge of 휕 (푣) − 푒 that is 푏-invariant in 퐺 − 푒 is also 푏-invariant in 퐺. Proof Let 퐶 := 휕 ( 푋), let 퐻 := (퐺 − 푒)/( 푋 → 푥) and let 퐺 1 := 퐺/( 푋 → 푥). Of these two, 퐻 is bipartite because 푋 is the bipartite shore of 퐶 − 푒, where 퐼 + 푥 and 퐵 are the two parts of its bipartition. The graph 퐺 1 is a brick, by Lemma 17.8. 17.12.1 Let 푒 ′ be any edge in 휕 (푣) −푒 that is removable in 퐻. Then 푒 ′ is a 푏-invariant edge of 퐺 − 푒. Furthermore, if the edge 푒 does not depend on 푒 ′ in 퐺, then 푒 ′ is a 푏-invariant edge of 퐺 itself. Proof The edge 푒 ′ is in 퐻 − 퐶. The two (퐶 − 푒)-contractions of 퐺 − 푒 − 푒 ′ are 퐻 − 푒 ′ and 퐺 1 − 푒. The graph 퐻 − 푒 ′ is bipartite and matching covered, because 퐻 is the bipartite (퐶 − 푒)-contraction of 퐺 − 푒 and 푒 ′ is removable in 퐻. The graph 퐺 1 − 푒 is a near-brick, because 푒 is 푏-invariant in 퐺 and 퐺 1 − 푒 is the nonbipartite (퐶 − 푒)-contraction of 퐺 − 푒. Thus, the graph 퐺 − 푒 − 푒 ′ is the splicing of 퐻 − 푒 ′ and 퐺 1 − 푒; hence 퐺 − 푒 − 푒 ′ is matching covered. Consequently, 푒 ′ is removable in 퐺 − 푒. Moreover, the cut 퐶 − 푒, a tight cut of 퐺 − 푒, is tight in 퐺 − 푒 − 푒 ′ . Thus, 퐺 − 푒 − 푒 ′ is a near-brick. As 퐺 − 푒 is a near-brick, we conclude that 푒 ′ is 푏-invariant in 퐺 − 푒. Now suppose that 푒 does not depend on 푒 ′ in 퐺. Then there is a perfect matching of 퐺 which contains 푒 but not 푒 ′ . It follows from Proposition 17.11 that 푒 ′ is removable in 퐺. As 푒 ′ is 푏-invariant in 퐺 − 푒, by the monotonicity of the function 푏 we deduce that 푏(퐺) ≤ 푏(퐺 − 푒 ′ ) ≤ 푏(퐺 − 푒 − 푒 ′ ) = 푏(퐺 − 푒) = 푏(퐺). Consequently, the edge 푒 ′ is 푏-invariant in 퐺 itself. This proves the second part of the assertion. ƒ

By Lemma 8.19, at most one edge of 휕 (푣) − 푒 is not removable in 퐻, hence at most one edge of 휕 (푣) − 푒 is not 푏-invariant in 퐺 − 푒. This completes the proof of the first item of the assertion. Also, clearly, edge 푒 could be dependent on at most one edge incident with 푣. Thus, at most one edge of 휕 (푣) − 푒 is 푏-invariant in 퐺 − 푒 but not 푏-invariant in 퐺. This proves the second part of the assertion. Suppose that 푒 has index two. Assume, contrary to the third item of the assertion, that 휕 (푣) − 푒 contains an edge, say 푒 ′ , that is 푏-invariant in 퐺 − 푒 but not 푏-invariant in 퐺. From statement 17.12.1 we infer that 푒 depends on 푒 ′ in 퐺. That is, every perfect matching of 퐺 that contains edge 푒 also contains edge 푒 ′ . Therefore, graph

17.2 Barriers and Tight Cuts in Near-Bricks

359

퐺 − 푒 ′ has barriers that include both ends of edge 푒. Let 퐵′ be a maximal barrier of 퐺 − 푒 ′ that includes both ends of 푒. Graph 퐺 − 푒 − 푒 ′ is matching covered; therefore 푒 is the only edge of 퐺 that depends on 푒 ′ . We conclude that in 퐺 no edge other than 푒 has both its ends in 퐵′ . As 푒 has index two, one of the ends of 푒, say, 푦, is in 퐺 1 − 푥. Thus, in 퐺 1 , the ends of 푒 are 푦 and the contraction vertex 푥. 17.12.2 The vertex 푦 is the only vertex of 퐺 1 − 푥 in 퐵′ . Proof Certainly, 푦, an end of 푒, is in 퐵′ . Let 푧 be any vertex of 퐺 1 − 푥 − 푦. The graph 퐺 1 is a brick; hence 퐺 1 − 푦 − 푧 has a perfect matching, 푀1 . Let 푓 be the edge of 푀1 incident with the contraction vertex 푥 of 퐺 1 . Thus, 푓 is the edge of 푀1 in 퐶. As 푦 is an end of 푒 in 퐺 1 , the matching 푀1 does not contain the edge 푒. Thus, 푓 is an edge of 퐶 − 푒. As 푒 ′ is not in 퐶, the edges 푒 ′ and 푓 are distinct. As 푒 ′ is removable in 퐻, the graph 퐻 − 푒 ′ has a perfect matching, 푀 퐻 , which contains the edge 푓 . Thus, 푀 := 푀1 ∪ 푀 퐻 is a perfect matching of 퐺 − 푦 − 푧 − 푒 ′ . Consequently, 푧 ∉ 퐵′ . This conclusion holds for each vertex 푧 of 퐺 1 − 푥 − 푦. We deduce that ƒ 푉 (퐺 1 − 푥) ∩ 퐵′ = {푦}. All the vertices of 퐺 1 − 푥 − 푦 are in components of 퐺 − 푒 ′ − 퐵′ . The graph 퐺 1 , a brick, is 3-connected. Thus, 퐺 1 − 푥 − 푦 is connected. We deduce that all the vertices of 퐺 1 − 푥 − 푦 are in the same component, 퐾, of 퐺 − 푒 ′ − 퐵′ . 17.12.3 Every neighbour of 푦 in 퐺 − 푒 is in 푉 (퐺 1 − 푥). Proof By the Three Case Lemma, 푦 is an isolated vertex of 퐺 1 − 푒 − 퐵2 , where 퐵2 is the maximal nontrivial barrier of 퐺 1 − 푒. As 퐺 1 is a brick, 퐵2 is not a barrier of 퐺 1 ; hence the end 푥 of 푒 in 퐺 1 is not in 퐵2 . Thus, every neighbour of 푦 in 퐺 − 푒 is ƒ in 푉 (퐺 1 − 푥). We conclude that all the neighbours of 푦 in 퐺 − 푒 are in 퐾. In that case, 퐵′′ := 퐵′ − 푦 is a nonempty barrier of 퐺 − 푒 ′ and 푉 (퐾) + 푦 is the set of vertices of an even component of 퐺 − 푒 ′ − 퐵′′ . This is not possible (by Proposition 2.6). We deduce that 푒 ′ is 푏-invariant in 퐺. This conclusion holds for each edge 푒 ′ ∈ 휕 (푣) − 푒 which is 푏-invariant in 퐺 − 푒, provided index(푒) = 2. ƒ Figure 17.8 depicts a brick 퐺, where 푣푤 is the only edge of 휕 (푣) which is removable in both 퐺 and 퐺 − 푒, thereby illustrating the necessity of index(푒) = 2 in Theorem 17.12. 푣

푒

푤 Fig. 17.8 Edge 푣푤 is the unique edge in 휕(푣 ) that is removable in both 퐺 and 퐺 − 푒.

360

17 Thin Edges in Bricks and Braces

Given any 푏-invariant edge 푒 of a brick 퐺, and a maximal barrier 퐵 of 퐺 − 푒 that is not thin, our objective is to show that there is some edge of 퐺 whose rank is higher than that of 푒. We shall in fact see in the next section that some edge incident with an isolated vertex of 퐺 − 푒 − 퐵 has higher rank than that of 푒. Our proof of this statement does not rely on the following proposition, whose proof is left as Exercise 17.2.6. Proposition 17.13 Let 푠 be any isolated vertex of 퐺 − 푒 − 퐵 having degree at least three in 퐺 − 푒, and let 푒 ′ be any edge incident with 푠 that is 푏-invariant in both 퐺 − 푒 and 퐺. Then, 푟 (푒 ′ ) ≥ 푟 (푒). ƒ

Exercises ∗ 17.2.1 Let 푒 be a 푏-invariant edge of a brick 퐺 and suppose that 퐵1 and 퐵2 are two maximal nontrivial barriers of 퐺 − 푒. Show that |퐵1 ∩ 퐼2 | = |퐵2 ∩ 퐼1 |. ∗ 17.2.2 Let 푒 be a 푏-invariant edge of a brick 퐺 of index two, and let 퐵1 and 퐵2 denote the two maximal nontrivial barriers of 퐺 − 푒. Suppose that the corresponding cuts 퐶1 and 퐶2 cross. Prove that (i) the graph 퐺 [푋1 ∩ 푋2 ] is equipartite and (ii) the sets 푋1 − 푋2 and 푋2 − 푋1 are bipartite shores of nontrivial tight cuts of 퐺 − 푒. 17.2.3 Consider the brick 퐺 shown in Figure 17.7. Determine the two graphs (퐺 − 푒)/( 푋1 → 푥1 )/(( 푋2 − 푋1 ) → 푥2 ) and (퐺 − 푒)/(( 푋1 − 푋2 ) → 푥1 )/( 푋2 → 푥2 ), and verify that they are isomorphic, up to multiple edges. 17.2.4 Prove Proposition 17.11. 17.2.5 Verify that the graph 퐺 in Figure 17.8 is a brick and also that 푣푤 is the only edge incident with 푣 which is removable in both 퐺 and 퐺 − 푒. ⊲17.2.6 Prove Proposition 17.13. Towards this goal, follow the two steps below: (i) Prove that 퐺 − 푒 − 푒 ′ has a tight cut decomposition which produces a brick which is equal to a brick of (퐺 − 푒)/( 푋 → 푥), having precisely the same set of contraction vertices. (ii) Prove that every brick of 퐺 − 푒 ′ has at least as many vertices as a brick of 퐺 − 푒 − 푒 ′ . Hint: for every cut 퐷 of 퐺, if 퐷 − 푒 ′ is tight in 퐺 − 푒 ′ then 퐷 − 푒 − 푒 ′ is tight in 퐺 − 푒 − 푒 ′ .

361

17.3 The Existence of Thin Edges in Bricks

17.3 The Existence of Thin Edges in Bricks ♯ In this section we shall present a proof of Theorem 17.3. By Theorem 15.1, any brick 퐺 different from 퐾4 , 퐶6 and P has a 푏-invariant edge 푒. If that edge 푒 is not itself thin, our strategy, as explained earlier, is to show that there exists another 푏-invariant edge 푓 of 퐺 such that 푟 ( 푓 ) > 푟 (푒). We start with a few pertinent observations before we turn to a proof of that statement. Thus suppose that 푒 is a 푏-invariant edge of 퐺 that is not thin. Then the near-brick 퐺 − 푒 has at least one nontrivial maximal barrier that is not thin. Let 퐵 be one such barrier of 퐺 − 푒 and let 푠 be a vertex of 퐼 that has degree greater than two in 퐺 − 푒. We have seen in the last section that at least one edge of 퐺 − 푒 incident with 푠 is 푏-invariant in both 퐺 and 퐺 − 푒. Let 푒 ′ be such an edge. Let 퐺 1 := 퐺/( 푋 → 푥) and 퐺 2 := 퐺/( 푋 → 푥) denote the two 퐶-contractions of 퐺. The following facts are easily observed: 17.13.1 The edge 푒 ′ has both its ends in 푋, whereas the edge 푒 has either one (Figure 17.9(a)) or two ends in 퐼 (Figure 17.9(b)). (The former case must occur when index(푒) ∈ {1, 2} and the latter case must occur when index(푒) = 3.) ƒ

퐵

퐵

푒′

푒′

푠

푠

퐼

푋

푋 푒

푒

푋 (푎)

Fig. 17.9 The two 푏-invariant edges 푒 and 푒

퐼

푋



(푏)

As 퐵 is a barrier of 퐺 − 푒, then it certainly is a barrier of 퐺 − 푒 − 푒 ′ . Conversely, since the edge 푒 ′ has an end in the barrier 퐵 of 퐺 − 푒, it follows that any barrier 퐵′ of 퐺 − 푒 − 푒 ′ which contains 퐵 is also a barrier of 퐺 − 푒. Since 퐵 is a maximal barrier of 퐺 − 푒, it follows that 퐵 is a maximal barrier of 퐺 − 푒 − 푒 ′ . The following proposition records this observation: Proposition 17.14 The set 퐵 is a maximal barrier of 퐺 − 푒 − 푒 ′ . The following lemma plays a crucial role in the proof of Theorem 17.3.

ƒ

362

17 Thin Edges in Bricks and Braces

Lemma 17.15 Let 퐶 ′ be a nontrivial cut of 퐺 such that 퐶 ′ − 푒 ′ is tight in 퐺 − 푒 ′ . If 퐶 ′ does not cross 퐶 then 푋 ′ is a proper subset of 푋. Proof The cut 퐶 ′ − 푒 ′ is tight in 퐺 − 푒 ′ but 퐶 − 푒 ′ = 퐶 is not. Thus, 퐶 ′ ≠ 퐶. As 퐶 ′ does not cross 퐶, the possible containment relations between 푋 and 푋 ′ are: (i) 푋 ′ ⊂ 푋, (ii) 푋 ⊂ 푋 ′ , (iii) 푋 ⊂ 푋 ′ , or (iv) 푋 ′ ⊂ 푋. Our task is to show that (iv) holds. We do this by showing that the other possibilities cannot occur. Edge 푒 ′ has at least one end in 푋 ′ , but has no end in 푋. Therefore (i) cannot hold. By Lemma 17.8, the graph 퐺 2 = 퐺/( 푋 → 푥) is a brick, implying that 퐺 [푋], which has neither 푒 nor 푒 ′ as an edge, is nonbipartite. But 퐺 [푋 ′ ] − 푒 − 푒 ′ is bipartite. Therefore (ii) does not hold either. One end of the edge 푒 ′ is in 퐵 and one end in 퐼. Therefore, if 푋 were a subset of ′ 푋 , an end of 푒 ′ would be in 퐵′ . But this would mean that 퐵′ , which is a nontrivial barrier of 퐺 − 푒 ′ is also a nontrivial barrier of the brick 퐺 itself. This is absurd. Thus, (iii) also cannot hold. ƒ We are now in position to prove the result concerning ranks that leads to a proof of Theorem 17.3. The rank augmentation theorem for bricks 17.16 Let 푒 be a 푏-invariant edge in a brick 퐺 and suppose that 푒 is not thin. Then 퐺 has a 푏-invariant edge 푓 such that 푟 ( 푓 ) > 푟 (푒). Proof As above, let 푠 be a vertex in 퐼 which has degree at least three in 퐺 − 푒, and let 푒 ′ be any edge incident with 푠 which is 푏-invariant in both 퐺 − 푒 and 퐺. We divide the proof of the assertion into two main cases. Case 1 For every maximal nontrivial barrier 퐵′ of 퐺 − 푒 ′ , the corresponding set 푋 ′ is a proper subset of 푋. By the Three Case Lemma, the brick of 퐺 − 푒 ′ is obtained from 퐺 − 푒 ′ by the contraction of proper subsets of 푋. As 푋 is nontrivial and odd, the set of vertices of the brick of 퐺 − 푒 ′ includes not only 푋, but also at least three more vertices. From this, by Proposition 17.10, 푟 (푒 ′ ) ≥ 푟 0 (푒 ′ ) ≥ | 푋 | + 3 ≥ 푟 (푒) + 2, implying that 푟 (푒 ′ ) > 푟 (푒). Thus, in this case, the assertion of the Theorem holds, with 푓 := 푒 ′ . (See Example 17.17 for an illustration.) Example 17.17 Consider the 푏-invariant edge 푒 = 푢푣 in the brick 퐺 shown in Figure 17.10. The index of 푒 is one, 푋 = {푎 1 , 푎 2 , 푎 3 , 푎 4 , 푏 1 , 푏 2 , 푏 3 , 푏 4 , 푏 5 }, and 푟 (푒) = 4. If 푒 ′ is the edge 푎 4 푏 2 , then its index is one, {푎 1 , 푎 2 , 푎 3 } is the unique nontrivial maximal barrier of 퐺 − 푒 ′ , and 푋 ′ = {푎 1 , 푎 2 , 푎 3 , 푏 1 , 푏 2 } is a proper subset of 푋. It can be verified that 푟 (푒 ′ ) = 8 > 푟 (푒). (Edge 푎 2 푏 2 , a 푏-invariant edge of index two, is another edge which satisfies the hypothesis of Case 1.)

363

17.3 The Existence of Thin Edges in Bricks 푎1

푎2

푎3

푎4 = 푢

푣

푒

푥

푏1

푏2

푏3

푏4

푦

푏5

Fig. 17.10 Illustration for Case 1, Theorem 17.16

Case 2 퐺 − 푒 ′ has a maximal nontrivial barrier 퐵′ such that its corresponding set 푋 ′ is not a proper subset of 푋. In this case, by Lemma 17.15, the corresponding cut 퐶 ′ crosses the cut 퐶. Lemma 17.18 Let 퐶 ′ be a cut of 퐺 such that 퐶 ′ − 푒 ′ is tight in 퐺 − 푒 ′ . Suppose that 퐶 and 퐶 ′ cross. Let 퐷 := 휕 ( 푋 − 푋 ′ ) and let 퐷 ′ := 휕 ( 푋 ′ − 푋). Then, 푋 ∩ 푋 ′ and 푋 ∩ 푋 ′ are both even and the following properties hold: (i) cut 퐷 is nontrivial and 퐷 − 푒 − 푒 ′ is tight in 퐺 − 푒 − 푒 ′ , 푋 − 푋 ′ is its bipartite shore, 퐵 − 푋 ′ the external part and 퐼 − 푋 ′ the internal part of 푋 − 푋 ′ , (ii) cut 퐷 ′ − 푒 is tight in 퐺 − 푒, 푋 ′ − 푋 is its bipartite shore, 퐵′ − 푋 the external part and 퐼 ′ − 푋 the internal part of 푋 ′ − 푋, and (iii) edge 푒 ′ has one end in 퐼 − 푋 ′ , the other in 퐵 ∩ 퐼 ′ . (See Figures 17.12 and 17.13.) Proof The cuts 퐶 − 푒 and 퐶 ′ − 푒 − 푒 ′ are both tight in 퐺 − 푒 − 푒 ′ . Therefore the four subgraphs of 퐺 − 푒 − 푒 ′ induced by 푋, 푋, 푋 ′ and 푋 ′ are connected. Thus, 퐺 − 푒 − 푒 ′ has four edges 푣 푖 푤 푖 , 1 ≤ 푖 ≤ 4, where 푣 1 ∈ 푋 ∩ 푋 ′ , 푤 1 ∈ 푋 ′ − 푋, 푣 2 ∈ 푋 − 푋 ′ , ′ 푤 2 ∈ 푋 ∩ 푋 ′ , 푣 3 ∈ 푋 ∩ 푋 ′ , 푤 3 ∈ 푋 − 푋 ′ , 푣 4 ∈ 푋 ′ − 푋 and 푤 4 ∈ 푋 ∩ 푋 . See Figure 17.11. The end in 푋 of any edge in the cut 퐶 − 푒 is in 퐵, and thus the two vertices 푣 1 and 푣 2 are in 퐵. For a similar reason, the two vertices 푣 3 and 푣 4 are in 퐵′ . 17.18.1 The two sets 퐵 and 퐵′ are disjoint.

ƒ

Proof Suppose that 퐵 ∩ 퐵′ ≠ ∅. By Proposition 17.14, the set 퐵 is a maximal barrier of 퐺 − 푒 − 푒 ′ . Clearly, 퐵′ is a barrier of 퐺 − 푒 − 푒 ′ . Thus, our assumption implies that 퐵′ ⊆ 퐵. This is a contradiction, because 퐵 is a subset of 푋, whereas 퐵′ contains the vertex 푣 4 in 푋 ′ − 푋. ƒ Let us now prove that 푋 ∩ 푋 ′ is even. Suppose that 푋 ∩ 푋 ′ is odd. As 퐶 − 푒 and are tight cuts of 퐺 − 푒 − 푒 ′ , then, by Theorem 4.12, cuts 휕 ( 푋 ∩ 푋 ′ ) − 푒 − 푒 ′ and 휕 ( 푋 ∩ 푋 ′ ) − 푒 − 푒 ′ are also tight cuts in 퐺 − 푒 − 푒 ′ , and each shore of these tight cuts induces a connected subgraph of 퐺 − 푒 − 푒 ′ . The subgraphs of 퐺 − 푒 − 푒 ′ induced

퐶 ′ − 푒 − 푒 ′

17 Thin Edges in Bricks and Braces

364 푋 ′

푣3

푤3

푋 푣1

푣2 퐶

푤1

푤2 푣4

푤4 퐶 ′

Fig. 17.11 The four edges 푣푖 푤푖 , 1 ≤ 푖 ≤ 4

by 푋 and by 푋 ′ are both bipartite; so too is (퐺 − 푒 − 푒 ′ ) [푋 ∩ 푋 ′ ]. Moreover, as (퐺 − 푒 − 푒 ′ ) [푋 ∩ 푋 ′ ] is connected, its bipartition is unique. Thus, {퐵 ∩ 푋 ′ , 퐼 ∩ 푋 ′ } = {퐵′ ∩ 푋, 퐼 ′ ∩ 푋 }. The vertex 푣 1 is in 퐵 and also in the external part of 푋 ∩ 푋 ′ . Thus, the external part of 푋 ∩ 푋 ′ is 퐵 ∩ 푋 ′ . Likewise, 푣 3 is in 퐵′ and also in the external part of 푋 ∩ 푋 ′ . Thus, the external part of 푋 ∩ 푋 ′ is 퐵′ ∩ 푋. We conclude that 퐵 ∩ 푋 ′ = 퐵′ ∩ 푋. This is a contradiction, because 퐵 and 퐵′ are disjoint by statement 17.18.1. We conclude that 푋 ∩ 푋 ′ is even. It follows that 푋 ∩ 푋 ′ is also even. By Theorem 4.12, 퐷 − 푒 − 푒 ′ and 퐷 ′ − 푒 are tight cuts of 퐺 − 푒 − 푒 ′ . See Figure 17.12. Clearly, 푋 − 푋 ′ induces a bipartite connected subgraph of 퐺 − 푒 − 푒 ′ . Thus, {퐵 − 푋 ′ , 퐼 − 푋 ′ } is the bipartition of (퐺 − 푒 − 푒 ′ ) [푋 − 푋 ′ ]. The vertex 푣 2 is in 퐵 and in the external part of 푋 − 푋 ′ . Thus, 퐵 − 푋 ′ is the external part of 푋 − 푋 ′ , and 퐼 − 푋 ′ is its internal part. Likewise, 푋 ′ − 푋 is the bipartite shore of 퐷 ′ − 푒. The set 퐵′ − 푋 is the external part of 푋 ′ − 푋, and 퐼 ′ − 푋 is its internal part.

To complete the proof, recall that edge 푒 ′ has its end 푠 in 퐼, and its other end, say 푡 ′ , in 퐵. We wish to show that 푠 is in 퐼 − 푋 ′ . Suppose, to the contrary, that 푠 is in 푋 ′ . The cut 퐶 ′ is nontrivial and 퐶 ′ − 푒 ′ is tight in 퐺 − 푒 ′ ; hence 푠 ∉ 퐵′ ; otherwise 퐶 ′ is tight in 퐺. It follows that 푠 ∈ 퐼 ′ . Then, 푠 ∈ 퐼 ∩ 퐼 ′ . Let 푓 be any edge of 퐺 − 푒 − 푒 ′ incident with 푠. The end of 푓 distinct from 푠 is in 퐵 ∩ 퐵′ , a contradiction, as 퐵 and 퐵′ are disjoint by statement 17.18.1. It follows that 푠 ∈ 푋 − 푋 ′ , and hence that 푠 ∈ 퐼 − 푋 ′ . As 퐼 − 푋 ′ is the internal part of 푋 − 푋 ′ , we deduce that 퐷 is nontrivial. The end 푡 ′ of 푒 ′ must be in 퐵. Moreover, at least one end of 푒 ′ is in 퐼 ′ . We conclude that 푡 ′ is in 퐵 ∩ 퐼 ′ . ƒ We have assumed that 퐺 − 푒 ′ has a maximal barrier 퐵′ whose correspoding cut 퐶 ′ crosses 퐶. So, we shall adopt the notation introduced in the statement of Lemma 17.18. In particular, 퐷 ′ denotes the cut 휕 ( 푋 ′ − 푋).

365

17.3 The Existence of Thin Edges in Bricks

Case 2.1 The cut 퐷 ′ is trivial. See Figure 17.12.

푋 ′

푢 = 푎1

푎2

푠 = 푎3

푎4

푒′ 푋 푏1

푏3

푡 ′ = 푏2

푏4

푏5

푒 퐶 푦 푥

푎5 푣 푧 퐶 ′ Fig. 17.12 Illustration for Case 2.1, Theorem 17.16.

As 푋 ∩ 푋 ′ is trivial, it follows that | 푋 | = | 푋 ∩ 푋 ′ | + 1. The set 푋 corresponds to the maximal nontrivial barrier 퐵 of 퐺 − 푒. By Proposition 17.10, 푟 (푒) ≤ | 푋 | + 1 = | 푋 ∩ 푋 ′ | + 2.

(17.3)

As 푋 ∩ 푋 ′ is nontrivial, | 푋 ′ | > | 푋 ∩ 푋 ′ | + 1. The set 푋 ′ corresponds to the maximal nontrivial barrier 퐵′ of 퐺 − 푒 ′ . Suppose that the index of 푒 ′ is not equal to two. By Proposition 17.10, 푟 (푒 ′ ) = | 푋 ′ | + 1 > | 푋 ∩ 푋 ′ | + 2. It then follows, from (17.3), that 푟 (푒 ′ ) > 푟 (푒).

We may thus assume that 푒 ′ has index two. Let 퐵′′ be the maximal nontrivial barrier of 퐺 − 푒 ′ distinct from 퐵′ . One of the ends of 푒 ′ is in 퐼 ′ , the other is in 퐼 ′′ . The end 푡 ′ of 푒 ′ is in 퐼 ′ ; hence the end 푠 of 푒 ′ is in 퐼 ′′ . The vertex 푠 is in 퐼; hence it is in 퐼 ∩ 퐼 ′′ . It follows that the neighbours of 푠 in 퐺 − 푒 − 푒 ′ are in 퐵 ∩ 퐵′′ ; hence 퐵′′ ∩ 퐵 is nonempty. By Proposition 17.14, 퐵 is a maximal barrier of 퐺 − 푒 − 푒 ′ . The set 퐵′′ , a barrier of 퐺 − 푒 ′ , is also a barrier of 퐺 − 푒 − 푒 ′ . As 퐵 and 퐵′′ are not disjoint, we conclude that 퐵′′ ⊆ 퐵. The set of isolated vertices of 퐺 − 푒 − 푒 ′ − 퐵 coincides with the set 퐼 of isolated vertices of 퐺 − 푒 − 퐵. Thus, the set of isolated vertices of 퐺 − 푒 − 푒 ′ − 퐵′′ is a subset of 퐼; hence 퐼 ′′ ⊆ 퐼. Consequently, 푋 ′′ ⊆ 푋; hence the second contraction is of the set 푋 ′′ − 푋 ′ , a subset of 푋 − 푋 ′ . We deduce that 푟 0 (푒 ′ ) ≥ | 푋 ∩ 푋 ′ | + 2. As 푒 ′ has index two, we have that

17 Thin Edges in Bricks and Braces

366

푟 (푒 ′ ) > | 푋 ∩ 푋 ′ | + 2.

(17.4)

From (17.4) and (17.3), we conclude that 푟 (푒 ′ ) > 푟 (푒). (See Example 17.19 for an illustration.) Example 17.19 Consider the brick 퐺 depicted in Figure 17.12. The edge 푒 = 푢푣 is a 푏-invariant edge of index two of 퐺 and the sets 퐵1 := {푏 1 , 푏 2 , 푏 3 , 푏 4 , 푏 5 } and 퐵2 := {푎 3 , 푎 4 , 푥, 푦} are the two nontrivial maximal barriers of 퐺 − 푒. The brick of 퐺 − 푒 is obtained by shrinking 푋1 = {푏 1 , 푏 2 , 푏 3 , 푏 4 , 푏 5 , 푎 1 , 푎 2 , 푎 3 , 푎 4 } to a single vertex, and the set 푋2 = {푥, 푦, 푣} to a single vertex. Thus 푒 is of index two, and 푟 (푒) = 5. The edge 푒 ′ = 푎 3 푏 2 incident with the vertex 푠 = 푎 3 in 퐼1 is an edge which is 푏-invariant in both 퐺 −푒 and 퐺. The index of 푒 ′ is two and 퐵′ = {푎 1 , 푎 2 , 푎 5 } is one of the two maximal barriers of 퐺 − 푒 ′ with 퐼 ′ = {푏 1 , 푏 2 } and 푋 ′ = {푎 1 , 푎 2 , 푎 5 , 푏 1 , 푏 2 }. The end 푡 ′ = 푏 2 of 푒 ′ is in 퐵 ∩ 퐼 ′ . The second maximal barrier of 퐺 − 푒 ′ is 퐵′′ = {푏 3 , 푏 4 , 푏 5 } with 퐼 ′′ = {푎 3 , 푎 4 } and 푋 ′′ = {푏 3 , 푏 4 , 푏 5 , 푎 3 , 푎 4 }. The brick of 퐺 − 푒 ′ has two contraction vertices in addition to the four vertices 푥, 푦, 푧 and 푣 of the set 푋 ∩ 푋 ′ . Thus 푟 (푒 ′ ) = 7. Case 2.2 The cut 퐷 ′ is nontrivial. See Figure 17.13. 푋 ′

푢 = 푎1

푎2

푠 = 푎3

푎4

푒′ 푋 푡 ′ = 푏2

푏1

푒

푏3

푏4

푏5 퐶

푎6 푎7

푎5

푣 = 푏6

푞

푝

푥

푏7

푦

퐶 ′ Fig. 17.13 Illustration for Case 2.2, Theorem 17.16. Both 푒 and 푒′ are 푏-invariant edges of index two with equal ranks. It can be verified that if 푀 is any perfect matching of 퐺 − {푎5 , 푎6 }, then {푒, 푒′ } is a subset of 푀. (Thus 푎5 and 푎6 play the roles of 푤1 and 푤2 in the analysis of Case 2.2.)

17.19.1 The edge 푒 has an end, 푣, in the internal part 퐼 ′ − 푋 of 푋 ′ − 푋. Consequently, 푒 has index two.

17.3 The Existence of Thin Edges in Bricks

367

Proof The cut 퐷 ′ − 푒 is a nontrivial tight cut of 퐺 − 푒; hence 푒 has an end, 푣, in the internal part 퐼 ′ − 푋 of 푋 ′ − 푋. This implies that 퐵′ − 푋 is a nontrivial barrier of ƒ 퐺 − 푒 disjoint with 퐵. Consequently, 푒 has index two. All vertices adjacent to 푣 in 퐺 − 푒 are in the external part 퐵′ − 푋 of 푋 ′ − 푋. Since 퐵′ − 푋 is nontrivial, vertex 푣, which is in the internal part of 푋 ′ − 푋, has at least two neighbours, say 푤 1 and 푤 2 , in 퐵′ − 푋. Let 푀 be a perfect matching of 퐺 − {푤 1 , 푤 2 }. Such a matching exists because 퐺 is a brick. A simple counting argument implies the following statement: 17.19.2 {푒, 푒 ′ } ⊂ 푀.

ƒ

17.19.3 There exist at least two edges incident with 푠 which are 푏-invariant in both 퐺 and 퐺 − 푒. Proof By definition of 푠, its degree in 퐺 − 푒 is at least three. The assertion follows from the fact that the index of 푒 is two, by Theorem 17.12. ƒ Let 푒 ′′ ≠ 푒 ′ be an edge incident with 푠 which is 푏-invariant in both 퐺 and 퐺 − 푒. We may now repeat the preceding analysis with 푒 ′′ playing the role of 푒 ′ . If, for every nontrivial barrier 퐵′′ of 퐺 − 푒 ′′ the corresponding set 푋 ′′ is a proper subset of 푋 then Case 1 applies and we would have 푟 (푒 ′′ ) > 푟 (푒). Thus, we may assume that 퐺 − 푒 ′′ has a maximal nontrivial barrier 퐵′′ such that the corresponding set 푋 ′′ is not a proper subset of 푋. Thus, Case 2 applies and the corresponding cut 퐶 ′′ crosses the cut 퐶. If the cut 퐷 ′′ := 휕 ( 푋 ′′ − 푋) is trivial, then Case 2.1 applies and we would again have 푟 (푒 ′′ ) > 푟 (푒). Thus, to conclude that 푟 (푒 ′′ ) > 푟 (푒), it suffices to show that the cut 퐷 ′′ of 퐺 is trivial. Suppose that 퐷 ′′ is not trivial. In this case, the end 푣 of 푒 is in 퐼 ′′ − 푋; hence all the neighbours of 푣 in 퐺 − 푒 are in 푋 ′′ − 푋. In particular, the vertices 푤 1 and 푤 2 are in 푋 ′′ − 푋 (as well as in 푋 ′ − 푋). As above, let 푀 be any perfect matching of 퐺 − {푤 1 , 푤 2 }. The same counting argument which implies that 푒 ′ ∈ 푀 also implies that 푒 ′′ ∈ 푀. But this is absurd because 푒 ′ and 푒 ′′ are distinct edges which are both incident with 푠 and 푀 is a matching. It follows that in the case of edge 푒 ′′ , one of the first two alternatives must occur. The proof of the rank augmentation theorem for bricks is now complete. (See Example 17.20 for an illustration. In that illustration the two vertices 푎 5 and 푎 6 play the roles of 푤 1 and 푤 2 .) ƒ Example 17.20 Let 퐺 be the brick shown in Figure 17.13. The edge 푒 = 푢푣 is a 푏invariant edge of index two of 퐺 and 퐵 = {푏 1 , 푏 2 , 푏 3 , 푏 4 , 푏 5 } is a maximal barrier of 퐺 − 푒, with 퐼 = {푎 1 , 푎 2 , 푎 3 , 푎 4 } and 푋 = {푏 1 , 푏 2 , 푏 3 , 푏 4 , 푏 5 , 푎 1 , 푎 2 , 푎 3 , 푎 4 }. The edge 푒 ′ = 푎 3 푏 2 , incident with 푎 3 ∈ 퐼 is an edge that is 푏-invariant in both 퐺 and 퐺 − 푒, and 퐵′ = {푎 1 , 푎 2 , 푎 5 , 푎 6 , 푎 7 } is a maximal barrier of 퐺 − 푒 ′ , with 퐼 ′ = {푏 1 , 푏 2 , 푏 6 , 푏 7 } and 푋 ′ = {푎 1 , 푎 2 , 푎 5 , 푎 6 , 푎 7 , 푏 1 , 푏 2 , 푏 6 , 푏 7 }.

The cuts 퐶 and 퐶 ′ cross, and 퐵′ − 푋 = {푎 5 , 푎 6 , 푎 7 } is a maximal barrier of (퐺 − 푒)/푋. Thus the index of 푒 is two and it is easy to see that 푟 (푒) = 7. The edge 푒 ′ is also of index two because {푏 3 , 푏 4 , 푏 5 }, which is disjoint from 퐵′ , is a maximal

368

17 Thin Edges in Bricks and Braces

barrier of (퐺 − 푒 ′ )/푋 ′ . Thus 푟 (푒 ′ ) = 푟 (푒) = 7. It is not difficult to see that any perfect matching 푀 of 퐺 − {푎 5 , 푎 6 } must contain both 푒 and 푒 ′ .

There are edges, distinct from 푒 ′ which are also incident with 푎 3 and are 푏invariant in both 퐺 and 퐺 − 푒. For example, 푎 3 푏 4 and 푎 3 푏 5 are two such edges. The graph 퐺 − 푎 3 푏 4 has one maximal nontrivial barrier, namely {푎 2 , 푎 4 }. The bipartite shore {푎 2 , 푏 4 , 푎 4 } of the tight cut associated with this barrier is a subset of 푋. So Case 1 applies, and it can be seen that 푟 (푎 3 푏 4 ) = 16. The graph 퐺 − 푎 3 푏 5 also has just one nontrivial barrier, namely {푎 4 , 푞}. The tight cut associated with this this barrier crosses 퐶 and the readers may verify that here Case 2.1 applies and that 푟 (푎 3 푏 5 ) = 16. Proof (of Theorem 17.3) Let 퐺 be a brick distinct from the three special bricks. By Theorem 15.1, 퐺 has 푏-invariant edges. By Theorem 17.16, a 푏-invariant edge of 퐺 ƒ having maximum rank is thin. In fact, the above proof of the existence of thin edges in bricks may be adapted to actually prove the following: Corollary 17.21 Let 퐺 be a brick, let 푒 be a 푏-invariant edge of 퐺 and let 퐻 be the underlying simple graph of any brick of 퐺 − 푒. If 푒 is not thin then 퐺 has a 푏-invariant edge 푒 ′ such that 푟 (푒 ′ ) > 푟 (푒), 푒 ′ is 푏-invariant in 퐺 − 푒 and 퐻 is the ƒ underlying simple graph of any brick of 퐺 − 푒 − 푒 ′ . Corollary 17.22 Let 퐺 be a brick, let 푒 be a 푏-invariant edge of 퐺 and let 퐽 be a simple brick. If 퐽 is a matching minor of 퐺 − 푒 then 퐺 has a thin edge 푓 such that 퐽 is a matching minor of 퐺 − 푓 . ƒ

Exercises ⊲17.3.1 Prove Corollary 17.21. ⊲17.3.2 Prove Corollary 17.22.

17.4 Building Bricks In this section we give precise definitions of the four operations alluded to in the introduction. We shall refer to these four operations as expansion operations. Given any brick 퐻, any one of these operations may be used to produce a new brick 퐺 from 퐻 such that |퐸 (퐺)| > |퐸 (퐻)|. Three of the four expansion operations involve the device of bisplitting vertices which was introduced in Chapter 12. We restate the definition of bisplitting here and introduce the notation, terminology and properties that we shall be using.

17.4 Building Bricks

369

Bisplitting a vertex (revisited) and constrained bisplitting Let 퐺 be a matching covered graph and let 푥 be a vertex of 퐺. The operation of bisplitting the vertex 푥 consists of first replacing 푥 by two vertices 푏 1 , 푏 2 and distributing the edges in 휕 (푥) between 푏 1 and 푏 2 so that each of these two vertices is incident with at least one edge in 휕 (푥), and then adding a vertex 푎 and joining 푎 to 푏 1 and to 푏 2 . We shall refer to 푏 1 and 푏 2 as the two new outer vertices and 푎 as the new inner vertex. In a special type of bisplitting, called a constrained bisplitting, we require that each outer vertex have at least three distinct neighbours. We denote by 퐺{푥 → (푎, 푏 1 , 푏 2 )} the graph resulting from a constrained bisplitting of vertex 푥 of a graph 퐺, having 푎 as the inner vertex, and 푏 1 and 푏 2 as the two outer vertices. As noted in Chapter 12, it is easy to check that 퐻 := 퐺{푥 → (푎, 푏 1 , 푏 2 )} is matching covered and that 퐺 can be recovered from 퐻 by bicontracting 푎. Furthermore, 푏(퐻) = 푏(퐺). We shall use the following property in the analysis of the four operations. Its proof is left as Exercise 17.4.1. Proposition 17.23 Let 퐺 be a matching covered graph, let 푥 be a vertex of 퐺, let 퐻 := 퐺{푥 → (푎, 푏 1 , 푏 2 )}, let 퐵 denote a maximal barrier of 퐻 and let 푋 := {푎, 푏 1 , 푏 2 }. Then: (i) if 퐵 and 푋 are disjoint then 퐵 is a barrier of 퐺, (ii) if 퐵 and {푏 1 , 푏 2 } have a vertex in common then {푏 1 , 푏 2 } ⊆ 퐵 and the set (퐵 − 푏 1 − 푏 2 ) + 푥 is a barrier of 퐺, and (iii) if 푎 ∈ 퐵 and 퐵 ≠ {푎} then the vertices 푏 1 and 푏 2 are in distinct components of 퐻 − 퐵 and 퐵 − 푎 is a barrier of 퐺. ƒ

17.4.1 The four expansion operations The simplest of the four operations is the one defined below: 1. Edge addition: Let 푢 and 푣 be two distinct vertices of a brick 퐻. Obtain 퐺 from 퐻 by adding a new edge 푒 joining 푢 and 푣. It is easy to see that 퐺 is a brick. The three remaining operations are more complicated and involve the device of constrained bisplitting vertices. Theorem 17.25 shows that an application of any one of these operations on a brick 퐻 results in a brick 퐺 of higher order. 2. Expansion of a vertex by a barrier of size two: Let 퐻 be a brick in which 푥 is a vertex of degree four or more, and let 퐻1 := 퐻{푥 → (푎, 푏 1 , 푏 2 )}.

(17.5)

Now obtain 퐺 from 퐻1 by joining the new inner vertex 푎 to a vertex 푤 of 퐻 − 푥 by an edge labelled 푒 (see Figure 17.14). We allow the possibility of 푤 and 푥 being

17 Thin Edges in Bricks and Braces

370

adjacent in 퐺. We shall refer to the graph 퐺 thus constructed as a graph obtained from 퐻 by an expansion of 푥 by a barrier of size two. 푎 푥

푏1

퐻 − 푥

푏2

퐻 − 푥

푒

푤

퐺 = 퐻1 + 푒

퐻 Fig. 17.14 Expansion of a vertex 푥 by a barrier of size two

3. Expansion of two vertices by barriers of size two: Let 퐻 be a brick in which 푥 and 푥 ′ are two distinct (possibly adjacent) vertices of degree at least four, and let 퐻2′ := 퐻{푥 → (푎, 푏 1 , 푏 2 )}

and

퐻2 := 퐻2′ {푥 ′ → (푎 ′ , 푏 ′1 , 푏 ′2 )}.

(17.6)

We require that each of the vertices 푏 1 , 푏 2 , 푏 ′1 and 푏 ′2 must have at least one neighbour in 푉 (퐻) − 푥 − 푥 ′ .

Now obtain 퐺 from 퐻2 by joining the two vertices 푎 and 푎 ′ by an edge labelled 푒 (see Figure 17.15). We shall refer to the graph 퐺 thus constructed as a graph obtained from 퐻 by expansions of 푥 and 푥 ′ by barriers of size two. 4. Expansion of a vertex by a barrier of size three: Let 퐻 be a brick containing a vertex 푥 of degree at least five, and let 퐻3′ := 퐻{푥 → (푎 1 , 푏 1 , 푡)}

and

퐻3 := 퐻3′ {푡 → (푎 2 , 푏 2 , 푏 3 )}.

(17.7)

Naturally, we require 푡 to be adjacent in 퐻3′ to three or more vertices in 푉 (퐻) − 푥 .

We adjust notation so that 푏 2 is the vertex in {푏 2 , 푏 3 } which is adjacent to 푎 1 . Now obtain 퐺 from 퐻3 by joining 푎 1 and 푎 2 by an edge labelled 푒 (see Figure 17.16). We shall refer to the graph 퐺 thus constructed as a graph obtained from 퐻 by an expansion of 푥 by a barrier of size three. We have already noted that adding any edge to a given brick 퐻 results in a brick 퐺 with one more edge. Now we turn to the task of showing that any application of the three above defined operations of expanding vertices of a given brick 퐻 results in a brick 퐺 of larger order.

17.4 Building Bricks

371 푎

푥 푏1

푏2 푒

퐻 − 푥 − 푥 ′

퐻 − 푥 − 푥 ′

푏1′

푏2′

푥 ′

푎 ′ 퐺 = 퐻2 + 푒

퐻 Fig. 17.15 Expansion of two vertices by barriers of size two

푒 푎1

푎2

푥 푏1

퐻 − 푥

퐻

푏2

푏3

퐻 − 푥

퐺 = 퐻3 + 푒

Fig. 17.16 Expansion of a vertex by a barrier of size three

In the ensuing discussion 퐻1 , 퐻2 and 퐻3 denote, respectively, the graphs defined in equations (17.5), (17.6), and (17.7). Our interest is in showing that, for 푖 = 1, 2, 3, the graph 퐺 = 퐻푖 + 푒 is a brick. Towards this end, it is clearly necessary for us to be able to assert that edge 푒 is matchable in 퐺. By Corollary 2.2, 푒 is matchable in 퐺 if and only if no barrier of 퐻푖 contains both the ends of 푒. With this in view, we proceed to determine the list of all maximal nontrivial barriers of 퐻푖 .

372

17 Thin Edges in Bricks and Braces

Lemma 17.24 The following statements concerning barriers in the three graphs 퐻1 , 퐻2 and 퐻3 hold: (i) the only nontrivial barrier in 퐻1 is {푏 1 , 푏 2 }; (ii) the only two nontrivial barriers in 퐻2 are {푏 1 , 푏 2 } and {푏 ′1 , 푏 ′2 }; and (iii) the only nontrivial maximal barrier in 퐻3 is {푏 1 , 푏 2 , 푏 3 }. Proof Let 퐵 denote a maximal barrier of 퐻푖 , where 푖 ∈ {1, 2, 3}. We consider the three cases separately, but their proofs have many common features. Case 1 The graph under consideration is 퐻1 . Let 푋1 := {푎, 푏 1 , 푏 2 } and let 퐵1 := {푏 1 , 푏 2 }. We consider separately three cases, (i) 퐵 and 푋1 are disjoint, (ii) 퐵 and 퐵1 are not disjoint and (iii) 푎 ∈ 퐵. In each of these three possibilities, we apply Proposition 17.23, with 퐻 playing the role of 퐺 and 퐻1 playing the role of 퐻. Suppose that 퐵 and 푋1 are disjoint. Then, by Proposition 17.23(i), 퐵 is s a barrier of 퐻. As 퐻, a brick, is bicritical, 퐵 is trivial. Suppose that 퐵 contains a vertex in 퐵1 . Then, by Proposition 17.23(ii), 퐵1 ⊆ 퐵 and 퐵′ := (퐵 − 퐵1 ) + 푥 is a barrier of 퐻. Again, as 퐻 is bicritical, we infer that 퐵′ = {푥}; hence 퐵 = {푏 1 , 푏 2 }. Finally, suppose that 푎 ∈ 퐵. Let us prove that 퐵 = {푎}. Suppose that 퐵′′ := 퐵 − 푎 is nonempty. Then, by Proposition 17.23(iii), the vertices 푏 1 and 푏 2 are in distinct components of 퐻1 − 퐵, respectively, 퐽1 and 퐽2 . Moreover, 퐵′′ is a barrier of 퐻. As 퐻 is a brick, then 퐵′′ is a singleton; let 푤 be its only vertex. Thus, 퐵 = {푎, 푤}. The vertex 푏 1 is adjacent to three or more vertices of 퐻1 . Thus, 푏 1 has a neighbour, 푐 1 , in the component 퐽1 . Likewise, 푏 2 has a neighbour, 푐 2 , in component 퐽2 . Consequently, the vertices 푐 1 and 푐 2 are in distinct components of the graph 퐻 − 푤 − 푥. This is not possible, as 퐻, a brick, is 3-connected. We conclude that 퐵1 is the only nontrivial barrier of 퐻1 . Case 2 The graph under consideration is 퐻2 . Let 푋2 := {푎, 푏 1 , 푏 2 }, let 퐵2 := {푏 1 , 푏 2 }, let 푋2′ := {푎 ′ , 푏 1′ , 푏 ′2 } and let 퐵′2 := {푏 ′1 , 푏 ′2 }. By definition of 퐻2′ (equation (17.6)), the outer vertices 푏 1 and 푏 2 have degree three or more in 퐻2′ . By Case 1, 퐵2 is the only nontrivial barrier of 퐻2′ . We consider separately three cases: (i) 퐵 ∩ 푋2 = ∅ or 퐵 ∩ 푋2′ = ∅, (ii) 퐵 ∩ (퐵2 ∪ 퐵′2 ) ≠ ∅, and (iii) {푎, 푎 ′ } ⊆ 퐵. In each of these three possibilities, we apply Proposition 17.23, with 퐻2′ , 푋2′ and 퐻2 playing respectively the role of 퐺, 푋 and 퐻. Suppose that 퐵∩ 푋2 = ∅ or 퐵∩ 푋2′ = ∅. The definition of 퐻2 is symmetric; without loss of generality suppose that 퐵 and 푋2′ are disjoint. By Proposition 17.23(i), 퐵 is a barrier of the graph 퐻2′ . We have seen that 퐵2 is the only nontrivial barrier of 퐻2′ . Thus, either 퐵 is a singleton or 퐵 = 퐵2 . Suppose now that 퐵 and 퐵2 ∪ 퐵2′ are not disjoint. Adjust notation so that 퐵 and 퐵′2 are not disjoint. By Proposition 17.23(ii), 퐵′2 ⊆ 퐵 and the set 퐵′ := (퐵 − 퐵′2 ) + 푥 ′ is a barrier of 퐻2′ . We have seen that 퐵2 is the only nontrivial barrier of 퐻2′ . The vertex 푥 ′ is in 퐵′ but is not in 퐵2 ; hence 퐵′ is trivial. That is, 퐵′ = {푥 ′ }, hence 퐵 = 퐵′2 .

17.4 Building Bricks

373

We may thus assume that 퐵 meets both 푋2 and 푋2′ , but it is disjoint with 퐵2 ∪ 퐵′2 . Thus, {푎, 푎 ′ } ⊆ 퐵. We shall prove that this is not possible. By Proposition 17.23(iii) the set 퐵′′ := 퐵 − 푎 ′ is a barrier of 퐻2′ . We have seen that 퐵2 is the only nontrivial barrier of 퐻2′ . Thus, 퐵′′ = {푎}; hence 퐵 = {푎, 푎 ′ }. Also by Proposition 17.23(iii), the vertices 푏 1′ and 푏 ′2 are in distinct components of 퐻2 − 퐵. For 푘 = 1, 2, let 퐽 푘 be the component of 퐻2 − 퐵 that contains the vertex 푏 ′푘 . By definition of the expansion, the vertex 푏 ′푘 is adjacent in 퐽 푘 to a vertex, 푐 푘 , which is not in {푏 1 , 푏 2 }. Thus, the graph 퐻2 − 푋2 − 푋2′ is not connected. Consequently, the graph 퐻 − 푥 − 푥 ′ is not connected. This is a contradiction, as 퐻, a brick, is 3-connected. We conclude that {푏 1 , 푏 2 } and {푏 1′ , 푏 2′ } are the only two nontrivial barriers of 퐻2 . Case 3 The graph under consideration is 퐻3 . By definition of 퐻3′ (equation (17.7)), the vertices 푏 1 and 푡 have three or more neighbours in 퐻3′ . Thus, Case 1 applies to 퐻3′ ; hence {푏 1 , 푡} is the only nontrivial barrier of 퐻3′ . Let 퐵3 := {푏 1 , 푏 2 , 푏 3 }. In order to prove that 퐵3 is the only maximal nontrivial barrier of 퐻3 , we shall consider three alternatives: (i) 퐵 ∩ 퐵3 ≠ ∅, (ii) either 푎 1 ∉ 퐵 or 푎 2 ∉ 퐵, and (iii) {푎 1 , 푎 2 } ⊆ 퐵. In possibilities (ii) and (iii) we shall apply Proposition 17.23, with 퐻3′ , {푎 2 , 푏 2 , 푏 3 } and 퐻3 playing respectively the role of 퐺, 푋 and 퐻. Suppose that 퐵 and 퐵3 are not disjoint. By the maximality of 퐵 and the Canonical Partition Theorem, 퐵3 ⊆ 퐵. The vertices 푎 1 and 푎 2 are isolated in 퐻3 − 퐵. Thus, 퐵′ := (퐵 − 퐵3 ) + 푥 is a barrier of 퐻. As 퐻 is a brick, 퐵′ = {푥}; hence 퐵 = 퐵3 .

We may thus assume that 퐵 and 퐵3 are disjoint. Consider now the case in which at least one of the vertices 푎 1 and 푎 2 is not in 퐵. By the symmetry of the definition of 퐻3 , we may assume, without loss of generality, that 푎 2 ∉ 퐵. As 퐵 and 퐵3 are disjoint, it follows that 퐵 and {푎 2 , 푏 2 , 푏 3 } are disjoint. By Proposition 17.23(i), 퐵 is a barrier of 퐻3′ . We have seen that {푏 1 , 푡} is the only nontrivial barrier of 퐻3′ . We have assumed that 퐵 and 퐵3 are disjoint; thus the vertex 푏 1 , a vertex of 퐵3 , is not in 퐵. We conclude that 퐵 is trivial in this case. Finally, suppose that {푎 1 , 푎 2 } ⊆ 퐵. We shall prove that this is not possible. By Proposition 17.23(iii), 퐵 − 푎 2 is a barrier of 퐻3′ . The only nontrivial barrier of 퐻3′ is {푏 1 , 푡}. As 푏 1 ∉ 퐵, we have that 퐵 − 푎 2 = {푎 1 }, hence 퐵 = {푎 1 , 푎 2 }. Also, by Proposition 17.23(iii), the vertices 푏 2 and 푏 3 are in distinct components of 퐻3 − 퐵. By symmetry, the vertices 푏 1 and 푏 2 are in distinct components of 퐻3 − 퐵. We deduce that one of the two components of 퐻3 − 퐵 contains the vertices 푏 1 and 푏 3 , whereas the other, say 퐾, contains the vertex 푏 2 . By definition of 퐻3 , the vertex 푏 2 has three or more neighbours in 퐻3 . Thus, 퐾 is nontrivial, hence 푏 2 is a cut vertex of 퐻3 . This is not possible, as 퐻3 is matching covered. We conclude that 퐵3 is the only maximal nontrivial barrier of 퐻3 .

ƒ

Theorem 17.25 Let 퐻 be a brick and let 퐺 be a graph obtained from 퐻 by one of the four operations of expansion. Then 퐺 is a brick.

374

17 Thin Edges in Bricks and Braces

Proof As we have already noted, any graph 퐺 which is obtained from 퐻 by simply adding an edge is a brick. Let us suppose that 퐺 is obtained from 퐻 by applying one of the three remaining expansion operations. The retract of 퐺 − 푒 is the brick 퐻; hence 퐺 − 푒 is a near-brick. By Lemma 17.24, in each case, no end of 푒 belongs to a barrier of 퐺 − 푒. Consequently, 퐺 is matching covered. Furthermore, by definition of 퐺 − 푒 and 퐺, every vertex has at least three distinct neighbours in 퐺. By the monotonicity of function 푏 (see Theorem 9.8), the graph 퐺 is a near-brick. We shall prove that 퐺 is bicritical. The required assertion would then follow because every bicritical near-brick is a brick (see Corollary 4.19). If possible suppose that 퐵 is a maximal nontrivial barrier of 퐺. As 퐺 is a nearbrick, 퐵 is special, by Proposition 17.6. If |퐵| = 2 then, as every vertex of 퐺 is adjacent to three or more vertices, we deduce that 퐵 is not special, a contradiction. We conclude that |퐵| ≥ 3. Since 퐺 −푒 is a spanning matching covered subgraph of 퐺, it follows that 퐵 is also a nontrivial barrier of 퐺 − 푒. By Lemma 17.24, 퐺 must have been obtained from 퐻 by the expansion of a vertex by a barrier of size three and 퐵 = {푏 1 , 푏 2 , 푏 3 }. The graph 퐺 is matching covered, 퐵 is a barrier of 퐺, yet 퐺 − 퐵 has an even component, induced by the pair {푎 1 , 푎 2 }. In all cases considered, the assumption that 퐵 is a maximal nontrivial barrier of 퐺 leads to a contradiction. We conclude that 퐺 is bicritical. Being a bicritical ƒ near-brick, 퐺 is a brick. We remark that the above proof also shows that, in every case, the edge 푒 is a thin edge of the brick 퐺 and that 퐻 is the brick of 퐺 − 푒. As a complement to the above result, we then state the following theorem, which is easily proved. (Exercise 17.4.3). Theorem 17.26 Every brick 퐻 different from the three basic bricks can be obtained from one of them by a sequence of applications of the four operations of expansion. ƒ See Figure 17.17 for an illustration. The graph 퐻 is obtained from the pentagonal prism by adding an edge. It is a plane brick with two odd faces, and hence it is 퐾4 -free. (See Theorem 12.19.) By expanding the two vertices 푥 and 푦 in 퐻 into barriers of size two, we obtain 퐺. (Note that 퐺 is a redrawing of the trellis depicted in Figure 12.10.) Conversely, by deleting the thin edge 푢 1 푢 5 from 퐺 and shrinking {푢 2 , 푢 1 , 푣 1 } to 푥 and {푢 4 , 푢 5 , 푣 5 } to 푦 we recover 퐻. A natural question that arises at this stage is whether all the four operations are really required for building bricks. This is indeed the case. (Exercise 17.4.7.)

(풃 + 풑)-invariant thin edges Let 퐺 be a brick that is not a Petersen brick. If 퐺 can be obtained from a Petersen brick by means of an expansion operation, then it can also be obtained from a brick that is not a Petersen brick by some expansion operation. This amounts to saying that if a brick 퐺 is not a Petersen brick and has a thin edge 푒 such that 퐺š − 푒 is a Petersen

17.4 Building Bricks

375

푣3

푣3

푢3 푣2

푣4 푥

푦

푢3

푣1

푣2

푣5

푢1

푢5

푢2 푠1

푠2

푡1

푢4 푠1

푡2

푠2 푡2

푡1

퐻

푣4

퐺

Fig. 17.17 Obtaining the trellis 퐺 from the graph 퐻 by expanding 푥 and 푦 into barriers of size two.

› brick, then it also has a thin edge 푓 such 퐺 − 푓 is not a Petersen brick. (Some cases of this assertion are illustrated in Figure 17.5.) Theorem 17.27 Every brick 퐺 that is not a Petersen brick and is distinct from 퐾4 ƒ and 퐶6 has a thin edge 푒 such that the brick of 퐺 − 푒 is not a Petersen brick. This theorem can be verified directly by straightforward case analysis similar to the one used in the proof of Corollary 15.19. We leave it to the readers to work out the details (Exercise 17.4.6). —————≀≀————— In view of Theorem 17.27, every brick that is not a Petersen brick may be obtained from 퐾4 and 퐶6 by a sequence of applications of the four operations of expansion. This is the sense in which any brick may be generated from one of the two bricks mentioned above, or, conversely, any brick may be reduced to one of them by deletions of thin edges and bicontractions! CLM (2006, [15]) used Theorem 17.27 as an inductive tool to show that the only simple solid planar bricks are the wheels. A simpler proof of this result was established by Kothari and Murty (2015, [50]) using the idea of ‘strictly thin edges’ which will be introduced in the next chapter.

Exercises ⊲17.4.1 Prove Proposition 17.23. Hint: in the proof of the last part of the assertion, if 푏 1 and 푏 2 are in the same component, 퐾, of 퐻 − 퐵, then the subgraph of 퐺 induced by 푉 (퐾) + 푎 is an even component of 퐺 − (퐵 − 푎).

17 Thin Edges in Bricks and Braces

376

⊲17.4.2 In the definition of the operation of expansion of two vertices by barriers of size two, there is a requirement that “each of 푏 1 , 푏 2 , 푏 ′1 and 푏 ′2 must have at least three neighbours, and at least one neighbour in 푉 (퐻 − 푥 − 푥 ′ ).” Explain why this requirement is necessary for the resulting graph 퐺 to be a brick. 17.4.3 Prove Theorem 17.26. Hint: use induction on |퐸 |. 17.4.4 Generate all the cubic bricks on four, six, and eight vertices.

푥

푥

퐺1

퐺2

푥 ′

푥

푥 ′

푥 퐺3

퐺4

Fig. 17.18 Expansions of four Petersen bricks–Exercise 17.4.5.

⊲17.4.5 Consider the four Petersen bricks shown in Figure 17.18. (i) In each of the following cases, find as many bricks as you can, up to isomorphism, that can be obtained by means of the indicated expansion operations: (a) expansion of vertex 푥 in 퐺 1 by a barrier of size two (four bricks); (b) expansion of vertex 푥 in 퐺 2 by a barrier of size three (two bricks); (c) expansion of vertices 푥 and 푥 ′ in 퐺 3 and 퐺 4 by barriers of size two (respectively, one and two bricks).

17.5 Thin Edges in Braces

377

(ii) For each brick 퐺 obtained by an expansion operation as in the first part, show that 퐺 has a thin edge 푓 such that the retract of 퐺 − 푓 is a brick that is not a Petersen brick. (See Exercise 17.1.8.) ⊲17.4.6 Prove Theorem 17.27. ⊲17.4.7 Verify that all the four operations are necessary for the generation of all bricks starting with the three basic bricks. For each of the four operations we give below a hint: (i) edge addition: consider any brick having minimum degree four or more; (ii) expansion of a vertex by a barrier of size two: consider the odd wheels of order six or more; (iii) expansion of a vertex by a barrier of size three: consider the tricorn H10 , shown in Figure 17.1(c); (iv) expansion of two vertices by barriers of size two: consider triangle-free cubic bricks distinct from the Petersen graph.

17.5 Thin Edges in Braces

Thin edges in braces and their indices The index of a removable edge 푒 in a brace 퐺 is the number of ends of 푒 having degree two in 퐺 − 푒. If 퐺 has order at least six, it is easy to verify that the index of 푒 is the number of contraction vertices in the retract 퐺š − 푒 of 퐺 − 푒.

A removable edge 푒 in a brace 퐺 is thin if the retract of 퐺 − 푒 is also a brace. Any removable edge in a brace of order two or four is a multiple edge and is clearly thin. Every edge in a brace of order six or more is removable (Theorem 8.3), but not all edges are necessarily thin. For example, in the brace 퐺 of order 12 shown in Figure 17.19, edge 푓 is one of the ten edges that are not thin; the edges 푒 0 , 푒 1 and 푒 2 in this brace are three of the nine edges which are thin, and their indices are zero, one and two, respectively.

In this section we develop a theory pertaining to thin edges in braces similar to the one related to thin edges in bricks described in the previous sections. The principal result (Theorem 17.34) we establish states that every brace distinct from 퐾2 and 퐶4 has at least two thin edges. A consequence of this assertion is that any brace may be obtained from either 퐾2 or 퐶4 by sequentially applying suitable expansion operations much the same way any brick may be obtained from one of 퐾4 , 퐶6 and P. Most of the ideas needed for establishing the existence of thin edges in braces are similar to the ones that arose in connection with the proof the existence of thin edges in bricks. One crucial difference arises in connection with the selection of shores of tight cuts. For example, if 푒 is a 푏-invariant edge of a brick 퐺, the graph 퐺 − 푒 is a near-brick, and one of the shores of a nontrivial tight cut 퐶 of 퐺 − 푒 is bipartite and

17 Thin Edges in Bricks and Braces

378

PSfrag 푒1

푒0

푒2

푓

Fig. 17.19 Examples of thin and nonthin edges in a brace: edges 푒0 , 푒1 and 푒2 are thin; edge 푓 is not thin.

the other is nonbipartite; hence the brick of 퐺 − 푒 may be obtained by a tight cut decomposition that starts by contracting the bipartite shore of 퐶. However, when 푒 is a removable edge of a brace 퐺, the graph 퐺 − 푒 is a bipartite matching covered graph, and both shores of any nontrivial tight cut 퐶 of 퐺 − 푒 are bipartite, and it is not clear which 퐶-contraction of 퐺 − 푒 yields the largest brace of 퐺 − 푒. Another important distinction is in the structure of the bipartite matching covered graphs obtained from a brace by the deletion of an edge. We shall address these differences first before turning to the proof of the existence of thin edges in braces.

17.5.1 Graphs obtained by the deletion of an edge from a brace Let 퐺 be a brace on six or more vertices, and let 푒 := 푢푣 be an edge of 퐺. Then, by Theorem 8.3, 퐺 − 푒 is matching covered. We shall now establish the structure of a tight cut decomposition of 퐺 − 푒. Tight cut decompositions of 푮 − 풆 Suppose that 퐺 − 푒 is not a brace. Then, the ends 푢 and 푣 of 푒 belong to different shores of any nontrivial tight cut of 퐺 − 푒. Let C be the (laminar) family of tight cuts used by a tight cut decomposition of 퐺 − 푒. As indicated in Exercise 8.1.10, the shores of the cuts in C containing 푢 form a nested family 푋1 ⊂ 푋2 ⊂ · · · ⊂ 푋 푘 of proper subsets of 푉 (see Figure 17.20), each containing 푢, such that

17.5 Thin Edges in Braces

379

(퐺 − 푒)/ 푋1 is a brace, (퐺 − 푒)/푋푖 /푋푖+1 is a brace for 2 ≤ 푖 < 푘, (퐺 − 푒)/푋 푘 is a brace. Suppose that 퐺 −푒 is not a brace. Then, clearly the set 푋1 is the minimal nontrivial subset of 푉 such that 푢 ∈ 푋1 and 휕 ( 푋1) is a tight cut of 퐺. Hence 푋1 is unique. For a similar reason 푋 푘 is also unique. It follows that if 퐺 − 푒 has at most three braces, then it has a unique tight cut decomposition. However, in general, 퐺 − 푒 may have more that one tight cut decomposition, but all of them conform to the above pattern. See Exercise 17.5.3.

푢 푋1

푋2

푋3

푒 푣

Fig. 17.20 The nested structure of shores of tight cuts of 퐺 − 푒.

Removable edges in 푮 − 풆 The following lemma plays a role in the proof of the thin edge theorem for braces similar to the one that Lemma 8.19 played in the proof of the thin edge theorem for bricks. Lemma 17.28 Let 퐺 be brace of order at least six and let 푒 be an edge of 퐺. Then (see Exercise 8.4.1): (i) for any vertex 푣 that has degree at least three in 퐺 − 푒, at most one edge of 퐺 − 푒 incident with 푣 is not removable in 퐺 − 푒. (ii) if 퐶 := 휕 ( 푋) − 푒 is a nontrivial tight cut of 퐺 − 푒 then each edge of 퐶 is removable in 퐺 − 푒. ƒ Maximal shores of tight cuts of 푮 − 풆 Let 퐻 be the underlying simple graph of a brace of 퐺 − 푒. Let 푋 be a maximal set of vertices of 퐺 such that 퐶 := 휕 ( 푋) is nontrivial, 퐶 − 푒 is tight in 퐺 − 푒 and 퐻 is the underlying simple graph of a brace of 퐺 ′ − 푒, where 퐺 ′ := 퐺/푋.

17 Thin Edges in Bricks and Braces

380

Lemma 17.29 If 퐺 ′ − 푒 is not a brace then 푋 has a unique (proper) subset 푌 such that 퐻 is the underlying simple graph of 퐺 ′′ − 푒, where 퐺 ′′ := 퐺 ′ /푌 . Proof Suppose that 퐺 ′ − 푒 is not a brace. Consider any tight cut decomposition of 퐺 ′ − 푒. The maximality of 푋 implies the existence of a nontrivial cut 퐷 := 휕 (푌 ), where 푌 is a proper subset of 푋 and 퐺 ′′ := 퐺 ′ /푌 , such that 퐻 is the underlying simple graph of 퐺 ′′ − 푒. Our task is thus to prove that 푌 is unique. For this, we shall prove that every tight cut decomposition of 퐺 ′ − 푒 uses cut 퐷 − 푒. Note that as 퐻 is a brace, then 퐺 ′′ − 푒 is also a brace. In fact, let 푌 ′ be a subset of 푋, let 퐷 ′ := 휕 (푌 ′ ) be nontrivial and suppose that ′ 퐷 − 푒 is tight in 퐺 − 푒. The cut 퐷 ′ − 푒 is tight in 퐺 − 푒; therefore 푌 ′ contains an end of 푒, say 푣. As 퐶 − 푒 is tight in 퐺 − 푒, the vertex 푣 is the only end of 푒 in 푋. As 퐷 − 푒 is tight in 퐺 − 푒, it follows that 푣 is also a vertex of 푌 . We first prove that 퐷 ′ and 퐷 do not cross. For this, assume the contrary. Let 퐼 := 휕 (푌 ∩ 푌 ′ ) and 푈 := 휕 (푌 ∩ 푌 ′ ). The cuts 퐼 − 푒 and 푈 − 푒 are both tight in 퐺 − 푒 (Exercise 8.1.10). The cut 푈 − 푒 is a cut of 퐺 ′′ − 푒, in turn a brace. Thus, 푈 = 퐶; hence 푌 ∩ 푌 ′ = 푋. See Figure 17.21. 퐶 퐷

퐷 ′

푋

푒

푌 ′

푣

푌

Fig. 17.21 The cuts 퐶, 퐷 and 퐷 ′ .

It follows that 퐺/( 푋 ∪ (푌 − 푌 ′ ))/(푌 ∩ 푌 ′ ) is isomorphic, up to multiple edges, to 퐺 , in contradiction to the maximality of 푋. We deduce that 퐷 and 퐷 ′ do not cross. ′′

As 푣 ∈ 푌 ∩ 푌 ′ , it follows that one of 푌 and 푌 ′ is a subset of the other. The set 푌 is not a proper subset of 푌 ′ ; else 퐷 ′ − 푒 would be a nontrivial tight cut of 퐺 ′′ − 푒, in turn a brace, a contradiction. The set 푌 ′ is then a subset of 푌 . This conclusion holds for each subset 푌 ′ of 푋 such that 휕 (푌 ′ ) − 푒 is a nontrivial tight cut of 퐺 − 푒. It follows that every tight cut decomposition of 퐺 ′ − 푒 uses the cut 푌 . Indeed, 푌 is unique. ƒ

17.5 Thin Edges in Braces

381

Exercises 17.5.1 Find an edge 푒 in the brace 퐺 shown in Figure 17.22 that is not thin and display a tight cut decomposition of 퐺 − 푒.

Fig. 17.22 Brace 퐺 for Exercise 17.5.1.

17.5.2 Consider the brace 퐺 shown in Figure 17.30. The two edges 푒 = 푎 1 푏 5 and 푒 ′ = 푎 3 푏 2 happen to have the property that both 퐺 − 푒 and 퐺 − 푒 ′ are matching covered graphs with three braces. Find the unique tight cut decompositions of the two bipartite graphs 퐺 − 푒 and 퐺 − 푒 ′ . 17.5.3 Find two different tight cut decompositions of the matching covered graph obtained by deleting the edge 푢 1 푣 6 from the brace 퐺 shown in Figure 17.23.

푢1

푣1

푣2

푢2

푢3

푢4

푣3

푣4

푣5

푢5

푢6

푣6

Fig. 17.23 퐺 − 푢1 푣6 has two different tight cut decompositions.

17.5.4 Consider the Heawood graph with the labelling of its vertices as shown in Figure 17.24, and let 푒 := 푎 0 푏 0 . Show that the retract of the graph obtained by deleting edge 푒 from the Heawood graph is isomorphic to the biwheel B10 . (Since the Heawood graph is edge transitive, it follows that all its edges are thin.)

17 Thin Edges in Bricks and Braces

382 푎5 replacements 푎6

푎1 푏1

푎3

푏0

푏3

푎0

푎2 푏2

푎4 푏4

푏6

푏5 Fig. 17.24 Figure for Exercise 17.5.4.

17.6 The Existence of Thin Edges in Braces ♯ The rank of an edge Our proof technique is constructive and is based on the notion of rank of an edge. Given an edge 푒 of 퐺 which is not thin, we shall show that there exists an edge of higher rank than 푒. This leads us to the conclusion that an edge of 퐺 of maximum rank is a thin edge with the desired property. We used a similar technique for showing that every brick different from 퐾4 , 퐶6 , and the Petersen graph has a thin edge. Our objective is to define the notion of the rank of an edge and then show that an edge of maximum rank is a thin edge. But, first, we introduce a closely related function which we shall refer to as ‘pre-rank’. The pre-rank of 푒, denoted by 푟 0 (푒), is defined to be the maximum of the orders of all braces of 퐺 − 푒. As an example, consider the brace 퐺 shown in Figure 17.25. The graph 퐺 − 푎푤 has two braces, one being the cube and the other being 퐾3,3 . Thus, 푟 0 (푎푤) = 8. The graph 퐺 − 푎푏 has two braces: one has four vertices and the other has ten vertices. Thus 푟 0 (푎푏) = 10. In the brace shown in Figure 17.25, the edge 푎푏 has the largest possible pre-rank and is a thin edge. However, this is not in general true. It turns out that, in defining the rank of an edge, in addition to considering the number of vertices in a largest brace of 퐺 − 푒, we need also to consider the number of contraction vertices that brace has. To illustrate this point, consider the brace 퐺 in Figure 17.26. Both graphs 퐺 − 푒 and 퐺 − 푓 have 퐾3,3 as a brace, up to multiple edges. The graph 퐺 − 푒 has two braces isomorphic to 퐾3,3 , up to multiple edges (consider the tight cut 퐶 − 푒). Edge 푓 is thin and the graph 퐺 − 푓 has only one brace isomorphic to 퐾3,3 , up to multiple edges; it has two other braces, both of order four. Thus, both 푒 and 푓 have pre-rank six. But edge 푒 is not thin, whereas 푓 is thin. Note that the braces of 퐺 − 푒 have only one contraction vertex, whereas the brace of 퐺 − 푓 on six

383

17.6 The Existence of Thin Edges in Braces

푗

푧

푤

푖

푦 푎

푑

푘

푐

푥

푏

푡

Fig. 17.25 The pre-rank of an edge, example: 푟0 (푎푤) = 8, 푟0 (푎푏) = 10.

푓

퐶 푒

퐷 Fig. 17.26 푟0 ( 푓 ) = 푟0 (푒) = 6, but 푟 ( 푓 ) = 7 and 푟 (푒) = 6.

vertices has two contraction vertices. So, in defining the rank of an edge 푒, we take into account the number of contraction vertices in a largest brace of 퐺 − 푒, giving preference to braces having two contraction vertices. We thus define the rank 푟 (푒) of an edge 푒 removable in 퐺 to be:    푟 0 (푒) + 1, if 퐺 − 푒 has a brace of order 푟 0 (푒)  with two contraction vertices 푟 (푒) :=   푟 0 (푒), otherwise. 

17 Thin Edges in Bricks and Braces

384

We remark that the same edge 푒 may provide different choices for the brace of 퐺 − 푒. For example, in Figure 17.27, 퐺 − 푒 has two braces isomorphic to 퐾3,3 up to multiple edges. One brace contains the vertices in {1, 2, 7, 8, 9}, plus a contraction vertex. The other brace contains vertices in {3, 4, 푎, 푏}, plus two contraction vertices. Thus, the latter is responsible for the value seven for 푟 (푒). 푢1

푢2

푢3

푢4

푢5

푢6

푣3

푣4

푣5

푣6

푒

푣1

푣2

Fig. 17.27 The graph 퐺 − 푒 has two braces of order six

Theorem 17.30 (The Rank Augmentation Theorem for Braces) Let 푒 := 푢푣 be an edge in a simple brace 퐺 := [ 퐴, 퐵] of order six or more and suppose that 푒 is not thin. Then there exist two edges 푒 ′ and 푒 ′′ in 퐺 such that: (i) 푒 ′ and 푒 ′′ are adjacent to each other, but not to 푒; (ii) 푟 (푒 ′ ) ≥ 푟 (푒), 푟 (푒 ′′ ) ≥ 푟 (푒); and (iii) either 푟 (푒 ′ ) > 푟 (푒) or 푟 (푒 ′′ ) > 푟 (푒). Proof By the definition of 푟 (푒), there is a brace of 퐺 − 푒 of order 푟 0 (푒). As 푒 is not thin by hypothesis, 퐺 − 푒 has at least two braces, and 푟 (푒) < |푉 |. For every tight cut decomposition of 퐺 − 푒, every brace has at most two contraction vertices. Consider all the tight cut decompositions of 퐺 − 푒. Let G★ denote the set of those braces 퐺 ★ of 퐺 − 푒 that satisfy the following properties: (i) 퐺 ★ has order 푟 0 (푒); and (ii) if 푟 (푒) is odd then 퐺 ★ has two contraction vertices.

