Blossom algorithm

Last updated

In graph theory, the blossom algorithm is an algorithm for constructing maximum matchings on graphs. The algorithm was developed by Jack Edmonds in 1961, [1] and published in 1965. [2] Given a general graph G = (V, E), the algorithm finds a matching M such that each vertex in V is incident with at most one edge in M and |M| is maximized. The matching is constructed by iteratively improving an initial empty matching along augmenting paths in the graph. Unlike bipartite matching, the key new idea is that an odd-length cycle in the graph (blossom) is contracted to a single vertex, with the search continuing iteratively in the contracted graph.

Contents

The algorithm runs in time O(|E||V|2), where |E| is the number of edges of the graph and |V| is its number of vertices. A better running time of for the same task can be achieved with the much more complex algorithm of Micali and Vazirani. [3]

A major reason that the blossom algorithm is important is that it gave the first proof that a maximum-size matching could be found using a polynomial amount of computation time. Another reason is that it led to a linear programming polyhedral description of the matching polytope, yielding an algorithm for min-weight matching. [4] As elaborated by Alexander Schrijver, further significance of the result comes from the fact that this was the first polytope whose proof of integrality "does not simply follow just from total unimodularity, and its description was a breakthrough in polyhedral combinatorics." [5]

Augmenting paths

Given G = (V, E) and a matching M of G, a vertex v is exposed if no edge of M is incident with v. A path in G is an alternating path, if its edges are alternately not in M and in M (or in M and not in M). An augmenting pathP is an alternating path that starts and ends at two distinct exposed vertices. Note that the number of unmatched edges in an augmenting path is greater by one than the number of matched edges, and hence the total number of edges in an augmenting path is odd. A matching augmentation along an augmenting path P is the operation of replacing M with a new matching

.

Edmonds augmenting path.svg

By Berge's lemma, matching M is maximum if and only if there is no M-augmenting path in G. [6] [7] Hence, either a matching is maximum, or it can be augmented. Thus, starting from an initial matching, we can compute a maximum matching by augmenting the current matching with augmenting paths as long as we can find them, and return whenever no augmenting paths are left. We can formalize the algorithm as follows:

   INPUT:  Graph G, initial matching M on G    OUTPUT: maximum matching M* on G A1 functionfind_maximum_matching(G, M) : M* A2     Pfind_augmenting_path(G, M) A3     ifP is non-empty then A4         returnfind_maximum_matching(G, augment M along P) A5     else A6         return M A7     end if A8 end function

We still have to describe how augmenting paths can be found efficiently. The subroutine to find them uses blossoms and contractions.

Blossoms and contractions

Given G = (V, E) and a matching M of G, a blossom B is a cycle in G consisting of 2k + 1 edges of which exactly k belong to M, and where one of the vertices v of the cycle (the base) is such that there exists an alternating path of even length (the stem) from v to an exposed vertex w.

Finding Blossoms:

Define the contracted graphG' as the graph obtained from G by contracting every edge of B, and define the contracted matchingM' as the matching of G' corresponding to M.

Edmonds blossom.svg

G' has an M'-augmenting path if and only if G has an M-augmenting path, and that any M'-augmenting path P' in G' can be lifted to an M-augmenting path in G by undoing the contraction by B so that the segment of P' (if any) traversing through vB is replaced by an appropriate segment traversing through B. [8] In more detail:

Edmonds lifting path.svg

Edmonds lifting end point.svg

Thus blossoms can be contracted and search performed in the contracted graphs. This reduction is at the heart of Edmonds' algorithm.

Finding an augmenting path

The search for an augmenting path uses an auxiliary data structure consisting of a forest F whose individual trees correspond to specific portions of the graph G. In fact, the forest F is the same that would be used to find maximum matchings in bipartite graphs (without need for shrinking blossoms). In each iteration the algorithm either (1) finds an augmenting path, (2) finds a blossom and recurses onto the corresponding contracted graph, or (3) concludes there are no augmenting paths. The auxiliary structure is built by an incremental procedure discussed next. [8]

The construction procedure considers vertices v and edges e in G and incrementally updates F as appropriate. If v is in a tree T of the forest, we let root(v) denote the root of T. If both u and v are in the same tree T in F, we let distance(u,v) denote the length of the unique path from u to v in T.

