Chip-firing game

Last updated
Example graph with the state variables s(v) indicated Chip firing game example graph.svg
Example graph with the state variables s(v) indicated
A possible finite firing sequence, with the vertex to be fired in red - the game ends as each vertex has s smaller than its degree Chip firing game example sequence.svg
A possible finite firing sequence, with the vertex to be fired in red the game ends as each vertex has s smaller than its degree

The chip-firing game is a one-player game on a graph which was invented around 1983 and since has become an important part of the study of structural combinatorics.

Contents

Each vertex has the number of "chips" indicated by its state variable. On each firing, a vertex is selected and one of its chips is transferred to each neighbour (vertex it shares an edge with). The number of chips on each vertex cannot be negative. The game ends when no firing is possible.

Definition

Let the finite graph G be connected and loopless, with vertices V = {1, 2, . . . , n}. Let deg(v) be the degree of a vertex, and e(v,w) the number of edges between vertices v and w. A configuration or state of the game is defined by assigning each vertex a nonnegative integer s(v), representing the number of chips on this vertex. A move starts with selecting a vertex w which has at least as many chips as its degree: s(w) ≥ deg(w). The vertex w is fired, moving one chip from w along each incident edge to a neighbouring vertex, producing a new configuration defined by:

and for v ≠ w,

A state in which no further firing is possible is a stable state. Starting from an initial configuration, the game proceeds with the following results (on a connected graph).

For finite chip-firing games, the possible orders of firing events can be described by an antimatroid. It follows from the general properties of antimatroids that the number of times each vertex fires, and the eventual stable state, do not depend on the order of firing events. [1]

Dollar games

Some chip-firing games, known as dollar games, interpret the chips as dollars and the vertices as money borrowers and lenders. Two variants of dollar game are prominent in the literature:

Baker and Norine's variant

In this dollar game, negative integer values (representing debt) are assigned to some of the vertices of the finite graph G. A game is called winnable if there exists a state where all the vertices can be made positive. [2] A a graph-theoretic analogue of Riemann–Roch theorem can be used to characterize if a game is winnable or not. [2] [3]

Biggs's Variant

In a variant form of chip-firing closely related to the sandpile model, also known as the dollar game, a single special vertex q is designated as the bank, and is allowed to go into debt, taking a negative integer value s(q) < 0. If any other vertex can fire, the bank cannot fire, only collecting chips. Eventually, q will accumulate so many chips that no other vertex can fire: only in such a state, vertex q can fire chips to neighbouring vertices to "jump start the economy".

The set of states which are stable (i.e. for which only q can fire) and recurrent for this game can be given the structure of an abelian group which is isomorphic to the direct product of and the sandpile group (also referred to as Jacobian group or critical group). The order of the latter is the tree number of the graph. [4] [5]

See also

Related Research Articles

In combinatorics, Ramsey's theorem, in one of its graph-theoretic forms, states that one will find monochromatic cliques in any edge labelling (with colours) of a sufficiently large complete graph. To demonstrate the theorem for two colours (say, blue and red), let r and s be any two positive integers. Ramsey's theorem states that there exists a least positive integer R(r, s) for which every blue-red edge colouring of the complete graph on R(r, s) vertices contains a blue clique on r vertices or a red clique on s vertices. (Here R(r, s) signifies an integer that depends on both r and s.)

In mathematics, Hall's marriage theorem, proved by Philip Hall (1935), is a theorem with two equivalent formulations. In each case, the theorem gives a necessary and sufficient condition for an object to exist:

<span class="mw-page-title-main">Petersen graph</span> Cubic graph with 10 vertices and 15 edges

In the mathematical field of graph theory, the Petersen graph is an undirected graph with 10 vertices and 15 edges. It is a small graph that serves as a useful example and counterexample for many problems in graph theory. The Petersen graph is named after Julius Petersen, who in 1898 constructed it to be the smallest bridgeless cubic graph with no three-edge-coloring.

<span class="mw-page-title-main">Bipartite graph</span> Graph divided into two independent sets

In the mathematical field of graph theory, a bipartite graph is a graph whose vertices can be divided into two disjoint and independent sets and , that is, every edge connects a vertex in to one in . Vertex sets and are usually called the parts of the graph. Equivalently, a bipartite graph is a graph that does not contain any odd-length cycles.