We note that 푟 0 (푒) is at least four (all braces of any bipartite matching covered graph of order at least four have order at least four). Let G be a tight cut decomposition of 퐺 − 푒 that contains braces in G★. As 푒 is not thin, every brace in G ∗ has a contraction vertex that is the result of the contraction of a set having five or more vertices. In particular, every brace in G that lies in G★ has a contraction vertex that is the result of the contraction of a set containing five or more vertices. Let X be the set of the shores 푋 of tight cuts used by G such that | 푋 | ≥ 5 and (퐺 − 푒)/푋 contains a brace in G★. Let 푋 be a maximal set in X, and let 퐶 := 휕 ( 푋). The maximality of 푋 implies that the graph 퐻 := (퐺 − 푒)/( 푋 → 푥) contains one brace in G ∗ and 푥 is a contraction vertex of that brace.

385

17.6 The Existence of Thin Edges in Braces

We now consider two possibilities, depending on the parity of 푟 (푒). If 푟 (푒) is even then 퐻 = 퐺 ∗ and 푟 (푒) = 푟 0 (푒) = 1 + | 푋 |. Alternatively, if 푟 (푒) is odd then 푋 has a proper subset 푌 such that (i) the cut 휕 (푌 ) − 푒 is nontrivial and tight in 퐺 − 푒, (ii) the graph 퐺 ★ := 퐻/푌 is a brace in G ∗ and (iii) no brace of (퐺 − 푒)/푌 is in G ∗ ; moreover, 푟 (푒) = 1 + 푟 0 (푒) = 1 + (2 + | 푋 − 푌 |) ≤ | 푋 |, where the last inequality follows from the fact that |푌 | ≥ 3. In sum, 푟 (푒) ≤ 1 + | 푋 |,

(17.8)

with equality only if 퐺 ★ has precisely one contraction vertex. Note that edge 푒 has its ends in 푋− and 푋 − . Let 푢 denote the end of 푒 in 푋− and let 푣 be its other end. Recall that ( 퐴, 퐵) denotes the bipartition of 퐺. Adjust notation so that 푢 lies in 퐴. (See Figure 17.28.) 퐶 푠

푋−

푒′

푢

푋 +

퐴

푋 −

퐵

푒′′ 푒

푡 ′

푋+

푡 ′′

푣

Fig. 17.28 Edges 푒, 푒′ and 푒′′ .

As | 푋 | ≥ 5, it follows that 푋− contains two or more vertices. Let 푠 ∈ 푋− be a vertex distinct from 푢. Then, 푠 has degree at least three in 퐺 − 푒. By Lemma 17.28, there are edges 푒 ′ and 푒 ′′ , both incident with 푠 and removable in 퐺 −푒. Then 퐺 −푒−푒 ′ and 퐺 − 푒 − 푒 ′′ are matching covered subgraphs of 퐺 − 푒. Let 푡 ′ and 푡 ′′ denote the ends of 푒 ′ and 푒 ′′ in 퐵. The vertices of 퐺 adjacent to 푠 in 퐺 − 푒 lie all in 푋+ . Thus, both 푡 ′ and 푡 ′′ lie in 푋+ (Figure 17.28). This establishes that 푒 ′ and 푒 ′′ are adjacent to each other, but not to 푒. Since 퐺 is a brace, 퐺 has a perfect matching that contains edges 푒 and 푒 ′′ , but not 푒 ′ , hence 퐺 − 푒 ′ is matching covered, by Proposition 17.11. Likewise, 퐺 − 푒 ′′ is matching covered. We shall now prove that 푟 (푒 ′ ) ≥ 푟 (푒), and characterize under what conditions equality holds. For this, we consider some cases. In all cases except the last, we show that 푟 (푒 ′ ) > 푟 (푒). In the last case we conclude that 푟 (푒 ′ ) = 푟 (푒). A similar reasoning holds for 푒 ′′ . We then finally show that equalities 푟 (푒 ′ ) = 푟 (푒) and 푟 (푒 ′′ ) = 푟 (푒) cannot both hold. This part of the proof is remarkably similar to the proof of the Rank Augmentation Theorem for Bricks (Theorem 17.16), but there are some differences, especially due to the fact that in a near-brick every tight cut has precisely one bipartite shore, whereas in a bipartite graph clearly both shores of a tight cut are bipartite.

386

17 Thin Edges in Bricks and Braces

The next statement is similar to a property established in the proof of Theorem 17.16 and its proof is left as Exercise 17.6.1. 17.30.1 Let 퐶 ′ be a cut of 퐺 such that 퐶 ′ − 푒 ′ is tight in 퐺 − 푒 ′ . If 퐶 and 퐶 ′ do not cross then 푋 is a proper subset of a shore of 퐶 ′ . ƒ Case 1 No tight cut in 퐺 − 푒 ′ crosses 퐶. In this case, by statement 17.30.1, the set 푋 is properly contained in a shore of every tight cut in 퐺−푒 ′ . Then 푋 is properly contained in the vertex set of a brace 퐻 of 퐺−푒 ′ . It follows that 퐻 has at least | 푋 | + 3 vertices. Thus, 푟 (푒 ′ ) ≥ |푉 (퐻)| ≥ | 푋 | + 3 > 푟 (푒), establishing that 푟 (푒 ′ ) > 푟 (푒). Case 2 퐺 − 푒 ′ has tight cuts that cross 퐶. The proof of the next result is somewhat similar to the proof of Lemma 17.18 (Exercise 17.6.2). Lemma 17.31 Let 퐶 ′ := 휕 ( 푋 ′) be a cut of 퐺 such that 퐶 ′ − 푒 ′ is tight in 퐺 − 푒 ′ . Suppose that 퐶 and 퐶 ′ cross and also that the end 푡 ′ of 푒 ′ is in 푋 ∩ 푋 ′ . Let 퐷 := 휕 ( 푋 − 푋 ′ ) and let 퐷 ′ := 휕 ( 푋 ′ − 푋). Then, the following properties hold (see Figure 17.29): (i) the end 푠 of edge 푒 ′ is in 푋 − 푋 ′ , (ii) 푋+ ⊂ 퐵 and 푋+′ ⊂ 퐴, (iii) the cut 퐷 ′ − 푒 is tight in 퐺 − 푒, and (iv) the cut 퐷 is nontrivial and 퐷 − 푒 − 푒 ′ is tight in 퐺 − 푒 − 푒 ′ .

ƒ

Let 푋 ′ be a maximal subset of 푉 such that 푡 ′ ∈ 푋 ′ , 퐶 ′ := 휕 ( 푋 ′ ) is nontrivial, − 푒 ′ is tight in 퐺 − 푒 ′ and the cuts 퐶 and 퐶 ′ cross. By statement 17.30.1, the maximality of 푋 ′ implies that every nontrivial tight cut of (퐺 − 푒 ′ )/푋 ′ has a shore that includes the set 푋 ∩ 푋 ′ .

퐶 ′

Case 2.1 Cut 퐷 ′ is trivial. Recall that 푟 (푒) ≤ 1 + | 푋 | = 2 + | 푋 ∩ 푋 ′ |. Every nontrivial tight cut of (퐺 − 푒 ′ )/푋 ′ has a shore that includes the set 푋 ∩ 푋 ′ , hence 푟 0 (푒 ′ ) ≥ 2 + | 푋 ∩ 푋 ′ | ≥ 푟 (푒). If 퐷 − 푒 ′ is not tight then 푟 0 (푒 ′ ) > 2 + | 푋 ∩ 푋 ′ | ≥ 푟 (푒). Alternatively, if 퐷 − 푒 ′ is tight then 푟 0 (푒 ′ ) = 2 + | 푋 ∩ 푋 ′ | and the corresponding brick of 퐺 − 푒 ′ has two contraction vertices, hence 푟 (푒 ′ ) > 2 + | 푋 ∩ 푋 ′ | ≥ 푟 (푒). In both alternatives, 푟 (푒 ′ ) > 푟 (푒). (See Example 17.32 for an illustration.) Example 17.32 (Case 2.1, Theorem 17.30) Consider the brace 퐺 depicted in Figure 17.29. The edge 푒 = 푎 1 푏 5 is a removable edge of 퐺, the cut 퐶 := 휕 ( 푋), where 푋 := {푎 1 , 푎 2 , 푎 3 , 푏 1 , 푏 2 , 푏 3 , 푏 4 } is a tight cut of 퐺 − 푒. Both 퐶-contractions of 퐺 − 푒 happen to be braces of order eight, each with just one contraction vertex. Thus 푟 (푒) = 8. The edge 푒 ′ = 푎 3 푏 2 is removable in both 퐺 − 푒 and 퐺. The cut 퐶 ′ := 휕 ( 푋 ′), where 푋 ′ := {푎 1 .푎 2 , 푎 4 , 푏 1 , 푏 2 }, is a tight cut of 퐺 − 푒 ′ . The two cuts 퐶 and 퐶 ′ cross, and | 푋 ∩ 푋 ′ | = 1. The graph 퐺 − 푒 ′ has three braces of orders six, eight and four. The largest brace of 퐺 − 푒 ′ , which is of order eight, has two contraction vertices. Therefore, 푟 (푒 ′ ) = 9.

387

17.6 The Existence of Thin Edges in Braces 푋 ′

푎1

푎2

푠 = 푎3 푒′

푋 푏1

푒

푡 ′ = 푏2

푎4

푏3

푎5

푏5

푏4

푎6

푏6

푎7

푏7

Fig. 17.29 Illustration for Case 2.1, Theorem 17.30

Case 2.2 Cut 퐷 ′ is nontrivial. As 퐷 ′ − 푒 is tight in 퐺 − 푒, the edge 푒 has an end in the minority part of the shore 푋 ′ − 푋 of 퐷 ′ . The minority part of 푋 ′ − 푋 is a subset of the part 퐵 of the bipartition of 퐺. Let 푣 be the end of 푒 in 푋 ′ − 푋. The other end 푢 of 푒 is in 퐴 and also in the minority part of 푋. As 퐶 ′ − 푒 ′ is tight in 퐺 − 푒 ′ , we deduce that 푢 is in 푋 ∩ 푋 ′ . See Figure 17.30. As every tight cut of (퐺 − 푒 ′ )/푋 ′ has a shore that includes 푋 ∩ 푋 ′ , the graph 퐻 := (퐺 −푒 ′ )/푋 ′ /( 푋 − 푋 ′ ) is a brace. This brace 퐻 is isomorphic to 퐺/푋/( 푋 ′ − 푋), up to multiple edges. The cut 퐷 ′ − 푒 is a nontrivial cut of (퐺 − 푒)/푋; therefore 퐺 ∗ has two contraction vertices. We deduce that 푟 (푒) = 푟 (푒 ′ ) = | 푋 ∩ 푋 ′ | + 3. Let 푌 := 푋 ′ − 푋. Note that 푌 is a proper subset of 푋, such that (퐺 − 푒)/푋/푌 is 퐺 ∗ . By Lemma 17.29, 푌 is unique. A similar reasoning may then be applied to 푒 ′′ . The only case in which the inequality 푟 (푒 ′′ ) ≥ 푟 (푒) holds with equality occurs when 퐺 has a cut 퐶 ′′ := 휕 ( 푋 ′′ ) that crosses 퐶, 퐶 ′′ − 푒 ′′ is tight and nontrivial in 퐺 − 푒 ′′ and 푌 = 푋 ′′ − 푋. Thus, 푋 ′ − 푋 = 푌 = 푋 ′′ − 푋; therefore there exist two vertices in 푌 ∩ 퐴, say 푎 ′ and ′′ 푎 , both of which are adjacent to the end 푣 of 푒. The vertex 푠 has a neighbour, 푏 ′ , distinct from both 푡 ′ and 푡 ′′ . The vertex 푏 ′ is in 퐵 − ( 푋 ′ ∪ 푋 ′′ ). Finally, in addition to its two contraction vertices, the brace 퐺 ∗ has at least two vertices in the equipartite set 푋 ∩ 푋 ′ , one of which, say 푏 ′′ , is in 퐵. In fact, 푏 ′′ ∈ 퐵 − ( 푋 ′ ∪ 푋 ′′ ). In sum, 푎 ′ and 푎 ′′ are two vertices in 퐴 ∩ ( 푋 ′ ∩ 푋 ′′ ) and the vertices 푏 ′ and 푏 ′′ are two vertices in 퐵 − ( 푋 ′ ∪ 푋 ′′ ). The graph 퐺 − 푎 ′ − 푎 ′′ − 푏 ′ − 푏 ′′ has a perfect matching, 푀. It is necessary for 푀 to contain both 푒 ′ and 푒 ′′ , which is a contradiction, as 푒 ′ and 푒 ′′ are adjacent. The proof of the Rank Augmentation Theorem for Braces is complete. ƒ

17 Thin Edges in Bricks and Braces

388 푋 ′

푢 = 푎1

푎2

푠 = 푎3 푒′ 푒′′

푋 푏1

푒

푡 ′ = 푏2

푎4

푎5

푎6

푣 = 푏5

푏6

푒′′′

푡 ′′ = 푏3

푡 ′′′ = 푏4

푎7

푎8

푏7

푏8

Fig. 17.30 Illustration for Case 2.2, Theorem 17.30

Example 17.33 (Case 2.2, Theorem 17.30) Consider the brace 퐺 depicted in Figure 17.30 with the edge 푒 = 푎 1 푏 5 in it. The cut 퐶 := 휕 ( 푋), where 푋 := {푎 1 , 푎 2 , 푎 3 , 푏 1 , 푏 2 , 푏 3 , 푏 4 } is a tight cut of 퐺 − 푒. The graph obtained from 퐺 − 푒 by contracting the set 푋 to a single vertex 푥, and the set {푎 4 , 푎 5 , 푏 5 } to a single vertex 푦 is a brace of order eight of 퐺 − 푒 with two contraction vertices. It follows that 푟 (푒) = 9. (Note that the graph obtained from 퐺 − 푒 by contracting 푋 is also a brace of order eight, but that brace has only one contraction vertex.) The edge 푒 ′ = 푎 3 푏 2 is removable in both 퐺 − 푒 and 퐺. The cut 퐶 ′ := 휕 ( 푋 ′), where 푋 ′ := {푎 1 .푎 2 , 푎 4 , 푎 5 , 푏 1 , 푏 2 , 푏 5 }, is a tight cut of 퐺 − 푒 ′ . The two cuts 퐶 and 퐶 ′ cross, and | 푋 ∩ 푋 ′ | = 3. The graph obtained from 퐺 − 푒 ′ by contracting 푋 ′ to a single vertex 푥 ′ , and the set {푎 3 , 푏 3 , 푏 4 } to a single vertex 푦 ′ is a brace of order eight with two contraction vertices. Thus the rank 푟 (푒 ′ ) of 푒 ′ is also 9. However, it can be verified that both the other edges 푒 ′′ = 푎 3 푏 3 and 푒 ′′′ = 푎 3 푏 4 incident with the vertex 푎 3 have higher rank (Exercise 17.6.3). If we set 푎 ′ := 푎 4 and 푎 ′′ := 푎 5 , and 푏 ′ := 푏 4 and 푏 ′′ = 푏 8 , then any perfect matching 푀 of 퐺 − 푎 ′ − 푎 ′′ − 푏 ′ − 푏 ′′ must contain both the edges 푒 and 푒 ′ . It cannot contain any edge incident with 푠 other than 푒 ′ . (The contradiction that both 푒 ′ and 푒 ′′ belong to 푀 was arrived at by assuming that 푟 (푒 ′′ ) = 푟 (푒 ′ ) = 푟 (푒). But this situation does not arise.) As a corollary of the above theorem we have:

17.6 The Existence of Thin Edges in Braces

389

Thin edge theorem for braces 17.34 Every brace 퐺 distinct from 퐾2 and 퐶4 has at least two thin edges. Proof Each member of a pair of multiple edges of 퐺 is thin. We may thus assume that 퐺 is simple. The braces 퐾2 and 퐶4 are the only simple braces on fewer than six vertices. We may thus assume that 퐺 has six or more vertices, in which case every edge of 퐺 is removable. If every edge of 퐺 is thin then the assertion holds. We may thus assume that 퐺 has edges that are not thin. Let 푟 ′ be the maximum rank of edges of 퐺 that are not thin. Let 푒 be an edge of 퐺 such that 푟 (푒) = 푟 ′ . By the Rank Augmentation Theorem for Braces, 퐺 has two adjacent removable edges, 푒 ′ and 푒 ′′ , such that at least one of 푒 ′ and 푒 ′′ is thin. If 푒 ′ and 푒 ′′ are both thin then the assertion holds. We may thus assume that precisely one of 푒 ′ and 푒 ′′ , say, 푒 ′ , is thin. In that case, 푟 (푒 ′′ ) = 푟 (푒) = 푟 ′ . We then apply the Rank Augmentation Theorem for Braces with 푒 ′′ playing the role of 푒, thereby obtaining a thin edge that is not adjacent to 푒 ′′ , hence distinct from 푒 ′ . Indeed, 퐺 has at least two thin edges. ƒ

Exercises 17.6.1 Prove statement 17.30.1. ∗ 17.6.2 Prove Lemma 17.31. Hint: see the proof of Lemma 17.18. 17.6.3 In the brace considered in Example 17.33, show that 푟 (푒 ′′ ) = 12 and 푟 (푒 ′′′ ) = 14. ∗ 17.6.4 Prove the following strengthening of Theorem 17.16 for solid bricks, whose statement is similar to that of Theorem 17.30: Theorem 17.35 Let 푒 be a 푏-invariant edge in a simple solid brick 퐺 and suppose that 푒 is not thin. Then there exist two 푏-invariant edges 푒 ′ and 푒 ′′ in 퐺 such that: (i) 푒 ′ and 푒 ′′ are adjacent to each other; (ii) 푟 (푒 ′ ) ≥ 푟 (푒), 푟 (푒 ′′ ) ≥ 푟 (푒); (iii) either 푟 (푒 ′ ) > 푟 (푒) or 푟 (푒 ′′ ) > 푟 (푒); and (iv) if the common end of 푒 ′ and 푒 ′′ is also an end of 푒 then 푟 (푒 ′ ) > 푟 (푒) and 푟 (푒 ′′ ) > 푟 (푒). Hint: use the exchange property of solid graphs (Theorem 10.23) to prove the existence of the two 푏-invariant edges 푒 ′ and 푒 ′′ . ∗ 17.6.5 Use Theorem 17.35 to prove that every solid brick distinct from 퐾4 has two thin edges. Hint: apply Theorem 9.16 to obtain two removable edges and then imitate the proof of Theorem 17.34.

390

17 Thin Edges in Bricks and Braces

17.7 Building Braces In this section we give a precise definition of the three operations necessary for building all braces from 퐾2 and 퐶4 . These operations are applicable to every brace 퐻, producing a brace 퐺. We refer to these three operations as expansions of a brace. Let 퐻 be a brace on which the operation will be applied and let ( 퐴, 퐵) denote the bipartition of 퐻. The resulting graph 퐺 will be a brace. Edge addition: Let 푥 ∈ 퐴 and 푦 ∈ 퐵 be vertices of 퐻. Obtain 퐺 from 퐻 by adding a new edge joining 푥 and 푦. It is easy to see that 퐺 is a brace. Expansion of a vertex by a barrier of size two: Let 퐻 be a brace and let 푥 be a vertex of 퐵 whose degree is at least four. Let 퐻 ′ := 퐻{푥 → (푎, 푏 1 , 푏 2 )} be such that, in the underlying simple graph of 퐻 ′ , 푏 1 and 푏 2 have degree at least three. Now obtain 퐺 from 퐻 ′ by joining 푎 to a vertex 푤 of 퐵 − 푥 by an edge labelled 푒. We shall refer to the graph 퐺 thus constructed as a graph obtained from 퐻 by an expansion of 푥 by a barrier of size two. Expansion of two vertices by barriers of size two: Let 퐻 be a brace and let 푥 ∈ 퐵 and 푦 ∈ 퐴 be two vertices of 퐻 whose degrees are at least four. Let 퐻 ′′ := 퐻{푥 → (푎, 푏 1 , 푏 2 )}{푦 → (푏, 푎 1 , 푎 2 )} be such that, in the underlying simple graph of 퐻 ′′ , each of 푏 1 , 푏 2 , 푎 1 and 푎 2 has degree at least three and, if 퐻 has more than two vertices, each of 푏 1 , 푏 2 , 푎 1 and 푎 2 has at least one neighbour in 푉 (퐻 − 푥 − 푦). Now obtain 퐺 from 퐻 ′′ by joining 푎 and 푏 by an edge labelled 푒. We shall refer to the graph 퐺 thus constructed as a graph obtained from 퐻 by an expansion of 푥 and 푦 by barriers of size two. (Note that the order in which the two vertices 푥 and 푦 are split is immaterial.) Figure 17.31 shows how the cube can be obtained from a brace of order four by applying the above defined operation. Theorem 17.36 Let 퐻 be a brace and let 퐺 be a graph obtained from 퐻 by one of the three operations of expansion. Then 퐺 is a brace. ƒ We remark that the above result also indicates that, in every expansion operation, the edge 푒 is a thin edge of the brace 퐺 and that 퐻 is the brace of 퐺 − 푒. As a complement to the above result, we then state the following theorem, which is easily proved. (Exercise 17.7.3). Theorem 17.37 Every brace 퐺 different from the special braces 퐾2 and 퐶4 can be obtained from one of them by a sequence of applications of the three operations of expansion. ƒ A natural question that arises at this stage is whether all the three operations are really required for building braces. This is indeed the case. (Exercise 17.7.5.) CLM (2015, [18]) proved the following strengthening of Theorem 17.37: Theorem 17.38 Let 퐺 be a brace and let 퐽 be a simple brace that is a matching minor of 퐺. If 퐺 and 퐽 are not isomorphic then 퐺 has a thin edge 푒 such that 퐽 is a matching minor of 퐺 − 푒. ƒ

391

17.7 Building Braces 푒2

푒2

푥 푒3

푒3 푒4

푒1 푒6

푒5

푦

푒

푒4

푒1

푒5

푒6

푒7 퐻

푒7 퐺

Fig. 17.31 Obtaining the cube (G) from a brace (H) of order four by expansion of two vertices by barriers of size two

Corollary 17.39 Let 퐺 be a brace and let 퐽 be a simple brace that is a matching minor of 퐺. Then 퐺 may be obtained from 퐽 by repeated applications of the three expansion operations. ƒ

Exercises ⊲17.7.1 In the definition of the operation of expansion of two vertices by barriers of size two, there is a requirement that “if 퐻 has more than two vertices, each of 푏 1 , 푏 2 , 푎 1 and 푎 2 has at least one neighbour in 푉 (퐻 − 푥 − 푦).” Explain why this requirement is necessary for the resulting graph 퐺 to be a brace if 퐻 has more than two vertices. ⊲17.7.2 Prove Theorem 17.36. 17.7.3 Prove Theorem 17.37. Hint: use induction on |퐸 |. 17.7.4 Generate all the cubic braces on eight or fewer vertices. ⊲17.7.5 Verify that all the three operations are necessary for the generation of all braces starting with the two basic braces. For each of the three operations we give below a hint: (i) edge addition: consider any brace having minimum degree four or more; (ii) expansion of a vertex by a barrier of size two: consider the biwheels of order six or more; (iii) expansion of two vertices by barriers of size two: consider cubic braces of order six or more.

392

17 Thin Edges in Bricks and Braces

This exercise is relevant to the subject of the next chapter. ⊲17.7.6 The families of graphs the following questions pertain to are all described in Section 2.5; some of these graphs are bricks and some are braces. (Hint: The statement in Exercise 8.2.4 is useful in finding the answers to some of the following questions. We recommend to the readers to first work with specific examples of graphs with around ten vertices.) (i) Show that: (a) the thin edges of the wheel 푊2푘+1 , 푘 ≥ 2, are precisely its 2푘 + 1 spokes; (b) each edge of the biwheel B8 (the cube) is a thin edge and, for 푛 ≥ 5, the thin edges of the biwheel B2푛 are precisely the 2푛 − 2 edges incident with its two hubs ℎ1 and ℎ2 ; (c) the thin edges of the truncated biwheel T2푛 , 푛 ≥ 4 are precisely the 2푛 − 4 edges incident with its two hubs ℎ1 and ℎ2 other than ℎ 1 푣 1 , ℎ1 푣 2푛−2 , ℎ2 푣 1 and ℎ2 푣 2푛−2 ; (d) the thin edges of the prism P2푛 , 푛 ≥ 5 are precisely its 푛 rungs 푢 1 푣 1 , 푢 2 푣 2 , . . . , 푢 푛 푣 푛 ; (e) the thin edges of the M¨obius ladder M2푛 , 푛 ≥ 5, are precisely its rungs 푢 1 푣 1 , 푢 2 푣 2 , . . . , 푢 푛 푣 푛 ; and (f) the thin edges of the staircase S2푛 , 푛 ≥ 4, are precisely its rungs 푢 2 푣 2 , . . . , 푢 푛−1 푣 푛−1 . (ii) In each of the above cases, show that the retract of the graph obtained by deleting a thin edge is not simple. (In the terminology to be introduced in the next chapter, such thin edges are not strictly thin.)

17.8 Notes Kothari (2016, [47]), (2019, [48]) proved that, with the exception of 퐾4 and 퐶6 , every near-bipartite brick 퐺 has a thin edge 푒 such that the retract 퐻 of 퐺 − 푒 is also near-bipartite. In fact, he proved that one may fix the removable doubleton 푅 of 퐺 so that 푅 remains a removable doubleton of 퐻.

Chapter 18

Strictly Thin Edges in Bricks and Braces

Contents 18.1 18.2

18.3

18.4 18.5 18.6

Strictly Thin Edges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 393 18.1.1 Families of graphs without strictly thin edges . . . . . . . . . . . 395 Multiple Edges in Retracts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 396 18.2.1 How do multiple edges in the retract of 퐺 − 푒 arise? . . . . . 397 18.2.2 What is the status of those edges in 퐺 itself? . . . . . . . . . . . . 399 18.2.3 Two key lemmas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 399 Characterization of G . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 404 18.3.1 Graphs 퐺 with index(퐺) = 3 . . . . . . . . . . . . . . . . . . . . . . . . . 405 18.3.2 Graphs 퐺 with index(퐺) = 1 . . . . . . . . . . . . . . . . . . . . . . . . . 405 18.3.3 Graphs 퐺 with index(퐺) = 2 . . . . . . . . . . . . . . . . . . . . . . . . . 408 18.3.4 The characterization of G . . . . . . . . . . . . . . . . . . . . . . . . . . . . 410 Generating Simple Braces and Bricks . . . . . . . . . . . . . . . . . . . . . . . . . 411 Splitter Versions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 412 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 413

18.1 Strictly Thin Edges A brick is solid if and only if its underlying simple graph is solid. A brick or brace has a specified simple matching covered graph 퐽 as a matching minor if and and only if its underlying simple graph has 퐽 as a matching minor. As these examples show, the properties of bricks and braces that are of interest to us are mostly those which are enjoyed by simple bricks and braces. This, as explained below, brings into question the efficacy of using thin edges as inductive tools for establishing properties of bricks and braces.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7_18

393

394

18 Strictly Thin Edges in Bricks and Braces

Consider as an example the problem of characterizing simple planar solid bricks. It turns out that every such brick 퐺 is an odd wheel. How do we go about proving that this is indeed the case? If the order 푛 of 퐺 is four, then it is clearly the complete graph 퐾4 which is the wheel 푊3 . So, suppose that 푛 ≥ 6 and assume inductively that every simple planar solid brick with fewer than 푚 edges is a wheel. As both 퐶6 and P are not solid, 퐺 is not a special brick. Thin Edge Theorem 17.3 implies that 퐺 has an edge 푒 such that 퐺 ′ := 퐺š − 푒 is a brick. Using the fact that 퐺 is planar and solid, it is easy to deduce that 퐺 ′ is a planar solid brick (Exercise 18.1.1). As 퐺 ′ clearly has fewer edges than 퐺, it follows from the induction hypothesis that the underlying simple graph of 퐺 ′ is a wheel and, furthermore, that 퐺 itself is obtained from such a graph by an expansion operation. There would be not too many cases to contend with if 퐺 ′ were simple. (See Exercise 18.1.2.) However, the Thin Edge Theorem does not guarantee that 퐺 ′ is simple. This necessitates the consideration of many more cases in order to be able to deduce that 퐺 is a wheel. (The authors confess to ploughing through the needed case analysis in the paper by CLM (2016, [15]). This was before they knew the contents of this chapter.) The above example is just one instance where it would be advantageous to have recursive procedures for generating simple bricks and braces within the realm of simple graphs. (See Exercise 18.1.3 for a second example.) In order to establish the existence of such procedures, one needs the notion of a strictly thin edge in a matching covered graph which is either a brick or a brace. Strictly thin edges in bricks and braces A thin edge 푒 of a simple brick 퐺 is strictly thin if the retract of 퐺 − 푒 is a simple brick. Likewise, a thin edge of a simple brace is strictly thin if the retract of 퐺 − 푒 is a simple brace. As we learnt in the last chapter, all bricks and braces, with a few isolated exceptions, have thin edges. But there exist infinite families of bricks and braces which do not have any strictly thin edges. However, we shall learn that these families of graphs can be identified and that they are precisely the ones described in Exercise 17.7.6. We list them below in the order that will be found to be convenient for later purposes: (i) (ii) (iii) (iv) (v) (vi)

wheels, biwheels, truncated biwheels, prisms, M¨obius ladders and staircases.

18.1 Strictly Thin Edges

395

18.1.1 Families of graphs without strictly thin edges

McCuaig braces In his pioneering paper entitled Brace generation, McCuaig (2001, [68]), identified the three families of braces that do not have strictly thin edges. They are: (i) biwheels, (ii) prisms P푛 , 푛 ≡ 0 (mod 4), and (iii) M¨obius ladders M푛 , 푛 ≡ 2 (mod 4). We shall refer to the union of these families as the McCuaig family of braces. As noted before, no brace in this family has strictly thin edges. More significantly, the following assertion was established in [68]: Theorem 18.1 (Strictly Thin Edge Theorem for Braces) Every simple brace of order six or more that is not a member of the McCuaig family of braces has a strictly thin edge.

Norine-Thomas bricks In a paper entitled Brick generation, Norine and Thomas (2008, [79]) identified all the infinite families of simple bricks that are free of strictly thin edges. They are: (i) wheels, (ii) truncated biwheels, (iii) prisms P푛 , 푛 ≡ 2 (mod 4), (iv) M¨obius ladders M푛 , 푛 ≡ 0 (mod 4), and (v) staircases. (Note that 퐾4 is the wheel 푊3 , and 퐶6 is the prism P6 .) We shall refer to the union of these families of bricks as the Norine-Thomas family of bricks. The theorem below follows from the results established in [79]: Theorem 18.2 (Strictly Thin Edge Theorem for Bricks) Every simple brick which is different from the Petersen graph P and which is not a member of the Norine-Thomas family of bricks has a strictly thin edge. It should be noted that the papers [68] and [78], mentioned in the definition of the families, contain more general results than the ones stated above. A description of those results will be given at the end of the chapter. In the previous chapter, we presented separate proofs for Thin Edge Theorems (17.3 and 17.34) for bricks and braces. Relying on those two theorems, and the fact that bricks and braces are the only matching covered graphs that are free of nontrivial tight cuts, we are able to prove the two Strictly Thin Edge Theorems (18.1 and 18.2) together by considering the class of graphs consisting of both bricks and braces: Definition 18.3 A matching covered graph is indecomposable if it is either a brick or a brace. We shall present a characterization of indecomposable graphs which are free of strictly thin edges and deduce from it the above-mentioned theorems of McCuaig, and of Norine and Thomas.

396

18 Strictly Thin Edges in Bricks and Braces

Exercises − 푒 is a planar 18.1.1 Let 푒 be a thin edge in a planar solid brick 퐺. Show that 퐺š solid brick. (Hint: Use Theorems 10.14 and 7.1.) 18.1.2 Find all the simple planar bricks that can be obtained from the wheels 푊5 and 푊7 by means of the four expansion operations, and identify those among the resulting bricks which are solid. 18.1.3

− 푒 (i) Let 퐺 be a simple brick and let 푒 be a strictly thin edge of 퐺 such that 퐺š is the Petersen graph P. Show that there exists another strictly thin edge 푓 of 퐺 › such that 퐺 − 푓 is not the Petersen graph. (ii) Assuming the Strictly Thin Edge Theorem for Bricks, and using the result in the first part, show that every brick 퐺 that is not a Petersen brick contains a − 푒 is not a Petersen brick. (This provides a proof of thin edge 푒 such that 퐺š Theorem 17.27.)

18.2 Multiple Edges in Retracts Let G denote the class of graphs consisting of 퐾2 , 퐶4 , the Petersen graph and the members of the six infinite families described in Section 18.1.1. Our principal objective is to to show that any simple indecomposable graph that is not a member of G has a strictly thin edge. The two basic braces, 퐾2 and 퐶4 , as well as the three special bricks, 퐾4 , 퐶6 and P, are members of G. Thus, by Theorems 17.3 and 17.34, every simple indecomposable graph that does not belong to G has at least one thin edge. In the remaining part of this section, we shall use the notation specified in the inset below: Notation: 퐺, 푒, 퐻, 푓 and 푔 • 퐺 is a simple indecomposable graph which does not have any strictly thin edges, (our objective is to show that 퐺 ∈ G); • 푒 is a thin edge of 퐺 (which is not strictly thin); • 퐻 := 퐺š − 푒 is the retract of 퐺 − 푒; and • 푓 and 푔 are two edges which are parallel to each other in 퐻. Since 퐺 itself is simple, 푓 and 푔 must be incident in 퐻 with at least one contraction vertex (that is, one resulting from the bicontraction of an end of 푒). Let us begin with a brief review of the conditions under which bicontractions of vertices of degree two in 퐺 − 푒 result in a graph with multiple edges. As 푒 is thin but not strictly thin in 퐺, we note that 퐻 ≠ 퐺 − 푒. Thus, the index of 푒 is either one, two or three, as defined in Chapter 13 (and reviewed in Chapter 17). In each of these

18.2 Multiple Edges in Retracts

397

cases, as described in the following inset, we can infer that either one or both ends of 푒 have degree three in 퐺. Vertices of degree three and the solid dot convention If the index of 푒 is one, then exactly one end of 푒 has degree three in 퐺, and if the index of 푒 is two, then both ends of 푒 have degree three in 퐺 and they have no common neighbour. Finally, if the index of 푒 is three, both ends of 푒 have degree three in 퐺 and they have a common neighbour. This property holds for any thin edge of 퐺. In proving that 퐺 must be a member of G, we shall make repeated use of the above-stated property to deduce that 퐺 has at most two vertices whose degrees are greater than three. In our illustrations we shall indicate vertices whose degrees can be inferred to be three by solid dots.

18.2.1 How do multiple edges in the retract of 푮 − 풆 arise? ♯ The index of the chosen thin edge 푒 of 퐺 may be either one, two or three. In what follows we shall examine each of these cases separately to discern how parallel edges in 퐻 = 퐺š − 푒 might arise.

Case in which the index of 푒 is one: Suppose that 푒 = 푢푣 is a thin edge of index one. Let 푣 be the end of 푒 that has degree three in 퐺, and let 푣 1 and 푣 2 be its neighbours distinct from 푢. The retract of 퐺 − 푒 has only one contraction vertex and that vertex is the result of the bicontraction of 푣 in 퐺 − 푒. Thus if 푓 and 푔 are parallel edges in the retract 퐻 of 퐺 − 푒, they must be adjacent edges in 퐺 such that one of them is incident with 푣 1 and the other with 푣 2 . Without loss of generality, we may thus assume that 푓 := 푣 1 푤 and 푔 := 푣 2 푤 are incident with a common vertex 푤 which is not in {푣, 푣 1 , 푣 2 }. See Figure 18.1. The degree of 푤 is at least four in 퐺 and it is possible for 푢 and 푤 to be the same vertex (in which case 푤 would have degree five or more in 퐺). 푣

푣1

푣2 푓

푔 푤 푢

Fig. 18.1 Parallel edges in 퐻, index(푒) = 1

푒

18 Strictly Thin Edges in Bricks and Braces

398

Case in which the index of 푒 is two: Suppose that 푒 = 푢푣 is a thin edge of index two in 퐺. The degrees of both 푢 and 푣 are three in 퐺, and are two in 퐺 − 푒. Let 푢 1 , 푢 2 and 푣 be the three neighbours of 푢, and let 푣 1 , 푣 2 and 푢 be the three neighbours of 푣 in 퐺. As the index of 푒 is two, 푢 and 푣 do not have a common neighbour, implying that {푢 1 , 푢 2 } ∩ {푣 1 , 푣 2 } = ∅. There are essentially three possible situations under which 푓 and 푔 become parallel edges when 푢 and 푣 are bicontracted in 퐺 − 푒. These three possibilities are illustrated in Figure 18.2. 푣

푣

푣1

푣2 푓

푒

푣

푣1

푣2

푣1

푣2

푓

푔

푢1

푢2

푔 푒

푤 푢1

푓

푢2

푒

푔

푢1

푢2

푢

푢

푢

Fig. 18.2 Parallel edges in 퐻, index(푒) = 2

Case in which the index of 푒 is three: Suppose now that 푒 = 푢푣 has index three. In this case, the ends 푢 and 푣 of 푒 have a common neighbour in 퐺. Let 푣 1 and 푣 2 be the neighbours of 푣 in 퐺 − 푒, and let 푣 2 and 푣 3 be the neighbours of 푢 in 퐺 − 푒. The bicontractions of 푢 and 푣 in 퐺 − 푒 result in a single contraction vertex. So, if 푓 and 푔 are parallel edges in the retract 퐻 of 퐺 − 푒, they must be incident with distinct vertices in {푣 1 , 푣 2 , 푣 3 } and must have a common end, say 푤, which is not in {푢, 푣, 푣 1 , 푣 2 , 푣 3 }. The common end 푤 of 푓 and 푔 has degree at least four in 퐺. In Figure 18.3, which illustrates such an occurrence, edges 푓 and 푔 are indicated by solid lines. 푣

푣1

푒

푣2

푔

푓 푤 Fig. 18.3 Parallel edges in 퐻, index(푒) = 3

푢

푣3

18.2 Multiple Edges in Retracts

399

18.2.2 What is the status of those edges in 푮 itself? Recall that 퐺 is a simple indecomposable graph in which the edge 푒 is thin, and that 푓 and 푔 are edges in the retract 퐻 of 퐺 − 푒 which are parallel to each other. Being multiple edges of 퐻, both 푓 and 푔 are clearly removable in 퐻. As 퐻 is the retract of 퐺 − 푒, it is obtained from 퐺 − 푒 by either one or two bicontractions. If 퐶 is the tight cut associated with a bicontraction, and if 푓 belongs to 퐶, it can be seen that 푓 is a multiple edge (though not necessarily parallel to 푔) in both the 퐶-contractions of 퐺 − 푒. Using this observation, it is easy to verify the following assertion. (We leave the details as Exercise 18.2.1). Proposition 18.4 Edge 푓 (and similarly edge 푔) is removable in 퐺 − 푒.

ƒ

If 퐺 is a brace (of order six or more), 푓 and 푔 are also removable in 퐺, but they may or may not be thin (see Exercise 18.2.2). On the other hand, if 퐺 is a brick, neither 푓 nor 푔 needs to be removable in 퐺. For example, in the bicorn H8 shown in Figure 18.4 the edge 푒 is thin, the edges 푓 and 푔 are parallel edges in the retract of H8 − 푒, but neither of them is removable in H8 .

푓

푒

푔

Fig. 18.4 Edges 푓 and 푔 are removable in H8 − 푒 but not in H8

The graph 퐻 is indecomposable and 푓 and 푔 are parallel edges in 퐻; thus the graph 퐻 − 푓 is indecomposable. As 푓 is a removable edge in 퐺 − 푒, the graph 퐺 − 푒 − 푓 is matching covered; and in fact 퐻 − 푓 is the retract of 퐺 − 푒 − 푓 . The following proposition is thus obvious: Proposition 18.5 Up to multiple edges tight cut decompositions of 퐺 − 푒 − 푓 and of 퐺 − 푒 produce the same list of indecomposable graphs. ƒ

18.2.3 Two key lemmas Various possibilities occur depending on the status of 푓 and 푔 as to their removability or thinness in 퐺. We shall examine all those possibilities and conclude that, in each case, the constraint that 퐺 has no strictly thin edges implies that 퐺 must contain certain configurations in the vicinity of the ends of 푒. In the next section we shall show

18 Strictly Thin Edges in Bricks and Braces

400

that those local configurations must replicate themselves, and close up eventually, implying that 퐺 belongs to the family G. The configuration defined below features prominently in the first lemma. Houses A house with 푒 as its floor and 푓 as as its ceiling is an induced subgraph 퐹 of 퐺 consisting of a pentagon with precisely one chord where that chord is the edge 푓 , and 푒 is the edge of the pentagon disjoint from 푓 , with the additional property that the ends of 푒 and 푓 have degree three in 퐺. See Figure 18.5. (In H8 shown in Figure 18.4, there is a house with 푒 as its floor and 푓 as its ceiling, and one with 푒 as its floor and 푔 as its ceiling.)

푓

푒 Fig. 18.5 A house with floor 푒 and ceiling 푓

The case in which 푓 is not removable in 퐺 Lemma 18.6 Let 퐺 be a simple brick, let 푒 := 푢푣 be a thin edge of 퐺 of index one or two and let 푓 be a multiple edge in the retract 퐻 of 퐺 − 푒. If 푓 is not removable in 퐺 then the following properties hold: (i) if index(푒) = 1 then 푓 has an end of degree three which is adjacent to both ends of 푒, and (ii) if index(푒) = 2 then 푓 is the ceiling of a house whose floor is 푒. See Figure 18.6. Proof As 퐺 is a brick and 푒 is thin, the retract 퐻 of 퐺 − 푒 is also a brick. Moreover, as 푓 is a multiple edge of 퐻, the graph 퐻 − 푓 is also a brick. We have seen that 퐺 − 푒 − 푓 is matching covered. Moreover, the retract of 퐺 − 푒 − 푓 is 퐻 − 푓 , a brick. Thus, 퐺 − 푒 − 푓 is a near-brick. Assume that 푓 is not removable in 퐺, implying that 퐺 − 푓 is not matching covered. But, 퐺 − 푒 − 푓 is matching covered; hence 푒 is the only unmatchable edge in 퐺 − 푓 (equivalently, 푒 is the only edge that depends on 푓 ). It follows that 퐺 − 푒 − 푓 has a maximal barrier, say 퐵, such that 푒 is the only edge of 퐺 that has both ends in 퐵. Edge 푓 is the only edge which has ends in different components of 퐺 − 푓 − 퐵. As 퐺 − 푒 − 푓 is a near-brick, the maximality of 퐵 implies that 퐵 is special, by Proposition 17.6.

18.2 Multiple Edges in Retracts

퐵

푢

푒

푠

푓

401

푣

푡

푢

푠

푒

푣

푓

푡

푤

퐾

퐵

퐾

(푎)

(푏)

Fig. 18.6 Lemma 18.6: (a) index(푒) = 1; (b) index(푒) = 2

Suppose that index(푒) ∈ {1, 2}. Let 퐾 denote the nontrivial component of 퐺 − 푓 − 퐵. As 퐻 − 푓 , a brick, is the retract of 퐺 − 푒 − 푓 , up to multiple edges 퐻 is a brick of 퐺/푉 (퐾); hence As 퐵 is special,

|푉 (퐻)| ≤ |푉 (퐾)| + 1. 2|퐵| − 1 = |푉 (퐺)| − |푉 (퐾)|.

As 푒 is thin and index(푒) ∈ {1, 2}, |푉 (퐺)| − |푉 (퐻)| = 2 · index(푒).

Addition of the inequalities above and simplification yields |퐵| ≤ index(푒) + 1.

(18.1)

Let 퐼 be the set of |퐵| − 1 isolated vertices of 퐺 − 푓 − 퐵. We shall now analyse two cases, which turn out to be correspond to the two items in the assertion of the lemma. Case 1 The set 퐼 contains a vertex, 푠, which is adjacent to both ends of 푒. Then, 푒 is in a triangle of 퐺. We have assumed that index(푒) ∈ {1, 2}. If index(푒) = 2 then the edge 푒 is not in any triangle of 퐺. Thus, index(푒) = 1. In that case, from (18.1) we infer that |퐵| = 2; hence 퐵 = {푢, 푣}. The vertex 푠 is adjacent in 퐺 to three or more vertices. We conclude that 푠 must then be an end of 푓 and have degree three in 퐺. Hence the assertion holds in this case. Case 2 No vertex of 퐼 is adjacent to both ends of 푒. Let 푠 be a vertex in 퐼. It has degree three or more in 퐺. At least two of the neighbours of 푠 are in 퐵 and one of them is not an end of 푒. Thus, |퐵| ≥ 3. We have assumed that index(푒) ∈ {1, 2}. From (18.1) we infer that |퐵| = 3 and index(푒) = 2.

18 Strictly Thin Edges in Bricks and Braces

402

Let 푤 be the vertex of 퐵 − 푢 − 푣. Then, 푠 is adjacent to one of the ends of 푒, and to 푤. Moreover, 푠 is an end of 푓 and has degree three in 퐺. Likewise, the vertex 푡 ∈ 퐼 − 푠 is also adjacent to an end of 푒 and to 푤, it is an end of 푓 and has degree three. Finally, as index(푒) = 2, no triangle of 퐺 contains an end of 푒. Thus, one of 푠 and 푡 is adjacent to 푢, the other is adjacent to 푣. We deduce that 퐺 has a house with 푒 as ƒ its floor and 푓 as its ceiling. The case in which 푓 is removable but not thin in 퐺 Lemma 18.7 Let 퐺 be a simple brace on ten or more vertices, or a simple brick, let 푒 := 푢푣 be a thin edge of 퐺, let 퐻 be the retract of 퐺 − 푒 and let 푓 and 푔 be two parallel edges of 퐻. Assume that 푓 is removable in 퐺 but that it is not thin. Then, the following properties hold: (i) index(푒) = 2 and (ii) edge 푔 is thin in 퐺 and index(푔) ≤ 1. Proof Suppose that the edge 푓 is removable in 퐺. For any cut 퐶 of 퐺, if 퐶 − 푓 is a tight cut of 퐺 − 푓 then 퐶 −푒 − 푓 is a tight cut of 퐺 −푒 − 푓 . Thus by starting with a tight cut decomposition of 퐺 − 푓 and refining it we may obtain a tight cut decomposition of 퐺 − 푒 − 푓 . Let F be the set of indecomposable graphs obtained from a tight cut decomposition of 퐺 − 푓 . We may refine F to a tight cut decomposition of 퐺 − 푒 − 푓 . The graph 퐻 − 푓 is the retract of 퐺 − 푒 − 푓 . Thus, F contains a graph 퐹 such that a tight cut decomposition of 퐹 − 푒 produces a graph isomorphic to 퐻, up to multiple edges. Thus, |푉 (퐹)| ≥ |푉 (퐻)|; hence |푉 (퐺)| − 4 ≥ |푉 (퐹)| ≥ |푉 (퐻)| ≥ |푉 (퐺)| − 4,

(18.2)

where the leftmost inequality follows from the fact that 푓 is not thin, and the rightmost inequality comes from the fact that 퐻 is the retract of 퐺 −푒 and 푒 is thin. We conclude that |푉 (퐹)| = |푉 (퐻)| = |푉 (퐺)| − 4. Consequently, index(푒) ∈ {2, 3}; hence both ends of 푒 have degree three in 퐺. If 퐹 has two contraction vertices then the tight cut decomposition F of 퐺 − 푓 would contain, apart from 퐹, two 퐶4 ’s. If this were the case, 푓 would be a thin edge of index two, a contradiction. We conclude that 퐹 has just one contraction vertex. Let 퐶 := 휕 ( 푋) then be a nontrivial cut of 퐺 such that 퐶 − 푓 is tight in 퐺 − 푓 , the graph 퐺 1 := (퐺 − 푓 )/( 푋 → 푥) is bipartite and has order six, and 퐹 = (퐺 − 푓 )/( 푋 → 푥). See Figure 18.7. Let 퐵 denote the barrier of 퐺 − 푓 associated with 퐶 and let 퐼 be the set of isolated vertices of 퐺 − 푓 − 퐵. The edge 푒 has no end in 푋, because both ends of 푒 have degree three and a tight cut decomposition of 퐹 −푒 would produce an indecomposable graph of order |푉 (퐻)| < |푉 (퐹)|. Thus, 푒 has both ends in 푋. Adjust notation so that 푢 ∈ 퐼 and 푣 ∈ 퐵. As index(푒) ∈ {2, 3}, no edge of 퐺 − 푒 incident with an end of 푒 is an edge of 퐻, yet 푓 is an edge of 퐻. Thus, 푓 is not incident with 푢. It follows that 퐵 consists of the three vertices of 퐺 adjacent to 푢. One of the vertices of 퐵 is 푣; the

18.2 Multiple Edges in Retracts 푣

403 푢1

푢2

PSfrag replacements

퐵

푣2

푒 푔

퐶 푋 푓 푢

푣1

푤

Fig. 18.7 Edge 푓 is removable but not thin in 퐺.

two vertices of 퐵 − 푣 are the two vertices of 퐺 − 푒 adjacent to 푢, say, 푢 1 and 푢 2 . It follows that no triangle of 퐺 contains edge 푒; therefore index(푒) < 3. We have seen that index(푒) ∈ {2, 3}, hence index(푒) = 2. This proves the first part of the assertion. Let 푣 1 and 푣 2 be the two neighbours of 푣 in 퐺 − 푒. Edge 푓 is not incident with any vertex of 퐵. In particular, 푓 is not incident with 푢 1 , nor with 푢 2 . Thus, 푓 and 푔 have a common end, 푤, not in {푢, 푣, 푢 1 , 푢 2 , 푣 1 , 푣 2 } (see Figure 18.2). Moreover, one of 푓 and 푔 joins 푣 1 to 푤; the other joins 푣 2 to 푤. Adjust notation so that 푓 = 푣 1 푤 and 푔 = 푣 2 푤. The vertex 푤, the end of 푓 distinct from 푣 1 , is in 푋. The edge 푔 joins 푣 2 to 푤. Thus, both ends of 푔 are in 푋. In sum, the vertices 푣 2 and 푤 are in 푋 and 푣 1 is in 퐼. Note that 푁 (푣 1 ) ⊆ 퐵 + 푤. As the edge 푔 is not incident with any vertex in 퐵, it follows that no end of 푔 is adjacent to vertex 푢. We deduce that there cannot be a house with 푒 as the floor and 푔 as the ceiling. Therefore, by Lemma 18.6, edge 푔 is removable in 퐺. Assume, to the contrary, that 푔 is not thin. We may then apply to 푔 the argument that we applied to 푓 , with 푣 2 playing the role of 푣 1 and conclude that 푁 (푣 2 ), just like 푁 (푣 1 ), is a subset of 퐵 + 푤. In 퐺 − 퐵 − 푤, the vertices 푢, 푣 1 and 푣 2 are isolated; therefore 퐵 + 푤 is a barrier of 퐺. This is a contradiction, because, no brick, or a brace on ten or more vertices, can have a barrier of size four. Thus, 푔 is thin. If 퐺 is a brick then 퐹 is a brick and 푤 is adjacent to three or more vertices in 퐹. Alternatively, if 퐺 is a brace then by hypothesis it has order ten or more; hence 퐹 is a brace of order |푉 (퐹)| = |푉 (퐺)| − 4 ≥ 6. In both alternatives, the end 푤 of 푔 is adjacent to three or more vertices in 퐹; hence it is adjacent to four or more vertices in 퐺. We conclude that index(푔) ≤ 1. ƒ

18 Strictly Thin Edges in Bricks and Braces

404

Exercises ∗ 18.2.1 Let 퐺 be an indecomposable graph, let 푒 be a thin edge of 퐺, and let 푓 and 푔 be two parallel edges of 퐺š − 푒. Show that both 푓 and 푔 are removable in the matching covered graph 퐺 − 푒. 18.2.2 Consider the brace 퐺 shown in Figure 18.8. Verify the following: (i) 푒 is a thin edge of 퐺, (ii) 푓 and 푔 are multiple edges in 퐺š − 푒, and (iii) 푓 is a thin edge of 퐺, but 푔 is not. 푔

푓

푒

› Fig. 18.8 Edge 푒 is thin and edges 푓 and 푔 are multiple edges in 퐻 = 퐺 − 푒. Edge 푓 is thin in 퐺, but 푔 is not. (This brace is the same as the one in Figure 17.25.)

18.3 Characterization of G ♯ To establish the stated characterization of G, we shall find it convenient to introduce a new parameter which is defined in the following inset. Definition: index(퐺) For an indecomposable graph 퐺 which has thin edges, we define index(퐺) to be equal to the lowest index of a thin edge of 퐺. For example, since all the thin edges of the tricorn H10 (depicted in Figure 17.1(c)) have index three, index(H10 ) = 3. The brace 퐺 in Figure 18.8 has some thin edges of index one and some of index two, and none of index zero. Thus index(퐺) = 1. It is important to notice that every multiple edge of 퐺 is strictly thin. Thus, if index(퐺) > 0 or if 퐺 is free of strictly thin edges then 퐺 is simple.

18.3 Characterization of G

405

18.3.1 Graphs 푮 with index(푮) = 3 It turns out that establishing the stated characterization of G, thin edges of index three are easy to deal with. Let us first note that if index(퐺) = 3, then 퐺 is a brick. Lemma 18.8 Let 퐺 be a brick. If index(퐺) = 3 then every thin edge of 퐺 is strictly thin. Proof Let 푒 be a thin edge of 퐺. Let 퐻 be the retract of 퐺 − 푒. Assume, to the contrary, that 퐻 is not simple. Let 푓 and 푔 denote two multiple edges of 퐻. All the ends of 푒 and its neighbours are contracted to a single vertex in the case where index(푒) = 3. Therefore, 푓 and 푔 have a common end, 푤, not adjacent to any end of 푒. See Figure 18.3. Graph 퐻 is a brick; therefore vertex 푤 is adjacent in 퐻 to three or more vertices. We deduce that 푤 has degree four or more in 퐺. Let 푀 be a perfect matching of 퐺 that contains edge 푒. As 푓 and 푔 are adjacent in 퐺, it follows that at most one of 푓 and 푔 lies in 푀. Adjust notation so that 푓 does not lie in 푀. The edge 푓 is removable in 퐺 − 푒 (Exercise 18.2.1). By Proposition 17.11, edge 푓 is removable in 퐺. By Lemma 18.7, edge 푓 is thin in 퐺. But the end 푤 of 푓 has degree four or more; therefore index( 푓 ) < 2. This is a contradiction to the hypothesis that index(퐺) = 3. We conclude that 푒 is strictly thin. This conclusion ƒ holds for each thin edge 푒 of 퐺. In view of the above lemma, we may now restrict our attention to indecomposable graphs with index = 1 or 2 which have thin edges, but are free of strictly thin edges. The structure of such a graph neatly depends on the value of that parameter. We divide the proof accordingly.

18.3.2 Graphs 푮 with index(푮) = 1 In this case, odd wheels, biwheels and truncated biwheels emerge as the families of indecomposable graphs without strictly thin edges. The following lemma establishes the existence of a certain configuration in the vicinity of thin edges of index one in graphs with index = 1. It is a simple consequence of Lemmas 18.6 and 18.7. We shall state it in a notation that is convenient for describing how this configuration replicates itself. Lemma 18.9 Let 퐺 be a brace on ten or more vertices, or a brick. Assume that 퐺 is free of strictly thin edges. Let 푒 := ℎ푣 2 be a thin edge of index one of 퐺, where 푣 2 is the end of 푒 with degree three. Let 푣 1 and 푣 3 denote the neighbours of 푣 2 distinct from ℎ. Then, vertices 푣 1 and 푣 3 have both degree three in 퐺 and there is a vertex ℎ′ of degree four or more, possibly equal to ℎ, that is adjacent to both 푣 1 and 푣 3 (Figure 18.9). Moreover: (i) if 푓 := ℎ′ 푣 1 is not removable in 퐺, then ℎ and ℎ′ are distinct and 푣 1 is adjacent to ℎ (Figure 18.9(a)), and

18 Strictly Thin Edges in Bricks and Braces

406

(ii) if 푓 := ℎ′ 푣 1 is removable in 퐺, then it is thin of index one (Figures 18.9(b) and 18.9(c)). (Similar statements also apply to 푔 := ℎ′ 푣 3 .) ℎ′

ℎ′

푓

푓

푣1

푣2

푣3

푣1 푣2

푣1

푣2

푓 푒

푣3

푣3 푒

푒

ℎ (a)

ℎ (b)

ℎ = ℎ′ (c)

Fig. 18.9 Lemma 18.9: (a) 푓 is not removable; (b) 푓 is thin and ℎ′ ≠ ℎ; (c) 푓 is thin and ℎ′ = ℎ

Proof Let 퐻 denote the indecomposable graph obtained from 퐺 − 푒 by the bicontraction of 푣 2 . By hypothesis, 푒 is not strictly thin. Therefore, 퐻 has multiple edges. This implies that 퐺 has a vertex ℎ′ , possibly equal to ℎ, but distinct from 푣 2 , that is adjacent to both 푣 1 and 푣 3 . Let us now prove that vertex ℎ′ has degree four or more in 퐺. If 퐺 is a brick then ′ ℎ is adjacent to three or more vertices in 퐻. If 퐺 is a brace, then, by hypothesis, 퐺 has eight or more vertices; therefore 퐻 is a brace on six or more vertices. In both alternatives, ℎ′ is adjacent to three or more vertices in 퐻. We conclude that ℎ′ has degree four or more in 퐺. To complete the proof, we consider separately two cases, depending on whether or not 푓 = ℎ′ 푣 1 is removable in 퐺. In the case in which 푓 is not removable, then, by Lemma 18.6, the edges 푒 and 푓 are not adjacent; hence ℎ and ℎ′ are distinct. Moreover, vertex 푣 1 has degree three and is adjacent to both ℎ and 푣 2 (Figure 18.9(a)). The assertion holds in this case. Thus, suppose that 푓 is removable in 퐺. By Lemma 18.7, if the edge 푓 were not thin, the index of 푒 would have to be two. Thus, 푓 is thin in 퐺. By hypothesis, 퐺 is free of strictly thin edges. Therefore, index( 푓 ) > 0. One end of 푓 , namely ℎ′ , has degree four or more. As index( 푓 ) ≥ 1, we deduce that the other end 푣 1 of 푓 has degree three, and that edge 푓 has index one. By applying similar arguments to the edge 푔 := ℎ′ 푣 3 , we may conclude that 푣 3 also has degree three, and also similar properties for 푔. ƒ Theorem 18.10 Let 퐺 be a brace on ten vertices or more, or a brick. Suppose that 퐺 is free of strictly thin edges. If index(퐺) = 1, then 퐺 is either an odd wheel, a biwheel or a truncated biwheel.

18.3 Characterization of G

407

Proof By hypothesis, 퐺 has a thin edge 푒 := ℎ푣 2 of index one. Adjust notation so that 푣 2 has degree three. Then, ℎ has degree four or more. Let 푣 1 and 푣 3 be the two neighbours of 푣 2 distinct from ℎ. By Lemma 18.9, 퐺 has a vertex ℎ′ of degree four or more, possibly equal to ℎ, that is adjacent to both vertices 푣 1 and 푣 3 . Moreover, vertices 푣 1 and 푣 3 have both degree three. Let thus 푃 := 푣 1 푣 2 . . . 푣 푡 , 푡 ≥ 3, be a path of maximum length in 퐺 − ℎ − ℎ′ that has the following properties (see Figure 18.10): (i) every vertex 푣 푖 of 푃 has degree three, and is adjacent to ℎ if 푖 is even, and to ℎ′ if 푖 is odd, and (ii) for every internal vertex 푣 푖 of 푃, the edge of 퐺 that joins 푣 푖 to one of ℎ and ℎ′ is thin of index one. ℎ′

푣1

푣2

푣3

푣4

푣5

푣6

푣7

ℎ Fig. 18.10 An illustration for Theorem 18.10.

Let 푓 := ℎ′ 푣 1 , and 푓 ′ := 푥푣 푡 , where 푥 = ℎ′ if 푡 is odd and 푥 = ℎ if 푡 is even. To draw the desired conclusion, we shall now consider the status of the edges 푓 and 푓 ′ . Case 1 One of the edges 푓 and 푓 ′ is removable in 퐺. Adjust notation so that 푓 is removable in 퐺. By Lemma 18.9, we deduce that 푓 is a thin edge of index one. Also, by Lemma 18.9, with 푓 playing the role of 푒, we deduce that the neighbours of 푣 1 , distinct from ℎ′ , have degree three in 퐺. Thus 푣 1 has a neighbour 푣 0 of degree three in 퐺 which is different from 푣 2 . By the maximality of 푃, vertex 푣 0 lies in 푉 (푃). All the vertices of 푃 have degree three. Therefore, 푣 0 cannot be an internal vertex of 푃. We deduce that 푣 0 = 푣 푡 . Hence the subgraph of 퐺 induced by 푉 (푃) is a cycle. Moreover, every vertex of 푃 is adjacent only to vertices of 푉 (푃) ∪ {ℎ, ℎ′ }. Graph 퐺, a brick or a brace on more than four vertices, is 3-connected. Therefore, 푉 (퐺) = 푉 (푃) ∪ {ℎ, ℎ′ }. If 푡 is odd then ℎ = ℎ′ and 퐺 is an odd wheel, on six or more vertices, with ℎ as its hub. If 푡 is even then ℎ and ℎ′ are

408

18 Strictly Thin Edges in Bricks and Braces

distinct and 퐺 is a biwheel, on ten or more vertices, with ℎ and ℎ′ as its two hubs, In both alternatives, edge 푓 ′ is also removable in 퐺. (Figure 2.9 depicts the biwheel B10 .) Case 2 Neither 푓 nor 푓 ′ is removable in 퐺. By applying Lemma 18.9, and using the fact that 푓 is not removable, we deduce that ℎ and ℎ′ are distinct and 푣 1 is joined to both ℎ and ℎ′ . Likewise, 푣 푡 is joined to both ℎ and ℎ′ . The set {ℎ, 푣 1 , 푣 2 } induces a triangle; hence 퐺 is not bipartite. Thus, 퐺 is a brick. Moreover, every vertex of 푃 is adjacent only to vertices of 푉 (푃) ∪ {ℎ, ℎ′ }. Graph 퐺, a brick, is 3-connected. Therefore, 푉 (퐺) = 푉 (푃) ∪ {ℎ, ℎ′ }. As ℎ and ℎ′ are distinct, it follows that 푡 is even. Then, as the degrees of ℎ and ℎ′ are equal and greater than three, it follows that 푡 ≥ 6. We deduce that 퐺 is a truncated biwheel. (Figure 2.10 depicts the truncated biwheel T10 .) ƒ

18.3.3 Graphs 푮 with index(푮) = 2 Now we turn to indecomposable graphs 퐺 in which there are no strictly thin edges, and index(퐺) = 2. Every such graph turns out to be either a prism, or a M¨obius ladder, or a staircase. As in the case index(퐺) = 1, we start with a lemma which establishes the existence of a certain configuration in the vicinity of every thin edge of index two. Lemma 18.11 Let 퐺 be a brace on ten vertices or more, or a brick. Suppose that index(퐺) = 2 and that 퐺 is free of strictly thin edges. Let 푒 := 푢 2 푣 2 be a thin edge of index two. Let 푢 1 and 푢 3 be the two neighbours of 푢 2 in 퐺 − 푒 and let 푣 1 and 푣 3 be the two neighbours of 푣 2 in 퐺 − 푒. Then, the vertices 푢 1 , 푢 3 , 푣 1 and 푣 3 are distinct and have degree three. Moreover, there are precisely two edges joining vertices in {푢 1 , 푢 3 } to vertices in {푣 1 , 푣 3 } and those two edges are nonadjacent (see Figure 18.11(c)). Proof Let 퐻 be the retract of 퐺 − 푒. By hypothesis, 푒 is not strictly thin. Therefore, 퐻 has multiple edges. Let 푓 and 푔 denote two multiple edges of 퐻. Then, both edges have ends that are adjacent to the same end of 푒 (see Figure 18.2). Adjust notation so that 푓 is incident with vertex 푢 1 , whereas 푔 is incident with vertex 푢 3 . 18.11.1 Both ends of 푓 have degree three in 퐺. Moreover, if 푓 is removable in 퐺 then 푓 is thin. (A similar statement holds for 푔.) Proof If 푓 is not removable in 퐺 then, by Lemma 18.6, edge 푓 is the ceiling of a house whose floor is edge 푒; in that case, both ends of 푓 have degree three. Suppose thus that 푓 is removable in 퐺. If 푓 is not thin then, by Lemma 18.7, 푔 is thin and has index one. This is a contradiction to the hypothesis that index(퐺) = 2. Thus, 푓 is thin; hence index( 푓 ) ∈ {2, 3}. Consequently, both ends of 푓 have degree three. A similar conclusion holds for edge 푔. ƒ

18.3 Characterization of G

409

푤

푓 푢1

푔 푢2 푒

푢2

푢1

푢3

푢3

푔

푓

푢1

푓

푢2

푒

푢3

푔

푒 푣2 푣1 푣3 (a) 푤 not adjacent to 푣2

푣1

푣2 푣3 (b) 푤 adjacent to 푣2

푣1 푣2 푣3 (c) 푓 and 푔 nonadjacent

Fig. 18.11 Illustration for Lemma 18.11.

18.11.2 The edges 푓 and 푔 are not adjacent. Proof Assume that 푓 and 푔 share a common vertex, 푤. Suppose that 푤 is not adjacent to 푣 2 (see Figure 18.11(a)). Then, 푤 is not adjacent to any end of 푒. By Lemma 18.6, edges 푓 and 푔 are both removable in 퐺. As index(퐺) = 2, it follows, by Lemma 18.7, that 푓 and 푔 are both thin. In 퐻, a brace on six or more vertices, or a brick, vertex 푤 is adjacent to three or more vertices. Therefore, 푤 has degree four or more in 퐺. It follows that index( 푓 ), index(푔) ≤ 1, in contradiction to the hypothesis that index(퐺) = 2. Suppose thus that 푤 ∈ {푣 1 , 푣 3 }. Adjust notation so that 푤 = 푣 1 (see Figure 18.11(b)). The vertex 푣 1 , an end of 푓 , has degree three and it is adjacent to the three vertices of 퐵 := {푣 2 , 푢 1 , 푢 3 }. In 퐺 − 퐵, the vertices 푣 1 and 푢 2 are both isolated. Hence 퐵 is a barrier of 퐺. This conclusion implies that 퐺 is a brace on six vertices, a contradiction to the hypothesis that if 퐺 is a brace then it has ten or more vertices. Indeed, 푓 and 푔 are nonadjacent. ƒ It follows that both 푓 and 푔 join vertices in {푢 1 , 푢 3 } to vertices in {푣 1 , 푣 3 }. Moreover, if we adjust notation so that 푓 = 푢 1 푣 1 , it follows that 푔 = 푢 3 푣 3 (see Figure 18.11(c)). We conclude that the four vertices in {푢 1 , 푢 3 , 푣 1 , 푣 3 } have degree three. ƒ Theorem 18.12 Let 퐺 be a brace on ten or more vertices, or a brick. Assume also that 퐺 is free of strictly thin edges. If index(퐺) = 2 then 퐺 is either a prism, a M¨obius ladder or a staircase. Proof By hypothesis, 퐺 has a thin edge 푢 2 푣 2 of index two. Let 푢 1 and 푢 3 denote the two neighbours of 푢 2 distinct from 푣 2 , let 푣 1 and 푣 3 denote the two neighbours of 푣 2 distinct from 푢 2 . By hypothesis, 푒 is not strictly thin. Therefore, by Lemma 18.11, the vertices in {푢 1 , 푢 3 , 푣 1 , 푣 3 } all have degree three. Moreover, there are precisely two edges that join 푢 1 and 푢 3 to 푣 1 and 푣 3 , and the two edges are nonadjacent. Adjust notation so that 푢 1 is adjacent to 푣 1 and 푢 3 to 푣 3 .

18 Strictly Thin Edges in Bricks and Braces

410

Let 푃 := 푢 1 푢 2 · · · 푢 푡 and 푄 := 푣 1 푣 2 · · · 푣 푡 , 푡 ≥ 3, be two disjoint paths in 퐺 of maximum length, such that for 푖 = 1, 2 . . . , 푡, the vertices 푢 푖 and 푣 푖 are adjacent and both have degree three, and, if 1 < 푖 < 푡 then edge 푢 푖 푣 푖 is thin of index two (see Figure 18.12.). 푢1

푢2

푢3

푢4

푣1

푣2

푣3

푣4

Fig. 18.12 Illustration for Theorem 18.12.

Case 1 At least one of 푢 1 푣 1 and 푢 푡 푣 푡 is removable in 퐺. Adjust notation so that 푢 1 푣 1 is removable in 퐺. By Lemma 18.7, 푢 1 푣 1 is thin in 퐺. The index of 푢 1 푣 1 is two or three, because index(퐺) = 2. But the index of 푢 1 푣 1 cannot be three; otherwise there would be a vertex 푦 adjacent to both 푢 1 and 푣 1 , and consequently edge 푢 2 푣 2 would depend on edge 푢 1 푣 1 , a contradiction. Indeed, 푢 1 푣 1 is thin of index two. By Lemma 18.11, vertex 푢 1 has a neighbour 푢 0 distinct from 푢 2 and 푣 1 , and 푣 1 has a neighbour 푣 0 distinct from 푢 1 and 푣 2 , such that 푢 0 and 푣 0 are adjacent and have degree three. By the maximality of 푡, it follows that {푢 0 , 푣 0 } = {푢 푡 , 푣 푡 }. If 푢 0 = 푢 푡 then 푣 0 = 푣 푡 and 퐺 is a prism. If 푢 0 = 푣 푡 then 푣 0 = 푢 푡 and 퐺 is a M¨obius ladder (Figure 2.7 depicts the prism P10 and the M¨obius ladder M10 ). In both alternatives, 푢 푡 푣 푡 is also removable in 퐺. Case 2 Neither 푢 1 푣 1 nor 푢 푡 푣 푡 is removable in 퐺. By Lemma 18.6, with 푢 2 푣 2 playing the role of 푒, we have that 퐺 has a vertex, 푤 1 , that, together with vertices 푢 1 , 푢 2 , 푣 1 , 푣 2 , span a house, where 푢 1 푣 1 is the ceiling and 푢 2 푣 2 the floor. Likewise, 퐺 has a vertex, 푤 푡 , which is adjacent to both 푢 푡 and 푣 푡 . Graph 퐺, a brick or a brace, is 3-connected. Therefore, 푉 (퐺) = 푉 (푃) ∪ 푉 (푄) ∪ {푤 1 , 푤 푡 }. Moreover, as 퐺 has an even number of vertices, it follows that 푤 1 and 푤 푡 are distinct. Finally, 푤 1 and 푤 푡 must be adjacent, in order to have minimum degree three. We conclude that 퐺 is a staircase (Figure 2.8(b) depicts the staircase S10 ). ƒ

18.3.4 The characterization of G ♯ The following result combines the results of McCuaig [68] and Norine and Thomas [78].

18.4 Generating Simple Braces and Bricks

411

Theorem 18.13 Let 퐺 be a brace on six or more vertices or a brick. If 퐺 has no strictly thin edge then 퐺 is either the Petersen graph or is an odd wheel, a biwheel, a truncated biwheel, a prism, a M¨obius ladder or a staircase. Proof Suppose that 퐺 has no thin edges. If 퐺 is a brace then 퐺 is either 퐾2 or 퐶4 , both braces with fewer than six vertices. Thus, 퐺 is a brick, hence 퐺 is either 퐾4 , 퐶6 or the Petersen graph. The graph 퐾4 is an odd wheel and 퐶6 a prism (and a truncated biwheel). We may thus assume that 퐺 has thin edges. If index(퐺) = 3 then, by Lemma 18.8, every thin edge of 퐺 is strictly thin. Thus, index(퐺) ≤ 2. As 퐺 has no strictly thin edges, its thin edges have positive index. Thus, index(퐺) ∈ {1, 2}. The only simple brace on six vertices is 퐾3,3 , a M¨obius ladder. Therefore we may assume that if 퐺 is a brace then it has eight or more vertices. Every brace on eight vertices other than the cube has a thin edge of index zero. The cube is a prism. We may thus assume that if 퐺 is a brace then 퐺 has ten vertices or more. If index(퐺) = 1 then, by Theorem 18.10, 퐺 is either an odd wheel, a biwheel or a truncated biwheel. If index(퐺) = 2 then, by Theorem 18.12, 퐺 is either a prism, a M¨obius ladder or a staircase. ƒ

Exercises ⊲18.3.1 Why, in the statement of Lemma 18.7, it is required that the brace have ten or more vertices? And in Lemma 18.9? And in Theorem 18.10?

18.4 Generating Simple Braces and Bricks We conclude the chapter with a brief observation concerning the generation procedures for simple braces and bricks. These involve the four expansion operations described in Chapter 17. Suppose that 퐺 is a simple brick and 푒 is a strictly thin edge of 퐺, and 퐻, the retract of 퐺 − 푒, is a simple brick. If index(푒) = 0 then 퐺 = 퐻 + 푒. If index(푒) = 1 then 퐺 is an expansion of 퐻 by a barrier of size two. If index(푒) = 2 then 퐺 is an expansion of 퐻 by two barriers of size two. Finally, if index(푒) = 3 then 퐺 is an expansion of 퐻 by a barrier of size three. The expansion operations for bricks were described in Section 17.4. Theorem 18.13 now implies the following result of Norine and Thomas (2007, [78]). Theorem 18.14 Given any simple brick 퐺, there exists a sequence 퐺 1 , 퐺 2 , . . . , 퐺 푘 of simple bricks such that: (i) 퐺 1 is either an odd wheel, or a truncated biwheel, or a prism, or a M¨obius ladder, or a staircase, or the Petersen graph, and 퐺 푘 = 퐺, and (ii) for 2 ≤ 푖 ≤ 푘, 퐺 푖 is obtained from 퐺 푖−1 by an expansion. ƒ

412

18 Strictly Thin Edges in Bricks and Braces

With the exception of the expansion by a barrier of size three, all the other expansions have similar expansions for braces. The expansion operations for braces were described in Section 17.7. Theorem 18.13 also implies the following result of McCuaig (2001,[68]). Theorem 18.15 Given any simple brace 퐺 of order six or more, there exists a sequence 퐺 1 , 퐺 2 , . . . , 퐺 푘 of simple braces such that: (i) 퐺 1 is either a biwheel, or a prism, or a M¨obius ladder, and 퐺 푘 = 퐺, and (ii) for 2 ≤ 푖 ≤ 푘, 퐺 푖 is obtained from 퐺 푖−1 by an expansion. ƒ

Exercises 18.4.1 Using Theorem 18.15, show that every brace of order six or more has a matching minor which, up to isomorphism, is either 퐾3,3 (which is a M¨obius ladder) or the cube (which is both a prism and a biwheel).