    INPUT:  Graph G, matching M on G     OUTPUT: augmenting path P in G or empty path if none found B01 functionfind_augmenting_path(G, M) : P B02    F ← empty forest B03    unmark all vertices and edges in G, mark all edges of M B05    for each exposed vertex vdo B06        create a singleton tree { v } and add the tree to F B07    end for B08    while there is an unmarked vertex v in F with distance(v, root(v)) even do B09        while there exists an unmarked edge e = { v, w } do B10            ifw is not in Fthen                    // w is matched, so add e and w's matched edge to F B11                x ← vertex matched to w in M B12                add edges { v, w } and { w, x } to the tree of v B13            else B14                ifdistance(w, root(w)) is odd then                        // Do nothing. B15                else B16                    ifroot(v)root(w)then                            // Report an augmenting path in F  { e }. B17                        P ← path (root(v) → ... → v) → (w → ... → root(w)) B18                        returnP B19                    else                            // Contract a blossom in G and look for the path in the contracted graph. B20                        B ← blossom formed by e and edges on the path vw in T B21                        G’, M’ ← contract G and M by B B22                        P’find_augmenting_path(G’, M’) B23                        P ← lift P’ to G B24                        returnP B25                    end if B26                end if B27            end if B28            mark edge e B29        end while B30        mark vertex v B31    end while B32    return empty path B33 end function

Examples

The following four figures illustrate the execution of the algorithm. Dashed lines indicate edges that are currently not present in the forest. First, the algorithm processes an out-of-forest edge that causes the expansion of the current forest (lines B10 B12).

Forest expansion.png

Next, it detects a blossom and contracts the graph (lines B20 B21).

Blossom contraction.png

Finally, it locates an augmenting path P′ in the contracted graph (line B22) and lifts it to the original graph (line B23). Note that the ability of the algorithm to contract blossoms is crucial here; the algorithm cannot find P in the original graph directly because only out-of-forest edges between vertices at even distances from the roots are considered on line B17 of the algorithm.

Path detection.png

Path lifting.png

Analysis

The forest F constructed by the find_augmenting_path() function is an alternating forest. [9]

Each iteration of the loop starting at line B09 either adds to a tree T in F (line B10) or finds an augmenting path (line B17) or finds a blossom (line B20). It is easy to see that the running time is .

Bipartite matching

When G is bipartite, there are no odd cycles in G. In that case, blossoms will never be found and one can simply remove lines B20 B24 of the algorithm. The algorithm thus reduces to the standard algorithm to construct maximum cardinality matchings in bipartite graphs [7] where we repeatedly search for an augmenting path by a simple graph traversal: this is for instance the case of the Ford–Fulkerson algorithm.

Weighted matching

The matching problem can be generalized by assigning weights to edges in G and asking for a set M that produces a matching of maximum (minimum) total weight: this is the maximum weight matching problem. This problem can be solved by a combinatorial algorithm that uses the unweighted Edmonds's algorithm as a subroutine. [6] Kolmogorov provides an efficient C++ implementation of this. [10]

Related Research Articles

<span class="mw-page-title-main">Minimum spanning tree</span> Least-weight tree connecting graph vertices

A minimum spanning tree (MST) or minimum weight spanning tree is a subset of the edges of a connected, edge-weighted undirected graph that connects all the vertices together, without any cycles and with the minimum possible total edge weight. That is, it is a spanning tree whose sum of edge weights is as small as possible. More generally, any edge-weighted undirected graph has a minimum spanning forest, which is a union of the minimum spanning trees for its connected components.

<span class="mw-page-title-main">Breadth-first search</span> Algorithm to search the nodes of a graph

Breadth-first search (BFS) is an algorithm for searching a tree data structure for a node that satisfies a given property. It starts at the tree root and explores all nodes at the present depth prior to moving on to the nodes at the next depth level. Extra memory, usually a queue, is needed to keep track of the child nodes that were encountered but not yet explored.

