To save content items to your account,
please confirm that you agree to abide by our usage policies.
If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account.
Find out more about saving content to .
To save content items to your Kindle, first ensure no-reply@cambridge.org
is added to your Approved Personal Document E-mail List under your Personal Document Settings
on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part
of your Kindle email address below.
Find out more about saving to your Kindle.
Note you can select to save to either the @free.kindle.com or @kindle.com variations.
‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi.
‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.
The subject of graph decompositions is a vast and sprawling topic, one which we certainly cannot begin to cover in a paper of this length. Indeed, recently a number of survey articles and several books have appeared, each devoted to a particular subtopic within this domain (e.g., see [Fi-Wi], [Gr-Rot-Sp],[So 1],[Do-Ro]).
What we will attempt to do in this report is twofold. First, we will try to give a brief overall view of the landscape, mentioning various points of interest (to us) along the way. When possible, we will provide the reader with references in which much more detailed discussions can be found. Second, we will focus more closely on a few specific topics and results, usually for which significant progress has been made within the past few years. We will also list throughout various problems, questions and conjectures which we feel are interesting and/or contribute to a clearer understanding of some of the current obstacles remaining in the subject.
Notation
By a graph G we will mean a (finite) set V = V(G), called the vertices of G together with a set E = E(G) of (unordered) pairs of vertices of G, called the edges of G.
Let H denote a family of graphs. By an H-decomposition of G we mean a partition of E(G) into disjoint sets E(H.) such that each of the graphs Hi induced by the edge set E(Hi) is isomorphic to a graph in H.
The aim of this review is to highlight some of the fundamental results about random graphs, mostly in areas I am particularly interested in. Though a fair number of references are given, the review is far from complete even in the topics it covers. Furthermore, very few of the proofs are indicated. The exception is the last section, which concerns random regular graphs. This section contains some very recent results and we present some proofs in a slightly simplified form.
The study of random graphs was started by Erdòs [33], who applied random graph techniques to show the existence of a graph of large chromatic number and large girth. A little later Erdös and Rényi [38] investigated random graphs for their own sake. They viewed a graph as an organism that develops by acquiring more and more edges in a random fashion. The question is at what stage of development a graph is likely to have a given property. The main discovery of Erdös and Rényi was that many properties appear rather suddenly. In the last twenty years many papers have been written about random graphs; some of them, in the vein of [33], tackle traditional graph problems by the use of random graphs, and others, in fact the majority, study the standard invariants of random graphs in the vein of [38]. Of course the two trends cannot really be separated for deep applications are impossible without detailed knowledge of random graphs.
The enumeration of graphs and other structures satisfying a duality condition, such as self-complementarity, is surveyedo A modification of Burnside's lemma due to de Bruijn is presented in order to unify and simplify the treatment of such problems. Some new equalities between classes of graphs and digraphs are found which seem not to be explained by natural 1-1 correspondences. Also some new natural 1–1 correspondences are derived using the modified Burnside lemma. Asymptotic analyses of the exact numbers are reviewed, and some recent results described.
Introduction
Methods for counting graphs and related structures which satisfy a duality condition are well-established in the literature. A duality condition is defined by invariance up to isomorphism under some operation. Complementation is an operation which has often been considered. Self-complementary structures enumerated include graphs and digraphs [Re63], tournaments [Sr70], n-plexes [Pa73a], m-ary relations [Wi74], multigraphs [Wi78], eulerian graphs [Ro69], bipartite graphs [Qu79], sets [B59 and B64], and boolean functions [Ni59, El60, Ha63, Ha64, and PaR-A]. Closely related are 2-colored or signed structures invariant under color interchange or sign interchange. These take in 2-colored graphs [HP63 and Ha7 9], graphs in which points, lines, or points and lines are signed [HPRS77], signed graphs under weak isomorphism [So80], 2-colored polyhedra [R27 and KnPR75], and necklaces [PS77 and Mi78]. The converse of a digraph results when all orientations of arcs are reversed.
Every undergraduate course in graph theory mentions basic results about finite planar graphs – Kuratowski's criterion for embeddability, Euler's Theorem, and so on. However the corresponding results for infinite graphs seem to be little known. It turns out that the concept of embeddability in the plane has many ramifications and variants in the infinite case, and one of the purposes of this exposition is to survey these. For the most part results will only be quoted and no proofs given – for these the reader is referred to the literature listed in the bibliography.