Informally, the reconstruction conjecture in graph theory says that graphs are determined uniquely by their subgraphs. It is due to Kelly and Ulam.

<span class="mw-page-title-main">Turán graph</span> Balanced complete multipartite graph

The Turán graph, denoted by , is a complete multipartite graph; it is formed by partitioning a set of vertices into subsets, with sizes as equal as possible, and then connecting two vertices by an edge if and only if they belong to different subsets. Where and are the quotient and remainder of dividing by , the graph is of the form , and the number of edges is

<span class="mw-page-title-main">Spanning tree</span> Tree which includes all vertices of a graph

In the mathematical field of graph theory, a spanning treeT of an undirected graph G is a subgraph that is a tree which includes all of the vertices of G. In general, a graph may have several spanning trees, but a graph that is not connected will not contain a spanning tree. If all of the edges of G are also edges of a spanning tree T of G, then G is a tree and is identical to T.

<span class="mw-page-title-main">Antimatroid</span> Mathematical system of orderings or sets

In mathematics, an antimatroid is a formal system that describes processes in which a set is built up by including elements one at a time, and in which an element, once available for inclusion, remains available until it is included. Antimatroids are commonly axiomatized in two equivalent ways, either as a set system modeling the possible states of such a process, or as a formal language modeling the different sequences in which elements may be included. Dilworth (1940) was the first to study antimatroids, using yet another axiomatization based on lattice theory, and they have been frequently rediscovered in other contexts.

In mathematics, in the areas of order theory and combinatorics, Dilworth's theorem characterizes the width of any finite partially ordered set in terms of a partition of the order into a minimum number of chains. It is named for the mathematician Robert P. Dilworth (1950).

<span class="mw-page-title-main">Degree (graph theory)</span> Number of edges touching a vertex in a graph

In graph theory, the degree of a vertex of a graph is the number of edges that are incident to the vertex; in a multigraph, a loop contributes 2 to a vertex's degree, for the two ends of the edge. The degree of a vertex is denoted or . The maximum degree of a graph , denoted by , and the minimum degree of a graph, denoted by , are the maximum and minimum of its vertices' degrees. In the multigraph shown on the right, the maximum degree is 5 and the minimum degree is 0.

In the mathematical field of graph theory, the Laplacian matrix, also called the graph Laplacian, admittance matrix, Kirchhoff matrix or discrete Laplacian, is a matrix representation of a graph. Named after Pierre-Simon Laplace, the graph Laplacian matrix can be viewed as a matrix form of the negative discrete Laplace operator on a graph approximating the negative continuous Laplacian obtained by the finite difference method.

Graph pebbling is a mathematical game played on a graph with zero or more pebbles on each of its vertices. 'Game play' is composed of a series of pebbling moves. A pebbling move on a graph consists of choosing a vertex with at least two pebbles, removing two pebbles from it, and adding one to an adjacent vertex (the second removed pebble is discarded from play). π(G), the pebbling number of a graph G, is the lowest natural number n that satisfies the following condition:

Given any target or 'root' vertex in the graph and any initial configuration of n pebbles on the graph, it is possible, after a series of pebbling moves, to reach a new configuration in which the designated root vertex has one or more pebbles.

<span class="mw-page-title-main">Rado graph</span> Infinite graph containing all countable graphs

In the mathematical field of graph theory, the Rado graph, Erdős–Rényi graph, or random graph is a countably infinite graph that can be constructed by choosing independently at random for each pair of its vertices whether to connect the vertices by an edge. The names of this graph honor Richard Rado, Paul Erdős, and Alfréd Rényi, mathematicians who studied it in the early 1960s; it appears even earlier in the work of Wilhelm Ackermann (1937). The Rado graph can also be constructed non-randomly, by symmetrizing the membership relation of the hereditarily finite sets, by applying the BIT predicate to the binary representations of the natural numbers, or as an infinite Paley graph that has edges connecting pairs of prime numbers congruent to 1 mod 4 that are quadratic residues modulo each other.