18.5 Splitter Versions As noted earlier, Norine and Thomas [78] and McCuaig [68] proved generalizations of Theorems 18.14 and 18.15, respectively. They refer to those generalizations as ‘splitter-versions’, as they are motivated by a theorem on 3-connected graphs, due to P.D. Seymour, which is known as the splitter theorem. For the sake of clarity, we shall restrict ourselves here to bricks. Theorem 18.16, below, is the splitter-version of Theorem 18.14. Recall (Chapter 12) that a matching covered graph 퐻 is a matching minor of another matching covered graph 퐺 if it can be obtained from a conformal subgraph of 퐺 by means of a sequence of bicontractions of vertices of degree two. For example, by Theorem 12.13, every brick contains either 퐾4 or 퐶6 as a matching minor. In order to be able to state the generalization of Theorem 18.14 proved by Norine and Thomas [78], we need to define a new class of graphs. A prismoid is either a truncated biwheel or one obtained from it by adding an edge joining its two hubs. Theorem 18.16 Let 퐺 be a brick which is not an odd wheel, a prismoid, a prism, a M¨obius ladder, a staircase, or the Petersen graph, and let 퐻 ∉ {퐾4 , 퐶6 } be a simple brick that is a matching minor of 퐺. Then there there exists a sequence 퐺 1 , 퐺 2 , . . . , 퐺 푘 of simple bricks distinct from the Petersen graph such that: (i) 퐺 1 = 퐻, 퐺 푘 = 퐺, and (ii) for 2 ≤ 푖 ≤ 푘, 퐺 푖 is obtained from 퐺 푖−1 by an expansion. ƒ An analogue of the above theorem for braces, due to McCuaig [68] is also described in [78]. Another proof of McCuaig’s splitter theorem for braces was given by CLM [18]. Unfortunately, a proof of the more general theorem stated above is beyond the scope of this book.

18.6 Notes

413

18.6 Notes Kothari [47] proved that every simple near-bipartite brick 퐺, except for those in eight infinite families, has a strictly thin edge 푒 such that the retract of 퐺 − 푒 is also near-bipartite. (If one wants to fix the removable doubleton and preserve it, then 11 infinite families are required.) This theorem and its proof also appear in Kothari and Carvalho (2020, [49]).

Part III

Pfaffian Orientations

Chapter 19

Pfaffian Orientations

Contents 19.1 19.2

19.3 19.4

The Pfaffian Orientation Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 417 Equivalent Formulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 420 19.2.1 Oddly oriented cycles and conformal subgraphs . . . . . . . . . 421 19.2.2 An application: the even directed cycle problem . . . . . . . . . 422 19.2.3 Non-Pfaffian graphs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 423 Pfaffian Orientations of Planar Graphs . . . . . . . . . . . . . . . . . . . . . . . . 432 19.3.1 Odd orientations of plane graphs . . . . . . . . . . . . . . . . . . . . . 432 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 434

19.1 The Pfaffian Orientation Problem This chapter is dedicated mainly to some interesting questions arising from Tutte’s original proof of his theorem, presented in Section 1.5, using Pfaffian identities. For the convenience of the readers, we recall briefly the relevant definitions. According to Cayley’s Theorem (1.9), the determinant of a skew-symmetric matrix A = (푎 푖 푗 ) is the square of a polynomial in the entries of A, called the Pfaffian of A. As the determinant of a skew-symmetric matrix of odd order is always zero, we restrict our attention here to graphs of even order. Thus let 퐺 be a graph of order 2푘 and let 퐷 be an orientation of 퐺. Then the adjacency matrix A = (푎 푖 푗 ) of 퐷 is a skew-symmetric matrix. Let 푀 := {푒 1 , 푒 2 , . . . , 푒 푘 } be any perfect matching of 퐺, and suppose that in the orientation 퐷 of 퐺, the tail and the head of the edge 푒 푖 , for 1 ≤ 푖 ≤ 푘, are 푢 푖 and 푣 푖 , respectively. We may then associate with 푀 the permutation 휋(푀) as defined below:   1 2 3 4 . . . 2푘 − 1 2푘 . 휋(푀) := 푢 1 푣 1 푢 2 푣 2 . . . 푢 푘 푣 푘 © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7_19

417

418

19 Pfaffian Orientations

The sign of 휋(푀) is equal to (−1) 푟 , where 푟 is the number of inversions of 휋(푀). The sign of 휋(푀) does not depend on the order in which the edges of 푀 are listed, but does depend on the orientation 퐷 of 퐺. We define the sign of 푀, denoted sign(푀), to be the sign of 휋(푀). The Pfaffian Pf (A) of the adjacency matrix A is then defined as the following polynomial in the entries of A: Õ Pf (A) := (19.1) sign(푀) 푎 푢1 푣1 푎 푢2 푣2 . . . 푎 푢푘 푣푘 where the sum is taken over all 푀 in M, the set of perfect matchings of 퐺. Since, by the adopted convention that, for 1 ≤ 푖 ≤ 푘, vertex 푢 푖 is the tail and 푣 푖 is the head in 퐷 for any edge 푒 푖 of 푀, it follows that the product 푎 푢1 푣1 푎 푢2 푣2 . . . 푎 푢푘 푣푘 is always equal to one. Therefore, if the signs of all perfect matchings of 퐺 with respect to an orientation 퐷 were the same, then the absolute value of Pf(A) would simply be the number of perfect matchings of 퐷. In this case, that number could be computed efficiently by computing the determinant of A and taking its square root. Motivated by this observation, we now introduce the concept that is the object of our main focus in this part of the book. Pfaffian digraphs and Pfaffian graphs A digraph is Pfaffian if all its perfect matchings have the same sign. In the same vein, an (undirected) graph is Pfaffian if it admits an orientation that is Pfaffian. We shall regard every digraph as an orientation of its underlying (undirected) graph. The same graph may admit one orientation that is Pfaffian and one which is not. For example, the second orientation of 퐾4 shown in Figure 1.8 is a Pfaffian orientation, but the first orientation is not. Although our main interest is in orientations of matching covered graphs, we shall find it convenient to consider orientations of arbitrary undirected graphs which may not even be matchable. (Any orientation of an undirected graph with at most one perfect matching is a Pfaffian orientation!) It should be noted that the signs of perfect matchings of a digraph do depend on the chosen enumeration of its vertices. However, the property of the digraph being Pfaffian or non-Pfaffian is independent of that enumeration. More specifically, two isomorphic digraphs are either both Pfaffian or both non-Pfaffian. (See Exercise 19.1.1.) The above definitions lead to the following important decision problems: Problem 19.1 (The Pfaffian recognition problem) Given: A digraph 퐷; Decide: whether 퐷 is Pfaffian. Problem 19.2 (The Pfaffian orientation problem) Given: A graph 퐺; Decide: whether 퐺 admits a Pfaffian orientation.

19.1 The Pfaffian Orientation Problem

419

Both these problems are known to be ‘difficult’ in general. However, we shall see that both of them are in co-NP and are polynomial-time equivalent. (The first proof of these results, due to Vazirani and Yanakakis, appeared in 1989 [91]. The CLM proof, using a different approach, appeared in 2005 [14]. This proof is presented in Section 20.4.) But in some special classes of graphs (such as planar graphs and bipartite graphs), polynomial-time algorithms have been devised to resolve these problems. These discoveries have some surprising and spectacular applications in the theory of algorithms. As noted before, the idea of using Pfaffians in matching theory goes back to Tutte (1947, [89]). In his mathematical autobiography Graph Theory as I have known it (1998, [90]), he describes the thought process that led him to the discovery of the relationship between Pfaffians and perfect matchings in graphs. The notion of Pfaffian orientations is, however, due to Kasteleyn (1963, [44]). He established the fundamental result that every planar graph has a Pfaffian orientation and used it to count the number of perfect matchings in certain classes of planar graphs. His work was motivated by problems in statistical physics. We shall present a proof of Kasteleyn’s Theorem in Section 19.3.

Exercises 19.1.1 Let 퐷 and 퐷 ′ be two isomorphic digraphs on {1, 2, . . . , 2푛 − 1, 2푛}, and let 휎 denote an isomorphism from 퐷 to 퐷 ′ . Show that there exists a constant 푐 ∈ {1, −1} such that for any perfect matching 푀 of the underlying graph of 퐷, sign(휋 퐷 ′ (휎(푀))) = 푐 · sign(휋 퐷 (푀)). 19.1.2 The adjacency matrix of the second orientation of 퐾4 , shown in Figure 1.8 is:    0 +1 +1 +1   −1 0 −1 +1     −1 +1 0 −1       −1 −1 +1 0 

Verify that the determinant of this matrix is nine, which is the square of the number of perfect matchings of 퐾4 . 19.1.3 Is the orientation of 퐾3,3 shown in Figure 19.1 a Pfaffian orientation? (Hint: Consider signs of the two perfect matchings 푀1 := {12, 34, 56} and 푀2 := {12, 63, 45}.) ⊲19.1.4 Let 푅 be a removable class of a matching covered graph 퐺 and let 퐷 be an orientation of 퐺. Suppose that 퐷 − 푅 is Pfaffian, and all perfect matchings including 푅 have equal sign. Prove that 퐺 is Pfaffian.

420

19 Pfaffian Orientations 1

3

5

2

4

6

Fig. 19.1 An orientation of 퐾3,3

19.2 Equivalent Formulations In this section we shall present several equivalent definitions of Pfaffian orientations in purely graph theoretical terms by using the terminology and notation described in the following inset: Oddly and evenly oriented trails Let 퐷 be an orientation of a graph 퐺 and let 푇 be a trail of 퐺. An arc of 푇 is a forward arc of 푇 if it is traversed from tail to head; and is a reverse arc of 푇 if it is traversed from head to tail. The set of forward arcs of 푇 is denoted fw퐷 (푇) and the set of reverse arcs of 푇 is denoted rv퐷 (푇). (If 퐷 is understood, the subscript 퐷 is omitted.) We say that 푇 is oddly oriented if |fw(푇)| is odd and is evenly oriented if |fw(푇)| is even. Note that |fw(푇)| + |rv(푇)| is equal to the length of 푇. Consequently, if 푇 has even length then |fw(푇)| ≡ |rv(푇)| (mod 2). In that case, 푇 is oddly oriented if and only its reverse is oddly oriented. In particular, this property holds for every even cycle and every path of even length of 퐷. In the first orientation of 퐾4 shown in Figure 1.8, the cycle (1, 3, 4, 2, 1) is oddly oriented and the cycle (1, 3, 2, 4, 1) is evenly oriented. The characterization of Pfaffian orientations of even cycles is a very important simple result, which has a fundamental, easily proved consequence (Exercises 19.2.1 and 19.2.2). Proposition 19.3 Let 퐷 be a digraph whose underlying graph is an even cycle, 푄. The digraph 퐷 is Pfaffian if and only if 푄 is oddly oriented in 퐷. ƒ Corollary 19.4 Let 퐷 be a digraph, let 푀1 and 푀2 be two perfect matchings of 퐺 and let 푘 denote the number of evenly oriented (푀1 , 푀2 )-alternating cycles in 퐷. Then, sign(푀1 ) · sign(푀2 ) = (−1) 푘 . ƒ

19.2 Equivalent Formulations

421

19.2.1 Oddly oriented cycles and conformal subgraphs Recall that a matchable subgraph 퐻 of a graph 퐺 is conformal if 퐺 − 푉 (퐻) is matchable. For example, a 6-cycle in the Petersen graph is not conformal, but an 8-cycle is; see Figure 19.2.

Fig. 19.2 Even cycles in the Petersen graph

Lemma 19.5 Let 퐺 be a graph, let 퐷 be an orientation of 퐺 and let 푀 be a perfect matching of 퐺. The following assertions are equivalent: (i) the orientation 퐷 of 퐺 is Pfaffian, (ii) for every conformal subgraph 퐻 of 퐺, the restriction 퐷 퐻 of 퐷 to 퐻 is Pfaffian, (iii) every conformal cycle of 퐷 is oddly oriented, (iv) every 푀-alternating cycle of 퐷 is oddly oriented. Proof (i)⇒(ii): Suppose that 퐷 is a Pfaffian orientation of 퐺, let 퐻 be a conformal subgraph of 퐺, let 퐷 퐻 be the restriction of 퐷 to 퐻, and let 푁1 and 푁2 be two perfect matchings of 퐻. As 퐻 is conformal, 퐺 − 푉 (퐻) has a perfect matching, 푁 ′ . For 푖 = 1, 2, let 푀푖 := 푁푖 ∪ 푁 ′ . Then, 푀1 and 푀2 are perfect matchings of 퐺. As 퐷 is a Pfaffian orientation of 퐺, sign(푀1 ) = sign(푀2 ). We may relabel the vertices of 퐺 so that 푉 (퐻) = {1, 2, . . . , |푉 (퐻)|} (see Exercise 19.1.1). It follows that 푁1 and 푁2 have the same sign in 퐷 퐻 . We deduce that any two perfect matchings of 퐻 have the same sign in 퐷 퐻 ; hence 퐷 퐻 is a Pfaffian orientation of 퐻. (ii)⇒(iii): Assume that for every conformal subgraph 퐻 of 퐺 the restriction 퐷 퐻 of 퐷 to 퐻 is Pfaffian. In particular, for every conformal cycle 퐶 of 퐺, 퐷 퐶 is Pfaffian. By Proposition 19.3, 퐷 퐶 is oddly oriented. (iii)⇒(iv): Assume that every conformal cycle of 퐷 is oddly oriented. Every 푀alternating cycle 퐶 of 퐺 is conformal, because 푀 − 퐸 (퐶) is a perfect matching of 퐺 − 푉 (퐶). Thus, every 푀-alternating cycle of 퐷 is oddly oriented.

(iv)⇒(i): Suppose that every 푀-alternating cycle of 퐷 is oddly oriented. Let 푁 be a perfect matching of 퐺. Every (푀, 푁)-alternating cycle of 퐷 is an 푀-alternating cycle of 퐷; hence it is oddly oriented. By Corollary 19.4, sign(푁) = sign(푀). It follows that every perfect matching of 퐷 has the same sign that 푀 has; hence 퐷 is a Pfaffian orientation of 퐺. ƒ

422

19 Pfaffian Orientations

In light of the above lemma, one may deduce that the digraph 퐷 in Figure 19.1 is not Pfaffian simply from the fact that the cycle (1, 2, 3, 4, 1) is conformal and evenly oriented. This, of course, does not immediately imply that 퐾3,3 is nonPfaffian. However, 퐾3,3 is non-Pfaffian (see Exercise 19.2.11(iii)) and, indeed, it is the smallest non-Pfaffian graph. From Lemma 19.5 it is also easy to deduce the following consequence (Exercise 19.2.3). Corollary 19.6 A matching covered graph 퐺 is Pfaffian if and only if each of its ƒ conformal subgraphs is Pfaffian.

19.2.2 An application: the even directed cycle problem A variety of algorithmic problems, not necessarily in graph theory and matrix theory, are reducible to the problem of recognizing whether or not a given directed bipartite graph is Pfaffian; see the paper by McCuaig (2004, [69]). One such problem is the following: Problem 19.7 (The even directed cycle problem) Given: a digraph 퐷; Decide: whether or not 퐷 has a directed cycle of even length. The above problem may be reduced to the Pfaffian Recognition Problem as follows. Let 퐷 be any digraph. Obtain the bipartite digraph 퐵(퐷) from 퐷 as described below. • First split each vertex 푣 of 퐷 into two new vertices 푣 + and 푣 − , so that all the arcs that leave 푣 in 퐷 leave 푣 + in 퐵(퐷), and that all the arcs that enter 푣 in 퐷 enter 푣 − in 퐵(퐷). (Thus, if an arc 푎 in 퐷 leaves a vertex 푥 and enters a vertex 푦, then, in 퐵(퐷), arc 푎 leaves 푥 + and enters 푦 − .) • Then, for each 푣 ∈ 푉 (퐷), add an arc directed from 푣 + to 푣 − . The graph 퐵(퐷) is the associated bipartite digraph of 퐷. See Figure 19.3. The following assertion is left as Exercise 19.2.8: Theorem 19.8 A digraph 퐷 has an even directed cycle if and only if the associated bipartite digraph is not Pfaffian. ƒ The problem of deciding whether or not a given digraph has an odd directed cycle is easy (Exercise 19.2.13). For a long time, the complexity status of the even directed cycle problem remained unresolved. As we have seen above, that problem is equivalent to the problem of recognizing whether or not a given bipartite digraph is Pfaffian. In Chapter 21 we shall describe a polynomial-time algorithm to determine whether a bipartite graph is Pfaffian.

19.2 Equivalent Formulations

423 푣 −

푣



푒푣

푒푣 ∈ 푀

푣 + 퐷

퐵(퐷)

Fig. 19.3 The construction of 퐵(퐷)

19.2.3 Non-Pfaffian graphs How do we show that a given matching covered graph 퐺 does not admit a Pfaffian orientation, if that happens to be the case? A na¨ıve approach would be to consider all possible orientations of 퐺 and in each of those orientations exhibit a conformal cycle that is not oddly oriented. This is clearly not feasible even for moderately large graphs. Little (1973, [52]) proposed an ingenious approach in which we choose an orientation 퐷 of 퐺 and try to show that no re-orientation of 퐷 could be a Pfaffian orientation of 퐺. This may at first seem to be no easier than trying out all possible orientations. But Little showed that, when 퐺 is non-Pfaffian, there exists a collection of conformal cycles in 퐺 whose orientations in 퐷 satisfy certain parity conditions and, more significantly, such a collection of cycles serves as a certificate to the effect that no orientation of 퐺 could be a Pfaffian orientation. Intractable collections of conformal cycles Let 퐺 be a matching covered graph. A collection C of conformal cycles of 퐺 is intractable if it satisfies the following two properties: (i) every edge of 퐺 is in an even number of cycles of C, and (ii) an odd number of cycles in C are evenly oriented in some orientation 퐷 of 퐺. Proposition 19.9 If a matching covered graph 퐺 has an intractable collection C of conformal cycles then, for every orientation 퐷 of 퐺, an odd number of cycles in C is evenly oriented in 퐷. Consequently, 퐺 is not Pfaffian. Proof Suppose that 퐺 has an intractable collection C of conformal cycles, and let 퐷 0 be an orientation of 퐺 such that an odd number of cycles in C is evenly oriented in 퐷 0 . Let 퐷 be an orientation of 퐺. Clearly, there is a sequence 퐷 0 , 퐷 1 , . . . , 퐷 푟 = 퐷,

푟 ≥ 0

such that for every 푖, 0 ≤ 푖 < 푟, 퐷 푖+1 is obtained from 퐷 푖 by reversing the orientation of an edge. By definition, an odd number of cycles in C is evenly oriented in 퐷 0 .

424

19 Pfaffian Orientations

Also by definition, each edge of 퐺 is in an even number of cycles in C. A simple inductive argument then shows that C contains an odd number of cycles which are evenly oriented in 퐷 푖 , for 푖 = 0, 1, . . . , 푟. In particular, an odd number of cycles in C are evenly oriented in 퐷. We deduce that 퐷 is not a Pfaffian orientation of 퐺. This conclusion holds for each orientation 퐷 of 퐺; hence 퐺 is not Pfaffian. ƒ We now illustrate the concept of intractable collections of conformal cycles with two examples, proving that neither 퐾3,3 nor the Petersen graph is Pfaffian. Lemma 19.10 The graph 퐾3,3 is not Pfaffian. Proof Consider the graph 퐾3,3 , depicted in Figure 19.4(a), and the orientation 퐷 of 퐾3,3 , shown in Figure 19.4(b). 푣1

푣3

푣5

푣2

푣4 푣1

푣2

푣4

푣3

푣5

푣6 푣6

(푎)

(푏)

Fig. 19.4 The graph 퐾3,3 and the orientation 퐷 of 퐾3,3

The figure indicates a set C := {퐶1 , 퐶2 , 퐶3 } of three quadrilaterals, which are conformal cycles of 퐾3,3 : 퐶1 := (푣 1 , 푣 2 , 푣 3 , 푣 4 , 푣 1 ) = 퐾3,3 − 푣 5 − 푣 6 퐶2 := (푣 1 , 푣 4 , 푣 3 , 푣 6 , 푣 1 ) = 퐾3,3 − 푣 5 − 푣 2 퐶3 := (푣 1 , 푣 6 , 푣 3 , 푣 2 , 푣 1 ) = 퐾3,3 − 푣 5 − 푣 4 The three cycles of C are evenly oriented. Every edge of 퐾3,3 is in an even number of cycles in C: the three edges incident with 푣 5 are not in any cycle in C, each one of the other six edges is in precisely two of the three quadrilaterals. Thus, C is intractable; hence 퐾3,3 is not Pfaffian. ƒ Lemma 19.11 The Petersen graph is not Pfaffian. Proof Figure 19.5 depicts the orientation 퐷 of the Petersen graph. Let 푋 := {푣 0 , 푣 1 , 푣 2 , 푣 3 , 푣 4 }. The cut 휕 ( 푋) is a directed cut in 퐷; the pentagons 퐶 ′ := (푣 0 , 푣 1 , 푣 2 , 푣 3 , 푣 4 , 푣 0 ) and 퐶 ′′ := (푣 5 , 푣 6 , 푣 7 , 푣 8 , 푣 9 , 푣 5 ) are directed cycles in 퐷. Figure 19.5 also shows the octagon 퐶1 := (푣 1 , 푣 2 , 푣 9 , 푣 8 , 푣 4 , 푣 3 , 푣 6 , 푣 7 , 푣 1 ), which is a conformal cycle of P, it is equal to P − 푣 0 − 푣 5 .

425

19.2 Equivalent Formulations 푣0

푣5 푣4

푣8

푣7

푣6

푣1

푣9

푣3

푣2

Fig. 19.5 The orientation 퐷 of the Petersen graph

The cycle 퐶1 has four edges in the directed cut 휕 ( 푋); two of these edges are forward edges and two are reverse edges. The cycle 퐶1 has two edges in the cycle 퐶 ′ ; one is a forward edge; the other is a reverse edge. Likewise, it has two edges in the cycle 퐶 ′′ ; one is a forward edge; one is a reverse edge. Thus, 퐶1 is evenly oriented in 퐷. Consider now the automorphism 휑 := (푣 0 , 푣 1 , 푣 2 , 푣 3 , 푣 4 ) (푣 5 , 푣 7 , 푣 9 , 푣 6 , 푣 8 ) The orbit of 퐶1 under 휑, denoted C, consists of five octagons, listed below: 퐶1 퐶2 퐶3 퐶4 퐶5

:= := := := :=

(푣 1 , 푣 2 , 푣 9 , 푣 8 , 푣 4 , 푣 3 , 푣 6 , 푣 7 , 푣 1 ) (푣 2 , 푣 3 , 푣 6 , 푣 5 , 푣 0 , 푣 4 , 푣 8 , 푣 9 , 푣 2 ) (푣 3 , 푣 4 , 푣 8 , 푣 7 , 푣 1 , 푣 0 , 푣 5 , 푣 6 , 푣 3 ) (푣 4 , 푣 0 , 푣 5 , 푣 9 , 푣 2 , 푣 1 , 푣 7 , 푣 8 , 푣 4 ) (푣 0 , 푣 1 , 푣 7 , 푣 6 , 푣 3 , 푣 2 , 푣 9 , 푣 5 , 푣 0 )

= P − 푣 0 − 푣 5 = P − 푣 1 − 푣 7 = P − 푣 2 − 푣 9 = P − 푣 3 − 푣 6 = P − 푣 4 − 푣 8

Each one of the five octagons in C is evenly oriented. Figure 19.6 shows views of 퐶1 and 퐶2 . Each edge of 퐶 ′ is in two octagons of C. Likewise, every edge of 퐶 ′′ is in two octagons of C. Finally, every edge of 휕 ( 푋) is in four octagons of C. In sum, every edge of P is in an even number of octagons of C. We deduce that C is intractable; hence P is not Pfaffian. ƒ The fact that we were able to find intractable collections of conformal cycles for both 퐾3,3 and P is not a coincidence. The following theorem shows that this is always possible.

426

19 Pfaffian Orientations 푣0

푣1

푣2

푣3

푣4

푣5

푣7

푣9

푣6

푣8

(a) 퐶1 푣1

푣2

푣3

푣4

푣0

푣7

푣9

푣6

푣8

푣5

(b) 퐶2 Fig. 19.6 Views of 퐶1 and 퐶2

Theorem 19.12 (Little (1973, [52])) A matching covered graph 퐺 is non-Pfaffian if and only if it has an intractable collection of conformal cycles. Proof Let 퐷 be an orientation of 퐺. We have seen in Proposition 19.9 that the intractability of a collection of conformal cycles does not depend on the orientation 퐷. For each conformal cycle 퐶 of 퐺 let  0 if 퐶 is oddly oriented in 퐷, 푡(퐶) := 1 if 퐶 is evenly oriented in 퐷. Consider the following system of linear equations over GF(2), where the set of variables is {푥 푒 : 푒 ∈ 퐸 (퐺)}: Õ 푥 푒 = 푡(퐶) (one for each conformal cycle 퐶 of 퐺). (19.2) 푒∈퐸 (퐶 )

19.12.1 Suppose that the system (19.2) has a solution. Then, reversal in 퐷 of the edges in 푅 := {푒 : 푥 푒 ≡ 1 (mod 2)} yields a Pfaffian orientation 퐹 of 퐺. Proof Let 퐶 be a conformal cycle of 퐺. As the system of equations (19.2) has a solution, it follows that |푅 ∩ 퐸 (퐶)| ≡ 푡(퐶) (mod 2). That is, if 퐶 is oddly oriented in 퐷 then |푅 ∩ 퐸 (퐶)| is even and if 퐶 is evenly oriented in 퐷 then |푅 ∩ 퐸 (퐶)| is odd. We deduce that 퐶 is oddly oriented in 퐹. This conclusion holds for each conformal cycle 퐶 of 퐺; hence 퐹 is a Pfaffian orientation of 퐺. ƒ

19.2 Equivalent Formulations

427

19.12.2 Suppose that 퐺 has a Pfaffian orientation, 퐹, and let 푆 be the set of edges of 퐺 whose orientations in 퐹 and in 퐷 are distinct. Then,  1, if 푒 ∈ 푆, 푥 푒 := 0, 표푡ℎ푒푟푤푖푠푒 is a solution of the linear system (19.2). Proof Let 퐶 be a conformal cycle of 퐺. By hypothesis, 퐹 is a Pfaffian orientation of 퐺; thus, 퐶 is oddly oriented in 퐹. If 퐶 is oddly oriented in 퐷 then 푡(퐶) = 0 and |푆 ∩ 퐸 (퐶)| is even, whereas if 퐶 is evenly oriented in 퐷 then 푡(퐶) = 1 and |푆 ∩ 퐸 (퐶)| is odd. In both alternatives, |푆 ∩ 퐸 (퐶)| ≡ 푡(퐶) (mod 2). In other words, Í ƒ 푒∈퐸 (퐶 ) 푥 푒 ≡ 푡(퐶) (mod 2). Indeed 푥 is a solution to the system (19.2). It follows that 퐺 is non-Pfaffian if and only if the system (19.2) has no solution. By the Theorem of Rouch´e-Capelli, the system has no solution if and only if the rank in GF(2) of the coefficient matrix is smaller than the rank in GF(2) of the augmented matrix. Thus, the system has no solution if andÍonly if 퐺 has a collection C of conformal cycles such that △퐶 ∈ C 퐸 (퐶) = ∅ and 퐶 ∈ C 푡(퐶) ≡ 1 (mod 2). That is, the system has no solution if and only if 퐺 has an intractable collection of conformal cycles. In other words, 퐺 is not Pfaffian if and only if it has an intractable collection of conformal cycles. ƒ

In Section 20.4 we shall show that given any matching covered graph 퐺 one can find an orientation 퐷 of 퐺, which we refer to as a characteristic orientation of 퐺, with the property that 퐺 is Pfaffian if and only if 퐷 is Pfaffian. Thus, if 퐺 happens to be non-Pfaffian, it suffices to show that 퐷 is non-Pfaffian. In order to do this, it suffices to point to a conformal cycle in 퐺 that is evenly oriented in 퐷.

Exercises ∗ 19.2.1 Prove Proposition 19.3. Hint: (i) enumerate the 2푛 vertices of 푄 so that 푄 = 1, 2, . . . , 2푛, 1 and let 푀1 and 푀2 be the two perfect matchings of 푄; (ii) verify that if rv(푄) is empty then 푀1 and 푀2 have distinct signs; (iii) prove that sign(푀1 ) · sign(푀2 ) = (−1) |rv(푄) |+1 , by induction on |rv(푄)|. 19.2.2 Deduce Corollary 19.4 from Proposition 19.3. ⊲19.2.3 Deduce Corollary 19.6 from Lemma 19.5. ⊲19.2.4 Let 퐺 be a matchable, but not necessarily matching covered graph. Let 푁 be the set of nonmatchable edges of 퐺 and let 퐻 := 퐺 − 푁. Prove the following properties: (i) every component of 퐻 is matching covered, (ii) an orientation 퐷 of 퐺 is Pfaffian if and only if its restriction to each component of 퐻 is Pfaffian, and

428

19 Pfaffian Orientations

(iii) the graph 퐺 is Pfaffian if and only if each component of 퐻 is Pfaffian. Hint: use Lemma 19.5. ⊲19.2.5 Figure 19.7 depicts an orientation of 푊5 − 푒 obtained by deleting a spoke 푒 from the 5-wheel 푊5 . By verifying that each conformal even cycle in 푊5 − 푒 is oddly oriented, deduce that this orientation is a Pfaffian orientation of 푊5 − 푒. Then show that there is precisely one way of extending the given orientation of 푊5 − 푒 to a Pfaffian orientation of 푊5 .

푒

Fig. 19.7 Pfaffian orientations of 푊5

⊲19.2.6 Consider the conformal 4-cycle 퐶4 := (1, 2, 3, 4, 1) in 퐾4 shown in Figure 19.8 with the indicated orientation which makes it an oddly oriented cycle. Find all the ways in which this orientation of 퐶4 may be extended to a Pfaffian orientation of 퐾4 .

1

2

4

3

Fig. 19.8 Pfaffian orientations of 퐾4

429

19.2 Equivalent Formulations

⊲19.2.7 Figure19.9 depicts 퐶6 and two orientations, 퐷 0 and 퐷. Show that (i) both these are Pfaffian orientations of 퐶6 , (ii) both 퐶-contractions of 퐷 are Pfaffian orientations of 퐾4 and that this is not the case with the orientation 퐷 0 . 퐶

푣1

푣2

푢1

푢2

푤1

퐶

푣1

푣2

푢1

푤2 퐷0

푢2

푤1

푤2 퐷

Fig. 19.9 Two Pfaffian orientations of 퐶6

⊲19.2.8 Prove Theorem 19.8. Hint: 푀 := {푣 + 푣 − : 푣 ∈ 푉 (퐷)} is a perfect matching of 퐵(퐷) and every perfect matching of 퐵(퐷) has the same sign of 푀 if and only if 퐷 has no directed cycle of even length. 19.2.9 The graph shown in Figure 19.10, known as the Heawood graph, is the smallest cubic graph of girth six. It has no conformal cycles of length 0 (mod 4). Consider the orientation of the Heawood graph in which all edges are oriented downwards. Show that this orientation is Pfaffian.

Fig. 19.10 The Heawood graph

19.2.10 Consider the digraph 퐷, known as the Koh-Tindell digraph, shown in Figure 19.11. Show that 퐵(퐷) is a Pfaffian orientation of the Heawood graph (see Exercise 19.2.9). Deduce that the Koh-Tindell digraph has no directed even cycles. ⊲19.2.11 (C. H. C. Little) Consider the graph 퐺 := 퐾2,푘 , where 푘 is an integer, 푘 ≥ 3. Let 푤 1 and 푤 2 be the two vertices of degree 푘, let 푣 1 , 푣 2 , . . . , 푣 푘 be the 푘 vertices of degree two of 퐺. For 푖 = 1, 2, . . . , 푘, let 푄 푖 be the quadrilateral induced by the set {푣 푖 , 푣 푖+1 , 푤 1 , 푤 2 }, where the indices are taken modulo 푘. Figure 19.12 depicts the particular case in which 푘 = 4.

430

19 Pfaffian Orientations

Fig. 19.11 The Koh-Tindell digraph 푤2

푣1

푄1

푣2 푄2

푣3

푄3

푣4

푄4

푤1 Fig. 19.12 Graph 퐾2,4

(i) Prove that for every orientation of 퐺 the number 푟 of oddly oriented quadrilaterals in {푄 푖 : 1 ≤ 푖 ≤ 푘} is even. Hint: observe that reversal of the orientation of any edge preserves the parity of 푟; hence the parity of 푟 does not depend on the orientation of 퐺; consider then the orientation in which 푤 1 is a source and 푤 2 is a sink. (ii) In Chapter 3, we defined a bisubdivision of a graph 퐺 to be a graph obtained from 퐺 by replacing each edge by a path of odd length (possibly equal to one). Suppose that a graph 퐺 has a subgraph 퐻 which is a bisubdivision of 퐾2,3 . Then, 퐻 has three quadrilaterals 푄 푖 , 푖 = 1, 2, 3, which are bisubdivisions of 퐶2 .

19.2 Equivalent Formulations

431

Show that if each of the three cycles is a conformal subgraph of 퐺 then 퐺 is not Pfaffian. (iii) Use part (ii) to prove that 퐾3,3 is not Pfaffian. ⊲19.2.12 (C. H. C. Little (1973, [52])) Let 퐺 be a matching covered graph and let 퐷 be an orientation of 퐺. A collection C of perfect matchings of 퐺 is 퐷-intractable if it satisfies the following two properties: (i) every edge of 퐺 is in an even number of perfect matchings in C, and (ii) an odd number of perfect matchings in C has sign -1 in 퐷 (or sign 1). Property (i) implies that |C| is even, because for any vertex 푣 of 퐺, |C| is equal to the sum, over all edges 푒 of 휕 (푣), of the number of perfect matchings in C that contain edge 푒; property (i) also implies that property (ii) does not depend on the orientation 퐷. For each perfect matching 푀 of 퐺, let  0 if sign(푀) = 1, 푡(푀) := 1 if sign(푀) = −1. Consider the following system of linear equations over GF(2), where the set of variables is {푥 푒 : 푒 ∈ 퐸 (퐺)}: Õ 푥 푒 = 푡(푀) (one for each perfect matching 푀 of 퐺). (19.3) 푒∈ 푀

(i) Prove that the intractability of C does not depend on the orientation 퐷, that is, prove that C is 퐷-intractable if and only if it is 퐹-intractable, for each orientation 퐹 of 퐺. (ii) Show that if the system (19.3) has a solution then reversal in 퐷 of the edges 푒 such that 푥 푒 ≡ 1 (mod 2) yields a Pfaffian orientation of 퐺. (iii) Prove that if 퐺 has a Pfaffian orientation, 퐹, then  1 if the orientations of 푒 in 퐷 and 퐹 are distinct, 푥 푒 := 0 표푡ℎ푒푟푤푖푠푒 is a solution of the linear system (19.3). (iv) Conclude that a graph is non-Pfaffian if and only if it has an intractable collection of perfect matchings. Hint: apply the criterion for the solvability of linear equations to show that the system has no solution if and Í only if 퐺 has a collection C of perfect matchings such that △ 푀 ∈ C 푀 = ∅ and 푀 ∈ C 푡(푀) ≡ 1 (mod 2).

19.2.13

(i) Show that a digraph has a directed cycle of odd length if and only if it has a closed directed walk of odd length. (ii) Describe a polynomial-time algorithm for finding a directed closed walk of odd length, if it has one.

432

19 Pfaffian Orientations

19.3 Pfaffian Orientations of Planar Graphs In this section we give a proof of Kasteleyn’s Theorem which states that every planar graph is Pfaffian [44]. As our main interest is in Pfaffian orientations of matching covered graphs, we shall restrict ourselves to 2-connected graphs. A plane graph is a planar graph that is embedded in the plane. We shall make use of the classical result of Whitney (1932, [93]) which says that in any 2-connected plane graph 퐺 all faces are bounded by cycles ([3, Theorem 10.7]). Any cycle in 퐺 which bounds a face is a facial cycle. Given an embedding 퐺 of a planar graph, there is one face in that embedding which is unbounded. That face is referred to as the outer face, and all other faces as inner faces. In general, the bounding cycle of the outer face in one embedding 퐺 of a planar graph need not be the bounding cycle of the outer face in a different embedding of the same planar graph (see Exercise 19.3.1). However, given any planar embedding 퐺 of a planar graph and an inner face 푓 of 퐺, it is possible to find another planar embedding 퐺˜ of that planar graph such that all facial cycles of 퐺 are ˜ (See also facial cycles of 퐺˜ and, in addition, the image 푓˜ of 푓 is the outer face of 퐺. Theorem 9.3 in [2].) In order to prove Kasteleyn’s Theorem, we shall first establish that every plane graph admits a special type of orientation, which we refer as an odd orientation, and then show that any odd orientation of a matching covered plane graph is a Pfaffian orientation of that graph.

19.3.1 Odd orientations of plane graphs Let 퐺 be a 2-connected plane graph. We shall adopt the convention that every cycle of 퐺 is traversed in the clockwise order, and an orientation of a cycle 퐶 of 퐺 is odd if there are an odd number of edges of the cycle which are oriented in that direction, and is even otherwise. Odd orientations An orientation 퐷 of a plane graph 퐺 is an odd orientation if all the facial cycles are odd, with the possible exception of the cycle bounding the outer face. We shall need the following strengthening of the ear decomposition theorem [3, Theorem 5.8] to plane graphs. Proposition 19.13 A 2-connected plane graph 퐺 has a sequence (퐺 0 , 퐺 1 , 퐺 2 , . . . , 퐺 푟 ) of 2-connected plane subgraphs satisfying the following conditions: (i) 퐺 푟 = 퐺 and 퐺 0 is a facial cycle of 퐺; (ii) for 0 ≤ 푖 < 푟, there is a path 푃푖 which has its ends in 푉 (퐺 푖 ), is internally-disjoint from 퐺 푖 , and lies in the outer face of 퐺 푖 , such that 퐺 푖+1 = 퐺 푖 ∪ 푃푖 .

19.3 Pfaffian Orientations of Planar Graphs

433

Lemma 19.14 Every plane graph admits an odd orientation. Proof by induction on the number of edges. Let (퐺 0 , 퐺 1 , 퐺 2 , . . . , 퐺 푟 ) be the sequence defined in the statement of Proposition 19.13. The assertion is obviously true if 퐺 = 퐺 0 is itself a cycle. Let 푖 be an integer such that 0 < 푖 < 푟, and assume inductively that 퐺 푖 has an odd orientation 퐷 푖 . We shall now show how to find an odd orientation 퐷 푖+1 of 퐺 푖+1 . Note that 퐺 푖+1 is obtained from 퐺 푖 by adding a path 푃 joining two vertices on the bounding cycle of the outer face 퐹. The addition of 푃 divides the face 퐹 into two regions which become faces of 퐺 푖+1 . One of these faces is a face of 퐺, bounded by a cycle, 푄. The other face is the outer face of 퐺 푖+1 . See Figure 19.13.

퐹

푄 푃

퐺푖

퐺푖+1

Fig. 19.13 The outer face 퐹, the path 푃 and the cycle 푄

Extend 퐷 푖 to an orientation 퐷 푖+1 of 퐺 푖+1 by orienting the edges of 푃 such that 푄 is oddly oriented. Clearly, 퐷 푖+1 is an odd orientation of 퐺 푖+1 . The assertion is valid, by induction. ƒ Note that underlying the above inductive proof there is an algorithm for finding an odd orientation of any given 2-connected plane graph. Kasteleyn’s Theorem Theorem 19.15 Any odd orientation of a 2-connected matching covered plane graph is Pfaffian. Proof Let 퐺 be a matching covered plane graph and consider an odd orientation 퐷 of 퐺 (with the adopted convention that each cycle is traversed clockwise). By the definition of odd orientation, if 푄 is any cycle of 퐺, then the bounding cycles of all the faces in the interior of 푄 are oddly oriented in 퐷.

434

19 Pfaffian Orientations

19.15.1 Let 푄 be any cycle of 퐺 and let 푝 be the number of vertices in the interior of 푄. Then, |fw(푄)| ≡ 푝 + 1 (mod 2), where fw(푄) is the number of forward arcs of 푄. Proof Let F , 푞 and ℓ denote, respectively, the collection of faces inside 푄, the number of arcs in the interior of 푄 and the length of 푄. For each face 퐹 in F , let 푄 퐹 denote the bounding cycle of 퐹. Each one of the 푞 arcs in the interior of 푄 is traversed twice by the bounding cycles of the faces in F , once in each direction. Moreover, by hypothesis, |fw(푄 퐹 )| is odd, for each face 퐹 ∈ F . Thus, Õ |fw(푄)| = (19.4) |fw(푄 퐹 )| − 푞 ≡ |F | − 푞 (mod 2). 퐹 ∈ F

By Euler’s formula, (|F | + 1) + ( 푝 + ℓ) − (푞 + ℓ) = 2, hence 푝 + 1 ≡ |F | − 푞 (mod 2). From this and equation (19.4), we conclude that ƒ |fw(푄)| ≡ 푝 + 1 (mod 2), as asserted. By hypothesis, graph 퐺 is matching covered and 2-connected. Let 푀 be a perfect matching of 퐺. Let 푄 be an 푀-alternating cycle. Clearly, every vertex in the interior of 푄 is matched by an edge of 푀 to another vertex in the interior of 푄. It follows that the number 푝 of vertices of 퐺 in the interior of 푄 must be even. By statement 19.15.1, the cycle 푄 is oddly oriented. This conclusion holds for each 푀-alternating cycle 푄 ƒ of 퐺. From Lemma 19.5 we infer that the odd orientation 퐷 of 퐺 is Pfaffian. The following more general assertion may be easily derived from the above lemma (Exercise 19.3.3). Corollary 19.16 Every planar graph, regardless of whether it is matching covered or not, is Pfaffian. ƒ

Exercises 19.3.1 Find a planar embedding of the pentagonal prism in which the outer face is a square and find a Pfaffian orientation of the pentagonal prism. 19.3.2 Supply the missing details in the proof of Lemma 19.14. 19.3.3 Give a proof of Corollary 19.16.

19.4 Notes There is a generalization of the notion of Pfaffian graphs, which is the notion of 푘-Pfaffian graphs. Let 퐺 be a graph, let M be the collection of perfect matchings of 퐺.

435

19.4 Notes

A 푘-orientation of 퐺 is a sequence D := (퐷 1 , 퐷 2 , . . . , 퐷 푘 ) of 푘 orientations of 퐺. For each perfect matching 푀 ∈ M, the sign of 푀 in D is the line vector signD (푀) := (sign퐷1 (푀), sign퐷2 (푀), . . . , sign퐷푘 (푀)). Let signD be the |M | × 푘 matrix, in which each line is the sign vector in D of a perfect matching of 퐺. The 푘-orientation D is Pfaffian if the system signD x = 1 has a solution, where x is a column vector of 푘 rows and 1 is a column vector of |M | 1’s. The graph 퐺 is 푘-Pfaffian if it has a Pfaffian 푘-orientation. Clearly, 퐺 is 1-Pfaffian if and only if 퐺 is Pfaffian. The graph 퐾3,3 , for example, is 4-Pfaffian. The Pfaffian number of a graph 퐺 is the smallest integer 푘 such that 퐺 is 푘-Pfaffian. Galluccio and Loebel (1999, [37]) and, independently, Tesler (2000, [88]) proved that if a graph is embeddable in an orientable surface of genus 푔 then its Pfaffian number is at most 4푔 . Norine (1999, [77]) proved that a graph is 4-Pfaffian if and only if it has an immersion in the torus in which every perfect matching has an even number of crossings. In that same article, Norine proved that every 5-Pfaffian graph is 4-Pfaffian and every 3-Pfaffian graph is Pfaffian. Norine also conjectured that the Pfaffian number of every graph is a power of four. Miranda and Lucchesi (2011, [75]) found a counterexample, a graph whose Pfaffian number is six. There are interesting open questions about this generalization. For example, we may think of the problems analogous to Problems 19.1 and 19.2. It is not even known whether these generalizations are polynomial-time equivalent.

Chapter 20

Similarity of Orientations and Characteristic Orientations

Contents 20.1

20.2 20.3

20.4

20.5

Similarity of Orientations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 437 20.1.1 Similarity and Pfaffian orientations . . . . . . . . . . . . . . . . . . . . 439 20.1.2 Similarity and removable ears . . . . . . . . . . . . . . . . . . . . . . . . 439 Pfaffian Orientations and Tight Cuts . . . . . . . . . . . . . . . . . . . . . . . . . . 442 20.2.1 Pfaffian orientations and separating cuts . . . . . . . . . . . . . . . 442 The Number of Similarity Classes . . . . . . . . . . . . . . . . . . . . . . . . . . . . 449 20.3.1 An upper bound for 휏(퐺) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 450 20.3.2 A lower bound for 휏(퐺) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 451 Characteristic Orientations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 452 20.4.1 Pfaffian extensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 453 20.4.2 Finding characteristic orientations . . . . . . . . . . . . . . . . . . . . 454 20.4.3 The equivalence of Problems 19.1 and 19.2 . . . . . . . . . . . . . 455 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 459

20.1 Similarity of Orientations In this section we introduce an equivalence relation on the set of all orientations of a graph. The study of equivalence classes with respect to this relation turns out to play an important role in developing algorithms for finding Pfaffian orientations of matching graphs which are known to be Pfaffian and in showing that the Pfaffian Recognition Problem (19.1) and the Pfaffian Orientation Problem (19.2) are both in co-NP and are polynomial-time equivalent.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7_20

437

438

20 Similarity of Orientations and Characteristic Orientations

Similar orientations Let 퐷 be a digraph, let 푆 be a set of arcs of 퐷. We denote by 퐷 rev 푆 the digraph obtained from 퐷 by reversing the orientations of the edges of 푆. Two orientations 퐷 1 and 퐷 2 of a graph 퐺 are said to be similar if 퐺 has a cut 퐶 such that 퐷 2 = 퐷 1 rev 퐶. We now record some basic properties of similarity of orientations. The proofs of some of these properties are left as Exercises 20.1.1 and 20.1.2. Proposition 20.1 (i) For any two sets 푋 and 푌 of vertices of a graph, 휕 ( 푋 △ 푌 ) = 휕 ( 푋) △ 휕 (푌 ). (ii) A set 퐶 of edges of a graph 퐺 is a cut if and only if |퐶 ∩ 퐸 (푄)| is even, for every cycle 푄 of 퐺. ƒ Proposition 20.2 Similarity is an equivalence relation on the collection of orientations of a graph. ƒ

Determining whether a set of edges of a matching covered graph is a cut We now describe a linear-time algorithm to determine whether a set of edges of a graph is a cut. Determining whether a set is a cut Algorithm 20.3 Input: a connected graph 퐺 and a set 푅 of edges of 퐺; Output: either (i) a set 푋 ⊆ 푉 such that 푅 = 휕 ( 푋) or (ii) a cycle 푄 of 퐺 that contains an odd number of edges in 푅. The objective of the algorithm is to determine a 2-colouring of the vertices of 퐺 so that the ends of an edge have distinct colours if and only if the edge is in 푅. To look for such a 2-colouring, we use a depth-first-search (dfs). Traverse the graph by a dfs, while 2-colouring the vertices, and constructing a tree, 푇. • Arbitrarily choose an initial vertex as root and give to it a colour of your choice; • when adding an edge to 푇, colour the new vertex so that the edge satisfies the requirement of the 2-colouring; • when examining an edge 푒 with both ends in 푇: – if the ends of 푒 do not satisfy the requirement of the 2-colouring then return the cycle 푄 in 푇 + 푒, which can be easily retrieved from the dfs structure; clearly, 푄 has an odd number of edges in 푅; – if the ends of 푒 satisfy the requirement of the 2-colouring, continue. • if all the edges satisfy the 2-colouring requirement then return 푋, the set of vertices coloured with one of the two colours: 푅 = 휕 ( 푋). Proposition 20.4 The Algorithm 20.3 runs in linear time.

ƒ

20.1 Similarity of Orientations

439

20.1.1 Similarity and Pfaffian orientations Proposition 20.5 Let 퐷 1 and 퐷 2 be two similar orientations of a graph 퐺. Then, 퐷 1 is Pfaffian if and only if 퐷 2 is Pfaffian. Proof Let 퐶 be the cut of 퐺 such that 퐷 2 = 퐷 1 rev 퐶 and let 푀 be a perfect matching of 퐺 and let 푄 be an 푀-alternating cycle of 퐺. Reversal of any edge of 푄 changes the parity of 푄. By Proposition 20.1(ii), 푄 has an even number of edges in 퐶. Thus, reversal of all the edges of 퐸 (푄) ∩ 퐶 preserves the parity of 푄. We deduce that 푄 is oddly oriented in 퐷 1 if and only if it is oddly oriented in 퐷 2 . This conclusion holds for each 푀-alternating cycle 푄 of 퐺. By Lemma 19.5, 퐷 1 is Pfaffian if and only if 퐷 2 is Pfaffian. ƒ The above proposition implies that if an equivalence class of orientations of a graph 퐺 includes a Pfaffian orientation of 퐺, then all members of that equivalence class are Pfaffian orientations of 퐺. It turns out, as we shall see later, that if 퐺 is a Pfaffian bipartite graph then all its Pfaffian orientations belong to a single equivalence class. However, this is not true in general. See Exercise 20.1.3. Proposition 20.6 Let 퐺 be a matching covered graph and let 푣 be a vertex of 퐺. Each Pfaffian orientation of 퐺 has a similar orientation 퐷 such that 푣 is a source (or a sink) of 퐷. Proof Let 퐷 0 be a Pfaffian orientation of 퐺 and let 휕 − (푣) denote the set of arcs of 퐷 0 that enter vertex 푣. Consider a pair 푅 of parallel edges of 퐺. Clearly, 푅 induces a conformal even cycle, 푄, of 퐺. As 퐷 0 is Pfaffian, the cycle 푄 is oddly oriented in 퐹; hence the two edges of 푅 are parallel arcs in 퐷 0 . This conclusion holds for each pair of parallel edges of 퐺. In particular, it holds for each pair of parallel edges incident with 푣. Consequently, if we let 푋 be the set of tails of arcs in 휕 − (푣) and let 퐶 := 휕 ( 푋), it follows that 퐶 ∩ 휕 (푣) = 휕 − (푣). Thus, in the similar orientation 퐷 := 퐷 0 rev 퐶 of 퐷 0 , the vertex 푣 is a source. Moreover, in the similar orientation ƒ 퐷 ′ := 퐷 rev 휕 (푣) of 퐷 0 , the vertex 푣 is a sink.

20.1.2 Similarity and removable ears The following property is easy to be verified. This result will play a crucial role in the proof of Lemmas 20.8 and 20.9. Proposition 20.7 Let 푃 be a path of odd length and let 퐷 be an orientation of 푃 such that 푃 is oddly oriented in 퐷. Then, there exists a set 푋 of internal vertices of 푃 such that 푃 rev 휕 ( 푋) is a directed path. Proof As 푃 is oddly oriented in 퐷, the number of forward arcs of 푃 in 퐷 is odd. As 푃 has odd length, the number of reverse arcs of 푃 in 퐷 is even.

440

20 Similarity of Orientations and Characteristic Orientations

Let 푃 := 푣 1 푣 2 . . . 푣 푘 . For 푖 = 1, 2, . . . , 푘, let 푟 푖 denote the number of reverse arcs of 퐷 in the traversal of the subpath 푣 1 푣 2 . . . 푣 푖 of 푃. Let 푋 := {푣 푖 : 푟 푖 is odd}. For each forward arc 푒, either both ends of 푒 are in 푋 or neither end of 푒 is in 푋. For each reverse arc 푓 , precisely one of the ends of 푓 is in 푋. Thus, 푃 rev 휕 ( 푋) is a directed path from 푣 1 to 푣 푘 . Moreover, 푟 푘 is the number of reverse arcs of 푃, which is even, and 푟 1 is obviously even; hence each vertex in 푋 is an internal vertex of 푃. ƒ Lemma 20.8 Let 푅 be a removable single ear of a matching covered graph 퐺, let 푄 be a conformal cycle of 퐺 that contains all the edges of 푅, and let 퐷 1 and 퐷 2 be two orientations of 퐺. If 푄 is oddly oriented in both 퐷 1 and 퐷 2 and if 퐷 1 − 푅 and 퐷 2 − 푅 are similar then 퐷 1 and 퐷 2 are also similar. Proof As 퐷 1 − 푅 and 퐷 2 − 푅 are similar, 푉 (퐺 − 푅) has a subset 푋 such that 퐷 2 − 푅 = (퐷 1 − 푅) rev 휕퐺−푅 ( 푋). Let 퐶 := 휕퐺 ( 푋) and let 퐷 1′ := 퐷 1 rev 퐶. Then, 휕퐺−푅 ( 푋) = 퐶 − 푅; hence 퐷 2 − 푅 = (퐷 1 − 푅) rev (퐶 − 푅) = (퐷 1 rev 퐶) − 푅 = 퐷 ′1 − 푅. By Proposition 20.1(ii), 푄 has an even number of edges in 퐶. As 푄 is oddly oriented in 퐷 1 , we infer that 푄 is oddly oriented in 퐷 1′ . Thus, 푄 is oddly oriented in both 퐷 1′ and 퐷 2 . Adjust the direction of the traversal of 푄 so that the segment 푃 of 푄 induced by 퐸 (푅) is oddly oriented in 퐷 1′ . We have seen that 퐷 1′ − 푅 = 퐷 2 − 푅. As 푄 is oddly oriented in 퐷 2 , it follows that 푃 is also oddly oriented in 퐷 2 . By Proposition 20.7, there exists a set 푋1 of internal ′ vertices of 푃 such that 푃 is a directed path in 퐷 ′′ 1 := 퐷 1 rev 휕 ( 푋1 ). Likewise, there exists a set 푋2 of internal vertices of 푃 such that 푃 is a directed path in 퐷 ′2 := 퐷 2 rev 휕 ( 푋2). As 푋1 and 푋2 consist solely of internal vertices of 푃, it follows that 퐷 1′′ − 푅 = 퐷 1′ − 푅 = 퐷 2 − 푅 = 퐷 2′ − 푅. ′ As 푃 is a directed path in both 퐷 1′′ and 퐷 ′2 , we deduce that 퐷 ′′ 1 = 퐷 2 . The orientation ′ ′′ 퐷 1 is similar to the orientation 퐷 1 , which in turn is similar to 퐷 1 . The orientation 퐷 ′2 is similar to the orientation 퐷 2 . As 퐷 1′′ = 퐷 ′2 , we conclude that 퐷 1 and 퐷 2 are similar. ƒ

Lemma 20.9 Let 푅 be a removable double ear of a matching covered graph 퐺, let 푄 be a conformal cycle of 퐺 that contains all the edges of 푅 and let C := {퐷 1 , 퐷 2 , 퐷 3 } be a collection of three orientations of 퐺. If 푄 is oddly oriented in each orientation in C and if the orientations in D := {퐷 − 푅 : 퐷 ∈ C} are pairwise similar then at least two of the orientations in C are similar. Proof As the orientations in D are similar, we infer that, for 푖 = 1, 2, 푉 (퐺 − 푅) has subsets 푋푖 such that 퐷 3 − 푅 = (퐷 푖 − 푅) rev 휕퐺−푅 ( 푋푖 ). Let 퐶푖 := 휕퐺 ( 푋푖 ) and let 퐷 푖′ := 퐷 푖 rev 퐶푖 . In order to have a uniform notation, let 퐷 ′3 := 퐷 3 . Then, 휕퐺−푅 ( 푋푖 ) = 퐶푖 − 푅; hence

20.1 Similarity of Orientations

441

퐷 ′3 − 푅 = 퐷 3 − 푅 = (퐷 푖 − 푅) rev (퐶푖 − 푅) = (퐷 푖 rev 퐶푖 ) − 푅 = 퐷 ′푖 − 푅. Let 푅1 and 푅2 be the two constituent paths of 푅. For 푗 = 1, 2, let 푃 푗 be the subpath of 푄 induced by 퐸 (푅 푗 ). By Proposition 20.7, for 푖 = 1, 2, 3 and for 푗 = 1, 2 there exists a set 푌푖 푗 of internal vertices of 푃 푗 such that, in 퐷 ′푖 rev 휕 (푌푖 푗 ), 푃 푗 is a directed path if 푃 푗 is oddly oriented in 퐷 ′푖 ; otherwise the reverse of 푃 푗 is a directed path. For 푖 = 1, 2, 3, let 퐷 푖′′ := 퐷 ′푖 rev 휕 (푌푖1 ∪ 푌푖2 ). We have seen that 퐷 ′1 − 푅 = 퐷 2′ − 푅 = 퐷 3′ − 푅. As each 푌푖 푗 is a set of internal vertices of 푃 푗 , we conclude that for 푖 = 1, 2, 퐷 3′′ − 푅 = 퐷 ′3 − 푅 = 퐷 3 − 푅 = 퐷 푖′ − 푅 = 퐷 푖′′ − 푅. ′′ ′′ The orientations of 푃1 in at least two of 퐷 ′′ 1 , 퐷 2 and 퐷 3 coincide. Thus, let ′′ 1 ≤ 푖 < 푗 ≤ 3 be such that the orientations of 푃1 in 퐷 푖 and in 퐷 ′′푗 coincide. ′′ ′′ ′′ Then, the orientations of 푃2 in 퐷 ′′ 푖 and in 퐷 푗 also coincide. In that case, 퐷 푖 = 퐷 푗 . ′′ ′ The orientation 퐷 푖 is similar to the orientation 퐷 푖 , which in turn is similar to the orientation 퐷 푖 . Likewise, the orientation 퐷 ′′푗 is similar to the orientation 퐷 ′푗 , which in turn is similar to the orientation 퐷 푗 . We conclude that 퐷 푖 and 퐷 푗 are similar. ƒ

Exercises ⊲20.1.1 Prove Proposition 20.1. ∗ 20.1.2 (Similarity is an equivalence relation) (a) Let 퐶1 and 퐶2 be two cuts of a digraph 퐷. Show that (퐷 rev 퐶1 ) rev 퐶2 = 퐷 rev (퐶1 △ 퐶2 ). (b) Using part (a) and Proposition 20.1(i), prove Proposition 20.2. 20.1.3 Find two dissimilar Pfaffian orientations 퐷 1 and 퐷 2 of 퐾4 . Explain why 퐷 1 and 퐷 2 are dissimilar. (Hint: Let 퐷 1 denote any Pfaffian orientation of 퐾4 and let 퐷 2 be the converse of 퐷 1 , that is, one which is obtained from 퐷 1 by reversing the orientations of all edges of 퐾4 .) 20.1.4 (Extensions of Orientations of Spanning Trees) (a) Suppose that 퐺 is a Pfaffian graph and that 푇 is a spanning tree of 퐺. Show that every orientation of 푇 can be extended to a Pfaffian orientation of 퐺. (Consider similar Pfaffian orientations of 퐺 and show that for each edge 푒 of 푇, 퐺 has a cut 퐶 such that 퐶 ∩ 퐸 (푇) = {푒}.) (b) Suppose that (i) 퐺 is a Pfaffian graph, (ii) 푇0 is a spanning tree of 퐺 and (iii) 푇 := 푇0 + 푒 is not bipartite. Show that every orientation 퐷 푇 of 푇 can be extended to a Pfaffian orientation of 퐺. (Firstly, apply the reasoning of part (a) to obtain a Pfaffian orientation 퐷 of 퐺 which is an extension of 퐷 푇 − 푒. If the orientations of 푒 in 퐷 and in 퐷 푇 do not coincide, let 퐷 ′ := (퐷 rev 퐸 (퐺)) rev 휕 ( 퐴), where ( 퐴, 퐵) is the bipartition of 푇0 .)

442

20 Similarity of Orientations and Characteristic Orientations

(c) Use part (a) to show that 퐾3,3 is not Pfaffian. 20.1.5 Show that the Petersen graph is not Pfaffian. (This is not quite as straightforward as showing that 퐾3,3 is not Pfaffian.)

20.2 Pfaffian Orientations and Tight Cuts We shall now use similarity of orientations to prove the following fundamental result due to Little and Rendl (1991, [55]). Theorem 20.10 Let 퐺 be a matching covered graph and let 퐶 := 휕 ( 푋) be a tight cut of 퐺. The graph 퐺 is Pfaffian if and only if both 퐶-contractions of 퐺 are Pfaffian. In what follows we shall first analyse a relationship between Pfaffian orientations of a graph and those of the contractions of that graph with respect to a separating cut, and then deduce the above assertion from that analysis.

20.2.1 Pfaffian orientations and separating cuts We first proceed to show that if 퐶 is a tight cut of a matching covered graph 퐺 and 퐷 is an orientation of 퐺 such that 퐶 is a directed cut in 퐷 and both 퐶-contractions of 퐷 are Pfaffian, then 퐷 is Pfaffian. One might think that given a tight cut 퐶 of 퐺 and an orientation 퐷 of 퐺 such that both 퐶-contractions of 퐷 are Pfaffian then 퐷 would be Pfaffian. This is not the case. For a simple example, consider the orientation 퐷 of a hexagon depicted in Figure 20.1. The cut 퐶 is tight in 퐺, both 퐶-contractions of 퐷 are Pfaffian, yet 퐷 is not Pfaffian. In order for 퐷 to be Pfaffian, we require 퐶 to be a directed cut in 퐷.

퐶

Fig. 20.1 The cut 퐶 is tight, both 퐶-contractions of 퐷 are Pfaffian, yet 퐷 is not Pfaffian

We prove a similar statement, where 퐶 is a separating cut, which clearly implies the result mentioned above, if 퐶 is tight.

20.2 Pfaffian Orientations and Tight Cuts

443

Lemma 20.11 Let 퐶 := 휕 ( 푋) be a separating cut of a matching covered graph 퐺 and let 퐷 be an orientation of 퐺 such that (i) 퐶 is a directed cut in 퐷 and (ii) both 퐶-contractions of 퐷 are Pfaffian. Then, all the perfect matchings of 퐺 that contain only one edge in 퐶 have the same sign in 퐷. Proof By hypothesis, the cut 퐶 is directed in 퐷. Adjust notation so that 푋 is the shore of 퐶 which contains the heads of all the arcs of 퐶. Let 퐺 1 := 퐺/( 푋 → 푥) and let 퐺 2 := 퐺/( 푋 → 푥) be the two 퐶-contractions of 퐺. Let 퐷 1 := 퐷/( 푋 → 푥) and let 퐷 2 := 퐷/( 푋 → 푥) be the two 퐶-contractions of 퐷. By hypothesis, 퐷 1 and 퐷 2 are Pfaffian. Let 푀 and 푁 be two perfect matchings of 퐺 and let 푒 and 푓 be two edges in 퐶 such that 푀 ∩ 퐶 = {푒} and 푁 ∩ 퐶 = { 푓 }. We shall prove that 푀 and 푁 have the same sign by showing that every (푀, 푁)-alternating cycle of 퐺 is oddly oriented. Thus, let 푄 be any (푀, 푁)-alternating cycle of 퐺. As 푀 and 푁 both have just one edge in 퐶, it follows that for 푖 = 1, 2, the sets 푀푖 := 푀 ∩ 퐸 (퐺 푖 ) and 푁푖 := 푁 ∩ 퐸 (퐺 푖 ) are perfect matchings of 퐺 푖 . If 푄 and 퐶 are disjoint then for some 푖 ∈ {1, 2}, 푄 is an (푀푖 , 푁푖 )-alternating cycle in 퐺 푖 . As 퐷 푖 is Pfaffian, 푄 is oddly oriented in 퐷 푖 , hence it is oddly oriented in 퐷. We may thus assume that 푄 and 퐶 have edges in common. Then, 푄 necessarily contains the edge 푒 of 푀 ∩퐶 and the edge 푓 of 푁 ∩퐶. Moreover, the edges 푒 and 푓 are the only two edges of 푄 in 퐶. For 푖 = 1, 2, the set 퐸 (푄) ∩ 퐸 (퐺 푖 ) induces an 푀푖 -alternating cycle, say 푄 푖 , in 퐺 푖 . As 퐷 푖 is Pfaffian, 푄 푖 is oddly oriented in 퐷 푖 . Consider traversing 푄, 푄 1 and 푄 2 so that 푒 is a forward edge in each of the three traversals (and 푓 is a reverse edge in each of the three traversals). Then, fw(푄 1 ) ∩ fw(푄 2 ) = {푒}. It follows that |fw(푄)| = |fw(푄 1 )| + |fw(푄 2 )| − 1 ≡ 1

(mod 2).

We conclude that 푄 is oddly oriented in 퐷. Regardless of whether 푄 and 퐶 are disjoint, the cycle 푄 is oddly oriented in 퐷. This conclusion holds for each (푀, 푁)-alternating cycle 푄 of 퐺. By Corollary 19.4, the perfect matchings 푀 and 푁 have the same sign. This conclusion holds for every pair {푀, 푁 } of perfect matchings of 퐺 that contain just one edge in 퐶. ƒ Our next step towards the proof of Theorem 20.10 is to consider a ‘converse’ of Lemma 20.11. We thus proceed to show that if 퐺 is a matching covered graph that admits a Pfaffian orientation and 퐶 is a separating cut of 퐺, then both 퐶-contractions of 퐺 also admit Pfaffian orientations. Before we present a proof of this lemma it is worth noting one subtle point. Suppose that 퐺 is a Pfaffian graph and 퐶 is a separating cut of 퐺. The graph 퐺, being Pfaffian, admits a Pfaffian orientation, say 퐷. One might think that for any such orientation 퐷 of 퐺, the two 퐶-contractions of 퐷 are Pfaffian orientations of the corresponding 퐶-contractions of 퐺. But this is not the case. The Pfaffian orientation 퐷 0 of 퐶6 shown in Exercise 19.2.7 illustrates this possibility. Another illustration is given by Example 20.12.

444

20 Similarity of Orientations and Characteristic Orientations

Example 20.12 Let 퐺 be the graph obtained by splicing 퐾4 and the cube, and let 퐷 be the orientation of 퐺 depicted in Figure 20.2. 푣1

푥1

푋

푋

푣2

푤1

푤2

푢1

푢3

푢2

퐶

푥2

푢4

Fig. 20.2 A tight cut 퐶 of 퐺 and a Pfaffian orientation 퐷 of 퐺 where neither 퐶-contraction of 퐷 is Pfaffian

The cut 퐶 := 휕 ( 푋) is a tight cut of 퐺. It may be verified that 퐷 is a Pfaffian orientation of 퐺, but that neither 퐶-contraction of 퐷 is Pfaffian. However, the orientation 퐷 ′ := 퐷 rev 휕 ({푤 2 , 푥1 , 푥2 }) is similar to 퐷, hence it is Pfaffian and is such that 퐶 is a directed cut of 퐷 ′ and the two 퐶-contractions of 퐷 ′ are Pfaffian. See Exercise 20.2.2. Lemma 20.13 (CLM (2012, [16])) Let 퐶 := 휕 ( 푋) be a separating cut of a matching covered graph 퐺 and let 퐷 0 be a Pfaffian orientation of 퐺. Then 퐺 has a (Pfaffian) orientation 퐷 similar to 퐷 0 such that 퐶 is directed in 퐷 and both 퐶-contractions of 퐷 are Pfaffian. Proof We wish to show that there is an orientation 퐷 of 퐺 which is similar to 퐷 0 such that 퐶 is a directed cut in 퐷. From this, using Lemma 19.5 we shall deduce that the two 퐶-contractions of 퐷 are Pfaffian orientations of the corresponding 퐶-contractions of 퐺. With this as our objective, we now proceed to describe a procedure for determining a subset 푆 of the ends of edges in 퐶 and show that 퐷 := 퐷 0 rev 휕 (푆) has the asserted properties. (Example 20.14, which is a continuation of Example 20.12, illustrates this step of the proof.) First note that 퐷, being similar to 퐷 0 , is Pfaffian by Proposition 20.5.

20.2 Pfaffian Orientations and Tight Cuts

445

Let 퐺 1 := 퐺/( 푋 → 푥) and 퐺 2 := 퐺/( 푋 → 푥) be the two 퐶-contractions of 퐺. Also, let 푉1 be the set of vertices of 푋 that are ends of edges of 퐶 and let 푉2 be the set of vertices of 푋 that are ends of edges of 퐶. Choose any edge 푒 := 푣 1 푣 2 of cut 퐶 where 푣 1 ∈ 푉1 and 푣 2 ∈ 푉2 . As 퐶 is a separating cut of 퐺, there exists a perfect matching of 퐺 which has just the edge 푒 in 퐶. Let 푀 be such a perfect matching of 퐺. For 푖 = 1, 2, let 푀푖 be the perfect matching 푀 ∩ 퐸 (퐺 푖 ) of 퐺 푖 that contains the edge 푒. See Figure 20.3.

푄

푋

푣1

푒

푣2

푃 (푤2 )

푄1

푤1

푓

푋

푤2

Fig. 20.3 Proof of Lemma 20.13

20.13.1 For any vertex 푤 2 in 푉2 , the graph 퐺 [푋] has an 푀2 -alternating path 푃(푤 2 ) of even length from 푣 2 to 푤 2 . See Figure 20.3. Proof If 푤 2 = 푣 2 then let 푃(푤 2 ) := 푣 2 . We may thus assume that 푤 2 ≠ 푣 2 . As 푤 2 ∈ 푉2 , 푤 2 is incident with an edge in 퐶, say 푓 . As 퐶 is separating, 퐺 has a perfect matching, say 푁, such that 푁 ∩ 퐶 = { 푓 }. Let 푄 be the (푀, 푁)-alternating cycle in 퐺 that contains edge 푓 . Then, this path 푄 meets 퐶 in precisely the two edges 푒 and 푓 . Let 푃(푤 2 ) denote the segment of 푄 in 퐺 2 that joins 푣 2 and 푤 2 . Clearly, 푃(푤 2 ) is an (푀, 푁)-alternating path in 퐺 2 − 푥, hence 푃(푤 2 ) is 푀2 -alternating. The edge of 푃(푤 2 ) incident with 푣 2 is in 푁, and the edge of 푃(푤 2 ) incident with 푤 2 is in 푀, hence 푃(푤 2 ) has even length. ƒ Likewise, for each vertex 푤 1 in 푉1 , define a path 푃(푤 1 ) to be an 푀1 -alternating path of even length in 퐺 [푋] that connects 푣 1 and 푤 1 . Thus, for any vertex 푤 in 푉1 ∪ 푉2 , there is an 푀-alternating path 푃(푤). We now define 푆 as the subset of 푉1 ∪ 푉2 consisting of those vertices 푤 such that 푃(푤) is oddly oriented in 퐷 0 . Recall that we have defined 퐷 to be equal to 퐷 0 rev 휕푆. 20.13.2 For each vertex 푤 ∈ 푉1 ∪ 푉2 , the path 푃(푤) is evenly oriented in 퐷. Proof The ends 푣 1 and 푣 2 of 푒 are not in 푆. Thus, reversal of the edges of 휕 (푆) preserves the parity of 푃(푤) if and only if 푤 ∉ 푆. The vertex 푤 is in 푆 if and only if 푃(푤) is oddly oriented in 퐷 0 . We deduce that 푃(푤) is evenly oriented in 퐷. This conclusion holds for each vertex 푤 ∈ 푉1 ∪ 푉2 . ƒ

446

20 Similarity of Orientations and Characteristic Orientations

For each 푤 in 푉1 ∪ 푉2 , the path 푃(푤) has even length and is evenly oriented in 퐷. Thus, the reverse of 푃(푤) is also evenly oriented in 퐷. 20.13.3 The cut 퐶 is directed in 퐷. Proof Let 푓 := 푤 1 푤 2 be any edge of 퐶 − 푒. The set {푒, 푓 } ∪ 퐸 (푃(푤 1 )) ∪ 퐸 (푃(푤 2 )) induces an 푀-alternating cycle, 푄, of 퐺. Let us traverse 푄 in 퐷 so that 푒 is a forward edge of 푄. The orientation 퐷 of 퐺 is similar to the Pfaffian orientation 퐷 0 of 퐺, hence it is Pfaffian. Thus, 푄 is oddly oriented. The paths 푃(푤 1 ) and 푃(푤 2 ) are evenly oriented. Thus, 푓 is a reverse edge of 푄 in 퐷. It follows that the heads of 푒 and 푓 in 퐷 are in the same shore of 퐶. This conclusion holds for each edge 푓 of 퐶 − 푒. Indeed, 퐶 is a directed cut in 퐷. ƒ For 푖 = 1, 2, let 퐷 푖 be the restriction of 퐷 to 퐺 푖 . 20.13.4 The orientation 퐷 1 of 퐺 1 is Pfaffian. Proof Let 푄 1 be an 푀1 -alternating cycle of 퐺 1 . To prove that 퐷 1 is Pfaffian, we make use of Lemma 19.5, and show that 푄 1 is oddly oriented in 퐷 1 . If 푄 1 has no edge in 퐶 then 푄 1 is an 푀-alternating cycle of 퐺. As 퐷 is Pfaffian, 푄 1 is oddly oriented in 퐷, hence it is also oddly oriented in 퐷 1 . We may thus assume that 푄 1 contains edges in 퐶. Clearly, 푄 1 contains precisely two edges in 퐶, one of which is 푒. Let 푓 be the other edge of 푄 1 in 퐶. Let 푤 2 be the end of 푓 in 푋. Let 푄 be the cycle of 퐺 induced by 퐸 (푄 1 ) ∪ 퐸 (푃(푤 2 )). The cycle 푄 is 푀-alternating in 퐺. As 퐷 is Pfaffian, 푄 is oddly oriented in 퐷. The path 푃(푤 2 ) is evenly oriented in 퐷 and the cut 퐶 is directed in 퐷. Thus, 푄 1 is oddly oriented in 퐷 1 . This conclusion holds for every 푀1 -alternating cycle 푄 1 of 퐺 1 . It follows that 퐷 1 is Pfaffian. ƒ A similar reasoning may be used to prove that 퐷 2 is Pfaffian. The proof of Lemma 20.13 is complete. ƒ Example 20.14 In the digraph depicted in Figure 20.2, where the shore 푋 of 퐶 is {푣 1 , 푤 1 , 푥1 }, 푀 := {푣 1 푣 2 , 푤 1 푥1 , 푢 1 푤 2 , 푢 2 푢 3 , 푢 4 푥2 }, 푁 := {푤 1 푤 2 , 푣 1 푥1 , 푢 1 푣 2 , 푢 2 푢 3 , 푢 4 푥2 }, and 퐹 := {푥1 푥2 , 푣 1 푤 1 , 푢 3 푤 2 , 푢 1 푢 2 , 푢 4 푣 2 } are perfect matchings such that the only edge in 퐶 ∩ 푀 is 푒 := 푣 1 푣 2 , the only edge of 퐶 ∩ 푁 is 푤 1 푤 2 , and the only edge of 퐶 ∩ 퐹 is 푥1 푥2 . In the shore 푋 of 퐶, the (푀2 , 푁2 )-alternating path from 푣 2 to 푤 2 is oddly oriented and the (푀2 , 퐹2 )-alternating path from 푣 2 to 푥2 is also oddly oriented. In the shore 푋 of 퐶, the (푀1 , 퐹1 )-alternating path from 푣 1 to 푥1 is oddly oriented, but the (푀1 , 푁1 )-alternating path from 푣 1 to 푤 1 is not. Thus 푆 = {푤 2 , 푥1 , 푥2 }, as indicated in Example 20.12.

