Reachability

Last updated

In graph theory, reachability refers to the ability to get from one vertex to another within a graph. A vertex can reach a vertex (and is reachable from ) if there exists a sequence of adjacent vertices (i.e. a walk) which starts with and ends with .

Contents

In an undirected graph, reachability between all pairs of vertices can be determined by identifying the connected components of the graph. Any pair of vertices in such a graph can reach each other if and only if they belong to the same connected component; therefore, in such a graph, reachability is symmetric ( reaches iff reaches ). The connected components of an undirected graph can be identified in linear time. The remainder of this article focuses on the more difficult problem of determining pairwise reachability in a directed graph (which, incidentally, need not be symmetric).

Definition

For a directed graph , with vertex set and edge set , the reachability relation of is the transitive closure of , which is to say the set of all ordered pairs of vertices in for which there exists a sequence of vertices such that the edge is in for all . [1]

If is acyclic, then its reachability relation is a partial order; any partial order may be defined in this way, for instance as the reachability relation of its transitive reduction. [2] A noteworthy consequence of this is that since partial orders are anti-symmetric, if can reach , then we know that cannot reach . Intuitively, if we could travel from to and back to , then would contain a cycle, contradicting that it is acyclic. If is directed but not acyclic (i.e. it contains at least one cycle), then its reachability relation will correspond to a preorder instead of a partial order. [3]

Algorithms

Algorithms for determining reachability fall into two classes: those that require preprocessing and those that do not.

If you have only one (or a few) queries to make, it may be more efficient to forgo the use of more complex data structures and compute the reachability of the desired pair directly. This can be accomplished in linear time using algorithms such as breadth first search or iterative deepening depth-first search. [4]

If you will be making many queries, then a more sophisticated method may be used; the exact choice of method depends on the nature of the graph being analysed. In exchange for preprocessing time and some extra storage space, we can create a data structure which can then answer reachability queries on any pair of vertices in as low as time. Three different algorithms and data structures for three different, increasingly specialized situations are outlined below.

Floyd–Warshall Algorithm

The Floyd–Warshall algorithm [5] can be used to compute the transitive closure of any directed graph, which gives rise to the reachability relation as in the definition, above.

The algorithm requires time and space in the worst case. This algorithm is not solely interested in reachability as it also computes the shortest path distance between all pairs of vertices. For graphs containing negative cycles, shortest paths may be undefined, but reachability between pairs can still be noted.

Thorup's Algorithm

For planar digraphs, a much faster method is available, as described by Mikkel Thorup in 2004. [6] This method can answer reachability queries on a planar graph in time after spending preprocessing time to create a data structure of size. This algorithm can also supply approximate shortest path distances, as well as route information.

The overall approach is to associate with each vertex a relatively small set of so-called separator paths such that any path from a vertex to any other vertex must go through at least one of the separators associated with or . An outline of the reachability related sections follows.

Given a graph , the algorithm begins by organizing the vertices into layers starting from an arbitrary vertex . The layers are built in alternating steps by first considering all vertices reachable from the previous step (starting with just ) and then all vertices which reach to the previous step until all vertices have been assigned to a layer. By construction of the layers, every vertex appears at most two layers, and every directed path, or dipath, in is contained within two adjacent layers and . Let be the last layer created, that is, the lowest value for such that .

The graph is then re-expressed as a series of digraphs where each and where is the contraction of all previous levels into a single vertex. Because every dipath appears in at most two consecutive layers, and because each is formed by two consecutive layers, every dipath in appears in its entirety in at least one (and no more than 2 consecutive such graphs)

For each , three separators are identified which, when removed, break the graph into three components which each contain at most the vertices of the original. As is built from two layers of opposed dipaths, each separator may consist of up to 2 dipaths, for a total of up to 6 dipaths over all of the separators. Let be this set of dipaths. The proof that such separators can always be found is related to the Planar Separator Theorem of Lipton and Tarjan, and these separators can be located in linear time.

For each , the directed nature of provides for a natural indexing of its vertices from the start to the end of the path. For each vertex in , we locate the first vertex in reachable by , and the last vertex in that reaches to . That is, we are looking at how early into we can get from , and how far we can stay in and still get back to . This information is stored with each . Then for any pair of vertices and , can reach via if connects to earlier than connects from .

Every vertex is labelled as above for each step of the recursion which builds . As this recursion has logarithmic depth, a total of extra information is stored per vertex. From this point, a logarithmic time query for reachability is as simple as looking over each pair of labels for a common, suitable . The original paper then works to tune the query time down to .