<span class="mw-page-title-main">Handshaking lemma</span> Every graph has evenly many odd vertices

In graph theory, a branch of mathematics, the handshaking lemma is the statement that, in every finite undirected graph, the number of vertices that touch an odd number of edges is even. For example, if there is a party of people who shake hands, the number of people who shake an odd number of other people's hands is even. The handshaking lemma is a consequence of the degree sum formula, also sometimes called the handshaking lemma, according to which the sum of the degrees equals twice the number of edges in the graph. Both results were proven by Leonhard Euler (1736) in his famous paper on the Seven Bridges of Königsberg that began the study of graph theory.

<span class="mw-page-title-main">Directed graph</span> Graph with oriented edges

In mathematics, and more specifically in graph theory, a directed graph is a graph that is made up of a set of vertices connected by directed edges, often called arcs.

<span class="mw-page-title-main">Abelian sandpile model</span> Cellular automaton

The Abelian sandpile model (ASM) is the more popular name of the original Bak–Tang–Wiesenfeld model (BTW). BTW model was the first discovered example of a dynamical system displaying self-organized criticality. It was introduced by Per Bak, Chao Tang and Kurt Wiesenfeld in a 1987 paper.

In graph theory, a partial cube is a graph that is isometric to a subgraph of a hypercube. In other words, a partial cube can be identified with a subgraph of a hypercube in such a way that the distance between any two vertices in the partial cube is the same as the distance between those vertices in the hypercube. Equivalently, a partial cube is a graph whose vertices can be labeled with bit strings of equal length in such a way that the distance between two vertices in the graph is equal to the Hamming distance between their labels. Such a labeling is called a Hamming labeling; it represents an isometric embedding of the partial cube into a hypercube.

The Havel–Hakimi algorithm is an algorithm in graph theory solving the graph realization problem. That is, it answers the following question: Given a finite list of nonnegative integers in non-increasing order, is there a simple graph such that its degree sequence is exactly this list? A simple graph contains no double edges or loops. The degree sequence is a list of numbers in nonincreasing order indicating the number of edges incident to each vertex in the graph. If a simple graph exists for exactly the given degree sequence, the list of integers is called graphic. The Havel-Hakimi algorithm constructs a special solution if a simple graph for the given degree sequence exists, or proves that one cannot find a positive answer. This construction is based on a recursive algorithm. The algorithm was published by Havel (1955), and later by Hakimi (1962).

<span class="mw-page-title-main">Locally linear graph</span> Graph where every edge is in one triangle

In graph theory, a locally linear graph is an undirected graph in which every edge belongs to exactly one triangle. Equivalently, for each vertex of the graph, its neighbors are each adjacent to exactly one other neighbor, so the neighbors can be paired up into an induced matching. Locally linear graphs have also been called locally matched graphs. Their triangles form the hyperedges of triangle-free 3-uniform linear hypergraphs and the blocks of certain partial Steiner triple systems, and the locally linear graphs are exactly the Gaifman graphs of these hypergraphs or partial Steiner systems.

<i>The Mathematics of Chip-Firing</i>

The Mathematics of Chip-Firing is a textbook in mathematics on chip-firing games and abelian sandpile models. It was written by Caroline Klivans, and published in 2018 by the CRC Press.

References

  1. Björner, Anders; Lovász, László; Shor, Peter W. (1991). "Chip-firing games on graphs". European Journal of Combinatorics . 12 (4): 283–291. doi: 10.1016/S0195-6698(13)80111-4 . MR   1120415.Knauer, Kolja (2009). "Chip-firing, antimatroids, and polyhedra". European Conference on Combinatorics, Graph Theory and Applications (EuroComb 2009). Electronic Notes in Discrete Mathematics. Vol. 34. pp. 9–13. doi:10.1016/j.endm.2009.07.002. MR   2591410.
  2. 1 2 Baker, Matthew; Norine, Serguei (2007). "Riemann–Roch and Abel–Jacobi theory on a finite graph". Advances in Mathematics . 215 (2): 766–788. doi: 10.1016/j.aim.2007.04.012 .
  3. Lamboglia, Sara; Ulirsch, Martin. "From the dollar game to the Riemann-Roch Theorem". Snapshots of Modern Mathematics from Oberwolfach. doi:10.14760/SNAP-2021-001-EN.
  4. Biggs, Norman L. (25 June 1997). "Chip-Firing and the Critical Group of a Graph" (PDF). Journal of Algebraic Combinatorics: 25–45. Retrieved 10 May 2014.
  5. wikidot. "Chip-firing references" . Retrieved 19 May 2014.