Proof of Theorem 20.10 Let 퐺 1 = 퐺/( 푋 → 푥) and let 퐺 2 := 퐺/( 푋 → 푥) be the two 퐶-contractions of 퐺. Suppose that 퐺 is Pfaffian. By Lemma 20.13, 퐺 has a Pfaffian orientation whose 퐶-contractions are both Pfaffian. Thus, 퐺 1 and 퐺 2 are both Pfaffian.

20.2 Pfaffian Orientations and Tight Cuts

447

Conversely, suppose that 퐺 1 and 퐺 2 are Pfaffian. By Proposition 20.6, 퐺 1 has a Pfaffian orientation, say 퐷 1 , such that its contraction vertex 푥 is a source in 퐷 1 . Likewise, 퐺 2 has a Pfaffian orientation, say 퐷 2 , such that its contraction vertex 푥 is a sink. Let 퐷 be the common extension of 퐷 1 and 퐷 2 to an orientation of 퐺. Clearly, 퐶 is a directed cut of 퐷. By Lemma 20.11, all the perfect matchings of 퐺 that contains just one edge in 퐶 have the same sign in 퐷. As 퐶 is tight, every perfect matching of 퐺 contains just one edge in 퐶. Thus, 퐷 is Pfaffian. We conclude that 퐺 is Pfaffian. ƒ It should be noted that the converse of Lemma 20.13 is false. That is, there are matching covered graphs that are not Pfaffian but have a separating cut 퐶 such that both its 퐶-contractions are Pfaffian. For example, the Petersen graph is not Pfaffian but it is the splicing of two 5-wheels, which are Pfaffian. Another non-Pfaffian splicing of two 5-wheels is presented in Figure 20.4.

Fig. 20.4 A non-Pfaffian graph which is a splicing of two 5-wheels

In Chapter 21 we will exhibit other counterexamples of the converse of Lemma 20.13. In fact, we shall present an infinite family of such graphs.

Some immediate consequences of Theorem 20.10 Corollary 20.15 A matching covered graph is Pfaffian if and only if all of its bricks and braces are Pfaffian. ƒ Pfaffian Orientations and Matching Minors Recall that a graph 퐻 is a matching minor of graph 퐺 if 퐻 can be obtained from a conformal subgraph of 퐺 by means of a sequence of bicontractions. The following assertion may be easily deduced (Exercise 20.2.5) from Theorem 20.10.

448

20 Similarity of Orientations and Characteristic Orientations

Corollary 20.16 A matching covered graph 퐺 is Pfaffian if and only if each of ƒ its matching minors (as defined in Chapter 12) is also Pfaffian. A non-Pfaffian matching covered graph 퐻 is matching-minor-minimal if every proper matching-minor of 퐻 is Pfaffian. (For example, 퐾3,3 is such a graph.) This leads us to contemplate the possibility of ‘excluded-matchingminor’ characterisations of Pfaffian matching covered graphs which is analogous to Kuratowski’s characterization of planar graphs. We shall explore this idea extensively in the next chapter by introducing the notion of a minor of a matching covered graph which is more general than that of a matching-minor. Corollary 20.17 Every optimal removable single ear of a Pfaffian matching covered graph is 푏-invariant. Proof Let 퐺 be a Pfaffian matching covered graph and let 푅 be an optimal removable single ear of 퐺. The graph 퐺 is Pfaffian and the Petersen graph is not Pfaffian. By Corollary 20.15, no brick of 퐺 is a Petersen brick, hence 푝(퐺) = 0. Likewise, 퐺 − 푅, a conformal subgraph of 퐺, is Pfaffian, by Lemma 19.5(ii), hence 푝(퐺 − 푅) = 0. By the monotonicity of function 푏, it follows that (푏 + 푝) (퐺) = 푏(퐺) ≤ 푏(퐺 − 푅) ≤ (푏 + 푝) (퐺 − 푅) = (푏 + 푝) (퐺). Equality holds throughout, hence 푏(퐺 − 푅) = 푏(퐺). Indeed, 푅 is 푏-invariant. This conclusion holds for each optimal removable ear 푅 of every Pfaffian matching covered graph. ƒ In Chapter 16 we have shown (Corollary 16.10) that every matching covered 퐺 has an ear decomposition which uses just 푏 + 푝 double ears. Combined with the above result we thus have: Corollary 20.18 Every Pfaffian matching covered graph 퐺 has an ear decomposition which uses just 푏 double ears. ƒ

Exercises 20.2.1 Let 퐺 be a graph, let 퐷 be an orientation of 퐺 and let 푣 be a vertex of degree two in 퐺 which is adjacent to two distinct vertices. Suppose that 푣 is neither a source nor a sink in 퐷. Prove that 퐷 is Pfaffian if and only if the digraph obtained from 퐷 by the bicontraction of 푣 is Pfaffian. ⊲20.2.2 Show that: (i) the directed graph 퐷 depicted in Figure 20.2 is a Pfaffian orientation of the graph 퐺 obtained by splicing 퐾4 and the cube; (ii) neither 퐶-contraction of 퐷 is Pfaffian,

20.3 The Number of Similarity Classes

449

(iii) the digraph 퐷 ′ := 퐷 rev 휕 ({푤 2 , 푥1 , 푥2 }) is also a Pfaffian orientation of 퐺; and finally show that (iv) both 퐶-contractions of 퐷 ′ are Pfaffian. ⊲20.2.3 Prove that the graph depicted in Figure 20.4 is not Pfaffian. ⊲20.2.4 The objective of this exercise is to provide a proof of a weaker version of Lemma 20.13. Let 퐺 and 퐶 be as in the statement of Lemma 20.13. Let 퐺 1 and 퐺 2 be the two 퐶-contractions of 퐺, let 퐷 be a Pfaffian orientation of 퐺 and let 퐷 1 and 퐷 2 be the 퐶-contractions of 퐷 which are orientations of 퐺 1 and 퐺 2 , respectively. For each edge 푒 of 퐶, let M1 (푒) denote the set of perfect matchings of 퐺 1 that contain edge 푒. (i) Prove that, for each edge 푒 ∈ 퐶, all the perfect matchings in M1 (푒) have the same sign in 퐷 1 . Hint: consider a perfect matching 푀2 of 퐺 2 that contains edge 푒 and the collection M (푒) := {푀1 ∪ 푀2 : 푀1 ∈ M1 (푒)}. (ii) From the previous part, conclude that 퐺 1 is Pfaffian. Hint: define the set 푅 to be {푒 ∈ 퐶 : sign(푀1 ) = −1 ∀푀1 ∈ M1 (푒)} and deduce that 퐷 1 rev 푅 is Pfaffian. Thus, 퐺 1 is Pfaffian. Likewise, 퐺 2 is also Pfaffian. 20.2.5 Give a proof of Corollary 20.16.

20.3 The Number of Similarity Classes ♯ Recall that if 퐷 is a Pfaffian orientation of 퐺, then any orientation of 퐺 that is similar to 퐷 is also a Pfaffian orientation of 퐺. Moreover, similarity is an equivalence relation. The number 휏(퐺) of dissimilar Pfaffian orientations Let 휏(퐺) denote the number of equivalence classes of Pfaffian orientations of a graph 퐺 with respect to similarity. In this section we shall prove that 휏(퐺) = 2푏 (퐺) for every Pfaffian matching covered graph 퐺. (Thus, interestingly, 휏(퐺) depends only on the number 푏(퐺) of bricks of 퐺.) Theorem 20.19 The equality 휏(퐺) = 2푏 (퐺) holds for every Pfaffian matching covered graph 퐺. Proof The proof is divided in two parts, consisting of the proofs of Lemmas 20.20 and 20.23. ƒ

450

20 Similarity of Orientations and Characteristic Orientations

20.3.1 An upper bound for 흉(푮) Lemma: an upper bound for the number of dissimilar Pfaffian orientations 20.20 Let 퐺 be a Pfaffian matching covered graph. Then, 휏(퐺) ≤ 2푏 (퐺) . Proof by induction on |퐸 |. Adopt, as the induction hypothesis, that the inequality 휏(퐻) ≤ 2푏 (퐻 ) holds for every Pfaffian matching covered graph 퐻 having fewer edges than 퐺. Let C be a collection of 휏(퐺) dissimilar Pfaffian orientations of 퐺. If 퐺 = 퐾2 then the two orientations of 퐺 are Pfaffian and similar. In that case, 휏(퐺) = |C| = 1. We may thus assume that 퐺 ≠ 퐾2 . By Theorem 16.9, 퐺 has an optimal removable ear, 푅. Let 푄 be a conformal cycle of 퐺 that contains an edge of 푅. Then, 푄 contains all the edges of 푅. The orientations in C are Pfaffian; hence 푄 is oddly oriented in each orientation in C. Let D := {퐷 − 푅 : 퐷 ∈ C}. Each orientation in D is a Pfaffian orientation of 퐺 − 푅. By induction, 휏(퐺 − 푅) ≤ 2푏 (퐺−푅) . Suppose that 푅 is a single ear. By Corollary 20.17, 푅 is 푏-invariant; hence 푏(퐺 − 푅) = 푏(퐺). By Lemma 20.8, the 휏(퐺) orientations in D are pairwise dissimilar, because the orientations in C are pairwise dissimilar. Thus, 휏(퐺) ≤ 휏(퐺 − 푅). In sum, 휏(퐺) ≤ 휏(퐺 − 푅) ≤ 2푏 (퐺−푅) = 2푏 (퐺) . The assertion holds if 푅 is a single ear. Suppose now that 푅 is a double ear. Then, 푏(퐺 − 푅) = 푏(퐺) − 1. The orientations in C are pairwise dissimilar. By Lemma 20.9, no three orientations in D are pairwise similar. Thus, D has a subcollection of at least |D|/2 pairwise dissimilar Pfaffian orientations; hence |D|/2 ≤ 휏(퐺 − 푅). Thus, |D| ≤ 2휏(퐺 − 푅). We deduce that 휏(퐺) ≤ 2휏(퐺 − 푅). In sum, 휏(퐺) ≤ 2휏(퐺 − 푅) ≤ 2 · 2푏 (퐺−푅) = 2 · 2푏 (퐺) −1 = 2푏 (퐺) . The assertion also holds if 푅 is a double ear. The proof of Lemma 20.20 is complete.ƒ Corollary 20.21 Let 퐺 be a Pfaffian matching covered graph which is either bipartite or a near-brick. Then,  1, if 퐺 is bipartite 휏(퐺) = 2, if 퐺 is a near-brick Proof Suppose that 퐺 is bipartite. As 퐺 is Pfaffian, 1 ≤ 휏(퐺). By the Theorem, 휏(퐺) ≤ 2푏 (퐺) = 1. Thus, 휏(퐺) = 1. Suppose that 퐺 is a near-brick. By the Theorem, 휏(퐺) ≤ 2푏 (퐺) = 2. As 퐺 is Pfaffian, let 퐷 be a Pfaffian orientation of 퐺. The orientation 퐷 rev 퐸 is also a Pfaffian orientation of 퐺. Moreover, as 퐺 is not bipartite, the set 퐸 is not a cut of 퐺; hence 퐷 and 퐷 rev 퐸 are dissimilar. We deduce that 휏(퐺) = 2. ƒ

20.3 The Number of Similarity Classes

451

20.3.2 A lower bound for 흉(푮) In order to determine a lower bound for 휏(퐺), we need the following elementary property of matching covered graphs. Proposition 20.22 Let 퐺 be a matching covered graph and let 푣 be a vertex of 퐺. If 퐺 − 푣 is bipartite then 퐺 is also bipartite. Proof Suppose that 퐺 − 푣 is bipartite and let ( 퐴, 퐵) denote a bipartition of 퐺 − 푣. Let 푒 := 푣푤 be an edge of 퐺 incident with 푣 and let 푀 be a perfect matching of 퐺 that contains the edge 푒. The vertex 푤 is in 퐴 ∪ 퐵. Adjust notation so that 푤 ∈ 퐵. Every vertex of 퐴 + 푣 is matched with a vertex of 퐵; thus |퐵| = | 퐴 + 푣|. It follows that 퐴 + 푣 is a barrier of 퐺. As 퐺 is matching covered, 퐴 + 푣 is an independent set ƒ of vertices of 퐺. Thus, ( 퐴 + 푣, 퐵) is a bipartition of 퐺. Lemma: a lower bound for the number of dissimilar Pfaffian orientations 20.23 Let 퐺 be a Pfaffian matching covered graph. Then, 휏(퐺) ≥ 2푏 (퐺) . Proof by induction on |퐸 |. Adopt as the inductive hypothesis that the inequality 휏(퐻) ≥ 2푏 (퐻 ) holds for each matching covered graph 퐻 having fewer edges than 퐺. If 퐺 is a brace then, by Corollary 20.21, 휏(퐺) = 1 and if 퐺 is a brick then, again by Corollary 20.21, 휏(퐺) = 2. In both cases, the assertion holds. We may thus assume that 퐺 is neither a brace nor a brick; hence 퐺 has nontrivial tight cuts. Let 푋 be a minimal set of vertices of 퐺 such that 퐶 := 휕 ( 푋) is a nontrivial tight cut. Let 퐺 1 := 퐺/( 푋 → 푥) and 퐺 2 := 퐺/( 푋 → 푥) be the two 퐶-contractions of 퐺. For 푖 = 1, 2, let 푏(퐺 푖 ) be the number of bricks of 퐺 푖 . Then, 푏(퐺) = 푏(퐺 1 ) + 푏(퐺 2 ). The minimality of 푋 implies that 퐺 1 is either a brick or a brace. The graph 퐺 is Pfaffian. By Theorem 20.10, 퐺 1 and 퐺 2 are Pfaffian. Let 퐷 1 be a Pfaffian orientation of 퐺 1 . By induction and Lemma 20.20, 휏(퐺 2 ) = 2푏 (퐺2 ) . Let D2 be a collection of 2푏 (퐺2 ) pairwise dissimilar Pfaffian orientations of 퐺 2 . By Proposition 20.6, we may assume that the contraction vertex 푥 of 퐺 1 is a source in 퐷 1 and the contraction vertex 푥 of 퐺 2 is a sink in each orientation in D2 . For each orientation 퐷 2 in D2 , there is an orientation 퐷 of 퐺 which is the common extension of 퐷 1 and 퐷 2 . We say that 퐷 is generated by 퐷 1 and 퐷 2 . Let D be the collection of orientations of 퐺 generated by 퐷 1 and the orientations in D2 . Clearly, |D| = |D2 | = 2푏 (퐺2 ) . 20.23.1 The collection D consists of 2푏 (퐺2 ) pairwise dissimilar Pfaffian orientations of 퐺. Proof By Lemma 20.11, each orientation in D is a Pfaffian orientation of 퐺. Let us now prove that the orientations in D are pairwise dissimilar.

452

20 Similarity of Orientations and Characteristic Orientations

Assume, to the contrary, that D contains two similar orientations, 퐷 and 퐹, respectively generated by distinct orientations 퐷 2 and 퐹2 in D2 . Let 푅 be the set of edges of 퐺 whose orientations in 퐷 and in 퐹 differ. We have assumed that 퐷 and 퐹 are similar; hence 푅 is a cut of 퐺, say, 휕 (푌 ). For each edge in 퐸 (퐺 1 ), its orientations in 퐷 and in 퐹 coincide with its orientation in 퐷 1 . Thus, 푅 ⊆ 퐸 (퐺 2 ) − 퐶. We may then adjust notation, so that 푌 ⊆ 푉 (퐺 2 − 푥); hence 푅 is a cut of 퐺 2 . Moreover, 퐷 2 rev 푅 = 퐹2 ; hence 퐷 2 and 퐹2 are similar. This is a contradiction. We conclude that the 2푏 (퐺2 ) orientations in D are pairwise dissimilar Pfaffian orientations of 퐺.ƒ If 퐺 1 is a brace then 푏(퐺) = 푏(퐺 2 ); hence D is a collection of 2푏 (퐺) pairwise dissimilar Pfaffian orientations of 퐺. The assertion holds in this case. We may thus assume that 퐺 1 is a brick; hence 푏(퐺) = 푏(퐺 2 ) + 1. It follows that |D| = 2푏 (퐺) −1 . Let 퐷 1′ be the orientation 퐷 1 rev (퐸 (퐺 1 ) − 퐶) of 퐺 1 . Let D ′ be the collection of orientations of 퐺 which are generated by 퐷 1′ and the orientations in D2. Repeating the reasoning done above, we conclude that D ′ consists of 2푏 (퐺) −1 pairwise dissimilar Pfaffian orientations of 퐺. Let 퐷 ∈ D and 퐷 ′ ∈ D ′ and let 푅 be the set of edges whose orientations in 퐷 and in 퐷 ′ differ. Clearly, 퐸 (퐺 1 ) − 퐶 ⊆ 푅. As 퐺 1 is a brick, 퐺 1 − 푥 is nonbipartite, by Proposition 20.22. Thus, 퐺 1 − 퐶 has an odd cycle, 푄. Every edge of 푄 is in 푅. By Proposition 20.1(ii), 푅 is not a cut of 퐺; hence 퐷 and 퐷 ′ are dissimilar. This conclusion holds for each 퐷 ∈ D and each 퐷 ′ ∈ D ′ . We deduce that D ∪ D ′ is a collection of 2푏 (퐺) pairwise dissimilar Pfaffian orientations of 퐺. In all alternatives considered, we proved that 휏(퐺) ≥ 2푏 (퐺) . The proof of Lemma 20.23 completes the proof of Theorem 20.19. ƒ

Exercises 20.3.1 Find two dissimilar Pfaffian orientations of 퐺, for 퐺 = 퐾4 , 퐶6 , M8 . ⊲20.3.2 Consider the matching covered graph 퐺 with 푏(퐺) = 3 obtained by splicing a 퐾4 at three vertices which belong to one of the two colour classes of the cube. Find 23 = 8 pairwise dissimilar Pfaffian orientations of 퐺.

20.4 Characteristic Orientations

Characteristic orientations An orientation 퐷 of a (not necessarily matching covered) graph 퐺 is a characteristic orientation of 퐺 if it has the property that 퐺 is Pfaffian if and only if 퐷 is a Pfaffian orientation of 퐺.

20.4 Characteristic Orientations

453

In this section, based on Theorem 20.19, we shall deduce the existence of a polynomial-time algorithm for finding a characteristic orientation of a graph. This striking result makes it possible for us to deduce the polynomial-time equivalence of the Pfaffian Recognition Problem 19.1 and the Pfaffian Orientation Problem 19.2.

20.4.1 Pfaffian extensions We now present a fundamental result which is the basis of the polynomial-time algorithm that produces a characteristic orientation of a graph. The Pfaffian extension theorem 20.24 Let 퐺 be a matching covered graph, let 퐷 be an orientation of 퐺, let 푅 be an optimal removable ear of 퐺 and let 푄 be a conformal cycle of 퐺 that contains all the edges of 푅. Suppose that (i) 퐷 − 푅 is Pfaffian and (ii) 푄 is oddly oriented in 퐷. Then, 퐺 is Pfaffian if and only if 퐷 is Pfaffian. Proof If 퐷 is Pfaffian then certainly 퐺 is Pfaffian. To prove the converse, suppose that 퐺 is Pfaffian. Let 푡 := 2푏 (퐺) . By Theorem 20.19, the graph 퐺 has a collection D := {퐷 1 , 퐷 2 , . . . , 퐷 푡 } of 푡 dissimilar Pfaffian orientations. Let 퐷 0 := 퐷 and let D ′ := {퐷 푖 − 푅 : 0 ≤ 푖 ≤ 푡}. For each ℓ such that 1 ≤ ℓ ≤ 푡, the orientation 퐷 ℓ of 퐺 is Pfaffian; hence 퐷 ′ℓ is also a Pfaffian orientation of 퐺 − 푅. Moreover, the cycle 푄 is oddly oriented in 퐷 ℓ , because 푄 is a conformal cycle of 퐺. By definition, 퐷 0 − 푅 is a Pfaffian orientation of 퐺 − 푅 and 푄 is also oddly oriented in 퐷 0 . Case 1 푅 is a single ear. By hypothesis, 푅 is an optimal removable ear of 퐺. As 푅 is a single ear, it is (푏 + 푝)invariant. By Corollary 20.17, 푅 is 푏-invariant; hence 2푏 (퐺−푅) = 2푏 (퐺) = 푡. By Theorem 20.19, 푡 is the number of dissimilar Pfaffian orientations of 퐺 − 푅. As D ′ consists of 푡 + 1 Pfaffian orientations of 퐺 − 푅, it follows that D ′ contains two similar orientations. Let 퐷 푖 and 퐷 푗 be orientations in D + 퐷 0 , where 0 ≤ 푖 < 푗 ≤ 푡 and 퐷 푖 − 푅 and 퐷 푗 − 푅 are similar. Moreover, 푄 is oddly oriented in both 퐷 푖 and 퐷 푗 . By Lemma 20.8, 퐷 푖 and 퐷 푗 are similar. The orientations in D are not similar; hence 푖 = 0. We deduce that 퐷 is similar to some Pfaffian orientation in D. Thus, 퐷 is Pfaffian. Case 2 푅 is a double ear. From the fact that 푅 is a double ear we infer that 2푏 (퐺−푅) = 2푏 (퐺) −1 = 푡/2. By Theorem 20.19, 푡/2 is the number of dissimilar Pfaffian orientations of 퐺 − 푅. As D ′ consists of 푡 + 1 Pfaffian orientations of 퐺 − 푅, it follows that D + 퐷 0 contains three orientations, 퐷 푖 , 퐷 푗 and 퐷 푘 , where 0 ≤ 푖 < 푗 < 푘 ≤ 푡 and the three orientations 퐷 푖 − 푅, 퐷 푗 − 푅 and 퐷 푘 − 푅 are similar. Moreover, 푄 is oddly oriented in 퐷 푖 , 퐷 푗 and 퐷 푘 . By Lemma 20.9, two of the three orientations 퐷 푖 , 퐷 푗 and 퐷 푘 are similar. The orientations in D are not similar; hence 푖 = 0. We deduce that 퐷 is similar to some Pfaffian orientation in D. Thus, 퐷 is Pfaffian. ƒ

454

20 Similarity of Orientations and Characteristic Orientations

Corollary 20.25 Let 퐺 be a Pfaffian matching covered graph and let 푅 be an optimal removable ear of 퐺. Every Pfaffian orientation of 퐺 − 푅 has an extension to a Pfaffian ƒ orientation of 퐺.

20.4.2 Finding characteristic orientations We are now in position to describe a polynomial-time algorithm alluded to at the beginning of this section. This algorithm finds a characteristic orientation of a graph in polynomial time. We first give a special case, in which the given graph is matching covered. This algorithm is then used in the full version, in which the given graph is not necessarily matching covered. Finding a characteristic orientation Algorithm 20.26 Input: a (not necessarily matching covered) graph 퐺; Output: a characteristic orientation of 퐺. A fundamental case: 퐺 is matching covered. We first apply a polynomial-time algorithm to obtain an ear decomposition 퐺 1 = 퐾2 ⊂ 퐺 2 ⊂ · · · ⊂ 퐺 푟 = 퐺 where, for 푖 = 2, 3, . . . , 푟, 퐺 푖−1 = 퐺 푖 − 푅푖 and 푅푖 is an optimal removable ear of 퐺 푖 . (See Exercise 16.2.8.) We then proceed to obtain a sequence of orientations (퐷 1 , 퐷 2 , . . . , 퐷 푟 ) such that 퐷 푖 is an orientation of 퐺 푖 , for 푖 = 1, 2, . . . , 푟. We arbitrarily orient 퐺 1 , thereby obtaining 퐷 1 . To obtain 퐷 푖 for 푖 = 2, 3, . . . , 푟, we first determine a conformal cycle 푄 푖 of 퐺 푖 that contains all the edges of 푅푖 and then extend 퐷 푖−1 to the orientation 퐷 푖 of 퐺 푖 by orienting the edges of 푅푖 so that 푄 푖 is oddly oriented in 퐷 푖 . To determine the cycle 푄 푖 , we apply a polynomial-time algorithm to determine a perfect matching 푀푖 of 퐺 푖 that contains an edge 푒 푖 of 푅푖 . (See Exercise 2.1.2.) Then, we determine the 푀푖 -alternating cycle of 퐺 that contains an edge 푓푖 adjacent to 푒 푖 . Clearly, 푄 푖 is a conformal cycle of 퐺 푖 that contains all the edges of 푅푖 . The orientation 퐷 1 is a characteristic orientation of 퐺 1, because 퐺 1 is 퐾2 . With the help of Theorem 20.24, it is easy to prove by induction on 푖 that 퐷 푖 is a characteristic orientation of 퐺 푖 , for 푖 = 1, 2, . . . , 푟. The generic case: G is not necessarily matching covered. We first determine whether 퐺 is matchable, using any polynomial-time algorithm which finds a maximum matching of a graph. If 퐺 is not matchable then output an arbitrary orientation of 퐺. We may thus assume that 퐺 is matchable. In polynomial time we then determine the set 푁 of nonmatchable edges of 퐺 (see Exercise 2.1.2).

20.4 Characteristic Orientations

455

Let 퐻 := 퐺 − 푁. Each component of 퐻 is matching covered. Apply the algorithm to each (matching covered) component 퐾 of 퐻, thereby obtaining a characteristic orientation 퐷 퐾 of 퐾. Let 퐷 퐻 := ∪퐾 퐷 퐾 and let 퐷 be an arbitrary extension of 퐷 퐻 to an orientation of 퐺. Let us now see why 퐷 is a characteristic orientation of 퐺. If 퐺 is not Pfaffian then no orientation of 퐺 is Pfaffian. In particular, 퐷 is not Pfaffian. We may thus assume that 퐺 is Pfaffian and we must show that 퐷 is Pfaffian. Let 퐾 be a component of 퐻. As seen in Exercise 19.2.4, 퐾 is Pfaffian. As 퐷 퐾 is a characteristic orientation of 퐾, we deduce that 퐷 퐾 is Pfaffian. This conclusion holds for each component 퐾 of 퐻. Again, as seen in Exercise 19.2.4, 퐷 is Pfaffian. Corollary 20.27 Algorithm 20.26 determines a characteristic orientation of a graph ƒ in polynomial time.

20.4.3 The equivalence of Problems 19.1 and 19.2 Theorem 20.28 The Problem 19.1, of determining whether a given graph is Pfaffian, is polynomial-time equivalent to the Problem 19.2, of deciding whether a given digraph is Pfaffian. Proof Let us first describe a polynomial-time reduction of Problem 19.1, of determining whether a given graph is Pfaffian, to the Problem 19.2, of deciding whether a given digraph is Pfaffian. For this, let 퐺 be a graph. Apply Algorithm 20.26 to graph 퐺, thereby obtaining a characteristic orientation 퐷 of 퐺. The graph 퐺 is Pfaffian if and only if 퐷 is Pfaffian. Let us now describe a polynomial-time reduction of Problem 19.2, of determining whether a given digraph is Pfaffian, to the Problem 19.1, of deciding whether a given graph is Pfaffian. We first give a particular case, in which the given graph is matching covered. This algorithm is then used in the full version, in which the given graph is not necessarily matching covered. Verifying an orientation Algorithm 20.29 Input: an orientation 퐷 of a (not necessarily matching covered) graph 퐺; Output: it either determines that 퐷 may be generated by Algorithm 20.26 or it outputs a cycle which is conformal in 퐺 and evenly oriented in 퐷. A fundamental case: 퐺 is matching covered. Our task is done by Algorithm 20.26, in the case that 퐺 is matching covered, by trying to keep each digraph in the sequence of orientations 퐷 푖 of 퐺 as a subdigraph of 퐷. If, at any stage, it is not possible to extend an orientation 퐷 푖−1 to an orientation 퐷 푖 , it follows that the cycle 푄 푖 , a conformal cycle of 퐺, is evenly oriented in 퐷. In that case,

456

20 Similarity of Orientations and Characteristic Orientations

output 푄 푖 as a conformal cycle of 퐺 which is evenly oriented in 퐷. Alternatively, we succeed in determining that 퐷 may be generated by Algorithm 20.26. The generic case: 퐺 is not necessarily matching covered. As in Algorithm 20.26, determine in polynomial time whether 퐺 is matchable and, if 퐺 is matchable, determine the set 푁 of non matchable edges of 퐺. If 퐺 is not matchable then Algorithm 20.26 outputs an arbitrary orientation of 퐺. In particular, 퐷 may be generated by Algorithm 20.26. We may thus assume that 퐺 is matchable. Let 퐻 := 퐺 − 푁. Each component of 퐻 is matching covered. For each component 퐾 of 퐻, let 퐷 퐾 be the restriction of 퐷 to 퐾. Apply the algorithm to each component 퐾 of 퐻, to verify whether 퐷 퐾 may be generated by Algorithm 20.26. Suppose that the algorithm determines that 퐷 퐾 may be generated by Algorithm 20.26, for each component 퐾 of 퐻. Then, 퐷 is an extension of 퐷 퐻 := ∪퐾 퐷 퐾 , and 퐷 may be generated by Algorithm 20.26. Alternatively, suppose that, for some component 퐾, the algorithm produces a cycle 푄 which is conformal in 퐾 and evenly oriented in 퐷 퐾 . The graph 퐾 is a conformal subgraph of 퐻, which in turn is a conformal subgraph of 퐺. Thus, 푄 is conformal in 퐺. Moreover, 푄 is evenly oriented in 퐷. The cycle 푄 is then the output of the algorithm in this case. The polynomial-time reduction of Problem 19.2 to Problem 19.1 is done by Algorithm 20.29. Let 퐷 be an orientation of a graph 퐺. If the algorithm produces a conformal cycle which is evenly oriented in 퐷, then 퐷 is not Pfaffian. Alternatively, the algorithm determines that 퐷 may be generated by Algorithm 20.26. Then, by Corollary 20.27, 퐷 is a characteristic orientation of 퐺; hence 퐷 is Pfaffian if and ƒ only if 퐺 is Pfaffian. We end this section with an important property of Pfaffian graphs, which will be used in Chapter 21, to justify a polynomial-time algorithm to determine whether a near-bipartite graph is Pfaffian. The inclusion-exclusion property of perfect matchings 20.30 Let 퐺 be a matching covered graph, let 푅 be an equivalence class of the dependence relation on the edges of 퐺, let 퐷 be an orientation of 퐺 and let 푄 be a conformal cycle of 퐺 that contains the edges of 푅 and is oddly oriented in 퐷. Then, 퐷 is Pfaffian if and only if the digraphs 퐷 − 푅 and 퐷 − 푉 (푅) are both Pfaffian, where 푉 (푅) is the set of the ends of the edges of 푅. Proof We may clearly suppose that 퐺 is not 퐾2 ; hence 퐺 contains an edge adjacent to an edge of 푅. The set 푅 is an equivalence class of the dependence relation on the edges of 퐺. Thus, if a perfect matching 푀푅 contains an edge adjacent to an edge of 푅 then it does not contain any edge of 푅 and is thus a perfect matching of 퐺 − 푅. Alternatively, if a perfect matching 푀푅 contains an edge of 푅 then it contains all the edges of 푅. We deduce that the collection M of perfect matchings of 퐺 can be partitioned in two sets, M 푅 and M 푅 , respectively the collection of perfect matchings

20.4 Characteristic Orientations

457

of 퐺 that contain no edge of 푅 and the collection of those perfect matchings that contain all the edges of 푅. The graph 퐺 − 푅 is thus a (spanning) matchable subgraph of 퐺; hence 퐺 − 푅 is a conformal subgraph of 퐺. Likewise, for any matching 푀 in M 푅 , the set 푀 − 푅 is a perfect matching of 퐺 − 푉 (푅); hence the graph 퐺 − 푉 (푅) is a matchable subgraph of 퐺. Moreover, the set 푅 is a perfect matching of 퐺 − (퐺 − 푉 (푅)). Thus, 퐺 − 푉 (푅) is also a conformal subgraph of 퐺. In sum, the graphs 퐺 − 푅 and 퐺 − 푉 (푅) are both conformal subgraphs of 퐺. If 퐷 is Pfaffian then 퐷 − 푅 and 퐷 − 푉 (푅) are both Pfaffian. To prove the converse, suppose that 퐷 − 푅 and 퐷 − 푉 (푅) are both Pfaffian. Let 푀 be a perfect matching of 퐺. If 푀 is a perfect matching of 퐺 − 푅, that is, if 푀 ∈ M 푅 , then, as 퐷 − 푅 is Pfaffian, we infer that every 푀-alternating cycle of 퐺 − 푅 is oddly oriented in 퐷; hence all the perfect matchings of 퐺 − 푅 have the same sign, say 푠, in 퐷. Likewise, if 푀 ∈ M 푅 , then, as 푀 − 푅 is a perfect matching of 퐺 − 푉 (푅), it follows that every 푀-alternating cycle of 퐺 − 푉 (푅) is oddly oriented in 퐷; hence all the perfect matchings of 퐺 − 푉 (푅) have the same sign, say 푡, in 퐷. Finally, the cycle 푄 is conformal in 퐺, contains the edges of 푅 and is oddly oriented in 퐷. As 푄 is conformal, it is an 푀-alternating cycle for some perfect matching, 푀, of 퐺. The set 푁 := 푀 △ 퐸 (푄) is a perfect matching of 퐺. As 푄 is oddly oriented in 퐷, the matchings 푀 and 푁 have the same sign in 퐷. One of 푀 and 푁 is in M 푅 and has sign 푠 in 퐷, the other is in M 푅 and has sign 푡 in 퐷. We deduce that 푠 = 푡. This conclusion implies that all the perfect matchings of 퐺 have the same sign in 퐷; hence 퐷 is Pfaffian. ƒ

Exercises ⊲20.4.1 In the statement of Theorem 20.24, the ear 푅 of 퐺, if it is a single ear, is required to be (푏 + 푝)-invariant. In fact, if it is not (푏 + 푝)-invariant, then one of two things may happen: either 푝(퐺 − 푅) > 0 or 푝(퐺 − 푅) = 0 but 푏(퐺 − 푅) > 푏(퐺). In the first alternative, 퐺 is not Pfaffian, by Corollary 20.15, because Petersen bricks are not Pfaffian. In the second alternative, 푅 is not 푏-invariant. (i) Let 퐺 be the simple graph obtained from the Petersen graph by the addition of an edge, 푒. The edge 푒 is 푏-invariant, but not (푏 + 푝)-invariant. Using Algorithm 20.26 , determine a characteristic orientation 퐷 ′ of 퐺 − 푒, and extend it to a characteristic orientation 퐷 of 퐺. Verify that 퐷 is not Pfaffian. (ii) The orientation 퐷 of the brick 퐺 depicted in Figure 20.5 is Pfaffian. The edge 푒 is not 푏-invariant in 퐺. The pair 푅 := { 푓 , 푔} is a removable doubleton of the graph 퐺 − 푒. Thus, 퐺 − 푒 has two characteristic orientations: one is 퐷 − 푒 and the other is (퐷 −푒) rev 푅. Verify that the second characteristic orientation of 퐺 − 푒 is not extendable to a Pfaffian orientation of 퐺. Hint: extend (퐷 − 푒) rev 푅 to an orientation 퐷 ′ of 퐺 by arbitrarily orienting edge 푒 and then find two conformal cycles of 퐺, both containing edge 푒 and having distinct signs in 퐷 ′ . 20.4.2 Find characteristic orientations of 퐾3,3 and P and verify that they are not Pfaffian.

458

20 Similarity of Orientations and Characteristic Orientations 푔

푓

푒

Fig. 20.5 The brick 퐺 and its orientation 퐷

20.4.3 Find maximal collections of dissimilar Pfaffian orientations of those graphs that represent the five Platonic polyhedra: the tetrahedron, the cube, the octahedron, the dodecahedron and the icosahedron. ⊲20.4.4 Let 퐺 be a matching covered graph 퐺. Recall that we denote the numbers of vertices, edges, bricks, Petersen bricks, and the number of perfect matchings of 퐺 by 푛, 푚, 푏, 푝 and Φ, respectively. We say that 퐺 is Pfaffian-extremal if Φ(퐺) = 푚 − 푛 + 2 − (푏 + 푝). For example, for each edge 푒, the graphs 퐾3,3 − 푒 and P − 푒 are both Pfaffian-extremal. (i) Prove that if Φ(퐺) ≤ 푚 − 푛 + 2 − (푏 + 푝) then in fact 푝 = 0 and equality holds. (Recall, from Chapter 6, that Φ(퐺) ≥ 푚 − 푛 + 2 − 푏). (ii) Prove that if 퐺 is Pfaffian-extremal and 푅 is an optimal removable ear of 퐺, then 퐺 − 푅 is Pfaffian-extremal. (iii) Prove that every Pfaffian-extremal graph is Pfaffian and conclude that 퐾3,3 − 푒 and P − 푒 are Pfaffian, for each edge 푒. (Consider an optimal removable ear decomposition of 퐺.) ⊲20.4.5 Prove that Problem 19.1 is in co-N P. More precisely, show that given a non-Pfaffian digraph 퐷, there is a succinct certificate to show that 퐷 is not Pfaffian, which can be checked in polynomial-time. ⊲20.4.6 Prove that Problem 19.2 is in co-N P. More precisely, show that, given a non-Pfaffian matching covered graph 퐺, there is a succinct certificate to show that 퐺 is not Pfaffian, which can be checked in polynomial-time. Hint: use the result of Exercise 20.4.5 and Algorithm 20.26.

20.5 Notes

459

20.5 Notes The reduction of Pfaffian matching covered graphs to Pfaffian bricks and braces was first proved by Little and Rendl (1991, [55]); see Theorem 20.10. The proof we give here is based on similarity of orientations and was published by CLM more than 20 years later (2012, [16]). The number of dissimilar Pfaffian orientations of a matching covered graph was published by CLM (2005, [14]). The equivalence of Problems 19.1 and 19.2 was published by Vazirani and Yannakakis (1989, [91]).

Chapter 21

Excluded-Minor Characterizations of Pfaffian Graphs

Contents 21.1 21.2

21.3 21.4

21.5

21.6

21.7

푆-Minors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 461 푆-Minimal non-Pfaffian Graphs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 464 21.2.1 Properties of non-Pfaffian Graphs . . . . . . . . . . . . . . . . . . . . . 465 21.2.2 Monochromatic chords of even cycles . . . . . . . . . . . . . . . . . 468 Pfaffian Bipartite graphs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 471 Pfaffian Near-Bipartite Graphs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 473 21.4.1 Some immediate implications of the induction hypothesis . 473 21.4.2 Removable edges of 퐻 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 475 21.4.3 Degree of vertices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 475 21.4.4 There is a unique (푀1 , 푀2 )-alternating cycle in 퐺 . . . . . . . 475 21.4.5 Quadrilaterals in 퐻 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 476 21.4.6 Triangles in 퐺 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 477 21.4.7 The equality 푀1 ∩ 푀2 = 푅 = {푒, 푓 } . . . . . . . . . . . . . . . . . . . 479 21.4.8 The graphs Γ1 and Γ2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 483 Polynomial-time Algorithms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 485 21.5.1 Efficient recognition of Pfaffian bipartite graphs . . . . . . . . . 486 21.5.2 Efficient recognition of Pfaffian near-bipartite graphs . . . . . 489 An Infinite Family of Minimal Non-Pfaffian Graphs . . . . . . . . . . . . . 491 21.6.1 푁푇-minors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 491 21.6.2 Non-Pfaffian 푁푇-minimal graphs . . . . . . . . . . . . . . . . . . . . . 492 21.6.3 The Norine and Thomas family N T . . . . . . . . . . . . . . . . . . 492 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 495

21.1 푺-Minors As noted in Chapter 19, Kasteleyn (1963, [44]) not only pioneered the theory of Pfaffian orientations, but also established the first definitive result in this theory by showing that all planar graphs are Pfaffian. © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7_21

461

462

21 Excluded-Minor Characterizations of Pfaffian Graphs

By Corollary 20.16, a matching covered graph is Pfaffian if and only if all its matching minors are Pfaffian. Thus any graph which admits a known non-Pfaffian graph (such as 퐾3,3 ) as a matching minor is also non-Pfaffian. Little (1975, [54]) advanced the theory of Pfaffian orientations by showing that a bipartite matching covered graph is Pfaffian if and only if it does not have a matching minor isomorphic to 퐾3,3 . Some years later, in collaboration with Fischer (2001, [34]), Little participated in another advance by characterizing near-bipartite graphs which admit a Pfaffian orientation. A proper description of this characterization requires the notion of a minor that is more general than that of a matching minor. This general notion, which plays a crucial role in this chapter, is defined in the inset below: Definition: 푆-minor A separation-deletion minor of a matching covered graph 퐺, or simply an 푆-minor of 퐺, is a graph that is obtainable from 퐺, up to isomorphism, by means of deletions of removable classes and contractions of shores of separating cuts. In other words, 퐻 is an 푆-minor of 퐺 if there exists a sequence (퐺 1 , 퐺 2 , . . . , 퐺 푟 )

(푟 ≥ 1)

of graphs such that, 퐺 1 = 퐺, 퐺 푟  퐻 and, for 1 ≤ 푖 ≤ 푟 − 1, the graph 퐺 푖+1 is obtained from 퐺 푖 by either deleting a removable class or by contracting a shore of a separating cut to a single vertex. (We allow the possibility that 푟 = 1; that is, we regard a graph 퐺 to be an 푆-minor of itself.) As an example, consider the sequence (퐺 1 , 퐺 2 , 퐺 3 , 퐺 4 ) of graphs in Figure 21.1. The graph 퐺 1 is a nonsolid brick and the cut 퐶 is a (nontight) separating cut of 퐺 1 . The graph 퐺 2 is a 퐶-contraction of 퐺 1 ; it is a brick (which happens to be the solid brick Σ8 shown in Figure 7.2), and 푒 is a removable edge in it. The cut 퐶 in 퐺 3 := 퐺 2 − 푒 is a separating cut (in fact, a tight cut) and 퐺 4 is obtained from 퐺 3 by contracting one of the shores of 퐶. Thus 퐺 1 , 퐺 2 , 퐺 3 and 퐺 4 are 푆-minors of 퐺 1 . (Note that 퐾4 is also an 푆-minor of 퐺 1 .) It follows from the top-down version of the Ear Decomposition Theorem 11.3 that every conformal subgraph of a matching covered graph 퐺 is an 푆-minor of 퐺. Consequently, every matching minor of 퐺 is also an 푆-minor of 퐺. But not every 푆-minor of 퐺 is a matching minor of 퐺. For example, 퐾3,3 is an 푆-minor of 퐺 3 , but it is not a matching minor of that graph (see Exercise 12.1.6). From Corollary 19.6 and Lemma 20.13, we may now deduce the following important property of Pfaffian graphs, whose proof is left as Exercise 21.1.1. Theorem 21.1 A matching covered graph is Pfaffian if and only if all its 푆-minors are Pfaffian. ƒ

21.1 푆-Minors

463

퐶

퐶 푒

푒 (b) 퐺2 (Σ8 )

(a) 퐺1

(c) 퐺3

(d) 퐺4 Fig. 21.1 푆-minors

In light of the above theorem, to show that a given graph is non-Pfaffian, it suffices to produce an 푆-minor of that graph which is known to be non-Pfaffian. For example, since 퐾3,3 is known to be non-Pfaffian, and since it is a 푆-minor of each of the graphs in Figure 21.1, we may conclude that all of them are non-Pfaffian. This leads us to the notion of an 푆-minimal non-Pfaffian graph. A non-Pfaffian matching covered is 푆-minimal if each of its proper 푆-minors is Pfaffian. In Section 21.2 we describe some properties of minimal Pfaffian graphs which are pertinent to characterizations of Pfaffian bipartite and Pfaffian near-bipartite graphs presented in Sections 21.3 and 21.4, respectively. In the latter section we show that the brace 퐾3,3 is the only 푆-minimal non-Pfaffian bipartite graph and the two bricks Γ1 and Γ2 depicted in Figures 21.2 and 21.3, respectively, are the only 푆-minimal non-Pfaffian near-bipartite graphs. This, in essence, is the theorem of Fischer and Little. (In their paper they characterize all matching minor-minimal non-Pfaffian near-bipartite graphs.) 푓 1 8

2 푥1 푒

푦2

7

3 푥2

6

4 5

Fig. 21.2 The brick Γ1

푦1

21 Excluded-Minor Characterizations of Pfaffian Graphs

464

푓 1 8

2 푥1 푒

푦2

7

3

푦1

푥2

6

4 5

Fig. 21.3 The brick Γ2

Exercises 21.1.1 Prove Theorem 21.1. 21.1.2 Prove that every non-Pfaffian matching covered graph has order six or more. ∗ 21.1.3 The objective of this exercise is to prove that the cube P8 is the only simple brace of order eight that is Pfaffian. (i) Prove that the cube is the only cubic brace of order eight. (ii) Prove that the cube is a subgraph of each brace of order eight. (iii) Prove that every simple brace of order eight distinct from the cube is nonPfaffian. ⊲21.1.4 Prove that the graphs Γ1 and Γ2 are non-Pfaffian. Hint: for 푖 = 1, 2, find a characteristic orientation of Γ푖 − 푒 − 푓 , using Algorithm 20.26 and extend it to a characteristic orientation 퐷 푖 of Γ푖 . Verify that the cycle 푄 := (1, 2, . . . , 8, 1) is conformal in Γ푖 and evenly oriented in 퐷 푖 . (For 푖 = 1, 2: the graph Γ푖 − 푒 − 푓 has four vertices of degree two; the retract 퐻푖 of Γ푖 − 푒 − 푓 is 퐶4 , up to multiple edges; extend any Pfaffian orientation of 퐻푖 to an orientation of Γ푖 − 푒 − 푓 , by ensuring that each vertex of degree two is neither a source nor a sink.)

21.2 푺-Minimal non-Pfaffian Graphs Motivated by Theorem 21.1, we now introduce the important notion of an 푆-minimal minor of a non-Pfaffian graph.

21.2 푆-Minimal non-Pfaffian Graphs

465

Definition: 푆-minimal non-Pfaffian minors A non-Pfaffian 푆-minor of given matching covered graph is separation-deletion minimal, or simply 푆-minimal, if all its proper 푆-minors are Pfaffian. For example, suppose that the non-Pfaffian brace 퐾3,3 is an 푆-minor of some graph 퐺. Then 퐾3,3 is an 푆-minimal minor of 퐺 because all proper 푆-minors of 퐾3,3 , being planar, are Pfaffian. It follows from Corollary 20.15 that every 푆-minimal non-Pfaffian matching covered graph is either a brick or a brace. In addition, it is easy to see that such a graph is also simple. The only simple braces of order four or less are 퐾2 and 퐶4 . The only simple brick of order four is 퐾4 . It is easy to deduce that every non-Pfaffian graph has order at least six (Exercise 21.1.2). These observations imply: Proposition 21.2 All non-Pfaffian 푆-minimal minors of a non-Pfaffian graph are simple and are either bricks or braces of order six or more. ƒ We shall assume the properties of 푆-minimal minors stated in the above proposition without explicitly stating them.

21.2.1 Properties of non-Pfaffian Graphs The objective of this subsection is to establish a number of basic properties of non-Pfaffian matching covered graphs. Theorem 21.3 Let 퐷 be an orientation of a non-Pfaffian matching covered graph 퐺 and let 푅 be a removable class of 퐺. If 퐷 − 푅 is Pfaffian then the perfect matchings of 퐺 that contain 푅 do not all have the same sign. Proof As 퐷 − 푅 is a Pfaffian orientation of 퐺 − 푅, all perfect matchings of 퐺 − 푅 have the same sign, say 푠, in 퐷. If all the perfect matchings containing all the edges of 푅 also have sign 푠 in 퐷, then 퐷 itself would be a Pfaffian orientation of 퐺. And, if all the perfect matchings containing all the edges of 푅 have sign −푠 in 퐷, the digraph obtained from 퐷 by reversing the orientation of one of the edges of 푅 would be a Pfaffian orientation of 퐺. Both these cases are impossible because, by hypothesis, 퐺 is non-Pfaffian. We conclude that the perfect matchings of 퐺 that contain 푅 do not all have the same sign. ƒ In the following inset we introduce the notion of an 푅-triple which will play a crucial role in our proofs of the two theorems which provide characterizations of bipartite graphs and of near-bipartite graphs.

466

21 Excluded-Minor Characterizations of Pfaffian Graphs

푅-triples and Their Significance Definition: Let 푅 be a removable class of a matching covered graph 퐺, let 퐷 be an orientation of 퐺 such that 퐷 − 푅 is Pfaffian and let 푀1 and 푀2 be perfect matchings of 퐺 both containing all the edges of 푅, and having distinct signs in 퐷. Then, the orientation 퐷 of 퐺 and the perfect matchings 푀1 and 푀2 of 퐺 constitute, in that order, an 푅-triple. We denote such a triple by the ordered set (퐷, 푀1 , 푀2 ). If 푅 is a singleton, say, {푒}, then we simply refer to the triple (퐷, 푀1 , 푀2 ) as an 푒-triple. The notion of an R-triple plays a significant role in our presentations of the proofs of the above mentioned theorems. This, as we shall see below, is due to the fact that if 푅 happens to be an optimal removable class, then an associated 푅-triple serves as a certificate which demonstrates that 퐺 is non-Pfaffian. Lemma 21.4 Let 퐺 be a matching covered graph and let 푅 be an optimal removable class of 퐺. If 퐺 has an 푅-triple then 퐺 is not Pfaffian. Proof Suppose that 퐺 has an 푅-triple, (퐷, 푀1 , 푀2 ). By definition, the orientation 퐷 − 푅 of 퐺 − 푅 is Pfaffian, the two perfect matchings 푀1 and 푀2 have distinct signs in 퐷, and 푅 ⊆ 푀1 ∩ 푀2 . Reversal of the orientation of any edge of 푅 changes the sign of every perfect matching of 퐺 that contains that edge of 푅. It follows that 푀1 and 푀2 have distinct signs in every orientation of 퐺 that is an extension of the orientation 퐷 − 푅 of 퐺 − 푅, implying that no extension of the Pfaffian orientation 퐷 − 푅 of 퐺 − 푅 to an orientation of 퐺 could be Pfaffian. On the other hand, as 푅 is an optimal class, by Corollary 20.25, if 퐺 were Pfaffian, it should be possible to extend any Pfaffian orientation of 퐺 − 푅 to a Pfaffian orientation of 퐺. We conclude that 퐺 is not Pfaffian. ƒ The existence of an 푅-triple, without the requirement that 푅 is an optimal removable class of 퐺, does not necessarily imply that 퐺 is not Pfaffian. See Exercise 21.2.3. Example 21.5 Consider the orientation 퐷 of 퐾3,3 shown in Figure 19.1. If we take 푅 to consist of the edge ℎ := 56, then 퐷 − ℎ is a Pfaffian orientation of 퐾3,3 − ℎ. The triple (퐷, 푀1 , 푀2 ) is an ℎ-triple, where 푀1 := {12, 34, 56} and 푀2 := {14, 23, 56}, because 푀1 △ 푀2 induces an evenly oriented cycle. This is yet another proof that 퐾3,3 is not Pfaffian! Example 21.6 As a second example, consider the brick Γ1 with the orientation 퐷 depicted in Figure 21.4. The pair 푅 := {푒, 푓 } is a removable doubleton of Γ1 , where 푒 = 푥1 푥2 and 푓 = 푦 1 푦 2 . The digraph 퐷 − 푅 is Pfaffian. (See Exercise 21.2.1.) The perfect matchings 푀1 := {12, 34, 56, 78, 푥1 푥2 , 푦 1 푦 2 } and 푀2 := {23, 45, 67, 81, 푥1 푥2 , 푦 1 푦 2 } have distinct signs in 퐷, as the (푀1 , 푀2 )-alternating cycle (1, 2, 3, 4, 5, 6, 7, 8, 1) is evenly oriented. Thus, (퐷, 푀1 , 푀2 ) is an 푅-triple corresponding to the optimal removable class 푅. It follows that Γ1 is not Pfaffian.

21.2 푆-Minimal non-Pfaffian Graphs

467 푓 1

8

2 푥1 푒

푦2

7

3

푦1

푥2

6

4 5

Fig. 21.4 The orientation 퐷 of Γ1

Lemma 21.7 (Exchange Property: optimal classes) Let 퐺 be a non-Pfaffian matching covered graph, let 푅 be an optimal removable class of 퐺 and let (퐷, 푀1 , 푀2 ) be an 푅-triple. Suppose that 퐺 − 푅 has an optimal removable class 푆 which is disjoint with 푀1 ∪ 푀2 . Then, 푆 is an optimal removable class of 퐺, 푅 is an optimal removable class of 퐺 − 푆 and the graph 퐺 − 푆 is not Pfaffian. Proof By definition of 푅-triple, 푅 ⊆ 푀1 ∩ 푀2 . Thus, 푅 ⊆ 푀1 and 푀1 is a perfect matching of 퐺 − 푆. From Lemma 16.13, with 푀1 playing the role of 푀, we infer that 푆 is an optimal removable class of 퐺 and 푅 is an optimal removable class of 퐺 − 푆. By definition of 푅-triple, the digraph 퐷 − 푅 is Pfaffian; hence 퐷 − 푅 − 푆 is also Pfaffian. The sets 푀1 and 푀2 are perfect matchings of 퐺 − 푆 that have distinct signs in 퐷. Thus, 푀1 and 푀2 are perfect matchings of 퐺 − 푆 that have distinct signs in 퐷 − 푆. Moreover, the optimal removable class 푅 of 퐺 − 푆 is a subset of 푀1 ∩ 푀2 . In sum, (퐷 − 푆, 푀1 , 푀2 ) is an 푅-triple of 퐺 − 푆. By Lemma 21.4, the graph 퐺 − 푆 is not Pfaffian. ƒ In a bipartite graph all removable classes are singletons and are optimal (푏invariant). Based on this observation, it is easy to deduce the following corollary of the above lemma. It plays an important role in our proof of Little’s Theorem in the next section. Corollary 21.8 (Exchange Property: bipartite graphs) Let 퐺 be a bipartite nonPfaffian matching covered graph, let 푒 be a removable edge of 퐺, and let (퐷, 푀1 , 푀2 ) be an 푒-triple. Suppose that 퐺 − 푒 has a removable edge 푓 which does not belong to 푀1 ∪ 푀2 . Then 푓 is a removable edge of 퐺 and 퐺 − 푓 is not Pfaffian. ƒ Suppose that 퐺 is a non-Pfaffian graph, 푅 is a removable class of 퐺, and that (퐷, 푀1 , 푀2 ) is an 푅-triple of 퐺. Then, by definition, the two perfect matchings 푀1 and 푀2 have different signs in 퐷. It follows from Corollary 19.4 that there are an odd number of (푀1 , 푀2 )-alternating cycles which are evenly oriented in 퐷. In our approach to characterizations of 푆-minimal bipartite and near-bipartite non-Pfaffian

468

21 Excluded-Minor Characterizations of Pfaffian Graphs

graphs, it turns out to be convenient to ascertain that there is precisely one evenly oriented (푀1 , 푀2 )-alternating cycle. The following lemma describes a condition under which this is the case. Lemma 21.9 Let 퐺 be a non-Pfaffian matching covered graph and let 푅 be a removable class of 퐺. Let (퐷, 푀1 , 푀2 ) be an 푅-triple of 퐺. If 퐺 has two or more (푀1 , 푀2 )-alternating cycles then there exists an 푅-triple (퐷, 푀1′ , 푀2 ) such that 퐺 has an (푀1′ ∩ 푀2 )-alternating cycle. Proof Let C denote the collection of all (푀1 , 푀2 )-alternating cycles of 퐺. Suppose that C consists of two or more cycles. As (퐷, 푀1 , 푀2 ) is an 푅-triple, the two perfect matchings 푀1 and 푀2 have distinct signs in 퐷. By Corollary 19.4, it follows that the collection C includes an odd number of evenly oriented cycles (and it may also include some oddly oriented cycles). Let 푄 be an evenly oriented (푀1 , 푀2 )alternating cycle in C. Let 푇 := ∪퐶 ∈ C−푄 퐸 (퐶) and let 푀1′ := 푀1 △ 푇. It is easy to see that 푀1′ is a perfect matching of 퐺 and that 푄 is the only (푀1′ , 푀2 )-alternating cycle in 퐺. Thus, once again, by Corollary 19.4 it follows that 푀1′ and 푀2 have distinct signs in 퐷. Furthermore those edges which belong to 푀1 ∩ 푀2 also belong to 푀1′ ∩ 푀2 . Thus, as 푅 ⊆ 푀1 ∩ 푀2 , it follows that 푅 ⊆ 푀1′ ∩ 푀2 . In sum, 퐷 is a non-Pfaffian orientation of 퐺, the set 푅 is a removable class of 퐺, the digraph 퐷 − 푅 is Pfaffian, the perfect matchings 푀1′ and 푀2 have distinct signs in 퐷 and 푅 ⊆ 푀1′ ∩ 푀2 . Thus, (퐷, 푀1′ , 푀2 ) ƒ is also an 푅-triple of 퐺. The assertion holds. It is important to understand the significance of the above lemma that does not depend on the choice of notation. This assertion relates two properties an 푅-triple (퐷, 푀1 , 푀2 ) of a non-Pfaffian graph 퐺 may have: (i) the number of (푀1 , 푀2 )alternating cycles in 퐺 is just one; and (ii) there are (푀1 ∩ 푀2 )-alternating cycles in 퐺. What the lemma says is that if there is no 푅-triple of 퐺 with the second property, then all 푅-triples of graph 퐺 have the first property. As an example, let 푒 = 푎 1 푏 1 be a removable edge in a minimal non-Pfaffian bipartite graph 퐺. We shall show in the next section that, if (퐷, 푀1 , 푀2 ) is any arbitrary 푒-triple of 퐺, then 푀1 ∩ 푀2 = {푒}. This clearly implies that 퐺 has no (푀1 ∩ 푀2 )-alternating cycles. By the above lemma, it follows that if (퐷, 푀1 , 푀2 ) is any particular 푒-triple of 퐺, then 퐺 has precisely one (푀1 , 푀2 )-alternating cycle 푄 and that cycle is a spanning cycle of 퐺 − 푎 1 − 푏 1 .

21.2.2 Monochromatic chords of even cycles Let 푄 be an even cycle. A chord 푒 of 푄 is monochromatic if both ends of 푒 are in the same part of the bipartition of 푄. In other words, 푒 is monochromatic if and only if 푄 + 푒 is not bipartite. An 8-cycle in the Petersen graph, for example, has two chords and both of them are monochromatic. A Hamilton cycle in the pentagonal prism has five chords; two of them are monochromatic and the other three are not. The following result establishes conditions under which all chords of an even cycle in a non-Pfaffian graph are monochromatic.

21.2 푆-Minimal non-Pfaffian Graphs

469

Lemma 21.10 Let 퐺 be a non-Pfaffian simple matching covered graph and let 푅 be a removable class of 퐺 with the property that if (퐷, 푀1′ , 푀2′ ) is any 푅-triple of 퐺, then 푀1′ ∩ 푀2′ = 푅. Now let (퐷, 푀1 , 푀2 ) be an 푅-triple of 퐺, and let 푄 be any (푀1 , 푀2 )-alternating cycle in 퐺 which is evenly oriented in 퐷. Then, every chord of 푄 in 퐺 is monochromatic. Proof Assume, to the contrary, that 푄 has a chord, say 푢푤, which is not monochromatic. See Figure 21.5. 푢 푄 푀1 푄1

푄2

푀2

푤 Fig. 21.5 The chord 푢푤 of cycle 푄

As 푄 is bipartite, the vertices 푢 and 푤 define two segments of 푄 of odd length. And, as 퐺 is simple, each such segment has an odd length equal to three or more. These segments, plus the chord 푢푤, define two cycles, 푄 1 and 푄 2 . Using the fact that 푄 is evenly oriented in 퐷, it is easy to deduce that one of 푄 1 and 푄 2 is oddly oriented and the other is evenly oriented (Exercise 21.2.4(i)). Moreover, one of 푄 1 and 푄 2 is 푀1 -alternating and the other is 푀2 -alternating. Adjust notation so that 푄 1 is 푀1 alternating and oddly oriented. Let 푀1′ := 푀1 △ 퐸 (푄 1 ). Then, sign(푀1′ ) = sign(푀1 ) (Exercise 21.2.4(ii)). Moreover, 푅 ⊆ 푀1′ ∩ 푀2 ; hence (퐷, 푀1′ , 푀2 ) is an 푅-triple. As 푄 2 has length four or more, we deduce that (푀1′ ∩ 푀2 ) − 푅 is nonempty. This contradicts the hypothesis. It follows that every chord of 푄 is monochromatic. ƒ In the special case where 퐺 is a simple non-Pfaffian bipartite graph and 푅 = {푒} is a removable singleton of 퐺, the hypothesis of the above lemma implies that 푄 has no chords at all because the ends of any edge in 퐺 belong to different colour classes.

Exercises ⊲21.2.1 The objective of this exercise is to prove that Γ1 is an 푆-minimal non-Pfaffian graph. For the proof of some properties of Γ1 , it is convenient to note that the following permutations are automorphisms of Γ1 :

470

21 Excluded-Minor Characterizations of Pfaffian Graphs

(1 5) (2 4) (6 8) (3) (7) (푥1 ) (푥2 ) (푦 1 ) (푦 2 ) (1) (5) (2 8) (3 7) (4 6) (푥1 ) (푥2 ) (푦 1 푦 2 ) (i) Let 퐽 be the retract of the graph 퐻 := Γ1 − 푒 − 푓 . The underlying simple graph of 퐽 is a quadrilateral. Find a Pfaffian orientation 퐷 퐽 of 퐽. (ii) Extend 퐷 퐽 to a Pfaffian orientation 퐷 퐻 of 퐻. Hint: apply the result of Exercise 20.2.1 to each vertex of degree two of 퐻, ensuring that each of the four vertices of degree two in 퐻 is neither a source nor a sink in 퐷 퐻 . (iii) Prove that Γ1 is not Pfaffian. (The cycle (1, 2, . . . , 8, 1) is evenly oriented and conformal in Γ1 ). (iv) Prove that for 푔 ∈ {12, 23, 1푥1 , 3푥2 , 2푦 1 } the graph Γ1 −푔 is Pfaffian. Taking into account the two automorphisms given above, conclude that Γ1 − 푔 is Pfaffian, for each removable edge 푔 of Γ1 . (v) Let 푄 be a cycle in {(푥1 , 푥2 , 3, 2, 1, 푥1 ), (푦 1 , 푦 2 , 8, 1, 2, 푦 1 )}. Prove that Γ1 /푉 (푄) is Pfaffian. Taking into account the two automorphisms given above, conclude that for each nontrivial separating cut 퐶 of Γ1 , both 퐶-contractions of Γ1 are Pfaffian. ⊲21.2.2 The objective of this exercise is to prove that Γ2 , depicted in Figure 21.3, is an 푆-minimal non-Pfaffian matching covered graph. Hint: it is convenient to notice that (1 2 3 4 5 6 7 8) (푥1 푦 1 푥2 푦 2 ) is an automorphism of Γ2 . (i) Let 퐽 be the retract of the graph 퐻 := Γ2 − 푒 − 푓 . The underlying simple graph of 퐽 is a quadrilateral. Find a Pfaffian orientation 퐷 퐽 of 퐽. (ii) Extend 퐷 퐽 to a Pfaffian orientation 퐷 퐻 of 퐻. Hint: apply the result of Exercise 20.2.1 to each vertex of degree two of 퐻, ensuring that each of the four vertices of degree two in 퐻 is neither a source nor a sink in 퐷 퐻 . (iii) Prove that Γ2 is not Pfaffian. (The cycle (1, 2, . . . , 8, 1) is evenly oriented and conformal in Γ2 ). (iv) Prove that for 푔 ∈ {12, 1푥1 } the graph Γ2 − 푔 is Pfaffian. Taking into account the automorphism given above, conclude that Γ2 − 푔 is Pfaffian, for each removable edge 푔 of Γ2 . (v) Let 푄 be the cycle (푥1 , 푥2 , 3, 2, 1, 푥1 ). Prove that Γ2 /푉 (푄) is Pfaffian. Taking into account the automorphism given above, conclude that for each nontrivial separating cut 퐶 of Γ2 , both 퐶-contractions of Γ2 are Pfaffian. ⊲21.2.3 Lemma 21.4 requires that the removable class 푅 be an optimal removable class. Give an example of a Pfaffian graph 퐺, a removable class 푅 of 퐺 and an 푅-triple of 퐺. 21.2.4 Justify the following claims made in the proof of Lemma 21.10: (i) the cycles 푄 1 and 푄 2 have distinct parities in 퐷; (ii) the signs of the two perfect matchings 푀1′ and 푀1 are the same.

21.3 Pfaffian Bipartite graphs

471

21.3 Pfaffian Bipartite graphs In this section we present a proof of the classical result of Little on non-Pfaffian bipartite graphs. Little’s Theorem [54] 21.11 The brace 퐾3,3 is the only 푆-minimal non-Pfaffian bipartite matching covered graph. Proof We have seen that 퐾3,3 is not Pfaffian (Lemma 19.10). Furthermore, all proper 푆-minors of 퐾3,3 , being planar, are Pfaffian by Kasteleyn’s Theorem (19.15). Thus 퐾3,3 is an 푆-minimal non-Pfaffian bipartite matching covered graph. Now let 퐺 be any 푆-minimal non-Pfaffian bipartite matching covered graph. These properties of 퐺, as we have already noted (right after the definition of 푆-minimal non-Pfaffian minors), imply that 퐺 is a simple brace of order six or more. To establish the desired assertion, we must prove that 퐺 is, in fact, the brace 퐾3,3 . Towards this end, we first note that, by Theorem 8.3, every edge of 퐺 is removable. The following assertion may be deduced using Theorem 8.4 (see Exercise 8.4.1). 21.11.1 For every (removable) edge 푒 of 퐺 and for every vertex 푎 of degree three or more in 퐺 − 푒, at most one edge of 퐺 − 푒 incident with 푎 is not removable in 퐺 − 푒. ƒ Let 푒 be a (removable) edge of 퐺. The minimality of 퐺 implies that 퐺 − 푒 is Pfaffian. By Theorem 21.3, 퐺 has 푒-triples. Let (퐷, 푀1 , 푀2 ) be an 푒-triple of 퐺. 21.11.2 Every removable edge of 퐺 − 푒 is in 푀1 ∪ 푀2 . Proof Assume that 퐺 − 푒 has a removable edge, say 푓 , which is not in 푀1 ∪ 푀2 . By Corollary 21.8, edge 푓 is removable in 퐺 and 퐺 − 푓 is non-Pfaffian. This is in contradiction to the assumption that 퐺 is a minimal non-Pfaffian graph. ƒ 21.11.3 The graph 퐺 is cubic. Moreover, 푀1 ∩ 푀2 = {푒}. Proof Let 푣 be any vertex of 퐺. As 퐺 is a brace of order six or more, vertex 푣 has degree three or more. We consider two cases depending whether or not 푣 is incident with the edge 푒. Case 1 Vertex 푣 is incident with 푒. In this case, as 푒 ∈ 푀1 ∩ 푀2 , we infer that 휕 (푣) − 푒 and 푀1 ∪ 푀2 are disjoint. By statement 21.11.2, no edge of 휕 (푣) − 푒 is removable in 퐺 − 푒. By statement 21.11.1, 푣 has degree two in 퐺 − 푒; hence 푣 has degree three in 퐺.

21 Excluded-Minor Characterizations of Pfaffian Graphs

472

Case 2 Vertex 푣 is not incident with 푒. In this case, 푣 has degree three or more in 퐺 − 푒. By statement 21.11.1, at most one edge of 휕 (푣) is not removable in 퐺 − 푒; all the other edges of 휕 (푣) are removable in 퐺 − 푒. By statement 21.11.2, all the edges of 휕 (푣) removable in 퐺 − 푒 are in 푀1 ∪ 푀2 . We deduce that 푣 has degree three in 퐺. Moreover, 푣 is incident with two edges in 푀1 ∪ 푀2 ; hence 푣 is not incident with an edge in 푀1 ∩ 푀2 . In both cases, these conclusions hold for each vertex 푣 of 퐺 not incident with 푒. Indeed, 퐺 is cubic and 푒 is the only edge in 푀1 ∩ 푀2 . ƒ Let 푎 1 and 푏 1 be the ends of the edge 푒.

푏2

푏1

푒

푎2

푎3

푎1

푏3 Fig. 21.6 The graph 퐾3,3

21.11.4 The set 푀1 △ 푀2 induces a Hamilton cycle 푄 in the graph 퐺 − 푎 1 − 푏 1 . Moreover, 푄 has no chords. Proof By statement 21.11.3, 푀1 ∩ 푀2 = {푒}. This equality holds for each 푒-triple (퐷, 푀1 , 푀2 ) of 퐺. Thus, for any 푒-triple (퐷, 푀1 , 푀2 ), graph 퐺 is free of (푀1 ∩ 푀2 )alternating cycles. By Lemma 21.9, 퐺 has precisely one (푀1 , 푀2 )-alternating cycle, say 푄. Every vertex of 퐺 − 푎 1 − 푏 1 is incident with two edges in 푀1 ∪ 푀2 , hence it is a vertex of 푄. This conclusion holds for each vertex of 퐺 − 푎 1 − 푎 2 . It thus follows that 푄 is a Hamilton cycle of 퐺 − 푎 1 − 푏 1 . By Lemma 21.10, every chord of 푄 is monochromatic. However, as 퐺 is bipartite, no edge joins two vertices belonging to the same colour class. Hence, 푄 has no chords. ƒ The brace 퐺 has order six or more, hence 푄 has length four or more. Every vertex of 푄 has degree three and has only two neighbours in 푉 (푄). Thus, every vertex of 푄 is adjacent to (precisely) one of the ends of 푒. See Figure 21.6. As 퐺 is cubic and bipartite, it follows that the cycle 푄 has precisely four vertices. The two vertices of one part of 푄 are adjacent to 푎 1 , the two vertices of the other part are adjacent to 푏 1 . Indeed, as asserted, graph 퐺 is 퐾3,3 . ƒ

21.4 Pfaffian Near-Bipartite Graphs

473

Pfaffian solid bricks Every bipartite matching covered graph is solid, and many properties of bipartite graphs are valid more generally for solid graphs. This also happens to be the case in the context of Pfaffian orientations. In a paper [16] entitled A Generalization of Little’s Theorem on Pfaffian Orientations CLM showed that every minimal nonPfaffian brick must have a separating cut, and deduced that 퐾3,3 is the only minimal non-Pfaffian solid matching covered graph.

21.4 Pfaffian Near-Bipartite Graphs ♯ The Fischer-Little Theorem [34] may be restated as follows: The Fischer-Little Theorem 21.12 A near-bipartite graph is non-Pfaffian if and only if it contains one of the graphs 퐾3,3 , Γ1 and Γ2 as an 푆-minor. Proof The graphs 퐾3,3 , Γ1 and Γ2 are 푆-minimal non-Pfaffian (Theorem 21.11, Exercise 21.2.1 and Exercise 21.2.2). Let 퐺 be a non-Pfaffian near-bipartite graph. Assuming inductively that the statement of the theorem is valid for all near-bipartite graphs with fewer than |퐸 (퐺)| edges, we shall proceed to prove that 퐺 has one of 퐾3,3 , Γ1 and Γ2 as an 푆-minor.

21.4.1 Some immediate implications of the induction hypothesis If 퐺 has an 푆-minor 퐺 ′ that is bipartite and non-Pfaffian then, by Theorem 21.11, 퐺 ′ has 퐾3,3 as an 푆-minor and, hence, so does 퐺. On the other hand, if 퐺 has a proper 푆-minor 퐺 ′′ which is near-bipartite and non-Pfaffian then, by induction, one of the three graphs 퐾3,3 , Γ1 and Γ2 is an 푆-minor of 퐺 ′′ and, hence, of 퐺. We may thus assume that (i) every proper 푆-minor of 퐺 which is near-bipartite is Pfaffian and we may also assume that (ii) every bipartite 푆-minor of 퐺 is Pfaffian. Relying on these assumptions, we shall derive certain properties of 퐺 and deduce that 퐺 is either Γ1 or Γ2 . We shall adopt the notation defined below in our proof: Notation • As 퐺 is near-bipartite, it has removable doubletons. Let 푅 := {푒, 푓 }, where 푒 := 푥1 푥2 and 푓 := 푦 1 푦 2 be a removable doubleton of 퐺. • Let 푍 := {푥1 , 푥2 , 푦 1 , 푦 2 } be the set of the ends of edges 푒 and 푓 .