In this survey we aim to show how fruitful is the interaction between the theories of finite and of infinite planar graphs. Results from one of these fields often inspire nontrivial problems in the other, and frequently suggest analogous questions about graphs embeddable in manifolds of arbitrary genus.
Although it may seem foreign to the subject, especially to those only interested in problems of a strictly combinatorial or topological nature, quite a large part of what we shall do will be metrical in character. There are several reasons for this. For example, in our discussion of Euler's Theorem and its variants in Section 3, the results are not true unless we impose quite strong restrictions on the kinds of graph we are considering – and we only know how to formulate these restrictions in metrical terms.
In [7[ a functor Ext is defined in terms of C*-extensions. It is a covariant functor from the homotopy category of compact, metrizable spaces to abelian groups. Further details are given in [7, 8, 9, 11]. From [7, 14] Ext extends to a Steenrod homology theory, Ext*, which may be identified with the one associated with unitary K-theory. Since Lie groups are fundamental to K-theory (see [2, p. 24]) one might expect Ext(G) to be of interest when G is a Lie group.
In 1963 Mathematika published a note [2[ in which I “proved” that the equation f(x1, …, xm) = 1 could always be solved in algebraic integers f(x1, …, xm), whenever f(x1, …, xm) was a homogeneous polynomial of degree n ≥ 1, with algebraic integers of greatest common divisor 1 as coefficients. This “proof” was so good that it had to be corrected in [3]. Since neither I nor anyone else had any use for this result, these papers dropped into the decent obscurity reserved for dead ends in mathematical research. They presumably would have remained there had not Cantor [1] recently started looking at similar results. He discovered, to his surprise and mine, that the entire article [2] had been anticipated by Skolem in a 1934 monograph ]4] which, apparently, had also languished in obscurity. The only consolation I can draw from this is the observation that, if I was unaware of Skolem's article, he was unaware of Steinitz's work [5] of 1911, which he duplicated in Theorems 5 and 6 of [4]. The moral of this story is that any working mathematician would rather prove something himself than try to find it in any but the most accessible literature.
I am grateful to Professor K. Prachar for pointing out to me that there is a mistake in the proof of Theorem 2 in my paper “On the distribution of primes in short intervals” [Mathematika, 23 (1976), 4–9]. The mistake is in the assertion on p. 6 that, if 1 ≤ μ/λ < 4, the result is trivial. The corrected version reads as follows.
Consider a slab which is made from a homogeneous and isotropic thermoelastic material and which occupies the region 0 ≤ x ≤ a, where x, y, z are the usual rectangular cartesian coordinates. Suppose that the slab undergoes a motion in which the displacement vector is parallel to the x-axis and the displacement and the temperature are functions of the coordinate x and the time t ( ≥ 0) only. Suppose too that the faces of the slab are clamped, that the face x = 0 is maintained at a constant temperature, and that heat is supplied to unit area of the face x = a at a prescribed rate h(t).
This note contains characterizations of those sigma-fields for which sigma-finiteness is a necessary condition in the Radon-Nikodym Theorem.
Our purpose is to consider those σ-fields for which σ-finiteness is a necessary condition in the Radon–Nikodym Theorem. We first prove a measure theoretic equivalence in the general case, and then use this to obtain an algebraic characterization in the case when the σ-field is the Borel field of a locally compact separable metric space. For undefined terminology we refer the reader to [1] for measure theoretic and [2] for algebraic properties.
By a measure, we mean a countably additive function from σ-field of sets or a Boolean σ-algebra into the non-negative extended real numbers. We will say that a measure μ on a σ-field of sets Σ is RN provided each μ-continuous finite measure on Σ has a Radon–Nikodym derivative in L1(μ).
Much recent attention has been given to geometric representation of elements of the stable homotopy groups of spheres, π*s A particular example concerns non-singular bilinear maps ℝm+1 × ℝn+1 → ℝm+n+1−p; on restriction and normalisation these become biskew maps Sm × Sn ℝ Sm+n-p;. Now the Hopf construction ℋ applied to any map f: Sm × Sn → Sm+n-p yields
analytic and univalent in U = {z: |z| < 1} is said to be starlike there, if f(U) is f starshaped with respect to the origin, that is, if w ε f(U) implies tw ε f(U) for 0 ≤t ≤ 1. We denote by S* the class of all such functions. The Koebe function; k(z) = z(l – z)-2, z ε U, maps U onto the complex plane minus a slit along the I negative real axis from - ¼ to ∞, and thus belongs to the class S*. Recently Leung [4] has shown that, if