In summarizing the analysis of this method, first consider that the layering approach partitions the vertices so that each vertex is considered only times. The separator phase of the algorithm breaks the graph into components which are at most the size of the original graph, resulting in a logarithmic recursion depth. At each level of the recursion, only linear work is needed to identify the separators as well as the connections possible between vertices. The overall result is preprocessing time with only additional information stored for each vertex.

Kameda's Algorithm

A suitable digraph for Kameda's method with
s
{\displaystyle s}
and
t
{\displaystyle t}
added. Graph suitable for Kameda's method.svg
A suitable digraph for Kameda's method with and added.
The same graph as above after Kameda's algorithm has run, showing the DFS labels for each vertex Kameda's algorithm run.svg
The same graph as above after Kameda's algorithm has run, showing the DFS labels for each vertex

An even faster method for pre-processing, due to T. Kameda in 1975, [7] can be used if the graph is planar, acyclic, and also exhibits the following additional properties: all 0-indegree and all 0-outdegree vertices appear on the same face (often assumed to be the outer face), and it is possible to partition the boundary of that face into two parts such that all 0-indegree vertices appear on one part, and all 0-outdegree vertices appear on the other (i.e. the two types of vertices do not alternate).

If exhibits these properties, then we can preprocess the graph in only time, and store only extra bits per vertex, answering reachability queries for any pair of vertices in time with a simple comparison.

Preprocessing performs the following steps. We add a new vertex which has an edge to each 0-indegree vertex, and another new vertex with edges from each 0-outdegree vertex. Note that the properties of allow us to do so while maintaining planarity, that is, there will still be no edge crossings after these additions. For each vertex we store the list of adjacencies (out-edges) in order of the planarity of the graph (for example, clockwise with respect to the graph's embedding). We then initialize a counter and begin a Depth-First Traversal from . During this traversal, the adjacency list of each vertex is visited from left-to-right as needed. As vertices are popped from the traversal's stack, they are labelled with the value , and is then decremented. Note that is always labelled with the value and is always labelled with . The depth-first traversal is then repeated, but this time the adjacency list of each vertex is visited from right-to-left.

When completed, and , and their incident edges, are removed. Each remaining vertex stores a 2-dimensional label with values from to . Given two vertices and , and their labels and , we say that if and only if , , and there exists at least one component or which is strictly less than or , respectively.

The main result of this method then states that is reachable from if and only if , which is easily calculated in time.

A related problem is to solve reachability queries with some number of vertex failures. For example: "Can vertex still reach vertex even though vertices have failed and can no longer be used?" A similar problem may consider edge failures rather than vertex failures, or a mix of the two. The breadth-first search technique works just as well on such queries, but constructing an efficient oracle is more challenging. [8] [9]

Another problem related to reachability queries is in quickly recalculating changes to reachability relationships when some portion of the graph is changed. For example, this is a relevant concern to garbage collection which needs to balance the reclamation of memory (so that it may be reallocated) with the performance concerns of the running application.

See also

Related Research Articles

In graph theory, a planar graph is a graph that can be embedded in the plane, i.e., it can be drawn on the plane in such a way that its edges intersect only at their endpoints. In other words, it can be drawn in such a way that no edges cross each other. Such a drawing is called a plane graph or planar embedding of the graph. A plane graph can be defined as a planar graph with a mapping from every node to a point on a plane, and from every edge to a plane curve on that plane, such that the extreme points of each curve are the points mapped from its end nodes, and all curves are disjoint except on their extreme points.

In graph theory, an undirected graph H is called a minor of the graph G if H can be formed from G by deleting edges, vertices and by contracting edges.

<span class="mw-page-title-main">Graph coloring</span> Methodic assignment of colors to elements of a graph

In graph theory, graph coloring is a special case of graph labeling; it is an assignment of labels traditionally called "colors" to elements of a graph subject to certain constraints. In its simplest form, it is a way of coloring the vertices of a graph such that no two adjacent vertices are of the same color; this is called a vertex coloring. Similarly, an edge coloring assigns a color to each edge so that no two adjacent edges are of the same color, and a face coloring of a planar graph assigns a color to each face or region so that no two faces that share a boundary have the same color.

<span class="mw-page-title-main">Strongly connected component</span> Partition of a graph whose components are reachable from all vertices

In the mathematical theory of directed graphs, a graph is said to be strongly connected if every vertex is reachable from every other vertex. The strongly connected components of an arbitrary directed graph form a partition into subgraphs that are themselves strongly connected. It is possible to test the strong connectivity of a graph, or to find its strongly connected components, in linear time (that is, Θ(V + E)).