474

21 Excluded-Minor Characterizations of Pfaffian Graphs

• Let 퐻 be the bipartite graph 퐺 − 푅 and let ( 퐴, 퐵) denote the bipartition of 퐻. Adjust notation so that the ends 푥1 and 푥2 of 푒 are in 퐴, and the ends 푦 1 and 푦 2 of 푓 are in 퐵. • By Theorem 21.3 and Lemma 21.13, 퐺 has an 푅-triple. Let (퐷, 푀1 , 푀2 ) be an 푅-triple. Our approach to proving that 퐺 is either Γ1 or Γ2 has many features in common with the proof of Little’s Theorem presented in the previous section. For example, analogous to Statement 21.11.2 we shall prove that every removable edge of 퐻 is in 푀1 ∪ 푀2 ; and analogous to Statement 21.11.4 we shall prove that there is a unique (푀1 , 푀2 )-alternating cycle, that it is a Hamilton cycle of 퐺 − 푍, and that it is chordless. But establishing these results in the present context is technically much more involved. We begin with two simple observations. Lemma 21.13 The graph 퐻 := 퐺 − 푅 is Pfaffian. Proof The graph 퐺 − 푅 is a bipartite 푆-minor of 퐺, and we have assumed that bipartite 푆-minors of 퐺 are Pfaffian. ƒ Lemma 21.14 The graph 퐺 is a brick. Proof As 퐺 − 푅 is bipartite 푏(퐺 − 푅) = 0. It follows from the monotonicity property (Theorem 9.9) that 푏(퐺) = 1 (that is, 퐺 is a near-brick). Suppose that 퐺 is not a brick. Then, 퐺 has nontrivial tight cuts. Let 퐶 be a nontrivial tight cut of 퐺 and let 퐺 1 and 퐺 2 denote the two 퐶-contractions of 퐺. As 퐺 is a near-brick, one of the 퐶-contractions of 퐺 is bipartite and the other is a near-brick. Adjust notation so that 퐺 1 is bipartite and 퐺 2 is a near-brick. 21.14.1 The graph 퐺 2 is near-bipartite. Proof Every perfect matching of 퐺 − 푅 is a perfect matching of 퐺. The cut 퐶 is tight in 퐺; hence the cut 퐶 − 푅 is tight in 퐺 − 푅. As 퐺 − 푅 is bipartite, it follows that both (퐶 − 푅)-contractions of 퐺 − 푅 are bipartite. As 퐺 2 is a near-brick and since 퐺 2 − 푅 is bipartite, it follows that 푅 is a removable doubleton of 퐺 2 ; hence 퐺 2 is near-bipartite. ƒ The graph 퐺 1 is a bipartite 푆-minor of 퐺. We have assumed that bipartite 푆minors of 퐺 are Pfaffian; hence 퐺 1 is Pfaffian. The graph 퐺 2 is a proper 푆-minor of 퐺, because 퐶 is nontrivial and separating. Moreover, the graph 퐺 2 is near-bipartite. We have assumed that near-bipartite proper 푆-minors of 퐺 are Pfaffian; hence 퐺 2 is Pfaffian. In sum, both 퐶-contractions of 퐺 are Pfaffian. From Theorem 20.10 we infer that 퐺 is Pfaffian, a contradiction. We deduce that 퐺 is a brick. ƒ

21.4 Pfaffian Near-Bipartite Graphs

475

21.4.2 Removable edges of 푯 Lemma 21.15 Every removable edge of 퐻 is in 푀1 ∪ 푀2 . Proof Suppose that 퐻 has a removable edge ℎ which is not in 푀1 ∪ 푀2 . The doubleton 푅 is an optimal removable class of 퐺. As 퐺 − 푅 is bipartite, the edge ℎ is 푏-invariant in 퐺 − 푅; hence ℎ is an optimal removable edge of 퐺 − 푅. By Lemma 21.7 (Exchange property: optimal classes), ℎ is an optimal removable edge of 퐺, 푅 is an optimal removable class of 퐺 − ℎ and 퐺 − ℎ is not Pfaffian. However, 퐺 is a brick, ℎ is a 푏-invariant edge of 퐺 and 푅 is a removable doubleton of 퐺 − ℎ. Thus, 퐺 − ℎ is a near-brick and 푅 is a removable doubleton of 퐺 − ℎ; hence 퐺 − ℎ − 푅 is bipartite and matching covered. We deduce that 퐺 − ℎ is near-bipartite. Moreover, 퐺 − ℎ is not Pfaffian. We have assumed that every near-bipartite proper 푆-minor of 퐺 is Pfaffian; hence 퐺 − ℎ is Pfaffian, a contradiction! ƒ

21.4.3 Degree of vertices Lemma 21.16 Let 푣 be a vertex of 퐺 and let 푑 be its degree in 퐺. The following properties hold: (i) If 푣 is incident with two edges in 푀1 ∪ 푀2 then 3 ≤ 푑 ≤ 4. (ii) If 푣 is incident in 퐺 with an edge ℎ of 푀1 ∩ 푀2 , then 푑 = 3. (In particular, all the four vertices in 푍 have degree three.) Moreover, if ℎ ∉ 푅 then ℎ is the only edge of 휕 (푣) removable in 퐻. Proof By Lemma 8.6, at most two edges of 휕퐻 (푣) are not removable in 퐻. By Lemma 21.15, every removable edge of 퐻 is in 푀1 ∪ 푀2 . If 푣 is incident with two edges in 푀1 ∪ 푀2 then 푑 ≤ 4. Alternatively, suppose that 푣 is incident in 퐺 with an edge ℎ in 푀1 ∩ 푀2 . Then, 푑 = 3. Moreover, if ℎ ∉ 푅 then 푣 is incident in 퐻 with a removable edge, which must necessarily be edge ℎ. ƒ

21.4.4 There is a unique (푴1 , 푴2 ) -alternating cycle in 푮 Lemma 21.17 The graph 퐺 has no (푀1 ∩ 푀2 )-alternating cycle. Proof Assume, to the contrary, that 퐺 has an (푀1 ∩ 푀2 )-alternating cycle 퐶. For 푖 = 1, 2, let 푀푖′ := 푀푖 △ 퐸 (퐶). It follows that, for 푖 = 1, 2, the two perfect matchings 푀푖 and 푀푖′ have exactly the same edges in the set 퐸 (퐺) − 퐸 (퐶), and the cycle 퐶 is the only (푀푖 , 푀푖′ )-alternating cycle in 퐺. Thus, if 퐶 is oddly oriented in 퐷 then sign(푀푖′ ) = sign(푀푖 ), whereas if 퐶 is evenly oriented then sign(푀푖′ ) = −sign(푀푖 ). As 푀1 and 푀2 have distinct signs in 퐷 we conclude that in both alternatives the perfect matchings 푀1′ and 푀2′ of 퐺 have distinct signs in 퐷. As 퐷 − 푅 is Pfaffian, it follows that 푀1′ and 푀2′ both contain an edge of 푅. As 푅 is a removable doubleton

476

21 Excluded-Minor Characterizations of Pfaffian Graphs

of 퐺, we infer that 푅 ⊂ 푀1′ ∩ 푀2′ . Hence (퐷, 푀1′ , 푀2′ ) is an 푅-triple of 퐺. The doubleton 푅 is also a subset of 푀1 ∩ 푀2 ; hence 푅 and 퐸 (퐶) are disjoint. By Lemma 21.16, we have two contradictory conclusions. For suppose that (푒 1 , 푒 2 , . . . , 푒 2푘 ), in that cyclic order, are the edges of the cycle 퐶. Since this cycle is assumed to be (푀1 ∩ 푀2 )-alternating, we may suppose that the edges of 퐶 in 푀1 ∩ 푀2 are 푒 1 , 푒 3 , . . . , 푒 2푘−1 . Then by Lemma 21.16 it follows that the only edges of 퐶 that are removable in 퐻 are 푒 1 , 푒 3 , . . . , 푒 2푘−1 . On the other hand, by the definition of 푀1′ and 푀2′ , it follows that 퐶 is an (푀1′ ∩ 푀2′ )-alternating cycle and the edges of 퐶 that belong to 푀1′ ∩ 푀2′ are 푒 2 , 푒 4 , . . . , 푒 2푘 . In this case, by Lemma 21.16, with (퐷, 푀1′ , 푀2′ ) playing the role of (퐷, 푀1 , 푀2 ), it follows that the only edges of 퐶 that are removable in 퐻 are 푒 2 , 푒 4 , . . . , 푒 2푘 . This conclusion is a clear contradiction. ƒ The above result, together with Lemma 21.9, implies the following corollary which plays an important role in the proof of the Fischer-Little Theorem. Corollary 21.18 The graph 퐺 has precisely one (푀1 , 푀2 )-alternating cycle. Proof By Lemma 21.17, 퐺 has no (푀1 ∩ 푀2 )-alternating cycles. This assertion holds for each 푅-triple (퐷, 푀1 , 푀2 ). By Lemma 21.9, 퐺 has precisely one (푀1 , 푀2 )ƒ alternating cycle.

The cycle 푄 Let 푄 := (푣 1 , 푣 2 , . . . , 푣 ℓ , 푣 1 ) be the only (푀1 , 푀2 )-alternating cycle of 퐺. Since both edges 푒 and 푓 of 푅 belong to 푀1 ∩ 푀2 , none of the ends of these two edges, namely the vertices in 푍, lie on this cycle. Also, as 퐺 is simple, ℓ ≥ 4. Via a series of intermediate steps we shall show that 푄 is a Hamilton cycle of the graph 퐺 − 푍, and that 푄 has no chords. The first crucial step towards showing that 푄 is a Hamilton cycle of 퐺 − 푍 is to show that 푀1 ∩ 푀2 = 푅. The following two subsections establish results concerning short cycles in 퐺 which are required for achieving this objective.

21.4.5 Quadrilaterals in 푯 Lemma 21.19 No quadrilateral of 퐻 contains two or more vertices in the set 푍 := {푥1 , 푥2 , 푦 1 , 푦 2 }. Proof Assume, to the contrary, that 퐻 has a quadrilateral 푄 4 that contains two or more vertices in the set 푍. Let us first consider the possibility that 푄 4 contains both the ends 푦 1 and 푦 2 of edge 푓 . In this case, since the degrees of all vertices in 푍 are three by Lemma 21.16, and since 퐺 has at least six vertices, it follows that the set 푉 (푄 4 ) − {푦 1 , 푦 2 } is a 2-separation of 퐺, in contradiction to the hypothesis that 퐺 is a brick. See Figure 21.7(a). Likewise, 푄 4 does not contain both the ends 푥1 and 푥2 of edge 푒. Thus, 푄 4

21.4 Pfaffian Near-Bipartite Graphs

477

contains just one end of 푒 and just one end of 푓 . Adjust notation so that 푥1 and 푦 1 are the vertices of 푄 4 in 푍. Clearly, 푥1 and 푦 1 are adjacent in 푄 4 . Let 푢 ∈ 퐴 and 푤 ∈ 퐵 be the other two vertices of 푄 4 . That is, 푄 4 = (푢, 푤, 푥1 , 푦 1 , 푢). Let 푒 ′ := 푢푦 1 and 푓 ′ := 푤푥1 . (See Figure 21.7(b).) Now we proceed to show that 푅 ′ = {푒 ′ , 푓 ′ } is also a removable doubleton of 퐺, and use it to derive a contradiction. 푦1

푒′

푦1

푓 ′

푥1 (b)

푓

푦2

푢 푄4

푓 푤 푦2 (a)

푒

푥2

Fig. 21.7 A quadrilateral of 퐻 that contains two vertices in { 푥1 , 푥2 , 푦1 , 푢푦2 }

In the bipartite graph 퐻 = 퐺 − {푒, 푓 }, the ears (푢, 푤) and 푆 := (푢, 푦 1 , 푥1 , 푤) are parallel (that is, have the same ends); hence they are both removable in 퐻. Thus, 퐻 ′ := 퐻 − 푆 is matching covered. Moreover, 퐻 ′ is the same as the graph 퐻 − 푥1 − 푦 1 obtained from 퐻 by deleting 푥1 and 푦 1 . It follows that ( 퐴 − 푥1 , 퐵 − 푦 1 ) is the bipartition of 퐻 ′ . We may then add to 퐻 ′ the ear (푥2 , 푥1 , 푦 1 , 푦 2 ), thereby obtaining the bipartite matching covered graph 퐻 ′′ . The graph 퐻 ′′ is equal to 퐺 − 푒 ′ − 푓 ′ ; hence 푅 ′ := {푒 ′ , 푓 ′ } is a removable doubleton of 퐺. The graph 퐺 − 푅 ′ is a bipartite 푆-minor of 퐺. We have assumed that bipartite 푆-minors of 퐺 are Pfaffian. Thus, 퐺 − 푅 ′ is Pfaffian. By Theorem 21.3, 퐺 has two perfect matchings, 푀1′ and 푀2′ , and an orientation 퐷 ′ , such that (퐷 ′ , 푀1′ , 푀2′ ) is an 푅 ′ -triple of 퐺. The edges of the doubleton 푅 ′ belong to both 푀1′ and 푀2′ and, thus, the cycle 푄 4 is an (푀1′ ∩ 푀2′ )alternating cycle of 퐺. This is a contradiction to Lemma 21.17 (with the 푅 ′ -triple (퐷 ′ , 푀1′ , 푀2′ ) playing the role of the 푅-triple (퐷, 푀1 , 푀2 )). Indeed, no quadrilateral of 퐻 contains two or more vertices in the set 푍. ƒ

21.4.6 Triangles in 푮 Lemma 21.20 No end of an edge of 푀1 ∩ 푀2 is adjacent to both ends of an edge of 푅. Proof Suppose that 푀1 ∩ 푀2 contains an edge ℎ such that an end of ℎ, say 푤, is adjacent to both ends of an edge of 푅. By the hypothesis and Lemma 21.16, the degree of 푤 is three. Adjust notation so that 푤 is adjacent to the ends 푥1 and 푥2 of 푒. The subgraph of 퐺 induced by the vertices of the set 푇 := {푥1 , 푥2 , 푤} is a triangle in 퐺. Let 퐶 := 휕 (푇). We consider two cases depending on whether or not ℎ = 푓 . See Figure 21.8.

21 Excluded-Minor Characterizations of Pfaffian Graphs

478 푥1

푒

퐶

푤

푥2

푇

푥1

푒

퐶

푤

ℎ = 푓

푥2

푇

ℎ ≠ 푓 푀1 ∩ 푀2 (푎)

(푏)

Fig. 21.8 The triangle 푇

Case 1 The two edges ℎ and 푓 are the same (Figure 21.8(a)). The cut 퐶 − 푓 is tight in 퐻, because 푤 has degree two in 퐻. Moreover, every perfect matching of 퐺 that contains an edge in 푅 in fact contains precisely the edge 푓 in 퐶. Thus, 퐶 is tight in 퐺. This is a contradiction, as 퐺 is a brick having more than four vertices. Case 2 The edges ℎ and 푓 are different. (Figure 21.8(b)). Let 퐺 1 := 퐺/(푇 → 푡) and let 퐺 2 := 퐺/푇. We shall prove that 퐶 is separating in 퐺, and that the 퐶-contraction 퐺 1 of 퐺 is near-bipartite but is not Pfaffian. By Theorem 4.2, in order to show that 퐶 is a separating cut of 퐺 we must show that, given any edge of 퐺, there is a perfect matching of 퐺 containing that edge which meets 퐶 in precisely one edge. Firstly, note that edge ℎ is removable in 퐻 by Lemma 21.16. Thus, the cut 퐶 − ℎ is tight in 퐻 − ℎ because 푤 has degree two in 퐻 − ℎ. So, every edge of 퐻 − ℎ is in a perfect matching that meets 퐶 in just one edge. On the other hand, every perfect matching that contains an edge of 푅 contains the edge 푒 and consequently the edge ℎ. Thus, every edge of 퐺 is in a perfect matching of 퐺 that contains only one edge in 퐶. We deduce that 퐶 is a separating cut of 퐺. The graph 퐺 1 − 푓 − ℎ is the result of a bicontraction of vertex 푤 in 퐻 − ℎ; hence 퐺 1 − 푓 − ℎ is bipartite. As 퐺 is a brick having more than four vertices, the nontrivial separating cut 퐶 is not tight, therefore 퐺 1 is not bipartite. We conclude that { 푓 , ℎ} is a removable doubleton of 퐺 1 ; hence 퐺 1 is near-bipartite. Let 푋 be a subset of 푇 such that, in the digraph 퐷 ′ = 퐷 rev 휕 ( 푋), the cut 퐶 is a directed cut in 퐷 ′ . As 퐷 ′ is similar to 퐷 which is non-Pfaffian, it follows that 퐷 ′ is also non-Pfaffian. Cycle 푄 is an (푀1 , 푀2 )-alternating cycle, and, by definition of an 푅-triple, both 푒 and 푓 belong to 푀1 ∩ 푀2 . Hence 휕 ( 푋) and 퐸 (푄) are disjoint. It follows that the orientation of the edges of the cycle 푄 are not changed. Thus, 푄 is evenly oriented in 퐷 ′ and hence 푀1 and 푀2 have distinct signs in 퐷 ′ . Moreover, ℎ is the only edge of 푀1 and of 푀2 in the cut 퐶. By Lemma 20.11, one of the 퐶-contractions of 퐺 is not Pfaffian. As 퐺 2 , being 퐾4 , is Pfaffian, we conclude that 퐺 1 , the other 퐶-contraction of 퐺, is not Pfaffian. In sum, 퐺 1 is near-bipartite, non-

21.4 Pfaffian Near-Bipartite Graphs

479

Pfaffian and a proper 푆-minor of 퐺. This is a contradiction, as we have assumed that ƒ every near-bipartite proper 푆-minor of 퐺 is Pfaffian.

21.4.7 The equality 푴1 ∩ 푴2 = 푹 = {풆, 풇 } Our goal is to show that 푄 is a Hamilton cycle of the graph 퐺 − 푍 and also that 푄 has no chords. Then, we will be able to prove that the length ℓ of the cycle 푄 is eight and deduce that 퐺 is one of Γ1 and Γ2 , which is our ultimate goal. But, before proving that ℓ = 8, there is one more hurdle to cross, namely proving the following lemma, which relies on the Zigzag Lemma 8.20, and then deriving its consequences. Lemma 21.21 Let 푃 := (푢 1 , 푢 2 , 푢 3 , 푢 4 ) be a path in 퐻 such that the vertices 푢 2 and 푢 3 are adjacent to three or more vertices of 퐻. If no edge of 푃 is removable in 퐻 then 푢 2 푢 3 ∈ 푀1 ∪ 푀2 . Proof Suppose that no edge of 푃 is removable in 퐻. By Lemma 8.20, there are partitions ( 퐴1 , 퐴2 , 퐴3 , 퐴4 ) of 퐴 and (퐵1 , 퐵2 , 퐵3 , 퐵4 ) of 퐵 such that (i) (ii) (iii) (iv) (v)

푢 1 ∈ 퐴1 , 푢 2 ∈ 퐵2 , 푢 3 ∈ 퐴3 , 푢 4 ∈ 퐵4 , for 푖 = 1, 2, 3, 4, | 퐴푖 | = |퐵푖 |, 푢 1 푢 2 is the only edge that joins a vertex in 퐴1 to a vertex of 퐵 − 퐵1 , 푢 3 푢 4 is the only edge that joins a vertex of 퐴 − 퐴4 to a vertex of 퐵4 , and 푢 2 푢 3 is the only edge that joins a vertex of 퐵1 ∪ 퐵2 to a vertex of 퐴3 ∪ 퐴4 .

See Figure 21.9. 퐵1

퐵2

퐵3

퐵4

푢2

푢1 퐴1

푢4

푢3 퐴2

퐴3

퐴4

Fig. 21.9 The path 푃 and the partitions of 퐴 and 퐵

If we remove the | 퐴1 | + 1 vertices of the set 퐵1 + 푢 2 from 퐻, the vertices of 퐴1 become isolated. Thus, the cut 휕퐻 ( 퐴1 ∪ 퐵1 ∪ {푢 2 }) is nontrivial and tight in the graph 퐻 = 퐺 − 푅. But, as 퐺 is a brick, this cut cannot be tight in 퐺, and so the edge 푒 must have an end in 퐴1 (Exercise 21.4.2). Let 푥1 be the end of 푒 in 퐴1 . Similarly, if we remove the | 퐴3 | + | 퐴4 | + 1 vertices of the set 퐵3 ∪ 퐵4 ∪ {푢 2 } from 퐻, then the vertices of 퐴3 ∪ 퐴4 become isolated. Thus, the end 푥2 of 푒 has to be in 퐴3 ∪ 퐴4 (Exercise 21.4.2).

21 Excluded-Minor Characterizations of Pfaffian Graphs

480

In sum, the end 푥1 of 푒 is in 퐴1 and the end 푥2 of 푒 is in 퐴3 ∪ 퐴4 . Likewise, 푓 has an end, say 푦 1 , in 퐵1 ∪ 퐵2 and the other end 푦 2 in 퐵4 . 21.21.1 Let 푋 := 퐴1 ∪ 퐴2 ∪ 퐵1 ∪ 퐵2 , and let 퐶 denote the cut 휕 ( 푋) in 퐺 (see Figure 21.10). If 푢 2 푢 3 ∉ 푀1 ∪ 푀2 , then (i) 푒 and 푓 are the only two edges of 푀1 ∩ 푀2 in 퐶 and (ii) all the vertices of the cycle 푄 belong to one of the two shores of the cut 퐶. 푋 퐵1

푓

퐵2

퐵3

퐵4

푢2

푥1

푢4

푢1

푦2

푢3

퐴1

퐴2

퐴3

퐴4

푒 퐶 Fig. 21.10 The set 푋 and the cut 퐶

Proof As the cut 퐶 is even, every perfect matching of 퐺 meets 퐶 in an even number of edges; in particular, for 푖 = 1, 2, |푀푖 ∩ 퐶| is even. The edges 푒 and 푓 are in 퐶. The edge 푢 2 푢 3 is the only edge of 퐶 that joins a vertex in 푋 ∩ 퐵 to a vertex in 푋 ∩ 퐴. Suppose that 푢 2 푢 3 ∉ 푀1 ∪ 푀2 . A simple counting argument then shows that no edge in 푀1 ∪ 푀2 joins a vertex in 퐴2 to a vertex in 퐵3 (Exercise 21.4.2). We conclude that 푒 and 푓 are the only edges of 푀1 ∪ 푀2 in 퐶. It follows that 푉 (푄) is part of a shore of 퐶. ƒ Without loss of generality, assume that 푉 (푄) ⊆ 푋. In this case, the edges of 푀1 ∪ 푀2 in the graph 퐺 [푋] are edges in 푀1 ∩ 푀2 . Let 푁 be the set of edges of 푀1 ∪ 푀2 that have both ends in 푋. All the edges of 푁 are in fact edges of 푀1 ∩ 푀2 . Note that 푁 is a perfect matching of the bipartite graph 퐽 := 퐺 [푋] − 푥2 − 푦 2 . We shall in fact show that 퐽 is empty, that is, we shall prove that 푋 = {푥2 , 푦 2 }. By Lemma 21.17, 퐺 is free of (푀1 ∩ 푀2 )-alternating cycles; hence 푁 is the unique perfect matching of 퐽 (Exercise 21.4.2). The characterization of bipartite graphs with a unique perfect matching presented in Exercise 3.2.2 implies the following statement: 21.21.2 There is an enumeration 푣 1 , 푣 2 , . . . , 푣 푘 of the vertices of 퐴 ∩ 푉 (퐽) and an enumeration 푤 1 , 푤 2 , . . . , 푤 푘 of the vertices of 퐵 ∩ 푉 (퐽) such that (i) 푣 푖 푤 푖 ∈ 푁, for 푖 = 1, 2, . . . , 푘 and (ii) 푣 푖 푤 푗 ∉ 퐸 (퐽) for 1 ≤ 푖 < 푗 ≤ 푘. (See Figure 21.11.) ƒ

21.4 Pfaffian Near-Bipartite Graphs

481

퐶

푢2

푤1

푦2

푤2

푤3

푤푘 ... 푁

푣1

푥2

푣2

푣3

푣푘

푋 Fig. 21.11 The enumeration of 푉 ( 퐽 )

The vertex 푣 1 , being an end of an edge in 푀1 ∩ 푀2 , has degree three in 퐺 by Lemma 21.16. The only neighbour of 푣 1 in 퐽 is the vertex 푤 1 . Thus, 푁 (푣 1 ) = {푤 1 , 푦 2 , 푢 2 }. Consequently, 푣 1 = 푢 3 . See Figure 21.12. 퐶

푢2

퐶

푦2

푤1

푥2

푣1 = 푢3

푢2

푋 (a) 푘 = 1

푦2

푤1

푤2

푥2

푣1 = 푢3

푣2

푋 (b) 푘 ≥ 2

Fig. 21.12 The adjacencies of vertices in 퐴 ∩ 푉 ( 퐽 )

Suppose that 푘 = 1. Then, the vertex 푥2 , which has degree two in 퐻, must be adjacent to 푤 1 and to 푦 2 . See Figure 21.12(a). In this case (푣 1 , 푤 1 , 푥2 , 푦 2 , 푣 1 ) is a quadrilateral in 퐺 that contains two vertices that are ends of edges 푒 and 푓 . This is a contradiction to Lemma 21.19. Alternatively, suppose that 푘 ≥ 2. Then, 푁 (푣 2 ) = {푤 2 , 푤 1 , 푦 2 }, because 푣 2 is not adjacent to any vertex 푤 푗 , 푗 > 2. Let 푌 := {푣 1 , 푦 2 , 푣 2 }. The cut 휕 (푌 ) is tight in 퐻. The edges 푣 2 푤 1 and 푣 1 푤 1 are parallel in 퐻/푌 and 푣 2 푤 1 is not an edge of 퐻/푌 . Hence edge 푣 2 푤 1 is removable in 퐻. This is in contradiction to Lemma 21.15 because 푣 2 푤 1 does not belong to 푀1 ∪ 푀2 . In both alternatives, we derived a contradiction. Thus, 퐽 is the empty graph; hence 푋 = {푥2 , 푦 2 }. This is not possible, because 푋, which is 퐴3 ∪ 퐴4 ∪ 퐵3 ∪ 퐵4 , has four or more vertices. Indeed, if no edge of the path 푃 = (푢 1 , 푢 2 , 푢 3 , 푢 4 ) is removable in 퐻 then 푢 2 푢 3 is in 푀1 ∪ 푀2 . ƒ

21 Excluded-Minor Characterizations of Pfaffian Graphs

482

Lemma 21.22 푀1 ∩ 푀2 = 푅. Proof Suppose that (푀1 ∩ 푀2 ) − 푅 contains an edge, 푢 2 푣 2 . Let 푢 1 and 푢 3 be the two neighbours of 푢 2 distinct from 푣 2 , and let 푣 1 and 푣 3 be the two neighbours of 푣 2 distinct from 푢 2 . See Figure 21.13. 푢4

푢3

푔

푣3 푢2

푣2

푢1

푣1

Fig. 21.13 The edge 푢2 푣2 of 푀1 ∩ 푀2

By Lemma 21.20, at least one of 푢 1 and 푢 3 is not an end of an edge of 푅. Adjust notation so that 푢 3 is not an end of an edge of 푅. Then, 푢 3 has degree three or more in 퐻. Let 푔 be an edge of 휕 (푢 3) − 푢 2 푢 3 . Then, 푃 := (푢 1 , 푢 2 , 푢 3 , 푢 4 ) is a path of 퐻, where 푔 = 푢 3 푢 4 . The vertices 푢 2 and 푢 3 both have degree three or more in 퐻. The edges 푢 1 푢 2 and 푢 2 푢 3 are not in 푀1 ∪ 푀2 ; hence they are not removable in 퐻. In particular, 푢 2 푢 3 is not in 푀1 ∪ 푀2 . By Lemma 21.21, the edge 푔 is removable in 퐻. This conclusion holds for each edge 푔 in 휕 (푢 3) − 푢 2 푢 3 . By Lemma 21.16, every edge in 휕 (푢 3) − 푢 2 푢 3 is in 푀1 ∪ 푀2 ; hence 푢 3 has degree three and is in 푉 (푄). In sum, 푢 3 has degree three, is in 푉 (푄) and the two edges of 푄 incident with 푢 3 are removable in 퐻. Likewise, adjust notation so that 푣 3 is not an end of an edge of 푅. The vertex 푣 3 has degree three, is in 푉 (푄) and the two edges of 푄 incident with 푣 3 are removable in 퐻. The chordal path (푢 3 , 푢 2 , 푣 2 , 푣 3 ) defines, together with the cycle 푄, two cycles, 푄 1 and 푄 2 . See Figure 21.14. 푄 푣3

푣2 푄1

푀1 푄2

푀2

푢2 푔1 푢3 Fig. 21.14 The chordal path (푢3 , 푢2 , 푣2 , 푣3 ) and the cycles 푄1 and 푄2

21.4 Pfaffian Near-Bipartite Graphs

483

One of 푄 1 and 푄 2 is 푀1 -alternating and the other is 푀2 -alternating. One of 푄 1 and 푄 2 is oddly oriented and the other is evenly oriented. Adjust notation so that 푄 1 is 푀1 -alternating and oddly oriented. Let 푔1 be the edge of 푄 1 incident with 푢 3 and let 푀1′ := 푀1 △ 퐸 (푄 1 ). Then, sign(푀1′ ) = sign(푀1 ) and 푅 ⊂ 푀1′ ∩ 푀2 ; hence (퐷, 푀1′ , 푀2 ) is an 푅-triple. The edge 푔1 , a removable edge of 퐻, is not in 푀1′ ∪ 푀2 . This is a contradiction to Lemma 21.15, with 푀1′ playing the role of 푀1 . ƒ

21.4.8 The graphs 횪1 and 횪2 In order to prove that 퐺 is one of Γ1 and Γ2 , we need yet another auxiliary result, to establish that ℓ = 8. Lemma 21.23 The following properties hold: (i) the cycle 푄 has no chords, and (ii) no edge of 퐺 joins an end of 푒 to an end of 푓 . Proof (i): For each 푅-triple (퐷, 푀1 , 푀2 ) of 퐺, 푅 = 푀1 ∩ 푀2 , by Lemma 21.22. By Lemma 21.10, every chord of 푄 is monochromatic. The graph 퐺 − 푍 is bipartite. Thus, 푄 has no chords. (ii): Assume, to the contrary, that an edge of 퐺 joins an end of 푒 to an end of 푓 . Adjust notation so that 푥1 푦 1 is an edge of 퐺. By Lemma 21.17, 푥2 and 푦 2 are not adjacent. By Lemma 21.20, the vertices 푥1 and 푦 2 are not adjacent. Likewise, 푥2 and 푦 1 are not adjacent either. We shall now obtain a contradiction. The vertex 푥1 has degree three in 퐺; hence 휕 (푥1 ) has precisely one edge that joins 푥1 to a vertex of the cycle 푄. Adjust notation so that 푥1 is adjacent to vertex 푣 2 of 푄. By Lemma 21.19, no quadrilateral contains the vertices 푥1 and 푦 1 . Thus, 푦 1 is not adjacent to the vertices 푣 1 and 푣 3 . It follows that the length ℓ of 푄 is six or more. However, at most six edges join ends of edges 푒 and 푓 to vertices of 푄. As 푄 has no chords, it follows that 푄 has length six. Moreover, 푦 1 is adjacent to 푣 5 , 푦 2 is adjacent to 푣 1 and to 푣 3 , and 푥2 is adjacent to 푣 4 and to 푣 6 . See Figure 21.15. The graph 퐻 ′ := 퐻 + 푥2 푦 2 is bipartite, matching covered and planar. By Corollary 19.16, 퐻 ′ is Pfaffian. The orientation 퐷 − 푅 of 퐻 is Pfaffian and 푥2 푦 2 is an optimal removable edge of 퐻 ′ . By Corollary 20.25, 퐷 − 푅 has a Pfaffian extension to a Pfaffian orientation of 퐻 ′ , say, 퐹 ′ . The cycle 푄 is conformal in 퐻 ′ . Moreover, 푄 is evenly oriented in 퐷; hence it is evenly oriented in 퐹 ′ . This is a contradiction to Lemma 19.5(iii). We conclude that no edge of 퐺 joins an end of 푒 to an end of 푓 . ƒ Corollary 21.24 The cycle 푄 is a chordless Hamilton cycle of the graph 퐺 − 푍. Moreover, |휕 (푍)| = 8. ƒ

21 Excluded-Minor Characterizations of Pfaffian Graphs

484

푣2 PSfrag replacements 푣1

푣3

푥1

푦2 푒 푓

푣6

푥2

푣4

푦1

푣5 Fig. 21.15 The graph 퐺, where 푥1 and 푦1 are adjacent

The finish line Finally, we are now in a position to prove that the graph 퐺 is one of the graphs Γ1 and Γ2 . Let 퐶 := 휕 (푉 (푄)). By Corollary 21.24, 퐶 = 휕 (푍) and |퐶| = 8. Thus the length ℓ of 푄 is at most eight. As 퐺 is simple, ℓ ≥ 4; hence ℓ ∈ {4, 6, 8}. Let us now examine the possibilities. Let 퐻푒 := 퐺 − 푦 1 − 푦 2 and let 퐻 푓 := 퐺 − 푥1 − 푥2 . Suppose that ℓ ∈ {4, 6}. Then, 퐻푒 is planar, and it has an embedding in the plane in which the cycle 푄 bounds the infinite face. See Figures 21.16(a) and 21.16(b). Likewise, 퐻 푓 is planar and has an embedding in the plane such that 푄 bounds the infinite face. Thus, 퐺 is planar. In that case, 퐺 is Pfaffian, a contradiction. We deduce that in fact ℓ = 8. Thus, there are two possibilities for 퐻푒 , up to isomorphism (as well as for 퐻 푓 ). See Figures 21.16(c) and 21.16(d). In the first possibility, denoted 퐻푒′ and depicted in Figure 21.16(c), the graph 퐻푒′ is planar and has an embedding in which 푄 bounds its infinite face. Thus, one of 퐻푒 and 퐻 푓 must be isomorphic to the graph 퐻푒′′ , depicted in Figure 21.16(d). If 퐻푒 is 퐻푒′ then 퐻 푓 is isomorphic to 퐻푒′′ and the graph 퐺 is Γ1 . Alternatively, both 퐻푒 and 퐻 푓 may be isomorphic to graph 퐻푒′′ , in which case the graph 퐺 is Γ2 . Indeed, 퐺 is one of the graphs Γ1 and Γ2 . ƒ

Exercises 21.4.1 Find a Pfaffian orientation of the graph 퐺 ′ := 퐺 + 푥2 푦 2 , where 퐺 is the graph depicted in Figure 21.15. Hint: extend a Pfaffian orientation of the planar bipartite matching covered graph 퐺 ′ − 푅 to a Pfaffian orientation of 퐺 ′ making the 푅-alternating quadrilateral oddly oriented. 21.4.2 Supply the missing details in the proof of Lemma 21.21. ⊲21.4.3 Prove that, as defined above, if 퐻푒 = 퐻푒′ then 퐻 푓 is isomorphic to 퐻푒′′ and 퐺 is Γ1 .

21.5 Polynomial-time Algorithms

485

푣1

푣1

푣2

푣6 푣4

푥1

푒

푣2

푥2

푄

푥1

푒

푥2

푣5

푣3

푣3 (a) ℓ = 4

푣4 (b) ℓ = 6

푣1

푣1 푣2

푣8

푣7

푥1

푒

푣3

푥2

푣2

푣8

푄

푣4

푣6

푄

푣5 (c) 퐻푒′

푣7

푥1

푒

푣3

푥2

푄

푣4

푣6 푣5 (d) 퐻푒′′

Fig. 21.16 The possibilities of graph 퐻푒

⊲21.4.4 Prove that if neither 퐻푒 nor 퐻 푓 is isomorphic to 퐻푒′ then 퐺 is Γ2 . 21.4.5 Let us refer to a near-bipartite graph 퐺 with a removable doubleton 푅 as a brace-based near-bipartite graph if 퐺 − 푅 is a brace. Characterize brace-based non-Pfaffian near-bipartite graphs.

21.5 Polynomial-time Algorithms Little’s excluded minor characterization of Pfaffian bipartite graphs provides yet another way of showing the Pfaffian Orientation Problem for bipartite graphs is in co-NP. (To show that a bipartite graph is not Pfaffian, one just has to exhibit a matching minor of that graph which is isomorphic to 퐾3,3 .) Nearly 25 years later, Robertson, Seymour and Thomas [82] and, independently, McCuaig [68], succeeded in devising a polynomial-time algorithm which accepts any bipartite graph as input and, as output, it either displays a Pfaffian orientation of that graph or a matching minor isomorphic to 퐾3,3 . This important algorithm is rather hard and a full description

486

21 Excluded-Minor Characterizations of Pfaffian Graphs

of it is beyond the scope of this book. However, we do provide in this section a sketch of the main ideas involved in it. Some years later Miranda and Lucchesi [74] showed how the polynomial-time algorithm for bipartite graphs could be adapted to obtain polynomial-time algorithms for near-bipartite graphs. Their idea is described towards the end of this section.

21.5.1 Efficient recognition of Pfaffian bipartite graphs A polynomial-time algorithm for determining whether a bipartite matching covered graph is Pfaffian was discovered by McCuaig (2004, [69]) and, independently, by Robertson, Seymour and Thomas (1999, [82]). The concept of 4-cycle-sum, defined below, plays a central role in their algorithm. 4-Cycle sums Let 퐺 1 , 퐺 2 , . . . , 퐺 푛 (푛 ≥ 2) be bipartite graphs and let 푄 be a 4-cycle such that (i) each graph 퐺 푖 has order six or more, and (ii) for any two distinct 푖 and 푗, 1 ≤ 푖 < 푗 ≤ 푛, 퐺 푖 ∩ 퐺 푗 = 푄.

For each subset Ð푛 푅 of 퐸 (푄), the (푄, 푅)-sum of the 푛 summands 퐺 푖 is the 퐺 푖 ) − 푅. If 푄 and 푅 are understood, we simply say 4-cycle graph 퐺 := ( 푖=1 sum instead of (푄, 푅)-sum.

Figure 21.17 depicts the rotunda, a 4-cycle sum of three cubes, where 푅 is the set of the four edges of the quadrilateral 푄, common to the three summands.

Fig. 21.17 The rotunda is a 4-cycle sum of three cubes - the edges of 푄 are represented by dashed lines.

21.5 Polynomial-time Algorithms

487

The importance of 4-cycle sums is highlighted by the following three results, whose proofs are left as Exercises 21.5.1, 21.5.2 and 21.5.3. Lemma 21.25 If a 4-cycle sum of 푛 ≥ 2 bipartite graphs is a brace then each summand is a brace. ƒ Lemma 21.26 If a 4-cycle sum of 푛 ≥ 3 bipartite graphs is a Pfaffian brace then ƒ each summand is a Pfaffian brace. Lemma 21.27 Every brace which is a 4-cycle sum of 푛 ≥ 2 Pfaffian braces is ƒ Pfaffian. 4-Cycle decompositions and the Heawood graph A bipartite graph 퐺 is reducible if it is a 4-cycle sum of three or more bipartite graphs. The next simple result is the basis of a polynomial-time algorithm to determine whether a bipartite graph is reducible. Proposition 21.28 A bipartite graph 퐺 [ 퐴, 퐵] is reducible if and only there exists a set 푋 of vertices of 퐺 such that | 푋 ∩ 퐴| = 2 = | 푋 ∩ 퐵| and 퐺 − 푋 has three or more components. ƒ In view of Lemma 21.25, we may apply to a given brace 퐺 the following procedure, which produces a collection C of braces, called a 4-cycle decomposition of 퐺: if 퐺 is irreducible then let C := {퐺}; otherwise determine a 4-cycle sum of 퐺 of 푛 ≥ 3 summands 퐺 푖 , 푖 = 1, 2, . . . , 푛, and recursively apply the procedure to each summand 퐺 푖 , thereby obtaining a collection C푖 , and then let 푛 C . C := ∪푖=1 푖 The Heawood graph is Pfaffian and has a property, which does not seem to have a simple proof, that plays a crucial role in the algorithm: Theorem 21.29 The Heawood graph is the only simple Pfaffian irreducible nonplanar brace. ƒ

Description of the algorithm The Heawood graph has girth six, and each summand of a 4-cycle sum has a quadrilateral. Thus, the Heawood graph is not a summand of any 4-cycle sum. From Theorem 21.29 and Lemmas 21.26 and 21.27 we then infer the following consequence, which is the basis of the polynomial-time recognition of Pfaffian bipartite graphs. Corollary 21.30 Let 퐺 be a brace whose underlying simple graph is not the Heawood graph and let C be a 4-cycle decomposition of 퐺. The brace 퐺 is Pfaffian if and only if each brace in C is planar.

488

21 Excluded-Minor Characterizations of Pfaffian Graphs

Proof by induction on |푉 |. Let C be the 4-cycle decomposition of 퐺. Consider first the case in which 퐺 is irreducible. This case is the basis of the inductive proof. In that case, C = {퐺}. By hypothesis, the underlying simple graph of 퐺 is not the Heawood graph. By Theorem 21.29, 퐺 is Pfaffian if and only if it is planar. The assertion holds in this case. We may thus assume that 퐺 is reducible. Let G be a collection of 푛 ≥ 3 braces 퐺 푖 , 푖 = 1, 2, . . . , 푛 such that 퐺 is the 4-cycle sum of the 푛 graphs 퐺 푖 . By Lemmas 21.26 and 21.27, 퐺 is Pfaffian if and only if each 퐺 푖 is Pfaffian. For 푖 = 1, 2, . . . , 푛, let C푖 be a 4-cycle decomposition of 퐺 푖 . Then, C = ∪푖 C푖 . Each 퐺 푖 has a quadrilateral. The girth of the Heawood graph is six. Thus, the underlying simple graph of 퐺 푖 is not the Heawood graph. By induction, 퐺 푖 is Pfaffian if and only if the collection C푖 consists solely of planar braces. We deduce that 퐺 is Pfaffian if and only if C consists solely of planar braces. ƒ We now describe a polynomial-time algorithm to determine whether a given bipartite graph 퐺 is Pfaffian. Verifying whether a bipartite graph is Pfaffian Algorithm 21.31 Input: a (not necessarily matching covered) bipartite graph 퐺; Output: it checks whether 퐺 is Pfaffian. A fundamental case: 퐺 is matching covered. First we apply a tight cut decomposition procedure to 퐺, thereby obtaining a collection B of braces (see Exercise 5.4.5). In view of Theorem 20.10, 퐺 is Pfaffian if and only if each brace in B is Pfaffian. These observations reduce the problem to the recognition of Pfaffian braces. So, suppose that 퐺 is a brace. We first check whether the underlying simple graph 퐻 of 퐺 is the Heawood graph H14 . If 퐻 = H14 then 퐺 is Pfaffian. If 퐻 ≠ H14 then we apply a 4-cycle decomposition procedure to 퐺, thereby obtaining a collection C of braces. We then determine, in polynomial time, whether each graph in C is planar. The brace 퐺 is Pfaffian if and only each graph in C is planar, by Corollary 21.30. The generic case: 퐺 may not be matching covered. First we determine whether 퐺 is matchable, using any polynomial-time algorithm which determines a maximum matching of a graph. If 퐺 is not matchable then 퐺 is Pfaffian. We may thus assume that 퐺 is matchable. Then, in polynomial time, we determine the set 푁 of nonmatchable edges of 퐺 (Exercise 2.1.2). Let 퐻 := 퐺 − 푁. Each component of 퐻 is matching covered. The graph 퐺 is Pfaffian if and only if each component of 퐻 is Pfaffian (Exercise 19.2.4). So, the algorithm checks whether each component of 퐻, a matching covered graph, is Pfaffian.

21.5 Polynomial-time Algorithms

489

21.5.2 Efficient recognition of Pfaffian near-bipartite graphs Miranda, in his Ph. D. thesis (2009, [72]), written under the supervision of Lucchesi, discovered a polynomial-time algorithm for the recognition of Pfaffian near-bipartite graphs. This result was also published in a paper, by Miranda and Lucchesi (2008, [74]). We give here a description of that algorithm. Theorem 20.30 plays a crucial role in the analysis of the algorithm.

Description of the algorithm Efficient recognition of Pfaffian near-bipartite graphs Algorithm 21.32 Input: a near-bipartite graph 퐺; Output: it checks whether 퐺 is Pfaffian. Let 푅 be a removable doubleton of 퐺 and let 푉 (푅) denote the ends of the two edges of 푅. We first use Algorithm 21.31 to determine whether the (bipartite) graphs 퐺 − 푅 and 퐺 −푉 (푅) are Pfaffian. (Note that 퐺 −푉 (푅) is matchable but may not be matching covered.) Both these graphs are conformal subgraphs of 퐺. If one of them is not Pfaffian then 퐺 is not Pfaffian. We may thus assume that the graphs 퐺 − 푅 and 퐺 − 푉 (푅) are Pfaffian. The set 푅 is a removable doubleton of 퐺; hence 푅 is an optimal removable class of 퐺. We then apply Algorithm 20.26 to determine a characteristic orientation of 퐺 − 푅 and extend it to a characteristic orientation 퐷 of 퐺, by choosing a conformal cycle 푄 of 퐺 that contains the edges of 푅 and orienting the edges of 푅 so that 푄 is oddly oriented in 퐷. Finally, we apply Algorithm 20.29 to 퐷 − 푉 (푅). Consider first the case in which Algorithm 20.29 outputs a conformal cycle 퐶 of 퐺 − 푉 (푅) that is evenly oriented in 퐷. As 퐺 − 푉 (푅) is a conformal subgraph of 퐺, we infer that 퐶 is a conformal cycle of 퐺; hence 퐷 is not Pfaffian. As 퐷 is a characteristic orientation of 퐺 we conclude that 퐺 is not Pfaffian. Alternatively, suppose that 퐷 −푉 (푅) may be generated by Algorithm 20.26. Then, 퐷 − 푉 (푅) is a characteristic orientation of 퐺 − 푉 (푅), in turn a Pfaffian graph; hence 퐷 −푉 (푅) is Pfaffian. The graph 퐷 − 푅 is a characteristic orientation of 퐺 − 푅, in turn a Pfaffian graph; hence 퐷 − 푅 is Pfaffian. In sum, the digraphs 퐷 − 푅 and 퐷 − 푉 (푅) are Pfaffian and the cycle 푄 contains both edges of 푅 and is oddly oriented in 퐷. By Theorem 20.30, 퐷 is Pfaffian. We conclude in this case that 퐺 is Pfaffian.

490

21 Excluded-Minor Characterizations of Pfaffian Graphs

Exercises ∗ 21.5.1 Let 퐺 be a brace which is a (푄, 푅)-sum of 푛 ≥ 2 summands 퐺 푖 , 1 ≤ 푖 ≤ 푛. Prove that each summand is a brace. Hint: clearly, 퐺 + 푅 is a brace (Theorem 5.17), hence 2-extendable (Theorem 5.18). (i) Prove that each summand 퐺 푖 is connected. Hint: apply the result proved in Exercise 5.4.6 to the graph 퐺 − 푋, where 푋 := 푉 (푄) − 푣 and 푣 is a vertex of 푄. (ii) Prove that each summand 퐺 푖 is matchable. Hint: extend a perfect matching of 푄 to a perfect matching of 퐺 + 푅. (iii) Prove that each summand 퐺 푖 is a brace. Hint: let 푒 1 and 푒 2 be any two nonadjacent edges of 퐺 푖 and let 푀 be a perfect matching of 퐺 + 푅 that contains both 푒 1 and 푒 2 ; use part (ii) to prove that the set 푀 ∩ 퐸 (퐺 푖 ) may be extended to a perfect matching of 퐺 푖 by the addition of a subset of 퐸 (푄). ∗ 21.5.2 Suppose that a Pfaffian brace 퐺 is a (푄, 푅)-sum of 푛 ≥ 3 summands 퐺 1 , 퐺 2 , . . . , 퐺 푛 . Prove that each summand 퐺 푖 is Pfaffian. Hint: prove, by induction on |푅|, that the brace 퐺 + 푅 is Pfaffian. Then, deduce that 퐺 푖 is Pfaffian, because it is a conformal subgraph of 퐺 + 푅. Suppose that, for some edge 푒 ∈ 푅, the brace 퐺 + 푒 is not Pfaffian; let 퐷 be an arbitrary extension of a Pfaffian orientation of 퐺 to an orientation of 퐺 + 푒. (i) Prove that the brace 퐺 + 푒 has a perfect matching 푀 and an 푀-alternating cycle 퐶 such that 푒 ∈ 푀 and 퐶 is evenly oriented in 퐷 but does not contain the edge 푒. (ii) Prove that the set 퐸 (퐶) − 퐸 (푄) has edges in at most two summands of 퐺 + 푒, and use a third summand to show that in fact 퐺 has a perfect matching 푁 such that 퐶 is 푁-alternating in 퐺. (That is the reason for requiring three or more summands in the 4-cycle sum!) ∗ 21.5.3 Suppose that 퐺 is a brace which is a (푄, 푅)-sum of 푛 ≥ 2 summands 퐺 1 , 퐺 2 , . . . , 퐺 푛 which are all Pfaffian braces. Prove that 퐺 is Pfaffian. Hint: prove that 퐺 + 푅 is Pfaffian. Then, 퐺, a conformal subgraph of 퐺 + 푅, is also Pfaffian. (i) Prove that all the summands 퐺 푖 have Pfaffian orientations whose restrictions to 푄 coincide and are oddly oriented. Denote by 퐷 푖 such a Pfaffian orientation of 퐺 푖 . Let 퐷 := ∪푖 퐷 푖 . (ii) Let 푀 be a perfect matching of 퐺 + 푅 that contains two edges of 푄 and let 퐶 be an 푀-alternating cycle of 퐺 + 푅. If the set 퐸 (퐶) − 퐸 (푄) has edges in only one summand then certainly 퐶 is oddly oriented. Alternatively 퐸 (퐶) − 퐸 (푄) has edges in two summands and contains both edges of 푀 ∩ 퐸 (푄). In that case, the set 푁 := 푀 △ 퐸 (푄) is a perfect matching of 퐺 + 푅 and the edges of 퐸 (퐶) − 퐸 (푄) are partitioned into two 푁-alternating paths, each of which is in precisely one summand. A simple counting argument of forward edges then shows that 퐶 is oddly oriented in 퐷.

21.6 An Infinite Family of Minimal Non-Pfaffian Graphs

491

21.5.4 Prove Proposition 21.28. ⊲21.5.5 Simulate the execution of Algorithm 21.32 having graphs Γ1 and Γ2 as input. ∗ 21.5.6 (Generalization of Algorithm 21.32 – Miranda [72]) Let 퐻 [ 퐴, 퐵] be a matching covered bipartite graph; let 퐺 be a matching covered graph obtained from 퐻 by the addition of an edge 푒 having both ends in 퐴 and the addition of 푟 ≥ 1 edges 푓1 , 푓2 , . . . , 푓푟 having both ends in 퐵. Describe a polynomialtime algorithm to determine whether 퐺 is Pfaffian. Hint: for 푖 = 1, 2, . . . , 푟, the graph 퐻푖 := 퐻 + 푒 + 푓푖 is near-bipartite, and the set 푅푖 := {푒, 푓푖 } is a removable doubleton of 퐻푖 .

21.6 An Infinite Family of Minimal Non-Pfaffian Graphs ♯ In this section we describe N T , an infinite family of minimal non-Pfaffian graphs, discovered by Norine and Thomas (2008,[79]), who conjectured that the family is complete, in the sense that every minimal non-Pfaffian graph is in that family.

21.6.1 푵푻 -minors Although in this book we are concerned mostly with matching covered graphs, some authors develop theorems without restricting themselves to matching covered graphs. For example, in the definition of a Pfaffian digraph, we do not require the digraph to be matching covered. As such, Lemma 19.5 and Proposition 20.5 hold even if the graph is not matching covered. Indeed, we do not require the graph to be either matchable or connected, for that matter. In fact, if a digraph is not matchable then it is vacuously Pfaffian, because a non-Pfaffian digraph must have two perfect matchings having distinct signs. If a matchable graph is not connected then in order for it to be Pfaffian necessarily each of its components must be Pfaffian. These observations are summarized in the following result, and its proof is left as Exercise 21.6.1. Proposition 21.33 A graph 퐺 is non-Pfaffian if and only if it is matchable and at least one of its components is non-Pfaffian. ƒ Norine and Thomas [79] define a concept of minor, which we will shall refer to as an 푁푇-minor. A graph 퐻 is an 푁푇-minor of a graph 퐺 if 퐻 is obtained from 퐺 by repeated applications of three operations: (i) removal of an edge, (ii) bicontraction of a vertex, and (iii) contractions of the edges of an odd cycle. In other words, 퐻 is an 푁푇-minor of 퐺 if there exists a sequence (퐺 1 , 퐺 2 , . . . , 퐺 푟 )

(푟 ≥ 1)

492

21 Excluded-Minor Characterizations of Pfaffian Graphs

of graphs such that, 퐺 1 = 퐺, 퐺 푟  퐻 and, for 1 ≤ 푖 ≤ 푟 − 1, 퐺 푖+1 is obtained from 퐺 푖 by the application of one of the three operations mentioned above. (The Norine and Thomas graphs are simple, free of loops and multiple edges, so in their definition of bicontraction and contraction of an odd cycle they make sure to say that resulting loops and multiple edges are removed.) It can be shown that if 퐺 is Pfaffian then every 푁푇-minor of 퐺 is Pfaffian (Exercise 21.6.2). Thus, Proposition 21.34 A graph is Pfaffian if and only if each of its 푁푇-minors is Pfaffian. ƒ

21.6.2 Non-Pfaffian 푵푻 -minimal graphs A non-Pfaffian graph 퐺 is 푁푇-minimal (with respect to being non-Pfaffian) if application to 퐺 of any of the three operations mentioned earlier yields a Pfaffian graph. The notions of 푆-minimality and 푁푇-minimality of connected non-Pfaffian graphs are equivalent (Exercise 21.6.4). The equivalence of 푆-minimality and 푁푇-minimality Theorem 21.35 (Miranda (2008, [73])) A connected graph is 푁푇-minimal non-Pfaffian if and only if it is matching ƒ covered and 푆-minimal non-Pfaffian.

21.6.3 The Norine and Thomas family N T We begin with the definition of a subcollection of N T . The graphs 퐺 (푘, ∅) For a positive integer 푘, let 퐺 1 be the wheel 푊2푘+1 whose rim is (푠1 , 푠2 , . . . , 푠2푘+1 , 푠1 ), and let 퐺 2 be the graph 퐾2푘+1,3 with bipartition (푇, 푈), where 푇 := {푡1 , 푡2 , . . . , 푡2푘+1 } and 푈 := {푢 1 , 푢 2 , 푢 3 }. The graph 퐺 (푘, ∅) is defined to be the result of the splicing of 퐺 1 at its hub with 퐺 2 at vertex 푢 3 . See Figures 21.18 and 21.19.

21.6 An Infinite Family of Minimal Non-Pfaffian Graphs

493

푠2

푠1

푡1

푠3

푡2

푡3

푢1

푢2

Fig. 21.18 The graph 퐺 (1, ∅ )

푠2

푠1

푡1

푡2

푠3

푡3

푠4

푠5

푡4

푢1

푠6

푡5

푠7

푡6

푡7

푢2

Fig. 21.19 The graph 퐺 (3, ∅ )

The graphs 퐺 (푘, 푀) In the definition of 퐺 (푘, 푀) we use an auxiliary odd cycle 퐶2푘+1 := (푟 1 , 푟 2 , . . . , 푟 2푘+1 , 푟 1 ). For every integer 푘 ≥ 2 and every matching 푀 of the odd cycle 퐶2푘+1 , the graph 퐺 (푘, 푀) is obtained from 퐺 (푘, ∅) by the following procedure: For each edge 푟 푖 푟 푖+1 ∈ 푀, remove the edges 푠푖 푠푖+1 , 푡푖 푢 2 , 푡푖+1 푢 1 and add the edges 푠푖 푡푖+1 and 푠푖+1 푡푖 . See Figures 21.20 and 21.21. We are now able to define the family N T .

21 Excluded-Minor Characterizations of Pfaffian Graphs

494

푠2

푠1

푡1

푠3

푡2

푡3

푠4

푡4

푠5

푡5

푢1

푢2

Fig. 21.20 The graph 퐺 (2, {푟1 푟2 } ) 푠2

푠1

푡1

푠3

푡2

푡3

푠4

푡4

푢1

푠5

푡5

푢2

Fig. 21.21 The graph 퐺 (2, {푟1 푟2 , 푟3 푟4 } )

The family N T Let M 푘 denote the set of (possibly empty) matchings of the odd cycle 퐶2푘+1 := (푟 1 , 푟 2 , . . . , 푟 2푘+1 , 푟 1 ). Let   N T := 퐾3,3 , P, Γ2 , 퐺 (1, ∅) ∪ 퐺 (푘, 푀) : 푘 ≥ 2, 푀 ∈ M 푘 The reader may notice that Γ2 is explicitly added to the family but Γ1 is not. The reason for this discrepancy is that Γ1 is isomorphic to 퐺 (2, {푟 1 푟 2 , 푟 3 푟 4 }), depicted in Figure 21.21. (Exercise 21.6.5). The following result states that each graph in N T is a minimal non-Pfaffian graph, with one exception. Its proof is quite full of details; we refer the interested reader to Norine and Thomas’ paper [79]. Theorem 21.36 (Norine and Thomas [79]) Every graph in the family N T is nonPfaffian and, with the exception of 퐺 (1, ∅), is minimal. ƒ Conjecture 21.37 (Norine and Thomas [79]) Every connected minimal nonPfaffian graph is isomorphic to a graph in N T .

21.7 Notes

495

Exercises 21.6.1 (i) Prove Proposition 21.33. (ii) Give an example of a Pfaffian graph which has a non-Pfaffian component. ⊲21.6.2 Prove that if 퐺 is Pfaffian then every 푁푇-minor of 퐺 is Pfaffian. ⊲21.6.3 Prove that if 퐺 is a connected, non-Pfaffian graph which is not matching covered then one of the following alternatives holds: (i) either 퐺 has an edge whose removal yields a connected, non-Pfaffian graph, (ii) or 퐺 has a vertex of degree two. ∗ 21.6.4 Prove Theorem 21.35. 21.6.5 Prove that 퐺 := 퐺 (2, {푟 1 푟 2 , 푟 3 , 푟 4 }), depicted in Figure 21.21, is isomorphic to Γ1 .

21.7 Notes Miranda [71] observed that a partial result towards the proof of Conjecture 21.37 would be to prove that the graphs 퐾3,3 , Γ1 and Γ2 are the only cubic minimal non-Pfaffian graphs. Norine and Thomas [79] also consider two conjectures, in addition to Conjecture 21.37. For the first conjecture they define three operations, in addition to edge removal, vertex bicontraction and odd cycle contraction, which they call compressions of type 1,2,3. They then conjecture that a graph is non-Pfaffian if and only if it is either Γ2 or it may be reduced to 퐾3,3 or the Petersen graph by repeated applications of the above operations. For the second conjecture, they define an operation on a graph to be valid if the application of the operation to a Pfaffian graph yields a Pfaffian graph. They then conjecture that there exists a finite collection of valid operations such that every non-Pfaffian graph may be reduced to 퐾3,3 by repeated applications of operations in that collection. Robertson, Seymour and Thomas, in their paper describing the polynomial-time algorithm to recognize Pfaffian bipartite graphs [82], proved that every nonplanar brace contains an odd subdivision of either 퐾3,3 , the Heawood graph or the rotunda as a conformal subgraph. Another proof of this result and also the proof of the polynomial-time algorithm for the recognition of Pfaffian bipartite graphs, was given by P. Whalen in his Ph.D. Thesis, written under the supervision of Robin Thomas (2014, [92]).

Unsolved Problems

1. Characterize bricks, distinct from 퐾4 , 퐶6 and the Petersen graph, in which every 푏-invariant edge is solitary (page 320). 2. Characterize planar extremal bricks (page 136). 3. Every nonsolid extremal brick distinct from the Petersen brick is the result of a splicing of 퐾4 , possibly with multiple edges, and an extremal brick (page 136). 4. There exists a positive integer 푁 such that for every 푛 ≥ 푁, every brace of order 푛 has at least (푛 − 2) 2 /4 perfect matchings (page 137). 5. There exists a positive integer 푁 such that for every 푛 ≥ 푁, every brick of order 푛 has at least 푛 − 1 perfect matchings (page 137). 6. Is the problem of recognizing whether a given brick is solid in the complexity class NP? Is it in P? (page 142). 7. Do there exist solid bricks which are 4-connected? (page 153). 8. There exists a positive integer 푁 such that every simple solid matching covered graph of order 2푛 has at most 푛2 edges, for all 푛 ≥ 푁 (page 153). 9. There exist a positive real number 푐 and an integer 푁 such that any brick 퐺 of order 푛 ≥ 푁 has at least 푐푛 removable edges (page 206). 10. Characterize wheel-like bricks as a splicing of two bricks (page 217). 11. Every planar wheel-like brick is a wheel (page 217). 12. Characterize 퐾4 -free bricks (page 268). 13. Characterize 퐶6 -free bricks (page 268). 14. Every brace of order six or more has at least two nonadjacent thin edges (page 389). 15. There exists a positive constant 푐 such that every brace of order 푛 has at least 푐푛 thin edges (CLM, page 389). 16. Are the problems of recognizing a 푘-Pfaffian graph and a Pfaffian 푘-orientation polynomial-time equivalent? (page 435). 17. Every 푆-minimal non-Pfaffian graph is isomorphic to a graph in N T (Norine and Thomas, page 494). 18. The graphs 퐾3,3 , Γ1 and Γ2 are the only cubic 푆-minimal non-Pfaffian graphs (Miranda, page 495).

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7

497

498

Unsolved Problems

Some Classical Problems 1. Every planar 푟-graph is 푟-edge-colourable (Seymour, solved for 푟 ≤ 8, page 136). 2. Every 푟-graph is (푟 + 1)-edge-colourable. (Seymour, page 136). 3. For any 2-connected cubic graph 퐺, the vector 2 = (2, 2, . . . , 2) is in its perfect matching integer cone Int Cone(퐺) (Fulkerson, page 346). 4. For any matching covered graph 퐺, denote by 푐(퐺) the minimum number of perfect matchings of 퐺 whose union is 퐸 (퐺). For every 2-connected cubic graph 퐺, 푐(퐺) ≤ 5 (Berge, page 346).

Appendix A

Solutions to Selected Exercises

Contents A.1 A.2 A.3 A.4 A.5 A.6 A.7 A.8 A.9 A.10 A.11 A.12 A.13 A.14 A.15 A.16 A.17 A.18 A.19 A.20 A.21

Chapter 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 500 Chapter 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 502 Chapter 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 506 Chapter 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 509 Chapter 5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 514 Chapter 6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 517 Chapter 7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 523 Chapter 8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 525 Chapter 9 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 529 Chapter 10 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 534 Chapter 11 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 537 Chapter 12 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 539 Chapter 13 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 541 Chapter 14 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 541 Chapter 15 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 542 Chapter 16 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 543 Chapter 17 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 545 Chapter 18 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 547 Chapter 19 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 547 Chapter 20 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 548 Chapter 21 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 548

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7

499

Solutions to Selected Exercises

500

A.1 Chapter 1

Exercise 1.2.2 (i) Let 푀 be a matching and let 푆 be any subset of 푉. Then at least 표(퐺 − 푆) − |푆| odd components of 퐺 − 푆 have one or more vertices not covered by 푀. We conclude that def(푀) ≥ 표(퐺 − 푆) − |푆|. (ii) Suppose that equality holds for some set 푆. Let 푁 be any matching of 퐺. Then, def(푁) ≥ 표(퐺 − 푆) − |푆| = def (푀), hence def(푁) ≥ def(푀). This conclusion holds for each matching 푁 of 퐺, and so 푀 is maximum.

Exercise 1.3.3 Let 푀 be a matching of 퐺 and denote by Q the set of (푀0 , 푀)-alternating paths and cycles of 퐺. For each cycle 퐶 in Q, we deduce that |푀 ∩ 퐸 (퐶)| = |푀0 ∩ 퐸 (퐶)| ≥ 1. Let 푃 be any path in Q. As 푃 is (푀0 , 푀)-alternating, |푀 ∩ 퐸 (푃)| ≤ |푀0 ∩ 퐸 (푃)| + 1. Moreover, the matching 푀0 is maximal, thus 푀0 ∩ 퐸 (푃) is nonempty. Consequently, ∀ 푄 ∈ Q, It follows that

|푀 ∩ 퐸 (푄)| ≤ |푀0 ∩ 퐸 (푄)| + 1 ≤ 2|푀0 ∩ 퐸 (푄)|.

Í |푀 | = |푀 ∩ 푀0 | + 푄∈ Q |푀 ∩ 퐸 (푄)| Í ≤ 2|푀 ∩ 푀0 | + 푄∈ Q 2|푀0 ∩ 퐸 (푄)| = 2|푀0 |.

We conclude that |푀 | ≤ 2|푀0 |. This inequality holds for every matching 푀. In particular, |푀 ∗ | ≤ 2|푀0 |. Exercise 1.3.5 Start with a cycle 퐶 of length 푟 − 1. Let 퐴 be a set of 푟 − 2 new vertices (distinct from the vertices of 퐶) and join each vertex in 퐴 with all vertices in 퐶. The resulting graph, say 퐾, has an odd number of vertices. Every vertex of 퐴 has degree 푟 − 1 and every vertex of 퐶 has degree 푟. Graph 퐾 plays the role of an odd component in this construction. Let 푆 be a set of 푟 − 2 new vertices (distinct from 퐴 ∪ 퐶) and join each vertex of 푆 with one vertex in 퐴 forming a perfect matching in 푆 ∪ 퐴. Now, take 푟 − 1 copies of 퐾 and do the same links with the vertices in 푆. Figure A.1 illustrates this construction for 푟 = 5. The resulting graph 퐺 is an (푟 − 2)-edge connected 푟-regular graph on an

Exercises of Chapter 1

501

even number of vertices. Moreover, 퐺 − 푆 has |푆| + 2 odd components. By Tutte’s Theorem, 퐺 is not matchable.

Fig. A.1 A 3-edge connected 5-regular graph which is not matchable

Exercise 1.5.6 Using Jacobi’s identity, we shall now derive Tutte’s identity. We shall do this taking A to be the matrix of indeterminates associated with an orientation of the complete  graph. That would be an identity in 푛2 variables. If 퐺 is any subgraph of 퐾푛 , by setting the variables corresponding to ‘non-edges’ to be zero, we would obtain the identity that corresponds to 퐺. If A is nonsingular, as would be the case when 퐺 = 퐾푛 , (1.14) reduces to: det(C푅 ) = (det(A)) 푟 −1 det(A푅 )

(A.1)

We have proved this identity taking 푅 to be {1, 2, ..., 푟}, but the identity is clearly valid for an 푟-subset of {1, 2, ..., 푛}. Derivation of Tutte’s Pfaffian identity requires two applications of Jacobi’s identity, once with 푟 = 2, and once with 푟 = 4. Let us first take 푅 to be {푖, 푗 }. In this case, (A.1) reduces to:   0 푐 푖 푗 = det(A) det(A푖 푗 ). (A.2) det −푐 푖 푗 0 Thus, (푐 푖 푗 ) 2 = det(A) det(A푖 푗 ). Taking square roots, and noting that the Pfaffian of a skew-symmetric matrix is one of the square roots of its determinant, we have: 푐 푖 푗 = ± 푃 푃푖 푗 .

(A.3)

Now let us take 푅 to be {푥, 푦, 푧, 푤}. In this case, (A.1) reduces to: 0 푐 푥 푦 푐 푥푧 © 푐 푦푧 ­ −푐 푥 푦 0 det ­ ­ −푐 푥푧 −푐 푦푧 0 « −푐 푥푤 −푐 푦푤 −푐 푧푤

푐 푥푤 ª 푐 푦푤 ® ® = (det(A)) 3 det(A 푥 푦푧푤 ). 푐 푧푤 ® 0 ¬

(A.4)

Solutions to Selected Exercises

502

This yields: (푐 푥 푦 푐 푧푤 − 푐 푥푧 푐 푦푤 + 푐 푥푤 푐 푦푧 ) 2 = (det(A)) 3 det(A 푥 푦푧푤 ).

(A.5)

Using (A.3), this reduces to: 푃4 (± 푃 푥 푦 푃 푧푤 ± 푃 푥푧 푃 푦푤 ± 푃 푥푤 푃 푦푧 ) 2 = (det(A)) 3 det(A 푥 푦푧푤 ).

(A.6)

Since 푃4 = (det(A)) 2 , we get: (± 푃 푥 푦 푃 푧푤 ± 푃 푥푧 푃 푦푤 ± 푃 푥푤 푃 푦푧 ) 2 = (det(A)) det(A 푥 푦푧푤 ).

(A.7)

Now, upon taking square roots, we have Tutte’s Pfaffian identity: ± 푃 푥 푦 푃 푧푤 ± 푃 푥푧 푃 푦푤 ± 푃 푥푤 푃 푦푧 = 푃 푃 푥 푦푧푤 .

A.2 Chapter 2

Exercise 2.2.5 Clearly, 퐺 1 is connected; we must therefore prove that every edge of 퐺 1 is matchable. Assume, to the contrary, that 퐺 1 has an edge, 푒 := 푣푤, that is not matchable in 퐺 1 . Then, 퐺 1 has a set 푆 of vertices such that 푒 has both ends in 푆 and 표(퐺 1 − 푆) ≥ |푆|. If the contraction vertex 푥 of 퐺 1 is not in 푆 then 표(퐺 − 푆) = 표(퐺 1 − 푆) ≥ |푆|. In that case, as 푒 has both ends in 푆, it follows that 푒 is not matchable in 퐺. This is not possible, as 퐺 is matching covered, by hypothesis. We may thus assume that 푥 ∈ 푆. Let 푇 := 푆 − 푥 and let 퐻 := 퐺 [푋] = 퐺 1 − 푥. Clearly, 표(퐻 − 푇) = 표(퐺 1 − 푆) ≥ |푆| > |푇 |. As 푆 contains both ends of 푒, the set 푇 is nonempty; hence the graph 퐻 is not critical (Exercise 1.3.9). This is not possible, as 퐻, a graph that has an odd cycle as a hamiltonian cycle, is critical.

Exercise 2.2.6 (i) Suppose that 퐷 is strongly connected. Clearly, 퐻 is connected; hence 퐺 is connected. Let us show that every edge of 퐺 is matchable. Certainly, the edges of 푀 are matchable. Let 푎푏 denote an edge of 퐸 (퐺) −푀, where 푎 ∈ 퐴 and 푏 ∈ 퐵. As 퐷 is strongly connected, there is in 퐷 a directed path, 푃, from 푏 to 푎. Thus, 푃푎푏 is a directed cycle in 퐷. The corresponding cycle in 퐺 is 푀-alternating and contains the edge 푎푏, hence 푎푏 is matchable. This conclusion holds for each edge 푎푏 in 퐸 (퐺) − 푀. We deduce that every edge of 퐺 is matchable. As 퐺 is connected, in fact 퐺 is matching covered.

Exercises of Chapter 2

503

Conversely, suppose that 퐺 is matching covered. Clearly, 퐺 is connected; hence so too are 퐻 and 퐷. To prove that 퐷 is strongly connected we must prove that every arc 푢푣 of 퐷 is in a directed cycle. If 푢푣 is in 푀 or if 푢푣 is corresponds to an edge added to 퐻 as a parallel edge of an edge in 푀 then 푢푣 is in a directed cycle of order two. Assume then that 푢푣 corresponds to an edge of 퐸 (퐺) − 푀. As 퐺 is matching covered, the edge 푢푣 is matchable, hence 푢푣 is in an 푀-alternating cycle, 푄. The corresponding cycle in 퐷 is directed. Thus, 푢푣 is in a directed cycle of 퐷. We conclude that every arc of 퐷 is in a directed cycle. As 퐷 is connected, we conclude that 퐷 is strongly connected. (ii) Select an arbitrary node, 푣, of 퐷 and apply, for instance, two breadth-first searches [25], one determining the set 푆 of nodes of 퐷 that are reachable from 푣 by a directed path, the other determining the set 푇 of nodes of 퐷 that may reach 푣 by a directed path. The digraph 퐷 is strongly connected if and only if 푆 and 푇 are both equal to 푉 (퐷). It is also possible to determine whether 퐷 is strongly connected using only one depth-first search [25].

Exercise 2.3.1 If there exists a nonempty proper subset 푆 of 퐴 such that |푁 (푆)| = |푆| then every edge (if any) joining 푁 (푆) with 퐴 − 푆 is not matchable. If there is no such an edge then 퐺 would not be connected, and consequently, not matching covered. In any case, 퐺 is not matching covered. Conversely, suppose that 퐺 is not matching covered. By Theorem 2.9, 퐴 has a nonempty proper subset 푆 such that |푁 (푆)| ≤ |푆|. As 퐺 is matchable, |푁 (푆)| = |푆|, by Hall’s Theorem.

Exercise 2.3.2 Let 푆 be a minimal nonempty subset of 퐴 such that |푁 (푆)| = |푆|. It follows from Corollary 2.10 that the subgraph of 퐺 induced by 푆 ∪ 푁 (푆) is matching covered. Exercise 2.3.3 By induction on |푉 (퐺)|. By Proposition 2.11, 퐺 has a subset 푆 of 퐴 such that the subgraph of 퐺 induced by 푆 ∪ 푁 (푆) is matching covered. If 푆 = 퐴 then 퐺 is matching covered, and (퐴) and (퐵) constitute a Dulmage-Mendelsohn decomposition of 퐺. Otherwise, let 퐺 ′ := 퐺 − (푆 ∪ 푁 (푆)). Then, 퐺 ′ is nonempty. Let [ 퐴′ , 퐵′ ] denote the parts of 퐺 ′ , where 퐴′ ⊂ 퐴 and 퐵′ ⊂ 퐵. For any perfect matching 푀 of 퐺, the restriction of 푀 to 퐺 ′ is a perfect matching of 퐺 ′ . Then 퐺 ′ [ 퐴′ , 퐵′ ] is matchable. By induction, 퐺 ′ [ 퐴′ , 퐵′ ] has a Dulmage-Mendelsohn decomposition, say, a partition ( 퐴1 , . . . , 퐴푟 ) (푟 ≥ 1) of 퐴′ and a partition (퐵1 , . . . , 퐵푟 ) of 퐵′ . If we now set 퐴0 := 푆 and 퐵0 = 푁 (푆), then partition ( 퐴0 , . . . , 퐴푟 ) of 퐴 and partition (퐵0 , . . . , 퐵푟 ) of 퐵 constitute a Dulmage-Mendelsohn decomposition of 퐺.

504

Solutions to Selected Exercises

Exercise 2.4.1 Let 퐻 be the result of the splicing of two matching covered graphs, 퐺 1 and 퐺 2 , at vertices 푣 1 and 푣 2 , respectively. Let 퐶 denote the cut of 퐻 that coincides with 휕 (푣 푖 ) in graph 퐺 푖 , for 푖 = 1, 2. If 퐺 푖 has only two vertices then 퐺 푖 − 푣 푖 is connected. Alternatively, if 퐺 푖 has order four or more then, by Corollary 2.7, 퐺 푖 is 2-connected; hence 퐺 푖 − 푣 푖 is connected. In both alternatives, 퐺 푖 − 푣 푖 is connected; hence 퐻 is connected. Let 푒 be any edge of 퐻. Adjust notation so that 푒 is an edge of 퐺 1 . By hypothesis, 퐺 1 has a perfect matching, 푀1 , that contains edge 푒. Let 푓 denote the edge of 푀1 in 퐶. Let 푀2 be a perfect matching of 퐺 2 that contains edge 푓 . Then, 푀1 ∪ 푀2 is a perfect matching of 퐺 that contains edge 푒. Thus, edge 푒 is matchable in 퐻. That conclusion holds for each edge 푒 of 퐻. As 퐻 is connected, it follows that 퐻 is matching covered.

Exercise 2.4.2 Let 푣 푖 , for 1 ≤ 푖 ≤ 5, denote the vertices of the rim of the 5-wheel 퐺, such that 푣 1 푣 2 푣 3 푣 4 푣 5 occur in that cyclic order in the rim. Order the vertices 푤 푖 of the rim of 퐻, for 1 ≤ 푖 ≤ 5 analogously. The prism 푃10 may be obtained by defining 휃 (푣 푖 ) := 푤 푖 , for 푖 = 1, 2, . . . , 5. The Petersen graph may be obtained by defining 휙(푣 푖 ) := 푤 2푖−1 computed modulo 5. Now consider any splicing 퐺 ⊙ 퐻 that produces a graph distinct from the Petersen graph. Let 퐶 be the set of edges that join vertices of the rim of 퐺 to vertices of the rim of 퐻. We show that the hypothesis that 퐺 ⊙ 퐻 is not the Petersen graph implies that 퐶 contains two edges that join two adjacent vertices of the rim of 퐺 to two adjacent vertices of the rim of 퐻. To see this, assume the contrary. Then, given two adjacent vertices of the rim of 퐺, their neighbours in the rim of 퐻 must be at distance two. Adjust notation so that 푣 1 and 푤 1 are adjacent. Then, 푣 2 is adjacent to one of 푤 3 and 푤 4 . Without loss of generality, we may assume that 푣 2 is adjacent to 푤 3 . It follows that 푣 3 is adjacent to 푤 5 , 푣 4 is adjacent to 푤 2 and 푣 5 is adjacent to 푤 4 . We deduce that 퐺 ⊙ 퐻 is the Petersen graph, a contradiction. Without loss of generality, we may assume that an edge 푒 푖 of 퐶 joins 푣 푖 to 푤 푖 , for 푖 = 1, 2. Let 푀푖 be the perfect matching of 퐺 ⊙ 퐻 such that 푀푖 ∩ 퐶 = {푒 푖 }. Those two perfect matchings are disjoint. As 퐺 ⊙ 퐻 is cubic, it follows that it is 3-edge-colourable. Assume, to the contrary, that the Petersen graph is 3-edge-colourable. As P is cubic, it follows that 퐸 (P) has a partition in three perfect matchings. Every perfect matching of P must contain an odd number of edges in the cut 퐶. Thus, one of the three disjoint perfect matchings of P, say 푀, must contain precisely three edges in 퐶. The ends of the edges of 푀 ∩ 퐶 in the rim of 퐺 must be three cyclically consecutive vertices of the rim of 퐺 as a fourth edge of 푀 joins the remaining two vertices of the rim of 퐺. Likewise, the ends of the three edges of 푀 ∩ 퐶 must be three consecutive

Exercises of Chapter 2

505

vertices of the rim of 퐻. The Petersen graph does not contain three edges in 퐶 with these properties.