In computer science, the Edmonds–Karp algorithm is an implementation of the Ford–Fulkerson method for computing the maximum flow in a flow network in time. The algorithm was first published by Yefim Dinitz in 1970 and independently published by Jack Edmonds and Richard Karp in 1972. Dinic's algorithm includes additional techniques that reduce the running time to .

<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.

This is a glossary of graph theory. Graph theory is the study of graphs, systems of nodes or vertices connected in pairs by lines or edges.

<span class="mw-page-title-main">Maximum flow problem</span> Computational problem in graph theory

In optimization theory, maximum flow problems involve finding a feasible flow through a flow network that obtains the maximum possible flow rate.

In the mathematical discipline of graph theory, a matching or independent edge set in an undirected graph is a set of edges without common vertices. In other words, a subset of the edges is a matching if each vertex appears in at most one edge of that matching. Finding a matching in a bipartite graph can be treated as a network flow problem.

<span class="mw-page-title-main">Edge coloring</span> Problem of coloring a graphs edges such that meeting edges do not match

In graph theory, a proper edge coloring of a graph is an assignment of "colors" to the edges of the graph so that no two incident edges have the same color. For example, the figure to the right shows an edge coloring of a graph by the colors red, blue, and green. Edge colorings are one of several different types of graph coloring. The edge-coloring problem asks whether it is possible to color the edges of a given graph using at most k different colors, for a given value of k, or with the fewest possible colors. The minimum required number of colors for the edges of a given graph is called the chromatic index of the graph. For example, the edges of the graph in the illustration can be colored by three colors but cannot be colored by two colors, so the graph shown has chromatic index three.

The Hungarian method is a combinatorial optimization algorithm that solves the assignment problem in polynomial time and which anticipated later primal–dual methods. It was developed and published in 1955 by Harold Kuhn, who gave it the name "Hungarian method" because the algorithm was largely based on the earlier works of two Hungarian mathematicians, Dénes Kőnig and Jenő Egerváry. However, in 2006 it was discovered that Carl Gustav Jacobi had solved the assignment problem in the 19th century, and the solution had been published posthumously in 1890 in Latin.

In graph theory, the Dulmage–Mendelsohn decomposition is a partition of the vertices of a bipartite graph into subsets, with the property that two adjacent vertices belong to the same subset if and only if they are paired with each other in a perfect matching of the graph. It is named after A. L. Dulmage and Nathan Mendelsohn, who published it in 1958. A generalization to any graph is the Edmonds–Gallai decomposition, using the Blossom algorithm.

<span class="mw-page-title-main">Kőnig's theorem (graph theory)</span> Theorem showing that maximum matching and minimum vertex cover are equivalent for bipartite graphs

In the mathematical area of graph theory, Kőnig's theorem, proved by Dénes Kőnig (1931), describes an equivalence between the maximum matching problem and the minimum vertex cover problem in bipartite graphs. It was discovered independently, also in 1931, by Jenő Egerváry in the more general case of weighted graphs.

In computer science, the Hopcroft–Karp algorithm is an algorithm that takes a bipartite graph as input and produces a maximum-cardinality matching as output — a set of as many edges as possible with the property that no two edges share an endpoint. It runs in time in the worst case, where is set of edges in the graph, is set of vertices of the graph, and it is assumed that . In the case of dense graphs the time bound becomes , and for sparse random graphs it runs in time with high probability.

Maximum cardinality matching is a fundamental problem in graph theory. We are given a graph G, and the goal is to find a matching containing as many edges as possible; that is, a maximum cardinality subset of the edges such that each vertex is adjacent to at most one edge of the subset. As each edge will cover exactly two vertices, this problem is equivalent to the task of finding a matching that covers as many vertices as possible.

<span class="mw-page-title-main">Factor-critical graph</span> Graph of n vertices with a perfect matching for every subgraph of n-1 vertices

In graph theory, a mathematical discipline, a factor-critical graph is a graph with n vertices in which every subgraph of n − 1 vertices has a perfect matching.

<span class="mw-page-title-main">Claw-free graph</span> Graph without four-vertex star subgraphs

In graph theory, an area of mathematics, a claw-free graph is a graph that does not have a claw as an induced subgraph.