<span class="mw-page-title-main">Chordal graph</span> Graph where all long cycles have a chord

In the mathematical area of graph theory, a chordal graph is one in which all cycles of four or more vertices have a chord, which is an edge that is not part of the cycle but connects two vertices of the cycle. Equivalently, every induced cycle in the graph should have exactly three vertices. The chordal graphs may also be characterized as the graphs that have perfect elimination orderings, as the graphs in which each minimal separator is a clique, and as the intersection graphs of subtrees of a tree. They are sometimes also called rigid circuit graphs or triangulated graphs.

In computer science, a topological sort or topological ordering of a directed graph is a linear ordering of its vertices such that for every directed edge uv from vertex u to vertex v, u comes before v in the ordering. For instance, the vertices of the graph may represent tasks to be performed, and the edges may represent constraints that one task must be performed before another; in this application, a topological ordering is just a valid sequence for the tasks. Precisely, a topological sort is a graph traversal in which each node v is visited only after all its dependencies are visited. A topological ordering is possible if and only if the graph has no directed cycles, that is, if it is a directed acyclic graph (DAG). Any DAG has at least one topological ordering, and algorithms are known for constructing a topological ordering of any DAG in linear time. Topological sorting has many applications especially in ranking problems such as feedback arc set. Topological sorting is possible even when the DAG has disconnected components.

<span class="mw-page-title-main">Dominating set</span> Subset of a graphs nodes such that all other nodes link to at least one

In graph theory, a dominating set for a graph G is a subset D of its vertices, such that any vertex of G is either in D, or has a neighbor in D. The domination numberγ(G) is the number of vertices in a smallest dominating set for G.

<span class="mw-page-title-main">Maximal independent set</span> Independent set which is not a subset of any other independent set

In graph theory, a maximal independent set (MIS) or maximal stable set is an independent set that is not a subset of any other independent set. In other words, there is no vertex outside the independent set that may join it because it is maximal with respect to the independent set property.

<span class="mw-page-title-main">Feedback arc set</span> Edges that hit all cycles in a graph

In graph theory and graph algorithms, a feedback arc set or feedback edge set in a directed graph is a subset of the edges of the graph that contains at least one edge out of every cycle in the graph. Removing these edges from the graph breaks all of the cycles, producing a directed acyclic graph, an acyclic subgraph of the given graph. The feedback arc set with the fewest possible edges is the minimum feedback arc set and its removal leaves the maximum acyclic subgraph; weighted versions of these optimization problems are also used. If a feedback arc set is minimal, meaning that removing any edge from it produces a subset that is not a feedback arc set, then it has an additional property: reversing all of its edges, rather than removing them, produces a directed acyclic graph.

In graph theory, a path decomposition of a graph G is, informally, a representation of G as a "thickened" path graph, and the pathwidth of G is a number that measures how much the path was thickened to form G. More formally, a path-decomposition is a sequence of subsets of vertices of G such that the endpoints of each edge appear in one of the subsets and such that each vertex appears in a contiguous subsequence of the subsets, and the pathwidth is one less than the size of the largest set in such a decomposition. Pathwidth is also known as interval thickness, vertex separation number, or node searching number.

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.

In computer science, fractional cascading is a technique to speed up a sequence of binary searches for the same value in a sequence of related data structures. The first binary search in the sequence takes a logarithmic amount of time, as is standard for binary searches, but successive searches in the sequence are faster. The original version of fractional cascading, introduced in two papers by Chazelle and Guibas in 1986, combined the idea of cascading, originating in range searching data structures of Lueker (1978) and Willard (1978), with the idea of fractional sampling, which originated in Chazelle (1983). Later authors introduced more complex forms of fractional cascading that allow the data structure to be maintained as the data changes by a sequence of discrete insertion and deletion events.

<span class="mw-page-title-main">Laman graph</span>

In graph theory, the Laman graphs are a family of sparse graphs describing the minimally rigid systems of rods and joints in the plane. Formally, a Laman graph is a graph on n vertices such that, for all k, every k-vertex subgraph has at most 2k − 3 edges, and such that the whole graph has exactly 2n − 3 edges. Laman graphs are named after Gerard Laman, of the University of Amsterdam, who in 1970 used them to characterize rigid planar structures. This characterization, however, had already been discovered in 1927 by Hilda Geiringer.

In computer science, a kernelization is a technique for designing efficient algorithms that achieve their efficiency by a preprocessing stage in which inputs to the algorithm are replaced by a smaller input, called a "kernel". The result of solving the problem on the kernel should either be the same as on the original input, or it should be easy to transform the output on the kernel to the desired output for the original problem.

In graph theory, the planar separator theorem is a form of isoperimetric inequality for planar graphs, that states that any planar graph can be split into smaller pieces by removing a small number of vertices. Specifically, the removal of vertices from an n-vertex graph can partition the graph into disjoint subgraphs each of which has at most vertices.