Exercise 2.4.6 We now refer to the graphs depicted in Figure A.2. Clearly, G4 = {퐾4 } = {퐺 1 } and 퐺 1 is vertex-transitive. Thus, G6 consists of just one graph, 퐶6 = 퐺 2 . Again, 퐺 2 is vertex-transitive and G8 = {퐺 3 }. The graph 퐺 3 is not vertex-transitive. The orbit of every vertex contains one of the three black vertices. Thus, G10 consists of the three graphs 퐺 4 , 퐺 5 and 퐺 6 . 퐺1

퐺4

퐺2

퐺3

퐺5

퐺6

Fig. A.2 The six cubic graphs of the families G4 , G6 , G8 and G10

Exercise 2.5.4 The prism P2푛 is obtained from the ladder L2푛 by adding the edges 푢 1 푢 푛 and 푣 1 푣 푛 . Let us first compute Φ(P2푛 ) for 푛 ≥ 3 and odd. In this case, a perfect matching of P2푛 either contains none of 푢 1 푢 푛 and 푣 1 푣 푛 or both of them. Using (2.4), we have Φ(P2푛 ) = Φ(L2푛 ) + Φ(L2(푛−2) ) = 퐹 (푛 + 1) + 퐹 (푛 − 1). When 푛 ≥ 4 is even, a perfect matching of P2푛 may contain any subset of {푢 1 푢 푛 , 푣 1 푣 푛 }. Moreover, there are only two perfect matchings that contain precisely one of 푢 1 푢 푛 and 푣 1 푣 푛 . Using (2.4), we have Φ(P2푛 ) = Φ(L2푛 ) + Φ(L2(푛−2) ) + 2 = 퐹 (푛 + 1) + 퐹 (푛 − 1) + 2. The formula for Φ(M2푛 ) can be deduced analogously.

Solutions to Selected Exercises

506

Using the notation as in the definition of staircase, the formula for S2푛 can be deduced from the fact that there are 퐹 (푛) perfect matchings containing the edge 푥푦, and one perfect matching containing each one of the edges 푥푢 1 and 푥푣 1 .

A.3 Chapter 3

Exercise 3.1.3 Figure A.3 depicts graph 퐺 := 퐾3,3 ⊙ 퐾3,3 . By Corollary 3.5, every barrier of 퐺 is 퐴

푌

푋

퐵

Fig. A.3 Graph 퐾3,3 ⊙ 퐾3,3

a subset of one of 퐴 and 퐵. Every vertex of 퐺 is a barrier. No two vertices of 퐴 constitute a barrier. No four vertices of 퐴 constitute a barrier. The set 푋 is the only subset of 퐴 with just three vertices which is a barrier. Thus, the nontrivial barriers of 퐺 that are subsets of 퐴 are 푋 and 퐴. Likewise, the nontrivial barriers of 퐺 that are subsets of 퐵 are 푌 and 퐵.

Exercise 3.1.4 Let 퐵′ be a proper superset of 퐵, let us prove that 퐵′ is not a barrier of 퐺. Each component of 퐺 − 퐵 is critical, hence odd. Let K be the set of components of 퐺 − 퐵 that contain vertices of 퐵′ . For each 퐾 ∈ K, let 푆 퐾 := 퐵′ ∩ 푉 (퐾). As stated in Exercise 1.3.9, 표(퐾 − 푆 퐾 ) ≤ |푆 퐾 | − 1 for each 퐾 ∈ K. Thus, Í 표(퐺 − 퐵′ ) = 표(퐺 − 퐵) − |K | + 퐾 ∈ K 표(퐾 − 푆 퐾 ) Í ≤ |퐵| − |K | + 퐾 ∈ K [|푆 퐾 | − 1] = |퐵′ | − 2|K | < |퐵′ |. Thus, 퐵′ is not a barrier of 퐺. This conclusion holds for each proper superset 퐵′ of 퐵. We deduce that 퐵 is a maximal barrier of 퐺.

Exercises of Chapter 3

507

Exercise 3.1.7 If 퐺 is matching covered then 퐺 − 푎 − 푏 is matchable, for each 푎 ∈ 퐴 and each 푏 ∈ 퐵, by Corollary 3.6. Conversely, suppose that 퐺 − 푎 − 푏 is matchable, for each 푎 ∈ 퐴 and each 푏 ∈ 퐵. By hypothesis, 퐺 has an edge, say, 푢푣, 푢 ∈ 퐴, 푏 ∈ 퐵. The graph 퐺 − 푢 − 푣 is matchable, hence 퐺 is also matchable. Suppose that 퐺 is not matching covered. By Corollary 2.10, 퐴 has a nonempty proper subset 푆 such that |푁 (푆)| = |푆|. Then, 퐴 − 푆 is nonempty, let 푎 ∈ 퐴 − 푆. Also, 푁 (푆) is nonempty, let 푏 ∈ 푁 (푆). By Hall’s Theorem, the graph 퐺 − 푎 − 푏 is not matchable, a contradiction. We conclude that 퐺 is matching covered.

Exercise 3.2.2 (i) Let 푃 be a maximal 푀-alternating path in 퐻 and let 푢 denote the origin of 푃. Suppose that the edge 푢푣 of 푀 incident with 푢 is not in 푃. By the maximality of 푃, the vertex 푣 is in 푃. Thus, the edges of the subpath of 푃 extending from 푢 to 푣, plus the edge 푢푣, induce an 푀-alternating cycle of 퐻, in contradiction to the uniqueness of 푀. We deduce that the edge 푢푣 is in 푃. As 푃 is a path, it follows that 푢푣 is the first edge of 푃. Thus, 푃 starts with an edge in 푀. Suppose that 푢 has degree greater than one, let 푢푤 be an edge of 퐻 incident with 푢 but distinct from 푢푣. The maximality of 푃 implies that 푤 is in 푃. As 퐻 is bipartite, the edges of the subpath of 푃 extending from 푢 to 푤, plus edge 푢푤, induce an 푀-alternating cycle of 퐻, again contradicting the uniqueness of 푀. We deduce that the origin of 푃 has degree one and 푃 starts with an edge in 푀. Likewise, the end of 푃 has degree one and 푃 ends with an edge in 푀. The length of 푃 is odd, therefore it starts and ends in distinct parts of ( 퐴, 퐵). (ii) Let 퐵 denote a maximal barrier of 퐺. Consider the core H(퐵) of 퐺 associated with 퐵. The set 퐴 of contracted vertices of H constitutes, with 퐵, a bipartition of H(퐵). By Theorem 3.2, every component of 퐺 − 퐵 is odd and critical. It follows that every perfect matching of H(퐵) is extendable to a perfect matching of 퐺. By hypothesis, 퐺 is 2-edge-connected. Thus, every vertex has degree two or more in H(퐵). It follows, from the first part, that H(퐵) has at least two perfect matchings. Thus, 퐺 has at least two perfect matchings. (iii) By hypothesis, 퐺 has a (unique) perfect matching. Thus, 퐺 is matchable. Let us now consider a Dulmage-Mendelsohn decomposition of 퐺. Let ( 퐴1 , 퐴2 , . . . , 퐴푟 ) be a partition of 퐴 and let (퐵1 , 퐵2 , . . . , 퐵푟 ) be a partition of 퐵 such that (i) for 푖 = 1, 2, . . . , 푟, the graph 퐻푖 := 퐺 [ 퐴푖 ∪ 퐵푖 ] is matching covered and (ii) for every edge 푎 푖 푏 푗 of 퐺, with 푎 푖 ∈ 퐴푖 and 푏 푗 ∈ 퐵 푗 , 푖 ≥ 푗. For 푖 = 1, 2, . . . , 푟 the graph 퐻푖 has a perfect matching, 푀푖 . For any perfect matching 푁푖 of 퐻푖 , the set (푀 − 푀푖 ) ∪ 푁푖 is a perfect matching of 퐺. By hypothesis, 퐺 has a unique perfect matching. It follows that 퐻푖 has a unique perfect matching. As 퐻푖 is matching covered, 퐻푖 = 퐾2 . The assertion holds.

Solutions to Selected Exercises

508

Exercise 3.3.1 By hypothesis, 퐵 is a maximal barrier, therefore, by Theorem 3.2, every component of 퐺 − 퐵 is odd and critical. Thus, 퐾 − 푎 has a perfect matching, 푀퐾 . Likewise, for every component 퐿 of 퐺 − 퐵 distinct from 퐾, 푁 has precisely one edge, 푒 퐿 , incident 푀 퐿 , where 푣 퐿 denotes the end of 푒 퐿 with 푉 (퐿), and 퐿 Ð − 푣퐿Ðhas a perfect matching,  is matching of 퐺 − 푎 − 푏. 푀 a perfect in 푉 (퐿). Thus, 푁 퐽 퐽 ∈ O (퐺−퐵) Exercise 3.3.5 Let 퐺 be a nonbipartite matching covered graph. Let 퐵 be a nontrivial maximal barrier of 퐺. As 퐺 is nonbipartite, 퐺 − 퐵 has a nontrivial component, 퐾. By Theorem 3.2, 퐾 is critical. By Exercise 1.3.8, 퐾 has an odd cycle 퐶. Recall that every edge of a matching covered graph joins vertices in distinct parts of the canonical partition. It follows that the vertices of 퐶 lie in at least three parts of the canonical partition. Thus, the canonical partition of 퐺 has at least four parts. Let 푟 ≥ 4 be an integer. Let us now give an example of a matching covered graph 퐺 whose canonical partition has precisely 푟 parts. If 푟 is even then let 퐺 := 퐾푟 . If 푟 is odd, let 퐺 be the graph obtained from 퐾푟 − 푒 by the addition of a vertex of degree two adjacent to both ends of 푒. If 푟 is even then 퐺 is bicritical and has precisely 푟 vertices. If 푟 is odd then the two ends of 푒 constitute a barrier, all the other barriers are singletons. In both cases, the canonical partition of 퐺 consists of precisely 푟 barriers.

Exercise 3.3.10 For 푖 = 1, 2, let 퐵푖 denote the barrier in the canonical decomposition of 퐺 that contains vertex 푢 푖 . The following three statements are equivalent: 1. The graph 퐺 + 푒 푖 is not matching covered; 2. the graph 퐺 − 푢 푖 − 푣 푖 is not matchable; 3. vertex 푣 푖 also lies in 퐵푖 . Thus, the necessary and sufficient conditions are: (i) 푣 푖 lies in 퐵푖 , for 푖 = 1, 2 and (ii) 퐺 − 푢 1 − 푣 1 − 푢 2 − 푣 2 is matchable. Exercise 3.4.4 Consider first the case in which an ear decomposition of 퐻 may be extended to an ear decomposition of 퐺. Let 퐺 1 = 퐾2 , 퐺 2 , . . . , 퐺 푞 = 퐻, . . . , 퐺 푟 = 퐺, be an ear decomposition of 퐺 which is an extension of an ear decomposition of 퐻, where 1 ≤ 푞 ≤ 푟. By definition of ear decomposition, each 퐺 푖 is matching

Exercises of Chapter 4

509

covered, hence 퐻 is matchable. The set of even numbered edges of the ears 푃푖 , 푖 = 푞 + 1, 푞 + 2, . . . , 푟, is a perfect matching of 퐺 − 푉 (퐻). Thus, 퐻 is a conformal subgraph of 퐺. To prove the converse, assume that 퐻 is a conformal subgraph of 퐺. We prove that an ear decomposition of 퐻 may be extended to an ear decomposition of 퐺 by induction on |퐸 (퐺)| − |퐸 (퐻)|. If 퐻 = 퐺 then the assertion holds trivially. We may thus assume that 퐻 is a proper subgraph of 퐺. As 퐺, a matching covered graph, is connected, 퐺 has an edge 푓 that is incident with a vertex in 푉 (퐻) but does not lie in 퐻. Graph 퐻 is a conformal subgraph of 퐺, let 푀 be a perfect matching of 퐺 such that 푀 ∩ 퐸 (퐻) is a perfect matching of 퐻. Let 푀 푓 be a perfect matching of 퐺 that contains edge 푓 . As in the proof of the Bipartite Ear Decomposition Theorem (3.14), consider an (푀, 푀 푓 )-alternating cycle that contains edge 푓 and let 푔 denote the edge with which this cycle reenters 푉 (퐻) for the first time. Edge 푔 also lies in 푀 푓 . Thus, the segment 푃 of the cycle starting with 푓 and ending with 푔 has odd length and is internally disjoint with 퐻. Moreover, 퐻 ′ := 퐻 + 푃 is a conformal subgraph of 퐺, as 푀 − 퐸 (퐻 ′ ) is a perfect matching of 퐺 −푉 (퐻 ′ ). Moreover, the ends of 푃 lie in distinct parts of the bipartition of 퐻, thus 퐻 ′ is matching covered and bipartite. The assertion holds, by induction.

A.4 Chapter 4

Exercise 4.2.2 Let 푀 be any perfect matching of 퐺. As {푢, 푣} is a 2-separation of 퐺, every edge of 퐶 is incident with a vertex in {푢, 푣}. Thus, |푀 ∩ 퐶| ≤ 2. By definition, {푢, 푣} is not a barrier of 퐺. Thus, every connected component of 퐺 − 푢 − 푣 is even. This implies that cut 퐶 is odd. Thus, |푀 ∩ 퐶| is odd. It follows that 푀 contains precisely one edge in 퐶. This conclusion holds for each perfect matching 푀 of 퐺. We deduce that 퐶 is tight. Likewise, cut 퐷 is also tight.

Exercise 4.2.4 (i) Let 푒 be any edge of 퐺. We shall prove that 퐺 has a perfect matching that contains 푒 and just one edge in each of 퐶 and 퐷. Consider first the case in which 푒 is an edge of 퐺/푋. Let 푀1 be a perfect matching of 퐺/푋 that contains edge 푒 and just one edge in 퐷. Let 푓 be the edge of 푀1 in 퐶. By definition, 퐺/ 푋 is matching covered; let 푀2 be a perfect matching of 퐺/푋 that contains edge 푓 . Then, 푀1 ∪ 푀2 is a perfect matching of 퐺 that contains edge 푒 and just one edge in each of 퐶 and 퐷. Consider next the case in which 푒 lies in 퐺/푋. Let 푁1 be a perfect matching of 퐺/푋 that contains edge 푒. Let 푔 be the edge of 푁1 in 퐶. Let 푁2 be a perfect matching of 퐺/푋 that contains edge 푔 and just

Solutions to Selected Exercises

510

one edge in 퐷. Then 푁1 ∪ 푁2 is a perfect matching of 퐺 that contains edge 푒 and precisely one edge in each of 퐶 and 퐷. (ii) Let 푀 be any perfect matching of 퐺. By hypothesis, 퐶 is tight in 퐺; therefore 푀 contains precisely one edge in 퐶. It follows that the restriction of 푀 to 퐺/푋 is a perfect matching of 퐺/푋. By hypothesis, 퐷 is tight in 퐺/푋. Thus, 푀 contains precisely one edge in 퐷. This conclusion holds for each perfect matching 푀 of 퐺. (iii) A trivial example is obtained by taking 퐷 := 퐶 and 퐶 not tight in 퐺. A less trivial example is shown in Figure A.4. Cut 퐷 is a 2-separation cut in 퐺/푋, but the perfect matching of 퐺 that contains all five edges in 퐶 contains three edges in 퐷. 푋

퐷

퐶 Fig. A.4 Example for the solution to Exercise 4.2.4 4.2.4

(iv) Let 푃 be a pentagon of P, the Petersen graph. Consider another pentagon, 푄, which is neither 푃 nor P − 푉 (푃). The cuts 퐶 := 휕 (푃) and 퐷 := 휕 (푄) are both separating. Moreover, 퐶 and 퐷 have just one edge in common, and each perfect matching that contains that common edge has one edge in one of 퐶 and 퐷 and has five edges in the other.

Exercise 4.3.8 Suppose first that 퐶 and 퐷 do not cross and, without loss of generality, that 푋 ∩푌 = ∅. In this case 푋 ⊂ 푌 . Thus 퐼 = 휕 ( 푋) = 퐶 and 푈 = 휕 (푌 ) = 퐷, and hence the assertion holds. So, we may assume that 퐶 and 퐷 do cross. Let 푀 be any perfect matching of 퐺. Denote by 퐹 the set of edges of 퐺 that join a vertex in 푋 ∩ 푌 and a vertex in 푋 ∩ 푌 . Clearly, |푀 ∩ 퐶| + |푀 ∩ 퐷| = |푀 ∩ 퐼 | + |푀 ∩ 푈| + 2|푀 ∩ 퐹|. (A.8)

Exercises of Chapter 4

511

Let 푒 be any edge of 퐺. By hypothesis, {퐶, 퐷} is cohesive; thus 퐺 has a perfect matching, 푀푒 , that contains edge 푒 and just one edge in each of 퐶 and 퐷. The cuts 퐼 and 푈 are both odd; hence 푀푒 contains at least one edge in each cut in {퐼, 푈}. From (A.8) we have that 2 = |푀푒 ∩ 퐶| + |푀푒 ∩ 퐷| = |푀푒 ∩ 퐼 | + |푀푒 ∩푈| + 2|푀푒 ∩ 퐹| ≥ 2 + 2|푀푒 ∩ 퐹|. (A.9) We conclude that for each edge 푒 of 퐺 there is a perfect matching 푀푒 of 퐺 such that (a) 푒 ∈ 푀푒 , |푀푒 ∩ 퐹| = 0, (b) |푀푒 ∩ 퐶| = 1, |푀푒 ∩ 퐷| = 1, |푀푒 ∩ 퐼 | = 1 and |푀푒 ∩ 푈| = 1. From (a) we deduce that 푒 ∉ 퐹. This conclusion holds for each edge 푒 of 퐺. Thus, 퐹 is empty. That is, no edge of 퐺 joins a vertex in 푋 ∩ 푌 to a vertex in 푋 ∩ 푌 . As 퐹 is empty, from (A.8) we deduce the validity of equality (4.3). For each edge 푒, 퐺 has a perfect matching 푀푒 that satisfies (b). We deduce that {퐶, 퐷, 퐼, 푈} is cohesive. In particular, the cuts 퐼 and 푈 are both separating. Exercise 4.3.9 The cut 퐶 is separating and the cut 퐷 is tight. Thus, the pair {퐶, 퐷} is cohesive (Exercise 4.2.3). By Theorem 4.14, no edge of 퐺 has an end in 푋 ∩ 푌 and an end in 푋 ∩ 푌 and the cuts 퐼 and 푈 are separating. Let 푀 be a perfect matching of 퐺. By Theorem 4.14, equality (4.3) holds. The cut 퐷 is tight; hence |푀 ∩ 퐶| + 1 = |푀 ∩ 퐶| + |푀 ∩ 퐷| = |푀 ∩ 퐼 | + |푀 ∩ 푈| ≥ 2. If 퐶 is tight then |푀 ∩ 퐼 | + |푀 ∩ 푈| = 2; hence |푀 ∩ 퐼 | = 1 = |푀 ∩ 푈|; as this conclusion holds for each perfect matching 푀 of 퐺, we deduce that 퐼 and 푈 are both tight. Alternatively, if 퐶 is not tight then, for some perfect matching 푀 of 퐺, we have that |푀 ∩ 퐶| > 1. In that case, |푀 ∩ 퐼 | + |푀 ∩ 푈| > 2; hence at least one of the cuts 퐼 and 푈 is not tight.

Exercise 4.3.11 If 퐶 is tight then then 퐼 and 푈 are also tight. In that case, 휆(퐶) = ∞ = 휆(퐼) = 휆(푈). Thus, 휆(퐶) = min{휆(퐼), 휆(푈)}. We may thus assume that 퐶 is separating but not tight. Let us first prove that 휆(퐶) ≥ min{휆(퐼), 휆(푈)}. Let 푀 be a perfect matching of 퐺 that contains precisely 휆(퐶) edges in 퐶. As 퐷 is tight, we infer from (4.3) that 휆(퐶) + 1 = |푀 ∩ 퐼 | + |푀 ∩ 푈|. As 휆(퐶) > 1, at least one of 퐼 and 푈 has more than one edge in 푀. Without loss of generality, suppose that |푀 ∩ 퐼 | > 1. Then, |푀 ∩ 퐼 | ≥ 휆(퐼).

512

Solutions to Selected Exercises

As the cuts 퐼 and 푈 are both odd, we have that both 퐼 and 푈 have at least one edge in 푀; hence neither 퐼 nor 푈 has more than 휆(퐶) edges in 푀. In particular, 휆(퐶) ≥ |푀 ∩ 퐼 | ≥ 휆(퐼). We conclude that 휆(퐶) ≥ min{휆(퐼), 휆(푈)}. Let us now prove that 휆(퐶) ≤ min{휆(퐼), 휆(푈)}. As 퐶 is not tight, at least one of 퐼 and 푈 is not tight; hence min{휆(퐼), 휆(푈)} < ∞. Without loss of generality, suppose that 휆(퐼) = min{휆(퐼), 휆(푈)} < ∞. Let 푀퐼 be a perfect matching of 퐺 that contains 휆(퐼) edges in 퐼. Cut 퐷 is tight; therefore 푀퐼 has just one edge in 퐷, say 푓 . As cut 푈 is separating, 퐺 has a perfect matching, 푀푈 , that contains 푓 and just one edge in 푈. Then,     푀퐼 ∩ 퐸 (퐺 [푌 ]) ∪ 푀푈 ∩ 퐸 (퐺 [푌 ]) ∪ { 푓 } is a perfect matching of 퐺 that contains precisely 휆(퐼) edges in 퐶. It follows that 휆(퐶) ≤ 휆(퐼) = min{휆(퐼), 휆(푈)}. Exercise 4.5.2 Let us define a family G := {퐺 푘 : 푘 ≥ 2} and prove that each 퐺 푘 is a brick which has a separating cut whose 퐶-contractions are 퐾4 and a graph 퐻 푘 such that 푏(퐻 푘 ) = 푘 − 1. The graph 퐺 푘 has precisely 2푘 + 2 vertices: 푉 (퐺 푘 ) := {푣 푖 , 푤 푖 : 0 ≤ 푖 ≤ 푘}. Now let us add 4푘 + 1 edges in order to obtain 퐺 푘 (Figure A.5(a) depicts graph 퐺 5 ): 1. Form 푘 − 1 triangles by the addition of edges 푣 푖 푤 푖 , 푤 푖 푣 푖+1 and 푣 푖+1 푣 푖 , for 푖 = 1, 2, . . . , 푘 − 1. 2. Form another triangle, by the addition of edges 푣 0 푤 0 , 푤 0 푤 푘 and 푤 푘 푣 0 . 3. Add edges 푣 0 푤 푖 for 푖 = 1, 2, . . . , 푘 − 1. 4. Add edges 푣 1 푤 0 and 푣 푘 푤 푘 . We remark that 퐺 2 = 퐶6 . Let 푋 := {푣 0 , 푤 0 , 푤 푘 } and let 퐶 := 휕 ( 푋). The 퐶contraction 퐺 ′ := 퐺 푘 /푋 is 퐾4 , up to multiple edges. Let 퐺 ′′ := 퐺 푘 /( 푋 → 푥) be the other 퐶-contraction of 퐺 푘 . The set of 2-separations of 퐺 ′′ is {푥, 푣 푖 }, 2 ≤ 푖 ≤ 푘 − 1. The tight cut decomposition of 퐺 ′′ produces only 퐾4 ’s, up to multiple edges, and 푏(퐺 ′′ ) = 푘 − 1. Thus, 푏(퐺 ′ ) + 푏(퐺 ′′ ) = 푘. Moreover, 퐺 푘 is the splicing of 퐺 ′ and 퐺 ′′ . Thus, 퐺 푘 is matching covered. Let us now prove that 퐺 푘 is a brick, by induction on 푘. The basis of the induction is for 푘 = 2, in which case 퐺 2 = 퐶6 is a brick. Let us now assume that 푘 > 2. Let 푌 := {푣 푘−1 , 푣 푘 , 푤 푘−1 }. The set 푌 induces a triangle in 퐺 푘 . Let 퐷 := 휕 (푌 ) and let 퐻 := 퐺 푘 /(푌 → 푦) (Figure A.5(b)). It is easy to see that 퐻 is isomorphic to 퐺 푘−1 + 푒, where 푒 = 푣 0 푤 푘−2 . Let us now prove that 퐻 is a brick. By induction, 퐻 − 푒 is a brick. Let 퐹 be any nontrivial cut of 퐻. As 퐻 − 푒 is a brick, the graph 퐻 − 푒 has a perfect matching, 푀, that contains more than one edge in 퐹 − 푒. Thus, 푀 is a perfect matching of 퐻 that contains more than one edge in 퐹. We deduce that 퐹 is not tight. This conclusion

Exercises of Chapter 4

513 푣3 푌 푣2

푣4 푤2

푤3

푤4

푤1 푣1

푣5 퐷

푣0

푋

푤5

푤0

퐶

(a) 퐺5 푣3

푣2

푦 푤2

푤3 푒

푤1 푣1 푣0

푤0 (b) 퐺4 + 푒

푤5

Fig. A.5 Graphs 퐺5 and 퐺4 + 푒

holds for each nontrivial cut 퐹 of 퐻. Thus, 퐻 is either a brick or a brace. But 퐻 is not bipartite; hence 퐻 is a brick. Let us now use the subadditivity of function 푏 to prove that 퐺 푘 is a near-brick. The graph 퐻, a brick, is bicritical. Let 푀 be a perfect matching of 퐻 − 푣 0 − 푤 푘 . Addition of the edges 푣 0 푤 푘−1 and 푣 푘 푤 푘 to 푀 yields a perfect matching of 퐺 that contains three edges in 퐷. Thus, 퐷 is a separating cut of 퐺 which is not tight. Moreover, both 퐷-contractions of 퐺 are bricks. By the subadditivity of function 푏, the graph 퐺 푘 is a near-brick. Finally, let us prove that 퐺 푘 is free of nontrivial barrier cuts. We do this by showing that 퐺 푘 is bicritical. Let 푇 := 퐺 푘 [푌 ] be the triangle induced by the set 푌 . Let 푡1 and 푡2 denote two vertices of 퐺 푘 . If neither 푡1 nor 푡2 lies in 푇 then a perfect

514

Solutions to Selected Exercises

matching of 퐻 − 푡1 − 푡2 may be extended to a perfect matching of 퐺 푘 − 푡1 − 푡2 by the addition of the appropriate edge of 푇. Alternatively, if 푡1 and 푡2 lie both in 푇 then they are adjacent, and since 퐺 푘 is matching covered it follows that 푡1 푡2 is matchable, hence 퐺 푘 − 푡1 − 푡2 is matchable. We may thus assume that precisely one of 푡1 and 푡2 , say, 푡1 , lies in 푇. In that case, any perfect matching of 퐻 − 푦 − 푡2 may be extended to a perfect matching of 퐺 푘 − 푡1 − 푡2 by the addition of the appropriate edge of 푇. In sum, 퐺 푘 is a bicritical near-brick, that is, 퐺 푘 is a brick.

A.5 Chapter 5

Exercise 5.2.7 Suppose that 퐺 [푋] is bipartite. Let ( 퐴, 퐵) denote its bipartition. Assume, to the contrary, that 푋 contains two or more vertices. Graph 퐺, a brick, is 3-connected. Thus, the ends of the three edges of 퐶 in 푋 are distinct. Each end in 푋 of an edge in 퐶 is adjacent to two or more vertices in 푋 and a vertex of 푋 not incident with any edge of 퐶 is adjacent to three or more vertices in 푋. We conclude that | 퐴|, |퐵| ≥ 2. Let 푎 be the number of ends of edges of 퐶 in 퐴. Let 푏 denote the number of ends of edges of 퐶 in 퐵. Then, 푎 + 푏 = 3. Observe that, in 퐺 − 퐴, |퐵| − 푏 vertices of 퐵 are isolated. As 퐺 is bicritical and | 퐴| ≥ 2, we have that |퐵| − 푏 ≤ | 퐴| − 2. Likewise, | 퐴| − 푎 ≤ |퐵| − 2. Adding those two inequalities, we have that | 퐴| + |퐵| − (푎 + 푏) ≤ | 퐴| + |퐵| − 4, whence 푎 + 푏 ≥ 4, a contradiction.

Exercise 5.2.8 First obtain a perfect matching 푀 of 퐺, using the algorithm√ of Micali and Vazirani [70] for finding a maximum matching in a graph in 푂 (푚 푛)-time. Then, for each pair {푢, 푣} of vertices of 퐺, verify whether 퐺−푢−푣 is connected and matchable. Connectedness can be computed in linear time and checking if 퐺 − 푢 − 푣 has a perfect matching can also be computed in time 푂 (푚) by applying one iteration of Edmonds’ maximum matching algorithm (see Gabow and Tarjan (1985, [36])), to get an augmenting path between the two 푀-exposed vertices of 퐺 −푢 − 푣. If 퐺 −푢 − 푣 is not matchable then a Tutte set 푆 is obtained and 퐵 := 푆 + 푢 + 푣 is a (nontrivial) barrier of 퐺. In that case, as 퐺 is nonbipartite, at least one (odd) component, 퐾, of 퐺 − 퐵, is nontrivial and the cut 휕 (퐾) is a nontrivial tight cut of 퐺. Alternatively, if the graph 퐺 − 푢 − 푣 is matchable but not connected, it follows that each component of 퐺 − 푢 − 푣 is even. Let 퐿 be one such component. The cut 휕 (푉 (퐿) + 푢) is a nontrivial tight cut of 퐺. Finally, if 퐺 − 푢 − 푣 is matchable and connected for each pair {푢, 푣} of distinct vertices of 퐺 then √ 퐺 is a brick, by Theorem 5.7. The total running time of this algorithm is 푂 (푚 푛) + 푂 (푛2 ) × 푂 (푚) = 푂 (푚푛2 ).

Exercises of Chapter 5

515

Exercise 5.4.3 Suppose that 퐺 is a brace, let 푒 1 := 푢 1 푣 1 and 푒 2 := 푢 2 푣 2 be two nonadjacent edges of 퐺. By Theorem 5.17, 퐺 − 푢 1 − 푢 2 − 푣 1 − 푣 2 has a perfect matching, 푀. Then, 푀 + 푒 1 + 푒 2 is a perfect matching of 퐺 that contains both edges 푒 1 and 푒 2 . We conclude that each pair {푒 1 , 푒 2 } of nonadjacent edges of 퐺 is in a perfect matching of 퐺, hence 퐺 is 2-extendable. Conversely, suppose that 퐺 is 2-extendable. Let us first prove that 퐺 is matching covered. By hypothesis, 퐺 is connected; therefore it remains to be shown that every edge of 퐺 is matchable. Let 푒 1 := 푢 1 푣 1 , let us prove that 푒 1 is matchable. For this, we must prove that some edge of 퐺 − 푒 1 is nonadjacent to 푒 1 . For this, let 푢 2 ∈ 퐴 − 푢 1 and 푣 2 ∈ 퐵 − 푣 1 . If 푢 2 is not adjacent to 푣 1 then 푢 2 is adjacent to some vertex 푣 3 of 퐵 − 푣 1 ; in this case 푒 1 and 푢 2 푣 3 are nonadjacent edges of 퐺. We may thus assume that 푢 2 is adjacent to 푣 1 . Likewise, we may assume that 푣 2 is adjacent to 푢 1 . By hypothesis, 퐺 has a perfect matching, 푀, that contains the edges 푢 1 푣 2 and 푢 2 푣 1 . As 퐺 has order six or more, 퐴 − 푢 1 − 푢 2 contains a vertex, 푎, which is joined by an edge of 푀 to a vertex 푏 ∈ 퐵 − 푣 1 − 푣 2 . The edges 푒 1 and 푎푏 of 퐺 are nonadjacent. In all cases considered, we conclude that 퐺 contains an edge nonadjacent to 푒 1 . By hypothesis, 퐺 is 2-extendable, hence 푒 1 is matchable. This conclusion holds for each edge 푒 1 of 퐺, hence every edge of 퐺 is matchable. As 퐺 is connected, we deduce that 퐺 is matching covered. Let us now prove that 퐺 is a brace. For this, assume, to the contrary, that 퐺 has a nontrivial tight cut 퐶 := 휕 ( 푋). Adjust notation so that 푋+ ⊂ 퐵 and 푋− ⊂ 퐴, where | 푋+ | = | 푋− | + 1 (Theorem 4.8). If all the edges of 퐶 are incident with the same vertex, 푢, in 퐴 − 푋− , then 푁 ( 푋+ ) = 푋− + 푢 and |푁 ( 푋+ )| = | 푋+ |. In that case, 퐺 is not matching covered (Exercise 2.3.1). Likewise, if all the edges of 퐶 are incident with the same vertex, 푣, in 푋+ , then 푁 ( 푋+ − 푣) = 푋− and |푁 ( 푋+ − 푣)| = | 푋+ − 푣|. We may thus assume that 퐶 contains two nonadjacent edges, 푓1 and 푓2 . As 퐶 is tight, 퐺 has no perfect matching that contains both 푓1 and 푓2 , a contradiction to the hypothesis that 퐺 is 2-extendable. Indeed, 퐺 is a brace.

Exercise 5.4.4 Let ( 퐴, 퐵) be the bipartition of 퐺. Suppose that 푀 has an edge, 푒 := 푢푣, 푢 ∈ 퐴, 푣 ∈ 퐵, such that 퐻 := 퐺 − 푢 − 푣 is not matching covered. By hypothesis, the graph 퐺 has order six or more; hence 퐻 has order four or more. By Theorem 2.9, 퐴 − 푢 has a nonempty proper subset 푋 such that |푁 퐻 ( 푋)| ≤ | 푋 |. (A.10) The set 푀 − 푒 is a perfect matching of 퐻; therefore, 퐻 is matchable. We deduce that equality holds in (A.10). Let 푌 := 푋 ∪ 푁 퐻 ( 푋) ∪ {푣}. The set 퐶 := 휕 (푌 ) is a tight cut of 퐺. Moreover, 푋 is a proper subset of 퐴 − 푢. Thus, 푌 is nontrivial. We conclude that 퐶 is a nontrivial tight cut of 퐺. Hence, 퐺 is not a brace.

516

Solutions to Selected Exercises

Conversely, suppose that 퐺 − 푢 − 푣 is matching covered, for each 푢푣 ∈ 푀. To prove that 퐺 is a brace, let 푋 ⊂ 퐴 such that 0 < | 푋 | < | 퐴| − 1. We want to prove that |푁 ( 푋)| ≥ | 푋 | + 2. The graph 퐺 is matching covered, by hypothesis. From Theorem 2.9, we infer that |푁퐺 ( 푋)| ≥ | 푋 | + 1. Thus, 푀 contains an edge 푎푏 such that 푏 ∈ 푁퐺 ( 푋) and 푎 ∈ 퐴 − 푋. By hypothesis, the graph 퐽 := 퐺 − 푎 − 푏 is matching covered. As 0 < | 푋 | < | 퐴| − 1 = | 퐴 − 푎|, we deduce that the set 푋 is a nonnull proper subset of 퐴 − 푎. By Theorem 2.9, we infer that |푁퐺 ( 푋)| = |푁 퐽 ( 푋) + 푏| = |푁 퐽 ( 푋)| + 1 ≥ | 푋 | + 2. Thus, |푁퐺 ( 푋) ≥ | 푋 | + 2. This conclusion holds for each subset 푋 of 퐴 such that 0 < | 푋 | < | 퐴| − 1. By Theorem 5.17, 퐺 is a brace. Let us now describe a polynomial-time algorithm that either determines that 퐺 is a brace or, alternatively, determines a nontrivial tight cut of 퐺. For each edge 푒 := 푢푣 of 푀, 푢 ∈ 퐴 and 푣 ∈ 퐵, let 퐻푒 := 퐺 − 푢 − 푣 and let H := {퐻푒 : 푒 ∈ 푀 }. For each 퐻푒 in H , let 퐷 푒 be the directed graph as described in Exercise 2.2.6, with 퐻푒 playing the role of 퐺 and 퐷 푒 playing the role of 퐷. In linear time it is possible to determine whether 퐷 푒 is strongly connected. Thus, in linear time it is possible to determine whether 퐻푒 is matching covered. Verification of whether each 퐻푒 ∈ H is matching covered can thus be done in time 푂 ((푚 + 푛)푛). Moreover, suppose that 퐷 푒 is not strongly connected, for some 푒 ∈ 푀. The linear time search determines a directed cut 퐶 in 퐷 푒 . By definition of 퐷 푒 , it follows that 퐶 ⊆ 퐸 (퐻푒 ) − 푀. Thus, 퐶 is a cut of 퐻푒 which has a shore 푋 such that every edge of 퐶 is incident with a vertex in 푋 ∩ 퐵. Moreover, as 푀 − 푒 is a perfect matching of 퐻푒 , it follows that | 푋 ∩ 퐴| = | 푋 ∩ 퐵|. Consequently, the set 푌 := 푋 + 푣 is the shore of a nontrivial tight cut of 퐺. Exercise 5.4.5 Let 퐺 be the given matching covered graph. In linear time, determine whether 퐺 is bipartite. If 퐺 is bipartite, use the 푂 (푚푛)-time algorithm described in Exercise 5.4.4 to either attest that 퐺 is a brace, or alternatively, determine a nontrivial tight cut 퐶 of 퐺. Likewise, if 퐺 is nonbipartite, use the 푂 (푚푛2 )-time algorithm described in Exercise 5.2.8 to either attest that 퐺 is a brick or, alternatively, determine a nontrivial tight cut 퐶 of 퐺. If 퐺 is a brick or a brace then the tight cut decomposition consists solely of 퐺. Alternatively, if the nontrivial tight cut 퐶 is found, then recursively determine the tight cuts decompositions of both 퐶-contractions of 퐺 and combine the results. The number 푟 of nontrivial tight cuts of the tight cut decomposition is at most 푛/2 − 2 (Exercise 4.3.5). Thus, the two algorithms which are used to find nontrivial tight cuts are executed at most 2푟 + 1 ≤ 푛 − 3 times. Thus, the algorithm runs in time 푂 (푚푛3 ).

Exercises of Chapter 6

517

Exercise 5.4.6 Let 푆 be a set of two or three vertices of 퐺 such that neither 퐴 nor 퐵 contains three vertices of 푆. Adjust notation so that |푆 ∩ 퐴| ≥ |푆 ∩ 퐵|. Case 1 푆 = {푢 1 , 푣 1 }, where 푢 1 ∈ 퐴 and 푣 1 ∈ 퐵. Let 퐻 := 퐺 − 푆 and let 푎 ∈ 퐴 − 푢 1 and 푏 ∈ 퐵 − 푣 1 . By hypothesis, 퐺 is a brace. By Theorem 5.17, 퐻 − 푎 − 푏 has a perfect matching, 푀. As 퐺 has order six or more, it follows that 푀 has at least one edge. Thus, 퐻 has at least one edge and 퐻 − 푎 − 푏 is matchable, for each 푎 ∈ 퐴 and 푏 ∈ 퐵. As stated in Exercise 3.1.7, 퐻 is matching covered; hence 퐺 − 푆 = 퐻 is 2-connected. Case 2 푆 = {푢 1 , 푢 2 , 푣 1 }, where 푢 1 , 푢 2 ∈ 퐴 and 푣 1 ∈ 퐵. By Case 1, 퐺 − 푢 1 − 푣 1 is 2-connected, hence 퐺 − 푆 is connected. Case 3 푆 = {푢 1 , 푢 2 } ⊂ 퐴. Let 푣 1 be a vertex in 퐵. By Case 2, 퐺 − 푢 1 − 푢 2 − 푣 1 is connected. As 퐺, a brace, has order six or more, the vertex 푣 1 is adjacent to a vertex in 퐴 − 푆; otherwise 푆 + 푣 1 is the shore of a nontrivial tight cut of 퐺. It follows that 퐺 − 푆 is connected.

A.6 Chapter 6

Exercise 6.1.2 For every perfect matching 푀, clearly 휒 푀 satisfies the non-negativity and degree constraints. For every odd set 푆 of vertices, 푀 contains an odd number of edges in 휕 (푆). Thus, 휒 푀 also satisfies the odd set constraints. Consider the linear combination Õ x := 훼 푀 휒 푀 , 0 ≤ 훼 푀 . 푀 ∈ M

Then, x satisfies the non-negativity constraints. Moreover, for every vertex 푣, we have that x(휕 (푣)) = 푠, and for every odd set 푆 of vertices, we have that x(휕 (푆)) ≥ 푠, Í where 푠 := 푀 ∈ M 훼 푀 . In particular, if the linear combination is convex then 푠 = 1 and x satisfies the three constraints. Exercise 6.1.4 Figure A.6 depicts w1 expressed as a convex linear combination of perfect matchings of 퐺 1 , Figure A.7 depicts w2 expressed as a convex linear combination of perfect matchings of 퐺 2 and Figure A.8 depicts w expressed as a convex linear combination.

Solutions to Selected Exercises

518

1 2 1 3

1 6

1 6

+

+

1 2

1 3

Fig. A.6 w1 expressed as a convex linear combination 1 6

1 3

1 6

1 6 1 3

+

1 6

+

1 3

1 6

+ 1 6

1 3

1 3 1 3

Fig. A.7 w2 expressed as a convex linear combination 1 6 1 6

1 6

+

1 3 1 3

1 3

1 3

1 6

1 6

1 3 1 3 1 6

+

1 6 1 6

Fig. A.8 w expressed as a convex linear combination

1 3

1 3

+

Exercises of Chapter 6

519

Exercise 6.1.7 Õ Let x := 훼 푀 휒 푀 . Our task is to prove that w = x. For this, we shall prove that 푀 ∈ M

w( 푓 ) = x( 푓 ), for each edge 푓 in 퐸. Adjust notation so that 푓 ∈ 퐸 (퐺 1 ). Let M1 ( 푓 ) denote the set of perfect matchings of 퐺 1 that contain edge 푓 and let N ( 푓 ) denote the set of those matchings in N that contain edge 푓 . For each edge 푒 of 퐶, let M1 (푒, 푓 ) denote the set of those matchings in M1 ( 푓 ) that contain edge 푒, let N (푒, 푓 ) denote the set of those matchings of N ( 푓 ) that contain edge 푒 and let M2 (푒) denote the set of perfect matchings of 퐺 2 that contain edge 푒. Clearly, {M1 (푒, 푓 ) : 푒 ∈ 퐶} is a partition of M1 ( 푓 ) and {N (푒, 푓 ) : 푒 ∈ 퐶} is a partition of N ( 푓 ). Thus, w( 푓 ) = w1 ( 푓 ) = w(푒) = w2 (푒) = x( 푓 ) = =

Õ

Õ

푀 ∈ N ( 푓 )

Õ

Õ

훽 푀1 =

훼 푀 =

훾 푀2

Õ

Õ

Õ

훽 푀1 ,

(A.11)

푒∈퐶 푀1 ∈ M1 (푒, 푓 )

푀1 ∈ M1 ( 푓 ) 푀2 ∈ M2 (푒)

Õ

Õ

Õ

(∀푒 ∈ 퐶),

(A.12)

훼 푀

푒∈퐶 푀 ∈ N (푒, 푓 )

푒∈퐶 푀1 ∈ M1 (푒, 푓 ) 푀2 ∈ M2 (푒)

훼 푀1 ∪푀2 .

(A.13)

From (A.11), (A.12) and (A.13), we then deduce that w( 푓 ) = =

Õ

Õ

Õ

Õ

푒∈퐶 푀1 ∈ M1 (푒, 푓 )

Õ 훽 푀1 훾 푀2 w(푒) 푀2 ∈ M2 (푒) Õ 훼 푀1 ∪푀2 = x( 푓 ).

푒∈퐶 푀1 ∈ M1 (푒, 푓 ) 푀2 ∈ M2 (푒)

Exercise 6.1.12 If 푆 is empty then x = 0, a vector in Cone(퐺). We may thus assume that 푆 is nonempty. The non-negativity of x then implies that 푟 > 0. (i) Let 퐻 := 퐺 [푆] and let 푋 ⊆ 푉 (퐻). For each 퐾 ∈ O(퐺 − 푋), x(휕 (퐾)) ≥ 푟. Thus, Õ Õ 푟 · 표(퐺 − 푋) ≤ x(휕 (퐾)) ≤ x(휕 ( 푋)) ≤ x(푣) = 푟 · | 푋 |. 퐾 ∈ O (퐺−푋)

푣 ∈푋

As 푟 > 0, we deduce that 표(퐺 − 푋) ≤ | 푋 |. This conclusion holds for each subset 푋 of 푉 (퐻); hence 퐻 is matchable, by Tutte’s Theorem. (ii) The proof in this case is like Case 1 of the proof of Theorem 6.1.

Solutions to Selected Exercises

520

(iii) Let 푀 be a perfect matching of 퐺 [푆]. We prove that the inequalities (6.6a), (6.6b) and (6.6c) hold for y and 푟 − 휖 playing respectively the roles of x and 푟. By definition, 휖 ≤ x(푒), for each edge 푒. Thus, y(푒) = x(푒) − 휖 ≥ 0. This establishes the validity of inequality (6.6a). For each vertex 푣, y(휕 (푣)) = x(푣) − 휖 휒 푀 (휕 (푣)) = 푟 − 휖. The validity of inequality (6.6b) is established. For each odd cut 퐶 such that |푀 ∩ 퐶| = 1, y(퐶) = x(퐶) − 휖 휒 푀 (퐶) ≥ 푟 − 휖. For each odd cut 퐶 such that |푀 ∩ 퐶| > 1, by definition of 휖, y(퐶) = x(퐶) − 휖 휒 푀 (퐶) ≥ x(퐶) − 푟 − Note that if 휖 =

x(퐶) − 푟 ≥ 푟 − 휖. |푀 ∩ 퐶| − 1

x(퐶) − 푟 |푀 ∩ 퐶| = |푀 ∩ 퐶| − 1

x(퐶) − 푟 then y(퐶) = 푟 −휖. The validity of inequality (6.6c) |푀 ∩ 퐶| − 1

is established. By definition of 휖, either 휖 = x(푒) for some edge 푒 of 푆, or y(퐶) = 푟 − 휖, for some (nontrivial) odd cut 퐶 such that |푀 ∩ 퐶| > 1. In the former case, the support of y is a proper subset of 푆 and the assertion follows by induction. In the latter case, the proof proceeds as in (ii).

Exercise 6.2.1 Let x be a 1-regular vector and let y :=

Í

푀 ∈ M

|M | −1 휒 푀 .

Let C := {퐶 : 퐶 is an odd cut, x(퐶) < 1 < y(퐶)} and let 퐹 := {푒 ∈ 퐸 : x(푒) < 0}. If 퐹 ∪ C is empty then x ∈ Poly(퐺). We may thus assume that 퐹 ∪ C is nonempty. By definition of y and 퐹, x(푒) < 0 < y(푒), for each edge 푒 ∈ 퐹. By definition of y and C, x(퐶) < 1 < y(퐶), for each cut 퐶 ∈ C. Thus, 휖 := min



 y(퐶) − 1  y(푒) : 푒 ∈ 퐹 ∪ : 퐶 ∈ C y(푒) − x(푒) y(퐶) − x(퐶)

satisfies the inequalities 0 < 휖 < 1. Let z := (1 − 휖)y + 휖x. We must now prove that z ∈ Poly(퐺). Let 푒 be an edge of 퐺. By definition of y, we have that y(푒) > 0. If 푒 ∉ 퐹 then x(푒) ≥ 0; hence z(푒) = (1 − 휖)y(푒) + 휖x(푒) > 0. Alternatively, if 푒 ∈ 퐹 then 0 < 휖 ≤

y(푒) ; y(푒) − x(푒)

hence z(푒) = y(푒) − 휖 (y(푒) − x(푒)) ≥ 0.

In both alternatives, z satisfies the inequalities (6.3a). Let 퐶 be a tight cut of 퐺. As x and y are both 1-regular, we deduce that z(퐶) = 1; hence z is 1-regular. Thus, x satisfies the inequalities (6.3b).

Exercises of Chapter 6

521

Let 퐶 be an odd cut of 퐺. If 퐶 ∉ C then x(퐶), y(퐶) ≥ 1; hence z(퐶) = (1 − 휖)y(퐶) + 휖x(퐶) ≥ 1. Alternatively, if 퐶 ∈ C then 0 < 휖 ≤

y(퐶) − 1 ; y(퐶) − x(퐶)

hence z(퐶) = y(퐶) − 휖 (y(퐶) − x(퐶)) ≥ 1.

In both alternatives, z satisfies the inequalities (6.3c). Indeed, z ∈ Poly(퐺). Exercise 6.2.5 Let B0 denote any basis of Lin0 (퐺) and let y be any 1-regular vector of R퐸 . Let us first show that B := B0 ∪ {y} is linearly independent. For this, consider a linear combination Õ 훼b b + 훽y = 0. b∈ B0

Note first that 훽 = 0; else y can be expressed as a linear combination of vectors in B0 , a contradiction, as B0 is a basis of Lin0 (퐺) and y ∉ Lin0 (퐺). As 훽 = 0, the linear independence of B0 then implies that 훼b = 0, for each b ∈ B0 . We conclude that B is linearly independent. Let us now prove that B spans Lin(퐺). For this, let x ∈ Lin(퐺). Then, x is 푟-regular, for some 푟 ∈ R. If 푟 = 0 then x ∈ Lin0 (퐺) and x is equal to some linear combination of vectors in B0 . Alternatively, if 푟 ≠ 0 then 1 푟 x − y ∈ Lin0 (퐺); therefore x is equal to some linear combination of vectors in B. In both alternatives we deduce that B spans Lin(퐺). Exercise 6.2.6 (i) Let x ∈ R퐸 . By Theorem 6.7, x ∈ Lin(퐺) iff x is regular. Thus, for any specified vertex 푢, x ∈ Lin(퐺) iff 휒퐶 x − 휒 휕(푢) x = 0 for each 퐶 ∈ T − {휕 (푢)}. That is, x ∈ Lin(퐺) iff Bx = 0. It follows that the rows of B span (Lin(퐺)) ⊥ . Í (ii) Let p ∈ 푉 R such that 푣 ∈푉 p(푣) = 0. Let x ∈ R퐸 be defined by the equality x(푢푣) := p(푢) + p(푣) for Í each edge 푢푣 of 퐺. LetÍ푀 be any perfect matching of 퐺. Then, x( 휒 푀 ) = 푢푣 ∈ 푀 [p(푢) + p(푣)] = 푣 ∈푉 p(푣) = 0. Indeed, x is matching orthogonal. (iii) Assume that 퐺 has at most one brick. From Exercise 6.2.3(ii), we know that x is regular if and only if x(푢) = x(푣), for any two vertices 푢 and 푣 of 퐺. In that case, we may use the 푛 ×푚 incidence matrix A of 퐺, instead of T in the previous items. Let x ∈ (Lin(퐺)) ⊥ . Then, x may be expressed as a linear combination of rows of B: Õ   x= 푝(푣) 휒 휕(푣) − 휒 휕(푢) . 푣 ∈푉 −푢

522

Let 푝(푢) : = − Then,

Õ

Solutions to Selected Exercises

Õ

푝(푣).

푣 ∈푉 −푢

푝(푣) = 0 and

푣 ∈푉

x=

Õ

푣 ∈푉 −푢

푝(푣) 휒 휕(푣) + 푝(푢) 휒 휕(푢) =

Õ

푝(푣) 휒 휕(푣) .

푣 ∈푉

It follows that for every edge 푣푤 of 퐺, x(푣푤) = 푝(푣) + 푝(푤).

Exercise 6.2.7 For each vertex 푣 of 퐺, let A푣 denote the row of A corresponding to 푣. Likewise, for every edge 푒 of 퐺, let A푡푒 denote the column of A that corresponds to edge 푒. Let us first prove that (A.14) rank(A) ≥ 푛 − 1. For this, we need the following auxiliary result. A.1 Let 푇 be any nontrivial subtree of 퐺. The set 푆(푇) := {A푡푒 : 푒 ∈ 퐸 (푇)} is linearly independent. Proof by induction on |퐸 (푇)|. As 푇 is nontrivial, it has a vertex of degree one, say 푣. Let 푒 be the edge of 푇 incident with 푣. Consider any linear combination on the set 푆(푇) that produces the vector 0. The entry A푣푒 is the only nonnull entry in the row A푣 of A. Thus, the coefficient corresponding to 푒 is zero. If 푒 is the only edge of 푇 then we are done. Otherwise, the restriction of the linear combination to 푆(푇 − 푒) also produces 0; therefore all its coefficients are equal to zero, by induction. Indeed, ƒ the set 푆(푇) is linearly independent. The graph 퐺 is connected; therefore it has a spanning tree, 푇. The set 푆(푇) of the 푛 − 1 columns corresponding to the edges of 푇 is linearly independent. We then deduce the validity of (A.14). Let us first consider the case in which 퐺 has a bipartition, ( 푋, 푌 ). Let us show that the 푛 rows of A are not linearly independent. For this, consider the linear combination Õ Õ y := A푣 − A푣 . 푣 ∈푋

푣 ∈푌

For each edge 푒 of 퐺, we have that y(푒) = 0. Thus, y = 0. We conclude that rank(A) ≤ 푛 − 1. From (A.14), we deduce that rank(A) = 푛 − 1, if 퐺 is bipartite. Let us consider next the case in which 퐺 is not bipartite. For some odd positive integer 푟, let 푣 1 푣 2 . . . 푣 푟 푣 1 be an odd cycle of 퐺. For each vertex 푣 of 퐺, let 훼푣 ∈ R; let Õ x := 훼푣 A푣 . (A.15) 푣 ∈푉

Exercises of Chapter 7

523

Assume that x = 0. For 푖 = 1, 2, . . . , 푟 − 1, we must have that x(푣 푖 푣 푖+1 ) = 0, hence 훼푣푖 = −훼푣푖+1 . As the length 푟 is odd, it follows that 훼푣푖 = 0, for 푖 = 1, 2, . . . , 푟. Thus, some coefficients 훼푣 in (A.15) are equal to zero. From (A.14), we deduce that all coefficients are equal to zero. That is, the set of rows of A is linearly independent. We deduce that rank(A) = 푛, if 퐺 is not bipartite.

Exercise 6.3.3 Consider any ear decomposition (퐺 1 , 퐺 2 , . . . , 퐺 푟 ) of 퐺, where 푟 ≥ 1, 퐺 1 = 퐾2 and 퐺 푟 = 퐺. The proof is by induction on 푟. If 푟 = 1 then the assertion holds immediately, with 푀1 := 퐸 (퐺 1 ). We may thus assume that 푟 > 1. Let 푃푟 denote the ear added to 퐺 푟 −1 to obtain 퐺 푟 . By induction, 퐺 푟 −1 has 푟 − 1 perfect matchings 푁푖 , 1 ≤ 푖 ≤ 푟 − 1, such that every integer vector in Lin(퐺 푟 −1 ) is an integral linear combination of { 휒푖푁 : 1 ≤ 푖 ≤ 푟 − 1}. Each perfect matching 푁푖 may be extended to a perfect matching 푀푖 of 퐺 by adding the even numbered edges of 푃푟 . In addition, the odd numbered edges of 푃푟 are contained in some perfect matching 푀푟 of 퐺. Let x be an integral vector in Lin(퐺). Let 푑 := x(푢) for some vertex 푢 of 퐺. Then, x(푣) = 푑, for each vertex 푣 of 퐺. In particular, as this equality applies to the internal vertices of 푃푟 , it follows that for some integer 푑 ′ , (i) x(푒) = 푑 ′ if 푒 is an odd numbered edge of 푃푟 and (ii) x(푒) = 푑 − 푑 ′ , if 푒 is an even numbered edge of 푃푟 . Then, the restriction y of x − 푑 ′ 휒 푀푟 to 퐺 푟 −1 is an integral vector of Lin(퐺 푟 −1 ), because y is (푑 − 푑 ′ )-regular. By induction, y is an integral linear combination of the incidence vectors of 푁푖 , 1 ≤ 푖 ≤ 푟 − 1. It follows that x − 푑 ′ 휒 푀푟 is an integral linear combination of 휒 푀푖 , for 1 ≤ 푖 ≤ 푟 − 1. Thus, x is an integral linear combination of 휒 푀푖 , for 1 ≤ 푖 ≤ 푟.

A.7 Chapter 7

Exercise 7.2.3 (i) By hypothesis, the vector x is 1-regular and non-negative. Moreover, x(퐷) < 1; thus Õ x(휕 (퐾)) ≤ x(휕 (푆)) = |푆| − 1 + x(퐷) − 2x(퐸 (퐺 1 [푆])) < |푆|. 퐾 ∈ O (퐺1 −푆)

As 표(퐺 1 − 푆) ≥ |푆|, it follows that Í

퐾 ∈ O (퐺1 −푆)

x(휕 (퐾))

표(퐺 1 − 푆)

Hence x(휕 (퐾)) < 1 for some 퐾 ∈ O(퐺 1 − 푆).

< 1.

Solutions to Selected Exercises

524

(ii) Let 퐹 := 휕 (퐾). Let us prove that 퐹 ≺ 퐷. For this, let 푀 be a perfect matching of 퐺. For each odd component 퐽 of 퐺 1 − 푆, the cut 휕 (퐽) has at least one edge in 푀. Thus,

≤ =

|푀 ∩ 퐹| + 표(퐺 1 − 푆) − 1 Õ |푀 ∩ 휕 (퐽)| ≤ |푀 ∩ 휕 (푆)|

퐽 ∈ O (퐺1 −푆)

|푀 ∩ 퐷| + |푆| − 1 − 2|푀 ∩ 퐸 (퐺 1 [푆])|

.

As |푆| ≤ 표(퐺 1 − 푆), it follows that |푀 ∩ 퐹| ≤ |푀 ∩ 퐷|,

(A.16)

with equality only if |푆| = 표(퐺 1 − 푆) and 푀 ∩ 퐸 (퐺 1 [푆]) is empty. If |푆| < 표(퐺 1 − 푆) then |푀 ∩ 퐹| < |푀 ∩ 퐷| and this inequality holds for each perfect matching 푀 of 퐺. In that case, 퐹 ≺ 퐷. We may thus assume that |푆| = 표(퐺 1 −푆). By hypothesis, 퐺 contains an edge, 푒, having both ends in 푆. From (A.16) we infer that |푀 ∩ 퐹| ≤ |푀 ∩ 퐷|, with equality only if 푒 ∉ 푀. As 퐺 is matching covered, it has a perfect matching that contains edge 푒. We conclude that 퐹 ≺ 퐷.

Exercise 7.3.3 Let w be an extreme point of P, let 푆 be the support of w, let 퐻 := P[푆] and let 퐾 be a component of 퐻. By Theorem 7.7, 퐾 is either 퐾2 or an odd cycle. If 퐾 contains only one edge, 푒, then w(푒) = 1, whereas if 퐾 is an odd cycle then w(푒) = 1/2, for each edge 푒 of 퐾. As P has an even number of vertices, an even number of odd cycles are components of 퐻. The Petersen graph has girth five; therefore, either 푆 is a perfect matching of P or 퐻 consists of two vertex-disjoint pentagons. The set M of perfect matchings of P consists of precisely six perfect matchings. Each perfect matching 푀 is a separating cut whose shores induce two edge-disjoint pentagons 푄 1 and 푄 2 , where 퐸 (푄 1 ) ∪ 퐸 (푄 2 ) = 퐸 − 푀. Thus, the set of extreme points of Poly(P) is a subset of S := { 휒 푀 , 1/2 · 휒 퐸− 푀 : 푀 ∈ M}. Let 푀 ∈ M. The vector 휒 푀 is the only vector in S whose support does not contain any edge in 퐸 − 푀 and the vector 휒 퐸− 푀 is the only vector in S whose support does not contain any edge in 푀. Thus, both 휒 푀 and 1/2 · 휒 퐸− 푀 are in S. This conclusion holds for each 푀 ∈ M. Thus, S is the set of extreme points of Poly(P).

Exercises of Chapter 8

525

Exercise 7.3.5 Let {퐶1 , 퐶2 } be any pair of vertex-disjoint disjoint odd cycles of 퐺. We shall show that 퐺 − (푉 (퐶1 ) ∪ 푉 (퐶2 )) is not matchable and then use the Reed-Wakabayashi Theorem 7.8 to deduce that 퐺 is solid. Let 퐴 := {0, 1, 3, 5, ℎ 2 , 푣 2 } and let 퐵 := {2, 4, 6, ℎ 1 , ℎ3 , 푣 1 }. Note that ( 퐴, 퐵) is an equipartition of the set 푉 (퐺). The set 퐸 퐴 := {01, ℎ2 푣 2 } is the set of edges of 퐺 having both ends in 퐴 and the set 퐸 퐵 := {ℎ1 푣 1 } is the set of edges of 퐺 having both ends in 퐵. Suppose that one of the cycles 퐶1 and 퐶2 , say 퐶1 , contains edge ℎ1 푣 1 . Then, it also contains a vertex in {0, 1} and a vertex in {ℎ 2 , 푣 2 }, which implies that 퐶1 contains an end of each of the three edges of 퐸 퐴 ∪ 퐸 퐵 . This implies that 퐺 − 푉 (퐶1 ) is bipartite, a contradiction. Thus, no cycle in {퐶1 , 퐶2 } contains edge ℎ1 푣 1 . That edge is the only edge of 퐸 퐵 . Moreover, 퐸 퐴 is a doubleton; thus each cycle in {퐶1 , 퐶2 } contains one of the edges of 퐸 퐴. As {퐴, 퐵} is an equipartition of 푉 (퐺), the graph 퐺 − (푉 (퐶1 ) ∪ 푉 (퐶2 )) is not matchable. It now follows from the Reed-Wakabayashi Theorem 7.8 that 퐺 is solid.

A.8 Chapter 8

Exercise 8.1.7(iii) The vectors 휒 푒 and 휒 푓 are incidence vectors of the singletons {푒} and { 푓 }. Thus the vector x = 휒 푒 − 휒 푓 has precisely two nonzero entries: x(푒) = 1, x( 푓 ) = −1, and x(푔) = 0 for each edge 푔 ∉ {푒, 푓 }. Since every perfect matching of the brick 퐺 which contains the edge 푒 also contains the edge 푓 , it follows that x is a matching orthogonal vector of 퐺. By part (ii), it follows that there is a real-valued function 푝 Í on the vertex set of 퐺 (a labelling of the vertices of 퐺) such that (i) 푣 ∈푉 푝(푣) = 0 and, (ii) x(푠푡) = 푝(푠) + 푝(푡), for any edge 푠푡 of 퐺. Thus, in particular, if 푠푡 is any edge different from 푒 and 푓 , then 푝(푠) = −푝(푡). If possible, let 푤 1 푤 2 . . . 푤 2푘+1 푤 1 be an odd cycle in the graph 퐺 − 푒 − 푓 . Then, the above observation implies that 푝(푤 1 ) = −푝(푤 1 ). This is possible only if 푝(푤 푖 ) = 0 for every vertex on this odd cycle. Also since 퐺 − 푒 − 푓 is matching covered, it is connected, and any two of its vertices are connected by a path. Thus, if 퐺 − 푒 − 푓 has an odd cycle, it would follows that 푝(푣) = 0 for each 푣 ∈ 푉. But this impossible because, if 푒 = 푢 1 푢 2 , then 푝(푢 1 ) + 푝(푢 2) has to be equal to one. We conclude 퐺 −푒− 푓 is bipartite.

Exercise 8.1.10 Let ( 퐴, 퐵) be the bipartition of 퐺 and adjust notation so that 푎 ∈ 퐴 and 푏 ∈ 퐵. Suppose that 퐺 − 푒 is not a brace; then 푎 and 푏 belong to the minority parts of different shores of any nontrivial tight cut of 퐺 − 푒.

Solutions to Selected Exercises

526

(i) By hypothesis, the vertex 푎 is in 푋 ∩ 푌 . As the cuts 퐶 − 푒 and 퐷 − 푒 are both nontrivial and tight in 퐺 − 푒, it follows that 푎 ∈ 푋− ∩ 푌− and 푏 lies in ( 푋) − ∩ (푌 ) − . As 푎 ∈ 퐴 and 푏 ∈ 퐵, we deduce that 푋+ and 푌+ are both subsets of 퐵. Now consider the quadrant 푋 ∩ 푌 . If it is empty then 푋 is a subset of 푌 ; in that case, 퐼 = 퐶 and the assertion holds immediately. We may thus assume that 푋 ∩ 푌 is nonempty. As 퐶 − 푒 is tight in 퐺 − 푒, the subgraph of 퐺 − 푒 induced by 푋 is connected, therefore there is an edge in 퐺 − 푒 that joins a vertex 푎 1 ∈ 푋 ∩ 푌 to a vertex 푏 1 ∈ 푋 ∩ 푌 . See Figure A.9. 푌

푎 푋

푋 ∩ 푌

푋 ∩ 푌 푏1

푏2

푎1

퐶 푎2

푋 ∩ 푌

푋 ∩ 푌 퐷

Fig. A.9 The four quadrants.

The vertex 푏 1 is in 푌+ , in turn a subset of 퐵; hence 푎 1 ∈ 퐴. Likewise, the subgraph of 퐺 − 푒 induced by 푌 is connected; hence 퐺 − 푒 has an edge that joins a vertex 푏 2 in 푋 ∩ 푌 to a vertex 푎 2 in 푋 ∩ 푌 . The vertex 푏 2 is in 푋+ , in turn a subset of 퐵. Thus, 휕 ( 푋 ∩ 푌 ) − 푒 contains an edge incident with a vertex in 퐴 ∩ 푋 ∩ 푌 and an edge incident with a vertex in 퐵 ∩ 푋 ∩ 푌 . Therefore, 휕 ( 푋 ∩ 푌) − 푒 is not tight in 퐺 − 푒. By Theorem 4.12, one of the cuts 퐼 − 푒 and 휕 ( 푋 ∩ 푌 ) − 푒 is tight in 퐺 − 푒. We conclude that 퐼 − 푒 is tight in 퐺 − 푒. The vertex 푎 is in 푋 ∩ 푌 ∩ 퐴 and the vertex 푏 1 is in 푋 ∩ 푌 ∩ 퐵. Thus, 푋 ∩ 푌 is nontrivial. Likewise, the vertex 푎 1 is in 푋 ∩ 푌 ∩ 퐴 and the vertex 푏 2 is in 푋 ∩ 푌 ∩ 퐵. Thus, both shores of 퐼 are nontrivial. Indeed, 퐼 − 푒 is a nontrivial tight cut of 퐺 − 푒. A similar reasoning may be used to prove that 푈 − 푒 is a nontrivial tight cut of 퐺 − 푒. (ii) Let C be the family of tight cuts corresponding to a tight cut decomposition of 퐺 − 푒. For each nontrivial cut 퐶 := 휕 ( 푋) of 퐺 such that 퐶 − 푒 is tight in 퐺 − 푒, one of 푎 and 푏 is in 푋− , the other is in ( 푋) − . Let X denote the collection of the shores 푋 of cuts in C such that 푎 ∈ 푋 (and 푏 ∈ 푋). For each 푋 ∈ X, | 푋+ | = | 푋− | + 1 and 푎푏 is the only edge of 퐺 that joins a vertex in 푋− to a vertex in ( 푋) − . A.1.1 For distinct 푋 and 푌 in X, one of 푋 and 푌 is a (proper) subset of the other.

Exercises of Chapter 8

527

Proof For distinct 푋 and 푌 in X, we have that 푎 ∈ 푋 ∩ 푌 and 푏 ∈ 푋 ∩ 푌 . As 휕 ( 푋) and 휕 (푌 ) do not cross, it follows that either 푋 ∩ 푌 is empty or 푋 ∩ 푌 is empty. Consequently, one of 푋 and 푌 is a (proper) subset of the other. ƒ Let 푘 := |X|. Define now inductively the sequence 푋1 , 푋2 , . . . , 푋 푘 of the sets in X as follows: having defined 푋1 , 푋2 , . . . , 푋푖−1 , 1 ≤ 푖 ≤ 푘, let 푋푖 be a minimal set in X − {푋1 , 푋2 , . . . , 푋푖−1 }. We shall now prove that 푋1 ⊂ 푋2 ⊂ · · · ⊂ 푋푖 , for 푖 = 1, 2, . . . , 푘, by induction on 푖. The basis of the inductive proof is for 푖 = 1, and the assertion holds trivially. Let us now consider the inductive step, suppose that 1 < 푖. By (A.1.1), one of 푋푖−1 and 푋푖 is a proper subset of the other. By the minimality of 푋푖−1 in the collection {푋푖−1 , 푋푖 . . . , 푋 푘 }, we conclude that 푋푖−1 ⊂ 푋푖 . By induction, 푋1 ⊂ 푋2 ⊂ · · · ⊂ 푋푖−1 . Thus, 푋1 ⊂ 푋2 ⊂ . . . 푋푖−1 ⊂ 푋푖 . We conclude that 푋1 ⊂ 푋2 ⊂ · · · ⊂ 푋 푘 .

(A.17)

See Figure A.10. Let us now prove (8.1). From (A.17), we infer that 휕 ( 푋1 ) is the

푎 푋1

푋2

푋3

푒 푏 Fig. A.10 An illustration for Exercise 8.1.10, with 푘 = 3

only cut of C in 퐻1 := (퐺 − 푒)/푋1 , and it is trivial in 퐻1 , hence 퐻1 is a brace, by definition of C. Likewise, for each 푖 such that 2 ≤ 푖 < 푘, we infer from (A.17) that 휕 ( 푋푖 ) and 휕 ( 푋푖+1 ) are the only two cuts of C in 퐻푖 := (퐺 − 푒)/푋푖 /푋푖+1 , and they are both trivial in 퐻푖 , hence 퐻푖 is a brace, by definition of C. Finally, from (A.17) we infer that 휕 ( 푋 푘 ) is the only cut of C in 퐻 푘 := (퐺 − 푒)/푋 푘 , and it is trivial in 퐻 푘 , hence 퐻 푘 is a brace, by definition of C. (iii) The graph 퐻1 , which is (퐺 − 푒)/푋1 , is a brace. Thus, it is free of nontrivial tight cuts. Let 푌 be a shore of a nontrivial tight cut of 퐺 − 푒 such that 푎 ∈ 푌− . The cut 휕 ( 푋1 ∩ 푌 ) − 푒 is nontrivial and tight in 퐺 − 푒, therefore it is a trivial cut of 퐻1 . It follows that 푋1 ⊆ 푌 . This conclusion holds for each shore 푌 of a nontrivial tight cut of 퐺 − 푒 that contains the vertex 푎. Thus, 푋1 is unique. A similar reasoning may be used to prove the uniqueness of 푋 푘 . Consequently, if 퐺 − 푒 has three or less braces then X is unique.

Solutions to Selected Exercises

528

Exercise 8.3.8 The proof is by induction on |C|. Let us first consider the basis case, in which C is empty. The maximality of C implies that 퐺 is free of even 2-cuts. Let 푣 be a vertex of 퐺 having degree Δ(퐺) and let 휕 (푣) := {푒 푖 : 1 ≤ 푖 ≤ Δ(퐺)}. For each 푒 푖 , let 푅푖 denote a minimal dependence class induced by 푒 푖 . As 퐺 is free of even 2-cuts, it follows that 퐺 − 푅푖 is matching covered. That is, 푅푖 is a removable class of 퐺. Clearly, 푅푖 and 푅 푗 are distinct, for 1 ≤ 푖 < 푗 ≤ Δ(퐺). The assertion holds in this case. Let us now consider the induction step. Let {푒, 푓 } := 휕 ( 푋) be an even 2-cut in C. Let 푢 1 and 푢 2 denote the two ends of 푒 and let 푣 1 and 푣 2 denote the two ends of 푓 . Adjust notation so that 푢 1 and 푣 1 are both in 푋. Let 퐺 1 denote the graph obtained from 퐺 [푋] by the addition of edge 푢 1 푣 1 . Likewise, let 퐺 2 denote the graph obtained from 퐺 [ 푋] by the addition of edge 푢 2 푣 2 . For 푖 = 1, 2, note that 퐺 푖 is isomorphic to a 퐶푖 -contraction of 퐺, where 퐶푖 is one of the two barrier cuts associated with the barrier {푢 1 , 푣 2 } of 퐺. Thus, 퐺 1 and 퐺 2 are matching covered. Moreover, for 푖 = 1, 2, we have the following properties: 1. 훿(퐺 푖 ) ≥ 3. 2. The restriction C푖 of C − {푒, 푓 } to 퐸 (퐺 푖 ) is a maximal laminar collection of even 2-cuts of 퐺 푖 . 3. Every removable class of 퐺 푖 that does not contain edge 푢 푖 푣 푖 is a removable class of 퐺. Moreover, Δ(퐺) = max{Δ(퐺 푖 ) : 푖 = 1, 2}. Without loss of generality, assume that Δ(퐺) = Δ(퐺 1 ). Let 푟 푖 denote the number of removable classes of 퐺 푖 . Let 푟 be the number of removable classes of 퐺. By induction, 푟 1 ≥ Δ(퐺 1 ) + |C1 | 푟 2 ≥ Δ(퐺 2 ) + |C2 | 푟 ≥ (푟 1 − 1) + (푟 2 − 1). Addition of the three inequalities and simplification yields   푟 ≥ Δ(퐺) + Δ(퐺 2 ) + |C| − 1 − 2 ≥ Δ(퐺) + |C|. Exercise 8.4.1 (i) Suppose that 푣 is a vertex of degree three or more in 퐺 − 푒. Suppose, to the contrary, that two edges, 푒 1 and 푒 2 , both incident with 푣, are not removable in 퐺 − 푒. Apply Theorem 8.4, with 푟 = 2. Edge 푒 is not incident with vertices in both 퐵1 and 퐵2 . Adjust notation so that 푒 is not incident with any vertex in 퐵1 . As 퐺 is a brace of order six or more, vertex 푏 1 is adjacent to three or more vertices in 퐺. As 푒 is not incident with any vertex in 퐵1 , in particular 푏 1 is adjacent to two or more vertices of 퐴1 . Thus, | 퐴1 | ≥ 2. As 푒 is not incident with

Exercises of Chapter 9

529

any vertex of 퐵1 , we conclude that 휕 ( 퐴1 ∪ (퐵1 − 푏 1 )) is a nontrivial tight cut of 퐺, a contradiction to the hypothesis that 퐺 is a brace. (ii) Let 퐻 be a 퐶-contraction of 퐺 − 푒, say, 퐻 := (퐺 − 푒)/( 푋 → 푥). Let ( 퐴, 퐵) denote the bipartition of 퐻, where the contraction vertex 푥 of 퐻 is in 퐴. As 퐶 is nontrivial, then the edge 푒 joins a vertex in 푋− to a vertex in ( 푋) − . Thus, edge 푒 is not incident with a vertex in 퐵. Let 푓 ∈ 퐶. Then, 푥 is an end of 푓 in 퐻. Let 푏 denote the other end of 푓 in 퐻. Assume, to the contrary, that 푓 is not removable in 퐻. By Lemma 8.2, 퐴 has a nonempty proper subset 퐴1 and 퐵 a nonempty proper subset 퐵1 such that | 퐴1 | = |퐵1 | and 푓 = 푏푥 is the only edge of 퐻 that joins a vertex of 퐵1 to a vertex not in 퐴1 . The edge 푒 is not incident with any vertex in 퐵. In particular, 푏 is not an end of 푒. Thus, 푏 is adjacent in 퐺 − 푒 to three or more vertices. Consequently, 푏 is adjacent to two or more vertices of 퐴1 . Thus, | 퐴1 | ≥ 2. As 푒 is not incident with any vertex in 퐵, the set 퐴1 ∪ (퐵1 − 푏) is the shore of a nontrivial tight cut of 퐺, a contradiction. We conclude that 푓 is removable in 퐻, a 퐶-contraction of 퐺 − 푒. This conclusion holds for both 퐶-contractions of 퐺 − 푒. Thus, 푓 is removable in 퐺 − 푒.