In graph theory, a branch of mathematics, a skew-symmetric graph is a directed graph that is isomorphic to its own transpose graph, the graph formed by reversing all of its edges, under an isomorphism that is an involution without any fixed points. Skew-symmetric graphs are identical to the double covering graphs of bidirected graphs.

In theoretical computer science and network routing, Suurballe's algorithm is an algorithm for finding two disjoint paths in a nonnegatively-weighted directed graph, so that both paths connect the same pair of vertices and have minimum total length. The algorithm was conceived by John W. Suurballe and published in 1974. The main idea of Suurballe's algorithm is to use Dijkstra's algorithm to find one path, to modify the weights of the graph edges, and then to run Dijkstra's algorithm a second time. The output of the algorithm is formed by combining these two paths, discarding edges that are traversed in opposite directions by the paths, and using the remaining edges to form the two paths to return as the output. The modification to the weights is similar to the weight modification in Johnson's algorithm, and preserves the non-negativity of the weights while allowing the second instance of Dijkstra's algorithm to find the correct second path.

The expected linear time MST algorithm is a randomized algorithm for computing the minimum spanning forest of a weighted graph with no isolated vertices. It was developed by David Karger, Philip Klein, and Robert Tarjan. The algorithm relies on techniques from Borůvka's algorithm along with an algorithm for verifying a minimum spanning tree in linear time. It combines the design paradigms of divide and conquer algorithms, greedy algorithms, and randomized algorithms to achieve expected linear performance.

<span class="mw-page-title-main">Gallai–Edmonds decomposition</span> Partition of the vertices of a graph giving information on the structure of maximum matchings

In graph theory, the Gallai–Edmonds decomposition is a partition of the vertices of a graph into three subsets which provides information on the structure of maximum matchings in the graph. Tibor Gallai and Jack Edmonds independently discovered it and proved its key properties.

In graph theory, a Hall violator is a set of vertices in a graph, that violate the condition to Hall's marriage theorem.

References

  1. Edmonds, Jack (1991), "A glimpse of heaven", in J.K. Lenstra; A.H.G. Rinnooy Kan; A. Schrijver (eds.), History of Mathematical Programming --- A Collection of Personal Reminiscences, CWI, Amsterdam and North-Holland, Amsterdam, pp. 32–54
  2. Edmonds, Jack (1965). "Paths, trees, and flowers". Can. J. Math. 17: 449–467. doi: 10.4153/CJM-1965-045-4 .
  3. Micali, Silvio; Vazirani, Vijay (1980). An O(V1/2E) algorithm for finding maximum matching in general graphs. 21st Annual Symposium on Foundations of Computer Science. IEEE Computer Society Press, New York. pp. 17–27.
  4. Edmonds, Jack (1965). "Maximum matching and a polyhedron with 0,1-vertices". Journal of Research of the National Bureau of Standards Section B. 69: 125–130. doi: 10.6028/jres.069B.013 .
  5. Schrijver, Alexander (2003). Combinatorial Optimization: Polyhedra and Efficiency. Algorithms and Combinatorics. Berlin Heidelberg: Springer-Verlag. ISBN   9783540443896.
  6. 1 2 Lovász, László; Plummer, Michael (1986). Matching Theory. Akadémiai Kiadó. ISBN   963-05-4168-8.
  7. 1 2 Karp, Richard, "Edmonds's Non-Bipartite Matching Algorithm", Course Notes. U. C. Berkeley (PDF), archived from the original (PDF) on 2008-12-30
  8. 1 2 Tarjan, Robert, "Sketchy Notes on Edmonds' Incredible Shrinking Blossom Algorithm for General Matching", Course Notes, Department of Computer Science, Princeton University (PDF)
  9. Kenyon, Claire; Lovász, László, "Algorithmic Discrete Mathematics", Technical Report CS-TR-251-90, Department of Computer Science, Princeton University
  10. Kolmogorov, Vladimir (2009), "Blossom V: A new implementation of a minimum cost perfect matching algorithm", Mathematical Programming Computation, 1 (1): 43–67, doi:10.1007/s12532-009-0002-8