In computer science, a property testing algorithm for a decision problem is an algorithm whose query complexity to its input is much smaller than the instance size of the problem. Typically property testing algorithms are used to distinguish if some combinatorial structure S satisfies some property P, or is "far" from having this property, using only a small number of "local" queries to the object.

In theoretical computer science, the Aanderaa–Karp–Rosenberg conjecture is a group of related conjectures about the number of questions of the form "Is there an edge between vertex and vertex ?" that have to be answered to determine whether or not an undirected graph has a particular property such as planarity or bipartiteness. They are named after Stål Aanderaa, Richard M. Karp, and Arnold L. Rosenberg. According to the conjecture, for a wide class of properties, no algorithm can guarantee that it will be able to skip any questions: any algorithm for determining whether the graph has the property, no matter how clever, might need to examine every pair of vertices before it can give its answer. A property satisfying this conjecture is called evasive.

In computer science, the method of contraction hierarchies is a speed-up technique for finding the shortest-path in a graph. The most intuitive applications are car-navigation systems: a user wants to drive from to using the quickest possible route. The metric optimized here is the travel time. Intersections are represented by vertices, the road sections connecting them by edges. The edge weights represent the time it takes to drive along this segment of the road. A path from to is a sequence of edges ; the shortest path is the one with the minimal sum of edge weights among all possible paths. The shortest path in a graph can be computed using Dijkstra's algorithm but, given that road networks consist of tens of millions of vertices, this is impractical. Contraction hierarchies is a speed-up method optimized to exploit properties of graphs representing road networks. The speed-up is achieved by creating shortcuts in a preprocessing phase which are then used during a shortest-path query to skip over "unimportant" vertices. This is based on the observation that road networks are highly hierarchical. Some intersections, for example highway junctions, are "more important" and higher up in the hierarchy than for example a junction leading into a dead end. Shortcuts can be used to save the precomputed distance between two important junctions such that the algorithm doesn't have to consider the full path between these junctions at query time. Contraction hierarchies do not know about which roads humans consider "important", but they are provided with the graph as input and are able to assign importance to vertices using heuristics.

<span class="mw-page-title-main">Queue number</span> Invariant in graph theory

In the mathematical field of graph theory, the queue number of a graph is a graph invariant defined analogously to stack number using first-in first-out (queue) orderings in place of last-in first-out (stack) orderings.

A central problem in algorithmic graph theory is the shortest path problem. One of the generalizations of the shortest path problem is known as the single-source-shortest-paths (SSSP) problem, which consists of finding the shortest paths from a source vertex to all other vertices in the graph. There are classical sequential algorithms which solve this problem, such as Dijkstra's algorithm. In this article, however, we present two parallel algorithms solving this problem.

References

  1. Skiena, Steven S. (2011), "15.5 Transitive Closure and Reduction", The Algorithm Design Manual (2nd ed.), Springer, pp. 495–497, ISBN   9781848000698 .
  2. Cohn, Paul Moritz (2003), Basic Algebra: Groups, Rings, and Fields, Springer, p. 17, ISBN   9781852335878 .
  3. Schmidt, Gunther (2010), Relational Mathematics, Encyclopedia of Mathematics and Its Applications, vol. 132, Cambridge University Press, p. 77, ISBN   9780521762687 .
  4. Gersting, Judith L. (2006), Mathematical Structures for Computer Science (6th ed.), Macmillan, p. 519, ISBN   9780716768647 .
  5. Cormen, Thomas H.; Leiserson, Charles E.; Rivest, Ronald L.; Stein, Clifford (2001), "Transitive closure of a directed graph", Introduction to Algorithms (2nd ed.), MIT Press and McGraw-Hill, pp. 632–634, ISBN   0-262-03293-7 .
  6. Thorup, Mikkel (2004), "Compact oracles for reachability and approximate distances in planar digraphs", Journal of the ACM , 51 (6): 993–1024, doi:10.1145/1039488.1039493, MR   2145261, S2CID   18864647 .
  7. Kameda, T (1975), "On the vector representation of the reachability in planar directed graphs", Information Processing Letters , 3 (3): 75–77, doi:10.1016/0020-0190(75)90019-8 .
  8. Demetrescu, Camil; Thorup, Mikkel; Chowdhury, Rezaul Alam; Ramachandran, Vijaya (2008), "Oracles for distances avoiding a failed node or link", SIAM Journal on Computing , 37 (5): 1299–1318, CiteSeerX   10.1.1.329.5435 , doi:10.1137/S0097539705429847, MR   2386269 .
  9. Halftermeyer, Pierre, Connectivity in Networks and Compact Labeling Schemes for Emergency Planning, Universite de Bordeaux.