A.9 Chapter 9

Exercise 9.1.6 (i) Assume that two cuts in C, say, 퐶 := 휕 ( 푋) and 퐷 := 휕 (푌 ), cross. We now consider the four quadrants 푋 ∩푌 , 푋∩푌 , 푋 ∩푌 and 푋∩푌, which are, respectively, shores of the cuts 퐶푋푌 , 퐶푋푌 , 퐶푋푌 , 퐶푋푌 . See Figure A.11. 푌

푋

푋 ∩ 푌

푋 ∩ 푌

푋 ∩ 푌

푋 ∩ 푌

Fig. A.11 Two crossing cuts in the collection C

Let 퐹 be the set of edges that join a vertex in 푋 ∩ 푌 to a vertex in 푋 ∩ 푌 or join a vertex in 푋 ∩ 푌 to a vertex in 푋 ∩ 푌 . By definition, 푅 푘 = 퐶 ∩ 퐷, hence 푅 푘 = 퐹. Thus, |퐶푋푌 | + |퐶푋푌 | + |퐶푋푌 | + |퐶푋푌 | = 2|퐶| + 2|퐷| − 2|퐹| = 12.

Solutions to Selected Exercises

530

As 퐺, a brick, is 3-connected, we deduce that each of the four cuts is a 3-cut. No edge of 푅 푘 has both ends in any of the four quadrants; hence the subgraph of 퐺 induced by any one of the quadrants is bipartite. As seen in Exercise 5.2.7, each quadrant is a singleton; hence 퐺 has order four and is thus 퐾4 . We conclude that if C is not laminar then 퐺 is 퐾4 . (ii) Suppose that 퐺 is not 퐾4 . By the previous item, C is laminar. Let 퐶 := 휕 ( 푋) and 퐷 := 휕 (푌 ) be two cuts in C, where 푋, 푌 ∈ X. By definition, 푣 푘 ∈ 푋 ∩ 푌 . As 푒 푘 , an edge of 푅 푘 , is in 퐶 ∩ 퐷, the vertex 푤 푘 is in 푋 ∩ 푌 . As 퐶 and 퐷 do not cross, we deduce that either 푋 ∩ 푌 = ∅ or 푋 ∩ 푌 = ∅. Thus, one of 푋 and 푌 is a subset of the other. Moreover, 퐶 and 퐷 are distinct; hence the inclusion is proper. This conclusion holds for any two distinct cuts 퐶 and 퐷 of C. Thus, the shores in X of the cuts in C form a nested collection. (iii) For 푖 = 1, 2, . . . , 푘 − 1, let 퐶푖 := 휕 ( 푋푖 ). Define the graphs 퐻푖 , 푖 = 1, 2, . . . , 푘 as follows: a. let 퐻1 := 퐺 [푋1 ], b. for 푖 = 2, 3, . . . , 푘 − 1, let 퐻푖 := 퐺 [푋푖 − 푋푖−1 ], and c. let 퐻 푘 := 퐺 [푋 푘−1 ]. (See Figure 9.3.) For 푖 = 1, 2, . . . , 푘, the cut 휕 (퐻푖 ) is equal to 푅푖−1 ∪ 푅푖 ; thus, it is a 4-cut. If 퐻푖 were not connected then 휕 (퐻푖 ) would not be a bond; hence 퐺 would not be 3-edge-connected. This is not possible. Thus, the 푘 graphs 퐻푖 are the components of 퐺 − 푅. Each 퐻푖 is a subgraph of 퐺 − 푅 푘 , in turn a bipartite (matching covered) graph. Thus, each 퐻푖 is bipartite. To prove that 퐻푖 is matchable, note that as 푅푖−1 is a removable doubleton of 퐺, the graph 퐺 − 푅푖−1 is a bipartite matching covered graph. Let 푣 be the end in 퐻푖 of an edge of 푅푖 . The vertex 푣 has degree two or more in 퐺 − 푅푖−1 ; hence 푣 is incident with an edge, 푒, which is not in 푅푖 . Let 푀푒 be a perfect matching of 퐺 − 푅푖−1 that contains the edge 푒. Then, 푀푒 ∩ 퐸 (퐻푖 ) is a perfect matching of 퐻푖 . We conclude that 퐻푖 is matchable.

Exercise 9.2.3 The brick shown in Figure A.12 has two disjoint triangles. It is clearly not a prism or a M¨obius ladder or a staircase. It has precisely two removable doubletons.

Fig. A.12 A brick with just two mutually exclusive doubletons

Exercises of Chapter 9

531

Exercise 9.2.4 (i) By Lemma 9.12, the doubletons 푅1 and 푅2 induce an aitch configuration. Adjust notation so that 푒 1 and 푒 2 share a common end, 푢 12 , of degree three. Then, 푓1 and 푓2 also share a common end, 푣 12 , also of degree three, and adjacent to 푢 12 . Let ℎ be the edge 푢 12 푣 12 . See Figure 9.5. (ii) Clearly, 퐶 contains also the edges 푒 2 , 푓1 and 푓2 , Assume, to the contrary, that 퐶 is not hamiltonian. The symmetric difference 푀1 △ 푀2 induces one or more edge-disjoint (푀1 , 푀2 )-alternating cycles. It cannot induce two or more cycles; otherwise one of the cycles, say 퐶 ′ , distinct from 퐶, does not contain any of the four edges of 푅1 ∪ 푅2 ; hence the set 푀1′ := 푀1 △ 퐸 (퐶 ′ ) is a perfect matching of 퐺 that also contains the two edges of 푅1 . This is in contradiction to the uniqueness of 푀1 . Thus, 퐶 is the only cycle induced by 푀1 △ 푀2 . If 퐶 is not hamiltonian then some vertex of 퐺 not in 퐶 is incident with an edge of 푀1 ∩ 푀2 . This is a contradiction to the fact that 푀1 and 푀2 are disjoint. Indeed, 퐶 is hamiltonian. The edge ℎ is a chord of 퐶. (iii) Assume, to the contrary, that 퐶 has a chord ℎ′ , distinct from ℎ, that joins two vertices in 푃. The vertices 푢 12 and 푣 12 both have degree three and are joined by the chord ℎ. Thus, ℎ′ joins internal vertices of 푃. Let 푤 푖 and 푤 푗 denote the ends of ℎ′ , where 푖 < 푗. Let 푃′ denote the subpath of 푃 from 푤 푖 to 푤 푗 . Then, 퐸 (푃′ ) and 푅1 ∪ 푅2 are disjoint. Adjust notation so that the first edge of 푃′ is in 푀1 . As 퐺 − 푅1 is bipartite, it follows that 푤 푖 and 푤 푗 lie in distinct parts of the bipartition of 퐺 − 푅1 . Thus, the last edge of 푃′ also lies in 푀1 ; hence the cycle 퐶 ′ obtained by the addition of the chord ℎ′ to the path 푃′ is 푀1 -alternating. We conclude that 푀1 △ 퐸 (퐶 ′ ) is a perfect matching of 퐺 that includes both edges of 푅1 , in contradiction to the uniqueness of 푀1 . We deduce that no chord of 퐶, except ℎ, has both ends in 푉 (푃). Likewise, ℎ is the only chord of 퐶 that has both ends in 푉 (푄). (iv) A.1.2 Let 푤 푖 and 푥 푗 be internal vertices of 푃 and 푄, respectively and suppose that 푤 푖 and 푥 푗 are adjacent. The edge 푒 푖 of 푃 that precedes 푤 푖 is in 푀1 if and only if the edge 푓 푗 of 푄 that precedes 푥 푗 is in 푀2 . Proof We first recall that 푅1 and 푅2 are distinct equivalence classes of the dependence relation on the edges of 퐺. Consequently, for 푘 ∈ {1, 2}, an 푀 푘 alternating cycle that contains an edge of 푅 푘 contains in fact both edges of 푅 푘 . Let 푃′ be the subpath of 푃 from 푢 12 to 푤 푖 and let 푄 ′ be the subpath of 푄 from 푢 12 to 푥 푗 . Let 퐶 ′ be the cycle 푃′ 푤 푖 푥 푗 푅(푄 ′ ), where 푅(푄 ′ ) is the reverse of 푄 ′ . If both edges 푒 푖 and 푓 푗 are in 푀1 then 퐶 ′ is an 푀1 -alternating cycle that contains only one edge of 푅1 . Likewise, if 푒 푖 and 푓 푗 are both in 푀2 then 퐶 ′ is an 푀2 -alternating cycle that contains only one edge of 푅2 . In both alternatives we derive a contradiction. ƒ We prove this part by induction on 푖. Adopt, as the inductive hypothesis, that for every 푗 such that 1 ≤ 푗 < 푖, the vertices 푤 푗 and 푥 푗 have degree three and are adjacent.

Solutions to Selected Exercises

532

Assume, to the contrary, that either (i) 푤 푖 and 푥푖 are not adjacent or (ii) 푤 푖 has degree four or more. By induction, for each 푟 such that 1 ≤ 푟 < 푖, the vertices 푤 푟 and 푥푟 have degree three and are adjacent. If 푤 푖 and 푥푖 are not adjacent then 푤 푖 is adjacent to some vertex 푥 푗 , 푗 > 푖. Alternatively, if 푤 푖 has degree four or more then it is also adjacent to some vertex 푥 푗 , 푗 > 푖. In both alternatives, 푤 푖 is adjacent to some 푥 푗 , 푗 > 푖. The subgraph 퐺 − 푢 12 − 푣 12 is connected and bipartite. Thus, 푗 ≥ 푖 + 2. Moreover, the vertex 푥 푗 −1 is not adjacent to 푤 푖 . See Figure A.13. 푢12 푓1

푒1

푥1

푤1

푥2

푤2

...

...



푥푖−1

푤푖−1 ...

푥 푗 −1

푤푖 ... 푥 푗

...

...

푤ℓ

푣12 Fig. A.13 Illustration for item (iv)

In fact, 푥 푗 −1 is not adjacent to any vertex of 푃 that precedes 푤 푖 . Thus, 푥 푗 −1 is adjacent to some vertex 푤 ℓ of 푃, where ℓ > 푖, and 푖 and ℓ have distinct parities. Let 푃′ the subpath of 푃 from 푤 푖 to 푤 ℓ and let 푘 ∈ {1, 2} be such that the first edge of 푃′ is in 푀 푘 . Then, the last edge of 푃′ also lies in 푀 푘 , because 푖 and ℓ have distinct parities. The edge that precedes 푤 ℓ in 푃 is in 푀 푘 . Thus, by statement (A.1.2), the edge 푥 푗 −1 푥 푗 is also in 푀 푘 . We deduce that the cycle 퐶 ′′ , obtained by adding the path 푤 ℓ 푥 푗 −1 푥 푗 푤 푖 to 푃′ , is an 푀 푘 -alternating cycle that does not contain any edge in 푅 푘 . Thus, 푀 푘 △ 퐸 (퐶 ′′ ) is a perfect matching of 퐺 that contains both edges in 푅 푘 and is distinct from 푀 푘 . This is a contradiction to the uniqueness of 푀 푘 . The assertion holds, by induction. 5. We conclude that 퐺 is cubic and the paths 푃 and 푄 have the same length. Moreover, each internal vertex 푤 푖 of 푃 is joined to the corresponding internal vertex 푥푖 of 푄 by a chord of 퐶. If 퐺 has only four vertices then it is 퐾4 . If 퐺 has order six or more, then it is a staircase.

Exercises of Chapter 9

533

Exercise 9.2.5 (i) Let 푣 be a vertex of 퐺. As 퐺 is a brick, then 푣 has degree 푑 ≥ 3. Let 휕 (푣) := {푒 푖 : 1 ≤ 푖 ≤ 푑}. For 푖 = 1, 2, . . . , 푑, let 푅푖 be a minimal equivalence class of the dependence relation induced by 푒 푖 . Those 푑 classes are mutually exclusive and removable. If two or them are singletons then 퐺 has two removable edges. We conclude that at most one of those classes is a singleton. Then, at least two or these classes are removable doubletons, say, 푅1 and 푅2 . Moreover, 푅1 and 푅2 are mutually exclusive. (ii) Let 푒 be a removable edge of 퐻푖 which is not in 푀푖 . The graph 퐻푖 − 푒 is then matching covered and 푀푖 is a perfect matching of 퐺 − 푒 that contains both edges of 푅푖 . Thus, the edges of 푅푖 are matchable in 퐺 − 푒; hence 퐺 − 푒 is matching covered. In other words, 푒 is removable in 퐺. This conclusion holds for each edge 푒 which is removable in 퐻푖 but is not in 푀푖 . (iii) Suppose that 푀푖 − 푅푖 contains an edge 푔 := 푣푤 which is not removable in 퐻푖 . The edge 푔 is in 푀푖 − 푅푖 . Thus, 푣 is not incident with any edge in 푅푖 . It follows that 푣 has degree three or more in 퐻푖 . By Lemma 8.6, at most two edges of 휕 (푣) are not removable in 퐻푖 . One of these edges is 푔. As 푣 has degree three or more in 퐻푖 , we deduce that 휕 (푣) − 푔 contains an edge, 푒, which is removable in 퐻푖 . Likewise, 휕 (푤) − 푔 contains an edge, 푓 , which is removable in 퐻푖 . As 퐺 is simple, the edges 푒 and 푓 are distinct. In sum, 푒 and 푓 are distinct removable edges of 퐻푖 neither of which is in 푀푖 . From part (ii) we infer that 푒 and 푓 are distinct removable edges of 퐺, a contradiction. (iv) Suppose that 푀푖 is not unique and let 푁푖 be a perfect matching of 퐺 distinct from 푀푖 that contains the edges of 푅푖 . Let 퐶 be an (푀푖 , 푁푖 )-alternating cycle of 퐺. As 퐺 is simple, the cycle 푄 has length four or more. Thus, 퐶 ∩ 푁푖 contains two edges, 푒 and 푓 . Those two edges are in 푁푖 − 푀푖 . From parts (iii) and (ii), we infer that 푒 and 푓 are both removable in 퐺. (v) Suppose that 푀1 ∩ 푀2 contains an edge, 푒. As 푒 ∈ 푀1 , the edge 푒 is removable in 퐻1 . The graph 퐻1 −푒 contains both edges of 푅2 . Let 푁2 be a perfect matching of 퐻1 − 푒 that contains edge 푒 2 . Clearly, 푁2 is a perfect matching of 퐺; hence it contains both edges of 푅2 . In sum, 푀2 and 푁2 are perfect matchings of 퐺 that contain both edges of 푅2 . But the edge 푒 is in 푀2 − 푁2 ; hence 푀2 and 푁2 are distinct. This is a contradiction to the uniqueness of 푀2 . (vi) By Theorem 9.15, the brick 퐺 is either 퐾4 or a staircase. (vii) Let 푛 denote the order of 퐺. If 푛 = 4 then 퐺 is 퐾4 . We may thus assume that 푛 ≥ 6. Then, 퐺 is a staircase. The graph 퐺 has a hamiltonian cycle, 퐶. As 퐺 is cubic, 퐶 has precisely 푛/2 chords. One of the chords joins the common vertices of both removable doubletons. Two other chords are edges of the two triangles of 퐺. The remaining 푛/2 − 3 chords are removable. Thus, 푛/2 − 3 ≤ 1; hence 푛 ≤ 8. We conclude that 푛 ∈ {6, 8}. If 푛 = 6 then 퐺 = 퐶6 , whereas if 푛 = 8 then 퐺 = H8 .

534

Solutions to Selected Exercises

A.10 Chapter 10

Exercise 10.1.6 Proof (of Corollary 10.7) Let 푣 be a vertex of 퐺. Assume that for 푖 = 1, 2, 3, edge 푣푣 푖 is not removable in 퐺. Then, for 푖 = 1, 2, 3, 푉 has an equipartition (퐵푖 , 퐼푖 ) such that 퐵1 ∩퐵2 = {푣 3 }, 퐵1 = 퐼3 −푣+푣 2 and 퐵2 = 퐼3 −푣+푣 1 . Then, 퐼3 −푣 = 퐵1 ∩퐵2 = {푣 3 }, whence 퐼3 = {푣, 푣 3 }. In that case, as 퐼3 contains half the vertices of 퐺, it follows that ƒ 퐺 contains precisely four vertices. Proof (of Corollary 10.8) Let 푣 be a vertex of 퐺. Suppose that neither 푣푣 1 nor 푣푣 2 is in a removable class of 퐺. Then, in particular, neither 푣푣 1 nor 푣푣 2 is removable in 퐺. By Theorem 10.5, 푣 has precisely three neighbours, 푣 1 , 푣 2 and 푣 3 . For 푖 = 1, 2, let 퐵푖 , 퐼푖 and 퐻푖 be as in the statement of Theorem 10.5. For 푖 = 1, 2, let 푏 푖 denote the number of edges of 퐺 that have both ends in 퐵푖 . As |퐵푖 | = |퐼푖 | and since 푣푣 푖 has both ends in 퐼푖 , it follows that 푏 푖 > 0; otherwise 푣푣 푖 would not be matchable in 퐻푖 . Moreover, all edges with both ends in 퐵1 , and all edges with both ends in 퐵2 , are incident with 푣 3 . By hypothesis, the degree of 푣 3 is three or four. As 푣 3 is adjacent to 푣, which is a vertex of 퐼1 ∩ 퐼2 , it follows that 푏 1 + 푏 2 ≤ 3; it follows that at least one of 푏 1 and 푏 2 is equal to one. Adjust notation so that 푏 1 = 1. Let 푓1 denote the only edge of 퐺 having both ends in 퐵1 . Then, 퐻1 = 퐺 − 푣푣 1 − 푓1 . Neither 푣푣 1 is matchable in 퐺 − 푓1 nor 푓1 is matchable in 퐺 − 푣푣 1 . We deduce that {푣푣 1 , 푓1 } is a removable doubleton of 퐺. This contradicts the assumption that neither 푣푣 1 nor 푣푣 2 lies in a removable class of 퐺. ƒ

Exercise 10.2.2 Let 퐺 be a solid matching covered graph and let 푅 be a set of edges of 퐺 such that 퐺 − 푅 is matching covered. We prove that 퐺 − 푅 is solid by induction on |퐸 |. If 푅 is empty then there is nothing to be proved. We may thus assume that 푅 is nonempty. If 퐺 is bipartite then 퐺 − 푅 is also bipartite, hence solid. Suppose that 퐺 is a brick. As 퐺 − 푅 is matching covered, it follows that 푅 includes a removable class, 푆, of 퐺. If 푆 is a singleton then 퐺 − 푆 is solid, by Theorem 10.14. Alternatively, if 푆 is a doubleton then 퐺 − 푆 is bipartite, hence solid. In both alternatives, 퐺 − 푆 is solid. By induction, (퐺 − 푆) − (푅 − 푆) is solid. We may thus assume that 퐺 has a nontrivial tight cut, 퐶 := 휕 ( 푋). Let 퐺 1 and 퐺 2 be the two 퐶-contractions of 퐺. The cut 퐶 − 푅 is tight in 퐺 − 푅. By Theorem 7.1, each brick of 퐺 is solid; hence each brick of 퐺 푖 is solid, for 푖 = 1, 2. Consequently, 퐺 1 and 퐺 2 are both solid. The graphs 퐺 푖 − 푅 are matching covered. By induction, 퐺 1 − 푅 and 퐺 2 − 푅 are both solid. Thus, all the bricks of 퐺 1 − 푅 and all the bricks of 퐺 2 − 푅 are solid. Consequently, all the bricks of 퐺 − 푅 are solid; hence 퐺 − 푅 is solid.

Exercises of Chapter 10

535

Exercise 10.2.3 Let 푒 be a removable edge of a solid matching covered graph 퐺. We prove that 푏(퐺 − 푒) = 푏(퐺) by induction on |퐸 |. If 퐺 is bipartite then 퐺 − 푒 is also bipartite and 푏(퐺 −푒) = 0 = 푏(퐺), hence the assertion. If 퐺 is a brick then, by Theorem 10.16, 푏(퐺 − 푒) = 1 = 푏(퐺). We may thus assume that 퐺 has a nontrivial tight cut, 퐶 := 휕 ( 푋). Let 퐺 1 and 퐺 2 be the two 퐶-contractions of 퐺. By Theorem 7.1, each brick of 퐺 is solid; hence each brick of 퐺 푖 is solid, for 푖 = 1, 2. Consequently, 퐺 1 and 퐺 2 are both solid. By induction, 푏(퐺 푖 − 푒) = 푏(퐺 푖 ), for 푖 = 1, 2. Clearly, 퐶 − 푒 is tight in 퐺 − 푒. Thus, 푏(퐺 − 푒) = 푏(퐺 1 − 푒) + 푏(퐺 2 − 푒) = 푏(퐺 1 ) + 푏(퐺 2 ) = 푏(퐺). In all cases considered, 푒 is 푏-invariant in 퐺. This conclusion holds for each removable edge 푒 of 퐺. By Theorem 7.1, each brick of a solid matching covered graph is solid. The Petersen graph is not solid; thus, no brick of a solid matching covered graph is a Petersen brick. Hence, 푝(퐺) = 0. Moreover, by Theorem 10.15, 퐺 − 푒 is also solid. Thus, 푝(퐺 − 푒) = 0. We conclude that (푏 + 푝) (퐺 − 푒) = 푏(퐺 − 푒) = 푏(퐺) = (푏 + 푝) (퐺). Exercise 10.2.4 Let 퐶 := 휕 ( 푋) ∈ C and assume, to the contrary, that 퐶 is not separating in 퐺. Then, a 퐶-contraction of 퐺, say, 퐺 1 := 퐺/( 푋 → 푥), is not matching covered. By definition of C, the cut 퐶 − 푅 is separating in 퐺 − 푅. Hence, 퐺 1 − 푅 is matching covered. Thus, only edges of 푅 are not matchable in 퐺 1 . Let 푒 be an edge of 푅 which is not matchable in 퐺 1 and let 퐵 denote a maximal barrier of 퐺 1 that contains both ends of 푒 in 퐺 1 . As 푒 is matchable in 퐺, the set 퐵 is not a barrier of 퐺. Thus, the contraction vertex 푥 of 퐺 1 is in 퐵. Let 퐸 퐵 denote the set of edges of 퐺 1 that contain both ends in 퐵. Adopt the notation used in the proof of Lemma 10.13. As 퐺 1 − 푅 is matching covered, 퐸 퐵 ⊆ 푅. Let 푀 be a perfect matching of 퐺. As in equation 10.2, we have that Õ |푀 ∩ 퐶퐾 | = |푀 ∩ 퐶| + (|O| − 1) − 2|푀 ∩ 퐸 퐵 |. (A.18) 퐾 ∈ O

Let 퐽 ∈ O. Then, |푀 ∩ 퐶 퐽 | +

Õ

퐾 ∈ O− 퐽

|푀 ∩ 퐶퐾 | = |O| − 1 ≤

Õ

|푀 ∩ 퐶퐾 |,

Õ

퐾 ∈ O

퐾 ∈ O− 퐽

|푀 ∩ 퐶퐾 |.

Addition of equations (A.18)–(A.20) and simplification yields

(A.19) (A.20)

Solutions to Selected Exercises

536

|푀 ∩ 퐶 퐽 | ≤ |푀 ∩ 퐶| − 2|푀 ∩ 퐸 퐵 |. We conclude that |푀 ∩ 퐶 퐽 | ≤ |푀 ∩ 퐶|, with equality only if 푀 and 퐸 퐵 are disjoint. We deduce that 퐶 퐽 ≺ 퐶. Let us now prove that 퐶 퐽 − 푅 is separating in 퐺 − 푅. Let 푓 be any edge of 퐺 − 푅. As 퐶 − 푅 is separating in 퐺 − 푅, there exists a perfect matching 푀 푓 of 퐺 − 푅 that contains edge 푓 and just one edge in 퐶 − 푅. As 퐶 퐽 ≺ 퐶, we deduce that 푀 푓 contains just one edge in 퐶 퐽 . In sum, every edge of 퐺 − 푅 is in a perfect matching of 퐺 − 푅 that contains just one edge in 퐶 퐽 − 푅. We conclude that 퐶 퐽 − 푅 is separating in 퐺 − 푅. So far we have proved that for every component 퐽 in O, 퐶 퐽 ≺ 퐶 and 퐶 퐽 − 푅 is separating in 퐺 − 푅. We now get a contradiction to the minimality of 퐶, by showing that 휆퐺−푅 (퐶 퐽 − 푅) ≤ 휆퐺−푅 (퐶 − 푅) for some component 퐽 in O. If 퐶 − 푅 is tight in 퐺 − 푅 then 퐶 퐽 − 푅 is tight in 퐺 − 푅; in that case, 휆퐺−푅 (퐶 퐽 − 푅) = ∞ = 휆퐺−푅 (퐶 − 푅). We may thus assume that 퐶 − 푅 is not tight in 퐺 − 푅. Let 푀 be a perfect matching of 퐺 − 푅 that contains precisely 휆퐺−푅 (퐶 − 푅) > 1 edges in 퐶 − 푅. As 퐸 퐵 ⊆ 푅 and since 푀 is a perfect matching of 퐺 − 푅, from (A.18) we infer that O contains a component 퐽 that satisfies the inequalities 1 < |푀 ∩ 퐶 퐽 | ≤ |푀 ∩ 퐶| = 휆퐺−푅 (퐶 − 푅). We deduce that 휆퐺−푅 (퐶 퐽 − 푅) ≤ 휆퐺−푅 (퐶 − 푅). We proved that O has a component 퐽 such that 퐶 퐽 ≺ 퐶, 퐶 퐽 − 푅 is separating in 퐺 − 푅 and 휆퐺−푅 (퐶 퐽 − 푅) ≤ 휆퐺−푅 (퐶 − 푅). This is a contradiction to the minimality of 퐶 in the relation ≺. Exercise 10.2.5 (i) Suppose that 푅 is a singleton, {푒}. By hypothesis, 퐺 − 푅 − 푆 is matching covered; hence the edge 푒 is removable in 퐺 − 푆. The graph 퐺 is solid and 푒 is removable in 퐺; hence, by Theorem 10.17. 푏(퐺 − 푒) = 푏(퐺). By Theorems 9.8, and 9.9, 푏(퐺 − 푆) ≤ 푏(퐺 − 푒 − 푆) = 푏(퐺 − 푒) − 1 = 푏(퐺) − 1. Thus, equality holds. By Corollary 10.18, the set 푆 is a removable doubleton of 퐺. (ii) Consider next the case in which 푅 is a doubleton, {푒 1 , 푒 2 }. The graph 퐺 − 푅 − 푆 is matching covered by hypothesis. Thus, either 푅 is a removable doubleton of 퐺 − 푆 or one of the edges, say 푒 1 , is removable in 퐺 − 푆 and the other edge of 푅 is removable in 퐺 − 푆 − 푒 1 . If 푅 is a removable doubleton in 퐺 − 푆 then 푏(퐺 − 푆) = 푏(퐺 − 푅 − 푆) + 1, by Theorem 9.9. Alternatively, if 푒 1 is removable in 퐺 − 푆 and 푒 2 is removable in 퐺 − 푒 1 − 푆 then 푏(퐺 − 푆) < 푏(퐺 − 푅 − 푆) + 1, by Theorem 9.8. In both alternatives, 푏(퐺 −푆) ≤ 푏(퐺 − 푅 −푆) +1. By Theorems 9.8 and 9.9, 푏(퐺 − 푆) ≤ 푏(퐺 − 푅 − 푆) + 1 = 푏(퐺 − 푅) = 푏(퐺) − 1. Thus, 푏(퐺 − 푆) = 푏(퐺) − 1 and 푆 is a removable doubleton of 퐺.

Exercises of Chapter 11

537

Exercise 10.2.6 (a) If 퐺−푅 is solid then 휆(퐺) ≤ ∞ = 휆(퐺−푅) and the inequality holds immediately. We may thus assume that 퐺 − 푅 is not solid. Let 퐷 be a cut of 퐺 such that 퐷 − 푅 is separating in 퐺 − 푅 and 휆퐺−푅 (퐷 − 푅) = 휆(퐺 − 푅). Using the result in Exercise 10.2.4, we infer that 퐺 has a separating cut 퐶 such that 퐶 − 푅 is separating in 퐺 − 푅 and 휆퐺−푅 (퐶 − 푅) ≤ 휆퐺−푅 (퐷 − 푅). By definition of 퐷, 휆퐺−푅 (퐷 − 푅) = 휆(퐺 − 푅); hence 휆퐺−푅 (퐶 − 푅) = 휆(퐺 − 푅). Let 푀 be a perfect matching of 퐺 − 푅 such that |푀 ∩ (퐶 − 푅)| = 휆(퐺 − 푅). The set 푀 is a perfect matching of 퐺 and 퐶 is separating in 퐺; hence 휆(퐺) ≤ |푀 ∩ 퐶| = |푀 ∩ (퐶 − 푅)| = 휆(퐺 − 푅), hence the inequality. (b) If 퐺 is solid then 휆(퐺) = ∞. By the first part, 휆(퐺 − 푅) ≥ 휆(퐺). Thus, 휆(퐺 − 푅) = ∞; hence 퐺 − 푅 is solid.

A.11 Chapter 11

Exercise 11.5.7 Let us prove the two assertions by induction on 푚. In both (i) and (ii), suppose b of 퐺. Let 푚 b, b 푛 that the graph 퐺 has vertices of degree two. Consider the retract 퐺 b Clearly, and b 푏 denote, respectively, the number of edges, vertices and bricks of 퐺. b is neither 푚 b−b 푛 = 푚 − 푛 and b 푏 = 푏. If 퐺 is neither 퐾2 nor an even cycle then 퐺 b Moreover, 퐾2 nor an even cycle. If 퐺 is not a bisubdivision of 퐾4 then neither is 퐺. bicontractions preserve the solidity of 퐺. In both assertions, the result follows by induction and Theorem 11.21. We may thus assume that the minimum degree 훿 of vertices of 퐺 is three or more, in both (i) and (ii). (i): We assume that 훿 ≥ 3. Case 1 Graph 퐺 is a brace. If 퐺 has order two then, as 퐺 is not 퐾2 , all its edges are removable. Likewise, if 퐺 has order six or more then all its edges are removable. The assertion thus holds if 푛 ≠ 4. Suppose now that 푛 = 4. As 퐺 is free of vertices of degree two, it has a perfect matching whose edges are both multiple edges of 퐺. Thus, at most two edges of 퐺 are not removable; hence the number of removable edges of 퐺 is at least 푚−2 = 푚−푛+2. Case 2 Graph 퐺 is a brick. If 퐺 is 퐾4 then it has three removable doubletons; hence the assertion holds, as 푚 − 푛 + 2 − 푏 = 3. We may thus assume that 퐺 is not 퐾4 . The only bricks that do not have removable edges are 퐾4 and 퐶6 . Moreover, 퐶6 is not solid. Thus, 퐺 has a removable edge, 푒. By Theorem 10.14, the graph 퐺 − 푒 is solid. As 퐺 is solid, we infer from Theorem 10.16 that the edge 푒 is 푏-invariant; hence the graph 퐺 − 푒 is a

538

Solutions to Selected Exercises

near-brick. By induction, 퐺 − 푒 has at least 푚 − 푛 edge-disjoint removable ears. Let 푃 := 푣 0 푣 1 . . . 푣 ℓ , ℓ odd, be a constituent path of a removable ear of 퐺 − 푒. If ℓ ≥ 3 then, as 훿 ≥ 3, it follows that 푒 joins the vertices 푣 1 and 푣 2 . This is not possible, because in that case the set {푣 0 , 푣 1 , 푣 2 } is the shore of a nontrivial tight cut of 퐺. Thus, 퐺 − 푒 has at least 푚 − 푛 edge-disjoint removable classes. Each removable class of 퐺 − 푒 includes a removable class of 퐺, by Theorem 10.23. Consequently, in addition to edge 푒, 퐺 has at least 푚 − 푛 edge-disjoint removable classes, none of which contains the edge 푒. The assertion holds. Case 3 Graph 퐺 has a nontrivial tight cut 퐶 := 휕 ( 푋). Let 퐺 1 := 퐺/( 푋 → 푥) and 퐺 2 := 퐺/( 푋 → 푥) be the two 퐶-contractions of 퐺. Every brick of 퐺 is solid; hence every brick of 퐺 1 is also solid. Thus, 퐺 1 is solid. Likewise, 퐺 2 is solid. For 푖 = 1, 2, let 푚 푖 , 푛푖 and 푏 푖 denote respectively the number of edges, of vertices and of bricks of 퐺 푖 . By induction, 퐺 푖 has at least 푟 푖 := 푚 푖 − 푛푖 + 2 − 푏 푖 edge-disjoint removable ears. Each vertex of 퐺 1 distinct from the contraction vertex 푥 has degree three or more in 퐺 1 . Thus, the 푟 1 edge-disjoint removable ears of 퐺 1 are in fact removable classes of 퐺 1 . Likewise, the 푟 2 edge-disjoint removable ears of 퐺 2 are removable classes of 퐺 2 . Each removable class of 퐺 푖 disjoint with 퐶 is a removable class of 퐺. For each edge 푒 of 퐶 and for 푖 = 1, 2, let  0, if no removable class of 퐺 푖 contains the edge 푒 푠푖 (푒) := 1, otherwise. A removable class 푅1 of 퐺 1 and a removable class 푅2 of 퐺 2 may share an edge in 퐶. In that case, 퐺−(푅1 ∪푅2 ) is matching covered; hence 푅1∪푅2 includes a removable class of 퐺. Alternatively, 푠1 (푒) + 푠2 (푒) = 1 implies that the removable class of 퐺 1 or of 퐺 2 that contains the edge 푒 is not removable Í in 퐺. In sum, the number 푟 of edge-disjoint removable classes of 퐺 is at least 푟 1 +푟 2 − 푒∈퐶 min{1, (푠1 (푒) +푠2 (푒)} ≥ 푟 1 +푟 2 −|퐶|. Thus, 푟 ≥ 푟 1 + 푟 2 − |퐶| = (푚 1 − 푛1 + 2 − 푏 1 ) + (푚 2 − 푛2 + 2 − 푏 2 ) − |퐶| = (푚 1 + 푚 2 − |퐶|) − (푛1 + 푛2 − 2) + 2 − (푏 1 + 푏 2 ) = 푚 − 푛 + 2 − 푏. In all cases considered, we deduce that 퐺 has at least 푚 − 푛 + 2 − 푏 edge-disjoint removable ears. In the particular case of bipartite graphs, each removable ear is a single ear and 푏 = 0; hence in that case 퐺 has at least 푚 − 푛 + 2 edge-disjoint removable single ears. (ii): We assume that 훿 ≥ 3 and also that the near-brick 퐺 is not 퐾4 . Case 1 The graph 퐺 is a brick. As in the previous analysis, 퐺 has a removable edge, 푒 and the graph 퐺 − 푒 is a solid near-brick. If 퐺 − 푒 is not a bisubdivision of 퐾4 then, by induction, 퐺 − 푒 has at least 푚 − 푛 − 2 edge-disjoint removable single ears. As in the analysis of (i), the 푚 − 푛 − 2 removable single ears of 퐺 − 푒 are in fact removable edges of 퐺 − 푒. Each

Exercises of Chapter 12

539

removable edge of 퐺 − 푒 is a removable edge of 퐺, by Theorem 10.23. Consequently, in addition to edge 푒, 퐺 has at least 푚 − 푛 − 2 removable edges. The assertion holds. Alternatively, the graph 퐺 − 푒 is a bisubdivision of 퐾4 . As 퐺 is a brick, the edge 푒 is a multiple edge of 퐺 and 퐺 − 푒 is 퐾4 . Thus, 퐺 has two removable edges and 푚 − 푛 − 1 = 7 − 4 − 1 = 2. The assertion holds. Case 2 The graph 퐺 has a nontrivial tight cut 퐶 := 휕 ( 푋). As in the analysis of (i), define, for 푖 = 1, 2, 퐺 푖 , 푚 푖 and 푛푖 . One of 퐺 1 and 퐺 2 is bipartite, the other is a near-brick. Adjust notation so that 퐺 1 is bipartite. As in the analysis of (i), the removable ears of 퐺 푖 are in fact removable classes of 퐺 푖 . By part (i), 퐺 1 has at least 푟 1 = 푚 1 − 푛1 + 2 removable edges. If 퐺 2 is not a bisubdivision of 퐾4 then, by induction, 퐺 2 has at least 푟 2 = 푚 2 − 푛2 − 1 removable edges. In that case, as in the analysis of (i), the number of removable edges of 퐺 is at least 푟 1 + 푟 2 − |퐶| = 푚 − 푛 − 1. Finally, suppose that 퐺 2 is a bisubdivision of 퐾4 . As every vertex of 퐺 2 distinct from its contraction vertex 푥 has degree three or more, it follows that 퐺 2 is 퐾4 . In that case, 푟 2 = 0 and 푟 1 + 푟 2 − |퐶| = 푚 1 − 푛1 + 2 − 3 = 푚 1 − 푛1 − 1. But 푚 − 푛 = 푚 1 + 3 − (푛1 + 3) = 푚 1 − 푛1 . The assertion holds.

A.12 Chapter 12

Exercise 12.2.7 (i) Let 퐻 be the retract of 퐺 − 푒 and let 퐽 be the underlying simple graph of 퐻. Let 푣 0 and 푤 0 be the two ends of 푒. We have assumed that 퐺 is a simple brace of order ten or more. If the minimum degree 훿(퐽) of the vertices of 퐽 is three or more, then, for every nontrivial subset 푋 of 푉 such that 휕퐽 ( 푋) is tight in 퐺, a vertex of 푋− has degree three or more; hence | 푋 | ≥ 5. It then follows that 퐺 − 푒 has a brace of order six or more. Assume thus, to the contrary, that 퐽 has a vertex, 푥, of degree two. Necessarily, that vertex is a contraction vertex of 퐻. Adjust notation so that 푥 is the result of the contraction of the only two neighbours 푣 1 and 푣 2 of 푣 0 in 퐺 − 푒. For 푖 = 1, 2, let 푁 (푣 푖 ) denote the set of neighbours of 푣 푖 in 퐺 and let 퐵 := 푁 (푣 1 ) ∪ 푁 (푣 2 ). The vertices 푣 1 and 푣 2 are isolated in 퐺 − 퐵. If |퐵| ≤ 3 then 퐵 is a barrier of 퐺, and we derive a contradiction, as 퐺 is a brace having more than six vertices. See Figures A.14(a) and A.14(b). Another possibility for 푥 to have degree two in 퐽 is the fact that 푤 0 also has degree three in 퐺, and 퐵 consists of four vertices, two of which are the only two neighbours 푤 1 and 푤 2 of 푤 0 in 퐺 − 푒. Again, 퐵 is a barrier of 퐺, because in this case the vertices 푣 1 , 푣 2 and 푤 0 are isolated in 퐺 − 퐵. We have a contradiction, as 퐺 is a brace of order ten or more.

Solutions to Selected Exercises

540 푣0

푣0

푣0

푒

replacements 푣1

푣2

푣1

푣2

푣1

푣2

푒 푤2

푤1

(a)

푒 푤2

푤1

푤0

푤0

(b)

(c)

Fig. A.14 The possible causes for 훿 ( 퐽 ) = 2

(ii) Every vertex of a brace of order six or more has degree three or more. Thus, 퐾3,3 is the only simple brace of order six. Let 퐺 := 퐺 [ 퐴, 퐵] be a simple cubic graph of order eight. No two vertices 푣 and 푤 of 퐴 have the same set of three neighbours; otherwise the vertex of 퐵 not adjacent to either 푣 or 푤 would have degree two, a contradiction. It follows that each vertex 푣 of 퐴 corresponds to one of the four sets of three vertices of 퐵. Consequently, the cube is the only simple cubic graph of order eight. Let us now consider any simple graph 퐺 := 퐺 [ 퐴, 퐵] of order eight and minimum degree three or more. Let 푀 be a maximal matching of 퐺 joining vertices of degree four of 퐺. It is easy to see that the subgraph of 퐺 induced by 푀 consists of all the vertices of degree four. Thus, 퐺 − 푀 is the cube. Addition of any edge to a brace yields another brace. Consequently, 퐺 is a brace. Moreover, the cube is a spanning subgraph of 퐺. (iii) Let us now prove the result by induction on |퐸 |. If 퐺 has a pair of multiple edges then remove an edge from that pair and the assertion holds. We may thus assume that 퐺 is simple. If 퐺 has only six vertices then certainly it is 퐾3,3 . Alternatively, if 퐺 has order eight then it has the cube as a spanning subgraph; thus, the cube is a conformal subgraph of 퐺. Let us assume then that 퐺 has order ten or more. Let 푒 be an edge of 퐺. By the first part, 퐺 − 푒 has a brace, 퐽, of order six or more. By induction, 퐽 has a conformal minor 퐿 which is either 퐾3,3 or the cube. Those two graphs are cubic; thus, 퐿 is a matching minor of 퐽, by Corollary 12.11. By Exercise 12.1.8, 퐿 is a matching minor of 퐺. Hence, again by Corollary 12.11, 퐿 is a conformal minor of 퐺.

Exercises of Chapter 14

541

A.13 Chapter 13

Exercise 13.2.2 (i) The 퐶-contraction 퐺/ 푋 is the wheel 푊5 , whereas the 퐶-contraction 퐺/푋 is 푊5 + 푒, a brick. Thus, 퐶 is separating. The set {푢 1 푣 1 , 푢 3 푣 3 , 푢 4 푣 4 , 푣 2 푣 5 , 푒} is a perfect matching of 퐺 that contains precisely three edges in 퐶. Thus, 휆(퐶) = 3. (ii) The graph 퐺 − 푒 is P, the Petersen graph, hence (푏 + 푝) (퐺 − 푒) = 2. Figure A.15 depicts the retract 퐻 of 퐺 − 푓1 − 푓2 , where 푥 is the contraction vertex of 퐻. 푢3

푢2

푢5

푢1

푥

푢4

푣4

푣1

Fig. A.15 The retract of 퐺 − 푓1 − 푓2

The graph 퐻 is a brick, it is H8 , the bicorn. The retract of 퐺 − 푓1 is the graph 퐻 + 푓2 , hence also a brick. Likewise, the retract of 퐺 − 푓2 is 퐻 + 푓1 , a brick. Moreover, 퐻 + 푓1 and 퐻 + 푓2 have only eight vertices, hence they are not Petersen bricks. We deduce that (푏 + 푝) (퐺 − 푓1 ) = 1 = (푏 + 푝) (퐺 − 푓2 ). (iii) There are eight (푏 + 푝)-invariant edges of 퐺: 푢 1 푢 2 , 푢 1 푢 5 , 푢 2 푢 3 , 푢 4 푢 5 , 푢 2 푣 2 , 푢 5 푣 5 , 푣 1 푣 3 , and 푣 1 푣 4 . The edge 푢 1 푣 1 is nonremovable and the edge 푒 is 푏-invariant, but not (푏 + 푝)-invariant. There are six edges which are not 푏-invariant : 푢 3 푣 3 , 푢 4 푣 4 , 푢 3 푢 4 , 푣 2 푣 4 , 푣 3 푣 5 and 푣 2 푣 5 .

A.14 Chapter 14

Exercise 14.1.6

(i): Firstly, suppose that 퐺 has a nontrivial separating cut. Then, by Corollary 14.4, 퐺 has a robust cut, 퐶 := 휕 ( 푋). By Corollary 14.5, 푋 has a subset 푋 ′ and 푋 has a subset 푋 ′′ such that the graph 퐻 := 퐺/( 푋 ′ → 푥 ′ )/( 푋 ′′ → 푥 ′′ ) is bipartite and matching covered, with the contraction vertices 푥 ′ and 푥 ′′ in distinct parts of the bipartition of 퐻. Furthermore, the graphs 퐺/푋 ′ and 퐺/푋 ′′ are both bricks.

Solutions to Selected Exercises

542

The sets 푋 ′ and 푋 ′′ are both critical subsets of 푉. The graph 퐺 − ( 푋 ′ ∪ 푋 ′′ ) is equal to the graph 퐻 − 푥 ′ − 푥 ′′ , which is matchable, by Corollary 3.6. Conversely, suppose that 퐺 is a brick and it has disjoint critical sets 푋 and 푌 such that 퐺 − ( 푋 ∪ 푌 ) has a perfect matching, 푀. For each vertex 푣 of 푋, let 푀 (푣) denote a perfect matching of 퐺 [푋] − 푣. Likewise, for each vertex 푣 of 푌 , let 푁 (푣) denote a perfect matching of 퐺 [푌 ] − 푣. Let x :=

Õ 1 1 Õ 푁 (푣) 휒 푀 (푣) + 휒 + 휒 푀 . | 푋 | − 1 푣 ∈푋 |푌 | − 1 푣 ∈푌

Note that x is nonnegative and 1-regular. Moreover, 푋 is odd and x(휕 ( 푋)) = 0. By Theorem 6.1, x ∉ Poly(퐺). By Theorem 7.5, 퐺 is not solid.

(ii): Suppose that 퐺 has two disjoint odd cycles 푄 1 and 푄 2 such that 퐺 ′ := 퐺 − (푉 (푄 1 ) ∪ 푉 (푄 1 )) is matchable. The sets 푉 (푄 1 ) and 푉 (푄 2 ) are both critical and disjoint. Moreover, 퐺 ′ is matchable. By the first part, 퐺 is nonsolid. Conversely, suppose that 퐺 ′ is a nonsolid brick. By the first part, 퐺 has two disjoint critical sets, 푋 and 푌 , such that 퐻 := 퐺 − ( 푋 ∪ 푌 ) has a perfect matching, 푀 퐻 . The graph 퐺 [푋] is critical; thus it has an ear decomposition 퐺 1 , 퐺 2 , . . . , 퐺 푟 , 푟 ≥ 2, such that each 퐺 푖 is a conformal critical subgraph of 퐺 [푋]. In particular, 퐺 2 is a cycle, a conformal subgraph of 퐺 [푋] which is a cycle. Let 푋 ′ := 푉 (퐺 2 ). The set 푋 ′ is critical in 퐺 [푋]; let 푀푋 denote a perfect matching of 퐺 [푋] − 푋 ′ . Likewise, 퐺 [푌 ] has a critical subgraph which is an odd cycle. Let 푌 ′ be the set of vertices of the odd cycle and let 푀푌 denote a perfect matching of 퐺 [푌 ] − 푌 ′ . The sets 푋 ′ and 푌 ′ are both critical and 푀 퐻 ∪ 푀푋 ∪ 푀푌 is a perfect matching of 퐺 − ( 푋 ′ ∪ 푌 ′ ). Exercise 14.3.1 If 퐶 is robust then certainly it is separating. To prove the converse, suppose that 퐶 is separating. Let 퐺 1 and 퐺 2 denote the two 퐶-contractions of 퐺. As 퐶 is separating, the graphs 퐺 1 and 퐺 2 are matching covered. Moreover, as 퐶 is not tight in 퐺, the graphs 퐺 1 and 퐺 2 are not bipartite. By the monotonicity of function 푏, 1 ≤ 푏(퐺 1 ) ≤ 푏(퐺 1 −푒). Likewise, 1 ≤ 푏(퐺 2 ) ≤ 푏(퐺 2 −푒). As 푒 is quasi 푏-invariant and 퐶 − 푒 is tight in 퐺 − 푒, it follows that 푏(퐺 1 − 푒) + 푏(퐺 2 − 푒) = 푏(퐺 − 푒) = 2. Consequently, the graphs 퐺 1 and 퐺 2 are near-bricks. In sum, both 퐶-contractions of 퐺 are near-bricks. It follows that 퐶 is robust.

A.15 Chapter 15

Exercise 15.2.6 Let 퐶 be a tight cut of 퐺 − 푒. As 푒 is 푏-invariant, one of the shores of 퐶 in 퐺 − 푒 is bipartite. By Lemma 8.19(i), every edge of 퐶 is removable in the bipartite 퐶-

Exercises of Chapter 16

543

contraction of 퐺 − 푒. This implies that every edge of the brick 퐻 incident with a contraction vertex of 퐻 is in a tight cut 퐶 of 퐺 − 푒 and is removable in the bipartite 퐶-contraction of 퐺 − 푒. It follows that every removable class of 퐻 is also removable in 퐺 − 푒. Suppose that 퐺 is regular and 푅 is a removable doubleton of 퐻. By the previous part, 푅 is a removable doubleton of 퐺 − 푒. As 푒 is 푏-invariant, the graph 퐺 − 푒 is a near-brick. As 푅 is a removable doubleton of 퐺 − 푒, the graph 퐺 − 푒 − 푅 is bipartite and matching covered. Let ( 퐴, 퐵) be the bipartition of 퐺 − 푒 − 푅. One of the edges of 푅 has both ends in 퐴, the other has both ends in 퐵. As 퐺 is regular, a simple counting argument then shows that 푒 has one end in 퐴 an one end in 퐵. By Corollary 3.7, the graph 퐺 − 푅 is also bipartite and matching covered. We conclude that 푅 is a removable doubleton of 퐺.

A.16 Chapter 16

Exercise 16.2.6 Let us assume that 퐺 has a perfect matching, 푀, which contains the edges of 푅 and is disjoint with 푆. The set 푅 is a removable class of 퐺 and the set 푆 is a removable class of 퐺 − 푅; hence 퐺 − 푅 − 푆 is matching covered. That is, every edge of 퐺 − 푅 − 푆 is matchable and 퐺 − 푅 − 푆 is connected. Clearly, 퐺 − 푆 is thus connected. Moreover, every edge of 퐺 − 푅 − 푆 is matchable in 퐺 − 푆 and 푀 is a perfect matching of 퐺 − 푆 that contains the edges of 푅; hence each edge of 퐺 − 푆 is matchable in 퐺 − 푆. We conclude that 퐺 − 푆 is matching covered. Let 퐻 be a matching covered graph and let 푇 ⊂ 퐸 (퐻) such that 퐺 − 푇 is matching covered and 푇 is either a singleton or a doubleton. If 푇 is a singleton then the monotonicity of function 푏 implies that 푏(퐻 − 푇) ≥ 푏(퐻), with equality if and only if 푇 is an optimal removable class of 퐺. If 푇 is a doubleton, the monotonicity of function 푏 implies that 푏(퐺 − 푇) ≥ 푏(퐺) − 1, with equality if and only if 푇 is a removable doubleton of 퐺. In sum, 푏(퐻 − 푇) ≥ 푏(퐻) − (|푇 | − 1),

(A.21)

with equality if and only if 푇 is an optimal removable class of 퐻. We have seen that 퐺 − 푆 is matching covered and 푆 is either a singleton or a doubleton. From equation (A.21) we infer that 푏(퐺 − 푆) ≥ 푏(퐺) − (|푆| − 1),

(A.22)

with equality only if 푆 is an optimal removable class of 퐺. Likewise, 푏(퐺 − 푅 − 푆) ≥ 푏(퐺 − 푆) − (|푅| − 1),

(A.23)

Solutions to Selected Exercises

544

with equality only if 푅 is an optimal removable class of 퐺 − 푆. From equations (A.23) and (A.22) we deduce that 푏(퐺 − 푅 − 푆) ≥ 푏(퐺) − (|푅| − 1) − (|푆| − 1),

(A.24)

with equality only if 푆 is an optimal removable class of 퐺 and 푅 is an optimal removable class of 퐺 − 푆. As 푅 is an optimal removable class of 퐺 and 푆 is an optimal removable class of 퐺 − 푅, from equation (A.21) we infer that 푏(퐺 − 푅 − 푆) = 푏(퐺 − 푅) − (|푆| − 1) = 푏(퐺) − (|푅| − 1) − (|푆| − 1). Thus, equality holds in equation (A.24). We conclude that 푆 is an optimal removable class of 퐺 and 푅 is an optimal removable class of 퐺 − 푆. Exercise 16.2.8 Algorithm A.2 is a description of a polynomial-time algorithm to determine an optimal ear decomposition of a matching covered graph and the corresponding sequence of optimal removable ears. Algorithm A.2 Given a matching covered graph 퐺, determine an optimal ear decomposition of 퐺 and the corresponding sequence of optimal removable ears. If 퐺 = 퐾2 then just return the sequence (퐺) as the ear decomposition and the empty sequence of removable ears. If 퐺 ≠ 퐾2 , obtain an optimal removable ear 푅 of 퐺 using Algorithm A.3. Recursively determine an optimal ear decomposition G of 퐺 − 푅 and the corresponding sequence R of optimal removable ears. Add 퐺 as the last element of the sequence G and 푅 as the last element of the sequence R. Algorithm A.3 Given a matching covered graph 퐺 distinct from 퐾2 , determine an optimal removable ear of 퐺. First determine the retract 퐻 of 퐺. This is obtained by contracting the set of vertices of each even lobe and by replacing each odd lobe by an edge. See Theorem 11.21. Then, determine the set 푅1 of removable edges of 퐻. To do this, for each edge 푒 of 퐻, determine whether 퐻 − 푒 is matching covered (See Exercise 2.2.2). Then, for each pair {푒 1 , 푒 2 } of edges of 퐻 not in 푅1 , determine whether 퐺 − 푒 1 − 푒 2 is matching covered. If a removable doubleton {푒 1 , 푒 2 } is found, the odd lobes 푃1 and 푃2 of 퐺 which were replaced by 푒 1 and 푒 2 are the two constituent paths of a removable double ear of 퐺. So, suppose that 퐻 has no removable doubleton. By Theorem 16.9, 푅1 must contain an edge that is (푏 + 푝)-invariant. Our task now is to find such an edge. First obtain a tight cut decomposition C of 퐻 (See Exercise 5.4.5). Then, the parameter (푏 + 푝) (퐻) is determined, by checking, for each graph 퐽 ∈ C whether it is bipartite, and, if not bipartite, whether the underlying simple graph of 퐽 is the Petersen graph.

Exercises of Chapter 17

545

Likewise, for each removable edge 푒 of 퐻 (an edge in 푅1), determine the parameter (푏+푝) (퐻−푒) of 퐻−푒. There must be an edge 푒 such that (푏 + 푝) (퐻 − 푒) = (푏 + 푝) (퐻). In other words, 푒 is a (푏 + 푝)-invariant edge of 퐻. The odd lobe of 퐺 which was replaced by 푒 is thus a (푏 + 푝)-invariant ear of 퐺.

A.17 Chapter 17

Exercise 17.2.1 By hypothesis, 퐵1 and 퐵2 are two maximal nontrivial barriers of 퐺 − 푒. By Theorem 3.11, 퐵1 and 퐵2 are disjoint. Thus, 퐼1 and 퐼2 are also disjoint; hence 푋1 ∩ 푋2 = (퐵1 ∩ 퐼2 ) ∪ (퐵2 ∩ 퐼1 ). By the Three Case Lemma, index(푒) = 2 and 푋2 − 푋1 is the bipartite shore of a nontrivial tight cut. Thus, 1 + |퐼2 − 퐵1 | = 1 + |퐼2 − 푋1 | = |퐵2 − 푋1 | = |퐵2 − 퐼1 |. Furthermore, 퐵2 is a special barrier of 퐺 − 푒; hence 1 + |퐼2 | = |퐵2 |. Subtraction of the first equality from the second one yields the asserted equality.

Exercise 17.2.2 We have seen, in the solution of Exercise 17.2.1, the equality 푋1 ∩ 푋2 = (퐵1 ∩ 퐼2 ) ∪ (퐵2 ∩ 퐼1 ). We also saw that |퐵1 ∩ 퐼2 | = |퐵2 ∩ 퐼1 |. Thus, 퐺 [푋1 ∩ 푋2 ] is equipartite. The second part of the assertion of the exercise comes immediately from the Three Case Lemma.

Exercise 17.6.2 By definition, both ends of 푒 ′ are in 푋 and its end 푡 ′ is in 푋 ′ . Thus, the end 푠 of 푒 ′ is in 푋 − 푋 ′ . By definition, the end 푢 of 푒 in 푋 is in 퐴. Thus, 푋− ⊂ 퐴 and 푋+ ⊂ 퐵. Also by definition, the end 푡 ′ of 푒 ′ is in 푋 ′ and in 퐵. Thus, 푋+′ ⊂ 퐴. Let 퐻 := 퐺 − 푒 − 푒 ′ and let 퐹 := 휕 ( 푋 ∩ 푋 ′ ). Let us prove that 퐹 − 푒 − 푒 ′ is not tight in 퐻. For this, assume the contrary. The subset of 퐻 induced by 푋 ∩ 푋 ′ is then connected. Thus, ( 퐴 ∩ 푋 ∩ 푋 ′ , 퐵 ∩ 푋 ∩ 푋 ′ ) is the only bipartition of the subgraph of 퐻 induced by 푋 ∩ 푋 ′ .

546

Solutions to Selected Exercises

The cuts 퐶 − 푒 and 퐶 ′ − 푒 − 푒 ′ are both tight in 퐻. Therefore the four subgraphs of 퐻 induced by 푋, 푋, 푋 ′ and 푋 ′ are connected. Thus, 퐻 has four edges 푣 푖 푤 푖 , 1 ≤ 푖 ≤ 4, where 푣 1 ∈ 푋 ∩ 푋 ′ , 푤 1 ∈ 푋 ′ − 푋, 푣 2 ∈ 푋 − 푋 ′ , 푤 2 ∈ 푋 ∩ 푋 ′ , 푣 3 ∈ 푋 ∩ 푋 ′ , ′ 푤 3 ∈ 푋 − 푋 ′ , 푣 4 ∈ 푋 ′ − 푋 and 푤 4 ∈ 푋 ∩ 푋 . See Figure 17.11. As 푋+ ⊂ 퐵, the vertices 푣 1 and 푣 2 are in 퐵. Likewise, 푋+′ ⊂ 퐴; thus, the vertices 푣 3 and 푣 4 are in 퐴. Thus, 푤 3 ∈ 퐵. In sum the cut 퐹 − 푒 − 푒 ′ has two edges, one incident with 푣 1 and the other incident with 푣 3 . Thus, 퐹 − 푒 − 푒 ′ has two edges, one incident with a vertex in one of the parts of 퐻 [푋 ∩ 푋 ′ ], the other incident with the other part of 퐻 [푋 ∩ 푋 ′ ]. This is a contradiction. We conclude that 퐹 − 푒 − 푒 ′ is not tight in 퐺 − 푒 − 푒 ′ . By Theorem 4.12, 퐷 − 푒 − 푒 ′ and 퐷 ′ − 푒 − 푒 ′ are tight cuts of 퐺 − 푒 − 푒 ′ . The edge 푒 ′ has no end in 푋. Thus, 퐷 ′ − 푒 is tight in 퐺 − 푒. Let us now prove that 퐷 is nontrivial. The pair ( 퐴 ∩ ( 푋 − 푋 ′ ), 퐵 ∩ ( 푋 − 푋 ′ )) is the unique bipartition of (퐺 − 푒 − 푒 ′ ) [푋 − 푋 ′ ]. The edge 푣 3 푤 3 is an edge of 퐷 − 푒 − 푒 ′ which is incident with 푤 3 , a vertex of 퐵 ∩ ( 푋 − 푋 ′ ). Thus, ( 푋 − 푋 ′ )+ ⊂ 퐵. The end 푠 of 푒 ′ is in 퐴 ∩ ( 푋 − 푋 ′ ), in turn a set equal to ( 푋 − 푋 ′ ) − . We deduce that 퐷 is nontrivial.

Exercise 17.6.4 By hypothesis, 푒 is 푏-invariant but not thin. Thus, 퐺 has a nontrivial cut 퐶 := 휕 ( 푋) such that 퐶 − 푒 is tight in 퐺 − 푒, 푋 is the bipartite shore of 퐶 − 푒, 퐵 denotes the barrier of 퐺 − 푒 associated with 퐶 − 푒, and the set 퐼 of isolated vertices of 퐺 − 푒 − 퐵 has a vertex, 푠, which has degree three or more in 퐺 − 푒. Let 퐻 denote the bipartite graph (퐺 − 푒)/푋. At most one edge of 휕 (푠) − 푒 is not removable in 퐻. Every edge of 휕 (푠) − 푒 that is removable in 퐻 is removable in 퐺 − 푒. By the exchange property of solid graphs (Theorem 10.23), every edge that is removable in 퐺 − 푒 is removable in 퐺. Thus, at least two edges of 휕 (푠) − 푒, say, 푒 ′ and 푒 ′′ , are removable in 퐺. As 퐺 is solid, both 푒 ′ and 푒 ′′ are 푏-invariant in 퐺. By the proof of Theorem 17.16, 푟 (푒 ′ ) ≥ 푟 (푒), 푟 (푒 ′′ ) ≥ 푟 (푒) and at least one of these inequalities is strict. Suppose now that equality holds for one of these inequalities; suppose that 푟 (푒 ′ ) = 푟 (푒). By the proof of Theorem 17.16, Case 2.2 applies, 퐺 has a cut 퐶 ′ := 휕 ( 푋 ′) such that 퐶 ′ − 푒 ′ is tight in 퐺 − 푒 ′ , 퐶 and 퐶 ′ cross, 푠 ∈ 푋 − 푋 ′ and 푒 has an end in 푋 ′ − 푋 and an end in 푋 ∩ 푋 ′ . Thus, 푠 is not an end of 푒.

Exercise 17.6.5 Let 퐺 be a solid brick distinct from 퐾4 . Each member of a pair of multiple edges of 퐺 is thin. We may thus assume that 퐺 is simple. The bricks 퐾4 , 퐶6 and H8 are the only bricks with fewer than two removable edges (Theorem 9.16). The bricks 퐶6 and H8 are not solid. Furthermore, by hypothesis, 퐺 is not 퐾4 . Thus, 퐺 has at least two removable edges. If every removable edge of 퐺 is thin then the assertion holds. We may thus assume that 퐺 has removable edges that are not thin.

Exercises of Chapter 19

547

By Theorem 10.16, every removable edge of 퐺 is 푏-invariant. Thus, 퐺 has at least two 푏-invariant edges. Let 푟 ′ be the maximum rank of 푏-invariant edges of 퐺 that are not thin. Let 푒 be a 푏-invariant edge of 퐺 such that 푟 (푒) = 푟 ′ . By Theorem 17.35, 퐺 has two adjacent 푏-invariant edges, 푒 ′ and 푒 ′′ , such that at least one of 푒 ′ and 푒 ′′ is thin. Let 푠 be the common end of 푒 ′ and 푒 ′′ . If 푒 ′ and 푒 ′′ are both thin then the assertion holds. We may thus assume that precisely one of 푒 ′ and 푒 ′′ , say, 푒 ′ , is thin. In that case, 푟 (푒 ′′ ) = 푟 (푒) = 푟 ′ . We then apply Theorem 17.35 with 푒 ′′ playing the role of 푒, thereby obtaining a thin edge 푓 . If 푓 and 푒 ′′ do not share the common end 푠 of 푒 ′ and 푒 ′′ then 푓 and 푒 ′ are distinct; hence 퐺 has two thin edges. Alternatively, if 푓 is incident with 푠 then 휕 (푠) − 푒 ′′ has two or more edges having maximum rank, hence thin.

A.18 Chapter 18

Exercise 18.2.1 If index(푒) = 0 then 퐺š − 푒 = 퐺 − 푒. In that case, the edges 푓 and 푔 are parallel in 퐺 − 푒, hence removable in 퐺 − 푒. We may thus assume that index(푒) > 0. Let 퐶 := 휕퐺−푒 ( 푋) be a nontrivial tight cut of 퐺 − 푒 such that the set 푋 is − 푒. If index(푒) ∈ {1, 2} then contracted to a single vertex which is a vertex of 퐺š 푋 := 푁퐺−푒 (푣) + 푣, where 푣 is an end of 푒. Alternatively, if index(푒) = 3 then 푋 = 푁퐺−푒 ({푣 1 , 푣 2 }) + 푣 1 + 푣 2 , where 푣 1 and 푣 2 are the two ends of 푒. Let 퐻 := (퐺 − 푒)/ 푋. The retract of 퐻 has just two vertices and 퐶 is its set of edges. Thus, each edge of 퐶 is removable in 퐻. We deduce that every edge that is removable in 퐺š − 푒 is removable in 퐺 − 푒. In particular, the edges 푓 and 푔 are removable in 퐺 − 푒.

A.19 Chapter 19

Exercise 19.2.1 Let us fix a sense of traversal of 퐷. Adjust notation so that the sequence of vertices of 푄 in this sense of traversal is 1, 2, 3, . . . , 2푛, 1 (see Exercise 19.1.1). The edge set 퐸 (푄) of 푄 has a partition into two perfect matchings, say 푀1 and 푀2 , where 푀1 := {{1, 2}, {3, 4}, . . . , {2푛 −1, 2푛}}, and 푀2 := {{2, 3}, {4, 5}, . . . , {2푛 −2, 2푛 − 1}, {2푛, 1}}. Let us now prove that sign(푀1 ) · sign(푀2 ) = (−1) |rv(푄) |+1 by induction on |rv(푄)|. If rv(푄) = ∅, then 푄 is a directed cycle. It is easily seen that, in this case, sign(푀1 ) = 1 and sign(푀2 ) = −1, and thus that 푀1 and 푀2 have distinct signs. It follows that 푄 is evenly oriented and 퐷 is not Pfaffian. The assertion holds in this case.

Solutions to Selected Exercises

548

Let us now consider the case in which rv(푄) contains an edge, say 푒. Let 퐷 ′ be obtained from 퐷 by reversing the orientation of the edge 푒. For 푖 = 1, 2, denote by sign′ (푀푖 ) the sign of 푀푖 in 퐷 ′ . Precisely one of 푀1 or 푀2 , which contains the edge 푒, has its sign changed. Moreover, |rv(푄 ′ )| = |rv(푄)| − 1. By induction, ′

sign(푀1 ) · sign(푀2 ) = − sign′ (푀1 ) · sign′ (푀2 ) = −(−1) |rv(푄 ) |+1 = (−1) |rv(푄) |+1 . The assertion holds, by induction. It follows that sign(푀1 ) = sign(푀2 ) if and only if |rv(푄)| is odd. That is, 퐷 is Pfaffian if and only if 푄 is oddly oriented.

A.20 Chapter 20

Exercise 20.1.2 (a) An arc 푒 of 퐷 is reversed in 퐷 ′ := (퐷 rev 퐶1 ) rev 퐶2 if and only if either (i) 푒 is in 퐶1 but not in 퐶2 or (ii) 푒 is in 퐶2 but not in 퐶1 . Thus, the set of arcs that have distinct orientations in 퐷 and in 퐷 ′ is equal to 퐶1 △ 퐶2 . That is, (퐷 rev 퐶1 ) rev 퐶2 = 퐷 rev (퐶1 △ 퐶2 ). (b) The relation of similarity is certainly reflexive, because the empty set is a cut of any graph. The relation is also symmetric. For this, suppose that 퐷 1 and 퐷 2 are two orientations of 퐺 such that 퐷 2 is similar to 퐷 1 . Then, 퐺 has a cut 퐶 such that 퐷 2 = 퐷 1 rev 퐶. In that case, 퐷 1 = (퐷 1 rev 퐶) rev 퐶 = 퐷 2 rev 퐶; hence 퐷 1 is similar to 퐷 2 . Finally, let us prove that the relation of similarity is transitive. Let 퐷 1 , 퐷 2 and 퐷 3 be three orientations of a graph 퐺, such that 퐷 1 is similar to 퐷 2 and 퐷 2 is similar to 퐷 3 . Let 퐶1 be the cut of 퐺 such that 퐷 2 = 퐷 1 rev 퐶1 and let 퐶2 be the cut of 퐺 such that 퐷 3 = 퐷 2 rev 퐶2 . Then, 퐷 3 = 퐷 2 rev 퐶2 = (퐷 1 rev 퐶1 ) rev 퐶2 = 퐷 1 rev (퐶1 △ 퐶2 ), where the last equality follows from the first part. By Proposition 20.1(i), 퐶1 △퐶2 is a cut of 퐺; hence 퐷 1 and 퐷 3 are similar.

A.21 Chapter 21

Exercise 21.1.3 (i) Let 퐺 be a cubic brace of order eight and let ( 퐴, 퐵) denote the bipartition of 퐺. For every vertex 푣 of 퐴, the set 푁 (푣) of neighbours of 푣 consists of precisely three vertices of 퐵; otherwise 푁 (푣) + 푣 would be the shore of a nontrivial tight cut of 퐺. Thus, 퐺 is simple. Let 푓 be a mapping from 퐴 to 퐵 such that

Exercises of Chapter 21

549

푓 (푣) is the vertex of 퐵 − 푁 (푣), for each 푣 in 퐴. The mapping 푓 is a bijection, for if 퐵 − 푁 (푣) = 퐵 − 푁 (푤) for any two distinct vertices 푣 and 푤 of 퐴 then 푁 (푣) ∪ {푣, 푤} is the shore of a nontrivial tight cut of 퐺. As 푓 is a bijection, the graph 퐺 is isomorphic to the cube. (ii) Let 퐺 be a brace of order eight. We shall prove that the cube is a subgraph of 퐺, by induction on |퐸 |. If 퐺 has a multiple edge 푒, then, by induction, the cube is a subgraph of 퐺 − 푒, in turn a subgraph of 퐺. We may thus assume that 퐺 is simple; let ( 퐴, 퐵) denote the bipartition of 퐺. Every vertex of 퐺 has degree three or four. If 퐺 is cubic then, by the first part, 퐺 is the cube. We may thus assume that 퐺 has vertices of degree four. Clearly, the number of vertices of degree four in 퐴 is equal to the number of vertices of degree four in 퐵. Let 푣 be a vertex of degree four in 퐴; let 푤 be a vertex of degree four in 퐵. Let 퐻 := 퐺 − 푣푤. Assume, to the contrary, that 퐻 is not a brace. Let 푋 be a set of vertices of 퐻 which is a shore of a nontrivial tight cut of 퐻. Adjust notation, by replacing 푋 by the complement 푋 if necessary, so that 푋+ = 푋 ∩ 퐵. The vertex 푣 is in 푋− and the vertex 푤 is in 푋 − . Thus, the vertex 푤 is adjacent in 퐺 to each vertex of 푋− . As 푣푤 is the only edge of 퐺 that joins a vertex in 푋− to a vertex not in 푋, we deduce that 푋− is the singleton, {푣}. Then, | 푋+ | = 2; hence 푣 has degree three in 퐺, a contradiction. We conclude that 퐻 is a brace. By induction, the cube is a subgraph of 퐻, in turn a subgraph of 퐺. (iii) Let 퐺 be the simple brace obtained from the cube by the addition of an edge 푣푤. Let 퐷 be a Pfaffian orientation of the cube. The orientation is unique, up to similarity. Let 푀 be a perfect matching of 퐺 that contains the edge 푣푤. The graph 퐺 − 푣 − 푤 is a cycle which is evenly oriented (and 푀-alternating). Thus, 퐺 is not Pfaffian, by Lemma 21.4. Every simple brace of order eight distinct from the cube has the graph 퐺 as a subgraph; hence it is not Pfaffian.

Exercise 21.5.1 (i) To prove that each summand 퐺 푖 is connected, note that the graph 퐺 − 푋 is connected (Exercise 5.4.6). Every path in 퐺 − 푋 which joins a vertex in 푉 (퐺 푖 ) − 푋 to vertex 푣 is clearly a path in 퐺 푖 − 푋. Thus, 퐺 푖 − 푋 is connected; hence 퐺 푖 is connected. (ii) To prove that each summand 퐺 푖 is matchable, let 푁 be a perfect matching of 푄; extend it to a perfect matching 푀 of 퐺 + 푅. Then, 푀 ∩ 퐸 (퐺 푖 ) is a perfect matching of 퐺 푖 . (iii) To prove that each summand 퐺 푖 is a brace, let 푒 1 and 푒 2 be any two nonadjacent edges of 퐺 푖 . Let 푀 be a perfect matching of 퐺 + 푅 that contains both edges 푒 1 and 푒 2 . As 퐺 푖 is matchable, |푉 (퐺 푖 )| is even; hence the set 푀 contains an even number of edges in 퐶 := 휕퐺 (푉 (푄)). Moreover, as 퐺 푖 is bipartite, half of the edges of 푀 ∩ 퐶 are incident with vertices of one part of the bipartition of 퐺 푖 ; the other half are incident with vertices of the other part. Thus, 푀 ∩ 퐸 (퐺 푖 ) is a matching of 퐺 푖 which may be extended to a perfect matching of 퐺 푖 by the

Solutions to Selected Exercises

550

addition of at most two edges of 푄. The graph 퐺 푖 is connected and 2-extendable; hence it is a brace.

Exercise 21.5.2 To prove that each summand 퐺 푖 is Pfaffian, we prove that 퐺 + 푅 is Pfaffian, by induction on |푅|. Then, 퐺 푖 is a conformal subgraph of 퐺 + 푅; hence 퐺 푖 is Pfaffian. If 푅 is empty then 퐺 + 푅 = 퐺 and 퐺 + 푅 is Pfaffian. We may thus assume that 푅 is not empty. Let 푄 := (푞 1 , 푞 2 , 푞 3 , 푞 4 , 푞 1 ); let 푒 be an edge of 푅 and let 퐻 := 퐺 + 푒. Adjust notation so that 푒 = 푞 1 푞 2 . Extend a Pfaffian orientation of 퐺 to an orientation 퐷 of 퐻 by arbitrarily orienting edge 푒. If all the perfect matchings of 퐻 that contain the edge 푒 have the same sign in 퐷 then either 퐷 is Pfaffian or the reversal of edge 푒 in 퐷 yields a Pfaffian orientation of 퐻. Assume, to the contrary, that 퐻 contains two perfect matchings that contain the edge 푒 and have distinct signs in 퐷. Then, 퐻 has a perfect matching 푀 that contains the edge 푒 and an 푀-alternating cycle 퐶 which is evenly oriented in 퐷 and does not contain the edge 푒. The set 퐸 (퐶) − 퐸 (푄) contains edges in at most two summands. As 푛 ≥ 3, let 퐺 푗 be a summand of 퐺 that does not contain any edge in 퐸 (퐶) − 퐸 (푄). As 퐺 푗 is a brace, it has a perfect matching, 푀 푗 , such that 푀 푗 ∩ 퐸 (푄) = 푞 3 푞 4 . Replace the edges of 푀 ∩ (퐸 (퐺 푗 ) − 푞 3 푞 4 ) by the edges of 푀 푗 − 푞 3 푞 4 . This operation yields a perfect matching 푁 of 퐺 such that the cycle 퐶 is 푁-alternating and evenly oriented in 퐷 − 푒. Moreover, 푒 ∉ 푁. This is a contradiction, as 퐷 − 푒 is a Pfaffian orientation of 퐺. Exercise 21.5.3 Let 푄 := (푞 1 , 푞 2 , 푞 3 , 푞 4 , 푞 1 ). To prove that 퐺 + 푅 is Pfaffian, note that each summand 퐺 푖 is Pfaffian and let 퐷 푖 be a Pfaffian orientation of 퐺 푖 . We may adjust notation, replacing 퐷 푖 by a similar orientation if necessary, so that 푞 1 is a source of 퐷 푖 ∩ 퐸 (푄) and 푞 4 is a sink of 퐷 푖 ∩ 퐸 (푄). The cycle 푄 is a conformal subgraph of the brace 퐺 푖 ; hence 푄 is oddly oriented in 퐷 푖 . We deduce that (푞 1 , 푞 4 ) and (푞 1 , 푞 2 , 푞 3 , 푞 4 ) are directed paths in 퐷 푖 . Thus, 퐷 := ∪푖 퐷 푖 is an orientation of 퐺 + 푅. See Figure A.16(a). 푞1

푒1

푃1

푄 푞4

푒2 (a)

푞1

푞2

푞3

Fig. A.16 The orientation 퐷 and the cycle 퐶

푒1

푞2 푃2

푄 푞4

푒2 (b)

푞3

Exercises of Chapter 21

551

Let us now prove that 퐷 is Pfaffian. The graph 퐺 is a brace; therefore 퐺 + 푅 is also a brace. Let 푀 be a perfect matching of 퐺 + 푅 that contains the edges 푒 1 := 푞 1 푞 2 and 푒 2 := 푞 3 푞 4 . Let 퐶 be an 푀-alternating cycle of 퐺 + 푅. Suppose that all the edges of 퐶 are in the same summand of 퐺 + 푅, say, 퐺 1 . Then, as 퐷 1 is a Pfaffian orientation of 퐺 1 , and since 푀1 := 푀 ∩ 퐸 (퐺 1 ) is a perfect matching of 퐺 1 , we deduce that 퐶 is oddly oriented in 퐷 1 . As 퐷 is an extension of 퐷 1 , we conclude that 퐶 is oddly oriented in 퐷. We may thus assume that 퐸 (퐶) − 퐸 (푄) has edges in more than one summand of 퐺 + 푅. Then, both 푒 1 and 푒 2 are in 퐶 and 퐸 (퐶) − 퐸 (푄) has edges in precisely two summands of 퐺 + 푅, say, 퐺 1 and 퐺 2 . Then, 퐶 has two segments 푃1 and 푃2 , where 푃1 is a path in 퐺 1 − 퐸 (푄) from 푞 4 to 푞 1 , 푃2 is a path in 퐺 2 − 퐸 (푄) from 푞 2 to 푞 3 , and 퐶 = 푃1 · (푞 1 , 푞 2 ) · 푃2 · (푞 3 , 푞 4 ). See Figure A.16(b). The set 푁 := 푀 △ 퐸 (푄) is a perfect matching of 퐺 + 푅, 퐶1 := 푃1 · (푞 1 , 푞 4 ) is an 푁-alternating cycle of 퐺 1 , and the cycle 퐶2 := 푃2 · (푞 3 , 푞 2 ) is an 푁-alternating cycle of 퐺 2 . By the previous case, both 퐶1 and 퐶2 are oddly oriented in 퐷. Moreover, |fw(퐶1 )| + |fw(푄)| + |fw(퐶2 )| ≡ |fw(퐶)|

(mod 2).

We conclude that 퐶 is oddly oriented in 퐷. This conclusion holds for each 푀alternating cycle of 퐺 + 푅. Thus, 퐺 + 푅 is Pfaffian; therefore 퐺 is also Pfaffian. Exercise 21.5.6 First apply Algorithm 21.31 to 퐻 and Algorithm 21.32 to each of the near-bipartite graphs 퐻1 , 퐻2 , . . . , 퐻푟 . Each of these 푟 + 1 graphs is a conformal subgraph of 퐺. If any one of these graphs is not Pfaffian then 퐺 is not Pfaffian. We may thus assume that 퐻, 퐻1 , 퐻2 , . . . , 퐻푟 are Pfaffian. We shall now produce a Pfaffian orientation of 퐺. Apply Algorithm 20.26 to determine a characteristic (Pfaffian) orientation 퐷 퐻 of 퐻 and extend it to a characteristic (Pfaffian) orientation 퐷 푖 of 퐻푖 , by making a conformal cycle of 퐻푖 that contains the edges of 푅푖 to be oddly oriented in 퐷 푖 . Note that reversal of both edges of 푅푖 yields another Pfaffian orientation of 퐻푖 . We may thus adjust notation so that the orientation of edge 푒 is the same in all the 푟 orientations 퐷 푖 . Let 퐷 := ∪푖 퐷 푖 . Let us now prove that 퐷 is Pfaffian. For this, let 푄 be a conformal cycle of 퐺. If 푒 is not in 푄 then 푄 is a conformal cycle of 퐻, and 푄 is thus oddly oriented in 퐷 퐻 , hence oddly oriented in 퐷. Alternatively, suppose that 푒 is in 푄. Then, precisely one of the edges 푓1 , 푓2 , . . . , 푓푟 is in 푄, say, 푓푖 . Then, 푄 is a conformal cycle of 퐻푖 . Thus, 푄 is oddly oriented in 퐷 푖 ; hence 푄 is oddly oriented in 퐷. In all cases considered, 푄 is oddly oriented in 퐷. Thus, 퐷 is Pfaffian; hence 퐺 is Pfaffian.

Exercise 21.6.4 We must prove that a connected graph is 푁푇-minimal non-Pfaffian if and only if it is matching covered and 푆-minimal non-Pfaffian.

552

Solutions to Selected Exercises

if part: Suppose that a matching covered graph 퐺 is non-Pfaffian, but every proper 푆-minor of 퐺 is Pfaffian. Let us prove that every proper 푁푇-minor of 퐺 is Pfaffian. We begin by proving that 퐺 is free of nontrivial tight cuts. For each nontrivial tight cut 퐶 of 퐺, both 퐶-contractions of 퐺 are proper 푆-minors of 퐺; hence Pfaffian. As 퐺 is non-Pfaffian we deduce, by Theorem 20.10, that 퐺 is free of nontrivial tight cuts. Consequently, every vertex of 퐺 is adjacent to three or more vertices. Let 퐻 be a graph obtained from 퐺 by the application of either edge removal or contraction of edges of an odd cycle. We prove that in each case 퐻 is Pfaffian. Let us first analyse the case in which 퐻 = 퐺 − 푒, for some edge 푒. Suppose that 퐺 is a brace. As 퐺 is not Pfaffian, it has six or more vertices; in that case, 푒 is removable in 퐺 and 퐻 is matching covered. Thus, 퐻 is a proper 푆-minor of 퐺; hence 퐻 is Pfaffian. We may thus assume that 퐺 is a brick. Let 푅 be an equivalence class of the dependence relation induced by 푒. The edge 푒 may be in 푅, it in fact may even be the only edge of 푅. By Corollary 9.3, 푅 is a removable class of 퐺. The graph 퐺 − 푅, a proper 푆-minor of 퐺, is Pfaffian, hence (퐺 − 푅) − 푒 is Pfaffian. The edges of 푅 − 푒, if any, are not matchable in 퐺 − 푒, hence 퐺 − 푒 is Pfaffian. This conclusion holds for each edge 푒 of 퐺. In particular, we infer that 퐺 is free of multiple edges. Suppose that 퐻 is obtained from 퐺 by the contraction of the edges of an odd cycle 푄. Let 푋 := 푉 (푄), and let 퐶 := 휕 ( 푋). Then, 퐻 := 퐺/( 푋 → 푥). Let us prove that 퐻 is Pfaffian. Suppose that 퐻 is not matching covered. It is certainly connected, as it is the result of the contraction of the edges of an odd cycle of the (connected) graph 퐺. Thus, 퐻 has a nonmatchable edge, 푒. The graph 퐻 − 푒 is an 푁푇-minor of 퐺 − 푒. We have seen above that 퐺 − 푒 is Pfaffian; hence 퐻 − 푒 is Pfaffian. As 푒 is not matchable in 퐻, we deduce that 퐻 is Pfaffian. We may thus assume that 퐻 is matching covered. The graph 퐻 ′ := 퐺/( 푋 → 푥) is also matching covered (Exercise 2.2.5). In sum, both 퐶-contractions of 퐺 are matching covered and one of them, 퐻, a proper 푆-minor of 퐺, is Pfaffian. In the two possibilities analysed, the graph 퐻 is Pfaffian. Indeed, the graph 퐻, obtained from 퐺 by either removing an edge or by contracting the set of edges of an odd cycle, is Pfaffian. Moreover, 퐺 is free of vertices of degree two. We deduce that every proper 푁푇-minor of 퐺 is Pfaffian. only if part: Suppose that 퐺 is a connected non-Pfaffian graph such that every proper 푁푇-minor of 퐺 is Pfaffian. We shall prove that (i) 퐺 is matching covered and (ii) every proper 푆-minor of 퐺 is Pfaffian. Let us first prove that 퐺 is matching covered. Let 푒 be any edge of 퐺. The graph 퐺 − 푒, a proper 푁푇-minor of 퐺, is Pfaffian. As 퐺 is not Pfaffian, we deduce that 푒 is matchable in 퐺. This conclusion holds for each edge 푒 of 퐺. Moreover, by hypothesis, 퐺 is connected; hence 퐺 is matching covered. Let us now proceed to prove that every proper 푆-minor of 퐺 is Pfaffian. Let 푅 be a removable class of 퐺; let us prove that 퐺 − 푅 is Pfaffian. For any edge 푒 ∈ 푅, the graph 퐺 − 푒, a proper 푁푇-minor of 퐺, is Pfaffian; hence the graph 퐺 − 푅 is Pfaffian. Let us now prove that for each nontrivial separating cut 퐶 := 휕 ( 푋), the 퐶contraction 퐻 := 퐺/( 푋 → 푥) is Pfaffian. We do this by induction on | 푋 |. Let 퐺 1 := 퐺/( 푋 → 푥).

Exercises of Chapter 21

553

Consider first the case in which 퐺 1 has a nontrivial separating cut, say, 퐷 := 휕퐺1 (푌 ). Adjust notation, replacing 푌 by 푉 (퐺 1 )−푌 if necessary, so that the contraction vertex 푥 of 퐺 1 is not in 푌 . The collection {퐶, 퐷} is cohesive (Exercise 4.2.4). Thus, 퐷 is a separating cut of 퐺. Moreover, 퐶 is a separating cut of the 퐷-contraction 퐺 ′ := 퐺/(푌 → 푦) and 퐻 is a 퐶-contraction of 퐺 ′ . By induction, 퐺 ′ is Pfaffian. By Lemma 20.13, 퐻 is Pfaffian. Consider next the case in which 퐺 1 has a removable class 푅 disjoint from 퐶. The cut 퐶 is nontrivial and separating in 퐺 − 푅. We have seen above that 퐺 − 푅 is Pfaffian. By Lemma 20.13, 퐻 is Pfaffian. We may thus assume that 퐺 1 is free of nontrivial separating cuts. In particular, this implies that 퐺 1 is free of nontrivial tight cuts. Thus, 퐺 1 is either a brace or a solid brick. We may also assume that every removable class of 퐺 1 contains an edge in 퐶. Let us now prove that 퐺 1 is not a brace. Assume the contrary, let ( 퐴, 퐵) denote the bipartition of 퐺 1 , and adjust notation so that the contraction vertex 푥 is in 퐴. Let 푣 be a vertex of 퐴 − 푥. We have assumed that every removable class of 퐺 1 contains an edge of 퐶. This implies that no edge of 퐺 1 incident with 푣 is removable. By Theorem 8.3, 퐺 1 has only four vertices. As 퐺 is free of multiple edges, 푣 has degree two; hence 퐻 is the graph obtained from 퐺 by the bicontraction of 푣. Moreover, 퐶 is tight in 퐺. Then, as 퐺 is not Pfaffian and 퐺 1 is Pfaffian, we infer from Theorem 20.10 that 퐻 is not Pfaffian. This is a contradiction, as 퐻 is a proper 푁푇-minor of 퐺 which is not Pfaffian. We conclude that 퐺 1 is a solid brick. Every removable class of 퐺 1 contains an edge in 퐶. By Corollary 10.11, 퐺 1 is an odd wheel, up to multiple spokes, having 푥 as its hub. In that case, the graph 퐻 is obtained by the contraction of the odd cycle which is the rim of 퐻1 . Thus, 퐻 is a proper 푁푇-minor of 퐺, hence 퐻 is Pfaffian. We conclude that 퐺 is matching covered and each proper 푆-minor of 퐺 is Pfaffian.

References

[1] E. Balas. Integer and fractional matchings. In P. Hansen, editor, Studies on Graphs and Discrete Programming, pages 1–13. North-Holland, 1981. [2] J. A. Bondy and U. S. R. Murty. Graph Theory with Applications. Elsevier North Holland, 1976. [3] J. A. Bondy and U. S. R. Murty. Graph Theory. Springer, 2008. [4] R. L. Brooks, C. A. B. Smith, A. H. Stone, and W. T. Tutte. The dissection of rectangles into squares. Duke Math. J., 7:312–340, 1940. [5] C. N. Campos and C. L. Lucchesi. On the relation between the Petersen graph and the characteristic of separating cuts in matching covered graphs. Technical Report 22, Institute of Computing, University of Campinas, Brazil, 2000. ´ em Orelhas para Grafos Matching [6] M. H. Carvalho. Decomposic¸a˜ o Otima Covered. PhD thesis, Instituto de Computac¸a˜ o – unicamp, 1996. in Portuguese. [7] M. H. Carvalho and C. L. Lucchesi. Matching covered graphs and subdivisions of 퐾4 and 퐶 6 . J. Combin. Theory Ser. B, 66:263–268, 1996. [8] M. H. Carvalho, C. L. Lucchesi, and U. S. R. Murty. Ear decompositions of matching covered graphs. Combinatorica, 19:151–174, 1999. [9] M. H. Carvalho, C. L. Lucchesi, and U. S. R. Murty. On a conjecture of Lov´asz concerning bricks. I. The characteristic of a matching covered graph. J. Combin. Theory Ser. B, 85:94–136, 2002. [10] M. H. Carvalho, C. L. Lucchesi, and U. S. R. Murty. On a conjecture of Lov´asz concerning bricks. II. Bricks of finite characteristic. J. Combin. Theory Ser. B, 85:137–180, 2002. [11] M. H. Carvalho, C. L. Lucchesi, and U. S. R. Murty. Optimal ear decompositions of matching covered graphs and bases for the matching lattice. J. Combin. Theory Ser. B, 85:59–93, 2002. [12] M. H. Carvalho, C. L. Lucchesi, and U. S. R. Murty. The perfect matching polytope and solid bricks. J. Combin. Theory Ser. B, 92:319–324, 2004. [13] M. H. Carvalho, C. L. Lucchesi, and U. S. R. Murty. Graphs with independent perfect matchings. J. Graph Theory, 48:19–50, 2005. [14] M. H. Carvalho, C. L. Lucchesi, and U. S. R. Murty. On the number of dissimilar Pfaffian orientations of graphs. RAIRO Theor. Inform. Appl., 39:93–113, 2005. © The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7

555

556

References

[15] M. H. Carvalho, C. L. Lucchesi, and U. S. R. Murty. How to build a brick. Discrete Math., 306:2383–2410, 2006. [16] M. H. Carvalho, C. L. Lucchesi, and U. S. R. Murty. A generalization of Little’s theorem on Pfaffian orientations. J. Combin. Theory Ser. B, 102:1241–1266, 2012. [17] M. H. Carvalho, C. L. Lucchesi, and U. S. R. Murty. On the number of perfect matchings in a bipartite graph. SIAM J. Discrete Math., 27:940–958, 2013. [18] M. H. Carvalho, C. L. Lucchesi, and U. S. R. Murty. Thin edges in braces. Electron. J.Combin., 22(#P4.14):38, 2015. [19] M. H. Carvalho, C. L. Lucchesi, and U. S. R. Murty. On tight cuts in matching covered graphs. J. Comb., 9:163–184, 2018. [20] G. Chen, X. Feng, F. Lu, C. L. Lucchesi, and L. Zhang. Laminar tight cuts in matching covered graphs. J. Combin. Theory Ser. B, 150:177–194, 2021. [21] G. Chen, X. Feng, F. Lu, and L. Zhang. Disjoint odd cycles in cubic solid bricks. SIAM J. Discrete Math., 33:393–397, 2019. [22] M. Chudnovsky, K. Edwards, K. Kawarabayashi, and P. D. Seymour. Edgecolouring seven-regular planar graphs. J. Combin. Theory Ser. B, 115:276–302, 2015. [23] M. Chudnovsky, K. Edwards, and P. D. Seymour. Edge-colouring eight-regular planar graphs. J. Combin. Theory Ser. B, 115:303–338, 2016. [24] A. Cobham. The intrinsic computational difficulty of functions. In Y. BarHillel, editor, Logic, Methodology and Phylosophy of Science, Proceeding of the 1964 International Congress, Studies in Logic and the Foundation of Mathematics, pages 24–30. North-Holland, 1965. [25] T. H. Cormen, C. E. Leiserson, R. L. Rivest, and C. Stein. Introduction to Algorithms. The MIT Press, England, 2009. [26] M. H. de Carvalho, N. Kothari, C. L. Lucchesi, and U. S. R. Murty. Bricks with several removable doubletons. Unpublished, 2015. [27] J. C. de Pina, C. G. Fernandes, B. A. Reed, and Y. Wakabayashi. A characterization of solid bricks. Personal Communication, 2003. [28] A. L. Dulmage and N. S. Mendelsohn. Coverings of bipartite graphs. Canad. J. Math., 10:517–534, 1958. [29] Z. Dvoˇra´ k, K. Kawarabayashi, and D. Kr´al. Packing six T-joins in plane graphs. J. Combin. Theory Ser. B, 116:287–305, 2016. [30] J. Edmonds. Maximum matchings and a polyhedron with (0,1) vertices. J. Res. Nat. Bur. Standards Sect B, 69B:125–130, 1965. [31] J. Edmonds. Paths, trees, and flowers. Canad. J. Math., 17:449–467, 1965. [32] J. Edmonds, L. Lov´asz, and W. R. Pulleyblank. Brick decomposition and the matching rank of graphs. Combinatorica, 2:247–274, 1982. [33] L. Esperet, F. Kardoˇs, A. D. King, D. Kr´al’, and S. Norine. Exponentially many perfect matchings in cubic graphs. Adv. Math., 227:1646–1664, 2011. [34] I. Fischer and C. H. C. Little. A characterisation of Pfaffian near bipartite graphs. J. Combin. Theory Ser. B, 82:175–222, 2001. ¨ zerlegbare Determinanten. Sitzungber. K¨onig. Preuss. [35] G. Frobenius. Uber Akad. Wiss., XVIII:274–277, 1917.

References

557

[36] H. N. Gabow and R. E. Tarjan. A linear time algorithm for a special case of disjoint set union. J. Comput. System Sci., 30:209–221, 1985. [37] A. Galluccio and M. Loebl. On the theory of Pfaffian orientations. I. Perfect matchings and permanents. Electron. J.Combin., 6, 1999. [38] C. D. Godsil. Algebraic Combinatorics. Chapman and Hall, 1993. [39] B. Guenin. Packing 푇-joins and edge colouring in planar graphs. Technical Report CORR 2003-09, Dept of Combinatorics and Optimization, University of Waterloo, 2003. [40] P. Hall. On representatives of subsets. J. London Math. Soc., 10:26–30, 1935. [41] J. H. Halton. A combinatorial proof of Cayley’s Theorem on Pfaffians. J. Combin. Theory Ser. B, 1:333–337, 1966. [42] J. H. Halton. An identity of the Jacobi Type for Pfaffians. J. Combin. Theory Ser. B, 1:224–232, 1966. [43] J. E. Hopcroft and R. M. Karp. An 푛5/2 algorithm for maximum matchings in bipartite graphs. SIAM J. Comput., 2:225–231, 1973. [44] P. W. Kasteleyn. Dimer statistics and phase transitions. J. Math. Phys., 4:287– 293, 1963. [45] K. Kawarabayashi and K. Ozeki. A simpler proof for the two disjoint odd cycles theorem. J. Combin. Theory Ser. B, 103:313–319, 2013. [46] N. Kothari. Private communication. 2019. [47] N. Kothari. Brick Generation and Conformal Subgraphs. PhD thesis, University of Waterloo, Waterloo, Ontario, Canada, 2016. [48] N. Kothari. Generating near-bipartite bricks. J. Graph Theory, 90:565–590, 2019. [49] N. Kothari and M. H. de Carvalho. Generating simple near-bipartite bricks. J. Graph Theory, 95:594–637, 2020. [50] N. Kothari and U. S. R. Murty. 퐾4 -free and 퐶6 -free planar matching covered graphs. J. Graph Theory, 82:5–32, 2015. [51] C. H. C. Little. The parity of the number of 1-factors of a graph. Discrete Math., 2:179–181, 1972. [52] C. H. C. Little. Kasteleyn’s theorem and arbitrary graphs. Canad. J. Math., 25:758–764, 1973. [53] C. H. C. Little. A theorem on connected graphs in which every edge belongs to a 1-factor. J. Austral. Math. Soc., 18:450–452, 1974. [54] C. H. C. Little. A characterization of convertible (0, 1)-matrices. J. Combin. Theory Ser. B, 18:187–208, 1975. [55] C. H. C. Little and F. Rendl. Operations preserving the Pfaffian property of a graph. J. Aust. Math. Soc. (Series A), 50:248–257, 1991. [56] L. Lov´asz. Three short proofs in graph theory. J. Combin. Theory Ser. B, 19:269–271, 1975. [57] L. Lov´asz. Ear decompositions of matching-covered graphs. Combinatorica, 3:105–117, 1983. [58] L. Lov´asz. Matching structure and the matching lattice. J. Combin. Theory Ser. B, 43:187–222, 1987. [59] L. Lov´asz and M. D. Plummer. Matching Theory. Number 29 in Annals of Discrete Mathematics. Elsevier Science, 1986.

558

References

[60] L. Lov´asz and M. D. Plummer. Matching Theory. Reprint of the 1986 original. AMS Chelsea Publishing, 2009. [61] L. Lov´asz and S. Vempala. On removable edges in matching covered graphs. Unpublished Manuscript. [62] F. Lu, N. Kothari, X. Feng, and L. Zhang. Equivalence classes in matching covered graphs. Discrete Math., 343, 2020. [63] C. L. Lucchesi. On a theorem of Fischer and Little on Pfaffian orientations of near-bipartite graphs. Submitted for publication, 2023. [64] C. L. Lucchesi, M. H. de Carvalho, N. Kothari, and U. S. R. Murty. On two unsolved problems concerning matching covered graphs. SIAM J. Discrete Math., 32:1478–1504, 2018. [65] Y. Ma, D. Mattiolo, E. Steffen, and I. H. Wolf. Pairwise disjoint perfect matchings in 푟-edge-connected 푟-regular graphs. SIAM J. Discrete Math., 37:1548–1565, 2023. [66] G. Mazzuoccolo. The equivalence of two conjectures of Berge and Fulkerson. J. Graph Theory, 68:125–128, 2011. [67] D. McCarthy and R. G. Stanton, editors. Selected Papers of W. T. Tutte. Charles Babbage Research Centre, Manitoba, Canada, 1979. (in two volumes). [68] W. McCuaig. Brace generation. J. Graph Theory, 38:124–169, 2001. [69] W. McCuaig. P´olya’s permanent problem. Electron. J.Combin., 11, 2004. p [70] S. Micali and V. V. Vazirani. An 표( |푉 ||푒|) algorithm for finding maximum matching in general graphs. Proc. 21st IEEE FOCS, pages 17–27, 1980. [71] A. A. A. Miranda. Private communication, 2008. [72] A. A. A. Miranda. Grafos Pfaffianos e Problemas Relacionados. PhD thesis, Institute of Computing, University of Campinas, Brazil, 2009. In Portuguese. [73] A. A. A. Miranda and C. L. Lucchesi. Private communication, 2008. [74] A. A. A. Miranda and C. L. Lucchesi. A polynomial time algorithm for recognizing near-bipartite Pfaffian graphs. Electron. Notes Discrete Math., 30:171–176, 2008. IV Latin-American Algorithms, Graphs and Optimization Symposium - lagos 2007, Chile. [75] A. A. A. Miranda and C. L. Lucchesi. Matching signatures and pfaffian graphs. Discrete Math., 311:289–294, 2011. [76] D. Naddef. Rank of maximum matchings in a graph. Math. Program., 22:52– 70, 1982. [77] S. Norine. Drawing 4-Pfaffian graphs on the torus. Combinatorica, 29:109– 119, 2009. [78] S. Norine and R. Thomas. Generating bricks. J. Combin. Theory Ser. B, 97:769–817, 2007. [79] S. Norine and R. Thomas. Minimally non-Pfaffian graphs. J. Combin. Theory Ser. B, 98:1038–1055, 2008. [80] J. Petersen. Die Theorie der regul¨aren Graphs. Acta Math., 15:193–220, 1891. [81] R. Rizzi. Indecomposable 푟-graphs and some other counterexamples. J. Graph Theory, 32:1–15, 1999. [82] N. Robertson, P. D. Seymour, and R. Thomas. Permanents, Pfaffian orientations and even directed circuits. Ann. of Math., 150:929–975, 1999.

References

559

[83] A. Schrijver. Theory of Linear and Integer Programming. Wiley-Interscience Series in Discrete Mathematics and Optimization, 1998 (paperback edition). [84] A. Schrijver. Combinatorial Optimization: Polyhedra and Efficiency. Springer, 2003. [85] P. D. Seymour. On multi-colourings of cubic graphs and conjectures of Fulkerson and Tutte. Proc. London Math. Soc, 38:423–460, 1979. [86] Z. Szigeti. The two ear theorem on matching-covered graphs. J. Combin. Theory Ser. B, 74:104–109, 1998. [87] Z. Szigeti. Perfect matchings versus odd cuts. Combinatorica, 22:575–589, 2002. [88] G. Tesler. Matchings in graphs on non-orientable surfaces. J. Combin. Theory Ser. B, 78:198–231, 2000. [89] W. T. Tutte. The factorization of linear graphs. J. London Math. Soc., 22:107– 111, 1947. [90] W. T. Tutte. Graph Theory as I Have Known It. Number 11 in Oxford Lecture Series in Mathematics and its Applications. Clarendon Press, Oxford, 1998. [91] V. V. Vazirani and M. Yannakakis. Pfaffian orientation of graphs, 0-1 permanents, and even cycles in directed graphs. Discrete Appl. Math., 25:179–190, 1989. [92] P. Whalen. Pfaffian Orientations, Flat Embeddings, and Steinberg’s Conjecture. PhD thesis, Georgia Institute of Technology, 2014. [93] H. Whitney. Non-separable and planar graphs. Trans. Amer. Math. Soc., 34:339–362, 1932. [94] N. C. Wormald. Classifying 퐾-connected cubic graphs. In A. F. Horadam and W. D. Wallis, editors, Combinatorial Mathematics VI, volume 748 of Lecture Notes in Mathematics, Berlin, Heidelberg, 1979. Springer. [95] D. H. Younger. William Thomas Tutte. Biogr. Mems Fell. R. Soc., 58:283–297, 2012.

List of Figures

1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9

(a) The Petersen graph; (b) The Sylvester graph . . . . . . . . . . . . . . . . . |푆| = 5, and 표(퐺 − 푆) = 7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The case in which 퐽 is not complete . . . . . . . . . . . . . . . . . . . . . . . . . . . Illustration of the proof of Tutte’s Theorem . . . . . . . . . . . . . . . . . . . . . A 4-graph which is not 4-edge-colourable . . . . . . . . . . . . . . . . . . . . . . Proof of Hall’s Theorem using Tutte’s Theorem . . . . . . . . . . . . . . . . A skew-symmetric matrix and the corresponding weighted digraph The signs of the perfect matchings in two orientations of 퐾4 . . . . . . A digraph and its adjacency matrix . . . . . . . . . . . . . . . . . . . . . . . . . . .

4 6 8 9 12 13 16 17 21

2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 2.10 2.11

Barriers in a matchable graph . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The trio of important cubic matching covered graphs . . . . . . . . . . . . The subgraph of 퐺 [ 퐴, 퐵] induced by 푆 ∪ 푁 (푆) is matching covered (a) 퐺; (b) 퐻; (c) (퐺 ⊙ 퐻)푢,푣, 휃 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (a) 퐾4 ; (b) 퐶6 ; (c) H8 , the bicorn; (d) H10 , the tricorn . . . . . . . . . . . . The Meredith graph . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (a) prism P10 ; (b) M¨obius ladder M10 . . . . . . . . . . . . . . . . . . . . . . . . . (a) Ladder L8 ; (b) Staircase S10 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Two drawings of the biwheel B10 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Two drawings of the truncated biwheel T10 . . . . . . . . . . . . . . . . . . . . . The graph 퐴10 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

24 28 30 32 33 34 35 36 37 37 39

3.1 3.2 3.3 3.4 3.5 3.6 3.7

The core H(퐵) of 퐺 with respect to the barrier 퐵 . . . . . . . . . . . . . . . . Bisubdivision of an edge 푒 = 푢푣 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . A bisubdivision of 퐾4 contained in the Petersen graph . . . . . . . . . . . Figure for Exercise 3.3.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . A bicritical graph of order six . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . An ear decomposition of the cube . . . . . . . . . . . . . . . . . . . . . . . . . . . . Graph for Exercise 3.4.3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

45 47 48 49 49 52 54

4.1 4.2

Examples of separating cuts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58 A special barrier cut . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7

561

List of Figures

562

4.3 4.4 4.5 4.6 4.7 4.8 4.9 4.10 4.11 4.12

Two 2-separation cuts in a matching covered graph . . . . . . . . . . . . . . The case in which the 3-cut 퐶 is a matching . . . . . . . . . . . . . . . . . . . . A tight cut that is neither a barrier cut nor a 2-separation cut . . . . . . A tight cut in a bipartite matching covered graph . . . . . . . . . . . . . . . . A graph 퐺 and its two 휕 ( 푋)-contractions . . . . . . . . . . . . . . . . . . . . . . Crossing cuts – the four quadrants . . . . . . . . . . . . . . . . . . . . . . . . . . . . Isomorphic contractions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Two cuts that cross (Exercise 4.3.6) . . . . . . . . . . . . . . . . . . . . . . . . . . . Graph for Exercise 4.3.10 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Illustration for the proof of Theorem 4.20 . . . . . . . . . . . . . . . . . . . . . .

5.1 5.2 5.3 5.4 5.5 5.6 5.7 5.8 5.9 5.10 5.11 5.12 5.13 5.14 5.15 5.16

The cut 퐶 is not an ELP cut, but 퐷 is a 2-separation cut . . . . . . . . . . 82 An edge 푢푣 in 퐶 such that 퐺 [ 푋 − 푣] is connected . . . . . . . . . . . . . . . 84 A nonperipheral cut in a matching covered graph . . . . . . . . . . . . . . . . 84 휕 (푢) ∩ 푋 = {푣} and 휕 (푣) ∩ 푋 = {푢} . . . . . . . . . . . . . . . . . . . . . . . . . . 88 The graphs 퐺 and 퐺 ′′ ; dashed lines indicate removed edges . . . . . . 89 Vertex 푢 is in an odd component 퐾 of 퐺 ′′ − 퐵′′ . . . . . . . . . . . . . . . . . 90 휕 (푢) ∩ 푋 = {푣} and 휕 (푣) ∩ 푋 = {푢} . . . . . . . . . . . . . . . . . . . . . . . . . . 91 The graph 퐺 is bicritical, but one of its 퐶-contractions is not . . . . . . 95 A graph obtained from the cube by two edge-extensions . . . . . . . . . . 96 Profile of a tight cut in a bipartite matching covered graph . . . . . . . . 96 Essentially 2-separation cuts in a graph . . . . . . . . . . . . . . . . . . . . . . . . 100 퐶-avoiding barriers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101 The graph 퐺 and its cut contractions . . . . . . . . . . . . . . . . . . . . . . . . . . 104 The bipartition ( 퐴∗ , 퐵∗ ) of 퐺 [푌 ] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106 The graph of Example 5.26 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 The cocktail party graph of order eight . . . . . . . . . . . . . . . . . . . . . . . . 109

6.1 6.2 6.3 6.4 6.5 6.6 6.7 6.8 6.9 6.10

Two vectors in the edge space of 퐶6 (zero entries not shown) . . . . . . 112 Proof of Edmonds’ Theorem, Case 1 . . . . . . . . . . . . . . . . . . . . . . . . . . 117 If an odd cycle 푄 is an end block of 퐺 [w], then w(휕 (푄)) < 1 . . . . . 118 Two 1-regular vectors in the edge space of 퐶6 . . . . . . . . . . . . . . . . . . 122 The matrix T(퐺) (the first six rows constitute A(퐺)) . . . . . . . . . . . . 124 A matching covered graph 퐺 with two bricks . . . . . . . . . . . . . . . . . . . 126 Merger of bases of tight cut-contractions . . . . . . . . . . . . . . . . . . . . . . . 129 Merger across a separating cut . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134 The brick of Exercise 6.3.8. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135 Extremal cubic brick of order 16 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136

7.1 7.2 7.3 7.4 7.5 7.6 7.7 7.8

A projective planar odd-intercyclic graph . . . . . . . . . . . . . . . . . . . . . . 141 Σ8 , a solid brick of order eight . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141 A brick which does not admit a unique separating cut decomposition142 Σ8 ⊙ Σ8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143 퐶 precedes 퐷 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145 Finding a separating cut using precedence . . . . . . . . . . . . . . . . . . . . . 146 Proof of Theorem 7.8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149 퐺 20 : a cubic solid brick with disjoint odd cycles . . . . . . . . . . . . . . . . 150

61 62 63 64 67 68 70 72 73 78

List of Figures

563

7.9 7.10

The solid brick 퐻12 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152 A cubic solid brick of order 14 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152

8.1 8.2 8.3 8.4 8.5 8.6 8.7 8.8 8.9 8.10 8.11 8.12 8.13 8.14 8.15 8.16 8.17 8.18 8.19 8.20 8.21 8.22

A nonremovable edge in a bipartite graph . . . . . . . . . . . . . . . . . . . . . . 157 Illustration for the case 푟 = 2 and 푑 = 3 (Theorem 8.4) . . . . . . . . . . . 159 Edge 푒 is removable in H8 but not in one of its 퐶-contraction . . . . . 162 Graph obtained by deleting an edge from the Petersen graph . . . . . . 163 The bicorn H8 and the associated dependence digraph 퐷 (H8 ) . . . . . 165 Illustration for Lemma 8.12 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166 Edge 푣푤 depends on the edge 푢푢 ′ . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168 b (H8 ) corresponding to the graph H8 in Figure 8.5 . . . 169 The digraph 퐷 An edge that induces several minimal classes . . . . . . . . . . . . . . . . . . . 169 A minimal class with eight edges (indicated by dotted lines) . . . . . . 171 The four scenarios which give rise to even 2-cuts (Case 1) . . . . . . . . 173 The case in which 푄 1 = {푒 1 , 푒 2 } and 푄 2 = {푒 2 , 푒 3 } . . . . . . . . . . . . . . 174 Barrier 퐵′′ of 퐺 − 푒 3 containing the two ends of 푒 1 . . . . . . . . . . . . . . 174 A dependence class which is not minimal . . . . . . . . . . . . . . . . . . . . . . 175 The doubleton {푒, 푓 } is a nonremovable dependence class . . . . . . . . 176 Illustration for Lemma 8.19 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177 The path 푃 and the partitions of 퐴 and 퐵 . . . . . . . . . . . . . . . . . . . . . . 179 The partitions ( 퐴12 , 퐴3 , 퐴4 ) and (퐵12 , 퐵3 , 퐵4 ) . . . . . . . . . . . . . . . . . . 179 The graph 퐻 and the partitions of 퐴 퐻 and 퐵 퐻 . . . . . . . . . . . . . . . . . . 180 The expansion of vertex 푥 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180 An illustration for Theorem 8.21. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181 A 3-connected graph with a dependence class of size three . . . . . . . 182

9.1 9.2 9.3

퐺 − 퐵 has one nontrivial component 퐾 and 퐶퐾 ∩ 푅 = {푒} . . . . . . . . 189 A removable doubleton and a 2-separation cut . . . . . . . . . . . . . . . . . . 190 The general structure of 퐺, where 푘 = 4, 푅4 = {푒 4 , 푓4 } and 퐶푖 := 푅푖 ∪ 푅4 , for 푖 = 1, 2, 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190 Mutually exclusive equivalence classes–examples . . . . . . . . . . . . . . . 192 The doubletons {푒 1 , 푓1 } and {푒 2 , 푓2 } induce an aitch configuration . 193 An extreme form of aitch graph induced by 퐾4 . . . . . . . . . . . . . . . . . . 193 The cut 푄 1 ∪ 푄 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194 Edges 푓1 and 푓2 adjacent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195 The case in which ℎ ∈ 푄 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196 The case in which ℎ ∉ 푄 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197 The adjacencies of 푎 1 and 푏 푘 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201 The adjacencies of 푎 1 and 푎 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201 The adjacencies of 푏 푘−1 and 푏 푘 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202 The adjacencies of 푢 1 , 푢 2 , 푣 1 and 푣 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . 203 A brick in which all removable edges are incident with one vertex. . 204 Edge ℎ is removable in 퐺 − 푒 − 푓 , but not in 퐺. . . . . . . . . . . . . . . . . 204

9.4 9.5 9.6 9.7 9.8 9.9 9.10 9.11 9.12 9.13 9.14 9.15 9.16 10.1 10.2 10.3

A DM barrier 퐵 in 퐺 − 푓 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209 A nonremovable edge in a bicritical graph . . . . . . . . . . . . . . . . . . . . . 211 The solid brick 퐺 and the pairs 퐵1 , 퐼1 and 퐵2 , 퐼2 . . . . . . . . . . . . . . . . 212

List of Figures

564

The case 푡 = 5 and 푠 = 9. Dashed lines indicate the possibility of an edge. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215 10.5 The brick 퐺 must be simple – heavy lines indicate the edges of 푀0 215 10.6 The wheel-like brick 퐺. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217 10.7 Illustration for Lemma 10.13 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219 10.8 Illustration for Example 10.19 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222 10.9 Illustration for Example 10.21 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223 10.10 The edge 푒 of 푅 depends on edge 푓 of 푆. . . . . . . . . . . . . . . . . . . . . . . 224 10.4

11.1 11.2 11.3 11.4 11.5 11.6 11.7 11.8 11.9 11.10

A bottom-up ear decomposition of the M¨obius ladder M8 . . . . . . . . 231 Simultaneous addition of vertex-disjoint single ears . . . . . . . . . . . . . 232 Examples of removable ears . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234 An ear decomposition of the Petersen graph . . . . . . . . . . . . . . . . . . . . 239 (a) 퐺 := P + 47; (b) 퐻, a bisubdivision of 퐶6 in 퐺 . . . . . . . . . . . . . . . 244 Bicontraction of vertex 푣 0 : (a) 퐺, (b) 퐺/( 푋 → 푥) . . . . . . . . . . . . . . . 245 b of a matching covered graph 퐺 . . . . . . . . . . . . . . . . . . . 246 The retract 퐺 Graph for Exercise 11.5.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248 An ear decomposition of a critical graph . . . . . . . . . . . . . . . . . . . . . . . 250 The graph 퐺 푟 −1 , largest block 퐿 of 퐺 푟 −1 , and the last ear 푃푢푣 . . . . . 252

12.1 12.2 12.3 12.4 12.5 12.6 12.7 12.8 12.9 12.10 12.11 12.12

The wheels 푊5 and 푊3 = 퐾4 are matching minors of H10 . . . . . . . . . 256 Bisplitting a vertex 푥 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257 The case in which 퐺 is not a brick but both 퐻1 and 퐻2 are . . . . . . . . 260 The case in which 퐻1 has a vertex of degree two . . . . . . . . . . . . . . . . 261 The Blanuˇsa snark . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263 Conformal minors: (a) 퐾4 of P; (b) 퐶6 of P + 푒; (c) 퐶6 of 퐾4 ⊙ P . . 264 Primary Ear Decomposition (Exercise 12.2.2) . . . . . . . . . . . . . . . . . . 266 퐶6 is a conformal minor of S12 + 푒 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267 A planar matching covered graph that is 퐾4 -free . . . . . . . . . . . . . . . . 269 Trellis, the smallest nonplanar 퐾4 -free brick . . . . . . . . . . . . . . . . . . . . 270 The bricks of 퐺 − 푢 1 푢 2 and 퐺 − 푢 2 푣 2 . . . . . . . . . . . . . . . . . . . . . . . . . 271 The Heawood graph (a nonplanar 퐾3,3 -free brace) . . . . . . . . . . . . . . . 273

13.1 13.2 13.3 13.4 13.5

푏-Invariant edges, Exercise 13.1.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 280 The edges 푓1 and 푓2 are (푏 + 푝)-invariant in P + 푒 . . . . . . . . . . . . . . . 281 The three cases of Lemma 13.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283 Index of 푒 is three; (a) |푀푒 ∩ 퐶| = 1; (b) |푀푒′ ∩ 퐶| = 3 . . . . . . . . . . 286 Index of 푒 is one; (a) |푀푒 ∩ 퐶| = 1; (b) |푀푒′ ∩ 퐶| = 3 . . . . . . . . . . . . 286

14.1 14.2 14.3 14.4 14.5

Cut 퐶 is robust but 퐷 is not . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 290 2-separation cuts 퐷 1 and 퐷 2 in 퐺 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . 290 퐺 1 := 퐺/( 푋 → 푥) (Lemma 14.2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 292 Peripheral robust cuts (Corollary 14.5) . . . . . . . . . . . . . . . . . . . . . . . . 294 Minimality of | 푋 | in Lemma 14.6 cannot be replaced by minimality of 푋 as a set. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 297 The case in which (퐶 − 푒, 퐷 − 푒) is a barrier cut pair of 퐺 − 푒 . . . . 299 Essentially 2-separation cuts 퐶 − 푒 and 퐷 − 푒 of 퐺 − 푒. . . . . . . . . . . 300

14.6 14.7

List of Figures

14.8

565

The pair (퐶 − 푒), 퐷 − 푒) is an essentially 2-separation cut pair, but 퐷 is not separating in 퐺. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 301

Edge 푒 is in 퐶 and not 푏-invariant in 퐺 1 . Solid lines indicate the edges of 푀0 ∩ 퐶 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 306 15.2 (a) Cuts 퐶 and 퐷 in 퐺; (b) 퐻 = 퐺 2 /푌 . (If 푣 ≠ 푦 is any vertex in 퐵, then some edge incident with 푣 is 푏-invariant in both 퐺 2 and 퐺.) . . 308 15.3 Edge 푒 is the only 푏-invariant edge of the bicorn . . . . . . . . . . . . . . . . 309 15.4 Graph 퐺 in the case 퐺 2 = P. Solid lines indicate 푏-invariant edges of 퐺 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 309 15.5 The case in which (퐶 − 푒, 퐷 − 푒) is a barrier cut pair of 퐺 − 푒. The graph 퐻 = (퐺 − 푒)/( 푋 → 푥)/(푌 → 푦) = (퐺 2 − 푒)/(푌 → 푦) is bipartite. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 312 15.6 The graphs 퐺 − 푒 and 퐺 2 − 푒 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 313 15.7 The two (푈 − 푒)-contractions of 퐺 2 − 푒 . . . . . . . . . . . . . . . . . . . . . . . 314 15.8 The perfect matchings 푀0 and 푀1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 318 15.9 Solid dots represent vertices known to have degree three. (a) The ′′ ′ ′′ vertices 푣 ′2 , 푣 ′3 , 푣 ′′ 2 and 푣 3 . (b) the vertices 푣 2 and 푣 2 coincide. . . . . 319 ′ ′ 15.10 (a) The vertices 푣 2 and 푣 3 , where solid dots represent vertices known to have degree three; (b) the Petersen graph. . . . . . . . . . . . . . 320 15.1

16.1 16.2 16.3 16.4 16.5 16.6 16.7 16.8 16.9

(a) A 7-regular vector x in Z퐸 ; (b) the 9-regular vector y = x + 2휒 푁 328 The pentagonal prism 퐺 := P10 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 330 The brick for Exercise 16.1.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 331 (a) A near-brick; (b) its retract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 333 A graph with two bricks 퐺/푋 and 퐺/푌 , and a brace . . . . . . . . . . . . . 334 A near-brick with 푏 + 푝 = 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 336 Examples of matching integral vectors . . . . . . . . . . . . . . . . . . . . . . . . 338 A 5-cycle vector w in the edge space of P. . . . . . . . . . . . . . . . . . . . . . 339 A vector in the edge space of the pentagonal prism. . . . . . . . . . . . . . . 341

17.1 17.2 17.3 17.4 17.5 17.6 17.7

Examples of thin edges in three different bricks . . . . . . . . . . . . . . . . . 348 Illustration for Exercise 17.1.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 351 The brick 퐺 in Exercise 17.1.3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 351 Two edges of the same pre-rank, one thin and the other not . . . . . . . 352 Thin edges in four bricks related to the Petersen graph . . . . . . . . . . . 353 Cuts and their shores in near-bricks . . . . . . . . . . . . . . . . . . . . . . . . . . . 353 The case in which the tight cuts associated with two maximal barriers of 퐺 − 푒 cross . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 356 Edge 푣푤 is the unique edge in 휕 (푣) that is removable in both 퐺 and 퐺 − 푒. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 359 The two 푏-invariant edges 푒 and 푒 ′ . . . . . . . . . . . . . . . . . . . . . . . . . . . 361 Illustration for Case 1, Theorem 17.16 . . . . . . . . . . . . . . . . . . . . . . . . . 363 The four edges 푣 푖 푤 푖 , 1 ≤ 푖 ≤ 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 364 Illustration for Case 2.1, Theorem 17.16. . . . . . . . . . . . . . . . . . . . . . . 365 Illustration for Case 2.2, Theorem 17.16 . . . . . . . . . . . . . . . . . . . . . . . 366 Expansion of a vertex 푥 by a barrier of size two . . . . . . . . . . . . . . . . . 370

17.8 17.9 17.10 17.11 17.12 17.13 17.14

List of Figures

566

17.15 Expansion of two vertices by barriers of size two . . . . . . . . . . . . . . . . 371 17.16 Expansion of a vertex by a barrier of size three . . . . . . . . . . . . . . . . . 371 17.17 Obtaining the trellis 퐺 from the graph 퐻 by expanding 푥 and 푦 into barriers of size two. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 375 17.18 Expansions of four Petersen bricks–Exercise 17.4.5. . . . . . . . . . . . . . 376 17.19 Examples of thin and nonthin edges in a brace: edges 푒 0 , 푒 1 and 푒 2 are thin; edge 푓 is not thin. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 378 17.20 The nested structure of shores of tight cuts of 퐺 − 푒. . . . . . . . . . . . . . 379 17.21 The cuts 퐶, 퐷 and 퐷 ′ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 380 17.22 Brace 퐺 for Exercise 17.5.1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 381 17.23 퐺 − 푢 1 푣 6 has two different tight cut decompositions. . . . . . . . . . . . . . 381 17.24 Figure for Exercise 17.5.4. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 382 17.25 The pre-rank of an edge, example: 푟 0 (푎푤) = 8, 푟 0 (푎푏) = 10. . . . . . . 383 17.26 푟 0 ( 푓 ) = 푟 0 (푒) = 6, but 푟 ( 푓 ) = 7 and 푟 (푒) = 6. . . . . . . . . . . . . . . . . . . 383 17.27 The graph 퐺 − 푒 has two braces of order six . . . . . . . . . . . . . . . . . . . . 384 17.28 Edges 푒, 푒 ′ and 푒 ′′ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 385 17.29 Illustration for Case 2.1, Theorem 17.30 . . . . . . . . . . . . . . . . . . . . . . . 387 17.30 Illustration for Case 2.2, Theorem 17.30 . . . . . . . . . . . . . . . . . . . . . . . 388 17.31 Obtaining the cube (G) from a brace (H) of order four by expansion of two vertices by barriers of size two . . . . . . . . . . . . . . . . 391 Parallel edges in 퐻, index(푒) = 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 397 Parallel edges in 퐻, index(푒) = 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 398 Parallel edges in 퐻, index(푒) = 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 398 Edges 푓 and 푔 are removable in H8 − 푒 but not in H8 . . . . . . . . . . . . 399 A house with floor 푒 and ceiling 푓 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 400 Lemma 18.6: (a) index(푒) = 1; (b) index(푒) = 2 . . . . . . . . . . . . . . . . 401 Edge 푓 is removable but not thin in 퐺. . . . . . . . . . . . . . . . . . . . . . . . . 403 − 푒. Edge 푒 is thin and edges 푓 and 푔 are multiple edges in 퐻 = 퐺š Edge 푓 is thin in 퐺, but 푔 is not. (This brace is the same as the one in Figure 17.25.) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 404 18.9 Lemma 18.9: (a) 푓 is not removable; (b) 푓 is thin and ℎ′ ≠ ℎ; (c) 푓 is thin and ℎ′ = ℎ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 406 18.10 An illustration for Theorem 18.10. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 407 18.11 Illustration for Lemma 18.11. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 409 18.12 Illustration for Theorem 18.12. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 410

18.1 18.2 18.3 18.4 18.5 18.6 18.7 18.8

19.1 19.2 19.3 19.4 19.5 19.6 19.7 19.8 19.9 19.10

An orientation of 퐾3,3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 420 Even cycles in the Petersen graph . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 421 The construction of 퐵(퐷) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 423 The graph 퐾3,3 and the orientation 퐷 of 퐾3,3 . . . . . . . . . . . . . . . . . . . 424 The orientation 퐷 of the Petersen graph . . . . . . . . . . . . . . . . . . . . . . . 425 Views of 퐶1 and 퐶2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 426 Pfaffian orientations of 푊5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 428 Pfaffian orientations of 퐾4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 428 Two Pfaffian orientations of 퐶6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 429 The Heawood graph . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 429

List of Figures

567

19.11 The Koh-Tindell digraph . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 430 19.12 Graph 퐾2,4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 430 19.13 The outer face 퐹, the path 푃 and the cycle 푄 . . . . . . . . . . . . . . . . . . . 433 20.1 20.2 20.3 20.4 20.5 21.1 21.2 21.3 21.4 21.5 21.6 21.7 21.8 21.9 21.10 21.11 21.12 21.13 21.14 21.15 21.16 21.17 21.18 21.19 21.20 21.21 A.1 A.2 A.3 A.4 A.5 A.6 A.7 A.8 A.9 A.10 A.11 A.12

The cut 퐶 is tight, both 퐶-contractions of 퐷 are Pfaffian, yet 퐷 is not Pfaffian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 442 A tight cut 퐶 of 퐺 and a Pfaffian orientation 퐷 of 퐺 where neither 퐶-contraction of 퐷 is Pfaffian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 444 Proof of Lemma 20.13 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 445 A non-Pfaffian graph which is a splicing of two 5-wheels . . . . . . . . . 447 The brick 퐺 and its orientation 퐷 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 458 푆-minors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 463 The brick Γ1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 463 The brick Γ2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 464 The orientation 퐷 of Γ1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 467 The chord 푢푤 of cycle 푄 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 469 The graph 퐾3,3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 472 A quadrilateral of 퐻 that contains two vertices in {푥1 , 푥2 , 푦 1 , 푢푦 2 } . 477 The triangle 푇 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 478 The path 푃 and the partitions of 퐴 and 퐵 . . . . . . . . . . . . . . . . . . . . . . 479 The set 푋 and the cut 퐶 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 480 The enumeration of 푉 (퐽) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 481 The adjacencies of vertices in 퐴 ∩ 푉 (퐽) . . . . . . . . . . . . . . . . . . . . . . . 481 The edge 푢 2 푣 2 of 푀1 ∩ 푀2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 482 The chordal path (푢 3 , 푢 2 , 푣 2 , 푣 3 ) and the cycles 푄 1 and 푄 2 . . . . . . . . 482 The graph 퐺, where 푥1 and 푦 1 are adjacent . . . . . . . . . . . . . . . . . . . . . 484 The possibilities of graph 퐻푒 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 485 The rotunda is a 4-cycle sum of three cubes - the edges of 푄 are represented by dashed lines. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 486 The graph 퐺 (1, ∅) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 493 The graph 퐺 (3, ∅) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 493 The graph 퐺 (2, {푟 1 푟 2 }) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 494 The graph 퐺 (2, {푟 1 푟 2 , 푟 3 푟 4 }) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 494 A 3-edge connected 5-regular graph which is not matchable . . . . . . 501 The six cubic graphs of the families G4 , G6 , G8 and G10 . . . . . . . . . . 505 Graph 퐾3,3 ⊙ 퐾3,3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 506 Example for the solution to Exercise 4.2.4 4.2.4 . . . . . . . . . . . . . . . . . 510 Graphs 퐺 5 and 퐺 4 + 푒 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 513 w1 expressed as a convex linear combination . . . . . . . . . . . . . . . . . . . 518 w2 expressed as a convex linear combination . . . . . . . . . . . . . . . . . . . 518 w expressed as a convex linear combination . . . . . . . . . . . . . . . . . . . . 518 The four quadrants. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 526 An illustration for Exercise 8.1.10, with 푘 = 3 . . . . . . . . . . . . . . . . . . 527 Two crossing cuts in the collection C . . . . . . . . . . . . . . . . . . . . . . . . . . 529 A brick with just two mutually exclusive doubletons . . . . . . . . . . . . . 530

List of Figures

568

A.13 A.14 A.15 A.16

Illustration for item (iv) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 532 The possible causes for 훿(퐽) = 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 540 The retract of 퐺 − 푓1 − 푓2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 541 The orientation 퐷 and the cycle 퐶 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 550

Glossary

(퐺 ⊙ 퐻)푢,푣, 휃 (퐺 ⊙ 퐾4 ) (퐺 ⊙ 퐾4 )푢 (푏 + 푝) (퐺) 퐴 + 푥 퐴 − 푥 퐵퐺 (퐶) 퐶 ≺ 퐷 퐶  퐷 퐶퐺 (퐵) 퐷 rev 푆 퐸 퐸 even (푅) 퐸 odd (푅) 퐹 (푛) 퐺 − 푆 퐺 − 푒 퐺 − 푣 퐺 [ 퐴, 퐵] 퐺 ⊙ 퐻 퐼 (퐺) 퐼퐺 (퐵) 퐼퐺 (퐶) 퐾4

the splicing of 퐺 at 푢 and 퐻 at 푣 with respect to 31 휃 32 the splicing of vertex-transitive 퐺 and 퐾4 32 the splicing of 퐺 at 푢 and 퐾4 186 푏(퐺) + 푝(퐺) addition of element 푥 to set 퐴, commonly repre- xxi sented by 퐴 ∪ {푥} deletion of element 푥 from set 퐴, commonly rep- xxi resented by 퐴 \ {푥} 354 the majority part of 푋퐺 (퐶) 145 cut 퐶 strictly precedes cut 퐷 cut 퐶 precedes cut 퐷 145 354 the cut 휕퐺 ( 푋퐺 (퐵)) the digraph obtained from 퐷 by reversing the 438 orientations of the edges in 푆 the set of edges of a graph xxii the set of even numbered edges of path(s) of 푅 231 the set of odd numbered edges of path(s) of 푅 231 the 푛푡 ℎ Fibonacci number 38 the graph obtained from 퐺 by removing the set 푆 xxii (of edges or of vertices) the graph obtained from 퐺 by the deletion of edge xxii 푒 the graph obtained from 퐺 by the deletion of xxii vertex 푣 a graph 퐺 with bipartition ( 퐴, 퐵) 13 the splicing of 퐺 and 퐻 31 15 the set of isolated vertices of a graph 퐺 the set of isolated vertices of 퐺 − 퐵 354 the minority part of 푋퐺 (퐶) 354 27 the complete graph on four vertices

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7

569

570

푁 (푆) 푁 (푣) 푁퐺 (푆)

Glossary

xxii the neighbourhood of a set 푆 of vertices the neighbourhood of vertex 푣 xxii the neighbourhood of a set 푆 of vertices in xxi graph 퐺 푁퐺 (푣) the neighbourhood of vertex 푣 in graph 퐺 xxi set difference, commonly represented by 푆 \ 푇 xxi 푆 − 푇 the set of vertiocs of a graph xxii 푉 28 wheel whose rim has order 푘 푊 푘 the majority part of a bipartite graph, where 푋 is 63 푋+ its set of vertices the minority part of a bipartite graph, where 푋 is 63 푋− its set of vertices the set 퐵 ∪ 퐼퐺 (퐵) 푋퐺 (퐵) 354 the bipartite shore of tight cut 퐶 of near-brick 퐺 354 푋퐺 (퐶) Int Cone(퐺) 132 the integer cone of 퐺 the perfect matching cone of 퐺 Cone(퐺) 120 퐺/푋 the graph obtained from 퐺 by contracting a xxiii nonempty subset 푋 of 푉 to a new single vertex and removing all resulting loops the graph obtained from 퐺 by contracting a xxiii 퐺/( 푋 → 푥) nonempty subset 푋 of 푉 to a new single vertex 푥 and removing all resulting loops xxiii the maximum degree of the vertices of 퐺 Δ(퐺) the graph Γ1 , see Figure 21.2 463 Γ1 Γ2 464 the graph Γ2 , see Figure 21.3 115 Poly(퐺) the perfect matching polytope of 퐺 144 Poly∗ (퐺) the fractional perfect matching polytope of 퐺 the tight-cut perfect matching polytope of 퐺 144 Poly푡 (퐺) 17 Pfaffian of A Pf (A) Φ(퐺) the number of perfect matchings of 퐺 38, 136 111 R퐸 the edge space of graph 퐺 := (푉, 퐸) 퐺{푥 → (푎, 푏 1 , 푏 2 )} constrained bisplitting vertex 푥 in graph 퐺 369 H8 the bicorn 32 the biwheel of order 푛 36 B푛 the incidence vector of a set 퐹 of edges 112 휒 퐹 xxiii the minimum degree of the vertices of 퐺 훿(퐺) det(A) 16 the determinant of A 420 the set of forward edges of a trail 푇 fw(푇) xxii, the edge-connectivity of graph 퐺 휅 ′ (퐺) xxiii the connectivity of graph 퐺 xxii 휅(퐺) 35 L푛 the ladder of order 푛 the characteristic of cut 퐶 in a graph 휆(퐶) 73 휆(퐺) the characteristic of graph 퐺 73, 74 Lat(퐺) the matching lattice of 퐺 133 135, 341 the matching lattice of 퐺 over F Lat(퐺, F)

Glossary

Lat∗ (퐺) Lin(퐺) Lin(퐺, 퐶) Lin(퐺, F) H(퐵) P B B1 ∨ B2 ML M P ST TB T (퐺) M푛 퐶6 푋 휕퐺 (퐻) 휕퐺 (푆) 휕퐺 (푣) P푛 rv(푇) sign(푀) S푛 휏(퐺) H10 T푛 x◦y x(퐹) 푏(퐺) 푑 (G) 푒 ⇔ 푓 푒 ⇒ 푓 푚 푛 푝(퐺) O(퐺) co-NP

def(푀) 표(퐺)

571

the set of matching integral vectors of 퐺 337 121 the perfect matching space of 퐺 131 a subspace of Lin(퐺) the matching space of 퐺 over F 135, 341 the core of a graph with respect to barrier 퐵 44 27 the Petersen graph 36 the family of biwheels 128 the merger operation of B1 and B2 35 the family of M¨obius ladders 35 the set of perfect matchings of a graph 34 the family of prisms the family of staircases 36 37 the family of truncated biwheels the set of tight cuts of 퐺 121 34 the M¨obius ladder of order 푛 the triangular prism 27 the complement 푉 − 푋 of 푋 58 the edge cut 휕퐺 (푉 (퐻)) xxii the edge cut of 퐺 associated with 푆 xxii the edge cut of 퐺 associated with {푣} xxii the prism of order 푛 34 the set of reverse edges of a trail 푇 420 the sign of a perfect matching 푀 16 the staircase of order 푛 36 the number of equivalence classes of Pfaffian ori- 449 entations of 퐺 with respect to similarity the tricorn 32 the truncated biwheel of order 푛 36 the scalar product of x and y 163 Í 112 푒∈퐹 x(푒) the number of bricks of 퐺 76 the number of double ears of ear decomposi- 240 tion G edges 푒 and 푓 are mutually dependent 164 edge 푒 depends on (implies) edge 푓 164 the size of a graph xxii the order of a graph xxii the number of Petersen bricks of 퐺 186 the set of odd components of graph 퐺 6 the class of decision problems whose comple- 10 ments are solvable by nondeterministic algorithms in polynomial time the deficiency of a matching 푀 7 the number of odd components of graph 퐺 6

572

Glossary

NP

the class of decision problems solvable by non- 10 deterministic algorithms in polynomial time

P

the class of decision problems solvable in poly- 10 nomial time

Index

(푏 + 푝)-invariant edge, 280 single ear, 332 퐶-avoiding barrier, 101 set, 107 퐶-contractions, 58 퐶-sheltered set, 107 퐽-based graphs, 267 퐽-free graphs, 267 퐾4 , 퐶6 , H8 , 199 퐾4 , 퐶6 , 195 퐾4 -free planar bricks, 269 퐾4 -planar bricks a characterization, 269 푁푇-minimal, 492 푁푇-minor, 491 푅-triples, 466 푆-minimal, 464 푆-minor, 462 퐶6 -free planar bricks, 271 the list, 271 푏-invariant ear, 278 edge, 206, 278 index, 283 푏-invariant edges existence, 304 proof, 304 푓 -factors, 22

푘-edge-connected graph, xxii 푘-extendable graph, 39, 108 푘-regular bipartite graphs, 14 푘-edge-colourable, 14 푟-graphs, 12, 27 푟-regular vector, 121 푣 0 -wheel, 213 1-extendable graph, 108 1-extendable graphs, 39 1-factors, 3, 22 2-extendable graph, 108 2-extendable bipartite graph, 96, 98 2-separation, 61 cut, 61 3-edge cut, 59, 61, 66, 92, 93 4-cycle sums, 486 addition of ear, 50 addition of edge in bipartite graph, 43 adjugate matrix, 21 aitch configuration, 193 induced by doubletons, 193 alternating cycles, 11 alternating paths, 11 augmenting paths, 11 Balas, 148 barrier cut, 60 associated with a barrier, 82

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2024 C. L. Lucchesi, U. S. R. Murty, Perfect Matchings, Algorithms and Computation in Mathematics 31, https://doi.org/10.1007/978-3-031-47504-7

573

574

Index

barriers, 23, 41 of a graph, 74 basis matrix of Lin(퐺), 132 of a separating cut, 73 Berge, 11 characteristic orientation, 452 number, 346 characterization of braces, 96–98 bicontraction of a vertex, 245 2-extendable, 98 bicontractions Hall-type, 97 restricted, 264 characterization of bricks, 81, 90 bicorn, 32, 36, 136 Chen, Feng, Lu and Zhang, 149 bicritical graph, 46 Chen, Feng, Lu, Lucchesi and Zhang, bisplitting a vertex, 256, 369 106 constrained, 369 chord bisubdivision of a M¨obius ladder, 35 of a graph, 47 Chudnovsky, 136 of an edge, 47 CLM, xvi biwheels, 36 Cobham, 10 Bletchley Park, 5 cocktail party graph, 109 bond, xxii, 59 cocktail party graph, 108 Bondy and Murty, xxi, xxii, 3, 8, 14, cofactor matrix, 21 22 cohesive collection of cuts, 65 brace, 66 conformal characterization, 96–98 minor, 54 2-extendable, 98 subgraph, 54 Hall-type, 97 conformal minors, 264 branch vertex, 48, 246 unavoidable, 265 brick, 46, 66 conformal subgraphs, 231, 250 characterization, 81 connectivity of a graph, xxii special, 304 contraction vertex, xxiii building a brace convex hulls, 113 expansion operations, 390 convex linear combinations, 113 building a brick convex polytopes, 113 expansion operations, 368–370 convex sets, 113 extreme points, 113 Campos and Lucchesi, 302 core with respect to a barrier, 44 canonical partition, 46 Cormen, 503 Carvalho, 48 Cormen, Leiserson, Rivest and Stein, 22 Ph. D. Thesis, 304, 321 critical graphs, 12, 15 Carvalho, Kothari, Lucchesi crossing cuts, 68 and Murty, 190 cube, 34, 36 Carvalho, Lucchesi and Murty, xvi, 99, cubic 106, 136, 137, 153, 181, 182, 207, graph, xxiii 266, 271, 277, 287, 301, 304 matchable graph, 10 Carvalho and Lucchesi, 266 matching covered graph, 27, 93 Cayley’s Theorem, 18 brace, 93 central subgraph, 55 cut, xxii certificate for the nonremovability of an contractions, 58 edge, 158 characteristic

Index

de Pina, Fernandes, Reed and Wakabayashi, 147, 148 deletion of an edge, 156 dependence classes, 164, 167 in bricks, 183 mutually exclusive, 192, 195 size in bricks, 184 digraph, 164 minimal dependence class, 168 mutual, 164 in bricks, 183 reduced dependence digraph, 168 relation, 164 dimension formula, 127, 131 for braces, 125 for bricks, 125 DM-barrier, 85 Dulmage-Mendelsohn barrier, 82, 85 barriers and nonremovable edges, 208 and separating cuts, 207 decomposition, 30 Dvoˇra´ k, 136 ear, 50 addition, 50, 230 closed ear, 250 constituent path, 235 decomposition, 237 bottom up, 237 of a bipartite graph, 50, 51 of critical graphs, 249, 250 optimal, 243, 244 partial, 238 top down, 237 decompositions canonical, 266 primary, 266 deletion, 233, 234 double ear, 230 ear path, 229 even ear, 50 odd ear, 50 open ear, 250 optimal decomposition

575

of solid graphs, 244 path, 50 removable, 236 double ear, 234 ear, 234 single ear, 234 single ear, 50, 229 the two ear theorem, 235 edge dependence, 164 implication, 164 matchable, 24 mutual dependence, 164 unmatchable, 24 edge cut, xxii edge space of a graph, 111 edge-connectivity of a graph, xxii edge-extension, 95 Edmonds, 10, 115, 119 characterization of Poly(퐺), 115 Edmonds, Lov´asz and Pulleyblank, 80, 82, 90, 131 Theorem, 87 Edwards, 136 elementary graph, 39 ELP cut, 82 ultimately, 108 Theorem, 82, 87 Laminar, 106, 107 tight cut decomposition, 108 Esperet, Kardoˇs, King, Kr´al’, Norine, 137 essential 2-separation, 99 essentially 2-separation cut, 100 essentially 4-edge-connected graph, 66, 93 even directed cycle problem, 422 evenly oriented trail, 420 exchange property, 222 bipartite graphs, 163 of solid graphs, 223 external part of a bipartite shore of a tight cut, 352 extremal graph, 136 extreme point

576

of Poly∗ (퐺), 148

Index

Kotzig, 43 Kr´al, 136

factor-critical graphs, 12 Fibonacci numbers, 38 ladders, 35 Fischer-Little laminar Theorem, 473 collection of cuts, 68 forward arc, 420 ELP Theorem, 106, 107 fractional perfect matching polytope, lattices, 324 144 bases, 324 Frobenius, 5 generating sets, 324 Fulkerson’s Conjecture, 346 Leiserson, 503 Lemma on wheels, 213 Gabow and Tarjan, 514 Little, 238, 239, 277, 423 Guenin, 136 Theorem, 471 lobes, 246 half-space, 114 even lobes, 246 Hall, 5, 13 odd lobes, 246 Hall’s Theorem, 13, 14 Lov´asz, 6, 8, 39, 43, 68, 74, 80, 132, Heawood graph, 429 133, 161, 163, 183, 206, 207, 251, Hopcroft and Karp, 10 265, 277, 278, 287 hub of a wheel, 28 Conjecture, 303, 321 hubs of a biwheel, 36 Lov´asz and Plummer, 8, 22, 39, 43, 55, hyperplane, 113 80, 109, 112, 137, 170, 235 hypomatchable graphs, 12 Lov´asz and Vempala, 158, 161, 199 Lu, Kothari, Feng and Zhang, 182 incidence vector, 112 Lucchesi, Carvalho, Kothari indecomposable graph, 395 and Murty, 149, 272, 304 index of a 푏-invariant edge, 283, 349 of a removable edge, 377 induced cut by another cut, 290 inheritance of a near-brick from a brick, 354 inner vertex, 257 internal part of a bipartite shore of a tight cut, 352 intractable collection of conformal cycles, 423 of perfect matchings, 431 Jacobi’s Identity, 21 Kasteleyn’s Theorem, 432 Kawarabayashi and Ozeki, 141 Kawarabyashi, 136 Kothari, 152, 269 Kothari and Murty, 269, 272

M¨obius ladder, 34 brace, 66 brick, 66 supergraph of a ladder, 35 majority part, 63 matchable edge, 24 and barriers, 25 graph, 4, 7 matching equivalent cuts, 145 integral vectors, 337 lattice, see perfect matching lattice lattices basis matrices, 337 orthogonal vector, 127, 338 over a field, 342 space, see perfect matching space spaces

Index

over a field, 341 over a field, 135 matching covered graph, 11, 23, 26 solid, 139 matching minors, 255 and tight cuts, 258 matching-minor-closed class of graphs, 258 matchings, 3 matching covers a vertex, 14 maximum matching bipartite graphs, 15 characterization, 7 perfect matchings, 3 matrix of indeterminates, 18 maximal barrier, 41, 42 in bipartite graphs, 43 Mazzuoccolo, 346 McCarthy and Stanton, 10 McCuaig, 34, 55, 137, 277 family of graphs, 395 Menger, xxii Meredith graph, 33, 34 merger, 128 operation, 128 Micali and Vazirani, 10, 514 minimal class, see minimal dependence class minimal cut with respect to , 145 minimal dependence class, 168 induced by an edge, 168 minimal non-Pfaffian graphs, 491 Minkowski and Weyl, 114 minority part, 63 monotonicity of function 휆, 227 of function 푏, 186, 187, 236 revisited, 220 of function 푏 + 푝, 186, 187, 236 mutually exclusive doubletons in a brick, 193 Naddef, 125, 131 near-bipartite graph, 160 near-brick, 76 near-perfect matching, 250

577

neighbour set of a set of vertices, xxi of a vertex, xxi nice subgraph, 55 nonremovable edges and DM-barriers, 208 in bicritical graphs, 208 Norine and Thomas, 34, 37, 255, 277 bricks, 92 family N T of non-Pfaffian graphs, 492 family of bricks without strictly thin edges, 395 Theorem, 494 number of bricks, 76 number of perfect matchings, 38 of bipartite graphs, 39 of biwheels, 38 of ladders, 38 of M¨obius ladders, 38 of prisms, 38 of staircases, 38 of truncated biwheels, 38 odd orientation, 432 odd wheels, 28, 91 odd-intercyclic brick, 140 graph, 140 projective planar graph, 140 oddly oriented trail, 420 optimal ear decompositions, 332 removable ears, 332 classes, 332 removable classes, 319 order of a graph, xxii outer vertex, 257 parity of a trail, 420 pentagonal prism, 31, 32, 34 perfect matching cone, 120 fractional polytope, 144 integer cone, 132 lattice, 132, 133

578

Index

lattices, 323 precedence relation on cuts, 145, 218 with 푝 = 0, 326 minimal cut, 145 with 푝 = 1, 327 prism, 34 polytope, 115, 144 brace, 66 degree constraints, 115 brick, 66 non-negativity constraints, 115 supergraphs of a ladder, 35 odd set constraints, 115 prismoid, 37, 412 space, 121 quadrants, 68 characterization, 123 quasi 푏-invariant edges, 297 regular vector, 123 tight-cut polytope, 144 rank peripheral of a 푏-invariant edge, 350 bricks, 293 a removable edge of a brace, 383 of robust cuts, 293 pre-rank tight cut, 83 of a 푏-invariant edge, 350 Petersen, 4 of a removable edge of a brace, 382 brick, 186, 278 reference matching, 53 graph, 3, 5, 27, 31, 33, 136 regular vector, 121 Theorem, 10 removable Pfaffian, 6, 15–17, 22 classes, 168, 169, 174 digraph, 418 number in bricks, 185 graph, 418 size of, 170, 174 identity, 19 ear, 234 double orientation, 18, 273 156, 160 doubletons, bipartite graph, 471 with, 163 graphs near-bipartite graph, 473 191 in bricks, 185, planar graph, 432 ears, 234 problem, 418 edges, 156 recognition problem, 418 푏-invariant, 206 plane graph, 268 and separating cuts, 218 Plummer, 251 a bipartite graph, 157–159 in polynomial-time algorithm 158 in a brace, recognition of braces, 98 graphs, 208 in bicritical recognition of bricks, 90 199 197, in bricks, recognition of matchable edges, 25 in near-bipartite bricks, 199 recognition of matching covered graphs, in solid bricks, 210, 218 28 number, in bricks, 185, 197 bipartite, 29 existence of a removable class, 236 recognition of Pfaffian sets of edges, 156 near-bipartite graphs, 489 and separating cuts, 161 recognition of Pfaffian bipartite graphs, and tight cuts, 161 486 in a bipartite graph, 156 tight cut decomposition, 98 minimal, 156 to find a maximum matching, 10 single ear, 234 of a bipartite graph, 10 retract polytopes, 113 of a graph, 246

Index

579

the retract is cubic, 249 cut, 60, 99 of a path in a graph, 246 in a near-brick, 76 reverse arc, 420 special bricks, 304 rim splicing of of a biwheel, 36 bicritical graphs, 94 of a wheel, 28 bricks, 94 Rivest, 503 graphs, 31, 32 Rizzi, 136 spokes of a wheel, 28 Robertson, Seymour and Thomas, 55, staircases, 36, 198 277 Stein, 503 robust cut, 79 stepwise in a near-brick, 79 canonical basis, 240 robust cuts, 289 descending order having minimum shores, 295 matchings, 240 peripheral, 293 strictly thin edges, 394 subadditivity of function 푏, 77 rungs of a ladder, 35 subdivision Schrijver, 22, 112, 119, 330 even, 47 separating cut, 57, 58 odd, 47 characterization, 59 of an edge, 47 decomposition, 142 vertex, 48 in a bipartite graph, 65 subdivision vertices of a graph, 246 with a bipartite shore, 64 support of a vector in R퐸 , 112 separation-deletion minimal, 464 Sylvester graph, 3–5 separation-deletion minor, 462 Szigeti, 82 series reduction of a path, 235 Tait, 3 Seymour, 132, 136 the three case lemma, 282, 355 shores of a cut, xxii the two ear theorem, 235 sign thin of a perfect matching, 16 barriers, 348 similar orientations, 438 edges size of a graph, xxii in braces, 377 skew-symmetric matrices, 15 in bricks, 347 and weighted digraphs, 16 tight cut, 57, 59 solid 3-edge cut, 61, 92 brick, 140 companion, 100 infinite family, 149 decomposition, 67, 74 recognition, 142 of a bicritical graph, 93 matching covered graph, 139, 144 uniqueness, 68, 74 solid graphs in bipartite graphs, 64 optimal ear decomposition, 244 in graphs with two bricks, 99, 100 solid matching covered graphs perfect matching polytope, 144 exchange property, 221 triangular prism, 136 hereditary property, 220 triangular prism, 27, 34, 36, 37 invariance of function 푏, 220 tricorn, 32, 136 solitary edge, 320 trivial barriers, 23 special barrier, 25, 99

580

trivial cut, xxii truncated biwheels, 36 Tutte, 5, 6 Tutte’s Theorem, 7 ultimately ELP cut, 108 uncrossing cohesive pairs, 71 separating cuts, 70, 73 tight cuts, 68 unmatchable edge, 24 unsolved problems, 497 Vempala, 287 vertex covered by a matching, 14 vertex cut, xxii well-fitted subgraph, 55 wheels, 28 Lemma on, 213 wheel-like brick, 216 Younger, 5, 15

Index